Theory for the negative longitudinal magnetoresistance in the quantum limit of Kramers Weyl semimetals
Abstract
Negative magnetoresistance is rare in non-magnetic materials. Recently, a negative magnetoresistance has been observed in the quantum limit of -Ag2Se, where only one band of Landau levels is occupied in a strong magnetic field parallel to the applied current. -Ag2Se is a material that host a Kramers Weyl cone with band degeneracy near the Fermi energy. Kramers Weyl cones exist at time-reversal invariant momenta in all symmorphic chiral crystals, and at a subset of these momenta, including the point, in non-symmorphic chiral crystals. Here, we present a theory for the negative magnetoresistance in the quantum limit of Kramers Weyl semimetals. We show that, although there is a band touching similar to those in Weyl semimetals, negative magnetoresistance can exist without a chiral anomaly. We find that it requires screened Coulomb scattering potentials between electrons and impurities, which is naturally the case in -Ag2Se.
I Introduction
Magnetoresistance measures the change of the electric resistance of a solid due to magnetic fields. In magnetic materials, randomly-oriented ferromagnetic domains suppress the tunneling of electrons and increase the resistance. A magnetic field can align the domains and thus lower the resistance, leading to a negative magnetoresistance. In contrast, negative magnetoresistance is rare in non-magnetic materials Argyres and Adams (1956), because the Lorentz force imposed by the magnetic field prevents electrons from moving forward. One of the mechanisms of negative magnetoresistance is weak localization Lee and Ramakrishnan (1985), which is induced by quantum interference and thus only survives at extremely low temperatures. Recently, negative magnetoresistance at higher temperatures in non-magnetic topological insulators Wang et al. (2012); He et al. (2013); Wiedmann et al. (2016); Wang et al. (2015); Breunig et al. (2017); Assaf et al. (2017) and semimetals Kim et al. (2013, 2014); Li et al. (2016a); Zhang et al. (2016a); Huang et al. (2015); Xiong et al. (2015); Li et al. (2015); Zhang et al. (2017a); Li et al. (2016b); Arnold et al. (2016); Yang et al. (2015a, b); Wang et al. (2016) has attracted tremendous interest. In topological semimetals, the negative magnetoresistance is widely believed to be interpretable as a manifestation of the chiral anomaly, that is, the violation of chiral symmetry by quantum effects Adler (1969); Bell and Jackiw (1969); Nielsen and Ninomiya (1981). In topological insulators, the negative magnetoresistance is found to be related to the anomalous velocity induced by a nontrivial distribution of Berry curvature Dai et al. (2017). Because of the nontrivial mechanism behind each of the cases, a negative magnetoresistance observed in novel systems warrants a detailed study.
In a recent experiment on a single-crystalline silver chalcogenide material -Ag2Se Zhang et al. (2017b), a negative magnetoresistance has been observed when the magnetic field is parallel to the current. -Ag2Se is among the material candidates for a class of systems called Kramers Weyl semimetals, in which Weyl nodes are pinned to time-reversal invariant momenta (TRIMs) in the Brillouin zone because of time-reversal symmetry Bradlyn et al. (2016); Chang et al. (2018). The Kramers Weyl nodes generically appear in all chiral crystals, i.e., is crystals that lack any roto-inversion symmetries and thus have a sense of handedness Chang et al. (2018). Specifically, in symmorphic chiral crystals, every Kramers pair of bands at every TRIM is guaranteed to host a Weyl cone; while in non-symmorphic chiral crystals, it is true for a subset of TRIMs only, which however always includes the () point. -Ag2Se provides an instance of the latter case. In -Ag2Se, a negative magnetoresistance of about was observed at a magnetic field of about 9 T. At such a strong magnetic field, the system has entered the quantum limit, i.e., only the lowest Landau band crosses the Fermi energy. In the quantum limit, the magnetoresistance depends subtly on scattering mechanisms Lu et al. (2015); Goswami et al. (2015); Zhang et al. (2016b). Besides, since the Kramers Weyl nodes are protected from symmetries in chiral space groups, it provides a new platform for investigating the scattering mechanisms.
In this work, we present a theory for the longitudinal magnetoresistance in the quantum limit of a Kramers Weyl semimetal in strong parallel magnetic fields. We start with a generic model with one Kramers Weyl cone and use the standard Kubo formalism to calculate the conductivity, considering impurity scattering with screened Coulomb potentials and Gaussian potentials. We show that in the quantum limit the resistance has a dependence in the presence of impurities with screened Coulomb potential Dubinskaya (1969), and thus indeed gives rise to a negative magnetoresistance. In many Weyl semimetals that emerge from band inversion, Weyl nodes of opposite chirality are degenerate in energy, e.g. due to some mirror symmetry. This is generically not the case for the Kramers Weyl nodes in chiral crystals, so the inter-valley charge pumping and relaxation is absent in our calculations, suggesting that the negative magnetoresistance in the Weyl semimetal can exist without any apparent link to the chiral anomaly. Although magnetoresistance in the quantum limit has been systematically studied for a number of models and potentials Lu et al. (2015); Goswami et al. (2015); Zhang et al. (2016b), the case we study here has not been addressed before.
To justify our results, we will present the step-by-step details of the calculation. The paper is organized as follows. In Sec. II, we introduce a generic model for Kramers Weyl semimetals. In Secs. III and IV, we present the calculations for the transport time in the presence of the screened Coulomb scattering potential and longitudinal conductivity in the quantum limit, respectively. To justify the isotropic model and screened Coulomb scattering potential we used, we discuss the effect of anisotropy and Gaussian scattering potential in Secs. V and VI, respectively. Finally, we summarize and discuss the results in Sec. VII.
II Model and Landau Bands
A Kramers Weyl cone can be described by the effective model proposed in Zhang et al. (2017b),
(1) |
where we have suppressed the anisotropy of the model parameters and present in the original model. We will find that this isotropic model can capture the main physics of negative magnetoresistance. In contrast to the band-inversion Weyl cone, the parabolic term overwhelms the linear term and the energy difference between paired Weyl nodes can be much larger than the temperature scale in Kramers Weyl semimetals. It is reasonable to consider only one Weyl cone as the longitudinal conductivity is dominated by electrons near the Fermi surface.
The energy spectrum of this model has two bands
(2) |
which are schematically shown in Fig. 1 (b) in the plane . At , a Kramers Weyl node forms as band touches band . A similar Weyl cone is also present in BiTeI Murakami et al. (2003). Weyl nodes always come in pairs because of the fermion-doubling theorem Nielsen and Ninomiya (1983). For Weyl semimetals in which these pairs are (nearly) degenerate in energy, charges can be pumped from one Weyl cone to another Weyl cone of the opposite chirality in an external electric field, and both inter- and intra-cone scattering has to be considered. However, as this kind of energetic degeneracy of Weyl cones is not found in Kramers Weyl semimetals, we can assume that the intra-node scattering dominates. Later, we will show that a negative magnetoresistance can arise in this one-node Kramers Weyl semimetal in the presence of impurities with a screened Coulomb scattering potential.

In a magnetic field, the energy spectrum is quantized into a set of one-dimensional (1D) bands of Landau levels. Because the present model is isotropic, we can assume that the magnetic field is applied along the direction, that is, . Along the direction of the magnetic field the wave vector is a good quantum number, so the 1D Landau bands disperse with . We adopt the Landau gauge in which the vector potential is . Under the Peierls replacement, the wave vector becomes , and the Hamiltonian in the applied magnetic field becomes
(3) |
The chosen vector potential breaks the translational symmetry along the -direction but not along the - or -direction, and are still good quantum numbers. Introducing the ladder operators Shen et al. (2004); Shen (2004)
(4) | |||||
(5) |
we can replace the wave vectors , , , where is the magnetic length, , and we have defined . Then the Hamiltonian in Eq. (3) becomes
(6) |
where and . With the trial wave functions for , and for , where indexes the Hermite polynomial wave functions given below, the eigenenergies can be obtained from the secular equation
(7) |
for ; and for , where . The eigenenergies are found as
where . This gives the spectrum of a set of Landau energy bands ( as band index), which is shown in Fig. 1 (c). The eigenstates found for are
(9) | |||||
and the eigenstate for is
(10) |
where . The wave functions in real-space are found to be
(11) |
for and
(12) |
for . Here, , and are the lengths of the sample in the and directions, respectively, , is the guiding center and are the Hermite polynomials. does not appear in the energy spectrum explicitly, because different give rise to the Landau degeneracy per unit area in the x-y plane.
In what follows, we only concentrate on the quantum limit, which means only the lowest () Landau band crosses the Fermi energy [see Fig. 1 (b)]. The analytical solution of the Landau bands allows us to determine the analytical solution to the longitudinal magnetoresistance in the quantum limit of the Kramers Weyl cone.
III Screened Coulomb Scattering Potential and Transport Time
To calculate the longitudinal conductivity, we need to know the transport time of carriers Zhang et al. (2016b); Goswami et al. (2015), which sensitively depends on the impurity scattering potential. Following Abrikosov’s treatment when exploring the linear magnetoresistance in the quantum limit of Ag2Se under perpendicular magnetic fields Abrikosov (1998), we adopt a screened Coulomb scattering potential for the impurities,
(13) |
with
(14) |
where is the dielectric constant, is the electron charge, and is the screening length. The screened Coulomb potential is justified because excess silver atoms form clusters doping the rest of the materialAbrikosov (1998). This potential is different from the ionic potential under the Thomas-Fermi approximation Goswami et al. (2015). Following the standard random phase approximation Abrikosov (1998); Bruus and Flensberg (2004),
(15) | |||||
where , are the Matsubara frequencies of fermions, is the Fermi energy, and the energy spectrum of the Landau band has been given in Eq. (II). After summing over the Matsubara frequencies, we have
(16) |
Substituting this into Eq. (15), we have
(17) |
where is the Fermi velocity of the Landau band. To calculate the transport time, we need to calculate the scattering matrix elements,
(18) |
where is defined in Eq. (13), and are indices for the Landau bands. In the quantum limit, due to the strong magnetic field, the energy spectrum has split into 1D Landau bands, and the spacing between the Landau bands is large. Moreover, because the Fermi energy crosses only the Landau band, the occupation of electrons for bands vanishes and the Landau band is partially filled as shown in Fig. 1 (c). Therefore, we only need to take into account the impurity scatterings between Landau states. Using Eq. (13)
(19) | |||||
where the integral
describes the probability amplitude of charge carriers scattered from state to by an impurity located at position in real space. We may then write the impurity scattering potential as
(20) |
where the Fourier transform is given by
(21) |
Substituting this into Eq. (III) and integrating along the and directions gives
(22) | |||||
where is the Kronecker symbol. In terms of , the absolute value squared of the scattering matrix element between the states of the Landau band can be written as
(23) |
After averaging over impurity configurations we obtain
where we have used the random impurity approximation
(25) |
where is the density of impurities over the full sample. Using the energy dispersion of the Landau bands in Eq. (II), the velocity is found as
Measured from the band bottom at , the carrier density of the Landau band is given by
(27) |
Combining the above two relations, we have
(28) |
so is proportional to for a fixed .
The transport time of electrons in the Landau band is defined as
As shown in Fig. 1 (b), the parabolic Landau band crosses the Fermi energy at two points of wave vector, denoted as , from which we can simplify Eq. (III) as
(30) |
where is an extra correction factor introduced to avoid the van Hove singularity at the band edge of the 1D Landau band Lu et al. (2015). Equation (30) suggests that only the backscattering survives, i.e., . Substituting Eq. (23) into Eq. (30), and considering in strong magnetic fields, we have
(31) |
By using Eq. (17), we arrive at the transport time of the band in strong magnetic fields
(32) |
IV Longitudinal Magnetoconductivity
With the transport time, we are ready to calculate the longitudinal conductivity of the band. Along the -direction, the semiclassical Drude conductivity can be calculated as
(33) |
where is the electron charge, is the sample volume, is the velocity along the -direction for a state with wave vector in the Landau band,
(34) |
are the retarded/advanced Green’s functions with the lifetime of a state with wave vector and in the Landau band. In the diffusive regime, can be replaced by
(35) |
Changing the summations into integrals through
(36) |
where the integral covers the entire Brillouin zone and the integral is confined by the degeneracy of the Landau levels, the conductivity can be expressed as
(37) |
By using Eq. (35), we have
(38) |
where the Landau degeneracy has already been taken into account. Now using the transport time in Eq. (32), we finally arrive at
(39) |
Considering the in Eq. (39) and the velocity along z-direction obtained in Eq. (28), shows a positive linear dependence on magnetic field, i.e.,
(40) |
In a parallel magnetic field, there is no Hall effect, thus the resistivity is the inverse of the conductivity. Hence,
(41) |
which means that the resistivity drops with increasing magnetic field. In other words, we find a negative longitudinal magnetoresistance in the quantum limit of the Kramers Weyl cone in the presence of a screened Coulomb impurity scattering potential.
V Gaussian scattering potential
Another common choice of the scattering potential is the Gaussian one. We can show that there is no negative magnetoresistance in the presence of the Gaussian scattering potential in the quantum limit of the Kramers Weyl cone. The Hamiltonian of the Gaussian scattering reads Goswami et al. (2015); Zhang et al. (2016b)
(42) |
or in terms of its Fourier transform
(43) |
where
(44) |
and measures the scattering strength at . A major difference here is that the range of the potential is not a function of the magnetic field. As shrinks to zero, the potential reduces to a delta potential.
Following the same procedure that was used to obtain Eq. (32), the transport time for the Gaussian scattering potential is found to be
(45) |
Inserting Eq. (45) into (38), one obtains the conductivity
(46) |
According to Eqs. (III) and (28), , so will decrease with increasing , which cannot give a negative magnetoresistance. We conclude that the screened Coulomb scattering potential is an essential ingredient for a negative magnetoresistance.
VI Anisotropic Kramers Weyl cone

The original model for the Kramers Weyl cone in -Ag2Se is anisotropic. Now we show that the anisotropy does not qualitatively change the negative magnetoresistance. The original model reads Zhang et al. (2017b)
(47) |
where the model parameters eVÅ2, eVÅ2, eVÅ2, eVÅ, eVÅ, eVÅ. Because of the anisotropy, there is no analytical solution for the Landau bands. We numerically solve the energy spectrum of the Landau bands. Figure 2 shows that the structure of the Landau spectrum does not change qualitatively in the presence of the anisotropy, so one can also expect that a negative magnetoresistance is obtained in the anisotropic case. This justifies our calculation using the isotropic model.
VII Conclusion and Discussion
In conclusion, we presented a theory for the negative magnetoresistance observed in the quantum limit of -Ag2Se, a paradigmatic Kramers-Weyl semimetal, in parallel magnetic fields Zhang et al. (2017b). It requires several ingredients: (1) a Kramers Weyl cone; (2) impurities with a screened Coulomb potential; (3) a fixed carrier density; (4) the quantum limit under parallel magnetic fields. These conditions naturally exist in -Ag2Se.
The negative magnetoresistance has been previously studied in the quantum limit of the band-inversion Weyl semimetals in parallel magnetic fields Lu et al. (2015); Goswami et al. (2015); Zhang et al. (2016b). In the presence of the charge-neutral Gaussian potential, the conductivity is linearly proportional to the magnetic field Goswami et al. (2015); Zhang et al. (2016b). In the ionic potential under the Thomas-Fermi approximation, the conductivity is proportional to in all cases Goswami et al. (2015). The novel mechanism for the negative magnetoresistance that we present in this work has not been covered in these previous studies. In particular, here the negative magnetoresistance is not related to the chiral anomaly, which has been used to explain the effect in band-inversion induced Weyl semimetals.
In band-inversion Weyl semimetals, two Weyl cones of opposite chirality have to appear in pairs. In a strong magnetic field, the Landau bands from the two cones have opposite velocities inherited from the chirality of the Weyl cones. In contrast, in Kramers Weyl semimetals, the energy difference between Kramers Weyl nodes of opposite chirality can be very large, because, e.g., one is located at the point and the other at the Brillouin zone corner. As the Fermi energy crosses the vicinity of one of the Weyl nodes, the contributions to transport coming from the other Weyl node can be safely ignored. For instance, the band of only one Weyl node in Fig. 1 (b) has no sense of the inherited chirality, although it looks similar to the quantum limit of the band-inversion Weyl semimetal. Therefore, a negative magnetoresistance of the single Kramers Weyl cone has nothing to do with the chiral anomaly.
According to the above result for the hole carriers, the key ingredients of the negative magnetoresistance are the parabolic dispersion of the 0th Landau band and the screened Coulomb scattering potential. According to the DFT band structure in Ref. [26], there is another electron pocket on the Fermi surface and its dispersion looks quite conventional. The conventional electron pocket is expected to have the parabolic dispersion for the 0th Landau band. Therefore, along with the same screened Coulomb scattering potential, the electron pocket is also expected to give the negative magnetoresistance.
Acknowledgements.
We thank helpful discussions with Cheng-Long Zhang, Su-Yang Xu, and Shuang Jia. This work was supported by Guangdong Innovative and Entrepreneurial Research Team Program (Grant No. 2016ZT06D348), the National Key R & D Program (Grant No. 2016YFA0301700), the National Natural Science Foundation of China (Grants No. 11525417 and No. 11574127), and the Science, Technology and Innovation Commission of Shenzhen Municipality (Grant No. ZDSYS20170303165926217 and JCYJ20170412152620376). F.S. and T.N. acknowledge support from the Swiss National Science Foundation (Grant No. 200021_169061) and from the European Unions Horizon 2020 research and innovation program (ERC-StG-Neupert-757867-PARATOP). X.W. also acknowledges support by the Fundamental Research Funds for the Central Universities (No. 020414380085).References
- Argyres and Adams (1956) P. N. Argyres and E. N. Adams, “Longitudinal magnetoresistance in the quantum limit,” Phys. Rev. 104, 900 (1956).
- Lee and Ramakrishnan (1985) Patrick A. Lee and T. V. Ramakrishnan, “Disordered electronic systems,” Rev. Mod. Phys. 57, 287–337 (1985).
- Wang et al. (2012) Jian Wang, Handong Li, Cuizu Chang, Ke He, JoonSue Lee, Haizhou Lu, Yi Sun, Xucun Ma, Nitin Samarth, Shunqing Shen, Qikun Xue, Maohai Xie, and Moses H.W. Chan, “Anomalous anisotropic magnetoresistance in topological insulator films,” Nano Research 5, 739–746 (2012).
- He et al. (2013) H. T. He, H. C. Liu, B. K. Li, X. Guo, Z. J. Xu, M. H. Xie, and J. N. Wang, “Disorder-induced linear magnetoresistance in (221) topological insulator Bi2Se3 films,” Appl. Phys. Lett. 103, 031606 (2013).
- Wiedmann et al. (2016) S. Wiedmann, A. Jost, B. Fauqué, J. van Dijk, M. J. Meijer, T. Khouri, S. Pezzini, S. Grauer, S. Schreyeck, C. Brüne, H. Buhmann, L. W. Molenkamp, and N. E. Hussey, “Anisotropic and strong negative magnetoresistance in the three-dimensional topological insulator Bi2Se3,” Phys. Rev. B 94, 081302 (2016).
- Wang et al. (2015) Li-Xian Wang, Yuan Yan, Liang Zhang, Zhi-Min Liao, Han-Chun Wu, and Da-Peng Yu, “Zeeman effect on surface electron transport in topological insulator Bi2Se3 nanoribbons,” Nanoscale 7, 16687–16694 (2015).
- Breunig et al. (2017) Oliver Breunig, Zhiwei Wang, A. A. Taskin, Jonathan Lux, Achim Rosch, and Yoichi Ando, “Gigantic negative magnetoresistance in the bulk of a disordered topological insulator,” Nature Commun. 8, 15545 (2017).
- Assaf et al. (2017) B. A. Assaf, T. Phuphachong, E. Kampert, V. V. Volobuev, P. S. Mandal, J. Sánchez-Barriga, O. Rader, G. Bauer, G. Springholz, L. A. de Vaulchier, and Y. Guldner, “Negative longitudinal magnetoresistance from the anomalous landau level in topological materials,” Phys. Rev. Lett. 119, 106602 (2017).
- Kim et al. (2013) H. J. Kim, K. S. Kim, J. F. Wang, M. Sasaki, N. Satoh, A. Ohnishi, M. Kitaura, M. Yang, and L. Li, “Dirac versus Weyl fermions in topological insulators: Adler-Bell-Jackiw anomaly in transport phenomena,” Phys. Rev. Lett. 111, 246603 (2013).
- Kim et al. (2014) Ki-Seok Kim, Heon-Jung Kim, and M. Sasaki, “Boltzmann equation approach to anomalous transport in a Weyl metal,” Phys. Rev. B 89, 195137 (2014).
- Li et al. (2016a) 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, “Chiral magnetic effect in ,” Nature Phys. 12, 550–554 (2016a).
- Zhang et al. (2016a) C. L. Zhang, S. Y. Xu, I. Belopolski, Z. Yuan, Z. Lin, B. Tong, N. Alidoust, C. C. Lee, S. M. Huang, T. R. Chang, H. T. Jeng, H. Lin, M. Neupane, D. S. Sanchez, H. Zheng, G. Bian, J. Wang, C. Zhang, H. Z. Lu, S. Q. Shen, T. Neupert, M. Z. Hasan, and S. Jia, “Signatures of the Adler-Bell-Jackiw chiral anomaly in a Weyl Fermion semimetal,” Nat. Commun. 7, 10735 (2016a).
- Huang et al. (2015) X. C. Huang, L. X. Zhao, Y. J. Long, P. P. Wang, D. Chen, Z. H. Yang, H. Liang, M. Q. Xue, H. M. Weng, Z. Fang, X. Dai, and G. F. Chen, “Observation of the chiral-anomaly-induced negative magnetoresistance in 3D Weyl semimetal TaAs,” Phys. Rev. X 5, 031023 (2015).
- Xiong et al. (2015) J. Xiong, S. K. Kushwaha, T. Liang, J. W. Krizan, M. Hirschberger, W. Wang, R. J. Cava, and N. P. Ong, “Evidence for the chiral anomaly in the Dirac semimetal ,” Science 350, 413 (2015).
- Li et al. (2015) C. Z. Li, L. X. Wang, H. W. Liu, J. Wang, Z. M. Liao, and D. P. Yu, “Giant negative magnetoresistance induced by the chiral anomaly in individual nanowires,” Nature Commun. 6, 10137 (2015).
- Zhang et al. (2017a) Cheng Zhang, Enze Zhang, Weiyi Wang, Yanwen Liu, Zhi-Gang Chen, Shiheng Lu, Sihang Liang, Junzhi Cao, Xiang Yuan, Lei Tang, Qian Li, Chao Zhou, Teng Gu, Yizheng Wu, Jin Zou, and Faxian Xiu, “Room-temperature chiral charge pumping in Dirac semimetals,” Nat. Commun. 8, 13741 (2017a).
- Li et al. (2016b) H. Li, H. T. He, H. Z. Lu, H. C. Zhang, H. C. Liu, R. Ma, Z. Y. Fan, S. Q. Shen, and J. N. Wang, “Negative magnetoresistance in Dirac semimetal ,” Nature Commun. 7, 10301 (2016b).
- Arnold et al. (2016) Frank Arnold, Chandra Shekhar, Shu-Chun Wu, Yan Sun, Ricardo Donizeth dos Reis, Nitesh Kumar, Marcel Naumann, Mukkattu O. Ajeesh, Marcus Schmidt, Adolfo G. Grushin, Jens H. Bardarson, Michael Baenitz, Dmitry Sokolov, Horst Borrmann, Michael Nicklas, Claudia Felser, Elena Hassinger, and Binghai Yan, “Negative magnetoresistance without well-defined chirality in the Weyl semimetal TaP,” Nature Commun. 7, 11615 (2016).
- Yang et al. (2015a) X. J. Yang, Y. P. Liu, Z. Wang, Y. Zheng, and Z. A. Xu, “Chiral anomaly induced negative magnetoresistance in topological Weyl semimetal NbAs,” arXiv:1506.03190 (2015a).
- Yang et al. (2015b) Xiaojun Yang, Yupeng Li, Zhen Wang, Yi Zhen, and Zhu-An Xu, “Observation of negative magnetoresistance and nontrivial Berry’s phase in 3D Weyl semi-metal NbAs,” arXiv:1506.02283 (2015b).
- Wang et al. (2016) Huichao Wang, Chao-Kai Li, Haiwen Liu, Jiaqiang Yan, Junfeng Wang, Jun Liu, Ziquan Lin, Yanan Li, Yong Wang, Liang Li, David Mandrus, X. C. Xie, Ji Feng, and Jian Wang, “Chiral anomaly and ultrahigh mobility in crystalline HfTe5,” Phys. Rev. B 93, 165127 (2016).
- Adler (1969) S. L. Adler, “Axial-vector vertex in spinor electrodynamics,” Phys. Rev. 177, 2426–2438 (1969).
- Bell and Jackiw (1969) J. S. Bell and R. Jackiw, “A PCAC puzzle: in the -model,” Il Nuovo Cimento A 60, 47–61 (1969).
- Nielsen and Ninomiya (1981) H. B. Nielsen and M. Ninomiya, “Absence of neutrinos on a lattice: (i). Proof by homotopy theory,” Nuclear Physics B 185, 20 – 40 (1981).
- Dai et al. (2017) Xin Dai, Z. Z. Du, and Hai-Zhou Lu, “Negative magnetoresistance without chiral anomaly in topological insulators,” Phys. Rev. Lett. 119, 166601 (2017).
- Zhang et al. (2017b) Cheng-Long Zhang, Frank Schindler, Haiwen Liu, Tay-Rong Chang, Su-Yang Xu, Guoqing Chang, Wei Hua, Hua Jiang, Zhujun Yuan, Junliang Sun, Horng-Tay Jeng, Hai-Zhou Lu, Hsin Lin, M. Zahid Hasan, X. C. Xie, Titus Neupert, and Shuang Jia, “Ultraquantum magnetoresistance in the kramers-weyl semimetal candidate -Ag2Se,” Phys. Rev. B 96, 165148 (2017b).
- Bradlyn et al. (2016) Barry Bradlyn, Jennifer Cano, Zhijun Wang, M. G. Vergniory, C. Felser, R. J. Cava, and B. Andrei Bernevig, “Beyond Dirac and Weyl fermions: Unconventional quasiparticles in conventional crystals,” Science 353, 045203 (2016).
- Chang et al. (2018) Guoqing Chang, Benjamin J Wieder, Frank Schindler, Daniel S Sanchez, Ilya Belopolski, Shin-Ming Huang, Bahadur Singh, Di Wu, Tay-Rong Chang, Titus Neupert, et al., “Topological quantum properties of chiral crystals,” Nature materials 17, 978 (2018).
- Lu et al. (2015) H. Z. Lu, S. B. Zhang, and S. Q. Shen, “High-field magnetoconductivity of topological semimetals with short-range potential,” Phys. Rev. B 92, 045203 (2015).
- Goswami et al. (2015) P. Goswami, J. H. Pixley, and S. Das Sarma, “Axial anomaly and longitudinal magnetoresistance of a generic three-dimensional metal,” Phys. Rev. B 92, 075205 (2015).
- Zhang et al. (2016b) Song-Bo Zhang, Hai-Zhou Lu, and Shun-Qing Shen, “Linear magnetoconductivity in an intrinsic topological Weyl semimetal,” New J. Phys. 18, 053039 (2016b).
- Dubinskaya (1969) Y. L. Dubinskaya, “Negative longitunidal magnetoresistance associated with scattering from ionized impurity centers,” Sov. Phys. JETP 29, 436 (1969).
- Murakami et al. (2003) Shuichi Murakami, Naoto Nagaosa, and Shou-Cheng Zhang, “Dissipationless quantum spin current at room temperature,” Science 301, 1348–1351 (2003).
- Nielsen and Ninomiya (1983) H.B. Nielsen and Masao Ninomiya, “The Adler-Bell-Jackiw anomaly and Weyl fermions in a crystal,” Physics Letters B 130, 389 – 396 (1983).
- Shen et al. (2004) S. Q. Shen, M. Ma, X. C Xie, and F. C. Zhang, “Resonant spin Hall conductance in two-dimensional electron systems with a Rashba interaction in a perpendicular magnetic field,” Phys. Rev. Lett. 92, 256603 (2004).
- Shen (2004) Shun-Qing Shen, “Spin Hall effect and berry phase in two-dimensional electron gas,” Phys. Rev. B 70, 081311(R) (2004).
- Abrikosov (1998) A. A. Abrikosov, “Quantum magnetoresistance,” Phys. Rev. B 58, 2788–2794 (1998).
- Bruus and Flensberg (2004) Henrik Bruus and Karsten Flensberg, Many-Body Quantum Theory in Condensed Matter Physics: An Introduction (Oxford University Press, Oxford, UK, 2004).