0804.1705/EE.tex
1: \documentclass[pra,twocolumn,superscriptaddress,showpacs]{revtex4}
2: \usepackage{amsmath,amssymb,amsfonts,bbm,graphicx}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \newcommand{\ket}[1]{\vert #1 \rangle} \newcommand{\bra}[1]{\langle #1 \vert}
5: \newcommand{\dket}[1]{\vert #1 \rangle\rangle}
6: \newcommand{\dbra}[1]{\langle\langle #1 \vert}
7: \newcommand{\braket}[2]{\langle #1 \vert #2 \rangle}
8: \newcommand{\dbraket}[2]{\langle\langle #1 \vert #2 \rangle\rangle}
9: \newcommand{\ketbra}[2]{\vert #1 \rangle \! \langle #2 \vert}
10: \newcommand{\sandwich}[3]{\left \langle #1 | #2 | #3 \right\rangle}
11: \newcommand{\overlap}[2]{\left \langle #1 | #2 \right\rangle}
12: \newcommand{\average}[1]{\left \langle #1  \right\rangle}
13: \newcommand{\be}{\begin{equation}}
14: \newcommand{\ee}{\end{equation}}
15: \newcommand{\bae}{\begin{eqnarray}} \newcommand{\eae}{\end{eqnarray}}
16: \newcommand{\nnum}{\nonumber \\} \newcommand{\media}[1]{\langle #1 \rangle}
17: \newcommand{\real}[1]{\Re\mbox{e}[#1]} \newcommand{\immag}[1]{\Im\mbox{m}[#1]}
18: \newcommand{\calpha}{\alpha^*} \newcommand{\cbeta}{\beta^*} \newcommand{\cxi}{\xi^*}
19: \newcommand{\czeta}{\zeta^*} \newcommand{\cv}{v^*} \newcommand{\cw}{w^*}
20: \newcommand{\wtA}{\widetilde{A}} \newcommand{\wtB}{\widetilde{B}}
21: \newcommand{\bmsigma}{\boldsymbol \sigma} \newcommand{\bmSigma}{\boldsymbol \Sigma}
22: \newcommand{\bmX}{\boldsymbol X} \newcommand{\bmA}{\boldsymbol A}
23: \newcommand{\bmB}{\boldsymbol B} \newcommand{\bmC}{\boldsymbol C}
24: \newcommand{\bmS}{\boldsymbol S} \newcommand{\bmalpha}{\boldsymbol \alpha}
25: \newcommand{\bmLambda}{\boldsymbol \Lambda}
26: \newcommand{\bmlambda}{\boldsymbol \lambda}
27: \newcommand{\bbGamma}{{\rm I}\!\Gamma}
28: \newcommand{\calN}{{\cal N}} \newcommand{\calA}{{\cal A}}
29: \newcommand{\calB}{{\cal B}} \newcommand{\calC}{{\cal C}}
30: \newcommand{\calH}{{\cal H}} \newcommand{\sfx}{{\sf x}} \newcommand{\sfy}{{\sf y}}
31: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
32: \def\({\left(} \def\){\right)} \def\[{\left[} \def\]{\right]}
33: \def\v{\mathbf} \def\ni{\nu} \def\mi{\mu}
34: \def\Tr{\hbox{Tr}} \def\sigmaCM{\boldsymbol{\sigma}}
35: \def\X{\overline{\boldsymbol{X}}} \def\lambdaB{\boldsymbol{\lambda}}
36: \def\D{\mathcal{D}} \def\F{\mathcal{F}} 
37: \def\I{\mathcal{I}}
38: \def\elu{\epsilon_{\hbox{\tiny U}}}
39: \def\eln{\epsilon_{\hbox{\tiny N}}}
40: \def\eps{\epsilon}
41: \def\E{\epsilon}
42: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
43: \begin{document}
44: \title{Optimal estimation of entanglement}
45: \author{Marco G. Genoni}
46: \affiliation{Dipartimento di Fisica, Universit\`a di Milano, I-20133 Milano, Italia}
47: \affiliation{CNISM, UdR Milano, I-20133 Milano, Italia}
48: \author{Paolo Giorda}
49: \affiliation{I.S.I. Foundation, I-10133 Torino, Italia}
50: \author{Matteo G. A. Paris}
51: \affiliation{Dipartimento di Fisica, Universit\`a di Milano, I-20133 Milano, Italia}
52: \affiliation{CNISM, UdR Milano, I-20133 Milano, Italia}
53: \affiliation{I.S.I. Foundation, I-10133 Torino, Italia}
54: \date{\today}
55: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
56: \begin{abstract}
57: Entanglement does not correspond to any observable and its evaluation
58: always corresponds to an estimation procedure where the amount of
59: entanglement is inferred from the measurements of one or more proper
60: observables.  Here we address optimal estimation of entanglement in the
61: framework of local quantum estimation theory and derive the optimal
62: observable in terms of the symmetric logarithmic derivative. We evaluate
63: the quantum Fisher information and, in turn, the ultimate bound to
64: precision for several families of bipartite states, either for qubits or
65: continuous variable systems, and for different measures of entanglement.
66: We found that for discrete variables, entanglement may be efficiently
67: estimated when it is large, whereas the estimation of weakly entangled
68: states is an inherently inefficient procedure. For continuous variable
69: Gaussian systems the effectiveness of entanglement estimation strongly
70: depends on the chosen entanglement measure.  Our analysis makes an
71: important point of principle and may be relevant in the design of
72: quantum information protocols based on the entanglement content of
73: quantum states.  
74: \end{abstract}
75: \pacs{03.67.Mn, 03.65.Ta}
76: \maketitle
77: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
78: \section{Introduction}\label{s:intro}
79: Entanglement is perhaps the most distinctive feature of quantum
80: mechanics, and definitely the most relevant resource for quantum
81: information processing \cite{EntR}. Indeed, quantification of
82: entanglement and schemes for its measurement have been the subject of
83: extensive efforts in the last decade
84: \cite{Aud06,Che03,Eis07,Guh04,Guh07,
85: Sun07,Lou06,Wal06,Ren04,Nav08,Hor03,Aci00,Hor02,Dur01,Bar03,Guh03,Dar03,
86: Pla02,Pit03}. Indeed, the entanglement content of a quantum state is a
87: crucial piece of information in the design of quantum information
88: protocols, and a question naturally arises on whether quantum mechanics
89: itself poses limits to the precision of its determination.  As a matter
90: of fact, any quantitative measure of entanglement corresponds to a
91: nonlinear function of the density operator and thus cannot be associated
92: to a quantum observable.  As a consequence, any procedure aimed to
93: evaluate the amount of entanglement of a quantum state is ultimately a
94: parameter estimation problem, where the value of entanglement is
95: indirectly inferred from the measurement of one or more proper
96: observables. An optimization problem thus naturally arises, which may be
97: properly addressed in the framework of quantum estimation theory (QET)
98: \cite{QET}, which provides analytical tools to find the optimal
99: measurement according to some given criterion. 
100: \par
101: Our aim is indeed to evaluate the ultimate bounds to precision posed by
102: quantum mechanics, {\em i.e} the smallest value of the
103: entanglement that can be discriminated, and to determine the optimal
104: measurements achieving those bounds. Being entanglement an intrinsic
105: property of quantum states we adopt local quantum estimation theory,
106: where the optimal estimators are those maximizing the Fisher information
107: \cite{Hel67,BC94,BC96} and in turn minimizing the variance at fixed
108: value of entanglement.  Local QET provides any family of quantum states
109: with a geometric structure based on distinguishability and accounts for
110: the optimal measurement that can be performed on the quantum system as
111: well as the optimal data processing of the outcomes of the measurement.
112: \par
113: Local QET has been applied to the estimation of quantum optical phase
114: \cite{Mon06} as well as to estimation problems involving non unitary processes
115: in open quantum systems \cite{Sar06}, either in finite dimensional
116: systems \cite{Hot06} or continuous variable ones \cite{Mon07}. This
117: includes the optimal estimation of the noise parameter of depolarizing
118: \cite{Fuj01} or amplitude-damping \cite{Zhe06,Mon07} channels.
119: Recently, the geometric structure induced by the Fisher
120: information itself has been exploited to give a quantitative operational
121: interpretation for multipartite entanglement \cite{Boi08} and to assess
122: quantum criticality as a resource for quantum estimation \cite{ZP07}. 
123: \par
124: In this paper we systematically apply local QET to the problem of
125: efficiently estimate the amount of entanglement of a quantum state.  We
126: consider several families of bipartite states, either for qubits or
127: Gaussian states, and evaluate the symmetric logarithmic derivative to
128: estimate entanglement through different measures, {\em e.g.} negativity
129: or linear entropy.  Then we explicitly calculate the quantum Fisher
130: information and derive the ultimate bounds to the precision of
131: estimation. Overall, we found that, both for qubits and Gaussian states,
132: entanglement may be efficiently estimated when it is large. On the other
133: hand the estimation of a small amount of entanglement for qubits is an
134: inherently inefficient procedure, {\em i.e} the signal-to-noise ratio is
135: vanishing for vanishing entanglement, whereas for Gaussian states it
136: depends on the chosen measure of entanglement.  We also found that the
137: presence of other free parameters besides entanglement does not
138: generally influence the estimation precision, thus preventing the
139: possibility of further optimizing the estimation procedure.
140: \par
141: The paper is structured as follows: in the next Section we give some
142: basic elements of quantum estimation theory and introduce the quantum 
143: signal-to-noise to assess the estimability of a parameter. In Section
144: \ref{s:qubits} we analyze the
145: estimation of entanglement by means of negativity \cite{NVW}
146: and linear entropy for the family of pure two-qubit states as well as
147: for two families of entangled mixtures. In Section \ref{s:PPT} we
148: address a family of PPT bound-entangled states \cite{HorodPPT} for
149: two-qutrit systems as an example of states with an inherently small
150: amount of entanglement.  In Section \ref{s:CV} we address entanglement
151: estimation for Gaussian states, either pure states (twin-beams) or
152: entangled mixtures.  Section \ref{s:conclusions} closes the paper with
153: some concluding remarks. 
154: %%%
155: \section{Quantum estimation theory}\label{s:qet}
156: In an estimation problem one tries to infer the value the of a parameter
157: $\lambda$ by measuring a different quantity $X$, which is somehow
158: related to $\lambda$. An estimator $\hat\lambda \equiv\lambda 
159: (x_1, x_2,..)$ for $\lambda$ is a real function of the outcomes of 
160: measurement. The Cramer-Rao theorem
161: \cite{Cra46} establishes a lower bound for the variance
162: ${\mathrm{Var}}(\lambda)$ of any unbiased estimator 
163: \begin{align}
164: {\mathrm{Var}}(\lambda) \geq \frac{1}{M F(\lambda)} \label{eq:CramerRao}
165: \end{align}
166: in terms of the number of measurements $M$ and the so-called 
167: Fisher Information (FI) 
168: \begin{align}
169: F(\lambda) &=\sum_x  p(x\vert \lambda)\, [\partial_\lambda \ln p(x\vert
170: \lambda)]^2 
171: = \sum_x \frac{[\partial_\lambda  p(x\vert
172: \lambda)]^2}{p(x\vert \lambda)}  , \label{eq:ClassicalFisher}
173: \end{align}
174: where $p(x\vert\lambda)$ denotes the conditional probability of
175: obtaining the value $x$ when the parameter has the value $\lambda$.
176: \par
177: In quantum mechanics, according to the Born rule we have
178: $p(x\vert \lambda) = \Tr[E_x \varrho_\lambda]$ where $\left\{E_x\right\}$
179: are the elements of a positive operator-valued measure (POVM) and
180: $\varrho_\lambda$ is the density operator parametrized by the quantity
181: we want to estimate. Introducing the Symmetric Logarithmic
182: Derivative (SLD) $L_\lambda$ as the operator satisfying
183: the equation \begin{align}
184: \frac{L_\lambda \varrho_\lambda + \varrho_\lambda L_\lambda}{2} =
185: \frac{\partial \varrho_\lambda}{\partial \lambda} \label{eq:SLD}
186: \end{align}
187: we have that
188: $\partial_\lambda p(x\vert\lambda) = \Tr[ \partial_\lambda\varrho_\lambda E_x]
189: = \hbox{Re}( \Tr[\varrho_\lambda L_\lambda E_x])$, and 
190: the Fisher Information in Eq. (\ref{eq:ClassicalFisher}) may 
191: be rewritten as
192: \begin{align}
193: F(\lambda) = \sum_x \frac{\hbox{Re}(\Tr[\varrho_\lambda L_\lambda E_x])^2}
194: {\Tr[\varrho_\lambda E_x]} .\label{eq:CQFisher} 
195: \end{align}
196: Starting from Eq. (\ref{eq:CQFisher}) one may prove that $F(\lambda)$
197: is upper bounded by the so-called \emph{Quantum Fisher Information}
198: (QFI) \cite{BC94, BC96} 
199: \begin{align}
200: F(\lambda) \leq H(\lambda) \equiv \Tr[\varrho_\lambda L_\lambda^2]\:,
201: \label{eq:QuantumFisher} \end{align}
202: and, in turn, that $\hbox{Var} (\lambda) \geq [M H(\lambda)]^{-1}$
203: represents the quantum version of the Cramer-Rao theorem, {\em i.e.}
204: the ultimate bound to precision for any quantum measurement aimed 
205: to estimate the parameter $\lambda$. The SLD itself provides an optimal 
206: measurement, that is, using a measurement described by the projectors 
207: over the eigenbasis of $L_\lambda$ we saturate inequality (\ref{eq:QuantumFisher}). 
208: \par
209: Upon diagonalizing $\varrho_{\lambda} = \sum_k p_k 
210: |\psi_k\rangle\langle\psi_k|$ and using  Eqs. (\ref{eq:SLD}) 
211: and (\ref{eq:QuantumFisher}) we obtain
212: \begin{align}\label{defs}
213: L_{\lambda} & = 2 \sum_{n,m}
214: \frac{\langle\psi_n | \partial_\lambda \varrho_\lambda|\psi_m\rangle }
215: {p_n + p_m} |\psi_n\rangle\langle\psi_m| \\ 
216: H (\lambda ) & = \sum_n \frac{(\partial_\lambda p_n)^2}{p_n} + 2\sum_{n,m}
217: \frac{(p_n-p_m)^2}{p_n+p_m} |\langle\psi_n|\partial_\lambda\psi_m\rangle|^2
218: \nonumber\, 
219: \end{align}
220: which, for a family of pure states $\varrho_\lambda=
221: |\psi_\lambda\rangle\langle\psi_\lambda|$ reduces to 
222: \begin{align}\label{defsp}
223: L_\lambda & =2\,\partial_\lambda |\psi_\lambda\rangle\langle\psi_\lambda|
224: \\ H(\lambda) & = 4 \left[
225: \langle\partial_\lambda\psi_\lambda
226: |\partial_\lambda\psi_\lambda\rangle
227: + \langle\partial_\lambda\psi_\lambda |
228: \psi_\lambda\rangle^2
229: \right]
230: \nonumber\,.
231: \end{align}
232: When more than a parameter are involved we have quantum states
233: $\varrho_{\Lambda}$ depending on a set of $N$ parameters $\Lambda=\{
234: \lambda_j \}$, $j=1,\dots,N$. In this case the geometry of the
235: estimation problem is contained in the QFI matrix, whose elements are
236: defined as ${\boldsymbol H}(\Lambda)_{ij} =
237: \frac{1}{2}\Tr[\varrho_{\Lambda} \{ L_i,L_j \}]$ where $L_i$ is the SLD
238: that corresponds to the parameter $\lambda_i$ and $\{A,B\}=AB + BA$
239: denotes anti-commutator. The explicit formula for the QI matrix reads as
240: follows
241: \begin{align}
242: &H(\Lambda)_{ij} = \sum_n \frac{(\partial_i p_n) (\partial_j p_n)}{p_n} + 
243: \sum_{n,m}
244: \frac{(p_n-p_m)^2}{p_n + p_m} \times \nonumber \\
245: &\times (\langle \psi_n|\partial_i \psi_m\rangle
246: \langle \partial_j \psi_m| \psi_n \rangle +
247: \langle \psi_n|\partial_j \psi_m\rangle
248: \langle \partial_i \psi_m| \psi_n \rangle )
249: \end{align}
250: The inverse of the Fisher matrix provides a lower bound ${\boldsymbol \gamma} 
251: \geq H^{-1}$, on the covariance matrix ${\boldsymbol \gamma}_{ij}= 
252: \langle \lambda_i \lambda_j \rangle - \langle \lambda_i\rangle\langle
253: \lambda_j\rangle$ of global estimators of $\Lambda$, which is not 
254: generally achievable. On the other
255: hand, the diagonal elements of the inverse Fisher matrix provide
256: achievable bounds for the variances of single parameter estimators
257: (at fixed value of the others)
258: \begin{align}
259: {\mathrm{Var}}(\lambda_i) = {\boldsymbol \gamma}_{ii} 
260: \geq ({\boldsymbol H}^{-1})_{ii}.
261: \label{eq:QCRMulti}
262: \end{align}
263: Let us now suppose to reparametrize the family of quantum states with
264: a new set of parameters $\tilde{\Lambda}= \{\tilde{\lambda}_j=
265: \tilde{\lambda}_j(\Lambda) \}$. We have
266: $\tilde{\partial}_j = \sum_i B_{ij} \partial_i$ with $B_{ij} =
267: \partial \lambda_i/\partial \tilde{\lambda}_j$ and, in turn  
268: \begin{align}
269: \tilde{L}_j & = \sum_i B_{ij} L_i \nonumber \\ 
270: H(\tilde{\Lambda})_{ij} & = \sum_{r,s}
271: B_{ir}
272: H(\Lambda)_{rs} 
273: B_{js} \label{eq:Hchangevar}
274: \end{align}
275: {\em i.e.} $\widetilde{{\boldsymbol H}} = {\boldsymbol B} 
276: {\boldsymbol H} {\boldsymbol B}^T$. \par
277: In the following we address the problem of finding the bounds to the
278: estimation of the entanglement between the two subsystems $A,B$ of a
279: family of bipartite quantum states. The general strategy will be to
280: start from the expression of the family of states in terms of a given
281: set of "natural" parameters and then make a change of variable in order
282: to write the state directly in terms of the chosen entanglement
283: monotone $\lambda=\epsilon$.
284: Once this is achieved, the results obtained hold, at every fixed value of
285: the entanglement,  for the whole orbit of states that can be obtained
286: by acting on the given one with local unitary operators $U_{AB}\doteq U_A\otimes U_B$.
287: Indeed, the latter, acting locally on the two subsystems $A$ and $B$,
288: they do not change the entanglement of the state; furthermore they do
289: not change the value of the QFI.  This follows from the observation
290: that, if a given $L_{\epsilon}$ is the solution of (\ref{eq:SLD}), the
291: SLD that will correspond to $\tilde{\varrho}_{\epsilon}=
292: U_{AB}\varrho_{\epsilon}U_{AB}^\dagger$ is given by
293: $\tilde{L}_{\epsilon}=U_{AB}L_{E}U_{AB}^\dagger$ and, due to the cyclic property
294: of the trace,
295: $H(\epsilon)=\Tr[\varrho_{\epsilon}L_{\epsilon}^2]=
296: \Tr[\tilde{\varrho}_{\epsilon}\tilde{L}_{\epsilon}^2]$.
297: \par
298: We finally notice that in order to assess the estimability of a given
299: parameter the relevant figure of merit is given by the signal-to-noise
300: ratio
301: \begin{align}
302: R(\lambda) &= \frac{\lambda^2}{\mathrm{Var}(\lambda)} \le
303: Q (\lambda)  = \lambda^2 H(\lambda) \label{eq:SNR def}\;,
304: \end{align}
305: rather than the variance itself. In particular the SNR of an estimator
306: is relevant to assess its performances in estimating small values of the
307: parameter. Eq. (\ref{eq:SNR def}) shows that
308: the SNR $R(\lambda)$ is bounded by the quantum signal-to-noise ratio
309: (QSNR) $Q(\lambda)$ expressed in terms of the QFI.
310: Upon taking into account repeated measurements we have that 
311: the number of measurements leading to a $99.9 \%$ ($3\sigma$) 
312: confidence interval corresponds to a relative
313: error 
314: $$
315: \delta^2 = \frac{9 \hbox{Var}(\lambda)}{M\lambda^2} = \frac{9}{M}
316: \frac1Q(\lambda) = \frac{9}{M\lambda^2 H(\lambda)}
317: $$
318: Therefore, the number of measurements needed to achieve a $99.9\%$
319: confidence interval with a relative error $\delta$ scales as
320: \begin{equation}
321: M_\delta(\lambda) = \frac9{\delta^2} \frac1{Q(\lambda)} \label{eq:Mdelta}
322: \end{equation}
323: In other words, a vanishing $Q(\lambda)$ implies a diverging number 
324: of measurements to achieve a given relative error, whereas for a 
325: finite  $Q(\lambda)$ the number of measurements is determined by the
326: desired level of precision. 
327: In order to have a non-vanishing $Q(\lambda)$ for small value
328: of the parameter $\lambda$, the QFI should diverge at least as
329: $H(\lambda) \sim \lambda^{-2}$ for vanishing $\lambda$.
330: We notice that a similar quantity, namely $\lambda H(\lambda)$,
331: has been used in to
332: asses estimation strategy for the parameter of a qubit depolarizing
333: channel \cite{Fuj03}.
334: %%%
335: \section{Two-qubit systems} \label{s:qubits}
336: In this section we analyze estimation of entanglement for families of
337: two-qubit states and evaluate limits to precision using the formalism
338: developed in the previous section. At first we address the set of pure
339: states and then consider families of entangled mixtures. In both cases 
340: we consider different measures of entanglement.
341: %%%
342: \subsection{Pure states}
343: We start by considering the set of pure states of two qubits. 
344: Upon exploiting the Schmidt decomposition
345: \be
346: \ket{\Psi_q}=\sqrt{q}\ket{0}_A\ket{0}_B+\sqrt{1-q}
347: \ket{1}_A\ket{1}_B\:, \label{eq: Psi_q}
348: \ee
349: the whole family of pure states can be parametrized by a single parameter:
350: the Schmidt coefficient $q$. Since for two-qubit pure states $q$ is
351: itself an entanglement monotone, all measures of entanglement can be
352: expressed as a monotone function $\epsilon=\epsilon(q)$. As a
353: consequence, in order to determine the precision of estimation it 
354: suffices to evaluate the QFI $H(q)$ and then use the rule for
355: repametrization in Eq. (\ref{eq:Hchangevar}), {\em i.e.}
356: $H(\epsilon)=H[q(\epsilon)][\partial_\epsilon q(\epsilon)]^2$.
357: Since the states are pure, the SLD may be evaluated as $L_{q}=2\partial_q
358: \ketbra{\Psi_q}{\Psi_q}$; the resulting Cramer-Rao bound and the QSNR 
359: read as follows 
360: \begin{align}
361: {\mathrm{Var}}(q)& \geq H(q)^{-1}=q(1-q) \\ 
362: Q(q) &= \frac{q}{1-q} \stackrel{q\rightarrow 0}{\sim} q \:. 
363: \label{eq:QSNR twoqubitpure}
364: \end{align}
365: $Q(q)$ vanishes for vanishing $q$, thus indicating that any
366: estimator of the Schmidt coefficient $q$ becomes
367: less and less precise for vanishing $q$. It is worth noting that
368: the Schmidt coefficient $q$ coincides with the only independent
369: eigenvalue of the reduced density matrix $\varrho_{A(B)}=\hbox{diag}\{q,1-q\}$,
370: which is diagonal in the Schmidt basis. Therefore the QSNR in Eq.
371: (\ref{eq:QSNR twoqubitpure}) also imposes bound to the determination
372: of the eigenvalue of $\varrho_{A(B)}$. Indeed, the same bound could have
373: been obtained by applying the estimation machinery directly
374: to $\varrho_{A(B)}$.
375: \par
376: Let us now consider two different measures of entanglement for 
377: pure two-qubit states, {\em i.e} the negativity $\epsilon_{\hbox{\tiny N}}$
378: \cite{NVW} and the (normalized) linear entropy
379: $\epsilon_{\hbox{\tiny L}}=2(1-\hbox{Tr}[\varrho_A^2])$. In terms
380: of the Schmidt coefficient $q$ we have
381: \be
382: \epsilon_{\hbox{\tiny N}}=
383: \sqrt{\epsilon_{\hbox{\tiny L}}}=2\sqrt{q(1-q)}
384: \ee
385: We recall that the negativity is a good measure of entanglement for
386: generic two-qubit states, {\em i.e} it is an entanglement monotone
387: and it differs from zero iff the state is entangled, whereas 
388: the linear entropy is a good entanglement monotone only iff the state is pure.
389: Upon expressing the Schmidt coefficient as 
390: $q=\frac12 (1-\sqrt{1-\epsilon_{\hbox{\tiny N}}^2})$ 
391: and using (\ref{eq:Hchangevar}), we have
392: $\mathrm{Var}(\epsilon_{\hbox{\tiny N}}) \geq
393: H(\epsilon_{\hbox{\tiny N}})^{-1}=
394: 1-\epsilon_{\hbox{\tiny N}}^2$, 
395: $\mathrm{Var}(\epsilon_{\hbox{\tiny N}}) \geq 
396: H(\epsilon_{\hbox{\tiny L}})=
397: 4\epsilon_{\hbox{\tiny L}}(
398: 1-\epsilon_{\hbox{\tiny L}})$ and, in turn, 
399: \begin{align}
400: Q(\epsilon_{\hbox{\tiny N}}) & =\frac{\epsilon_{\hbox{\tiny
401: N}}^2}{1-\epsilon_{\hbox{\tiny N}}^2}\stackrel{\epsilon_{\hbox{\tiny
402: N}}\rightarrow 0}{\sim} \epsilon_{\hbox{\tiny N}}^2 \\
403: Q(\epsilon_{\hbox{\tiny L}}) & =\frac{\epsilon_{\hbox{\tiny
404: L}}}{4(1-\epsilon_{\hbox{\tiny L}})} \stackrel{\epsilon_{\hbox{\tiny
405: L}}\rightarrow 0}{\sim} \epsilon_{\hbox{\tiny L}}/4\:.
406: \end{align}
407: The optimal estimator for the Schmidt coefficient has a variance
408: ${\mathrm{Var}}(q)$ which is minimum for $q=0,1$ (product state) and
409: maximum for $q=1/2$ (Bell state) whereas for the two entanglement
410: measure we have that ${\mathrm{Var}}(\epsilon_{\hbox{\tiny N}})$ is
411: monotonically decreasing with $\epsilon_{\hbox{\tiny N}}$;
412: ${\mathrm{Var}}(\epsilon_{\hbox{\tiny L}})$ is minimum when the state is
413: either in a product form ($\epsilon_{\hbox{\tiny L}}=0$) or is maximally
414: entangled ($\epsilon_{\hbox{\tiny L}}=1$) and is  maximum in the
415: "intermediate" case ($\epsilon_{\hbox{\tiny L}}=1/2$).  Despite the
416: variances behave quite differently we have the same qualitative
417: behavior of the quantum signal-to-noise ratios and of the number of
418: measurement necessary at fixed relative error $M_\delta$. Indeed, in all
419: cases the QSNR is an increasing function of the parameter and it
420: diverges when the latter takes its maximum value ($q=1$,
421: $\epsilon_{\hbox{\tiny N}}=\epsilon_{\hbox{\tiny L}}=1$);
422: $M_\delta(\epsilon)$ diverges for $\epsilon=\epsilon_{\hbox{\tiny N}}
423: =\epsilon_{\hbox{\tiny L}}=0$ and then decreases monotonically, going to
424: zero for the maximum value of entanglement
425: $\epsilon=\epsilon_{\hbox{\tiny N}}=\epsilon_{\hbox{\tiny L}}=1$ .
426: Moreover for vanishing entanglement the QNSR of the linear entropy
427: estimator is vanishing slower than the corresponding quantity for the
428: negativity.  We conclude that the linear entropy is a more efficient
429: entanglement estimator though, being the QSNR vanishing, the estimation
430: is anyway inherently inefficient.
431: \par
432: The above result can obviously generalized to the case of systems
433: composed by a qubit and an $N$-level system; indeed only two of the
434: dimensions of the latter can be used to express the state: the reduced
435: density matrices of both subsystems have only two non zero eigenvalues.
436: %%%
437: \subsection{Entangled mixtures}
438: We now consider few families of mixed entangled states with
439: different properties and show that they exhibit a common
440: behavior concerning estimation of entanglement.
441: The first family is described by the set of density matrices
442: \be
443: \varrho=U(q) \nu_p U^\dag (q)\label{eq: Mixedstate1}
444: \ee
445: where 
446: \begin{align}
447: \nu_p & =p\ketbra{0,0}{0,0}+(1-p)\ketbra{1,1}{1,1} \nonumber \\ 
448: U(q) & =\exp\{i\,\hbox{arcos}\, q\, \sigma_x\otimes \sigma_x\} \nonumber \,.
449: \end{align}
450: These states depend on two parameters $(q,p)$ and
451: are obtained via the action of the entangling operator
452: $U(q)$ on the classically correlated state $\nu_p$. Upon varying the
453: parameter $p$ we may control the purity $\mu(p)=1-2p(1-p)$
454: of the state, while varying
455: $q$ we tune the amount of entanglement. The QFI matrix is
456: diagonal with elements:
457: \bae
458: {\boldsymbol H}(p,q) =
459: \hbox{diag}\left(\frac{1}{p(1-p)},\frac{(1-2p)^2}{q(1-q)}\right)
460: \eae
461: The negativity of the state  (\ref{eq: Mixedstate1}) is given by
462: the unique
463: negative eigenvalue of the partially transposed state $\varrho^{T_A}$:
464: \be
465: \epsilon_{\hbox{\tiny N}}=2\sqrt{q(1-q)} (1-2p)
466: \ee
467: Upon inverting the above relation and the one for the purity, 
468: we reparametrize the set of
469: states in terms of the new parameter $(p,q) \rightarrow 
470: (\mu,\epsilon_{\hbox{\tiny N}})$.
471: The transfer matrix is given by
472: \begin{equation}
473: {\boldsymbol B}=\left(\begin{array}{cc}
474: -\frac{1}{2\sqrt{2\mu -1}} & \frac{\epsilon_{\hbox{\tiny N}}^2}
475: {2\sqrt{(2\mu -1)^2(2\mu -1 -\epsilon_{\hbox{\tiny N}}^2)}}\\
476: 0 & \frac{\epsilon_{\hbox{\tiny N}}}
477: {2\sqrt{(2\mu -1)(2\mu -1 -\epsilon_{\hbox{\tiny N}}^2)}}
478: \end{array}\right)
479: \end{equation}
480: and the inverse QFI matrix:
481: \begin{equation}
482: {\boldsymbol H}(\mu,\epsilon_{\hbox{\tiny N}})^{-1}=\left(\begin{array}{cc}
483: -4\mu^2 + 6\mu -2 
484: &2\epsilon_{\hbox{\tiny N}}(1-\mu)\\
485: 2\epsilon_{\hbox{\tiny N}}(1-\mu) & 1-\epsilon_{\hbox{\tiny N}}^2
486: \end{array}\right)
487: \end{equation}
488: The corresponding bound on the variance is thus given by:
489: \be
490: {\mathrm{Var}}(\epsilon_{\hbox{\tiny N}})\ge
491: \left({\boldsymbol H}(\mu,\epsilon_{\hbox{\tiny N}})^{-1}\right)_{22} =
492: 1-\epsilon_{\hbox{\tiny N}}^2\,
493: \label{varn}
494: \ee
495: which represents the bound to the precision of any entanglement
496: (negativity) estimation procedure performed at fixed purity $\mu$.
497: The result in Eq. (\ref{varn}) is independent on the purity, 
498: no optimization procedure may be pursued, and it coincides with
499: the bound obtained and discussed in the previous
500: subsection for pure states.
501: \par
502: Let us now consider the family of Werner-like states
503: \be
504: \varrho_{pq} = \frac{1-p}{4}\openone \otimes \openone + p \ketbra{\Psi_q}{\Psi_q}
505: \ee
506: obtained by depolarizing an entangled state $\ket{\Psi_q}$ 
507: of the form given in Eq. (\ref{eq: Psi_q}).  This set of states 
508: depends on two parameters $p,q$. As in the previous example upon varying
509: the parameter $p$ we may control the purity $\mu(p)=(1+3p^2)/4$ of the state,
510: while the amount of entanglement depends on both parameters.
511: The eigenvalues of $\varrho_{pq}$ depends only on $p$ whereas
512: the eigenvectors depends only on $q$. The QFI matrix is thus given 
513: by the diagonal form
514: \be
515: {\boldsymbol H}(p,q)=\mbox{diag}\left \{ \frac{3}{1+(2-3p)p},\, \,
516: \frac{p^2}{q(1-q)(1+p)} \right \}\,
517: \ee
518: and the inverses of the diagonal elements correspond
519: to the ultimate bounds to ${\mathrm{Var}}(p)$ and ${\mathrm{Var}}(q)$
520: of any estimator of $p$ and $q$, either at fixed value of the other
521: parameter or in a joint estimation procedure. Entanglement of Werner 
522: states may be evaluated in terms of negativity,
523: \be
524: \epsilon_{\hbox{\tiny N}}=\max \left\{0,
525: \frac12 \left[p\left((1+4\sqrt{q(q-1)}\right)-1\right]\right\}
526: \label{wen}
527: \:,
528: \ee
529: which implies that Werner states are entangled for $(1 > p > [1 
530: + 4q (1 - q)]^{-1}$.
531: Upon inverting Eq. (\ref{wen}) for $p$ or $q$ we may parametrize 
532: the Werner states using $(\epsilon_{\hbox{\tiny N}},q)$ or 
533: $(p,\epsilon_{\hbox{\tiny N}})$ and evaluate the QFI matrices 
534: $H(\epsilon_{\hbox{\tiny N}},q)$ and $H(p,\epsilon_{\hbox{\tiny N}})$, 
535: their inverses and, in turn, the corresponding bounds to the precision
536: of entanglement (negativity) estimation. 
537: The main results are that the ultimate bounds to the variance, 
538: and thus to the QSNR depend very slightly on the other free
539: parameter ($q$ or $p$). In other words, estimation procedures
540: performed at fixed value of $p$ or $q$ respectively shows 
541: different precision, but the differences are negligible in the whole
542: range of variations of the parameters.
543: We do not report here the analytic expression of $Q(\epsilon_{\hbox{\tiny N}})$
544: at fixed $p$ or $q$ which is quite cumbersome. Rather, we show the behavior 
545: of $Q(\epsilon_{\hbox{\tiny N}})$ in Fig. \ref{fig:SNR_Wer_Neg}.
546: On the left we show the QSNR $Q_{q=0.5}(\epsilon_{\hbox{\tiny N}}) 
547: $ for $q=0.5$, whereas on the right we show the ratio 
548: $Q_{p}(\epsilon_{\hbox{\tiny N}})/Q_{q=0.5}(\epsilon_{\hbox{\tiny N}})$
549: for different value of $p$ (a similar behavior may be observed 
550: upon varying $q$).
551: As it is apparent from the main panel the QSNR is a growing function 
552: of $\epsilon_{\hbox{\tiny N}}$, vanishes for
553: vanishing negativity and diverges for maximally entangled states
554: $\epsilon_{\hbox{\tiny N}}=1$.  The inset shows that 
555: there is almost no
556: dependence on the actual value of $p$ and $q$ respectively and
557: this prevents any possible optimization of the estimation 
558: procedure. For small $\eln$ we have 
559: $Q (\eln) \simeq f(q) \eln^2$ and $Q(\eln) \simeq g(p)\eln^2$
560: respectively, where both the functions $f(q)\simeq 1$ and 
561: $g(p)\simeq 1$ are again very close to unit value for the whole ranges of
562: variation of $q$ and $p$. 
563: %%%
564: \begin{figure}[h!]
565: \includegraphics[width=0.46\textwidth]{WQ.eps}
566: \caption{Quantum signal-to-noise ratio for the estimation of
567: entanglement (negativity) of a two-qubit Werner state as a
568: function of the negativity at fixed $q=0.5$. The inset shows 
569: the ratio $Q_{p}(\epsilon_{\hbox{\tiny N}})/
570: Q_{q=0.5}(\epsilon_{\hbox{\tiny N}})$ for different value 
571: of $p$.}
572: \label{fig:SNR_Wer_Neg}
573: \end{figure}
574: %%%
575: \par
576: In all the cases we have considered, the QSNR is small for the most
577: part of entanglement range and starts growing only for highly 
578: entangled states. In other words, estimation of
579: entanglement is, on average, an inefficient procedure.
580: %%%
581: \section{Two-qutrit bound entangled states} \label{s:PPT}
582: In the previous section we have seen how the estimation of entanglement,
583: as measured by negativity is a fairly inefficient procedure for weakly
584: entangled states. Here we want to test how the QFI and the related
585: bounds behave when one considers states that have inherently small
586: amount of entanglement. A paradigmatic examples of such states are the
587: so called {\em bound entangled states}, which exhibit non-classical 
588: correlations even if they satisfy the separability criterion based 
589: on partial transposition of the density matrix 
590: \cite{HorodPPT,PeresSep,HorodSep}. The first example of bound entangled
591: states is given by the following family of two spin-1 states 
592: \cite{HorodPPT} 
593: \begin{eqnarray}
594: \label{eq:bound}
595: \varrho_a&=& \frac{a}{1+ 8 a} \big(
596:    | \downarrow 0 \rangle\langle \downarrow 0  | +
597:    |\downarrow \uparrow \rangle\langle \downarrow  \uparrow | + 
598:    |0 \downarrow \rangle\langle  0 \downarrow | 
599:    \nonumber \\ &+&
600:    |0 \uparrow \rangle\langle  0 \uparrow | +
601:    |\uparrow 0 \rangle\langle \uparrow 0 | \big)
602:   + \frac{3 a}{1+ 8 a}
603:    |E \rangle\langle E | 
604: \nonumber \\ &+&
605:     \frac{1}{1+ 8 a}
606:    |\Pi \rangle\langle \Pi |
607:  \end{eqnarray}
608: where
609: \begin{align}
610: | E \rangle & =
611: \frac{1}{\sqrt{3}}\left(
612: | \downarrow\downarrow
613: \rangle + | 00 \rangle + |\uparrow\uparrow\rangle
614: \right) 
615: \nonumber \\ 
616: |\Pi \rangle & =
617: \sqrt{\frac{1+a}{2}} | \uparrow\downarrow \rangle +
618: \sqrt{\frac{1-a}{2}} |\uparrow\uparrow \rangle.
619: \nonumber  
620: \end{align}
621: Since for all values of the parameter $a$ $\; \varrho_a$ has a
622: positive partial transpose (PPT), negativity  cannot be used as a
623: measure of the quantum correlations present in state. In order to
624: estimate the entanglement we will use the scheme proposed in
625: \cite{HoffmanLUR,HoffmanLURPPT}. The latter is based on the following
626: considerations. Given the sets of $n$ non-commuting operators $\{A_i\}$
627: and $\{B_i\}$ acting locally on the subsystem A and B respectively one
628: has a lower bound for the sum of the local uncertainties relations
629: (LUR): 
630: \be
631: \sum_i \delta A_i^2 > U_A \mbox{ and } \sum_i \delta B_i^2 > U_B
632: \ee
633: where $\delta O_i^2 = \average{O_i^2}-\average{O_i}^2$ is the variance
634: of the operator $O_i$. Since for all separable states on has that
635: $\sum_i \delta (A_i +B_i)^2> U_A+U_B$, the latter inequality set a
636: necessary condition for a state to be entangled. The relative violation
637: of the inequality defined as 
638: \be
639: \elu=1-\frac{\sum_i \delta (A_i +B_i)^2}{U_A+U_B}
640: \ee
641: can then be used as a measure of the quantum correlations present in the
642: given state. The violation is necessary condition for the presence of
643: the entanglement, thus, in order to effectively have and maximize such
644: violation one can judiciously choose and optimize the choice of the sets
645: $\{A_i\}$ and $\{B_i\}$. The result of a possible optimization for the
646: state $\varrho_a$ is given in \cite{HoffmanLURPPT} and the
647: corresponding relative violation depends on the parameter $a$ and is
648: given by: 
649: \be
650: \elu(\varrho_a)=\frac{3 a^2(1-a)}{4 (2+a)(1+8a)^2}
651: \ee
652: The latter expression can be used to parameterize the state
653: $\varrho_a$ in terms of $\elu(\varrho_a)$ and then
654: apply the QFI machinery in order to obtain the desired bound on the
655: estimation of relative violation of the LUR. The results 
656: are shown in Fig. \ref{fig:M_CLUR_PPT}.
657: We first note that the relative violation of the LUR is small, {\em i.e} 
658: $\elu(\varrho_a)\in[0,2/1125]$. Nonetheless, the number of
659: measurements $M_\delta(\elu)$ is of the same order of those needed in the 
660: qubit case $M_\delta(\epsilon_{\hbox{\tiny N}})$ in similar conditions,
661: and thus the overall efficiency of the estimation process is comparable
662: for most part of the entanglement range. 
663: Finally, we notice that also for this family of qutrits the number 
664: of measurements $M_\delta(\elu)$ diverges for vanishing entanglement.
665: %%%
666: \begin{figure}[h!]
667: \includegraphics[width=0.22\textwidth]{PPTTrialsCLUR0M.eps}
668: \includegraphics[width=0.22\textwidth]{PPTTrialsCLURM1.eps}
669: \caption{Number of measurement $M_\delta(\elu)$ needed to achieve a
670: given relative error $\delta=10^{-1}$ (red), $\delta=10^{-2}$ (green), 
671: $\delta=10^{-3}$ (blue) in the the estimation of LUR entanglement
672: measure $\elu$ of PPT state $\varrho_a$ in Eq.(\ref{eq:bound}).
673: Left: $a \in [0,4/13]$, Right: $a \in [4/13,1]$}\label{fig:M_CLUR_PPT}
674: \end{figure}
675: %%%
676: \section{Gaussian states}\label{s:CV}
677: In this section analyze continuous variable systems and derive the
678: bounds for the estimation of entanglement of two-mode Gaussian states 
679: \cite{GaussiansAOP}.
680: After a brief introduction  we consider both pure states, {\em i.e.}
681: twin-beam and the family of mixed states represented by squeezed thermal
682: states (STS). Different measures of entanglement will be considered.
683: \par 
684: The characteristic function of the state $\varrho$ is defined as 
685: $\chi[\varrho]({\bf \Gamma})= \hbox{Tr}\left[\varrho\, 
686: D(\Gamma)\right]$ where $D(\Gamma)=\exp{ \left [i \textbf{R}^T 
687: {\bf \Omega} {\bf \Gamma} \right ] }$ is the displacement
688: operator, defined in terms of the 
689: symplectic matrix
690: \be
691: {\bf \Omega}=\bigoplus_{k=1}^n \left( \begin{array}{cc}
692: 0 & 1 \\
693: -1 & 0 \\
694: \end{array}\right )_n
695: \ee
696: and the vector $\textbf{R}^T=(q_1,p_1,\cdots,q_n,p_n)$ of canonical 
697: operators. ${\bf \Gamma}^T=(a_1,b_1,\cdots,a_n,b_n)$ denotes the Cartesian 
698: coordinates and we have $\left [ R_k,R_h\right ]=i {\bf \Omega}_{kh}$. 
699: Two-mode Gaussian state are those with a characteristic function of the
700: form
701: \be
702: \chi[\varrho]({\bf \Gamma})=\exp{\left( -\frac{1}{2}{\bf \Gamma}^T{\bf \sigma}{\bf \Gamma}
703: +i{\bf \Gamma}^T{\bf \overline{X }}\right)}
704: \ee
705: where
706: \be
707: {\bf \sigma}_{kh}=\frac{1}{2}\langle\{R_k,R_h\}\rangle-\langle R_k\rangle\langle R_h\rangle
708: \ee
709: is the covariance matrix of the second moments and ${\bf \overline{X }}$
710: denotes mean values. The second term in the exponential does not contain 
711: any information about entanglement, and can be set to zero via local
712: operation.  The covariance matrix completely characterize the
713: state and by means of local symplectic transformations 
714: can be transformed into the standard block form :
715: ${\sigma}=  \left(\begin{array}{cc}
716: A & C \\ C & B 
717: \end{array}\right )$ 
718: where $A=\hbox{Diag}(a,a)$, $B=\hbox{Diag}(b,b)$, and
719: $C=\hbox{Diag}(c_+,c_-)$.
720: For two-mode Gaussian states PPT condition is necessary and sufficient
721: for separability and thus the entanglement properties of the state are 
722: encoded in the least symplectic eigenvalue \cite{GaussianCrit}
723: \be
724: \tilde{d}_-=\sqrt{(a-c_+)(a+c_-)}
725: \ee
726: where the symplectic spectrum of ${\bf \sigma }^{T_A}$ can be evaluated by finding the
727: eigenvalues of ${\bf \Omega^{-1}} {\bf \sigma}^{T_A} .$ In this framework the PPT
728: criterion can be cast in term of the smallest symplectic eigenvalue i.e.,
729: $\varrho$ is separable iff $\tilde{d}_- \ge 1/2$.
730: Indeed, $\tilde{d}_-$ is itself an entanglement monotone. Furthermore,
731: all the different entanglement measures for symmetric Gaussian states
732: that have been proposed \cite{NVW, EoFGiedke,EntBuresMarian1}
733: turns out to be a monotone function of the smallest symplectic
734: eigenvalue. Since we are interested in the estimation of entanglement,
735: we may first study the estimation of the symplectic eigenvalue $\tilde{d}_{-}$
736: and then use repametrization in order to asses the performances of other 
737: entanglement monotones. In particular, we will focus on the following
738: measures:
739: \begin{align}
740: \epsilon_{\hbox{\tiny N}}(\tilde{d}_-) &= \max{\{0, -\ln{2}\tilde{d}_-\}} 
741: \label{eq:LogNeg} \\
742: \epsilon_{\hbox{\tiny L}}(\tilde{d}_-) &= 1 - \frac{4\tilde{d}_-}{1+4\tilde{d}_-^2} 
743: \label{eq:LinEnt} \\
744: \epsilon_{\hbox{\tiny S}}(\tilde{d}_{-}) &= 1 -2 \tilde{d}  
745: \label{eq:BuresEnt} \\
746: \epsilon_{\hbox{\tiny B}}(\tilde{d}_{-}) &= \frac{(1-\sqrt{2 \tilde{d}})^2}
747: {1+2\tilde{d}}. 
748: \label{eq:SEnt}
749: \end{align}
750: where the expressions for the
751: linear entropy $\epsilon_{\hbox{\tiny L}}$ has been obtained
752: in the pure state case. In particular $\epsilon_{\hbox{\tiny L}}$
753: and the logarithmic negativity $\epsilon_{\hbox{\tiny N}}$ will be 
754: used for pure 
755: states, whereas the Bures distance-based measure 
756: $\epsilon_{\hbox{\tiny B}}$ and $\epsilon_{\hbox{\tiny S}}$ 
757: \cite{EntBuresMarian1} 
758: will be used for mixed states.
759: Notice that $\epsilon_{\hbox{\tiny B}}$ is a good measure of entanglement
760: just for symmetric two-mode Gaussian states and that Eq.  (\ref{eq:BuresEnt}) 
761: has been obtained for this particular class of states.
762: In order to evaluate the QFI, one has first to determine the actual
763: expression of $\tilde{d}_{-}$. Then, one expresses the elements of the
764: covariance matrix in terms of the chosen entanglement monotone and
765: proceeds with the repametrization rules described in Section
766: \ref{s:qet}. 
767: %%%%
768: \subsection{Pure states}
769: Here we address estimation of the entanglement of pure two
770: modes Gaussian states, {\em i.e.} twin-beam. These are defined by the 
771: following relations between the elements of the covariance matrix: 
772: $a=b$ i.e., the states are symmetric and $c_+=-c_-=\sqrt{a^2-1/4}$. 
773: The states can thus be described by the single parameter $a$ or, by 
774: inverting $\tilde{d}_-= a -\sqrt{a^2-1/4}$ they can be completely 
775: described by their entanglement content. For Gaussian states the 
776: evaluation of the SLD and the QFI, besides the use of Eqs. (\ref{defs}), 
777: may be pursued using phase-space techniques.
778: In fact, for pure states we have
779: $L_\E=2\partial_\E \varrho_\E$ and this allows to directly evaluate 
780: the characteristic function of the SLD as follows 
781: \be
782: \chi[L_\E] (\bf\Gamma )=2\partial_\E\chi[\varrho_\E]=
783: -{\bf \Gamma}^T \partial_\E\sigma_\E{\bf \Gamma}\ \chi[\varrho_\E]\:.
784: \ee
785: where ${\bf \Gamma}^T=({\bf \Gamma_1}^T, {\bf \Gamma_2}^T)$.
786: The corresponding QFI
787: $H(\epsilon) = \mbox{Tr}[\varrho\, L_\E^2]$ is given by
788: \bae
789: H(\epsilon)&=&
790: \int\!\!\int
791: \frac{d^{2}{\bf \Gamma}_1 }{2\pi}\frac{d^{2}
792: {\bf \Gamma}_2}{2\pi}\, \mathcal{I}
793: \left(\E,{\bf \Gamma}_1,{\bf \Gamma}_2\right)
794: \eae 
795: where the integrand function reads as follows
796: \bae
797: \mathcal{I}&=& \chi[L_\E]({\bf \Gamma}_1)\chi[L_\E]({\bf \Gamma}_2)
798: \mbox{ Tr }[\varrho D^\dagger({\bf \Gamma}_1) D^\dagger({\bf \Gamma}_2)]
799: \label{eq: IGauss1}
800: \eae
801: We now use the relations $D[{\bf \Gamma}]^\dagger=D[-{\bf \Gamma}]$ 
802: and $D({\bf \Gamma}_1) D({\bf \Gamma}_2)=D({\bf \Gamma}_1+
803: {\bf \Gamma}_2)g({\bf \Gamma}_1,{\bf \Gamma}_2)$ and we 
804: rewrite (\ref{eq: IGauss1}) as
805: \bae
806: \mathcal{I}&=&
807: \left({\bf \Gamma}_1^T \partial_\E\sigma_\E{\bf \Gamma}_1\right)
808: \left({\bf \Gamma}_2^T\partial_\E\sigma_\E{\bf \Gamma}_2\right)
809: \cdot\nonumber \\ &&
810: \chi_{\varrho_\E}({\bf \Gamma}_1)
811:  \chi_{\varrho_\E}
812: ({\bf \Gamma}_2) \chi_{\varrho_\E:}(-{\bf \Gamma}_1-{\bf \Gamma}_2)
813: g({\bf \Gamma}_1,{\bf \Gamma}_2)
814: \label{eq: IGauss2}
815: \nonumber \\ &=&
816: \left ({\bf \Gamma}^T{\bf \Sigma_1}{\bf \Gamma}\right )
817: \ \left ( {\bf \Gamma}^T{\bf \Sigma_2}{\bf \Gamma} \right )
818: \ \exp{\left (-\frac{1}{2}{\bf \Gamma}^T{\bf \Delta}{\bf \Gamma} \right )}
819: \eae
820: where we have introduced  the matrices:
821: \be
822: {\bf \Sigma}=  \left( \begin{array}{cc}
823: 2\sigma &\sigma \\
824: \sigma & 2\sigma
825: \end{array}\right ),
826: {\bf \Sigma_1}=  \left( \begin{array}{cc}
827: \partial_\E\sigma &0 \\
828: 0 & 0
829: \end{array}\right ),
830: {\bf \Sigma_2}=  \left( \begin{array}{cc}
831: 0 &0 \\
832: 0 & \partial_\E \sigma
833: \end{array}\right )\nonumber
834: \ee
835: and where $\Delta= {\bf \Sigma} + i \Upsilon$, with $\Upsilon=\frac12\,
836: \sigma_y \otimes {\mathbbm I} \otimes \sigma_y$ in terms of Pauli matrices.  
837: The result of the
838: integration QFI is a function of $\E,\partial_\E \sigma_\E$ and can be
839: expressed in various ways depending on the entanglement monotone that
840: one chooses to estimate.  By setting
841: $a=\frac{1}{2}\cosh\epsilon_{\hbox{\tiny N}}$ in the covariance matrix,
842: one can use the logarithmic negativity $\epsilon_{\hbox{\tiny N}}$. In
843: this case one finds that the QFI is independent on the entanglement
844: content of the state.  \par If one uses directly the symplectic
845: eigenvalue $\tilde{d}_-$ the QFI now depends on the entanglement
846: monotone:
847: \be
848: H(\tilde{d}_-)=\tilde{d}_-^{-2}
849: \label{Gnu_}
850: \ee 
851: and the minimal variance in the estimation can be obtained in the limit
852: of infinite entanglement i.e., $\tilde{d}_-=0$. Moreover we observe
853: that, while for the logarithmic negativity one sees that the QSNR
854: $Q(\eln)$ is simply proportional to $\eln^2$, if we consider the least
855: symplectic eigenvalue we have $Q(\tilde{d}_-)=1$ over the whole range of
856: variation, \emph{i.e.} the estimation procedure can be done efficiently
857: either for highly entangled states and weakly entangled ones.
858: \par
859: As a matter of fact, twin-beam may be also written in the Fock basis as 
860: $\ket{\Psi} = \sum_n f_n \ket{n}\ket{n}$
861: where, in terms of the log-negativity or the linear entropy 
862: one may write
863: \begin{align}
864: f_n  = \sqrt{2 \frac{(\cosh \epsilon_{\hbox{\tiny
865: N}}-1)^n}{(\cosh \epsilon_{\hbox{\tiny N}}+1)^{1+n)}}}\: 
866: = \sqrt{2\frac{
867: \epsilon_{\hbox{\tiny L}}^n
868: (1-\epsilon_{\hbox{\tiny L}})
869: }{
870: (2-\epsilon_{\hbox{\tiny L}})^{1+n}
871: }}\,.
872: \end{align}
873: Using this representation we may directly exploit Eqs. (\ref{defsp}):
874: for a generic parameter $x$ and $f_n(x)\in \mathbb{R}$
875: we have $\braket{\Psi}{\partial_x \Psi}=0$ and thus
876: \begin{equation}
877: H(x)=\braket{\partial_x \Psi}{\partial_x \Psi}=\sum_n (\partial_x
878: f_n(x))^2\:.
879: \end{equation}
880: Using the above equation one recover the result for the 
881: log-negativity and may evaluate the QFI in terms of the
882: linear entropy obtaining
883: \begin{equation}
884: H(\epsilon_{\hbox{\tiny L}}) =
885: \left (4(2 -\epsilon_{\hbox{\tiny L}})(\epsilon_{\hbox{\tiny L}}-1)^2 
886: \epsilon_{\hbox{\tiny L}}  \right )^{-1} .
887: \end{equation}
888: %%%%
889: \subsection{Entangled mixtures}
890: We now analyze entanglement estimation for a relevant 
891: family of mixed Gaussian states labeled by two independent 
892: parameters. The symmetric two-mode squeezed thermal states 
893: (STS) are given by
894: \begin{align}
895: \varrho_{ST} = S_{12}(r,\phi) (\nu_{N_t}\otimes\nu_{N_t}) S_{12}^{\dag}(r,\phi)
896: \label{eq:ST}
897: \end{align}
898: and represents a two parameter family 
899: $\varrho_{ST}=\varrho_{ST}(N_t,r)$ obtained from symmetric 
900: two-mode thermal state with $N_{t}$
901: thermal photons for each mode by the action of the two-mode 
902: squeezing operator
903: $S_{12}(r) = \exp \{ r (a_1^{\dag}a_2^{\dag} - a_1a_2)\}$. 
904: The family in Eq. (\ref{eq:ST}) occurs when one considers
905: the propagation of twin-beam in a noisy channel or the 
906: generation of entanglement from a noisy background \cite{ntw}, 
907: and represents the CV generalization of the family of entangled 
908: mixed states introduced in Eq. (\ref{eq: Mixedstate1}).
909: We evaluate the corresponding Fisher information matrix 
910: ${\boldsymbol H}(r,N_t)$ and obtain a diagonal matrix
911: \begin{align}
912: {\boldsymbol H}(r,N_t) =
913: \hbox{diag}\left( 8 - \frac{4}{1+ 2 N_t(1+N_t)},\frac{2}{N_t(1+N_t)}\right) .
914: \end{align}
915: Let us now consider the smallest symplectic eigenvalue 
916: $\tilde{d}_{-}(r,N_t)$ and the purity of the state $\mu(N_t)$
917: \begin{align}
918: \tilde{d}_{-} = \frac{e^{-2r}}{2} (1 + 2 N_t) \qquad
919: \mu = \frac{1}{2 N_t +1}. \nonumber
920: \end{align}
921: Upon inverting the above equations  
922: we may reparametrize the set of states in terms of
923: the new parameters $(\tilde{d}_{-},\mu)$, the transfer matrix 
924: ${\boldsymbol B}$ being given by  
925: \begin{align}
926: {\boldsymbol B} &=
927: \left(
928: \begin{array}{c c}
929: -\frac{1}{2 \tilde{d}_{-}} & 0 \\
930: -\frac{1}{2 \mu} & -\frac{1-\mu}{2\mu^2} - \frac{1}{2\mu}
931: \end{array}
932: \right) \nonumber
933: \end{align}
934: The new QFI matrix ${\boldsymbol H}(\tilde{d}_{-},\mu)$
935: is calculated by means of Eq. (\ref{eq:Hchangevar})
936: and the bound on the covariance matrix 
937: ${\boldsymbol \gamma}(\tilde{d}_{-},\mu) \geq {\boldsymbol
938: H}(\tilde{d}_{-},\mu)^{-1}$ is established by its inverse
939: \begin{align}
940: {\boldsymbol H}(\tilde{d}_{-},\mu)^{-1} &=
941: \left(
942: \begin{array}{c c}
943: \tilde{d}_{-}^2 & -\frac{\tilde{d}_{-}\mu(1-\mu^2)}{2} \\
944: -\frac{\tilde{d}_{-}\mu(1-\mu^2)}{2} & \frac{\mu^2}{2}(1-\mu^2)
945: \end{array}
946: \right)
947: \end{align}
948: The lower bounds on the variance for the symplectic eigenvalue
949: is given by
950: \begin{align}
951: {\mathrm{Var}}(\tilde{d}_{-}) &\geq ({\boldsymbol H}
952: (\tilde{d}_{-},\mu)^{-1})_{11} = \tilde{d}_{-}^2
953: \end{align}
954: and represents the limit to the precision of any estimator of $\tilde{d}_{-}$
955: at fixed purity $\mu$. In particular, we observe that this bound does not 
956: depend on the purity, and coincides with the bound  in Eq. (\ref{Gnu_})
957: obtained for pure states. Therefore also for this class of states
958: the QSNR is $Q(\tilde{d}_{-})=1$ and hence $\tilde{d}_-$ can be always 
959: estimated efficiently. 
960: \par
961: Let us now consider a generic measure of entanglement 
962: $\epsilon=\epsilon(\tilde{d}_{-})$. Upon using Eq. (\ref{eq:Hchangevar})
963: we may show that the reparametrization $(\tilde{d}_{-},\mu)\rightarrow
964: (\epsilon, \mu)$ leads to
965: \begin{align}
966: H(\epsilon,\mu)_{11}
967: & = \left( \frac{\partial \tilde{d}}{\partial \epsilon}
968: \right)^2 H(\tilde{d}_{-},\mu)_{11} \label{eq:FishFunct} \\
969: {\mathrm{Var}}(\epsilon) &\geq
970: ({\boldsymbol H}(\epsilon,\mu)^{-1})_{11}
971: =
972: \left( \frac{\partial \tilde{d}}{\partial \epsilon} \right)^{-2}
973: ({\boldsymbol H}(\tilde{d}_{-},\mu)^{-1})_{11} \label{eq:CRBFunct}
974: \end{align}
975: Let us consider the two monotone functions of the symplectic eigenvalue
976: $\epsilon_{\hbox{\tiny S}}(\tilde{d}_-)$
977: and $\epsilon_{\hbox{\tiny B}}(\tilde{d}_-)$ introduced in Eqs. (\ref{eq:SEnt}) 
978: and (\ref{eq:BuresEnt}).
979: The symplectic eigenvalue can be expressed in terms of the measures as
980: \begin{align}
981: \tilde{d}(\epsilon_{\hbox{\tiny S}}) &= \frac{1 - \epsilon_{\hbox{\tiny S}}}{2}
982: \nonumber \\
983: \tilde{d}(\epsilon_{\hbox{\tiny B}}) &= \frac{1 + 2 \epsilon_{\hbox{\tiny B}} -
984:  \epsilon_{\hbox{\tiny B}}^2 -2\sqrt{2 \epsilon_{\hbox{\tiny B}} - 
985:  \epsilon_{\hbox{\tiny B}}^2}}{2(1-\epsilon_{\hbox{\tiny B}})^2}.
986: \nonumber
987: \end{align}
988: thus leading to
989: \begin{align}
990: {\mathrm{Var}}(\epsilon_{\hbox{\tiny S}}) &\geq (1-\epsilon_{\hbox{\tiny S}})^2
991: \nonumber \\
992: {\mathrm{Var}}(\epsilon_{\hbox{\tiny B}}) &\geq \frac{\epsilon_{\hbox{\tiny B}}
993: (2-\epsilon_{\hbox{\tiny B}})(1-\epsilon_{\hbox{\tiny B}})^2}{4} \nonumber
994: \end{align}
995: We notice that ${\mathrm{Var}}(\epsilon_{\hbox{\tiny S}})$ 
996: and ${\mathrm{Var}}(\epsilon_{\hbox{\tiny B}})$ 
997: show different behavior; in particular, while the bound
998: on ${\mathrm{Var}}(\epsilon_{\hbox{\tiny S}})$ vanishes 
999: only when $\epsilon_{\hbox{\tiny S}}$
1000: is maximum ($\epsilon_{\hbox{\tiny S}}=1$), the bound on
1001: ${\mathrm{Var}}(\epsilon_{\hbox{\tiny B}})$ reaches zero both when
1002: $\epsilon_{\hbox{\tiny B}}$ is maximum ($\epsilon_{\hbox{\tiny B}}=1$) and when
1003: is minimum ($\epsilon_{\hbox{\tiny B}}=0$) and presents
1004:  a maximum for $\epsilon_{\hbox{\tiny B}}=1-1/\sqrt{2}$.
1005: \par
1006: We finally evaluate the QSNR for the measures of entanglement
1007: introduced, obtaining
1008: \begin{align}
1009: Q(\epsilon_{\hbox{\tiny S}}) &\leq \frac{\epsilon_{\hbox{\tiny S}}^2}
1010: {(1-\epsilon_{\hbox{\tiny S}})^2} \stackrel{\epsilon_{\hbox{\tiny
1011: S}}\rightarrow 0}{\sim} \epsilon_{\hbox{\tiny S}}^2  \\
1012: Q(\epsilon_{\hbox{\tiny B}}) &\leq \frac{4 \epsilon_{\hbox{\tiny B}}}
1013: {(1-\epsilon_{\hbox{\tiny B}})^2(2-\epsilon_{\hbox{\tiny B}})}
1014: \stackrel{\epsilon_{\hbox{\tiny
1015: B}}\rightarrow 0}{\sim} 2\epsilon_{\hbox{\tiny B}}  \nonumber
1016: \end{align}
1017: The two QSNRs are increasing function of entanglement, vanish 
1018: for zero entanglement and diverge for maximally entangled states. 
1019: In turn, the numbers of measurements  $M_\delta(\epsilon_{\hbox{
1020: \tiny B}})$ and  $M_\delta(\epsilon_{\hbox{\tiny S}})$ vanish for 
1021: maximum entanglement and diverge for vanishing entanglement. 
1022: The QSNR of $\epsilon_{\hbox{\tiny B}}$ is vanishing slower than the 
1023: corresponding quantity for $\epsilon_{\hbox{\tiny S}}$
1024: and therefore we conclude that the measure based on the
1025: Bures distance is more efficiently estimable 
1026: compared to the linear measure $\epsilon_{\hbox{\tiny S}}$.
1027: On the other hand, being 
1028: the QSNR vanishing, the estimation is
1029: anyway inherently inefficient. 
1030: %
1031: \section{Conclusions} \label{s:conclusions}
1032: Entanglement of quantum states is not an observable quantity. 
1033: On the other hand, the amount of entanglement can be indirectly 
1034: inferred by an estimation procedure, {\em i.e.} by measuring 
1035: some proper observable and then processing the outcomes by a 
1036: suitable estimators. In this paper we have established a first 
1037: approach to the estimation of the entanglement content of a 
1038: quantum state and to the search of optimal quantum estimators, 
1039: {\em i.e} those with minimum variance. Our approach is based 
1040: on the theory of local quantum estimation and allows, upon the 
1041: evaluation of the quantum Fisher information, to derive the 
1042: ultimate bounds to precision imposed by quantum mechanics. 
1043: We have applied our analysis to several families of quantum 
1044: states either describing finite size systems or continuous 
1045: variables ones, and have considered different measures in order
1046: to quantify the amount of entanglement. \par
1047: For the case two-qubit pure state we have found that 
1048: any procedure to estimate entanglement (either quantified by 
1049: negativity or by linear entropy) is efficient only for maximally
1050: or near maximally entangled states, whereas it becomes inherently 
1051: inefficient for weakly entangled states. In particular, the number 
1052: of measurements needed to achieve a $99.9\%$ confidence interval 
1053: withing a given relative error diverges as far as the value of 
1054: entanglement becomes small. The same results hold also for 
1055: families of mixed states, remarkably for the orbit of an entangling
1056: unitary an for a general class of Werner-like states. 
1057: Indeed in all the examples we have considered the presence of 
1058: other free parameters besides entanglement, though changing the 
1059: QFI, does not affect the estimation precision, \emph{i.e.} the value
1060: of the relevant element of the inverse QFI matrix. In turn, this also
1061: prevents the possibility of further optimizing the estimation procedure.
1062: \par
1063: On the other hand, we have showed that for an important class of
1064: states whose entanglement of distillation is zero (PPT bound entangled
1065: states), the use of an optimized measure of quantum correlation i.e.,
1066: the relative violation of local unitary relations introduced in
1067: \cite{HoffmanLURPPT}, results in a more efficient estimation
1068: procedure, with precision comparable with those achievable in the
1069: estimation of entanglement through negativity. 
1070: \par
1071: In the case of continuous variable Gaussian states we have shown 
1072: that the estimation of the least symplectic eigenvalue $\tilde{d}_-$ of
1073: the covariance matrix may be performed with arbitrary precision 
1074: at fixed number of measurements, independently on the value of $\tilde{d}_-$
1075: itself and for both pure states and mixed states. 
1076: If we rather introduce other measures of entanglement proposed in
1077: literature, in particular the logarithmic negativity for pure
1078: states and the one based on the Bures distance \cite{EntBuresMarian1}
1079: for the symmetric squeezed thermal (mixed) states, we observe the same behavior
1080: obtained in the discrete variable case: the estimation is efficient only
1081: for maximally entangled state and inherently inefficient for weakly
1082: entangled states.  Therefore it is apparent that for continuous variable
1083: systems, the efficiency of the estimation strongly depends on the
1084: measure one decides to adopt.
1085: \par
1086: In conclusion, upon exploiting the geometric theory of quantum
1087: estimation we have quantitatively evaluated the ultimate bounds posed by
1088: quantum mechanics to the precision of entanglement estimation for
1089: several families of quantum states. To this aim we used the quantum
1090: Cramer-Rao theorem and the explicit evaluation of the quantum Fisher
1091: information matrix.  We have also given a recipe to build the
1092: observable achieving the ultimate precision in terms of the symmetric
1093: logarithmic derivative.  The analysis reported in this paper  makes an
1094: important point of principle and may be relevant in the design of
1095: quantum information protocols based on the entanglement content of
1096: quantum states.  Finally, we notice that our approach may be generalized
1097: and applied to the estimation of other quantities not corresponding to
1098: proper quantum observables, as the purity of a state or the coupling 
1099: constant of an interaction Hamiltonian \cite{ZP07,MK08}.  Work along 
1100: this lines is in progress and results will
1101: be reported elsewhere. 
1102: %%%%%%%%%%%%%
1103: \section*{Acknowledgments}
1104: This work as been partially supported by the CNR-CNISM convention.
1105: %%%%%%%%%%%%%
1106: \begin{thebibliography}{99}
1107: \bibitem{EntR} S. L. Braunstein et al., Rev. Mod. Phys. {\bf 77}, 513 (2005);
1108: R Horodecki et al., arXiv:quant-ph/0702225.
1109: \bibitem{Aud06}	K. Audenaert et al., New J. Phys. {\bf 8}, 266 (2006).
1110: \bibitem{Che03} K. Chen, Q. Inf. Comp. {\bf 3}, 193 (2003).
1111: \bibitem{Eis07} J. Eisert et al., New J. Phys. {\bf 9}, 46 (2006)
1112: \bibitem{Guh04} O. Guhne et al., Phys. Rev. Lett. {\bf 92}, 117903 (2004)
1113: \bibitem{Guh07} O. Guhne et al., Phys. Rev. Lett. {\bf 98}, 110502 (2007)
1114: \bibitem{Sun07} F. W. Sun et al., Phys. Rev. A {\bf 76}, 052303 (2007).
1115: \bibitem{Lou06} P. Lougovski et al., Eur. Phys. J. D {\bf 38}, 423 (2006).
1116: \bibitem{Wal06} S. P. Walborn et al., Nature {\bf 440}, 1022 (2006).
1117: \bibitem{Ren04} D. M. Ren, Comm. Theor. Phys. {\bf 42}, 33 (2004).
1118: \bibitem{Nav08} M. Navascues, Phys. Rev. Lett. {\bf 100}, 070503 (2008).
1119: \bibitem{Hor03} P. Horodecki, Phys. Lett. A {\bf 319}, 1 (2003).
1120: \bibitem{Aci00} A. Acin et al., Phys. Rev. A {\bf 61}, 062307  (2000);
1121: J. M. Sancho et al., Phys. Rev. A {\bf 61}, 042303 (2000).
1122: \bibitem{Hor02} P. Horodecki et al., Phys. Rev. Lett. {\bf 89}, 127902 (2002).
1123: \bibitem{Dur01} W. Dur et al., J. Phys. A {\bf 34}, 6837 (2001).
1124: \bibitem{Bar03} M. Barbieri et al., Phys. Rev. Lett. {\bf 91}, 227901 (2003).
1125: \bibitem{Guh03} O. Guhne et al., J. Mod. Opt. {\bf 50}, 1079 (2003).
1126: \bibitem{Dar03} G. M. D'Ariano et al., Phys. Rev. A {\bf 67}, 04230 (2003).
1127: \bibitem{Pla02} F. Plastina et al., J. Mod. Opt. {\bf 49}, 1389 (2002).
1128: \bibitem{Pit03} A. O. Pittenger et al., Phys. Rev. A {\bf 67}, 012327 (2003).
1129: \bibitem{QET} C. W. Helstrom, {\em Quantum Detection and Estimation 
1130: Theory} (Academic Press, New York, 1976); A.S. Holevo, {\em Statistical 
1131: Structure of Quantum Theory}, Lect. Not. Phys {\bf 61}, (Springer, Berlin,
1132: 2001).
1133: \bibitem{Hel67} C. W. Helstrom, Phys. Lett. A {\bf 25}, 1012 (1967). 
1134: \bibitem{BC94} S. Braunstein and C. Caves, Phys. Rev. Lett. {\bf 72}, 3439 (1994).
1135: \bibitem{BC96} S. Braunstein, C. Caves, and G. Milburn, Ann. Phys. {\bf 247}, 135 (1996).
1136: \bibitem{Mon06} A. Monras, Phys. Rev. A {\bf 73}, 033821 (2006).
1137: \bibitem{Sar06} M. Sarovar and G. Milburn, J. Phys. A {\bf 39}, 8487 (2006).
1138: \bibitem{Hot06} M. Hotta et al., Phys. Rev. A {\bf 72}, 052334 (2006).
1139: \bibitem{Mon07} A. Monras, M. G. A. Paris, Phys. Rev. Lett. {\bf 98}, 160401 (2007).
1140: \bibitem{Fuj01} A. Fujiwara, Phys. Rev. A {\bf 63},  (2001).
1141: \bibitem{Zhe06} J. Zhenfeng et al., preprint LANL quant-ph/0610060
1142: \bibitem{Boi08} S. Boixo, A. Monras, Phys. Rev. Lett. {\bf 100}, 100503 (2008).
1143: \bibitem{ZP07} P. Zanardi, M. G A Paris, arXiv:0708.1089
1144: \bibitem{NVW} G. Vidal and R. Werner, Phys. Rev. A {\bf 65}, 032314 (2002).
1145: \bibitem{HorodPPT} P. Horodecki, Phys. Lett. A {\bf 232}, 333 (1997). 
1146: \bibitem{Cra46} H. Cramer, {\em Mathematical Methods of Statistics} 
1147: \bibitem{Fuj03} A. Fujiwara and H. Imai, J. Phys. A {\bf 36}, 8093 (2003).
1148: \bibitem{HorodSep} M. Horodecki et al., Phys. Lett. A {\bf 223}, 1 (1996).
1149: \bibitem{PeresSep} A. Peres, Phys. Rev. A {\bf 54}, 2685 (1996). 
1150: \bibitem{HoffmanLUR} H. F. Hofmann and S. Takeuchi, Phys. Rev. A {\bf 68}, 032103 (2003). 
1151: \bibitem{HoffmanLURPPT} H. F. Hofmann, Phys. Rev. A {\bf 68}, 034307 (2003).
1152: \bibitem{GaussiansAOP} A. Ferraro, S. Olivares, and M. G. A. Paris, {\em Gaussian 
1153: States in Quantum Information} ((Napoli Series on Physics and Astrophysics. 2005), ).
1154: \bibitem{GaussianCrit} R. Simon, Phys. Rev. Lett. {\bf 84} 2726 (2000); 
1155: L. Duan et al., Phys. Rev. Lett. {\bf 84}, 2722 (2000).
1156: \bibitem{EntBuresMarian1} P. Marian, T. A. Marian, arXiv:quant-ph/0705.1138v2
1157: \bibitem{EoFGiedke} G. Giedke et al., Phys. Rev. Lett. {\bf 91}, 107901 (2003). 
1158: (Princeton University Press, 1946).
1159: \bibitem{ntw} A. Serafini et al., Phys. Rev A {\bf 69}, 022318 (2004); 
1160: S. Olivares, M. G. A. Paris, J. Opt. B {\bf 7}, 392 (2005).
1161: \bibitem{MK08} M. Korbman, C. Invernizzi, L. Campos, M. G. A. Paris, in
1162: preparation. 
1163: \end{thebibliography}
1164: %%%%%%%%%%%%%%%%%%%
1165: \end{document}
1166: