2003 Book QuantumInformationWithContinuo PDF
2003 Book QuantumInformationWithContinuo PDF
Quantum Information
with Continuous Variables
Edited by
SAMUEL L. BRAUNSTEIN
School of Informatics,
University of Wales, Bangor. United Kingdom
and
ARUN K. PAT!
Institute of Physics, Orissa, India
and
Theoretical Physics Division,
BARe, Mumbai, India
ISBN 978-90-481-6255-0
Cover illustration:
reprinted with permission from Nature [Nature 413, 400-403 (2001)]
Copyright (2001) Macmillan Pub1ishers Ltd.
Softcover reprint of the hardcover 1st edition 2001
Preface Xl
18. Atomic continuous variable processing and light-atoms quantum interface 231
Alex Kuzmich and Eugene S. Polzik
Index 423
Preface
SAMUEL L. BRAUNSTEIN
ARUN K. PAT!
About the Editors
xiii
I
QUANTUM COMPUTING
Chapter 1
Samuel L. Braunstein
Informatics, Bangor University, Bangor LL57 I UT, United Kingdom
schmuel @sees.bangor.ac.uk
Arun K. Pati
Institute of Physics, Bhubaneswar-75IO05, Orissa, INDIA
Theoretical Physics Division, BARe, Mumbai, INDIA
akpati @iopb.res.in
This book will introduce the reader to the area of quantum information
processing with continuous variables. However, to put it into some context
with the "conventional" approach to quantum computing with qubits we shall
give a brief introduction to its basic principles skipping all details. Briefly,
we will touch on notions of bits, qubits, quantum parallelism, and quantum
algorithms such as the Deutsch-Jozsa, the Shor factoring problem, and the
Grover quantum search.
Quantum information theory is a marriage between two scientific pillars
of modern science, namely, quantum theory and classical information theory.
Quantum theory as developed by Planck, Einstein, Schrodinger, Dirac, Heisen-
berg and many others in the early part of the last century is one of the finest
theories that explains phenomena ranging from molecules to electrons, protons,
neutrons and other microscopic particles. The mathematical theory of classi-
cal information was put forth by Shannon in the mid part of the last century.
Whatever revolution in information technology we see at present is partly due
to the ground breaking work by Shannon, Turing, Church and others.
When the ideas from information theory are carried over to quantum theory
there emerges a revolution in our ability to process information. Ultimately, the
basic ways of expressing and manipulating information require physical states
3
SoL Braunstein and A.K. Pati (eds.), Quantum ['!!ormation with Continuous Variables, 3-8.
© 2003 Kluwer Academic Publishers.
4 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
and processes. In quantum theory we know that physical processes are funda-
mentally different than those of classical physics. Therefore manipulation of
information based on quantum physical processes has also to be fundamentally
different than their classical counterparts. It was first realised by Feynman that
simulating quantum systems on a classical computer would be very inefficient
[1]. However, if one utilizes quantum systems than one can do much more. For
this reason, quantum information is distinguished from conventional classical
information.
1. QUANTUM COMPUTATION
The physics of information and computation are intimately related. Informa-
tion is encoded in the state of a physical system, whereas computation involves
the processing of this information through actions on the physical system. This
processing must obey physical law. Therefore, the study of information and
computation are linked through the study of underlying physical processes. If
the physical processes obey the rules of classical physics, the corresponding
computation is dubbed "classical." If on the other hand, the underlying pro-
cesses are subjected to quantum mechanical rules, the resulting computation
will be called "quantum computation." The logic that lies at the heart of conven-
tional computers and quantum computers is therefore fundamentally different.
Quantum computation is a particular way of processing information which uti-
lizes the principles of the linearity of quantum mechanics, out of which comes
quantum superposition, quantum entanglement and quantum parallelism. This
was first suggested by Deutsch [2].
In a conventional computer information is stored as binary digits (bits)
usually (logically) labeled 0 and 1. To represent a bit, one may use any
physical system that one likes provided it allows two distinct states. Two
bits of information can be stored in any system allowing 22 = 4 possible
distinct states. Similarly, n bits of information may be represented in any
system capable of providing 2n distinct states. In each case, there is only one
configuration at a time of the logical bits (e.g., 110 for 3 bits). Information
stored in these binary digits can be manipulated using elementary logic gates
that obey Boolean algebra. For example, in a conventional classical computer
one can manipulate information using sequence of logical operations such
as AND, OR, NOT, and XOR gates. These gates may be built into circuits
constructing any possible Boolean functions [3].
1.1 QUBITS
Let us represent a bit, 0 or 1, by saying that the spin of a neutron is up
or down, or we could sayan atom is in ground or in an excited state, or a
photon is horizontally or vertically polarized. All these systems are examples
Quantum computing with qubits 5
of two-state quantum systems because the two states are orthogonal and hence
logically distinct from each other. When a quantum system is in a given basis
state it may be said to carry classical information. However, quantum theory
also allows states which are in linear superpositions of these basis states. If
we use the logical label for each of the basis states then the most general pure
quantum state for a "two-state" system is given by
for some complex numbers a and,6 which satisfy lal 2 + 1,61 2 = 1. According
to Dirac a quantum state 1/1 is denoted by a "ket" 11/1} E 1-£ = ((J (a two-
dimensional Hilbert space). Such a system contains a quantum bit or "qubit"
of information [4].
A single qubit allows two inputs to be stored (and possibly processed) si-
multaneously. As we add more qubits the number of simultaneous possibilities
grows very quickly. For example, for two qubits, which have logically distinct
states labeled by 00,01,10 and 11 the most general state may be written as a
superposition of these four possibilities
For n qubits with 2n possible distinct logical basis states the most general pure
state will take the form of a simultaneous superposition of this exponential
number of possibilities, such as
(1.3)
1 N-l
on any given input x, with the input located in the first register and the output
stored in a second register. Then by the linearity of quantum mechanics, the
action of Uf on the equal superposition of the input register plus an extra output
register will produce
1 N-l 1 N-l
Uf : .;N ~ Ix)IO) -t .;N ~ Ix)lf(x)) . (1.6)
That is, all possible function evaluations have been done in a single step. This
massive parallelism comes for free if one could ever build a quantum computer.
Unfortunately, we are unable to extract all this information in a single ob-
servation of the resulting quantum state. Instead, we will only be able to
probabilistically extract the information encoded in a single "branch" or term
in this superposition. Naively then the quantum parallelism is all lost at the
Quantum computing with qubits 7
References
[1] R. Feynman, Found. of Physics 16, 507 (1986).
[2] D. Deutsch, Proc. R. Soc. London. A, 400, 97 (1985).
[3] G. Boole, An Investigation of The Laws of Thought reprinted as (Dover,
New York, 1958).
[4] B. Schumacher, Phys. Rev. A 51,2738 (1995).
[5] J. Bell, Speakables and Unspeakables in Quantum Theory (Oxford Uni-
versity Press).
[6] D. Deutsch and R. Jozsa, Proc. R. Society (London) A 439,553 (1992).
[7] P. Shor, in Proc 35th Annual Symp. on Found. of Compo Sci. (IEEE
Computer Society Press, 1994).
[8] L. K. Grover, Phys. Rev. Lett. 79, 325 (1997).
[9] M. A. Nielsen and 1. L. Chuang, Quantum Computation and Quantum
Information (Cambridge University Press, Cambridge, England, 2000).
[10] H. K. Lo and S. L. Braunstein (Eds), Scalable Quantum Computers
(Wiley-VCH Verlag, Berlin, 2001).
Chapter 2
QUANTUM COMPUTATION
OVER CONTINUOUS VARIABLES*
Seth Lloyd
MIT Department of Mechanical Engineering
MIT 3-160, Cambridge, Mass. 02139, USA
[email protected]
Samuel L. Braunstein
Informatics, Bangor University, Bangor LL57 I UT, United Kingdom
[email protected]
Abstract This paper provides necessary and sufficient conditions for constructing a uni-
versal quantum computer over continuous variables. As an example, it is shown
how a universal quantum computer for the amplitudes of the electromagnetic
field might be constructed using simple linear devices such as beam splitters and
phase shifters, together with squeezers and nonlinear devices such as Kerr-effect
fibers and atoms in optical cavities. Such a device could in principle perform
"quantum floating point" computations. Problems of noise, finite precision, and
error correction are discussed.
9
10 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
in the limit that 8t --+ 0, the result is the same as if one had applied the Hamil-
tonian irA, B] for time 8t 2 . In general, if one can apply a set of Hamiltonians
{±Hi}, one can construct any Hamiltonian that is a linear combination of
Hamiltonians of the form ±i[Hi, Hj], ±[Hi' [Hj, Hk]], etc. [9,10,11,12,13],
and no other Hamiltonians. That is, one can construct the Hamiltonians in the
algebra generated from the original set by commutation. This key point, orig-
inally derived in the context of quantum control and discrete quantum logic,
makes it relatively straightforward to determine the set of Hamiltonians that
can be constructed from simpler operations.
Now apply this result to the continuous variables introduced above. Since
[X, P] = i, the application of the translations ±X and ±P for short periods
of time clearly allows the construction of any Hamiltonian aX + bP + c that
is linear in X and P; this is all that it allows. To construct more complicated
Hamiltonians one must also be able to perform operations that are higher order
polynomials in X and P. Suppose now that one can apply the quadratic
Hamiltonian H = (X2 + p 2)/2. Since P = i[H, P] = X, X = i[H, X] =
- P, application of this Hamiltonian for time t takes X --+ cos tX - sin tP,
P --+ cos tP + sin tX. If X and P are quadrature amplitudes of a mode of
the electromagnetic field, then H is just the Hamiltonian of the mode (with
frequency w = 1) and corresponds to a phase shifter. Hamiltonians of this
form can be enacted by letting the system evolve on its own or by inserting
artificial phase delays. Note that since eiHt is periodic with period 1/47r, one
can effectively apply - H for a time 8t by applying H for a time 47r - 8t. The
simple commutation relations between H, X and P imply that the addition
of ±H to the set of operations that can be applied allows the construction of
Hamiltonians of the form aH + bX + cP + d.
12 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Suppose that in addition to translations and phase shifts one can a{lply
the quadratic Hamiltonian ±S = ±(XP + PX)/2. S has the effect X =
i[S, X] = X, P = i[S,P] = -P, i.e., applying +S takes X -t etX,
P -t e- t P: S "stretches" X and "squeezes" P by some amount. Similarly
-S squeezes X and stretches P. In the case of the electromagnetic field,
S corresponds to a squeezer operating in the linear regime. It can easily be
verified that [H, S] = i(X2 - p2). Looking at the algebra generated from
X, P, H and S by commutation, one sees that translations, phase shifts, and
squeezers allow the construction of any Hamiltonian that is quadratic in X and
P, and of no Hamiltonian of higher order.
To construct higher order Hamiltonians, nonlinear operations are required.
One such operation is the "Kerr" Hamiltonian H2 = (X2 + p2)2, corre-
sponding to a X3 process in nonlinear optics. This higher order Hamilto-
nian has the key feature that whereas commuting the previous Hamiltoni-
ans, X, P, H, S with some polynomial in X and P resulted in a polyno-
mial with the same or lower order, commuting H2 with a polynomial in X
and P typically increases its order. By evaluating a few commutators, e.g.,
[H2, X] = i(X2p+PX 2 +2p3)/2, [H 2,P] = -i(P2X +XP2 +2X 3)/2,
[X, [H2, S]] = p3, [P, [H2, S]] = X 3 one sees that the algebra generated
by X, P, H, Sand H2 by commutation includes all third order polynomials
in X and P. A simple inductive proof now shows that one can construct
Hamiltonians that are arbitrary Hermitian polynomials in any order of X and
P. Suppose that one can construct any polynomial of order M or less, where
M is of degree at least 3. Then since [P 3, pm xn] = ip m+2X n- 1 + lower
order terms, and [X 3, pm xn] = ipm-l xn+2+ lower order terms, one can
by judicious commutation of X 3 and p 3 with monomials of order M construct
any monomial of order M + 1. Since any polynomial of order M + 1 can
be constructed from monomials of order M + 1 and lower, by applying linear
operations and a single nonlinear operation a finite number of times one can
construct polynomials of arbitrary order in X and P to any desired degree of
accuracy. Comparison with similar results for the discrete case [14] shows that
the number of operations required grows as a small polynomial in the order of
the polynomial to be created, the accuracy to which that polynomial is to be
enacted, and the time over which it is to be applied.
The use of the Kerr Hamiltonian H2 was not essential: any higher order
Hamiltonian will do the trick. Note that commutation of a polynomial in X
and P with X and P themselves (which have order 1) always reduces the order
of the polynomial by at least 1, commutation with H and S (which have order
2) never increases the order, and commutation with a polynomial of order 3
or higher typically increases the order by at least 1. Judicious commutation
of X, P, Hand S with an applied Hamiltonian of order 3 or higher therefore
Quantum computation over continuous variables 13
might provide the long-lived, lossless states required for quantum computation
over continuous variables.
Noise poses a difficult problem for quantum computation [18, 19,20], and
continuous variables are more susceptible to noise than discrete variables.
Since an uncountably infinite number of things can go wrong with a contin-
uous variable, it might at first seem that continuous error correction routines
would require infinite redundancy. In fact, continuous quantum error correc-
tion routines exist and require no greater redundancy than conventional routines
[3, 4, 5]. Such routines are capable of correcting for noise and decoherence
in principle: In practice, measurement noise, losses, and the lack of perfect
squeezing will lead to imperfect error correction [5]. Surprisingly, continuous
quantum error correction routines are in some sense easier to enact than discrete
quantum error correction routines, in that the continuous routines can be imple-
mented using only linear operations together with classical feedback [5]. The
relative simplicity of such routines suggests that robust, fault-tolerant quantum
computation may in principle be possible for continuous quantum variables
as well as for qubits. (A scheme for quantum computation is fault-tolerant
if quantum computations can be carried out even in the presence of noise and
errors [21, 22]. A fault-tolerant scheme that allows for arbitrarily long quantum
computations to be carried out is said to be robust [23].) If this is indeed the
case then quantum computation over continuous variables, despite its intrinsic
difficulties, may be an experimentally viable form of quantum information pro-
cessing. Continuous variables might be used to simulate continuous quantum
systems such as quantum field theories. Even in the absence of fault tolerance,
the large bandwidths available to continuous quantum computation make it
potentially useful for quantum communications and cryptography [24].
S.L. would like to thank H. Haus and H. J. Kimble for useful discussions. This
work was supported by DARPA under the QUIC initiative.
References
[1] D. DiVincenzo, Science 270,255 (1995).
[2] S. Lloyd, Sci. Am. 273, 140 (1995).
[3] S. Lloyd and J. J.-E. Slotine, Phys. Rev. Lett. 80,4088 (1998).
[4] S. L. Braunstein, Phys. Rev. Lett. 80, 4084 (1998).
[5] S. L. Braunstein, Nature (London) 394, 47 (1998).
[6] S. L. Braunstein and H. J. Kimble, Phys. Rev. Lett. 80, 869 (1998).
[7] A. Furusawa, et at, Science 282, 706 (1998).
[8] This definition of quantum computation corresponds to the normal "cir-
cuit" definition of quantum computation as in, e.g., D. Deutsch, Proc.
16 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Samuel L. Braunstein
Informatics, Bangor University, Bangor LL57 1 UT, United Kingdom
[email protected]
Abstract We propose an error correction coding algorithm for continuous quantum vari-
ables. We use this algorithm to construct a highly efficient 5-wave-packet code
which can correct arbitrary single wave-packet errors. We show that this class of
continuous variable codes is robust against imprecision in the error syndromes.
A potential implementation of the scheme is presented.
Quantum computers hold the promise for efficiently factoring large integers
[1]. However, to do this beyond a most modest scale they will require quantum
error correction [2]. The theory of quantum error correction is already well
studied in two-level or spin-! systems (in terms of qubits or quantum bits)
[2, 3, 4, 5, 6, 7]. Some of these results have been generalized to higher-spin
systems [8, 9, 10, 11]. This work applies to discrete systems like the hyperfine
levels in ions but is not suitable for systems with continuous spectra, such as
unbound wave packets. Simultaneously with this paper, Lloyd and Slotine
present the first treatment of a quantum error correction code for continuous
quantum variables [12], demonstrating a 9-wave-packet code in analogy with
Shor's 9-qubit coding scheme [2]. Such codes hold exciting prospects for
the complete manipulation of quantum systems, including both discrete and
continuous degrees of freedom, in the presence of inevitable noise [13].
In this Letter we consider a highly efficient and compact error correction cod-
ing algorithm for continuous quantum variables. As an example, we construct
a 5-wave-packet code which can correct arbitrary single-wave-packet errors.
We show that such continuous variable codes are robust against imprecision in
19
20 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
the error syndromes and discuss potential implementation of the scheme. This
paper is restricted to one-dimensional wave-packets which might represent the
}\lave function of a nonrelativistic one-dimensional particle or the state of a
single polarization of a transverse mode of electromagnetic radiation. We shall
henceforth refer to such descriptions by the generic term wave packets [14].
Rather than starting from scratch we shall use some of the theory that has
already been given for error correction on qubits. In particular, Steane has
noted that the Hadamard transform,
(3.1)
maps phase flips into bit flips and can therefore be used to form a class of
quantum error correction codes that consist of a pair of classical codes, one for
each type of "flip" [3]. This mapping between phase and amplitude bases is
achieved with a rotation about the y-axis by 7f /2 radians in the Bloch sphere
representation of the state. In analogy, the position and momentum bases of a
continuous quantum state may be transformed into each other by 7f /2 rotations
in phase-space. This transition is implemented by substituting the Hadamard
rotation in the Bloch sphere by a Fourier transform between position and
momentum in phase-space. This suggests that we could develop the analogous
quantum error correction codes for continuous systems [15].
We shall find it convenient to use a units-free notation where
(3.3)
To avoid confusion we shall work in the position basis throughout and so define
the Fourier transform as an active operation on a state by
(3.4)
where both x and yare variables in the position basis. Note that Eqs. (3.3)
and (3.4) correspond to a change of representation and a physical change of the
state, respectively.
Error correction/or continuous quantum variables 21
By removing the cyclic structure of the XOR gate we have produced a gate
which is no longer its own inverse. Thus, in addition to the Fourier transform
and this generalized XOR gate we include their inverses to our list of useful
gates. This generalized XOR operation performs translations over the entire
real line, which are related to the infinite additive group on JR.. The characters
X of this group satisfy the multiplicative property X(x + y) = X(x)x(y) for all
x, y E JR. and obey the sum rule
11
-
7f
00
-00
dx X(x) = J(x) , (3.6)
where X( x) = e2ix . Interestingly, this sum rule has the same form as that found
by Chau in higher-spin codes [10). Once we have recognized the parallel, it
is sufficient to take the code of a spin-~ system as a basis for our continuous-
variable code.
Based on these parallel group properties, we are tempted to speculate a
much more general and fundamental relation: We conjecture that n-qubit error
correction codes can be paralleled with n-wave-packet codes by replacing
the discrete-variable operations (Hadamard transform and XOR gate) by their
continuous-variable analogs (Fourier transform, generalized-XOR and their
inverses). As a last remark before embarking on the necessary substitutions (in
a specific example), we point out that the substituti?n conjecture is only valid
for qubit codes whose circuits involve only these (H and XOR) elements. We
shall therefore restrict our attention to this class of codes.
An example of a suitable 5-qubit code was given by Laflamme et al. [16). We
show an equivalent circuit in Fig. 3.1 [17]. As we perform the substitutions,
we must determine which qubit-XOR gates to replace with the generalized-
XOR and which with its inverse. To resolve this ambiguity, two conditions
are imposed. First, we demand that the code retain its properties under the
parity operation (on each wave packet). We conclude that either gate may be
chosen for the first operation on initially zero-position eigenstates. Ambiguity
remains for the last four XOR substitutions. As a second step, the necessary
and sufficient condition for quantum error correction [5,6],
22 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
I~)
10)
10)
10)
10)
Figure 3.1 Quantum error correction circuit from [17]. The qubit I'I/J) is rotated into a 5-particle
subspace by the unitary operations represented by the operations shown in this circuit. Note that
the 3-qubit gates are simply pairs of XORs.
In the case of a single wave-packet error, for our 5-wave-packet code, it turns
out that amongst the conditions ofEq. (3.7) only (x~n odeltlo. t5{3IXencode)'
having errors on wave packets 4 and 5, is affected by t~e ambiguity (see detail
below). An explicit calculation of all the conditions shows that the circuit of
Fig. 3.2 yields a satisfactory quantum error correction code (as do variations of
this circuit due to the extra freedom with respect to the choice of operator acting
on wave-packets 1-3). By analogy with the results for higher-spin codes, we
know that this code is optimal (though not perfect) and that no four-wave-packet
code would suffice [10]. The code thus constructed has the form
Ix encod)
e
= _1_
7[3/2
J dw dy dz e2i (wy+xz)
xlz,y+X,w+X,w-z,y-z). (3.8)
Error correction/or continuous quantum variables 23
----------------~F ~.--
10)
10)
10)
10)
Figure 3.2 This "circuit" unitarily maps a one-dimensional single-wave-packet state 1'1/1) into
a 5-wave-packet error correction code. Here the auxiliary wave packets 10) are initially zero-
position eigenstates. For degrees-of-freedom larger than qubits the ideal XOR is not its own
inverse; here the daggers on the XOR gates represent the inverse operation.
= Z wy +xz-w '"
- 1 / dw' dy' dz' dw dy dz e2"( y -x z ')
11"3
For the other cases we find by explicit calculation, for wave-packets j =I- k, that
I At A _ I
(xencodel£ja £k,Blxencode) - <5(x - x) )..a,B • (3.11)
(3.12)
x Iw, w, w, y, y, y, z, z, z) , (3.13)
where parity alone removes all ambiguity. (This code has been independently
obtained by Lloyd and Slotine [12].) Since this 9-wave-packet code corrects
position errors and momentum errors separately, it is sufficient to study the
subcode
designed to correct position errors on a single wave packet. The most general
position error (on a single wave packet) is given by some function of the
momentum of that system E(p) and need not be unitary on the code subspace
[Eq. (3.7)]. The action of such an error on a wave packet may be written in the
position basis as
where £(x) is the Fourier transform of £(p). Thus the most general position
error looks like a convolution of the wave packet's ket with some unknown
(though not completely arbitrary) function. Suppose this error occurs on wave
packet 1 in the repetition code (3.14). Further, let us use auxiliary wave-packets
(so-called ancillae) and compute the syndrome as shown in Fig. 3.3, then the
resulting state may be written as
! dy£(y)lx-y,x,x,-y,O,y). (3.16)
Error correction/or continuous quantum variables 25
t
10)
t
10) '- '-1./
t
10) I'~ I'
'-
Figure 3.3 Syndrome calculation and measurement: A state with a sing1e-wave-packet position
error (here on wave packet 1) enters and the differences of each pair of positions is computed.
The syndrome {81, 82,83} may now be directly measured in the position basis.
Everything up till now has been unitary and assumed ideal. Now measure
the syndrome: Ideally it would be { -y, 0, y} collapsing the wave packet for
a specific y. Correcting the error is now easy, because we know the location,
value and sign of the error. Shifting the first wave packet by the amount y
retrieves the correctly encoded state lx, x, x) Note that this procedure uses only
very simple wave-packet-gates: The comparison stage is done classically, in
contrast to the scheme of Lloyd and Slotine, where the comparison is performed
at the amplitude level and involves significantly more complicated interactions
[12].
It is now easy to see what imprecise measurements of the syndromes will do.
Suppose each measured value of a syndrome sj is distributed randomly about
the true value sj according to the distribution Pmeas (sj - sj). We find two
conditions for error-correction to proceed smoothly. First, Pmeas (x) must be
narrow compared to any important length scales in t (x). This guarantees that
the chance for "correcting" the wrong wave packet is negligible and reduces the
position-error operator to an uninteresting prefactor. If the original unencoded
J
state had been dx ~ (x) Ix) then after error correction we would obtain the
mixed state
Thus, unless Pmeas (x) is also narrow compared to any important length scales
in ~ (x), decoherence will appear in the off-diagonal terms for wave packet 1
of the corrected state (3.17). This second condition is also seen in the quantum
teleportation of continuous variables due to inaccuracies caused by measure-
26 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
ment [13]. These conditions roughly match those described by Lloyd and
Slotine [12]. We note that any syndrome imprecision will degrade the encoded
states, though this precision may be improved by repeated measurements of the
syndromes. For our 5-wave-packet example (3.8), syndromes consist of sums
of two or more wave-packet positions or momenta and are measured similarly.
It should be noted that Chau's higher-spin code [10] could have been imme-
diately taken over into a quantum error correction code for continuous quantum
variables in accordance with our substitution procedure. However, we have pro-
duced an equivalent code with a more efficient circuit prescription: Whereas
Chau gives a procedure for constructing his higher-spin code using 9 general-
ized XOR operations, the circuit in Fig. 3.2 requires only 7 such gates or their
inverses. In fact, we could run this substitution backwards to obtain a cleaner
5-particle higher-spin code based on Eq. (3.8).
In order to consider potential implementations of the above code let us restrict
our attention to a situation where the wave packets are sitting in background
harmonic-oscillator potentials. By the virial theorem the form of a wave packet
in such a potential is preserved up to a trivial rotation in phase-space with
time. The two operations required may be implemented simply as follows:
The rotation in phase-space, Eq. (3.4), may be obtained by delaying the phase
of one wave packet relative to the others, and the XOR operation, Eq. (3.5),
should be implemented via a quantum non-demolition (QND) coupling. There
exists extensive experimental literature on these operations both for optical
fields and for trapped ions [13, 18, 19, 20, 21].
The conjecture put forth in this Letter leads to a simple, 2-step design of
error correction codes for continuous quantum variables. According to this
conjecture, any qubit code, whose circuit operations include only a specific
Hadamard transformation, its inverse and the ideal XOR, may be translated
to a continuous quantum-variable code, by substituting these operators with
their continuous analogs and then imposing two criteria - parity invariance
and the error-correction condition - which remove any ambiguities in the
choice of operators. We demonstrate the success of this coding procedure
in two examples (one based on Shor's 9-qubit code [2], and a second based
on a variation of the Laflamme et al. 5-qubit code [16, 17]). The 5-wave-
pa.:;ket code presented here is the optimal continuous encoding of a single
one-dimensional wave packet that protects against arbitrary single-wave-packet
errors. We show that this code (and in fact the entire class of codes derived
in this manner) are robust against imprecision in the error syndromes. The
potential implementation of the proposed class of circuits in optical-field and
ion-trap set-ups is an additional incentive for further investigation of the robust
manipulation of continuous quantum variables.
Error correctionjor continuous quantum variables 27
This work was funded in part by EPSRC grant GR/L91344. The author appreci-
ated discussions with N. Cohen, H. J. Kimble, D. Gottesman and S. Schneider.
Appendix: Addendum
The above work demonstrates that an error correction code may be designed,
though for implementation it requires some kind of QND couplings. Soon after
this work was complete the author realized that much of the same work could
be done with straight linear optics, in particular coupling based upon beam
splitter operations [22].
First, consider two light fields Ix) and Iy) incident on an "ideal" (phase-free)
beam splitter. The output light fields are given by
where the subscripts 12 refer to the wave packets acted upon. From these ideal
beam splitters we construct a 3-port device called a tritter [23]
~ ~ ~ -1 1
7123 = 8 23 (11"/4) 8 12 {cos y'3), (3.A.2)
from which we may construct the three wave-packet subcode of Shor's 9-wave-
packet code via
(3.A.3)
In fact, the only difference between this code and the "ideal" version of it
Ix, x, x) is a simple common scaling in each of the code's wave packets.
Finally, we combine these elements together with this freedom to scale, in
order to produce our encoding device out of linear optics. A suitable choice for
the 9-wave-packet code is a 9-port beam-splitter, which we call a nona-splitter
(3.A.4)
References
[1] P. W. Shor in Proc. 35th Annual Symposium on the Foundations of Com-
puter Science, edited by S. Goldwasser (IEEE Computer Society Press,
Los Alamitos, California, 1994), p.124.
[2] P. W. Shor, Phys. Rev. A 52, R2493 (1995).
[3] A. M. Steane, Proc. Roy. Soc. London 452, 2551 (1996).
[4] A. R. Calderbank and P. W. Shor, Phys. Rev. A 54, 1098 (1996).
[5] c. H. Bennett, D. P. DiVincenzo, J. A. Smolin and W. K. Wootters, Phys.
Rev. A 54,3824 (1996).
[6] E. Knill and R. Laflamme, Phys. Rev. A, 55,900 (1997).
[7] A. R. Calderbank, E. M. Rains, P. W. Shor and N. J. A. Sloane, Phys. Rev.
Lett. 78, 405 (1997).
[8] E. Knill, LANL report LAUR-96-2717, preprint quant-phl9608048.
[9] H. F. Chau, Phys. Rev. A 55, R839 (1997).
[10] H. F. Chau, Phys. Rev. A 56, Rl (1997).
[11] E. M. Rains, LANL preprint quant-phl9703048.
[12] S. Lloyd and J.-J. E. Slotine, LANL preprint quant-phl9711021.
[13] S. L. Braunstein and H. J. Kimble, "Teleportation of continuous quantum
variables," Phys. Rev. Lett., submitted; S. L. Braunstein, H. J. Kimble, Y.
Sorensen, A. Furusawa and N. Ph. Georiades, "Teleportation of continu-
ous quantum variables," IQEC 1998, abstract submitted.
[14] Although we consider multi-wave-packet states as one-dimensional we
would not want them to physically overlap so they could, for example, be
displaced one from another in an orthogonal direction.
[15] An example using as few as 7 qubits to correct an arbitrary single-qubit
error in a l-qubit encoded state can be found in [3].
[16] R. Laflamme, C. Miquel, J. P. Paz and W. H. Zurek, Phys. Rev. Lett. 77,
198 (1996).
[17] S. L. Braunstein and J. A. Smolin, Phys. Rev. A 55,945 (1997).
[18] S. F. Pereira, Z. Y. Ou and H. J. Kimble, Phys. Rev. Lett. 72, 214 (1994).
[19] K. Bencheikh, j. A. Levenson, P. Grangier and O. Lopez, Phys. Rev. Lett.
75,3422 (1995).
[20] R. L. de Matos Filho and W. Vogel, Phys. Rev. Lett. 76, 4520 (1996).
[21] R. Bruckmeier, H. Hansen and S. Schiller, Phys. Rev. Lett. 79, 1463
(1997).
[22] S. L. Braunstein, Nature (London) 394, 47 (1998).
Error correctionjor continuous quantum variables 29
DEUTSCH-JOZSA ALGORITHM
FOR CONTINUOUS VARIABLES
Arun K. Pati
Institute of Physics, Bhubaneswar-751005, Orissa, INDIA
Theoretical Physics Division, BARe, Mumbai, INDIA
Samuel L. Braunstein
Informatics, Bangor University, Bangor LL57 1 Ur, United Kingdom
[email protected]
°
box" or "oracle query" that computes a binary function f (i) : En -t E.
Further, the function f which only takes values or 1 is promised to be either
constant or balanced (with an equal number of each type of outcome over all
input strings). The aim is to determine this property for f, i.e., whether it is
constant or balanced. On a classical computer in the worst case the oracle query
requires O(2n) function evaluations. However, if one calculates the function
using reversible quantum operations then only a single function evaluation is
required to achieve the goal [6, 21].
In the continuous variable setting we pose the problem in the following
way. Suppose there is a particle located somewhere along the x-axis. Since
x E lR. is a continuous variable it can take value from -00 to +00 (in practice
it may be from - L to L, where L is some length scale involved, but still the
number of possible values of x is infinite). Suppose there are two persons
Alice and Bob playing a game [22]. Alice tells Bob a value of x and Bob
calculates some function f(x) which takes values 0 or 1. Further, Bob has
promised Alice that he will use a function which is either constant or balanced.
A constant function is 0 or 1 for all values of x E (-L, +L). For a balanced
function, f (x) = 0 or 1 for exactly half of the cases. One can define the
balanced function more precisely in the following manner. Imagine that the
interval for the continuous variable x has been divided into n sub-intervals.
Let p, be the Lebesgue measure on R A function f(x) is balanced provided
the Lebesgue measure of the support for where the function is zero is identical
to the Lebesgue measure of the support for where the function is one, i.e.,
Deutsch-Iozsa algorithm for continuous variables 33
where both x and y are in the position basis. This has been used in developing
error correction codes [16, 18] and Grover's algorithm for continuous variables
[20]. This Fourier transformation can be easily applied in physical situations.
For example, when Ix) represents quadrature eigenstate of a mode of the
electromagnetic field, Fix) is simply an eigenstate of the conjugate quadrature
produced by a 'If' /2 phase delay.
Another useful gate on a continuous variable quantum computer is XOR
gate (analogous to the controlled NOT gate for qubits but without the cyclic
condition) defined as [18]
Further, we assume that given a classical circuit for computing f (x) there is a
quantum circuit which can compute a unitary transformation Uj on a continuous
variable quantum computer. If a quantum circuit exists that transforms
then by linearity it can also act on any superposition of qunat states. For
example, if we evaluate the function on a state (4.1) along with another qunat
state Jz), we have
This shows that using quantum parallelism for idealized qunat computers one
can evaluate all possible values of a function simultaneously with one applica-
tion ofUj.
34 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(i) Alice stores her query in a qunat register prepared in an ideal position
eigenstate Ixo) and attaches another qunat in a position eigenstate 17r/2). So
the two quants are in the state Ixo)I7r/2)
(ii) She creates superpositions of qunat states by applying the Fourier trans-
formation to the query qunat and the target qunat. The reSUlting state is given
by
(iii) Bob evaluates the function using the unitary operator U!. The state
transforms as
Here, the key role is played by the ancilla qunat state 17r /2). To see how the
function evaluation takes place consider the intermediate steps given by
If the function f (x) = 0 there is no sign change and if f (x) = 1 there is a sign
change. After the third step performed by Alice, she has a quant state in which
the result of Bob's function evaluation is encoded in the amplitude of the qunat
superposition state given in (4.6). To know the nature of the function she now
performs an inverse Fourier transformation on her qunat state.
(iv) The qunat states after the inverse Fourier tranform is given by
(v) Alice measures her qunat by projecting onto the original position eigen-
state Ixo). In an ideal continuous variable scheme the correct projection oper-
ator is defined as [23]
xo-l:!.xo/2
dyly)(yl· (4.9)
As has been explained in [20, 23] if the observable has a continuous spectrum
then the measurement cannot be performed precisely but must involve some
spread L\xo. Therefore, the action of projection onto the qunat state after step
(iv) is given by
(4.10)
Now consider two possibilities. If the function is constant then the above
equation r~duces to ±lxo)FI.1T /2). [In simplifying we ~eed to use th~ Di~ac
delta function (1/71') f dx e2~x(xO-Y) = 8(xo - y).] ThiS means that if Allce
measures Ixo) she is sure that f(x) is definitely constant. In the other case, i.e.,
when the function is balanced she will not get the measurement outcome to be
Ixo). In fact, in the balanced case the outcome is orthogonal to the constant
case as the result gives zero. Therefore, a single function evaluation (follwed
by a measurement onto Ixo)) in a qunat quantum computer can decide whether
the promised function is constant or balanced. Unlike the qubit case, in the
idealized continuous variable case the reduction in the number of query calls
is from infinity to one.
In conclusion, we have generalised the primitive quantum algorithm (Deutsch-
Jozsa algorithm) from the discrete case to the idealized continuous case. It may
be worth mentioning that as in error correction codes for continuous-variablees
[16], if one replaces the Hadamard transform and XOR gate by their continuous-
variable analogs in original Deutsch-Jozsa algorithm for qubit case, then the
idealized algorithm works perfectly. This theoretically demonstrates the power
of quantum computers to exploit the superposition principle giving an infinite
speed up compared to classical scenario. This idealized analysis has not con-
sidered the affects of finite precision in measurement or state construction and
so whether it may be implemented remains an open question for further study.
Part of the difficulty in extending this work in this direction is that defining an
oracle for continuous variables appears to be a difficult task, one that we have
carefully avoided here. An alternate way forward might be to consider some
sort of "hybrid" approach involving both qunats and qubits. This is precisely
what Seth Lloyd considers in the following chapter.
36 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
AKP thanks P. van Loock and R. Simon for useful feedback. AKP also thanks
G. Giedke for discussions during Benasque Science Center-2000 in Spain on
defining balanced function for continuous variables.
References
[1] S. Lloyd, Science, 261, 1569 (1993).
[2] S. Lloyd, Science, 273, 1073 (1996).
[3] P. Benioff, Phys. Rev. Lett. 48, 1581 (1982).
[4] R. Feynman, Int. J. Theor. Phys. 21, 467 (1982).
[5] D. Deutsch, Proc. R. Soc. London A 400, 97 (1985).
[6] D. Deutsch and R. Jozsa, Proc. R. Soc. London, A 439,553 (1992).
[7] E. Bernstein and U. Vazirani, in Proc. of the 25th Annual Symposium on
the Theory of Computing (ACM Press, New York, 1993), p. 11-20.
[8] D. R. Simon, in Proc. of the 35th Annual Symposium on Foundations
of Computer Science, edited by S. Goldwasser (IEEE Computer Society,
Los Alamitos, CA, 1994), p. 116-123.
[9] D. DiVincenzo, Science, 270, 255 (1995).
[10] P. W. Shor, in Proceedings of the 37th Annual Symposium on Foundations
of Computer Science (IEEE Computer Society Press, Los Alamitos, CA,
1996), pp. 56-65.
[11] A. Ekert and R. Jozsa, Rev. Mod. Phys. 68, 733 (1996).
[12] L. K. Grover, Phys. Rev. Lett. 79, 325 (1997).
[13] L. K. Grover, Phys. Rev. Lett. 80, 4329 (1998).
[14] A. K. Pati, "Grover's algorithm, time-dependent search and unitary per-
turbation" (preprint), (1999).
[15] S. L. Braunstein and H. J. Kimble, Phys. Rev. Lett. 80, 869 (1998).
[16] S. L. Braunstein, Phys. Rev. Lett. 80,4084 (1998).
[17] S. Lloyd and J.-J. E. Slotine, Phys. Rev. Lett. 80, 4088 (1998).
[18] S. L. Braunstein, Nature (London) 394, 47 (1998).
[19] S. Lloyd and S. L. Braunstein, Phys. Rev. Lett. 82,1784 (1999).
[20] A. K. Pati, S. L. Braunstein and S. Lloyd, quant-ph/0002082.
[21] R. Cleve, A. Ekert, C. Macciavello and M. Mosca, Proc. R. Soc. London
A 454, 339 (1998).
[22] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum
Information, (Cambridge University Press, 2000).
[23] C. Cohen-Tannoudji, B. Diu and F. Laloe, Quantum mechanics (John
Wiley & Sons, New York, 1977).
Chapter 5
Seth Lloyd
d'Arbeloff Laboratory for Information Systems and Technology
Department of Mechanical Engineering
Massachusetts Institute of Technology
Cambridge, Massachusetts 02139
[email protected]
Abstract Necessary and sufficient conditions are given for the construction of a hybrid
quantum computer that operates on both continuous and discrete quantum vari-
ables. Such hybrid computers are shown to be more efficient than conventional
quantum computers for performing a variety of quantum algorithms, such as
computing eigenvectors and eigenvalues.
nuisance: either they figure as sources of noise and decoherence, as in the case
of environmental baths of harmonic oscillators, or they must be restricted to a
discrete set of states by cooling, as in the case of the oscillatory modes of ions
in ion-trap quantum computers. In hybrid quantum computation, by contrast,
the full range of continuous quantum variables can be put to use.
The basic model for performing quantum computation using a hybrid of
continuous and discrete variables follows the normal model for performing
quantum computation using discrete or continuous variables on their own [7,
15]. Assume that one has the ability to 'tum on' and 'tum off' the members of a
set Hamiltonian operators {±Hj}, corresponding to the ability to apply unitary
transformations of the form e±~Hjt. The set of transformations that can be
constructed in this fashion is the set of transformations of the form e- iHt where
His a member of the algebra generated from theHj via commutation: i.e., since
eiH2teiHlte-iH2te-iHlt = e-[Hl,H2]t 2 + O(t 3 ), the ability to tum on and tum
off ±H1 and ±H2 allows one effectively to tum on and off H = ±i[H1 , H 2],
etc. Transformations of the form e- iHt for non-inifinitesimal t can then be
built up from infinitesimal transformations to any desired degree of accuracy.
For the sake of ease of exposition, concentrate here on discrete variables
(qubits) that are spins, characterized by the usual Pauli operators (Yx, (Y y, (Y z and
to continuous variables (qunats) that are harmonic oscillators characterized by
the usual annihilation and creation operators a, at ([a, at] = 1), and by the
'position' and 'momentum' operators X = (a + at )j2, P = (a - a t )j2i,
([X, P] = i). It is convenient to think of the harmonic oscillators as modes
of the electromagnetic field with X and P proportional to the quadrature
amplitudes of the mode. The generalization to discrete variables with more
than two states and to other forms of continuous variable is straightforward and
will be discussed below.
To perform quantum computations one must be able to prepare one's vari-
ables in a desired state, perform quantum logic operations, and read out the
results. Assume that it is possible to prepare the discrete variables in the state
10) == I t)z, and the continuous variables in the vacuum state 10): alO) = O.
Assume that it is possible to measure (Y z for the discrete variables and X for
the continuous variables.
N ow look at performing transformations of the variables. Begin with just a
pair - one spin and one oscillator. Suppose that one can tum on and tum off
the Hamiltonians
(5.1)
As will now be seen, the ability to tum on and off Hamiltonians from this set
allows one to enact Hamiltonians that are arbitrary polynomials of the (Y's, X
and P. Note that these Hamiltonians all represent interactions between qubits
and oscillators: this is physically realistic in the sense that transformations on
Hybrid quantum computing 39
physical spins or atoms are accomplished by making the spins interact with the
electromagnetic field, and vice versa. In physically realizable situations, such
as the ion traps and optical cavities discussed below, the interactions in 5.1 are
turned on and off by applying laser or microwave pulses to couple discrete to
continuous degrees of freedom.
Now investigate what can be accomplished by turning on and off these inter-
actions. If the spin is prepared in the state IO}, then turning on the Hamiltonian
a zP is equivalent to turning on the Hamiltonian P for the oscillator on its own.
The Hamiltonian X can be turned on in a similar fashion. In order to apply this
Hamiltonian for a finite amount of time, the spin must be constantly reprepared
in the state IO} or new spins in this state must be supplied. This operation allows
the construction of coherent states of the oscillator.
Now start constructing effective Hamiltonians by the method of commutation
above. Since i[P, axX] = ax, we can effectively tum on the Hamiltonian ax.
Similarly for the Hamiltonian ±i[P, azX] = ±az . And since i[az , ax] = 2a y ,
any single qubit transformation e-iO"t E SU(2) can be enacted by turning on
and off Hamiltonians in the set. Since i[azP, azX] = 2, an arbitrary overall
phase can also be turned on and off. That is, we can enact arbitrary single qubit
transformations.
Now systematically build up higher order transformations. Since i[azX,
axX] = 2ayX2, and i[ayX2, axX] = 2az X 3 , etc., we can effectively tum on
and off Hamiltonians of the form aX n , for arbitrary a, n. Similarly, we can
tum on and off Hamiltonians of the form a pn. By preparing the spin in the state
IO} and turning on and off the Hamiltonians azxm, azpn, we can enact sin-
gle oscillator transformations corresponding to Hamiltonians that are arbitrary
Hermitian polynomials in X and P. (Not all such Hamiltonians are bounded.
Nonetheless, one can build up infinitesimal versions of such Hamiltonians and
apply them for finite time to states for which they are bounded.)
So the simple set of Hamiltonians above allows the construction of arbitrary
single qubit transformations and arbitrary polynomial transformations of the
continuous variable, along with arbitrary interactions between the spin and the
oscillator. Let us now look at more than one spin and one oscillator.
Since i[a~P, a;X] = a~a;, we can tum on the interaction Hamiltonian
a~a; between two spins 1 and 2 by making them both interact with the same
oscillator. But the ability to tum on this Hamiltonian together with the ability to
tum on arbitrary single-spin Hamiltonian translates into the ability to perform
arbitrary transformations on sets of spins: that is, one can perform arbitrary
quantum logic operations on the qubits alone.
Similarly, since i[ayXl' a x X 2] = 2azX1X2, the ability to make two os-
cillators interact with the same spin, initially in the state IO}, allows one to
tum on the Hamiltonian X 1 X2 between the two oscillators 1 and 2. But this
ability, together with the ability to tum on single oscillator Hamiltonians that
40 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
they include in the computation states and degrees of freedom that would
normally be sources of noise, decoherence, and loss.
Now tum to applications of hybrid quantum computers. Where does the
ability to perform manipulations of continuous variables as well as qubits give
an advantage? The first point to note in constructing hybrid algorithms is that
we must be careful to assume physically reasonable uses of hybrid variables-
i.e., uses that do not require infinite or exponentially high precision. Even in
the classical case, the use of continuous variables can give remarkable compu-
tational speed ups (the ability to solve NP-complete problems in polynomial
time, the ability to find the the answer to uncomputable problems in finite
time, etc.) if one allows arbitrary precision in manipulating and measuring
continuous variables. By giving an explicit construction of the operations that
can be used to perform continuous variable and hybrid quantum computation,
however, we have implicitly avoided the use of infinite or excessive precision:
all such operations would require infinite or excessive computational resources
to construct, manipulate, and measure the desired over-precise states.
With this caveat in mind, tum to the operations that are relatively easy to
perform using continuous quantum variables. A particularly useful subroutine
in a variety of quantum algorithms is the quantum Fourier transform: Ix) -+
z=~= 1 e ixy Iy). In the case of discrete quantum variables the quantum Fourier
transform on N qubits takes on the order of N quantum logic operations to
perform. Although this is an efficient algorithm it is nonetheless difficult at
present to perform quantum Fourier transforms on more than a few qubits (the
current record is three) [17]. By contrast, in the case of the continuous quantum
variables X and P, the quantum Fourier transform is trivial. If the eigenstates
of X with eigenvalue x are written Ix), then theeigenstates of P with eigenvalue
p can be written Ip) = (1/V2n) f~oo eiPXlx)dx. That is, the eigenstates of
P are the quantum Fourier transform of the eigenstates of X. Coupling to P
instead of X then allows immediate access to the Fourier transformed variable.
The quantum Fourier transform on a continuous variable is accomplished by a
zero-step operation. The ease of performing the quantum Fourier transform on
continuous variables suggests that in devising algorithms for hybrid quantum
computers we look for problems in which the quantum Fourier transform plays
a central role.
Perhaps the best known quantum algorithm in which the quantum Fourier
transform plays a central role is Shor's algorithm for factoring large numbers
[4]. Setting aside the difficulty of performing the other operations in this
algorithm (such as modular exponentiation), it is immediately clear that using
a continuous variable as the register on which to perform the quantum Fourier
transform in Shor's algorithm would require an exponentially high precision in
the preparation and manipulation of the continuous variable. (Hybrid quantum
42 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
1'I/J)lx = 0)
-+ e-iHPtl'I/J)lx = 0)
= e- iEj EjIEj){EjIPt L: j 'l/JjIEj) IX = 0)
L:j e-iEjtP'l/JjIEj)lx = 0)
= L:j 'l/JiIEj) Ix = Ejt) , (5.2)
since e -iPt Ix) = Ix + t). Clearly, at this point, a measurement of the variable
X on the continuous variable will yield the result x = tEi with probability
'l/Ji, leaving the system in the state lEi = x/t). That is, one can sample
the spectral decomposition of I'I/J), obtaining the eigenvalues Ei together with
their corresponding weights l'l/Jil 2 and eigenvectors lEi). The process is highly
efficient, requiring only the ability to prepare the initial squeezed state Ix = 0)
and to apply the Hamiltonian H P.
The hybrid eigenvalue and eigenvector finding algorithm using a continuous
variable to register the eigenvalue is more efficient than the corresponding
algorithm using qubits to register the eigenvalue. Since the quantum Fourier
transform is performed implicitly in the continuous register, fewer steps are
required in the hybrid algorithm. In addition, unlike the conventional version
of the algorithm, the hybrid version is insensitive to approximate decoherence
of the register in the course of the computation: measuring the value x of the
register in the course of the coupling does not affect the ability of the algorithm
to find eigenvectors and eigenvalues.
The requirement that the initial state of the continuous variable be perfectly
squeezed can also be relaxed. Suppose that the initial state is in a Gaussian state
J e-.B x2 / 2 Ix)dx. For example, {3 = 1 gives the un squeezed n = 0 vacuum
state, while {3 > 1 gives partial squeezing in X. With this initial state for the
continuous variable, after the algorithm has been run, the continuous variable
and the system are in the state
~
J
f e-.B x2 / 2 IE j) Ix + Ejt)dx. (5.3)
That is, the eigenvalues and eigenvectors are resolved to within an accuracy
1/t..J!3. By coupling the system to the continuous variable for a sufficiently
long time, the eigenvectors and eigenvalues of H may be determined to an
arbitrary degree of accuracy, even when the initial state is unsqueezed. Note
that resolving the eigenvalues of a system with an exponentially large number of
states requires exponential squeezing of the pointer state. But as noted in [21],
this algorithm still provides a potentially exponential speedup over classical
44 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
References
[1] R. P. Feynman, Optics News 11, 11 (1985); reprinted in Found. Phys. 16,
507 (1986).
[2] D. Deutsch, Proc. R. Soc. London A 400,97-117 (1985).
[3] S. Lloyd, Science 261, 1569-1571 (1993).
[4] P. Shor, in Proceedings 0/ the 35th Annual Symposium on Foundations
o/Computer Science, edited by S. Goldwasser, (IEEE Computer Society,
Los Alamitos, CA, 1994), pp. 124-134.
[5] S. Lloyd, Sci. Am. 273, 140-145 (1995).
[6] D. Divincenzo, Science 270,255-261 (1995).
[7] S. Lloyd, Phys. Rev. Lett. 75, 346 (1995).
[8] J. I. Cirac and P. Zoller, Phys. Rev. Lett. 74, 4091-4094 (1995).
[9] Q. A. Turchette, C. J. Hood, W. Lange, H. Mabuchi and H. J. Kimble,
Phys. Rev. Lett. 75, 4710-4713 (1995).
[10] C. Monroe, D. M. Meekhof, B. E. King, W. M. Itano and D. J. Wineland,
Phys. Rev. Lett. 75, 4714-4717 (1995).
[11] T. Pellizzari, S. A. Gardiner, J. I. Cirac and P. Zoller, Phys. Rev. Lett. 75,
3788-3791 (1995).
[12] S. Lloyd and J. J.-E. Slotine, Phys. Rev. Lett. 80,4088 (1998).
[13] S. L. Braunstein, Phys. Rev. Lett. 80,4084 (1998).
[14] S. L. Braunstein, Nature 394,47 (1998).
[15] S. Lloyd, S. L. Braunstein, Phys. Rev. Lett. 82, 1784-1787 (1999).
[16] J. Preskill and A. Kitaev, to be published.
[17] Y. Weinstein, S. Lloyd and D. Cory, quant-ph/9906059.
Hybrid quantum computing 45
Abstract We obtain sufficient conditions for the efficient simulation of a continuous vari-
able quantum algorithm or process on a classical computer. The resulting theorem
is an extension of the Gottesman-Knill theorem to continuous variable quantum
information. For a collection of harmonic oscillators, any quantum process that
begins with unentangled Gaussian states, performs only transformations gener-
ated by Hamiltonians that are quadratic in the canonical operators, and involves
only measurements of canonical operators (including finite losses) and suitable
operations conditioned on these measurements can be simulated efficiently on a
classical computer.
• S. D. Bartlett, B. C. Sanders, S. L. Braunstein and K. Nemoto, Physical Review Letters 88, 097904/1-4
(2002).
Copyright (2002) by the American Physical Society.
47
48 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(6.1)
X(q)ls) = Is + q) ,
~
Z(P)ls) = exp(/iPs) Is) . (6.3)
(6.4)
Efficient classical simulation ofcontinuous variable quantum information processes 51
This gate is an interaction gate operation on the Pauli group (12 for two systems.
Referring to the definition (6.1) for the Pauli operators for a single system, the
action of this gate on the 92 Pauli operators is given by
F : X(q) -+ Z(q) ,
Z(p) -+ X(p)-l. (6.7)
[The operator P(",) is called the phase gate, in analogy to the discrete-variable
phase gate P [15], because of its similar action on the Pauli operators.]
For discrete variables, it is possible to generate the Clifford group using only
the SUM, F, and P gates [15]. However, for the CV definitions above, the
operators SUM, F, and P(",) are all elements of Sp(2n, lR); they are generated
by homogeneous quadratic Hamiltonians only. Thus, they are in a subgroup of
the Clifford group. In order to generate the entire Clifford group, one requires
a continuous HW(I) transformation [Le., a linear Hamiltonian, that generates a
one-parameter subgroup of HW(I)] such as the Pauli operator X(q). This set
{SUM, F, P(",), X(q); "', q E lR} generates the Clifford group.
We now have the necessary components to prove the main theorem of this
paper regarding efficient classical simulation of a CV process. We employ the
stabilizer formalism used for discrete variables and follow the evolution of the
52 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Pauli operators rather than the states. To start with, let us consider the ideal
case of a system with an initial state in the computational basis of the form
Iq1, q2, ... , qn}. This state may be fully characterized by the eigenvalues of
the generators of n Pauli operators {q1, q2, . .. , qn}. Any continuous variable
process or algorithm that is expressed in terms of Clifford group transformations
can then be modeled by following the evolution of the generators of these n
Pauli operators, rather than by following the evolution of the states in the Hilbert
space £2 (Rn ). The Clifford group maps linear combinations of Pauli operator
generators to linear combinations of Pauli operator generators (each qi and Pi
is mapped to sums of qj, Pj, j = 1, ... ,n in the Heisenberg picture). For
each of the n generators describing the initial state, one must keep track of 2n
real coefficients describing this linear combination. To simulate such a system,
then, requires following the evolution of 2n2 real numbers.
In the simplest case, measurements (in the computational basis) are per-
formed at the end of the computation. An efficient classical simulation involves
simulating the statistics of linear combinations of Pauli operator generators. In
terms of the Heisenberg evolution, the qj are described by their initial eigen-
values, and the pj in the sum by a uniform random number. This prescription
reproduces the statistics of all multi-mode correlations for measurements of
these operators.
Measurement in the computational basis plus feed-forward during the com-
putation may also be easily simulated for a sufficiently restricted class of feed-
forward operations; in particular, operations corresponding to feed-forward
displacement (not rotation or squeezing, though this restriction will be dropped
below) by an amount proportional to the measurement result. Such feed-
forward operations may be simulated by the Hamiltonian that generates the
SUM gate with measurement in the computational basis delayed until the end
of the computation. In other words, feed-forward from measurement can be
treated by employing conditional unitary operations with delayed measure-
ment [2], thus reducing feed-forward to the case already treated.
In practice, infinitely squeezed input states are not available. Instead, the
initial states will be of the form
(6.10)
glement between systems may still satisfy the conditions of the theorem and
thus may be simulated efficiently on a classical computer; included are those
used for CV quantum teleportation [4], quantum cryptography [6,7,8,9], and
error correction for CV quantum computing [12, 13]. Although these processes
are of a fundamentally quantum nature and involve entanglement between sys-
tems, this theorem demonstrates that they do not provide any speedup over a
classical process. Thus, our theorem provides a valuable tool in assessing the
classical complexity of simulating these quantum processes.
As shown in [10], in order to generate all unitary transformations given by
an arbitrary polynomial Hamiltonian (as is necessary to perform universal CV
quantum computation), one must include a gate described by a Hamiltonian
other than an inhomogeneous quadratic in the canonical operators, such as a
cubic or higher-order polynomial. Transformations generated by these Hamil-
tonians do not preserve the Pauli group, and thus cannot be described by the
stabilizer formalism. Moreover, any such Hamiltonian is sufficient [10]. One
example would be to include an optical Kerr nonlinearity [17], but there is a lack
of sufficiently strong nonlinear materials with low absorption. Alternatively,
it has recently been proposed that a measurement-induced nonlinearity (using
ideal photodetection) could be used in an optical scheme without the need for
nonlinear materials in the computation [15, 18]. The physical realization of
such nonlinearities is an important quest for quantum information theory over
continuous variables. These nonlinear transformations can be used in CV al-
gorithms that do not satisfy the criteria of this theorem, and which may provide
a significant speedup over any classical process.
References
[1] D. Gottesman, in Proceedings of the XXII International Colloquium on
Group Theoretical Methods in Physics, edited by S. P. Corney et ai.,
(International Press, Cambridge, MA, 1999), p. 32.
[2] M. A. Nielsen and 1. L. Chuang, Quantum Computation and Quantum In-
formation (Cambridge University Press, Cambridge, U.K., 2000), p. 464.
[3] E. Knill, quant-ph/0108033; B. M. Terhal and D. P. DiVincenzo, quant-
ph/O 1080 10.
[4] S. L. Braunstein and H. J. Kimble, Phys. Rev. Lett. 80, 869 (1998).
[5] A. Furusawa et al., Science 282, 706 (1998).
[6] T. C. Ralph, Phys. Rev. A 61, 010303(R) (2000).
Efficient classical simulation ofcontinuous variable quantum information processes 55
QUANTUM ENTANGLEMENT
Chapter 7
INTRODUCTION TO ENTANGLEMENT-BASED
PROTOCOLS
Samuel L. Braunstein
Informatics, Bangor University, Bangor LL57 1 UT, United Kingdom
[email protected]
Arun K. Pati
Institute of Physics, Bhubaneswar-751005, Orissa, INDIA
Theoretical Physics Division, BARC, Mumbai, INDIA
[email protected]
1. INTRODUCTION
Quantum entanglement was first introduced by SchrOdinger and explored by
Einstein-Podolsky-Rosen (EPR) in their famous paper on the incompleteness of
quantum theory [1]. As SchrOdinger put it, the phenomenon of entanglement
is one of the quintessential features in quantum mechanics that has no ana-
logue in classical physics. Typically, when two quantum systems interact, the
wavefunction of one can get intertwined with the wavefunction of other. The
combined wavefunction of the composite system can show strong correlations
even though they may be widely separated. In an attempt to understand these
correlations Bell constructed an inequality that must be satisfied for all local
realistic models [2]. Surprisingly, he noted that entangled states violate such
inequalities. Thus demonstrating that indeed entanglement resists any classical
explanation.
Over the last decade it has been realised that quantum entanglement is an
important resource in quantum information and computation. In particular,
59
S.L Braunstein andA.K. Pati (eds.J, Quantum Information with Continuous Variables, 59-66.
© 2003 Kluwer Academic Publishers.
60 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
there are numerous quantum communication protocols that require shared en-
tanglement between a pair of parties, Alice and Bob. Utilizing this shared
resource together with local operations and classical communication (LOCC),
Alice and Bob can perform various feats: Alice can teleport an unknown state
to Bob using an entangled pair and two classical bits [3]; dense coding of
classical information is possible with shared entanglement between sender and
receiver [4]; and one may create entanglement between a pair of systems that
have never interacted in the past, called entanglement swapping [5].
For these protocols to work perfectly one requires ideal EPR pairs. This
is a problem since such states are highly prone to errors during transportation
or storage. The resulting imperfect EPR pairs behave like a noisy channel
in the above protocols and no longer work in an ideal manner. Fortunately,
there is yet another protocol that is capable of purifying such imperfect EPR
pairs (provided they are still entangled). In purification, two parties start with
a given number of imperfect pairs and by performing LOCC operations they
produce a smaller number of more highly entangled pairs. This process may
be repeated, provided the resource has not been exhausted, until the EPR pairs
are sufficiently ideal for carrying out the desired protocol.
In this article we briefly discuss entanglement, teleportation, entanglement
swapping and purification protocols for finite-dimensional systems (especially
for qubits). We hope this will motivate readers for the continuous variable
generalizations discussed in the following chapters.
2. QUANTUM ENTANGLEMENT
Entanglement is always a property of a composite system viewed as a system
consisting of subsystems. If the whole system is always viewed as a single
object then we would not bother about the role of entanglement. Entanglement
is a feature of states in Hilbert spaces which have been given a tensor product
structure. In the case of a composite system consisting of two subsystems A
and B (called a bipartite system) the combined state lives in llAB = llA 01lB.
If such a system can be described by a state which can be written as
(7.1)
where I~)A and I¢)B are states of the subsystems A and B, respectively, then
the state is not entangled (which is synonymous with it being separable). If
one cannot assign a definite pure state to each sub-system then it is entangled.
This defines entanglement for pure states.
Introduction to entanglement-based protocols 61
where {IXi)} and {IYj)} are orthonormal basis states in the Hilbert space 1-lA
and 1-lB, respectively. Now the Schmidt decomposition theorem tells us that
by performing a singular-value decomposition of the complex matrix (Cij), we
may write any arbitrary bipartite state as [6]
min(NA,NB)
IW)AB = L JPili)Ali')B , (7.3)
where li)A and li')B are the singular-value transformed bases and the Pi'S are
non-zero eigenvalues of the reduced density matrices of the subsystem A or
B. We note the.../Pi are called Schmidt numbers and EiPi = 1. One can see
then that for pure states if there is more than one Schmidt number, then the
state is entangled. It should be remembered that the Schmidt decomposition
theorem holds only for a bipartite system and does not hold for tripartite
systems or higher. (However, there are special conditions under which Schmidt
decomposition theorem may be found for tripartite systems [7].)
More generally, states of a composite system need not be pure. If the
composite system is described by a mixed state, then we need a different
definition for entanglement. In general, we say that a state p AB is separable
(not entangled) on 1-lA ® 1-lB if it can be written as [8]
(7.4)
where p~) and p~) are states in 1-lA and 1-lB, respectively and the qi are non-
negative real weights. Such states display only classically correlations. This is
because two parties may always prepare such states from locally created states
plus classical communication.
One may ask how much entanglement such states contain. But to answer this
question we must define a measure of entanglement. It has been proposed that
any measure of entanglement E(p) for a state p (whether p is pure or mixed)
should satisfy three conditions [9]:
1. E(p) should be zero if and only if p is separable.
2. E (p) should be invariant under local unitary operations. This means that
if p -+ (UA ® UB)p(UA ® UB)t = p', then E(p) = E(p').
62 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(7.5)
where Pi = AipA! and the Ai are local Kraus operators [11] satisfying
L::i A! Ai = 1.
and has units of entangled bits or "ebits." This can reach maximum value
when all the Pi are equal. Hence the maximum amount of entanglement in this
system is log N assuming N = min( N A, N B). One can define a maximally
entangled state I\[! max) AB to be
(7.7)
(7.S)
with 10), 11) being the computational basis states for qubits. One can easily
check that if we adopt the measure of entanglement as the von Neumann entropy
given in Eq. (7.6), then an EPR state contains exactly 1 ebit of entanglement
which is the maximum amount of entanglement that any 2 x 2-dimensional
Hilbert space can admit. Hence, this EPR pair is also a maximally entangled
state of two-qubits. It may be worth observing that, if we have a bipartite sate
in an N x N-dimensional Hilbert space with N = 2n , then the state I\[! max) is
equivalent to n (qubit-based) EPR pairs.
We now consider a number of communication protocols that may be achieved
with shared entanglement.
Introduction to entanglement-based protocols 63
3. QUANTUM TELEPORTATION
Suppose that Alice and Bob are at widely separated locations. Alice wants
to send an unknown qubit to Bob without physically sending it! How is it
possible?
Let Alice and Bob share an EPR pair IEPR) 12 and have access to particles 1
and 2, respectively. If Victor gives a qubit to Alice (whose identity is unknown
to her) of the form
(7.9)
then the combined state of the input and EPR pair is just I'l/J) a ® IW-) 12. This
may be expressed in terms of a basis of "Bell-states" for particles a and 1 as
where R(l) = iO"y, R(2) = O"x, R(3) = o"z are Pauli matrices and R(4) = IT.
Here the states I<p±) and IW±) are the canonical Bell-states first introduced in
Ref. [12] as
~(IO)IO) ± 11)11))
~(IO)ll) ± 11)10)) . (7.11)
4. DENSE CODING
Usually, from a transmitted qubit one can extract only one classical bit upon
measurement. However, if the sender and receiver share prior entanglement
then sending a suitably modulated version of the shared half (corresponding to
one qubit) to receiver, one can now extract two classical bits of information.
This doubling of classical capacity of quantum channel (with the assistance of
entanglement) is called dense coding or sometimes super dense coding.
Let us imagine that Alice and Bob share a canonical Bell-state
(7.12)
One important property of the Bell-states is that one can convert anyone of
them to any other by acting locally on one half of the system. In particular,
Alice can apply a set of local unitary operations giving:
After applying any of these four local operations, Alice sends her particle to
Bob. Then, by performing a Bell-measurement Bob can distinguish each case
thus, extracting two classical bits of information. It may be noted that dense
coding is closely related to quantum teleportation. In teleportation, one sends
a single qubit via two cbits plus shared entanglement, whereas in dense coding
one sends two cbits via one qubit plus shared entanglement.
5. ENTANGLEMENT SWAPPING
Entanglement swapping is a method to create entanglement between two
independent particles that have never interacted in the past. Though at first
glance this sounds strange, it is indeed possible in the quantum world [5].
Let us consider two independent sources each emitting an EPR pair 1,2 and
3,4, respectively. This means particles 1 and 2 are entangled and similarly for
particles 3 and 4. However, neither 1 and 4 nor 2 and 3 are entangled. Let Alice
Introduction to entanglement-based protocols 65
take particle 1, Charlie take particle 4 and Bob, the man in the middle, will take
particles 2 and 3. The combined state of the system can then be written as
6. ENTANGLEMENT PURIFICATION
As we have already noted, in practice ideal EPR pairs may undergo deco-
herence and become mixed entangled states. In the absence of a maximally
entangled resource one can no longer faithfully perform teleportation, dense
coding or entanglement swapping. This suggests that it would be useful if we
could improve the quality of imperfect EPR pairs through some other protocol.
Fortunately, this is possible through "entanglement purification" [13]. Since
we wish to restrict this protocol to LOCC operations we will not be able to win
on average. However, provided we have some entanglement there is still the
possibility of probabilistic ally purifying our imperfect EPR pairs.
Suppose Alice and Bob share multiple copies of an imperfectly entangled
EPR pair of the form
(7.15)
This state is a Werner mixture with the spin-singlet appearing with probability
f. This mixture is entangled provided f > ~ [13]. The simplest protocol
works on pairs of such imperfect states at a time. By performing suitable
LOCC operations [13] Alice and Bob will be able to post-select a single pair
66 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
having the form PAB(J/) with f' > f provided f > ~. This process only
occurs with a finite probability, however, every time it works a higher quality
entangled resource is purified out of several lesser quality states. Provided pairs
are available, this process may be repeated until a suitably high quality EPR
pair is created. In principle, the fidelity relative to an ideal EPR can approach
unity.
7. CONCLUSION
We have briefly touched on basic concepts of quantum entanglement and
its utility in quantum information processing. (More details may be found in
Refs. [11, 14].) Of these protocols, only entanglement purification does not
straightforwardly go over to continuous variable schemes, as we shall see in
the following chapters.
References
[1] A. Einstein, B. Podolsky, and N. Rosen, Phys. Rev. 47, 777 (1935)
[2] J. S. Bell, Physics 1, 195 (1964)
[3] c. H. Bennett, G. Brassard, C. Crepeau, R. Jozsa, A. Peres and W. K.
Wootters, Phys. Rev. Lett. 70, 1895 (1993).
[4] C. H. Bennett and S. J. Wiesner, Phys. Rev. Lett. 69, 2881 (1992).
[5] M. Zukowski, A. Zeilinger, M. A. Horne and E. Ekert, Phys. Rev. Lett.
71,4278 (1993).
[6] A. Peres, Quantum Theory: Concepts and Method (Kluwer Academic
Publisher, 1995).
[7] A. K. Pati, Phys. Lett. A 278, 118 (2000), and references therein.
[8] R. F. Werner, Phys. Rev. A 40, 4277 (1989).
[9] V. Vedral and M. Plenio, Phys. Rev. A 57, 1619 (1998).
[10] S. Popescu and D. Rohrlich, Phys. Rev. A 56, R3319 (1997).
[11] M. A. Nielsen and 1. L. Chuang, Quantum Computation and Quantum
Information, (Cambridge University Press, Cambridge, England, 2000).
[12] S. L. Braunstein, A. Mann and M. Revzen, Phys. Rev. Lett. 68, 3259
(1992).
[13] c. H. Bennett, H. J. Bernstein, S. Popescu, B. Schumacher, J. A. Smolin
and W. K. Wootters, Phys. Rev. Lett. 76, 722 (1996).
[14] J. Preskill, Lecture Notes, available at
http://www.theory.caltech.edu/people/preskill/ph229
Chapter 8
TELEPORTATION OF CONTINUOUS
QUANTUM VARIABLES*
Samuel L. Braunstein
Informatics, Bangor University, Bangor LL57 I Ur, United Kingdom
[email protected]
H. J. Kimble
Norman Bridge Laboratory of Physics 12-33
California Institute of Technology
Pasadena, CA 91125
Abstract Quantum teleportation is analyzed for states of dynamical variables with con-
tinuous spectra, in contrast to previous work with discrete (spin) variables. The
entanglement fidelity of the scheme is computed, including the roles of finite
quantum correlation and nonideal detection efficiency. A protocol is presented
for teleporting the wave function of a single mode of the electromagnetic field
with high fidelity using squeezed-state entanglement and current experimental
capability.
67
68 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Ollt
2
W;,,(a) 1
Figure 8.1 Scheme for quantum teleportation of an (unknown) input state Win (a) from Alice's
sending station S to Bob's remote receiving terminal R, resulting in the teleported output state
Wout(a).
where aj = Xj + ipj. Here, the real quantities (Xj, pj) correspond to canoni-
cally conjugate variables for the relevant pathways and describe, for example,
position and momentum for a massive particle, and quadrature amplitudes for
the electromagnetic field. Note that for r -+ 00, the state described by Eq. (8.1)
becomes precisely the EPR state of Ref. [3] employed by Vaidman [5] and pro-
vides an ideal entangled "pair" shared between the teleportation sending and
receiving stations, albeit with divergent energy in this limit.
As for the protocol itself, the first step in teleporting the (unknown) state
Win(ain) is to form new variables {3a,b along paths (a, b) which are linear
superpositions of those of the initially independent pathways in and 1 at the
sending station S of Fig. 8.1, namely {3a,b = ~(al ± ain). The resulting
Wigner function in the variables ({3a; (3b; (2) exhibits "entanglement" between
the paths (a, b) and the remote path 2. Step 2 at S is then to measure the
observables corresponding to Re{3a = ~(Xl + Xin) == Xa and Im{3b =
~(Pl - Pin) == Pb at the detectors (Da, Db) shown in Fig. 8.1, with the
reSUlting classical outcomes denoted by (ixa' i pb ), respectively. We define
ideal measurement of (Xa,Pb) to be that for which the distribution Pab( iXa; i pb )
is identical to the associated Wigner function Wab(Xa;Pb). With the entangled
70 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
where the variance T = sech2r /2. Note that as r --* 00, G T (a) quickly
approaches a delta-function, while Gv(a) describes a broad background state.
Thus, for large r, the reduced state of mode 2 is described by a broad pedestal
with negligible probability upon which sits a randomly located peak at 0:'2 ~
.J2(i xa - i i pb ) closely mimicing the incoming state Win(a). The location
of this random "displacement" is distributed according to Eq. (8.2), and is the
classical information that Alice sends to Bob.
By way of the actuator Ax,p shown in Fig. 8.1, Bob thus performs linear
displacements of the real and imaginery components of the complex amplitude
a2 to produce aout = a2 + .J2(i xa - i i pb ), where the quantities (ixa' i pb ) are
scaled to (Xa,Pb). Integrating out iXa and ipb yields the ensemble description
of states produced at the output of the teleportation device on an ensemble of
input states Win, namely
noise contributions arises from Alice's attempt to measure both (Xin,Pin) [14],
while the second comes from Bob's use of this necessarily noisy information
to generate a coherent state at V2( iXa - i i pb )' In this way quantum mechanics
extracts two tariffs (one at each instance of the border crossing between quantum
and classical domains), each of which we term the quantum duty (or quduty).
Note that the limit r = 0 corresponds to what might be considered "classical"
teleportation for which the "best measurement" of the coherent amplitude of
the unknown state is made [14] and sent to the receiving station, where it is
used to produce a coherent state of that classical amplitude. For any r > 0, our
quantum teleportation protocol beats this classical scheme.
Before calculating an actual figure of merit for our protocol, we now spe-
cialize from general continuous variables to the case of a single mode of the
electromagnetic field and thereby to actual physical implementations of the
various transformations shown in Fig. 8.1. Beginning with the EPR state itself,
we note that such a state can be generated by nondegenerate parametric ampli-
fication with the quantities (x j, pj) as the quadrature-phase amplitudes of the
field [6], as has been experimentally confirmed via Type II down-conversion
[7]. The linear transformation i3a,b = ~(al ± ain) is accomplished by the
simple superposition of modes in and 1 at a 50/50 beam splitter. The detectors
(D a, Db) of Fig. 8.1 are now just balanced homodyne detectors with the phases
of their respective local oscillators set to record (Xa,Pb) in the observed pho-
tocurrents (ixa' i pb )' Note that for unit efficiency, homodyne detection provides
an ideal quantum measurement of the quadrature amplitudes required for our
protocol [15, 16, 17].
Non-ideal detectors, each having (amplitude) efficiency 'f/, may be modeled
by using a pair of auxiliary beam splitters at (Da, Db) to introduce noise from a
pair of vacuum modes described by annihilation operators (Ca,b, da,b) [15, 18].
It is then convenient to introduce annihilation operators corresponding to the
"modes" of the photocurrents described by
~ ~
Za,b = 'f/i3a,b + V~-2-(Ca,b + da,b) ,
A A A
(8.5)
where G( has variance ( = (1 - 'f/2)/2'f/2, which goes to zero for 'f/ --+ 1 in
correspondence with the ideal character of homodyne detection. Substituting
72 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(8.9)
where Pin is the original state being teleported and b(a) is the displacement
operator. The dynamics associated with Eq. (8.9) were first studied by Glauber
[19] and Lachs [20] for an "incoming" vacuum state p = 10)(0 Iand for squeezed
vacuum by Vourdas and Weiner [21]. The detailed behavior of the photocount
statistics under this dynamics was investigated by Musslimani et ai. [22].
These references also relate the development of the convolutional formalism
used here (see also Refs. [23,24]).
To illustrate the protocol, consider teleportation of the coherent superposition
state
(8.10)
0.4 0.4
0.2 0.2
0 0
-0.2 -0.2
-0.4 -0.4
a) W.In b) W
out
-0.6 -0.6
2 2
0
-2 -1 0 -1
Figure 8.2 (a) Wigner function Win(a) for the input state ofEq. (8.10) with a = 1.5i and
¢;= 7r. (b) Teleported output state Wout(a) for r = 1.15 and 1/ 2 = 0.99.
where X . (~) = tr D(i~) Pin is the characteristic function for the incoming
Vl1n
state's Wigner function.
For the coherent superposition of Eq. (8.10) direct substitution yields a
fidelity of entanglement Fe of
(8.12)
For the state shown in Fig. 8.2(b) this fidelity is 0.6285 for r = 1.15 and rt 2 =
0.99 compared to 0.2487 for r = 0 and the same detector efficiency. This latter
fidelity precludes observation of any quantum features in the classically tele-
ported state, while the former case yields observable quantum characteristics
as seen in Fig. 8.2.
74 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(8.13)
where IDoal 2 == (laI 2) - l(a)12 averaged over Win(a). Thus, the condition
for high fidelity teleportation (i.e., 1 - Fe « 1) becomes I/IDoal 2 » a. Now
IDoal 2 is just the number of photons (plus ~) in the incoming state after it
has been shifted so as to have no coherent amplitude. Roughly speaking it is
the maximal rms spread of the Wigner function of the unknown quantum state
being teleported and so its reciprocal bounds the size of "important" small scale
features in that state, though there can indeed be smaller features. Apparently
then the condition for high entanglement fidelity says that features in the Wigner
function smaller than I/IDoal do not give a significant contribution to the state's
identity.
In conclusion, our analysis suggests that existing experimental capabilities
should suffice to teleport manifestly quantum or nonclassical states of the
electromagnetic field with reasonable fidelity. For such experiments, extensions
of our analysis to the teleportation of broad bandwidth information must be
made and will be discussed elsewhere. In qualitative terms, our scheme should
allow efficient teleportation every inverse bandwidth, in sharp constrast to
relatively rare transfers for proposals involving weak down conversion for spin
degrees of freedom. Although our analysis is the first to obtain explicitly the
fidelity of entanglement on an infinite dimensional Hilbert space, an unresolved
issue is whether or not our protocol is "optimum," either with respect to this
measure or with regard to other criteria in the area of quantum communication
(e.g., the ability to teleport optimally an "alphabet" {j} of orthogonal states
Win)' More generally, the work presented here is part of a larger program
to extend classical communication with complex amplitutes into the quantum
domain.
S.L.B. was funded in part by EPSRC Grant No. GRlL91344 and a Humboldt
Fellowship. H.J.K. acknowledges support from DARPA via the QUIC Insti-
tute administered by ARO, from the Office of Naval Research, and from the
National Science Foundation. Both appreciate the hospitality of the Institute
for Theoretical Physics under National Science Foundation Grant No. PHY94-
07194.
Teleportation of continuous quantum variables 75
References
[1] A. Steane, LANL Report No. quant-phl9708022; A. S. Holevo, LANL
Report No. quant-phl9708046.
[2] C. H. Bennett, G. Brassard, C. Crepeau, R. Jozsa, A. Peres and W. K.
Wootters, Phys. Rev. Lett. 70, 1895 (1993).
[3] A. Einstein, B. Podolsky and N. Rosen, Phys. Rev. 47, 777 (1935).
[4] J. S. Bell, in Speakable and Unspeakable in Quantum Mechanics (Cam-
bridge Univ. Press, 1988), p. 196.
[5] L. Vaidman, Phys. Rev. A 49, 1473 (1994).
[6] M. D. Reid and P. D. Drummond, Phys. Rev. Lett. 60, 2731 (1988); M.
D. Reid, Phys. Rev. A 40,913 (1989).
[7] (a) z. Y. Ou, S. F. Pereira, H. J. Kimble and K. C. Peng, Phys. Rev. Lett.
68,3663 (1992); (b) Appl. Phys. B 55, 265 (1992).
[8] L. Davidovich, N. Zagury, M. Brune, J. M. Raimond and S. Haroche,
Phys. Rev. A 50, R895 (1984).
[9] J.1. Cirac and A. S. Parkins, Phys. Rev. A 50, R4441 (1994).
[10] T. Sleator and H. Weinfurter, Ann. N. Y. Acad. Sci. 755, 715 (1995).
[11] S. L. Braunstein and A. Mann, Phys. Rev. A 51, RI727 (1995); 53, 630(E)
(1996).
[12] D. Boumeester et al., Nature (London) 390, 575 (1997).
[13] B. Yurke, S. L. McCall and J. R. Klauder, Phys. Rev. A 33,4033 (1986).
[14] E. Arthurs and J. L. Kelly Jr., Bell. Syst. Tech. J. 44, 725 (1965).
[15] H. P. Yuen and J. H. Shapiro, IEEE Trans. Inf. Theory 26, 78 (1980).
[16] S. L. Braunstein, Phys. Rev. A 42,474 (1990).
[17] Z. Y. Ou and H. J. Kimble, Phys. Rev. A 52,3126 (1995).
[18] K. Banaszek and K. W6dkiewicz, Phys. Phys. ASS, 3117 (1997).
[19] R. J. Glauber, Phys. Rev. 131,2766 (1963).
[20] G. Lachs, Phys. Rev. 138, B1012 (1965).
[21] A. Vourdas and R. M. Weiner, Phys. Rev. A 36, 5866 (1987).
[22] Z. H. Musslimani, S. L. Braunstein, A. Mann, and M. Revzen, Phys. Rev.
A 51, 4967 (1995).
[23] M. S. Kim and N. Imoto, Phys. Rev. A 52,2401 (1995).
[24] K. Banaszek and K. W6dkiewicz, Phys. Rev. Lett. 76, 4344 (1996).
[25] E. S. Polzik, J. Carri and H. J. Kimble, Phys. Rev. Lett. 68, 3020 (1992);
(b) Appl. Phys. B 55, 279 (1992).
[26] B. Schumacher, Phys. Rev. A 54,2614 (1996).
Chapter 9
EXPERIMENTAL REALIZATION
OF CONTINUOUS VARIABLE
TELEPORTATION
Akira Furusawa
Department of Applied Physics, University of Tokyo
7-3-1 Hongo, Bunkyo-ku, Tokyo113-8656, Japan
H. J. Kimble
Norman Bridge Laboratory of Physics
California Institute of Technology, mcl2-33
Caltech, Pasadena, CA9II25, USA
1. INTRODUCTION
Quantum teleportation is a method of quantum state transportation with a
classical channel and a quantum channel [1]. In this technique, the "informa-
tion" contained in a quantum state is transferred from a sending station (Alice)
to a receiving station (Bob), with the original quantum state thereby recon-
structed at Bob's place with the received information and previously shared
entanglement. Note that it is impossible to perform the state transformation
represented by quantum teleportation only with a classical channel, which can
be qualitatively explained as follows. If one attempts to obtain complete in-
formation with some particular measurement on an unknown quantum state of
motion, for example, then both position and momentum (canonically conjugate
variables) must be determined simultaneously with negligible error, which is
of course impossible [2]. It is thus impossible for Alice to obtain complete
information on the unknown quantum state, so that she certainly cannot send
enough information for the reconstruction of the state to Bob. He then is unable
to reconstruct the complete state at his place. By contrast, in quantum tele-
portation, Alice and Bob neatly circumvent constraints that would otherwise
be imposed on Alice's state measurement and Bob's state generation, and are
thereby able to reconstruct the original state at Bob's place.
77
S.L. Braunstein and A.K. Pati (eds.), Quantum Information with Continuous Variables, 77-93.
© 2003 Kluwer Academic Publishers.
78 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
The essential character for this trick is quantum entanglement. For the case
of continuous quantum variables, the quantum entanglement is that discussed
by Einstein, Podolsky, and Rosen, in now the famous "EPR paradox" [3] .
The sharing of quantum entanglement between Alice and Bob is the critical
resource employed to enable the kind of quantum information transfer implicit
in quantum teleportation.
Unitary
Bell measurement transfonnation
Source of
Alice entanglement Bob
Victor
Classical infonnation
2. EXPERIMENTAL REALIZATION
OF CONTINUOUS VARIABLE TELEPORTATION
2.1 EPR CORRELATION
The quantum entanglement relevant to our continuous variable teleportation
experiment is the EPR correlation as in the original 1935 paper [3], which
relates to the canonically conjugate position and momentum variables. As
was originally pointed out by Reid and Drummond [13], the EPR gedanken
experiment can be realized via the quadrature-phase amplitudes of a single-
mode of the electromagnetic field, since these variables are also canonically
conjugate in direct correspondence to position and momentum. Indeed, the first
realization of the EPR experiment was accomplished by way of the quadrature-
phase amplitudes using nondegenerate parametric down-conversion process
80 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
with type-II phase matching [14]. An equivalent EPR state can also be produced
with type-I phase matching and a half beam splitter [12]. This state is called a
two-mode squeezed vacuum. In the Heisenberg representation, the quadrature-
phase amplitude operators (Xj, pj) are transformed as follows [15, 16].
1 r ,,-(0) 1 -r ,,-(0)
Xl = yl2e Xl + .,fie X2 ,
1 -r ,,-(0) 1 r ,,-(0)
PI = yl2e PI + yl2e P2
(9.1)
°
certainly not infinite in any real experiment; however, entanglement exists for
any finite value of r > (even infinitesimal r and even in the presence of loss).
[17, 18]
EPR
1 " be:uns
"
Figure 9.2 Experimental setup for the test of the EPR correlation of continuous variables.
Experimental realization of continuous variable teleportation 81
Figure 9.2 shows the experimental setup for the realization ofEPR correlation
of continuous variables [12, 14]. Two squeezed vacuum beams are created by
parametric down conversion in a subthreshold optical parametric oscillator
(OPO). The OPO consists of a ring cavity and a potassium niobate K Nb0 3
crystal and has two independent opposite circulations. Two pump beams
(430nm) drive the OPO in the opposite directions; these pump beams are
created by way of second harmonic generation with another K Nb0 3 crystal
in an external buildup cavity which is itself driven by the fundamental 860nm-
output of a titanium:sapphire Ti : Al 2 0 3 laser. The two squeezed vacuum
beams are combined with a half beam splitter, where the relative phases of
the input squeezed vacuum beams are locked to be ~ with respect to each
other. The beams emerging from the beam splitter are in a two-mode squeezed
vacuum, thereby realizing entangled EPR beams. One of the EPR beams goes
to Alice's station and the other beam goes to Bob's station. Alice and Bob's
stations consist in the first instance of homodyne detectors which enable us to
confirm the correlation (and hence entanglement [17, 18]) between the beams
by way of the difference of the photocurrent from the detectors. This difference
signal corresponds to (Xl - X2) or uh +'fJ2) depending upon the relative phase
chosen between Alice and Bob's local oscillators (LO Alice, LOBob) [14].
Operationally speaking, the electromagnetic fields employed in experiments
such as these are not single-mode but rather have finite bandwidth. The single-
mode treatment must be generalized to the case of multimode fields of finite
bandwidth, as discussed in detail in Ref. [14]. In this situation, the relevant
quantities are the spectral components (x(O),p(O)) of the quadrature-phase
amplitudes, where a general quadratue-phase amplitude at phase 8 is defined
by
11!1+Ll!1
2(0,8) == -2 dO' [a(O') exp( -i8) + at (-0') exp( +i8)] (9.2)
!1-Ll!1
with a(a t ) as the annihilation (creation) operator for the field at offset 0 from
the optical carrier, with (x(O),p(O)) = (2(0,0),2(0,11"/2)), and with the
integration extending over a small interval .6.0 about O. Then, the spectral
density of photocurrent fluctuations \]! (0) obeys the following relation,
(9.3)
In simple terms, the overall treatment for the multimode case remains essen-
tially unchanged from the single-mode case. However, now the relevant state
describes the electromagnetic field at frequency offset ±O within a bandwidth
.6.0 about the carrier WL (laser frequency); that is to say, AM and FM modula-
tion sidebands.
Figure 9.3 shows an experimental result for the setup shown in Figure 9.2.
In Figure 9.3, the horizontal axis corresponds to the relative phase between
82 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Tim. (sec)
Figure 9.3 EPR correlation for Alice and Bob beyond the vacuum-state limit.
Alice and Bob's local oscillators. Actually the phase of Alice's local oscillator
(LO) is locked to the phase of the EPR beam, and the phase of Bob's LO is
scanned. The vertical axis represents the photocurrent fluctuation W(O)EPR
for the difference output between Alice and Bob's photocurrents. Note that
W(D) E P R = OdE == W0 corresponds to the classical limit of the difference
output without quantum correlation. It is determined experimentally without
inputs to Alice and Bob's homodyne detectors (i.e., with vacuum inputs to their
detectors), and is the level arising from uncorrelated vacuum fluctuations at each
of the balanced detectors (i.e., two units of vacuum noise). Most importantly
is to examine the minima for wmin(D)EPR, which corresponds to a quantum
correlation between Alice and Bob's output [(:h - 1:2) or (lh +P2)] if and only
if Wmin < wo, which is indeed the case in Figure 9.3. In fact, the work of Refs.
[17, 18] ensures that the observation W min < W 0 is sufficient as to guarantee
that the beams are entangled. Of course, for the ideal EPR state, the level Wmin
would become arbitrarily small as compared to Wo (tending to -00 when
expressed on a logarithmic scale as in Figure 9.3). The maximum value Wmax
of the periodical curve indicates the level of anticorrelation implicit in the EPR
state. The shape of the curve rising and falling between W min and W max shows
the appearance of correlation and an anticorrelation between the quadrature
amplitudes at Alice and Bob's station as the relative phase between Alice and
Experimental realization of continuous variable teleportation 83
Figure 9.4 Experimental setup for the continuous variable teleportation [12].
Pu
Pv (9.4)
(a) 30
25
o 20 40 60 80
Time [msec)
X'2 '
Xin - v 1n2 x 2 - V1n2'
L:e -r"(O) L:X u ,
Xtel X2 + g..,fixu,
Ptel P2 + g..,fipv. (9.6)
In the absence of losses ('fJ = 1) and for unity gain (g = 1), the quadrature
operators associated with the teleported state become
1n2 -r ,,(0)
Xtel Xin - Y X2 ,
A A
~e
Ptel
A
= Pin
A + Y1n2~e -r"-(O)
PI . (9.7)
10
. .. . •• ... ~
.....•.-.
5 ... . . .-•
--............. -
-2 o 2 4 6 A
Gain g? [dB)
Figure 9.6 The variation of the coherent amplitude Aout and of the variance ow with gain l
without EPR beams. The input amplitude Ain is +21dB above the vacuum-state limit in this
particular case. The solid lines are the theoretical curves for ( = 1.
we can detennine 9 = 1. From the Figure 9.6, we can see aw = 4.8dB for
9 = 1, whose meaning will be presented later.
Moving then to the case of teleportation in the presence of entangled EPR
beams, Bob combines his modulated beam with his EPR beam and reconstructs
the state to be teleported. In this process, the "noise" arising from the EPR
beam is effectively "subtracted" from Bob's modulated beam by destructive
interference at mBob.
Experimental results from this protocol are shown in Figure 9.7. The hori-
zontal axis corresponds to the phase of Victor's local oscillator, which is being
swept in time. The vertical axis 1lTVictor corresponds to the spectral density
of photocurrent fluctuations associated with the quadrature amplitudes Xtel (n)
and Pte I (n) measured by Victor for a fixed (but arbitrary) phase for the input
state. The maximum value of the periodic curve corresponds to a coherent
amplitude for the output state approximately 25dB above the vacuum-state
level <p~ictor; here, the gain has been set to be 9 ~ 1 as in the previous discus-
sion. This result shows the classical phase-space displacement is successfully
reconstructed.
88 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(a)
-5
6
. y Victor
(b) 5
o
4 --' --- . --~- .-. ~ .~ ---~-~ _._.,- AVictor
3
,
2
0 -- -'~~--~-------- __
._- - <l>Viclor
-1 o
0 20 40 60 80
Time [msec]
Figure 9.7 Bob's output verified by Victor. The part (b) is the expanded view wi th a ten-trace
average for the vacuum input for Alice. [12]
The minima of the trace for WVi ctor correspond to the variance of the output
state for the quadrature orthogonal to that of the coherent amplitude, and are
equivalent to the level AVictor shown by the labeled flat trace. The various
phase-independent traces in the figure correspond to the quantum noise levels
with the EPR beams present for Alice and Bob (A Victor), without these EPR
beams at both locations (Y6i ctor) , and with a vacuum-state input to Victor's
Experimental realization of continuous variable teleportation 89
homodyne detector (<p ~ ictor). Of course, "without the EPR beams" means
that vacuum noise (r = 0) invades Alice and Bob's stations, leading to a
degradation of the "quality" of teleportation.
Indeed, for teleportation of coherent states in the absence of shared entan-
glement between Alice and Bob (no EPR beams), Equation (9.7) shows that the
quantum noise for Bob's output becomes three units of vacuum noise (in either
quadrature, (f:l!i;~el)' (f:lP~el)). One unit comes from the original quantum noise
of the input coherent state, and the other two units correspond to successive
"quantum duties", the first being to cross the boundary from the quantum to
classical world (Alice's attempt to detect both quadrature amplitudes) and the
second from the classical to quantum (Bob's generation of a coherent displace-
ment) [11]. The experimental result T~ictor ~ 4.8dB in correspondence to a
factor of 3 above the vacuum-state limit in Figures 9.6 and 9.7 indicates almost
perfect performance of the "classical" teleportation with near unity detection
efficiency (recall ( = 0.97). As discussed in more detail in Ref. [8, 9], T~ictor
is the limit of "classical" teleportation, where explicitly we mean teleportation
without shared entanglement.
From Figure 9.7 and similar measurements, we determine that AVictor lies
1.ldB-lower than T~ictor. This means that quantum te1eportation is suc-
cessfully performed beyond the classical limit, as clarified by the following
discussion. To quantify the "quality" of the teleportation for a pure state l.,pin) ,
we calculate the teleportation fidelity F == (.,pin IPout l.,pin) [8,9]. For the case
of teleportation of coherent states, the boundary between classical and quantum
teleportation has been shown to be fidelity F = 0.50 [8, 9]. We stress that this
limit applies to the specific case of coherent states and only to the distinction
between what Alice and Bob can accomplish with and without shared entangle-
ment. Teleportation to accomplish other tasks in quantum information science
requires yet higher values for the fidelity.
Nonetheless, when the input state is a coherent state, the fidelity F of the
teleported output can be represented as follows [15]:
where aQ and a~ are the variances of the Q function of the teleported field
for the corresponding quadratures. The relevant variances aQ and ~ can be
determined from the measured efficiency factors in the experiment and are
90 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
where rx,p are the squeezing parameters for the respective quadrature compo-
nents, 6,2 characterize the (amplitude) efficiency with which the EPR beams
are propagated and detected along paths(1,2), and'fJ gives the (amplitude) effi-
ciency for detection of the unknown input state by Alice. We stress that all of
these quantities can be directly measured, so that the comparison of theory as
in the above equation and the experimentally recorded variances can be made
with no adjustable parameters.
(Jw X.P 6
(dB] 7
Without EPR beams
6
I~-L~~~~~~~~~~~
-.. -3 -2 -I 0 234
Figure 9.8 Variances a~l of the teleported field measured by Victor [12]. Open and filled
symbols in the figure are experimental results. The open squares represent the results for the case
with the slight imbalance of amount of squeezing in the two-mode squeezed vacuum. The filled
squares and triangles represent the results for the case of the balanced amount of squeezing. The
solid lines represent the theoretical predictions of Equation (9.9).
and a~ together with the independently measured values for the gain g, we can
use Equation (9.8) to arrive at an experimental estimate of the fidelity Fexp,
with the results shown by the points in Figure 9.8 for the cases with and without
the EPR beams present. We can also calculate Ftheory by way of Equations
(9.8, 9.9), with this theoretical prediction shown by the curves in Figure 9.9.
The agreement between theory and experiment is evidently quite good.
0.7
\
Qi
'U 0.3
u:
1"
/
0.2
0 .1 ,I
.!ol.
~.
0.0
~ ·3 ·2
-
.,
,IV'"
0 2 3
Gain 92 [dB]
Figure 9.9 Fidelity F inferred from the measurement of Victor [12]. Open and filled squares
in the figure are experimental estimates of the fidelity F e xp. The open squares represent the
Fe x p for the case with the slight imbalance of amount of squeezing in the two-mode squeezed
vacuum. The filled squares represent the F ex p for the case of the balanced amount of squeezing.
The solid lines represent the theoretical predictions Ftheory.
From the Figure 9.9, we see that the fidelity Fexp for the case with EPR
beams exceeds the classical limit Fo = 0.50 for 9 = 1 (OdE), with the
maximum value F exp = 0.58 ± 0.02 obtained. Fexp > Fo is an unambiguous
demonstration of the quantum character of the teleportation protocol.
3. SUMMARY
The fidelity F exp = 0.58±0.02 has been obtained in an experiment with con-
tinuous variable quantum teleportation. This value exceeds the classical limit
for the fidelity for the teleportation of coherent states of the electromagnetic
field. As discussed in more detail in Ref. [8, 9], this is the first demonstration
92 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
of "bona fide" quantum teleportation for which every state entering Alice's
sending station is actually teleported to emerge from Bob's receiving station
with a fidelity exceeding that which would be possible if Alice and Bob shared
only a classical communication channel.
Acknowledgments
The authors express their gratitude to S. L. Braunstein, J. L. Sorensen, C. A
Fuchs, E. S. Polzik, and S. J. van Enk. This work was supported by the NSF,
by the ONR, and by DARPA via the QUIC Institute administered by the ARO.
AF acknowledges T. Ide for preparing the figures.
References
[1] Bennett, C. H., et al.,(1993) Teleporting an unknown quantum state via
dual classical and Einstein-Podolsky-Rosen Channels, Phys. Rev. Lett. 70,
1895.
[2] Arthurs, E. and Kelly Jr., J. L., (1965) On the simultaneous measurement
of a pair of conjugate variables, Bell. Syst. Tech. J. 44, 725.
[3] Einstein, A, Podolsky, B., and Rosen, N.,(1935) Can quantum-
mechanical description of physical reality be considered complete?, Phys.
Rev. 47, 777.
[4] Boschi, D., et al.,(1998) Experimental realization of teleporting an un-
known pure quantum state via dual classical and Einstein-Podolsky-Rosen
channels, Phys. Rev. Lett. 80, 1121.
[5] Bouwmeester, D., et al.,(1997) Experimental quantum teleportation, Na-
ture 390, 575.
[6] Braunstein, S. L., and Kimble, H. J., (1998) A posteriori teleportation,
Nature 394, 840.
[7] Kok, P., and Braunstein, S. L., (2000) Postselected versus nonpostselected
quantum teleportation using parametric down-conversion, Phys. Rev. A 61,
042304.
[8] Braunstein, S. L., Fuchs, C. A, and Kimble, H. J., (2000) Criteria for
continuous-variable quantum teleportation, J. Mod. Opt. 47, 267.
[9] Braunstein, S. L., Fuchs, C. A, Kimble, H. J., and van Loock, P., (2001)
Quantum versus classical domains for teleportation with continuous vari-
ables, Phys. Rev. A 64, 022321.
[10] Vaidman, L., (1994) Teleportation of quantum states, Phys. Rev. A 49,
1473.
[11] Braunstein, S. L., and Kimble, H. J., (1998) Teleportation of continuous
quantum variables, Phys. Rev. Lett. 80, 869.
Experimental realization of continuous variable teleportation 93
Samuel L. Braunstein
Informatics, Bangor University, Bangor LL57 1 UT, United Kingdom
[email protected]
H. J. Kimble
Norman Bridge Laboratory of Physics 12-33
California Institute of Technology, Pasadena, CA 91125
Abstract A scheme to achieve dense quantum coding for the quadrature amplitudes of the
electromagnetic field is presented. The protocol utilizes shared entanglement
provided by nondegenerate parametric down-conversion in the limit of large
gain to attain high efficiency. For a constraint in the mean number of photons
fj, associated with modulation in the signal channel, the channel capacity for
dense coding is found to be In(1 + fj, + fj,2), which always beats coherent-state
communication and surpasses squeezed-state communication for fi > 1. For
fi ~ 1, the dense coding capacity approaches twice that of either scheme.
Stated more explicitly, if a classical signal a taken from the ensemble Po.
is to be transmitted as a quantum state Po., then Holevo's bound for a bosonic
quantum channel says that the mutual information H (A: B) between the sender
A (Alice) and receiver B (Bob) is bounded by [1]
(10.1)
where S(p) is the von Neumann entropy associated with the density operator
P = f d2 a Po. Po. for the mean channel state.
By contrast, if Alice and Bob share a quantum resource in the form if
an ensemble of entangled states, then quantum mechanics enables protocols
for communication that can circumvent the aforementioned bound on channel
capacity. For example, as shown originally by Bennett and Wiesner [3], Alice
and Bob can beat the Holevo limit by exploiting their shared entanglement to
achieve dense quantum coding. Here, the signal is encoded at Alice's sending
station and transmitted via one component of a pair of entangled quantum states,
with then the second component of the entangled pair exploited for decoding
the signal at Bob's receiving station. In this scheme, the cost of distributing
the entangled states to Alice and Bob is not figured into the accounting of
constraints on the quantum channel (e.g., the mean energy). Such neglect
of the distribution cost of entanglement is sensible in some situations, as for
example, if the entanglement were to be sent during off-peak times when the
communication channel is otherwise under utilized, or if it had been conveyed
by other means to Alice and Bob in advance (e.g., via a pair of quantum
CDs with stored, entangled quantum states). Note that in general, no signal
modulation is applied to the second (i.e., Bob's) component of the entangled
state, so that it carries no information by itself.
Although quantum dense coding has most often been discussed within the
setting of discrete quantum variables (e.g., qubits) [3,4], in this paper we show
that highly efficient dense coding is possible for continuous quantum variables.
As in our prior work on quantum teleportation [5, 6, 7], our scheme for achieving
quantum dense coding exploits squeezed-state entanglement, and therefore
should allow unconditional signal transmission with high efficiency, in contrast
to the conditional transmission with extremely low efficiency achieved in Ref.
[4]. More specifically, for signal states a associated with the complex amplitude
of the electromagnetic field, the channel capacity for dense coding is found to
be In(1 + n + n 2 ), where n is the mean photon number for modulation in
the signal channel. The channel capacity for dense coding in our scheme
thus always beats coherent-state communication and surpasses squeezed-state
communication for n > 1. For n » 1, the dense coding capacity approaches
twice that of either scheme.
Dense coding for continuous variables 97
Alice
Quantum Channel
Entanglement Distribution
Figure 10.1 Illustration of the scheme for achieving super-dense quantum coding for signal
states over the complex amplitude ex = x + ip of the electromagnetic field. The quantum
resource that enables dense coding is the EPR source that generates entangled beams (1,2)
shared by Alice and Bob.
As illustrated in Fig. 10.1, the relevant continuous variables for our protocol
are the quadrature amplitudes (x,]3) of the electromagnetic field, with the
classical signal a = (x) +i(p) then associated with the quantum state Pex drawn
from the phase space for a single mode of the field. The entangled resource
shared by Alice and Bob is a pair of EPR beams with quantum correlations
between canonically conjugate variables (X,]3)(1,2) as were first described by
Einstein, Podolsky, and Rosen (EPR [8]), and which can be efficiently generated
via the nonlinear optical process of parametric down conversion, resulting in
a highly squeezed two-mode state of the electromagnetic field [9, 10]. In the
ideal case, the correlations between quadrature-phase amplitudes for the two
beams (1, 2) are such that
(10.2)
Beam /12
(a) Ppz (b) a
EPRBeam2
Figure 10.2 Depiction of signal decoding at Bob's receiving station. (a) At Bob's 50 - 50
beam splitter mb, the displaced EPR beam 1 is combined with the component 2 to yield
two independent squeezed beams, with the /31,2 beams having fluctuations reduced below the
vacuum-state limit along (Xf3llPf3J. Homodyne detection at (d""d p ) (Fig. 10.1) with LO
phases set to measure (Xf31' Pf3J, respectively, then yields the complex signal amplitude Gout
with variance set by the associated squeezed states. (b) The net effect of the dense coding
protocol is the transmission and detection of states of complex amplitUde G with an effective
uncertainty below the vacuum-state limit (indicated by the dashed circle).
Dense coding for continuous variables 99
Consider the specific case of EPR beams (1,2) approximated by the two-
mode squeezed state with Wigner function
where the subscripts R and I refer to real and imaginary parts of the field
amplitude a, respectively (i.e., aR,I = x,p). Note that for r -t 00, the field
state becomes the ideal EPR state as described in Eq. (10.2), namely,
(10.4)
where (3 = (31R + i(32J and represents a highly peaked distribution about the
complex displacement a/ V2. For large squeezing parameter r this allows us
to extract the original signal a which we choose to be distributed as
1
Pa = -2
7ra-
exp( -Ial
2
/a- 2 ) . (10.6)
100 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Note that mode 1 of this displaced state has a mean number of photons given
by
(10.7)
In order to compute the quantity of information that may be sent through this
dense coding channel we note the unconditioned probability for the homodyne
statistics is given by
p _ 2 ex ( -21,61 2 ) (10.8)
(,6) - 7r((]"2 + e-2r) p (]"2 + e-2r .
(10.9)
(10.10)
C rv 2r , (10.13)
for large squeezing r. This is just one-half of the asymptotic dense coding
mutual information, see Eq. (10.11). Thus asymptotically, at least, the dense
coding scheme allows twice as much information to be encoded within a given
Dense coding for continuous variables 101
state, although it has an extra expense (not included within the simple constraint
n) of requiring shared entanglement.
It is worth noting that this dense coding scheme does not always beat the
optimal single channel capacity. Indeed, for small squeezing it is worse. The
break-even squeezing required for dense coding to equal the capacity of the
optimal single channel communication is
which is beaten by the dense coding scheme ofEq. (10.10) for n > 1, i.e., the
break-even squeezing required is
References
[1] A. S. Holevo, IEEE Trans. Info. Theory 44,269 (1998).
[2] B. Schumacher and M. D. Westmoreland, Phys. Rev. A 56, 131 (1997).
[3] c. H. Bennett and S. J. Wiesner, Phys. Rev. Lett. 69, 2881 (1992).
[4] K. MattIe, H. Weinfurter, P. G. Kwiat and A. Zeilinger, Phys. Rev. Lett.
76,4656 (1996).
[5] S. L. Braunstein and H. J. Kimble, Phys. Rev. Lett. 80, 869 (1998).
[6] P. van Loock, S. L. Braunstein, and H. J. Kimble, "Broadband teleporta-
tion," LANL preprint, quant-phl9902030.
[7] A. Furusawa, J. Sl<1rensen, S. L. Braunstein, C. Fuchs, H. J. Kimble, and
E. S. Polzik, Science 282, 706 (1998).
[8] A. Einstein, B. Podolsky, N. Rosen, Phys. Rev. 47, 777 (1935).
[9] M. D. Reid and P. D. Drummond, Phys. Rev. Lett. 60,2731 (1988); M.
D. Reid, Phys. Rev. A 40,913 (1989).
[10] Z. Y. Ou, S. F. Pereira, H. J. Kimble, and K. C. Peng, Phys. Rev. Lett. 68,
3663 (1992); Appl. Phys. B: Photophys. Laser Chern. 55, 265 (1992).
[11] L. Mandel and E. Wolf, Optical Coherence and Quantum Optics, (Cam-
bridge University Press, Cambridge, England, 1995).
Dense coding for continuous variables 103
[12] C. M. Caves and P. D. Drummond, Rev. Mod. Phys. 66, 481 (1994).
[13] H. P. Yuen and M. Ozawa, Phys. Rev. Lett. 70, 363 (1993).
[14] J. P. Gordon, Proc. IRE 50, 1898 (1962).
[15] C. Y. She, IEEE Trans. Inf. Theory IT·14, 32 (1968).
[16] Y. Yamamoto and H. A. Haus, Rev. Mod. Phys. 58, 1001 (1986).
[17] H. Weinfurter (private communication).
[18] S. Lloyd and S. L. Braunstein, Phys. Rev. Lett. 82, 1784 (1999).
[19] S. Lloyd and J. J.-E. Slotine, Phys. Rev. Lett. 80, 4088 (1998).
[20] S. L. Braunstein, Phys. Rev. Lett. 80, 4084 (1998).
[21] S. L. Braunstein, Nature (London) 394, 47 (1998).
[22] C. H. Bennett et al., Phys. Rev. Lett. 70, 1895 (1993).
[23] L. Vaidman, Phys. Rev. A49, 1473 (1994).
[24] A. S. Parkins and H. J. Kimble, e-print quant-phl9904062.
[25] A. S. Parkins and H. J. Kimble, e-print quant-phl9907049.
Chapter 11
MULTIPARTITE
GREENBERGER-HORNE-ZEILINGER
PARADOXES FOR CONTINUOUS VARIABLES
- x d - pL
x = y'7fL an p = y'7f , (11.1)
which follows from [x,jJ] = i/7r and the identity eAe B = e[A,BleBe A (valid
if A and B commute with their commutator). The continuous variable GHZ
paradoxes will be built out of these operators.
Let us first consider the case of three spatially separated parties, A, B,
C, each of which possess one part of an entangled system described by the
canonical variables XA,PA, XB,PB, xc and pc. Consider the operators X[7r
and Yj±7r acting on the space of party j (j = A, B, C). Since a{3 = ±7r 2 ,
it follows from (11.3), that these operators obey the commutations relations
X[7rYj±7r = - Yj±7r X[7r. Using these operators let us construct the following
four GHZ operators:
These four operators give rise to a GHZ paradox as we now show. First note
that the following two properties hold:
1. VI, 112, V3, V4 all commute. Thus they can be simultaneously diagonal-
ized (in fact there exists a complete set of common eigenvectors).
These properties are easily proven using the commutations relations Xf7r 1j±7r =
_1j±7r Xf7r. Any common eigenstate of VI, V2, V3, V4 will give rise to a GHZ
paradox. Indeed suppose that the parties measure the hermitian operators x j
or Pj, j = A, B, C on this common eigenstate. The result of the measurement
associates a complex number of unit norm to either the Xj or Yj unitary op-
erators. If one of the combinations of operators that occurs in eq. (11.4) is
measured, a value can be assigned to one of the operators Vi, V2, V3, V4. Quan-
tum mechanics imposes that this value is equal to the corresponding eigenvalue.
Moreover - due to property 2 - the product of the eigenvalues is -1.
But this is in contradiction with local hidden variables theories. Indeed in
a local hidden theory one must assign, prior to the measurement, a complex
number of unit norm to all the operators Xj and Yj. Then taking the product of
the four c-numbers assigned simultaneously to VI, V2, V3, V4 yields +1 instead
of -l.
Remark that all other tests of non-locality for continuous variable systems [1,
2, 3, 4, 7] use measurements with a discrete spectrum (such as the parity photon
number) or involving only a discrete set of outcome (such as the probability that
x > 0 or x < 0). In our version of the GHZ paradox for continuous variables
this discrete character doesn't seem to appear at first sight. However it turn out
that it is also the case thought in a subtle way because eq. (11.4) can be viewed
as an infinite set of 2 dimensional paradoxes (see [6] for more details).
In [10], GHZ paradoxes for many parties and multidimensional systems
where constructed. These paradoxes where build using d-dimensional unitary
operators with commutation relations:
(11.5)
the operators x±q, yq and y-3q where q = 7r / J2. They obey the commu-
tation relation x±qyq = e±i7r /2yq x±q and X±qy-3q = e±i7r /2y-3q x±q.
Consider now the six unitary operators
WI X1 Xq
W2 X A YB~q
q
W3 Yl Xi/ (11.6)
W4 Yl y~
W5 Yl y~
W,6 y-3q yq
A . B
One easily shows that these six unitary operators commute and that their product
is minus the identity operator. Furthermore if one assigns a classical value to
x j and to Pj for j = A, B, C, D, E, then the product of the operators takes the
value + 1. Hence, using the same argument as in the three party case, we have
a contradiction.
There is a slight difference between the paradox (11.6) and the 4-dimensional
paradox described in [10]. The origin of this difference is that in a d-
dimensional Hilbert space, if unitary operators X, Y obey XY = e i7r / dy X,
then X d = yd = I (up to a phase which can be set to 1), or equivalently,
X d- I = xt and yd-I = yt. In the continuous case these relations no longer
hold and the GHZ operators Wi'S must be slightly modified, i.e. the operator
X- q = x qt and y-3q = y3qt have to be explicitly introduced in order for
the product of the Wi'S to give minus the identity. Note that the same remark
applies for the previous paradox (11.4) where in the descrete 2-dimensional
version xt = X and yt = y.
As we mentioned earlier the GHZ states are not Gaussian states. A detailed
analysis of the common eigenstates of VI, V2 , V3 , V4 is given in [6]. Let us give
an example of such an eigenstate. Define the following coherent superpositions
of infinitely squeezed states:
1
L
00 _
L
00
1
I-!-) J2 (Ii = 2k) - iii = 2k + 1)) , (11.7)
k=-oo
Acknowledgments
We would like to thank N. Cerf for helpful discussions. We acknowledges
funding by the European Union under project EQUIP (lST-FET program).
S.M. is a research associate of the Belgian National Research Foundation.
References
[1] K. Banaszek and K. W6dkiewicz, Phys. Rev. A 58,4345 (1998).
[2] A. Kuzmich, I. A. Walmsley and L. Mandel, Phys. Rev. Lett. 85, 1349
(2000).
[3] Z. Chen, 1. Pan, G. Hou and Y. Zhang, Phys. Rev. Lett. 88040406 (2002).
[4] B. Yurke, M. Hillery and D. Stoler, Phys. Rev. A 60,3444 (1999).
[5] R. Clifton, Phys. Lett. A 271, 1 (2000).
[6] S. Massar and S. Pironio, Phys. Rev. A 64 062108 (2001).
[7] Z. Chen and Y. Zhang, Phys. Rev. A 65 044102 (2001).
[8] D. M. Greenberger, M. Home, A. Zeilinger, in Bell's Theorem, Quan-
tum Theory, and Conceptions of the Universe, M. Kafatos, ed., Kluwer,
Dordrecht, The Netherlands (1989), p. 69.
[9] N. D. Mermin, Phys. Rev. Lett. 65, 3373 (1990) and Phys. Today, 43(6),
9 (1990).
[10] N. Cerf, S. Massar, S. Pironio, Greenberger-Home-Zeilinger paradoxes
for many qudits, quant-phlOl07031, to be published in Phys. Rev. Lett.
[11] D. Gottesman, A. Kitaev and J. Preskill, Phys. Rev. A 64 012310 (2001).
Chapter 12
MULTIPARTITE ENTANGLEMENT
FOR CONTINUOUS VARIABLES
Samuel L. Braunstein
lnfonnatics, Bangor University, Bangor LL57 1Ur, United Kingdom
[email protected]
Abstract First, we show how the quantum circuits for generating and measuring mUlti-party
entanglement of qubits can be translated to continuous quantum variables. We
derive sufficient inseparability criteria for N -party continuous-variable states and
discuss their applicability. Then, we consider a family of multipartite entangled
states (multi-party multi-mode states with one mode per party) described by
continuous quantum variables and analyze their properties. These states can be
efficiently generated using squeezed light and linear optics.
1. INTRODUCTION
What is the main motivation to deal with continuous variables for quantum
communication purposes? Quantum communication schemes rely on state
preparation, local unitary transformations, measurements, and classical com-
munication. In addition, sometimes shared entanglement is part of the protocol.
Within the framework of quantum optics, these ingredients can be efficiently
implemented when they are applied to the continuous quadrature amplitudes
of electromagnetic modes. For example, the tools for measuring a quadrature
with near-unit efficiency or for displacing an optical mode in phase space are
provided by homodyne detection and feed-forward techniques, respectively.
Continuous-variable entanglement can be efficiently produced using squeezed
111
S.L. Braunstein and A.K. Pati (eds.). Quantum Information with Continuous Variables, 111-143.
© 2003 Kluwer Academic Publishers.
112 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
light and linear optics. In this chapter, we consider a rather general manifes-
tation of continuous-variable entanglement, namely that between an arbitrary
number of parties (modes). We will see that even those N-party entangled
states where none of the N parties can be separated from the others in the
total state vector are comparatively "cheap" in tenns of the resources needed:
their generation only requires one single-mode squeezed state and N - 1 beam
splitters.
2. MULTIPARTITE ENTANGLEMENT
The main subject of this section is mUlti-party entanglement of infinite-
dimensional states described by continuous variables. After a few general
remarks on entanglement between two and more parties in arbitrary dimen-
sions, we will show how the quantum circuits for creating and measuring qubit
entanglement may be translated to continuous variables. Then we derive in-
equalities that may serve as sufficient multi-party inseparability criteria for
continuous-variable states. These are applicable both for a theoretical test and
for an indirect experimental verification of multi-party entanglement. Finally,
we focus on a family of genuinely multi-party entangled continuous-variable
states whose members are fully inseparable with respect to all their parties.
(12.1)
n
where the summation goes over the smaller of the dimensionalities of the two
subsystems. The Schmidt coefficients Cn are real and non-negative, and satisfy
I:n c~ = 1. The Schmidt decomposition may be obtained by transforming the
expansion of an arbitrary pure bipartite state as
mk nmk n
with Cnn == Cn. In the first step, the matrix a with complex elements amk
is diagonalised, a = ucv T , where u and v are unitary matrices and c is a
diagonal matrix with non-negative elements. In the second step, we defined
lun ) == I: mumnlm) and Ivn) == I:k vknl k ) which form orthonormal sets due
to the unitarity of u and v and the orthonormality of 1m) and Ik). A pure state of
two d-Ievel systems ("qudits") is now maximally entangled when the Schmidt
Multipartite entanglement/or continuous variables 113
coefficients of its total state vector are all equal. Since the eigenvalues of the
reduced density operator upon tracing out one half of a bipartite state are the
Schmidt coefficients squared,
tracing out either qudit of a maximally entangled state leaves the other half
in the maximally mixed state n/ d. A pure two-party state is factorizable (not
entangled) if and only if the number of nonzero Schmidt coefficients is one. Any
Schmidt number greater than one indicates entanglement. Thus, the "majority"
of pure state vectors in the Hilbert space of two parties are nonmaximally
entangled. Furthermore, any pure two-party state is entangled if and only if
for suitably chosen observables, it yields a violation of inequalities imposed by
local realistic theories [2]. A unique measure of bipartite entanglement for pure
states is given by the partial von Neumann entropy, the von Neumann entropy
(-Tr,olog,o) of the remaining system after tracing out either subsystem [3]:
Ev.N. = -Tr,ollogd,ol = - L:n c~ logd c~, ranging between zero and one (in
units of "edits").
Mixed states are more subtle, even for only two parties. As for the quan-
tification of bipartite mixed-state entanglement, there are various measures
available such as the entanglement of formation and distillation [4]. Only for
pure states, these measures coincide and equal the partial von Neumann en-
tropy. The definition of pure-state entanglement via the non-factorizability of
the total state vector is generalized to mixed states through non-separability
(or inseparability) of the total density operator. A general quantum state of a
two-party system is separable if its total density operator is a mixture (a convex
sum) of product states [5],
(12.4)
two-mode Gaussian states, see below), this condition is both necessary and suf-
ficient. For any other dimension, negative partial transpose is only sufficient for
inseparability [7] 2. Other sufficient inseparability criteria include violations
of inequalities imposed by local realistic theories (though mixed inseparable
states do not necessarily lead to such violations), an entropic inequality [namely
Ev.N. (PI) > Ev.N. (h2), again with PI = Tr 2P12] [11], and a condition based
on the theory of majorization [12]. Concluding the discussion of two-party
entanglement, we emphasize that both the pure-state Schmidt decomposition
and the partial transpose criterion for mixed states are also applicable to infinite
dimensions. An example for the infinite-dimensional Schmidt decomposition
is the two-mode squeezed vacuum state in the Fock (photon number) basis
[13]. The unphysical operation (a positive, but not completely positive map)
that corresponds to the transposition is time reversal [14]: in terms of contin-
uous variables, any separable two-party state remains a legitimate state after
the transformation (XI,PI, X2,P2) -+ (Xl, -PI, X2,P2), where (Xi,Pi) are the
phase-space variables (positions and momenta) for example in the Wigner
representation. However, arbitrary inseparable states may be turned into un-
physical states, and furthermore, inseparable two-party two-mode Gaussian
states always become unphysical via this transformation [14].
Multipartite entanglement, the entanglement shared by more than two
parties, is a more complex issue. For pure multi-party states, a Schmidt
decomposition does not exist in general. The total state vector then cannot be
written as a single sum over orthonormal basis states. There is, however, one
very important representative of multipartite entanglement which does have the
form of a mUlti-party Schmidt decomposition, namely the Greenberger-Horne-
Zeilinger (GHZ) state [15]
For the case of three qubits, any pure and fully entangled state can be
transformed to either the GHZ state or the so-called W state [17],
is inseparable which can be verified by taking the partial transpose [the eigen-
values are 1/3, 1/3, (1 ± v'5)/6]. This is in contrast to the GHZ state where
tracing out one party yields the separable two-qubit state
Here, I<p±) are two of the four Bell states, I<p±) = (100) ± 111) )/J2, IW±) =
(101) ± 11O))/J2·
What can be said about arbitrary mixed entangled states of more than two
parties? There is of course an immense variety of inequivalent classes of multi-
party mixed states [e.g., five classes of three-qubit states of which the extreme
cases are the fully separable (p = :Ei Pi Pil ®Pi2®Pi3) and the fully (genuinely)
116 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
1 1
10) ---+ v'2 (10) + 11)) , 11) ---+ v'2(IO) -11)) , (12.11)
100) ---+ 100) , 101) ---+ 101) , 110) ---+ 111) , 111) ---+ 110) .(12.12)
The first qubit (control qubit) remains unchanged under the C-NOT. The second
qubit (target qubit) is flipped if the control qubit is set to 1, and is left unchanged
otherwise. Equivalently, we can describe the action of the C-NOT gate by
IYl, Y2) -+ IYl, Yl EEl Y2) with Yl, Y2 = 0,1 and the addition modulo two EEl.
The N -partite entangled output state of the circuit (see Fig. 12.1) is the N -qubit
GHZ state.
Let us translate the qubit quantum circuit to continuous variables [19].
The position and momentum variables x and p (units-free with n = 1,
[Xl,Pk] = idlk/2) may correspond to the quadrature amplitudes of a single
electromagnetic mode, i.e., the real and imaginary part of the single mode's
annihilation operator: a = x + ip. At this stage, it is convenient to consider
position and momentum eigenstates. We may now replace the Hadamard by a
Multipartite entanglementfor continuous variables 117
10)
C-NOT
10)
10)
I
I
10) ED
Figure 12.1 Quantum circuit for generating the N-qubit GHZ state. The gates (unitary trans-
formations) are a Hadamard gate ("H") and pairwise acting C-Naf gates.
Fourier transform,
Flx)position = ~
y 7r
1 00
-00
dy e2ixYIY)position = Ip = X)momentum , (12.13)
and the C-NOT gates by appropriate beam splitter operations 3. The input
states are taken to be zero-position eigenstates Ix = 0). The sequence of beam
splitter operations Bjk (()) is provided by a network of ideal phase-free beam
splitters (with typically asymmetric transmittance and reflectivity) acting on
the position eigenstates as
I
to a zero-momentum eigenstate Ip = 0) oc dx Ix) of mode 1 (the Fourier
transformed zero-position eigenstate) and N - 1 zero-position eigenstates
Ix = 0) in modes 2 through N. We obtain the entangled N-mode state
I dx lx, x, . .. ,x). This state is an eigenstate with total momentum zero and
all relative positions Xi - Xj = 0 (i, j = 1,2, ... ,N). It is clearly an analogue
to the qubit GHZ state with perfect correlations among the quadratures. How-
ever, it is an unphysical and unnormalizable state (e.g., for two modes, it cor-
responds to the maximally entangled, infinitely squeezed two-mode squeezed
vacuum state with infinite energy). Rather than sending infinitely squeezed
position eigenstates through the entanglement-generating circuit, we will now
use finitely squeezed states.
118 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
sin ()
( C'k ) -_ ( cos cos () ) ( Ck ) (12.16)
c
1 (} - sin (} Cl .
(12.18)
with i = 2,3, ... ,N and vacuum modes labeled by the superscript '(0)'. In
terms of the input quadratures, we have
(12.20)
for o'j = Xj + ipj (j = 1,2, ... , N). The squeezing parameters rl and r2
determine the degree of squeezing of the momentum-squeezed and the N - 1
position-squeezed modes, respectively. The correlations between the output
quadratures are revealed by the arbitrarily small noise in the relative positions
and the total momentum for sufficiently large squeezing rl and r2,
e- 2T2 /2 ,
Ne- 2'fl /4, (12.21)
for k i- l (k, l = 1,2, ... , N) and O,~ = x~ + ip~. Note that all modes involved
have zero mean values, thus the variances and the second moments are identical.
Multipartite entanglement for continuous variables 119
where n, ml, m2, ... , mN-I = 0,1. The projection onto the basis states
{I wn,m1 ,m2 , ... ,mN -I)} is accomplished when the output states of the inverted
circuit (see Fig. 12.1),
are measured in the computational basis. Eventually, {IW n ,m1,m2, ... ,mN_J}
are distinguished via the measured output states
..;:;r 1
00
-1 dx e2ivX lx) ®Ix - U1)
-00
Since J~oo Ix) (xl = n and (xix') = J(x - x'), they form a complete,
= n®N, (12.26)
and orthogonal,
(w( v, U1, U2, ... , UN -1) Iw( v', U~, U;, ... , U~ -1))
= J(v - V')J(U1 - U~)J(U2 - U;)··· J(UN-1 - U~_l)' (12.27)
(12.28)
1 (A A A)
AI
PI - V3 PI + P2 + P3 ,
AI
X2 V(L 1 (A A)
3"Xl - v'6 X2 + X3 ,
1 (A A)
AI
X3 V2 X2 - X3 , (12.29)
Eq. (12.25). More precisely, the fidelity ofthe state discrimination can be arbi-
trarily high for sufficiently good accuracy of the homodyne detectors. We may
conclude that the requirements of such a "GHZ state analyzer" for continuous
variables are easily met by current experimental capabilities. This is in contrast
to the GHZ state analyzer for photonic qubits [capable of discriminating or
measuring states like those in Eq. (12.22)]. Although arbitrarily high fidelity
can be approached in principle using linear optics and photon number detec-
tors, one would need sufficiently many, highly entangled auxiliary photons and
detectors resolving correspondingly large photon numbers [21, 20]. Neither of
these requirements is met by current technology. Of course, the C-NOT gates
of a qubit GHZ state measurement device can in principle be implemented
via the so-called cross Kerr effect using nonlinear optics. However, on the
single-photon level, this would require optical nonlinearities of exotic strength.
In this section, we have shown how measurements onto the maximally
entangled continuous-variable GHZ basis can be realized using linear optics and
quadrature detections. These schemes are an extension of the well-known two-
party case, where the continuous-variable Bell basis [Eq. (12.25) with N = 2] is
the analogue to the qubit Bell states [Eq. (12.22) with N = 2]. The continuous-
variable and the qubit Bell states form those measurement bases that were used
in the quantum teleportation experiments [22] and [23, 24], respectively. The
extension of measurements onto the maximally entangled basis to more than
two parties and their potential optical realization in the continuous-variable
realm might be relevant to mUlti-party quantum communication protocols such
as the multi-party generalization of entanglement swapping [25]. However, the
entanglement resources in a continuous-variable protocol, namely the entangled
continuous-variable states that are producible with squeezed light and beam
splitters, exhibit only imperfect entanglement due to the finite degree of the
squeezing. When can we actually be sure that they are multi-party entangled at
all? In the next section, we will address this question and discuss criteria for the
theoretical and the experimental verification of mUltipartite continuous-variable
entanglement.
can both be accomplished by equal means, namely a single beam splitter and
two homodyne detectors. In other words, the effectively inverse circuit for
the generation of bipartite entanglement provides the recipe for both measur-
ing maximum entanglement and verifying nonmaximum entanglement. When
looking for multi-party inseparability criteria for arbitrarily many modes, it
seems to be natural to pursue a similar strategy. We are therefore aiming at a
criterion which is based on the variances of those quadrature combinations that
are the measured observables in a continuous-variable GHZ measurement.
Let us consider three modes. According to Eq. (12.29), we define the
operators
(12.31)
(12.32)
Using this state, we can calculate the total variance of the operators in Eq. (12.31),
where (... )i means the average in the product state Pil 0 Pi2 0 Pi3. Similar
to the derivation in Ref. [26], we can apply the Cauchy-Schwarz inequality
2:i Pi(u); 2: (2:i Pi l(u)iI)2, and see that the last two lines in Eq. (12.33) are
bounded below by zero. Also taking into account the sum uncertainty relation
((b.Xj)2)i + ((b.pj)2)i 2: I[xj,pjll = 1/2 (j = 1,2,3), we find that the total
variance itself is bounded below by 1 (using 2:i Pi = 1). Any total variance
smaller than this boundary of 1 would imply that the quantum state concerned
is not fully separable as in Eq. (12.32). But would this also imply that the
quantum state is genuinely tripartite entangled in the sense that none of the
parties can be separated from the others (as, for example, in the pure qubit
states IGHZ) and IW)? This is obviously not the case and a total variance
below 1 does not rule out the possibility of partial separability. The quantum
state might still not be a genuine tripartite entangled state, since it might be
written in one or more of the following forms 4 [18]:
(12.35)
hI, h2' ... , hN which isfully separable, P = I:i Pi Pil ® Pi2 ® ... ® PiN, obeys
the inequality
(12.36)
(12.37)
(12.38)
for any N and i, j. Similarly, the output quadratures of the inverse N -splitter
yield, for instance, for N = 3,
Both criteria in Eq. (12.36) and Eq. (12.37) represent necessary conditions
for full separability, though they are not entirely equivalent [i.e., there are
partially inseparable states that violate Eq. (12.37), but satisfy Eq. (12.36),
see below]. Moreover, the criterion in Eq. (12.37) contains in some sense
redundant observables. As we know from the previous section, N observables
suffice to measure an N -party GHZ entangled state. These N observables are
suitably chosen quadratures of the N output modes of an inverse N -splitter.
Their detection simultaneously determines the total momentum and the N - 1
relative positions Xl -X2, X2 -X3, ... , and XN-I -XN [Eq. (12.30)] 5. However,
in Eq. (12.37), there are 1 + [N(N -1)]/2 different operators. Nevertheless, for
two parties and modes, the conditions in Eq. (12.36) and Eq. (12.37) coincide
and correspond to the necessary separability condition for arbitrary bipartite
states given in Ref. [26].
In summary, we have shown in this section that the circuit for measur-
ing GHZ entanglement also provides a sufficient inseparability criterion for
Multipartite entanglementfor continuous variables 125
_1 (1_e+ 2r + e- 2r ) + _
1 > _
1 (12.41)
4 3 6-2'
as a necessary condition for full separability. We find that equality holds for
r = 0 which doesn't tell us anything, though we know, of course, that the
state is fully separable in this case and must obey Eq. (12.36). For some finite
squeezing r, the inequality Eq. (12.41) is illustrated in Fig. 12.2. Similarly,
application of the criterion in Eq. (12.37) to the above state leads to
1 3
"4 (3e- 2r + cosh2r + 2) ~ "2 . (12.42)
Again, equality holds for r = O. In Fig. 12.2, also this condition is depicted
for some finite squeezing r.
126 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
1.6~--------------~--~
1.5~---------------,~
0.55
1.4
0.5~---------T--------~
1.3
0.45
1.2
Figure 12.2 Application of the necessary conditions for full three-party separability. On the
left: the inequality Eq. (12.42) with the boundary 3/2 as a function of the squeezing r; for
o < r < 1, the inequality is violated nearly everywhere. On the right: the inequality Eq. (12.41)
with the boundary 1/2 as a function of the squeezing r; for 0 < r < 1, the inequality is satisfied
nearly as much as it is violated.
N -splitter means
Ut(N) (a~ a2 ~I
aN
)T (12.43)
U t (N) U(N) (al a2 ~
aN )T ,
with aI, a2, ... , aN fromEq. (12.19). SinceUt(N)U(N) = I (identity matrix),
the squeezed quadratures ofEq. (12.20) can be directly inserted into Eq. (12.36)
yielding a violation of (e- 2r1 + e- 2r2 )/4 ~ 1/2 for any Tl > 0 or T2 > o.
Alternatively, using Eq. (12.21), the criterion in Eq. (12.37) becomes
This condition is also violated for any Tl > 0 or T2 > O. Due to their purity
and total symmetry we conclude that the members of the family of states which
emerge from the N -splitter circuit are genuinely mUlti-party entangled for any
Tl > 0 or T2 > O. This applies in particular to the case where Tl > 0 and
T2 = 0, i.e., when only one squeezed light mode is required for the creation of
genuine multipartite entanglement.
Independent of the inequalities Eq. (12.36) and Eq. (12.37), there are also
other ways to see that these particular states are genuinely multi-party entangled.
One simply has to find some form of entanglement in these states. For example,
by tracing out modes 2 through N of the pure N -mode state given in Eq. (12.18),
one finds that the remaining one-mode state is mixed, provided Tl > 0 or T2 > 0
[28]. Thus, the pure N -mode state is somehow entangled and hence genuinely
mUlti-party entangled due to its complete symmetry. Note that in order to infer
even only partial entanglement via tracing out parties, the N -mode state here
has to be pure. In contrast, the mUlti-party inseparability criteria of section 2.4
may verify partial inseparability for any N -mode state.
From a conceptual point of view, it is very illuminating to analyze which
states of the above family of N -mode states can be transformed into each other
via local squeezing operations [29]. For example, by applying local squeezers
with squeezing 81 and 82 to the two modes of the bipartite state generated with
only one squeezer [Eq. (12.20) for N = 2 with T2 = 0], we obtain
(12.45)
With the choice of 81 = 82 = TI/2 == T, the state in Eq. (12.45) is identical to
a two-mode squeezed state built from two equally squeezed states [Eq. (12.20)
128 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
for N = 2 with rl == r, r2 == r]. The latter and the state produced with only
one squeezer [Eq. (12.20) with rl = 2r and r2 = 0] are equivalent under local
squeezing operations. This means that Alice and Bob sharing the state produced
with one squeezer rl = 2r have access to the same amount of entanglement as
in the "canonical" two-mode squeezed state with squeezing r = rI/2, Ev.N. =
[cosh(rI/2)j2 log[cosh(rI/2)j2 - [sinh(rI/2)j2 log[sinh(rI/2)j2 [30]. For a
given amount of entanglement, however, the canonical two-mode squeezed
vacuum state has the least mean photon number. Conversely, for a given mean
energy, the canonical two-mode squeezed vacuum state contains the maximum
amount of entanglement possible.
Similar arguments apply to the states of more than two modes. From the
family of N -mode states, the state with the least mean photon number is
determined by the relation
(12.47)
We see that in order to produce the minimum-energy N-mode state, the single
rl -squeezer is, in terms of the squeezing factor, N - 1 times as much squeezed
as each r2-squeezer. However, also in this general N -mode case, the other
N -mode states of the family can be converted into the minimum-energy states
via local squeezing operations. This applies in particular to the N-mode states
produced with just a single squeezer and to those built from N equally squeezed
states. As a result, due to the equivalence under local entanglement-preserving
operations, with a single sufficiently squeezed state and beam splitters, arbi-
trarily many genuinely multi-party entangled modes can be created just as well
as with N squeezers and beam splitters.
In contrast to the three-mode state given by Eq. (12.40), the output states of
the N -splitter are totally symmetric under interchange of parties. This becomes
more transparent when we look at the states in the Wigner representation. For
simplicity, let us assume r = rl = r2. The position-squeezed input states of
the N -splitter circuit, for instance, have the Wigner function
(12.48)
Multipartite entanglement/or continuous variables 129
Through the linear N -splitter operation, the total input Wigner function to the
N -splitter (one momentum-squeezed and N - 1 position-squeezed vacuum
modes),
Here we have used x = (Xl, X2, ... , XN) and p = (PI,P2, .. ·,PN). The pure-
state Wigner function Wout (x, p) is always positive, symmetric among the N
modes, and becomes peaked at Xi - Xj = 0 (i,j = 1,2, ... , N) and PI +
P2 + ... + PN = 0 for large squeezing r. For N = 2, it exactly equals the
well-known two-mode squeezed vacuum state Wigner function [13], which is
proportional to 8(XI - x2)8(PI + P2) in the limit of infinite squeezing. As
discussed previously, the state Wout (x, p) is genuinely N -partite entangled for
any squeezing r > O. The quantum nature of the cross correlations XiXj and
PiPj appearing in Wout (x, p) for any r > 0 is also confirmed by the purity of
this state. This purity is guaranteed, since beam splitters tum pure states into
pure states (it can also be checked via the correlation matrix of the Gaussian
state Wout (x, p) [28]).
A nice example for a mUlti-party entangled state which is not a member
of the above family of states and not totally symmetric with respect to all its
modes is the (M + I)-mode state described by the Wigner function
130 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(sm()OPI +
. COS ()o M+I )2
_2e+ 2T1 ..JM ~ Pi
1 . () {Jf;
VM + 1 <- sm 0 <_ . M + 1 '
- r=:=::=== (12.52)
e- 2T! == ..JM sin ()o - cos ()o , -2T2 _ ..JM cos ()o - sin ()o
..JM sin ()o + cos ()o e = ..JM cos ()o + sin ()o .
The significance of this (M + I)-mode state is that it represents a kind of
multiuser quantum channel ("MQC") enabling optimal 1 -+ M "telec1oning"
of arbitrary coherent states from one sender to M receivers [31]. Though
not completely symmetric with respect to all M + 1 modes (but to modes 2
through M + 1), it is a pure Gaussian state which is indeed genuinely multi-
party entangled. This can be seen, because none of the modes can be factored
out of the total Wigner function. Despite its "asymmetry", this state is not
only partially multi-party entangled as is the asymmetric pure three-mode state
given by Eq. (12.40). Of course, the bipartite entanglement between mode 1
on one side and modes 2 through M + 1 on the other side is the most important
property of WMQc(X, p) in order to be useful for 1 -+ M telec10ning [31].
The generation of the state WMQc(X, p) is very similar to that of the above
family of multi-party entangled states produced with an N -splitter: first make
a bipartite entangled state by combining two squeezed vacua, one squeezed in
P with Tl and the other one squeezed in X with T2, at a phase-free beam splitter
with reflectivity/transmittance parameter () = eo.
Then keep one half (the
mode 1) and send the other half together with M - 1 vacuum modes through an
Multipartite entanglement for continuous variables 131
M -splitter. The annihilation operators of the initial modes aj before the beam
splitters, j = 1,2, ... , M + 1, are then given by
with
The first beam splitter, acting on modes 0,1 and 0,2, has reflectivity/transmittance
parameter B == Bo. The remaining beam splitters represent an M -splitter. In
Eq. (12.54), the output modes bj correspond to the M + 1 modes of the
MQC state described by WMQC in Eq. (12.51). Let us now return to the
totally symmetric multipartite entangled states given by Eq. (12.18) and explore
some of their properties. For simplicity, we will thereby focus on those states
emerging from the N -splitter circuit which are created with input states equally
squeezed in momentum and position, r = rl = r2.
(12.56)
where 0 = x + ip = (aI, a2, ... , aN) and 11(0) is the quantum expectation
value of the operator
N N
fr(o) = @fri(ai) = Q9Di(ai)(-l)TtiD!(ad . (12.57)
i=l i=l
(12.59)
Di(ai) (12.60)
k=O
where o-(aN) = ±1 and o-(a',y) = ±1 describe two possible outcomes for two
possible measurement settings (denoted by aN and a',y) of measurements on
the Nth particle. Note, the expressions B~ are equivalent to BN but with all
the ai and a~ swapped. Provided that B N -1 = ±2 and B~ -1 = ±2, Equation
(12.62) is trivially true for a single run of measurements where 0-( aN) is either
+1 or -1 and similarly for o-(a',y). Induction proves Eq. (12.62) for any N
when we take
B2 == [0-(a1) + 0-(aD]0-(a2)
+[o-(ad - o-(a~)]o-(a~) = ±2 . (12.63)
(12.65)
Following Bell [2], an always positive Wigner function can serve as the hidden-
variable probability distribution with respect to measurements corresponding
to any linear combination of x and p. In this sense, the finitely squeezed two-
mode squeezed state Wigner function could prevent the CHSH inequality from
134 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
where a = (aI, a2, ... , aN). However, for parity measurements on each mode
with possible results ±1 for each differing displacement, this would require
unbounded 8-functions for the local objective quantities a(ai' Ai) [34], as in
this case we have
This relation directly relates the correlation function to the Wigner function
and is indeed crucial for the nonlocality proof of the continuous-variable states
in Eq. (12.50).
Let us begin by analyzing the nonlocal correlations exhibited by the entan-
gled two-party state. For this state, the two-mode squeezed state in Eq. (12.50)
with N = 2, we may investigate the combination [34]
which according to Eq. (12.65) satisfies 182 1:::; 2 for local realistic theories.
Here, we have chosen the displacement settings al = a2 = 0 and a~ = a,
a2'-(3
- .
t
Writing the states in Eq. (12.50) as
~IC(al' a2, a3, a~) + C(al' a2, a~, a4) + C(al' a~, a3, a4)
+C(a~,a2,a3,a4) + C(al,a2,a~,a~) + C(al,a~,a3,a~)
+C(a~,a2,a3,a~) + C(al,a~,a~,a4) + C(a~,a2,a~,a4)
+C(a~,a~,a3,a4) - C(a~,a~,a~,a4) - C(a~,a;,a3,a~)
-C(a~,a2,a~,a~) - C(al,a;,a~,a~) - C(al,a2,a3,a4)
-C(a~,a;,a~,a~)1 ~ 2. (12.75)
136 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
It is symmetric among allfour parties as any inequality derived from Eq. (12.62)-
(12.64) is symmetric among all parties. For the settings al = a2 = a3 = a4 =
° and a~ = a, a~ = {3, a~ = "(, a~ = 8, complying with local realism means
184 1~ 2 where
1
84 = 2" [II(O, 0, 0, 8) + II(O, 0, ,,(, 0) + II(O, {3, 0, 0)
+II(a, 0, 0, 0) + II(O, 0, ,,(, 8) + II(O, {3, 0, 8)
+II(a, 0, 0, 8) + II(O, {3, ,,(, 0) + II(a, 0, ,,(, 0)
+II(a, {3, 0, 0) - II(a, {3, ,,(, 0) - II(a, {3, 0, 8)
-II(a, 0,,,(, 8) - II(0,{3,,,(,8) - II(O,O,O,O)
-II(a, {3, ,,(, 8)] . (12.76)
which has to satisfy 185 1~ 2 and contains the same settings as for N = 4, but
°
in addition we have chosen a5 = and a~ = E.
We can now use the entangled states of Eq. (12.70) with N = 4 and N =
5 and apply the inequalities to them. For the same reason as for N = 3
(symmetry among all modes in the states and in the inequalities), the choice
a = {3 = "( = 8 = 10 = iVY appears to be optimal (maximizes positive terms
and minimizes negative contributions).
With this choice, we obtain
which is clearly a separable mixed state (and indeed not the maximally mixed
J
state ex: dx dx' lx, x')(x, x'I). More interesting is the behaviour of a regular-
J
ized version of dx lx, x, x). In order to apply bipartite inseparability criteria,
let us trace out (integrate out) one mode of the Wigner function W out (x, p) in
Eq. (12.50) for N = 3,
(12.80)
From the resulting Gaussian two-mode Wigner function, we can extract the
inverse correlation matrix. For Gaussian N -mode states with zero mean values,
the Wigner function is given by
where b.~i = ~i - (~i) = ~i for zero mean values. The last equality defines
the correlation matrix for any quantum state, but for Gaussian states of the
form Eq. (12.81), the Wigner function is completely determined by the second-
moment correlation matrix. Now we can calculate the bipartite correlation
matrix of the state in Eq. (12.80),
+ 2C"
)
( e+2>- 0 2 sinh2r 0
V=~ 0
e- 2r + 2e+ 2r 0 -2 sinh2r
12 2 sinh2r 0 e+ 2r + 2e- 2r 0
0 -2sinh2r 0 e- 2r + 2e+ 2r
(12.84)
We could have also obtained this two-mode correlation matrix by extracting the
three-mode correlation matrix V of the state Wout(x, p) in Eq. (12.50) with
N = 3 and ignoring all entries involving mode 1 [or equivalently by explicitly
calculating the correlations between modes 2 and 3 with the Heisenberg opera-
tors in Eq. (12.18) for r = rl = r2 and N = 3]. The resulting two-mode state
is a (mixed) inseparable state for any nonzero squeezing r > O. Note that, for
instance, the total variance in Eq. (12.37) with N = 2 becomes for this state
(5e- 2r + e+ 2r )/6, which drops below the boundary of 1 only for sufficiently
small nonzero squeezing, but approaches infinity as the squeezing increases.
However, we can easily verify the state's inseparability for any r > 0 by look-
ing at the necessary two-party separability condition in product form given in
Ref. [37]. We find that
(12.85)
which drops below the separability boundary of 1/4 for any r > O. Of course,
also the necessary and sufficient partial transpose criterion from Ref. [14]
indicates entanglement for any r > 0 [28]. Recall that by first taking the
"infinite-squeezing limit" and then tracing out one mode, we had obtained a
separable state [Eq. (12.79)]. That was what we expected according to the
result for the maximally entangled qubit state IGHZ).
SO after all, we confirm what we had intuitively expected: the tripartite
state W out (x, p) for finite squeezing is a nonmaximally entangled state like the
qubit state IW). Only for infinite squeezing does it approach the maximally
entangled state f dx lx, x, x), the analogue of IGHZ). This result reflects
what is known for two parties. The two-mode squeezed state W out (x, p) with
N = 2 becomes a maximally entangled state f dx lx, x), such as the Bell state
(100) + 111))/ y'2, only for infinite squeezing. For finite squeezing, it is known
to be nonmaximally entangled.
140 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
3. CONCLUSIONS
Let us conclude by asking whether we were able to find answers to the
questions posed at the beginning of this chapter: how can we generate, measure,
and (theoretically and experimentally) verify genuine multipartite entangled
states for continuous variables? How do the continuous-variable states compare
to their qubit counterparts with respect to various properties?
As for the generation, we demonstrated that genuinely N -party entangled
states are producible with squeezed light resources and beam splitters. In
particular, one sufficiently squeezed light mode is in principle the only resource
needed to create any degree of genuine mUlti-party entanglement by means of
linear optics. The resulting states, though genuinely multi-party entangled, are
always nonmaximally entangled mUlti-party states due to the finite amount of
the squeezing. They behave like the N -party versions of the qubit state IW).
First, they also contain bipartite entanglement readily available between any
pair of modes (just as IW) and as opposed to the qubit state IGHZ). Secondly,
they yield a non-exponential increase of violations of mUlti-party Bell-type
inequalities for growing number of parties (as for IW) and different from the
qubit state IGHZ) for which the increase is exponential).
Furthermore, we have seen that by inverting the circuits for generating
genuine but nonmaximum mUlti-party entanglement, one can perform projec-
tion measurements onto the maximally entangled mUlti-party (GHZ) basis for
continuous variables. In contrast to the difficulties in performing such measure-
ments for photonic qubits within the framework of linear optics, continuous-
variable GHZ measurements only require beam splitters and homo dyne detec-
tors. In addition, we showed that the circuits for measuring maximum GHZ
entanglement are also applicable to the theoretical and experimental verifica-
tion of the nonmaximum entanglement of the multi-party states (which are
those producible in the laboratory). The circuits provide a necessary condition
for full separability of any N-partite N-mode state (pure or mixed, Gaussian
or non-Gaussian) with any number of modes N. However, this condition is
not sufficient for full separability and, more importantly, its violation does not
verify genuine but only partial multipartite entanglement. For the theoretical
verification of genuine multipartite entanglement, additional assumptions have
to be taken into account such as the total symmetry of the relevant states. There-
fore, an unambiguous experimental proof of genuine multipartite entanglement
of continuous-variable states was not proposed in this chapter. A possible ap-
proach to this would be to consider the violation of stricter N-party Bell-type
inequalities which cannot be violated by only partially entangled states. How-
ever, the experimental nonlocality test would then rely on observables such as
the photon number parity, and hence become unfeasible with current technol-
Multipartite entanglementfor continuous variables 141
ogy. More desirable would be a test for genuine multipartite entanglement that
is solely based on linear optics and efficient homodyne detections.
Notes
1. Separable states also exhibit correlations, but those are purely classical. For instance, compare the
separable state P == ~(10)(010 10)(01 + 11)(11 0 11)(11) to the pure maximally entangled "Bell state"
I<J>+) == 0(10) 010) + 11) 011)) = 0(1+) 01+) + H 01-)) with the conjugate basis states
I±) = 0(10) ± 11)). The separable state p is classically correlated only with respect to the predetermined
basis {I 0), II)}. However, the Bell state I<J> +) is a priori quantum correlated in both bases {I 0), II)} and
{I +), 1-)}, and may become a posteriori classically correlated depending on the particular basis choice in a
local measurement. Similarly, we will see later that the inseparability criteria for continuous variables need
to be expressed in terms of the positions and their conjugate momenta.
2. Inseparable states with positive partial transpose cannot be distilled to a maximally entangled state
via local operations and classical communication. They are so-called "bound entangled" [8]. The converse,
however, does not hold. An explicit example of a bound entangled state with negative partial transpose was
given in Ref. [9]. In other words, not all entangled states that reveal their inseparability through negative
partial transpose are distillable or "free entangled". On the other hand, any state P12 that violates the
so-called reduction criterion, PI 0 !l - P12 2: 0 or !l 0 ih - Ih2 2: 0, is both inseparable and distillable
[10]. This reduction criterion is in general weaker than the partial transpose criterion and the two criteria
are equivalent in the (2 X 2)- and (2 x 3)-dimensional cases.
3. A possible continuous-variable generalization of the C-NOT gate is lXI, X2) -+ lXI, Xl + X2),
where the addition modulo two of the qubit C-NOT, IYl, YZ) -+ IYl, Yl ED Y2) with Yl, Y2 = 0,1, has
been replaced by the normal addition. However, for the quantum circuit here, a beam splitter operation as
described by Eq. (12.14) is a suitable substitute for the generalized C-NOT gate.
4. A full classification of tripartite Gaussian states is given in Ref. [27] in analogy to that for qubits
from Ref. [18]. In addition, necessary and sufficient three-mode inseparability criteria for Gaussian states
are proposed in Ref. [27].
5. The variances of the N - 1 relative positions :h - X2, X2 - X3, ... , and XN -1 - XN are also
available via the variances of the output quadratures of the inverse N -splitter. First, the variance of
x~ == 0 (x N -1 - XN ) corresponding to the last line in Eq. (12.30) is directly measurable. In addition,
by converting the measured photocurrent into a light amplitude and "displacing" (feed-forward) x~_l
according to x~_l -+ x'fv_l = x~_l - Jsx~ == !f(XN-2 - XN-l), one can directly measure
the variance of XN -2 - XN -1 etc. Similarly, one would also employ this feed-forward technique in a
mUlti-party quantum communication protocol that relies on the N classical results of an N -mode GHZ state
measurement.
6. note that we use the term "partially entangled" here for states which are not genuinely multi-
party inseparable. In the literature, sometimes "partial entanglement" is also referred to as nonmaximum
entanglement of two or more parties (in the sense that for two parties the Schmidt coefficients are not all
equal). As discussed later, also the genuinely multi-party entangled continuous-variable states are only
nonmaximally entangled due to the finite degree of the squeezing.
7. This choice of two equal settings leads to the same result as that of Banaszek and Wodkiewicz [34]
who used opposite signs: Q = v:J and f3 = -v:J.
8. However, remember that this is not the maximally mixed state for two qubits. Only when tracing
out two parties do we end up having the maximally mixed one-qubit state.
References
As with discrete systems, quantum entanglement also plays the basic role in
quantum information protocols with continuous variables. A problem of great
importance is then to check whether a continuous variable state, generally
mixed, is entangled (inseparable). For discrete systems, there is the Peres-
Horodecki inseparability criterion [1, 2], based on the negativity of the partial
transpose of the composite density operator. This negativity provides a neces-
sary and sufficient condition for inseparability of 2 x 2 or 2 x 3-dimensional
systems. In this section, we will describe an entirely different inseparability
criterion for continuous variable states, which was first proposed in Ref. [3].
The Peres-Horodecki criterion was also successfully extended to the continu-
ous variable systems shortly afterwards, which will be described in the next
section by Simon.
The inseparability criterion described here is based on the total variance
of a pair of Einstein-Podolsky-Rosen (EPR) type operators. For any separable
continuous variable states, this total variance is bounded from below by a certain
value resulting from the uncertainty relation, whereas for entangled states
this bound can be exceeded. So violation of this bound provides a sufficient
condition for inseparability of any continuous variable state. Furthermore, for
the set of Gaussian states, which are of great practical importance, this criterion
turns out to be a necessary and sufficient condition for inseparability. In fact,
for any Gaussian state the compliance with the low bound by a certain pair
of EPR type operators guarantees that the state has a P-representation with
positive distribution, so the state must be separable.
145
S.L. Braunstein and A.K. Pari (eds.), Quantum Information with Continuous Variables, 145-153.
© 2003 Kluwer Academic Publishers.
146 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
where we assume Pil and Pi2 to be normalized states of the modes 1 and 2,
respectively, and Pi ~ 0 to satisfy L:
Pi = l.
i
A maximally entangled continuous variable state can be expressed as a co-
eigenstate of a pair of EPR-type operators [4], such as Xl + X2 and Pi - P2.
SO the total variance of these two operators reduces to zero for maximally
entangled continuous variable states. Of course, the maximally entangled
continuous variable states are not physical, but for the physical entangled
continuous variable states-the two-mode squeezed states [5], this variance will
rapidly tend to zero by increasing the degree of squeezing. Interestingly, we
find that for any separable state, there exists a lower bound to the total variance.
To be more general, we consider the following type of EPR-like operators:
u
~
= Ia I~Xl + -X2,
1~ (13.2)
a
v
~
= Iapi
I~ 1~
- -P2, (13.3)
a
u v
Proof. We can directly calculate the total variance of the and operators
using the decomposition (13.1) of the density operator p, and finally get the
following expression:
Inseparability criterion for continuous variable systems 147
+2 1: 1
(~Pi (Xl)i (X2)i - ~ Pi (Pi)i (P2)i) - (U)~ - (v)~ (13.5)
~ ~
+ :2 (( (LlX2)2) + ((Llp2)2) J]
i
the last line of Eq. (13.5) is bounded from below by zero. Hence, the total
u v
variance of the two EPR-like operators and is bounded from below by
a 2 + -%- for any separable state. This completes the proof of the theorem.
a
Note that this theorem in fact gives a set of inequalities for separable states.
The operators Xj ,Pj (j = 1,2) in the definition (13.1) can be any local operators
satisfying the commutators [Xj,Pj'] = ~8jj'. In particular, if we apply an
u v,
arbitrary local unitary operation Ul Q9 U2 to the operators and the ineqUality
(13.4) remains unchanged. Note also that without loss of generality we have
taken the operators Xj and Pj dimensionless.
For inseparable states, the total variance of the u and v operators is required
by the uncertainty relation to be larger than or equal to ~ Ia 2 - ~ I ' which
reduces to zero for a = 1. For separable states the much stronger bound given
by Eq. (13.4) must be satisfied. A natural question is then how strong the
bound is. Is it strong enough to ensure that if some inequality in the form of Eq.
(13.4) is satisfied, the state necessarily becomes separable? Of course, it will be
very difficult to consider this problem for arbitrary continuous variable states.
However, in recent experiments and protocols for quantum communication
[6, 7, 8, 9, 10, 11, 12, 13], continuous variable entanglement is generated by
two-mode squeezing or by beam splitters, and the communication noise results
148 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
from photon absorption and thermal photon emission. All these processes
lead to Gaussian states. So, we will limit ourselves to consider Gaussian
states, which are of great practical importance. We find that the inequality
(13.4) indeed gives a necessary and sufficient inseparability criterion for all the
Gaussian states. To present and prove our main theorem, we need first mention
some notations and results for Gaussian states.
It is convenient to represent a Gaussian state by its Wigner characteristic
function. A two-mode state with the density operator p has the following
Wigner characteristic function [5]
tr [pexp !
(AlaI - Ai a + A2 a2 - A;an]
tr {pexp [2i (A{X1 + Afp1 + A~X2 + A~P2)J XP.6)
where the parameters Aj = Af + iA§, and the annihilation operators aj =
Xj + ipj, with the quadrature amplitudes Xj,Pj satisfying the commutators
[Xj ,Pj'] = ~ 15j j' (j, j' = 1, 2). For a Gaussian state, the Wigner character-
istic function x(w) (AI, A2) is a Gaussian function of Af and A§ [5]. Without
loss of generality, we can write x(w) (AI, A2) in the form
x(W) (AI, A2) = exp [-~ (Ai, Af, A~, A~) M (Ai, Af, A~, A~)T] (13.7)
In Eq. (13.7), linear terms in the exponent are not included since they can
be easily removed by some local displacements of Xj, pj and thus have no
influence on separability or inseparability of the state. The correlation property
of the Gaussian state is completely determined by the 4 x 4 real symmetric
correlation matrix M, which can be expressed as
n c
MI = (
scm :
n
) ,(n,m ~ 1) (13.9)
c'
Proof. A LLUBO on the state PG transforms the correlation matrix M in the
Wigner characteristic function in the following way
(13.10)
where VI and V2 are real matrices with det VI = det V2 = 1. Since the
matrices GI and G 2 in Eq. (13.8) are real symmetric, we can choose first
a LLUBO with orthogonal VI and V2 which diagonalize G I and G2, and
then a local squeezing operation which transforms the diagonalized G I and
G 2 into the matrices G~ = nI2 and G~ = mI2, respectively, where 12 is
the 2 x 2 unit matrix. After these two steps of operations, we assume the
matrix 0 in Eq. (13.8) is changed into 0', which always has a singular value
decomposition, thus it can be diagonalized by another LLUBO with suitable
orthogonal VI and V2. The last orthogonal LLUBO does not influence G~
and G~ any more since they are proportional to the unit matrix. Hence, any
Gaussian state can be transformed by three-step LLUBOs to the standard form
I. The four parameters n, m, c, and c' in the standard form I are related to
the four invariants det G I, det G 2 , det C, and det M of the correlation matrix
under LLUBOs by the equations det GI = n 2 , det G2 = m 2 , det C = cc',
and det M = (nm - c2) (nm - c'2) .
Lemma 2 (standard form ll): Any Gaussian state PG can be transformed
through LLUBOs into the standard form II with the correlation matrix given by
(13.11)
nI-1 n2 -1
, (13.12)
mI-1 m2- 1
hl-l c21 J(ni -1) (mi -1) - J(n2 - 1) (m2 -1).(13.13)
Proof. First, any Gaussian state can be tranformed through LLUBOs to the
standard form I. We then apply two additional local squeezing operations on
the standard form I, and get the state with the following correlation matrix
150 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
nrl J r lr2 C ,
, B..
Tl
c
.jTIT2
M= (13.14)
J r lr2 e , mr2
c m
.jTIT2 T2
where rl and r2 are arbitrary squeezing parameters. M' in Eq. (13.14) has the
standard form M! I (13 .11) if rl and r2 satisfy the following two equations
B..-1 m_1
Tl = _T'-"2'---_, (13.15)
nrl -1 mr2-1
Our task remains to prove that Eqs. (13.15) and (13.16) are indeed satisfied by
some positive rl and r2 for arbitrary Gaussian states. Without loss of generality,
we assume lei ~ Ie' I and n ~ m. From Eq. (13.15), r2 can be expressed as a
continuous function of rl with r2 (rl = 1) = 1 and r2 (rl) Tl--+f m. Substitut-
ing this expression r2 (rl) into Eq. (13.16), we construct a function f (rl) by
subtracting the right hand side ofEq. (13.16) from the left hand side. Obviously,
f (rl = 1) = lei-Ie' I ~ 0, and f (rl) T~ Jrlm (lcl- In (m - !)) s
0, where the inequality J
lei s n (m - rk) results from the physical condition
((~11:0)2) + ((~VO)2) ~ 1[11:0,0'0]1 with 11:0 = Jm-
!Xl - rcrvnX2 and
0'0 = '(;: P2. It follows from continuity that there must exist a ri E [1, (0)
which makes f (rl = ri) = 0. So Eqs. (13.15) and (13.16) have at least one
solution. This proves lemma 2.
We remark that corresponding to a given standard form I or II, there are
a class of Gaussian states, which are equivalent under LLUBOs. Note that
separability or inseparability is a property not influenced by LLUBOs, so all
the Gaussian states with the same standard forms have the same separability
or inseparability property. With the above preparations, now we present the
following main theorem:
Theorem 2 (necessary and sufficient inseparability criterion for Gaus-
sian states): A Gaussian state Pa is separable if and only if when expressed
in its standard form II, the inequality (13.4) is satisfied by the following two
EPR-type operators
Inseparability criterion for continuous variable systems 151
CI
u = aOxI- --X2,
~
Icd ao
1~
(13.17)
C2
v = aOPI- --P2,
~ 1~
(13.18)
IC21 ao
2 =
where a o Jm1-1 = Jm -1.
nl-l n2-1
2
Proof. The 'only if' part follows directly from theorem 1. We only need
to prove the 'if' part. From lemma 2, we can first transform the Gaussian
state through LLUBOs to the standard form n. The state after transformation
is denoted by pM.Then, substituting the expression (13.17,13.18) of and u
vinto the inequality (13.4), and calculating ((~u)2) + ((~v)2) using the
correlation matrix MIl, we get the following inequality
2nl + n2 ml + m2 1 1 (2 1)
ao 4 + 4a5 - 2' (I cII + Ic21) 2: 2' ao + a5 ' (13.19)
where P (a, (3) is the Fourier transformation of X(~) (AI, A2) and thus is a
positive Gaussian function. Eq. (13.23) shows pg
is separable. Since the
original Gaussian state PG differs from pM
by only some LLUBOs, it must
also be separable. This completes the proof of theorem 2.
152 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Now we have a necessary and sufficient inseparability criterion for all the
Gaussian states. We conclude the paper by applying this criterion to a simple
example. Consider a two-mode squeezed vacuum state e 1 2
r(atat -a1( 2 )
Ivae)
with the squeezing parameter r. This state has been used in recent experiment
for continuous variable quantum teleportation [13]. Suppose that the two
optical modes are subject to independent thermal noise during transmission
with the same damping coefficient denoted by 'fJ and the same mean thermal
photon number denoted by n. It is easy to show that after time t, the standard
correlation matrix for this Gaussian state has the form of Eq. (13.9) with n =
m = cosh (2r) e- 21/t + (2n + 1) (1 - e- 21/t ) and e = -e' = sinh (2r) e- 21/t
[15]. So the inseparability criterion means that if the transmission time t
satisfies
References
[1] A. Peres, Phys. Rev. Lett. 77, 1413 (1996).
[2] M. Horodecki, P. Horodecki, and R. Horodecki, Phys. Lett. A 223, 1
(1996).
[3] L. M. Duan, G. Giedke, J. I. Cirac, and P. Zoller, Phys. Rev. Lett. 84, 2722
(2000).
[4] A. Einstein, B. Podolsky, and R. Rosen, Phys. Rev. 47, 777 (1935).
[5] C. W. Gardiner and P. Zoller, Quantum Noise (2nd. Ed.), Springer-Verlag
(1999).
[6] L. Vaidman, Phys. Rev. A 49, 1473 (1994).
[7] S. L. Braunstein, H. J. Kimble, Phys. Rev. Lett. 80, 869 (1998).
[8] S. L. Braunstein, Nature 394, 47 (1998).
[9] S. L. Braunstein and S. Lloyd, Phys. Rev. Lett. 82,1789 (1999).
[10] G. J. Milburn and S. L. Braunstein, quant-phl9812018.
[11] P. Loack, A. L. Braunstein, and H. J. Kimble, quant-phl9902030.
[12] A. S. Parkins, and H. J. Kimble, quant-phl9904062.
Inseparability criterion/or continuous variable systems 153
SEPARABILITY CRITERION
FOR GAUSSIAN STATES
R. Simon
The Institute of Mathematical Sciences, Tharamani, Chennai 600 113, India
Abstract The PPT (positivity under partial transpose) criterion is studied in the context of
separability of continuous variable bipartite states. The partial transpose oper-
ation admits, in the Wigner representation of quantum mechanics, a geometric
interpretation as momentum reversl or mirror reflection in phase space. This
recognition leads to uncertainty principles, stronger than the traditional ones, to
be obeyed by all PPT (separable as well as bound entangled) states. In the special
case of bipartite two-mode systems, the PPT crrterion turns out to be necessary
and sufficient condition for separability, for all Gaussian states: a 1 + 1 syatem
has no bound entangled Gaussian state. The symplectic group oflinear canonical
transformations and the representation of these transformations through (meta-
plectic) unitary Hilbert space operators play an important role in our ananysis.
transposition turns out to be necessary and sufficient for separability for all
states of 2 x 2 and 2 x 3 dimensional bipartite systems, but ceases to be so in
higher dimensions as shown by Horodecki [10]. Entanglements which are not
witnessed by the partial transpose operation cannot be distilled, and for this
reason such entanglements have come to be known as bound entanglements.
With increasing Hilbert space dimension, tests for separability will be ex-
pected to become more and more difficult to implement in practice, and less
and less definitive in their outcome. On the other hand, in the infinite dimen-
sional case corresponding to continuous variable systems Gaussian states are
of particular interest from the point of view of experiments, and for this reason
a test for separability which is decisive for all Gaussian states could be of
considerable value in the continuous case, even if it fails to be decisive for non-
Gaussian states. Hence is the importance of the result establised in Ref. [11]
that the partial transpose condition proves both necessary and sufficient (NS)
condition for separability, for all Guassian states of a bipartite system of two
harmonic oscillators. We should hasten to add that the partial transpose crite-
rion of separability ceases to be NS, even for Gaussian states, when both sides
of the system (Alice and Bob) have two or more oscillators each, as shown by
Werner and Wolf [12].
An interesting approach to separability of Gaussian states, based on the total
variance of a pair of Einstein-Podolsky-Rosen type operators, was indepen-
dently formulated by Duan et al. [13]. This approach too leads to a criterion
which proves NS in the 1 + 1 case. The same authors have proposed also
an entanglement purification protocol to generate maximally entangled states
from two-mode squeezed states or from mixed Gaussian entangled states [14].
The origin of the fact that the issue of separability is easily tractable in
the case of Gaussin states can be traced as follows. First, a Guassian state is
fully determined by its first and second moments. The Weyl group of unitary
operators which effect translations in phase space are manifestly separable,
and hence their action does not affect the separability or otherwise of a state.
We may thus assume, wihout loss of generality, that the first moments of the
state of our interest all vanish. In other words, separability of a Gaussian state
is determined entirely by the variance (or covariance) matrix of the state. The
problem thus gets reduced from study of the infnite dimensional density matrix
to analysis of the finite dimensional variance matrix.
Secondly, under the unitary action of the symplectic group of real linear
canonical transformation the moments of a quantum state in general, and the
variance matrix in particular, undergo simple geometric changes. One may
thus use local symplectic transformtions (local transformations have no effect
on separability) to convert the variance matrix into a canonical form in which
the separability issue becomes particularly easy to settle.
Separability criterionjor Gaussian states 157
ql iiI
PI PI
q2 ii2
~= P2 t= P2 (14.1)
qn qn
Pn Pn
The complete set of canonical commutation relations
where
J 0 0 o
o J 0 o
o 0 J o (14.4)
000 J
Linear homogeneous canonical transformations in the 2n-dimensional phase
space of our n-oscillator system act as Hilbert space unitary operators in the
quantum description. With a set of canonical coordinates chosen for the phase
space, such a transformation is identified by its matrix S:
S: ~ -+ e = S~. (14.5)
158 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
snsT = n, (14.6)
which is recognized as the defining condition for Sp(2n, R), the group of
symplectic matrices. Thus, linear canonical transformations are in one-to-one
correspondence with elements of the symplectic group Sp(2n, R). It is useful
to note that det S = 1, \:IS E Sp(2n, R), and that S E Sp(2n, R) implies
sr E Sp(2n, R).
Let U (S) be a Hilbert space unitary operator corresponding to the canonical
transformation S. This means we have the following evolution for the operators
t
2n
U(S): ta -+ t~ =U(S)t taU(S) = "L Sa/3t- (14.7)
/3=1
We could have written this in the abreviated form t' = U(S)t tU(S) = S t-
In order that the unitaty operators U(S) give a representation of Sp(2n, R),
we should be able to choose the former in such a way that U(S)U(S') =
U(SS'), \:IS, s' E Sp(2n.R). Clearly, U(S)t tU(S) = S t allows, in the cor-
respondence S -+ U(S), freedom of a S-dependent phase factor exp( i¢(S) )
mUltiplying U(S). Can we make a one-to-one choice S -+ ¢(S) such that
U(S)U(S') = U(SS'), \:IS, s' E Sp(2n.R)? The answer turns out to be
in the negative. The maximum simplification one can obtain in this regard is
a one-to-two correspondence S -+ ±U(S). In other words this correspon-
dence yields a two-valued unitary 'representation' of the symplectic group.
Conversely, these unitary operators close to form the defining representation
of a new group called the metaplectic group Mp(2n, R). This group is a
double cover of the symplectic group. A detailed discussion of the symplectic-
metaplectic connection, with further references, can be found in Ref. [16]. The
situation is somewhat analogous to the well known connection between SO(3)
and SU(2). The major diference is that SO(3) is doubly connected and so its
double cover SU(2) is simply connected (and hence is the universal cover of
SO(3) ); but Sp(2n, R) is infinitely connected, and so also is its double cover.
The point being made is that what is natural for the quantum description of a
system of oscillators at the state vector level is the metaplectic group rather than
the symplectic group. The symlectic group is recoverd in the description at
the density operator level (or, equivalently, in the Wigner description), though.
This subtle aspect can be appreciated by considering just one oscillator. After
evolution through one period, which corresponds to the identity element of
Sp(2, R), the position and momentum - and indeed the entire phase space-
return to their original configuration. But the unitary evolution of a statevector
17J;) of the same oscillator through one period returns - 17J;) rather than 17J;)
Separability criterion for Gaussian states 159
itself. The state vector is recovered in full only after two periods. Originating
in this subtlety is not only the zero-point energy, but also the Gouy phase picked
up by a light beam as it crosses a focus [17], the 7r/4 phase picked up by the
WKB wavefunction at the turning points, and the Maslov index.
W(q,p) = 7r- nJd nq, (q - q'l p Iq + q') exp(2i q' . p). (14.8)
where q = (ql, q2, ... qn) andp = (Pl,P2, ... Pn). We will often write W(e)
in place of W (q, p). It is clear that this map is invertible, and thus the function
W (q, p) captures the dendity matrix in its entirety. In most aspects of structure
and dynamics, the Wigner distribution function W(q,p) behaves exactly like
the phase space density in classical statistical mechanics, but W (q, p) is not
pointwise nonnegative for most quantum states. Indeed, the Wigner function
of a pure state is pointwise nonnegative if and only if the state is Gaussian [19]!
The defining properties of Wigner distributions are transcriptions, through
the map (14.8), of those of density operators. Thus, hermiticity of density
operator p is equivalent to the corresponding phase space distribution being
real, while the condition trp = 1 transcribes into Jd2n e W(e) = 1. The
nonnegativity of p, which can be stated as the requirement that tr(pp') ~ 0 for
every den sty operator p', gets translated into the condition
(14.10)
these is the action of canonical transformations. We have noted that the sym-
plectic group Sp(2n, R) acts unitarily and irreducibly on the n-mode Hilbert
space [20]. The (infinite dimensional) unitary operator U(S) corresponding
to S E Sp(2n, R) transforms the state vector i"p) to i"p') = U(S)i"p), and
hence the density operator p to p' = U (S) PU (S) t. This transformation takes
a strikingly simple form in the Wigner description, and this is one reason for
the effectiveness of the Wigner picture in handling canonical transformations
[20]:
S: p ---7 U(S) pU(S)t ¢:::::? W(e) ---7 W'(e) = W(S-le). (14.11)
That is, W'(Se) = W(e) for every canonical transformation S E Sp(2n, R),
and the Wigner function transforms as a Sp(2n, R) scalar field.
Secondly, the transpose map T and the partial transpose map PT take tran-
parent geometric form in the Wigner description. Indeed, it follows from the
definition of Wigner distribution that transpose operation on the density opera-
tor, which is equivalent to complex conjugation of the elements of the density
matrix in the position representation, transcribes faithfully into momentum
reversal operation in the Wigner description:
T: W(q,p) -+ W'(q,p) = W(q, - p) = W(Ae),
A = diag(I,-I; 1. -1; ... ; 1,-1)
= a3 EEl a3 EEl ••• EEl a3· (14.12)
(14.13)
Thus, f31. = tr(Vn)21., i = 1,2, ... , n are symplectic invariants. And these
are the only inepenent invariants of V, (Vn)k being traceless for odd values
of k.
It is elear that not every rotation in the 2n-dimensional phase space is a
canonical transformation. Indeed, the set of all phase space rotations forming
a subgroup of Sp(2n, R) is isimorphic to the n 2 -dmensional unitary group
U(n). And n 2 is smaller than 2n2 - n, the dimension of the full roration
group SO(2n), for n > 1. We do not, therefore, expect in general to be
able to diagonalize a real symmetric matrix V using symplectic congruence
V -+ V' = sv ffI', S E Sp(2n, R). Williamson theorem [21, 15, 22]
guaranttes that if V possesses the additional quality of being positive definite,
it can be diagonalized through symplectic congruence. That is, V > 0 implies
that there exists an Sv E Sp(2n, R) such that
Sv: V -+ Vwc = Sv V SJ = diag( ~l, ~l; ~2, ~2; ... ; ~n, ~n X.l4.19)
Comparing with the Gaussian Wigner distribution in (14.15), it is clear that
the Williamson canonical form Vwc corresponds to product of single-mode
thermal state Wigner distributions, the temperature parameter of the j -th mode
being determined by ~j. The invariant traces are asily computed to be
n
f3<
n -
.-
- 2 "'(~
~ J.)21. , i -
-I" 2 ... ,n. (14.20)
j=l
Proposition 2: The variance matrix of a state has the special form V = ~ SST,
with S E Sp(2n, R), if and only if the state under consideration is a Gaussian
pure state.
Consistent with this notation we may write the symplectic metric n as a direct
sum of the symplectic metrics of the two subsystems:
o
Q= ( (14.22)
o
Let P be a bipartite density matrix, and W(e) = W(eA, eB) the correspond-
ing Wigner distribution. The partial transpose map with respect to the Bob
subsystem corresponds to
W(e) -+ W'(e) = W(A'e) = W(eA,ABeB),
P= LPkPAk0PBk, (14.24)
k
with nonnegative Pk'S, where PAk'S and PBk'S are density operators of the
subsystems of Alice and Bob respectively. Clearly, product states correspond
to product Wigner distributions and separable states separable Wigner distribu-
tions:
W(e) = LPk WAk(eA)WBk(eB). (14.25)
k
o
V= ( (14.26)
o VB
Separability criterion for Gaussian states 165
where VA and VB are the variance matrices of the subsystems. Let VA k EI1 VB k
be the variance matrix of the product state Wk(e) = WAk(eA)WB k(eB),
and let e(k) be its first moments. Then the variance matrix of the separable
state (14.25) is given by [recall that (e) has been assumed to vanish for W(e) ]
(L Pk VAk) EEl (L Pk VB k)
k k
1 T T
> "2 SA SA EElSBSB· (14.29)
Since the second sum on the right hand side of (14.27) is a manifestly nonneg-
ative matrix, we have estalished a necessary condition on the variance matrix
of a separable state.
1 T T
V ~ "2 SASA EI1 SBSB· (14.30)
The validity of this condition is quite general: the number of modes in the
possession of Alice and Bob can be arbitrary, and the state under reference
need not be Gaussian.
If the state under consideration is Gaussian, the above result can be con-
siderbly strengthened. To this end, we begin by noting the following.
d et v(res)]
class J dTt." '
where dTe indicates integraton over the range of Vclass, and WVGPp(e - e) is
the Wigner distribution of the GPP state. Thus we have
V + in>
2 -,
0 n= A'nA' = (nA0 0)
-nB '
(14.35)
6. TWO-MODE SYSTEMS
In this case wherein Alice and Bob hold a single mode each, the set of all
homogeneous real linear canonical transformations constitute the ten-parameter
real symplectic group Sp( 4, R), and nand becomen
n= ( JO)
0J
- (J0 -J0) .
' n= (14.38)
It is possible and desirable to cast the stronger uncertainy principle (14.37) for
separable states in an Sp(2, R) ® Sp(2, R) invariant form.
The congruence V -+ Slocal V ~cal by the local group changes the blocks
of V in the following manner:
To prove this result, first note that the two inequalities are equivalent for variance
matrices of the special form
Vo = (14.42)
But any variance matrix can be brought to this special form by effecting
a suitable local canonical transformation corresponding to some element of
Sp(2, R) x Sp(2, R). In veiw of the manifest Sp(2, R) ® Sp(2, R) invariant
structure of (14.41), it follows that the two inequalities are indeed equivalent
for all two-mode variance matrices.
Under partial transpose or momentum reversal on Bob's side represented by
the phase space mirror reflectin A' = Id EEl C73, we have V -+ V = A'V A'.
That is, C -+ CC73 and B -+ C73BC73, while A remains unchanged [C73 is the
diagonal Pauli matrix: C73 = diag (1, -1)]. As a consequence, 13 = det C flips
signature while h, hand 14 remain unchanged. Thus, condition (14.34) for V
takes a form identical to (14.41) with only the signature in front of det C in the
second term on the left hand side reversed. Thus the PPT requirement that the
variance matrix of a separable state has to obey A'V A' + ~n ;::: 0 inaddition to
the fundamental uncertainty principle V + ~n ;::: 0, takes the form
This is the final form of the implication of PPT on the variance matrix of a
two-mode bipartite state. This condition is invariant not only under phase space
mirror reflection (partial transpose) A', but also under Sp(2, R) ® Sp(2, R), as
it should be! It constitutes a complete description of the implication PPT has
for the second moments of any state.
For the standard form Va, our condition (14.43) reads
(14.44)
But the point is that the separability (PPT) check (14.43) can be applied directly
on V, with no need to transform it to the special form Vo.
To summarise, conditions (14.33), (14.41), and V + ~n 2: 0 are equivalent
statements of the fundamental uncertainty principle, and hence will be satisfied
by every physical state. The mutually equivalent statements (14.37), (14.43),
and (14.34) constitute the PPT criterion at the level ofthe second moments, and
should necessarily be satisfied by every separable state, Gaussian or otherwise.
As the reader may anticipate, we can make stronger statements for Gaussan
states: separability and PPT become equivalent in the two-mode Gaussian case.
To this end, note that states with det C 2: 0 definitely satisfy the PPT condi-
tion (14.43), which in this case is subsumed by the physical condition (14.41).
We will begin by establishing that such Gaussian states are indeed separable.
(14.45)
PI, P2 planes:
Vto' -+ Vt"
0 = d'lag (~+, " );
~+, ~-, ~_ .
(1446)
Proof: We consider in turn the two distinct cases det C < 0 and det C ~ O.
Suppose det C < O. Then there are two possibilities. If (14.33) [or (14.34)]
is violated, then the Gaussian state is definitely entangled since (14.33) is a
necessary condition for separability. If (14.33) is respected, then the mirror
reflected state is a physical Gaussian state with det C > 0 (recall that mirror
reflection flips the signature of det C), and is separable by the above lemma.
This implies separability of the original state, since a mirror reflected separable
state is separable. Finally, suppose det C ~ O. Condition (14.33) is definitely
satisfied since it is subsumed by the uncertainty principle V + ~n ~ 0 in the
det C ~ 0 case. By our lemma, the state is separable. This completes proof of
the theorem.
Separability criterion for Gaussian states 171
2 0 0 0 1 0 0 0
0 1 0 0 0 0 0 -1
0 0 2 0 0 0 -1 0
0 0 0 1 0 -1 0 0
V=~ (14.49)
2 1 0 0 0 2 0 0 0
0 0 0 -1 0 4 0 0
0 0 -1 0 0 0 2 0
0 -1 0 0 0 0 0 4
Their proof that this Gaussian state is PPT, and that it is not separable, is
distinguished by effective use of symmetry arguments.
To conclude, we note that a recent work of Giedke et. al. [23] formulates a
necessary and sufficient condition for separability of Gaussian states of bipartite
systems of arbitrary number of modes. These authors show that all bipartite
Gaussian states with nonpositive partial transpose are distillable.
References
[1] L. Vaidman, Phys. Rev. A 49, 1473 (1994); L. Vaidman and N. Yoran,
Phys. Rev. A 59, 116 (1999).
[2] A. S. Parkins and H. J. Kimble, quant-phl9904062; quant-phl9907049;
quant-phl9909021.
[3] S. L. Braunstein, Nature 394,47 (1998); quant-phl9904002; S. Lloyd and
S. L. Braunstein, Phys. Rev. Lett. 82, 1784 (1999) ..
[4] G. J. Milburn and S. L. Braunstein, quant-phl9812018.
[5] P. van Loock, S. L. Braunstein, and H. J. Kimble, quant-phl9902030; P.
van Loock and S. L. Braunstein, quant-phl9906021; quant-phl9906075.
[6] S. L. Braunstein and H. J. Kimble, Phys. Rev. Lett. 88, 86991998).
172 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
1. ENTANGLEMENT PURIFICATION
Entanglement (cf. Ch. II.7) is one of the most striking and characteristic
features of quantum mechanics. In quantum communication it is also a valuable
resource: it is necessary for many applications such as quantum teleportation
[1] or quantum cryptography[2], and is typically used up in those processes.
Ideally, the communicating parties A(lice) and B(ob) share maximally entan-
gled states, which enable them to perfectly implement the desired protocols.
In reality, however, due to loss and decoherence in the channel connecting A
and B, it is only possible to generate mixed, partially entangled states between
distant locations, which are often not directly useful for the task at hand. A way
to overcome this problem is offered by entanglement purification or distillation
[3,4,5], which describes the process of obtaining at least one maximally en-
tangled state out of many partially entangled ones. In general, any sequence of
local quantum operations (possibly correlated by classical communication) that
allows A and B to transform a (sufficiently large) number of entangled states
173
S.L. Braunstein andA.K. Pati (eds.), Quantum 11!formation with Continuous Variables, 173-192.
© 2003 Kluwer Academic Publishers.
174 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
then the state p is distillable. Here, trB stands for the partial trace with respect
to the second subsystem. An important aspect of this criterion is that if one can
find a state I~) satisfying (15.1), then one can explicitly construct a protocol to
distill p [12].
The question of distillability of general CV states is a formidable task that
we do not attempt to tackle here. Rather we consider in the remainder of this
chapter a simple but important class of CV states, namely Gaussian states of two
modes. There are several reasons to focus on these states. First, most of the CV
states that can currently be prepared in the lab are Gaussian states [13]. Second,
important CV quantum communication protocols are based on Gaussian states
(teleportation [14, 15], cryptography [16]), and finally, Gaussian states are
mathematically well understood, e.g., there exists an inseparability criterion.
Thus the largest set of potentially distillable states is easily characterized. In
the next section we will show that indeed all inseparable two-mode Gaussian
states can be distilled.
Distillability and entanglement purification for Gaussian states 175
2. GAUSSIAN STATES
We consider states that are defined on the Hilbert space L2(JR) (or N copies
»,
thereof: L2(JR)®N = rJ(JR N that is, e.g., states of one or more modes of
the electromagnetic field. (We will in the following often use quantum optical
terms, like "modes" or "beam splitters", because quantum optics currently
offers the most promising setting for the realization of CV systems; this does
not limit the presented results to quantum optical systems.)
Gaussian states of two modes Any Gaussian state of two modes can be
transformed into what we called the standard form, using local quasifree trans-
formations (LQT) only [21, 22]. For a state in standard form the corresponding
characteristic function has displacement d = 0 and the correlation matrix M
has the simple form
(15.7)
where the block matrices are diagonal; and M a , Mb proportional to the identity:
The four real parameters (na, nb, kx, kp) fully characterize a Gaussian state
up to LQTs. They can be easily calculated from the LQT-invariants IMAI,
IMBI, IMABI and IMI via:
(IS.9a)
(IS.9b)
Inseparability On the other hand, in [22] it was shown that a Gaussian state is
entangled iff it does not transform into a proper state under partial transposition.
Starting from this result it is easy to show that a Gaussian state is entangled iff
the corresponding parameters satisfy [28]
(15.11)
which still satisfies (15.1) but now with 11/-1) = I~~) := IN 2:f=llk, k), the
symmetric maximally entangled state of two N-Ievel systems. In this case,
(15.1) implies tr(p Iq,~) (q,~j) > liN.
A state satisfying this inequality can be distilled by a generalization of the
protocol of Ref. [3], which consists of two steps: depolarization and joint
measurements.
(ii) Applying an operation of the form U ® U* (U a randomly chosen unitary)
depolarizes p, i.e. transforms it into a mixture of the maximally entangled state
Iq,~) (which is invariant under transformations of the form U ® U*) and the
178 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
completely mixed state ~:n.; the overlap of p with IcI>~) remains unchanged.
(iii) Taking two entangled pairs in this depolarized form, both A and B perform
the generalized XOR gate XORN : Ik) Il) t--+ Ik) I(l + k)modN) on their
respective systems. Then both measure the state of their second system in
the basis Ik). The first pair is kept, if they get the same result otherwise it
is discarded (as the second pair always is). The resulting state has a density
matrix p', which has a larger overlap with the maximally entangled state 1cI>~)
than the original p. Iterating the last two steps sufficiently often, the overlap
between the resulting state and 1cI>~) approaches 1, that is, the distilled stated
converges to the maximally entangled state 1cI>~).
(15.13)
Let P be a symmetric Gaussian state in standard form and 'I/; the pure two-
mode squeezed state 1'1/;) = co;h T l:n tanh n r Inn). This is itself a Gaussian
state and the four parameters (15.9) are na = nb = cosh r, kx = -kp = sinh r.
In the limit of large r (keeping only the leading terms in eT ) Ineq. (15.13)
becomes after some simple algebra
(15.14)
But Ineq. (15.14) is - for symmetric states - implied by the inseparability
criterion: in that case, (15.11) simplifies to In 2 - kxkp - 11 < n(k x - kp),
which implies Ineq. (15.14), proving that all symmetric inseparable Gaussian
states are distillable.
If the state is not symmetric, it means that the reduced state at one of the
two sides has larger entropy than the other. This suggests to let a pure state
interact with the "hotter" side to cool it down. To do this without destroying the
Distillability and entanglement purification for Gaussian states 179
(Na,Nb,Kx,Kp) = (nb,na,-kp,-kx)/JiMT·
A state in the standard form (15.8) can be brought into the Wigner standard
form by local squeezing operations.
Now assume that Nb < N a, i.e., B is the hotter side. B takes an ancilla
mode in the vacuum state and couples it to its mode using a beam splitter with
transmittivity cos 2 (}. After a homodyne measurement of the ancilla results a
state p with Wigner correlation matrix M with2
M _.!. ( c2 Na + 8 2 Dx 0 )
A - v 0 c2 Na + 8 2 NaNb '
M _ .!. ( cKx 0 )
AB - v 0 cKpv '
Checking (15.11) for M one easily sees that the sign of the left-hand side does
not change; therefore the transformed state is inseparable iff the original one
was inseparable. It remains to be shown that there always exists a (} to satisfy
(15.15), i.e., that the right hand side of Eq. (15.15) is positive. The numerator
is positive since Nb < N a, the denominator is positive for all states since the
second part of condition (15.10) implies that (Na - DxNb) > 0 and the first
part assures that (Na - DxNb)(Nb - DpNa ) ;::: (NaKx + NbKp) 2 ;::: 0, hence
all Gaussian states can be symmetrized this way. But since every Gaussian state
can be brought into Wigner standard form by local unitaries, this completes the
proof. _
180 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
This implies that the protocol of Ref. [12] can in principle be used to obtain
maximally entangled states in a finite dimensional Hilbert space from any given
inseparable Gaussian two-mode state.
Before discussing EPPs in more detail in the following, we end this section
on distillability by mentioning an interesting recent result. The equivalence of
inseparability and distillability of Gaussian states holds only for the case of two
modes. Werner and Wolf showed in [19] that if both A and B possess more that
one mode, there exist Gaussian states that are bound entangled, namely states
that are entangled but whose density matrices remain positive under partial
transposition. In later work [23] it was shown that for any number of modes
bipartite Gaussian states are distillable iff their partial transpose is not positive.
Each iteration of these two steps brings the state closer to a maximally entan-
gled state in the Hilbert space of dimension N, where (N + 1)2 is the last result
of the total photon number measurement. Hence with finite probability one can
get arbitrarily close to a maximally entangled state in any finite dimensional
space provided the initial supply of states P is sufficiently large.
states from a relevant class of mixed Gaussian states which result from losses
in the light transmission. Furthermore, we propose and analyze how to imple-
ment this protocol experimentally using high finesse cavities and cross-Kerr
nonlinearities. We begin by describing the entanglement purification protocol
for pure two-mode squeezed states, and then extend the protocol to include
mixed Gaussian CV states, and describe a physical implementation in the next
subsection.
For pure states First assume that we have generated m entangled pairs Ai, Bi
(i = 1,2,··· m) between two distant sides A and B. Each pair of modes Ai, Bi
are prepared in the pure two-mode squeezed state I'll) AiBi' which in the number
basis has the form
00
where.x = tanh (r) ,and r is the squeezing parameter [18]. The entanglement
BJ
E (I'll) Ai of the two-component state (15.16) is uniquely quantified by the
von Neumann entropy of the reduced density operator of one subsystem. The
joint state I'll) (AiBd of the m entangled pairs is simply the product of all the
I\lI) Ai Bi' which can be rewritten as
1'lI)(AiBd = (1 - .x2) T f J
j:=:O
.xj fjm) 1J)(AiBd ' (15.17)
. f(m).
The fu nctIOn j m E q. (15 . 17) an d (15 . 17)·IS gIven
. b y f(m)
j = (j+m-l)!
j!(m-l)! .
Note that the state Ij) (A;Bi} represents a maximally entangled state in the
subspace corresponding to a local photon number of j at both sides. To con-
centrate the entanglement of these m entangled pairs, we perform a QND
measurement of the total photon number N Al + N A2 + ... + N Am on the A
side (we will describe later how to implement this measurement experimen-
tally). The QND measurement projects the state 1'lI)(AiBd onto the two-party
maximally entangled state Ij)(Ai B ,} with probability
0.3 0.3
O.ZI 0.2.\
0.2 0.2
c:
.
O.IS 0.1.
0.1 0.\ ,
,
0.05 0.01
0.'
Figure 15.1 The purification success probability versus entanglement increase ratio for two
(left) and four (right) pairs. Dotted line for the squeezing parameter r = 0.5, dashed line for
r = 1.0, and solid line for r = 1.5.
rj = E (lj)(AiBd) IE (IW)AiBJ
defines the entanglement increase ratio, and ifrj > 1, we get a more entangled
state. Even with a small number m, the probability of obtaining a more
entangled state is quite high. Figs. 15.1a,b show the probability of achieving an
entanglement increase ratio r j for various values of initial entanglement and
initial number of pairs.
To measure how efficient the scheme is, we define the entanglement transfer
efficiency Y with the expression
Y= f ______
p;m) E (Jj) (AiBd)
~J=_o ______
~
(15.19)
mE (IW)AiBJ
It is the ratio of the average entanglement after concentration measurement
to the total initial entanglement contained in the m pairs. Obviously, Y :S 1
should always hold. With the squeezing parameter r = 0.5,1.0 or 1.5, the
entanglement transfer efficiency versus the number of pairs m is shown in Fig.
15.2.
From this figure, we see that the entanglement transfer efficiency is near to
1 for a large number of pairs. In fact, it can be proven that as m -+ 00, we
would get with unit probability a maximally entangled state with entanglement
mE (I w) Ai Bi ). To show this, we calculate the mean value and the variance of
184 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
0.9
0.8 ...
0.7
0.6
0.5
0.4
0.3
0.2
0.1
10 12 14 16 18 20
m
Figure 15.2 The entanglement transfer efficiency versus the number of pairs m in simultaneous
concentration. Dotted line for r = 0.5, dashed line for r = 1.0, and solid line for r = 1.5.
(IS.20)
(1 - ).2)2'
so the entanglement transfer efficiency tends to unity. This proves that the
purification method described above is optimal in the asymptotic limit (m -+
(0), analogous to the purification protocol presented in [S] for the qubit case.
For any finite number of entangled pairs, this purification protocol is more
efficient than that in [S], since it takes advantage of the special relations between
the coefficients in the two-mode squeezed state.
An interesting feature of this entanglement purification protocol is that for
any measurement outcome j of. 0, we always get a useful maximally entangled
state in some finite Hilbert space, though the entanglement of the outcome state
I)) (AiBd does not necessarily exceed that of the original state I\II) A,B, if j is
small.
where p is the density operator of the m entangled pairs with p (0) = 1\l1)(Ai Bd
(\l1I, the pure two-mode squeezed state (15.16), and the effective Hamiltonian
(15.23)
In Eqs. (15.22) and (15.23), aai denotes the annihilation operator of the mode
ai, (a = A or B), and we have assumed that the damping rates 17 A and 17Bare
the same for all the m entangled pairs, but 17A and 17B may be different from
each other. Eq. (15.22) describes a situation in which single photon absorption
is the only relevant source of noise as is typically the case at optical frequencies.
Small Noise In many practical cases, it is reasonable to assume that the
light transmission noise is small. Let T denote the transmission time, then
17AT and 17BT are small factors. To the first order in 17AT and 17BT the fi-
nal state of the m entangled pairs is in the language of quantum trajectories
[18] either 1\l1(O))(Ai B;} DC e-iHeffT 1\l1)(Ai B;} (no quantum jumps occurred)
or 1\l1(a;))(Ai Bd DC y'17aTaai 1\l1)(Ai Bd (a jump occurred in the ai channel
(a = A,B andi = 1,2,··· ,m). The final density operator is a mixture of all
these possible states. To distill entanglement from this mixed state, we perform
QND measurements of the total photon number on both sides A and B, with
results jA and jB, respectively. We then compare jA and jB through classical
communication (eC), and keep the outcome state if and only if jA = jB.
Let pJ) and piP
denote the projections onto the eigenspaces of the corre-
m m
sponding total number operators L a~aAi and L a1.aBi with eigenvalue
i=l' i=l'
j, respectively. It is easy to show that
So if j A = jB = j, the outcome state is the maximally entangled state Ij) (Ai B;}
with entanglement log (tjm)). The probability to get the state jj) (AiBd is
now given by pj = (1- ,\2)m ,\2j fjm)e-('TJA+'TJB)Tj. It should be noted that
the projection operators pi!) piP cannot eliminate the states obtained from the
186 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
initial state IW)(AiBd by a quantum jump on each side A and B. The total
probability of this kind of quantum jumps is proportional to m 2n 2'T1A'TIBT2. So
the condition for small transmission noise requires m 2n 2'T1A'TIBT2 « 1, where
n = sinh2 (r) is the mean photon number for a single mode.
Asymmetric Noise In the purification of mixed entanglement, we need CC to
confirm that the measurement results on both sides are the same, and during this
CC, we implicitly assume that the storage noise for the modes is negligible.
In fact, that the storage noise is much smaller than the transmission noise
is a common assumption made in all the entanglement purification schemes
which need the help of repeated CC [3]. If we make this assumption for
continuous variable systems, there exists another simple configuration in which
the purification protocol works. Let the two-mode squeezed states be generated
at side A. After state generation, we keep the modes Ai on side A with a very
small storage loss rate 'TIA, and at the same time the modes Bi are transmitted
to the distant side B with a loss rate 'TIB » 'TIA. We call this a configuration
with an asymmetric transmission noise. In this configuration, the purification
protocol is exactly the same as that described in the above paragraph. We note
that the component in the final mixed density operator which is kept by the
projection P)1) pijl should be subject to the same times of quantum jumps
on each side A and B. We want this component to be a maximally entangled
state. This requires that the total probability of the same nonzero number of
quantum jumps on both sides to be very small. Clearly, this total probability is
always smaller than n'TIAT, no matter how large the damping rate 'TIB is. So the
working condition of the purification protocol in the asymmetric transmission
noise configuration is given by n'TIAT « 1. The loss rate 'TIB can be large. The
probability to obtain the maximally entangled state Ij) (Ai B ;} is still given by
pj = (1 - ,\2)m ,\2j fjm)e-(TJA+TJB)Tj.
A-posteriori Purification For continuous variable systems the assumption
of storage with a very small loss rate is typically unrealistic. In that case, we
can use the following simple method to circumvent the storage problem. Note
that the purpose of distilling maximally entangled states often is to directly use
them in some quantum communication protocol, such as quantum cryptography
or quantum teleportation. So we can modify the above purification protocol
by the following procedure: immediately after the state generation, we make a
QND measurement of the total photon number on side A (measurement result
jA). Then we do not store the resulting state on side A, but immediately
use it (e.g., perform the corresponding measurement as required by a quantum
cryptography protocol [16]). During this process, the modes Bi are being sent to
the distant side B, and when they arrive, we make another QND measurement
of the modes Bi and get a outcome j B. The resulting state on side Bean
be directly used (for quantum cryptography, for instance) if j A = j B, and
Distillability and entanglement purification for Gaussian states 187
A B
Figure 15.3 Schematic setup for generating Gaussian entangled states between two distant
cavities.
KH------H
KI+-......... II
Figure 15.4 Schematic experimental setup to measure the total photon number NI + N2
contained in the cavities I and II. The cavities I and II, each with a smaIl damping rate K, and
with a cross Kerr medium inside, are put respectively in a bigger ring cavity. The ring cavities
with the damping rate 'Yare used to enhance the cross Kerr interactions. A strong continuous
coherent driving light bi! (t) is incident on the first ring cavity, whose output bOI is directed to
the second ring cavity. The output bo 2 (t) of the second ring cavity is continuously observed
through a homodyne detection.
medium, and thus is negligible [33, 34]. In the frame rotating at the optical
frequency, the Langevin equations describing the dynamics in the two ring
cavities have the form
with the boundary conditions (see Fig. 1) bi2 = b01 = b~l + g..,fY + ..,fYb 1
and bo2 = bi2 +..,fYbz. In the realistic case "( » X (Ni ) , (i = 1, 2), we
can adiabatically eliminate the cavity modes bi, and express the final output
bo2 of the second ring cavity as an operator function of the observable Nl +
N 2 . The experimentally measured quantity is the integration of the homodyne
photocurrent over the measurement time T. Choosing the phase of the driving
Distillability and entanglement purification for Gaussian states 189
XT
(15.26)
(15.27)
This condition seems to be feasible with the present technology. For ex-
ample, if we assume the cross-Kerr interaction is provided by the resonantly
enhanced Kerr nonlinearity as considered and demonstrated in [33, 34], the
Kerr coefficient X/27r '"" O.lMHz would be obtainable4 . We can choose the
decay rates K,/27r '"" 4MHz and, /27r '"" 100MHz, and let the dimensionless
factor 9 '"" 100 (for a cavity with cross area S '"" 0.5 x 1O-4 cm2 , 9 '"" 100
corresponds a coherent driving light with intensity about 40mWcm- 2 ). The
mean photon number (N1 ) = (N2 ) = sinh2 (r) '"" 1.4 for a practical squeez-
ing parameter r '"" 1.0. With the above parameters, Eq. (15.27) can be easily
satisfied if we choose the measuring time T '"" 8ns. More favorable values for
the parameters are certainly possible.
To bring the above proposal into a real experiment, there are several imper-
fections which should be considered. These include phase instability of the
driving field, imbalance between the two ring cavities, light absorption in the
Kerr medium and the mirrors, self phase modulation effects, light transmission
loss between the ring cavities, and inefficiency of the detectors. To realize
a QND measurement, the imperfections should be small enough. We have
deduced quantitative requirements for all the imperfections listed above [31].
190 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
With the parameters given in the above paragraph, all these requirements can
be met experimentally.
Acknowledgments
G.G. acknowledges financial support by the German Friedrich-Naumann-
Stiftung. This work was supported by the Austrian Science Foundation under
the SFB "Control and Measurement of Coherent Quantum Systems" (Project
11), the European Union under the TMR network ERB-FMRX-CT96-0087
and the project EQUIP (contract IST-1999-11 053), the European Science Foun-
dation, and the Institute for Quantum Information GmbH, Innsbruck.
Notes
1. S is called symplectic if ST JS = J.
2. See [35] for the general formalism to describe measurements on Gaussian states.
3. To be more precise: this equivalence holds on the infinite dimensional space, when XOR:ln, m) >-+
In, m + n). For states in a N dimensional subspace (as obtained after the first step) this equivalence is only
true for measurement outcomes NOt. :os: N
4. In fact, Ref. [33] considered a configuration, yielding a Kerr coefficient X ~ lOOMHz, to realize
a single-photon turnstile device. But the estimation there puts a stringent limit on the required cavity
parameters [K. M. Gheri et at., Phys. Rev. A 60, R2673, 1999]. We take a much more moderate estimation
of the relevant parameters and find X/27r rv O.lMHz is obtainable. This value of the Kerr coefficient
is large enough for performing the QND measurement, though it is certainly not enough for realizing a
single-photon turnstile device.
References
[3] N. Gisin, Phys. Lett. A 210, 151 (1996); C.H. Bennett, G. Brassard, S.
Popescu, B. Schumacher, J.A. Smolin, W.K. Wootters, Phys. Rev. Lett.
76, 722 (1996).
[4] C.H. Bennett, D.P. DiVincenzo, J.A. Smolin, W.K. Wootters, Phys. Rev.
A 543824 (1996).
[5] C.H. Bennett, H.J. Bernstein, S. Popescu, B. Schumacher, Phys. Rev. A
53,2046 (1996).
[6] H.-J. Briegel, W. Dur, J.I. Cirac, P. Zoller, Phys. Rev. Lett. 81, 5932
(1998).
Distil/ability and entanglement purification/or Gaussian states 191
[29] T. Opatrny, G. Kurizki, D.-G. Welsch, Phys. Rev. A 61, 032302 (1999).
[30] L.-M. Duan, G. Giedke, J.I. Cirac, P. Zoller, Phys. Rev. Lett. 84, 4002
(2000).
[31] L.-M. Duan, G. Giedke, J.I. Cirac, P. Zoller, Phys. Rev. A 62, 032304
(2000).
[32] A. S. Parkins, H. J. Kimble, Phys. Rev. A 61, 052104 (2000).
[33] A. Imamoglu, H. Schmid, G. Woods, M. Deutsch, Phys. Rev. Lett. 79,
1467 (1997); 81,2836 (1998).
[34] L.v. Hau, S.E. Harris, Z. Dutton, C.H. Behroozi, Nature 397,594 (1999).
[35] G. Giedke and J. I. Cirac, quant-phl0204085.
Chapter 16
ENTANGLEMENT PURIFICATION
VIA ENTANGLEMENT SWAPPING
S. Parker
Optics Section, The Blackett Laboratory, Imperial College
London, England, SW72BW
[email protected]
S. Bose
Centre for Quantum Computing, Clarendon Laboratory,
University of Oxford, England, OX] 3DU
[email protected]
M. B. Plenio
Optics Section, The Blackett Laboratory, Imperial College
London, England, SW72BW
[email protected]
1. INTRODUCTION
The experimental [1, 2, 3, 4] and theoretical [5] realization of teleportation
in continuous variable systems has demonstrated that entanglement between
separated systems is essential in order to be able to perform teleportation and
193
S.L. Braunstein and A.K. Pati (eds.), Quantum Information with Continuous Variables, 193-209.
© 2003 Kluwer Academic Publishers.
194 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
continuous analogue of the former single "particle" gate is the Fourier trans-
form:
whose inverse is obtained by replacing the + with a - sign on the right hand
side.
We can now define the 'entangling' operation and its inverse:
(16.4)
i:
C12F1I Ga(x1))1IG,6(X2))2
exp [:2 ( _x 2a 2 -
IX)l Ix + y + X2)
~: + 2iX1X) ]
dxdy
IB a,6(x1' X2))12' (16.5)
196 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
. "~'"
0.08
. '"
-":"'" ...
... ,....
0.06
W 0.02
-0.02
-0.04
-0.06
10
10
x
Figure 16.1 Wigner function for a Schrbdinger cat state, 1/V2 (I+a) + I-a» with coherent
states la) = 13).
Our generalizations we will call two-mode cat states [16, 17], as they are
Schrodinger cat states of two modes whose locations are correlated with each
Entanglement purification via entanglement swapping 197
jC(d))12 = i:
other quantum mechanically:
[Aoe-CX-d)2_CY+d)2
1 f;00
-00
1 2
[A j e-(X-C-l)i d) -(y+C-l)id) ]
2
The complex coefficients, A j , are such that jAoj2 + jAlj2 = 1 (so the state
is not normalized correctly). This state is a superposition of the first particle
being located around d and the second around -d and vice versa. The scale
length does not appear here (it is set to unity) as an increase in scale length is
equivalent to a decrease in the value of d.
3. QUANTIFICATION OF ENTANGLEMENT
The measure of entanglement for pure bipartite states, the entropy of entan-
glement, E(PI2), is just the Von Neumann entropy, S, of either partial density
operator, PI or P2, of the system [6]
(16.7)
where the Ai'S are the eigenvalues of PI or P2. We can very simply generalize
this to continuous variable systems but we need to know a little about the
eigenvalues of such systems. Let us express a continuous variable state as
follows
j'l/Jh2 = J
'l/J(x, y) jxh jY)2 dxdy, (16.8)
and find the partial density operator of one of the particles (the first, say) by
tracing out the other
(16.10)
198 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
only one of these, tracing out the second particle. The integral eigenvalue
equation corresponding to this partial trace (aside from normalization) is
! [1(
exp a 2 -
(2
a 1)
+ 2(32 (x 2 + X ,2 )
It can then be shown [9] that this eigenvalue equation is equivalent to (in that it
has the same set of eigenvalues, aside from normalization) the following one
K(x,x')
)..¢(x) (16.12)
where K(x, x') is the kernel of our integral eigenvalue equation representing
the elements of the density operator and
(16.13)
Notice that the eigenvalues are independent of the scale length, a, and the
value of Xl or X2. It is only dependent on the product of widths of the original
Gaussian distributions with respect to a.
Similarly we calculate the partial trace of the two-mode cat state and find
that it's corresponding integral eigenvalue equation is
¢(X') dx'
v
K(x,x')
)..¢(x).(16.14)
where the outer sum is over the set of eigenvalues and the sums over s and t
are to normalize the set of eigenvalues as the Ppq is not normalized.
What we are doing here is sampling the spectra of eigenvalues over discrete
ranges. If we were to take 8 -+ 0 and n -+ 00 we should converge to the exact
value of the entropy of entanglement as in Eq. (16.7). In fact, for these states
there is an analytical result for the entropy of entanglement [3] which we will
discuss in the next section. In practice for most values of a. and (3, 2n + 1 = 201
and w around 10 standard deviations from the mean were sufficient to produce
results in agreement with the analytic result accurate to 6 significant figures.
The numerical results for the entanglement are shown in figure 16.2 for
varying values of a. and (3. They were generated using numerical procedures
for eigenvalue problems from the NAG library.
We expect the entanglement to increase when the parameters a. and (3 are
reduced as this corresponds to a reduction in the spread or uncertainty in
the wavefunction of the two particles before the entangling operations were
performed. As these parameters approach zero the states become like the
entangled states of Vaidman [15]. These are simultaneous eigenstates of the
operators Xl + X2 and ih - 'P2 and are therefore maximally correlated. The
entanglement is then infinite and convergence in the numerical procedure is
difficult to achieve for these small values of a. and (3.
5,------,-------,------,------,-------,------,
4.5
3.5
3
E
Q)
E
Q)
c;, 2.5
~ 2
1.5
Figure 16.2 Entanglement of a partially correlated state in terms of a and (3, the widths of the
Gaussians from which they are formed.
After taking the partial trace of this state the integral eigenvalue equation can
also be transfonned into the fonn of Eq. (16.11) with a parameter [9]
This tells us a number of things, firstly that our partially correlated states are
just a generalization of two-mode squeezed states and secondly, as the state can
be written analytically in the number basis [3]
1 00
(16.21)
This in tum gives us an analytical result for our partially correlated states via
the substitution 2r = arcsinh(1/ a,B) into Eq. (16.21). This follows directly
from Eqs. (16.19) and (16.13).
202 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
. ~ ..
'
1.5
o 0
d
Figure 16.3 Entanglement of the cat states against (half) the distance between Gaussians, d,
and the coefficient Ao. The entanglement is greatest for high values of d where the Gaussians
become orthogonal, and for IAol2 = 0.5 as with discrete entanglement.
Note that the entanglement for any given value of d is maximum when
lAo 12= 1/2 and that the entanglement increases with d for given values of Ao,
approaching the limit E = -lAo 1210g2 lAo 12 - IAll210g21All2 as d ~ 00,
where the separated Gaussians become orthogonal. Note also that only two
eigenvalues dominated the contribution to the entropy. These observations,
therefore, tell us that these states behave very much like discrete 2-level entan-
gled systems.
Entanglement purification via entanglement swapping 203
(16.22)
and 11/1) 34. Then a Bell state measurement on particles 2 and 3 will, for certain
measurement results, leave particles 1 and 4 in states
(16.23)
with probability (a 4 + (34)/2, which is less entangled (provided the two pairs
are not already maximally entangled (a = (3 = 1/v'2», and states
(16.24)
• ~
....
Bell state projection •
_------------I initial
•
Alice Bob
• • final
• •
Figure 16.4 The entanglement swapping procedure. Bob, holding a copy of the entangled
state shared by himself and Alice, performs a Bell state measurement on a particle from each
pair and, for certain measurement outcomes, the entanglement of the final shared pair is higher
than that of the initial shared pair.
where
(16.26)
Has the entanglement increased? Again we can take the partial trace of this
state and transform the kernel into the form of Eq. (16.11). This gives us a
parameter Pswap = 2[(a 4 + a 2,82 + 1)(,84 + ,82a2 + 1) - 1] where before
the swapping process the parameter was Po = 2a 4 or 2,84. Note firstly that
P swap does not depend on the measurement results a and b and is therefore not
probabilistic so any increase in entanglement would be deterministic thereby
breaking laws of the conservation of entanglement. However Pswap ~ Po and
the entanglement is strictly decreasing with increase in P so the swapped pair
has less entanglement and we have not achieved purification.
Entanglement purification via entanglement swapping 205
Of course, our final projections in this method were onto the unphysical states
la) and Ib) but further calculations indicate that with finite width projections
(performed by projecting onto the Gaussian states of Eq. (16.4» the parameter
P still increases. Such calculations involve 6th degree polynomials in the
width parameters (O! and (3 etc.) so proving that P increases for all values of
these parameters is difficult and we have not been able to do so analytically.
However, numerical results indicate that this is true.
= f ,2: 1
A j A k e-(x-(-1)jd)2_(Y+(-1)k d)2
J,k=O
x e( dbh(tt) ( (-l)j +(1+1.12)( _l)k )+2d20jk)
x e(iadh(tt)((1+tt 2)(-1)L(-1)k)) IX)1IY)4 dxdy (16.27)
where
(16.28)
Writing this state out in full in the high precision measurement limit, I-' = 0
e-2d2+2db e-(x-d)2_(y+d)2
11/J}14 = / (AoAo
e-(x-d)2_(y-d)2
+AoA1 e 2iad
e- 2iad e-(x+d)2_(y+d)2
+AIAo
e-2d2-2db e -(X+d)2_(y-d)2)
+A1A1
IX)l IY)4 dxdy (16.29)
and looking at the particular case where the probabilistic measured values are
a = b = 0 we can see purification for high values of d as the middle two
terms now dominate and have coefficients of equal magnitude. As d -+ 00
they become maximally entangled. This is again very much like the action
of discrete entanglement under purification procedures: the coefficients of the
states have changed, not the states themselves.
For the results of figure 16.5 and 16.6 we have chosen the values Ao = y'Q.3
and d = 1.0. They show the entanglement of the resulting state (16.26) for a
range of values of a and b with I-' = 0 and 0.5 respectively.
206 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
o.g
0.7
.,c
E
~0.6
c
:l!
oll 05
0.4
0.3
0.2
-0.5
-0.5
0.5 -1
b
a/x
Figure 16.5 Entanglement of swapped cat states with J-t = O. Above the level of the plane
purification has been achieved.
The horizontal planes are at the level of entanglement of either cat state before
the purification procedure is performed, that is, above this plane purification
has been achieved.
6. SUMMARY
We should now address why it is that we have observed purification via
entanglement swapping in only one of the sets of states that we have considered.
As mentioned above the two-mode cat states have many characteristics like
those of discrete two-level entanglement and so it is not surprising that the
method of entanglement swapping generalized from these systems is successful.
The failure of the procedure for the partially correlated states is more difficult
to explain. These states are the kind of states used in continuous variable
teleportation experiments [1, 2, 3, 4, 5] and a simple method of purification
would be ideal. However, we were unable to find any simple continuous
procedure that would produce purification, indeed no procedure was found
where the final entanglement was in anyway probabilistic in the measurement
outcomes, an essential ingredient if a successful procedure is not to violate
conservation laws of entanglement. In fact it has been brought to our attention
Entanglement purification via entanglement swapping 207
0'9h~~~
o,e
0.7
c
~ 0.6
0.
§c: 0.5
w
;,. :: . ......... .
0.5
-0.5
o
0,5 _1
b
a/x
Figure 16.6 Entanglement of swapped cat states with p, = 0.5. Again purification has been
achieved above the level of the plane, but with the inaccuracy in the measurement part of the
entanglement swapping process the amount of purification is reduced.
[27] that such methods could never increase the amount of entanglement for
Gaussian states. When completing this chapter it was found rigorously, that
Gaussian operations never allow entanglement distillation of Gaussian states,
see Ref. [28].
Fortunately, more recent work has now achieved entanglement purification
in Gaussian states [10] using non-demolition measurements with a number
of entangled pairs. However, what exactly the key difference is between the
two types of states presented here which allows purification by our methods in
one class but not the other is still unclear. There are obvious correspondences
between the form of entanglement in the two mode cat states and discrete
systems and it would be interesting to find a condition for continuous variable
purification, as it has been attempted here, which a state undergoing purification
must obey. The fact that purification has been demonstrated here and elsewhere
in continuous systems, however, are interesting results.
208 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Acknowledgments
This work is supported by the United Kingdom Engineering and Physical
Sciences Research Council (EPSRC), the Inlaks Foundation, The Leverhulme
Trust, the EU TMR-networks ERB 4061PL95-1412 and ERB FMRXCT96-
0066, the EU project EQUIP and the European Science Foundation programme
on quantum information theory.
References
[1] S. L. Braunstein, H. J. Kimble, Phys. Rev. Lett. 80, 869 (1998).
[2] G. J. Milburn, S. L. Braunstein, Phys. Rev. A 60, 937 (1999).
[3] SJ. van Enk, Phys. Rev. A 60,5095 (1999).
[4] T. C. Ralph, P. K. Lam, Phys. Rev. Lett. 81, 5668 (1998).
[5] A. Furusawa, J. L. Sj2Irensen, S. L. Braunstein, C. A. Fuchs, H. J. Kimble,
E. S. Polzik, Science, 282, 706 (1998).
[6] C. H. Bennett, H. J. Herbert, S. Popescu, B. Schumacher, Phys. Rev. A 53,
2046 (1996); S. Popescu, D. Rohrlich, Phys. Rev. A 56, R3319 (1997).
[7] C. H. Bennett, D. P. DiVincenzo, J. A. Smolin, W. K. Wootters, Phys.
Rev. A 54,3824 (1996).
[8] M. B. Plenio and V. Vedral, Contemp.Phys. 39,431 (1998)
[9] S. Parker, S. Bose, M. B. Plenio, Phys. Rev. A 61,032305 (2000).
[10] Lu-Ming Duan, G. Giedke, J.I. Cirac, P. Zoller, Phy~. Rev. A 62, 032304
(2000); Lu-Ming Duan, G. Giedke, J.I. Cirac, P. Zoller, Physical imple-
mentation for entanglement purification of Gaussian continuous variable
quantum states, LANL eprint: quant-ph 0003116.; Geza Giedke, Lu-Ming
Duan, J. Ignacio Cirac, Peter Zoller, All inseparable two-mode Gaussian
continuous variable states are distillable Lanl e-print quant-ph/0007061
[11] M. Lewenstein and P. Horodecki, lanl e-print quant-ph/OOO1035
[12] H. -K. Lo, S. Popescu, Concentrating entanglement by local actions -
beyond mean values, LANL eprint: quant-ph 9707038; M. A. Nielsen,
Phys. Rev. Lett. 83, 436 (1999); G. Vidal, Phys. Rev. Lett. 83, 1046
(1999);D. Jonathan, M. B. Plenio, Phys. Rev. Lett. 83, 3566 (1999); L.
Hardy, Phys. Rev. A 60, 1912 (1999).
[13] M. Zukowski, A. Zeilinger, M. A. Home, A. K. Ekert, Phys. Rev. Lett.
71,4287 (1993); J -W. Pan, D. Bouwmeester, H. Weinfurter, A. Zeilinger,
Phys. Rev. Lett. 80, 3891 (1998); S. Bose, V. Vedral, P. L. Knight, Phys.
Rev. A 57,822 (1998).
[14] S. Braunstein, Phys. Rev. Lett. 80,4084 (1998).
[15] L. Vaidman, Phys, Rev. A 49, 1473 (1994).
Entanglement purification via entanglement swapping 209
Abstract We discuss the notion of bound entanglement (BE) for continuous variables (CV).
We show that the set of non-distillable states (NDS) for CV is nowhere dense
in the set of all states, i.e., the states of infinite-dimensional bipartite systems
are generically distillable. This automatically implies that the sets of separable
states, entangled states with positive partial transpose, and bound entangled states
are also nowhere dense in the set of all states. All these properties significantly
distinguish quantum CV systems from the spin like ones. The aspects of the
definition of BE for CV is also analysed, especially in context of Schmidt numbers
theory. In particular the main result is generalised by means of arbitrary Schmidt
number and single copy regime.
1. INTRODUCTION
Bound entanglement [1] is the entanglement which cannot be distilled (pu-
rified), i.e. no pure state entanglement can be obtained from it by means of
local operations and classical communication (LOCC)[2]. So far, it has been
studied mainly for spin like systems. These studies has allowed to discover
many interesting properties of bound entanglement, for both bipartite[3], and
multiparticle systems[4]. Recently, much attention has been devoted contin-
uous variable (CV) systems (c.f. [5]). Bound entanglement has also been
considered for continuous variables (CV), and the first nontrivial examples of
BES for CV have been constructed[6] (see also [7]). Once we have some ex-
amples of BES for CV, it is interesting to ask how frequent is the phenomenon
211
S.L. Braunstein and A.K. Pati (eds.), Quantum Information with Continuous Variables, 211-228.
© 2003 Kluwer Academic Publishers.
212 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
of bound entanglement, i. e. how many states of that kind are in set of all CV
states?
The question of "how many quantum states having some interesting property
are there?" is very natural. In the context of entanglement it was first considered
in Ref. [8], where the problem of the volume of the subset of separable (non-
entangled) states in the set of all bipartite states of spin systems was considered.
Numerical evidence has shown that the volume of the set of separable states
approaches zero when the size of the spin goes to infinity. It was also shown
that for any finite spin system the volume of separable states is nonzero due to
the existence of a separable neighborhood, i.e. an open ball of separable states
in the vicinity of the maximally mixed state in arbitrary dimension. Further first
analystical bounds on the size of neighbourhood have been provided [9]. All
this raised a series of questions concerning the interpretation of experiments
of quantum computing based on high temperature NMR; many interesting
analyses have been performed in this context [to, 11].
The question of the "size" of the set representing separable states has been
recently answered [12] for CV; it has been show that for bipartite states this
subset is nowhere dense (relative to the trace-norm topology). This implies that
this set does not contain any open ball and also that CV states are generically
non-separable. On the other hand there exists another subset that is of interest
in the context of entanglement. This is the subset of of non-distillable states
(NDS), i.e. states that cannot be distilled. This subset contains the separable
states, and therefore it might well be that an appreciable fraction of all states
are in such a subset. In this paper we show that this is not the case; that is, the
subset of NDS is nowhere dense in the set of bipartite states of Cv. We present
two different proofs of this fact. One uses the uniform topology, and the other
one the trace-norm topology.
We also perform analsysis how one can relax conditions of NDS in context
of CV in comparison with the standard definition, and prove stronger version
of the main result with help of Schmidt numbers theory [13] and single copy
regime (see [15, 14]).
There are several results which follow from our proofs. In particular, since
the subset of NDS contains the subset of BES, we have that that subset is also
nowhere dense. The same thing occurs with the subset of states with positive
partial transpose (PPT) [1] and therefore with with those PPT states that are
entangled. Moreover, since the subset of separable states is also contained in
the one of NDS, our results include the ones given in Ref. [12].
Bound entanglement/or continuous variables is a rare phenomenon 213
(17.1)
(17.2)
where -1/2 :S Ak :S 1, i.e. IAk I :S 1 and the I¢k) are also in the same 2 x 2
subspace.
(17.5)
Bound entanglement/or continuous variables is a rare phenomenon 215
and the fact that both Ilpll and IIp'll are smaller than one. Thus, if", < lEI!4n,
we see that for any p' E B'T/ (p),
p = LPnlwn)(wnl, (17.7)
n=1
where we have chosen PI 2: P2, .... Note that since p is a trace class operator,
the sequence Pn converges monotonically to zero. On the other hand, we can
write the Schmidt decomposition of each Iw n ) as
00
where again we have chosen An,k 2: An,k+l 2: 0 and An,k converges monoton-
ically to zero as k -+ 00. Now, we define
N
PN == LPnIWN,n)(WN,nl, (17.9)
n=1
where
N
IWN,n) = L VAn,klun,k,Vn,k). (17.10)
k=1
(17.13)
The map <I> : S --+ 7 is continuous (in the norm II . liT) and onto. In particular
it maps dense subsets onto dense subsets (see [12] for explanation).
Consider the set X of all vectors ul23 = A®I ®I(VI23) for all A E B(1lI).
The vector Vl23 is called i-cyclic (see [12]) if the closure of X in the norm II·IIT
turns out to be the whole space 111 ® 112 ® 1l3. The physical interpretation of
I-cyclic vectors in both finite dimensional, as well as in the CV case, is that
those are the vectors which have maximal possible Schmidt rank. Note, that
according to Lemma 2 of Ref. [25] they form a dense set in 111 ® 112 ® 1l3.
Now we consider the following simple
Observation 1.- Let the set NV be (i) a proper closed (in II . liT norm)
subset of the set of states 7 which is (ii) invariant under the operations A ® I.
Bound entanglement/or continuous variables is a rare phenomenon 217
defined for arbitrary A, such that Ilv'll = 1. We shall show first that ~(v') also
belongs to NV. Indeed (see [12]) we have ~(v') = A ® 1~(v)At ® I and
(because the norm of v' is one) the trace ~(v') is one. But, because the set NV
is closed under the operation A ® 1(·)At ® I, we see that ~(v') still belongs
to the set.
Now suppose that v were I-cyclic. Then, that the set M of all vectors v'
would be dense in the unit sphere S of all normalized vectors belonging to
1i ® 1i3. As the map ~ is continuous and onto, it certainly would map M onto
some new set denoted by ~(M), which would be dense in set of all bipartite
states T. Thus closure of ~(M) must have give all T. But, on the other hand
any element of ~ (M) (which is defined as ~ (v') for some vector v' of the form
(17.14» belongs to NV. As the latter is closed, the closure of ~(M) would
have to be a subset of NV. But Nv was supposed to be closed and strictly
smaller than the set T, so the closure of ~(M) cannot be equal to T. This
gives the required contradiction. The above reasoning follows the lines of the
proof of Ref. [12]. The only difference is that instead of the specific set of
separable states considered there, here we have considered an abstract set NV,
which has some special properties. Note, that the assumptions of Observation
1 and the fact that the I-cyclic vectors form a dense set in S imply that the
closed set NV is nowhere dense. If it had contained a open set, then, following
continuity of ~ this open set would have had to be an image of open subset
of 1i 1 ® 1i2 ® 1i3, which would have had to contain a ball, an thus a I-cyclic
vector. Now, to show that the set of NDS states is nowhere dense we have to
show that it is (i) invariant under local operations of the type A ® I, (ii) closed
in the trace norm II . liT. The first property (i) is immediate, since a NDS
cannot be converted into a free entangled state by means of local operations.
The second one is not so obvious for continuous variables, but it follows from
the results of the previous subsection. We thus have:
Observation 2.- The property of non-distillability is invariant under the one
side local action A ® IOAt ® I.
The proof is simple - the arguments of Ref. [1] can be applied (see also
[6]) to show that any local separable superoperator cannot cause that the state
looses the non-distillability property.
Observation 3. - The set of all NDS is closed in the norm II . liT.
218 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
To prove the closeness of the set of NDS, we prove that its complement, i.e.
the set of distillable states D is open in the trace norm. To this aim we repeat
the arguments of subsection A and consider some p E D, for which there exists
some finite integer number n, PA, PB, rank two projectors acting on H A, H B,
and a rank two vector Iw) E HA ® HB such that Eq. (17.2) is fulfilled. We
consider now an open ball in the trace norm, i.e. i3.,., (p) = {p', liP' - piiT < 'f/}.
Note that if p' E i3.,., (p) then the operatornorm fulfills lip' - pil :::; lip' - piiT < 'f/
[16]. Using the same argument as before we show that for 'f/ < IEI/4n,
i3.,., (p) CD, which completes the proof. 0
Combining the Observations 1.-3. we see that the set of NDS states is
nowhere dense in the trace norm, which implies the same property for the BES,
PPT states, and separable states.
(17.16)
The above state is build from infinitely many "copies" of the same 3 ® 3 1 BES
CJ labeled by CJ n . Each of CJ n has the matrix elements of the original CJ, but in
the basis Sn = {li,j)}7,j~~n' Here {Pdi=l is an infinite sequence of nonzero
probabilities, L~l Pi = 1. The bound entanglement of the CV state a is in a
Bound entanglement/or continuous variables is a rare phenomenon 219
(17.18)
for n < m with (in general) complex an and en, such that (i) 0 < ICn+l1 <
Icnl < 1, (ii) E~=l E:>n IIwmnl1 2 is finite. The latter condition can be
achieved for example by setting an = an, en = cn, for some 0 < a < c < 1,
see [6]. Physically, the vector Iw), when normalized, may describe a state
of two modes of the quantized electromagnetic field, or more generally two
harmonic oscillators. The state (17.17) has the following properties: (i) it is
bound entangled, as it has the PPT property (i. e. it has the positive partial
transpose); (ii) it is not a simple "direct sum" of finite spin BES in a sense of
the "spurious" examples discussed above (Eq. (17.16».
Recently considerable attention has been devoted to the so called Gaussian
states. In systems of two harmonic oscillator modes (one of Alice, one of Bob),
i.e. in the, so called, Gaussian 1 x 1 case, it has been shown that no bound
entanglement exists - such Gaussian states are either separable [18, 19], or
distillable [20]. In another words, in this case PPT property is a necessary and
sufficient condition for separability, and non-distillability. This result can be
220 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
extended to the case 1 x N. Soon after realizing this facts Werner and Wolf
have found an example of a Gaussian BES with PPT property [7]. This result
has been achieved by considering first covariance matrices of Gaussian states
and their null subspaces. It was noted that the Gaussian state is separable,
iff its covariance matrix can be minorized by some block diagonal covariance
matrix. Second, the characterization of PPT states in terms of covariance
matrix has been found. The BES has been constructed using an elegant explicit
construction, performed using the analysis of the range and the "subtraction
method" first developed for spin systems in Refs. [21, 22, 23, 24]. In the
terminology of Refs. [21,22,23,24] the states found in Ref. [7] are examples
of the so called "edge states". The approach of Ref. [7] can be used further
to analyze multiparticle entanglement. In particular, one can try to "split" the
covariance matrix of n x n state in a way to get m x m x m state with some
bound entanglement properties. Indeed, we have recently managed to solve the
separability problem for the case of tripartite system with one mode per each
party [25]. The result of Werner and Wolf appeared first a little surprising in
the view of Refs. [18, 19, 20]. Recently, some of us have been able to clarify
this and solve ultimately the separability [26] and distillability [27] problems
for Gaussian states of two parties sharing arbitrary number of modes. While
the PPT property remains a valid necessary and sufficient condition for non-
distillability, the separability criterion has a complex form of a nonlinear map
for covariance matrices.
Finally, it is worth mentioning that it is not known yet whether there exist
BES which do not have PPT, even though there is a strong indication of this
fact [28]. If this were finally true, this would have important implications in
the context of distillation [29], since it may well happen that by mixing two
NDS one obtains a distillable one.
it should mean that the pure states with infinite Schmidt rank are necessarily
involved in the mixed state representation. One possible definition would be
that a generic CV state with infinite Schmidt rank should be necessarily of the
form (! = EiPilwi)(Wil, with IWi) not necessarily orthogonal, but with at
least one IWi) of infinite rank. Such states obviously exist - take for instance
one pure state of infinte Schmidt rank, or a convex combination of two such
states. However, in the above definition the precise notion of the decomposition
in the CV case in the sense of Ref. [30] has to be specified. Another possible
definition (which seems to be significantly weaker) would be to require the
generic CV state to be the limit of n ® n states of Schmidt rank nOt for some
O<a~1.
Concerning BES - we do not know whether there exists any BES for CV
with PPT property, having at the same time the feature of being a generic CV
state, whatever it would mean. It is worth stressing at this point that according
to the results of Ref. [31], PPT entangled state in n ® n space are expected
to have Schmidt number smaller than n. In fact, in the Appendix A, we
present the arguments analogous to those used in Ref. [31] that the typical PPT
bound entangled states in n ® n space either have the Schmidt number of order
0(1), or their partial transpose have this property. It is possible, however,
that the recently introduced Gaussian bound entangled states [7, 26] satisfy
all requirements as far as the CV genericity is concerned. It would thus be
interesting to analyze the Schmidt number of those states.
As we shall see below this leads to more complicated issue. We will not
give definite answers here. However further methods of investigation will be
suggested.
Again, as in previous section, one of proposed definitions could be the
following:
A. The state (! represents ''fully CV free (nondistillable)" entanglement if
and only if it is possible (impossible) to distill nonzero amount of pure states
with infinite Schmidt rank from state {! .
Nonzero amount is here understood in sense of usual distillation yeld (i.
e. as a nonzero amount of pairs). Note that to qualify distilled entanglement
in finite dimensions the condition of asymptotic approaching the maximally
entangled state Iw+) = Jm l:~llei' ei) was required. It is known however
that there is no maximally entangled states of infinite Schmidt rank. Thus in
place of W+ one would have probably use some fixed pure state Woo having the
reduced density matrix non singular or at least of infinite rank (this is eqivalent
to the infinite Schmidt rank of woo).
Another interesting (weaker) defintion would be more in spirit of Ref. [6]
where increasing sequences of finite Schmidt rank were used. Namely one can
propose:
B. The state (! represents CV free (nondistillable) entangled if and only if
it is possible (impossible) to distill nonzero amount 'T/p > 0 of p-Schmidt rank
states with limsupp'T/p > 0 .
The main difficulty dealing with Schmidt rank in those definitions is that the
operational methods of its detection in context of CV are not enough developed.
For example it is not known whether the proposal B above is equivalent to
the following generalisation of the "two-qubit subspace"(see sec. II A):
the state is fully free (bound) entangled iff there is (no) n and the family of
bilocalfilters Ap 0 Bp such that the new n-copy states
(17.19)
violate the p-Schmidt rank test via positive map i. e. [ll 0 Ap] ((!~) is not positive
matrix. The map Ap(X) == Tr(X)J - (p _1)-1 X is p - I-positive but not
p-positive and was used to detect p-Schmidt rank of isotropic entangled states
[13].
Dealing with the state (17.19) is not easy because even in the case of finite
dimensions the possibility of asymptotically singular denominator in formulas
like (17.19) leads to suprising effects (see [15]). Nevertheless, after simplifi-
cation we shall utilise the above point of view. In particular, putting Ap 0 Bp
equal identity we shall generalise the results of section II.
Bound entanglementfor continuous variables is a rare phenomenon 223
Now suppose that g' E B1/(g) = {g', IIg - fl'IIT < 1]}. Then
I(wl[rr ® Ap](g - g')lw)1 = l(wl(gA ® I - gA ® I
+(p _1)-I(g - g')lw)1 :::; ITr[gW(gA - gAJ ® III +
l(wl(P _1)-I(g - g')lw)1 :::; Tr[g~lgA - gAil +
(p -1)-II(wlg - g'lw)1 :::; 1](1 + (P _1)-1) (17.21)
states which completes the proof. Remarkable that the line of the proof remains
completely correct for "uniform" assumption [i.e., (17.20) satisfied with one
€ for all natural p some l\]f) = 1\]f(P))] but than the assumption itself can be
easily shown to be false in the sense that no state can satisfy it.
4. CONCLUSIONS
In this paper we have considered non-distillable states for continuous vari-
ables. In the main part of the paper we have proven that the subset of non-
distillable states is nowhere dense in the set of all CV states. This is a much
stronger result than the recent one by Clifton and Halvorson [12], which prove
the same result for the set of separable states, since that one is contained in the
set of NDS. Moreover, our results imply that the subsets of BES and PPT states
are also nowhere dense. Thus, generic CV states are distillable. We have also
presented some examples of BES and discussed their genericity from the point
of view of CV and their Schmidt number. In the Appendix A we have presented
an evidence that all PPT BES in n ® n systems either have Schmidt number
smaller that 0 (1), or their partial transposes have this property. Finally we have
analysed the genericity of CV entanglement in context of Schmidt number. In
particular, we have studied the assumptions of the main theorem and proved
more general result that nowhere dense is the set of all states from which it is
impossible to produce p-Schmidt rank state from a single copy in some (well
defined) way. The latter involves single copy protocol (provided in Appendix
B) being a generalisation of that obtained with help of reduction cirterion.
By provinding some proposals of definitions of what can be treated as "fully
CV" we have shown that further investigation of genericity in context of CV is
desirable.
We thank Anna Sanpera, Dagmar BroS Geza Giedke and Otfried Giihne
for useful discussions. This work has been supported by the DFG (SFB 407
and Schwerpunkt "Quanteninformationsverarbeitung"), the Austrian Science
Foundation (SFB "control and measurement of coherent quantum systems"),
the European Union Programme "EQUIP" (IST-1999-11 053) and the Institute
for Quantum Information GmbH. Part of the work was completed at The
Erwin SchrOdinger International Institute for Mathematical Physics, during the
Program "Quantum Measurement and Information", Vienna 2000.
• It is enough to show the conjecture for the, so called, edge states [22, 23,
24], i.e. the PPT states 8 such that there exist no product vector Ie, 1)
in their range, such that le*, 1) is in the range of the partially transposed
operator 8TA .
• Let r (p) denotes the rank of p. It is likely that it is enough to prove the
conjecture for the edge states of maximal ranks [22], i.e. those whose
ranks fulfill r(8) + r(8 TA ) = 2n 2 - 2n + 1. We expect that such states
are dense in the set of ~l edge states. To show the latter statement, we
consider an edge state 8 which does not have maximal ranks. We can
always add to it infinitesimal amount of projectors on product vectors
destroying the edge property. The resulting state p would have more
product states in its range, than the product states used to destroy the
edge property. Subtracting projector on product states different from
the latter ones, would typically allow to construct an ec!.ge state 8 with
maximal ranks, which would be infinitesimally close to 8 in any norm.
• Let R(A), K(A) denotes the range and kernel of A, respectively. The
canonical form of an non-decomposable entanglement witness that de-
tects the edge state 8 is ([23, 24], see also [32])
(17.A.1)
Let us therefore try to construct the desired vector I~S) of Schmidt number
s. In general such (unnormalized) vector will have a form
S
where li are arbitrary complex coefficients for i = 1, ... ,s, and lei, h) are
linearly independent product vectors for i = 1, ... ,s. Note, that the vector
(17.A.2) depends on s complex parameters li for i = 1, ... ,s, whereas each of
the s vectors lei), Ih) depends themselves of n-l relevant complex parameters.
Let r(P) = kl' and r(Q) = 2n - 1 - kl. Since we want to prove the
conjecture either for the edge state 8, or for its partial transpose, without
loosing the generality, we may assume that kl 2: 1. We may then single
out one projector out of P, and write P = PI + Iw)(wl, where PI 2: 0,
r(PI) = r(P) - 1, and Iw) is in the range of P. We can choose then lei, h)
in such a way that Qlei,h) = 0, and Pllel,h) = 0. These are effectively
2n - 2 equations for vectors lei, h) which depend on 2n - 2 parameters, so
that we expect a finite, but quite large number of solutions (c.f. [22]). At
the same time, ('ljJsIQI'ljJS) will become a quadratic hennitian form of li'S with
vanishing diagonal elements. Such a hermitian form has typically more than
one dimensional subspace N of negative eigenvalues for large s. But, one has
to fulfill also the last equation implied by ('ljJslw) = 0; this limits the values
of li to a hyperplane, which should have at least one dimensional common
subspace with the subspace of negative eigenvalues N. This would prove that
either the Schmidt number of 8 or of 8T A is of the order of 1.
Note that for a given 8, if the presented construction can be shown to be
successful for every witness of 8, then it provides a sufficient condition for the
state 8 to have the Schmidt number smaller than s.
(17.B.1)
(l7.B.2)
Bound entanglement/or continuous variables is a rare phenomenon 227
Now instead of I'll) we put I'll') being a normalised projection of I'll) on support
of (lm. In this way we get the inequality identical to the one of [14] with the
only difference that p was equal 2 there. This allows to repeat the reasoning of
Ref. [14]: after the application of suitable local filtering and U ® U* twirling to
(lm one produces the m ® m isotropic state (lis = (1- q) ~ + q Iw+) (w + I with
the fidelity F == (W+I(lislw+) > p~l. But the latter implies (see [13]) that the
final state (produced from initial (l by means of local operations and classical
communication - LOCC) has the Schmidt number at least p. This concludes
the analysis. Note that further steps of recurrence protocol with generalised
XOR ([14]) can be applied.
Notes
1. Subsequently we shall denote by n (8) n states the states of quantum systems defined on the Hilbert
=
space 1{ en (8) en. The space will be sometimes called "n (8) n space".
References
[1] M. Horodecki, P. Horodecki and R. Horodecki, Phys. Rev. Lett. 80, 5239
(1998).
[2] C. H. Bennett, G. Brassard, S. Popescu, B. Schumacher, J. Smolin and W.
K. WooUers, Phys. Rev. Lett. 76, 722 (1996).
[3] P. Horodecki, M. Horodecki and R. Horodecki, Phys. Rev. Lett. 82, 1046
(1999).
[4] M. Murao, V. Vedral, Phys. Rev. Lett. 86, 352 (2001); P. W. Shor, J. A.
Smolin, and A. V. Thapliyal, quant-ph/0005U7. W. Diir, J. I. Cirac, and
R. Tarrach, Phys. Rev. Lett. 83, 3562 (1999); W. Diir and J. I. Cirac, Phys.
Rev. A 62, 22302 (2000).
[5] See S. Braunstein and J. Kimble, quant-ph/991 0010, and references
therein; for experiments see A. Furusawa et al. Science 282, 706 (1998).
[6] P. Horodecki and M. Lewenstein, Phys. Rev. Lett. 85,2657 (2000).
[7] R. F. Werner and M. M. Wolf, Phys. Rev. Lett. 86, 3658 (2001).
[8] K. Zyczkowski, P. Horodecki, A. Sanpera and M. Lewenstein, Phys. Rev.
A 58833 (1998); K. Zyczkowski, Phys. Rev. A 60 3496 (1999).
[9] G. Vidal, R. Tarrach, Phys. Rev. A 59141 (1999).
[10] S. L. Braunstein, C. M. Caves, R. Jozsa, N. Linden, S. Popescu and R.
Schack: Phys. Rev. Lett. 83, 1054 (1999); R. Schack and C. M. Caves,
Phys. Rev. A 60, 4354 (1999).
[11] N. Linden and S. Popescu, Phys. Rev. Lett. 87, 047901 (2001).
[12] R. Clifton and H. Halvorson, Phys. Rev. A 61,012108 (2000).
228 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Alex Kuzmich
Norman Bridge Laboratory of Physics 12-33
California Institute of Technology, California 91125, USA
[email protected]
Eugene S. Polzik
QUANTOP - Quantum Optics Center
Danish National Research Foundation
Institute of Physics and Astronomy, University of Aarhus, Denmark
[email protected]
Abstract In this Chapter methods for generation of squeezed and entangled states of atomic
ensembles are described along with the protocols for quantum state exchange
between light and atomic samples. Realization of these protocols provides the
means to store/retrieve quantum information transmitted by light in/from atomic
samples, which can be used for processing of this information. Polarization
variables (Stokes parameters) of a multi-photon light pulse and spin components
of a multi-atom atomic ensemble are the continuous variables employed in these
protocols. Two different methods for a quantum state exchange are analyzed:
(a) mapping of non-classical states of light onto atomic spins via complete
absorption of resonant light, (b) teleportation-like transfer of a quantum state via
a QND-type interaction between off-resonant light and atoms.
1. INTRODUCTION
The ability to implement quantum interfacing between light and matter is cru-
cial for many aspects of quantum information processing including distributed
computing, quantum networks, computational complexity, eavesdropping in
quantum cryptography, to name a few. Continuous field and atomic variables
have a dramatic practical advantage for this type of operations compared to
231
S.L. Braunstein andA.K. Pati (eds.), Quantum Information with Continuous Variables, 231-265.
© 2003 Kluwer Academic Publishers.
232 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
single-particle qubits, namely efficient quantum state exchange for the former
can be achieved without high-Q cavities necessary for operations with the latter.
Whereas efficient quantum state exchange between single photons and single
atoms requires cavity QED with strong coupling [1], efficient state exchange
between collective quantum variables of light and atoms can be achieved in
free space with very weak interaction on a single photon/single atom level.
"Strong" interaction in case of continuous variables for multiphoton pulses and
multi-atom samples reduces to high on-resonance optical density of the atomic
sample, which is easy to achieve experimentally.
An efficient quantum state interaction between light and atoms at the level
of continuous variables allows to carry out a number of quantum state/quantum
information processing protocols. In this chapter we will describe some of them
with the emphasis on experimental implementation. First we will define con-
tinuous variables for atomic spin polarized ensembles and polarized light and
introduce coherent, squeezed and entangled spin states. We will then consider
mapping of a quantum state of light onto atomic state via complete absorption
of resonant squeezed light. This method of mapping a quantum state of reso-
nant light onto atoms (Section 2) has been proposed in [2], further investigated
in [3,4], and experimentally implemented for generation of spin squeezing in
[5]. Next we will discuss an off-resonant dispersive interaction of light and
atomic ensembles. This approach proved to be especially successful in several
aspects. We will show how it allows for a quantum non-demolition (QND)
measurement of spin projection leading to generation of a spin squeezed state
[6]. An important problem of atomic teleportation for continuous variables will
be considered. Teleportation requires a resource of entanglement, more specif-
ically distant entangled atomic samples are needed. Whereas both discrete
and continuous variable entangled (EPR) states of light have been successfully
generated by several groups [7,8,9], entanglement of distant atomic systems
has been experimentally achieved only recently [10] with continuous variables
following the proposals [11, 12]. Its applications to atomic teleportation and
interspecies light-atoms teleportation will be considered.
z axis, with commutation relations [ai(t), aj(t')] = OijO(t - t'), i,j = +,-.
The Stokes parameters obey the standard spin commutation relations,
(18.1)
In the following we will assume that light is polarized in the classical sense
along the x axis, i.e. that only Sx has a non-zero mean value. All the interactions
considered below introduce only minute changes in the mean polarization, and
therefore we substitute Sx with its classical mean value (Sx).
Analogous to polarized light, a quantum state of a collection of N spin 112
atoms can be described by its collective spin F == ~f=l F(k) with commuta-
tion relations
(18.2)
As with light, we will assume that atoms are spin polarized in the classical sense
along the x axis, and that this mean polarization does not change much. There-
fore the quantum states of polarization, both for light and atoms, considered
here are generated by small rotations of the collective spin (Stokes vector for
light) around its mean value. A formal analogy between the mUlti-particle spin
ensembles considered here and standard position/momentum-like continuous
variables in the phase plane comes from the commutation relations (18.1,18.2).
Under the assumption of small spin rotations, the right hand sides of these equa-
tions are constants. Dividing the equations by the absolute values of their right
hand sides, one obtains standard commutation relations of position/momentum
for normalized spin variables. The procedure of substituting small rotations on
the Bloch sphere with displacements on the plane is known [13] as contraction
of the group SU(2) to group U(l). For example, for small rotations around
the IF, - F) state (that is, the state with mean polarization along the z-axis),
this is achieved with the following correspondence between the generators of
the groups:
At
F+ -t a,
F_ -t a, (18.3)
(Fz + F) -t aAtAa,
(analogous relations hold for the Stokes vector of light, of course). This
contraction of the group SU(2) onto the group U(l) is at the heart of the
approaches to continuous quantum information processing that we describe
below. The spin formalism allows us to treat light and atoms on an equal
footing. In the following, when it is not specified otherwise, "spin" refers to
both light and atoms.
234 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(lS.4)
(lS.5)
Commutation relations (1S.2) for the spin F have Heisenberg uncertainty rela-
tionship associated with them
(1S.6)
(1S.7)
is less than 1 for some ((), ¢). Here Fl. is spin component in a direction
orthogonal to ((), ¢). The squeezing parameter 'fJ determines the accuracy of
phase measurement 8¢:
8¢=~. (lS.S)
Atomic continuous variable processing and light-atoms quantum interface 235
where Fxv , Fyv are x- and y- spin components of an auxiliary spin in a CSS
directed along the z-axis (we assume that all three spins have the same length
F). r is the parameter that characterizes the strength of the quantum correlations
between these two spatially separated spins.
For two spins with opposite classical orientation, (FzI ) = -(Fz2 ), the
necessary and sufficient condition for entanglement can be formulated in terms
of the variances of measured quantities [19] as
(18.11)
236 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
The interpretation of this condition follows from the fact that for both spins
in CSSs the following equality holds: ((Ll(Fx,y))2) = !(Fz ). Entanglement
between the elementary spins of the two samples is, therefore, according to the
above condition equivalent to the spin variances smaller than that for samples in
a CSSs, which is characterized by uncorrelated individual atoms. The entangled
state of this type is a two-mode squeezed state for the continuous spin variables.
3.1 THEORY
Let us consider the propagation of quantum correlated e. -m. fields through an
atomic medium which fully absorbs (and spontaneously re-emits) the incoming
light field. In the process of propagation some quantum statistical properties
of the field are altered by the interaction with atoms. The atoms acquire
certain quantum features of the field via the absorption mechanism, and at
the same time, spontaneously emit photons. In the steady state the scattered
and the incoming light intensities are identical and they are proportional to
the total excited state population of the atomic medium. In the limit of weak
fields, the emission is also coherent; both the scattered field amplitude, and the
atomic dipole amplitude are proportional to the incoming field amplitude. This
suggests that certain collective atomic operators exhibit new and non-classical
noise properties due to the interaction with the fields. We consider a V-transition
with each arm 0 ++ 1(2) interacting with a separate quantum field, 121(2) (see
Fig. 18.1). The two excited states can be two magnetic sublevels (say, with
m = 1 and m = -1) in which case the spin will correspond to the orientation
Atomic continuous variable processing and light-atoms quantum interface 237
Excited state
m=-l m=+l
cr+-
••••
• • ••
0-
~ m=O,
Atomic ground state
sample
Figure lB. 1 Spin squeezing of an excited state via complete absorption of squeezed light.
Linearly polarized squeezed light can be viewed as a superposition of quantum correlated
modes in right- and left-circular polarizations. When these modes excite the atomic state as
shown in the Figure, quantum correlations (squeezing) are partially transferred to the atomic
state. Spontaneous emission which prevents perfect transfer is also shown.
and/or alignment of the excited ensemble. Alternatively, the two states can be
two hyperfine sublevels. We treat the case when the 0 - 2 transition interacts
with a squeezed vacuum mode, and the 0-1 transition interacts with a coherent
field mode. To analyze the noise properties of the collective upper-states atomic
coherence FI2 we assume that the coherent amplitude a is much stronger than
the squeezed vacuum fluctuations, and subsequently that the coherent part of
FOi is much greater than the fluctuating quantity F02 . The mean spin is along
the z-direction and has a value equal to half of the total atomic population in
the excited state 11):
(Fx) = (Fy) = O. The components where we shall look for squeezing are
Fx,y. Detailed calculations [2] result in the collective atomic operator FI2 at
frequency .6. being expressed simply in terms of the incident and the transmitted
field operators:
,-2.6.
a* . a2(in) (WI +.6.) +
,-2.6. d
a* . 2(in) (WI + .6.)08.13)
squeezed vacuum field, d2 (in) (WI + b..) is the annihilation operator describing
the bath of the spontaneous decay modes.
Let us examine if the Fx,y are squeezed. We find from Eq.(18.13) that the
variances of the atomic spin components are linked to the quadrature phase
variances X;,y of the input field ih:
(18.14)
For the vacuum input field 0.2 the variances 4(X~,_) = 1 and the x, y spin
variances are equal to ~(Fz) as would be expected for the coherent spin state.
With broadband squeezed vacuum as input field 0.2, one of the variances van-
ishes, e.g. (X;) = 0, and we obtain the spin squeezed state with 50% degree
. • A2 _ 1 A
of squeezmg. (Fx) - 4: (Fz ).
In the example above spin squeezing is achieved in an off-diagonal element
of the excited state orientation. By slightly changing the geometry of the ex-
periment one can also achieve squeezing in the diagonal elements, i.e. in the
population difference between the two excited states. In order to achieve this
goal the coherent and squeezed vacuum fields with orthogonal polarizations
should be mixed at a polarizer to produce quantum correlated left- and right-
circularly polarized fields 0.1 and 0.2. With the average atomic spin oriented
along the direction of the coherent polarization x, either Fy or Fz compo-
nents can be squeezed by 50%, depending on the phase difference between
the squeezed vacuum and the coherent beam. The latter case ("longitudinal
spin squeezing") corresponds to sub-Poissonian fluctuations in the population
difference between the two excited states.
3.2 EXPERIMENT
The experimental realization of this proposal has been carried out with a
collection of cold Cesium atoms confined in a magneto-optical trap [5, 20].
Atoms are excited with a weak quantum pump (Fig.l8.2) from the ground state
to the excited state 6P3/2, F = 5 (hereby abbreviated as 6P). The pump is a
mixture of the x-polarized coherent beam and y -polarized squeezed vacuum
produced by a frequency tunable optical parametric oscillator below threshold.
A weak linearly polarized probe in the configuration shown in Fig.18.2 is
sensitive to the following three operators: Fz, F; - F;, and FxFy + FyFy
[21]. These are spin operators describing the entire 6P ensemble. They obey
the commutation relation ~ [Fz, FxFy + FyFx] = F; - F;. We consider these
three components as components of a quasi-spin operator. With the coherent
component of the excitation polarized along the x axis the only component
of the spin polarization with a non-zero mean is (F; - F;). The uncertainty
relation following from the commutation relation is: 8(Fz)8(FxFy + FyFx) ~
Atomic continuous variable processing and light-atoms quantum inteiface 239
Lock-In (DC)
Lock-In (RF)
~
Cesium
6D
m
~~~~prObe
I,lt+--t'--·+-----, 6P312 F=5 -
852nm Quantum
6S F=4 pump
112
I(F; - F;) I .Since the operators describe the whole ensemble, the right hand
side of this inequality is proportional to the number of atoms N in the 6P state.
Excitation of the atoms with coherent light will lead to the quantum noise of
the z component o(Fz)coh = V51(F; - FJ)I/(2F + 3) '" VN due to the
uncorrelated noise of individual atomic spins. Following the theory described
in the previous section, one should expect that injection of the squeezed e.-m.
vacuum into the other port of the polarizer PBS 1 (Fig.18.2) should lead to
squeezed o(Fz)sq < o(Fz)coh or anti squeezed o(Fz)antisq > o(Fz)coh spin
states depending on the phase of the squeezed vacuum field.
Turning now to the experiment, about 109 Cs atoms have been collected in a
magneto-optical trap. The spin noise ofthe 6P state has been detected using the
probe polarization noise technique described in detail in Ref.[22](Fig. 18.2).
The change in the probe differential photocurrent noise caused by atoms is
';) 0,40
......
o
"0 0,30
'-'
..'
,,-....
8'.
'.
-0,04
-
$:I)
-0,06 Jg
L.--r-.....--.-----.---,--.--"'T""'=--.....---,--l-O,08 ~
-10 -5 0 5 10 '-'
(18.15)
sxj 1 ftx
Off-resonant :::I===:::J=t=======:::=:=::=_
<::'
linearly polarized
light Polarized Atoms
a continuous XOR gate, one of the cornerstones for the quantum information
processing with continuous variables. To show this, let us derive equations of
motion for the field and atomic variables. Let the light pulse with the mean
polarization along the x-axis propagate along the z-axis through the atomic
sample also with the mean spin along the x-axis (Fig. 18.4). In the Heisenberg
picture transformations of the spin operators can be written as
- sin{aS
A
cos{aSz )
z) 0) (Fx
0
A
t:y
)
~n)
, (18.16)
o 1 Fz
whereas the z-components of the spin are constants of motion. Here the
Stokes vector S of the light pulse has the length (Sin) = ~x, where n is the
244 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
number of photons in the pulse. As can be seen from the above equations, the
interaction Hamiltonian (1S.15) provides a continuous analog of XOR gate if
aJFnN/2 = 1 (F is the value of the spin of one atom). The rotations of
individual qunats are achieved for light with the help of A/2 and A/4 plates and
for atomic spins by applying dc magnetic fields.
Many algorithms of the field of discrete quantum information have their
continuous analogs if every qubit is replaced with a qunat, every XOR-gate
is replaced with a continuous XOR-gate, and every single qubit rotation is re-
placed with a corresponding rotation of the qunat in the quadrature plane [25].
Using collections of atomic spins and the field modes as qunats is a suitable ar-
chitecture for implementation of continuous quantum information algorithms.
Below we will show how using such philosophy one can, for example, achieve
exact mapping of quantum states between light and atomic spins, in both direc-
tions. But first we will describe the experiment demonstrating the feasibility
of the above Hamiltonian for continuous quantum information applications,
namely for generation of spin squeezed states of atoms.
We find
N a
= -2 cos n -2'
A
(F)
x
N 1
-4 + -2na
- 2 (n(1 - cos N a) + (1 + cos N a))
N . a N-l a
- sm - cos - (1S.19)
a 2 2
exp[~/2l
7] ~ JN~ , (1S.20)
takes care of the random variation from trial to trial of the direction of (F).
observable,
(18.21)
(18.22)
We obtain
(18.23)
where F(i) is the total spin of one atom (F(i) = 4 in our experiment), T == dlv,
N = pdvTL is the total number of atoms passing through the interaction region
during time T and P is the optical power of the probe beam. We obtain for the
second moment
(18.24)
where the first and second terms are due to (atomic) spin noise and photon shot
noise, respectively. Now we can calculate the measurement accuracy of the
spin rotation angle from Eq.(18.22) and we find
(18.25)
When (XT)~PB « 1, the photon shot noise corresponding to the last term
under the square root sign is much smaller than the atomic noise and can
be neglected. In this case, for low frequencies n « 1I T we find that the
measurement accuracy is given by the SQL, O¢SQL = 1/vF(i)N. From
Eq.(18.25) it follows that, as the frequency n of the applied spin rotation
increases, the phase uncertainty o¢ falls below the SQL. The limit on the
achievable uncertainty is due to the photon shot noise. If the collective atom-
photon coupling is high enough (through the use of either dense atomic samples
Atomic continuous variable processing and light-atoms quantum inteiface 247
or optical cavities for the probing beam), the limit is detennined by the back-
action of the probe photons onto the collective atomic spin. The measurement
accuracy of the spin rotation angle is then the Heisenberg limit 1/N [26].
An outline of the experimental setup is shown in Fig. 18.5. A paraffin-
coated glass cell contains Cs atoms. Under the conditions of our experiment,
the measured lifetime of spin polarization in the cell is on the order of 1 s,
which means that the atoms undergo thousands of collisions with the walls
without being depolarized.
Figure 18.5 Outline of the experimental setup. Not shown are 3 pairs of coils to produce dc
magnetic fields and a coil producing rf magnetic field. See text for details.
Initially it is important to prepare the atomic sample in a CSS. There are two
important features of a CSS: (a) it is fully spin-polarized along some direction
(and therefore has maximum coherence) and (b) spin fluctuations of the sample
J
are at the shot-noise level given by F /2, where F is the value ofthe collective
spin.
In our experiment two 852 nm diode lasers DL1 and DL2 are used to spin-
polarize the atomic medium. Both light beams are circularly polarized with the
same helicity and propagate along the direction of the magnetic field of several
Gauss. The laser DL1 of about 3 mW average power (adjusted with a neutral
density filter) is tuned to the 681 / 2 , F = 3 --+ 6P3/2 transition(s), while the laser
DL2 ofless than 100 f-lW average power is tuned to the 681 / 2 , F = 4 --+ 6P3/2
transition(s). A small portion of the linearly polarized light from DL3 is split off
and used to monitor the degree of spin polarization. This probe beam propagates
through the cell along the direction of spin polarization. By scanning the
frequency across the Doppler profile, we measure the dispersive profile of the
polarization rotation angle and the Lorentzian profile for absorption. The major
248 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
part of the light from laser DL3 is used as our QND-probe (see below). We align
the magnetic field (and therefore also the spin polarization) orthogonally to the
probe beam with the help of the Faraday rotation signal in the dc channels of our
homodyne detectors on the QND-probe beam. The degree of spin-polarization
of the atoms is found to be greater than 95%.
The QND probe beam is detuned about three Doppler widths from the center
of the 681/2 , F = 4 -+ 6P3/2 transition(s). Al mW light beam with Gaussian
waist of 100 p,m is focused into the gas cell. The optical pumping beams and
the probe beam do not spatially overlap. The. probe beam is then refocused by
lens L2 onto a Glan-Thompson polarizing beam splitter (PBS) whose outputs
are measured by two (85% quantum efficiency) photodiodes D1 and D 2 • The
outputs of the photodiodes are amplified, and the difference signal is fed into
a spectrum analyzer 8 A, where noise spectra between 3 and 20 MHz are
recorded.
An important aspect of the spin squeezing experiment is the need to identify
the spin shot-noise level. By analogy with the use of thermal light for the
photon shot noise normalization, we make use of the unpolarized atoms as
benchmark for the spin shot noise. We have shown earlier in this Chapter, that
the spin noise of unpolarized atoms in a cell scales linearly with the number
of atoms, as expected for an uncorrelated sample of atoms. Because our Cs
atoms have total spin F = 4 instead of elementary I/2-spins, the spin noise
of an unpolarized collection of atoms is j(F + 1) times larger than the spin
noise of the CSS with an equal number of atoms. As we argued above, the spin
shot-noise can be measured at n = O. From the experimental point of view,
however, the difficulty of measuring it in this way is connected with the fact
that our probe laser has a large amount of excess amplitude and phase noise at
frequencies below 1 MHz. Because of this, we prefer to normalize with respect
to the shot noise, with the magnetic field applied in a direction perpendicular
to the QND-probe propagation. In this case the spin noise manifests itself at
the Larmor precession frequency WL, while the amount of the observed noise
is reduced to half. We find the peak shot atomic noise to be about four times
the photon shot noise.
To obtain the shot noise level for the CSS, we make use of the results of
the spin noise measurement with unpolarized atoms. We take into account the
factor of 2 (F + 1) /3 for the difference between the shot noises of an unpolarized
sample of atoms and the CSS, and also of the fact that after optical pumping
we have 16/9 times more atoms in the F=4 state than in the unpumped cell.
That was also confirmed experimentally from absorption measurements. The
spin noise level for the CSS obtained in this way is shown in Fig. 18.6 (broken
line).
Next, we reduce the magnetic field back to 2 Gs and apply a rf magnetic
field along the y-axis in order to produce a small rotation. Fig. 18.6 (solid
Atomic continuous variable processing and light-atoms quantum interface 249
Cil
SOL
.2>
IJ)
4
Q)
IJ)
o 3
C
-0
Q)
.~ 2
ctl
E
"-
o 1
c
o ~~~~~~~~~~-L~~~~
8 10 12 14 16 18
rf frequency (MHz)
Figure 18.6 Measured spin noise spectrum. The dashed line indicates the SQL accuracy of
spin rotation measurement. The peak at 16 MHz is due to spin rotation by the applied rf coils
and is a demonstration of sub-shot noise atomic interferometry.
line) shows the spin noise spectrum observed under these conditions. The solid
line is the SQL of the spin projection measurement for a CSS when the shot
noise of the probe is taken into account. The peak at 16 MHz is the measured
spin rotation due to the applied rf field. This demonstrates the sub-shot noise
performance of the atomic spin interferometer.
As the next step, we pass the pulse through the cell along the y-axis. As a
result, we obtain
S(out-2) = S(out) = s(in)
z y y
+ p(in)
z' (18.27)
Since the light is initially in a squeezed state, with ((b.S~in))2) « n/4, we can
neglect the first tenn in the Eq.(18.27). Thus, we have achieved exact mapping
of atomic spin state onto polarized light.
The first line describes the Faraday effect (polarization rotation of the probe),
whereas the second line shows the back action of light on atoms, i. e., spin
Atomic continuous variable processing and light-atoms quantum inteiface 251
The spin components Fi:;) are now defined in the frame rotating at the fre-
quency S1 around the magnetic field axis, x. Measuring the cos{S1t)j sin{S1t)
component of the photocurrent we perform a projection measurement of the
z / y component of the total spin for the two samples. This measurement entan-
gles two atomic ensembles. The degree of entanglement depends on the ratio
252 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
of the second "atomic" term in Eq.(18.30) to the Fourier component of the first
term, S~in) (t) cos(Ot) or S~in) (t) sin(Ot) . The latter is just the component of
the shot noise of the probe at the frequency O.
......
pumping pulse pulse
-
&~~. .~. .~.
B-field
~~.
i
Enlangnng and
• •
:
verifying pulses ______
•
: ' _ " - -_ _ _ _ _---.,
.... O.5ms-< Trme
6. 2 S2 = (S~~~s (0 )) 2 + (S~~fn (0 ) ) 2
8S; + '" { (F;in) + F;~n))2 + (F;in) + F;~n))2} . (18.31)
Atomic continuous variable processing and light-atoms quantum interface 253
:::i 40 •
~
<l
•
of
en
::; 30
•
0..
Q)
..c
e
0..
Q)
:5 20
15
~
c:
CI'l
.~ 10
~
t5
Q)
0.. J x [1012]
U)
04-~--~~--~~--~~-,--~~--~-,
o 2 4 6
Collective spin of the atomic sample
Figure 18.9 Nonnalized spectral variance showing entangled state of the two cells.
tocurrents from the two pulses are subtracted electronically and the variance of
the difference, 81p R' is measured. The vanishing 81p R corresponds to two
repeated measurements on the total spin state of the two samples producing
the same results, i.e. to a perfect knowledge of both z and y total components
for the two samples, and therefore to a perfectly entangled state. In the experi-
ment the correlation between the entangling and verifying pulse measurements
is imperfect for several reasons. First, the two optical pulses possess quan-
tum(shot) noise which does not cancel out. Second, decoherence processes
change the spin state between the two measurements. The first imperfection
can be reduced by increasing the size of the atomic portion of 8 2 compared
to the shot noise of the probe. The results of measurements with the delay
of O.5m sec between preparation of the entangled state and its verification are
shown in Fig.18.9. The results are normalized to the CSS limit (the linear fit
in Fig.18.8). This limit thus corresponds to the unity level in Fig.18.9. The
raw experimental data for the entangled state are shown as stars. The values
below the unity level verify that the entangled state of the two atomic samples
has been generated and maintained for O.5m sec. For detailed derivation of
the degree of entanglement from the data in Fig.18.9 we refer to Ref. [10].
The degree of entanglement calculated operationally from the data without
additional assumptions is (35 ± 7)%. The degree of entanglement useful for
teleportation calculated using an additional, experimentally proven assumption
of the initially CSS for both samples is higher, (52±7)%. The predicted fidelity
Atomic continuous variable processing and light-atoms quantum inteiface 255
(18.32)
where n is the number of photons in each of the coherent fields (for simplicity
assumed to be equal for both fields). That is, the spin components of the Stokes
vectors of the output states of light are given by the corresponding quadrature
operators of the OPO output. Since the latter obey relations of the form of
Eqs.(18.l0), we obtain relations Eqs.(18.1O) for the z- and y -spin components
of the output states of light.
As was pointed out in the original teleportation proposal [29], it is imperative
to be able to perform joint measurements on the quantum state to be teleported
and one of the EPR states. Off-resonant atom-photon interaction described
by Eq.(18.15) allows us to achieve this goal. The inter~ction leads to rotation
of the polarization of the field that is proportional to F z . Subsequent optical
polarization measurements provide the classical information in our protocols.
Let Alice pass her bright EPR beam 1 with the Stokes vector 8(1) through
the cell containing polarized atomic vapor along the z-axis. The unitary time
evolution operator of Eq.(l8.15) in the case of small spin fluctuations in F and
256 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Memory -
Atoms
state
Actuator -
phase and amplitude
modulator
Figure lB. 10 Outline of the scheme to teleport atomic quantum state on light.
These equations for interaction of the collective atomic spin with light re-
semble the beam splitter relations used for teleportation of a mode of electro-
magnetic field in Refs.[25, 30]. An insightful analysis of why a linear device
of a beamsplitter allows to perform joint measurements necessary for quantum
teleportation has been given by Vaidman and Yoran [31].
The nonlinear atom-light interaction of Eq.(18.15) is analogous to the QND-
type interaction for the optical quadratures rv XaXb . Unlike the beam splitter,
which mixes both the X and the Y quadratures of the two input beams in
the same fashion, interaction rv XaXb mixes only one pair of quadratures (Ya
and Yb), leaving the other pair (Xa and Xb) unchanged. Relations (18.33)
correspond to a 50:50 beam splitter of the optical teleportation when
(18.34)
(18.35)
The first term is negligible if ncoh/nl » 1. After Alice sends the results of
(l)(out) (coh)(out)
measurements of Sy and Sz to Bob, reconstruction of F m in
A A A (. )
§ (2) is achieved by Bob rotating the latter around the z-axis by the angle (PI =
_ _2_S(1)(out)and around the y-axis by the angle ¢2 = _ 2y'nl S(coh)(out)
yfnln2 y n co hy'Ti2 z
. The rotations are equivalent to displacements because the angles are small.
We obtain using Eqs.(18.1O)
S(2)(out)
y
S(2)(out) (18.36)
z
The last terms in these expressions are due to the extra noise introduced by
the imperfect EPR spin state. This noise goes to zero as r goes to infinity.
The factor of square root of two in the last terms corresponds to the "quantum
duty" of the quantum teleportation [25], as seen explicitly in the case of zero
parametric gain, r = o. Eqs.(18.36) show that the Stokes vector §~ut is
identical to the initial collective spin vector of Alice's atoms when
(18.37)
FAy + SUz
'" I '" A/ " '"
FAz = FAz, FAy =
Next Alice's atomic state is rotated around the direction of the mean spin,
x , so that F~z -+ F~y and F~y -+ - F~z. A strong coherent beam with the
mean polarization along x axis is then sent through Alice's ensemble along the
z axis and its y Stokes parameter is measured by the detector D A2 with the
result
In the above equations n, nc are the mean photon numbers of the unknown
state of light and the strong coherent beam respectively, and we assume n / nc «
1.
The mapping of the unknown state of light onto Bob's ensemble is now
completed by displacing the z, y component of his spin by dAI, - :c dA2 to
obtain
two macroscopic spin systems (Alice's and Bob's) in the initial states given
by collective spin operators FA, FB with mean polarizations (FAx) = (FBx)
and with the other two projections FAy, FBy,FAz, FBz. We also have at our
disposal a source of the EPR spin states of light as described above. The
Stokes' vector of the Alice's EPR beam is rotated by 7r 12 around the x-axis
so that the Stokes operators are BAz = -BBy, BAy = -BBz. The protocol
begins with Alice sending her EPR beam along the z-axis and measuring its
y -Stokes parameter with the detector D AI, and Bob sending a coherent x-
polarized pulse B]3h containing ne photons along the y-axis and measuring its
z-Stokes parameter with the detector D BI. The resulting atomic states of Alice
and Bob are described by the following operators:
I ,.. .... ,
= = FAy + SAz,
A A ,..
(18.39)
(18.40)
(18.41)
+ FBz
A. " A/
DB2 = SBy (18.42)
(18.43)
To complete the teleportation Bob now rotates his atomic state. We use
displacements instead of rotations to simplify the expressions. Bob's state is
260 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
FBz
te1e =
-u
F Bz - DA2 - DB2 = -FAy
-
(18.44)
FBy
te1e =
~"
F By +DAI- DEl = FAz
A
and the teleportation of the unknown state of the Alice's collection of atomic
spins onto Bob's atoms is proven (within a rotation of 1r around the x-axis). In
the above equations we assumed perfect entanglement between the EPR beam
components (r -* 00 in Eqs.(18.1O». We have also used the fact that the
measured values FBz,By of operators FBz,By are obtained with the QND-type
Hamiltonian, and therefore FBz,By - FBz,By = 0 (same is true for 8 13;).
A ;., I A.
(18.46)
and
sr'z "/
(18.48)
sr'
y -
AI
-Sy = -Sy
A
Here we used Eqs. (18.45, 18.46). After interacting with the atoms the two
EPR beams are detected by photodetectors D 1 , D 2 • The Stokes parameter S~
(18.46) is measured for the first beam and the Stokes parameter S~' (18.48)
for the second. The results of the measurements are used to rotate the atomic
spins in order to achieve the teleportation. The projections F~z and F~z are
displaced by the value S~ obtained from D 1 , and the projections F~y and F~y
are displaced by the value S~/. The results are
.... 11 I ,..
ptele = Fz - Sy = -FBz, (18.49)
Az
FAy
te1e = fr; - S~' = -FBy.
FABz
swap
= - Az
FA
(18.50)
FAswap
By = -
FA
Ay
Eqs.(18.49),(18.50) prove that the initial collective quantum states of the two
samples have been exchanged.
A different method for atomic continuous variable teleportation, using co-
herent light and entangled atomic samples, was proposed in Ref. [12].
5. SUMMARY
One of the most striking features of quantum information processing with
continuous variables is the relative simplicity of realization of the light-atoms
quantum interface. As opposed to the case of discrete variables (single pho-
tons and atoms) such interface for continuous variables does not require strong
coupling via cavity QED. Free space interaction of light with atomic ensem-
bles provides enough coupling for quantum state exchange and entanglement
of continuous variables. We have presented two approaches to continuous
quantum information processing, one based on resonant interaction, the other
262 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Acknowledgments
We are indebted to many of our collaborators who contributed to the work
reviewed in this chapter, particularly to L. Mandel, N. P. Bigelow, I. Cirac,
L.-M. Duan, B. Julsgaard, J. Hald, A. Kozhekin, K.Mj1jlmer, J.L. Sj1jrensen, and
P. Zoller.
Appendix: A
Let us assume that we have a running-wave cavity with a single spatial mode
of the electromagnetic field interacting with a collection of N atoms. First,
we consider an idealized case of alkali-like atoms with zero nuclear spin. The
interaction Hamiltonian desribing the atoms-field systems can be written as
2 N
HI = Ii L L1 9 exp(i Wii +1 zJL)ai ah+1 + h.c., (l8.A.I)
.
~=1 JL=
C
where a+, a_ are annihilation operators for the right- and left-hand polarized
components of the field. 9 is the coupling constant, zJL is the position of the
p,-th atom in the sample. For large detuning ~ » 'Y the dynamics of the system
can be described by the effective interaction Hamiltonian (see, e.g. [36])
~ ~ g2 + ~ ~ ~ 1 ~
He!! = Ii ~ 1. ai aiJii = M2(SzJz + 4,'nN). (l8.A.2)
i=l 'lq +D
The second term is spin independent and can be omitted. He!! as it is written
above is not Hermitian. In order to make the Hamiltonian to be Hermitian, one
Atomic continuous variable processing and light-atoms quantum interface 263
must add terms that describe interaction with the external system responsible
for relaxation (in our case, the bath of spontaneous emission modes of the atoms
serves as such external system). However, for our purposes, we can neglect this
issue and simply take the Hermitian part of the Hamiltonian (18.A.2). From
the experimental point of view, a very interesting situation is of free-space
atom light coupling. One can show [26, 12] that the effective Hamiltonian
still has the form of Eq. (18.A.2). The anti-Hermitian part of the Hamiltonian
(18.A.2), allows us to write down the expression for the coupling constant 9
through directly measurable quantities: photon-atom atomic cross-section (J"
and the transverse cross-sectional area of the light beam A. The on-resonance
absorption length la is given by ((J" N / AL ) -1. Using the equations of motion for
the electric field that Hamiltonian (18.A.2) results in, we obtain the following
expression for La:
C'Y
la = 2g2N
Comparing the two expressions, we find g2 = ((J"c'Y) / (2LA). If instead of the
Dl transition we consider the D2 one, we would find a similar result, with an
extra factor of ~ present in the expression for the Hamiltonian. The unitary
evolution (; operator is obtained by exponentiating the Hermitian part of the
Hamiltonian: (; = exp( -ifhL/c).
The isotopes of interest of the alkali atoms have non-zero nuclear spin.
In this case we need to express the "average" value of the z-component of
the electronic angular momentum through the total angular momentum of the
atom. The following relation can be obtained [37]:
javg = F F(F + 1) + 3/4 - J(J + 1) = ±F 1
Z Z 2F(F+1) z(I+1/2)
Combining the formulas, we arrive at the Eq.(18.15) for the unitary evolution
operator. It is interesting to note that a more rigorous derivation for atoms
possesing nuclear spin [38] gives the same result as the intuitive procedure of
averaging electronic spin.
References
[1] A. S. Parkins and H. J. Kimble, J. Opt. B 1, 496 (1999); S. Bose, P. L.
Knight, M. B. Plenio, and V. Vedral, Phys. Rev. Lett. 83, 5158 (1999).
[2] A. Kuzmich, K. Mjljlmer, andE. S. Polzik, Phys. Rev. Lett. 79, 4782 (1997).
[3] A. E. Kozhekin, K. Mjljlmer, E. S. Polzik, Phys. Rev. A 62, 033809 (2000).
[4] E. S. Polzik, Phys. Rev. A 59,4202 (1999).
[5] J. Hald, J. L. Sjljrensen, C. Schori, and E. S. Polzik, Phys. Rev. Lett. 83,
1319 (1999).
264 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
[6] A. Kuzmich, L. Mandel, and N. P. Bigelow, Phys. Rev. Lett. 85, 1594
(2000).
[7] Z. Y. Ou, S. F. Pereira, 1. H. Kimble, and K. C. Peng, Phys. Rev. Lett. 68,
3663 (1992).
[8] Ch. Silberhorn, P. K. Lam, O. Weiss, F. Konig, N. Korolkova, and G.
Leuchs, Phys. Rev. Lett.86, 4267 (2001).
[9] C. Schori, J. L. Sj/lrensen, E. S. Polzik, quant-phl0205015.
[10] B. Julsgaard, A. Kozhekin, and E. S. Polzik, Nature 413,400 (2001).
[11] A. Kuzmich and E. S. Polzik, Phys. Rev. Lett. 85, 5639 (2000).
[12] Lu-Ming Duan, J.I. Cirac, P. Zoller, and E. S. Polzik, Phys. Rev. Lett. 85,
5643 (2000).
[13] F. T. Arecchi, E. Courtens, R. Gilmore, and H. Thomas, Phys. Rev. A 6,
2211 (1972).
[14] I. Cirac, unpublished.
[15] M. Kitagawa and M. Ueda, Phys. Rev. Lett. 67, 1852 (1991); Phys. Rev.
A 47, 5138 (1993).
[16] D. J. Wineland, J. J. Bollinger, W. M. !tano, and F. L. Moore, Phys. Rev.
A 46, R6797 (1992);
[17] D. J. Wineland, J. J. Bollinger, W. M. !tano, and D. J. Heinzen, Phys. Rev.
A 50, 67 (1994).
[18] G. S. Agarwal and R. R. Puri, Phys. Rev. A, 41,3782
[19] L.-M. Duan, G. Giedke, J. I. Cirac, P. Zoller, Phys. Rev. Lett. 84, 2722
(2000).
[20] J. Hald and E. S. Polzik. Special Issue: Quantum Coherence and Entan-
glement. Journal of Optics B: Quantum and Semiclassical Optics, 3, S83
(2001).
[21] F. Laloe, M. Leduc, and P. Miguzzi, J. Phys. 30, 277 (1969).
[22] J. L. S!Ilrensen, J. Hald, and E. S. Polzik, Phys. Rev. Lett. 80, 3487 (1998).
[23] W. Happer, Rev. Mod. Phys. 44, 169 (1972).
[24] W. Happer and B. S. Mathur, Phys. Rev. Lett. 18,577 (1967).
[25] S. L. Braunstein, Nature, 394, 47 (1998); S. Lloyd and S. L. Braunstein,
Phys. Rev. Lett. 82, 1784 (1999).
[26] A. Kuzmich, N. P. Bigelow, and L. Mandel, Europhys. Lett. A 42,481
(1998).
[27] K. M!Illmer, Eur. Phys. J. D 5, 301 (1999); Y. Takahashi, K. Honda, N.
Tanaka, K. Toyoda, K.Ishikava, and T. Yabuzaki, Phys. Rev. A 60, 4974
(1999).
Atomic continuous variable processing and light-atoms quantum interface 265
[28] J. -Ph. Poizat, J. -F. Roch, and P. Grangier, Ann. Phys. Fr. 19,265 (1994).
[29] C. H. Bennett, G. Brassard, C. Crepeau, R. Jozsa, A. Peres, and W. K.
Wooters, Phys. Rev. Lett. 70, 1895 (1993).
[30] A. Furusawa, J. L. Sj2!rensen, S. L. Braunstein, C. A. Fuchs, H. J. Kimble,
and E. S. Polzik, Science 282, 706 (1998).
[31] L. Vaidman and N. Yoran, Phys. Rev. A 59, 116 (1999).
[32] C. Schori, B. Julsgaard, J. L. Sj2!rensen, and E. S. Polzik, quant-
phl0203023.
[33] S. L. Braunstein, Phys. Rev. Lett. 80, 4084 (1998).
[34] P. van Loock and S. L. Braunstein, Phys. Rev. Lett. 84, 3482 (2000).
[35] C. Ahn, A. C. Doherty, and A. J. Landahl, quant-phlOl10111.
[36] M. Brune, S. Haroche, J. M. Raimond, L. Davidovich, and N. Zagury,
Phys. Rev. A 45,5193 (1992).
[37] L. D. Landau and E. M. Lifshitz, Quantum mechanics: non-relativistic
theory (Pergamon Press, New York, 1991).
[38] W. Happer and B. S. Mathur, Phys. Rev. 163, 12 (1965).
IV
LIMITS ON QUANTUM INFORMATION
AND CRYPTOGRAPHY
Chapter 19
Samuel L. Braunstein
Informatics, Bangor University, Bangor LL57 I UT, United Kingdom
[email protected]
Arun K. Pati
Institute of Physics, Bhubaneswar-75I005, Orissa, INDIA
Theoretical Physics Division, BARe, Mumbai, INDIA
[email protected]
Abstract In this chapter we briefly discuss some of the limitations to processing discrete
quantum information, such as no-cloning, no-complementing and no-deleting.
The no-cloning principle, in particular, has important practical implications; for
example, quantum cryptography uses it to guarantee security against eavesdrop-
pers.
1. INTRODUCTION
Quantum information differs from classical information in a variety of ways.
Knowing these differences is crucial to understanding the ultimate limits to our
ability to store, process and extract useful information from quantum states.
These are some of the primary tasks for any protocol in information theory.
Among the key differences between classical and quantum information are are
no-cloning [1,2], no-complementing [3] and no-deleting [4]. These limitations
may appear as hurdles, but with the right insight they can be turned into practical
applications. For example, the no-cloning principle lies at the heart of quantum
cryptography.
269
SL Braunstein andA.K. Pati (eds.), Quantum Information with Continuous Variables, 269-276.
© 2003 Kluwer Academic Publishers.
270 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Firstly, one must appreciate the fact that (or the lack of) knowledge about a
quantum state plays an important role. Suppose we are given an arbitrary qubit
I1/;) = ala) + (311) , (19.1)
where a and (3 are complex numbers. Because one can ignore the overall
phase factor we may always choose a to be real. Thus, an arbitrary qubit
may be represented by a point on a two-dimensional sphere with the two real
parameters Band ¢, where a = cos(B/2) and (3 = sin(B/2)exp(i¢) with
a S B S 11" and a S ¢ S 211".
If a qubit is prepared by a third party, then it may be completely unknown to
us. Our lack of knowledge about the preparation procedure translates into our
inability to pinpoint the state on this sphere. In principle, to precisely determine
such a completely unknown state might require a vast amount of information
corresponding to the many bits needed to specify the two parameters B and
¢ [5]. If we had prepared the qubit ourselves then we would have known its
location and hence we would not have lacked any information. In studying
the fundmentallimitations to processing quantum information this preparation
knowledge is important.
More generally, a state may be prepared from a limited alphabet of possible
choices. In particular, the state P is assumed to be selected from one of a set
of states {Pi} each of which occur with respective (and known) probabilities
Pi. If the alphabet (consisting of the set {Pi} and the associated probabilities
Pi) is known, then only the specific choice from this set remains unknown. For
example, for continuous variables, we often consider a restricted alphabet of
the set of coherent states for representing information.
How would this machine act on orthogonal inputs 10) and II)? We would
expect them to transformation according to
10hlEhiAh -+ 10hlohlAo)s
11hlEhiAh -+ 11hllhlAlh . (19.3)
However, these transformations are enough to specify the actions of our
cloner machine on an arbitrary qubit I¢) as input. In particular, by the linearity
of quantum theory we have
Thus, the cloning machine must fail for completely unknown states. If we
knew the state we could rotate it back to 10) or 11) and perform the cloning
operation.
For a restricted alphabet of quantum states such as two non-orthogonal
states from a set S = {¢d, k = 1,2, ... K, the proof of no-cloning theorem
follows just from unitarity of the evolution. Thus the deterministic cloning
of non-orthogonal states is impossible [8, 9, 10]. However, one can give up
either exactness of the clone or determinism of the process in the cloning
operation. It may be shown that approximate cloning is possible using a
unitary operation or that exact cloning is possible with a unitary operation plus
a suitable measurement outcome.
The possibility of producing approximate copies of a quantum state has been
considered by Buzek and Hillery [11]. Here, one demands that all inputs should
be copied equally well (so the performance should have a universal character)
and that the fidelity of cloning should be maximal. For example, the simplest
universal cloner of a qubit making 1 -+ 2 copies can attain an optimal [12]
fidelity F = 5/6. An explicit circuit for achieving this universal cloning has
been determined [13]. More generally, universal cloning has been considered
for 1 -+ M copies [14,15]. The fidelity goes as F1--tM = (2M + 1)/(3M) and
has been shown to be consistent with a no-superluminal-signaling condition
[16, 17]. Further, the fidelity of universal cloner for qudits (d-dimensional
generalizations ofqubits) for N -+ M copies goes as FN--tM(d) = [N(d-l)+
M(N + 1)]/[M(N +d)J. Notice that when d -+ 00 then FN--tM(OO) = N/M.
This was also obtained in Ref. [18] while discussing quantum information
distribution and continuous variable quantum cloning.
In addition, it is possible to design probabilistic cloning machines [19]. A
state selected from a set of non-orthogonal, linearly-independent states S =
{¢d, k = 1,2, ... ,K, can be exactly cloned with a finite probability of
272 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
success. For any two non-orthogonal pairs (i, j) the success probability is
!(Pi +Pj) ~ 1/(1 + I('l/Ji I'l/Jj) I)· Further, it was shown that linearly independent
quantum states can evolve into a linear superposition of multiple copies with a
branch for failure [20]. Probabilistic and deterministic cloning transformations
were shown to be special cases of this operation. A cloning machine that
interpolates between approximate and exact clones has also been proposed
[21].
The impossibility of cloning has been used to argue the impossibility of
completely determining the wavefunction of a single unknown quantum system
[9]. However, we might ask to what extent can one determine the state given a
finite number of copies. This is related to optimal state estimation. For the case
of qubits the best fidelity for recreating a state based on measurement results
from N copies [6] is F = (N + 1)/(N + 2). For ad-dimensional qudits this
optimal fidelity of state estimation [7] is F = (N + 1) / (N + d). Thus, given a
single qubit we can extract out only ~ 's worth of the fidelity. One can see that
if we are given an infinite number of copies will the state estimation fidelity
approaches unity.
Finally, one may ask whether it is possible to clone a quantum state by
supplementing extra information (either quantum or classical)? Classically we
can always determine the state, so no further information is necessary. Surpris-
ingly, however, for cloning a quantum system the supplementary information
must be sufficient to manufacture the clone! This provides us with a stronger
version of the no-cloning theorem [22].
3. NO·COMPLEMENTING PRINCIPLE
We know that a classical bit 0 can be complemented to 1 or vice-versa. In
the quantum world, if a qubit is in a computational state 10) or 11) it can also be
complemented, for example, by applying the Pauli operator ax. However, is it
possible to complement an arbitrary qubit? If we could complement a qubit by
a physical operation, then we would take I'I/J) to I~), where ('l/JI~) = O. So the
complementing operation should be defined by
4. NO-DELETION PRINCIPLE
Classically we may delete one copy against others, uncopying it in a perfectly
reversible manner. However, in the quantum world this is not the case. This
limitation complements no-cloning principle which says that given two (or
more) copies of an unknown quantum state one cannot delete a copy acting
jointly on both the copies [4].
If we could have a quantum deleting machine it would act on two initially
identical qubits in some unknown state I~) and an ancilla in some initial state
IA). This machine is supposed to delete one of the copies and replace it with
some standard qubit IL;). The quantum deleting therefore would yield
(19.6)
where IA1jJ) is the final state of the ancilla. The only solution to this equation is to
swap the second and third qubits. However, this obviously should be excluded
for any sensible definition of a deleting machine, since the information has only
been moved to another location.
Similar to the no-cloning theorem, deletion of a copy from two non-orthogonal
states is impossible [22]. Recently, it has been shown that probabilistic deletion
of copies of linearly independent quantum states is allowed [23, 24, 25]. How-
ever, it remains an open question how to design a universal and approximate
quantum deleting machine.
Next we discuss how these limitations can be turned around into positive
applications such as cryptography.
5. CRYPTOGRAPHY
Cryptography is the art of secretely sending messages from sender to receiver.
Usually the sender and receiver establish a key and using it they can encrypt
and decrypt a message. The security of a cryptic message thus depends on how
well the secrecy of the decryption key is maintained.
The first provably secure encryption scheme was invented by Vernam in
1917 and is ofen called the one-time pad. Here a secret key of random numbers
is added (using modulo arithmetic) to a simple numerical representation of the
message. For each number from the message a new number from the key must
be used. It can be shown that in the absence of knowledge of the key, the
encoded message carries no information about the original. A subtle point in
this result is that the key must not be reused, hence the name one-time pad.
Because of this the main problem with this encoding scheme is the requirement
for distributing a copy of the (long) key to the receiver without its falling into
hostile hands.
It should be noted that one way around the key-distribution problem has
been "solved" by public key cryptosystems. The most well-known is the RSA
274 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
scheme invented by Rivest, Shamir and Adelman [26]. Here the encoding
key is placed in the public domain, whereas the decoding key is kept secret.
Security relies on the computational difficulty of extracting the decoding key
from knowledge of the encoding key alone. It is generally accepted that any
procedure for determining the former relies on factoring the large numbers
appearing in the latter. Because factoring is apparently computationally in-
tractable, the RSA scheme is believed to provide suitable security. Of course,
all this would change if scalable quantum computers capable of running Shor's
algorithm became available.
Public key cryptography solves the key distribution problem at the expense
of absolute security. To guarantee absolute security we must go back to the
Vemam cipher and find a way of guaranteeing the security of distribution of
the key. Here is where quantum mechanics comes in. Quantum cryptography
is really a method of creating pairs of correlated keys between sender and
receiver, without the need for transporting these keys through the intervening
space. Quantum cryptography is sometimes more precisely called quantum
key distribution.
In quantum key distribution, the key is created through a protocol based on
the transfer of quantum states between sender and receiver (Alice and Bob,
respectively). Suppose an eavesdropper (Eve), who wishes to circumvent the
security of this protocol, couples an ancilla to the states transmitted between
Alice and Bob and evolves as
Her intention is to extract information about the states sent by measuring the
final state of the ancilla and distinguishing the outcomes. However, unitarity of
this coupling implies (1fJ111fJ2) = (1fJ111fJ2) (AIIA2). Thus, if the protocol utilizes
non-orthogonal states then the final states of the ancilla become identical. As a
consequence if Eve is to extract information, she can only do so by disturbing
the state transmitted. In essence, this is really nothing more than the no-cloning
principle at work.
For qubit-based quantum cryptography there have been two basic schemes.
Those involving the sending of states from non-orthogonal bases, such as the
original protocol Bennett and Brassard invented in 1984 (called BB84) [27],
and those relying on sharing entanglement between sender and receiver, such
as Ekert's scheme [28]. We shall not discuss the potential advantages of either
scheme here. Below we give a brief discussion to the BB84 protocol.
First Alice generates a random sequence of O's and 1 'So Then she randomly
chooses between two different bases to encode the information. In one basis,
she encodes 0 as 10) and 1 as 11). In the second bassis, she encodes 0 as ~ (10) +
Limitations on discrete quantum information and cryptography 275
11)) and 1 as ~ (10) - 11)) . (Note that these two bases are incompatible.) At
the receiving end, Bob randomly chooses between these bases to make his
measurements. Only when Alice and Bob happen to be using compatible bases
will their bits be correlated. Next, Alice and Bob communicate over a public
channel. They reveal their choice of bases for each round, thus discovering
during which rounds their results should agree. Of these they make a small
selection to "sacrifice" in order to check the fidelity of this correlation. Any
deviation from the ideal is a signal of potential information gained by Eve.
When they are happy that Eve has been excluded, they may use the remaining
(unrevealed) bits from their compatible choice of bases. Thus, this protocol
has managed to create, in separated locations, pairs of random numbers - the
key.
There has been tremendous developments in the area of quantum cryptogra-
phy. For example, to exclude Eve, one may use privacy amplification or even
quantum privacy amplification in order to improve the robustness of the proto-
col [29]. Finally, an absolute prove of the security of BB84 has been given by
Shor and Preskill [30]. For a recent account on discrete quantum cryptography
see Ref. [31].
References
[1] W. K. wootters and W. H. Zurek, Nature (London) 299,802 (1982).
[2] D. Dieks, Phys. Lett. A 92,271 (1982).
[3] V. Buzek and M. Hillery, Phys. Rev. A 60, R2626 (1999).
[4] A. K. Pati and S. L. Braunstein, Nature 404, 164 (2000).
[5] R. Jozsa, in Geometric Issues in the Foundations of Science, Eds. S.
Huggett et aI, Oxford University Press, 1997.
[6] S. Massar and S. Popescu, Phys. Rev. Lett. 74, 1259 (1995).
[7] R. Derka, V. Buzek and A. Ekert, Phys. Rev. Lett. 80, 1571 (1997).
[8] H. P. Yuen, Phys. Lett. A 113, 405 (1986).
[9] G. M. D'Ariano and H. P. Yuen, Phys. Rev. Lett. 76, 2832 (1996).
[10] H. Barnum et aI, Phys. Rev. Lett.76, 2818 (1996).
[11] V. Buzek and M. H. Hillery, Phys. Rev.A54, 1844 (1996).
[12] D. Brub et aI, Phys. Rev. A 57, 2368 (1998).
[13] V. Buzek etaI, Phy. Rev. A. 56, 3446 (1997).
[14] N. Gisin and S. Massar, Phys. Rev. Lett. 79, 2153 (1997).
[15] M. Keyl and R. F. Werner, J. Math. Phys. 40, 3283 (1999).
[16] N. Gisin, Phys. Lett. A 242, 1 (1998).
[17] S. Ghosh, G. Kar and A. Roy, Phys. Lett. A 261, 17 (1999).
276 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
QUANTUM CLONING
WITH CONTINUOUS VARIABLES
Nicolas J. Cerf
Ecole Poly technique, CP 165, Universite Libre de Bruxelles, 1050 Brussels, Belgium
[email protected]
1. INTRODUCTION
Quantum information theory has developed dramatically over the past de-
cade, driven by the prospects of quantum-enhanced communication and com-
putation systems. Among the most striking successes, one finds for example
the discovery of quantum factoring, quantum key distribution, or quantum tele-
portation. Most of these concepts were initially developed for discrete quantum
variables, in particular quantum bits, which have now become the symbol of
quantum information. Recently, however, a lot of attention has been devoted to
investigating the use of continuous-variable systems in quantum informational
or computational processes. Continuous-spectrum quantum variables, for ex-
ample the quadrature components of a light mode, may be easier to manipulate
than quantum bits. It is actually sufficient to process squeezed states of light
into linear optics circuits in order to perform various quantum information pro-
cesses over continuous variables [1]. As reported in the present book, variables
with a continuous spectrum have been shown to be useful to carry out quantum
teleportation, quantum entanglement purification, quantum error correction, or
even quantum computation.
In this Chapter, the issue of cloning a continuous-variable quantum system
will be analyzed, and a Gaussian cloning transformation will be introduced.
Cloning machines, that is, transformations that achieve the best approximate
copying of a quantum state compatible with the no-cloning theorem, have
been a fundamental research topic over the last five years (see e.g., [2] for an
overview). This question is of particular significance given the close connection
between quantum cloning and quantum cryptography: using an optimal cloner
generally makes it possible to obtain a tight bound on the best individual
eavesdropping strategy in a quantum cryptosystem. This provides a strong
277
S.L. Braunstein and A.K. Pati (eds.), Quantum Information with Continuous Variables, 277-293.
© 2003 Kluwer Academic Publishers.
278 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
p the two quadratures of the input mode, and by X and P the corresponding
jointly measured output quadratures, we have
X = x+nx (20.1a)
P=p+np (20.1b)
where nx and np stand for the excess noise that we have on the measured
quadratures. Since we consider a joint measurement, the variables X and P
must commute: they can be viewed respectively as the x and p quadratures of
two distinct modes. Thus, we have
(20.2)
Assuming that the excess noises nx and np are independent of the input quadra-
tures, i.e., [x, n p] = [nx,p] = 0, we get [nx, n p] = -i, implying that nx and
np must obey an uncertainty relation. Specifically, any attempt to measure x
and p simultaneously on a quantum system is constrained by the inequality
(20.3)
where Llni: and Lln~ denote the variances of the excess noises originating from
the joint measurement device. If the variances of the x and p quadratures of
the input state are denoted by bx 2 and bp2, respectively, we thus have for the
variances of the measured values LlX2 = bx 2 + Llni: and LlP2 = bp2 + Lln~.
As a consequence, the Heisenberg uncertainty relation bx bp 2: 1/2 together
with inequality (20.3) implies the relation [9]
LlX LlP 2: 1 (20.4)
where we have used the inequality a 2 + b2 2: 2)a 2b2. Thus, the best possible
joint measurement of x and p with a same precision on both quadratures of a
coherent state (bx 2 = bp2 = 1/2) gives
LlX2 = LlP2 =1 (20.5)
Compared with the vacuum noise, we note that the joint measurement of x and
p effects an additional noise of minimum variance 112, so that the measured
values suffer twice the vacuum noise.
Inequality (20.3) immediately translates into a lower bound on the cloning-
induced noise variance [8]. If we assume that the device that is used in order
to perform the joint measurement of x and p is actually a cloning machine
followed by two measuring apparatuses (x being measured on one clone and
p on the other clone), we conclude that the variance of the noise added by this
cloning machine cannot be lower than 112 in order to comply with Eq. (20.3),
that is
(20.6)
280 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(We require here the same noise level on x and p.) This can also be shown
explicitly by writing the canonical transformation of the c10ner [10]. Denoting
by Xa(b) and Pa(b) the two quadratures ofthe output mode a (resp. b), we have
Xa = x+nx,a (20.7a)
Pa =p+np,a (20.7b)
Xb = x+nx,b (20.7c)
Pb =p+np,b (20.7d)
where x and p are the two quadratures of the input mode and nx/p,a/b stand for
the excess noises. Since the clones are carried by different modes (a and b), we
have [Xa, Pb] = [Xb, Pal = O. Assuming, as before, that the excess noises are
independent of the input mode, we get [nx,a, np,b] = [nx,b, np,a] = -i. This
gives rise to two no-cloning uncertainty relations
which constrain the excess noise variances ~n;/p,a/b of the two clones [5, 10].
Consequently, if the cloning process induces a small position (momentum)
error on the first copy, then the second copy is necessarily affected by a large
momentum (position) error. The Gaussian cloner we will discuss in the next
Session saturates these inequalities and is symmetric in a and b (and in x and
p):
(20.9)
10>
10>
. :
!. -.-----.-----..-.--------------.. ----.---~
ancilla
Figure 20.1 Quantum circuit for the continuous-variable cloning transformation. It consists of
four C-NOT gates preceeded by a preparation stage. Here, the ancillae are prepared in the state
given by Eg. (20.19). See [5, 12].
!i:
The two auxiliary variables must be initially prepared in the joint state
Iw(x,p)) = 1
v27r
to=
1 00
-00
dx' e~Px
.
Ix')lx' + x)
I
(20.12)
are the EPR states (the maximally-entangled states of two continuous variables).
The cloning transformation is defined as
(20.13)
originating from the fact that a continuous C-NOT gate is not equal to its
!I:
inverse. After applying U to the state 11/Ih Ixh,3, we get the joint state
where variables 1 and 2 are taken as the two outputs of the cloner (clones a
and b), while variable 3 (the ancilla) must simply be traced over. This is a
peculiar state in that it can be reexpressed in a similar form by exchanging the
!I:
two clones, namely
II:
with
being the two-dimensional Fourier transform of f (x, p). The reSUlting state of
!I:
the individual clones can then be written as
!I:
Pa = dxdp If(x,p)1 2 b(x,p) 11/1) (1/11 bt(x,p) (20.17a)
which is consistent with tracing Eq. (20.10) over anyone of the clones. Thus,
the clones are affected by position and momentum errors that are distributed
according to If (x, p) 12 and I9 (x, p) 12. A central point here is that interchanging
the two clones amounts to substitute the function f with its two-dimensional
Fourier transform g. This property is crucial as it ensures that the two copies
suffer from complementary position and momentum errors. Indeed, one can
II:
check [5] that the four excess noise variances defined as
L1n~,a= II:
!I:
dxdpp 2 If(x,p)1 2 (20. 18b)
obey the no-cloning inequalities (20.8a) and (20.8b). (Here, we assume that
the first-order moments of If(x,p)1 2 and Ig(x,p)1 2 vanish, that is, the clones
are not biased.)
Quantum cloning with continuous variables 283
IX)2,3 = i:
the two auxiliary variables must be prepared in the state
which is simply the product vacuum state 10hiOh processed by a C-NOT gate
e-iX2P3. The resulting transformation effected by U on an input position state
Ix) is thus given by
Ix}t Ixh,3
Ix + y}t Ix + Z)2 Ix + y + zh (20.20)
where the three variables denote the two clones and the ancilla, respectively.
For an arbitrary input state 11/;), it is readily checked that this transformation
outputs two clones whose individual states are Gaussian distributed with a
variance a 2 = 1/2, namely
Pa = Pb = ..!.
7[
j.J-oo
(Xl dx dp e-(x 2+p2) b(x,p) I1/;) (1/;1 bt(x,p), (20.21)
This cloning fidelity does not depend on a, so this Gaussian cloner copies all
coherent states with the same fidelity 2/3. It can be viewed as the continuous
counterpart of the universal qubit cloner [13], as its cloning fidelity is invariant
under rotations in phase space. The physical origin of the cloning noise be-
comes, however, much more evident in the case of continuous variables: the
Gaussian noise that affects the clones can simply be traced back to the Gaussian
wave function of the two ancillary modes, see (20.19). This suggests that the
noise that inevitably arises when cloning is intrinsically linked to the vacuum
fluctuations of the auxiliary modes.
284 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Note finally that this formalism can easily be extended to the cloning of
squeezed state instead of coherent states [5]. One simply unsqueeze the state
before cloning and then squeeze the clones again. For any value of the squeezing
parameter r, one can then define a Gaussian cloner that copies with fidelity
2/3 all squeezed states of which the same quadrature is squeezed by the same
amount r. In contrast, cloning these squeezed states using the rotation-covariant
cloner defined above results in a fidelity that decreases as r increases.
4. OPTICAL IMPLEMENTATION
It is very instructive to write the cloning transformation in the Heisenberg
picture, that is, following the evolution of the annihilation operators associated
with the modes that are involved. Again, mode 1 denotes the input mode, and
modes 2 and 3 the ancillary modes. Mode l' and 2' stand for the two clones,
while 3' is the ancilla that is traced over after cloning. Here, aj = (x j +ipj) / v'2
stands for the annihilation operator for mode j. We require that the cloning
transformation conserves the mean values, i.e., (a~) = (a~) = (al), so that
the clones are centered on the original coherent state. We also require that the
cloning transformation is covariant under rotations in phase space. It is shown
in [14] that the optimal transformation satisfying these requirements is
, a2 a1
al al+-+- (20.23a)
v'2 v'2
a2 a3t
a'2 al--+- (20.23b)
v'2 v'2
a'3 at + v'2 a 3 (20.23c)
a'
Vacuum 2
j
Pump
"""'"
'3
al
. Linear
Ampli
a out
/
/
BS
a'
(G=2)
a'
3
AnclUa
(20.24)
with mode 3 denoting the idler mode. This amplifier is limited by the quantum
noise so it naturally leads to an optimal cloner. A gain G = 2 is needed since
the cloner doubles the energy by creating two clones with the same energy as
the input state. One then produces these two clones simply by processing the
output signal of the amplifier through a 50:50 phase-free beam splitter,
(20.25)
F MN (20.27)
N,M~MN+M_N
The proof is connected to quantum state estimation theory, the key idea being
that cloning should not be a way of circumventing the noise limitation encoun-
tered in any measuring process. More specifically, concatenating a N -+ M
cloner with a M -+ L cloner results in a N -+ L cloner that cannot be better that
the optimal N -+ L cloner. We then make use of the fact that the excess noise
variance of this N -+ L cloner simply is the sum of the excess noise variances
of the two component cloners [8]. Denoting by O"Jv Mthe excess noise variance
of the optimal N -+ M cloner, we get the inequ~lity O"Jv,L ~ O"Jv,M + O"k,L.
In particular, if L -+ 00, we have
(20.28)
fM ao+JM
VN N
-lat
z'
(20.30a)
J~ -1 ab + az · I¥r (20.30b)
BlankM-N
aM_I
-
,
. OulpUIM-1
,
aN
Blank I OulpulN
aN_I
-
D
,
loputN-I OutputN-I
I
! a, C :
.
loput I Oulpull
a0
lopulO Output 0
Linear
Ancillu
a, '-- AmpJi -
1
Figure 20.3 Implementation of an N -t M continuous-variable cloning machine based on a
phase-insensitive linear amplifier. Here, C stands for the amplitude concentration stage while
D refers to amplitude distribution. Both can be realized using a network of beam-splitters that
achieve a DFf. See [14].
Alice reveals the quadrature she squeezed (and displaced) and Bob rejects the
cases where he measured the wrong quadrature, this discussion being made
over an authenticated public channel (this procedure is known as sifting). The
subset of states that are accepted by Bob then constitutes a Gaussian raw key
(correlated Gaussian data at Alice's and Bob's side). Indeed, denoting as v
the variance of the quadrature that is squeezed by Alice, Bob gets for his mea-
sured quadrature an outcome r' that is Gaussian distributed around r with a
variance v (assuming for the moment that the quantum channel is perfect and
that there is no eavesdropping). If the variance of the random displacements r
imposed by Alice is noted V, then this raw key shared by Alice and Bob can be
viewed as resulting from a Gaussian additive-noise channel characterized by a
signal-to-noise ratio of Vlv.
The maximum amount of shared key bits that can be extracted from this
Gaussian raw key can be analyzed by applying some standard notions of Shan-
non theory for continuous channels (see e.g. [20]). Consider a discrete-time
continuous channel that adds a Gaussian noise of variance v to the signal. If
the input r of the channel is a Gaussian signal of variance V, the uncertainty on
r can be measured by its Shannon entropy h(r) = 2- 1 10g2 (21f e V) bits. Con-
ditionally on r, the output r' is distributed as a Gaussian of variance v, so that
the entropy of r' conditionally on r becomes h(r'lr) = 2- 1 10g2 (21f e v) bits.
Now, the overall distribution of r' is of course the convolution of these two
distributions, i. e., a Gaussian of variance V + v, so that the output entropy
is h(r') = 2- 1 10g2 (21fe (V + v)) bits. According to Shannon theory, the
information processed through this noisy channel r --+ r' can be expressed as
the amount by which the uncertainty on r' is reduced by knowing r, that is
where Vlv is the signal-to-noise ratio. This is Shannon's famous formula for
the capacity of a Gaussian additive-noise channel. It is worth noticing that this
capacity is achieved in the case where the input is distributed as a Gaussian,
which is precisely the case under consideration here.
In the protocol analyzed in [3], the variances v and V are related by the
constraint that Alice's choice of encoding the key into either x or p should be
invisible to a potential eavesdropper. In the first case, Alice applies a Gaussian-
distributed displacement D(r, 0) on a squeezed state whose x quadrature has
a variance v, so that the quadratures x and p of this Gaussian mixture have a
variance V + v and 1/(4v), respectively. In the second case, Alice applies a
displacement D(O, r) on a squeezed state in p, resulting in a Gaussian mixture
with variances 1I (4v) and V + v for x and p. These two Gaussian mixtures are
required to be indistinguishable, which simply translates into the requirement
290 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
1
V+v=- (20.33)
4v
This gives for the information
I = log2 1/2)
( --:;;- (20.34)
which measures the maximum number of key bits that can be extracted asymp-
totically (at the limit of long sequences) per use of the channel. (The factor
1/2 here is just the vacuum noise, so we see that this protocol requires squeez-
ing, that is, v < 1/2.) The actual methods that may be used to discretize the
Gaussian raw key and correct the resulting errors so as to extract a common bit
string are known as reconciliation protocols [21].
Let us now consider the information that is transmitted in the presence of
an eavesdropper. We assume that the eavesdropper (Eve) processes each key
element into a Gaussian cloning machine, keeps one clone, and sends the other
one to Bob. Once the quadrature that contains the key (x or p) is revealed by
Alice and Bob, Eve properly measures her clone. Clearly, Eve needs to use
an asymmetric version of the Gaussian cloner described above as she must be
able to tune the information she gains, and therefore the disturbance she effects
in the transmission. (A possible implementation of this asymmetric Gaussian
cloner is discussed in [17].) Thus, Eve adds some extra noise on the quadrature
encoding the key, which results in a reduced signal-to-noise ratio on Alice-Bob
channel. Remember here, that the quality of the two clones obey a no-cloning
uncertainty relation akin to the Heisenberg relation, implying that the product
of the x-error variance on the first clone times the p-error variance on the second
one remains bounded by (1/2)2; see Eqs. (20.8a) and (20.8b). In particular, if
x and p are treated symmetrically, we have
(20.35)
(20. 36a)
(20.36b)
Quantum cloning with continuous variables 291
which gives
(20.37)
One can then show that lAB + I AE - I :::; 0 by checking that the quantity inside
the logarithm is less or equal to one. This simplifies to the condition
(20.38)
which is indeed true as a consequence of Eq. (20.35) and v < 1/2. Con-
sequently, we have proven that, in this quantum cryptographic protocol, the
no-cloning uncertainty relation translates into an information exclusion princi-
ple [3]
(20.39)
Acknowledgments
I would like to thank: S. L. Braunstein, S.lblisdir, P. van Loock, S. Massar,
and G. Van Assche for their contribution to the work reported on here.
References
[1] S. L. Braunstein. Quantum error correction for communication with linear
optics. Nature 394, 47 (1998).
[2] S. L. Braunstein, V. Buzek, and M. Hillery. Quantum-information distribu-
tors: Quantum network for symmetric and asymmetric cloning in arbitrary
dimension and continuous limit. Phys. Rev. A 63, 052313 (2001).
[3] N. J. Cerf, M. Levy, and G. Van Assche. Quantum distribution of Gaussian
keys using squeezed states. Phys. Rev. A 63, 052311 (2001).
[4] F. Grosshans and P. Grangier. Continuous variable quantum cryptography
using coherent states. Phys. Rev. Lett. 88, 057902 (2002).
[5] N. J. Cerf, A. Ipe, and X. Rottenberg. Cloning of continuous quantum
variables. Phys. Rev. Lett. 85, 1754 (2000).
[6] W. K. Wootters and W. H. Zurek. A single quantum cannot be cloned.
Nature 299, 802 (1982).
[7] D. Dieks. Communication by EPR devices. Phys. Lett. A 92, 271 (1982).
[8] N. J. Cerf and S.lblisdir. Optimal N -to-M cloning of conjugate quantum
variables. Phys. Rev. A 62, 040301 (2000).
[9] E. Arthurs and J. L. Kelly, Jr. On the simultaneous measurement of a pair
of conjugate observables. Bell Syst. Tech. J. 44, 725 (1965).
[10] F. Grosshans and P. Grangier. Quantum cloning and teleportation criteria
for continuous quantum variables. Phys. Rev. A 64, 010301 (2001).
[11] S. L. Braunstein. Error correction for continuous variables. Phys. Rev.
Lett. 80, 4084 (1998).
[12] N. J. Cerf and S. Iblisdir. Universal copying of coherent states: a Gaus-
sian cloning machine. In Quantum Communication, Computing, and
Measurement 3, (Kluwer Academic, New York, 2001), pp. 11-14.
[13] V. Buzek and M. Hillery. Quantum copying: Beyond the no-cloning
theorem. Phys. Rev. A 54, 1844 (1996).
[14] S. L. Braunstein, N. J. Cerf, S. Iblisdir, P. van Loock, and S. Massar. Op-
timal cloning of coherent states with a linear amplifier and beam splitters.
Phys. Rev. Lett. 86,4438 (2001).
Quantum cloning with continuous variables 293
T. C. Ralph
Department of Physics, Centre for Quantum Computer Technology,
University of Queensland, St Lucia 4072, QLD, Australia
[email protected]
Abstract We discuss a quantum key distribution scheme in which small phase and ampli-
tude modulations of quantum limited, CW light beams carry the key information.
We identify universal constraints on the level of shared information between the
intended receiver (Bob) and any eavesdropper (Eve) and use this to make a
general evaluation of the security and efficiency of the scheme.
1. INTRODUCTION
The distribution of random number keys for cryptographic purposes can
be made secure by using the fundamental properties of quantum mechanics
to ensure that any interception of the key information can be detected. This
was first discussed for discrete systems in Refs. [1, 2, 3]. Experimental
demonstrations have been carried out using low photon number, optical sources
[4,5].
The basic mechanism used in quantum cryptographic schemes is the fact that
the act of measurement in quantum mechanics inevitably disturbs the system.
This measurement back-action exists for both discrete and continuous quan-
tum mechanical variables. Thus it is natural to ask if quantum cryptographic
schemes based on continuous variables are possible. There are a number of
practical disadvantages with discrete quantum cryptographic schemes, mainly
associated with the lack of true single photon sources. Also it is of fundamental
interest to quantum information research to investigate links between discrete
variable, single photon phenomena and continuous variable, multi-photon ef-
fects. This has motivated a consideration of quantum cryptographic schemes
using multi-photon light modes [6, 7,8,9, to, 11, 12, 13, 14].
295
S.L. Braunstein andA.K. Pati (eds.), Quantum Information with Continuous Variables, 295-316.
© 2003 Kluwer Academic Publishers.
296 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Most of these schemes use squeezed light [15] in their protocols, either by
producing entanglement from the squeezing [11, 9, 10] or using the squeezing
directly [8, 12]. In contrast to these, schemes based on coherent states have also
been discussed [7, 14]. The signals from which the key material is obtained
are encoded in various ways in the different schemes.
The question of optimum protocols and eavesdropper strategies has been
studied in detail for the single quanta case [16], leading to general proofs
of security for discrete systems [17, 18]. A general proof of the optimum
eavesdropper strategy for individual attacks in a simple continuous variable
scheme was presented in Ref. [11]. Physical implementations saturating this
optimum strategy were discussed in Refs [11, 13, 14]. A general proof of
absolute security for a more sophisticated scheme was presented in Ref. [12].
In this chapter we will analyse in some detail quantum key distribution proto-
cols based on the optical coherent state and squeezed state schemes introduced
in Refs. [7, 11]. Our emphasis will be on specific implementations that Alice
and Bob might use rather than general limits. The particular implementations
have been chosen mostly for their simplicity rather than their optimality. Eve
on the other hand is always assumed to be employing the optimum eavesdrop-
ping strategies allowed by quantum physics [19]. We estimate the efficiency of
the two schemes and hence secure key transmission rates under conditions of
negligible and non-negligible losses.
In Section I we review the encoding of information on light with small am-
plitude and phase modulations and introduce a particular encoding scheme. In
Section II we find the minimum disturbance that an optimum eavesdropping
scheme will introduce. The coherent state cryptographic scheme is introduced
in Section III and the minimum error rates that an optimum Eve will intro-
duce are calculated. In Section IV the concepts of mutual information, data
reconciliation and privacy amplification are introduced and specific examples
are applied to the coherent state scheme. The security and efficiency of the
scheme are evaluated. The squeezed state cryptographic scheme is introduced,
analysed and evaluated in Section V. In section VI we discuss a physical im-
plementation of the optimal eavesdropper strategy and we conclude in Section
VII.
in excess of about a MHz will suffice. That quantum mechanics must impose
limits in this situation is because the amplitude and and phase quadrature
amplitudes of the beam are the analogues of position and momentum variables.
Hence they are continuous, non-commuting variables that exhibit uncertainty
relations.
We can represent our light field via
(21.3)
and
(21.4)
where Vn+ (Vn-) is the amplitude (phase) quantum noise power whilst V/ CV~-)
is the amplitude (phase) signal power. The tilde indicates a Fourier transform.
The amount of information that can be carried on a Gaussian, additive noise,
communication channel, such as we will consider here, depends on the signal
to noise [20]. For a fixed bandwidth, any reduction in the signal to noise will
inevitably lead to increased errors in the transmission. In our cryptographic
scheme signals will be encoded on both quadratures but read out from only
one, randomly chosen. This will force any eavesdroppers to monitor both the
amplitude and phase quadratures simultaneously. For these non-commuting
observables the information that can be obtained in this way is strictly limited by
the generalized uncertainty principle for simultaneous measurements [21,22].
We will discuss this principle in detail in the next section. Here let us consider
a simple example. Suppose we try to observe both quadratures by dividing the
beam in two at a 50:50 beamsplitter and detecting the amplitude quadrature of
298 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
one beam and the phase quadrature of the other. Originally the signal to noises
are given by
(SIN)± = ~: (21.5)
Vn
However the signal to noises detected after the beamsplitter are
(SIN)±' = ( Vn±
szm Vn± + vJ )SIN± = T±SIN± (21.6)
8 = 1 1~
-erfc- -SIN (21.7)
222
Suppose our signal to noise is initially 13 dB. From Eq.21.7 direct detection of
a single quadrature will retrieve its pulse train with a bit error rate of 1%. If
the beam is in a coherent state and we simultaneously detect both quadratures
then Eq.21.6 tells us that the signal to noise is halved. Eq.7 then predicts the
error rate will rise to 5 %.
Alternatively, and more efficiently, signals can be encoded as coherent sig-
nals. Given the ability to phase lock to the signal frequency a mathematical
equivalence exists between the situation of signal side bands on a QNL back-
ground and that of a DC coherent state of the same amplitude. We can thus
represent a coherent signal, of amplitude a by the state ket la}. If the infor-
mation is sent as a Gaussian distribution of such states, then in principle the
Shannon channel capacity [20]
(21.8)
vI + vI OR Vi + Vi
vI
Vi
Figure 21.1 Schematic of general set-up. Alice sends information encoded in the amplitude
(VI) and phase (Vi) spectra. Bob makes measurements of either the amplitude or phase
quadrature. Some additional noise is present on his measurements, vi. Eve does not know
which quadrature Bob will measure thus she needs to be able to extract information about both
quadratures from her intercepted material. This leads to strict bound on the allowed values of
the additional noise which must appear on her measurements (Vi)
v+v-
M M-
>1 (21.9)
where V~ are the measurement penalties for the amplitude (+) and phase
( -) quadratures, normalized to the amplification gain between the system
observables and the measuring apparatus. For example suppose an attempt to
measure the amplitude quadrature variance of a system vt
returned the result
300 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
v+v- >1
E E - (21.10)
Now suppose Bob makes an ideal (no noise added) amplitude measurement
on the beam he receives. In order to satisfy Eq.21.9 it must be true that the
noise penalty carried on the amplitude quadrature of this beam Vi
due to Eve's
intervention, is sufficiently large such that
v+v,-
B E -
>1 (21.11)
Similarly, Bob can also choose to make ideal measurements of the phase
quadrature so we must also have
y+V-
E B
>
-
1 (21.12)
Eqs.21. 10,21. 11,21. 12 set strict quantum mechanical limits on the minimum
disturbance Eve can cause to Bob's information given a particular maximum
quality of the information she receives. This applies regardless of the method
she uses to eavesdrop. Note that quantum memory does not negate the above
results provided we insist that Alice and Bob do not exchange any potentially
revealing classical information until Alice is sure that Bob has received and
measured her signals.
These relations could form the basis of a security analysis of any continuous
variable quantum cryptographic scheme in which a single quantum beam is
exchanged. However the ramifications of a particular level of disturbance will
vary for different schemes. In the following section we will analyse the security
of a very simple scheme based on the exchange of a beam in a coherent state.
subset of Bob's data to be the test and the rest to be the key. For example,
they may pick the amplitude quadrature as the test signal. They would then
compare results for the times that Bob was looking at the amplitude quadrature.
If Bob's results agreed with what Alice sent, to within some acceptable error
rate, they would consider the transmission secure. They would then use the
undisclosed phase quadrature signals, sent whilst Bob was observing the phase
quadrature, to create their key. By randomly swapping which quadrature is key
and which is test throughout the data comparison an increased error rate on
either quadrature will immediately be obvious.
random number
generators
homodyne detection
randomly measuring
either phase or
amplitude signal power
I------;~ - - - - - ------i.~D
coherent AM PM
source
ALICE BOB
Before making a general analysis of security let us first consider some specific
strategies an eavesdropper could adopt. Eve could guess which quadrature Bob
is going to measure and measure it herself. She could then reproduce the digital
signal of that quadrature and impress it on another coherent beam which she
would send on to Bob. She would learn nothing about the other quadrature
through her measurement and would have to guess her own random string of
numbers to place on it. When Eve guesses the right quadrature to measure Bob
and Alice will be none the wiser, however, on average 50% of the time Eve
will guess wrong. Then Bob will receive a random string from Eve unrelated
to the one sent by Alice. These will agree only 50% ofthe time. Thus Bob and
Alice would see a 25% bit error rate in the test transmission if Eve was using
this strategy. This is analogous to the result for single quanta schemes in which
this type of strategy is the most readily available. Another single measurement
strategy Eve could use is to do homodyne detection at a quadrature angle
half-way between phase and amplitude. This fails because the signals become
mixed. Thus Eve can tell when both signals are 0 or both are 1 but she cannot
tell the difference between 1,0 and 0,1. This again leads to a 25% bit error rate.
However, for bright beams it is possible to make simultaneous measurements
of the quadratures, with the caveat that there will be some loss of information.
So a second strategy that Eve could follow would be to split the beam in half,
302 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
measure both quadratures and impose the information obtained on the respective
quadratures of another coherent beam which she sends to Bob. How well will
this strategy work? We performed this calculation at the end of section I using
Eq.21.7. The halving of signal to noise imposed by the 50:50 beamsplitter
means the information Eve intercepts and subsequently passes on to Bob will
have an error probability of 5% (for the particular case of bandwidth limited
binary pulse code modulation). This is clearly a superior strategy and would
be less easily detected. Further more Eve could adopt a third strategy of only
intercepting a small amount of the beam and doing simultaneous detection on
it. For example, by intercepting 16% of the beam, Eve could gain information
about both quadratures with an error rate of 25% whilst Bob and Alice would
observe only a small increase of their error rate to 1.7%. In other words Eve
could obtain about the same amount of information about the key that she could
obtain using the "guessing" strategy, whilst being more difficult to detect.
Now let us analyze this coherent state scheme using Eqs.21.1O,21. 11,21. 12.
We choose to couch our evaluation in terms of bit error rates because they
represent an unambiguous, directly observable measure of the extent to which
Eve can intercept information and the resulting corruption of Bob's information.
This connection will be developed in Section IV. Depending on the particular
technique Eve uses Bob and Alice may be able to gain additional evidence for
Eve's presence by making a more detailed comparison ofthe sent and received
signals. This can only increase the security of the system. By considering a
general limit on error rates we can find a minimum guaranteed security against
eavesdropping regardless of the technique Eve employs.
The signal transfer coefficients for Bob and Eve will be given by
T+ =
(SjN)tve Vi!
E
(SjN)in Vi! + vt
T- (SjN);ve Vi;;:
E = =
(SjN)in Vi;;: + Vi
T+ (SjN)tb Vi!
B
(SjN) in Vi! + V;
T- =
(SjN)"bob Vi;;: (21.13)
B
(SjN)in Vi;;: + Vi
Substituting Eqs.21.13 into Eqs.21.1O,21.11,21.12 and using the factthat Vi! =
1 we find
Ti +TE < 1
Ti +T"B < 1
Tit +TE < 1 (21.14)
Quantum key distribution with continuous variables in optics 303
Eqs.21.14 clearly show that any attempt by Eve to get a good signal to noise
on one quadrature (e.g. Ti -+ 1) results not only in a poor signal to noise in
her information of the other quadrature (e.g. Ti -+ 0) but also a poor signal to
noise for Bob on that quadrature (e.g. Tii -+ 0), making her presence obvious.
This is the general limit of the guessing strategy presented in the last section
and leads to the same error rates.
Because of the symmetry of Bob's readout technique Eve's best approach
is a symmetric attack on both quadratures. Eqs.21.14 then reduces to two
equations
(21.15)
If Eve extracts her maximum allowable signal to noise transfer, T~ = 0.5, then
ideally Bob suffers the same penalty Ti= 0.5. This is the general limit of the
second strategy of the previous section. The same reduction in Bob's signal
to noise occurs as in the specific implementation thus this implementation can
be identified as an optimum eavesdropper strategy for obtaining maximum
simultaneous information about both quadratures.
Eve's best strategy is to intercept only as much information as she can
without being detected. The system will be secure if that level of information
can be made negligible. Suppose, as in the last section, Eve only intercepts
a signal transfer of T~ = .08. From Eq.21.15 this means Bob can receive
at most a signal transfer of Ti = .92. This is greater than the result for the
specific implementation discussed in the last section, thus that implementation
is not an optimum eavesdropper strategy. Using the optimum eavesdropper
strategy the error rates for the specific encoding scheme discussed in the last
section will be: if Eve intercepts information with an error probability of 25%,
then the minimum error rate in Bob's information will be 1.4%.
In Fig. 21.3 we represent the general situation by plotting the minimum
error rate Bob and Alice can observe against the error rate in Eve's intercepted
information using Eq.21.7 and 21.15. An error rate of 50% (i.e. completely
random) represents no information about the data. Two traces are shown,
representing different initial signal to noises in Alice's data. This graph shows
that in principle any incursion by Eve will result in some increase in Bob and
Alice's error rate. However one could also argue that any finite resolution in
Bob and Alice's determination of their error rate will allow Eve to do better
than the random result. In order to assess whether this system can be made
secure we need to introduce the concepts of mutual information and privacy
amplification.
304 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
12
10
Bob's 8
Error
6
Probability
% 4
(a)
2
10 20 30 40 50
Eve's Error
Probability %
Figure 21.3 Minimum allowable error probabilities in the data of Bob and Eve are plotted for
two signal to noise levels of Alice's beam. Trace (a) is for a signal to noise of 13dB whilst trace
(b) is for a signal to noise of lOdB.
(21.16)
where Px and 1 - Px are the probabilities of the two outcomes. Similarly party
2' s data, y, has binary entropy
(21.17)
If the two data strings x and yare random then H{x) = H{y) = 1. Suppose
the error probability between the data strings is 8, then the joint probabilities
are given by Po,o = PI,1 = 1 - 8 and PO,1 = PI,O = 8. Thus we find
Suppose A is Alice's data string, B is Bob's data string and E is Eve's data
string. Maurer has shown [24] that provided H(A : B) > H(A : E) then it
is in principle possible for Alice and Bob to extract a secret key from the data.
Eve's mutual information with this secret key can be made arbitrarily small.
From Eq.21.20 we see that this condition will be satisfied provided Bob's error
rate is less than Eve's. From Fig. 21.3 we see that provided Alice and Bob's
error rate does not exceed 5% for case (a) or 12% for case (b) then secret key
generation is in principle possible. In the following we will look at a simple
specific example of a secret key generation protocol and evaluate its efficiency.
Because of the transmission errors (and possibly the actions of Eve) Alice
and Bob won't share the same data string. However techniques exist for data
reconciliation which allow Alice and Bob to select with high probability a subset
of their data which is error free, whilst giving Eve minimal extra knowledge.
As a simple example Alice and Bob could perform a parity check on randomly
chosen pairs of bits. If the error rate between Bob and Alice is low then the
probability of both bits being wrong is very low. Thus discarding all pairs which
fail the parity check will lead to a big reduction in errors in the shared data
whilst not revealing the values of the individual bits to Eve. A series of parity
checks will lead with high probability to zero errors. Eve can also remove the
pairs that Bob removes and in a worse case scenario may remove up to the same
number of errors as Bob. But if Eve initially had significantly more errors than
Bob then she will still have significant errors after the reconciliation, whilst
Bob and Alice will have virtually none. The data string length will be reduced
by a factor of approximately 1 - 2BB, where BB is Bob's error probability.
In order to reduce Eve's mutual information to a negligible amount the
technique of privacy amplification is employed [25]. This involves the random
hashing or block coding of the reconciled key into a shorter key. As a simple
example Alice and Bob could randomly pick data strings of length n from
the reconciled key and form a new key from the sum, modulo 2, of each n
unit block. It is important that the privacy amplification is "orthogonal" to
the reconciliation protocol. That is none of the pairs used in the parity checks
should appear together in the privacy amplification blocks. The length of the
new key will be reduced by a factor of 1/ n. The error probability in the new
key will be given by
where B is the error probability of the original string. If B ~ 0, as for Bob and
Alice, then this process introduces virtually no errors. But when Eve copies
this process her errors will be "amplified", hopefully te the point where her
mutual information is negligible. Some caution is required in evaluating Eve's
mutual information now. Just as Bob and Alice were able to select a sub-set
306 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
of results they knew were correct in the reconciliation process, so Eve can also
obtain a (smaller) subset of results for which she has greater confidence. We
make the worst case assumption that after privacy amplification Eve is left with
some small probability, Pr, of possessing certain bits that she knows are right,
and a large probability, 1- Pr, of possessing bits which are completely random.
In such a situation it is appropriate to set Eve's mutual information as
H (A : E) = Pr = 1 - 2Bpae (21.22)
Mutual
Information
20 25 30
Block Length
Figure 21.4 Decay of Eve mutual information as a function of the block length, n, in Alice and
Bob's privacy amplification protocol is plotted for two signal to noise levels of Alice's beam.
Trace (a) is for a signal to noise of 13dB whilst trace (b) is for a signal to noise of IOdB. The
solid traces are exponential fits.
from Bob's error rate as if there was no loss, but we must set our error threshold
quite high because the losses will drive up Bob's errors. In this simple approach
it is clear that loss of 50% or more can not be tolerated because Eve's and Bob's
error rates become equal at this point. Indeed as losses approach 50% the
expenditure of resources by Bob and Alice needed to reconcile and privacy
amplify will increase rapidly.
Let us estimate by what factor the length of the final secure key would be
reduced over the length of the original string sent by Alice in a system with
25% loss. Consider an original signal to noise of about lOdB, leading to a base
error rate with 25% loss of 7.7%. Setting as before our maximum error rate
1.5% above the base rate at 9.3% we can bound Eve's error rate at ~ 16.3%.
Bob and Alice sacrifice half their data in this step. Reconciliation will reduce
Bob and Alice's data string by a factor of 0.81 and leave Eve with an error
rate ~7%. If we require that Eve's mutual information be ~0.001 for the
transmission to be considered secure then we find a block length of n = 46 is
required in the privacy amplification step. Thus the secure key will be reduced
by a factor of 0.5 x 0.81 x 0.02 = 0.01. (A similar estimate for the optimum
no loss case gives a reduction factor of 0.025) Data transmission via rf signals
is a mature technology and bit transmission rates of 100 MHz would seem
quite reasonable. Thus secure key transmission rates of a a MHz would seem
practical under these conditions. This is about three orders of magnitude better
than what is presently achievable with single photon schemes. On the other
hand single quanta schemes can tolerate much higher losses [4].
The loss problem can be circumvented by using postselection techniques [28]
or reverse reconcilliation [29]. Both these techniques exploit the asymmetry
in the correlations between Eve and Alice's data and Eve and Bob's data, ie
308 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
the fact that H(B : E) < H(A : E). Using these techniques there is no in
principle loss bound, just as in the discrete case.
ALICE BOB
,----------1 r--------i
I
---~:
+ - : I I
t---c>l'--'-----.----- - - - HD I
I amplitude I
sum or I
phase I
difference I
Figure 21.5 Schematic of squeezed light cryptographic set-up. Sqza and sqzb are phase locked
squeezed light sources. Rna and Rnb are independent random number sources. Bs and pbs
are non-polarizing and polarizing beamsplitters respectively. Half-wave plates to rotate the
polarizations are indicated by >../2 and optical amplification by A. The 7r /2 phase shift is also
indicated. HD stands for homodyne detection system.
The set-up is shown in Fig. 21.5. Once again Alice encodes her number
strings digitally, but now she impresses them on the amplitude quadratures
of two, phase locked, amplitude squeezed beams, a and b, one on each. A
7r /2 phase shift is imposed on beam b and then they are mixed on a 50:50
beam splitter. The resulting output modes, c and d, are given by
c /{(a + ib)
d = /{(a-ib) (21.23)
These beams are now in an entangled state which will exhibit Einstein, Podol-
sky, Rosen (EPR) type correlations [30, 32]. Negligible information about the
signals can be extracted from the beams individually because the large fluctua-
tions of the anti-squeezed quadratures are now mixed with the signal carrying
squeezed quadratures. One of the beams, say c, is transmitted to Bob. The
other beam, d, Alice retains and uses homodyne detection to measure either
its amplitude or phase fluctuations, with respect to a local oscillator in phase
with the original beams a and b. She randomly swaps which quadrature she
measures, and stores the results. Bob, upon receiving beam c, also randomly
Quantum key distribution with continuous variables in optics 309
chooses to measure either its amplitude or phase quadrature and stores his
results. After the transmission is complete Alice sends the results of her mea-
surements on beam d to Bob on an open channel. About half the time Alice
will have measured a different quadrature to Bob in a particular time window.
Bob discards these results. The rest of the data corresponds to times when
they both measured the same quadratures. If they both measured the amplitude
quadratures of each beam Bob adds them together, in which case he can obtain
the power spectrum
where the tilde indicate Fourier transforms. Thus he obtains the data string
impressed on beam a, Vs,a, imposed on the sub-QNL noise floor of beam a,
Vn\ . Alternatively if they both measured the phase quadratures of each beam,
Bob subtracts them, in which case he can obtain the power spectrum
i.e. he obtains the data string impressed on beam b, Vs,b, imposed on the sub-
QNL noise floor of beam b, Vn\ . Thus the signals lie on conjugate quadratures
but both have sub-QNL noise fl'oors. This is the hallmark of the EPR correlation
[33]. As for the coherent state case Alice and Bob now compare some sub-set
of their shared data and check for errors. If the error rate is sufficiently low they
deem their transmission secure and use reconciliation and privacy amplification
on the undisclosed sub-set of their data to produce a secure key.
Consider now eavesdropper strategies. Eve must intercept beam c if she is
to extract any useful information about the signals from the classical channel
(containing Alice's measurements of beam d) sent later. She can adopt the
guessing strategy by detecting a particular quadrature of beam c and then using
a similar apparatus to Alice's to re-send the beam and a corresponding classical
channel later. As before she will only guess correctly what Bob will measure
half the time thus introducing a BER of 25%. Instead she may try simultaneous
detection of both quadratures of beam c. As in the coherent case the noise she
introduces into her own measurement (Vi') and that she introduces into Bob's
(Vf) are in general limited according to Eqs.21.l0,21.11 and 21.12. However
now the consequences of these noise limits on the signal to noise transfers that
Eve and Bob can obtain behave quite differently because the signals they are
trying to extract lie on sub-QNL backgrounds. The maximum signal transfer
310 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
TE+ --
2Vn~a Vn~b + vt (Vn~a + Vn~b)
(Vi + 2Vn~a) vn~b
(21.26)
T+
B
2Vn~a vn~b + VI (Vn~a + Vn~b)
(Vi + 2Vn~a)Vn~b
TB = (21.27)
Vn~a +O.5Vt
Vn\ ,
T-
E (21.28)
T+
B
Vn~a + O.5VI
V/b,
TB = (21.29)
For the squeezed noise floors the same (Vn~a = Vn~b = Vn ) we find the signal
transfers are restricted via
2 1 1
411: ( - - 1)(- - 1) > 1 (21.30)
n T+ T- -
E E
2 1 1
411: ( - - 1)(- - 1) > 1 (21.31)
n T+ T- -
E B
Quantum key distribution with continuous variables in optics 311
2 1 1
4Vn (T+ - l)(T- - 1) ~ 1 (21.32)
B E
It is straightforward to show that a symmetric attack on both quadratures is
Eve's best strategy as it leads to a minimum disturbance in both her and Bob's
measurements. Using this symmetry to simplify Eq.21.30 leads to the following
general restriction on the signal transfer Eve can obtain:
(21.34)
If squeezing is strong then almost any level of interception by Eve will result
in very poor signal transfer to Bob. In Fig. 21.6 we show plots of error rates
of Bob versus minimum error rates of Eve for various levels of squeezing.
In comparison with the coherent scheme (Fig. 21.3) it can be seen that larger
disturbances are caused in Bob's information for the same quality of Eve's
interception. As a numerical example consider the specific encoding scheme
of section I and suppose the squeezing is 10 dB (Vn = 0.1). Assuming no loss
and using the same assumptions as those used to evaluate the coherent scheme
in the last section we find that a secure key of length 0.07 times the original
data string length can be generated. That is an efficiency of about 1/14, to be
compared to the coherent case of 1/40, a clear improvement.
As for the coherent scheme losses of 50% or more cannot be tolerated in
this simple approach however the more sophisticated protocols mentioned in
the coherent state context can also be implemented with squeezing to achieve
high loss operation.
7. TELEPORTATION AS AN OPTIMUM
EAVESDROPPER STRATEGY
It is interesting to consider what physical techniques Eve could use to realize
the optimal attack strategy we have assumed her capable of throughout this
discussion. Firstly she would need to replace the lossy transmission line that
Bob and Alice are using with her own transmission line of negligible loss. Given
that Bob and Alice will presumably employ the most efficient transmission line
312 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
20
90%
Bob's 15
Error
Probability 10 75%
%
5
10 20 30 40 50
Eve's Error
Probability %
Figure 21.6 Minimum allowable error probabilities in the data of Bob and Eve are plotted for
various levels of squeezing.
they can obtain, Eve's job is not trivial. A conceptually simple strategy for
Eve is a completely passive intervention; also known as the beam splitter attack
[14]. That is she takes only the lost light. This she holds in "quantum memory"
until Bob has revealed the bases in which his measurements were made. At
this point she then measures her held portion of the beam in the relevant bases.
In the presence of transmission efficiency TJ Bob's noise penalty will be
± I-TJ
VB = - - (21.35)
TJ
whilst if Eve obtains all the lost light she will have a noise penalty
(21.36)
(21.37)
where lin is the annihilation operator of the input to the teleporter and = 31
VGV1 + VG - 1v~ is the annihilation operator for one of the entangled beams.
The Vi are the vacuum mode inputs to the squeezers, G is the parametric gain
of the squeezers and K > > 1 is the measurement amplification factor. Being a
classical channel simultaneous measurements of both quadratures can be made
without additional penalty thus immediately Eve's measurement penalty is
Vf = 2G-l (21.38)
(21.40)
314 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Eve
/
"
Entanglement V sq
1 + ~~
Aopt = 1 _ V2 (21.41)
sq
8. CONCLUSION
In this chapter we have investigated continuous variable quantum cryptog-
raphy as it could be realized in optics by analysing the security and efficiency
of specific implementations of two systems based on coherent and squeezed
state light respectively. An Eve employing an optimal eavesdropper attack is
assumed throughout. Possible optimal attack strategies that Eve could employ
are outlined.
We find that the coherent scheme can be made secure, but is not very efficient.
None-the-Iess, given the maturity of optical communication technology based
on rf modulation, this system may prove competitive with discrete schemes.
The squeezed state scheme can also be made secure and in principle is more
efficient than the coherent state system.
We have looked at simple protocols throughout this analysis which we
hope clearly illustrate the basic principles. However we have noted that more
sophisticated encoding, reconciliation and privacy amplification techniques
would lead to significant improvements in performance.
Quantum key distribution with continuous variables in optics 315
Acknowledgements
We thank Michael Nielsen, Christine Silberhorn and Philippe Grangier for
useful discussions. This work was supported by the Australian Research Coun-
cil.
References
[1] S. Wiesner, Sigact News, 15, 78 (1983), C. H. Bennett and G. Brassard,
Proc. IEEE Int. Conf. on Computers, Systems and Signal Processing (Ban-
galore), 175 (1984).
[2] C. H. Bennett, Phys. Rev. Lett.68, 3121 (1992).
[3] A. K. Ekart, Phys. Rev. Lett.67, 661 (1991).
[4] W. T. Buttler et al, Phys. Rev. A57, 2379 (1998).
[5] H. Zbinden et aI, Appl.Phys.B 67, 743 (1998).
[6] Y. Mu et aI, Opt.Comm. 123, 344 (1996).
[7] T. C. Ralph, Phys. Rev. A61 010303(R) (1999).
[8] M. Hillery, Phys. Rev. A61 022309 (2000).
[9] M. D. Reid, Phys. Rev. A62 062308 (2000).
[10] Ch .Silberhorn, N .Korolkova and G .Leuchs, Phys. Rev. Lett.88, 167902
(2002).
[11] T. C. Ralph, Phys. Rev. A62 062306 (2000).
[12] D. Gottesman and J. Preskill, Phys. Rev. A63, 022309 (2001).
[13] N. J. Cerf, M. Levy, G. Van Assche, Phys. Rev. A63, 052311 (2001).
[14] F. Grosshans and P. Grangier, Phys. Rev. Lett.88 057902 (2002).
[15] D. F. Walls and G. J. Milburn, Quantum Optics (Springer-Verlag, Berlin,
1994).
[16] C. A. Fuchs and A. Peres, Phys. Rev. A53, 2038 (1996), c. A. Fuchs,
N. Gisin, R. B. Griffiths, C. -So Niu and A. Peres, Phys. Rev. A56, 1163
(1997), I. Cirac and N. Gisin, Phys.Lett.A 229, 1 (1997).
[17] D. Mayers, Advances in Cryptology, Proceedings of Crypto '96, 343
(Springer-Verlag, 1996).
[18] H.-K. Lo and H. F. Chau, Science 283,2050 (1999).
[19] We only explicitly consider individual eavesdropper attacks here.
[20] C. E. Shannon, Bell System Tech. J. 27, 623 (1948).
[21] Y. Yamamoto and H. A. Haus, Rev.Mod.Phys., 58, 1001 (1986).
[22] E. Arthurs and M. S. Goodman, Phys. Rev. Lett.60, 2447 (1988).
316 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Daniel Gottesman
EECS: Computer Science Div., University of California, Berkeley, CA 94720, USA
[email protected]
John Preskill
California Institute of Technology, Pasadena, CA 91125, USA
[email protected]
Abstract We prove the security of a quantum key distribution scheme based on transmission
of squeezed quantum states of a harmonic oscillator. Our proof employs quantum
error-correcting codes that encode a finite-dimensional quantum system in the
infinite-dimensional Hilbert space of an oscillator, and protect against errors that
shift the canonical variables p and q. If the noise in the quantum channel is weak,
squeezing signal states by 2.51 dB (a squeeze factor e T = 1.34) is sufficient
in principle to ensure the security of a protocol that is suitably enhanced by
classical error correction and privacy amplification. Secure key distribution can
be achieved over distances comparable to the attenuation length of the quantum
channel.
1. INTRODUCTION
Two of the most important ideas to emerge from recent studies of quantum
information are the concepts of quantum error correction and quantum,key
distribution. Quantum error correction allows us to protect unknown quantum
states from the ravages of the environment. Quantum key distribution allows
us to conceal our private discourse from potential eavesdroppers.
In fact these two concepts are more closely related than is commonly appre-
ciated. A quantum error correction protocol must be able to reverse the effects
of both bit flip errors, which reflect the polarization state of a qubit about the
x-axis, and phase errors, which reflect the polarization about the z-axis. By
317
S.L. Braunstein and A.K. Pati (eds.), Quantum Information with Continuous Variables, 317-356.
© 2003 Kluwer Academic Publishers.
318 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
reversing both types of errors, the protocol removes any entanglement between
the protected state and the environment, thus restoring the purity of the state.
In a quantum key distribution protocol, two communicating parties verify
that qubits polarized along both the x-axis and the z-axis can be transmitted
with an acceptably small probability of error. An eavesdropper who monitors
the x-polarized qubits would necessarily disturb the z-polarized qubits, while
an eavesdropper who monitors the z-polarized qubits would necessarily disturb
the x-polarized qubits. Therefore, a successful verification test can show that
the communication is reasonably private, and the privacy can then be amplified
via classical protocols.
In quantum key distribution, the eavesdropper collects information by entan-
gling her probe with the transmitted qubits. Thus both error correction and key
distribution share the goal of protecting quantum states against entanglement
with the outside world.
Recently, this analogy between quantum error correction and quantum key
distribution has been sharpened into a precise connection, and used as the basis
of a new proof of security against all possible eavesdropping strategies [18].
Earlier proofs of security (first by Mayers [13, 14], and later by Biham et
al. [3]) made no explicit reference to quantum error correction; nevertheless,
the connection between quantum error correction and quantum key distribution
is a powerful tool, enabling us to invoke the sophisticated formalism of quantum
error-correcting codes in an analysis of the security of quantum key distribution
protocols.
Also recently, new quantum error-correcting codes have been proposed that
encode a finite-dimensional quantum system in the infinite-dimensional Hilbert
space of a quantum system described by continuous variables [8]. In this paper,
we will apply these new codes to the analysis of the security of quantum key
distribution protocols. By this method, we prove the security of a protocol that
is based on the transmission of squeezed quantum states of an oscillator. The
protocol is secure against all eavesdropping strategies allowed by the principles
of quantum mechanics.
In our protocol, the sending party, Alice, chooses at random to send either a
state with a well defined position q or momentum p. Then Alice chooses a value
of q or p by sampling a probability distribution, prepares a narrow wave packet
centered at that value, and sends the wave packet to the receiving party, Bob.
Bob decides at random to measure either q or p. Through public discussion,
Alice and Bob discard their data for the cases in which Bob measured in a
different basis than Alice used for her preparation, and retain the rest. To
correct for possible errors, which could be due to eavesdropping, to noise
in the channel, or to intrinsic imperfections in Alice's preparation and Bob's
measurement, Alice and Bob apply a classical error correction and privacy
Secure quantum key distribution using squeezed states 319
amplification scheme, extracting from the raw data for n oscillators a number
k < n of key bits.
Alice and Bob also sacrifice some of their data to perform a verification test to
detect potential eavesdroppers. When verification succeeds, the probability is
exponentially small in n that any eavesdropper has more than an exponentially
small amount of information about the key. Intuitively, this protocol is secure
because an eavesdropper who monitors the observable q necessarily causes a
detectable disturbance of the complementary observable p (and vice versa).
Since preparing squeezed states is technically challenging, it is important
to know how much squeezing is needed to ensure the security of the protocol.
The answer depends on how heavily the wave packets are damaged during
transmission. When the noise in the channel is weak, we show that it suffices in
principle for the squeezed state to have a width smaller by the factor e- r = .749
than the natural width of a coherent state (corresponding to an improvement
by 2.51 dB in the noise power for the squeezed observable, relative to vacuum
noise). It is also important to know that security can be maintained under
realistic assumptions about the noise and loss in the channel. Our proof of se-
curity applies if the protocol is imperfectly implemented, and shows that secure
key distribution can be achieved over distances comparable to the attenuation
length of the channel. Squeezed-state key distribution protocols may have some
practical advantages over single-qubit protocols, in that neither single-photon
sources nor very efficient photodetectors are needed.
Key distribution protocols using continuous variable quantum systems have
been described previously by others [16, 9, 17], but ours is the first complete
discussion of error correction and privacy amplification, and the first proof of
security against arbitrary attacks.
In §2. we review continuous variable quantum error-correcting codes [8] and
in §3. we review the argument [18] exploiting quantum error-correcting codes
to demonstrate the security of the BB84 quantum key distribution scheme [1].
This argument is extended to apply to continuous variable key distribution
schemes in §4. and §5. Estimates of how much squeezing is required to ensure
security of the protocol are presented in §6. The effects on security of losses
due to photon absorption are analyzed in §7., and §8. contains conclusions.
10) ex: L
8=-00
Iq = (28) . Vi)
00
ex: L
8=-00
Ip = 8· Vi),
00
11) ex: L
8=-00
Iq = (28 + 1) . ../if)
L
00
The operators
e- i (y0i')p , (22.3)
commute with the stabilizer generators and so preserve the code subspace; they
act on the basis eq. (22.1) according to
This code is designed to protect against errors that induce shifts in the values
of q and p. To correct such errors, we measure the values of the stabilizer
generators to determine the values of q and p modulo y7f, and then apply a
shift transformation to adjust q and p to the nearest integer multiples of ../if. If
the errors induce shifts 6.q, 6.p that satisfy
S p -- e-2rrirjJp
, (22.6)
Secure quantum key distribution using squeezed states 321
can be obtained from the <pq = <pp = 0 code by applying the phase space
translation operator
(22.7)
the angular variables <pq and <pp E (-1/2,1/2] denote the allowed values of
q/ ...;rr
and p / ...;rr
modulo an integer. In this code space, the encoded operations
Z and X (which square to the identity) can be chosen to be
(22.8)
The code with stabilizereq. (22.1) can be generalized in a variety of ways [8].
For example, we can increase the dimension of the protected code space, and
we can modify the code to protect against shifts that are asymmetric in q and
in p. If we choose the stabilizer to be
where d is a positive integer and a is a positive real number, then the code has
dimension d and protects against shifts that satisfy
16} ::::::
(4)
:;
1/4
1 00
-00 dq Iq} e-HA~)q2
00
x L e- ~(q_2sy'1r)2 / A~
::::::
s=-oo
- 1
'Jr1/4
1-00
00
dp Ip} e - 21(A2)
q P
2
x L00
e-~(p-sy'1r)2/A~ j (22.11)
s=-oo
If b. q and b. p are small, then in principle these shifts can be corrected with high
probability: e.g, for b. q = b.. p == b., the probability that a shift in q or p causes
an uncorrectable error is no worse than the probability that the size of the shift
exceeds ...fif/2, or
/2
dq e- q2 /6. 2
2b..
< -exp(-1f/4b. 2 ). (22.12)
1f
For the d = 2 code with ai-I, this same estimate of the error probability
applies if we rescale the widths appropriately,
b. q = b.. . a , b.. p = b../ a . (22.13)
We can concatenate a shift-resistant code with an [[n, k, d]] stabilizer quan-
tum code. That is, first we encode (say) a qubit in each of n oscillators; then k
better protected qubits are embedded in the block of n. If the typical shifts are
small, then the qubit error rate will be small in each of the n oscillators, and
the error rate in the k protected qubits will be much smaller. The quantum key
distribution protocols that we propose are based on such concatenated codes.
We note quantum codes for continuous quantum variables with an infinite-
dimensional code space were described earlier by Braunstein [5], and by Lloyd
and Slotine [11]. Entanglement distillation protocols for continuous variable
systems have also been proposed [15, 7].
In the case of the BB84 key distribution invented by Bennett and Bras-
sard [1], the necessary error correction and privacy amplification are entirely
classical. Nevertheless, the formalism of quantum error correction can be use-
fully invoked to show that the error correction and privacy amplification work
effectively [18]. The key point is that if Alice and Bob carry out the BB84
protocol, we can show that the eavesdropper is no better off than if they had
executed a protocol that applies quantum error correction to the transmitted
quantum states. Appealing to the observation that Alice and Bob could have
applied quantum error correction (even though they didn't really apply it), we
place limits on what Eve can know about the key.
(22.14)
where 14>+) is the Bell state (100) + Ill))/Y2; however, the pairs are noisy,
approximating 1<I?(n)) with imperfect fidelity. They wish to extract k < n pairs
that are less noisy.
For this purpose, they have agreed in advance to use a particular [[n, k, dJ] sta-
bilizer code. The code space can be characterized as a simultaneous eigenspace
of a set of mutually commuting stabilizer generators {Mi' i = 1, 2, ... , n - k}.
Each Mi is a "Pauli operator," a tensor product of n single-qubit operators where
324 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
The operations {Xa, Za, a = 1,2, ... , k} acting on the encoded qubits are
Pauli operators that commute with all of the Mi.
The Bell state 11>+) is the simultaneous eigenstate with eigenvalue one of
the two commuting operators X A (8) X Band Z A (8) Z B (where subscripts A and
B indicate whether the operator acts on Alice's or Bob's qUbit). Thus the state
1<I>(n)) is the simultaneous eigenstate with eigenvalue one of the commuting
operators
Now suppose that Alice and Bob both measure the n - k commuting Mi'S. If
the state they measure is precisely 1<I>(n)), then Alice and Bob obtain identical
measurement outcomes. Furthermore, since their measurements do not disturb
the encoded operations Xa and Za, their measurement would prepare the
encoded state 1<I>(k)) == I<i>+)®k, the encoded state with
3.2 VERIFICATION
If the initial pairs are too noisy, either because of the intervention of an
eavesdropper or for other reasons, then the purification protocol might not
succeed. Alice and Bob need to sacrifice some of their EPR pairs to verify that
purification is likely to work. If verification fails, they can abort the protocol.
Under what conditions will purification succeed? If their pairs were perfect,
each would be in the state Iif>+), the simultaneous eigenstate with eigenvalue one
of the two commuting observables X ® X and Z ® Z. Suppose for a moment,
that each of the pairs is a simultaneous eigenstate of these observables (a Bell
state), but not necessarily with the right eigenvalues: in fact no more than tx
of the n pairs have X ® X = -1, and no more than tz of the n pairs have
Z ® Z = -1. Then, if Alice and Bob use a stabilizer code that can correct
up to t z bit flip errors and up to t x phase errors, the purification protocol will
work perfectly - it will yield the encoded state 1<I>(k)) = I¢+)®k with fidelity
F=1.
Now, the initial n pairs might not all be in Bell states. But suppose that Alice
and Bob were able to perform a Bell measurement on each pair, projecting it
onto a simultaneous eigenstate of X ® X and Z ® Z. Of course, since Alice and
Bob are far apart from one another, they cannot really do this Bell measurement.
But let's nevertheless imagine that they first perform a Bell measurement on
each pair, and then proceed with the purification protocol. Purification works
if the Bell measurement yields no more than tx pairs with X ® X = -1 and
no more than t z pairs with Z ® Z = -1. Therefore, ifthe initial state of the
n pairs has the property that Bell measurement applied to all the pairs will,
with very high probability, produce pairs with no more than t z bit flip errors
and no more than t x phase errors, then we are assured that Bell measurement
followed by purification will produce a very high fidelity approximation to the
encoded state 1<I>(k)).
But what if Alice and Bob execute the purification protocol without first per-
forming the Bell measurement? We know that the purification works perfectly
applied to the space 'Hgood spanned by Bell pairs that differ from Iif>+)®n by
no more than t z bit flip errors and no more than t x phase errors. Let II denote
the projection onto 'Hgood . Then if the protocol is applied to an initial density
operator p of the n pairs, the final density operator p' approximates 1<I>(k)) with
fidelity
(22.18)
(22.19)
(22.20)
Since the states IWgood)SER and IWbad)SER are orthogonal, the unitary re-
covery operation USR ® IE maps them to states IW~ood)SER and IW~ad)SER
that are also orthogonal to one another. Furthermore, since recovery works
perfectly on the space 1-lgood' we have
, - (k)
IWgood)SER = leI> )s ® ljunk)ER, (22.21)
ER(junkljunk)ER = SER(W~oodlw~ood)SER
= SER(WgoodIWgood)SER = tr(ITp) . (22.22)
where
(22.24)
eq. (22.18) then follows. The key point is that, because of eq. (22.21), and
because IW~ood)SER and IW~ad)SER are orthogonal, there is no "good-bad"
cross term in eq. (22.22).
Our arguments so far show that Alice and Bob can be assured that entangle-
ment purification will work very well if they know that it is highly unlikely that
Secure quantum key distribution using squeezed states 327
more than t z bit flip errors or more than t x phase errors would have been found
if they had projected their pairs onto the Bell basis. While they have no way
of directly checking whether this condition is satisfied, they can conduct a test
that, if successful, will provide them with high statistical confidence. We must
now suppose that Alice and Bob start out with more than n pairs; to be definite,
suppose they have about 2n to start, and that they are willing to sacrifice about
half of them to conduct their verification test. Alice randomly decides which
pairs are for verification (the "check pairs") and which are for key distribution
(the "key pairs"), and for each of her check qubits, she randomly decides to
measure either X or Z. Then Alice publicly announces which are the check
pairs, whether she measured X or Z on her half of each check pair, and the
results of those measurements (in addition to the results of her measurements
of the stabilizer generators).
Upon hearing of Alice's choices, Bob measures X or Z on his half of each of
the check pairs; thus Alice and Bob are able to measure X (8) X on about half of
their check pairs, and they measure Z (8) Z on the remaining check pairs. Now
since the check pairs were randomly chosen, the eavesdropper Eve has no way
of knowing which are the check pairs, and she can't treat them any differently
than the key pairs; hence the measured error rate found for the check pairs will
be representative of the error rate that would have been found on the key pairs if
Alice and Bob had projected the key pairs onto the Bell basis. Therefore, Alice
and Bob can use their check data and classical sampling theory to estimate how
many bit flip and phase errors would have been expected if they had measured
the key pairs.
For example, in a sample of N pairs, suppose that if Alice and Bob both
measured Z for all the pairs, a fraction p of their measurements would disagree,
indicating bit flip errors. Then if they randomly sample M < N of the pairs,
the probability distribution for the number M (p - c:) of errors found would be2
P(c:) <exp(-Mc: 2 /2p(1-p)). (22.25)
If Alice and Bob have no a priori knowledge of the value of p, then by Bayes'
theorem, the conditional probability that the total number of errors in the
population is pN, given that there are pzM errors in the sample, is the same as
the probability that there are pzM errors in the sample given that there are pN
errors in the total popUlation. Writing p = pz + c:, the number of errors on the
N-MuntestedpairsisNp-Mpz = (N-M)pz+Nc: = (N-M)·(pz+c:'),
where c:' = N c: / (N - M). Expressing P (c:) in terms of c:' we find
, ( M(N - M)2c:,2)
P(c:) < exp - 2N2 pz (1 - pz) , (22.26)
a bound on the probability that the fraction of the untested pairs with errors is
larger than pz + c:'. In particular, ifthey test about M = n/2 pairs for bit flip
328 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
errors out of a total of about N = n + n/2 pairs, the probability that a fraction
p z + c' of the remaining N - M = n pairs have bit flip errors is
and since PAB is very nearly pure, S(PAB) and J(AB; E) are very close to zero.
Specifically, if the fidelity of PAB is F = 1 - 8, then the largest eigenvalue of
PAB is at least 1 - 8. For a system with dimension D, the density matrix with
largest eigenvalue 1 - 8 that has the maximal Von Neumann entropy is
for which
S(Pmax) = -(1 - 6) log2(1- 6) - 610g2(6/(D - 1))
Taking D = 22k (the total dimension of Alice's and Bob's code spaces), we
conclude that
Finally, we have shown that if the verification test succeeds, then with proba-
bilityexponentially close to one (the probability that the error rate inferred from
the check sample is not seriously misleading), Eve's mutual information with
the key is exponentially small (because the state of the key bits approximates
I<i>(k)) with fidelity exponentially close to one). This proof of security applies
to any conceivable eavesdropping strategy adopted by Eve.
The proof relies on the ability of quantum error-correcting codes to reverse
the errors caused by interactions between the key pairs and Eve's probe. Hence
it may seem odd that the proof works for arbitrary attacks by Eve, since
quantum error correction works effectively only for a restricted class of error
superoperators. Specifically, the error superoperator acting on a block of n
qubits can be expanded in terms of a basis of "Pauli error operators," where in
each term of the expansion bit flip errors and/or phase errors are inflicted on
specified qubits within the block. The encoded quantum information is well
protected only if the error superoperator has nearly all of its support on Pauli
operators that can be corrected by the code, e.g., those with no more than tz
bit flip errors and t x phase errors.
If Eve's probe interacts collectively with many qubits, it may cause more
bit flip or phase errors than the code can correct. But the crucial point is that,
with high probability, an attack that causes many errors on the key bits will
also cause many errors on the check bits, and Alice and Bob will detect Eve's
presence.
Alice also decides at random which of her qubits will be used for key
distribution and which will be used for verification. For each of the check
bits, she decides at random whether to send an X eigenstate (with random
eigenvalue) or a Z eigenstate (with random eigenvalue).
Bob receives the qubits sent by Alice, carefully deposits them in his quantum
memory, and publicly announces that the qubits have been received. Alice then
publicly reveals which qubits were used for the key, and which qubits are the
check qubits. She announces the stabilizer eigenValues that she chose to encode
her state, and for each check qubit, she announces whether it was prepared as
an X or Z eigenstate, and with what eigenvalue.
Once Bob learns which qubits carry the encoded key information, he mea-
sures the stabilizer operators and compares his results with Alice's to obtain
a relative error syndrome. He then performs error recovery and measures the
encoded state to decipher the key.
Bob also measures the check qubits and compares the outcomes to the values
announced by Alice, to obtain an estimate of the error rate. If the error rate is
low enough, error recovery applied to the encoded key bits will succeed with
high probability, and Alice and Bob can be confident in the security of the key.
If the error rate is too high, Bob informs Alice and they abort the protocol.
As described so far, the protocol requires that Alice and Bob have quantum
memories and quantum computers that are used to store the qubits, measure
stabilizer generators, and correct errors. But ifthey use a stabilizer code of the
CSS (Calderbank-Shor-Steane) type [6, 19], then the protocol can be simplified
further. The crucial property of the CSS codes is that there is a clean separation
between the syndrome information needed to correct bit flip errors and the
syndrome information needed to correct phase errors.
A CSS quantum stabilizer code is associated with a classical binary linear
code CIon n bits, and a subcode C2 C C 1 . Let HI denote the parity check
matrix of C 1 and H2 the generator matrix for the code C2 (and hence the parity
check matrix of the dual code Cf). The stabilizer generators of the code are
of two types. Associated with the i th row of the matrix HI is a "Z -generator,"
the tensor product of I's and Z's
(22.32)
and associated with the ith row of H2 is an "X -generator," the tensor product
of I's and X's
(22.33)
From measurements of the Z generators, bit flip errors can be diagnosed, and
from measurement of the X generators, phase errors can be diagnosed.
The elements of a basis for the code space with eigenvalues of stabilizer
generators
(22.35)
(22.36)
Thus, to distribute the key, Alice chooses x and z at random, encodes one of
the 11fJ (v )) X,z 's, and sends the state to Bob. After Bob confirms receipt, Alice
broadcasts the values of x and z. Bob compares Alice's values to his own
measurements of the stabilizer generators to infer a relative syndrome, and
he performs error correction. Then Bob measures Z of each of his n qubits,
obtaining a bit string v + w + x. Finally, he subtracts x and applies H2 to
compute H 2 v, from which he can infer the coset represented by v and hence
the key.
Now notice that Bob extracts the encoded key information by measuring Z
of each of the qubits that Alice sends. Thus Bob can correctly decipher the
key information by correcting any bit flip errors that occur during transmission.
Bob does not need to correct phase errors, and therefore he has no use for the
phase syndrome information; hence there is no need for Alice to send it.
Without in any way weakening the effectiveness of the protocol, Alice can
prepare the encoded state 11fJ (v ) ) x ,z, but discard her value of z, rather then
transmitting it; thus we can consider the state sent by Alice to be averaged over
the value of z. Averaging over the phase (-1 y.w destroys the coherence of
the sum over w E O2 in 11fJ(v))x,z; in effect, then, Alice is preparing n qubits
as Z eigenstates, in the state Iv + w + x), sending the state to Bob, and later
broadcasting the value of x. We can just as well say that Alice sends a random
string u, and later broadcasts the value of u + v. Bob receives u + e (where e
has support on the bits that flip due to errors) extracts v + e, corrects it to the
nearest 01 codeword, and infers the key, the coset v + O2 •
Alice and Bob can carry out this protocol even if Bob has no quantum
memory. Alice decides at random to prepare her qubits as X or Z eigenstates,
with random eigenvalues, and Bob decides at random to measure in the X
or Z basis. After public discussion, Alice and Bob discard the results in the
332 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
cases where they used different bases and retain the results where they used
the same basis. Thus the protocol we have described is just the BB84 protocol
invented by Bennett and Brassard [1], accompanied by classical error correction
(adjusting v + e to a C1 codeword) and privacy amplification (extracting the
coset v + C2).
What error rate is acceptable? In a random CSS code, about half of the
n - k generators correct bit flips, and about half correct phase flips. Suppose
that the verification test finds that bit flip errors (ZA ® ZB = -1) occur with
probability pz and phase errors (XA ® XB = -1) occur with probability px.
Classical coding theory shows that a random CSS code can correct the bit flips
with high probability if the number of typical errors on n bits is much smaller
than the number of possible bit flip error syndromes, which holds provided that
k
R= - < 1-2H2(pz) ,
n
k
R =- < 1 - 2H2(PX) . (22.38)
n
This upper bound on R crosses zero at pz (or px) = .1100. We conclude that
secure key distribution is possible if p x,z < 11 %.
The random coding argument applies if the errors in the key qubits are
randomly distributed. To assure that this is so, we can direct Alice to perfonn
a random pennutation of the qubits before sending them to Bob. After Bob
confinns receipt, Alice can broadcast the pennutation she perfonned, and Bob
can invert it.
Again, the essence of this argument is that the amount of infonnation that an
eavesdropper could acquire is limited by how successfully we could have carried
out quantum error correction if we had chosen to - and that this relation holds
irrespective of whether we really implemented the quantum error correction or
not.
Other proofs of the security of the BB84 protocol have been presented [13, 3],
which don't make direct use of this connection with quantum error-correcting
codes. However, these proofs do use classical error correction and privacy
amplification, and they implicitly exploit the structure of the CSS codes.
Secure quantum key distribution using squeezed states 333
Similarly, to prepare an X -state she measures in the basis {I +), I-)}, sending
one of
A(+lpABI+)A
P+ tr (A(+lpABI+)A) ,
A(-lpABI-)A
P- = tr (A(-lpABI-)A) ,
(22.41)
Prob(+) tr(A(+lpABI+)A) ,
Prob(-) tr (A(-lpABI-)A) . (22.42)
Unless the state PAB is precisely the pure state Icf;+), Alice's source isn't doing
exactly what it is supposed to do. Depending on how PAB is chosen, the source
might be biased; for example it might send Po with higher probability than Pl.
And the states Po and PI need not be the perfectly prepared 10) and 11) that the
protocol calls for.
Now suppose that Alice's source always emits one of the states Po , PI, P+, P_,
and that after the qubits emerge from the source, Eve is free to probe them any
way she pleases. Even though Alice's source is flawed, Alice and Bob can
perform verification, error correction, and privacy amplification just as in the
BB84 protocol. To verify, Bob measures Z or X, as before; if he measures Z,
say, they check to see whether Bob's outcome 10) or 11) agrees with whether
Alice sent Po or PI (even though the state that Alice sent may not have been a
Z eigenstate). Thereby, Alice and Bob estimate error rates p z and p x. If both
error rates are below 11 %, then the protocol is secure.
We emphasize again that the security criterion p x, P z < 11 % applies not
to all sources, but only to the restricted class of imperfect sources that can
be simulated by measuring half of a (possible noisy) entangled state. To give
an extreme example of a type of source to which the security proof does not
apply, suppose that Alice always sends the Z -state 10) or the X -state 1+).
Secure quantum key distribution using squeezed states 335
Clearly the key distribution protocol will fail, even if Bob's bits always agree
with Alice's! Indeed, a source with these properties cannot be obtained by
measuring half of any two-qubit state PAB. Rather, if the source is obtained by
such a measurement, then a heavy bias when we send a Z-state would require
that the error probability be large when we send an X -state.
which also commute with the stabilizer generators that Alice measured. Thus
Alice's measurement has prepared an encoded Bell pair in the code space
labeled by (cpq, cpp), the state
Of course the initial EPR pair shared by Alice and Bob might be imperfect,
and then the encoded state produced by Alice's measurement will also have
errors. But if the EPR pair is not too noisy, they can correct the errors with high
probability. Alice broadcasts her measured values of the stabilizer generators
to Bob; Bob also measures the stabilizer generators and compares his values to
those reported by Alice, obtaining a relative syndrome
(22.49)
That is, the relative syndrome determines the value of qA - qB (mod Vi), and
PA + PB (mod y'i). Using this information, Bob can shift his oscillator's q and
P (by an amount between -Vi/2 and Vi/2) to adjust qA - qB (mod Vi), and
PA + PB (mod y'i) both to zero. The result is that Alice and Bob now share a
bipartite state in the code subspace labeled by (cpq, cpp).
If the initial noisy EPR state differs from the ideal EPR state only by relative
shifts of Bob's oscillator relative to Alice's that satisfy I~ql, l~pl < Vi/2,
then the shifts will be corrected perfectly. And if larger shifts are highly
unlikely, then Alice and Bob will obtain a state that approximates the desired
encoded Bell pair 14>+) with good fidelity. This procedure is a "distillation"
protocol in that Alice and Bob start out with a noisy entangled state in a tensor
product of infinite dimensional Hilbert spaces, and "distill" from it a far cleaner
entangled state in a tensor product of two-dimensional subspaces.
Once Alice and Bob have distilled an encoded Bell pair, they can use it to
generate a key bit, via the usual EPR key distribution protocol: Alice decides
at random to measure either X or Z, and then publicly reveals what she chose
to measure but not the measurement outcome. Bob then measures the same
observable and obtains the same outcome - that outcome is the shared key bit.
How do they measure X or Z? If Alice (say) wishes to measure Z, she
can measure q, and then subtract cPq from the outcome. The value of Z is
determined by whether the result is an even (Z = 1) or an odd (Z = -1)
multiple of Vi. Similarly, if Alice wants to measure X, she measures P and
subtracts CPP - The value of X is determined by whether the result is an even
(X = 1) or an odd (X = -1) multiple of Vi.
Imperfections in the initial EPR pairs are inescapable not just because of
experimental realities, but also because the ideal EPR pairs are unphysical
nonnormalizable states. Likewise, the stabilizer operators cannot even in prin-
ciple be measured with arbitrary precision (the result would be an infinite bit
Secure quantum key distribution using squeezed states 337
string), but only to some finite m-bit accuracy. Still, if the EPR pairs have rea-
sonably good fidelity, and the measurements have reasonably good resolution,
entanglement purification will be successful.
To summarize, Alice and Bob can generate a shared bit by using the contin-
uous variable code for entanglement purification, carrying out this protocol:
To carry out this protocol, Alice requires sophisticated tools that enable her to
prepare the approximate codewords, and Bob needs a quantum memory to store
the state that he receives until he hears Alice's classical broadcast. However,
we can reduce the protocol to one that is much less technically demanding.
When Bob extracts the key bit by measuring (say) q, he needs Alice's value
of q modulo ..,fi, but he does not need her value of the other stabilizer generator.
Therefore, there is no need for Alice to send it; surely, the eavesdropper will
be no better off if Alice sends less classical information. If she doesn't send
the value of Sp, then we can consider the protocol averaged over the unknown
value of this generator. Formally, for perfect (nonnormalizable) codewords the
density matrix describing the state that is accessible to a potential eavesdropper
then has a definite value of Sq but is averaged over all possible values of Sp - it is
a (nonnormalizable) equally weighted superposition of all position eigenstates
with a specified value of q mod ...(if; e.g. in the case where Alice prepares a Z
eigenstate, we have
p(¢q, Z = 1)
ex L Iq = (28 + ¢q)v-rr)(q = (28 + ¢q)v-rrl ,
s
p(¢q, Z = -1)
ex L Iq = (28 + 1 + ¢q)v-rr)(q = (28 + 1 + ¢q)J7f1 .
s
(22.50)
they distill k encoded Bell pairs of much better fidelity, and then generate a key
by measuring the encoded Bell pairs.
By once again following the chain of reductions recounted in §3. and §4.,
we arrive at an equivalent protocol involving transmission of squeezed states.
The complete protocol, including verification, error correction, and privacy
amplification, becomes:
Continuous-variable QKD
Our analysis of the BB84 protocol indicates that the squeezed state protocol
is secure provided that p z and p x are both below 11 %, and assuming that
Alice and Bob scramble and unscramble the oscillators (by applying a random
permutation and its inverse).
However, as noted in §3.4, the proof and the security criterion Pz, Px < 11%
apply only if Alice's source can be simulated by measuring half of an entangled
state of two oscillators. In particular, we may imagine that Alice has many
pairs of oscillators identically prepared in the state PAB, and that she prepares
the state that she sends to Bob by measuring oscillator A. When she measures
in the q basis, she sends the state
(22.51)
with probability
(22.52)
and when she measures in the p basis, she sends the state
(22.53)
with probability
(22.54)
Thus, the states that Alice sends need not be perfect position or momentum
eigenstates for the proof of security to work, and Alice's source might even
have a bias so that the raw key bit carried by an oscillator is more likely to be
a 0 than a 1. Still, for a source of this type, if Alice and Bob verify that the
error rate for the raw key bits is below 11 % in both bases, then the protocol is
provably secure. We will discuss examples in §6. and §7.
Intuitively, the squeezed state protocol is secure because the eavesdropper
cannot monitor the value of q (or p) transmitted without introducing a detectable
disturbance in the complementary observable p (or q). As shown in Fig. 22.1,
the Wigner functions of the signal states squeezed in p and in q overlap, so that
the states cannot be reliably distinguished.
6. GAUSSIAN STATES
Perfectly squeezed states (position or momentum eigenstates) are unphysical
nonnormalizable states, so the protocol will actually be carried out with imper-
fectly squeezed states. Furthermore, engineering a source that produces highly
squeezed states would be quite technically demanding. How much squeezing
342 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
r------+q
Figure 22.1 One-sigma contours of the Wigner functions for typical squeezed states used in
the quantum key distribution protocol, with squeeze factor Li = e- r = 1/2. The signal states
squeezed in p and in q overlap with one another, preventing Eve from learning about one without
disturbing the other.
is really needed for the protocol to be secure? A related question is, how must
we choose the probability distributions Ppos (q) and Pmom (p) that govern the
center of the squeezed state?
We will analyze the most favorable case, in which the squeezed states are
Gaussian wave packets and the probability distributions are also Gaussian. We
will begin again with a description of how the code is used for entanglement
purification, but where Alice and Bob start with many copies of a Gaussian
entangled pair of oscillators that is an approximate eigenstate of qA - qB and
PA + PB · If we imagine that Alice measures half of each pair before she
sends the other half to Bob, then we obtain a protocol in which Alice sends
imperfectly squeezed states governed by a particular probability distribution.
The initial Gaussian entangled state of the two oscillators is
(22.55)
Secure quantum key distribution using squeezed states 343
(22.56)
we may assume without loss of generality (changing the sign of the position and
momentum of Bob's oscillator if necessary), that 0 < ~2 ~ 2. In the limiting
case ~2 = 2, I'¢'(~))AB becomes the product of two oscillator vacuum states.
For ~ 2 < 2, it is an entangled state. The amount of entanglement shared
between the oscillators, in "ebits," is defined as
(22.57)
where
(22.59)
In this entangled state, if Alice measures the position of her oscillator and
obtains the outcome qA, she prepares for Bob the Gaussian state
where
(22.61)
and
- 2 ~2
~ = 1 + !~4 . (22.62)
4
(22.63)
344 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
and we can easily see from eq. (22.54) that if Alice and Bob both measure q,
then the difference of their outcomes is governed by the probability distribution
(22.65)
and similarly for px (the probability that X ® X = -1). For the values of Ll
that are typically of interest (e.g. Ll < 1), the error probability is dominated by
values of qA - qB (or PA + PB) lying in the range [..Jff/2, 3Vir/2], so that the
estimate of the error probability can be sharpened to
2 J3..;7i /2 2 b. 2
pz,px rv - - dq e- q / • (22.66)
V7rLl2 ..;7i/2
After error correction and measurement in the encoded Bell basis, the initial
bipartite pure state of two oscillators, with entanglement E given by eq. (22.57)
and (22.59), is reduced to a bipartite mixed state, diagonal in the encoded Bell
basis, with fidelity F = (1 - Pz ) (1 - Px); this encoded state has entanglement
of formation [2]
(22.67)
distill qubits of arbitrarily good fidelity at a finite asymptotic rate provided that
pz and Px are both below 11 %; from eq. (22.66) we find that this condition is
satisfied for ~ < .784 (which should be compared with the value ~ = v'2 cor-
responding to a product of two oscillators each in its vacuum state). Thus secure
EPR key distribution is possible in principle with two-mode squeezed states
provided that the squeeze parameter r satisfies r > -loge(.784/v'2) = .590;
from eq. (22.57) and (22.67), ~ = .784 corresponds to E = 1.19 ebits carried
by each oscillator pair, which is reduced by error correction and encoded Bell
measurement to E = .450 ebits carried by each of the encoded Bell pairs.
Now consider the reduction of this entanglement distillation protocol to a
protocol in which Alice prepares a squeezed state and sends it to Bob. In the
squeezed-state scheme, Alice sends the state 11/J(qA)} with probability P(qA).
The width A of the state that Alice sends is related to the parameter ~ appearing
in the estimated error probability according to
(22.68)
, ,
, ,
,,
,,
,,
.,
,,
q or p
-4 4
Figure 22.2 Probability distributions for the squeezed quantum key distribution protocol, with
squeeze factor Ii = 1/2. The dotted line is the probability distribution P (a Gaussian with
variance (1/21i 2) . (1 -Ii 4) that Alice samples to determine the center of the squeezed signal
that she sends. The solid lines are the probability distributions in position or momentum of the
squeezed states (Gaussians with variance Ii 2 /2, shown with a different vertical scale than P)
centered at -,ji, 0, and ,ji. The intrinsic error probability due to imperfect squeezing (prior
to binary error correction and privacy amplification) is 1.2%.
If this is the only consequence of the noise, the squeezing exiting the channel
should still satisfy !:::. < .784 for the protocol to be secure, as we discuss in
more detail in §7. Otherwise, the errors due to imperfect squeezing must be
added to errors from other causes to determine the overall error rate.
So far we have described the case where the p states and the q states are
squeezed by equal amounts. The protocol works just as well in the case of
unequal squeezing, if we adjust the error correction procedure accordingly.
Consider carrying out the entanglement distillation using the code with general
parameter Q rather than Q = 1. The error rates are unaffected if the squeezing
in q and p is suitably rescaled, so that the width of the q and p states becomes
In this modified protocol, Alice broadcasts the value of q modulo ..,fir. Q or the
value of p modulo ..,fir/ Q. Bob subtracts the value broadcast by Alice from his
own measurement outcome, and then adjusts the difference he obtains to the
nearest multiple of ..,fir . Q or ..,fir/ Q. The key bit is determined by whether the
multiple of..,fir . Q, or ..,fir/ Q, is even or odd.
Thus, for example, the error rate sustained due to imperfect squeezing will
have the same (acceptably small) value irrespective of whether Alice sends
states with !:::.q = !:::.p = 1/2, or !:::.q = 1 and !:::.p = 1/4; Alice can afford to
send coherent states about half the time if she increases the squeezing of her
other transmissions by a compensating amount.
Secure quantum key distribution using squeezed states 347
amplitude; this measurement may be less sensitive to detector defects than the
single-photon measurement required in BB84.
But, as in the BB84 protocol, losses due to the absorption of photons in the
channel will enhance the error rate in squeezed-state quantum key distribution,
and so will limit the distance over which secure key exchange is possible. We
study this effect by modeling the loss as a damping channel described by the
master equation
p=r(apat-~atap-~pata) j (22.70)
(22.71)
where
(22.73)
(22.74)
(22.75)
(22.76)
Now let's revisit the analysis of §6., taking into account the effects of losses.
We imagine that Alice prepares entangled pairs of oscillators in the state (22.54),
Secure quantum key distribution using squeezed states 349
and sends one oscillator to Bob through the lossy channel; then they perform
entanglement purification. This protocol reduces to one in which Alice prepares
a squeezed state that is transmitted to Bob. In the squeezed-state protocol, Alice
decides what squeezed state to send by sampling the probability distribution
P(qA) given in eq. (22.63); if she chooses the value qA, then she prepares and
sends the state 1'I/J(qA)} in eq. (22.59). When it enters the channel, this state is
governed by the probability distribution
(22.77)
and when Bob receives the state this distribution has, according to eq. (22.76),
evolved to
(22.78)
where
~ = e-K.d/2 , (22.81)
where d is the length of the channel and 1\:-1 is its attenuation length (typically
of the order of 10 km in an optical fiber).
The protocol is secure if the error rate in both bases is below 11%; as in
§6., this condition is satisfied for b.~ < .784. Thus we can calculate, as a
function of the initial squeezing parameter Li, the maximum distance d max that
the signal states can be transmitted without compromising the security of the
protocol.
For Li « 1, we find
Figure 22.3 The effect of channel losses on the security of quantum key distribution using
squeezed states. The maximum length r;,d max of the channel (in units of the attenuation length)
is plotted as a function of the width t. of the squeezed state that enters the channel. For a longer
channel, the error rate due to losses is too large and the proof of security breaks down. The
curve labeled "with amplification" applies to the protocol in which the signal is amplified prior
to detection in order to compensate for the losses; the curve labeled "without amplification"
applies to the protocol in which the signal is not amplified.
Thus, the more highly squeezed the input signal, the less we can tolerate the
losses in the channel. This feature, which sounds surprising on first hearing,
arises because the amount of squeezing is linked with the size of the range
in qA that Alice samples. Errors are not unlikely if losses cause the value
of qB to decay by an amount comparable to yI1r /2. In our protocol, if the
squeezed states have a small width Li, then the typical states prepared by Alice
are centered at a large value qA "'-' Li -1; therefore, a small fractional decay can
cause an error.
On the other hand, even without losses, Alice needs to send states with
.6. < .749 to attain a low enough error rate, and as Li approaches .749 from
below, again only a small loss is required to push the error probability over
11 %. Thus there is an intermediate value of .6. that optimizes the value of d max ,
as shown in Fig. 22.3. This optimal distance,
(q) -t e- (q) ,
1
(22.86)
Error rates in the q and p bases are below 11 %, and the protocol is provably
secure, for (b..~) amp < .784.
By solving (b..~)amp = .784. we can find the maximum distance d (where
e- 2 = eK.d) for which our proof of security holds; the result is plotted in
352 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Fig. 22.3. When the squeezed input is narrow, is. < < 1, the solution becomes
e = exp (11; dmax ) = 1.307 + O(~- 2 ) ,
-2 _
(22.87)
or
Comparing the two curves in Fig. 22.3, we see that the protocol with am-
plification remains secure out to longer distances than the protocol without
amplification, if the input is highly squeezed. In that case, the error rate in the
protocol without amplification is dominated by the decay of the signal, which
can be corrected by the amplifier. But if the input is less highly squeezed, then
the protocol without amplification remains secure to longer distances. In that
case, the nonzero width of the signal state contributes significantly to the error
rate; the amplifier noise broadens the state further.
With more sophisticated protocols that incorporate some form of quantum
error correction, continuous-variable quantum key distribution can be extended
to longer distances. For example, if Alice and Bob share some noisy pairs
of oscillators, they can purify the entanglement using protocols that require
two-way classical communication [15, 7]. After pairs with improved fidelity
are distilled, Alice, by measuring a quadrature amplitude in her laboratory,
prepares a squeezed state in Bob's; the key bits can be extracted using the
same error correction and privacy amplification schemes that we have already
described.
Our proof of security applies to the case where squeezed states are carried
by a lossy channel (assuming a low enough error rate), because this scenario
can be obtained as a reduction of a protocol in which Alice and Bob apply
entanglement distillation to noisy entangled pairs of oscillators that they share.
More generally, the proof applies to any imperfections that can be accurately
modeled as a quantum operation that acts on the shared pairs before Alice and
Bob measure them. As one example, suppose that when Alice prepares the
squeezed state, it is not really the q or p squeezed state that the protocol calls
for, but is instead slightly rotated in the quadrature plane. And suppose that
when Bob performs his homodyne measurement, he does not really measure
q or p, but actually measures a slightly rotated quadrature amplitude. In the
entanglement-distillation scenario, the imperfection of Alice's preparation can
be modeled as a superoperator that acts on her oscillator before she makes a
perfect quadrature measurement, and the misalignment of Bob's measurement
can likewise be modeled by a superoperator acting on his oscillator before
he makes a perfect quadrature measurement. Therefore, the squeezed state
protocol with this type of imperfect preparation and measurement is secure, as
long as the error rate is below 11 % in both bases. Of course, this error rate
Secure quantum key distribution using squeezed states 353
includes both errors caused by the channel and errors due to the imperfection
of the preparation and measurement.
We also recall that in the protocols of §5., Alice's preparation and Bob's
measurement were performed to m bits of accuracy. In the entanglement
distillation scenario, this finite resolution can likewise be well modeled by a
quantum operation that shifts the oscillators by an amount of order 2- m before
Alice and Bob perform their measurements. Thus the proof applies, with the
finite resolution included among the effects contributing to the permissible
11 % error rate. The finite accuracy causes trouble only when Alice's and Bob's
results lie a distance apart that is within about 2- m of ..j7r/2; thus, just a few
bits of accuracy should be enough to make this additional source of error quite
small.
8. CONCLUSIONS
We have described a secure protocol for quantum key distribution based on
the transmission of squeezed states of a harmonic oscillator. Conceptually,
our protocol resembles the BB84 protocol, in which single qubit states are
transmitted. The BB84 protocol is secure because monitoring the observable
Z causes a detectable disturbance in the observable X, and vice versa. The
squeezed state protocol is secure because monitoring the observable q causes a
detectable disturbance in the observable p, and vice versa. Security is ensured
even if the adversary uses the most general eavesdropping strategies allowed
by the principles of quantum mechanics.
In secure versions of the BB84 scheme, Alice's source should emit single-
photons that Bob detects. Since the preparation of single-photon states is
difficult, and photon detectors are inefficient, at least in some settings the
squeezed-state protocol may have practical advantages, perhaps including a
higher rate of key production. Squeezing is also technically challenging, but
the amount of squeezing required to ensure security is relatively modest.
The protocol we have described in detail uses each transmitted oscillator to
carry one raw key bit. An obvious generalization is a protocol based on the code
with stabilizer generators given in eq. (22.8), which encodes ad-dimensional
protected Hilbert space in each oscillator. Then a secure key can be generated
more efficiently, but more squeezing is required to achieve an acceptable error
rate.
Our protocols, including their classical error correction and privacy amplifi-
cation, are based on CSS codes: each of the stabilizer generators is either of the
"q"-type (the exponential of a linear combination of n q's) or of the "p-type"
(the exponential of a linear combination of n p's). The particular CSS codes
that we have described in detail belong to a restricted class: they are concate-
nated codes such that each oscillator encodes a single qubit, and then a block of
354 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Acknowledgments
We thank Andrew Doherty, Steven van Enk, Jim Harrington, Jeff Kimble,
and especially Hoi-Kwong Lo for useful discussions and comments. This
work has been supported in part by the Department of Energy under Grant No.
DE-FG03-92-ER40701, and by DARPA through the Quantum Information and
Computation (QUIC) project administered by the Army Research Office under
Grant No. DAAH04-96-1-0386. Some of this work was done at the Aspen
Center for Physics. This paper first appeared in Physical Review A
Secure quantum key distribution using squeezed states 355
Notes
I. We implicitly assume that Eve uses a strategy that passes the verification test with nonnegligible
probability, so that the rate of key generation is not exponentially small. If, for example, Eve were to
intercept all qubits sent by Alice and resend them to Bob, then she would almost certainly be detected, and
key bits would not be likely to be generated. But in the rare event that she is not detected and some key bits
are generated, Eve would know a lot about them.
2. This bound is not tight. It applies if the sample of M pairs is chosen from the population of N with
replacement. In fact the sample is chosen without replacement, which suppresses the fluctuations. A better
bound was quoted by Shor and Preskill [18].
References
[1] Bennett, C. H., and G. Brassard (1984), "Quantum cryptography: Public-
key distribution and coin tossing," in Proceedings of IEEE International
Conference on Computers, Systems and Signal Processing (Bangalore,
India), pp. 175-179; C. H. Bennett and G. Brassard (1985), "Quantum
public key distribution," IBM Technical Disclosure Bulletin 28, 3153-
3163.
[2] Bennett, C. H., D. P. DiVincenzo, J. A. Smolin and W. K. Wootters (1996),
"Mixed state entanglement and quantum error correction," Phys. Rev. A
54,3824-3851, quant-phl9604024.
[3] Biham, E., M. Boyer, P. O. Boykin, T. Mor and V. Roychowdhury (2000),
"A proof of the security of quantum key distribution," in Proceedings of
the Thirty-Second Annual ACM Symposium on Theory of Computing, pp
715-724. New York: ACM Press, quant-phl9912053.
[4] Brassard, G., N. Liitkenhaus, T. Mor, and B. C. Sanders (2000), "Lim-
itations on practical quantum cryptography," Phys. Rev. Lett. 85, 1330,
quant-phl9911054.
[5] Braunstein, S. (1998), "Error correction for continuous quantum vari-
ables," Phys. Rev. Lett. 80, 4084, quant-phl9711049.
[6] Calderbank, A. R., and P. W. Shor (1996), "Good quantum error correcting
codes exist," Phys. Rev. A 54, 1098-1105, quant-phl9512032.
[7] Duan, L. M., G. Giedke, J. I. Cirac, and P. Zoller (1999), "Entangle-
ment purification of Gaussian continuous variable quantum states," quant-
ph/9912017; L. M. Duan, G. Giedke, J. I. Cirac, and P. Zoller (2000),
"Physical implementation for entanglement purification of Gaussian con-
tinuous variable quantum systems," quant-phl0003116.
[8] Gottesman, D., A. Kitaev, and J. Preskill (2000), "Encoding a qudit in an
oscillator," quant-phl0008040.
[9] Hillery, M. (1999), "Quantum cryptography with squeezed states,"
quant-phl9909006.
356 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
EXPERIMENTAL DEMONSTRATION OF
QUANTUM DENSE CODING AND
QUANTUM CRYPTOGRAPHY WITH
CONTINUOUS VARIABLES
Abstract In this paper we will present the experimental demonstrations of quantum dense
coding and quantum cryptography using continuous electromagnetic field with
Einstein-Podolsky-Rosen(EPR) correlations. The bright EPR optical beams
with the quantum correlations between the amplitude and phase quadratures
are produced from a nondegenerate optical parametric amplifier. The direct
detection technology of the Bell-state is ultilized in the measurements of the
quantum correlations and the signals modulated on the quadratures instead of
usual homodyne detection. Usability of experimentally accessible squeezed-state
entanglements, high efficiencies of bit transmission and information detection,
relatively straightforward systems and operating procedures, and security directly
provided by quantum correlations make the presented schemes valuable to be
applied to the developing quantum information science.
1. INTRODUCTION
Recently, more and more investigation interests in quantum information
science have focus on exploiting the quantum system which possesses contin-
uous spectra [1-19]. The successful experiments on the quantum teleportation
[20] and quantum dense coding [21] using EPR correlations of continuous
electromagnetic fields provided possible technologies for quantum information
processing based on continuous variables. High bit transmission rates and
357
S.L Braunstein andA.K. Pati (eds.), Quantum Information with Continuous Variables, 357-378.
© 2003 Kluwer Academic Publishers.
358 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
high detection efficiency are the key advantages of continuous variable systems
better than discrete systems.
In this paper we will present the experimental schemes and results on con-
tinuous variable quantum dense coding and quantum cryptography. The ex-
perimental technologies for the generation of EPR correlated beams and the
detection of the Bell-state will be described firstly.
and (23.1)
Figure 23.1 Experimental setup for the generation of EPR beams: YAP-Nd:YAI0 3 , F.R-
Faraday Rotator, EOM-Electric Optical modulator, PBS-polarizing beam splitter.
crystal (10mm long), the front face of which is coated to be used as the input
coupler( the transmission>95% at 540nm wavelength and 0.5% at lOS0nm)
and the other face is coated with the dual-band antireflection at both 540nm
and lOS0nm, as well as a concave mirror of 50mm-curvature radius, which is
used as the output coupler of EPR beam at lOSOnm (the transmission of 5%
at lOS0 and high reflectivity at 540nm). The output coupler is mounted on a
piezoelectric transducer to lock actively the cavity length on resonance with the
injected seed wave at lOS0nm using the FM sideband technique. By fine tuning
the crystal temperature the birefringence between signal and idler waves in KTP
is compensated and the simultaneous resonance in the cavity is reached. The
process of adjusting temperature to meet double resonance can be monitored
with an oscilloscope during scanning the length of cavity. Once the double
resonance is completed the NOPA is locked on the frequency of the injected seed
wave [25]. The measured finesse, the free spectral range, and the line-width of
the parametric oscillator are 110, 2.SG, and 26 MHz, respectively. The pump
source of NOPA is a home-made all-solid-state intracavity frequency-doubled
and frequency-stabilized CW ring Nd:YAP (Nd:YAI03, Yttrium-Aluminum-
Perovskite) laser [27]. The output second-harmonic wave at 540nm and the
leaking fundamental wave at 1080nm from the laser serve as the pump light and
the seed wave respectively. The laser-diode pumped all-solid-state laser and the
semi-monolithic F-P configuration of the parametric cavity ensure the stability
of system, so the frequency and phase of light waves can be well-locked during
the experiments.
The NOPA is pumped by the harmonic wave at 540nm, that is controlled
just below the oscillation threshold of the NOPA, and the polarization of that
is along the b axis of the KTP crystal. Due to the large transmission (>95%)
of input coupler at 540nm, the pump field only passes the cavity twice without
resonating. After the seed beam at 1080nm polarized at 45° relative to the b
axis of the KTP crystal is injected into the cavity, it is decomposed to signal
and idler seed waves with identical intensity and the orthogonal polarization
along the b and c axes, respectively, which correspond to the vertical and
horizontal polarization. The temperature of KTP crystal placed in a special
designed oven is actively controlled around the temperature for achieving type
II noncritical phase matching (63°C) with a broad full width of about 30°C
[25]. An electronics feedback circuit is employed to stabilize actively the
temperature of crystal to a few mK.
Locking the relative phase between the pump laser and the injected seed wave
of NOPA to 2mr or (2n+ 1)-11", where n is integer, to enforce the NOPA operating
at amplification or deamplification, the entangled EPR beam with the quantum
correlation of amplitude(or phase) quadratures and the quantum anticorrelation
of phase(or amplitude) quadratures was generated [21,25]. The two halves of
the EPR beams are just the signal and idler modes of the subharmonic wave field
360 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
produced from type II parametric down conversion, thus they have orthogonal
polarizations originally. The output signal and idler beams are separated by
a polarizing-bearn-splitter (PBS2) to be a pair of EPR beams, (X1,Y1) and
(X2, Y 2), which possess the feature of nonlocal quantum entanglement of the
amplitude and phase quadratures.
RF
Splitter
DJ
Ya-Yb
d
RF
Splitter
b D2
Figure 23.2 Direct measurement of the Bell state. BS: 50% beam splitter. DI and D2:
detectors.
Fig. 23.2 is the diagram of the direct detection system. Two bright coherent
beams with identical intensities are expressed by the annihilation operators a
and b. A phase shift of 7r/2 is imposed on beam a, and then the beams are
mixed on a 50% beam splitter. The resulting output beams c and d are given by:
d= V;(a-ib). (23.2)
Experimental demonstration of dense coding and quantum cryptography 361
We define upper-case operators in the rotating frame about the center frequency
wo,
The fields are now described as functions of the modulation frequency O. Thus
we can consider any field as a carrier 0(0) oscillating at frequency Wo with an
average value equal to the steady state field, surrounded by "noise side-bands"
0(0) oscillating at frequency Wo ± 0 with zero average values. The amplitude
and phase quadrature can be written as
The two bright output beams can be directly detected by Dl and D2. The
discussed photocurrents are normalized with the average value of the field.
The normalized output photocurrents spectra are given by
1
= "2 (Xa(O) + Ya(O) - Yb(O) + Xb(O)) (23.6)
1
"2(Xa(O) - Ya(O) + Yb(O) + Xb(O))
The Eq. (23.6) shows that the photocurrent of the each arm of the 50% beam-
splitter consists of two parts, part one (term 1 and term 4) is self terms of
two input fields Xa(O) and Xb(O) at the beamsplitter, the part 2 comprises
the interference terms (2 and 3) including phase quadratures Ya(O) and Yb(O)
deriving from the 7f /2 phase shift. Comparing ic and ict
in Eq. (23.6), it is
obvious that the self terms of two arms are correlated (or in phase), and the
interference terms are anticorrelated (or out of phase). Each of photocurrents
is divided into two parts through the RF power splitter. The sum and difference
of the divided photocurrents are
1 ~ ~ 1
= v'2(i c (O) + id(O)) = v'2(Xa(O) + Xb(O)) (23.7)
We can see that the sum i+ of photocurrents of two arms c, d only leaves
the self tenns which include the amplitude quadrature of the signal field a
and b; the difference photocurrent i_ leaves the interference tenns, which
gives the infonnation of their phase quadratures. Thus a combined Bell-
state measurement of beams and is achieved with this simple self-homodyne
detector. In our dense coding and cryptography experiments, the detected
a
fields and b are the signal and idler modes, (X 1,Y1) and (X2,Y2), of the
output field from the NOPA operating at deamplification, so there are quantum
anticorrelation and correlation between their amplitude and phase quadratures
(see the inequalities (23.1) ).
The correlations measured by self-homodyne detector between the quad-
rature-phase amplitudes of the two halves of EPR beam are show in Fig. 23.3.
Both variances of (8(X1 + X2)2) (Fig. 23.3(a» and (8(Y1 - Y2)2) (Fig.23.3
(b» measured directly are -4dB below that of the SNL (considering the electron-
ics noise that is 8dB below the SNL, the actual fluctuation should be -5.4dB be-
low that of the SNL). The product of the correspondent conditional variances of
the EPR beam is (8(X1 + X2)2) (8(Y1 - y 2)2)=0.63. The bright EPR beam
of -70j.tmW(the correspondent photon-number flow is about 3.8 x 1014 8- 1 )
was obtained at the following operation parameters ofNOPA: the pump power
150mW just below the power of the oscillation threshold of 175mW and the
polarization of that was along the b axis of the KTP crystal. The power of the
injected seed wave was 10mW before entering the input coupler of the cavity
and that was polarized at 45° relative to the b axis.
-90
-92
t~
-94 <O(X t + X2)2>
j "l'f~N\-V\J"\Ji'..v\ ",vJ\\fl;V'.;'!,;VV'~V/,1t~\:".\{V\"v;y~/;f'i".;"\
-96
-98 +--~--'r--~---r-~--r--~--r-~-..,
time (a)
-86
~oo~
I~2
51 -94
'(5
Z
time (b)
Figure 23.3 Spectral densities of photocurrent fluctuations (c5(Xl + X2)2) (a) and (c5(Yl +
Y2?) (b), SNL-the Shot Noise Limit. Acquisition parameters: radio frequency (rf)
!1127r=2MHz, resolution bandwidth t..!1127r=30KHz, Video bandwidth O.1KHz, the electronics
noise is 8dB below the SNL,
(ETX500 InGaAs), and then each photocurrent of Dl and D2 was divide into
two parts through the power splitters_ The sum and difference of the divided
photocurrent were nothing else but the transmitted amplitude and phase signals
from Alice to Bob, which are expressed by [28]
364 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(23.S)
where X s (n) and Y s (n) are the modulated amplitude and phase signals on the
first half of EPR beam at the sender Alice. With perfect EPR entangled beam
(X1 (n) + X2(n)) -+ 0 and (Y1 (n) - Y2 (n)) -+ 0, Eqs. (23.7) and (23.S) are
simplified:
(23.9)
(23.10)
It means that under ideal conditions, the signals Xs(n) and Ys(n) encoded on
the amplitude quadrature and phase quadrature of beam 1 are simultaneously
retrieved at the receiver Bob without any error. In general, the both encoded
signals will be recovered with a sensitivity beyond that of the SNL when the
beam 1 and beam 2 are quantum entangled. In fact, the sum of the amplitude
quadratures between the two halves ofEPR beam commutes with the difference
of its phase quadrature, therefore the detected variances in them can be below
that of the SNL simultaneously, and not violate the uncertainty principle [24].
Fig. 23.5 shows the directly measured amplitude [Fig. 23.5(a)] and phase
[Fig. 23.5(b)] signals at Bob, which are the signals of 2MHz modulated on
the first half of EPR beam (beam 1) by the amplitude and phase modulators
at Alice. It can be seen that the original signals are retrieved with the high
signal to noise ratio of -4dB and -3.6dB beyond that of the SNL under the
help of the other half of EPR beam (beam 2)(accounting for the electronics
noise of -SdB below the SNL, the actual value should be -5.4dB and -4.SdB
respectively) . Compared with the previously completed sub shot noise limit
optical measurements and quantum non-demolition measurements for a signal,
in which the squeezed state light have been applied, our experiments have
achieved the simultaneous measurements of two signals modulated on the
amplitude and phase quardratures respectively with the precision beyond that
of the SNL by means of exploiting the EPR entanglement. Although when
sending and modulating two quadratures of a coherent beam a factor of two in
channel capacity may also be gained, the signal-to-noise ratios of measurements
are not able to breakthrough the SNL.
As well-discussed in Ref. [15], in which a signal is transmitted via two
quantum channels of EPR pair, in our scheme the individual signal channel
Experimental demonstration of dense coding and quantum cryptography 365
:
540nrn
Laser 1080nm
AM PM
Figure 23.4 Schematic ofthe experimental apparatus for dense coding for continuous variables.
Two bits of classical information X sand Y s are encoded on the amplitude and phase quadratures
of a half of EPR beam (beam 1) at Alice, then are decoded by the other half of EPR beam (beam
2) at Bob.
5. QUANTUM CRYPTOGRAPHY
In recent years the cryptographic schemes employing continuous coherent
and nonclassical light fields have been suggested [6-19]. A lot of interest has
arisen in continuous variable quantum cryptography (CVQC) with EPR beams
due to that the experimental demonstrations of quantum teleportation [20]
and quantum dense coding [21] for continuous variables (CV). J.Kimble and
his colleagues completed the quantum communication of dual channels with
correlated nonclassical states of light [15]. However in this case, an eavesdrop-
per (Eve) could simultaneously access the signal and idler beams without the
366 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
I~+---~-.--~---.------,---~-.
10 1.5 2C 15
Frequoncy(MHz}
(a)
·M
I ~; +---~--.-------.---~--.---~~
I .U 1~ 2.U :.1.1,)I.'
(b) Frequency(MHz)
Figure 23.5 Measured amplitude (a) and phase signal (b) at Bob, when EPR beam 1 is phase
and amplitude modulated at 2M Hz at Alice. SNL-the Shot Noise Limit (black line). Acquisition
parameter: measured frequency range LOMHz-3.0MHz, resolution bandwidth 30KHz, video
bandwidth 0.1KHz, the electronics noise is 8dB below the SNL.
-80
Frequency(MHz)
Figure 23.6 Spectral density of photo current fluctuations of beam 1 with the modulation signals
(trace(a», the modulated signals are submerged in the noise background. SNL-the Shot Noise
Limit (trace(b» Acquisition parameter: measured frequency range l.OMHz-3.0MHz, resolution
bandwidth 30KHz, vides bandwidth O.IKHz; the electronics noise is 5.6dB below the SNL.
bits [13]. All transmitted bits are used for constituting the key string without
bit rejection, thus the transmission efficiency of 100% may be achieved in
principle.
The quadrature amplitudes of the two output field modes from a NOPA
operating at deamplification are [22,23,26]:
(23.12)
368 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(23.13)
Here we have assumed that the two modes are totally balanced during the
process of measurements and this requirement is easily achieved in the exper-
iments. Generally, r is a function of the noise frequencies. The modulated
signals at given radio frequencies (rf) can be considered as a noise 8Xs(O)
and 8Ys(f!). When the powers of 8Xs(O) and 8Ys(f!) are smaller than the
original quantum noise power of the signal beam (Eqs. (23.12» and larger
than the correlation noise power (Eqs. (23.13», the signals are submerged in
the noise background of the signal beam ( Xl, YI ) and may be decoded by
the correspondent idler beam ( X 2, Y2 ) of the EPR beams. Thus, the strength
(8 2 Xs) of the modulated signals should satisfy the following inequalities:
Fig. 23.7 shows the functions of the variances (8 2 (XI)) = (8 2 (YI)) (curve I)
and (8 2(Xl + X 2)) = (8 2(YI - Y2 )) (curve II) versus the correlation param-
eters . At the cross point A of the curves 1 and II, r = rl = 0.275, the variance
is 2.4dB below the normalized SNL of EPR beams (line ii) and when r > rl the
variances of signal beam (curve I) are higher than the correlation fluctuations
(curve II). The line of variance=1 (line i) stands for the SNL of the noise power
of signal beam which is 3dB below that of both EPR correlated beams (line ii).
The larger the correlation parameter is, the higher the noise power of the signal
beam is and the lower the correlation variance is. For a perfect transmission
line of noiseless the correlation parameter of the NOPA should be larger than
rl = 0.275 at least for accomplishing the quantum cryptography. The quantum
correlation of -2.4dB is a kind of boundary where the security becomes more
favorable due to that the interception of Eve becomes harder. The height of the
modulated signals should be between curve (I) and curve (II) for a given value
r > rl.
At Bob station, the signals 8Xs(f!) and 8Ys(f!) modulated on the amplitude
and phase quadratures of signal beam ( Xl, YI ) are decoded by the retained
idler beam by means of the direct detection of photocurrents and two rf beam
splitters. The normalized photocurrents measured with the positive (+) and
negative (-) power combiners are :
Experimental demonstration of dense coding and quantum cryptography 369
Variance
r
0.2 rl r2 0.4 0.6 0.8
Figure 23.7 The variances of W(Xl)) = (c\'2(YI)) (curve I) and W(X 1 +X2)) = W(Yl -
Y 2 )) versus the correlation parameter r. The line (i) and (ii) are the SNL of signal beam and
EPR beams, respectively. At the point A (r = r' ~ 0.275) the variance is ....... 2.4dBm below the
SNL of EPR beams (ii).
1
yI2{[X1 (O) + X2(O)] + Xs(O) * BV} (23.15)
1 --
yI2{[Yl(O) - Y2(O] + Ys(O) * BV}
Where BV and BV stand for the bit values 1 and O. BV is the NOT of BV,
that means when BV equals "I" the BV must be "O",vice versa.
Fig. 23.8 is the schematic of the quantum cryptography system. The EPR
source, NOPA, is set inside the Bob receiving station. The one of the bright
EPR correlated beams, the signal beam ( XI, Yl), is sent to the Alice sending
station where Alice encodes the transmitted information on the amplitude and
phase quadratures by the choice of the modulation types with the binary bit
values, for example the modulated amplitude signals stands for "I" and the
phase signals for "0". Then the encoded signal beam is transmitted back to
Bob where the information is decoded by the retained other one of the EPR
correlated beams.
370 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Figure 23.8 The schematic of the quantum cryptography using the EPR beams, AM-amplitude
modulator, PM-phase modulator, RNG-random number generator, SA-spectrum analyzer, PS-
phase shifter, BS-beamsplitter of 50%, PBS-polarization beamsplitter, NOPA-nondegenerate
optical parametric amplifier.
T T. Ts
Alice PM PM AM AM PM
AQ
Bob
PQ
Key o o o
Table 23.1 The generation of the predetermined secret key string, AQ-Amplitude modulation
signals, PQ-Phase modulation signals, CD-The correlation degrees between signal and idler
beams measured by Bob. Key-The secret keys, upper traces-The SNL of EPR beams, middle
traces-the noise spectra of signal beam, lower traces-the noise spectra of (8 2(Xl + X 2)) and
(8 2(Yl - Y2)).
the additional loss caused by a partial tapping of the signal channel, perhaps
by Eve. If Eve wants to perform the quantum-nondemonlition-measurment
(QND) on one of the amplitude-phase quadratures of the signal beam, the
other quadrature must be disturbed and the disturbance must be reflected on
the measured results of Bob. The deviation would indicate the possibility that
Eve have performed a QND. More quantitative analyses on protecting against
the optical tap attack of Eve have been presented in the previous publications
[11-15] and can be used in the discussion to the presented protocol.
For the ideal case of the EPR beams with perfect quantum correlation, the
SNR on the signal channel goes to zero due to the fact that the quantum noise
of the signal beam goes to infinite [15,28]. Thus, the attacking from optical tap
or the optical QND on the signal beam has no possibility to extract information
hidden in the infinite quantum noise without the help of the other one of the EPR
correlated beams. However, for the imperfect correlations the signal channel
has large but finite quantum noise background. In this case, the eavesdroppers
can enhance the measurement SNR by narrowing the resolution bandwidth
(RBW) of the spectral analyzer (SA) according to [33]:
<i~)
SNR = (iJv) = 'fJls/BFo (23.16)
where <i~) and <iJv) are the mean square signal and noise photocurrents
respectively, Is is the photon number flux of the signal beam, "I is the quantum
efficiency of the detection system and Fo is the Fano factor of the signal beam,
Fo -+ 00 for perfect correlated EPR beams and Fo = 1 without quantum
correlation(coherent state light). For the given "I, Is and F o, the usable RBW
of the measurement is limited by the requirement of SNR;::: 1,i.e.
B0 "lIs
<
- Fo
- (23.17)
It means that the duration of measurement should be larger than BOI. For
secure communication Alice should encode the message with a bit rate larger
than Bo decided by Eq. (23.17), so that Eve can not intercept information
due to having no enough interval to accomplish the measurement. In our
experiment when B = 30kHz, the signals have been totally submerged in the
noise background (the middle traces in Table 23.1), thus the condition is easily
satisfied in the practical optical communication.
A possible eavesdropping scheme using fake EPR beams is shown in Fig. 23.9.
Eve intercepts totally the signal beam (Xl, Yl ) and transmits a fake signal beam
(Xf, Y{) produced by a fake EPR source (EPR2) in her station to Alice. Alice
has no ability to recognize the fake beam. She modulates the information on it
and then sends out as usual. Eve intercepts the beam with messages again and
Experimental demonstration of dense coding and quantum cryptography 373
Bob
Detector 1 t====1
Eve
(X/,Yj)
(Xt',Y/)
Alice
Figure 23.9 Diagram of eavesdropping Scheme using fake EPR beams, EPR1, EPR2-Sources
of EPR beams, AM, PM-Amplitude and phase modulators, SA-spectral analyzer, OSl, OS2-
Oscilloscopes, Detector 1, 2-Detection systems of Bell-state, Do-photoelectric detector, RNG-
random number generator.
decodes the infonnation with the other one of the fake EPR correlated beams
(X~, Yf) retained by her. At the same time she modulates the real signal beam
(XI'yI) according to the intercepted bit values and sends it back to Bob who
also does not know that the infonnation has been intercepted. To reveal this
type of quantum interception Alice may randomly block the signal beam with
374 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
I
965-. 56-18~
-965.5r 184 ••
Figure 23.10 The fluctuation shapes of signal (I) and idler (2) beams recorded by the oscillo-
scopes OSl and OS2 at Alice and Bob.
6. CONCLUSION
We have experimentally demonstrated the quantum dense coding and quan-
tum cryptography with continuous variables. Due to using the bright EPR
correlated beams, the transmitted signals are directly modulated on the am-
plitude and phase quadratures of signal beam with the amplitude and phase
modulators compatible with that used in the present optical communication
system operating at very high rates. The application of the direct detection
technology of the Bell-state simplifies the measurement system and the align-
ing procedures, and further improves the detection efficiency. Available high
bits transmission rates and high detection efficiency are favorable features of the
presented schemes. Besides, we have proved that the experimentally achiev-
able quantum correlation and squeezing level may be used for performing the
quantum communication. Unlike most of proposed protocols for quantum
cryptography based on BB84 [34], our scheme does not require any randomly
chooses of measuring components, so there is no the usual 50% bit rejec-
tion. The predetermined secret key string is directly modulated on the optical
beam(the carrier of signals) just like that used in traditional optical communi-
cation systems. The partial compatibility with classical optical communication
376 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
References
[1] Braunstein, S. L. (1998) Error correction for continuous quantum vari-
ables. Phys. Rev. lett. 80,4084-4087.
[2] Braunstein, S. L. (1998) Quantum error correction for communication
with linear optics. Nature (London) 394. 47-49
[3] Braunstrin, S. L. & Kimble, H. J. (2000) Dense coding for continuous
variables. Phys. Rev. A 61, 042302
[4] Ban, M. (1999) Quantum dense coding via a two-mode squeezed-vacuum
state. J. opt. B: Quantum Semic1ass. Opt. 1 L9-Lll
[5] Loock, P. Van &Braunstein, S. L. (2000) Uconditional entanglement
swapping for continuous variables. Phys. Rev. A 61, 10302(R)
[6] Gottesman, D., & Preskill, J. (2001) Secure quantum key distribution
using squeezed states. Phys. Rev. A 63, 022309
[7] Hillery, M., (2000) Quantum cryptography with squeezed states. Phys.
Rev. A 61, 022309
[8] Cerf, N. J., Levy, M., & Assche, G. v., (2000) Quantum distribution of
Gaussian keys using squeezed states. Phys. Rev. A 63,052311
[9] Cerf, N. J., Iblisdir, S., & Assche, G. V., (2001) Cloning AND
Cryptography with Quantum Continuous Variables. quant-phlO107077
[Eur.Phys.J.D(to be published)]
[10] Assche G. V. et al., (2001) Reconciliation of a Quantum-Distributed Gaus-
sian Key. cS.CRlO107030 (to be published)
[11] Ralph, T. C. (2000) Continuous variable quantum cryptography. Phys.
Rev. A 61, 010303(R)
[12] Ralph, T. c., (2000) Security of continuous-variable quantum cryptogra-
phy. Phys. Rev. A 62, 062306
[13] Reid, M. D., (2000) Quantum cryptography with a predetermined key,
using continuous-variable Einstein-Podolsky-Rosen correlations. Phys.
Rev. A 62, 062308
[14] Silberhom, c., Korolkova, N., & Leuchs, G., (2001) Quantum key distri-
bution with bright entangled beams. quant-phlO109009
Experimental demonstration of dense coding and quantum cryptography 377
[15] Pereira, S. E, Ou, Z. Y., & Kimble, H. J., (2000) Quantum communication
with correlated nonclassical states. Phys. Rev. A 62, 042311; Kimble, H.
J., Ou, Z. Y., & Pereira, S. E, Method and Apparatus for Quantum Com-
munication Employing Nonclassical Correlations of Quadrature-Phase
Amplitudes. u.s. Patent No. 5,339,182, Issued 8/16/94.
[16] Bencheikh, K. et al., (2001) Quantum key distribution with continuous
variables. J. Mod. Opt. 48, 1903
[17] Lorenz, S. et al., (2001) Squeezed light from microstructured fibres: to-
wards free space quantum cryptography. quant-phlOl09018
[18] Navez, P. et aI., (2001) A "quantum public key" based cryptographic
scheme for continuous variables. quant-phlOlO1113
[19] Grosshans, E, & Grangier, P., (2002) Continuous Variable Quantum Cryp-
tography Using Coherent States. Phys. Rev. Lett. 88,057902
[20] Furusawa, A. et al. (1998) Unconditional quantum teleportation. Science
282, 706-709
[21] Li, X. Y. et ai. Quantum Dense Coding Exploiting a Bright Einstein-
Podolsky-Rosen Beam. Phys. Rev. Lett. 88,047904 (2002)
[22] Reid, M. D., & Drammond, P. D. (1988) Quantum Correlations of Phase
in Nondegenerate Parametric Oscillation. Phys. Rev. Lett. 60, 2731-2733
[23] Reid, M. D. (1989) Demonstration of the Einstein-Podolsky-Rosen para-
dox using nondegenerate parametric amplification. Phys. Rev. A 40,913
[24] Ou, Z. Y.,Pereira, S. E, & Kimble, H. J. (1992) Realization of the Einstein-
Podolsky-Rosen paradox for continuous variables in nondegenerate opti-
cal parametric amplifier. Appi. Phys. B 55, 265
[25] Zhang, Y. et al. (2000) Experimental generation of bright two-mode
quadrature squeezed light from a narrow-band nondegenerate optical para-
metric amplifier. Phys. Rev. A 62,023813
[26] Zhang, Y., Su, H., Xie, C. D. & Peng, K. C. (1999) Quantum variances
and squeezing of output field from NOPA. Phys. Lett. A 259, 171
[27] Li,X. Y., Pan, Q., Jing, J. T., Xie, C. D., & Peng, K. C.,(2001) LD pumped
intracavity frequency-doubled and frequency-stabilized Nd:YAP/KTP
laser with 1.1 w output at 540nm, Optics Communications, (01) 01685-6
[28] Zhang, J. & Peng, K. C. (2000) Quantum teleportation and dense coding
by means of bright amplitude-squeezed light and direct measurement of
a Bell state. Phys. Rev. A 62, 064302
[29] Jing, J.,Pan, Q., Xie, C. D. & Peng, K. C. (2002) Quantum Cryptography
Using Einstein-Podolsky-Rosen Correlations of Continuous Variables.
quant-phl0204111
378 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
[30] Pan, Q., Zhang, Y., Zhang, T. c., Xie, C. D. & Peng, K. c., Experimental
inverstigation of intensity fifference squeezing using Nd: YAP laser as
pump source, J. Phys. D: Appl. Phys.30(l997) 1588-1590
[31] Julsgaard, B., Kozhekin, A., & Polzik, E. S. (2001) Experimental long-
lived entanglement of two macroscopic objects. Nature 413, 400-403
[32] Parkins, A. S. & Kimble, R. J. (2000) Position-momentum Einstein-
Podolsky-Rosen state of distantly separated trapped atoms. Phys. Rev. A
61,052104
[33] Li, Y. Q., Lynam, P., Xiao, M., & Edwards, P. J. Sub-Shot-Noise laser
Doppler Anemometry with Amplitude-Squeezed Light. Phys. Rev. Lett.
78, 3105 (1997)
[34] Bennett, C. R., & Brassard, G. Quantum Cryptography: Public Key Dis-
tribution and Coin Tossing. Proc.IEEE Int. Conf. On Computers, Systems
and Signal Processing(Bangalore), 175-179 (IEEE, New York,1984)
Chapter 24
1. INTRODUCTION
Non-local properties of continuous variables have been the focus of the
historic discussion which today is considered the starting point of quantum in-
formation [1]. At the time, however, the disputants did not have any application
in mind. They were rather struggling for the proper interpretation of quantum
theory. Einstein, Podolsky, and Rosen called the scenario they described [2]
a Gedanken-experiment for a good reason. Nobody knew how to create con-
tinuous variable quantum systems in an entangled state which could be used
in the laboratory to test for the non-locality of quantum theory. Then in 1951,
Bohm reformulated the EPR-Gedanken-experiment in terms of discrete spin
variables [3] laying the foundation for the laboratory studies to come. The in-
centive for more refined experiments rose when Bell derived a general criterion
based on which the non-locality of quantum theory could be put to a stringent
test [4]. The experiments by Clauser et al. [5] and Aspect et al. [6] which
followed all aimed at a better understanding of the foundations of quantum
379
S.L Braunstein and A.K. Pati (eds.), Quantum Information with Continuous Variables, 379-421.
© 2003 Kluwer Academic Publishers.
380 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
theory. One reason for their success was that dichotomic quantum systems can
be prepared experimentally in a state of close-to-perfect entanglement. And
even without any entanglement a discrete quantum system has the additional
advantage that the detector sends 'on-off' messages conveniently providing a
built-in discriminator function. It is therefore not surprising (in retrospect) that
when proposing quantum teleportation and quantum key distribution Bennett et
al. [7] used the example of the discrete quantum variables which had proven to
be so successful. The numerous experiments which followed were pioneered
by Zeilinger et al. [9]. Soon cryptography will be the first real world appli-
cation of quantum information [8]. However, the production process for the
entangled photon pairs used in these experiments is spontaneous parametric
down conversion and is therefore probabilistic. You never know ahead of time
at which point the experiment will be successful. For cryptographic key distri-
bution the probabilistic production of single photons may be less of a concern
but for other cases such as teleportation the concern seems more serious. Along
with the probabilistic production, the signal rate is fairly low.
Ou et al. [10] were the first to demonstrate that entanglement can be pro-
duced experimentally also with the continuous variable field quadratures of
two different modes of light. Ou et al. used an optical parametric amplifier
to produce entangled light beams at a very low light level. Such continuous
variable entanglement can be seen to be quite complementary to entangled pairs
of photons. Continuous variable entanglement is less perfect and more fragile
with respect to attenuation but it is not probabilistic. It is there at your disposal.
This and the much higher transmission rates may well be favourable in some
applications. Recently bright light entanglement was achieved using the in-
terference scheme [11] and using a nondegenerate optical parametric amplifier
[12], an additional advantage being the availability of efficient direct detection.
Furthennore, nonlinear coupling between bright light fields has been demon-
strated to lead to quantum correlation and entanglement [13, 14, 15] which has
not yet been achieved for single interacting photons. It should be noted here
that future developments of cavity quantum electrodynamics may lead to deter-
ministic sources of single photons [16, 17] and maybe also of entangled photon
pairs. Commercial tools for fibre-optic test and measurement are beginning to
exploit the ultrasensitivity of photon counting techniques at telecommunication
wavelength [18, 19], paving the way towards photon-counting communication
technology.
In the following sections of this chapter we discuss the steps towards ex-
perimental quantum information processing with an example of experiments
using fibre optical soliton pulses. The nonlinear interaction of solitons in a
fibre combined with linear optical elements such as beam splitters is enough
to assemble non-trivial basic building blocks for continuous variable quantum
communication.
Quantum solitons: towards experimental quantum communication 381
[27], where the QND interaction of coherent input fields was discussed as well
as QND coupling of squeezed beams aiming at improved teleportation quality.
(24.1)
with
(24.2)
(24.3)
(24.4)
(24.5)
The upper signs are used if the amplitudes are anti-correlated and the phases
are correlated and the lower signs are for the reversed situation.
Note, that for the optimal gains gx = l;r, gy = g?t the inference errors
(24.5) in the derivation of the EPR-paradox are equivalent to the conditional
variances (24.1,24.3):
(24.6)
Consider now the quantum noise limit for the measurement of conjugate
quadratures of a single field which is given by the Heisenberg relation:
V(X)V(Y) ~ 1. (24.7)
(24.8)
(24.9)
This specifies the ability to infer" at a distance" either of the two non-commuting
signal observables with a precision below the shot noise level of the signal
beam [24], (24.4). Condition (24.9) defines the EPR boundary for a non-
local quantum correlation and provides a sufficient condition for a state to be
entangled. Hence, if the quantum correlation is stronger than a certain threshold
one refers to EPR-entanglement and this is closely related to the idea of a QND
measurement.
384 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
(24.10)
where Pil, Pi2 denote density matrices of two subsystems 1 and 2 (see also
[30]). In contrast, an entangled state is a non-separable quantum state of a
system which cannot be represented as a convex sum of product states of two
subsystems. For discrete variables, i.e. for the Hilbert space of 2 x 2 and
2 x 3 dimensions, a necessary and sufficient criterion for separability is the
requirement that the partial transpose pT of the density matrix is positive which
is known as the Peres-Horodecki criterion.
Recently, the Peres-Horodecki criterion was extended to higher dimensions
of the Hilbert space, i.e. for continuous variables, using two different ap-
proaches (for details see [31, 32] and chapters 13 and 14 of the present book).
The analysis of Simon [31] is accomplished using the formalism of the phase
space Wigner function W(x,p) [21, 33]. It is based on the recognition that
the partial transpose pT is equivalent to a mirror reflection in phase space.
If p is separable, its Wigner distribution necessarily goes over into a Wigner
distribution under the phase space mirror reflection [31]:
Here x, p are the canonical variables position and momentum. This recognition
(24.11) leads to uncertainty principles to be obeyed by all separable states. For
all bipartite Gaussian states, the continuous variable Peres-Horodecki criterion
can then be written, which turns out to be a necessary and sufficient condition
for separability [31]. The criterion of Simon corresponds to the necessary and
Quantum solitons: towards experimental quantum communication 385
sufficient form of the criterion derived by Duan et al [32]. The rest of this
section is devoted to this latter criterion first in its sufficient and then in its
necessary and sufficient form as it allows one to use the conventional tools
of experimental quantum optics and is more convenient for the experimental
verification of entanglement of intense beams.
Duan et al [32] are deriving the non-separability criterion in the spirit of
two-mode squeezing [33]:
(24.12)
They introduce the EPR-like operators in the form of joint variables:
AliA
U= a Xl + -alAX2, AliA
V= a PI - -alAP2; (24.13)
[xj,lh] = i8j k (j, k = 1,2) (24.14)
where a is an arbitrary non-zero real number. Then the total variance of u
and v is derived under the assumption that (24.10) holds. The lower bound
for this total variance is calculated which represents a separability condition.
A sufficient criterion for non-separability of an arbitrary two-mode state is
obtained whenever the following inequality is fulfilled [32]:
be characterized by the normalized squeezing variances v;,~ (.X), v;,~ CY) which
are known from the context of two-mode squeezing [33]:
(24.16)
(24.17)
The field modes are denoted by the respective subscripts and SN denotes the
shot noise limit of the respective beam and gx = gy = 9 by the argument of
symmetry. Apart for a difference in the normalization, the variances in (24.15)
correspond to the squeezing variances in (24.16, 24.17). The joint variables
(24.13) for the optical beams are expressed in terms of amplitude and phase
quadratures.
To rewrite the non-separability criterion (24.15) in terms of measured vari-
ances of quadrature operators, let us consider the amplitude quadratures Xj
first. If we are dealing with bright beams, we can write:
n·"'n·+a·8X·
J'" J J J (24.18)
(24.19)
The variance of the combined variables read out by means of photo detection
is given by (24.18):
(24.21)
Quantum solitons: towards experimental quantum communication 387
The variance Yst (.X) corresponding to that of the joint variables of (24.13)
[32] is then obtained in the form:
For two beams 1 and 2 of equal intensity it takes a simple form discussed by
Reid in 1989 while relating their EPR criterion [24] for continuous variables
amplitude and phase to two-mode squeezing:
(24.23)
V+(X) = 2 V(U)
sq a2 + Q;21 for a= ff£
-
90.2
1
(24.24)
where the factor 2 comes from different normalizations used: [Xj,Pk] = iOjk
in (24.13) and [Xj, Yk] = 2iojk for quadrature operators, i.e. V(Xj) =
V ( ~ ). The latter normalization for quadratures is convenient in the current
context while it corresponds to the Heisenberg uncertainty relation bounded by
unity and the two-mode squeezing condition of Yst(X) < 1. Analogously,
V-(Y) = 2 V(V)
sq a2 + Q;21 for a= ff£
-.
90.2
1
(24.25)
(24.26)
example, the case for bright entangled beams generated in our experiment
[11] (Sec. 3.) and it corresponds to a particular choice of the Xj, Yj basis
ensuring maximal available correlations and to the optimized gain gOpt [32].
In general, a Gaussian state of n modes is completely characterized by a
covariance matrix [34] or a correlation matrix [35], the elements of which
are measurable quantities, covariances of Gaussian probability distributions
describing the correlations between all relevant conjugate variables in and
between all involved modes of a Gaussian bipartite state. Separability and
the positivity of the partial transpose of any Gaussian bipartite state can be
characterized in terms of the correlation matrix [34, 35]. The necessary and
sufficient condition for separability [32, 35] and distillability criterion [35] for
all Gaussian bipartite states have been recently derived using this formalism
(see also chapters 13-15 ofthis book). The experimental determination of the
correlation matrix for bright beams is rather evolved as it requires the use of
strong, phase-matched local oscillators or other techniques for tomographic
measurements of all quadrature operators. In our current experiments, we thus
restrict ourselves to the entanglement characterization using Eq. (24.26) [32].
However, the correlation matrix is an important tool for describing Gaussian
states. The development of experimental methods to record it are in progress.
QND (spectral shift); non-linear !:lip ex: ns (pp ex: ns) QND
Konig et al. [14]. !:lis ex: np (Ps ex: np)
non-separability criterion:
±
3. ~q (X) Vs~ (Y) < 4 a sufficient criterion
A A
Gedankenexperiment [24]
squeezed-state entanglement
EPR entanglement
limit:
(24.29)
(24.30)
To test experimentally for EPR entanglement one has to measure the sum
and difference photo currents, but a different normalization has to be used
compared to (24.16), (24.17). The sum and difference photo currents for both
amplitude and phase quadratures are normalized to the shot noise variance of
the beam described by the first variable in the argument of Vcond.
where the sign "«" emphasizes the ability to reliably differentiate between the
measured correlation level and the quantum/classical boundary V(X, YSN).
The limit of maximally entangled beams would imply
(24.32)
oj b)
o 2
Figure 24.1 Different boundaries for the continuous variable entanglement and respective
entanglement regions in terms of conditional and squeezing variances: (a) Demonstration of the
EPR-Gedankenexperiment [24] (dark grey region below the curve) and EPR-entanglement (light
grey square); (a) Non-separable quantum states [32, 31] (dark grey region below the line) and
squeezed-state entanglement (light grey square). The gain is taken to be gx = gv = 9 = 1 to
facilitate the comparison. Note, however, that this gain corresponds to the optimal one only on
the diagonal of the squares where Vcond(5XlIJX2)=Vcond(5i\15Y2) and v;.t(5X)=v;.~(5Y).
Thus all the depicted boundaries are sufficient conditions for a quantum state to be entangled.
The discussed limits are summarized in Table 24.2. The mutual relations
between these differerit boundaries is illustrated in Fig. 24.1. Figure 24.1
assumes l;r= gfft = g, as well as the whole discussion in this paragraph
does. Note that for entangled beams asymmetric in terms of uncertainties,
different optimal gains might be required for different quadratures. Moreover,
to be able to verify the non-separability condition 24.26 in its necessary and
sufficient form, local linear unitary Bogoliubov operations should be applied
to the two-mode quantum state in this case, i.e. local squeezing transformation
together with some rotations [32].
The boundaries (24.27), (24.28), (24.29), (24.30) seem to be the most con-
servative and reliable experimental criteria for the evaluation of the continuous
variable entanglement for quantum communication purposes.
(24.34)
Quantum solitons: towards experimental quantum communication 393
have to be satisfied for the relevant signal variable( s) where iI = iI0 + iII is the
total Hamiltonian of the measurement interaction (for explanations and details
about a QND measurement see part 3 of this chapter and references therein). An
example of such entanglement and its application in quantum communication is
the quantum teleportation scheme based on QND-entanglement between field
quadratures of bright light [27]. Further examples are considered thoroughly in
part 3, where the QND-entanglement emerging in soliton collisions in different
schemes is presented.
In analogy to the definitions of EPR versus squeezed-state entanglement,
consider the situation when the condition (24.8) is replaced by (24.27), (24.28):
(24.35)
That would imply a quantum correlation between signal and probe variables,
however, without necessarily the possibility to infer the signal variable As with
a precision below the quantum noise limit by a measurement of the probe Sp.
In this case the generated entanglement might be too weak for obtaining a QND
readout.
The interesting question which still has to be answered is how the EPR-
entanglement and the QND-entanglement are related to each other (see Fig.24.2).
One can introduce two sub-classes of the EPR-entanglement important for un-
derstanding EPR and QND properties. Two-way EPR-entanglement of some
conjugate pairs of variables Xl, Y1 ; X 2, Y2 obeys not only the EPR conditions
stated as:
(24.36)
Squeezed-state
entanglement
one-way EPR-entanglement
Figure 24.2 Relation between the different types of continuous variable entanglement.
394 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Signal Probe
measurement action
"QND
observable"
"Probe
observable"
back action
Figure 24.3 QND interaction as one-way EPR-entanglement. Solid and dashed lines corre-
spond to two possible pairs of correlated variables in Eq. 24.38,24.39.
or
(24.37)
but also the same relations under exchange of indices ("second way"). The
schemes for continuous entanglement generation involving OPO [10, 24] and
interference of squeezed beams [11,22] all are examples of two-way entangle-
ment.
For one-way EPR-entanglement holds (note the order of the indices I):
(24.38)
or
(24.39)
the signal phase (see Table 1). However, there is a certain asymmetry which
comes from the QND-requirements set on the signal variable. It means, one
can conclude on the signal photon number from the measurement of the probe
phase. It is also possible to infer the information of the probe phase from the
measurement of the signal photon number but it is not possible to do that with a
precision below the vacuum noise level. That would contradict the conditions
(24.34) set on the QND-variable of the signal. Thus this QND entanglement
represents the one-way EPR entanglement between conjugate variables.
For the aim of the QND measurement itself, however, it is not important
whether the EPR condition is satisfied or not. What is of interest is the entan-
glement between the relevant variable of the signal and the observable of the
probe. Therefore, in many QND experiments the behaviour of the conjugate
pair of the variables was not at all addressed. Are there always two entangled
pairs of conjugate variables arising in a QND interaction? Note, that even
if there appears to be only one entangled pair, e.g. Xl ex X2 but YI is not
quantum correlated with Y2 , the situation may change by moving into the other
basis, for example, XI cos () + YI sin () and X2 cos <p + Y2 sin <p. In general, the
quantum state can be non-separable in only one variable (in each subsystem)
and in this sense it can be entangled in one variable. This can be readily seen
from the non-separability criterion (24.26) which can be satisfied if only one
of the conditional variances obeys the condition (24.33) for EPR-like quantum
correlations.
An example of a QND scheme where there seems to be only one entangled
pair of variables, i.e. which is not an EPR-entanglement, is a QND measurement
of an amplitude quadrature of the signal field via an interference on a beam
splitter with an amplitude-squeezed probe field [37]. In the limit of perfect
squeezing and assuming a 50/50 beam splitter, the correlation coefficients for
both quadratures are equal to unity, x~ut ex x;n=out and Ysout ex y.;ut, but the
corresponding conditional variances are given by:
v;conds
(xin=outlxout)
p
-+ 0 , v;cond (y'0utlyout)
p s -+ 1 (24.40)
to review the bright entanglement schemes which can deliver a high degree of
entanglement, making full use of the advantages discussed in the introduction
and exhibiting experimental feasibility with the prospects for technological
applications.
Y
4
=ex + 8a
A 1\
A
Figure 24.4 Generation of EPR entanglement: interference of two amplitude squeezed beams
where the beam splitter is taken to be symmetric (t = 1/V2. r = i/V2).
In Fig. 24.4 this mapping is illustrated with the help of the input phasors in
phase direction (large arrows) spanning the uncertainty regions at the outputs.
The classical fields add vectorially and form a new amplitude direction 45°
to the input fields. The projections of the phase uncertainties of the incident
fields onto the new amplitude direction of the output beams show strict anti-
correlations, those onto the new phase direction strict correlations. This would
correspond to perfect entanglement. However, the finite amplitude uncertainties
of the squeezed input beams reduce the correlations as indicated by the small
light circles at the outputs. According to this the sum variance of the two
entangled output beams, reflecting the amplitude uncertainties, is decreasing
linearly with the squeezing at the inputs. Respectively, the difference variance
for the phase uncertainties is growing linearly. The phase diagrams in Fig 24.4
describe the interference of two single mode fields. The reasoning leading to
entanglement also holds in the more complex case of squeezed multi-mode
fields such as amplitude squeezed pulses [23].
VA
c:::J---88-b . . ., :. . . . ,.: . .
').J2 ••••::.
(.:~.:~:. :. ::~o
IL---f
Cr:YAG
laser
.....~....
squeezed
beam #2
Figure 24.5 Schematic of the experimental setup of a two-in-one-squeezer for the generation
of two squeezed beams. VA: variable attenuater, ),,/2: half-wave plate, G: gradient index lens,
PBS: polarizing beam splitter, 90110: beam splitter with 90% reflectivity. Insert shows the
polarization direction of the beam at the input of the fibre with respect to the main axes.
ensures the independence of the squeezing of the two output beams, which was
checked experimentally.
The Kerr nonlinearity induces an intensity dependent phase shift during the
propagation of the pulses along the fibre. This in tum provides a relative phase
shift between the strong and weak counter-propagating pulses having the same
polarization and it influences the quantum characteristics of the strong pulse. In
a single mode model the circular shaped phase space uncertainty of a coherent
light field is formed into an ellipse by the Kerr effectKerr,effect during the
propagation along the fibre [42, 43]. However, in order to be able to detect
the amplitude-squeezing in direct detection, the uncertainty ellipse has to be
realigned. The interference of the strong pulse with the weak one at the output
of a fibre Sagnac interferometer implements this re-orientation due to the Kerr-
induced relative phase shift between the strong and weak counter-propagating
pulses [42]. Still, the intensity dependent phase of the interference has to be
matched and hence the directly detectable amplitude squeezing depends on the
pulse energy. To detect the squeezing, the light is equally distributed to a pair of
balanced photo diodes of high efficiency and the generated photo currents are
added and subtracted. The noise variances of the sum and difference signals
give the noise of the output beam and the corresponding shot noise level.
Fig. 24.6 shows the results for the output ports of the described "two-in-one"
squeezer. The noise powers in Fig. 24.6 are recorded as a function of the input
energy and the grey areas indicate energy ranges where squeezing occurs.
~.
I'
~'
"2 .
I
1/. · \
I/ ~
":. \
E E
co ~ . III
... if
~ ~
....
Q)
·70 Q) -70 _ 1_-.-,
~ . : ~ ,..._--r
\j'\j
~.
, i
1\ ;1
c.. Co
, ..."
.. L - - - " r Q)
V\J
Q)
III III
'0
c: -75 'g ·75
,
;' , I
80
40 60 eo 100 120 140 160 ~~~0~4~0~6~0~e~0~1~00~1~~~1~~~1~6~O~
20
input energy [pJ] input energy [pJ]
Figure 24.6 Noise powers of the outputs of a "two-in-one" squeezer as a function of the input
energy; doted line: noise of the output beams; solid line: corresponding shot noise levels as
derived from the difference signal and verified by attenuation measurements. Inserts show the
normalized variances of the respective traces.
400 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
iii' 10
3:!..
CI)
o
~ 5
.s:;
<II
>
Q>
Ul
'0
c:
"C
.~ ·S
"iij
E
o .io,_ _ _-=-__ ~ __ __;_--_..,~-__""':'
c: o 6
Figure 24.7 Calculated noise variances of the output beams at the beam splitter versus the
squeezing of the input beams for minimum uncertainty states.
~I
I:
10
g.•
0
c .•
·50
E' .55
i
~ .60 i 1
{ ............. .1. .......... ....
~ ·65 i;."/
0.·70 SN
CI>
<I)
'0 ·75
1"' ................,
c
·80
20 40 60 80 100 120 140 160
input energy [pJ]
Figure 24.8 Experimental data: noise measurements of the output beams at the beam splitter;
insert shows normalized variance of the sum of the output beams demonstrating amplitude
quantum correlation
the trace labeled SN gives the shot noise level, which corresponds to the optical
power of the combined output beams. The grey areas mark energy regions,
where the fields entering the input ports of the beam splitter are amplitude-
squeezed (see Figs. 24.7 and 24.8).
The calculations (see Fig. 24.7) predict no correlations for coherent states
with shot noise limited uncertainty for amplitude and phase. In the experimental
data of Fig. 24.6 and Fig. 24.8 these states should be found at the boundaries of
402 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
the grey areas, which correspond to 0 dB input squeezing. However, in Fig. 24.8
the noise levels of the sum, difference and single beam signals (traces 1 - 4) at
odB input squeezing differ significantly from each other indicating still existing
classical anti-correlation of the amplitude quadratures. This phenomenon can
be understood by additional noise in the phase quadrature of the input beams
due to the interaction of the photon pulse with thermal phonons in the fibre,
which leads to classical amplitude anti-correlation in addition to the quantum
effects. This classical anti-correlation can result in high correlation coefficients,
but the conditional variance as the figure of merit for the quantum correlation
will never drop below unity. The enlarged phase noise is present for all energies
and explains the high noise powers of the individual output beams as well as
the noise trace for the difference photo current. In contrast, calculations for
the sum photo current show, that the noise variance of the sum is not affected
by the additional phase noise. The noise power of the sum photo currents was
observed to drop up to 4.0 ± 0.2 dB below the shot noise level, the quantum
limit for the anti-correlation. As expected, this quantum anti-correlation is
limited by the squeezing of the input beams, as seen by the comparison of Fig.
24.6 and Fig. 24.7.
Simultaneous squeezing of the input beams is also recorded for higher ener-
gies. However, in that region the sum variance of the amplitudes lies above the
quantum limit. There are two possible explanations. Firstly, the interference
was optimized for lower energies and, secondly, the detection of the quantum
amplitude anti-correlation is very sensitive to the balance of the photo-detector
pair. In the ideal situation traces 3 and 4 for the single beams in Fig. 24.8
should coincide for all energies, but they separate for the high energy range.
This can be attributed to an imbalance of the photo detectors which where
calibrated for the lower energy band around 100 pJ.
For the complete determination of the EPR-entanglement in addition the
quantum phase correlation has to be investigated. In order to directly measure
the phase quadrature variances, strong optical local oscillators with matched
spectra and wavefronts are required. To circumvent the experimental difficulties
involved when applying this technique to bright light pulses an interferometric
measurement scheme for the characterization of the state was invented.
In Fig. 24.9 the source of the EPR-entanglement is treated as a black box
with two output beams. These beams are made to interfere at yet another SO/50
beam splitter with the relative phase of the beams being scanned (Fig. 24.10).
To prove the non-separability of the state it is now sufficient to record the
amplitude noise variance in addition to the shot noise of one of the resulting
output beams c and d. Calculations show, that for a suitable interference
phase the reduced amplitude noise of a single output beam is equivalent to
the measurement of i: (V(Xa + Xb) + VCYa - Yb)). The shot noise levels
Quantum solitons: towards experimental quantum communication 403
balanced
Scan of detector
Interference
Phase e
Figure 24.9 Schematic for indirect measurement for the phase quadrature correlation
V(X1,SN + X2 ,SN) and V(Y1,SN + Y2 ,SN) are both equivalent to twice the shot
noise ~ (Xc/ d,S N ) of cor d. Thus the normalized amplitude noise of the beams
cand d is given by
That means the described setup permits the direct detection of the Peres-
Horodecki non-separability criterion for continuous variables (see Eq.24.26):
+
Vsq (X) + Vs~ (Y) < 2.
A A
In the experiment a value of vst(X) + Vsq (Y) = 0.80 ± 0.03 was observed,
demonstrating the high non-separability of the state. Fig.24.10 represents the
corresponding experimental data.
a
Provided the variance of the sum of the amplitude of beams and bis already
known, the quantum phase correlation can be inferred from the measured
amplitude noise of c or d. A quantum phase correlation of up to 4.0 ± 0.4 dB
was verified from the data of Fig. 24.10.
Note, that all indicated experimental values describe the noise statistics of
the light without correction for linear losses. The photo-detectors showed a
detection efficiency of around 92 ± 5% and the noise powers were recorded
for a detection frequency of 10 MHz with a resolution bandwidth of 300 kHz.
The measured data are corrected for the electronic noise of the photo-detectors
and the spectrum analysers. For quantum information applications the dark
noise can limit the availability of the entanglement. If one corrects only the
shot noise level for electronic noise to evaluate the quantum limit and for an
uncorrected signal noise, the value of 1.12 ± 0.06 < 2 was observed in the
experiment for the Peres-Horodecki criterion.
404 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
...... - + channel
E
m
......
"0
...CD -6
~
0
a..
CD
III
'0
- shot noise level =
Z quantum limit
noise of the input fields in this case the optimal gain for (24.8) was lIt = 1.
Similarly, for the phase quadrature difference squeezing (24.28, 24.17) of also
up to 4.0 ± 004 dB was associated with V- CY) = 0040 ± 0.04 < 1. This
indicates a quantum correlation with the same non-classical conditional vari-
ance of Vcond(YalYb) = V~~-J~b)
a,SN
= 0.80 ± 0.07 < 1 (the optimal gain is
again g?Jt = 1) . Thus the EPR-paradox for continuous variables (24.9) was
demonstrated:
(24.41)
Quantum solitons: towards experimental quantum communication 405
(24.42)
.9 ~ 0.9 ~
.3
#
"
Q) 0.3
~
"
Q)
-0.3~ -0.3 ~
.~ .~
-3.2 -1 .6
-0.9 Q
F 0 16 -0.9 Q
0.4
reqUency ITHz 3.2
Figure 24.11 Time and frequency domain interaction of two fundamental solitons. The in-
tensity of the field is displayed in a reference frame moving with the center of mass of the
solitons.
(24.43)
it can immediately be seen that the signal photon number is preserved, i.e.
[Hs, ns] = O. If the photon numbers are as large as 108 , a linearized description
of the field fluctuations can be used [60]. In terms of quadratures the photon
number or amplitude quadrature can then be written as Xs = ns/(2) (ns))
and the conjugate phase quadrature as Ys = J(ns)~s.3
The Kerr effect couples two copropagating and interacting modes in a fibre
by an intensity dependent phase modulation (XPM). In soliton QND schemes
XPM coupling is achieved by a soliton interaction, i.e. the collision of a signal
with a probe soliton. This is depicted in Fig. 24.11 in the time and in the
frequency domain. The colliding pair is described by an unbound solution
of the nonlinear SchrOdinger equation. Maximum pulse overlap occurs in the
center of the collision at z = O. Before the pulses begin to overlap they
propagate with unaltered shape as expected from single solitons. The collision
transiently disturbs the envelopes of the pulses, which retrieve their pulse
envelope, energy, and momentum after the interaction. Therefore only two of
the four soliton parameters, the photon number and momentum, are conserved.
The photon number preservation during the interaction can be verified with
the Hamiltonian of Eq. 24.43, [HI, ns ] = O. From the pulse trajectories an
acceleration of the pulses towards each other is visible, causing a subsisting
shift in phase and position, which depends on the intensity of the other soliton.
The probe soliton phase shift is a possible readout for the signal amplitude,
because HI is a function of ns and [HI, Yp] # 0 ([yp, np] # 0) [25,63] . For
a small phase shift, Yp represents the probe phase shift, which can be read
out by an interference with an additional phase reference soliton. This type of
QND interaction is explained in section 4.3. Transient changes in the solitons
408 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
during the collision such as the change in relative velocity of the pulses can
be used for a QND readout as well: Because of the chromatic dispersion of
the fibre a relative group-velocity increase corresponds to a transient increase
in the spectral separation of the pulses (Fig.24.11). Note, that in the center
of the collision the spectral density at line center approaches zero. This holds
for any relative velocity and phase of the solitons provided they have equal
amplitudes. Detecting the spectral shift of the probe pulse serves as a QND
readout of the amplitude of the signal soliton, as explained in section 4.4. To
date, these two are the only schemes for a QND measurement with solitons in
fibres. XPM provides the cross talk mechanism to write the signal intensity
information into the probe system without degrading the signal-to-noise ratio
of the signal. Thus XPM generates the desired entanglement of the signal and
the probe system. The back-action noise introduced by the measurement is fed
into the signal phase and position only (Fig. 24.3).
QND measurements based on the Kerr effect were among the first proposals
for QND measurements in optics [50,51,52]. The first QND experiment was
realized using the Kerr effect in fibres [64]. Fibre-optical experiments aiming
at back-action evading measurements of a quantum-limited signal intensity
are listed in Table 24.3. The experiments used shot-noise limited intensity
Table 24.3 Experiments aiming at fibre QND detection of a shot noise limited signal beam.
Pump light denotes the signaUprobe power or pulse energy and center wavelength. The corre-
lation coefficient C is the fraction of probe readout photocurrent rms noise due to signal shot
noise. Leff is the effective interaction length, different from fibre length due to resonator finesse
[65] or pulse walk-off [14,26,66].
(24.44)
Here 'Y is the QND gain and a the self phase modulation (SPM) coefficient.
Note that both depend on the interaction lengths. The probe readout quadrature
oY.;ut contains the signal oX;n. The inevitable back-action noise due to the
measurement is channeled into the quadrature Ys, conjugate to Xs. However,
one typically finds that a > 'Y and the probe observable is contaminated by
SPMnoise.
The corresponding fibre QND scheme is shown in Fig. 24.12. In order
to evaluate the performance of a certain QND device, transfer coefficients
are practical quantities for the determination of the signal degradation and
information transfer to the probe observable [28, 70, 71]. They are measureable
quantities even in a single stage BAE measurement. In the case of uncorrelated
input states, i.e. (x;nYsin)sym = 0, the transfer coefficients can be derived
directly from the correlation coefficients [28]:
(24.45)
410 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
0.9 ~
"
c:8 c:If
~
-0.3.fY -0.3 .fY
.f!} .f!}
-3.2 -1 .6
-0.9 Q -0.9 Q
0.4 F 0 16
reqUency ITHz 3.2
Figure 24.12 QND scheme using XPM induced phase changes in optical fibres. The signal-
intensity dependent refractive index changes the phase of a probe pulse and vice versa. A QND
measurement of the signal amplitude can be performed via a probe phase detection, leaving the
signal amplitude unperturbed (after Sizmann and Leuchs [43]).
Here sin, sout and pout refer to the observables in a general QND measure-
ment. The inequality
(24.46)
characterizes a quantum measurement, in which the signal information is noise-
lesslyamplified. Using Eq. 24.44 we can calculate the transfer coefficients for
a coherent signal and a coherent probe input (V (x~n) = V (x~n) = V cv;n) =
1):
quadratures [72, 73]. The probe output fluctuations are then measured in the
quadrature Bp:
8B;ut('Ij;) = 8X;ut cos('Ij;) + 8y;ut sin('Ij;)
and for optimized angle 'lj;o [43]:
The probe observable, 8B~ut then contains only the probe input phase quadra-
ture noise and the QND copy of the signal. In analogy to Eq.24.47 the sum of
the transfer coefficients is:
,2
T s +Tp = 1 + - 12' (24.49)
+,
This semiclassical analysis shows that an SPM-noise-free QND measurement
is possible when using a combined amplitude and phase detection.
Negative Dispersion
+
Probe Reference Optical Fiber,
"
400m
-----J\~A~/\~O
From Two-Color "
O~
Soliton Source Signal
Figure 24.13 Outline of a soliton QND measurement. After the probe-signal collision in the
fibre, the phase difference between probe and reference is a readout of the signal soliton photon
number. For phase detection, probe and reference interfere with a 11"/2 relative phase delay in
addition to the group delay (after Friberg et al. [26]).
only a different pulse sequence and a longer fibre to realize two collisions. In
the experiment it was shown that probe and reference, when both colliding
with the signal, experience the same XPM to within 13 dB. However, repeated
BAE detection requires individual readouts of the two probe pulses of the
two collisions, which can be implemented in this scheme with two additional
solitons which do not collide with the signal soliton.
The soliton QND experiments discussed so far were limited by detection
efficiency and by SPM noise of the probe system. The two limitations were
inherent in the Mach-Zehnder and the pulse delay detection scheme. The SPM
noise limitation leads to a predicted conditional variance which is at best 0.8,
corresponding to Ts + Tp = 1.2 in this case [72]. To get closer to an ideal
soliton QND experiment the probe would have to be detected with a linear
combination of amplitude and phase quadratures [72, 77, 78]. A different
detection scheme [73] could reduce both classical GAWBS noise and quantum
SPM noise simultaneously.
Quantum solitons: towards experimental quantum communication 413
-eco
.~ ~.,.:-+--:.>4.J.L...,..~""-l
c
OJ
"0
1i!
uOJ
C.
en
14SO 1550
Figure 24.14 The QND experiment using frequency shift detection (left). Experimental spectra
(right) (a) of two colliding pulses in the center of the collision and (b) of the pulses without a
collision.
solid-state laser emitting 150-fs pulses with 163 MHz repetition rate at the
wavelength of 1.5 p,m. Each laser pulse is separated spatially by a dichroic
mirror into two shot noise limited and uncorrelated signal and probe pulses,
each 14 nm in spectral width and spectrally separated by 15 nm. Because of the
short length of the pulses only a short interaction length is required to produce
spectral correlations [85]. The timing of the pulses is adjusted such that they
fully overlap at the end of the fibre. This locates the center of the collision.
Output spectra of colliding pulses (a) as well as of non-colliding pulses (b) are
displayed in Fig. 24.14 (right). The spectral shift of the respective signal and
probe spectra is clearly visible and amounts to approximately half the spectral
width of the single pulses (7 nm). In addition to this a modulation appears in
the spectrum due to interference with a dispersive background wave created by
the slight chirp of the input pulses.
After the pulses leave the fibre they are separated using a grating. In addition,
the probe pulse is spectrally filtered by a subsequent spatial slit. Note that the
photon number of the probe pulse is preserved in the interaction. The spectral
filtering then induces losses to the probe pulse and the transmitted photon
number is taken as the probe output. Correlations between the signal and
the probe outputs are detected as follows: The photocurrent of the probe is
attenuated by an amount a (variable gain 9 in Fig. 24.14) and added to the
signal photocurrent. The input-output relations can be written as
proportional to:
-
(l)
"0 6
'-
Q)
4
~
0
0.
2
Q)
en
·0
c 0
Q)
'-
-2
0.0 0.3
relative attenuation a
Figure 24.15 Reduction of the signal noise power below the signal shot noise level. A
conditional variance of 0.73 results.
Acknowledgments
The authors are grateful to R. Loudon, D. Ostrowsky, R. Werner, B. Buchler,
S. Lorenz and T. C. Ralph for fruitful discussions and their support.
Notes
1. In principle deterministic perturbations of X. are allowed [29.54].
2. In fibres the absorption is typically as low as 0.2 dBlkm at >. = 1.5 j.tm.
References
[1] E. Schr6dinger, Naturwiss., 23, 807 (1935).
[2] A. Einstein, B. Podolsky, and N. Rosen, Phys. Rev. 47, 777 (1935).
[3] D. Bohm, Quantum Theory. Prentice Hall, Englewood Cliffs, NJ, 1951.
[4] J. S. Bell, Physics 1,195 (1964).
[5] J. F. Clauser, M. A. Home, A. Shimony, and R. A. Holt, Rev. Prog. Phys.
23, 880 (1969); J. F. Clauser and A. Shimony, Rev. Prog. Phys. 41, 1881
Quantum solitons: towards experimental quantum communication 417
(1978).
[6] A. Aspect, P. Grangier, and G. Roger, Phys. Rev. Lett. 47, 460 (1981);
A. Aspect, J. Dalibard, and G. Roger, Phys. Rev. Lett. 49, 1804 (1982).
[7] C. H. Bennett, G. Brassard, C. Crepeau, R. Jozsa, A. Peres, and
W. K. Wootters, Phys. Rev. Lett. 70,1895 (1993); C. H. Benett, F. Bessette,
G. Brassard, L. Salvail, and J. Smolin, 1. Cryptology 5,3 (1992).
[8] For a review see W. Tittel, G. Ribordy, and N. Gisin, Physics World,41
(March 1998); N. Gisin, G. Ribordy, W. Tittel, and H. Zbinden, Rev. Mod.
Phys. 74, 145-195 (2002).
[9] D. Bouwmeester, A. Ekert, A. Zeilinger (Eds.), Physics of Quantum In-
formation. Springer, Berlin, 2000.
[10] Z. Y. Ou, S. F. Pereira, H. J. Kimble, K. C. Peng, Phys. Rev. Lett. 68,
3663 (1992).
[11] Ch. Silberhorn, P. K. Lam, O. Weill, F. Konig, N. Korolkova, and
G. Leuchs, Phys. Rev. Lett. 86,4267 (2001) and quant-phlOI03002.
[12] Y. Zhang, H. Wang, X. Li, J. Jing, C. Xie, and K. Peng, Phys. Rev. A 62,
023813 (2000).
[13] K. Bencheikh, J. A. Levenson, Ph. Grangier, O. Lopez, Phys. Rev. Lett.
75,3422 (1995).
[14] F. Konig, B. Buchler, T. Rechtenwald, G. Leuchs, and A. Sizmann,
"Soliton back-action evading measurement using spectral filtering", Phys.
Rev. A, submitted; F. Konig, T. Rechtenwald, M. A. Zielonka, R. Steidl,
G. Leuchs, and A. Sizmann, "Quantum-nondemolition measurement us-
ing spectral correlations between fibre-optical pulses." Conference on
Lasers and Electro-Optics/ Quantum Electronics and Laser Science Con-
ference CLEO/QELS'2000, San Fransisco, California, May 7-12, 2000,
Technical digest, QThI28, p. 206.
[15] F. Konig, M. A. Zielonka, and A. Sizmann, "Transient photon-number
correlations of interacting solitons", Phys. Rev. A 66, in print (2002);
A. Sizmann, F. Konig, M. A. Zielonka, R. Steidl, and T. Rechtenwald,
"Quantum correlations of colliding solitons", in Massive WDM and TDM
Soliton Transmission Systems (A ROSC Symposium). A. Hasegawa (Ed.),
Kluwer Academic Publishers, Dodrecht 2000, p. 289-298.
[16] A. Kuhn, M. Hennrich, T. Bondo, and G. Rempe, Appl. Phys., 69, 373
(1999); M. Hennrich, T. Legero, A. Kuhn, and G. Rempe, Phys. Rev. Lett.
85, 4872 (2000).
[17] c. K. Law and H. J. Kimble, 1. Mod. Opt. 44, 2067 (1997).
[18] B. Huttner and J. Brendel, "Photon-counting techniques for fiber mea-
surements", Lightwave, August 2000, pp. 112-120.
418 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Abrams-Lloyd, 42 Decoherence, 40
Algorithm Dense coding, 64, 95
Abrams-Lloyd, 42 Deutsch-lozsa, 7, 32
Deutsch-lozsa, 7, 32 Displacement-covariant, 280
Grover, 8, 33 Distillability, 174
Shor, 7, 41 Down conversion, 79
Anticlone, 284 Eavesdropper, 290
BB84 protocol, 332 Efficient simulation, 47
Beam splitter, 27, 112, 117, 256, 285-286 Entanglement, 60, 250
Bell inequality, 5 bipartite, 112
Bell-state, 63, 78, 86 bound, 174, 180,211,213,218
Bit error rate, 298, 302 distillation, 173,207,323
Bit, 4 entropy, 194
Cauchy-Schwarz inequality, 123 fidelity,73
Characteristic function, 175 measure, 61, 194
Check pairs, 327 mixed,115
CHSH inequality, 133 multipartite, 111-112, 114, 125
Circuit partial, 141
quantum, 116 purification, 65, 173, 193, 323
Classical teieportation, 89 quantification, 194
Clifford swapping, 64
algebra, 49 EPR,59,68,255,382,390,396,400,404
group, 49-50 beams, 81,97
Cloning, 277 pair,62,78
fidelity, 283, 286 state, 69, 281
CNOT, 48, 50, 116,281 Error
Coherent correction, 19
amplitude, 71 probability, 298, 305
state, 71, 300 Excess noise, 279
Commutation relation, 106 Exponential speedup, 43
Completely-positive, 280 Factorize, 7
Computational complexity, 7 Feed-forward, III
Concatenation of c\oners, 286 Fidelity, 68, 79, 89
Conditional variance, 382 cloning, 283, 286
Conjugate observables, 278 Fourier, 117
Correlation matrix, 138, 175 Gate
Covariance maxtix, 175 Pauli,48
Covariant, 280 phase, 48, 51
CSS code, 330 SUM, 50, 52
Damping channel, 348 Gaussian, 43, 70, 219
423
424 QUANTUM INFORMATION WITH CONTINUOUS VARIABLES
Squeezed Transform
light, 112, 308 discrete Fourier, 287
sUrte, 42, 112, 180,236,284,319,388,413 Fourier, 20, 282
vacuum, 13,81 Hadamard, 20
Stabilizer, 48 Transpose
generators, 320 partial, 113-114, 180,218
Standard form, 176 Tritter,27
State swapping, 260 Two-mode cat sUrte, 202
SUM gate, 50, 52 Vacuum fluctuations, 82, 283
Symplectic, 176 Virial theorem, 26
Syndrome, 25-26 Von Neumann entropy, 62, 96, 113, 116
Telecloning, 130 W state, lIS
Teleportation, 63,67, 77,255,258 Wave-packet, 22
classical, 89 Weyloperator, 175
Theorem Wigner
Gottesman-Knill, 47 function, 68, 73, 99, 128, 384, 396
no-cloning, 278 representation, 128
virial,26 XOR, 21, 33, 243-244
Transfer coefficient, 409 XPM coupling, 407