1: \documentclass{ws-ijqi}
2: \usepackage{bbm,amsmath}
3: \begin{document}
4: %%%%%%%%%%%%%%%%%%
5: \def\set#1{{\cal #1}}\def\sH{\set{H}}
6: \def\Tr{\hbox{Tr}}
7: \def\sT{{\scriptscriptstyle T}}
8: \def\kket#1{|#1\rangle\rangle}
9: \def\bbra#1{\langle\langle #1|}
10: \newcommand{\ket}[1]{ |#1 \rangle}
11: \newcommand{\bra}[1]{ \langle #1|}
12: \newcommand{\dket}[1]{ | #1 \rangle\!\rangle}
13: %%%%%%%%%%%%%%%%%%
14: \markboth{Matteo G A Paris}{Local Quantum Estimation}
15: \title{QUANTUM ESTIMATION for QUANTUM TECHNOLOGY}
16: \author{MATTEO G A PARIS}
17: \address{Dipartimento di Fisica dell'Universit\`a di Milano, I-20133
18: Milano, Italia \\ CNSIM, Udr Milano, I-20133 Milano, Italia \\ ISI
19: Foundation, I-10133 Torino, Italia}
20: \date{\today}
21: \maketitle
22: %%%%%%%%%%%%%%%%%%
23: \begin{abstract}
24: Several quantities of interest in quantum information, including
25: entanglement and purity, are nonlinear functions of the density matrix
26: and cannot, even in principle, correspond to proper quantum observables.
27: Any method aimed to determine the value of these quantities should
28: resort to indirect measurements and thus corresponds to a parameter
29: estimation problem whose solution, {\em i.e} the determination of the
30: most precise estimator, unavoidably involves an optimization procedure.
31: We review local quantum estimation theory and present explicit formulas
32: for the symmetric logarithmic derivative and the quantum Fisher
33: information of relevant families of quantum states. Estimability of a
34: parameter is defined in terms of the quantum signal-to-noise ratio and
35: the number of measurements needed to achieve a given relative error.
36: The connections between the optmization procedure and the geometry
37: of quantum statistical models are discussed. Our analysis allows to
38: quantify quantum noise in the measurements of non observable quantities
39: and provides a tools for the characterization of signals and
40: devices in quantum technology.
41: \end{abstract}
42: %%%%%%%%%%%%%%%%%%
43: \section{Introduction}\label{s:intro}
44: Many quantities of interest in physics are not directly accessible,
45: either in principle or due to experimental impediments. This is
46: particolarly true for quantum mechanical systems where relevant
47: quantities like entanglement and purity are nonlinear functions of the
48: density matrix and cannot, even in principle, correspond to proper
49: quantum observables.
50: In these situations one should resort to indirect
51: measurements, inferring the value of the quantity of interest by
52: inspecting a set of data coming from the measurement of a different
53: obeservable, or a set of observables. This is basically a parameter
54: estimation problem which may be properly addressed in the framework of
55: quantum estimation theory (QET) \cite{QET}, which provides
56: analytical tools to find the optimal measurement according to some
57: given criterion. In turn, there are two main paradigms in QET:
58: {\em Global} QET looks for the POVM minimizing a suitable cost
59: functional, averaged over all possible values of the parameter to
60: be estimated. The result of a global optimization is thus a single
61: POVM, independent on the value of the parameter. On the other
62: hand, {\em local} QET looks for the POVM maximizing the Fisher
63: information, thus minimizing the variance of the estimator, at a
64: fixed value of the parameter \cite{hel67,yl73,hk74,BC94,BC96}. Roughly speaking,
65: one may expect local QET to provide better performances since the
66: optimization concerns a specific value of the parameter, with some
67: adaptive or feedback mechanism assuring the achievability of the
68: ultimate bound \cite{gil00}. Global QET has been mostly applied
69: to find optimal measurements and to evaluate lower bounds on
70: precision for the estimation of parameters imposed by unitary
71: transformations. For bosonic systems these include single-mode
72: phase \cite{Hol79,Dar98}, displacement \cite{Hel74}, squeezing
73: \cite{Mil94,Chi06} as well as two-mode transformations, e.g.
74: bilinear coupling \cite{per01}. Local QET has been applied to the
75: estimation of quantum phase \cite{Mon06} and to estimation
76: problems with open quantum systems and non unitary processes
77: \cite{Sar06}: to finite dimensional systems \cite{Hot06}, to
78: optimally estimate the noise parameter of depolarizing
79: \cite{Fuj01} or amplitude-damping \cite{Zhe06}, and for continuous
80: variable systems to estimate the loss parameter of a quantum
81: channel \cite{Dau06,Gra87,Pol92,Mon07} as well as the position of a single
82: photon \cite{fri07}. Recently, the geometric structure induced
83: by the Fisher information itself has been exploited to give a
84: quantitative operational interpretation for multipartite
85: entanglement \cite{Boi08} and to assess quantum criticality as
86: a resource for quantum estimation \cite{ZP07}.
87: \par
88: In this paper we review local quantum estimation theory and present
89: explicit formulas for the symmetric logarithmic derivative and the
90: quantum Fisher information of relevant families of quantum states. We
91: are interested in evaluating the ultimate bound on precision
92: (sensitivity), {\em i.e} the smallest value of the parameter that can be
93: discriminated, and to determine the optimal measurement achieving those
94: bounds. Estimability of a parameter will be then defined in terms of
95: the quantum signal-to-noise ratio and the number of measurements needed
96: to achieve a given relative error.
97: \par
98: The paper is structured as follows. In the next Section we review local
99: quantum estimation theory and report the solution of the
100: optimization problem, {\em i.e.} the determination of the optimal
101: quantum estimator in terms of the symmetric logarithmic derivative,
102: as well as the ultimate bounds to precision in terms of the quantum
103: Fisher information. General formulas for the symmetric logarithmic
104: derivative and the quantum Fisher information are derived.
105: In Section \ref{s:qep} we address the quantification of estimability
106: of a parameter put forward the quantum signal-to-noise ratio and
107: the number of measurements needed to achieve a given relative error
108: as the suitable figures of merit.
109: In Section \ref{s:exa} we present explicit formulas for sets of pure
110: states and the generic unitary family. We also consider the multiparamer
111: case and the problem of repametrization. In Section \ref{s:geo}
112: we discuss the connections between estimability of a set of parameters,
113: the optmization procedure and the geometry of quantum statistical models.
114: Section \ref{s:out} closes the paper with some concluding remarks.
115: %%%%
116: \section{Local quantum estimation theory}\label{s:qet}
117: The solution of a parameter estimation problem amounts to find
118: an estimator, {\em i.e} a mapping $\hat\lambda=\hat\lambda
119: (x_1,x_2,...)$ from the set $\chi$ of measurement outcomes into
120: the space of parameters.
121: Optimal estimators in classical estimation theory are those
122: saturating the Cramer-Rao inequality \cite{Cra46}
123: \begin{align}
124: {\mathrm{V}}(\lambda) \geq \frac{1}{M F(\lambda)}
125: \label{eq:CramerRao}
126: \end{align}
127: which establishes a lower bound on the mean square error $V(\lambda)
128: = E_{\lambda} [(\hat\lambda(\{x\}) - \lambda)^2]$ of any estimator
129: of the parameter $\lambda$.
130: In Eq. (\ref{eq:CramerRao}) $M$ is the number of measurments
131: and $F(\lambda)$ is the so-called
132: Fisher Information (FI)
133: \begin{align}
134: F(\lambda) = \int\!\!dx\, p(x\vert \lambda) \left(
135: \frac{\partial \ln p(x\vert \lambda)}{\partial \lambda}
136: \right)^2
137: = \int\!\!dx\, \frac{1}{p(x\vert \lambda)} \left(
138: \frac{\partial p(x\vert \lambda)}{\partial \lambda}
139: \right)^2 . \label{eq:ClassicalFisher}
140: \end{align}
141: where $p(x\vert\lambda)$ denotes the conditional probability of
142: obtaining the value $x$ when the parameter has the value $\lambda$.
143: For unbiased estimators,
144: as those we will deal with, the mean square error is equal to the
145: variance $\hbox{Var}(\lambda)= E_{\lambda} [\hat\lambda^2] -
146: E_{\lambda} [\hat\lambda]^2$.
147: \par
148: When quantum systems are involved any estimation problem may be stated
149: by considering a family of quantum states $\varrho_\lambda$ which are
150: defined on a given Hilbert space ${\cal H}$ and labeled by a parameter
151: $\lambda$ living on a $d$-dimensional manifold ${\cal M}$, with the
152: mapping $\lambda \mapsto \varrho_\lambda$ providing a coordinate system.
153: This is sometimes referred to as a quantum statistical model. The
154: parameter $\lambda$ does not, in general, correspond to a quantum
155: observable and our aim is to estimate its values through the measurement
156: of some observable on $\varrho_\lambda$. In turn, a quantum estimator
157: $O_\lambda$ for $\lambda$ is a selfadjoint operator, which describe a
158: quantum measurement followed by any classical data processing performed
159: on the outcomes. The indirect procedure of parameter estimation implies
160: an additional uncertainty for the measured value, that cannot be avoided
161: even in optimal conditions. The aim of quantum estimation theory is to
162: optimize the inference procedure by minimizing this additional
163: uncertainty.
164: \par
165: In quantum mechanics, according to the Born rule we have
166: $p(x\vert \lambda) = \Tr[\Pi_x \varrho_\lambda]$ where
167: $\left\{\Pi_x\right\}$, $\int\!\!dx\, \Pi_x = {\mathbbm I}$,
168: are the elements of a positive operator-valued measure (POVM) and
169: $\varrho_\lambda$ is the density operator parametrized by the quantity
170: we want to estimate. Introducing the Symmetric Logarithmic
171: Derivative (SLD) $L_\lambda$ as the selfadjoint operator satistying
172: the equation \begin{align}
173: \frac{L_\lambda \varrho_\lambda + \varrho_\lambda L_\lambda}{2} =
174: \frac{\partial \varrho_\lambda}{\partial \lambda} \label{eq:SLD}
175: \end{align}
176: we have that
177: $\partial_\lambda p(x\vert\lambda) = \Tr[ \partial_\lambda\varrho_\lambda \Pi_x]
178: = \hbox{Re}( \Tr[\varrho_\lambda \Pi_x L_\lambda ] )$.
179: The Fisher Information (\ref{eq:ClassicalFisher}) is then rewritten as
180: \begin{align}
181: F(\lambda) = \int\!\!dx\, \frac{\hbox{Re}\left(\Tr\left[\varrho_\lambda \Pi_x
182: L_\lambda\right]\right)^2} {\Tr[\varrho_\lambda \Pi_x]}
183: \:. \label{eq:CQFisher}
184: \end{align}
185: For a given quantum measurement, {\em i.e.} a POVM $\{\Pi_x\}$, Eqs.
186: (\ref{eq:ClassicalFisher}) and (\ref{eq:CQFisher}) establish the
187: classical bound on precision, which may be achieved by a proper
188: data processing, {\em e.g.} by maximum likelihood, which is
189: known to provide an asymptotically efficient estimator.
190: On the other hand, in order to evaluate the ultimate bounds to
191: precision we have now to maximize the Fisher information over
192: the quantum measurements.
193: Following Refs. \cite{yl73,hk74,BC94,BC96} we have
194: \begin{align}
195: F(\lambda) & \leq
196: \int\!\!dx\, \left|\frac{\Tr\left[\varrho_\lambda \Pi_x
197: L_\lambda\right]}{\sqrt{\Tr[\varrho_\lambda \Pi_x]}}\right|^2
198: \label{uno} \\
199: & = \int\!\!dx\, \left|
200: \Tr\left[ \frac{\sqrt{\varrho_\lambda}
201: \sqrt{\Pi_x}}{\sqrt{\Tr\left[\varrho_\lambda \Pi_x\right]}}\, \sqrt{\Pi_x} L_\lambda
202: \sqrt{\varrho_\lambda} \right] \right|^2 \nonumber \\
203: & \leq \int\!\!dx\, \Tr\left[\Pi_x L_\lambda \varrho_\lambda L_\lambda\right]
204: \label{due}\\
205: & =\Tr[L_\lambda \varrho_\lambda L_\lambda ] = \Tr[\varrho_\lambda L_\lambda^2]
206: \nonumber
207: \end{align}
208: The above chain of inequalities prove that the Fisher information
209: $F(\lambda)$ of any quantum measurement is bounded by the so-called
210: \emph{Quantum Fisher Information} (QFI)
211: \begin{align}
212: F(\lambda) \leq H(\lambda)
213: \equiv \Tr[\varrho_\lambda L_\lambda^2]
214: = \Tr[\partial_\lambda \varrho_\lambda L_\lambda]
215: \label{eq:QuantumFisher}
216: \end{align}
217: leading the quantum Cramer-Rao bound
218: \begin{align}
219: {\mathrm{Var}}(\lambda) \geq \frac{1}{M H(\lambda)} \label{QCR}
220: \end{align}
221: to the variance of any estimator. The quantum version of the
222: Cramer-Rao theorem provides an ultimate bound: it does depend on
223: the geometrical structure of the quantum statistical model and
224: does not depend on the measurement.
225: Optimal quantum measurements for the estimation of $\lambda$
226: thus corresponds to POVM with a Fisher information equal to the quantum Fisher
227: information, {\em i.e} those saturating both inequalities (\ref{uno})
228: and (\ref{due}). The first one is saturated when
229: $\Tr[\varrho_\lambda \Pi_x L_\lambda]$ is a real number $\forall \lambda$.
230: On the other hand, Ineq. (\ref{due}) is based on the Schwartz inequality
231: $|\Tr[A^\dag B]|^2 \leq \Tr[A^\dag A] \Tr[B^\dag B]$ applied to
232: $A^\dag= \sqrt{\varrho_\lambda}
233: \sqrt{\Pi_x}/\sqrt{\Tr\left[\varrho_\lambda \Pi_x\right]}$ and
234: $B=\sqrt{\Pi_x} L_\lambda \sqrt{\varrho_\lambda}$ and it is saturated when
235: \begin{equation}
236: \frac{\sqrt{\Pi_x}\sqrt{\varrho_\lambda}}
237: {\Tr\left[\varrho_\lambda \Pi_x\right]}
238: =
239: \frac{\sqrt{\Pi_x} L_\lambda \sqrt{\varrho_\lambda}}
240: {\Tr[\varrho_\lambda \Pi_x L_\lambda]}\qquad \forall \lambda\:,
241: \label{opt}
242: \end{equation}
243: The operatorial condition in Eq. (\ref{opt}) is satisfied iff $\{\Pi_x\}$
244: is made by the set of projectors over the eigenstates of $L_\lambda$,
245: which, in turn, represents the optimal POVM to estimate the parameter
246: $\lambda$. Notice, however, that $L_\lambda$ itself may not represent
247: the optimal observable to be measured. In fact, Eq. (\ref{opt})
248: determines the POVM and not the estimator {\em i.e} the function of the
249: eigenvalues of $L_\lambda$. As we have already mentioned above, this
250: corresponds to a classical
251: post-processing of data aimed to saturate the Cramer-Rao inequality
252: (\ref{eq:CramerRao}) and may be pursued by maximum likelihood, which is
253: known to provide an asymptotically efficient estimator. Using the fact
254: that $\Tr[\varrho_\lambda L_\lambda]=0$ an explicit form for the
255: optimal quantum estimator is given by
256: \begin{equation}
257: O_\lambda = \lambda {\mathbbm I}
258: + \frac{L_\lambda}{H(\lambda)}
259: \label{QEst}
260: \end{equation}
261: for which we have
262: $$\Tr[\varrho_\lambda O_\lambda] = \lambda\,, \:\: \Tr[\varrho_\lambda
263: O_\lambda^2] = \lambda ^2 + \frac{\Tr[\varrho_\lambda
264: L^2_\lambda]}{H^2(\lambda)}\,,\:\: \hbox{and thus }
265: \langle \Delta O^2_\lambda \rangle = 1/H(\lambda)\:.$$
266: \par
267: Eq. (\ref{eq:SLD}) is Lyapunov matrix equation to be solved
268: for the SLD $L_\lambda$. The general solution may be written as
269: \begin{equation}
270: L_\lambda = 2 \! \int_0^\infty\!\! dt\, \exp\{-\varrho_\lambda t\}\, \partial_\lambda
271: \varrho_\lambda \exp\{-\varrho_\lambda t\}
272: \label{LE}
273: \end{equation}
274: which, upon writing $\varrho_\lambda$ in its eigenbasis
275: $\varrho_\lambda = \sum_n \varrho_n |\psi_n\rangle\langle\psi_n |$,
276: leads to
277: \begin{equation}
278: L_\lambda = 2 \sum_{nm} \frac{\langle \psi_m| \partial_\lambda
279: \varrho_\lambda | \psi_n\rangle}{\varrho_n+ \varrho_m}
280: |\psi_m\rangle\langle\psi_n | \:,
281: \label{LL}
282: \end{equation}
283: where the sums include only terms with $\varrho_n+\varrho_m \neq 0$.
284: The quantum Fisher information is thus given by
285: \begin{equation}
286: H (\lambda) = 2 \sum_{nm} \frac{\left|\langle \psi_m| \partial_\lambda
287: \varrho_\lambda | \psi_n\rangle\right|^2}{\varrho_n+ \varrho_m}\:,
288: \label{HH}
289: \end{equation}
290: or, in a basis independent form,
291: \begin{equation}
292: H (\lambda) = 2 \! \int_0^\infty\!\! dt\,
293: \Tr\left[
294: \partial_\lambda \varrho_\lambda
295: \exp\{-\varrho_\lambda t\}\,
296: \partial_\lambda \varrho_\lambda
297: \exp\{-\varrho_\lambda t\}
298: \right]
299: \label{HHbi}\:.
300: \end{equation}
301: Notice that the SLD is defined only on the support of
302: $\varrho_\lambda$ and that both the eigenvalues $\varrho_n$ and the
303: eigenvectors $|\psi_n\rangle$ may depend on the parameter.
304: In order to separate the two contribution to the QFI we explicitly
305: evaluate $\partial_\lambda\varrho_\lambda$
306: \begin{equation}
307: \partial_\lambda\varrho_\lambda = \sum_p
308: \partial_\lambda \varrho_p |\psi_p\rangle\langle \psi_p | +
309: \varrho_p |\partial_\lambda \psi_p\rangle\langle \psi_p | +
310: \varrho_p |\psi_p\rangle\langle\partial_\lambda \psi_p | \label{der}
311: \end{equation}
312: The symbol $|\partial_\lambda \psi_n\rangle$ denotes
313: the ket $|\partial_\lambda \psi_n\rangle = \sum_k
314: \partial_\lambda \psi_{nk} | k\rangle$, where $\psi_{nk}$
315: are obtained expanding $|\psi_n\rangle$ in arbitrary basis
316: $\{|k\rangle\}$ independent on $\lambda$.
317: Since $\langle \psi_n | \psi_m\rangle = \delta_{nm}$ we
318: have $\partial_\lambda \langle \psi_n | \psi_m\rangle
319: \equiv \langle\partial_\lambda \psi_n |\psi_m\rangle +
320: \langle \psi_n |\partial_\lambda \psi_m\rangle = 0$
321: and therefore
322: $$ \hbox{Re} \langle\partial_\lambda \psi_n |\psi_m\rangle = 0
323: \qquad \langle\partial_\lambda \psi_n |\psi_m\rangle = -
324: \langle \psi_n |\partial_\lambda \psi_m\rangle = 0\:. $$
325: Using Eq. (\ref{der}) and the above identities we have
326: \begin{equation}
327: L_\lambda =
328: \sum_p \frac{\partial_\lambda \varrho_p}{\varrho_p}
329: |\psi_p\rangle\langle \psi_p | +
330: 2 \sum_{n\neq m} \frac{\varrho_n - \varrho_m}{\varrho_n + \varrho_m}
331: \langle \psi_m |\partial_\lambda \psi_n\rangle
332: \: |\psi_m\rangle\langle \psi_n |
333: \label{Le}
334: \end{equation}
335: and in turn
336: \begin{equation}
337: \label{He}
338: H (\lambda) =
339: \sum_p \frac{\left(\partial_\lambda \varrho_p\right)^2}{\varrho_p}
340: + 2 \sum_{n\neq m} \sigma_{nm}
341: \left|\langle \psi_m |\partial_\lambda \psi_n\rangle \right|^2
342: \end{equation}
343: where
344: \begin{equation}\label{sg1}
345: \sigma_{nm} = \frac{(\varrho_n - \varrho_m)^2}{\varrho_n+\varrho_m} +
346: \hbox{any antisymmetric term}\:,
347: \end{equation}
348: as for example
349: \begin{equation}\label{sg2}
350: \sigma_{nm} = 2 \varrho_n \frac{\varrho_n - \varrho_m}{\varrho_n+\varrho_m}
351: \qquad
352: \sigma_{nm} = 2 \varrho_n \left(\frac{\varrho_n
353: - \varrho_m}{\varrho_n+\varrho_m}\right)^2
354: \end{equation}
355: The first term in Eq. (\ref{He}) represents the classical Fisher
356: information of the distribution $\{\varrho_p\}$ whereas the second
357: term contains the truly quantum contribution. The second term vanishes
358: when the eigenvectors of $\varrho_\lambda$ do not depend. In this case
359: $[\varrho_\lambda, \partial_\lambda\varrho_\lambda]=0$ and Eq. (\ref{LE})
360: reduces to $L_\lambda = \partial_\lambda \log \varrho_\lambda$.
361: \par
362: Finally, upon substituting the above Eqs. in Eq. (\ref{QEst}),
363: we obtain the corresponding optimal quantum estimator
364: \begin{align}
365: O_\lambda =
366: \sum_p \left( \lambda + \frac{\partial_\lambda \varrho_p}{\varrho_p}
367: \right)
368: |\psi_p\rangle\langle \psi_p | +
369: \frac{2}{H(\lambda)} \sum_{n\neq m} \frac{\varrho_n - \varrho_m}{\varrho_n + \varrho_m}
370: \langle \psi_m |\partial_\lambda \psi_n\rangle
371: \: |\psi_m\rangle\langle \psi_n |\:.
372: \end{align}
373: %%%
374: So far we have considered the case of a parameter with a fixed given
375: value. A question arises on whether a bound for estimator variance may
376: be established also for a parameter having an {\em a priori} distribution
377: $z(\lambda)$. The answer is positive and given by the Van Trees inequality
378: \cite{Van67,Gil95} which provides a bound for the average variance
379: $$
380: \overline{\hbox{Var}(\lambda)}
381: = \int\!\! dx\!\! \int\!\! d\lambda \:z(\lambda)\:
382: \left[\hat\lambda(\{x\})-\lambda)\right]^2
383: $$ of any unbiased estimator of the random parameter $\lambda$.
384: Van Trees inequality states that
385: \begin{equation}
386: \overline{\hbox{Var}(\lambda)} \geq \frac1{Z_F}
387: \end{equation}
388: where the generalized Fisher information $Z_F$ is given by
389: \begin{equation}
390: Z_F= \int\!\! dx\!\! \int\!\! d\lambda \: p(x,\lambda) \left[\partial_\lambda
391: \log p(x,\lambda)\right]^2 \:, \label{ZZ}
392: \end{equation}
393: $p(x,\lambda)$ being the joint probability distribution of the outcomes
394: and the parameter of interest. Upon writing the joint distribution
395: as $p(x,\lambda) = p(x|\lambda) z(\lambda)$ Eq. (\ref{ZZ}) may be
396: rewritten as
397: \begin{equation}
398: Z_F= \int\!\! d\lambda\: z(\lambda)\: F (\lambda) + M
399: \int\!\! d\lambda\: z(\lambda) \left[\partial_\lambda
400: \log z(\lambda)\right]^2 \:.
401: \label{ZZ1}
402: \end{equation}
403: Eq. (\ref{ZZ1}) says that the generalized Fisher information is the sum
404: of two terms, the first is simply the average of the Fisher information
405: over the {\em a priori} distribution whereas the second term is the
406: Fisher information of the priori distribution itself. As expecteed, in
407: the asymptotic limit of many measurements the {\em a priori} distribution
408: is no longer relevant. The
409: quantity $Z_F$ is upper bounded by the analogue expression $Z_H$ where
410: the average of the Fisher information is replaced by the average of the
411: QFI $H(\lambda)$ The resulting quantum Van Trees bound may be easily
412: written as
413: \begin{equation}
414: \overline{\hbox{Var}(\lambda)} \geq \frac1{Z_H}\:.
415: \end{equation}
416: %%%%
417: \section{Estimability of a parameter}\label{s:qep}
418: A large signal is easily estimated whereas a quantity with
419: a vanishing value may be inferred only if the corresponding
420: estimator is very {\em precise} {\em i.e} characterized by
421: a small variance. This intuitive statement indicates that
422: in assessing the performances of an estimator and, in turn,
423: the overall estimability of a parameter, the relevant figure
424: of merit is the scaling of the variance with the mean value
425: rather than its absolute value. This feature may be quantified
426: by means of the signal-to-noise ratio (for a single measurement)
427: $$
428: R_\lambda = \frac{\lambda^2}{\hbox{Var}(\lambda)}
429: $$
430: which is larger for better estimators.
431: Using the quantum Cramer-Rao bound one easily
432: derives that the signal-to-noise ratio of any estimator
433: is bounded by the quantity
434: $$
435: R_\lambda \leq Q_\lambda \equiv \lambda^2 H(\lambda)
436: $$
437: which we refer to as the quantum signal-to-noise ratio.
438: We say that a given parameter $\lambda$ is effectively estimable
439: quantum-mechanically when the corresponding $Q_\lambda$ is large.
440: \par
441: Upon taking into account repeated measurements we have that
442: the number of measurements leading to a $99.9 \%$ ($3\sigma$)
443: confidence interval corresponds to a relative
444: error
445: $$
446: \delta^2 = \frac{9 \hbox{Var}(\lambda)}{M\lambda^2} = \frac{9}{M}
447: \frac1Q_\lambda = \frac{9}{M\lambda^2 H(\lambda)}
448: $$
449: Therefore, the number of measurements needed to achieve a $99.9\%$
450: confidence interval with a relative errro $\delta$ scales as
451: $$
452: M_\delta = \frac9{\delta^2} \frac1Q_\lambda
453: $$
454: In other words, a vanishing $Q_\lambda$ implies a diverging number of
455: measurements to achieve a given relative error, whereas a finite value
456: allows estimation with arbitrary precision at finite number of
457: measurements.
458: %%%
459: \section{Examples}\label{s:exa}
460: In this section we provide explicit evaluation of the symmetric
461: logarithmic derivative and the quantum Fisher information for relevant
462: families of quantum states, including sets of pure states and the
463: generic unitary family. We also consider the multiparamer case and
464: the problem of repametrization.
465: \subsection{Unitary families and the pure state model}
466: Let us consider the case where the parameter of interest is
467: the amplitude of a unitary perturbation imposed to a given
468: initial state $\varrho_0$. The family of quantum states we are dealing
469: with may be expressed as $\varrho_\lambda = U_\lambda \varrho_0
470: U^\dag_\lambda$ where $U_\lambda = \exp\{-i\lambda G\}$ is a unitary
471: operator and $G$ is the corresponding Hermitian generator.
472: Upon expanding the unperturbed state in its eigenbasis $\varrho_0 = \sum
473: \varrho_n |\varphi_n\rangle\langle\varphi_n |$ we have $\varrho_\lambda
474: = \sum_n \varrho_n |\psi_n\rangle\langle\psi_n |$ where
475: $|\psi_n\rangle = U_\lambda |\varphi_n\rangle$. As a consequence
476: we have
477: $$\partial_\lambda \varrho_\lambda = i U_\lambda
478: [G,\varrho_0]U^\dag_\lambda\,.$$
479: and the SLD is may be written as
480: $L_\lambda = U_\lambda L_0 U^\dag_\lambda$ where $L_0$ is given by
481: \begin{align}
482: L_0 & = 2 i \sum_{n,m} \frac{\langle\varphi_m| [G,\varrho_0] |
483: \varphi_n\rangle}{\varrho_n+\varrho_m}\,
484: |\varphi_n\rangle\langle\varphi_m | =
485: 2 i \sum_{n\neq m}\langle\varphi_m| G |
486: \varphi_n\rangle \frac{\varrho_n-\varrho_m}{\varrho_n+\varrho_m}\,
487: |\varphi_n\rangle\langle\varphi_m |\,.
488: \end{align}
489: The corresponding quantum Fisher information
490: is independent on the value of parameter and may be written
491: in compact form as
492: $$
493: H
494: = \hbox{Tr}\left[ \varrho_0\, L_0^2 \right]
495: = \hbox{Tr}\left[ \varrho_0\, [L_0,G] \right]
496: = \hbox{Tr}\left[ L_0\, [G,\varrho_0] \right]
497: = \hbox{Tr}\left[ G\, [\varrho_0, L_0] \right]
498: $$
499: or, more explicitly, as
500: $$
501: H = 2 \sum_{n\neq m} \sigma_{nm} G^2_{nm}
502: $$
503: where the elements $\sigma_{nm}$ are given in Eq. (\ref{sg1}), or
504: equivalently (\ref{sg2}), and $G_{nm}=\langle\varphi_n | G |
505: \varphi_m\rangle=\langle\psi_n | G |
506: \psi_m\rangle$ denote the matrix element
507: of the generator $G$ in either the eigenbasis of $\varrho_0$
508: or $\varrho_\lambda$.
509: \par
510: For a generic family of pure states we have $\varrho_\lambda =
511: |\psi_\lambda\rangle\langle\psi_\lambda |$. Since $\varrho_\lambda^2 =
512: \varrho_\lambda$ we have $\partial_\lambda\varrho_\lambda
513: = \partial_\lambda \varrho_\lambda\,\varrho_\lambda
514: + \varrho_\lambda \partial_\lambda \varrho_\lambda$ and thus
515: $L_\lambda = 2 \partial_\lambda \varrho_\lambda
516: =|\psi_\lambda\rangle\langle\partial_\lambda\psi_\lambda |
517: +|\partial_\lambda\psi_\lambda\rangle\langle\psi_\lambda |
518: $.
519: Finally we have
520: \begin{equation}\label{pure}
521: H(\lambda) = 4
522: \left[
523: \langle\partial_\lambda\psi_\lambda
524: |\partial_\lambda\psi_\lambda\rangle
525: + \left( \langle\partial_\lambda\psi_\lambda |
526: \psi_\lambda\rangle
527: \right)^2
528: \right]
529: \end{equation}
530: \par
531: For a unitary family of pure states $|\psi_\lambda\rangle = U_\lambda
532: |\psi_0\rangle$ we have
533: \begin{align}
534: |\partial_\lambda \psi_\lambda\rangle & = - i G
535: U_\lambda |\psi_0\rangle = -i G |\psi_\lambda\rangle
536: \:,\nonumber \\
537: \langle\partial_\lambda\psi_\lambda
538: |\partial_\lambda\psi_\lambda\rangle & =
539: \langle\psi_0| G^2 | \psi_0\rangle
540: \:, \nonumber \\
541: \langle\partial_\lambda\psi_\lambda |
542: \psi_\lambda\rangle & = -i
543: \langle\psi_0| G | \psi_0\rangle
544: \nonumber\:.
545: \end{align}
546: The quantum Fisher information thus reduces to the simple form
547: \begin{equation}
548: H = 4 \langle\psi_0| \Delta G^2 | \psi_0\rangle
549: \label{HUnPure}
550: \end{equation}
551: which is independent on $\lambda$ and proportional to the fluctuations
552: of the generator on the unperturbed state. Using Eq. (\ref{HUnPure})
553: the quantum Cramer-Rao bound in (\ref{QCR}) rewrites in the appealing
554: form \cite{mac06}
555: \begin{equation}
556: \hbox{Var}(\lambda) \langle\Delta G^2 \rangle \geq \frac1{4M}\,,
557: \label{parUN}
558: \end{equation}
559: which represents a parameter-based uncertainty relation which applies
560: also when the shift parameter $\lambda$ in the unitary $U_\lambda =
561: e^{- i \lambda G}$ {\em does not} correspond to the observable canonically
562: conjugate to $G$.
563: When the unperturbed state is not pure the QFI may be written as
564: \begin{align}
565: H & = 4\,\Tr\left[\Delta G^2 \varrho_0\right]
566: + 4 \sum_n \varrho_n \langle \varphi_n | \langle G\rangle ^2 - 2
567: G K^{(n)} G | \varphi_n\rangle
568: \label{HUnMix} \\
569: K^{(n)} &= \sum_{m} \frac{\varrho_m}{\varrho_n+\varrho_m}
570: |\varphi_m\rangle\langle\varphi_m | \stackrel{\varrho_0 \rightarrow
571: |\varphi_0\rangle\langle\varphi_0 |}{\longrightarrow}
572: \frac12|\varphi_0\rangle\langle\varphi_0 |
573: \end{align}
574: and Eq. (\ref{parUN}) becomes
575: \begin{align}
576: \hbox{Var}(\lambda) \langle\Delta G^2 \rangle \geq \frac1{4M}
577: \left[1+ \sum_n \varrho_n \langle \varphi_n | \langle G\rangle ^2 -2
578: G K^{(n)} G | \varphi_n\rangle
579: \right]^{-1}\:. \label{parUNm}
580: \end{align}
581: The second term in Eqs. (\ref{HUnMix}) and (\ref{parUNm}) thus
582: represents the {\em classical} contribution to uncertainty due to
583: the mixing of the initial signal.
584: \par
585: As we have seen, for unitary families of quantum states the
586: QFI is independent on the value of the parameter. As a consequence
587: the quantum signal-to-noise ratio
588: $Q_\lambda$ vanishes for vanishing $\lambda$ and thus the number of
589: measurements needed to achieve a relative error $\delta$ diverges as
590: $M_\delta \sim (\delta \lambda)^{-2}$.
591: %%
592: \subsection{Quantum operations}
593: Let us now consider a family of quantum states obtained from a given
594: inital state $\varrho_0$ by the action of a generic quantum operation
595: $\varrho_\lambda = {\cal E}_\lambda (\varrho_0) = \sum_k M_{k\lambda}
596: \varrho_0 M^\dag_{k\lambda}$. Upon writing the initial and the evolved
597: states in terms of their eigenbasis
598: $\varrho_0 = \sum_s \varrho_{0s} |\varphi_s\rangle\langle\varphi_s |$,
599: $\varrho_\lambda = \sum_s \varrho_{n} |\psi_n\rangle\langle\psi_n |$
600: we may evaluate the SLD and the
601: quantum Fisher information using Eqs. (\ref{LL}) and (\ref{HH}) where
602: \begin{align}
603: \varrho_n =& \sum_{ks} \varrho_{0s} \left|
604: \langle\psi_n|M_{k\lambda}|\varphi_s\rangle\right|^2 \\
605: \langle\psi_m|\partial_\lambda\varrho_\lambda |\psi_n\rangle =& \sum_{ks} \varrho_{0s}
606: \left[
607: \langle\psi_m|\partial_\lambda M_{k\lambda}|\varphi_s\rangle
608: \langle\varphi_s|M_{k\lambda}^\dag|\psi_n\rangle
609: \right. \nonumber \\ &+ \left.
610: %+
611: \langle\psi_m|M_{k\lambda}|\varphi_s\rangle
612: \langle\varphi_s|\partial_\lambda M_{k\lambda}^\dag|\psi_n\rangle
613: \right] \:.
614: \end{align}
615: For a pure state at the input $\varrho_0=|\psi_0\rangle\langle\psi_0 |$
616: the above equation rewrites without the sum over $s$.
617: %%%
618: \subsection{Multiparametric models and reparametrization}
619: In situations where more than a parameter is involved the family
620: of quantum states $\varrho_{\boldsymbol\lambda}$ depends on a set
621: ${\boldsymbol\lambda}=\{ \lambda_\mu \}$, $\mu=1,\dots,N$.
622: In this cases the relevant object in the estimation problem
623: is given by the so-called quantum Fisher information matrix,
624: whose elements are defined as
625: \begin{align}
626: {\boldsymbol H}({\boldsymbol\lambda})_{\mu\nu} =&
627: \Tr\left[\varrho_{{\boldsymbol\lambda}} \frac{L_\mu L_\nu +L_\nu
628: L_\mu}{2}\right]
629: = \Tr[\partial_\nu \varrho_{\boldsymbol \lambda} L_\mu]
630: = \Tr[\partial_\mu \varrho_{\boldsymbol \lambda} L_\nu]
631: \nonumber \\= &
632: \sum_n \frac{(\partial_\mu \varrho_n) (\partial_\nu \varrho_n)}{\varrho_n}
633: + \sum_{n\neq m}
634: \frac{(\varrho_n-\varrho_m)^2}{\varrho_n + \varrho_m} \times \nonumber \\
635: &\times \left[\langle \psi_n|\partial_\mu \psi_m\rangle
636: \langle \partial_\nu \psi_m| \psi_n \rangle +
637: \langle \psi_n|\partial_\nu \psi_m\rangle
638: \langle \partial_\mu \psi_m| \psi_n \rangle\right]
639: \end{align}
640: where $L_\mu$ is the SLD corresponding to the parameter $\lambda_\mu$.
641: The Cramer-Rao theorem for multiparameter estimation says that the inverse
642: of the Fisher matrix provides a lower bound
643: on the covariance matrix $\hbox{Cov}[\boldsymbol\gamma]_{ij}=
644: \langle \lambda_i \lambda_j \rangle -
645: \langle \lambda_i\rangle\langle\lambda_j\rangle$, {\em i.e}
646: $$
647: \hbox{Cov}[\boldsymbol\gamma] \geq \frac1M
648: {\boldsymbol H}({\boldsymbol\lambda})^{-1}
649: $$
650: The above relation is a matric inequality and the corresponding
651: bound may not be achievable achievable in a multiparameter estimation.
652: On the other hand, the diagonal elements of the inverse Fisher matrix provide
653: achievable bounds for the variances of single parameter estimators
654: {\em at fixed value} of the others, in formula
655: \begin{align}
656: {\mathrm{Var}}(\lambda_\mu) = \gamma_{\mu\mu} \geq \frac1M ({\boldsymbol
657: H}^{-1})_{\mu\mu}.
658: \label{eq:QCRMulti}
659: \end{align}
660: Of course, for a diagonal Fisher matrix ${\mathrm{Var}}(\lambda_\mu)
661: \geq 1/{\boldsymbol H}_{\mu\mu}$.
662: \par
663: Let us now suppose that the quantity of interest $g$ is a known function
664: $g(\boldsymbol\lambda)$ of the parameters used to label the family of states.
665: In this case we need to reparametrize the familiy with a new set of parameters
666: $\widetilde{{\boldsymbol\lambda}}= \{\widetilde{\lambda}_j=
667: \widetilde{\lambda}_j({\boldsymbol\lambda})$ that includes the
668: quantity of interest, {\em e.g} $\widetilde{\lambda}_1 \equiv
669: g(\boldsymbol\lambda)$. Since
670: $\widetilde{\partial}_\mu = \sum_\nu B_{\mu\nu} \partial_\nu$
671: where
672: $B_{\mu\nu} = \partial\lambda_\nu/\partial \widetilde{\lambda}_\mu$
673: it is easy to prove that
674: $$\widetilde{L}_\mu = \sum_\nu B_{\mu\nu}L_\nu
675: \qquad\widetilde{\boldsymbol H} = {\boldsymbol B} {\boldsymbol H}
676: {\boldsymbol B}^T \:.$$
677: The ultimate precision on the estimation of $g$ at fixed values
678: of the other parameters is thus given by
679: $$
680: \hbox{Var}(g) \geq \frac{1}{M} (\widetilde{\boldsymbol H}^{-1})_{11}
681: $$
682: %%%
683: \section{Geometry of quantum estimation}\label{s:geo}
684: The estimability of a set of parameters labelling the family
685: of quantum states $\{\varrho_{\boldsymbol\lambda}\}$ is naturally
686: related to the distinguishability of the states within the quantum
687: statistical model {\em i.e.} with the notions of distance.
688: On the manifold of quantum states, however, different distances
689: may be defined and a question arises on which of them captures
690: the notion of estimation measure. As it can be easily proved
691: it turns out that the Bures distance
692: \cite{bur69,uhl76,jos94,hub92,sla96,hal98,dit99}
693: is the proper quantity to
694: be taken into account. This may be seen as follows.
695: The Bures distance between two density matrices is defined as
696: $D_B^2 (\varrho,\sigma)= 2[1-\sqrt{F(\varrho,\sigma)}]$ where
697: $F(\varrho,\sigma) = \left(
698: \Tr\left[\sqrt{\sqrt{\varrho}\sigma\sqrt{\varrho}}\right]\right)^2$
699: is the fidelity. The Bures metric $g_{\mu\nu}$ is obtained upon
700: considering the distance for two states obtained by an infinitesimal
701: change in the value of the parameter
702: $$
703: d^2_{B} = D^2_B (\varrho_{\boldsymbol \lambda}, \varrho_{\boldsymbol
704: \lambda + d \boldsymbol\lambda})
705: = g_{\mu\nu} d\lambda_\mu d\lambda_\nu\:.
706: $$
707: By explicitly evaluating the Bures distance \cite{som03} one arrives at
708: $g_{\mu\nu}= \frac14 {\boldsymbol H}_{\mu\nu} (\boldsymbol\lambda)$, {\em i.e.}
709: the Bures metric is simply proportional to the QFI, which itself
710: is symmetric, real and positive semidefinite, {\em i.e.}
711: represents a metric for the manifold underlying the quantum
712: statistical model.
713: Indeed, a large QFI for a given $\lambda$ implies that the quantum
714: states $\varrho_{\boldsymbol \lambda}$ and $\varrho_{\boldsymbol
715: \lambda + d \boldsymbol\lambda}$ should be statistically distinguishable
716: more effectively than the analogue states for a value $\lambda$
717: corresponding to smaller QFI. In other words, one confirms
718: the intuitive picture in which optimal estimability (that
719: is, a diverging QFI) corresponds to quantum states that are sent
720: far apart upon infinitesimal variations of the parameters.
721: \begin{figure}[h]
722: \centerline{\includegraphics[width=0.9\textwidth]{geo.eps}}
723: \caption{Geometry of quantum estimation}
724: \label{f:geo}
725: \end{figure} \\
726: The structures described above are pictorially described in Fig.
727: \ref{f:geo}. The idea is that any measurement aimed to estimate
728: the parameters ${\boldsymbol \lambda}$
729: turns the set of parameters into a statistical differential
730: manifold endowed with the Fisher metric $\boldsymbol F_{\mu\nu}
731: ({\boldsymbol\lambda})$. On the other hand, when the parameters
732: are mapped into the manifold of quantum states the statistical
733: distance is expressed in terms of the Bures metric. The connection
734: between the two constructions is provided by the optimization of
735: the estimation procedure over quantum measurements, which
736: shows that the Quantum Fisher metric $\boldsymbol H_{\mu\nu}
737: (\boldsymbol \lambda)$ is the bound to $\boldsymbol F_{\mu\nu}
738: ({\boldsymbol\lambda})$ and coincides, apart from a factor four,
739: with the Bures metric.
740: %%%
741: \section{Conclusions and outlooks}\label{s:out}
742: As a matter of fact, there are many quantities of interest that do not
743: correspond to any quantum observable. Among these, we mention the amount
744: of entanglement and the purity of a quantum state and the coupling
745: constant of an interaction Hamiltonian or a quantum operation. In these
746: situations, the values of the quantity of interest can be indirectly
747: inferred by an estimation procedure, {\em i.e.} by measuring one or more
748: proper observables, a quantum estimator, and then manipulating the
749: outcomes by a suitable classical processing.
750: \par
751: In this paper, upon exploiting the geometric theory of quantum
752: estimation, we have described a general method to solve a quantum
753: statistical model, {\em i.e} to find the optimal quantum estimator and
754: to evaluate the corresponding bounds to precision. To this aim we used
755: the quantum Cramer-Rao theorem and the explicit evaluation of the
756: quantum Fisher information matrix. We have derived the explicit form of
757: the optimal observable in terms of the symmetric logarithmic derivative
758: and evaluated the corresponding bounds to precision, which represent the
759: ultimate bound posed by quantum mechanics to the precision of parameter
760: estimation. For unitary families of quantum states the bounds may
761: expressed in the form of a parameter-based uncertainty relation.
762: \par
763: The analysis reported in this paper has a fundamental interest and
764: represents a relevant tool in the design of realistic
765: quantum information protocols. The approach here outlined is currently
766: being applied to the estimation of entanglement \cite{ee08} and the
767: coupling constant of an interaction Hamiltonian \cite{ZP07,MK08}.
768: %%%
769: \section*{Acknowledgments}
770: The author thanks Paolo Giorda, Alex Monras, Paolo Zanardi, Marco
771: Genoni, Michael Korbman, Carmen Invernizzi and Stefano Olivares for
772: stimulating discussions.
773: %%%%%%%%%%%%%%%%%%
774: \begin{thebibliography}{99}
775: \bibitem{QET} C. W. Helstrom, {\em Quantum Detection and Estimation
776: Theory} (Academic Press, New York, 1976); A.S. Holevo, {\em Statistical
777: Structure of Quantum Theory}, Lect. Not. Phys {\bf 61}, (Springer, Berlin,
778: 2001).
779: \bibitem{hel67} C. W. Helstrom, Phys. Lett. A {\bf 25}, 1012 (1967).
780: \bibitem{yl73} H. P. Yuen, M. Lax, IEEE Trans. Inf. Th. {\bf 19}, 740
781: (1973).
782: \bibitem{hk74} C. W. Helstrom, R. S. Kennedy, IEEE Trans. Inf. Th. {\bf
783: 20}, 16 (1974).
784: \bibitem{BC94} S. Braunstein and C. Caves, Phys. Rev. Lett. {\bf 72}, 3439 (1994).
785: \bibitem{BC96} S. Braunstein, C. Caves, and G. Milburn, Ann. Phys. {\bf 247}, 135 (1996).
786: \bibitem{gil00} O. E. Barndorff-Nielsen, R. D. Gill, J. Phys. A {\bf
787: 33}, 4481 (2000).
788: \bibitem{Hol79} A. S. Holevo, Rep. Math. Phys. {\bf 16}, 385 (1979).
789: \bibitem{Dar98} M. D'Ariano et al., Phys. Lett. A {\bf 248}, 103 (1998).
790: \bibitem{Hel74} C. W. Helstrom, Found. Phys. {\bf 4}, 453 (1974).
791: \bibitem{Mil94} G. J. Milburn et al., Phys. Rev. A {\bf 50}, 801 (1994).
792: \bibitem{Chi06} G. Chiribella et al., Phys. Rev. A {\bf 73}, 062103 (2006).
793: \bibitem{per01} G. M. D'Ariano, M. G. A. Paris, P. Perinotti,
794: J. Opt. B {\bf 3}, 337 (2001).
795: \bibitem{Mon06} A. Monras, Phys. Rev. A {\bf 73}, 033821 (2006).
796: \bibitem{Sar06} M. Sarovar, G. J. Milburn, J. Phys. A {\bf 39}, 8487
797: (2006).
798: \bibitem{Hot06} M. Hotta et al., Phys. Rev. A {\bf 72}, 052334 (2005);
799: J. Phys. A {\bf 39} (2006).
800: \bibitem{Fuj01} A. Fujiwara, Phys. Rev. A {\bf 63}, 042304 (2001);
801: A. Fujiwara, H. Imai, J. Phys. A {\bf 36}, 8093 (2003).
802: \bibitem{Zhe06} J. Zhenfeng et al., preprint LANL quant-ph/0610060
803: \bibitem{Mon07} {A. Monras, M. G. A. Paris} Phys. Rev. Lett. {\bf 98}, 160401 (2007).
804: \bibitem{Dau06} V. D'Auria et al.,J. Phys. B {\bf 39}, 1187 (2006).
805: \bibitem{Gra87} P. Grangier et al., Phys. Rev. Lett. {\bf 59}, 2153 (1987).
806: \bibitem{Pol92} E. S. Polzik et al., Phys. Rev. Lett. {\bf 68}, 3020 (1992).
807: \bibitem{fri07} B. R. Frieden, Opt. Comm. {\bf 271}, 7 (2007).
808: \bibitem{Boi08} S. Boixo, A. Monras, Phys. Rev. Lett. {\bf 100}, 100503 (2008).
809: \bibitem{ZP07} P. Zanardi, M. G A Paris, arXiv:0708.1089
810: \bibitem{Cra46} H. Cramer, {\em Mathematical methods of statistics},
811: (Princeton University Press, 1946).
812: \bibitem{mac06} L. Maccone, Phys. Rev. A {\bf 73}, 042307, (2006).
813: \bibitem{Van67} H. L. Van Trees, {\em Detection, Estimation, modulation
814: theory} (Wiley, New York, 1967).
815: \bibitem{Gil95} R. D. Gill. B. Y. Levit. Bernoulli {\bf 1}, 59 (1995)
816: \bibitem{bur69} D. J. C. Bures, Trans. Am. Math. Phys. {\bf 135}, 199
817: (1969).
818: \bibitem{uhl76} A. Uhlmann, Rep. Math. Phys. {\bf 9}, 273 (1976).
819: \bibitem{jos94} R. Josza, J. Mod. Opt. {\bf 41}, 2315 (1994).
820: \bibitem{hub92} M. H\"ubner, Phys. Lett. A {\bf 163}, 239 (1992).
821: \bibitem{sla96} P. B. Slater, J. Phys. A {\bf 29}, L271 (1996); Phys.
822: Lett. A {\bf 244}, 35 (1998).
823: \bibitem{hal98} M. J. W. Hall. Phys. Lett. A {\bf 242}, 123 (1998).
824: \bibitem{dit99} J. Dittmann, J. Phys. A {\bf 32}, 2663 (1999).
825: \bibitem{som03} H-J. Sommers et al., J. Phys. A {\bf 36}, 10083 (2003).
826: \bibitem{ee08} M. G. Genoni, P. Giorda, M. G. A. Paris,
827: preprint arXiv:0804.1705
828: \bibitem{MK08} C. Invernizzi, M. Korbman, L. Campos, M. G. A. Paris,
829: preprint arXiv:0807.3213
830: \end{thebibliography}
831: %%%%%%%%%%%%%%%%%%
832: \end{document}
833: