ArticlePDF Available

Intrinsic optical conductivity of modified-Dirac fermion systems

Authors:

Abstract and Figures

We analytically calculate the intrinsic longitudinal and transverse optical conductivities of electronic systems which govern by a modified-Dirac fermion model Hamiltonian for materials beyond graphene such as monolayer MoS$_2$ and ultrathin film of the topological insulator. We analyze the effect of a topological term in the Hamiltonian on the optical conductivity and transmittance. We show that the optical response enhances in the non-trivial phase of the ultrathin film of the topological insulator and the optical Hall conductivity changes sign at transition from trivial to non-trivial phases which has significant consequences on a circular polarization and optical absorption of the system.
Content may be subject to copyright.
PHYSICAL REVIEW B 89, 115413 (2014)
Intrinsic optical conductivity of modified Dirac fermion systems
Habib Rostami and Reza Asgari*
School of Physics, Institute for Research in Fundamental Sciences (IPM), Tehran 19395-5531, Iran
(Received 25 December 2013; revised manuscript received 24 February 2014; published 12 March 2014)
We analytically calculate the intrinsic longitudinal and transverse optical conductivities of electronic systems
which govern by a modified Dirac fermion model Hamiltonian for materials beyond graphene such as monolayer
MoS2and ultrathin film of the topological insulator. We analyze the effect of a topological term in the Hamiltonian
on the optical conductivity and transmittance. We show that the optical response enhances in the nontrivial phase
of the ultrathin film of the topological insulator and the optical Hall conductivity changes sign at transition from
trivial to nontrivial phases which has significant consequences on the circular dichroism property and optical
absorption of the system.
DOI: 10.1103/PhysRevB.89.115413 PACS number(s): 72.20.i,78.67.n,78.20.e
I. INTRODUCTION
Two-dimensional (2D) materials have been one of the most
interesting subjects in condensed matter physics for potential
applications due to the wealth of unusual physical phenomena
that occur when charge, spin, and heat transport are confined
to a 2D plane [1]. These materials can be mainly classified in
different classes which can be prepared as a single atom thick
layer namely, layered van der Waals materials, layered ionic
solids, surface growth of monolayer materials, 2D topological
insulator solids, and finally 2D artificial systems and they
exhibit novel correlated electronic phenomena ranging from
high-temperature superconductivity, quantum valley or spin
Hall effect to other enormously rich physics phenomena. Two-
dimensional materials can be mostly exfoliated into individual
thin layers from stacks of strongly bonded layers with weak
interlayer interaction and a famous example is graphene and
hexagonal boron nitride [2]. The 2D exfoliates versions of
transition metal dichalcogenides that exhibit properties that
are complementary to and distinct from those in graphene [3].
Optical spectroscopy is a broad field and useful to explore
the electronic properties of solids. Optical properties can be
tuned by varying the Fermi energy or the electronic band
structure of 2D systems. Recently, developed 2D systems such
as gapped graphene [4], thin film of the topological insulator
[5,6], and monolayer of transition metal dichalcogenides [3]
provide the electronic structures with direct band gap signa-
tures. The optical response of semiconductors with direct band
gap is strong and easy to explore experimentally since photons
with energy greater than the energy gap can be absorbed or
omitted. The thin film of the topological insulator, on the other
hand, has been fabricated experimentally by using Sb2Te3slab
[7] and has been shown that a direct band gap can be formed
owning to the hybridization of top and bottom surface states.
Furthermore, a nontrivial quantum spin Hall phase has been
realized experimentally which was predicted previously in this
system [810]. Although pristine graphene and surface states
of the topological insulator reveal massless Dirac fermion
physics, by opening an energy gap they become formed as
massive Dirac fermions. The thin film of the topological
insulator and monolayer transition metal dichacogenides can
*asgari@ipm.ir
be described by a modified Dirac Hamiltonian. A monolayer
of the molybdenum disulfide (ML-MoS2) is a direct band gap
semiconductor [11], however, its multilayer and bulk show
indirect band gap [3]. This feature causes the optical response
in ML-MoS2to increase in comparison with its bulk and
multilayer structures [1216].
One of the main properties of ML-MoS2is a circular
dichroism aspect responding to a circular polarized light where
the left- or right-handed polarization of the light couples only
to the Kor Kvalley and it provides an opportunity to
induce a valley polarized excitation which can profoundly
be of interest in the application for valleytronics [1719].
Another peculiarity of ML-MoS2is the coupled spin valley
in the electronic structure which is owing to the strong
spin-orbit coupling originating from the existence of a heavy
transition metal in the lattice structure and the broken inversion
symmetry, too [20]. These two aspects are captured in a minima
massive Dirac-like Hamiltonian introduced by Xiao et al. [20].
However, it has been shown, based on the tight binding [21,22]
and k.p method [23], that other terms like an effective mass
asymmetry, a trigonal warping, and a diagonal quadratic term
might be included in the massive Dirac-like Hamiltonian. The
effect of the diagonal quadratic term is very important; for
instance, if the system is exposed by a perpendicular magnetic
field, it will induce a valley degeneracy breaking term [21].
The optical properties of ML-MoS2have been evaluated by ab
initio calculations [24] and studied theoretically based on the
simplified massive Dirac-like model Hamiltonian [25], which
is by itself valid only near the main absorbtion edge. A part
of the model Hamiltonian which describes the dynamic of
massive Dirac fermions are known in graphene committee to
have an optical response quite different from that of a standard
2D electron gas. Thus it would be worthwhile to generalize
the optical properties of such systems by using the modified
Dirac fermion model Hamiltonian.
The modified Hamiltonian for ML-MoS2without trigonal
warping effect at Kpoint is very similar to the modified
Dirac equation which has been studied for an ultrathin
film of the topological insulator (UTF-TI) around point
[8,26]. The modified Dirac Hamiltonian reveals nontrivial
quantum spin hall (QSH) and trivial phases corresponding
to the existence and absence of the edge states, respectively.
Those phases have been predicted theoretically [810,27] and
recently observed by experiment [7]. An enhancement of the
1098-0121/2014/89(11)/115413(12) 115413-1 ©2014 American Physical Society
HABIB ROSTAMI AND REZA ASGARI PHYSICAL REVIEW B 89, 115413 (2014)
optical response of UTF-TI has been obtained in the nontrivial
phase [28] and a band crossing is observed in the presence
of the structure inversion asymmetry induced by substrate
[29]. Since the modified Dirac Hamiltonian incorporates an
energy gap and a quadratic term in momentum which both
have topological meaning, it is natural to expect that the
topological term of the Hamiltonian plays an important role in
the optical conductivity. In this paper, we analytically calculate
the intrinsic longitudinal and transverse optical conductivities
of the modified Dirac Hamiltonian as a function of photon
energy. This model Hamiltonian covers the main physical
properties of ML-MoS2and UTF-TI systems in the regime
where interband transition plays a main role. We analyze the
effect of the topological term in the Hamiltonian on the optical
conductivity and transmittance. Furthermore, we show that the
UTF-TI system has a nontrivial phase and its optical response
enhances; in addition, the optical Hall conductivity changes
sign at a phase boundary, when the energy gap is zero. This
changing of the sign has a significant consequence on the
circular polarization and the optical absorbtion of the system.
The paper is organized as follows. We introduce the low-
energy model Hamiltonian of ML-MoS2and UTF-TI systems
and then the dynamical conductivity is calculated analytically
by using Kubo formula in Sec. II. The numerical results for
the optical Hall and longitudinal conductivities and optical
transmittance are reported, and we also provide discussions
with circular dichroism in both systems in Sec. III.Abrief
summary of results is given in Sec. IV.
II. THEORY AND METHOD
The low-energy properties of the ML-MoS2and other
transition metal dicalcogenide materials can be described by a
modified Dirac equation [2123] and the Hamiltonian around
the Kand Kpoints is given by
Hτs =λ
2τs +λτ s
2σz+t0a0q·στ+
2|q|2
4m0
(α+βσz),
(1)
where the Pauli matrices stand for a pseudospin which indi-
cates the conduction and valence band degrees of freedom, τ=
±denotes the two independent valleys in the first Brillouin
zone, q=(qx,qy) and στ=(τσ
xy). The numerical values
of the parameters will be given in Sec. III.
The UTF-TI system, on the other hand, can be described
by a modified Dirac Hamiltonian around the point with two
independent hyperbola (isospin) degrees of freedom [8,26] and
thus the Hamiltonian reads
Hτ=0+τ
2σz+t0a0q·σ+
2|q|2
4m0
(α+τβσz).(2)
Note that the Pauli matrices in this Hamiltonian stand for
the real spin where spin is rotated by operator U=diag[1,i]
which results in UσxU=−σyand UσyU=σxand the
isospin index of τindicates two independent solutions of
UTF-TI which are degenerated in the absence of the structure
inversion asymmetry and can be assumed as an internal isospin
(spin, valley or sublattice) degree of freedom. Two mentioned
models, Eqs. (1) and (2), are similar to some extent and
describe similar physical properties.
Generally, the Hamiltonian around the (τ=+) and
K(τ=+) points for UTF-TI and monolayer MoS2systems,
respectively, can be rewritten as
H=a1+b(α+β)q2cq
cq a2+b(αβ)q2,(3)
where a1=/2+0,a
2=−/2+0for UTF-TI and
a1=/2,a
2=−/2+λs for ML-MoS2. Note that b=
2/4m0a2
0,c=t0, and we set a0qq. The eigenvalue and
eigenvector of the Hamiltonian, Eq. (3), can be obtained as
|ψc,v= 1
Dc,v cq
hc,v ,
hc,v =dd2+c2q2,d=a1a2
2+ q 2,
(4)
Dc,v =c2q2+h2
c,v,
εc,v =a1+b(α+β)q2hc,v,
and velocity operators along the xand ydirections are
vx=∂H
∂qx=x+2bαqx+2 q xσz,
(5)
vy=∂H
∂qy=y+2bαqy+2 q yσz.
The intrinsic optical conductivity can be calculated by using
the Kubo formula [3032] in a clean sample and it is given by
σxy (ω)=−ie2
2πh d2qf(εc)f(εv)
εcεvψc|vx|ψvψv|vy|ψc
ω+εcεv+i0++ψv|vx|ψcψc|vy|ψv
ω+εvεc+i0+,
(6)
σxx(ω)=−ie2
2πh d2qf(εc)f(εv)
εcεvψc|vx|ψvψv|vx|ψc
ω+εcεv+i0++ψv|vx|ψcψc|vx|ψv
ω+εvεc+i0+,
where f(ω) is the Fermi distribution function. We include only the interband transitions and the contribution of the intraband
transitions, which leads to the fact that the Drude-like term is no longer relevant in this study since the momentum relaxation time
is assumed to be infinite. This approximation is valid at low temperature and a clean sample where defect, impurity, and phonon
scattering mechanisms are ignorable. We also do not consider the bound state of exciton in the systems. After straightforward
calculations (details can be found in Appendix A), the real and imaginary parts of diagonal and off-diagonal components of the
115413-2
INTRINSIC OPTICAL CONDUCTIVITY OF MODIFIED . . . PHYSICAL REVIEW B 89, 115413 (2014)
conductivity tensor at τ=+are given by
σRe
xy (ω)=2e2
hqdq(f(εc)f(εv)) ×c2
d2+c2q2(d2 q 2)P1
(ω)2(εcεv)2,
σIm
xy (ω)=πe2
hqdq f(εc)f(εv)
εcεv×c2
d2+c2q2(d2 q 2){δ(ω+εvεc)δ(ω+εcεv)},
(7)
σIm
xx (ω)=−2e2
h
ωqdq f(εc)f(εv)
εcεv×c2c2q2
d2+c2q2c2
2+ (a1a2)P1
(ω)2(εcεv)2,
σRe
xx (ω)=−πe2
hqdq f(εc)f(εv)
εcεv×c2c2q2
d2+c2q2c2
2+ (a1a2){δ(ω+εvεc)+δ(ω+εcεv)},
where Re and Im refer to the real and imaginary parts of σand Pdenotes the principle value. It is worthwhile mentioning that the
conductivity for ML-MoS2for τ=−can be found by implementing px→−pxand λ→−λ. Using these transformations, the
velocity matrix elements around the Kpoint can be calculated by taking the complex conjugation of the corresponding results
for the τ=+case. Furthermore, for the UTF-TI case system, we must replace and βby their opposite signs which lead to
the same results in comparison with the ML-MoS2case around Kpoint. More details in this regard are given in Appendix A.
A. Optical conductivity of ML-MoS2
Having obtained the general expressions of the conductivity for the modified Dirac fermion systems, the conductivity of two
examples namely the ML-MoS2and UTF-TI could be obtained. Here, we would like to focus on the ML-MoS2case and explore
its optical properties, although all results can be generalized to the UTF-TI system as well. Therefore, the optical conductivity for
each spin and valley components of ML-MoS2can be obtained by using appropriate substitution in Eq. (7) and results are written as
σRe s
xy (ω)=2e2
h
Pdq(f(εc)f(εv)) ×τ(
τsqβq3)
(
τs +βq2)2+q2[4((
τs +βq2)2+q2)(ω/t0)2],
σIm s
xy (ω)=πe2
2hdq(f(εc)f(εv)) ×τ(
τsqβq3)
(
τs +βq2)2+q2δ(ω/t02(
τs +βq2)2+q2),
σIm s
xx (ω)=−2e2
h
ωPdq(f(εc)f(εv)) ×q
(
τs +βq2)2+q2[4((
τs +βq2)2+q2)(ω/t0)2]
q31
2+2β
τs
((
τs +βq2)2+q2)3/2[4((
τs +βq2)2+q2)(ω/t0)2],
σRe s
xx (ω)=−πe2
2hdq(f(εc)f(εv)) ×q
(
τs +βq2)2+q2q31
2+2β
τs
((
τs +βq2)2+q2)3/2
×δ(ω/t02(
τs +βq2)2+q2),(8)
where
τs =(λτ s)/2t0,α=bα/t0,β= /t0,στ,s
xy =
σRe s
xy +Im s
xy , and στ,s
xx =σRe s
xx +Im s
xx .
Note that in the case of UTF-TI, there is no extra spin index
of sas a degree of freedom and
τs might be replaced by
=/2t0; consequently we have στ
xy and στ
xx rather than
στs
xy and στs
xx .Tobemoreprecise,λτ s , which is located out
of the radical in Eq. (8), might be replaced by 0in the c,v
to achieve desirable results corresponding to the UTF-TI. It is
clear that the dynamical charge Hall conductivity vanishes in
both the UTF-TI and ML-MoS2systems due to the presence
of the time reversal symmetry. For the MoS2case, the spin and
valley transverse ac conductivity are given by
σs
xy =
2e
τστ,
xy στ,
xy
v
xy =1
e
sσK,s
xy σK,s
xy ,
(9)
and for the longitudinal ac-conductivity case, an electric
field can only induce a charge current and corresponding
conductivity is given as
σxx =
τστ,
xx +στ,
xx .(10)
Moreover, the longitudinal conductivity is the same as ex-
pression given by Eq. (10) for the UTF-TI case, however, the
Hall conductivity is slightly changed. Owing to the coupling
between the isospin and the spin indexes, the hyperbola Hall
conductivity is a spin Hall conductivity [7,8] and it is thus
given by
σhyp
xy =1
eσ+
xy σ
xy .(11)
115413-3
HABIB ROSTAMI AND REZA ASGARI PHYSICAL REVIEW B 89, 115413 (2014)
B. Intrinsic dc conductivity
To find the static conductivity in a clean sample, we
set ω=0 and thus the interband longitudinal conductivity
vanishes. Consequently, we calculate only the transverse
conductivity in this case. At zero temperature, the Fermi
distribution function is given by a step function, i.e., f(εc,v)=
(εFεc,v). We derive the optical conductivities for the case
of ML-MoS2and results corresponding to the UTF-TI can
be deduced from those after appropriate substitutions. Most
of the interesting transport properties of ML-MoS2originates
from its spin-splitting band structure for the hole doped case.
Therefore, for the later case, when the upper spin-split band
contributes to the Fermi level state, the dc conductivity is
given by
σK
xy =−σK
xy =−e2
2hqc
qF
(
Kqβq3)dq
((
K+βq2)2+q2)3
2
=−e2
2hCK+e2
2h
2μ+2b(αβ)q2
F
λ+2μ+2bαq2
F
,(12)
and for the spin-down component we thus have
σK
xy =−σK
xy =−e2
2hqc
0
(
Kqβq3)dq
((
K+βq2)2+q2)3
2
=−e2
2hCK,(13)
where qcis the ultraviolate cutoff and μ/t0=
(
K+βq2
F)2+q2
F
Kαq2
Fstands for the chemical
potential and it is easy to show that CKs =sgn(
λs)sgn(β) at large cutoff values. In a precise definition,
CKs terms are the Chern numbers for each spin and valley
degrees of freedom and the total Chern number is zero owing
to the time reversal symmetry. Intriguingly, the quadratic
term in Eq. (3), β, leads to a new topological characteristic.
When β > 0, with >λ, the system has a trivial phase
with no edge mode closing the energy gap, however, for
the case that β < 0, the topological phase of the system
is a nontrivial with edge modes closing the energy gap. In
the case of the ML-MoS2, the tight binding model [21,33]
predicts the trivial phase (β>0) with CKs =0. However, a
nontrivial phase is expected by Refs. [22,23] (where β<0)
which leads to CKs =2. In other words, the term proportional
to βhas a topological meaning in Z2symmetry invariant like
the UTF-TI system [8] and the sign of βplays important
role.
The transverse intrinsic dc conductivity for the hole doped
ML-MoS2case, is given by
σs
xy =
eσK
xy σK
xy =e
2π
μ+b(αβ)q2
F
λ+2μ+2bαq2
F
,
σv
xy =2
eσK
xy +σK
xy =−e
hCK+2
σs
xy ,(14)
where, at large cutoff, CK=[sgn(λ)+sgn(+λ)]/2
sgn(β) stands for the valley Chern number and it equals
to zero or 2 corresponding to the nontrivial or trivial band
structure, respectively. In the case of the UTF-TI, the isospin
Hall conductivity is
σhyp
xy =−e
hC+2e
h
μ+b(αβ)q2
F
+2μ+2bαq2
F
,(15)
where μ/t0=(+βq2
F)2+q2
F0αq2
Fand
C=sgn()sgn(β) at large cutoff. This result is consistent
with that result obtained by Lu et al. [8]. It should be noted
that in the absence of the diagonal quadratic term, the nonzero
valley Chern number at zero doping predicts a valley Hall
conductivity, which is proportional to sign(). Therefore, the
exitance of edge states, which can carry the valley current, is
anticipated. However, Z2symmetry prevents the edge modes
from existing. Since the Z2topological invariant is zero when
the gap is caused only by the inversion symmetry breaking
[34], thus the topology of the band structure is trivial and there
are no edge states to carry the valley current when the chemical
potential is located inside the energy gap. Therefore, we can
ignore the valley Chern number in σv
xy and thus the results are
consistent with those results reported by Xiao et al. [20]ata
low doping rate where μλ.
C. Intrinsic dynamical conductivity
In this section, we analytically calculate the dynamical
conductivity of the modified Dirac Hamiltonian which results
in the trivial and nontrivial phases. Using the two-band
Hamiltonian, including the quadratic term in momentum, the
optical Hall conductivity for each spin and valley components
are given by
σRe s
xy (ω)=τe2
h[Gτs(ω,qF)Gτs(ω,qc)],
σIm s
xy (ω)=τπe2
2h
τs βq2
0 s
ωn(ω)2ε
Fλτs
2αq2
0 s ω(ω→−ω)
×(n(ω)(1 +2β
τs)),(16)
where Re and Im indicate to the real and imaginary parts,
respectively, and Gτs(ω,q) reads as below (details are given
in Appendix B):
Gτs(ω,q)
=
τs
ωn(ω)ln
ωm(q)
n(ω)2(
τs +βq2)2+q2
ωm(q)
n(ω)+2(
τs +βq2)2+q2
+1
4βωn(ω)ln
ωm(q)
n(ω)2(
τs +βq2)2+q2
ωm(q)
n(ω)+2(
τs +βq2)2+q2
1
4βωln
ω2(
τs +βq2)2+q2
ω+2(
τs +βq2)2+q2
,(17)
where m(q)=1+2β
τs +2β2q2,n(ω)=
1+4β
τs +β2(ω)2,ω=ω/t0,ε
F=εF/t0,
and λ=λ/t0.Thevalueofq0 s can be evaluated from
115413-4
INTRINSIC OPTICAL CONDUCTIVITY OF MODIFIED . . . PHYSICAL REVIEW B 89, 115413 (2014)
m(q0 s )=n(ω). Note that qc, the ultraviolate cutoff, is
assumed to be equal to 1/a0. Some special attentions might
be taken for the situation in which there is no intersection
between the Fermi energy and the band energy, for instance,
in a low doping hole case of the ML-MoS2in which the Fermi
energy lies in the spin-orbit splitting interval. In this case,
the Fermi wave vector (qF, which has no contribution to the
Fermi level) vanishes.
The quadratic terms can also affect profoundly on the
longitudinal dynamical conductivity which plays main role
in the optical response when the time reversal symmetry is
preserved. In this case, one can find
σRe s
xx (ω)=−πe2
4h
1
n(ω)11+4β
τs
22q0 s
ω2
×2ε
Fλτs 2αq2
0 s ω
(ω→−ω)(n(ω)(1 +2β
τs))
σIm s
xx (ω)=−e2
h[Hτs(ω,qF)Hτs(ω,qc)],(18)
where Hτs(ω,q) is given by (details are given in Appendix B)
Hτs(ω,q)
=(1 +2β
τs)m(q)(1 +4β
τs)
2β2ω(
τs +βq2)2+q2
+1+4β
τs
2β2(ω)2ln
ω
2(
τs +βq2)2+q2
ω
2+(
τs +βq2)2+q2
+(1 +2β
τs)(1 +4β
τs)+β2(ω)2
2β2(ω)2n(ω)
×ln
ω
2
m(q)
n(ω)(
τs +βq2)2+q2
ω
2
m(q)
n(ω)+(
τs +βq2)2+q2
.(19)
It is worthwhile mentioning that the Gand Hfunctions do not
depend on the αterm given in Eq. (3). For β=0inEq.(3),
we have m(q)/n(ω)1, 1/n(ω)12β
τs, therefore
Gτs(ω,q) reduces to gτs(ω,q) and in a similar way, Hτs reduces
to hτs.Heregτs and hτs read as below:
gτs(ω,q)=λτ s
4ωln
ω(λτ s)2+4t2
0q2
ω+(λτ s)2+4t2
0q2
,
hτs(ω,q)=λτ s
2ω
λτ s
(λτ s)2+4t2
0q2
+1
41+λτ s
ω2
×ln
ω(λτ s)2+4t2
0q2
ω+(λτ s)2+4t2
0q2
.(20)
Using Eqs. (8) and (20), the conductivity simplifies when β=
0 and the results are
σRe s
xy (ω)=τe2
h[gτs(ω,qF)gτs(ω,qc)],
σIm s
xy (ω)=τπe2
4h
λτ s
ω[(2εFλτ s ω)
(ω→−ω)](ω(λτ s)).(21)
The longitudinal conductivity for the case of β=0isgiven
by the following relations for the electron doped case:
σRe s
xx (ω)=−πe2
8h1+λτ s
ω2(ω(λτ s))
×[(2εFλτ s ω)(ω→−ω)]
σIm s
xx (ω)=−e2
h[hτs(ω,qF)hτs(ω,qc)].(22)
These relations are consistent with those results reported in
Ref. [25]. Furthermore, dropping the λterm gives rise to the
optical conductivity of gapped graphene and the result is in
good agreement with the universal conductivity of graphene
[35]for=λ=α=β=0.
III. NUMERICAL RESULTS
In most numerical results, we use set0:λ=
0.08 eV, =1.9eV,t0=1.68 eV =m0/m+=0.43 =
m0/m4m0v2/(λ)=2.21 where m±=memh/
(mh±me) and v=t0a0/. These values have been obtained
in Ref. [21]. Moreover, for the sake of completeness,
we introduce two other sets of the parameters as
t0=1.51 eV =1.77, and another set t0=2.02 eV; β=0
corresponding to the same effective masses (α=0for
me=−mh=0.5m0) for electron and hole bands. These
parameters are calculated by using the procedure reported
in Ref. [21]. The later comparison helps us to perceive the
validity of the effective mass approximation for the ML-MoS2
system and for this purpose, we assume the same effective
masses for electron and hole bands to compare the spin Hall
conductivity resulted from the Dirac-like and modified Dirac
Hamiltonians. Notice that all energies are measured from the
center of the energy gap.
The real part of the optical Hall and longitudinal
conductivities for the two sets of parameters, with and without
quadratic terms, are illustrated in Figs. 1and 2where top
and bottom panels indicate electron and hole doped systems,
respectively. The effect of the mass asymmetry between
the effective masses of the electron and hole (α) bands
is neglected and it will be discussed later. It is clear that
the quadratic term βcauses a reduction of the intensity
of the optical Hall conductivity with no changing of the
position of peaks for both electron and hole doped cases. The
position of peaks in the real part of Hall conductivity is given
by ω=(λτ s)2+4t2
0qFs2for the β=0 case and
ωm(qFs)n(ω)12(
τs +βqFs2)2+qFs 2=0 and
ω2(
τs +βqFs2)2+qFs 2=0 for each spin
component with corresponding Fermi wave vector qFs
and for the case that β= 0. Surprisingly, the last two
equations for the latter case simultaneously fulfilled the
115413-5
HABIB ROSTAMI AND REZA ASGARI PHYSICAL REVIEW B 89, 115413 (2014)
0 1 2 3 4
¯(eV)
0.05
0.00
0.05
0.10
0.15
0.20
σxy(e2/¯h)
(a)
s=1 =0
s=1 =1.77
s=1 =0
s=1 =1.77
0 1 2 3 4
¯
(
eV
)
0.05
0.00
0.05
0.10
0.15
0.20
σxy(e2/¯h)
(b)
s=1 =0
s=1 =1.77
s=1 =0
s=1 =1.77
FIG. 1. (Color online) Real part of the Hall conductivity (in units
of e2/)for(a)electronwithεF=1 eV and (b) hole with εF=
1eV+λdoped cases as a function of photon energy (in units
of eV) around the Kpoint. Electron and hole masses are set to
be 0.5m0and for two sets of parameters, β=0,t0=2.02 eV and
β=1.77,t0=1.51.
equation m(qFs)=n(ω) in frequency. In the energy range
shown in the figures, the numerical value of the peak position
for both cases are approximately equal and it indicates that
the position of peaks and steplike shape do not change due
to the βterm in a certain Fermi energy. It should be noticed
that the intensity of the real part of σxx decreases with the
quadratic term. Consequently, it indicates that the effective
mass approximation of the Hamiltonian for the ML-MoS2is
not completely valid because two sets of parameters with the
same effective masses are showing distinct results.
A. Mass asymmetry between electron and hole
In this subsection, we consider the mass asymmetry
between electron and hole bands and then the conductivity
of the ML-MoS2is calculated for the Hamiltonian given in
Eq. (3). The results are illustrated in Figs. 3and 4around the
Kpoint. Due to the mass asymmetry, a small splitting between
0 1 2 3 4
¯(eV)
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
σ
xx
(e
2
/¯h)
(a)
s=1 =0
s=1 =1.77
s=1 =0
s=1 =1.77
0 1 2 3 4
¯(eV)
0.00
0.02
0.04
0.06
0.08
0.10
0.12
0.14
σ
xx
(e
2
/¯h)
(b)
s=1 =0
s=1 =1.77
s=1 =0
s=1 =1.77
FIG. 2. (Color online) Real part of the longitudinal conductivity
(in units of e2/) for (a) electron with εF=1 eV and (b) hole with
εF=−1eV+λdoped cases as a function of photon energy (in units
of eV) around the Kpoint. Electron and hole masses are set to
be 0.5m0and for two sets of parameters, β=0,t0=2.02 eV and
β=1.77,t0=1.51.
electron and hole doped cases takes place in the spin-up
component. On the other hand, there is considerable splitting
between electron and hole doped cases due to both spin-orbit
coupling and mass asymmetry for the spin-down case. We also
note a sharp onset in the imaginary part of the conductivity,
minimum energy associated with the possible interband optical
transition. Moreover, corresponding to the onset in σIm
xy (σRe
xx )
where there is a peak in its real (imaginary) part at the same
energy as they are related by the Kramers-Kroning relations.
The position of peaks or steplike configuration of the
dynamical conductivity, can be controlled by the doping rate.
Figure 5shows the difference between the position of those
peaks, δω =ωω, around the Kpoint for electron and
hole doped cases corresponding to the real part of the Hall
conductivity for each spin component. As it is clearly shown
in this figure, δω increases linearly from a negative value to a
positive one up to a saturation value (2λ) for the hole doped
115413-6
INTRINSIC OPTICAL CONDUCTIVITY OF MODIFIED . . . PHYSICAL REVIEW B 89, 115413 (2014)
1.8 1.9 2.0 2.1 2.2
¯(eV)
0.00
0.05
0.10
0.15
0.20
0.25
σxy(e2/¯h)
(a)
s= +1, electron doped
s=1, electron doped
s= +1, hole doped
s=1, hole doped
1.8 2.0 2.2 2.4 2.6 2.8 3.0
¯(eV)
0.00
0.02
0.04
0.06
0.08
0.10
σxy(e2/¯h)
(b)
s= +1, electron doped
s=1, electron doped
s= +1, hole doped
s=1, hole doped
FIG. 3. (Color online) (a) Real and (b) imaginary parts of the
optical Hall conductivity (in units of e2/) as a function of photon
energy (in units of eV) around the Kpoint. Red (blue) color stands
for the electron (hole) doped case with εF=1eV(εF=−1eV+λ)
and solid (dashed) line indicates the spin-up (-down).
case. The linear part of the result originates from the spin
splitting in the valence band and the fact that there are two
Fermi wave vectors in which one component spin has zero
Fermi wave vector and does not change by increasing the
doping rate. Finally, by increasing the Fermi energy, two Fermi
wave vectors contribute to the calculations and the position of
both peaks move in the same way and lead to a saturation value
for δω.
B. Circular dichroism and optical transmittance
One of the main optical properties of the monolayer tran-
sition metal dichalcogenide system is the circular dichroism
when it is exposed by a circularly polarized light in which left-
or right-handed light can be absorbed only by the Kor K
valley and it makes the material promising for the valleytronic
field. This effect originates from the broken inversion sym-
metry and it can be understood by calculating the interband
optical selection rule P±=m0ψc|vx±ivy|ψvfor incident
1.8 2.0 2.2 2.4 2.6 2.8 3.0
¯(eV)
0.00
0.02
0.04
0.06
0.08
0.10
σxx(e2/¯h)
(a)
s= +1, electron doped
s=1, electron doped
s= +1, hole doped
s=1, hole doped
1.8 1.9 2.0 2.1 2.2
¯(eV)
0.0
0.1
0.2
0.3
0.4
0.5
σxx(e2/¯h)
(b)
s= +1, electron doped
s=1, electron doped
s= +1, hole doped
s=1, hole doped
FIG. 4. (Color online) (a) Real and (b) imaginary parts of the
optical longitudinal conductivity (in units of e2/) as a function
of photon energy (in units of eV) around the Kpoint. Red(blue)
color stands for the electron (hole) doped case with εF=1eV
(εF=−1eV +λ) and the solid (dashed) line indicates the spin-up
(-down).
right-(+) and left-()handed light. The photoluminescence
probability for the modified Dirac fermion Hamiltonian is
|P±|=m0t0a0
1±τd2 q 2
d2+c2q2,(23)
where q2=q2
x+q2
y. Notice that the mass asymmetry term α
has no effect on the optical selection rule. The selection rule
can simply prove the circular dichroism in the ML-MoS2.
Another approach which helps us to understand this effect is
to calculate the optical conductivity around the Kpoint of two
kinds of light polarizations as σ±=s{σKs
xx ±σKs
xy }, which
has been calculated by using the Dirac-like model [19,25], and
now we modify that by using the modified Dirac Hamiltonian.
Figure 6shows the coupling of the light and valleys. Note
that Re[σ] is large and comparable in size for either spin-up
or -down while Re[σ+] is small in comparison. The valley
around the Kpoint can couple only to the left-handed light
115413-7
HABIB ROSTAMI AND REZA ASGARI PHYSICAL REVIEW B 89, 115413 (2014)
0.0 0.1 0.2 0.3 0.4 0.5
|μμ0|(eV)
0.2
0.1
0.0
0.1
0.2
¯ω(eV)
hole doped
electron doped
FIG. 5. (Color online) Difference between the position of the
peak in the real part of the Hall conductivity, δω =ωω,for
both spin components for the electron doped case including mass
asymmetry as a function of the chemical potential. Note that μ0,
which is the band edge in the conduction and valence bands, is 0.95 eV
and 0.87 eV for the electron and hole doped case, respectively.
and this effect is washed up by increasing the frequency
of the light and the result is in good agreement with recent
experimental measurements [12].
Furthermore, the optical transmittance is an important
physical quantity and it can be evaluated stemming from the
conductivity. The optical transmittance of a free standing thin
film exposed by a linear polarized light is given by [36]
T(ω)=1
2
2
2+Z0σ+(ω)
2
+
2
2+Z0σ(ω)
2,(24)
1.0 1.5 2.0 2.5 3.0 3.5 4.0
¯
(
eV
)
0.00
0.05
0.10
0.15
0.20
0.25
Re[σ]
Re[σ]
Re[σ+]
FIG. 6. (Color online) Real part of the optical conductivity
around Kpoint, for left-(solid) hand and right-(dashed) handed
light. It indicates the appearance of the circular dichroism effect for
the modified Dirac equation. The electron (εF=1 eV) doped case
including mass asymmetry.
1.50 1.75 2.00 2.25 2.50
¯(eV)
0.90
0.92
0.94
0.96
0.98
1.00
Transmittance
hole doped
electron doped
FIG. 7. (Color online) Optical transmittance in a finite frequency
for the electron (εF=1 eV) and hole (εF=−1eV+λ) doped cases
including mass asymmetry.
where Z0=376.73and σ±(ω)=σxx(ω)±xy are the
vacuum impedance and the optical conductivity of the thin
film, respectively. For the ML-MoS2case, the total Hall
conductivity in the presence of the time reversal symmetry is
zero and the total longitudinal conductivity is given by σxx =
2(σK
xx +σK
xx ). The optical transmittance of the multilayer
of MoS2systems has been recently measured [33] and it is
about 94.5% for each layer in the optical frequency range. The
optical transmittance of the ML-MoS2is displayed in Fig. 7for
both electron and hole doped cases using the numerical value
defined as set0. The result shows that the optical transmittance
is about 98% for the frequency range in which both spin
components are active for giving response to the incident
light. Importantly, for the electron dope case, there are two
minimums with distance about 0.16 eV/in frequency which
mostly indicates the spin-orbit splitting (2λ) in the valence
band and it is consistent with the results illustrated in Fig. 5.
The optical transmittance for the electron doped case is about
98% in all frequency ranges. Moreover, for the hole dope case
as it is shown in Fig. 5, the optical transmittance changes
by the tuning doping rate. Interestingly, at μ=−0.942 eV
the difference between the position of peaks of two spin
components, δω, is approximately zero. Consequently, the
total optical conductivity enhances in this resonating doping
rate which has significant effect on the optical transmittance of
the system where the transmittance decreases and particularly
reaches to a value less than 90% at the resonance frequency
when δω 0. Our numerical calculations show that the
hole doped ML-MoS2is darker than the electron doped one
especially close to the resonance frequency. Furthermore, this
feature provides an opportunity for measuring the spin-orbit
coupling by an optical transmittance measurement.
C. Optical response in the nontrivial phase
The modified Dirac Hamiltonian shows a nontrivial phase
when β < 0 and it has been numerically shown that in this
phase a light matter interaction enhances due to the change
115413-8
INTRINSIC OPTICAL CONDUCTIVITY OF MODIFIED . . . PHYSICAL REVIEW B 89, 115413 (2014)
TABLE I. Numerical parameter for the ultrathin film of a
topological insulator [8].
L(˚
A) (eV) t0(eV) αβ
20 0.14 2.22 1.05 23.67
25 0.0 2.21 2.37 18.41
32 0.04 2.20 3.94 6.31
of the parabolic band dispersion into the shape of a Mexican
hat with two extrema [28]. To fulfill such a band dispersion, a
negative value β with a large absolute value is required and it
is accessible for an ultrathin film of the topological insulator.
The sign and the absolute values of the parameters can be
manipulated by the thickness of the thin film, while in the case
of the ML-MoS2, to the best of our knowledge, it is barely
possible to create a Mexican-hat-like dispersion relation even
for the model Hamiltonian with a nontrivial topology phase
[22,23]. In this case, we plot the optical Hall and longitudinal
conductivities of the UTF-TI in its trivial and nontrivial phases.
In the UTF-TI [37] system, in which only in-plan components
of momentum are relevant, one can find the Hamiltonian given
by Eq. (2) where the numerical value of the model parameters
depends on the thickness of the thin films [8,26]. We consider
three different thicknesses for which three sets of parameters
[8] are listed in Table I. We also neglect the value of 0which
is just a constant shift in the energy.
As can be seen from Table I, a sample with L=20 ˚
Aor
L=32 ˚
A indicates the trivial or nontrivial phases, respec-
tively. However, for a sample with L=25 ˚
A the energy gap
vanishes and thus at critical thickness, L=25 ˚
A, the trivial to
nontrivial phase transition takes place. Hereafter, we call that
a phase boundary.
Now, we calculate the real part of the Hall and longitudinal
conductivities for τ=+ and the results are illustrated in
Fig. 8. It shows that the conductivity enhances in the nontrivial
phase which is consistent with previous numerical work
[28]. More interestingly, we are now showing that the Hall
conductivity changes sign through changing the thickness and
it is very important in the circular dichroism effect. This
changing of the sign means a different helicity of the light
can be coupled to the system. It is worth mentioning that the
circular dichroism effect on the electronic system governing
modified Dirac Hamiltonian is also possible when energy
gap is zero [19,25]. The selection rule equation reads as
|P±|=m0t0a0
(1 τbβq/
( q )2+c2) for the case of zero
gap. This expression indicates that the circular polarization is
achievable away from the point even in the absence of the
energy gap. It might be emphasized that the peak in the optical
conductivity at zero energy gap originates from a nonzero
Fermi energy in which the low energy part of phase space
is no longer available for a photon absorbtion process [38]
based on the Pauli exclusion principle. More precisely, there
is a peak at energy point ω2εFin the topological insulator
case and it can be seen from Eqs. (16) and (18). Therefore, the
peak disappears at zero Fermi energy for a gapless system. In
Fig. 9, we show the optical conductivity for the two helicities of
light for τ=+. The results show that the circular polarization
0.0 0.1 0.2 0.3 0.4 0.5
¯(eV)
0.05
0.00
0.05
0.10
0.15
0.20
σxy(e2/¯h)
(a)
×(1)
L=20˚
A
L=25˚
A
L=32˚
A
0.0 0.1 0.2 0.3 0.4 0.5
¯
(
eV
)
0.00
0.02
0.04
0.06
0.08
0.10
σxx(e2/¯h)
(b) L=20˚
A
L=25˚
A
L=32˚
A
FIG. 8. (Color online) Real part of the Hall (a) and longitudinal
(b) conductivity for τ=1 and different values of film thickness.
It is clear that in the nontrivial phase the optical response of the
system is stronger than that of its trivial one. The Fermi energy is
εF=||/2+0.03 eV.
changes sign for the negative value of the gap and it gets more
strength in the nontrivial phase rather than the trivial phase.
IV. SUMMARY
We have analytically calculated the intrinsic conductivity of
the electronic systems which govern a modified Dirac Hamil-
tonian by using the Kubo formula. We have studied the effect
of the quadratic term in momentum β, which has been recently
predicted, and found the different optical responses. This
discrepancy originates from the different topological structures
of the systems. Our calculations show that the βterm has no
effect on the position of the peak of the optical conductivity
but it has considerable effect on its magnitude. Therefore, it
shows that the same effective mass approximation for electron
and hole bands for monolayer MoS2cannot fully describe
the optical properties. The effect of the strong spin-orbit
interaction can be traced by the difference of the energy interval
between the position of the peak in the optical conductivity for
115413-9
HABIB ROSTAMI AND REZA ASGARI PHYSICAL REVIEW B 89, 115413 (2014)
0.0 0.1 0.2 0.3 0.4
¯(eV)
0.00
0.05
0.10
0.15
0.20
Re[σ]
(a)
L=20˚
A, Re[σ]
L=20˚
A, Re[σ+]
0.0 0.1 0.2 0.3 0.4
¯
(
eV
)
0.00
0.05
0.10
0.15
0.20
0.25
Re[σ]
(b) L=25˚
A, Re[σ]
L=25˚
A, Re[σ+]
L=32˚
A, Re[σ]
L=32˚
A, Re[σ+]
FIG. 9. (Color online) Circular dichroism effect for different
values of the thickness. The real part of the optical conductivity
around the Kpoint is shown for (a) L=20 ˚
Aand(b)L=25 ˚
Aand
31 ˚
A. The Fermi energy is εF=||/2+0.03 eV.
the two spin components in electron and hole doped cases.
We have shown that this interval for the electron doped case
is approximately constant while for the hole doped case, it
increases from a negative value to a positive one, and then it
increases linearly up to a saturation value. The effect of the
mass asymmetry in monolayer MoS2induces a small splitting
between the conductivity spectrum for the electron and hole
doped cases. The circular dichroism effect is investigated for
the modified Dirac Hamiltonian of the monolayer MoS2by
calculating the selection rule and the optical conductivity. We
have also obtained the optical transmittance of the monolayer
MoS2for the hole and electron doped cases and the results
show that the valence band spin splitting has considerable
effect on the intensity of the transmittance.
We have also studied the effect of the quantum phase
transition, which occurs owing to the reducing of the thickness,
on the optical conductivity of the thin film of the topological
insulator. We have shown that at the phase boundary, when
the energy gap is zero, the diagonal quadratic term plays
a significant role on the optical conductivity and selection
rule. Moreover, we have illustrated that the optical response
enhances and the optical Hall conductivity changes sign in the
nontrivial phase (QSH) and the phase boundary.
ACKNOWLEDGMENTS
R.A. would like to thank the Institute for Material Research
in Tohoku University for its hospitality during the period when
the last part of this work was carried out.
APPENDIX A
In this Appendix, the details of the calculations deriving
Eq. (8) are presented. Since ψc|ψv=0, we get
ψc|vx|ψv=cψc|σx|ψv+2 qxψc|σz|ψv,
(A1)
ψv|vy|ψc=cψv|σy|ψc+2 qyψv|σz|ψc,
owing to the fact that the mass asymmetry parameter αplays
no role in the velocity matrix elements. Using hchv=−c2q2
we have
ψc|σx|ψv= c
DcDv
[qhv+qhc],
ψv|σy|ψc= ic
DcDv
[qhcqhv],(A2)
ψc|σz|ψv=ψv|σz|ψc= 2c2q2
DcDv
.
In this case
ψc|vx|ψv= c2
DcDv{−[qhv+qhc]+4bβq xq2},
(A3)
ψv|vy|ψc= c2
DcDv{i[qhcqhv]+4bβq yq2}.
Consequently, the product of the velocity matrix elements are
ψc|vx|ψvψv|vy|ψc
=c4
(DcDv)2{−i(qhv+qhc)(qhcqhv)+(4 q 2)2qxqy
+4 q 2(qy(qhv+qhc)+iqx(qhcqhv))},
ψc|vx|ψvψv|vx|ψc
=c4
(DcDv)2{|qhv+qhc|2+(4bqxβq2)2
4bqxβq2(qhv+qhc+qhv+qhc)}.(A4)
Using tan φ=qy/qx, one can find
(qhv+qhc)(qhcqhv)
=−2ic2q4sin 2φ4q2dd2+c2q2
qy(qhv+qhc)+iqx(qhcqhv)
=2q2[id2+c2q2+dsin 2φ],
qhv+qhc+qhv+qhc=4qd cos φ,
|qhv+qhc|2=4q2(d2+c2q2sin φ2),
(DcDv)2=4c2q2[d2+c2q2].(A5)
115413-10
INTRINSIC OPTICAL CONDUCTIVITY OF MODIFIED . . . PHYSICAL REVIEW B 89, 115413 (2014)
After substituting Eq. (A5) into Eq. (A4), we get
ψc|vx|ψvψv|vy|ψc=c2q2sin 2φ
d2+c2q2c2
2+2 ( q 2+d)+ic2
d2+c2q2{d2 q 2},
(A6)
ψc|vx|ψvψv|vx|ψc=c2c2q2cos φ2
d2+c2q2{c2+2 (a1a2)}.
Using sin 2φ=0,cos φ2=π, one can find
σxy =e2
hqdq f(εc)f(εv)
εcεv×c2
d2+c2q2(d2 q 2)1
ω+εcεv+i0+1
ω+εvεc+i0+,
σxx =−ie2
hqdq f(εc)f(εv)
εcεv×c2c2q2
d2+c2q2c2
2+ (a1a2) 1
ω+εcεv+i0++1
ω+εvεc+i0+.
(A7)
Using (x+i0+)1=Px1iπδ(x) where Pstands for principal value, it is easy to show that the real and imaginary parts of
diagonal and off-diagonal components of the conductivity tensor read as below:
σRe
xy =2e2
hqdq(f(εc)f(εv)) ×c2
d2+c2q2(d2 q 2)P1
(ω)2(εcεv)2,
σIm
xy =πe2
hqdq f(εc)f(εv)
εcεv×c2
d2+c2q2(d2 q 2){δ(ω+εvεc)δ(ω+εcεv)},
(A8)
σIm
xx =−2e2
h
ωqdq f(εc)f(εv)
εcεv×c2c2q2
d2+c2q2c2
2+ (a1a2)P1
(ω)2(εcεv)2,
σRe
xx =−πe2
hqdq f(εc)f(εv)
εcεv×c2c2q2
d2+c2q2c2
2+ (a1a2){δ(ω+εvεc)+δ(ω+εcεv)}.
To find the conductivity around the Kpoint we must
implement the following changes: px→−pxand λ→−λ.
Using these transformations, the velocity matrix elements
around the Kpoint can be calculated by taking complex
conjugation of the corresponding results around the Kpoint.
Moreover, according to the following dimensionless parame-
ters, εcεv=2d2+c2q2, and thus δ(ω+εcεv)0
for positive frequency in the absorbtion process. Thus Eq. (8)
for the dynamical transverse and longitudinal conductivity is
obtained.
APPENDIX B
In this Appendix, the details of calculations for some
integrals which appear in our model are presented. Using
new variables y=βq2+
τs +(2β)1and a2=
τs+
(4β2)1, it is easy to show that (
τs +βq2)2+q2=y2a2
and we have
Gτs(ω,q)=1
β2
τs +1
2βI1I2,
Hτs(ω,q)=
ω
βI12
τs +1
2βI3
+2
τs +1
2β
τs +1
2βI4,(B1)
where I1,I2,I3, and I4are given by
I1=Pdy
y2a2[4(y2a2)(ω)2],
I2=Pydy
y2a2[4(y2a2)(ω)2],
(B2)
I3=Pydy
(y2a2)3
2[4(y2a2)(ω)2],
I4=Pdy
(y2a2)3
2[4(y2a2)(ω)2].
I1and I4can be calculated by defining uas a new variable
where y=a
1u2and it leads to
I1=1
2ω4a2+(ω)2ln
uω
4a2+(ω)2
u+ω
4a2+(ω)2
,
I4=1
a2I1+1
(ω)2u+4a2+(ω)2
(ω)3
×ln
uω
4a2+(ω)2
u+ω
4a2+(ω)2
.(B3)
115413-11
HABIB ROSTAMI AND REZA ASGARI PHYSICAL REVIEW B 89, 115413 (2014)
By defining y2=u2+a2,I2and I3are obtained as
I2=1
4ωln
uω
2
u+ω
2
,I
3=1
(ω)2u+1
(ω)3ln
uω
2
u+ω
2
.(B4)
Using the above expressions for I1,I2,I3, and I4,itiseasytoproveEqs.(17) and (19).
[1] M. Xu, T. Liang, M. Shi, and H. Chen, Chem. Rev. 113,3766
(2013).
[2] A. K. Geim and I. V. Grigorrieva, Nature (London) 499,419
(2013).
[3] Q. H. Wang, K. Kalantar-Zadeh, A. Kis, J. N. Coleman, and
M. S. Strano, Nat. Nanotechnol. 7,699 (2012).
[4] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov,
and A. K. Geim, Rev. Mod. Phys. 81,109 (2009).
[5] M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82,3045 (2010).
[6] Shun-Qing Shen, Topological Insulator: Dirac Equation in
Condensed Matters (Springer, New York, 2012).
[7] T. Zhang, J. Ha, N. Levy, Y. Kuk, and J. Stroscio, Phys. Rev.
Lett. 111,056803 (2013).
[8] H.-Z.Lu,W.-Y.Shan,W.Yao,Q.Niu,andS.-Q.Shen,Phys.
Rev. B 81,115407 (2010).
[9] H. Li, L. Sheng, D. N. Sheng, and D. Y. Xing, Phys.Rev.B82,
165104 (2010).
[10] H. Li, L. Sheng, and D. Y. Xing, Phys.Rev.B85,045118 (2012).
[11] K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, Phys. Rev.
Lett. 105,136805 (2010).
[12] K. F. Mak, K. He, J. Shan, and T. F. Heinz, Nat. Nanotechnol.
7,494 (2012).
[13] K. F. Mak, K. He, C. Lee, G. H. Lee, J. Hone, T. F. Heinz, and
J. Shan, Nat. Mater. 12,207 (2013).
[14] H. Zeng, J. Dai, W. Yao, D. Xiao, and X. Cui, Nat. Nanotechnol.
7,490 (2012).
[15] T. Cao, G. Wang, W. Han, H. Ye, C. Zhu, J. Shi, Q. Niu, P. Tan,
E. Wang, B. Liu, and J. Feng, Nat. Commun. 3,887 (2012).
[16] S. Wu, J. S. Ross, G. B. Liu, G. Aivazian, A. Jones, Z. Fei,
W. Zhu, D. Xiao, W. Yao, D. Cobden, and X. Xu, Nat. Phys. 9,
149 (2013).
[17] A. Rycerz, J. Tworzydlo, and C. W. J. Beenakker, Nat. Phys. 3,
172 (2007).
[18] D. Xiao, W. Yao, and Q. Niu, Phys. Rev. Lett. 99,236809
(2007).
[19] W. Yao, D. Xiao, and Q. Niu, Phys.Rev.B77,235406
(2008).
[20] Di Xiao, Gui-Bin Liu, W. Feng, X. Xu, and W. Yao, Phys. Rev.
Lett. 108,196802 (2012).
[21] H. Rostami, A. G. Moghaddam, and R. Asgari, Phys. Rev. B 88,
085440 (2013).
[22] G.-B. Liu, W.-Y. Shan, Y. Yao, W. Yao, and D. Xiao, Phys. Rev.
B88,085433 (2013).
[23] A. Kormanyos, V. Zolyomi, N. D. Drummond, P. Rakyta,
G. Burkard, and V. I. Fal’ko, Phys. Rev. B 88,045416 (2013).
[24] A. Carvalho, R. M. Ribeiro, and A. H. Castro Neto, Phys. Rev.
B88,115205 (2013).
[25] Zhou Li and J. P. Carbotte, Phys. Rev. B 86,205425 (2012).
[26] W.-Y. Shan, H.-Z. Lu, and S.-Q. Shen, New J. Phys. 12,043048
(2010).
[27] M. Kim, C. H. Kim, H.-S. Kim, and J. Ihm, Proc. Natl. Acad.
Sci. USA 109,671 (2012).
[28] N. M. R. Peres and J. E. Santos, J. Phys.: Condens. Matter 25,
305801 (2013).
[29] Hai-Zhou Lu, An Zhao, and Shun-Qing Shen, Phys. Rev. Lett.
111,146802 (2013).
[30] T. Stauber, N. M. R. Peres, and A. K. Geim, Phys. Rev. B 78,
085432 (2008).
[31] Wang-Kong Tse and A. H. MacDonald, Phys. Rev. B 84,205327
(2011).
[32] Steven G. Louie and Marvin L. Cohen, Conceptual Foundations
of Materials: A Standard Model for Ground- and Excited-State
Properties (Elsevier, Amsterdam, 2006).
[33] A. C.- Gomez, R. Rold´
an, E. Cappelluti, M. Buscema, F. Guinea,
H. S. J. van der Zant, and G. A. Steele, Nano Lett. 13,5361
(2013).
[34] Wang Yao, Shengyuan A. Yang, and Qian Niu, Phys. Rev. Lett.
102,096801 (2009).
[35] K. Ziegler, Phys.Rev.B75,233407 (2007).
[36] A. Ferreira, J. Viana-Gomes, Y. V. Bludov, V. Pereira, N. M. R.
Peres, and A. H. Castro Neto, Phys.Rev.B84,235410 (2011).
[37] H. J. Zhang, C. X. Liu, X. L. Qi, X. Dai, Z. Fang, and S. C.
Zhang, Nat. Phys. 5,438 (2009).
[38] Kh. Jahanbani and R. Asgari, Eur.Phys.J.B73,247 (2010).
115413-12
... A particularly important physical property of any material is the optical conductivity whose real part relates to absorption of photons and provides a powerful approach for extracting band structures and optical properties. The optical conductivity of Dirac materials has received intensive research both theoretically and experimentally, for graphene [47][48][49], α-(BEDT-TTF) 2 I 3 [29], silicene [30], 8-Pmmn borophene [31], 1H-MoS 2 [50][51][52], and topological insulators [53][54][55]. ...
... Without applying the vertical electric field (α = 0), the energy bands of 1T -MoS 2 are fully gapped and spindegenerate such that the total optical conductivity Reσ j j (ω) = g v g s Reσ +s j j (ω) with g v and g s the degeneracy factors of valley and spin. As shown in Fig. 2, Reσ xx (ω) and Reσ yy (ω) exhibit a strong anisotropy because the dispersion is tilted along the k y direction but untilted along the k x direction, similar to that in 8-Pmmn borophene [31] but unlike that in 1H-MoS 2 [50][51][52]. ...
... When the vertical electric field is applied (α > 0), the valley-spin-polarized bands and gaps of 1T -MoS 2 result in many interesting changes in the longitudinal optical conductivity. As shown in Figs. 3, 4, and 5, Reσ +s xx (ω) and Reσ +s yy (ω) are generally anisotropic and valley-spin-polarized, which is strongly different from that in 1H-MoS 2 [50][51][52]. When we set v − = v 2 = 0 and v + = v 1 = v F , our results restore that in silicene [30]. ...
Article
Full-text available
1T′−MoS2 exhibits valley-spin-polarized tilted Dirac bands in the presence of external vertical electric field and undergoes a topological phase transition between the topological insulator and band insulator around the critical value of the electric field. Within the linear response theory, we theoretically investigate the anisotropic longitudinal optical conductivities of tilted Dirac bands in both undoped and doped 1T′−MoS2, including the effects of the vertical electric field. The influence of the spin-orbit coupling gap, band tilting, and vertical electric field on the optical conductivities of tilted Dirac bands is revealed. A theoretical scheme for probing the topological phase transition in 1T′−MoS2 via exotic behaviors of longitudinal optical conductivities is proposed. The results for 1T′−MoS2 are expected to be qualitatively valid for other monolayer tilted gapped Dirac materials, such as α−SnS2, TaCoTe2, and TaIrTe4, due to the similarity in their band structures.
... Figures 8(a) and 8(b) present the real and imaginary parts of σ xx and σ xy of monolayer MoS 2 , for spin-up and spin-down electrons in the K valleys, respectively. As shown, two jumps can be seen in the conductivity spectra at the optical excitation frequencies of the spin-up and spin-down electrons of the K valley[49]. Similarly, the optical conductivities σ xx and σ xy of monolayer MoS 2 in the K valley in Figs. ...
Article
Full-text available
We investigate the photonic spin Hall effect (PSHE) of light beams reflected from surfaces of various two-dimensional crystalline structures while considering their associated time-reversal 𝒯 and inversion ℐ symmetries. Using the modified Haldane model Hamiltonian with tunable parameters as a generic system, we explore longitudinal and transverse spin separations of the reflected beam in both topologically nontrivial and trivial systems. The PSHE observed in these materials is attributed to their topology. Topological phase transitions in buckled Xene monolayer materials are demonstrated through the PSHE, showing the manipulation of spin-orbit coupling and external electric fields. Moreover, we investigate spatial shifts in the PSHE of monolayer transition metal dichalcogenides, suggesting that the spin and valley degrees of freedom of charge carriers provide a promising avenue to manipulate the PSHE in both classes of these materials. The study suggests that the PSHE in Haldane model materials can serve as a metrological tool for characterizing topological phase transitions through quantum weak value measurement techniques.
... The optical conductivity provides a powerful method for extracting the information of energy band structure, and has been extensively investigated theoretically and experimentally in the 2D untilted Dirac bands [83][84][85][86][87][88][89][90][91][92][93][94][95][96] and undertilted Dirac bands [22,[26][27][28][29][30][31]. Specifically, the exotic behaviors of longitudinal optical conductivity (LOC) can be used to characterize the topological phase transitions in both silicene [90] and 1T -MoS 2 [30]. ...
Article
Full-text available
Lifshitz transition is a kind of topological phase transition in which the Fermi surface is reconstructed. It can occur in the two-dimensional (2D) tilted Dirac materials when the energy bands change between the type-I phase (0<t<1) and the type-II phase (t>1) through the type-III phase (t=1), where different tilts are parametrized by the values of t. In order to characterize the Lifshitz transition therein, we theoretically investigate the longitudinal optical conductivities (LOCs) in type-I, type-II, and type-III Dirac materials within linear response theory. In the undoped case, the LOCs are constants either independent of the tilt parameter in both type-I and type-III phases or determined by the tilt parameter in the type-II phase. In the doped case, the LOCs are anisotropic and possess two critical frequencies determined by ω=ω1(t) and ω=ω2(t), which are also confirmed by the joint density of state. The tilt parameter and chemical potential can be extracted from optical experiments by measuring the positions of these two critical boundaries and their separation Δω(t)=ω2(t)−ω1(t). With increasing the tilting, the separation becomes larger in the type-I phase whereas smaller in the type-II phase. The LOCs in the regime of large photon energy are exactly the same as that in the undoped case. The type of 2D tilted Dirac bands can be determined by the asymptotic background values, critical boundaries, and their separation in the LOCs. These can therefore be taken as signatures of Lifshitz transition therein. The results of this work are expected to be qualitatively valid for a large number of 2D tilted Dirac materials, such as 8-Pmmn borophene monolayer, α−SnS2, TaCoTe2, TaIrTe4, and 1T′ transition metal dichalcogenides, due to the underlying intrinsic similarities of 2D tilted Dirac bands.
... Similarly in a monolayer of the molybdenum disulfide (ML-MoS 2 ) this term induces a valley degeneracy breaking term [19]. For ML-MoS 2 optical conductivity peaks magnitude is considerably effected by the modified Dirac Hamiltonian [20]. ...
Article
Full-text available
In this work we investigate the influence of quadratic in momentum term (Schrodinger term) on magneto-transport properties of thin film topological insulators. The Schrodinger term modifies the Dirac cones into an hourglass shape which results in inter and intraband Landau levels crossings. Breaking of the particle-hole symmetry in Landau level spectrum in the presence of k2 term leads to asymmetrical density of states profile. We calculate collisional and Hall conductivity for mixed Dirac-Schrodinger system in linear response regime and show oscillatory behavior in collisional con- ductivity, while Zeeman and hybridization terms provide a doubly split peak structure in collisional conductivity for the case m/me → ∞. We calculate Hall conductivity analytically and show that for mixed system filling factor is not symmetric about Fermi energy unlike symmetic plateaus for pure Dirac case.
... In contrast, ACM arises from the non-adiabatic dynamics of carriers and can be excited by a protocols that activates inter-band transitions between the bands. Much as AV displays the static dc response, ACM is the physical manifestation of the dynamical response of anomalous Hall materials [16][17][18][19][20]. ...
Article
Full-text available
Non-trivial Bloch band overlaps endow rich phenomena to a wide variety of quantum materials. The most prominent example is a transverse current in the absence of a magnetic field (i.e. the anomalous Hall effect). Here we show that, in addition to a dc Hall effect, anomalous Hall materials possess circulating currents and cyclotron motion without magnetic field. These are generated from the intricate wavefunction dynamics within the unit cell. Curiously, anomalous cyclotron motion exhibits an intrinsic decay in time (even in pristine materials) displaying a characteristic power law decay. This reveals an intrinsic dephasing similar to that of inhomogeneous broadening of spins. Circulating currents can manifest as the emission of circularly polarized light pulses in response to an incident linearly polarized (pulsed) electric field, and provide a direct means of interrogating a type of Zitterbewegung of quantum materials with broken time reversal symmetry.
... The parameter m is for absorption scaling; here the value m 1 is used. The imaginary part of the interband surface conductivity, Eq. (8), can be found by the Kramers-Kronig relations [27]. The total surface conductivity used in this Letter is given by σ MoS 2 σ inter σ res . ...
Article
Full-text available
We propose a nanophotonic structure that gives high-degree quantum interference (QI) in the spontaneous emission (SE) of a quantum emitter (QE) in conjunction with strong light-matter coupling and non-Markovian dynamics. Specifically, we study the SE dynamics of a three-level V-type QE close to a MoS 2 nanodisk (ND). We combine quantum dynamics calculations with electromagnetic calculations and find reversible population dynamics in the QE, together with high-degree QI created by the ND. A rich population dynamics is obtained, depending on the energy of the QE with respect to the energies of the exciton-polariton resonances of the MoS 2 ND, the distance of the QE from the ND, and the initial state of the QE. Our results have potential applications in emergent quantum technologies.
Article
The generalized Drude formula is used to study conductivity properties of the two-dimensional weakly doped Holstein model with a free-electron-like dispersion. The relaxation processes associated with the scattering of conduction electrons by optical phonons are described in terms of a frequency- and temperature-dependent memory function. The imaginary and real parts of the memory function are analyzed in detail in the regime when characteristic energy scales of the problem, i.e., electron Fermi energy and optical phonon energy, are comparable in size. Results obtained at zero temperature and at finite temperatures are used to determine temperature effects in the real part of the dynamical conductivity as well as the frequency dependence of the optical electron mass and the electron relaxation rate. Finally, the characteristic fingerprints of a Holstein system with multiple phonon branches are identified in the dynamical electron conductivity.
Article
Full-text available
The semi-Dirac point of type-II semi-Dirac (SD) materials is a merging of triple Dirac points, distinguishing from the conventional type-I SD with double-Dirac point merging, and exhibits unique topological property. Here, we investigate its longitudinal and transverse optical conductivities. By controlling the evolution of the SD point with a perturbation parameter Δ, we present large longitudinal optical conductivity at the Van Hove singularity, not only in linear but also in parabolic directions. Furthermore, we find the nonzero dynamical Hall conductivity, which is sensitive to the Fermi energy and Dirac mass. Through introducing a momentum-dependent mass term, e.g., irradiating with circularly polarized light, the dynamical Hall conductivity exhibits more featured structures, depending on the parameter Δ, due to opening new channels of interband transitions. It is found that the frequency-dependent Kerr/Faraday angle can present all features of the dynamical Hall conductivity at characteristic frequencies. By detecting the Kerr or Faraday spectra, it is helpful to understand the physics of evolution of the SD to Dirac regime and further to extract the systemic parameters of SD materials from characteristic frequencies.
Article
We study the dynamics of entanglement for a composite two-qubit system, where each qubit is close to a molybdenum disulfide nanodisk, by combining quantum dynamics and classical electromagnetic calculations. We assume that each qubit interacts locally with its own photonic environment and its free-space decay time is in the picosecond regime. We observe non-Markovian entanglement dynamics indicating strong light-matter interaction between the qubits and the photonic reservoirs. We investigate the entanglement dynamics using entanglement of formation as a measure of entanglement by varying the initial state of the composite system, which always remains an extended Werner-like state, the transition frequency and the orientation of the transition dipole moment of each qubit, the radius of the molybdenum disulfide nanodisk, and the distance of a qubit from its corresponding nanodisk. For several values of the above parameters, we observe interesting phenomena during the time evolution of entanglement of formation, such as entanglement sudden death, revival of entanglement and entanglement trapping.
Article
Full-text available
We measured the response of the surface state spectrum of epitaxial Sb_{2}Te_{3} thin films to applied gate electric fields by low temperature scanning tunneling microscopy. The gate dependent shift of the Fermi level and the screening effect from bulk carriers vary as a function of film thickness. We observed a gap opening at the Dirac point for films thinner than four quintuple layers, due to the coupling of the top and bottom surfaces. Moreover, the top surface state band gap of the three quintuple layer films was found to be tunable by a back gate, indicating the possibility of observing a topological phase transition in this system. Our results are well explained by an effective model of 3D topological insulator thin films with structure inversion asymmetry, indicating that three quintuple layer Sb_{2}Te_{3} films are topologically nontrivial and belong to the quantum spin Hall insulator class.
Article
Full-text available
The bulk states of some materials, such as topological insulators, are described by a modified Dirac equation. Such systems may have trivial and non-trivial phases. In this article, we show that in the non-trivial phase a strong light-matter interaction exists in a two-dimensional system, which leads to an optical conductivity at least one order of magnitude larger than that of graphene.
Article
Full-text available
We show that inversion symmetry breaking together with spin-orbit coupling leads to coupled spin and valley physics in monolayer MoS2 and group-VI dichalcogenides, making possible controls of spin and valley in these 2D materials. The spin-valley coupling at the valence band edges suppresses spin and valley relaxation, as flip of each index alone is forbidden by the 0.1 eV valley contrasting spin splitting. Valley Hall and spin Hall effects coexist in both electron-doped and hole-doped systems. Optical interband transitions have frequency-dependent polarization selection rules which allow selective photoexcitation of carriers with various combination of valley and spin indices. Photo-induced spin Hall and valley Hall effects can generate long lived spin and valley accumulations on sample boundaries. The physics discussed here provides a route towards the integration of valleytronics and spintronics in multi-valley materials with strong spin-orbit coupling and inversion symmetry breaking.
Article
Full-text available
We present a three-band tight-binding (TB) model for describing the low-energy physics in monolayers of group-VIB transition metal dichalcogenides $MX_2$ ($M$=Mo, W; $X$=S, Se, Te). As the conduction and valence band edges are predominantly contributed by the $d_{z^{2}}$, $d_{xy}$, and $d_{x^{2}-y^{2}}$ orbitals of $M$ atoms, the TB model is constructed using these three orbitals based on the symmetries of the monolayers. Parameters of the TB model are fitted from the first-principles energy bands for all $MX_2$ monolayers. The TB model involving only the nearest-neighbor $M$-$M$ hoppings is sufficient to capture the band-edge properties in the $\pm K$ valleys, including the energy dispersions as well as the Berry curvatures. The TB model involving up to the third-nearest-neighbor $M$-$M$ hoppings can well reproduce the energy bands in the entire Brillouin zone. Spin-orbit coupling in valence bands is well accounted for by including the on-site spin-orbit interactions of $M$ atoms. The conduction band also exhibits a small valley-dependent spin splitting which has an overall sign difference between Mo$X_{2}$ and W$X_{2}$. We discuss the origins of these corrections to the three-band model. The three-band TB model developed here is efficient to account for low-energy physics in $MX_2$ monolayers, and its simplicity can be particularly useful in the study of many-body physics and physics of edge states.
Book
From the Contents: Introduction.- Starting from the Dirac equation.- Minimal lattice model for topological insulator.- Topological invariants.- Topological phases in one dimension.- Quantum spin Hall effect.- Three dimensional topological insulators.- Impurities and defects in topological insulators.- Topological superconductors and superfluids.- Majorana fermions in topological insulators.- Topological Anderson Insulator.- Summary: Symmetry and Topological Classification.
Article
The experimental observation of the long-sought quantum anomalous Hall effect was recently reported in magnetically doped topological insulator thin films [Chang et al., Science 340, 167 (2013)]. An intriguing observation is a rapid decrease from the quantized plateau in the Hall conductance, accompanied by a peak in the longitudinal conductance as a function of the gate voltage. Here, we present a quantum transport theory with an effective model for magnetic topological insulator thin films. The good agreement between theory and experiment reveals that the measured transport originates from a topologically nontrivial conduction band which, near its band edge, has concentrated Berry curvature and a local maximum in group velocity. The indispensable roles of the broken structure inversion and particle-hole symmetries are also revealed. The results are instructive for future experiments and transport studies based on first-principles calculations.
Article
We investigate transport properties of the surface states of three-dimensional topological insulator thin films in the presence of an electrostatic potential γ and a spin-splitting Zeeman field g. It is shown that there exist a quantum pseudospin Hall phase, a quantum anomalous Hall phase, and a common insulator phase in this system. By tuning g and γ, three quantum phase transitions between them can occur accompanied with energy-band closing, each of which is characterized by a change of the quantized pseudospin Chern numbers.
Article
Research on graphene and other two-dimensional atomic crystals is intense and is likely to remain one of the leading topics in condensed matter physics and materials science for many years. Looking beyond this field, isolated atomic planes can also be reassembled into designer heterostructures made layer by layer in a precisely chosen sequence. The first, already remarkably complex, such heterostructures (often referred to as 'van der Waals') have recently been fabricated and investigated, revealing unusual properties and new phenomena. Here we review this emerging research area and identify possible future directions. With steady improvement in fabrication techniques and using graphene's springboard, van der Waals heterostructures should develop into a large field of their own.
Article
Tuning the electronic properties of a material by subjecting it to strain constitutes an important strategy to enhance the performance of semiconducting electronic devices. Using local strain, confinement potentials for excitons can be engineered, with exciting possibilities for trapping excitons for quantum optics and for efficient collection of solar energy. Two-dimensional materials are able to withstand large strains before rupture, offering a unique opportunity to introduce large local strains. Here, we study atomically thin MoS2 layers with large local strains of up to 2.5% induced by controlled delamination from a substrate. Using simultaneous scanning Raman and photoluminescence imaging, we spatially resolve a direct bandgap reduction of up to 90 meV induced by local strain. We observe a funnel effect in which excitons drift hundreds of nanometers to lower bandgap regions before recombining, demonstrating exciton confinement by local strain. The observations are supported by an atomistic tight-binding model developed to predict the effect of inhomogeneous strain on the local electronic states in MoS2. The possibility of generating large strain-induced variations in exciton trapping potentials opens the door for a variety of applications in atomically thin materials including photovoltaics, quantum optics and two-dimensional optoelectronic devices.
Article
We have studied the optical conductivity of two-dimensional (2D) semiconducting transition metal dichalcogenides (STMDC) using ab-initio density functional theory (DFT). We find that this class of materials presents large optical response due to the phenomenon of band nesting. The tendency towards band nesting is enhanced by the presence of van Hove singularities in the band structure of these materials. Given that 2D crystals are atomically thin and naturally transparent, our results show that it is possible to have strong photon-electron interactions even in 2D