weyl semimetals from noncentrosymmetric topological...

10
PHYSICAL REVIEW B 90, 155316 (2014) Weyl semimetals from noncentrosymmetric topological insulators Jianpeng Liu and David Vanderbilt Department of Physics and Astronomy, Rutgers University, Piscataway, New Jersey 08854-8019, USA (Received 22 September 2014; published 28 October 2014) We study the problem of phase transitions from three-dimensional topological to normal insulators without inversion symmetry. In contrast with the conclusions of some previous work, we show that a Weyl semimetal always exists as an intermediate phase regardless of any constriant from lattice symmetries, although the interval of the critical region is sensitive to the choice of path in the parameter space and can be very narrow. We demonstrate this behavior by carrying out first-principles calculations on the noncentrosymmetric topological insulators LaBiTe 3 and LuBiTe 3 and the trivial insulator BiTeI. We find that a robust Weyl-semimetal phase exists in the solid solutions LaBi 1x Sb x Te 3 and LuBi 1x Sb x Te 3 for x 38.5%–41.9% and x 40.5%–45.1%, respectively. A low-energy effective model is also constructed to describe the critical behavior in these two materials. In BiTeI, a Weyl semimetal also appears with applied pressure, but only within a very small pressure range, which may explain why it has not been experimentally observed. DOI: 10.1103/PhysRevB.90.155316 PACS number(s): 73.43.Nq, 73.20.At, 78.40.Kc I. INTRODUCTION The significance of topology in determining electronic properties has became widely appreciated with the discovery of the integer quantum Hall effect and been highlighted further by the recent interest in topological insulators (TIs) [14]. In topological band theory, a topological index, such as the Chern number or the Z 2 index, is well defined only for gapped systems, and the topological character is signaled by the presence of novel gapless surface states which cannot exist in any isolated two-dimensional (2D) system [3,4]. Recently, the concept of topological phases is further generalized to three- dimensional (3D) bulk gapless systems, whose topological behavior is protected by lattice translational symmetry, known as the Weyl semimetal (WSM) [59]. A Weyl semimetal is characterized by a Fermi energy that intersects the bulk bands only at one or more pairs of band-touching points (BTPs) between nondegenerate valence and conduction bands. This can occur in the presence of spin-orbit coupling (SOC), typically in a crystal with broken time-reversal or inversion symmetry but not both, so that the pairs are of the form (k 0 , k 0 ) in the Brillouin zone (BZ). The effective Hamiltonian around a single BTP k 0 can be written as H (k) = f 0 (k) + f (k) · σ , where f 0 and f are scalar and vector functions, respectively, of wave vector in the BZ and the σ j are the Pauli matrices acting in the two-band space. If one expands the coefficient f (k) to linear order around k 0 , one gets a Hamiltonian having the form of the Weyl Hamil- tonian in relativistic quantum mechanics after a coordinate transformation in k space. If the sign of the determinant of the Jacobian that describes the coordinate transformation is positive (negative), we call the BTP a Weyl node with positive (negative) chirality, and the low-energy excitations around such a Weyl node provide a condensed-matter realization of left-handed (right-handed) Weyl fermions. These pairs of Weyl nodes are topologically protected in the sense that they are robust against small perturbations, which can be seen from the codimension argument as follows. One can introduce a parameter λ that acts as a perturbation on the BTP, and let both f 0 and f be dependent on λ. In order to get a band touching at (k 0 0 ), the three coefficients f (f x ,f y ,f z ) have to vanish. However, since there are four degrees of freedom, if λ 0 λ 0 + δλ, instead of opening a gap, the Weyl node would just shift slightly in momentum to compensate for the perturbation. In fact, there is no way to remove a Weyl node unless two Weyl nodes with opposite chirality annihilate each other. If the two Weyl nodes are aligned in energy due to either time-reversal or some lattice symmetry, and the bands are filled right up to the Weyl nodes, then the Fermi energy would be locked there regardless of weak perturbations. That is, the Fermi level could be slightly shifted upward (downward) due to some weak perturbation, such that there is an electronlike (holelike) Fermi surface, then there must also be a holelike (electronlike) Fermi surface to conserve the total number of electrons, which is impossible in such a a semimetal. It follows that the low-energy physics in the Weyl semimetal is completely dominated by the linearly dispersing states around the Weyl nodes, which leads to interesting surface states and transport properties. The presence of Weyl nodes in the bulk band structure is responsible for the presence of Fermi arcs at the surface, which can be understood as follows [7,8]. Consider a small loop in the 2D surface BZ that encloses the projection along k z of one Weyl point. When translated along k z , this loop traces out a surface in the 3D BZ, and the application of Gauss’s theorem implies that the Chern number on this surface must equal the chirality of the enclosed Weyl node. It follows that as (k x ,k y ) is carried around the loop, a single electron is pumped up to (or down from) the top surface, and this is only consistent with charge conservation if a single surface state crosses the Fermi energy E F during the cycle. Since this argument applies for an arbitrary loop, surface states must exist at E F along some arc emerging from the surface-projected Weyl point. If there is another Weyl node with opposite chirality, then the Chern number can vanish once the cylinder encloses both of the nodes, such that the Fermi arc would only extend between the two projected Weyl nodes [7,8]. It is also interesting to note that the surface states residing on the top and bottom surfaces disperse linearly in opposite directions and remain robust even when the Fermi energy deviates from the bulk Weyl nodes [10]. 1098-0121/2014/90(15)/155316(10) 155316-1 ©2014 American Physical Society

Upload: others

Post on 10-Feb-2021

7 views

Category:

Documents


0 download

TRANSCRIPT

  • PHYSICAL REVIEW B 90, 155316 (2014)

    Weyl semimetals from noncentrosymmetric topological insulators

    Jianpeng Liu and David VanderbiltDepartment of Physics and Astronomy, Rutgers University, Piscataway, New Jersey 08854-8019, USA

    (Received 22 September 2014; published 28 October 2014)

    We study the problem of phase transitions from three-dimensional topological to normal insulators withoutinversion symmetry. In contrast with the conclusions of some previous work, we show that a Weyl semimetalalways exists as an intermediate phase regardless of any constriant from lattice symmetries, although the intervalof the critical region is sensitive to the choice of path in the parameter space and can be very narrow. Wedemonstrate this behavior by carrying out first-principles calculations on the noncentrosymmetric topologicalinsulators LaBiTe3 and LuBiTe3 and the trivial insulator BiTeI. We find that a robust Weyl-semimetal phaseexists in the solid solutions LaBi1−xSbxTe3 and LuBi1−xSbxTe3 for x ≈ 38.5%–41.9% and x ≈40.5%–45.1%,respectively. A low-energy effective model is also constructed to describe the critical behavior in these twomaterials. In BiTeI, a Weyl semimetal also appears with applied pressure, but only within a very small pressurerange, which may explain why it has not been experimentally observed.

    DOI: 10.1103/PhysRevB.90.155316 PACS number(s): 73.43.Nq, 73.20.At, 78.40.Kc

    I. INTRODUCTION

    The significance of topology in determining electronicproperties has became widely appreciated with the discoveryof the integer quantum Hall effect and been highlighted furtherby the recent interest in topological insulators (TIs) [1–4].In topological band theory, a topological index, such as theChern number or the Z2 index, is well defined only for gappedsystems, and the topological character is signaled by thepresence of novel gapless surface states which cannot exist inany isolated two-dimensional (2D) system [3,4]. Recently, theconcept of topological phases is further generalized to three-dimensional (3D) bulk gapless systems, whose topologicalbehavior is protected by lattice translational symmetry, knownas the Weyl semimetal (WSM) [5–9].

    A Weyl semimetal is characterized by a Fermi energythat intersects the bulk bands only at one or more pairs ofband-touching points (BTPs) between nondegenerate valenceand conduction bands. This can occur in the presence ofspin-orbit coupling (SOC), typically in a crystal with brokentime-reversal or inversion symmetry but not both, so that thepairs are of the form (k0, −k0) in the Brillouin zone (BZ).The effective Hamiltonian around a single BTP k0 can bewritten as H (k) = f0(k) + f(k) · σ , where f0 and f are scalarand vector functions, respectively, of wave vector in the BZand the σj are the Pauli matrices acting in the two-band space.If one expands the coefficient f(k) to linear order around k0,one gets a Hamiltonian having the form of the Weyl Hamil-tonian in relativistic quantum mechanics after a coordinatetransformation in k space. If the sign of the determinant ofthe Jacobian that describes the coordinate transformation ispositive (negative), we call the BTP a Weyl node with positive(negative) chirality, and the low-energy excitations aroundsuch a Weyl node provide a condensed-matter realization ofleft-handed (right-handed) Weyl fermions.

    These pairs of Weyl nodes are topologically protected inthe sense that they are robust against small perturbations,which can be seen from the codimension argument as follows.One can introduce a parameter λ that acts as a perturbationon the BTP, and let both f0 and f be dependent on λ. Inorder to get a band touching at (k0,λ0), the three coefficients

    f ≡ (fx,fy,fz) have to vanish. However, since there are fourdegrees of freedom, if λ0 → λ0 + δλ, instead of opening agap, the Weyl node would just shift slightly in momentum tocompensate for the perturbation. In fact, there is no way toremove a Weyl node unless two Weyl nodes with oppositechirality annihilate each other.

    If the two Weyl nodes are aligned in energy due to eithertime-reversal or some lattice symmetry, and the bands arefilled right up to the Weyl nodes, then the Fermi energywould be locked there regardless of weak perturbations.That is, the Fermi level could be slightly shifted upward(downward) due to some weak perturbation, such that there isan electronlike (holelike) Fermi surface, then there must alsobe a holelike (electronlike) Fermi surface to conserve the totalnumber of electrons, which is impossible in such a a semimetal.It follows that the low-energy physics in the Weyl semimetal iscompletely dominated by the linearly dispersing states aroundthe Weyl nodes, which leads to interesting surface states andtransport properties.

    The presence of Weyl nodes in the bulk band structure isresponsible for the presence of Fermi arcs at the surface, whichcan be understood as follows [7,8]. Consider a small loop inthe 2D surface BZ that encloses the projection along kz of oneWeyl point. When translated along kz, this loop traces out asurface in the 3D BZ, and the application of Gauss’s theoremimplies that the Chern number on this surface must equal thechirality of the enclosed Weyl node. It follows that as (kx,ky)is carried around the loop, a single electron is pumped up to(or down from) the top surface, and this is only consistentwith charge conservation if a single surface state crosses theFermi energy EF during the cycle. Since this argument appliesfor an arbitrary loop, surface states must exist at EF alongsome arc emerging from the surface-projected Weyl point. Ifthere is another Weyl node with opposite chirality, then theChern number can vanish once the cylinder encloses both ofthe nodes, such that the Fermi arc would only extend betweenthe two projected Weyl nodes [7,8]. It is also interesting tonote that the surface states residing on the top and bottomsurfaces disperse linearly in opposite directions and remainrobust even when the Fermi energy deviates from the bulk Weylnodes [10].

    1098-0121/2014/90(15)/155316(10) 155316-1 ©2014 American Physical Society

    http://dx.doi.org/10.1103/PhysRevB.90.155316

  • JIANPENG LIU AND DAVID VANDERBILT PHYSICAL REVIEW B 90, 155316 (2014)

    In a WSM with broken time-reversal (TR) symmetry, thereis also a nonzero anomalous Hall conductivity (AHC) that isclosely related to the positions of the Weyl nodes [11–13].Consider the AHC σxy with the crystal oriented such that thethird primitive reciprocal vector is along z. If we track theChern number Cz of a 2D slice normal to z in the 3D BZ, wemust find that it changes by ±1 whenever kz passes a Weylnode, with the sign depending on the chirality of the node. Inthe simplest case, if Cz =1 for kz between the two Weyl nodesand zero elsewhere, then the AHC is just proportional to theseparation of the two Weyl nodes in kz. This is also interpretedas a consequence of the “chiral anomaly” in a WSM [14].Other interesting transport phenomena can arise due to thechiral anomaly. For example, if a magnetic field B is appliedto a WSM in the z direction, Landau levels will be formed in the(x,y) plane. The zeroth Landau level disperses linearly alongkz, but in opposite directions for Weyl nodes with oppositechirality. As a result, if an electric field E is applied along z,electrons would be pumped from one Weyl node to the otherat a rate proportional to E · B, with the Fermi arcs serving asa conduit [5,8,13,15].

    As discussed above, a WSM requires the breaking of eitherTR or inversion symmetry. Many of the previous works arefocused on WSMs without TR symmetry, such as in py-rochlore iridates [7], magnetically doped TI multilayers [12],and Hg1−x−yCdxMnyTe [16]. In this paper, we study theWSM with preserved TR symmetry but broken inversionsymmetry.

    It was argued some time ago that the Z2-odd and Z2-evenphases of a noncentrosymmetric insulator should always bebridged by a critical WSM phase [17,18]. If the transition isdescribed by some adiabatic parameter λ, then as λ increasesone expects first the appearance of m higher-order BTPs inthe half BZ (and another m at the time-reversed points), wherem=1 is typical of low-symmetry systems, while m>1 canoccur when, e.g., rotational symmetries are present. Thesehigher-order BTPs generally have quadratic dispersion in onedirection while remaining linear in the other two, and arenonchiral; we refer to such a point henceforth as a “quadraticBTP.” As λ increases, each quadratic BTP splits to form a pairof Weyl nodes (4m altogether), which then migrate throughthe BZ and eventually annihilate at a second critical value ofλ after exchanging partners. The previous work demonstratedthat this process inverts the strongZ2 index if m is odd [17,18].Recently, however, Yang et al. claimed that for systems withcertain high-symmetry lines in the BZ, the phase transitioncould occur at a unique critical value of λ at which thebands would touch and immediately reopen, instead of oversome finite interval in λ, even when inversion symmetry isabsent [19]. These authors suggested that BiTeI under pressurecould serve as an example to support their claim [19,20].

    In this paper, we address this issue carefully. We show thatan intermediate critical WSM phase should always exist forany topological phase transition (TPT) between a normal anda Z2-odd insulating phase. We find, however, that the widthof the critical WSM phase can be sensitive to the choice ofpath in parameter space and can sometimes be very small. Tojustify our conclusions, we take specific materials as examples.We first study the TPT in the solid solutions LaBi1−xSbxTe3and LuBi1−xSbxTe3 using the virtual crystal approximation,

    where the phase transition is driven by Sb substitution. Theparent compounds at x =100%, LaBiTe3 and LuBiTe3, arehypothetical noncentrosymmetric materials that are predictedto be strong topological insulators in Ref. [21] and in thepresent work, respectively. Instead, the end members LaSbTe3and LuSbTe3 at x =0% are trivial insulators [21]. We find thata WSM phase is obtained when x is in the range of about38.5%–41.9% for LaBiTe3 and 40.5%–45.1% for LuBiTe3.We further construct a low-energy effective model to describethe topological and phase-transitional behavior in this classof materials. We also revisit the TPT of BiTeI driven byapplied pressure, where a WSM phase has not previouslybeen observed [20,22]. Based on our calculations, we findthat a small interval of WSM phase does actually intervene asincreasing pressure drives the system from the trivial to thetopological phase.

    The paper is organized as follows. In Sec. II we derive thegeneral behavior of TPTs in noncentrosymmetric insulatorsand point out some deficiencies in the discussion of BiTeIby Yang et al. [19] In Sec. III we describe the latticestructures and basic topological properties of the materials, aswell as the numerical methods used in the realistic-materialcalculations, especially the methods used in modeling thealloyed and pressurized systems and in searching for BTPs inthe BZ. In Sec. IV we present the results for LaBi1−xSbxTe3,LuBi1−xSbxTe3, and BiTeI, and discuss the sensitivity to thechoice of path. In Sec. V, we summarize our work.

    II. TOPOLOGICAL TRANSITION INNONCENTROSYMMETRIC INSULATORS

    A. General behavior

    We consider the problem of TPTs in noncentrosymmetricinsulators in the most general case. In the space of the twobands which touch at the TPT, the system can be described bythe effective Hamiltonian

    H (k,λ) = fx(k,λ)σx + fy(k,λ)σy + fz(k,λ)σz, (1)where λ is the parameter that drives the TPT and σx,y,z arethe three Pauli matrices defined in the space spanned by thehighest occupied and the lowest unoccupied states at k. Sincewe study the TPT between two insulating phases, we canassume without loss of generality that the system is gappedfor λ

  • WEYL SEMIMETALS FROM NONCENTROSYMMETRIC . . . PHYSICAL REVIEW B 90, 155316 (2014)

    the assumption that the system is insulating for λ

  • JIANPENG LIU AND DAVID VANDERBILT PHYSICAL REVIEW B 90, 155316 (2014)

    FIG. 1. (Color online) (a) The lattice structure of LaBiTe3,LuBiTe3, LaSbTe3, and LuSbTe3. (b) The BZ of La(Lu)Bi(Sb)Te3.(c) The lattice structure of BiTeI. (d) The BZ of BiTeI.

    As has been verified in Ref. [19], the Jacobian does remainof rank two on these lines in BiTeI, and we shall show belowin Sec. IV B that an intermediate WSM phase does occur. Wealso point out that Fig. 2 of Ref. [19] does not demonstrate theabsence of the Weyl nodes, since they are expected to lie offthe (kx,kz) plane on which the dispersion was plotted.

    Yang et al. [19] gave another argument in favor of theinsulator-insulator scenario in BiTeI as follows. They notedthat the band touching first takes place on the A-H line,which is invariant under the combination of time-reversal andmirror operations. This imposes some constraints on the formof the effective Hamiltonian around the BTP, and from thesethe authors concluded that, if Weyl nodes do appear, theyshould migrate along trajectories of the form p1 ∼ ±δλ1/2,p2 ∼ ±δλ3/2, and p3 ∼ δλ. Such a curve in 3D space possessesnonzero torsion, so that the trajectories of the two Weyl nodesemerging from one quadratic BTP could never join againand form a closed curve. This implies that if the WSM isformed by such an event, then it would remain permanently,contradicting the fact that BiTeI clearly becomes a globallygapped TI at higher pressures. Based on this reasoning, theyconcluded that the TPT in BiTeI must be an insulator-insulatortransition without an intermediate WSM.

    However, this argument neglects the fact that the C3vsymmetry means that there are several A-H lines in the BZ ofBiTeI, and the gap first closes by the simultaneous appearanceof quadratic BTPs at equivalent positions on all of these lines.Even though the two Weyl nodes which emerge from a singlequadratic BTP cannot meet each other, as shown from thetorsion of their trajectories, the Weyl nodes from differentBTPs can interchange partners and eventually annihilate eachother in such a way as to form a closed curve in the BZ. Thisis exactly the mechanism of the topological phase transitionin noncentrosymmetric TIs [17,18]. As will be discussed inSec. IV B, there are actually six quadratic BTPs in the fullBZ that appear simultaneously, according to the crystallineand TR symmetries. These six Dirac nodes split into 12 Weylnodes, which are eventually gapped out by annihilation afterexchanging partners.

    In the following section, we will study the TPTs invarious inversion asymmetric materials by first-principlescalculations. We predict LaBi1−xSbxTe3 and LuBi1−xSbxTe3to be WSM candidates within a certain range of impuritycomposition x. We also revisit the case of BiTeI, and find thata WSM phase emerges when external pressure is applied toBiTeI, but only within a small pressure interval.

    III. PRELIMINARIES

    A. Lattice structures and basic topological properties

    The assumed crystal structures of LaBiTe3 and LuBiTe3are very similar to Bi2Te3, where five atomic monolayersstack in the [111] direction in an A-B-C-A-... sequenceforming quintuple layers (QLs) as shown in Fig. 1(a). The onlydifference is that one of the two Bi atoms in the primitive unitcell is replaced by a La or Lu atom, which breaks the inversionsymmetry. The lattice structure of LaSbTe3 and LuSbTe3 isthe same as for LaBiTe3 and LuBiTe3, except that all the Biatoms are substituted by Sb. The in-plane hexagonal lattice pa-rameters for LaBiTe3 and LuBiTe3 are a=4.39 Å and 4.18 Å,respectively, while the size of a QL along c is 10.07 and 10.29Å, respectively. The lattice parameters of LaSbTe3 are slightlydifferent from LaBiTe3, with a = 4.24 Å and c = 10.13 Å. Thelattice parameters for LuSbTe3 have not been reported before,so we use those from LuBiTe3. Among these four hypotheticalmaterials, LaBiTe3 has been previously reported as a candidatefor an inversion-asymmetric TI [21]. LuBiTe3 is first reportedas a TI candidate in this paper; the nontrivial band topology isconfirmed by calculating the bulk Z2 index [23] and checkingthe existence of topological surface states. On the other hand,LaSbTe3 and LuSbTe3 are trivial insulators.

    As shown in Fig. 1(c), BiTeI has a hexagonal latticestructure with three atoms in the primitive cell stacked asA-B-C-A-... along the z direction. The lattice parameters inplane and along the hexagonal axis are a = 4.339 Å andc = 6.854 Å. BiTeI itself is a trivial insulator with a largeRashba spin splitting in the bulk [24], but it can be driveninto a TI state by applying pressure. Previous studies havesuggested that the transition to the topological phase is notmediated by a WSM phase [19,20], but we revisit this issue inSec. IV B and come to different conclusions.

    B. First-principles methodology

    We carry out the bulk first-principles calculations usingthe VASP package including SOC [25,26]. The generalized-gradient approximation is used to treat the exchange-correlation functional [27,28]. The BZ is sampled on an8×8×8 Monkhorst-Pack [29] k mesh and an energy cutoff of340 eV is used. The output from the first-principles plane-wavecalculations are then interfaced to the WANNIER90 package [30]to construct realistic tight-binding (TB) models for thesematerials [31].

    To describe the electronic structure of LaBi1−xSbxTe3 andLuBi1−xSbxTe3, we adopt the virtual crystal approximation(VCA) in which each Bi or Sb is replaced by a “virtual”atom whose properties are a weighted average of the twoconstituents. The VCA treatment typically gives a reasonabledescription for solid-solution systems in which the dopant and

    155316-4

  • WEYL SEMIMETALS FROM NONCENTROSYMMETRIC . . . PHYSICAL REVIEW B 90, 155316 (2014)

    host atoms have a similar chemical character. For example, theVCA was shown to work well in describing Sb substitutionin Bi2Se3, because of the similar orbital character of Sb5p and Bi 6p, but not for In substitution, where In 5sorbitals become involved [32]. The VCA is implemented inthe Wannier basis by constructing separate 36-band modelsfor LaBiTe3 (LuBiTe3) and LaSbTe3 (LuSbTe3), includingall the valence p orbitals of the cations and anions, as wellas the 5d and 6s orbitals of the rare-earth elements [33]. Inthe solid solution, the Hamiltonian matrix elements are thentaken as a linear interpolation in impurity composition x of thecorresponding matrix elements of the parent materials. Thatis, we take H VCAmn = (1 − x)H Bimn + xH Sbmn, where H Bimn and H Sbmndenote the matrix elements of the TB models of LaBiTe3 andLaSbTe3. It worth noting that when generating the Wannierfunctions (WFs) for the VCA treatment, the Wannier basisfunctions have to be chosen as similar as possible before theaveraging [32]. We therefore use WFs that are constructedsimply by projecting the Bloch states onto the same setof atomiclike trial orbitals without applying a subsequentmaximal-localization procedure [34,35].

    Similarly, to study the pressure-induced TPT in BiTeI,we carry out first-principles calculations for the system atthe zero-pressure volume, where it is topologically normal,and also at 85.4% of the original volume, a value chosensomewhat arbitrarily to be well inside the TI region [20].We denote these two states as η = 0 and η = 1, respectively.Then from the Wannier representation we again construct arealistic Hamiltonian for each system, denoted as H0 andH1, respectively, including all the valence p orbitals of Bi,Te, and I. Finally we linearly interpolate these as H (η)=(1 − η)H0 + ηH1, treating η as an adiabatic parameter thattunes the system through the topological phase transition.

    Using these Wannierized effective TB models, we cansearch for BTPs very efficiently over the entire BZ. We firstsample the irreducible BZ using a relatively sparse k mesh,e.g., 20×20×20, and find the point k0 having the smallestdirect band gap on this mesh. A second-round search isconducted by scanning over a denser k mesh within a spherecentered on k0. We then repeat the procedure iteratively untilconvergence is reached. All of the trajectories of Weyl nodespresented in Sec. IV are obtained using this approach.

    IV. RESULTS

    A. LaBi1−xSbxTe3 and LuBi1−xSbxTe31. Band gap and Weyl chirality

    For each of these materials we scan over a mesh incomposition x, and for each x we construct the WannierizedHamiltonian for the corresponding solid solution within theVCA. We then use the methods of the previous section tosearch for the BTPs in the entire irreducible BZ. Plots ofthe smallest direct band gap in the BZ vs x are presented inFig. 2. Clearly the gap remains closed over a finite range ofx in both cases, from 38.5% to 41.9% for LaBi1−xSbxTe3and 40.5%to 45.1% for LuBi1−xSbxTe3. By checking thedispersion around the gap-closure point, we confirm that thesystem is semimetallic with the Fermi level lying at a set ofdegenerate Weyl BTPs over this entire range.

    0

    2

    4

    6

    8

    10

    Gap

    (meV

    )

    0.38 0.40 0.42 0.44 0.460

    2

    4

    6

    8

    Gap

    (meV

    )

    (a)

    (b)

    Composition x

    FIG. 2. (Color online) Smallest direct band gap in the BZ vscomposition x for (a) LaBi1−xSbxTe3 and (b) LuBi1−xSbxTe3. Dashedgray line marks our chosen threshold of 0.2 meV to signal a gapclosure.

    To illustrate the topological character, we further calculatethe chirality of the BTPs, which is given by the determinantof the Jacobian matrix Jij = ∂fi/∂kj . Figure 3 shows howdet(J) varies with x for the BTPs in LaBi1−xSbxTe3 andLuBi1−xSbxTe3. The red and black open circles mark thevalues of det(J) for the BTPs with positive and negativechirality, which are mapped into each other by mirroroperations about the kx =0 and other equivalent mirror planes.One can see that at the beginning of the band touching, thechirality starts at zero, indicating the creation of a quadratic

    −0.5

    0

    0.5

    det (

    J)

    0.39 0.40 0.41 0.42 0.43 0.44 0.45

    −0.5

    0

    0.5

    det (

    J)

    Composition x

    (a)

    (b)

    FIG. 3. (Color online) Determinant of the Jacobian matrix eval-uated at Weyl nodes with positive (red) and negative (black) chiralityvs composition x, for (a) LaBi1−xSbxTe3 and (b) LuBi1−xSbxTe3.

    155316-5

  • JIANPENG LIU AND DAVID VANDERBILT PHYSICAL REVIEW B 90, 155316 (2014)

    BTP. As x increases, each quadratic BTP splits into two Weylnodes with opposite chirality. These then migrate through theBZ and eventually annihilate each other at the point where thechirality returns to zero.

    2. Symmetry considerations

    As mentioned earlier, the point group of this class ofmaterials is C3v , which has a threefold rotation axis along kzand three mirror planes that contain the kz axis and intersect thekz =0 plane on the lines kx =0 and ky =±kx/

    √3. We define

    an azimuthal angle θ that measures the rotation of (kx,ky) fromthe +ky axis in the clockwise direction as shown in Fig. 4(a).As a result of the threefold rotational symmetry, if a Weylnode with positive chirality appears at some θ in the region0� θ � π/3 and at some kz in the upper half BZ, then theremust be another two nodes with the same chirality and the samekz located at θ + 2π/3 and θ − 2π/3. Taking into account themirror symmetry, these must have negative-chirality partnersat the same kz but at −θ , −θ + 2π/3, and −θ − 2π/3.

    −0.05 0 0.05

    −0.05

    0

    0.05

    −0.05 0 0.05

    −0.05

    0

    0.05

    (a)

    (b)

    kx

    kx

    ky

    ky

    FIG. 4. (Color online) Trajectories of Weyl nodes in the (kx,ky)plane (in units of Å−1). Dashed red lines indicate Weyl nodes ofpositive chirality; solid black lines are negative. The “*” and “⊕”denote, respectively, the points of creation or annihilation of Weylnodes. (a) For LaBi1−xSbxTe3. (b) For LuBi1−xSbxTe3.

    Finally, because of TR symmetry, each Weyl node at k isalways accompanied by another at −k with the same chirality,giving six more Weyl nodes in the lower half BZ. We thusgenerically expect a total of 12 Weyl nodes in the entire BZfor compositions x in the region of the WSM phase.

    3. Weyl trajectories

    Figure 4 shows the trajectories of the Weyl nodes inLaBi1−xSbxTe3 and LuBi1−xSbxTe3 projected onto the (kx,ky)plane as x passes through the critical region. The red dashedline represents the trajectory of Weyl nodes with positivechirality, while the solid black one denotes those with negativechirality, and the “*” and “⊕” denote the creation andannihilation points of the Weyl nodes, respectively. As xincreases, six quadratic BTPs are simultaneously created in themirror planes; this occurs at xc1 =38.5% for LaBi1−xSbxTe3and 40.5% for LuBi1−xSbxTe3. Each quadratic BTP then splitsinto two Weyl nodes of opposite chirality, and these 12 nodesmigrate along the solid black and dashed red lines shown in thefigure. Eventually, after exchanging partners, the Weyl nodesmeet and annihilate each other in another set of high-symmetryplanes (ky =0 and other equivalent planes), at xc2 =41.9% forLaBi1−xSbxTe3 and 45.1% for LuBi1−xSbxTe3.

    Figures 5(a) and 5(b) show the trajectory of the Weyl nodesin the kz direction. At x =xc1, six quadratic BTPs are created,three in the top half-BZ and three in the bottom half-BZ, butall of them fairly close to the BZ boundary plane at kz =±π/c.As x increases, the six BTPs split to form 12 Weyl nodes, andthese begin to move toward the above-mentioned BZ boundaryplane. Finally, after interchanging partners, Weyl nodes ofopposite chirality annihilate in pairs at xc2 on the BZ boundary

    0.40

    0.42

    0.44

    −0.02

    −0.01

    0

    0.01

    0.02

    0.40

    0.44

    −0.02

    −0.01

    0

    0.01

    0.02(a) (b)

    (c) (d)

    FIG. 5. (Color online) (a), (b) Trajectories of Weyl nodes inthe kz direction (in units of Å−1) for (a) LaBi1−xSbxTe3 and(b) LuBi1−xSbxTe3. Dashed red (solid black) lines refer to the Weylnodes with positive (negative) chirality. θ is the azimuthal angle inthe (kx,ky) plane, as indicated in Fig. 4(a). The “*” and “⊕” denotethe creation and annihilation point of the Weyl nodes, respectively.(c), (d) Trajectories of Weyl nodes in the direction of impuritycomposition x for (c) LaBi1−xSbxTe3 and (d) LuBi1−xSbxTe3.

    155316-6

  • WEYL SEMIMETALS FROM NONCENTROSYMMETRIC . . . PHYSICAL REVIEW B 90, 155316 (2014)

    plane at kz = ±π/c. For x > xc2 a global gap opens up andthe system is again an insulator but with an inverted Z2 index.

    The locus of Weyl points can be regarded as forming a loopin the four-dimensional space of (kx,ky,kz,x), and just as thisloop can be projected onto kz as in Figs. 5(a) and 5(b), it canalso be projected onto the direction of impurity compositionx as shown in Figs. 5(c) and 5(d). Again, it is clear thatthe Weyl nodes are created at xc1 in the mirror planes andannihilated at xc2 at θ = ±π/6. These plots may also behelpful in seeing how the high sixfold symmetry contributesto the narrowness of the WSM region. If the symmetry of thesystem were lower, the period of oscillation in θ in Figs. 5(c)and 5(d) would be longer, which would allow the Weyl nodesto oscillate farther in the x direction, giving a wider windowof concentration for the WSM phase. In contrast, a fictitioussystem with an N -fold rotational symmetry would force thewidth of the WSM region to vanish as N → ∞. Here we haveN = 6, which is evidently large enough to limit the WSMphase to a rather small interval in x.

    4. Surface Fermi arcs

    One of the most characteristic features of WSMs is theexistence of Fermi arcs in the surface band structure. Here wecalculate the surface states using the surface Green’s-functiontechnique [36], which is implemented in the context of theVCA effective Hamiltonian in the Wannier basis. The surfaceBZ is sampled by a 64×64 k mesh, and the surface spectralfunctions calculated on this mesh are then linearly interpolatedto fit a 128×128 k mesh. Figure 6 shows the normalized

    −0.05

    0

    0.05

    0.2

    0.4

    0.6

    0.8

    1

    −0.05 0 0.05

    −0.05

    0

    0.05

    0.2

    0.4

    0.6

    0.8

    1

    (a)

    (b)

    kx

    ky

    ky

    FIG. 6. (Color online) Surface spectral function averaged aroundthe Fermi level (kx and ky in units of Å−1) for (a) LaBi1−xSbxTe3 atx = 0.405 and (b) LuBi1−xSbxTe3 at x = 0.43.

    surface spectral functions averaged around the Fermi levelfor LaBi1−xSbxTe3 at x = 0.405 and for LuBi1−xSbxTe3 atx = 0.43. The averaging is done over an energy window of±4.5 meV around the Fermi energy, which is determinedby the position of the bulk Weyl nodes. Six Fermi arcsconnecting the projected Weyl nodes of opposite chirality arevisible, confirming the existence of the WSM phase in thesetwo solid-solution systems. Note that because of the smallprojected bulk gap on the loops where the Fermi arcs reside,some non-negligible spectral weight is visible even outsidethe Fermi arcs in Fig. 6, coming from the artificial smearingof the Green’s functions.

    5. Simplified six-band model

    In order to capture the essential physics in these materials,we construct a six-band TB model to describe the interestingcritical behavior. From the band-structure plots presented inFig. 7, it is clear that the band inversion occurs around theZ point of LaBiTe3 and LuBiTe3, so we focus our attentionon the six states at Z closest to the Fermi level. A symmetryanalysis shows that these six states belong to two copies ofthe two-dimensional Z6 irreducible representation (irrep) ofthe C3v group at Z, plus a Kramers pair of one-dimensionalcomplex-conjugate Z4 and Z5 irreps corresponding to linearcombinations of jz = ±3/2 orbitals.

    We thus build our six-band TB model out of basis stateshaving the symmetry of |pz, ↑〉 and |pz, ↓〉 on the Teatoms at the top and bottom of the quintuple layer, and

    −2

    −1

    0

    1

    2

    Ene

    rgy(

    eV)

    −2

    −1

    0

    1

    2

    Ene

    rgy(

    eV)

    (a)

    (b)

    FIG. 7. (Color online) Bulk band structures of (a) LaBiTe3 and(b) LuBiTe3.

    155316-7

  • JIANPENG LIU AND DAVID VANDERBILT PHYSICAL REVIEW B 90, 155316 (2014)

    FIG. 8. (Color online) Top: Schematic diagram of the interlayerspin-independent hopping terms in the six-band model. Orbitals onsites Te1, Te2, and Te1′ make up a quintuple layer; A, B, and Clabel in-plane hexagonal positions. Bottom: Phase diagram for thetopological behavior of the six-band model.

    |px + ipy, ↑〉 and |px − ipy, ↓〉 combinations located on thecentral Te atoms. A schematic illustration of the six-bandmodel is shown in Fig. 8, where the top, bottom, and centralTe atoms are denoted by Te1, Te1′, and Te2, respectively.First of all, six interlayer spin-independent hopping terms areincluded in the model. As shown in Fig. 8, we consider thefirst-neighbor hopping between the central and top (bottom)Te atoms t1 (t2), the inter-QL (intra-QL) hopping between thetop and bottom Te atoms t3 (t4), and some further-neighborhoppings tu and tv that are crucial in obtaining a nontrivialZ2 index. Second, to capture the Rashba spin- splitting inthe first-principles band structure, in-plane Rashba-like spin-dependent hoppings within the top and bottom Te monolayersare included and are denoted by λ1 and λ2, respectively. Forcompleteness, the interlayer first-neighbor Te1-Te2 (λ3) andTe1′-Te2 (λ4) Rashba-like hopping terms are also included.Lastly, to reproduce the first-principles band structure better,we also introduce first-neighbor spin-independent hoppingterms within the Te1, Te2, and Te1′ monolayers, denoted by v1,v2, and v3, respectively. The on-site energies are also differentand are labeled by E1 for Te1, E2 for Te2, and E3 for Te1′. Asour model is only intended to be semiquantitative, we use thesame model parameters to describe both LaBiTe3 and LuBiTe3.

    We take all of the parameters in the model to depend on ascaling parameter δ that drives the TPT. When δ is zero, thesystem is a trivial insulator; as δ increases, the system becomesa topological insulator by going through a critical WSM. Thedependence of the parameters on δ defines a path in parameterspace. It is important to note that the width of the critical WSMregion can be highly sensitive to this path, with an improperchoice sometimes leading to an extremely narrow WSM phase.Our choice is specified in Table I.

    Following the path we have chosen, a WSM phase isobtained for 0.067 eV < δ < 0.074 eV. As shown in Fig. 9(b),the smallest direct band gap in the BZ vanishes when 0.067 eV< δ < 0.074 eV, indicating the existence of BTPs in BZ. If onefurther checks the position of the BTPs, one finds that whenδ ≈ 0.067 eV, six quadratic BTPs are created in the mirrorplanes, which then split into 12 Weyl nodes and propagatein the BZ following the solid black and dashed red lines in

    TABLE I. Parameters of the six-band model (in eV).

    t1 0.2 − δ/4 λ3 0.15 − δ/2 v3 0t2 0.15 − δ/4 λ4 0.12 − δ/2 E1 0.1 + δ − 6v1t3 δ tu 0.12 + δ/2 E2 −6v2t4 0.1 − δ/4 tv 0.06 − δ/2 E3 −0.1 − δλ1 0.24 − δ/2 v1 0.05λ2 0.2 − δ/2 v2 0.1

    Figs. 9(c) and 9(d). These Weyl nodes eventually annihilatewith each other at δ ≈ 0.074 eV after exchanging partners,which qualitatively reproduces the phase-transition behaviorof the VCA effective Hamiltonians very well. When δ > 0.074eV, the system becomes a strong TI. The bulk band structureat δ = 0.09 eV in the TI phase is shown in Fig. 9(a), whichvery well captures the low-energy dispersions around Z thatwere found in the first-principles calculations.

    6. Discussion

    To conclude this section, we would like to comment thatthe width of the WSM phase depends on two ingredients. Onone hand, as discussed above, it depends on the symmetry ofthe system; other things being equal, the WSM interval tends tobe wider in systems with lower symmetry. On the other hand,even for fixed symmetry, it also depends on the the detailedchoice of path connecting the topological and trivial phases.Choosing a different path may broaden or reduce the WSMregion. For example, if one artificially changes the strengthof the atomic SOC strength in LaBiTe3 and LuBiTe3 in theWannierized TB models, and scales the variation of the actualatomic SOC by a single scaling parameter λ, then we find

    −1

    0

    1

    Ene

    rgy

    (eV

    )

    0.067 0.070 0.0730

    2

    4

    6

    Gap

    (m

    eV)

    −0.1 −0.05 0 0.05 0.1−0.1

    −0.05

    0

    0.05

    0.1

    −0.05

    0

    0.05

    (a) (b)

    (c) (d)

    FIG. 9. (Color online) (a) Bulk bandstructure of the six-bandmodel at δ=0.09 eV. (b) Smallest direct band gap in the BZ vs. δ.(c) Trajectory of Weyl nodes projected onto the (kx,ky) plane. Dashedred (solid black) line refers to the Weyl node with positive (negative)chirality. The “*” and “⊕” denote the creation and annihilation pointof the Weyl nodes respectively. θ is the azimuthal angle in the (kx,ky)plane. (d) Trajectory of Weyl nodes along kz. Units of kx , ky and kzare Å−1.

    155316-8

  • WEYL SEMIMETALS FROM NONCENTROSYMMETRIC . . . PHYSICAL REVIEW B 90, 155316 (2014)

    that the WSM region only shows up for λ in the range of76.8%–77.3%, which is significantly narrower than for theVCA case. However, if an average SOC is applied to the entiresystem, such that the SOC strength on Te is artificially highand that on Bi is artificially low, we find that a much widerWSM region results. Thus, it may potentially be possible toengineer the width of a critical WSM phase if one can modifythe transformation path, as by epitaxial strain, pressure, oradditional chemical?brk?> substitution.

    B. BiTeI: Revisited

    In order to justify the discussion in Sec. II B, we revisitthe TPT in BiTeI driven by pressure. In our calculations, thepressure is applied by compressing the volume of the primitivecell. The fully compressed volume V is taken to be 85.4% ofthe original volume V0, such that the former is well insidethe topological region [20], and both the lattice vectors andatomic positions are relaxed at the compressed volume. Asdiscussed in Sec. III B, we searched for BTPs over the entireirreducible BZ for a transitional Hamiltonian scaled as H (η)=(1 − η)H0 + ηH1 for 0�η�1, where H0 and H1 representthe Hamiltonians of the uncompressed and fully compressedBiTeI, with even and odd Z2 indices, respectively. As shownin Fig. 10(a), as the pressure is increased from 0% to 100%(alternatively, as V is decreased from 100% to 85.4% of V0),a semimetallic phase emerges for η in the range of about54%–56%.

    0.53 0.54 0.55 0.56 0.570

    1

    2

    3

    4

    5

    6

    7

    8

    Gap

    (m

    eV)

    −0.05 0 0.05

    −0.05

    0

    0.05

    0

    0.2

    0.4

    0.6

    0.8

    1

    (a)

    (b)

    FIG. 10. (Color online) (a) Smallest direct band gap in the BZ ofBiTeI vs the pressure-scaling variable η. (b) Surface spectral functionof BiTeI in the WSM phase (at 55% of the full pressure).

    −0.04 0 0.04

    −0.04

    0

    0.04

    −0.015

    0

    0.015(a) (b)

    FIG. 11. (Color online) (a) Trajectories of Weyl nodes in the(kx,ky) plane (in units of Å−1). Dashed red (solid black) lines indicatethe trajectories of Weyl nodes with positive (negative) chirality. The“*” and “⊕” denote the creation and annihilation point of the Weylnodes, respectively. (b) Trajectory of Weyl nodes in the kz direction(units of Å−1).

    The point group of BiTeI is the same as for LaBiTe3 andLuBiTe3, namely, C3v . Therefore, as explained in Sec. IV A,one would expect the emergence of 12 Weyl nodes in theentire BZ during the phase-transition process. The trajectoriesof the Weyl nodes are plotted in Figs. 11(a) and 11(b).When η≈54%, six quadratic BTPs are first created at theBZ boundary kz = π/c in the ky = 0 and other equivalenthigh-symmetry planes. These BTPs then split into 12 Weylnodes which propagate along the directions indicated by solidblack (antimonopoles) and dashed red (monopoles) lines.They annihilate each other in the three mirror planes afterexchanging partners. Note that in this case the system goesfrom a normal to topological insulator as η increases, which isthe reverse of the LaBi1−xSbxTe3 and LuBi1−xSbxTe3 cases.

    The results shown in Fig. 11 support our conclusions inSec. II B. In particular, even though the torsion argumentimplies that the trajectories of the two Weyl nodes which splitoff from a given quadratic BTP would never meet each other,a closed curve is still formed in the 3D BZ of BiTeI throughthe interchange of partners among the Weyl nodes.

    Figure 10(b) shows the surface spectral function of BiTeIaveraged around the Fermi level for η=0.55, in the WSMphase. It is clear that there are six Fermi arcs extending betweenthe six pairs of projected Weyl nodes, which is again thehallmark of a WSM phase.

    We therefore conclude that a WSM phase does exist in theTPT of BiTeI, but it occurs only within a narrow pressure range.If η is changed by 2.5%, the volume is only changed by 0.39%,which might be difficult to measure experimentally. Again, thenarrowness of the WSM interval can be attributed in part tothe high symmetry of the system. However, as emphasized inthe previous section, the width of the critical WSM is alsosensitive to the choice of path in parameter space. The criticalWSM could get broadened by choosing a different path, asfor example by applying uniaxial pressure. We leave this for afuture study.

    V. SUMMARY

    In this paper, we have investigated the nature of the TPTin a noncentrosymmetric TI in the most general case. We find

    155316-9

  • JIANPENG LIU AND DAVID VANDERBILT PHYSICAL REVIEW B 90, 155316 (2014)

    that an intermediate WSM phase is always present, regardlessof other lattice symmetries, as long as inversion symmetryis absent. We discussed separately the cases in which theJacobian matrix is rank one or rank two when the gap firstcloses. In the rank-two case, each quadratic BTP would alwayssplit into a pair of Weyl nodes, which annihilate each other afterexchanging partners. If the rank of the Jacobian is one, thenthe doubly quadratic BTP in this case would either split intofour Weyl nodes, or else immediately be gapped out again,corresponding to an “insulator-insulator transition.” However,in the latter case, the bulk Z2 indices are not expected tochange. Therefore, we conclude that Z2-even and Z2-oddphases of a noncentrosymmetric insulator must always beseparated by a region of WSM phase, even if other symmetriesare present.

    To illustrate our conclusions, we have carried out cal-culations on specific noncentrosymmetric insulators. ForLaBi1−xSbxTe3 and LuBi1−xSbxTe3 we have used Wannier-ized VCA Hamiltonians to find a WSM phase in the regionx ≈ 38.5%–41.9% and x ≈ 40.5%–45.1%, respectively. Asix-band TB model was also constructed to describe thetopological and critical behavior in these materials. We foundthat the width of the critical WSM phase can be highly sensitiveto the choice of path in the parameter space, suggesting thatthere is flexibility to engineer the WSM phase.

    We have also revisited the TPT of BiTeI as a function ofpressure, where previous work suggested the absence of aWSM phase [19]. Using a carefully constructed algorithm tosearch for the minimum gap in the full three-dimensional BZ,we found that a WSM phase is indeed present over a narrowinterval of pressure, although this range may be so narrow asto make its experimental observation difficult.

    In summary, we have clarified the theory of a generalZ2-even to Z2-odd topological phase transition in a three-dimensional time-reversal-invariant insulator with brokeninversion symmetry, and demonstrated that an intermediateWSM phase must always be present. We have also detailed thebehavior of LaBi1−xSbxTe3 and LuBi1−xSbxTe3 as promisingcandidates for WSMs of this kind. While we have notconsidered disorder or interactions explicitly, we expect ourconclusions to survive at least for weak disorder or interac-tions. Our work is a step forward in the general understandingof topological phase transitions, and may provide usefulguidelines for the experimental realization of new classes ofWeyl semimentals.

    ACKNOWLEDGMENT

    This work is supported by NSF Grants No. DMR-10-05838and No. 14-08838.

    [1] K. v. Klitzing, G. Dorda, and M. Pepper, Phys. Rev. Lett. 45,494 (1980).

    [2] D. J. Thouless, M. Kohmoto, M. P. Nightingale, and M.den Nijs, Phys. Rev. Lett. 49, 405 (1982).

    [3] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045(2010).

    [4] X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057 (2011).[5] H. B. Nielsen and M. Ninomiya, Phys. Lett. B 130, 389 (1983).[6] A. A. Burkov, M. D. Hook, and L. Balents, Phys. Rev. B 84,

    235126 (2011).[7] X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov,

    Phys. Rev. B 83, 205101 (2011).[8] A. M. Turner and A. Vishwanath, in Topological Insulators,

    edited by M. Franz and L. Molenkamp (Elsevier, Amsterdam,2013), Chap. 11, pp. 293–322.

    [9] P. Hosur and X. Qi, C. R. Phys. 14, 857 (2013).[10] R. Okugawa and S. Murakami, Phys. Rev. B 89, 235315 (2014).[11] K.-Y. Yang, Y.-M. Lu, and Y. Ran, Phys. Rev. B 84, 075129

    (2011).[12] A. A. Burkov and L. Balents, Phys. Rev. Lett. 107, 127205

    (2011).[13] F. D. M. Haldane, arXiv:1401.0529.[14] A. A. Zyuzin and A. A. Burkov, Phys. Rev. B 86, 115133 (2012).[15] V. Aji, Phys. Rev. B 85, 241101 (2012).[16] D. Bulmash, C.-X. Liu, and X.-L. Qi, Phys. Rev. B 89, 081106

    (2014).[17] S. Murakami and S.-i. Kuga, Phys. Rev. B 78, 165313 (2008).[18] S. Murakami, New J. Phys. 9, 356 (2007).[19] B.-J. Yang, M. S. Bahramy, R. Arita, H. Isobe, E.-G. Moon, and

    N. Nagaosa, Phys. Rev. Lett. 110, 086402 (2013).[20] M. Bahramy, B.-J. Yang, R. Arita, and N. Nagaosa,

    Nat. Commun. 3, 679 (2012).

    [21] B. Yan, H.-J. Zhang, C.-X. Liu, X.-L. Qi, T. Frauenheim, andS.-C. Zhang, Phys. Rev. B 82, 161108 (2010).

    [22] X. Xi, C. Ma, Z. Liu, Z. Chen, W. Ku, H. Berger, C. Martin, D. B.Tanner, and G. L. Carr, Phys. Rev. Lett. 111, 155701 (2013).

    [23] A. A. Soluyanov and D. Vanderbilt, Phys. Rev. B 83, 235401(2011).

    [24] K. Ishizaka, M. Bahramy, H. Murakawa, M. Sakano,T. Shimojima, T. Sonobe, K. Koizumi, S. Shin, H. Miyahara, A.Kimura, K. Miyamoto, T. Okuda, H. Namatame, M. Taniguchi,R. Arita, N. Nagaosa, Y. Kobayashi, K Murakami, R. Kumai,Y. Kaneko, Y. Onose, and Y. Tokura, Nat. Mater. 10, 521 (2011).

    [25] G. Kresse and J. Furthmüller, Phys. Rev. B 54, 11169 (1996).[26] G. Kresse and J. Furthmüller, Comput. Mater. Sci. 6, 15 (1996).[27] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 77,

    3865 (1996).[28] J. P. Perdew, K. Burke, and M. Ernzerhof, Phys. Rev. Lett. 78,

    1396 (1997).[29] H. J. Monkhorst and J. D. Pack, Phys. Rev. B 13, 5188 (1976).[30] A. A. Mostofi, J. R. Yates, Y. S. Lee, I. Souza, D. Vanderbilt,

    and N. Marzari, Comput. Phys. Commun. 178, 685 (2008).[31] The TB models from WANNIER90 are realistic in the sense that

    the Wannier-interpolated band structure can reproduce the first-principles band energies exactly in a specified energy windowwhich is chosen here to be centered around the Fermi level.

    [32] J. Liu and D. Vanderbilt, Phys. Rev. B 88, 224202 (2013).[33] The fully occupied f shell of Lu has very little influence on the

    electronic structure around the Fermi level.[34] N. Marzari and D. Vanderbilt, Phys. Rev. B 56, 12847 (1997).[35] I. Souza, N. Marzari, and D. Vanderbilt, Phys. Rev. B 65, 035109

    (2001).[36] M. L. Sancho, J. L. Sancho, J. L. Sancho, and J. Rubio, J. Phys.

    F 15, 851 (1985).

    155316-10

    http://dx.doi.org/10.1103/PhysRevLett.45.494http://dx.doi.org/10.1103/PhysRevLett.45.494http://dx.doi.org/10.1103/PhysRevLett.45.494http://dx.doi.org/10.1103/PhysRevLett.45.494http://dx.doi.org/10.1103/PhysRevLett.49.405http://dx.doi.org/10.1103/PhysRevLett.49.405http://dx.doi.org/10.1103/PhysRevLett.49.405http://dx.doi.org/10.1103/PhysRevLett.49.405http://dx.doi.org/10.1103/RevModPhys.82.3045http://dx.doi.org/10.1103/RevModPhys.82.3045http://dx.doi.org/10.1103/RevModPhys.82.3045http://dx.doi.org/10.1103/RevModPhys.82.3045http://dx.doi.org/10.1103/RevModPhys.83.1057http://dx.doi.org/10.1103/RevModPhys.83.1057http://dx.doi.org/10.1103/RevModPhys.83.1057http://dx.doi.org/10.1103/RevModPhys.83.1057http://dx.doi.org/10.1016/0370-2693(83)91529-0http://dx.doi.org/10.1016/0370-2693(83)91529-0http://dx.doi.org/10.1016/0370-2693(83)91529-0http://dx.doi.org/10.1016/0370-2693(83)91529-0http://dx.doi.org/10.1103/PhysRevB.84.235126http://dx.doi.org/10.1103/PhysRevB.84.235126http://dx.doi.org/10.1103/PhysRevB.84.235126http://dx.doi.org/10.1103/PhysRevB.84.235126http://dx.doi.org/10.1103/PhysRevB.83.205101http://dx.doi.org/10.1103/PhysRevB.83.205101http://dx.doi.org/10.1103/PhysRevB.83.205101http://dx.doi.org/10.1103/PhysRevB.83.205101http://dx.doi.org/10.1016/j.crhy.2013.10.010http://dx.doi.org/10.1016/j.crhy.2013.10.010http://dx.doi.org/10.1016/j.crhy.2013.10.010http://dx.doi.org/10.1016/j.crhy.2013.10.010http://dx.doi.org/10.1103/PhysRevB.89.235315http://dx.doi.org/10.1103/PhysRevB.89.235315http://dx.doi.org/10.1103/PhysRevB.89.235315http://dx.doi.org/10.1103/PhysRevB.89.235315http://dx.doi.org/10.1103/PhysRevB.84.075129http://dx.doi.org/10.1103/PhysRevB.84.075129http://dx.doi.org/10.1103/PhysRevB.84.075129http://dx.doi.org/10.1103/PhysRevB.84.075129http://dx.doi.org/10.1103/PhysRevLett.107.127205http://dx.doi.org/10.1103/PhysRevLett.107.127205http://dx.doi.org/10.1103/PhysRevLett.107.127205http://dx.doi.org/10.1103/PhysRevLett.107.127205http://arxiv.org/abs/arXiv:1401.0529http://dx.doi.org/10.1103/PhysRevB.86.115133http://dx.doi.org/10.1103/PhysRevB.86.115133http://dx.doi.org/10.1103/PhysRevB.86.115133http://dx.doi.org/10.1103/PhysRevB.86.115133http://dx.doi.org/10.1103/PhysRevB.85.241101http://dx.doi.org/10.1103/PhysRevB.85.241101http://dx.doi.org/10.1103/PhysRevB.85.241101http://dx.doi.org/10.1103/PhysRevB.85.241101http://dx.doi.org/10.1103/PhysRevB.89.081106http://dx.doi.org/10.1103/PhysRevB.89.081106http://dx.doi.org/10.1103/PhysRevB.89.081106http://dx.doi.org/10.1103/PhysRevB.89.081106http://dx.doi.org/10.1103/PhysRevB.78.165313http://dx.doi.org/10.1103/PhysRevB.78.165313http://dx.doi.org/10.1103/PhysRevB.78.165313http://dx.doi.org/10.1103/PhysRevB.78.165313http://dx.doi.org/10.1088/1367-2630/9/9/356http://dx.doi.org/10.1088/1367-2630/9/9/356http://dx.doi.org/10.1088/1367-2630/9/9/356http://dx.doi.org/10.1088/1367-2630/9/9/356http://dx.doi.org/10.1103/PhysRevLett.110.086402http://dx.doi.org/10.1103/PhysRevLett.110.086402http://dx.doi.org/10.1103/PhysRevLett.110.086402http://dx.doi.org/10.1103/PhysRevLett.110.086402http://dx.doi.org/10.1038/ncomms1679http://dx.doi.org/10.1038/ncomms1679http://dx.doi.org/10.1038/ncomms1679http://dx.doi.org/10.1038/ncomms1679http://dx.doi.org/10.1103/PhysRevB.82.161108http://dx.doi.org/10.1103/PhysRevB.82.161108http://dx.doi.org/10.1103/PhysRevB.82.161108http://dx.doi.org/10.1103/PhysRevB.82.161108http://dx.doi.org/10.1103/PhysRevLett.111.155701http://dx.doi.org/10.1103/PhysRevLett.111.155701http://dx.doi.org/10.1103/PhysRevLett.111.155701http://dx.doi.org/10.1103/PhysRevLett.111.155701http://dx.doi.org/10.1103/PhysRevB.83.235401http://dx.doi.org/10.1103/PhysRevB.83.235401http://dx.doi.org/10.1103/PhysRevB.83.235401http://dx.doi.org/10.1103/PhysRevB.83.235401http://dx.doi.org/10.1038/nmat3051http://dx.doi.org/10.1038/nmat3051http://dx.doi.org/10.1038/nmat3051http://dx.doi.org/10.1038/nmat3051http://dx.doi.org/10.1103/PhysRevB.54.11169http://dx.doi.org/10.1103/PhysRevB.54.11169http://dx.doi.org/10.1103/PhysRevB.54.11169http://dx.doi.org/10.1103/PhysRevB.54.11169http://dx.doi.org/10.1016/0927-0256(96)00008-0http://dx.doi.org/10.1016/0927-0256(96)00008-0http://dx.doi.org/10.1016/0927-0256(96)00008-0http://dx.doi.org/10.1016/0927-0256(96)00008-0http://dx.doi.org/10.1103/PhysRevLett.77.3865http://dx.doi.org/10.1103/PhysRevLett.77.3865http://dx.doi.org/10.1103/PhysRevLett.77.3865http://dx.doi.org/10.1103/PhysRevLett.77.3865http://dx.doi.org/10.1103/PhysRevLett.78.1396http://dx.doi.org/10.1103/PhysRevLett.78.1396http://dx.doi.org/10.1103/PhysRevLett.78.1396http://dx.doi.org/10.1103/PhysRevLett.78.1396http://dx.doi.org/10.1103/PhysRevB.13.5188http://dx.doi.org/10.1103/PhysRevB.13.5188http://dx.doi.org/10.1103/PhysRevB.13.5188http://dx.doi.org/10.1103/PhysRevB.13.5188http://dx.doi.org/10.1016/j.cpc.2007.11.016http://dx.doi.org/10.1016/j.cpc.2007.11.016http://dx.doi.org/10.1016/j.cpc.2007.11.016http://dx.doi.org/10.1016/j.cpc.2007.11.016http://dx.doi.org/10.1103/PhysRevB.88.224202http://dx.doi.org/10.1103/PhysRevB.88.224202http://dx.doi.org/10.1103/PhysRevB.88.224202http://dx.doi.org/10.1103/PhysRevB.88.224202http://dx.doi.org/10.1103/PhysRevB.56.12847http://dx.doi.org/10.1103/PhysRevB.56.12847http://dx.doi.org/10.1103/PhysRevB.56.12847http://dx.doi.org/10.1103/PhysRevB.56.12847http://dx.doi.org/10.1103/PhysRevB.65.035109http://dx.doi.org/10.1103/PhysRevB.65.035109http://dx.doi.org/10.1103/PhysRevB.65.035109http://dx.doi.org/10.1103/PhysRevB.65.035109http://dx.doi.org/10.1088/0305-4608/15/4/009http://dx.doi.org/10.1088/0305-4608/15/4/009http://dx.doi.org/10.1088/0305-4608/15/4/009http://dx.doi.org/10.1088/0305-4608/15/4/009