1: % a sample file for Journal of Quantum Information and Computation (QIC) in
2: % LaTex2e by inputing macro file "qic.sty" with command \usepackage{qic},
3: % all the macros have been defined in the style file, so it is no need to
4: % put many macros at the beginning of the text file
5:
6: \documentclass[twoside]{article}
7: \usepackage{qic,epsfig}
8: \usepackage{theorem}
9: \usepackage{enumerate}
10: \usepackage{amsmath}
11: \usepackage{amssymb}
12: %\usepackage{rsfs}
13:
14: \newcommand{\lv}{\left \vert}
15: \newcommand{\rv}{\right \vert}
16: \newcommand{\la}{\left \langle}
17: \newcommand{\ra}{\right \rangle}
18: \newcommand{\ket}[1]{\lv #1 \ra}
19: \newcommand{\bra}[1]{\la #1 \rv}
20:
21: \newcommand{\B}{\mathfrak{B}}
22: \newcommand{\Hi}{\mathcal{H}}
23: \newcommand{\Ki}{\mathcal{K}}
24: \newcommand{\Tr}{\mathrm{tr}}
25: \theorembodyfont{\rmfamily}
26: \newtheorem{Theorem}{\textit{Theorem} }
27: \newtheorem{Remark}{Remark}
28: \newtheorem{Definition}{\textit{Definition} }
29: \newtheorem{Corollary}{\textit{Corollary}}
30: \newtheorem{Lemma}{\textit{Lemma}}
31: \newtheorem{Proof}{\textit{Proof}}
32: \renewcommand{\theProof}{}
33: \textwidth=5.6truein
34: \textheight=8.0truein
35:
36: \renewcommand{\thefootnote}{\fnsymbol{footnote}} %use symbolic footnote
37:
38: %%%%%%% starting the text file
39:
40: \begin{document}
41: \setlength{\textheight}{8.0truein} %FOR 2ND PAGE ONWARDS
42:
43: \runninghead{$\epsilon$-convertibility of entangled states and
44: extension of Schmidt rank in infinite-dimensional systems}
45: {M. Owari, S. L. Braunstein, K. Nemoto, M. Murao}
46:
47: \normalsize\textlineskip
48: \thispagestyle{empty}
49: \setcounter{page}{1}
50:
51: %\copyrightheading{Vol.}{No.}{Year}{Page Nos.}
52: %\copyrightheading{0}{0}{2003}{000--000}
53:
54: \vspace*{0.88truein}
55:
56: \alphfootnote
57:
58: \fpage{1}
59:
60: \centerline{\bf {\Large $\epsilon$}-CONVERTIBILITY OF ENTANGLED STATES AND EXTENSION}
61: % convertibility of entangled states }
62: \vspace*{0.035truein} \centerline{\bf
63: OF SCHMIDT
64: %and extension of Schmidt
65: RANK IN INFINITE-DIMENSIONAL SYSTEMS} \vspace*{0.37truein}
66: \centerline{\footnotesize Masaki Owari} \vspace*{0.015truein}
67: \centerline{\footnotesize\it Collaborative Institute for Nano Quantum Information Electronics, The University
68: of Tokyo\footnote{The major part of this work was done when MO was in Department of Physics, Graduate School of Science, The University of Tokyo.}}
69: \baselineskip=10pt \centerline{\footnotesize\it Tokyo
70: 113-0033, Japan} \vspace*{10pt} \centerline{\footnotesize Samuel
71: L. Braunstein} \vspace*{0.015truein} \centerline{\footnotesize\it
72: Computer Science, University of York} \baselineskip=10pt
73: \centerline{\footnotesize\it York YO10 5DD, United Kingdom}
74: \vspace*{10pt} \centerline{\footnotesize Kae Nemoto}
75: \vspace*{0.015truein} \centerline{\footnotesize\it National
76: Institute of Informatics } \baselineskip=10pt
77: \centerline{\footnotesize\it Tokyo 101-8430, Japan}
78: \vspace*{10pt} \centerline{\footnotesize Mio Murao}
79: \vspace*{0.015truein} \centerline{\footnotesize\it Department of
80: Physics, The University of Tokyo } \baselineskip=10pt
81: \centerline{\footnotesize\it Tokyo 113-0033, Japan}
82: \centerline{\footnotesize\it PRESTO, JST } \baselineskip=10pt
83: \centerline{\footnotesize\it Kawaguchi, Saitama 332-0012, Japan}
84:
85: \vspace*{0.225truein}
86: %\publisher{(received date)}{(revised date)}
87:
88: \vspace*{0.21truein}
89:
90: %% \abstracts{first paragraph}{second paragraph}{third paragraph}
91: %% If there is only one paragraph, just keep the second and third empty
92: %% like the following one
93: \abstracts{By introducing the concept of
94: $\epsilon$-convertibility, we extend Nielsen's and Vidal's
95: theorems to the entanglement transformation of
96: infinite-dimensional systems. Using an infinite-dimensional
97: version of Vidal's theorem we derive a new stochastic-LOCC (SLOCC)
98: monotone which can be considered as an extension of the Schmidt
99: rank. We show that states with polynomially-damped Schmidt
100: coefficients belong to a higher rank of entanglement class in
101: terms of SLOCC convertibility. For the case of Hilbert spaces of
102: countable, but infinite dimensionality, we show that there are
103: actually an uncountable number of classes of pure
104: non-interconvertible bipartite entangled states. }{}{}
105:
106: \vspace*{10pt}
107:
108: \keywords{Entanglement, LOCC, Infinite dimension, Continuous
109: variable} \vspace*{3pt}
110: % \communicate{to be filled by the Editorial}
111:
112: \vspace*{1pt}\textlineskip %) USE THIS MEASUREMENT WHEN THERE IS
113: %) A SECTION HEADING
114: %\vspace*{-0.5pt}
115: %\noindent
116: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
117: %put the text of the paper here
118: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
119:
120: \section{Introduction}
121: %%not yet finished%%
122: %
123: Entanglement is one of the central topics in quantum information, and
124: has both physical and information scientific aspects. In
125: particular, entanglement involves quantum non-local correlations
126: which have been of interest in physics \cite{epr}, and also acts
127: as a resource of quantum information processing for informatics
128: \cite{resource, quantum communication}. Thus, the characterization
129: of the entanglement of physical systems is important from both a
130: physical and information scientific viewpoint.
131:
132: Mathematically, physical systems can be categorized into two
133: classes, that is, finite-dimensional systems which can be treated
134: in the framework of conventional linear algebra, and
135: infinite-dimensional systems which need to be treated in the
136: framework of functional analysis \cite{neumann, functional
137: analysis}. It is therefore worthwhile to know whether such a
138: mathematical difference of systems can make an essential
139: difference in the properties of entanglement in these systems.
140: Indeed, this may provide an answer to the question of what is the
141: essential difference between the physics of finite-dimensional
142: systems and the physics of infinite-dimensional systems from the
143: viewpoint of non-local correlations. Moreover, from the
144: information-theoretic viewpoint, if such a difference exists,
145: there may be the possibility that we can achieve an information
146: processing in infinite-dimensional systems which cannot be
147: achieved in finite-dimensional systems.
148:
149: In this paper, we mainly focus on seeking a difference between the
150: properties of entanglement of finite-dimensional systems and those of
151: infinite-dimensional systems. Since much work on the
152: characterization of bipartite entangled states has been done in
153: finite-dimensional systems \cite{bennett, majorization, vidal,
154: monotone}, we concentrate our efforts on the characterization of
155: bipartite entangled states in infinite-dimensional systems, and
156: try to find a difference in the properties of their entanglement.
157:
158: So far, research on the characterization of entanglement in
159: infinite dimensions has been done in the form of separability
160: criteria \cite{separability}, Gaussian LOCC convertibility
161: \cite{gaussian, gaussian distill}, and entanglement measures
162: \cite{entanglement measure}. The separability criteria gives us a
163: way of judging whether or not a given state is entangled. The
164: Gaussian LOCC convertibility gives the detailed structure of the
165: strength of entanglement of Gaussian states. Entanglement measures
166: give an approximated strength of entanglement in the limit of an
167: asymptotic infinite number of copies.
168:
169: The above research mainly is concerned with Gaussian states and
170: Gaussian operations, and unique properties of infinite-dimensional
171: entanglement do not appear clearly in this regime. Therefore, in
172: order to find a property unique to infinite-dimensional
173: entanglement it is important to investigate the strength of
174: entanglement more precisely for a broader class of states and
175: operations.
176:
177: The strength of entanglement is defined by means of the
178: convertibility between entangled states under local operations,
179: \textit{e.g.}, local operations and classical communication
180: (LOCC), stochastic-LOCC (SLOCC), or the positive partial transpose
181: (PPT) operation \cite{LOCC,uniqueness}. Among such local
182: operations we mainly focus in this paper on SLOCC, and investigate
183: the SLOCC convertibility of entangled states in
184: infinite-dimensional systems without any assumption for states or
185: operations to find a unique property of entanglement in
186: infinite-dimensional systems.
187:
188: When we consider SLOCC convertibility for infinite-dimensional
189: systems, there are at least two difficulties, namely, the problem
190: of continuity and the problem of a potentially infinite cost for
191: classical communication. In order to avoid such difficulties, we
192: propose a new definition of state convertibility that we call
193: $\epsilon$-convertibility. We define $\epsilon$-convertibility as
194: the convertibility of states in an approximated setting by means
195: of the trace norm. Then, within the framework of
196: $\epsilon$-convertibility, we investigate SLOCC convertibility in
197: infinite-dimensional systems and show a fundamental difference of
198: SLOCC convertibility between infinite and finite-dimensional
199: systems.
200:
201: The paper is organized as follows: in section \ref{infinite
202: Nielsen}, we define the $\epsilon$-convertibility of LOCC and
203: SLOCC, and show how to avoid the problems of discontinuity and
204: infinitely-costly classical communication. Then within the
205: framework of $\epsilon$-convertibility, we give the
206: infinite-dimensional extensions and proofs of Nielsen's and
207: Vidal's theorem, which give the necessary and sufficient
208: conditions of LOCC and SLOCC convertibility, respectively. In
209: section \ref{Extension}, first we define monotones (monotonic
210: functions) of SLOCC convertibility, which can be considered an
211: extension of the Schmidt rank for infinite-dimensional systems,
212: then by means of this monotone, we investigate the SLOCC
213: convertibility for infinite-dimensional systems. We show that the
214: cardinal number of the quotient set of states by SLOCC
215: convertibility is greater than or equal to the cardinal number of
216: the continuum, and also show that however many (finite) copies of
217: exponentially-damped states (states with exponentially damped
218: Schmidt coefficients) there are, they cannot be converted into
219: even a single copy of a polynomially-damped state (a state with
220: polynomially-damped Schmidt coefficients). Such properties do not
221: exist in finite-dimensional systems and are actually unique to
222: infinite-dimensional systems.
223:
224: %%%%%%%%%%%%%%%%%%% epsilon-convertibility %%%%%%%%%%%%%%%%%%%%
225:
226: \section{$\epsilon$-convertibility} \label{infinite Nielsen}
227:
228: In this paper we consider the bipartite infinite-dimensional
229: system $\Hi = \Hi _A \otimes \Hi _B$ where $\dim \Hi _A = \dim \Hi
230: _B = \infty$ and we shall assume that $\Hi _A$ and $\Hi _B$ are
231: separable. By $\B (\Hi)$ we denote the Banach space of all bounded
232: operators on $\Hi$. If we use the term LOCC, we will always assume
233: that operations succeed \textit{with unit probability}
234: \cite{uniqueness}. On the other hand we use the term SLOCC in the
235: case where operations work with a finite probability less than
236: unity. For simplicity we use at most countably infinite POVMs as
237: the element of an LOCC (or SLOCC), $\{A_i\}_{i=1}^{\infty}$,
238: $A_i\in\B(\Hi)$, $\sum_{i\in \mathbb{N}}A ^{\dagger}A =
239: (\textbf{or} \leq )I$ (corresponding to ultra-weak convergence).
240:
241: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
242:
243: \subsection{$\epsilon$-convertibility for LOCC and SLOCC}\label{epsilon}
244:
245: As mentioned in the introduction to give a detailed discussion of
246: SLOCC convertibility in infinite-dimensional systems, there are at
247: least two difficulties, namely, discontinuity and infinite
248: classical communication costs. In this subsection we define
249: $\epsilon$-convertibility and see that it allows us to avoid the
250: difficulty of discontinuity. The other difficulty is addressed in
251: the following subsection.
252:
253: In infinite-dimensional systems we cannot deny the possibility
254: that $\ket{\Psi}$ is SLOCC convertible to any neighborhood of
255: $\ket{\Phi}$ (in terms of strong, or weak topology), but not to
256: $\ket{\Phi}$ itself. To avoid such a discontinuity, when
257: considering convertibility among genuinely infinite-dimensional
258: states (i.e., states with infinitely many non-zero Schmidt
259: coefficients), we shall identify these neighborhoods with the
260: state itself. To achieve this we shall extend the definition of
261: LOCC and SLOCC convertibility to satisfy the above requirement.
262: Mathematically, we redefine LOCC convertibility as follows:
263: $\ket{\Psi}$ can be converted to $\ket{\Phi}$ by LOCC, if and only
264: if for any neighborhood of $\ket{\Phi}$, there exists an LOCC
265: operation by which $\ket{\Psi}$ is transformed to a state in the
266: neighborhood of $\ket{\Phi}$. We call this new definition of
267: convertibility $\epsilon$-convertibility. Below we rigorously
268: define $\epsilon$-convertibility for LOCC, then we show that this
269: definition recovers the continuity of convertible probability at
270: least with some suitable weak meaning.
271:
272: Before we give the definition of $\epsilon$-convertibility, we
273: need to choose a topology of the convergence which we use in our
274: definition. Actually, it is well known that there are many
275: different topologies defined by associated norms in
276: infinite-dimensional systems. Therefore, we need to take care to
277: choose our `distance'. Since we introduced
278: $\epsilon$-convertibility because of the fundamental impossibility
279: for discriminating a state $\ket{\Phi}$ from states within
280: infinitely small neighborhoods of $\ket{\Phi}$, the distance we
281: consider needs to echo this difficulty with discrimination. We can
282: easily see that the trace norm possesses such a property as
283: follows. Suppose $M$ is an arbitrary POVM element and $\lim
284: _{n\rightarrow\infty}\|\rho-\rho_n\|_{\rm tr}=0$. Then, $\lim _{n
285: \rightarrow \infty} | \Tr\; \rho M - \Tr\; \rho _n M| \le \lim _{n
286: \rightarrow \infty} \| \rho - \rho _n \| _{\rm tr}\|M\|_{\rm op} =
287: 0$, where $\| \cdot \| _{\rm op}$ is the operator norm. Thus, for
288: all measurements the resulting probability distributions for
289: $\rho_n$ converge to the resulting probability distribution for
290: $\rho$. That is, if $\rho_n$ converges $\rho$ in the trace norm,
291: there is no way to discriminate $\rho$ from $\rho _n$ for
292: sufficiently large $n$. But this is just the property required of
293: the distance needed to deal with the discontinuity difficulty in
294: the definition of $\epsilon$-convertibility. Therefore, we shall
295: use the trace norm as our distance measure for
296: $\epsilon$-convertibility.
297:
298: Following this discussion we rigorously define
299: $\epsilon$-convertibility for LOCC as:
300: \begin{Definition}
301: We say that $\ket{\Psi}$ is $\epsilon$-convertible to $\ket{\Phi}$
302: by LOCC, if for any $\epsilon > 0$, there exists an LOCC operation
303: $\Lambda$ which satisfies the condition
304: $\|\Lambda(\ket{\Psi}\bra{\Psi})-\ket{\Phi}\bra{\Phi}\|_{\rm
305: tr}<\epsilon$ where $\|\cdot\|_{\rm tr}$ is the trace norm.
306: \end{Definition}
307: Similarly, we define $\epsilon$-convertibility for SLOCC as:
308: \begin{Definition}
309: \label{epsilon SLOCC} We say that $\ket{\Psi}$ is
310: $\epsilon$-convertible to $\ket{\Phi}$ by SLOCC with probability
311: $p > 0$ if for any $\epsilon
312: >0$, there exists an SLOCC operation $\Lambda $ which satisfies the following
313: condition, $\| \Lambda ( \ket{\Psi} \bra{\Psi} )/\Tr\; \Lambda (
314: \ket{\Psi} \bra{\Psi} ) - \ket{\Phi} \bra{\Phi} \| _{\rm tr} <
315: \epsilon$ and $\Tr\; \Lambda ( \ket{\Psi} \bra{\Psi} ) \ge p$.
316: \end{Definition}
317: This definition of $\epsilon$-convertibility under SLOCC means
318: that {\itshape with more than some fixed non-zero probability
319: $p$}, $\ket{\Psi}$ can be converted to any neighborhood of
320: $\ket{\Phi}$ by SLOCC.
321:
322: Here, we prove by means of $\epsilon$-convertibility that we can
323: recover enough continuity to achieve a classification of states by
324: SLOCC convertibility. In infinite-dimensional systems when we
325: consider SLOCC convertibility it might happen that $\ket{\Psi}$
326: cannot be converted to $\ket{\Phi}$ and yet there exists a
327: sequence of $\ket{\Phi _n}$ such that $\ket{\Psi}$ can be
328: converted to $\ket{\Phi _n}$ with probability $p_n$ and $\lim _{n
329: \rightarrow \infty} p_n > 0$; however, we cannot discriminate
330: $\ket{\Phi }$ and $\ket{\Phi _n}$ for large $n$. If this were to
331: happen then it would be nonsense that $\ket{\Psi}$ could not
332: converted to $\ket{\Phi}$ by SLOCC.\fnm{{\it a}}\fnt{}{\ To avoid confusion, we add a remark.
333: We do not know an explicit example of this discontinuity.
334: However, in infinite dimensional systems,
335: it is not trivial whether such discontinuity occurs or not.
336: Hence, it is better to introduce a modified definition of convertibility under which
337: we can trivially avoid the discontinuity.}
338: However, by means of our new
339: definition of convertibility, we avoid such a discontinuity. That
340: is, we can easily show the following continuity property of
341: $\epsilon$-convertibility.
342: \begin{Lemma}\label{epsilon SLOCC lemma}
343: If $\ket{\Psi}$ is not $\epsilon$-convertible to $\ket{\Phi}$ by
344: SLOCC, but $\ket{\Psi}$ is $\epsilon$-convertible to $\ket{\Phi
345: _n}$ by SLOCC with probability $p_n$ for all $n$, where $\lim _{n
346: \rightarrow \infty} \ket{\Phi _n} \bra{\Phi _n} = \ket{\Phi}
347: \bra{\Phi}$ by the trace norm, then $\lim _{n \rightarrow \infty}
348: p_n = 0$.
349: \end{Lemma}
350: \begin{Proof}
351: We prove this lemma by contradiction. We assume the following
352: condition; $\ket{\Psi}$ is not $\epsilon$-convertible to $\ket{\Phi}$,
353: $\ket{\Psi}$ is $\epsilon$-convertible to $\ket{\Phi _n}$ with
354: probability $p_n > 0$, where $\lim _{n \rightarrow \infty}
355: \ket{\Phi _n} = \ket{\Phi }$. Moreover, if we add one condition;
356: $\limsup _{n \rightarrow \infty} p_n >0$, then we can show the
357: contradiction as follows.
358:
359: Since $\limsup _{n \rightarrow \infty} p_n >0$, there exists a
360: subsequence of $p_{n}$, such that $\lim p_{n(k)} = p >0$ and
361: $p_{n(k)} >p/2$ for all $k \in \mathbb{N}$. Then, since $\ket{\Psi
362: }$ is $\epsilon$-convertible to $\ket{\Phi _n}$ with probability
363: $p_n$, for any $\epsilon >0$ and for any $k \in \mathbb{N}$, there
364: exists an SLOCC operation $\Lambda _{\epsilon, n(k)}$ such that
365: $\| \Lambda_{\epsilon, n(k)}(\ket{\Psi}\bra{\Psi})/\Tr
366: \Lambda_{\epsilon, n(k)}(\ket{\Psi}\bra{\Psi}) - \ket{\Phi
367: _{n(k)}}\bra{\Phi _{n(k)}} \| < \epsilon$ and $\Tr
368: \Lambda_{\epsilon, n(k)} (\ket{\Psi}\bra{\Psi}) \ge p_{n(k)} >
369: p/2$. Moreover, since $\lim _{n \rightarrow \infty}\ket{\Phi _n}
370: =\ket{\Phi}$, for any $\epsilon >0$, there exists an $N_{\epsilon}
371: \in \mathbb{N}$ such that for any $n \ge N_{\epsilon}$, $\|
372: \ket{\Phi _n}\bra{\Phi _n} - \ket{\Phi}\bra{\Phi} \| < \epsilon$.
373: Therefore, for any $2 \epsilon >0$, by choosing $k \in \mathbb{N}$
374: as $n(k) \ge N_{\epsilon}$,
375: \begin{eqnarray*}
376: &\quad & \| \Lambda _{\epsilon, n(k)}/\Tr \Lambda_{\epsilon, n(k)}
377: -
378: \ket{\Phi}\bra{\Phi} \| \\
379: &\le & \| \Lambda_{\epsilon, n(k)}/\Tr \Lambda_{\epsilon, n(k)} -
380: \ket{\Phi _{n(k)}}\bra{\Phi _{n(k)}} \| + \| \ket{\Phi
381: _{n(k)}}\bra{\Phi _{n(k)}} - \ket{\Phi}\bra{\Phi}\| \\
382: &\le & 2 \epsilon.
383: \end{eqnarray*}
384: Moreover, $\Tr \Lambda_{\epsilon, n(k)}(\ket{\Psi}\bra{\Psi}) \ge
385: p_{n(k)} >p/2$. Therefore, $\ket{\Psi}$ is $\epsilon$-convertible
386: to $\ket{\Phi}$ with probability $p/2$. This is a contradiction.
387: Therefore, if $\ket{\Psi}$ is not $\epsilon$-convertible to
388: $\ket{\Phi} $, and if $\ket{\Psi}$ is $\epsilon$-convertible to
389: $\ket{\Phi _n}$ with probability $p_n$ where $\lim _{n \rightarrow
390: \infty} \ket{\Phi_n} = \ket{\Phi} $, then, $\lim _{n \rightarrow
391: \infty} p_n = 0$. \hfill $square$
392: \end{Proof}
393: This lemma means that if $\ket{\Psi}$ cannot be converted to
394: $\ket{\Phi}$ by SLOCC, then $\ket{\Psi}$ is also almost certainly
395: inconvertible to states near to $\ket{\Phi}$. Therefore, our
396: definition of $\epsilon$-convertibility preserves continuity of
397: the theory (at least sufficiently for the purposes of the
398: classification of states), and we can avoid the discontinuity
399: difficulty mentioned above.
400:
401: %%%%%%%%%%%%%%%%% Nielsen's and Vidal's theorem %%%%%%%%%%%%%%%%%%%%%%%%%%%%
402:
403: \subsection{Nielsen's and Vidal's theorems for infinite-dimensional systems}
404: \label{nielsen vidal}
405:
406: In this subsection we reconstruct Nielsen's and Vidal's theorems
407: for infinite-dimensional systems by means of
408: $\epsilon$-convertibility. As a result, we will see that we can
409: also avoid the difficulty of a potentially infinite cost for
410: classical communication by our convertibility, that is, only a
411: finite amount of classical communication is actually necessary for
412: our theory of convertibility. As is well known, Nielsen's and
413: Vidal's theorems give the necessary and sufficient conditions of
414: LOCC and SLOCC, respectively. Therefore, by proving these theorems
415: rigorously we may obtain a firm foundation for the analysis of
416: SLOCC convertibility for infinite-dimensional systems, which we
417: shall consider in the next section. Since Vidal's theorem is a
418: generalization of Nielsen's theorem, we shall first discuss
419: Nielsen's theorem and then go on to consider Vidal's theorem.
420:
421: In finite-dimensional systems Nielsen's theorem gives the
422: necessary and sufficient conditions for LOCC convertibility
423: between a pair of bipartite pure states $\ket{\Phi}$ and
424: $\ket{\Psi}$ as follows
425: \begin{equation}\label{nielsenfinite}
426: \ket{\Psi} \rightarrow \ket{\Phi}
427: ~~ \Leftrightarrow ~~
428: \mathbf{\lambda} \prec \mathbf{\mu}\;,
429: \end{equation}
430: where the arrow $\rightarrow$ represents convertibility under
431: LOCC, and $\mathbf{\lambda}$ and $\mathbf{\mu}$ represent
432: sequences of Schmidt coefficients (in descending order) of the
433: states $\ket{\Psi}$ and $\ket{\Phi}$, respectively, and $\prec$
434: denotes majorization of the sequences \cite{majorization} (if
435: $\mathbf{\lambda} \prec \mathbf{\mu}$, we say ``$\mathbf{\lambda}$
436: is majorized by $\mathbf{\mu}$''). In infinite-dimensional
437: systems we can show that Eq.~(\ref{nielsenfinite}) is still valid
438: where we replace the meaning of $\rightarrow$ by
439: $\epsilon$-convertibility under LOCC. Nielsen's theorem then takes
440: the following form for infinite-dimensional systems:
441: \begin{Theorem} \label{epsilon Nielsen}
442: $\ket{\Psi}$ is $\epsilon$-convertible to $\ket{\Phi}$, if and
443: only if $\lambda \prec \mu $, where $\prec$ means majorization in
444: infinite-dimensional systems (see Appendix \ref{Majorization}),
445: and $\mathbf{\lambda}$ and $\mathbf{\mu}$ are the Schmidt
446: coefficients of $\ket{\Psi}$ and $\ket{\Phi}$, respectively.
447: \end{Theorem}
448: Since the proof of theorem \ref{epsilon Nielsen} is long, we have
449: placed the rigorous proof of this theorem in Appendix \ref{Proof
450: of Nielsen}. Below we only give a sketch of the proof:
451:
452: \textbf{Sketch of Proof}
453:
454: 1) The necessary part: We can directly extend the proof of
455: necessity of the original theorem to infinite-dimensional systems.
456: The necessary condition part of the original theorem is
457: constructed using the Lo-Popescu theorem (Theorem
458: \ref{Lo-Popescu}) \cite{lo-popescu} and Uhlmann's theorem (Theorem
459: \ref{Uhlmann}). Since these two theorems can themselves be
460: extended to infinite-dimensional systems (see Appendix A and B).
461: The same proof for finite-dimensional systems still holds in
462: infinite-dimensional systems.
463:
464: 2) The sufficient part: In the proof of sufficiency, our
465: definition of $\epsilon$-convertibility plays a crucial role in
466: extending the proof of Nielsen's theorem. Our proof is based on
467: the proof for finite-dimensional systems in Ref.~\cite{uniqueness}
468: and is extended to genuine infinite-dimensional states by means of
469: $\epsilon$-convertibility. We can show that for any $N$, there
470: exists a state $ \ket{\Phi'}$ (which depends on $N$)
471: such that its first $N$ Schmidt coefficients are equal to the
472: Schmidt coefficients of $\ket{\Phi}$ and where the Schmidt
473: coefficients of $\ket{\Psi }$ are majorized by the Schmidt
474: coefficients of $\ket{\Phi'}$. Therefore, for every neighborhood of
475: $\ket{\Phi}$, we can always find a state to which $\ket{\Psi}$ can
476: be converted under LOCC. \hfill $\square$
477:
478: By means of Nielsen's theorem in infinite-dimensional systems we
479: can extend Vidal's theorem for SLOCC convertibility \cite{vidal},
480: which gives the necessary and sufficient condition of SLOCC
481: convertibility with a probability $p$ to infinite-dimensional
482: systems using $\epsilon$-convertibility. Vidal's theorem states
483: that a bipartite pure state $\ket{\Psi}$ can be converted to
484: another bipartite pure state $\ket{\Phi}$ under SLOCC with
485: probability at least $p$ if and only if $\lambda\prec ^{\omega} p
486: \mu$ [here $\prec ^{\omega}$ denotes super-majorization and is defined
487: in Appendix A1, Eq. (\ref{super-majorization}) of Definition 4].
488: The generalization of Vidal's theorem for
489: $\epsilon$-convertibility can then be written:
490: \begin{Theorem} \label{epsilon Vidal}
491: $\ket{\Psi}$ is $\epsilon$-convertible to $\ket{\Phi}$ by SLOCC
492: with probability $p$, if and only if $\lambda \prec ^{\omega} p
493: \mu$ are satisfied where $\lambda$ and $\mu$ are the Schmidt
494: coefficients of $\ket{\Psi}$ and $\ket{\Phi}$ respectively.
495: \end{Theorem}
496: \begin{Proof}
497: The proof of this theorem is in appendix \ref{Proof of Vidal}.
498: \end{Proof}
499: Therefore, the extension of Vidal's theorem also applies to
500: $\epsilon$-convertibility.
501:
502: Although infinite amounts of classical information do not exist in
503: the real world, an infinite amount of classical communication is
504: necessary to convert one genuine infinite-dimensional state to
505: another by LOCC and SLOCC in the conventional theory of
506: convertibility. From the proof of Theorem \ref{epsilon Nielsen},
507: we can show that we can avoid such infinite costs of classical
508: communication in LOCC convertibility by our new definition of
509: $\epsilon$-convertibility. In the proof of this theorem, we showed
510: that there exists a natural number $M$ such that $\ket{\Phi'}$
511: satisfies the condition $\mu'_N = \lambda _N$ for $N \ge M$.
512: Therefore, the LOCC operation by which $\ket{\Psi }$ can be
513: converted into $\ket{\Phi'}$ is actually an LOCC operation
514: requiring only a finite amount of classical communication. We can
515: also show a similar result for SLOCC convertibility. By the proof
516: of Theorem \ref{epsilon Vidal} it is easily seen that we can
517: construct the protocol of SLOCC with only a finite amount of
518: classical communication in a manner similar to LOCC. As a result,
519: in our definition of $\epsilon$-convertibility of LOCC and SLOCC,
520: we can convert states with any finite accuracy by only a finite
521: amount of communication, and only when this error goes to zero
522: does the amount of classical communication go to infinity.
523: Therefore, our definition of $\epsilon$-convertibility yields a
524: theory of single-copy LOCC and SLOCC convertibility requiring only
525: a finite amount of classical communication even in the
526: infinite-dimensional setting.
527:
528: Here we need to add two final remarks about our framework of
529: $\epsilon$-convertibility. From the proofs of Theorems
530: \ref{epsilon Nielsen} and~\ref{epsilon Vidal}, we can derive
531: another interpretation of $\epsilon$-convertibility. First, in the
532: case of LOCC, that is, Nielsen's theorem, since the state
533: $\ket{\Phi '} = \sum _{k=1}^{\infty} \sqrt{\mu '} \ket{i} \otimes
534: \ket{i}$ is also majorized by $\ket{\Phi}$ in the proof of Theorem
535: \ref{epsilon Nielsen} (Appendix \ref{Proof of Nielsen}), we can
536: immediately see the following fact: If $\lambda \prec \mu $ where
537: $\lambda$ and $\mu$ are the Schmidt coefficients of $\ket{\Phi}$
538: and $\ket{\Psi}$ respectively, then there exists a sequence of
539: LOCC $\{ \Lambda _n \} _{n=1}^{\infty}$ such that for all $n \in
540: \mathbb{N}$, $\Lambda_n \cdots \Lambda _1 (\ket{\Psi} \bra{\Psi}
541: )$ has the same Schmidt basis and their Schmidt coefficients are
542: majorized by those of $\ket{\Phi}$ and they also satisfy $\lim _{n
543: \rightarrow \infty} \Lambda _n \cdots \Lambda _1 (\ket{\Psi}
544: \bra{\Psi} ) = \ket{\Phi} \bra{\Phi}$. Thus, we can interpret the
545: above sequence of LOCC as the LOCC with an infinite number of
546: steps of classical communication. Second, since in the proof of
547: the above theorem in appendix \ref{Proof of Vidal}, we constructed
548: a sequence of SLOCC $\{ \Lambda _n \}_{n=1}^{\infty}$ such that
549: $\lim _{n \rightarrow \infty} \Lambda _n \cdots \Lambda _1
550: (\ket{\Psi} \bra{\Psi} )/\Tr\; \Lambda _n \cdots \Lambda _1
551: (\ket{\Psi} \bra{\Psi} ) = \ket{\Phi}\bra{\Phi}$, we can consider
552: that Vidal's theorem is also naturally extended to
553: infinite-dimensional systems by the redefinition of LOCC
554: convertibility including an infinite number of steps of classical
555: communication. Therefore, we can also say that both the Nielsen
556: and Vidal theorems can be extended to infinite-dimensional systems
557: if we allow for an infinite number of steps of LOCC.
558:
559: In this section, we proposed a new definition of convertibility,
560: {\itshape $\epsilon$-convertibility}, to treat entanglement
561: convertibility between genuine infinite-dimensional states. This
562: redefinition is suitable from both the technical and realistic
563: viewpoints, that is, to avoid the difficulties of both
564: discontinuity and infinite cost in classical communication for
565: infinite-dimensional systems. Then, by means of
566: $\epsilon$-convertibility we proved the Nielsen and Vidal theorems
567: which are the fundamental theorems of LOCC and SLOCC
568: convertibility in infinite-dimensional systems. As a result under
569: our change of definition the framework of entanglement
570: convertibility is preserved in the context of infinite-dimensional
571: systems, and therefore our definition of $\epsilon$-convertibility
572: for LOCC is suitable and sufficient for realistic conditions of
573: quantum information processing in infinite-dimensional systems.
574:
575: %%%%%%%%%%%%%%%% Extension of Schmidt rank %%%%%%%%%%%%%%%%%%%%
576:
577: \section{Extension of Schmidt rank} \label{Extension}
578: \subsection{Definition and its basic property} \label{definition}
579:
580: In this section we discuss the SLOCC convertibility of
581: infinite-dimensional systems and show that there are many
582: important differences between the structure of the SLOCC
583: classification of genuinely infinite-dimensional states and that
584: of finite-dimensional states. For this purpose in this subsection
585: we first define a pair of new SLOCC monotones which can be
586: considered as extensions of the Schmidt rank, and then we analyze
587: their properties. Finally, we show that there are continuously
588: many classes of states under SLOCC convertibility in
589: infinite-dimensional systems. In the following we always consider
590: SLOCC convertibility in the meaning of $\epsilon$-convertibility
591: defined above. Therefore, we henceforth omit the qualifier
592: `$\epsilon$'.
593:
594:
595: To study convertibility in detail, monotones of convertibility are
596: crucially important. In finite systems, the Schmidt rank (the rank
597: of the reduced density matrix) gives the necessary and sufficient
598: condition of SLOCC convertibility. On the other hand, in
599: infinite-dimensional systems, since almost all states have
600: infinite Schmidt rank, the classification by Schmidt rank is not
601: useful. Therefore, proposals for new SLOCC monotones for genuine
602: infinite-dimensional entangled states are essential for analysis
603: of SLOCC convertibility between genuinely infinite-dimensional
604: states. In the followings, we give a definition of a pair of new
605: SLOCC monotones $R^-$ and $R^+$, which can be considered an
606: extension of the Schmidt rank for infinite-dimensional systems.
607:
608: Since the usual Schmidt rank represents how quickly Schmidt
609: coefficients vanish, when we consider their extension to genuine
610: infinite-dimensional states it is natural to define the extension
611: of this concept as a function which represents how quickly a
612: sequence of Schmidt coefficients converge to zero. In Vidal's
613: theorem Schmidt coefficients always appear in the form of a sum
614: from $n$ to $\infty$ which is an LOCC monotone for all $n \in
615: \mathbb{N}$ and is called ``{\it Vidal's monotone}''
616: \cite{monotone}. Therefore, rather than studying the direct
617: convergence of Schmidt coefficients $\{ \lambda _n
618: \}_{n=1}^{\infty}$ we shall study the convergence of Vidal's
619: monotones $\{ \sum _{i=n}^{\infty} \lambda _i \}_{n=1}^{\infty}$.
620: To measure the speed of convergence of Vidal's monotone we compare
621: a sequence of Vidal's monotones with some real parameterized class
622: of sequences. Thus, we define the new monotones as follows:
623: \begin{Definition} \label{def of monotone}
624: For a parameterized class of sequences $\{ f_r (n) \} _{r \in
625: (a,b)}$ which satisfy the following conditions:
626: \begin{list}{}{}
627: \item $\forall r \in (a,b)$, $\lim _{n \rightarrow \infty} f_r(n) =0$
628: \item $\forall r \in (a,b)$,
629: $n_1 < n_2 \Rightarrow f_r(n_1) > f_r(n_2) > 0$
630: \item $\forall n \in \mathbb{N}$,
631: $r_1 < r_2 \Rightarrow \lim _{n \rightarrow \infty}
632: \frac{f_{r_1} (n) }{f_{r_2} (n) } =0$\;,
633: \end{list}
634: where $n \in \mathbb{N}$, $0 \le a < b \le \infty$, we define a
635: pair of functions $R^+_{f_r} (\ket{\Psi} )$ and $R^-_{f_r}
636: (\ket{\Psi})$ by
637: \begin{eqnarray}
638: R^+ _{f_r}(\ket{\Psi}) &=& \inf \{ r \in (a,b) | \lim _{n
639: \rightarrow \infty}
640: \frac{\sum _{i=n}^{\infty} \lambda _i}{f_r(n)} = 0 \} \\
641: R^- _{f_r}(\ket{\Psi}) &=& \inf \{ r \in (a,b) | \underline{\lim}
642: _{n \rightarrow \infty}
643: \frac{\sum _{i=n}^{\infty} \lambda _i}{f_r(n)} = 0 \}\;.
644: \end{eqnarray}
645: If for all $r \in (a,b) $, $\overline{\lim} _{n \rightarrow
646: \infty} {\sum _{i=n}^{\infty} \lambda _i}/{f_r(n)} > 0$, then we
647: define $R^+_{f_r} (\ket{\Psi}) =b$. Here, we use the notation of
648: $\overline{\lim} = \limsup$ and $\underline{\lim} = \liminf$.
649: \end{Definition}
650: For the definition of $R^+_{f_r}(\ket{\Psi})$, we could also have
651: defined this function as
652: \begin{equation*}
653: R^+ _{f_r}(\ket{\Psi}) = \inf \{ r \in (a,b) | \overline{\lim} _{n
654: \rightarrow \infty}
655: \frac{\sum _{i=n}^{\infty} \lambda _i}{f_r(n)} = 0 \}\;,
656: \end{equation*}
657: however this definition is the same as the previous definition, since
658: $\overline{\lim} _{n \rightarrow \infty} {\sum _{i=n}^{\infty}
659: \lambda _i}/{f_r(n)} = 0$ guarantees
660: $\lim_{n \rightarrow \infty}
661: {\sum _{i=n}^{\infty} \lambda _i}/{f_r(n)} = 0$.
662: Note that the limits
663: $\overline{\lim} _{n \rightarrow \infty} \sum _{i=n}^{\infty}
664: \lambda _i / f_r(n)$ and $\underline{\lim} _{n \rightarrow \infty}
665: \sum _{i=n}^{\infty} \lambda _i /f_r(n)$ do not generally
666: coincide. Thus, to measure the speed of the convergence, we need
667: two functions $R^+_{f_r}(\ket{\Psi})$ and $R^-_{f_r}(\ket{\Psi})$
668: corresponding to these different approaches of the limit as given
669: above. By their definition, we can easily see that $R^+ _{f_r}$
670: and $R^- _{f_r}$ satisfy $R^- _{f_r} (\ket{\Psi}) \le R^+ _{f_r}
671: (\ket{\Psi})$ for all $\ket{\Psi}$. As we might expect, both $R^+
672: _{f_r}$ and $R^- _{f_r}$ are SLOCC monotones, and moreover, the
673: sufficient condition of monotonicity is also partially valid as
674: given by the statement of the following theorem. (Note, that when
675: the choice of $f_r(x)$ is clear we write simply $R^+$ and $R^-$
676: for the monotones.)
677: \begin{Theorem} \label{theorem of monotone}
678: For all $f_r$ which satisfy the condition in Definition \ref{def
679: of monotone},
680: \begin{enumerate}
681: \item If $\ket{\Psi}$ can be converted to $\ket{\Phi}$ by SLOCC then
682: $R^+_{f_r} (\ket{\Psi}) \ge R^+_{f_r} (\ket{\Phi})$ and $R^-_{f_r}
683: (\ket{\Psi}) \ge R^-_{f_r} (\ket{\Phi})$. \item If
684: $R^+_{f_r}(\ket{\Phi}) < R^-_{f_r}(\ket{\Psi})$, then $\ket{\Psi}$
685: can be converted to $\ket{\Phi}$ by SLOCC.
686: \end{enumerate}\end{Theorem}
687: \begin{Proof}
688: Proof of part 1: \\
689: We only prove this for the case of $R^+ _{f_r}$ since the proof
690: for $R^- _{f_r}$ is identical. Suppose $R^+ _{f_r} (\ket{\Phi}) >
691: R^+ _{f_r} (\ket{\Psi})$ then for all $R^+ _{f_r} (\ket{\Psi}) < r
692: < R^+ _{f_r} (\ket{\Phi})$, $\overline{\lim} _{n \rightarrow
693: \infty} \sum _{i=n}^{\infty} \lambda _i/f_r (n) = 0$ and \newline
694: $\overline{\lim} _{n \rightarrow \infty} \sum _{i=n}^{\infty} \mu
695: _i / f_r (n) > 0$, where $\{ \lambda _i \}_{i=0}^{\infty}$ and $\{
696: \mu _i \}_{i=0}^{\infty}$ are Schmidt coefficients of $\ket{\Psi
697: }$ and $\ket{\Phi }$, respectively. Thus, for all $\delta
698: > 0$ there exists an $N_0 (\delta)$ such that if $n > N _0 (\delta
699: )$, then $\sum _{i=n}^{\infty} \lambda _i / f_r(n) < \delta $.
700: Suppose $a \stackrel{def}{=} \overline{\lim} _{n \rightarrow
701: \infty} \sum _{i=n}^{\infty} \mu _i / f_r(n) >0$, then there
702: exists a partial sequence of $\sum _{i=n}^{\infty} \mu _i /
703: f_r(n)$, say $\sum _{i=k(n)}^{\infty} \mu _i / f_r(k(n))$, such
704: that $\lim _{n \rightarrow \infty} \sum _{i=k(n)}^{\infty} \mu _i
705: / f_r(k(n)) = a
706: >0$. Then there exists an $N_1 \in \mathbb{N}$ such that for all $n > N_1$,
707: $\sum _{i=k(n)}^{\infty} \mu _i / f_r(k(n)) > a/2$. Therefore if
708: we define $N_2(\delta)$ as $N_2(\delta) = \max (N_1, \min \{n \in
709: N |k(n) \ge N_0 \})$, then for all $n > N_2(\delta)$, $\sum
710: _{i=k(n)}^{\infty} \lambda _i / f_r(k(n)) < \delta$ and $
711: f_r(k(n)) / \sum _{i=k(n)}^{\infty} \mu _i < 2/a$. That is, $\sum
712: _{i=k(n)}^{\infty} \lambda _i / \sum _{i=k(n)}^{\infty} \mu _i <
713: 2\delta / a$. This means $\underline{\lim}_{n \rightarrow
714: \infty} \sum _{i=n}^{\infty} \lambda _i / \sum _{i=n}^{\infty} \mu
715: _i =0$, which means $\ket{\Psi }$ cannot be convertible to
716: $\ket{\Phi }$ by SLOCC from Vidal's Theorem.
717:
718: Proof of part 2: \\
719: If $R^+(\ket{\Phi}) < R^-(\ket{\Psi})$, then for all
720: $R^+(\ket{\Phi}) < r < R^-(\ket{\Psi})$, $a \stackrel{def}{=}
721: \underline{\lim} _{n \rightarrow \infty} \sum _{i=n}^{\infty}
722: \lambda _i / f_r(n) >0$ and $\lim _{n \rightarrow \infty} \sum
723: _{i=n}^{\infty} \mu _i / f_r(n) =0$. Then, for all $\delta > 0$
724: there exists an $N_0(\delta)$ such that if $n > N_0(\delta)$,
725: $\sum _{i=n}^{\infty} \lambda _i / \sum _{i=n}^{\infty} \mu _i
726: > a / 2\delta$. That is $\lim _{n \rightarrow \infty} \sum
727: _{i=n}^{\infty} \lambda _i / \sum _{i=n}^{\infty} \mu _i =
728: \infty$. From Vidal's theorem, this means that $\ket{\Psi }$ can
729: be converted to $\ket{\Phi }$ by SLOCC. \hfill $\square$
730: \end{Proof}
731: Hence, % not only the monotonicity for SLOCC,
732: this SLOCC monotone satisfies the sufficient condition of
733: convertibility of SLOCC at least with the above meaning. As we
734: shall see in the following sections by using
735: $R^+_{f_r}(\ket{\Psi})$ and $R^-_{f_r}(\ket{\Psi})$ together, we
736: can determine the classification of SLOCC convertibility better
737: than in the case of using the other SLOCC monotones, although we
738: also need both $R^+_{f_r}$ and $R^-_{f_r}$ to lead to the
739: sufficient condition. Therefore, we can consider this pair of
740: monotones as extensions of the Schmidt rank.
741:
742: In the last part of this subsection we note one important fact
743: which we can easily see from Theorem \ref{theorem of monotone},
744: that is that ``{\it in infinite dimensional systems there are at
745: least continuously infinitely many different classes of SLOCC
746: convertibility.}'' Since $R^+(\ket{\Psi})$ (or $R^- (\ket{\Psi})$)
747: is an SLOCC monotone whose range is a non-trivially connected set
748: (interval) of real numbers, if $\ket{\Psi _r}$ satisfies
749: $R^+(\ket{\Psi _r}) =r$, each $\ket{\Psi _r}$ should belong to
750: different classes of SLOCC convertibility for every different
751: value of $r$. That is, there exists an injective map from a
752: non-trivially connected set of real numbers to the quotient set of
753: states by SLOCC. Therefore, in infinite-dimensional bipartite
754: systems, the cardinal number of the quotient set of states by
755: SLOCC convertibility is greater than or equal to the cardinal
756: number of the {\it continuum}, (where the cardinal number of the
757: continuum is equal to the cardinal number of an arbitrary interval
758: of real numbers)\cite{kolmogorov}. Comparing this to the
759: finite-dimensional systems case, where the cardinal number of the
760: quotient set of states by SLOCC convertibility is equal to the
761: dimension of the local systems. This fact is remarkable, that is,
762: the cardinal number of such classes is actually larger than the
763: local dimension (which is only {\it countably infinite}) in
764: infinite-dimensional systems.
765:
766: \subsection{Examples of $R^+(\ket{\Psi})$ and $R^-(\ket{\Psi})$}
767: \label{exsample}
768:
769: In this subsection we construct important examples of the SLOCC
770: monotones $R^+(\ket{\Psi})$ and $R^-(\ket{\Psi})$, and analyze
771: SLOCC convertibility between some interesting classes of genuinely
772: infinite-dimensional states. One is a class of states with
773: polynomially-damped Schmidt coefficients; another is the class of
774: two-mode squeezed states. Since these new monotones depend on a
775: real parameterized family of sequences $\{ f_r(n) \}_{r \in
776: (a,b)}$, we need to choose this family suitably to analyze SLOCC
777: convertibility among particular states. For this purpose it is
778: convenient to derive the reference-states class $\{ \ket{\Psi _r}
779: \}_{r \in (a,b)}$ for particular $\{ f_r(n) \}_{r \in (a,b)}$, as
780: the states which satisfy the condition $R^+(\ket{\Psi _r}) =
781: R^-(\ket{\Psi _r}) =r$. Therefore, at first we construct a way of
782: finding the reference class $\ket{\Psi _r}$ from $f_r(n)$. The
783: following corollary gives a method.
784: \begin{Corollary} \label{reference}
785: If $\{ f_r (n) \} _{r \in (a,b), n \in \mathbb{N}}$ satisfies
786: following conditions:
787: \begin{enumerate}
788: \item $\forall r \in (a,b), n_1 \le n_2 \Rightarrow
789: f_r(n_1) > f_r(n_2)$ (monotonically decreasing)
790: % f_r(n_2) < f_r(n_1)$ (monotonically decreasing)
791: \label{monotonically decreasing}
792: \item $\forall r \in (a,b)$ and $\forall n \in \mathbb{N},
793: f_r(n) + f_r(n+2) \ge 2f_r(n+1)$ (convexity) \label{convexity}
794: \item $\forall m \in \mathbb{N}$, $r_1 \le r_2 \Leftrightarrow
795: \lim _{n \rightarrow \infty}
796: \frac{f_{r_1}(n)}{f_{r_2}(n+m)} =0$ (monotonicity)\;,
797: \label{monotonisity}
798: \end{enumerate}
799: then, $\ket{\Psi _r} = \frac{1}{c_r} \sum _{n=1}^{\infty} \sqrt{-
800: f^{'}_r (n)} \ket{n} \otimes \ket{n}$, where $c_r= \sum _{n
801: =1}^{\infty} -f^{'}_r(n)$, satisfies $R^+(\ket{\Psi _r}) =
802: R^-(\ket{\Psi _r}) = r$ which are made from $\{ f_r (n) \} _{r \in
803: \mathbb{R}, n \in \mathbb{N}}$, and where $f_r'(x)$ denotes the
804: derivative of $f_r(x)$ with respect to $x$.
805:
806: \end{Corollary}
807: \begin{Proof}
808: From conditions \ref{monotonically decreasing} and \ref{convexity}
809: above, there exists a class of doubly differentiable functions $\{
810: f_r(x) \} _{r \in (a,b), x \in \mathbb{R^+ }}$ which are an
811: extension of the sequences $\{ f_r (n) \} _{r \in (a,b), n \in
812: \mathbb{N}}$ such that they satisfy $f^{'}_r(x) < 0$ and
813: $f^{''}_r(x) \ge 0$. Therefore, a class of states $\{ \ket{\Psi
814: _r} \}_{r \in (a,b)}$ is well defined and their Schmidt
815: coefficients are $\{ -f^{'}_r(n) /c_r \} _{n=1}^{\infty}$ in
816: decreasing order. By definition then
817: \begin{eqnarray}
818: \int _{n}^{\infty} \frac{-f^{'}_r(x)}{c_r} dx & \le & \sum
819: _{k=n}^{\infty} \frac{-f^{'}_r(n)}{c_r} \le \int _{n-1}^{\infty}
820: \frac{-f^{'}_r(x)}{c_r} dx
821: \nonumber \\
822: \frac{f_r(n)}{c_rf_{r_1}(n)} & \le & \sum _{k=n}^{\infty}
823: \frac{-f^{'}_r(n)}{c_rf_{r_1}(n)} \le
824: \frac{f_r(n-1)}{c_rf_{r_1}(n)}\;. \nonumber
825: \end{eqnarray}
826: If $r<r_1$ then
827: \begin{equation}
828: \lim _{n \rightarrow \infty} \sum _{k=n}^{\infty}
829: \frac{-f^{'}_r(n)}{c_rf_{r_1}(n)} \le \lim _{n \rightarrow \infty}
830: \frac{f_r(n-1)}{c_rf_{r_1}(n)} = 0\;,\nonumber
831: \end{equation}
832: and if $r>r_2$ then
833: \begin{equation}
834: \lim _{n \rightarrow \infty} \sum _{k=n}^{\infty}
835: \frac{-f^{'}_r(n)}{c_rf_{r_1}(n)} \ge \lim _{n \rightarrow \infty}
836: \frac{f_r(n)}{c_rf_{r_1}(n)} = +\infty \;.\nonumber
837: \end{equation}
838: Thus, $R^+(\ket{\Psi _r}) = R^-(\ket{\Psi _r}) = r$. \hfill
839: $\square$
840: \end{Proof}
841: This Corollary means that with the above three additional
842: conditions for $f_r(n)$ we may always derive a class of reference
843: states which correspond to each value of $R^-(\ket{\Psi})$ and
844: $R^+(\ket{\Psi})$.
845:
846: In what follows we construct examples of SLOCC monotones by means
847: of the above Corollary and analyze two remarkable classes of
848: states. One corresponds to the states which belong a higher class
849: of SLOCC convertibility and the other to the well-known two-mode
850: squeezed states.
851:
852: As a first example consider $R^-(\ket{\Psi})$ and
853: $R^+(\ket{\Psi})$ made from $\{ f_r(n) = n^{-(\frac{1}{r}-1)} \}
854: _{r \in (0,1)}$. By Corollary \ref{reference}, their class is
855: \begin{equation}
856: \ket{\Psi _r} = \frac{1}{\sqrt{\zeta (1/r)}} \sum _{n=1}^{\infty}
857: n^{-\frac{1}{2r}} \ket{n} \otimes \ket{n}\;,
858: \end{equation}
859: where $\zeta (x)$ is the Riemann zeta function as a normalization
860: factor. By definition, $R^-(\ket{\Psi})$ and $R^+(\ket{\Psi})$
861: represent how quickly the Schmidt coefficients of $\ket{\Psi}$
862: converge to $0$ as a polynomially-damped function. Thus, this
863: function has a strictly positive value for states with
864: polynomially-damped Schmidt coefficients. Similarly, for all
865: states $\ket{\Psi}$ for which the Schmidt coefficients damp
866: exponentially like two-mode squeezed states, we have
867: $R^-(\ket{\Psi}) = R^+(\ket{\Psi}) =0$. Because the Schmidt
868: coefficients can never be proportional to $1/n$ asymptotically in
869: infinite-dimensional systems (since $\sum _{n=1}^{\infty} 1/n =
870: \infty$), for small $\epsilon > 0$, $\ket{\Psi _r}$ with $r=1-
871: \epsilon$ can be converted to almost any state. In the above sense
872: we can say that they belong to a ``higher rank'' of entanglement
873: class in terms of single-copy SLOCC. On the other hand, the above
874: state with small $\epsilon$ is not of the ``{\it highest
875: states}''. That is, we can consider a class of states which belong
876: to a higher rank of entanglement class than $\{ \ket{\Psi _r}
877: \}_{r \in (0,1)}$ as follows. For the states
878: \begin{equation}
879: \ket{\Psi _t} = \frac{1}{C_t} \sum_{n=1}^{\infty}
880: \frac{1}{\sqrt{x(\log x)^t}} \ket{n} \otimes \ket{n}\;,
881: \end{equation}
882: with $t>0$, $R^-(\ket{\Psi _t}) = R^+(\ket{\Psi _t})=1$, and we
883: can easily see that for all $t >0$, $\ket{\Psi _r}$ can not be
884: converted to $\ket{\Psi _t}$ by SLOCC. In a similar manner, for all
885: one-parameter classes of states we can always define a class of
886: states which belong to a higher rank and can define a new pair of
887: monotones from this class of states. Therefore, there does not
888: exist a highest one-parameter class of states within the SLOCC
889: classification.
890:
891: As a next example consider $f_q(n) = e^{2n \log q} = q^{2n}$, $q
892: \in (0,1)$. In this case, the reference class is $\ket{\Psi _q}=
893: \frac{1}{c_q} \sum _{n=1}^{\infty} q^n \ket{n} \otimes \ket{n}$,
894: that is the well-known two-mode squeezed states with
895: $\frac{1}{2}\log \frac{1+q}{1-q}$ being the squeezing parameter.
896: Therefore, $R^+_{f_q}(\ket{\Psi})$ and
897: $R^-_{f_q}(\ket{\Psi})$ can be regarded as being analogs of
898: squeezing parameters for any entangled states.
899:
900: The above two examples also show that the classification of SLOCC
901: is quite different from the classification by the amount of
902: entanglement, that is, the classification of asymptotic
903: (infinite-copy) LOCC in infinite-dimensional systems. In
904: infinite-dimensional systems we often consider the class of
905: two-mode squeezed states
906: $\ket{\Psi_q}=\frac{1}{c_q}\sum_{n=1}^{\infty}q^n\ket{n}\otimes\ket{n}$
907: with $q = 1 - \epsilon$ instead of the maximally entangled states.
908: Because $ \lim _{q \rightarrow 1} E(\ket{\Psi _q}) = \infty$ this
909: state converts to almost any state asymptotically by infinite-copy
910: LOCC with unit probability. However by single-copy SLOCC with
911: non-zero probability they cannot be converted to states
912: $\ket{\Psi}$ with $R^+(\ket{\Psi})>0$, where the monotone
913: $R^+(\ket{\Psi})$ is made from $f_r (n) = n^{-(\frac{1}{r}-1)}$.
914: On the other hand, if we consider the class of states $\ket{\Psi
915: _r}= \sum _{n=1}^{\infty} n^{-\frac{1}{2r}} \ket{n} \otimes
916: \ket{n}/\sqrt{\zeta (1/r)}$ as we have already seen for small
917: $\epsilon >0$, $\ket{\Psi _{1-\epsilon}}$ can be converted to
918: almost any state by single-copy SLOCC with non-zero probability.
919: Although the amount of entanglement for both $\{ \ket{\Psi _q} \}$
920: and $\{ \ket{\Psi _r } \}$ tend to infinity in the limit, $\{
921: \ket{\Psi _r} \}$ belongs to a higher class than $\{ \ket{\Psi _q}
922: \}$ in the single-copy scenario.
923:
924: We add one final remark here: Although, we have only presented two
925: examples for $f_r(n)$, there may be many other examples which are
926: important in some situations. Generally speaking, for any given
927: states, we can find a suitable function $f_r(n)$ for the analysis
928: of the states. For example, if we deal with states whose Schmidt
929: coefficients damp exponentially we can chose $f_r(n) = \exp
930: (n^{-1/r}), \exp (\exp (n^{-1/r})), $ etc, as the
931: coefficients damp quickly enough to evaluate the monotones for the
932: states.
933:
934: \subsection{Strong inhibition law}
935: \label{inhibition law}
936: %%%%%%%%%%% Stronger statement of inhibition law for SLOCC convertibility %%%%%
937:
938: So far we have emphasized the difference between states with
939: exponentially-damped Schmidt coefficients and those with
940: polynomially-damped coefficients and shown that
941: exponentially-damped states cannot be converted into
942: polynomially-damped states no matter how large their measure of
943: entanglement is. Here we shall give one more fact which will
944: demonstrate the remarkable difference between exponentially and
945: polynomially-damped states. That is, ``\textit{However finitely
946: many copies there are, exponentially-damped states cannot be
947: converted into polynomially-damped states}.'' This fact can be
948: showed as the follows: Suppose $\ket{\Psi}$ is an
949: exponentially-damped state and $\ket{\Phi}$ is a
950: polynomially-damped one, then rigorously speaking, there exists a
951: real number $r$ and a polynomial $p(n)$ which satisfy $\lim _{n
952: \rightarrow \infty} \frac{g_{\ket{\Psi}}(n)}{e^{-rn}}=0$ and
953: $\underline{\lim}_{n\rightarrow \infty}
954: \frac{p(n)}{g_{\ket{\Phi}}(n)} =0$, where $g_{\ket{\Psi}}(n)$ is
955: Vidal's monotone of $\ket{\Psi}$. Define $\ket{\xi _r} =
956: \frac{1}{C_r} \sum _{n=1}^{\infty} e^{-rn} \ket{n}\otimes\ket{n}$,
957: then we have
958: \begin{eqnarray}
959: \ket{\xi _r} ^{\otimes p}&=& \frac{1}{C_r^p} \sum _{n_1, n_2,
960: \cdots ,n_p}^{\infty} e^{-r(n_1+n_2+ \cdots +n_p)} \ket{n_1, n_2,
961: \cdots , n_p}\otimes \left \vert n_1, n_2,\cdots,n_p\right\rangle
962: \nonumber \;.\!\!
963: \end{eqnarray}
964: If we reorder the Schmidt terms to the form $\ket{\xi _r}
965: ^{\otimes p} = \frac{1}{C} \sum _{k=1}^{\infty} f(k)
966: \ket{k}\otimes \ket{k}$ we can see by easy calculation that $f(k)
967: \le e^{-r[ (p! k )^{1/p} +1]}$. Thus, we have $\lim _{n
968: \rightarrow \infty} {f(n)}/{p(n)} =0$ and this means $\ket{\Psi}
969: ^{\otimes p}$ cannot converted into $\ket{\Phi}$ for any $p \in
970: \mathbb{N}$. This result shows that in infinite-dimensional
971: systems some classes of states (like states with finite Schmidt
972: ranks, with exponentially-damped Schmidt coefficients, and with
973: polynomially-damped Schmidt coefficients) can be distinguished
974: from each other more strongly than the case of finite-dimensional
975: systems by SLOCC classification. Thus, \textit{with arbitrary
976: finitely-many copies}, we also cannot convert the states from
977: finite rank to infinite rank, and similarly from exponentially
978: damped to polynomially damped. In finite-dimensional systems,
979: there is no feature like this. Therefore, these properties of
980: entanglement are genuine for infinite-dimensional systems and show
981: the special strong position of states with polynomially-damped
982: Schmidt coefficients from the view of finite-copy transformations.
983:
984: As a final remark for this section we must discuss the energy of
985: such long-tailed states. In realistic situations the set of states
986: which can be produced experimentally will be limited by some bound
987: in energy. Therefore, it is essential to consider the subset of
988: states which consist of states restricted to that bounded energy.
989: However, for several states with polynomially-damped Schmidt
990: coefficients, the mean value of a polynomial Hamiltonian, like for
991: example the harmonic oscillator, diverges. Therefore, generally
992: only a fraction of polynomially-damped states can be created in
993: laboratories.
994:
995: %%%%%%%%%%% summary %%%%%%%%%%%
996:
997: \section{Summary}
998: \noindent
999: In this paper in order to avoid the difficulties of discontinuity
1000: and infinite amounts of classical communication in the theory of
1001: SLOCC convertibility of infinite-dimensional systems, we proposed
1002: a new definition of convertibility, {\it
1003: $\epsilon$-convertibility}, as the convertibility of states in an
1004: approximated setting by means of the trace norm. In the Section
1005: \ref{infinite Nielsen} we showed that this definition guarantees
1006: at least weak continuity for SLOCC convertibility (Lemma
1007: \ref{epsilon SLOCC lemma}), and guarantees that the protocol only
1008: uses finite amounts of classical communication. Then, we
1009: reconstructed the basic theorems of single-copy LOCC and SLOCC
1010: transformation, Neilsen's and Vidal's theorem in the
1011: infinite-dimensional pure state space (Theorems \ref{epsilon
1012: Nielsen} and~\ref{epsilon Vidal}). As a result we showed that
1013: under this change of definition the framework of entanglement
1014: convertibility is preserved for infinite-dimensional systems, and
1015: therefore, our definition of $\epsilon$-convertibility for LOCC is
1016: suitable and sufficient for realistic conditions of quantum
1017: information processing in infinite-dimensional systems.
1018:
1019: In Section \ref{Extension} in order to study SLOCC convertibility
1020: in infinite-dimensional systems, we constructed a pair of SLOCC
1021: monotones which can be considered as extensions of the Schmidt
1022: rank to infinite-dimensional spaces. By these monotones we showed
1023: that states with polynomially-damped Schmidt coefficients belong
1024: to a higher rank of entanglement class than other states in terms
1025: of single-copy SLOCC convertibility.
1026:
1027: In the last Section \ref{inhibition law} we showed that arbitrary
1028: finitely many copies of exponentially-damped states cannot be
1029: converted to even a single copy of polynomially-damped states.
1030: Since such differences of classes do not exist in the
1031: finite-dimensional setting, the SLOCC classification of
1032: infinite-dimensional states has a much richer structure than for
1033: finite-dimensional ones. Therefore, these new features of
1034: entanglement have the potential to produce new quantum information
1035: protocols which are impossible for finite-dimensional systems.
1036: Finally, we stress that in infinite-dimensional systems, there
1037: remain important problems that are yet to be solved even for the
1038: simplest bipartite pure states.
1039:
1040:
1041: \nonumsection{Acknowledgements} \noindent
1042: MO is grateful to Professor M.\ Ozawa, Professor M.B.\ Plenio,
1043: Professor K.\ Matsumoto, Professor M.\ Hayashi, and Dr.\ A.\
1044: Miyake for discussions. This work has been supported by the Asashi
1045: Grass Foundation, the Sumitomo Foundation, the Japan Society of
1046: Promotion of Science, the Japan Scholarship Foundation, the
1047: Japan Science and Technology Agency,
1048: and the Special Coordination Funds for Promoting Science and Technology.
1049:
1050:
1051:
1052: \nonumsection{References} \noindent
1053: \vspace{-0.5cm}
1054: \begin{thebibliography}{000}
1055: \bibitem{epr} A.\ Einstein, B.\ Podolsky, and N.\ Rosen {\it
1056: Phys.\ Rev.\/} {\bf 47}, 777 (1935); J.S.\ Bell {\it Physics} {\bf
1057: 1}, 195 (1964); J.F.\ Clauser, M.A.\ Horne, A.\ Shimony. R.A.\
1058: Holt {\it Phys.\ Rev.\ Lett.\/} {\bf 23}, 880 (1969); R.F.\
1059: Werner, {\it Phys.\ Rev.\ A} {\bf 40} 4277 (1989).
1060:
1061: \bibitem{resource} A.K.\ Ekert, {\it Phys.\ Rev.\ Lett.\/} {\bf
1062: 68}, 661 (1991); P.W.\ Shor, Proc.\ {\it 35nd Annual Symposium on
1063: Foundations of Computer Science}, (IEEE Computer Society Press,
1064: 1994), 124-134.
1065:
1066: \bibitem{quantum communication} C.H.\ Bennett and S.J.\ Wiesner
1067: {\it Phys.\ Rev.\ Lett.\/} {\bf 69}, 2884 (1992); C.H.\ Bennett,
1068: G.\ Brassard, C.\ Crepeau, R.\ Jozsa, A.\ Peres, and W.K.\
1069: Wootters, {\it Phys.\ Rev.\ Lett.\/} {\bf 70}, 1895 (1993).
1070:
1071: \bibitem{neumann} J.\ von Neumann, {\itshape Mathematical
1072: Foundations of Quantum Mechanics} (Princeton University Press,
1073: Princeton, New Jersey, 1955).
1074:
1075: \bibitem{functional analysis} M.\ Reed, B.\ Simon {\it Functional
1076: Analysis
1077: (Methods of Modern Mathematical Physics)} (Academic Press, 1980).
1078:
1079: \bibitem{bennett} C.H.\ Bennett, D.P.\ DiVincenzo, J.A.\ Smolin
1080: and W.K.\ Wootters, {\it Phys.\ Rev.\ A} {\bf 54}, 3824 (1996);
1081: C.H.\ Bennett, G.\ Brassard, S.\ Popescu, B.\ Schumacher, J.A.\
1082: Smolin and W.K.\ Wootters, {\it Phys.\ Rev.\ Lett.\/} {\bf 76},
1083: 722 (1996); C.H.\ Bennett, H.J.\ Berstein, S.\ Popescu and B.\
1084: Schumacher, {\it Phys.\ Rev.\ A} {\bf 53}, 2046 (1996).
1085:
1086: \bibitem{majorization} M.A.\ Nielsen, {\it Phys.\ Rev.\ Lett.\/}
1087: {\bf 83}, 436 (1999).
1088:
1089: \bibitem{vidal} G.\ Vidal, {\it Phys.\ Rev.\ Lett.\/} {\bf 83},
1090: 1046 (1999).
1091:
1092: \bibitem{monotone} G.\ Vidal {\it J.\ Mod.\ Opt.\/} {\bf 47}, 355
1093: (2000).
1094:
1095:
1096: \bibitem{separability}L.-M.\ Duan, G.\ Giedke, J.I.\ Cirac, and
1097: P.\ Zoller, {\it Phys.\ Rev.\ Lett.\/} {\bf 84}, 2722 (2000); R.\
1098: Simon, {\it Phys.\ Rev.\ Lett.\/} 84, 2726 (2000); G.\ Giedke, B.\
1099: Kraus, M.\ Lewenstein, J.I.\ Cirac, {\it Phys.\ Rev.\ A} {\bf 64},
1100: 052303 (2001).
1101:
1102: \bibitem{gaussian} G.\ Giedke, J.\ Eisert, J.I.\ Cirac and M.B.\
1103: Plenio, {\it Quant.\ Inf.\ Comp.\/} {\bf 3}, 211 (2003); G.\
1104: Giedke, M.M.\ Wolf, O.\ Kruger, R.F.\ Werner and J.I.\ Cirac, {\it
1105: Phys.\ Rev.\ Lett.\/} {\bf 91}, 107901 (2003); M.M.\ Wolf, G.\
1106: Giedke, O.\ Kruger, R.F.\ Werner and J.I.\ Cirac, {\it
1107: quant-ph/}0306177 (2003).
1108:
1109: \bibitem{gaussian distill} G.\ Giedke and J.I.\ Cirac {\it Phys.\
1110: Rev.\ A} {\bf 66}, 032316 (2002); J.\ Fiurasek {\it Phys.\ Rev.\
1111: Lett.\/} {\bf 89}, 137904 (2002); J.\ Eisert, S.\ Scheel and M.B.\
1112: Plenio, {\it Phys.\ Rev.\ Lett.\/} {\bf 89}, 137903 (2002); J.\
1113: Eisert, D.\ Browne, S.\ Scheel and M.B.\ Plenio, {\it Ann.\
1114: Phys.\/} (NY) {\bf 311}, 431 (2004).
1115:
1116: \bibitem{entanglement measure} J.\ Eisert, C.\ Simon, and M.B.\
1117: Plenio, {\it J.\ Phys.\ A} {\bf 35}, 3911, (2002); M.\ Keyl, D.\
1118: Schlingemann, and R.F.\ Werner, {\it Quant.\ Inf.\ Comp.\/} {\bf
1119: 3}, 281 (2003); M.M.\ Wolf, G.\ Giedke, O.\ Krueger, R.F.\ Werner,
1120: J.I.\ Cirac {\it Phys.\ Rev.\ A} {\bf 69}, 052320 (2004).
1121:
1122:
1123: \bibitem{LOCC} V.\ Vedral, M.B.\ Plenio, M.A.\ Rippin, P.L.\
1124: Knight, {\it Phys.\ Rev.\ Lett.\/} {\bf 78}, 2275 (1997); E.M.\
1125: Rains, {\it IEEE Trans.\ Inf.\ Tec.\/} {\bf 47}, 2921 (2001).
1126:
1127: \bibitem{uniqueness} M.J.\ Donald, M.\ Horodecki, O.\ Rudolph,
1128: {\it J.\ Math.\ Phys.\/} {\bf 43}, 4252 (2002).
1129:
1130: \bibitem{lo-popescu} H.-K.\ Lo and S.\ Popescu, {\it Phys.\ Rev.\
1131: A} {\bf 63}, 022301 (2001).
1132:
1133: \bibitem{kolmogorov}A.N.\ Kolmogorov and S.V.\ Fomin, {\it
1134: Introductory real analysis} (Dover Publications, Inc, 1975)
1135:
1136: \bibitem{Bhartia}R.\ Bhatia, {\it Matrix analysis}
1137: (Springer-Verlag, New York, 1997).
1138:
1139:
1140: \bibitem{D'ariano} G.M.\ D'Ariano, M.F.\ Sacchi, {\it Phys.\
1141: Rev.\ A} {\bf 67}, 042312 (2003).
1142:
1143: \bibitem{markus} A.S.\ Markus, {\it Russian Math.\ Surveys} {\bf 19}, 91 (1964)
1144:
1145: \end{thebibliography}
1146:
1147: \appendix{\ Schmidt decomposition and Lo-Popescu's
1148: Theorem in infinite dimensional systems} \label{Lo-Popescu section}
1149:
1150: %%%%%%%%%%% Schmidt decomposition and Lo-Popescu's Theorem %%%%%%%
1151:
1152:
1153: In this appendix and the next, as a preparation for the proofs
1154: of Nielsen's and Vidal's theorem in infinite dimensional systems,
1155: we will see how we can extend basic theorems about LOCC and
1156: majorization \cite{majorization, vidal} to infinite dimensional
1157: systems.
1158:
1159: At first, we extend the concept of Schmidt decomposition and
1160: Schmidt coefficients in infinite-dimensional systems:
1161: \begin{Theorem}{(Schmidt decomposition)}
1162: For any $\ket{\Psi} \in \Hi = \Hi _A \otimes \Hi _B$, there exist
1163: orthonormal sets ({\it but not necessarily basis sets}) $\{
1164: \ket{e_i} \} _{i=1}^{\infty} $ and $\{ \ket{f_i} \}
1165: _{i=1}^{\infty} $ of $\Hi _A$, and $\Hi _B$, respectively, such
1166: that
1167: \begin{equation} \label{Schmidt inf}
1168: \ket{\Psi} = \sum _{i=1}^{\infty} \sqrt{\lambda _i} \ket{e_i}
1169: \otimes \ket{f_i} \:,
1170: \end{equation}
1171: where $\lambda _i \ge 0$, $\lambda _i \ge \lambda _{i+1} $ and
1172: $\sum _{i=1}^{\infty} \lambda _i = 1$. The representation of a
1173: state $\ket{\Psi}$ in the form of Eq.(\ref{Schmidt inf}) is called
1174: a Schmidt decomposition and $\{ \lambda _i \} _{i=1}^{\infty}$ are
1175: called Schmidt coefficients in infinite-dimensional systems.
1176: \end{Theorem}
1177: \begin{Proof}
1178: We use the singular value decomposition given as follows in an
1179: infinite dimensional system: For a compact operator $M$ from $\Hi
1180: _A$ onto $\Hi _B$, there exist orthonormal sets (but not necessarily
1181: basis sets) $\{ \ket{e_i} \}_{i=1}^{\infty} \subset \Hi _A$
1182: and $\{ \ket{f_i} \}_{i=1}^{\infty} \subset \Hi _B$ and positive
1183: real numbers $\{ \lambda _i \}_{i=1}^{\infty}$ with
1184: $\sqrt{\lambda _n} \rightarrow 0$ such that
1185: \begin{equation}\label{Schmidt eq}
1186: M = \sum _{i=1}^{\infty} \sqrt{\lambda _i}\ket{e _i} \bra{f _i},
1187: \end{equation}
1188: where the above sum converges in the operator norm \cite{functional
1189: analysis}. In particular, if $M$ is a Hilbert-Schmidt class
1190: operator, $\{ \sqrt{\lambda _i } \}_{i=1}^{\infty}$ satisfy $\sum
1191: _{i=1}^{\infty} \lambda _i = (\| M \|_2)^2 \stackrel{\rm def}{=}
1192: \Tr M^{\dagger}M$, where $\| \cdot \|_2$ is the Hilbert-Schmidt
1193: norm \cite{functional analysis}. Thus, we derive Eq.(\ref{Schmidt
1194: inf}) from Eq.(\ref{Schmidt eq}), because the linear map
1195: $\ket{e_i} \bra{f_j} \mapsto \ket{e_i} \otimes \ket{f_j}$ gives an
1196: isomorphism from the Hilbert-Schmidt space $\mathfrak{C} _2(\Hi
1197: _A, \Hi _B)$ (the Hilbert space of all Hilbert-Schmidt class
1198: operators between $\Hi_A$ and $\Hi _B$ with the inner product
1199: $\left(M |N \right) \stackrel{\rm def}{=} \Tr M^{\dagger} N $) to
1200: the Hilbert space $\Hi = \Hi _A \otimes \Hi _B$ \cite{functional
1201: analysis}. \hfill $\square$
1202: \end{Proof}
1203:
1204: In finite $d$-dimensional bipartite systems, the Schmidt
1205: decomposition of a state $\ket{\psi}$ is given by
1206: \begin{equation} \label{Schmidt}
1207: \ket{\psi} = \sum _{i=1}^{d} \sqrt{\lambda _i} \ket{e_i}
1208: \otimes \ket{f_i}\;,
1209: \end{equation}
1210: where $\{ \ket{e_i} \} _{i=1}^{d}$ and $\{ \ket{f_i} \}
1211: _{i=1}^{d}$ are the basis sets. Therefore, convertibility of
1212: states under local {\it unitary} operations are determined by the
1213: Schmidt coefficients $\{ \lambda _i \} _{i=1}^{d}$ of states,
1214: namely, the two states are convertible to each other under local
1215: unitary operations if and only if the two states have same
1216: Schmidt coefficients. In infinite-dimensional systems, the
1217: Schmidt coefficients determine convertibility of states under
1218: local {\it partial isometry} instead of local unitary operations.
1219: That is, if the Schmidt coefficients of $ \ket{\Psi}$ and $\ket{\Phi}$
1220: are the same, then there exist local partial isometries $U_A$ and $U_B$
1221: and $\ket{\Psi} = U_A \otimes U_B \ket{\Phi}$ is satisfied.
1222:
1223: Partial isometry is defined as a unitary operator between
1224: subspaces. If we had defined the Schmidt coefficients to be a
1225: sequence including the dimension of the kernel of the reduced
1226: density matrix (of the given state), we could make Schmidt
1227: coefficients indicating the convertibility under local unitary
1228: operations. However, to develop the theory of LOCC and SLOCC, (which
1229: include local partial isometries), convertibility for
1230: infinite-dimensional systems, the former definition is more
1231: suitable than the latter, so we take the definition of Eq.
1232: {\ref{Schmidt inf}}. This is because states are convertible to
1233: each other by LOCC, if and only if they are convertible to each
1234: other by local partial isometries (we can show this fact from
1235: Theorem \ref{infinite Nielsen}); moreover, there exists a pair of
1236: states which are convertible to each other by local partial isometries,
1237: but not by local unitaries. The following example satisfies such
1238: a condition. Suppose states $\ket{\Psi}$ and $\ket{\Phi}$ on $\Hi _A
1239: \otimes \Hi _B$ are defined as
1240: \begin{eqnarray}
1241: \ket{\Psi} &=& \sum _{i=1}^{\infty} \sqrt{\lambda _i} \ket{e_i}
1242: \otimes \ket{f_i} \label{def of psi} \\
1243: \ket{\Phi} &=& \sum _{i=1}^{\infty} \sqrt{\lambda _i} \ket{e_{2i}}
1244: \otimes \ket{f_{2i}}, \label{def of phi}
1245: \end{eqnarray}
1246: where $\{ \ket{e_i } \}_{i=1}^{\infty}$ and $\{ \ket{f_i}
1247: \}^{\infty}$ are orthonormal basis sets of $\Hi _A$ and $\Hi _B$,
1248: respectively. In this case, $\ket{\Phi }$ and $\ket{\Psi }$ can be
1249: convertible to each other by LOCC: $\ket{\Phi }$ can be
1250: convertible to $\ket{\Psi }$ by a local partial isometry $\sum
1251: _{i=1}^{\infty} \ket{e_{i}}\bra{e_{2i}} \otimes
1252: \ket{f_i}\bra{f_{2i}}$, and $\ket{\Psi }$ can be convertible to
1253: $\ket{\Phi }$ by a local isometry $\sum _{i=1}^{\infty}
1254: \ket{e_{2i}}\bra{e_{i}} \otimes \ket{f_{2i}}\bra{f_{i}}$. However,
1255: their are not convertible by local unitaries. This is because a
1256: subspace spanned by $\{ \ket{e_{2i}} \}_{i=1}^{\infty}$ and a
1257: subspace spanned by $\{ \ket{f_{2i}}_{i=1}^{\infty}$ should be mapped to $\Hi
1258: _A$ and $\Hi _B$, respectively; a proper subspace should be mapped
1259: to a whole space. This map is obviously impossible by unitary
1260: operators, which are bijections and always map a whole space to a
1261: whole space.
1262:
1263: Actually, from a physical point of view, convertibility under local
1264: partial isometry can be understood that we may need additional
1265: independent ancilla systems for each subspace to convert
1266: $\ket{\Psi}$ to $\ket{\Phi}$. For example, $\ket{\Phi}$ defined as
1267: Eq.(\ref{def of phi}) can be transformed to $\ket{\Psi}$ defined
1268: as Eq.(\ref{def of psi}) by the following protocol: First, attach
1269: one-qubit ancilla systems $\Hi _{A'}$ and $\Hi _{B'}$ to both
1270: local systems $\Hi _A$ and $\Hi _B$, and prepare the ancilla
1271: systems in $\ket{0}_{A'}$ and $\ket{0}_{B'}$. Second, apply a
1272: local unitary transformation $U_{AA'} \otimes U_{BB'}$ to the
1273: states $\ket{\Phi}_{AB} \otimes \ket{0}_{A'} \otimes
1274: \ket{0}_{B'}$, where $U_{AA'}$ and $U_{BB'}$ are unitary
1275: transformations on $\Hi _A \otimes \Hi _{A'}$ and $\Hi _B \otimes
1276: \Hi _{B'}$ defined as
1277: \begin{eqnarray*}
1278: U_{AA'} &\stackrel{\rm def}{=}& \sum _{n=1}^{\infty} \big(
1279: |n \rangle_{A} |0 \rangle _{A'} {}_{A} \langle 2n | {}_{A'} \langle 0| +
1280: | 2n+1 \rangle _A |1 \rangle _{A'} {}_{A}\langle 2n+1| {}_{A'}\langle 0| \\
1281: &\quad & \qquad + |2n \rangle_A |1 \rangle _{A'} {}_A \langle n|
1282: {}_{A'}\langle 1| \big) \\
1283: U_{BB'} &\stackrel{\rm def}{=}& \sum _{n=1}^{\infty} \big(
1284: |n\rangle_{B} |0\rangle_{B'} {}_B \langle 2n| {}_{B'}\langle 0| +
1285: | 2n+1 \rangle_B |1\rangle_{B'} {}_B\langle 2n+1| {}_{B'}\langle 0| \\
1286: &\quad & \qquad + |2n\rangle_B |1\rangle_{B'} {}_B\langle n| {}_{B'}\langle 1| \big).
1287: \end{eqnarray*}
1288: After this local unitary transformation, the state is changed to
1289: $\ket{\Psi} _{AB} \otimes \ket{0}_{A'} \otimes \ket{0}_{B'}$.
1290: Finally, by removing the ancilla system $\Hi _{A'}$ and $\Hi
1291: _{B'}$, we derive $\ket{\Psi}$ on the systems $\Hi _A \otimes \Hi
1292: _B$.
1293:
1294: For finite-dimensional systems, Lo and Popescu proved
1295: that if $\ket{\Psi}$ can be converted into $\ket{\Phi}$ by LOCC,
1296: there exists a one-way classical communication LOCC which consists
1297: of a local measurement of one of the local spaces and a unitary
1298: operation of the other depending on the result of the measurement
1299: \cite{lo-popescu}. This theorem is called the Lo-Popescu theorem.
1300: Intuitively speaking, the Schmidt decomposition denotes the existence
1301: of symmetry between local subspaces, and Lo-Popescu's theorem is
1302: the reflection of the symmetry of subsystems. As we have shown that
1303: the Schmidt decomposition in infinite-dimensional systems is
1304: weaker (indicating equivalence under partial isometries instead of
1305: unitary operations) than finite-dimensional systems, the
1306: corresponding Lo-Popescu theorem is slightly modified as
1307: follows.
1308:
1309: \begin{Theorem}[Lo-Popescu's] \label{Lo-Popescu}
1310: In the separable Hilbert space $\Hi _A \otimes \Hi _B$ , for any
1311: given state $\ket{\Psi}$ and bounded operator $M \in \B (\Hi _B)$
1312: there exist a bounded operator $N \in \B (\Hi _A)$ and {\it partial
1313: isometry} $U \in \B (\Hi _B)$ which satisfy $ I \otimes M
1314: \ket{\Psi} = N \otimes U \ket{\Psi}$.
1315: \end{Theorem}
1316:
1317: \begin{Proof}
1318: Suppose $\ket{\Psi} = \sum _{i=1}^{\infty} \sqrt{\mu _i} \ket{a_i}
1319: \otimes \ket{b_i}$. Define a partial isometry $U$ as $U= \sum
1320: _{i=1}^{\infty} \ket{b_i} \bra{a_i}$, then we have
1321: \begin{equation}
1322: I \otimes M \ket{\Psi} =\sum _{i=1}^{\infty} \sqrt{\mu _i}
1323: \ket{a_i} \otimes M \ket{b_i} \nonumber \;,
1324: \end{equation}
1325: and
1326: \begin{equation}
1327: MU\otimes I \ket{\Psi} =\sum _{i=1}^{\infty} \sqrt{\mu _i} M
1328: \ket{b_i} \otimes \ket{b_i}\nonumber \;.
1329: \end{equation}
1330: Thus, we obtain
1331: \begin{eqnarray}
1332: \Tr\; _A I \otimes M \ket{\Psi} \bra{\Psi} I\otimes M^{\dagger} =
1333: \Tr\; _B MU \otimes I \ket{\Psi} \bra{\Psi} U^{\dagger}
1334: M^{\dagger} \otimes I\;.\nonumber
1335: \end{eqnarray}
1336: By our definition of Schmidt decomposition, $\rho _A$ and $\rho
1337: _B$ are partial isometry equivalent for any state $\ket{\Psi}$.
1338: Therefore, there exist partial isometries $U_A$ and $U_B$ which
1339: satisfy
1340: \begin{equation}
1341: U_A \otimes U_B (MU \otimes I) \ket{\Psi} = I \otimes M
1342: \ket{\Psi}\nonumber\;.
1343: \end{equation}
1344: Defining $N \stackrel{\rm def}{=} U_A MU \otimes U_B $, then the
1345: theorem has been proven. \hfill $\square$
1346: \end{Proof}
1347: %
1348: %If one of the party tries to take place of the measurement of the
1349: %other party has to perform a partial isometric operation instead
1350: %of a unitary operation, to compensate the operation of Alice.
1351:
1352:
1353: %%%%%%%%%%% Majorization in infinite-dimensional systems %%%%%%%%%
1354:
1355: \appendix{ \quad Majorization in infinite-dimensional systems}
1356: \label{Majorization}
1357:
1358: In finite-dimensional systems, majorization is a pseudo partial
1359: ordering on the whole vector space \cite{Bhartia}. On the other
1360: hand in infinite-dimensional system, majorization can be defined
1361: on only a subset of the whole vector space. To formulate our
1362: definition of Schmidt coefficients in infinite-dimensional
1363: systems, we define a majorization on
1364: \begin{equation}
1365: l _1 = \{ \{ x_i \} _{i=1}^{\infty} | \sum _{i=1}^{\infty} |x_i| <
1366: \infty \}\;.
1367: \end{equation}
1368: For mathematical simplicity, we only define majorization on
1369: \begin{eqnarray}
1370: l_1 ^+ = \{ \{ x \} _{i=1}^{\infty } \in l_1 | x_i \ge 0,
1371: \sharp \{ i | x_i=0 \} < \infty \nonumber \vee \ \sharp \{ i | x_i > 0 \} < \infty \}\;,
1372: \end{eqnarray}
1373: where $\sharp$ denotes the cardinality of a set. In this case
1374: $l_1^+$ is a convex cone of $l_1$ and identifies the set of all
1375: permutations of Schmidt coefficients.
1376: Thus, $l_1^+$ is enough for our purposes, and we do not need
1377: to define majorization for all element of $l_1$. Moreover,
1378: in the following definition of majorization, we use decreasing reordering
1379: of $x \in l_1$. However, if $x$ is not in $l_1^+$,
1380: it is difficult to rearrange elements of
1381: $x$ in decreasing order,
1382: and we need to extend the definition of majorization
1383: so as to include sequences which are not in $l_1^+$, but in $l_1$.\fnm{{\it a}}\fnt{}{\
1384: The general definition of majorization of sequences can be found in
1385: \cite{markus}. The all propositions in this subsection can be extended to this general case
1386: by appropriate modification of proofs.} This is also the reason why we
1387: define majorization only on $l_1^+$.
1388:
1389: Now, we define majorization
1390: of Schmidt coefficients in infinite-dimensional systems as
1391: follows:
1392: \begin{Definition}
1393: For any $x , y \in l_1^+$, $x \prec _{\omega} y $ (or $x$ is
1394: sub-majorized by $y$) is defined, if and only if
1395: \begin{equation}
1396: \sum _{i=1}^k x^{\downarrow} _i \le
1397: \sum _{i=1}^k y^{\downarrow}_i\;,
1398: \end{equation}
1399: for $k \in \mathbb{N}$, where $x^{\downarrow} _i$ is given by
1400: $x^{\downarrow} _i = x_{P(i)}$, and $P$ is an element of an infinite
1401: symmetry group satisfying $x^{\downarrow} _i \ge x^{\downarrow}
1402: _{i+1}$ in decreasing reordering of $x$. Similarly, if $x$ and $y$
1403: satisfy,
1404: \begin{equation} \label{super-majorization}
1405: \sum _{i=k}^{\infty} x^{\downarrow} _i \ge \sum _{i=k}^{\infty}
1406: x^{\downarrow} _i\;,
1407: \end{equation}
1408: then, we write $x \prec ^{\omega} y$ and say $x$ is
1409: super-majorized by $y$.
1410:
1411: Additional to the sub-majorization or super-majorization conditions,
1412: (both conditions lead to the same majorization condition) if $x$ and
1413: $y$ satisfy the normalization condition
1414: \begin{equation}
1415: \sum _{i=1}^{\infty} x^{\downarrow} _i
1416: = \sum_{i=1}^{\infty} y^{\downarrow} _i\;,
1417: \end{equation}
1418: then we write $x \prec y$ and say $x$ is majorized by $y$.
1419: \end{Definition}
1420: The sub-majorization condition does not require the normalization
1421: condition, but it is easily proven that $\mathbf{x} \prec
1422: _{\omega} \mathbf{y} $ is equivalent to
1423: \begin{eqnarray}
1424: \sum _{j=1}^{\infty} (\mathbf{x_j^{\downarrow} } -t)^+ \le
1425: \sum _{j=1}^{\infty} (\mathbf{y_j^{\downarrow} } -t)^+\;,
1426: \end{eqnarray}
1427: for all real $t$, where $z^+ = \max (z,0) $ is the positive part of
1428: any real number.
1429:
1430: Uhlmann's theorem relates operations on quantum states and
1431: majorization conditions. This theorem is one of the essential items
1432: for proving Nielsen's theorem for LOCC convertibility. To prove
1433: Uhlmann's theorem in infinite-dimensional systems, we define a
1434: doubly stochastic matrix in infinite-dimensional systems. It is
1435: similar to the one for finite-dimensional systems. For all double
1436: sequences $\{ d_{ij} \}_{ij = 1}^{\infty}$ which satisfy
1437: $\sum_{i=1}^{\infty} d_{ij} =1$ for all $j$, and $\sum
1438: _{j=1}^{\infty} d_{ij} = 1$ for all $i$, we can define a bounded
1439: linear operator $D \in \B (l_1)$ by
1440: \begin{eqnarray}
1441: D \mathbf{x} = \{ \sum_{j=1}^{\infty} d_{ij}
1442: \mathbf{x_j} \} _{i=1}^{\infty}\;,
1443: \end{eqnarray}
1444: for all $\mathbf{x} \in l_1$. These operators are called doubly
1445: stochastic matrices on $l_1$. We can easily see that the operator
1446: norm of a doubly stochastic matrix is $1$.
1447:
1448: The defined doubly stochastic matrices are related to majorization
1449: as follows: If $D$ is doubly stochastic, then, for all $\mathbf{x} \in
1450: S$, we have
1451: \begin{eqnarray} \label{ds}
1452: D \mathbf{x} \prec \mathbf{x}\;,
1453: \end{eqnarray}
1454: since
1455: \begin{eqnarray}
1456: \sum_{i=1}^{\infty} [(Dx)_i - t ]^+ &=&
1457: \sum_{i=1}^{\infty}
1458: \Bigl[ \sum_{j=1}^{\infty} D_{ij} ( x_j - t ) \Bigr]^+ \nonumber\\
1459: &\le& \sum_{i=1}^{\infty}
1460: \sum_{j=1}^{\infty} D_{ij} ( x_j - t )^+ \nonumber\\
1461: &=& \sum_{j=1}^{\infty} (x_j - t )^+\;,
1462: \end{eqnarray}
1463: for any real $t$, due to the convexity of $( )^+$. On the other hand,
1464: $\sum_{i=1}^{\infty} (Dx)_i = \sum_{i=1}^{\infty} x_i$ is trivial.
1465: Thus, the necessary condition (majorization condition) for the doubly
1466: stochastic condition is proven. We note that Eq.(\ref{ds}) is
1467: also valid for weaker conditions of $D$ than the doubly stochastic
1468: matrix, for example, $D$ satisfying $\sum _{j=1}^{\infty} d_{ij}
1469: \le 1$.
1470:
1471: Now we are ready to extend Uhlmann's theorem for
1472: infinite-dimensional systems:
1473: \begin{Theorem}[Uhlmann] \label{Uhlmann}
1474: If the two density operators $\rho _1$ and $\rho _2$ on the infinite
1475: separable Hilbert space $\Hi$ satisfy the following relation,
1476: \begin{eqnarray}
1477: \rho _1 = \sum _{j=1}^{\infty} p_j U_j \rho _2 U_j^{\dagger}, \label{Uhlmann eq}\\
1478: \qquad \sum _{j=1}^{\infty} p_j = 1 ,
1479: \end{eqnarray}
1480: where $U_j$ are partial isometries whose initial space includes
1481: closure of the range of $\rho _2$, ${\rm ker U_j}^{\perp} \supset
1482: \overline{\rm Ran \rho _2} $, then the non-zero eigenvalue of $\rho
1483: _1$ is majorized by $\rho _2$.
1484: \end{Theorem}
1485:
1486: \begin{Proof}
1487: Suppose $\rho _1 = \sum _{k=1}^{\infty} \mu _k \ket{e_k}
1488: \bra{e_k}$ and $\rho _2 = \sum _{i=1}^{\infty} \lambda _i \ket{i}
1489: \bra{i}$. Define a partial isometry as $V = \sum _{i=1}^{\infty}
1490: \ket{i} \bra{e_i}$, whose initial space is given by $\overline{\rm
1491: Ran \rho _1}$, and whose final space is given by $\overline{\rm Ran
1492: \rho _2}$. Then $U_j V$ is a partial isometry whose initial and
1493: final space are given by $\overline{\rm Ran \rho _1}$ and $U_j(
1494: \overline{\rm Ran \rho _2} )$, respectively. Actually, it is
1495: trivial that $U_j V$ is a zero operator on ${ \overline{\rm Ran
1496: \rho} }^{\perp}$. Now suppose $\ket{\Phi} \in \overline{\rm Ran
1497: \rho _1}$, then we obtain $\| U_j V \ket{\Phi} \| = \| \ket{\Phi}
1498: \|$ from the condition $V \ket{\Phi} \in \overline{\rm Ran \rho
1499: _2}$.
1500:
1501: Next, define Fourier's coefficients of $U_j \ket{i}$ as $u
1502: _{ji}^h$, that is, $U_j \ket{i} = U_j V \ket{e_i} =\sum
1503: _{h=1}^{\infty} u _{ji}^h \ket{e_h}$, and
1504: \begin{equation} \label{normalization of u}
1505: \sum _{h=1}^{\infty} | u _{ji}^h | ^2 =1.
1506: \end{equation}
1507: Then, we can rewrite Eq.(\ref{Uhlmann eq}) as,
1508: \begin{equation}
1509: \sum _{k=1}^{\infty} \mu _k \ket{e_k} \bra{e_k} = \sum _{ij \in
1510: N^2} p_j \lambda _i (\sum _{h=1}^{\infty} u _{ji}^h \ket{e_h} )
1511: (\sum _{l=1}^{\infty} {u^{\ast}} _{ji}^l \bra{e_l} )\;.
1512: \end{equation}
1513: Taking the inner product between $\ket{e_n}$,
1514: \begin{equation}
1515: \mu _n = \sum _{ij \in \mathbb{N}^2} p_j \lambda _i | u _{ji}^n | ^2
1516: = \sum _{i=1}^{\infty} \lambda _i \sum _{j=1}^{\infty} p_j
1517: | u_{ji}^n | ^2\;.
1518: \end{equation}
1519: We define $D _{ni}$ as $D _{ni} = \sum _{j=1}^{\infty} p_j |
1520: u_{ji}^n | ^2$. (\ref{normalization of u}) guarantees that $\sum
1521: _{n=1}^{\infty} D _{ni} = 1$. Since
1522: \begin{eqnarray}
1523: \sum _{i=1}^{\infty} | u_{ji}^n |^2 &=& \sum _{i=1}^{\infty}
1524: | \bra{e_n} U_j V \ket{e_i} |^2 \le \| U_j V \ket{e_n} \| ^2 \nonumber\\
1525: &\le& \| U_j \|_{\rm op} ^2 \| V \|_{\rm op} ^2 \| \ket{e_n} \| ^2 =1\; ,
1526: \end{eqnarray}
1527: where $\| \cdot \|_{\rm op}$ is the operator norm and $\sum
1528: _{i=1}^{\infty} D _{ni} \le 1$.
1529: Therefore, using the
1530: necessary condition of the double stochastic matrices and the
1531: weaker condition of (\ref{ds}) for $D$, we derive $\mu = D \lambda
1532: \prec \lambda$. The theorem is proven.
1533: \end{Proof}
1534: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1535:
1536: \newpage
1537: \appendix{ \quad Proof of Theorem \ref{epsilon Nielsen} (Nielsen's theorem
1538: in infinite-dimensional systems)} \label{Proof of Nielsen}
1539:
1540: In this appendix, based on Appendix A and B, we prove Nielsen's
1541: \cite{majorization} theorem for infinite dimensional systems.
1542:
1543: Before we show the proof of Nielsen's theorem of
1544: $\epsilon$-convertibility in infinite-dimensional systems, we
1545: first show that the necessary part of the original proof
1546: of Nielsen's theorem \cite{majorization} can be directly extended
1547: to infinite-dimensional systems by means of the Lo-Popescu and
1548: Uhlmann theorems we have already proven in the last section.
1549: \begin{Lemma} \label{necessity Nielsen}
1550: The necessary condition for convertibility of an
1551: infinite-dimensional state $\ket{\Psi}$ to another state
1552: $\ket{\Phi}$ under LOCC operations is given by $\lambda \prec
1553: \mu$, where $\lambda = \{ \lambda \} _{i=1}^{\infty}$ and $\mu =
1554: \{ \mu \} _{i=1}^{\infty} $ are the sequences of Schmidt
1555: coefficients of the states $\ket{\Psi}$ and $\ket{\Phi}$,
1556: respectively.
1557: \end{Lemma}
1558: \begin{Proof}
1559: Suppose $\ket{\Psi}$ can be converted to $\ket{\Phi}$ by LOCC,
1560: then By Lo-Popescu's theorem, $\rho _{\Phi} = p_m M_m \rho _{\Psi}
1561: M _m^{\dagger}$ where $\sum _{m=1}^{\infty } p_m = 1$, $\rho
1562: _{\Psi} = \Tr\; _B ( \ket{\Psi} \bra{\Psi} )$ and $\rho _{\Phi}
1563: =\Tr\; _B ( \ket{\Phi} \bra{\Phi} )$. Then, according to the same
1564: method of Nielsen's original proof, we derive
1565: $\rho _{\Psi} = \sum
1566: _{m=1}^{\infty } p_m U_m \rho _{\Phi } U_m^{\dagger }$, where
1567: $U_m$ is a partial isometry originating in the polar decomposition
1568: of $M_m \sqrt{\rho _{\Psi }}$. Since $\ker U_m ^{\perp } = \ker
1569: M_m \sqrt{\rho _{\Psi }}^{\perp } \supset \overline{\rm Ran \rho
1570: _{\psi } }$ \cite{functional analysis}, by Uhlmann's theorem we
1571: get $\lambda \prec \mu$. \hfill $\square$
1572: \end{Proof}
1573:
1574: By means of this lemma, we can prove the necessary part
1575: of Nielsen's theorem in infinite-dimensional systems. For the
1576: sufficient condition, we fully use the properties of
1577: $\epsilon$-convertibility.
1578: \begin{Proof}{(Theorem \ref{epsilon Nielsen})}
1579:
1580: Only if part: If $\ket{\Psi}$ is $\epsilon$-convertible to
1581: $\ket{\Phi}$ for any $\epsilon > 0$, there exists a sequence of
1582: states $\{ \ket{\Phi' _n} \} _{n=1}^{\infty}$ which strongly
1583: converges to $\ket{\Phi}$ (for pure states the topology of the trace norm
1584: is stronger than the strong topology of Hilbert space). Then, from
1585: Lemma \ref{necessity Nielsen}, $\lambda \prec \mu ^{'} _n$ where
1586: $\mu ^{'} _n$ and $\lambda$ are the Schmidt coefficients of
1587: $\ket{\Phi' _n}$ and $\ket{\Psi}$ for all $n \in \mathbb{N}$.
1588: Because Schmidt coefficients are continuous in the strong topology,
1589: $\sum _{i=1}^{n} \mu ^{'} _{n,i} \ge \sum _{i=1}^{n} \lambda _i$
1590: means $\sum _{i=1}^{n} \mu _i \ge \sum _{i=1}^{n} \lambda _i$
1591: where $\mu _i$ are the Schmidt coefficients of $\ket{\Phi}$.
1592:
1593:
1594: If part: When the Schmidt ranks (the number of non-zero Schmidt
1595: coefficients) of both $\ket{\Psi}$ and $\ket{\Phi}$ are finite,
1596: the proof is identical to the one for finite-dimensional systems.
1597: By the definition of Schmidt decomposition, we can assume
1598: $\ket{\Psi}$ and $\ket{\Phi}$ have the same Schmidt basis without
1599: loss of generality. In what follows, we divide the proof for
1600: the remaining cases into two parts: the case
1601: where $\ket{\Psi}$ has finite Schmidt rank, and
1602: the case where both of the states have infinite Schmidt ranks.
1603:
1604: 1) The case that $\ket{\Psi}$ has infinite Schmidt rank and
1605: $\ket{\Phi}$ has finite Schmidt rank:\\
1606: Suppose the Schmidt rank of $\ket{\Phi}$ is given by $N$. In what
1607: follows, we assume $\epsilon$ is arbitrary, but satisfies
1608: $\epsilon < \mu _N$. Since for any Schmidt coefficient $\{ \lambda
1609: _i \}_{i=1}^{\infty}$, we have $\lim _{n \rightarrow \infty} n
1610: \lambda _n = 0$. Therefore, there exists an $N_1 (\epsilon ) $ such
1611: that $n \lambda _n < \epsilon /2$ for any $ n \ge N_1$. On the
1612: other hand, since $\sum _{i=1}^{\infty} \lambda _i =1$ , there
1613: exists an $N_2 (\epsilon )$ such that $\sum _{i=n}^{\infty} \lambda
1614: _i < \epsilon /2$ for any $ n \ge N_1$. Suppose $M = \max (N_1,
1615: N_2, N)$, then we define $\{ {\mu '}_i \}$ as follows:
1616: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1617: \begin{tabbing}
1618: \qquad \= For \= $1 \le i \le N-1 $ \quad \= : ${\mu '}_i = \mu _i$ \\
1619: \> \> $i=N$ \> : ${\mu '}_N = \mu _N - \left ((M-N)
1620: \lambda _M + \sum _{n=M+1}^{\infty} \lambda _n \right)$ \\
1621: \> \> $N+1 \le i \le M$ \> : ${\mu '}_i = \lambda _M$ \\
1622: \> \> $M+1 \le i $ \> : ${\mu ' }_i = \lambda _i$
1623: \end{tabbing}
1624: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1625: % \begin{list}{}{}
1626: % \item \ For $1 \le i \le N-1 $ \;: ${\mu '}_i = \mu _i$
1627: % \item \qquad $i=N$ \quad \ \qquad : ${\mu '}_N = \mu _N -((M-N)
1628: % \lambda _M + \sum _{n=M+1}^{\infty} \lambda _n )$
1629: % \item \qquad $N+1 \le i \le M$ \ : ${\mu '}_i = \lambda _M$
1630: % \item \qquad $M+1 \le i $ \qquad \quad \!\!\!\! : ${\mu ' }_i
1631: %= \lambda _i$\;.
1632: % \end{list}
1633: We define $\ket{\Phi '}$ as $\ket{\Phi '} = \sum _{i=1}^{\infty}
1634: \sqrt{{\mu '}_i} \ket{i} \otimes \ket{i}$. Then, by definition,
1635: $\lambda \prec \mu^{'} \prec \mu $. Moreover, we
1636: obtain
1637: \begin{eqnarray}
1638: \| \ket{\Phi} - \ket{\Phi '} \| ^2 &=& |(M-N)\lambda _M + \sum
1639: _{n=M+1}^{\infty} n \lambda _n | ^2 \nonumber \\
1640: &=& |(M-N)\lambda _M|^2
1641: + |\sum _{i=M+1}^{\infty} \lambda_i|^2 \nonumber \\
1642: &\le & \epsilon ^2\;.
1643: \end{eqnarray}
1644: Therefore, for any neighborhood of $\ket{\Phi}$, we can find
1645: a $\ket{\Phi'}$ such that $\ket{\Psi} \rightarrow \ket{\Phi'}$.
1646:
1647: 2) The case that the Schmidt ranks of $\ket{\Psi}$
1648: and $\ket{\Phi}$ are infinity:\\
1649: By easy calculation we can show that for any $\epsilon$, there
1650: exists a natural number $N_1 (\epsilon )$ such that if $\ket{\Phi
1651: '} = \sum _{i=1}^{\infty} \sqrt{{\mu '} _i} \ket{i} \otimes
1652: \ket{i}$ satisfies $\mu ' _i= \mu$, $i \le N_1 (\epsilon ) $,
1653: then $\| \ket{\Phi}\bra{\Phi} - \ket{\Phi '}\bra{\Phi'} \|_{\rm
1654: tr} < \epsilon $. Since $\sum _{i=1}^n \lambda _i < 1$ and $\lim
1655: _{n \rightarrow \infty} n \lambda _n = 0$ for any $n \in
1656: \mathbb{N}$,
1657: \begin{equation}
1658: \lim _{N_2 \rightarrow \infty} \Bigl[\sum _{k=1}^{N_2 (\epsilon )} \lambda _k
1659: - (N_2 - N_1 )\lambda _{N_2} \Bigr] = 1\;.
1660: \end{equation}
1661: Thus, there exists a natural number $N_2 (\epsilon ) \ge N_1
1662: (\epsilon ) + 1 $ such that
1663: \begin{equation} \label{eq N_2}
1664: \sum _{k=1}^{N_2 (\epsilon )} \lambda _k
1665: - (N_2 - N _1 ) \lambda _{\lambda _{N_2}}
1666: \ge \sum _{i=1}^{N_1 (\epsilon ) } \mu _i\;,
1667: \end{equation}
1668: and
1669: \begin{equation} \label{epsilon 1}
1670: \sum _{k=1}^{N_2 (\epsilon ) - 1} \lambda _k
1671: - (N_2 - N _1 -1) \lambda _{N_2 - 1}
1672: \le \sum _{i=1}^{N_1 (\epsilon ) } \mu _i\;.
1673: \end{equation}
1674: We examine in the two cases $N_2(\epsilon ) = N_1(\epsilon ) +1$
1675: and $N_2 (\epsilon ) > N_1(\epsilon ) + 1$ separately.
1676:
1677: a) The case that $N_2(\epsilon ) = N_1(\epsilon ) +1$: The
1678: inequalities (\ref{eq N_2}) and (\ref{epsilon 1}) guarantee $\sum
1679: _{k=1}^{N_1 (\epsilon )} \lambda _k = \sum _{k=1}^{N_1(\epsilon )}
1680: \mu _k$. If we define $1 \le k \le N_1(\epsilon )$, ${\mu '}_k =
1681: \mu _k$, $k \ge N_1(\epsilon ) + 1$ and ${\mu '}_k = \lambda _k$,
1682: $\{ {\mu '}_i \}_{i=1}^{\infty}$ satisfies $\sum _{i=1}^k \lambda
1683: _i \le \sum _{i=1}^k {\mu '}_i \le \sum _{i=1}^k \mu _i$ for all
1684: $k \in \mathbb{N}$.
1685:
1686: b) The case that $N_2 (\epsilon ) > N_1(\epsilon ) + 1$: We define
1687: $\delta$ as
1688: \begin{eqnarray}
1689: \delta = \Big[ \sum _{i=1}^{N_1 (\epsilon )} \lambda _i
1690: + \sum _{k=N_1 (\epsilon) +1}^{N_2 -1} (\lambda _k - \lambda _{N_2 })
1691: -\sum
1692: _{i=1}^{N_1 (\epsilon ) } \mu _i \Big] /( N_2(\epsilon ) -
1693: N_1(\epsilon ) -1) \:.
1694: \end{eqnarray}
1695: Then, since $\delta \ge 0$, we can define ${\mu '}_k$ as the
1696: following,
1697: \begin{tabbing}
1698: \qquad \= ${\mu '}_k = \mu _k$ \qquad \qquad \= for $1 \le k \le N_1 (\epsilon )$, \\
1699: \> ${\mu '}_k = \lambda _{N_2(\epsilon )} + \delta$ \> for $N_1(\epsilon )+1 \le k \le N_2(\epsilon ) -1$, \\
1700: \> ${\mu '}_k = \lambda _k$ \> for $N_2(\epsilon ) \le k$
1701: \end{tabbing}
1702: %\begin{list}{}{}
1703: % \item ${\mu '}_k = \mu _k$ for $1 \le k \le N_1 (\epsilon )$\;,
1704: % \item ${\mu '}_k = \lambda _{N_2(\epsilon )} + \delta$
1705: % for $N_1(\epsilon )+1 \le k \le N_2(\epsilon ) -1 $ \;,
1706: % \item ${\mu '}_k = \lambda _k$ for $N_2(\epsilon ) \le k$ \;,
1707: %\end{list}
1708: then we have
1709: \begin{eqnarray}
1710: \sum _{i=1}^{\infty} {\mu '}_i &=&
1711: \sum _{i=1}^{N_1 (\epsilon )}
1712: \mu _i + \sum _{k=N_1 (\epsilon )+1}^{N_2 (\epsilon )-1}
1713: (\lambda_{N_2(\epsilon )} + \delta )
1714: +\sum _{k=N_2(\epsilon )}^{\infty}
1715: \lambda _k \nonumber\\
1716: &=&1\;.
1717: \end{eqnarray}
1718: Therefore, the $\{ \mu '_i \}_{i=1}^{\infty}$ are well defined Schmidt
1719: coefficients.
1720:
1721: First, we show $\sum _{i=1}^{N} \mu '_i \le \sum _{i=1}^{N} \mu
1722: _i$ for all $N $. Since this condition is trivial for $N \le N_1$
1723: and $N_2 \le N$ by definition of $\{ \mu' _i \}_{i=0}^{\infty}$,
1724: we only need to check this condition for $ N_1 + 1 \le N \le N_2 -
1725: 1$. Suppose there exists an $N'$ such that $N_1 +1 \le N' \le N_2 -1$
1726: and $\sum _{i= N_1 +1}^{N'} \mu'_i > \sum _{i = N_1 +1}^{N'}\mu
1727: _i$. Then, since $\mu'_i = \lambda _{N_2} + \delta$ for all $N_1
1728: +1 \le i \le N_2 - 1$ and $\mu_i \ge \mu_{i+1} $, we can easily
1729: see $\mu' _{N'}
1730: > \mu _{N'}$. That is, $\sum _{i= N_1 +1}^{N} \mu'_i > \sum _{i = N_1 +1}^{N}\mu
1731: _i$ for all $N' \le N \le N _2 -1$, and we can conclude $\sum _{i
1732: = N_1 +1}^{N_2 -1} \mu'_i > \sum _{i=N_1}^{N_2 -1} \mu _i$ which
1733: is a contradiction. Therefore, for all $N_1 + 1 \le N \le N_2 -1 $,
1734: $\sum _{i = N_1 +1}^{N} \mu'_i \le \sum _{i=N_1}^{N} \mu _i$.
1735:
1736: Second, we show $\sum _{i=1}^N \lambda _i \le \sum _{i=1}^N \mu'
1737: _i$ for all $N$. Since this condition is trivial for $N \le N_1$
1738: and $N_2 \le N$, we only check for $N_1 +1 \le N \le N_2 -1$. In
1739: this case, we calculate
1740: \begin{eqnarray}
1741: \sum _{k=1}^N {\mu '}_k - \sum _{k=1}^N \lambda _k &=& \sum
1742: _{k=1}^{N_1} \mu _k + \sum _{k=N_1+1}^N (\lambda _{N_2} + \delta)
1743: - \sum _{k=1}^N \lambda _k \nonumber \\ &=& \frac{N_2 - N -
1744: 1}{N_2 - N_1 - 1} [(\sum _{k=1}^{N_1} \mu _k - \sum _{k=1}^{N_1}
1745: \lambda _k ) - \sum _{k=N_1 +1}^N \lambda _k ] \nonumber \\
1746: &\ & + \frac{N-N_1}{N_2 - N_1 -1} \sum _{k=N+1}^{N_2-1} \lambda
1747: _k\;. \label{epsilon 3}
1748: \end{eqnarray}
1749: From Eq.~(\ref{epsilon 1}), we obtain
1750: \begin{equation}
1751: \sum _{k=1}^{N_1} \lambda _k + \sum _{k=N_1 + 1}^{N_2 -2} (\lambda _k
1752: - \lambda _{N_2} ) \le
1753: \sum _{k=1}^{N_1} \mu _k\;.
1754: \end{equation}
1755: % \begin{widetext}
1756: Thus, for the Eq.(\ref{epsilon 3}) may be bounded by
1757: \begin{eqnarray*}
1758: & \quad & \sum _{k=1}^N \mu' _k - \sum _{k=1}^N \lambda _k \\
1759: &\ge &\frac{N_2 - N -1 }{N_2 - N_1 -1} \left[ \sum _{k= N_1
1760: +1}^{N_2 -2} (\lambda _k - \lambda _{N_2 -1}) - \sum _{k= N_1 +
1761: 1}^{N} \lambda _k \right] + \frac{N - N_1}{N_2 - N_1 -1} \sum _{k=
1762: N+1}^{N_2 -1}
1763: \lambda _k. \\
1764: &=& \sum _{k= N+1}^{N_2 -2} \lambda _k - (N_2 -N -2 )\lambda
1765: _{N_2 -1} \\
1766: &\ge& 0.
1767: \end{eqnarray*}
1768: Thus, for $N_1 +1 \le N \le N_2 -1$, we have $\sum _{k=1}^N \lambda _k
1769: \le
1770: \sum _{k=1}^N {\mu '}_k$.
1771: % \end{widetext}
1772:
1773: Finally, for any natural number $N$, we obtain
1774: \begin{equation}\label{}
1775: \sum _{k=1}^N \lambda _k \le \sum _{k=1}^N {\mu '}_k \le \sum
1776: _{k=1}^N \mu _k\;.
1777: \end{equation}
1778: We define $\ket{\Phi'}$ by using Schmidt coefficients $\{ \mu' _i
1779: \}_{i=1}^{\infty}$. Then, since $\mu'_k = \lambda _k$ for all $k
1780: \ge N_2(\epsilon )$, we can convert $\ket{\Psi } $ to
1781: $\ket{\Phi'}$ by means of the LOCC operation which is derived by
1782: the original (finite dimensional) Nielsen's
1783: theorem \cite{majorization}. Moreover, since $\mu' _k = \mu _k$ for
1784: all $k \le N_1(\epsilon )$, $\ket{\Phi'}$ satisfies $\|
1785: \ket{\Phi'}\bra{\Phi'} - \ket{\Phi}\bra{\Phi} \|_{\rm tr} <
1786: \epsilon$. Therefore, for any neighborhood of $\ket{\Phi}$, we can
1787: find a $\ket{\Phi'}$ such that $\ket{\Psi} \rightarrow \ket{\Phi'}$,
1788: where $\ket{\Phi'}$ is defined by using Schmidt coefficients $\{
1789: \mu' _i \}_{i=1}^{\infty}$.
1790:
1791:
1792: \hfill $\square$
1793: \end{Proof}
1794:
1795: \appendix{ \quad Proof of Theorem \ref{epsilon Vidal}} \label{Proof of Vidal}
1796: In this appendix, based on Appendix A, B and C, we prove Vidal's
1797: \cite{vidal} theorem infinite dimensional systems.
1798:
1799:
1800: \begin{Proof}
1801: If part: The proof of Vidal's theorem in \cite{D'ariano} is most
1802: suitable to extend this part to infinite-dimensional systems.
1803: Suppose $\lambda \prec ^{\omega} p \mu$, then $\{ \nu _i \}
1804: _{i=1}^{\infty}$ defined by the condition $\nu _1 = 1-p(1-\mu_1)$
1805: and $\nu _i = p \mu _i$ for $i \neq 1$ satisfies the conditions
1806: $\lambda \prec \nu$ and $\nu \le \mu$. If we define the state
1807: $\ket{\Omega}$ as the state whose Schmidt coefficients are $\{
1808: \nu _i \}$ and whose Schmidt basis is same as $\ket{\Phi}$'s,
1809: then, by Nielsen's theorem in infinite-dimensional systems, for
1810: any small $\epsilon >0$, $\ket{\Psi}$ can be transformed to
1811: $\ket{\Omega'}$ by LOCC with certainty, where $\|
1812: \ket{\Omega'}\bra{\Omega'} - \ket{\Omega}\bra{\Omega} \| \le
1813: \epsilon $. Then, $\ket{\Omega'}$ can also be transformed to
1814: $\ket{\Phi'}$ by the local measurement $E = \sum _{i=1}^{\infty}
1815: \sqrt{\frac{\nu _i}{\mu _i}} \ket{i} \bra{i}$ with probability
1816: $p$, where $\| \ket{\Phi'}\bra{\Phi'} - \ket{\Phi}\bra{\Phi} \|
1817: _{\rm tr} \le \epsilon $.
1818:
1819: Only If part: At first, since the first half of theorem 2 of
1820: \cite{vidal} is directly extended to infinite-dimensional systems,
1821: Vidal's monotone $E_n(\ket{\Psi}) = \sum _{i=n}^{\infty} \lambda
1822: _i$ is a monotonic function of LOCC in mean value. In other words,
1823: if $\ket{\Psi}$ can be transformed to $\ket{\Phi _i}$ with
1824: probability $p_i$ in LOCC, then $E_n(\ket{\Psi}) \ge \sum
1825: _{i=1}^{\infty} p_i E_n(\ket{\Phi _i})$.
1826:
1827: If for a set of $\ket{\Psi}$ and $\ket{\Phi}$, and any $\epsilon
1828: _1 > 0$, there exists a $\ket{\Phi'}$ such that $\|
1829: \ket{\Phi'} \bra{\Phi'} - \ket{\Phi} \bra{\Phi} \|_{\rm tr} <
1830: \epsilon _1$ and $\ket{\Psi}$ can be converted to $\ket{\Phi'}$ by
1831: some SLOCC with probability $p'$ which satisfies $p' \ge p$ and
1832: $\lambda \prec ^{\omega} p\mu$ is not true, that is for some $k_1$
1833: $E_{k_1} (\ket{\Psi}) < p E_{k_1} (\ket{\Phi})$. Then, there exists
1834: a sequence $\ket{\Phi _n}$ and $p_n$ which satisfies $\lim _{n
1835: \rightarrow \infty} \ket{\Phi _n } = \ket{\Phi}$ and $p_n \ge p$ for all $n
1836: \in \mathbb{N}$. Moreover, there also exists a sequence of SLOCC
1837: operations which transforms $\ket{\Psi}$ to $\ket{\Phi _n}$. Then,
1838: by monotonicity of $E_k (\ket{\Psi})$ we have $E_k (\ket{\Psi})
1839: \ge p_n E_k (\ket{\Phi _n})$ for all $k \in \mathbb{N}$. Since $1
1840: - E_n (\ket{\Psi})$ is finite sum of eigenvalues of the reduced
1841: density operator, $E_n(\ket{\Psi})$ is continuous in the whole
1842: Hilbert space and for all $n$. Thus, by taking the limit of the
1843: inequality, we have $E_k (\ket{\Psi}) \ge p E_k (\ket{\Phi})$ for
1844: all $k \in \mathbb{N}$. This is a contradiction. \hfill $\square$
1845: \end{Proof}
1846:
1847:
1848: \end{document}
1849: