PhysRevB 99 245204

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

PHYSICAL REVIEW B 99, 245204 (2019)

Nonlinear spin-current generation in quantum wells with arbitrary


Rashba-Dresselhaus spin-orbit interactions
Aniruddha Pan and D. C. Marinescu
Department of Physics and Astronomy, Clemson University, Clemson, South Carolina 29634, USA

(Received 1 April 2019; revised manuscript received 28 May 2019; published 7 June 2019)

The problem of the nonlinear spin-currents generated by an electric field E and a temperature gradient ∇T in
spin-orbit coupled systems is revisited in a different formalism. Here, the second-order correction to the particle
distribution function δ f (2) is derived in a semiclassical approximation that takes into account the local change
in the equilibrium distribution function induced by the external fields. Our approach departs significantly from
the present theory, where δ f (2) is written as an iterative solution to the Boltzmann transport equation in the
relaxation-time approximation. As we show, such an expression does not actually satisfy the collision term of
the equation, and therefore it is not self-consistent. We apply our formalism to the case of a quantum well with
arbitrary values of the linear Rashba α and Dresselhaus β interactions. For the whole range of α versus β values,
we obtain analytic results for all the spin currents that can be driven in the system, proportional with E2 , ∇T 2 ,
or with E · ∇T . The magnitude of these currents is smaller than previously anticipated.

DOI: 10.1103/PhysRevB.99.245204

I. INTRODUCTION to have, in both cases, significant values, thus heralding a new


paradigm for experimental realization.
Studied intensively for the past decade, spin-orbit (SO)
More recently, the same computational algorithm was
coupled systems continue to provide an interesting explo- applied to the case of a semiconductor quantum well with
ration ground for theoretical and experimental investigations Rashba and Dresselhaus spin-orbit interaction in Ref. [13],
focused on detecting phenomena that can lead to the establish- where spin currents quadratic in the applied electric field
ment of a spin-dependent transport paradigm for applications were evaluated. Analytic expressions were derived when only
[1,2]. Of all systems endowed with some form of spin-orbit one interaction was present, either α = 0 or β = 0, while
interaction, two-dimensional III-V semiconductor quantum numerical results were obtained for the case of both couplings
wells with inversion asymmetry occupy a preeminent spot, being considered. Thus, it was found that pure spin currents,
given their well-understood behavior as normal Fermi sys- whose amplitude is proportional with α or β, are driven
tems, which, coupled with the high degree of electric control simultaneously along parallel and perpendicular directions
over the SO parameters, makes them suitable for a large on the electric field. In each case, the spin polarization is
range of experiments [3–9]. There, the spin-orbit interaction perpendicular to the direction of propagation. The calculated
is described by the linear coupling between the spin and mo- magnitude of the parallel and perpendicular currents is differ-
mentum with strength α for the Rashba term, which originates ent, with the current along the direction of the electric field
in the inversion asymmetry of the well [10], and with strength about five times larger, but in both cases it was significant
β for the Dresselhaus term, which originates in the inversion enough to exceed the values obtained by other electric means.
asymmetry of the crystal [11]. Since the two terms rotate the A numerical study of the case α = 0 and β = 0 indicated that
spin in opposite directions, their simultaneous consideration both currents disappear in the limit α = β.
in theoretical investigations is somewhat challenging, and In this paper, we reformulate the theory of nonlinear spin
often solutions are given only in numerical formalisms. currents by using a different expression for δ f (2) . This is nec-
The search for the realization of stable spin currents in essary since, as we demonstrate below, a second-order distri-
two-dimensional systems has received an important impetus bution function cannot be derived iteratively in the relaxation-
from the predicted nonlinear generation of such currents in time approximation as it does not satisfy self-consistently the
anisotropic Fermi pockets that relied on using the second- collision term of the Boltzmann transport equation. Instead,
order correction to the distribution function induced by the we calculate δ f (2) from a power expansion of the Fermi
electric field or temperature gradient, δ f (2) [12]. The idea distribution function written in a semiclassical approximation
behind this approach exploits the even parity in the mo- for a local energy configuration that reflects the change in the
mentum space of δ f (2) written as an iterative solution of electron energy in the presence of the electric field and the
the Boltzmann transport equation (BTE) in the relaxation- change in the Boltzmann factor in the presence of a temper-
time approximation, which is juxtaposed on the anisotropy ature gradient. The spatial validity of this approximation is
of the single-particle energy spectrum to create a nonzero established by the average distance between two successive
differential valley or spin-current effect. In the presence of collisions, which is proportional to the relaxation time. This
an electric field E and a temperature gradient ∇T , spin and limitation is imposed by the requirement that the first-order
charge currents proportional to E2 and (∇T )2 were calculated correction to the distribution function δ f (1) derived in this way

2469-9950/2019/99(24)/245204(7) 245204-1 ©2019 American Physical Society


ANIRUDDHA PAN AND D. C. MARINESCU PHYSICAL REVIEW B 99, 245204 (2019)

coincides with the solution of the Boltzmann equation linear from Hamilton’s equations of motion. Therefore, the general
in the external fields. iterative equation satisfied is
Our nonlinear transport formalism is applied to the case of
ṗ · ∇p δ fp(n−1) + ṙ · ∇r δ fp(n−1)
a two-dimensional electron gas in a quantum well with Rashba
  
and Dresselhaus spin-orbit interactions. Analytic expressions =− Pp,p δ fp(n) − δ fp(n)
 , n  1. (5)
for the nonlinear spin currents are calculated for arbitrary val- p
ues of the Rashba and Dresselhaus coupling constants in the
∂ fp0
presence of an electric field and a temperature gradient. The In equilibrium, of course, ( )
∂t coll.
= 0.
spin polarization of the currents is parallel or perpendicular
to the direction of propagation depending on the geometric A. Present theory
distribution of the applied perturbations. Their amplitude is The present theory of nonlinear spin and charge currents
proportional to the spin-orbit coupling constants and cancels discussed in Refs. [12,13] is based on deriving δ fp(2) , the
in the limit α = β regardless of the directions of the applied distribution function that is quadratic in the applied fields, by
fields and propagation. approximating the collision term in Eq. (5) for n = 2 “in the
relaxation time approximation.” Thus, generalizing for the nth
II. SECOND-ORDER DISTRIBUTION FUNCTION order,
We consider a simple two-dimensional Fermi system de- δ fp(n)
scribed by a single-particle Hamiltonian H whose eigenvalues ṗ · ∇p δ fp(n−1) + ṙ · ∇r δ fp(n−1) = − (n  1). (6)
τ
p are functions of the momentum p. Other quantum numbers,
such as spin (or valley index), might be present, but since τ is the energy-independent relaxation time, the average time
they are not relevant at this point, they will not be explicitly between two consecutive collisions.
declared. In thermodynamic equilibrium, the statistical occu- In the following considerations, we assume for simplicity
pancy of a single-particle state is established by the Fermi dis- that only an in-plane electric field is applied. Then, the first-
 − order correction to the distribution function δ fp(1) is obtained
tribution function at temperature T , fp0 = [1 + exp pkB T F ]−1 ,
from Eq. (6) when n = 1 and ṗ = −eE, in the relaxation-time
where F is the Fermi energy.
approximation as
In the presence of a time-independent perturbation,
f (p, r), function of position r and momentum p, describes ∂ fp0
the particle distribution and satisfies the general Boltzmann δ fp(1) = eτ E , (7)
∂ p
transport equation (BTE),
    where p is the projection of the electron momentum along
∂ f (p, r) ∂ f (p, r) the direction of the electric field (p⊥ is the component of the
= . (1)
∂t ∂t coll. linear momentum perpendicular to the field). This is simply
This equality expresses the conservation of the number of the well-known linear textbook solution for the BTE.
particles in a volume in the phase space, as particles leave the For n = 2, the second-order correction to the particle dis-
trajectory prescribed by Hamilton’s equations only as a result tribution function follows immediately from (6) and (7) as
of collisions. Introducing δ fp (r) as the deviation equilibrium ∂ 2 fp0
of the distribution function, a solution of the BTE is δ fp(2) = (eτ E )2 , (8)
∂ p2
f (r, p) = fp0 + δ fp (r). (2)
while the nth-order generalization is [13]
Henceforth, the explicit dependence on r of δ fp (r) will not
∂ n fp0
be declared, considering that it does not play a role in the δ fpn = (eτ E )n . (9)
collision integral, the central part of our analysis. With this, ∂ pn
the collision term can be written as Using (8), nonlinear currents are obtained by summing
  
∂ f (r, p) a generic (spin or valley-charge) velocity operator matrix
=− Pp,p [ fp (1 − fp ) − fp (1 − fp )]
∂t coll.
element ṽp multiplied by δ fp(2) over the two-dimensional
p
 momentum space. Symmetry considerations require that a
=− Pp,p (δ fp − δ fp ), (3) nonzero result is obtained only if the two terms have the same
p parity, even in the momentum. Thus,
where Pp,p is the probability of scattering between states of   ∂ 2 fp0
δj = ṽp δ fp(2) = (eτ E )2 ṽp . (10)
momenta p and p . For a scattering matrix element Vp,p (θpp ), ∂ p2
p p
which depends at most on the angle between the initial and
final momentum states θp,p , the Fermi golden rule generates When implemented for a 2D electron system with either
Rashba or Dresselhaus interactions, this algorithm leads to

Pp,p = |Vpp (θpp )|2 δ(p − p ). (4) analytic expressions predicting the existence of spin currents
h̄ whose magnitude is proportional with α or β, respectively.
When Eq. (1) is written for the deviation from equilibrium Numerical results are obtained when both spin-orbit couplings
of the distribution function in all orders in the external fields are present. In the limit of α = β, when the two spin-orbit
δ fp(n) , an additional perturbation is introduced via ṗ and ṙ interactions are equal, no spin currents are found [13].

245204-2
NONLINEAR SPIN-CURRENT GENERATION IN QUANTUM … PHYSICAL REVIEW B 99, 245204 (2019)

As we demonstrate below, however, using Eq. (6) in deriv- When one solves Eq. (5) for n = 2, the above procedure
ing a second-order correction to the particle density distribu- has to be repeated. Therefore, a second-order solution should
tion function is not appropriate since, for n  2, the collision be of the form
term in Eq. (5) cannot be approximated in the relaxation-time
δ fp(2) = C(p )(eE · ∇p )δ fp1 = C(p )τ (eE · ∇p )2 fp0 , (16)
approximation,
 where as before C(p ) is a constant, at most dependent on
  δ fp(n)
Pp,p δ fp(n) − δ fp(n)
 = , n  2. (11) the energy. With this choice, the collision integral in Eq. (5)
p
τ becomes
  
Pp,p δ fp(2) − δ fp(2)

B. The relaxation-time approximation p
to the Boltzmann equation
ν0 2π ∂ 2 fp0 ∂ 2 fp0
When Eq. (5) is written for n = 1, the first-order correction = C(p )τ (eE )2 dθp |V (θp,p )|2 − .
to the equilibrium particle distribution function satisfies h̄ 0 ∂ p2 ∂ (p )2

∂ fp0    (17)
eE · vp = Ppp δ fp(1) − δ fp(1)
 , (12)
∂p p
A straightforward calculation yields
 
where vp = ∇p p is the velocity of the particle. ∂ 2 fp0 ∂ ∂p ∂ fp0 ∂ 2 p ∂ fp0 ∂p 2 ∂ 2 fp0
= = +
Solving Eq. (12) for δ fp(1) starts by proposing a self- ∂ p2 ∂ p ∂ p ∂p ∂ p2 ∂p ∂ p ∂p2
consistent solution of the form  
∂ 2 p ∂ fp0 1 ∂p 2 p − F ∂ fp0
∂ fp0 = − tanh .
δ fp(1) = C(p )(eE · vp ) , (13) ∂ p2 ∂p kB T ∂ p 2kB T ∂p
∂p
(18)
where C(p ) is a constant, at most dependent on energy. With
this choice, the right-hand side of Eq. (12) becomes When Eq. (18) is used to evaluate the collision integral, we
   obtain
Ppp δ fp(1) − δ fp(1)

  
p
Pp,p δ fp(2) − δ fp(2)


p
2π    
= |V (θp,p )|2 δ(p − p ) δ fp(1) − δ fp(1)
 ∂ fp0 ν0 2π
h̄ p = C(p )τ (eE )2 dθp |V (θp,p )|2
∂p h̄ 0
∂ fp0 ν0 2π
= eC(p )E vp dθp,p |V (θp,p )|2 ∂ 2 p ∂ 2  p 1  p − F
∂p h̄ 0 × − − vp2 tanh
∂ p2 ∂ (p )2 kB T 2kB T
× [cos θp,E − cos θp ,E ] 

ν0 2π
× dθp |V (θp,p )|2 (cos2 θp,E − cos2 θp ,E )
= δ fp(1) dθp,p |V (θp,p )|2 (1 − cos θp,p ) 0
h̄ 0
δ fp(2)
δ fp(1) = , (19)
= . (14) τ
τ
where τ has to be the same as in Eq. (15).
The end result is possible because the scattering matrix ele-
Consequently, a second-order solution cannot be obtained
ment |Vp,p |2 is an even function of θp,p implying that the odd
in the “relaxation-time approximation” since it does not sat-
part of cos θp ,E = cos(θp,E + θp,p ) does not contribute to the
isfy consistently the right-hand side of the BTE, i.e., the colli-
angular integral. Therefore C(p ) = τ , the electron relaxation
sion integral. Any other higher-order approximation cannot be
time, same as in Eq. (6), given by
obtained in this way, either. The gist of this argument is that in
h̄ order for p Pp,p δ fp(n)  ∼ δ fp(n) , δ fp(n) has to have an angular
= ν0 |Vp,p |2 (1 − cos θpp )dθpp , (15) dependence that is given exactly by vp · E, which is true only
τ
in first order in the perturbative fields.
with ν0 the density of states at the Fermi surface. The final It is clear, therefore, that the generalization Eq. (9) is inac-
form of δ fp(1) is also obtained in a quantum-mechanical for- curate as it neglects the second term of the collision integral,
malism that evaluates the conductivity as a linear-response  (n)
 Pp,p  δ f  . In reality, this term is of the same order as
function to an electric field. It is equivalent to the renormaliza- p p
p Pp,p δ f p = δ f p /τ (assuming isotropic scattering), so
(n) (n)
tion of the electron velocity (current vertex renormalization)
its disappearance is not justified.
on account of impurity-mediated scattering [14].
It is important to accentuate the idea that the transport time
τ appears in the expression of δ fp(1) precisely because it was C. The local approximation to the particle distribution function
assumed that δ fp(1) is of the form (13) when the collision term Here we discuss a different algorithm that generates
in the BTE was calculated. the second-order distribution function in a semiclassical

245204-3
ANIRUDDHA PAN AND D. C. MARINESCU PHYSICAL REVIEW B 99, 245204 (2019)

approximation, based on the local perturbation of the single- Eq. (9),


particle distribution function induced by the external fields.
The addition of an electrostatic potential V (r) = −E · r mod- 1 ∂ n fp0
δ fpn = (eτ E · vp )n n . (23)
ifies locally the electron energy to ˜ p = p + eE · r, while n! ∂p
a temperature gradient changes the Boltzmann factor to
1/kB (T + ∇T · r). If these changes are weak compared with III. SPIN CURRENTS
the Fermi energy, the electron distribution function is just the
Fermi function written for the local energy and Boltzmann Using the second-order distribution function derived
factor, above, we calculate the quadratic spin currents driven by an
   electric field and a temperature gradient in a quantum well
p − F + eE · r −1 with arbitrary Rashba and Dresselhaus interactions.
fp (r) = 1 + exp
˜ . (20)
kB (T + ∇T · r)
Equation (20) can be expanded in a series of terms pro- A. Single-particle Hamiltonian
portional to powers of the electric field and the temperature We consider a 2D electron system in a [100] quantum well
gradient, respectively. When the linear terms are constrained described in a system of reference with the ŷ perpendicular on
to replicate the solution of the BTE, it is found that r = vp τ , a the plane. The x̂-ẑ in-plane axes are rotated by π /4 from the
result that establishes the spatial range of the approximation. usual directions such that x  [110] and z  [11̄0] [15]. This
With this input, the expansion of the distribution function choice of axes takes advantage of the existence of a privileged
becomes in-plane direction ẑ, which becomes the quantization axis of
   0 Sz when α = β. When α < 0, this system of coordinates is ro-
∇T ∂ fp
fp = fp + (eτ E · vp ) − (p − F ) τ vp ·
˜ 0 tated by π /2, such that when α + β = 0, the spin quantization
T ∂p occurs along the new ẑ axis.
  2
1 ∇T The single-particle Hamiltonian of an electron of mo-
+ (eτ E · vp )2 + (p − F )2 τ vp · mentum p = (p cos ϕ, 0, p sin ϕ), spin σ = (σx , σy , σz ), and
2 T effective mass m∗ is written as
  2 0
∇T ∂ fp p2
− 2(p − F )(eτ E · vp ) τ vp · H= − (α + β )px σz + (α − β )pz σx . (24)
T ∂p2 2m∗
  The eigenvalues of the Hamiltonian are
∇T 2
+ 2(p − F ) τ vp · − 2(eτ E · vp ) p2
T ± = ±p , (25)
   2m∗
∇T ∂ fp0 where
× τ vp · . (21)
T ∂p 
= (α + β )2 cos2 ϕ + (α − β )2 sin2 ϕ. (26)
Immediately, the second-order distribution function is
therefore The associated eigenstates are, respectively,
  
1 ∇T 2  
δ fp =
(2)
(eτ E · vp )2 + (p − F )2 τ vp · ψ+ = cos |↑ + sin |↓ ,
2 T 2 2
(27)
  2 0  
∇T ∂ fp ψ− = − sin |↑ + cos |↓ ,
− 2(p − F )(eτ E · vp ) τ vp · 2 2
T ∂p2
where
 
∇T 2 (α − β )
+ 2(p − F ) τ vp · − 2(eτ E · vp ) tan  = − tan ϕ. (28)
T (α + β )
  
∇T ∂ fp0 The Fermi energy of the system is reached in each subband
× τ vp · . (22) at maximum values of the momenta given by
T ∂p 
pF ± = 2m∗ F + (m∗ )2 ∓ m∗ (29)
Obtained in this way, δ fp(2) is not required to satisfy
the Boltzmann transport equation since its derivation is not if F > 0, the standard case in a 2D electron system.
associated with an evaluation of the collision integral. In The Fermi energy of the system is established by the
the degenerate Fermi gas approximation, the only term in number of particles since
Eq. (21) that does not cancel is the one proportional to both   2π pF ξ
1
electric field and temperature gradient. Evaluating the current n= θ (p − pξ ) = dϕ pdp
integrals requires the Sommerfeld algorithm for all of the ξ =± p
(2π h̄)2 ξ =± 0 0
remaining cases.
Moreover, the nth-order deviation from equilibrium in- m ∗ F (m∗ )2 2
= + (α + β 2 ), (30)
duced by an applied electric field is, in contrast with π h̄ 2
π h̄2

245204-4
NONLINEAR SPIN-CURRENT GENERATION IN QUANTUM … PHYSICAL REVIEW B 99, 245204 (2019)

where θ (pF − p) is the Heaviside distribution in the momen- from Eqs. (33) and (34),
tum space. Consequently, 
(eτ Eν )2  2π ∞
d p μ̄
δ jμμ̄ = dϕ dξ p ṽ (vξ ,ν )2
nπ h̄2 16kB T π 2 h̄ ξ =± 0 0 dξ ξ ,μ
F = − m∗ (α 2 + β 2 ). (31)
m∗ 
ξ − F ∂ fξ0
In a typical GaAs quantum well with particle density n = × tanh −
2kB T ∂ξ
7 × 1015 m−2 and β = 4.77 × 102 m/s [9], the condition

F > 0 is satisfied up to αmax = 2.5 × 105 m/s, a limit much (eτ Eν )2 kB T  2π d2 d p μ̄
larger than one of the highest achieved values, α = 1.6 × 104 = dϕ 2 p ṽξ ,μ (vξ ,ν )2
96h̄ 0 d ξ d ξ
m/s (or equivalently h̄α = 0.1 eV Å) [16]. The variation of ξ =±
α is possible on account of its dependence on the potential 
ξ − F
applied across the well, while β is a characteristic of the actual × tanh
2kB T ξ =F
structure since it is roughly proportional to the square inverse
 
width of the quantum well multiplied by the bulk Dresselhaus (eτ Eν )2  2π d d p μ̄
constant [9]. = dϕ p ṽ (vξ ,ν ) 2
,
96h̄ ξ =± 0 dξ dξ ξ ,μ ξ =F
The symmetrized spin-velocity operator is given by ṽˆ ij =
( ∂ pi σ j + σ j ∂∂Hpi )/2, where the lower index i corresponds to the
∂H
(35)
direction of propagation, while the upper index j indicates
where in deriving the last two lines we used the Sommerfeld
the polarization direction. Its matrix elements are evaluated
algorithm to evaluate the integrals over the energy.
immediately using Eq. (27),
Thus we obtain
vx σx + σx vx px (eτ Ex(z) )2 2π
sin2 2ϕ
ṽxx = ±| |± =± sin , (32a) δ jxz = ±(α + β )(α − β )2 dϕ
2 m∗ 48h̄ 0
2

vx σz + σx vx px (eτ Ex(z) )2
ṽxz = ±| |± = ± ∗ cos  − (α + β ), (32b) = ±π (α + β )(α − β )2 . (36)
2 m 48h̄ max(α 2 , β 2 )
vz σx + σx vz pz
ṽzx = ±| |± = ± ∗ sin  + (α − β ), (32c) Similarly,
2 m
vz σz + σz vz pz (eτ Ex(z) )2
ṽzz = ±| |± = ± ∗ cos . (32d) δ jzx = ±π (α − β )(α + β )2 . (37)
2 m 48h̄ max(α 2 , β 2 )
The spin currents are obtained by summing all the spin The polarization of these currents is positive if driven by Ex
velocities multiplied by δ fp(2) for all values of the momentum and negative if driven by Ez .
in both minibands, When α  β, both currents have the same amplitude,
 μ̄ 
δ j  = απ (eτ Eν )2 /48h̄, (38)
h̄   px μ
δ jxx = ξ sin δ fp(2) , (33a) a value that represents only about 60% of previous estimates
2 ξ =± p m∗
for the same system parameters [13]. The currents cancel
 
h̄   px when α = β, a result of the paramagnetic nature of the
δ jx =
z
ξ ∗ cos  − (α + β ) δ fp(2) , (33b) electron system that is spin-polarized along the ẑ axis by the
2 ξ =± p m
combined spin-orbit interaction, which behaves like a de facto
 
h̄   pz magnetic field of magnitude proportional with α + β [17].
δ jz =
x
ξ ∗ sin  + (α − β ) δ fp(2) , (33c) When the electric field has components along both x̂ and ẑ
2 ξ =± p m
directions, Ex and Ez , respectively, the second-order distribu-
h̄   pz tion function symmetry permits the existence of spin currents
δ jzz = ξ cos δ fp(2) . (33d) whose polarization is along the direction of propagation. We
2 ξ =± p m∗
use the general algorithm in Eq. (35) to compute δ jzz and jxx .
We obtain
 
B. Electric-field-driven spin currents π (α + β )e2 τ 2 Ex Ez α2 + β 2
δ jzz = 1− ,
In the presence of an electric field E, the second-order 12h̄ 2 max(α 2 , β 2 )
distribution function is, from Eq. (22),  
π (α − β )e2 τ 2 Ex Ez α2 + β 2
δ jxx = − 1− . (39)
1 p − F ∂ fp0 12h̄ 2 max(α 2 , β 2 )
δ fp(2) = (eτ E · vp )2 tanh − . (34)
2kB T 2kB T ∂p When β = 0, the currents are
α(eτ )2 Ex Ez
For an electric field aligned with the system axes, the spin δ jzz = −δ jxx = , (40)
currents are polarized perpendicular to the direction of prop- 24h̄
agation. In this case, the general computational algorithm of twice as large as that obtained when the field acts along a
a spin current jμμ̄ driven by an electric field Eν using δ fp(2) is, single axis. When α = β, the currents cancel.

245204-5
ANIRUDDHA PAN AND D. C. MARINESCU PHYSICAL REVIEW B 99, 245204 (2019)

FIG. 1. The spin currents expressed in units of h̄E /2e are plotted
against the ratio α/β for an intensity of the applied electric field of
104 V/m. The linear Dresselhaus coefficient β is 477 m/s.
FIG. 2. The spin currents expressed in units of h̄E /2e are plotted
against the ratio α/β for an applied temperature gradient ∇T /T of
104 m−1 . The linear Dresselhaus coefficient β is 477 m/s.
We illustrate our results in the case of a GaAs quantum
well with electron density n = 7 × 1015 m−2 , β = 4.77 × 102
m/s, and τ = 2.6 × 10−12 s [9]. The currents are plotted in
given by
units of E h̄/2e, since the remaining quantity has dimensions
of an electric conductivity, measured in −1 . Of course, the  
resulting spin-conductivity coefficient is a linear function of 1  p − F
δ fp(2) = (eEx τ )(τ ∇ z T /T ) 1 − (p − F ) tanh
the electric field, so it scales proportionally with it. In Fig. 1 kB T 2kB T
we plot the currents as a function of the α/β ratio for an
∂ fp0
electric field of 104 V/m. When α = 1.6 × 104 m/s and β = × − . (42)
4.77 × 102 m/s, the nonlinear spin-conductivity in Eq. (37) ∂p
is 3.33 × 10−9 E −1 . For an electric field of 105 V/m, it
becomes 3.33 × 10−4 −1 . In this case, the only spin currents that can be driven are
In the limit α  β, the ratio of a nonlinear spin conductiv- jzz and jxx from Eqs. (33d) and (33a). Using the same general
ity σs2 to the usual charge conductivity σc(1) = ne2 /m∗ can be algorithm as in Eq. (35), we obtain
written as, with input from Eq. (30),
 
π 1 F (α + β )(eτ Ex )(τ ∇z T /T )
σs(2) eτ E α δ jzz = − −
=η , (41) 12 2π h̄
σc(1) F + m∗ α 2 /2   
α2 + β 2 m∗ (α 2 − β 2 )2
× 1− 1 − ,
where η is a numerical coefficient equal to π 2 /48 for spin 2 max(α 2 , β 2 ) 2F
 
currents polarized perpendicular to the direction of propaga- π 1 F (α − β )(eτ Ex )(τ ∇z T /T )
tion and to π 2 /24 for those polarized along the direction of δ jx =
x

12 2π h̄
propagation. For the same system characteristics as above and   
for a field of 1 × 105 V/m, the ratios of the two conductivities α2 + β 2 m∗ (α 2 − β 2 )2
× 1− 1 − . (43)
are 2.7 × 10−2 and 5.4 × 10−2 , respectively. 2 max(α 2 , β 2 ) 2F

In Fig. 2 we plot the ẑ-direction spin-current that can


C. Thermoelectric spin currents be driven by an electric field and temperature gradient ap-
If the temperature gradient and the electric field are applied plied along perpendicular directions for the same system
along perpendicular directions, the angular symmetry of the parameters as before. In the limit of α = β, the current
second-order distribution function allows the existence of cancels.
quadratic spin currents whose polarization is parallel to the For α = 1.6 × 104 m/s and β = 4.77 × 102 m/s, the
direction of propagation. spin-conductivity for the thermoelectric effect is −2.85 ×
In a geometry with the electric field parallel to the x̂ 10−9 ∇T /T −1 . The second type of thermoelectric spin cur-
axis and the temperature gradient parallel to ẑ, δ fp(2) is rents are those driven by the quadratic temperature gradient.

245204-6
NONLINEAR SPIN-CURRENT GENERATION IN QUANTUM … PHYSICAL REVIEW B 99, 245204 (2019)

From Eq. (21), An an in-plane temperature gradient with ∇Tx and ∇Tz
components can drive spin-currents polarized along the
1 p − F ∂ fp0 direction of propagation,
δ fp(2) = (p − F )2 tanh −  
2kB T 2kBT ∂p π (α + β )e2 τ 2 ∇Tx ∇Tz α2 + β 2
δ jzz = − 1− ,
  3h̄ 2 max(α 2 , β 2 )
∇T 2 ∂ fp0  
− (p − F ) τ vp · − . (44) π (α − β )e2 τ 2 ∇Tx ∇Tz α2 + β 2
T ∂p δ jx =
x
1− . (46)
3h̄ 2 max(α 2 , β 2 )
As before, the evaluation of current integral kernels has to pro-
∂2 f 0 IV. CONCLUSION
ceed in the Sommerfeld algorithm. Since ∂pp ∼ (p − F ),
there is no contribution from the first term in (44). In a quadratic expansion of the Fermi distribution func-
Expressions for the spin-currents driven by temperature tion written for a local energy configuration, we obtain a
gradients applied along the system axes are derived following second-order correction to the distribution function that is
the general algorithm described in Eq. (35), proportional to the square values of the perturbative fields.
The range of validity of this approximation is set equal to
(α + β )(α − β )2 (kB τ ∇x(z) T )2 the distance between two electron collisions, which is pro-
δ jxz = ∓π , portional to τ , the relaxation time, such that the first-order
12h̄ max(α 2 , β 2 )
(45) correction to the function satisfies the Boltzmann transport
(α − β )(α + β )2 (kB τ ∇x(z) T )2 equation. When applied to spin transport in a quantum well
δ jzx = ∓π ,
12h̄ max(α 2 , β 2 ) with competing Rashba-Dresselhaus linear interactions, our
formalism generates analytic expressions for the nonlinear
where the spin polarization is negative when the currents are spin currents driven by electric fields and/or temperature
driven by temperature gradients applied along the x̂ direction gradients. The magnitude of the spin currents generated by
and positive when the temperature gradients are along the ẑ an electric field is found to be approximately 0.6 of previous
direction. predictions. All currents cancel in the α = β limit.

[1] J. Schliemann, Rev. Mod. Phys. 89, 011001 (2017). [9] P. J. Weigele, D. C. Marinescu, F. Dettwiler, J. Fu, S. Mack,
[2] I. Žutić, J. Fabian, and S. Das Sarma, Rev. Mod. Phys. 76, 323 J. C. Egues, D. D. Awschalom, and D. M. Zumbühl,
(2004). arXiv:1801.05657.
[3] J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki, Phys. Rev. [10] Y. Bychkov and E. I. Rashba, JETP Lett. 39, 78 (1984).
Lett. 78, 1335 (1997). [11] G. Dresselhaus, Phys. Rev. 100, 580 (1955).
[4] S. J. Papadakis, E. P. D. Poortere, H. C. Manoharan, [12] H. Yu, Y. Wu, G.-B. Liu, X. Xu, and W. Yao, Phys. Rev. Lett.
M. Shayegan, and R. Winkler, Science 283, 2056 (1999). 113, 156603 (2014).
[5] D. Grundler, Phys. Rev. Lett. 84, 6074 (2000). [13] K. Hamamoto, M. Ezawa, K. W. Kim, T. Morimoto,
[6] T. Koga, J. Nitta, T. Akazaki, and H. Takayanagi, Phys. Rev. and N. Nagaosa, Phys. Rev. B 95, 224430
Lett. 89, 046801 (2002). (2017).
[7] M. Kohda, V. Lechner, Y. Kunihashi, T. Dollinger, P. Olbrich, [14] J. Rammer, Quantum Transport Theory (Perseus Books,
C. Schönhuber, I. Caspers, V. V. Bel’kov, L. E. Golub, D. Weiss, Reading, MA, 1998).
K. Richter, J. Nitta, and S. D. Ganichev, Phys. Rev. B 86, [15] D. C. Marinescu, Phys. Rev. B 96, 115109 (2017).
081306(R) (2012). [16] R. A. Simmons, S. R. Jin, S. J. Sweeney, and S. K. Clowes,
[8] F. Dettwiler, J. Fu, S. Mack, P. J. Weigele, J. C. Egues, D. D. Appl. Phys. Lett. 107, 142401 (2015).
Awschalom, and D. M. Zumbühl, Phys. Rev. X 7, 031010 [17] D. C. Marinescu, Physica E 69, 34 (2015).
(2017).

245204-7

You might also like