ArticlePDF Available

Light-matter entanglement over 50 km of optical fibre

Authors:

Abstract and Figures

When shared between remote locations, entanglement opens up fundamentally new capabilities for science and technology. Envisioned quantum networks use light to distribute entanglement between their remote matter-based quantum nodes. Here we report on the observation of entanglement between matter (a trapped ion) and light (a photon) over 50 km of optical fibre: two orders of magnitude further than the state of the art and a practical distance to start building large-scale quantum networks. Our methods include an efficient source of ion–photon entanglement via cavity-QED techniques (0.5 probability on-demand fibre-coupled photon from the ion) and a single photon entanglement-preserving quantum frequency converter to the 1550 nm telecom C band (0.25 device efficiency). Modestly optimising and duplicating our system would already allow for 100 km-spaced ion–ion heralded entanglement at rates of over 1 Hz. We show therefore a direct path to entangling 100 km-spaced registers of quantum-logic capable trapped-ion qubits, and the optical atomic clock transitions that they contain.
Content may be subject to copyright.
ARTICLE OPEN
Light-matter entanglement over 50 km of optical bre
V. Krutyanskiy
1
, M. Meraner
1,2
, J. Schupp
1,2
, V. Krcmarsky
1,2
, H. Hainzer
1,2
and B. P. Lanyon
1,2
When shared between remote locations, entanglement opens up fundamentally new capabilities for science and technology.
Envisioned quantum networks use light to distribute entanglement between their remote matter-based quantum nodes. Here we
report on the observation of entanglement between matter (a trapped ion) and light (a photon) over 50 km of optical bre: two
orders of magnitude further than the state of the art and a practical distance to start building large-scale quantum networks. Our
methods include an efcient source of ionphoton entanglement via cavity-QED techniques (0.5 probability on-demand bre-
coupled photon from the ion) and a single photon entanglement-preserving quantum frequency converter to the 1550 nm telecom
C band (0.25 device efciency). Modestly optimising and duplicating our system would already allow for 100 km-spaced ionion
heralded entanglement at rates of over 1 Hz. We show therefore a direct path to entangling 100 km-spaced registers of quantum-
logic capable trapped-ion qubits, and the optical atomic clock transitions that they contain.
npj Quantum Information (2019) 5:72 ; https://doi.org/10.1038/s41534-019-0186-3
INTRODUCTION
Envisioned quantum networks
1
consist of distributed matter-
based quantum nodes, for the storage, manipulation and
application of quantum information, which are interconnected
with photonic links to establish entanglement between the nodes.
While the most ambitious form of a quantum network is a
collection of remote quantum computers, far simpler networks
with a handful of qubits at each node could already enable
powerful applications in quantum enhanced distributed sensing,
timekeeping, cryptography and multiparty protocols.
2
Entanglement has been achieved between two atoms in traps a
few 10 m apart,
3
between two ions in traps a few metres apart
4
and recently between two nitrogen-vacancy centres 1.3 km apart.
5
In these experiments, photon-matter entanglement is rst
generated, then detection of one or two photons heralds remote
matter-matter entanglement (entanglement is swappedfrom
matter-light to matter-matter). A current goal is to signicantly
scale up the distance over which quantum matter can be
entangled to a 100 km or more, which are practical internode
spacings to enable large-scale quantum networks.
Some key challenges to entangling matter over such distances are
now described. First, the aforementioned matter systems emit
photons at wavelengths that are strongly absorbed in optical
waveguides (such as optical bre), limiting the internode distance to
a few kilometres. For example, in the present work 854 nm photons
are collected from a trapped atomic ion. While the ~3 dB per km
losses suffered by 854nm photons through state-of-the-art optical
bre allows for few kilometre internode distances, transmission over
50 km of bre would be 10
15
. Single-photon quantum frequency
conversion to the telecom C band (1550nm) would offer a powerful
solution: this wavelength suffers the minimum bre transmission
losses (~0.18 dB per km, yielding 10% transmission over 50 km) and
is therefore an ideal choice for a standard interfacing wavelength for
quantum networking. Photons from solid-state memories,
6
cold gas
memories,
7,8
quantum dots and nitrogen-vacancy centres
9
have
been converted to telecom wavelengths. Frequency conversion of
photons from ions has very recently been performed, including to
the telecom C band (without entanglement),
10
to the telecom O
band with entanglement over 80 m
11
and directly to an atomic
Rubidium line at 780 nm.
12
The use of photon conversion to extend
the distance over which light-matter and matter-matter entangle-
ment can be distributed has not previously been achieved.
A second challenge to long distance matter entanglement is to
preserve entanglement when such long photonic channels are
involved. Uncontrolled decoherence processes that act on the photon
as it travels along its path, and those that act on the quantum matter
during the photon travel time, can easily destroy entanglement. For
example, the entanglement-carrying photon signal, which attenuates
exponentially with distance in any lossy waveguide, can be
overwhelmed by added photon noise from the photon frequency
conversion process or dark counts of the photon detectors. The inter-
node photon travel time also imposes a minimum coherence time for
matter, which for e.g. 50 km of optical bre is already signicant at
250 μs (and 500 μs to allow for the classical signal of a successful
herald to return). Moreover, quantum networking applications require
distributed entanglement of a quality above certain thresholds, for
which the required matter coherence times and photon signal to
noise ratio are far more challenging.
A third challenge comes again from the photon travel time. The
shortest time required to entangle remote matter (or indeed light)
in two places is the light travel time between them. The 500 μs
wait time over 50 km of optical bre yields a maximum attempt
rate of only 2 kHz: one must wait 500 μs to learn if an individual
attempt to distribute remote entanglement has been successful.
To achieve practical entanglement distribution rates in the face of
such a restriction, one can work on achieving a high probability for
individual attempts to succeed and (or) to run many attempts in
parallel (as discussed later).
Received: 29 March 2019 Accepted: 5 August 2019
1
Institut für Quantenoptik und Quanteninformation, Österreichische Akademie der Wissenschaften, Technikerstr. 21A, 6020 Innsbruck, Austria and
2
Institut für
Experimentalphysik, Universität Innsbruck, Technikerstr. 25, 6020 Innsbruck, Austria
Correspondence: B. P. Lanyon (ben.lanyon@uibk.ac.at)
These authors contributed equally: V. Krutyanskiy, M. Meraner, J. Schupp
www.nature.com/npjqi
Published in partnership with The University of New South Wales
In this work, entanglement between a trapped-ion qubit and a
photon that has travelled over 50 km of optical bre is achieved.
The quality of the entanglement is sufciently high to allow for a
clear violation of a Bell inequalityas required for some of the
most challenging device-independent quantum network applica-
tions.
13
Furthermore, when modestly optimised, the achieved rate
is expected to allow for entanglement distribution between
100 km-spaced trapped ions at rates over 1 Hz. The paper is
organised as follows. First, there is a short motivation for quantum
networking trapped ions. Second, a brief overview of the
experimental methods is given, with much detail left for the
Supplementary Material. Third, the tomographically reconstructed
entangled state, of the ion qubit and photon polarisation qubit
after 50 km, is presented and the achieved delity, efciency and
rate are analysed. Fourth, the ion qubit is shown to provide a
quantum information storage time (coherence time) of more than
20 ms, allowing for future entanglement distribution over
thousands of kilometers. Finally, the prospects for 100 km ionion
entanglement are presented as well as a path to signicantly
increase the rate via multi-mode and hybrid quantum networking.
Trapped ions are particularly powerful systems to enable
quantum networking and the envisioned applications. For
example, a complete set of tools for deterministic universal
manipulation of quantum information encoded into registers of
trapped ions is readily available and of a quality near fault tolerant
thresholds,
1416
as required for arbitrary distance quantum
networking via the quantum repeater approach.
17,18
Key quantum
networking functionalities have been demonstrated between ions
over a few meters, including remote state teleportation
19
and
multi-ion protocols.
20
Trapped ions are also some of the most
sensitive measurement probes yet developed. For example, many
ion species, including the one used in this work, contain optical
atomic clock transitions and therefore entangling them over
distance enables the ideas presented in
21,22
to be explored.
RESULTS
Our network node consists of a
40
Ca
+
ion in a radio-frequency
linear Paul trap with an optical cavity that enhances photon
collection on the 854 nm electronic dipole transition (Fig. 1). A
Raman laser pulse at 393 nm triggers emission, by the ion, of a
photon into the cavity via a bichromatic cavity-mediated
Raman transition (CMRT).
23
Two indistinguishable processes are
driven in the CMRT, each leading to the generation of a
cavity photon and resulting in entanglement between photon
polarisation and the electronic qubit state of the ion of the form
1=
ffiffi
2
pðDJ¼5=2;mj¼5=2;VþDJ¼5=2;mj¼3=2;HÞ, with horizontal (H)and
vertical (V) photon polarisation and two metastable Zeeman states
of the ion ðDJ;mjÞ, see Supplementary Fig. 3. The total measured
probability of obtaining an on-demand free-space photon out of
the ion vacuum chamber (entangled with the ion) is P
out
=0.5 ±
0.1 (Secction II, Supplementary material of this paper), enabled by
the novel low-loss cavity in our setup.
The CMRT yields an entangled state with a frequency-
degenerate photon qubit (the two polarisation components have
the same frequency to within the cavity linewidth
23
), providing a
signicant benet for long distance networking: the phase of the
light-matter entangled state does not depend on the time at
which the photon detection event occurs at a given distance from
the ion. Photon detection time uctuates due to the intrinsic nite
temporal extent of the photon wavepacket and in the case of
optical path length changes, which could be signicant over tens
of kilometres of deployed optical bre. Our photons are generated
over several tens of microseconds, with a corresponding
bandwidth of tens of kilohertz. This unusually narrow bandwidth
allows for strong frequency ltering, which we exploit in the
photon conversion process and could have further benets in
future deployed networks, e.g to enable co-propagating classical
and quantum light. Furthermore, the corresponding photon
coherence-length is potentially thousands of metres, allowing
for essentially path-length-insensitive entanglement swapping
between remote ions via Hong-Ou-Mandel interference.
4,24,25
Single-mode bre-coupled photons from the ion are injected
into a polarisation-preserving photon conversion system (pre-
viously characterised using classical light
26
). In summary, a χ
(2)
optical nonlinearity is used to realise difference frequency
generation, whereby the energy of the 854 nm photon is reduced
by that of a pump-laser photon at 1902 nm, yielding 1550 nm.
Two commercially available free-space and crossed PPLN ridge
waveguide crystals are used, one to convert each polarisation, in a
self-stable polarisation interferometer. The total bre-coupled
device conversion efciency here is 25 ± 0.02%, for an added
white noise of 40 photons per second, within the ltering
bandwidth of 250 MHz centred at 1550 nm. As discussed in,
26
the
854 nm line in
40
Ca
+
is almost unique amongst trapped-ion
transitions in its potential for low-noise, highly-efcient single-step
frequency conversion to the telecom C band.
Following conversion, the telecom photon is injected into a
50.47 km SMF28single-mode bre spool with 0.181 dB per km
loss (10.4 ± 0.5% measured total transmission probability). The
spool is not actively stabilised. Polarisation dynamics in an
unspooled bre could be actively controlled using methods
developed in the eld of quantum cryptography (e.g.,
27
). Finally,
free-space projective polarisation analysis is performed and the
photon is detected using a telecom solid-state photon detector
with an efciency of 0.10 ± 0.01 and free-running dark count rate
Fig. 1 Simplied experiment schematic. From left to right: a single atomic ion (red sphere) in the centre of a radio-frequency linear Paul trap
(gold electrodes) and a vacuum anti-node of an optical cavity. A Raman laser pulse triggers emission of an 854 nm photon into the cavity,
which exits to the right. The photon, polarisation-entangled with two electronic qubit states of the ion (two Zeeman states of the D
J
=5/2
manifold, not shown), is then wavelength-converted to 1550 nm using difference frequency generation involving ridge-waveguide-integrated
periodically-poled lithium niobate (PPLN) chips and a strong (~1 W ) pump laser at 1902 nm.
26
The photon then passes through a 50 km single-
mode bre spool, is ltered with a 250 MHz bandwidth etalon (SF) to reduce noise from the conversion stage,
26
and is polarisation-analysed
using waveplates, a polarising beam splitter (PBS) and a solid-state single photon counting module (SPCM, InGaAs ID230 from IDQuantique).
The electronic state of the ion is measured (not shown), conditional on the detection of a photon. Additional photon conversion lters are not
shown.
26
For further details see, Supplementary material of this paper section I
V. Krutyanskiy et al.
2
npj Quantum Information (2019) 72 Published in partnership with The University of New South Wales
1234567890():,;
of 2 counts per second (cps). Measurement of the ion-qubit state
is performed conditional on the detection of a 50 km photon
within a 30 μs time window: the Zeeman ion qubit is mapped into
the established
40
Ca
+
optical quadrupole clock qubit
28
via laser
pulses at 729 nm, followed by standard uorescence state
detection (see Methods).
Quantum state tomography is performed to reconstruct the two-
qubit (ion qubit and photon polarisation qubit) state, Supplemen-
tary material of this paper section III. The 247 μs photon travel time
through the bre limits the maximum attempt rate for generating a
photon from the ion to 4 kHz (2 kHz if the bre was stretched out
away from our ion to force an additional delay for the classical
signal photon clickto return). Here, until photon detection occurs,
photon generation is (Raman laser pulses are) performed every
453 μs, yielding an attempt rate of 2.2 kHz. For the complete
experimental sequence see Methods. All error bars on quantities
derived from the tomographically-reconstructed states (density
matrices) are based on simulated uncertainties due to nite
measurement statistics (see Supplementary Material section III).
A strongly entangled ionphoton state is observed (Fig. 2)over
50 km, quantied by a concurrence
29
C=0.75 ± 0.05 and state
delity F
m
=0.86 ± 0.03 with a maximally entangled state (C=1).
Simulating a CHSH Bell inequality test
30
on our tomographic data
yields a value of 2.304 ± 0.125, thereby exceeding the classical bound
(of 2) by 2.4 standard deviations. Using a shorter detection window
(rst 2/3 of the full photon wavepacket) increases the signal to noise
ratio and yields F
m
=0.90 ± 0.03 and CHSH Bell inequality violation
by 4.8 standard deviations at the expense of an efciency decrease
of only 10%. The quality of our light-matter entangled state therefore
surpasses this stringent threshold for its subsequent application.
For a detailed analysis of the sources of indelity in the
entangled state see, Supplementary material of this paper section
IV; here now is a short summary. In a second experiment, the
telecom entangled state is reconstructed right after the conversion
stage (without the 50 km spool), yielding F
m
=0.92 ± 0.02. The
drop in delity when adding the 50 km spool can, to within
statistical uncertainty, be entirely explained by our telecom photon
detector dark counts (2cps). In a third experiment, the 854 nm
entangled state is reconstructed right out of the vacuum chamber
(without conversion), yielding F
m
=0.967 ± 0.006. The observed
drop in delity through the conversion stage alone is dominated
by a drop in photon signal to noise signal. Here the noise consists
of comparable rates of telecom detector dark counts and
conversion noise (caused by Anti-Stokes Raman scattering of the
pump laser
26
) and the signal is reduced by the nite conversion
setup efciency and the lower telecom detectorsefciency
compared to the 854 nm one. The indelity in the 854 nm
photon-ion entangled state is consistent with that achieved in.
23
The total probability that a Raman pulse led to the detection of
a photon after 50 km was P=5.3 × 10
4
, which given an attempt
rate of 2.2 kHz yielded a click rate of 1 cps. Photon loss
mechanisms in our experiment are discussed in, Supplementary
material of this paper section II. In summary, the 50 km bre
transmission (0.1) and our current telecom detector efciency (0.1)
limit the maximum click probability to P=0.01. The majority of
other losses are in passive optical elements, and could largely be
eliminated by e.g. more careful attention to coupling into optical
bres and photon conversion waveguides. In combination with
state-of-the-art telecom detectors (e.g. Scontel recently supplied
us superconducting nanowire detectors providing 0.8 cps dark
count rate and 77% efciency according to the companys
calibration), a total 50 km efciency of P0.01 would be expected
and a corresponding click rate of 20 cps.
One of the functions played by matter in a quantum network is
as a memory to store established entanglement, while entangle-
ment is being made or processed in other parts of the network.
Decoherence processes in the matter qubit will limit the distance
over which it is possible to distribute quantum entanglement (the
distance a photon could possibly travel in the coherence timeof
the matter qubit). In our 50 km experiment, the ion qubit is
already stored for the 250 μs photon travel time through the
50 km bre, with no statistically signicant reduction in the
ionphoton entanglement quality (this was achieved by installing
a mu-metal shield around the ion-trap vacuum chamber to
attenuate ambient magnetic eld uctuations).
Additional tomographic measurements are performed to see for
how long ionphoton entanglement could be stored in our ion-
trap network node before decoherence in the ion-qubit would
destroy it. Specically, state tomography is performed for increas-
ing delays introduced between measurements of the telecom
photon polarisation state (0 km bre travel distance) and measure-
ments of the state of the ion-qubit. This is equivalent to introducing
an additional storage time for the ion-qubit. The results show that
strong entanglement is still present after 20 ms wait time (F
m
=
0.77 ± 0.04, C=0.57 ± 0.08), the longest wait time employed. This
already opens up the possibility of distributing entanglement over
several thousands of kilometers and the time to perform hundreds
of single and multi-qubit ion quantum logic gates.
31
A dominant source of decoherence of our ion-qubit are
uncontrolled uctuating energy-level shifts due to intensity
uctuations of the 806 nm laser eld used to lock the cavity
around the ion. Further attention to minimising the absolute size
of these uctuations should lead to entanglement storage times
of more than 100 ms and therefore the possibility to distribute
entanglement to the other side of the earth. Beyond this, the ion-
qubit could be transferred to hyperne clock transitions within
different co-trapped ion species that offer coherence times of
many seconds and longer.
32
DISCUSSION
The rates for future 100 km-spaced photon-detection heralded
ionion entanglement using our methods are now discussed (see
Fig. 3). A modestly optimised version of our experimental system
in this work is now considered (see dashed box in Fig. 3), that
achieves an on-demand detected 50 km photon click probability
of P=0.01 and operates at an attempt rate of R=2 kHz (the two-
way light travel time). By duplicating our optimised system (Fig. 3),
and following a two-photon click heralding scheme,
24
the
probability of heralding a 100 km-spaced ionion entangled state
would be H2¼1
2P2¼5´105, at an average click rate of H
2
×R=
0.1 cps (comparable with the rst rates achieved over a few
meters
33
of 0.03 cps. Following instead a one-photon click
Time (μs)
260 280 300 320 340 360 380
0
0.3
Photon click
probability (x10-4)
Fig. 2 Observation of ion-photon entanglement over 50 km of
optical bre. (i) 2D red bar chart: histogram of photon detection
times (photon wavepacket in dashed box), following the generation
of an 854 nm photon with a 30 μs Raman laser pulse (R) 250 μs
earlier, repeated at 2.2 kHz. Ionphoton state tomography is
performed for photon detection events recorded in the dashed
box (total contained probability P=5.3 × 10
4
). (ii) 3D bar chart:
absolute value of experimentally-reconstructed density matrix of the
telecom photonic polarisation qubit (Hand Vare Horizontal and
Vertical, respectively) and ion-qubit state (0¼DJ¼5=2;mj¼3=2,
1¼DJ¼5=2;mj¼5=2)
V. Krutyanskiy et al.
3
Published in partnership with The University of New South Wales npj Quantum Information (2019) 72
heralding scheme,
24
the probability of heralding a 100 km-spaced
ionion entangled state would be H
1
=2P× 0.1 =0.002, at an
average click rate of H
1
×R=4 cps, where 0.1 is the reduced
photon generation probability at each node (as required for this
scheme). The factor 40 improvement (H
1
/H
2
) of the one-photon
scheme over the two-photon scheme comes at the expense of the
need to interferometrically stabilise the optical path length across
the 100 km network.
An approach to signicantly increase the remote entanglement
heralding rate is multi-mode quantum networking, where many
photons are sent, each entangled with different matter qubits. In
this way, of running many such processes in parallel, the
probability of at least one successful heralding event occurring
can be made arbitrarily high. In our setup, for example, multiple
ions can be trapped and it may be possible to produce a train of
photons, each entangled with a different ion. In this case, a higher
rate of photon production can be employed, as the time between
photons in the train is not limited by the light travel time.
Furthermore, multi-mode networking could be realised using
inhomogenously broadened ensemble based solid-state quantum
memories.
34
Such memories could be quantum-networked with
ions via a photon conversion interface
35
to form a powerful hybrid
system for long distance quantum networking.
The 50 km photon in our experiments is entangled with the
729 nm optical-qubit clock transition in
40
Ca
+
, over which a
fractional frequency uncertainty of 1 × 10
15
has been achieved
(comparable with the Cs standard).
36
Furthermore,
40
Ca
+
can be
co-trapped with Al
+
,
37
which contains a clock transition for which
a fractional systematic frequency uncertainty at the 1 × 10
18
level
was recently achieved.
38,39
Transfer of the remote
40
Ca
+
entanglement to a co-trapped Al
+
ion could be done via quantum
logic techniques.
39,40
As such, our work provides a direct path to
realise entangled networks of state-of-the-art atomic clocks over
large distances.
21
Entangling clocks provides a way to perform
more sensitive measurements of their average ticking frequen-
cies
21
and to overcome current limits to their synchronisation.
22
METHODS
Trapped ion node
We use a 3D radio-frequency linear Paul trap with a DC endcap to ion
separation of 2.5 mm and ion to blade distance of 0.8 mm. The trap
electrodes are made of titanium, coated with gold and are mounted on
Sapphire holders. The trap drive frequency is 23.4 MHz. The radial secular
frequencies are ω
x
ω
y
=2π× 2.0 MHz, split by approximately 10 kHz and
the axial frequency is ω
z
=2π× 0.927 MHz. Atoms are loaded from a
resistively heated atomic oven and ionised via a two photon process
involving 375 and 422 nm laser light.
The optical cavity around the ion is near-concentric with a length l=
19.9057 ± 0.0003 mm and radii of curvature ROC =9.9841 ± 0.0007 mm,
determined from simultaneous measurements of the free spectral range
(FSR) and higher-order TEM mode spacing (assuming identical mirror
geometries).
41
From this we calculate an expected cavity waist of ω
0
=
12.31 ± 0.07 μm and a maximum ion-cavity coupling rate of g
max
=2π×
1.53 ± 0.01 MHz. The nesse of the cavity (at 854 nm) is
2π
L¼54000 ± 1000, with the total cavity losses T1þT2þL1þ2¼
116 ± 2 ppm, determined from measurements of the cavity ringdown time.
This gives the cavity linewidth 2κ=2π× 140 ± 3 kHz, κbeing the half-
width at half maximum. Taking into account the spontaneous scattering
rate of the P
3/2
state of the ion (γ=2π× 11.45 MHz, half-width) the
expected cooperativity is C¼g2
max
2κγ ¼1:47 ± 0:03.
The transmission T
1,2
of our cavity mirrors was measured to be T
1
=
2.2 ± 0.3 ppm, T
2
=97 ± 4 ppm, that yields expected probability of
extracting a photon from the cavity of Pmax
out ¼T2=ðT1þT2þL1þ2Þ¼
0:83 ± 0:03 (polishing of the mirror substrates done by Perkins Precision
Development, Boulder (Colorado), coating done by Advanced Thin Films).
The optical cavity axis is close to perpendicular to the principle ion trap axis
(~5° difference). A magnetic eld of 4.22 G is applied perpendicular to the
cavity axis and at an angle of 45° to the principal ion trap axis (Supplementary
Fig. 1). The Raman photon generation beam is circularly polarised and parallel
to the magnetic eld (to maximise the coupling on the relevant dipole
transition SJ¼1=2;mj¼1=2$PJ¼3=2;mj¼3=2, see Supplementary Fig. 3).
Pulse sequence for 50 km experiment
First, a 30 μsinitialisationlaser pulse at 393 nm is applied, measured by a
photodiode in transmission of the ion-trap chamber, which allows for
intensity stabilisation of the subsequent 393 nm photon generation Raman
pulse via a sample and hold system. The initialisation pulse is followed by a
1500 μs Doppler cooling pulse. Next, a loop starts in which single photons
are generated (see Supplementary Fig. 2). This loop consists of an additional
Doppler cooling pulse (50 μs), optical pumping to the S¼SJ¼1=2;mj¼1=2
state via circularly polarised 397 nm sigmalaser light (60 μs), and a 393 nm
photon generation Raman pulse (30 μs). This is followed by a wait time for
the photon to travel through the 50 km bre and a subsequent photon
detection window. This sequence loops until a photon is detected.
In the case of a photon detection (detector click), the state of the ion is
measured. To perform an ion state measurement, the DDJ¼5=2;mj¼5=2
electron population is rst mapped to the S¼SJ¼1=2;mj¼1=2state via a
729 nm πpulse (Supplementary Figs. 2 and 3). That is, the D-manifold qubit
is mapped into an optical qubit (with logical states S¼SJ¼1=2;mj¼1=2and
D¼DJ¼5=2;mj¼3=2). In order to measure which of these states the electron
is in, the standard electron shelving technique is used. We perform this
measurement for a detection time(397 nm photon collection time) of
1500 μs, which is sufcient to distinguish bright (scattering) and dark (non-
scattering) ions with an error of less than 1%. The aforementioned process
implements a projective measurement into the eigenstates of the σ
z
basis
(Pauli spin-1/2 operator).
To perform measurements in other bases e.g σ
x
(σ
y
), as required for full
quantum state tomography, an additional π/2 pulse on the Smj¼1=2to
Dmj¼3=2with a 0 (π/2) phase is applied after the πpulse and before the
397 nm pulse, to rotate the ion-qubit measurement basis. The scheme of
the experimental sequence is given in Supplementary Fig. 2.
State characterisation
To reconstruct the ionphoton state, a full state tomography of the two-
qubit system is performed. On the photon polarisation qubit side, the state
is projected to one of 6 states (horizontal, vertical, diagonal, anti-diagonal,
right circular and left circular) by waveplates and a polariser. This is
equivalent to performing projective measurements in three bases
described by the Pauli spin-1/2 operators. For example, horizontal and
vertical are the eigenstates of the Pauli σ
z
operator. On the ion qubit side,
measurement is performed in the three Pauli bases as described in the
previous section. From these measurementsoutcomes probabilities we
reconstruct the 2-qubit state density matrix by linear search with
subsequent Maximum Likelihood method.
42
The values of delity,
concurrence and other measures presented in the Results section are
calculated using reconstructed density matrices for each of the experi-
ments. The error bars for all quantities provided in the Results section
A1
D1 D2
QFC
B1B2
A2
QFC
QFC
854 nm 1550 nm
50 km
BS
100 km
This work
Q. Node Q. Node
QFC
Fig. 3 Path to 100 km matter-matter entanglement. This work:
quantum frequency conversion (QFC) converts a photon, emitted
on-demand from and entangled with an ion qubit (A
1
) in node A, to
the telecom C band at 1550 nm. The photon then travels through
50 km of optical bre before detection (D1 or D2). Future work:
duplicating the system, interfering the two photonic channels on a
beamsplitter (BS). Single or two photon detection heralds the
projection of ions A
1
and B
1
into an entangled state.
24
Deterministic
intra-node quantum logic and measurement between e.g. B
1
and B
2
and A
1
and A
2
can swap the entanglement over larger distances
(quantum repeater). Additional qubits in nodes are available for
entanglement purication. Nodes could as well contain solid-state
memories,
35
NV centres
9
or neutral atoms.
7,8
V. Krutyanskiy et al.
4
npj Quantum Information (2019) 72 Published in partnership with The University of New South Wales
represent one standard deviation of distribution of these quantities over
randomised set of data following the Monte-Carlo approach, for more
details see Supplementary material of this paper section III.
DATA AVAILABILITY
The data and code that support the ndings of this study are available from the
corresponding author upon reasonable request.
ACKNOWLEDGEMENTS
We thank the staff at IQOQI Innsbruck; Rainer Blatt for providing encouragement,
laboratory space and the environment and group support in which to develop our work;
Daniel Heinrich, Klemens Schüppert, Tiffany Brydges, Christine Maier and Tracy Northup
for their support. This work was supported by the START prize of the Austrian FWF project
Y 849-N20, the Army Research Laboratory Center for Distributed Quantum Information via
the project SciNet under Cooperative Agreement Number W911NF-15-2-0060, the
Institute for Quantum Optics and Quantum Information (IQOQI) of the Austrian Academy
Of Sciences (OEAW) and the European Unions Horizon 2020 research and innovation
programme under grant agreement No 820445 and project name Quantum Internet
Alliance. The European Commission is not responsible for any use that may be made of
the information this paper contains.
AUTHOR CONTRIBUTIONS
All authors contributed to the design, development and characterisation of the
experimental systems. In particular, J.S. focused on the ion trap and optical cavity,
M.M. on the photon conversion system, V. Krc. on the ion trap, H.H. on laser
frequency stabilisation and V. Kru. and B.P.L. on all aspects. Experimental data taking
was done by V. Kru., V. Krc., M.M. and J.S. Data analysis and interpretation was done
by V. Kru., J.S., M.M. and B.P.L. All authors contributed to the paper writing. The
project was conceived and supervised by B.P.L.
ADDITIONAL INFORMATION
Supplementary information accompanies the paper on the npj Quantum
Information website (https://doi.org/10.1038/s41534-019-0186-3).
Competing interests: The authors declare no competing interests.
Publishers note: Springer Nature remains neutral with regard to jurisdictional claims
in published maps and institutional afliations.
REFERENCES
1. Kimble, H. J. The quantum internet. Nature 453, 1023 (2008).
2. Wehner, S., Elkouss, D. & Hanson, R. Quantum internet: a vision for the road
ahead. Science 362, 6412 (2018).
3. Ritter, S. et al. An elementary quantum network of single atoms in optical cavities.
Nature 484, 195 (2012).
4. Moehring, D. L. et al. Entanglement of single-atom quantum bits at a distance.
Nature 449, 68 (2007).
5. Hensen, B. et al. Loophole-free bell inequality violation using electron spins
separated by 1.3 kilometres. Nature 526, 682 (2015).
6. Maring, N. et al. Photonic quantum state transfer between a cold atomic gas and
a crystal. Nature 551, 485 (2017).
7. Radnaev, A. G. et al. A quantum memory with telecom-wavelength conversion.
Nat. Phys. 6, 894 EP (2010).
8. Albrecht, B., Farrera, P., Fernandez-Gonzalvo, X., Cristiani, M. & de Riedmatten, H.
A waveguide frequency converter connecting rubidium-based quantum mem-
ories to the telecom c-band. Nat. Commun. 5, 3376 (2014).
9. Dréau, A., Tchebotareva, A., Mahdaoui, A. E., Bonato, C. & Hanson, R. Quantum
frequency conversion of single photons from a nitrogen-vacancy center in dia-
mond to telecommunication wavelengths. Phys. Rev. Appl. 9, 064031 (2018).
10. Walker, T. et al. Long-distance single photon transmission from a trapped ion via
quantum frequency conversion. Phys. Rev. Lett. 120, 203601 (2018).
11. Bock, M. et al. High-delity entanglement between a trapped ion and a telecom
photon via quantum frequency conversion. Nat. Commun. 9, 1998 (2018).
12. Siverns, J. D., Hannegan, J. & Quraishi, Q. Neutral atom wavelength compatible
780 nm single photons from a trapped ion via quantum frequency conversion.
Preprint arXiv:1801.01193 (2018).
13. Lo, H.-K., Curty, M. & Qi, B. Measurement-device-independent quantum key dis-
tribution. Phys. Rev. Lett. 108, 130503 (2012).
14. Schindler, P. et al. Experimental repetitive quantum error correction. Science 332,
1059 (2011).
15. Linke, N. M. et al. Fault-tolerant quantum error detection. Sci. Adv. 3, 10 (2017).
16. Benhelm, J., Kirchmair, G., Roos, C. F. & Blatt, R. Towards fault-tolerant quantum
computing with trapped ions. Nat. Phys. 4, 463 (2008).
17. Briegel, H.-J., Dür, W., Cirac, J. I. & Zoller, P. Quantum repeaters: the role of
imperfect local operations in quantum communication. Phys. Rev. Lett. 81, 5932
(1998).
18. Sangouard, N., Dubessy, R. & Simon, C. Quantum repeaters based on single
trapped ions. Phys. Rev. A 79, 042340 (2009).
19. Olmschenk, S. et al. Quantum teleportation between distant matter qubits. Sci-
ence 323, 486 (2009).
20. Hucul, D. et al. Modular entanglement of atomic qubits using photons and
phonons. Nat. Phys. 11, 37 (2014).
21. Kómár, P. et al. A quantum network of clocks. Nat. Phys. 10, 582 (2014).
22. Ilo-Okeke, E. O., Tessler, L., Dowling, J. P. & Byrnes, T. Remote quantum clock
synchronization without synchronized clocks. npj Quantum Inf. 4, 40 (2018).
23. Stute, A. et al. Tunable ion-photon entanglement in an optical cavity. Nature 485,
482 (2012).
24. Luo, L. et al. Protocols and techniques for a scalable atom-photon quantum
network. Fortschr. Phys. 57, 1133 (2009).
25. Hong, C. K., Ou, Z. Y. & Mandel, L. Measurement of subpicosecond time intervals
between two photons by interference. Phys. Rev. Lett. 59, 2044 (1987).
26. Krutyanskiy, V., Meraner, M., Schupp, J. & Lanyon, B. P. Polarisation-preserving
photon frequency conversion from a trapped-ion-compatible wavelength to the
telecom c-band. Appl. Phys. B 123, 228 (2017).
27. Treiber, A. et al. A fully automated entanglement-based quantum cryptography
system for telecom ber networks. New J. Phys. 11, 045013 (2009).
28. Schindler, P. et al. A quantum information processor with trapped ions. New J.
Phys. 15, 123012 (2013).
29. Wootters, W. K. Entanglement of formation of an arbitrary state of two qubits.
Phys. Rev. Lett. 80, 2245 (1998).
30. Clauser, J. F., Horne, M. A., Shimony, A. & Holt, R. A. Proposed experiment to test
local hidden-variable theories. Phys. Rev. Lett. 23, 880 (1969).
31. Lanyon, B. P. et al. Universal digital quantum simulation with trapped ions. Sci-
ence 334, 57 (2011).
32. Wang, Y. et al. Single-qubit quantum memory exceeding ten-minute coherence
time. Nat. Photonics 11, 646 (2017).
33. Matsukevich, D. N., Maunz, P., Moehring, D. L., Olmschenk, S. & Monroe, C. Bell
inequality violation with two remote atomic qubits. Phys. Rev. Lett. 100, 150404
(2008).
34. Simon, C. et al. Quantum repeaters with photon pair sources and multimode
memories. Phys. Rev. Lett. 98, 190503 (2007).
35. Seri, A. et al. Quantum correlations between single telecom photons and a
multimode on-demand solid-state quantum memory. Phys. Rev. X 7, 021028
(2017).
36. Chwalla, M. et al. Absolute frequency measurement of the
40
Ca
+
4s
2
s
1/2
3d
2
d
5/2
clock transition. Phys. Rev. Lett. 102, 023002 (2009).
37. Guggemos, M., Heinrich, D., Herrera-Sancho, O. A., Blatt, R. & Roos, C. F. Sympa-
thetic cooling and detection of a hot trapped ion by a cold one. New J. Phys. 17,
103001 (2015).
38. Brewer, S. M. et al.
27
Al
+
Quantum-logic clock with a systematic uncertainty
below 10
18
,Phys. Rev. Lett. 123, 033201 (2019).
39. Chen, J.-S. Ticking near the Zero-Point Energy: Towards 1 × 10
18
Accuracy in Al
+
Optical Clocks. Ph.D thesis, University of Colorado Boulder (2017).
40. Schmidt, P. O. et al. Spectroscopy using quantum logic. Science 309, 749 (2005).
41. Siegman, A. Lasers. University Science Books (1986).
42. Hradil, Z., Summhammer, J. & Rauch, H. Quantum tomography as normalization
of incompatible observations. Phys. Lett. A 261, 20 (1999).
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative
Commons license, and indicate if changes were made. The images or other third party
material in this article are included in the articles Creative Commons license, unless
indicated otherwise in a credit line to the material. If material is not included in the
articles Creative Commons license and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly
from the copyright holder. To view a copy of this license, visit http://creativecommons.
org/licenses/by/4.0/.
© The Author(s) 2019
V. Krutyanskiy et al.
5
Published in partnership with The University of New South Wales npj Quantum Information (2019) 72
... The remote entanglement rate between trapped ion qubit modules scales with the photon emission rate [10]. To enhance the photon emission rate, one could place the ion inside an optical cavity [12,13]. The enhanced photon emission probability into the cavity output scales with the cooperativity of the system, while the reduced cavity length increases the emission rate [14]. ...
Preprint
The miniaturization of ion trap and the precise placement of its electrodes are necessary for the integration of a microcavity to facilitate efficient ion-cavity coupling. We present a miniature monolithic ion trap made of gold-coated fused silica with high numerical aperture access. A laser writing method referred to as selective laser etching is employed to extract a trap structure from a block of fused silica. The fully monolithic structure eliminates the need for any post-fabrication alignment. Trenches are integrated into this structure such that the various electrodes on the monolithic device remain electrically isolated following their metalization via evaporative coating. We give details of the trap design and production, along with the demonstration of successful trapping of ions and characerization of the trap.
... ParameterBaseline value Server efficiency (= emit × freq. conversion) 0.1325 (= 0.53[57] × 0.25[58]) ...
Article
Full-text available
In blind quantum computing (BQC), a user with a simple client device can perform a quantum computation on a remote quantum server such that the server cannot gain knowledge about the computation. Here, we numerically investigate hardware requirements for verifiable BQC using an ion trap as server and a distant measurement-only client. While the client has no direct access to quantum-computing resources, it can remotely execute quantum programs on the server by measuring photons emitted by the trapped ion. We introduce a numerical model for trapped-ion quantum devices in NetSquid, a discrete-event simulator for quantum networks. Using this, we determine the minimal hardware requirements on a per-parameter basis to perform the verifiable BQC protocol. We benchmark these for a five-qubit linear graph state, with which any single-qubit rotation can be performed, where client and server are separated by 50 km. Current state-of-the-art ion traps satisfy the minimal requirements on a per-parameter basis, but all current imperfections combined make it impossible to perform the blind computation securely over 50 km using existing technology. Using a genetic algorithm, we determine the set of hardware parameters that minimises the total improvements required, finding directions along which to improve hardware to reach our threshold error probability that would enable experimental demonstration. In this way, we lay a path for the near-term experimental progress required to realise the implementation of verifiable BQC over a 50 km distance.
... Heralded entanglement generation requires agreement between neighboring network nodes to trigger entanglement generation in precise time-bins [48], organized into a network schedule [49] that dictates when nodes make entanglement. It is a technological challenge to manage the interdependencies between the schedule of local operations, and the networked operations, since in all current processing node implementations [23,50], entanglement generation cannot be performed simultaneously with local operations [23,51]. While interdependencies may be mitigated in the future [52], this implies that we cannot schedule (i.e. ...
Preprint
Full-text available
The goal of future quantum networks is to enable new internet applications that are impossible to achieve using solely classical communication. Up to now, demonstrations of quantum network applications and functionalities on quantum processors have been performed in ad-hoc software that was specific to the experimental setup, programmed to perform one single task (the application experiment) directly into low-level control devices using expertise in experimental physics. Here, we report on the design and implementation of the first architecture capable of executing quantum network applications on quantum processors in platform-independent high-level software. We demonstrate the architecture's capability to execute applications in high-level software, by implementing it as a quantum network operating system -- QNodeOS -- and executing test programs including a delegated computation from a client to a server on two quantum network nodes based on nitrogen-vacancy (NV) centers in diamond. We show how our architecture allows us to maximize the use of quantum network hardware, by multitasking different applications on a quantum network for the first time. Our architecture can be used to execute programs on any quantum processor platform corresponding to our system model, which we illustrate by demonstrating an additional driver for QNodeOS for a trapped-ion quantum network node based on a single 40Ca+^{40}\text{Ca}^+ atom. Our architecture lays the groundwork for computer science research in the domain of quantum network programming, and paves the way for the development of software that can bring quantum network technology to society.
... This natural assumption allows us to substantially reduce the complexity of unitary certification by removing the need for high overall detection efficiencies without requiring to trust the calibration of the certification devices. We use this tool to set the first calibration-independent bound on the unitarity of a channel-a state-of-the-art polarization-preserving quantum frequency converter (QFC) [31][32][33][34] . We employ a trapped-ion platform as source of light-matter entanglement between an atomic Zeeman qubit and the polarization state of a spontaneously emitted photon 35 . ...
Article
Full-text available
We report on a method to certify a unitary operation with the help of source and measurement apparatuses whose calibration throughout the certification process needs not be trusted. As in the device-independent paradigm our certification method relies on a Bell test and requires no assumption on the underlying Hilbert space dimension, but it removes the need for high detection efficiencies by including the single additional assumption that non-detected events are independent of the measurement settings. The relevance of the proposed method is demonstrated experimentally by bounding the unitarity of a quantum frequency converter. The experiment starts with the heralded creation of a maximally entangled two-qubit state between a single ⁴⁰ Ca ⁺ ion and a 854 nm photon. Entanglement preserving frequency conversion to the telecom band is then realized with a non-linear waveguide embedded in a Sagnac interferometer. The resulting ion-telecom photon entangled state is assessed by means of a Bell-CHSH test from which the quality of the frequency conversion is quantified. We demonstrate frequency conversion with an average certified fidelity of ≥84% and an efficiency ≥3.1 × 10 ⁻⁶ at a confidence level of 99%. This ensures the suitability of the converter for integration in quantum networks from a trustful characterization procedure.
Article
A key challenge toward future quantum internet technology is connecting quantum processors at metropolitan scale. Here, we report on heralded entanglement between two independently operated quantum network nodes separated by 10 kilometers. The two nodes hosting diamond spin qubits are linked with a midpoint station via 25 kilometers of deployed optical fiber. We minimize the effects of fiber photon loss by quantum frequency conversion of the qubit-native photons to the telecom L-band and by embedding the link in an extensible phase-stabilized architecture enabling the use of the loss-resilient single-click entangling protocol. By capitalizing on the full heralding capabilities of the network link in combination with real-time feedback logic on the long-lived qubits, we demonstrate the delivery of a predefined entangled state on the nodes irrespective of the heralding detection pattern. Addressing key scaling challenges and being compatible with different qubit systems, our architecture establishes a generic platform for exploring metropolitan-scale quantum networks.
Article
The ability to distribute entanglement between quantum nodes may unlock new capabilities in the future that include teleporting information across multinode networks, higher resolution detection via entangled sensor arrays, and measurements beyond the quantum limit enabled by networked atomic clocks. These new quantum networks also hold promise for the Aerospace community in areas such as deep space exploration, improved satellite communication, and synchronizing drone swarms. Although exciting, these applications are a long way off from providing a “real-world” benefit, as they have only been theoretically explored or demonstrated in small-scale experiments. An outstanding challenge is to identify near-term use cases for quantum networks; this may be an intriguing new area of interest for the aerospace community, as the quantum networking field would benefit from more multidisciplinary collaborations. This paper introduces quantum networking, discusses the difficulties in distributing entanglement within these networks, highlights recent progress toward this endeavor, and features two current case studies on mobile quantum nodes and an entangled clock network, both of which are relevant to the aerospace community.
Article
The surface code is a quantum error-correcting code for one logical qubit, protected by spatially localized parity checks in two dimensions. Due to fundamental constraints from spatial locality, storing more logical qubits requires either sacrificing the robustness of the surface code against errors or increasing the number of physical qubits. We bound the minimal number of spatially nonlocal parity checks necessary to add logical qubits to a surface code while maintaining, or improving, robustness to errors. We saturate the lower limit of this bound, when the number of added logical qubits is a constant, using a family of hypergraph product codes, interpolating between the surface code and constant-rate low-density parity-check codes. Fault-tolerant protocols for logical gates in the quantum code can be inherited from its classical parent codes. We provide near-term practical implementations of this code for hardware based on trapped ions or neutral atoms in mobile optical tweezers. Long-range-enhanced surface codes outperform conventional surface codes using hundreds of physical qubits, and they represent a practical strategy to enhance the robustness of logical qubits to errors in near-term devices.
Article
Full-text available
The stages of a quantum internet As indispensable as the internet has become in our daily lives, it still has many shortcomings, not least of which is that communication can be intercepted and information stolen. If, however, the internet attained the capability of transmitting quantum information—qubits—many of these security concerns would be addressed. Wehner et al. review what it will take to achieve this so-called quantum internet and propose stages of development that each correspond to increasingly powerful applications. Although a full-blown quantum internet, with functional quantum computers as nodes connected through quantum communication channels, is still some ways away, the first long-range quantum networks are already being planned. Science , this issue p. eaam9288
Article
Full-text available
We report on the conversion to telecom wavelength of single photons emitted by a nitrogen-vacancy (NV) defect in diamond. By means of difference frequency generation, we convert spin-selective photons at 637 nm, associated with the coherent NV zero-phonon-line, to the target wavelength of 1588 nm in the L-telecom band. The successful conversion is evidenced by time-resolved detection revealing a telecom photon lifetime identical to that of the original 637 nm photon. Furthermore, we show by second-order correlation measurements that the single-photon statistics are preserved. The overall efficiency of this one-step conversion reaches 17\% in our current setup, along with a signal-to-noise ratio of \approx7 despite the low probability of an incident 637 nm photon. This result shows the potential for efficient telecom photon - NV center interfaces and marks an important step towards future long-range entanglement-based quantum networks.
Article
Full-text available
Interfacing fundamentally different quantum systems is key to building future hybrid quantum networks. Such heterogeneous networks offer capabilities superior to those of their homogeneous counterparts, as they merge the individual advantages of disparate quantum nodes in a single network architecture. However, few investigations of optical hybrid interconnections have been carried out, owing to fundamental and technological challenges such as wavelength and bandwidth matching of the interfacing photons. Here we report optical quantum interconnection of two disparate matter quantum systems with photon storage capabilities. We show that a quantum state can be transferred faithfully between a cold atomic ensemble and a rare-earth-doped crystal by means of a single photon at 1,552nanometre telecommunication wavelength, using cascaded quantum frequency conversion. We demonstrate that quantum correlations between a photon and a single collective spin excitation in the cold atomic ensemble can be transferred to the solid-state system. We also show that single-photon time-bin qubits generated in the cold atomic ensemble can be converted, stored and retrieved from the crystal with a conditional qubit fidelity of more than 85 per cent. Our results open up the prospect of optically connecting quantum nodes with different capabilities and represent an important step towards the realization of large-scale hybrid quantum networks.
Article
Full-text available
Quantum computers will eventually reach a size at which quantum error correction becomes imperative. Quantum information can be protected from qubit imperfections and flawed control operations by encoding a single logical qubit in multiple physical qubits. This redundancy allows the extraction of error syndromes and the subsequent detection or correction of errors without destroying the logical state itself through direct measurement. We show the encoding and syndrome measurement of a fault-tolerantly prepared logical qubit via an error detection protocol on four physical qubits, represented by trapped atomic ions. This demonstrates the robustness of a logical qubit to imperfections in the very operations used to encode it. The advantage persists in the face of large added error rates and experimental calibration errors.
Article
Full-text available
Entanglement between a stationary quantum system and a flying qubit is an essential ingredient of a quantum-repeater network. It has been demonstrated for trapped ions, trapped atoms, color centers in diamond, or quantum dots. These systems have transition wavelengths in the blue, red or near-infrared spectral regions, whereas long-range fiber-communication requires wavelengths in the low-loss, low-dispersion telecom regime. A proven tool to interconnect flying qubits at visible/NIR wavelengths to the telecom bands is quantum frequency conversion. Here we use an efficient polarization-preserving frequency converter connecting 854\,nm to the telecom O-band at 1310\,nm to demonstrate entanglement between a trapped 40^{40}Ca+^{+} ion and the polarization state of a telecom photon with a high fidelity of 98.2 ±\pm 0.2%\%. The unique combination of 99.75 ±\pm 0.18%\% process fidelity in the polarization-state conversion, 26.5%\% external frequency conversion efficiency and only 11.4 photons/s conversion-induced unconditional background makes the converter a powerful ion-telecom quantum interface.
Article
Full-text available
A major outstanding problem for many quantum clock synchronization protocols is the hidden assumption of the availability of synchronized clocks within the protocol. In general, quantum operations between two parties do not have consistent phase definitions of quantum states, which introduce an unknown systematic phase error. We show that despite prior arguments to the contrary, it is possible to remove this unknown phase via entanglement purification. This closes the loophole for entanglement based quantum clock synchronization protocols, which are most compatible with current photon based long-distance entanglement distribution schemes. Starting with noisy Bell pairs, we show that the scheme produces a singlet state for any combination of (i) differing basis conventions for Alice and Bob; (ii) an overall time offset in the execution of the purification algorithm; and (iii) the presence of a noisy channel. Error estimates reveal that better performance than existing classical Einstein synchronization protocols should be achievable using current technology.
Article
Full-text available
We demonstrate polarisation-preserving frequency conversion of single-photon-level light at 854 nm, resonant with a trapped-ion transition and qubit, to the 1550-nm telecom C band. A total photon in / fiber-coupled photon out efficiency of ∼ 30% is achieved, for a free-running photon noise rate of ∼ 60 Hz. This performance would enable telecom conversion of 854 nm polarisation qubits, produced in existing trapped-ion systems, with a signal-to-noise ratio greater than 1. In combination with near-future trapped-ion systems, our converter would enable the observation of entanglement between an ion and a photon that has travelled more than 100 km in optical fiber: three orders of magnitude further than the state-of-the-art.
Article
We describe an optical atomic clock based on quantum-logic spectroscopy of the S01↔P30 transition in Al+27 with a systematic uncertainty of 9.4×10−19 and a frequency stability of 1.2×10−15/τ. A Mg+25 ion is simultaneously trapped with the Al+27 ion and used for sympathetic cooling and state readout. Improvements in a new trap have led to reduced secular motion heating, compared to previous Al+27 clocks, enabling clock operation with ion secular motion near the three-dimensional ground state. Operating the clock with a lower trap drive frequency has reduced excess micromotion compared to previous Al+27 clocks. Both of these improvements have led to a reduced time-dilation shift uncertainty. Other systematic uncertainties including those due to blackbody radiation and the second-order Zeeman effect have also been reduced.
Article
Quantum networks consisting of quantum memories and photonic interconnects can be used for entanglement distribution (L.-M.Duan and H. J. Kimble, PRL 90, 253601 (2003), H. J. Kimble, Nat. 453, 1023 EP (2008)), quantum teleportation (S.Pirandola et.al, Nat. Photon. 9, 641 (2015)), and distributed quantum computing (T.Spiller, et.al., New J. Phys. 8, 30 (2006). Remotely connected two-node networks have been demonstrated using memories of the same type: trapped ion systems (D.Hucul, et.al, Nat. Phys. 11, 37 (2014)), quantum dots (W. B. Gao et. al, Nat. Photon. 9, 363 (2015)), and nitrogen vacancy centers (W. B. Gao et. al, Nat. Photon. 9, 363 (2015), B. Hensen, et. al., Nat. 526, 682 (2015)). Hybrid systems constrained by the need to use photons with the native emission wavelength of the memory, have been demonstrated between a trapped ion and quantum dot (H.Meyer et.al PRL 114, 123001 (2015)) and between a single neutral atom and a Bose-Einstein Condensate (M.Lettner et. al, PRL 106, 210503 (2011)). Most quantum systems operate at disparate and incompatible wavelengths to each other so such two-node systems have never been demonstrated. Here, we use a trapped 138Ba+^{+} ion and a periodically poled lithium niobate (PPLN) waveguide, with a fiber coupled output, to demonstrate 19% end-to-end efficient quantum frequency conversion (QFC) of single photons from 493 nm to 780 nm. At the optimal signal-to-noise operational parameter, we use fluorescence of the ion to produce light resonant with the 87^{87}Rb D2D_2 transition. To demonstrate the quantum nature of both the unconverted 493 nm photons and the converted photons near 780 nm, we observe strong quantum statics in their respective second order intensity correlations. This work extends the range of intra-lab networking between ions and networking and communication between disparate quantum memories. A
Article
Trapped atomic ions are ideal single photon emitters with long lived internal states which can be entangled with emitted photons. Coupling the ion to an optical cavity enables efficient emission of single photons into a single spatial mode and grants control over their temporal shape. These features are key for quantum information processing and quantum communication. However, the photons emitted by these systems are unsuitable for long-distance transmission due to their wavelengths. Here we report the transmission of single photons from a single 40Ca+^{40}\text{Ca}^{+} ion coupled to an optical cavity over a 10 km optical fibre via frequency conversion from 866 nm to the telecom C-band at 1,530 nm. We observe non-classical photon statistics of the direct cavity emission, the converted photons and the 10 km transmitted photons, as well as the preservation of the photons' temporal shape throughout. This telecommunication ready system can be a key component for long-distance quantum communication as well as future cloud quantum computation.