ArticlePDF Available

Abstract

It has been proposed that valence-band holes can form robust spin qubits1, 2, 3, 4 owing to their weaker hyperfine coupling compared with electrons5, 6. However, it was demonstrated recently7, 8, 9, 10, 11 that the hole hyperfine interaction is not negligible, although a consistent picture of the mechanism controlling its magnitude is still lacking. Here we address this problem by measuring the hole hyperfine constant independently for each chemical element in InGaAs/GaAs, InP/GaInP and GaAs/AlGaAs quantum dots. Contrary to existing models10, 11 we find that the hole hyperfine constant has opposite signs for cations and anions and ranges from −15% to +15% relative to that for electrons. We attribute such changes to the competing positive contributions of p-symmetry atomic orbitals and the negative contributions of d-orbitals. These findings yield information on the orbital composition of the valence band12 and enable a fundamentally new approach for verification of computed Bloch wavefunctions in semiconductor nanostructures13. Furthermore, we show that the contribution of cationic d-orbitals leads to a new mechanism of hole spin decoherence.
arXiv:1109.0733v2 [cond-mat.mes-hall] 18 Oct 2011
Isotope sensitive measurement of the hole-nuclear spin interaction in quantum dots
E. A. Chekhovich1, A. B. Krysa2, M. Hopkinson2, P. Senellart3, A. Lemaˆıtre3, M. S. Skolnick1, A. I. Tartakovskii1
1Department of Physics and Astronomy, University of Sheffield, Sheffield S3 7RH, UK
2Department of Electronic and Electrical Engineering, University of Sheffield, Sheffield S1 3JD, UK
3Laboratoire de Photonique et de Nanostructures, Route de Nozay, 91460 Marcoussis, France
(Dated: October 19, 2011)
Decoherence caused by the hyperfine interaction with
nuclear spins is known to be the main obstacle on the
way to realization of quantum computation using sin-
gle electron spins [13], which led to proposals to use
valence band holes having a weaker hyperfine coupling
[4–9]. Although the hole hyperfine interaction has been
addressed recently both theoretically [10, 11] and ex-
perimentally [1214], full understanding of the the un-
derlying physics is still lacking. Here we report on ex-
perimental measurements of the hole hyperfine inter-
action strength in three different material systems: un-
strained GaAs/AlGaAs quantum dots (QDs), and self-
assembled InGaAs/GaAs and InP/GaInP QDs. In con-
trast to previous studies we use resonant radio fre-
quency (rf) excitation to achieve selective measurement
of the hole hyperfine constant for individual isotopes.
This allows to avoid the ambiguity of previous mea-
surements relying on detection of the total Overhauser
shifts including contributions of all isotopes. We find
that the hole hyperfine constant (normalized by the
electron hyperfine constant) changes sign for different
isotopes and ranges from -15% for indium to +15%
for arsenic, revealing strong anisotropy of the dipole-
dipole hyperfine interaction. Moreover, the hole hy-
perfine constant varies for the same isotope in differ-
ent materials opening the way for better understanding
and possible optimization of the hyperfine interaction
for applications using single hole spins.
Due to the s-type of the wavefunction, the hyperfine
interaction of the conduction band electrons has a sim-
ple isotropic form (the Fermi contact interaction) and is
described by a single constant A, depending on isotope
and material [15]. Since all nuclei in III-V semiconduc-
tors have positive nuclear moments, their electron hyper-
fine constants are also positive: A > 0. By contrast, for the
p-type valence band holes, the contact interaction vanishes,
and the non-local dipole-dipole interaction dominates. As
a result, the hole hyperfine interaction is anisotropic and
its strength depends strongly on the actual form of the
Bloch wavefunction, which is usually difficult to estimate
with sufficient precision. Indeed, for a long time, it was
wrongly assumed that the hole hyperfine interaction is neg-
ligibly small. Only recently it was shown first theoretically
[10, 11] and then experimentally [13, 14] that it can be as
large as 10% of that for the electron. However, despite
its simple fundamental nature, understanding of the hole-
nuclear interaction is far from being complete.
In this work we address this problem experimentally. We
perform direct measurements of the hole hyperfine con-
stants by independently detecting electron and hole hy-
perfine shifts. This is achieved by using high resolution
photoluminescence (PL) spectroscopy of optically forbid-
den (”dark”) exciton states in single neutral quantum dots
[14]. In contrast to previous work, where similar tech-
niques were used [14], we now also apply radio-frequency
(rf) excitation, which allows selective saturation of nuclear
polarization of different isotopes. This opens the way to
isotope-sensitive probing of the valence band hole hyper-
fine interaction, revealing that in all studied materials group
III isotopes (gallium, indium) have negative hole hyperfine
constant, while it is positive for group V arsenic.
Our experiments were performed on undoped
GaAs/AlGaAs [16] and InGaAs/GaAs QD samples
without electric gates. PL of neutral QDs was measured
at T= 4.2K, in external magnetic field Bznormal to the
sample surface. QD PL was analyzed with a 1 m double
spectrometer and a CCD.
In a neutral dot electrons () with spin se
z=±1/2
and heavy holes () with momentum jh
z=±3/2par-
allel (antiparallel) to the growth axis Oz can form either
optically-forbidden (”dark”) excitons |⇑↑i (|⇓↓i) with the
spin projection Jz= +2(2), or bright” excitons |⇑↓i
(|⇓↑i) with Jz= +1(1) optically allowed in σ+(σ)
polarization. QD axis misorientation or symmetry reduc-
tion leads to weak mixing of ”bright” and ”dark” states: as
a result the latter are observed in PL [17, 18].
Non-zero average nuclear spin polarization hIzialong
the Oz axis acts as an additional magnetic field on the
electron and hole spins. It is convenient to introduce the
hole pseudospin Sh
z=±1/2corresponding to the ()
heavy hole state. Coupling of the electron to the nuclear
spin of isotope iis described by the hyperfine constant
Ai, whereas for the heavy hole the dipole-dipole interac-
tion with hIzi[10, 11] is described using a constant Ci
expressed in terms of the normalized heavy-hole hyperfine
constant γias Ci=γiAi. The expression for the exciton
energy taking into account the shift due to non-zero aver-
age nuclear spin polarization can be written as:
E[Sh
z, se
z] = EQD +E0[Sh
z, se
z] +
+se
zX
i
ρiAihIi
zi+Sh
zX
i
ρiγiAihIi
zi,(1)
where the quantum dot band-gap EQD and shift
E0[Sh
z, se
z]determined by the Zeeman and exchange en-
2
60 80 100
probe & PL detection
rf pulse
(e)
Experiment cycle
pump
75As 71Ga
69Ga
115In
rf power spectral
density (arb.units)
Band Z
Band Y
Band X
(d)
Frequency (MHz)
1.7130 1.7135
BZ=3.2 T
GaAs/AlGaAs
(a)
|›`Ò
|flcÒ
|›cÒ
|fl`Ò
PL energy (eV)
PL Intensity (arb. units)
1.3575 1.3580 1.3585
PL energy (eV)
IZ < 0
IZ > 0
BZ=8 T
InGaAs/GaAs
(b)
|›`Ò
|flcÒ
|›cÒ
|fl`Ò
645
650
655 57.75 58.00
Spectral splitting (µeV)
(c)
75As
81.00 81.25
Frequency (MHz)
69Ga
103.00 103.25
71Ga
FIG. 1. (a,b) Photoluminescence spectra of a single GaAs/AlGaAs QD at Bz3.2 T (a) and an InGaAs/GaAs QD at Bz8.0 T
(b) measured at negative (open symbols) and positive (solid symbols) nuclear spin polarization hIziinduced on the dot. Both bright
excitons |⇑↓i,|⇓↑i and both dark excitons |⇑↑i,|⇓↓i are observed. (c) Optically detected NMR spectrum of a single GaAs QD
at Bz8.0 T. Gallium resonances with widths of 30 kHz are observed at 81.21 MHz and 103.17 MHz for 69 Ga and 71Ga
respectively. Arsenic resonance observed at 57.9 MHz has a linewidth of 100 kHz, which is determined by residual elastic strain.
(d) Schematic diagram of the radio-frequency excitation spectrum used to erase nuclear polarization of different isotopes in InGaAs
QDs at Bz8.0 T. The solid vertical bars show resonance frequencies of gallium and arsenic derived from (c), while the dashed line
shows the calculated central frequency 74.14 MHz of 115In. Bands X and Z are used to destroy nuclear polarization of 75 As and
71Ga respectively. Band Y is used to erase polarization of both 115In and 69Ga simultaneously. (e) Timing diagram of the pump-probe
experiment used in the measurements of the hole hyperfine constants.
ergy [17] do not depend on nuclear polarization. The sum-
mation goes over all isotopes icontributing to the Over-
hauser shift. The relative concentration of each isotope is
given by ρi. We note that in Eq. 1 we neglect any pos-
sible variation of average nuclear spin polarization within
the volume of the QD. Also we neglect any difference of
isotope concentrations within the electron and hole local-
ization volumes.
Since mixing of the ”dark” and ”bright” excitonic states
is weak, the oscillator strength of the ”dark” states is small,
leading to their saturation at relatively low laser powers. As
a result, all four exciton states can be observed in PL only
at low excitation power. However, at this low power, opti-
cally induced nuclear spin polarization is small and weakly
depends on polarization of photoexcitation [18], and thus
the shifts of the hole spin states due to the interaction with
the nuclei cannot be measured accurately. To avoid this
problem, we use pump-probe techniques [19]. The experi-
3
ment cycle shown in Fig. 1 (e) is similar to that used in our
previous work [14]: nuclear spin polarization is prepared
with a long (6 s) high power pump pulse. Following this,
the sample is excited with a low power probe pulse, during
which the PL spectrum of both bright and dark excitons is
measured. However, in contrast to previous work, we now
add an rf pulse between the pump and probe pulses. This
pulse is formed by an oscillating magnetic field perpendic-
ular to Oz at a frequency tuned to erase transitions between
nuclear spin states of a chosen isotope (isotopes).
The direct and simultaneous measurement of the hole
and electron energy shifts due to the hyperfine interaction
is carried out by detecting the probe spectra recorded at
different magnitudes of hIi
ziprepared by the pump. Typ-
ical probe spectra of GaAs and InGaAs QDs are shown
in Figs. 1 (a) and (b) respectively. If the rf pulse con-
tains frequency components resonant only with isotope k,
then the exciton energies Ek[Sh
z, se
z]detected by the probe
pulse will be given by Eq. 1 with the sum going only
over the isotopes for which i6=k. We also perform an
experiment with no rf pulse which gives exciton energies
E[Sh
z, se
z]corresponding to initial polarization. Calculat-
ing the difference of the exciton splittings for experiments
with and without rf pulse allows to extract the electron and
hole hyperfine shifts for the k-th isotope. For example,
according to Eq. 1, the experimentally measured quan-
tity Ek
hf,e = (E[,]E[,]) (Ek[,]Ek[
,]) gives the magnitude of the electron hyperfine shift
ρkAkhIk
ziinduced only by nuclei depolarized by the rf
pulse (and thus corresponding to isotope kpolarization),
whereas the hole hyperfine shift ρkγkAkhIk
ziis given by
Ek
hf,h = (E[,]E[,]) (Ek[,]Ek[,]).
From Eq. 1 we find that when nuclear spin polarization is
varied, Ek
hf,h and Ek
hf,e (expressed via the experimen-
tally measured exciton energies) depend linearly on each
other with the proportionality constant given by the hole
hyperfine constant γk.
We start with the analysis of the results for
GaAs/AlGaAs interface fluctuation quantum dots.
Optically detected nuclear magnetic resonance (ODNMR)
is well studied for GaAs/AlGaAs QDs [16, 20]. A typical
NMR spectrum at Bz8T is shown in Fig. 1 (c).
Resonances corresponding to all three isotopes (75As,
69Ga, 71 Ga) are clearly observed. We find no contribution
from 27Al isotope of the quantum well barrier and estimate
that contribution of this isotope into the total Overhauser
shift is less then 3% and can be neglected.
In order to measure the hole hyperfine interaction with
75As, we perform experiments with the rf pulse of a rect-
angular shaped spectral band 600 kHz wide with the central
frequency corresponding to the NMR resonance frequency
of 75As. Inside the band, the rf signal has a constant spec-
tral power density (a white noise type), with the power den-
sity outside the bands 1000 smaller than inside the band.
Application of such a pulse results in complete depolariza-
tion of arsenic spins, whereas the gallium polarization re-
mains unaffected. The dependence of Ek
hf,h on Ek
hf,e
for k=75As is shown in Fig. 2 (a) with squares for QD
A1.
Since both gallium isotopes have equal chemical prop-
erties (i. e. equal electron wavefunctions), we can assume
that they have the same values of the relative hole hyper-
fine interaction constants γ69 Ga =γ71 Ga . Thus measure-
ment of γGa can be accomplished by performing experi-
ment with rf pulse erasing both 69Ga and 71 Ga polarization,
achieved by applying the rf pulse consisting of two equal
spectral bands centered at corresponding resonant frequen-
cies. The result of this experiment for the same QD is
shown in Fig. 2 (a) with circles. It can be seen that de-
pendencies for both Ga and As follow linear pattern pre-
dicted by Eq. 1. Fitting gives the following values for the
hole hyperfine constants γGa =7.0±4.0% and γAs =
+15.0±4.5%. Similar measurements were performed on
3 other GaAs QDs. The resulting values are given in Table
I. Since variation between different dots is within the ex-
perimental error, we take average values for all dots yield-
ing γGa =7.5±2.0% and γAs = +16.0±2.5%. We
thus conclude that different isotopes have opposite signs of
the hole hyperfine constants: it is positive for arsenic and
negative for gallium. This is in contrast to previous reports
[13, 14] where negative values of γfor InP and InGaAs
QDs have been derived.
We have also performed isotope-sensitive measurements
of the hole nuclear interaction in InGaAs/GaAs QDs. How-
ever, these QDs have a more complicated nuclear spin sys-
tem. This is due to significant lattice mismatch result-
ing in strain-induced quadrupolar shifts [21]. Quadrupolar
effects shift NMR frequencies causing significant broad-
ening of the resonance [15]. The magnitudes of these
shifts for typical values of strain 0.02 % can be esti-
mated [22] using the known values of the tensor relating
electric field gradient and elastic strain [23]: f(75As)
3MHz, f(115In)5MHz, f(69 Ga)1.5MHz,
f(71Ga)1MHz. This frequency shift can vary
strongly within the QD volume, therefore in order to erase
nuclear polarization, broadband rf excitation must be used.
At Bz= 8 T we used three different bands of rf excita-
tion shown in Fig. 1 (d). Bands X and Z are used to erase
selectively polarization of 75As and 71 Ga: the widths of
these bands are chosen to be several times the quadrupo-
lar broadening fof the targeted isotope while leaving
the other isotopes unaffected. However, the frequencies of
115In and 69 Ga are too close for these isotopes to be ad-
dressed individually (Fig. 1 (d)). For that reason, we use rf
excitation with the broad band Y, which erases polarization
of both isotopes [24].
The dependencies of Ek
hf,h on Ek
hf,e for InGaAs QD
B1 are shown in Fig. 2 (b) for 71Ga (circles, using rf band
Z) and 75As (squares, rf band X). As in GaAs, we find that
arsenic has a positive hole hyperfine constant while gallium
has a negative one. We also measured EIn+69 Ga
hf,h as a
4
-40 -20 0 20 40
-10
-5
0
5
10
-40 -20 0 20 40 60
Hole HF shift E k
hf, h (µeV)
(a)
As
Ga
GaAs/AlGaAs InGaAs/GaAs
(b)
As
71Ga
In & 69Ga
Electron HF shift Ek
hf, e (µeV)
FIG. 2. The dependence of the hole hyperfine shift Ek
hf,h on the electron hyperfine shift Ek
hf,e for different isotopes in GaAs QD A1
(a) and InGaAs QD B1 (b). the electron hyperfine shift for isotope kis found as a difference of the spectral splitting (E[,]E[,])
measured without rf-excitation and the splitting (Ek[,]Ek[,]) measured after erasing nuclear polarization corresponding to the k-
th isotope by the rf pulse. In the same way, the hole hyperfine shift is measured as Ek
hf,h = (E[,]E[,])(Ek[,]Ek[,]).
Solid lines show fitting, their slopes are given by the corresponding relative hole-nuclear hyperfine constants γk. We find γGa 7.0%,
γAs +15.0% for GaAs QD A1 and γGa 6.5%,γAs +10.5% for InGaAs QD B1. Since NMR resonances of 69Ga and 115 In
in InGaAs cannot be resolved, we measure the total hyperfine shifts EI n+69Ga
hf,e and EI n+69Ga
hf,h produced by these isotopes. Fitting
(see text) gives γIn 16.0% for QD B1. Dashed line is a guide for an eye.
function of EIn+69Ga
hf,e with rf band Y depolarizing both
115In and 69Ga nuclei. This is shown with triangles in Fig.
2 (b). It can be seen that the experimental dependency for
TABLE I. Experimentally measured hole hyperfine constants γi
for different isotopes iin several GaAs and InGaAs QDs. Error
estimates give 90% confidence trust regions. Average values for
each isotope in each material are given at the bottom of the table,
the value for indium in InP is taken from Ref. [14]
Material/QD γGa,%γI n,%γAs,%
GaAs/AlGaAs:
QD A1 -7.0±4.0 - +15.0±4.5
QD A2 -8.5±3.5 - +17.0±5.0
QD A3 -5.5±4.5 - +15.0±4.0
QD A4 -7.5±4.5 - +18.5±5.5
InGaAs/GaAs:
QD B1 -6.5±5.5 -16.0±4.0 +10.5±2.0
QD B2 -3.0±6.5 -15.5±5.0 +10.0±3.0
QD B3 -5.5±5.0 -16.0±4.0 +8.0±2.0
QD B4 -4.5±7.0 -13.0±4.5 +8.5±3.0
Average:
InGaAs/GaAs -5.0±3.0 -15.0±2.0 +9.0±1.0
GaAs/AlGaAs -7.5±2.0 - +16.0±2.5
InP/GaInP - -10.5±1.0 -
115In and 69Ga has a negative slope, with the absolute value
exceeding that of 71 Ga. Consequently, we conclude that
γIn <0and γI n < γGa.
Fitting using Eq. 1 gives the following values for the hy-
perfine constants γGa =6.5±5.5% and γAs = +10.5±
2.0%. Similar measurements were performed on 3 other
InGaAs QDs. The resulting values are given in Table I.
Since the variation between different dots is within the ex-
perimental error, we take average values for all dots yield-
ing γGa =5.0±3.0% and γAs = +9.0±1.0%. The
hyperfine constant of indium can also be estimated from the
experimental results. This requires an additional assump-
tion that both gallium isotopes have the same degrees of
spin polarization hI69 Ga
zi=hI71Ga
zias a result of nuclear
spin pumping. Such assumption is justified by the fact that
both isotopes have the same spin I= 3/2and both become
polarized due to the hyperfine interaction with the optically
polarized electrons. Since γ69 Ga =γ71 Ga , we can calcu-
late the Overhauser shifts of 69 Ga from the measured shifts
of 71Ga. For that we need to take into account the ratio of
natural abundances of this isotopes, ρ69Ga 71Ga 1.5,
and the ratio of the absolute magnitudes of the electron hy-
perfine constants, A69 Ga /A71Ga, equal to the ratio of the
magnetic moments µ69 Ga71Ga 0.79. Thus the elec-
tron (hole) hyperfine shifts of indium can be written as
EIn
hf,e(h)= EI n+69Ga
hf,e(h)ρ69 Ga
ρ71Ga
µ69Ga
µ71Ga E71 Ga
hf,e(h). Using
this expression for the fitting, we find γIn =16.0±4.0%
for QD B1 with an average for 4 QDs of γIn =15.0±
5
2.0%.
We can now compare these results with previous reports
[1214]. In our previous work, [14] we used InP/GaInP
QDs. Using the rf-induced depolarization techniques re-
ported here for InGaAs QDs we found that the contribution
of gallium and phosphorus isotopes into the Overhauser
shift is small compared to the contribution of indium (less
than 10%). Therefore, the value γ=10.5±1.0% re-
ported previously corresponds to the indium hole hyperfine
constant γIn in InP. On the other hand, Fallahi et al [13] re-
ported negative γ=9.0±2.0% for InGaAs QDs (aver-
age for all isotopes). For InGaAs QDs studied in this work
we find a much smaller average hyperfine constant (mea-
sured without isotopic sensitivity) γ 2.0%. This can
be explained if we take into account that InGaAs QDs used
in this work have large abundance of gallium with smaller
absolute value of γcompared to indium. This is revealed
by the short PL wavelength (915 nm) in our sample com-
pared to the longer wavelengths (950 nm) reported by
Fallahi et al. [13] As a result, the negative average hole
hyperfine interaction found in Ref. [13] can be explained
by a significant contribution of indium, dominating due to
its large magnetic moment 3.5 times greater than that for
arsenic which has a positive hyperfine constant.
The hyperfine interaction of the valence band holes has
been considered in several theoretical papers [10, 11, 25].
The hyperfine interaction of the conduction band electrons
has a contact (Fermi) form and therefore depends only on
the electron wavefunction density at the nucleus site [15].
By contrast for the valence band holes, the hyperfine cou-
pling is dominated by the dipole-dipole interaction. As a
result, calculations of the hole hyperfine constants mea-
sured in this work requires averaging over the spatial co-
ordinates using an explicit expression for the Bloch wave-
functions, which are not known. One approach is to use
the spherical approximation taking p-type atomic orbitals
to approximate the real Bloch wavefunctions [11, 25]. The
hole hyperfine constants γcalculated in this way have the
same signs for all isotopes in contradiction with our experi-
mental observations. The exact reason for this discrepancy
is not yet fully understood. A possible explanation is that
the real Bloch wavefunction with the symmetry imposed
by the crystal symmetry deviates strongly from the spheri-
cal approximation resulting in variation of γ, including the
sign reversal, for the isotopes with the opposite charges.
In conclusion, we have employed optical spectroscopy
of single quantum dots to measure hole hyperfine constants
γfor individual isotopes in tree types of III-V semicon-
ductor quantum dots. Strong variation of γfor different
isotopes and for different material systems has been found.
This opens the way for improved modeling of microscopic
wavefunctions and deeper insight into fundamental proper-
ties of semiconductors on the atomic scale. Better under-
standing of these properties will allow to engineer mate-
rials with the valence band hyperfine interaction optimized
for future applications requiring highly coherent hole spins.
The authors are thankful to M. M. Glazov for fruitful
discussions, and D. Martrou for help with the GaAs sam-
ple growth. This work has been supported by EPSRC Pro-
gramme Grant No. EP/G601642/1, the Royal Society, and
ITN Spin-Optronics.
[1] A. V. Khaetskii, D. Loss, L. Glazman, Phys. Rev. Lett. 88,
186802 (2002).
[2] H. Bluhm, et al., Nature Physics 7, 109 (2011).
[3] S. Foletti, H. Bluhm, D. Mahalu, V. Umansky, A. Yacoby,
Nature Physics 5, 903 (2009).
[4] B. D. Gerardot, et al., Natrue 451, 441 (2007).
[5] D. Brunner, et al., Science 325, 70 (2009).
[6] D. Heiss, et al., Phys. Rev. B 76, 241306 (2007).
[7] T. M. Godden et al, arXiv:1106.6282 (2011).
[8] A. Greilich et al, arXiv:1107.0211 (2011).
[9] K. De Greve et al, arXiv:1106.5676 (2011).
[10] C. Testelin, F. Bernardot, B. Eble, M. Chamarro, Phys. Rev.
B79, 195440 (2009).
[11] J. Fischer, W. A. Coish, D. V. Bulaev, D. Loss, Phys. Rev. B
78, 155329 (2008).
[12] B. Eble, et al., Phys. Rev. Lett. 102, 146601 (2009).
[13] P. Fallahi, S. T. Y ılmaz, A. Imamo˘glu, Phys. Rev. Lett. 105,
257402 (2010).
[14] E. A. Chekhovich, A. B. Krysa, M. S. Skolnick, A. I. Tar-
takovskii, Phys. Rev. Lett. 106, 027402 (2011).
[15] A. Abragam, The principles of Nuclear Magnetism (Oxford
University Press, London, 1961).
[16] M. N. Makhonin, et al., Phys. Rev. B 82, 161309 (2010).
[17] M. Bayer, et al., Phys. Rev. B 65, 195315 (2002).
[18] E. A. Chekhovich, A. B. Krysa, M. S. Skolnick, A. I. Tar-
takovskii, Phys. Rev. B 83, 125318 (2011).
[19] E. A. Chekhovich, et al., Phys. Rev. B 81, 245308 (2010).
[20] D. Gammon, et al., Science 277, 85 (1997).
[21] K. Flisinski, et al., Phys. Rev. B 82, 081308 (2010).
[22] P. Maletinsky, M. Kroner, A. Imamoglu, Nature Physics 5,
407 (2009).
[23] R. K. Sundfors, Phys. Rev. B 10, 4244 (1974).
[24] In order to ensure that rf excitation applied to a targeted iso-
tope does not affect nuclear polarization of other isotopes
we have performed additional calibration experiments. For
k=71Ga and k=75 As the hyperfine shift of the k-th iso-
tope Ek
hf,e was measured as a function of the spectral
width wexc of the rf excitation centered at a frequency of
that isotope. As expected Ek
hf,e first increases and then
saturates for a certain wexc that exceeds the total width of
the NMR resonance of the k-th isotope. This way we deter-
mined bands X and Z [see Fig. 1 (d)] that erase only nuclear
polarization of 75As and 71 Ga respectively and found that
69Ga and 115 In resonances can not be separated. Thus band
Y was designed to erase only nuclear polarization of both
69Ga and 115 In.
[25] E. I. Gryncharova, V. I. Perel, Sov. Phys. Semicond. 11, 997
(1977).
... where ∆E PL,0 is the photoluminescence doublet splitting at zero nuclear spin polarization. Valence band hole hyperfine interaction is of the order of 10% of the electron hyperfine interaction [45]. ...
... For valence band holes the contact (Fermi) contribution vanishes, leaving the weaker dipoledipole terms to dominate the hyperfine interaction. Compared to the valence band electrons, the coupling has a more complicated non-Ising form [45]. The effect of the net nuclear polarization on the heavy-hole spin splitting can be captured by a simplified expression: ...
... whereĵ z is the z component of the hole spin momentum operator. The valence band hyperfine material constants C (j) are sensitive to heavy-light hole mixing and both their signs and magnitudes depend on the material [45]. ...
Preprint
Full-text available
Magnetic noise of atomic nuclear spins is a major problem for solid state spin qubits. Highly-polarized nuclei would not only overcome this obstacle, but also make nuclear spins a useful quantum information resource. However, achieving sufficiently high nuclear polarizations has remained an evasive goal. Here we implement a nuclear spin polarization protocol which combines strong optical pumping and fast electron tunneling. Polarizations well above 95% are generated in GaAs semiconductor quantum dots on a timescale of 1 minute. The technique is compatible with standard quantum dot device designs, where highly-polarized nuclear spins can simplify implementations of quantum bits and memories, as well as offer a testbed for studies of many-body quantum dynamics and magnetism.
... The hyperfine interaction of the valence band holes is an order of magnitude smaller [49] and can be ignored in the context of this work. ...
Preprint
Full-text available
The measurement problem dates back to the dawn of quantum mechanics. Here, we measure a quantum dot electron spin qubit through off-resonant coupling with thousands of redundant nuclear spin ancillae. We show that the link from quantum to classical can be made without any "wavefunction collapse", in agreement with the Quantum Darwinism concept. Large ancilla redundancy allows for single-shot readout with high fidelity $\approx99.85\%$. Repeated measurements enable heralded initialization of the qubit and probing of the equilibrium electron spin dynamics. Quantum jumps are observed and attributed to burst-like fluctuations in a thermally populated phonon bath.
... The performance of the qubits can be estimated by the quality factor Q = ω R T 2 /2π, which is the operation times of a full rotation before the qubit states decohere. In silicon, the dephasing of hole spin qubit due to the hyperfine interaction can be reduced by isotopic purification [62][63][64]. Thus, the pure dephasing of the acceptor qubit is mainly induced by the charge noise. ...
Preprint
Full-text available
Long coherence time and compatibility with semiconductor fabrication make spin qubits in silicon an attractive platform for quantum computing. In recent years, hole spin qubits are being developed as they have the advantages of weak coupling to nuclear spin noise and strong spin-orbit coupling (SOC), in constructing high-fidelity quantum gates. However, there are relatively few studies on the hole spin qubits in a single acceptor, which requires only low density of the metallic gates. In particular, the investigation of flexible tunability using controllable strain for fault-tolerant quantum gates of acceptor-based qubits is still lacking. Here, we study the tunability of electric dipole spin resonance (EDSR) of acceptor-based hole spin qubits with controllable strain. The flexible tunability of LH-HH splitting and spin-hole coupling (SHC) with the two kinds of strain can avoid high electric field at the "sweet spot", and the operation performance of the acceptor qubits could be optimized. Longer relaxation time or stronger EDSR coupling at low electric field can be obtained. Moreover, with asymmetric strain, two "sweet spots" are induced and may merge together, and form a second-order "sweet spot". As a result, the quality factor $Q$ can reach $10^{4}$ for single-qubit operation, with high tolerance for the electric field variation. Furthermore, the two-qubit operation of acceptor qubits based on dipole-dipole interaction is discussed for high-fidelity two-qubit gates. The tunability of spin qubit properties in acceptor via strain could provide promising routes for spin-based quantum computing.
Article
Full-text available
The measurement problem dates back to the dawn of quantum mechanics. Here, we measure a quantum dot electron spin qubit through off-resonant coupling with a highly redundant ancilla, consisting of thousands of nuclear spins. Large redundancy allows for single-shot measurement with high fidelity ≈99.85%. Repeated measurements enable heralded initialization of the qubit and backaction-free detection of electron spin quantum jumps, attributed to burstlike fluctuations in a thermally populated phonon bath. Based on these results we argue that the measurement, linking quantum states to classical observables, can be made without any “wave function collapse” in agreement with the Quantum Darwinism concept.
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
Long coherence time and compatibility with semiconductor fabrication make spin qubits in silicon an attractive platform for quantum computing. In recent years, hole spin qubits are being developed as they have the advantages of weak coupling to nuclear spin noise and strong spin-orbit coupling (SOC), in constructing high-fidelity quantum gates. However, there are relatively few studies on the hole spin qubits in a single acceptor, which requires only low density of the metallic gates. In particular, the investigation of flexible tunability using controllable strain for fault-tolerant quantum gates of acceptor-based qubits is still lacking. Here, we study the tunability of electric dipole spin resonance (EDSR) of acceptor-based hole spin qubits with controllable strain. The flexible tunability of heavy hole-light hole splitting and spin-hole coupling (SHC) with the two kinds of strain can avoid a high electric field at the “sweet spot”, and the operation performance of the acceptor qubits could be optimized. Longer relaxation time or stronger EDSR coupling at a low electric field can be obtained. Moreover, with asymmetric strain, two sweet spots are induced and may merge together, and form a second-order sweet spot. As a result, the quality factor Q can reach 104 for a single-qubit operation, with a high tolerance for the electric field variation. Furthermore, the two-qubit operation of acceptor qubits based on dipole-dipole interaction is discussed for high-fidelity two-qubit gates. The quality factors of single-qubit gates and two-qubit gates can be enhanced by 100 and 7 times respectively with tunable strain. The tunability of spin qubit properties in an acceptor via strain could provide promising routes for spin-based quantum computing.
Article
Full-text available
In recent years, hole-spin qubits based on semiconductor quantum dots have advanced at a rapid pace. We first review the main potential advantages of these hole-spin qubits with respect to their electron-spin counterparts, and give a general theoretical framework describing them. The basic features of spin-orbit coupling and hyperfine interaction in the valence band are discussed, together with consequences on coherence and spin manipulation. In the second part of the article we provide a survey of experimental realizations, which spans a relatively broad spectrum of devices based on GaAs, Si, or Si/Ge heterostructures. We conclude with a brief outlook.
Article
Spin-orbit interaction (SOI) plays a fundamental role in many low-dimensional semiconductor and hybrid quantum devices. In the rapidly evolving field of semiconductor spin qubits, SOI is an essential ingredient that can allow for ultrafast qubit control. The exact manifestation of SOI in a given device is, however, often both hard to predict theoretically and probe experimentally. Here, we develop a detailed theoretical connection between the leakage current through a double quantum dot in Pauli spin blockade and the underlying SOI in the system. We present a general analytic expression for the leakage current, which allows to connect experimentally observable features to both the magnitude and orientation of an effective spin-orbit field acting on the moving carriers. Motivated by the large recent interest in hole-based quantum devices, we further zoom in on the case of Pauli blockade of hole spins, assuming a strong transverse confinement potential. In this limit we also find an analytic expression for the current at low external magnetic field, that includes the effect of hyperfine coupling of the hole spins to randomly fluctuating nuclear spin baths. This result can be used to extract information about both hyperfine and spin-orbit coupling parameters for hole spins in devices with a significant fraction of nonzero nuclear spins.
Article
Full-text available
Strained semiconductor nanostructures can be used to make single-photon sources 1 , detectors 2 and photovoltaic devices 3 , and could potentially be used to create quantum logic devices 4,5. The development of such applications requires techniques capable of nanoscale structural analysis, but the microscopy methods 6–8 typically used to analyse these materials are destructive. NMR techniques can provide non-invasive structural analysis, but have been restricted to strain-free semiconductor nanostructures 9–11 because of the significant strain-induced quadrupole broadening of the NMR spectra 12–14. Here, we show that optically detected NMR spectroscopy can be used to analyse individual strained quantum dots. Our approach uses continuous-wave broadband radiofrequency excitation with a specially designed spectral pattern and can probe individual strained nanostructures containing only 1 3 10 5 quadrupole nuclear spins. With this technique, we are able to measure the strain distribution and chemical composition of quantum dots in the volume occupied by the single confined electron. The approach could also be used to address problems in quantum information processing such as the precise control of nuclear spins 15–17 in the presence of strong quadrupole effects 18–21. Our optically detected nuclear magnetic resonance (ODNMR) technique is used to examine two different types of strained semiconductor nanostructures: self-assembled InP/GaInP and InGaAs/GaAs quantum dots (for details on their structure and growth see Supplementary Section S1). All measurements were performed at T ¼ 4.2 K in an external magnetic field B z normal to the sample surface. Under excitation with circularly polarized light, nuclear spins become strongly polarized as a result of spin transfer from electrons via the hyperfine interaction 22. The resulting nuclear spin polarization on the dot is detected in photoluminescence of excitons in single quantum dots, as shown in Fig. 1a for InGaAs and Fig. 1b for InP dots in a high magnetic field of B z. 5 T. Each spectrum consists of an exciton Zeeman doublet with splitting E z. The detection of changes in E z allows measurement of the electron Overhauser shift 22,23
Article
Full-text available
Most zinc blende semiconductors have a single anion-like s state near the bottom of the valence band, found in density-of-states (DOS) calculations, and seen in photoemission. Here, we discuss the case where two s-like peaks appear, due to strong s-d coupling. Indeed, away from the k=0 Brillouin zone center, cation d states and anion s states can couple in zinc blende symmetry. Depending on the energy difference ΔEsd=Esanion-Edcation, this interaction can lead to either a single or two s-like peaks in the DOS and photoemission. We find four types of behaviors. (i) In GaP, GaAs, InP, and InAs, ΔEsd is large, giving rise to a single cation d peak well below the single anion s peak. (ii) Similarly, in CdS, CdSe, ZnS, ZnSe, and ZnTe, we see also a single s peak, but now the cation d is above the anion s. In both (i) and (ii) the s-d coupling is very weak. (iii) In GaN and InN, the local density approximation (LDA) predicts two s-like peaks bracketing below and above the cation d-like state. Correcting the too low binding energies of LDA by LDA+SIC (self-interaction correction) still leaves the two s-like peaks. The occurrence of two s-like peaks represents the fingerprint of strong s-d coupling. (iv) In CdTe, LDA predicts a single s-like peak just as in case (ii) above. However, LDA+SIC correction shifts down the cation d state closer to the anion s band, enhancing the s-d coupling, and leading to the appearance of two s-like peaks. Case (iv) is a remarkable situation where LDA errors cause not only quantitative energetic errors, but actually leads to a qualitative effect of a DOS peak that exists in LDA+SIC but is missing in LDA. We predict that the double-s peak should be observed in photoemission for GaN, InN, and CdTe.
Article
Full-text available
We study experimentally nuclear spin-pumping mechanisms in neutral InP/GaInP quantum dots under nonresonant optical excitation. We find two distinct regimes of dynamic nuclear polarization. At low optical powers when the dot is populated with “dark” excitons we observe nuclear spin polarization up to ~10% with direction insensitive to polarization and wavelength of light. Measurements of photoluminescence of both “dark” and “bright” excitons in single dots reveal that at low optical power nuclear spin pumping occurs via a virtual spin-flip transition between these states accompanied by photon emission. Under these conditions the sign of the nuclear spin polarization is determined by asymmetry in the exciton energy spectrum rather than by the sign of the exciton spin polarization. By contrast at high optical powers resulting in saturation of the quantum dot and suppression of exciton photoluminescence we detect nuclear spin polarization with direction and degree (up to ~50%) determined by the polarization of light.
Article
The longitudinal relaxation time T//1 of nuclear spins is calculated under the assumption that it is governed by the dipole-dipole interaction with holes in a semiconductor. The relaxation time T//1 is calculated for cubic and unaxial crystals with nondegenerate and degenerate carriers. The calculated results agree well with the available experimental data for silicon. It is shown that, for uniaxial semiconductors, the relaxation of a nuclear spin parallel to the c axis due to the scattering from holes is much greater than for a spin perpendicular to this direction.
Article
An empirical tight-binding method for tetrahedrally coordinated, cubic materials is presented and applied to group-IV and III-V semiconductors. The present spds * method extends existing calculations by the inclusion of all five d orbitals per atom in the basis set. On-site energies and two-center integrals between nearest neighbors in the Hamiltonian are fitted to measured energies, pseudopotential results, and the free-electron band structure. We demonstrate excellent agreement with pseudopotential calculations up to about 6 eV above the valence-band maximum even without inclusion of interactions with more distant atoms and three-center integrals. The symmetry character of the Bloch functions at the X point is considerably improved by the inclusion of d orbitals. Density of states, reduced masses, and deformation potentials are correctly reproduced.
Article
Modern electronic and optoelectronic devices are approaching nanometric dimensions where microscopic details cannot be treated in an effective way. Atomistic approaches become necessary for modelling structural, electronic and optical properties of such nanostructured devices. On the other hand, theoretical developments and numerical optimizations make device modelling approachable by atomistic methods. The purpose of this review is to report on microscopic theories to describe these nanostructured semiconductor devices. Empirical and density functional tight-binding as well as pseudopotential approaches are applied to the study of organic and inorganic semiconductor nanostructures and nanostructured devices. We show how these microscopic methods overcome the limitations imposed by the simplified approaches based on envelope function approximations and in the meantime keep the computational cost low. Typical calculations are shown for one-, two- and three-dimensional confined nanostructured devices, and comparisons with other approaches are outlined.
Article
Qubits, the quantum mechanical bits required for quantum computing, must retain their quantum states for times long enough to allow the information contained in them to be processed. In many types of electron-spin qubits, the primary source of information loss is decoherence due to the interaction with nuclear spins of the host lattice. For electrons in gate-defined GaAs quantum dots, spin-echo measurements have revealed coherence times of about 1mus at magnetic fields below 100mT (refs 1, 2). Here, we show that coherence in such devices can survive much longer, and provide a detailed understanding of the measured nuclear-spin-induced decoherence. At fields above a few hundred millitesla, the coherence time measured using a single-pulse spin echo is 30mus. At lower fields, the echo first collapses, but then revives at times determined by the relative Larmor precession of different nuclear species. This behaviour was recently predicted, and can, as we show, be quantitatively accounted for by a semiclassical model for the dynamics of electron and nuclear spins. Using a multiple-pulse Carr-Purcell-Meiboom-Gillecho sequence, the decoherence time can be extended to more than 200mus, an improvement by two orders of magnitude compared with previous measurements.
Article
The angular momentum contents of Bloch functions for a number of group IV and zincblende crystals are examined using a pseudopotential approach. The large amplitudes for d and f character functions found even for valence electron wave functions are related to the overlap of p and s symmetry states on different atoms. In zincblende crystals the strength of the d and f components increases around cations and decreases around anions for the valence bands. The f-like component is appreciable for the Γ1 conduction band particularly around a cation.
Article
Two sets of sharp emission lines, associated with the photoluminescence spectrum of the isoelectronic nitrogen trap in GaP, are unambiguously identified as the recombination of a second exciton bound to this center with an energy of 10 meV. This is the first observation of an excitonic molecule bound at a defect. The two electrons and two holes within this complex combine to form two antisymmetric states of angular momentum Jt=0 and Jt=2. The Jt=2 state is split 0.16 meV by the cubic crystal field, and the Jt=0 state lies 0.17 meV above the center of gravity of this doublet. Transitions from these states to the A and B states of a single exciton bound to nitrogen are seen. The complex Zeeman splittings predicted by this model agree in detail with experiment. The excitonic molecule is stable at low temperature; at 1.5°K, the excitonic molecule emission lines increase as the square of the single exciton intensity with increase in pumping power. However, nonradiative Auger recombination reduces the over-all nitrogen emission by a factor of 3 below the intensity at 4.2°K. The binding energy of the second exciton at the nitrogen trap is nearly equal to that of the single exciton. This remarkable fact may be possible only for an isoelectronic trap.
Article
Electronic structure and optical spectra of GaAs nanocrystals for a wide range of sizes are studied by using both sp3s* and sp3s*d5 nearest-neighbor tight-binding models. Our results show that the inclusion of d orbitals into a minimal basis set is necessary for a proper description of the lowest electron states, especially in the strong confinement regime. For dot sizes below 2.5 nm, the ground electron state is primarily built of L-point bulk band states, giving the nanocrystals indirect-gap character. Simpler sp3s* models yield an incorrect description of electron states made from bulk band states away from the Brillouin zone center. In contrast, sp3s*d5 models are able to provide a consistent picture of the main optical features in agreement with experiments.