ArticlePDF Available

Generation of a Maximally Entangled State Using Collective Optical Pumping

Authors:

Abstract

We propose and implement a novel scheme for dissipatively pumping two qubits into a singlet Bell state. The method relies on a process of collective optical pumping to an excited level, to which all states apart from the singlet are coupled. We apply the method to deterministically entangle two trapped ^{40}Ca^{+} ions. Within 16 pumping cycles, an initially separable state is transformed into one with 83(1)% singlet fidelity, and states with initial fidelity of ⪆70% converge onto a fidelity of 93(1)%. We theoretically analyze the performance and error susceptibility of the scheme and find it to be insensitive to a large class of experimentally relevant noise sources.
Generation of a Maximally Entangled State Using Collective Optical Pumping
M. Malinowski ,1,* C. Zhang ,1V. Negnevitsky,1I. Rojkov ,1F. Reiter ,1T.-L. Nguyen ,1
M. Stadler ,1D. Kienzler ,1K. K. Mehta,1and J. P. Home1,2,
1Institute for Quantum Electronics, ETH Zürich, 8093 Zürich, Switzerland
2Quantum center, ETH Zürich, 8093 Zürich, Switzerland
(Received 10 August 2021; revised 10 December 2021; accepted 11 January 2022; published 22 February 2022)
We propose and implement a novel scheme for dissipatively pumping two qubits into a singlet Bell state.
The method relies on a process of collective optical pumping to an excited level, to which all states apart
from the singlet are coupled. We apply the method to deterministically entangle two trapped 40Caþions.
Within 16 pumping cycles, an initially separable state is transformed into one with 83(1)% singlet fidelity,
and states with initial fidelity of 70% converge onto a fidelity of 93(1)%. We theoretically analyze the
performance and error susceptibility of the scheme and find it to be insensitive to a large class of
experimentally relevant noise sources.
DOI: 10.1103/PhysRevLett.128.080503
Quantum entanglement is a resource for quantum com-
putation [1], communication [2], cryptography [3], and
metrology [4]. Entangled states are typically prepared using
a two-step process, the first involving initialization of a
separable state by optical pumping, followed by a unitary
transformation which generates entanglement [5]. In such
an open-loop process the final state is sensitive to the
parameters of the drive used to create it and is not protected
from future errors. An alternative mode of operation is to
use a closed-loop process, where feedback from a low-
entropy reference system drives the system continuously
toward the desired state or subspace. This can be done using
measurement-conditioned classical control (e.g., quantum
error correction or outcome heralding) or through dissipative
engineering [610]. Dissipation engineering allows useful
quantum states to be created in the steady state, making the
process self-correcting with regard to transient errors
[11,12], and resulting in a resource state or subspace which
is continuously available. Entanglement of qubits using
dissipative engineering has previously been demonstrated
using trapped ions [13,14],atomicensembles[15],and
superconducting circuits [1618]. Beyond qubit-based
approaches, reservoir engineering has been used to create
and stabilize nonclassical states of bosonic systems [6,19,20]
as well as to perform quantum error correction [20,21].
A widely used strategy for dissipation engineering is to
rely on engineered resonances, whereby pumping into the
desired entangled state is achieved by resonant drives,
while leakage processes out of the desired state are off
resonant [11,14,22,23]. This approach has proven to be
versatile, and has been theoretically extended to the
generation of multiqubit states [2426], quantum error
correction [27,28], and quantum simulation [29]. However,
these protocols can be slow to converge. This is because, in
order to suppress leakage processes, the drives need to be
weak compared to the splittings of the resonances. The
resulting competition with additional uncontrolled dissipa-
tion channels limits the achievable fidelities. It has been
proposed that this issue could be overcome by dissipative
schemes based on symmetry [3034].
In this Letter, we present a method for dissipatively
generating two-body entanglement using a deterministic
collective optical pumping process which does not couple
to the target entangled state: the singlet Bell state jΨi
ðj↑↓ij↓↑=
ffiffi
2
p. Unlike previous demonstrations, our
method relies on symmetry, involving only global fields
which couple equally to each system. We thereby overcome
the speed limitations of previous schemes, achieving a
faster convergence. Our scheme is robust to global error
processes. We implement the protocol using two trapped
ions in a surface-electrode trap with integrated optical
control fields [35], achieving a 93(1)% fidelity with jΨi.
Compared to earlier trapped-ion approaches, our method
does not require ground-state cooling.
The scheme is illustrated in Fig. 1(a). We consider a spin
ground-state manifold consisting of the collective states
j↓↓i;j↑↑i;jΨþiðj↑↓iþj↓↑=
ffiffi
2
p(spin triplet) and
jΨi(spin singlet), as well as excited states, of which
the most important for our purposes consists of both
systems in a particular excited state jei. Three elements
define the pumping process. The first is a collective
excitation (A) from j↓↓ito jeei. Its collective nature
Published by the American Physical Society under the terms of
the Creative Commons Attribution 4.0 International license.
Further distribution of this work must maintain attribution to
the author(s) and the published articles title, journal citation,
and DOI.
PHYSICAL REVIEW LETTERS 128, 080503 (2022)
0031-9007=22=128(8)=080503(7) 080503-1 Published by the American Physical Society
means that it does not couple to the other states in the
ground-state manifold. The state jeeiis quenched through a
decay channel (B), redistributing its population into all four
spin states. (A) and (B) together provide a collective optical
pumping process which moves the population from j↓↓ito
the other ground states. To prepare only the singlet, this is
supplemented by a symmetric drive (C) which resonantly
drives both spins with equal amplitude and phase. Because
of its symmetry, this drive cycles population within the
triplet subspace, while leaving the singlet untouched. Thus
the triplet states have a chance of being repumped through
the collective pumping, while population in the singlet is
dark to all drives. jΨithen becomes the steady state of the
system. The protocol can be implemented continuously or
by sequentially applying each component. For our imple-
mentation, we expect the latter to be more robust to
experimental imperfections and analyze it below. The
continuous implementation is analyzed in detail in
Supplemental Material (SM), Sec. VI [36].
To identify optimal settings, we optimize a superoperator
which combines the three drives. For the collective exci-
tation this is derived from a unitary,
UAðΦÞ¼eiΦS2
x;e ;ð1Þ
with Sx;e ¼σx;e1þ1σx;e,σx;e¼jeihjþjihej,
and 1 is a 3×3identity operator. This provides a full
transfer from j↓↓ito jeeifor Φ¼π=4. Drive (B) repumps
the population from jeiwith branching ratios parametrized
by pe=pe¼tan2ðγÞ. Drive (C) is described by a
unitary UCðθÞ¼exp½iðθ=2Þσxexp½iðθ=2Þσx, where
σx¼jihjþjihj. After N1cycles, the singlet
fidelity, defined as FðjΨiÞ¼hΨjρjΨifor the system
density matrix ρ, approaches unity as FðjΨ
1expðN=N0Þ. Through eigenvalue analysis we find
the most rapid convergence for Φ¼π=4,θ0.72π,and
γ0.22π, where N0¼7.62 cycles (SM, Sec. I [36]). The
steady state is insensitive to the values of Φ,γ,andθ, i ndicating
that these parameters do not require precise calibration.
We implement the protocol on a pair of 40
Caþions
confined in the surface-electrode trap described in
Ref. [35]. The qubit is encoded into ground-state Zeeman
sublevels ji¼jS1=2;m
j¼1=2iand ji¼jS1=2;m
j¼
þ1=2iwhich have a frequency splitting of 2π×16.5MHz
in the applied magnetic field of 0.59 mT. We use an
ancilliary state jei¼jD5=2;m
j¼1=2i. Narrow-linewidth
laser light at 729 nm is delivered through trap-integrated
photonics, and coherently drives the S1=2D5=2transi-
tions. Free-space laser beams are used for cooling, repump-
ing, and readout. The jijitransition is driven by
resonant radio-frequency magnetic fields.
The collective excitation step (A) is implemented using a
bichromatic 729 nm laser field with Rabi frequency Ωand
two frequency components detuned by δ¼2π×14.7kHz
from the red and blue motional sidebands of the jijei
transition, using the axial stretch mode (where ions oscillate
out of phase) at ωm2π×2.4MHz for which the Lamb-
Dicke parameter η¼0.026. This results in a Hamiltonian
HA¼1
2ηΩSx;eðˆ
aeiδtþˆ
aeiδtÞwhich implements a force
on the oscillator whose phase depends on the eigenstate
of Sx;e.ThisiscommonlyreferredtoasaMølmer-Sørensen
drive, and is one of the primary methods for performing two-
qubit gates with trapped ions [3739]. A pulse of duration t
then results in the unitary
UA¼e½αðtÞˆaαðtÞˆaSx;e eiΦðtÞS2
x;e ;ð2Þ
where αðtÞ¼iðηΩ=δÞeiδt=2sinðδt=2Þis an oscillator
phase-space displacement amplitude and ΦðtÞ¼
ðη2Ω2=4δ2Þ½δtsinðδtÞ is a collective phase factor.
Equation (2) reduces to a pure S2
x;e coupling of the form of
Eq. (1) in two cases. The first, appropriate to a continuous
implementation, is when jδjηΩand so the oscillator
excitation can be neglected [40]. The second, which is the
main focus of this Letter, is when t¼2nπ=δwith nZ,for
which Φ¼nπη2Ω2=ð2δ2Þ[41]. Repump (B) is implemented
using a laser at 854 nm, which couples all D5=2sublevels to
the short-lived P3=2states, which primarily decay into the
ground-state manifold. A second decay channel to the D3=2
states is repumped using a laser at 866 nm. After 5μs, we
measure a probability of leaving D5=2of >0.9999 with a
branching ratio of γ0.3π. The symmetric drive (C) is
implemented by passing a current through a track on a circuit
board at around 1 mm distance from the ions.
We employ a number of error mitigation techniques. The
collective excitation step (A) is implemented as a sequence
of two pulses (t¼2nπ=δwith n¼2) with ηΩ¼δ=2,
resulting in Φ¼π=4and a drive time of t¼150 μs.
(a) (b)
FIG. 1. (a) High-level description of the protocol. When drives
(A), (B), and (C) are switched on, the system is pumped into a
maximally entangled state jΨi. (b) Atomic transitions and drives
in 40Caþused in this work. (A) and (B) are driven by laser beams,
while (C) is implemented by an oscillating Bfield. Dashed lines
denote motional sidebands of the jijeitransition.
PHYSICAL REVIEW LETTERS 128, 080503 (2022)
080503-2
The phase of the force acting on the oscillator is shifted by
πfor the second pulse, thus canceling any residual
displacement produced by a single pulse [42]. To our
surprise the ac Stark shift of the jijitransition
produced by the collective drive was different by 2π×
2.5kHz on the two ions, causing a near-complete failure of
the protocol (SM, Sec. IV [36]). To mitigate this, we
replace the optimal value of θapplied in each cycle with
two values, θ1¼πapplied in odd cycles (drive time
tC¼6.4μs) and θ¼π=2(tC¼3.2μs) applied in even
cycles. This has the desired effect of a spin echo, but at the
cost that high-fidelity singlet states are produced only after
even cycles. One cycle of the protocol takes 165 μson
average. In the absence of other errors, the protocol
produces jΨiregardless of the ionstemperature.
However, finite temperature amplifies existing errors asso-
ciated with residual oscillator excitation [i.e., when
jαðtÞj >0]. For this reason, we cool the ion close to the
motional ground state.
We measure the Pð↓↓Þ,Pð↓↑ÞþPð↑↓Þ,andPð↑↑Þ
populations by shelving jiinto ancillary D5=2sublevels,
followed by state-dependent fluorescence [43]. This allows
us to extract the ground-state parity hσzσzi,whilehσxσxiand
hσyσyiare obtained by measuring the parity following radio-
frequency spin rotations exp½iðπ=2Þσxexp½iðπ=2Þσx
and exp½iðπ=2Þσyexp½iðπ=2Þσy, respectively. These
are combined to estimate the singlet state fidelity, using
FðjΨ ¼ 1
4ð1hσxσxihσyσyihσzσz.
Figure 2(a) shows the measured fidelity as a function of
the number of cycles of the protocol, applied to a range of
initial states, showing the expected convergence toward the
singlet. Different starting states were created by initializing
the ions to j↓↓iand mapping it to a mixture of singlet and
triplet states (SM. Sec. V [36]). Figure 2(b) illustrates how,
after 16 cycles of the protocol, states with initial fidelities
0.75 are mapped onto output states with the same final
fidelity. Averaged over all the data, we find a fidelity of
93(1)% at 16 cycles.
We analyze the noise robustness of the protocol by
considering the bichromatic drive (A) as the dominant
source of errors. Assuming jiis spectroscopically
decoupled from (A), we can describe all errors through
16 elementary error channels fIe;X
e;Ye;Z
eg2. These act
as Pauli operators on the fji;jeig subspace, and as
identity on ji(SM, Sec. II [36]). The effect of those
errors acting with probability pper cycle in a depolarizing
model is shown in Fig. 3(a). We find that the final fidelity is
independent of all global errors (such as XeXeor XeZe).
On the other hand, all local errors (such as XeIeor IeZe)
become amplified. A particularly experimentally relevant
class of errors is correlated local errors. These include
correlated bit-flip errors (corresponding to an application of
the operator IeXeþXeIe) which arise due to residual spin-
motion entanglement at the end of the collective excitation
step [i.e., when αðtÞ0] or off-resonant excitation of
spectatortransitions (i.e., nearby undesired resonances).
Magnetic-field fluctuations common to both ions would
produce a correlated phase-flip error (IeZeþZeIe). We
find that such correlations increase the fidelity of the
collective optical pumping compared with uncorrelated
errors with similar constituent operators. Results of sim-
ulations showing this are displayed in Fig. 3(b).For
example, a bit-flip error with probability pper cycle
reduces the singlet fidelity by 5.2pwhen uncorrelated
and 3.2pwhen correlated. Correlated phase-flip errors
leave the fidelity unaffected since jΨiresides in a
decoherence-free subspace.
These insights are matched by simulations of the
dynamics of the collective optical pumping in the presence
of experimentally relevant error sources. These reveal that,
compared to either a single entangling gate or a two-
loop phase-modulated entangling gate [44] based on the
Hamiltonian HA, our protocol reduces the effect of qubit
frequency errors and Rabi frequency errors. For motional
frequency errors and fast (Markovian) optical qubit dephas-
ing, our protocol does not provide benefits. Details and
discussion of practical applicability of these results are
presented in SM, Sec. III [36].
It is challenging to exactly account for the measured
error from first principles due to a number of setup-specific
imperfections. We experience kilohertz-level drifts in
(a)
(b)
FIG. 2. Entanglement generation. (a) The effect of applying up
to 16 cycles of the protocol. Measurements for different initial
states are shown in different colors and connected by dashed
lines. All cases converge toward jΨi. (b) Comparison of the
singlet fidelity before and after the protocol is applied. The phase
ϕprep of the preparation pulse (SM, Sec. V [36]) sets the input
fidelity to FðjΨ 0.88cos2ϕprep (green dashed line). After 16
cycles, states with fidelities 0.75 converge onto the same state
of FðjΨ 0.93 (blue dashed line). Statistical 1σerror bars
are smaller than data points, and the result spread is dominated by
experimental drifts.
PHYSICAL REVIEW LETTERS 128, 080503 (2022)
080503-3
motional mode frequencies due to charging of the trap
surface by light shining through the integrated waveguides
[45]. These occasionally lead to a mode spectrum where the
collective excitation step off-resonantly excites spectator
optical transitions. This error can be corrected by tuning
mode frequencies, but it is challenging to estimate its
magnitude between calibrations. Mode frequency drifts
associated with laser power changes were also the primary
reason we worked with a fixed number of protocol cycles
(N¼16, corresponding to 3ms of 729 nm light per
shot). High heating rates mean that each cycle of the
protocol starts with a higher occupancy of motional modes
(0.5quanta per cycle on a 1 MHz center-of-mass mode),
leading to an increase in a correlated bit-flip error during
the drive (A) as the protocol progresses. The combined
effect of all the error sources is a bit-flip error probability of
p0.01 for the first cycle, and p0.02 after 16 cycles.
The measured 16-cycle fidelity 93(1)% is consistent with
the value of 13.2ppredicted by the correlated bit-flip
error model. The obtained fidelity is significantly reduced
compared to the unitary gate based on the same
Hamiltonian HA, which produces ðj↓↓iijee=
ffiffi
2
pwith
fidelity of 99% [35], though it does improve our ability
to prepare jΨi, which is currently limited by errors in
single-ion addressing in our coherent implementation.
In order to verify that the correlated bit-flip error model
captures the essential performance limitations of the pro-
tocol, we measure the 16-cycle fidelity, as well as the
bit-flip probability in the collective excitation step, for a
range of experimental miscalibrations. For each parameter,
we experimentally approximate the steady-state fidelity
FðjΨ by first preparing jΨiwith fidelity around 0.75
using unitary methods, and then applying 16 cycles of the
protocol. The bit-flip probability pfor the last cycle is
independently estimated by applying 16 cycles of the
protocol, followed by optical pumping and a single round
of drive (A) (SM, Sec. V [36]). The comparison between ;
FðjΨ and the bit-flip model prediction is shown in
Fig. 3(c). We find qualitative agreement, suggesting that the
bit-flip model accurately captures the errors of the protocol.
Figure 3(d) illustrates the challenge associated with spectral
crowding. We modify the spectator mode spectrum by
adding an additional quadrupole potential with eigenaxes at
45° to the trap surface in the radial plane. This changes
the radial mode orientations, frequencies, and temperatures,
while keeping the (axial) gate mode frequency approx-
imately constant. We find that the spectator spectrum is
clear only for a narrow range of curvatures (here between
5.7×107and 6.1×107V=m2), which needs recalibrating
every few hours.
All of the limitations listed above are setup specific and
do not pose a fundamental limitation to the protocol.
Coupling to spectator transitions could be suppressed by
increasing the magnetic field. The heating rate observed in
this trap is particularly high, exceeding levels observed in
cryogenic traps with comparable ion-electrode distances
by a factor of 100 [46,47]. Reducing it to more typical
levels, combined with better shielding of nearby dielectrics
[48,49], would suppress drifts within each collective
pumping sequence and hence allow the protocol to reach
(a)
(b) (d)
(c)
FIG. 3. (a) Simulated steady-state error associated with individual error channels of probability p. (b) Comparison of simulated steady-
state errors associated with uncorrelated (solid lines) and correlated (dashed lines) errors. (c),(d) Experimentally measured values of
FðjΨ (blue dots) compared to the prediction (13.2p) of a correlated bit-flip error model (gray lines), with pobtained from
independent experimental measurements. Error bars (smaller than most data points) show a 1σconfidence interval. In several data
points in (d) the measured values of pare large, and we can no longer apply the linear approximation. Inset in (d) highlights the typical
operation region
PHYSICAL REVIEW LETTERS 128, 080503 (2022)
080503-4
its true steady state. Alternatively, motional mode temper-
ature could be stabilized throughout the protocol by
sympathetic cooling [50].
We have presented and implemented a novel protocol
for collective optical pumping into a maximally entangled
two-qubit state. We measure a singlet fidelity of 93(1)%
after 16 cycles of the protocol (at which point a quasisteady
state has been achieved), to our knowledge exceeding
previously reported dissipative methods, and slightly below
a simultaneous work by Cole et al. [34]. The observed
infidelity is consistent with measured bit-flip errors of the
effective S2
xdrive which for our implementation is mediated
by a motional mode. The protocol can be practically
beneficial in experiments limited by global errors,
especially as a method of purifying lower-fidelity Bell
states. Dissipative generation of high-fidelity entangled
states could find application in a variety of quantum
information processing tasks, such as dissipative encoding
[51], error-corrected quantum sensing [52], and as a
supply of entangled resource states for quantum gate
teleportation [53].
While the analysis in this Letter focused on a specific
implementation in 40
Caþions, the protocol is general, and we
anticipate it might be applied in a wide range of platforms
where collective excitation may be engineered. These
include nitrogen-vacancy centers (via direct spin-spin inter-
actions [54]), neutral atom platforms (via Rydberg dressing
[55]) or superconductors (via parametric drives [32]).
We acknowledge funding from the Swiss National
Science Foundation (Grant No. 200020_165555), the
National Centre of Competence in Research for Quantum
Science and Technology (QSIT), ETH Zürich, and the
Intelligence Advanced Research Projects Activity
(IARPA) via the U.S. Army Research Office Grant
No. W911NF-16-1-0070. F. R. and I. R. acknowledge finan-
cial support from the Swiss National Science Foundation
(Ambizione Grant No. PZ00P2_186040). J. P. H. devised the
scheme, which was then simulated and analyzed by M. M.
and V. N. M .M., C. Z., and K. K. M. carried out the experi-
ments presented here in an apparatus with significant
contributions from M. M, C. Z., K. K. M., T.-L. N., and
M. S. M. M. analyzed the data and constructed the error
model. I. R. and F. R. developed an analytic approach to
describe the protocol. V. N. performed an initial experimen-
tal investigation. J. P. H., K. K. M., and D. K. supervised the
work. The Letter was written by M. M., F. R., and J. P. H.
with input from all authors.
*maciejm@phys.ethz.ch
jhome@phys.ethz.ch
[1] R. Jozsa and N. Linden, On the role of entanglement in
quantum-computational speed-up, Proc. R. Soc. A 459,
2011 (2003).
[2] C. H. Bennett and S. J. Wiesner, Communication via One-
and Two-Particle Operators on Einstein-Podolsky-Rosen
States, Phys. Rev. Lett. 69, 2881 (1992).
[3] A. K. Ekert, Quantum Cryptography Based on Bells
Theorem, Phys. Rev. Lett. 67, 661 (1991).
[4] G. Tóth and I. Apellaniz, Quantum metrology from a
quantum information science perspective, J. Phys. A 47,
424006 (2014).
[5] D. P. DiVincenzo, The physical implementation of quantum
computation, Fortschr. Phys. 48, 771 (2000).
[6] J. F. Poyatos, J. I. Cirac, and P. Zoller, Quantum Reservoir
Engineering with Laser Cooled Trapped Ions, Phys. Rev.
Lett. 77, 4728 (1996).
[7] B. Kraus, H. P. Büchler, S. Diehl, A. Kantian, A. Micheli,
and P. Zoller, Preparation of entangled states by quantum
Markov processes, Phys. Rev. A 78, 042307 (2008).
[8] M. B. Plenio, S. F. Huelga, A. Beige, and P. L. Knight,
Cavity-loss-induced generation of entangled atoms, Phys.
Rev. A 59, 2468 (1999).
[9] F. Verstraete, M. M. Wolf, and J. Ignacio Cirac, Quantum
computation and quantum-state engineering driven by dis-
sipation, Nat. Phys. 5, 633 (2009).
[10] F. Ticozzi and L. Viola, Steady-state entanglement by
engineered quasi-local Markovian dissipation, Quantum
Inf. Comput. 14, 265 (2014).
[11] M. J. Kastoryano, F. Reiter, and A. S. Sørensen, Dissipative
Preparation of Entanglement in Optical Cavities, Phys. Rev.
Lett. 106, 090502 (2011).
[12] G. Morigi, J. Eschner, C. Cormick, Y. Lin, D. Leibfried, and
D. J. Wineland, Dissipative Quantum Control of a Spin
Chain, Phys. Rev. Lett. 115, 200502 (2015).
[13] J. T. Barreiro, M. Müller, P. Schindler, D. Nigg, T. Monz, M.
Chwalla, M. Hennrich, C. F. Roos, P. Zoller, and R. Blatt,
An open-system quantum simulator with trapped ions,
Nature (London) 470, 486 (2011).
[14] Y. Lin, J. P. Gaebler, F. Reiter, T. R. Tan, R. Bowler, A. S.
Sørensen, D. Leibfried, and D. J. Wineland, Dissipative
production of a maximally entangled steady state of two
quantum bits, Nature (London) 504, 415 (2013).
[15] H. Krauter, C. A. Muschik, K. Jensen, W. Wasilewski, J. M.
Petersen, J. I. Cirac, and E. S. Polzik, Entanglement Gen-
erated by Dissipation and Steady State Entanglement of
Two Macroscopic Objects, Phys. Rev. Lett. 107, 080503
(2011).
[16] S. Shankar, M. Hatridge, Z. Leghtas, K. M. Sliwa, A. Narla,
U. Vool, S. M. Girvin, L. Frunzio, M. Mirrahimi, and M. H.
Devoret, Autonomously stabilized entanglement between
two superconducting quantum bits, Nature (London) 504,
419 (2013).
[17] Y. Liu, S. Shankar, N. Ofek, M. Hatridge, A. Narla, K. M.
Sliwa, L. Frunzio, R. J. Schoelkopf, and M. H. Devoret,
Comparing and Combining Measurement-Based and
Driven-Dissipative Entanglement Stabilization, Phys. Rev.
X6, 011022 (2016).
[18] M. E. Kimchi-Schwartz, L. Martin, E. Flurin, C. Aron,
M. Kulkarni, H. E. Tureci, and I. Siddiqi, Stabilizing
Entanglement via Symmetry-Selective Bath Engineering
in Superconducting Qubits, Phys. Rev. Lett. 116, 240503
(2016).
PHYSICAL REVIEW LETTERS 128, 080503 (2022)
080503-5
[19] D. Kienzler, H. Y. Lo, V. Negnevitsky, C. Flühmann, M.
Marinelli, and J. P. Home, Quantum Harmonic Oscillator
State Control in a Squeezed Fock Basis, Phys. Rev. Lett.
119, 033602 (2017).
[20] B. de Neeve, T. L. Nguyen, T. Behrle, and J. Home, Error
correction of a logical grid state qubit by dissipative
pumping, arXiv:2010.09681.
[21] J. M. Gertler, B. Baker, J. Li, S. Shirol, J. Koch, and C.
Wang, Protecting a bosonic qubit with autonomous quan-
tum error correction, Nature (London) 590, 243 (2021).
[22] G. Vacanti and A. Beige, Cooling atoms into entangled
states, New J. Phys. 11, 083008 (2009).
[23] F. Reiter and A. S. Sørensen, Effective operator formalism
for open quantum systems, Phys. Rev. A 85, 032111
(2012).
[24] J. Cho, S. Bose, and M. S. Kim, Optical Pumping into
Many-Body Entanglement, Phys. Rev. Lett. 106, 020504
(2011).
[25] Y. Lin, J. P. Gaebler, F. Reiter, T. R. Tan, R. Bowler, Y. Wan,
A. Keith, E. Knill, S. Glancy, K. Coakley, A. S. Sørensen, D.
Leibfried, and D. J. Wineland, Preparation of Entangled
States through Hilbert Space Engineering, Phys. Rev. Lett.
117, 140502 (2016).
[26] F. Reiter, D. Reeb, and A. S. Sørensen, Scalable Dissipative
Preparation of Many-Body Entanglement, Phys. Rev. Lett.
117, 040501 (2016).
[27] J. Cohen and M. Mirrahimi, Dissipation-induced continuous
quantum error correction for superconducting circuits, Phys.
Rev. A 90, 062344 (2014).
[28] F. Reiter, A. S. Sørensen, P. Zoller, and C. A. Muschik,
Dissipative quantum error correction and application to
quantum sensing with trapped ions, Nat. Commun. 8,1
(2017).
[29] F. Reiter, F. Lange, S. Jain, M. Grau, J. P. Home, and Z.
Lenarčič, Engineering generalized Gibbs ensembles with
trapped ions, Phys. Rev. Research 3, 033142 (2021).
[30] C. D. B. Bentley, A. R. R. Carvalho, D. Kielpinski, and J. J.
Hope, Detection-Enhanced Steady State Entanglement with
Ions, Phys. Rev. Lett. 113, 040501 (2014).
[31] K. P. Horn, F. Reiter, Y. Lin, D. Leibfried, and C. P. Koch,
Quantum optimal control of the dissipative production
of a maximally entangled state, New J. Phys. 20, 123010
(2018).
[32] E. Doucet, F. Reiter, L. Ranzani, and A. Kamal, High
fidelity dissipation engineering using parametric inter-
actions, Phys. Rev. Research 2, 023370 (2020).
[33] D. C. Cole, J. J. Wu, S. D. Erickson, P.-Y. Hou, A. C.
Wilson, D. Leibfried, and F. Reiter, Dissipative preparation
of W states in trapped ion systems, New J. Phys. 23, 073001
(2021).
[34] D. C. Cole, S. D. Erickson, G. Zarantonello, K. P. Horn,
P.-Y. Hou, J. J. Wu, D. H. Slichter, F. Reiter, C. P. Koch, and
D. Leibfried, preceding Letter, Resource-Efficient Dissipa-
tive Entanglement of Two Trapped-Ion Qubits, Phys. Rev.
Lett. 128, 080502 (2022).
[35] K. K. Mehta, C. Zhang, M. Malinowski, T. L. Nguyen, M.
Stadler, and J. P. Home, Integrated optical multi-ion quan-
tum logic, Nature (London) 586, 533 (2020).
[36] See Supplemental Material at http://link.aps.org/
supplemental/10.1103/PhysRevLett.128.080503 for ex-
tended discussion of the protocol and the experiment.
[37] A. Sørensen and K. Mølmer, Quantum Computation
with Ions in Thermal Motion, Phys. Rev. Lett. 82, 1971
(1999).
[38] J. Benhelm, G. Kirchmair, C. F. Roos, and R. Blatt, Towards
fault-tolerant quantum computing with trapped ions, Nat.
Phys. 4, 463 (2008).
[39] J. P. Gaebler, T. R. Tan, Y. Lin, Y. Wan, R. Bowler, A. C.
Keith, S. Glancy, K. Coakley, E. Knill, D. Leibfried, and
D. J. Wineland, High-Fidelity Universal Gate Set for 9Beþ
Ion Qubits, Phys. Rev. Lett. 117, 060505 (2016).
[40] K. Kim, M. S. Chang, S. Korenblit, R. Islam, E. E. Edwards,
J. K. Freericks, G. D. Lin, L. M. Duan, and C. Monroe,
Quantum simulation of frustrated Ising spins with trapped
ions, Nature (London) 465, 590 (2010).
[41] G. Kirchmair, J. Benhelm, F. Zähringer, R. Gerritsma,
C. F. Roos, and R. Blatt, Deterministic entanglement of
ions in thermal states of motion, New J. Phys. 11, 023002
(2009).
[42] D. Hayes, S. M. Clark, S. Debnath, D. Hucul, I. V. Inlek,
K. W. Lee, Q. Quraishi, and C. Monroe, Coherent Error
Suppression in Multiqubit Entangling Gates, Phys. Rev.
Lett. 109, 020503 (2012).
[43] 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, High-Fidelity Readout of
Trapped-Ion Qubits, Phys. Rev. Lett. 100, 200502 (2008).
[44] A. R. Milne, C. L. Edmunds, C. Hempel, F. Roy, S.
Mavadia, and M. J. Biercuk, Phase-Modulated Entangling
Gates Robust to Static and Time-Varying Errors, Phys. Rev.
Applied 13, 024022 (2020).
[45] M. Harlander, M. Brownnutt, W. Hänsel, and R. Blatt,
Trapped-ion probing of light-induced charging effects on
dielectrics, New J. Phys. 12, 093035 (2010).
[46] M. Brownnutt, M. Kumph, P. Rabl, and R. Blatt, Ion-trap
measurements of electric-field noise near surfaces, Rev.
Mod. Phys. 87, 1419 (2015).
[47] J. A. Sedlacek, J. Stuart, D. H. Slichter, C. D. Bruzewicz, R.
McConnell, J. M. Sage, and J. Chiaverini, Evidence for
multiple mechanisms underlying surface electric-field noise
in ion traps, Phys. Rev. A 98, 063430 (2018).
[48] S. Hong, Y. Kwon, C. Jung, M. Lee, T. Kim, and D. I. D.
Cho, A new microfabrication method for ion-trap chips that
reduces exposure of dielectric surfaces to trapped ions,
J. Microelectromech. Syst. 27, 28 (2018).
[49] S. Ragg, C. Decaroli, T. Lutz, and J. P. Home, Segmented
ion-trap fabrication using high precision stacked wafers,
Rev. Sci. Instrum. 90, 103203 (2019).
[50] J. P. Home, M. J. McDonnell, D. J. Szwer, B. C. Keitch,
D. M. Lucas, D. N. Stacey, and A. M. Steane, Memory
coherence of a sympathetically cooled trapped-ion qubit,
Phys. Rev. A 79, 050305 (2009).
[51] G. Baggio, F. Ticozzi, P. D. Johnson, and L. Viola,
Dissipative encoding of quantum information, arXiv:2102
.04531.
[52] I. Rojkov, D. Layden, P. Cappellaro, J. Home, and F.
Reiter, Bias in error-corrected quantum sensing, arXiv:
2101.05817.
PHYSICAL REVIEW LETTERS 128, 080503 (2022)
080503-6
[53] D. Gottesman and I. L. Chuang, Demonstrating the viability
of universal quantum computation using teleportation
and single-qubit operations, Nature (London) 402, 390
(1999).
[54] J. R. Maze, A. Gali, E. Togan, Y. Chu, A. Trifonov, E.
Kaxiras, and M. D. Lukin, Properties of nitrogen-vacancy
centers in diamond: The group theoretic approach, New J.
Phys. 13, 025025 (2011).
[55] A. Mitra, M. J. Martin, G. W. Biedermann, A. M. Marino,
P. M. Poggi, and I. H. Deutsch, Robust Mølmer-Sørensen
gate for neutral atoms using rapid adiabatic Rydberg
dressing, Phys. Rev. A 101, 030301(R) (2020).
PHYSICAL REVIEW LETTERS 128, 080503 (2022)
080503-7
... For example, quantum bath engineering [5] has been shown to enable efficient processing with applications in stabilized state preparation[6-9], quantum error correction (QEC) [10][11][12], and dissipationassisted phase transition [13][14][15][16]. Consequently, bath engineering has evolved into a fruitful research field, encompassing applications such as cooling various quantum systems [17][18][19][20] and entanglement stabilization in both trapped ions [6,21,22] and superconducting circuits [23][24][25]. This non-unitary technique for state stabilization is appealing because it directs a driven-dissipative system toward steady states of interest by engineering dissipative channels, making it insensitive to initial states and noise [22]. ...
... For example, quantum bath engineering [5] has been shown to enable efficient processing with applications in stabilized state preparation [6][7][8][9], quantum error correction (QEC) [10][11][12], and dissipationassisted phase transition [13][14][15][16]. Consequently, bath engineering has evolved into a fruitful research field, encompassing applications such as cooling various quantum systems [17][18][19][20] and entanglement stabilization in both trapped ions [6,21,22] and superconducting circuits [23][24][25]. This non-unitary technique for state stabilization is appealing because it directs a driven-dissipative system toward steady states of interest by engineering dissipative channels, making it insensitive to initial states and noise [22]. ...
... Consequently, bath engineering has evolved into a fruitful research field, encompassing applications such as cooling various quantum systems [17][18][19][20] and entanglement stabilization in both trapped ions [6,21,22] and superconducting circuits [23][24][25]. This non-unitary technique for state stabilization is appealing because it directs a driven-dissipative system toward steady states of interest by engineering dissipative channels, making it insensitive to initial states and noise [22]. ...
Preprint
Generation and preservation of quantum entanglement are among the primary tasks in quantum information processing. State stabilization via quantum bath engineering offers a resource-efficient approach to achieve this objective. However, current methods for engineering dissipative channels to stabilize target entangled states often require specialized hardware designs, complicating experimental realization and hindering their compatibility with scalable quantum computation architectures. In this work, we propose and experimentally demonstrate a stabilization protocol readily implementable in the mainstream integrated superconducting quantum circuits. The approach utilizes a Raman process involving a resonant (or nearly resonant) superconducting qubit array and their dedicated readout resonators to effectively emerge nonlocal dissipative channels. Leveraging individual controllability of the qubits and resonators, the protocol stabilizes two-qubit Bell states with a fidelity of 90.7%90.7\%, marking the highest reported value in solid-state platforms to date. Furthermore, by extending this strategy to include three qubits, an entangled W state is achieved with a fidelity of 86.2%86.2\%, which has not been experimentally investigated before. Notably, the protocol is of practical interest since it only utilizes existing hardware common to standard operations in the underlying superconducting circuits, thereby facilitating the exploration of many-body quantum entanglement with dissipative resources.
... Non-unital channels differ fundamentally from unital channels. Non-unital channels can increase the purity of quantum states and can have more interesting fixed points such as entangled states [16,17,23,35,42,44,51,61]. In the context of parameterized quantum circuits unital channels lead to noise-induced barren plateaus while non-unital channels can help avoid them [75]. ...
... The discipline of dissipation engineering, also known as reservoir engineering, leverages the interaction between a quantum system and environmental degrees of freedom to perform quantum information processing tasks. [1,4,16,17,23,32,35,38,42,44,51,61,63,65,66,81]. Dissipation towards the environment lifts the requirement for classical measurement and feedback and has scaling and robustness advantages over unitary approaches [35,42,51]. ...
Preprint
We propose incorporating multi-qubit nonunitary operations in Variational Quantum Thermalizers (VQTs). VQTs are hybrid quantum-classical algorithms that generate the thermal (Gibbs) state of a given Hamiltonian, with applications in quantum algorithms and simulations. However, current algorithms struggle at intermediate temperatures, where the target state is nonpure but exhibits entanglement. We devise multi-qubit nonunitary operations that harness weak symmetries and thereby improve the performance of the algorithm. Utilizing dissipation engineering, we create these nonunitary multi-qubit operations without the need for measurements or additional qubits. To train the ansatz, we develop and benchmark novel methods for entropy estimation of quantum states, expanding the toolbox for quantum state characterization. We demonstrate that our approach can prepare thermal states of paradigmatic spin models at all temperatures. Our work thus creates new opportunities for simulating open quantum many-body systems.
... An alternative approach is to deliver light using optical waveguides directly integrated into the trap structure. This provides compact routing of light to the zones of interest, where grating couplers allow it to be focused directly on the trapped particles [2,[13][14][15][16][17][18]. ...
Article
Full-text available
Multiplexed operations and extended coherent control over multiple trapping sites are fundamental requirements for a trapped-ion processor in a large-scale architecture. Here, we demonstrate these building blocks using a surface-electrode trap with integrated photonic components which are scalable to larger numbers of zones. We implement a Ramsey sequence using the integrated light in two zones, separated by 375 μ m , performing transport of the ion from one zone to the other in 200 μ s between pulses. In order to achieve low motional excitation during transport, we develop techniques to measure and mitigate the effect of the exposed dielectric surfaces used to deliver the integrated light to the ion. We also demonstrate simultaneous control of two ions in separate zones with low optical crosstalk and use this to perform simultaneous spectroscopy to correlate field noise between the two sites. Our work demonstrates the first transport and coherent multizone operations in integrated photonic ion trap systems, forming the basis for further scaling in the trapped-ion quantum charge-coupled device architecture. Published by the American Physical Society 2025
... Dissipation is typically viewed as detrimental to quantum information, but recent advances in experimental control have begun to harness its potential to create well-controlled nontrivial open quantum systems [1][2][3][4][5][6][7]. This raises an intriguing question: can an open quantum system host exotic quantum phases that are inherently mixed and nonthermal [8]? ...
Preprint
Full-text available
Recent experimental progress in controlling open quantum systems enables the pursuit of mixed-state nonequilibrium quantum phases. We investigate whether open quantum systems hosting mixed-state symmetry-protected topological states as steady states retain this property under symmetric perturbations. Focusing on the decohered cluster state -- a mixed-state symmetry-protected topological state protected by a combined strong and weak symmetry -- we construct a parent Lindbladian that hosts it as a steady state. This Lindbladian can be mapped onto exactly solvable reaction-diffusion dynamics, even in the presence of certain perturbations, allowing us to solve the parent Lindbladian in detail and reveal previously-unknown steady states. Using both analytical and numerical methods, we find that typical symmetric perturbations cause strong-to-weak spontaneous symmetry breaking at arbitrarily small perturbations, destabilize the steady-state mixed-state symmetry-protected topological order. However, when perturbations introduce only weak symmetry defects, the steady-state mixed-state symmetry-protected topological order remains stable. Additionally, we construct a quantum channel which replicates the essential physics of the Lindbladian and can be efficiently simulated using only Clifford gates, Pauli measurements, and feedback.
... In the context of control of quantum systems, it is an essential resource for pumping into quantum states or manifolds, however often constitutes a hindrance to quantum coherence. The last decade has seen the development of increasingly intricate methods for engineering dissipation to directly produce quantum states (including entangled states) [1][2][3][4][5][6][7][8][9][10][11], as well as to achieve protected manifolds of states required for quantum error correction [12][13][14][15][16][17][18][19]. These methods involve engineering the coupling to a bath system which removes entropy, and are known as reservoir engineering [20]. ...
Preprint
Full-text available
We introduce a novel reservoir engineering approach for stabilizing multi-component Schr\"odinger's cat manifolds. The fundamental principle of the method lies in the destructive interference at crossings of gain and loss Hamiltonian terms in the coupling of an oscillator to a zero-temperature auxiliary system, which are nonlinear with respect to the oscillator's energy. The nature of these gain and loss terms is found to determine the rotational symmetry, energy distributions, and degeneracy of the resulting stabilized manifolds. Considering these systems as bosonic error-correction codes, we analyze their properties with respect to a variety of errors, including both autonomous and passive error correction, where we find that our formalism gives straightforward insights into the nature of the correction. We give example implementations using the anharmonic laser-ion coupling of a trapped ion outside the Lamb-Dicke regime as well as nonlinear superconducting circuits. Beyond the dissipative stabilization of standard cat manifolds and novel rotation symmetric codes, we demonstrate that our formalism allows for the stabilization of bosonic codes linked to cat states through unitary transformations, such as quadrature-squeezed cats. Our work establishes a design approach for creating and utilizing codes using nonlinearity, providing access to novel quantum states and processes across a range of physical systems.
... Thus, large-scale highly uniform fabrication of multi-layer qubit control structures is relatively straightforward. While integrated photonics requires sub-micron dielectric features [85], their uniformity needs can be considerably relaxed by carrying out dissipative operation in the saturated steady-state regime [86][87][88]. Finally, the qubit coupling inhomogeneities caused by residual chip manufacturing and design imperfections can be shimmed out using local tuning electrodes. ...
Preprint
The central challenge of quantum computing is implementing high-fidelity quantum gates at scale. However, many existing approaches to qubit control suffer from a scale-performance trade-off, impeding progress towards the creation of useful devices. Here, we present a vision for an electronically controlled trapped-ion quantum computer that alleviates this bottleneck. Our architecture utilizes shared current-carrying traces and local tuning electrodes in a microfabricated chip to perform quantum gates with low noise and crosstalk regardless of device size. To verify our approach, we experimentally demonstrate low-noise site-selective single- and two-qubit gates in a seven-zone ion trap that can control up to 10 qubits. We implement electronic single-qubit gates with 99.99916(7)% fidelity, and demonstrate consistent performance with low crosstalk across the device. We also electronically generate two-qubit maximally entangled states with 99.97(1)% fidelity and long-term stable performance over continuous system operation. These state-of-the-art results validate the path to directly scaling these techniques to large-scale quantum computers based on electronically controlled trapped-ion qubits.
... Implementation of Rydberg antiblockade gates is one of the current mainstream protocols towards fast and robust neutral-atom quantum computation [1][2][3][4][5][6], as the traditional Rydberg antiblockade regime with enhanced dissipative dynamics can lead to the preparation of high-fidelity steady entanglement in versatile systems [7][8][9][10][11][12][13][14][15][16][17]. Beyond applications in quantum entanglement, constructing a quantum logic gate via the antiblockade effect [18][19][20] is more straightforward because the simultaneous multiatom excitation establishes an effective model which exactly avoids the occupancy of lossy singly-excited Rydberg states. ...
Article
Position error is treated as the leading obstacle that prevents Rydberg antiblockade gates from being experimentally realizable, because of the inevitable fluctuations in the relative motion between two atoms, invalidating the antiblockade condition. In this work we report progress towards a high-tolerance antiblockade-based Rydberg swap gate enabled by the use of a modified antiblockade condition combined with carefully optimized laser pulses. Depending on the optimization of diverse pulse shapes, our protocol shows that the amount of time spent in the double Rydberg state can be shortened by more than 70% with respect to the case using the perfect antiblockade condition, which significantly reduces this position error. Moreover, we benchmark the robustness of the gate by taking into account the technical noises, such as the Doppler dephasing due to atomic thermal motion, the fluctuations in laser intensity and laser phase, and the intensity inhomogeneity. Compared to other existing antiblockade-gate schemes, we are able to maintain a predicted gate fidelity above 0.91 after a very conservative estimation of various experimental imperfections, especially considered for a realistic interaction deviation of δV/V≈5.92% at T∼20 µK. Our work creates an opportunity for the experimental demonstration of Rydberg antiblockade gates in the near future.
Article
Full-text available
In this paper, we investigate how the evolution of the states of two qubits initially in a direct product state can be controlled by the optical field in a Tavis-Cummings (TC) model. For the two qubits initially in the direct product state, we find that their matrix elements at any moment can be modulated by the coefficients of the optical field initial states in the number state space. We propose a method for preparing an X-type state of two qubits. Subsequently, for descriptive convenience, we divide the Bell states of the two qubits into two kinds in the paper. When both qubits are initially in the ground state, we find that the two qubits can be controlled to produce the first type of Bell state by the superposition state optical field that is initially in the next-nearest-neighbor number state and that the production of any of the first type of Bell states can be controlled by controlling the phase between the two next-nearest-neighbor number states. When one of the two qubits is in the ground state, and the other is in the excited state, we can control the two qubits to produce the second type of Bell state by the single-photon number state optical field. Finally, we study the generation of Werner states by controlling two qubits initially, both in the ground state, using an optical field.
Article
We demonstrate a novel experimental tool set that enables irreversible multiqubit operations on a quantum platform. To exemplify our approach, we realize two elementary nonunitary operations: the or and nor gates. The electronic states of two trapped Ca40+ ions encode the logical information, and a cotrapped Sr88+ ion provides the irreversibility of the gate by a dissipation channel through sideband cooling. We measure 87% and 81% success rates for the or and nor gates, respectively. The presented methods are a stepping stone toward other nonunitary operations such as in quantum error correction and quantum machine learning.
Article
Full-text available
The concept of generalized Gibbs ensembles (GGEs) has been introduced to describe steady states of integrable models. Recent advances show that GGEs can also be stabilized in nearly integrable quantum systems when driven by external fields and open. Here, we present a weakly dissipative dynamics that drives towards a steady-state GGE and is realistic to implement in systems of trapped ions. We outline the engineering of the desired dissipation by a combination of couplings which can be realized with ion-trap setups and discuss the experimental observables needed to detect a deviation from a thermal state. We present a mixed-species motional mode engineering technique in an array of microtraps and demonstrate the possibility to use sympathetic cooling to construct many-body dissipators. Our paper provides a blueprint for experimental observation of GGEs in open systems and opens an avenue for quantum simulation of driven-dissipative quantum many-body problems.
Article
Full-text available
We present protocols for dissipative entanglement of three trapped-ion qubits and discuss a scheme that uses sympathetic cooling as the dissipation mechanism. This scheme relies on tailored destructive interference to generate any one of six entangled W states in a three-ion qubit space. Using a beryllium–magnesium ion crystal as an example system, we theoretically investigate the protocol’s performance and the effects of likely error sources, including thermal secular motion of the ion crystal, calibration imperfections, and spontaneous photon scattering. We estimate that a fidelity of ∼98% may be achieved in typical trapped ion experiments with ∼1 ms interaction time. These protocols avoid timescale hierarchies for faster preparation of entangled states.
Article
Full-text available
To build a universal quantum computer from fragile physical qubits, effective implementation of quantum error correction (QEC)¹ is an essential requirement and a central challenge. Existing demonstrations of QEC are based on an active schedule of error-syndrome measurements and adaptive recovery operations2,3,4,5,6,7 that are hardware intensive and prone to introducing and propagating errors. In principle, QEC can be realized autonomously and continuously by tailoring dissipation within the quantum system1,8,9,10,11,12,13,14, but so far it has remained challenging to achieve the specific form of dissipation required to counter the most prominent errors in a physical platform. Here we encode a logical qubit in Schrödinger cat-like multiphoton states¹⁵ of a superconducting cavity, and demonstrate a corrective dissipation process that stabilizes an error-syndrome operator: the photon number parity. Implemented with continuous-wave control fields only, this passive protocol protects the quantum information by autonomously correcting single-photon-loss errors and boosts the coherence time of the bosonic qubit by over a factor of two. Notably, QEC is realized in a modest hardware setup with neither high-fidelity readout nor fast digital feedback, in contrast to the technological sophistication required for prior QEC demonstrations. Compatible with additional phase-stabilization and fault-tolerant techniques16,17,18, our experiment suggests quantum dissipation engineering as a resource-efficient alternative or supplement to active QEC in future quantum computing architectures.
Article
Full-text available
Practical and useful quantum information processing requires substantial improvements with respect to current systems, both in the error rates of basic operations and in scale. The fundamental qualities of individual trapped-ion¹ qubits are promising for long-term systems², but the optics involved in their precise control are a barrier to scaling³. Planar-fabricated optics integrated within ion-trap devices can make such systems simultaneously more robust and parallelizable, as suggested by previous work with single ions⁴. Here we use scalable optics co-fabricated with a surface-electrode ion trap to achieve high-fidelity multi-ion quantum logic gates, which are often the limiting elements in building up the precise, large-scale entanglement that is essential to quantum computation. Light is efficiently delivered to a trap chip in a cryogenic environment via direct fibre coupling on multiple channels, eliminating the need for beam alignment into vacuum systems and cryostats and lending robustness to vibrations and beam-pointing drifts. This allows us to perform ground-state laser cooling of ion motion and to implement gates generating two-ion entangled states with fidelities greater than 99.3(2) per cent. This work demonstrates hardware that reduces noise and drifts in sensitive quantum logic, and simultaneously offers a route to practical parallelization for high-fidelity quantum processors⁵. Similar devices may also find applications in atom- and ion-based quantum sensing and timekeeping⁶.
Article
Full-text available
Quantum reservoir engineering provides a versatile framework for quantum state preparation and control, with improved robustness to decoherence. However, established methods for dissipative state preparation typically rely on resolving resonances, limiting the target state fidelity due to a competition between the stabilization mechanism and uncontrolled dissipation. We propose a new framework for engineering dissipation that combines the advantages of static dispersive couplings with strong parametric driving and show how it can realize high fidelity and fast entanglement stabilization devoid of such constraints. In addition, the phase sensitivity of parametric couplings allows arbitrary state preparation and continuous control of the stabilized state within a fixed parity manifold. The proposed protocol is readily accessible with the state-of-the-art superconducting qubit technology and holds promise for fast preparation of large entangled resource states.
Article
We demonstrate a simplified method for dissipative generation of an entangled state of two trapped-ion qubits. Our implementation produces its target state faster and with higher fidelity than previous demonstrations of dissipative entanglement generation and eliminates the need for auxiliary ions. The entangled singlet state is generated in ∼7 ms with a fidelity of 0.949(4). The dominant source of infidelity is photon scattering. We discuss this error source and strategies for its mitigation.
Article
The Rydberg blockade mechanism is now routinely considered for entangling qubits encoded in clock states of neutral atoms. Challenges towards implementing entangling gates with high fidelity include errors due to thermal motion of atoms, laser amplitude inhomogeneities, and imperfect Rydberg blockade. We show that adiabatic rapid passage by Rydberg dressing provides a mechanism for implementing two-qubit entangling gates by accumulating phases that are robust to these imperfections. We find that the typical error in implementing a two-qubit gate, such as the controlled phase gate, is dominated by errors in the single-atom light shift, and that this can be easily corrected using adiabatic dressing interleaved with a simple spin echo sequence. This results in a two-qubit Mølmer-Sørensen gate. A gate fidelity ∼0.995 is achievable with modest experimental parameters and a path to higher fidelities is possible for Rydberg states in atoms with a stronger blockade, longer lifetimes, and larger Rabi frequencies.
Article
Entangling operations are among the most important primitive gates employed in quantum computing, and it is crucial to ensure high-fidelity implementations as systems are scaled up. We experimentally realize and characterize a simple scheme to minimize errors in entangling operations related to the residual excitation of mediating bosonic oscillator modes that both improves gate robustness and provides scaling benefits in larger systems. The technique employs discrete phase shifts in the control field driving the gate operation, determined either analytically or numerically, to ensure all modes are de-excited at arbitrary user-defined times. We demonstrate an average gate fidelity of 99.4(2)% across a wide range of parameters in a system of Yb171+ trapped ion qubits, and observe a reduction of gate error in the presence of common experimental error sources. Our approach provides a unified framework to achieve robustness against both static and time-varying laser amplitude and frequency detuning errors. We verify these capabilities through system-identification experiments revealing improvements in error susceptibility achieved in phase-modulated gates.
Article
We describe the use of laser-enhanced etching of fused silica in order to build multilayer ion traps. This technique offers high precision of both machining and alignment of adjacent wafers. As examples of designs taking advantage of this possibility, we describe traps for realizing two key elements of scaling trapped ion systems. The first is a trap for a cavity-QED interface between single ions and photons, in which the fabrication allows shapes that provide good electrostatic shielding of the ion from charge buildup on the mirror surfaces. The second incorporates two X-junctions allowing two-dimensional shuttling of ions. Here, we are able to investigate designs which explore a trade-off between pseudopotential barriers and confinement at the junction center. In both cases, we illustrate the design constraints arising from the fabrication.
Article
Electric-field noise from ion-trap electrode surfaces can limit the fidelity of multiqubit entangling operations in trapped-ion quantum information processors and can give rise to systematic errors in trapped-ion optical clocks. The underlying mechanism for this noise is unknown, but it has been shown that the noise amplitude can be reduced by energetic ion bombardment, or “ion milling,” of the trap electrode surfaces. Using a single trapped Sr+88 ion as a sensor, we investigate the temperature dependence of this noise both before and after ex situ ion milling of the trap electrodes. Making measurements over a trap electrode temperature range of 4 K to 295 K in both sputtered niobium and electroplated gold traps, we see a marked change in the temperature scaling of the electric-field noise after ion milling: power-law behavior in untreated surfaces is transformed to Arrhenius behavior after treatment. The temperature scaling becomes material-dependent after treatment as well, strongly suggesting that different noise mechanisms are at work before and after ion milling. To constrain potential noise mechanisms, we measure the frequency dependence of the electric-field noise, as well as its dependence on ion-electrode distance, for niobium traps at room temperature both before and after ion milling. These scalings are unchanged by ion milling.