You are on page 1of 42

frog. Quanr. Elecfr. 1995, Vol. 19. pp.

89-130
0 1995. Elsevier Science Ltd
Pergamon 0@7!3-6727(94)00007-7 Printed in Great Britain. All rights reserved.
0079-6727195 $29.00

MEASURING THE QUANTUM STATE OF LIGHT

U. LEONHARDT and H. PAUL


Arbeitsgruppe ‘Nichtklassische Strahlung’ der Max-Planck-Gesellschaft an der
Humboldt-Universitlt zu Berlin, Rudower Chaussee 5, 12484 Berlin, Germany

Abstract-We analyse several schemes for measuring the quantum state of a single light mode.
These schemes are useful for experimentalists to gain the maximum information about non-classical
light fields which they investigate. Measured or reconstructed from measurements are quasiprob-
ability distributions for quadrature components. We address beam splitting, amplification and
heterodyning as experimental schemes for measuring the Q function. We show that the squared
Wigner function can be directly measured as a probability distribution. We analyse Optical
Homodyne Tomography where the Wigner function is reconstructed from measured quadrature
distributions. Finally, we discuss the influence of detection inefficiencies on the measured or
reconstructed quasiprobability distributions. We show how this influence can be reduced.

CONTENTS

1. Introduction 89
2. Quantum Mechanical Description of Light 92
2.1. Basic concepts 92
2.2. Wave function 93
2.3. Wigner function 96
2.4. Quasiprobability distributions 98
3. Experimental Tools 101
3.1. Photodetector 101
3.2. Beam splitter 102
3.3. Homodyne detector 103
3.4. Realistic homodyne detector 104
3.5. Amplifier 107
4. Measurement Schemes 109
4.1. Phase measurement and Q function 109
4.2. Determination of the Q function via amplification 113
4.3. Measuring the Q function in heterodyne detection 114
4.4. Squared Wigner function 116
4.5. Optical Homodyne Tomography 118
5. Inefficient Detection 122
5.1. Beam splitting 122
5.2. Amplification 123
5.3. Optical Homodyne Tomography 125
5.4. High-accuracy optical homodyne detection 126
6. Summary 128
Acknowledgements 128
References 129

1. INTRODUCTION

Most physicists have felt, and still feel, that an unsatisfactory feature of quantum theory,
in contrast to classical wave theory, is that it relies on the concept of a wave function that
lacks direct physical meaning. Apparently, this higher level of abstraction in the description
of physical reality is the price to be paid for the formal synthesis of the particle and the wave
picture which are incompatible in classical physics. The quantum-mechanical wave function
89
90 U. Leonhardt and H. Paul

cannot be measured on principle. On the other hand, it contains the maximum information
one can have about a quantum system, and there are well-known rules that allow us to predict
the outcomes of measurements, such as the probability of finding a given value as the result
of the measurement of a certain observable, the average value (quantum mechanical
expectation value) of an observable or its mean square deviation. In this context, it should
be noted that all quantum mechanical predictions are of a statistical nature. Accordingly, a
wave function can be attributed only to an ensemble of equally prepared systems but not to
an individual system.
The question then arises as to how one can determine the wave function if it cannot be
measured? The usual procedure is to utilise theory itself for this purpose: Since it associates
a given possible value of an observable with a definite wave function which is the
eigenfunction of the corresponding eigenvalue problem, the wave function can be inferred
from the process of preparing the system. For instance, knowing the detailed energy level
scheme for a given atom, one can excite a definite energy state via light absorption by
appropriately choosing the light frequency, or one concludes from spin conservation that the
two photons emitted subsequently by an atom in a j = O-r1 +O transition must be in an
eigenstate, with eigenvalue 0, of the total spin. (The latter state has recently played an essential
role in the experimental realization of the famous Gedanken experiment by Einstein,
Podolsky and Rosenc4).)
Having thus assigned a definite state to the system, one may follow its evolution which
results from an interaction with a second system whose initial state is also known, by solving
the corresponding SchrGdinger equation. In this way more complex wave functions come into
‘existence’.
Such coupled systems also offer a new opportunity for measurement. This is because a
measurement on a single system to prepare the system in a definite state (in an eigenstate of
the Hermitian operator corresponding to the observable in question) is of little practical use
here, since in the measuring process the system usually gets lost in the sense it is no longer
available for further experiments, or it is even destroyed completely as is the case in
photodetection. However, when we are dealing with a complex system composed of two or
more subsystems and known to be in a specified pure state which is entangled, we can infer
from a measurement on one subsystem the wave function of the second subsystem without
in any way disturbing the latter. This type of measurement, in fact, lies at the heart of the
Einstein-Podolsky-Rosen paradox. (23)An example is given by the resonant interaction of a
single-mode field in a microwave cavity with an excited atom traversing the cavity. Measuring
the energy state of the atom after it has left the cavity where it is certainly completely
decoupled from the field yields information on the field state, as modified by the interaction.@)
It should be emphasised that in all cases considered thus far the wave function is a
mathematical construct and it is obtained from the theory by solving an eigenvalue or an
initial value problem. In the case of a compound system, the reduction of the total wave
function is the formal mathematical technique that yields the wave function for the
non-observed subsystem.
On the other hand, the squared modulus of Schradinger’s wave function for a particle does
have a well-defined physical meaning: It gives us the probability density of finding a particle
at a given position and, hence, is directly measurable. What one still needs for a complete
characterisation of the quantum mechanical state is therefore the information on the phase
of the wave function. We thus arrive at the basic question “Can experimental schemes be
devised that provide us with this additional information?” Since the phase is not directly
measurable, its determination requires indirect techniques. Thus, we will have to reconstruct
the wave function mathematically from experimental data provided by measuring different,
presumably canonically conjugate, variables. It is important to note that this poses no
experimental problem: The measurements of different variables can be performed on different
Measuring the quantum state of light 91

subensembles selected at will from the whole ensemble for which the wave function is
characteristic. In fact, the problem of reconstructing the wave function from the distributions
for position and momentum, respectively, had already been addressed in the early days of
quantum mechanics. (58)To date, however, no practicable scheme for the solution of this
problem has been devised. Only recently, the situation has drastically improved, however, in
quantum optics. (F.)In fact, one of the exciting features of this field is that it greatly enlarges
the realm of feasible experiments to test quantum physics. One of the most powerful
techniques in quantum optics is balanced homodyne detection (“J’.~‘)in which the field under
study is optically mixed with a strong coherent reference beam commonly referred to as the
local oscillator. For a single-mode field it becomes possible with this technique to measure
not only two canonically conjugate field quadrature components that are the analogues of
position and momentum of a material harmonic oscillator, but any linear combination of
them, corresponding to a rotation by an angle 0. Experimentally, this rotation is simply
brought about by phase-shifting the local oscillator by 0. So a large (in principle infinite)
manifold of distribution functions we(xe) can readily be measured, where x6 is the
quadrature component observed when the local oscillator phase has been set to 0. This
wealth of experimental information makes it, in fact, possible to perform the quantum-state
reconstruction not only for pure states but also for mixed states. Thus, the problem can be
solved in its full generality: It has been shown theoretically (79)that the Wigner function which
contains the complete information on the quantum state of the system can be reconstructed
from a set of distributon functions w,(x,), with 0 gradually varying from 0 to n. This scheme
closely resembles the tomographic technique known from medical examinations and, hence,
has been termed Optical Homodyne Tomography. It is only recently that experiments along
those lines have successfully been performed.(9,‘0~6’~7’-73)
While the tomographic technique is certainly rather involved, one can get information on
the field state also with much less effort, namely directly from appropriate measurements, at
the price, however, of measurement accuracy. It has been shown theoretically(25~38,39J4J that two
canonically conjugate field quadrature components can be determined simultaneously,
however, in a noisy (fuzzy) measurement. This is achieved by utilising a beam splitter to split
the field in two. Each of the two variables is then measured on a different outgoing field from
the beam splitter. The distribution function for such simultaneous measurements was
predicted(25,38*39*84)
to be given by a function that is well known from theory, namely the
so-called Q function. This is a special quasiprobability distribution which, in contrast to the
Wigner function, has the merit of being non-negative. (The latter property is, of course, a
requirement to be met by any probability distribution.) Actually, the Q function can be
interpreted as a smoothed Wigner function. This smoothing process, on the one hand, is
responsible for the positive-definiteness of the Q function since it causes negative values the
Wigner function may take on to disappear. On the other hand, it leads to a certain loss of
information where the finer details of the Wigner function are wiped out as well. (In fact,
negative values in the Wigner function is by itself such a detail which reflects an intrinsically
quantum mechanical feature.) The smoothing thus indicates, in a rather intuitive way, the
noisy character of the measurement originating from beam splitting. (Indeed, it is the vacuum
noise which inevitably enters the unused port of the beam splitter, that is the noise source.)
What has been said thus far about the experimental possibilities of reconstructing the
Wigner function from measurements or directly measuring the Q function relies, however,
critically on the assumption that ideal (this means unit-efficiency) photodetectors are
employed. Strictly speaking, one has to require that the overall detection efficiency which
comprises any kind of losses (in particular, those due to mode mismatch) the field undergoes
until it is ultimately detected, attains the optimum value unity. In practice, this is certainly
not the case, and this in fact causes an additional deterioration of the measurements. In a
recent systematic study’42’it was shown that this undesired effect is correctly accounted for
92 LJ. Leonhardt and H. Paul

by further smoothing those distributions that would be measured, or reconstructed, with ideal
detectors. It is interesting to note that the smoothed distribution functions in question fit
precisely in a theoretical scheme which already exists, namely that of generalised quasiprob-
ability distributions. As early as 1969, Cahill and Glauber (‘4~‘5)
introduced such a concept into
quantum optics, with the intention to embed the three well known quasiprobability
distributions-the P function, the Wigner function and the Q function-in a larger manifold
of functions which they termed s-parametrised quasiprobability distributions. In fact, the
smoothed distribution functions mentioned above were identifiedc4’) with certain s-parame-
trised quasiprobability functions, whereby the parameters is related to the (overall) detection
efficiency 11in a rather simple way. So it is just a variety of those unfamiliar distributions
(depending on the actual parameter q) which can be observed directly, or reconstructed, in
realistic measurements, rather than the Q and the Wigner function (corresponding to s = - 1
and s = 0, respectively) that have found widespread interest in theoretical quantum optics.
The present paper is devoted to the exciting potential of quantum optics to determine
experimentally quasiprobability distributions. Such measurements constitute, in fact, a new
type of observations which provide us with a holistic picture of the system, whereas
conventional measurements yield detailed information on just one physical variable (for
instance, the photon number) leaving the conjugate variable in obscurity. Similar experiments
are possible using atoms for probing the quantum state of light.(22,26*86) However, they are
much more difficult to perform.
After introducing the formal concepts for the description of the field state (wave function,
density operator and quasiprobability distributions) in Section 2, we discuss in Section 3 the
main experimental tools based on photodetection. We will briefly describe the action of a
photodetector with arbitrary efficiency and then focus on the balanced homodyne detection
scheme. In addition, we analyse the action of a beam splitter and of an optical amplifier on
the quantum state of the system. In Section 4 we describe the experimental schemes for
measuring quasiprobability distributions. We will address beam splitting, amp@ication and
heterodyning as techniques for measuring the Q function of the field. Additionally, we will
show that the squared Wigner function can be directly measured for certain states. Further,
we describe Optical Homodyne Tomography as a means for reconstructing the Wigner
function from experimental data for all states. Finally, in Section 5 we analyse the degrading
effect of inefficient detectors and show that this can be avoided in balanced homodyne
detection when the signal is parametrically preamplified.(47)

2. QUANTUM MECHANICAL DESCRIPTION OF LIGHT

2.1. Basic concepts


The quantisation procedure for the electromagnetic field rests on a perfect formal analogy
between a monochromatic field mode and a harmonic oscillator. It amounts to replacing the
amplitude and its complex conjugate, A and A *, in the expression for the electric field strength
of a free single-mode field
E(r, t) = c(r, t)A + c*(r, t)A* (1)

by the photon annihilation operator and its Hermitian conjugate, the photon creation
operator, a and 6+. Hence the operator for the electric field strength is given by

B(r, t) = C[c(r, t)ci + c*(r, t)ci+]. (2)

The function c(r, t) = c(r)exp( - iot) (co circular frequency) describes the spatial field
distribution (for fixed time t) and its propagation (for varying t). It is the same in both
classical and quantum theories. The constant C is a normalisation factor. Usually one
Measuring the quantum state of light 93

identifies the field modes with classical eigenmodes which obey either physical boundary
conditions, as dictated by the presence of a cavity, or fictitious ones that select spatially
periodic solutions of Maxwell’s equations. Hence the field modes are monochromatic. We
would like to emphasise, however, that Eqns (1) and (2) hold true also for coherent
nonmonochromatic waves,(“) in particular for short laser pulses with given pulse shape.
The photon annihilation and creation operators obey the commutation rule

[h, ci+]
= 1, (3)

and their normally ordered product

fi =B’i (4)

has the meaning of the photon number operator. Then fi = hwtitci is the monochromatic-field
Hamiltonian, and, hence, the time-evolution operator is given by

O(t) = exp( - iwtci+&). (5)

The effect of temporal evolution is just a phase shift A0 = wbt because only the combination
wz + 0 enters the expession (1) for the electric field strength, where - 0 is the phase of the
complex amplitude A. Hence we can identify the operator

6(Q) = exp( - iOB+ci) (6)

with the phase-shifting operator. It shifts the phase of the photon annihilation operator Li
according to
ri+(@)ciir(O) = L?eeie. (7)

The eigenstates of fi, the so-called Fock states In) where n = 0, 1,2, . . . , provide a
complete orthonormal basis for the Hilbert space of the single-mode field under consider-
ation. This enables us to expand any pure field state in the Fock basis

2.2. Wave function


The formalism based on photon annihilation and creation operators is certainly well suited
to the study of interaction processes with atoms and molecules, since in this case annihilation
and creation of photons actually takes place. There are, however, different experimental
situations (and later on we will be dealing with them) in which it proves advantageous to
represent the field state by a Schrodinger wave function. In fact, this is what is usually done
in quantum mechanics of material systems. There, the basic variables are the position and
its canonical conjugate, the momentum, and the argument of the Schriidinger function is the
coordinate in ordinary space X, or speaking more pecisely, the continuous eigenvalue
spectrum of the position operator. So in order to treat a field mode in a similar way, we first
have to seek the analogs of position and momentum. We will denote these new variables by
x and p in order to stress the analogy to mechanics.
It is easily shown that they are given by two orthogonal field quadrature components that
come into play when we pass from the complex representation of the electric field strength
(1) to the real one, i.e. when we write in place of Eqn. (1)

JW, f) = xc,@,
t) +pc,(r, t), (9)
94 U. Leonhardt and H. Paul

where cl@, t) and c2(r, t) are the real and imaginary parts of c*(r, t) apart from a factor of
21’29
q(r, t) = 2”’ Re{c*(r, t)>, c2(r, t) = 2’12Im{c*(r, t)}. (10)

Similarly x and p correspond to the real and imaginary parts of the complex amplitude A

x=2”‘ReA =2-“‘(A*+A),

p = 2’/2 Im A = 2-“‘i(A * - A). (11)

In order to elucidate the physical meaning of x and p, let us consider a running plane wave
with circular frequency o and wave vector k. Then the electric field strength takes the form

E(r, t) = E,[x cos(ot - kr) + p sin(ot - kr)]. (12)

One learns from Eqn. (12) that, in comparison with a reference field oscillating as
cos(ot - kr), x represents the in-phase field component and p the out-of-phase component.
Translating Eqn (11) into quantum mechanics, we obtain

_i! = 2-“‘(dt + &), fi = 2-1/2i(it _ a).


(13)

Inverting those relations, we can also express the photon creation and annihilation operators
through C?and p in the form
ci = 2-“‘(X? + i$), 6’ = 2-‘12(f _ ifi). (14)

It follows immediately from Eqn. (13) that the operators 2 and fi, in fact, obey the canonical
commutation rule
[i?,$1 = il. (15)

We would like to mention that, the definition (11) of the two quadrature components is not
free from arbitrariness. This is easily seen from Eqn. (12). We may choose, with equal right,
the reference field to be phase-shifted by an arbitrary value 0, thus writing, in place of Eqn.
(12),
E(r, t) = E,,[x@ cos(wt - kr - 0) + ps sin(ot - kr - O)]. (16)

The so defined quadrature components xg and pe are then related to A and A* through the
relations
xs = 2-“‘[A * exp(i@) + A exp( - iO)],

ps = 2-“‘i[A * exp(i@) - A exp( -iO)]. (17)

Alternatively, we may express x@andp, through the original quadrature components x(=x,,)
and p( rp,) in the form
xQ = cos Ox + sin Op,

ps = -sin Ox + cos Op. (18)

Obviously, this transformation describes a rotation by an angle 0 in the phase space


(x,p-plane). It follows from Eqn. (18) that pB = x@+.,~. Hence x@ comprises the whole
manifold of quadrature components, when we allow 0 to vary from 0 to II. Passing to
Measuring the quantum state of light 95

quantum mechanics, Eqns (17) and (18) remain valid as operator relations. Hence we can
write
1
x@= 2-‘12[(i+exp(i@) + ci exp( -iO)],

be = 22”‘i[B+ exp(i@) - ~5exp( -iO)] (19)

or, utilising the phase-shifting operator (6),

(20)

(see Eqn. (7)) and


& = cos O$ + sin @$,

$e = -sin 02 + cos OF. (21)

In this context, we would like to emphasise that for different parameters 0, the observables
&, correspond to physically d@rent measurements. It is one of the exciting features of
quantum optics that it provides us with the capability to actually measure a whole set of
quadrature components xe which in principle can be infinitely large. This is achieved
by homodyne detection of the field. So in quantum optics a wealth of information can be
gained experimentally in a simple way, in contrast to quantum mechanics where it is rather
difficult to measure an observable that is defined as a linear combination of position and
momentum.(62)
Let us now return to the question how to describe a single-mode field by a Schrbdinger
wave function. Similarly to quantum mechanics we will introduce the latter as a complex
function of the variable x that corresponds to the operator i. Now, what makes such a wave
function especially attractive? There are, in fact, two reasons. First, relevant physical states
such as Glauber states, the analogues of classical field states, and squeezed states, the most
prominent representatives of non-classical field states, are distinguished by extraordinarily
simple Schrodinger representations: Their wave functions are just Gaussians

e,(x)
[
= (s/7c)‘14exp -f (x - 2”25)z + i2”2rlx
1. (22)

Here, 5 and rl characterise the displacement, and s is the squeezing parameter. The wave
function (22) for squeezed states includes that of a Glauber state Ia) as a special case
corresponding to s = 1 and CI= 4’+ iv. For r = rl = 0, Eqn. (22) describes the so-called
squeezed vacuum (which has energetically nothing to do with vacuum!). Second, the
observables 2 and b-and, more generally, their superpositions (21ban actually be
measured very precisely with the help of the optical homodyne detection technique (see
Section 3).
Now, it is well known that a wave function is characteristic of a pure state, while the most
general state is a statistical mixture. The latter can only be described by a density operator
6. In the photon number representation, for instance, 6 is given by a matrix (nldln’). but
we may choose also the x representation (x 16Ix’}. For the special case of a pure state, the
density matrix (x]&j]x’), which is a complex function of x and x’, factorises into the wave
function and its complex conjugate

(xldlx’> = II/(x)$*(x’). (23)


96 U. Leonhardt and H. Paul

2.3. Wigner function


A different kind of state description that will prove to be especially advantageous in
quantum optics, is based on a closer analogy to classical statistical mechanics. In fact, a
striking difference between classical and quantum statistics is that in the latter a (usually
complex) function of just one variable x corresponding to position, the Schrddinger wave
function, comprises the whole (statistical) information one can have about a system, whereas
in classical statistical mechanics a joint distribution function for both position x and
momentump is needed to fully characterise the system. The root of this discrepancy lies, at the
very heart of quantum mechanics: Since one cannot measure x and p simultaneously because
of Heisenberg’s uncertainty principle, a distribution function for both variables seems to
make no sense in quantum mechanics. However, this is not the last word!
In fact, as early as 1932 E. P. Wigner suggested a representation of a quantum state by
means of a joint probability distribution for coordinates and momenta, thus establishing a
formal analogy to the classical description. (33*76*85)
This distribution, now known as the Wigner
distribution or the Wigner function, was introduced to study quantum corrections to classical
statistical mechanics and has found many applications also in quantum chemistry and
quantum optics. The main properties of the Wigner function W(x, p) are as follows. (In view
of later application to quantum optics, we consider a one-dimensional system only and
replace A by unity.)
(i) W(x, p) is real, it may, however, take negative values in certain domains of the phase
space. (Actually, for pure states only Gaussian Wigner functions are positive
definite.(‘o)
(ii) The marginal distributions for x and p

w(x) =
s+m

-m
Wx, P ) dp (24)

and

w(P)=
s+CC

--co
W(x,P) dx, (25)

give us the correct quantum mechanical distributions for position and momentum.
(iii) The quantum mechanical expectation value for an operator function F(_-Z,p) which
is symmetrically ordered with respect to 2 and 6 agrees with the corresponding
‘classical’ average, whereby W(x,p) plays the role of the weight function

(F(Z,$))= +* +* W(x, P IV-b, P 1 dx dp. (26)


s -m s -m

Here, the classical function f is obtained from the operator function F by replacing
_%by x and J? by p.
(iv) The Wigner function is correctly normalised

ss +m

-m
+m

-m
W(x,p)dx dp=l. (27)
Measuring the quantum state of light 97

w The squared modulus of the scalar product of any two (pure) quantum states Jcp)
and I+) equals, apart from a factor of 27r, the corresponding integral in phase space

I((PIICI>I~
= 2~ +m +noW,(x,p)&(x,p) dx dp, (28)
s --co s --Do
where W, and W, are the Wigner functions for the states 1~7)and I$), respectively.
04 An immediate consequence of the relationship (28) is a normalisation property of the
squared Wigner function. In fact, putting 1~) = I$) in Eqn. (28) we obtain
+m +m
W’(x,p)dx dp =(2x)-‘. (29)
s --oo s -Cc
It should be noticed that this relation holds true only for a pure state. The
generalisation to a mixed state i.e. a statistical mixture is’76’
+m +m
W2(x, p) dx dp = (271))’ Tr(d2), (30)
s -m s -m
where @denotes the density operator.
(vii) The phase-shifting operation (6) transforms the Wigner function of a single-mode
field W(x,p) into

W(x,p;O)= W(xcosO-psin@,xsinO+pcos@). (31)

This means phase shifting (or time evolution) is equivalent to simply rotating the
Wigner function in the phase space (x, p-plane).
(viii) The recipe for constructing the Wigner function from the wave function IL(x) or,
more generally, the density operator 4 is given by

W(x,p)=?r-’ ‘mexpt2ipy)~(x-~)Jt*(.r +y)dy (32)


s -a,

or
+m
W(x,p) = n-’ ev(2ipyKx -Y IdIx + Y > dy, (33)
s -m

respectively.
Finally, Eqn. (33) can be easily inverted to obtain the density matrix in terms of the
Wigner function

(XIB Ix’> = +03exp[i(x -x’)p] W&x + x’),pJ dp. (34)


s -cc
Apart from its similarities to classical statistical mechanics, the formalism based on the
Wigner function differs distinctly from a true classical description. Firstly, since it normally
takes on negative values in certain domains, the Wigner function cannot be interpreted as
a classical distribution function which is non-negative by necessity. Secondly, it has an even
more severe defect, compared to a classical distribution. In fact, the latter can be used to
calculate not only the average of an observable B, but also all of its moments (the averages
of the powers of B). The same, however, does not hold true for the Wigner function (apart
from the special case that B is a linear combination of x and p). This is a consequence of
98 U. Leonhardt and H. Paul

the fact that the powers of B are not symmetrically ordered when B is, so that Eqn. (26) does
not apply to the powers of B. Nevertheless, the Wigner function has proved to be a convenient
and powerful tool for quantum mechanical calculations. (Later on we will present examples
in the field of quantum optics.) We would like to emphasise that it contains, the full
information on the quantum state. This is easily seen from Eqn. (34) which reveals that the
density operator in x representation can be obtained from the Wigner function by performing
a Fourier transformation. Certainly, the Wigner function is an ingenious construct which,
apparently, “in itself is without physical meaning.““@ Contrary to this widespread opinion,
we will show, however, that actually contact can be made with feasible experiments to
determine it. In particular, the squared Wigner function of a pure quantum state for a
single-mode field will turn out to be directly measurable.
2.3.1. Examples of Wignerfunctions in quantum optics. The Gaussian character of the wave
function for squeezed states (see Eqn. (22)) is preserved by the Fourier transformation in the
expression Eqn. (32) for the Wigner function which is readily found to be

WS,(x,p)=iexp[-s(x -2’125)2-s-‘(p -2”*rj)*]. (35)

Specialising to s = 1, we obtain the Wigner function for a Glauber state, and putting,
moreover, < = r] = 0 we arrive at the Wigner function for the vacuum state. For a Fock state
In) the Wigner function takes the more complicated forrn(*‘)

wFock(x,P)=(-l) - ev[-(x2 +P*W,P(~*+P*I,


71

where L, denotes the n-th Laguerre polynomial. Evidently, this Wigner function takes on
negative values in certain domains.

2.4. Quasiprobability distributions


The concept of quasiprobability distributions is, in fact, more general that that of the
Wigner function. In particular, in quantum optics two other distribution functions have
become known, Glauber’s P function and the Q function. Similar to Eqn. (26) they allow
us to express quantum mechanical expectation values as classical averages, with the respective
distribution function as the weight function. An important requirement for such a relation-
ship to hold is a definite type of ordering of the operators in the quantum mechanical operator
function. The ordering rules are readily formulated in terms of the photon creation and
annihilation operators, tit and ci, respectively. We speak of a normally ordered product of m
creation and n annihilation operators when all the annihilation operators stand to the right
of the creation operators. The opposite holds true for antinormal ordering and symmetric
ordering, as required in case of the Wigner function, means that the above-mentioned product
has to be written as the average over all possible differently ordered’products, e.g.

{B+*B}symm= f(Li+*Ci+ tit&+ + rid+*}. (37)

Then the following generalisation of the relationship (26) holds true

(F(b+, ii)) =
ss
+cc

-m
i-m

-zz
W(x,PY(x, P) dx dp,

where W(x, p) stands for any of the three quasiprobability distributions under consideration,
(38)

and the following ordering rules have to be obeyed: Normal ordering has to be associated
Measuring the quantum state of light 99

with the P function, antinormal ordering with the Q function, and symmetric ordering with
the Wigner function. Given the operator function F(ri+, Li) in the required order, the
corresponding classical function f(x,p) is obtained from it by the following replacements

bdcr - 2-“z(x + ip), &+,a * s 2-‘!2(_x - ip), (39)

in accordance with Eqn. (13).


Let us now compare the P function and the Q function with the Wigner function. Like
the latter, the P function may take on also negative values. Whereas the Wigner distribution
is always a well-behaved function, the P function, on the contrary, may become singular. For
some states it is, in fact, so strongly singular that it is not even a tempered distribution. So
the usability of the P function for the quantum mechanical description of light is, in fact,
restricted. The Q function, on the contrary, is well-behaved and moreover distinguished by
the property that it is non-negative. This follows immediately from its definition as the
squared modulus of the projection of the state II(/) on a Glauber state ]z),

Q(x,P> = (2~)-‘I(Ic/lol)12, c( = z-“~(x + ip) (40)

which generalises in case of a statistical mixture represented by a density operator 4, to

Q(x,p) = (2rc-‘(albla), a = 2-“2(x + ip). (41)

Hence there is no formal obstacle that would prevent the interpretation of the Q function
as a probability distribution function. We will, in fact, specify in Section 4 feasible
experimental schemes that allow us to observe it directly.
Actually, there exists a deeper relationship between the three quasiprobability distributions
in question which makes understandable the differences in their behaviour: Both the Wigner
function and the Q function can be interpreted as smoothed P functions, whereby the Q
function is more strongly smoothed than the Wigner function, and hence, can be considered
also as a smoothed Wigner function. In this way, any singularities that the P function may
have are made to disappear when passing to the Wigner function so that the latter becomes
well-behaved, and similarly any negative values the Wigner function may take disappear.
when passing to the Q function. In explicit terms, the smoothing process is described as a
convolution with a Gaussian, and the quantitative relationship between the different
distribution functions can be written in the compact form””

W(x,p;s)=n-‘(t -s)-’
ss +4

-0c
+m

-cc
we’, p’; t>

x exp( -(t - s))‘[(x - x’)~ + 0, -p’)‘]} dx’ dp’ (42)

with s -Ct.
Here a parameter s in W(x,p; s> has been introduced to allow us to distinguish the three
distribution functions in the following way: when s = + 1, 0, - 1 W(x;p; s) corresponds to
Glauber’s P function, to the Wigner function, and to the Q function, respectively. Putting,
for instance, s = - 1 and t = 0 in Eqn. (42), we get the representation of the Q function as
a smoothed Wigner function

Q(x,p)=n-’
ss +=
--co -cc
+= W(x',p')exp{ -[(x -x’)‘+ (p -p’)‘]} dx’ dp’. (43)
100 U. Leonhardt and H. Paul

Actually, Cahill and Glauber(‘5) generalised the concept of quasiprobability distributions even
further by introducing so-called s-parametrised quasiprobability distributions W(x, p ; s) with
s being a continuous variable (that may even be chosen complex). These generalised
distributions, in particular, interpolate between Glauber’s P function (s = + l), the Wigner
function (s = 0) and the Q function (s = - 1).
With respect to later applications, we are especially interested in the case t = 0, i.e. in
quasiprobabilities that can be interpreted as smoothed Wigner functions

f’w +m
W(x,p;s) = -i W(x’, p’)exp f [(x -x’)’ + (p - P’)~] dx’ dp’(s < 0). (44)
ns s _a, s --

It is interesting to note that Cahill and Glaubero4) succeeded even in generalising the
ordering concept, defining a general s-ordering of creation and annihilation operators such
that a relationship of the kind (38) applies also to the generalised quasiprobability
distributions.
While Cahill and Glauber probably did not think of any physical applications of
generalised distribution functions, we will, in fact, establish a connection with feasible
experiments in quantum optics. In Section 5 we will show that the effect of low-efficiency
detectors in optical homodyne measurements can be correctly described by appropriately
smoothing the distribution function that would be measured with the help of unit-efficiency
detectors. So what is actually measured in experiments designed to measure the Q function
is a certain s-parametrised quasiprobability distribution with s < - 1, and, similarly, what is
actually reconstructed in Optical Homodyne Tomography is not the Wigner function but a
smoothed Wigner function W(x,p; s) with s < 0. In both cases, the value of the parameter
s is related to the detector efficiency in a simple way.
Finally, we should like to address the question as to whether all generalised quasiproba-
bility distributions are physically equivalent, i.e. contain the full information on the quantum
state.c3@The answer is certainly affirmative when the s parameter lies between 1 and 0, since
then the relationship (42) enables us to calculate from W(x,p; t), 1 > t 2 0, the Wigner
function W(x,p; 0) which, in fact, has the desired property, as has been mentioned in the
preceding section (see Eqn. (34)). For negative values of the s parameter, the answer depends
on whether the distribution W(x,p; s), s c 0, is known as a mathematical function or as a
result of measurements. In the first case one can, in fact, find the Wigner function by
deconvolution. In the second case, however, such a procedure is not feasible since it would
require the Fourier transform of the smoothed distribution function

W,rl;s)=
ss
+m

-m --co
+m
TX, P ; s)ev[ - i(5x + ~~11dx dp

to be known with virtually unlimited accuracy especially in the high-frequency domain,


(45)

5’ + q2+ co. This can be seen from Fourier transforming Eqn. (42) which gives us, according
to the convolution theorem, for t = 0 and s < 0

@(5, ‘I; s) = *(L rl; 0)exp[s(S2 + q2)/4]. (46)

From this relationship it becomes evident that the Fourier transform of the Wigner function
m(1(5,V; 0) we are looking for, is obtained from the Fourier transform of the smoothed
distribution @‘(<,V; s) (s < 0) by multiplication with the increasing Gaussian
exp( -s(t2 + q2)/4) (note that s is negative). In this way uncertainties in the measurement will
become drastically enhanced for t2 + q 2-co. Hence Eqn. (46) gives us reliable values for
@(5, V; 0) in a restricted frequency range only, say r2 + v2 < R2. Fourier transforming this
Measuring the quantum state of light 101

truncated function yields a coarse-grained Wigner function rather than the true one.(45)Hence
we conclude that in experimentally determined generalised distribution functions with s < 0
information on the fine details of the quantum state is irretrievably lost. The physical reason
for this defect will be found in Section 5 with the study of undesired noise in the measuring
apparatus. In particular, it will turn out that any scheme that allows us to measure the Q
function (corresponding to s = - 1) is inherently noisy, even when unit-efficiency detectors
are employed.

3. EXPERIMENTAL TOOLS

3.1. Photodetector
Any direct experimental investigation of light is based on photodetection. Moreover, high
efficiency photodetectors are required for measuring subtle quantum features. When a single
mode is excited a perfect photodetector should measure the intensity

j = 6’6. (47)

(Here we consider the intensity in units of photon numbers.) However, perfect photon
counting cannot be achieved in practice. The present state of the art is an efficiency of 76%
in the single-photon counting regimeC3’)and about 85% for intense photocurrents.“‘) Any
other losses such as absorption degrade the signal additionally. To understand the influence
of inefficiencies we have to resort to the theory of photodetection, see for instance
Refs (35,5 1,63,65,66,74). It turns out ~3~)that there is a concise way of accounting for these
inefficiencies and losses in photodetection. In Eqn. (47) the annihilation operator ci of the
signal field has to be replaced by the quantity

lY=J;Ici+JiT$-. (48)

Here n (with 0 < q < 1) comprises the overall detection efficiency including all kinds of losses.
The operator 6 represents the annihilation operator of a fictitious bosonic mode being in the
vacuum state. We obtain from Eqn. (48) that the mean intensity

(ci’+ti’) = tj ($ri) (49)

is reduced by a factor of I], as expected. On average q photons are counted and 1 - 4 photons
are ignored. However, since a single photon is either counted or ignored as a whole, the
outcome of a single measurement is unpredictable. We thus have an additional source of
irregularity expressed by the fictitious vacuum-noise field which is represented by b’ in Eqn.
(48). Formally, 6 must be introduced in order to conserve the bosonic commutation rules of
the detected field. In fact, from

[ii, f?+]= [6,0] = 1 (50)

we also obtain

[ci’, a’t] = 1. (51)

The additional vacuum mode plays the role of Langevin noise in damping according to the
dissipation-fluctuation theorem. This extra noise reduces our possible knowledge about the
quantum state of the signal field.
102 U. Leonhardtand H. Paul

3.2. Beam splitter


There is another intuitive picture for formula (48) describing losses in photodetection. The
ignored photons may be regarded as being scattered away by a fictitious beam splitter and
absorbed (see Fig. 1). In the quantum regime, a beam splitter combines always two ingoing
beams to produce two emerging fields. Even when one beam is split into two, it interferes
with the vacuum fluctuations entering the unused second input port of the beam splitter. In
inefficient photodetection modelled by a fictitious beam splitter in front of an ideal detector,
the additional detection-noise mode is modelled as the vacuum noise entering the unused port.
Supplementing formula (48) by a fictitious second output mode 6’ we may write

=(-_A T)(4)-
(%:> (52)

Again, Eqn. (52) has been constructed such as to conserve the bosonic commutation rules
for the annihilation operators 6’ and 6’. One may ask, what is the most general description
of a lossless beam splitter? We need to suppose only that two incoming modes interfere with
each other to yield two emerging fields (see Fig. 2):

(53)

Equation (53) describes a two-mode interferometer, a polariser or a linear coupler as well.


As a consequence of the commutation rules

pi, cii’]= [cii,(i;+] = 6,l (54)

we find the requirements

I& I’+ I&l2 = 14, I2+ IB,,I’ = 1, B;F,& + WA, = 0. (55)

second output
\ s;
/
second input
vacuum 62

detector I
(yt; first output
i(;
FIG. 1. Effective model for an inefficient photodetector.
A fictitious beam splitter is placed in front of an ideal FIG. 2. Lossless beam splitter. Two input fields interfere
photodetector. The transmittance of this beam splitter linearly with each other to yield two emerging output
equals the detection efficiency. fields.
Measuring the quantum state of light 103

Thus, the B-matrix has to be unitary. Unitary 2-dimensional matrices have a simple structure:

(56)

We may redefine the phases of the incoming and outgoing fields

In this way we reduce our formula (53) to a rotation of the mode operators
^, cos c? -sin@ b,
0 (
aI
ri;
=
sin u cos X >( Liz) .
(58)

Mixing of two modes by a rotation is the basic operation of a lossless beam splitter as well
as a two-mode interferometer, a polariser or a linear coupler. Setting cos c1= r] and
-sinu = JT 1 r] we obtain Eqn. (52) describing losses in photodetection. The transmittance
t of the beam splitter is given by cos* cx while the reflectance r equals sin* LX.The identity
cos’a + sin’s = 1 guarantees energy conservation. For the special case of a balanced (or
50 : 50) beam splitter we have

(59)

Equation (58) describes the input/output relation for a beam splitter as a transformation of
operators. This corresponds to the Heisenberg picture in quantum theory. What would we
get in the Schrijdinger picture? How is the quantum state transformed when the operators
remain fixed? The answer is quite simple. (4’1The total Schrddinger wave function of both
incident modes $(x, , x2) or, in the momentum representation, $(p,, p2) is rotated as well,
but in the opposite direction as the mode operators in the Heisenberg picture

II//(X’,Xi) = @(xi cos u + xi sin a, -xi sin x + xi cos cz),

~‘(P;,P;)=~(P;cosrx+p;sincr, -p;sincc+p;coscl). (60)

The arguments of the Wigner function are also rotated

W’(x;,x;,p;,p;)= W(x;coscc +x;sincc, -x;sina +x;cosa,

p;coscr +p;sinz, -pisin@ +p;cosa). (61)

In this way, beam splitting is easily described. Other methods can be found for instance in
Ref. (16).

3.3. Homodyne detector


Our basic detection instrument is the balanced homodyne detectorC9” (see Fig. 3). It consists
of a 50: 50 beam splitter, two photodetectors, a reference beam having a well-defined phase
with respect to the signal field, and some electronics. The signal and the reference beam (also
called the ‘local oscillator’) are optically mixed at the beam splitter. The emerging beams are
POE 19,2-B
104 U. Leonhardt and H. Paul

detector

detector

local
oscillator
(aLo)

FIG. 3. Balanced homodyne detector. The signal is optically mixed with a strong coherent local
oscillator using a 50: 50 beam splitter. The emerging fields are detected and the photocurrents are
electronically subtracted to yield the measured quantity.

detected and the measured photocurrents are electronically subtracted to yield the measured
quantity

Al = G;+ci;- ci;+ci;= ;(s+ + citc)(ci + a,) - :(Lj+- &)(b - 6,) = ri+B,, +&lo. (62)

When the local oscillator is coherent and intense with respect to the signal we may describe
it classically. We simply substitute the annihilation operator (iLo by the complex amplitude
aL0 :
1
h_O-+aLo = bLo le’* (63)

and obtain for the photocurrent difference

A3 = [aLoI(ri ebe + tit eie). (64)

(A more refined quantum-statistical theory of homodyne detection can be found for instance
in Refs (12,81). However, our simple formula (64) remains correct for a local oscillator in
a highly excited coherent state.) Hence the balanced homodyne detector measures a
quadrature component
Al = (aLoI&. ie (65)

i.e. any linear combination of position 2 E &, and momentum $ = f,,, corresponding to a
rotation Z(Q) = f cos 0 +@ sin 0 (see Eqns (19) and (21)). The rotation angle is defined by
the local oscillator phase 0 (or, to be more precise, by the phase difference between local
oscillator and signal). Shifting this phase varies the mixing angle between position and
momentum. Thus homodyne detection drastically enlarges our measuring capabilities in a
simple way. In the case of quantum-mechancial matter waves it is much more difficult to
measure combinations of position and momentum.@2’

3.4. Realistic homodyne detector


As already mentioned, photodetection is never perfectly efficient in practice. In order to
understand the effect of inefficiencies on homodyne detection we use our simple model for
Measuring the quantum state of light 105

detector

vacuum

signal
/
B
detector
/\
V
local
oscillator
(aLo)

w
detector

f
vacuum
I ii:’

signal
/
c &’
detector
/\
V
local
I oscillator
(aLo)

FIG. 4. (a) Realistic homodyne detector. Fictitious beam splitters are placed in front of the
photodetectors to account for losses in detection efficiency (Fig. 3 combined with Fig. 1). (b)
Equivalent model for a realistic homodyne detector. One effective beam splitter in front of an ideal
homodyne detector represents losses in overall detection efficiency. It also comprises losses due to
mode mismatch.

inefficient direct photodetection. We imagine fictitious beam splitters to be placed in front


of the two detectors in the measurement set-up (see Fig. 4a). For the annihilation operators
of the detected fields we find according to Eqn. (48)
106 U. Leonhardt and H. Paul

Hence we obtain for the measured photocurrent difference

Aj= ri;+&’ - ci;+ri; = (Jtlri;+ + fi6l)(J;;ci; + ,/i-=&)

= ,/(ri;+ci;- li;tci;)+J;ro(ci;+b, + Gfci;+ a;+6,- 6;s;)

+(l - 49(6f6*- 616,). (67)

The annihilation operators 6; and ri; describe the fields emerging from the 50 : 50 beam splitter
where the signal is optically mixed with the local oscillator. Hence ci; and d; are mixtures
of the annihilation operators for the signal beam B and the local oscillator ci,:

The local oscillator has to be a strong coherent field. Hence we may replace the annihilation
operator &e by the complex amplitude aLo. We need to consider only the leading terms in
Eqn. (67) with respect to aLo, and obtain

Aj = q(a$ii + aLoci+)
+ ,/i?=8 (69)

We define the annihilation operator

6 = 62- 6, (70)
-yiT
It corresponds to the optical mixing of the fictitious vacuum-noise modes 6, and 62.Since
the mixing of vacuum with vacuum yields vacuum, the fluctuation mode 6 can be regarded
as a bosonic mode being in the vacuum state as well. Hence the measured photocurrent
difference takes the form

Aj = &&,(J;;c~ + J--J) + H.c. (71)

This formula has a simple interpretation: Similar to direct photon counting (see Eqn. (48)
or Fig. 1) we also find here in homodyne detection that a fictitious vacuum field has to be
added to the intensity-reduced signal. This means we may replace the arrangement of two
fictitious beam splitters in front of the photodetectors (see Fig. 4a) by just one efictive beam
splitter in front of an ideal homodyne detector (see Fig. 4b). This effective beam splitter can
also be used to account for other kind of losses, such as those due to mode mismatch. The
measured quadrature distribution is therefore the quadrature distribution for a mode which
is attenuated to some extent. To be more quantitative, we utilise the fact that the marginals
of the Wigner function yield the quadrature distributions, e.g.

%(Xe) =
sss
+m

-m
+m

-m
+a,

-m
W’(x, , pe ,R,, le 1dpedfe dh . (72)
Measuring the quantum state of light 107

Here W’ denotes the Wigner function after beam splitting. According to our beam-splitter
theory, Eqn. (61), W’ is given by

with W(x@,p,) and W,,,(.f,,p,) being the Wigner function of the incident signal and the
Wigner function of the vacuum field entering the unused port of the fictitious beam splitter,
respectively,

(74)

according to Eqn. (35) for s = 1, 5 = q = 0. Performing the substitutions

and introducing the quadrature distribution for the original field

wf(x;) =
s +m

-cc
dp;,W(xia>~L) (76)

we find the measured quadrature distribution to be given by finally

exp[ -$---(xb - sy]dx;. (77)

Hence, a realistic homodyne detector measures an appropriately smoothed quadrature


distribution. We remark that a quantum-statistical theory of inefficient homodyne detec-
tion@‘) yields the same result.

3.5. Amplzjier
We will see later that amplifiers also provide useful tools for measuring the quantum state
of light. A linear single-mode amplifier magnifies the complex amplitude of the mode at a
rate proportional to the amplitude

%_?$2(6). (78)

The solution of Eqn. (78) is

(f?(t)) = e(Yiz)r(ci). (79)

Now we translate the solution (79) for the average amplitude (i(t)) into at, equation for the
annihilation operator i(t). We have to be aware that an additional Langevin-type noise
operator t is needed in order to maintain the canonic commutation rules for a(t)

If (t ) = &/z)ri + i.
108 U. Leonhardt and H. Paul

Conservation of the commutation rules requires that t is expressed as

031)
with
[S, 6’1 = 1. (82)

Hence 6 may be regarded as a bosonic mode operator. It represents the amplz$cation noise,
for instance the spontaneous-emission noise. For phase-insensitive amplification this noise is
minimized when the &mode is in the vacuum state. (17)For phase-sensitive amplification the
amplification noise may be reduced below the vacuum level for one quadrature component
and enhanced for the other using a squeezed or rigged noise reseNoir.(‘9-2’.32.48.78)
The simplest
realisation of a linear amplifier is a parametric amplifier. Here the ri mode corresponds to
the signal beam ci, and the 6 mode to the idler ci,. The amplifier performs a transformation
of the operators(4g*m)
69; coshg sinhg 6,
= (83)
0 a; ( sinhg coshg >( fi2) ’

where g denotes the product of the effective coupling constant and the interaction time.
Setting
cash* g = ey’ (84)

we see that, in fact, a parametric amplifier fits into our linear-amplifier model. On the other
hand it has been shown(43) that any linear amplification caused by Gaussian reservoirs is
correky described using the parametric-amplifier model (83). Similar to inefficient photo-
detection, a linear amplifier can be modelled by a fictitious parametric amplifier, where the
idler mode represents the amplification noise. Using a real parametric amplifier, however, has
the advantage that the idler (the ‘noise mode’) can be easily manipulatedo6) and also observed
after amplification in contrast to the fictitious ‘amplification-noise mode’.
Equation (83) describes the transformation of the operators 6, and d2 in the Heisenberg
picture. What would we get in the Schiidinger picture? It is easy to showo4,43)that the total
wave function of both modes $(x1, x2) or, in the momentum representation, &(p, ,p2) is
Lorentz transformed

$‘(x;,x;)=~(x;coshg-xX;sinhg, -x;sinhg+x;coshg), (85)

$‘(p;,p;)= &(P;coshg +p$sinhg,p;sinhg +p;coshg). (86)

Note that the momentum wave function is oppositely transformed as the position wave
function in order to avoid a violation of the uncertainty relation between position and
momentum. (A Lorentz transformation in the x,, x,-plane performs a contraction in one
diagonal direction and an expansion in the other. (43)This is compensated for by the opposite
transformation in the p, , p,-plane.) The Wigner function is transformed accordingly

W’(x;,x;,p;,p;)= W(x;coshg -x;sinhg, -x;sinhg +x;coshg,

pi coshg +p; sinhg,p; sinhg +p;coshg). (87)

So far we have studied linear amplification or non-degenerate parametric amplification.


There is, however, another possibility for conserving the commutation rules after amplication:
Measuring the quantum state of light 109

degenerate parametric amplification. Then the ‘fluctuation mode’ 6 and the signal mode B are
identical, and no additional noise mode is required. In this case we obtain

ci’ = cash gd + sinh gcit (88)

or decomposing ci and h’ in terms of position and momentum quadratures (see Eqn. (13)) we
get
2’ = @$, p’ = e-gp. (89)

Hence the momentum wave function is squeezed

$‘(p’) = eq(e”p’) (90)

and the position wave function is anti-squeezed

*‘(XI) = e-g/*$f(e--g,yr), (91)

provided g is positive. Otherwise the position wave function is squeezed and the momentum
wave function is anti-squeezed. The prefactors maintain the normalisations of the wave
functions. The Wigner function is transformed accordingly:

W’(x’,p’) = W(eegx’, epp’). (92)

A degenerate parametric amplifier squeezes one quadrature and anti-squeezes the other. (The
squeezing effect has been observed experimentally since 1985.(69))

4. MEASUREMENT SCHEMES

4.1. Phase measurement and Q function


One of the most delicate problems in quantum optics is the proper description of the phase
of a single-mode radiation field. There were numberous attempts to treat phase on the same
footing as usual observables, namely to construct a corresponding Hermitian operator.
However, up to the present no fully satisfactory solution could be found. It was only by
resorting to an infinite sequence of finite-dimensional Hilbert spaces that a strictly Hermitian
phase operator could be defined. (7,59*6o)
From a practical point of view, it will be felt even more
unsatisfactory that no experimental scheme for an ‘ideal’ phase measurement could be
devised. Actually, both the theoretical and the experimental difficulties have a common root:
Since an electromagnetic field couples to matter via the electric field strength which comprises
both (real) amplitude and phase, the phase cannot be measured directly, on principle.
In view of this unsatisfactory situation, one might be inclined to turn the tables, namely
to first devise a feasible experimental scheme for measuring phase properties, thus giving an
operational definition of phase, and afterwards search for the proper quantum-mechanical
description of the experimental procedure in question. The first to carry out such a program
successfully were Bandilla and Paul. (5)Their basic idea was to amplify, with the help of a
quantum amplifier (laser or optical parametric amplifier), the microscopic field under study
to a macroscopic level, where classical phase measurement techniques could readily be
applied. It is evident that such a measurement cannot be an ‘ideal’ one, since the amplifier
necessarily introduces noise. Nevertheless, it will yield valuable, even if not perfectly accurate,
information on phase properties. What one will intuitively expect is that the phase
distribution obtained from many repeated measurements will be somewhat broader than the
‘true’ one, as a result of the noisy charcter of the measurement.
110 U. Leonhardt and H. Paul

A reduced measurement accuracy is, in fact, a common feature of realistic phase


measurement schemes presently known. Besides the above mentioned amplification scheme,
these are a proposal by Shapiro and Wagner (67)based on heterodyne detection of the signal
field, and an operational approach analysed theoretically and studied experimentally by Noh,
Fougeres and Mande1.(53-55’The latter scheme has been also investigated earlier by Walker
and Carroll.@2-84)It is, in fact, very instructive. On the one hand, it has a sound physical basis
in classical optics, and on the other hand, it clearly shows up the specific difficulties due to
the quantum nature of the radiation field. Hence we will focus our attention on this scheme.
However, we simplify the measurement by employing a strong local oscillator in the balanced
homodyne detection scheme.
Already in classical optics we are confronted with the problem that the phase cp of a
quasimonochromatic field cannot be determined from any single measurement, since the
latter yields combined information on both phase and amplitude. Hence at least two different
measurements on the same field are needed to determine the phase and the amplitude
separately. A simple way to perform those measurements is suggested by decomposing the
electric field strength E = E. cos(wt - cp- kr) into two parts that are in phase or out of phase
compared to a reference field, according to Eqn. (12). From this equation it is readily inferred
that cos cp and sin cp can be expressed through the quadrature components x and p as

(93)

(see also Eqn. (11)).


Putting it more simply, the phase angle cp is just the polar angle in the x,p-plane (phase
space). The formulas (93) suggest the following scheme for phase measurement: measure
simultaneously x and p and calculate cos cp and sin cp with their help. In practice, one will
use a SO:50 beam splitter and measure x and p separately on the two outgoing beams, see
Fig. 5a.
Let us now discuss such a measuring scheme from the quantum-mechanical point of view.
Actually, we have no difficulties in performing the required measurements, since a very
efficient technique for observing x and p is provided by balanced homodyne detection (see
Fig. Sb), as has been pointed out in Section 3. What is different, however, between the classical
and the quantum pictures, is the effect of the beam splitter. While this device is a perfectly
harmless optical element in the classical theory in that it divides the field strength of the
incident beam into two identical parts, in the quantum theory it actually disturbs the outgoing
fields by adding noise.
Let us describe the action of a 50: 50 beam splitter in some detail. It is advantageous to
represent the state of the incident field by a Wigner function IV,(x, ,p,). According to Section
3.2 we have to take into account the fact that the vacuum enters the unused port of the beam
splitter as an input field. Hence the Wigner function for the ingoing fields is given by

Wl(xI,~,)Wvac(x2,~2)= Wl(xlypl)~ exp[-bi+p31, (94)

where we have introduced the Wigner function for the vacuum state (Eqn. (35) for
s = 1, c = q = 0) on the right-hand side. According to Section 3.2 we get the Wigner function
for the outgoing fields WoUl(x;, x;;p;,p;) by substituting in Eqn. (94)
Measuring the quantum state of light Ill

(4 signal

4
urement
ofP;

measurement
of 2;

signal pi measurement

I L7

-I pplate

c3
z; measurement local oscillator

FIG. 5. (a) Beam-splitting scheme. The initial signal field is split into two parts using a 50: 50 beam
splitter. The quadrature x; is measured on one emerging beam and the quadrature pi is measured
on the other. (b) Eight-port homodyne detector for realising the beam-splitting scheme depicted
in (a). Both the signal and the local oscillator are split and the emerging beams are distributed to
two homodyne detectors. The I./4 plate performs a n/2 phase shift in order to measure pi on one
homodyne detector while xi is measured on the other.
112 U. Leonhardt and H. Paul

Let us assume that xi and pi are measured (with the help of two separate balanced homodyne
detection apparata). Then we integrate over the remaining variables x; and pi in order to
get the probability distribution w(x; ,p;) for this simultaneous measurement, Considering the
equations for x, and p, in (95) as substitutions that allow us to integrate over x, , p, instead
of x;,p;, we find the integral in question to be given by

w(x;,P;) =

ss
4-m

-co

2 +m +m
+a3

-m
W,,,(X;,X;,p;,p;)dx;dp;

=-- W, (x,,p,) exp[-(x, - 2”*x;)*- @, - 2”*p;)*] (96)


= s -m s -m

x dx, dp, . (97)

This is a convolution of the original Wigner function with a Gaussian, and we learn from
Eqn. (43) that it equals just the appropriately scaled Q function for the incident field

w(x; ,p;) = 2Q(2”*x;, 2”2x;).

From Eqn. (98) we readily obtain a phase distribution as a marginal distribution, i.e. by
introducing polar coordinates (cf. Eqn. (38))

xi=@ coscp pi=@ sincp (99)

and integrating over the amplitude e. Actually, a phase distribution contains the full
information on the phase properties of the field under study. With its help one can, in
particular, calculate the average of any function of cp, e.g. cos cp, cos* cp, sin 40 . cos cp etc.
It is of great physical interest that the other two phase measuring schemes mentioned before
also amount to measure the Q function. @u’) Hence, the conclusion can be drawn that they
all are physically fully equivalent. In all three cases, the smoothing of the Wigner function
that leads to the Q function is due to the introduction of noise which originates from different
physical processes but has, nevertheless, similar characteristics. (In the following sections we
will discuss in some detail the amplification scheme(5~6~30~3’~57~64)
and heterodyne detection.@“)
In the present paper, we are actually not so much interested in the experimental determination
of phase distributions but in the underlying technique that allows us to directly measure one
of the quasiprobability distributions well known in quantum optics, the Q function. (Another
quasiprobability distribution, the positive P representation, can be reduced to the Q function,
and hence measured in a similar way. (‘s’~))In this context, we have seen that such an
apparently simple device as a beam splitter provides a feasible solution to the long-standing
problem of simultaneously measuring two canonically conjugate variables.(2~‘3~‘8~7~~*7~Of course,
according to Heisenberg’s uncertainty principle an accurate measurement of this kind is
impossible. Precisely speaking, we are not measuring the two conjugate variables (two
orthogonal field quadrature components), but modified variables which contain noise
contributions originating from beam splitting due to the intrusion of vacuum noise into the
unused input port of the beam splitter. In fact, according to Eqn. (59) the annihilation
operators for the signal, 6, , and the vacuum mode, ci, transform, in case of a 50: 50 beam
splitter, as
Measuring the quantum state of light 113

Passing to the corresponding relationship for quadrature components J? and p by virtue of


Eqn. (13a), we find the measurement of xi to be represented by the operator

+‘(a, -a,) (101)


Jz
and the measurement of pi by

(cf. Ref. (40)).


It is readily seen that the modified observables a; and fi; commute, in accordance with the
experimental fact that they are measured on separate beams so that the measurements do not
disturb each other. Hence the measurements of 2; and p; can be performed with arbitrarily
high precision, in principle. With respect to the original observables 2,) fi, , these are, however
imprecise measurements. Quantitatively, the loss of accuracy is expressed by Eqn. (98)
indicating that the measured distribution w(x; , pi) gives us not the Wigner function for the
original field, but a smoothed Wigner function, the Q function, in which information on finer
quantum features is lost (see Section 2.4). The reduction in measurement accuracy can, in fact,
be characterised in a simple and instructive way: The uncertainties for xi and p;, as taken
directly from the Q function, satisfy the inequality’3,75,88.~)

Ax;Ap;> 1. (103)

In comparison with Heisenberg’s uncertainty relation, the right-hand side of the inequality
is larger by a factor of two.

4.2. Determination of the Q Function via ampl@cation


Linear amplification of a signal field can be described in close analogy to beam splitting,
as pointed out in Section 3.5. Quite generally, a linear amplifier can be modeled as a
parametric amplifier with an idler mode representing the amplification noise.(43)Hence it is
sufficient to understand the action of a parametric amplifier. Let the field under investigation
be injected into a parametric amplifier as a signal, and the idler field be initially in the vacuum
state. Then, the Wigner function for the total field, at the starting point of the amplification
process, is again given by Eqn. (94). To describe amplification, the arguments xl, x2, p, , p2
have now to be substituted according to

(see Eqn. (87)). Here, the following abbreviations have been introduced

S = sinh g, C = cash g. (105)

Thus the Wigner function for the amplified signal field is given by tracing over the unobserved
(amplified) idler

Klln,,(X;,P;)=~-’
ssfao

-Cc
+‘x

-02
w,(x,,p,)exp{-(x:+~:)}dx;dp;. (106)
114 U. Leonhardt and H. Paul

Here, the arguments x, , p, , x2, p2 have to be replaced according to the relations (104). It is
advantageous to consider the equations for x1 and p, in (104) as substitutions that allow us
to integrate over x, and p, instead of xi and pi. Then Eqn. (106) can be written as

JJ
+m +m
w,,,,(x;,p;) = n:-‘F2 W,(X,9P,)
--co -03

x exp{ - C2Se2[(x, - xi/C)’ + (p, - pi/C)‘]}

x dx, dp,. (107)

According to Eqn. (44) this integral can be identified with an s-parametrised quasiprobability
distribution, where s is given by s = -C2/S2,

(108)

For strong amplification, the s-parameter tends to - 1, this means, the distribution becomes
the Q function for the original signal field. The scaling is such that the amplification of the
field is properly taken into account: The Q function is magnified.
Strong amplification, C B 1, actually brings the field to a classical level. Under these
conditions, there is practically no difference between the Wigner function and Glauber’s P
function, and the latter can be identified with a classical distribution function for the complex
field amplitude. (In a formally different approach based on studying the evolution of
Glauber’s P function due to amplification, it was shown directly that the P function for the
amplified field is given by the appropriately scaled Q function for the original field.‘“))
Therefore we arrive at the important result that the Q function of the original field can be
inferred from measurements on the strongly amplified, and hence classical, field. What one
has to do is to determine the classical distribution function and rescale it properly. The
corresponding measurements are, of course, of classical character, too. For instance, the beam
splitting scheme discussed in Section 4.1 can be applied, however, the introduction of noise
in this device has no longer a noticeable effect in view of the high field intensities, and the
same holds true for nonunit detection efficiencies (see Section 5). So amplification offers,
indeed, a feasible scheme for determining the Q function of a microscopic field.

4.3. Measuring the Q function in heterodyne detection


Shapiro and Wagner@‘) studied the scheme depicted in Fig. 6 in the context of measuring
phase and amplitude uncertainties: The signal wave at frequency o is mixed with a strong
local oscillator field at w + 6 via a very weakly reflecting mirror. (It is also possible to use
a balanced heterodyne detection scheme. @)) Both fields are directed to a photodetector and
the beat signal, which is an alternating current oscillating at the frequency 6, is electronically
extracted from the photocurrent. In classical optics, the complex amplitude of the beat signal
is proportional to the complex amplitude of the signal wave to be measured. In quantum
optics, however, an image mode at frequency o + 26 has to be considered as well. The image
mode interferes with the local oscillator and contributes to the beat signal, even when it is
in the vacuum state. The image mode plays essentially the same role as the vacuum field
entering the unused port of the beam splitter in the measurement scheme described in
Section 4.1.
Let us study the heterodyning scheme more quantitatively. We denote the annihilation
operators for the signal and the image mode by ci and 8,, respectively. We assume the local
oscillator field to be still strong enough for justifying a classical description (the replacement
Measuring the quantum state of light 115

“image”
LO 6

signal 6 quadrature demodulator


3

cos bt

local oscillator

FIG. 6. Heterodyne detector. The signal field (frequency CO)is mixed with the local oscillator
(frequency o + 6) and detected. The neat signal (frequency 6) is extracted from the measured
photocurrent using a quadrature demodulator. The image mode (frequency o + 26) is assumed to
be in a vacuum state and contributes to the beat signal as well.

of the corresponding annihilation operator B,c by the complex amplitude aLo). The
photocurrent j is proportional to the quantity

j K (d t eiwr + c1$ ei(o + 6)f + 2 1 ei(w+ 29 x (2 e - IWI + aLo e i(w + 6)r + 6, e - I(UJ + 2&r)

a(btri +ritcr,, eei6’ + citci, e-“6’+cr$2 eis’+ I~l~o(~


+ a$B, em;”
1
+ cifa,, eid’+ if&).
+Bfa er2&’ (109)

Hence, we obtain for the amplitude a of the beat signal at frequency 6

A aa$d + aLoci:. (110)

We may adjust the local-oscillator phase in such a way that

Aaii -ii:. (111)

In the classical regime, where the value of B is large compared with the tiny quantum-noise
term df, the amplitude R is proportional to the complex field amplitude CLIn the quantum
regime, the noise term CiJplays a double role. On the one hand, the image mode introduces
quantum noise, on the other hand, it guarantees that a and at commute

[A+, A]a[at -B,, ci - Lff] = 0. (112)

Hence, the real and the imaginary part of the complex amplitude ci can be measured
simultaneously. They are the real amplitudes of those parts of the signal that oscillate as
cos 6t and sin St, respectively. They can be measured using a photocurrent quadrature
demodulator.@” Similar to Eqn. (13b), we denote the real and the imaginary part of the
operator 2 by _C’and fi’, respectively. We obtain from formula (111)

i-‘ai - ZI, @‘afi + 0,. (113)


116 U. Leonhardt and H. Paul

Hence in heterodyning essentially the same observables are measured as in the beam-splitting
scheme (see Eqns (101, 102)). As in that case, we can interpret the measurement of z?’ and
8’ as a simultaneous, however noisy, measurement of the quadrature components x and p
of the original field. Thus heterodyning is an approximative simultaneous measurement
scheme for position and momentum as well. Since both schemes are completely equivalent, the
probability distribution of the x’- and p/-values is the Q function for the signal mode.

4.4. Squared Wigner function


All schemes for the direct measurement of the Q function are schemes for the simultaneous
yet noisy measurement of position and momentum. The physical origin of this noise is the
vacuum noise associated with an additional degree of freedom needed to obtain commuting
observables which approximate the original position and momentum variables. Can we reduce
the effect of this noise? We can squeeze the noise when the additional system is in a squeezed
vacuum state. Then the extra noise is below the vacuum level for one quadrature and above
the vacuum level for the conjugate quadrature. The result is that one quadrature is measured
more accurately then the other. We can also squeeze the additional noise when we change
the ratio in which it is distributed to the detectors. For instance, we can use an unbalanced
first beam splitter in the beam-splitting scheme of Fig. 5. This will transmit less vacuum noise
to one homodyne detector while reflecting respectively more vacuum noise to the other. In
this way a squeezed Q function will be measured. c4’@)However, the total amount of undesired
noise is not changed.
In the context of phase measurement Hradi1(34)mentioned that the signal field could be
mixed with itself in order to reduce the phase dispersion. When the field is mixed with
itself or, more generally and more precisely, when the second system involved is in a state
related to the state of the signal the extra quantum noise may be tailored to reduce its total
effect.
Let us assume the signal to be in a pure state described by the wave function $(x,) and
the second system in the phase-conjugated state t,b*(x2). Let us first study the beam-splitting
scheme. Both fields are optically mixed using a 50: 50 beam splitter. According to Eqn. (60)
the total wave function for the emerging beams is given by

$‘(x, ) x2) = $[2-“7X, + x2)1$ *[2-“2(-x, + x2)]. (114)

We propose to measure x2 on the beam 2 and p, on the beam 1. The probability distribution
w(x2,p,) for the measured x2 and p, values is the modulus squared of the wave function

ev( - ip,xl M ‘(xl x2)dx,.


9 (115)

Inserting Eqn. (114) into Eqn. (115) and substituting y = x, /,/2 we obtain for the probability
distribution w (x2, p, )

(116)
Measuring the quantum state of light 117

According to Section 2.3, Eqn. (32), the distribution w(+, p, ) is proportional to the modulus
squared of the Wigner function. Since the Wigner function is a real function the probability
distribution w(x~,P,) is proportional to the squared Wigner function

w(x,,p,)=7rW2 -+fi . (117)


( JZJZ )

Hence, in the beam-splitting scheme the squared Wigner function can be directly measured
as a probability distribution when the signal field being in the pure state +(x, ) interferes with
a second field being in the phase-conjugated state J/*(x2).
A similar measurement can be performed using a parametric amplifier. Again, we assume
the initial signal field to be in the state $(x, ) and the initial idler in the state $*(x2). According
to Eqn. (85) the total wave function after amplification reads

$ ‘(x, , x2) = II/(x, cash g - x2 sinh g)+ *(-x, sinh g + x2 cash g) (118)

with g being the product of the amplification constant and the interaction time. The
probability distribution w (x2, p, ) for position measurements performed on the amplified idler
and momentum measurements on the amplified signal is given by

exp(-iP,xl)~'(x,,x2)dx, (119)

We substitute y = $x,/2 - x, sinh g and obtain

W(Xz,P,)'
2rc silnh’g l[~~exp(-i&y)

x )[x,(%cothg -sinhg)-y cothg]$*(x2y+y)dyr. (120)

In the limit of strong amplification the gain G 3 cosh2 g is large and because of coth g -+ 1
we get

where sinh g -&/2-+cosh g = ,/6 has been utilised. Similar to the beam-splitting scheme the
probability distribution for a position measurement on the idler and a momentum measure-
ment on the signal is proportional to the modulus squared of the Wigner function. Since the
Wigner function is real we obtain the final result

“(x,,P,)=~ w’($j&$ (122)

In the amplification scheme as well as in the beam-splitting scheme a scaled squared Wigner
function can be directly measured as a probability distribution. However, in both schemes
a second field has to be prepared in the phase-conjugated state of the first field. This cannot
118 U. Leonhardt and H. Paul

be done using a phase-conjugating mirror since similar to optical amplification phase


conjugation is inherently noisy. cz7)Hence, our method is restricted to pure states having a real
wave function for a certain quadrature. But nevertheless we have shown that the squared
Wigner function is a directly measurable quantity for a broad class of pure states. (For more
details see Ref. (46).)
We would like to add that even the Wigner function itself can be directly measured for
a certain class of states, namely squeezed states. In fact, one learns from Eqn. (35) that in
this special case the Wigner function is nothing but the product of the x and the p distribution
(the corresponding marginals of the Wigner function). Hence, the Wigner function of a
squeezed state can indeed be determined directly by simply measuring the x distribution on
one half of the statistical ensemble and the p distribution on the other half. This result is of
actual experimental interest, since squeezed states are not only prominent representations of
nonclassical states of light from the theoretical point of view, but can also be generated with
existing techniques.

4.5. Optical homodyne tomography


Measuring the Q function does not yield the full information on the system (see Section
2.4). It is, one may say, coarse-grained information which nevertheless will be useful in
predicting gross features of the state. For instance, one will get a hint whether one is dealing
with a squeezed state, and how the uncertainty ellipse is orientated. On the other hand, a
desirable goal is to extract the full information from experimental data. Surprisingly, such
an ambitious program can actually be carried out in quantum optics. It relies on the potential
of balanced homodyne detection to measure not only two quadrature components x and p,
the analogs of position and momentum of a material harmonic oscillator, but actually an,
in principle, infinite set of variables that are linear combinations of x and p, corresponding
to a rotation by an angle 0. In Section 2.2 we introduced those variables x, and pe and in
Section 3.3 we pointed out that they can be measured directly with the help of the balanced
homodyne detection technique by varying the phase angle 0 between the signal and the
(strong) local oscillator. So one can actually determine experimentally a whole set of
distribution functions we(x,). It was shown in a fundamental paper by Vogel and Risken(79)
that one can reconstruct the Wigner function from such a set. (Bertrand and Bertrand”” used
a similar property in order to define the Wigner function.)
First, Vogel and Risken showed that we(xe) is given by the following integral over the
Wigner function

w&e) =
s +U2

-m
W(X, cos o -pe sin 0, xe sin 0 + pe cos 0 > dp, (123)

called the Radon transformation. To understand this formula, we recall that the quadrature
xe is measured for a shifted local-oscillator phase 0. Phase shifting rotates the Wigner
function according to Eqn. (31). Further, we recall that the quadrature distribution
w,(x,; 0) is given by the Wigner function averaged over the unobserved canonically
conjugate variable pe. In this way we obtain formula (123).
Second, Vogel and Risken (79)showed that this forumla can be inverted. Probably the easiest
way to derive the inversion is Fourier transformation. We denote the Fourier transformed
Wigner function (also called the characteristic function) by

W,v)=
ss+co

-m
+m

-cc
W(x, p)exp( - i{x - iqp) dx dp. (124)
Measuring the quantum state of light 119

According to Eqn. (123) we obtain for the Fourier transformed quadrature distribution

wo =
s +C=2

-m
webe )ev( - ibe 1dxe

=
ss+m

-m
+m

--a,
W(X, cos 0 -pB sin 0, x, sin 0 + pB cos 0 >

x exp( - icx,) dx, dp, .


(125)

(126)

Now we substitute simply

x =x,cos@ -p,sin@, p =x@sinO +p,coSO (127)

and obtain

%3(r)=
ss +m

-cc
+m

--gi
W(x, p)exp[ - ii (x cos 0 + p sin 0)] dx dp

or, according to the definition (124) of the characteristic function,


(128)

Ge([) = W([ cos O,c sin 0). (129)

We see that the Fourier transformed quadrature distributions fi@(i) are equivalent to the
characteristic function expressed in polar co-ordinates. Given these distributions, we get the
Wigner function by the inverse Fourier transformation

1 +co +a,
@(5, q)exp(i<x +irlp) dr d?, (130)
w(xYp)=(2x)2 s _m s _-co

or, written in polar co-ordinates,

cos 0, c sin @)jc]exp(il:(x cos 0 +p sin 0)) d@ d[. (131)

We substitute @‘(c cos 0, c sin 0) by i?@(l), use the definition (125) of G’,, and arrive at the
final result

1 +cc I +m
w,(x~)][ ]exp(ic(x cos 0 +p sin 0 - x@))dx, d@ d[. (132)
w(x,p)= (2X)2 _m ss0 _m
-s

Formula (132) is the desired inversion of the relation (123) between the Wigner function
W(x, p) and the whole set of quadrature distributions ws(xe). Note that a simpler inversion
formula named Abel inversion formula can be derived for quantum states with random
phase.c”)
Hence, if such a set of distributions has been measured by homodyne detection, the Wigner
function can be reproduced using the inversion formula (132). In practice, however, not all
of the we(xe) distributions within 0 < 0 < n can be measured, but a discrete set of them.
In this case, the Wigner function can be numerically estimated using software developed for
tomographic imaging. In fact, the Radon transform (123) and its inversion (132) relating the
Wigner function to quadrature distributions is also used in Computer Assisted Tomography
JPQE
1912-c
120 U. Leonhardt and H. Paul

2
P

cl

-2

-4
-4 -2 0 2 4

FIG. 7. Optical Homodyne Tomography. The Wigner function of a tight mode is tomographically
reconstructed from a set of measured distributions we(xe). Mathematically, a quadrature
distribution w&x@) is obtained by line integrals (dashed line) of the Wigner function orthogonal
to the x,-quadrature axis in phase space (line). This relation is numerically inverted to yield the
Wigner function.

known from medical examinations. (52)In Transmission Computer&d Tomography, for


instance, a cross-section of the human body is scanned by a thin X-ray beam whose
attenuated intensity is recorded by a detector. The ‘object’ (or the apparatus) is rotated to
yield several intensity distributions depending on the respective rotation angle. Finally, a
computer processes the data for producing an image of the ‘object’ in form of a spatial
distribution of the absorption coefficient. In our case, the role of the spatially dependent
transmission coefficient of the ‘object’ plays the Wigner function, the recorded intensity
distribution is replaced by the quadrature distribution We, and the phase 8 corresponds
to the rotation angle. The X-ray beam is replaced by a line orthogonal to the x,-quadrature
axis in phase space (see Fig. 7). Equation (123) means that the value of w,(xe) (the ‘intensity’)
is given by the integrated Wigner function (the integrated ‘transmission coefficient’) along this
line. Finally, a computer has to calculate the Wigner function using the inversion formula
(132), provided the whole set of distributions we(xe) is known. For a finite number of
distributions We, software used in Computer Assisted Tomography can be applied to
reconstruct the Wigner function (for instance, the standard filtered back projection algorithm
for parallel-beam sampling geometry(s2)).

FIG. 8. Apparatus for balanced homodyne measurement of quadrature amplitude. The crystals
are oriented at 45” with respect to the polar&r (PBS 1) axes in order to produce the squeezed field.
Prisms (not shown) in front of each detector remove the pump 532 run beam from the 1064 mn
signal beam. (Reproduced from Ref. (71))
Measuring the quantum state of light 121

In this way, first pioneering experiments have been recently performed by Raymer et
In view of the similarity to Computer
aZ.(9*‘o~7’-73~6’) Assisted Tomography they called their
method Optical Homodyne Tomograpy. The first Optical-Homodyne-Tomography exper-
iment”‘) is sketched in Fig. 8. Pulses of squeezed vacuum were generated by using a walk-off
compensated, travelling-wave optical parametric amplifier consisting of two type II phase-
matched KTiOPO, crystals.(70) This is a non-degenerate parametric amplifier producing two
highly correlated light beams (signal and idler) having orthogonal polarisations. However,
by mixing the signal and idler fields both become disentangled squeezed-vacuum fields. The
mixing is achieved by the polariser PBS 1. It also serves for introducing the local-oscillator
beam. The polariser PBS 2 combined with the L/2 wave plate HWP mixes the local oscillator
with the signal, and the emerging beams are detected, as required in balanced homodyne
measurement. Every pulse represents an individual quantum system within a statistical
ensemble whose quantum state (here the squeezed-vacuum state) is to be measured.
According to a fundamental theorem of Titulaer and Glauber,“‘) pulses can be understood
as single modes with spatially and temporally varying mode shapes ~(r, t), provided they obey
first-order coherence. Hence the set-up sketched in Fig. 8 measures a single-mode quantum
state on a set of optical pulses. In the experiment, 4000 repeated measurements at 27 values
of the local-oscillator phase 4 were performed. The measured quadrature distributions are
depicted in Fig. 9. The data were processed by a computer to reconstruct the Wigner function

1.57
Phase cp
FIG. 9. (a) Measured quadrature-amplitude distributions at various values of local oscillator
phase. Note that since these distributions are normalised, a decreasing width of a particular
distribution is accompanied by an increase in its peak height. (b) Variances of quadrature amplitude
vs. LO phase: circles, squeezed state; triangles, vacuum state. (Reproduced from Ref. (71).)
122 U. Leonhardt and H. Paul

-2%oV -1.0 0; . 2.0 -2%ob7EGTT-


. 2.0
x .
FIG. 10. Measured Wigner distributions for (a), (b) a squeezed state and (c), (d) a vacuum state,
viewed in 3D and as contour plots, with equal numbers of constant-height contours. Squeezing of
the noise distribution is clearly seen in (b). (Reproduced from Ref. (71).)

of the squeezed light (see Fig. lOa, lob) via the standard filtered back projection algorithm
for parallel-beam sampling geometry. (s2)By blocking the squeezed beam before it enters the
apparatus the Wigner function of vacuum was reconstructed as well, as can be seen from
Figs IOc, 10d. They visualise the vacuum fluctuations, while Figs IOa, lob show the
anisotropic fluctuations of a squeezed vacuum. In later experiments Wigner functions of
coherent fields(lO) and of phase- or amplitude-randomised coherent states were recon-
structed.@‘)
Optical Homodyne Tomography has been proved to be a powerful method for
reconstructing the Wigner function i.e. the quantum state of light from measured quadrature
distributions.

5. INEFFICIENT DETECTION

So far we have implicitly assumed that perfect detectors are used in our measurement
schemes for quasiprobability distributions. In practice, however, losses in photodetection or
due to mode mismatch cannot be completely avoided. In this Section we analyse the influence
of homodyne-detection losses on some schemes for measuring the Q function and on Optical
Homodyne Tomography. Intuitively, we may expect that inefficiencies smooth the quasiprob-
ability distributions as they smooth quadrature distributions measured by inefficient homo-
dyne detectors. We will identify these smoothed distributions as s-parametrised quasi-
probability distributions. (Is) Finally we will show that homodyne-detection losses can be
reduced when the signal is parameirically preamplified before it has been detected.
5.1. Beam splitting
The Q function of the signal beam can be measured using an eight-port homodyne detector
(see Section 4.1, Fig. 5). The essence of the measurement scheme is the splitting of the signal
into two beams. Then the x1 quadrature is measured on one beam and the p2 quadrature on
Measuring the quantum state of light 123

the other. The statistical distribution function for the measured x, and p2 values is the (scaled)
Q function of the signal

(133)

provided, perfect homodyne detectors have been employed in measuring xl and p2. (We use
the notation of s-parametrised quasiprobability distributions, see Section 2.4. The Q function
corresponds to s = - 1.) Now we study the influence of detection losses on the measured
distribution w(x) ,p2). As pointed out in Section 3.4, inefficient homodyne detectors can be
visualised as perfect homodyne detectors with fictitious beam splitters placed in front of them.
We obtain from the general formula (77) that the measured distribution w(x, ,p2) is a
smoothed ideal distribution Wid(x\ , p2)

whP2)=
ss
+co

-cc
i-cc

--oo
Wid(X;,p;)~
w
1
-rl)

Xe.,[ -&((x;-~)?+(p;-$)]dx;dp;. (134)

We utilise the fact that the ideal distribution Wid(X1,p2) is a scaled Q function (see Eqn. (133)),
then substitute x; = xi fi, p; = p2a, and find

W(XbP2) =
ss
+a0

-m
+co

--co
w(xy,p;; - 1)
1
n(l - rl)

xexp[-A((,;-x,#+(p;--p2J)31dx;.dP;. (135)

This equation shows clearly that the measured distribution function is a smoothed Q function.
According to Cahill and Glauber(“) we can express w (x, , p2) in terms of an s-parametrised
quasiprobability distribution using the general relationship (42) between different quasiprob-
ability distributions parametrised by s and t. In fact, setting c = - 1 and (t - s)-’ =
q /[2( 1 - q)] we get s = - (2 - q)/q, and obtain the final result

WhP2) = i w(x,&p2J ; -3. (136)

We see again that in general the Q function is not measured, but an s-parametrised
quasiprobability distribution with s less than - 1 for 0 < q < 1 and such a distribution is
smoother than the Q function. The eight-port homodyne detector requires perfect detection
efficiency for measuring the Q function.

5.2. Amplzjication
To measure the Q function the signal field can be linearly amplified, as shown in Section
4.2. This is, in fact, the oldest scheme for realistic measurement of phase.c5) Let us now study
the influence of detection inefficiencies on the measured distribution. After amplification the
distribution for the complex field amplitude should be measured. For instance, we could use
the eight-port homodyne detector described in Section 4.1. According to Eqn. (133) it
measures a scaled Q function, provided perfect detectors are used. Equation (43) relates the
124 U. Leonhardt and H. Paul

Q function to the Wigner function. We use the expression (108) for the Wigner function of
the amplified field and obtain

Wid(XI,P2)=2
ss
--co
+m +a,

-02
c-2w(x;/c,p;lc; -S21C2)

xkexp[-(x;-x,fi)2-(p;--p2,/$2]dx{dp;, (137)

where C and S are abbreviations of cash g and sinh g, respectively. Substituting x; = xi/C
and p; =p;/C we get
+m +m
wid(xI,P2)=2 w(x;,p;; -S2/C’)
s -m I -cc

x i exp[ - C2(x ;’- x, 8/C)’ - C’(p; -p2&C)l dx; dp;. (138)

Now we utilise the general relationship (42) between different quasiprobability distributions
parametrised by s and t, respectively. We set t = - S2/C2 and (t - s)-’ = C2, use C2 - S2 = 1,
and obtain s = - 1. Hence the ideal distribution function reads

WidCxl 3P2 ) =g w(x,~,P2&; -I), (139

being a scaled Q function. Here G = C2 denotes the gain of the amplifier. To understand the
influence of detection losses we use the same trick as in the preceding paragraph. We
convolute w. (x,,p,) with the Gaussian n(1 - rl-’ exp[-q(l - q)-‘(~;~+p;~)], substitute
x; = xi&, p; =p;@, and obtain

w(x, 9P2) =
ss
+m

-co
+m

-*
w(x;,p;; - 1)
1
n(l - ?)

x exp[ -&((x; -x,6>’ + (p; -pz&>‘>l dx;dA’. (140)

Now we utilise again the relationship (42) between different quasiprobability distributions.
Setting c = - 1 and (t - s)-’ = qGl[2(1 - q)] we get s = - 1 - 2(1 - q)/(rlG), and obtain the
result
wbP2)=‘I 2w(x,&,P2&; -1-359). (141)

Equation (141) shows that the s-parameter approaches - 1 when the gain G is large, whatever
the detection efficiency might be. Clearly, the reason is that a strongly amplified signal reaches
a macroscopic level where it is no longer affected by the quantum fluctuations of the detection
noise. However, we must account for the fact that the amplifier bringing the signal to that
level is not noiseless. Fortunately, the amplification noise and the beam-splitting noise are
equivalent. Both kinds of noise are the minimum noise sources needed for guaranteeing that
the split or the amplified beam remain bosonic light modes. Both noise sources play the role
of Lungeuin noise and hence they have the same effect. In this way we understand that linear
amplification of the signal is a means for measuring the Q function without using perfect
detectors.
Measuring the quantum state of light 125

5.3. Optical homodyne tomography


We may expect that losses in homodyne-detection efficiency influence Optical Homodyne
Tomography as well. Here the Wigner function is reconstructed from a set of measured
quadrature distributions provided they have been perfectly measured. To understand the
effect of inefficiencies we recall from Section 3.4 that they can be modeled as an effective beam
splitter placed in front of an ideal homodyne detector (see Fig. 4b). This fictitious beam
splitter transmits a part of the signal and introduces detection noise which is the vacuum noise
entering its unused port. In our model, we have separated the detection noise from the specific
type of measurement. The effective beam splitter comprises all kind of losses, while the
following ideal homodyne detector in Fig. 4b can be assumed to be perfect. Hence with
imperfect detectors quadrature values of the transmited beam are perfectly measured. Thus,
in Optical Homodyne Tomography the Wigner function of an attenuated field is recon-
structed. This Wigner function IV’@,p) is easily calculated. We express w’(x, p) in terms of
the total Wigner function for both output fields of the effective beam splitter

WY4PI =
I
+oc

s
-02 -a,

and obtain from Section 3.4, Eqns (73) and (74), that
fee
W'(x, p, X, jj) dl djj (142)

w’k P, %F) = &/ix - J-G% Jt?P - J=lp,

x 1 exp[-(fix + J+)’ - (&p + &)2]. (143)


lr

Substituting

we get

WY.%
P) =
ss
+a0

--co
+a3

-_oo
Wx’, P') ~ ’
41 -tl) -p{-&[(X’-k=+(p.--$=)iljdx’dp’.
(145)

According to Eqn. (44) we obtain the final compact result

1-V .
W’(x,p)=$
w $5;
( -?_
>
(146)

We see that due to detection losses an s-parametrised quasiprobability distribution instead


of the Wigner function is reconstructed. It is interesting to note that for 50% detection
efficiency the Q function is obtained which is never negative. Counting every second photon
thus wipes out any negativity the Wigner function may have. For 50% efficiency the amount
of introduced detection noise is exactly the same as the introduced vacuum noise in the
beam-splitting scheme for measuring the Q function directly with perfect detectors.
Quite generally, the parameter s in the reconstructed or directly measured quasiprobability
distribution can be understood as the amount of total detection noise. It comprises,
for instance, the beam-splitting or amplification noise or the noise due to inefficient
detection.
126 U. Leonhardt and H. Paul

5.4. High-accuracy optical homodyne detection


Is there a feasible way of overcoming losses in balanced homodyne detection? Of course,
the photodetectors used can be improved and the mode matching of the local oscillator and
the signal field can be enhanced as well. When the limits of present-day technology have been
reached, is there still a way to go a step further? We have learnt in Section 5.2 that an optical
amplifier can bring the signal field to a macroscopic level where it is no longer effected by
the extra quantum noise due to inefficient detection. So why not amplify the field? The other
side of the coin is that an amplifier adds noise to the signal. However, in homodyne detection
only one quadrature component x@ is measured. The conjugate component pe is left in
‘obscurity’. By allowing the latter to be deamplified we see that a squeezer, e.g. a degenerate
parametric amplifier will do just this required task. In fact, it squeezes one quadrature
component pe while anti-squeezing (i.e. amplifying) the other quadrature X~ which should
be measured. Hence we propose the set-up depicted in Fig. 11. An ‘anti-squeezer’ is used as
an preamplifier for a realistic homodyne measurement. According to the model discussed in
Section 3.4, an inefficient homodyne detector consists of an effective beam splitter to represent
losses and an ideal homodyne detector. In order to describe the<action of the apparatus we
use Wigner functions. First, the degenerate parametric amplifier squeezes the Wigner function
W(xe ,pe) of the signal field

w,(x,,,p,,) = WX,~G-“~,P,,G”*) (147)

with G being the gain of the amplifier (see Eqn. (3.46)). Second, the beam splitter provides
a coupling with a vacuum mode via its unused input port. We start with the Wigner function

wx,ex*e;Pw?P2,)=
7 wsq(x,,~Ple)w”,,(x*e,P2e), (148)

where W, (x,, , p,@) and Wvac(xz8, pze) = IZ-’ exp[ - (x:, + p:,)] are the Wigner functions for
the incident squeezed beam and the vacuum mode, respectively. According to Section 3.2 the
action of the beam splitter is simply described by a rotation in the x,,, x2@- and in the
plop he -plane
cos ci sin a
-sina cos a

Pie = cosa sin a Pie


6) (
28 -sin a cos a >6 ;63>*
(149)

anti-squeezer losses

.
G

detector 2

I
I I
vacuum local oscillator
FIG. 11. Schematic diagram of the proposed setup: an anti-squeezer, followed by a fictitious beam
splitter modelling losses in detection efficiency, and an ideal homodyne detector.
Measuring the quantum state of light 127

Here cos* a is to be identified with the transmittance of the fictitious beam splitter which
equals the (overall) detection efficiency u. Combining Eqns (148, 149), we find the Wigner
function for the total field emerging from the beam splitter to be given by

where the variables x1@, x2@,p,s and pze are to be substituted according to Eqn. (149). In
order to obtain the Wigner function for beam 1, W, (xie,pis), we trace over the variables
xie and p&, of the unobserved beam 2. Utilising the equations for xle and pie in (149) as
substitutions in the integral, we readily find W, (xiQ,pie) to be given by the following
convolution

w,(x;B,P;e)=~-‘(l -VI-’
ss +m

-cc -m
+m
WXW~PW)

x exp[-q(l -~)-‘G(x,,-~-~‘*G-~~~x;~)*]

xexP[-?(l -?)-‘G-‘&J -@2G’!2p;~)2]dx,,dp,,, (151)

where the transmittance cos2 a has been replaced by the detection efficiency q. Calculating the
marginal distribution for xie following from Eqn. (151) we obtain

wp(x;@) =
s +m

-m
W,(x;,,p;e) dp;,

=7-C -l/2(1 _ rl)-“2


s+‘X

-9
w&,e)

x exp[-q(l - q)-‘G(x,@ - 1-112G-1’2x;e)2]dx,8, (154

where the relation we(x,,) = J?z W(x ,e,p,e) dple has been utilised. We see that the
measured quadrature distribution w$)(x;@) is a smoothed ideal distribution w,(x,,), quite
similar to the distribution measured without preamplification. However, the smoothing is
reduced. Moreover, in the limit of strong amplification where the gain G is large, the Gaussian
in Eqn. (152) approaches a delta function. Consequently the measured distribution becomes
a scaled ideal distribution
w$‘(x&) = (~G)-“2~,((~G)-“2~;8). (153)

In this case the use of the ‘anti-squeezer’ as a preamplifier in homodyne detection allows one
to measure perfectly quadratures with imperfect detectors. Only an appropriate resealing of
the measured distribution is needed according to Eqn. (153).
Apart from other applications, the proposed set-up can be used in Optical Homodyne
Tomography. Here, of course, the phase of the pumping field has to be properly adjusted
to the phase of the local oscillator, in order to ensure that just the measured quadrature
becomes anti-squeezed. However, this is easily achieved when both the local oscillator and
the pump originate from a common laser beam (with the pump beam being frequency-
doubled), as it is usually the case. In this way the measured distributions w$‘(x&,) are
always affected by the same detection noise. They are convolutions with the
Gaussian rc- li2(l - ~)-“2exp[-~(l - q)-’ Gx:,]. Hence, the tomographically reconstructed
128 U. Leonhardt and H. Paul

quasiprobability distribution w’(x, p) is a two-dimensional convolution with the same


Gaussian. According to Eqn. (44) this is an s-parametrised quasiprobability distribution

w x,p;
Wf(x,p)=$
(J&m --1-V . (154)
qG>
We see that W’(x, p) approaches a scaled Wigner function in the limit of strong amplification

w(-&,-&;o)=--$
wyx,p)=$ W(--j$,-j==) for G-too, (155)

as we may expect, since in this case the quadratures are perfectly measured. Note that this
tomographically reconstructed Wigner function is distinct from the Wigner functions
W,(x,, pe) and W, (xi@, pi@) for the squeezed field and the detected field, respectively. The
use of the anti-squeezed field quadratures only and the phase-shifting of the local oscillator
with respect to the pump allows us to get the best out of the signal. In this way we can indeed
reconstruct the Wigner function.
On the other hand, to measure the Q function we could use the beam-splitting set-up
described in Section 4. I with our improved homodyne detector. However, in that case it may
be easier to utilise the amplification scheme proposed by Bandilla and Paul (see Section 4.2).
This scheme uses phase-insensitive amplification of the field by a quantum amplifier before
the quadrature measurement, and lower detection efficiencies are tolerated.

6. SUMMARY

Measuring the quantum state of a single light mode can be done in several schemes mostly
based on homodyne detection. Measured or reconstructed from measurements are quasiprob-
ability distributions for quadrature components. We have addressed beam splitting, amp@-
cation and heterodyning as experimental schemes for measuring the Q function. We suggested
ways of measuring the squared Wigner function directly, and we discussed Optical Homodyne
Tomography as a means for reconstructing the Wigner function from quadrature distri-
butions. Finally, we discussed the influence of detection inefficiencies on the measured or
reconstructed quasiprobability distributions. We showed how this influence can be reduced.
The community of theoreticians has the pleasure to see that s-parametrised quasiprobability
distributions are not only quite useful mathematical constructions for calculations, but have
indeed a physical meaning. They are associated with relatively simple measurement schemes.
The parameter s comprises the overall detection noise of a particular detection scheme. We
have also shown that approximate simultaneous measurements of position and momentum can
be indeed performed in quantum optics. Again, quantum optics has proved its potential for
simple yet fundamental experiments of quantum physics.
Experimentalists will gain from the measurement schemes more information on nonclassi-
cal light fields than expressed in single bare parameters, like the squeezing parameter or
Mandel’s Q, as was done previously. In fact, the schemes offer the-great possibility of gaining
the maximal information allowed by the very principles of quantum physics. We believe that
they are becoming an experimental standard. Experimentalists all over the world are invited
to use them.
Acknowledgemenrs-We profited very much from countless discussions with our colleagues H. A. Bachor, A.
Bandilla, S. M. Bamett, M. Freyberger, R. J. Glauber, M. Grabowski, U. Herzog, Z. Hradil, I. Jex, P. L. Knight,
L. Knoll, H. Kuhn, R. Loudon, A. Orlowski, M. G. Raymer, Th. Richter, H. Risken, H.-H. Ritze, W. Schleich,
S. Stenholm, J. A. Vaccaro, K. Vogel, W. Vogel, D.-G. Welsch and A. Wtinsche. We would like to thank them for
their help and advice. We am especially indebted to M. G. Raymer for giving us the original figures of the pioneering
Optical-Homodyne-Tomography experiment.
Measuring the quantum state of light 129

REFERENCES

I. G. S. Agarwal and S. Chaturvedi, Phys. Rev. A49, R665 (1994).


2. E. Arthurs and J. L. Kelly, Jr, Bell iysr. Tech. J. 44, 725 (1965).
3. E. Arthurs and M. S. Goodman. Phvs. Rev. Lefr. 60. 2447 (1988).
4. A. Aspect, Ph. Grangier and G.‘Roger, Phys. Rev. .&t. 49,‘91 (1982)
5. A. Bandilla and H. Paul, Ann. Phys. (Leipzig) 23, 323 (1969).
6. A. Bandilla and H. Paul, Ann. Phys. (Leipzig) 24, 119 (1970).
7. S. M. Bamett and D. T. Pegg, J. Mod. Opt. 36, 7 (1989).
8. S. M. Bamett, Physics Worhz: July 1993, pp. 28.
9. M. Beck. D. T. Smithev and M. G. Ravmer. Phvs. Reo. A48. R890 (1993)
10. M. Beck; D. T. Smithe;, J. Cooper and M. G. Rawer, Opt.’ Left. 1‘8, 1259 (1993).
11. J. Bertrand and P. Bertrand, Found. Phys. 17, 397 (1987).
12. S. L. Braustein, Phys. Rev. A42, 474 (1990).
13. S. L. Braunstein, C. M. Caves and G. J. Milburn, Phys. Rev. A43, 1153 (1991).
14. K. E. Cahill and R. J. Gauber, Phys. Rev. 177, 1857 (1969).
15. K. E. Cahill and R. J. Glauber, Phys. Rev. 177, 1882 (1969).
16. R. A. Campos, B. E. A. Saleh and M. C. Teich, Phys. Rev. A40, 1371 (1989).
17. C. Caves, Phys. Rev. D26, 1817 (1982).
18. F. A. M. de Oliveira, Phys. Rev. A45, 3113 (1992).
19. M.-A. Dupertius and S. Stenholm, J. Opt. Sot. Am. B4, 1094 (1987).
20. M.-A. Dupertius, S. M. Bamett and S. Stenholm, J. Opt. Sot. Am. B4, 1102 (1987).
21. M.-A. Dupertius, S. M. Bamett and S. Stenholm, J. Opt. Sot. Am. B4, 1124 (1987).
22. S. M. Dutra, P. L. Knight and H. Moya-Cessa, Phys. Rev. A48, 3168 (1993).
23. A. Einstein, B. Podolsky and N. Rosen, Phys. Rev. 47, 777 (1935).
24. A. K. Ekert and P. L. Knight, Am. J. Phys. 57, 692 (1989).
25. M. Freyberger, K. Vogel and W. Schleich, Phys. Left. A176, 41 (1993).
26. M. Freyberger and A. M. Herkommer, Phys. Rev. Left. 72, 1952 (1994).
27. A. L. Gaeta and R. W. Boyd, Phys. Rev. Leti. 60, 2618 (1988).
28. C. W. Gardiner, Quantum Noise, Chap. 4.4.4., Springer, Berlin (1991).
29. C. W. Gardiner, Quantum Noise, Chap. 6.4., Springer, Berlin (1991).
30. H. Gerhardt, H. Welling and D. Friilich, Appl. Phys. 2, 91 (1973).
31. H. Gerhardt, U. Biichler and G. Litfin, Phys. Left. 49A, 119 (1974).
32. C. R. Gilson, S. M. Barnett and S. Stenholm, J. Mod. Opt. 34, 949 (1987).
33. M. Hillery, R. F. O’Connell, M. 0. Scully and E. P. Wigner, Phys. Rep. 106, 121 (1984).
34. Z. Hradil, Quanrum Opt. 4, 93 (1992).
35. P. L. Kelley and W. H. Kleiner, Phys. Rev. 136, A316 (1964).
36. H. Kuhn, D.-G. Welsch and W. Vogel, J. Mod. Optics (in press).
37. P. G. Kwiat, A. M. Steinberg, R. Y. Chiao, P. H. Eberhard and M. D. Petroff, Phys. Rev. A48, R867 (1993).
38. Y. Lai and H. A. Haus, Quantum Opt. 1, 99 (1989).
39. U. Leonhardt and H. Paul, Phys. Rev. A47, R2460 (1993).
40. U. Leonhardt and H. Paul, .I. Mod. Optics 40, 1745 (1993).
41. U. Leonhardt, Phys. Rev. A48, 3265 (1993).
42. U. Leonhardt and H. Paul, Phys. Rev. A48, 4598 (1993).
43. U. Leonhardt, Phys. Rev. A49, 1231 (1994).
44. U. Leonhardt and I. Jex, Phys. Rev. A49, R1555 (1994).
45. U. Leonhardt and H. Paul, J. Mod. Optics 41, 1427 (1994).
46. U. Leonhardt and H. Paul, Phys. Let?. A193, 117 (1994).
47. U. Leonhardt and H. Paul, Phys. Rev. L&t. 72, 4086 (1994).
48. M. A. M. Marte. H. Ritsch and D. F. Walls. Phvs. Rev. ASS. 3577 (1988).
49. B. R. Mollow and R. J. Glauber, Phys. Reu.‘l60; 1076 (1967). ~ ’
50. B. R. Mollow and R. J. Glauber, Phys. Rev. 160, 1097 (1967).
51. B. R. Mollow, Phys. Rev. 168, 1896 (1968).
52. F. Natterer, The Mathematics of Computerized Tomography. Wiley, New York (1986).
53. J. W. Noh, A. Fougeres and L. Mandel, Phys. Rev. Lett. 67, 1426 (1991).
54. J. W. Noh, A. Fougeres and L. Mandel, Phys. Rev. A45, 424 (1992).
55. J. W. Noh, A. Fougbres and L. Mandel, Phys. Rev. A46, 2840 (1992).
56. Z. Y. Ou, S. F. Pereira and H. J. Kimble, Phys. Rev. Lett. 70, 3239 (1993).
57. H. Paul, Fortschr. Phys. 22, 657 (1974).
58. W. Pauli, In: Die aiigemeinen Prinzipien der Wellenmechanik, in Handbuch der Physik, Vol. 24. part 1, H. Geiger
and K. Scheel (eds), Springer, Berlin (1993); re-edited in Handbuch der Physik, Vol. 5, Part 1, S. Fliigge (ed.),
Springer, Berlin (1958); English translation: W. Pauli, General Principles of Quanmm Mechnics, p. 17, Springer,
Berlin (1980).
59. D. T. Pegg and S. M. Barnett, Europhys. Left. 6, 483 (1988).
60. D. T. Pegg and S. M. Barnett, Phys. Rev. A39, 1665 (1989).
61. M. G. Raymer, D. T. Smithey, M. Beck, M. Anderson and D. F. McAlister, Measurement of the Wigner
function in quantum optics in the Proceedings of the Third International Wigner Symposium, Int. J. Mod. Phys.
R, Vol. 8 (1994).
130 U. Leonhardt and H. Paul

62. M. G. Raymer, M. Beck and D. F. McAlister, Phys. Rev. Lerr. 72, 1137 (1994).
63. M. Rousseau, J. Phys. Math. Gen. 10, 1043 (1977).
64. W. Schleich, A. Bandilla and H. Paul, Phys. Rev. A45, 6652 (1992).
65. M. 0. Scully and W. E. Lamb, Jr, Phys. Rev. 179, 368 (1969).
66. A. Selloni, P. Schwedimann, A. Quattropani and H. P. Baltes, J. Phys. Math. Gen. 11, 1427 (1978).
67. J. H. Shapiro and S. S. Wagner, IEEE J. Quanrum Electron. QEZO, 803 (1984).
68. J. H. Shapiro, IEEE J. Quantum Elecrron. QE-21, 237 (1985).
69. R. E. Slusher, L. W. Hollberg, B. Yurke, J. C. Mertz and J. F. Valley, Phys. Rev. Lerr. 55, 2409 (1985).
70. R. E. Slusher, P. Grangier, A. LaPorta, B. Yurke and M. Potasek, Phys. Rev. Left. 59, 2566 (1987).
71. D. T. Smithey, M. Beck, M. G. Raymer and A. Faridani, Phys. Rev. Lerr. 70, 1244 (1993).
72. D. T. Smithey, M. Beck, J. Cooper and M. G. Raymer, Phys. Ser. T48, 35 (1993).
73. D. T. Smithey, M. Beck, J. Cooper and M. G. Raymer, Phys. Rev. A48, 3159 (1993).
74. M. D. Srinivas and E. B. Davies, Opt. Acra 28, 981 (1981).
75. S. Stenholm, Ann. Phys. (N.Y.) 218, 233 (1992).
76. V. I. Tatarskii, 50s. Phys. Usp. 26, 311 (1983).
77. V. M. Titulaer and R. J. Gauber, Phys. Rev. 145, 1041 (1966).
78. J. A. Vaccaro and D. T. Pegg, J. Mod. Opt. 34, 855 (1987).
79. K. Vogel and H. Risken, Phys. Rev. A40, 2847 (1989).
80. K. Vogel, V. M. Akulin and W. P. Schleich, Phys. Rev. L-err. 71, 1816 (1993).
81. W. Vogel and J. Grabow, Phys. Rev. A47, 4427 (1993).
82. N. G. Walker and J. E. Carroll, Elecrron. Lerr. 20, 981 (1984).
83. N. G. Walker and J. E. Carroll, Opr. Quantum Elecrron. 18, 355 (1986).
84. N. G. Walker, J. Mod. Optics 34, 15 (1987).
85. E. P. Wigner, Phys. Rev. 40, 749 (1932).
86. M. Wilkens and P. Meystre, Phys. Rev. A43, 3832 (1991).
87. K. Wbdkiewicz, Phys. Rev. Len. 52, 1064 (1984).
88. K. Wodkiewicz, Phys. Lerr. A124, 207 (1987).
89. H. P. Yuen and J. H. Shapiro, In: Coherence and Quanrum Oprics IV, L. Mandel and E. Wolf (eds), p. 719.
Plenum, New York (1978).
90. H. P. Yuen, Phys. L&t. 91A, 101 (1982).
91. H. P. Yuen and V. W. S. Chan, Opt. Lert. 8, 177 (1983).

You might also like