ArticlePDF Available

Atomic clock transitions in silicon-based spin qubits

Authors:

Abstract and Figures

A major challenge in using spins in the solid state for quantum technologies is protecting them from sources of decoherence. This is particularly important in nanodevices where the proximity of material interfaces, and their associated defects, can play a limiting role. Spin decoherence can be addressed to varying degrees by improving material purity or isotopic composition, for example, or active error correction methods such as dynamic decoupling (or even combinations of the two). However, a powerful method applied to trapped ions in the context of atomic clocks is the use of particular spin transitions that are inherently robust to external perturbations. Here, we show that such 'clock transitions' can be observed for electron spins in the solid state, in particular using bismuth donors in silicon. This leads to dramatic enhancements in the electron spin coherence time, exceeding seconds. We find that electron spin qubits based on clock transitions become less sensitive to the local magnetic environment, including the presence of (29)Si nuclear spins as found in natural silicon. We expect the use of such clock transitions will be of additional significance for donor spins in nanodevices, mitigating the effects of magnetic or electric field noise arising from nearby interfaces and gates.
Content may be subject to copyright.
Atomic clock transitions in silicon-based spin qubits
Gary Wolfowicz,
1, 2,
Alexei M. Tyryshkin,
3
Richard E. George,
1
Helge Riemann,
4
Nikolai V. Abrosimov,
4
Peter Becker,
5
Hans-Joachim Pohl,
6
Mike L. W. Thewalt,
7
Stephen A. Lyon,
3
and John J. L. Morton
1, 8,
1
London Centre for Nanotechnology, University College London, London WC1H 0AH, UK
2
Dept. of Materials, Oxford University, Oxford OX1 3PH, UK
3
Dept. of Electrical Engineering, Princeton University, Princeton, New Jersey 08544, USA
4
Institute for Crystal Growth, Max-Born Strasse 2, D-12489 Berlin, Germany
5
Physikalisch-Technische Bundesanstalt, D-38116 Braunschweig, Germany
6
Vitcon Projectconsult GmbH, 07745 Jena, Germany
7
Dept. of Physics, Simon Fraser University, Burnaby, British Columbia V5A 1S6, Canada
8
Dept. of Electronic & Electrical Engineering, University College London, London WC1E 7JE, UK
(Dated: January 10, 2014)
A major challenge in using spins in the solid state for quantum technologies is protecting them
from sources of decoherence. This can be addressed, to varying degrees, by improving material
purity or isotopic composition [1, 2] for example, or active error correction methods such as dynamic
decoupling [3, 4], or even combinations of the two [5, 6]. However, a powerful method applied to
trapped ions in the context of frequency standards and atomic clocks [7, 8], is the use of particular
spin transitions which are inherently robust to external perturbations. Here we show that such ‘clock
transitions’ (CTs) can be observed for electron spins in the solid state, in particular using bismuth
donors in silicon [9, 10]. This leads to dramatic enhancements in the electron spin coherence time,
exceeding seconds. We find that electron spin qubits based on CTs become less sensitive to the local
magnetic environment, including the presence of
29
Si nuclear spins as found in natural silicon. We
expect the use of such CTs will be of additional importance for donor spins in future devices [11],
mitigating the effects of magnetic or electric field noise arising from nearby interfaces.
Out of the various candidates for solid state qubits,
spins have been of particular interest due to their rel-
ative robustness to decoherence compared to other de-
grees of freedom such as charge. So far, the most coher-
ent solid state systems investigated have been the spins
of well-isolated donors in bulk 28-silicon, with coherence
times (T
2
) of up to seconds (extrapolated) for the electron
spin [1] and minutes for the nuclear spin [5], comparable
to those of ion trap qubits [12, 13]. However, in practical
devices, spin coherence times are likely to be limited by
factors such as coupling to nearby qubits and magnetic
or electric field noise from the environment. For exam-
ple, cross-talk with other donors 100 nm away limits the
electron spin T
2e
to a few milliseconds [1], while a nearby
interface can limit the donor electron spin T
2e
to 0.3 ms
at 5.2 K [14]. Finally, without isotopic enrichment, the
5% natural abundance of
29
Si limits the electron spin
T
2e
to less than 1 ms [9, 10].
An approach to creating more robust qubits is to tune
free parameters of the system Hamiltonian to obtain in-
sensitivity to specific sources of decoherence. This has
been extensively used in ion trap qubits to protect against
magnetic field fluctuations [12, 13], building on work on
atomic clocks where hyperfine states, used as frequency
standards, must remain stable against such variations.
These so-called “clock transitions” (CTs) have a transi-
tion frequency (f ) which is insensitive to magnetic field
(B) variations, at least to first-order (in other words
df/dB = 0). More recently, superconducting circuit
qubits have also taken advantage of a tuned Hamiltonian
to remain immune to charge, flux or current noise [15, 16].
Nuclear spin CTs in rare-earth dopants (nuclear spins
I > 5/2) have been studied in the context of optical quan-
tum memories [17, 18] leading to a 600-fold improvement
of the coherence times to 150 ms, limited by second-order
effects, while recent experiments on phosphorus donor
nuclear spins also exploited a CT [5]. For electron spins
in the solid-state, CTs remain relatively unused due in
part to the requirement of a spin Hamiltonian of sufficient
complexity. One of the richest single-defect spin systems
is the bismuth donor in silicon (Si:Bi), which possesses an
electron spin S = 1/2 coupled to a nuclear spin I = 9/2.
The electron spin decoherence rates for Si:Bi have been
found to follow df/dB in both natural silicon [9], and
isotopically enriched
28
Si [19]. These results, combined
with the identification of a number of CTs in the spin
Hamiltonian of Si:Bi [20, 21], motivate the study of spin
coherence times around CTs in Si:Bi, where df/dB 0.
In this Letter, we investigate one such CT in Si:Bi, at
7.0317 GHz, using both natural silicon and
28
Si.
When describing the states of coupled electron and nu-
clear spins, two basis conventions are typically used: in
the high magnetic field limit, the electron and nuclear
spin projections m
S
and m
I
are good quantum numbers,
while in the zero-field limit, the total spin F (= I ± S)
and its projection m
F
(= m
S
+ m
I
) are used. CTs are
often found in an intermediate regime [22], nevertheless
it is possible to categorize them as nuclear magnetic reso-
nance (NMR)- or electron spin resonance (ESR)-type, on
the basis of whether the transition couples primarily to
S
x
or I
x
, where these are the electron and nuclear spin op-
erators perpendicular to the applied magnetic field. The
arXiv:1301.6567v1 [quant-ph] 28 Jan 2013
2
-10
10
0
Energy (GHz)
A B
|df/dB| (γ
e
)
0 100 200 300 400 500 600
0
0.2
1.0
0.4
0.6
0.8
0
20
100
40
60
80
|df/dB| (γ
n
)
70 74 78 82 86 90
7.0315
7.0320
7.0325
7.0330
7.0335
Transition frequency (GHz)
Transition
frequency (GHz)
5
10
–1 –2
F = 5
F = 4
m
F
=
ESR-type CTs
NMR-type CTs
X-band
FIG. 1. Electron spin resonance (ESR)-type clock transitions (CTs) of Si:Bi. A, The eigenstate energies (top) of
Si:Bi as function of magnetic field, the ESR- (black) and NMR-type (grey) transition frequencies between these states (middle),
and the first-order magnetic field dependence (df/dB) of these transition frequencies (bottom). ESR-type CTs (blue lines and
open circles) are found at 27, 80, 133 and 188 mT, and appear in the spectrum as doublets F m
F
= ±1 separated by up to
3 MHz. NMR-type CTs are found above 300 mT (red lines and open circles). B, Electron spin echo-detected magnetic field
sweeps around the 80 mT CT measured at microwave frequencies 7.0315 GHz. The transition probablities for ∆Fm
F
= +1
(dark blue) and 1 (light blue) transitions are equal near the CT.
ESR-type CTs which we investigate in this manuscript
involve states which are close to pure in the |F, m
F
i ba-
sis and hence for convenience we label them according to
the dominant |F, m
F
i component (full details are given
in the Supplementary Material and in Ref [20]).
For bismuth donors in silicon, NMR-type CTs can
be found at high field (> 350 mT) with frequencies
around 1 GHz as shown in red in Figure 1A. At low
field (< 200 mT), four ESR-type CTs are present with
frequencies in the range 5.2 to 7.3 GHz as shown in blue
in the same figure. We will focus here on the ESR-type
CTs, which possess only slightly reduced spin manipu-
lation time compared to free electron spins as well as a
large energy splitting even at low magnetic field (which
has interesting applications for use in hybrid supercon-
ducting circuits [9, 23, 24]).
In the silicon samples we study here, Bi donors were
introduced during crystal growth using the method de-
veloped in Ref [25], with concentrations ranging from
3.6 × 10
14
cm
3
to 4.4 × 10
15
cm
3
. Pulsed-ESR experi-
ments were performed using a spectrometer based around
a modified Bruker Elexsys E580 system with a 7 GHz
loop-gap cavity (for the CT) and 9.75 GHz dielectric res-
onator.
Figure 1B shows ESR spectra measured using mi-
crowave frequencies between 7.031 and 7.034 GHz, by
plotting electron spin echo intensity as a function of mag-
netic field. The spectra show two transitions correspond-
ing to [{F, m
F
} = 1, ±1}] and [{F, m
F
} =
1, 1}]; for brevity, these transitions can be distin-
guished by the value of the product F m
F
= ±1.
Together, they offer a controllable two-qubit subsystem
with low sensitivity to magnetic field fluctuations (see
inset of Figure 1B).
We model the ESR spectra using an isotropic spin
Hamiltonian common for group V donors in silicon:
H
0
= B
0
(γ
e
S
z
1 γ
n
1 I
z
) + A
~
S.
~
I (1)
where the two first terms correspond to the electronic
(S) and nuclear (I) spin Zeeman interactions with an
external field B
0
and the last term corresponds to the
hyperfine coupling A. A common way to estimate Hamil-
tonian parameters such as the electron and nuclear gyro-
magnetic ratios (γ
e
and γ
n
) and the hyperfine constant is
by measuring the magnetic field dependences of the spin
transition frequencies. We use the opportunity provided
by the CT (with df /dB 0) to extract a measure of the
hyperfine constant A = 1.47517(6) GHz with high preci-
sion, because uncertainties in the magnetic field become
irrelevant. In our simulations, we additionally use the
previously reported value of γ
e
= 27.997(1) GHz/T [26]
and the generic value of γ
n
= 7 MHz/T for
209
Bi [9].
Figure 1B shows that the ESR linewidth in the mag-
netic field domain increases around the CT: the deriva-
tive df /dB tends to zero hence its inverse, dB/df, di-
verges until it becomes limited by the non-linear terms
3
Resonant Bi
Central Bi
O-resonant Bi
209
Bi
SD (ID)dFF
{
29
Si
{
SD
{
z
S
z
S,
y
S
y
S+
x
S
x
S
z
I
z
S
{
SD (T
1e
, iFF)
z
S
z
S
T
2e
(s)
1
0.1
0.01
110
-1
10
-2
10
-3
|df/dB| (γ
e
)
10
3.6×10
14
cm
-3
2.0×10
15
cm
-3
4.4×10
15
cm
-3
dFF
iFF
ID
9.75 GHz
7 GHz
A
B
1 20 3 4 5
6
4
2
0
1/T
2e
(s
-1
)
Concentration (10
15
cm
-3
)
SD: spectral diusion
ID: instantaneous diusion
dFF/iFF: direct/indirect ip-op
FIG. 2. Decoherence mechanisms of Bi donors in sil-
icon and their dependence on df/dB. A, In the central
spin representation, a Bi donor is coupled to neighbouring Bi
donors as well as
29
Si spins. At the ESR CT, all spectral dif-
fusion (SD) contributions to decoherence are essentially elim-
inated, leaving only the direct flip-flop term (dFF) between
the central spin and a neighbouring, resonant Bi spin. B, T
2e
measurements at 4.8 K show a strong dependence on df/dB,
as shown for 3 different donor concentrations in
28
Si:Bi. Mea-
surements close to df /dB = γ
e
were taken using the ten X-
band ESR transitions, while the remaining points were taken
close to the CT. For each concentration, the dependence on
df/dB is modeled using contributions from ID, FF and iFF,
as shown separately in dashed lines for the lowest concen-
tration. Inset shows the limit of 1/T
2e
when approaching to
the exact CT as a function of donor concentration, showing a
nearly linear dependence, as expected for dFF.
in f(B
0
). These spectra are all well fit assuming a con-
stant linewidth in the frequency domain of 270 kHz. This
linewidth can be attributed to a distribution in the hy-
perfine constant of around 60 kHz, using f =
df
dA
A
at the CT. Fourier-Transform ESR performed at a range
of frequencies confirmed that the ESR linewidth in fre-
quency domain is indeed magnetic field independent (see
Supplementary Material).
We now examine the decoherence mechanisms which
affect the electron spin of donors in silicon. At sufficiently
low temperature (< 5 K), spin-lattice relaxation T
1e
can
be mostly neglected, and dipolar interactions ( 2S
z
S
z
(S
x
S
x
+ S
y
S
y
)) with neighbouring spins are the primary
source of decoherence. In a central spin representation,
as shown in Figure 2A, the surrounding spins can be
divided into three categories: i) resonant spins affected
by microwave excitation; ii) off-resonant spins of the same
species, i.e. Bi spins in m
F
levels not addressed by the
microwaves; and iii) other spin species such as
29
Si. Away
from CTs, the limiting factor for electron spin coherence
times is spectral diffusion (SD) from the S
z
S
z
term of the
dipolar interaction. This term can be assimilated into
effective fluctuations in the magnetic field environment
of the central spin. SD is independent of any frequency
detuning between spins and thus is valid between the
central spin and any others.
In the static case, dipolar couplings to (ii) and (iii)
can be refocused with a microwave π-pulse such as in
the Hahn echo sequence. However, this does not cor-
rect for the dipolar coupling between resonant spins (i)
as both spins are simultaneously flipped by the π-pulse.
This is called “instantaneous diffusion” (ID) and lim-
its T
2e
to 10 100 ms for typical donor concentra-
tions (> 10
14
cm
3
) [1, 19] [27]. Furthermore, dynamic
changes from spin flips in the environment cannot be refo-
cused. At high temperature, such flips arise from phonon
scattering but at low temperature, this is due to flip-flops
(FF) from the S
x
S
x
+ S
y
S
y
term of the dipolar interac-
tion. FF are energy conserving and as such are only rel-
evant between spins that have similar transition frequen-
cies. In natural silicon, the dominant decoherence mech-
anism is SD from
29
Si FF, while in isotopically enriched
28
Si, it arises from FF between resonant Bi spin pairs. In
the latter case, we distinguish between FF which involve
the central spin (direct FF, dFF), and those which do
not (indirect FF, iFF).
We begin by discussing results on samples of isotopi-
cally enriched
28
Si (100 ppm
29
Si). At the CT the transi-
tion frequency is insensitive to magnetic field fluctuations
in first order, so we expect SD to have little effect, leav-
ing only the dipolar coupling between resonant spin pairs.
With reference to Figure 2A, this implies then that all
terms apart from dFF vanish. In Figure 2B, measure-
ments of electron spin coherence times (T
2e
) are shown
for three different concentrations over a wide range of
df/dB. The data includes values measured at X-band as
well as those near the CT (∆F m
F
= +1) at 79.8 mT,
7.0317 GHz. Measurements at the CT shown here were
taken at 4.8 K where T
1e
= 9 s, however no increase in
T
2e
was seen at lower temperature.
For each sample, enhancements of about two orders of
magnitude are seen at the CT, compared to the case for a
free electron g-factor, such as that of phosphorus donors.
As shown in Figure 2B, the dependence of the measured
T
2e
on df/dB arises from two factors: the effect on ID,
and on iFF. ID has a known quadratic dependence on the
4
420 6
0
1
Echo signal (a.u.)
Time, 2τ (s)
0.2
0.4
0.6
0.8
T
2e
= 2.7 s
0.20.10 0.3
T
2e
= 93 ms
28
Si:Bi
nat
Si:Bi
A B
π
2
π
ττ
Time, 2τ (s)
FIG. 3. Hahn echo decay at the CT. A,
28
Si:Bi at 4.3 K
with a Bi concentration of 3.6 × 10
14
cm
3
. B,
nat
Si:Bi at
4.8 K with a Bi concentration of 10
15
cm
3
. The decay in
natural Si is a stretched exponential, and therefore T
2e
is de-
fined as the time when the amplitude reaches 1/e. Magnitude
detection was used to eliminate instrumental noise, likely due
to phase noise in the microwave source.
gyromagnetic ratio of the central spin [19, 28], and be-
comes a negligible effect for df/dB < 0.1γ
e
. Indirect FF
dephase the central spin through the S
z
S
z
term, giving
a linear dependence of T
2e
on df /dB. Direct FF, on the
other hand, are not eliminated at the CT, and provide
an upper bound on T
2e
for a given donor spin concentra-
tion, as plotted in the inset of Figure 2B. For the lowest
concentration sample, electron spin coherence times of
up to 2.7 s were measured from simple two-pulse Hahn
echo decays, as shown in Figure 3A.
We now turn to measurements on Bi-doped natural sil-
icon (
nat
Si:Bi), which has 5%
29
Si. Away from the CT
the effect of the
29
Si (I = 1/2) is both to broaden the
ESR linewidth to about 0.4 mT (equivalent to 12 MHz
in the frequency domain for a free electron) due to unre-
solved
29
Si hyperfine, as well as to limit the T
2e
to about
0.8 ms due to SD [9]. At the CT we find that the ESR
linewidth reduces to 500 kHz (see Supplementary Ma-
terial), within a factor of two of the value for enriched
28
Si material, while T
2e
increases by over two orders of
magnitude to about 90 ms (Figure 3B). The effect of
the suppression of SD around the CT has been simu-
lated for
nat
Si:Bi using cluster expansion methods [29],
though further refinements are required in the simula-
tion before a quantitative comparison can be made. The
stretched-exponential decay implies that T
2e
is still lim-
ited at 93 ms by SD from
29
Si due to the second order
term (d
2
f/dB
2
6= 0). For modest
28
Si enrichment (e.g.
[
29
Si] 1000 ppm), T
2e
should already exceed seconds,
and indeed there may be an optimal
28
Si purity above
which T
2e
at the CT drops, due to the role of
29
Si or
30
Si
in detuning otherwise-identical spins [30].
We have shown how CTs in Si:Bi can be used to pro-
duce magnetic field-insensitive spin qubits with directly
measured coherence times of several seconds. Such qubits
would be insensitive to magnetic field noise arising, for
example, from fluctuating dangling-bond spins at the
Si/SiO
2
interface. Conversely, if electric field noise is
dominant, this can couple to donor spins via the hyper-
fine interaction and cause decoherence. Again, CTs can
be designed to be immune from electric charge noise by
selecting points where df/dA 0 (see Supplementary
Material). Through the use of CTs, it is likely that the
seconds-long electron spin coherence times measured in
the bulk can be harnessed for spins in practical quantum
devices.
We thank Stephanie Simmons, Tania Monteiro and Se-
trak Balian for fruitful discussions. This research is sup-
ported by the EPSRC through the Materials World Net-
work (EP/I035536/1) and a DTA, as well as by the Eu-
ropean Research Council under the European Commu-
nity’s Seventh Framework Programme (FP7/2007-2013)
/ ERC grant agreement no. 279781. Work at Princeton
was supported by NSF through Materials World Net-
work (DMR-1107606) and through the Princeton MR-
SEC (DMR-0819860), and also by NSA/LPS through
LBNL (6970579). J.J.L.M. is supported by the Royal
Society.
gary.wolfowicz@materials.ox.ac.uk
jjl.morton@ucl.ac.uk
[1] A. M. Tyryshkin et al., Nature Materials 11, 143 (2012).
[2] G. Balasubramanian et al., Nature materials 8, 383
(2009).
[3] L. Viola and S. Lloyd, Physical Review A 58, 2733
(1998).
[4] H. Bluhm et al., Nature Physics 7, 109 (2011).
[5] M. Steger et al., Science (New York, N.Y.) 336, 1280
(2012).
[6] P. C. Maurer et al., Science (New York, N.Y.) 336, 1283
(2012).
[7] J. Bollinger, J. Prestage, W. Itano, and D. Wineland,
Physical Review Letters 54, 1000 (1985).
[8] P. Fisk et al., IEEE Transactions on Instrumentation and
Measurement 44, 113 (1995).
[9] R. E. George et al., Phys. Rev. Lett. 105, 67601 (2010).
[10] G. W. Morley et al., Nature Materials 9, 725 (2010).
[11] J. J. Pla et al., Nature 489, 541 (2012).
[12] P. Haljan et al., Physical Review A 72, 062316 (2005).
[13] C. Langer et al., Physical Review Letters 95, 060502
(2005).
[14] T. Schenkel et al., Applied Physics Letters 88, 112101
(2006).
[15] D. Vion et al., Science (New York, N.Y.) 296, 886 (2002).
[16] J. Koch et al., Physical Review A 76, 042319 (2007).
[17] J. Longdell, A. Alexander, and M. Sellars, Physical Re-
view B 74, 195101 (2006).
[18] D. McAuslan, J. Bartholomew, M. Sellars, and J.
Longdell, Physical Review A 85, 032339 (2012).
[19] G. Wolfowicz et al., Physical Review B 86, 245301
(2012).
[20] M. Mohammady, G. W. Morley, and T. S. Monteiro,
Phys. Rev. Lett. 105, 067602 (2010).
[21] M. H. Mohammady, G. W. Morley, A. Nazir, and T. S.
5
Monteiro, Phys. Rev. B 85, 094404 (2012).
[22] G. W. Morley et al., Nature Materials 11, 1 (2012).
[23] D. Schuster et al., Phys. Rev. Lett. 105, 140501 (2010).
[24] Y. Kubo et al., Phys. Rev. A 85, 012333 (2012).
[25] H. Riemann, N. Abrosimov, and N. Noetzel, ECS Trans-
actions 3, 53 (2006).
[26] G. Feher, Phys. Rev. 114, 1219 (1959).
[27] By reducing the microwave power and effectively flipping
only a small part of the resonant spins, it is possible to
obtain an extrapolation of coherence times in the limit of
no ID [31, 32]. However, it is not a solution to overcoming
the effect of ID in practice.
[28] A. Schweiger and G. Jeschke, Principles of pulse elec-
tron paramagnetic resonance (Oxford University Press,
Oxford, 2001).
[29] S. J. Balian et al., Phys. Rev. B 86, 104428 (2012).
[30] W. Witzel et al., Physical Review Letters 105, 187602
(2010).
[31] J. Klauder and P. Anderson, Physical Review 125, 912
(1962).
[32] K. Salikhov, Journal of Magnetic Resonance (1969) 42,
255 (1981).
Supplementary Material
ESR- AND NMR-TYPE MAGNETIC FIELD “CLOCK” TRANSITIONS (CT)
A B
F + 1
F
m
F
m
F
-1
ESR-type CTs
Magnetic eld (mT)
0 50 100 150 200 250
-5
5
0
Energy (GHz)
-2
-4
-6
-8
ESR-type CTs
L-Z anticrossings
= +1
F
mF
1=
F
mF
FIG. 1. Description of ESR-type CTs. A, The eigenstate energies of Si:Bi as function of
magnetic field. The color scale shows the logarithmic distance to pure Bell states (Landau-
Zener (L-Z) anticrossings) in the |m
S
, m
I
i basis, defined as log(|θ π/4|) for an eigenstate
|Φi = cos(θ)
1
2
, m
I
±
1
2
± sin(θ)
±
1
2
, m
I
1
2
. The Bell state at 0 mT is barely visible here
due to degeneracy. B, ESR-type CTs with F m
F
= +1 in dark blue and 1 in light blue. The
four involved eigenstates, which are two pairs of hyperfine coupled states in the |m
S
, m
I
i basis,
form a subspace of the Hilbert space.
Below we discuss the general requirements for ESR and NMR-type CTs for systems with
electron spin S = 1/2 and nuclear spin I and assuming an isotropic hyperfine coupling,
with a particular focus on Group V donors in silicon. This letter is the first measurement of
ESR-type CTs to our knowledge, though they have been theoretically described by Moham-
mady et.al. [1, 2] for donors in Si, in particular Bi. On the other hand, NMR-type CTs have
been used in various systems in the past [3, 4], including in phosphorus donors in silicon [5],
to reduce sensitivity to magnetic field inhomogeneities (i.e. increase frequency resolution)
or to increase nuclear coherence times.
In the basis of the electron and nuclear spin |m
S
, m
I
i, the isotropic hyperfine interaction
(A
~
I ·
~
S) couples pairs of states within the Hilbert space such that [∆m
S
= ±1, m
I
= 1].
In the strongly coupled electron-nuclear spin basis |F, m
F
i (F = I ±S, m
F
= m
S
+m
I
), these
pairs of states share the same m
F
value. When the static magnetic field is increased, the
Zeeman energy rises to the same order of magnitude as the hyperfine interaction, resulting
1
in avoided Landau-Zener crossings between states with m
F
0 as shown in Figure 1A.
ESR-type CTs are located between pairs of these avoided crossings.
At the intermediate fields relevant to CTs and Landau-Zener anticrossing, the eigenstates
can be expressed either as being close to Bell states in the |m
S
, m
I
i basis, or still quite pure
in the |F, m
F
i basis. For example, the CT at 7.0317 GHz, explored in the main text,
connects the following pairs of states:
0.74
1
2
,
5
2
+ 0.67
1
2
,
3
2
0.74
1
2
,
3
2
+ 0.67
1
2
,
1
2
in the |m
S
, m
I
i basis
0.99 |4, 2i + 0.15 |5, 2i 0.15 |4, 1i + 0.99 |5, 1i in the |F, m
F
i basis
Hence, for convenience, we refer to these states by the dominant term in the |F, m
F
i basis
(i.e. |4, 2i and |5, 1i in the example above). The ∆F m
F
= +1 and ∆F m
F
= 1 CTs
are each transitions between one state of the first Landau-Zener crossing to a second state
of the second crossing, forming a 4-dimensional subspace of the Hilbert space (Figure 1B).
For non-integer [integer] values of I, there are I + 1/2 [I] Landau-Zener anticrossings and
consequently 2(I 1/2) [2I] ESR-type CTs. The minimum complexity required is thus a
nuclear spin I 1 to have at least two pairs of hyperfine coupled states, which is not the
case for phosphorus donors (
31
P has I = 1/2).
Arsenic (
75
As, I = 3/2) and antimony (
121
Sb, I = 5/2 and
123
Sb, I = 7/2) have sufficient
nuclear spin to permit ESR-type CTs, however the hyperfine coupling is relatively weak
(A 198, 186 and 101 MHz, respectively), so the avoided crossings are found at low
magnetic field and transition frequencies (see Table I). Bismuth
209
Bi with I = 9/2 and
A = 1.475 GHz is thus the optimal Group V donor in silicon from the point of view of CTs,
possessing four of them at GHz frequencies.
75
As (I = 3/2)
121
Sb (I = 5/2)
123
Sb (I = 7/2)
209
Bi (I = 9/2)
F = +1, m
F
= 1 0 1 0 2 1 1 0 2 1 3 2 1 0 2 1 3 2 4 3
Magnetic field (mT) 3.8 3.4 10.4 1.8 5.5 9.3 26.6 79.8 133.3 187.8
Frequency (GHz) 0.384 0.552 0.482 0.403 0.376 0.314 7.338 7.032 6.372 5.214
TABLE I. Summary of ESR-type magnetic-field CTs in donors in silicon. CTs exist in
pairs (∆F m
F
= ±1) separated by less than 0.15 mT in magnetic field and 3 MHz in frequency.
Phosphorus does not possess any CT due to its small nuclear spin (I = 1/2).
As df/dB 0, the next figure of merit is the electron transition probability amplitude,
2
which is always 50% of the high field limit for ESR-types. This means that manipulation
times are only slightly reduced while electron coherence times T
2e
are drastically increased.
Conversely, as flip-flops between two donor electron spins follow the square of this probability
amplitude, the flip-flopping rate remains strong, limiting T
2e
as observed and explained in
the main text.
NMR-type CTs occur at strong magnetic field (in Si:Bi, 0.3 T < B
0
< 5 T, from m
I
=
9/2 at low field to m
I
= 9/2 at high field) and as such are transitions between quasi-
pure states in the (m
S
, m
I
) basis. NMR-type CTs possess a change in nuclear spin state
of m
I
= 1 and can be manipulated in a conventional electron nuclear double resonance
(ENDOR) or NMR experiment. Amongst the NMR-type CTs, those at higher magnetic
fields have a smaller electron spin component. This leads to a reduced coupling to the
environment, but also increased spin manipulation times, converging to that of a regular
NMR transition (typically 10 µs).
ESR LINEWIDTHS
Measurements of spin linewidths provide important details regarding the spin environ-
ment, yielding information on crystalline defects and strains amongst other properties. Line
broadening arises from a variation in either the hyperfine interaction (∆A) or the magnetic
field (∆B) (or g-tensor) across the sample. They can either be measured in a magnetic
field-swept spectrum where (to first order):
B
T otal
= ∆B +
dB
df
df
dA
A (1)
or in Fourier-Transform (FT) ESR where the FT of the free induction decay after a π/2
rotation gives the spectrum in the frequency domain. In this case, (to first order):
f =
df
dB
B +
df
dA
A (2)
At the CT where dB/df , the linewidth in the magnetic field domain broadens
strongly until it becomes limited by the second order dependence on magnetic field. Both
B and A can be identified by varying the transition frequency and fitting numerically
knowing
dB
df
df
dA
. To confirm that this change in linewidth is strictly related to the term df/dB,
we can make use of the FT ESR technique which simplifies at the CT to ∆f =
df
dA
A, where
3
0.5 1 1.5 2 2.5
0.2
0.4
0.6
0.8
Frequency (MHz)
FT−ESR intensity (a.u.)
80.4 mT, 7.0306 GHz
45.7 mT, 7.0970 GHz
0
1
= +1
F
mF
1=
F
mF
0.2
0.4
0.6
0.8
FT−ESR intensity (a.u.)
0
1
2 4 6 80
Frequency (MHz)
0 3
28
Si:Bi
nat
Si:Bi
80.4 mT, 7.0270 GHz
= +1
F
mF
1=
F
mF
BA
= 0µ
= 1µ
= 2µ
Sum
Model (normalized):
= 3µ
FIG. 2. FT-ESR around the CT. Each spectrum is the FT of the free induction decay taken
using a microwave frequency slightly below resonance (values given in legend). A, For the case of
28
Si:Bi, we observe two peaks in the ESR spectrum around the CT, corresponding to the transitions
F m
F
= ±1. Spectra are shown as measured at two settings of magnetic field/microwave
frequency. In the magnetic field domain, the ESR linewidths in these two cases are 1.6 mT close to
the CT and 0.07 mT farther away (see Figure 1 of the main manuscript), however in the frequency
domain as shown above, the ESR linewidths are constant. B, In
nat
Si:Bi, these two primary ESR
transitions are further split into sub-peaks, corresponding to a mass-effect from nearest neighbour
Si atoms. Each shift of one neutron mass (∆µ) yields a shift of 1.7 MHz in transition frequency
(or 0.024% change in the hyperfine coupling A). Dashed lines show simulated peaks whose intensity
is calculated from a trinomial distribution of
30
Si,
29
Si and
28
Si isotopes in
nat
Si (with respective
concentration 3.1%, 4.7% and 92%). As the FT-ESR is derived from the free induction decay, the
intensities are normalized by the FT of the inhomogeneous decay T
2e
(Lorentzian) and the cavity
bandwidth.
df/dA is quasi-constant around the CT. In Figure 2A, the linewidth is indeed constant about
270 kHz.
In natural silicon, the elimination of the B term dramatically reduces the ESR
linewidth. Away form the CT (e.g. at X-band), B is normally around 4 G due to unre-
solved coupling to
29
Si nuclear spins. In the frequency domain, this would be equivalent to
nearly 12 MHz, hiding multiple spectral features. First, the two transitions F m
F
= ±1
would not be resolvable. Second, as show in Figure 2B, we observe several other peaks (ab-
sent in isotopically pure
28
Si samples) which arise from variations in the hyperfine coupling
4
due to the total mass of nearest-neighbour silicon atoms (
28
Si,
29
Si and
30
Si). This effect is
described in full detail, including ENDOR experiments, in a forthcoming work [6].
ELECTRIC FIELD CLOCK TRANSITIONS
-10
10
0
Energy (GHz)
5
5
15
-15
Magnetic eld (mT)
0 200 400 600 800 1000
ESR-type CTs
NMR-type CTs
0
dA
df
FIG. 3. Clock transitions in Si:Bi where df/dA = 0 which should be robust to electric
field noise. Both ESR- and NMR-type CTs can be observed, as for magnetic-field CTs. One
further ESR-CT is found at higher magnetic fields (2.6 T, not shown).
In the experiments reported in the main text, CTs were used to reduce the sensitivity
of electron spin to magnetic field variations, as quantified by df/dB. While magnetic field
noise is indeed the main decoherence mechanism in bulk materials, this may not be the case
in nanoscale devices where the electric field at interfaces could couple strongly with both
the donor electron and nuclear spins through the hyperfine interaction (and also, to lesser
extent, through a modulation in the electron spin g-factor). The sensitivity of a spin to
this effect can be quantified by the gradient of the frequency with respect to the hyperfine
constant df/dA, combined with values for the DC Stark effect for donors in silicon (which
for Group V donors is in the order of 10
3
µm
2
/V
2
, as a fractional change in the hyperfine
coupling [7, 8]). Those CTs which will be most robust to electric field noise (df/dA 0)
are identified in Si:Bi in Figure 3 and in Table II for all Group V donors in silicon.
In practice, both magnetic and electric field fluctuations will participate to the donor spin
decoherence. There will thus be an optimal CT, at a specific magnetic field and frequency,
where the coherence time would be maximum. For example, the magnetic field CT near
188 mT in Bi has the lowest value of df /dA out of the four possible CTs. In other scenarios,
it might be advantageous to minimise the inhomogeneous broadening as much as possible
5
75
As (I = 3/2)
121
Sb (I = 5/2)
123
Sb (I = 7/2)
209
Bi (I = 9/2)
m
S
= +1, m
I
= 1/2 1/2 3/2 1/2 3/2 5/2 1/2 3/2 5/2 7/2
Magnetic field (mT) 53 117 39 114 38 23 2607 868 519 369
Frequency (GHz) 1.43 3.21 0.92 3.17 0.98 0.49 72.64 23.18 12.57 7.30
TABLE II. Summary of ESR-type electric-field CTs in donors in silicon. At the given
magnetic fields, the electron and nuclear spins are weakly coupled and the eigenstates must thus
be expressed in the |m
S
, m
I
i basis. The [∆m
S
= ±1, m
I
= 2] (∆F m
F
= 1) transitions
are nearly completely forbidden here; they would have been found at the same magnetic field as
the [∆m
S
= ±1, m
I
= 0] (∆F m
F
= +1) transitions, but separated by less than 40 MHz in
frequency.
(e.g. for coupling a spin ensemble to a microwave resonator), and this would also require
different optimal operating points within the Hilbert space of the bismuth electron and
nuclear spins.
[1] M. Mohammady, G. W. Morley, and T. S. Monteiro, Phys. Rev. Lett. 105, 067602 (2010).
[2] M. H. Mohammady, G. W. Morley, A. Nazir, and T. S. Monteiro, Phys. Rev. B 85, 094404
(2012).
[3] W. Hardy, A. Berlinsky, and L. Whitehead, Physical Review Letters 42, 1042 (1979).
[4] J. Longdell, A. Alexander, and M. Sellars, Physical Review B 74, 195101 (2006).
[5] M. Steger et al., Journal of Applied Physics 109, 102411 (2011).
[6] A. Tyryshkin et al., in preparation (2013).
[7] F. Bradbury et al., Physical Review Letters 97, 176404 (2006).
[8] R. Rahman et al., Physical Review Letters 99, 36403 (2007).
6
... For example, 209 Bi is a spin-9 / 2 particle with nearly 100% natural abundance. The parallel component of the hyperfine coupling between the electron and Bi nuclear spins, 1.27 GHz, is similar to Bi donors in Si 29 . Hence, the combined electron-nuclear system exhibits 30 energy levels (see Fig. 5a) that are separately addressable in experiments, providing a broad space of spin states accessible for the design of quantum technologies. ...
... The strong electron-nuclear spin coupling in Bi Ca V O À leads to a set of avoided crossings between energy levels as a function of the magnetic field. The spin transitions between these levels, known as "clock transitions" (CT) 29 , are remarkably robust to external perturbations, and thus the coherence time of qubits operating at a CT can be substantially increased 30,31 . Using the cluster-correlation expansion method (CCE), implemented in the PyCCE code 21 , we computed the coherence of the spin qubit Bi Ca V O À near CTs (see Fig. 5b). ...
Article
Full-text available
Virtually noiseless due to the scarcity of spinful nuclei in the lattice, simple oxides hold promise as hosts of solid-state spin qubits. However, no suitable spin defect has yet been found in these systems. Using high-throughput first-principles calculations, we predict spin defects in calcium oxide with electronic properties remarkably similar to those of the NV center in diamond. These defects are charged complexes where a dopant atom — Sb, Bi, or I — occupies the volume vacated by adjacent cation and anion vacancies. The predicted zero phonon line shows that the Bi complex emits in the telecommunication range, and the computed many-body energy levels suggest a viable optical cycle required for qubit initialization. Notably, the high-spin nucleus of each dopant strongly couples to the electron spin, leading to many controllable quantum levels and the emergence of atomic clock-like transitions that are well protected from environmental noise. Specifically, the Hanh-echo coherence time increases beyond seconds at the clock-like transition in the defect with ²⁰⁹Bi. Our results pave the way to designing quantum states with long coherence times in simple oxides, making them attractive platforms for quantum technologies.
... The sensitivity of quantum sensors typically scales with the sensor's coherence time and is thus also limited by decoherence [5]. While fabrication efforts focus on minimizing noise in quantum devices [6] and quantum error correction techniques allow detection and correction of noise-induced errors [7,8], several important strategies, such as decoherence-free spaces [9], clock transitions [10], dynamical decoupling [11], and composite pulses [12][13][14][15][16][17][18][19], reduce the effect of noise, lowering decoherence and control error rates [20]. ...
Article
Full-text available
Decoherence and imperfect control are crucial challenges for quantum technologies. Common protection strategies rely on noise temporal autocorrelation, which is not optimal if other correlations are present. We develop and demonstrate experimentally a strategy that uses the cross-correlation of two noise sources. Utilizing destructive interference of cross-correlated noise extends the coherence time tenfold, improves control fidelity, and surpasses the state-of-the-art sensitivity for high frequency quantum sensing, significantly expanding the applicability of noise protection strategies. Published by the American Physical Society 2024
... GHz. This creates interesting spin physics phenomena related to the strong electron-nuclear mixing 33,36 . However, in single-donor experiments where the nuclear spin is read out via the ancilla electron, the stronger hyperfine coupling introduces a measurement back-action (in other words, a deviation from the QND measurement condition) of order ðA=γ e B 0 Þ 2 37 , which is thus over two orders of magnitude larger in 209 Bi compared to 123 Sb. ...
Article
Full-text available
Efficient scaling and flexible control are key aspects of useful quantum computing hardware. Spins in semiconductors combine quantum information processing with electrons, holes or nuclei, control with electric or magnetic fields, and scalable coupling via exchange or dipole interaction. However, accessing large Hilbert space dimensions has remained challenging, due to the short-distance nature of the interactions. Here, we present an atom-based semiconductor platform where a 16-dimensional Hilbert space is built by the combined electron-nuclear states of a single antimony donor in silicon. We demonstrate the ability to navigate this large Hilbert space using both electric and magnetic fields, with gate fidelity exceeding 99.8% on the nuclear spin, and unveil fine details of the system Hamiltonian and its susceptibility to control and noise fields. These results establish high-spin donors as a rich platform for practical quantum information and to explore quantum foundations.
Article
A relativistic magnetic hyperfine interaction Hamiltonian based on the Douglas–Kroll–Hess (DKH) theory up to the second order is implemented within the ab initio multireference methods, including spin–orbit coupling in the Molcas/OpenMolcas package. This implementation is applied to calculate relativistic hyperfine coupling (HFC) parameters for atomic systems and diatomic radicals with valence s or d orbitals by systematically varying active space size in the restricted active space self-consistent field formalism with restricted active space state interaction for spin–orbit coupling. The DKH relativistic treatment of the hyperfine interaction reduces the Fermi contact contribution to the HFC due to the presence of kinetic factors that regularize the singularity of the Dirac delta function in the nonrelativistic Fermi contact operator. This effect is more prominent for heavier nuclei. As the active space size increases, the relativistic correction of the Fermi contact contribution converges well to the experimental data for light and moderately heavy nuclei. The relativistic correction, however, does not significantly affect the spin-dipole contribution to the hyperfine interaction. In addition to the atomic and molecular systems, the implementation is applied to calculate the relativistic HFC parameters for large trivalent and divalent Tb-based single-molecule magnets (SMMs), such as Tb(III)Pc2 and Tb(II)(CpiPr5)2 without ligand truncation using well-converged basis sets. In particular, for the divalent SMM, which has an unpaired valence 6s/5d hybrid orbital, the relativistic treatment of HFC is crucial for a proper description of the Fermi contact contribution. Even with the relativistic hyperfine Hamiltonian, the divalent SMM is shown to exhibit strong tunability of HFC via an external electric field (i.e., strong hyperfine Stark effect).
Article
Silicon is a key semiconducting material for electrical devices and hybrid quantum systems where low temperatures and zero-spin isotopic purity can enhance quantum coherence. Electrical conductivity in Si is characterized by carrier freeze out at around 40 K allowing microwave transmission, which is a key component for addressing spins efficiently in silicon quantum technologies. In this work, we report an additional sharp transition of the electrical conductivity in a Si-28 cylindrical cavity at around 1 K. This is observed by measuring microwave resonator whispering gallery mode frequencies and Q factors with changing temperature and comparing these results with finite-element models. We attribute this change to a transition from a relaxation mechanism-dominated to a resonant phononless absorption-dominated hopping conduction regime. Characterizing this regime change represents a deeper understanding of a physical phenomenon in a material of high interest to the quantum technology and semiconductor device community and the impact of these results is discussed.
Preprint
Full-text available
The usefulness of solid-state spins in quantum technologies depends on how long they can remain in a coherent superposition of quantum states. This Colloquium discusses how first-principles simulations can predict spin dynamics for different types of solid-state electron spins, helping design novel and improved platforms for quantum computing, networking, and sensing. We begin by outlining the necessary concepts of the noise affecting generic quantum systems. We then focus on recent advances in predicting spin-phonon relaxation of the spin-defect qubits. Next, we discuss cluster methods as a means of simulating quantum decoherence induced by spin-spin interactions, emphasizing the critical role of validation in ensuring the accuracy of these simulations. We highlight how validated cluster methods can be instrumental in interpreting recent experimental results and, more importantly, predicting the coherence properties of novel spin-based quantum platforms, guiding the development of next-generation quantum technologies.
Article
Full-text available
The intrinsic quantum nature of molecules opens exciting opportunities for developing the field of quantum information science. In this context, porphyrins stand out as ideal building blocks for quantum technologies thanks to their unique optical and electrical properties as well as their capacity to accommodate metal atoms and ions. This review bridges the chemistry and physics of porphyrins, providing an overview of recent advances in porphyrin‐based molecular qubits. Starting from qubits, the review explores the potential of porphyrin units to combine, leading to the formation of quantum logic gates and hierarchical higher‐dimensional structures. Next, the exploitation of porphyrins' unique photophysical properties for realizing long‐lived high spin states is examined. These states are promising for the photogeneration of multi‐level systems and the optical initialization and control of molecular qubits. With a critical eye on the current state‐of‐the‐art, the review elucidates the future perspectives of porphyrins for advancing quantum technologies.
Article
Molecular lanthanide (Ln) complexes are promising candidates for development of next-generation quantum technologies. High-symmetry structures incorporating integer spin Ln ions can give rise to well-isolated crystal field quasi-doublet ground states, i.e., quantum two-level systems that may serve as the basis for magnetic qubits. Recent work has shown that symmetry lowering of the coordination environment around the Ln ion can produce an avoided crossing, or clock transition within the ground doublet, leading to significantly enhanced coherence. Here, we employ single-crystal high-frequency electron paramagnetic resonance spectroscopy and high-level ab initio calculations to carry out a detailed investigation of the nine-coordinate complexes, [Ho3+L1L2], where L1= 1,4,7,10-tetrakis(2-pyridylmethyl)-1,4,7,10-tetraaza-cyclododecane) and L2 = F– (1) or[MeCN]0 (2). Meanwhile, off-diagonal crystal field interactions give rise to a giant 116.4±1.0 GHz clock transition within this doublet. We then demonstrate targeted crystal field engineering of the clock transition by replacing F– with neutral MeCN (2), resulting in an increase in the clock transition frequency by a factor of 2.2. This tunability is highly desirable because decoherence caused by 2nd-order sensitivity to magnetic noise scales inversely with the clock transition frequency.
Article
Full-text available
Electron paramagnetic resonance on chromium doped SrTiO3 samples grown using the Verneuil technique shows the presence of charge-compensated Cr3+-VO as one of the dominant chromium centers. The spin-lattice relaxation processes have been investigated in samples with both isotropic Cr3+ and Cr3+-VO centers in heavily doped SrTiO3. The relaxation of longitudinal magnetization was dominated by the sum of two exponentials with two time constants (i.e., a slow and a fast constant) at liquid-helium temperatures. The results of fitting the temperature variation of T1 suggest that the dominant exponential contribution is related to the spin-phonon relaxation time arising from the local phonon mode.
Article
Full-text available
We implanted ultralow doses (2×1011 cm−2) of antimony ions () into isotopically enriched silicon () and find high degrees of electrical activation and low levels of dopant diffusion after rapid thermal annealing. Pulsed electron spin resonance shows that spin echo decay is sensitive to the dopant depths, and the interface quality. At 5.2 K, a spin decoherence time, T2, of 0.3 ms is found for profiles peaking 50 nm below a Si/SiO2 interface, increasing to 0.75 ms when the surface is passivated with hydrogen. These measurements provide benchmark data for the development of devices in which quantum information is encoded in donor electron spins.
Article
Full-text available
Pulsed magnetic resonance allows the quantum state of electronic and nuclear spins to be controlled on the timescale of nanoseconds and microseconds respectively. The time required to flip dilute spins is orders of magnitude shorter than their coherence times, leading to several schemes for quantum information processing with spin qubits. Instead, we investigate 'hybrid nuclear-electronic' qubits consisting of near 50:50 superpositions of the electronic and nuclear spin states. Using bismuth-doped silicon, we demonstrate quantum control over these states in 32 ns, which is orders of magnitude faster than previous experiments using pure nuclear states. The coherence times of up to 4 ms are five orders of magnitude longer than the manipulation times, and are limited only by naturally occurring (29)Si nuclear spin impurities. We find a quantitative agreement between our experiments and an analytical theory for the resonance positions, as well as their relative intensities and Rabi oscillation frequencies. These results bring spins in a solid material a step closer to research on ion-trap qubits.
Article
Monocrystalline, high purity FZ-Silicon doped with bismuth is a promising laser material for the not jet opened up THz- band. With pill-doping during the common FZ-process only a narrow axial part of the crystal could be doped at a sufficient level. Crystals with FZ-like purity and a longer, highly doped section were grown after the well known pedestal technique using a specially developed pill-doping method out from axial holes in the pedestal feed rod being caped by the molten zone. Also doping with lithium and magnesium was done In this way for the first time. The crystals had diameters of 15-20mm using slim 35mm feed rods. Therefore up-scaling of the growth process and, at least for Bi doping, a dislocation-free structure appear achievable.
Article
We present characterization of the hyperfine interaction for the europium in hydrated europium chloride and as a dopant in yttrium orthosilicate. The Zeeman and pseudoquadrupole tensors were determined by measuring the hyperfine splittings while rotating the direction of a weak (˜300G) magnetic field. The hyperfine spectra were recorded using Raman-heterodyne spectroscopy, an radio-frequency optical double resonance technique. For both materials magnetic field values were identified where there is no first order Zeeman shift in a europium hyperfine transition. These field insensitive transitions will have application in achieving very long hyperfine coherence times.
Article
The electron spin-echo (ESE) signal decay of paramagnetic centers induced by their dipole-dipole interaction modulated simultaneously by the processes of random spin flips (the spectral diffusion mechanism) and microwave pulses (the instantaneous diffusion mechanism) is analyzed. The decay kinetics is calculated numerically for the cases in which the spin-flip rate W and the time of the decay kinetics observation τ are interconnected as Wτ ≈ 1; analytical expressions are obtained for Wτ ⪢1, Wτ ⪡ 1. Analogous calculations are carried out for spin-flip rate distributions. The calculated data are compared with experiment for the decay kinetics of SO4− ion radicals studied over a wide range of Tτ values. The calculations show a fairly good fit to experiment.
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
In spin resonance experiments random flipping by T1 or T2 processes of nearby, nonresonant spins introduces fluctuations into the precessional frequency of the observed spins. These fluctuations may be described by means of a stochastic model, and for wide classes of both Markoffian and non-Markoffian distributions we make predictions for the line shape, for the free induction decay, and for various spin-echo signals. If the homogeneous broadening of the line is due to a dipolar interaction term, then we find that the conditional distribution for the precessional frequency has the shape of a Lorentzian with a cutoff on the wings, rather than a Gaussian shape as commonly assumed. The causes and consequences of Lorentzian diffusion are analyzed in detail for samples in which T1 processes control the source of local frequency fluctuations and for samples in which T2 processes dominate. Recent two- and three-pulse spin-echo experiments of Mims et al. dramatically confirm the predictions of Lorentzian diffusion for electron paramagnetic resonances in samples with temperature-dependent diffusion, as well as with temperature-independent diffusion. "Instantaneous" diffusion caused by the action of the applied pulses is predicted by our model and explains features of Mims' data. The generality of our principal results still permits the outcome of various resonance experiments to be predicted, even when a simple dipolar interaction is no longer an adequate model.
Article
The ground-state wave function of the antimony, phosphorus, and arsenic impurities in silicon has been investigated by means of the electron nuclear double resonance (ENDOR) method. By this method the hyperfine interactions of the donor electron with the Si29 nuclei situated at different lattice sites were obtained. The isotropic part of the hyperfine interaction agreed with the theory of Kohn and Luttinger to better than 50%. From a comparison of the experimental results with their theory a value for the conduction band minimum in silicon of k0k=0.85+/-0.03 was obtained. So far no satisfactory theory exists to account quantitatively for the observed anisotropic part of the hyperfine interaction. The observed line shape agreed with the shape predicted by summing up the individual hyperfine interactions which are the cause of the broadening. The behavior of an inhomogeneously broadened line observed under adiabatic fast passage conditions is discussed in an appendix. The electronic g-values were measured with respect to the free carriers in a degenerate n-type silicon sample. The g-value of the free carriers was found to be 1.99875+/-0.00010. The deviations of the donor g-values from the above value is several parts in 104 and increases monotonically with increasing ionization energy of the donors. Besides the shallow donors Sb, P, and As, several other centers were investigated, but in considerably less detail. They include the chemical impurities Bi, Li, Fe, centers associated with the surface of the sample and with the heat treatment of silicon. The influence of substitutional germanium atoms on the resonance line in phosphorus-doped silicon has also been investigated.
Article
Magnetic resonance studies are reported for a low-density gas of hydrogen atoms for 4.2<T<77 K. The resonance observed is between the two lowest hyperfine levels in a field of 6.5 kG, where this splitting has its minimum value (765 MHz). Information is obtained about translational diffusion through He4 and H2 buffer gases, frequency shifts and broadening due to wall collisions, and the effects of spin-exchange broadening.