ArticlePDF Available

Hybrid quantum logic and a test of Bell's inequality using two different atomic species

Authors:

Abstract

Entanglement is one of the most fundamental properties of quantum mechanics, and is the key resource for quantum information processing. Bipartite entangled states of identical particles have been generated and studied in several experiments, and post-selected entangled states involving pairs of photons, or single photons and single atoms, have also been produced. Here, we deterministically generate a "hybrid" entangled state of two different species of trapped-ion qubit, perform full tomography of the state produced, and make the first test of Bell's inequality with non-identical atoms. We use a laser-driven two-qubit quantum logic gate, whose mechanism is insensitive to the qubits' energy splittings, to produce a maximally-entangled state of one Ca40 qubit and one Ca43 qubit, held in the same ion trap, with 99.8(5)% fidelity. We make a test of Bell's inequality for this novel entangled state, and find that it is violated by 15 sigma. Mixed-species quantum logic is an essential technique for the construction of a quantum computer based on trapped ions, as it allows protection of memory qubits while other qubits undergo logic operations, or are used as photonic interfaces to other processing units.
Hybrid quantum logic and a test of Bell’s inequality
using two different atomic isotopes
C. J. Ballance, V. M. Sch¨afer, J. P. Home, D. J. Szwer, S. C. Webster, D. T. C. Allcock,
N. M. Linke, T. P. Harty, D. P. L. Aude Craik, D. N. Stacey, A. M. Steane and D. M. Lucas
Department of Physics, University of Oxford, Clarendon Laboratory, Parks Road, Oxford OX1 3PU, U.K.
(Dated: 27 November 2015, v40.43)
Entanglement is one of the most fundamental
properties of quantum mechanics [13], and is the
key resource for quantum information process-
ing [4, 5]. Bipartite entangled states of identical
particles have been generated and studied in sev-
eral experiments, and post-selected or heralded
entangled states involving pairs of photons, sin-
gle photons and single atoms, or different nuclei
in the solid state, have also been produced [6–
12]. Here, we use a deterministic quantum logic
gate to generate a “hybrid” entangled state of
two trapped-ion qubits held in different isotopes
of calcium, perform full tomography of the state
produced, and make a test of Bell’s inequality
with non-identical atoms. We use a laser-driven
two-qubit gate [13], whose mechanism is insensi-
tive to the qubits’ energy splittings, to produce
a maximally-entangled state of one 40Ca+qubit
and one 43Ca+qubit, held 3.5 µm apart in the
same ion trap, with 99.8(6)% fidelity. We test the
Clauser-Horne-Shimony-Holt [14] (CHSH) ver-
sion of Bell’s inequality for this novel entangled
state and find that it is violated by 15 stan-
dard deviations; in this test, we close the detec-
tion loophole [8] but not the locality loophole [7].
Mixed-species quantum logic is a powerful tech-
nique for the construction of a quantum computer
based on trapped ions, as it allows protection of
memory qubits while other qubits undergo logic
operations, or are used as photonic interfaces to
other processing units [15, 16]. The entangling
gate mechanism used here can also be applied to
qubits stored in different atomic elements; this
would allow both memory and logic gate errors
due to photon scattering to be reduced below the
levels required for fault-tolerant quantum error
correction, which is an essential pre-requisite for
general-purpose quantum computing.
For Schr¨odinger, entanglement was the characteris-
tic trait of quantum mechanics” [1] and it has been at
the heart of debates about the foundations of quantum
mechanics since the framing of the Einstein-Podolsky-
Rosen paradox [2]. The theoretical work of Bell [3], and
of Clauser et al. [14], established an experimental test
which could be used to rule out local hidden-variable
theories on the basis of correlations between measured
properties of entangled particles, and numerous experi-
ments, starting with that of Freedman and Clauser, have
confirmed the predictions of quantum mechanics [6–10].
Tests of Bell’s inequality with trapped ions were the first
to close the so-called “detection loophole”; hitherto these
trapped-ion tests have exclusively been carried out with
identical atoms [8, 17, 18]. The entanglement explored
in tests of Bell’s inequality is typically an entanglement
between distinguishable particles, in the strict quantum
mechanical sense, but when the particles are identical in
their internal structure and state, they are distinguish-
able only through their spatial localization. By employ-
ing different isotopes, our experiments involve entities
that are also distinguishable by many internal properties,
such as baryon number, mass, spin, resonant frequencies,
and so on.
Apart from its intrinsic interest, entanglement is a
central resource for quantum information applications,
such as quantum cryptography [5] and quantum comput-
ing [4]. Trapped atomic ions are one of the most promis-
ing technologies for the implementation of quantum com-
putation; several demonstrations of simple multi-qubit
algorithms have been made [19] and the elementary set
of quantum logic operations has recently been demon-
strated with the precision required for the implemen-
tation of fault-tolerant techniques [20, 21]. Scaling up
trapped-ion systems to the large numbers of qubits re-
quired for useful quantum information processing and
quantum simulation will almost certainly require the use
of more than one species of ion, both for the purpose
of sympathetic laser-cooling (which allows independent
control of the external and internal atomic degrees of
freedom) [15, 22, 23] and for providing robust mem-
ory qubits. The best memory qubits reside in hyperfine
ground states [20, 24], which have essentially infinite life-
times against spontaneous decay, but are vulnerable to
the scattering of a single photon of resonant laser light.
In a complex, multi-zone, ion trap processor it will be dif-
ficult to shield the memory qubits sufficiently well from
resonant laser beams, hence it will be useful to employ
different species of ion, for example as memory and logic
qubits, and a high-fidelity entangling gate operation be-
tween the two species will be invaluable. A significant
initial step was the demonstration of coherent state trans-
fer between different species in the context of precision
metrology [25, 26]. The relative merits of using different
isotopes versus different elements are discussed below.
In the present work, we entangle qubits stored in
two different isotopes of calcium. The 40Ca+qubit
is stored in the Zeeman-split ground level, (|↓i,|↑i) =
(4S1/2
1/2,4S+1/2
1/2), and the 43Ca+qubit is stored in the hy-
perfine ground states (|⇓i,|⇑i) = (4S4,+4
1/2,4S3,+3
1/2), see fig-
arXiv:1505.04014v2 [quant-ph] 27 Nov 2015
2
FIG. 1: Calcium ion energy levels and experimental geome-
try. (a) Qubit states and Raman transitions in 43Ca+and
40Ca+. The two Raman beams have a mean detuning of
= 1.04 THz from the 4S1/24P1/2(397 nm) transition,
and a difference frequency of δ=fz+δg2 MHz. (b) Ra-
man gate beam geometry. The two perpendicular beams are
aligned to set the lattice k-vector parallel to the trap axis ˆz.
The beams have waist radii w= 27 µm, a power of 5 mW
each, and orthogonal linear polarizations as indicated. A
third, π-polarized, Raman beam (not shown) co-propagates
with the σbeam and is used for sub-Doppler sideband cool-
ing and single-qubit operations on 40Ca+. The quantization
axis is set by a magnetic field B0.2 mT. The diagram is
not to scale: the ions are separated by 3.5 µm, which is 12 1
2
periods of the standing wave, and around 20,000 times the
atomic radius of calcium.
ure 1. The qubit energy splittings differ by some three
orders of magnitude (fl5.4 MHz, fm3.2 GHz), but
they may nevertheless be efficiently coupled via the two-
qubit gate mechanism of Leibfried et al. [13], in which
the “travelling standing wave” from a pair of far-detuned
laser beams exerts a qubit-state-dependent force on the
ions whose magnitude Fis largely independent of the
qubit frequency. The force originates from a spatially-
varying light shift, oscillates at the difference frequency
δbetween the two beams and, when δ=fz+δgis set
close to the resonant frequency fzof a normal mode of
motion of the two-ion crystal, a two-qubit phase gate may
be implemented by applying the force for a time (1g).
An advantage of this type of gate is that the phase of the
optical field does not need to be referenced to either of
the qubit phases (see Methods); this makes scaling the
system easier because the relative optical phase does not
need to be controlled between different trap zones, or
during time delays between gates.
An important difference in the gate mechanism com-
pared with the case of identical ions is that the forces on
corresponding qubit states differ (F6=Fand F6=F)
so that, in general, the four possible qubit states (↑⇑,
↑⇓,↓⇑,↓⇓) each acquire different phases. We choose to
implement the gate operation in two halves, each of du-
ration tg/2 = 1g, separated by spin-flip operations (π-
200 100 0 100 200
0
0.2
0.4
0.6
0.8
1
Probability
Analysis phase (degrees)
P()
P()
P()+P()
Probability
Analysis phase (degrees)
P()
P()
P()+P()
a
b
Analysis phase (degrees)
Probability
tg/2
FIG. 2: Entangling gate sequence and results. (a) Gate se-
quence, showing the operations applied to the 40Ca+(upper
line) and 43Ca+(lower line) qubits. The final state analy-
sis (tomography) π/2-pulses shown in green are optional; by
scanning their phase φwe can diagnose the state produced
by the gate. (b) Qubit populations and parity signal after
correcting for readout errors (see Methods). The individual
qubit populations are consistent with 1
2, as expected for the
Bell state (|↓⇓i+|↑⇑i)/2. The parity signal P(↑⇓)+P(↓⇑),
i.e. the probability of the two qubits being in opposite states,
should oscillate between 0 and 1 as sin(2φ) for a perfect Bell
state. From the contrast of the parity signal and a measure-
ment of the populations without the analysis pulses, we infer
a Bell state fidelity of 99.8(6)%. The error bars show 1σsta-
tistical errors.
pulses) on the qubits (figure 2a). This symmetrizes the
gate operation such that the relative phases acquired by
the four states are (0,Φ,Φ,0). By setting the laser power
(i.e., effective Rabi frequency) and gate detuning δgap-
propriately, such that Φ = π/2, and enclosing the gate
operation in a Ramsey interferometer (two pairs of π/2-
pulses), we can generate the maximally-entangled Bell
state (|↓⇓i +|↑⇑i)/2 from the initial state |↓⇓i. The
π-pulses also protect the qubits against dephasing due to
slow (tg) variations in magnetic fields.
In our experiment, we implement the gate using the
in-phase axial motional mode (at fz= 2.00 MHz) of a
linear Paul trap [27], with the ion separation (3.5 µm)
equal to a half-integer number of standing wavelengths,
thus exciting the motion maximally for the |↑⇓i and |↓⇑i
states. The Lamb-Dicke parameters for the two different
isotopes are η40 = 0.121 and η43 = 0.126. After initial
Doppler cooling, both axial modes are cooled close to
their ground states (mean occupation number ¯n < 0.1)
by Raman sideband cooling applied to the 40Ca+ion,
which sympathetically cools the 43Ca+ion [28]. Both
qubits are initialized by optical pumping, after which we
3
0
0.1
0.2
0.3
0.4
0.5
Re(ρ)
0
0.1
0.2
0.3
0.4
0.5
Im(ρ)
FIG. 3: Density matrix of the mixed-isotope Bell state. The
plots show the real (left) and imaginary (right) parts of the
density matrix, after correcting for qubit readout errors (see
Methods). The measurements were made by rotating each
qubit independently to perform full quantum state tomog-
raphy. We used a maximum likelihood method to find the
density matrix that best represents the experimental data.
This gives a separate estimate of the gate fidelity, 99(1)%.
apply the gate sequence shown in figure 2a, using a gate
duration tg= 27.4µs. Single-qubit π/2- and π-pulses,
for the spin-echo and tomography operations, are applied
using co-propagating Raman beams (for 40 Ca+) and mi-
crowaves (for 43Ca+). The ordering of the ion pair in the
trap was kept constant over the time taken to acquire
the full data set, to guard against systematic effects as-
sociated with ion position (see Methods). We implement
individual single-shot qubit readout by state-selectively
shelving both ions to the 3D5/2level simultaneously, then
detecting the ions’ fluorescence sequentially in two pho-
tomultiplier counting periods (see Methods).
From the contrast of the parity fringes shown in fig-
ure 2, and a measurement of the qubit populations be-
fore the analysis pulses [13], we estimate the fidelity of
the Bell state produced by the gate to be F= 99.8(6)%,
where the error is dominated by statistical uncertainty.
Known contributions to the gate error are significantly
smaller [27] than the statistical uncertainty; for example
the photon scattering error at the = 1.04 THz Ra-
man detuning used is estimated to be 0.1%. Since the
two qubits may be rotated independently by addressing
them in frequency space, we can also perform full tomog-
raphy of the entangled state and extract the density ma-
trix (figure 3); the density matrix is consistent with that
for the desired Bell state, to within the systematic errors
from the imperfect tomography pulses, and gives a sepa-
rate estimate of the fidelity F= 99(1)%. In both cases,
Frepresents the fidelity of the entangling gate operation;
it excludes errors due to state preparation and readout,
which we characterize in independent experiments (see
Methods).
To perform a test of the CHSH version of Bell’s in-
equality, we follow the gate sequence with further inde-
pendent single-qubit rotations and measurements. The
single qubit rotations have constant phase φbut varying
rotation angle θ. From these measurements we deter-
mine the two-particle correlation functions with results
θa(40Ca+)π/4 3π/4π/4 3π/4
θb(43Ca+)π/2π/2 0 0
E(θa, θb) 0.565(7) 0.530(7) 0.560(7) 0.573(8)
TABLE I: Bell/CHSH inequality test results, using the mixed-
isotope entangled state. The qubits aand bare independently
rotated through angles (θa, θ0
a) = (π/4,3π/4) and (θb, θ 0
b) =
(π/2,0), and for each combination of angles the correlation
function E(θa, θb) is measured, with results shown. (Eis
defined as in ref.[17].) The CHSH parameter is given by S=
|E(θa, θb)+E(θ0
a, θb)|+|E(θa, θ0
b)E(θ0
a, θ0
b)|= 2.228(15) >2,
thus violating Bell’s inequality for this system of non-identical
atoms. The state detection errors are sufficiently small (6%,
see Methods), that it is not necessary to make a fair-sampling
assumption. For each angle setting 4,000 measurements were
made.
shown in table I. As is well known, the maximal CHSH
parameter Sallowed by local hidden-variable theories is
2, whereas quantum mechanics allows S22. In order
to avoid having to make a fair-sampling assumption, we
do not correct for qubit readout errors in these experi-
ments. The finite detection error then limits the CHSH
parameter to a detectable maximum Smax = 2.236(7) for
a perfect Bell state; our results give S= 2.228(15), con-
sistent with Smax to within the stated uncertainties, and
violating the CHSH inequality by 15σ.
The mixed-species quantum logic gate that we have
demonstrated has allowed us to create a novel entangled
state, leading to the first test of a Bell inequality violation
between isolated non-identical atoms. As an application,
the two isotopes used here could be employed for scal-
able quantum computing architectures based on trapped
ions; hyperfine qubits in 43Ca+at present constitute the
best single-qubit memories (T
21 min) [20], whereas the
simpler atomic structure of 40Ca+is well suited for use as
a “photonic interconnect” qubit [16]. There are technical
advantages to using ions of similar mass for sympathetic
cooling and ion transport in multi-zone traps. However,
while the relatively small isotope shifts (1 GHz) allow
the convenient use of the same laser systems for manipu-
lation of both species, they may provide insufficient pro-
tection of qubits from stray resonant light unless tightly
focussed beams are used [18, 28]. Therefore in the long
term it may be necessary to use different atomic ele-
ments [22]. The gate mechanism employed here is inde-
pendent of the qubit frequency and thus can also be used
to couple qubits stored in different elements, provided
that the Raman laser fields exert sufficient force on both
qubits. We note that Ca+and Sr+ions are an attractive
choice in this respect: the 4S1/24P1/2transition in
Ca+is separated from the 4S1/24P3/2transition in
Sr+by 20 THz. A Raman laser detuning = 8 THz
(comparable to that used in our recent 43Ca+43Ca+
two-qubit gate experiments [21]) would enable the imple-
mentation of a mixed-species logic gate with a photon-
scattering error of 104, significantly below the error
threshold for fault-tolerant operations [29].
4
Similar experiments using trapped-ion qubits stored in
two different elements (9Be+and 25Mg+) have recently
been carried out in the NIST Ion Storage Group [30].
Subsequent to the submission of our manuscript, a
CHSH-Bell test which closes both detection and locality
loopholes, using heralded entanglement of remote elec-
tron spins, has been reported [31].
METHODS
Ion crystal order. The 40 Ca+43Ca+ion crystal
ordering is kept constant during the experiments to
control systematic errors. The principal error which
would arise if the ion order were not controlled is due to
an (undesired) axial magnetic field gradient that causes
the magnetic field between the two ions to differ by
0.18 µT. This means that the qubit frequencies for the
two possible ion orders differ by 5 kHz, which would
lead to errors in single-qubit rotations. We measure the
frequency of each qubit using slow (typically 100 µs)
carrier π-pulses, interleaved with the main experimental
pulse sequence, which allows us to detect and to correct
for both common-mode qubit frequency changes (due
to drift in the global magnetic field B) and differential
changes (due to incorrect ion crystal ordering). If the
ion order is wrong, we randomly reorder the crystal
until the order is correct with a short period of Doppler
heating to melt the crystal, followed by a short period
of Doppler cooling.
Single-qubit phases and light shifts. Despite the
qubits having very different frequencies, no special phase
control is needed to implement the entangling gate. The
43Ca+qubit phase is tracked by the microwave local os-
cillator, and the 40 Ca+qubit phase is tracked by the
difference phase of the co-propagating Raman beams,
in turn referenced to a radio-frequency local oscillator.
The phases of the Raman beams that implement the en-
tangling gate have no relationship to either of the qubit
phases. However, the travelling standing wave result-
ing from the interference of the Raman gate beams also
generates an isotope-dependent differential light shift on
each qubit with an amplitude that oscillates at the Ra-
man difference frequency δ. Over the course of the gate
operation this light shift adds phase shifts to the qubits
that depend on the (uncontrolled) optical phase differ-
ence of the Raman beams. These uncontrolled phase
shifts reduce the fidelity of the gate operation. We greatly
reduce this light shift error by shaping the turn-on and
turn-off of the Raman laser intensities with a character-
istic time of 1 µs; we estimate that without this pulse-
shaping the light shift would lead to an average gate error
of up to 5% (see ref.[27]).
We adjust the polarisation of each Raman beam
individually to null the differential light shift from each
single beam on the 40 Ca+qubit. (The interference of the
two gate beams nevertheless gives rise to a polarization
modulation which provides the state-dependent force.)
Due to the difference in atomic structure there is a
residual light shift on the 43Ca+qubit of 0.2% of
the light shift for a purely circularly-polarised beam of
the same intensity and frequency. This small light shift
does not cause any significant issues in the experiments
reported here; if necessary it could be suppressed further
by increasing the Raman detuning at the expense of
requiring more Raman beam power.
State preparation and measurement errors. To
perform individual single-shot qubit readout, we selec-
tively shelve one qubit state of each ion to the 3D5/2
level, then apply the Doppler cooling lasers sequentially
in time first for one isotope, then for the other. If an
ion was not shelved it fluoresces, and this is detected
with a photomultiplier. We simultaneously shelve the
two isotopes using a weak 393nm beam resonant with
the 43Ca+4S4,+4
1/24P5,+5
3/2transition, with a 1.94 GHz
EOM sideband which drives the 40Ca+4S+1/2
1/24P+3/2
3/2
transition. An intense 850nm beam resonant with the
40Ca+3D3/24P3/2transition makes the shelving for
this isotope state-selective, through an EIT process in-
volving the two transitions [32]. The 43Ca+shelving is
state-selective due to the 3.2 GHz splitting between the
two qubit states [33]. Both these shelving processes have
a maximum theoretical efficiency of 90% due to leak-
age to 3D3/2(which for 43Ca+could be eliminated using a
further 850 nm beam if required [33]), leading to readout
errors of ¯5% when averaged over both qubit states.
From independent experiments (similar to those we de-
scribe in ref. [20]), we estimate the state-preparation er-
ror to be 0.1%, which is negligible compared with the
readout error.
We measure the readout errors for each qubit state of
each isotope, by preparing and measuring each state typ-
ically 10,000 times. Since the qubits are measured indi-
vidually, it is then straightforward to calculate the linear
mapping that corrects for the readout errors, provided
that they remain constant. The readout errors relevant
to the entangling gate experiment (figure 2) were mea-
sured to be ¯40 = 7.7(2)% for 40 Ca+and ¯43 = 4.4(2)%
for 43Ca+(averaged over both qubit states). Measure-
ments of the readout errors were interleaved with the gate
experimental runs, to check for systematic drifts, and
were made using the mixed-isotope crystal, to avoid sys-
tematic effects associated with ion position. We estimate
the systematic uncertainty in determining the readout er-
rors to be 0.1%, less than the statistical error in these
measurements. If we did not correct for readout errors,
the apparent infidelity in the Bell state would increase
by 3
240 + ¯43)18%. For the CHSH test, we do
not correct for readout errors, but we nevertheless mea-
sure them in order to calculate the maximum attainable
CHSH parameter (Smax).
5
Acknowledgements This work was supported by
the U.K. EPSRC “Networked Quantum Information
Technology” Hub and the U.S. Army Research Office
(contract W911NF-14-1-0217). D.M.L. would like to
thank Aida and Esther Andrade Castillo for their
hospitality while revising the manuscript.
Author contributions All authors contributed to the
development of the apparatus and/or the design of the
experiments. J.P.H. and D.M.L. conceived the experi-
ments and took preliminary data. C.J.B. and V.M.S.
designed and performed the experiments described
here, analysed data and produced the figures. C.J.B.
and D.M.L. wrote the manuscript, which all authors
discussed.
Author information Correspondence should be ad-
dressed to C.J.B. (c.ballance@physics.ox.ac.uk) or
D.M.L. (d.lucas@physics.ox.ac.uk).
[1] E. Schr¨odinger. Mathematical Proceedings of the Cam-
bridge Philosophical Society 31, 555–563 (1935).
[2] A. Einstein, B. Podolsky, and N. Rosen. Physical Review
47, 777–780 (1935).
[3] J. S. Bell. Physics 1, 195–200 (1964).
[4] D. Deutsch. Proceedings of the Royal Society A 400, 97–
117 (1985).
[5] A. K. Ekert. Physical Review Letters 67, 661–663 (1991).
[6] S. J. Freedman and J. F. Clauser. Physical Review Let-
ters 28, 938–941 (1972).
[7] A. Aspect, P. Grangier, and G. Roger. Physical Review
Letters 49, 91–94 (1982).
[8] M. A. Rowe, D. Kielpinski, V. Meyer, C. A. Sackett,
W. M. Itano, C. Monroe, and D. J. Wineland. Nature
409, 791–794 (2001).
[9] D. L. Moehring, M. J. Madsen, B. B. Blinov, and C. Mon-
roe. Physical Review Letters 93, 090410 (2004).
[10] M. Giustina, A. Mech, S. Ramelow, B. Wittmann,
J. Kofler, J. Beyer, A. Lita, B. Calkins, T. Gerrits, S. W.
Nam, R. Ursin, and A. Zeilinger. Nature 497, 227–230
(2013).
[11] B. G. Christensen, K. T. McCusker, J. Altepeter,
B. Calkins, T. Gerrits, A. E. Lita, A. Miller, L. K. Shalm,
Y. Zhang, S. W. Nam, N. Brunner, C. C. W. Lim,
N. Gisin, and P. G. Kwiat. Physical Review Letters 111,
130406 (2013).
[12] W. Pfaff, T. H. Taminiau, L. Robledo, H. Bernien,
M. Markham, D. J. Twitchen, and R. Hanson. Nature
Physics 9, 29–33 (2013).
[13] D. Leibfried, B. DeMarco, V. Meyer, D. M. Lucas, M. D.
Barrett, J. W. Britton, W. M. Itano, B. Jelenkovi´c, C. E.
Langer, T. Rosenband, and D. J. Wineland. Nature 422,
412–415 (2003).
[14] J. F. Clauser, M. A. Horne, A. Shimony, and R. A. Holt.
Physical Review Letters 23, 880–884 (1969).
[15] D. J. Wineland, C. Monroe, W. M. Itano, D. Leibfried,
B. E. King, and D. M. Meekhof. Journal Of Research
Of The National Institute Of Standards And Technology
103, 259–328 (1998).
[16] C. Monroe and J. Kim. Science 339, 1164–1169 (2013).
[17] D. N. Matsukevich, P. Maunz, D. L. Moehring, S. Olm-
schenk, and C. Monroe. Physical Review Letters 100,
150404 (2008).
[18] B. P. Lanyon, M. Zwerger, P. Jurcevic, C. Hempel,
W. ur, H. J. Briegel, R. Blatt, and C. F. Roos. Physical
Review Letters 112, 100403 (2014).
[19] R. Blatt and D. J. Wineland. Nature 453, 1008–1015
(2008).
[20] T. P. Harty, D. T. C. Allcock, C. J. Ballance, L. Guidoni,
H. A. Janacek, N. M. Linke, D. N. Stacey, and D. M.
Lucas. Physical Review Letters 113, 220501 (2014).
[21] C. J. Ballance, T. P. Harty, N. M. Linke, and D. M.
Lucas. arXiv:1406.5473.
[22] M. D. Barrett, B. DeMarco, T. Schaetz, V. Meyer,
D. Leibfried, J. W. Britton, J. Chiaverini, W. M. Itano,
B. Jelenkovi´c, J. D. Jost, C. Langer, T. Rosenband, and
D. J. Wineland. Physical Review A 68, 042302 (2003).
[23] J. P. Home, D. Hanneke, J. D. Jost, J. M. Amini,
D. Leibfried, and D. J. Wineland. Science 325, 1227–
1230 (2009).
[24] C. Langer, R. Ozeri, J. D. Jost, J. Chiaverini, B. De-
Marco, A. Ben-Kish, R. B. Blakestad, J. Britton, D. B.
Hume, W. M. Itano, D. Leibfried, R. Reichle, T. Rosen-
band, T. Schaetz, P. O. Schmidt, and D. J. Wineland.
Physical Review Letters 95, 060502 (2005).
[25] P. O. Schmidt, T. Rosenband, C. Langer, W. M. Itano,
J. C. Bergquist, and D. J. Wineland. Science 309, 749–
752 (2005).
[26] D. B. Hume, T. Rosenband, and D. J. Wineland. Physical
Review Letters 99, 120502 (2007).
[27] C. J. Ballance. D.Phil. thesis, University of Oxford
(2014).
[28] J. P. Home, M. J. McDonnell, D. J. Szwer, B. C. Keitch,
D. M. Lucas, D. N. Stacey, and A. M. Steane. Physical
Review A 79, 050305 (2009).
[29] A. G. Fowler, M. Mariantoni, J. M. Martinis, and A. N.
Cleland, Physical Review A 86, 032324 (2012).
[30] T. R. Tan, J. P. Gaebler, Y. Lin, Y. Wan, R. Bowler,
D. Leibfried, and D. J. Wineland. arXiv:1508.03392.
[31] B. Hensen, H. Bernien, A. E. Dr´eau, A. Reiserer,
N. Kalb, M. S. Blok, J. Ruitenberg, R. F. L. Vermeulen,
R. N. Schouten, C. Abell´an, W. Amaya, V. Pruneri,
M. W. Mitchell, M. Markham, D. J. Twitchen, D. Elk-
ouss, S. Wehner, T. H. Taminiau, and R. Hanson.
arXiv:1508.05949.
[32] M. J. McDonnell, J.-P. Stacey, S. C. Webster, J. P. Home,
A. Ramos, D. M. Lucas, D. N. Stacey, and A. M. Steane.
Physical Review Letters 93, 153601 (2004).
[33] A. H. Myerson, D. J. Szwer, S. C. Webster, D. T. C.
Allcock, M. J. Curtis, G. Imreh, J. A. Sherman, D. N.
Stacey, A. M. Steane, and D. M. Lucas. Physical Review
Letters 100, 200502 (2008).
... One solution used to overcome this involves utilizing different types of qubits within a node. For example, trapping various ion species allows for individual addressing based on their distinct electronic transition frequencies [207], [208], [209]. Similarly, carbon-13 nuclear spins near a diamond N-V center offer a robust register of memory qubits that do not interact with the laser control fields on the N-V electron spin [210]. ...
Preprint
Full-text available
With the recent advancements and developments in quantum technologies, the emerging field of quantum communication and networking has gained the attention of the researchers. Owing to the unique properties of quantum mechanics, viz., quantum superposition and entanglement, this new area of quantum communication has shown potential to replace modernday communication technologies. The enhanced security and high information sharing ability using principles of quantum mechanics has encouraged networking engineers and physicists to develop this technology for next generation wireless systems. However, a conceptual bridge between the fundamentals of quantum mechanics, photonics and the deployability of a quantum communication infrastructure is not well founded in the current literature. This paper aims to fill this gap by merging the theoretical concepts from quantum physics to the engineering and computing perspectives of quantum technology. This paper builds the fundamental concepts required for understanding quantum communication, reviews the key concepts and demonstrates how these concepts can be leveraged for accomplishing successful communication. The paper delves into implementation advancements for executing quantum communication protocols, explaining how hardware implementation enables the achievement of all basic quantum computing operations. Finally, the paper provides a comprehensive and critical review of the state-of-the-art advancements in the field of quantum communication and quantum internet; and points out the recent trends, challenges and open problems for the real-world realization of next generation networking systems.
... The MS and LS gates have also been applied to the multi-species atomic ions [103][104][105][106][107][108][109][110]. The Oxford group has achieved gate fidelity of 99.8 % between a 43 Ca + hyperfine qubit and a 88 Sr + Zeeman qubit [109]. ...
Article
The trapped-ion system has been a leading platform for practical quantum computation and quantum simulation since the first scheme of a quantum gate was proposed by Cirac and Zoller (Phys Rev Lett 74:4091, 1995). Quantum gates with trapped ions have shown the highest fidelity among all physical platforms. Recently, sophisticated schemes of quantum gates such as amplitude, phase, frequency modulation, or multi-frequency application, have been developed to make the gates fast, robust to many types of imperfections, and applicable to multiple qubits. Here, we review the basic principle and recent development of quantum gates with trapped ions.
Preprint
Full-text available
We present a practical scheme for the efficient preparation of laser-cooled 43Ca+ ions in an ion trap. Our approach integrates two well-established methods: isotope-selective photoionization and isotope-specific parametric excitation. Drawing inspiration from the individual merits of each method, we have successfully integrated these techniques to prepare extended chains of 43Ca+ ions, overcoming the challenge posed by their low natural abundance of 0.135% in a natural source. Furthermore, we explore the subtleties of our scheme, focusing on the influence of different factors on the purification process. Our investigation contributes to a broader understanding of the technique and highlights the adaptability of established methods in addressing specific isotopic challenges.
Article
Ion trap is one of the leading physical platforms to implement quantum computation. Currently, high-fidelity elementary quantum operations above the fault-tolerant threshold, including state preparation, measurement and universal gates, have been demonstrated for tens of ionic qubits. One important future research direction is to further enlarge the qubit number to the scale required for solving practical problems while maintaining the high performance of individual qubits. This paper introduces the current mainstream schemes for scalable ion trap quantum computation like quantum charge-coupled device (QCCD) and ion-photon quantum network, and describes the main limiting factors in current research. Then we further explore new schemes to scale up the qubit number like two-dimensional ion crystals and dual-type qubit, and discuss the future research directions.
Article
Full-text available
Quantum computers have made extraordinary progress over the past decade, and significant milestones have been achieved along the path of pursuing universal fault-tolerant quantum computers. Quantum advantage, the tipping point heralding the quantum era, has been accomplished along with several waves of breakthroughs. Quantum hardware has become more integrated and architectural compared to its toddler days. The controlling precision of various physical systems is pushed beyond the fault-tolerant threshold. Meanwhile, quantum computation research has established a new norm by embracing industrialization and commercialization. The joint power of governments, private investors, and tech companies has significantly shaped a new vibrant environment that accelerates the development of this field, now at the beginning of the noisy intermediate-scale quantum era. Here, we first discuss the progress achieved in the field of quantum computation by reviewing the most important algorithms and advances in the most promising technical routes, and then summarizing the next-stage challenges. Furthermore, we illustrate our confidence that solid foundations have been built for the fault-tolerant quantum computer and our optimism that the emergence of quantum killer applications essential for human society shall happen in the future.
Article
We experimentally investigate nonlocal contextual relations between complementary photon polarizations by adapting the entanglement and the local polarizations of a two-photon state to satisfy three deterministic conditions demonstrating both quantum contextuality and nonlocality. The key component of this adaptive input state control is the variable degree of entanglement of the photon-pair source. Local polarization rotations can optimize two of the three correlations, and the variation of the entanglement optimizes the third correlation. Our results demonstrate that quantum contextuality is based on a nontrivial trade-off between local complementarity and quantum correlations.
Article
Full-text available
Entanglement witnesses are one of the most effective methods to detect entanglement. It is known that nonlinear entanglement witnesses provide better entanglement detection than their linear counterparts, in that the former detect a strictly larger subset of entangled states than the latter. Whether linear or nonlinear, the method is measurement-device dependent, so that imperfect measurements may cause false certification of entanglement in a shared state. Measurement-device-independent entanglement witnesses provide an escape from such measurement dependence of the entanglement detection for linear entanglement witnesses. Here we present measurement-device-independent nonlinear entanglement witnesses for non-positive partial transpose entangled states as well as for bound entangled states with positive partial transpose. Although the witness considered herein does not detect a larger set of entangled states than other nonlinear entanglement witnesses, it is more efficient in that it never leads to a false detection corresponding to wrong measurements. The constructed measurement-device-independent nonlinear entanglement witnesses certify the entanglement of the same sets of entangled states as their device-dependent parents do, and therefore are better than the linear entanglement witnesses, device-independent or otherwise.
Preprint
We experimentally investigate non-local contextual relations between complementary photon polarizations by adapting the entanglement and the local polarizations of a two-photon state to satisfy three deterministic conditions demonstrating both quantum contextuality and non-locality. The key component of this adaptive input state control is the variable degree of entanglement of the photon source. Local polarization rotations can optimize two of the three correlations, and the variation of the entanglement optimizes the third correlation. Our results demonstrate that quantum contextuality is based on a non-trivial trade-off between local complementarity and quantum correlations.
Article
The high-fidelity storage of quantum information is crucial for quantum computation and communication. Many experimental platforms for these applications exhibit highly biased noise, with good resilience to spin depolarization undermined by high dephasing rates. In this work, we demonstrate that the memory performance of a noise-biased trapped-ion-qubit memory can be greatly improved by incorporating error correction of dephasing errors through teleportation of the information between two repetition codes written on a pair of qubit registers in the same trap. While the technical requirements of error correction are often considerable, we show that our protocol can be achieved with a single global entangling phase gate of remarkably low fidelity, leveraging the fact that the gate errors are also dominated by dephasing-type processes. By rebalancing the logical spin-flip and dephasing error rates, we show that for realistic parameters our memory can exhibit error rates up to two orders of magnitude lower than the unprotected physical qubits, thus providing a useful means of improving memory performance in trapped-ion systems where field-insensitive qubits are not available.
Article
Full-text available
Precision control over hybrid physical systems at the quantum level is important for the realization of many quantum-based technologies. For trapped-ions, a hybrid system formed of different species introduces extra degrees of freedom that can be exploited to expand and refine the control of the system. Here, we demonstrate an entangling gate between two atomic ions of different elements that can serve as an important building block of quantum information processing (QIP), quantum networking, precision spectroscopy, metrology, and quantum simulation. An entangling geometric phase gate between a $^9$Be$^+$ ion and a $^{25}$Mg$^+$ ion is realized through an effective spin-spin interaction generated by state-dependent forces. Combined with single-qubit gates and same-species entangling gates, this mixed-element entangling gate provides a complete set of gates over such a hybrid system for universal QIP. Using a sequence of such gates, we demonstrate a Controlled-NOT (CNOT) gate and a SWAP gate. We also perform a CHSH-type Bell-inequality test on Bell states composed of different ion species.
Article
Full-text available
We report on the experimental violation of multipartite Bell inequalities by entangled states of trapped ions. First, we consider resource states for measurement-based quantum computation of between 3 and 7 ions and show that all strongly violate a Bell-type inequality for graph states, where the criterion for violation is a sufficiently high fidelity. Second, we analyze Greenberger-Horne-Zeilinger states of up to 14 ions generated in a previous experiment using stronger Mermin-Klyshko inequalities, and show that in this case the violation of local realism increases exponentially with system size. These experiments represent a violation of multipartite Bell-type inequalities of deterministically prepared entangled states. In addition, the detection loophole is closed.
Article
Full-text available
The great potential of quantum computing requires two essential ingredients for its realization: high-fidelity quantum logic operations and a physical implementation which can be scaled up to large numbers of quantum bits. We introduce a trapped-ion qubit stored in ultrastable "atomic clock" states of $^{43}$Ca$^+$, in which we implement all single-qubit operations with fidelities sufficient for fault-tolerant quantum computing. We measure a combined qubit state preparation and single-shot readout fidelity of 99.93%, a memory coherence time of T$_2^*$ = 50 seconds, and an average single-qubit gate fidelity of 99.9999%. These results are achieved in a room-temperature device without the use of magnetic field shielding or dynamic decoupling techniques to overcome technical noise. The surface-electrode ion trap chip incorporates integrated resonators and waveguides for coherent manipulation of the qubit using near-field microwaves. Two-qubit gates and individual qubit addressing have already been demonstrated using this approach, which is scalable for a many-qubit architecture.
Article
Full-text available
We have measured the linear polarization correlation of the photons emitted in an atomic cascade of calcium. It has been shown by a generalization of Bell's inequality that the existence of local hidden variables imposes restrictions on this correlation in conflict with the predictions of quantum mechanics. Our data, in agreement with quantum mechanics, violate these restrictions to high statistical accuracy, thus providing strong evidence against local hidden-variable theories.
Article
Full-text available
We present a source of entangled photons that violates a Bell inequality free of the "fair-sampling" assumption, by over 7 standard deviations. This violation is the first experiment with photons to close the detection loophole, and we demonstrate enough "efficiency" overhead to eventually perform a fully loophole-free test of local realism. The entanglement quality is verified by maximally violating additional Bell tests, testing the upper limit of quantum correlations. Finally, we use the source to generate secure private quantum random numbers at rates over 4 orders of magnitude beyond previous experiments.
Article
More than 50 years ago, John Bell proved that no theory of nature that obeys locality and realism can reproduce all the predictions of quantum theory: in any local-realist theory, the correlations between outcomes of measurements on distant particles satisfy an inequality that can be violated if the particles are entangled. Numerous Bell inequality tests have been reported; however, all experiments reported so far required additional assumptions to obtain a contradiction with local realism, resulting in 'loopholes'. Here we report a Bell experiment that is free of any such additional assumption and thus directly tests the principles underlying Bell's inequality. We use an event-ready scheme that enables the generation of robust entanglement between distant electron spins (estimated state fidelity of 0.92 ± 0.03). Efficient spin read-out avoids the fair-sampling assumption (detection loophole), while the use of fast random-basis selection and spin read-out combined with a spatial separation of 1.3 kilometres ensure the required locality conditions. We performed 245 trials that tested the CHSH-Bell inequality S ≤ 2 and found S = 2.42 ± 0.20 (where S quantifies the correlation between measurement outcomes). A null-hypothesis test yields a probability of at most P = 0.039 that a local-realist model for space-like separated sites could produce data with a violation at least as large as we observe, even when allowing for memory in the devices. Our data hence imply statistically significant rejection of the local-realist null hypothesis. This conclusion may be further consolidated in future experiments; for instance, reaching a value of P = 0.001 would require approximately 700 trials for an observed S = 2.4. With improvements, our experiment could be used for testing less-conventional theories, and for implementing device-independent quantum-secure communication and randomness certification.
Article
We demonstrate the cooling of a two species ion crystal consisting of one ${}^{9}{\mathrm{Be}}^{+}$ and one ${}^{24}{\mathrm{Mg}}^{+}$ ion. Since the respective cooling transitions of these two species are separated by more than 30 nm, laser manipulation of one ion has negligible effect on the other even when the ions are not individually addressed. As such this is a useful system for reinitializing the motional state in an ion trap quantum computer without affecting the qubit information. Additionally, we have found that the mass difference between ions enables a method for detecting and subsequently eliminating the effects of radio frequency micromotion.
Article
We study the speed/fidelity trade-off for a two-qubit phase gate implemented in $^{43}$Ca$^+$ hyperfine trapped-ion qubits. We characterize various error sources contributing to the measured fidelity, allowing us to account for errors due to single-qubit state preparation, rotation and measurement (each at the $\sim0.1\%$ level), and to identify the leading sources of error in the two-qubit entangling operation. We achieve gate fidelities ranging between $97.1(2)\%$ (for a gate time $t_g=3.8\mu$s) and $99.9(1)\%$ (for $t_g=100\mu$s), representing respectively the fastest and lowest-error two-qubit gates reported between trapped-ion qubits by nearly an order of magnitude in each case.