You are on page 1of 11

PHYSICAL REVIEW B 96, 045307 (2017)

Quantum transport in Weyl semimetal thin films in the presence of spin-orbit coupled impurities
Weizhe Edward Liu (),1 Ewelina M. Hankiewicz,2 and Dimitrie Culcer1,*
1
School of Physics and Australian Research Council Centre of Excellence in Low-Energy Electronics Technologies, UNSW Node,
The University of New South Wales, Sydney 2052, Australia
2
Institute for Theoretical Physics and Astrophysics, Würzburg University, Am Hubland, 97074 Würzburg, Germany
(Received 21 May 2017; published 26 July 2017)

Topological semimetals have been at the forefront of experimental and theoretical attention in condensed-matter
physics. Among these, recently discovered Weyl semimetals have a dispersion described by a three-dimensional
Dirac cone, which is at the root of exotic physics such as the chiral anomaly in magnetotransport. In a time-reversal
symmetric (TRS) Weyl semimetal film, the confinement gap gives the quasiparticles a mass, while TRS is
preserved by having an even number of valleys with opposite masses. The film can be tuned through a topological
phase transition by a gate electric field. In this work, we present a theoretical study of the quantum corrections
to the conductivity of a topological semimetal thin film, which is governed by the complex interplay of the
chiral band structure, mass term, and scalar and spin-orbit scattering. We study scalar and spin-orbit scattering
mechanisms on the same footing, demonstrating that they have a strong qualitative and quantitative impact on
the conductivity correction. We show that, due to the spin structure of the matrix Green’s functions, terms linear
in the extrinsic spin-orbit scattering are present in the Bloch and momentum relaxation times, whereas previous
works had identified corrections starting from the second order. In the limit of small quasiparticle mass, the
terms linear in the impurity spin-orbit coupling lead to a potentially observable density dependence in the weak
antilocalization correction. Moreover, when the mass term is around 30% of the linear Dirac terms, the system
is in the unitary symmetry class with zero quantum correction, and switching the extrinsic spin-orbit scattering
drives the system to the weak antilocalization. We discuss the crossover between the weak localization and
weak antilocalization regimes in terms of the singlet and triplet Cooperon channels, and analyze this transition
as a function of the mass and spin-orbit scattering strength. Experimental schemes to detect this transition are
discussed.

DOI: 10.1103/PhysRevB.96.045307

I. INTRODUCTION a gap ), and then the system [Fig. 1(b)] exhibits precisely
the same two-dimensional (2D) massive Dirac fermion states
In 1929 Hermann Weyl theoretically predicted the existence
[57,58] as those associated with the quantum anomalous Hall
of chiral massless fermions with a linear dispersion by identi-
effect [59,60]. In a similar manner, a Dirac semimetal thin film
fying a specific solution of the Dirac equation. Over the past 80
may host the quantum spin Hall effect, since it can be regarded
years, the search for Weyl fermions has stimulated numerous
as consisting of two Weyl semimetals that are time-reversed
studies in high-energy, condensed-matter, and mathematical
pairs [7]. In contrast to previously studied systems [61–63],
physics [1–3]. Recently, first-principles theoretical predic-
the quantum spin and anomalous Hall effects may be observed
tions followed by angle-resolved photoemission spectroscopy
in a single-compound device, a fact that has stimulated a
(ARPES) experiments have confirmed the existence of chiral
considerable amount of recent interest in quantum transport
massless Weyl fermions [4] in topological Dirac semimetals,
in thin films of topological semimetals.
[5–22], as well as in type-I [23–32] and type-II [33–37] Weyl
Topological semimetals can demonstrate a variety of trans-
semimetals (WSM).
port phenomena [64–73]. The ability to observe the quantum
Weyl fermion semimetals are three-dimensional topolog-
spin and anomalous Hall effects requires the Fermi energy
ical states of matter [38–41] in which the conduction and
to lie inside the gap opened by the quasiparticle mass. When
valence bands touch linearly at an even number of nodes
this happens then a set of one-dimensional edge states is well
[Fig. 1(a)], which appear in pairs with opposite chirality. This
defined inside the gap, whose ballistic transport is responsible
band structure is referred to as topological [42–52] because it
for the quantized conductance observed experimentally. When
is equivalent to a two-dimensional band insulator with the
the Fermi energy is not in the mass gap, quantum transport at
gap controlled by the momentum (kz ) along the direction
low temperatures is dominated by weak localization (WL) or
connecting two nodes. As kz changes from outside to between
the paired Weyl nodes, the system undergoes a topological antilocalization (WAL) effects [74–76]. These corrections to
phase transition from a trivial phase to a quantum anomalous the conductivity are noticeable when the quasiparticle mean
Hall phase [53]. As a result, there are kz -dependent topological free path is much shorter than the phase coherence length
edge states [54], forming the Fermi arcs that connect the paired [77] and arise as a result of the quantum interference between
Weyl fermions [55,56]. In an ultrathin film of topological closed, time-reversed loops that circle regions in which one or
WSM, kz is quantized, giving rise to a mass (or equivalently, more impurities are present [78]. Since the interference effects
leading to WL/WAL disappear in weak external magnetic
fields, these corrections can typically be identified straight-
forwardly in an experiment and are frequently used to char-
* acterize samples, in particular, transport in novel materials.
d.culcer@unsw.edu.au

2469-9950/2017/96(4)/045307(11) 045307-1 ©2017 American Physical Society


LIU, HANKIEWICZ, AND CULCER PHYSICAL REVIEW B 96, 045307 (2017)

(a) Weyl Semimetals (b) WSM thin films in the strength of the extrinsic spin-orbit scattering potential,
which has a strong angular dependence. This term appears
in the Bloch lifetime of the quasiparticles, in the transport
relaxation time, the spin relaxation time, and the Cooperon, and
energy
gives rise to a nontrivial density dependence of the quantum
εF correction to the conductivity, which may be observable when
the quasiparticle mass is very small. Taking into account
ky Δ terms up to second order in the extrinsic spin-orbit scattering,
kx we analyze quantitatively the carrier density dependence of
the electrical conductivity and provide a fitting formula that
may be used to extract the strength of the impurity spin-orbit
interactions, as outlined in Sec. III B. Furthermore, the first-
order spin-orbit scattering contribution suppresses the weak
FIG. 1. (a) A minimal sketch of the energy dispersion of a Weyl localization channel in the massive limit much more efficiently
semimetal. We have defined k2 = kx2 + ky2 , while (kx ,ky ,kz ) represents than the second-order term studied previously, and therefore
the three-dimensional wave vector. kz points along the preferred has a strong qualitative and quantitative effect on the phase
direction. The conductance and valence bands cross linearly at the diagram of the weak-localization to weak-antilocalization
Weyl nodes, i.e., the left and right cones, and the nodes appear in transition. With this insight, we determine the full phase
pairs with opposite chirality number. A Dirac node appears when two diagram as a function of the strengths of the extrinsic spin-orbit
oppositely chiral Weyl nodes merge together. (b) A schematic picture coupling and of the mass term in the quasiparticle dispersion,
of the band structure in WSM thin films, where  is the band gap which may be regarded as the central result of this work.
and εF is the Fermi energy. This paper is organized as follows. In Sec. II we describe the
generic model and theoretical formalism utilized throughout
They provide valuable information about the system [79,80], this article. In Sec. III we briefly display the conductivity
such as symmetries of the system, the phase coherence length dependence as a function of the external magnetic field, the
[81,82], and the mass of the Dirac fermions [83]. carrier density, and the sign of the σy term in the band
Thin films of topological semimetals provide a new Hamiltonian. In Sec. IV we discuss the physical implications
platform to understand weak localization and antilocalization of our results and their potential experimental applications.
behavior in generic 2D Dirac fermion systems in which the In Sec. V we summarize our conclusions briefly and discuss
mass may be taken as a parameter [83–87]. In the seminal possible future research directions.
work by Hikami et al. for conventional electrons (with a
parabolic dispersion εp = p2 /2m), weak localization and an-
tilocalization effects are classified according to the orthogonal, II. MODEL AND FORMALISM
symplectic, and unitary symmetry classes, corresponding to In this section, we illustrate our model and develop our
scalar, spin-orbit, and magnetic impurities, respectively [88]. formalism by Keldysh Green’s functions. Then we present
In the context of the 2D massive Dirac fermions, Shan et al. expressions for the Drude conductivity and the quantum
have considered all these impurity classes [89], where weak interference conductivity, along with a solution to the Bethe-
spin-orbit scattering, in analogy with the treatment in Ref. [88], Salpeter equation in a matrix form.
is included by retaining only the second-order terms in the
strength of the impurity spin-orbit coupling.
In this article we formulate a complete theory describing A. Band Hamiltonian
weak localization and antilocalization of 2D massive Dirac In WSM thin films, the effective band Hamiltonian in the
fermions in topological semimetal thin films. We focus on a vicinity of a Weyl node takes the form [90,91]
Weyl semimetal thin film as a prototype system. We stress
that in these materials, in which the band structure spin-orbit
H0k = A(σx kx + σy ky ) + Mσz , (1)
interactions are exceedingly strong, spin-orbit coupling in the
impurity scattering potentials is also expected to be sizable.
It is, therefore, necessary to treat scalar and spin-dependent where σx,y,z are Pauli matrices, A = h̄v is a material-specific
scattering on the same footing, which requires one to retain the constant, v is the effective velocity, k = (kx ,ky ) is the in-plane
matrix structure of all the Green’s functions and impurity po- wave vector measured from the Weyl node, and M is the effec-
tentials. We have developed a transparent theory that accounts tive mass due to the quantum confinement, which may also be
fully for the matrix structure of the Dirac/Weyl system within viewed as an effective Zeeman energy. H0k  describes two chiral
the framework of the Keldysh Green’s function formalism. bands whose eigenvalues are εk± = ± A2 k 2 + M 2 ≡ ±εk ,
The resulting weak localization/antilocalization behavior is where + and − indicate the conduction and valence bands,
consistent with the universality classes in the massless and respectively. This is seen in Fig. 1(b), in which the gap is
massive limits, respectively, and with the symmetries of the  = 2M. In this work, we assume that the√Fermi level is
system under chirality reversal. located in the conduction band and ε ∼ εF = A2 kF2 + M 2 at
Thanks to the matrix structure inherent in our formalism, a low temperatures, where kF is the Fermi wave vector. Since
major theoretical advance of this study compared to previous transport properties only rely on the Fermi level, we introduce
works [89] is our revelation of the existence of a linear term a = AkF /εF and b = M/εF where a 2 + b2 = 1.

045307-2
QUANTUM TRANSPORT IN WEYL SEMIMETAL THIN . . . PHYSICAL REVIEW B 96, 045307 (2017)

B. Impurity potential where ∂t1 ≡ ∂/∂t1 and the overhead left arrow means that the
For short-range impurities, the matrix elements of a single derivative acts on the left part. Here we use abbreviations
impurity potential in the reciprocal space, including spin-orbit 1 ≡ (t1 ,r 1 ) and δ(1 − 2) = δ(r 1 − r 2 )δ(t1 − t2 ), while H =
coupling, are given by H0 + V is the total Hamiltonian, with H0 the band Hamiltonian
and V the total impurity potential as introduced above.
Ukk = Ukk (1 + iλσz sin γ ), γ = θ  − θ, (2) The Dyson equation for Ĝ may be written as
where Ukk ≡ U is taken to represent the reciprocal-space
matrix element of a short-range impurity potential, θ (θ  ) is the Ĝ(1,2) = ĝ(1,2) + d3 d4 ĝ(1,3)(3,4) ˆ Ĝ(4,2), (8)
polar angle of the vector k(k ), 1 represents the 2×2 identity
matrix, and i is the imaginary unit. We assume that λ < 1, where ĝ is the bare contour-ordered Green’s function matrix
allowing us to do perturbation theory in this parameter in what that does not consider the impurity potential V , and d3 =
follows. In this work we use a simplified notation for the ˆ is
dr 3 dt3 , etc. The self-energy matrix 
spin-orbit contribution to the impurity potential as introduced  K 
  R −1
in Eq. (2); however, we note that λσz sin γ represents a =R
ˆ R , (9)
A 0
term that is frequently written as λ0 σ · k × k , where λ0 is
a material-specific constant and |k| = |k | = kF . Hence it is where  K ,  R , and  A are the Keldysh, retarded, and
important to emphasize the fact that λ ∝ kF2 = 4π ne is a linear advanced self-energies. After the average over impurity
function of the electron density ne , which experimentally can configurations, the retarded component of the self-energy
be tuned by changing the gate voltage or the temperature [92]. in the Born approximation is shown in Fig. 2(c), while the
The total impurity potential in the real space is Keldysh components of the self-energy leading to quantum

V (r) = U (r − RI ), (3)
I

where I is a summation over all impurities and R I is
the impurity coordinate. Following an average over random
impurity configurations, the total impurity potential in the
αβ αβ ηζ
reciprocal space Vkk becomes Vkk = 0 and Vkk Vk1 k =
1
αβ ηζ
ni Ukk Uk1 k , where α, β, η, and ζ are spin indices, and ni
1
represents the impurity concentration. For strong screening
[93–96] we may assume U = Z/NF , where NF = εF /(2π A2 )
is the density of states at εF and Z ≡ 1 is the atomic number
of the impurities.

C. Summary of the Keldysh formalism


In the following, we follow a version of the well-established
formalism due to Keldysh [97] to calculate the electrical
conductivity. The central quantity in the Keldysh formalism
is the Keldysh Green’s function GK , whose physical content
is analogous to that of the density matrix ρ in the quantum
Liouville equation. In the Keldysh representation, the Keldysh
Green’s function GK , together with the retarded and advanced
Green’s functions GR and GA , form a 2×2 matrix:
 
0 GA
Ǧ = . (4)
GR GK
In the following, we use τi (i = x,y,z) instead of σi to
designate the Pauli matrices in Keldysh space, in order to
avoid the confusion with the spin space. The Green’s function
matrix Ǧ is related to the contour-ordered Green’s function Ĝ
through the equation FIG. 2. The diagrams for the weak (anti-)localization conductiv-
ity of Dirac fermions. (a) Define Green’s function as arrowed solid
Ĝ = R Ǧ R −1 , (5) lines in which Greek letters are spin indices. (b) Definition of dashed
√ lines: impurities lines expressed in both retarded and advanced cases.
with R = (1 + iτy )/ 2, where the contour takes into account
(c) The retarded self-energy in the first-order Born approximation,
time evolution in both directions. In real space, the time where GR0 is the bare retarded Green’s function. (d, e) Keldysh
evolution of Ĝ is given by K,b
self-energies (k,γ K,R
  δ and  k,αβ ) of maximally crossed diagrams in
i h̄ ∂t1 − H r 1 t1 Ĝ(1,2) = τz δ(1 − 2), (6) the bare and the retarded dressed cases, respectively, where  is the
 ←− − 
← Cooperon structure factor. (f) The Bethe-Salpeter equation for the
Ĝ(1,2) −i h̄ ∂ t2 − H r 2 t2 = τz δ(1 − 2), (7) twisted Cooperon structure factor . ˜

045307-3
LIU, HANKIEWICZ, AND CULCER PHYSICAL REVIEW B 96, 045307 (2017)

interference are in Figs. 2(d) and 2(e) [98]. In Fig. 2, the We have retained the matrix structure of all Green’s
spin indices are kept, not the indices in the Green’s function functions and note that (i) a linear-in-λ term appears in the
matrix. The diagrams appearing in Fig. 2 are important for elastic scattering time τ and transport time τtr , which is missed
the upcoming calculations of the Drude conductivity and the if the matrix structure is neglected and has not been discussed
quantum-interference correction to the conductivity. previously in Dirac/Weyl semimetals; (ii) a nontrivial coupling
Based on the above, omitting a series of cumbersome between the singlet and the triplet channels in the Cooperon
technical details, in an external electric field E that is constant emerges below when first-order impurity spin-orbit coupling
and uniform, the quantum kinetic equation for GK may be terms are taken into account in backscattering processes;
written in the form [99] (iii) in order to obtain the correct result in (i) and (ii), it is neces-

sary to consider explicitly the angular dependence of the spin-
εk + (i/h̄) H0k ,Gεk + Jεk = (e/h̄)E · ∂ k Gεk ,
∂t GK K K
(10)
orbit coupling terms in the impurity potential, in contrast to
where ∂k ≡ ∂/∂ k, e is the elementary charge (thus the electron the conventional Hikami-Larkin-Nagaoka approach in which
charge is −e), ε is an intermediate energy variable that is the square of these terms is averaged over the Fermi surface.
integrated over (representing nonlocality in time), and the
scattering term Jεk is given by E. Drude conductivity
 R K  R,A,K
Jεk = (i/h̄) εk Gεk − GK εk εk + εk Gεk − Gεk εk . (11)
A K A R K With the self-energies Bn in the Born approximation,
in which the advanced and Keldysh cases are similar to the
We adopt the form GK εk = χ k (Gεk −Gεk ) with χ k a scalar,
R A
retarded case in Fig. 2(c), the Born-approximation scattering
following the reasoning of Ref. [100]. Note that the Wigner term becomes JBn (χk ) = −2 iπ gk χk /τtr , where the transport
transformation is applied on Ĝ(1,2) in order to find the single- scattering time (momentum relaxation time) is
particle Green’s function Ĝεk .
In the following, we will introduce the disorder averaged τtr = 2τ0 /[1 + 3b2 + 2a 2 λ + (5 + 3b2 )λ2 /4]. (16)
Green’s functions in Sec. II D, the Drude conductivity in √
The elastic scattering length is e = Dτ , where D = vF2 τtr /2
Sec. II E, and quantum-interference conductivity in Sec. II F. is the diffusion constant and vF = av is the Fermi velocity.
We assume that e is much shorter than the phase coherence
D. Self-consistent retarded and advanced Green’s functions length φ as required in diffusive quantum transport. To the
The arrowed lines in Fig. 2 are Green’s functions that are leading order of τtr−1 the solution of Eq. (10) is χEk = [2e E ·
defined in Fig. 2(a), and the dashed lines in Fig. 2 correspond k̂ τtr /h̄]δ(k − kF ), where k̂ is the unit vector along k, and the
to the impurity scattering, whose definitions in the retarded Drude conductivity takes the form
and the advanced cases are displayed in Fig 2(b). The bare Dr
σxx = (e2 / h)(vakF τtr /2), (17)
Green’s function is
ε1 + H0k ε→εk 1 + H0k /εk gk where h is the Planck constant.
G0k = 2 −−−→ = , (12)
ε − εk2 2(ε − εk ) ε − εk
F. Quantum interference correction to the conductivity
where gk = [1 + H0k /εk ]/2. Under the first-order Born ap-
proximation, the retarded self-energy Bn R
[see Fig. 2(c)] is The quantum interference between two time-reversed
expressed by closed trajectories will generate three different contributions to
the conductivity, whose Keldysh self-energies are (bare case)
dk  dθ  bK , (retarded dressed case) RK , and (advanced dressed case)
R
Bn = Ukk GR0,k Uk k
2π 2π AK . The Keldysh self-energy diagrams of bK and RK are
 2 shown in Figs. 2(d) and 2(e), respectively. In these diagrams,
(1 + βσz ) 1 + λ2 + αλ(σx cos θ + σy sin θ ) the maximally crossed diagram (Cooperon structure factor) 
= , (13)
2iπ h̄/ni |U|2 NF is proportional to 1/| Q|2 , where Q is the sum of the incoming
and the disorder-averaged (retarded and advanced) Green’s and the outgoing wave vector. The divergence of  Q at
functions are | Q| → 0 indicates the primary contribution into the quantum-

gk
∗ interference conductivity comes from the backscattering. The
R −1 R
GRk = 1 − GR0k Bn G0k = = GAk , (14) quantum-interference self-energies (bK , RK , and AK ) will
ε − εk + 2τ
ih̄
generate the quantum-interference scattering term Jqi (χEk )
where GR0k = gk /(ε − εk + i 0+ ), ∗ denotes the Hermitian that is balanced by JBn (χqi,k ) as follows:
conjugate, and GAk is the advanced Green’s function. The JBn (χqi,k ) = −Jqi (χEk ), (18)
elastic scattering time τ is
where the right-hand side plays the role of a driving term.
τ = τ0 /[(1 + λ2 /2)(1 + b2 ) + λa 2 ], (15) The term χqi,k found from this equation leads to the quantum-
interference conductivity correction
where 1/τ0 = π ni |U|2 NF /h̄. In Eq. (15), the linear λ term
comes from the chirality interplay between the band Hamilto-  2 2 3 
b
σxx = − e vF τ NF ηv /h̄
qi 2 2
C Q + 2C RQ , (19)
nian and the spin-orbit impurities and only appears when we Q
maintain the matrix structure. Compared with the topological
insulator case [101], the opposite chirality number in the band where Q ≡ d Q/(2π ) , and ηv = τtr /τ is the current vertex
2

Hamiltonian gives the opposite sign of the linear term. renormalization factor. C bQ and C RQ are the bare Cooperon

045307-4
QUANTUM TRANSPORT IN WEYL SEMIMETAL THIN . . . PHYSICAL REVIEW B 96, 045307 (2017)

TABLE I. Up to the order of λ2 , all nonvanishing angular- TABLE II. Explicit expressions of αi,j and γi,j components.
dependent Cooperon structure factor components mn with their The definitions of fi (b) are f1 (b) = [67 + 227b2 + 143b4 − 225b6 −
expressions. 679b8 − 1127b10 − 427b12 − 27b14 ]/(1 + b2 )4 , f2 (b) = [11 −
252b2 − 56b3 − 58b4 − 16b5 + 428b6 − 24b7 − 289b8 ]/(1 + b2 ),
O(λ) Expression O(λ2 ) Expression f3 (b) = 7 + 5b + 5b2 − b3 , f4 (b) = [4 + 7b + 11b2 + 5b3 + 5b4 ]/
(1 + b2 ), f5 (b) = 41 + 154b2 − 17b4 − 52b6 − 729b8 − 358b10 −
[(1 + b1 N0 )b1 N1 −1,−1 b1 N1 0,−1 63b12 , and f6 (b) = [43 − 152b + 215b2 − 688b3 − 170b4 − 64b5 −
−1,0
+b1 N2 b−1 N−1 ]00 −1,1 b1 N1 0,1 890b6 + 944b7 − 641b8 − 808b9 + 163b10 ]/[(1 + b2 )2 (1 + b)2 ].
[(1 + b−1 N0 )b−1 N−1 1,1 b−1 N−1 0,1
1,0 j =0 j =1 j =2
+b−1 N−2 b1 N1 ]00 1,−1 b−1 N−1 0,−1
a4 4a 2 b2 (3 + b2 ) b2 f1 (b)
00 [N−1 b1 (1 + N0 b1 ) 2,0 b−2 N−2 00 α1,j − −
0,1
+N1 b−1 N−2 b1 ] −2,0 b2 N2 00 2(1 + 3b2 ) (1 + 3b2 )3 (1 + 3b2 )4
4b2 (1 + b2 ) 2(1 − b)b2 f3 (b) b2 f2 (b)
00 [N1 b−1 (1 + N0 b−1 ) 0,−2 00 N2 b−2 α2,j −
0,−1 (1 + 3b2 )2 (1 + 3b2 )3 (1 + 3b2 )4
+N−1 b1 N2 b−1 ] 0,2 00 N−2 b2
4b2 (1 + b2 ) 2(1 + b)b2 f3 (−b) b2 f2 (−b)
α3,j −
(1 + 3b2 )2 (1 + 3b2 )3 (1 + 3b2 )4
2a 2 b2 2b (1 − 14b2 − 3b4 )
2
b2 f5 (b)
and the retarded-dressed Cooperon, respectively, where C Q = γ1,j − − 2
K αγ (1 + b2 )2 (1 + b2 )3 2a (1 + b2 )6
Re[k,γ α g k ] in general, and α and γ are spin indices. The 8(1 − b)2 b2 8(1 − b)b3 f4 (b) b2 f6 (b)
Einstein summation rule over repeated indices is used through- γ2,j −
(1 + b)2 (1 + 3b2 ) (1 + b)2 (1 + 3b2 )2 (1 + 3b2 )3
out this work. Note that the advanced dressed Cooperon
contributes the exactly same amount as its retarded dressed 8(1 + b)2 b2 8(1 + b)b3 f4 (−b) b2 f6 (−b)
γ3,j − −
counterpart. (1 − b)2 (1 + 3b2 ) (1 − b)2 (1 + 3b2 )2 (1 + 3b2 )3
qi
In order to solve for σxx explicitly, the Cooperon structure
factor is twisted, and the resulting structure factor satisfies
δρ
˜ νβ (k1 ,k, Q) = νρδβ
(k1 , Q − k, Q), since the time-reversal over n from 2e /2B to 2φ /2B gives rise to the field-dependent
qi Dr
symmetry is preserved and here we reverse the GA lines. δ, ν, conductivity σxx (B). Since σxx is nearly independent of B, the
qi qi
β, and ρ are spin indices. From Fig. 2(f), the twisted Cooperon magnetoconductivity σ (B) = σxx (B) − σxx (0) is described
structure factor ˜ is ladderlike and satisfies the Bethe-Salpeter by a general expression [102]
equation,  αi e2  Bφ,i 1

Bφ,i

σ (B) =  + − ln , (22)
˜ mn = b−m δm,−n + b−m N−m+l ˜ ln , (20) i=1,2,3
πh |B| 2 |B|

where (k ˜ 1 ,k, Q) is expanded as ˜ mn eimθ1 +inθ with where Bφ,i = (h̄/4e)(1/2φ + 1/2i ) is the effective phase co-

θ1 = θk1 , Nn = k einθk GRk ⊗ GAQ−k , and b(k1 ,k, Q) ≈ herence field, and  is the digamma function. When x 1,
Uk1 k ⊗ U−k1 ,−k = bn ein(θ−θ1 ) is the bare vertex. In Eq. (20), y(x) = √
(1/|x| + 1/2) − ln (1/|x|) ∝ x 2 , and when x  10,
we neglected the spin indices and introduced [A ⊗ B]δγ =
αβ y(x) ∝ x [90].
Aαδ Bβγ for simplicity. The leading solution in Eq. (20) is ˜ 00
and, up to O(λ2 ), all angular-dependent Cooperon structure III. RESULTS
factor components are listed in Table I. It is more convenient In the following, the results for the magnetoconductivity,
to transform Eq. (20) into the singlet-triplet (ST) basis and zero-field conductivity, and sign reversal of the σy term
then to extract the singlet and triplet contributions separately. in Eq. (1) are discussed in Secs. III A, III B, and III C,
Also, note that it is crucial to include all off-diagonal terms respectively.
in [˜ 00 /b0 ]−1 , which are related to the correlations between
singlet and triplet channels.
A. Magnetoconductivity
Zero-field conductivity.—The zero-field conductivity can
be obtained by integrating Q ≡ | Q| between 1/e and 1/φ : We would like to shed light on the role and importance
of the linear-in-λ terms in the conductivity, which have
 αi e2 1/2i + 1/2φ not been discussed previously. The linear λ terms come
qi
σxx (0) = ln , (21)
i=1,2,3
πh 1/2i + 1/2e from the interplay between angular dependences of the band
Hamiltonian and of the spin-orbit scattering. To this end, in
where i = 1,2, and 3 are the singlet, the triplet-up, and the this section we first artificially turn off the angular dependence
triplet-down channel indices, respectively, αi = αij λj is the of the spin-orbit scattering in order to reproduce known results
weight of channel i, and 1/2i = γi /(vF τ )2 is the effective [89] and then turn them back on and analyze their effect.
coherence length with γi = γij λj . The explicit expressions of At first, we turn off the angular dependence of the spin-orbit
αi,j and γi,j are listed in Table II. In the massless limit (b = 0), scattering. This means that we will reduce the θ  integral of the
α1 = −1/2 and γ1 = 0, matching the findings of Ref. [101]. product
of the impurity potentials and the Green’s function,
Magnetoconductivity formula. In an out-of-plane magnetic see dθ  /2π in Eq. (13) for example, into the product of
field B, Q2 will be quantized as Q2n = (n √+ 1/2)/B , where
2
the θ  integral of the impurity potentials (which corresponds
n is the quantization number and lB = h̄/4eB. Summing to the angular-independent contribution of the spin-orbit

045307-5
LIU, HANKIEWICZ, AND CULCER PHYSICAL REVIEW B 96, 045307 (2017)

impurity scattering) and that of the Green’s function. In this


way, Ukk GR0,k Uk k = ni |U|2 k [GR0,k + σz GR0,k σz λ2 /2] in the
Born self-energy and bn=0 → 0 in the bare vertex. As a result, 0.5
the scattering time is τ → τ0 /[(1 + λ2 /2)(1 + b2 )] and the 1000 meV
transport time is

Δσ (e2/h)
100 meV
τ tr → 2τ0 /[1 + 3b2 + (3 + b2 )λ2 /2], (23) 0.0 40 meV

which agrees with Ref. [89]. However, 1/τ tr , due to angular 20 meV
structure of the diffusion, is very different between Ref. [89] 10 meV
and our paper. Even more significantly, Ref. [89] misses linear −0.5
M= 0
in λ terms in both scattering and transport times, which as we
argue below leads to nontrivial physics.
The magnetoconductivity obtained without considering the 0.0 0.1 0.2 0.3 0.4 0.5
angular dependence of the spin-orbit scattering is plotted in
B(T)
Fig. 3. In Fig. 3(a), in the absence of the spin-orbit scattering,
the crossover between the WAL and WL is displayed when FIG. 4. The magnetoconductivity σ (B) plots at λ = 0.5 for
modulating M, as featured in Ref. [83]. In Fig. 3(b), the different masses M with considering the angular-dependent part in the
suppression of weak localization in the presence of the spin- spin-orbit scattering. Compared with Fig. 3(b), it is shown that there
orbit impurities is shown. Moreover, in Fig. 3, the blue line that is more suppression of the WL channel when the angular-dependent
corresponds to the massless limit changes noticeably when part of the spin-orbit scattering is taken into account.
the spin-orbit scattering is included. In the end, comparison
between Figs. 3(a) and 3(b) confirms the suppressed WL due
to the spin-independent part of the spin-orbit scattering, as M. Moreover, the WAL channel number (α1 ) in the massless
argued in Ref. [89]. Noticeably, in this approach, the WAL limit is exactly 1/2, which also satisfies the universal condition
channel number (α1 ) that emerges in the massless limit is protected by time-reversal symmetry.

α1 ≡ −1/2. (24)
B. Zero-field conductivity
The absence of any λ dependence in Eq. (24) does not At zero magnetic field, the electrical conductivity includes
violates the universality requirement of the symplectic the classical Drude conductivity σ Dr and the quantum-
magnetoconductivity: the weak antilocalization channel num- interference conductivity σ qi (0), with σ Dr the dominant term.
ber is expected to be universal up to the second order in λ Zero quantum conductivity. The zero-field quantum con-
in the massless limit. However, this cannot in general justify ductivity σ qi (0) can be used to locate the crossover between
neglecting the linear-λ term. WAL and WL, and then separate out the regimes in which
Next, we switch on the λ-linear term in the Green’s they occur, as shown in Fig. 5. In Fig. 5 the λ-linear terms
functions, with the results shown in Fig. 4. Comparing Fig. 4 are taken into account from the start. A positive/negative
with Fig. 3(b), a much stronger suppression of the WL channel σ qi (0) corresponds to WAL/WL, respectively. From Fig. 5, in
(α2 ) is seen due to the λ-linear terms, in particular for large the absence of spin-orbit scattering, the WAL/WL transition

FIG. 3. The magnetoconductivity plots obtained without consid-


ering the angular-dependent (∝ λ) part in the Green’s functions.
(a, b) Magnetoconductivity σ (B) for different masses M at weak
spin-orbit scattering (λ = 0) and strong spin-orbit scattering (λ =
0.5). Comparison between the two panels confirms the suppression
of WL due to spin-orbit scattering. For the effect of the linear-
in-λ terms, see Fig. 4(b). Parameters are A = 300 meV nm, ne = FIG. 5. Zero-field quantum conductivity σ qi (0) in terms of a/b
0.01 nm−2 ,ni = 0.000 1 nm−2 , and φ = 500 nm. Our results by the and λ. The unit of the conductivity is e2 / h with the color bar on the
Keldysh formalism are qualitatively consistent with those by the Kubo right. The blue dashed line separates the WAL and WL regimes. The
formula [89], verifying the validity of our approach. parameters here are the same as in Fig. 3.

045307-6
QUANTUM TRANSPORT IN WEYL SEMIMETAL THIN . . . PHYSICAL REVIEW B 96, 045307 (2017)

occurs at AkF /M = a/b ∼ 3.3. Naively, one would expect that


the unitary symmetry point occurs for AkF = M. However, this
limit is shifted due to nontrivial coupling between singlet and
triplet Cooperon channels. As the strength of the spin-orbit
scattering is increased, the WAL/WL transition happens at
smaller a/b. The spin-orbit scattering pushes the WAL/WL
boundary, so the WAL regime becomes far broader. On the
other hand, the behavior at large mass is similar to the 2D
conventional electron gas case where the spin-orbit scattering
drives the system from the WL to WAL regimes.
Carrier density dependence. One interesting feature of
Ref. [86] is the detailed study of the carrier density depen-
dence. Although those studies were only for the 2D TIs in
Ref. [86], it motivated authors to investigate the carrier density
dependence of the zero-field conductivity. Note that a back
gate is usually applied to adjust the carrier density, whereas
Hall measurements are performed to measure the density.
We introduce λ = λc ne , since λ ∝ kF2 gives a linear density
dependence. Note that λc = 4π λ0 where λ0 is introduced in
Sec. II B.
At zero temperature, the residual conductivity [103] is
qi
σ (0) = σxx
Dr
+ σxx , and φ in Eq. (21) should be replaced by the
sample size L, because the former will be divergent if T → 0.
By tuning the gate voltage, the carrier density dependence
of σ (0) can be experimentally measured and the strength of
the spin-orbit scattering can be extracted from σ (0) . This is
FIG. 6. The carrier density dependence of the total conductivity
one of the central arguments of this work. To date in the σ tot = σ (0) at the massless (M = 0) (a) and massive (M = 100 meV)
literature only λ2 terms in τ or τtr have been identified, which (b) limits. Parameters are the same as Fig. 3, and L is set to be equal
arise when the spin-orbit scattering terms are averaged over to φ for simplicity. Note that λ = λc ne and AkF = 106 meV for
directions in momentum space assuming a circular Fermi ne = 0.01 nm−2 .
surface. It contributes a negligible n2e dependence to σ (0) ,
which is effectively linear in ne due to the Drude part. When the
nontrivial linear-λ term is taken into account in τ and τtr , which small density is displayed, because the effect of the mass term
is caused by the noncommutativity of the band structure and becomes more significant when the carrier density decreases
the random spin-orbit impurity field, the WAL/WL correction and the WAL channel is suppressed. In the large mass limit [see
has a more pronounced dependence on the carrier number Fig. 7(c)], a negative conductivity correction (WL) is expected.
density. This provides a new possibility to extract the spin-orbit
scattering constant from the density dependence of σ (0) .
The carrier density dependence of the total conductivity C. Sign reversal of the σ y term in Eq. (1)
for different values of the spin-orbit scattering strength is Reference [104] suggests an alternative model for describ-
shown in Fig. 6. In Fig. 6(a), the massless limit is considered ing Dirac fermions:
and the density dependence deviates considerably from the
linear form of the Drude contribution (in the approximation H0k = A(σx kx − σy ky ) + Mσz . (25)
use here of short-range impurities). In Fig. 6(b), the mass term
(M = 100 meV) suppresses the density dependence due to the Compared with Eq. (1), Eq. (25) gives the same energy
extrinsic spin-orbit scattering. dispersion but with an opposite sign on the ky momentum,
In light of this finding, and of the fact that the WL/WAL which means a chirality reversal. Since the angular dependence
contribution can be separated from the Drude contribution of the spin-orbit scattering is taken into account, the change
in an experiment, it is enlightening to plot the carrier in the angular dependence of the band Hamiltonian H0k will
density dependence of the quantum-interference part of the also affect the results in Secs. III A and III B. In this section,
conductivity alone. The carrier density dependence of the we will show the effect of the sign reversal of the σy term,
quantum-interference part of the conductivity is plotted in and with a comparison between two cases, the effect of
Fig. 7 for three different masses: M = 0, 10 meV, and the angular-dependent part of the spin-orbit scattering might
100 meV. In the massless limit [Fig. 7(a)], for λc = 0 the become more clear.
qi
contribution σxx follows the logarithmic dependence expected For the model described by Eq. (25), the scattering time is
of a two-dimensional electron gas. As λc increases this
logarithm becomes nearly flat at large enough densities. This τ = τ0 /[(1 + λ2 /2)(1 + b2 ) − λa 2 ], (26)
altered density dependence arises from the linear terms in
the scattering and transport times. In the small mass limit which matches Ref. [101] in the massless limit and has an
[Fig. 7(b)], a sharp suppression of the conductivity at the opposite sign in the linear λ term when compared with Eq. (15),

045307-7
LIU, HANKIEWICZ, AND CULCER PHYSICAL REVIEW B 96, 045307 (2017)

IV. DISCUSSION
The focus of our work is to analyze the strong effect of linear
terms in the spin-orbit scattering strength on weak localization
and antilocalization in WSM thin films. In contrast to previous
works [89], where only the angle-averaged (second-order in λ)
contribution is considered in the scattering time τ , we have
demonstrated that a correction to first order in the spin-orbit
scattering strength is present, which has a strong dependence
on the relative angle between the incoming and outgoing wave
vectors. It emerges from the angular integration of the chiral
Dirac fermion Hamiltonian and the disorder Hamiltonian
describing spin-orbit scattering. This correction is 1 order of
magnitude larger than that found in Ref. [89], making it a
significant perturbation.
In the presence of strong spin-orbit scattering, the sup-
pression of the WL channel is expected on general grounds,
since this channel corresponds to one of the spin-polarized
triplet channels. Spin-orbit scattering randomizes the spin
polarization in the triplet channels and eventually suppresses
the contribution due to these. The suppression is already seen,
even when only the isotropic part of the spin-orbit scattering
(∝ λ2 ) is taken into account, as in Ref. [89]. The λ linear terms
enhance this suppression, as shown in Fig. 4.
An interesting qualitative change due to the extrinsic
spin-orbit scattering is observed in the transition from WL
to WAL, occurring as a function of the mass. This may be
seen in Fig. 5. At small and large values of the mass, as
expected, the extrinsic spin-orbit scattering has little effect
on the WL/WAL correction. At large mass, moreover, the
extrinsic spin-orbit scattering also plays a negligible role in
the momentum relaxation time. However, when the system
is close to the unitary symmetry class, the extrinsic spin-
orbit scattering is critical in determining whether the system
experiences WL or WAL. A rough estimate of WL/WAL
transition line in Fig. 5 is a/b ≈ 1/(0.3 + 0.8 · λ), which
FIG. 7. The carrier density dependence of the quantum- exists when ne ∈ [0.01,0.02] nm−2 and φ ∈ [200,500] nm.
interference conductivity σ qi at the massless (M = 0) (a), small
This fitting equation can provide a semiempirical formula to
mass (M = 10 meV) (b), and large mass (M = 100 meV) (c) cases.
extract the λ, and the extraction of λ can be examined by
Parameters here are the same as Fig. 6.
another method as introduced in the following.
At small mass, the extrinsic spin-orbit scattering makes
and the transport time becomes an important contribution to the momentum relaxation time,
causing it to acquire a strong density dependence. This
τtr = 2τ0 /[1 + 3b2 − 2a 2 λ + (5 + 3b2 )λ2 /4], (27) manifests itself in a flattening of the 2D WAL correction as
a function of carrier number density. To extract the spin-orbit
except for an opposite sign in the linear λ term from Eq. (16).
scattering strength λ from σ qi (0) in the massless limit, it is
These opposite signs of linear λ terms come from the opposite
possible to follow the fitting equation
sign of the σy term in the band Hamiltonian, which implies
that linear λ terms in τ and τtr are due to the interplay between σ qi (0) = a0 ln ne + b0 ne . (28)
the band Hamiltonian and the spin-orbit scattering.
Furthermore, the WAL channel number (α1 ) remains the Here we assume zero temperature, so φ is replaced by the
same, while the rest of the channel numbers (i = 2,3) become sample size L as mentioned above. The extracted coefficient b0
αi = αi,j (−λ)j and the effective coherence length coefficients yields the spin-orbit scattering strength λ = −b0 ne /(3e2 /2π h)
become γi = γi,j (−λ)j , where αi,j and γi,j are listed in when λ 1. The short-range impurity potential that we have
Table II. Since j  2, the changes only happen on the linear λ used in this article needs to be replaced by a long-range
term. The unchanged WAL channel number indicates that the Coulomb potential (or a generic long-range potential). In
sign reversal of the σy term does not affect the WAL behavior, the latter case, the Bethe-Salpeter equation (20) will involve
because there is no net spin in the singlet channel. The λ → −λ an additional angular integration over the impurity potential,
phenomena of triplet channel numbers under the sign reversal in which case it is no longer solvable in closed form.
of the σy term is expected because of the net spin of triplet One possible rough estimate for σ qi (0) can be obtained by
channels. retaining the form of Eq. (21) while substituting the same

045307-8
QUANTUM TRANSPORT IN WEYL SEMIMETAL THIN . . . PHYSICAL REVIEW B 96, 045307 (2017)

τ for the long-range potential, as was found above for the WL or WAL and strongly affect the density dependence of
short-range case. In the long-range case, the normal impurity the WAL correction in the massless limit. In addition, these
potential U for 2D massless Dirac fermions becomes Ulong = terms restore the universality of the WAL channel number in
Ze2 /[2εr kF sin2 (γ /2)] [105], where εr is the material-specific the massless limit, which appears to be violated when only the
dielectric constant. Thus, the long-range potential Ulong will second-order terms are considered.
have the same ne dependence as the short-range case, and the
fitting equation (28) will be still correct. ACKNOWLEDGMENTS
This research was supported by the Australian Re-
search Council Centre of Excellence in Future Low-Energy
V. CONCLUSIONS
Electronics Technologies (Project No. CE170100039) and
In conclusion, we have studied theoretically the weak funded by the Australian Government, NBRPC (Grant No.
localization/antilocalization correction to the conductivity of 2015CB921102), the National Key R & D Program (Grant
ultrathin films of Weyl semimetals. We have considered scalar No. 2016YFA0301700), and the National Natural Science
and spin-orbit scattering mechanisms on the same footing. Foundation of China (Grants No. 11534001, No. 11474005,
The noncommutativity of the matrix Green’s functions and and No. 11574127). E.M.H. acknowledges financial support
spin-dependent impurity potential gives rise to terms linear by the German Science Foundation (DFG) via SFB 1170 “To-
in the extrinsic spin-orbit scattering strength, which play an CoTronics” and the ENB (Elite Netzwerk Bayern) Graduate
important role in determining whether the system experiences School on Topological Insulators.

[1] L. Balents, Physics 4, 36 (2011). [16] X. Xiao, S. A. Yang, Z. Liu, H. Li, and G. Zhou, Sci. Rep. 5,
[2] A. Vishwanath, Physics 8, 84 (2015). 7898 (2015).
[3] R. Cianci, L. Fabbri, and S. Vignolo, Eur. Phys. J. C 75, 478 [17] C.-Z. Li, L.-X. Wang, H. Liu, J. Wang, Z.-M. Liao, and D.-P.
(2015). Yu, Nat. Commun. 6, 10137 (2015).
[4] D. Xiao, M. C. Chang, and Q. Niu, Rev. Mod. Phys. 82, 1959 [18] H. Li, H. He, H.-Z. Lu, H. Zhang, H. Liu, R. Ma, Z. Fan, S.-Q.
(2010). Shen, and J. Wang, Nat. Commun. 7, 10301 (2016).
[5] S. M. Young, S. Zaheer, J. C. Y. Teo, C. L. Kane, E. J. Mele, [19] A. A. Burkov and Y. B. Kim, Phys. Rev. Lett. 117, 136602
and A. M. Rappe, Phys. Rev. Lett. 108, 140405 (2012). (2016).
[6] Z. Wang, Y. Sun, X. Q. Chen, C. Franchini, G. Xu, H. Weng, [20] L.-X. Wang, C.-Z. Li, D.-P. Yu, and Z.-M. Liao, Nat. Commun.
X. Dai, and Z. Fang, Phys. Rev. B 85, 195320 (2012). 7, 10769 (2016).
[7] Z. Wang, H. Weng, Q. Wu, X. Dai, and Z. Fang, Phys. Rev. B [21] J. Xiong, S. Kushwaha, J. Krizan, T. Liang, R. J. Cava, and
88, 125427 (2013). N. P. Ong, Europhys. Lett. 114, 27002 (2016).
[8] Z. K. Liu, B. Zhou, Y. Zhang, Z. J. Wang, H. M. Weng, [22] B. Zhao, P. Cheng, H. Pan, S. Zhang, B. Wang, G. Wang, F.
D. Prabhakaran, S.-K. Mo, Z. X. Shen, Z. Fang, X. Dai, Z. Xiu, and F. Song, Sci. Rep. 6, 22377 (2016).
Hussain, and Y. L. Chen, Science 343, 864 (2014). [23] X. Wan, A. M. Turner, A. Vishwanath, and S. Y. Savrasov,
[9] Z. K. Liu, J. Jiang, B. Zhou, Z. J. Wang, Y. Zhang, H. M. Phys. Rev. B 83, 205101 (2011).
Weng, D. Prabhakaran, S.-K. Mo, H. Peng, P. Dudin, T. Kim, [24] K. Y. Yang, Y. M. Lu, and Y. Ran, Phys. Rev. B 84, 075129
M. Hoesch, Z. Fang, X. Dai, Z. X. Shen, D. L. Feng, Z. Hussain, (2011).
and Y. L. Chen, Nat. Mater. 13, 677 (2014). [25] B. Singh, A. Sharma, H. Lin, M. Z. Hasan, R. Prasad, and A.
[10] M. Neupane, S.-Y. Xu, R. Sankar, N. Alidoust, G. Bian, C. Bansil, Phys. Rev. B 86, 115208 (2012).
Liu, I. Belopolski, T.-R. Chang, H.-T. Jeng, H. Lin, A. Bansil, [26] D. Bulmash, C.-X. Liu, and X.-L. Qi, Phys. Rev. B 89, 081106
F. Chou, and M. Z. Hasan, Nat. Commun. 5, 3786 (2014). (2014).
[11] S. Borisenko, Q. Gibson, D. Evtushinsky, V. Zabolotnyy, B. [27] S.-Y. Xu, I. Belopolski, N. Alidoust, M. Neupane, G. Bian, C.
Büchner, and R. J. Cava, Phys. Rev. Lett. 113, 027603 (2014). Zhang, R. Sankar, G. Chang, Z. Yuan, C.-C. Lee, S.-M. Huang,
[12] H. Yi, Z. Wang, C. Chen, Y. Shi, Y. Feng, A. Liang, Z. Xie, S. H. Zheng, J. Ma, D. S. Sanchez, B. Wang, A. Bansil, F. Chou,
He, J. He, Y. Peng, X. Liu, Y. Liu, L. Zhao, G. Liu, X. Dong, J. P. P. Shibayev, H. Lin, S. Jia, and M. Z. Hasan, Science 349,
Zhang, M. Nakatake, M. Arita, K. Shimada, H. Namatame, M. 613 (2015).
Taniguchi, Z. Xu, C. Chen, X. Dai, Z. Fang, and X. J. Zhou, [28] H. Weng, C. Fang, Z. Fang, B. A. Bernevig, and X. Dai,
Sci. Rep. 4, 6106 (2014). Phys. Rev. X 5, 011029 (2015).
[13] J. Xiong, S. K. Kushwaha, T. Liang, J. W. Krizan, M. [29] S. M. Huang, S. Y. Xu, I. Belopolski, C. C. Lee, G. Chang,
Hirschberger, W. Wang, R. J. Cava, and N. P. Ong, Science B. K. Wang, N. Alidoust, G. Bian, M. Neupane, C. Zhang,
350, 413 (2015). S. Jia, A. Bansil, H. Lin, and M. Z. Hasan, Nat. Commun. 6,
[14] J. H. Pixley, P. Goswami, and S. Das Sarma, Phys. Rev. Lett. 7373 (2015).
115, 076601 (2015). [30] B. Q. Lv, H. M. Weng, B. B. Fu, X. P. Wang, H. Miao, J. Ma,
[15] S. K. Kushwaha, J. W. Krizan, B. E. Feldman, A. Gyenis, P. Richard, X. C. Huang, L. X. Zhao, G. F. Chen, Z. Fang, X.
M. T. Randeria, J. Xiong, S.-Y. Xu, N. Alidoust, I. Belopolski, Dai, T. Qian, and H. Ding, Phys. Rev. X 5, 031013 (2015).
T. Liang, M. Z. Hasan, N. P. Ong, A. Yazdani, and R. J. Cava, [31] C.-L. Zhang, S.-Y. Xu, I. Belopolski, Z. Yuan, Z. Lin, B. Tong,
APL Mater. 3, 041504 (2015). G. Bian, N. Alidoust, C.-C. Lee, S.-M. Huang, T.-R. Chang,

045307-9
LIU, HANKIEWICZ, AND CULCER PHYSICAL REVIEW B 96, 045307 (2017)

G. Chang, C.-H. Hsu, H.-T. Jeng, M. Neupane, D. S. Sanchez, [59] N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald, and N. P.
H. Zheng, J. Wang, H. Lin, C. Zhang, H.-Z. Lu, S.-Q. Shen, T. Ong, Rev. Mod. Phys. 82, 1539 (2010).
Neupert, M. Zahid Hasan, and S. Jia, Nat. Commun. 7, 10735 [60] A. A. Burkov, Phys. Rev. Lett. 113, 187202 (2014).
(2016). [61] R. Yu, W. Zhang, H. J. Zhang, S. C. Zhang, X. Dai, and Z.
[32] D. Wu, J. Liao, W. Yi, X. Wang, P. Li, H. Weng, Y. Shi, Y. Fang, Science 329, 61 (2010).
Li, J. Luo, X. Dai, and Z. Fang, Appl. Phys. Lett. 108, 042105 [62] C.-Z. Chang, J. Zhang, X. Feng, J. Shen, Z. Zhang, M. Guo,
(2016). K. Li, Y. Ou, P. Wei, L.-L. Wang, Z.-Q. Ji, Y. Feng, S. Ji, X.
[33] A. A. Soluyanov, D. Gresch, Z. Wang, Q. Wu, M. Troyer, X. Chen, J. Jia, X. Dai, Z. Fang, S.-C. Zhang, K. He, Y. Wang, L.
Dai, and B. A. Bernevig, Nature (London) 527, 495 (2015). Lu, X.-C. Ma, and Q.-K. Xue, Science 340, 167 (2013).
[34] Y. Wang, E. Liu, H. Liu, Y. Pan, L. Zhang, J. Zeng, Y. Fu, M. [63] S. A. Yang, H. Pan, W.-K. Tse, and Q. Niu, arXiv:0911.3134.
Wang, K. Xu, Z. Huang, Z. Wang, H.-Z. Lu, D. Xing, B. Wang, [64] M. Liu, C.-Z. Chang, Z. Zhang, Y. Zhang, W. Ruan, K. He,
X. Wan, and F. Miao, Nat. Commun. 7, 13142 (2016). L.-l. Wang, X. Chen, J.-F. Jia, S.-C. Zhang, Q.-K. Xue, X. Ma,
[35] K. Deng, G. Wan, P. Deng, K. Zhang, S. Ding, E. Wang, M. and Y. Wang, Phys. Rev. B 83, 165440 (2011).
Yan, H. Huang, H. Zhang, Z. Xu, J. Denlinger, A. Fedorov, H. [65] A. A. Burkov, Phys. Rev. Lett. 113, 247203 (2014).
Yang, W. Duan, H. Yao, Y. Wu, S. Fan, H. Zhang, X. Chen, [66] X. C. Huang, L. X. Zhao, Y. J. Long, P. P. Wang, D. Chen, Z. H.
and S. Zhou, Nat. Phys. 12, 1105 (2016). Yang, H. Liang, M. Q. Xue, H. M. Weng, Z. Fang, X. Dai, and
[36] Y. Wang, K. Wang, J. Reutt-Robey, J. Paglione, and M. S. G. F. Chen, Phys. Rev. X 5, 031023 (2015).
Fuhrer, Phys. Rev. B 93, 121108 (2016). [67] H. Z. Lu, S. B. Zhang, and S. Q. Shen, Phys. Rev. B 92, 045203
[37] S. Borisenko, D. Evtushinsky, Q. Gibson, A. Yaresko, T. (2015).
Kim, M. N. Ali, B. Buechner, M. Hoesch, and R. J. Cava, [68] A. A. Burkov, Phys. Rev. B 91, 245157 (2015).
arXiv:1507.04847. [69] P. Goswami, J. H. Pixley, and S. Das Sarma, Phys. Rev. B 92,
[38] A. C. Potter, I. Kimchi, and A. Vishwanath, Nat. Commun. 5, 075205 (2015).
5161 (2014). [70] S.-B. Zhang, H.-Z. Lu, and S.-Q. Shen, New J. Phys. 18,
[39] A. A. Burkov, Nat. Mater. 15, 1145 (2016). 053039 (2016).
[40] J. H. Pixley, D. A. Huse, and S. Das Sarma, Phys. Rev. X 6, [71] Y. Li, Z. Wang, P. Li, X. Yang, Z. Shen, F. Sheng, X. Li, Y. Lu,
021042 (2016); 6, 039901(E) (2016). Y. Zheng, and Z.-A. Xu, Front. Phys. 12, 127205 (2017).
[41] S. A. Yang, SPIN 06, 1640003 (2016). [72] T. Louvet, D. Carpentier, and A. A. Fedorenko, Phys. Rev. B
[42] M. König, S. Wiedmann, C. Brüne, A. Roth, H. Buhmann, 95, 014204 (2017).
L. W. Molenkamp, X.-L. Qi, and S.-C. Zhang, Science 318, [73] A. C. Niemann, J. Gooth, S.-C. Wu, S. Bäßler, P. Sergelius,
766 (2007). R. Hühne, B. Rellinghaus, C. Shekhar, V. Süß, M. Schmidt, C.
[43] D. Culcer and R. Winkler, Phys. Rev. B 79, 165422 (2009). Felser, B. Yan, and K. Nielsch, Sci. Rep. 7, 43394 (2017).
[44] J. E. Moore, Nature (London) 464, 194 (2010). [74] Q. I. Yang, M. Dolev, L. Zhang, J. Zhao, A. D. Fried, E.
[45] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010). Schemm, M. Liu, A. Palevski, A. F. Marshall, S. H. Risbud,
[46] X.-L. Qi and S.-C. Zhang, Rev. Mod. Phys. 83, 1057 (2011). and A. Kapitulnik, Phys. Rev. B 88, 081407 (2013).
[47] D. Culcer, Phys. Rev. B 84, 235411 (2011). [75] H.-Z. Lu and S.-Q. Shen, Chin. Phys. B 25, 117202 (2016).
[48] X.-G. Li, G.-F. Zhang, G.-F. Wu, H. Chen, D. Culcer, and Z.-Y. [76] M. Kammermeier, P. Wenk, J. Schliemann, S. Heedt, and T.
Zhang, Chin. Phys. B 22, 097306 (2013). Schäpers, Phys. Rev. B 93, 205306 (2016).
[49] Z. Zhang, X. Feng, M. Guo, K. Li, J. Zhang, Y. Ou, Y. Feng, [77] F. Evers and A. D. Mirlin, Rev. Mod. Phys. 80, 1355 (2008).
L. Wang, X. Chen, K. He, X. Ma, Q. Xue, and Y. Wang, [78] P. A. Lee and T. V. Ramakrishnan, Rev. Mod. Phys. 57, 287
Nat. Commun. 5, 4915 (2014). (1985).
[50] X. Peng, Y. Yang, R. R. P. Singh, S. Y. Savrasov, and D. Yu, [79] T. Koga, J. Nitta, T. Akazaki, and H. Takayanagi, Phys. Rev.
Nat. Commun. 7, 10878 (2016). Lett. 89, 046801 (2002).
[51] H. Liu, W. E. Liu, and D. Culcer, Phys. E (Amsterdam, Neth.) [80] D. Kim, P. Syers, N. P. Butch, J. Paglione, and M. S. Fuhrer,
79, 72 (2016). Nat. Commun. 4, 2040 (2013).
[52] H. Liu, W. E. Liu, and D. Culcer, Phys. Rev. B 95, 205435 [81] M. Mühlbauer, A. Budewitz, B. Büttner, G. Tkachov, E. M.
(2017). Hankiewicz, C. Brüne, H. Buhmann, and L. W. Molenkamp,
[53] D. Culcer, E. H. Hwang, T. D. Stanescu, and S. Das Sarma, Phys. Rev. Lett. 112, 146803 (2014).
Phys. Rev. B 82, 155457 (2010). [82] H.-Z. Lu and S.-Q. Shen, Phys. Rev. Lett. 112, 146601 (2014).
[54] S. Oh, Science 340, 153 (2013). [83] H.-Z. Lu, J. Shi, and S.-Q. Shen, Phys. Rev. Lett. 107, 076801
[55] S.-Y. Xu, C. Liu, S. K. Kushwaha, R. Sankar, J. W. Krizan, I. (2011).
Belopolski, M. Neupane, G. Bian, N. Alidoust, T.-R. Chang, [84] K.-I. Imura, Y. Kuramoto, and K. Nomura, Phys. Rev. B 80,
H.-T. Jeng, C.-Y. Huang, W.-F. Tsai, H. Lin, P. P. Shibayev, 085119 (2009).
F.-C. Chou, R. J. Cava, and M. Z. Hasan, Science 347, 294 [85] H.-T. He, G. Wang, T. Zhang, I.-K. Sou, G. K. L. Wong, J.-N.
(2015). Wang, H.-Z. Lu, S.-Q. Shen, and F.-C. Zhang, Phys. Rev. Lett.
[56] A. Lau, J. van den Brink, and C. Ortix, arXiv:1701.01660. 106, 166805 (2011).
[57] H. Z. Lu, W. Y. Shan, W. Yao, Q. Niu, and S. Q. Shen, [86] G. Tkachov and E. M. Hankiewicz, Phys. Rev. B 84, 035444
Phys. Rev. B 81, 115407 (2010). (2011).
[58] D. A. Abanin and D. A. Pesin, Phys. Rev. Lett. 106, 136802 [87] Y. Yoshimura, W. Onishi, K. Kobayashi, T. Ohtsuki, and K.-I.
(2011). Imura, Phys. Rev. B 94, 235414 (2016).

045307-10
QUANTUM TRANSPORT IN WEYL SEMIMETAL THIN . . . PHYSICAL REVIEW B 96, 045307 (2017)

[88] S. Hikami, A. I. Larkin, and Y. Nagaoka, Prog. Theor. Phys. [97] L. V. Keldysh, Sov. Phys. JETP 20, 1018 (1965).
63, 707 (1980). [98] F. T. Vasko and O. E. Raichev, Quantum Kinetic Theory
[89] W.-Y. Shan, H.-Z. Lu, and S.-Q. Shen, Phys. Rev. B 86, 125303 and Applications (Springer Science and Business Media,
(2012). New York, 2005).
[90] H.-Z. Lu and S.-Q. Shen, Front. Phys. 12, 127201 (2016). [99] J. Rammer and H. Smith, Rev. Mod. Phys. 58, 323
[91] D. Culcer, Y. Yao, A. H. MacDonald, and Q. Niu, Phys. Rev. (1986).
B 72, 045215 (2005). [100] P. Schwab, R. Raimondi, and C. Gorini, Europhys. Lett. 93,
[92] C. Shekhar, A. K. Nayak, Y. Sun, M. Schmidt, M. Nicklas, 67004 (2011).
I. Leermakers, U. Zeitler, Y. Skourski, J. Wosnitza, Z. Liu, Y. [101] P. Adroguer, W. E. Liu, D. Culcer, and E. M. Hankiewicz,
Chen, W. Schnelle, H. Borrmann, Y. Grin, C. Felser, and B. Phys. Rev. B 92, 241402 (2015).
Yan, Nat. Phys. 11, 645 (2015). [102] J. Hellerstedt, M. T. Edmonds, N. Ramakrishnan, C. Liu, B.
[93] T. Ando, J. Phys. Soc. Jpn. 75, 074716 (2006). Weber, A. Tadich, K. M. O’Donnell, S. Adam, and M. S.
[94] I. L. Aleiner and K. B. Efetov, Phys. Rev. Lett. 97, 236801 Fuhrer, Nano Lett. 16, 3210 (2016).
(2006). [103] E. Rossi, J. H. Bardarson, M. S. Fuhrer, and S. Das Sarma,
[95] S. Adam, E. H. Hwang, and S. Das Sarma, Phys. Rev. B 85, Phys. Rev. Lett. 109, 096801 (2012).
235413 (2012). [104] H. Pan, M. Wu, Y. Liu, and S. A. Yang, Sci. Rep. 5, 14639
[96] S. Das Sarma, E. H. Hwang, and H. Min, Phys. Rev. B 91, (2015).
035201 (2015). [105] D. Culcer and R. Winkler, Phys. Rev. B 78, 235417 (2008).

045307-11

You might also like