1: \documentclass[prl,showpacs,twocolumn,superscriptaddress]{revtex4}
2: %\documentclass{article}
3: \usepackage{amsmath,amssymb,amsfonts}
4: \usepackage{amsthm}
5: \usepackage{graphicx}
6: \usepackage{pmat}
7:
8: %Useful quantum definitions
9: \newcommand{\bra}[1]{\ensuremath{\langle #1 |}}
10: \newcommand{\ket}[1]{\ensuremath{| #1 \rangle}}
11: \newcommand{\bigket}[1]{\left|#1\right\rangle}
12: \newcommand{\bigbra}[1]{\left\langle #1\right|}
13: \newcommand{\bk}[2]{\langle #1|#2\rangle}
14: \newcommand{\kb}[1]{| #1 \rangle\langle #1|}
15: \newcommand{\kbb}[2]{| #1 \rangle\langle #2|}
16:
17: %Math operators
18: \DeclareMathOperator{\tr}{tr}
19: \DeclareMathOperator{\sgn}{sgn}
20: \DeclareMathOperator{\arcsinh}{arcsinh}
21: \DeclareMathOperator{\arccosh}{arccosh}
22: \DeclareMathOperator{\arctanh}{arctanh}
23:
24: \def\one{\leavevmode\hbox{\small1\normalsize\kern-.33em1}}
25: \newcommand{\CC}{\mathbb{C}}
26: \newcommand{\RR}{\mathbb{R}}
27: \newcommand{\ZZ}{\mathbb{Z}}
28:
29:
30: \newcommand{\BB}[1]{\mathcal{B}_\mathrm{#1}}
31: \newcommand{\YY}[2]{\left\langle {{Y}}_\mathrm{#1} {{Y}}_\mathrm{#2} \right\rangle}
32:
33: \newcommand{\AB}{\langle \mathcal{B}_{\text{AB}}\rangle}
34: \newcommand{\AC}{\langle \mathcal{B}_{\text{AC}}\rangle}
35: \newcommand{\BC}{\langle \mathcal{B}_{\text{BC}}\rangle}
36:
37: \def\PP{{P}}
38:
39: \def \qedman {\hfill \rule{0.2cm}{0.2cm}\vspace{3mm}}
40: \def \qedme {\rule{0.2cm}{0.2cm}}
41: \def \fixme {}
42:
43: %Theorems
44: \newtheorem{theorem}{Theorem}
45: \newtheorem{lemma}[theorem]{Lemma}
46: \newtheorem{cor}[theorem]{Corollary}
47: \newtheorem{prop}[theorem]{Proposition}
48: \newtheorem{protocol}[theorem]{Protocol}
49:
50: \theoremstyle{definition}
51: \newtheorem{definition}[theorem]{Definition}
52:
53: \theoremstyle{remark}
54:
55: %comments
56: \newcommand{\comment}[1]{\begin{quote}\sf [*** #1 ***]\end{quote}}
57:
58: \begin{document}
59: \title{Monogamy of Bell correlations and \uppercase{T}sirelson's bound}
60: \author{Benjamin F. Toner}
61: \email{Ben.Toner@cwi.nl}
62: \affiliation{Institute for Quantum Information, California Institute of
63: Technology, Pasadena, CA 91125, USA }
64: \affiliation{Centrum voor Wiskunde en Informatica, Kruislaan 413, 1098 SJ Amsterdam, The Netherlands}
65: \date{\today}
66: \author{Frank Verstraete}
67: \email{frank.verstraete@univie.ac.at}
68: \affiliation{Institute for Quantum Information, California
69: Institute of Technology, Pasadena, CA 91125, USA } \affiliation{Fakult\"at f\"ur Physik, Universit\"at Wien,
70: Austria}
71: \date{\today}
72:
73: %%%%%%%%%%%% Abstract %%%%%%%%%%%%%%%%%%%%%%%%%%%
74:
75: \begin{abstract}
76: We consider three parties, A, B, and C, each performing one of two
77: local measurements on a shared quantum state of arbitrary dimension.
78: We characterize the trade-off between the nonlocality of the Bell
79: correlations observed by AB and of those observed by AC. This
80: generalizes Tsirelson's bound on the quantum value of the CHSH
81: inequality, the latter being recovered when C is completely
82: uncorrelated with AB. We also discuss the trade-off between Bell
83: violations and local expectation values of observables that
84: anticommute with the ones used in the Bell test.
85: \end{abstract}
86:
87: \pacs{03.65.Ud, 03.65.Ta, 03.67.-a}
88:
89: \maketitle
90:
91: The existence of Bell inequalities~\cite{Bell:64a,Clauser:69a} and
92: their observed violation in experiments has had a very deep impact on
93: the way we look at quantum mechanics. On the one hand, it has led to a
94: study of the precise meaning of nonlocality and opened up the field of
95: entanglement theory. On the other, it has led to the observation that
96: Bell violations can be exploited in the design of cryptographic
97: protocols~\cite{PhysRevLett.67.661}.
98: In that case, an eavesdropper (C) tries to gain access to some quantum
99: correlations shared by Alice (A) and Bob (B). If the Bell correlations
100: between A and B are strong, it can happen that C's outcomes will be
101: almost uncorrelated with them, and A and B will be able to execute a
102: purification protocol so as to create private randomness. In the
103: present paper, we will make a precise quantitative statement about the
104: following monogamy property: Suppose A and B violate a Bell inequality
105: by a certain amount. How does that bound the possible Bell
106: correlations between A and C? This is also interesting from the point
107: of view of entanglement theory, as it provides monogamy relations
108: independent of the size of the local Hilbert spaces. For the
109: Clauser-Horne-Shimony-Holt (CHSH) inequality, the region of accessible
110: Bell correlations between AB and AC turns out to be very simple (see Figure~1).
111:
112: \begin{figure}[b] \centering
113: \includegraphics{fig1arxiv.eps}
114: \caption{Accessible values of $\AB$ and $\AC$ for classical theories (interior of square), quantum theory (interior of
115: circle), and no-signalling theories (interior of diamond). Note that both quantum and no-signalling theories obey
116: monogamy constraints; classical local hidden variable theories do not.} \label{fig:1}
117: \end{figure}
118:
119:
120:
121: In the setting where two parties, A and B, share a quantum state $\rho$, and each has the choice of two local
122: measurements, there is just one relevant Bell inequality, the CHSH
123: inequality~\cite{Clauser:69a}. Define the {CHSH operator}
124: \begin{equation}
125: \label{eq:1}
126: {\cal B}_\text{AB} = {A}_1 \otimes \left({B}_1 + {B}_2\right) +
127: {A}_2 \otimes \left({B}_1 - {B}_2\right),
128: \end{equation}
129: where ${A}_1$ and ${A}_2$ (${B}_1$ and ${B}_2$) are A's (B's) observables and are Hermitian operators with
130: spectrum in $[-1,+1]$. For particular measurements and a particular state $\rho$, the quantum value of the CHSH
131: inequality is defined as $\AB = \tr \left({\cal B}_\text{AB} \rho\right)$. All correlations described by local
132: hidden variable (LHV) models satisfy the CHSH inequality, $\left|\AB_\text{LHV} \right| \leq 2 $, but in the case
133: of entangled quantum systems, this bound can be violated. For example, on the singlet state of two qubits there exist
134: operators $A_i,B_i$ such that $\AB = \bra{\psi^-} {\cal B}_\text{AB}\ket{\psi^-} = 2\sqrt2 $.
135: %Bell inequality violation has been observed in many experiments~\cite{note:Bellexp2}, although all experiments performed to date
136: %are susceptible to one or more loopholes.
137:
138: We do not yet know how to calculate a bound on the maximum quantum value of an arbitrary Bell inequality, but a
139: number of ad hoc techniques have been
140: developed~\cite{Tsirelson:80a,Tsirelson:85b,toner:_monog,navascues:_bound}.
141: In the case of the CHSH inequality, Tsirelson has proved that $ \left|\tr \left( {\cal B}_\text{AB}\rho\right)
142: \right| \leq 2\sqrt2 $ for all observables ${A}_1$, ${A}_2$, ${B}_1$, ${B}_2$, and all states
143: $\rho$~\cite{Tsirelson:80a}. This Tsirelson bound can itself be violated if we consider more general hypothetical
144: {\em no-signalling} theories: a {\em nonlocal box} violates the CHSH inequality maximally, $\AB_\text{NL} =
145: 4$~\cite{Popescu:94a}. Tsirelson's bound is a simple mathematical consequence of the axioms of
146: quantum theory, but is there some deeper reason why a violation greater than $2 \sqrt 2$ is unphysical? For
147: example, a violation greater than $\sqrt{32/3}\approx 3.27$ would imply that any communication complexity problem
148: can be solved using a constant amount of communication~\cite{dam05:_implaus_conseq_super_nonloc}.
149: As will be clear from the following results, the bound $2\sqrt{2}$ is very natural if one considers the possible
150: Bell violations in a three-party setup.
151:
152: We establish the following monogamy trade-off:
153: \begin{theorem}
154: \label{thm:1}
155: Suppose that three parties, A, B, and C, share a quantum
156: state (of arbitrary dimension) and each chooses to measure one of two
157: observables. Then
158: \begin{equation}
159: \label{eq:2}
160: \AB^2 + \AC ^2 \leq 8.
161: \end{equation}
162: \end{theorem}
163: Here, $\mathcal{B}_{\text{AC}}$ is defined as in Eq.~(\ref{eq:1}), but with B's observables replaced by C's. Note
164: that we obtain Tsirelson's bound, $\AB^2 \leq 8$, as a simple corollary. Note also that A's measurements are the
165: same in $\AB$ and $\AC$: otherwise we could have $\AB = \AC = 2\sqrt 2$ and there would be no trade-off.
166: Theorem~\ref{thm:1} is analogous to the Coffman-Kundu-Wootters theorem that describes the trade-off between how
167: entangled A is with B, and how entangled A is with C~\cite{Coffman:00a}. Eq.~(\ref{eq:2}) is the best
168: possible bound: there are states and measurements achieving any values of $\AB$ and $\AC$ that satisfy it.
169: Previously the best bound known was
170: $
171: \left|\AB\right| + \left|\AC\right| \leq 4$,
172: which is tight for correlations that arise from no-signalling theories~\cite{toner:_monog,masanes05:_gener_nonsig_theor}.
173: We illustrate the monogamy trade-offs for various theories in Figure~\ref{fig:1}.
174:
175: We prove Theorem~\ref{thm:1} in two parts. We first show that is
176: sufficient to restrict to states with support on a qubit at each site.
177: We can then relax the requirement that A's measurements be the same in
178: $\AB$ and $\AC$, maximizing over the measurements in $\AB$ and $\AC$
179: separately, but keeping the state fixed. Our proof suggests a
180: connection between anticommutation and Bell inequality violation,
181: which we then explore more deeply.
182:
183: {\em Dimensional reduction}.---We start by establishing a bound on the dimension of the quantum state required to
184: maximally violate certain Bell inequalities. A similar result was proved by Masanes~\cite{masanes05:_extrem}. The
185: main ingredient---a canonical decomposition for a pair of subspaces of $\CC^n$---is described in more detail in,
186: e.g., Ref.~\cite{bhatia96:_matrix_analy}.
187:
188: \begin{lemma} \label{lemma:circle1}
189: Consider any Bell inequality in the setting where $m$
190: parties each choose from two two-outcome measurements. Then the
191: maximum quantum value of the Bell inequality is achieved by a state
192: that has support on a qubit at each site. Furthermore, we can assume
193: this state has real coefficients and that the observables are real
194: and traceless.
195: \end{lemma}
196:
197: \begin{proof}
198: For $i \in \{1,2\}$, assume party $k$ has observables ${M}_{k,i}$,
199: acting on a Hilbert space ${\cal H}_k$. By
200: extending the local Hilbert spaces ${\cal H}_k$, we can assume for all
201: $k$ and for all $i = 1,2$ that (i) ${\cal H}_k = \CC^{2d}$ for some
202: fixed $d$, (ii)
203: ${M}_{k,i}$ has eigenvalues $\pm 1$, and (iii) $\tr {M}_{k,i}
204: = 0$. The first condition states that all local spaces have the same
205: dimension $2d$, the latter two that each observable corresponds to
206: a projective measurement onto a $d$-dimensional subspace and its
207: complement. We also define ${ M}_{k,0} = \one_{2d}$, the identity operator on ${\cal H}_k$. We can
208: write a generic Bell operator in the setting stated in the lemma as
209: \begin{equation}
210: {\cal B} = \sum_{i_1=0}^2 \sum_{i_2=0}^2 \cdots \sum_{i_m=0}^2 c_{i_1i_2\cdots i_m}\bigotimes_{k=1}^m {M}_{k,i_k},
211: \label{circle-eq:3}
212: \end{equation}
213: where the coefficients $c_{i_1i_2\cdots i_m}$ are arbitrary real numbers.
214: Our goal is find the quantum value of this Bell operator, which is
215: maximum of $B \equiv \bra \psi \cal B \ket \psi$ over states $\ket \psi$ and
216: measurements ${M}_{k,i}$.
217:
218: We now choose a local basis for each ${\cal H}_k$ such that party $k$'s
219: observables have a simple form. We start by taking
220: ${M}_{k,1} = \begin{pmat}[{|}]\one_d & {0} \cr\- {0}
221: & -\one_d \cr
222: \end{pmat}$. This leaves us the freedom to specify the basis within
223: the two $d\times d$ blocks on which ${M}_{k,1}$ is constant. Let
224: ${M}_{k,2} = 2 \PP \PP^\dagger - \one_{2d}$ (we suppress the
225: dependence on $k$), where $\PP$ is a $2d \times d$ matrix with
226: orthonormal columns, which span the $+1$--eigenspace of
227: ${M}_{k,2}$). Write $\PP = \begin{pmat}[{}]\PP_1 \cr \- \PP_2
228: \cr\end{pmat}$, where $\PP_1$ and $\PP_2$ are $d \times d$
229: matrices. The rows of $P$ are orthonormal, which implies ${
230: P}^\dagger {P} = { P}_1^\dagger {P}_1 + { P}_2^\dagger {P}_2 =
231: \one_d$, so $ { P}_1^\dagger {P}_1$ and ${ P}_2^\dagger {P}_2$ are
232: simultaneously diagonalizable. This means there is a singular
233: value decomposition of the form $ \PP_1 = {U}_1^\dagger {D}_1{V}$,
234: $\PP_2 = {U}_2^\dagger {D}_2{V}$, where ${U}_1$, ${U}_2$ and $V$ are
235: $d\times d$ unitary matrices and ${D}_1$ and ${D}_2 = \sqrt{\one_d -
236: {D}_1^2}$ are nonnegative (real) diagonal matrices.
237: %% Next,
238: %% we take a QR-decomposition of ${P}_2 {V}$, giving ${P}_2 {V} =
239: %% Q R$, where $Q$
240: %% is a $d \times d$ unitary, and $R$ is an upper-triangular matrix
241: %% with nonnegative (real) entries on the diagonal. By permuting the
242: %% rows of $R$ (which can be done by left-multiplication by a unitary
243: %% matrix), we can ensure that if a diagonal entry of $R$ is zero, then
244: %% so are all other entries in that row. Assume this has been done.
245: %% We now have ${P} = {U} \oplus {Q} \begin{pmat}[{}]{\bf
246: %% D}\cr \- {R} \cr \end{pmat}
247: %% {V}^\dagger$. Since the rows of $P$ are orthonormal, ${\bf
248: %% P}^\dagger {P} =
249: %% \one_d$, which implies ${D}^2 + {R}^\dagger {R} = \one_d$. It follows
250: %% that ${R}^\dagger {R}$ is diagonal, and this in turn implies
251: %% that ${R}$
252: %% itself is diagonal (the proof is by induction on the rows of ${R}$).
253: %% Assume the
254: %% first $p-1$ rows of $R$ are consistent with $R$ being diagonal. If
255: %% $R_{pp} = 0$, then we are done, by the property of $R$ given above;
256: %% if $R_{pp}\neq 0$, it follows that $R_{pq} = 0$ for all $q\geq p$.)
257: %% Therefore ${R} = \sqrt{ \one_d - {D}^2}$.
258: Changing basis according to
259: the unitary ${U}_1 \oplus {U}_2$, which leaves ${M}_{k,1}$ invariant, it
260: follows that
261: %% ${M}_{k,2}
262: %% = \begin{pmat}[{|}]2{D}^2 - \one_d & 2 {D} \sqrt{\one_d -
263: %% {D}^2}\cr \- 2 {D}
264: %% \sqrt{\one_d - {D}^2}& \one_d - 2{D}^2 \cr \end{pmat}$,
265: ${M}_{k,2}
266: = \begin{pmat}[{|}]2{D}_1^2 - \one_d & 2 {D}_1 {D}_2\cr \-
267: 2 {D}_1 {D}_2& 2{D}_2^2 - \one_d \cr \end{pmat}$,
268: where each of
269: the $d \times d$ blocks is diagonal. We relabel
270: our basis vectors so that ${M}_{k,1} = \bigoplus_{j=1}^d {Z}$, $M_{k,2} =
271: \bigoplus_{j=1}^d \left(\cos \theta_j {Z} + \sin \theta_j
272: {X} \right)$, where $2{D}_1^2 - \one_d
273: = \mathrm{diag}(\cos \theta_1, \cos \theta_2, \ldots, \cos
274: \theta_d)$ and $X$ and $Z$ are the usual Pauli operators. Hence our operators are real and preserve a $\oplus_{j = 1}^d \CC^2$
275: subspace of ${\cal H}_k$. They are traceless on each $\CC^2$ space.
276:
277: We wish to maximize $B = \bra \psi \cal B \ket \psi$ over the state
278: $\ket \psi$ and the
279: measurements ${M}_{k,i}$. Fix $k$, and let $\rho_{k,j}$ be the reduced density
280: matrix obtained by projecting $\ket \psi$ onto the $j$'th $\CC^2$
281: factor of the $\oplus_{j = 1}^d \CC^2$ subspace induced by
282: $M_{k,1}$ and $M_{k,2}$ at site $k$.
283: %% Then $B$ is unchanged if we decohere
284: %% the state $\ket \psi$ onto the $\oplus_{j = 1}^d \CC^2$ subspace induced by
285: %% $M_{k,1}$ and $M_{k,2}$ at site $k$, whereupon
286: Then $B = \sum_{j=1}^d \tr
287: {\cal B} \rho_{k,j}$ is a convex sum over the $\CC^2$ factors, whereupon it follows that the maximum is
288: achieved by a state with support on a qubit at site $k$. Since this
289: argument works for all $k$, the maximum of $B$ is
290: achieved by a state that has support on a qubit on each site.
291:
292: Finally, write $\ket \psi = \ket {\psi_1} + i \ket {\psi_2}$, where
293: $\ket {\psi_1}$ and $\ket {\psi_2}$ are real. Then $\bra \psi {\cal
294: B} \ket \psi = \bra {\psi_1} {\cal B} \ket {\psi_1} + \bra {\psi_2}
295: {\cal B} \ket {\psi_2}$ since $\cal B$ is real, which is the same
296: expression we would obtain if the state were a real mixture of $\ket
297: {\psi_1}$ and $\ket {\psi_2}$. Hence the maximum of $B$ is achieved
298: by a state with real coefficients.
299: \end{proof}
300:
301: {\em Monogamy trade-off relation.}---The region $\cal R$ of allowed
302: values of $\left(\AB, \AC \right)$ is convex and can therefore be
303: described by an (infinite) family of half-space inequalities,
304: \begin{equation}
305: c_{\text{AB}}\AB + c_{\text{AC}}\AC \leq d,
306: \label{circle-eq:4}
307: \end{equation}
308: with $c_{\text{AB}}, c_{\text{AC}}, d \in \RR$. The left-hand side of
309: Eq.~(\ref{circle-eq:4}) is a Bell operator, as defined in Eq.~(\ref{circle-eq:3}),
310: which means we can apply Lemma~\ref{lemma:circle1} to conclude that extreme points of
311: $\cal R$ are achieved by real states on three qubits, with measurements of the form ${M} = \cos \theta {Z} +
312: \sin \theta {X}$. Theorem~1 will emerge as a corollary of:
313: \begin{lemma}
314: \label{lemma:circle2}
315: Let $\ket \psi$ be a pure state in $\CC^2 \otimes \CC^2 \otimes
316: \CC^2$ with real coefficients. Then the maximum of $\left\langle\BB{AB}\right\rangle$ over
317: real traceless observables ${A}_1, {A}_2, {B}_1, { B}_2$ is
318: \begin{equation}
319: \label{circle-eq:8}
320: 2 \sqrt {1 + \YY{A}B^2-\YY{A}C^2-\YY{B}C^2},
321: \end{equation}
322: where $Y$ is the usual Pauli operator, $\YY{A}B = \tr \left(Y_A
323: \otimes Y_B \otimes \one\, \rho\right)$, and so on. Cyclic
324: permutations of Eq.~(\ref{circle-eq:8}) hold for $\langle\BB{AC}\rangle$ and
325: $\langle\BB{BC}\rangle$.
326: %\begin{equation}
327: %\frac{1}{4}\left(\begin{matrix}\BB{AB}^2\\\BB{AC}^2\\\BB{BC}^2\end{matrix}\right) =
328: %\left(\begin{matrix}1\\1\\1\end{matrix}\right) +
329: %\left(\begin{matrix}1&-1&-1\\-1&1&-1\\-1&-1&1\end{matrix}\right)
330: %\left(\begin{matrix}\YY{A}B^2\\\YY{A}C^2\\\YY{B}C^2\end{matrix}\right).
331: %\end{equation}
332: \end{lemma}
333:
334: \begin{proof}We consider $\rho_{AB} = \tr_C \kb \psi$, which is a real
335: state on $\CC^2 \otimes \CC^2$. Horodecki and family have calculated the maximum quantum value of the CHSH
336: operator for a state on $\CC^2 \otimes \CC^2$~\cite{horodecki96:_separ}. Their analysis simplifies in our case
337: because the state and measurements are real. Define
338: \begin{equation}
339: \label{eq:12}
340: T_\text{AB} = \left[\begin{matrix}\langle X_A X_B\rangle & \langle X_A Z_B\rangle \\
341: \langle Z_A X_B\rangle & \langle Z_A Z_B\rangle
342: \end{matrix}\right].
343: \end{equation}
344: For $i=1,2$, write ${A}_i = \hat {a}_i \cdot \vec {\sigma}_r$, ${B}_i = {\hat b}_i \cdot \vec \sigma_r$, where
345: $\hat a_i$ and $\hat b_i$ are two-dimensional unit vectors and $\vec \sigma_r = ({X}, {Z})$. Define
346: \begin{equation}
347: \label{circle-eq:5}
348: \hat {b}_1 + \hat {b}_2 = 2 \cos \theta \hat {d}_1, \ \ \hat {b}_1 - \hat {b}_2 = 2 \sin \theta \hat {d}_2,
349: \end{equation}
350: where $\theta \in [0, \pi/2]$ and $\hat {d}_1$ and $\hat {d}_1$ are orthogonal unit vectors. %Then
351: %\begin{equation}
352: %\BB{AB} = \cos \theta \hat {a}_1^T T \hat {c}_1 + \sin \theta \hat {a}_2^T T \hat {c}_2.
353: %\end{equation}
354: Then
355: \begin{eqnarray}
356: \frac12 \max_{{A}_i, {B}_j} \left \langle\BB{AB}\right \rangle &=& \max_{\hat {d}_i, \theta, \hat {a}_i} \cos \theta \hat {a}_1^t T_\text{AB} \hat {d}_1 + \sin \theta \hat {a}_2^t T_\text{AB} \hat {d}_2 \nonumber \\
357: &=& \max_{\hat {d}_i, \theta} \cos \theta \left\|T_\text{AB} \hat {d}_1\right\| + \sin \theta \left\| T_\text{AB} \hat {d}_2\right\| \nonumber \\
358: &=& \max_{\hat {d}_i} \sqrt{\left\|T_\text{AB} \hat {d}_1\right\|^2 + \left\| T_\text{AB} \hat {d}_2\right\|^2}\label{eq:d}\\
359: &=& \sqrt {\tr \left({T_\text{AB}} {T_\text{AB}}^t\right)}.%\\
360: %&=& \sqrt {\langle X_A X_B\rangle^2 + \langle X_A Z_B\rangle^2 +
361: %\langle Z_A X_B\rangle^2 + \langle Z_A Z_B\rangle^2}. \nonumber
362: \end{eqnarray}
363: This is just the Frobenius norm of $T_\text{AB}$ and it is straightforward to check that, for pure states on
364: $\CC^2 \otimes \CC^2 \otimes \CC^2$ with real coefficients, it is equal to half of
365: Eq.~(\ref{circle-eq:8}).\end{proof}
366: \begin{lemma}
367: \label{sec:dimens-reduct-1}
368: For a pure state $\ket \psi$ with real coefficients in $\CC^2 \otimes \CC^2 \otimes \CC^2$,
369: \begin{equation}
370: \label{eq:9}
371: \max_{{A}_i, {B}_j, {C}_k} \left \langle \BB{AB}\right \rangle^2 + \left \langle \BB{AC}\right\rangle^2 = 8 \left(1 - \YY{B}C^2 \right).
372: \end{equation}
373: \end{lemma}
374: \begin{proof}
375: Lemma~\ref{lemma:circle2}, applied to $\AB$ and $\AC$ separately, immediately implies:
376: \begin{eqnarray}
377: \label{eq:10}
378: \max_{{A}_i, {B}_j, {C}_k} \left \langle \BB{AB}\right \rangle^2 + \left \langle \BB{AC}\right\rangle^2 &\leq&
379: \max_{{A}_i, {B}_j} \left \langle \BB{AB}\right \rangle^2 + \max_{{A}_i, {C}_k} \left \langle \BB{AC}\right\rangle^2\nonumber \\
380: &=& 8 \left(1 - \YY{B}C^2 \right).
381: \end{eqnarray}
382: The reason we do not have equality is that the measurements $A_i$ achieving the maximum in $\AB$ and $\AC$ may be different. We have to show they can be chosen to be the same. Define $T_\text{AC}$ in analogy with Eq.~(\ref{eq:12}) and write the vectors corresponding to C's measurements as
383: \begin{equation}
384: \label{eq:13}
385: \hat {c}_1 + \hat {c}_2 = 2 \cos \theta \hat {e}_1, \ \ \hat {c}_1 - \hat {c}_2 = 2 \sin \theta \hat {e}_2,
386: \end{equation}
387: in analogy with Eq.~(\ref{circle-eq:5}) for B's observables. One can check that $\left[ T_\text{AB} T_\text{AB}^t, T_\text{AC} T_\text{AC}^t\right] = 0$ for all pure states $\ket \psi$ with real coefficients. Hence there are orthonormal vectors $a'_1$ and $a'_2$ that are simultaneous eigenvectors of $T_\text{AB} T_\text{AB}^t$ and $T_\text{AC} T_\text{AC}^t$. Next, note that the term being maximized in Eq.~(\ref{eq:d}), $\|T_\text{AB} \hat d_1
388: \|^2 + \|T_\text{AB} \hat d_2 \|^2$, is actually independent of the
389: $\hat d_i$ (recall that $\hat d_1 \cdot \hat d_2 =0$), so we are free to choose the $\hat d_i$ as we please. Take $\hat d_i = T^t_\text{AB} \hat a_i'$ for $i = 1, 2$ and, similarly, take $\hat e_i = T^t_\text{AC} \hat a_i'$. Alice's measurement vector $\hat a_i$ in the AB maximization of the previous lemma was taken to be the unit vector along $T_\text{AB} \hat d_i$, but this is $T_\text{AB} T^t_\text{AB} \hat a'_i \propto \hat a'_i$ so $\hat a_i = \hat a'_i$. The same will hold in the AC maximization. Hence we can choose A's measurement vectors to be the same in both cases, and we have equality in Eq.~(\ref{eq:9}).
390: % in particular we can
391: %choose take them to be eigenvectors of $T_\text{AB}^t T_\text{AB}$.
392: %Alice's vectors $\hat a_i$ are then the normalized eigenvectors of $T_\text{AB} T_\text{AB}^t$.
393: \end{proof}
394:
395:
396:
397: Combining Lemmas~\ref{lemma:circle1} and~\ref{sec:dimens-reduct-1}, we obtain Theorem~\ref{thm:1}.
398:
399: {\em The monogamy trade-off is tight}.---Lemma~\ref{sec:dimens-reduct-1} also implies that any $\AB$ and $\AC$ compatible with Eq.~(\ref{eq:9}) are achievable. In particular,
400: the state
401: \begin{equation}
402: \label{eq:14}
403: \ket \psi = c_- \left(\ket {010} + \ket {011} \right) + c_+ \left( \ket {100} + \ket {101}\right),
404: \end{equation}
405: where
406: \begin{equation}
407: \label{eq:15}
408: c_\pm = \frac12 \, \sqrt{1 \pm \sqrt 2 \sin t},
409: \end{equation}
410: and $0 \leq t \leq \pi/4$, gives $\AB = 2 \sqrt2 \, \cos t$, $\AC = 2 \sqrt2 \, \sin t$.
411:
412: {\em Extensions.}---In the case of the CHSH inequality we can, in principle, obtain monogamy trade-offs
413: when there are more than three parties via Lemma~\ref{lemma:circle1}, which converts the problem into a finite
414: optimization problem. In the three-party setting, if we are interested in $\BC$ as well as $\AB$ and $\AC$ then
415: we can obtain the trade-off surface numerically. The technique of Lemma~\ref{sec:dimens-reduct-1}---to allow A's
416: measurements to be different in $\AB$ and $\AC$ and then show that they could be the same anyway---does not work.
417: It predicts that the trade-off surface be the intersection of the three cylinders, $\AB^2 + \AC^2 \leq 8$, $\AB^2 +
418: \BC^2 \leq 8$, and $\AC^2 + \BC^2 \leq 8$, but one can show, for example by using the multipartite generalization of
419: Navascues, Pironio and Ac{\'\i}n's semidefinite programming bounds~\cite{navascues:_bound}, that there are points on
420: this surface that are not achievable. It would be interesting to extend the semidefinite programming technique to
421: obtain monogamy inequalities for other Bell inequalities.
422:
423:
424: %{Lemma~2}: For real states on $\CC^2 \otimes \CC^2
425: %\otimes \CC^2$, allowed values of the YY correlation described by the tetrahedron
426: %\begin{equation}
427: % \label{circle-eq:9}
428: %(-1)^{x} \YY AB + (-1)^y \YY AC - (-1)^{x+y} \YY BC \leq 1
429: %\end{equation}
430: %where $x, y = \pm 1$.
431: %% {\it Proof}. Any point inside the octahedron is acheivable: the state
432: %% $\ket{\psi^-}_{AB} \ket 0_C$ has $\YY A B = -1$, and so each of the
433: %% vertices of the octahedron is achievable by symmetry. Next suppose there is a state $\rho$ that violates Eq.(\ref{circle-eq:9}). We write
434:
435: %{\it Proof}. Write the state as
436: %\begin{equation}
437: % \label{circle-eq:10}
438: % \rho =\frac18 \sum_{ijk} s_{ijk} \sigma_i \otimes \sigma_j \otimes \sigma_k
439: %\end{equation}
440: %with $\sigma_0 = {I}, \sigma_1 = {X}, \sigma_2 = {Y}, \sigma_3 = {Z}$. Note that $s_{000} = 1$, $s_{220} = \YY AB$, $s_{202} = \YY AC$, $s_{022}= \YY BC$. We now transform the state so all other coefficients vanish. Taking a mixture $\rho' = \left( \rho + {X}_A{Z}_A \rho {Z}_A{X}_A \right)$ preserves the coefficients of interest, but sets $s'_{ijk} = 0$ if $i = 1$ or $3$. Similar convex sums for other parties. Note that $s_{002} = s_{020} = s_{200} = s_{222} = 0$ because $\rho$ is real. Therefore $(\YY AB, \YY AC, \YY BC)$ is allowed iff
441: %\begin{equation}
442: % \label{circle-eq:11}
443: %% I_AI_BI_C + \YY AB Y_A Y_B I_C + \YY AC Y_A I_B Y_C + \YY BC I_A Y_B Y_C
444: % III + \YY AB Y Y I + \YY AC Y I Y + \YY BC I Y Y \nonumber
445: %\end{equation}
446: %is positive. Taking eigenvalues yields the condition stated in the lemma.\qed
447:
448: {\em Bell inequality violation and anticommutation}.---The precise form of Eq.~(\ref{eq:9}) suggests a
449: general connection between the trade-off of Bell inequality violation and the expectation values of
450: anticommuting observables; indeed, the operator $Y_BY_C$ anticommutes with all observables $B_j,C_k$ measured in
451: the Bell test. We now give a more general result, restricting for simplicity to the two party case.
452: \begin{theorem}
453: \label{thm:n}
454: Let ${\cal B} = \sum_{i,j=1}^n p_{ij} A_i \otimes B_j$ be a 2-party 2-outcome correlation Bell operator such
455: that $\tr \rho {\cal B} \leq {\cal Q}$ for all shared states $\rho$ and all observables $A_i$, $B_j$ (with spectrum in $[-1,+1]$). Let $W$ be any observable (with spectrum in $[-1,+1]$) on Bob's Hilbert space that
456: anticommutes with $B_j$ for all $j$. Then
457: \begin{equation}
458: \tr \rho {\cal B} \leq {\cal Q}\sqrt {1 - \left( \tr \rho W\right)^2 }.
459: \end{equation}
460: \end{theorem}
461: \begin{proof}
462: We start by noting that it is sufficient to restrict to the case
463: where $\rho$ is pure, i.e., $\rho = \ket \psi \bra \psi$. The general case then follows by applying
464: Jensen's inequality to the concave function $f(x) = \sqrt{1-x^2}$.
465: % \begin{eqnarray}
466: % \label{eq:5}
467: %\tr \rho {\cal B} &\leq& {\cal Q} \sum_i p_i \sqrt {1 - \bra{\psi_i}W\ket{\psi_i}^2 }\\
468: %&\leq& Q \sqrt {1 - \left( \sum_i p_i \bra{\psi_i}W\ket{\psi_i}\right) ^2 }
469: % \end{eqnarray}
470:
471: % We can assume that there is some operator $B$ that anticommutes with $W$.
472: % Then $[B^2, W]=0$ so $B^2$ and $W$ have common eigenvalues.
473: If $\ket w$ is an eigenvector of $W$ with eigenvalue $w$, $B_j\ket
474: w$ is either 0 or an eigenvector of $W$ with eigenvalue $-w$, since
475: $B_j$ and $W$ anticommute. This means that there is a decomposition
476: ${\cal H}_B = {\cal W}_0 \oplus {\cal W}_1 \oplus {\cal W}_2$, where
477: $W$ annihilates vectors in ${\cal W}_0$, every $B_j$ annihilates
478: vectors in ${\cal W}_1$, and ${\cal W}_2$ is spanned by eigenvectors of
479: $W$ with nonzero eigenvalues, which occur in positive/negative pairs.
480:
481: Denote the distinct positive
482: eigenvalues associated with eigenvectors of $W$ in ${\cal W}_2$ as
483: $0 < w_2 < w_3 < \cdots < w_m \leq 1$ and
484: let $V_i^\pm$ be the subspace in ${\cal W}_2$ corresponding to eigenvalue
485: $\pm w_i$. Decompose $\ket \psi$ as $\ket \psi = \sum_{i=0}^m\sqrt{p_i}\ket {\psi_i}$, where $\sum_i p_i = 1$, $\ket {\psi_0} \in {\cal W}_0$, $\ket {\psi_1} \in {\cal W}_1$, $\bk {\psi_0} {\psi_0} = \bk {\psi_1}{\psi_1} = 1$, and for $i \geq 2$, $\ket {\psi_i} = \left( \cos \theta_i \ket{w_i^+} +
486: \sin \theta_i\ket{w_i^-}\right)$, $\ket {w_i^\pm} \in V_i^\pm$, and
487: $\bk {w_i^\pm}{w_i^\pm}=1$. Then
488: ${\cal B}\ket {w_i^\pm} \in V_i^\mp$ for $i \geq 2$. %since each $B_j$ anticommutes with $W$.
489: It follows that
490: \begin{eqnarray}
491: \bra \psi {\cal B} \ket\psi &=& p_0 \bra {\psi_0} {\cal B} \ket {\psi_0} + \sum_{i=2}^m p_i \mathrm{Re}\left(\bra {w_i^+} {\cal B} \ket {w_i^-} \right)\sin 2 \theta_i,\nonumber\\
492: \bra \psi W \ket\psi &=& p_1 \bra {\psi_1} W \ket {\psi_1} + \sum_{i=2}^m p_i w_i^2 \cos 2 \theta_i.
493: \end{eqnarray}
494: But these are the same expressions we would obtain with the mixed state $\rho = \sum_i p_i \ket {\psi_i} \bra {\psi_i}$. It therefore follows from our initial remark that it is sufficient to prove the claim for each state $\ket {\psi_i}$. For $i=0,1$, the result is trivial. Fix $i\geq 2$, set $\chi = \mathrm{Re}\left(\bra {w_i^+} {\cal B} \ket {w_i^-} \right)$ and let $x \in \{\pm1\}$ be the sign of $\chi$. Set $\ket \phi =
495: \frac{1}{\sqrt2} \ket {w_i^-} + x \frac{1}{\sqrt2} \ket {w_i^+}$.
496: Then $\bra \phi {\cal B} \ket\phi = \chi < Q$ by assumption, while
497: \begin{eqnarray}
498: \label{eq:3}
499: \bra {\psi_i} {\cal B} \ket{\psi_i} &\leq& \chi \sin 2 \theta_i\\
500: &\leq& \chi \sqrt{1 - \bra {\psi_i} W \ket{\psi_i}^2 }\\
501: &\leq& Q \sqrt{1 - \bra {\psi_i} W \ket{\psi_i}^2 }.
502: \end{eqnarray}
503: %Observe that
504: %\begin{equation}
505: % \label{eq:6}
506: %\sum_{ij} |a_i^+a_j^-||a_i^-a_j^+| \leq \sqrt{\sum_{ij} (a_i^+ a_j^-)^2} \sqrt{\sum_{ij} (a_i^- a_j^+)^2} = \sum_{ij} (a_i^- a_j^+)^2
507: %\end{equation}
508: %by Cauchy-Schwartz.
509: This completes the proof.
510: %\begin{eqnarray}
511: % \label{eq:3}
512: %\Delta &\equiv& \bra \phi {\cal B} \ket\phi^2 - \bra \psi {\cal B} \ket\psi^2 \\ &\geq& m^2 \left [ 1 - 4 \sum_i \left | a_i^+ a_i^-\right| \sum_j \left| a_j^+ a_j^- \right|\right].
513: %\end{eqnarray}
514: %Then
515: %\begin{eqnarray}
516: %\label{eq:4}
517: %\bra \psi W \ket\psi &=& \sum_i w_i^2 \left[(a_i^+)^2 - (a_i^-)^2\right]\\ &\leq& \left| \sum_i (a_i^+)^2 - (a_i^-)^2\right|.
518: %\end{eqnarray}
519: %Hence
520: %\begin{eqnarray}
521: % \label{eq:7}
522: %\Delta &\geq& m^2 \left[\sum_i \left((a_i^+)^2 + (a_i^-)^2\right) \sum_j \left((a_j^+)^2 + (a_j^-)^2 \right) - 4 \sum_{ij} (a_i^- a_i^+)^2 \right]
523: %\\&\geq& m^2 \langle W \rangle ^2, \end{eqnarray}
524: %by Eq.~(\ref{eq:4}). The claim follows.
525: \end{proof}
526:
527: We now apply Theorem~\ref{thm:n} to the CHSH inequality. If $B_1$ and $B_2$ are
528: observables with $\pm 1$ eigenvalues, then they both anticommute with
529: their commutator $i [ B_1, B_2]/2$ (the factor of $i/2$ makes this an
530: observable). Applying Theorem~\ref{thm:n} to both Alice and Bob's
531: observables, it follows that
532: \begin{eqnarray}
533: \label{eq:8}
534: \AB &\leq& 2 \sqrt {2 - \left|\langle [B_1, B_2] \rangle\right|^2 }. \\
535: \AB &\leq& 2 \sqrt {2 - \left|\langle [A_1, A_2] \rangle\right|^2 }.
536: \end{eqnarray}
537: These are local analogues of Tsirelson's bound~\cite{Tsirelson:80a},
538: \begin{equation}
539: \AB \leq \sqrt {4 + \left|\langle [A_1,A_2][B_1, B_2] \rangle\right| }.\\
540: \end{equation}
541: In particular, for maximal quantum violation of the CHSH inequality, the local observables corresponding to the
542: commutators must be locally random $\langle [A_1, A_2] \rangle = \langle [B_1, B_2] \rangle = 0$ but perfectly
543: correlated $\left|\langle [A_1,A_2][B_1, B_2] \rangle\right| = 4$. This is a clear manifestation of the fact
544: that entanglement goes hand in hand with local randomness.
545:
546: In conclusion, we investigated Bell correlations in a tripartite setting and obtained tight monogamy bounds on
547: the trade-off between them. The main message is depicted in Figure~1, which gives a universal picture of nonlocal
548: correlations valid for quantum systems of any dimension.
549:
550:
551:
552: {\em Acknowledgments}.---We thank Richard Cleve, Wim van Dam, Andrew Doherty, Nicolas Gisin, and Oded Regev for discussions. This
553: work was supported in part by the NSF under Grants PHY-0456720 and CCF-0524828, the ARO under Grant W911NF-05-1-0294, EU project QAP 015848, and NWO VICI project 639-023-302.
554:
555: %\bibliography{../../newref}
556:
557: \begin{thebibliography}{24}
558: \expandafter\ifx\csname natexlab\endcsname\relax\def\natexlab#1{#1}\fi
559: \expandafter\ifx\csname bibnamefont\endcsname\relax
560: \def\bibnamefont#1{#1}\fi
561: \expandafter\ifx\csname bibfnamefont\endcsname\relax
562: \def\bibfnamefont#1{#1}\fi
563: \expandafter\ifx\csname citenamefont\endcsname\relax
564: \def\citenamefont#1{#1}\fi
565: \expandafter\ifx\csname url\endcsname\relax
566: \def\url#1{\texttt{#1}}\fi
567: \expandafter\ifx\csname urlprefix\endcsname\relax\def\urlprefix{URL }\fi
568: \providecommand{\bibinfo}[2]{#2}
569: \providecommand{\eprint}[2][]{\url{#2}}
570:
571: \bibitem[{\citenamefont{Bell}(1964)}]{Bell:64a}
572: \bibinfo{author}{\bibfnamefont{J.~S.} \bibnamefont{Bell}},
573: \bibinfo{journal}{Physics} \textbf{\bibinfo{volume}{1}}, \bibinfo{pages}{195}
574: (\bibinfo{year}{1964}).
575:
576: \bibitem[{\citenamefont{Clauser et~al.}(1969)\citenamefont{Clauser, Horne,
577: Shimony, and Holt}}]{Clauser:69a}
578: \bibinfo{author}{\bibfnamefont{J.~F.} \bibnamefont{Clauser}},
579: \bibinfo{author}{\bibfnamefont{M.~A.} \bibnamefont{Horne}},
580: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Shimony}}, \bibnamefont{and}
581: \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{Holt}},
582: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{23}},
583: \bibinfo{pages}{880} (\bibinfo{year}{1969}).
584:
585: \bibitem[{\citenamefont{Ekert}(1991)}]{PhysRevLett.67.661}
586: \bibinfo{author}{\bibfnamefont{A.~K.} \bibnamefont{Ekert}},
587: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{67}},
588: \bibinfo{pages}{661} (\bibinfo{year}{1991});
589: \bibinfo{author}{\bibfnamefont{C.~H.} \bibnamefont{Bennett}} \bibnamefont{and}
590: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Brassard}}, in
591: \emph{\bibinfo{booktitle}{Proceedings of IEEE International Conference on
592: Computers, Systems, and Signal Processing}} (\bibinfo{publisher}{IEEE},
593: \bibinfo{year}{1984}), pp. \bibinfo{pages}{175--179};
594: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Barrett}},
595: \bibinfo{author}{\bibfnamefont{L.}~\bibnamefont{Hardy}}, \bibnamefont{and}
596: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Kent}},
597: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{95}},
598: \bibinfo{pages}{010503} (\bibinfo{year}{2005});
599: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Ac{\'\i}n}},
600: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Gisin}}, \bibnamefont{and}
601: \bibinfo{author}{\bibfnamefont{Ll.}~\bibnamefont{Masanes}}, \bibinfo{note}{quant-ph/0510094}.
602:
603: \bibitem[{\citenamefont{Cirel'son}(1980)}]{Tsirelson:80a}
604: \bibinfo{author}{\bibfnamefont{B.~S.} \bibnamefont{Cirel'son}},
605: \bibinfo{journal}{Lett. Math. Phys.} \textbf{\bibinfo{volume}{4}},
606: \bibinfo{pages}{93} (\bibinfo{year}{1980}).
607:
608: \bibitem[{\citenamefont{Tsirelson}(1987)}]{Tsirelson:85b}
609: \bibinfo{author}{\bibfnamefont{B.~S.} \bibnamefont{Tsirelson}},
610: \bibinfo{journal}{J. Soviet Math.} \textbf{\bibinfo{volume}{36}},
611: \bibinfo{pages}{557} (\bibinfo{year}{1987});
612: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Tsirelson}},
613: \bibinfo{journal}{Hadronic J. Suppl.} \textbf{\bibinfo{volume}{8}},
614: \bibinfo{pages}{329} (\bibinfo{year}{1993});
615: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Cleve}},
616: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{H{\o}yer}},
617: \bibinfo{author}{\bibfnamefont{B.}~\bibnamefont{Toner}}, \bibnamefont{and}
618: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Watrous}}, in
619: \emph{\bibinfo{booktitle}{Proceedings of the 19th IEEE Conference on
620: Computational Complexity (CCC 2004)}}, pp.
621: \bibinfo{pages}{236--249};
622: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Buhrman}} \bibnamefont{and}
623: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Massar}}, \bibinfo{note}{quant-ph/0409066};
624: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Wehner}},
625: \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{73}},
626: \bibinfo{eid}{022110} (\bibinfo{year}{2006}).
627:
628: \bibitem[{\citenamefont{Toner}(2006)}]{toner:_monog}
629: \bibinfo{author}{\bibfnamefont{B.~F.} \bibnamefont{Toner}}, \bibinfo{note}{quant-ph/0601172}.
630:
631: \bibitem[{\citenamefont{Navascues et~al.}()\citenamefont{Navascues, Pironio,
632: and Acin}}]{navascues:_bound}
633: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Navascues}},
634: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Pironio}}, \bibnamefont{and}
635: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Acin}},
636: \bibinfo{note}{quant-ph/0607119}.
637:
638: \bibitem[{\citenamefont{Popescu and Rohrlich}(1994)}]{Popescu:94a}
639: \bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Popescu}} \bibnamefont{and}
640: \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Rohrlich}},
641: \bibinfo{journal}{Found. Phys.} \textbf{\bibinfo{volume}{24}},
642: \bibinfo{pages}{379} (\bibinfo{year}{1994});
643: \bibinfo{author}{\bibfnamefont{L.~A.} \bibnamefont{Khalfin}} \bibnamefont{and}
644: \bibinfo{author}{\bibfnamefont{B.~S.} \bibnamefont{Tsirelson}}, in
645: \emph{\bibinfo{booktitle}{Symposium on the Foundations of Modern Physics}},
646: edited by \bibinfo{editor}{\bibfnamefont{P.}~\bibnamefont{Lahti}}
647: \bibnamefont{and}
648: \bibinfo{editor}{\bibfnamefont{P.}~\bibnamefont{Mittelstaedt}}
649: (\bibinfo{publisher}{World Scientific}, \bibinfo{address}{Singapore},
650: \bibinfo{year}{1985}), pp. \bibinfo{pages}{441--460}.
651:
652:
653: \bibitem[{\citenamefont{{van Dam}}(2005)}]{dam05:_implaus_conseq_super_nonloc}
654: \bibinfo{author}{\bibfnamefont{W.}~\bibnamefont{{van Dam}}}, \bibinfo{note}{quant-ph/0501159};
655: \bibinfo{author}{\bibfnamefont{G.}~\bibnamefont{Brassard}},
656: \bibinfo{author}{\bibfnamefont{H.}~\bibnamefont{Buhrman}},
657: \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Linden}},
658: \bibinfo{author}{\bibfnamefont{A.~A.} \bibnamefont{Methot}},
659: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Tapp}}, \bibnamefont{and}
660: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Unger}},
661: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{96}},
662: \bibinfo{eid}{250401} (\bibinfo{year}{2006}).
663:
664: \bibitem[{\citenamefont{Coffman et~al.}(2000)\citenamefont{Coffman, Kundu, and
665: Wootters}}]{Coffman:00a}
666: \bibinfo{author}{\bibfnamefont{V.}~\bibnamefont{Coffman}},
667: \bibinfo{author}{\bibfnamefont{J.}~\bibnamefont{Kundu}}, \bibnamefont{and}
668: \bibinfo{author}{\bibfnamefont{W.~K.} \bibnamefont{Wootters}},
669: \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{61}},
670: \bibinfo{pages}{052306} (\bibinfo{year}{2000});
671: \bibinfo{author}{\bibfnamefont{T.~J.} \bibnamefont{Osborne}} \bibnamefont{and}
672: \bibinfo{author}{\bibfnamefont{F.}~\bibnamefont{Verstraete}},
673: \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{96}},
674: \bibinfo{eid}{220503} (\bibinfo{year}{2006}).
675:
676: \bibitem[{\citenamefont{Masanes et~al.}(2006)\citenamefont{Masanes, Ac{\'\i}n,
677: and Gisin}}]{masanes05:_gener_nonsig_theor}
678: \bibinfo{author}{\bibfnamefont{Ll.}~\bibnamefont{Masanes}},
679: \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Ac{\'\i}n}},
680: \bibnamefont{and} \bibinfo{author}{\bibfnamefont{N.}~\bibnamefont{Gisin}},
681: \bibinfo{journal}{Phys. Rev. A} \textbf{\bibinfo{volume}{73}},
682: \bibinfo{eid}{012112} (\bibinfo{year}{2006}).
683:
684: \bibitem[{\citenamefont{Masanes}(2005)}]{masanes05:_extrem}
685: \bibinfo{author}{\bibfnamefont{Ll.}~\bibnamefont{Masanes}}, \bibinfo{note}{quant-ph/0512100}.
686:
687: \bibitem[{\citenamefont{Bhatia}(1996)}]{bhatia96:_matrix_analy}
688: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Bhatia}},
689: \emph{\bibinfo{title}{Matrix Analysis}}, vol. \bibinfo{volume}{169} of
690: \emph{\bibinfo{series}{Graduate texts in mathematics}}
691: (\bibinfo{publisher}{Springer-Verlag}, \bibinfo{address}{New York},
692: \bibinfo{year}{1996}), \bibinfo{note}{section VII.1}.
693:
694: \bibitem[{\citenamefont{Horodecki et~al.}(1996)\citenamefont{Horodecki,
695: Horodecki, and Horodecki}}]{horodecki96:_separ}
696: \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Horodecki}},
697: \bibinfo{author}{\bibfnamefont{P.}~\bibnamefont{Horodecki}},
698: \bibnamefont{and}
699: \bibinfo{author}{\bibfnamefont{R.}~\bibnamefont{Horodecki}},
700: \bibinfo{journal}{Phys. Lett. A} \textbf{\bibinfo{volume}{223}},
701: \bibinfo{pages}{1} (\bibinfo{year}{1996}).
702:
703: %\bibitem[{\citenamefont{Popescu and Rohrlich}(1992)}]{popescu92:_which}
704: %\bibinfo{author}{\bibfnamefont{S.}~\bibnamefont{Popescu}} \bibnamefont{and}
705: % \bibinfo{author}{\bibfnamefont{D.}~\bibnamefont{Rohrlich}},
706: % \bibinfo{journal}{Phys. Lett. A} \textbf{\bibinfo{volume}{169}},
707: % \bibinfo{pages}{411} (\bibinfo{year}{1992});
708: %\bibinfo{author}{\bibfnamefont{S.~L.} \bibnamefont{Braunstein}},
709: % \bibinfo{author}{\bibfnamefont{A.}~\bibnamefont{Mann}}, \bibnamefont{and}
710: % \bibinfo{author}{\bibfnamefont{M.}~\bibnamefont{Revzen}},
711: % \bibinfo{journal}{Phys. Rev. Lett.} \textbf{\bibinfo{volume}{68}},
712: % \bibinfo{pages}{3259} (\bibinfo{year}{1992}).
713:
714:
715: \end{thebibliography}
716:
717:
718:
719: \end{document}
720: