Paper The following article is Open access

Hidden Weyl points in centrosymmetric paramagnetic metals

, , and

Published 2 March 2017 © 2017 IOP Publishing Ltd and Deutsche Physikalische Gesellschaft
, , Focus on Topological Semimetals Citation Dominik Gresch et al 2017 New J. Phys. 19 035001 DOI 10.1088/1367-2630/aa5de7

1367-2630/19/3/035001

Abstract

The transition metal dipnictides TaAs2, TaSb2, NbAs2 and NbSb2 have recently sparked interest for exhibiting giant magnetoresistance. While the exact nature of the magnetoresistance in these materials is still under active investigation, there are experimental results indicating that it is of the anisotropic negative variety. We study the effect of magnetic fields on the band structure topology of these materials by applying Zeeman splitting. In the absence of a magnetic field, we find that the materials are weak topological insulators, which is in agreement with previous studies. When the magnetic field is applied, we find that type-II Weyl points form. This result is found first from a symmetry argument, and then numerically for a ${\bf{k}}\cdot {\bf{p}}$ model of TaAs2 and a tight-binding model of NbSb2. This effect could be of help in the search for an explanation of the anomalous magnetoresistance in these materials.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Weyl nodes are the point-like crossings of two energy bands with linear dispersion. Locally, they can be described by a Hamiltonian of the form

Equation (1)

where $i\in \{x,y,z\}$ and $j\in \{0,x,y,z\}$. Topologically, a Weyl node can be characterized as being a quantized source or sink of Berry curvature, depending on its chirality [1]. Due to their quantized nature, Weyl points can only be created or annihilated in pairs of opposite chirality.

It was recently shown [2] that Weyl nodes come in two types. Type-I Weyl fermions have a point-like Fermi surface. When a magnetic field is applied, they exhibit a chiral Landau level [35] regardless of the magnetic field direction. When this chiral Landau level crosses the Fermi level, it can be a source of reduced magnetoresistance [611]. Type-II Weyl fermions, on the other hand, have an energy spectrum that is tilted by a strong ${\sigma }_{0}$ contribution to the Hamiltonian (equation (1)). As a consequence, the Fermi surface opens, and the chiral anomaly is anisotropic, appearing only for certain magnetic field directions.

The inversion symmetry P maps a Weyl node at point ${\bf{k}}$ onto a Weyl node of opposite chirality at $-{\bf{k}}$. Similarly, the time-reversal symmetry ${ \mathcal T }$ maps a Weyl point at ${\bf{k}}$ onto one at $-{\bf{k}}$, but without changing its chirality. Consequently, in the presence of the product symmetry ${ \mathcal T }\ast P$, Weyl nodes are mapped onto themselves but with opposite chirality. This four-fold degenerate crossing, consisting of two superimposed Weyl points of opposite chirality, is known as a Dirac node. Unlike Weyl points, they are not protected from gapping by any quantized topological charge. Consequently, additional symmetries are needed to stabilize Dirac nodes.

In centrosymmetric non-magnetic materials, the presence of both inversion and time-reversal symmetry only allows for Dirac nodes to form. Weyl nodes are not possible unless the product symmetry ${ \mathcal T }\ast P$ is broken.

Recently, transition metal dipnictides of the type AB2 (A ∈{Ta, Nb}, B ∈{As, Sb}) have gained a lot of attention [1219] for their giant magnetoresistance. These materials are semimetals, but without a direct closure of the band gap. Consequently, they do not host any Weyl or Dirac points.

The exact nature of magnetoresistance in these materials—especially the dependency on the direction of the magnetic field—is still under active investigation. Negative magnetoresistance has been observed experimentally for NbAs2 [17, 18], TaAs2 [17] and TaSb2 [17, 20]. Anomalous, albeit not negative, magnetoresistance has been observed for NbSb2 [12] and TaAs2 [15]. However, there are also experiments which point to the opposite, which is that there is no negative magnetoresistance in these materials. In [16], negative magnetoresistance was observed at first, but was then determined to be an artifact of the measurement setup.

In the following, we propose a mechanism for Weyl nodes to appear in these materials under the influence of a magnetic field. The chiral anomaly associated with these Weyl nodes is a possible source of negative magnetoresistance. Such an appearance of Weyl points under a magnetic field has recently been proposed in [21]. The mechanism with which the Weyl points appear, however, is a different one: in this work, the Weyl points appear from a previously gapped state, while [21] discusses Weyl points arising from the splitting of a four-fold crossing.

The paper is structured as follows: in the first section, the atomic and electronic structure of the four compounds is described. A four-band Hamiltonian for TaAs2 is derived from the symmetry considerations and fitted to the band structure. In the second section, the topology of the band structure is studied, first without a magnetic field and then by applying a Zeeman term. We find that this leads to the appearance of Weyl points.

2. Atomic and electronic structure of AB2 compounds

2.1. Atomic structure

In the following, the atomic structure of TaAs2 [22], TaSb2 [23], NbSb2 [24] and NbAs2 [25] is described.

The reduced unit cell of AB2 compounds has the general form

Equation (2)

with the parameters as given in table 1 [22, 24].

Table 1.  Unit cell dimensions (in Å) for AB2 compounds.

  ${\rm{a}}$ ${\rm{b}}$ ${\rm{c}}$ ${\rm{d}}$
TaAs2   4.6655 1.6915 3.8420 6.7330
TaSb2 5.11 1.822 4.1950 7.1502
NbAs2 4.684 1.698 3.8309 6.7933
NbSb2 5.1198 1.8159 4.1705 7.2134

Each unit cell contains two formula units. The atoms are located at the general Wyckoff positions $(x,-x,y)\,\,{\rm{a}}{\rm{n}}{\rm{d}}\,(-x,x,-y)$ for $(x,y)$, as shown in table 2 [22, 24].

Table 2.  Atomic positions $(x,y)$.

  A B1 B2
TaAs2   $(0.157,0.1959)$ $(0.4054,0.1082)$ $(0.1389,0.5265)$
TaSb2 $(0.152,0.19)$ $(0.405,0.113)$ $(0.147,0.535)$
NbAs2 $(0.1574,0.1965)$ $(0.4059,0.1084)$ $(0.14,0.528)$
NbSb2 $(0.1521,0.1903)$ $(0.4051,0.1127)$ $(0.1475,0.5346)$

Figure 1 shows the reduced unit cell and first BZ of TaAs2. The k-point path along which the bandstructure calculations are performed is indicated. On the basis reciprocal to that of equation (2), the special k-points are given by

Equation (3)

Figure 1.

Figure 1. (a) The reduced unit cell of TaAs2. (b) The first BZ of TaAs2. The k-point path and its projection onto the 010 surface are indicated.

Standard image High-resolution image

2.2. Electronic structure

The electronic structure calculations were performed in VASP [26], with the projector augmented-wave (PAW) [27, 28] pseudopotentials. The PBE approximation [29] was used, and spin–orbit coupling was included in the potentials. The self-consistent field (SCF) calculations were performed on a $11\times 11\times 5$ Γ-centered grid for TaAs2, and a $10\times 10\times 5$ Γ-centered grid for NbSb2. The energy cut-off given in the potential files was used, which is $293.2\,\mathrm{eV}$ for NbAs2 and NbSb2, and $223.7\,\mathrm{eV}$ for TaAs2 and TaSb2.

Additionally, the PBE calculations were tested against the accurate HSE06 hybrid functional [30, 31]. The hybrid SCF calculations for the band structures were performed on a Γ-centered $6\times 6\times 4$ grid for all materials. For the generation of the Wannier tight-binding model of NbSb2, a Γ-centered $10\times 10\times 5$ grid was used.

The band structure of TaAs2 and NbSb2 is shown in figure 2. Both materials exhibit a pair of electron and hole pockets near the M-point, where the minimum band gap is about $318\,\mathrm{meV}$ ($120\,\mathrm{meV}$ without hybrid functionals) in the case of TaAs2, $151\,\mathrm{meV}$ ($98\,\mathrm{meV}$) for TaSb2, $261\,\mathrm{meV}$ ($22\,\mathrm{meV}$) for NbAs2, and $67\,\mathrm{meV}$ ($18\,\mathrm{meV}$) in the case of NbSb2. A more complete calculation of the band structure can be found, for example, in [32].

Figure 2.

Figure 2. The band structures of AB2 compounds. The inset shows electron and hole pockets around M. The orange line represents the calculations using hybrid functionals. (a and b) TaAs2, (c and d) TaSb2, (e and f) NbAs2, (g and h) NbSb2.

Standard image High-resolution image

2.3. Symmetry operations and the ${\bf{k}}\cdot {\bf{p}}$ model

The AB2 compounds studied here have $C2/m$ symmetry (space group 12). The rotation axis is along the Cartesian y-axis. In reduced coordinates, the symmetry matrices are as follows:

  • Identity $E=\,{{\mathbb{I}}}_{3\times 3}$
  • Rotation ${C}_{2y}=\left(\begin{array}{ccc}0 & 1 & 0\\ 1 & 0 & 0\\ 0 & 0 & -\,1\end{array}\right)$
  • Parity $P=\,-{{\mathbb{I}}}_{3\times 3}$
  • Mirror ${M}_{y}={{PC}}_{2y}=\left(\begin{array}{ccc}0 & -\,1 & 0\\ -\,1 & 0 & 0\\ 0 & 0 & 1\end{array}\right).$

From the first-principles wave-functions, the representations corresponding to the two highest valence and two lowest conduction bands at the M-point were determined using the WIEN2k code [34, 35]. They were found to be ${{\rm{\Gamma }}}_{3}^{+},{{\rm{\Gamma }}}_{4}^{+}$ and ${{\rm{\Gamma }}}_{3}^{-},{{\rm{\Gamma }}}_{4}^{-}$, respectively. Their characters are shown in table 3, which comes from table 15 on page 35 in Koster et al [33]. Consequently, the symmetry representations in these four bands are given by

  • $\mathrm{Identity}\,E=\,{{\mathbb{I}}}_{4\times 4}$
  • Rotation ${C}_{2y}=\left(\begin{array}{cccc}i & 0 & 0 & 0\\ 0 & -\,i & 0 & 0\\ 0 & 0 & i & 0\\ 0 & 0 & 0 & -\,i\end{array}\right)$
  • Parity $P=\left(\begin{array}{cccc}1 & 0 & 0 & 0\\ 0 & 1 & 0 & 0\\ 0 & 0 & -\,1 & 0\\ 0 & 0 & 0 & -\,1\end{array}\right)$
  • Mirror ${M}_{y}={{PC}}_{2y}=\left(\begin{array}{cccc}i & 0 & 0 & 0\\ 0 & -\,i & 0 & 0\\ 0 & 0 & -\,i & 0\\ 0 & 0 & 0 & i\end{array}\right)$
  • Time-reversal ${ \mathcal T }=\left(\begin{array}{cccc}0 & -\,1 & 0 & 0\\ 1 & 0 & 0 & 0\\ 0 & 0 & 0 & -\,1\\ 0 & 0 & 1 & 0\end{array}\right)\hat{K}.$

For each of the symmetry operations g, the constraint

Equation (4)

is imposed on the 4 × 4 Hamiltonian, where D(g) is the symmetry representation. By applying these constraints on the general form of a four-band Hamiltonian

Equation (5)

we find the Hamiltonian to be of the form

Equation (6)

where the ${C}_{{ij}}({\bf{k}})$ are given up to the second order in ${{\bf{k}}}^{* }={\bf{k}}-M$ (in reduced coordinates) by

Equation (7)

Equation (8)

Equation (9)

Equation (10)

Equation (11)

Equation (12)

Table 3.  Character table for the relevant double group representations of C2m [33].

  E C2y P  My
${{\boldsymbol{\Gamma }}}_{3}^{+}$ 1 $\,i$ $\,1$ i
${{\boldsymbol{\Gamma }}}_{4}^{+}$ 1 $-i$ $\,1$ $-i$
${{\boldsymbol{\Gamma }}}_{3}^{-}$ 1 $\,i$ −1 $-i$
${{\boldsymbol{\Gamma }}}_{4}^{-}$ 1 $-i$ −1 i

These 16 parameters were numerically fitted to the band structure of TaAs2 using the SciPy [36] package to obtain the values in table 4. The resulting band structure around the M-point is shown in figure 3. Comparing it to the band structure obtained from first-principles reveals that the approximation is accurate in the immediate vicinity of the M-point, but breaks down at around $6 \% $ of the distance along the line MA. Importantly, the minimum band gap is not preserved in this model. Nevertheless, the model can be used to qualitatively study effects in TaAs2, owing to the fact that it contains the correct symmetry representations.

Figure 3.

Figure 3. The TaAs2 band structure of the ${\bf{k}}\cdot {\bf{p}}$ model (thick orange line), compared to the first-principles result (black lines).

Standard image High-resolution image

Table 4.  Parameters of the 4 × 4 Hamiltonian of TaAs2 around M up to the second order.

$[\mathrm{eV}]$ ${C}_{00}^{1}$ $=\,7.066$ ${C}_{z0}^{1}$ $=\,-0.224$
$[\mathrm{eV}\mathring{\rm A} ]$ ${C}_{{xz}}^{x+y}$ $=\,1.272$ ${C}_{y0}^{x+y}$ $=\,1.270$
  ${C}_{{xx}}^{x-y}$ $=\,-0.061$ ${C}_{{xy}}^{x-y}$ $=\,-1.999$
  ${C}_{{xx}}^{z}$ $=\,-0.554$     ${C}_{{xy}}^{z}$ $=\,-0.253$
$[\mathrm{eV}{\mathring{\rm A} }^{2}]$   ${C}_{00}^{{x}^{2}+{y}^{2}}$ $=\,-71.21$ ${C}_{z0}^{{x}^{2}+{y}^{2}}$ $=\,56.30$
  ${C}_{00}^{{xy}}$ $=\,-137.1$ ${C}_{z0}^{{xy}}$ $=\,123.1$
  ${C}_{00}^{{xz}-{yz}}$ $=\,1.52$ ${C}_{z0}^{{xz}-{yz}}$ $=\,-1.49$
  ${C}_{00}^{{z}^{2}}$ $=\,-0.84$ ${C}_{z0}^{{z}^{2}}$ $=\,-1.88$

3. Band structure topology

In this section, we describe the band structure topology and the influence of the magnetic field. First, we describe the topology in the absence of the magnetic field for all four compounds. Then, we show that Weyl points appear under a sufficient magnetic field. This result is shown first for the ${\bf{k}}\cdot {\bf{p}}$ model of TaAs2 derived in section 2.3, and then for a tight-binding model of NbSb2 derived from first-principles.

3.1. Band structure topology without magnetic fields

In the absence of a magnetic field, there is no direct band gap closure in AB2 compounds. Since the valence bands thus form a well-defined manifold, they can be classified, just like insulators, according to the topology of these valence bands. Because time-reversal symmetry is fulfilled, a ${{\mathbb{Z}}}_{2}$ classification is possible.

All compounds were found to be weak topological insulators, with ${{\mathbb{Z}}}_{2}$ indices $0;(111)$. That is, all time-reversal invariant planes ${k}_{i}=0,\,0.5$ have a non-trivial ${{\mathbb{Z}}}_{2}$ index ${\rm{\Delta }}=1$. This result was derived from first-principles using the Z2Pack code [37], and agrees with previous studies [15, 17, 32]. The corresponding evolution of the Wannier charge centers is shown, for the case of TaAs2, in figure 4.

Figure 4.

Figure 4. The Wannier charge center evolution for the time-reversal invariant planes of TaAs2. (a) kx = 0, (b) kx = 0.5, (c) ky = 0, (d) ky = 0.5, (e) kz = 0 and (f) kz = 0.5.

Standard image High-resolution image

Figure 5 shows the surface density of states for a slab of TaAs2, with surfaces parallel to the mirror plane perpendicular to the Cartesian y-axis (the light blue plane shown in figure 1). The presence of topological surface states confirms the conclusion that the material is a weak topological insulator. The surface spectrum was calculated by the iterative Green's function [38], which was implemented in WannierTools [39].

Figure 5.

Figure 5. The surface density of states of TaAs2 on the 010 surface, along the k-point path shown in figure 1.

Standard image High-resolution image

3.2. The effect of Zeeman splitting on the ${\bf{k}}\cdot {\bf{p}}$ model for TaAs2

Here we study the effects of the magnetic field on TaAs2 by adding a Zeeman splitting term to the ${\bf{k}}\cdot {\bf{p}}$ model derived in section 2.3 (equation (6)). The splitting term is given by

Equation (13)

where ci is the strength of the Zeeman splitting induced by the magnetic field in that direction, that is

Equation (14)

This assumes that the g-factor is equal for all bands. The limitations of this approximation are discussed in section 3.4.

3.2.1. Magnetic field along the rotation axis $\hat{y}$

When a magnetic field is applied along the rotation axis $\hat{y}$, the Zeeman term equation (13) takes the form

Equation (15)

This term preserves all spatial symmetries of the system, breaking only time-reversal.

Along the MA line, the Cxx and Cyy contributions to the Hamiltonian vanish since ${k}_{x}^{* }={k}_{y}^{* }$ and ${k}_{z}^{* }=0$. Consequently, the energy eigenvalues are given by

Equation (16)

The Zeeman term counteracts the original splitting (square root term), such that for a sufficient magnetic field there will be a direct band gap closure. Away from the MA line, the band gap remains open, giving rise to a Weyl point.

When the Zeeman splitting is gradually switched on, two pairs of Weyl points form at about ${c}_{y}=0.11\,\mathrm{eV}$. Increasing the Zeeman splitting leads to a separation between the two nodes in a pair, with one node each moving towards the M-point. Finally, at ${c}_{y}\approx 0.25\,\mathrm{eV}$, these two nodes meet at M and annihilate. This process is shown in figure 6.

Figure 6.

Figure 6. The band gap of TaAs2 in the ${k}_{z}^{* }=0$ plane for different values of the magnetic field in the y-direction, calculated from the ${\bf{k}}\cdot {\bf{p}}$ model. A dark spot indicates the presence of a Weyl point. (a) No magnetic field; there are no Weyl points present. (b) ${c}_{y}=0.11\,\mathrm{eV}$; two pairs of Weyl points have appeared on the kx = ky line. (c) ${c}_{y}=0.2\,\mathrm{eV}$; the pair of Weyl points moves further apart. (d) ${c}_{y}=0.25\,\mathrm{eV}$; one pair of Weyl points has annihilated at M, leaving two Weyl points.

Standard image High-resolution image

The existence of these Weyl points was confirmed by verifying that the nodes are a source or a sink of Berry curvature. For this purpose, the Chern number of spheres surrounding the points was calculated by tracking the hybrid Wannier charge centers (HWCCs) on loops around the sphere [2, 37, 40, 41], using the Z2Pack software [37]. Figure 7 shows the evolution of the sum of the HWCC for two of the four nodes found at ${c}_{y}=0.12\,\mathrm{eV}$, demonstrating that the two points are Weyl nodes of opposite chirality.

Figure 7.

Figure 7. The evolution of the sum of the HWCC on spheres surrounding the Weyl points at ${c}_{y}=0.12\,\mathrm{eV}$ splitting. (a) The Weyl point at ${\bf{k}}=(0.5247,0.5247,0.5)$, having positive chirality $C=+1$, (b) the Weyl point at ${\bf{k}}=(0.53258,0.53258,0.5)$ with negative chirality $C=-1$.

Standard image High-resolution image

3.2.2. General magnetic field direction

Finally, the effects of a magnetic field in a general direction were studied. It turns out that even though such a field breaks the spatial symmetries of the system, Weyl nodes still appear under a strong enough magnetic field. When a magnetic field is applied in an $\hat{x}$- or $\hat{z}$-direction, a single pair of Weyl points emerges from the M-point. These Weyl nodes are located on the ${k}_{x}=-{k}_{y}$ plane, as shown in table 5.

Table 5.  The position ${{\bf{k}}}^{* }={\bf{k}}-M$ (in reduced coordinates) and chirality of the Weyl points for Zeeman splitting in the $\hat{x}$- and $\hat{z}$-direction.

Splitting $[\mathrm{eV}]$   Weyl position ${{\bf{k}}}^{* }$ Chirality
cx = 0.225 $(-0.0042,0.0042,0.00093)$   −1
  $(0.0042,-0.0042,-0.00093)$   1
cx = 0.25 $(-0.025,0.025,0.0054)$ −1
  $(0.025,-0.025,-0.0054)$ 1
cx = 0.3 $(-0.044,0.044,0.0098)$ −1
  $(0.044,-0.044,-0.0098)$ 1
cz = 0.225 $(0.0011,-0.0011,-0.018)$   −1
  $(-0.0011,0.0011,0.018)$   1
cz = 0.25 $(0.0066,-0.0066,-0.11)$ −1
  $(-0.0066,0.0066,0.11)$ 1
cz = 0.3 $(0.012,-0.012,-0.18)$ −1
  $(-0.012,0.012,0.18)$ 1

Figure 8 shows the number of Weyl points as a function of the Zeeman splitting. To obtain this phase diagram, the candidate Weyl points were identified using a quasi-Newton algorithm to find the minima in the band gap (using SciPy.optimize.minimize [36]), for different initial guesses. In the second step, the Chern number on the small sphere (radius ${10}^{-4}\,{\mathring{\rm A} }^{-1}$) surrounding the candidate points was evaluated (using Z2Pack [37]), keeping only the points with a non-zero Chern number. Finally, duplicate points were eliminated by checking whether two points lie within the diameter of the sphere of one another.

Figure 8.

Figure 8. A phase diagram showing the number of Weyl points in the ${\bf{k}}\cdot {\bf{p}}$ model of the TaAs2 as a function of the Zeeman splitting (in $\,\mathrm{eV}$).

Standard image High-resolution image

3.3. The effect of Zeeman splitting on the tight-binding model for NbSb2

Having studied the effects of Zeeman splitting on the ${\bf{k}}\cdot {\bf{p}}$ model for TaAs2, we now study a more realistic tight-binding model for NbSb2, derived from a first-principles calculation with hybrid functionals using the Wannier90 code [42, 43]. NbSb2 was chosen because it has the smallest direct band gap of the four materials, making it the most promising candidate for hosting Weyl points at realistic magnetic field strength.

The Zeeman splitting for this model can again be expressed by adding the corresponding terms to the Hamiltonian

Equation (17)

where the change in the splitting terms (compared to equation (13)) is due to the different orbital basis used for the tight-binding model. We search for Weyl points between the last valence band and the first conduction band.

First, we study the effect of applying a magnetic field in the y-direction. Figure 9 shows the effect of this splitting along the MA line. For ${c}_{y}\approx 0.06\,\mathrm{eV}$, two pairs of Weyl points appear close to the MA line. The reason these points are not exactly on the line is because the crystal symmetry is broken when constructing the Wannier-based tight-binding model [42, 43]. Apart from the numerical difference, this effect is analogous to the case of the ${\bf{k}}\cdot {\bf{p}}$ model for TaAs2, where the two pairs of Weyl points appeared at ${c}_{y}=0.11\,\mathrm{eV}$.

Figure 9.

Figure 9. The band structure of the tight-binding model for NbSb2 along the MA line (a) without Zeeman splitting, and (b) with ${c}_{y}=0.06\,\mathrm{eV}$ Zeeman splitting.

Standard image High-resolution image

Table 6 shows the Weyl point positions, chirality and type for selected values of Zeeman splitting. It shows that Weyl points appear even at smaller values of cy away from the MA line. This is a crucial difference from the ${\bf{k}}\cdot {\bf{p}}$ model, which is only valid near the M-point. Furthermore, all Weyl points found for these splitting values are of type-II [2]. Type-II Weyl points have a tilted energy spectrum, making their Fermi surface open instead of point-like. As a consequence, their chiral anomaly—and their effect on magnetoresistance—is expected to be anisotropic.

Table 6.  The Weyl point positions (in reduced coordinates), chirality and type for different values of Zeeman splitting in the tight-binding model for NbSb2.

Splitting $[\mathrm{eV}]$   Position ${\bf{k}}$ Chirality Type
cx = 0.045 (0.4393, 0.4460, 0.5004) +1 II
  $(0.4359,0.4444,0.5026)$ −1 II
  $(0.5641,0.5556,0.4974)$ +1 II
  $(0.5607,0.5540,0.4996)$ −1 II
cy = 0.03 $(0.3670,0.5141,0.0977)$   +1 II
  $(0.3655,0.5142,0.1004)$ −1 II
  $(0.6345,0.4858,0.8997)$ +1 II
  $(0.6330,0.4858,0.9023)$ −1 II
cy = 0.04 $(0.3724,0.5116,0.0890)$ +1 II
  $(0.3627,0.5135,0.1055)$ −1 II
  $(0.6373,0.4865,0.8945)$ +1 II
  $(0.6276,0.4884,0.9110)$ −1 II
  $(0.9028,0.0340,0.5451)$ +1 II
  $(0.9018,0.0354,0.5390)$ −1 II
  $(0.0982,0.9646,0.4610)$ +1 II
  $(0.0974,0.9658,0.4545)$ −1 II
cy = 0.06 $(0.3791,0.5068,0.0775)$ +1 II
  $(0.3592,0.5131,0.1108)$ −1 II
  $(0.6407,0.4869,0.8892)$ +1 II
  $(0.6211,0.4929,0.9222)$ −1 II
  $(0.9033,0.0328,0.5532)$ +1 II
  $(0.9006,0.0364,0.5314)$ −1 II
  $(0.0994,0.9636,0.4686)$ +1 II
  $(0.0968,0.9671,0.4467)$ −1 II
  $(0.4493,0.4555,0.5031)$ +1 II
  $(0.4309,0.4320,0.4825)$ −1 II
  $(0.5691,0.5680,0.5175)$ +1 II
  $(0.5507,0.5445,0.4969)$ −1 II
cz = 0.0475 $(0.4494,0.4384,0.4853)$ +1 II
  $(0.4420,0.4366,0.4816)$ −1 II
  $(0.5580,0.5634,0.5184)$ +1 II
  $(0.5506,0.5616,0.5147)$ −1 II

Finally, a phase diagram was calculated showing the number of Weyl points as a function of the magnetic field (see figure 10). Unlike for the ${\bf{k}}\cdot {\bf{p}}$ model, the number of Weyl points keeps increasing when the applied Zeeman term grows stronger. Again, the reason for this difference is that Weyl points also form far away from the M-point, where the ${\bf{k}}\cdot {\bf{p}}$ approximation is no longer applicable.

Figure 10.

Figure 10. The phase diagram showing the number of Weyl points as a function of Zeeman splitting (in eV) for the tight-binding model for NbSb2.

Standard image High-resolution image

For some values of splitting, the phase diagram shows an odd number of Weyl points, which is physically impossible. The reason for this is that the numerical procedure used to identify the number of Weyl points may not find a Weyl point if it is too close to another one. Since this problem only occurs rarely (see figure 11), the phase diagram is still valid overall. Also, the procedure ensures that no Weyl point can be counted twice, so the phase diagram represents a lower limit for the real number of Weyl points. Thus, the general result, in which the number of Weyl points increases with stronger Zeeman splitting, remains valid.

Figure 11.

Figure 11. A phase diagram showing whether the number of Weyl points shown in figure 10 is even (physically possible) and odd (non-physical). The odd phases are a relict of the numerical evaluation of the phase.

Standard image High-resolution image

3.4. Limitations of the model for the magnetic field

In the previous sections, the effect of the magnetic field was modeled by applying Zeeman splitting to the model Hamiltonian. The discussion was simplified by assuming that the g-factor was equal for all energy bands, and independent of ${\bf{k}}$. Here, we discuss how the results might change if this assumption is not made.

If the g-factor is k-dependent, but still the same for all energy bands, the results above will change quantitatively, but not qualitatively. The reason for this is that the Weyl node that appears at a specific k-point will still be there, but for a different magnetic field. That is, the order in which the Weyl nodes at different k-points appear might change, but not the overall picture of an increasing number of Weyl points with a stronger magnetic field.

The same is true if the g-factor varies for different energy bands, as long as the sign of the g-factor remains the same. Because the appearance of Weyl points is due to the relative Zeeman splitting between the last valence and first electron bands, it does not matter how much the splitting of each band contributes.

If the g-factors in the relevant bands have opposite signs, however, there is a qualitative change in behavior. This is illustrated in the following with the example of the ${\bf{k}}\cdot {\bf{p}}$ model of TaAs2, discussed in sections 2.3, 3.2. To account for the opposite sign of the g-factor for the valence and conduction bands, the Zeeman splitting term (equation (13)) is changed to

Equation (18)

With cy splitting, the energy bands on the mirror plane are then given by

Equation (19)

As in equation (16), the Zeeman term counteracts the original splitting. The difference from the previous case is that this equation holds on an entire plane in reciprocal space instead of just a line. As a consequence, we can expect the appearance of a nodal line with sufficient Zeeman splitting. Indeed, a nodal line appears for ${c}_{y}\gtrsim 0.2242\,\mathrm{eV}$, as shown in figure 12. The Berry phase on a closed path around this nodal line was calculated to be π, using the Z2Pack [37] software. This verifies the topological nature of the nodal line.

Figure 12.

Figure 12. The band gap of TaAs2 on the mirror plane with Zeeman splitting as given in equation (18). (a) At ${c}_{y}\approx 0.2242\,\mathrm{eV}$, a nodal line appears at the M-point. (b) The nodal line expands for stronger splitting (${c}_{y}=0.23\,\mathrm{eV}$).

Standard image High-resolution image

In conclusion, the qualitative result obtained above remains intact when the g-factors are assumed to be k-dependent, and different for the valence and conduction bands, as long as they keep the same sign. A more adequate model of the magnetic field is needed to establish the exact qualitative and quantitative nature of the topological phases with the applied magnetic field. The current results indicate that Weyl nodes will appear, at least for some directions of magnetic field.

4. Conclusions

We studied the topological phase of transition metal dipnictides of the type AB2 (A ∈{Ta, Nb}, B ∈{As, Sb}), with and without an external magnetic field. In the absence of a magnetic field, we found—in accordance with previous results [15, 17]—that these materials can be classified as weak topological insulators, despite having an indirect band gap closure.

The effect of the magnetic field was studied by applying Zeeman splitting, and we found that Weyl points can appear. We showed this result first from theoretical considerations on a four-band ${\bf{k}}\cdot {\bf{p}}$ model, and numerically on a ${\bf{k}}\cdot {\bf{p}}$ model of TaAs2 and a tight-binding model of NbSb2. In the tight-binding model, we found the number of Weyl points to increase with a growing magnetic field. For specific values of Zeeman splitting, the type of Weyl points in the tight-binding model was studied, and they were all found to be of type-II.

The appearance of such field-induced Weyl points could help explain the reduced or negative magnetoresistivity in these materials. However, it is unclear whether the Weyl points studied here appear at a magnetic field that is realistic to observe in experiments. Further studies, in particular, to obtain a realistic g-factor and more reliable data for the direct band gap, are required to accurately estimate the required magnetic field. Furthermore, it is known that modeling a strong magnetic field with only Zeeman splitting is not sufficient, and a more accurate model should be considered. Finally, the effect of these Weyl points on the magnetoresistance should be calculated. This is influenced by the orientation of the type-II Weyl points, and their distance from the Fermi level.

Consequently, there are three open questions which require further investigation: first, whether the appearance of field-induced Weyl points is realistic in these AB2 compounds; second, if these Weyl points do appear, whether they alone are responsible for the experimentally observed behavior of magnetoresistance or if there are other effects; finally, whether there are other compounds which contain the same kind of field-induced Weyl points, possibly appearing already at a weaker magnetic field.

Acknowledgments

We would like to thank D Rodic and M Könz for helpful discussions. The authors were supported by ERC Advanced Grant SIMCOFE, Microsoft Research, and the Swiss National Science Foundation through the National Competence Centers in Research MARVEL and QSIT.  Calculations were performed on the Mönch cluster of ETH Zurich.

Please wait… references are loading.
10.1088/1367-2630/aa5de7