ArticlePDF Available

Spin-orbit splitting of valence subbands in semiconductor nanostructures

Authors:

Abstract and Figures

We propose the 14-band $\mathbf k \cdot \mathbf p$ model to calculate spin-orbit splittings of the valence subbands in semiconductor quantum wells. The reduced symmetry of quantum well interfaces is incorporated by means of additional terms in the boundary conditions which mix the $\Gamma_{15}$ conduction and valence Bloch functions at the interfaces. It is demonstrated that the interface-induced effect makes the dominating contribution to the heavy-hole spin splitting. A simple analytical expression for the interface contribution is derived. In contrast to the 4$\times$4 effective Hamiltonian model, where the problem of treating the $V_z k_z^3$ term seems to be unsolvable, the 14-band model naturally avoids and overcomes this problem. Our results are in agreement with the recent atomistic calculations [J.-W. Luo et al., Phys. Rev. Lett. {\bf 104}, 066405 (2010)].
Content may be subject to copyright.
Spin-orbit splitting of valence subbands in semiconductor nanostructures
M. V. Durnev, M. M. Glazov, and E. L. Ivchenko
Ioffe Physical-Technical Institute of the RAS, 194021 St. Petersburg, Russia
We propose the 14-band k·pmodel to calculate spin-orbit splittings of the valence subbands in
semiconductor quantum wells. The reduced symmetry of quantum well interfaces is incorporated
by means of additional terms in the boundary conditions which mix the Γ15 conduction and valence
Bloch functions at the interfaces. It is demonstrated that the interface-induced effect makes the
dominating contribution to the heavy-hole spin splitting. A simple analytical expression for the
interface contribution is derived. In contrast to the 4×4 effective Hamiltonian model, where the
problem of treating the Vzk3
zterm seems to be unsolvable, the 14-band model naturally avoids and
overcomes this problem. Our results are in agreement with the recent atomistic calculations [J.-W.
Luo et al., Phys. Rev. Lett. 104, 066405 (2010)].
PACS numbers:
I. INTRODUCTION
As follows from the time inversion symmetry and the
Kramers theorem, the electronic states in centrosymmet-
ric systems are at least doubly-degenerate. On the other
hand, in the three-, two- and one-dimensional systems
lacking a center of space inversion, the degeneracy of free
Bloch single-electron states is removed with exception
of particular points and directions of the Brillouin zone.
The removal of spin degeneracy occurs due to spatially
antisymmetric part of the single-particle periodic Hamil-
tonian H(r,ˆ
p)with allowance for the spin-orbit interac-
tion. In terms of the effective Hamiltonian H(k)the in-
teraction appears as a spin-dependent contribution odd
in the electron wave vector k. This contribution is re-
sponsible for a number of fascinating and important ef-
fects being actively studied nowadays, see, e.g., Refs. [1–
8].
In nano- and heterostructures, the quantum confine-
ment strongly modifies free-carrier dispersion. Partic-
ularly in quantum wells (QWs), the parabolic conduc-
tion band turns into series of two-dimensional subbands
shifted, in parallel, along the energy axis. In the absence
of inversion center, the spin-orbit interaction splits each
subband with the splitting described by linear and, some-
times, cubic kterms in the 2×2 effective Hamiltonian.9,10
Because of the more complex valence band structure, the
dispersion of holes is also much more complicated than
that in the conduction band. Rashba and Sherman11
were the first to calculate the spin splitting of the top-
most heavy- and light-hole subbands in QWs grown along
zk[001] from zinc-blende lattice semiconductors by us-
ing the bulk effective Hamiltonian (four-by-four matrix)
consisting of the conventional Luttinger Hamiltonian and
spin-dependent terms of the order of k3. They imposed
the simplest conditions for the four-component hole enve-
lope wave function, ψ= 0, on the boundaries of the QW
and obtained k-linear terms in the effective Hamiltoni-
ans of two-dimensional hole subbands. Direct extension
of the procedure developed in Ref. 11 for the realistic
models of quantum confinement, particularly, including
the effects of finite barrier, deems impossible owing to
the presence of k3
zspin-dependent term in the bulk 4×4
Hamiltonian. This term makes consistent matching of
the valence-band wave functions really challenging, since
it leads to a k2
zcontribution to the velocity operator ˆvz
and, therefore, to a singularity of the flux ˆvzψat the
interface.
Here we develop the 14-band k·pmodel to calculate
spin splittings of hole subbands in QWs, which allows
us to avoid the k3
z-term problem. Moreover, we propose
additional terms in the boundary conditions for the 14-
component envelope which naturally describe the inter-
face heavy-light hole mixing arising due to anisotropy of
chemical bonds at the interfaces.12–17 The developed k·p
approach presents an independent alternative to atom-
istic calculations of the spin-orbit splittings in QWs.18,19
The paper is organized as follows: Sec. II presents the
symmetry analysis of the spin-dependent terms in the
valence band Hamiltonian, Sec. III outlines the 14-band
k·pmodel and the spin-orbit splittings in the bulk semi-
conductor as well as boundary conditions for the QW
structures; numerical results and analytical approxima-
tions are presented in Sec. IV, and Sec. V contains brief
conclusions.
II. SYMMETRY CONSIDERATIONS
A. Bulk zinc-blende-lattice crystals
We begin the symmetry analysis by reminding that,
in a bulk zinc-blende-lattice semiconductor, the expan-
sion of spin-dependent part H(3)
cof the electron effective
Hamiltonian in the conduction band Γ6starts from the
nonzero cubic term
H(3)
c=γc(σκ
κ
κ),(1)
where γcis the band parameter, σand κ
κ
κare pseudovec-
tors composed of the Pauli matrices in the coordinate
system xk[100],yk[010],zk[001] and the cubic com-
binations κx=kx(k2
yk2
z)etc. The expansion of the
4×4 effective Hamiltonian in the Γ8valence band starts
arXiv:1311.5747v1 [cond-mat.mes-hall] 22 Nov 2013
2
from the first order term H(1)
v= (4k0/3)V k, where
Vx={Jx, J 2
yJ2
z}s,{A, B}s= (AB +BA)/2,Jαare
the angular momentum matrices in the basis of spherical
harmonics Y3/2,m.20,21 Although the Tdpoint symmetry
allows this term, the coefficient k0is nonzero only due to
the k·pmixing between the valence states and the re-
mote Γ3states, it is quite small and may be neglected for
the GaAs-based systems.20,21 In the Γ8band, the cubic-
kterm contains three linearly independent contributions,
as follows,
H(3)
v=γv(Jκ
κ
κ)
+1
2δγv"a2X
α
J3
ακα+a3X
α
Vαkαk2
α1
3k2#,
(2)
where, for further convenience, two of the three coeffi-
cients are presented as products of the parameter δγv
which has the dimension of γc, γvand dimensionless fac-
tors a2/2and a3/2.20 It is noteworthy that the first term
is non-relativistic in its origin, it is symmetry-allowed for
the Γ15 band. Two terms in the second line of Eq. (2) are
relativistic and one can see that the last summand con-
tains the “dangerous” contribution proportional to k3
z.
B. Quantum well structures
In the following symmetry analysis we consider only
symmetrical QW structures grown along the crystallo-
graphic axes [001], [110] or [111] and having the point
symmetries (i) D2d, (ii) C2vand (iii) C3v, respectively.
The Cartesian coordinates are conveniently chosen along
the axes (i) xk[100], y k[010], z k[001] or x1k[1¯
10], y1k
[110], z1k[001], (ii) x2k[1¯
10], y2k[001], z2k[110], and
(iii) x3k[11¯
2], y3k[¯
110], z3k[111]. In a QW struc-
ture the Γ8valence band is split into the heavy- and
light-hole-like states. In the following we choose the ba-
sic states at the Γ-point (kxj=kyj= 0) transforming
under the symmetry operations as the Bloch functions
ψ(1)
hh ≡ |Γ8,3/2i=(XjiYj)/2,(3)
ψ(2)
hh ≡ |Γ8,3/2i=− ↑ (Xj+ iYj)/2
or
ψ(1)
lh ≡ |Γ8,1/2i= [2 ↑ Zj− ↓ (Xj+ iYj)]/6,(4)
ψ(2)
lh ≡ |Γ8,1/2i= [2 ↓ Zj+(XjiYj)]/6.
Here ,are the spin-up and spin-down two-component
columns, Xj,Yjand Zjare the periodic orbital Bloch
functions in the chosen coordinate system xj, yj, zj(j=
1,2,3), and for simplicity we omit the index jin the spin
columns.
1. Growth direction [001]
The states in the heavy-hole subbands hh1, hh3. . .
and light-hole subbands lh2, lh4. . . transform accord-
ing to the Γ6spinor representation of the point group
D2dwhereas the eigenstates lh1, lh3. . . and hh2, hh4. . .
form the bases for the Γ7representation. In the method
of invariants21, the 2×2 matrix effective Hamiltonian in
each hole subband is decomposed into a linear combina-
tion of four basis matrices. Since both direct products
Γ6×Γ6and Γ7×Γ7reduce to the same direct sum of
irreducible representations Γ1+ Γ2+ Γ5, the basis ma-
trices can be chosen common for the subbands of Γ6and
Γ7symmetries. If the basic functions of the spinor rep-
resentations are chosen in the form (3), (4), then the set
of basic matrices includes the identity matrix (Γ1rep-
resentation), pseudospin matrix σz1(Γ2representation)
and two pseudospin matrices σx1, σy1transforming as the
pair of wave vector components ky1, kx1(Γ5representa-
tion). This allows one to write the linear-kterm in the
effective Hamiltonian as
H[001]
n=β(n)
1(σx1ky1+σy1kx1) = β(n)
1(σxkxσyky),(5)
where n=hhν or lhν,ν= 1,2. . . , and β(n)
1are the
subband parameter. In Eq. (5), the effective Hamilto-
nian is presented in the two coordinate systems x, y, z
and x1, y1, z1relevant for the (001) structures. We stress
that the same form of the effective Hamiltonian for the
heavy-hole and light-hole subbands results from the spe-
cial order of the Bloch functions in Eqs. (3) and (4).
2. Growth direction [110]
Both heavy- and light-hole states transform according
to the same spinor representation Γ5of the group C2v.
Among components kx2, ky2,σαjonly kx2and σz2trans-
form according to equivalent representations. As a result,
the linear-kterm has the form
H[110]
n=β(n)
2σz2kx2,(6)
with β(n)
2being the subband parameters.
3. Growth direction [111]
The pair of functions (3) and the states hhν transform
according to the reducible representation Γ5+ Γ6of the
C3vpoint group. The direct product 5+ Γ6)×5+
Γ6) = 2Γ1+ 2Γ2does not contain the Γ3representation
which means that the k-linear splitting of the heavy-hole
states is symmetry-forbidden. The first non-vanishing
spin-dependent contribution to the heavy-hole effective
Hamiltonian is cubic in kand has the form22
H[111]
hhν =γ(ν)
1σx3ky3k2
y33k2
x3(7)
+γ(ν)
2σy3kx3k2
x33k2
y3+γ(ν)
3σz3ky3k2
y33k2
x3
3
with three independent parameters γ(ν)
1,γ(ν)
2and γ(ν)
3.
By contrast, k-linear terms are allowed in the dis-
persion of light-hole subbands. Indeed, the functions
(4) and the light-hole states lhν transform according
to the two-dimensional representation Γ4. The product
Γ4×Γ4= Γ12+Γ3contains a Γ3representation mean-
ing that the k-linear light-hole splitting is described by
H[111]
lhν =β(ν)
3(σx3ky3σy3kx3),(8)
with a single parameter β(ν)
3.
III. 14-BAND MODEL
A. Energy spectrum and spin splittings in bulk
semiconductor
The 14-band k·pmodel, called sometimes 5-level k·p
model or the 14×14 extended Kane model, displays the
full symmetry of a zinc-blende-lattice crystal and de-
scribes in detail the electron dispersion in the vicinity of
the Γpoint in materials.9,20,23–26 The model includes the
Γ8vand Γ7vvalence bands formed from the orbital Bloch
functions X,Y,Z(Γ15 representation in the coordinate
system x, y, z), the lowest conduction band Γ6cformed
from the invariant orbital function S(Γ1symmetry) and
the higher conduction bands Γ8cand Γ7coriginating from
the Γ15-symmetry orbital functions X0,Y0,Z0. For the
spinor Γ-point Bloch functions |Ni(N= 1 . . . 14), we use
the notations
|Γ6c, mi,|Γ8v, m0i,|Γ7v, mi,|Γ8c, m0i,|Γ7c, mi,(9)
where m=±1/2and m0=±1/2,±3/2, the Γ6basis is
taken in the form ↑S,↓S, and the Γ8basis is given by
Eqs. (3) and (4), the Γ7basic functions are also taken in
the canonical form.10 The model contains eight param-
eters of the 14-band model, namely, the band gap Eg,
the energy distance E0
gbetween the Γ7cand Γ6states,
spin-orbit splittings and 0of the valence and higher
conduction bands, interband matrix elements of the mo-
mentum operator
P= i ~
m0hS |ˆpx|Xi ,(10)
P0= i ~
m0hS |ˆpx|X0i,
Q= i ~
m0hX0|ˆpy|Zi ,
and, finally, the interband matrix element of the spin-
orbit interaction between the valence and higher conduc-
tion bands defined by
= 3hΓ8c, m0|Hso|Γ8v, m0i=3
2hΓ7c, m|Hso|Γ7v, mi,
(11)
TABLE I: Analytic expressions for the effective mass of an
electron in the Γ6conduction band and the Luttinger param-
eters γ1,γ2and γ3for the Γ8vband.9
m0
me
=2m0P2
~2Eg11
3
Eg+ 2m0P02
~2E0
g12
3
0
E0
g+ ∆0
γ1=2m0
3~2P2
Eg
+Q2
Eg+E0
g
+Q2
Eg+E0
g+ ∆0
γ2=m0
3~2P2
Eg
Q2
Eg+E0
g
γ3=m0
3~2P2
Eg
+Q2
Eg+E0
g
where Hso =~2([σ× ∇U(r)]p)/4m2
0c2is the spin-
orbit Hamiltonian, U(r)is the spin-independent single-
electron periodic potential, cis the speed of light, and
m0is the free-electron mass. Note that hereafter we ig-
nore the difference between the generalized momentum
operator πand the operator p=i~because the
k·(πp)correction is usually negligibly small.27 The
14×14 Hamiltonian matrix H(14)
N0Nis a sum of the diag-
onal matrix elements E0
NδN0Nand off-diagonal matrix
elements linear either in or in k. As frequently used
in the simplified multiband k·pmodels,9we ignore the
free-electron term (~2k2/2m0)δN0N, which in the case of
QW structures reduces the number of boundary condi-
tions at an interface from 28 to 14 and simplifies numer-
ical calculations.
The diagonalization of the 14-band Hamiltonian yields
the electron energy spectrum in the bulk material. Ta-
ble I summarizes four fundamental band parameters, the
electron effective mass in the Γ6conduction band and the
three Luttinger parameters for the Γ8vband, calculated
in the second order of the k·pperturbation theory. Ta-
ble II shows three different parametrizations of the 14-
band model used in Refs. 20, 25 and 28. One can see
that these parametrizations provide close values of the
conduction-band effective mass and Luttinger parame-
ters but as demonstrated below quite different values of
k-dependent spin-orbit splittings.
Following Ref. [20] we present the coefficients γcand γv
in Eqs. (1) and (2) as sums γc=γc0+δγcand γv=γv0+
a1δγv/2, respectively, which allows one to separate the
k·pthird-order contributions (γc0, γv0) from the fourth
order contributions (δγc, δγv), which include one order in
. The third order contributions were found in Ref. [20]
with the result
γc0=4
3P P 0Q∆(E0
g+ ∆0)+∆0Eg
EgE0
g(Eg+ ∆)(E0
g+ ∆0),(12)
γv0=4
3P P 0QEg+E0
g+ ∆0/2
Eg(Eg+E0
g)(Eg+E0
g+ ∆0).(13)
Note that, as compared with Eqs. (18) and (20) of
Ref. [20], we use here the different sign for Q, include
the factor i~/m0in our definition of Pand P0, and re-
fer E0
gnot to the Γ8cbut to Γ7cband. The fourth-order
4
TABLE II: Three parametrizations of k·pmodel for GaAs used in the literature. Energy gaps are given in eV, matrix elements
P,P0and Qare given in eV˚
A. Conduction band effective mass and Luttinger parameters are calculated after expressions given
in Tab. I.
Parametrization EgE0
g0P P 0Q me/m0γ1γ2γ3
Ref. [28] (I) 1.52 0.341 3.02 0.2 -0.17 9.88 0.41 8.68 0.063 8.51 2.08 3.53
Ref. [20] (II) 1.52 0.34 2.93 0.17 -0.1 10.3 3.3 6 0.062 7.51 2.7 3.4
Ref. [25] (III) 1.52 0.341 2.97 0.17 -0.061 10.31 3 7.7 0.061 8.42 2.48 3.63
TABLE III: The constants of spin-orbit splitting calculated after Eqs. (12)–(15) (in eV˚
A3), corrections to the parameters
ajcalculated after Eq. (17), and spin-orbit constants for kk[110] calculated after Eqs. (18) (in eV˚
A3) for three different
parameterizations introduced in Tab. II.
Parametrization γc0δγcγcγv0δγvδa1δa2δa3γlh γhh
(I) -2.39 -22.0 -24.4 6.65 76.7 0.883 -0.431 -0.258 96.4 13.2
(II) -13.9 -10.6 -24.5 39.5 34.7 0.396 -0.193 -0.116 79.3 5.84
(III) -16.0 -8.13 -24.1 45.7 26.9 0.645 -0.315 -0.189 80.5 8.79
correction to γcreads
δγc=4
9QP2(3E0
g+ 2∆0) + P02(3Eg+ ∆)
EgE0
g(Eg+ ∆)(E0
g+ ∆0).(14)
The fourth-order corrections δγvai/2in the Γ8valence
band are conveniently written as (a(0)
i+δai)δγv/2(i=
1,2,3), where
δγv=4
9
P2Q[∆ + 2∆0+ 3(Eg+E0
g)]
Eg(Eg+E0
g+ ∆)(∆0+Eg+E0
g).(15)
and
a(0)
1=13
4, a(0)
2=1, a(0)
3= 1 .(16)
The terms proportional to a(0)
irepresent P2Qcon-
tribution, they were derived in Ref. [20]. Equation (15)
differs from Eq. (21a) in Ref. [20] by extra in the nu-
merator and denominator. In addition to a(0)
ithere are
other contributions proportional to Q3which are dis-
regarded in Ref. [20] and can be presented in the form
δai=ci
Q2Eg
P2(Eg+E0
g),(17)
c1=41
12 , c2=5
3, c3=1.
These contributions may play an important role because
Pand Qare of the same order of magnitude. It is worth
to mention that additional contributions P02Qare
negligibly small as compared to those in Eq. (17).
For completeness, below we write down the expressions
for the coefficients of the k3splittings hh(k) = γhh k3,
lh(k) = γlh k3of the heavy- and light-hole subbands for
the particular direction kk[110]:
γhh =1
2
γv 13ξ1
p1+3ξ2!+δγv" 12ξ
p1+3ξ2!
+1
2δa1 13ξ1
p1+3ξ2!+1
8δa2 721ξ13
p1+3ξ2!
+1
4δa3 1 + ξ
p1+3ξ2!#
,(18a)
γlh =1
2
γv 1 + 3ξ1
p1+3ξ2!+δγv" 1 + 2ξ
p1+3ξ2!
+1
2δa1 1 + 3ξ1
p1+3ξ2!+1
8δa2 7 + 21ξ13
p1+3ξ2!
+1
4δa3 1ξ
p1+3ξ2!#
,(18b)
where the ratio ξ=γ32characterizes the valence band
warping. Table III summarizes the values of spin-orbit
splitting constants for the conduction and valence bands
calculated, again for different parameterizations of the
k·pmodel. It is seen that the fourth-order contribu-
tions are comparable with and can be even larger than
the third order ones. Moreover, the inclusion of the
corrections δaiin Eqs. (18) significantly increases the
spin splitting of heavy-hole states. For example, in the
parametrization (I) (Ref. 28) the omission of δaiterms
yields γhh 4.6eV˚
A3while the corrected value is larger
by almost a factor of 3.
5
B. Boundary conditions and electronic states in
QWs
Let us now apply the 14-band model to calculate the
energy spectrum in a symmetric QW grown along the
zk[001] direction. In addition to the basis (9), we
use another set of basic functions |l, si(l= 1 . . . 7, s =
±1/2), where |l, 1/2i=↑Rlfor the spin s= 1/2and
|l, 1/2i=↓Rlfor s=1/2, and Rlare the orbital
Bloch functions S,X,Y,Z,X0,Y0,Z0. This basis con-
sisting of products of the up- and down-spinors and the
orbital functions is more convenient for the formulation
and analysis of boundary conditions at the interfaces.
Within the 14-band approach each electron state Ψn,j in
a QW is described by 14 envelope functions fnj,ls in the
expansion
Ψnj =eikkρ
SX
ls
fnj,ls(z)|l, si.(19)
Here zis the growth axis, kkis the in-plane wave vec-
tor with two components kx, ky,Sis the normaliza-
tion area, the subscript ndenotes the subband, e.g.,
n=e1, hh1, lh1etc., and jis a pseudospin index enu-
merating two states in each subband ndegenerated in
the Γ-point (kk= 0). The energy spectrum Enj (kk)in
the n-th electronic subband in k-space is obtained from
the numeric solution of the Schr¨odinger equation
H(14) kk,ˆ
kzΨnj =Enj (kknj ,(20)
where kzis replaced by a differential operator ˆ
kz=
i∂/∂z acting on the envelopes fnj,ls(z).
Equation (20) should be supplemented with the bound-
ary conditions. The key requirement is associated with
the conservation of the particle flux. For a 14-component
envelope fls the flux is given by
S=1
~X
l0s0ls
f
l0s0
H(14)
l0s0,ls(k)
kfls ,(21)
or explicitly one has, e.g., for the flux x-component
Sx=i
~hPˆ
f
Xˆ
fSˆ
f
Sˆ
fX+P0ˆ
f
X0ˆ
fSˆ
f
Sˆ
fX0
+Qˆ
f
Z0ˆ
fXˆ
f
Xˆ
fZ0+ˆ
f
X0ˆ
fZˆ
f
Zˆ
fX0i ,(22)
where two-component spinor envelopes
ˆ
fl="fl,1/2
fl,1/2#
are introduced for brevity.
Since the k·pHamiltonian contains kzterms only of
the first order, it is enough to impose one condition per
envelope fnj,ls. In what follows we assume that the inter-
band matrix elements P,P0,Q, and are the same in
the QW and barrier materials, so that solely the diagonal
elements E0
N, i.e. the positions of bands at k= 0, expe-
rience discontinuities at heterointerfaces. In this case the
simplest and intuitively natural set of boundary condi-
tions conserving the flux could be merely the continuity of
all envelopes at the interface, see Eq. (22). However, such
boundary conditions do not account for the reduced mi-
croscopic symmetry of a single interface described by the
C2vpoint group and caused by the anisotropy of chem-
ical bonds in the (001) plane.12–14,29 Therefore, we are
interested in boundary conditions as simple as possible
but those which conserve the flux and make allowance
for the interface heavy-light hole mixing. We recall that
in calculations based on 4×4 Luttinger Hamiltonian and
four-component envelope Φthis kind of state mixing is
described by an extra term in the boundary conditions13
ΦA= ΦB,(23)
ˆ
M1
AΦ
∂z A
=ˆ
M1
BΦ
∂z B
+2
3
tl-h
a0m0{JxJy}sΦ,
where the matrix Mis diagonal and comprises the val-
ues of heavy-hole, mhh =m0/(γ12γ2), and light-hole,
mlh =m0/(γ1+ 2γ2), effective masses in the [001] direc-
tion, tl-his a real coefficient, and a0is the lattice con-
stant. In order to include the hole mixing effect into the
14-band k·pmodel, we provide the minimal generaliza-
tion of natural boundary conditions for envelopes fnj,ls as
the requirement of continuity of the five envelopes fnj,ls
corresponding to Rl=S,X,Y,Zand Z0and the follow-
ing discontinuity of the remaining envelopes, as follows,
ˆ
fnj,X0A=ˆ
fnj,X0B+˜
tˆ
fnj,XB,(24)
ˆ
fnj,Y0A=ˆ
fnj,Y0B+˜
tˆ
fnj,YB,
where ˜
tis a real dimensionless interface-mixing parame-
ter. One can readily see that the proposed conditions (24)
are in agreement with the flux continuity. The bound-
ary conditions for the 4×4model can be obtained from
Eqs. (24) taking into account that the Qkzoff-diagonal
matrix elements in the 14×14 Hamiltonian couple ˆ
fnj,X0
with ˆ
fnj,Yand ˆ
fnj,Y0with ˆ
fnj,X. As a result we arrive in
the linear-kzapproximation at the second boundary con-
dition (23) with the heavy-light hole mixing coefficient
tl-h2m0a0
3~2Q˜
t . (25)
In what follows we use tl-has an independent parameter
of our theory. It is worthwhile to stress that an inclusion
of other extra terms in the 14-band boundary conditions
modifies Eqs. (23) to a more complicated form of Ref. 30.
IV. RESULTS AND DISCUSSION
Figures 1 and 2 display the main results of our 14-
band model calculation, with the generalized boundary
6
k || [100] k || [110]
(b)
Spin-orbit splitting (meV)
0
0.2
0.4
0.6
0.8
Wave vector (106 cm-1)
1.5 1.0 0.5 0 0.5 1.0 1.5
k || [100] k || [110]
hh1
h+
h
e1
(d)
Spin-orbit splitting (meV)
0
0.2
0.4
0.6
0.8
Wave vector (106 cm-1)
1.5 1.0 0.5 0 0.5 1.0 1.5
k || [100] k || [110]
parametrization (I)
GaAs/Al0.35Ga0.65As
a = 100 Å
tl-h = 0
hh1
h+ (lh1)
h (hh2)
(a)
Hole energy (meV)
50
40
30
20
10
Wave vector (106 cm-1)
1.5 1.0 0.5 0 0.5 1.0 1.5
k || [100] k || [110]
GaAs/Al0.35Ga0.65As
a = 100 Å
tl-h = 0
parametrization (II)
hh1
h+
h
(c)
Hole energy (meV)
40
30
20
10
Wave vector (106 cm-1)
1.5 1.0 0.5 0 0.5 1.0 1.5
FIG. 1: Dispersion (a, c) and spin splitting (b, d) of valence subbands for GaAs/Al0.35Ga0.65 As QW. The calculations are done
for two parametrizations (see Tab. II): (I) [panels (a) and (b)] and (II) [panels (c) and (d)]. The spin splitting of conduction
subband e1is presented in (b) and (d) for comparison.
conditions (24), of valence band structure in a 100-˚
A-
thick GaAs/Al0.35Ga0.65As QW. The energy dispersion
of three topmost valence subbands is obtained neglecting,
Fig. 1, and taking into account, Fig. 2, the heavy-light
hole interface mixing. Each figure contains four panels.
Panels (a), (b) are calculated for the parametrization (I),
see Tab. II, panels (c), (d) represent the parametriza-
tion (II). Solid and dashed lines in panels (a), (c) show
the subband dispersion, while the wave vector depen-
dence of the spin splittings is depicted in panels (b) and
(d). As mentioned above we assume the same values of P,
P0,Q,,0and for the well and barrier materials,
the band gap of AlxGa1xAs solid solution is taken from
the quadratic equation31 Eg(x) = Eg(0)+1.04x+0.46x2.
The standard ratio 2/3 is used for the Γ8v- and Γ6c-band
off-sets, Evand Ec. The off-set Ec0for the higher
conduction band is chosen to equal Evsince all pro-
posed parametrizations give practically the same value
for the sum Eg+E0
gin GaAs and AlAs.
We begin the discussion from comparison of the Γ-
point positions of the valence subbands calculated by us-
ing the 14-band model and 4×4 Luttinger Hamiltonian
in the absence of interface mixing, tl-h= 0. In the four-
band Luttinger model the heavy- and light-hole envelope
functions and Γ-point energies Ehhν ,Elhν (ν= 1,2. . . )
are found from
~2
2
d
dz
1
mhh
d
dz +V(z)Φhhν =Ehhν Φhhν ,
~2
2
d
dz
1
mlh
d
dz +V(z)Φlhν =Elhν Φlhν ,(26)
where V(z)is the confinement potential determined by
the valence band offsets, and functions Φ(z)satisfy the
Bastard boundary conditions given by the Eqs. (24) at
tl-h= 0. It turns out that the Γ-point energies calcu-
lated in the 14-band and Luttinger models agree with
each other within 5% accuracy. It is seen from Figs. 1(a)
and 1(c) that the two sets of parameters result in sig-
nificantly different positions of the two lower subbands
at kk= 0 despite relatively small (.30%) difference of
7
k || [100] k || [110]
GaAs/Al0.35Ga0.65As
a = 100 Å
tl-h = 0.5
parametrization (II)(c) hh1
h+
h
Hole energy (meV)
40
30
20
10
Wave vector (106 cm-1)
1.5 1.0 0.5 0 0.5 1.0 1.5
k || [100] k || [110]
GaAs/Al0.35Ga0.65As
a = 100 Å
tl-h = 0.5
parametrization (I)
(a)
hh1
h+
h
Hole energy (meV)
50
40
30
20
10
Wave vector (106 cm-1)
1.5 1.0 0.5 0 0.5 1.0 1.5
k || [100] k || [110]
hh1
h+
h
(b)
Spin-orbit splitting (meV)
0
1
2
3
4
5
Wave vector (106 cm-1)
1.5 1.0 0.5 0 0.5 1.0 1.5
k || [100] k || [110]
(d)
Spin-orbit splitting (meV)
0
1
2
3
4
Wave vector (106 cm-1)
1.5 1.0 0.5 0 0.5 1.0 1.5
FIG. 2: Dispersion (a, c) and spin splitting (b, d) of valence subbands for GaAs/Al0.35Ga0.65As QW with account for the
interface mixing of heavy and light holes [tl-h= 0.5corresponding to ˜
t0.07 for the set (I) and ˜
t0.1for the set (II)]. The
calculations are done for two parametrizations (see Tab. II): (I) [panels (a) and (b)] and (II) [panels (c) and (d)].
the Luttinger parameters. For the parameter set (II),
these subbands are much closer in energy than for the
set (I). Moreover, the calculation carried out within the
Luttinger Hamiltonian model for the parametrization (II)
gives the crossing between the hh2and lh1states at
a95 ˚
A, see Ref. 34 for details. For the 100 ˚
A-thick
QW, the pure-state energies still lie very close to each
other, the mixing is remarkable and we use for these Γ
states in Figs. 1 and 2 the notation h+, hinstead of lh1
and hh2, respectively.
The heavy-light hole mixing is crucial both for the h±
states and the spin-orbit splitting of the valence sub-
bands. In the 4×4 effective Hamiltonian approach one
could expect to get the hh2-lh1mixing by including
the Vzˆ
k3
zterm of Eq. (2). The matrix Vzindeed mixes
the |Γ8,3/2istate with the |Γ8,1/2istate as well as
|Γ8,3/2iwith |Γ8,1/2i. The inclusion of the last term
in Eq. (2) into the effective Hamiltonian makes however
the problem unsolvable, neither rigorously nor approxi-
mately, even for the infinitely high barriers in which case
the boundary conditions reduce to vanishing of the en-
velopes at the both interfaces. Firstly, strictly speaking,
the order of differential equations increases, and addi-
tional unphysical solutions appear. Secondly, an attempt
to take the operator Vzˆ
k3
zinto account as a perturbation
encounters the problem of a nonhermitian nature of the
ˆ
k3
zoperator, hhh2|ˆ
k3
z|lh1i 6=hlh1|ˆ
k3
z|hh2i, defined on the
space of envelopes satisfying Eqs. (26) and vanishing at
the interfaces. It is worth to mention that, due to the spe-
cial reason,37 the linear-kterms in the valence subbands
hhν, lhν related to the Vzˆ
k3
zoperator and found as a first-
order correction to the energy spectrum in Ref. [11] are
finite and can be compared with the 14-band calculations,
see below. For barriers of finite height the inconsistency
related to the Vzˆ
k3
zoperator seems insurmountable for
finding both the Γ-point energies and the linear-kdisper-
sion. On the other hand, the 14-band k·pmodel under
study allows one to derive the cubic terms of Eq. (2) for
bulk materials and to compute the QW valence eigen-
states comprising an admixture of the Bloch functions
8
|Γ8,3/2iand |Γ8,1/2ior |Γ8,3/2iwith |Γ8,1/2iat
the point kx=ky= 0.
The curves in Fig. 2 are calculated for the same set of
parameters as in the previous figure, with one exception:
Now the interface mixing parameter ˜
tin the boundary
conditions (24) is nonzero and corresponds to a reason-
able value of the parameter tl-h= 0.5related with ˜
tby
Eq. (25). Comparing Figs. 1 and 2 we observe striking
effects of the interface mixing. The splitting between the
h+and hstates in Fig. 2(c) tremendously increases.
Furthermore, all the spin splittings in Fig. 2 are enhanced
by about an order of magnitude. The positions of the h+
and hstates at the Γ-point can perfectly be evaluated
in the framework of Luttinger Hamiltonian and general-
ized boundary conditions (23). In the first order in tl-h
the role of extra term in the boundary conditions (23)
can be reduced to an effective matrix element
l-h=tl-h~2
m0a0
Φhh2(zilh1(zi)(27)
that mixes the lh1and hh2states at kk= 0. Here zi
is the coordinate of the right-hand interface. The mixed
state energies are given by
Eh±=Ehh2+Elh1
2±sEhh2Elh1
22
+ ∆2
l-h.(28)
At the crossing point, where Ehh2=Elh1, the splitting
between the h+and heigenstates equals 2|l-h|and
each of them is an equal admixture of the hh2and lh1
pure states. The comparison of Figs. 1(c) and 2(c) shows
that, in parametrization II, the “bulk” hh2-lh1mixing
inside the QW layer is by an order of magnitude smaller
than that due to the interface effect.
GaAs/AlAs
a = 85 Å
parametrization (II)
parametrization (I)
hh1
pseudopotential calculations (J.-W. Luo et al.)
|β(hh1)
1| (meV Å)
0
50
100
Interface mixing parameter tl-h
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
FIG. 3: The results of calculations for heavy-hole (hh1) spin
splitting as a function of the interface mixing strength for a
GaAs/AlAs 85 ˚
A well. The dashed line indicates the result
of pseudo potential calculations obtained for the same QW in
Ref. 19.
The 14-band model developed here automatically pro-
vides the spin-orbit splitting of conduction and valence
subbands for kk6= 0. Like the energy dispersion pre-
sented in panels (a,c), the spin splittings are given for
two directions of the in-plane wavevector kkk[100] and
kkk[110]. It is noteworthy that the anisotropy of spin
splittings becomes pronounced at kk106cm2. One
can see in Figs. 1(b,d) that, even in the absence of in-
terface mixing, the k-linear spin splitting of hh1sub-
band is comparable with that for e1conduction subband
also shown (by triangles). The other notable feature is
a huge linear-in-kspin splitting of hh2and lh1(or h+
and h) subbands which is particularly pronounced for
the set of parameters (II) where these states are close in
energy. For this set of parameters the linear terms for
h±states markedly exceed those for e1electron and hh1
valence subbands. Such a behavior was uncovered by
Rashba and Sherman11 in the model of infinite barriers:
It is caused by (i) the heavy-light-hole mixing by the off-
diagonal elements of Luttinger Hamiltonian proportional
to ˆ
kz(kx±iky)and (ii) Vzˆ
k3
zterm in the bulk spin-orbit
Hamiltonian (2). The heavy-light-hole interface mixing
results in the considerable enhancement of spin splittings
of the valence subbands but has only a small effect on the
spin splitting in the e1conduction subband.32
To analyze further the effect of interface mixing, we
present in Fig. 3 the absolute value |β(hh1)
1|as a function
of tl-hfor a 85 ˚
A-thick GaAs/AlAs QW. The spin split-
ting constant β(hh1)
1in Eq. (5) vanishes at a particular
value of tl-h0.2 either 0.5 for the parametrizations (II)
and (I), respectively, where the bulk-inversion asymme-
try and interface-inversion asymmetry contributions to
the splitting cancel each other. Here, it is worth to re-
mark that the cancellation takes place for positive values
of tl-h, while for tlh<0the absolute value of the split-
ting monotonously increases with an increase of |tlh|.
The sign of this parameter lies beyond the limits of this
work. Atomistic calculation of the spin-orbit splitting
performed in Ref. [19] predicted huge values of β(hh1)
1for
the hh1subband reaching 115 meV˚
A for a GaAs/AlAs
structure with the GaAs layer thickness of 85 ˚
A, as shown
in Fig. 3 by the horizontal dashed line. For this particu-
lar structure the 14-band model yields the same value of
hh1spin-orbit splitting assuming tl-h= 1.2÷1.6. Rel-
atively high values of interface-mixing parameter meet
the expectation of a monotonous increase of tl-hwith
the content xof the heteropair GaAs/AlxGa1xAs. The
dependence of β(hh1)
1on the QW width ais shown in
Fig. 4(a). One can see that for the both sets of param-
eters the interface mixing dominantly contributes to the
spin splitting. In the selected well-width range, β(hh1)
1is
a monotonous function of a. For smaller values of the
width this coefficient reaches a maximum and then de-
creases with the increasing penetration of the wavefunc-
tion into the barriers. For completeness, the variation
of spin-splitting coefficients β(h±)
1with ain the h±sub-
bands is included in Fig. 4(a), see the inset.
9
tl-h = 0.5
tl-h = 0
parametrization (I)
parametrization (II)
|β(hh1)
1| (meV Å)
0
50
100
Well width (Å)
60 80 100 120 140
interface Eq. (30)
infinite barriers
|β(hh1)
1| (meV Å)
0
50
100
Well width (Å)
60 80 100 120 140
(a) (b)
FIG. 4: Spin-orbit k-linear term β(n)
1for the hh1subband in a GaAs/Al0.35Ga0.65 As QW. (a) 14-band numerical calculation
is shown for two sets of parameters (solid and dashed lines) and for two values of interface mixing parameter: tl-h= 0 and
tl-h= 0.5. The inset represents the results for h+and hsubbands at tl-h= 0 for the parameterization (I). (b) Analytical
calculation of β(hh1)
1. Three bottom curves are obtained in the limit of infinitely-high barriers from Eq. (8) of Ref. 11: the solid
curve represents the parametrization (I), the dotted and dashed curves are calculated for the parametrization (II) neglecting
and taking into account the corrections δaiin Eq. (17), respectively. Two top curves demonstrate the interface-induced spin
splitting according to Eq. (30) with tl-h= 0.5.
The numerical results presented above can be inter-
preted in terms of three independent contributions to
the spin-orbit splitting of the valence subbands. First
one is similar to that in the conduction band and, for the
heavy-hole subband hh1, originates from the PαJ3
ακα
and (Vxkx+Vyky)k2
zterms in the spin-orbit Hamilto-
nian for the Γ8band, Eq. (2), averaged over the size-
quantization wavefunction. The second contribution re-
sults from the interference of the Vzk3
zterm in Eq. (2)
and off-diagonal elements Hof the Luttinger Hamilto-
nian. In evaluation of this second contribution one en-
counters the “dangerous” k3
zmatrix element which can
be calculated only in the limit of infinite barriers. In this
limit, the sum of two contributions is given by Eq. (8)
of Ref. 11. The third contribution to the k-linear split-
ting of the heavy-hole subband arises from the interface-
induced heavy-light-hole mixing and becomes dominant
for |tl-h|&1. It can be evaluated within the 4-band
model using the Luttinger Hamiltonian and the bound-
ary conditions Eq. (23) and taking into account that for
tl-h= 0 the heavy-hole wave functions can be presented
as33
Ψ±3/2= Φhh1(z)|Γ8,±3/2i±i(kx±iky)Slh(z)|Γ8,±1/2i.
Here the admixture of |Γ8,±1/2istates is considered in
the first order in kk, and the function Slh is found from34
~2
2
d
dz
1
mlh
d
dz +V(z)Ehh1Slh(z) =
=3~2
m0a0γ3
d
dz s
Φhh1(z).(29)
Here Ehh1is the energy of hh1subband in the Γ-point,
Eq. (26), and as before the curly brackets assume sym-
metrization of operators. Allowance for the tl-h6= 0 in
Eq. (23) gives rise to the interface inversion asymmetry
contribution to the hh1subband, which in the first order
in heavy-light hole interface mixing reads15,35
β(hh1)
1;int =2tl-h~2
m0a0
aΦhh1(zi)Slh(zi).(30)
The results of analytical calculations of the above con-
tributions to the hh1spin splitting are presented in
Fig. 4(b). The two sets of curves are depicted: the set
of three bottom curves corresponds to the “bulk” con-
tribution calculated in the limit of infinite-barrier well,
Ref. 11, and the set of two top curves represents the
interface-induced contribution calculated after Eq. (30),
for tl-h= 0.5and the finite barriers corresponding to a
GaAs/Al0.35Ga0.65 As QW. For calculation of the “bulk”
contribution, bottom dashed and solid curves, we used
parameters aiwith inclusion of corrections Eq. (17). For
comparison, the dotted curve in Fig. 4(b) is calculated
for ai=a(0)
iaccording to Eq. (8) of Ref. 11 for the set of
parameters (II). From the bottom curves in Fig. 4(a) and
Fig. 4(b) it is clearly seen that the infinite-barrier model
strongly overestimates the “bulk” contribution to the spin
splitting found within the 14-band model for tl-h= 0.
This can be attributed mainly to (i) significantly larger
values of the ˆ
k2
zoperator averaged over the heavy-hole en-
velope Φhh1found in the infinite-barrier well compared
to the case of finite barriers, and (ii) the overestimation
of Vzk3
zeffect.
10
TABLE IV: Valence-band spin splittings for a
100 ˚
A GaAs/Al0.35Ga0.65 As QW.
n β(n)
1(meV˚
A) γ(n)
1(eV˚
A3)γ(n)
2(eV˚
A3)
tl-h= 0 hh112.2 82 31
h+13.5 153 54
h67 140 78
tl-h= 0.5hh136 55 29
h+186 412 677
h230 475 475
Similar analytical procedure can also be used to cal-
culate the interface induced k-linear spin-orbit splitting
of the h±subbands. The detailed discussion of the en-
ergy spectrum for this pair of subbands will be presented
elsewhere, here we resort to a simple resonant approxi-
mation which neglects all energy bands but h+and h.
In this case, the dominant contribution results from the
interface-induced mixing of the heavy- and light- hole
states and reads17
β(h±)
1=±23~2
m0
l-hDlh1nγ3ˆ
kzoshh2E
p(Ehh2Elh1)2+ 4∆2
l-h
,(31)
Dlh1nγ3ˆ
kzoshh2E=ZΦlh1(z)nγ3ˆ
kzosΦhh2(z)dz .
Equation (31) closely reproduces the results of numerical
calculation of β(h±)
1for parametrization (II) where the
subbands h±are particularly close in energy.
Above we have paid the main attention to the k-linear
spin splitting of valence subbands. However, it follows
from the 14-band calculations presented in Fig. 1(b, d)
and Fig. 2(b, d) that, at kk106cm1, cubic in kterms
begin to play essential role resulting in the anisotropy
of the spin splitting. The k3contribution of n-th hole
subband in [001]-grown QWs contains two independent
parameters γ(n)
1and γ(n)
236
H[001]
n=γ(n)
1σx1k3
y1+σy1k3
x1
+γ(n)
2σx1k2
x1ky1+σy1k2
y1kx1.(32)
The parameters of Hamiltonian Eq. (32) extracted from
numerical simulation of a 100˚
A-GaAs/Al0.35Ga0.65 As
QW are listed in Tab. IV. It is worth to stress that (i)
interface mixing makes a significant contribution both to
k-linear and k3terms in the valence band effective Hamil-
tonian, and (ii) the k-linear terms given by β(n)
1indeed
exceed by far the k-linear terms in the bulk valence-band
Hamiltonian.
V. CONCLUSIONS
To conclude, we have presented here the 14-band k·p
model extended to allow for the reduced microscopic sym-
metry of QW interfaces which makes it possible to calcu-
late the spin-orbit splitting of hole subbands in QWs. We
proposed a simple boundary condition, Eq. (24), which
takes into account heavy-light hole mixing at the inter-
face due to anisotropic orientation of interface chemical
bonds. Main contributions to the hole spin splitting are
identified. The developed model has been applied to cal-
culate the valence-band spin splittings in (001) QWs, but
it can be used as well for QWs of any crystallographic
orientation including the (110) and (111) orientations.
The results of numerical calculations are well described
by the developed analytical theory. Moreover, we have
demonstrated that the large values of the spin splitting
for the topmost heavy-hole subband predicted in Ref. 19
on the basis of atomistic calculations can be ascribed to
the relatively strong interface-induced mixing of heavy-
and light-hole states.
Acknowledgments
We are grateful to M.O. Nestoklon and E.Ya. Sherman
for valuable discussions.
This work was supported by RFBR, Dynasty Founda-
tion, as well as EU project POLAPHEN.
1F. Meier and B. Zakharchenya, eds., Optical orientation
(1984).
2I. ˇ
Zuti´c, J. Fabian, and S. Sarma, Rev. Mod. Phys. 76, 323
(2004).
3J. Fabian, A. Matos-Abiague, C. Ertler, P. Stano, and I.
ˇ
Zuti´c, Acta Phys. Slovaca 57, 565 (2007).
4M. I. Dyakonov, ed., Spin physics in semiconductors
(Springer-Verlag: Berlin, Heidelberg, 2008).
5Yu. Kusraev and G. Landwehr, eds., Semicond. Sci. Tech-
nol. 23, N11 (2008), special issue on Optical Orientation.
6Semiconductors 42, N8 (2008), special issue in memory of
V.I. Perel.
7L. W. Molenkamp and J. Nitta, eds., Semicond. Sci. Tech-
nol. 24, N6 (2009), special issue on effects of spin-orbit
interaction on charge transport.
8M.W. Wu, J.H. Jiang, and M.Q. Weng, Phys. Rep. 493,
61 (2010).
9R. Winkler, Spin-Orbit Coupling Effects in Two-
11
Dimensional Electron and Hole Systems (Springer, 2003).
10 E.L. Ivchenko, Optical spectroscopy of semiconductor
nanostructures (Alpha Science International, Harrow, UK,
2005).
11 E. I. Rashba and E. Y. Sherman, Phys. Lett. A 129, 175
(1988).
12 I. L. Aleiner and E. L. Ivchenko, JETP Letters 55, 692
(1992).
13 E.L. Ivchenko, A.Yu. Kaminski, and U. Roessler, Phys.
Rev. B 54, 5852 (1996).
14 O. Krebs and P. Voisin, Phys. Rev. Lett. 77, 1829 (1996).
15 L. Vervoort, R. Ferreira, and P. Voisin, Phys. Rev. B. 56,
12744 (1997); Semicond. Sci. Technol. 14, 227 (1999).
16 T. Guettler, A. L. C. Triques, L. Vervoort, R. Ferreira,
Ph. Roussignol, P. Voisin, D. Rondi, J. C. Harmand, Phys.
Rev. B 58, 10179 (1998).
17 A. A. Toropov, E. L. Ivchenko, O. Krebs, S. Cortez, P.
Voisin, J. L. Gentner, Phys. Rev. B 63, 035302 (2000).
18 B. Foreman, Phys. Rev. B 76, 045326 (2007); 76, 045327
(2007).
19 J.-W. Luo, A. N. Chantis, M. van Schilfgaarde, G. Bester,
and A. Zunger, Phys. Rev. Lett. 104, 066405 (2010).
20 G.E. Pikus, V. A. Maruschak, and A. N. Titkov, Sov. Phys.
Semicond. 22, 115 (1988).
21 E. L. Ivchenko and G. E. Pikus, Superlattices and
otherheterostructures. Symmetry and optical phenomena
(Springer, 1997).
22 L. Wang and M. W. Wu, Phys. Rev. B 85, 235308 (2012).
23 U. R¨ossler, Solid State Commun. 49, 943 (1984).
24 M. Cardona, N.E. Christensen, and G. Fasol, Phys. Rev.
B38, 1806 (1988).
25 P. Pfeffer and W. Zawadski, Phys. Rev. B 41, 1561 (1990);
Phys. Rev. B 53, 12813 (1996).
26 Loc C. Lew Yan Voon and Morten Willatzen, The
k·pMethod: Electronic Properties of Semiconductors
(Springer, 2009).
27 Z. Liu, M. O. Nestoklon, J.L. Cheng, E.L. Ivchenko, and
M.W. Wu, Phys. Solid State, 55, 1619 (2013).
28 J.-M. Jancu, R. Scholz, E. A. de Andrada e Silva, and G.
C. La Rocca, Phys. Rev. B 72, 193201 (2005).
29 M. O. Nestoklon, L. E. Golub, and E. L. Ivchenko, Phys.
Rev. B 73, 235334 (2006).
30 G.F. Glinskii and K.O. Kravchenko, Phys. Solid State 40,
803 (1998).
31 A.A. Kiselev, L.V. Moiseev, Phys. Solid State 38, 866
(1996).
32 U. Roessler and J. Kainz, Solid State Commun. 121, 313
(2002).
33 I. A. Merkulov, V. I. Perel’, M. E. Portnoi, JETP 72, 669
(1991).
34 M.V. Durnev, in press.
35 L. E. Golub, Phys. Rev. B 67, 235320 (2003).
36 X. Cartoix`a, L.-W. Wang, D.Z.-Y. Ting, and Y.-C. Chang,
Phys. Rev. B 73, 205341 (2006).
37 The spin-orbit splitting of the hh1subband contains sym-
metrized combinations hhh1|k2
z|lhνi+hl|k2
z|hh1iwhich
are meaningful for the envelope functions found in the limit
of infinite barriers.
Article
Hole qubits in germanium quantum dots are promising candidates for coherent control and manipulation of the spin degree of freedom through electric dipole spin resonance. We theoretically study the time dynamics of a single heavy-hole qubit in a laser-driven planar germanium quantum dot confined laterally by a harmonic potential in the presence of linear and cubic Rashba spin-orbit couplings and an out-of-plane magnetic field. We obtain an approximate analytical formula of the Rabi frequency using a Schrieffer-Wolff transformation and establish a connection of our model with the ESDR results obtained for this system. For stronger beams, we employ different methods such as unitary transformation and Floquet theory to study the time evolution numerically. We observe that high radiation intensity is not suitable for the qubit rotation due to the presence of high-frequency noise superimposed on the Rabi oscillations. We display the Floquet spectrum and highlight the quasienergy levels responsible for the Rabi oscillations in the Floquet picture. We study the interplay of both the types of Rashba couplings and show that the Rabi oscillations, which are brought about by the linear Rashba coupling, vanish for typical values of the cubic Rashba coupling in this system.
Article
Full-text available
Spin photocurrent spectra induced by Rashba- and Dresselhaus-type circular photogalvanic effect (CPGE) at inter-band excitation have been experimentally investigated in InGaAs/AlGaAs quantum wells at a temperature range of 80 to 290 K. It is found that, the sign of Rashba-type current reverses at low temperatures, while that of Dresselhaus-type remains unchanged. The temperature dependence of ratio of Rashba and Dresselhaus spin-orbit coupling parameters, increasing from −6.7 to 17.9, is obtained, and the possible reasons are discussed. We also develop a model to extract the Rashba-type effective electric field at different temperatures. It is demonstrated that excitonic effect will significantly influence the Rashba-type CPGE, while it has little effect on Dresselhaus-type CPGE.
Article
Silicon hole quantum dots have been the subject of considerable attention thanks to their strong spin-orbit coupling enabling electrical control, a feature that has been demonstrated in recent experiments combined with the prospects for scalable fabrication in CMOS (complementary metal-oxide-semiconductor) foundries. The physics of silicon holes is qualitatively different from germanium holes and requires a separate theoretical description, since many aspects differ substantially: the effective masses, cubic symmetry terms, spin-orbit energy scales, magnetic field response, and the role of the split-off band and strain. In this work, we theoretically study the electrical control and coherence properties of silicon hole dots with different magnetic field orientations, using a combined analytical and numerical approach. We discuss possible experimental configurations required to obtain a sweet spot in the qubit Larmor frequency to optimize the electric dipole spin resonance (EDSR) Rabi time, the phonon relaxation time, and the dephasing due to random telegraph noise. Our main findings are as follows. (i) The in-plane g factor is strongly influenced by the presence of the split-off band, as well as by any shear strain that is typically present in the sample. The g factor is a nonmonotonic function of the top gate electric field, in agreement with recent experiments. This enables coherence sweet spots at specific values of the top gate field and specific magnetic field orientations. (ii) Even a small ellipticity (aspect ratios ∼1.2) causes significant anisotropy in the in-plane g factor, which can vary by 50%–100% as the magnetic field is rotated in the plane. This is again consistent with experimental observations. (iii) EDSR Rabi frequencies are comparable to Ge and the ratio between the relaxation time and the EDSR Rabi time ∼105. For an out-of-plane magnetic field the EDSR Rabi frequency is anisotropic with respect to the orientation of the driving electric field, varying by ≈20% as the driving field is rotated in the plane. Our work aims to stimulate experiments by providing guidelines on optimizing configurations and geometries to achieve robust, fast, and long-lived hole spin qubits in silicon.
Article
We theoretically study gate-defined one-dimensional channels in planar Ge hole gases as a potential platform for non-Abelian Majorana zero modes. We model the valence band holes in the Ge channel by adding appropriate confinement potentials to the 3D Luttinger-Kohn Hamiltonian, additionally taking into account a magnetic field applied parallel to the channel, an out-of-plane electric field, as well as the effect of compressive strain in the parent quantum well. Assuming that the Ge channel is proximitized by an s-wave superconductor (such as Al) we calculate the topological phase diagrams for different channel geometries, showing that sufficiently narrow Ge hole channels can indeed enter a topological superconducting phase with Majorana zero modes at the channel ends. We estimate the size of the topological gap and its dependence on various system parameters such as channel width, strain, and the applied out-of-plane electric field, allowing us to critically discuss under which conditions Ge hole channels may manifest Majorana zero modes. Since ultraclean Ge quantum wells with hole mobilities exceeding one million and mean-free paths on the order of many microns already exist, gate-defined Ge hole channels may be able to overcome some of the problems caused by the presence of substantial disorder in more conventional Majorana platforms.
Article
Full-text available
Quantum computing offers the potential to revolutionize information processing by exploiting the principles of quantum mechanics. Among the diverse quantum bit (qubit) technologies, silicon‐based semiconductor spin qubits have emerged as a promising contender due to their potential scalability and compatibility with existing semiconductor technologies. In this paper, the latest developments of spin qubits in gate‐defined semiconducting nanostructures made of silicon and germanium, starting from the basic properties of electron and hole states in group‐IV semiconductors, are reviewed. Specifically, various nanostructures that exploit their unique microscopic properties for qubit implementations, elaborating on the advances and challenges in experiments, are discussed. Strategies for enhancing qubit performance, such as designing new nanostructures and identifying suitable operating points, particularly those involving the valleys of electron qubits and the heavy‐hole–light‐hole mixing of hole qubits, are also highlighted. This comprehensive review thus provides valuable insights into the current state‐of‐the‐art in semiconductor quantum computing and suggests avenues for future research.
Article
Hole spin qubits in group-IV semiconductors, especially Ge and Si, are actively investigated as platforms for ultrafast electrical spin manipulation thanks to their strong spin-orbit coupling. Nevertheless, the theoretical understanding of spin dynamics in these systems is in the early stages of development, particularly for in-plane magnetic fields as used in the vast majority of experiments. In this work, we present a comprehensive theory of spin physics in planar Ge hole quantum dots in an in-plane magnetic field, where the orbital terms play a dominant role in qubit physics, and provide a brief comparison with experimental measurements of the angular dependence of electrically driven spin resonance. We focus the theoretical analysis on electrical spin operation, phonon-induced relaxation, and the existence of coherence sweet spots. We find that the choice of magnetic field orientation makes a substantial difference for the properties of hole spin qubits. Specifically, we find that (i) EDSR for in-plane magnetic fields varies nonlinearly with the field strength and weaker than for perpendicular magnetic fields. (ii) The EDSR Rabi frequency is maximized when the a.c. electric field is aligned parallel to the magnetic field, and vanishes when the two are perpendicular. (iii) The orbital magnetic field terms make the in-plane g-factor strongly anisotropic in a squeezed dot, in excellent agreement with experimental measurements. (iv) Focusing on random telegraph noise, we show that the effect of noise in an in-plane magnetic field cannot be fully mitigated, as the orbital magnetic field terms expose the qubit to all components of the defect electric field. These findings will provide a guideline for experiments to design ultrafast, highly coherent hole spin qubits in Ge.
Article
We investigate the existence of linear-in-momentum spin orbit interactions in the valence band of Ge/GeSi heterostructures using an atomistic tight-binding method. We show that symmetry breaking at the Ge/GeSi interfaces gives rise to a linear Dresselhaus-type interaction for heavy holes. This interaction results from the heavy-hole/light-hole mixings induced by the interfaces and can be captured by a suitable correction to the minimal Luttinger-Kohn, four bands k·p Hamiltonian. It is dependent on the steepness of the Ge/GeSi interfaces, and is suppressed if interdiffusion is strong enough. Besides the Dresselhaus interaction, the Ge/GeSi interfaces also make a contribution to the in-plane gyromagnetic g factors of the holes. The tight-binding calculations also highlight the existence of a small linear Rashba interaction resulting from the couplings between the heavy-hole/light-hole manifold and the conduction band enabled by the low structural symmetry of Ge/GeSi heterostructures. These interactions can be leveraged to drive the hole spin. The linear Dresselhaus interaction may, in particular, dominate the physics of the devices for out-of-plane magnetic fields. When the magnetic field lies in-plane, it is, however, usually far less efficient than the g-tensor modulation mechanisms arising from the motion of the dot in nonseparable, inhomogeneous electric fields and strains.
Article
We have used a polarized spatially resolved microluminescence technique to investigate photocarrier charge and spin transport at 6 K in a GaAs nanowire (NW; n-type doping level ≈1017cm−3). Because of the difference in expansion coefficients of the NW and of its SiO2 substrate, the NW is under strain, as revealed by the splitting between light- and heavy-hole emissions in the luminescence intensity spectrum. Light valence levels lie above the heavy valence ones, which is attributed as being caused by a tensile strain along both the axial and the lateral directions of the NW, equivalent to a compressive strain in the direction of light excitation. The symmetry group of the perturbed nanowire is then lowered to C2v. No spin polarization can be evidenced for the heavy valence levels. The electron spin polarization decays up to a distance of 5µm from the excitation spot, because of spin relaxation, and stays constant for larger distances because of the increased value of the drift velocity. Remarkably, the light-hole spin polarization exhibits damped spatial oscillations over as much as 5µm. Analysis of the effect of strain on valence states shows that these oscillations are caused by the spin-orbit interaction in the light valence level. It is found that, for the C2v point group, the corresponding Hamiltonian is linear in momentum. This spin-orbit interaction causes coherent oscillations rather than a spin relaxation process since transport essentially has a drift character in the internal electric field. The equivalent effective magnetic field induced by spin-orbit interaction and strain is, taking a light-hole g factor of 1, of the order of 60mT.
Article
Nanowires (NWs) of III-V semiconductor materials have been of interest to researchers for the last two decades. Knowledge of the subband spectrum of charge carriers in NWs and NW-based structures is very important for current applications. The electronic subband spectrum in NWs is currently known in detail, while for holes it is found with significant simplifications. One or more of the following crucial features are usually neglected: the real NW cross section shape, the crystal orientation of the NW, an accounting for the real anisotropic Hamiltonian of the bulk host material, and contributions that are due to the lack of an inversion center in the crystal lattice. Here we present a detailed calculation of hole subbands in GaAs NWs with the [111] orientation with a zinc blende crystal lattice, taking into account all the above four features. The spectrum of hole subbands based on the 4×4 Luttinger Hamiltonian is numerically calculated taking into account two main contributions arising from the lack of inversion symmetry (the Td point group) in the lattice of the host crystal. Accounting for these contributions leads to the appearance of spin splitting only for some subbands, in accordance with symmetry considerations. However, a significant rearrangement also occurs in the spectrum of nonsplit subbands. The hole densities are visualized, and it is shown that the contribution of terms with Td symmetry significantly changes the structure of the multicomponent wave function. Thus, taking into account the lack of an inversion center is essential for the spectrum of hole subbands and wave functions in GaAs NWs. This can be more pronounced for NWs of III-V materials constituted by heavy elements, such as InSb, where spin-orbit interaction is stronger. The effect of a transverse electric field leading to so-called Rashba spin splitting is considered as well.
Article
Full-text available
Coherent optical spectroscopy of semiconductor nanostructures is a well-established field with several decades of history. This contribution discusses the selected topics of imaging spectroscopy, speckle analysis, and four-wave mixing. Imaging spectroscopy is getting increasingly popular due to the availability of suited detectors, and for the intuition for the ,physics behind the data provoked by a two- or three-dimensional view on the measured quantities. A general discussion on the optical imaging and detector requirements is presented, followed by examples concerning microcavity polaritons in spectral and time domain as well as in real space and reciprocal space. Speckle analysis is a linear optical technique conceived a decade ago capable of extracting microscopic properties such as dephasing from ensembles of localized optical excitations even in the presence of inhomogeneous broadening. Four-wave mixing is a well-known technique of coherent spectroscopy, which in the last decade has been improved by the introduction of heterodyne detection and spectral interferometry, enabling to investigate quantum wires and quantum dots, both in ensembles and also for individual localized transitions.
Article
Full-text available
Coherent optical spectroscopy of semiconductor nanostructures is a well-established field with several decades of history. This contribution discusses the selected topics of imaging spectroscopy, speckle analysis, and four-wave mixing. Imaging spectroscopy is getting increasingly popular due to the availability of suited detectors, and for the intuition for the physics behind the data, provoked by a, two- or three-dimensional view on the measured quantities. A general discussion on the optical imaging and detector requirements is presented, followed by examples concerning microcavity polaritons in spectral and time domain as well as in real space and reciprocal space. Speckle analysis is a linear optical technique conceived a decade ago capable of extracting microscopic properties such as dephasing from ensembles of localized optical excitations even in the presence of inhomogeneous broadening. Four-wave mixing is a well-known technique of coherent spectroscopy, which in the last decade has been improved by the introduction of heterodyne detection and spectral interferometry, enabling to investigate quantum wires and quantum dots, both in ensembles and also for individual localized transitions.
Article
Full-text available
The effect of an atomically sharp impenetrable interface on the spin splitting of the spectrum of two-dimensional electrons in heterostructures based on (001) III–V has been analyzed. To this end, the single-band Hamiltonian Γ6 for envelope functions is supplemented by a general boundary condition taking into account the possibility of the existence of Tamm states. This boundary condition also takes into account the spin-orbit interaction, the asymmetry of a quantum well, and the noncentrosymmetricity of the crystal and contains the single phenomenological length R characterizing the structure of the interface at atomic scales. The model of a quasi-triangular well created by the field F has been considered. After the unitary transformation to zero boundary conditions, the modified Hamiltonian contains an interface contribution from which the two-dimensional spin Hamiltonian is obtained through averaging over the fast motion along the normal. In the absence of magnetic field B, this contribution is the sum of the Dresselhaus and Bychkov-Rashba terms with the constants renormalized owing to the interface contribution. In the field B containing the quantizing component B z , the off-diagonal (in cubic axes) components of the g factor tensor are linear functions of |B z | and the number of the Landau level N. The results are in qualitative agreement with the experimental data.
Book
Full-text available
Springer Series in Solid-State Sciences Approx. 485 p. 140 illus., Hardcover ISBN: 978-3-540-78819-5 Written for: Libraries, scientists, graduate students
Article
Full-text available
The theory for light-hole Zeeman splitting developed in [Physica E 44, 797 (2012)] is compared with experimental data found in literature for GaAs/AlGaAs, InGaAs/InP and CdTe/CdMgTe quantum wells. It is shown that the description of experiments is possible with account for excitonic effects and peculiarities of the hole energy spectrum in a quantum well including complex structure of the valence band and the interface mixing of light and heavy holes. It is demonstrated that the absolute values and the sign of the light-hole $g$-factor are extremely sensitive to the parametrization of the Luttinger Hamiltonian.
Article
Superlattices and Other Heterostructures deals with optical properties of superlattices and quantum-well structures with emphasis on phenomena governed by crystal symmetries. After a brief introduction to group theory and symmetries, methods to calculate spectra of electrons, excitions and phonons in heterostructures are discussed. Further chapters cover absorption and reflection of light under interband transitions, cyclotron and electron spin-resoncance, light scattering by free and bound carriers as well as by optical and acoustic phonons, polarized photoluminescence, optical spin orientation of electrons and excitions, and nonlinear optical and photogalvanic effects.
Article
Introduction.- Band Structure of Semiconductors.- The Extended Kane Model.- Electron and Hole States in Quasi 2D Systems.- Origin of Spin-Orbit Coupling Effects.- Inversion Asymmetry Induced Spin Splitting.- Anisotropic Zeeman Splitting in Quasi 2D Systems.- Landau Levels and Cyclotron Resonance.- Anomalous Magneto-Oscillations.- Conclusions.- Notation and Symbols.- Quasi Degenerate Perturbation Theory.- The Extended Kane Model: Tables.- Band Structure Parameters.
Article
Experimental data on the anisotropy of electron spin resonance in a GaAs/AlGaAs quantum well have been interpreted. In has been shown that the spin-orbit interaction in quantum wells includes, in addition to the isotropic Bychkov-Rashba and anisotropic volume Dresselhaus contributions, the anisotropic contribution determined by the structure of interfaces.
Article
The spin-dependent electron-phonon scattering in the L and Γ valleys of germanium crystals has been investigated theoretically. For this purpose, the 16 × 16 k · p Hamiltonian correctly describing the electron dispersion in the vicinity of the L point of the Brillouin zone in germanium in the lowest conduction bands and the highest valence bands has been constructed. This Hamiltonian facilitates the analysis of the spin-dependent properties of conduction electrons. Then, the electron scatterings by phonons in the L and Γ valleys, i.e., intra- L valley, intra-Γ valley, inter- L-Γ valley, and inter- L-L valley scatterings, have been considered successively. The scattering matrix expanded in powers of the electron wave vectors counted from the centers of the valleys has been constructed using the invariant method for each type of processes. The numerical coefficients in these matrices have been found by the pseudopotential method. The partial contributions of the Elliott and Yafet mechanisms to the spin-dependent electron scattering have been analyzed. The obtained results can be used in studying the optical orientation and relaxation of hot electrons in germanium.
Article
The dispersion of the conduction band in GaAs is calculated using k·p models which in different ways take into account the coupling to the p-bonding and p-antibonding states. Nonparabolicity, warping and spin-splitting are accurately described up to energies about 50 meV above the conduction band minimum by the 8×8 Kane model. For higher energies a 14×14 matrix is required.