Resource requirements for efficient quantum communication
using all-photonic graph states generated from a few matter
qubits
Paul Hilaire, Edwin Barnes, and Sophia E. Economou
Department of Physics, Virginia Tech, Blacksburg, Virginia 24061, USA
Quantum communication technologies
show great promise for applications rang-
ing from the secure transmission of secret
messages to distributed quantum comput-
ing. Due to fiber losses, long-distance
quantum communication requires the use
of quantum repeaters, for which there ex-
ist quantum memory-based schemes and
all-photonic schemes. While all-photonic
approaches based on graph states gen-
erated from linear optics avoid coher-
ence time issues associated with memo-
ries, they outperform repeater-less pro-
tocols only at the expense of a pro-
hibitively large overhead in resources.
Here, we consider using matter qubits to
produce the photonic graph states and an-
alyze in detail the trade-off between re-
sources and performance, as character-
ized by the achievable secret key rate
per matter qubit. We show that fast
two-qubit entangling gates between mat-
ter qubits and high photon collection and
detection efficiencies are the main ingre-
dients needed for the all-photonic proto-
col to outperform both repeater-less and
memory-based schemes.
The ability to share entangled states over long
distances is a major milestone for the realiza-
tion of a fully-functional quantum internet [1, 2].
Beyond secure communications via quantum key
distribution (QKD) [35], the implementation of
such a quantum internet would also have various
applications ranging from distributed quantum
computing [6], secure access to a remote quan-
tum computer [7, 8], accurate clock synchroniza-
tion [9], and improved telescope observations [10].
However, enabling world-wide quantum com-
munication requires addressing the major prob-
Paul Hilaire: paulhilaire@vt.edu
lem of photonic losses, which significantly reduces
the range of quantum information transfer. Even
though direct amplification of a quantum state
is made impossible due to the no-cloning theo-
rem [11, 12], this exponential photon loss can still
be overcome through the realization of quantum
repeaters (QR) [13, 14].
Several QR approaches have been proposed
and can be divided into two main categories de-
pending on the method used to propagate quan-
tum information between two adjacent repeater
nodes [15]. The first approach proposed, re-
ferred to here as a "memory-based" approach,
relies on heralded-entanglement generation [16
24] between two quantum memories (QM) situ-
ated at adjacent repeater nodes. Examples of
memory-based schemes are depicted schemati-
cally in Figs. 1(a),(b). The heralded entangle-
ment generally succeeds when a detector situated
at an intermediate node measures photons emit-
ted by the QMs [2527]. A classical signal car-
rying the outcome of an entanglement generation
attempt must be sent back to the QMs, thus lim-
iting the repetition rate of these protocols. When
successful, entanglement swapping transfers the
entanglement through these repeater nodes.
The second category of repeater [2834] relies
on logically-encoded multi-photon states [32, 35],
resistant both to photonic losses and errors, to
transfer quantum information across a network.
One particularly promising approach of this type
was put forward in Ref. [31], which proposed an
all-optical QR protocol based on repeater graph
states (RGS). An example of an RGS with logical
encoding is shown in Fig. 1(c), while the corre-
sponding communication protocol is summarized
in Fig. 1(d) and will be detailed later. Ref. [31]
also proposed a method to construct these states
probabilistically using single-photon sources, lin-
ear optics, and detectors [36]. However, this
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 1
arXiv:2005.07198v4 [quant-ph] 11 Feb 2021
L
0
L
0
Alice Bob
QR node
Measurement
node
(a) 2-quantum-memory protocol
L
0
Classical channel
Photon
Entanglement
Swapping
HEG
Multiplexed QR
(b) Multiplexed QR
2N
=
b
0
=3
b
1
=2
m=3
(c) RGS
(d) RGS protocol
Measurement
node
RGS node
L
0
L
0
Matter qubit
Alice Bob
Matter qubit
Figure 1: (a) 2-quantum-memory QR protocol. (b) Multiplexed quantum repeater composed of 2N memories per
repeater with heralded entanglement generation (HEG) at measurement nodes. (c) Left side: Example of a repeater
graph state. Vertices represent qubits (blue: physical qubits, orange: logical qubits) while edges represent CZ gates.
Right side: Example of logical encoding using a depth-3 tree graph. (d) Quantum communication scheme using the
RGS protocol. Matter qubits are also included.
generation procedure requires about 10
6
single-
photon sources per repeater node to just barely
outperform direct fiber transmission [37]. These
findings suggest that the RGS protocol may be
well beyond the reach of current and near-future
technological capabilities.
However, it remains unclear whether the RGS
protocol can become more feasible if alternative
methods to create the states are used. Ref. [38]
proposed a deterministic generation procedure to
produce a linear graph state using one quantum
emitter which emits spin-entangled photons [39];
this procedure was later demonstrated experi-
mentally [40]. Ref. [41] showed theoretically
that emitters that undergo entangling gates can
also be used to generate two-dimensional cluster
states. Refs. [42, 43] built on this protocol and
demonstrated that entanglement between emit-
ters can be harnessed for the generation of more
complex photonic graph states. These ideas have
been extended further to general prescriptions
and to protocols tailored to specific physical sys-
tems [4447]. Ref. [42] introduced a protocol for
producing an arbitrary-sized RGS using only a
few matter qubits, thus significantly decreasing
the resource overhead required for RGS genera-
tion, making it deterministic in principle. Us-
ing this generation technique, the RGS protocol
only necessitates a few-qubit processor at each
repeater node, which would ease its practical im-
plementation compared to other error-correction-
based proposals that generally require several
hundreds of qubits per repeater node [29, 30]
(with the notable exception of Ref. [48] which
uses techniques introduced in Ref. [42] and re-
quires deterministic spin-photon Bell measure-
ments). Although the RGS protocol with de-
terministic state generation seems promising, a
systematic and detailed evaluation of its perfor-
mance and resource requirements has not been
carried out.
In this paper, we compare the resource-
efficiency—characterized by the achievable secret
key rate per matter qubit—of this protocol to di-
rect fiber transmission and to QR schemes based
on memories and heralded entanglement gener-
ation. We first show that the rate per matter
qubit has a fundamental upper bound in the case
of memory-based QRs. We then review the RGS
protocol and how RGSs can be generated using a
few matter qubits. We evaluate the performance
of this scheme, show that its rate per matter qubit
does not have a theoretical upper-bound, and find
the conditions under which it outperforms both
the repeater-less and the memory-based QR ap-
proaches. These conditions depend sensitively on
the speed with which two-qubit gates between the
matter qubits can be executed and on the col-
lection and detection efficiencies of the photons
emitted by these matter qubits.
1 Upper bound on rate for memory-
based repeater schemes
In this section, we show that there is a theoret-
ical upper bound, R
(QM)
max
, on the rate per mat-
ter qubit for protocols based on quantum mem-
ories and heralded entanglement generation. In
such protocols, the total distance L between Al-
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 2
ice and Bob is divided into smaller distances L
0
by N
QR
= L/L
0
1 repeater nodes. Quantum
memories at adjacent repeater nodes are entan-
gled via a heralded entanglement procedure (see
Fig. 1(a)). When a repeater node shares entan-
glement connections with its two adjacent nodes,
entanglement swapping is performed on the two
memory qubits within that node to create a di-
rect entanglement connection between memories
on the adjacent nodes. This procedure can be
repeated until Alice and Bob share an entangled
qubit pair.
It is clear that creating an entangled pair
between Alice and Bob requires generating an
entanglement connection between each adjacent
pair of repeaters. This means that the overall pro-
tocol rate R
(QM)
is limited by the entanglement
generation rate hT
ent
i
1
between two adjacent re-
peaters. Here, we use this fact to derive an upper
bound on the rate per matter qubit for QR pro-
tocols that use quantum memories and heralded
entanglement generation.
To determine an upper bound on hT
ent
i
1
, we
focus for concreteness on the protocol presented
in Fig. 1(a), which uses two quantum memories
per node. A quantum memory should emit a pho-
ton that is maximally entangled with one of its
degrees of freedom. Two photons generated at
adjacent repeater nodes arrive at the same mea-
surement node situated halfway between the two
repeaters, where they are measured in a Bell state
basis. Because a photon Bell state measurement
using only linear optics succeeds with probabil-
ity at best 1/2 (without ancillary qubits or QND
measurements [4954]), the overall success prob-
ability of the distant heralded entanglement gen-
eration is P
ent
1/2. It is worth mentioning
that a method for achieving heralded entangle-
ment generation with higher success probability
has been proposed [55], but its efficacy is re-
stricted to qubits separated by a short distance,
so we exclude this from our analysis (see Supple-
mentary Materials for more details). A classical
signal must then inform the repeater nodes of the
success or failure of the Bell state measurement.
Because the distance from the measurement node
to the repeater nodes is L
0
/2, this means that
the overall heralded entanglement generation at-
tempt takes total time T
trial
L
0
/c. This in-
cludes the time L
0
/(2c) for the single-photon
transfer from the repeater to the measurement
Notation Definition
L Total distance between Alice and Bob.
L
0
Distance between adjacent nodes.
N
QR
Number of repeater nodes.
2m Number of arms of the RGS.
~
b Branching vector of an error-correction
tree (
~
b = (b
0
, b
1
, . . . , b
n1
)).
T
CZ
CZ gate time.
η
t
(l) Transmission of a fiber of length l (η
t
(l) =
exp(l/L
att
)).
η
c
, η
d
In-fiber collection and detection efficien-
cies of photons.
L
att
, c Attenuation distance of the fiber and
speed of light in a fiber. (L
att
20km)
t
att
Average time of flight of photons in the
fiber (t
att
= L
att
/c).
T
RGS
Generation time of an RGS.
R, R
m
Rate and rate per matter qubit of the pro-
tocol.
Single-photon error rate.
Table 1: Table of notations
node and also the time L
0
/(2c) for the classical
signaling in the opposite direction. Here, we are
neglecting the time it takes to prepare and pump
the quantum memories. Throughout this proto-
col, a QM can be maximally entangled with at
most one other qubit (either a photon or another
QM). Therefore, it cannot emit another spin-
entangled photon before receiving the classical
signal carrying the information about the success
or failure of the Bell measurement, hence limiting
the repetition rate of the protocol. To generate
an entanglement connection between the two ad-
jacent nodes, the procedure must be repeated on
average P
1
ent
times. The entanglement generation
rate is therefore hT
ent
i
1
= P
ent
/T
trial
< c/(2L
0
).
We derived this result for a specific heralded en-
tanglement generation protocol but it also holds
for all known protocols [2527] (see Supplemen-
tary Materials).
From these results, we can show that the rate
per matter qubit (where the number of matter
qubits is N
m
= 2(N
QR
+ 1) = 2L/L
0
) has a the-
oretical upper bound, R
(QM)
max
:
R
(QM)
N
m
hT
ent
i
1
N
m
c
4L
= R
(QM)
max
. (1)
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 3
This theoretical upper bound also holds if there
are more than two QMs at each repeater node
(see Fig. 1(b)) as the rate would linearly in-
crease with the number of matter qubits. There-
fore, we have derived a general theoretical upper
bound for memory-based protocols based on her-
alded entanglement generation. It is worth not-
ing that the fundamental reason for this upper
bound comes from the need for classical signal-
ing in these protocols. Such classical signaling
is not required for RGS protocols, enabling them
to surpass this limit, as we show below. In the
Supplementary Materials, we also show that a
tighter bound, R
(QM)
max
= c/7L, can be obtained
for memory-based schemes in which there are two
quantum memories per repeater node, and her-
alded entanglement swapping is used.
We emphasize that the upper bound derived in
this section holds for QR protocols that are based
on quantum memories and distant heralded en-
tanglement generation. This corresponds to the
first and second generations of QRs, as catego-
rized in Ref. [15]. Consequently, in the following,
the RGS-based protocol will be compared only
to these categories of QRs, for which the perfor-
mance is limited by classical signaling.
2 Rate of the RGS protocol with de-
terministic graph state generation
In this section, we review the RGS protocol as
introduced in Ref. [31] and the deterministic gen-
eration of RGSs using a few matter qubits as
proposed in Ref. [42]. We show how the rate of
the RGS protocol depends on various parameters
in the case where deterministic state generation
methods are used.
2.1 RGS protocol and rate
An RGS is a quantum state |Gi that can con-
veniently be represented in the form of a graph
G = (V, E) with V vertices and E edges. Each
vertex corresponds to a photonic qubit prepared
in the |+i state, and each edge corresponds to the
application of a CZ gate between the two qubits
it connects:
|Gi =
Y
(i,j)E
CZ
ij
|+i
V
. (2)
An example of the graph representing an RGS
is shown in Fig. 1(c). These states include 2m
inner photonic qubits that are referred to as the
first-leaf qubits. All the first-leaf qubits are fully
connected to each other and each of them is also
connected to one additional qubit, referred to
as a second-leaf qubit. The first-leaf qubits are
logically-encoded using tree graph states; further
details on this are given below.
In an RGS protocol (see Fig. 1(d)), the distance
L separating Alice and Bob is also divided into
smaller steps L
0
by N
QR
= L/L
0
1 source nodes
where the RGSs are created. The RGS is divided
into two equal parts, each containing m arms,
and one part is sent to the left adjacent measure-
ment node and the other to the right. Thus, half
of one RGS meets half of another RGS at each
measurement node, where each second-leaf qubit
from one of the half-RGSs undergoes a Bell mea-
surement with its counterpart from the other half-
RGS. Further details about this entanglement
swapping procedure at the measurement node are
given later, but it is important to note that the
RGS protocol does not use quantum memories at
all and thus cannot store the information. This
implies that an entanglement connection between
Alice and Bob should be realized in only one trial
with a probability P
AB
much higher than direct
fiber transmission: P
AB
η
t
(L). The genera-
tion of an entanglement connection between Alice
and Bob requires the realization of successful en-
tanglement connections between all the adjacent
RGSs. It is important to note that the measure-
ments performed at each measurement node do
not require information from other measurement
nodes so that, in contrast to memory-based ap-
proaches, the RGS protocol does not require any
classical signaling while entanglement is being ex-
tended through the network. Classical signaling
is needed only once at the end of the protocol to
recover the Pauli frame of the Bell pair shared
by Alice and Bob, i.e. to determine which local
Pauli rotations Alice and Bob’s qubits should un-
dergo. This means that in the RGS protocol, it
is not necessary to wait for any classical signaling
before proceeding to generate the next batch of
RGSs needed to create the next Bell pair shared
between Alice and Bob (see Supplementary Ma-
terials for more details). Consequently, unlike the
memory-based scheme, the rate of the RGS pro-
tocol does not depend on the time it takes for the
photon to get from one node to the next. It is
limited only by the generation time T
RGS
of an
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 4
RGS:
R
(RGS)
=
P
RGSRGS
L/L
0
T
RGS
, (3)
with P
RGSRGS
the probability to generate
an entanglement connection between two RGSs.
The main notations used in this work are defined
in Table 1.
A successful entanglement link can be gener-
ated if at least one of the Bell measurements at
a measurement node succeeds. In that case, the
two first-leaf qubits attached to the second-leaf
qubits that underwent the successful Bell mea-
surement are measured in the X basis, while the
remaining 2m 2 first-leaf qubits are measured
in the Z basis. The X measurements transfer
the entanglement connection to the next two ad-
jacent measurement nodes, while the Z measure-
ments disentangle all the excess qubits associated
with failed Bell measurements. All these first-leaf
qubit measurements must be successful in order
to reliably create an entanglement link. There-
fore, the probability to successfully create an en-
tanglement link between two RGSs is given by:
P
RGSRGS
= (1 (1 P
Bell
)
m
)
× Pr(M
X,`
)
2
Pr(M
Z,`
)
2m2
,
(4)
where P
Bell
= P
ph
2
/2 is the probability of a
successful Bell measurement. This depends on
P
ph
= η
c
η
d
η
t
(L
0
/2), which is the probability
that a single photon is emitted and collected
into the fiber (η
c
), is transmitted to the mea-
surement node (η
t
(L
0
/2)) and detected (η
d
).
Pr(M
X,`
) and Pr(M
Z,`
) are the probabilities that
the logical X and Z measurements on the first-
leaf qubits succeed. Note that if the first-leaf
qubits were not logically encoded, we would have
Pr(M
X
)
2
Pr(M
Z
)
2m2
= P
ph
2m
η
t
(L
0
), and it
would be impossible to have an advantage over
direct fiber transmission. Therefore, the loss-
tolerance of logically-encoded qubits is crucial for
this protocol. Next, we review this encoding,
which was introduced in Refs. [35] and [31].
2.2 Loss-tolerance with tree graph states
We review how the probabilities Pr(M
X,`
) and
Pr(M
Z,`
) depend on the single-photon transfer
probability P
ph
and on the shape of the tree graph
state used for the logical encoding. Ref. [35]
demonstrated that this encoding remains loss-
tolerant as long as P
ph
is above 50%.
We consider the calculation of the probabili-
ties of successful measurements of the logically
encoded qubits in the presence of loss errors on
a tree graph state. A tree is characterized by
its branching vector
~
b = (b
0
, b
1
, ..., b
n1
) (see
Fig. 1(c)), which describes the connectivity be-
tween the different levels of the tree. To perform
a Z measurement M
Z,k
on a qubit at level k, it is
possible to either perform a direct measurement
on this qubit (with success probability P
ph
) or, if
it fails (with probability 1 P
ph
), perform an in-
direct measurement (with probability r
k
). Thus,
the overall success probability of a Z measure-
ment at level k is:
Pr(M
Z,k
) = P
ph
+ (1 P
ph
)r
k
. (5)
To perform an indirect measurement on a qubit
(call it A) at level k, one can use the stabilizing
property of a graph state [56]. It is possible to
deduce the outcome of the Z measurement on
A by performing an X measurement on another
qubit (B) at level k + 1 and a Z measurement on
all the qubits, C
i
, that are in the neighborhood
of B at level k + 2 (see Fig. 2(a)). This works
because of the invariance of graph states when
they are acted upon by their stabilizers:
|Gi = X
B
jN(B)
Z
j
|Gi
= X
B
Z
A
i∈{1,b
k+1
}
Z
C
i
|Gi ,
(6)
so we have:
Z
A
|Gi = X
B
i∈{1,b
k+1
}
Z
C
i
|Gi . (7)
A single indirect measurement has a success
probability s
k
. Note, however, that the tree
structure allows b
k
indirect measurement at-
tempts, and only one needs to succeed to indi-
rectly measure a qubit at level k. So the proba-
bility that at least one indirect measurement suc-
ceeds is
r
k
= 1 (1 s
k
)
b
k
, (8)
with
s
k
= P
ph
Pr(M
Z,k+2
)
b
k+1
. (9)
It is thus possible to derive the success probabil-
ity of a measurement recursively, given that the
qubits at the lowest level can only be measured
directly: Pr(M
Z,n
) = P
ph
. Logical measurements
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 5
A
B
C
2
C
1
Level k
Level k+1
Level k+2
(a) (b)
Figure 2: (a) Indirect measurement of qubit A at the level k using a stabilizer based on qubit B at level k + 1. (b)
Protocol for the deterministic generation of an RGS with logical encoding using matter qubits.
in the X or Z basis are given by [31]:
Pr(M
X,`
) = r
0
,
Pr(M
Z,`
) = (P
ph
+ (1 P
ph
)r
1
)
b
0
= Pr(M
Z,1
)
b
0
.
(10)
It is interesting to note that logical encoding with
tree graph states can also correct single-qubit er-
rors, as shown in Ref. [31] and described in the
Supplementary Materials. This will be used later
when we evaluate the sensitivity of the RGS pro-
tocol to errors.
2.3 Generation of an RGS
The achievable rate between Alice and Bob also
depends on the repetition rate of the protocol,
which is given by the generation time of the RGS.
Because it is impossible to realize determinis-
tic two-qubit gates on photons with linear op-
tics, such a graph state can either be generated
probabilistically by the recursive fusion of smaller
graphs using linear optics and Bell state measure-
ments as shown in Ref. [31] and [37], or deter-
ministically using a few matter qubits as shown
in Ref. [42]. We now review the latter.
An arbitrary-sized RGS can be generated de-
terministically by following a given sequence
based on four operations on matter qubits: the
emission of a photon maximally entangled with
the matter qubit E
ph
, the Hadamard gate H,
measurements in the Pauli bases M
X
, M
Y
, M
Z
and the CZ gate. The generation of an RGS with
2m arms and a tree graph encoding with branch-
ing vector
~
b = (b
0
, b
1
, ..., b
n1
) requires n+1 mat-
ter qubits Q
1
, ..., Q
n+1
, and is given by the se-
quence (see also Fig. 2(b))
M
Y,Q
1
(M
X,Q
3
M
X,Q
2
CZ
Q
1
,Q
3
CZ
Q
2
,Q
3
E
ph,Q
3
G
3
b
0
)
2m
with G
k
= M
Z,Q
k
H
Q
k
E
ph,Q
k
CZ
Q
k1
,Q
k
G
k+1
b
k2
and G
n+2
= E
ph,Q
n+1
,
(11)
where, for simplicity, we have omitted the single
photonic qubit rotations.
The overall generation time of an RGS using
this procedure is therefore
T
RGS
= 2m
1 + f(
~
b, n 1)
T
E
ph
+ T
M
+ 2m
2 + f(
~
b, n 2)
(T
M
+ T
CZ
)
+ 2mf(
~
b, n 2)T
H
,
(12)
with f (
~
b, k) =
P
k
i=0
Q
i
j=0
b
j
and with T
E
ph
, T
M
,
T
H
and T
CZ
the times for photon emission, mat-
ter qubit measurement, Hadamard and CZ gates,
respectively.
In the following, we make the realistic assump-
tion that the CZ gate time T
CZ
is much longer
than the durations of the other operations, and so
we set T
H
= T
M
= T
E
ph
= 0 for simplicity. With
this assumption, the generation time T
RGS
only
depends on the number of CZ gates and their
duration:
T
RGS
= 2m
2 +
n2
X
k=0
k
Y
j=0
b
j
T
CZ
. (13)
This is for an RGS with 2m arms and a log-
ical tree encoding with branching vector
~
b =
(b
0
, ..., b
n1
). In the following, we will assume
a depth-two tree graph state (
~
b = (b
0
, b
1
)), so
that the full RGS can be generated with only
three matter qubits (n = 2). The total num-
ber of matter qubits in the network is therefore
N
m
= (L/L
0
1)(n + 1) = 3(L/L
0
1).
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 6
3 Performance comparisons
We now evaluate the performance of the RGS
protocol when it is generated deterministically
by a few matter qubits. We compare our re-
sults to the direct fiber transmission limit derived
by Refs. [57, 58] and to the memory-based upper
bound R
(QM)
max
= c/4L found in Sec. 1.
3.1 Optimizing the RGS protocol
(a) (b)
(d)(c)
Figure 3: Maximum achievable rate per matter qubit
R
(RGS)
m
T
CZ
(here normalized by the parameter T
CZ
1
)
and optimal node separation L
0
for the RGS protocol
with depth-2 logical tree encoding with the total dis-
tance fixed at L = 50L
att
for a range of RGS parameters
b
0
, b
1
, m. (a) Three-dimensional plot showing optimal
rate and node separation as a function of b
0
, b
1
and
m. For each point in the plot, the value of L
0
is op-
timized to maximize R
(RGS)
m
T
CZ
. Point sizes represent
maximized R
(RGS)
m
T
CZ
values while point colors repre-
sent the optimal values of L
0
(indicated with the color
scale). The RGS parameters that achieve the largest
value of R
(RGS)
m
T
CZ
in this case are indicated with
dashed lines. The corresponding maximal R
(RGS)
max
T
CZ
is given in Table 2. (b,c,d) Three different orthogonal
two-dimensional slices of the plot shown in panel (a).
In the following, we show how the RGS proto-
col parameters can be optimized to maximize the
overall rate per matter qubit, R
(RGS)
m
, or in the
presence of errors, the secret key rate per matter
qubit. For the moment, we assume perfect pho-
ton collection and detection efficiencies (η
c
η
d
= 1)
and no single-photon errors ( = 0); we take these
effects into account later on. From Eqs. (3), (4),
and (13), the achievable rate per matter qubit
R
(RGS)
m
of the RGS protocol for a total distance
L is inversely proportional to the CZ gate time
T
CZ
, but it depends non-linearly on the separa-
L/L
att
L
0
/L
att
m b
0
b
1
R
(RGS)
max
T
CZ
10 0.23 11 8 4 2.4 × 10
5
25 0.21 13 10 5 6.8 × 10
6
50 0.19 14 10 5 2.8 × 10
6
100 0.17 15 10 5 1.1 × 10
6
150 0.15 16 10 5 6.8 × 10
7
Table 2: Optimal RGS parameters m, b
0
, b
1
and node
separation L
0
for several different total network dis-
tances L. Here, L
att
is the attenuation length of optical
fibers.
tion distance, L
0
, between two RGS nodes and
the RGS shape (number of arms 2m and the tree
branching vector
~
b = (b
0
, b
1
)). Therefore, for each
choice of the total distance L, there is a certain
node separation and RGS shape that maximize
the achievable rate.
The optimization of the rate per matter qubit
for a total distance L = 50L
att
( 1000 km) is
shown in Fig. 3. The position of each point cor-
responds to a specific RGS shape, the color of the
point indicates the node separation L
0
that op-
timizes the rate for that shape, and the size of
the point represents the maximal rate per mat-
ter qubit for these parameters. This optimization
converges, allowing us to extract the optimal RGS
shape and distance L
0
for this particular choice of
the total distance L. The optimization can be re-
peated for various choices of L, and the extracted
optimal parameters are recorded in Table 2.
We compare the RGS protocol to direct fiber
transmission in Fig. 4(a). The maximum achiev-
able rate per matter qubit for the RGS proto-
col, R
(RGS)
max
, is shown as a function of the total
distance L/L
att
. In the case of direct transmis-
sion, we show seven different curves correspond-
ing to the achievable rate for seven different val-
ues of the single-photon source repetition rate.
We see that the RGS protocol outperforms direct
transmission with the highest repetition rate for
L & 30L
att
. In the same figure, we also show how
well the protocol works if we keep the optimized
parameters fixed and change the total distance
L. To demonstrate this, we fix the RGS param-
eters and node separation L
0
to the values that
optimize the rate for a total distance L
tar
. We
then adjust L away from L
tar
without changing
the RGS parameters or L
0
, and we calculate the
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 7
new rate for each value of L. In Fig. 4(a), we
show the resulting rates as a function of L/L
0
for five different choices of L
tar
. We see that the
RGS protocol continues to work well over a broad
range of total distance L when we use parameters
that are optimized for a large total distance L
tar
.
For long distances L in the range 50L
att
200L
att
, R
(RGS)
max
scales approximately as L
1.27
,
and thus the scaling with distance is slightly
worse than the upper bound scaling for memory-
based schemes, R
(QM)
max
L
1
, obtained in Sec. 1.
However, contrary to the memory-based schemes,
there are no fundamental upper limits imposed by
classical signaling on the maximal achievable rate
R
(RGS)
max
, as the latter is inversely proportional to
the CZ gate time, T
CZ
. If T
CZ
is sufficiently
small, then the RGS protocol should surpass the
upper bound of memory-based schemes. This is
illustrated in Fig. 4(b), where the red region of
the plot indicates the regime where the RGS pro-
tocol outperforms memory-based schemes.
From our calculations, it seems that the max-
imum achievable rate per matter qubit, R
(RGS)
max
,
is in principle unbounded since it is inversely pro-
portional to the CZ gate time. We should recall
however that these results are based on the as-
sumption that the CZ gate takes much longer
than the other operations O made on the mat-
ter qubits: T
CZ
T
O
. Decreasing T
CZ
ini-
tially increases the rate, but if T
CZ
becomes small
enough, neglecting the durations of other opera-
tions eventually becomes invalid. Consequently,
to increase the rate further, not only the CZ
gate time but all the operation times should be
reduced simultaneously. If the durations of all
these operations could be made arbitrarily small,
then the rate per matter qubit would increase
to infinity. In practice however, the operation
times could also have an intrinsic lower bound
that limits the performance of the RGS proto-
col, but these lower bounds would depend on the
specific system on which the protocol is imple-
mented, while the limit imposed by classical sig-
naling is much more stringent and general.
For distances below 200L
att
4000 km, a gate
time T
CZ
below 6 × 10
4
t
att
60 ns is suffi-
ciently short to outperform any memory-based
scheme. For this range of total distance, the
optimal value of the node separation L
0
ranges
from 0.15 0.19L
att
3 3.8 km. As an
illustration of the RGS performance, for L =
1000 km and T
CZ
= 10 ns, the total rate is
R = 220 kHz for N
m
= 786 matter qubits used,
leading to R
(RGS)
max
= 276 Hz per matter qubit,
while R
(QM)
max
= 50 Hz per matter qubit for this
distance.
3.2 Sensitivity to errors
So far, we have considered the generation of a
perfect RGS, which is a pure entangled state of
many photons. Generating such a perfect state
is not feasible experimentally, and so we need to
evaluate the sensitivity of the RGS protocol to
errors. To do so, we now consider single-photon
loss and single-photon errors in our optimization
process.
The single-photon loss (other than the fiber
losses) depends on the probability that an emit-
ted photon is neither collected into the fiber
nor detected; this corresponds to the case where
η
c
η
d
6= 1. Fig. 4(c) shows the optimized rate per
matter qubit for different values of η
c
η
d
. Simi-
larly, Fig. 4(d) compares the performance of the
RGS protocol with the upper bound for memory-
based protocols, R
(QM)
max
, as a function of T
CZ
and
η
c
η
d
for a total distance L = 50L
att
. These re-
sults show that the photon losses must be below
15% (η
c
η
d
> 85%) in order for the RGS proto-
col to outperform both the direct fiber transmis-
sion and memory-based protocols. This rather
stringent requirement is mainly due to the loss-
tolerance of the tree graph state, which only
works when the total photon loss (including fiber
losses) is below 50%.
Apart from the photon losses, the RGS pro-
tocol also depends on other kinds of errors that
reduce the fidelity F
AB
of the final entangled pho-
ton pair shared by Alice and Bob. These er-
rors include, but are not limited to, single-photon
measurement errors, photon Bell-measurement
errors, and depolarization errors. In the present
case where the RGS is generated using matter
qubits, the limited coherence time of the matter
qubits should also limit the fidelity of the entan-
gled photons. For QKD applications, the final
secret key rate depends on the total rate R and
the fidelity F
AB
[59, 60]:
R
skr
= R(1 2h(F
AB
)), (14)
where h is the binary entropy function:
h(F ) = F log
2
(F ) (1 F ) log
2
(1 F ). (15)
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 8
max
R
(RGS)
(QM)
max
R-
max
R
(RGS)
(QM)
max
R+
(a)
(b)
(c)
(d)
=50
(f)
=50
max
R
(RGS)
(Q
)
max
R=
(e)
Figure 4: (a) The maximum achievable secret key rate per matter qubit as a function of the total distance L for
the RGS protocol with depth-2 logical tree encoding compared to direct fiber transmission. The dashed red line
is the rate for the RGS protocol, with all RGS parameters and L
0
optimized for each value of L. The seven red
curves correspond to rates for direct fiber transmission [58] for seven different choices of the single-photon source
repetition rate: 10
i
T
CZ
1
(for i=0,...,6). The solid yellow, green, and blue curves show the performance of the RGS
protocol when L is changed while keeping all RGS parameters and L
0
fixed to the optimal values obtained for total
distance L
tar
. (b) Relative difference between the maximum achievable rates of the RGS protocol, R
(RGS)
max
, and
memory-based protocols, R
(QM)
max
, as a function of gate time T
CZ
and total distance L. The red region corresponds
to R
(RGS)
max
> R
(QM)
max
. (c,d) Optimized rate of the RGS protocol for different values of photon loss probability η
c
η
d
compared to (c) direct fiber transmission and (d) memory-based limit. (e,f) Similar to (c,d) but now showing results
for the single-photon error probability . In (c,e), the RGS rate is maximized for each value of L.
In the remainder of this paper, we now use
R
(RGS)
m
to refer to the secret key rate per matter
qubit, which coincides with the rate per matter
qubit in the absence of errors. We model these
imperfections by using a single-qubit error prob-
ability that affects all the photonic qubits. This
model works well for qubit measurement errors as
well as depolarization errors. In Ref. [38], it was
also shown that decoherence errors on quantum
emitters induce only local errors in the emitted
cluster state. The detailed derivation of F
AB
is
given in the Supplementary Materials. Fig. 4(e)
compares the optimized secret key rate of the
RGS protocol against that of direct fiber trans-
mission as a function of the single-qubit error
probability and the total distance L. Fig. 4(f)
compares the performance of the RGS protocol to
the upper bound for memory-based protocols for
L = 50L
att
. We find that the single-qubit error
should be below < 10
4
to outperform both the
direct fiber transmission and the memory-based
protocol upper bound, R
(QM)
max
.
One might expect a better tolerance to single-
qubit errors since they are also corrected by the
logical encoding of the 1
st
-leaf qubits. This is
not the case because the 2
nd
-leaf qubits are not
logically encoded and thus are very sensitive to
errors. The dependence on single-qubit errors is
actually similar to the protocols using heralded
entanglement generation. It is also possible to
use entanglement purification [6163] to improve
the fidelity F
AB
and maximize the secret key rate.
From the maximum single-qubit error rate, we
can derive an order-of-magnitude estimate for
the coherence time requirements for the matter
qubits. To do so, we use two strong assumptions.
We first assume that the fidelity F
RGS
of an RGS
depends on the ratio between the generation time
T
RGS
of the RGS and the coherence time T
2
of
the matter qubits:
F
RGS
e
3T
RGS
T
2
(16)
The coefficient 3 corresponds to the number of
matter qubits that degrade the RGS fidelity.
Here, we have made the simplifying assumption
that the main source of infidelity is the matter
qubit decoherence, and we have assumed that the
effect of this decoherence is to reduce the overall
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 9
Matter
qubit
T
CZ
F T
2
β
Coupled
QDs
1ns [64]
[47, 65]
80%
[65]
3µs
[66, 67]
99.7%
[68]
defects
(diamond)
120ns
[6971]
99.2%
[72]
0.6s
[73]
99%
[23]
Trapped
ions
10µs
[74]
99.9%
[74]
10min
[75]
89%
[76]
Table 3: State-of-the-art characteristics of candidate
quantum emitters. T
CZ
is the duration of a CZ gate,
T
2
is the coherence time, β is the single-mode photon
emission probability, and F the two-qubit gate fidelity.
Blue (red) color means that the condition is (not) ful-
filled with this system, while orange means that it is
theoretically possible but has not been realized experi-
mentally. Note that all these results were not achieved
on the same system at the same time.
fidelity of the graph state by e
T
RGS
/T
2
, which de-
cays exponentially with the time during which the
qubit is attached to the RGS (T
RGS
) divided by
its coherence time. This assumption may or may
not be justified depending on the system used
to implement the protocol, but we use it sim-
ply to get an order of magnitude estimate of the
typical coherence time that is required for this
protocol. We also assume that the errors are ho-
mogeneously distributed over the N
ph
photons of
an RGS, such that F
RGS
= (1 )
N
ph
. This as-
sumption is consistent with Ref. [38], which shows
that errors generated by a quantum emitter and
imprinted onto the generated graph state are lo-
calized. With these assumptions, the coherence
time should satisfy T
2
2500T
CZ
.
4 Discussion
In the previous section, we derived a set of re-
quirements that need to be fulfilled in order for
the RGS protocol to have a clear advantage over
both memory-based QRs and repeater-less proto-
cols:
(i) T
CZ
T
lim
(L) 60ns,
(ii) T
2
2500T
CZ
,
(iii) η
c
η
d
> 85%.
It is important to note that these bounds were
obtained using RGSs with a depth-2 logical tree
encoding. We may expect an increase in the rate
per matter qubit if deeper trees were used but
the improvement would certainly not be large as
T
RGS
grows quickly with tree depth. Checking
this directly is challenging because the process
of optimizing the RGS protocol becomes signif-
icantly more complex. We leave this intensive
numerical analysis to future work.
Table 3 shows the current state-of-the-art for
three candidate systems that can be used to gen-
erate RGSs. So far, the characteristics of the cur-
rent systems are not good enough yet to exceed
the memory-based resource-efficient upper bound
R
(QM)
max
. The main limitation comes from long
two-qubit gate times, since fast and high-fidelity
gates are critical for the performance of the RGS
protocol. Currently, most of the two-qubit gates
that have been experimentally demonstrated are
slow because they rely on local interactions, such
as the hyperfine coupling, which are generally
weak. Efficient light-matter interfaces using cav-
ities can also be used to realize deterministic,
photon-mediated CZ gates between two matter
qubits [70], using spin-photon gates [7779]. The
crucial advantage of this technique is that it only
requires optical interactions, which are intrin-
sically much faster than hyperfine interactions.
Progress in the experimental realization of strong
spin-dependent phase shifts has been made re-
cently [23, 8082]. The second condition (ii) will
also be easier to achieve if the CZ gate time is
reduced.
The third condition (iii) requires the efficient
collection of single photons emitted by a single
quantum emitter. The collection of single pho-
tons needs to be facilitated by increasing the
emission of photons into a single optical mode.
This is generally realized using waveguides or
cavities with single-mode emission probabilities
β that can reach 99.7% [68]. This single mode
should then be efficiently coupled to a single-
mode fiber. Finally, superconducting photon de-
tectors can be used to achieve high detection ef-
ficiencies [83].
In conclusion, we have identified the main re-
quirements that must be met in order for all-
photonic repeater protocols based on determin-
istically generated graph states to outperform
memory-based protocols. We did this by find-
ing the graph state structures that maximize the
secret key rate and by comparing this to an up-
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 10
per bound on the rate for memory-based schemes
that we derived. We found that two-qubit gate
times and the efficient collection and detection
of emitted photons are the most important fac-
tors in determining whether the all-photonic ap-
proach is superior to memory-based ones. An in-
teresting future direction would be to improve the
protocol for the emitter-based generation of pho-
tonic graph states by incorporating atom-photon
CZ gates [44, 79, 84], which would both sim-
plify the procedure and decrease gate times, and
thus enhance the performance of deterministic,
all-photonic repeaters.
Acknowledgements
We thank Yuan Zhan and Shuo Sun for their
contribution in solving issues in the code. This
research was supported by the NSF (Grant No.
1741656) and by the EU Horizon 2020 programme
(GA 862035 QLUSTER).
Methods
The numerical model used for the results in this
article is available here:
https://github.com/Paulhilaire/
performance_repeater_graph_state.
References
[1] H. J. Kimble. The quantum internet.
Nature, 453(7198):1023–1030, 2008. DOI:
10.1038/nature07127.
[2] Stephanie Wehner, David Elkouss, and
Ronald Hanson. Quantum internet: A
vision for the road ahead. Science, 362
(6412):eaam9288, 2018. DOI: 10.1126/sci-
ence.aam9288.
[3] Charles H Bennett and Gilles Brassard.
An update on quantum cryptography. In
Workshop on the Theory and Application
of Cryptographic Techniques, pages 475–480.
Springer, 1984. DOI: 10.1007/3-540-39568-
7_39.
[4] Charles H. Bennett, François Bessette, Gilles
Brassard, Louis Salvail, and John Smolin.
Experimental quantum cryptography. Jour-
nal of Cryptology, 5:3–28, 1992. DOI:
10.1007/BF00191318.
[5] Thomas. Jennewein, C. Simon, Gregor
Weihs, Harald Weinfurter, and Anton
Zeilinger. Quantum cryprography using
entangled photons. Physical Review Let-
ters, 84:4729, 2000. DOI: 10.1103/Phys-
RevLett.84.4729.
[6] Anne Broadbent, Joseph Fitzsimons, and El-
ham Kashefi. Universal blind quantum com-
putation. In 2009 50th Annual IEEE Sym-
posium on Foundations of Computer Sci-
ence, pages 517–526. IEEE, 2009. DOI:
10.1109/FOCS.2009.36.
[7] Lov K Grover. Quantum telecomputation.
arXiv preprint quant-ph/9704012, 1997.
[8] Naomi H Nickerson, Joseph F Fitzsimons,
and Simon C Benjamin. Freely scalable
quantum technologies using cells of 5-to-50
qubits with very lossy and noisy photonic
links. Physical Review X, 4(4):041041, 2014.
DOI: 10.1103/PhysRevX.4.041041.
[9] Peter Komar, Eric M Kessler, Michael
Bishof, Liang Jiang, Anders S Sørensen, Jun
Ye, and Mikhail D Lukin. A quantum net-
work of clocks. Nature Physics, 10(8):582,
2014. DOI: 10.1038/nphys3000.
[10] Daniel Gottesman, Thomas Jennewein, and
Sarah Croke. Longer-baseline telescopes
using quantum repeaters. Physical Re-
view Letters, 109(7):070503, 2012. DOI:
10.1103/PhysRevLett.109.070503.
[11] W. K. Wootters and W. H. Zurek. A single
quantum cannot be cloned. Nature, 299:802–
803, 1982. DOI: 10.1038/299802a0.
[12] D. Dieks. Communication by EPR devices.
Physics Letters A, 92:271–272, 1982. DOI:
10.1016/0375-9601(82)90084-6.
[13] H-J Briegel, Wolfgang Dür, Juan I Cirac,
and Peter Zoller. Quantum repeaters: the
role of imperfect local operations in quan-
tum communication. Physical Review Let-
ters, 81(26):5932, 1998. DOI: 10.1103/Phys-
RevLett.81.5932.
[14] W Dür, H-J Briegel, JI Cirac, and P Zoller.
Quantum repeaters based on entanglement
purification. Physical Review A, 59(1):169,
1999. DOI: 10.1103/PhysRevA.59.169.
[15] Sreraman Muralidharan, Linshu Li,
Jungsang Kim, Norbert Lütkenhaus,
Mikhail D Lukin, and Liang Jiang. Optimal
architectures for long distance quantum
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 11
communication. Scientific reports, 6:20463,
2016. DOI: 10.1038/srep20463.
[16] L. M. Duan, M. D. Lukin, J. I. Cirac,
and P. Zoller. Long-distance quantum
communication with atomic ensembles and
linear optics. Nature, 414(6862):413–418,
November 2001. ISSN 0028-0836. DOI:
10.1038/35106500.
[17] Lilian Childress, JM Taylor, Anders Sønd-
berg Sørensen, and MD Lukin. Fault-
tolerant quantum communication based on
solid-state photon emitters. Physical Re-
view Letters, 96(7):070504, 2006. DOI:
10.1103/PhysRevLett.96.070504.
[18] Lorenz Hartmann, Barbara Kraus, H-J
Briegel, and W Dür. Role of memory er-
rors in quantum repeaters. Physical Review
A, 75(3):032310, 2007. DOI: 10.1103/Phys-
RevA.75.032310.
[19] OA Collins, SD Jenkins, A Kuzmich,
and TAB Kennedy. Multiplexed memory-
insensitive quantum repeaters. Physical
review letters, 98(6):060502, 2007. DOI:
10.1103/PhysRevLett.98.060502.
[20] Nicolas Sangouard, Christoph Simon,
Hugues De Riedmatten, and Nicolas Gisin.
Quantum repeaters based on atomic en-
sembles and linear optics. Reviews of
Modern Physics, 83(1):33, 2011. DOI:
10.1103/RevModPhys.83.33.
[21] Scott E Vinay and Pieter Kok. Practi-
cal repeaters for ultralong-distance quan-
tum communication. Physical Review A,
95(5):052336, 2017. DOI: 10.1103/Phys-
RevA.95.052336.
[22] Filip Rozpędek, Raja Yehia, Kenneth
Goodenough, Maximilian Ruf, Peter C
Humphreys, Ronald Hanson, Stephanie
Wehner, and David Elkouss. Near-
term quantum-repeater experiments with
nitrogen-vacancy centers: Overcoming the
limitations of direct transmission. Physi-
cal Review A, 99(5):052330, 2019. DOI:
10.1103/PhysRevA.99.052330.
[23] Mihir K Bhaskar, Ralf Riedinger,
Bartholomeus Machielse, David S Lev-
onian, Christian T Nguyen, Erik N Knall,
Hongkun Park, Dirk Englund, Marko
Lončar, Denis D Sukachev, et al. Experi-
mental demonstration of memory-enhanced
quantum communication. Nature, 580
(7801):60–64, 2020. DOI: 10.1038/s41586-
020-2103-5.
[24] Sumeet Khatri, Corey T Matyas, Aliza U
Siddiqui, and Jonathan P Dowling. Prac-
tical figures of merit and thresholds for
entanglement distribution in quantum net-
works. Physical Review Research, 1(2):
023032, 2019. DOI: 10.1103/PhysRevRe-
search.1.023032.
[25] C Cabrillo, JI Cirac, P Garcia-Fernandez,
and P Zoller. Creation of entangled states
of distant atoms by interference. Phys-
ical Review A, 59(2):1025, 1999. DOI:
10.1103/PhysRevA.59.1025.
[26] Sean D Barrett and Pieter Kok. Effi-
cient high-fidelity quantum computation us-
ing matter qubits and linear optics. Phys-
ical Review A, 71(6):060310, 2005. DOI:
10.1103/PhysRevA.71.060310.
[27] L.-M. Duan and H. J. Kimble. Scalable
photonic quantum computation through
cavity-assisted interactions. Physical Re-
view Letters, 92(12):127902, 2004. DOI:
10.1103/PhysRevLett.92.127902.
[28] Liang Jiang, Jacob M Taylor, Kae Nemoto,
William J Munro, Rodney Van Meter, and
Mikhail D Lukin. Quantum repeater with
encoding. Physical Review A, 79(3):032325,
2009. DOI: 10.1103/PhysRevA.79.032325.
[29] William J Munro, Ashley M Stephens, Si-
mon J Devitt, Keith A Harrison, and Kae
Nemoto. Quantum communication with-
out the necessity of quantum memories.
Nature Photonics, 6(11):777, 2012. DOI:
10.1038/nphoton.2012.243.
[30] Sreraman Muralidharan, Jungsang Kim,
Norbert Lütkenhaus, Mikhail D Lukin,
and Liang Jiang. Ultrafast and fault-
tolerant quantum communication across
long distances. Physical review letters, 112
(25):250501, 2014. DOI: 10.1103/Phys-
RevLett.112.250501.
[31] Koji Azuma, Kiyoshi Tamaki, and Hoi-
Kwong Lo. All-photonic quantum repeaters.
Nature communications, 6:6787, 2015. DOI:
10.1038/ncomms7787.
[32] Fabian Ewert, Marcel Bergmann, and Pe-
ter van Loock. Ultrafast long-distance
quantum communication with static lin-
ear optics. Physical review letters, 117
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 12
(21):210501, 2016. DOI: 10.1103/Phys-
RevLett.117.210501.
[33] Yasushi Hasegawa, Rikizo Ikuta, Nobuyuki
Matsuda, Kiyoshi Tamaki, Hoi-Kwong Lo,
Takashi Yamamoto, Koji Azuma, and
Nobuyuki Imoto. Experimental time-
reversed adaptive bell measurement towards
all-photonic quantum repeaters. Nature
communications, 10(1):378, 2019. DOI:
10.1038/s41467-018-08099-5.
[34] Zheng-Da Li, Rui Zhang, Xu-Fei Yin, Li-
Zheng Liu, Yi Hu, Yu-Qiang Fang, Yue-Yang
Fei, Xiao Jiang, Jun Zhang, Li Li, et al. Ex-
perimental quantum repeater without quan-
tum memory. Nature Photonics, page 1,
2019. DOI: 10.1038/s41566-019-0468-5.
[35] Michael Varnava, Daniel E Browne, and
Terry Rudolph. Loss tolerance in one-
way quantum computation via counterfac-
tual error correction. Physical review letters,
97(12):120501, 2006. DOI: 10.1103/Phys-
RevLett.97.120501.
[36] Daniel E Browne and Terry Rudolph.
Resource-efficient linear optical quantum
computation. Physical Review Letters, 95
(1):010501, 2005. DOI: 10.1103/Phys-
RevLett.95.010501.
[37] Mihir Pant, Hari Krovi, Dirk Englund, and
Saikat Guha. Rate-distance tradeoff and
resource costs for all-optical quantum re-
peaters. Physical Review A, 95(1):012304,
2017. DOI: 10.1103/PhysRevA.95.012304.
[38] Netanel H Lindner and Terry Rudolph. Pro-
posal for pulsed on-demand sources of pho-
tonic cluster state strings. Physical Re-
view Letters, 103(11):113602, 2009. DOI:
10.1103/PhysRevLett.103.113602.
[39] C Saavedra, KM Gheri, P Törmä, JI Cirac,
and P Zoller. Controlled source of entan-
gled photonic qubits. Physical Review A,
61(6):062311, 2000. DOI: 10.1103/Phys-
RevA.61.062311.
[40] Ido Schwartz, Dan Cogan, Emma R
Schmidgall, Yaroslav Don, Liron Gantz,
Oded Kenneth, Netanel H Lindner, and
David Gershoni. Deterministic generation of
a cluster state of entangled photons. Sci-
ence, 354:434–437, 2016. DOI: 10.1126/sci-
ence.aah4758.
[41] Sophia E Economou, Netanel Lindner, and
Terry Rudolph. Optically generated 2-
dimensional photonic cluster state from cou-
pled quantum dots. Physical review letters,
105(9):093601, 2010. DOI: 10.1103/Phys-
RevLett.105.093601.
[42] Donovan Buterakos, Edwin Barnes, and
Sophia E Economou. Deterministic gen-
eration of all-photonic quantum repeaters
from solid-state emitters. Physical Review
X, 7(4):041023, 2017. DOI: 10.1103/Phys-
RevX.7.041023.
[43] Antonio Russo, Edwin Barnes, and Sophia E
Economou. Generation of arbitrary all-
photonic graph states from quantum emit-
ters. New Journal of Physics, 21(5):055002,
2019. DOI: 10.1088/1367-2630/ab193d.
[44] Hannes Pichler, Soonwon Choi, Peter Zoller,
and Mikhail D Lukin. Universal photonic
quantum computation via time-delayed feed-
back. Proceedings of the National Academy
of Sciences, 114(43):11362–11367, 2017.
DOI: 10.1073/pnas.1711003114.
[45] Antonio Russo, Edwin Barnes, and Sophia E
Economou. Photonic graph state genera-
tion from quantum dots and color centers for
quantum communications. Physical Review
B, 98(8):085303, 2018. DOI: 10.1103/Phys-
RevB.98.085303.
[46] Ming Lai Chan. Optimized protocol
to create repeater graph states for all-
photonic quantum repeater. arXiv preprint
arXiv:1811.10214, 2018.
[47] Mercedes Gimeno-Segovia, Terry Rudolph,
and Sophia E Economou. Determinis-
tic generation of large-scale entangled pho-
tonic cluster state from interacting solid
state emitters. Physical review letters,
123(7):070501, 2019. DOI: 10.1103/Phys-
RevLett.123.070501.
[48] Johannes Borregaard, Hannes Pichler, Tim
Schröder, Mikhail D Lukin, Peter Lodahl,
and Anders S Sørensen. One-way quan-
tum repeater based on near-deterministic
photon-emitter interfaces. Physical Review
X, 10(2):021071, 2020. DOI: 10.1103/Phys-
RevX.10.021071.
[49] S Lloyd, MS Shahriar, JH Shapiro, and
PR Hemmer. Long distance, uncondi-
tional teleportation of atomic states via com-
plete bell state measurements. Physical Re-
view Letters, 87(16):167903, 2001. DOI:
10.1103/PhysRevLett.87.167903.
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 13
[50] Yoon-Ho Kim, Sergei P Kulik, and Yan-
hua Shih. Quantum teleportation of a
polarization state with a complete bell
state measurement. Physical Review Let-
ters, 86(7):1370, 2001. DOI: 10.1103/Phys-
RevLett.86.1370.
[51] Yoon-Ho Kim, SERGEI KULIK, and Yan-
hua Shih. Quantum teleportation with a
complete bell state measurement. Journal of
Modern Optics, 49(1-2):221–236, 2002. DOI:
10.1080/09500340110087633.
[52] Warren P Grice. Arbitrarily complete bell-
state measurement using only linear optical
elements. Physical Review A, 84(4):042331,
2011. DOI: 10.1103/PhysRevA.84.042331.
[53] Fabian Ewert and Peter van Loock. 3/4-
efficient bell measurement with passive lin-
ear optics and unentangled ancillae. Physical
review letters, 113(14):140403, 2014. DOI:
10.1103/PhysRevLett.113.140403.
[54] Stephen Wein, Khabat Heshami, Christo-
pher A Fuchs, Hari Krovi, Zachary Dut-
ton, Wolfgang Tittel, and Christoph Si-
mon. Efficiency of an enhanced linear opti-
cal bell-state measurement scheme with re-
alistic imperfections. Physical Review A,
94(3):032332, 2016. DOI: 10.1103/Phys-
RevA.94.032332.
[55] Leigh S Martin and K Birgitta Whaley.
Single-shot deterministic entanglement be-
tween non-interacting systems with linear
optics. arXiv preprint arXiv:1912.00067,
2019.
[56] Marc Hein, Jens Eisert, and Hans J Briegel.
Multiparty entanglement in graph states.
Physical Review A, 69(6):062311, 2004. DOI:
10.1103/PhysRevA.69.062311.
[57] Masahiro Takeoka, Saikat Guha, and
Mark M Wilde. Fundamental rate-loss trade-
off for optical quantum key distribution. Na-
ture communications, 5(1):1–7, 2014. DOI:
10.1038/ncomms6235.
[58] Stefano Pirandola, Riccardo Laurenza, Carlo
Ottaviani, and Leonardo Banchi. Funda-
mental limits of repeaterless quantum com-
munications. Nature communications, 8(1):
1–15, 2017. DOI: 10.1038/ncomms15043.
[59] Peter W Shor and John Preskill. Simple
proof of security of the bb84 quantum key
distribution protocol. Physical review let-
ters, 85(2):441, 2000. DOI: 10.1103/Phys-
RevLett.85.441.
[60] Valerio Scarani, Helle Bechmann-
Pasquinucci, Nicolas J Cerf, Miloslav
Dušek, Norbert Lütkenhaus, and Momtchil
Peev. The security of practical quantum key
distribution. Reviews of modern physics,
81(3):1301, 2009. DOI: 10.1103/RevMod-
Phys.81.1301.
[61] Jian-Wei Pan, Christoph Simon, Časlav
Brukner, and Anton Zeilinger. Entangle-
ment purification for quantum communica-
tion. Nature, 410(6832):1067, 2001. DOI:
10.1038/35074041.
[62] M Zwerger, W Dür, and HJ Briegel.
Measurement-based quantum repeaters.
Physical Review A, 85(6):062326, 2012.
DOI: 10.1103/PhysRevA.85.062326.
[63] M Zwerger, HJ Briegel, and W Dür.
Measurement-based quantum communica-
tion. Applied Physics B, 122(3):50, 2016.
DOI: 10.1007/s00340-015-6285-8.
[64] Alex Greilich, Samuel G. Carter, Danny
Kim, Allan S. Bracker, and Daniel Gammon.
Optical control of one and two hole spins in
interacting quantum dots. Nature Photon-
ics, 5(11):702–708, November 2011. DOI:
10.1038/nphoton.2011.237.
[65] Danny Kim, Samuel G Carter, Alex Greilich,
Allan S Bracker, and Daniel Gammon.
Ultrafast optical control of entanglement
between two quantum-dot spins. Na-
ture Physics, 7(3):223–229, 2011. DOI:
10.1038/nphys1863.
[66] A. Greilich, A. Shabaev, D. R. Yakovlev,
Al. L. Efros, I. A. Yugova, D. Reuter, A. D.
Wieck, and M. Bayer. Nuclei-induced fre-
quency focusing of electron spin coherence.
Science, 317(5846):1896–1899, 2007. DOI:
10.1126/science.1146850.
[67] Xiaoya Judy Wang, Stefano Chesi, and
W. A. Coish. Spin-echo dynamics of a
heavy hole in a quantum dot. Phys. Rev.
Lett., 109:237601, 2012. DOI: 10.1103/Phys-
RevLett.109.237601.
[68] Daniel Najer, Immo Söllner, Pavel Sekatski,
Vincent Dolique, Matthias C Löbl, Daniel
Riedel, Rüdiger Schott, Sebastian Starosi-
elec, Sascha R Valentin, Andreas D Wieck,
et al. A gated quantum dot strongly coupled
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 14
to an optical microcavity. Nature, pages 1–1,
2019. DOI: 10.1038/s41586-019-1709-y.
[69] GD Fuchs, Guido Burkard, PV Klimov, and
DD Awschalom. A quantum memory intrin-
sic to single nitrogen–vacancy centres in dia-
mond. Nature Physics, 7(10):789–793, 2011.
DOI: 10.1038/nphys2026.
[70] Dmitry Solenov, Sophia E Economou, and
Thomas L Reinecke. Two-qubit quantum
gates for defect qubits in diamond and
similar systems. Physical Review B, 88
(16):161403, 2013. DOI: 10.1103/Phys-
RevB.88.161403.
[71] ML Goldman, TL Patti, D Levonian,
SF Yelin, and MD Lukin. Optical control of a
single nuclear spin in the solid state. Physical
Review Letters, 124(15):153203, 2020. DOI:
10.1103/PhysRevLett.124.153203.
[72] Xing Rong, Jianpei Geng, Fazhan Shi, Ying
Liu, Kebiao Xu, Wenchao Ma, Fei Kong,
Zhen Jiang, Yang Wu, and Jiangfeng Du.
Experimental fault-tolerant universal quan-
tum gates with solid-state spins under am-
bient conditions. Nature communications,
6(1):1–7, 2015. DOI: 10.1038/ncomms9748
(2015).
[73] Nir Bar-Gill, Linh M Pham, Andrejs
Jarmola, Dmitry Budker, and Ronald L
Walsworth. Solid-state electronic spin co-
herence time approaching one second. Na-
ture communications, 4:1743, 2013. DOI:
10.1038/ncomms2771.
[74] CJ Ballance, TP Harty, NM Linke, MA Se-
piol, and DM Lucas. High-fidelity quan-
tum logic gates using trapped-ion hyper-
fine qubits. Physical review letters, 117
(6):060504, 2016. DOI: 10.1103/Phys-
RevLett.117.060504.
[75] Ye Wang, Mark Um, Junhua Zhang, Shuom-
ing An, Ming Lyu, Jing-Ning Zhang, L-
M Duan, Dahyun Yum, and Kihwan Kim.
Single-qubit quantum memory exceeding
ten-minute coherence time. Nature Photon-
ics, 11(10):646, 2017. DOI: 10.1038/s41566-
017-0007-1.
[76] Severin Daiss, Stephan Welte, Bastian
Hacker, Lin Li, and Gerhard Rempe. Single-
photon distillation via a photonic parity
measurement using cavity qed. Physical re-
view letters, 122(13):133603, 2019. DOI:
10.1103/PhysRevLett.122.133603.
[77] C. Y. Hu, A. Young, J. L. O’Brien, W. J.
Munro, and J. G. Rarity. Giant optical Fara-
day rotation induced by a single-electron
spin in a quantum dot: Applications to
entangling remote spins via a single pho-
ton. Physical Review B, 78(8):085307, Aug
2008. DOI: 10.1103/PhysRevB.78.085307.
URL http://prb.aps.org/abstract/PRB/
v78/i8/e085307.
[78] Kazuki Koshino, Satoshi Ishizaka, and Ya-
sunobu Nakamura. Deterministic photon-
photon swap gate using a λ system. Phys-
ical Review A, 82(1):010301, 2010. DOI:
10.1103/PhysRevA.82.010301.
[79] Serge Rosenblum, Scott Parkins, and Barak
Dayan. Photon routing in cavity QED: Be-
yond the fundamental limit of photon block-
ade. Physical Review A, 84:033854, 2011.
DOI: 10.1103/PhysRevA.84.033854.
[80] Shuo Sun, Hyochul Kim, Glenn S Solomon,
and Edo Waks. A quantum phase switch
between a single solid-state spin and a pho-
ton. Nature Nanotechnology, 11(6):539–544,
2016. DOI: 10.1038/nnano.2015.334.
[81] LM Wells, Sokratis Kalliakos, Bruno Villa,
DJP Ellis, RM Stevenson, AJ Bennett, Ian
Farrer, DA Ritchie, and AJ Shields. Pho-
ton phase shift at the few-photon level and
optical switching by a quantum dot in a mi-
crocavity. Physical Review Applied, 11(6):
061001, 2019. DOI: 10.1103/PhysRevAp-
plied.11.061001.
[82] Petros Androvitsaneas, Andrew Young,
Joseph Lennon, Christian Schneider, Sebas-
tian Maier, Janna Hinchliff, George Atkin-
son, Edmund Harbord, Martin Kamp, Sven
Höfling, et al. An efficient quantum photonic
phase shift in a low q-factor regime. ACS
Photonics, 2019. DOI: 10.1021/acsphoton-
ics.8b01380.
[83] Robert H Hadfield. Single-photon detec-
tors for optical quantum information appli-
cations. Nature photonics, 3(12):696, 2009.
DOI: 10.1038/nphoton.2009.230.
[84] Yuan Zhan and Shuo Sun. Deterministic
generation of loss-tolerant photonic cluster
states with a single quantum emitter. Phys-
ical Review Letters, 125(22):223601, 2020.
DOI: 10.1103/PhysRevLett.125.223601.
[85] C. Y. Hu, A. Young, J. L. O’Brien, W. J.
Munro, and J. G. Rarity. Giant optical Fara-
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 15
day rotation induced by a single-electron
spin in a quantum dot: Applications to en-
tangling remote spins via a single photon.
Physical Review B, 78(8):085307, 2008. DOI:
10.1103/PhysRevB.78.085307.
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 16
Supplementary Materials: Resource requirements for efficient quantum com-
munication using all-photonic graph states generated from a few matter qubits
Bounds for memory-based repeater protocols
As discussed in the main text, the rate of a memory-based quantum repeater based on heralded
entanglement generation is limited by the rate hT
QM
i
1
at which a given memory qubit can be re-
used. hT
QM
i is given by the sum of the average time required for generating entanglement hT
ent
i and
the time during which the entanglement is stored hT
store
i. hT
ent
i is given by T
trial
/P
ent
, where T
trial
is
the time required for a heralded entanglement generation attempt, and P
ent
is the probability that an
attempt is successful. Let’s consider three heralded entanglement procedures [2527]:
In Ref. [26], T
trial
= L
0
/c, and the probability is P
ent
= e
L
0
/L
att
/2 (assuming perfect photon
emission and detection probability).
In Ref. [27], T
trial
= 2L
0
/c, and the probability is P
ent
= e
L
0
/L
att
.
In Ref. [25], T
trial
= L
0
/c, and the probability is P
ent
e
L
0
/(2L
att
)
. However, the protocol needs
to be implemented in a regime where the probability to emit a photon is very small to avoid
two-photon emission, which degrades the entanglement fidelity. Consequently, P
ent
1/2.
Thus in all these cases, if we consider small values of L
0
L
att
, we have:
hT
ent
i
2L
0
c
e
L
0
/L
att
2L
0
c
. (17)
This limitation can be understood from the fact that the photonic qubit and the classical signal need
to travel back and forth either to a measurement node (twice the distance L
0
/2) or to the next QR
(twice the distance L
0
). If the photon travels to a quantum repeater, it is possible to deterministically
entangle the QM and the photon, so the entanglement probability is only bounded by the photonic
losses. If the photon travels to an intermediate node, the entanglement generation procedure uses
linear optics and has a success probability that cannot reach more than 1/2 (unless ancillary photonic
qubits are used, but this requires another quantum emitter to produce them [5254]).
From this limitation, we find the upper bound R
(QM)
max
max(R/N
m
) in the case of QR protocols
with two QMs. In this case, the rate R is limited by hT
QM
i
1
hT
ent
i
1
= c/(2L
0
). The number
of matter qubits is given by twice the number of repeaters plus Alice and Bob’s matter qubits, thus
N
m
= 2(N
QR
+ 1) = 2L/L
0
, therefore:
R
(QM)
max
= c/4L. (18)
In the case of multiplexed repeaters, by multiplying the number of memories per repeater by N, the
entangling rate between two repeaters is bounded by Nc/(2L
0
), but the number of matter qubits is
now N
m
= 2NL/L
0
, so that the total rate is still bounded by R
(QM)
max
.
Ref. [55] also proposes a method to generate heralded entanglement deterministically by detecting
photons. The idea is to excite the two quantum emitters (each consisting of two ground states and one
excited state that couples to only one of the ground states) such that they both emit a photon whose
which-path information is erased using linear optics. When the first photon is detected a signal is sent
to each quantum emitter to perform a rotation between the ground states. This rotation needs to be
performed before the second photon is emitted. Therefore, the latency between the photon detection
and the rotation should be much smaller than the lifetime of the quantum emitter’s excited state. A
direct implication of this is that the photon detectors should be placed very close to the quantum
emitters, which is impractical. For example, for quantum emitters with a lifetime under 10 ns (which
is typical for QDs and defects in diamond), the position of the detectors should be much smaller than
1 m. In addition, to obtain high fidelities with this scheme, the two quantum emitters are excited for
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 17
a typical time T
1
which should be much larger than the classical signaling time L
0
/c, and thus the
previous analysis should still hold in that case.
In the case of memory-based schemes using heralded entanglement and 2 QMs per QR, it is possible
to find a stricter upper bound. To derive this bound, consider two repeater nodes labeled 1 and 2.
Node 1 contains two memory qubits QM1 and QM1
0
, while node 2 contains QM2 and QM2
0
. Suppose
that we succeed in entangling QM1 and QM2 via a Bell measurement at an intermediate measurement
node. How long do we have to maintain this entanglement before further entanglement swappings at
neighboring nodes free up QM1 and QM2, allowing them to be re-used for creating the next Bell pair
to be shared between Alice and Bob? The answer to this question is what we called hT
store
i above,
the average amount of time we have to store the entanglement. The upper bound on the rate is then
hT
QM
i
1
, where hT
QM
i = hT
ent
i + hT
store
i. We obtained a bound for hT
ent
i above and used this to
derive the bound on the rate given in Eq. (18). There, we neglected the role of hT
store
i. Here, we
reinstate this quantity to obtain a stricter bound.
To determine hT
store
i, we first note that the entanglement between QM1 and QM2 needs to be
maintained until QM1
0
and QM2
0
become entangled with memory qubits on other nodes. Once this
happens, the entanglement can then be swapped, leaving QM1 and QM2 completely disentangled and
ready for re-use. Of course, QM1
0
and QM2
0
could already be entangled with other nodes by the time
QM1 and QM2 become entangled. Thus, there are four equally probable cases to consider:
If QM1
0
and QM2
0
are already connected to other nodes, we immediately swap the entanglement
and thus do not wait any longer. In this case hT
case1
store
i = 0.
If one QM (either QM1
0
or QM2
0
) is already connected to other nodes, we need to wait an average
time of hT
ent
i for the second QM to become entangled as well. In this case, hT
case2
store
i = hT
case3
store
i =
hT
ent
i.
If neither QM1
0
or QM2
0
are connected, we need to wait for both to become connected. In this
case, hT
case4
store
i =
32P
ent
2P
ent
× hT
ent
i.
The expression for hT
case4
store
i above requires some explanation. It comes from the fact that both QM1
0
and QM2
0
become successfully entangled with other nodes after n attempts with probability
Pr(n, P
ent
) = [1 (1 P
ent
)
n
]
2
. (19)
The average storage time in this case then depends on the average number of attempts needed before
entanglement is successfully created and on the time T
trial
each attempt takes:
hT
case4
store
i = T
trial
X
n=0
n[Pr(n, P
ent
) Pr(n 1, P
ent
)] =
3 2P
ent
2 P
ent
hT
ent
i. (20)
This implies that the overall average storage time is
hT
store
i =
hT
case1
store
i + hT
case2
store
i + hT
case3
store
i + hT
case4
store
i
4
=
7/4 P
ent
2 P
ent
hT
ent
i. (21)
The upper bound on the rate per matter qubit is therefore
R
(2QM)
max
=
1
(hT
ent
i + hT
store
i)N
m
=
2 P
ent
15 8P
ent
c
L
, (22)
where we set hT
ent
i 2L
0
/c and N
m
= 2L/L
0
as appropriate for the memory-based scheme. If we set
P
ent
1 [85], this becomes
R
(2QM)
max
=
c
7L
for P
ent
= 1. (23)
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 18
Measurement
node
Alice Bob
(a) (b)
(c)
Figure 5: (a) Recursive definition of the quantum state |RGS
k
i. (b) Realization of a bridge between two RGSs
|RGS
k,bridge
i. (c) Linear cluster state shared between Alice, Bob and all the measurement nodes.
Classical signaling does not limit the repetition rate of the RGS protocol
In this section, we show that classical signaling does not limit the repetition rate of the RGS protocol.
This is important to consider because some of the X and Z measurements that are performed on the
RGS will, depending on the measurement outcomes, leave the remaining photons in a state that is
only locally equivalent to a graph state. In this case, additional single-qubit gates must be applied
to bring the state to a proper graph state before the next measurements are performed. Otherwise,
the bases for these subsequent measurements will be effectively rotated, and the measurements will
not modify the state in the intended way. In the following, we show that the extra single-qubit gates
only need to be applied on qubits located at the same measurement node where the measurements
on which they depend occur. Thus, classical communication between nodes is not needed. The only
exception to this occurs at the very end of the protocol, when we are left with a linear cluster state
in which the two end qubits are those of Alice and Bob. The final step is to turn this state into a
Bell pair shared by Alice and Bob by performing X measurements on the inner qubits of the linear
cluster state. Afterward, Alice and Bob’s qubits form a Bell pair whose Pauli frame depends on the
X measurement outcomes. This information therefore needs to be transferred to them using classical
signaling. However, the protocol can be repeated before this classical signal is received and thus, the
classical signaling does not limit the repetition rate. The latter is only limited by the generation time
of RGSs.
Recursive construction of a repeater graph state.
We denote the RGS with k arms by |RGS
k
i. For the purposes of this argument, we allow k to be
either odd or even. We want to define this quantum state recursively. As illustrated in Fig. 5(a), we
can make this state by starting with an RGS with k 1 arms and then adding one new 1
st
-leaf qubit
and one new 2
nd
-leaf qubit by applying CZ gates between these two new qubits and between the new
1
st
-leaf qubit and all the existing 1
st
-leaf qubits in |RGS
k1
i. In other words, we can start from a
2-qubit cluster state and |RGS
k1
i in a product state: (|0
k
1
, +
k
2
i + |1
k
1
,
k
2
i) |RGS
k1
i. Here, k
1
and k
2
denote the 1
st
-leaf qubit and the 2
nd
-leaf qubit from the k
th
arm. We ignore the normalization
factors troughout this section. The remaining step is to apply CZ gates between k
1
and all the 1
st
-leaf
qubits in |RGS
k1
i, which yields
|RGS
k
i = |0
k
1
, +
k
2
i |RGS
k1
i + |1
k
1
,
k
2
i
RGS
k1
E
, (24)
where we have introduced the notation
RGS
k
E
k
Y
i=1
Z
i
1
|RGS
k
i . (25)
Here, Z
i
j
is the Pauli Z operator applied to the j
th
-leaf qubit of the i
th
arm of the RGS.
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 19
Note that if the RGS has a logical tree encoding, then instead of attaching a 2-qubit cluster state to
|RGS
k1
i, we need to attach a star graph with b
0
arms (in the case of a depth-two tree), where b
0
is the
first branching parameter of the tree. To create |RGS
k
i, each arm of the star graph needs to connect
to every 1
st
-leaf qubit of |RGS
k1
i; this requires a total of b
0
(k 1) CZ gates. The resulting RGS
can still be decomposed as in Eq. (24), but we need to replace |0
k
1
i by a sum over all the even-parity
bit strings on b
0
qubits, and |1
k
1
i needs to be replaced by all the odd-parity bit strings on b
0
qubits.
In what follows, we will continue to use Eq. (24) with the understanding that the k
1
qubit states need
to be replaced in this manner if a logical tree encoding is included. This does not otherwise alter the
analysis below.
Bell state measurement
Let’s now consider the effect of a Bell measurement on 2
nd
-leaf qubits from two different RGSs. We
suppose that one of these qubits resides in arm k of one RGS, and the other in arm k
0
of the other RGS.
To analyze this measurement, we expand the two-RGS product state using the recursive definition of
an RGS shown above:
|RGS
k
i
RGS
0
k
= |RGS
k1
i |0
k
1
, +
k
2
, +
k
0
2
, 0
k
0
1
i
RGS
0
k1
+ |RGS
k1
i |0
k
1
, +
k
2
,
k
0
2
, 1
k
0
1
i
RGS
0
k1
E
+
RGS
k1
E
|1
k
1
,
k
2
, +
k
0
2
, 0
k
0
1
i
RGS
0
k1
+
RGS
k1
E
|1
k
1
,
k
2
,
k
0
2
, 1
k
0
1
i
RGS
0
k1
E
.
(26)
We want to perform a Bell state measurement on qubits k
2
and k
0
2
. After the Bell measurement, we
want to obtain the state:
|RGS
k,bridge
i = |RGS
k1
i |0
k
1
, 0
k
0
1
i
RGS
0
k1
+ |RGS
k1
i |0
k
1
, 1
k
0
1
i
RGS
0
k1
E
+
RGS
k1
E
|1
k
1
, 0
k
0
1
i
RGS
0
k1
RGS
k1
E
|1
k
1
, 1
k
0
1
i
RGS
0
k1
E
,
(27)
which corresponds to the state represented in Fig. 5(b) where an edge replaces a pair of 2
nd
-leaf qubits
shared by two RGSs.
The Bell state measurements are realized in the basis:
φ
±
= |0, +i ± |1, −i ,
ψ
±
= |1, +i ± |0, −i .
(28)
These are not the standard Bell states but are related to them by a Hadamard gate on the second
qubit. By applying such a Hadamard gate, a standard Bell state analyzer can be used. Using this
basis for the 2
nd
-leaf qubits k
2
and k
0
2
, we can rewrite the state before the Bell measurement as
|RGS
k
i
RGS
0
k
= |RGS
k1
i |0i
k
1
φ
+
E
+
φ
+
ψ
+
E
+
ψ
k
2
,k
0
2
|0i
k
0
1
RGS
0
k1
+ |RGS
k1
i |0i
k
1
φ
+
E
φ
+
ψ
+
E
ψ
k
2
,k
0
2
|1i
k
0
1
RGS
0
k1
E
+
RGS
k1
E
|1i
k
1
φ
+
E
+
φ
ψ
+
E
ψ
k
2
,k
0
2
|0i
k
0
1
RGS
0
k1
+
RGS
k1
E
|1i
k
1
φ
+
E
+
φ
+
ψ
+
E
ψ
k
2
,k
0
2
|1i
k
0
1
RGS
0
k1
E
.
(29)
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 20
Suppose that we can resolve only the two states |ψ
±
i with a linear optics Bell state analyzer.
Depending on the measurement outcomes, the state becomes:
ψ
Z
k
1
Z
k
0
1
|RGS
k,bridge
i ,
ψ
+
E
Z
k
1
|RGS
k,bridge
i .
(30)
Thus, we find that if the outcome is
ψ
+
(respectively |ψ
i) we need to apply a Z rotation on qubit
k
1
(on qubits k
1
and k
0
1
) to recover the desired state. Note that this qubit is at the measurement node
where we perform the Bell measurement.
Z measurements
If the measurement fails or if one Bell measurement has already succeeded, we need to perform Z
measurements on 1
st
-leaf qubits to disconnect them and their associated 2
nd
-leaf qubits from the
graph. Depending on the outcomes of these measurements, the resulting state is one of four possible
states:
|RGS
k1
i
RGS
0
k1
,
RGS
k1
E
RGS
0
k1
,
|RGS
k1
i
RGS
0
k1
E
,
RGS
k1
E
RGS
0
k1
E
.
(31)
Note that if a Bell measurement has already succeeded, then each of the above outcomes needs to be
acted upon by a CZ gate on the two 1
st
-leaf qubits associated with the successful measurement. We
leave this possible CZ gate implicit in the following. Ideally, we want to recover |RGS
k1
i
RGS
0
k1
E
.
Therefore, we need to be able to convert a state
RGS
k1
E
into |RGS
k1
i by performing measurements
only on qubits present in the measurement node. Let’s consider any other arm l of the RGS that is
present at the same measurement node and rewrite the state as
|RGS
k1
i = |0
l
1
, +
l
2
i |RGS
k2
i + |1
l
1
,
l
2
i
RGS
k2
E
. (32)
By applying Z gates on all the 1
st
-leaf qubits, we also have:
RGS
k1
E
= |0
l
1
, +
l
2
i
RGS
k2
E
|1
l
1
,
l
2
i) |RGS
k2
i . (33)
A comparison of these two results leads to the identity:
Y
l
1
Z
l
2
RGS
k1
E
= i |RGS
k1
i . (34)
By applying local rotations of qubits situated at the same measurement node, it is therefore possible to
recover the desired state. The last two Z measurements on the two qubits denoted k
1
and k
0
1
requires
special care as the only two other qubits, l
1
and l
0
1
, positioned at this measurement node, are the
one which connect the two RGSs (if a Bell measurement has succeeded) and should undergo an X
measurement. In that case, it is still possible to use l
1
and l
0
1
to recover the desired graph state by
performing single-qubit rotations that depends on the measurement outcomes of k
1
and k
0
1
:
0
k
1
, 0
k
0
1
E
I
l
1
I
l
0
1
,
0
k
1
, 1
k
0
1
E
iZ
l
1
Y
l
0
1
,
1
k
1
, 0
k
0
1
E
iY
l
1
Z
l
0
1
,
1
k
1
, 1
k
0
1
E
X
l
1
X
l
0
1
,
(35)
where I
i
denotes the identity on qubit i.
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 21
X measurements and classical signaling
If at least one Bell measurement succeeds at each measurement node and if all the required Z mea-
surements succeed, we are left with a linear cluster state with two 1
st
-leaf qubits at each measurement
node (see Fig. 5(c)). The next step is to perform X measurements on these qubits. Performing X
measurements on two connected qubits of a graph state has the effect of connecting each qubit in the
neighborhood of the first qubit to all the qubits in the neighborhood of the second one [35]. Therefore,
after the X measurements of all the qubits at each measurement node, the two remaining qubits shared
by Alice and Bob are in a maximally entangled state. More precisely, this state is equivalent to a Bell
pair up to local qubit rotations. It is important to note that all the X measurements can be performed
independently of each other. Although the state can deviate from a proper graph state depending
on the measurement outcomes, it is straightforward to show that the state only differs by at most
local Z gates. This can effectively convert some of the X measurements to Y measurements, but both
types of measurements have a similar effect on the state, and a Bell pair is still obtained once all the
measurements are completed. However, the precise state depends on all the measurement outcomes of
each X measurement. This information needs to be sent via a classical channel to Alice and Bob to
recover the desired state.
This classical signal does not limit the repetition rate of the protocol, which can either be used for
quantum teleportation of quantum states or for QKD. Indeed, for quantum teleportation, two quantum
memories situated at Alice’s and Bob’s nodes can be used to store the Bell pair while waiting for the
classical signal to arrive. In that case, the protocol can still be repeated at a high rate as long as
there are enough quantum memories at Alice’s and Bob’s nodes to store the states that are already
transferred. It is still the case that the protocol does not require quantum memories at the nodes
inside the network, which is a clear resource improvement compared to the memory-based approach.
For QKD, Alice’s and Bob’s qubits can be measured in a random basis upon arrival at their nodes,
and they don’t need to be stored in quantum memories until the classical signaling is received. The
classical signaling is still required to recover the Pauli frame of the measured qubits and to check if
Alice’s and Bob’s qubits were measured in compatible basis for QKD.
Error correction using tree graph states
In the main text, we have shown that logical encoding with tree graph states allows measuring a qubit
with a probability Pr(M
Z,`
) or Pr(M
X,`
) that is higher than the single-photon measurement probability
P
ph
. We now show that the logical encoding based on tree graphs also increases the tolerance of the
RGS protocol to single-qubit errors. Indeed, if m
k
indirect measurements are performed on a qubit
at level k in the tree, it is possible to use a majority vote of the measurements to reduce the effect
of errors. Suppose that all b
k
of the possible indirect measurements are performed, where b
k
is the
k
th
branching parameter of the tree. The probability that exactly m
k
of these measurements succeed
(perhaps with errors) is
p
k
(m
k
) =
b
k
m
k
!
s
k
m
k
(1 s
k
)
b
k
m
k
, (36)
where s
k
is the probability that a single indirect measurement succeeds.
The average error after b
k
indirect measurement attempts at level k is:
e
I
k
=
1
r
k
b
k
X
m
k
=1
p
k
(m
k
)e
I
k
|m
k
, (37)
where e
I
k
|m
k
is the average error in the case of m
k
successful measurements, and r
k
is the probability
that at least one indirect measurement succeeds. For m
k
indirect measurements, an error still occurs
in the majority vote if more than half of the indirect measurements (m
k
/2) are faulty. Thus, if m
k
is
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 22
odd:
e
I
k
|m
k
= Pr(N
errors
> m
k
/2)
=
m
k
X
j=dm
k
/2e
m
k
j
!
(e
I
k
|1
)
j
(1 e
I
k
|1
)
m
k
j
,
(38)
where e
I
k
|1
is the average error of a single indirect measurement. If m
k
is even, we have:
e
I
k
|m
k
=
m
k
1
X
j=dm
k
/2e
m
k
1
j
!
(e
I
k
|1
)
j
(1 e
I
k
|1
)
m
k
1j
. (39)
The sum goes only up to m
k
1 because we can randomly remove one result to return to the odd-number
case and re-use the above formula for the probability.
The single indirect Z measurement error e
I
k
|1
of a qubit A at level k depends on the error of the
X measurement of qubit B at level k + 1 (which we take to be a constant error probability ) and on
the errors of the b
k+1
measurements on the qubits at level k + 2, denoted C
i
in Fig. 2(a). These can
either be direct (with error ) or indirect (with error e
I
k+2
). Let n
k
be the number of qubits that are
only measured directly (with error ) while b
k+1
n
k
qubits are measured indirectly (with error e
I
k+2
).
e
I
k
|1
is thus given by:
e
I
k
|1
=
b
k+1
X
n
k
=0
b
k+1
n
k
!
Pr(I
Z,k+2
|M
Z,k+2
)
b
k+1
n
k
(1 Pr(I
Z,k+2
|M
Z,k+2
))
n
k
e
n
k
, (40)
where e
n
k
is the error in the case of n
k
direct qubit measurements, and
Pr(I
Z,k
|M
Z,k
) =
Pr(I
Z,k
)Pr(M
Z,k
|I
Z,k
)
Pr(M
Z,k
)
=
r
k
Pr(M
Z,k
)
. (41)
Here, I
Z,k
denotes a successful indirect Z measurement of a photon at level k and thus Pr(I
Z,k
) is
what is called r
k
in the main text. Therefore, Pr(I
Z,k
|M
Z,k
) is the probability that a photon at level k
was successfully measured indirectly (I
Z,k
) given that it has been successfully measured (M
Z,k
). Note
that a photon can be both successfully directly and indirectly measured. In that case, we keep the
indirect measurement outcome which should have a smaller error e
I
k
compared to the intrinsic single
qubit error , thanks to the error correction.
The only remaining quantity to determine is the error e
n
k
. The X measurement on qubit B and the
n
k
direct Z measurements on the qubits C
i
are performed with error , while the remaining b
k+1
n
k
indirect Z measurements are performed with error e
I
k+2
. Note that the indirect measurement outcome
is given by the net parity of all the X and Z measurement outcomes. Thus, the indirect measurement
can still yield the correct outcome even if there are errors in the X and Z measurements, provided
there are an even number of errors so that the parity remains unchanged. Only an odd number of
errors causes the indirect measurement to fail. Thus:
e
n
k
=
n
k
+1
X
i=0
n
k
+ 1
i
!
i
(1 )
n
k
+1i
b
k+1
n
k
X
j=0,
i+j=1[2]
b
k+1
n
k
j
!
e
I
k+2
j
(1 e
I
k+2
)
b
k+1
n
k
j
(42)
In this equation, we take into account the fact that only odd numbers of measurement errors con-
tribute to e
n
k
by restricting the second sum to only values of i + j = 1[2]. This is different than what
was obtained in Ref. [31], perhaps due to additional approximations in that reference.
As shown in Fig. 6(a), the logical Z measurement M
Z,`
succeeds if all the 1
st
-level qubits are measured
directly or indirectly in the Z basis. The logical X measurement M
X,`
succeeds if any of the 1
st
level
qubits is measured in the X basis and all its corresponding 2
nd
level qubits are successfully measured
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 23
(a) (b)
Figure 6: (a) Logical Z measurement M
Z,`
at the physical level. (b) Logical X measurement M
X,`
at the physical
level. The parity of the Z measurements on the qubits in the green area is the same as the logical M
X,`
measure-
ment on the logical qubit (purple) and is the same as the parity measurement on the two light blue areas (where
measurement bases are indicated next to each qubit).
in the Z basis either directly or indirectly, as shown in Ref. [31]. To understand this, let’s denote by
L the logical qubit that is connected to a set A = {A
1
, A
2
, ...} of other qubits, represented by the
green area in Fig. 6(b). Because X
L
Z
A
= X
L
i
Z
A
i
is a stabilizer of the RGS at the logical level, the
logical X measurement of qubit L, M
X,`
, has the same outcome as Z
A
, i.e. the parity of the product
of all Z measurements on the qubits in A. At the physical level, the logical qubit is encoded with
a tree graph state where we consider a first-level qubit B and its set of neighbor qubits in the tree
C(B). X
B
Z
C(B)
Z
A
is a stabilizer of the RGS at the physical level and thus X
B
Z
C(B)
(represented by
either of the light blue areas in Fig. 6(b)) has the same parity as Z
A
and therefore constitutes an X
measurement of the logical qubit L.
The logical X and Z errors are thus given by:
¯e
X
= e
I
0
,
¯e
Z
=
b
0
X
n=0
b
0
n
!
Pr(I
1
|Z
1
)
b
0
n
(1 Pr(I
1
|Z
1
))
n
e
n
,
with e
n
=
n
X
i=0
n
i
!
i
(1 )
ni
b
0
n
X
j=0,
i+j=1[2]
b
0
n
j
!
e
I
1
j
(1 e
I
1
)
b
0
nj
.
(43)
The derivation of ¯e
Z
follows the same idea as e
I
k
|1
(for "k = 1") except that no X measurements are
needed here.
Fidelity of the full protocol
We have used the results already derived in the supplementary materials of Ref. [31] to calculate the
global fidelity of the generated entangled pair of photons shared by Alice and Bob. The total fidelity
F
AB
is given by
F
AB
= 1 (E
x
+ E
y
+ E
z
), (44)
with
E
x
= E
z
=
1
4
1 (1 2)
2(N
QR
+1)
(1 2¯e
X
)
2N
QR
,
E
y
=
1
4
1 + (1 2)
2(N
QR
+1)
(1 2¯e
X
)
2N
QR
2(1 2)
2(N
QR
+1)
(1 2¯e
X
)
N
QR
(1 2¯e
Z
)
(2m2)N
QR
.
(45)
Accepted in Quantum 2021-02-03, click title to verify. Published under CC-BY 4.0. 24