1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: %
3: % ce.tex
4: %
5: % Appendix
6: %
7: % 18 Feb 2002 Masahide Sasaki
8: % 20 Mar 2002 Sasaki <--- Revision by Barnett
9: %
10: %
11: \documentclass[twocolumn,aps,showpacs,superscriptaddress,floatfix]
12: {revtex4}
13: %\documentclass[preprint,showpacs,superscriptaddress,floatfix]{revtex4}
14: \usepackage[dvips]{graphicx}
15: %\usepackage[dvipdfm,backref]{hyperref}
16: %
17: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
18: %
19: % definitions
20: %
21: % \let\begin=\begin
22: \newcommand{\bsquare}{\hbox{\rule{6pt}{6pt}}}
23: \newcommand{\reals}{\mbox{I$\!$R}}
24: \newcommand{\nums}[1]{\mbox{Z}_{#1}}
25: \newcommand{\ket}[1]{\left | #1 \right \rangle}
26: \newcommand{\bra}[1]{\left \langle #1 \right |}
27: \newcommand{\amp}[2]{\left \langle #1 | #2 \right \rangle}
28: \newcommand{\proj}[1]{\ket{#1} \bra{#1}}
29: \newcommand{\tr}{\mathrm{Tr}}
30: \newcommand{\be}{\begin{equation}}
31: \newcommand{\ee}{\end{equation}}
32: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
33: %
34: % Front matters
35: %
36: \begin{document}
37: %
38: \title{%
39: Optimal parameter estimation of depolarizing channel}
40: %
41: \author{Masahide Sasaki}
42: \email{psasaki@crl.go.jp}
43: \affiliation{Communications Research Laboratory,
44: Koganei, Tokyo 184-8795, Japan}
45: \affiliation{CREST, Japan Science and Technology Agency}
46: \author{Masashi Ban}
47: \affiliation{Advanced Research Laboratory, Hitachi Ltd, 1-280,
48: Higashi-Koigakubo, Kokubunnji, Tokyo
49: 185-8601, Japan}
50: \author{Stephen M. Barnett}
51: \affiliation{Department of Physics and Applied Physics,
52: University of Strathclyde, Glasgow G4 0NG, Scotland}
53: %
54: \begin{abstract}
55: We investigate strategies for estimating a depolarizing
56: channel for a finite dimensional system.
57: Our analysis addresses the double optimization problem of selecting
58: the best input probe state and the measurement strategy that minimizes
59: the Bayes cost of a quadratic function.
60: In the qubit case, we derive the Bayes optimal strategy for any finite
61: number of input probe particles when bipartite entanglement can be
62: formed in the probe particles.
63: \end{abstract}
64: %
65: \pacs{03.67.Hk, 03.65.Ta, 42.50.--p}
66: % 03.67.Hk Quantum communication
67: % 03.65.Ta Foundations of quantum mechanics; measurement theory
68: % 42.50.-p Quantum optics
69: %
70: \date{\today}
71: %
72: \maketitle
73: %
74: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
75: %
76: \section{Introduction}
77: \label{sec:intro}
78: %
79: In order to design a reliable communication system one requires
80: a priori knowledge of the property of a channel.
81: Precise knowledge of the channel allows us to devise
82: appropriate coding, modulation, and filtering schemes.
83: In general, the channel property is not stationary,
84: so one should first acquire and then track the optimal operating point
85: of each device by monitoring the condition of the channel.
86: It is important, therefore to know how to estimate the channel
87: property in an efficient way, that is,
88: as precisely as possible with minimum resources.
89:
90:
91: A reasonable assumption is that we know that the channel belongs to a
92: certain parameterized family, and only the values of the parameters
93: are not known. To know them one may input a probe system in an
94: appropriate state into the channel and make a measurement on the
95: output state.
96: Only when an infinite amount of input resource is available,
97: one can determine the channel parameters with perfect accuracy.
98: In the quantum domain, however, the resource is often restricted for
99: various reasons. For example, when one is to monitor a fast quantum
100: dymanics at cryogenic temperatures, the input probe power should be
101: kept as low as possible so as to prevent the system from heating
102: up while obtaining meaningful data in a short time.
103: This restricts the available amount of probe particles.
104: Furthermore, preparing the probe in an appropriate quantum state
105: is usually an elaborate process.
106: Thus to find the efficient estimation strategy relying only on
107: a restricted amount of input resource is of practical importance.
108:
109:
110: In estimating a quantum channel parameter, given a finite amount of
111: input resource,
112: both the input probe state and the measurement of the output state
113: need to be optimized.
114: This double maximization problem has been studied in the context of
115: estimation of SU($d$) unitary operation~
116: \cite{Acin01}.
117: Estimating a noisy quantum channel has been discussed in the
118: literature~
119: \cite{Fujiwara01,Fischer01,Cirone01}.
120: In ref. \cite{Fujiwara01}, the locally unbiased estimator and the
121: Cram\'er-Rao bound are extensively discussed for the depolarizing
122: channel for a qubit system. The locally optimal strategy, which
123: achieves the Cram\'er-Rao bound at a local point of the parameter
124: space was derived when two qubits at most are used.
125: This result would be useful in the limit of large ensemble of the
126: input probe. In such a limit, of course,
127: one can establish the channel parameter with a very high degree of
128: accuracy.
129: To improve the rate at which
130: the estimation accuracy grows with the number of probe particles,
131: one may first apply some preliminary estimation using a part of probe
132: particles to establish the most likely value of the parameter,
133: and then use the locally optimal strategy around this value to get
134: the final estimate~
135: \cite{Barndorff98,Gill00,Hayashi02}.
136: Refs. \cite{Fischer01,Cirone01} focus on several noisy qubit channels.
137: They study some reasonable, although not optimal, strategies based on
138: maximum likelyhood estimator, and derive the asymptotic behavior of
139: the cost as a function of the number of input probe qubits.
140:
141:
142: In contrast, we are concerned here with the Bayes optimal strategy
143: which minimizes the \textit{average} cost.
144: The scenario we have in mind is that one has no particular
145: knowledge about the a priori parameter distribution, and
146: the available number of probe particles is strictly limited.
147: We then take into account the possibility of rather large errors.
148: We seek the strategy that works equally well for all
149: possible values of the parameter on average,
150: that is, the strategy which is more universal for various possible
151: situations.
152:
153:
154: It seems difficult for us to study this problem for the most general
155: probe state.
156: In this paper we deal with the depolarizing channel by assuming that
157: we dispose of $M$ pairs of probe particles and
158: only bipartite entanglement can be formed in each pair.
159: This might be a practically sensible assumption from the view point of
160: optical implementation given current technology.
161: Our problem is to find the best estimation strategy to
162: minimize the average cost.
163: We consider the quadratic of a cost function.
164:
165:
166:
167:
168: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
169: \section{Qubit case}
170: \label{sec:qubit}
171: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
172:
173: Let $\hat\rho$ be a density operator in the 2 dimensional Hilbert
174: space ${\cal H}_2$.
175: The depolarizing channel ${\cal L}_\theta$ maps a density operator
176: $\hat\rho$ to a density operator which is a mixture of $\hat\rho$
177: and the maximally mixed state,
178: \be
179: {\cal L}_\theta \hat\rho = \theta \hat\rho
180: + \frac{1-\theta}{2}\hat I.
181: \ee
182: The parameter $\theta$ represents the degree of randomization of
183: polarization. For the map ${\cal L}_\theta$ to be completely
184: positive, the parameter $\theta$ must lie in the interval
185: $-{1\over3}\le\theta\le1$.
186:
187:
188: Let us start with two qubit systems as the input probe.
189: For simplicity we only consider a pure state family of the probe
190: $\hat\Psi=\proj\Psi$.
191: This may be represented in the Schmidt decomposition
192: \be
193: \ket{\Psi}=\sqrt{x}\ket0\otimes\ket{e_0}
194: +\sqrt{1-x}\ket1\otimes\ket{e_1},
195: \ee
196: where $\{\ket0,\ket1\}$ and $\{\ket{e_0},\ket{e_1}\}$ are orthonormal
197: basis sets for the first and second probe particle, respectively.
198: What is the best way to use this state?
199: There are two possibilities to consider;
200: \begin{itemize}
201: \item[(a)] \enskip
202: Input one qubit of the pair into the channel keeping the other
203: untouched leading to the output state
204: \be
205: \hat\Psi_1(\theta)
206: \equiv({\cal L}_\theta\otimes\hat I)\proj{\Psi},
207: \ee
208: \item[(b)] \enskip
209: Input both qubits into the channel and have the output state
210: \be
211: \hat\Psi_2(\theta)
212: \equiv({\cal L}_\theta\otimes{\cal L}_\theta)\proj{\Psi}.
213: \ee
214: \end{itemize}
215: A measurement is described by a probability operator measure (POM)
216: $\hat\Pi(\theta)$~
217: \cite{Helstrom_QDET,Holevo_book},
218: also referred to as a positive operator valued measure (POVM)
219: \cite{Peres_book}.
220: The average cost for the quadratic cost function is given by
221: \be
222: \bar C_i(x)
223: =\int_{-{1\over3}}^1d\tilde\theta
224: \int_{-{1\over3}}^1d\theta
225: (\tilde\theta-\theta)^2 z(\theta)
226: \tr\left[ \hat\Pi(\tilde\theta) \hat\Psi_i(\theta) \right],
227: \ee
228: where
229: $z(\theta)$ is the a priori probability distribution of $\theta$,
230: and $\int_{-{1\over3}}^1d\tilde\theta \hat\Pi(\tilde\theta)=\hat I$.
231: It is assumed that we have no a priori knowledge about $\theta$,
232: that is, $z(\theta)={3\over4}$.
233: Given the channel ${\cal L}_\theta$, we are to find the optimal
234: probe $\ket{\Psi}$ and the POM $\hat\Pi(\theta)$ minimizing
235: the average cost $\bar C(x)$.
236:
237:
238: It is convenient to introduce the \textit{risk} operator
239: \begin{eqnarray}
240: \hat W(\theta)
241: &=&{3\over4}\int_{-{1\over3}}^1d\theta'
242: (\theta-\theta')^2 \hat\Psi_i(\theta'), \\
243: &=&\hat W^{(2)} - 2\theta \hat W^{(1)} + \theta^2 \hat W^{(0)},
244: \end{eqnarray}
245: where
246: $\hat W^{(k)}\equiv{3\over4}\int_{-{1\over3}}^1d\theta
247: \theta^k \hat\Psi_i(\theta)$.
248: The average cost is then
249: \be
250: \bar C(x)=\tr\hat\Gamma, \quad
251: \hat\Gamma
252: \equiv
253: \int_{-{1\over3}}^1d\theta \hat\Pi(\theta) \hat W(\theta).
254: \ee
255: For a fixed probe state $\ket{\Psi}$,
256: the optimal POM $\hat\Pi(\theta)$ is derived
257: from the necessary and sufficient conditions to minimize the average
258: cost~
259: \cite{Holevo73_condition,YuenKennedyLax75}:
260: \begin{itemize}
261: \item[(i)] \enskip
262: $\hat\Gamma=\hat\Gamma^\dagger$, and
263: $\left[\hat W(\theta)-\hat\Gamma\right]\hat\Pi(\theta)=0$
264: for all $\theta$,
265: \item[(ii)] \enskip
266: $\hat W(\theta)-\hat\Gamma\ge0$ for all $\theta$.
267: \end{itemize}
268: The optimal solution for a single parameter estimation with a
269: quadratic cost is well known~
270: \cite{Personick71b,Helstrom_QDET}.
271: The optimal POM is constructed by finding the eigenstate
272: $\ket\theta$ of the \textit{minimizing} operator $\hat\Theta$
273: which is defined by
274: \be
275: \hat\Theta \hat W^{(0)} + \hat W^{(0)} \hat\Theta = 2 \hat W^{(1)},
276: \ee
277: that is, $\hat\Pi(\theta)=\proj\theta$
278: so that $\hat\Theta\ket\theta=\theta\ket\theta$.
279: We then have
280: $\hat\Gamma=\hat W^{(2)}-\hat\Theta \hat W^{(0)} \hat\Theta$
281: from which the conditions (i) and (ii) are easily verified.
282:
283:
284: For a discrete system, one can find the optimal POM with
285: finite elements.
286: Let the spectral decomposition of $\hat W^{(0)}$ for our two-qubit
287: system be
288: \be\label{W0}
289: \hat W^{(0)}=\sum_{i=1}^4 \omega_i \proj{\omega_i}.
290: \ee
291: Then the minimizing operator is
292: \be
293: \hat\Theta=\sum_{i,j=1}^4 \frac{2}{\omega_i+\omega_j}
294: \proj{\omega_i} \hat W^{(1)} \proj{\omega_j}.
295: \label{minmizing_op}
296: \ee
297: Let the spectral decomposition of $\hat\Theta$ be
298: \be
299: \hat\Theta=\sum_{i=1}^4 \theta_i \proj{\theta_i}.
300: \ee
301: The optimal POM is then given by
302: \be
303: \hat\Pi(\theta)=\sum_{i=1}^4 \delta(\theta-\theta_i) \proj{\theta_i}.
304: \ee
305: This implies that the measurement has 4 outputs at most and
306: we then estimate the channel parameter as one of 4 $\theta_i$'s.
307: Before going on to derive the optimal strategies, let us define some
308: notations. As seen below the output states $\hat\Psi_i(\theta)$'s
309: can be written as a direct sum
310: \be
311: \hat\Psi_i(\theta)=\hat\psi_i(\theta)\oplus\hat\phi_i(\theta),
312: \ee
313: where $\hat\psi_i(\theta)$ is in the subspace $\cal{H}_\psi$ spanned
314: by
315: $\ket{\mu_1}\equiv\ket0\otimes\ket{f_0}$ and
316: $\ket{\mu_2}\equiv\ket1\otimes\ket{f_1}$,
317: and $\hat\phi_i(\theta)$ in the subspace $\cal{H}_\phi$ spanned by
318: $\ket{\nu_1}\equiv\ket0\otimes\ket{f_1}$ and
319: $\ket{\nu_2}\equiv\ket1\otimes\ket{f_0}$.
320: In the following all 2$\times$2 matrices represent density operators
321: in $\cal{H}_\psi$ with
322: $\ket{\mu_1}=\left(\begin{array}{c} 1 \\ 0 \end{array}\right)$ and
323: $\ket{\mu_2}=\left(\begin{array}{c} 0 \\ 1 \end{array}\right)$.
324:
325:
326: \noindent
327: \textbf{Case (a)}:
328:
329: The output state $\hat\Psi_1(\theta)$ is given by
330: \begin{eqnarray}
331: \hat\psi_1(\theta)
332: &=&
333: {1\over2}
334: \left[
335: \begin{array}{cc}
336: (1+\theta)x & 2\theta\sqrt{x(1-x)} \\
337: 2\theta\sqrt{x(1-x)} & (1+\theta)(1-x)
338: \end{array}
339: \right],
340: \\
341: \hat\phi_1(\theta)
342: &=& \frac{1-\theta}{2}
343: \Bigl[ (1-x)\proj{\nu_1} + x \proj{\nu_2} \Bigr].
344: \end{eqnarray}
345: The elements of the risk operator are
346: \begin{eqnarray}
347: \hat W^{(0)}
348: &=&{1\over3}
349: \Bigl(
350: \left[
351: \begin{array}{cc}
352: 2x & \sqrt{x(1-x)} \\
353: \sqrt{x(1-x)} & 2(1-x)
354: \end{array}
355: \right]
356: \oplus \hat\varphi_1
357: \Bigr),
358: \\
359: \hat W^{(1)}&=&{1\over{27}}
360: \Bigl(
361: \left[
362: \begin{array}{cc}
363: 8x & 7\sqrt{x(1-x)} \\
364: 7\sqrt{x(1-x)} & 8(1-x)
365: \end{array}
366: \right]
367: \oplus \hat\varphi_1
368: \Bigr),
369: \\
370: \hat W^{(2)}&=&{1\over{27}}
371: \Bigl(
372: \left[
373: \begin{array}{cc}
374: 6x & 5\sqrt{x(1-x)} \\
375: 5\sqrt{x(1-x)} & 6(1-x)
376: \end{array}
377: \right]
378: \oplus \hat\varphi_1
379: \Bigr),
380: \end{eqnarray}
381: where
382: $\hat\varphi_1=(1-x)\proj{\nu_1} + x \proj{\nu_2}$.
383: After a lengthy but straightforward calculation
384: (see Appendix \ref{app_a})
385: we have
386: \be\label{Theta_a}
387: \hat\Theta
388: ={2\over9}
389: \left[
390: \begin{array}{cc}
391: 1+x & 2\sqrt{x(1-x)} \\
392: 2\sqrt{x(1-x)} & 2-x
393: \end{array}
394: \right]
395: \oplus {1\over9}\hat I_\phi.
396: \ee
397: To diagonalize $\Theta$ we introduce $r=\sqrt{1+12x(1-x)}$ and
398: \be
399: \mathrm{cos}\gamma=\sqrt{\frac{r-1+2x}{2r}},
400: \quad
401: \mathrm{sin}\gamma=\sqrt{\frac{r+1-2x}{2r}}.
402: \ee
403: The eigenstates and eigenvalues are then
404: \be\label{eigenvec1}
405: \begin{array}{lll}
406: \ket{\theta_1}&=\mathrm{cos}\gamma\ket{\mu_1}
407: +\mathrm{sin}\gamma\ket{\mu_2}, \quad
408: &\theta_1=(3+r)/9, \\
409: \ket{\theta_2}&=-\mathrm{sin}\gamma\ket{\mu_1}
410: +\mathrm{cos}\gamma\ket{\mu_2}, \quad
411: &\theta_2=(3-r)/9, \\
412: \ket{\theta_3}&=\ket{\nu_1}, \quad
413: &\theta_3=1/9, \\
414: \ket{\theta_4}&=\ket{\nu_2}, \quad
415: &\theta_4=1/9.
416: \end{array}
417: \ee
418: The average cost finally reads
419: \be
420: \bar C_1(x)=\tr(\hat W^{(2)}-\Theta\hat W^{(0)}\Theta)
421: ={8\over{81}}\left[1+(x-{1\over2})^2\right].
422: \ee
423: This is minimized by the maximally entangled state input
424: \be
425: \ket{\Psi}=\frac{1}{\sqrt2}
426: \left(\ket0\otimes\ket{f_0}+\ket1\otimes\ket{f_1}\right),
427: \ee
428: for which $\theta_1={5\over9}$ and
429: $\theta_2=\theta_3=\theta_4={1\over9}$. Therefore the optimal
430: measurement is actually constructed by the two projectors
431: \be\label{optimal POM}
432: \hat\Pi_1=\proj{\Psi}, \quad \hat\Pi_2=\hat I - \proj{\Psi},
433: \ee
434: with the associated guesses
435: $\theta_1={5\over9}$ and $\theta_2={1\over9}$, respectively.
436: The minimum average cost is $\bar C_{1\mathrm{min}}={8\over81}$.
437:
438:
439:
440: \noindent
441: \textbf{Case (b)}:
442:
443: The output state
444: $\hat\Psi_2(\theta)=\hat\psi_2(\theta)\oplus\hat\phi_2(\theta)$
445: is given by
446: \begin{eqnarray}
447: \hat\psi_2(\theta)
448: &=&
449: \left[
450: \begin{array}{cc}
451: {1\over4}-({1\over2}-x)\theta+\theta^2 & \theta^2\sqrt{x(1-x)} \\
452: \theta^2\sqrt{x(1-x)} & {1\over4}+({1\over2}-x)\theta+\theta^2
453: \end{array}
454: \right],
455: \\
456: \hat\phi_2(\theta)
457: &=& \frac{1-\theta^2}{4}\hat I_\phi.
458: \end{eqnarray}
459: The elements of the risk operator are
460: \begin{eqnarray}
461: \hat W^{(0)}&=&{1\over{27}}
462: \left[
463: \begin{array}{cc}
464: 4+9x & 7\sqrt{x(1-x)} \\
465: 7\sqrt{x(1-x)} & 13-9x
466: \end{array}
467: \right]
468: \nonumber\\
469: &\oplus& {5\over{27}}\hat I_\phi,
470: \label{case_b_W0}
471: \\
472: \hat W^{(1)}&=&{1\over{27}}
473: \left[
474: \begin{array}{cc}
475: 7x & 5\sqrt{x(1-x)} \\
476: 5\sqrt{x(1-x)} & 7(1-x)
477: \end{array}
478: \right]
479: \nonumber\\
480: &\oplus& {1\over{27}}\hat I_\phi,
481: \\
482: \hat W^{(2)}
483: &=&{1\over{405}}
484: \left[
485: \begin{array}{cc}
486: 4+75x & 61\sqrt{x(1-x)} \\
487: 61\sqrt{x(1-x)} & 79-75x
488: \end{array}
489: \right]
490: \nonumber\\
491: &\oplus&\frac{11}{405}\hat I_\phi.
492: \end{eqnarray}
493: The minimizing operator is (see Appendix \ref{ap_b})
494: \be\label{Theta_b}
495: \hat\Theta
496: =\frac{1}{17[13+8x(1-x)]}
497: \left[
498: \begin{array}{cc}
499: a & c \\
500: c & b
501: \end{array}
502: \right]
503: \oplus {1\over5}\hat I_\phi,
504: \ee
505: where
506: \be
507: \begin{array}{lll}
508: a&=&7x(35-20x+2x^2), \\
509: b&=&7x(17+16x+2x^2), \\
510: c&=&9[9-2x(1-x)]\sqrt{x(1-x)}.
511: \end{array}
512: \ee
513: To diagonalize it we use
514: $r=\sqrt{(a-b)^2+4c^2}$ and
515: \be
516: \mathrm{cos}\gamma=\sqrt{\frac{r+a-b}{2r}},
517: \quad
518: \mathrm{sin}\gamma=\sqrt{\frac{r-a+b}{2r}}.
519: \ee
520: We then have the similar eigenstates to Eq. (\ref{eigenvec1}) and
521: the eigenvalues $\theta_1=\theta_+$, $\theta_2=\theta_-$, and
522: $\theta_3=\theta_4={1\over5}$ with
523: \be
524: \theta_\pm=\frac{119[1+2x(1-x)] \pm r}{34[13+8x(1-x)]}.
525: \ee
526: The average cost is then
527: \be
528: \bar C_2(x)=\frac{8[391+606x(1-x)-10x^2(1-x)^2]}{2295[13+8x(1-x)]}.
529: \ee
530: This is an upward convex function, symmetric with respect to
531: $x={1\over2}$. The minimum is attained at $x=0,1$, that is, by
532: separable input states. This reads
533: $\bar C_{2\mathrm{min}}=\frac{184}{1755}$.
534: %%%%%%%%%%%%%%%%%%%%
535: \begin{figure}
536: \begin{center}
537: \includegraphics[width=0.41\textwidth]{Cx.eps}
538: \end{center}
539: \caption{\label{fig:Cx}
540: The average costs as a function of $x$.
541: }
542: \end{figure}
543: %%%%%%%%%%%%%%%%%%%%
544: %%%%%%%%%%%%%%%%%%%%
545: \begin{figure}
546: \begin{center}
547: \includegraphics[width=0.43\textwidth]{scheme.eps}
548: \end{center}
549: \caption{\label{fig:scheme}
550: The optimal estimation strategy using two probe qubits.
551: $\ket\Psi$ is the maximally entangled state. The output state is
552: projected onto $\{\hat\Pi_1, \hat\Pi_2\}$. We guess the channel
553: parameter as $\theta={5\over9}$ for the outcome $\hat\Pi_1$ and
554: $\theta={1\over9}$ otherwise.
555: }
556: \end{figure}
557: %%%%%%%%%%%%%%%%%%%%
558:
559: The average costs for cases (a) and (b) are shown in
560: Fig.~\ref{fig:Cx}:
561: $\bar C_1(x)$ (solid line) and $\bar C_2(x)$ (dashed line).
562: We see that $\bar C_{1\mathrm{min}}<\bar C_{2\mathrm{min}}$ so that
563: the optimal estimation strategy, using two probe qubits, is to
564: prepare them as a maximally entangled pair and to input one qubit
565: of the pair into the channel keeping the other untouched.
566: The estimation is then obtained by applying the two element POM,
567: Eq. (\ref{optimal POM}), as described in Case (a).
568: This strategy is represented schematically in
569: Fig. \ref{fig:scheme}.
570:
571:
572:
573: When $M$ maximally entangled pairs $\ket\Psi^{\otimes M}$ are
574: available, it is best to use them so as to have the output
575: $[({\cal L}_\theta\otimes\hat I)\proj{\Psi}]^{\otimes M}$.
576: The optimal measurement for this can be derived straightforwardly.
577: This is discussed in the next section as a part of an arbitrary
578: finite dimensional case.
579:
580:
581:
582: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
583: \section{$d$-dimensional case}
584: \label{sec:d-dim}
585: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
586:
587:
588: The action of the depolarizing channel on a $d$ dimensional system
589: is described by
590: \be
591: {\cal L}_\theta \hat\rho = \theta \hat\rho
592: + \frac{1-\theta}{d}\hat I.
593: \ee
594: Complete positivity then implies $-{1\over{d^2-1}}\le\theta\le1$.
595: For $d\ge3$, we have not succeeded in finding the optimal probe state,
596: even when we restrict ourselves to a pure state.
597: In this section we focus on the most plausible input state,
598: that is, the maximally entangled state, and consider the estimation
599: using $M$ entangled pairs.
600: Only for $d=2$, is the optimality ensured.
601:
602:
603: It might be interesting to compare the three cases specified by the
604: three different outputs;
605: \begin{itemize}
606: \item[(a)] \enskip
607: $M$ product states of the pair
608: \begin{eqnarray}
609: \hat\Psi_1(\theta)
610: &=&({\cal L}_\theta\otimes\hat I) \proj{\Psi},
611: \nonumber\\
612: &=&\theta\proj{\Psi}+\frac{1-\theta}{d^2}\hat I\otimes\hat I,
613: \label{output-a-1}
614: \end{eqnarray}
615: where $\ket{\Psi}$ is the maximally entangled state,
616: \item[(b)] \enskip
617: $M$ product states of the pair
618: \begin{eqnarray}
619: \hat\Psi_2(\theta)
620: &=&({\cal L}_\theta\otimes{\cal L}_\theta) \proj{\Psi},
621: \nonumber\\
622: &=&\theta^2\proj{\Psi}
623: +\frac{1-\theta^2}{d^2}\hat I\otimes\hat I,
624: \end{eqnarray}
625: \item[(c)] \enskip
626: $2M$ product states of
627: \begin{eqnarray}
628: \hat\psi(\theta)
629: &=&{\cal L}_\theta\proj{0},
630: \nonumber\\
631: &=&\theta\proj{0}+\frac{1-\theta}{d}\hat I.
632: \end{eqnarray}
633: \end{itemize}
634: (The input state in case (c) can be any pure state in the $d$
635: dimensional space.)
636: Let us first consider the case (a).
637: We denote Eq. (\ref{output-a-1}) as
638: \be
639: \hat\Psi(\theta)=f_0(\theta) \hat a_0 + f_1(\theta) \hat a_1,
640: \label{output-a-2}
641: \ee
642: where
643: \be
644: \hat a_0\equiv\proj{\Psi}, \quad
645: \hat a_1\equiv\hat I-\proj{\Psi},
646: \ee
647: and
648: \be
649: f_0(\theta)=\theta+\frac{1-\theta}{d^2}, \quad
650: f_1(\theta)=\frac{1-\theta}{d^2}.
651: \ee
652: The output state can then be represented as
653: \be
654: \hat\Psi(\theta)^{\otimes M}=\sum_{m=0}^M
655: f_0(\theta)^{M-m} f_1(\theta)^{m} \hat A_m,
656: \ee
657: where
658: \be
659: \hat A_m=\sum_{(i_1+...i_M=m)}
660: \hat a_{i_1}\otimes\cdots\otimes\hat a_{i_M},
661: \label{projector-A}
662: \ee
663: is the projector onto the symmetric subspace.
664: The risk operator is
665: \be
666: \hat W(\theta)=\sum_{m=0}^M
667: [\omega_m^{(2)}-2\theta\omega_m^{(1)}+\theta^2\omega_m^{(0)}]
668: \hat A_m,
669: \ee
670: where
671: $\omega_m^{(k)}\equiv
672: \int_{-{1\over3}}^1d\theta
673: \theta^k f_0(\theta)^{M-m} f_1(\theta)^{m}$.
674: The optimal POM is
675: \be
676: \hat\Pi(\theta)=\sum_{m=0}^M \delta(\theta-\theta_m) \hat A_m,
677: \label{POM_M}
678: \ee
679: where
680: \be
681: \theta_m\equiv\frac{\omega_m^{(1)}}{\omega_m^{(0)}},
682: \label{estimate}
683: \ee
684: We then note that
685: \be
686: \hat W(\theta)-\hat\Gamma
687: =\sum_{m=0}^M (\theta-\theta_m)^2 \omega_m^{(0)} \hat A_m\ge0,
688: \ee
689: from which it can easily be seen that the conditions (i) and (ii)
690: hold.
691: The minimum average cost is
692: \be
693: \bar C_1(M)=\sum_{m=0}^M
694: \left[
695: \omega_m^{(2)}-\frac{(\omega_m^{(1)})^2}{\omega_m^{(0)}}
696: \right]
697: \left(\begin{array}{c}
698: M\\m
699: \end{array}
700: \right)
701: (d^2-1)^m.
702: \label{C1_M}
703: \ee
704:
705:
706: The other cases can be dealt with in a similar manner.
707: In the case (b), we just put
708: \be
709: f_0(\theta)=\theta^2+\frac{1-\theta^2}{d^2}, \quad
710: f_1(\theta)=\frac{1-\theta^2}{d^2}.
711: \label{f0f1_C2}
712: \ee
713: The minimum average cost $\bar C_2(M)$ is then given by the same
714: expression as Eq. (\ref{C1_M}) with $\omega_m^{(k)}$'s defined by
715: $f_0(\theta)$ and $f_1(\theta)$ of Eq. (\ref{f0f1_C2}).
716:
717:
718: In the case (c), we use
719: \be
720: \hat a_0\equiv\proj{0}, \quad
721: \hat a_1\equiv\hat I -\proj{0},
722: \ee
723: and
724: \be
725: f_0(\theta)=\theta+\frac{1-\theta}{d}, \quad
726: f_1(\theta)=\frac{1-\theta}{d}.
727: \label{f0f1_Cs}
728: \ee
729: The minimum average cost is
730: \be
731: \bar C_\mathrm{SEP}(M)=\sum_{m=0}^{2M}
732: \left[
733: \omega_m^{(2)}-\frac{(\omega_m^{(1)})^2}{\omega_m^{(0)}}
734: \right]
735: \left(\begin{array}{c}
736: 2M\\m
737: \end{array}
738: \right)
739: (d-1)^m.
740: \label{Csepmin_M}
741: \ee
742:
743:
744: The three costs $\bar C_1(M)$, $\bar C_2(M)$, and
745: $\bar C_\mathrm{SEP}(M)$
746: are plotted in Fig. \ref{fig:CMd2} $(d=2)$,
747: Fig. \ref{fig:CMd3} $(d=3)$, and Fig. \ref{fig:CMd10} $(d=10)$.
748: In the figures another average cost $\bar C_\mathrm{ML}(M)$
749: is also plotted.
750: This cost is by the strategy belonging to the case (c), but unlike
751: the one attaining $\bar C_\mathrm{SEP}(M)$, the estimator is made by
752: the maximum likelyhood principle for which
753: \be
754: \theta_m=\frac{md}{2M(d-1)},
755: \ee
756: instead of Eq. (\ref{estimate}), and leads to the analytic expression
757: \be
758: \bar C_\mathrm{ML}(M)=\frac{1}{2M}\frac{d^5(d+3)}{6(d^2-1)^3}
759: \ee
760: It is this strategy that was used in ref. \cite{Cirone01} for the
761: case of $d=2$.
762:
763:
764: For $d\ge3$, the minimum average cost is always attained by a
765: separable probe state.
766: Only in the two dimensional case, is it the bipartite entangled probe
767: that attains the minimum average cost.
768: It is worth mentioning the depolarizing channel with the narrower
769: parameter region $0\le\theta\le1$, which is a more commonly
770: used model with an well defined interpretation of randomized
771: \textit{probability} of $\theta$.
772: We found that the best probe in this model is always a separable
773: state.
774: In this sense a separable state is generally an adequate probe state
775: for the depolarizing channel estimation as far as the comparison with
776: a bipartite entangled probe state is concerned.
777:
778:
779:
780: %%%%%%%%%%%%%%%%%%%%
781: \begin{figure}
782: \begin{center}
783: \includegraphics[width=0.41\textwidth]{CMd2.eps}
784: \end{center}
785: \caption{\label{fig:CMd2}
786: The average costs as a function of the number of pairs.
787: }
788: \end{figure}
789: \begin{figure}
790: \begin{center}
791: \includegraphics[width=0.41\textwidth]{CMd3.eps}
792: \end{center}
793: \caption{\label{fig:CMd3}
794: The average costs as a function of the number of pairs.
795: }
796: \end{figure}
797: \begin{figure}
798: \begin{center}
799: \includegraphics[width=0.41\textwidth]{CMd10.eps}
800: \end{center}
801: \caption{\label{fig:CMd10}
802: The average costs as a function of the number of pairs.
803: }
804: \end{figure}
805: %%%%%%%%%%%%%%%%%%%%
806:
807:
808:
809: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
810: \section{Concluding remark}
811: \label{sec:remark}
812: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
813:
814:
815: When we have several identical samples at our disposal,
816: it might be desirable to apply the best
817: \textit{collective} measurement on the whole system.
818: This means preparing a single multi-qubit state followed by an
819: optimized measurement.
820: We might also consider performing
821: a preliminary measurement on a part of the system and then
822: feedback this back to deal with the remaining part.
823: But in the case of the previous section, the collective measurement on
824: $M$ identical output pairs or $2M$ identical output particles is not
825: necessary.
826: The action of the depolarizing channel on a maximally entangled
827: state always results in a statistical mixture between
828: the input state and its orthogonal complement
829: (Eq. (\ref{output-a-2})).
830: Estimating the channel parameter is nothing but determining this
831: mixing ratio, which is a \textit{classical} distribution.
832: Therefore the optimal measurement is realized by a separable type
833: constructed
834: by the binary orthogonal projectors $\{\hat a_0, \hat a_1 \}$
835: according to Eq. (\ref{projector-A}).
836: In the case where the output state includes the channel parameter as
837: a quantum distribution, that is, the parameter appears in the off
838: diagonal components in the density matrix, the optimal measurement
839: would be a collective measurement.
840: When the channel includes a unitary opreration, we will have to face
841: this problem. Channel estimation for such a case is a future problem.
842:
843:
844: It is a remaining problem to see how effective the multipartite
845: entangled probe is.
846: However, in the estimation of decoherence channel under the power
847: constraint scenario,
848: that is, under a given and fixed number of probe particles,
849: it seems more common that entanglement is not necessary.
850: In fact, in the cases of the amplitude damping channel and
851: dephasing channel, there is no merit to use entangled probe.
852: In the amplitude damping channel, for example,
853: the best probe is to input the most highly excited state.
854: An entangled probe is rather wasteful because this includes the state
855: components other than the excited state and these components are less
856: sensitive to the damping.
857:
858:
859:
860: Finally it might be interesting to study the multi parameter case,
861: such as the Pauli channel estimation.
862: We may then ask how to optimaize (in Bayesian sense) the simultaneous
863: measurement on the noncommuting observables as well as searching for
864: appropriate probe states.
865:
866:
867:
868:
869: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
870: %
871: \begin{acknowledgments}
872: We are grateful to Mr. K. Usami, Dr. Y. Tsuda, and Dr. K. Matsumoto
873: for helpful discussions.
874: This work was supported, in part, by the British Council,
875: the Royal Society of Edinburgh, and by the Scottish
876: Executive Education and Lifelong Learning Department.
877: \end{acknowledgments}
878: %
879: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
880: \appendix*
881: %
882: %
883: \section{Derivation of Eq. (\ref{Theta_a})}
884: \label{app_a}
885:
886:
887: For obtaining the minimizing operator $\Theta$ in Eq. (\ref{Theta_a}),
888: we first diagonalize $\hat W^{(0)}$ by
889: $\hat U_0=\hat u_0\oplus\hat I_\phi$ where
890: \be
891: \hat u_0
892: = \left[
893: \begin{array}{cc}
894: \mathrm{cos}\gamma_0 & -\mathrm{sin}\gamma_0 \\
895: \mathrm{sin}\gamma_0 & \mathrm{cos}\gamma_0
896: \end{array}
897: \right],
898: \ee
899: with $r_0=\sqrt{1-3x(1-x)}$ and
900: \be
901: \mathrm{cos}\gamma_0=\sqrt{\frac{r_0-1+2x}{2r_0}},
902: \quad
903: \mathrm{sin}\gamma_0=\sqrt{\frac{r_0+1-2x}{2r_0}}.
904: \ee
905: The spectral decomposition
906: \be
907: \hat W^{(0)}=\sum_{i=1}^4 \omega_i \proj{\omega_i}.
908: \ee
909: is given by
910: \be\label{eigenvec0}
911: \begin{array}{lll}
912: \ket{\omega_1}&=\hat u_0\ket{\mu_1}, \quad
913: &\omega_1=(1+r_0)/3, \\
914: \ket{\omega_2}&=\hat u_0\ket{\mu_2}, \quad
915: &\omega_2=(1-r_0)/3, \\
916: \ket{\omega_3}&=\ket{\nu_1}, \quad
917: &\omega_3=(1-x)/3, \\
918: \ket{\omega_4}&=\ket{\nu_2}, \quad
919: &\omega_4=x/3.
920: \end{array}
921: \ee
922: We then calculate
923: \be
924: \tilde\Theta=\sum_{i,j=1}^4 \frac{2}{\omega_i+\omega_j}
925: \ket{i}\bra{\omega_i} \hat W^{(1)} \ket{\omega_j}\bra{j},
926: \label{minmizing_op}
927: \ee
928: where
929: \be
930: \begin{array}{lll}
931: \ket{1}&=\ket{\mu_1}, \\
932: \ket{2}&=\ket{\mu_2}, \\
933: \ket{3}&=\ket{\nu_1}, \\
934: \ket{4}&=\ket{\nu_2}.
935: \end{array}
936: \ee
937: This gives
938: \be
939: \tilde\Theta=
940: \tilde\Theta_\psi \oplus {1\over9}\hat I_\phi,
941: \ee
942: where
943: \be
944: \tilde\Theta_\psi
945: = {1\over9}
946: \left[
947: \begin{array}{cc}
948: \frac{4r_0(1+r_0)+3x(1-x)}{r_0(1+r_0)} &
949: -\frac{3(1-2x)\sqrt{x(1-x)}}{r_0} \\
950: -\frac{3(1-2x)\sqrt{x(1-x)}}{r_0} &
951: \frac{4r_0(1-r_0)-3x(1-x)}{r_0(1-r_0)}
952: \end{array}
953: \right].
954: \ee
955: The minimizing operator is given by
956: $\hat\Theta=\hat U_0\tilde\Theta\hat U_0^\dagger$
957: which results in Eq. (\ref{Theta_a}).
958:
959:
960: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
961: \section{Derivation of Eq. (\ref{Theta_b})}
962: \label{app_b}
963:
964:
965: The unitary operator for diagonalizing $\hat W^{(0)}$ in
966: Eq. (\ref{case_b_W0}) is
967: $\hat U_0=\hat u_0\oplus\hat I_\phi$ where
968: \be
969: \hat u_0
970: = \left[
971: \begin{array}{cc}
972: \mathrm{cos}\gamma_0 & -\mathrm{sin}\gamma_0 \\
973: \mathrm{sin}\gamma_0 & \mathrm{cos}\gamma_0
974: \end{array}
975: \right],
976: \ee
977: with $r_0=\sqrt{81-128x(1-x)}$ and
978: \be
979: \mathrm{cos}\gamma_0=\sqrt{\frac{r_0-9(1-2x)}{2r_0}},
980: \quad
981: \mathrm{sin}\gamma_0=\sqrt{\frac{r_0+9(1-2x)}{2r_0}}.
982: \ee
983: The spectral decomposition
984: \be
985: \hat W^{(0)}=\sum_{i=1}^4 \omega_i \proj{\omega_i}.
986: \ee
987: is given by
988: \be\label{eigenvec0}
989: \begin{array}{lll}
990: \ket{\omega_1}&=\hat u_0\ket{\mu_1}, \quad
991: &\omega_1=(17+r_0)/54, \\
992: \ket{\omega_2}&=\hat u_0\ket{\mu_2}, \quad
993: &\omega_2=(17-r_0)/54, \\
994: \ket{\omega_3}&=\ket{\nu_1}, \quad
995: &\omega_3=5/27, \\
996: \ket{\omega_4}&=\ket{\nu_2}, \quad
997: &\omega_4=5/27.
998: \end{array}
999: \ee
1000: We then have
1001: \be
1002: \tilde\Theta=\tilde\Theta_\psi \oplus {1\over5}\hat I_\phi,
1003: \ee
1004: where
1005: \be
1006: \tilde\Theta_\psi
1007: =
1008: \left[
1009: \begin{array}{cc}
1010: \frac{7[r_0+9-16x(1-x)]}{r_0(17+r_0)} &
1011: \frac{8(1-2x)\sqrt{x(1-x)}}{17r_0} \\
1012: \frac{8(1-2x)\sqrt{x(1-x)}}{17r_0} &
1013: \frac{7[r_0-9+16x(1-x)]}{r_0(17-r_0)}
1014: \end{array}
1015: \right].
1016: \ee
1017: Substituting this to
1018: $\hat\Theta=\hat U_0\tilde\Theta\hat U_0^\dagger$, we have
1019: Eq. (\ref{Theta_b}).
1020:
1021:
1022: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1023: %
1024: % references
1025: %
1026: \begin{thebibliography}
1027: \raggedright
1028: %
1029: \bibitem{Acin01}
1030: A.~Ac{\'i}n, E.~Jan{\'e}, and G.~Vidal,
1031: Phys.\ Rev.\ A\,\textbf{64}, 050302(R) (2001).
1032: \bibitem{Fujiwara01}
1033: A. Fujiwara,
1034: Phys.\ Rev.\ A\,\textbf{63}, 042304 (2001).
1035: \bibitem{Fischer01}
1036: D. G.~Fischer, H.~Mack, M. A.~Cirone, and M. Freyberger,
1037: Phys.\ Rev.\ A\,\textbf{64}, 022309 (2001).
1038: \bibitem{Cirone01}
1039: M. A.~Cirone, A.~Delgado, D. G.~Fischer, M. Freyberger, H.~Mack,
1040: and M. Mussinger, quant-ph/0108037.
1041: \bibitem{Barndorff98}
1042: O. E. Barndorff-Nielsen and R. D. Gill, quant-ph/9808009.
1043: \bibitem{Gill00}
1044: R. D. Gill and S. Massar,
1045: Phys.\ Rev.\ A\,\textbf{61}, 042312 (2000).
1046: \bibitem{Hayashi02}
1047: M. Hayashi, quant-ph/0202003.
1048: \bibitem{Helstrom_QDET}
1049: C. W.~Helstrom, \textit{Quantum Detection and Estimation Theory}
1050: (Academic Press, New York, 1976).
1051: \bibitem{Holevo_book}
1052: A. S.~Holevo : \textit{Probabilistic and Statistical Aspects of
1053: Quantum Theory} (North-Holland, Amsterdam, 1982).
1054: \bibitem{Peres_book}
1055: A.~Peres: \textit{Quantum Theory: concepts and methods}, 279 %--289
1056: (Kluwer Academic Publishers, Dortrecht, 1993).
1057: \bibitem{Holevo73_condition}
1058: A. S.~Holevo, J.~Multivar.\ Anal.\ \textbf{3}, 337 (1973).
1059: \bibitem{YuenKennedyLax75}
1060: H. P.~Yuen, R.S.~Kennedy, and M.~Lax,
1061: IEEE Trans.\ Inf.\ Theory \textbf{IT-21}(2), 125 (1975). %--134
1062: \bibitem{Personick71b}
1063: S. D. Personick,
1064: IEEE Trans.\ Inf.\ Theory \textbf{IT-17}(5), 240 (1971). %--246
1065: % \texttt{ArXiv:quant-ph/0008028}
1066: %
1067: \end{thebibliography}
1068: %
1069: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1070: %
1071: \end{document}
1072: %
1073: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1074:
1075:
1076:
1077: