Quantum oscillations and three-dimensional quantum Hall effect in ZrTe5
Abstract
Recent experiments have reported a lot of spectacular transport properties in topological materials, such as quantum oscillations and three-dimensional (3D) quantum Hall effect (QHE) in ZrTe5. In this paper, by using a strong topological insulator model to describe ZrTe5, we study the magnetotransport property of the 3D system. With fixed carrier density, we find that there exists a deferring effect in the chemical potential, which favors distinguishing the saddle points of the inverted LLs. On the other hand, with fixed chemical potential, the features of 3D QHE are demonstrated and we attribute the underlying mechanisms to the interplay between Dirac fermions, magnetic field and impurity scatterings.
I Introduction
In the past two decades, the search of various topological phases in materials has ignited ongoing interest in condensed matter physics M.Z.Hasan ; X.L.Qi . Compared to conventional trivial phases, the most exotic property of topological phases is that they can support the surface states, which show robustness to the weak disorder or defects. To detect the topological phases in experiments, one may capture the surface states by tuning the chemical potential inside the gap B.Q.Lv ; J.A.Sobota . This works well for wide-gap topological insulators (TIs), such as Bi2Se3 and Bi2Te3, with a gap larger than meV Y.Xia ; H.Zhang ; Y.L.Chen ; C.X.Liu . But for a narrow-gap material, it is not easy to tune the chemical potential inside the gap.
Three-dimensional (3D) ZrTe5 is such a narrow-gap topological material H.Weng . Recent experiments show that it can exhibit a variety of unique transport properties, including chiral magnetic effect Q.Li , anomalous Hall effect T.Liang2018 , saturating thermoelectric Hall effect J.L.Zhang ; W.Zhang , and gigantic magnetochiral anisotropy Y.Wang . Instead of the surface states, an alternative route to probe the topological phase in ZrTe5 was proposed through detecting the saddle points of the inverted bands J.Wang . Under a magnetic field, electrons are forced to move on quantized curved orbits-the Landau levels (LLs), which, if inverted, can exhibit two saddle points in the Brillouin zone (BZ) Y.Jiang ; L.You , the point and the point. When the magnetic field changes, the chemical potential will cross the LLs one by one, leading to quantum oscillations (QOs) in transport and thermodynamics D.Shoeberg ; G.M.Monterio ; S.Kaushik , in which the signatures of the saddle points need further study.
3D quantum Hall effect (QHE) was another intriguing phenomenon observed recently in ZrTe5 F.Tang ; S.Galeski2020 and HfTe5 P.Wang ; S.Galeski2020 . The Hall resistivity was reported to exhibit a quasi-quantized plateau in the quantum limit. The quantum limit means that only the zeroth LL is occupied, which can be reached at a critical magnetic field of about 1.3 T in ZrTe5 F.Tang and 1.8 T in HfTe5 P.Wang , both indicating the low carrier densities in these 3D crystals. More importantly, the plateau scales as , with being the Fermi wavevector. It was proposed that the enhanced interactions in the quantum limit can induce the Fermi surface instability, which in turn drives the charge density wave (CDW) that would open a gap around the Fermi level F.Tang ; P.Wang . In a 3D system, applying a magnetic field can reduce the energy spectrum to one dimension (1D). Since the 1D system features an almost perfect nesting Fermi surface, its ground state favors the CDW states G.Gruner , thus it seems that the CDW picture may be supported in explaining 3D QHE F.Qin .
Nevertheless, there exist several facts against such a CDW scenario: (i) if the Fermi energy lies in the gap, the quasi-quantized Hall plateaus should correspond to the vanishing longitudinal resistivity , which, however, remains finite as observed in ZrTe5 F.Tang ; S.Galeski2020 and HfTe5 P.Wang ; S.Galeski2020 ; (ii) the CDW picture can only account for the plateau in the quantum limit, but not for the other plateaus in the QO regime F.Tang ; S.Galeski2020 ; P.Wang ; and (iii) in a recent study S.Galeski2021 , Galeski et al., corroborated that the CDW states were absent in the ZrTe5 crystal through various measurements, based on which they suggested that 3D QHE emerges from the interplay between the intrinsic properties of ZrTe5 and its Dirac-type semimetal character. Thus, the physical mechanism behind 3D QHE also requires further study.
Motivated by this progress, in this paper, we use the strong TI model to describe the low-energy excitations in ZrTe5 and systematically investigate its bulk magneotransport, with a focus on the problems of QOs and 3D QHE. We will consider both conditions of fixed carrier density and fixed chemical potential. The former condition was assumed in understanding the saturating thermoelectric effect J.L.Zhang , anomalous resistivity at finite temperature B.Fu ; C.Wang in ZrTe5, whereas the latter one was used to explain the optical spectroscopy B.Xu ; E.Martino and magnetoinfared spectroscopy measurements R.Y.Chen ; Y.Jiang in ZrTe5. We will calculate the density of states (DOS), the oscillating chemical potential or carrier density, as well as the conductivities and resistivities in the 3D system. In the calculations, the impurity scatterings are included in the system phenomenologically by introducing the linewidth broadening to the LLs.
Our results show that under the condition of fixed carrier density, there exists a deferring effect in the chemical potential, which favors distinguishing the saddle points of the inverted LLs. At weak impurity scatterings, the anomalous peak in the quantum limit and the fourfold peaks in the QO regime can be captured in the DOS as well as the longitudinal resistivity , which can help identify the band inversions of the LLs. Under the condition of fixed chemical potential, the characteristics of 3D QHE are demonstrated, including the quasi-quantized plateau in the Hall resistivity and the nonvanishing dips in . We attribute 3D QHE to the interplay between Dirac fermions, magnetic field, and impurity scatterings. Our paper could provide more insights to understand the peculiar transport properties of ZrTe5 in experiments.
II Model
In the study of ZrTe5, a simple two-band model was proposed to calculate the experimental observables Y.X.Wang2020a ; Z.Rukelj , and also a four-band model based on the ab initio approach was used to investigate the optical conductivity in the system C.Morice . Since it is generally believed that the ground state of 3D ZrTe5 lies in the proximity of the phase boundary between a strong TI and a weak TI H.Weng , here we adopt the strong TI model to describe its low-energy excitations. The strong TI phase features the in-plane and out-of-plane band inversions, and has been supported by many experimental evidences in ZrTe5 G.Manzoni ; B.Xu ; Y.Jiang ; J.Wang ; Z.G.Chen .
In the four-component basis , the Hamiltonian reads () Y.Jiang ; J.Wang ; H.Zhang ; C.X.Liu
(1) |
where and are the Pauli matrices acting on the spin and orbit degrees of freedom, respectively. and are the Fermi velocities, and are the band inversion parameters, and denotes the Dirac mass. In the following calculations, we take the parameters from the magnetoinfrared measurements Y.Jiang : m/s, meV nm2, meV. With these parameters, besides the point, owns the second saddle point point, which is located at , with . In addition, can support more topological phases when the parameters are taken in a broad range, such as single Dirac point phase, Dirac ring phase Y.X.Wang2021 , weak TI phase W.Wu and Dirac semimetal phase.
Under a perpendicular magnetic field , the quantized LLs are formed, with dispersions along the direction. The magnetic field is included in the system with the help of the Peierls substitution, , where the vector potential is chosen as in the Landau gauge. We then introduce the ladder operators and , with the magnetic length nm. By using the trial wavefunction , in which the state is defined as and denote the coefficients, the energies for the zeroth and LLs are obtained as L.You
(2) | ||||
(3) |
respectively, where the index denotes the conduction/valence band, and the two branches representing the Zeeman splitting between up spin and down spin L.You . With increasing magnetic field, we see that the zeroth LL moves to zero energy, while the LLs move away from zero energy. The latter is because in Eq. (3), the second term under the square root is much larger than the first squared term.
The band inversions can also exist in LLs. The inverted points are located at
(4) | |||
(5) |
for the zeroth and LL, respectively. Note that in , the weak term has been neglected.
At a strong magnetic field, the inverted LLs may become noninverted L.You , implying that the point would be merged with the point. For the zeroth LL, the critical magnetic field is T, and for the LL, T. Below we see that the chosen magnetic fields are much weaker than these critical values and thus the LLs are always inverted.
III Method
III.1 Density of states and carrier density
We calculate the DOS of 3D ZrTe5 under a magnetic field. With the help of the retarded Green’s function Y.X.Wang2021 , the DOS is defined as
(6) |
where is system size in the direction, is the LL degeneracy in the - plane and can be denoted as the uniform DOS, and represents the linewidth broadening phenomenologically that is induced by impurity scatterings. Consider the short-range pointlike impurities, , and the impurity concentration , then in the Born approximation, the scattering time P.Hosur ; A.A.Burkov2014 ; J.Klier ; H.W.Wang , with characterizing the strength of the impurity potential, and the DOS at the Fermi energy. Note that when the band inversions are absent, , the divergence in the DOS is one over the square-root type S.Kaushik ; J.Klier ; Y.X.Wang2019 . Now the LL energies show complicated dependencies on , leading to the intricated behavior of the DOS, as shown below.
On the other hand, the carrier density is related to the DOS by Ashby
(7) |
where is the Fermi distribution function, with being the chemical potential, the Boltzmann constant, and the temperature. In the following, we focus on zero temperature.
III.2 Conductivity and resistivity
We also calculate the conductivity and resistivity of the system. The conductivity tensors can be derived from the linear-response Kubo-Streda formula L.Smrcka ; G.D.Mahan ,
(8) |
where is the system volume, is the current density operator, and is the advanced Green’s function. In the transport theory S.Datta , the relaxation time denotes the time that an electron experiences from the undisturbed state to the equilibrium state through impurity scatterings, while the inverse LL broadening is related to the lifetime that an electron stays in the state. Here we assume that both quantities are equal to the scattering time S.Galeski2021 . Such an assumption has the advantage that the conductivities can be directly determined from the band structures.
After a straightforward calculation, we obtain the longitudinal conductivity and transverse Hall conductivity , which are expressed as a summation over the LL index (see Appendix A),
(9) |
and
(10) |
where is the unit of the quantum conductivity, is the step function, the summation sign is
(11) |
and the matrix element is
(12) |
Note that the nonvanishing matrix element determines the common selection rules Y.Jiang ; L.You .
In the transport experiments, since the measured quantities are the resistivities , we need to convert conductivities into resistivities. Using the relation , we obtain
(13) |
IV Main Results and Discussions
IV.1 Fixed carrier density
In this section, we study the magnetotransport in ZrTe5 under the condition of fixed low carrier density, which is set as cm-3.

The calculated DOS and chemical potential are plotted in Fig. 1(a). We see that when T-1, the system lies in the quantum limit, with an anomalous peak in the DOS. When T-1, the system lies in the QO regime. In this regime, the DOS exhibits the fourfold peak structure and the chemical potential oscillates with descending amplitudes. They both show a period T-1. When is asymptotically large (not shown here), the system will enter the semiclassical regime H.W.Wang , in which the energy spacing between two adjacent LLs becomes smaller than the linewidth broadening .
The oscillation period can be understood with the Onsager’s relation L.Onsager ; D.Shoeberg . It tells us that the period is inversely proportional to the extremal cross-sectional Fermi surface area in the plane perpendicular to the magnetic field:
(14) |
When the magnetic field is absent, the chemical potential for the fixed is meV (see Appendix B), as indicated by the red dotted line in Fig. 1(a). The corresponding extremal Fermi surface is plotted in the inset of Fig. 1(a). The Fermi surface is almost an ellipse with the area nm-2. Then we obtain T-1, agreeing well with the period extracted from Fig. 1(a).
To illustrate the chemical potential oscillations, we show how crosses the LLs. The LL dispersions are plotted in Figs. 1(b) and 1(c), with the magnitudes of corresponding to the blue lines and in Fig. 1(a), respectively. Actually with increasing , is determined by the competition between the local DOS gaining due to the 1D LL dispersion and the uniform DOS dropping induced by the magnetic field (i.e., the LL degeneracy ). In Fig. 1(b), when T-1, we have meV, which meets the point of the LL [see Fig. 1(b), inset]. Since the local DOS gaining surpasses the uniform DOS dropping, to conserve the carrier density, will decrease and thus meV behaves as a peak. In Fig. 1(c), when T-1, the chemical potential decreases to meV, which lies above the point of the LL [see Fig. 1(c), inset]. With further increasing , the local DOS gaining is inferior to the uniform DOS dropping and thus meV behaves as a valley. After that, goes into the next oscillation. Importantly, in the local region that includes four saddle points of the -th LLs, the decreasing chemical potential with can defer its crossing over these saddle points. Such a deferring effect favors distinguishing the saddle points with weak energy difference.
In the DOS, since the peaks are related to the saddle points of the bands, here the anomalous peak in the quantum limit is caused by the point of the zeroth LL, and the fourfold peaks in the LLs are attributed to the fact that the branches are splitted and each branch owns two saddle points, the point and the point.

We discuss the saddle point sequence that the chemical potential sweeps with increasing . If
(15) |
the saddle point is swept in the sequence ---. By contrast, if
(16) |
the sequence will be ---. For example, in the inset of Fig. 1(b), we have , thus the saddle points are swept in the former sequence, which is also valid for other LLs. In a previous work J.Wang , the carrier density of the ZrTe5 crystal is reported to be above cm-3 and thus the quantum limit is reached at the critical magnetic field T. Such a strong magnetic field results in the well-separated branches and the latter sequence ---.
If we include the spin Zeeman effect into the system, the Hamiltonian is
(17) |
where denotes the Landé -factor and is the Bohr magneton. Taking the value that is reported in ZrTe5 R.Y.Chen ; Z.G.Chen , the total Zeeman splittings will get strengthened, leading to the further separation of the two branches. Thus, for the lower LLs, the saddle point sequence may become ---, and for the higher LLs, the sequence remains unchanged. This means that with increasing LL index , there exists a transition of the saddle point sequence, which depends heavily on the carrier density and the -factor of the LLs.
Next we study the conductivities. The results are plotted as a function of the inverse magnetic field in Figs. 2(a) and 2(b). We analyze the longitudinal conductivity in Fig. 2(a). In the limiting clean case , which corresponds to the infinite scattering time, vanishes. The impurity-induced scatterings can give rise to the drift current along the electric field direction, leading to the increasing of with weak . When meV, exhibits clear oscillations and fourfold peaks that are similar to the DOS. When meV, the fourfold peaks are smoothened and become broad. Further increasing to be as strong as meV, we observe that the oscillations in are smeared out, implying that the LL structures are broken. Moreover, in the quantum limit, will get further increased, while in the QO regime, decreases. Such an effect of impurity scatterings on is consistent with our self-consistent Born approximation study of the diagonal disorder in a Weyl semimetal system Y.X.Wang2020 .
We also analyze the Hall conductivity in Fig. 2(b). When , is proportional to the inverse magnetic field , which retrieves the classical Hall conductivity expression V.Konye ,
(18) |
With the extracted slope m cm-1 T, the carrier density is obtained as cm-3, which is consistent with the chosen carrier density. For the weak impurity scattering meV, can be ignored in Eq. (10), thus shows certain robustness to weak . When increases, the robustness disappears and will get suppressed from the higher LLs to lower ones Y.X.Wang2019 . At strong meV, even becomes negative at large . This is because when the chemical potential lies between the -th and -th LLs, the dominate contributions to come from the LL transition , which is similar to two-dimensional (2D) Dirac fermion in graphene V.P.Gusynin . Since the energy spacing between the higher LLs is smaller, the associated will get suppressed at first, which will occur successively for the lower LLs. In the case when the energy spacing is less than the strong , becomes negative.

Then we study the resistivity. To make direct comparisons with experiments, the resistivities are plotted as a function of the magnetic field in Figs. 2(c) and 2(d). With weak meV, the conductivities satisfy , and correspondingly, the longitudinal resistivity and Hall resistivity are given as
(19) |
respectively. Thus in Fig. 2(c), exhibits the SdH oscillations on an increasing background, which keeps the features of . The fourfold peak structure of is more clearly seen in the inset of Fig. 2(c). On the other hand, shows a linear dependence on . The linear was derived by Abrikosov for Dirac electrons occupying the lowest zeroth LL Abrikosov , where the large and nonsaturated magnetoresistivity was understood with fixed carrier density. In experiment, such a linear was observed in the Weyl semimetal NbP C.Shekhar and the Dirac semimetal Cd3As2 L.P.He ; T.Liang2015 even at a very low magnetic field.
At strong meV, becomes comparable to . When the chemical potential crosses the th LLs, the fourfold peaks in merge into the single peak structure. Now in Fig. 2(d), for , the SdH oscillations are blurred but the kinks are seen. For , it exhibits a ramp in the quantum limit that gradually increases with . Note that shows a plateau in a broad range T T. In fact, such a plateau originates from the plateaus in and that are induced by impurity scatterings, as indicated by the arrows in Figs. 2(a) and 2(b). These results strongly suggest that 3D QHE cannot be explained with fixed carrier density.
IV.2 Fixed chemical potential

In this section, we study the magnetotransport of ZrTe5 under the condition of fixed chemical potential, which is set as meV.
The calculated DOS and carrier density are plotted in Fig. 3(a). The quantum limit is reached at the critical magnetic field T. Now as the point of the zeroth LL is always below the chemical potential,
(20) |
the anomalous peak that appears in Fig. 1(a) is absent. In the QO regime, we can see that the fourfold peaks of the DOS are overlapping and thus are difficult to be distinguished. On the other hand, the carrier density oscillates with a period T, implying that the charging energies exist in the system. The period can also be understood with the Onsager’s relation. Without a magnetic field, the extremal Fermi surface in the - plane is plotted in the inset of Fig. 3(a), with its area nm-2. Then, according to Eq. (14), we have T-1, which shows good consistency.
The competition between the local DOS and the uniform DOS that was used to analyze the chemical potential oscillations in the previous section can be extended to illustrate the carrier density oscillations here. It is worth pointing out that the oscillations in Figs. 1(a) and 3(a) are out of phase. For example, in the insets of Figs. 1(b) and 3(b), although both the chemical potentials meet the point of the LL, in Fig. 1(a), behaves as a peak, whereas in Fig. 3(a), the corresponding carrier density behaves as a valley.
Next we turn to the conductivity and resistivity. Figures 4(a) and 4(b) show that the effects of the increasing impurity scatterings on the conductivities and are the same as those in Figs. 2(a) and 2(b), respectively. Figure 4(c) shows that with weak linewidth meV, in the longitudinal resistivity , the SdH oscillations are evident, but the fourfold peaks can only be distinguished for the LLs and are overlapping for other LLs, with details seen in the inset of Fig. 4(c). This means that the saddle points of the inverted LLs cannot be well distinguished with fixed chemical potential L.You , even when the spin Zeeman effect is included. In the Hall resistivity , the quasi-quantized plateaus are clearly observable.
Actually, a 3D Hall system under a magnetic field can be regarded as the 2D slice stacking A.A.Burkov2011 ; Y.X.Wang2019 . Then the Hall conductivity (in unit of ) is a summation of the Chern numbers of the occupied LLs (see Appendix C),
(21) |
in which the Chern number of the LL at wave vector is for Dirac fermions. In the quantum limit, the Fermi wave vector is obtained as
(22) |
in which the -dependent term is much weaker than and can be neglected. Then we have , leading to the quasi-quantized plateau in that is estimated to be m cm, as shown in Fig. 4(c). For the LL, if occupied, the Fermi wavevector is obtained as
(23) |
Here, as the -dependent term cannot be neglected, decreases with . Thus in Fig. 4(c), the plateau of , which is given as , with the factor accounting for the two LL branches, increases slowly with . For the kink of around T in Figs. 4(c) and 4(d), it occurs when the branch is occupied and the branch is unoccupied, with the magnitude .
In Fig. 4(d) when increases to 1 meV, we observe that in , the fourfold peaks no longer exist, but the SdH oscillations are still discernible, while in , the Hall plateaus are preserved, with the neighboring plateau transitions becoming smooth. More importantly, the nonvanishing dips in correspond to the plateaus in , and the peaks in correspond to the plateau transition regions in . Although the peak heights in and the plateau magnitudes in depend heavily on the model parameters, the numerical results are qualitatively consistent with the behaviors of and observed experimentally in ZrTe5 F.Tang and HfTe5 P.Wang , which thus capture the main features of 3D QHE.
In a 2D electron system, the magnetotransport is often described by fixed chemical potential, resulting in the well-defined Hall plateaus V.P.Gusynin ; D.N.Sheng . Here also under the condition of fixed chemical potential, we can understand 3D QHE as a close analog to 2D QHE, in which the quasi-quantized plateaus in are determined by the dependence of the Fermi wavevectors on the magnetic field. Moreover, the impurity scatterings play indispensable roles in driving 3D QHE, since they can enhance the peak heights in as well as smoothen the plateau transitions in . Although our results based on a non-interacting model are quite different from the interaction-induced CDW scenario F.Tang ; P.Wang ; F.Qin , but are supported by a recent study S.Galeski2021 , in which the CDW states are demonstrated to be absent in ZrTe5.
Note that in a Dirac semimetal with fixed chemical potential, the Fermi wavevectors are solved as
(24) |
for the LL, and
(25) |
for the LL. We see that the Fermi wavevectors in Eqs. (24) and (25) show similar dependence on the magnetic field as those in Eqs. (22) and (23). Based on these results, we suggest that the proposed mechanism for 3D QHE is justified for a 3D Dirac fermion system, no matter the ground state is a strong TI ( and ), a weak TI ( and ), or even a Dirac semimetal.
V Conclusions
To summarize, we have numerically studied QOs and 3D QHE in ZrTe5 by using the strong TI model. Under the conditions of fixed carrier density and fixed chemical potential, although both QOs can be explained in a similar way and the impurity scatterings show a similar effects on the conductivities, the main conclusions are totally different.
With fixed carrier density, we find that there exists a deferring effect in the chemical potential, thus the saddle points of the inverted LLs are well distinguished by tuning the magnetic field magnitude. This can provide an effective route beyond the surface states to identify the nontrivial topological phase. On the other hand, with fixed chemical potential, we demonstrate that 3D QHE originates from the interplay between Dirac fermions, magnetic field and impurity scatterings, which may represent a universal mechanism of the 3D Dirac fermion systems and needs more studies in the future.
VI Acknowledgment
This work was supported by the Natural Science Foundation of China (Grant No. 11804122), the China Postdoctoral Science Foundation (Grant No. 2021M690970).
VII Appendix
VII.1 Derivation of Eqs. (9) and (10) in the main text
The conductivity tensors can be derived from the Kubo-Streda formula in Eq. (8). At zero temperature, the longitudinal conductivity is written as
(A1) |
in which the identity is used in the second line. Note that the expression of in Eq. (A1) is the same as those used in previous studies H.W.Wang ; Y.X.Wang2020 . Consider the LL degeneracy , and insert the energies and wavefunctions of the LLs, then we have
(A2) |

Similarly, the Hall conductivity is
(A3) |
In previous literatures, the expressions of are much more complicated H.W.Wang ; B.Fu ; C.Wang and include a normal part and an anomalous part, where the former is determined by the states around the Fermi level and the latter reflects the thermodynamic property of the system. Here Eq. (A3) is quite concise and thus are facilitate to the numerical calculations.
After calculating the current density matrix elements , we obtain
(A4) |
and
(A5) |
where the summation signs are
(A6) |
and the matrix element is given in Eq. (12).
We note that the components of the conductivity satisfy the following relations:
(A7) | |||
(A8) |
implying that the contributions to from the LL transition and from are equal, while those to are opposite. According to these properties, we can change the index , , and in the second summation of Eqs. (A4) and (A5), and obtain the final expressions in Eqs. (9) and (10) in the main text.
VII.2 Density of states and chemical potential with no magnetic field
When the magnetic field is absent, the energies for the Hamiltonian in Eq. (1) can be obtained directly, which are given as
(A9) |
where denotes the conduction/valence band. Note that both the conduction and valence bands own twofold degeneracy.
The DOS is calculated as,
(A10) |
where denotes the twofold degeneracy. To do the integrations, we make the following substitutions Y.X.Wang2020 : , , , and , , and change the Cartesian into the Cylindrical system: , . After completing the integration over the azimuth angle , we have
(A11) |
On the other hand, the carrier density is defined in Eq. (7), which, at zero temperature, is written as
(A12) |
Evidently, the integrands in Eqs. (A11) and (A12) show complicated dependencies on and , thus the integration cannot be solved analytically. But we can resort to the numerical calculations. In Figs. 5(a) and 5(b), we plot the calculated DOS as a function of the energy and the carrier density as a function of the chemical potential , respectively. We see that the DOS increase with and increases with , both in a monotonous tendency. In Fig. 5(b), the red dashed lines denote that the chosen carrier density cm-3 in Sec. IVA corresponds to the chemical potential meV.
VII.3 Derivation of Eq. (21) in the main text
Here we give a detailed derivation of Eq. (21). The 2D Chern number of the -th band is defined as an integration over the BZ,
(A13) |
where is the Berry curvature and is the Berry connection. Then we have
(A14) |
where and are the energy and wave function of the -th band. Note that in the second line of Eq. (A14), the identity is used. Comparing Eq. (A14) with Eq. (A3), we obtain Eq. (21) in the main text.
References
- (1) M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).
- (2) X. L. Qi and S. C. Zhang, Rev. Mod. Phys. 83, 1057 (2011).
- (3) B. Q. Lv, T. Qian, and H. Ding, Rev. Mod. Phys. 93, 025002 (2021).
- (4) J. A. Sobota, Y. He, and Z. X. Shen, Rev. Mod. Phys. 93, 025006 (2021).
- (5) Y. Xia, D. Qian, D. Hsieh, L. Wray, A. Pal, H. Lin, A. Bansil, D. Grauer, Y. S. Hor, R. J. Cava, and M. Z. Hasan, Nat. Phys. 5, 398 (2009).
- (6) Y. L. Chen, J. G. Analytis, J. H. Chu, Z. K. Liu, S. K. Mo, X. L. Qi, H. J. Zhang, D. H. Lu, X. Dai, Z. Fang, S. C. Zhang, I. R. Fisher, Z. Hussain, Z. X. Shen, Science 325, 178 (2009).
- (7) H. Zhang, C. X. Liu, X. L. Qi, X. Dai, Z. Fang, and S. C. Zhang, Nat. Phys. 5, 438 (2009).
- (8) C. X. Liu, X. L. Qi, H. J. Zhang, X. Dai, Z. Fang, and S. C. Zhang, Phys. Rev. B 82, 045122 (2010).
- (9) H. Weng, X. Dai, and Z. Fang, Phys. Rev. X 4, 011002 (2014).
- (10) Q. Li, D. E. Kharzeev, C. Zhang, Y. Huang, I. Pletikosic, A. V. Fedorov, R. D. Zhong, J. A. Schneeloch, G. D. Gu, and T. Valla, Nat. Phys. 12, 550 (2016).
- (11) T. Liang, J. Lin, Q. Gibson, S. Kushwaha, M. Liu, W. Wang, H. Xiong, J. A. Sobota, M. Hashimoto, P. S. Kirchmann, Z. X. Shen, R. J. Cava and N. P. Ong, Nat. Phys. 14, 451 (2018).
- (12) J. L. Zhang, C. M. Wang, C. Y. Guo, X. D. Zhu, Y. Zhang, J. Y. Yang, Y. Q. Wang, Z. Qu, L. Pi, H. Z. Lu, and M. L. Tian, Phys. Rev. Lett. 123, 196602 (2019).
- (13) W. Zhang, P. Wang, B. Skinner, R. Bi, V. Kozii, C. W. Cho, R. Zhong, J. Schneeloch, D. Yu, G. Gu, L. Fu, X. Wu, and L. Zhang, Nat. Commun. 11, 1046 (2020).
- (14) Y. Wang, H. F. Legg, T. Bömerich, J. Park, S. Biesenkamp, A. A. Taskin, M. Braden, A. Rosch, and Y. Ando, Phys. Rev. Lett. 128, 176602 (2022).
- (15) J. Wang, Y. Jiang, T. Zhao, Z. Dun, A. L. Miettinen, X. Wu, M. Mourigal, H. Zhou, W. Pan, D. Smirnov, and Z. Jiang, Nat. Comm. 12, 6758 (2021).
- (16) Y. Jiang, J. Wang, T. Zhao, Z. L. Dun, Q. Huang, X. S. Wu, M. Mourigal, H. D. Zhou, W. Pan, M. Ozerov, D. Smirnov, and Z. Jiang, Phys. Rev. Lett. 125, 046403 (2020).
- (17) L. You, Z. Y. Zhang, and Y. X. Wang, New J. Phys. 23, 123033 (2021).
- (18) D. Shoenberg, Magnetic Oscillations in Metals, (Cambridge University Press, Cambridge, 1984).
- (19) G. M. Monteiro, A. G. Abanov, and D. E. Kharzeev, Phys. Rev. B 92, 165109 (2015).
- (20) S. Kaushik, D. E. Kharzeev, Phys. Rev. B 95, 235136 (2017).
- (21) F. Tang, Y. Ren, P. Wang, R. Zhong, J. Schneeloch, S. A. Yang, K. Yang, P. A. Lee, G. Gu, Z. Qiao, and L. Zhang, Nature 569, 537 (2019).
- (22) S. Galeski, X. Zhao, R. Wawrzyńczak, T. Meng, T. Förster, P. M. Lozano, S. Honnali, N. Lamba, T. Ehmcke, A. Markou, Q. Li, G. Gu, W. Zhu, J. Wosnitza, C. Felser, G. F. Chen, J. Gooth, Nat. Commun. 11, 1 (2020).
- (23) P. Wang, Y. Ren, F. Tang, P. Wang, T. Hou, H. Zeng, L. Zhang, and Z. Qiao, Phys. Rev. B 101, 161201(R) (2020).
- (24) G. Grüner, Density Waves in Solids, (Perseus, Cambridge, 2000).
- (25) F. Qin, S. Li, Z. Z. Du, C. M. Wang, W. Zhang, D. Yu, H. Z. Lu, and X. C. Xie, Phys. Rev. Lett. 125, 206601 (2020).
- (26) S. Galeski, T. Ehmcke, R. Wawrzyńczak, P. M. Lozano, K. Cho, A. Sharma, S. Das, F. Küster, P. Sessi, M. Brando, R. Küchler, A. Markou, M. König, P. Swekis, C. Felser, Y. Sassa, Q. Li, G. Gu, M. V. Zimmermann, O. Ivashko, D. I. Gorbunov, S. Zherlitsyn, T. Förster, S. S. P. Parkin, J. Wosnitza, T. Meng, and J. Gooth, Nat. Commun. 12, 3197 (2021).
- (27) B. Fu, H. W. Wang, and S. Q. Shen, Phys. Rev. Lett. 125, 256601 (2020).
- (28) C. Wang, Phys. Rev. Lett. 126, 126601 (2021).
- (29) B. Xu, L. X. Zhao, P. Marsik, E. Sheveleva, F. Lyzwa, Y. M. Dai, G. F. Chen, X. G. Qiu, and C. Bernhard, Phys. Rev. Lett. 121, 187401 (2018).
- (30) E. Martino, I. Crassee, G. Eguchi, D. Santos-Cottin, R. D. Zhong, G. D. Gu, H. Berger, Z. Rukelj, M. Orlita, C. C. Homes, and A. Akrap, Phys. Rev. Lett. 122, 217402 (2019).
- (31) R. Y. Chen, Z. G. Chen, X. Y. Song, J. A. Schneeloch, G. D. Gu, F. Wang, and N. L. Wang, Phys. Rev. Lett. 115, 176404 (2015).
- (32) Y. X. Wang and F. Li, Phys. Rev. B 101, 195201 (2020).
- (33) Z. Rukelj, C. C. Homes, M. Orlita, and A. Akrap, Phys. Rev. B 102, 125201 (2020).
- (34) C. Morice, E. Lettl, T. Kopp, and A. P. Kampf, Phys. Rev. B 102, 155138 (2020).
- (35) Z. G. Chen, R. Y. Chen, R. D. Zhong, J. Schneeloch, C. Zhang, Y. Huang, F. Qu, R. Yu, Q. Li, G. D. Gu, and N. L. Wang, Proc. Natl. Acad. Sci. 114, 816 (2017).
- (36) G. Manzoni, L. Gragnaniello, G. Autés, T. Kuhn, A. Sterzi, F. Cilento, M. Zacchigna, V. Enenkel, I. Vobornik, L. Barba, F. Bisti, Ph. Bugnon, A. Magrez, V. N. Strocov, H. Berger, O. V. Yazyev, M. Fonin, F. Parmigiani, and A. Crepaldi, Phys. Rev. Lett. 117, 237601 (2016).
- (37) Y. X. Wang, and F. Li, Phys. Rev. B 103, 115202 (2021).
- (38) W. Wu, Z. Shi, Y. Du, Y. Wang, F. Qin, X. Meng, B. Liu, Y. Ma, Z. Yan, M. Ozerov, C. Zhang, H. Z. Lu, J. Chu and X. Yuan, Nat. Mat. 22, 84 (2023).
- (39) P. Hosur, S. A. Parameswaran, and A. Vishwanath, Phys. Rev. Lett. 108, 046602 (2012).
- (40) A. A. Burkov, Phys. Rev. Lett. 113, 187202 (2014).
- (41) J. Klier, I. V. Gornyi, and A. D. Mirlin, Phys. Rev. B 92, 205113 (2015).
- (42) H. W. Wang, B. Fu, and S. Q. Shen, Phys. Rev. B 98, 081202(R) (2018).
- (43) Y. X. Wang, EPL 126, 67005 (2019).
- (44) P. E. C. Ashby, J. P. Carbotte, Phys. Rev. B 87, 245131 (2013).
- (45) L. Smrcka and P. Streda, J. Phys. C: Solid State Phys. 10, 2153 (1977).
- (46) G. D. Mahan, Many-Particle Physics, 3rd ed. (Plenum, New York, 2000).
- (47) S. Datta, Electronic Transport in Mesoscopic Systems, (Cambridge University Press, Cambridge, 1995).
- (48) L. Onsager, Philos. Mag. 43, 1006 (1952).
- (49) Y. X. Wang and F. Li, Phys. Rev. B 101, 085201 (2020).
- (50) V. Könye and M. Ogata, Phys. Rev. B 98, 195420 (2018).
- (51) V. P. Gusynin and S. G. Sharapov, Phys. Rev. B 73, 245411 (2006).
- (52) A. A. Abrikosov, Phys. Rev. B 58, 2788 (1998).
- (53) C. Shekhar, A. K. Nayak, Y. Sun, M. Schmidt, M. Nicklas, I. Leermakers, U. Zeitler, Y. Skourski, J. Wosnitza, Z. Liu, Y. Chen, W. Schnelle, H. Borrmann, Y. Grin, C. Felser and B. Yan, Nat. Phys. 11, 645 (2015).
- (54) L. P. He, X. C. Hong, J. K. Dong, J. Pan, Z. Zhang, J. Zhang, and S. Y. Li, Phys. Rev. Lett. 113, 246402 (2014).
- (55) T. Liang, Q. Gibson, M. N. Ali, M. Liu, R. J. Cava and N. P. Ong, Nat. Mat. 14, 280 (2015).
- (56) A. A. Burkov and L. Balents, Phys. Rev. Lett. 107, 127205 (2011).
- (57) D. N. Sheng, L. Sheng, and Z. Y. Weng, Phys. Rev. B 73, 233406 (2006).