Recently, the spectral localizer framework has emerged as an efficient approach for classifying topology in photonic systems featuring local nonlinearities and radiative environments. In nonlinear systems, this framework provides rigorous definitions for concepts such as topological solitons and topological dynamics, where a system’s occupation induces a local change in its topology due to nonlinearity. For systems embedded in radiative environments that do not possess a shared bulk spectral gap, this framework enables the identification of local topology and shows that local topological protection is preserved despite the lack of a common gap. However, as the spectral localizer framework is rooted in the mathematics of C*-algebras, and not vector bundles, understanding and using this framework requires developing intuition for a somewhat different set of underlying concepts than those that appear in traditional approaches for classifying material topology. In this tutorial, we introduce the spectral localizer framework from a ground-up perspective and provide physically motivated arguments for understanding its local topological markers and associated local measure of topological protection. In doing so, we provide numerous examples of the framework’s application to a variety of topological classes, including crystalline and higher-order topology. We then show how Maxwell’s equations can be reformulated to be compatible with the spectral localizer framework, including the possibility of radiative boundary conditions. To aid in this introduction, we also provide a physics-oriented introduction to multi-operator pseudospectral methods and numerical K-theory, two mathematical concepts that form the foundation for the spectral localizer framework. Finally, we provide some mathematically oriented comments on the C*-algebraic origins of this framework, including a discussion of real C*-algebras and graded C*-algebras that are necessary for incorporating physical symmetries. Looking forward, we hope that this tutorial will serve as an approachable starting point for learning the foundations of the spectral localizer framework.
I. INTRODUCTION
Since the seminal works of Haldane and Raghu1,2 demonstrated that topological phenomena can manifest in any material system governed by a wave equation, the ideas of topological physics have excited the photonics community,3 both to use photonic platforms to explore new fundamental concepts and for leveraging topology’s benefits in photonic devices. In particular, topological photonic systems are both guaranteed to exhibit localized states at their boundaries or corners, and these states’ existence are robust against fabrication imperfections, yielding an enticing suite of properties for enhancing light–matter interactions4 and routing quantum information.5 Early, and still ongoing, efforts in the field of topological photonics have focused on finding systems that realize non-reciprocal topological phenomena at technologically relevant wavelengths6–18 despite the present lack of materials with a strong magneto-optical response at those same wavelengths. In addition, the community has also explored other classes of topology that do not require breaking time-reversal symmetry,19–42 but whose protected states are either reciprocal or zero-dimensional cavity states. More recently, the field has begun to shift its focus toward using topology in photonic devices, creating topological lasers43–51 as well as systems for controlling quantum light.23,24,33,34,38,40,41,52
Traditionally, topological phenomena are identified via invariants calculated using the band structure and Bloch eigenstates of a gapped (i.e., insulating) crystalline system, and these invariants cannot change without first closing the bulk bandgap.53,54 As such, a system’s invariants are protected against disorder that is not strong enough to close the bandgap. Moreover, bulk–boundary correspondence55,56 guarantees that a system with a non-trivial invariant exhibits some set of boundary- or corner-localized states. Thus, topological band theory is ideally suited to predicting the edge transport properties of large systems without needing to ensure any particular edge termination and regardless of strong disorder along the boundary or interface. In other words, such an approach is ideal for predicting the behavior of condensed matter systems.
However, the frontiers of topological photonics are diverging from those of condensed matter physics due to the different physical phenomena available in, as well as the different desired applications of, each platform. In particular, photonic systems can exhibit local mean-field nonlinearities,57 arbitrarily tailorable geometry that enables small system sizes,58,59 and non-Hermiticity both through radiative losses as well as gain and absorption.60–63 Moreover, these phenomena are central to many photonic devices seeking to leverage topological phenomena. For example, it can be advantageous to reduce a device’s on-chip footprint, which necessitates understanding finite size effects, or make surface emitting light sources, which requires the inclusion of radiative losses to a gapless medium. Unfortunately, photonic systems exhibiting these phenomena are ill-suited to be classified using topological band theory, either because the effects are local, or the system is small, such that a band theoretic description is not applicable; or because the full photonic system, including the radiative environment, lacks a spectral gap so band theory would not predict any global topological protection.
In the past few years, a topological classification framework based on the system’s spectral localizer has emerged,64–66 which is able to identify photonic and material topology in realistic systems67,68 regardless of whether a system has radiative boundaries69,70 or local nonlinearities.71 This framework provides a position-space picture of topology, yielding spatially resolved local topological markers and a local measure of topological protection. As such, the spectral localizer framework offers the possibility of aiding in the advancement of the field of topological photonics as it explores nonlinearities and finite system-size effects. However, the spectral localizer and its associated local markers are rooted in the mathematics of C*-algebras and cannot be expressed in the theories of geometry and vector bundles that are traditionally used to describe material topology. As such, the relevant formulas for using the spectral localizer framework initially appear quite foreign and do not suggest an immediate physical interpretation despite their utility. Moreover, the development of the spectral localizer framework is spread out over the physics and mathematics literature, yielding a daunting task for anyone interested in advancing its underlying theory or making use of the approach.
In this tutorial, we provide a physically motivated description of the spectral localizer framework and show examples of its use across a variety of different classes of topology, including those from the Altland–Zirnbauer classification72–74 and those of crystalline origin. In doing so, we show both how this framework provides quantitative predictions of a system’s topological robustness and how bulk–boundary correspondence manifests. As part of this tutorial, we also introduce two main mathematical topics key to the spectral localizer: multi-operator pseudospectral methods and numerical K-theory. Multi-operator pseudospectral methods are an approach to finding approximate joint eigenvectors for non-commuting operators, and these methods form the basis for understanding bulk–boundary correspondence in the spectral localizer framework, although they have utility beyond the study of topological materials. Numerical K-theory is the study of deriving formula for topological invariants that yield efficient computational implementations; the local markers calculated using the spectral localizer are an example of numerical K-theory. Looking forward, we hope that this tutorial provides an approachable on-ramp for members of both the physics and mathematics communities to use the spectral localizer framework to classify topological systems.
A. Preliminaries and terminology
Overall, this tutorial assumes that the reader has some familiarity with the study of topological physics, but makes no assumptions on the reader’s mathematical background beyond a standard introduction to linear algebra and the usual facts about linear operators and Hilbert spaces used in modern physics. Proofs for many mathematical details are instead provided in the associated references. The only exception to this assumption is in the self-contained Sec. VI, where some of the C*-algebraic underpinnings of the spectral localizer framework are presented in a form for a reader with a mathematically oriented background.
There are also two overarching terminology issues that arise in the study of topological photonics. First, this tutorial uses the phrase material topology, or variants of this, to refer to the possible topological behavior of any natural or artificial material, regardless of whether it is comprised of atoms, molecules, artificial atoms, or any other form of decoration that serves as a change in a system’s spatial potential. Second, this tutorial will use the condensed matter physics terminology of a system’s occupied states or bands to refer to those states or bands of a system whose eigenenergies are less than some chosen energy of interest. Of course, in natural materials whose topology arises in their many-body electron wave function, the energy of interest is the Fermi energy and all the (single particle) states with energies less than this are necessarily occupied at zero temperature. In contrast, in photonic systems, a structure’s topological phenomena can be observed by exciting the system at a single frequency, without carefully populating the system at other frequencies. Indeed, as photons do not directly interact, their system Hamiltonians are linear (for linear materials) and do not change based on the system’s occupation. Thus, a photonic system may exhibit several disparate spectral ranges with different non-trivial topologies, any of which may be experimentally accessible. Nevertheless, as many formulas for topological invariants at a given energy E are defined in terms of a system’s states |ψm⟩ with Em ≤ E, it is useful to retain terminology that refers to this set of lower-energy states.
B. Structure of the tutorial
In Sec. II, we provide a brief review of topological band theory and related position-space topological invariants. In Sec. III, we first provide an intuitive picture for building up the spectral localizer framework, including its local markers and measure of protection. This section then provides a set of general definitions for the framework and provides a number of examples across a variety of different classes of topology and system dimensions. In Sec. IV, we provide full discussions of the hyper-parameter κ that appears in the spectral localizer framework, as well as the manifestation of bulk–boundary correspondence, and efficient numerical implementations through numerical K-theory. In Sec. V, we show how the spectral localizer framework can be applied to photonic systems and discuss some of the challenges associated with considering systems described by differential operators such as Maxwell’s equations. In Sec. VI, we provide some self-contained comments on the mathematical background of the spectral localizer framework. Finally, in Sec. VII, we provide some concluding remarks. Throughout the tutorial, we also directly acknowledge open questions (marked as such) and other areas that require further study.
II. BRIEF REVIEW OF TOPOLOGICAL CLASSIFICATION THEORIES
In this section, we provide a very brief review of standard methods for classifying material topology: topological band theory, global position-space invariants, and local topological markers. This review is intended as a reminder, not a detailed introduction (instead see Refs. 53,54 and 75,76), and is included so that similarities and differences of these methods with the spectral localizer framework can be highlighted in later sections. As such, this brief review focuses solely on how each framework identifies Chern materials (2D class A in the Altland–Zirnbauer classification table72–74), as the distinctions between these approaches are similar across every topological class. Similarly, this section may also be safely skipped if a reader is already familiar with, or uninterested in, this background.
A. Topological band theory
Traditionally, material topology is classified using topological invariants defined in terms of quantities derived from a system’s band structure, such as the Chern number,77 winding number,54 or multipole moment.78,79 Such band theory-based invariants are only defined for materials that possess a bulk bandgap (i.e., insulators), and these invariants can only change when a sufficiently strong perturbation is added to the system that closes the bulk bandgap. Thus, by definition, standard band theoretic topological invariants are global properties of a crystalline material, as they assume the material is infinite and periodic so that Bloch’s theorem can be applied. Materials possessing non-trivial bulk topological invariants necessarily exhibit associated boundary-localized states, whose appearance is guaranteed by bulk–boundary correspondence.80–82
B. Global position-space invariants
It is also possible to classify material topology using a system’s position-space description, rather than its momentum-space description. Heuristically, such an approach must be possible, as the two descriptions are related by a Fourier transform, and Fourier transforms neither add nor remove information,83 they simply rearrange it. Thus, the aspects of a crystalline material that give rise to non-trivial topology, as determined by topological band theory, must also be detectable using invariants that leverage a system’s position-space Hamiltonian H rather than its Bloch periodic Hamiltonian H(k). In general, there are two different approaches to classifying material topology in position space: global invariants and local markers. Global position-space invariants always make use of a system Hamiltonian with periodic boundary conditions (PBC), but this Hamiltonian represents a large volume of the material rather than a single unit cell under Bloch periodic boundary conditions as is used by topological band theory. Theories of local topological markers can be constructed using either systems with open boundary conditions (OBC) or PBC, and classify the system’s topology at a chosen location.
When calculating a periodic material’s topology using Eq. (2), the size of H that is typically required (before reaching the periodic boundary condition) to guarantee accurate classification is related to ensuring that the material’s topological band inversion can be resolved. In other words, a finite H with PBC is sampling wavevector space every δkj = 2π/lj, where lj is the length of the system in the jth direction, so lj must be large enough to ensure the region of wavevectors where the bands have inverted90 is sufficiently sampled by the Bott index. Thus, heuristically, the Bott index is trading repeated calculations of the Bloch Hamiltonian H(k)’s spectrum and eigenstates at different k that are required to perform the full integral in Eq. (1) with a single calculation of the spectrum and eigenstates of the larger H. In addition, the matrices UX and UY are generally dense.
C. Local topological markers
In contrast to topological band theory and global position-space invariants, local topological markers are calculated at a specified position x, as well as a specified energy. Thus, a non-uniform system’s local markers can vary across the system, indicating regions with different material topology. As such, theories of local markers can be applied to aperiodic materials, such as quasicrystals91–94 and amorphous structures,95,96 as well as disordered systems or heterostructures,97 without alteration. The first local marker was derived by Kitaev in 2006,98 and Bianco and Resta produced a seminal study on the topic in 2011.99 Both of these initial local markers identify Chern topology in finite systems with open boundaries, but a variety of local markers for different classes of topology have been subsequently derived.100–116
Despite their different formulations, the Kitaev marker and the Bianco-Resta marker both exhibit similar properties. Unlike band theoretic invariants or global position-space invariants, neither are guaranteed to be integers for finite choices of integration areas [i.e., A, B, and C for Eq. (4) or D for Eq. (8)], although both converge to the material’s Chern number as the integration area is increased. Moreover, both local markers, as well as the Bott index, typically involve calculating all of a system’s eigenstates with energies below E, which can be numerically expensive given that these formulations require using Hamiltonians that represent large volumes of material. In some cases, it is possible to circumvent finding a system’s states through the kernel polynomial method to achieve more efficient algorithms,118 although this approach’s efficiency gains decrease as the bulk spectral gap decreases. Finally, none come equipped with an independent measure of the system’s topological protection and instead default to defining this protection in terms of the width of the relevant spectral gap.
III. INTRODUCTION TO THE SPECTRAL LOCALIZER FRAMEWORK
The spectral localizer framework can be thought of as a mathematical probe for understanding a material’s properties near the probe’s location, as shown in Fig. 2. To classify material topology, the framework yields a constellation of local markers that can identify a broad range of material topology, including all ten Altland-Zirnbauer classes in every physical dimension,64–66 crystalline topology,68 Weyl semimetals,119 and some forms of non-Hermitian topology.120–122 In addition, the framework provides a quantitative measure of topological protection. In contrast to other theories of local markers, the spectral localizer framework’s markers are guaranteed to be integer valued, although the theory still requires a choice of hyper-parameter that plays a similar role to the choice of integration area required for the Kitaev and Bianco-Resta markers (see Sec. IV A). Computationally, a key property of the spectral localizer framework is that it does not require finding a system’s eigenstates, nor make use of a system’s projector onto an occupied subspace and thus can leverage significant numerical speedups using matrix factorization techniques (see Sec. IV D). The theory was originally discovered by Loring in 2015,64 with substantial subsequent mathematical developments by Loring and Schulz-Baldes.65
In the remainder of this section, we first provide an argument for how the spectral localizer framework classifies Chern topology for 2D materials based on an analysis of the position-space behavior of atomic limits and discuss how a measure of topological protection naturally appears. In Sec. III C, we provide the general definition of the spectral localizer and discuss how it can classify topology in other symmetry classes. Later sections give further discussion on how the spectral localizer framework is applied to odd-dimensional systems, as well as how the framework can be applied to classify crystalline topology and higher-order topology. Finally, Sec. III G examines how the spectral localizer framework is generalized to the thermodynamic limit.
A. Intuitive picture of the spectral localizer framework for identifying 2D Chern materials
Before proceeding to a more formal introduction, here we begin by introducing the spectral localizer framework from a ground-up perspective for identifying Chern topology in 2D materials. The goal is to provide a complete intuitive picture of this framework, while pushing some details to later sections. Although there are different ways to understand why the spectral localizer framework can successfully classify material topology, as shown in Sec. VI, this section introduces the spectral localizer framework as a method for diagnosing whether a given system can be connected to an atomic limit without closing a spectral gap or violating a relevant symmetry.
An important concept in the modern study of topological materials is the idea of an atomic limit—the limit in which the constituent elements of a material are decoupled into individual atoms, molecules, or meta-atoms so that the system is simply a collection of those isolated elements. A crystal in an atomic limit exhibits a band structure that is completely flat.73 Atomic limits also possess a complete basis of localized Wannier functions123 that exhibit all of the same symmetries as the underlying system. Atomic limits are intimately connected to material topology because topologically non-trivial systems either do not possess a localized Wannier basis or their localized Wannier basis does not obey all of the symmetries of the original material. For example, 2D systems with non-zero Chern numbers do not possess a localized Wannier basis,124 which, in crystalline insulators, is a direct consequence of the fact that such systems have an obstruction that prohibits the choice of a smooth gauge for the Bloch wavefunctions across the (first) Brillouin zone. Similarly, electronic systems that exhibit the quantum spin Hall effect, a form of topology protected by fermionic time-reversal symmetry, possess a localized Wannier basis, but this basis does not obey fermionic time-reversal symmetry.125
Crucially, these statements are bijections, e.g., if a 2D material is found to not possess a localized Wannier basis, it necessarily has a non-zero Chern number. If a material does not possess a localized Wannier basis that exhibits all the same symmetries as the original system, the system is generally non-trivial with respect to a class of topology protected by those symmetries missing from the Wannier basis. In either case, the material cannot be continued to an atomic limit without breaking a symmetry or closing a spectral gap (which would allow the system to change its topology). Here, by “continued to an atomic limit,” one is asking whether a path of matrices Hτ with τ ∈ [0, 1] can be found, with H0 = H being the original system’s Hamiltonian and H1 = H(AL) being an atomic limit, such that every Hτ possesses a bulk spectral gap at the chosen energy E, remains somewhat local73 (i.e., sites separated by a sufficiently large distance are not coupled), and obeys the same symmetries as H (e.g., if a system is chiral symmetric with HΠ = −ΠH, then for every τ, HτΠ = −ΠHτ). As such, it is possible to flip the paradigm for identifying topological systems: rather than finding materials, calculating their topological invariants, and inferring the existence and properties of the system’s localized Wannier basis; one can instead choose a material, ask if it exhibits a localized Wannier basis that obeys all the same symmetries as the original material and infer its topology. Indeed, the recently developed framework for classifying crystalline materials of topological quantum chemistry126–133 takes precisely this approach, using a material’s band structure to understand whether it can be continued to an atomic limit and then inferring its topological properties.
However, it is also possible to ascertain whether a system can be continued to an atomic limit directly from position-space description, rather than through its band structure or Wannier basis. The key mathematical observation for this shift in perspective follows from the definition of an atomic limit: since all of an atomic limit’s constituent atoms or molecules are decoupled, the system does not possess any kinetic energy associated with this decoration-to-decoration coupling; moreover, the spacing between adjacent atoms or molecules is assumed to be large compared to the spacing between the elements within a single molecule. Thus, the Hamiltonian of an atomic limit commutes with its associated position operators , i.e., ; see Fig. 3. As such, from a position-space perspective, the question of whether a given system can be continued to an atomic limit is equivalent to asking whether the non-commuting Hamiltonian H and position operators Xj of the original system, [H, Xj] ≠ 0, can nevertheless be path continued via some set of Hτ and Xj,τ to be commuting while preserving all of the relevant symmetries and the relevant bulk spectral gap. It should be noted that in this position-space picture, the effect of symmetries must be considered on both Hτ and Xj,τ; while local symmetries such as time-reversal trivially commute with Xj,τ, crystalline symmetries might not, and the path of Xj,τ must preserve whatever the original relationship is. In addition, at this juncture, we are being purposefully vague about how to guarantee that the position-space path preserves the relevant bulk spectral gap; for a finite system with OBC, edge effects, topological or otherwise, as well as internal defects may result in H not exhibiting any global spectral gap even if its ordered crystalline counterpart exhibits bulk bandgaps. Exactly what spectral gap must be preserved by Hτ and Xj,τ will be made rigorous in Sec. III B.
Having reduced the question of classifying material topology to determining whether a set of non-commuting matrices can be appropriately path-continued to be commuting, a method must be argued for actually performing this determination. For a 2D Chern insulator, we can understand whether a given material can be path continued to the atomic limit using two theorems from the mathematics literature. The first theorem provides a way to identify the homotopy class of an invertible Hermitian matrix, i.e., it provides a method for identifying whether two invertible Hermitian matrices can be path connected.
Two n-by-n invertible Hermitian matrices L and L′ can be connected by a path of invertible Hermitian matrices if and only if sig[L] = sig[L′], where sig[L] is the signature of L, its number of positive eigenvalues minus its number of negative eigenvalues. (See Appendix for proof.)
While this theorem uses language that is not standard for photonics, the result is quite intuitive when depicted graphically: any attempt to connect the spectra of two invertible Hermitian matrices with different signatures will necessarily force at least one eigenvalue in the connecting path of Hermitian matrices Lτ to be 0, at which point that Lτ is not invertible, as shown in Fig. 4. With knowledge of what matrices can be connected, the second theorem then defines which of these homotopy classes contain atomic limits.
Overall, the argument presented in this section shows how the 2D spectral localizer can be used to classify whether a material at a given choice of (x, E) exhibits non-trivial local Chern topology by understanding that if a system can locally be path continued to an atomic limit. In addition, the calculation of a system’s local Chern marker can be computationally quite efficient and need not involve finding any eigenvalues of by taking advantage of sparse matrix factorization techniques; see Sec. IV D.
B. Defining the local gap
In Sec. III A, we argued that for a system to be topologically trivial, the path of matrices Hτ and Xj,τ connecting a given system to an atomic limit need to both respect all of the original system’s (relevant) symmetries and maintain a spectral gap. Then, we stated that the 2D spectral localizer must be applied to finite systems with OBC. However, this seems to present a problem: the Hamiltonian of a topological system with OBC will not generally exhibit a spectral gap; such a gap will instead by populated by boundary-localized states. Moreover, even topologically trivial systems may exhibit edge effects or possess internal defects that preclude its Hamiltonian from exhibiting a spectral gap.
Generally, when using the spectral localizer to classify the topology of a periodic systems with a bulk bandgap, the topological protection predicted by for x chosen in a system’s bulk and E in its spectral gap is quantitatively similar to the topological protection predicted by its bandgap.67 In particular, if a system’s bulk spectral gap is between a lower energy El and an upper energy Eu, is typically within of min[E − El, Eu − E]. However, for systems or heterostructures lacking bulk spectral gaps, the spectral localizer framework can still predict non-zero values for topological protection.69,137,138
To summarize the picture of the spectral localizer framework, for a given system defined by H and X, one can first construct the spectral localizer and then iterate over different choices of position and energy to map out the system’s local topology and corresponding topological protection. As L(x,E) is usually sparse, this iterative process can be made quite efficient; see Sec. IV D for more details. An example application of the spectral localizer framework to a tight-binding topological heterostructure is shown in Fig. 5. Here, a finite region of a Haldane lattice139 in a topological phase is embedded in a trivial insulator and, ultimately, surrounded by open boundaries. The density of states and local density of states (LDOS) confirm that the heterostructure exhibits bulk spectral gap that is populated by chiral edge states localized to the interface between the two constituent materials, as shown in Figs. 5(b) and 5(c). Furthermore, at the material interface, the spectral localizer framework shows that the local gap so that the local Chern marker can change its value and identify the inner material as topological; see Figs. 5(d) and 5(e). The spectrum of the spectral localizer spec[L(x,E)] shows a single eigenvalue crossing 0 that captures the change in the system’s local topology as (x, E) is varied across the heterostructure with E in the bulk spectral gap; see Fig. 5(f). This flow of the eigenvalue is called the spectral flow of L(x,E). Finally, this example also demonstrates that (x, E) can be chosen to be any real numbers, with the edges of the plots showing values for x outside of the heterostructure.
C. General definition of the spectral localizer in arbitrary dimensions
Intuitively, the use of a Clifford representation in the spectral localizer framework can be understood from the need to preserve the “orthogonality” of the information present in the system’s Hamiltonian and position operators while still combining these operators into a single L(x,E). This is akin to how the Pauli matrices form a complete basis for all 2-by-2 Hermitian matrices with spectra centered at 0. Rigorously, underlying the need for a Clifford representation in the spectral localizer framework lurks Clifford algebras; see Sec. VI.
Moreover, regardless of the specific form of any of the local markers defined using the spectral localizer, all of these markers are fundamentally connected to the spectrum of the spectral localizer and cannot change their value without first closing the local gap, . Thus, Eq. (14) always provides the measure of topological protection for any class of topology in the spectral localizer framework. In addition, it has been proven that changing the irreducible Clifford representation used in Eq. (19) does not alter the local gap; see Ref. 140, Lemma 1.2, e.g., and both yield the same .
Finally, we would like to be able to provide more guidance on how to pick the irreducible Clifford representation for Eq. (19), but this is currently an open topic. To illustrate the present difficulty, consider the 2D spectral localizer defined using −σx, −σy, and −σz instead of their positive counterparts. The negative Pauli matrices still form an irreducible Clifford representation, but the spectrum of the resulting spectral localizer will be multiplied by −1 relative to the standard choice of 2D spectral localizer given by Eq. (12), yielding a sign flip of the local Chern marker Eq. (13). Thus, the spectral localizer framework always predicts the correct number of chiral edge modes and is internally consistent for a given choice of Clifford representation. However, the spectral localizer framework can have a sign ambiguity when compared against other topological frameworks. Open question: further study is necessary to understand if there is an argument for constructively fixing the sign ambiguity, i.e., without an appeal to other topological invariants.
D. Identifying topology in 1D chiral symmetric systems using the spectral localizer framework
Altogether, through a judicious choice of elements of the Clifford representation used for an odd-dimensional system, all the necessary spectral information for defining local topological markers is contained in a single off-diagonal block of the spectral localizer. As always, the topological protection associated with these local markers is still given by Eq. (14), which is also equivalent to finding the smallest singular value of the reduced spectral localizer. An example of the spectral localizer framework applied to an SSH model with chiral-preserving long-range couplings is shown in Fig. 6. Here, the long-range couplings yield a local winding number of 2 in the lattice’s interior bulk, and the spectral flow of responsible for this change in the local index is shown in Fig. 6(c).
E. Classifying crystalline topology
A key benefit of the spectral localizer framework is that it is agnostic to the physical meaning of the matrices being used to construct it and its associated local markers. In particular, this means that if another system’s symmetry can be found that is outside the standard Altland-Zirnbauer classes and yet exhibits equivalent relations on the system’s Hamiltonian and position operators, then the local marker for this symmetry-protected topology can be immediately constructed.68
There are two subtleties associated with the crystalline winding number. First, just as any topologically protected states in the SSH model must exist at E = 0, the crystalline winding number identifies states at x = 0, which is the center of the crystalline symmetry. Thus, this class of markers fixes x and sweeps E to identify material topology. Second, as the spectral localizer works with systems that possess OBC, the relevant crystalline symmetry in Eq. (29) is a global symmetry, not a single unit cell operator. Open question: it remains an active area of research to explore whether the spectral localizer can be used to define topology with respect to single-unit-cell versions of a system’s critical symmetries, rather than their global counterparts.
An example of using the spectral localizer framework to classify crystalline topology is shown in Fig. 7, where we consider an inversion-symmetric SSH model with a defect site at the center. Thus, for this system, and , where is the inversion operator. The observation that is taking half-integer values is not a mistake, but is instead a direct consequence of the fact that this system has an odd number of sites in conjunction with the 1/2 in Eq. (29). These half-integers are not a problem; as differences in the local marker at different E are still integer valued, the spectral flow of across this range of E is the same integer, and this integer corresponds to the number of inversion-center localized topological states exist. (See Sec. IV C for a discussion of bulk–boundary correspondence.)
F. Dimensional reduction and higher-order topology
The spectral localizer framework can also dimensionally reduce a system and consider its topology in a lower dimension. As such, the framework can be used to identify higher-order topology,79,88,145–148 protected either by crystalline symmetry68 or by chiral symmetry.137 Mathematically, this dimensional reduction is achieved by simply omitting one or more position operators from the spectral localizer and choosing a Clifford representation suitable for the effective lower-dimensional system. In doing so, the full system is still used, i.e., the Hamiltonian remains unchanged. Instead, this process is projecting the system into the lower-dimensional space and simply forgetting about their position in the omitted dimension(s). Physically, changes in a system’s local topology can only occur at locations in position-energy space, where , which also guarantees that the system exhibits a nearby states; see Sec. IV C. Thus, by projecting the system into a suitable lower-dimensional space, one can ensure that a path in (x, E) taken by the spectral localizer always crosses a location where the local gap closes so the topology can change, precluding the possibility of choosing a path that goes around the local gap closure by omitting that dimension.
An example of using the spectral localizer to identify higher-order topology in a 2D breathing honeycomb lattice25 is shown in Fig. 8. The 2D breathing honeycomb lattice is chiral symmetric (class AIII), but in 2D, this local symmetry class is always trivial. Instead, this lattice’s higher-order topology and associated corner-localized states can be identified by projecting the lattice into a lower dimensional space and using the local winding number defined in Eq. (27). In Figs. 8(d)–8(g), the lattice is projected onto the x axis. As x is varied from negative to positive for E = 0, the local winding number changes four times at positions that correspond to the projected locations of the system’s six corners onto the x axis. The first and last changes in are by ±1, as these x locations intersect a single lattice corner. The middle two changes in the local marker are by ±2, as the path is crossing two projected corners simultaneously. Crucially, because the corresponding topological corner-localized zero-energy modes have the same chiral charge, i.e., they predominantly have support on the same sublattice, their combined contribution to the local winding number is ±2. In contrast, projecting the lattice instead onto the y axis yields a trivial local marker for every choice of y; see Figs. 8(h)–8(k). Here, although the local gap closes three times where the six corners are projected onto the y axis, each closing corresponds to the path crossing two corners with opposite chiral charge, yielding no change in .
There is no requirement that the axis that the system is projected onto for dimensional reduction is associated with one of the original position operators. For example, the 2D breathing honeycomb lattice shown in Fig. 8(a) could be projected onto an arbitrary choice of axis w = ax + by with associated position operator W = aX + bY. For most such choices, the associated local marker would change six times, each by ±1.
Dimensional reduction plays an important role in the application of the spectral localizer framework to realistic photonic systems, as many technologically relevant photonic systems are 2.5D systems, such as photonic crystal slabs and metasurfaces. Rigorously, such systems are 3D, but the desired topological boundary-localized states in these systems reside in the same planar slab as the structure. Thus, dimensional reduction enables the spectral localizer to rigorously define local markers associated with 2D material topology for such 3D planar systems. See Ref. 70 for an example where the spectral localizer framework is used to identify a Chern photonic crystal slab.
G. Behavior in the thermodynamic limit
The structure of the spectral localizer’s local markers presented in Sec. III C seems to suggest that these markers can only be applied to systems described by finite, but arbitrarily large, systems with bounded spectra where the signature, Pfaffian, and determinant operations are well-defined because, in principle, the entire spectrum of the spectral localizer can be found. (In practice, it is not recommended to ever compute the local markers this way; instead see Sec. IV D.) Nevertheless, the definitions of the spectral localizer framework are still well-defined for some infinite-dimensional operators149 such as those describing physical systems in the thermodynamic limit. The key is that formulas such as Eq. (13) can be generalized to infinite-dimensional L(x,E) by considering its spectral flow as (x, E) moves along a path in position-energy space.150
Consider a semi-infinite 2D system with a bounded spectrum. For x chosen far outside of the system, one can prove that the spectrum of its spectral localizer is balanced. Intuitively, this is straightforward to see, as if x moves away from the edge of the half lattice, κ(X − x1) ⊗ σx + κ(Y − y1) ⊗ σy has a growing spectral gap that (H − E1) ⊗ σz is of too-small a norm to close. Heuristically, we want this to mean as would be the case if the system were finite, but the signature is not defined for a semi-infinite system. To be rigorous, we instead need to use spectral flow and the so-called η-invariant.66,150 Then, as x is varied so that it crosses into the bulk of the system, the spectral localizer’s spectrum near 0 can be monitored for any crossings, i.e., one can track its spectral flow in an analogous manner to the examples shown in Figs. 5–8. Crucially, despite the fact that the spectral localizer is infinite-dimensional (referring to the Hilbert space’s dimension, not the number of physical dimensions), in many cases, its spectral flow is well-defined and yields a local topological marker. In other words, in these cases spectral localizer has a “signature” that is twice the index of a related Fredholm operator.65 The indices relate to the secondary indices that can occur when the index of a Fredholm operator is zero.143 Altogether, this means that the spectral localizer framework can be rigorously generalized to the thermodynamic limit without issue.
(At present, there is a technical restriction in defining the spectral flow of an infinite-dimensional operator, as we need to know that the spectral localizer has discrete spectrum, so now we cannot work directly with a differential operator such as those needed for photonic systems; see Sec. V. Open question: we conjecture that the methods of spectral truncation151,152 in noncommutative geometry are likely to provide a way around this limitation.)
IV. UNDERPINNINGS OF THE SPECTRAL LOCALIZER FRAMEWORK
Having introduced the spectral localizer framework and provided a number of examples of its application to identifying material topology across a variety of systems, this section discusses three topics that underpin how the spectral localizer framework functions. First, we provide a detailed discussion of the hyper-parameter κ and provide some guidance for choosing κ. Then, we turn to introducing multi-operator pseudospectral methods, which form the foundation for proving bulk-boundary correspondence in the spectral localizer framework. Finally, we discuss how to efficiently implement the spectral localizer’s local markers.
A. The role of the hyper-parameter κ
The hyper-parameter κ serves two critical roles in the spectral localizer framework. First, κ adjusts the units of the position operators in L(x,E) to have dimensions of energy so that they can be combined with the Hamiltonian. Mathematically, this choice is arbitrary, the spectral localizer could be defined in units of length or made dimensionless through a second hyper-parameter and the structure of the local markers would not change. Physically, this choice is useful, as it enables direct comparison between the local gap and a system’s bulk bandgap or other spectral gap. Second, κ balances the spectral weight of the Hamiltonian relative to the position operators. In other words, κ is chosen so that the eigenvalues of L(x,E) are similarly sensitive to changes in H − E1 and changes in X − x1, either because the choice of (x, E) is shifted or because the system is perturbed H → H + δH. Intuitively, this second function of κ is playing a similar role as the choice of region areas in the Kitaev marker, Eq. (4), or the choice of integration disk radius in the Bianco-Resta marker, Eq. (8). For example, larger values of κ are comparable with smaller integration disk radii in the Bianco-Resta marker and generally enable greater specificity in changes of x where the spectral localizer is evaluated, but potentially at the cost of being too insensitive to spectral information and mis-classifying the system’s topology. Similarly, smaller values of κ are comparable with larger integration disk radii and generally yield correct material classification in the bulk until the corresponding length scale ∝κ−1 is too large, similar to how the Bianco-Resta Chern marker is always trivial if the integration disk contains the entire lattice, as shown in Fig. 9.
In between these two limits, it is reasonable to expect that structures for which the local gap closes in position-energy space interpolate between these two distributions. However, for a system with non-trivial topology, a closed surface in (x, E)-space always appears, for which as part of the interpolative process, as shown in Figs. 9(b)–9(d). For locations outside of the system, or choices of energy outside of the bounded spectrum, the local markers are provably trivial (see Ref. 64, Lemma 7.4). Therefore, if there is a topologically non-trivial region in position and energy inside the system, any path in position-energy space starting in the non-trivial region and ending outside the system must possess at least one location, where so the local markers can change their values, yielding a closed surface.
Finally, as can change for different choices of κ, the spectral localizer framework’s best estimate for the topological protection at (x, E) is the maximum local gap over all κ that yield the same local markers. It should be noted that different choices of (x, E) might achieve their largest local gap for different values of κ.
B. Overview of multi-operator pseudospectral methods
In Sec. III B, we introduced the local gap and discussed how this provided a measure of topological protection for a system. Moreover, in Secs. III C and IV A, we argued that in systems with non-trivial topology, the local gap must close everywhere on a closed surface in position-energy space surrounding topological regions in a system. Thus, given that the local gap in a bulk region is quantitatively connected to the bulk spectral gap in that region, it is reasonable to argue that the (x, E) surface over which must be near the interface between the topological and trivial regions in a system. Yet, looking at some of the examples considered in this tutorial, such as Figs. 5(c) and 5(d), closings of the local gap appear to be co-located with positions of the associated topological edge-localized states in a spectral gap. In this section and the next, we make this connection between a system’s LDOS and local gap rigorous and ultimately show how bulk-boundary correspondence manifests in the spectral localizer framework.
More broadly, multi-operator pseudospectral methods are an approach to understanding whether an arbitrary number of non-commuting matrices nevertheless exhibit approximate joint eigenvectors, and a variety of other multi-operator pseudospectra have been proposed.154,155 Moreover, the multi-operator pseudospectral methods that we discuss here can be considered as a generalization of “two-operator” pseudospectra156,157 that have previously been used to study a variety of physical systems.158–163 It should be noted although that both multi-operator pseudospectral methods and two-operator pseudospectral methods are unrelated to “pseudospectral methods” as an alternate name for discrete variable representation methods used in the solution of partial differential equations.164
C. Bulk–boundary correspondence in position space
Thus, bulk-boundary correspondence naturally appears as a consequence of Eq. (39). As discussed in Sec. III C, changes in any of the spectral localizer’s topological markers can only occur at locations in the (x, E)-space where . Moreover, material systems are generally local, such that κ‖[Xj, H]‖ ∼ κa is small relative to the system’s energy scale, where a is the lattice constant. Thus, at locations where a system’s local topology changes, Eq. (39) guarantees that is small, such that there must be a nearby state of the system due to Eq. (37). In addition, it should be noted that the need for the spectral localizer framework to be applied to finite systems with OBC can be viewed as a consequence of Eq. (39); if PBC were allowed, then κ‖[Xj, H]‖ ∼ κlj with lj being the length of the system in the jth dimension, which is generally large, and thus a bulk–boundary correspondence would not generally exist.
However, Eqs. (37) and (39) have broader utility beyond guaranteeing boundary-localized topological states; they can be used to predict the properties of all of a system’s states. For example, the LDOS, local gap, and quadratic local gap are compared for both trivial and topological 1D SSH chains shown in Fig. 11. As can be seen, both and quantitatively resemble the system’s LDOS at all positions and energies. Open question: however, it should be noted that the (Clifford) local gap becomes smaller than the quadratic local gap where the system exhibits states. Thus, it appears that should provide a sharper estimate for position-energy locations where a system exhibits an approximately localized state, regardless of whether the state is of topological or trivial origin. Yet, at present, the best known bounds on a state’s location and localization are given in terms of via Eq. (37). Further study is required to understand whether better bounds can be derived in terms of the Clifford local gap.
D. Efficient algorithms and numerical K-theory
It is tempting, but inadvisable, to calculate the spectral localizer’s local markers with the forms given in Eq. (25) using a naïve approach, e.g., first calculating the determinant or Pfaffian of L(x,E) and then taking its sign, or first calculating the spectral localizer’s full spectrum and then calculating its signature. Unfortunately, even for modestly sized systems, such a naïve approach will yield, at best, a slow numerical implementation, and might, at worst, mis-classify the system. Fundamentally, these kinds of approaches fail to take advantage of two separate properties of the spectral localizer framework. First, such naïve approaches do not leverage the fact that L(x,E) is guaranteed to be sparse, which is due to Xj being diagonal and H being reasonably local.73 Second, in viewing the local marker formulas as specifying a sequential algorithm, e.g., find the determinant then take its sign, they miss significant speed-ups available from using matrix factorizations that may not preserve L(x,E)’s spectrum, but do preserve the local marker.
Finally, we note that the local gap can be efficiently computed using sparse matrix methods, which can find the single eigenvalue with the smallest magnitude without determining the full spectrum of the spectral localizer.
Altogether, the local topological markers provided by the spectral localizer framework are examples of numerical K-theory, which can broadly be defined as the study and development of the formula for K-theoretic invariants, such as invariants that classify material topology, which are amenable to efficient numerical calculation. Other examples of numerical K-theory outside of the spectral localizer framework have been developed by Prodan,169 Fulga,170 and Quinn and Bal.171
V. APPLYING THE SPECTRAL LOCALIZER TO PHOTONIC SYSTEMS
Having introduced the spectral localizer framework in general, in this section, we now turn to its application to realistic photonic systems. Overall, the process of adapting Maxwell’s equations for use with the spectral localizer is relatively straightforward and can be accomplished with standard discretization techniques, including both finite-difference67,69 and finite-element70 methods. However, in doing so, two issues arise that are not found in the kinds of simple tight-binding models considered in Secs. III and IV. First, Maxwell’s equations are a set of differential equations with an unbounded spectrum, i.e., these equations exhibit an infinite number of eigenvalues that become infinitely large. Thus, prior to discretization, quantities such as ‖H‖ are undefined. While discretizing the system yields a finite Hamiltonian matrix with a bounded spectrum, ‖H‖ is then determined by the modes at the extremes of H’s spectrum that are never in the frequency range of interest because they are not well-described by the discretization (if these frequencies are of interest, a finer discretization is needed). As such, the spectral localizer framework’s predictions for topological protection (Sec. III B) and bounds on κ (Sec. IV A) need to be adapted. Second, photonic systems commonly feature radiative boundary conditions,172 which are non-Hermitian. Thus, the spectral localizer framework’s local markers need to be adapted for non-Hermitian systems featuring line gaps.
In this section, we first remind a reader how Maxwell’s equations can be reformulated into an ordinary eigenvalue equation and subsequently discretized to be used in the spectral localizer framework. Then, in Sec. V C, we discuss how to use a generalized eigenvalue problem in the spectral localizer, which is an approach suited for use with finite-element methods. In Sec. V D, we show how to generalize the spectral localizer and local Chern marker to line-gapped non-Hermitian systems to allow for the inclusion of radiative boundary conditions. In Sec. V E, we consider how the spectral localizer framework’s local measure of protection needs to be altered for systems with unbounded, or effectively unbounded, spectra. Finally, in Sec. V F, we comment on using the spectral localizer to address nonlinear systems.
A. Review of Maxwell’s equations as an eigenvalue equation
Here, it should be noted that we have reviewed Maxwell’s equations from the perspective of coupled first-order differential equations. However, it is also possible to use the second-order differential equation form in the spectral localizer framework.68 Indeed, the second-order form has some advantages for some forms of topology as it does not require handling both ordinary vectors such as the electric field E(x) and co-vectors (also called pseudovectors) such as the magnetic field H(x) in the same eigenvalue equation as they transform differently under some symmetries.
At this point, there are two different ways to proceed. One can reformulate Eq. (51) as an ordinary eigenvalue equation, which preserves ω as the eigenvalue but places requirements on the constituent materials. Alternatively, one can work with Eq. (51) directly, which does not have any material constraints, but slightly alters how the frequency eigenvalue is handled.
B. Ordinary eigenvalue problem approach
Mathematically, there are other possibilities for the properties of Mdisc(ω) that would allow for a unique square root to be defined, such as if it is negative semi-definite. Physically, these cases are not especially relevant, as they generally correspond to systems that are completely formed of metals that are likely to be highly absorbing at technologically relevant frequencies.
C. Generalized eigenvalue problem approach
It should be noted that this approach has some subtleties in directly comparing perturbations in the effective Hamiltonian against the local gap that requires considering . Open question: it may be possible to instead define an alternative local gap based on ∑j(Xj − xj1)M ⊗ Γj + (Wdisc − ωMdisc(ω)) ⊗ Γd+1, which is guaranteed to have the same signature as a spectral localizer based on Eq. (52) and would not involve calculating a matrix inverse.
D. Accounting for non-Hermitian phenomena
Even in the absence of material absorption, many photonic systems exhibit radiative losses resulting in a Hamiltonian that is non-Hermitian. For example, a topological heterostructure constructed in a photonic crystal slab will exhibit out-of-plane radiation for states whose in-plane momenta k‖ are above the light line ω = c|k‖|.175 Mathematically, this creates some challenges for the spectral localizer framework because the kinds of homotopy arguments that the framework relies on typically demand Hermiticity. To illustrate this point, it should be recalled that Sec. III A and Fig. 4 discussed how two invertible Hermitian matrices can only be connected by a path of invertible Hermitian matrices if they have the same signature. However, just as Fig. 4(b) showed how invertible Hermitian matrices with different signatures can be connected by a path in which some of the Hermitian matrices were non-invertible, a similar connection can be made through a path that contains invertible non-Hermitian matrices. Thus, to handle non-Hermitian physical systems, the underlying homotopy arguments must either be replaced or expanded for the spectra localizer framework to remain applicable.
Numerically, the switch to non-Hermitian systems also yields problems, as for non-Hermitian M can no longer be calculated using the LDLT decomposition and Sylvester’s law of inertia as detailed in Sec. IV D and Eq. (42). Instead, progress has been made by starting with a Hermitian variant of the system and gradually turning on the absorbing boundary condition while monitoring , which is still efficient to calculate using sparse methods. In the Hermitian variant, the topology can be quickly calculated using techniques from Sec. IV D, and then, if the local gap remains open as the absorbing boundary condition is turned on, the topology cannot change. Indeed, this approach has proven sufficient in realistic photonic systems using both finite-difference69 and finite-element70 discretizations.
Open questions: there remains a great deal of work to do in this area. First, it is not known whether the local markers for any other class of topology in the Altland-Zirnbauer symmetry classification can be similarly generalized to non-Hermitian systems. Similarly, the non-Hermitian generalization of crystalline topology is not known. This problem is particularly acute for the odd-dimensional classes of topology, where the local markers for Hermitian systems are dependent on a symmetry reduced spectral localizer such as Eq. (28) that only includes a single copy of the system’s Hamiltonian. In addition, it is unknown whether there are efficient numerical approaches for directly calculating Eq. (55).
The spectral localizer framework has also been extended to consider other forms of non-Hermitian topology by Fulga and colleagues.120–122
E. Local protection for unbounded operators
Using the spectral localizer framework on systems described by differential equations with unbounded spectra, i.e., systems that exhibit an infinite number of eigenenergies that become infinitely large, presents a problem for the framework’s measure of topological protection. Regardless of which specific approach is taken to insert the system’s Hamiltonian into the spectral localizer, the norm of any perturbation diverges. Heuristically, this problem stems from the fact that the differential operator W is unbounded, and any perturbation to the material response matrix M(ω) → M(ω) + δM(ω) still ends up multiplying W for comparison with the local gap, so the full Hamiltonian perturbation remains unbounded. Thus, the criteria for the system to change its local topology is always trivially satisfied. Intuitively, the problem is that is a provably local quantity in both x and E (see Ref. 180, Sec. 7), reflecting whether the system exhibits a nearby state approximately localized at those position-energy coordinates (as discussed in Sec. IV C). In contrast, ‖δH‖ is a global quantity, and is controlled by the system’s response at high energies.
Open questions: however, there is a pressing need to develop an exact bound for the topological protection of realistic systems, possibly in terms of a resolvent, rather than such an approximate bound. Indeed, one of the strengths of the spectral localizer framework is its ability to be applied directly to experimentally realizable systems, but those systems are described by wave equations where the difficulty discussed in this section will arise. Ideally, this bound would be formulated in terms of quantities that can be efficiently computed for sparse matrices. Alternatively, the introduction of an alternative local gap, as discussed in Sec. V C, may also solve the difficulty discussed in this section by separating out the perturbation to the material response matrix yielding a bound on ‖δM‖, which is generally finite.
F. Classifying local nonlinear topology
Nonlinear topological systems represent an exciting frontier in photonics, providing a path to exploring phenomena beyond what can be found in electronic topological materials,57 such as bulk182–189 and edge190–192 solitons with topological properties, nonlinearly induced topological phase transitions,193–195 and topological multi-wave mixing.196–200 Moreover, the spectral localizer framework and its local picture of material topology appears to be extremely well positioned to study these systems,71,201 as its local markers are able to resolve a system’s local change in topology due to its occupation.
While the definition of local topological protection provided by remains unchanged for nonlinear systems, it acquires new physical meaning for such systems.71 As discussed in Sec. III B, for any perturbation to a linear system described by a bounded Hamiltonian H → H + δH to change the local topology at (x, E), the perturbation must be at least strong enough to close the local gap . However, in nonlinear systems, a small perturbation to the Hamiltonian can change whether a given solution to the nonlinear eigenvalue equation exists. In other words, given a self-consistent solution to Eq. (58), one can try to follow this solution as the strength of a perturbation is increased, but it is possible that for a sufficiently strong perturbation that the solution curve may simply terminate. Nevertheless, if this self-consistent solution is topological, i.e., it induces a local change in the system’s topology at some (x, E), then guarantees that the solution curve cannot disappear until the perturbation is strong enough to close the local gap, as the self-consistent solution disappearing causes a change in the system’s local topology.
VI. C*-ALGEBRAIC BACKGROUND TO THE SPECTRAL LOCALIZER
Overall, this tutorial has focused on providing an understanding of how the spectral localizer framework is used to classify material topology and its associated robustness in physical systems, with the goal of providing a reader with the necessary equations and sufficient intuition to analyze their system of interest. Nevertheless, in this penultimate section of the tutorial, we now turn to providing some explanation of the mathematical underpinnings of the spectral localizer framework. This section is primarily intended for a reader with a mathematics background, with the goal of providing some guidance on the relevant concepts needed to advance the mathematics of the spectral localizer. A reader uninterested in this topic may safely skip this section.
A. C*-algebras
Those familiar with band theory might not realize that the momentum-space picture , of a tight-binding Hamiltonian H, is an element of a C*-algebra. In particular, is an element of , which is the set of continuous functions from the d-torus to the N-by-N complex matrices. The elements of naturally act on the momentum Hilbert space , so we could consider as a subalgebra of . In general, denotes the algebra of all bounded linear operators on . However, if one wants to know the spectrum of the original Hamiltonian H, acting on position space , for example, one can instead compute the spectrum of , not as an operator on , but as an element of the C*-algebra . The spectrum is then simply the union of the spectra of all the matrices H(k). In band theory, this relation of band structure to the spectrum is explained by explicit calculations, essentially using the Plancherel (Fourier) transformation. However, this can also be explained in C*-algebra language, as an application of spectral permanence; see, for example, Murphy,204 Theorem 2.1.11.
More broadly, the most basic C*-algebra encountered in physics, and the one most central to numerical studies, is simply the n-by-n matrices . This is essentially the same as (isomorphic to) with . In finite dimensions (as in, n is finite), boundedness is automatic. Thus, the model used to arrive at a definition of a C*-algebra is just equipped with the linear structure, operator composition (essentially matrix-matrix multiplication), the spectral norm, and the adjoint. The spectral norm ‖M‖ of a finite matrix can be described as either the largest singular value of M, or the maximum of ‖Mv‖ subject to the constraint ‖v‖ = 1.
Another instance where a physicist will have implicitly performed calculations within a C*-algebra is when functions are applied to an operator. If H is Hermitian with spectrum Ω, then the C*-algebra C*(H, 1), formed by taking limits of expressions such as 2H + iH3, will be isomorphic to C(Ω), the algebra of all complex-valued continuous functions on Ω. The isomorphism sets up an intuitive correspondence, sending H to the function h(λ) = λ, and the unitary propagator eitH to the function u(λ) = eitλ.
B. C*-algebras with extra symmetries
Theoretical work in the K-theory of C*-algebras almost always works with homotopy classes of elements that have been spectrally flattened. That is, one looks at homotopy properties of Hermitian elements with spectrum in {−1, 1} or unitary matrices, whose spectrum must lie in the unit circle. However, numerically, spectral flattening is slow and typically results in a dense matrix. One generally cannot apply formulas from pure math papers on K-theory, unmodified, and expect fast algorithms.
C. Clifford algebras
Clifford algebras operate behind these scenes to help categorize all the possible irreducible Clifford representations.140 Moreover, it can be useful to consider multiple Clifford representations in the same calculation. For example, two very different Clifford representations can be used to prove that various symmetries in (x, E) lead to a corresponding symmetry in the Clifford pseudospectrum of (x, E).140
Let d denote one less than the number of Clifford matrices needed, since we usually use d + 1 Clifford matrices when there are d physical dimensions. The key thing to know when d is odd is that any two Clifford representations of minimal size will be related via conjugation by a single unitary. When d is even, this is false, as the Clifford representations come in two flavors. For d = 2, the two flavors can be distinguished by checking which of Γ1Γ2 = ±iΓ3 holds. Within one of the flavors, all irreducible representations are related via conjugation by a single unitary.
VII. SUMMARY AND OUTLOOK
In this tutorial, we have endeavored to provide a physically motivated introduction to the spectral localizer framework to facilitate its use across the community to address challenges at the frontiers of topological photonics. As this framework provides local markers of material topology and comes equipped with a local measure of protection, it is able to analyze systems that are either difficult or impossible to consider using traditional approaches, such as topological phase transitions induced via local nonlinearities, effects dependent on finite system sizes, and the appearance of topological phenomena despite the absence of a spectral gap. Moreover, due to the mathematical formulation of the framework’s invariants and local gap, these quantities can be computed efficiently even for realistic systems governed by differential operators and numerically described using finite-difference or finite-element methods. In addition, the framework’s generality has also enabled its application in plenty of condensed matter settings.181,210,211 As part of this introduction, we have outlined the mathematical concepts of multi-operator pseudospectral methods, which allow for the prediction of approximate joint eigenvectors of non-commuting matrices and form the basis of bulk-boundary correspondence in the spectral localizer framework, as well as numerical K-theory, the concept that underpins the framework’s numerical efficiency. Finally, we have provided some guidance to any interested mathematically oriented reader for how to continue to develop the associated possibly real, possibly graded C*-algebras.
Looking forward, substantial opportunities remain in both the development of the spectral localizer framework and its application to novel physical systems to predict new phenomena. Throughout this tutorial, we have marked open questions where the framework would benefit from additional results proving a generalization to better address significant classes of physical systems. Moreover, the ability to use the spectral localizer framework in conjunction with models of realistic systems beyond photonics, while including the possibility of aperiodicity or disorder, may yield fruitful results across a range of fields of study. In addition, as the field of topological photonics turns to designing novel device architectures, where miniaturization is at a premium, the framework may also find utility in providing a better understanding of topological protection in the presence of finite system size effects. However, more broadly, we are hopeful that given a physically motivated introduction to the subject, the community will find even more applications of the spectral localizer framework.
ACKNOWLEDGMENTS
We thank Wladimir Benalcazar and Stephan Wong for providing feedback on this tutorial. We thank Florian Sterl for the development of Fig. 2. A.C. acknowledges support from the U.S. Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering. T.A.L. acknowledges support from the National Science Foundation under Grant No. DMS-2349959. This work was performed, in part, at the Center for Integrated Nanotechnologies, an Office of Science User Facility operated for the U.S. Department of Energy (DOE) Office of Science. Sandia National Laboratories is a multimission laboratory managed and operated by National Technology and Engineering Solutions of Sandia, LLC, a wholly owned subsidiary of Honeywell International, Inc., for the U.S. DOE’s National Nuclear Security Administration under Contract No. DE-NA-0003525. The views expressed in the article do not necessarily represent the views of the U.S. DOE or the United States Government.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Alexander Cerjan: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Software (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Terry A. Loring: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Software (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).
DATA AVAILABILITY
Data sharing is not applicable to this article as no new data were created or analyzed in this study.
APPENDIX: HOMOTOPY RESULTS
Here, we discuss the essential results on the homotopy classification of specific classes of matrices of a fixed size. We start with proving the theorem in Sec. III A.
The argument to show that invertible Hermitian, n-by-n matrices with different signatures cannot be connected by a Hermitian path that remains invertible is essentially the argument in the caption to Fig. 4. It is possible to find a continuous path, perhaps by interpolation, of Hermitian matrices between the two, but at some point along the path, an eigenvalue must cross zero and that leads to non-invertibility.
Now, we get a subtle situation, that of fermionic parity. Kitaev73 discusses how this applies to 0D systems in Class D. In addition, see Ref. 212 for a more detailed discussion of fermionic parity for a coupled pair of quantum dots in a superconducting setting.
Suppose n is even. Two n-by-n invertible Hermitian skew-symmetric matrices H and H′ can be connected by a path of invertible Hermitian skew-symmetric matrices if and only if .
Two n-by-n invertible real matrices A and A′ can be connected by a path of invertible real matrices if and only if .
Suppose H is real and invertible. The determinant of a real matrix is always real and cannot be zero when H is invertible. Since the determinant is continuous as a function of H, we see that two such matrices with determinants of opposite sign cannot be connected in this space of matrices.
To prove that all such matrices of a given sign of determinant are homotopic, we first need to reduce to the case where A is real orthogonal. We do this by utilizing the path . We see that A0 = A and that A1 is real orthogonal. This construction is continuous in A, a fact that can be used to show that two real orthogonal matrices that are homotopic in the larger space of invertible real matrices must be homotopic in the smaller space. Real orthogonal matrices form a group O(n) that has two connected components, SU(n) and the space of real orthogonal matrices of determinant minus one. One can understand this intuitively due to the nature of the eigenvalues of A when A is real orthogonal. The spectrum of A has three parts. There are eigenvalues at 1 that do not matter in the sign of the determinant. There are conjugate pairs on the unit circle that also do not matter since their product is positive. It is the eigenvalues at −1 that can lead to det(A) = −1. Any two of these can be deformed as a conjugate pair that ends up with both as +1. It is the solo eigenvalue at −1 that cannot be moved.