Quantum repeaters with individual rare-earth ions at telecom-
munication wavelengths
F. Kimiaee Asadi, N. Lauk, S. Wein, N. Sinclair, C. O’Brien, and C. Simon
Institute for Quantum Science and Technology, and Department of Physics & Astronomy, University of Calgary, 2500 University Drive NW,
Calgary, Alberta T2N 1N4, Canada
August 30, 2018
We present a quantum repeater scheme that is
based on individual erbium and europium ions.
erbium ions are attractive because they emit
photons at telecommunication wavelength, while
europium ions offer exceptional spin coherence
for long-term storage. Entanglement between
distant erbium ions is created by photon detec-
tion. The photon emission rate of each erbium
ion is enhanced by a microcavity with high Pur-
cell factor, as has recently been demonstrated.
Entanglement is then transferred to nearby eu-
ropium ions for storage. Gate operations be-
tween nearby ions are performed using dynam-
ically controlled electric-dipole coupling. These
gate operations allow entanglement swapping to
be employed in order to extend the distance
over which entanglement is distributed. The de-
terministic character of the gate operations al-
lows improved entanglement distribution rates
in comparison to atomic ensemble-based proto-
cols. We also propose an approach that utilizes
multiplexing in order to enhance the entangle-
ment distribution rate.
1 Introduction
Entanglement is a key requirement for many applica-
tions of quantum science. These include, for example,
quantum key distribution [1], global clock networks [2],
long-baseline telescopes [3], and the quantum internet
[4, 5]. However, due to transmission loss the direct
transmission of entanglement over distances of more
than several hundred kilometers is practically impos-
sible using current technology. The use of a quantum
repeater has been suggested to reduce (or eliminate)
the impact of loss in order to establish entanglement be-
tween distant locations. [6]. In many quantum repeater
schemes, entanglement is first distributed between two
locations that are separated by a short distance, referred
to as an elementary link. Then, the range of entangle-
ment is extended to successively longer distances by per-
forming entanglement swapping operations between the
entangled states that span each elementary link. Due to
the availability and diversity of component systems and
the strong light-matter coupling offered by atomic en-
sembles, many quantum repeater proposals use sources
of entanglement, ensemble-based quantum memories,
and linear optics-based entanglement swapping oper-
ations [7]. However, the success probability of linear
optics-based entanglement swapping (without auxiliary
photons) cannot exceed 50%, which has a compounding
effect for more complex quantum networks. The use of
auxiliary photons to improve the entanglement swap-
ping probability is possible [8, 9], but adds complex-
ity and compounds errors, thereby restricting their use
in practice. Single-emitter-based quantum repeaters,
on the other hand, offer the possibility to outperform
ensemble-based repeaters by using deterministic swap-
ping operations [10]. Impressive demonstrations of cer-
tain parts of a single-emitter quantum repeater schemes
have been performed using atom-cavity systems [11, 12],
color centers in diamond [13, 14], trapped ions [15, 16],
donor qubits in silicon [17], as well as quantum dots
[18, 19].
Over the years, crystals doped with rare earth (RE)
ions have attracted considerable attention for their use
in electromagnetic signal processing applications that
range from quantum memories [20, 21] to biological
imaging [22]. Narrow optical and spin homogeneous
linewidths, their convenient wavelengths as well as their
ability to be doped into solid-state crystals are some of
the most desired properties of RE ions. Compared to
nitrogen-vacancy centers in diamond [23] and quantum
dots [24], they also exhibit smaller spectral diffusion
[2527], and have a reduced sensitivity to phonons.
Among the different RE ions, Er
3+
is attractive due
to its well-known optical transition (around 1.5 µm)
in the conventional telecommunication wavelength win-
dow, in which absorption losses in optical fibers are min-
imal. Another unique aspect of certain RE ions is the
presence of hyperfine levels that feature long lifetimes,
which allows for long-term quantum state storage [28
31]. In particular, a coherence lifetime of six hours
Accepted in Quantum 2018-08-27, click title to verify 1
arXiv:1712.05356v3 [quant-ph] 29 Aug 2018
was reported in a europium-doped yttrium orthosili-
cate crystal (
151
Eu
3+
:Y
2
SiO
5
) [32]. Motivated by these
properties, we propose and analyze a quantum repeater
protocol that is based on individual RE ions in which
Er
3+
ions are used to establish entanglement over ele-
mentary links and
151
Eu
3+
ions are employed to store
this entanglement.
One disadvantage of RE ions is their weak light-
matter coupling, which has mostly precluded their use
as single quantum emitters. However, optical detec-
tion and addressing of single RE ions has recently been
shown by multiple groups [3337]. Moreover, very re-
cently Dibos et al. [34] demonstrated an enhancement
in the emission rate of a single Er
3+
ion in Y
2
SiO
5
by
a factor of more that 300 using a silicon nanophotonic
cavity. Strong coupling of ensembles of Nd ions was pre-
viously demonstrated using nanophotonic cavities fabri-
cated from Y
2
SiO
5
[38] and yttrium orthovanadate [39]
hosts. In light of these results, for our scheme we pro-
pose to couple single RE ions to a high-finesse cavity in
order to enhance the light-matter coupling and thus to
increase the collection efficiency as well as the indistin-
guishability of the emitted single photons.
The paper is organized as follows. In Sec. 2 we intro-
duce our proposal and discuss the required components
as well as the underlying principles. The entanglement
distribution rates and possible implementations are dis-
cussed in Secs. 3 and 4. We conclude and provide an
outlook in Sec. 5.
2 Quantum repeater protocol
The goal of our repeater scheme is to generate entan-
glement between nodes that are separated by a long
distance by swapping entanglement that is established
between nodes that are separated by smaller distances.
Each node consists of a Y
2
SiO
5
crystal containing an
Er
3+
ion that features an enhanced emission rate of
its telecommunication transition due to a high-finesse
cavity, and a nearby
151
Eu
3+
ion that acts as a quan-
tum memory due to its long spin coherence lifetime.
Entanglement is transferred, or gates are performed,
between nearby atoms using electric dipole-dipole cou-
pling [40, 41]. Our proposal has some similarity to a
scheme that involves the excitation of electron spins in
nitrogen-vacancy centers to generate entanglement be-
tween nodes, nuclear spins of carbon atoms for storage,
and gates that are performed using a magnetic dipole-
dipole coupling [42]. We emphasize that our (electric
dipole) approach to implementing gates is quite differ-
ent than previous approaches, as are many of the physi-
cal properties of the system (in addition to the emission
wavelength).
In the first step, entanglement between pairs of neigh-
boring Er
3+
ions is created (see Fig.1 (a)) which illus-
trates this step for the i-th and i+1-th ions) by first cre-
ating local entanglement between the spin state of each
ion and a spontaneously-emitted single photon. Then,
by performing a joint Bell-state measurement (BSM) on
photons that are spontaneously-emitted by neighboring
Er
3+
ions, the spin state of these ions is projected onto a
maximally-entangled state. This procedure is the same
as that used to create entanglement between two remote
nitrogen-vacancy centers in diamond [13, 43].
(i) (i+1) (i+2)(i-1)
a.
b.
c.
d.
Er
3+
Eu
3+
m
s
= 1 / 2
m
s
= 1 / 2
7
F
0
5
D
0
4
I
13/2
4
I
15/2
Map
Er-Eu
Swap
Er-Eu
Fail
Fail
Fail
Fail
e.
m
I
= 3 / 2
m
I
= 3 / 2
Figure 1: In each cavity, the blue and black dots represent
individual
151
Eu
3+
and Er
3+
ions, respectively. In order to dis-
tribute entanglement to the nodes that are situated at each end
of a long channel, first a) entanglement between two neighbor-
ing Er
3+
ions is created via the detection of two photons; b)
the quantum state of each Er
3+
ion is then mapped to a nearby
151
Eu
3+
ion using an electric dipole interaction; c) using two-
photon detection, Er
3+
ions are used again to establish entan-
glement between the other neighboring Er
3+
ions. Finally, we
extend the entanglement distance by performing entanglement
swapping using an electric dipole interaction between nearby
ions (red circle). d) Flow-chart that depicts our scheme. e)
Simplified energy level structure of Er
3+
and
151
Eu
3+
ions in
the presence of an external magnetic field. The electronic Zee-
man levels of Er
3+
, and nuclear hyperfine levels of
151
Eu
3+
,
are split. We encode qubit states |↑i and |↓i in the m
s
=
1
2
and m
s
=
1
2
m
I
=
3
2
and m
I
=
3
2
levels for the Er
3+
(
151
Eu
3+
) ion, respectively.
Accepted in Quantum 2018-08-27, click title to verify 2
Next, the quantum state of each Er
3+
ion is mapped
to a nearby
151
Eu
3+
ion for storage by exploiting a non-
vanishing permanent electric dipole moment, a common
feature of many RE ions that are doped into solids.
This mapping is achieved using a mutual electric dipole-
dipole interaction between close-lying Er
3+
and
151
Eu
3+
ions. The small nuclear magnetic moment of
151
Eu
3+
results in a magnetic-dipole coupling that is orders of
magnitude smaller than the electric-dipole coupling,
thus our scheme allows for much shorter gate durations
in comparison to those based on magnetic interactions
[42]. Another advantage of the proposed scheme is the
ability to dynamically control the interaction optically
by bringing the ion to the excited state and back to the
ground state. This allows for the realization of deter-
ministic two-qubit gate operations.
The mapping allows the Er
3+
ions to be re-initialized
so that new elementary links can be created between
them. Fig.1(c) illustrates this processes for the (i 1)th
and (i)th nodes. Immediately after generating entan-
glement between (the other) neighboring Er
3+
ions, the
entanglement distance is extended by performing en-
tanglement swapping between each of the closely-lying
Er
3+
and
151
Eu
3+
ions. As a result, the outer nodes
becoming entangled. Fig. 1 (d) depicts a flow chart of
our scheme. Photon loss and errors might cause some
steps in the protocol to fail.
As will be discussed in Sec. II A–C, the generation
of entanglement between Er
3+
ions, the mapping of
entanglement to the
151
Eu
3+
ions, and the entangle-
ment swapping steps, all rely on the detection of sin-
gle photons that are spontaneously emitted from the
Er
3+
ions. However, the lifetime of the
4
I
15/2
4
I
13/2
telecommunication transition of Er
3+
:Y
2
SiO
5
is rela-
tively long T
1
= 11.4 ms [44] with radiative lifetime
being even longer T
rad
= 54 ms [45], which necessi-
tates the need for a high-finesse cavity to enhance the
emission rate. Nonetheless, the cavity will also signifi-
cantly increase the probability of collecting the emitted
photons as compared to emission into free space. We
emphasize that the repetition rate of the protocol (i.e.,
the number of attempts of the protocol per unit time)
is limited by the communication time between distant
nodes, which means that there is no need for very fast
emission. Next we discuss each of the steps of the pro-
tocol in detail.
2.1 Entanglement generation
Our scheme considers several remote cavities that each
contain a single pair of close-lying Er
3+
and
151
Eu
3+
ions. Note that there may be other RE ions within each
cavity, but we assume that we can address a single such
close-lying pair. See also section4 below. An externally-
applied magnetic field splits the degenerate electronic
ground levels of Er
3+
via the Zeeman effect. We refer
to the resultant m
s
=
1
2
and m
s
=
1
2
Zeeman levels
as the qubit states |↑i and |↓i, respectively (see Fig. 1
(e)).
To generate entanglement between distant Er
3+
ions
that are separated by a long distance L
0
, we follow
the scheme of Barret and Kok [13, 43] (see also Fig.
1 (d)). First, each Er
3+
ion is prepared in one of the
qubit states (here m
s
=
1
2
; denoted |↑i). After this
initialization step, a
π
2
microwave (MW) pulse rotates
each Er
3+
ion into a superposition of |↑i and |↓i states.
The application of a brief laser pulse that is resonant
with the |↑i |e
2
i transition, followed by spontaneous
emission, will entangle each qubit state with the emit-
ted photon number. That is, when the qubit state is
|↑i (|↓i) there will be 1 (0) emitted photon(s). The
spontaneously-emitted photons are then directed to a
beam splitter located in between the ions using opti-
cal fibers. A single-photon detection at one of the two
beam splitter output ports projects the Er
3+
ions onto
an entangled state.
A possible loss in the fiber can lead to a situation
where both Er
3+
ions emit a photon but only one pho-
ton is detected while the other is lost. In this case, the
ions are left in a product state rather than an entangled
state. To exclude this possibility, immediately after the
first excitation-emission step of each Er
3+
ion, a π MW
pulse inverts each qubit state. Then a second excitation
pulse is applied. The detection of two consecutive single
photons at the beam splitter will leave the qubits in an
entangled state:
|ψi
±
Er
=
1
2
(|↑↓i ± |↓↑i). (1)
Here the + (-) sign corresponds to the case in which the
same (different) detector(s) received a photon.
2.2 Controlled logic
After successful entanglement of the two distant Er
3+
ions we transfer this entanglement to the neighboring
151
Eu
3+
ions for a long-term storage. Efficient entan-
glement mapping between neighboring rare-earth ions
can be employed by performing CNOT gate operations
and single qubit rotations and read-out. Here we first
explain the underlying mechanism and general scheme
for how to implement a CNOT gate in our system, and
in Sec.2.3 we will discuss in more detail how to swap
entanglement between the ions using this gate. Due to
a lack of site symmetry when doped into a crystal, a
RE ion can have a permanent electric dipole moment
that is different depending on whether the ion is in its
ground or optically excited state. The difference in the
Accepted in Quantum 2018-08-27, click title to verify 3
1 2 34
Δ
υ
ω
e, Eu
0
ω
, Eu
ω
Eu
ω
e, Er
0
ω
, Er
ω
Er
3
`
Figure 2: General pulse sequence of π-rotations to perform
a controlled logic gate between nearby Er
3+
and
151
Eu
3+
ions. The numbers indicate the sequential time ordering of
the pulses. For this gate, the Er
3+
spin state acts as the con-
trol qubit and the
151
Eu
3+
spin state acts as the target. A π
pulse excites the Er
3+
ion if it is in the state |↑i. When this
occurs, pulses 2, 3 and 4 are not resonant with the
151
Eu
3+
ion, leaving its state unaffected. Pulse 5 then brings the Er
3+
ion back to its original state. On the other hand, when Er
3+
is
in the state |↓i, pulses 1 and 5 will be ineffective and hence it
will remain in the ground state. Instead pulses 2, 3 and 4 will
now be resonant with the
151
Eu
3+
ion and (optically) swap its
spin-state
permanent dipole moments affects the optical transition
frequency of other nearby RE ions via the Stark effect
due to a modified local electric field environment. It is
possible to dynamically control the shift in the transi-
tion frequency of one ion by optically exciting its neigh-
boring ion. Based on this interaction, we can perform a
controlled-NOT (CNOT) operation between nearby RE
ions [46]. The modification of the transition frequency
ν of a
151
Eu
3+
ion by an Er
3+
ion due to the mutual
electric dipole-dipole interaction is given by [47]:
ν =
µ
Er
µ
Eu
4π
0
hr
3
((ˆµ
Er
·ˆµ
Eu
) 3 (ˆµ
Er
·ˆr) (ˆµ
Eu
·ˆr)) ,
(2)
where r is the distance between the ions, and µ is the
change of the permanent electric dipole moment of each
ion, h is the Planck constant,
0
is vacuum permittivity,
and is the dielectric constant.
To perform a CNOT gate between the nearby Er
(control) ion and Eu (target) ion, a sequence of five
π pulses is applied (see Fig.(2)). First a π pulse is ap-
plied to the Er
3+
ion on resonance with the |↑i |e
1
i
transition. From here, two cases must be considered:
either the state of the Er
3+
ion is |↑i or it is |↓i.
If the Er
3+
ion is in state |↑i, it will be excited by
pulse 1. This changes the permanent electric dipole
moment of the Er
3+
ion, and thus its local electric field.
Consequently, the transition frequency of the nearby
151
Eu
3+
ion will be shifted by ν. For the case that
this frequency shift is large enough (that is, the
151
Eu
3+
ion is sufficiently close to the Er
3+
ion), the
151
Eu
3+
ion will be unaffected by pulses 2, 3 and 4, thereby
remaining in its initial ground state (see Sec. 4.2 for
more discussion). Finally, pulse 5 will bring Er
3+
ion
back to its initial state. Hence, in this case, the pulse
sequence does not modify the initial state of the ion pair
system.
On the other hand, if the Er
3+
ion is initially in |↓i,
then pulses 1 and 5 will have no effect on the Er
3+
ion.
Since the Er
3+
ion will not be excited, optical pulses
2, 3 and 4 will be resonant with the transitions of the
151
Eu
3+
ion, and the pulse sequence will optically flip
the two spin states of the
151
Eu
3+
ion.
2.3 Entanglement mapping and distribution
Both the ground and excited states of
151
Eu
3+
have
three doubly-degenerate nuclear hyperfine levels. With
the application of a magnetic field, each doublet
m
I
= ±
1
2
, ±
3
2
, ±
5
2
will be split. For our proposal, we
denote the m
I
=
3
2
and
3
2
hyperfine levels [32] as the
151
Eu
3+
ion qubit states, |↑i and |↓i, respectively (see
Fig. 1 (e)). To map the state of each Er
3+
ion onto a
nearby Eu ion, we first perform a CNOT gate between
them. In our scheme, the
151
Eu
3+
ion is initially pre-
pared in one of the ground state levels (here |↑i) using
optical pumping. In this special case, pulse 4 does not
affect the system and can be neglected. This reduces
the total gate time and improves the state mapping fi-
delity (see Sec. 4.2).
After considering the phases that are acquired by
performing the CNOT gate on the ions in neighbor-
ing nodes (labeled here as 1 and 2 instead of (i) and
(i + 1)), the final state is
|ψi
±
Er,Eu
=
1
2
|↑↓i
Er
1
,Er
2
|↑↓i
Eu
1
,Eu
2
±e
|↓↑i
Er
1
,Er
2
|↓↑i
Eu
1
,Eu
2
, (3)
where
φ = (ω
,Er
2
ω
,Er
1
)τ + (ω
Eu
2
ω
Eu
1
)τ
2
+ (ω
,Eu
2
ω
,Eu
1
)(τ
3
+ τ
4
)
+ k
Eu
1
x
Eu
1
k
Eu
2
x
Eu
2
, (4)
and x is the distance the photon travels between each
151
Eu
3+
ion and the beam splitter, k
Eu
=
ω
,Eu
c
is the
wavenumber, τ
j
is the time elapsed after application of
the j
th
pulse and τ is the total time duration that is
needed to perform a CNOT gate. Since these phases
are known, they can be compensated by applying local
operations on the ions.
To conclude the mapping step, a
π
2
microwave (MW)
pulse is applied to rotate each Er
3+
qubit. This is fol-
lowed by a state measurement of both nearest-neighbor
Er
3+
ions, see also Sec. IV below. This projects the
Accepted in Quantum 2018-08-27, click title to verify 4
151
Eu
3+
ions onto an entangled state. Depending on the
outcome of these measurements, the entangled state be-
tween remote
151
Eu
3+
ions would be |ψ
+
i or |ψ
i. For
example, in the case that we begin with |ψi
+
Er
(given
in Eq. 1), after performing the gate, MW rotation, and
measurement, if both Er ions are found in the state
|↑i, the entangled state between remote
151
Eu
3+
ions is
1
2
|↑↓i
Eu
1
,Eu
2
+ |↓↑i
Eu
1
,Eu
2
.
Once entanglement is established in neighboring ele-
mentary links, we perform a joint measurement on both
ions at each intermediate node to distribute entangle-
ment (see Fig. 1 (c)). In our scheme, it is possible to
perform entanglement swapping deterministically using
the permanent electric dipole-dipole interaction.
To perform the desired entanglement swapping, first
a CNOT gate is applied in which Er
3+
serves as the con-
trol qubit and
151
Eu
3+
as the target qubit (similarly as
before). Here, both ions are in a superposition state
and so all 5 pulses of the CNOT gate sequence are re-
quired. Then the Er
3+
ion is measured in the diagonal
(X) basis. Next, another CNOT operation is performed
but now the target and control qubits are exchanged.
Finally the Er
3+
ion is measured in the logical (Z) ba-
sis. This ‘reverse’ CNOT gate is performed in order to
avoid directly measuring the spin state of
151
Eu
3+
ion
optically (the optical lifetime of
151
Eu
3+
is longer than
the Er
3+
spin coherence lifetime). Fig.(2) shows the
pulse sequence needed to perform the first CNOT gate
between the ions.
Based on the outcomes of the measurements on each
Er
3+
ion, the outer nodes will be projected onto one of
the four Bell states. To be more precise, when perform-
ing entanglement swapping between Er
i
and Eu
i
, if only
one of the state measurements of Er
i
is |↑i, the state of
Er
i1
and Eu
i+1
will project onto the |ψ
+
i or the |ψ
i
Bell state. On the other hand, if both measurements of
Er
i
are |↑i or both |↓i, the state will project onto the
|φ
+
i or the |φ
i Bell state.
To verify or utilize the entanglement that is gener-
ated between the distant ions, a measurement must be
performed on the endpoint nodes. Our protocol results
in an entangled state of the Er
3+
ion at the first node
with the
151
Eu
3+
ion at the last node. A measurement
of the Er
3+
ion can be performed by direct optical exci-
tation. A measurement of the
151
Eu
3+
ion spin can be
performed by using the nearby Er
3+
as a readout ion.
This is done by mapping the
151
Eu
3+
state to Er
3+
using gate operations (also see section 4.3 for further
details).
3 Entanglement generation rates and
multiplexing
In our scheme, the success probability of generating an
entangled state between two neighboring Er
3+
ions that
are separated by a distance L
0
is p
t
=
1
2
η
2
t
p
2
η
2
d
, where
η
t
= e
L
0
2L
att
is the transmission probability of a photon
through optical fiber and L
att
22 km (which cor-
responds to a loss of 0.2 dB/km), η
d
is the detection
efficiency, and p is the success probability of emitting a
single photon into a cavity mode. Similarly, the success
probability of mapping the state of an Er
3+
ion onto
a nearby
151
Eu
3+
ion is p
m
d
, because it requires
one optical read-out of each Er
3+
ion. Therefore, the
average time to generate entanglement in an elementary
link and perform the state mapping steps is
hT i
L
0
=
L
0
c
1
p
t
p
2
m
, (5)
where c = 2 × 10
8
m
s
is the speed of light in fiber.
The average time that is required to distribute entan-
glement over two neighboring elementary links, which
corresponds to a distance L = 2L
0
, can be estimated
as follows. First, entangled pairs of
151
Eu
3+
ions are
generated over an elementary link and, once successful,
Er
3+
ions are then generated in a neighboring elemen-
tary link. Thus, the probability of establishing entan-
glement for both links is
p
0
=
1
P
A
+
1
P
B
1
, (6)
where P
A
= p
t
and P
B
= p
t
p
2
m
represent the success
probability of generating entanglement between Er
3+
and
151
Eu
3+
ions, respectively. Then, entanglement is
extended by performing a BSM between the close-lying
Er
3+
and
151
Eu
3+
ions at the center node. The success
probability for the entanglement swapping step is p
s
p
2
η
2
d
because it requires two optically-induced spin read-
outs of an Er
3+
ion. Consequently, the average time to
distribute entanglement over a distance 2L
0
is
hT i
2L
0
=
L
0
c
1
p
0
p
s
. (7)
Accordingly, the average time to distribute an entan-
gled pair over a distance L = 2
n
L
0
, with n denoting the
number of nesting levels, is [7, 10]:
hT i
L
=
3
2
n1
L
0
c
1
p
0
p
n
s
. (8)
The factor 3/2 for each of the next nesting levels, which
is a good approximation for the exact result [7], can be
understood in following way.
Accepted in Quantum 2018-08-27, click title to verify 5
In contrast to the first nesting level, a successful en-
tanglement distribution for higher nesting levels does
not require waiting for a success in one link before the
establishment of entanglement in another link can be
attempted; rather, the establishment of entanglement
can be attempted in both links simultaneously. For ex-
ample, if a goal is to distribute entanglement over the
distance 4L
0
, entanglement must be generated in two
neighboring links each of length 2L
0
before entangle-
ment swapping is performed. If the average waiting
time for a success in one link of length 2L
0
is hT i
2L
0
,
entanglement will be established in one of two links af-
ter hT i
2L
0
/2. Then, after another hT i
2L
0
time dura-
tion, entanglement will be established in the other link.
Hence, the average time to establish entanglement in
the two neighboring links is 3 hT i
2L
0
/2. The same ar-
guments hold for the next nesting levels, resulting in a
factor
3
2
n1
for n nesting levels.
Multiplexing can be employed to significantly en-
hance the rate of entanglement distribution. Referring
to the encoding of several individual qubits, each into
their own distinguishable modes, multiplexing has been
utilized for some quantum repeater proposals [7, 48
50]. As outlined in Fig. 3, we consider an array of
m cavities in which each cavity emits a photon that
features a distinguishable resonance frequency from the
rest. This can be accomplished using frequency trans-
lation (see Sec. IV for more details). Using a coupler
device, each photon is directed into a common fiber and
is directed towards a BSM station that is composed of
similarly-designed decouplers and single-photon detec-
tors. Benefiting from the distribution of entanglement
into m parallel modes, the probability that at least one
entangled state is distributed over the entire channel is
1(1P
t
)
m
, which can be made unity for a sufficiently
high m. Here, P
t
= (2/3)
n1
p
0
p
n
s
is the success prob-
ability of distributing entanglement over a distance L
using a qubit that is encoded into a single mode. Fur-
ther details of this set-up are described in Sec.4.4.
The entanglement distribution rate of our scheme is
plotted as a function of distance for n = 3 in Fig. 4
and compared to that employing the well-known DLCZ
scheme [51], which uses ensemble-based memories, as
well as that using direct transmission with a single-
photon source which produces photons at 10 GHz. This
10 GHz photon rate is an optimistic rate for the direct
transmission of photons. The rate that we assume for
a single photon source is much faster than the rate for
the proposed repeater because they have different lim-
itations. Even though the photon rate of our scheme
(which depends on the cavity characteristics and the
optical lifetime of the Er
3+
ion) is much lower, the time
scale for the repeater is actually determined by the
communication time L
0
/c which is even longer. Note
MOTIVATION
Why+using+rare-earth+ions?+
Storing(and(retrieving(photons(at(conventional(telecom(wavelength(
Long(spin(coherence(times,(up(to(6(hours,(are(feasible(
Provide(scalab le(architecture
PROPOSAL
TEXT+BOXES
Contact+Information+
Email:((faezeh.kimiaeeasadi@ucalgary.ca(
Email:((nikolai.lauk@ucalgary.ca
Quantum+repeaters+with+single+rare-earth+ions+at+
telecommunication+wavelengths
Faezeh+Kimiaee+Asadi,+Nikolai+Lauk,+Stephen+Wein,+Chris+O’Brien,+Neil+Sinclair+and+Christoph+Simon
+
Department(of(Physics,(Institute(for(Quantum(Science(and(Technology,(University(of(Calgary,(Canada
Eu
3+
Er
3+
1.+Creating+heralded+entanglement+between+distant+Er+ions+using+Barrett-Kok+scheme++
2.+Mapping+Er+states+to+nearby+Eu+ions+using+dipole-dipole+interaction
3.+Er+ions+are+used+again+to+create+new+entangled+pairs
4.+Performing+deterministic+entanglement+swapping+between+Er +and+Eu+ions+
TEXT+BOXES
ENTANGLEMENT+GENERATION+RATES+AND+MULTIPLEXING
Tel ecom(
C-band
1535(nm
4
I
15/2
m
s
=1/2
m
s
= 1/2
| "i
| #i
4
I
13/2
|e
1
i
|e
2
i
cavity(enhanced(
transition
|ei
5
D
0
m
I
= 3/2
m
I
=3/2
7
F
0
| "i
| #i
Erbium Europium
|ei
Er
|ei
Eu
| #i
Er
| #i
Eu
| "i
Eu
| "i
Er
1
4
2
3
ij
ij
=
µ
i
µ
j
4⇡✏
0
hr
ij
µ
i
· ˆµ
j
3(ˆµ
i
· ˆr)(ˆµ
j
· ˆr))
dipole-dipole(shift
ij
10 MHZ
10 kHZ
r
ij
1nm
10 nm
State(after(performing(CNOT(gate(
and(measuring(Er(ions(in(X-basis
| i = |"#i
Er
1
,Er
2
±|#"i
Er
1
,Er
2
| i = |"#i
Eu
1
,Eu
2
±|#"i
Eu
1
,Eu
2
|ei
Er
|ei
Eu
| #i
Er
| #i
Eu
| "i
Eu
| "i
Er
1
5
2
3
ij
4
Perform(CNOT(gate(using(Er(as(
control(qubit(
Measure(Er(in(X-basis(
Perform(CNOT(gate(using(Eu(as(
control(qubit(
Measure(Er(in(X-basis
Procedure(analogous(to(Step(1.
Initialize(Er(state(
Excite(and(let(emit(a(photon(
Flip(Er(qubit(and(excite(again
State(after(detecting(two(photons
General+scheme+to+generate+and+distribute+entanglement+
Initialize(
Eu
Initialize(
Er
Entangle(
Er-Er(
Map(
Er-Eu
Initialize(
Er
Entangle(
Er-Er(
Swap
Fail
Fail
Fail
Fail
Possible+imperfections:+
transmission(efficiency((
detector(efficiency((
probability(of(emittin g(into(cavity(mode(
small(deviation(from(the(rotation(angle
d
t
p
P
Er-Er
1
2
(1
3
2
4
)
t
p
2
2
d
Probability(to(entangle(Er(ions
Probability(to(map(Er(state(to(Eu
Entanglement(swapping
P
m
(1
2
2
)p
d
P
s
(1
2
)p
2
2
d
=)
Using(these(quantities(we(can(calculate:
1.+Average+time+required+to+entangle+two+Eu+ions+separated+by+the+distance+
L
0
hT i
L
0
=
L
0
c
1
P
Er-Er
P
2
m
=
L
0
c
1
P
Eu-Eu
2.+Average+time+required+to+distribute+entanglement+over+the+distance++
2L
0
hT i
2L
0
=
L
0
c
1
P
s
P
0
=
L
0
c
1
P
s
1
P
Er-Er
+
1
P
Eu-Eu
(1 P
Er-Er
)(1 P
Eu-Eu
)
P
Er-Er
+ P
Eu-Eu
P
Er-Er
P
Eu-Eu
3.+Average+time+required+to+distribute+entanglement+over+the+distance++
hT i
2
n
L
0
=
L
0
c
3
2
n1
1
P
n
s
P
0
=
L
0
c
1
P
t
2
n
L
0
Enhancing+the+rate+by+parallelizing+and+multiplexing
Success(probability(
for(m(parallel(links
P =1 (1 P
t
)
m
Rate+comparison:+
direct(transmission(
proposed(scheme(with(8(
elementary(links(and(
parallelized(version(for(m=100((
Here(we(assumed(
p =
d
= 90%.
OUTLOOK
Quantify(decoherence(effects(
Calculate(overall(fidelity(of(the(distributed(entangled(state(
Using(proposed(scheme(it(might(be(possible(to(design(a(
scalable(quantum(computing(architecture(for(rare-earth(ions
!
1
!
2
!
m
!
m1
!
1
!
2
!
m
!
m1
MUX
MUX
DEMUX
DEMUX
BS
Figure 3: Our multiplexed scheme consists of 2
n
+1 nodes that
span the total channel distance L. Each node consists of an
array of m cavities that emit photons which feature differing
carrier frequencies. The carrier frequency of each photon is de-
termined by a frequency translation device. A coupling element
(COUPLER) ensures that each photon traverses a common
channel to a Bell-state measurement station that consists of
a beam splitter, decoupling element (DECOUPLER), and 2m
single photon detectors. This set-up allows m entanglement
generation protocols to be operated in parallel.
that the direct transmission scheme can also be inter-
preted as the Pirandola bound [52] for a repetition rate
of 10/1.44=6.9 GHz. The performance of our proto-
col and the DLCZ scheme with m = 100 multiplexed
channels is also shown Fig. 4. For more information
on other approaches see Ref. [7]. In this review paper,
the multiplexed DLCZ outperformed the other repeater
protocols considered (see Fig. 18 of ref.[7]). As will
be discussed in Sec. IV B, the use of even more than
m = 100 parallel spectral channels is possible.
Fig. 4 shows that our use of deterministic gates is
advantageous for the rate. The scaling of the DLCZ
scheme with distance is slightly better than that of our
scheme due to the requirement of detecting only one
photon for each elementary entanglement creation step
(rather than two photons for our scheme). However,
this comes at the significant expense of requiring phase
stabilization for the long-distance fiber links.
4 Implementation
For our scheme, we consider ion beam-milled Y
2
SiO
5
photonic crystal cavity systems that have been weakly-
doped with Er
3+
and
151
Eu
3+
ions [38, 39, 53]. Co-
doped crystals may be grown from the melt [53], or in-
dividual RE ions may be implanted into single Y
2
SiO
5
crystals [54]. After a milling step, the output of the
cavity can be coupled to an optical waveguide using,
e.g., microscopy [55], bonding [56] or a pick-and-place
technique [57], with the latter having been used to het-
erogeneously interface InAs/InP quantum dots with Si
waveguides. Despite the lack of on-demand control of
Accepted in Quantum 2018-08-27, click title to verify 6
A
B
C
D
E
Figure 4: A comparison of the entanglement distribution rate
for various schemes. Direct transmission scheme using a 10
GHz single-photon source (E) is compared with original DLCZ
protocol (A) and our scheme (B) using 8 elementary links, each
of length L
0
. Also shown are the rates that correspond to the
multiplexed versions of DLCZ (C) as well as our scheme (D)
with m = 100 using the same number of elementary links. We
assume that the detection efficiency and the success probability
of emitting a single photon by an ion is p = η
d
= 0.9, and in the
case of DLCZ, a storage efficiency of η
m
= 0.9. The repetition
rate for each repeater protocol is set by the communication
time L
0
/c.
the position and relative orientation of each ion [54], a
suitable Er-Eu ion pair can be identified by performing
laser-induced fluorescence spectroscopy. This involves
exciting each Er
3+
(
151
Eu
3+
) ion with a laser and mea-
suring the spectral shift of the resultant fluorescence
from other
151
Eu
3+
(Er
3+
) ions due to the aforemen-
tioned electric-dipole coupling (see Sec. 2.3) [53]. Note
that measurements using a 0.02%Er:1%Eu:Y
2
SiO
5
bulk
crystal revealed the optical transition frequency of sets
of Er
3+
ions that lie within approximately one nanome-
ter from adjacent Eu
3+
ions [53]. These results suggest
that the transition frequencies of suitable Er-Eu pairs
can be rapidly distinguished from other spectator ions.
The magnetic field applied in the D
1
D
2
plane at
135 degrees relative to the D
1
axis results in the decay of
the excited Er
3+
ion back into the initial spin state via
spontaneous emission with a probability of higher than
90%[58], and a Zeeman-level lifetime of about 130ms
was measured for a magnetic field of 1.2 mT at 2.1 K
temperature[58]. At large external magnetic fields of 1
T or more and temperatures below 3 K, one-phonon di-
rect process is the dominant spin-relaxation mechanism
and the temperature and magnetic field dependence of
the relaxation rate could be approximated by [59]:
R(B) = R
0
+ α
D
g
3
B
5
coth(
gµ
B
B
2kT
) (9)
where α
D
is the anisotropic constant, µ
B
is the Bohr
magneton, and k is the Boltzmann constant. The field
independent contribution R
0
can be attributed to cross-
relaxations with paramagnetic impurities in the crystal
and hence depends strongly on the crystal purity. Ex-
trapolating from Refs [58, 59], the spin relaxation time
would be about 40ms at 20mK for an external magnetic
field of 1T at the 135
in D
1
D
2
plane. At this mag-
netic field, however, the splitting of the Zeeman levels
would be too large to address it with microwave pulses
and optical Raman pulses should be applied instead.
For an ensemble-doped Er
3+
:Y
2
SiO
5
the spin coher-
ence lifetime at mK temperatures and external mag-
netic fields of few hundred mT can be as short as 7 µs
[60]. However, for an ensemble-doped Er
3+
:Y
2
SiO
5
crystal spin flip-flop processes are one of the main
sources of decoherence, the spin coherence lifetime of
a single Er
3+
ion in Y
2
SiO
5
is largely determined by
spin-spin magnetic dipole interactions. The magnetic
moments of the constituent spins of Y
2
SiO
5
are small:
0.137µ
N
, 0.5µ
N
, and 1.89µ
N
for
89
Y,
29
Si, and
17
O respectively. Compared to yttrium ions,
29
Si and
17
O have low isotopic natural abundances, so we assume
the contribution of these isotopes to be negligible and
only consider the Er-Y interactions. In a large enough
magnetic field (a few hundred milli-tesla or more) com-
pared to the Er-Y coupling strength, the magnetic mo-
ment of Er
3+
will detune the closest-proximity Y ions
from resonance with those further away. This well-
known effect referred to as the “frozen core” has been
observed and results in weaker decoherence of the Er
3+
ion by nearby Y ions [6163].
Hence, for a single Er
3+
in Y
2
SiO
5
, the presence of a
strong magnetic field may increase the spin coherence
lifetime into the milliseconds range. To further increase
the spin coherence lifetime, dynamical decoupling im-
mediately after the second optical excitation step is
necessary. While this has not been demonstrated in
Er
3+
:Y
2
SiO
5
, it is a widely-employed method, and has
been used to extend coherence lifetimes in Ce-doped yt-
trium aluminum garnet [64] as well as nitrogen-vacancy
centers [42].
Spin polarization of up to 90% in Er
3+
:Y
2
SiO
5
has
been realized by using stimulated emission and spin-
mixing methods [65]. The efficiency of spin polarization
is determined by a competition between decay of the
ground-level population relative to the optical pumping
efficiency of the ion. Since the latter will be enhanced
due to the cavity-induced decay rate, we expect a near-
unity Er
3+
-spin polarization for our scheme. Since the
ground state lifetime of
151
Eu
3+
is several hours, spin
polarization can be achieved by performing continuous
optical pumping of all but one of the ground states.
In Y
2
SiO
5
, RE ions can occupy two crystallographi-
cally inequivalent sites with C
1
symmetry. Due to this
Accepted in Quantum 2018-08-27, click title to verify 7
lack of symmetry, the orientation of the dipole moments
and their magnitudes, and hence their dipole-dipole in-
teraction strengths, are unknown. Previous measure-
ments of the linear Stark shift of Er
3+
:Y
2
SiO
5
[66] al-
low the calculation of the projection of the electric-
dipole moment difference onto the direction of the
externally applied electric field to be approximately
0.84 × 10
31
Cm [67]. The dipole moment difference
for
151
Eu
3+
:Y
2
SiO
5
can be as high as µ
Eu
= 0.81 ×
10
31
Cm [68], resulting in the shift of the transition fre-
quency of 10 and 0.01 MHz for r
ij
= 1 and 10 nm, re-
spectively. As a comparison, the magnetic dipole-dipole
interaction between Er
3+
magnetic moment, which can
be as high as µ
Er
= 14.65 µ
B
[69], and
151
Eu magnetic
moment with its intrinsic value of µ
Eu
= 3.42 µ
N
[70]
is much weaker and amounts to 363 and 0.363 kHz for
r
ij
= 1 and 10 nm, respectively.
4.1 Cavity
The cavity serves three main purposes in this proposal.
It improves the quantum efficiency, enhances the single-
photon indistinguishability, and increases the rate of
Er
3+
emission. To achieve these benefits, the cavity
must provide a significant Purcell enhancement to the
|e
2
i |↑i transition of Er
3+
.
In the context of RE ions, Purcell factors of several
hundred have been achieved [34, 38], and up to 10
3
seems to be a reasonable goal [71]. The Purcell factor
can be written: P = (γ
r
)C, where 1 is the excited
state lifetime, γ
r
is the radiative decay rate, and C is
the cavity cooperativity. The cavity-enhanced quan-
tum efficiency (probability of emitting a photon into
the cavity mode) is then given by: p = ηP/(1 + ηP ),
where η = βγ
r
is the Er
3+
spin-conserving quantum
efficiency, and β is the probability of an excited ion
to relax into the initial spin state via the spontaneous
emission. For Er
3+
:Y
2
SiO
5
, β = 0.9, γ = 2π × 14 Hz,
and γ
r
= 2π × 3 Hz [45], resulting in η = 0.19. For
P = 100, a cavity quantum efficiency of p = 0.95 is pos-
sible. Increasing the Purcell factor to P = 1000 allows
p = 0.995.
The single-photon indistinguishability is a metric
that quantifies the quality of interference between pho-
tons that originate from the same quantum emitter.
Without a cavity, the single-photon indistinguishabilty
can be defined as I
1
= T
2
/2T
1
[72]. For an ensemble
of 0.0015% Er
3+
:Y
2
SiO
5
we have T
1
= 11.4ms [44] and
optical T
2
is around 200µs at 1 T and a temperature of
a few K [73]. This would imply I
1
= 0.009 without a
cavity, which would require significant spectral filtering
to achieve successful entanglement generation. With
a cavity, the single-photon indistinguishability can be
approximated by: I
1
= (1 + ηP )/(ζ + 1 + ηP ), where
ζ = 2T
1
/T
?
2
= 2T
1
/T
2
1 is the dephasing ratio. There-
fore, with Purcell factors P = 1000 and P = 20, 000 the
single-photon indistinguishability would be I
1
= 0.63
and I
1
= 0.97, respectively. In a large magnetic field,
however, T
2
= 4 ms was measured at a few K [73]. For
an ensemble-doped Er
3+
:Y
2
SiO
5
spin flip-flop processes
are the dominant decoherence mechanism and applying
a large magnetic field can freeze these processes. As a
result we have a long optical T
2
. The flip-flop process
can also be suppressed by reducing the Er
3+
concentra-
tion. Therefore in our system, where we assume very
low dopant concetration, we still can expect to have op-
tical coherence times of the order of a few ms even at
lower magnetic fields (hundreds of mT). With P = 100
and T
2
= 4ms, I
1
= 0.82 is possible, which could be fur-
ther improved by attempting to spectrally filter the nar-
row 2π × 1.4 kHz bandwidth photons. With P = 1000,
I
1
= 0.98 could be achieved without spectrally filtering
the 2π × 14 kHz photons (corresponding to a duration
of 11 µs).
4.2 CNOT gate
To perform the CNOT gate, it is necessary for Er
3+
to
remain excited for a time that is long enough to apply
three π-pulses to
151
Eu
3+
. This implies that each Er
3+
ion must be excited to a different Zeeman level (|e
1
i)
than the level that is coupled to the cavity (|e
2
i). This
can be done if the Zeeman splitting between |e
1
i and
|e
2
i is much larger than the cavity linewidth. In this
section, for Er
3+
we use |ei
Er
= |e
1
i.
To quantify the fidelity of the CNOT gate and state
transfer procedures, we use a model nine-level system
where both Er
3+
and
151
Eu
3+
are represented by three-
level systems with two lower-energy spin states (|↑i, |↓i)
and a single excited state (|ei) (see Fig. 2). We use the
dipole-interaction Hamiltonian:
H = ∆νσ
+
Er
σ
Er
σ
+
Eu
σ
Eu
+
X
k,l
k,l
2
σ
+
k,l
+ σ
k,l
,
(10)
where ν is the dipole interaction strength and
k,l
is
the Rabi frequency for transition {k, l} for l {↑, ↓},
k {Er, Eu}. Without loss of generality, we choose
ν > 0 and
k,l
> 0. The master equation is
˙ρ = i[H, ρ] +
X
k,l
γ
k,l
D(σ
k,l
)ρ +
γ
?
k
2
D(σ
+
k,l
σ
k,l
)ρ
+
X
k
χ
k
2
D(σ
z,k
)ρ,
(11)
where D(σ)ρ = σρσ
σ
σρ/2ρσ
σ/2. The operators
are defined as σ
+
k,l
= |eihl|
k
, σ
k,l
= (σ
+
k,l
)
, and σ
z,k
=
(|↑ih↑|
k
|↓ih↓|
k
)/2. The rate γ
k,l
is the decay rate for
Accepted in Quantum 2018-08-27, click title to verify 8
transition {k, l}, γ
?
k
is the optical pure dephasing rate,
and χ
k
is the spin decoherence rate for ion k.
To solve for the Er-Eu state after applying the five-
pulse CNOT gate sequence, we first assume that each
π-pulse is a square pulse and that they are applied to the
system sequentially with no time delay between pulses.
In this case, the time taken to apply each π-pulse is
given by T
k,l
= π/
k,l
. We also assume that
k,l
and
ν are much larger than any dissipative rate so that we
can treat the dissipation perturbatively.
We then define the zero-order (reversible) superop-
erator L
0
where L
0
ρ = i[H, ρ]. Then we define the
first-order (irreversible) perturbation superoperator L
1
as L
1
ρ = ˙ρ L
0
ρ. From this, we define the rotation
superoperator R
k,l
(θ) corresponding to the pulse where
k,l
6= 0 and
k
0
,l
0
= 0 for all k
0
6= k and l
0
6= l. This
superoperator is given by
R
k,l
(θ) = e
L
0
θ
k,l
"
1 +
Z
θ
k,l
0
e
−L
0
τ
L
1
e
L
0
τ
#
, (12)
which is accurate to first-order in γ
k
0
,l
0
, γ
?
k
0
, χ
k
0
k,l
, ν for all k
0
and l
0
.
To simulate the entanglement swapping between
151
Eu
3+
and Er
3+
, we begin with an initial superpo-
sition state ρ
0
where |ψi = (|↑i + |↓i)
Er
(|↑i + |↓i)
Eu
and ρ
0
= |ψihψ|. Then after applying the CNOT pulse
scheme, the final state is
ρ = R
Er,
(θ)R
Eu,
(θ)R
Eu,
(θ)R
Eu,
(θ)R
Er,
(θ)ρ
0
,
where ideally θ = π; however, we set θ = π + to
simulate a π-pulse with a small rotation deviation ||
π caused by a non-ideal pulse area. The total time taken
to apply this sequence is T
gate
= 2T
Er,
+2T
Eu,
+T
Eu,
.
For a high fidelity gate, the pulse durations must be
small so that the gate is fast and qubit states do not
dephase during the sequence. This favors larger Rabi
frequencies. On the other hand, it is also necessary that
the three Eu pulses do not excite the
151
Eu
3+
spin state
if Er
3+
is excited. One way to achieve this is to set
Eu,l
to satisfy the detuning condition
2
Eu,l
/ν
2
1 so that
there is little chance for off-resonant excitation. How-
ever, since ν cannot be arbitrarily large (the separa-
tion between Er and Eu cannot be arbitrarily small), the
detuning criteria necessitates a very slow gate, which
cannot provide a high fidelity.
Alternatively, the detuning condition can be circum-
vented if
Eu,l
is chosen so that the Eu pulses per-
form an effective 2π rotation on the Eu spin state when
Er is excited, but still perform a π-pulse when Er is
not excited. This can be accomplished by requiring
that
Eu,l
satisfies the effective Rabi frequency relation
q
ν
2
+
2
Eu,l
= 2Ω
Eu,l
. This sets
Eu,l
= ∆ν/
3. As
a consequence of fixing the Rabi frequency, an accurate
characterization of the dipole interaction strength for
each Er-Eu pair is necessary to achieve a high fidelity.
This is because any mischaracterization δν from the
true value ν δν will cause a deviation from the de-
sired 2π rotation. To account for this, we also consider a
perturbation of the fidelity for deviation ν ν δν
where we assume |δν| ν
k,l
.
The effective 2π pulse leaves a relative phase between
the
151
Eu
3+
ground-state spins of ϕ = π(2
3)/2.
If Er
3+
is in |↓i, then only π-rotations are applied to
151
Eu
3+
. Hence, in this case, there is no acquired rel-
ative phase from the detuned pulses. However, if the
state is |↑↓i,
151
Eu
3+
will be affected by the third pulse
performing an effective 2π rotation, and so |↑↓i will ac-
quire a relative phase of ϕ. Likewise, if the state is
|↑↑i,
151
Eu
3+
will be affected by the second and fourth
pulses performing an effective 2π rotation, and so |↑↑i
will acquire a relative phase of 2ϕ. Hence, in the ab-
sence of dissipation and using perfect square π-pulses,
the expected final state is
|ψ
f
i =
1
2
|↓↓i + |↓↑i+ e
|↑↓i + e
i2ϕ
|↑↑i
(13)
Since the expected acquired phase ϕ = π(2
3)/2
is known, and independent of the dipole interaction
strength ν, it can be tracked or corrected and so we
use |ψ
f
i as the final state when calculating the fidelity.
We use the above expressions for ρ and |ψ
f
i to com-
pute the fidelity F
CNOT
= |hψ
f
|ρ |ψ
f
i|. For simplicity,
we choose to set
Er,l
= Ω
Eu,l
= Ω = ∆ν/
3; however,
Er,l
is not restricted by the dipole interaction strength
and could be made larger than to further decrease
the total gate time and increase fidelity. The solution
F
CNOT
to first-order in γ
k,l
, γ
?
k
, χ
k
ν for all
k, l and second-order in π and ξ = δν/ν 1 is
F
CNOT
' 1 T
CNOT
Γ
2
13π
16
ξ
43π
2
128
ξ
2
(14)
Here T
CNOT
= 5π/Ω = 5π
3/ν is the total gate time
and Γ is the effective dissipation rate:
Γ '
1
80
[31γ
Er
+ 17γ
?
Er
+ 8χ
Er
+ 11γ
Eu
+ 8γ
?
Eu
+ 17χ
Eu
] ,
(15)
where we define γ
k
= γ
k
+ γ
k
. To obtain the reverse
CNOT fidelity, it is only necessary to swap the dissipa-
tive parameters for Er
3+
and
151
Eu
3+
.
For a dipole interaction strength of ν = 2π × 46
kHz corresponding to an Er-Eu separation of about 6
nm, the CNOT gate time can be as small as T
CNOT
=
94 µs. To estimate the fidelity, we use the parameters
γ
Er
= 2π × 3 Hz, γ
?
Er
= 2π × 8 Hz, γ
Eu
= 2π × 1.3
Hz, and γ
?
Eu
= 2π × 19 Hz estimated from Ref. [45];
Accepted in Quantum 2018-08-27, click title to verify 9
also χ
Er
' 2π × 80 Hz and χ
Eu
' 0. In this case,
for a small π-pulse over-rotation of ' π/64 and a
dipole interaction strength over-estimation of 2% (ξ '
0.02), the fidelity is F
CNOT
= 0.986. The reverse CNOT
fidelity is slightly smaller due to the spin dephasing of
Er: F
R-CNOT
= 0.980.
This method can also be used to compute the fidelity
of state transfer from Er
3+
to
151
Eu
3+
. In this case,
we remove the fourth pulse (R
Eu,
) and use the initial
state |ψi = (|↑i + |↓i)
Er
|↑i
Eu
. The expected final state
is |ψ
f
i = |↓↓i + e
i2ϕ
|↑↑i. In this case, the fidelity is
similar to the CNOT gate (Eq. (14)):
F
ST
' 1 T
ST
Γ
5
8
2
3π
16
ξ
21π
2
256
ξ
2
(16)
where T
ST
= 4π/ = 4π
3/ν. The effective dissipa-
tion rate was found to be very similar to Eq. (15), but
depends less strongly on
151
Eu
3+
dissipation:
Γ '
1
80
[29γ
Er
+ 16γ
?
Er
+ 9χ
Er
+ 11γ
Eu
+ 7γ
?
Eu
+ 9χ
Eu
] .
(17)
Using the same parameters as above, the fidelity of state
transfer is F
ST
= 0.989.
4.3 State measurement
The spin readout of Er
3+
is performed by optical exci-
tation. Since only the |↑i|e
2
i transition is coupled to
the cavity, optical excitation will result in a presence or
absence of a photon emission depending on the state of
the Er
3+
ion. The measurement of the
151
Eu
3+
ion can
be performed through its nearby Er
3+
ion as a readout
ion, provided that the Er
3+
ion is initially prepared in
the |↑i state. Performing spin readout of the
151
Eu
3+
ion in the Z basis is then achieved by optically exciting
the
151
Eu
3+
ion (|↑i to |ei transition) followed by ex-
citing the Er
3+
ion (|↑i to |e
1
i transition). The Er
3+
ion will excite to |e
1
i (remain in (|↑i) if the state of the
151
Eu
3+
ion is |↓i (|↑i) due to the permanent electric
dipole-dipole interaction which shifts the Er
3+
optical
resonance. To readout in the X basis, a
π
2
MW pulse
should be applied to the
151
Eu
3+
ion in order to rotate
the ground-state spins before the optical excitation step.
State measurement of each Er
3+
ion requires the de-
tection of an emitted photon. Due to the Purcell effect,
as discussed in Sec. 4.1, the emission of a single photon
from the ion is highly preferential into the cavity mode.
For a high quality cavity, the probability that the pho-
ton emits into the cavity tends to unity. Therefore, the
detection probability will be limited by coupling losses
and single photon detectors, which can have detection
efficiency as high as 95%, as has been demonstrated in
superconducting detectors [74, 75]. To do better than
this limit, it is necessary to pump the Er
3+
ion into a
cycling transition such that many photons will be emit-
ted by the cavity, and eventually detected. Using such a
cycling transition, detection probability can be as high
as 98.7% [71]. The detection efficiency is not 100% in
this case because there is a small chance for the ion to
decay into a different state than the initial state, thus
ending the photon cycling [76]. This chance grows lin-
early with the number of cycles before detection.
4.4 Spectrally-multiplexed implementation
Our spectral multiplexing scheme relies on the possibil-
ity that many spectral channels can be operated in par-
allel. This requires that, in one node, different cavities
emit photons that feature different carrier frequencies.
This can be accomplished by frequency translation.
Noise-free translation over tens of gigahertz [50, 77]
can be achieved by using voltage-swept, commercially-
available, waveguide electro-optic modulators that can
be optically-coupled to the output port of each cavity.
After frequency translation, the output can be coupled
to a common spatial mode (e.g. a waveguide or fiber)
by using a tunable ring resonator filter that features
resonance linewidths as narrow as 1 MHz [78]. Arrayed
waveguides or fiber-Bragg gratings may also be used,
however they are bulky and their resonance linewidths
are currently not at the MHz level. The modulators
and filters may be fabricated on a single chip, offer-
ing the possibility of low loss and up to 10
4
spectral
modes. The Bell-state measurement station consists of
a beamsplitter, two sets of ring resonator filters which
are identical to that used at the nodes, and an array of
superconducting nanowire-based photon detectors, cho-
sen due to their combination of high-efficiency and low
noise properties [75].
5 Conclusion
Our proposal for a quantum repeater which is based on
individual RE ions promises the deterministic establish-
ment of high-fidelity entanglement over long-distances
at a rate which exceeds that corresponding to the di-
rect transmission of photons. Our scheme utilizes some
of the most desirable features of RE-ion-doped crys-
tals, specifically emission within the low-loss telecom-
munications window (Er
3+
) and the hours-long nuclear
spin coherence lifetime (
151
Eu
3+
:Y
2
SiO
5
) that is needed
to perform long-distance transmission and swapping of
entanglement. Moreover, control logic gates between
close-lying individual
151
Eu
3+
and Er
3+
ions allow the
quasi-deterministic swapping of entanglement by means
of a permanent electric dipole-dipole interaction. The
multiplexed version of our scheme improves the entan-
glement distribution rate by at least a factor of 100 over
Accepted in Quantum 2018-08-27, click title to verify 10
that of the single-mode version of our repeater.
Looking forward, it is interesting to consider the pos-
sibility of employing individual
167
Er
3+
ions instead of
Er-Eu ion pairs for a telecommunication wavelength
quantum repeater. In the presence of strong magnetic
fields,
167
Er
3+
:Y
2
SiO
5
features a nuclear spin coherence
lifetime in the one-second range [79], allowing the pos-
sibility of entanglement generation and storage using
the same ion, pairs (or small ensembles) of Er
3+
ions.
One of the main challenges for future work in this direc-
tion is to devise a scheme whereby individual Er
3+
ions
may be addressed and coupled within a single cavity.
This could be achieved by using magnetic dipole-dipole
interactions in a similar spirit to what has been demon-
strated using nitrogen vacancy centers and carbon spins
in diamond [42], or by using the cavity mode to mediate
the interaction. Another interesting direction is the pos-
sibility of long-term storage using host-ion spins such as
yttrium in Y
2
SiO
5
[80].
ACKNOWLED GMENTS
The authors would like to thank M. Afzelius, P. Barclay,
P. A. Bushev, C. W. Thiel and W. Tittel for helpful
discussions. This work was supported by the Natural
Sciences and Engineering Research Council (NSERC)
through its Discovery Grant program, the CREATE
grant ’Quanta’, and through graduate scholarships, by
the University of Calgary through an Eyes High post-
doctoral fellowship, by Alberta Innovates Technology
Futures (AITF) through graduate scholarships, and
by the Defense Advanced Research Projects Agency
(DARPA) through the Quiness program (Contract No.
W31P4Q-13-1-0004).
References
[1] T. Jennewein and B. Higgins, Physics World 26,
52 (2013).
[2] P. Komar, E. M. Kessler, M. Bishof, L. Jiang, A. S.
Sørensen, J. Ye, and M. D. Lukin, Nat. Phys. 10,
582 (2014).
[3] D. Gottesman, T. Jennewein, and S. Croke, Phys.
Rev. Lett. 109, 070503 (2012).
[4] H. J. Kimble, Nature 453, 1023 (2008).
[5] C. Simon, Nat. Photonics 11, 678 (2017).
[6] H.-J. Briegel, W. ur, J. I. Cirac, and P. Zoller,
Phys. Rev. Lett. 81, 5932 (1998).
[7] N. Sangouard, C. Simon, H. de Riedmatten, and
N. Gisin, Rev. Mod. Phys. 83, 33 (2011).
[8] W. P. Grice, Phys. Rev. A 84, 042331 (2011).
[9] S. Wein, K. Heshami, C. A. Fuchs, H. Krovi,
Z. Dutton, W. Tittel, and C. Simon, Phys. Rev.
A 94, 032332 (2016).
[10] N. Sangouard, R. Dubessy, and C. Simon, Phys.
Rev. A 79, 042340 (2009).
[11] S. Ritter, C. Nolleke, C. Hahn, A. Reiserer,
A. Neuzner, M. Uphoff, M. Mucke, E. Figueroa,
J. Bochmann, and G. Rempe, Nature 484, 195
(2012).
[12] A. Reiserer and G. Rempe, Rev. Mod. Phys. 87,
1379 (2015).
[13] H. Bernien, B. Hensen, W. Pfaff, G. Koolstra,
M. Blok, L. Robledo, T. Taminiau, M. Markham,
D. Twitchen, L. Childress, et al., Nature 497, 86
(2013).
[14] B. Hensen, H. Bernien, A. E. Dreau, A. Reiserer,
N. Kalb, M. S. Blok, J. Ruitenberg, R. F. L. Ver-
meulen, R. N. Schouten, C. Abellan, W. Amaya,
V. Pruneri, M. W. Mitchell, M. Markham, D. J.
Twitchen, D. Elkouss, S. Wehner, T. H. Taminiau,
and R. Hanson, Nature 526, 682 (2015).
[15] D. L. Moehring, P. Maunz, S. Olmschenk, K. C.
Younge, D. N. Matsukevich, L. M. Duan, and
C. Monroe, Nature 449, 68 (2007).
[16] L. Slodiˇcka, G. H´etet, N. ock, P. Schindler,
M. Hennrich, and R. Blatt, Phys. Rev. Lett. 110,
083603 (2013).
[17] K. J. Morse, R. J. Abraham, A. DeAbreu, C. Bow-
ness, T. S. Richards, H. Riemann, N. V. Abrosi-
mov, P. Becker, H.-J. Pohl, M. L. Thewalt, et al.,
Science advances 3, e1700930 (2017).
[18] A. Delteil, Z. Sun, W. Gao, E. Togan, S. Faelt, and
A. Imamoglu, Nat. Phys. 12, 218 (2016).
[19] R. Stockill, M. J. Stanley, L. Huthmacher,
E. Clarke, M. Hugues, A. J. Miller, C. Matthiesen,
C. Le Gall, and M. Atat¨ure, Phys. Rev. Lett. 119,
010503 (2017).
[20] W. Tittel, M. Afzelius, T. Chaneliere, R. L. Cone,
S. Kr¨oll, S. A. Moiseev, and M. Sellars, Laser &
Photonics Rev. 4, 244 (2010).
[21] F. Bussi`eres, N. Sangouard, M. Afzelius,
H. de Riedmatten, C. Simon, and W. Tit-
tel, J. Mod. Opt. 60, 1519 (2013).
[22] S. Wu, G. Han, D. J. Milliron, S. Aloni, V. Altoe,
D. V. Talapin, B. E. Cohen, and P. J. Schuck,
Proc. Natl. Acad. Sci. 106, 10917 (2009).
[23] Y. Chu, N. de Leon, B. Shields, B. Hausmann,
R. Evans, E. Togan, M. J. Burek, M. Markham,
A. Stacey, A. Zibrov, A. Yacoby, D. Twitchen,
M. Loncar, H. Park, P. Maletinsky, and M. Lukin,
Nano Letters 14, 1982 (2014).
[24] D. Brunner, B. D. Gerardot, P. A. Dalgarno,
G. W¨ust, K. Karrai, N. G. Stoltz, P. M. Petroff,
and R. J. Warburton, Science 325, 70 (2009).
Accepted in Quantum 2018-08-27, click title to verify 11
[25] R. G. Neuhauser, K. T. Shimizu, W. K. Woo, S. A.
Empedocles, and M. G. Bawendi, Phys. Rev. Lett.
85, 3301 (2000).
[26] R. de Sousa and S. Das Sarma, Phys. Rev. B 68,
115322 (2003).
[27] K. Xia, R. Kolesov, Y. Wang, P. Siyushev,
R. Reuter, T. Kornher, N. Kukharchyk, A. D.
Wieck, B. Villa, S. Yang, and J. Wrachtrup, Phys.
Rev. Lett. 115, 093602 (2015).
[28] B. Lauritzen, J. Min´aˇr, H. De Riedmatten,
M. Afzelius, N. Sangouard, C. Simon, and
N. Gisin, Phys. Rev. Lett. 104, 080502 (2010).
[29] M. Afzelius, I. Usmani, A. Amari, B. Lauritzen,
A. Walther, C. Simon, N. Sangouard, J. Min´aˇr,
H. De Riedmatten, N. Gisin, et al., Phys. Rev.
Lett. 104, 040503 (2010).
[30] G. Heinze, C. Hubrich, and T. Halfmann, Phys.
Rev. Lett. 111, 033601 (2013).
[31] C. Laplane, P. Jobez, J. Etesse, N. Gisin, and
M. Afzelius, Phys. Rev. Lett. 118, 210501 (2017).
[32] M. Zhong, M. P. Hedges, R. L. Ahlefeldt, J. G.
Bartholomew, S. E. Beavan, S. M. Wittig, J. J.
Longdell, and M. J. Sellars, Nature 517, 177
(2015).
[33] T. Zhong, J. M. Kindem, J. G. Bartholomew,
J. Rochman, I. Craiciu, V. Verma, S. W. Nam,
F. Marsili, M. D. Shaw, A. D. Beyer, et al., arXiv
preprint arXiv:1803.07520 (2018).
[34] A. M. Dibos, M. Raha, C. M. Phenicie, and J. D.
Thompson, Phys. Rev. Lett. 120, 243601 (2018).
[35] R. Kolesov, K. Xia, R. Reuter, R. St¨ohr, A. Zappe,
J. Meijer, P. Hemmer, and J. Wrachtrup, Nat.
Commun. 3, 1029 (2012).
[36] C. Yin, M. Rancic, G. G. de Boo, N. Stavrias, J. C.
McCallum, M. J. Sellars, and S. Rogge, Nature
497, 91 (2013).
[37] T. Utikal, E. Eichhammer, L. Petersen, A. Renn,
S. otzinger, and V. Sandoghdar, Nat. Commun.
5, 3627 (2014).
[38] T. Zhong, J. M. Kindem, E. Miyazono, and
A. Faraon, Nat. Commun. 6, 8206 (2015).
[39] T. Zhong, J. M. Kindem, J. G. Bartholomew,
J. Rochman, I. Craiciu, E. Miyazono, M. Bettinelli,
E. Cavalli, V. Verma, S. W. Nam, F. Marsili, M. D.
Shaw, A. D. Beyer, and A. Faraon, Science 357,
1392 (2017).
[40] J. J. Longdell and M. J. Sellars, Phys. Rev. A 69,
032307 (2004).
[41] J. J. Longdell, M. J. Sellars, and N. B. Manson,
Phys. Rev. Lett. 93, 130503 (2004).
[42] A. Reiserer, N. Kalb, M. S. Blok, K. J. van Bemme-
len, T. H. Taminiau, R. Hanson, D. J. Twitchen,
and M. Markham, Phys. Rev. X 6, 021040 (2016).
[43] S. D. Barrett and P. Kok, Phys. Rev. A 71, 060310
(2005).
[44] G. Liu and B. Jacquier, Spectroscopic properties of
rare earths in optical materials, Vol. 83 (Springer
Science & Business Media, 2006).
[45] D. McAuslan, J. J. Longdell, and M. Sellars, Phys.
Rev. A 80, 062307 (2009).
[46] N. Ohlsson, R. K. Mohan, and S. Kr¨oll, Opt. Com-
mun. 201, 71 (2002).
[47] S. Altner, G. Zumofen, U. Wild, and M. Mit-
sunaga, Phys. Rev. B 54, 17493 (1996).
[48] C. Simon, H. de Riedmatten, M. Afzelius, N. San-
gouard, H. Zbinden, and N. Gisin, Phys. Rev. Lett.
98, 190503 (2007).
[49] O. A. Collins, S. D. Jenkins, A. Kuzmich, and
T. A. B. Kennedy, Phys. Rev. Lett. 98, 060502
(2007).
[50] N. Sinclair, E. Saglamyurek, H. Mallahzadeh, J. A.
Slater, M. George, R. Ricken, M. P. Hedges,
D. Oblak, C. Simon, W. Sohler, and W. Tittel,
Phys. Rev. Lett. 113, 053603 (2014).
[51] L.-M. Duan, M. Lukin, J. I. Cirac, and P. Zoller,
Nature 414, 413 (2001).
[52] S. Pirandola, R. Laurenza, C. Ottaviani, and
L. Banchi, Nat. Commun. 8 (2017).
[53] C. W. Thiel, W. R. Babbitt, and R. L. Cone, Phys.
Rev. B 85, 174302 (2012).
[54] T. Kornher, K. Xia, R. Kolesov, N. Kukharchyk,
R. Reuter, P. Siyushev, R. St¨ohr, M. Schreck,
H. W. Becker, B. Villa, A. D. Wieck, and
J. Wrachtrup, Appl. Phys. Lett. 108, 053108
(2016).
[55] M. Davanco, J. Liu, L. Sapienza, C.-Z. Zhang, J. V.
De Miranda Cardoso, V. Verma, R. Mirin, S. W.
Nam, L. Liu, and K. Srinivasan, Nat. Commun. 8,
889 (2017).
[56] E. Murray, D. J. P. Ellis, T. Meany, F. F. Floether,
J. P. Lee, J. P. Griffiths, G. A. C. Jones, I. Farrer,
D. A. Ritchie, A. J. Bennett, and A. J. Shields,
Appl. Phys. Lett. 107, 171108 (2015).
[57] J.-H. Kim, S. Aghaeimeibodi, C. J. K. Richardson,
R. P. Leavitt, D. Englund, and E. Waks, Nano
Letters 17, 7394 (2017).
[58] S. R. Hastings-Simon, B. Lauritzen, M. U. Staudt,
J. L. M. van Mechelen, C. Simon, H. de Riedmat-
ten, M. Afzelius, and N. Gisin, Phys. Rev. B 78,
085410 (2008).
[59] T. ottger, C. W. Thiel, Y. Sun, and R. L. Cone,
Phys. Rev. B 73, 075101 (2006).
[60] S. Probst, H. Rotzinger, A. V. Ustinov, and P. A.
Bushev, Phys. Rev. B 92, 014421 (2015).
[61] O. Guillot-No¨el, H. Vezin, P. Goldner, F. Beau-
doux, J. Vincent, J. Lejay, and I. Lorger´e, Phys.
Rev. B 76, 180408 (2007).
Accepted in Quantum 2018-08-27, click title to verify 12
[62] E. Fraval, M. J. Sellars, A. Morrison, and A. Ferris,
J. Lumin. 107, 347 (2004).
[63] R. M. Macfarlane and R. M. Shelby, Spectroscopy
of Solids Containing Rare Earth Ions, edited by
A. A. Kaplyanskii and R. M. Macfarlane (North
Holland, Amsterdam, 1987).
[64] P. Siyushev, K. Xia, R. Reuter, M. Jamali,
N. Zhao, N. Yang, C. Duan, N. Kukharchyk,
A. Wieck, R. Kolesov, et al., Nat. Commun. 5
(2014).
[65] B. Lauritzen, S. R. Hastings-Simon, H. de Ried-
matten, M. Afzelius, and N. Gisin, Phys. Rev. A
78, 043402 (2008).
[66] J. Min´aˇr, B. Lauritzen, H. de Riedmatten,
M. Afzelius, C. Simon, and N. Gisin, New J. Phys.
11, 113019 (2009).
[67] R. M. Macfarlane, J. Lumin. 125, 156 (2007),
festschrift in Honor of Academician Alexander A.
Kaplyanskii.
[68] J. H. Wesenberg, K. Mølmer, L. Rippe, and
S. Kr¨oll, Phys. Rev. A 75, 012304 (2007).
[69] Y. Sun, T. ottger, C. W. Thiel, and R. L. Cone,
Phys. Rev. B 77, 085124 (2008).
[70] F. M. Pichanick, P. G. H. Sandars, and G. K.
Woodgate, Proc. Royal Soc. A 257, 277 (1960).
[71] C. O’Brien, T. Zhong, A. Faraon, and C. Simon,
Phys. Rev. A 94, 043807 (2016).
[72] T. Grange, G. Hornecker, D. Hunger, J.-P. Poizat,
J.-M. G´erard, P. Senellart, and A. Auff`eves, Phys.
Rev. Lett. 114, 193601 (2015).
[73] T. ottger, C. W. Thiel, R. L. Cone, and Y. Sun,
Phys. Rev. B 79, 115104 (2009).
[74] A. E. Lita, A. J. Miller, and S. W. Nam, Opt.
Express 16, 3032 (2008).
[75] F. Marsili, V. B. Verma, J. A. Stern, S. Harring-
ton, A. E. Lita, T. Gerrits, I. Vayshenker, B. Baek,
M. D. Shaw, R. P. Mirin, et al., Nat. Photonics 7,
210 (2013).
[76] J. G. Bartholomew, R. L. Ahlefeldt, and M. J.
Sellars, Phys. Rev. B 93, 014401 (2016).
[77] M. Grimau Puigibert, G. H. Aguilar, Q. Zhou,
F. Marsili, M. D. Shaw, V. B. Verma, S. W. Nam,
D. Oblak, and W. Tittel, Phys. Rev. Lett. 119,
083601 (2017).
[78] K. Y. Yang, D. Y. Oh, S. H. Lee, Q.-F. Yang, X. Yi,
B. Shen, H. Wang, and K. Vahala, Nat. Photonics
12, 297 (2018).
[79] M. Ranˇci´c, M. P. Hedges, R. L. Ahle-
feldt, and M. J. Sellars, Nat. Phys. (2017),
10.1038/nphys4254.
[80] B. Car, L. Veissier, A. Louchet-Chauvet, J.-L.
Le Gou¨et, and T. Chaneli`ere, Phys. Rev. Lett.
120, 197401 (2018).
Accepted in Quantum 2018-08-27, click title to verify 13