This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Magnetic field driven Lifshitz transition and one-dimensional Weyl nodes in three-dimensional pentatellurides

Zhigang Cai School of Science, Jiangnan University, Wuxi 214122, China    Yi-Xiang Wang wangyixiang@jiangnan.edu.cn School of Science, Jiangnan University, Wuxi 214122, China School of Physics and Electronics, Hunan University, Changsha 410082, China
Abstract

Recent experiments reported that the magnetic field can drive the Lifshitz transition and one-dimensional (1D) Weyl nodes in the quantum limit of three-dimensional pentatellurides, as they own low carrier densities and can achieve the extreme quantum limit at a low magnetic field. In this paper, we will investigate the conditions for the existence of the 1D Weyl nodes and their dc transport properties. We find that in the strong topological insulator (TI) phase of ZrTe5, the formation of the Weyl nodes depends heavily on the carrier density; while in the weak TI phase of HfTe5, the Weyl nodes are more likely to appear. These behaviors are attributed to the fact that in the strong and weak TI phases, the zeroth Landau levels exhibit opposite evolutions with the magnetic field. Moreover, the signatures of the critical fields that characterize the distinct behaviors of the system can be directly captured in the conductivities.

I Introduction

When a perpendicular magnetic field is applied on a two-dimensional (2D) or three-dimensional (3D) electronic system, the electrons will be confined to move on curved orbits due to the Lorentz force L.D.Landau , and, consequently, a set of discrete energy levels, i.e., the Landau levels (LLs) will form. If the magnetic field is strong and a few LLs are occupied, the system lies in the quantum Hall regime K.V.Klitzing ; D.C.Tsui ; D.Yoshioka . Further increasing the magnetic field, if all electrons occupy only the zeroth LLs, the extreme quantum limit will be reached. For most semiconductor materials, realizing the quantum limit seems impossible as the required magnetic field strength would be inaccessible in the experiment.

In the past 20 years, the developments of topological materials have provided insights to investigate the interactions between the magnetic field and electronic systems M.Z.Hasan ; X.L.Qi ; Lv2021 ; J.A.Sobota . Among the emerging topological materials, 3D pentatellurides Q.Li ; T.Liang2018 including ZrTe5 and HfTe5 are representatives and can exhibit a number of desirable features: (i) the low-energy bands are well described by the effective noninteracting models that are topologically nontrivial H.Weng ; (ii) they have narrow gaps and small band masses, leading to the sizable cyclotron frequencies even for a weak magnetic field E.Martino ; Z.G.Chen ; and (iii) the crystal sample owns a high purity, with the electron mobility reaching the order of 10510^{5} cm2V-1s-1, and the low carrier density of about 1016101710^{16}\sim 10^{17} cm-3 at a low temperature F.Tang ; E.Martino ; P.Wang ; S.Galeski2020 . Therefore, the quantum limit may be reached in 3D pentatellurides at a weak magnetic field, which makes its study possible.

Recent experiments in pentatellurides reported that the system can indeed reach the quantum limit at the magnetic field of several Tesla S.Galeski2022 ; W.Wu . More importantly, there exists a magnetic field driven Lifshitz transition to the 1D Weyl regime: the crossing points of the zeroth LLs can be regarded as 1D Weyl nodes, because the band structure and spin texture are analogous to those of the Weyl nodes that are formed by Bloch band crossings X.Wan ; S.Y.Xu ; Lv2015 . In this regime, since the Fermi surface is perfectly nested, the electronic states are unstable to the interactions G.Gruner ; F.Qin , which makes it a good platform to investigate the strongly correlated electronic states. In ZrTe5, the Lifshitz transition was characterized by the combined dc electric transport and ultrasound measurements S.Galeski2022 . Moreover, the chemical potential was found to meet and cross the Weyl nodes, resulting in the enhancement of the hole band occupation and the weakening of the electron band occupation S.Galeski2022 . In HfTe5, the Lifshitz transition was demonstrated by the magneto-infrared spectroscopy W.Wu , where a highly unusual reduction of optical activity and the variation of the accompanying resonant peaks were observed when the chemical potential crosses the zeroth LLs. Based on these observations, we propose the following questions: Do the Weyl nodes always exist in pentatellurides through modulating the magnetic field? If not, what are the conditions for their existence?

In this paper, we will systematically explore the above questions in theory. Although the ground states of ZrTe5 and HfTe5 depend heavily on the specific experimental conditions, such as the growth method, temperature B.Xu , and so on, it is widely believed that their ground states are located close to the phase boundary between a strong topological insulator (TI) and a weak TI H.Weng . In the strong TI, the band inversions can occur in all three directions; whereas in the weak TI, the band inversion occurs only in one direction or in one plane. More importantly, the two phases can be characterized by the Z2 topological invariant index L.Fu2007a ; L.Fu2007b . Here we will take the ground state in ZrTe5 and HfTe5 as the strong and weak TI, respectively, which are supported by many experimental studies Z.G.Chen ; G.Manzoni ; Y.Jiang ; J.Wang ; W.Wu .

Under the condition of fixed carrier density, we calculate the chemical potential as a function of the magnetic field BB. The chemical potential variation, together with the LL movements, will lead to several critical fields that characterize the distinct behaviors of the system. We also study the dc transport property in the quantum limit by calculating the longitudinal conductivity σxx\sigma_{xx} and Hall conductivity σxy\sigma_{xy} by using the Kubo-Streda formula. We find that in the strong TI phase of ZrTe5, the magnetic field can drive the zeroth LLs from crossing to be separated, thus the appearance of the 1D Weyl nodes depends heavily on the carrier density; whereas in the weak TI phase of HfTe5, the magnetic field drives the zeroth LLs from separated to cross each other, which makes the 1D Weyl nodes more likely to appear. The critical fields can be captured by the peaks of σxx\sigma_{xx} and the vanishing σxy\sigma_{xy}. Moreover, the linear dependence of σxy\sigma_{xy} on the inverse magnetic field, σxyB1\sigma_{xy}\sim B^{-1}, holds in the quantum oscillation regime but does not in the quantum limit. We attribute this observation to the g2g_{2}-spin Zeeman term that breaks the particle-hole symmetry. Our paper can help one better understand the quantum limit and the 1D Weyl nodes that are driven by the magnetic field in 3D pentatelluride experiments.

II Model and Methods

We use the effective 𝒌𝒑\bm{k}\cdot\bm{p} model to describe the low-energy excitations in 3D pentatellurides. In the four-component basis (|+,|,|+,|,)T\begin{pmatrix}|+,\uparrow\rangle&|-,\uparrow\rangle&|+,\downarrow\rangle&|-,\downarrow\rangle\end{pmatrix}^{T}, the Hamiltonian is written as (=1\hbar=1) E.Martino ; Y.Jiang ; Z.Rukelj ; J.Wang

H(𝒌)=\displaystyle H(\bm{k})= v(kxσzτx+kyIτy)+vzkzσxτx\displaystyle v(k_{x}\sigma_{z}\otimes\tau_{x}+k_{y}I\otimes\tau_{y})+v_{z}k_{z}\sigma_{x}\otimes\tau_{x}
+[Mξ(kx2+ky2)ξzkz2]Iτz.\displaystyle+[M-\xi(k_{x}^{2}+k_{y}^{2})-\xi_{z}k_{z}^{2}]I\otimes\tau_{z}. (1)

Here σ\sigma and τ\tau are the Pauli matrices acting on the spin and orbit degrees of freedom, respectively. vv and vzv_{z} are the Fermi velocities, ξ\xi and ξz\xi_{z} are the band inversion parameters, and MM denotes the Dirac mass. When taking vz=0v_{z}=0, H(𝒌)H(\bm{k}) becomes decoupled, and the upspin and downspin are good quantum numbers. In the 3D system, there exist the inversion symmetry (IS) 1H(𝒌)=H(𝒌){\cal I}^{-1}H(-\bm{k}){\cal I}=H(\bm{k}) with the operator =τz{\cal I}=\tau_{z}, the time-reversal symmetry (TRS) 𝒯1H0(𝒌)𝒯=H0(𝒌){\cal T}^{-1}H_{0}(-\bm{k}){\cal T}=H_{0}(\bm{k}) with 𝒯=iσyK{\cal T}=i\sigma_{y}K and KK being the complex conjugation, the particle-hole symmetry (PHS) 𝒫1H(𝒌)𝒫=H(𝒌){\cal P}^{-1}H(\bm{k}){\cal P}=-H(\bm{k}) with 𝒫=iσyτx{\cal P}=i\sigma_{y}\tau_{x}, as well as the chiral symmetry (CS) 𝒞1H(𝒌)𝒞=H(𝒌){\cal C}^{-1}H({\bm{k}}){\cal C}=-H(-{\bm{k}}) with 𝒞=σyτy{\cal C}=\sigma_{y}\otimes\tau_{y}. Thus according to the Altland and Zirnbauer notations, the topological states described by H(𝒌)H(\bm{k}) belong to the chiral symplectic class CII A.P.Schnyner ; C.K.Chiu .

When a uniform magnetic field 𝑩=B𝒆z\bm{B}=B\bm{e}_{z} acts on the 3D system, the 1D Landau bands will form, with the dispersions along the magnetic field direction. To solve the LLs, we choose the vector potential in the Landau gauge as 𝑨=By𝒆x\bm{A}=-By\bm{e}_{x} and make the Peierls substitution 𝝅=𝒌e𝑨\bm{\pi}=\bm{k}-e\bm{A}. Then we define the raising and lowering operators a=lB2(πx+iπy)a^{\dagger}=\frac{l_{B}}{\sqrt{2}}(\pi_{x}+i\pi_{y}) and a=lB2(πxiπy)a=\frac{l_{B}}{\sqrt{2}}(\pi_{x}-i\pi_{y}), with [a,a]=1[a,a^{\dagger}]=1 and the magnetic length lB=1eB=25.6Bl_{B}=\frac{1}{\sqrt{eB}}=\frac{25.6}{\sqrt{B}} nm. Besides the orbital effect, the magnetic field can also cause the spin Zeeman splitting, which is described as

HZ=12g1μBBσz12g2μBBσzτz.\displaystyle H_{Z}=-\frac{1}{2}g_{1}\mu_{B}B\sigma_{z}-\frac{1}{2}g_{2}\mu_{B}B\sigma_{z}\tau_{z}. (2)

Here μB\mu_{B} denotes the Bohr magneton, and g1g_{1} and g2g_{2} are the Landé gg-factors. Since we have 𝒫1σz𝒫=σz{\cal P}^{-1}\sigma_{z}{\cal P}=-\sigma_{z} and 𝒫1σzτz𝒫=σzτz{\cal P}^{-1}\sigma_{z}\tau_{z}{\cal P}=\sigma_{z}\tau_{z}, the PHS is preserved by the g1g_{1} term but is broken by the g2g_{2} term.

With the trial wavefunction ψn=(cn1|n,cn2|n1,cn3|n1,cn4|n)T\psi_{n}=(c_{n}^{1}|n\rangle,c_{n}^{2}|n-1\rangle,c_{n}^{3}|n-1\rangle,c_{n}^{4}|n\rangle)^{T}, where the harmonic oscillator state |n|n\rangle is defined by aa|n=n|na^{\dagger}a|n\rangle=n|n\rangle and cn1,,4c_{n}^{1,\cdots,4} are the coefficients, we obtain the energies for the zeroth and n1n\geq 1 LLs L.You

ε0λ(kz)=λ(Mξzkz2ξlB2+12g1μBB)+12g2μBB,\displaystyle\varepsilon_{0\lambda}(k_{z})=-\lambda\big{(}M-\xi_{z}k_{z}^{2}-\frac{\xi}{l_{B}^{2}}+\frac{1}{2}g_{1}\mu_{B}B\big{)}+\frac{1}{2}g_{2}\mu_{B}B, (3)
εnsλ(kz)=s[(M2nξlB2ξzkz2λ2g2μBB)2+2nv2lB2]12\displaystyle\varepsilon_{ns\lambda}(k_{z})=s\Big{[}\big{(}M-\frac{2n\xi}{l_{B}^{2}}-\xi_{z}k_{z}^{2}-\frac{\lambda}{2}g_{2}\mu_{B}B\big{)}^{2}+\frac{2nv^{2}}{l_{B}^{2}}\Big{]}^{\frac{1}{2}}
+λ(ξlB212g1μBB),\displaystyle\qquad\qquad\quad+\lambda\big{(}\frac{\xi}{l_{B}^{2}}-\frac{1}{2}g_{1}\mu_{B}B\big{)}, (4)

respectively. Here the index s=±1s=\pm 1 denotes the conduction/valence band, and λ=±1\lambda=\pm 1 characterizes the upspin/downspin branch.

With the density of states (DOS) D(ε)D(\varepsilon), the chemical potential μ\mu is determined by the carrier density n0n_{0} as B.Fu ; C.Wang

n0=\displaystyle n_{0}= 0𝑑εD(ε)f(εμ)+0𝑑εD(ε)[f(εμ)1],\displaystyle\int_{0}^{\infty}d\varepsilon D(\varepsilon)f(\varepsilon-\mu)+\int_{-\infty}^{0}d\varepsilon D(\varepsilon)[f(\varepsilon-\mu)-1], (5)

where f(x)=1exp(βx)+1f(x)=\frac{1}{\text{exp}(\beta x)+1} is the Fermi-Dirac distribution function with β=1kBT\beta=\frac{1}{k_{B}T} the inverse temperature, and the charge neutrality is taken at the zero energy.

The contribution of the zeroth LL to the DOS is given analytically as

D0λ(ε)=g2π[ξz(C+λελ2g2μBB)]12,\displaystyle D_{0\lambda}(\varepsilon)=\frac{g}{2\pi}\Big{[}\xi_{z}\big{(}C+\lambda\varepsilon-\frac{\lambda}{2}g_{2}\mu_{B}B\big{)}\Big{]}^{-\frac{1}{2}}, (6)

where g=12πlB2g=\frac{1}{2\pi l_{B}^{2}} is the LL degeneracy in the xx-yy plane and can be denoted as the uniform DOS, C=MξlB2+12g1μBBC=M-\frac{\xi}{l_{B}^{2}}+\frac{1}{2}g_{1}\mu_{B}B. We see that D0λD_{0\lambda} exhibits the square-root singularity, leading to the asymmetric peaks at ε=λC+12g2μBB\varepsilon=-\lambda C+\frac{1}{2}g_{2}\mu_{B}B.

We further study the dc transport properties of the system, as the conductivities or resistivities can be measured directly in the experiments to help judge the electronic states. The conductivity tensors are calculated by using the Kubo-Streda formula L.Smrcka ; G.D.Mahan ,

σαβ=\displaystyle\sigma_{\alpha\beta}= 12πV𝒌dεf(εμ)[Tr(JαdGRdεJβ(GAGR)\displaystyle\frac{1}{2\pi V}\sum_{\bm{k}}\int_{-\infty}^{\infty}d\varepsilon f(\varepsilon-\mu)\Big{[}\text{Tr}\Big{(}J_{\alpha}\frac{dG^{R}}{d\varepsilon}J_{\beta}(G^{A}-G^{R})
Jα(GAGR)JβdGAdε)],\displaystyle-J_{\alpha}(G^{A}-G^{R})J_{\beta}\frac{dG^{A}}{d\varepsilon}\Big{)}\Big{]}, (7)

where VV is the volume of the 3D system, Jα=eHkαJ_{\alpha}=e\frac{\partial H}{\partial k_{\alpha}} is the current density operator along the α\alpha direction, and GR/A(ε,η)=(εH±iη)1G^{R/A}(\varepsilon,\eta)=(\varepsilon-H\pm i\eta)^{-1} is the retarded/advanced Green’s function, with η\eta representing the LL linewidth broadening that is introduced phenomenologically to represent the impurity scatterings and will be taken as a constant for simplicity. In the following, we focus on the zero temperature.

With the help of the LL energies and wavefunctions, the longitudinal conductivity σxx\sigma_{xx} and Hall conductivity σxy\sigma_{xy} can be derived directly. The selection rules nn±1n\rightarrow n\pm 1 are determined from the nonvanishing matrix element of the current density, and there is no limit on the ss and λ\lambda index. We note that the conductivity components satisfy the following relations:

σxx(nsλn+1,sλ)=σxx(n+1,sλnsλ),\displaystyle\sigma_{xx}(ns\lambda\rightarrow n+1,s^{\prime}\lambda^{\prime})=\sigma_{xx}(n+1,s^{\prime}\lambda^{\prime}\rightarrow ns\lambda), (8)
σxy(nsλn+1,sλ)=σxy(n+1,sλnsλ),\displaystyle\sigma_{xy}(ns\lambda\rightarrow n+1,s^{\prime}\lambda^{\prime})=-\sigma_{xy}(n+1,s^{\prime}\lambda^{\prime}\rightarrow ns\lambda), (9)

meaning that the contributions to σxx\sigma_{xx} from the LL transition nsλ(n+1,sλ)ns\lambda\rightarrow(n+1,s^{\prime}\lambda^{\prime}) and from (n+1,sλ)nsλ(n+1,s^{\prime}\lambda^{\prime})\rightarrow ns\lambda are equal, while those to σxy\sigma_{xy} are opposite. The expressions of σxx\sigma_{xx} and σxy\sigma_{xy} are obtained as Y.X.Wang2023

σxx=\displaystyle\sigma_{xx}= σ0η2π2lB2𝑑kzn0s,sλMns,λ;n+1,s,λ2[(μεns,λ)2+η2][(μεn+1,s,λ)2+η2],\displaystyle\frac{\sigma_{0}\eta^{2}}{\pi^{2}l_{B}^{2}}\int_{-\infty}^{\infty}dk_{z}\sum_{n\geq 0}\sum_{s,s^{\prime}}\sum_{\lambda}\frac{M_{ns,\lambda;n+1,s^{\prime},\lambda}^{2}}{[(\mu-\varepsilon_{ns,\lambda})^{2}+\eta^{2}][(\mu-\varepsilon_{n+1,s^{\prime},\lambda})^{2}+\eta^{2}]}, (10)
σxy=\displaystyle\sigma_{xy}= σ0πlB2dkzn0s,sλ[(εns,λεn+1,s,λ)2η2]Mns,λ;n+1,s,λ2[(εns,λεn+1,s,λ)2+η2]2[θ(μεns)θ(εn+1,s,λμ)\displaystyle\frac{\sigma_{0}}{\pi l_{B}^{2}}\int_{-\infty}^{\infty}dk_{z}\sum_{n\geq 0}\sum_{s,s^{\prime}}\sum_{\lambda}\frac{[(\varepsilon_{ns,\lambda}-\varepsilon_{n+1,s^{\prime},\lambda})^{2}-\eta^{2}]M_{ns,\lambda;n+1,s^{\prime},\lambda}^{2}}{[(\varepsilon_{ns,\lambda}-\varepsilon_{n+1,s^{\prime},\lambda})^{2}+\eta^{2}]^{2}}[\theta(\mu-\varepsilon_{ns})\theta(\varepsilon_{n+1,s^{\prime},\lambda}-\mu)
θ(μεn+1,s,λ)θ(εns,λμ)],\displaystyle-\theta(\mu-\varepsilon_{n+1,s^{\prime},\lambda})\theta(\varepsilon_{ns,\lambda}-\mu)], (11)

where σ0=e22π\sigma_{0}=\frac{e^{2}}{2\pi} is the unit of the quantum conductivity, and θ(x)\theta(x) is the step function. Explicitly, the matrix elements are given as

Mns,1;n+1,s,1=2nξlBcns,11cn+1,s,11+2(n+1)ξlBcns,12cn+1,s,12+vcns,12cn+1,s,11,\displaystyle M_{ns,1;n+1,s^{\prime},1}=-\frac{\sqrt{2n}\xi}{l_{B}}c_{ns,1}^{1}c_{n+1,s^{\prime},1}^{1}+\frac{\sqrt{2(n+1)}\xi}{l_{B}}c_{ns,1}^{2}c_{n+1,s^{\prime},1}^{2}+vc_{ns,1}^{2}c_{n+1,s^{\prime},1}^{1}, (12)
Mns,1;n+1,s,1=2(n+1)ξlBcns,13cn+1,s,13+2nξlBcns,14cn+1,s,14vcns,13cn+1,s,14.\displaystyle M_{ns,-1;n+1,s^{\prime},-1}=-\frac{\sqrt{2(n+1)}\xi}{l_{B}}c_{ns,-1}^{3}c_{n+1,s^{\prime},-1}^{3}+\frac{\sqrt{2n}\xi}{l_{B}}c_{ns,-1}^{4}c_{n+1,s^{\prime},-1}^{4}-vc_{ns,-1}^{3}c_{n+1,s^{\prime},-1}^{4}. (13)

Due to the complicated form of the matrix elements Mns,λ;n+1,s,λM_{ns,\lambda;n+1,s^{\prime},\lambda}, the integrations over kzk_{z} in Eqs. (10) and (11) may not be completed analytically and need to be solved with numerics.

III Strong topological insulator in ZrTe5

Refer to caption
Figure 1: (Color online) (a) The LL spectra and the chemical potential μ\mu in the strong TI, with the LL index (nsλ)(ns\lambda) being labeled. The carrier density is fixed at n0=6.76×1016n_{0}=6.76\times 10^{16} cm-3 and the critical fields are Bs0=2B_{s}^{0}=2 T, Bs1=3.85B_{s}^{1}=3.85 T, Bs2=7.25B_{s}^{2}=7.25 T, Bs3=11.75B_{s}^{3}=11.75 T and Bs4=13B_{s}^{4}=13 T, as indicated by the dotted lines. The dashed blue line helps to judge Bs2B_{s}^{2} and the inset shows the enlarged plot around zero energy. (b)-(f) The LL dispersions at the critical fields Bs0,1,2,3,4B_{s}^{0,1,2,3,4}, respectively, in which the dotted red lines denote the positions of μ\mu.

In this section, we study ZrTe5 and take the model parameters from the experiment Y.Jiang : M=5M=5 meV, (v,vz)=(6,0)×105(v,v_{z})=(6,0)\times 10^{5} m/s, (ξ,ξz)=(100,200)(\xi,\xi_{z})=(100,200) meV nm2, g1=8g_{1}=-8 and g2=10g_{2}=10. In fact, when the Fermi velocity vzv_{z} is nonvanishing, a gap may be opened with the magnitude Y.X.Wang2021

Δ=2Mvz2ξzvz44ξz2.\displaystyle\Delta=2\sqrt{\frac{Mv_{z}^{2}}{\xi_{z}}-\frac{v_{z}^{4}}{4\xi_{z}^{2}}}. (14)

Here to observe the 1D Weyl nodes, we simply take vz=0v_{z}=0, so there is no gap opening and the system behaves as a semimetal. Since the features of band inversions in the xx-yy plane as well as the zz direction are kept and a finite vzv_{z} will not affect the main conclusions of this paper (see Appendix A), we still regard the ground state of the system as a strong TI. At a weak magnetic field, the band inversion in the zz direction leads to the zeroth LLs crossing at the momentum kz=±kck_{z}=\pm k_{c}, with

kc=(MξzeξBξz+g1μBB2ξz)12.\displaystyle k_{c}=\Big{(}\frac{M}{\xi_{z}}-\frac{e\xi B}{\xi_{z}}+\frac{g_{1}\mu_{B}B}{2\xi_{z}}\Big{)}^{\frac{1}{2}}. (15)

We fix the carrier density at n0=6.76×1016n_{0}=6.76\times 10^{16} cm-3 and calculate the chemical potential μ\mu through solving Eq. (5) self-consistently. The results are displayed in Fig. 1(a) as a function of the magnetic field BB, in which μ\mu exhibits a non-monotonous variation. In fact, μ\mu is determined by the interplay between the local DOS due to the 1D LL dispersion and the uniform DOS gg that is proportional to BB. When BB is weak, μ\mu lies above the n1n\geq 1 LLs and the quantum oscillations are clearly visible in μ\mu. According to the Onsager’s relation, the oscillation period is closely related to the Fermi surface area D.Shoenberg , as analyzed by us in a previous study Y.X.Wang2023 . When BB increases, the critical fields that are caused by the combined effects of the chemical potential variation and the LL movements will appear successively [Fig. 1(a)].

First, at the critical field Bs0B_{s}^{0}, the system enters into the quantum limit, with all electrons confined to the zeroth LLs [Fig. 1(b)]. When BB increases, the 0+0+ and 00- LLs move upwards and downwards, respectively. Since the uniform DOS gaining surpasses the local DOS dropping, μ\mu decreases with BB.

At the critical field Bs1B_{s}^{1}, μ\mu intersects the 00- LL [Fig. 1(c)], with

Bs1=Mμeξ(g1+g2)μB/2.\displaystyle B_{s}^{1}=\frac{M-\mu}{e\xi-(g_{1}+g_{2})\mu_{B}/2}. (16)

Then the Fermi surface varies from incorporating two points to four points. Correspondingly, the crossing points at kz=±kck_{z}=\pm k_{c} own opposite chiralities and behave as two 1D Weyl nodes. The effective Hamiltonian eff{\cal H}^{\text{eff}} is written as

eff=vF(kzkc)σz,\displaystyle{\cal H}^{\text{eff}}=v_{F}(k_{z}\mp k_{c})\sigma_{z}, (17)

where the Fermi velocity is

vF=2ξz12(MeξB+g1μBB2)12.\displaystyle v_{F}=2\xi_{z}^{\frac{1}{2}}\Big{(}M-e\xi B+\frac{g_{1}\mu_{B}B}{2}\Big{)}^{\frac{1}{2}}. (18)

Thus in the strong TI phase of ZrTe5, the magnetic field drives the Lifshitz transition and the system lies in the 1D Weyl regime.

When BB further increases, μ\mu decreases steadily, meaning that the magnetic field can effectively modulate the position of μ\mu with respect to the Weyl nodes. We see that μ\mu will meet the Weyl nodes at the critical field Bs2B_{s}^{2} [Fig. 1(d)], with

Bs2=2μg2μB,\displaystyle B_{s}^{2}=\frac{2\mu}{g_{2}\mu_{B}}, (19)

in which the g2g_{2}-spin Zeeman term plays a decisive role in driving Bs2B_{s}^{2}; without g2g_{2}, μ\mu cannot meet the Weyl nodes. In the experiment, Bs2B_{s}^{2} is detectable only when g2g_{2} is strong enough. The decreasing μ\mu is followed by an upturn [Fig. 1(a), inset]. This is because the Γ\Gamma point of the 0+0+ LL moves to be above the zero energy, leading to the carriers turning from holes to electrons.

Refer to caption
Figure 2: (Color online) The zeroth LLs in the quantum limit, and the chemical potential μ\mu for a set of carrier densities n0=(2.751,2.143,1.122)×1017n_{0}=(2.751,2.143,1.122)\times 10^{17} cm-3. The inset plots the characteristic carrier density nsn_{s}, with the horizontal lines denoting different n0n_{0}.
Refer to caption
Figure 3: (Color online) The DOS (a), longitudinal conductivity σxx\sigma_{xx} (b) and Hall conductivity σxy\sigma_{xy} (c) versus the magnetic field BB in the strong TI, with the parameters the same as Fig. 1(a) and the critical fields labeled by the dotted lines. (b) and (c) are plotted for different linewidths η\eta. The arrow in (c) marks the kink structure in σxy\sigma_{xy}. The inset in (c) plots σxy\sigma_{xy} versus the inverse magnetic field B1B^{-1} at η=0\eta=0, where the red dashed line shows the linear relation of σxyB1\sigma_{xy}\sim B^{-1}, with the slope extracted as k=0.108k=0.108 mΩ1\Omega^{-1}cm-1T.

After that, μ\mu will be close to the Γ\Gamma point of the 0+0+ LL and crosses it at the critical field Bs3B_{s}^{3} [Fig. 1(e)], with

Bs3=M+μeξ(g1g2)μB/2.\displaystyle B_{s}^{3}=\frac{M+\mu}{e\xi-(g_{1}-g_{2})\mu_{B}/2}. (20)

When B>Bs3B>B_{s}^{3}, only the 00- LL is occupied and the Weyl nodes do not exist.

Finally, at the critical field Bs4B_{s}^{4}, with

Bs4=Meξg1μB/2,\displaystyle B_{s}^{4}=\frac{M}{e\xi-g_{1}\mu_{B}/2}, (21)

the zeroth LLs touch each other at the Γ\Gamma point [Fig. 1(f)]. Further increasing BB, the zeroth LLs will be separated, and the system becomes a trivial insulator.

According to the above analysis, in the strong TI phase, the magnetic field can drive the zeroth LLs from crossing to be separated. Correspondingly, the system turns from topological nontrivial to trivial. Based on this observation, we suggest that the 1D Weyl nodes do not always appear in the strong TI phase; their appearance depends heavily on the carrier density of the system. To further clarify this point, we calculate the chemical potential μ\mu for a set of the carrier density n0n_{0} and plot the results in Fig. 2. Since a gap is opened when B>Bs4B>B_{s}^{4}, if μ\mu lies in the gap, only the 00- LL is occupied. The characteristic carrier density nsn_{s} is determined as

ns(B)=0μ=ε0(Γ)D0(ε)𝑑ε=gπC+g2μBB/2ξz.\displaystyle n_{s}(B)=\int_{0}^{\mu=\varepsilon_{0-}(\Gamma)}D_{0-}(\varepsilon)d\varepsilon=\frac{g}{\pi}\sqrt{\frac{C+g_{2}\mu_{B}B/2}{\xi_{z}}}. (22)

The inset of Fig. 2 shows that nsn_{s} increases with BB monotonously.

In Fig. 2, we see that there exist three cases for the 1D Weyl nodes: (i) When the carrier density n0<ns(Bs4)n_{0}<n_{s}(B_{s}^{4}), e.g., n0=1.122×1017n_{0}=1.122\times 10^{17} cm-3, the chemical potential μ\mu intersects the 00- LL before the zeroth LLs getting separated, which enables the formation of the Weyl nodes. This is similar to the case in Fig. 1(a). (ii) When ns(Bs4)<n0<ns(Bm)n_{s}(B_{s}^{4})<n_{0}<n_{s}(B_{m}), e.g., n0=2.143×1017n_{0}=2.143\times 10^{17} cm-3, μ\mu can only meet the Γ\Gamma point of the 0+0+ LL after a gap is opened in the system. Here BmB_{m} represents the maximum magnetic field that would be accessible in the experiment and is taken as Bm=30B_{m}=30 T. Then μ\mu will drop across the gap and lies below the Γ\Gamma point of the 00- LL, leading to the absence of the Weyl nodes. (iii) When n0>ns(Bm)n_{0}>n_{s}(B_{m}), e.g., n0=2.751×1017n_{0}=2.751\times 10^{17} cm-3, μ\mu cannot intersect the 00- LLs even after a gap is opened, and no Weyl nodes exist. Therefore, in the strong TI phase, the carrier density of the crystal sample directly determines the Weyl nodes that are driven by the magnetic field. For comparison, in Ref. W.Wu , a schematic plot of the chemical potential variation with BB is presented for a fixed carrier density in the strong TI phase, which is just case (iii).

To find the signatures of the critical fields in the quantum limit, we calculate the DOS as well as the conductivity. With the parameters chosen the same as Fig. 1(a), we plot the results in Fig. 3 as functions of the magnetic field BB, in which the critical fields are labeled by the dotted lines.

In Fig. 3(a), in the quantum limit, three asymmetric peaks are exhibited at the critical fields Bs0B_{s}^{0}, Bs1B_{s}^{1} and Bs3B_{s}^{3}, where μ\mu crosses the (1+)(1+-), 00-, and 0+0+ LL, respectively. Explicitly, the peaks at Bs0B_{s}^{0}, Bs1B_{s}^{1}, and Bs3B_{s}^{3} have long tails towards B<Bs0B<B_{s}^{0}, B>Bs1B>B_{s}^{1}, and B<Bs3B<B_{s}^{3}, respectively. Moreover, the DOS remains finite when μ\mu meets the Weyl nodes at Bs2B_{s}^{2}. This is different from the Weyl nodes that are formed by breaking the TRS or IS in a Dirac semimetal X.Wan ; S.Y.Xu ; Lv2015 , where the DOS vanishes at the Weyl nodes.

Refer to caption
Figure 4: (Color online) (a) The LL spectra and the chemical potential μ\mu in the weak TI, with the LL index (nsλ)(ns\lambda) being labeled. The carrier density is fixed at n0=1.25×1017n_{0}=1.25\times 10^{17} cm-3 and the critical fields are Bw0=4B_{w}^{0}=4 T, Bw1=7B_{w}^{1}=7 T, Bw2=8.6B_{w}^{2}=8.6 T, and Bw3=10.85B_{w}^{3}=10.85 T, as labeled by the dotted lines. The dashed blue line helps to judge Bw3B_{w}^{3}, and the inset shows the enlarged plot around zero energy. (b)-(e) The LL dispersions at the critical fields Bw0,1,2,3B_{w}^{0,1,2,3}, respectively, in which the red blue lines denote the positions of μ\mu.

In Figs. 3(b) and 3(c), the conductivities are shown, with the linewidth η\eta included to represent the effect of the impurity scatterings. In the quantum limit, we observe that: (i) The longitudinal conductivity σxx\sigma_{xx} increases with η\eta, whereas the Hall conductivity σxy\sigma_{xy} shows certain robustness to η\eta. These results are consistent with our previous studies Y.X.Wang2020 . (ii) In σxx\sigma_{xx}, corresponding to the DOS, three peaks are exhibited at the critical fields Bs0,Bs1B_{s}^{0},B_{s}^{1} and Bs3B_{s}^{3}. Such peaks are distinguishable when η\eta is weak but would be smeared at a strong η=10\eta=10 meV. (iii) σxy\sigma_{xy} decreases smoothly with BB except a kink as marked by the arrow [Fig. 3(c)], which is attributed to the upturn behavior of μ\mu. (iv) σxy\sigma_{xy} vanishes at the critical field Bs2B_{s}^{2} and will reverse its sign when the carriers change from electrons to holes. The vanishing σxy\sigma_{xy} shows certain robustness to η\eta, which favors the experimental observations. Thus the signatures of Bs0,1,3B_{s}^{0,1,3} and Bs2B_{s}^{2} can be captured by the peaks of σxx\sigma_{xx} and the zero value of σxy\sigma_{xy}, respectively. But no signatures of Bs4B_{s}^{4} are found in the conductivities, since the conductivities reflect the properties of the Fermi surface or Fermi sea in the system, and μ\mu cannot meet the touching Γ\Gamma point of the zeroth LLs unless the carrier density is exactly n0=n0s(Bs4)n_{0}=n_{0}^{s}(B_{s}^{4}). In the experiment, the signatures of Bs1B_{s}^{1} and Bs2B_{s}^{2} have been reported in the dc transport measurements of ZrTe5 S.Galeski2022 .

In the inset of Fig. 3(c), we plot σxy\sigma_{xy} as a function of the inverse magnetic field B1B^{-1} in the clean case η=0\eta=0. We see that in the quantum oscillation regime, the classical linear dependence of σxy\sigma_{xy} on B1B^{-1} A.A.Abrikosov ; V.Konye ,

σxy=n0eB,\displaystyle\sigma_{xy}=\frac{n_{0}e}{B}, (23)

is retrieved. With the extracted slope k=0.108k=0.108 mΩ1\Omega^{-1} cm-1 T, the carrier density is obtained as n0=6.75×1016n_{0}=6.75\times 10^{16} cm-3, which agrees well with the chosen value. But in the quantum limit, σxy\sigma_{xy} shows evident deviations from the linear relation. Intuitively, if the linear relation holds in the quantum limit, σxy\sigma_{xy} would vanish at an infinite magnetic field; now since σxy\sigma_{xy} vanishes at the critical field Bs2B_{s}^{2}, the linear relation can no longer hold.

To further understand the above behaviors, we investigate the role of the PHS in σxy\sigma_{xy} by choosing a set of the Zeeman splittings g1g_{1} and g2g_{2} and plot the results in the inset of Fig. 3(c). When g1=g2=0g_{1}=g_{2}=0, the system owns the PHS and the Weyl nodes are located at the zero energy. We see that the linear dependence of σxy\sigma_{xy} on B1B^{-1} holds in the quantum limit. When g1=8g_{1}=-8, the 0+0+ and 00- LLs will move upwards and downwards, respectively, but keep crossing each other. As the PHS is preserved, the Weyl nodes remain located at the zero energy and the linear dependence holds in the quantum limit. However, at a finite g2=10g_{2}=10, the PHS is broken and thus the Weyl nodes are shifted in energy. Consequently, the linear dependence will be destroyed in the quantum limit. In a recent magnetotransport experiment of ZrTe5, the Hall resistivity ρxy\rho_{xy} exhibited a similar dependence on the magnetic field S.Galeski2021 , which supports our theoretical analysis.

Actually, when the chemical potential μ>0\mu>0 lies between the nnth and (n+1)(n+1)th LLs, the dominant contributions to σxy\sigma_{xy} come from the LL transition (n,1,λ)(n+1,1,λ)(n,1,\lambda)\rightarrow(n+1,1,\lambda). This is reminiscent of the 2D Dirac fermion behavior in graphene V.P.Gusynin . When the PHS is broken, the 0+0+ and 00- LLs will move asymmetrically in energy. On one hand, such movements will not change the topological property of the LLs, which can explain the vanishing σxy\sigma_{xy} at the Weyl nodes. On the other hand, in the quantum oscillation regime, σxy\sigma_{xy} is related to the transition of n1n\geq 1 LLs, thus the LL movements will not affect σxy\sigma_{xy} and the linear relation; whereas in the quantum limit, σxy\sigma_{xy} is related to the transition of the zeroth LL to the n=1n=1 LL, and the LL movements will change σxy\sigma_{xy} and destroy the linear relation.

IV Weak topological insulator in HfTe5

In this section, we study HfTe5 and take the model parameters from the experiments W.Wu : M=2.5M=2.5 meV, (v,vz)=(4.5,0)×105(v,v_{z})=(4.5,0)\times 10^{5} m/s, (ξ,ξz)=(120,200)(\xi,\xi_{z})=(120,-200) meV nm2, g1=6g_{1}=-6 and g2=10g_{2}=10. Now the ground state of the system is a weak TI that features the band inversion only in the xyx-y plane. Under a weak magnetic field, the topological trivial bands in the zz direction will lead to a gap between the zeroth LLs and the system behaves as a gapped insulator. At a fixed carrier density n0=1.25×1017n_{0}=1.25\times 10^{17} cm-3, the calculated chemical potential μ\mu is displayed in Fig. 4. The increasing magnetic field BB will drive the emergence of several critical fields [Fig. 4(a)].

Refer to caption
Figure 5: (Color online) The DOS (a), longitudinal conductivity σxx\sigma_{xx} (b) and Hall conductivity σxy\sigma_{xy} (c) versus the magnetic field BB in the weak TI, with the parameters the same as Fig. 4(a) and the critical fields labeled by the dotted lines. (b) and (c) are plotted for different linewidths η\eta. The inset in (c) plots σxy\sigma_{xy} versus the inverse magnetic field B1B^{-1} at η=0\eta=0, where the red dashed line shows the linear relation of σxyB1\sigma_{xy}\sim B^{-1}, with the slope extracted as k=0.2055k=0.2055 mΩ1\Omega^{-1}cm-1T.

Firstly, the quantum limit is achieved at the critical field Bw0B_{w}^{0} [Fig. 4(b)]. We see that the 0+/0+/- LL lies in the valence/conduction band, which is different from the strong TI. With increasing BB, the 0+0+ and 00- LLs will move upwards and downwards, respectively. Since both the uniform DOS and local DOS increase with BB, μ\mu also decreases.

At the critical field Bw1B_{w}^{1}, the zeroth LLs touch each other at the Γ\Gamma point so that the gap is closed [Fig. 4(c)]. Then the zeroth LLs cross each other at kz=±kck_{z}=\pm k_{c}, with the gap being persistently closed. Note that kck_{c} has the same expression as Eq. (15). When BB further increases, the decreasing μ\mu will meet the 0+0+ LL at the critical field Bw2B_{w}^{2} [Fig. 4(d)]. Meantime, the Lifshitz transition occurs in the weak TI phase, where the Fermi surface varies from incorporating two points to four points. Correspondingly, the crossing points act as two 1D Weyl nodes. When B>Bw2B>B_{w}^{2}, μ\mu shows an upturn and then decreases slowly. This is because the local DOS drops when μ\mu crosses the Γ\Gamma point of the 0+0+ LL. Finally, μ\mu will meet the Weyl nodes at the critical field Bw3B_{w}^{3} [Fig. 4(e)].

According to the above analysis, in the weak TI phase, the magnetic field drives the zeroth LLs from being gapped to cross each other. Correspondingly, the system turns from topological trivial to nontrivial, which is opposite to the strong TI. This is also seen from the fact that the critical fields Bw1B_{w}^{1}, Bw2B_{w}^{2}, and Bw3B_{w}^{3} have the same expressions as Bs4B_{s}^{4}, Bs3B_{s}^{3} and Bs2B_{s}^{2}, respectively. Note that chemical potential variations with BB in the weak TI phase are consistent with Ref. W.Wu .

It is interesting to ask whether there exists the critical field Bw4B_{w}^{4} at which μ\mu crosses the 00- LL. When it happens, Bw4B_{w}^{4} will have the same expression as Bs1B_{s}^{1}. Now, since only the 0+0+ LL is occupied, the characteristic carrier density nw1n_{w}^{1} is determined as

nw1(B)\displaystyle n_{w}^{1}(B) =0μ=ε0(Γ)D0+(ε)𝑑ε\displaystyle=\int_{0}^{\mu=\varepsilon_{0-}(\Gamma)}D_{0+}(\varepsilon)d\varepsilon
=gπ(Cg2μBB/2ξz2Cξz),\displaystyle=\frac{g}{\pi}\Big{(}\sqrt{\frac{C-g_{2}\mu_{B}B/2}{\xi_{z}}}-\sqrt{\frac{2C}{\xi_{z}}}\Big{)}, (24)

which decreases with BB. Clearly, Bw4B_{w}^{4} will appear if the carrier density satisfies n0<nw1(B)n_{0}<n_{w}^{1}(B) at a certain BB. For the magnetic field Bw1<B<BmB_{w}^{1}<B<B_{m}, the calculations show that nw1n_{w}^{1} lies in the range (0.965.45)×1016(0.96\sim 5.45)\times 10^{16} cm-3, which is far below the chosen carrier density, thus Bw4B_{w}^{4} would not appear.

We investigate the conditions for the Weyl nodes in the weak TI phase. When the zeroth LLs get crossed, if the chemical potential μ\mu meets the Γ\Gamma point of the 0+0+ LL, the characteristic carrier density is

nw2(B)=\displaystyle n_{w}^{2}(B)= 0μ=ε0+(Γ)[D0+(ε)+D0(ε)]𝑑ε\displaystyle\int_{0}^{\mu=\varepsilon_{0+}(\Gamma)}[D_{0+}(\varepsilon)+D_{0-}(\varepsilon)]d\varepsilon
=\displaystyle= gπ(Cg2μBB/2ξz+2Cξz),\displaystyle\frac{g}{\pi}\Big{(}\sqrt{\frac{C-g_{2}\mu_{B}B/2}{\xi_{z}}}+\sqrt{\frac{2C}{\xi_{z}}}\Big{)}, (25)

which increases with the magnetic field. For the magnetic field Bw1<B<BmB_{w}^{1}<B<B_{m}, nw2n_{w}^{2} is given in the range (5.45133.87)×1016(5.45\sim 133.87)\times 10^{16} cm-3, which is quite broad. In the weak TI phase, the condition for the Weyl nodes is that the carrier density satisfies nw1(B)<n0<nw2(B)n_{w}^{1}(B)<n_{0}<n_{w}^{2}(B) at a certain BB. Since this condition is easily satisfied, the Weyl nodes are more likely to appear in the weak TI phase.

Next we study the DOS and conductivities of the system. The results are displayed in Fig. 5, where the critical fields are labeled by the dotted lines. At the critical fields Bw0B_{w}^{0} and Bw2B_{w}^{2}, the chemical potential crosses the (1+)(1+-) and 0+0+ LL, respectively, which leads to the asymmetric peaks in the DOS as well as in the longitudinal conductivity σxx\sigma_{xx} [Figs. 5(a) and 5(b)]. The Hall conductivity σxy\sigma_{xy} decreases with BB smoothly and vanishes at the critical field Bw3B_{w}^{3} [Fig. 5(c)]. Moreover, the effects of the linewidth η\eta on the conductivities are the same as those in Fig. 3. Therefore, the signatures of the critical field Bw0,2,3B_{w}^{0,2,3} are clearly seen in the conductivities. In the inset of Fig. 5(c), we plot σxy\sigma_{xy} as a function of the inverse magnetic field B1B^{-1} at η=0\eta=0. In the quantum oscillation regime, the linear dependence of σxy\sigma_{xy} on the inverse magnetic field B1B^{-1} is seen. With the extracted slope k=0.1993k=0.1993 mΩ1\Omega^{-1}cm-1T, the carrier density is obtained as n0=ke=1.244×1017n_{0}=\frac{k}{e}=1.244\times 10^{17} cm-3, which agrees well with the chosen carrier density. But in the quantum limit, σxy\sigma_{xy} shows evident deviations from the linear dependence. This observation can also be attributed to the broken PHS caused by the g2g_{2}-spin Zeeman term in the system, which is similar to the above strong TI phase analysis. In Ref. W.Wu of the longitudinal resistivity RxxR_{xx} measurements, a prominent linear behavior before saturating at high fields was observed but there were no evident peaks for the critical fields, thus more dc transport measurements in HfTe5 are expected in the future.

V Discussions and Conclusions

Refer to caption
Figure 6: (Color online) The LL dispersions versus kzk_{z} when the magnetic field B=2B=2 T in (a) and B=7B=7 T in (b), with the LL index (nsλ)(ns\lambda) being labeled. The Fermi velocity vz=5×104v_{z}=5\times 10^{4} m/s, and the other parameters are the same as Fig. 1(a). The saddle points of the zeroth LL, Γ\Gamma and ζ0\zeta_{0}, are marked as asterisks.

In this paper, we focus on fixed carrier density n0n_{0}. Now we discuss the case of fixed chemical potential μ0\mu_{0}, since it can provide more insights to understand the exotic transport behavior in pentatellurides, such as the 3D quantum Hall effect S.Galeski2021 ; F.Xiong ; Y.X.Wang2023 . In both strong and weak TIs, with fixed μ0\mu_{0}, the magnetic field can also drive the critical fields. In the strong TI, the Weyl nodes appear when μ0\mu_{0} satisfies ε0+(Γ)<μ0<ε0(Γ)\varepsilon_{0+}(\Gamma)<\mu_{0}<\varepsilon_{0-}(\Gamma), which strongly depends on the carrier density; while in the weak TI, the Weyl nodes appear when μ0\mu_{0} satisfies ε0(Γ)<μ0<ε0+(Γ)\varepsilon_{0-}(\Gamma)<\mu_{0}<\varepsilon_{0+}(\Gamma). The latter condition is easily satisfied, as the magnetic field drives the zeroth LLs to cross each other and the 0+0+ LL move upwards. Therefore we suggest that the conclusions for the 1D Weyl nodes with fixed μ0\mu_{0} are similar to those with fixed n0n_{0}.

When the Fermi velocity vzv_{z} is nonzero, the following consequences will be induced: (i) the 1D Weyl nodes will become gapped, thus the behavior of the zeroth LLs mimicks the physics of the 1D massive Weyl nodes; (ii) the upspin and downspin will be mixed and are no longer good quantum numbers; and (iii) in the strong TI phase, the zeroth LLs may avoid crossing each other, leading to the additional saddle points Y.Jiang ; J.Wang . Thus in the DOS and longitudinal conductivity σxx\sigma_{xx}, more peaks will be found when the chemical potential crosses such saddle points. The details are presented in Appendix A. In addition, the effect of temperature on the magneotransport is briefly discussed in Appendix B.

To summarize, our work explores the conditions for the magnetic field-driven Weyl nodes and the dc transport property in the quantum limit of the 3D pentatellurides. Although the quantitative results depend on the model parameters, they are qualitatively valid and can show guiding significance for the experiments. We hope that the 3D pentatellurides under a magnetic field can open an avenue for studying the interactions of 1D Weyl fermions as well as the resulting various strongly correlated electronic states.

VI Acknowledgment

This work was supported by the Natural Science Foundation of China (Grant No. 11804122), and the China Postdoctoral Science Foundation (Grant No. 2021M690970).

VII Appendix

VII.1 LLs with nonzero vzv_{z}

Refer to caption
Figure 7: (Color online) The chemical potential μ\mu (a), longitudinal conductivity σxx\sigma_{xx} (b) and Hall conductivity σxy\sigma_{xy} (c) versus the magnetic field BB for different temperatures TT. The parameters are the same as Fig. 1(a) and the critical fields are labeled by the dotted lines. The legends are the same in all figures.

For a finite Fermi velocity vzv_{z}, the zeroth LLs can still be obtained analytically, with the energies

ε0λ(kz)=\displaystyle\varepsilon_{0\lambda}(k_{z})= λ[(Mξzkz2ξlB2+12g1μBB)2+vz2kz2]12\displaystyle\lambda\Big{[}\big{(}M-\xi_{z}k_{z}^{2}-\frac{\xi}{l_{B}^{2}}+\frac{1}{2}g_{1}\mu_{B}B\big{)}^{2}+v_{z}^{2}k_{z}^{2}\Big{]}^{\frac{1}{2}}
+12g2μBB,\displaystyle+\frac{1}{2}g_{2}\mu_{B}B, (A1)

while the n1n\geq 1 LLs need to be solved numerically. Now although the 1D Weyl nodes are gapped, the LL spectra at kz=0k_{z}=0 are unaffected by the finite vzv_{z} in both strong and weak TIs [Figs. 1(a) and 4(a)], thus the evolution of the zeroth LLs with the magnetic field as well as the determined critical fields remain unchanged.

In the strong TI, besides the saddle point Γ\Gamma at kz=0k_{z}=0, the additional saddle points ζn\zeta_{n} may appear Y.Jiang ; J.Wang ; L.You . For example, when B=2B=2 T [Fig. 6(a)], in the zeroth LLs, the additional saddle points ζ0\zeta_{0} can be found and are located at kz=±(M+g1μB/2ξzξξzlB2vz22ξz2)12k_{z}=\pm\big{(}\frac{M+g_{1}\mu_{B}/2}{\xi_{z}}-\frac{\xi}{\xi_{z}l_{B}^{2}}-\frac{v_{z}^{2}}{2\xi_{z}^{2}}\big{)}^{\frac{1}{2}}. With increasing BB, ζ0\zeta_{0} moves to Γ\Gamma and will finally merge with it at the critical field B0c=Mvz2/2ξzeξg1μB/26B_{0c}=\frac{M-v_{z}^{2}/2\xi_{z}}{e\xi-g_{1}\mu_{B}/2}\simeq 6 T. When B>B0cB>B_{0c}, there are no additional saddle points in the zeroth LLs [Fig. 6(b)].

VII.2 Effect of temperature

Here we study the effect of temperature on the magnetotransport in pentatellurides. At a finite temperature TT, the conductivity σαβ\sigma_{\alpha\beta} will be modified in two aspects: one is that TT can shift the chemical potential μ\mu; and another is that TT directly enters σαβ\sigma_{\alpha\beta} via the Fermi-Dirac distribution function f(x)f(x). By multiplying the zero-temperature conductivity σαβ(T=0)\sigma_{\alpha\beta}(T=0) by 𝑑ϵδ(ϵμ)\int_{-\infty}^{\infty}d\epsilon\delta(\epsilon-\mu), the finite-temperature conductivity σαβ(T)\sigma_{\alpha\beta}(T) can be expressed as a weighted integration of σαβ(T=0)\sigma_{\alpha\beta}(T=0) around the chemical potential μ\mu and is written as L.Smrcka

σαβ(T)=𝑑ϵσαβ(T=0,ϵ)[f(ϵμ)ϵ],\displaystyle\sigma_{\alpha\beta}(T)=\int_{-\infty}^{\infty}d\epsilon\sigma_{\alpha\beta}(T=0,\epsilon)\big{[}-\frac{\partial f(\epsilon-\mu)}{\partial\epsilon}\big{]}, (A2)

where the derivative of the Fermi-Dirac distribution function is f(ϵμ)ϵ=12kBT(1+coshϵμkBT)-\frac{\partial f(\epsilon-\mu)}{\partial\epsilon}=\frac{1}{2k_{B}T\big{(}1+\text{cosh}\frac{\epsilon-\mu}{k_{B}T}\big{)}}.

In Fig. 7, with the parameters the same as Fig. 1(a), we display the results of the chemical potential μ\mu, longitudinal conductivity σxx\sigma_{xx}, and Hall conductivity σxy\sigma_{xy} for different temperatures TT. In Fig. 7(a), we see that with increasing TT, when B<Bs0B<B_{s}^{0} in the quantum oscillation regime, the oscillations of μ\mu are weakened; when B>Bs0B>B_{s}^{0} in the quantum limit, μ\mu is less affected at T20T\leq 20 K. When the temperature increases to T=100T=100 K, μ\mu even becomes negative, meaning that μ\mu shifts from the conduction band to the valence band. This result is consistent with the experimental observations Y.Zhang2017a ; Y.Zhang2017b and was believed to be the underlying physical mechanism of the anomalous resistivity peak at a finite TT in pentatellurides C.Wang . In Fig. 7(b), the peaks of σxx\sigma_{xx} are smeared by temperature, and in Fig. 7(c), σxy\sigma_{xy} is gradually suppressed. Note that when T=100T=100 K, σxx\sigma_{xx} becomes a smooth curve, indicating that the quantized LLs do not exist and the system enters the semiclassical diffusive region.

References

  • (1) L. D. Landau and E. M. Lifshitz, Quantum Mechanics: Non-relativistic Theory, Course of Theoretical Physics Vol. 3 (Elsevier, New York, 2013).
  • (2) K. V. Klitzing, G. Dorda, and M. Pepper, Phys. Rev. Lett. 45, 494 (1980).
  • (3) D. C. Tsui, H. L. Stormer, and A. C. Gossard, Phys. Rev. Lett. 48, 1559 (1982).
  • (4) D. Yoshioka, The Quantum Hall Effect, (Springer, New York, 2013), Vol. 133..
  • (5) M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).
  • (6) X. L. Qi and S. C. Zhang, Rev. Mod. Phys. 83, 1057 (2011).
  • (7) B. Q. Lv, T. Qian, and H. Ding, Rev. Mod. Phys. 93, 025002 (2021).
  • (8) J. A. Sobota, Y. He, and Z. X. Shen, Rev. Mod. Phys. 93, 025006 (2021).
  • (9) 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).
  • (10) 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).
  • (11) H. Weng, X. Dai, and Z. Fang, Phys. Rev. X 4, 011002 (2014).
  • (12) 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).
  • (13) 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).
  • (14) 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).
  • (15) P. Wang, Y. Ren, F. Tang, P. Wang, T. Hou, H. Zeng, L. Zhang, and Z. Qiao, Phys. Rev. B 101, 161201(R) (2020).
  • (16) 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, and J. Gooth, Nat. Commun. 11, 5926 (2020).
  • (17) S. Galeski, H. F. Legg, R. Wawrzyńczak, T. Förster, S. Zherlitsyn, D. Gorbunov, M. Uhlarz, P. M. Lozano, Q. Li, G. D. Gu, C. Felser, J. Wosnitza, T. Meng, and J. Gooth, Nat. Commun. 13, 7418 (2022).
  • (18) 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).
  • (19) X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov, Phys. Rev. B 83, 205101 (2011).
  • (20) S. Y. Xu, I. Belopolski, N. Alidoust, M. Neupane, G. Bian, C. Zhang, R. Sankar, G. Chang, Z. Yuan, C. C. Lee, S. M. Huang, H. Zheng, J. Ma, D. S. Sanchez, B. Wang, A. Bansil, F. Chou, P. P. Shibayev, H. Lin, S. Jia, M. Z. Hasan, Science 349, 613 (2015).
  • (21) B. Q. Lv, H. M. Weng, B. B. Fu, X. P. Wang, H. Miao, J. Ma, P. Richard, X. C. Huang, L. X. Zhao, G. F. Chen, Z. Fang, X. Dai, T. Qian, and H. Ding, Phys. Rev. X 5, 031013 (2015).
  • (22) G. Grüner, Density Waves in Solids (Perseus, Cambridge, 2000).
  • (23) 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).
  • (24) 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).
  • (25) L. Fu, C. L. Kane, and E. J. Mele, Phys. Rev. Lett. 98, 106803 (2007).
  • (26) L. Fu and C. L. Kane, Phys. Rev. B 76 045302 (2007).
  • (27) 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).
  • (28) 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).
  • (29) 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).
  • (30) Z. Rukelj, C. C. Homes, M. Orlita, and A. Akrap, Phys. Rev. B 102, 125201 (2020).
  • (31) A. P. Schnyder, S. Ryu, A. Furusaki, and A. W. W. Ludwig, Phys. Rev. B 78, 195125 (2008).
  • (32) C. K. Chiu, J. C. Y. Teo, A. P. Schnyder, and S. Ryu, Rev. Mod. Phys. 88, 035005 (2016).
  • (33) L. You, Z. Y. Zhang, and Y. X. Wang, New J. Phys. 23, 123033 (2021).
  • (34) B. Fu, H. W. Wang, and S. Q. Shen, Phys. Rev. Lett. 125, 256601 (2020).
  • (35) C. Wang, Phys. Rev. Lett. 126, 126601 (2021).
  • (36) L. Smrcka and P. Streda, J. Phys. C: Solid State Phys. 10, 2153 (1977).
  • (37) G. D. Mahan, Many-Particle Physics, 3rd ed. (Plenum, New York, 2000).
  • (38) Y. X. Wang and Z. Cai, Phys. Rev. B 107, 125203 (2023).
  • (39) Y. X. Wang and F. Li, Phys. Rev. B 103, 115202 (2021).
  • (40) D. Shoenberg, Magnetic Oscillations in Metals (Cambridge University Press, Cambridge, 1984).
  • (41) Y. X. Wang and F. Li, Phys. Rev. B 101, 085201 (2020).
  • (42) A. A. Abrikosov, Phys. Rev. B 58, 2788 (1998).
  • (43) V. Könye and M. Ogata, Phys. Rev. B 98, 195420 (2018).
  • (44) 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).
  • (45) V. P. Gusynin and S. G. Sharapov, Phys. Rev. B 73, 245411 (2006).
  • (46) F. Xiong, C. Honerkamp, D. M. Kennes, and T. Nag, Phys. Rev. B 106, 045424 (2022).
  • (47) Y. Zhang, C. Wang, L. Yu, G. Liu, A. Liang, J. Huang, S. Nie, X. Sun, Y. Zhang, B. Shen, J. Liu, H. Weng, L. Zhao, G. Chen, X. Jia, C. Hu, Y. Ding, W. Zhao, Q. Gao, C. Li, S. He, L. Zhao, F. Zhang, S. Zhang, F. Yang, Z. Wang, Q. Peng, X. Dai, Z. Fang, Z. Xu, C. Chen, and X. J. Zhou, Nat. Commun. 8, 15512 (2017).
  • (48) Y. Zhang, C. Wang, G. Liu, A. Liang, L. Zhao, J. Huang, Q. Gao, B. Shen, J. Liu, C. Hu, W. Zhao, G. Chen, X. Jia, L. Yu, L. Zhao, S. He, F. Zhang, S. Zhang, F. Yang, Z. Wang, Q. Peng, Z. Xu, C. Chen, X. Zhou, Sci. Bull. 62, 950 (2017).