1: \documentclass{iopart}
2: \usepackage{amsmath,graphicx,bbm,iopams,mathrsfs,psfrag}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: \newcommand{\bra}[1]{\langle #1 \vert}
5: \newcommand{\ket}[1]{ \vert#1\rangle }
6: \newcommand{\bmsigma}{\boldsymbol \sigma}
7: \newcommand{\h}{{\rm h}}
8: \newcommand{\bfp}{\boldsymbol p}
9: \newcommand{\dmat}{{\varrho}}
10: \newcommand{\qobs}{{A}}
11: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
12: \begin{document}
13: \title{Bayesian estimation of one-parameter qubit gates}
14: \author{Berihu Teklu}%\email{berihu.gebrehiwot@unimi.it}
15: \address{Dipartimento di Fisica, Universit\`a di Milano,
16: I-20133 Milano, Italy}
17: \author{Stefano Olivares}%\email{stefano.olivares@mi.infn.it}
18: \address{ CNISM, UdR Milano Universit\`a, I-20133 Milano, Italy
19: \\ Dipartimento di Fisica, Universit\`a di Milano, I-20133 Milano, Italy}
20: \author{Matteo G. A. Paris}%\email{matteo.paris@unimi.it}
21: \address{
22: Dipartimento di Fisica, Universit\`a di Milano, I-20133 Milano, Italy
23: \\CNISM, UdR Milano Universit\`a, I-20133 Milano, Italy
24: \\Institute for Scientific Interchange Foundation, I-10133 Torino, Italy}
25: \date{\today}
26: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
27: \begin{abstract}
28: We address estimation of one-parameter unitary gates for qubit systems
29: and seek for optimal probes and measurements. Single- and two-qubit
30: probes are analyzed in details focusing on precision and stability of
31: the estimation procedure. Bayesian inference is employed and compared
32: with the ultimate quantum limits to precision, taking into account the
33: biased nature of Bayes estimator in the non asymptotic regime. Besides,
34: through the evaluation of the asymptotic {\em a posteriori} distribution
35: for the gate parameter and the comparison with the results of Monte
36: Carlo simulated experiments, we show that asymptotic optimality of Bayes
37: estimator is actually achieved after a limited number of runs. The
38: robustness of the estimation procedure against fluctuations of the
39: measurement settings is investigated and the use of entanglement to
40: improve the overall stability of the estimation scheme is also analyzed
41: in some details.
42: \end{abstract}
43: %\pacs{03.65.Wj, 03.65.Ta, 02.50.Tt}
44: %\maketitle
45: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
46: \section{Introduction}\label{s:intro}
47: Let us consider a system prepared in a known quantum state which enters
48: an apparatus performing an operation on the state. The evolution imposed
49: by the apparatus depends on the value of some parameters and the
50: experimenter is interested in the estimation of those parameters.
51: A natural strategy to obtain the parameter is to detect the state at
52: the output and infer the value of the
53: parameters from the global sample coming from a number of repeated
54: measurements. The optimization of this strategy, {\em i.e.} the choice
55: of the best probe, measurement and data processing, is the subject of quantum
56: parameter estimation, which is a relevant subject in the quantum
57: characterization of states and operations \cite{Hel76,Hol82}. The
58: operation on the state may be unitary or not \cite{ChuNie,PoyCirZol97}
59: and may depend on one or more unknown parameters, which, in turn, may
60: correspond to quantities that are not directly observable.
61: The parameters of interest may be the amplitude of the carrier
62: signal, the position and orientation of an object, or the strength of an
63: external fields. Communications, image analysis and precision metrology
64: provide relevant examples.
65: In the simplest scenario, a parameter estimation problem
66: consists in the determination of the value of the interaction parameter
67: $\theta$ for unitaries of the form $U_{\theta} =\lbrace -i \theta
68: G\rbrace$ where $G$ is a Hamiltonian operator that generates the
69: transformation.
70: \par
71: Generally speaking, quantum estimation is concerned with the problem of
72: finding optimal ways to estimate quantum states and processes. In turn,
73: it has recently attracted much interest in quantum information
74: \cite{Braunstein1994, Hol2004,Gill,Barnd,Fuji1995} as a tool for
75: characterization of signal and gates at the quantum level
76: \cite{qsm,Braunstein1,Braunstein2,Braunstein3,gentomo,mkq,Boixo,EKnill,Giovannetti}.
77: The canonical way to address estimation of states and operation is by
78: quantum tomography, (see \cite{LNP649} for a review) {\em i.e} by
79: measuring a complete set (a {\em quorom} \cite{gentomo}) of observables,
80: which allows or the complete characterization. For single- and two-qubit
81: systems this involves the measurement of Pauli matrices and has been
82: realized for polarization qubits \cite{kw2,qbmi}. Process Tomography,
83: {\em i.e} the reconstruction of quantum operations \cite{op1,op2,op3},
84: is itself critical for verifying the actions of quantum logic gates
85: \cite{op4} and characterizing decoherence processes \cite{op5}.
86: On the other hand,
87: there are many situations where the full tomography of signals and
88: devices is not needed, either because the focus is on specific features
89: of the transformation, or the dynamics is partially known. In this cases
90: the relevant point is to find an optimal and stable way to achieve
91: quantum characterization by parameter estimation. For this reason, in
92: this paper we address estimation of one-parameter unitary gates for
93: qubit systems, {\em i.e.} transformation of the form $U_{\theta}
94: =\lbrace -i \theta G\rbrace$ where $G$ is a combination of Pauli
95: operators and $\theta$ is the parameter of interest. We consider the
96: gate probed either by one-qubit and two-qubit states and compare the
97: performances of standard measurements with the ultimate quantum limit to
98: precision (accuracy) of estimation. As we will see, ultimate bounds are
99: determined by the initial quantum state of the probe, the type of
100: interaction and the readout measurements that is used to extract
101: information from the probe. In particular, we are going to assess the
102: performances of Bayes estimators, which themselves play a central role
103: in many signal processing problems \cite{VTB}.
104: \par
105: The precision (variance) of any unbiased estimator of a parameter
106: $\theta$ is limited by the Cram\'er-Rao bound (CR), given by the inverse
107: of the Fisher information \cite{Gill,Cramer,Rao,Poor,qi,H1,H2}. Bayes
108: estimators are known to be asymptotically unbiased and, in turn, to
109: saturate CR asymptotically. For measurements that are related to the
110: unknown $\theta$ through a linear Gaussian model, the maximum likelihood
111: estimate of $\theta$ also achieves the CR. Furthermore, when $\theta$ is
112: estimated from independent identically distributed (iid) measurements,
113: under suitable regularity assumptions on the probability density, the
114: maximum likelihood estimator is asymptotically unbiased and achieves the
115: CR \cite{Rao,Leh}. On the other hand, being interested in realistic
116: measurement schemes, here we consider estimation procedure based on a
117: limited number of measurements. As a consequence, we have to take into
118: account the biased nature of Bayes estimators. The variance of any
119: estimator with a given bias is bounded by the biased CR
120: \cite{VTB1,Schi}, which is an extension of the CR taking into account
121: the {\em a priori} distribution of the parameter of interest. In turn,
122: it is a fundamental rule of estimation theory that the use of prior
123: knowledge leads to a more accurate estimator.
124: \par
125: In this paper we address the estimation of the interaction parameter of
126: unitary qubit transformations. We derive ultimate quantum limits to
127: precision and assess performances of Bayesian estimators \cite{zd,pz}.
128: In particular,
129: we focus our attention on measurement schemes as those in Fig.
130: \ref{fig:qubit} and Fig. \ref{fig:bell}, where a single-qubit gate is
131: probed by single- and two-qubit probes, respectively. We evaluate the a
132: posteriori distribution for the gate parameter, derive the ultimate
133: bound on precision, and compare the asymptotic performances of Bayes
134: estimator to that of Monte Carlo simulated experiments, thus showing
135: that asymptotic optimality is achieved after a limited number of runs.
136: The paper is structured as follows. In Section \ref{s:estimation} we
137: introduce notation and derive the a posteriori distribution, also
138: discussing the Bayesian version of the CRB. In Section \ref{s:single} we
139: discuss limits to precision in estimating unitary gates for qubit
140: systems. A comparison between single- and two-qubit entangled probes shows
141: that entanglement improves the overall stability of the estimation
142: procedure. We also compare the asymptotic a posteriori
143: distribution for the gate parameter to the results of the Monte Carlo
144: simulated experiments. Section \ref{s:remarks} closes the paper with
145: some concluding remarks.
146: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
147: \section{Parameter estimation of one-parameter qubit gates}
148: \label{s:estimation}
149: A generic unitary transformation acting on a qubit state can be written
150: as $U({\boldsymbol\theta}) = \exp\lbrace - \frac{i}{2}
151: {\boldsymbol\theta}\cdot{\bmsigma}\rbrace$,
152: where ${\bmsigma}=\lbrace\bmsigma_1,\bmsigma_2,\bmsigma_3\rbrace$ is
153: the vector of Pauli matrices and ${\boldsymbol\theta}$ is a vector
154: describing
155: the transformation. In the following we will assume that ${\boldsymbol\theta}$
156: depends on a single parameter and refer to it as to $\theta$. The
157: simplest scheme for the estimation of $\theta$ \cite{Hel76} consists in
158: choosing a {\em probe} pure qubit state $\ket{\psi_0}$ undergoing the
159: transformation and a measurement onto the evolved state
160: $\ket{\psi_\theta} \equiv U({\theta})\ket{\psi_0}$. Here we assume that
161: the measurement can be represented by the two {\em projectors} $\Pi_0$
162: and $\Pi_1$, so that the {\em conditional} probabilities to obtain the
163: outcomes ``0'' or ``1'' ({\em i.e.} given $\theta$) are $P(j|\theta) =
164: \bra{\psi_\theta}\Pi_j\ket{\psi_{\theta}}$, $j=0,1$.
165: After $M$ measurements on
166: equally prepared qubits we have the the sample $X = \{x_1,\ldots, x_M\}$
167: of outcomes, where the $x_k$'s can take the values ``0'' and ``1''. This
168: leads us to define the following sample probability or likelihood
169: function:
170: \begin{equation}
171: P(X|\theta) = \prod_{k=1}^{M} P(x_k|\theta).
172: \end{equation}
173: Our estimation problem is that of inferring the value of $\theta$ once
174: the sample of outcomes is assigned by the measurement; in other words,
175: we are interested in the conditional ({\em a posteriori}) probability
176: $P(\theta|X)$ of $\theta$ {\em given} the sample $X$. This can be easily
177: obtained by the Bayes theorem, which states that \cite{JMB,DM}
178: $P(\theta|X) P(\theta) = P(X|\theta) P(X)$ where $P(\theta)$ is the
179: prior probability and $P(X)$ is the overall (unconditional) probability
180: of the observed sample. Hence, the {\em a posteriori} distribution
181: may be written as
182: \begin{equation}\label{cond:th}
183: P(\theta|X) = \frac{1}{N} \prod_{k=1}^{M} P(x_k|\theta)\,, \qquad
184: {N} = \int_{\Omega}\!\!\! d\theta \prod_{k=1}^{M} P(x_k|\theta)\,,
185: \end{equation}
186: where ${N}$ is the normalization and $\Omega$ is the set of
187: possible values for $\theta$.
188: From (\ref{cond:th}) we may evaluate the expected value of
189: $\theta$ and the variance of the distribution
190: \begin{equation}
191: \overline{\theta} =
192: \int_{\Omega}\!\!\! d\theta \, \theta \, P(\theta|X),
193: \quad \quad
194: {\rm Var}[\theta] =
195: \int_{\Omega}\!\!\! d\theta \, (\theta-\overline{\theta})^2 \,
196: P(\theta|X)\:.
197: \end{equation}
198: The mean (expected) value $\overline{\theta}$ of the a posteriori
199: distribution is our Bayesian estimator.
200: \par
201: For a large number of measurements, $M \gg 1$, and assuming
202: that the true value of the parameter is $\theta^*$, the number of
203: times a factor $P(x|\theta)$, with $x = 0, 1$, appears in the product
204: (\ref{cond:th}) is approximately given by $P(x|\theta^*) M$. The
205: asymptotic a posteriori distribution for the parameter $\theta$, conditioned
206: on the true value $\theta^*$ is thus given by \cite{Hradil1995,Hradil1996}
207: \begin{equation} \label{LF}
208: P_M(\theta |\theta^{*}) = \frac{1}{N}\prod_{s=0,1}
209: P(s|\theta)^{P(s|\theta^{*}) M}\,.
210: \end{equation}
211: Since $\sum_{h=0,1} P(h|\theta) = 1$ we have
212: $ \partial_\theta P_M(\theta|\theta^{*})\left.
213: \vphantom{\frac{}{}}\right|_{\theta^{*}} =0$ and
214: $\partial^2_\theta P_M(\theta|\theta^{*})\left.
215: \vphantom{\frac{}{}}\right|_{\theta^{*}} < 0$, {\em i.e} the distribution
216: $P_M(\theta|\theta^*)$ has the desirable property of showing a maximum
217: at the true value of the parameter, {\em i.e.} Bayesian estimator is
218: asymptotically unbiased.
219: \par
220: According to the Laplace-Bernstein-von Mises theorem \cite{LeCam} the a posteriori
221: distribution of Eq.~(\ref{LF}) may be asymptotically approximated by a Gaussian
222: with variance given by $\sigma^2 = \left[M G(\theta^*)\right]^{-1}$
223: where we have introduced the {\em Fisher information} \cite{qi}:
224: \begin{equation}
225: G(\theta) = \sum_{s=0,1} \frac{1}{P(s|\theta)}\,
226: \left(\frac{d P(s|\theta)}{d\theta}\right)^{2}
227: \label{deffisher}
228: \end{equation}
229: The asymptotic a posteriori distribution is thus
230: completely characterized by its variance, or equivalently by
231: the Fisher information which itself gives a lower bound to
232: the variance of any unbiased estimator
233: $\hat\theta(X)$ via the Cram\'er-Rao inequality \cite{Gill,qi,H1,H2}:
234: \begin{equation}
235: {\rm Var}_\theta[\hat\theta] \ge \frac1{M G(\theta)}\:. \label{cramerrao}
236: \end{equation}
237: Any estimator saturating the inequality (\ref{cramerrao}) is referred
238: to as an {\em efficient} estimator. The relation $\sigma^2 = \left[
239: M G(\theta^*)\right]^{-1}$ thus says that Bayesian estimator is
240: asymptotically efficient, whereas this conclusion does not hold for
241: finite $M$. For finite $M$ Bayes estimator is biased and we have to
242: generalize Eq.~(\ref{cramerrao}) to take into account the bias. To this
243: aim one considers the conditional expectation of the error $B(\theta)=
244: \int dx\, [\hat\theta(x) -\theta]P(x,\theta)$, where $P(x,\theta)=P(x|\theta)
245: P(\theta)$ is the joint probability of the data and the parameter. Of
246: course, for unbiased estimators $B(\theta)=0$. Starting from the definition of
247: $B(\theta)$ one derives the so-called van Trees inequality \cite{VTB1} for
248: the mean squared error
249: \begin{equation}\label{VT:ineq}
250: \overline{{\rm Var[\hat \theta(x)-\theta]}}
251: = \int d\theta\,P(\theta) \int dx\, [\hat\theta(x)-\theta]^2 P(x|\theta)
252: \geqslant
253: \frac1{H_M(\theta)}\:,
254: \end{equation}
255: where we introduced the generalized Fisher information
256: $H_M(\theta)={F(\theta)+M\,G(\theta)}$, $G$ being the Fisher information of
257: Eq. (\ref{deffisher}), $M$ the number of repeated measurements, and $F$
258: the Fisher information of the prior, {\em i.e.} $F(\theta)=\int d\theta\,
259: \left[\partial_\theta \log P(\theta) \right]^{2}P(\theta)$.
260: Eq.~(\ref{VT:ineq}) takes into account the information due to the prior
261: and thus gives a lower bound than the CR one, which is anyway achieved
262: for $M\gg 1$. On the other hand, one may show that \cite{Schi}:
263: \begin{align}
264: H_M(\theta) &= \int\!\! dX\, P(X) \int\!\! d\theta\,
265: \left[ \partial_\theta \log P(\theta | X)\right]^2 P(\theta | X)
266: \nonumber \\
267: & \stackrel{M\gg 1}{=} F_M(\theta|\theta^*) \equiv \int\!\! d\theta
268: P_M(\theta|\theta^* ) \left[ \partial_\theta \log P_M(\theta|\theta^* )
269: \right]^2\,,
270: \end{align}
271: where $P(\theta | X)$ is the Bayesian probability distribution
272: of Eq. (\ref{cond:th}) and $P_M(\theta|\theta^* )$ its asymptotic
273: expression of Eq. (\ref{LF}). Thus, Eq.~(\ref{VT:ineq}) represents the
274: Bayesian counterpart of the CR.
275: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
276: \section{Bayesian estimation of one-parameter qubit gates}
277: \label{s:single}
278: In this section, without loss of generality, we address the case in which
279: $\boldsymbol{\theta}=(0,0,\theta)$, $\theta\in[0, \pi]$, i.e., $U_3(\theta) =
280: \exp(-\frac{i}{2}\theta\bmsigma_3)$. We first consider the gate probed
281: by a single-qubit state and jointly optimize the probe and the
282: measurement and then address the use of entanglement, showing that it
283: may be useful to improve the overall stability of the estimation
284: procedure.
285: \subsection{Estimation via sinqle-qubit probes}
286: In Fig.~\ref{fig:qubit} we schematically depict the estimation
287: procedure: a pure state $\ket{\psi_0}$ undergoes the unitary
288: transformation $U_3(\theta)$ and, then, is measured by means of a
289: projective two-outcome device. Our aim is that of optimizing
290: the estimation of $\theta$ by a suitable choice of both the probe state
291: and the projective measurement.
292: %%%%
293: \begin{figure}[h]
294: \includegraphics[width=0.6\textwidth]{Sch1.eps}
295: \caption{Estimation of qubit gates. The state $\lbrace\rho_{0}\rbrace$
296: is prepared, then enters the gate described by the unitary $U_3(\theta)$
297: with unknown $\theta$, and finally is detected by a detector described
298: by a two-value POVM. \label{fig:qubit}}
299: \end{figure}
300: %%%%
301: \par
302: Upon writing the pure qubit state in the standard way
303: \begin{equation}
304: |\psi_{0}\rangle=\cos \frac\alpha {2}|0\rangle
305: + e^{i \phi} \sin \frac \alpha {2}|1\rangle\:,
306: \end{equation}
307: where $|0\rangle$ and $|1\rangle$ are the two eigenvectors of
308: $\bmsigma_3$, the parameters $\alpha\in[0,\pi]$ and $ \phi\in[0,2\pi]$
309: uniquely determine $|\psi_{0}\rangle$ and the evolution under the unitary
310: $U_3(\theta)$ is straightforward.
311: Next, we perform the measurement described by the two projectors
312: \begin{align}
313: \Pi_0(\beta,\omega) =
314: \ket{\Psi(\beta,\omega)}\bra{\Psi(\beta,\omega)}\qquad
315: \Pi_1(\beta,\omega) = {\mathbbm 1} - \Pi_0(\beta,\omega),
316: \end{align}
317: where $|\Psi(\beta,\omega)\rangle=\cos \frac\beta {2}|0\rangle
318: + e^{i \omega} \sin \frac \beta {2}|1\rangle$. The probabilities of
319: the two outcomes are given by
320: $P(0|\theta) \equiv P_0(\alpha,\beta,\phi,\omega,\theta) =
321: |\langle 0|U_{\theta}|\psi_{0}\rangle|^{2}$ {\em i.e}
322: \begin{align}
323: \label{prob1}
324: P(0|\theta) &= \frac12 \left[1 + \cos\alpha\cos\beta - \cos
325: (\phi-\omega+\theta) \sin\alpha\sin\beta\right] \nonumber \\
326: P(1|\theta) &= 1-P(0|\theta)\:.
327: \end{align}
328: The Bayesian a posteriori distribution of Eq. (\ref{cond:th}) may be written as
329: \begin{align}
330: \label{ge}
331: P(\theta|M) = \frac1N \: P(0|\theta)^{m_0}\:P(1|\theta)^{m_1}\:,
332: \end{align}
333: where $m_j$ is the number of measurements with outcomes
334: $j=0,1$, $m_0+m_1=M$, in the observed sample.
335: For a large number of measurements $M\gg 1$, we can evaluate
336: $P_M(\theta|\theta^*)$ using Eq. (\ref{LF}) and, in turn, the
337: expectation $\overline\theta$ and the variance ${\rm Var}[\theta]$.
338: Upon expanding on $\alpha$ and $\beta$ up to second
339: order one sees that ${\rm Var}[\theta]$ achieves its minimum for the
340: choice $\alpha = \beta = \pi/2$, independently on the value of $\phi$ and
341: $\omega$. This is confirmed by the evaluation of the corresponding
342: Fisher information $G (\theta)$,
343: \begin{align}
344: G(\theta) = \frac{\sin^2 \alpha \sin^2 \beta
345: \sin^2(\theta+\phi-\omega)}{1-\cos^2 \alpha \cos^2 \beta -
346: \cos^2(\theta+\phi-\omega)\sin^2 \alpha \sin^2 \beta }\:,
347: \end{align}
348: which achieves its maximum $G=1$ for $\alpha = \beta = \pi/2$,
349: independently on the value of $\phi$ and $\omega$. Using this
350: results we have:
351: \begin{align}\label{ga}
352: P_{M}(\theta|\theta^{*})=\frac1{N}
353: \exp\left[M\left(\cos^{2}\frac{\theta^{*}}2\log \cos^{2}\frac{\theta}2
354: +\sin^{2}\frac{\theta^{*}}2\log\sin^{2}\frac{\theta}2\right)\right],
355: \end{align}
356: and the variance saturates the van Trees inequality, thus confirming
357: that Bayes estimator is asymptotically efficient.
358: In Fig.~\ref{f:uni} we report the ratio
359: $\overline{\theta}/\theta^{*}$ and variance
360: multiplied by ${H_M}$ [see Eq.~(\ref{VT:ineq})] as a function
361: of the number of measurements. As it apparent from the plots all
362: the curves approach one when the number of measurements increases.
363: In the asymptotic region $H_M\simeq M G(\theta) \simeq M$. \\
364: %%%%%%%%%%%%%
365: \begin{figure}[h!]
366: \psfrag{thst}{$\overline {\theta}/\theta^{*}$}
367: \psfrag{M}{$M$}
368: \psfrag{V}{${\rm Var}[\theta] {H_M}$}
369: \psfrag{300}[]{300} \psfrag{500}[]{500} \psfrag{700}[]{700}
370: %
371: \psfrag{1.0000}[]{1.000} \psfrag{1.0005}[]{}
372: \psfrag{1.0010}[]{1.001} \psfrag{1.0015}[]{}
373: \psfrag{1.0020}[]{1.002} \psfrag{1.0025}[]{} \psfrag{1.0030}[]{1.003}
374: \includegraphics[width=0.41\textwidth]{rav.eps}
375: \psfrag{M}{$M$} \psfrag{20}[]{20} \psfrag{40}[]{40} \psfrag{60}[]{60}
376: \psfrag{80}[]{80} \psfrag{100}[]{100}
377: %
378: \psfrag{1.}[]{1.} \psfrag{1.01}[]{1.01} \psfrag{1.02}[]{1.02}
379: \psfrag{1.03}[]{1.03} \psfrag{1.04}[]{1.04} \psfrag{1.05}[]{1.05}
380: $\qquad$\includegraphics[width=0.41\textwidth]{rvar.eps}
381: \caption{LogLinear plot of $\overline{\theta}/\theta^{*}$ (left)
382: and linear plot of ${\rm Var}[\theta]{H_M}$ (right) as a function
383: of the number of measurements $M$ for the estimation of the unitary
384: $U_3(\theta)$ and for different values of the true parameter. For both
385: plots the dotted-dashed line is for $\theta^{*}=0.8$, the dashed line
386: is for $\theta^{*}=1.2$ and the solid line is for $\theta^{*}=1.8$.}
387: \label{f:uni}
388: \end{figure} \par
389: %%%%%%%%%%%%%
390: Notice that the results here reported for $U_3(\theta)$ are actually valid for any other
391: unitary gate. This can be seen as follows: any unitary
392: $U_{\boldsymbol n}(\theta)=\exp\left( -i \bmsigma_n \theta \right)$ describing rotations
393: around the arbitrary axis $n$ may be written as $U_{\boldsymbol n}= O U_3(\theta)
394: O^\dag$ where $O$ is the rotation corresponding to the mapping
395: $3\rightarrow {\boldsymbol n}$. As a consequence, the optimal measurement
396: corresponds to the projectors $\Pi_0^\prime = O^\dag |\Psi(0,\omega)\rangle\langle
397: \Psi(0,\omega)| O$, $\Pi_1^\prime= 1 - \Pi_0^\prime$ and the optimal probe is
398: given by $O^\dag |0\rangle$.
399: \par
400: A question arises on whether the results reported above may be easily
401: implemented in practice. This concerns the stability of the measurement
402: rather than its precision. The point is the following: Suppose that for
403: some reasons the values of parameters $\alpha$ (probe) $\beta$ (measurement)
404: slightly deviate from the optimal settings. To which extent
405: the overall performances of the procedure are degraded? A convenient way
406: to address this issue is to make a perturbation analysis upon expanding
407: the Fisher information $G(\theta)$ around the optimal settings
408: $\alpha=\beta=\pi/2$ up to second order
409: \begin{align}
410: \label{expan}
411: G (\theta) & \simeq 1 -\frac1{\sin^2\theta}
412: \left[(\alpha-\pi/2)^2+(\beta-\pi/2)^2\right] + 2 \frac{\cos\theta}{\sin^2\theta}
413: (\alpha-\pi/2)(\beta-\pi/2)
414: \nonumber \\
415: & \stackrel{\theta\ll 1}{\simeq} 1 - \frac1{\theta^2}
416: (\alpha-\beta)^2\:.
417: \end{align}
418: \begin{figure}[ht]
419: \centerline{\includegraphics[width=0.8\textwidth]{gm.eps}}
420: \caption{Contour plot of the Fisher information $G(\theta)$ as a function
421: of the probe and measurement parameters $\alpha$ and $\beta$ for three
422: different values of the gate parameter $\theta^*=0.05,0.1,1$. Darker
423: areas corresponds to higher values.}
424: \label{f:expan}
425: \end{figure} \par
426: Eq. (\ref{expan}) shows the quadratic decrease of $G$ out of the
427: optimal setting and, especially for small values of the gate parameter
428: $\theta$, the dramatic effect of a mismatch between the values of $\alpha$
429: and $\beta$. The latter is well illustrated in Fig. \ref{f:expan} where
430: $G$ as a function of $\alpha$ and $\beta$ is shown for different values of
431: $\theta$. As it is apparent from the plots a mismatch
432: $|\alpha-\beta|\sim \theta$ of the order of the parameter to be
433: estimated is enough to make the whole procedure ineffective.
434: Fortunately, as we will see in Section \ref{ss:ent}, the stability issue may
435: be overcome by using entangled probes and the optimal performances
436: still achieved by a two-qubit probe configuration.
437: %%%%%%%%%%%%%%
438: \subsection{Monte Carlo simulated experiments}
439: Before addressing stability of the measurement let us compare the
440: asymptotic a {\em posteriori} distribution for the gate parameter with
441: the results of Monte Carlo simulated experiments. This is in order
442: to locate the asymptotic region and make quantitative statements on the
443: achievability of the ultimate bounds to precision.
444: We have simulated $M$ repeated measurements using the optimal probe/measurement
445: settings and inserted the resulting values of $m_j$, $j=0,1$ in Eq.
446: (\ref{ge}) to obtain the a posteriori distribution for the gate parameter.
447: In Fig. \ref{f:sim} we plot the rescaled variance of this distribution together
448: with the variance of the asymptotic a posteriori distribution of Eq.
449: (\ref{ga}). Remarkably the asymptotic region is achieved after a limited
450: number of runs, thus proving that Bayesian approach may useful in practical
451: applications. In the inset we report the full distribution, both the
452: experimental a posteriori and the asymptotic one, for $M=20$ and $M=500$.
453: Notice that: i) for $M=500$ the a posteriori experimental distribution
454: is already indistinguishable from the asymptotic one and ii) the
455: asymptotic is already unbiased, {\em i.e} efficiency has been achieved
456: for a limited number of runs.
457: \begin{figure}[h!]
458: %\psfrag{q}{$\theta$}
459: \psfrag{P}{\tiny $P(\theta)$}
460: %\includegraphics[width=0.5\textwidth]{Fig6_t.eps}
461: \includegraphics[width=0.5\textwidth]{VarMC.eps}
462: \caption{Rescaled variance (variance multiplied by the generalized
463: Fisher information) of the a posteriori distribution as a function of the
464: number of measurements. Black line is for the asymptotic distribution
465: whereas gray squares are for the experimental one.
466: Inset: asymptotic (dashed line) and experimental
467: (solid line) Bayesian a posteriori distributions
468: for $\theta^{*}=0.6$ as obtained using optimal
469: single-qubit probe and $M=20$ (left) or $M=500$ (right) repeated
470: measurements.} \label{f:sim}
471: \end{figure}
472: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
473: \subsection{Estimation via two-qubit entangled probes}\label{ss:ent}
474: An alternative scheme for gate estimation may be designed using entangled
475: states as depicted in Fig.~\ref{fig:bell}. In fact, the use of entanglement
476: may improve estimation \cite{Fujiwara,Matteo,gargpe,entdec}. In this section we
477: investigate whether this is the case for the present estimation problem.
478: Basically, the use of entanglement increases the dimension of Hilbert
479: space and thus the number of possible outcomes of a measurement
480: performed on the perturbed signal. The corresponding Fisher information
481: does not increase but the maximum value is achieved for a large class
482: of probe signals. The Bayesian estimator is able to exploit this fact to
483: increase the overall stability of the estimation
484: procedure of the gate parameter.
485: \begin{figure}[h!]
486: \includegraphics[width=0.6\textwidth]{Sch2.eps}
487: \caption{Schematic diagram of parameter estimation by entangled qubit
488: probe: $\lbrace|\psi_{0}\rangle\rangle\rbrace$ is prepared and is
489: subjected to the unitary transformation $U_{\theta}$ on one qubit.
490: Finally, a projective measurement $\lbrace\Pi_{\alpha}\rbrace$ is
491: performed.}\label{fig:bell}
492: \end{figure}
493: \par
494: In order to see this behavior in practice let us consider the
495: estimation of the parameter of the gate $U_3 (\theta)$ using a
496: generic pure state of the form
497: \begin{align}
498: |\psi_0\rangle\rangle = \frac{1}{\sqrt{2}}\sum_{k=0}^3 c_k |\sigma_k\rangle\rangle
499: \end{align}
500: where we used the matrix notation for states
501: $|A\rangle\rangle\doteq\sum_{ij}A_{ij}|i\rangle|j\rangle\equiv
502: A\otimes{\mathbbm I}|{\mathbbm I}\rangle\rangle$.
503: As a measurement we consider the "Bell" measurement
504: made of the four projectors $\Pi_k=\frac12
505: |\sigma_k\rangle\rangle\langle\langle\sigma_k |$, $k=0,1,2,3$,
506: over a set of maximally entangled states.
507: After the evolution under the unitary $U_3(\theta)$ the four possible
508: outcomes of the measurements occur with the probabilities
509: \begin{align}
510: P_{0}(\theta)&= c_0^2 \cos^2\frac{\theta}{2} + c_3^2
511: \sin^2\frac{\theta}{2}\:, \quad
512: P_{1}(\theta)= \left(c_1 \cos \frac{\theta}{2} - c_2
513: \sin\frac{\theta}{2}\right)^2\:,
514: \nonumber\\
515: P_{3}(\theta)&= c_0^2 \sin^2\frac{\theta}{2} + c_3^2 \cos^2\frac{\theta}{2}
516: \:, \quad
517: P_{2}(\theta)= \left(c_1 \sin \frac{\theta}{2} + c_2
518: \cos\frac{\theta}{2}\right)^2\:.
519: \end{align}
520: The corresponding Fisher information is given by
521: $$
522: G(\theta) = c_1^2 +c_2^2 + \frac{(c_0^2-c_3^2)(c_0^4-c_3^4)\sin^2
523: \theta}{(c_0^2+c_3^2)-(c_0^2-c_3^2)\cos^2\theta}\, $$
524: and achieves its maximum ($G=1$, $\forall \theta$) for any state
525: with $c_0=0$ or $c_3=0$. Therefore, having fixed the Bell measurement at
526: the output, a possible deviation in the probe preparations is not
527: degrading the performances of the estimation procedure.
528: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
529: \section{Conclusions}
530: \label{s:remarks}
531: In this paper we have analyzed estimation of one-parameter unitary gates
532: for qubit systems. We have addressed Bayesian estimation procedures and
533: compared their performances with the ultimate quantum limits to
534: precision. Bayes estimator is known to be asymptotically unbiased, but
535: for practical implementation is of interest to evaluate quantitatively
536: how many measurements are needed to achieve the asymptotic region. To
537: this aim, after the evaluation of the asymptotic {\em a posteriori}
538: distribution for the gate parameter, we have compared it to the
539: distribution obtained by Monte Carlo simulated experiments and full
540: Bayesian analysis and shown that asymptotic optimality of Bayes
541: estimator is achieved after a limited number of runs. We have also
542: addressed the issue of stability, {\em i.e} the robustness of the
543: optimal settings against fluctuations of the probe and measurement
544: parameters. It has been shown that the use of entanglement is useful to
545: improve stability. More explicitly, we have shown that, although the
546: Fisher information does not increase, its maximum value is achieved
547: for a large class of probe signals, thus making the procedure more
548: robust and increasing the overall stability of the estimation procedure.
549: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
550: \section*{Acknowledgments}
551: The authors thank M. Borrelli, M. Zaro, and N. Tomassoni for useful
552: discussions. This work has been partially supported by the CNR-CNISM
553: convention. This article was completed at a time of drastic cuts to
554: research budgets imposed by the Italian government \cite{nn08}; as a
555: result research is becoming increasingly difficult in Italian universities
556: and may in the near future be brought to a complete halt.
557: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
558: \section*{References}
559: \begin{thebibliography}{99}
560: \bibitem{Hel76} C.~W.~Helstrom, {\em Quantum Detection and Estimation Theory}
561: (Academic Press, New York, 1976).
562: \bibitem{Hol82} A.~S.~Holevo, {\em Probabilistic and Statistical Aspects
563: of Quantum Theory} (North-Holland, Amsterdam, 1982).
564: \bibitem{ChuNie}I.~L.~Chuang and M.~A.~Nielsen, J. Mod. Opt. {\bf 44}, 2455 (1997).
565: \bibitem{PoyCirZol97}J.~F.~Poyatos, et al., Phys. Rev. Lett. {\bf 78}, 390 (1997).
566: \bibitem{Braunstein1994} S.~L.~Braunstein and C.~M.~Caves, Phys. Rev. Lett.
567: {\bf 72}, 3439 (1994).
568: \bibitem{Hol2004} A.~S.~Holevo, Prob. Theory Appl. {\bf 49}, 207 (2004).
569: \bibitem{Gill} R.~D.~Gill and S.~Massar, Phys. Rev. A {\bf 61}, 042312 (2000).
570: \bibitem{Barnd} O.~E.~Barndorff-Nielsen and R.~D.~Gill,
571: J. Phys. A {\bf 30}, 4481 (2000).
572: \bibitem{Fuji1995} A.~Fujiwara and H.~Nagaoka, Phys. Lett. A {\bf 201},
573: 119 (1995).
574: \bibitem{qsm} M. G. A. Paris, Phys. Rev. A {\bf 53}, 2658 (1996).
575: \bibitem{Braunstein1} S. L. Braunstein and C. M. Caves, in
576: {\em Fundamental problems in quantum theory}, Vol. 755 of
577: Annals of the New York Academy of Sciences, edited by
578: D. Greenberger and A. Zellinger (New York Academy of Sciences,
579: New York, 1995), pp. 786-797.
580: \bibitem{Braunstein2} S. L. Braunstein and C. M. Caves in {\em Quantum
581: communication and Measurement}, edited by V. P. Belavkin, O. Hirota, and
582: R. L. Hudson (Plenum, New York, 1995), pp. 21-30.
583: \bibitem{Braunstein3}S. L. Braunstein, C. M. Caves, and G. J. Milburn,
584: Ann. Phys. (N. Y.) {\bf 247}, 135 (1996).
585: \bibitem{gentomo} G. M. D'Ariano, L. Maccone, M. G. A. Paris, J.
586: Phys. A, {\bf 34}, 93 (2001).
587: \bibitem{mkq} K. Banaszek, G.M. D'Ariano, M.G.A. Paris, and M.F. Sacchi,
588: Phys. Rev. A {\bf 61} 10304 (2000).
589: \bibitem{Boixo} S. Boixo, et al., Phys. Rev. Lett. {\bf 98}, 090401 (2007).
590: \bibitem{EKnill}E. Knill, G. Ortiz, and R. D. Somma, Phys. Rev. A {\bf 75}, 012328 (2007).
591: \bibitem{Giovannetti} V. Giovannetti, S. Lloyd, and L. Maccone, Phys.
592: Rev. Lett. {\bf 96}, 010401 (2006).
593: \bibitem{LNP649} M.~G.~A.~Paris and J.~\v Reh\' a\v cek, {\em Quantum
594: State Estimation}, Lect. Notes Phys. {\bf 649} (2004).
595: \bibitem{kw2} D. F. V. James, P. G. Kwiat, W. J. Munro,
596: and A.G. White, Phys. Rev. A {\bf 64}, 052312 (2001).
597: \bibitem{qbmi} {S. Cialdi, F. Castelli, I. Boscolo, M. G. A. Paris},
598: Appl. Opt. {\bf 47}, 1832 (2008).
599: \bibitem{op1} A. M. Childs et al., Phys. Rev. A {\bf 64} 012314 (2001).
600: \bibitem{op2} M. W. Mitchell et al., Phys. Rev. Lett. {\bf 91} 120402 (2003).
601: \bibitem{op3} G. M. D'Ariano, P. Lo Presti, Phys. Rev. Lett. {\bf 91} 047902 (2003).
602: \bibitem{op4} J. L O'Brien et al., Phys. Rev. Lett. {\bf 93}, 080502 (2004).
603: \bibitem{op5} S. H. Myrskog et al., Phys. Rev. A {\bf 72} 013615 (2005).
604: \bibitem{VTB} H.~L.~Van Trees and K.~L.~Bell,
605: {\em Bayesian Bounds for Parameter Estimation and Nonlinear Filtering/Tracking},
606: (IEEE Press and Wiley Interscience, 2007).
607: \bibitem{Cramer} H.~Cramer, {\em Mathematical Methods of Statistics},
608: (Princeton University Press, 1946).
609: \bibitem{Rao} C.~R.~Rao, Proc. Cambridge Phil. Soc., {\bf 43}, 280 (1946).
610: \bibitem{Poor} H.~V.~Poor, {\em An Introduction to Signal Detection and
611: Estimation} (New York, Wiley, 1973).
612: \bibitem{qi} T.~M.~Cover and J.~A.~Thomas, {\em Elements of Information
613: Theory} (Wiley, New York, 2006).
614: \bibitem{H1} M. Hayashi in {\em Quantum Communication, Computing and
615: Measurement,} edited by O. Hirota, A. S. Holevo, and C. M. Caves,
616: (Plenum Publishing, New York, 1997).
617: \bibitem{H2} M. Hotta and M. Ozawa, Phys. Rev. A
618: {\bf 70}, 022327 (2004).
619: \bibitem{Leh}E. L. Lehmann and G. Casella, {\em Theory of Point
620: Estimation}, 2nd.ed (Springer, New York, 1994).
621: \bibitem{VTB1} H. L. van Trees, {\em Detection, Estimation, and Modulation
622: Theory I}, (Wiley, New York, 2001).
623: \bibitem{Schi} M. P. Sch\"utzenberger, Bull. Amer. Math. Soc.
624: {\bf 63}, 142 (1957).
625: \bibitem{zd} Z. Hradil et al., Phys. Rev. Lett. {\bf 76}, 4295 (1996).
626: \bibitem{pz} L. Pezze et al., Phys. Rev. Lett. {\bf 99}, 223602 (2007).
627: \bibitem{JMB} J. M. Bernardo and A. F. M. Smith, {\em Bayesian Theory}
628: (Wiley, Chichester, 1994).
629: \bibitem{DM} D. Malakoff, Science {\bf 286}, 1460 (1990).
630: \bibitem{Hradil1995}Z. Hradil, Phys. Rev. A {\bf 51}, 1875 (1995).
631: \bibitem{Hradil1996}Z. Hradil et al., Phys. Rev. Lett. {\bf 76}, 4295 (1996).
632: \bibitem{LeCam} L. Le Cam, {\em Asymptotic Methods in Statistical
633: Decision Theory} (Springer, New York, 1986).
634: \bibitem{Fujiwara}A. Fujiwara, Phys. Rev. A {\bf 63}, 042304 (2001).
635: \bibitem{Matteo} G. M. D'Ariano, P. Lo Presti, M. G. A. Paris,
636: Phys. Rev. Lett. {\bf 87}, 3439 (2001).
637: \bibitem{gargpe} G. M. D'Ariano, M. G. A Paris, P. Perinotti,
638: J. Opt. B, {\bf 4}, S277 (2002).
639: \bibitem{entdec} G. M. D'Ariano, M. G. A. Paris, P. Perinotti,
640: Phys. Rev A {\bf 65} 062106 (2002).
641: \bibitem{nn08} {\em Cut-throat savings} Nature {\bf 455}, 835 October (2008).
642: \end{thebibliography}
643: %%%
644: \end{document}
645: