Topological properties of band structures in magnetic materials are systematically studied using symmetry representations.
Abstract
The properties of electrons in magnetically ordered crystals are of interest both from the viewpoint of realizing novel topological phases, such as magnetic Weyl semimetals, and from the application perspective of creating energy-efficient memories. A systematic study of symmetry and topology in magnetic materials has been challenging given that there are 1651 magnetic space groups (MSGs). By using an efficient representation of allowed band structures, we obtain a systematic description of several basic properties of free electrons in all MSGs in three dimensions, as well as in the 528 magnetic layer groups relevant to two-dimensional magnetic materials. We compute constraints on electron fillings and band connectivity compatible with insulating behavior. In addition, by contrasting with atomic insulators, we identify band topology entailed by the symmetry transformation of bands, as determined by the MSG alone. We provide an application of our results to identifying topological semimetals arising in periodic arrangements of hedgehog-like magnetic textures.
INTRODUCTION
The recent discovery of topological insulators and other topological phases (1) has revitalized the venerable subject of band theory. In addition to the explosion of understanding of different forms of band topology and how symmetries protect or prevent them, researchers are making rapid progress on several other fronts. For example, fundamental questions, such as the connection between electron count and insulating behavior (2–10), as well as the constraints imposed by crystal symmetries on the connectivity of bands (3–5, 10–17), have been resolved. New information has also been gleaned by contrasting real-space and momentum-space descriptions in all 230 crystal space groups (SGs) of nonmagnetic materials (3–5, 8–10, 15, 16, 18).
Increasingly, attention is turning to electronic systems that combine magnetism with band topology. Here, a further panoply of novel topological phenomena is anticipated. Examples that scientists have already realized, to name just a few, include the quantized anomalous Hall effect in magnetic topological insulators (19, 20) and the topological Hall effect in skyrmion crystals (21–25). Researchers are also now studying magnetic Weyl semimetal candidates intensively (26–28). In addition to the fundamental physics interest of these novel phases, current-driven magnetic textures, such as domain walls or skyrmions, could be harnessed in technological applications, for example, in the development of energy-efficient memory devices (29).
Despite these strong motivations, the pace of discovery of magnetic topological materials has been relatively slow compared to their nonmagnetic counterparts. There are at least two reasons for this: First, magnetic materials are necessarily correlated, making the prediction of the magnetic structure and properties more challenging. Consequently, the set of well-characterized magnetic materials is relatively small. Second, the sheer complexity of combining magnetic structures with SGs makes an exhaustive study of the theoretical possibilities daunting. There are a total of 1651 different magnetic SGs (MSGs), which were only tabulated in the 1950s. In addition, unlike for the 230 nonmagnetic SGs (2, 30), relevant group-representation information is not always readily available.
Here, we will tackle the second problem by providing a systematic understanding of electronic band structures (BSs) in the 1651 MSGs. Inspired by the recent synthesis of atomically thin magnets (31, 32), we also discuss the 528 magnetic layer groups (MLGs), relevant to two-dimensional (2D) magnetic materials. We report the results for three key properties, which are tabulated in section S1 and described below.
First, we determine electron fillings that can be compatible with insulating behavior in all MSGs. The electron count is a fundamental characteristic of an electronic crystal. The presence of nonsymmorphic symmetries, such as glides and screws, enforces connectivity of bands that raises the required fillings for realizing band insulators (3–5, 12). These conditions may be useful in the search for magnetic Weyl semimetals, as one can target fillings that, while forbidding band insulators, are nonetheless consistent with nodal-point Fermi surfaces (13, 33, 34).
Next, it has long been known that representations of energy levels at high-symmetry points must connect in specific ways in obtaining a set of isolated bands (11). We can view this as a refinement of the filling condition, which imposes additional constraints on BSs (14–17). The solutions to these constraints are most conveniently represented as a vector space (but with integer coefficients, so more accurately a “lattice” in mathematical terminology) (14, 15, 35) and are described using only a handful of basis vectors whose precise number dBS depends on the MSG.
Last, we contrast the general BSs defined in the momentum space with the subset of those obtained from “atomic insulators” (AIs), in which electrons are tightly bound to sites in the lattice furnishing orbitals with different symmetries. Significantly, we find that this produces a vector space of the same dimension, which in mathematical terms, is summarized by the equality dAI = dBS. However, the bases for BSs and AIs do not generally coincide, and we determine the classes of BSs that cannot be reduced to any atomic limit. This leads to an obstruction to finding symmetric localized Wannier states (36, 37) and corresponds to a band topology that one can diagnose without detailed information about the electronic wave functions. We give an example of how this can be used to diagnose topological semimetals and also an example of a nodal-line semimetal diagnosed through its filling. The latter consists of hedgehog-antihedgehog magnetic order with a lone electron at the core of these defects, leading to a gapless behavior.
We note that scientists have recently made a great deal of progress on related problems for nonmagnetic SGs (9, 10, 15, 16, 18). However, in the case of MSGs, these problems have only been attacked in certain restricted settings, for example, in generalizing the parity criterion (38) in order to identify Weyl semimetals, Chern insulators, and axion insulators (35, 39–42). In this study, we carry out a systematic study of all MSGs. In particular, it is important to emphasize that, although our approach has several features in common with K-theory–based classifications (14, 43–45), we do not seek to fully classify crystalline electronic phases here. As discussed in a recent insightful work (14) for 2D wallpaper groups, part of such a general classification includes band structure connectivities, a problem that has been posed since the early days of band theory (11). Using a result we proved below, we construct a framework that greatly simplifies these computations, which enables an extension of previous results (9, 10, 15) to the physically important case of MSGs.
In addition, topological distinctions revealed using our symmetry-based indicators fit naturally into well-established frameworks (14, 15, 43, 45) and remain stable upon the addition of trivial degrees of freedom. Yet, our approach is inherently symmetry-based, and so for crystals with low symmetry, that is, with only lattice translations, one cannot diagnose topological insulators without further knowledge of the wave functions. A similar caveat pertains to the complementary quantum chemistry approach of (16), which elaborates on the theory developed by (3), and takes a specific set of orbitals at fixed locations in the crystal as an input. While convenient for representing quantum chemistry information, it is less suited to capture stable topological distinctions that survive the inclusion of additional trivial bands. In addition, one forgoes the simplification arising from the vector space such as representation of energy bands that allowed us to generate results for all MSGs, while the work of Bradlyn et al. (16) is currently confined to just the 230 SGs.
MATERIALS AND METHODS
Magnetic BSs
Let us begin by reviewing some background materials concerning electronic BSs arising from a magnetic material. There are a total of 1651 MSGs and 528 MLGs (2, 30, 46–48). Among the 1651 MSGs, 230 of them were identical to SGs, in which only unitary spatial symmetries were considered (type I MSGs). All other MSGs had an equal number of unitary and antiunitary elements: . The unitary part was identical to 1 of the 230 SGs, and the antiunitary part could be generally written in the form , where is the product of a spatial operation g0 and time reversal (TR) . When g0 belongs to , the MSG is simply the direct product of an SG and . This led to another 230 MSGs, one for each SG, that are called type II. When g0 is not an element of , there are two types further differentiated by whether g0 is a pure translation (type IV) or not (type III). For type II to IV MSGs, the little group of k and the site-symmetry group of x may also have an antiunitary part, in addition to the usual unitary part. Note that in the literature [for example, (49)], our definition of type III and IV MSGs are sometimes called type IIIa and IIIb, respectively. In addition, there are two common labeling schemes for MSGs: Opechowski-Guccione and Belov-Neronova-Smirnova (BNS). In this study, we followed the BNS notation, where an MSG was labeled by a pair of integers written as S. L, with S corresponding to 1 of the 230 SG and L corresponding to an extra label to differentiate between different magnetic descendants. [Refer to, for example, (50) for the precise meaning of these numbers.]
Next, we introduced a general formalism for the efficient analysis of BS properties based on representations of the little group (14, 15, 35). Our main focus here was to address how the formalism developed for SGs could be readily applied to MSGs. We defined a BS as a set of bands isolated from others by band gaps above and below at all high-symmetry momenta, where by a high-symmetry momentum, we referred to a k in the Brillouin zone at which the unitary part of the little group, , is necessarily larger than the translation subgroup T. In 3D, they can be high-symmetry points, lines, or planes. A BS can be characterized by the set of nonnegative integers that count the number of times an irreducible representation (irrep) of appears. As we are interested in systems of spinful electrons, the irreps here are generally projective because of the spin-1/2 nature of electrons (also called “double-valued”). Since cannot be changed smoothly without gap closing or symmetry breaking, serves as “topological invariants” defined for each BS.
The integers cannot be chosen freely, since symmetries demand that they satisfy a collection of compatibility relations (2, 11). Since is a subgroup of , the full list of compatibility relation imposed on an -symmetric BS can be split into two sets, arising from and from . Let us denote by and {BS}phys the set of all n’s satisfying and , respectively. Here, the subscript “phys” indicates that all ’s in are non-negative, which is required for interpreting them as the multiplicities of irreps in a physical BS. We will introduce another set that relaxes this nonnegative condition later. Note that and {BS}phys differed only in the imposition of . In general, the antiunitary part requires a pairing of with another , unless b itself is already symmetric under . The pairing type can be easily determined using the Herring rule (2). (In section S4, we provided a more elaborated review on the compatibility relations and the Herring rule.)
Band topology
Having described some generalities about BSs, we now review how knowledge about the real space can inform band topology (9, 15). We defined the trivial class of BSs by the AIs, which were band insulators that were smoothly connected to a limit of vanishing hopping and hence were deformable to product states in real space. Equivalently, an AI admits symmetric, exponentially localized Wannier functions.
To specify an AI, one should choose a position x in real space at which electrons were localized and the type of the orbital put on that site. All inequivalent choices of the position x were classified by Wyckoff positions (51). The orbital can be chosen from the (co-)irreps of the site-symmetry group of x (section S5). Given these choices, an -invariant AI can be constructed by placing a symmetry-related orbital on each site of the -symmetric lattice and filling them by electrons. The AI has a specific combination of irreps in the momentum space, which automatically satisfies . We listed up all distinct n’s corresponding to an AI by varying x and the orbital type. We listed up all distinct n’s corresponding to an AI by varying the position x and the orbital type, and we obtained {AI}phys, a subset of {BS}phys. If one replaces above with , one gets the set of -symmetric AIs, .
Now, we are ready to tell which elements of {BS}phys must be topologically nontrivial and which elements can be trivial. This can be judged by contrasting the elements of {BS}phys with those in {AI}phys. Namely, any b ∈ {BS}phys not belonging to {AI}phys necessarily features nontrivial band topology because, by definition, there does not exist any atomic limit of the BS with the same combinations of irreps. This is a sufficient (but not necessary) condition to be topologically nontrivial: Here, we exclusively focused on the band topology that can be diagnosed by the set of irreps at high-symmetry momenta.
The simplest way of exploring the nontrivial elements of {BS}phys is thus to consider the complement of {AI}phys in {BS}phys, as in (15) and (52). However, this set has a complicated mathematical structure. To simplify the analysis, we allowed for the formal subtraction of bands and extended the values of to any integer, including the negative ones, à la a K-theory analysis. {BS}phys then becomes an abelian group (known as a “lattice” in the mathematical nomenclature) (14, 15, 35). In other words, there are dBS basis “vectors” , and {BS} can be expressed as . Similarly, by allowing negative integers when taking superposition of AIs, we got another abelian group , which is a subgroup of {BS}. The band topology we are interested in is now encoded in the quotient group
(1) |
dubbed the symmetry-based indicator of band topology (15). As we will see shortly, the quotient group is always a finite abelian group and hence must be a product of the form .
Constructing BSs from AIs
To compute XBS, the natural first step was to identify , the extra compatibility relations enforced by the antiunitary symmetries. Contrary to this expectation, we now show that, on the basis of our previous results on SGs, one can directly compute {BS} and XBS for any MSG without deriving . This served to demonstrate the power of the present approach: Symmetry content and connectivity of BSs could be readily extracted without the large overheads mandated by the conventional approach.
To this end, we first revisited the relevant aspects of the theory for an SG . By definition, is a subgroup of , and therefore, a priori, it could be the case that strictly. However, by an explicit computation for all the 230 SGs, Po et al. (15) found that always holds. This statement has an important implication: Every can be expanded as b = ∑iqiai with rational coefficients using a basis of . In other words, full knowledge of the group can be obtained from that of .
On the basis of this result for SGs, we will now prove the same statement, namely, dBS = dAI, for any MSG . This result enabled us to efficiently compute {BS} and XBS for all MSGs and MLGs using only information contained in {AI}, which could be readily extracted from the tabulated Wyckoff positions (46–48).
Our proof was centered on the following observation: Recall , where is unitary and is the antiunitary part generated by for some spatial symmetry g0. Note that, for type II MSGs, and can be chosen to be the identity, whereas for types III and IV, . Now, consider a -symmetric collection of fully filled local orbitals in real space, which defines an AI . This AI is not generally symmetric under ,; that is, it may not be invariant under the action of . However, if we stack it together with its -transformed copy, we will arrive at an -symmetric AI. Algebraically, this means .
In the momentum space, a similar symmetrization procedure could be performed on the representation content. Suppose that {|k, i〉} is a basis of an irrep of . Then, the -transformed copy, , forms a basis of an irrep of (section S4). When b represents a BS that contains -times, we denote by a BS that contains the same number of times. For any , satisfies the compatibility conditions and hence belongs to {BS}. We presented the explicit form of in section S6.
We are now ready to prove the statement. Observe that any B ∈ {BS} also belongs to , and as , it can be expanded on the basis of
(2) |
Now, we symmetrize both sides of Eq. 2 by . Since B is -invariant, , so
(3) |
As argued, (note that linearity was invoked when we extended the argument form to a general element of ). This proves that any B ∈ {BS} can be expanded in terms of {AI} (using rational coefficients), an equivalent statement of dAI = dBS. Furthermore, it implies that the quotient group XBS = {BS}/{AI} does not contain any -factor and hence is a finite abelian group of the form .
To summarize, we showed that the set of AIs and BSs was identical as far as their dimensionality goes. This is a powerful statement, since it means that we could simply focus on AIs, study their symmetry representations in momentum space, and then take rational combinations to generate all BSs. However, we caution that one has to properly rescale the entries of n when an irrep is paired with another copy of itself according to the Herring rule. For full details of the treatment, we refer interested readers to (15).
Following this strategy, we performed the first calculation of dBS, XBS, and νBS for all of the 1651 MSGs and 528 MLGs. The full list of the computation results are tabulated in tables S1 to S7. For readers’ convenience, we reproduced a few examples from these tables in Tables 1 and 2.
Table 1. Characterization of BSs in a MSG (excerpt from tables S1 to S6).
MSG* | d† | XBS‡ | νBS§ | |
2.4 | I | 9 | (2, 2, 2, 4) | 1 |
2.7 | IV | 5 | (2) | 2 |
3.4 | IV | 3 | (2) | 2 |
209.51 | IV | 3 | (1) | 2a |
*MSG number in the BNS notation, followed by a Roman numeral I, …, IV indicating its type.
†Number of linearly independent BSs.
‡Symmetry-based indicator of band topology, which takes the form , denoted by the collection of positive integers (n1, n2, ⋯).
§For most of the MSGs, the set of physical BS fillings {ν}BS and the set of AI fillings {ν}AI agree with each other, and they take the form . The superscript letter a indicates violation to this rule, as detailed in table S8.
Table 2. Characterization of MLGs through the corresponding MSG (excerpt from table S7).
MSG* | d† | XBS† | vBS† | |
2.5 (1) | II | 5 | (2) | 2 |
2.5 (2) | II | 5 | (2) | 2 |
2.5 (3) | II | 5 | (2) | 2 |
3.4 (1) | IV | 3 | (2) | 2 |
3.4 (2) | IV | 2 | (1) | 2 |
*The numbers in parentheses label the different ways to project the MSG down to 2D planes (section S7).
†Defined as in Table 1.
RESULTS
Representation-enforced semimetals
As an application of our results, here, we introduce new classes of representation-enforced semimetals (reSMs). A BS is said to be a reSM when (i) there are gap closings at some generic (that is, not high symmetry) momenta in the Brillouin zone and (ii) the gaplessness is mandated by the combination of irreps at high-symmetry momenta. Note that only unitary symmetries are incorporated in defining a high-symmetry momentum, and the notion of a BS only requires a continuous gap at all high-symmetry points, lines, and planes. reSMs arise when the topological band gap closings are buried in the low-symmetry regions of the Brillouin zone, rendering them hard to diagnose in conventional electronic structure calculations, but are nonetheless readily detectable and are robust against numerical uncertainty because of their representation-enforced nature. All previously known examples of reSMs concern TR-broken systems in 3D with either the inversion (35, 41) or rotation symmetry (42). The new classes we propose here are realized in spin-orbit–coupled systems in 2D or 3D and are associated with type III or IV MSGs.
Our example is the type IV MSG 3.4 (Pa112) that is generated by C2z (the π rotation about z axis) and (half translation by , followed by TR) apart from lattice translations. (The type III MSG 13.69 can also host a Dirac semimetal by means of a nearly identical mechanism; see section S8.) According to Table 1, this MSG has two classes of combinations of C2z eigenvalues, as indicated by . As we shall see now, the 3D BS belonging to this nontrivial class hosts at least four Weyl points, where the Weyl points with opposite charities are maximally separated in the k3 direction.
To understand this band topology step by step, let us first remove the lattice translation in z and take a single layer parallel to the xy plane. The MSG then reduces to the MLG pa112, which corresponds to the entry 3.4 (1) in Table 2 with . The rotation C2z is a symmetry at the four TR-invariant momenta (TRIMs) of the 2D BZ, whose eigenvalues are either ± i for spinful electrons. When k1 = 0, the antiunitary symmetry with demands the pairing of the two eigenvalues of C2z. In contrast, when k1 = π, , and no pairing is required. For this MLG, , and the topological index distinguishing the two classes is given by
(4) |
where ηn,k is the C2z eigenvalue of the n-th band. η = +1 for any AI, while η = −1 for nontrivial BSs. Note that this invariant is similar to those identified in some previous works (35, 38, 41, 42, 53).
Next, we show that the nontrivial BS with η = −1 corresponds to 2D magnetic Dirac reSMs (twofold degeneracy at the gapless point) enabled by a strong spin-rotation symmetry breaking, as illustrated in Fig. 1. Denoting the standard Pauli matrices by σ0,…,3, the combined symmetry dictates that the local 2-by-2 Hamiltonian h(k) near a gapless point takes the form
(5) |
that is, the σ3 term is missing. Since the number of tunable parameters k1 and k2 is the same as the number of relevant coefficients [g1(k) and g2(k)], the gaplessness is stable against perturbations respecting the symmetries (54). To see this more explicitly, we note that the symmetry quantizes the Berry phase to be 0 or π at any k1. The topological index η = −1 dictates that B(0) ≠ B(π), implying the presence of a gap closing somewhere in 0 < k1 < π. Using a k ⋅ p analysis, one can confirm that the bands disperse linearly, leading to a Dirac node. We note that, compared to the protection of Dirac node by an eigenvalue exchange of nonsymmorphic symmetries (34, 55–57), the present Dirac node has a more topological character.
Given this understanding of the 2D layer with MLG 3.4, let us now recover the third dimension by stacking the copies of 2D layers with generic interlayer couplings. The resulting BS, fully symmetric under MSG 3.4, host a total of four Weyl points: two on the k3 = 0 plane and the other two on the k3 = π plane, as depicted in Fig. 1D. Note that the two Weyl points on the k3 = 0 plane are related by C2z rotation (or equivalently in this setup, ) symmetry and so must have the same chirality. Consequently, the two Weyl points on the k3 = π plane have the opposite chirality. This large separation of the Weyl points with opposite chiralities dictates the presence of long Fermi arcs on the surface. Let us stress again that this nontrivial BS is indicated by the rotation eigenvalues at TRIMs via the formula in Eq. 4 applied to the k3 = 0 plane of the 3D BS. This illustrates how the ideas described in the present work can guide the search for ideal magnetic Weyl semimetals. We leave a full analysis of magnetic reSMs to future works.
Filling-enforced semimetals
As another application of our theory, we attack the following problem: Given an MSG , what are the electron fillings ν at which band insulators are allowed? One of the primary interests in solving this problem lies in the search of topological semimetals: If a material has an electron filling incompatible with any band insulator, it must have symmetry-protected gaplessness near the Fermi energy, provided that a band theory description is applicable. We refer to these systems as filling-enforced (semi)metals. Once the MSG describing the magnetic order is identified, one can readily predict these systems using our results on νBS. For stoichiometric compounds with perfect crystalline order and commensurate magnetism, one can further show that spatial symmetries quantize the physically allowed fillings, which in turn, allows one to further diagnose the fermiology of the filling-enforced (semi)metals. We relegate a detailed discussion to section S2 and summarize the results on the predicted fermiology in tables S9 to S12.
While one can perform a systematic analysis of filling-enforced semimetals (feSMs) using the tables mentioned above, it is instructive to discuss a physical picture linking a real-space description of the magnetic ordering to the momentum-space obstruction to realizing any band insulator. To this end, we first contrast the effect of topologically trivial and nontrivial magnetic orders on local electronic energy levels. When the surrounding magnetic moments order ferromagnetically, as in Fig. 2A, the electronic levels are Zeeman-split and nondegenerate. In contrast, when the moments are arranged into the hedgehog defect shown in Fig. 2B, the magnetic point group symmetries at the defect core demand that all single-particle local energy levels exhibit even degeneracies; that is, such a magnetic order forbids a gapped state when a lone electron is localized to the defect core.
To be compatible with lattice translations, however, there must be a balance between defects and antidefects in each unit cell. An interesting situation arises when the positions of the defects and antidefects are further related by symmetries prohibiting them from annihilation. As an example, consider the diamond lattice with one sublattice occupied by hedgehog defects, and the other sublattice, symmetry-related to the first by a glide mirror, occupied by antihedgehogs (Fig. 2C). The symmetries of this system are described by MSG 227.131 (type III). Suppose the defect core (that is, the diamond sites) corresponds to an atom with an odd atomic number, which leads to an electron filling of ν = 2 mod 4 per primitive unit cell. According to the previous argument, it should be impossible to obtain an AI by localizing all the electrons. This suggests that the electron filling might lead to an obstruction to forming a band insulator and hence enforcing a (semi)metallic behavior. Our result of νBS = 4 for this MSG (table S6) implies that all band insulators are ruled out at the specified filling of ν = 2 mod 4. In table S10, we further see that this corresponds to a feSM with movable nodal features.
Curiously, the hypothetical structure depicted in Fig. 2C could be relevant for materials with the chemical formula AB2O4 taking the spinel structure, where the A atoms, occupying the diamond sites, are surrounded by the B atoms sitting at the pyrochlore positions. If the B atoms are magnetically ordered into the all-in-all-out configuration (Fig. 2D), it can be viewed as a lattice realization of the described hedgehog-antihedgehog lattice in MSG 227.131. This suggests the following tight-binding model as an example of the feSM described
(6) |
where and represent the fermion operators corresponding to an s-orbital localized to the diamond site i (spin indices are suppressed), 〈ij〉 and 〈〈ij〉〉, respectively, denote first- and second-nearest-neighbor bonds, and dij is the vector connecting the site j to i (the lattice constant of the conventional cell is set to 1). In the term ∝ λ, we let l denote the common nearest neighbor of i and j. Note that one can physically interpret tJ as a modulation of the nearest-neighbor hopping by exchange coupling with the magnetic moment at pyrochlore sites, and λ parameterizes the symmetry-allowed spin-orbit coupling between the second-nearest-neighbor sites.
When tJ = 0, Eq. 6 is TR-symmetric (corresponding to the type II MSG 227.129) and reduces to the undistorted Fu-Kane-Mele model defined on the diamond lattice hosting three symmetry-related Dirac points (table S9) (58). Unlike lattice distortions, which lead to topological insulators (58), TR-symmetry breaking described by tJ ≠ 0 leads to a feSM. The Dirac points in the original Fu-Kane-Mele model are unstable toward this perturbation, and our results indicate that the semimetallic behavior is enforced by a more delicate band connectivity (table S10). This is explicitly verified in the example BSs plotted in Fig. 2 (E to H), showing how the Dirac points split into (doubly degenerate) nodal rings as tJ increases from 0.
This example serves only as a particular instance of the 421 MSGs we identified to be compatible with feSMs (section S2). A full list of these MSGs is presented in tables S9 and S10. Armed with this list, we studied the magnetic structures listed on the Bilbao Crystallographic Server (50) and found that the experimentally characterized magnetic materials YFe4Ge2 and LuFe4Ge2 (and related compounds) in the type III MSG 58.399 are realistic magnetic feSM candidates (59, 60). For these compounds, a nonmagnetic atom with an odd atomic number (Lu or Y) sits at a maximal symmetry site of multiplicity 2, whereas their surrounding magnetic moments (Fe) ordered into a pattern symmetric under the combined symmetry of spatial inversion and TR. Such a magnetic ordering falls into our broad description of nontrivial defect lattices, and these compounds are expected to feature Dirac points, pinned to high-symmetry momenta, near the Fermi energy (table S9).
We remark that the method described here can be readily applied to any other commensurate magnetic crystals where both the chemical formula and the MSG describing the magnetic order have been identified. In particular, a local-moment description is inessential: The filling criterion established on the basis of our computation of νBS applies equally well to systems exhibiting itinerant magnetism, as long as a band theory description remains applicable.
Indicator for antiferromagnetic topological insulator
As the last application of our results, let us make a connection of symmetry-based indicators of MSGs, fully computed in this work, to some previously studied topological insulators in TR-broken settings. As a canonical example, we discuss the type IV MSG 2.7 generated by the inversion symmetry I and the half translation, followed by TR, . This MSG has two classes of the combination of inversion parities, as indicated by in Table 1. The topological index distinguishing the two classes is given by
(7) |
Here, ξn,k is the parity of the n-th occupied band. To understand this formula, note first that even- and odd-parity bands form a Kramers pair at the TRIMs with kz = π. This is because I and anticommute when kz = π. Thus, the four TRIMs with kz = π never contribute to the index. On the other hand, when kz = 0, I and commute, and . This means that the same parity eigenvalue appears twice at TRIMs with kz = 0. In Fig. 3A, we illustrated an example of the parity combination with ξ = −1.
As demonstrated in (61), the index ξ = −1 implies a nontrivial band topology, the so-called “antiferromagnetic topological insulator” that supports gapless surface Dirac modes. The antiferromagnetic topological insulator can be understood as a staggered stacking of Chern insulators, in which a Chern insulator with C = +1 (described by red disks in Fig. 3B) is on every z = n () plane and the one with C = −1 (blue disks) is on () plane. All these Chern insulators are related by the symmetry.
If we introduce a perturbation that breaks the symmetry, the C = +1 Chern insulator at z = n (n ≠ 0) and the C = −1 Chern insulator at can pair-annihilate with each other in an inversion symmetric manner, leaving only the C = +1 Chern insulator on the z = 0 plane, as shown in Fig. 3C. This 1D chiral edge state on the surface of 3D topological insulators has been studied by Sitte et al. (62) and Zhang et al. (63) and its counterparts in many other symmetry settings have been found recently (64–71). Assuming that the inversion symmetry is unbroken, we see that one can diagnose this phase by another symmetry-based indicator for the type I MSG 2.4, which is simply the SG with only the inversion I apart from the lattice translation. According to Table 1, this MSG has classification, and the insulator hosting the 1D chiral edge mode constructed here belongs to the class (0, 0, 0, 2) in this classification.
CONCLUSION
Here, we revisit the old problem of assigning a global irrep label to each connected branch of a BS (11), generalizing our recent results on TR-symmetric systems (15) to all 1651 MSGs. Our central results are the computation of three fundamental quantities associated with magnetic BSs, which are tabulated in section S1: (i) dBS, which characterizes the number of independent building blocks of energy bands; (ii) XBS, which, akin to the Fu-Kane parity criterion (38), serves as a symmetry-based indicator of band topology; and (iii) νBS, which dictates the electron fillings at which band insulators are possible. We further demonstrate the utility of these results by applying them to the study of topological semimetals, focusing on cases where the absence of a band gap is either diagnosed through the symmetry representations at isolated high-symmetry points or mandated by the electron filling. In particular, we identify the exhaustive list of MSGs capable of hosting feSMs (section S2).
Although a full database on our computation of the basis vectors of the groups {BS} and {AI} is not included, we note that one can readily generate such a database (15): Our proof of dBS = dAI for all MSGs implies that full knowledge on {BS} is encoded in that of {AI}, and the latter can be computed by analyzing, for instance, hypothetical structures constructed by combining irreps of the site-symmetry groups to each of the Wyckoff positions. As long as all high-symmetry momenta are included in the analysis and the doubling of any irrep (according to the Herring rule) is incorporated, {BS} can be recovered by forming linear superpositions of the entries in {AI} using rational coefficients, subjected only to the constraint that the resultant irrep multiplicities must all be integer-valued. One can perform these calculations without deriving the compatibility relations, and this circumvents the careful convention-fixing required in both tabulating and using these results. In view of this, we refrain from providing a database on them.
We close by highlighting several interesting open questions that we leave for future studies: (i) performing a comprehensive study on the identification and characterization of reSMs in all 2D and 3D MSGs; (ii) developing a theory of quantized physical responses, or a proof of the absence thereof, for each nontrivial class in the exhaustive list of XBS we computed; and (iii) finding realistic topological materials based on systematic searches using the diagnostics our theory provides. Finally, we remark that some aspects of our filling criterion generalize to the interacting setting (72), similar to the corresponding relation between the results for TR symmetric crystals (8, 10).
Supplementary Material
Acknowledgments
H.W. thank S. Murakami and C. Fang for the useful discussion. Funding: A.V. and H.C.P. were supported by NSF DMR-1411343. A.V. acknowledges the support from a Simons Investigator Award. H.W. acknowledges support from Japan Society for the Promotion of Science KAKENHI grant number JP17K17678. Author contributions: All authors designed the research, performed the research, contributed new reagents/analytic tools, analyzed the data, and wrote the manuscript. Competing interests: The authors declare that they have no competing interests. Data and materials availability: All data needed to evaluate the conclusions in the paper are present in the paper and/or the Supplementary Materials. Additional data related to this paper may be requested from the authors.
SUPPLEMENTARY MATERIALS
REFERENCES AND NOTES
- 1.M. Franz, L. Molenkamp, Eds., Contemporary Concepts of Condensed Matter Science: Topological Insulators (Elsevier, 2013), vol. 6. [Google Scholar]
- 2.C. J. Bradley, A. P. Cracknell, The Mathematical Theory of Symmetry in Solids (Oxford Univ. Press, 1972). [Google Scholar]
- 3.Michel L., Zak J., Connectivity of energy bands in crystals. Phys. Rev. B 59, 5998–6001 (1999). [Google Scholar]
- 4.Michel L., Zak J., Elementary energy bands in crystalline solids. Europhys. Lett. 50, 519 (2000). [Google Scholar]
- 5.Michel L., Zak J., Elementary energy bands in crystals are connected. Phys. Rep. 341, 377–395 (2001). [Google Scholar]
- 6.Parameswaran S. A., Turner A. M., Arovas D. P., Vishwanath A., Topological order and absence of band insulators at integer filling in non-symmorphic crystals. Nat. Phys. 9, 299–303 (2013). [Google Scholar]
- 7.R. Roy, Space group symmetries and low lying excitations of many-body systems at integer fillings. arXiv:1212.2944 (2012).
- 8.Watanabe H., Po H. C., Vishwanath A., Zaletel M. P., Filling constraints for spin-orbit coupled insulators in symmorphic and nonsymmorphic crystals. Proc. Natl. Acad. Sci. U.S.A. 112, 14551 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Po H. C., Watanabe H., Zaletel M. P., Vishwanath A., Filling-enforced quantum band insulators in spin-orbit coupled crystals. Sci. Adv. 2, e1501782 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Watanabe H., Po H. C., Zaletel M. P., Vishwanath A., Filling-enforced gaplessness in band structures of the 230 space groups. Phys. Rev. Lett. 117, 096404 (2016). [DOI] [PubMed] [Google Scholar]
- 11.Bouckaert L. P., Smoluchowski R., Wigner E., Theory of Brillouin zones and symmetry properties of wave functions in crystals. Phys. Rev. 50, 58–67 (1936). [Google Scholar]
- 12.Herring C., Character tables for two space groups. J. Franklin Inst. 233, 525–543 (1942). [Google Scholar]
- 13.Wieder B. J., Kim Y., Rappe A. M., Kane C. L., Double dirac semimetals in three dimensions. Phys. Rev. Lett. 116, 186402 (2016). [DOI] [PubMed] [Google Scholar]
- 14.Kruthoff J., de Boer J., van Wezel J., Kane C. L., Slager R.-J., Topological classification of crystalline insulators through band structure combinatorics. Phys. Rev. X 7, 041069 (2017). [Google Scholar]
- 15.Po H. C., Vishwanath A., Watanabe H., Symmetry-based indicators of band topology in the 230 space groups. Nat. Commun. 8, 50 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Bradlyn B., Elcoro L., Cano J., Vergniory M. G., Wang Z., Felser C., Aroyo M. I., Bernevig B. A., Topological quantum chemistry. Nature 547, 298–305 (2017). [DOI] [PubMed] [Google Scholar]
- 17.Vergniory M. G., Elcoro L., Wang Z., Cano J., Felser C., Aroyo M. I., Bernevig B. A., Bradlyn B., Graph theory data for topological quantum chemistry. Phys. Rev. E 96, 023310 (2017). [DOI] [PubMed] [Google Scholar]
- 18.B. J. Wieder, B. Bradlyn, Z. Wang, J. Cano, Y. Kim, H.-S. D. Kim, A. M. Rappe, C. L. Kane, B. A. Bernevig, Wallpaper fermions and the topological Dirac insulator. arXiv:1705.01617 (2017). [DOI] [PubMed]
- 19.Yu R., Zhang W., Zhang H.-J., Zhang S.-C., Dai X., Fang Z., Quantized anomalous Hall effect in magnetic topological insulators. Science 329, 61–64 (2010). [DOI] [PubMed] [Google Scholar]
- 20.Chang C.-Z., Zhang J., Feng X., Shen J., Zhang Z., Guo M., Li K., Ou Y., Wei P., Wang L.-L., Ji Z.-Q., Feng Y., Ji S., Chen X., Jia J., Dai X., Fang Z., Zhang S.-C., He K., Wang Y., Lu L., Ma X.-C., Xue Q.-K., Experimental observation of the quantum anomalous Hall effect in a magnetic topological insulator. Science 340, 167–170 (2013). [DOI] [PubMed] [Google Scholar]
- 21.Ye J., Kim Y. B., Millis A. J., Shraiman B. I., Majumdar P., Tešanović Z., Berry phase theory of the anomalous Hall effect: Application to colossal magnetoresistance manganites. Phys. Rev. Lett. 83, 3737–3740 (1999). [Google Scholar]
- 22.Taguchi Y., Oohara Y., Yoshizawa H., Nagaosa N., Tokura Y., Spin chirality, berry phase, and anomalous Hall effect in a frustrated ferromagnet. Science 291, 2573–2576 (2001). [DOI] [PubMed] [Google Scholar]
- 23.Lee M., Onose Y., Tokura Y., Ong N. P., Hidden constant in the anomalous Hall effect of high-purity magnet MnSi. Phys. Rev. B 75, 172403 (2007). [Google Scholar]
- 24.Binz B., Vishwanath A., Chirality induced anomalous-Hall effect in helical spin crystals. Physica B Condens. Matter 403, 1336–1340 (2008). [Google Scholar]
- 25.Neubauer A., Pfleiderer C., Binz B., Rosch A., Ritz R., Niklowitz P. G., Böni P., Topological Hall effect in the A phase of MnSi. Phys. Rev. Lett. 102, 186602 (2009). [DOI] [PubMed] [Google Scholar]
- 26.Nakatsuji S., Kiyohara N., Higo T., Large anomalous Hall effect in a non-collinear antiferromagnet at room temperature. Nature 527, 212–215 (2015). [DOI] [PubMed] [Google Scholar]
- 27.Nayak A. K., Fischer J. E., Sun Y., Yan B., Karel J., Komarek A. C., Shekhar C., Kumar N., Schnelle W., Kübler J., Felser C., Parkin S. S. P., Large anomalous Hall effect driven by a nonvanishing berry curvature in the noncolinear antiferromagnet Mn3Ge. Sci. Adv. 2, e1501870 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Yang H., Sun Y., Zhang Y., Shi W.-J., Parkin S. S. P., Yan B., Topological Weyl semimetals in the chiral antiferromagnetic materials Mn3Ge and Mn3Sn. New J. Phys. 19, 015008 (2017). [Google Scholar]
- 29.Fert A., Cros V., Sampaio J., Skyrmions on the track. Nat. Nanotechnol. 8, 152–156 (2013). [DOI] [PubMed] [Google Scholar]
- 30.Aroyo M. I., Kirov A., Capillas C., Perez-Mato J. M., Wondratschek H., Bilbao crystallographic server. II. Representations of crystallographic point groups and space groups. Acta Crystallogr. A 62, 115–128 (2006). [DOI] [PubMed] [Google Scholar]
- 31.Gong C., Li L., Li Z., Ji H., Stern A., Xia Y., Cao T., Bao W., Wang C., Wang Y., Qiu Z. Q., Cava R. J., Louie S. G., Xia J., Zhang X., Discovery of intrinsic ferromagnetism in two-dimensional van der Waals crystals. Nature 546, 265–269 (2017). [DOI] [PubMed] [Google Scholar]
- 32.Huang B., Clark G., Navarro-Moratalla E., Klein D. R., Cheng R., Seyler K. L., Zhong D., Schmidgall E., McGuire M. A., Cobden D. H., Yao W., Xiao D., Jarillo-Herrero P., Xu X., Layer-dependent ferromagnetism in a van der Waals crystal down to the monolayer limit. Nature 546, 270–273 (2017). [DOI] [PubMed] [Google Scholar]
- 33.Chen R., Po H. C., Neaton J. B., Vishwanath A., Topological materials discovery using electron filling constraints. Nat. Phys. 14, 55–61 (2018). [Google Scholar]
- 34.Young S. M., Wieder B. J., Filling-enforced magnetic Dirac semimetals in two dimensions. Phys. Rev. Lett. 118, 186401 (2017). [DOI] [PubMed] [Google Scholar]
- 35.Turner A. M., Zhang Y., Mong R. S. K., Vishwanath A., Quantized response and topology of magnetic insulators with inversion symmetry. Phys. Rev. B 85, 165120 (2012). [Google Scholar]
- 36.Thouless D. J., Wannier functions for magnetic sub-bands. J. Phys. C Solid State Phys. 17, L325 (1984). [Google Scholar]
- 37.Thonhauser T., Vanderbilt D., Insulator/Chern-insulator transition in the Haldane model. Phys. Rev. B 74, 235111 (2006). [Google Scholar]
- 38.Fu L., Kane C. L., Topological insulators with inversion symmetry. Phys. Rev. B 76, 045302 (2007). [Google Scholar]
- 39.Turner A. M., Zhang Y., Vishwanath A., Entanglement and inversion symmetry in topological insulators. Phys. Rev. B 82, 241102 (2010). [Google Scholar]
- 40.Wan X., Turner A. M., Vishwanath A., Savrasov S. Y., Topological semimetal and Fermi-arc surface states in the electronic structure of pyrochlore iridates. Phys. Rev. B 83, 205101 (2011). [Google Scholar]
- 41.Hughes T. L., Prodan E., Bernevig B. A., Inversion-symmetric topological insulators. Phys. Rev. B 83, 245132 (2011). [Google Scholar]
- 42.Fang C., Gilbert M. J., Bernevig B. A., Bulk topological invariants in noninteracting point group symmetric insulators. Phys. Rev. B 86, 115112 (2012). [Google Scholar]
- 43.Freed D. S., Moore G. W., Twisted equivariant matter. Ann. Henri Poincaré 14, 1927–2023 (2013). [Google Scholar]
- 44.Read N., Compactly supported Wannier functions and algebraic K-theory. Phys. Rev. B 95, 115309 (2017). [Google Scholar]
- 45.Shiozaki K., Sato M., Gomi K., Topological crystalline materials: General formulation, module structure, and wallpaper groups. Phys. Rev. B 95, 235425 (2017). [Google Scholar]
- 46.Litvin D. B., Magnetic space-group types. Acta Crystallogr. A 57, 729–730 (2001). [DOI] [PubMed] [Google Scholar]
- 47.D. B. Litvin, Magnetic Group Tables: 1-, 2- and 3-Dimensional Magnetic Subperiodic Groups and Magnetic Space Groups (International Union of Crystallography, 2013). [Google Scholar]
- 48.H. T. Stokes, D. M. Hatch, B. J. Campbell, ISOTROPY Software Suite; http://iso.byu.edu/iso/isotropy.php.
- 49.Zhang R.-X., Liu C.-X., Topological magnetic crystalline insulators and corepresentation theory. Phys. Rev. B 91, 115317 (2015). [Google Scholar]
- 50.Gallego S. V., Perez-Mato J. M., Elcoro L., Tasci E. S., Hanson R. M., Momma K., Aroyo M. I., Madariaga G., MAGNDATA: Towards a database of magnetic structures. I. The commensurate case. J. Appl. Cryst. 49, 1750–1776 (2016). [Google Scholar]
- 51.T. Hahn, International Tables for Crystallography, Vol. A: Space-Group Symmetry (Springer, ed. 5, 2006). [Google Scholar]
- 52.H. C. Po, H. Watanabe, A. Vishwanath, Fragile topology and Wannier obstructions. arXiv:1709.06551 (2017). [DOI] [PubMed]
- 53.Kim Y., Wieder B. J., Kane C. L., Rappe A. M., Dirac line nodes in inversion-symmetric crystals. Phys. Rev. Lett. 115, 036806 (2015). [DOI] [PubMed] [Google Scholar]
- 54.Ahn J., Yang B.-J., Unconventional topological phase transition in two-dimensional systems with space-time inversion symmetry. Phys. Rev. Lett. 118, 156401 (2017). [DOI] [PubMed] [Google Scholar]
- 55.Young S. M., Kane C. L., Dirac semimetals in two dimensions. Phys. Rev. Lett. 115, 126803 (2015). [DOI] [PubMed] [Google Scholar]
- 56.Wang J., Antiferromagnetic Dirac semimetals in two dimensions. Phys. Rev. B 95, 115138 (2017). [Google Scholar]
- 57.Tsirkin S. S., Souza I., Vanderbilt D., Composite Weyl nodes stabilized by screw symmetry with and without time-reversal invariance. Phys. Rev. B 96, 045102 (2017). [Google Scholar]
- 58.Fu L., Kane C. L., Mele E. J., Topological insulators in three dimensions. Phys. Rev. Lett. 98, 106803 (2007). [DOI] [PubMed] [Google Scholar]
- 59.Schobinger-Papamantellos P., Rodríguez-Carvajal J., André G., Duong N. P., Buschow K. H. J., Tolédano P., Simultaneous structural and magnetic transitions in YFe4Ge2 studied by neutron diffraction and magnetic measurements. J. Magn. Magn. Mater. 236, 14–27 (2001). [Google Scholar]
- 60.Schobinger-Papamantellos P., Buschow K. H. J., Rodríguez-Carvajal J., Magnetoelastic phase transitions in the LuFe4Ge2 and YFe4Si2 compounds: A neutron diffraction study. J. Magn. Magn. Mater. 324, 3709–3715 (2012). [Google Scholar]
- 61.Mong R. S. K., Essin A. M., Moore J. E., Antiferromagnetic topological insulators. Phys. Rev. B 81, 245209 (2010). [Google Scholar]
- 62.Sitte M., Rosch A., Altman E., Fritz L., Topological insulators in magnetic fields: Quantum Hall effect and edge channels with a nonquantized θ term. Phys. Rev. Lett. 108, 126807 (2012). [DOI] [PubMed] [Google Scholar]
- 63.Zhang F., Kane C. L., Mele E. J., Surface state magnetization and chiral edge states on topological insulators. Phys. Rev. Lett. 110, 046404 (2013). [DOI] [PubMed] [Google Scholar]
- 64.Benalcazar W. A., Teo J. C. Y., Hughes T. L., Classification of two-dimensional topological crystalline superconductors and majorana bound states at disclinations. Phys. Rev. B 89, 224503 (2014). [DOI] [PubMed] [Google Scholar]
- 65.Benalcazar W. A., Bernevig B. A., Hughes T. L., Quantized electric multipole insulators. Science 357, 61–66 (2017). [DOI] [PubMed] [Google Scholar]
- 66.Hashimoto K., Wu X., Kimura T., Edge states at an intersection of edges of a topological material. Phys. Rev. B 95, 165443 (2017). [Google Scholar]
- 67.Song Z., Fang Z., Fang C., (d − 2)-dimensional edge states of rotation symmetry protected topological states. Phys. Rev. Lett. 119, 246402 (2017). [DOI] [PubMed] [Google Scholar]
- 68.Schindler F., Cook A. M., Vergniory M. G., Wang Z., Parkin S. S. P., Bernevig B. A., Neupert T., Higher-order topological insulators. Sci. Adv. 4, eaat0346 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Benalcazar W. A., Bernevig B. A., Hughes T. L., Electric multipole moments, topological multipole moment pumping, and chiral hinge states in crystalline insulators. Phys. Rev. B 96, 245115 (2017). [Google Scholar]
- 70.C. Fang, L. Fu, Rotation anomaly and topological crystalline insulators. arXiv:1709.01929 (2017). [DOI] [PMC free article] [PubMed]
- 71.Langbehn J., Peng Y., Trifunovic L., von Oppen F., Brouwer P. W., Reflection-symmetric second-order topological insulators and superconductors. Phys. Rev. Lett. 119, 246401 (2017). [DOI] [PubMed] [Google Scholar]
- 72.Watanabe H., Lieb-Schultz-Mattis-type filling constraints in the 1651 magnetic space groups. Phys. Rev. B 97, 165117 (2018). [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.