scispace - formally typeset
Open AccessJournal ArticleDOI

Coherent dynamics of a Josephson charge qubit

Reads0
Chats0
TLDR
In this article, a single-Cooper-pair box (SCB) was fabricated by capacitively coupling a SCB to an electrometer based upon a singleelectron transistor (SET) configured for radio-frequency readout (rf-SET).
Abstract
We have fabricated a Josephson charge qubit by capacitively coupling a single-Cooper-pair box (SCB) to an electrometer based upon a single-electron transistor (SET) configured for radio-frequency readout (rf-SET). Charge quantization of 2e is observed and microwave spectroscopy is used to extract the Josephson and charging energies of the box. We perform coherent manipulation of the SCB by using very fast dc pulses and observe quantum oscillations in time of the charge that persist to ≃10 ns. The observed contrast of the oscillations is high and agrees with that expected from the finite EJ/EC ratio and finite rise time of the dc pulses. In addition, we are able to demonstrate nearly 100% initial charge state polarization. We also present a method to determine the relaxation time T1 when it is shorter than the measurement time

read more

Content maybe subject to copyright    Report

Coherent dynamics of a Josephson charge qubit
T. Duty,
*
D. Gunnarsson, K. Bladh, and P. Delsing
Microtechnology and Nanoscience, MC2, Chalmers University of Technology, S-412 96 Go
¨
teborg, Sweden
Received 10 May 2003; revised manuscript received 31 October 2003; published 5 April 2004
We have fabricated a Josephson charge qubit by capacitively coupling a single-Cooper-pair box SCB to an
electrometer based upon a single-electron transistor SET configured for radio-frequency readout rf-SET.
Charge quantization of 2e is observed and microwave spectroscopy is used to extract the Josephson and
charging energies of the box. We perform coherent manipulation of the SCB by using very fast dc pulses and
observe quantum oscillations in time of the charge that persist to 10 ns. The observed contrast of the
oscillations is high and agrees with that expected from the finite E
J
/E
C
ratio and finite rise time of the dc
pulses. In addition, we are able to demonstrate nearly 100% initial charge state polarization. We also present
a method to determine the relaxation time T
1
when it is shorter than the measurement time T
meas
.
DOI: 10.1103/PhysRevB.69.140503 PACS numbers: 85.25.Hv, 74.40.k, 85.35.Gv
Although a large number of physical systems have been
suggested as potential implementations of qubits, solid-state
systems are attractive in that they offer a realistic possibility
of scaling to a large number of interacting qubits. Recently
there has been considerable experimental progress using su-
perconducting microelectronic circuits to construct artificial
two-level systems. A variety of relative Josephson and Cou-
lomb energy scales have been used to construct qubits based
upon a single-Cooper-pair box
1,2
and flux qubits based upon
a three-junction loop.
3,4
Coherence times of the order of
0.5
s have been achieved for a single-Cooper-pair box
qubit.
2
Rabi oscillations between energy levels of a single
large tunnel junction have also been observed.
5,6
Despite the
encouraging results, one aspect that is not well understood
concerns the contrast of the oscillations, which in all previ-
ously reported experiments is smaller than expected.
The experimental systems reported so far can also be dis-
tinguished by the readout method and the manner of per-
forming single-qubit rotations. The first demonstration of co-
herent control of a single-Cooper-pair box
1
SCB employed
a weakly coupled probe junction to determine the charge on
the island. In the more recent experiment reported by Vion
et al.,
2
the SCB was incorporated into a loop containing a
large tunnel junction, for which the switching current de-
pends on the state of the SCB. Switching current measure-
ments of superconducting quantum-interference devices
SQUID’s have also been used for flux and phase-type
qubits.
3–6
Nakamura et al.
1
performed single-qubit rotations
by applying very fast dc pulses to a gate lead in order to
quickly move the SCB into and away from the charge degen-
eracy point. This technique produces qubit rotations with an
operation time that can be of the order of the natural oscil-
lation period. Other experiments utilize microwave pulses to
perform NMR-like rotations of the qubit.
2,46
The latter ap-
proach requires less stringent microwave engineering, since
rf rotations can be accomplished with pulses that are more
than an order of magnitude slower in rise time and duration
than the natural oscillation period. By using fast pulses, how-
ever, single-qubit rotations can be performed 20 times
faster than with rf rotations.
In this Rapid Communication, we report measurements
made on a SCB-type qubit with very fast dc pulses used to
effect the qubit rotations, as in Nakamura et al.
1
For our
qubit, however, the readout system consists of a single-
electron transistor SET capacitively coupled to the SCB
and configured for radio-frequency readout rf-SET.Byin-
corporating a SET into a tank circuit, rf-SET electrometers
can be made that are both fast
7
and sensitive,
8
and are well
suited for SCB-qubit readout.
9,10
One advantage of using rf-
SET readout is that it can easily be turned on and off, al-
though for the measurements reported here it is operated in a
continuous way.
10
More significantly, the rf-SET is funda-
mentally different from qubit readout based on switching
currents in that it involves a weak measurement of the charge
in contrast to a yes or a no answer. This could be a significant
advantage for readout of multiple-qubit systems. For ex-
ample, a switching current measurement on one qubit would
also directly affect the other qubits due to its strong interac-
tion. Furthermore, switching current measurements are inher-
ently stochastic in time—an important consideration if one is
interested in measuring time correlations.
Electron-beam lithography and double-angle shadow
evaporation of aluminum films onto an oxidized silicon sub-
strate were used to fabricate the combined SCB-SET system
see Fig. 1.
11
The SCB box consists of a low-capacitance
superconducting island connected to a superconducting res-
ervoir by two parallel junctions that define a low-inductance
SQUID loop. An additional gate lead placed near the island
is used to change the electrostatic potential of the island by a
gate voltage V
g
through a gate capacitance C
g
. By adjusting
the magnetic flux through the loop, the effective Joseph-
son energy is tunable as E
J
E
J
max
cos(
/
0
)
, where
0
is
the magnetic-flux quantum (h/2e). Coupling of the SCB to
the SET was accomplished by extending part of the SET
island to the proximity of the box island and resulted in a
weak dimensionless coupling, C
I
/C
of 0.54 % for
samples with slightly different geometries and total SCB ca-
pacitances C
. C
I
is the capacitance between the SCB and
SET islands.
The samples were placed at the mixing chamber of a di-
lution refrigerator with a base temperature of 20 mK. An
external superconducting magnet was used to produce a large
field parallel to the plane of the samples and a small super-
conducting coil close to the sample was used to produce a
RAPID COMMUNICATIONS
PHYSICAL REVIEW B 69, 140503R兲共2004
0163-1829/2004/6914/1405034/$22.50 ©2004 The American Physical Society69 140503-1

field perpendicular to the SQUID loop. This allowed us to
independently suppress the superconducting gap of the alu-
minum film, and change the effective Josephson energy of
the Cooper-pair box. All control lines were filtered by a com-
bination of low-pass and stainless steel and copper-powder
filters. To present sharp pulses to the SCB gate, a high-
frequency coaxial line was used having 10 dB attenuation at
4.2 K and 20 dB at 1.5 K. The tank circuit had a resonant
frequency of 380 MHz and the rf-SET electrometer was bi-
ased near a feature in the IV vs V
g
SET
landscape known as the
double Josephson quasiparticle peak,
12
such that the bias cur-
rent through the SET was typically 200300 pA. Under
these conditions, the electrometer sensitivity for the com-
bined SCB-SET system was found to be 50 10
6
e/
Hz
with a bandwidth of 10 MHz.
To make an artificial two-level system, the energy scales
of the SCB must be chosen so that k
B
T E
C
, E
J
, where
E
C
is the energy cost required to add a single electron to the
island and is set by total capacitance of the island E
C
e
2
/2C
.If is large enough compared to E
C
, then the
ground states will be even parity states that differ only by the
average number of Cooper pairs on the island. The effective
Hamiltonian of the box, including the Josephson coupling, is
given by
H 4E
C
n
n n
g
2
n
典具
n
E
J
2
n
n 1
典具
n
n
典具
n 1
, 1
where we define gate charge as n
g
C
g
V
g
/2e n
0
, and n
0
is
the offset charge due to stray charges near the box. Figure
2a shows the ground-state and first excited-state energy
bands calculated using the Hamiltonian in Eq. 1.Onac-
count of the 2e periodicity of the system, we limit the dis-
cussion to gate charge 0n
g
1.
In thermodynamic equilibrium, the actual quantity that
determines the island parity is the even-odd free energy dif-
ference
˜
(T), which differs from due to entropic
considerations.
13
Measurements were made for several dif-
ferent samples with Coulomb energies E
C
/k
B
(E
C
/h) rang-
ing from 0.43 to 1.65 K 9–34 GHz and E
J
/E
C
ratios of 0.4
to 2.25. Coloumb staircases were measured by slowly ramp-
ing the box gate charge at a rate of 40e/s), and measuring
the power reflected from the rf-SET tank circuit. Examples
of the measured box charge Q
box
(n
g
) is shown in Fig. 2b.
Our samples show 2e periodicity when E
c
/k
B
1 K, well
below the gap /k
B
2.2 K. For samples with larger E
c
, the
FIG. 1. a Scanning electron micrograph of a sample. The de-
vice consists of a Cooper-pair box and SET electrometer and was
fabricated from an aluminum lighter regions evaporated onto an
oxidized Si substrate darker regions. b Circuit diagram of box
and electrometer.
FIG. 2. a Ground-state and first excited-state energies vs n
g
for
E
J
/E
C
0.7 solid line and E
J
0 dotted line. b Coulomb stair-
cases, Q
box
versus n
g
for a sample with E
c
/h 9.25 0.10 GHz
and E
J
max
/h 20.24 0.10 GHz. In i a staircase taken from a
sample under microwave radiation with a frequency of 35 GHz is
shown. In ii, we show data where E
J
/h has been suppressed to
11.4 GHz by applying a small magnetic field through the SQUID
loop. The dotted lines in i and ii show the calculated Q
box
vs n
g
for the respective Josephson energies. iii Measured Q
box
vs n
g
solid line when a fast pulse train is applied to the SCB gate having
an amplitude 0.88e, width t 130 ps, and repetition time T
R
130 ns here E
J
/h 4 GHz). The dashed line is the measured
Q
box
without the pulse train. c Avoided level crossing shown by
the positions of the spectral peaks and dips, n
g
(n
g
dip
n
g
peak
)/2, for varying microwave frequency. d Josephson energy
E
J
vs applied flux /
0
through the SQUID loop determined from
spectroscopic data circles and coherent oscillation data crosses.
This sample had E
c
/h 17.45 0.13 GHz and E
J
max
/h 15.5
0.2 GHz.
RAPID COMMUNICATIONS
T. DUTY, D. GUNNARSSON, K. BLADH, AND P. DELSING PHYSICAL REVIEW B 69, 140503R兲共2004
140503-2

staircases acquire an extra step around n
g
0.5 due to poi-
soning by nonequilibrium quasiparticles.
The characteristic energies E
C
and E
J
max
of the SCB can
be determined directly using microwave spectroscopy.
10
When monochromatic microwaves are applied to the SCB
gate, resonant peaks and dips occur in the measured Q
box
vs
n
g
staircase when the microwave photon energy matches
with the energy-level splitting
01
see Fig. 2b兲兴. As the
microwave frequency is reduced, the positions of these peaks
exhibit an avoided level crossing. Within the two-level ap-
proximation, the positions of the single-photon resonances
depend on the applied frequency
as n
g
h
2
2
E
J
2
/8E
C
, where we used n
g
(n
g
dip
n
g
peak
)/2.
An example of such an avoided level crossing is shown in
Fig. 2c where we have used least-squares fits to this equa-
tion to estimate the Coulomb and Josephson energies.
Measurements of time-resolved, coherent oscillations of
the charge on the SCB are made by slowly ramping the SCB
gate charge n
g
as before, while applying fast nonadiabatic
rectangular dc-voltage pulses of amplitude a in the form of a
pulse train to the SCB gate. When n
g
is far away from the
charge degeneracy and a such that n
g
a places the system
at the charge degeneracy, the leading and trailing edges of
each pulse act as successive Hadamard transformations. In
between, the system coherently oscillates between charge
eigenstates for a time t. The trailing edge of the pulse
returns the gate charge to n
g
with the system in a superposi-
tion of ground and excited states corresponding to the phase
of the oscillation acquired during t. After the pulse, the
excited-state component decays with a relaxation time T
1
.
The pulses are separated by a repetition time T
R
, typically
greater than T
1
.
Since the staircase is acquired by ramping n
g
on a time
scale much slower than T
R
, one measures an enhanced
charge Q
box
Q
box
on
Q
box
off
that is time averaged over the
pulse train and proportional to the probability of finding the
system in the excited state. Peaks in Q
box
vs n
g
will be
located at gate charges n
g
peak
such that approximately a half-
integer number of oscillations occur at gate charge n
g
peak
a during the time t. Figure 2b兲共bottom curve shows a
measured staircase using such a pulse train. When one mea-
sures such staircases for varying t, then time-domain oscil-
lations are evident in the t cross sections of Q
box
vs n
g
.
This is shown in Fig. 4a, where we plot the time evolution
near the charge degeneracy point. As T
R
is increased, the
pulse-induced charge contributes less to the time average. A
simple calculation which ignores coherence effects remain-
ing after a time T
R
leads to the dependence
Q
box
T
R
2n
0
T
1
T
R
1 e
T
R
/T
1
1 e
T
R
/T
1
, 2
where n
0
is the initial peak amplitude. One can estimate T
1
at the readout gate voltage by measuring the dependence of
the peak amplitude on T
R
and fitting the data to this formula.
Using such a pulse-train allows a measurement of the
excited-state probability even when the relaxation time T
1
is
shorter than the measurement time. Population averaging and
mixing limit the maximum observed amplitude of the Q
box
oscillations to 1e rather than 2e, and this occurs only for an
ideal rectangular pulse, short repetition times, and vanishing
E
J
/E
C
ratio. We have numerically integrated the Schro
¨
-
dinger equation using pulses with a finite rise time to study
the various factors affecting both the contrast of the oscilla-
tions and the degree of charge state polarization. The results
are best understood when plotted on the Bloch sphere as
shown in Fig. 3. The dash-dotted lines show the evolution in
time of the state vector starting from an initial condition
upward pointing arrows and evolving through a maximum
polarization of the charge state downward arrows. The ini-
tial state differs from pure S
z
due to a finite E
J
/E
C
. Figure
3a shows evolution for a pulse that takes the system to the
charge degeneracy. One finds that a finite pulse rise time
mimics a higher E
J
/E
C
ratio and reduces the oscillation con-
trast. Nonetheless, Fig. 3b shows that by pulsing past the
charge degeneracy, one can achieve nearly 100% polariza-
tion.
Measurements of the T
R
dependence of Q
box
are shown
in Fig. 4b, where we have used an appropriate pulse am-
plitude and duration to achieve nearly 100% initial polariza-
tion. We find T
1
108 20 ns at the readout point n
g
0.25
using a least-squares fit to Eq. 2. The data of Fig. 4a show
an oscillation contrast of 0.55e. Taking into account the
experimentally determined E
J
/E
C
and T
1
/T
R
, along with
the measured rise time
t35 ps of our pulse generator An-
ritsu MP1763C, we find good agreement between the ob-
served and expected initial oscillation contrast. Numerical
solutions of Schro
¨
dinger equation show that the reduction of
contrast due to the pulse rise time depends strongly on E
J
.
This was checked experimentally by varying E
J
and found to
be in agreement with the calculations. A fit of the data of Fig.
4a to an exponentially decaying sinusoid gives an oscilla-
tion frequency
0
3.6 GHz and decoherence time of 2.9 ns.
Using fast dc pulse trains we were able to observe oscil-
lations in time out to 10 ns. The dependence of decoherence
time on gate charge at the top of the pulse is shown in Fig.
4c and has a sharp maximum at the charge degeneracy.
This tells us that low-frequency charge noise is the dominant
source of dephasing in our qubit. As discussed and demon-
strated by Vion et al.,
2
dephasing due to charge fluctuations
is minimum at the charge degeneracy. The coherence time
FIG. 3. a Bloch sphere showing evolution about the charge
degeneracy for E
J
/h 6 GHz, E
C
/h 9 GHz, and a pulse rise
time of 35 ps. b Evolution for a pulse that takes the system past
the charge degeneracy and achieves 100% charge state polarization
downward pointing arrow.
RAPID COMMUNICATIONS
COHERENT DYNAMICS OF A JOSEPHSON CHARGE QUBIT PHYSICAL REVIEW B 69, 140503R兲共2004
140503-3

found by Vion et al. is much larger than that found in our
measurements although their initial oscillation amplitude is
much smaller. If we extrapolate the gate charge dependence
of T
1
not shown to the charge degeneracy, we expect T
1
10 ns. This implies that the decoherence time we measure
at the charge degeneracy is dominated by the short relaxation
time rather than pure dephasing.
Our measured T
1
was also found to be independent of
SET bias current in the subgap regime. Previously, a consid-
erably larger T
1
was measured in a similar sample using a
spectroscopic technique,
10
and found to agree with the theo-
retical T
1
1
s expected from quantum fluctuations of a
50 environment. The reduced T
1
observed here could be
attributed to several factors. One is coupling to the relatively
unfiltered high-frequency coaxial line, in addition to un-
wanted coupling to other metal traces and the microwave
environment of our sample. In practice it is difficult to esti-
mate the actual real part of the impedance on these leads at
such high transition frequencies. Another possibility con-
cerns quasiparticles. Our samples were fabricated without
quasiparticle traps and despite the observed static 2e period-
icity, there could be considerable nonequilibrium quasiparti-
cle dynamics occurring on these time scales. Finally, it may
be that the use of fast dc pulses leaves the bath of charge
fluctuators in a configuration having additional channels for
relaxtion of the qubit. Clearly what is needed is T
1
measure-
ments using a variety of techniques—rf rotations, spectros-
copy, as well as fast dc pulse—on the same sample.
In conclusion, we have fabricated and measured a solid-
state qubit based upon a SCB combined with a rf-SET read-
out system. Due to a relatively short T
1
, continuous mea-
surement of the SCB was employed. Fast dc pulses were
used to coherently manipulate the qubit and time-coherent
oscillations of the charge were observed. The initial contrast
of the oscillations is relatively large and quantitatively un-
derstood as due to finite E
J
/E
C
combined with the finite
pulse rise time. Furthermore, nearly 100% charge state po-
larization can be achieved. The oscillations show a maxi-
mum decoherence time at the charge degeneracy which indi-
cates that charge fluctuations dominate the dephasing rate.
We would like to acknowledge fruitful discussions with
R. Schoelkopf, K. Lehnert, A. Wallraff, G. Johansson,
A. Ka
¨
ck, G. Wendin, and Y. Nakamura. The samples were
made at the MC2 clean room. The work was supported by
the Swedish SSF and VR, by the Wallenberg and Go
¨
ran
Gustafsson foundations, and by the EU under the IST-
SQUBIT programme.
*
Electronic address: tim@mc2.chalmers.se
1
Y. Nakamura, Y.A. Pashkin, and J.S. Tsai, Nature London 398,
786 1999.
2
D. Vion, A. Aassime, A. Cottet, P. Joyez, H. Pothier, C. Urbina,
D. Esteve, and M.H. Devoret, Science 296, 886 2002.
3
C.H. van der Wal et al., Science 290, 773 2000.
4
I. Chiorescu, Y. Nakamura, C.J.P.M. Harmans, and J.E. Mooij,
Science 299, 1869 2003.
5
Y. Yu, S. Han, X. Chu, S.-I. Chu, and Y. Wang, Science 296, 889
2002.
6
J.M. Martinis, S. Nam, J. Aumentado, and C. Urbina, Phys. Rev.
Lett. 89, 117901 2002.
7
R.J. Schoelkopf, P. Walgren, A.A. Kozhevnikov, P. Delsing, and
D.E. Prober, Science 280, 1238 1998.
8
A. Aassime, D. Gunnarsson, K. Bladh, and P. Delsing, Appl.
Phys. Lett. 79, 4031 2001.
9
A. Aassime, G. Johansson, G. Wendin, R.J. Schoelkopf, and P.
Delsing, Phys. Rev. Lett. 86, 3376 2001.
10
K.W. Lehnert, K. Bladh, L.F. Spietz, D. Gunnarsson, D.I.
Schuster, P. Delsing, and R.J. Schoelkopf, Phys. Rev. Lett. 90,
027002 2003.
11
K. Bladh et al., Phys. Scr. T102, 167 2002.
12
A.A. Clerk, S.M. Girvin, A.K. Nguyen, and A.D. Stone, Phys.
Rev. Lett. 89, 176804 2002.
13
P. Lafarge, P. Joyez, D. Esteve, C. Urbina, and M.H. Devoret,
Nature London 365, 6445 1993.
FIG. 4. a兲⌬Q
box
vs t. The solid line is a least-squares fit to
an exponentially decaying sinusoid and gives a decay time T
2
2.9 ns and frequency
0
3.6 GHz. b Dependence of the peak
height Q
box
on repetition time T
R
for 100% initial polarization.
The solid line is a least-squares fit to Eq. 2 giving T
1
108
20 ns. c Decoherence time T
2
vs gate charge n
g
.
RAPID COMMUNICATIONS
T. DUTY, D. GUNNARSSON, K. BLADH, AND P. DELSING PHYSICAL REVIEW B 69, 140503R兲共2004
140503-4
Citations
More filters
Journal ArticleDOI

Superconducting quantum bits

TL;DR: Superconducting quantum bits (qubits) form the key component of these circuits and their quantum state is manipulated by using electromagnetic pulses to control the magnetic flux, the electric charge or the phase difference across a Josephson junction.
Journal ArticleDOI

A quantum engineer's guide to superconducting qubits

TL;DR: In this paper, the authors provide an introductory guide to the central concepts and challenges in the rapidly accelerating field of superconducting quantum circuits, including qubit design, noise properties, qubit control and readout techniques.
Journal ArticleDOI

Superconducting Circuits and Quantum Information

J. Q. You, +1 more
- 01 Nov 2005 - 
TL;DR: Superconducting circuits can behave like atoms making transitions between two levels as mentioned in this paper, and such circuits can test quantum mechanics at macroscopic scales and be used to conduct atomic-physics experiments on a silicon chip.
Journal ArticleDOI

A Quantum Engineer's Guide to Superconducting Qubits

TL;DR: In this article, the authors provide an introductory guide to the central concepts and challenges in the rapidly accelerating field of superconducting quantum circuits, including qubit design, noise properties, qubit control, and readout techniques.
Journal ArticleDOI

Decoherence in a superconducting quantum bit circuit

TL;DR: This work presents experiments, inspired from NMR, that characterize decoherence in a particular superconducting quantum bit circuit, the quantronium, and introduces a general framework for the analysis of decoherent, based on the spectral density of the noise sources coupled to the qubit.
Related Papers (5)
Frequently Asked Questions (1)
Q1. What have the authors contributed in "Coherent dynamics of a josephson charge qubit" ?

The authors perform coherent manipulation of the SCB by using very fast dc pulses and observe quantum oscillations in time of the charge that persist to. 10 ns. In addition, the authors are able to demonstrate nearly 100 % initial charge state polarization. The authors also present a method to determine the relaxation time T1 when it is shorter than the measurement time Tmeas.