Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Electric Dipole Induced Spin Resonance in Quantum Dots

Vitaly N. Golovach, Massoud Borhani, and Daniel Loss


Department of Physics and Astronomy, University of Basel, Klingelbergstrasse 82, 4056 Basel, Switzerland
(Dated: September 26, 2018)
arXiv:cond-mat/0601674v2 [cond-mat.mes-hall] 27 Aug 2006

An alternating electric field, applied to a quantum dot, couples to the electron spin via the spin-
orbit interaction. We analyze different types of spin-orbit coupling known in the literature and find
two efficient mechanisms of spin control in quantum dots. The linear in momentum Dresselhaus
and Rashba spin-orbit couplings give rise to a fully transverse effective magnetic field in the pres-
ence of a Zeeman splitting at lowest order in the spin-orbit interaction. The cubic in momentum
Dresselhaus terms are efficient in a quantum dot with non-harmonic confining potential and give
rise to a spin-electric coupling proportional to the orbital magnetic field. We derive an effective spin
Hamiltonian, which can be used to implement spin manipulation on a timescale of 10 ns with the
current experimental setups.

PACS numbers: 73.23.Hk, 73.63.Kv

I. INTRODUCTION portant practical implications in spintronics, where spins


of extended electrons are used as a resource to accom-
Coherent manipulation of electron spin is at the heart plish information processing. In the context of quantum
of spintronics1,2 and quantum computing with spins.3 In computing, however, interest arises in the spin of a lo-
the proposal of Ref. 3, the spin of an electron confined calized electron. A natural question is, therefore, “What
to a quantum dot is used as qubit to store and process is the microscopic mechanism of EDSR in quantum dots
quantum information. A quantum register consisting of and how strong is the EDSR effect?”
an array of such spin-1/2 quantum dots is operated by a EDSR has nearly half-a-century long history. It has
set of quantum gates that act on single spins and pairs of been first observed for extended electrons in bulk semi-
neighboring spins.3 Among the simplest quantum gates conductors,12,13 and studied more recently for donor-
are the spin rotations on the Bloch sphere. With the bound electrons in Cd1−x Mnx Se14 and extended elec-
help of only a static magnetic field and an electron-spin- trons in two-dimensional electron gases15,16 and epilay-
resonance (ESR) pulse, one can change the state of the ers.17 The “forbidden” electric-dipole transition between
spin qubit at will. It is important, however, that the ESR the electron spin-up and spin-down states becomes possi-
pulse can be applied locally to each of the quantum dots, ble in the presence of spin-orbit interaction. Absorption
ensuring that the spins are accessed independently from spectra of EDSR provide information about the value of
one another. For an ESR4,5 to occur, usually, the elec- the electron g factor and the strength of the spin-orbit
tron is exposed to an alternating magnetic field of a fre- coupling. In two-dimensional electron systems, one ex-
quency ωac that matches the electron Zeeman splitting. pects the Dresselhaus spin-orbit interaction10 to be en-
However, because strong local electric fields are easier to hanced compared to bulk semiconductors, because of the
obtain than strong local magnetic fields, interest arises confinement of electron motion in one direction. Further-
in spin resonance induced by electric fields. more, the Rashba spin-orbit interaction11 arises in het-
Recently, Kato et al.6 have demonstrated three- erostructures lacking inversion symmetry, such as, e.g.,
dimensional control of spins in a GaAs/Alx Ga1−x As het- heterojunctions. In some systems, the Rashba coupling
erostructure with the use of an alternating electric field. constant can be efficiently tuned by electric fields.18
The mechanism of spin coupling to the electric field re- In quantum dots,19 the spin-orbit interaction is gener-
lies on a specially engineered Landé g tensor in the het- ally suppressed due to complete localization of electron
erostructure, achieved by modulating the Al content dur- motion.20 Typically, the quantum dot lateral size λd is
ing the MBE growth.7 The resulting g tensor is both smaller than the spin-orbit length λSO , and any effect
anisotropic and space-dependent, and allows control over of the spin-orbit interaction is suppressed as a power of
the direction and magnitude of the spin precession fre- λd /λSO and therefore is expected to be weak. This ex-
quency.6,8 A g-factor modulation resonance (g-TMR) oc- pectation contrasts with the case of electrons in quantum
curs similarly to an ESR, when the frequency of the wells, where the EDSR meets most favorable conditions.9
electric field matches the Zeeman splitting.6 Rashba and The Zeeman interaction in quantum dots plays an impor-
Efros9 have further proposed to use the standard (Dres- tant role for observing spin-orbit effects.21,22 Without the
selhaus10 and Rashba11 ) spin-orbit couplings to achieve Zeeman interaction, the Rashba and linear in momentum
an electric dipole induced spin resonance (EDSR) in Dresselhaus spin-orbit terms do not contribute to spin-
quantum wells. Rashba and Efros9 have shown that related phenomena at the first order of spin-orbit interac-
the EDSR is highly efficient in quantum wells, promis- tion. This “absence of spin-orbit” at the leading order in
−1
ing electron spin control on a timescale ωR ≃ 100 ps, quantum dots has been discussed extensively in the liter-
4
where ωR is the Rabi frequency. These results have im- ature.20,21,22 Below, we show that a similar result arises
2

y ′ [1̄10] “unperturbed” Hamiltonian (see further). In particular,


for an electric field constant in space E(r, t) = E(t), the
E(t) potential energy reads V (r, t) = eE(t) · r.
For practical applications, it is a good idea to use two
S gates, as shown in Fig. 1, because this allows larger am-
φ plitudes of E(t) to be applied to the quantum dot, while
still maintaining the dot within the same Coulomb block-
Gate 1 QD Gate 2 ade valley. Ideally, the ac voltage drop is distributed
x′ [110] between the two gates symmetrically and the dot po-
tential is kept constant by counteracting potential shifts
FIG. 1: Schematic of a setup for electric field control of spin quadratic in the electric field. For a harmonic quantum
via the spin-orbit interaction. The quantum dot (QD) con- dot, the desired ac potential reads
tains a single electron with spin S = (~/2)σ, deep in the
Coulomb blockade valley. The gates 1 and 2 are used to gen- 2
(eE(t))
erate an alternating electric field E(t), which acts via the V (r, t) = eE(t) · r + , (2)
spin-orbit interaction on the electron spin. As a result, an
2me ω02
electric dipole spin resonance (EDSR) occurs if the frequency where me is the electron effective mass and ω0 is the
of E(t) is tuned to match the Larmor frequency ωZ = EZ /~.
oscillator frequency. Then the only effect of the ac signal
on the dot confinement is shifting the dot center as a
function of time by the amount
also for the cubic in momentum Dresselhaus terms in the eE(t)
case when the dot confining potential is quadratic and r 0 (t) = − . (3)
the perturbation is linear in the electron coordinates. me ω02
In this paper, we consider the use of EDSR for control The amplitude of r 0 (t) is going to be a relevant parame-
of individual electron spins in quantum dots. We derive ter in our following analysis. Therefore, setups in which
an effective spin Hamiltonian for a quantum-dot electron, the dot can be easily moved on the substrate by gates
subject to ac electric fields. We show that there are two are particularly interesting in the context of this paper.
major mechanisms of EDSR in quantum dots. One arises We discuss the case of r0 ∼ λSO in Sec. VI, whereas for
from the linear in momentum Dresselhaus and Rashba the bulk of the paper we restrict ourself to r0 ≪ λSO .
spin-orbit couplings in combination with the Zeeman in- The Hamiltonian H0 consists of several terms,
teraction. The other arises from the cubic Dresselhaus
terms in combination with the cyclotron frequency. We H0 = Hd + HZ + HSO , (4)
estimate the strengths of both EDSR effects and compare
them to the ordinary ESR. We find that despite a strong where Hd is the Hamiltonian of a confined electron,
suppression, compared to quantum wells, the EDSR in
quantum dots is still an efficient mechanism of spin ma- p2
Hd = + U (r), (5)
nipulation and can be used alone or together with ESR 2me
−1
to achieve control of spin on a timescale ωR ≃ 10 ns.
with p = −i~∂/∂r + (e/c)A(r) being the electron mo-
mentum, c the speed of light in vacuum, and U (r) the
II. EDSR SETUP quantum dot confining potential. We restrict our consid-
eration to quantum dots with strong confinement along
one axis, such as, e.g., quantum dots defined in a two-
We consider a quantum dot containing a single elec-
dimensional electron gas (2DEG). For GaAs, the 2DEG
tron with charge −e and spin S = (~/2)σ, where σ = lies, typically, in the crystallographic plane (001) and has
(σx , σy , σz ) are the Pauli matrices. The quantum dot is in
a width d ≃ 5 nm, which ensures a strong size quantiza-
the Coulomb blockade regime with extraction (U− ) and tion along z k [001]. The in-plane motion of the electron
addition (U+ ) energies that are large compared to the
is described by the Hamiltonian (5), where r = (x, y)
temperature, so that the dot occupation remains con- is the electron in-plane coordinate; whereas the trans-
stant. An external electric field E(r, t) is applied to the
verse motion (along z) has already been integrated out in
quantum dot. In practice, E(r, t) can be generated by a Eqs. (1)-(5). In the absence of external time-dependent
pair of gates, as sketched in Fig. 1, to which an ac signal
fields, A(r) accounts for the orbital effect of a static mag-
of frequency ωac is supplied from an external circuit (not netic field B. Assuming that B is constant in space, we
shown). The Hamiltonian describing the quantum dot
have A(r) = Bz (−y/2, x/2, 0) in the symmetric gauge.
electron in the external alternating field reads Note that the in-plane components Bx and By are not
H = H0 + V (r, t), (1) present in A(r),pbecause the motion along z is strongly
quantized (d ≪ ~c/eBx(y)).
where V (r, t) = e dr · E(r ′ , t) is the potential energy
Rr ′
The magnetic field B also induces a Zeeman splitting
of the electron in the external electric field and H0 is the EZ = gµB B and a spin quantization axis n = B/B via
3

the Zeeman interaction, relativistic nature of B(t). Furthermore, the proxim-


ity of the top gates to the 2DEG decrease the displace-
1 1 ment current enclosed by the magnetic field lines pen-
HZ = gµB B · σ = EZ n · σ, (6)
2 2 etrating the quantum dot. Using Eq. (9), we estimate
B = µκc−1 Lz ∂E/∂t ≈ 10−6 G, for a quantum dot which
where g is the electron g-factor and µB is the Bohr mag-
is Lz = 100 nm below the gates plane and an electric field
neton. In GaAs, the magnitude of the g-factor is anoma-
E(t) = E0 sin(ωac t), with amplitude µκE0 = 102 V/cm
lously small (g ≈ −0.44) compared to other AIII BV
and frequency ωac /2π = 109 Hz. As we show below, a
semiconductors. The Zeeman energy is, therefore, much
much stronger effective magnetic field arises from the
smaller than the cyclotron energy ~ωc = ~eBz /me c by a
EDSR effect in the present setup, and therefore, the dis-
factor gme /m ≪ 1 (with m being the electron mass in
placement current can be safely ignored.
vacuum), for magnetic fields applied transversely to the
Recently, a sizable ESR effect has been obtained with
2DEG. In Sec. V, we derive an efficient spin-electric cou-
the help of a wire placed on top of a GaAs double dot.27
pling that is proportional to ~ωc , but present only in non-
In this case, J 6= 0 and the magnitude of B(t) is estimated
harmonic quantum dots. We remark that the magnetic
from Eq. (9) to be B = πµc−1 I/(Ly + Lz ), where I is
field is an important ingredient in our EDSR scheme,
current in the wire and 2Ly is the lateral size of the wire.
since at B = 0 no spin-electric coupling can be obtained
For Ly = Lz = 100 nm and I = 1 mA, the magnetic field
at the first order of the spin-orbit interaction (see fur-
obtained in this setup is on the order of B ∼ 10 G.
ther).
In Eq. (4), HSO stands for the spin-orbit Hamiltonian.
We start with considering the so-called “linear in p” spin-
III. SPIN-ELECTRIC COUPLING
orbit interaction,

HSO = α(px σy − py σx ) + β(−px σx + py σy ), (7) Now, we focus on the electric-field component of


the ac-signal and show that, together with the spin-
which is the sum of the Rashba (α)11 and 2D Dresselhaus orbit interaction HSO and Zeeman splitting HZ , it suf-
(β)10,23 spin-orbit interactions. This type of spin-orbit fices to generate a sizable EDSR field in the quan-
interaction gives rise to a sizable phonon-induced spin tum dot. For simplicity, we set B(t) → 0 from now
relaxation rate 1/T1 ,21,22 of the same order of magnitude on and choose A(t) = 0. As a result, we retain
as experimentally measured.24,25,31 Moreover, in the 2D only a constant in space and time magnetic field B =
limit the linear in p spin-orbit interaction is dominant, B(cos ϕ sin ϑ, sin ϕ sin ϑ, cos ϑ) and an alternating electric
because β ∝ 1/d2 . field E(r, t) = (1/e)∇V (r, t). Thus, we consider further
In the standard EDR scheme, an alternating magnetic the Hamiltonian in Eq. (1), assuming that H0 is time-
field is generated by a current in a nearby conductor. In independent and describes the dot in the absence of ac
our setup (see Fig. 1), no charge flow is ideally present fields.
between the gates. However, the alternating electric field We aim at diagonalizing H0 using a Schrieffer-Wolff
E(r, z, t) gives rise to a displacement current, with the transformation, similar to Ref. 21. We thus look for
current density a transformation matrix S such that the transformed
Hamiltonian H̃0 = exp(S)H0 exp(−S) is fully diagonal,
κ ∂E(r, z, t) see Appendix A. At B = 0, the ground state of H0 (and
JD (r, z, t) = , (8)
4π ∂t also of H̃0 ) is a Kramers doublet, because the spin-orbit
interaction is always time-reversal symmetric (at B = 0).
where κ is the electric permittivity. The external mag-
We therefore choose to encode the qubit into the ground
netic field B acquires, thus, an ac-component, B →
state Kramers doublet of the quantum dot. Owing to the
B + B(t), where B(t) = ∇ × A(t). The vector potential
mixed spin and orbital nature of the states an alternat-
A(r, z, t) is obtained as usual from Ampere’s law26
ing potential V (r, t), such as in Eq. (1), couples to the
4πµ qubit. We proceed to derive this coupling, by calculat-
∇2 A = − (J + JD ) , (9) ing the transformation matrix S at the leading order of
c
spin-orbit interaction,
where µ is the magnetic permittivity and J is the charge
flow density (in our case J = 0). In Eq. (9), we adopted 1−P 2
S= HSO + O(HSO ), (10)
the Coulomb gauge ∇ · A = 0 and used the notation L̂d + L̂Z
∇ = (∂/∂r, ∂/∂z).
The magnetic field B(t) couples to the electron spin via where L̂d and L̂Z are Liouville superoperators, i.e.
the Zeeman interaction in Eq. (6) [with B → B + B(t)], L̂d A = [Hd , A] and L̂Z A = [HZ , A], ∀A. The projec-
giving rise to an ESR source, which can be used, in tor P projects onto the diagonal (or degenerate) part of
principle, for spin manipulation in quantum dots. How- the Hilbert space of Hd + HZ , which ensures applicabil-
ever, the amplitude of B(t) is, in practice, extremely ity of “non-degenerate” perturbation theory. The cou-
small; it is proportional to 1/c, as expected from the pling of spin to electric fields is then found by applying
4

the same Schrieffer-Wolff transformation to the potential The amplitude of the resulting effective magnetic field
V (r, t). We obtain the following effective Hamiltonian due to EDSR is found from Eq. (14) to be
for our qubit in the presence of an alternating potential
V (r, t), to leading order in the spin-orbit interaction, δB0 = 2B × Ω0 . (17)

Heff = HZ + hψ0 |[S, V (r, t)]|ψ0 i, (11) The maximal amplitude is obtained for B ⊥ Ω0 , which in
experiment can easily be arranged for by, e.g., choosing
where S is the transformation matrix in Eq. (10) and B k z. In-plane magnetic fields can also be used, pro-
|ψ0 i is the quantum dot ground state. For a quantum dot vided E(t) is linearly polarized. For example, an elec-
with a harmonic confining potential, U (r) = me ω02 r2 /2, tric field E(t) aligned with x′ generates, according to
the transformation matrix S was calculated in Ref. 28 to Eq. (15), a dimensionless field Ω(t) along y ′ . In this case,
all orders of the Zeeman interaction and the first order of B should be chosen along x′ for maximal spin-electric
the “linear in p” spin-orbit interaction (7). For simplicity, coupling.
we consider here only the linear in B terms, Using our previous estimate for Ω0 ∼ 10−3 , we ob-
tain from Eq. (17) that δB0 ∼ 1 mT for a magnetic field
EZ B = 1 T oriented transversely to Ω0 . In principle, the
S = iξ · σ − [n × ζ] · σ, (12)
me ω02 dimensionless factor Ω0 can be increased up to Ω0 ∼ 1.
ξ = (λ−1 ′ −1 ′ However, this requires a specially designed setup, where
− y , λ+ x , 0),
the value of the electron displacement r0 in Eq. (3) ap-
ζ = (λ−1 ′ −1 ′
− ∂/∂y , λ+ ∂/∂x , 0), proaches the spin-orbit length λSO .
Next we remark that Ω(t) in Eq. (15) can be written
where λ± = ~/me (β ± α) are the spin-orbit lengths, and
by the order of magnitude as Ω(t) ∼ r0 (t)/λSO . More
the vectors
√ ξ and in the coordinate frame x′ =
ζ are given √
rigorously, we rewrite Eq. (15) in the following form
(x + y)/ 2, y = −(x − y)/ 2, and z ′ = z (see Fig. 1).

The first term in Eq. (12) commutes with scalar poten- X


λ−1

Ωi (t) = SO ij r0j (t), (18)
tials and therefore drops out in Eq. (11). More generally,
j
for arbitrary confining potential, the first term is replaced
by i(1 − P)ξ · σ, resulting nevertheless in no coupling of where λ−1

SO ij is a tensor of inverse spin-orbit lengths,
spin to electric fields. The second term in Eq. (12), how-
ever, allows us to express the coupling of spin to charge 
0 1/λ−

λ−1

via the electric field E(t) = (1/e)hψ0 |∇V (r, z, t)|ψ0 i that SO ij = , (19)
1/λ+ 0
acts on the quantum dot electron. For the harmonic con-
fining potential, we obtain with 1/λ± = me (β ± α)/~ and the frame (x′ , y ′ ) was
1 1 used to represent the tensor. For order of magnitude
Heff = gµB B · σ + h(t) · σ, (13) estimates, it is useful to introduce the scalar
2 2
h(t) = 2gµB B × Ω(t), (14) 1 1
= √ kλ−1 k, (20)
Ω(t) =
−e
λ−1 Ey′ (t), λ−1
 λSO 2 SO
+ Ex′ (t), 0 . (15)
me ω02 −
where kλ−1SO k is the Frobenius norm of (λp
−1
SO )ij . In the
The dimensionless field Ω(t) describes a combined effect case of Eq. (19), we have 1/λSO = (me /~) α2 + β 2 .
of the spin-orbit interaction and electric fields (or more Despite the fact that Eq. (18) was obtained consider-
generally potential fluctuations) on the qubit. Ω(t) was ing the harmonic confining potential as an example, its
calculated in Ref. 21 for the phonon potential and in generality suggests that it should remain valid for quan-
Ref. 28 for the shot-noise of a QPC nearby the quantum tum dots of arbitrary confinement, provided, to a good
dot. In our case, Ω(t) is merely a classical driving field approximation, the ac-signal merely displaces the quan-
generated by the ac-signal. tum dot parallel to itself by a vector r0 (t) as a function
Considering further a constant in space (at least of time. Note that r0 (t) is the only available parameter
on the scale of the quantum dot), alternating elec- to be compared with λSO in the limit of strong confine-
tric field E(t) = E 0 sin(ωac t) of amplitude E 0 = ment (λd → 0). We extend the class of Hamiltonians
E0 (cos φ, sin φ, 0), where φ is the angle of E 0 with re- considered here to any combination of confinement and
spect to the axis x′ (see Fig. 1), we obtain explicitly ac-voltage potential that can be rewritten in the form
Ω(t) = Ω0 sin(ωac t), with
U (r) + V (r, t) = U (r − r0 (t)) + V0 (t), (21)
−eE0 −1
λ− sin φ, λ−1

Ω0 = 2 + cos φ, 0 . (16) where V0 (t) is independent of r. We note that, as be-
me ω 0
fore, the electron wave function extension λd is assumed
To give an estimate for the amplitude Ω0 in GaAs quan- to be small compared to the spin-orbit length λSO at
tum dots, we assume λ+ ≈ λ− ≈ λSO = 8 µm, ~ω0 = each moment in time. Equation (21) need not be satis-
1 meV, and E0 = 102 V/cm, which yields Ω0 ∼ 10−3 . fied exactly. Note that λd enters only in the definition of
5

r0 (t) and does appear alone as a parameter in Eq. (18). where δ = (ωZ − ωac )n gives the detuning from res-
Therefore, defining r0 (t) as the average electron position, onance. The relaxation tensor Γ̃ij is diagonal, with
2
r0 (t) = r |ψ(r, t)| d2 r, we expect Eq. (18) to be valid
R
Γ̃XX = Γ̃Y Y = 1/T2 and Γ̃ZZ = 1/T1 , and Υ̃i assumes
to leading order also when the electron probability den- Υ̃i = Γ̃ij SjT . Here, T1 and T2 are the relaxation and de-
2
sity |ψ(r, t)| changes shape, but the dot size changes coherence times in the absence of driving measured in ex-
weakly. periment,24,25,31 and S T = −(ng/2|g|) tanh(EZ /2kB T )
Equations (13), (14), and (18) form the basis of EDSR is the thermodynamic value of spin, with T being the
in quantum dots and can be used to efficiently manip- temperature.
ulate the electron spin by electrical gates. Finally, we The time-evolution of S̃ in Eq. (27) is simplest in a
remark that Eqs. (13)-(15) have been derived under the coordinate frame (X ′ , Y ′ , Z ′ ), with Z ′ k (δ + ωR ), and
assumption r0 ≪ λSO , and therefore, can be used only reads
for Ω0 ≪ 1. In Sec. VI we discuss the case of Ω0 ∼ 1 in
more detail. A further assumption in deriving Eqs. (13)-
q 
0 −t/T̃2 2 2
(15) was that the frequency spectrum of E(t) lies well S̃X (t) = S⊥ e
′ sin t δ + ωR + φ ,
below the size-quantization energy ~ω0 . This adiabatic- q 
ity constraint is generic to the spin-based quantum com- 0 −t/T̃2 2 2
S̃Y (t) = S⊥ e
′ cos t δ + ωR + φ ,
putation2,3 ; it guarantees that the electron is not excited
to higher in energy orbital levels. S̃Z ′ (t) = S̃T + (SZ0 ′ − S̃T )e−t/T̃1 , (28)
0
where S⊥ , SZ0 ′ , and φ give the initial spin state, hS(0)i ≡
IV. SPIN DYNAMICS AND COHERENCE 0 0
S̃(0) = (S⊥ sin φ, S⊥ cos φ, SZ0 ′ ), in the coordinate frame
′ ′ ′
(X , Y , Z ). Furthermore, the decay times T̃1 and T̃2
The electron spin obeys the Bloch equation29,30 read
hṠi = [ωZ + δω(t)] × hSi − ΓhSi + Υ, (22)  2 2

1 1 δ ωR
where ωZ = gµB B/~ is the Larmor spin-precession fre- = 2 2 + ,
T̃1 δ + ωR T1 T2
quency and δω(t) = h(t)/~. The spin relaxation tensor  2
2δ 2 + ωR
2

Γij and the inhomogeneous part Υi are due to the envi- 1 1 ωR
= 2) + . (29)
ronment and can be derived microscopically21,29 within T̃2 2(δ 2 + ωR T1 T2
the Born-Markov approximation. Strictly speaking, Γij
and Υi in Eq. (22) depend also on the driving. In particu- The stationary value of spin S̃T := S̃(t → ∞) to lead-
lar, Γij acquires, in general, a time-dependent part. How- ing order reads
ever, we neglect these effects here since the energy scales
are well separated. Indeed, from experiments24,25,31 and g (δ + ω R )δ
S̃T = − 2 tanh (EZ /2kB T ) . (30)
theory,21,22 we infer that Γij , Υi ∼ (102 − 106 ) s−1 , i.e. 2|g| δ 2 + (T1 /T2 )ωR
they are very small, so that the regime Γij , Υi ≪ δω ≪
ωZ holds. In this regime, the rotating wave approxima- Note that at resonance (δ = 0), the right-hand side in
tion30 is valid. We consider a completely general driving Eq. (30) vanishes. Therefore, in the vicinity of resonance,
field S̃T is determined by the subleading order term, which
δω(t) = δωa sin(ωac t) + δωb cos(ωac t), (23) can be obtained from Eq. (30) by replacing the numerator
(δ + ω R )δ → (1/T2 )[ωR × n]. Measurement of S̃T in the
which can be realized in practice by implementing two presence of driving provides information about the spin
independent electric fields at the quantum dot site. This lifetimes T1,2 . For instance, at resonance the relaxation
is, however, by no means necessary for our EDSR scheme. time T1 can be accessed at the leading order of Γij /ωR ≪
The Rabi frequency then reads 1,
1
ωR = (δωa × n − [δωb × n] × n) . (24) g ωR × n
2 S̃T (δ = 0) = − 2 tanh (EZ /2kB T ) . (31)
|g| 2T1 ωR
Here, we assume that ωac is not far from resonance, i.e.
|ωac − ωZ | < ωac /2. In a coordinate frame (X, Y, Z) with
Finally, we estimate the Rabi frequency ωR using
Z k B, the spin dynamics is approximated as follows
Eq. (24) and the parameters from Sec. III. For Ω0 ∼
hS± (t)i ≈ S̃± (t)e±iωac t (25) 10−3 , |g| = 0.44, and B = 10 T we obtain ωR ∼ 108 s−1 .
We conclude that, with the present quantum-dot setups,
hSZ (t)i ≈ S̃Z (t), (26)
EDSR enables one to manipulate the electron spin on
where S± = SX ± iSY . The spin S̃(t) obeys a simpler a time scale of 10 ns, which is considerably shorter than
(static) Bloch equation the spin lifetimes, for which values between 1 and 150 ms
(depending on the applied magnetic field) in gated GaAs
S̃˙ = (δ + ωR ) × S̃ − Γ̃S̃ + Υ̃, (27) quantum dots have been reported recently.24,31
6

V. p3 -DRESSELHAUS TERMS thus leave out L̂Z in Eq. (10) and consider a confining
potential U (r) that differs from a harmonic one by a
Next we consider the so-called p3 terms of the Dressel- function W (r),
haus spin-orbit interaction23 , X
U (r) = uij ri rj + W (r) ≡ UH (r) + W (r), (38)
γ ij
HSO = (py px py σx − px py px σy ) , (32)
2
p where uij are real coefficients and W (r) = O(r3 ). While
where γ = αc / 2m3e Eg is the spin-orbit coupling con- in general W (r) need not be small compared to HH =
stant, with αc (≈ 0.07 for GaAs23 ) being a dimensionless p2 /2me + UH (r), in the following we expand the denom-
constant defined in Ref. 32 and Eg the band gap. For inator of Eq. (10) in terms of W ≪ HH , considering
simplicity, we impose here the dipolar approximation for therefore only small deviations of U (r) from harmonic
the ac-signal, potentials. Then, using Eqs. (11), (13), and (33), we
Z r obtain at leading order in ωc
V (r, t) = e dr′ · E(r ′ , t) ≈ eE(t) · r. (33)
∂ 2 SH

0 hi (t)
= eE(t) · hψ0 |[R(r), ]|ψ0 i , (39)
ωc ∂ωc ∂σi ωc =0
Quite remarkably,
P if the quantum dot potential is har-
monic, U (r) = ij uij ri rj , then the spin does not cou- where we set ωc → 0 in the right-hand side of Eq. (39)
ple to E(t) at the first order of HSO and zeroth order after evaluating ∂SH /∂ωc in the symmetric gauge, with
of EZ . Indeed, the second term in Eq. (11) vanishes for SH defined as [HH , SH ] = HSO . The linear relation-
V (r, t) = eE(t) · r and S = L̂−1 d HSO because of the ship between
p hi and ωc holds for ωc ≪ ω0 , where
following two identities ω0 ≡ 2 det(u)/me Tr(u). In Eq. (39), the perturbation
W (r) enters via the function R(r) defined as follows
hψ0 |[L̂−1 −1
d HSO , r]|ψ0 i = hψ0 |[L̂d r, HSO ]|ψ0 i, (34)
 ∂W (r)
[L̂−1
X
d r, HSO ] = 0, ∀ HSO (p). (35) Ri (r) = u−1 ij ∂r
. (40)
j j
The latter is specific to Hd in Eq. (5) with a harmonic
U (r), for which the operator L̂−1d r can be expressed via
Note that hR/ri ∼ W0 λd /~ω0 λW is the small parameter
the components of p − (e/c)Bz × r. Note that, generally, of our expansion in terms of W (r), with W0 and λW .
[p, HSO ] = (e/c)[Bz ×r, HSO ] for any HSO that is a func- λd being, respectively, the characteristic amplitude and
tion of only p = (px , py ). Thus, for a harmonic confining length scale of the variation of W (r) over the quantum
potential, one is left with the same dominant mechanism dot size. It is important to note that the antisymmetric
as considered above for the ”linear in p” terms. Ex- part of W (r) drops out in Eq. (39) because HSO is also
panding in terms of the Zeeman interaction, we recover antisymmetric with respect to r → −r.
Eqs. (13) and (14) with Ω(t) given now by Next, as an example, we consider UH (r) = me ω02 r2 /2
and W (r) = ηr4 , and obtain
e X −1 
Ωi (t) = − λSO ij Ej (t), (36)
me ω02 j 1 eγη~2 ωc
h(t) · σ = (Ey (t)σx + Ex (t)σy ) . (41)
2 9me ω04
me ∂ 2 HSO
λ−1

SO ij = hψ0 | |ψ0 i, (37)
~ ∂σi ∂pj Here, we have used the deformation quantization the-
ory,33 which allowed us to considerably simplify the
where (λ−1SO )ij is a tensor of inverse spin-orbit lengths, derivation of Eq. (41) by performing most of the calcu-
and as before we consider U (r) = me ω02 r2 /2. For HSO lation in classical mechanics and only at the final stage
in Eq. (32), we obtain explicitly (λ−1 1 2
SO )ij = 4 γω0 me δij come back to quantum mechanics. We have also carried
and Ωi (t) = −γeme Ei (t)/4ω0 . To estimate the strength out a fully quantum derivation of Eq. (41) and recovered
of the resulting EDSR, we note that γ ∼ βd2 /~2 , and the same result.
therefore the amplitude of h(t) = 2gµB B × Ω(t) is down To estimate the strength of the resulting EDSR, we
now by a factor d2 /λ2 ≪ 1 compared to the p-terms. note that h ∼ ~ωc (λd /λSO )(eλd E0 /~ω0 )hR/ri, where
Next we consider a quantum dot with non-harmonic λSO = 4/γω0m2e (≈ λ2d /[0.01 nm] for GaAs) is the
potential U (r) and show that the p3 -terms in Eq. (32) spin-orbit length of the p3 -terms and the parameter
give rise to a spin-electric coupling proportional to the hR/ri ∼ W (λd )/~ω0 characterizes the deviation of the
cyclotron frequency ωc = eBz /me c. Since ~ωc differs quantum dot confinement from harmonic. In practice,
parametrically from EZ (EZ /~ωc = gme B/2mBz ), the hR/ri can be as large as unity, but here we assume
p3 -terms can be as significant as the p-terms, provided hR/ri ∼ ηλ4d /~ω0 = 0.1. For an electric field with am-
EZ /~ωc . d2 /λ2 , which is realistic for GaAs quantum plitude E0 = 102 V/cm and a GaAs quantum dot with
dots. Note that for the p-terms no spin-electric coupling ~ω0 = 1 meV, we obtain the equivalent of an ac mag-
proportional to ωc arises at the first order of HSO . We netic field δB(t) = h(t)/gµB that has an amplitude
7

δB0 ≈ 1 mT at Bz = 1 T and |g| = 0.44. In contrast valid only for ac electric fields E(t) that oscillate around
to the previous mechanism, δB(t) can have here also zero, whereas Eq. (42) holds also for static B-fields. The
a finite longitudinal component δBk (t) = n(n · δB(t)), reason why a static electric field E cannot be used in
which however vanishes if B k z. Eq. (44) will become clear after we explain the origin of
Finally, we note that the p3 -terms can also be relevant ν in Eq. (44).
for spin relaxation in quantum dots with non-harmonic The spin-electric moment ν arises because the dipolar
confining potential. Of course, the magnetic field has transitions in the quantum dot become allowed, e.g., for
to have an out-of-plane component for this spin-electric the ground state
coupling to dominate over the one considered in Sec. III.
hψ0↑ |r|ψ0↓ i 6= 0. (45)

VI. DISCUSSIONS The electron charge density operator ρ(r) = −eδ(r −


rel ), where rel is the electron coordinate, acquires spin-
dependent terms in the transformed basis,
The coupling of spin to electric fields that we have de-
rived above can be used in a variety of ways to access and |ψns i = e−S |ψn i|χs i, (46)
manipulate the electron spin in experiments. The effec-
tive Hamiltonian in Eq. (13) has the same form as the ρ̃(r) = eS ρ(r)e−S , (47)
Hamiltonian of an ESR effect. This shows that ESR and
where e−S is the transformation used in Section III and
EDSR are mutually interchangeable and the choice of the
studied in detail in Appendix A. One can present ρ̃(r)
effect to be used depends on the particular experimental
as a sum of two terms,
setup. In GaAs quantum dots, the spin-orbit interac-
tion is weak enough to ensure long coherence times and, ρ̃(r) = ρ̄(r) + δρ(r), (48)
at the same time, strong enough to allow room for spin
manipulation on an experimentally accessible timescale where ρ̄(r) is spin independent and δρ(r) is proportional
of ∼ 10 ns. Much shorter timescales can be achieved to the spin. Then the spin-electric moment can be writ-
in InAs quantum dots, because of a much stronger spin- ten as follows,
orbit coupling and a larger electron g-factor. In con- Z
trast, the EDSR effect is of little use in materials with ν = rδρ(r)dv, (49)
very weak or nearly absent spin-orbit interaction, such
as, e.g., carbon-nanotube quantum dots. To make order-
of-magnitude estimates easier, we draw an analogy be- where dv is the elementary volume of integration. Equa-
tween the EDSR and ESR effects in terms of the partic- tion (49) unveils the physical meaning of the spin-electric
ular way the B and E fields couple to the electron spin moment ν: due to the mixed spin and orbit nature of the
S = (~/2)σ. electron density, the electron spin couples to the first mo-
We recall that the ESR effect occurs as a result of ment (dipole moment) of the electron.
the Zeeman interaction of the electron spin with an ac While oscillating around an equilibrium position, the
magnetic field. It is convenient to write this interaction electron produces a time-dependent dipole moment, part
in the form of a magnetic dipole interaction, of which is proportional to the electron spin. Obviously,
in a static electric field, one can set to zero the electron
HESR = −µ · B(t), (42) dipole moment, because the new electron position can
be taken as the equilibrium one. Therefore, only the
where B(t) is the ac magnetic field and µ is the electron change of the moment as a function of time has a physical
magnetic moment, meaning for single electrons (as for any other monopoles).
In contrast, the electron magnetic moment µ couples to
1
µ = − gµB σ, (43) static magnetic fields, because one can view µ as arising
2 from a pair of Dirac monopoles of opposite signs, for
where g is, in general, a tensor, see Eq. (A18). which the relative distance between them has an absolute
By analogy with the ESR effect, the EDSR effect can meaning.
be viewed as arising from an interaction between the ac Equation (49) is written in a very general (operator)
electric field E(t) and a spin-electric moment ν. The form. After taking the expectation value in the orbital
respective spin-electric interaction is then analogous to ground state |ψ0 i, we obtain
Eq. (42) and reads
1
ν = − ν̄¯σ, (50)
HEDSR = −ν · E(t), (44) 2

where the spin-electric moment ν is due to an inter- ν̄¯ij = 2e hψ0 |eS ri e−S |ψ0 i, (51)
∂σj
play between the spin-orbit interaction and some time-
reversal breaking interaction, such as the Zeeman inter- where the derivative with respect to σj is defined as a
action. This analogy is not complete. Equation (44) is usual derivative of an expression that is linear in σ. Note
8

that Eq. (50) is analogous to Eq. (43) where the role of The third limitation of our theory is a small amplitude
gµB is played by the tensor ν̄¯. Using Eqs. (13)-(15) we of oscillation of the quantum dot, r0 /λSO ≪ 1. We have
obtain for the linear in momentum spin-orbit interaction, shown in Section III that the EDSR effect is proportional
to this small parameter. Thus, for breaking the time-
2egµB reversal symmetry by the Zeeman interaction we have
ν̄¯ij = − εjkl Bk λ−1

SO li . (52) (by order of magnitude)
me ω02
Similarly, for the p3 Dresselhaus terms we obtain from r0
ωR ∼ ωZ , (55)
Eq. (41) λSO
2eγη~2 ωc 0 1
 
¯ where ωZ = EZ /~. Similarly, for breaking the time-
ν̄ij = , (53)
9me ω04 1 0 reversal symmetry by the orbital B-field effect (Sec-
tion V), we have
where we use the coordinate frame (x, y) to represent
the tensor. Note that in both cases ν̄¯ij is proportional r0
to the magnetic field (or one of its components). The ωR ∼ ωc hR/ri , (56)
λSO
spin-orbit interaction produces no spin-electric coupling
at B = 0, because of the time-reversal symmetry of spin- where hR/ri is the small parameter of deviation of
orbit interaction. the quantum dot confinement from harmonic. We re-
The analogy between µ and ν is also seen in the pair- mark that our theory remains qualitatively valid also for
wise interaction between spins in separate (not tunnel r0 /λSO ∼ 1 and for hR/ri ∼ 1. Beyond these limits, we
coupled) quantum dots.34 For an unscreened Coulomb do not expect the Rabi frequency to grow indefinitely.
interaction between electrons, the spin-spin interaction The Rabi frequency is bound in the case of Eq. (55) by
is analogous to the magnetic dipole-dipole interaction,34 ωR ≤ ωZ , and in the case of Eq. (56) by ωR ≤ ωc . We
conclude that, by designing quantum dot setups that al-
2
X νi · νj rij − 3(νi · rij )(νj · rij ) low for large oscillation amplitudes r0 . λSO , the EDSR
Hdd = 5 , (54) effect can be strongly enhanced, beyond the numeric es-
i<j
κrij timates made in Sections III, IV, and V.
In conclusion, the EDSR mechanisms presented above
where rij = ri − rj is the distance between two quantum
provides a means of implementing local electrical control
dots (rij ≫ λd ) and κ is the electric permittivity of the
of electron spins in quantum dots.
material. For further detail and a microscopic derivation
of Eq. (54) we refer the reader to Ref. 34.
Next we discuss the limitations of our theory.
Acknowledgment
Throughout the paper, we have assumed that the spin
orbit interaction is weak compared to the dot level spac-
ing, or, in other words, that λd /λSO ≪ 1. This assump- We thank M. Trif, L.M.K. Vandersypen, and D.M.
tion allowed us to use the perturbation theory to find Zumbühl for discussions. We acknowledge support from
the unitary transformation exp(−S), see Appendix A. Of the Swiss NSF, NCCR Nanoscience, DARPA, ONR, and
course, this restriction was not necessary, if, e.g., we were JST ICORP.
to apply numerical methods to diagonalize the Hamil-
tonian. In particular, note that Eq. (49) is meaning-
ful whenever the unitary transformation exp(−S) exists. APPENDIX A: SCHRIEFFER-WOLFF
TRANSFORMATION AND FINE STRUCTURE
The latter is always the case, including also extended
states. Our perturbative results are qualitatively correct
for λd . λSO and can be used in experiments to estimate In this Appendix, we first work out the Schrieffer-
the strength of the EDSR effect. The case λd ≫ λSO is Wolff transformation to the third order of perturbation
more seldom and requires a separate theoretical investi- theory and for a general weak perturbation. Then, we
gation. consider an example Hamiltonian and use the Schrieffer-
As a second limitation, we would like to mention the Wolff transformation to partly diagonalize the Hamilto-
adiabaticity criterion. In Sections III and V, we have de- nian. Finally, we analyze the fine structure of the trans-
rived effective Hamiltonians for the low energy subspace formed Hamiltonian and complete its diagonalization by
of the quantum dot Hilbert space. For the validity of this an additional unitary transformation.
effective description, it is important that the switching As in standard perturbation theory, we consider a
(on and off) of the effective interaction occurs on a time Hamiltonian H = H0 + H1 , where H1 is a weak per-
scale that is larger than the inverse level spacing in the turbation with respect to H0 . For the matrix elements
quantum dot. Obviously, this criterion excludes applica- of H1 , we assume
bility of our theory to extended states. In practice, how-
ever, the finite temperature T imposes a more stringent hn|H1 |mi = 0, for En = Em , (A1)
criterion on the confinement energy, ~2 /me λ2d ≫ kB T . hn|H1 |mi ≪ En − Em , for En 6= Em , (A2)
9

where |ni and En are, respectively, the eigenstates and example, the transformation matrix at the lowest order
eigenvalues of H0 , and are obtained from H0 |ni = En |ni. reads S ′(1) = QL̂−1
0 H1 . For ∆H, we recover then the
The projector P, defined as follows perturbation theory expansion in a more familiar form,

∆H = −PH1 L̂−1 −1 −1
X
PA = Anm |nihm|, ∀A (A3) 0 H1 + PH1 L̂0 H1 L̂0 H1 + . . . , (A15)
nm
En =Em
with the usual convention, P L̂−1 0 A = 0, ∀A, adopted.
projects onto the diagonal or degenerate part of H0 . In Next, we remark that the fine structure of H̃ =
the particular case, when theP spectrum of H0 is non- H0 + ∆H can be addressed in each particular case by
degenerate, P assumes PA = n Ann |nihn|, ∀A. From means of degenerate perturbation theory. As an exam-
Eq. (A1) and the definition (A3), it follows that ple, we consider here the Hamiltonian H = H0 + H1 ,
with H0 = Hd + HZ and H1 = HSO . Here, Hd is given
PH1 = 0, (A4) in Eq. (5), with U (r) = me ω02 r2 /2, HZ is given in Eq. (6),
PH0 = H0 . (A5) and HSO is given in Eq. (7). Using the transformation
matrix S ′ = S, with S given in Eq. (12), we obtain a
Next we look for a unitary transformation that brings
diagonal Hamiltonian, H̃ = Hd + HZ , at the first order
the Hamiltonian H = H0 + H1 to a partly diagonal form,
of HSO . At the second order of HSO , however, a fine
structure in the energy spectrum arises. At B = 0, the
′ ′
H̃ = eS (H0 + H1 ) e−S = H0 + ∆H, (A6) transformed Hamiltonian reads

where the operator ∆H obeys P∆H = ∆H. Here, p2 me ω02 2 1


H̃ = + r + ∆SO ℓz σz , (A16)
S ′ = −S ′† is the transformation matrix. The unitary 2me 2 2
transformation in Eq. (A6) is called the Schrieffer-Wolff
where ℓz = −i(x∂/∂y − y∂/∂x) is the electron rotational
transformation.35 We expand S ′ and ∆H in terms of the
momentum and ∆SO = 2me (β 2 − α2 ). The Kramers
perturbation H1 :
doublets are identified, in this case, as the pairs of states
S ′ = S ′(1) + S ′(2) + S ′(3) + . . . , (A7) with quantum numbers (ℓz , σz ) and (−ℓz , −σz ). For ℓz >
(1) (2) (3) 0, the two-fold orbital degeneracy is lifted and a splitting
∆H = ∆H + ∆H + ∆H + ..., (A8) ∆SO ℓz arises. Note that the ground orbital state, which
where the superscripts give the order of perturbation the- has ℓz = 0, remains doubly degenerate in this case.
ory. Substituting Eqs. (A7) and (A8) into Eq. (A6), we At B 6= 0, the fine-structure interaction in Eq. (A16)
find a set of equations for S ′ , is modified by both the Zeeman energy EZ and the
cyclotron frequency ωc . For simplicity, we omit terms
[H0 , S ′(1) ] = H1 , (A9) ∼ ∆SO EZ /~ω0 , but keep terms ∼ ∆SO ωc /ω0 , assuming
Q ′(1) that EZ ≪ ~ωc . Then, the Hamiltonian (A16) acquires
[H0 , S ′(2) ] = [S , H1 ], (A10) two extra terms
2
Q ′(2) Q EZ ∆SO
[H0 , S ′(3) ] = [S , H1 ] + [S ′(1) , [S ′(1) , H1 ]] + n·σ+ σz Pr2 , (A17)
2 12 2 4λ2
Q ′(1)
[S , P[S ′(1) , H1 ]], (A11)
p
4 where λ = ~/me ωc is the magnetic length and we use
the symmetric gauge, A(r) = Bz (−y/2, x/2, 0). The last
where Q ≡ 1 − P. It is important to note that S is term in Eq. (A17) can be viewed as a renormalization of
defined in Eqs. (A9)-(A11) up to terms PM , where M the electron g-factor. Allowing for an anisotropic Zeeman
is arbitrary. Such terms drop out on the left-hand side interaction,
in Eqs. (A9)-(A11) because [H0 , PS ′ ] = 0. Thus, PS ′
can be chosen arbitrarily, which shows that there are 1 X
HZeff = µB gij σi Bj , (A18)
infinitely many transformation matrices S ′ that satisfy 2 ij
Eq. (A6). For simplicity, we choose PS ′ = 0 and address
the fine structure of H̃ = H0 + ∆H later on. For the we obtain that the tensor gij is diagonal in the main
operator ∆H, we obtain crystallographic frame, with
∆H (1) = 0, (A12) gxx = gyy = g,
P ′(1) m∆SO
∆H (2) = [S , H1 ], (A13) gzz = g + hψn |r2 |ψn i, (A19)
2 ~2
P ′(1) ′(1)
∆H (3) = [S , [S , H1 ]]. (A14) where m is the electron mass in vacuum and ψn is the
3
electron orbital state. For the ground orbital state, the
Introducing the Liouvillean L̂0 : L̂0 A = [H0 , A], ∀A, corrected
p g-factor reads gzz = g + m∆SO /me ~ω, where
we can formally solve Eqs. (A9)-(A11) one by one. For ω = ω02 + ωc2 /4. Note that the sign of the correction is
10

given by the sign of β 2 − α2 contained in ∆SO . The spin So far, we have considered a given orbital state ψn , for
quantization axis does not, in general, coincide with the which the tensor gij is given in Eq. (A19). The transfor-
magnetic field direction n and is given by the following mation above is also valid in general, provided ζ is un-
unit vector derstood as a diagonal operator, ζ = (m∆SO /g~2 )Pr2 .
n + ζnz We summarize by mentioning that the unitary trans-
ñ = p , (A20) formation in Eq. (A6) can, in principle, be adjusted to
1 + ζ(2 + ζ)n2z
give a fully diagonal H̃ = H0 + ∆H, i.e. we had not to
where ζ = (gzz − g)/g. An additional unitary transfor- require PS ′ = 0 in the first place. However, in practice,
mation can be used to diagonalize the 2 × 2 blocks of it is more convenient first to apply the non-degenerate
Zeeman-split Kramers doublets, perturbation theory, Eqs. (A9)-(A15), and then, at the
end, complete the diagonalization of H0 + ∆H by a sec-
′′ ′′ 1 ond unitary transformation. The latter is specific to each
H̃Zeff = eS HZeff e−S = ẼZ n · σ (A21)
2 particular case and amounts, in general, to applying the
p degenerate perturbation theory. For the sake of simplic-
where ẼZ = EZ 1 + ζ(2 + ζ)n2z is the renormalized Zee- ity, we shall refer to S in the main text of the paper as
man energy and to the full transformation matrix, despite the fact that
r r the respective unitary transformation comes, in practice,
′′ 1 + n · ñ ζ[n × nz ] · σ 1 − n · ñ as a product of two unitary transformations. Thus, we
e−S = − ip . ′ ′′
2 2 2 2
ζ nz (1 − nz ) 2 denote the product e−S e−S′′ by e−S in the main text.
(A22) Finally, we remark that e−S ≈ 1 + O(HSO 2
).

1 19
S.A. Wolf, D.D. Awschalom, R.A. Buhrman, J.M. L.P. Kouwenhoven, D.G. Austing, and S. Tarucha, Rep.
Daughton, S. von Molnár, M.L. Roukes, A.Y. Chtchelka- Prog. Phys. 64 701 (2001).
20
nova, and D.M. Treger, Science 294, 1488 (2001). A.V. Khaetskii and Yu.V. Nazarov, Phys. Rev. B 61, 12639
2
Semiconductor Spintronics and Quantum Computation, (2000); B.I. Halperin, A. Stern, Y. Oreg, J.N.H.J. Cremers,
D.D. Awschalom, D. Loss, and N. Samarth (eds.), J.A. Folk, and C.M. Marcus, Phys. Rev. Lett. 86, 2106
(Springer, Berlin, 2002). (2001); I.L. Aleiner and V.I. Fal’ko, Phys. Rev. Lett. 87,
3
D. Loss and D.P. DiVincenzo, Phys. Rev. A 57, 120 (1998). 256801 (2001).
4 21
N.M. Atherton, Electron Spin Resonance, (Ellis Horwod V.N. Golovach, A. Khaetskii, and D. Loss, Phys. Rev. Lett.
Limited, New York, 1973). 93, 016601 (2004).
5 22
H.-A. Engel and D. Loss, Phys. Rev. Lett. 86, 4648 (2001); A.V. Khaetskii and Yu.V. Nazarov, Phys. Rev. B 64,
Phys. Rev. B 65, 195321 (2002). 125316 (2001).
6 23
Y. Kato, R.C. Myers, D.C. Driscoll, A.C. Gossard, J. Levy, M.I. D’yakonov and V.Yu. Kachorovskii, Sov. Phys. Semi-
and D.D. Awschalom, Science 299, 1201 (2003). cond. 20, 110 (1986).
7 24
R.C. Miller, A.C. Gossard, D.A. Kleinman, and O. J.M. Elzerman, R. Hanson, L.H. Willems van Beveren, B.
Munteanu Phys. Rev. B 29, 3740 (1984). Witkamp, L.M.K. Vandersypen, and L.P. Kouwenhoven,
8
G. Salis, Y. Kato, K. Ensslin, D.C. Driscoll, A.C. Gossard, Nature (London) 430, 431 (2004).
25
and D.D. Awschalom, Nature (London) 414, 619 (2001). M. Kroutvar, Y. Ducommun, D. Heiss, M. Bichler, D.
9
E.I. Rashba and Al.L. Efros, Phys. Rev. Lett. 91, 126405 Schuh, G. Abstreiter, and J.J. Finley, Nature (London)
(2003); Appl. Phys. Lett. 83, 5295 (2003); E.I. Rashba, J. 432, 81 (2004).
26
Supercond. Incorp. Novel. Magn. 18, 137, (2005). J.D. Jackson, Classical Electrodynamics, (John Wiley &
10
G. Dresselhaus, Phys. Rev. 100, 580 (1955). Sons, Inc., New York, 1999).
11 27
Y. Bychkov and E. I. Rashba, J. Phys. C 17, 6039 (1984). F.H.L. Koppens, C. Buizert, K.J. Tielrooij, I.T. Vink,
12
R.L. Bell, Phys. Rev. Lett. 9, 52 (1962). K.C. Nowack, T. Meunier, L.P. Kouwenhoven, and
13
B.D. McCombe, S.G. Bishop, and R. Kaplan, Phys. Rev. L.M.K. Vandersypen, Nature (London) 442, 766 (2006).
28
Lett. 18, 748 (1967). M. Borhani, V.N. Golovach, and D. Loss, Phys. Rev. B 73,
14
M. Dobrowolska, H.D. Drew, J.K. Furdyna, T. Ichiguchi, 155311 (2006).
29
A. Witowski, and P.A. Wolff, Phys. Rev. Lett. 49, 845 F. Bloch, Phys. Rev. 70, 460 (1946); R. K. Wangsness and
(1982); M. Dobrowolska, A. Witowski, J.K. Furdyna, F. Bloch, Phys. Rev. 89, 728 (1953); F. Bloch, Phys. Rev.
T. Ichiguchi, H.D. Drew, and P.A. Wolff, Phys. Rev. B 105, 1206 (1957).
30
29, 6652 (1984). C. Cohen-Tannoudji, B. Diu, and F. Laloë, Quantum Me-
15
M. Schulte, J.G.S. Lok, G. Denninger, and W. Dietsche, chanics, Vol. II (John Wiley & Sons, New York, 1977).
31
Phys. Rev. Lett. 94, 137601 (2005). S. Amasha, K. MacLean, I. Radu, D.M. Zum-
16
M. Duckheim and D. Loss, Nature Physics 2, 195 (2006). buhl, M.A. Kastner, M.P. Hanson, and A.C. Gossard,
17
Y. Kato, R.C. Myers, A.C. Gossard, and D.D. Awschalom, cond-mat/0607110.
32
Nature (London) 427, 50 (2004). G.E. Pikus and A.N. Titkov, in Optical orientation, edited
18
J. Nitta, T. Akazaki, H. Takayanagi, and T. Enoki Phys. by F. Meier and B.P. Zakharchenya (North-Holland, Am-
Rev. Lett. 78, 1335 (1997). sterdam, 1984).
11

33 34
F. Bayen, M. Flato, C. Fronsdal, A. Lichnerowicz, and M. Trif, V.N. Golovach, and D. Loss, cond-mat/0608512.
35
D. Sternheimer, Ann. Phys. 111, 61 (1978); Ann. Phys. J.R. Schrieffer, P.A. Wolff, Phys. Rev. 149, 491 (1966).
111, 111 (1978).

You might also like