You are on page 1of 30

Gauge fields in spintronics

T. Fujita, M. B. A. Jalil, S. G. Tan, and S. Murakami


Citation: Journal of Applied Physics 110, 121301 (2011); doi: 10.1063/1.3665219
View online: http://dx.doi.org/10.1063/1.3665219
View Table of Contents: http://scitation.aip.org/content/aip/journal/jap/110/12?ver=pdfcov
Published by the AIP Publishing
Articles you may be interested in
Spintronics with graphene-hexagonal boron nitride van der Waals heterostructures
Appl. Phys. Lett. 105, 212405 (2014); 10.1063/1.4902814
Large-scale fabrication of BN tunnel barriers for graphene spintronics
J. Appl. Phys. 116, 074306 (2014); 10.1063/1.4893578
Magnetization reversal and spintronics of Ni/Graphene/Co induced by doped graphene
Appl. Phys. Lett. 102, 112403 (2013); 10.1063/1.4795764
Spintronic properties of graphene films grown on Ni(111) substrate
J. Appl. Phys. 110, 043704 (2011); 10.1063/1.3622618
Spintronic properties of zigzag-edged triangular graphene flakes
J. Appl. Phys. 108, 074301 (2010); 10.1063/1.3489919

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

JOURNAL OF APPLIED PHYSICS 110, 121301 (2011)

APPLIED PHYSICS REVIEWS

Gauge fields in spintronics


T. Fujita,1 M. B. A. Jalil,1,2,a) S. G. Tan,1,3 and S. Murakami4
1

Computational Nanoelectronics and Nano-device Laboratory, Electrical and Computer Engineering


Department, National University of Singapore, 4 Engineering Drive 3, 117576, Singapore
2
Information Storage Materials Laboratory, Electrical and Computer Engineering Department, National
University of Singapore, 4 Engineering Drive 3, 117576, Singapore
3
Data Storage Institute, A*STAR (Agency for Science, Technology and Research) DSI Building, 5 Engineering
Drive 1, 117608, Singapore
4
Department of Physics, Tokyo Institute of Technology, 2-12-1 Ookayama, Meguro-ku, Tokyo 152-8551, Japan

(Received 22 August 2011; accepted 9 November 2011; published online 22 December 2011)
We present an overview of gauge fields in spintronics, focusing on their origin and physical
consequences. Important topics, such as the Berry gauge field associated with adiabatic quantum
evolution as well as gauge fields arising from other non-adiabatic considerations, are discussed.
We examine the appearance and effects of gauge fields across three spaces, namely real-space,
momentum-space, and time, taking on a largely semiclassical approach. We seize the opportunity
to study other spin-like systems, including graphene, topological insulators, magnonics, and photonics, which emphasize the ubiquity and importance of gauge fields. We aim to provide an intuiC 2011 American
tive and pedagogical insight into the role played by gauge fields in spin transport. V
Institute of Physics. [doi:10.1063/1.3665219]

TABLE OF CONTENTS
I. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . .
II. TERMINOLOGY . . . . . . . . . . . . . . . . . . . . . . . . . . . .
III. SPIN-1/2 SYSTEMS IN THE PRESENCE OF
SPATIALLY VARYING MAGNETIC FIELD
TEXTURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
B. Berry phase: Theoretical approaches . . . . . . .
1. Berry (1984) . . . . . . . . . . . . . . . . . . . . . . . . .
2. Path integral formalism . . . . . . . . . . . . . . .
3. Unitary transformation . . . . . . . . . . . . . . . .
C. Physical consequences . . . . . . . . . . . . . . . . . . .
1. Spin-dependent forces: Chirality-driven
spin-Hall effect . . . . . . . . . . . . . . . . . . . . . .
2. Domain wall characterization . . . . . . . . . .
3. Spin torque in domain walls . . . . . . . . . . .
IV. REAL SPACE GAUGE FIELDS IN
GRAPHENE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
A. A primer on graphene . . . . . . . . . . . . . . . . . . .
B. Modeling the effects of strain. . . . . . . . . . . . .
C. Physical consequences . . . . . . . . . . . . . . . . . . .
1. Valley filtering . . . . . . . . . . . . . . . . . . . . . . .
2. Valley-dependent forces . . . . . . . . . . . . . . .
3. Edge states . . . . . . . . . . . . . . . . . . . . . . . . . .
V. SPIN-ORBIT COUPLING SYSTEMS: REAL
SPACE ANALYSIS . . . . . . . . . . . . . . . . . . . . . . . . .
a)

Electronic mail: elembaj@nus.edu.sg.

0021-8979/2011/110(12)/121301/29/$30.00

1
2

3
3
3
3
5
5
6
7
8
8
8
9
9
9
9
10
11
12

A. Spin-orbit coupling basics . . . . . . . . . . . . . . . .


B. Non-Abelian gauge field representation . . . .
C. Physical consequences . . . . . . . . . . . . . . . . . . .
1. Aharonov-Casher phase . . . . . . . . . . . . . . .
2. Spin-dependent transverse force . . . . . . . .
3. Quantum spin-Hall effect . . . . . . . . . . . . . .
4. Spin torque . . . . . . . . . . . . . . . . . . . . . . . . . .
D. Spatially nonuniform spin-orbit coupling . . .
VI. SPIN-ORBIT COUPLING SYSTEMS:
~
k-SPACE
ANALYSIS . . . . . . . . . . . . . . . . . . . . . . .
~
A. Derivation of k-space
Berry curvature . . . . .
B. Physical consequences . . . . . . . . . . . . . . . . . . .
1. Spin-Hall effect . . . . . . . . . . . . . . . . . . . . . .
2. Spin-Hall effect of light and optical
Magnus effect. . . . . . . . . . . . . . . . . . . . . . . .
3. Magnon-Hall effect in ferromagnetic
insulators . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4. Valley-Hall effect in graphene . . . . . . . . .
5. Topological insulators. . . . . . . . . . . . . . . . .
C. Hall conductivity. . . . . . . . . . . . . . . . . . . . . . . .
VII. TIME-DEPENDENT MAGNETIC SYSTEMS .
A. Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
B. Physical consequences . . . . . . . . . . . . . . . . . . .
1. Spin-Hall effect . . . . . . . . . . . . . . . . . . . . . .
2. Pseudospin-Hall effect in graphene . . . . .
3. Rayleigh scattering of polaritons. . . . . . . .
4. Spin motive force . . . . . . . . . . . . . . . . . . . .
~
C. Semiclassical connection with k-space
Berry curvature . . . . . . . . . . . . . . . . . . . . . . . . .
VIII. SUMMARY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

110, 121301-1

12
13
13
13
14
15
15
16
16
16
17
17
18
19
20
20
21
22
22
23
23
24
24
24
26
27

C 2011 American Institute of Physics


V

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-2

Fujita et al.

I. INTRODUCTION

Spintronics is the study of the quantum mechanical spin


property of carriers and its application to technology.1 There
are many avenues and tools available to study the motion of
spin species through the solid state, such as semiclassical
equations of motion, linear response Kubo formula, and
gauge theories. The latter entails the study of so-called gauge
fields, essentially generalizations of electromagnetism,
which naturally couple to the spin. In fact, the advantage of
working with gauge fields is that most of what we know
about usual magnetic and electric fields can be carried over
and applied to the richer realm of gauge fields. Thus, it
can be useful in deepening our understanding of certain
phenomena.
The use of gauge fields in spintronics kicked off in the
early 1980s with the study of the Aharonov-Casher (AC)
phase. The AC phase is essentially the spin-dependent version of the Aharonov-Bohm phase, whereby spin-orbit coupling can be regarded as a generalized (spin-dependent)
magnetic vector potential. This concept has recently been
revived in the modern language of Rashba spin-orbit coupling in semiconductor heterostructures. Observable consequences, such as spin separation due to generalized magnetic
Lorentz forces, have also been investigated. In 1984, around
the same time as the birth of the AC phase, the Berry phase
was discovered. The Berry phase results from cyclic, adiabatic transport of quantum states with respect to parameter
~ This fundaspace (e.g., real-space ~
r , momentum-space k).
mental result in quantum mechanics, which was previously
overlooked, has important consequences in spintronics. For
example, in the problem of spatially varying magnetic fields,
this theory predicts that carriers will experience spindependent trajectories, resulting in spin separation. Recent
works have also explored the possibility of domain wall
characterization and motion using these gauge fields. In the
separate context of spin-orbit coupling, the importance of the
Berry phase theory was realized in the early 2000s, following
the discovery of the intrinsic spin-Hall effect. The prediction
of ~
k-space gauge fields motivated efforts to find similar
effects in graphene, optics, and exciton systems. In this paper, we present a brief review of gauge fields in spintronics
with an emphasis on their origin and physical effects. As
such, much of the analysis is semiclassical, offering transparency to the reader.
This paper is organized as follows: In Sec. II, we introduce basic terminology, which is used throughout the paper,
by working through the simplest example of electromagnetism. In Sec. III, we introduce the concept of Berry gauge
fields. First, several theoretical approaches for studying the
Berry phase are presented. Then, the simple scenario of spatially varying magnetic fields is studied, which is shown to
give rise to a real-space (~
r) Berry gauge field. The physical
consequences of the ~
r-space gauge field are established,
including spin separation via spin-dependent forces and the
ability to impart spin torque on magnetic moments. We continue dealing with ~
r-space gauge fields in Sec. IV, although
we switch to a new platform, namely graphene. In particular,
we pay attention to strained graphene, which is modeled by

J. Appl. Phys. 110, 121301 (2011)

inclusion of a gauge field in the effective Hamiltonian


(though it is not Berry-like). The role of this gauge field in
so-called valley filtering and the quantum valley-Hall effect
are subsequently discussed. In Secs. V and VI, we discuss
gauge fields in the context of spin-orbit coupling, an important effect in the realm of spintronics. The former section
(V) views the spin-orbit coupling as a gauge field in ~
r-space;
analogous consequences to Sec. III occur, including
spin-dependent separation and spin torque. The latter section
(VI) describes the presence of a ~
k-space Berry gauge field in
general spin-orbit coupling systems. Physically, this drives
the intrinsic spin-Hall effect in semiconductors as well as
analogous effects in graphene, optical, magnonic, and topological insulator systems. Lastly, in Sec. VII, gauge fields in
time-space are discussed. This also drives the intrinsic spinHall effect in semiconductors and graphene. We close Sec.
VII with a discussion on the relationship between the two
intrinsic spin-Hall effect mechanisms in Secs. VI and VII.
Section VIII summarizes the paper.

II. TERMINOLOGY

We wish to introduce basic concepts of gauge fields by


examining the simplest example of electromagnetism. The
Hamiltonian of a free spin-12 particle in the presence of a
~ is given by
magnetic field B
H

2
1 
~ r  1 glB~
~
~
p eA~
r  B;
2m
2

(1)

where m is the effective carrier mass, ~


p hk~ is the momen~ r is the magnetic vector potentum, k~ is the wavevector, A~
~ B,
~ i.e., r  A
~ g is the Lande factor, lB is the
tial of B,
Bohr magneton quantifying the magnetic moment of the
intrinsic spin, and ~
r is the vector of Pauli spin matrices,






0 1
0 i
1 0
; ry
; rz
: (2)
rx
1 0
i 0
0 1
~ is a gauge field, and its curvature is
The vector potential A
~ defined through Bk lk @l A . One
the magnetic field B,
~ is not
~ the definition of A
should appreciate that, given B,
unique. For example, one could choose between the Landau
~ No matter the choice, the point is
or symmetric gauges for A.
~ is the same and the physics of the problem remains
that B
unchanged. For concreteness, let us consider the following
gauge transformation between two such possible values
~
for A,
~ rv;
~0 A
A

(3)

where v is a spatial scalar field. Since r  rv  0, it is evi~ B.


~0 r  A
~ Suppose the wavefunction
dent that r  A
~ gauge is jwi, while in the A
~0 gauge, it
of the system in the A
0
is jw i. If the physics were to remain invariant, we would
~! A
~0 corresponds
conjecture that the gauge transformation A
to a transformation of the wavefunctions jwi ! jw0 i
Ujwi, where U exp ik represents a change of phase.
Writing the Schrodinger equation, Hjwi jwi, in terms of

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-3

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

primed quantities ( is the energy eigenvalue), we get (ignoring the Zeeman term and after some algebra)
2
1 
~0  rv jw0 i jw0 i:
~
p
hrk eA
(4)
2m
Setting k hev, wethen have
2
1
~0 jw0 i jw0 i;
~
p eA
2m

(5)

which is of identical form to the original (unprimed) Schrodinger equation. Thus, the system remains invariant with
~! A
~ rv and
respect to
transformations A
 e the

jwi ! exp ihv jwi. In general, it is invariant to unitary
transformations (U belongs to the unitary group U1),
~0 A
~  ih UrU ;
A
e

(6)

jw0 i Ujwi:

(7)

Historically, the magnetic vector potential was considered


merely a mathematical instrument to describe the magnetic
~ represents the gauge invariant quantity.
field, since only B
~ itself is manifested as a phase, known as the
However, A
Aharonov-Bohm phase,2 determined by

~ r ;
r  A~
(8)
/AB d~
C

where C is a circuit in real space. This quantity is finite and


experimentally measurable when the interior of C contains a
~ but C itself lies exclusively in regions
magnetic field B,
~
where B is vanishing.
In the above, we have looked at the idea of gauge fields,
their curvatures, phases, and the idea of gauge transformations. This paper looks at generalizing the magnetic vector
potential for novel electronic and spintronic transport. For
example, in Sec. III, we shall look at the appearance of a special class of gauge fields, known as Berry gauge fields, and
~
their curvature. In a similar way that a real magnetic field B
produces physically observable effects, such as the Lorentz
force and electronic phase, we shall see, in Sec. III C, that
the Berry curvature produces analogous effects.
III. SPIN-12 SYSTEMS IN THE PRESENCE OF
SPATIALLY VARYING MAGNETIC FIELD TEXTURES

In 1984, Berry3 showed that, during adiabatic cyclic evolution of quantum states, a geometric phase necessarily accompa-

nies the evolution of the states (in addition to the usual


dynamic phase). This is known as the Berry phase. A special
case of this phase was actually discovered much earlier, in
1956, by Pancharatnam, who predicted and experimentally
verified that a phase is acquired by polarized light when it is
passed through a series of polarizers, such that the final and initial polarizations are the same4 (accordingly, the Berry phase is
also known as the Pancharatnam-Berry phase). The Berry
phase is topological in nature and is defined as the flux of the
Berry curvature cutting through the closed surface in the space
of parameters that the quantum states are evolving. The canonical example of the Berry curvature arises in the adiabatic transport of spins through a spatially non-uniform magnetic field
~ r. In this case, the states evolve with respect to B~
texture, B~
space, and we get a Berry curvature in that space. Of course, if
~ r is known,
the exact spatial magnetic field configuration B~
we can also define the Berry curvature in ~
r-space. Below, we
introduce three theoretical approaches of obtaining the curvature and Berry phase for the spatially inhomogeneous magnetic
field system.
A. System

The Hamiltonian of a free spin-12 particle in the presence


~ r is given by
of a spatially non-uniform magnetic field B~
2
1 
~ r  1 glB~
~ r;
~
p eA~
r  B~
(9)
H
2m
2
where m is the effective carrier mass, ~
p hk~is the momentum,
~ r is the magnetic vector potential of B,
~
k~is the wavevector, A~
~ B,
~ g is the Lande factor, and lB is the Bohr magi.e., r  A
neton quantifying the magnetic moment of the intrinsic spin.
B. Berry phase: Theoretical approaches
1. Berry (1984)

This approach takes on the spirit of Berrys original


paper.3 We study the cyclic, adiabatic spin evolution in a
~ r, described by
spatially inhomogeneous magnetic field B~
~
the Hamiltonian in Eq. (9). In the B-space,
one can conceive
a circuit C, illustrated in Fig. 1(a), traversed by the quantum
spin states with starting (ending) point at time t 0 (t T).
Adiabatic evolution corresponds to when C is traversed infinitesimally slowly, i.e., T ! 1. This implies that, when the
~
system is prepared in an initial eigenstate jwi B0i
(i f"; #g for the spin-12 system), it will evolve such that at

FIG. 1. (Color online) We illustrate the


Berry phase for the problem of inhomo~ (a) A Berry
geneous magnetic fields B.
phase is acquired when quantum states
undergo cyclic, adiabatic evolution C in
~
magnetic field B-space.
Adiabatic evolution corresponds to the limit T ! 1. (b)
~
The Berry curvature in Eq. (17) in Bspace is a Dirac monopole of strength
eg 12.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-4

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

~
time t > 0, it is in the state jwi Bti.
Specifically, the state
at time t can be written as having acquired a phase



i t 0 0
~
dt i t expici tjwi Bti;
(10)
jwti exp 
h 0
where the first factor is the usual dynamic phase (i is the
energy eigenvalue of jwi i). Berry pointed out that adiabatic
transport around a closed circuit in parameter space is necessarily accompanied by an additional non-zero phase ci . Since
jwti must obey Schrodingers equation, Hjwti i jwti
i
h@t jwti, we get, by direct substitution,



i t 0 0
i exp 
dt i t expici tjwi ti
h 0





i t 0 0 h
i
h exp 
dt i t expici tjw_ i ti
h 0




i
i t 0 0
_ expici tjwi ti i t exp 
dt i t
ict
h 0
~
 expici tjwi Bti;
(11)
where the over-dot notation symbolizes differentiation with
respect to time. Canceling the non-zero phase factors, simple
algebra yields
jw_ i ti ic_ i tjwi ti 0;

(12)

which, after applying the state bra hwi tj, gives the following relation satisfied by ci t:
c_ i t ihwi tjw_ i ti:

(13)

A well-defined phase is acquired by the state when it traver~


ses a circuit C in magnetic field B-space,

~  hwi Bjr
~ B jwi Bi:
~
(14)
ci C i dB

~ has the form of a Dirac magnetic


Remarkably, Xi B
monopole in Fig. 1(b), which we now show explicitly. The
two spin eigenstates of Eq. (9) are given by i "; #


1
expi/ cot h2
jw" i q
;
(18)
1
cot2 h2 1


1
 expi/ tan h2
;
(19)
jw# i q
1
tan2 h2 1
which correspond, respectively, to energy eigenvalues of
~ and # 1 glB jBj.
~ The angles h and /
"  12 glB jBj
2
are spherical angles, which parameterize the magnetic
~
~ Bx ; By ; Bz ; i.e.,q
h arccos B
and
field B
z =jBj
2
2
2
~
/ arctan By =Bx , where jBj Bx By Bz . The prefactors outside of the spinors ensure proper normalization of the
eigenstates. From the definition in Eq. (15), we can derive
~
the components of Aad:
" B for the up-spin eigenstate in
Eq. (18) as follows:
!
By
Bz
ad:
 1
;
ABx;" 
~
jBj
2 B2 B2
x

Aad:
By;"

!
Bz
 1
;
 
~
jBj
2 B2x B2y

(20)

Bx

Aad:
Bz;" 0:
Taking the curvature, XB rB  Aad:
B , we then arrive at the
following components for the Berry curvature:
XBx;"

Bx
~3
2jBj

; XBy;"

By
~3
2jBj

; XBz;"

Bz
~3
2jBj

(21)

that is, XB is the Dirac monopole field


~
X" B

~
B
~3
2jBj

(22)

Here, ci is the Berry phase. Intuitively, Eq. (14) looks like


the phase acquired by an electron when it encircles the vector potential of a magnetic field, i.e., the Aharonov-Bohm
effect in Eq. (8).
Accordingly, we can define the Berry gauge field in
~
B-space,
~
~
~
Aad:
i B ihwi BjrB jwi Bi;

(15)

where the superscript ad. reminds us of its adiabatic origin.


Using Stokess theorem, one can rewrite Eq. (14) in terms of
a surface integral,

~
(16)
ci d S~ rB  Aad:
i B;
where the quantity
~ rB  Aad: B
~
Xi B
i

(17)

~ has the meaning of


is the Berry curvature. Physically, Xi B
~
an effective magnetic field in B-space.

with strength eg 1=2. It is a simple exercise to show that


the Berry curvature for the down-spin eigenstate in Eq. (19)
is of equal magnitude, but opposite sign compared to the up~ X" B
~  B~ 3 . Emphasis should
spin curvature, X# B
~
2jBj
be placed on the role of degeneracies. The Berry curvature
exhibits monopole-like singularities at points in the parameter space where the system is degenerate.3 In the case of a
~ this corresponds to the origin B
~ 0 in
magnetic field B,
Eq. (22) where the Zeeman splitting vanishes.
~
The notion of the Berry curvature in B-space
is somewhat abstract and difficult to perceive physically. Since we
are dealing with spatially inhomogeneous magnetic fields,
~ B~
~ r , we can define an equivalent curvature in ~
B
r -space,
which has the physical meaning of a magnetic field (see
Sec. II). In particular, Eq. (14) for the Berry phase can be
written in terms of a circuit in Cr in real-space,

ci Cr i
d~
r  hwi ~
rjrr jwi ~
ri;
(23)
Cr

r
Aad:
i ~

ihwi ~
rjrr jwi ~
r i is the
where the quantity
~
r -space Berry gauge field which mimics the usual magnetic

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-5

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

vector potential. The curvature X~


r then represents a magnetic field.
2. Path integral formalism

The path integral approach in quantum mechanics perhaps gives us the most intuitive explanation for the Berry
phase. In this formalism, the Berry phase appears as a firstorder accumulated phase of overlapping eigenstates as they
evolve in space, ~
r ~
rt, where t parameterizes the motion.
In particular, we dissect the time parameter t into discrete
steps of size tn1  tn Dt, compute the evolution of the
states between steps (through propagator operators), then
finally apply the adiabatic limit Dt ! 0. The final eigenstate
is then found to contain the Berry phase upon completion of
a circuit in parameter space. Suppose that, at time tn , the system is in a spin eigenstate jwi ~
rn ; tn i of the magnetic field
~ rn (i f"; #g), where ~
rn ~
r tn denotes the correspondB~
ing position in real space. At a short time later, tn1 , the
wavefunction can be described by the path integral

rn1 ; tn1 d~
rn G~
rn1 tn1 ;~
rn tn wi ~
rn ; tn ;
(24)
wi ~
rn tn h~
rn1 jUtn1 tn j~
rn i is the propagawhere G~
rn1 tn1 ;~
tor between time tn and tn1 and Utn1 tn
expiHtn1  tn . The propagator can be readily evaluated as

! +
*

~
iDtl~
rB

h~
rn1 jUtn1 tn j~
rn i wi;n1 exp 
wi;n ;


h
!
~
iDtljBj
hwi;n1 jwi;n i exp 6
;
(25)
h
where the second line follows from first-order approximation
of the exponential, and the signs 6 refer to the up and down
spin eigenstates. The phase factor in Eq. (25) is just the usual
dynamic phase, which we shall ignore hereafter. Evaluating
the more interesting overlap between
evolved states,


we obtain hwi;n1 jwi;n i 1  hwi;n1 j jwi;n1 i  wi;n i  1


jw_ i  exp Dthw
jw_ i . We shall see that
Dthw
i;n1

i;n

i;n1

i;n

this overlap-induced phase is the Berry phase in the adiabatic


limit, i.e., Dt ! 0. Thus, the incremental evolution represents an additional phase factor of Dt  ihwi;n1 jw_ i;n i.
Evolution over an interval t 2 0; T, where ~
r0 ~
rT,

upon taking the limit Dt ! 0, results in an accumulated


phase factor of
T 

ci dt ihwi jw_ i i ;
0

(26)
d~
r  hwi ~
rjrr jwi ~
r i;
i
Cr

which is exactly the Berry phase in Eq. (23).


3. Unitary transformation

The Berry curvature can be alternatively derived from a


consideration of local unitary transformations. Another way
of looking at the spatially varying system described by
Eq. (9) is to consider locking the reference spin axis to the
magnetic field at each point in ~
r-space through a local transformation. In the laboratory (L) frame, the magnetic field
~ Bj
~ rotates as we move from one point in space
axis ~
n B=j
to another. Consider applying a local transformation to the
system which rotates the L-frame at each point ~
r such that
the reference spin axisthe ^z-axis by conventionpoints
along ~
n~
r. Then, the problem is effectively mapped to a
spatially uniform system; this is illustrated in Fig. 2. We
shall see, however, that this transformation is necessarily
accompanied by a gauge transformation of the coordinate
space ~
r . To facilitate the local rotation of L, we employ a
unitary matrix U U~
r , which satisfies U~
r~
nU rz at
every point ~
r. As one may have guessed, U is not unique,
and different Us are connected via gauge transformations
akin to Eqs. (6)(7). One such U is given by5
~~
Um
r;


(27)

h

~ sin h2 cos /; sin h2 sin /; cos 2 and h; / are


where m
~
spherical
angles
parameterizing
the
vector
n
sin h cos /; sin h sin /; cos h. Wavefunctions are transformed as jwi ! Ujwi (note here that jwi are twocomponent spinors), and the transformed Hamiltonian reads
H0 UHU

2
1 
~ r U  glB rz jB~
~ r j: (28)
U ~
p eA~
2m

Evaluating the first term on the right-hand side of Eq. (28) is


nontrivial, because ~
p ih@r represents spatial derivatives
and U explicitly depends on ~
r. One can readily show that the
effective Hamiltonian can be reformulated to read

FIG. 2. (Color online) A spatially varying magnetic field system (left), charac~ r , can be transformed via
terized by B~
local rotation to a locally uniform system (right). The effect of this is to modify the momentum of the carriers
~ where A
~ is a gauge field (see
~
p!~
p A,
~ is a true gauge field, in the
Eq. (28)). A
sense that it obeys the transformation
rule in Eq. (30) with respect to unitary
transformations.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-6

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

2
1 
~  ihUrU  1 glB rz jB~
~ rj;
~
p eU AU
2m
2
(29)
2 1
1 
~ hA~
~ r j;
~

p eU AU
r  glB rz jB~
2m
2

H0

where A  iUrU is a spatial gauge field induced by the


unitary transformation. Note that, unlike the usual vector
~ the components of A are 2  2 matrices which
potential A,
encode the spin state. Notably, A is an example of a YangMills or non-Abelian gauge field, whose components do not
commute. The gauge transformation rule of A is indicated
by Eq. (29) above to be
h
(30)
A0 UAU  i UrU ;
e
which generalizes Eq. (6). The Yang-Mills curvature of A is
defined as6
ie
Xk  Xl @l A  @ Al  A ; Al ;
(31)
ch
which is the usual curl portion in addition to noncommuting
corrections. Just like the usual magnetic field, X is invariant
with respect to A, which represents a freedom to choose the
gauge potential, and also manifests itself in physical effects.
Returning to our gauge field A iUrU , computation of
the curvature surprisingly results in X 0! This is because A
is a pure gauge.7 In order to induce a finite curvature, we consider the so-called adiabatic approximation, in which we
assume that carriers remain in the spin eigenstates of the quantum system as they travel between two points in ~
r-space and
that flipping between states is forbidden. Mathematically, this is
equivalent to diminishing the off-diagonal components of A to
zero and retaining only the diagonal (i.e., rz ) components of A.
The resulting adiabatic (ad.) gauge field, Aad: , has the form
1

1  cos hr/
0
ad:
2
A
;
(32)
0
121  cos hr/
which can be written in compact form Aad:
6
6121  cos hr/, where () denotes the up-spin "
eigenstate parallel (down-spin #, anti-parallel) to the local
~ r. By definition, Aad: is exactly the Berry
magnetic field B~
6
gauge field, and its curvature is the Berry curvature. Note
that, after taking the adiabatic approximation, the gauge field
becomes Abelian (i.e., whose components commute), the
noncommuting correction to Eq. (31) vanishes, and the curvature is prescribed by the usual curl, rr  Aad:
6 . In general,
however, the Berry gauge field can be non-Abelian; examples are discussed in Secs. VI B 1 and VI B 2. It is a simple
~
exercise to prove in the current example that Aad:
6 in B-space
describes the gauge potential of a magnetic monopole, in
agreement with the discussion in Sec. III B 1. In particular,
1
~
~ arccos Bz =jBj,
~
Aad:
hB
6 B 621  cos hrB /, where
q
~ arctan By =Bz , and jBj
~ B2x B2y B2z , and the
/B
~
corresponding Berry curvature in B-space
is
~ @Bl Aad: B
~  @B Aad: B
~ 6kl
Xl B
;6
l;6

Bk
~3
2jBj

; (33)

which is the expected result. As alluded to earlier, different


unitary matrices U can be defined to perform the required
local coordinate transformation, and each of these will inevitably derive a different adiabatic gauge potential A. The
gauge potentials are, however, connected by the gauge transformation rule in Eq. (30), and their curvatures are invariably
the Dirac monopole field.
In applying the adiabatic approximation above, we simply threw away the off-diagonal terms of A. Let us refine
this approximation a little, by first expanding the gauge field
as A Ax rx Ay ry Az rz . Obviously, in this notation,
Az Aad: represents the adiabatic Berry component of the
gauge field, while Ax and Ay represent off-diagonal components of A. Let us introduce the symbol AN to collectively
represent
the
off-diagonal
components,
i.e.,
A Az rz AN . Expanding the kinetic term, we have
(ignoring h for simplicity)
2
1
1 
~
~
p A 2
p Az r z AN ;
2m
2m
1
p Az rz 2
~

2m

1 
f~
p Az rz ; AN g AN 2 ;

2m

(34)

where fg denotes the anticommutator. In the second term on


the right-hand side, we ignore f~
p; AN g (rx ! 0, ry ! 0 in
the adiabatic limit), while fAz rz ; AN g 0, because of the
anticommutation property of the Pauli matrices. In the final
term, AN 2 A2x A2y Ax Ay frx ; ry g A2x A2y . From
our computed A, we find that the off-diagonal components
are characterized by
1
Ax  cos / sin hr/ sin /rh;
2
1
Ay  cos /rh  sin / sin hr/;
2

(35)
(36)

and so the effective Hamiltonian reads8,9


1
1
p A z r z 2
AN 2 ;
~
2m
2m

1
1 
p A z r z 2
rh2 sin2 hr/2 :
~

2m
8m

(37)
(38)

Thus, we find that the off-diagonal components of A contribute an additional scalar potential, U. Note that, in contrast to
the adiabatic vector potential in Eq. (32), the scalar potential
is spin-independent. Thus, in spintronic applications, the
term U is usually neglected (but see its effect on charge
transport in Eq. (172)).
C. Physical consequences

The Berry curvature due to a spatially inhomogeneous


~ r has physical and potentially practical
magnetic field B~
implications. The curvature itself acts analogously to a real
magnetic field, which is spin eigenstate-dependent and can
produce a Hall effect (Sec. III C 1) and assist in magnetization switching (Sec. III C 3). When examining such effects,
one must be careful to deal with the Berry curvature in the

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-7

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

~
~
r-space, X~
r (and not the B-space).
The curvature in real
space is the one that represents a real magnetic field, which
~
is felt by spin carriers. Although the B-space
curvature is
always a Dirac monopole, the form of the ~
r -space curvature
~ r . To evaludepends on the exact spatial configuration of B~
ate X~
r , one can write down the ~
r-space Berry gauge field
(e.g., in Eq. (32)) and then take the real-space curvature in
Eq. (31) (having in mind the explicit spatial dependence of
angles h and /). Alternatively, it is possible to transform
~ directly via the relation10,11
XB
!
~ r @ B~
~ r
@
B~
~ 
;
(39)
Xk ~
r XB

@ri
@rj


~
@~
n @~
n
n
6 

;
(40)
2 @ri @rj
~ Bj
~ is the unit vector pointing in the direction
where ~
n B=j
~
of B~
r and i; j; k are spatial coordinates.
1. Spin-dependent forces: Chirality-driven spin-Hall
effect

The Berry curvature X~


r acts as a spin eigenstatedependent magnetic field and gives rise to eigenstatedependent Lorentz forces. The quantum mechanical force
operator is defined by
~ m ~
t; H;
F
ih

(41)

where ~
t i1h~
r ; H is the velocity operator in the Heisenberg
~2
picture. Writing the Hamiltonian as H P
2m (we ignore
the Zeeman term, as
to the Stern-Gerlach force),
 this just leads

~ ~
~ Aad:  ~
p eAR contains all the
where P
pe A
gauge fields, the velocity operator (jth component) reads
i
1 h
rj ; P2x P2y P2z ;
2im
h
2 i
1 h 
rj ; pj eAR; j ;

2im
h

1   2

rj pj epj AR; j eAR; j pj e2 A2R; j


2im
h
 

tj

 p2j epj AR; j eAR; j pj e2 A2R; j rj :

(42)

In the above, rj ; A2R; j  0. Taking note that pj  ih@j is a


spatial derivative operator, we also have, by the product rule,
rj p2j  p2j rj rj pj pj  pj rj pj pj rj pj  pj pj rj ;
rj ; pj pj pj rj ; pj ;
2ihpj :

(43)

The four remaining first-order in pj terms in Eq. (42) can be


similarly evaluated and contribute a term 2iehAR; j . Thus, the
velocity operator equals
tj

 Pj
1
:
pj eAR; j
m
m

The force operator is then

(44)

1
Pj ; P2x P2y ;
2imh
1

Pj ; P2i  i 6 j;
2imh

1 
Pi ; Pj Pi Pi Pi ; Pj  :

2imh

Fj

(45)

Under the gauge fields, the covariant momenta Pi are


noncommuting,


Pi ; Pj   ieh @i AR; j  @j AR;i e2 AR;i ; AR; j ;
  iehXk ;

(46)

where the last line follows from Eq. (31). Substituting this
expression into Eq. (45), we obtain, for the force,
Fj eti Xk ~
rijk ;

(47)

which is the usual magnetic Lorentz force, but with contributions


~ and A. At this stage, a few remarks are in order.
from both A
Firstly, the exact meaning of the force operator in Eq. (47) is a
little ambiguous, because there is no concept of force in quantum
mechanics (due to Heisenbergs uncertainty principle). The definitive meaning can be obtained by deriving the real-space equations of motion of a wave packet eigenstate of the system; see,
for example, Ref. 12. It turns out that the force is indeed reflected
as a semiclassical acceleration term in the equations of motion.
Secondly, the Lorentz force due to the ~
r -space Berry curvature
can also be understood semiclassically from a consideration of
adiabatically evolving spin moments with respect to time, t.8
~
However, we shall defer the details to our discussion of the kspace Berry curvature, which follows analogously20 and is responsible for an array of spin-dependent phenomena. Thirdly,
the magnitude of the force in Eq. (47) is independent of the
~ rj. One can
strength of the spatially varying magnetic field, jB~
see this by observing the expression for X~
r in Eq. (40) and not~ The only requirement
ing that it carries no dependence on jBj.
~ is that it is sufficiently strong to guarantee adiabatic spin
for jBj
transport; once this is fulfilled, the resulting Berry curvature
depends only on how the field direction varies in ~
r-space and not
on its magnitude. At this point, one might ask: how strong is
strong? This can be answered more readily when we talk about
evolutions with respect to time, and we defer the conditions for
adiabaticity to Sec. VII (see Eq. (155)).
Chirality-driven electron dynamics constitutes the interesting topological transport effect.10 The Lorentz force in
Eq. (47) was used to describe the topological Hall effect by
Bruno et al.9,21 in two-dimensional electron gas (2DEG)
based nanostructures, with patterned ferromagnetic cylinders
providing the spatially nonuniform magnetic fields. The
authors in Ref. 9 assumed a fully spin-polarized 2DEG,
which corresponds to selecting only one of the gauge fields,
1
Aad:
" 21  cos hr/ in Eq. (32). This represents an
ordinary magnetic vector potential for the carriers, leading to
a Hall effect induced by the Lorentz force in Eq. (47). The
authors chose a magnetic field configuration with a zero spatial average (thus ruling out the ordinary Hall effect), but
whose nonuniformity results in a Berry curvature in Eq. (40)
r 6 0. To the first
with a non-zero spatial average, Xz;ave: ~

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-8

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

order, the resulting transverse force on the carriers is proportional to Xz;ave: ,



h
~
r
n  @y ~
n dxdy;
(48)
F / Xz;ave: ~
n  @x ~
eA0
~ Bj
~ in Eq. (40) and the integration is carried
where ~
n B=j
out over the unit cell area A0 . The factor h=e comes about by
converting the units for the curvature (1=r 2 ) to Teslas. The
application of an additional external magnetic field was
shown to result in a change of topology of the system, affecting the Hall conductivity in an unconventional way and providing an unambiguous signature of the effect. From the
semiclassical Drude theory, the Hall conductivity rxy (for
spin-up electrons) is proportional to the effective magnetic
field, which is given in Eq. (48),9
rxx eXz;ave: s
ne2 s
; rxx
;
rxy
m
m

(49)

chlores (compounds of form R2 Mo2 O7 , where R Nd, Gd,


Sm)1517 arising from carrier hopping in lattices with nontrivial spin textures. Experimental signatures of the topological
transport effect have been observed in Refs. 18 and 19 which
conclude that the spin chirality effect can play a significant
role in determining carrier transport.
2. Domain wall characterization

A domain wall (DW) refers to the interface between two


magnetic domains and is usually characterized by a smoothly
~ r across a finite distance. When
varying magnetization M~
electronic current is passed through a DW, the conduction
electron spins experience an exchange coupling represented
by the Hamiltonian
~ r ;
Hex JH ~
r  M~

As we have seen, the Berry gauge field in ~


r -space mimics
the usual magnetic vector potential. One possible application
r is its effect on the local magnetization in ferromagof Aad: ~
nets (FM). Bazaliy et al.23 considered the case of a FM system
~ r , such as a DW.
with a smoothly varying magnetization, M~
In the limit of large exchange-coupling JH ! 1, a charge
current will be fully spin polarized (in the spin-up eigenstate,
jw" i) within the DW and experience a Berry gauge field
Aad:
r hw" jA~
r jw" i. In Sec. III C 1, we discussed the
" ~
r on the orbital motion of conducting electrons.
effect of Aad: ~
Conventional electrodynamics governs that A~
r also
describes the interaction between the current ~
j and the local
~ r with the following energy density term:
magnetization M~
j"  Aad:
r;
int ~
" ~

where n is the electron density and s is the mean free time.


In a similar context, Tan et al.12 studied a 2DEG-based system with a mostly vertical magnetic field that has a small
superimposed spatial nonuniformity. In contrast to Ref. 9,
the weak magnetic field strength allowed both spin eigenstates to be considered, and the Lorentz force induced a Hall
separation of oppositely polarized spins (this was balanced
by another force, due to spin-orbit coupling, which is
described in Sec. V C 2).
An analogous situation has been proposed in ferromagnetic manganites (e.g., La2 (Ca,Pb)1 MnO3 )13,14 and pyro3

3. Spin torque in domain walls

(50)

where JH is the exchange coupling constant, which is positive


in ferromagnets. Clearly in the limit of strong exchange coupling, JH ! 1, the current will be fully spin-up polarized at
each spatial point within the DW. Thus, the adiabatic requirement is fulfilled in such cases and provides a suitable platform
for examining adiabatic spatial gauge fields. For example, Lee
et al.22 proposed a method for characterizing the spatial configuration of DWs based on Hall conductivity. Clearly, rxy in
Eq. (49) is highly sensitive to the spatial texture of a DW.
Thus, experimental measurement of rxy may provide a novel
approach for probing the magnetic configuration of DWs.22

(51)

where ~
j" is the electric current of the up-spin carriers. Intuitively, one would expect the localized spins to reorient themselves to minimize the interaction energy int . Thus, the local
magnetization dynamics should be governed by the gradient
~ r , i.e., they will experience
of the energy with respect to M~
an additional, topological switching field due to the Berry
r given by23
gauge field Aad:
" ~
~top 1 rn int ;
H
lM

h ji
@i ~
n~
n;
2M e

(52)
(53)

~
where ~
n M=M.
This additional switching field in the
presence of a current in magnets modifies the well-known
Landau-Lifshitz Gilbert equation,
~  M;
~
~_ ge H
M
2m

(54)

~ includes the contribution from H


~top in addition to
where H
the usual precession and damping terms. It is predicted that
the topological contribution in Eq. (53) can play an impor~ Bazaily et al.23 pretant role in the local dynamics of M.
~top in DWs can significantly deflect the local
dicted that H
magnetization within realistic values of current densities.

IV. REAL SPACE GAUGE FIELDS IN GRAPHENE

In Sec. III, we studied spin-12 systems, in which there was a


spatially non-uniform magnetic field. It was found that adiabatic transport of the spin eigenstates resulted in a Berry gauge
~ and, subsequently, ~
field in B
r-space, which had physical significance, such as a spin-dependent Lorentz force and Berry
phase. In this section, we review a gauge field in ~
r-space in graphene, which has a distinct origin from the Berry potential. In
particular, this gauge field is not associated with adiabatic transport and, indeed, does not require it. Instead, it arises when we
model the effect of strain in the graphene lattice. The point we
wish to make is that not all gauge potentials are Berry-like in
nature (we shall study a further instance in Sec. V). However,
in many aspects, their physical consequences mirror that of the

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-9

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

Berry curvature, such as the induced Lorentz force, which gives


rise to Hall effects.
A. A primer on graphene

Graphene is a single, atomically thin sheet of graphite.24


Since its experimental discovery in 2003,25,26 graphene has
become an important material, due to its remarkable electronic, optical, mechanical, and thermal properties. Plenty of
comprehensive details on graphene are available in the literature; see, for example, Ref. 27. Below, we give a primer on
basic graphene physics. Graphene refers to a single, isolated
sheet of graphite, whose hexagonal lattice is comprised of
two interpenetrating triangular sublattices, A and B, as illustrated inpFig.
3(a). The

p spatial lattice vectors are


is the
~
a2 a23;  3, where a 1:42 A
a1 a23; 3, ~
length of carboncarbon bonds. The three nearest neighbor
carbon p
atoms
are separated
by spatial vectors

p
~
d2 a21;  3; ~
d3 a1; 0. We can write
d1 a21; 3; ~
down the nearest neighbor tight-binding Hamiltonian as
X
ar;i br;j h:c:;
(55)
H t
hi;ji;r

where ar;i (ar;i ) is the electron creation (annihilation) opera~i on sublattice A for spin r (the b operator on lattice site R
tors are equivalent, but for sublattice B). The nearest
neighbor hopping integral in Eq. (55) has approximate value
t 2:8 eV. The energy dispersion of H is
q
~ 6t 3 f k
~;
(56)
6 k

p
p
~ 2 cos 3ky a 4 cos 3ky a cos3kx a and the
where f k
2
2
sign () indexes the conduction and valence bands,
respectively. Equation (56) has degeneracies at the six corners of the Brillouin zone (see Fig. 3(b)), of which there are
two inequivalent types K and K 0 . These are known as the valleys of graphene. Expanding the dispersion in Eq. (56)
around the two valleys, we obtain

6 ~
q 6tF j~
qj;

(57)

q k~ K, and j~
qj  jKj (or K 0 ).
where tF 106 ms1 , ~
The resulting conical dispersion suggests the presence of
massless Dirac fermions, whose velocity is independent of
energy, in stark contrast with conduction electrons in conventional semiconductors. This remarkable fact gives rise to
the many intriguing physical phenomena discovered in graphene. The corresponding low energy Hamiltonian of the
Dirac carriers in the K (K 0 ) valley are
H tF ~
r~
p;
0

p;
r ~
H tF~

(58)
(59)

where ~
r rx ; ry is a vector of Pauli matrices. Here,
the Pauli operators do not act on the electron spin, but rather
on the pseudospin, which indexes electron density on the
two sublattices A and B; e.g., the quantum state 1; 0T
denotes an electron situated on sublattice A.
B. Modeling the effects of strain

As an atomically thin sheet of carbon atoms, graphene can


be regarded as a stretchable, flexible membrane. In a practical
setting, strain has been introduced in graphene sheets via deposition onto stretchable substrates (polymers) and by suspending
the sheets across trenches defined by nanolithography. Mathematically, we can model the effect of strain by introducing a
perturbation to the local hopping parameter, t ! t dt. It has
been shown,28,29 that this is equivalent to the introduction of a
~S (S denotes strain),
real-space vector potential A


~S ~
A
r

;
(60)
H tF~
r ~
p  t1
F


~S r ;
H0 tF~
p t1
(61)
r0  ~
F A ~
~S has the form28
where A
1
ASx dt1 ~
r  dt2 ~
r dt3 ~
r ;
2
p
3
dt2 ~
r  dt3 ~
r :
ASy
2

(62)

r (i 1; 2; 3) are the local perturbations


In the above, the dti ~
to the three nearest neighbor hopping integrals, due to strain
~S has opposite signs in the
(see Fig. 3(a)). The gauge field A
two valleys: this preserves the overall time-reversal (TR)
symmetry of the system (since TR switches the valleys), distinguishing the strain-induced gauge field from the magnetic
~ of a magnetic field. However, A
~S can still
vector potential A
affect carriers as if it were a magnetic field, only now, carriers
in the two valleys K and K 0 experience the effect of oppositely directed magnetic fields. That is, the strain can induce a
valley-dependent magnetic field. Some possible uses of the
strain-induced gauge field are discussed below.
FIG. 3. (Color online) (a) Monolayer graphene comprised of two triangular
a2 are unit lattice vectors, and ~
di (i 1; 2; 3) are
sublattices A and B. ~
a1 and ~
the three nearest neighbor vectors. (b) The Brillouin zone of monolayer graphene is hexagonal, with two inequivalent corners K and K 0 , known as valleys. The energy spectrum is degenerate at the corners, and the low energy
Hamiltonian is centered about them in a Dirac cone configuration with slope
tF 1  106 ms1 .

C. Physical consequences
1. Valley filtering

The valley-dependent gauge fields arising from strain


can be utilized to facilitate valley filtering.30 A valley filter

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-10

Fujita et al.

preferentially transmits one of graphenes valleys over the


other, thereby producing at its output a valley-polarized current. Such a device would be useful in the area of graphene
valleytronics,33 which aims to utilize the valley degree of
freedom in technology in an analogy with the electron spin.
For concreteness, consider electron transmission across a
region of uniform uniaxial strain (length L1 ) along the armchair direction (^
y-direction) of monolayer graphene, as
shown in Fig. 4(a). From Eq. (62), the strain-induced gauge
~S dt^
x. In the spirit of traditional tunneling
field is simply A
problems, we begin by writing down the quantum mechanical wavefunctions in the three regions of the system. Translational invariance along x^ permits solutions in the K-valley
of the form Wx; y expikx xWy, where
 


1
1
iky
iky
Re
;
(63)
WI y e
ei/
ei/
 


1
1
iqy
iqy
WII y Ae
Be
;
(64)
eiu
eiu
 
1
iky
;
(65)
WIII y Te
ei/
and kx2 k2 2 kx dt2 q2 . For the K 0 -valley, the
replacements / ! / and u ! u are necessary. The
phases / and u are parameterized as kx  cos / and
kx dt  cos u, respectively, where the upper (lower) sign
corresponds to valley K (K 0 ) and are related via the conservation of kx across the first interface at y 0:
 cos /  cos u6dt (we assume dt > 0 and  > 0). The
transmission probability through the strained graphene is
characterized by T jTj2 , where T is the transmission coefficient in Eq. (65). Traditionally, the literature has studied
the wavevector filtering effect of magnetic vector potentials
due to magnetic barriers34 (the direction of the incident
wavevector is parameterized by the phase /). Here, we deal
with an analogous situation of TR symmetric gauge fields
due to strain. The filtering property of the present system can
be analyzed qualitatively from the following arguments:
From kx2 k2 2 , we require jkx j
 for non-vanishing
transmission. Similarly, from kx dt2 q2 2 , we get
6dt
kx
6dt. Combining the two inequalities, /
must satisfy 1 dt=
cos /
1 in the K-valley and
1
cos /
1  dt= in the K 0 -valley. Thus, high electron
transmission is restricted to valley-dependent windows of

FIG. 4. (Color online) Valley filter device comprised of (a) a region of uniform uniaxial strain along the armchair direction (strained bonds are highlighted) followed by (b) a magnetic barrier region.

J. Appl. Phys. 110, 121301 (2011)

/2

0; arccos1 dt=; K-valley


:
arccos1  dt=; p;
K 0 -valley

(66)

Evidently, when dt 0 (no strain), electron transmission


spans the entire spectrum of / 2 0; p for both valleys. The
onset of total separation occurs at  dt when each valley is
restricted to exactly half of the /-spectrum in Eq. (66). In
Fig. 5, we plot T as a function of / [for strain dt 25 meV
and L1 250 nm], confirming the remarkable separation of
valleys in /-space when the energy  dt. This separation
of the valleys in /-space forms the basis of a valley filter, for
one now simply needs to perform ordinary wavevector /-filtering using external magnetic fields (a possible setup is
illustrated in Fig. 4(b)); see, for example, Refs. 3537. By
exclusively selecting a single valley with an appropriate
/-filter, it was found that this method could produce a valley
polarization exceeding 99% within experimentally relevant
parameters.30 Other proposals of valley filtering using strains
in graphene are addressed in Refs. 31 and 32.
2. Valley-dependent forces

In Sec. III C 1, we established that real-space gauge


fields give rise to magnetic Lorentz forces, which act on carriers. The Lorentz force is responsible for the classical Hall
effect38 and, in the quantum limit, the quantum-Hall effect
(QHE).39 The latter is experimentally accessible in 2DEG
systems in the presence of large perpendicular magnetic
fields (> 5 T) and low temperatures. In the QHE, electrons
occupy Landau levels, which are insulating in the bulk, but
metallic at the edges (see Fig. 6(a)). These metallic edge
states are strictly unidirectional, one-dimensional channels,
in which backscattering is forbidden (the broken TR symmetry defines a definite chirality), giving rise to the precisely
quantized Hall conductivity. By the same token, we can
expect that real-space gauge fields will also lead to quantum
Hall states, provided that they correspond to large vertical

FIG. 5. (Color online) Transmission probability through the strained region


of Fig. 4(a) as a function of / (the direction of the incident wavevector).
Valley K (K 0 ) is denoted by solid (dashed) lines. Filtering in /-space, facilitated by magnetic field barriers shown in Fig. 4(b), therefore, results in valley filtering. The shaded region indicates the action of a /-filter, which
allows for valley K to be transmitted freely, while blocking valley K 0 .

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-11

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

FIG. 6. (Color online) (a) The quantum Hall state, induced by strong perpendicular magnetic fields in two-dimensional systems, comprised of an insulating
bulk and metallic edges. The edge states are chiral (uni-directional), owing to the broken TR symmetry, and consequently resist backscattering by impurities.
(b) The quantum valley-Hall effect arising in strained graphene is a TR symmetric cousin of the quantum Hall effect. Here, electrons in opposite valleys
(K and K 0 ) experience opposite vertical magnetic fields, which form two valley-resolved (chiral and anti-chiral) quantum Hall states. Backscattering of edge
states is only permitted with a corresponding switching of valleys. Thus, in the absence of TR-breaking impurities, the edge states will remain robust.

effective magnetic fields (defined, of course, by the curvature). Remarkably, it has been found experimentally that
strongly localized strains in graphene can result in a measurable effective magnetic field strength of up to 300 T,40 which
is well into the limit required by the QHE. Unlike the usual
QHE, however, the strained graphene system preserves TR
symmetry, which is reflected by the equal and opposite
effective magnetic fields experienced by carriers in opposite
~S B
~S K 0 . Due to the 2D nature
~S K r  A
valleys: B
of graphene, only the vertical ^z-component of this magnetic
field is finite,
!
@ASy @ASx
S
~

^z:
(67)
B
@x
@y
Intuitively, this suggests the presence of two valley-resolved
copies of the QHE; one comprised of chiral K quantum-Hall
states, and the other of antichiral K 0 states, as illustrated in
Fig. 6(b), which result in valley-filtered counter propagating
edge states. Provided intervalley (K $ K 0 ) scattering is
neglected, the two copies are completely decoupled and
form a helical quantum valley-Hall (QVH) liquid41,42 with a
perfectly quantized valley-Hall conductance of 2e2 =h. In
fact, this state of matter is an example of a topological
insulator (refer to details in Sec. VI B 5), which essentially
generalizes the quantum-Hall state for TR symmetric systems. Another such example, the quantum spin-Hall
state,4346 is examined in Secs. V C 3 and VI B 5.
3. Edge states

Strain-induced gauge fields result in a valley-dependent


~S , as prescribed by Eq. (67). Intermagnetic field, namely B
estingly, these effective magnetic fields couple to the pseudospin through the rz term and can create a finite pseudospin
polarization.28 To see how the gauge fields couple to the
pseudospin, we consider squaring the Hamiltonians in
Eqs. (60) and (61), which yield
S

~ 2
p  t1
ht2F rz BS ;
H2 t2F ~
F A  e

(68)

~ 2
H0 2 t2F ~
p t1
ht2F rz BS ;
F A  e

(69)

respectively. We see that a Zeeman-like term rz BS of the


same sign appears in both the K and K 0 valley effective
Hamiltonians, indicating a net pseudospin polarization. In
~ appears as a
contrast, a real external magnetic field B
z
Zeeman-like term r B of opposite sign in the two effective
Hamiltonians, implying zero net pseudospin polarization. Of
~ will result in a finite polarizacourse, a real magnetic field B
tion of the actual spins, though this polarization is quite negligible, because of the small Zeeman splitting in graphene.
~S field allows one
The pseudospin polarizing nature of the B
to characterize edge states in graphene strips.51 Finite-sized
graphene samples are terminated by edges, of which the two
main types are zig-zag and armchair edges, which correspond to the edges illustrated along the x^ and y^ axes in
Fig. 7(a), respectively.
For zig-zag edges, we can imagine cutting the sheet of
graphene along the perforated line shown in Fig. 7(a). Mathematically, we can represent this by setting the hopping integral perturbations dt1 t, while dt2 dt3 0 along y 0
(n.b., the gauge field description of deformed graphene is applicable only in the limit of small strain jdtj  t. Thus,
strictly speaking, the case of dt1 t lies beyond the gauge
field description of graphene edge states.) From Eq. (62), the
gauge field at the edge is
~zigzag t; 0:
A

(70)

~ @y Ax .
The effective magnetic field is calculated from B
Given that dt1 is actually an abrupt function of y, we obtain
~ < 0 for y < 0 (illustrated by ) and B
~ > 0 for y > 0 (illusB
trated by ). This implies that the graphene is pseudospin
(A; B) polarized on opposite sides of y 0 and that it supports edge states along the zig-zag edge.
For armchair edges, we cut along the dotted line x 0
shown in Fig. 7(b). This is modeled by setting dt1 0, while
dt2 dt3 t (it is convenient to consider the points
x 0 or x 0 ). The induced gauge field is then

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-12

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

FIG. 7. (Color online) Forming graphene edges by severing bonds as shown for (a) zig-zag and (b) armchair edges. This induces local gauge fields, whose curvatures represent effective magnetic fields capable of polarizing the pseudospin. Zig-zag edges are associated with a finite pseudospin polarization and, thus,
are able to support edge states. Armchair edges, on the other hand, do not polarize the pseudospin and, thus, cannot support edge states.

~armchair t; 0:
A

(71)

~ @y Ax vanishes
Apparently, the effective magnetic field B
for the armchair edge. Thus, we see that armchair edges
should not support edge states.
The above are consistent with prior theoretical studies
of graphene edges, which conclude that zigzag edges possess
localized states, while armchair edges do not.52,53 Further
details of the gauge field formalism to describe edge states
are found in Refs. 28 and 51.
V. SPIN-ORBIT COUPLING SYSTEMS: REAL SPACE
ANALYSIS
A. Spin-orbit coupling basics

The spin-orbit coupling (SOC) effect is ubiquitous in


semiconductor (SC) spintronics. The general Hamiltonian of
a free electron in the presence of SOC is given by54
H



~
p2
h2
V  2 2~
r  k~ rV ;
2m
4m c

(72)

where ~
p
hk~is the momentum, m is the electron mass, V is
the electrostatic potential, and c is the speed of light. The
spin-orbit Hamiltonian in Eq. (72) above is derived for
electrons in a vacuum upon reducing the relativistic Dirac
equation in the low energy limit.54 Qualitatively, the effect
can be understood from special relativity arguments; for an
~ rV
electron moving through a lattice, an electric field E
is Lorentz transformed to an effective magnetic field
~ in the rest frame of the electron.54,55 Since the SOC
k~ E
strength is inversely proportional to the relativistic energy
gap mc2  0:5 MeV, the effect in vacuum is highly suppressed. In SCs, however, the effect can be significantly
enhanced, as the energy gap can be of order 1 eV.54,56,57
Phenomenologically, Eq. (72) represents an electron in the
presence of a momentum-dependent effective magnetic
~~
field, B
k,
H

~
p2
~
~ k;
V  c~
r  B
2m

(73)

where c is the coupling strength. To avoid confusion


~ B~
~ r , we will denote
with an ordinary magnetic field B
~ throughout. For each k,
~ the
~ k
the spin-orbit field as B
spin degeneracy of electrons are split between two
eigenstates or subbands j6i with corresponding energy
eigenvalues of

6

h2 k~2
~
~ kj:
V6cjB
2m

(74)

In our notation, the ji state corresponds to the energetically


lower of the two subbands,  ; it corresponds to electrons
~ at each k.
~ The
~ k
with spins pointed in the direction of B
spin degeneracy of electrons and holes in generic semiconductors occurs due to the combined effect of spatial and time
reversal symmetry. In the absence of a real, external magnetic field, the time reversal (TR) symmetry of the SOC system is preserved. This implies that a momentum-dependent
~ B
~ On the
~ k
~ k.
internal magnetic field must satisfy B
~
~ k
other hand, spatial inversion symmetry implies B
~
~
~
~
Bk. Thus, when both symmetries are intact, Bk  0,
and the degeneracy is restored in Eq. (74). Clearly, the SOC
can be finite in systems and structures which break the spatial inversion symmetry.54 Two common instances are the
structural inversion asymmetry (SIA) in two-dimensional
electron gases (2DEGs) in SC heterostructures and bulk
inversion asymmetry (BIA) in certain crystal structures. Spatial inversion asymmetries can also arise under mechanical
strain applied to crystals. We describe these in more detail
below.
At the interface of a semiconductor heterostructure, carriers are confined in asymmetric quantum wells (they are triangular to first order) along the growth direction (^z, by
convention). This results in a two-dimensional electron gas
(2DEG) in the xy-plane of the interface that is subject to a
~ E^z. The SIA of the system
normal-to-plane electric field E
gives rise to the so-called Rashba SOC, governed by the
Hamiltonian58,59


(75)
H R a ky r x  kx r y ;
where a defines the strength of the effective Rashba field.
Apart from the built-in electric field, one can apply an external electric field to the 2DEG via conventional electrostatic
gating structures to significantly modify the strength of the
Rashba SOC. It has been experimentally demonstrated, for
example, that a in common semiconductor heterostructures
can be modulated by up to 50%.60 This provides a viable avenue for performing tunable spintronic functions through
purely electronic means. On the other hand, BIA is present
in crystals which lack a center of inversion, such as zinc
blende and hexagonal structures. The resulting spin-orbit
coupling is of the Dresselhaus type,61 described by the
k-cubic Hamiltonian

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-13

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)



HD g kx ky2  kz2 rx ky kz2  kx2 ry kz kx2  ky2 rz ;
(76)
where g is the coupling strength. Quite often, we are interested in low dimensional systems, such as in a 2DEG, in
which Eq. (76) collapses into a k-linear form,57,62


(77)
HD b kx rx  ky ry ;
where b  gp=dz 2 and dz is the width of the quantum well
along ^
z. Typically, b is around one order of magnitude
smaller than a in actual semiconductor samples. Often, however, we are interested in the scenario when b ! a. This
condition can be approached experimentally in extremely
narrow quantum wells, dz ! 0, since b / dz2 .
Less commonly studied is the strain-induced SOC.6365
Strain in zinc blende semiconductors, such as GaAs and
InSb, is represented by a symmetric tensor ij ji in the
Hamiltonian
C3
xy ky  xz kz rx zy kz  xy kx ry
2
zx kx  yz ky rz ;
(78)

HSt B  Tr

P
where Tr i ii and the constant Ch3 8  105 m/s in
GaAs. When we identify the off-diagonal components with
an electric field in the remaining direction, ij ! Ek , the
strain Hamiltonian is completely analogous to the generic
spin-orbit Hamiltonian in Eq. (72), with an additional contribution from the B-term.65
B. Non-Abelian gauge field representation

The SOC can be formulated within a real-space gauge


field framework by transforming the general Hamiltonian in
Eq. (72) to read6671
H

1 
e 2
~
p A ;
2m
c

(79)

h
rV  ~
r
4mce

(80)

where
A~
r

is a real-space gauge field, and we note that the transformation is not exact, but an approximation (second order terms
are neglected). The spin-orbit gauge field, A~
r above, is
another example of a gauge field which is not associated
with adiabatic transport of quantum states (see Sec. III B): it
appears merely as a result of rearranging the Hamiltonian.
For example, in the case of Rashba SOC, the gauge field has
the explicit form
r
AR ~

am
ry ; rx ; 0:
eh

(81)

Being a non-Abelian gauge field, the curvature of A~


r is
prescribed by the Yang-Mills curvature defined in Eq. (31).
Evaluating the curvature in Eq. (31), we find that only the
z-component is non-zero and equal to
^

XRz ~
r 

ie h R R i 2a2 m2 z
Ay ; Ax
r:
h
eh3

(82)

~
An analogous treatment can be applied to the k-linear
Dresselhaus SOC in 2DEGs, and we find that the curvature is given by
r 
XD
z ~

2b2 m2 z
r:
eh3

(83)

Interestingly, Eqs. (82) and (83) above represent vertical


magnetic fields which have different signs for spins polarized along ^z and ^z. This allows us to manipulate electrons
in a spin-dependent way. Some of the physical consequences
are described below.
C. Physical consequences
1. Aharonov-Casher phase

The Hamiltonian in Eq. (79) first gained popularity as


early as the 1980s in describing the spin-dependent version
of the Aharonov-Bohm effect [defined in Eq. (8)], namely
the Aharonov-Casher phase given by6670

/AC d~
r  A~
r:
(84)
In contrast to the AB effect, the gauge field in the AC effect
does not break TR symmetry,69 occurring in the absence of
magnetic fields. We briefly discuss the relationship between
the AC phase and the Berry phase (see, for example, Refs.
72 and 73). Firstly, we note that the AC phase is simply the
TR symmetric AB phase in Eq. (8). The well-known AB
phase comes from computing the flux threaded by a closed
~ It is inherloop (in ~
r-space) in a region of space with finite A.
ently a non-adiabatic phase factor with respect to quantum
state evolution. Such generalizations of the adiabatic Berry
phase to any (especially non-adiabatic) cyclic evolution of
quantum states is known as the Aharonov-Anandan (AA)
phase.74 Indeed, in the limit of slow cyclic evolutions, the
AA phase approaches the Berry phase.
Hatano et al.73 proposed a method to achieve perfect
spin filtering by utilizing the interplay between the AC phase
due to Rashba SOC in a semiconductor 2DEG and the AB
phase due to an externally applied magnetic field. According
to Eq. (82), spin-up and spin-down electrons experience
equal and opposite effective vertical magnetic fields. Therefore, they acquire equal but opposite AC phases in Eq. (84).
Hatano conceived a spatial circuit such that up (down) spin
electrons acquired an AC phase of p=2 (p=2). Furthermore,
a finite magnetic vector potential was assumed to be present
in the interior of the circuit, such that both spin-up and spindown electrons acquire an AB phase of p=2. In this scenario,
spin-up electrons acquire a total phase of expip 1,
which is completely destructive, while spin-down electrons
acquire a total phase factor of unity which is completely constructive. This results in a perfect spin filter whose output
only consists of spin-down electrons.73 The polarity of the
filter can be switched (such that the output consists of spinup electrons) by reversing the direction of the applied magnetic field.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-14

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

FIG. 8. (Color online) Transverse spin separation mediated by spin-orbit


coupling. Up and down spins experience opposing Lorentz forces, resulting
in a spin-Hall effect.

yt y0 

2. Spin-dependent transverse force

Being a gauge field in real-space, A~


r is expected to
result in spin-dependent transport due to the Lorentz force
defined in Eq. (47). This force was studied in detail by
Shen,71 who referred to it as the spin transverse force. It is
referred to as a spin force because it is spin-dependent,
and transverse refers to the fact that the response is perpendicular to an applied longitudinal spin current. In a 2DEG
with Rashba SOC in Eq. (75), a spin current jzx (read: drift velocity tx along x^ and spin polarized rz along ^z) experiences a
transverse force along the y^-direction in Eq. (47),
Fy 

2a2 m2 tx z
r:
h3

ent of the magnetic field strength. We also note that the spin
along the ^z-direction, with operator rz , is not a good quantum
number of the Rashba SOC system and that the spins will
undergo time evolution in the form of spin precession about
the effective Rashba field. This implies that both rz and Fy
in Eq. (85) are generally time-dependent. As ~
p is a good
quantum number, we put ~
p px x^ without loss of generality
and, for simplicity, assume an initial vertical spin-up state.
Then, the ^z-component of the spin as a function of time is
given by rz t cosxc t, where the Larmor frequency is
xc 2akx . Semiclassically, we then obtain the y^-position of
carriers by double integration of the acceleration obtained
from Eq. (85),

(85)

Clearly, Eq. (85) indicates that spin-up and spin-down polarized currents experience equal and opposite transverse
forces. This is illustrated in Fig. 8. One important note is that
this force is proportional to a2 , where a governs the strength
of the Rashba effective field. This is in stark contrast with
the force due to Berry curvature, which is always independ-

1
cosxc t;
2kx

(86)

where y0 is the initial y^-position. Eq. (86) above describes an


oscillatory motion of electrons due to SOC. The quantum
mechanical jitter motion of carriers is known in the literature
as zitterbewegung. The above shows that the zitterbewegung
of conduction electrons in Rashba SOC systems arises from
the coupling of the spin dynamics and the orbital motion.
This was further illuminated in Refs. 75 and 76, though these
works did not make explicit use of gauge fields.
The force expression in Eq. (85) indicates a Hall separation of spin carriers, shown in Fig. 8. This type of transverse
spin separation may be related to the spin-Hall effect (SHE).77
In Ref. 12, Tan et al. studied the competition of this
force in Eq. (85) with the topological force driven by a spatially varying magnetic field (discussed in Sec. III C 1). The
chosen magnetic field configuration (illustrated in Fig. 9)
was almost vertical (h  0), but with a small spatial
non-uniformity characterized by a crown-shaped distribution (with net chirality characterized by a solid angle X).
The real-space Berry curvature was calculated to be

FIG. 9. (Color online) (a) Illustration of proposed device, in which a transverse separation of spins occurs in response to a longitudinal charge current J. The
~SO (vertiseparation occurs heuristically as a result of spin-dependent forces due to (i) Rashba SOC, which is characterized by the perpendicular electric field, E
~ r. The directions of the spin-dependent force arising from the Rashba SOC, F
~SO , and from
cal, dark arrow), and (ii) a spatially nonuniform magnetic field, B~
~Berry , are indicated by arrows. We note that the forces from the two contributions act in opposite directions. The degree of cancellation
the Berry curvature, F
between the two forces can be modulated via a gate bias. This leads to the potential modulation of the transverse spin-current by purely electric means. (b) The
configuration of the spatially nonuniform magnetic field characterized by chirality h.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-15

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

h
z
Xz ~
r 2eR
2 r (where R is defined in Fig. 9(b)), which, for a
drift velocity tx along x^, produces a spin-dependent force
along y^ of

htx z
r:
Fy
2eR2

3. Quantum spin-Hall effect

From Eq. (82), a strong curvature equivalent, i.e., one


which gives rise to a magnetic field strength typically associated with quantum-Hall behavior (> 5 T), corresponds to a
Rashba parameter of
a2 > 5 

eh3
:
2m2

(88)

In semiconductors, the curvature of the conduction band


defines the effective mass m, which typically is of the order
m 0:1m0 , where m0 is the mass in vacuum. This yields
the requirement a > 4:3  1011 eVm, which lies in the
typical range of 1012 1011 eVm. Since the curvature in
Eq. (82) is spin-dependent, it seems to suggest that a quantum spin-Hall type behavior will be induced in the presence
of a sufficiently large a. Such a situation would be equivalent to two copies of the QHEchiral spin-up electrons and
antichiral spin-down electronsforming a helical quantum
spin-Hall (QSHE) liquid with counter propagating spinresolved edge states.45,78 However, the non-Abelian gauge
in the Rashba SOC system in Eq. (81) is non-diagonal and
does not couple directly to the momentum ~
p. Thus, the
effect of the Yang-Mills curvature is essentially different
from that of a real physical magnetic field and does not
give rise to any Landau levels. Since the Rashba SOC
system is a gapless system (assuming the Rashba SOC
coupling has no space dependence), the QSHE effect does
not arise, even in the presence of a large a, contrary to the
prediction in Ref. 78.
By contrast, Bernevig et al.45 proposed an alternative
SOC-induced QSHE, in which the two TR copies of the
QHE are completely decoupled. Here, the authors relied on
strain-induced SOC. In particular, they considered a strain
configuration xy 0, xz gy, and yz gx, where g quantifies a strain gradient (refer to Eq. (78)). In the presence of an
additional parabolic quantum well in the xy-plane, the effective Hamiltonian reads
H

~
p 2 C3 g

ykx  xky rz Dx2 y2 ;


2m
2

(89)

where D is the strength of the quantum confinement. When


D C23 g2 m=8
h2 , Eq. (89) can be written as a perfect square,

(90)

C3 gm z
r y; x; 0
2eh

(91)

where

(87)

~SO due to Rashba SOC in Eq. (85), on the other


The force F
hand, is corrected by a small factor of cos h. More importantly, the Rashba SOCinduced force can be varied in magnitude dynamically through the gate-tunable a parameter,60
while the chirality-driven force remains constant. This indicates that the degree of cancellation of the two forces could
be tuned in a carefully designed 2DEG-based device, resulting in a tunable source of spin-polarized current.12

1
~
p eA2 ;
2m

is a spin-dependent gauge field induced by the strain. Note


that, although its components are matrices, A is an Abelian
gauge field. Thus, its curvature is simply given by
X r  A and reads
r 
Xz ~

C3 gm z
r:
eh

(92)

We can anticipate the possibility of the QSHE provided that the


factor C3 gm is sufficiently large. In this scenario, the spin-up
and spin-down states are good quantum numbers (since they are
eigenstates of the Hamiltonian, Eq. (89)) and the two spinresolved copies of the QHE are completely decoupled. The chiral spin-up states contribute a charge Hall conductance quantized
in units of e2 =h, while the antichiral spin-down states contribute
a conductance quantized in e2 =h. Thus, the net charge Hall
conductance rc r" r# vanishes, but the spin-Hall conductance rs r"  r# is finite and quantized in units of 2e2 =h. The
exact quantization of the spin-Hall conductance in the QSHE is
only possible when sz is a good quantum number. However, the
QSH state remains robust, regardless of sz conservation, as is
guaranteed by the topological classification of the state.43
The QSHE is an example of a more general class of
solid state systems known as topological insulators, which
also includes the QHE. However, unlike the QHE, the QSHE
is time-reversal (TR) invariant. We shall return to the topic
~
of topological insulators in our discussions of the k-space
Berry curvature in Sec. VI B 5.
4. Spin torque

In Sec. III C 3, the real-space Berry gauge potential due


to a chiral magnetic system was shown to impart a spin tor~top .
que through an additional (topological) switching field, H
The fact that the SOC also induces a real-space gauge field
suggests that it may be used to provide another switching
field.7981 We consider again a ferromagnetic system with
~ r, but this time in the presnonuniform magnetization M~
ence of Rashba SOC; this could be realized in a 2DEG doped
with ferromagnetic dopants, namely a dilute magnetic semiconductor (DMS). The Hamiltonian of the system reads
H

1 
e 2
~ r ;
~
p AR JH ~
r  M~
2m
c

(93)

where JH is the exchange constant, and we have formulated


the Rashba SOC in the non-Abelian gauge field representation in Eq. (81). Performing a rotation of the local coordinates (recall Sec. III B 3), H becomes diagonalized at each
point in space (see Eq. (29)),


2
1
e
h
~ rj:
~
p UAR U  i UrU
JH rz jM~
H0
2m
c
e
(94)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-16

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

The square brackets contain two gauge fields, the first being
the SOC-induced non-Abelian gauge potential and the second being the transformation-induced potential. The latter,
~top
of course, gives rise to the topological switching field H
shown in Eq. (53). We are presently concerned with deriving
the switching field due to the former SOC-induced term.
Expanding this term, we obtain


~;
nE
(95)
UAR U arz ~
which is Abelian. In the limit of strong exchange coupling
JH ! 1, we assume a fully spin-polarized gas and apply the
adiabatic approximation. The corresponding interaction
~
energy between an electronic

 currentj and
 the local magnetj  UAR U j " , and the switching
ization is then int " j~
field defined by Eq. (52) is79
~so aji @nj Ek ijk :
H
l0 M @~
n

(96)

Let us consider the simple case of a current in the x^-direction,


~
j jx x^. In the case of Rashba SOC, only the ^z-component of
the electric field is non-zero and the switching field components are
Hso;x 0; Hso;y

jx Ez
; Hso;z 0:

l0 M

(97)

In a typical experimental setting with parameter values


jx 1011 Am2 , a 1012 eVm, M 5  103 Am1 , the
switching field is estimated to be Hso;y 6:5  103 Am1 .
This may be sufficiently large to affect a spin torque in ferromagnetic quantum wells with Rashba SOC. Indeed, such
results have already been demonstrated experimentally,82,83
which are consistent with the predictions in Eq. (97).83
D. Spatially nonuniform spin-orbit coupling

In the majority of cases, we study systems with spatially


homogeneous SOC. In general, the SOC will exhibit some
spatial variations. For concreteness, let us consider spatially
nonuniform Rashba SOC. The non-Abelian gauge field in
this case is
a~
rm
r
ry ; rx ; 0;
A ~
eh
R

(98)

where a~
r characterizes the spatially varying SOC, cf. Eq.
(81). The curvature, in this case, contains additional contributions from the curl, together with the non-Abelian term
(see Eq. (82)),


m @a x @a y
2a2 m2 z
R
r r
r
r:
(99)
Xz ~
eh @x
@y
eh3
This could affect various aspects of carrier transport, including
the zitterbewegung. Spatial discontinuity of SOC in multilayered structures can also produce interesting effects. Consider a
trilayer structure with a 2DEG channel, as illustrated in Fig.
10(a), in which the 2DEG exhibits Rashba SOC and the two
contact regions do not. In the simplest case, the spatial profile

FIG. 10. (Color online) (a) Trilayer structure in which the 2DEG channel
has Rashba spin-orbit coupling, while the two contact regions do not. (b)
The spatial discontinuity of the Rashba spin-orbit coupling induces spatially
narrow effective magnetic fields at the interfaces, which are spin-dependent,
rx 61.

of the Rashba SOC strength can be taken as a rectangular


function, ax a0 Hx  Hx  L, where Hx is the
unit-step function. The curvature can then be computed as
r
XRz ~

a0 m
dx  dx  Lrx
eh
2a2 m2
0 3 Hx  Hx  Lrz ;
eh

(100)

where dx is the delta-Dirac function. The first term is due


to the spatial discontinuity of the Rashba SOC in the trilayer
structure and physically represents the presence of narrow
magnetic field barriers centered at the interfaces. These interfacial fields are spin-dependent, rx 61 (refer to
Fig. 10(b)). Interestingly, many authors have studied the
electronic transport through similar systems, but where the
magnetic d-barriers are formed by actual magnets8487 and
are, therefore, spin-independent. However, controlling magnetic fields at the nanoscale is experimentally challenging.
The present system offers a more feasible platform to study
the transport properties of magnetic d-barriers; preliminary
calculations have been performed for a spin filter in Ref. 88.
VI. SPIN-ORBIT COUPLING SYSTEMS: ~
k -SPACE
ANALYSIS

In Sec. III, we examined spatially varying magnetic


systems and showed that, in the limit of adiabatic transport,
in which the spins remain aligned to the local magnetic
texture, a real-space Berry curvature arises in Eq. (40). In
this section, we examine a topic of significant current interest
~
in condensed matter, namely, the momentum-space or kspace Berry curvature. Not surprisingly, this is present in
~
systems with k-dependent
magnetic fields, i.e., systems with
finite spin-orbit coupling, such as Eq. (73).
A. Derivation of ~
k -space Berry curvature

The general spin-orbit Hamiltonian is given in Eq. (73).


~
To derive the k-space
Berry curvature, we shall use the
method of unitary transformations described in Sec. III B 3.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-17

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

~
Following that method, we proceed to lock the k-dependence by applying a local transformation to the system, which
~ This is facili~ k.
rotates the reference spin axis to lie along B
~
tated by a unitary matrix U Uk, given by Eq. (27),
~
where, in this scenario, the h; / are the k-dependent
spheri~ The trans~ k.
cal angles parameterizing the direction of B
formed Hamiltonian reads
H0 UHU

h2 k~2
~ UV~
~ kj
 crz jB
r U :
2m

(101)

The last potential energy term transforms nontrivially because


~
~
~
derivatives and U depends on k.
r i@k represents k-space
Considering the potential energy due to an applied constant
~ we have V~
~ ~
electric field, E,
r eE
r and20,89
~  irk iUrk U :
UVirk U eE

(102)

Thus, the local rotation of coordinates has transformed the


~ where
definition of the spatial operators ~
r !~
r  Ak

~
~
Ak  iUrk U is a 2  2 gauge field in k-space. In the
adiabatic limit, we are only interested in the diagonal compo~ which represent the two subbands of the SOC.
nents of Ak,
~
Following Eq. (32), we can write down the k-space
Berry
gauge fields as
1
~
~
~
Aad:
6 k 6 1  cos hkrk /k;
2

(103)

where the 6 indexes the two bands. Following the analogous


situation in real ~
r-space in Eq. (40), the Berry curvature in
~
k-space reads
!
~ @~
~
1
@~
n

k
n

k
~ 6 ~
~
Xk k
;
(104)

nk
2
@ki
@kj
~ Bj
~ is the unit vector pointing along the direcwhere ~
n B=j
~ As we shall see below, the
~ k.
tion of the spin-orbit field, B
~ represents an effective magnetic field in
curvature Xk
~
k-space, which can influence the orbital motion of carriers.
B. Physical consequences

In Sec. III C 1, we showed that the ~


r -space Berry curvature leads to Lorentz forces that are spin eigenstate-dependent,
~
as in Eq. (47). Here, we describe the physical effect of the kspace curvature. Without proof, we state the result derived by
Sundaram and Niu,90 which gives the semiclassical equations
~
of motion of a wave packet in the presence of Xk,
_
~ e~
hk~ eE
r_  X~
r;

(105)

1 @s ~_
~
~
r_
 k  Xk:
h @ ~
k

(106)

The first equation, Eq. (105), describes the force acting on car~
riers, which includes a contribution due to an electric field, E,
and the ~
r-space Berry curvature, X~
r (and magnetic field,
~ r , if present). The second equation, Eq. (106), describes
B~
the velocity of wave packets, which includes the usual group
velocity and also a contribution from X~
k. This contribution

is often called the anomalous velocity. Evidently, the anomalous velocity in Eq. (106) is of analogous form to the Lorentz
force in Eq. (105): the two equations are invariant upon inter~ Drawchanging the role of the position, ~
r , and momentum, k.
~
ing from this observation, we remark that Xk essentially
~
gives rise to a Lorentz force in k-space.
Historically, Eq.
(106) was derived by Karplus and Luttinger in 1954 (Ref. 91)
when they tried to explain intrinsic contributions to the anom~ arises from the
alous Hall effect in ferromagnets. Here, Xk
Bloch wavefunctions in crystal momentum space. The connection with topology, however, was made only recently.90
Equation (106) describes topological transport in a vast
array of condensed matter systems with a finite momentum
space Berry curvature, which are discussed below.
1. Spin-Hall effect

Earlier, in Fig. 8, we illustrated the idea of a Hall-like


separation of carriers based on their spin. Let us formalize
this below. We consider the situation where an electric field
Ei drives a separation of spins sk along the transverse Hall
^ In the spin-Hall effect (SHE), i;
^ j;
^ k^ are mutually
direction, j.
orthogonal coordinates. Thus, the spin-dependent force
expression in Eq. (85) provides an example of a SHE. The
strength of the SHE is characterized by the spin-Hall conductivity (SHC), which is defined as the ratio of the transverse
spin current and the driving electric field, namely,
rsH

jks;j
Ei

(107)

where the spin current js is calculated as jks;j h4hfrk ; tj gi. In


Sec. VI C, we shall derive the (spin) Hall conductivity for
generic SOC systems and show that it is intimately con~
nected to the k-space
Berry curvature. We focus here on the
intrinsic SHE, which arises from the properties of the band
~
structure (Berrys curvature in k-space),
as opposed to the
extrinsic SHE arising from spin anisotropic scattering due to
impurities (more details of the latter can be found, e.g., in
Refs. 77, 92, and 93). Although, it should be noted that a
clear distinction between intrinsic and extrinsic contributions
to the SHE has not yet been elucidated; certain extrinsic
SHE mechanisms have been found to exhibit characteristics
of intrinsic phenomena94,95 (independent of scattering rate,
impurity density, etc.) and are even related to the
momentum-space Berry curvature.96
By now, the intrinsic SHE has been quantified in a host
of systems, starting with p-doped semiconductors.89 This
system is modeled by the Luttinger Hamiltonian, which
describes the spin-orbit splitting in the doubly degenerate valence bands of common semiconductors,97
HLutt:



~
5
k2
~ 2 V~
c1 c  ck~ S

r;
2
2

(108)

where c1 and c are valence-band parameters defining the


effective hole masses and S~ is the vector of spin-32 matrices.
The holes described by Eq. (108) have a well-defined chiral~ which takes on values k 61=2 and
~ hjkj,
ity, k k~ S=
k 63=2. Because of the k-squared term in the Hamiltonian,

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-18

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

states with opposite signs of the chirality are degenerate, corresponding to the two-fold degenerate light-hole (LH,
jkj 1=2) and heavy-hole (HH, jkj 3=2) bands. Parameterizing the momentum vector as k~ j~
kjsin h cos /; sin h sin /;
cos h, we proceed to transform the Hamiltonian
H0 UHU


~
~ z,
with a 4  4 unitary matrix U, satisfying U k  S~ U jkjS
3
which is diagonal in the spin-2 space. The resulting 4  4
~ iUrk U , which, in the adiabatic
gauge field is Ak
limit, reduces to two 2  2 copies corresponding to LH and
HH states. Due to the degeneracy of the LH and HH bands,
the Berry gauge fields here are non-Abelian (their components
are non-commuting). However, we may make them Abelian
by removing their off-diagonal components, upon which the
exact form of the Berry gauge fields are reminiscent of Eq.
(103), but upon replacing the factor 12 ! jkj. The corresponding Berry curvature reads89
~
~ k k ;
Xk
(109)
j~
kj3
~
i.e., it is a Dirac monopole in k-space
with strength equal to
the hole helicity, eg k. Substituting the expression for the
curvature in Eq. (109) into the equation of motion, Eq. (106),
the anomalous velocity component is given by
k~
_
ta kk~ 3 ;
j~
kj

(110)

~
which is perpendicular to both the applied electric field E
_ ~
~
~
(since k / E) and the momentum vector k. Since the chirality of the holes has sign k > 0 < 0 for hole spins (anti-)
parallel to ~
k, the KL velocity is perpendicular to the spin S~
and points along opposite directions, depending on the sign
of the chirality. This transverse separation of the spins gives
rise to the SHE of holes in the Luttinger system.
A similar effect is predicted in n-doped zinc blende semiconductors with k3 -Dresselhaus SOC.98 From the Hamiltonian
in Eq. (76), we can easily compute the Berry curvature,
~ 6ijk
Xk k

kx2  ky2 kx2  kz2 ky2  kz2


kk ;
~ 3
~D kj
2jB

(111)

~ is the
~D k
where 6 represents the two SOC subbands and B
~D . The SHE
rB
effective Dresselhaus field, i.e., HD g~
results from the Berry curvature for electrons that are collimated, namely, traveling in a highly unidirectional manner.
Methods to achieved electron collimation are essentially analogous to beam collimation in optics, whereby the principle of
total internal reflection is employed, e.g., in optical fibers. In
electronic collimation, the role of the refractive index is played
by the local carrier density; for details, see Refs. 98100.
Assuming electron collimation along the ^z-axis, we have
jkx j; jky j  jkz j, and the Berry curvature above is reduced to
~ 6ijk
Xk k

kz4 kx2

 ky2
k:
~ 3 k

~D kj
2jB

values hrx i 6kx =kk and hry i ky =kk ,101 where


q
kk kx2 ky2 . Assuming an applied electric field along ^z,
the anomalous velocity from Eq. (106) has the form
~ y;
ta;x k_z Xy / Ez f ks

(113)

~ x;
ta;y k_z Xx / Ez f ks

(114)

kz4 kx2 ky2


~ 3
~D kj
2jB

~
where f k
kk , which indicates a spin-Hall effect in
both x^ and y^ directions.
Analogous Berry phase effects are predicted to be important for explaining the intrinsic SHE in metals, such as
platinum.102 Metals may be favorable as a platform for utilizing SHE over semiconductors due to several reasons,
including larger spin-Hall conductivity102,103 and the fact
that FM contacts to harness the generated spin current do not
suffer from poor injection difficulties.104

2. Spin-Hall effect of light and optical Magnus effect

An analogous situation occurs in optics. The effect is


called the spin-Hall effect of light or the Magnus effect of
photons. Here, the role of the spin is replaced by the circular
polarization of the light. In the spin-Hall effect of light (optical Magnus effect), left and right circularly polarized light
are deflected along opposite directions normal to the propagation direction. It was first described in 1990,105 and the
connection with the Berry curvature was established in
2004.106109
The semiclassical equations of motion of a wave packet
become similar to those for electrons (Eqs. (105) and (106))
~
p _
~
r_ t~
r ~
p  hzjX~
pjzi;
p

(115)

~
p_ rt~
rp;

(116)

jzi
_ i~
p_  A~
pjzi;

(117)

where t~
r is the velocity of light and jzi t z ; z  denotes
p and A~
p are the Berry curthe polarization state.106,107 X~
vature and the gauge field, whose expressions will be
explained in the next paragraph. As can be seen from these
equations, the electric field in the semiclassical equations of
motion for electrons is replaced by the spatial variation of
the refractive index.
In order to derive the equations of motion, one method
is to use the variational theory.107 Here, we discuss another
derivation, using an effective Hamiltonian. In the geometric
optics (GO) approximation, light propagation in a spatially
inhomogeneous medium is modeled by the effective
Hamiltonian
H

(112)

To see how this leads to the SHE, we consider the limit of


strong collimation. Then, the electron spins have expectation


1 2
~
p  n2 ~
r  Q;
2

(118)

where ~
p is the momentum, n~
r is the spatially varying refractive index, and Qij pi pj is a 3  3 matrix in the spin-1
space. The GO model is known to be excellent when the
wavelength of the light is much smaller than length scales

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-19

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

of spatially inhomogeneities. The GO Hamiltonian may be


diagonalized, giving rise to a ~
p-space gauge field,
A~
p iUrp U . The diagonalized Q matrix reads
UQU diag0; 0; p2 , indicating a pair of degenerate transverse modes (left and right polarization) and a single longitudinal mode. Applying the adiabatic approximation here
corresponds to neglecting terms within A, which connect
the transverse and longitudinal modes, namely, the terms
A13 , A23 , A31 , and A32 . This results in a 2  2 gauge field
in the transverse mode subspace and a regular U1 vector
gauge field for the longitudinal mode. Focusing on
the degenerate transverse modes, the Berry gauge field in
~
p-space reads108
!
p
p
p
p
y
z
x
z
;
; 0 ry ;
Aad: ~
p
(119)
pp2x p2y
pp2x p2y
where p

q
p2x p2y p2z . It is well known that, for 2-fold

degenerate bands, Aad: is generally a U2 non-Abelian


gauge field (which generalizes the usual Abelian U1 Berry
field), known as the Wilczek-Zee gauge field.121 From Eq.
(119), however, we can see that for photons, the Berry gauge
field depends only on ry and is therefore Abelian. Furthermore, by applying a rotation V to the system where
Vry V rz , the gauge field for the transverse modes is diagonalized and reads
!
py pz
px pz
ad:
;
;0 ;
(120)
p 6
A6 ~
pp2x p2y
pp2x p2y
i.e., it is reduced to two independent scalar U1 potentials,
corresponding to left and right polarization. The corresponding Berry curvature is106,108,109
X~
p

~
p
;
p3

(121)

where  and corresponds to the right and left circular


polarizations, respectively. This Berry curvature has a Dirac
monopole in ~
p-space with strength eg 1. This gives rise
to the anomalous velocity X~
p  ~
p_.106,108 The anomalous
_
velocity X~
p  ~
p is perpendicular to ~
p and rn.106 For
example, at an interface between different media, the refractive index has spatial dependence and gives rise to the spinHall effect of light. At the interface, the shift is perpendicular
to the incident plane and the shift is opposite for the left- and
right-circularly polarized light. This shift at the interface
reflection and refraction is called the Imbert-Fedorov shift,
which was theoretically proposed by Fedorov110 in 1955 and
experimentally observed by Imbert111 in 1972 for perfect
reflection. For partial reflection, it was observed by Hosten
and Kwiat112 in 2008. We note here that, in a strict sense, the
semiclassical equations of motion, Eqs. (115), (116), and
(117), apply only when the spatial variation of the system is
slow; they cannot be applied at an interface where the refractive index changes abruptly. Indeed, there is a small deviation from what has been predicted by the Berry curvature
theory.108,112

This anomalous velocity also gives rise to the separation


of left and right circularly polarized light in the Magnus
effect.108,109 The Magnus effect was experimentally
observed in 2008.113 These optical measurements remain an
important demonstration, because, in the optical system, the
effect of the anomalous velocity can be directly verified (i.e.,
by measuring the spatial displacement of the photons via
photodetectors). Typically, the displacement is small, but it
is possible to enhance the effect by the designing of the band
structure of the photons. Theoretically, it can be done using a
photonic crystal, i.e., the periodic array of dielectrics, so that
the band structure has band crossings, near which the Berry
curvature becomes enhanced.106
We here note the role of the Berry curvature in the
theory of spin-Hall effect of light. In earlier days, the ImbertFedorov shift was discovered as an effect from classical
electromagnetic theory. By introducing the notion of Berry
curvature, one can reinterpret such a known effect as a member of the spin-Hall effect family. Furthermore, the Berry
curvature enables us to predict further nontrivial effects,
such as optical Magnus effect and enhanced Imbert-Fedorov
shift in photonic crystals.
3. Magnon-Hall effect in ferromagnetic insulators

Similar types of spin-Hall effect are predicted for other


types of waves, such as magnons in ferromagnetic insulators.
Magnons (spin waves) are low-energy excitations in magnets
and have a band structure. Therefore, the Berry curvature
may arise from this band structure. The effect of the Berry
curvature is intuitively described in the semiclassical equations of motion for magnons,114,115
1 @n~k ~_
~
~
r_
 k  Xn k;
h @ k~
_
hk~ rU~
r ;

(122)
(123)

where U~
r is a confinement potential for magnons and n~k
denotes the magnon dispersion for the nth band. These
equations are quite similar to those for electrons, and similar
phenomena are theoretically predicted. One difference, however, comes from the fact that the magnon has no charge
and, therefore, cannot be driven by electric field. Instead, the
magnon wavepacket can be driven by temperature gradients.
The effect of temperature gradients can be described also
from the semiclassical equations of motion, Eqs. (122) and
(123), using the picture of the edge current.
Near the edge of the sample, the confining potential
U~
r gives rise to the anomalous velocity (the second term
of Eq. (122)) along the edge. The total current along the
edge forms an edge current, which is the same for all the
edges. (We note that, in contrast to edge modes in quantum
Hall systems, this current is not confined at the edge of the
sample, but it mixes with a bulk magnon current.) Because it
is a circulating current, it has no net transport. Such an edge
current depends on the temperature through the Bose distribution function. Therefore, if there is a temperature gradient,
the edge current is not equal between the two sides of the
sample, giving rise to a net transverse current. As a result,

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-20

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

the thermal Hall effect (Righi-Leduc effect) results. The


thermal Hall conductivity is obtained as
jxy

2kB2 T X
c2 qn Xz ~
k;
hV ~

(124)

n;k

log 1q
q

2

where
c2 q 1 q
log q 2Li2 q,
q qnk~ is the Bose distribution function and Li2 z is the
polylogarithm function.
Apart from these results based on semiclassical theory
of the Berry curvature, there have been some works on the
thermal Hall effect.116118 The formula presented in Refs.
116 and 117 by the linear response theory is different from
Eq. (124). This difference is attributed to the missing terms
in the linear response theory in Refs. 116 and 117, and the
modified linear response theory developed in Refs. 114 and
115 gives an identical result with Eq. (124).
The thermal Hall effect of magnons has been measured for
the pyrochlore ferromagnet Lu2 V2 O7 ,117 and the result agrees
well with that estimated from Eq. (124).114 In addition, because
the Berry curvature is associated with the wave nature, this
theory is also applicable to classical magnetostatic spin-wave
systems, such as yttrium-iron garnet (YIG). In YIG film, the
demagnetizing field gives rise to an anisotropy in the spinwave spectrum, giving rise to the coupling between the spin
and the orbital motion. This induces a nonzero Berry curvature,
resulting in thermal Hall effect and edge current of magnons.
The spin-wave transport in YIG has been intensively
studied in experiments,119,120 because of its long coherence
length and its potential application to spintronics. The wave
packet of spin wave can be observed optically, and the coherence length can be as long as centimeters. Therefore, YIG
is expected to offer a good stage for observing dynamics of
wave packets arising from the ~
k-space Berry curvature.
4. Valley-Hall effect in graphene

~
The Berry curvature in k-space
also arises in gra122,123
We consider monolayer graphene with a sublatphene.
tice asymmetry, namely, that the A and B sublattice sites are
energetically different. This can be represented by adding a
Zeeman-like term in the pseudospin space,
~ p Urz ;
r  B~
H tF ~

(125)

~ p px ; npy ; 0 is the effective pseudospin field in


where B~
Eqs. (58) and (59), n 61 for the valley K (K 0 ), and U
(U) is the electrostatic energy on sublattice A (B). It is
straightforward to derive the Berry curvature as
k
X z ~

nU
3

(126)

2p2x p2y U 2 2
The valley-dependence of Eq. (126) means that electrons
residing in opposite valleys will become separated upon
application of an electric field, leading to the valley-Hall
effect. This effect was discussed in Ref. 122. In particular, a
finite Hall voltage is predicted if there is an imbalance of
electron distribution in the two valleys; such a situation

might be achieved through the use of a valley filter, e.g., as


described in Sec. IV C 1.
5. Topological insulators

The Berry curvature structure is also present in topological insulators (TI). Let us begin with a short introduction to
TIs (otherwise, the reader may skip directly to the Hamiltonian in Eq. (127)); the interested reader is directed to the
comprehensive review article by Hasan and Kane.124 Currently, TIs represent a field attracting immense research interest due to the highly interesting physics, mathematics, and
potential device applications. The first example of a TI was
found in the quantum-Hall effect (QHE), which was discovered experimentally in 1980.39 Just like an ordinary insulator, the bulk quantum Hall state is gapped (the valence and
conduction bands are separated in energy) and nonconducting. But unlike a normal insulator, the quantum Hall
state supports electronic transport at the edges of the sample,
which have remarkable characteristics: they are chiral (they
traverse along a definite direction along each edge), gapless
(there are edge states at any Fermi energy lying within the
bulk gap), and are extremely robust to the presence of disorder or the geometry of the edge. The quantum Hall conductivity (QHC) can be expressed in terms of a Chern integer n,
a topological index which remains invariant to continuous
deformations of the system125 (this is detailed in Sec. VI C,
below). This gives the QHE state a topologically non-trivial
classification (in conventional band insulators, the index n
vanishes). Up until recently, it was thought that all topologically non-trivial insulators were inherently TR-breaking.
It was only until after the discovery of graphene that the
TR-symmetric topologically nontrivial insulating phase was
discovered. In 2006, Kane and Mele43,44 proposed that, when
spin-orbit coupling (SOC) was considered in graphene, the
electrons behaved as though two time-reversed copies of the
quantum Hall effect were placed one on top of the other. In
this case, the edge states exist in pairs; electrons of opposite
spins ("; #) traverse along opposite directions along each edge
(they are helical), they are gapless, and they are robust to backscattering by non-magnetic impurities (for backscattering is
only permitted when a spin is flipped). As mentioned earlier, a
similar theoretical proposal has been made by Bernevig and
Zhang in strained semiconductors.45 This two-dimensional
effect is known as the quantum spin-Hall effect (QSHE). The
2D QSHE was proposed to occur in the HgTe quantum well
system,46 which has been verified by transport measurement.47,48 In the HgTe quantum well, which is supposed to be
the quantum spin-Hall system, the measured conductance is
2e2 =h, which is a direct evidence that there are two perfectly
conducting channels corresponding to the edge states.
The 2D QSHE was generalized to 3D systems by Fu and
Kane.126 In 3D topological insulators (TIs), spin-filtered
surface states arise as opposed to edge states. These have no
quantum charge Hall effect analogs. At low energies, the
surface states form 2D Dirac cones, i.e., they are massless
Dirac fermions governed by the graphene-like Hamiltonian
r~
p;
H tF ~

(127)

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-21

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

where tF 105 ms1 and only a single valley exists, as


opposed to graphenes two valleys, c.f. Eqs. (58) and (59).
Also, in contrast to graphene, ~
r in Eq. (127) refers to the
actual electron spin rather than the pseudospin. This dispersion was observed in various systems, such as Bi2 Se3 ,50 as
has been proposed by first-principle calculations.49
In TIs, a finite Berry curvature can result if only a gap
is introduced in the Dirac spectrum. Lu et al.127 considered
ultra thin topological insulator films and found that tunneling
processes between the top and bottom surfaces open up a
gap in the energy spectrum (the energy gap is proportional to
1=L2 , where L is the film thickness). The corresponding
~
k-space Berry curvature was shown to result in a finite spinHall effect at the TI surface, which will lead to an observed
spin accumulation at the edges.89 Alternatively, the ordinary
TI surface states in the presence of an exchange field due to
a deposited FM film also gaps the spectrum, and we expect a
finite Berry curvature. Moreover, in this scheme, the Berry
curvature should carry an explicit dependence on the gap
size, which is proportional to the FM magnetization; this follows from an analogous situation in graphene when one
breaks sublattice asymmetry, giving rise to a Berry curvature
and the valley Hall effect.122 Having a tunable Berry curvature is desirable from the point of view of tunable electronic
(topological) transport, most notably, the Hall effect.



~ s ~
~ t k
~
tr
J

kP
kJ

kP
X
X
x
y
1
~

nt  ns k;
~ t k
~
A s;t6 ~
im  s k
k


(133)

~  nF s k
~ , the summawhere nt  ns ~
k nF t k
tion over n in the last line is calculated by the Matsubara
~ is the Fermi distribution function of subsum, and nF s k
band s. We now do the conversion from Matsubara Greens
function to the retarded Greens function in order to obtain
the Hall conductivity,
rxy lim

x;g!0 x

Qxy im x ig;

(134)



~ ~ ~ ~
i X X tr Jx kPs kJy kPt k

nt  ns ~
k;
~   s ~
A s;t6 ~
t k
k2
k

(135)


~
iX
f k
~
n  n k;
~
~ 2 
A ~  k   k

(136)

~   k
~ 2
where the denominator is evaluated to be  k
2 ~~ 2
4c jBkj from Eq. (74) and


~ kJ
~ y kP
~  ~
~  kJ
~ y kP
~ k
~ :
kJx kP
f ~
k tr Jx kP

C. Hall conductivity

(137)

We derive the Hall conductivity for the general spinorbit Hamiltonian in Eq. (73) using the Kubo method and
show that it is intimately connected to the ~
k-space Berry curvature. By definition, the single-particle retarded Greens
function is given by
~ x ixI  H1 :
Gk;

(128)

Inserting the Hamiltonian in Eq. (73), we get


1
1
P
P ;
(129)
ix  
ix  


~ ~
nk
r and
where 6 are as defined in Eq. (74), P6 12 I6~
~ Bj.
~ The Hall conductivity via the Kubo formula is
~
n B=j
given by128
~ x
Gk;

~ can be derived from the HamiltoThe current operator Ji k


nian in Eq. (73),
~
Ji k

~ 0 k
~
~
~a k
@Hk
@B

c
ra ;
@ki
@ki
@ki

(138)

where 0 is the kinetic energy. Substituting this into the


expression for f ~
k in Eq. (137), we obtain some algebra,


~ 2 nb @kx na @ky nc acb :
~ 2ic2 jB
~ kj
(139)
f k
Thus, rxy in Eq. (136) reads
rxy

1 X @na @nc
~
nb
acb n  n k:
@kx @ky
2A ~

(140)

rxy lim

x!0 x

Qxy x ig;

(130)


1 X  ~ ~
~ k;
~ ixn ;
tr Jx kGk; ixn m Jy kG
Qxy im
Ab ~
k;n

(131)
where A is the system area and xn and m are the fermionic
and bosonic Matsubara frequencies, respectively. Substituting
the Greens function in Eq. (129) into the above, we obtain


~ s ~
~ t k
~
tr
J

kP
kJ

kP
X
X
x
y
1


;
Qxy im
Ab s;t6 ~ ixn m  s k
~
~
ix



k
n
t
k;n
(132)

Converting the above to integral form over the first Brillouin


zone (FBZ) and assuming that the Fermi energy lies inside
the gap between the upper and the lower bands, we have for
the Hall conductivity,



Lx Ly 2 ~
1
@~
n
@~
n
~
d

k
n

;
(141)
rxy
2A FBZ 4p2
@kx @ky

1
~
kXz k;
d2 ~
(142)
2
8p FBZ
where Lx Ly A, and, in the last step, we have used Eq. (40).
Thus, we see that the Hall conductivity is the integral of the
~
k-space
Berry curvature, Xz ~
k, over the FBZ if the Fermi
energy lies inside the energy gap.

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-22

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

In the spin-Hall effect,89,98 the Berry curvature carries a


~ X# ~
spin-dependence, X"z k
z k. This gives equal and opposite Hall conductivities for up and down spin species,
implying that the spin-Hall conductivity in Eq. (107) is finite
and equal to
rsH r"xy  r#xy :

(143)

In systems where TR symmetry is preserved, e.g., in a SOC


environment with no applied magnetic field, the resulting
spin-Hall conductivity represents a pure transverse spin current with equal numbers of up and down spin carriers flowing in opposite directions. The net charge Hall conductivity
in such situations rxy r"xy r#xy vanishes. Thus, TR symmetric SOC systems are a potential candidate for a pure spin
current source. When the Fermi level of the system lies
between a gap so that all bands are either completely empty
or full, rxy in Eq. (142) is a topological invariant. If one
~ T 2 ! S2 as a mapping from the two-dimensional
views ~
nk:
FBZ to the unit sphere, the integral of ~
n  @kx ~
n  @ky ~
n is a
topologically quantized value 4pn, where n 2 Z is the winding number of ~
n.
The above formula for the Hall conductivity in terms of
~
the Berry curvature in k-space
also applies to the quantum
Hall effect (QHE), which takes place in the presence of
strong magnetic fields. In this context, the integer n is known
as the TKNN index,125 which physically counts the number
of occupied Landau levels below the Fermi level. Remarkably, although the Landau-level problem is vastly different
from the general SOC problem, the topological quantization
of the Hall conductivity in Eq. (142) in both cases arises
~
~
from common k-space
Berry phase effects.128 Indeed, the kspace Berry curvature in the QHE arises from the Bloch
~ defined by the periodicity of the latwavefunctions jun ki
~ The latter
tice, as they evolve adiabatically with respect to k.
is also responsible for intrinsic contributions to the anomalous Hall effect (AHE) in the low temperature clean limit;
see, for example, Refs. 129 and 130.
VII. TIME-DEPENDENT MAGNETIC SYSTEMS

In Secs. IIIVI, we have covered the origin of real-space


and momentum-space gauge fields and their physical consequences. We next consider gauge fields in one more space,
namely, time. Gauge fields in time-space, as it turns out, also
have extremely important physical consequences in spintronic systems. We review them below.
A. Derivation

The gauge field in time-space arises naturally in spintronic systems in which there is a time-dependent magnetic
~ The Hamiltonian of such a system takes the form
field, Bt.
2
1 
~ r  1 glB~
~
~
H
p eA~
r  Bt:
2m
2

(144)

Following the method in Sec. III B 3, we diagonalize the


Schrodinger equation, Hjwi jwi, at time t by applying a
unitary transformation Ut defined as in Eq. (27), with h

and / carrying an explicit time-dependence. Writing the


energy operator as a time derivative,  i@t , we obtain
UtHtU t Utih@t U t;


2
1
~ r  1 glB rz jBtj
~ ih@t ihUt@t U t;
~
p eA~
2m
2
   hA0 t:
(145)
On the right-hand side, we obtain from the time-dependence
of U a gauge field in time-space,
A0 t  iUt@t U t:

(146)

In order to extract the physical meaning of this gauge field,


~t  ~
r, where
we note that it can be expressed as A0 t A
1
_
_
~
~
~ Bj;
~ m
~ m
~ 2~
~ is
n~
n (Ref. 5) (~
n B=j
At m
n~
n A t  ~
defined in Eq. (25)). Thus, the term hA0 t represents an
effective Zeeman-like term, indicating the presence of an
effective magnetic field in the rotating frame. Let us elucidate the origin of this field. The unitary transformation Ut
~L t of the
~L x
defines the instantaneous angular velocity x
coordinates (in the laboratory L-frame) as they follow the
time-dependent magnetic field. In the rotating frame R, this
~R U ~
~L U . Since
~R , where ~
rx
rx
vector is given by x
131
1
L
~ , the Zeeman-like term hA0 t in Eq. (145)
rx
U_ 2iU~
~R , which corresponds to an effective
equals h=2~
rx
magnetic field of ~
xR (omitting a scaling factor) in the rotating frame R. This translates into an effective magnetic field
~t ~
xL in the laboratory frame. Furthermore, with
B
~
n~
nt denoting the unit vector pointing along the direction
~
of Bt
at time t, we have the equation of motion
~L  ~
~
n. Performing a post cross product on both sides
n_ x
by ~
n, we arrive at the expression for the angular velocity
~L ~
~L  ~
x
n~
n_ x
n~
n or, in terms of the effective magnetic field,


~t  ~
~t ~
n_  ~
n B
n ~
n:
(147)
B
Thus, the effective magnetic field in the laboratory frame repn_  ~
n
resentation arising from A0 t has a component along ~
and ~
n. It does not have any component along ~
n_. The unitary
matrix used here [defined in Eq. (27)] is not unique. Specifically, different rotation matrices Ui , each specifying distinct
~Li , can be used to align the reference
angular velocities x
~
^z-axis along the instantaneous magnetic field Bt;
the freedom of choice here lies in determining the trajectory of the
remaining x^-^
y axes, that is, the rotation about ~
n itself. The second term on the right-hand side of Eq. (147) reflects this freedom of choice of the gauge transformation. It is not an
invariant of the gauge transformation (its magnitude being
dependent on the particular gauge choice) and, therefore, does
not represent a physical field. However, the first component
~? ~
n_  ~
n
B

(148)

is invariant with respect to the particular choice of gauge


transformation, depending only on the time-dependence of
~ We illustrate this magnetic field component in Fig. 11.
Bt.
The same expression can be derived classically by directly
comparing the spin vector in adjacent time frames.8,20 The

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-23

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

correspondingly transform as jwI ti eiH0 t=h jwti and


evolve according to the new Schrodinger equation,
HI tjwI ti ih@t jwI ti;

(152)

where HI t eiH0 t=h H1 eiH0 t=h . Explicitly, the Hamiltonian


in the interaction picture is found to be
HI t
FIG. 11. (Color online) In the presence of a time-dependent magnetic field,
~ jBtj~
~ nt, an additional magnetic field B
~? ~
n_  ~
n (vertical arrow)
Bt
is seen by spins. The net instantaneous magnetic field felt by spins is the
~ and B
~? , denoted by the dashed arrow.
vector sum of Bt

~
n_  ~
n component represents a physical magnetic field which
couples to the electron spins.8,132
B. Physical consequences
1. Spin-Hall effect

The spin-Hall effect due to the ~


k-space Berry curvature
was examined in Sec. VI B 1. It was shown that the
momentum-space curvature results in an anomalous velocity
component which is spin-dependent, driving the spatial spin
separation of the SHE. Here, we show that the gauge field in
time-space, A0 t, also leads to an intrinsic SHE in systems
with SOC. We shall make a heuristic distinction between these
two mechanisms, the former arising from spin-dependent
anomalous velocities and the latter from momentumdependent spin polarization. Semiclassically, these two mechanisms can be reconciled, as discussed in Sec. VII C.
The two key ingredients for the SHE are SOC and an
applied electric field which drives a charge current. An elec_
tric field accelerates electrons, namely ~
k 6 0, making the
momentum k~ carry explicit time-dependence. This, in turn,
~ time-dependent. Conse~ k
makes the spin-orbit field B
quently, A0 t arises naturally in SHE systems, as in
Eq. (145). For the SOC system, we formalize this idea by
switching to the interaction quantum picture, where operators, such as the momentum, can carry time-dependence. In
this picture, the original SOC Hamiltonian in Eq. (73) is split
into two parts, H H0 H1 , where
~ ~
H 0 eE
r

(149)

governs the time evolution of the operators and


2

H1

~
p
~
~ k
 c~
r  B
2m

(150)

governs the time evolution of the states. In the usual sense,


an operator A in the Schrodinger picture is transformed
to the interaction picture (subscript I) as AI t
eiH0 t=h AeiH0 t=h , carrying an explicit time-dependence, as
defined by the Heisenberg relation, A_ I i1hAI ; H0 . In particular, the momentum operator in the interaction picture is
found to be
~
~
pI t ~
p  eEt;

(151)

i.e., with the expected linear time-dependence due to the


electric field. State vectors jwti in the Schrodinger picture

~
p2I
~ k~I t;
 c~
r  B
2m

(153)

where
2
2 2~ ~
~~
~  eEi t @ Bk e Ej El t @ Bk    
~ k
~ k~I t B
B
2
h @ki
@kl @kj
2h
(154)

and summation over repeated indices is implied. The Hamiltonian in Eq. (153) is that of an electron subject to an explic~ as
itly time-dependent magnetic field, which we denote Bt
in Eq. (144). The origin of the gauge potential A0 t then follows from Eq. (145).
The gauge field A0 t leads to the intrinsic SHE in the
Rashba SOC system, discovered by Sinova et al.133 For
Rashba SOC, the effective spin-orbit field hasqdirection

~
n p1 py ; px ; 0, as in Eq. (75), where p p2x p2y .
Assuming an electric field applied along the x^-direction,
~ Ex x^, we obtain ~
E
n_ p1 0; eEx ; 0 (of course, p here carries a time-dependence through px , but which is weaker than
that of px itself. To a first approximation, we assume a constant p). Since ~
n is strictly in-plane (i.e., it lies in the x^-^
y
~? term in Eq. (148) represented by
plane of the 2DEG), the B
A0 is an out-of-plane magnetic field, which is along the ^zdirection by convention.
The SHE is an adiabatic phenomena in SOC systems.89
Let us here attempt to quantify the requirement for adiabatic~
~ kj
ity. In the ideal adiabatic limit, the magnetic field jB
should be infinitely strong, such that the spins always remain
~ is finite and the rel~ kj
aligned to it. However, in practice, jB
evant adiabatic condition reads
~  jB
~? j;
jBj

(155)

~ but with a
~ k,
i.e., the electron spin is primarily aligned to B
~? . For the Rashba SOC system, the
small component along B
adiabatic condition is
ak2
 Ex :
e

(156)

Inserting typical values for the Rashba parameter a 1011


eVm and the Fermi wave-vector k 108 m1 , we arrive at
the condition Ex  105 Vm1 , which usually holds true in
experiments. Assuming that the spin of electrons follow the
~R , which is the
direction of the net effective magnetic field, B
~
~
~
sum of the Rashba field BR k and B? , the classical spin vector is given by
~
s6

~R
h B
;
~R j
2 jB

(157)

where 6 represents spin aligned parallel or anti-parallel


~R . Note that 6 also denotes the two Rashba SOC
 to B

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-24

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

subbands. The component of the spin along the out-of-plane


z-direction is
^

1 h  _
~
sz 6
n~
n  ^z;
(158)
~R j 2
jB
where, to be consistent in units, the magnetic field in the
denominator is defined in terms of its equivalent angular
velocity. Applying the adiabatic limit to Eq. (158), in which
~ we obtain for
~R k,
~R approaches that of B
the magnitude of B
the out-of-plane spin polarization

1 h  _
~
n~
n  ^z;
~R j 2
jB


h2 h
1
 2 eEx py ;
6
p
2ap 2

sz 6

eh3 py Ex
:
4ap3

~6
W

Equation (159) above describes a transverse separation of


spins in the Rashba system. For example, let us consider the
case for the  subband (refer to Eq. (74)). Since the spin
y-direction
polarization sz / py , electrons moving in the ^
are polarized out-of-plane along ^z, whereas those moving
in the ^
y-direction are polarized along ^z (the spin polarization sz does not violate TR symmetry because it is odd in py ).
For the other eigenstate  , the directions of the polarization
are reversed. Thus, there is a cancellation of the polarizations
when both eigenstates are present. However, at the Fermi
level, there are more electrons in the  band, giving rise to a
net transverse spin separation and, hence, the SHE described
in Ref. 133 (see Eqs. (5)(7) there). Summing over the Fermi
surfaces of the two eigenstates yields an intrinsic spin-Hall
conductivity, Eq. (107), of rsH e=8p.20,133,134
The gauge field A0 t also explains the intrinsic spinHall effect in other spintronic systems,134 including the
~
k-linear135137 and k~3 Dresselhaus SOC138 and hole systems
with Rashba SOC.139141

(161)

i.e., describing the electronic amplitude of electrons on the


h2 ~

jBBLG j. The contwo layers, with eigenvalues 6 62m
trolled doping of BLG can shift the Fermi level into either
the conduction ( state) or valence band ( ),25 defining
which band contributes to the Hall transport. When an
~ Ex x^ is applied to the system, an additional
electric field E
out-of-plane component accompanies the strictly in-plane
~BLG . In the adiabatic limit, the out-of-plane
pseudospin field B
component induces a pseudospin polarization given by
sz

(159)


~
w6 B
;
w6 A

2meEx ky
h2 k4

(162)

along 6^z. The polarization corresponds to charge transfer


between the two layers. This is an essential ingredient for the
technology known as pseudospintronics,143145 in which binary states are encoded by relative charge densities on the
two monolayers.
Physically, the effect of this polarization for the 
eigenstate is to separate electrons with py > 0 (py < 0) to the
bottom (top) graphene monolayer of the BLG system. For
the K-valley, the BLG eigenstates are of the form
~ 6 w6 A; w6 B,
~ 142 i.e., the order of the components
W
is reversed, and, thus, the pseudospin polarization affects
~
electrons in the opposite manner compared to the K-valley;
electrons with py > 0 (py < 0) are transferred to the top (bottom) monolayer. Thus, the net effect is expected to vanish
when contributions from both valleys are taken into consideration. A finite effect is expected when the electron distribution in the two valleys differs, e.g., at the output of a valley
filter. An example of such a filter was described in Sec. IV C
1. Furthermore, we assume that the applied electric field is
sufficiently weak, such that intervalley transitions can be
neglected within the BLG system.150 The proposed effect in
BLG is completely analogous to the SHE upon replacements
~
r $~
s. It may, therefore, be termed the intrinsic pseudospinHall effect in graphene.

2. Pseudospin-Hall effect in graphene

An analogous effect takes place in bilayer graphene


(BLG) systems.134 The BLG system is modeled as two
coupled monolayer graphene sheets, with each layer having
~ B~ and A; B in the top and
two inequivalent lattice sites; A;
bottom layers, respectively. We assume the Bernal stacking
~
(A-B)
configuration, as it is commonly adopted. In the low
energy limit, electrons in the BLG system are described by
an effective 2  2 Hamiltonian142
HBLG 

h2
~
~BLG k;
~
sB
2m

(160)

~BLG kx2  ky2 ; 2kx ky ; 0. Here, ~


s is the vector of
where B
Pauli matrices acting on the pseudospin rather than the actual
electron spin. The symmetry of the lattice supports two
~ on the hexagonal
inequivalent, degenerate points, K and K,
~
Brillouin zone. In the K-valley, the two-component BLG
eigenstates are of the form142

3. Rayleigh scattering of polaritons

An analogous effect in optics occurs when polaritons


undergo Rayleigh scattering in a semiconductor microcavity.146 The polariton polarization is quantified as a pseudospin ~
s, where the pseudospin field is exactly that of bilayer
graphene BBLG in Eq. (160) above. Upon scattering (which
~
changes the wavevector k~akin to the E-field
effect), polaritons
z
acquire a finite s component, corresponding to circular polarization, whose sign depends on the initial momenta. The
manifestation of this effect has been observed in
experiments,147149 which is a promising indication for the
detection of the effect in spintronic and graphene systems.
4. Spin motive force

We consider the dynamics of spin particle in the presence of a magnetic texture that is nonuniform in both space
and time. The single electron Schrodinger equation for the

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-25

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

system is (we ignore the spin-independent magnetic vector


potential for simplicity)

dependent part gives rise to the spin-motive force. This is


given by151,152

~
p2 1
~ r; t ;
 gl ~
r  B~
2m 2 B

ad:
E i F 0i @t Aad:
i  @i A 0 ;

(163)

where the momentum operator ~


p ihr and the energy
operator  i
h@t . Following the method outlined in
Sec. III B 3, we perform a unitary transformation U~
r; t to
rotate the reference local ^z-axis to be aligned along the
magnetic field direction at every point in space and time.
The effective Hamiltonian reads, Eqs. (29) and (145),
1
1
~   hA0 t;
~
p A~
r2  glB rz jBj
2m
2

(164)

where A~
r and A0 t are the spatial and temporal gauge
fields, respectively. We consider the adiabatic limit. Under
this approximation, the transformation-induced gauge potentials become (see Eq. (32))
1
r rz 1  cos hrr /;
Aad: ~
2

(165)

1 z
Aad:
0 t r 1  cos h@t /;
2

(166)

~ r; t and the
where rz denotes the two spin eigenstates of B~
Hamiltonian in Eq. (164) reads
2 1
1 
~ hAad:
~
p Aad: ~
r  glB rz jBj
0 t U ; (167)
2m
2
where U is the spin-independent potential which arises from
the quadratic terms of the non-diagonal elements of A~
r after applying the adiabatic approximation [see Eq. (38)]. This
can be written compactly as
U


1 
1
rh2 sin2 hr/2
r~
n2 ;
8m
8m

(168)

where ~
n sin h cos /; sin h sin /; cos h is the unit vector in
spin space and r~
n2 @i nj @i nj . To calculate all forces in
the system, it is instructive to lump all of the potential terms
together,

1
~ 1 r~
n2 :
A00 rz 1  cos h@t /  glB jBj
2
8m

(169)

We proceed to evaluate the curvature of the gauge fields. In


the 2  2 spin-12 space, the adiabatic approximation coincides
with the Abelian reduction. Thus, the non-Abelian contribution of the Yang-Mills Eq. (31) vanishes, and the curvature
is given simply by
ad:
F ij @i Aad:
j  @j Ai ;

(170)

where i; j ft; x; y; zg are the space-time coordinates. The


spatial curvatures are simply the Berry curvature of the Berry
(vector) potential in Eq. (40), which represents an effective
magnetic field. On the other hand, the temporal curvature is
equivalent to a generalized electric field, whose spin-

(171)

where i fx; y; zg. Substituting Eqs. (165) and (169) into


Eq. (171), the effective electric field is then given by

h
gl 
i
B
l
z sinh
_
_
~ 1 rr~
rjBj
hr/
 /rh
 rz
n2 :
E r
2
2
8m
(172)
The last term, which is the gradient of the spin-independent
potential U has been derived classically in Ref. 8 by considering a rotating frame of reference. It tends to repel the
charge from a spatial region, where the spin texture is varying most rapidly. The second term drives the Stern-Gerlach
force, which is proportional to the spatial gradient of the
magnetic field strength. Assuming that the external B-field is
constant in magnitude and changes only in direction (so as to
induce the spatial spin texture ~
n, we can neglect this term.
The spin-dependent force due to the time-dependent B-field
is then given by
h
i
_
_
~ eE 6 h sin h hr/
 /rh
;
(173)
F
2
where the additional factor of h=e is introduced to give E the
required units of V=m. The above force is called the spin
motive force (SMF),153155 and the 6 sign corresponds to the
force experienced by electrons with 6z spin. Due to the
monopolar form of Aad:
0 in Eq. (166), the SMF can be related
to the local spin texture ~
n via a Weiss-Zumino-like relation,
Eq. (40), i.e.,
~i 6 h ~
n  @t ~
n:
n:@i~
F
2

(174)

The SMF has been analyzed in domain wall configurations


with an applied magnetic field, which induces precession of
the local magnetic moments154,156 and in magnetic vortex
cores in magnetic disks.157 Moreover, the SMF results in upand down-spin carriers being separated spatially along the
direction of E, i.e., a spin current. In Ref. 158, it was shown
that, in the presence of SOC, this leads to the generation of a
charge current perpendicular to E. This is known as the
inverse spin-Hall effect.158160
We consider the effect of the SMF on the electron transport in the presence of the extrinsic SOC effect arising, e.g.,
from side-jump and skew-scattering processes. Since the material system is a semiconductor, its longitudinal conductivity
1 is related to the carrier density and, hence, spin accumulation, i.e., 1 r0 10 elr s  r0 10 r 1 , where l es=m
is the mobility. Due to the spin dependence of the conductivity, the presence of the spin-motive field E  in Eq. (173)
would drive a combination of spin and charge currents, i.e.,
j 0 r0 j  r 10 r0 1 r E 0 r0 E  r ;

(175)

where E 0 is the ordinary electric field. It is reasonable to


assume that the spin accumulation s is polarized along the

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-26

Fujita et al.

J. Appl. Phys. 110, 121301 (2011)

~
applied B-field
direction and, hence, that of the spin texture
n^. For simplicity, we assume the net n^ to be pointing in the
z-direction. Thus, from Eq. (175), one directly obtains the
spin and charge currents as
j 0 r0 j  r r0 10 E 0 1z E z rz 1z E 0 10 E z :

(176)

We now consider the Hall current due to the extrinsic SOC


effect. This is given by
r:^
n  E ) jH 0i r0 jH i r
j 0H r0 j H r 1SH ^
1SH eilk n^l rl E 0k r0 E zk rz ;

(177)

k m
where 1SH 1sjSH 1ss
SH 1D 4 es1 pmDqF u, where D is
the diffusion constant, qF is the density of states at the Fermi
level, and u is the impurity scattering potential.161,162 As
before, we assume the net spin texture to be in the vertical
direction (^
nl dlz ) and the applied ordinary electric field to
be in the x-direction. Then, the Hall currents are

jH 0y r0 jH zi rz 1SH eizk n^z rz E 0k r0 E zk rz


r0 eizk 1SH E zk rz 1SH E 0y :

(178)

Thus, the SMF drives an in-plane charge Hall current, while


the ordinary E-field drives a spin-Hall current, which agrees
with the microscopic analysis of Ref. 158.
C. Semiclassical connection with k -space Berry
curvature

In Sec. VI B 1, the intrinsic SHE due to the Berry curva~


ture in k-space
was discussed. The spin separation is
achieved via a spin-dependent anomalous velocity, Eq.
(106), which pushes opposite spin species in opposite transverse directions, e.g., in p-doped semiconductors. On the
other hand, in Rashba SOC systems, spin separation occurs
from a momentum-dependent magnetic field, Eq. (159),
which polarizes electrons along opposite directions out-ofplane, depending on their transverse momenta. Despite being
distinct, the two intrinsic mechanisms share the common
requirements of adiabaticity and time-dependence. This
prompts one to ask: Are the two mechanisms related?
To answer this question, we investigate possible connec~? field (given
tions between the anomalous velocity and the B
by Eq. (148)) in SHE systems.20 In this endeavor, we employ
~
the k-space
analogue of the analysis by Aharanov and Stern8
of the origin of Berrys curvature in real space. We begin with
the general spin-orbit Hamiltonian in Eq. (73). The velocity
along the i-th coordinate is given by Hamiltons equation,

ti

~
~ k
1
pi
@ B
ri ; H   c
~
r:
m
ih
@pi

(179)

~ jBj~
~ n, the partial derivative
Writing the spin-orbit field as B
in Eq. (179) can be expanded into its magnitude and directional parts as
~ @jBj
~
@B
n
~ @~
~

:
n jBj
@pi
@pi
@pi

(180)

Henceforth, we shall adopt a semiclassical approach,


whereby all operators are reduced to scalars. When the mag~ Bt,
~ the spins exnetic field carries a time-dependence, B
~
perience an additional field B? , as discussed in Sec. VII A.
~R ,
Assuming that the spins are polarized along the net field B
~
~
the classical unit spin vector is ~
r BR =jBR j. Then, we
obtain for the velocity
!
 @~
~ @jBj
~
pi
jBj
h  _
n
~
ti  c
;
(181)

n~
n 
~R j @pi 2c
m
@pi
jB
~? gives it the corwhere the conversion factor of h=2c for B
rect units of Tesla. The first term is the kinetic velocity and
is of no interest here. We focus on the second term. In the ad~  jB
~? j, the second term in Eq. (181)
iabatic limit, i.e., jBj
reads
ti c

 @~
~ h 
@jBj
n
 ~
:
n_  ~
n 
@pi 2
@pi

(182)

@~
n
By the chain rule, ~
n_ k_j @k
, where summation over j is
j
implied. Rearranging the terms, we obtain

ti c



~ k_j @~
@jBj
n @~
n
~
n:


2 @ki @kj
@pi

(183)

The first term represents a velocity component that is due to


~ in momentum
~ k
the inhomogeneity of the spin-orbit field B
space, i.e., it is the reciprocal space analogue of the SternGerlach force. The second term in Eq. (183) is the anomalous KL velocity of electrons due to Berrys curvature in ~
kspace.20 In fact, it can be written as


~
@~
n @~
n
n
~
ijk k_j Xk k;

(184)
tKL;i k_j 
2 @ki @kj
where the last equality follows from Eq. (104). This is precisely the KL velocity in Eq. (106) arising from the Berrys
~ in k-space.
~
curvature Xk
Moreover, the spin-dependence of

TABLE I. A summary of gauge fields, the context in which they occur, and their physical (measurable) consequences.
Space
Real space, ~
r

Momentum, k~
Time, t

Context

Physical consequences

 Spatially nonuniform magnetic field or magnetization


 Spin-orbit coupling
 Strain (in graphene)
 Spin-orbit coupling
 Bloch wavefunctions
 Time-dependent magnetic field or magnetization

 Spin torque, chirality-driven spin-Hall effect (SHE)


 Quantum spin-Hall effect, spin transverse force, spin torque
 Valley filtering, edge states, quantum valley Hall effect (in graphene)
 Intrinsic SHE
 Anomalous Hall effect
 Intrinsic SHE

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-27

Fujita et al.

the KL term can be obtained by letting the spin vector be


~R j
~R =jB
aligned anti-parallel to the net magnetic field, ~
r B
in Eq. (179) for the down-spin state. Thus, the anomalous velocity due to the ~
k-space Berry curvature physically arises
~? magnetic field component,
from the presence of the B
which in turn is related to the time component of the gauge
field, A0 t.
The above highlights the physical origin of the Berry
~
curvature in k-space.
We note that adiabaticity was translated
from a quantum mechanical to classical description, where,
in the former, we neglected off-diagonal interband elements
~
of the k-space
gauge field, while in the latter we considered
~? j
the limit of a strong magnetic field strength relative to jB
(see Eq. (155)).
VIII. SUMMARY

We have examined the origin and physical consequences of gauge fields in spintronic and other spin-like systems, such as photons and graphene. The gauge fields can lie
~
in three spaces, namely, real space ~
r, momentum space k,
and time t. Berry gauge fields are associated with the adiabatic transport of quantum states with respect to parameter
space. We summarize this paper in Table I below.
In ~
r -space, Berry gauge fields arise in the presence of
spatially nonuniform magnetic fields or magnetization (e.g.,
a domain wall); see Sec. III. Non-Abelian ~
r-space gauge
fields can arise from spin-orbit coupling (Sec. V) or, for
example, in graphene that has been mechanically strained
(Sec. IV). Generally, the ~
r -space gauge fields can be considered as magnetic vector potentials which affect carrier transport via Lorentz forces (see Secs. III C 1, IV C 2, and V C 2)
or exert spin torque (Secs. III C 3 and V C 4).
~
In k-space,
Berry gauge fields due to spin-orbit coupling
arise ubiquitously in spintronic, optical, and graphene systems (see Sec. VI). Such gauge fields can be considered as
~
describing magnetic fields in k-space
and give rise to an
anomalous velocity (analogous to the Lorentz force in
~
r-space). This leads to observable effects, such as the spinHall effect.
Finally, time-dependent gauge fields arise in the presence of time varying magnetic fields. Physically, they lead to
the spin-Hall effect and spin motive force (see Sec. VII B).

ACKNOWLEDGMENTS

We gratefully acknowledge financial support extended


by the National University of Singapore (NUS), the Science
and Engineering Research Council (SERC), the SMF-NUS
RHA (grant nos. R-263-000-632-592, R-263-000-632-646
and R-398-000-061-305), and the Agency for Science, Technology and Research of Singapore.
1

S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S. von


Molnar, M. L. Roukes, A. Y. Chtchelkanova, and D. M. Treger, Science
294, 1488 (2001).
2
Y. Aharonov and D. Bohm, Phys. Rev. 115, 485 (1959).
3
M. V. Berry, Proc. R. Soc. London, Ser. A 392, 45 (1984).
4
S. Pancharatnam, Proc. Indian Acad. Sci., Sect. A 44, 247 (1956).
5
G. Tatara, H. Kohno, and J. Shibata, Phys. Rep. 468, 213 (2008).

J. Appl. Phys. 110, 121301 (2011)


6

C. N. Yang and R. L. Mills, Phys. Rev. 96, 191 (1954).


M. Iyanaga, J. Math. Phys. 22, 2713 (1981).
8
Y. Aharonov and A. Stern, Phys. Rev. Lett. 69, 3593 (1992).
9
P. Bruno, V. K. Dugaev, and M. Taillefumier, Phys. Rev. Lett. 93,
096806 (2004).
10
K. Y. Bliokh and Y. P. Bliokh, Ann. Phys. 319, 13 (2005).
11
S. G. Tan, M. B. A. Jalil, and T. Fujita, Ann. Phys. 325, 1537 (2010).
12
S. G. Tan, M. B. A. Jalil, X.-J. Liu, and T. Fujita, Phys. Rev. B 78,
245321 (2008).
13
J. Ye, Y. B. Kim, A. J. Millis, B. I. Shraiman, P. Majumdar, and Z.
Tesanovic, Phys. Rev. Lett. 83, 3737 (1999).
14
S. H. Chun, M. B. Salamon, Y. Lyanda-Geller, P. M. Goldbart, and P. D.
Han, Phys. Rev. Lett. 84, 757 (2000).
15
Y. Taguchi and Y. Tokura, Phys. Rev. B 60, 10280 (1999).
16
K. Ohgushi, S. Murakami, and N. Nagaosa, Phys. Rev. B 62, 6065
(2000).
17
S. Onoda and N. Nagaosa, Phys. Rev. Lett. 90, 196602 (2003).
18
Y. Taguchi, Y. Oohara, H. Yoshizawa, N. Nagaosa, and Y. Tokura,
Science 291, 2573 (2001).
19
Y. Taguchi, T. Sasaki, S. Awaji, Y. Iwasa, T. Tayama, T. Sakakibara,
S. Iguchi, T. Ito, and Y. Tokura, Phys. Rev. Lett. 90, 257202 (2003).
20
T. Fujita, M. B. A. Jalil, and S. G. Tan, New J. Phys. 12, 013016 (2010).
21
G. Metalidis and P. Bruno, Phys. Rev. B 74, 045327 (2006).
22
C. H. Lee, S. G. Tan, M. B. A. Jalil, G. Wu, and N. Chen, 55th Annual
Conference on Magnetism and Magnetic Materials, Atlanta, GA,
November 1418, 2010. poster presentation DU-01.
23
Ya. B. Bazaliy, B. A. Jones, and S.-C. Zhang, Phys. Rev. B. 57, R3213
(1998).
24
R. Saito, G. Dresselhaus, and M. S. Dresselhaus, Physical Properties of
Carbon Nanotubes (Imperial College Press, London, 1998).
25
K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, S. V.
Dubonos, I. V. Grigorieva, and A. A. Firsov, Science 306, 666 (2004).
26
K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V. Khotkevich,
S. V. Morozov, and A. K. Geim, Proc. Natl. Acad. Sci. U.S.A. 102,
10451 (2005).
27
A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K.
Geim, Rev. Mod. Phys. 81, 109 (2009).
28
K.-I. Sasaki and R. Saito, Prog. Theor. Phys. Suppl. 176, 253 (2008).
29
V. M. Pereira and A. H. Castro Neto, Phys. Rev. Lett. 103, 046801
(2009).
30
T. Fujita, M. B. A. Jalil, and S. G. Tan, Appl. Phys. Lett. 97, 043508
(2010).
31
A. Chaves, L. Covaci, Kh. Yu. Rakhimov, G. A. Farias, and F. M. Peeters, Phys. Rev. Lett. 82, 205430 (2010).
32
Z. Wu, F. Zhai, F. M. Peeters, H. Q. Xu, and K. Chang, Phys. Rev. Lett.
106, 176802 (2010).
33
A. Rycerz, J. Tworzydlo, and C. W. J. Beenakker, Nat. Phys. 3, 172
(2007).
34
A. Matulis, F. M. Peeters, and P. Vasilopoulos, Phys. Rev. Lett. 72, 1518
(1994).
35
M. R. Masir, P. Vasilopoulos, and F. M. Peeters, Appl. Phys. Lett. 93,
242103 (2008).
36
L. DellAnna and A. De Martino, Phys. Rev. B 80, 155416 (2009).
37
S. Ghosh and M. Sharma, J. Phys.: Condens. Matter 21, 292204 (2009).
38
E. Hall, Am. J. Math. 2, 287 (1879).
39
K. von Klitzing, G. Dorda, and M. Pepper, Phys. Rev. Lett. 45, 494
(1980).
40
N. Levy, S. A. Burke, K. L. Meaker, M. Panlasigui, A. Zettl, F. Guinea,
A. H. Castro Neto, and M. F. Crommie, Nat. Phys. 329, 544 (2010).
41
T. Low and F. Guinea, Nano Lett. 10, 3551 (2010).
42
F. Guinea, M. I. Katsnelson, and A. K. Geim, Nat. Phys. 6, 30
(2010).
43
C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 226801 (2005).
44
C. L. Kane and E. J. Mele, Phys. Rev. Lett. 95, 146802 (2005).
45
D. N. Sheng, Z. Y. Weng, L. Sheng, and F. D. M. Haldane, Phys. Rev.
Lett. 97, 036808 (2006).
46
B. A. Bernevig, T. L. Hughes, and S. C. Zhang, Science 314, 1757
(2006).
47
M. Konig, S. Wiedmann, C. Brune, A. Roth, H. Buhmann, L. W. Molenkamp, X. L. Qi, and S.-C. Zhang, Science 318, 766 (2007).
48
D. Hsieh, D. Qian, L. Wray, Y. Xia, Y. S. Hor, R. J. Cava, and M. Z.
Hasan, Nature 452, 970 (2008).
49
H. J. Zhang, C. X. Liu, X. L. Qi, X. Dai, Z. Fang, and S.-C. Zhang, Nat.
Phys. 5, 438 (2009).
7

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-28
50

Fujita et al.

Y. Xia, D. Qian, D. Hsieh, L. Wray, A. Pal, H. Lin, A. Bansil, D.


Grauer, Y. S. Hor, R. J. Cava, and M. Z. Hasan, Nat. Phys. 5, 398
(2009).
51
K.-I. Sasaki and K. Wakabayashi, Phys. Rev. B 82, 035421 (2010).
52
M. Fujita, K. Wakabayashi, K. Nakada, and K. Kusakabe, J. Phys. Soc.
Jpn. 65, 1920 (1996).
53
K. Nakada, M. Fujita, G. Dresselhaus, and M. S. Dresselhaus, Phys. Rev.
B 54, 17954 (1996).
54
R. Winkler, Spin-Orbit Coupling Effects in Two-Dimensional Electron
and Hole Systems (Springer-Verlag, Berlin, 2003).
55
D. D. Awschalom and N. Samarth, Physics 2, 50 (2009).
56
E. I. Rashba, Physica E 34, 31 (2006).
57
E. I. Rashba, Physica E 20, 189 (2004).
58
E. I. Rashba, Fiz. Tverd. Tela (Leningrad) 2, 1224 (1960).
59
Y. A. Bychkov and E. I. Rashba, J. Phys. C 17, 6039 (1984).
60
J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki, Phys. Rev. Lett. 78,
1335 (1997).
61
G. Dresselhaus, Phys. Rev. 100, 580 (1955).
62
M. I. Dyakanov and V. Y. Kachorovskii, Sov. Phys. Semicond. 20, 110
(1986).
63
A. G. Malshukov, C. S. Tang, C. S. Chu, and K. A. Chao, Phys. Rev.
Lett. 95, 107203 (2005).
64
L. Jiang and M. W. Wu, Phys. Rev. B 72, 033311 (2005).
65
B. A. Bernevig and S. C. Zhang, Phys. Rev. Lett. 96, 106802 (2006).
66
Y. Aharonov and A. Casher, Phys. Rev. Lett. 53, 319 (1984).
67
J. Anandan, Phys. Lett. A 138, 347 (1989).
68
H. Mathur and A. D. Stone, Phys. Rev. Lett. 68, 2964 (1992).
69
Y. Oreg and O. Entin-Wohlman, Phys. Rev. B 46, 2393 (1992).
70
S. Oh and C.-M. Ryu, Phys. Rev. B 51, 13441 (1995).
71
S. Q. Shen, Phys. Rev. Lett. 95, 187203 (2005).
72
T.-Z. Qian and Z.-B. Su, Phys. Rev. Lett. 72, 2311 (1994).
73
N. Hatano, R. Shirasaki, and H. Nakamura, Phys. Rev. A 75, 032107
(2007).
74
Y. Aharonov and J. Anandan, Phys. Rev. Lett. 58, 1593 (1987).
75
J. Cserti and G. David, Phys. Rev. B 74, 172305 (2006).
76
G. David and J. Cserti, Phys. Rev. B 81, 121417 (2010).
77
J. Hirsch, Phys. Rev. Lett. 83, 1834 (1999).
78
B. Basu and P. Bandyopadhyay, Phys. Lett. A 373, 148 (2008).
79
S. G. Tan, M. B. A. Jalil, X.-J. Liu, and T. Fujita, Ann. Phys. 326, 207
(2011); S. G. Tan, M. B. A. Jalil, and X.-J. Liu, arXiv:0705.3502 (2007).
80
A. Manchon and S. Zhang, Phys. Rev. B 78, 212405 (2008).
81
A. Manchon and S. Zhang, Phys. Rev. B 79, 094422 (2009).
82
A. Chernyshov, M. Overby, X. Liu, J. K. Furdyna, Y. Lyanda-Geller and
L. P. Rokhinson, Nat. Phys. 5, 656 (2009).
83
I. M. Miron, G. Gaudin, S. Auffret, B. Rodmacq, A. Schuhl, S. Pizzini,
J. Vogel, and P. Gambardella, Nature Mater. 9, 230 (2010).
84
A. Majumdar, Phys. Rev. B 54, 11911 (1996).
85
H. Z. Xu and Y. Okada, Appl. Phys. Lett. 79, 3119 (2001).
86
Y. Jiang, M. B. A. Jalil, and T. S. Low, Appl. Phys. Lett. 80, 1673 (2002).
87
M. B. A. Jalil, S. G. Tan, T. Liew, K. L. Teo, and T. C. Chong, J. Appl.
Phys. 95, 7321 (2004).
88
T. Fujita, M. B. A. Jalil, and S. G. Tan, IEEE Trans. Magn. 46, 1323
(2010).
89
S. Murakami, N. Nagaosa, and S. Zhang, Science 301, 1348 (2003).
90
G. Sundaram and Q. Niu, Phys. Rev. B 59, 14915 (1999).
91
R. Karplus and J. M. Luttinger, Phys. Rev. 95, 1154 (1954).
92
M. I. Dyakanov and V. I. Perel, JETP Lett. 13, 467 (1971).
93
M. I. Dyakanov and V. I. Perel, Phys. Lett. A 35, 459 (1971).
94
N. A. Sinitsyn, J. Phys.: Condens. Matter 20, 023201 (2008).
95
N. Nagaosa, J. Sinova, S. Onoda, A. H. MacDonald and N. P. Ong, Rev.
Mod. Phys. 82, 1539 (2010).
96
E. I. Rashba, Semiconductors 42, 905 (2008).
97
J. M. Luttinger, Phys. Rev. 102, 1030 (1956).
98
T. Fujita, M. B. A. Jalil, and S. G. Tan, Ann. Phys. 324, 2265 (2009).
99
J. Spector, H. L. Stormer, K. W. Baldwin, L. N. Pfeiffer, and K. W. West,
Appl. Phys. Lett. 56, 1290 (1990).
100
M. Noguchi, H. Sakakibara, and T. Ikoma, Jpn. J. Appl. Phys. 32, 5014
(1993).
101
V. I. Perel, S. A. Tarasenko, I. N. Yassievich, S. D. Ganichev, V. V.
Belkov, and W. Prettl, Phys. Rev. B 67, 201304 (2003).
102
G. Y. Guo, S. Murakami, T.-W. Chen, and N. Nagaosa, Phys. Rev. Lett.
100, 096491 (2008).
103
T. Kimura, Y. Otani, T. Sato, S. Takahashi, and S. Maekawa, Phys. Rev.
Lett. 98, 156601 (2007).

J. Appl. Phys. 110, 121301 (2011)


104

G. Schmidt, D. Ferrand, L. W. Molenkamp, A. T. Filip, and B. J. van


Wees, Phys. Rev. B 62, R4790 (2000).
105
B. Ya. Zeldovich and V. S. Liberman, Sov. J. Quantum Electron. 20, 427
(1990).
106
M. Onoda, S. Murakami, and N. Nagaosa, Phys. Rev. Lett. 93, 083901
(2004).
107
M. Onoda, S. Murakami, and N. Nagaosa, Phys. Rev. E 74, 066610 (2006).
108
K. Yu. Bliokh and V. D. Freilikher, Phys. Rev. B 72, 035108 (2005).
109
K. Yu. Bliokh and Yu. P. Bliokh, Phys. Rev. E 70, 026605 (2004).
110
E. I. Fedorov, Dokl. Akad. Nauk SSSR 105, 465 (1955).
111
C. Imbert, Phys. Rev. D 5, 787 (1972).
112
O. Hosten and P. Kwiat, Science 319, 787 (2008).
113
K. Y. Bliokh, A. Niv, V. Kleiner, and E. Hasman, Nature Photon. 2, 748
(2008).
114
R. Matsumoto and S. Murakami, Phys. Rev. Lett. 106, 197292 (2011).
115
R. Matsumoto and S. Murakami, Phys. Rev. B 84, 184406 (2011).
116
H. Katsura, N. Nagaosa, and P. A. Lee, Phys. Rev. Lett. 104, 066403
(2010).
117
Y. Onose, T. Ideue, H. Katsura, Y. Shiomi, N. Nagaosa, and Y. Tokura,
Science 329, 297 (2010).
118
S. Fujimoto, Phys. Rev. Lett. 103, 047203 (2009).
119
Y. Kajiwara, K. Harii, S. Takahashi, J. Ohe, K. Uchida, M. Mizuguchi, H.
Umezawa, H. Kawai, K. Ando, K. Takanashi, S. Maekawa and E. Saitoh,
Nature 464, 262 (2010).
120
S. O. Demokritov, B. Hillebrands, and A. N. Slavin, Phys. Rep. 348, 441
(2001).
121
F. Wilczek and A. Zee, Phys. Rev. Lett. 52, 2111 (1984).
122
D. Xiao, W. Yao, and Q. Niu, Phys. Rev. Lett. 99, 236809 (2007).
123
P. Gosselin, A. Berard, and H. Mohrbach, Eur. Phys. J. C 59, 883 (2009).
124
M. Z. Hasan and C. L. Kane, Rev. Mod. Phys. 82, 3045 (2010).
125
D. J. Thouless, M. Kohmoto, M. P. Nightingale, and M. den Nijs, Phys.
Rev. Lett. 49, 405 (1982).
126
L. Fu and C. L. Kane, Phys. Rev. B 76, 045302 (2007).
127
H.-Z. Lu, W.-Y. Shan, W. Yao, Q. Niu, and S.-Q. Shen, Phys. Rev. B 81,
115407 (2010).
128
X.-L. Qi, Y.-S. Wu, and S.-C. Zhang, Phys. Rev. B 74, 085308 (2006).
129
M. Onoda and N. Nagaosa, J. Phys. Soc. Jpn. 71, 19 (2002).
130
Z. Fang, N. Nagaosa, K. S. Takahashi, A. Asamitsu, R. Mathieu, T. Ogasawara, H. Yamada, M. Kawasaki, Y. Tokura, and K. Terakura, Science
302, 92 (2003).
131
A. G. Wagh and V. C. Rakhecha, Phys. Rev. A 48, 1729 (1993).
132
J. Xiao, A. Zangwill, and M. D. Stiles, Phys. Rev. B 73, 054428 (2006).
133
J. Sinova, D. Culcer, Q. Niu, N. A. Sinitsyn, T. Jungwirth, and A. H.
MacDonald, Phys. Rev. Lett. 92, 126603 (2004).
134
T. Fujita, M. B. A. Jalil, and S. G. Tan, J. Phys. Soc. Jpn. 78, 104714
(2009).
135
T.-W. Chen, C.-M. Huang, and G. Y. Guo, Phys. Rev. B 73, 235309 (2006).
136
S.-Q. Shen, Phys. Rev. B 70, 081311 (2004).
137
N. A. Sinitsyn, E. M. Hankiewicz, W. Teizer, and J. Sinova, Phys. Rev. B
70, 081312 (2004).
138
B. A. Bernevig and S.-C. Zhang, arXiv:0412550 (2004).
139
J. Schlieman and D. Loss, Phys. Rev. B 71, 085308 (2005).
140
B. A. Bernevig and S.-C. Zhang, Phys. Rev. Lett. 95, 016801 (2005).
141
X. Dai, Z. Fang, Y.-G. Yao, and F.-C. Zhang, Phys. Rev. Lett. 96, 086802
(2006).
142
E. McCann and V. I. Falko, Phys. Rev. Lett. 96, 086805 (2006).
143
H. Min, G. Borghi, M. Polini, and A. H. MacDonald, Phys. Rev. B 77,
041407 (2008).
144
P. San-Jose, E. Prada, E. McCann, and H. Schomerus, Phys. Rev. Lett.
102, 247204 (2009).
145
S. G. Tan, M. B. A. Jalil, D. E. Koh, H. K. Lee, and Y. H. Wu, J. Magn.
Magn. Mater. 322, 2390 (2010).
146
A. Kavokin, G. Malpuech, and M. Glazov, Phys. Rev. Lett. 95, 136601
(2005).
147
M. Glazov and A. Kavokin, J. Lumin. 125, 118 (2007).
148
W. Langbein, Proceedings of 26th International Conference on Physics
of Semiconductors (Institute of Physics, Bristol, 2003), p. 112.
149
C. Leyder, M. Romanelli, J. Ph. Karr, E. Giacobino, T. C. H. Liew,
M. M. Glazov, A. V. Kavokin, G. Malpuech, and A. Bramati, Nat. Phys.
3, 628 (2007).
150
I. Martin, Y. M. Blanter, and A. F. Morpurgo, Phys. Rev. Lett. 100,
036804 (2008).
151
P.-Q. Jin, Y.-Q. Li, and F.-C. Zhang, J. Phys. A 39, 7115 (2006).
152
M. B. A. Jalil and S. G. Tan, IEEE Trans. Magn. 46, 1626 (2010).

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

121301-29
153

Fujita et al.

A. Stern, Phys. Rev. Lett. 68, 1022 (1992).


S. E. Barnes and S. Maekawa, Phys. Rev. Lett. 98, 246601 (2007).
155
Y. Tserkovnyak and M. Mecklenburg, Phys. Rev. B 77, 134407 (2008).
156
M. Stamenova, T. N. Todorov, and S. Sanvito, Phys. Rev. B 77, 054439
(2008).
157
J.-I. Ohe and S. Maekawa, J. Appl. Phys. 105, 07C706 (2009).
158
J. Shibata and H. Kohno, Phys. Rev. Lett. 102, 086603 (2009).
154

J. Appl. Phys. 110, 121301 (2011)


159

E. Saitoh, M. Ueda, H. Miyajima, and G. Tatara, Appl. Phys. Lett. 88,


182509 (2006).
160
L. K. Werake, B. A. Ruzicka, and H. Zhao, Phys. Rev. Lett. 106, 107205
(2011).
161
H.-A. Engel, B. I. Halperin, and E. I. Rashba, Phys. Rev. Lett. 95, 166605
(2005).
162
W.-K. Tse and S. D. Sarma, Phys. Rev. Lett. 96, 056601 (2006).

[This article is copyrighted as indicated in the article. Reuse of AIP content is subject to the terms at: http://scitation.aip.org/termsconditions. Downloaded to ] IP:
143.107.180.128 On: Tue, 02 Dec 2014 13:05:53

You might also like