cs0607139/cs0607139
1: \documentclass[10pt]{article}
2: \usepackage{amsmath}
3: \usepackage{amsthm}
4: \usepackage{amssymb}
5: \usepackage{subfigure}
6: \newtheorem{definition}{Definition}
7: 
8: \newtheorem{theorem}[definition]{Theorem}
9: \newtheorem{corollary}[definition]{Corollary}
10: \newtheorem{proposition}[definition]{Proposition}
11: \newtheorem{lemma}[definition]{Lemma}
12: \newtheorem{claim}[definition]{Claim}
13: \newtheorem{fact}[definition]{Fact}
14: \newtheorem{conjecture}[definition]{Conjecture}
15: 
16: \usepackage{typearea}
17: \newcommand{\SD}[2]{\|#1 - #2\|}
18: \newcommand*{\n}{\penalty 100\relax} 
19: \newcommand{\eps}{\varepsilon}
20: \newcommand{\B}{\backslash}
21: \newcommand{\xvec}{x^n}
22: \newcommand{\yvec}{y^n}
23: \newcommand{\xkvec}{x^k}
24: \newcommand{\ykvec}{y^k}
25: \newcommand{\ukvec}{u^k}
26: \newcommand{\vkvec}{v^k}
27: \newcommand*{\Pd}{\mathsf{P}}
28: \newcommand*{\cupdisj}{\mathbin{\dot{\cup}}}
29: \DeclareMathOperator{\cov}{cov}
30: \DeclareMathOperator*{\E}{E}
31: \DeclareMathOperator*{\val}{val}
32: \newcommand*{\bbN}{\mathbb{N}}
33: \newcommand*{\cbar}{\bar{c}}
34: \newcommand*{\ibar}{\bar{i}}
35: \newcommand*{\pbar}{\bar{p}}
36: \newcommand*{\vbar}{\bar{v}}
37: \newcommand*{\xbar}{\bar{x}}
38: \newcommand*{\ybar}{\bar{y}}
39: \newcommand*{\Abar}{\overline{A}}
40: \newcommand*{\Bbar}{\overline{B}}
41: \newcommand*{\Lbar}{\overline{L}}
42: \newcommand*{\Ubar}{\overline{U}}
43: \newcommand*{\Sbar}{\overline{S}}
44: \newcommand*{\Tbar}{\overline{T}}
45: \newcommand*{\Wbar}{\overline{W}}
46: \newcommand*{\Xbar}{\overline{X}}
47: \newcommand*{\Ybar}{\overline{Y}}
48: \newcommand*{\cA}{\mathcal{A}}
49: \newcommand*{\cB}{\mathcal{B}}
50: \newcommand*{\cE}{\mathcal{E}}
51: \newcommand*{\cR}{\mathcal{R}}
52: \newcommand*{\cS}{\mathcal{S}}
53: \newcommand*{\cT}{\mathcal{T}}
54: \newcommand*{\cU}{\mathcal{U}}
55: \newcommand*{\cV}{\mathcal{V}}
56: \newcommand*{\cX}{\mathcal{X}}
57: \newcommand*{\cY}{\mathcal{Y}}
58: \newcommand*{\cZ}{\mathcal{Z}}
59: \newcommand*{\Atilde}{\widetilde{A}}
60: \newcommand*{\Btilde}{\widetilde{B}}
61: \newcommand*{\Stilde}{\widetilde{S}}
62: \newcommand*{\Ttilde}{\widetilde{T}}
63: \newcommand*{\Utilde}{\widetilde{U}}
64: \newcommand*{\Vtilde}{\widetilde{V}}
65: \newcommand*{\Xtilde}{\widetilde{X}}
66: \newcommand*{\Ytilde}{\widetilde{Y}}
67: \newcommand{\mkA}{\mathfrak{A}}
68: \newcommand{\mkB}{\mathfrak{B}}
69: \newcommand{\mkG}{\mathfrak{G}}
70: \newcommand{\mkL}{\mathfrak{L}}
71: \newcommand{\mkN}{\mathfrak{N}}
72: \newcommand{\mkQ}{\mathfrak{Q}}
73: \newcommand{\mkW}{\mathfrak{W}}
74: \newcommand{\Wevent}{W}
75: \newcommand{\NoJ}{{(\backslash j)}}
76: \newcommand{\Tot}{\mathrm{Tot}}
77: 
78: \newcommand{\vns}{v_{ns}}
79: 
80: \newcommand{\cancel}[1]{}
81: \newcommand{\nocancel}{}
82: \title{Parallel Repetition: \\
83:   Simplifications and the No-Signaling Case}
84: \author{Thomas Holenstein\thanks{Microsoft Research, Silicon Valley; 
85:       \texttt{thomahol@microsoft.com}. This work was done while the author was at ETH Zurich.}}
86: \date{}
87: \begin{document}
88: \maketitle
89: \begin{abstract}
90:   Consider a game where a referee chooses~$(x,y)$ according
91:   to a publicly known distribution~$\Pd_{XY}$, sends~$x$ to Alice, and
92:   $y$ to Bob.  Without communicating with each other, Alice responds
93:   with a value~$a$ and Bob responds with a value~$b$.  Alice and Bob
94:   jointly win if a publicly known predicate~$Q(x,y,a,b)$ holds.
95:   
96:   Let such a game be given and assume that the maximum probability
97:   that Alice and Bob can win is~$v<1$.  Raz (SIAM J.~Comput.~27, 1998)
98:   shows that if the game is repeated~$n$ times in parallel, then the
99:   probability that Alice and Bob win \emph{all} games simultaneously
100:   is at most~$\vbar^{\tfrac{n}{\log(s)}}$, where $s$ is the maximal
101:   number of possible responses from Alice and Bob in the initial game,
102:   and $\vbar<1$ is a constant depending only on~$v$.  
103:   
104:   In this work, we simplify Raz's proof in various ways and thus
105:   shorten it significantly.  Further we study the case where Alice and
106:   Bob are not restricted to local computations and can use any
107:   strategy which does not imply communication among them.
108: \end{abstract}
109: \section{Introduction}
110: 
111: %\sloppy
112: 
113: The question how much parallel repetition of a game as in the abstract
114: reduces the winning probability of the players was motivated by the
115: study of two-prover interactive proofs, initiated by Ben-Or et al.
116: \cite{BGKW88}.  It was first conjectured that in a game which is
117: repeated~$n$ times in parallel, the probability that Alice and Bob win
118: all the games simultaneously is at most~$v^n$ (see \cite{FoRoSi94}).
119: However, later a counterexample to this conjecture was given
120: \cite{Fortno89}.
121: 
122: \paragraph{Related Work}
123: Various papers give upper bounds on the winning probability of a game
124: which is repeated~$n$ times in parallel
125: \cite{CaCoLi92,Feige91,LapSha95,Raz98,Verbit94}.  However, the upper
126: bound given by Raz \cite{Raz98} is the only explicit bound for
127: arbitrary distributions~$\Pd_{XY}$ (it is also quantitatively the
128: strongest).  Parnafes, Raz, and Wigderson \cite{PaRaWi97} modify Raz's
129: proof to show that the term~$\log(s)$ can be replaced by a parameter
130: which is much smaller for some games.
131: 
132: Games for which the~$n$-fold parallel repetition decreases the winning
133: probability less than from~$v$ to~$v^n$ were also constructed: Fortnow
134: \cite{Fortno89} gives\footnote{For readers not familiar with such
135:   counter-examples, a variation of Fortnow's game is reproduced in
136:   Appendix~\ref{app:nontriviality}.} a game for which the maximal
137: winning probability in two repetitions is larger than~$v^2$ (see also
138: \cite{FeiLov92}), Feige \cite{Feige91} constructs a game where the
139: winning probability in two parallel repetitions does not decrease at
140: all, and Feige and Verbitsky \cite{FeiVer02} give, for infinitely
141: many~$s$, a game where~$\Theta(\frac{\log(s)}{\log{\log(s)}})$
142: repetitions decrease the winning probability from at most~$\frac34$ to
143: at least~$\frac18$, where $s$ is the number of possible answers Alice
144: and Bob can give.  This last result shows that in general Raz's bound
145: is close to optimal.
146: 
147: \paragraph{No-signaling strategies}
148: No-signaling strategies are all those strategies which do not imply
149: communication.  Popescu and Rohrlich \cite{PopRoh94} give an example
150: of such a strategy: Alice receives a bit~$x$, Bob receives a bit~$y$,
151: and they respond with uniform random bits~$a$ and~$b$ such
152: that~$a\oplus b = x\land y$.  Note that even though we cannot
153: implement this strategy with shared randomness and without
154: communication, Alice and Bob cannot communicate if they only have
155: black-box access to such functionality.
156: 
157: The study of no-signaling strategies is motivated by the idea that if
158: Alice and Bob share some entangled quantum state, the set of possible
159: strategies they might use increases, but stays a subset of the
160: no-signaling strategies (this subset is strict: for example the above
161: strategy which achieves~$a\oplus b=x\land y$ from~$(x,y)$ cannot be
162: simulated perfectly using quantum mechanics \cite[Problem
163: 2.3]{NieChu00}, \cite{Cirels80} --- the corresponding game is called
164: the CHSH-game \cite{CHSH69}).  
165: 
166: We remark that there are games which can be won with probability 1
167: given a shared quantum state (and thus with a no-signaling strategy),
168: but not using local strategies.  Those are called ``pseudo-telepathy
169: games'' (see \cite{BrBrTa05} and the references therein).
170: 
171: A parallel repetition theorem for the case where Alice and Bob share
172: a quantum state and the decision of the referee only depends on the
173: XOR of the binary answers of Alice and Bob was recently given by
174: Cleve et al.~\cite{CSUU06}.
175: 
176: \paragraph{Contributions of this paper}
177: In this paper we simplify Raz's proof.  Most importantly, we replace a
178: large part (essentially Section 6) of Raz's paper with the simpler
179: Lemma~\ref{lem:LocallyComputableCommonPartPre}.  This also allows us to
180: give an explicit bound on the maximal winning probability of a game
181: repeated~$n$ times in parallel (Raz does not explicitly describe the
182: dependence of~$\vbar$ on~$v$).  
183: 
184: The use of Lemma~\ref{lem:LocallyComputableCommonPartPre} also makes
185: the rest of the argument simpler.  We shortly explain why:  The main
186: part of the proof consists of showing that the information the players
187: get in the~$n$-fold repetition does not help them to win the subgame
188: in some coordinate~$j$, even conditioned on the event that certain
189: other subgames are won.  This is done in three steps.  In two of these
190: steps the information does not help the players because they can
191: generate this information themselves with local computation only.
192: Lemma~\ref{lem:LocallyComputableCommonPartPre} shows that this also holds
193: for the third step.  This allows us to merge some of the steps, which
194: simplifies the overal structure.
195: 
196: We also study how much the term~$\log(s)$ in the exponent in the
197: parallel repetition theorem can be reduced.  In \cite{PaRaWi97} it is
198: shown that the logarithm of the partition number of the accepance
199: predicate can be used instead of~$\log(s)$.  Based on the ideas from
200: there, Theorem~\ref{thm:pr_local_strengthened} gives a bound which
201: might be stronger for some games.
202: 
203: Finally, we prove a parallel repetition theorem in case Alice and Bob
204: are restricted to no-signaling strategies (in both the given game and
205: the parallel repetition of it).
206: 
207: \section{Notation and Basic Facts}
208: \subsection{Probability Distributions}\label{sec:probdistr}
209: 
210: We use calligraphic letters to denote sets.  We denote random
211: variables using capital letters, and values with lower case letters.
212: We use superscripts to denote tuples, e.g., $X^n := (X_1, \ldots,
213: X_{n})$ and $x^n :=\n (x_1, \ldots, x_{n})$.
214: 
215: If a distribution $\Pd_{XY}$ over~$\cX \times \cY$ is given, we write
216: $\Pd_X$ or~$\Pd_Y$ to denote the marginal distribution, e.g.,
217: $\Pd_X(x) := \n\sum_{y\in\cY}\Pd_{XY}(x,y)$.  The conditional
218: distribution~$\Pd_{Y|X=x}$ is $\Pd_{Y|X=x}(y) :=
219: \Pd_{XY}(x,y)/\Pd_X(x)$.
220: 
221: Let~$\Pd_{X_0}$ be a distribution over~$\cX$ and~$\Pd_{Y_1|X_1=x}$ be
222: a conditional distribution over~$\cY$.  We define the
223: distribution~$\Pd_{X_0}\Pd_{Y_1|X_1}$ over~$\cX\times\cY$ as
224: \begin{align}\label{eq:29}
225:   (\Pd_{X_0} \Pd_{Y_1|X_1})(x,y) := \Pd_{X_0}(x) \cdot \Pd_{Y_1|X_1=x}(y).
226: \end{align}
227: For this, it is necessary that $\Pd_{Y_1|X_1=x}$ is defined for
228: every~$x\in\cX$.  We also use this notation when~$\Pd_{Y_1|X_1=x}$ is
229: defined as marginal of a given distribution $\Pd_{X_1Y_1}$.  In this
230: case, we define~$\Pd_{Y_1|X_1=x}$ in an arbitrary way
231: if~$\Pd_{X_1}(x)=0$.  This notation is used for example in
232: Corollary~\ref{cor:LocallyComputableCommonPart} in the form
233: $\Pd_{X_0Y_0}\Pd_{S|X}$, where it is understood as
234: $(\Pd_{X_0Y_0}\Pd_{S|X})(x,y,s) := \Pd_{X_0Y_0}(x,y)\Pd_{S|X=x}(s)$.
235: Note that the conditional distribution~$\Pd_{S|X=x}$ is defined
236: there by the marginal distribution~$\Pd_{SX}$ of the given
237: distribution~$\Pd_{SXY}$.  Our notation is not explicit since it does
238: not specify which random variables are associated with each other.
239: However, this will always be clear from the context.
240: 
241: For two probability distributions~$\Pd_{X_0}$ and~$\Pd_{X_1}$
242: over the same set~$\cX$ we define the statistical distance
243: \begin{align}
244:   \|\Pd_{X_0} - \Pd_{X_1}\| := 
245:   \frac12\sum_{x\in\cX} \bigl|\Pd_{X_0}(x)-\Pd_{X_1}(x)\bigr|.
246: \end{align}
247: \subsection{Games}
248: \begin{definition}\label{def:game}
249:   A \emph{game~$\mkG=(\Pd_{XY},Q)$
250:     over~$\cX\times\cY\times\cA\times\cB$} is a
251:   distribution~$\Pd_{XY}$ over~$\cX \times\cY$ and a predicate~$Q$
252:   over~$\cX\times\cY\times\cA\times\cB$.  The \emph{value}~$v(\mkG)$
253:   of a game is
254:   \begin{align*}
255:     v(\mkG) := \max_{h_a,h_b} \Pr_{XY}[Q(X,Y,h_a(X),h_b(Y))],
256:   \end{align*}
257:   where the maximization is over functions~$h_a: \cX \rightarrow \cA$
258:   and~$h_b: \cY\rightarrow\cB$.  
259:   A \emph{strategy}~$(h_a,h_b)$ for a game is a pair of such functions.
260: \end{definition}
261: 
262: Sometimes also randomized strategies for Alice and Bob are considered,
263: where~$h_a$ and~$h_b$ also depend on (the same) shared randomness~$r$ chosen
264: according to some distribution~$\Pd_{R}$.  However, there always exists
265: an~$r\in\cR$ such that
266: \begin{align}
267:   \Pr_{RXY}[Q(X,Y,h_a(X,R),h_b(Y,R))]
268:   &=
269:   \E_R\bigl[\Pr_{XY}[Q(X,Y,h_a(X,R),h_b(Y,R))]\bigr]\nonumber\\
270:   &\leq
271:   \Pr_{XY}[Q(X,Y,h_a(X,r),h_b(Y,r))],\label{eq:5}
272: \end{align}
273: and we see that the definition of the value is robust against such a
274: change.  Individual (local) randomness can be obtained from shared
275: randomness and is thus a special case of the above.
276: 
277: \begin{definition}\label{def:nfoldrepetition}
278:   The~\emph{$n$-fold parallel repetition}~$\mkG^{n}$ of a game~$\mkG =
279:   (\Pd_{XY},Q)$ over~$\cX\times\cY\times\cA\times\cB$ is the game over
280:   $\cX^n\times\cY^n\times\cA^n\times\cB^n$ which is given by~$\mkG^{n}
281:   := (\Pd_{X^nY^n}, Q^{\land n})$ where
282:   \begin{align*}
283:     \Pd_{X^nY^n}(x^n,y^n) &:=
284:     \prod_{i=1}^n\Pd_{XY}(x_i,y_i),\text{\quad and}\\
285:     Q^{\land n}(x^n,y^n,a^n,b^n) &:= \bigwedge_{i=1}^n Q(x_i,y_i,a_i,b_i).
286:   \end{align*}
287: \end{definition}  
288: If a strategy is given, the distribution~$\Pd_{X^nY^nA^nB^n}$ of
289: queries and answers is defined in the obvious way.  We further define,
290: for all~$i$, the event~$W_i$ which occurs if the~$i$th subgame is won.
291: \begin{definition}\label{def:randomvars}
292:   For a game~$\mkG^{n}$ and a strategy~$(h_a,h_b)$ the
293:   distribution~$\Pd_{X^n Y^n A^n B^n}$
294:   over~$\cX^n\times\cY^n\times\cA^n \times\cB^n$ is given by
295:   \begin{align*}
296:     \Pd_{X^n Y^n A^n B^n}(x^n, y^n,a^n,b^n)
297:     &:=
298:     \begin{cases}
299:        \Pd_{X^nY^n}(x^n,y^n) & \text{if $h_a(x^n)=a^n$ and
300:         $h_b(y^n) = b^n$}\\
301:       0& \text{otherwise.}\\
302:     \end{cases}
303:   \end{align*}
304:   Further, $W^n$ is the tuple of events~$(W_1,\ldots,W_n)$ where~$W_i
305:   :\iff Q(X_i,Y_i,A_i,B_i)$.
306: \end{definition} 
307: 
308: We prove the following version of the parallel repetition theorem.
309: \begin{theorem}[Parallel Repetition Theorem]\label{thm:pr_local}
310:   For any game~$\mkG$ with value~$v := v(\mkG)$ and any integer~$n$:
311:   \begin{align*}
312:     v(\mkG^{ n}) \leq
313:     \Bigl(1-\frac{(1-v)^3}{6000}\Bigr)^{\frac{n}{\log(|\cA||\cB|)}}.
314:   \end{align*}
315: \end{theorem}
316: The constant 6000 could be improved by a more carful analysis (we will
317: not optimize constants which would improve it during the proof).
318: However, we do not know whether the 3 in the exponent can be reduced.
319: 
320: In \cite{PaRaWi97} it is shown that in Raz's proof the term
321: $\log(|\cA||\cB|)$ in the exponent can be reduced to the maximum of
322: the logarithm of the partition number of $Q(x,y,\cdot,\cdot)$.  As
323: shown by Beame \cite{Beame06}, the argument can be adapted to work
324: with the proof given here.  We give a slightly different argument in
325: Section~\ref{sec:reducingtheexponent} which shows how the term can be
326: reduced to a quantity which is a lower bound on the logarithm of the
327: partition number.
328: 
329: \section{Proof Sketch}
330: Fix an arbitrary game~$\mkG$, its $n$-fold parallel
331: repetition~$\mkG^{n}$, and a strategy~$h_a$,~$h_b$ for~$\mkG^{n}$.
332: With the notation from Definition~\ref{def:randomvars}, the parallel
333: repetition theorem is simply an upper bound on $\Pr[W_1\land\dots\land
334: W_n]$.  To get such an upper bound, we show that for arbitrary
335: indices~$i_1,\ldots,i_m$ there exists an index~$j$ such that
336: \begin{align}
337:   \Pr[W_{j}|W_{i_1}\land\dots\land W_{i_m}]\leq v(\mkG) + \eps, \label{eq:1}
338: \end{align}
339: where $\eps$ depends on~$m$,~$n$, $\log(|\cA||\cB|)$, and
340: $\Pr[W_{i_1}\land\dots\land W_{i_m}]$ (this is
341: Lemma~\ref{lem:gameconditionedvalue}).  From~(\ref{eq:1}) a simple
342: induction gives the parallel repetition theorem, thus we now
343: concentrate on the proof of~(\ref{eq:1}).
344: 
345: \paragraph{Locally Computable Embeddings}
346: In order to prove~(\ref{eq:1}) we define the
347: distribution
348: \begin{align}\label{eq:10}
349:   \Pd_{\Xtilde^n \Ytilde^n} := \Pd_{X^nY^n|W_{i_1}\land\dots\land
350:     W_{i_m}}
351: \end{align}
352: (i.e., the distribution of the message which the referee sends to
353: Alice and Bob conditioned on the event that the games~$i_1$ to~$i_m$
354: are won). 
355: 
356: We show (Lemma~\ref{lem:locallyComputableGame}) that for some~$j$ the
357: following can be achieved by Alice and Bob without communication and
358: using shared randomness only:
359: \begin{enumerate}
360: \item Alice, on input~$x$, produces a tuple~$\xbar^n$
361:   with~$\xbar_j = x$.
362: \item Bob, on input~$y$, produces a tuple~$\ybar^n$ with~$\ybar_j
363:   = y$.
364: \item Let $\Pd_{\Xbar^n\Ybar^n}$ be the resulting joint distribution
365:   of the tuples~$(\xbar^n,\ybar^n)$, assuming that $(x,y)$ is chosen
366:   according to $\Pd_{XY}$.  Then
367:   \begin{align*}
368:     \| \Pd_{\Xbar^n\Ybar^n} - \Pd_{\Xtilde^n\Ytilde^n} \| \leq \eps.
369:   \end{align*}
370: \end{enumerate}
371: We say that~$(X,Y)$ can be $1-\eps$-embedded into
372: $(\Xtilde^n,\Ytilde^n)$ with~$(\Xtilde_j,\Ytilde_j)=(X,Y)$ by local
373: computation.  
374: 
375: If such an embedding is given, we can consider the following strategy
376: for the initial game~$\mkG$: Alice and Bob embed their inputs $(X,Y)$
377: in $(\Xtilde^n,\Ytilde^n)$ with~$(\Xtilde_j,\Ytilde_j)=(X,Y)$, and
378: answer with coordinate~$j$ of~$h_a(\Xtilde^n)$ and~$h_b(\Ytilde^n)$.
379: This strategy wins with probability at
380: least~$\Pr[W_j|W_{i_1}\land\dots\land W_{i_m}]-\eps$.  Since no
381: strategy for the initial game has higher winning probability than
382: $v(\mkG)$ this implies (\ref{eq:1}).
383: 
384: We remark that a necessary condition for such an embedding to exist is
385: that
386: \begin{align}\label{eq:2}
387:   \| \Pd_{XY} - \Pd_{\Xtilde_j \Ytilde_j} \| \leq \eps,
388: \end{align}
389: and indeed this follows from Lemma~\ref{lem:basicrepetitions} for~$U_j
390: = (X_j,Y_j)$ (of course this condition is not a sufficient one).
391: 
392: \paragraph{Constructing an Embedding}
393: We now give a more detailled explanation how Alice and Bob can
394: embed~$(X,Y)$ into~$(\Xtilde^n,\Ytilde^n)$
395: with~$(\Xtilde_j,\Ytilde_j)=(X,Y)$.  For this, given values~$(x,y)$
396: distributed according to~$\Pd_{XY}$, Alice and Bob proceed as follows:
397: \begin{enumerate}
398: \item Alice and Bob use shared randomness to produce queries and
399:   responses for all the won games, i.e., 
400:   values~$(x_{i_1},y_{i_1},a_{i_1},b_{i_1})$
401:   to~$(x_{i_m},y_{i_m},a_{i_m},b_{i_m})$.  Here, Alice and
402:   Bob \emph{both} produce \emph{all} these values.
403: \item For every index~$i \notin \{i_1,\ldots,i_m,j\}$, Alice and Bob
404:   examine a shared random bit~$d_i$.  If~$d_i = 1$ both locally
405:   produce~$x_i$, otherwise both locally produce~$y_i$.  Again, Alice
406:   and Bob both produce all these values.
407: \item Using individual randomness, Alice and Bob locally expand their
408:   information such that Alice gets~$x^n$ and Bob~$y^n$.
409: \end{enumerate}
410: In steps 1 and 2 we have to take care of two things: first, the values
411: produced should be distributed according to the the respective
412: marginal of the distribution
413: $\Pd_{\Atilde^n\Btilde^n\Xtilde^n\Ytilde^n|\Xtilde_j=x \land
414:   \Ytilde_j=y}$ (where $\Pd_{\Atilde^n\Btilde^n\Xtilde^n\Ytilde^n}$ is
415: defined analogously to (\ref{eq:10})).  Second, Alice and Bob should
416: produce \emph{equal values} (otherwise the resulting random
417: variables~$(\Xbar^n,\Ybar^n)$ will not have the correct overall
418: distribution).
419: 
420: For step 1 achieving both is simple: it follows from
421: Corollary~\ref{cor:disjointreps} that Alice and Bob can choose the
422: values~$(x_{i_1},y_{i_1},a_{i_1},b_{i_1}),\ldots,(x_{i_m},y_{i_m},a_{i_m},b_{i_m})$
423: independently of~$(x,y)$ according to
424: $\Pd_{\Xtilde_{i_1}\Ytilde_{i_1}\Atilde_{i_1}\Btilde_{i_1} \cdots
425:   \Xtilde_{i_m}\Ytilde_{i_m}\Atilde_{i_m}\Btilde_{i_m}}$.  Using
426: shared randomness this can be done such that both get the same tuple.
427: 
428: The second step is harder, as in this case the values cannot be chosen
429: independently of~$(x_j,y_j)$ anymore.\footnote{The values~$d_i$ can be
430:   chosen independently, but \emph{not} the values of $x_i$ respective
431:   $y_i$.  We quickly explain why this is impossible in general.
432:   Assume that the random variables~$X$ and~$Y$ contain a shared
433:   bit~$B$.  The game~$\mkG^n$ and the strategy~$(h_a,h_b)$ may be such
434:   that Alice and Bob win subgame~$i_1$ in case~$B_1\oplus\dots\oplus
435:   B_n=0$.  Generating the values independently of~$(x,y)$ would now
436:   produce a distribution with statistical distance at least~$\frac12$
437:   from the target distribution.  Therefore, a bit which is contained
438:   in both $x$ and~$y$ \emph{must} be considered when generating the
439:   values of~$x_i$ and~$y_i$.}  However, let~$\Stilde$ be the random
440: variables which Alice and Bob produce in this step.  It will follow
441: from Corollary~\ref{cor:disjointreps} that
442: $\|\Pd_{XY}\Pd_{\Stilde|\Xtilde_j} - \Pd_{XY\Stilde}\|$ and
443: $\|\Pd_{XY}\Pd_{\Stilde|\Ytilde_j} - \Pd_{XY\Stilde}\|$ are both
444: small, and Lemma~\ref{lem:LocallyComputableCommonPartPre} implies that
445: this is sufficient to generate~$\Stilde$ locally.
446: 
447: In fact, Corollary~\ref{cor:disjointreps}
448: and Lemma~\ref{lem:LocallyComputableCommonPartPre} are strong enough to do
449: steps 1 and 2 at the same time, and thus these steps are done
450: simultaneously in the proof of Lemma~\ref{lem:locallyComputableGame}.
451: 
452: Step 3 will be simpler to implement.  Because the players also
453: computed~$a_{i_\ell}$ and~$b_{i_\ell}$ in step 1, they can expand
454: their known values according to the given distributions and the
455: resulting distribution will be correct (this follows from
456: Lemma~\ref{lem:locallyComputableMarkov}, and a detailed explanation
457: is in the proof of Lemma~\ref{lem:locallyComputableGame}).
458: 
459: 
460: \section{Conditioned Distributions}
461: 
462: The following lemma is essentially Claim~5.1 in Raz's paper
463: \cite{Raz98} (and we use the proof given there).  It states that if
464: random variables $U_i$ are chosen independently, then conditioning on
465: an event does not change the individual distributions a lot on
466: average.
467: 
468: \begin{lemma}\label{lem:basicrepetitions}
469:   Let $\Pd_{U^k} := \Pd_{U_1} \dots \Pd_{U_k}$ be a probability
470:   distribution over~$\cU^k$, $\Wevent$ an event.  Then,
471:   \begin{align}\label{eq:50}
472:     \Pr[W] \leq 2^{-\sum_{j=1}^k (\SD{\Pd_{U_j|\Wevent}}{\Pd_{U_j}})^2}.
473:   \end{align}
474: \end{lemma}
475: As an example, let $U_i$ be uniform and independent bits and~$W$ be
476: the event that at least~$k(\frac{1}{2}+\eps)$ of these bits are one.
477: Then $\|P_{U_i|W} - \Pd_{U_i}\| \geq \eps$ and the lemma states that
478: $\Pr[W] \leq 2^{-k\eps^2}$, which is a version of Chernoff's
479: inequality (note that this implies that
480: Lemma~\ref{lem:basicrepetitions} is almost tight; see, for example,
481: \cite{HolRen06}).
482: 
483: Using $(\sum_{j=1}^k a_j)^2 \leq k \sum_{j=1}^k a_j^2$ one easily checks that
484: \eqref{eq:50} implies
485: \begin{align}\label{eq:52}
486:   \sum_{j=1}^k \| \Pd_{U_j|W} - \Pd_{U_j}\| \leq 
487:   \sqrt{k\log\Bigl(\frac{1}{\Pr[W]}\Bigr)}\;,
488: \end{align}
489: which is the form we use later.
490: 
491: \begin{proof}
492:   For two distributions~$\Pd_{S}$ and~$\Pd_{T}$ over the same
493:   set~$\cS$, the \emph{relative entropy} $D(\Pd_{S}\|\Pd_{T})$ is
494:   defined as
495:   \begin{align}\label{eq:8}
496:     D(\Pd_{S}\|\Pd_{T}) := \sum_{s \in \cS} \Pd_{S}(s)\log\Bigl(\frac{\Pd_{S}(s)}{\Pd_{T}(s)}\Bigr).
497:   \end{align}
498:   This quantity satisfies $D(\Pd_{S}\|\Pd_{T}) \geq
499:   \bigl(\|\Pd_{S}-\Pd_{T}\|\bigr)^2$ (see \cite[Lemma
500:   12.6.1]{CovTho91}).  Also, if~$\Pd_{U^k} = \Pd_{U_1}\dots
501:   \Pd_{U_k}$ and~$\Pd_{V^k}$ are distributions over the set~$\cU^k$,
502:   then $\sum_{j=1}^k D(\Pd_{V_j}\|\Pd_{U_j}) \leq
503:   D(\Pd_{V^k}\|\Pd_{U^k})$ (see Appendix~\ref{app:relentrlemma}).
504:   
505:   Using the above we get
506:   \begin{align*}
507:     \sum_{j=1}^k\Bigl(\SD{\Pd_{U_j|\Wevent}}{\Pd_{U_j}}\Bigr)^2
508:     &\leq \sum_{j=1}^k D(\Pd_{U_j|\Wevent}\|{\Pd_{U_j}})\\
509:     &\leq D(\Pd_{U^k|\Wevent}\|\Pd_{U^k})\displaybreak[2]\\
510:     &= \sum_{u^k} \Pd_{U^k|\Wevent}(u^k)
511:     \log\Bigl(\frac{\Pd_{U^k|\Wevent}(u^k)}{\Pd_{U^k}(u^k)}\Bigr)\\
512:     &= \sum_{u^k} \Pd_{U^k|\Wevent}(u^k)
513:     \log\Bigl(\frac{\Pr[\Wevent|U^k=u^k]}{\Pr[\Wevent]}\Bigr)\\
514:     &= \log\Bigl(\frac{1}{\Pr[\Wevent]}\Bigr) +
515:      \sum_{u^k} \Pd_{U^k|\Wevent}(u^k)
516:     \log\bigl(\Pr[\Wevent|U^k=u^k]\bigr)\\
517:     &\leq \log\Bigl(\frac{1}{\Pr[\Wevent]}\Bigr).\qedhere
518:   \end{align*}
519: \end{proof}
520: 
521: We now give a slight extension of this lemma (this makes it simpler to
522: apply later).  First, the~$U_j$ are independent given the value of an
523: additional random variable~$T$.  Second, an arbitrary third random
524: variable~$V$ with bounded alphabet size gives side information
525: about~$U_j$.  Then, choosing~$U_j$ without considering the fact that
526: an event~$W$ happened and ignoring~$V$ does not change the
527: distribution of~$U_j$ too much on average.  For the notation in the
528: following corollary we refer to Section~\ref{sec:probdistr}, equation
529: (\ref{eq:29}) and the subsequent remarks.
530: \begin{corollary}\label{cor:disjointreps}
531:   Let~$\Pd_{T U^k V} := \Pd_{T}
532:   \Pd_{U_1|T}\Pd_{U_2|T}\dots\Pd_{U_k|T} \Pd_{V|T U^k}$
533:   be a probability distribution over~$\cT\times\cU^k \times \cV$,~$\Wevent$
534:   be an event. Then,
535:   \begin{align*}
536:     \sum_{j=1}^k 
537:     \Bigl\| \Pd_{T U_j V|\Wevent} - 
538:     \Pd_{T V|\Wevent}\Pd_{U_j|T} \Bigr\|
539:     &\leq \sqrt{k} \sqrt{\log(|\cV^*|) +
540:       \log\Bigl(\frac{1}{\Pr[\Wevent]}\Bigr)},
541:   \end{align*}
542:   where $\cV^* := \{ v \in \cV | \Pd_{V|W}(v) > 0\}$.
543: \end{corollary}
544: The proof is essentially an application of Jensen's inequality on
545: Lemma~\ref{lem:basicrepetitions}.
546: \begin{proof}
547:   Fix a pair $(t,v)\in\cT\times\cV$ and consider the
548:   distributions~$\Pd_{U^k|T=t,V=v,\Wevent}$ and $\Pd_{U^k|T=t}$.  We
549:   apply Lemma~\ref{lem:basicrepetitions} (in the form given by
550:   \eqref{eq:52}) on these distributions (with the
551:   event~$(V\mathord{=}v) \land \Wevent$) and get
552:   \begin{align}
553:     \sum_{j=1}^k 
554:     \Bigl\| \Pd_{T U_j V|\Wevent} \!-\! 
555:     \Pd_{T V|\Wevent}\Pd_{U_j|T} \Bigr\|
556:     &=\sum_{t \in \cT,v\in\cV^*} \Pd_{TV|\Wevent}(t,v) \cdot
557:     \sum_{j=1}^k \Bigl\| \Pd_{U_j|T=t, V=v,\Wevent} - \Pd_{U_j|T=t} \Bigr\|\nonumber\\
558:     &\leq
559:     \sum_{t \in \cT,v\in\cV^*} \Pd_{TV|\Wevent}(t,v)
560:     \sqrt{k \log\Bigl(\frac{1}{\Pr[\Wevent\land V=v|T=t]}\Bigr)}\nonumber\\
561:     &\leq
562:     \sqrt{k\log\Bigl(\sum_{t \in \cT,v\in\cV^*} \Pd_{TV|\Wevent}(t,v)
563:       \frac{1}{\Pr[\Wevent\land V=v|T=t]}\Bigr)},\label{eq:17}
564:   \end{align}
565:   where the last inequality is Jensen's inequality applied
566:   on the function~$\sqrt{\log(\cdot)}$ which is concave on~$[1,\infty)$.  We compute
567:   \begin{align}
568:     \sum_{t \in \cT,v\in\cV^*}\!\!\!\!\!\!
569:     \Pd_{TV|\Wevent}(t,v) \frac{1}{\Pr[W\land V=v|T=t]} 
570:     &=
571:     \sum_{t \in \cT,v\in\cV^*} \frac{\Pr[{T=t} \land
572:       {V=v}|\Wevent]}{\Pr[\Wevent\land V=v|T=t]} \nonumber\\
573:     &=
574:     \sum_{t \in \cT,v\in\cV^*} 
575:     \frac{\Pr[{T=t} \land {V=v} \land \Wevent] \Pr[T=t]}{\Pr[\Wevent]
576:       \Pr[V=v \land T=t \land \Wevent]} \nonumber\\
577:     &=
578:     \sum_{t \in \cT,v\in\cV^*}
579:     \frac{\Pr[T=t]}{\Pr[\Wevent]}
580:     =
581:     \frac{|\cV^*|}{\Pr[\Wevent]}.\nonumber
582:   \end{align}
583:   Inserting this into (\ref{eq:17}) completes the proof.
584: \end{proof}
585: 
586: \section{Embedding by Local Computation}
587: We next study under what conditions random variables can be embedded
588: into other random variables by local computations.
589: \begin{definition}[Embeddable]
590:   For two distributions~$\Pd_{X_0 Y_0}$ and $\Pd_{X_1 S
591:     Y_1 T}$ we say that $(X_0, Y_0)$ is $1-\eps$-embeddable in
592:   $(X_1 S, Y_1 T)$ with~$(X_1,Y_1)=(X_0,Y_0)$ if 
593:   there exists a probability measure~$\Pd_{R}$ over a set~$\cR$ and
594:   functions~$f_A: \cX\times\cR\rightarrow\cS$, $f_B:\cY\times\cR
595:   \rightarrow \cT$, such that
596:   \begin{align*}
597:     \bigl\| \Pd_{X_0Y_0}\Pd_{F_A F_B|XY}
598:     - \Pd_{X_1Y_1S T}\bigr\| \leq \eps,
599:   \end{align*}
600:   where~$\Pd_{F_A F_B|X=xY=y}$ is the distribution defined by the
601:   random variable~$(f_A(x,R),\n f_B(y,R))$.
602: \end{definition}
603: The following lemma gives a condition under which $(X,Y)$ is
604: embeddable in~$(XS,\n YS)$.  It is one of the main contributions of
605: this paper.
606: \begin{lemma}\label{lem:LocallyComputableCommonPartPre}
607:   Let a distribution~$\Pd_{SXY}$ be given.  If 
608:   \begin{align}
609:     \| \Pd_{SXY}-\Pd_{X Y}\Pd_{S|X}\| \leq \eps_1 \label{eq:12}
610:     \intertext{and}
611:     \| \Pd_{SXY}-\Pd_{X Y}\Pd_{S|Y}\| \leq \eps_2 ,\label{eq:13}
612:   \end{align}
613:   then~$(X, Y)$ is $1-2\eps_1-2\eps_2$-embeddable\footnote{It is
614:     understood that the embedding satisfies $(X,Y)=(X,Y)$, i.e., that
615:     the original random variables will result if from the resulting
616:     $(XS,YS)$ the $S$-part is omitted.}
617:   in~$(XS, YS)$.
618: \end{lemma}
619: 
620: Even if~$\eps_1=\eps_2=0$, equations (\ref{eq:12}) and (\ref{eq:13})
621: do not imply that~$S$ is independent of~$X$ and~$Y$. For example, if
622: $X$ and~$Y$ contain the
623: same uniform random bit, then $S$ can depend on this bit. 
624: However, if~$\eps_1=\eps_2 = 0$ the lemma is obviously true: Alice
625: uses shared randomness to choose~$S$ according to~$\Pd_{S|X=x}$ (more
626: concretely: Alice chooses a uniform random real~$\rho \in [0,1]$ and
627: uses the smallest element~$s$ for which the cumulative distribution
628: function~$\sum_{s'\leq s}\Pd_{S|X=x}(s')$ is larger than~$\rho$).
629: Since Bob has the same distribution~$\Pd_{S|Y=y}$ he will find the
630: same value if he uses the same shared randomness.
631: 
632: In case~$\eps_1>0$ and~$\eps_2>0$, we have to overcome the following
633: problem: $\Pd_{S|Y=y}$ is unknown to Alice (since~$y$ is unknown to
634: Alice), and analogously $\Pd_{S|X=x}$ is unknown to Bob.  The solution
635: is to define the function $f_A: \cX \times \cR \rightarrow \cS$ with
636: the following process: Alice chooses, using shared randomness, a
637: uniform random element~$s$ from $\cS$ and a uniform random real
638: number~$\rho\in[0,1]$.  If $\Pd_{S|X=x}(s)>\rho$ she outputs~$s$,
639: otherwise Alice repeats the above.  The function~$f_B: \cY \times \cR
640: \rightarrow \cS$ is defined by the analogous process given~$y$.  It is
641: easy to see that Alice outputs elements according to the
642: distribution~$\Pd_{S|X=x}$, Bob according to~$\Pd_{S|Y=y}$.  We
643: further show that usually the output of~$f_A$ is equal to the output
644: of~$f_B$.
645: \begin{proof}
646:   Let~$\cR := (\cS\times[0,1])^\infty$ be the set of infinite
647:   sequences over~$\cS\times[0,1]$.  For a fixed~$x,y$ and a sequence $r
648:   := \{(s_i,\rho_i)\}_{i\geq 0}$, we define~$f_A(x,r) := s_i$ if~$i$
649:   is the smallest index for which $\Pd_{S|X=x}(s_i) > \rho_i$.
650:   Analogously, $f_B(y,r) := s_j$ if~$j$ is the smallest index
651:   with~$\Pd_{S|Y=y}(s_j)>\rho_j$ and\footnote{The use of~$f_{AB}$ in
652:     order to simplify the analysis was suggested by Anup Rao.}
653:   $f_{AB}(x,y,r) := s_k$ if~$k$ is the smallest index
654:   with~$\Pd_{S|X=xY=y}(s_k)>\rho_k$.  If no such index exist the
655:   respective function is defined in an arbitrary way (this happens
656:   with probability~$0$).
657:   
658:   Let $\Pd_{XYF_AF_BF_{AB}}$ be the joint distribution
659:   of~$(x,y,f_A(x,r), f_B(y,r),f_{AB}(x,y,r))$ where $(x,y)$ is chosen
660:   according to $\Pd_{XY}$ and $r$ uniformly from $\cR$.  We have
661:   $\Pd_{F_{AB}|X=xY=y}=\Pd_{S|X=xY=y}$, $\Pd_{F_A|X=x}=\Pd_{S|X=x}$
662:   and $\Pd_{F_B|Y=y}=\Pd_{S|Y=y}$, since these equalities hold
663:   conditioned on the event that the respective function accepts in 
664:   round~$i$, for any fixed~$i$.
665:   
666:   Further, we have $\Pr[F_A=F_{AB}|X=x,Y=y] \geq
667:   1-2\|\Pd_{F_{A}|X=x}-\Pd_{F_{AB}|X=xY=y}\|$: the two values
668:   $F_A,F_{AB}$ are equal if $\rho_j <
669:   \min(\Pd_{F_{A}|X=x}(s_j),\Pd_{F_{AB}|X=xY=y}(s_j))$ for the
670:   smallest $j$ for which $\rho_j <
671:   \max(\Pd_{F_{A}|X=x}(s_j),\Pd_{F_{AB}|X=xY=y}(s_j))$ is satisfied. 
672:   This happens with probability
673:   \begin{align*}
674:     \frac{\sum_{s}\min(\Pd_{F_{A}|X=x}(s_j),\Pd_{F_{AB}|X=xY=y}(s_j))}
675:     {\sum_{s}\max(\Pd_{F_{A}|X=x}(s_j),\Pd_{F_{AB}|X=xY=y}(s_j))}
676:     &=\frac{1-\|\Pd_{F_{A}|X=x}-\Pd_{F_{AB}|X=xY=y}\|}{1+\|\Pd_{F_{A}|X=x}-\Pd_{F_{AB}|X=xY=y}\|}\\
677:     &\geq 1-2\|\Pd_{F_{A}|X=x}-\Pd_{F_{AB}|X=xY=y}\|.
678:   \end{align*}
679:   This yields $\Pr[F_A=F_{AB}] \geq 1-2\eps_1$, and analogously we get
680:   $\Pr[F_B=F_{AB}] \geq 1-2\eps_2$, and thus $\Pr[F_A=F_B=F_{AB}]\geq
681:   1-2\eps_1-2\eps_2$.  
682:   This implies
683:   \begin{align*}
684:     \|\Pd_{XYSS} - \Pd_{XY}\Pd_{F_AF_B|XY}\|
685:     &=
686:     \|\Pd_{XYF_{AB}F_{AB}} - \Pd_{XYF_AF_B}\| \geq 1-2\eps_1-2\eps_2.\qedhere
687:   \end{align*}
688: \end{proof}
689: 
690: In the following corollary, the input distribution is changed
691: slightly.  This makes it a bit easier to apply later.
692: \begin{corollary}\label{cor:LocallyComputableCommonPart}
693:   Let distributions~$\Pd_{SXY}$ and~$\Pd_{X_0 Y_0}$ be given.  If 
694:   \begin{align}
695:     \| \Pd_{SXY}-\Pd_{X_0 Y_0}\Pd_{S|X}\| \leq \eps_1 \label{eq:48}
696:     \intertext{and}
697:     \| \Pd_{SXY}-\Pd_{X_0 Y_0}\Pd_{S|Y}\| \leq \eps_2 ,\label{eq:49}
698:   \end{align}
699:   then~$(X_0, Y_0)$ is $1-3\eps_1-2\eps_2$-embeddable\footnote{The
700:     statement could be made symmetric (i.e., $(X_0,Y_0)$ is
701:     $1-2\eps_1-2\eps_2 - \min(\eps_1,\eps_2)$-embeddable).}
702:   in~$(XS, YS)$ with~$(X,Y)=(X_0,Y_0)$.
703: \end{corollary}
704: \begin{proof}
705:   From (\ref{eq:48}) we get $\| \Pd_{XY}-\Pd_{X_0Y_0}\|\leq\eps_1$.
706:   One can now find a joint distribution $\Pd_{XYX_0Y_0}$ with
707:   $\Pr[(X,Y)=(X_0,Y_0)] \geq 1-\eps_1$.  The corollary now follows by
708:   applying $f_A$ and~$f_B$ from
709:   Lemma~\ref{lem:LocallyComputableCommonPartPre}.
710: \end{proof}
711: 
712: Random variables~$S,T,U$ form a Markov chain,
713: written~$S\leftrightarrow T\leftrightarrow U$ if $\Pd_{STU} =
714: \Pd_{T}\Pd_{S|T}\Pd_{U|T}$ (i.e., if given~$T$ the probability
715: distribution of $U$ does not depend on~$S$).  The following lemma is
716: essentially Lemma~4.1 in Raz's paper.
717: 
718: \begin{lemma}\label{lem:locallyComputableMarkov}
719:   Let~$\Pd_{XYST}$ be any distribution. If 
720:   \begin{align*}
721:     S \leftrightarrow X \leftrightarrow YT 
722:     \intertext{and}
723:     XS \leftrightarrow Y \leftrightarrow T 
724:   \end{align*}
725:   then~$(X,Y)$ is $1$-embeddable in~$(XS,YT)$.
726: \end{lemma}
727: \begin{proof}
728:   Using individual (non-shared) randomness, Alice computes~$S$
729:   according to~$\Pd_{S|X=x}$ and Bob computes~$T$ according
730:   to~$\Pd_{T|Y=y}$.  Since
731:   \begin{align}
732:     \Pd_{STXY} = \Pd_{XY}\Pd_{S|XY}\Pd_{T|SXY} = \Pd_{XY}\Pd_{S|X}\Pd_{T|Y}
733:   \end{align}
734:   this gives the correct (global) distribution.
735: \end{proof}
736: 
737: 
738: \section{Embeddings for Games}
739: Given a game~$\mkG$ and its~$n$-fold parallel repetition, we now show
740: that $(X,Y)$ can be embedded into~$(\Xtilde^n,\Ytilde^n)$,
741: where~$\Pd_{\Xtilde^n\Ytilde^n} := \Pd_{X^nY^n|W_{k+1}\land\dots\land W_n}$.
742: 
743: We need the following simple fact on statistical distance.
744: \begin{fact}\label{fact:statdistsplit}
745:   Let~$\Pd_{Z_0}$ and~$\Pd_{Z_1}$ be distributions over~$\cZ$.
746:   Let $\cS \subseteq \cZ$ be such that $\Pr[Z_0 \in \cS] = \Pr[Z_1 \in
747:   \cS] = \frac{1}{2}$.
748:   Then, 
749:   \begin{align*}
750:     \| \Pd_{Z_0|Z_0 \in \cS} - \Pd_{Z_1|Z_1\in\cS}\|
751:     \leq 
752:     2\| \Pd_{Z_0} - \Pd_{Z_1}\|\, .
753:   \end{align*}
754: \end{fact}
755: 
756: Also, we need the following statements about Markov chains.
757: 
758: \begin{claim}\label{claim:hmm}
759:   Let~$\Pd_{X_0 Y_0}\Pd_{X_1Y_1}$ be a distribution
760:   over~$\cX_0\times\cY_0\times\cX_1\times\cY_1$, $f:
761:   \cX_0\times\cX_1\rightarrow \cU$ and~$g:
762:   \cY_0\times\cY_1\rightarrow\cV$ be arbitrary.  Then,
763:   \begin{align}\label{eq:22}
764:     X_0 X_1 \leftrightarrow X_0 f(X_0,X_1) Y_1 g(Y_0,Y_1)
765:     \leftrightarrow Y_0 Y_1.
766:   \end{align}
767: \end{claim}
768: \begin{proof}
769:   It is sufficient to show this for all possible values~$x_0\in\cX_0$
770:   and~$y_1\in\cY_1$.  Let~$\Pd_{\Ytilde_0\Xtilde_1} :=
771:   \Pd_{Y_0X_1|X_0=x_0 Y_1=y_1} = \Pd_{Y_0|X_0=x_0}\Pd_{X_1|Y_1=y_1}$.
772:   In this case, (\ref{eq:22}) reduces to
773:   \begin{align*}
774:     \Xtilde_1 \leftrightarrow
775:     f(x_0,\Xtilde_1) g(\Ytilde_0,y_1)\leftrightarrow \Ytilde_0.
776:   \end{align*}
777:   Since $\Xtilde_1$ and~$\Ytilde_0$ are independent this is obvious.
778: \end{proof}
779: 
780: \begin{claim}\label{claim:markovcondition}
781:   Let~$\Pd_{TUV}$ be a distribution
782:   over~$\cT\times\cU\times\cV$ and $W$ an event with
783:   \begin{align*}
784:     T &\leftrightarrow U \leftrightarrow V,\\
785:     W &\leftrightarrow U \leftrightarrow TV.
786:   \end{align*}
787:   Then, for~$\Pd_{\Ttilde\Utilde\Vtilde} := \Pd_{TUV|W}$ we have
788:   \begin{align*}
789:     \Ttilde\leftrightarrow \Utilde \leftrightarrow \Vtilde.
790:   \end{align*}
791: \end{claim}
792: \begin{proof}
793:   \begin{align*}
794:     \Pd_{\Ttilde\Utilde\Vtilde}(t,u,v) &= \Pd_{TUV|W}(t,u,v)\\
795:     &=\Pd_{U|W}(u)\Pd_{TV|U=u,W}(t,v)\\
796:     &=\Pd_{U|W}(u)\Pd_{TV|U=u}(t,v)\\
797:     &=\Pd_{U|W}(u)\Pd_{T|U=u}(t)\Pd_{V|U=u}(v)\\
798:     &=\Pd_{U|W}(u)\Pd_{T|U=u,W}(t)\Pd_{V|U=u,W}(v).\qedhere
799:   \end{align*}
800: \end{proof}
801: 
802: \begin{lemma}\label{lem:locallyComputableGame}
803:   Let a game $\mkG^{ n}=(Q^n, (\Pd_{XY})^n)$, a
804:   strategy $(h_a,h_b)$, and $k \leq n$ be given.  Let
805:   \begin{align*}
806:     \Pd_{\Xtilde^n\Ytilde^n} :=
807:     \Pd_{X^nY^n|W_{k+1}\land\dots\land W_n}
808:   \end{align*}
809:   
810:   Then, for~$1 \leq j \leq k$, there exists~$\eps_j \geq 0$ such
811:   that~$(X,Y)$ is~$1-\eps_j$-embeddable in~$(\Xtilde^n,\Ytilde^n)$
812:   with~$(\Xtilde_j,\Ytilde_j)=(X,Y)$ and
813:   \begin{align}
814:     \sum_{j=1}^{k}
815:     \eps_j \leq 
816:     15\sqrt{k}\sqrt{(n-k)\log(|\cA|\,|\cB|)+
817:       \log\Bigl(\frac{1}{\Pr[W_{k+1}\land\dots\land W_n]}\Bigr)}.\label{eq:27}
818:   \end{align}
819: \end{lemma}
820: \begin{proof}
821:   As described in Definition~\ref{def:randomvars} we consider the
822:   distribution~$\Pd_{X^nY^nA^nB^nW^n}$ and the corresponding random
823:   variables.  Additionally, we let~$D_{1}, \ldots, D_{k}$ be uniform
824:   and independent bits.  For~$1\leq j\leq k$ we define
825:   \begin{align*}
826:     U_j &:= \begin{cases}
827:       X_j & \text{if $D_j = 0$}\\
828:       Y_j & \text{otherwise}
829:     \end{cases}\\
830:     \intertext{and}
831:     \Ubar_j &:= \begin{cases}
832:       Y_j & \text{if $D_j = 0$}\\
833:       X_j & \text{otherwise.}
834:     \end{cases}
835:   \end{align*}
836:   Also, we set
837:   \begin{align}
838:     T &:= (X_{k+1},\ldots,X_{n}, Y_{k+1},\ldots,Y_n, D^k, \Ubar^k),\label{eq:44}\\
839:     V &:=      (A_{k+1},\ldots,A_{n},B_{k+1},\ldots,B_n),\label{eq:45}
840:   \end{align}
841:   and define the event~$W := W_{k+1}\land\dots \land W_n$.
842:   
843:   From Corollary~\ref{cor:disjointreps}  we get
844:   \begin{align}
845:     \sum_{j=1}^k 
846:     \Bigl\| \Pd_{T U_jV|W} - 
847:     \Pd_{T V|W}\Pd_{U_j|T} \Bigr\|
848:     &\leq \eps_{\Tot}\label{eq:14}\,,
849:   \end{align}
850:   where we set \begin{align} \eps_{\Tot} := \sqrt{k}
851:     \sqrt{(n-k)\log(|\cA||\cB|)+
852:       \log\Bigl(\frac{1}{\Pr[W]}\Bigr)}
853:   \end{align}
854:   (we applied Corollary~\ref{cor:disjointreps} using~$|\cV^*| \leq |\cV|$).
855: 
856:   In (\ref{eq:14}), we condition on both sides on the event~$D_j=0$,
857:   which is, on both sides, a restriction on a subset which has
858:   probability~$\frac{1}{2}$.  Fact~\ref{fact:statdistsplit} implies
859:   \begin{align}
860:     \sum_{j=1}^k 
861:     \Bigl\| \Pd_{T U_jV|W \land (D_j=0)} - 
862:     \Pd_{T V|W\land (D_j=0)}\Pd_{U_j|T} \Bigr\|
863:     &\leq 2\eps_{\Tot}\,,\label{eq:19}
864:   \end{align}
865:   where we do not need to condition on~$D_j=0$ in~$\Pd_{U_j|T}$ since
866:   this is included in the given~$t$ anyhow; in fact we can now
867:   write~$\Pd_{X_j|Y_j}$ instead of~$\Pd_{U_j|T}$.
868: 
869:   For a fixed~$j$, define the random variable
870:   \begin{align}
871:     T^\NoJ &:= 
872:     (X_{k+1},\ldots,X_{n}, Y_{k+1},\ldots,Y_n,\nonumber\\
873:     &\qquad\qquad D_1,\ldots,D_{j-1},D_{j+1},\ldots,D_k, \nonumber\\
874:     &\qquad\qquad\Ubar_1,\ldots,\Ubar_{j-1},\Ubar_{j+1},\ldots,\Ubar_{k}).
875:   \end{align}
876:   With this notation (\ref{eq:19}) is equivalent to
877:   \begin{align}
878:     \sum_{j=1}^k 
879:     \Bigl\| \Pd_{T^{\NoJ}X_j Y_j V|W \land (D_j=0)} - 
880:     \Pd_{T^{\NoJ} Y_j V|W \land (D_j=0)}\Pd_{X_j|Y_j} \Bigr\|
881:     &\leq 2\eps_{\Tot}\,.
882:   \end{align}
883:   But now nothing depends on~$D_j = 0$ anymore, so this also means
884:   \begin{align}\label{eq:4}
885:     \sum_{j=1}^k 
886:     \Bigl\| \Pd_{T^{\NoJ} X_j Y_j V|W} - 
887:     \Pd_{T^{\NoJ} Y_j V|W}\Pd_{X_j|Y_j} \Bigr\|
888:     &\leq 2\eps_{\Tot}\,.
889:   \end{align}
890:   We set~$S := (T^\NoJ,V)$ and define the probability distribution
891:   \begin{align}
892:     \Pd_{\Stilde \Xtilde^n \Ytilde^n} := \Pd_{S X^nY^n|W}.
893:   \end{align}
894:   With this, (\ref{eq:4}) becomes
895:   \begin{align}
896:     \sum_{j=1}^k 
897:     \Bigl\| \Pd_{\Stilde \Xtilde_j \Ytilde_j} - 
898:     \Pd_{\Stilde \Ytilde_j}\Pd_{X_j|Y_j} \Bigr\|
899:     &\leq 2\eps_{\Tot}\,,
900:   \end{align}
901:   or, equivalently
902:   \begin{align}
903:     \sum_{j=1}^k 
904:     \Bigl\| \Pd_{\Stilde \Xtilde_j \Ytilde_j} - 
905:     \Pd_{\Ytilde_j}\Pd_{\Stilde|\Ytilde_j}\Pd_{X|Y} \Bigr\|
906:     &\leq 2\eps_{\Tot}\,.
907:   \end{align}
908:   Lemma~\ref{lem:basicrepetitions} implies 
909:   \begin{align}
910:     \sum_{j=1}^k 
911:     \Bigl\| \Pd_{\Ytilde_j} - \Pd_{Y} \Bigl\| \leq \eps_{\Tot},
912:   \end{align}
913:   and thus
914:   \begin{align}\label{eq:25}
915:     \sum_{j=1}^k 
916:     \Bigl\| \Pd_{\Stilde \Xtilde_j \Ytilde_j} - 
917:     \Pd_{XY}\Pd_{\Stilde|\Ytilde_j} \Bigr\|
918:     &\leq 3\eps_{\Tot}\,.
919:   \end{align}
920:   Symmetric reasoning yields
921:   \begin{align}\label{eq:26}
922:     \sum_{j=1}^k 
923:     \Bigl\| \Pd_{\Stilde \Xtilde_j \Ytilde_j} - 
924:     \Pd_{XY}\Pd_{\Stilde|\Xtilde_j} \Bigr\|
925:     &\leq 3\eps_{\Tot}\,.
926:   \end{align}
927:   From~(\ref{eq:25}) and (\ref{eq:26}),
928:   Corollary~\ref{cor:LocallyComputableCommonPart} implies that~$(X,Y)$
929:   is $1-\eps_j$-embeddable in $(\Xtilde_j\Stilde,\Ytilde_j\Stilde)$
930:   with $(\Xtilde_j,\Ytilde_j)=(X,Y)$ and such that $\sum_{j=1}^k
931:   {\eps_j} \leq 15\eps_{\Tot}$.
932: 
933:   We next show that
934:   \begin{align}\label{eq:28}
935:     X^k \leftrightarrow T V \leftrightarrow Y^k.
936:   \end{align}
937:   If the bits~$D^k$ and the values $X_{k+1},\ldots,X_n$,
938:   $Y_{k+1},\ldots,Y_n$ are fixed, this follows immediately from
939:   Claim~\ref{claim:hmm}.  Since it holds for all these values it must
940:   also hold overall.
941: 
942:   From (\ref{eq:28}) we easily get
943:   \begin{align*}
944:     X^n \leftrightarrow X_j S \leftrightarrow
945:     Y^n Y_j S\\
946:     X^n X_j S \leftrightarrow Y_j S \leftrightarrow
947:     Y^n.
948:   \end{align*}
949:   Claim~\ref{claim:markovcondition} yields
950:   \begin{align}
951:     \Xtilde^n \leftrightarrow \Xtilde_j \Stilde \leftrightarrow
952:     \Ytilde^n\Ytilde_j\Stilde \label{eq:20}\\
953:     \Xtilde^n\Xtilde_j\Stilde \leftrightarrow \Ytilde_j\Stilde \leftrightarrow
954:     \Ytilde^n.\label{eq:21}
955:   \end{align}
956:   
957:   Above we have seen that $(X,Y)$ is embeddable
958:   in~$(\Xtilde_j\Stilde,\Ytilde_j\Stilde)$
959:   with~$(\Xtilde_j,\Ytilde_j)=(X,Y)$.  
960:   Lemma~\ref{lem:locallyComputableMarkov} together with (\ref{eq:20})
961:   and (\ref{eq:21}) now implies that we can $1$-locally embed this in
962:   $(\Xtilde^n\Xtilde_j\Stilde, \Ytilde^n\Ytilde_j\Stilde)$.  Since
963:   Alice and Bob can then ignore part of the constructed information
964:   this completes the proof.
965: \end{proof}
966: 
967: \begin{lemma}\label{lem:gameconditionedvalue}
968:   Let a game~$\mkG=(Q,\Pd_{XY})$, its~$n$-fold
969:   repetition~$\mkG^{n}$, and a strategy $(h_a,h_b)$
970:   for~$\mkG^{n}$ be given.  Let indices~$i_1,\ldots,i_m$ be
971:   given.  Then, there exists an index~$i_{m+1}$ such
972:   that
973:   \begin{align}\label{eq:3}
974:     \Pr[W_{i_{m+1}} | &W_{i_1}\land\dots\land W_{i_m}] \nonumber \\
975:     &\leq v(\mkG) +
976:     15\sqrt{\frac{1}{n-m}}\sqrt{m\log(|\cA||\cB|) +
977:       \log\Bigl(\frac{1}{\Pr[W_{i_1}\land\dots\land W_{i_m}]}\Bigr)}.
978:   \end{align}
979: \end{lemma}
980: \begin{proof}
981:   First, we can assume that the given indices $i_\ell$, $1\leq\ell\leq
982:   m$, are pairwise different (otherwise we get a stronger statement).
983:   Given this we can even assume that~$i_\ell = n-\ell+1$ by
984:   appropriately redefining the functions~$(h_a,h_b)$.
985:   
986:   Define the distribution~$\Pd_{\Xtilde^n \Ytilde^n} := \Pd_{X^n Y^n|
987:     W_{n-m+1}\land\cdots\land W_{n}}$.
988:   Lemma~\ref{lem:locallyComputableGame} implies that there exists an
989:   index $j$ such that~$(X,Y)$ is~$1-\eps$-embeddable
990:   in~$(\Xtilde^n,\Ytilde^n)$ with $(\Xtilde_j,\Ytilde_j)=(X,Y)$ and
991:   \begin{align*}
992:     \eps := 15\sqrt{\frac{1}{n-m}}\sqrt{m\log(|\cA||\cB|)+
993:       \log\Bigl(\frac{1}{\Pr[W_{n-m+1}\land\dots\land W_{n}]}\Bigr)}.
994:   \end{align*}
995:   Consider the following strategy for~$\mkG$.  On input~$(X,Y)$ Alice
996:   and Bob~$1-\eps$-embed this into~$(\Xtilde^n,\Ytilde^n)$
997:   with~$(\Xtilde_j,\Ytilde_j)=(X,Y)$.  Since the resulting
998:   distribution has statistical distance at most~$\eps$
999:   from~$\Pd_{\Xtilde^n \Ytilde^n}$, if they output coordinate~$j$
1000:   of~$h_a(\Xtilde^n)$ and~$h_b(\Ytilde^n)$ they have probability at
1001:   least~$\Pr[W_j | W_{n-m+1}\land\dots\land W_{n}]-\eps$ to win the
1002:   initial game.  The shared randomness can be eliminated (see the
1003:   remark after Definition~\ref{def:game}), and thus
1004:   \begin{align*}
1005:     v(\mkG) &\geq \Pr[W_j | W_{n-m+1}\land\dots\land
1006:     W_{n}]-\eps.\qedhere
1007:   \end{align*}
1008: \end{proof}                                
1009: 
1010: 
1011: \section{Parallel Repetition Theorem}
1012: \begin{proof}[Proof (of Theorem~\ref{thm:pr_local})]
1013:   Fix a strategy~$(h_a,h_b)$ for~$\mkG^n$.  Then, repeatedly choose
1014:   the index~$i_{m+1}$ for which~$\Pr[W_{i_{m+1}}|W_{i_1}\land \dots
1015:   \land W_{i_{m}}]$ is minimized.  We set $p_0 := 1$ and $p_m :=
1016:   \Pr[W_{i_1}\land \dots \land W_{i_{m}}]$.
1017:   Lemma~\ref{lem:gameconditionedvalue} implies
1018:   \begin{align}
1019:     p_{m+1} \leq  p_{m} \cdot \biggl(v+15\sqrt{\frac{1}{n-m}}\sqrt{m\log(|\cA||\cB|) +
1020:       \log\Bigl(\frac{1}{p_{m}}\Bigr)}\biggr).\label{eq:33}
1021:   \end{align}
1022:   We show per induction that
1023:   \begin{align*}
1024:     p_m \leq \Bigl(\frac{1+v}{2}\Bigr)^{m},
1025:   \end{align*}
1026:   as long as $m \leq \frac{(1-v)^2 (n-m)}{2700\log(|\cA||\cB|)}$.  The
1027:   statement holds for~$m=0$ and we now make a step from~$m$ to~$m+1$.
1028:   First, we can assume that~$p_{m} \geq
1029:   \bigl(\frac{1+v}{2}\bigr)^{m+1} > \frac{1}{2}^{m+1}$, as otherwise
1030:   the induction step is trivial.  In this case, \eqref{eq:33} yields
1031:   \begin{align}
1032:     p_{m+1} 
1033:     &\leq  
1034:     p_{m} \cdot \biggl(v+15\sqrt{\frac{1}{n-m}}\sqrt{m\log(|\cA||\cB|) +
1035:       (m+1)}\biggr)\nonumber\\
1036:     & \leq
1037:     p_{m} \cdot \biggl(v+\sqrt{\frac{1}{n-m}}\sqrt{675 m\log(|\cA||\cB|)}\biggr)\label{eq:35}    
1038:   \end{align}
1039:   Since we assume~$m \leq \frac{(1-v)^2}{2700
1040:     \log(|\cA||\cB|)}(n-m)$ this proves the induction step.
1041: 
1042:   In total we get for~$m = \frac{n(1-v)^2}{3000\log(|\cA||\cB|)}$
1043:   \begin{align}\label{eq:38}
1044:     p_{m} \leq \Bigl(\frac{1+v}{2}\Bigr)^\frac{n(1-v)^2}{3000\log(|\cA||\cB|)}.
1045:   \end{align}
1046:   We have
1047:   \begin{align}
1048:     \Bigl(\frac{1+v}{2}\Bigr)^{\frac{(1-v)^2}{3000}}
1049:     &= \Bigl(1-\frac{1-v}{2}\Bigr)^{\frac{(1-v)^2}{3000}} \nonumber\\
1050:     &\leq 1-\frac{(1-v)^3}{6000},\label{eq:39}
1051:   \end{align}
1052:   where the last inequality follows from 
1053:   $(1-b)^a \leq 1-ab$ which holds for all $a \in [0,1]$, $b\leq1$. %%
1054: %%Formelbuch s. 50 oben
1055:   Since~$\Pr[W_1\land\dots\land W_n] \leq p_m$, (\ref{eq:38}) and
1056:   (\ref{eq:39}) imply the theorem.\footnote{The minimal value of the
1057:     sequence defined by $p_0 := 1$ and $p_{m+1} := p_m
1058:     \bigl(v+\sqrt{\frac{225}{n-m}}\sqrt{m \ell + \log(1/p_m)}\bigr)$
1059:     is indeed $\Bigl(1-\Theta((1-v)^3)\Bigr)^{\frac{n}{\ell}}$.  The
1060:     argument in the proof above shows that the minimal value can 
1061:     only be lower. On the other hand, the sequence given by $p'_0 := 1$,
1062:     $p'_{m+1} := p'_m\Bigl(v+\sqrt{\frac{m\ell}{n}}\Bigr)$
1063:     is strictly smaller than the sequence $\{p_j\}_{j\geq0}$.  This
1064:     sequence does not decrease anymore if~$m > m ' := n(1 - v)^2
1065:     /\ell$, and
1066:     \begin{align*}
1067:   p'_{m'} &= \prod_{i=0}^{m'-1} \Bigl(v+\sqrt{\frac{i\ell}{n}}\Bigr) 
1068:   =
1069:   \exp\left(\sum_{i=0}^{m'-1} \ln\Bigl(v+\sqrt{\frac{i\ell}{n}}\Bigr)\right)\\
1070:   &\approx
1071:   \exp\left(\int_{0}^{m'} \ln\Bigl(v+\sqrt{\frac{i\ell}{n}}\Bigr)\right)
1072:   =
1073:   \exp\Bigl((4v-1+2v^2 \ln(v) -
1074:   3v^2)\cdot\frac{n}{2\ell}\Bigr)
1075:   \approx
1076:   \Bigl(1-\frac{(1-v)^3}{2}\Bigr)^{\frac{3n}{4\ell}}.
1077: \end{align*}}
1078: \end{proof}
1079: 
1080: \section{Improving the Rate}\label{sec:reducingtheexponent}
1081: 
1082: Theorem~\ref{thm:pr_local} shows that the~$n$-fold parallel repetition
1083: reduces the winning probability from~$v(\mkG)$ to
1084: $(1-\Theta(1-v(\mkG))^3)^{\Omega(\frac{n}{\log(|\cA||\cB|)})}$.  As
1085: shown in \cite{PaRaWi97}, the term~$|\cA|\cdot|\cB|$ in the exponent
1086: can be reduced to the the maximum (over~$x$, $y$) number of
1087: (fractional) rectangles needed to cover the $1$-entries
1088: in~$Q(x,y,\cdot,\cdot)$.  Here, we show that it can be reduced to a
1089: quantity which is possibly smaller in some cases.
1090: 
1091: \begin{definition}[Exact Fractional Product Cover]
1092:   Let~$Q: \cA\times\cB\rightarrow\{0,1\}$ be an arbitrary predicate.
1093:   Two functions $f: \cA \times \{1,\ldots,\alpha\} \rightarrow [0,1]$
1094:   and~$g: \cB\times \{1,\ldots,\alpha\}\rightarrow [0,1]$ form an
1095:   \emph{exact fractional product cover of size~$\alpha$} for~$Q$ if for
1096:   all~$a,b$:
1097:   \begin{align*}
1098:     Q(a,b) = \sum_{i=1}^\alpha f(a,i)g(b,i).
1099:   \end{align*}
1100: \end{definition}
1101: 
1102: Clearly, any partition by rectangles  gives an exact fractional
1103: product cover (by definining $f(a,i)$ and~$g(b,i)$ as appropriate
1104: predicates).  We will prove the following strengthening of
1105: Theorem~\ref{thm:pr_local}.
1106: \begin{theorem}\label{thm:pr_local_strengthened}
1107:   Let~$\mkG = (\Pd_{XY},Q)$ be a game.  Let~$\alpha$ be such that for
1108:   all~$(x,y)$ there exists an exact fractional product cover of size~$\alpha$
1109:   for~$Q_{x,y}(a,b) := Q(x,y,a,b)$.
1110:   If~$\alpha > 1$ then
1111:   \begin{align}
1112:     v(\mkG^{ n}) \leq
1113:     \Bigl(1-\frac{(1-v)^3}{6000}\Bigr)^{\frac{n}{\log(\alpha)}},\label{eq:23}
1114:   \end{align}
1115:   and if~$\alpha = 1$ then
1116:   \begin{align}
1117:     v(\mkG^{ n}) \leq
1118:     \Bigl(1-\frac{(1-v)^2}{6000}\Bigr)^{n}.\label{eq:24}
1119:   \end{align}
1120: \end{theorem}
1121: 
1122: To prove Theorem~\ref{thm:pr_local_strengthened} we first need a
1123: characterization of fractional product covers by Markov chains.
1124: 
1125: \begin{lemma}\label{lem:fractionalCovEq}
1126:   Let a distribution $\Pd_{ABZ}=\Pd_{A}\Pd_{B}\Pd_{Z|AB}$ be given for which
1127:   there exists functions $f(a,z): \cA\times\cZ \rightarrow [0,1]$
1128:   and~$g(b,z): \cB\times\cZ\rightarrow[0,1]$ which satisfy
1129:   \begin{align}\label{eq:9}
1130:     \Pd_{Z|A=a B=b}(z) = f(a,z)\cdot g(b,z).
1131:   \end{align}
1132:   Then, $A\leftrightarrow Z \leftrightarrow B$.
1133: \end{lemma}
1134: Lemma~\ref{lem:fractionalCovEq} could be strengthened as follows: if
1135: $\Pd_{Z|AB}$ is such that $A\leftrightarrow Z\leftrightarrow B$ for
1136: all distributions $\Pd_{A}\Pd_{B}$, then $\Pd_{Z|AB}$ is of the form
1137: (\ref{eq:9}) for some functions $f$ and~$g$.  For completeness, we
1138: prove this in Appendix~\ref{app:fractEquivalence}.
1139: 
1140: Lemma~\ref{lem:fractionalCovEq} implies the following: if~$Q:
1141: \cA\times\cB \rightarrow \{0,1\}$ has a fractional product cover of
1142: size~$\alpha$, then there exists a random variable~$Z$ over some
1143: set~$\cZ$ given by a conditional distribution~$\Pd_{Z|AB}$ with the
1144: following properties:
1145: \begin{itemize}
1146: \item For any product distribution~$\Pd_{AB}=\Pd_{A}\Pd_{B}$ we
1147:   have~$A \leftrightarrow Z \leftrightarrow B$
1148: \item $|\{z \in \cZ|\exists a,b: Q(a,b)=1 \land \Pd_{Z|A=aB=b}(z) >
1149:   0\}| \leq \alpha$
1150: \item $Q(a,b)$ can be inferred from~$z$.
1151: \end{itemize}
1152: (Note that we do not restrict the alphabet size of~$Z$ in
1153: case~$Q(a,b)=0$, which means that in this case $z$ can be, for example,
1154: $(a,b)$.)
1155: 
1156: \begin{proof}[Proof (of Lemma~\ref{lem:fractionalCovEq})]
1157:   We get
1158:   \begin{align*}
1159:     \Pd_{A|B=b Z=z}(a) 
1160:     &=
1161:     \frac{\Pd_{ABZ}(a,b,z)}{\Pd_{BZ}(b,z)}\\
1162:     &=
1163:     \frac{\Pd_{A}(a)\Pd_{B}(b)f(a,z)g(b,z)}
1164:     {\sum_{a'}\Pd_{B}(b)\Pd_{A}(a')f(a',z)g(b,z)}\\
1165:     &=
1166:     \frac{\Pd_{A}(a)f(a,z)}
1167:     {\sum_{a'}\Pd_{A}(a')f(a',z)}\\
1168:     &=\frac{\sum_{b'} \Pd_{A}(a)\Pd_{B}(b')f(a,z)g(b',z)}
1169:     {\sum_{a',b'}\Pd_{A}(a')\Pd_{B}(b')f(a',z)g(b',z)}\\
1170:     &=\frac{\Pd_{AZ}(a,z)}{\Pd_{Z}(z)}\\
1171:     &=\Pd_{A|Z=z}(a),
1172:   \end{align*}
1173:   and thus $\Pd_{ABZ}(a,b,z) = \Pd_{Z}(z)\Pd_{B|Z=z}(b)
1174:   \Pd_{A|B=bZ=z}(a) = 
1175:   \Pd_{Z}(z)\Pd_{B|Z=z}(b)
1176:   \Pd_{A|Z=z}(a)$, which means that $A\leftrightarrow
1177:   Z \leftrightarrow B$.
1178: \end{proof}
1179: 
1180: Given the characterization from Lemma~\ref{lem:fractionalCovEq} we can
1181: now prove Theorem~\ref{thm:pr_local_strengthened}.
1182: 
1183: \begin{proof}[Proof (of Theorem~\ref{thm:pr_local_strengthened})]
1184:   We first show that Lemma~\ref{lem:locallyComputableGame} still holds
1185:   if we replace (\ref{eq:27}) by
1186:   \begin{align}
1187:     \sum_{j=1}^{k}
1188:     \eps_j \leq 
1189:     15\sqrt{k}\sqrt{(n-k)\log(\alpha)+
1190:       \log\Bigl(\frac{1}{\Pr[W_{k+1}\land\dots\land W_n]}\Bigr)}.\label{eq:34}
1191:   \end{align}
1192:   
1193:   For this, we define the random variables~$D^k$, $U^k$, $\Ubar^k$,
1194:   and $T$ exactly as in the proof of
1195:   Lemma~\ref{lem:locallyComputableGame}.  Instead of (\ref{eq:45}) we
1196:   now define
1197:   \begin{align}
1198:     V := (Z_{k+1},\ldots,Z_{n}),
1199:   \end{align}
1200:   where $Z_i$ is obtained from $(A_i,B_i,X_i,Y_i)$ by a channel that
1201:   has alphabet size at most~$\alpha$ in case~$W_i$, which ensures $A_i
1202:   \leftrightarrow X_iY_iZ_i\leftrightarrow B_i$ in case $A_i$
1203:   and~$B_i$ are independent, and for which $W_i$ can be inferred from
1204:   $(X_i,Y_i,Z_i)$.  The existence of such a random variable is ensured
1205:   by Lemma~\ref{lem:fractionalCovEq} and the fact that for
1206:   every~$(x,y)$ there exists a exact fractional product cover of
1207:   size~$\alpha$ for~$Q(x,y,\cdot,\cdot)$ (the alphabet size of~$Z$ in
1208:   case $Q(x,y,a,b)=0$ is irrelevant and $Z$ can be defined, for
1209:   example, as~$(A,B)$ in this case).
1210:   
1211:   From Corollary~\ref{cor:disjointreps} we now get
1212:   \begin{align}
1213:     \sum_{j=1}^k 
1214:     \Bigl\| \Pd_{T U_jV|W} - 
1215:     \Pd_{T V|W}\Pd_{U_j|T} \Bigr\|
1216:     &\leq \eps_{\Tot}\label{eq:14b}\,,
1217:   \end{align}
1218:   where we set
1219:   \begin{align} \eps_{\Tot} := \sqrt{k}
1220:     \sqrt{(n-k)\log(\alpha)+
1221:       \log\Bigl(\frac{1}{\Pr[W]}\Bigr)}.
1222:   \end{align}
1223:   
1224:   For a fixed~$j$ we define $T^\NoJ$ as in the proof of
1225:   Lemma~\ref{lem:locallyComputableGame} and obtain in exactly the same
1226:   way for $S := (T^\NoJ,V)$ and
1227:   \begin{align}
1228:     \Pd_{\Stilde \Xtilde^n \Ytilde^n} := \Pd_{S X^nY^n|W}
1229:   \end{align}
1230:   the equations
1231:   \begin{align}
1232:     \sum_{j=1}^k 
1233:     \Bigl\| \Pd_{\Stilde \Xtilde_j \Ytilde_j} - 
1234:     \Pd_{XY}\Pd_{\Stilde|\Ytilde_j} \Bigr\|
1235:     &\leq 3\eps_{\Tot}
1236:   \end{align}
1237:   and
1238:   \begin{align}
1239:     \sum_{j=1}^k 
1240:     \Bigl\| \Pd_{\Stilde \Xtilde_j \Ytilde_j} - 
1241:     \Pd_{XY}\Pd_{\Stilde|\Xtilde_j} \Bigr\|
1242:     &\leq 3\eps_{\Tot}\,.
1243:   \end{align}
1244:   Again, 
1245:   Corollary~\ref{cor:LocallyComputableCommonPart} implies that~$(X,Y)$
1246:   is $1-\eps_j$-embeddable in $(\Xtilde_j\Stilde,\Ytilde_j\Stilde)$
1247:   with $(\Xtilde_j,\Ytilde_j)=(X,Y)$ and such that $\sum_{j=1}^k
1248:   {\eps_j} \leq 15\eps_{\Tot}$.
1249: 
1250:   Again we get
1251:   \begin{align}
1252:     X^k \leftrightarrow T V \leftrightarrow Y^k,
1253:   \end{align}
1254:   now using the properties of the~$Z_i$. (This is done as follows:
1255:   clearly, $X^k \leftrightarrow T \leftrightarrow Y^k$, i.e. for a
1256:   fixed values~$t$ for~$T$ the $X^k$ and $Y^k$ are independent.  Now,
1257:   inductively adding $Z_i$ will not change this in any step.)
1258:   Claim~\ref{claim:markovcondition} now yields
1259:   \begin{align}
1260:     \Xtilde^n \leftrightarrow \Xtilde_j \Stilde \leftrightarrow
1261:     \Ytilde^n\Ytilde_j\Stilde \label{eq:42}\\
1262:     \Xtilde^n\Xtilde_j\Stilde \leftrightarrow \Ytilde_j\Stilde \leftrightarrow
1263:     \Ytilde^n,\label{eq:43}
1264:   \end{align}
1265:   and Lemma~\ref{lem:locallyComputableMarkov} completes the proof that
1266:   (\ref{eq:34}) can replace (\ref{eq:27}) in
1267:   Lemma~\ref{lem:locallyComputableGame}.
1268: 
1269:   From Lemma~\ref{lem:locallyComputableGame} where (\ref{eq:27}) is
1270:   replaced by~(\ref{eq:34}) we obtain~(\ref{eq:23}) exactly as in the
1271:   proof of Theorem~\ref{thm:pr_local}.  To get~(\ref{eq:24}) we note
1272:   first that in this case (\ref{eq:34}) reduces to
1273:   \begin{align}\label{eq:46}
1274:     \sum_{j=1}^{k}
1275:     \eps_j \leq 
1276:     15\sqrt{k}\sqrt{\log\Bigl(\frac{1}{\Pr[W_{k+1}\land\dots\land W_n]}\Bigr)}.
1277:   \end{align}
1278:   Using an analogous definition for~$p_m$ as previously, we get
1279:   \begin{align}\label{eq:6}
1280:     p_{m+1} \leq p_{m} \cdot
1281:     \Bigl(v+15\sqrt{\frac{1}{n-m}\log\Bigl(\frac{1}{p_{m}}\Bigr)}\Bigr).
1282:   \end{align}
1283:   Here, we show per induction that $p_m \leq (\frac{1+v}2)^m$ as long as
1284:   $m+1 \leq \frac{(n-m)(1-v)}{400}$.  To make a step from~$m$ to~$m+1$
1285:   we can assume~$p_m \geq (\frac{1+v}2)^{m+1}$, which implies $p_m
1286:   \geq 2^{-(1-v)(m+1)}$ (since $(1-\frac12)^{1-v} \leq
1287:   1-\frac{1-v}{2}$, see inequality below), which means that
1288:   \begin{align}
1289:     p_{m+1} &\leq p_{m}
1290:     \cdot\Bigl(v+\sqrt{\frac{225(m+1)(1-v)}{n-m}}\Bigr),
1291:   \end{align}
1292:   for relevant values of~$m$.  If $m+1 \leq \frac{(n-m)(1-v)}{900}$
1293:   this implies the hypothesis.  We thus get for~$m =
1294:   \frac{n(1-v)}{1800}$
1295:   \begin{align*}
1296:     p_m &\leq
1297:     \Bigl(\frac{1+v}{2}\Bigr)^{\frac{n(1-v)}{1800}}.
1298:   \end{align*}
1299:   Finally, 
1300:   $(\frac{1+v}{2})^{(1-v)/1800} = (1-\frac{(1-v)}{2})^{(1-v)/1800}
1301:   \leq 1-\frac{(1-v)^2}{3600}$,
1302:   again using $(1-b)^a\leq 1-ab$ for $a\in[0,1]$, $b\leq
1303:   1$.\footnote{Note that in case $p_m \leq
1304:     \Bigl(1-\Theta((1-v)^2)\Bigr)^{n}$ equation~(\ref{eq:6})
1305:     only implies $p_{m+1} \leq 
1306:     p_m (v + 15\sqrt{-\log(1-\Theta((1-v)^2))}) =
1307:     p_m (v + 15\sqrt{\Theta((1-v)^2)}) = 
1308:     p_m (v + 1-v) = p_m$,
1309:     and thus (\ref{eq:6}) cannot be
1310:     used to get a significantly stronger version of the theorem.}
1311: \end{proof}
1312: 
1313: \section{No-signaling Strategies}\label{sec:nosignal}
1314: 
1315: No-signaling strategies are those where the only restriction on the
1316: response of Alice and Bob is that they do not \emph{imply}
1317: communication.
1318: 
1319: \begin{definition}[No-signaling]
1320:   A pair~$(h_a,h_b)$ of functions is \emph{no-signaling} if~$h_a:
1321:   \cX\times\cY\times\cR \rightarrow \cA$ and~$h_b:
1322:   \cX\times\cY\times\cR\rightarrow\cB$ satisfy
1323:   \begin{align*}
1324:     \Pr_{R}[h_a(x,y,R)] &= \Pr_R[h_a(x,y',R)]\\
1325:     \Pr_{R}[h_b(x,y,R)] &= \Pr_R[h_b(x',y,R)],
1326:   \end{align*}
1327:   for all $x,x',y,y'$.
1328: \end{definition}
1329: \begin{definition}[No-signaling value]
1330:   The no-signaling value $\vns(\mkG)$ of a
1331:   \emph{game~$\mkG=(\Pd_{XY},Q)$
1332:     over~$\cX\times\cY\times\cA\times\cB$} is
1333:   \begin{align*}
1334:     \vns(\mkG) := \max \Pr_{XYR}[Q(X,Y,h_a(X,Y,R),h_b(X,Y,R))],
1335:   \end{align*}
1336:   where the maximum is over all no-signaling functions $(h_a,h_b)$.
1337: \end{definition}
1338: Clearly,~$v(\mkG) \leq \vns(\mkG)$, since any local strategy is a
1339: no-signaling strategy.  We further note that for no-signaling
1340: strategies $\vns(\mkG^2) > (\vns(\mkG))^2$ is also possible, similar to
1341: the local case (see Appendix~\ref{app:nontriviality}).
1342: 
1343: We will prove the following Theorem:
1344: \begin{theorem}\label{thm:pr_nosignaling}
1345:   For any game~$\mkG$ with no-signaling value~$\vns := \vns(\mkG)$ and any
1346:   integer~$n$:
1347:   \begin{align}\label{eq:7}
1348:     \vns(\mkG^n) \leq 
1349:     \Bigl(1-\frac{(1-\vns)^2}{6400}\Bigr)^{n}. %% 
1350:   \end{align}
1351: \end{theorem}
1352: We remark that the proof of this theorem will be much simpler than the
1353: proof of Theorem~\ref{thm:pr_local}.
1354: 
1355: We first show that if $\Pd_{XYST}$ a distribution which is close to
1356: no-signalling (i.e., $\|\Pd_{XYS} - \Pd_{XY}\Pd_{S|X}\|$ and
1357: $\Pd_{XYT}-\Pd_{XY}\Pd_{S|Y}$) then there exists a no-signalling
1358: strategy which produces value which are statistically close to~$S$
1359: and~$T$ from~$X$ and~$Y$.
1360: 
1361: \begin{lemma}\label{lem:no-signaling-close-helper}
1362:   Let~$\Pd_{ST}$, $\Pd_{S'}$ be arbitrary distributions
1363:   over~$\cS\times\cT$ and~$\cS$,~$\cS$ and~$\cT$ finite.
1364:   Then, there exists a distribution~$\Pd_{\Sbar\,\Tbar}$ such that
1365:   \begin{align}
1366:     \| \Pd_{\Sbar\,\Tbar} - \Pd_{ST} \| &\leq \| \Pd_{S'} - \Pd_{S} \|\label{eq:30}\\
1367:     \| \Pd_{\Sbar} - \Pd_{S'}\| &= 0\label{eq:31}\\
1368:     \| \Pd_{\Tbar} - \Pd_{T} \| &= 0.\label{eq:32}
1369:   \end{align}
1370: \end{lemma}
1371: \begin{proof}
1372:   We change~$\Pd_{ST}$ gradually to~$\Pd_{\Sbar\,\Tbar}$
1373:   such that in the end~(\ref{eq:31}) and (\ref{eq:32}) hold.
1374:   
1375:   For this, fix values~$s_0$ and~$s_1$ with
1376:   \begin{align}\label{eq:18}
1377:     \Pd_{S}(s_0) < \Pd_{S'}(s_0) \text{ and }
1378:     \Pd_{S}(s_1) > \Pd_{S'}(s_1).  
1379:   \end{align}
1380:   Then, as long as~(\ref{eq:18}) holds find a value~$t$ for
1381:   which~$\Pd_{ST}(s_1,t) > 0$.  Decrease $\Pd_{ST}(s_1,t)$ by~$\eps$
1382:   and increase $\Pd_{ST}(s_0,t)$ by $\eps$, such that afterwards
1383:   $\Pd_{ST}(s_1,t) = 0$ or (\ref{eq:18}) does not hold anymore
1384:   for~$s_0$, $s_1$.  After a finite number of repetitions (\ref{eq:18})
1385:   is not true anymore, and we start the process over again with new
1386:   values for~$s_0,s_1$.  However, this can also only happen a finite
1387:   number of times, thus the process terminates.
1388:   
1389:   If~(\ref{eq:18}) cannot be satisfied then clearly~(\ref{eq:31})
1390:   holds.  We never change~$\Pd_{T}(t)$ for any~$t$ which
1391:   implies~(\ref{eq:32}).  Finally, (\ref{eq:30}) is ensured by the
1392:   fact that we only decrease~$\|\Pd_{S'}-\Pd_{S}\|$ and do not change
1393:   $\Pd_{ST}$ more than~$\Pd_{S}$.
1394: \end{proof}
1395: 
1396: \begin{lemma}\label{lem:nosignalingclose}
1397:   Let~$\Pd_{X_0Y_0}$ and $\Pd_{XYST}$ be arbitrary distributions.  If
1398:   \begin{align}
1399:     \| \Pd_{X_0Y_0}\Pd_{S|X} - \Pd_{XYS} \| \leq \eps_1,\\
1400:     \| \Pd_{X_0Y_0}\Pd_{T|Y} - \Pd_{XYT} \| \leq \eps_2,
1401:   \end{align}
1402:   then there exists a conditional
1403:   distribution~$\Pd_{S'T'|X'=xY'=y}$ with $\Pd_{S'|X'=x Y'=y} =
1404:   \Pd_{S'|X'=x}$ and $\Pd_{T'|X'=xY'=y} = \Pd_{T'|Y'=y}$ such
1405:   that
1406:   \begin{align}
1407:     \| \Pd_{X_0Y_0}\Pd_{S'T'|XY} - \Pd_{XYST} \| \leq 3\eps_1 + 2\eps_2.
1408:   \end{align}
1409: \end{lemma}
1410: \begin{proof}
1411:   For fixed~$x,y$ we define~$\Pd_{S_0T_0|X=xY=y}$ using
1412:   Lemma~\ref{lem:no-signaling-close-helper} with the following properties:
1413:   \begin{align*}
1414:    \|\Pd_{S_0T_0|X=xY=y}-\Pd_{ST|X=xY=y}\|&\leq\|\Pd_{S|X=x}-\Pd_{S|X=xY=y}\|\\
1415:    \|\Pd_{S_0|X=xY=y}-\Pd_{S|X=x}\| & = 0\\
1416:    \|\Pd_{T_0|X=xY=y}-\Pd_{T|X=xY=y}\| &= 0.
1417:   \end{align*}
1418:   Then, again using Lemma~\ref{lem:no-signaling-close-helper} we
1419:   define $\Pd_{S'T'|X=xY=y}$ such that
1420:   \begin{align*}
1421:     \|\Pd_{S'T'|X=xY=y}-\Pd_{S_0T_0|X=xY=y}\| &\leq 
1422:     \|\Pd_{T_0|Y=y}-\Pd_{T_0|X=xY=y}\|\\
1423:     \|\Pd_{T'|X=xY=y}-\Pd_{T_0|Y=y}\| &= 0\\
1424:     \|\Pd_{S'|X=xY=y}-\Pd_{S_0|X=xY=y}\| & = 0.
1425:   \end{align*}
1426:   We see that for all pairs~$x,y$ we have $\Pd_{S'|X=xY=y} =
1427:   \Pd_{S'|X=x}$ and $\Pd_{T'|X=xY=y}=\Pd_{T'|Y=y}$.
1428:   
1429:   We further get
1430:   \begin{align*}
1431:     \|\Pd_{X_0Y_0}& \Pd_{S'T'|XY}-\Pd_{XYST}\| \\
1432:     &\leq
1433:     \eps_1+\|\Pd_{XY}\Pd_{S'T'|XY} - \Pd_{XYST}\|\\
1434:     &=
1435:     \eps_1+\sum_{x,y,s,t} \bigl|\Pd_{XY}(x,y)\Pd_{S'T'|X=xY=y}(s,t) 
1436:     - \Pd_{XY}(x,y)\Pd_{ST|X=xY=y}(s,t) \bigr|\\
1437:     &\leq
1438:     \eps_1+\sum_{x,y}\Pd_{XY}(x,y)\Bigl( \| \Pd_{S|X=x}-\Pd_{S|X=xY=y}\| +
1439:     \|\Pd_{T|Y=y}-\Pd_{T|X=xY=y}\|\Bigr)\\
1440:     &\leq
1441:     \eps_1+\|\Pd_{XY}\Pd_{S|X}-\Pd_{SXY}\| +
1442:     \|\Pd_{XY}\Pd_{T|Y}-\Pd_{TXY}\|\\
1443:     &\leq
1444:     3\eps_1+2\eps_2.\qedhere
1445:   \end{align*}
1446: \end{proof}
1447: 
1448: We can now prove a non-signaling analogue of
1449: Lemma~\ref{lem:gameconditionedvalue}.\footnote{A previous version of
1450:   the proof of this lemma contained an error, which was first noticed
1451:   by Oded Regev and Ricky Rosen.}
1452: \begin{lemma}\label{lem:ns_gameconditionedvalue}
1453:   Let a game~$\mkG=(Q,\Pd_{XY})$, its~$n$-fold repetition~$\mkG^{n}$,
1454:   and a no-signaling strategy $(h_a,h_b)$ for~$\mkG^{n}$ be given.  Let
1455:   indices~$i_1,\ldots,i_m$ be given.  Then, there exists an
1456:   index~$i_{m+1}$ such that
1457:   \begin{align}
1458:     \Pr[W_{i_{m+1}} | &W_{i_1}\land\dots\land W_{i_m}] \nonumber \\
1459:     &\leq \vns(\mkG) +
1460:     10\sqrt{\frac{1}{n-m}}\sqrt{\log\Bigl(\frac{1}{\Pr[W_{i_1}\land\dots\land W_{i_m}]}\Bigr)}.
1461:   \end{align}
1462: \end{lemma}
1463: \begin{proof}
1464:   As in the proof of Lemma~\ref{lem:gameconditionedvalue} we assume
1465:   that $i_\ell = n-\ell+1$ and we define~$W :=
1466:   W_{n-m+1}\land\dots\land W_{n}$.  The no-signaling property of
1467:   $(h_a, h_b)$ implies $\Pd_{X^nY^nA^n} =
1468:   \Pd_{X^n}\Pd_{Y^n|X^n}\Pd_{A^n | X^n} = \Pd_{A^nX^n}\Pd_{Y^n|X^n}$.
1469:   Thus, when we apply Corollary~\ref{cor:disjointreps} on this
1470:   distribution (with the event~$W$ and the random variables $T=(X^n,
1471:   A^n)$ and $U_j = Y_j$) we get
1472:   \begin{align*}
1473:     \sum_{j=1}^{n-m}\Bigl\| \Pd_{X^nA^nY_j|W} -
1474:     \Pd_{X^nA^n|W}\Pd_{Y_j|X_j}\Bigr\| &=
1475:     \sum_{j=1}^{n-m}\Bigl\| \Pd_{T Y_j|W} - \Pd_{T|W}\Pd_{Y_j|T} \Bigr\|\\
1476:     &\leq \sqrt{(n-m)\log\Bigl(\frac{1}{\Pr[W]}\Bigr)}.
1477:   \end{align*}
1478:   Taking appropriate marginals this gives
1479:   \begin{align*}
1480:    \sum_{j=1}^{n-m}\Bigl\| \Pd_{X_jY_jA_j|W} -
1481:     \Pd_{X_jA_j|W}\Pd_{Y_j|X_j}\Bigr\| 
1482:     \leq \sqrt{(n-m)\log\Bigl(\frac{1}{\Pr[W]}\Bigr)}.
1483:    \end{align*}
1484:   Applying Lemma~\ref{lem:basicrepetitions} once more and rearranging
1485:   we get
1486:   \begin{align}\label{eq:15}
1487:    \sum_{j=1}^{n-m}\Bigl\| \Pd_{X_jY_jA_j|W} -
1488:     \Pd_{XY}\Pd_{A_j|X_j W}\Bigr\| 
1489:     \leq 2\sqrt{(n-m)\log\Bigl(\frac{1}{\Pr[W]}\Bigr)}.
1490:   \end{align}
1491:   Symmetrically, we obtain
1492:   \begin{align}\label{eq:11}
1493:     \sum_{j=1}^{n-m} \Bigl\| 
1494:     \Pd_{X_j Y_j B_j|W} - 
1495:     \Pd_{XY}\Pd_{B_j|Y_j W}
1496:     \Bigr\| \leq 
1497:     2\sqrt{(n-m)\log\Bigl(\frac{1}{\Pr[W]}\Bigr) }.
1498:   \end{align}
1499:   From (\ref{eq:15}), (\ref{eq:11}), and
1500:   Lemma~\ref{lem:nosignalingclose} we get that there exists a
1501:   distribution $\Pd_{A_j'B_j'|XY}$ which can be implemented by
1502:   no-signaling functions and for which
1503:   \begin{align*}
1504:     \sum_{j=1}^{n-m}
1505:     \Bigl\| 
1506:     \Pd_{XY}\Pd_{A_j'B_j'|XY} -
1507:     \Pd_{X_jY_jA_jB_j|W}
1508:     \Bigr\| \leq 10
1509:     \sqrt{(n-m)\Bigl(\log\Bigl(\frac{1}{\Pr[W]}\Bigr)\Bigr)}
1510:       .
1511:   \end{align*}
1512:   Thus, if Alice and Bob use the strategy implied by~$\Pd_{A_j'B_j'|XY}$
1513:   (which is no-signaling) they can win the initial game with
1514:   probability $\Pr[W_j|W] - 10\sqrt{1/(n-m)}\sqrt{\log(1/\Pr[W])}$ for
1515:   some~$j$, which implies the lemma.
1516: \end{proof}
1517: 
1518: \begin{proof}[Proof (of Theorem~\ref{thm:pr_nosignaling})]
1519:   Fix a no-signaling strategy~$(h_a,h_b)$ for~$\mkG$.  As in the proof
1520:   of Theorem~\ref{thm:pr_local} we repeatedly select indices~$i_{m+1}$
1521:   such that~$\Pr[W_{i_{m+1}}|W_{i_1}\land\dots\land W_{i_m}]$ is
1522:   minimized.  Let~$p_m := \Pr[W_{i_1}\land\dots\land W_{i_m}]$.
1523:   Lemma~\ref{lem:ns_gameconditionedvalue} implies
1524:   \begin{align}\label{eq:36}
1525:     p_{m+1} \leq p_{m} \cdot
1526:     \Bigl(v+10\sqrt{\frac{1}{n-m}\log\Bigl(\frac{1}{p_{m}}\Bigr)}\Bigr).
1527:   \end{align}
1528:   From (\ref{eq:36}) we obtain (\ref{eq:7}) in the same way as we
1529:   obtained (\ref{eq:24}) from (\ref{eq:6}) in the proof of
1530:   Theorem~\ref{thm:pr_local_strengthened}.
1531: \end{proof}
1532: 
1533: \section*{Acknowledgments}
1534: I would like to thank Paul Beame, Georges Baatz, Ryan O'Donnell, Johan
1535: H\aa{}stad, Bartosz Przydatek, Anup Rao, Ran Raz, Renato Renner, Ricky
1536: Rosen, Stefano Tessaro, Stefan Wolf, and J\"urg Wullschleger for
1537: helpful discussions.  Paul Beame pointed out \cite{PaRaWi97} to me and
1538: explained how the strengthening given there can be adapted to the
1539: proof given here; he also simplified an argument in a previous version
1540: of this paper.  Anup Rao simplified an argument in the proof of
1541: Lemma~\ref{lem:LocallyComputableCommonPartPre}.  Oded Regev and Ricky
1542: Rosen found an error in a previous version of the proof of
1543: Lemma~\ref{lem:ns_gameconditionedvalue}.  I was supported by the Swiss
1544: National Science Foundation, project no.~200020-103847/1.
1545: 
1546: {
1547: \footnotesize
1548: \begin{thebibliography}{BOGKW88}
1549: 
1550: \bibitem[BBT05]{BrBrTa05}
1551: Gilles Brassard, Anne Broadbent, and Alain Tapp.
1552: \newblock Quantum pseudo-telepathy.
1553: \newblock {\em Foundations of Physics}, 35(11):1877--1907, 2005.
1554: \newblock (quant-ph/0407221).
1555: 
1556: \bibitem[Bea06]{Beame06}
1557: Paul Beame.
1558: \newblock Personal communication, 2006.
1559: 
1560: \bibitem[BOGKW88]{BGKW88}
1561: Michael Ben-Or, Shafi Goldwasser, Joe Kilian, and Avi Wigderson.
1562: \newblock Multi-prover interactive proofs: How to remove intractability
1563:   assumptions.
1564: \newblock In {\em Proceedings of the Twentieth Annual ACM Symposium on Theory
1565:   of Computing}, pages 113--132, 1988.
1566: 
1567: \bibitem[CCL92]{CaCoLi92}
1568: {Jin-yi} Cai, Anne Condon, and Richard~J. Lipton.
1569: \newblock On games of incomplete information.
1570: \newblock {\em Theoretical Computer Science}, 103(1):25--38, 1992.
1571: 
1572: \bibitem[CHSH69]{CHSH69}
1573: John~F. Clauser, Michael~A. Horne, Abner Shimony, and Richard~A. Holt.
1574: \newblock Proposed experiment to test local hidden-variable theories.
1575: \newblock {\em Physical Review Letters}, 23(15):880--884, 1969.
1576: 
1577: \bibitem[Cir80]{Cirels80}
1578: Boris~S. Cirel'son.
1579: \newblock Quantum generalizations of {B}ell's inequality.
1580: \newblock {\em Letters in Mathematical Physics}, 4(2):93--100, 1980.
1581: 
1582: \bibitem[CSUU06]{CSUU06}
1583: Richard Cleve, William Slofstra, Falk Unger, and Sarvagya Upadhyay.
1584: \newblock Strong parallel repetition theorem for quantum {XOR} proof systems,
1585:   2006.
1586: \newblock (quant-ph/0608146).
1587: 
1588: \bibitem[CT91]{CovTho91}
1589: Thomas~M. Cover and Joy~A. Thomas.
1590: \newblock {\em Elements of Information Theory}.
1591: \newblock John Wiley~\& Sons,~Inc., first edition, 1991.
1592: \newblock ISBN 0-471-06259-6.
1593: 
1594: \bibitem[Fei91]{Feige91}
1595: Uriel Feige.
1596: \newblock On the success probability of the two provers in one-round proof
1597:   systems.
1598: \newblock In {\em Proceedings of the sixth Annual Structure in Complexity
1599:   Theory Conference}, pages 116--123, 1991.
1600: 
1601: \bibitem[FL92]{FeiLov92}
1602: Uriel Feige and L{\'a}szl{\'o} L{\'o}vasz.
1603: \newblock Two-prover one-round proof systems: Their power and their problems.
1604: \newblock In {\em Proceedings of the Twenty-Fourth Annual ACM Symposium on
1605:   Theory of Computing}, pages 733--744, 1992.
1606: 
1607: \bibitem[For89]{Fortno89}
1608: Lance Fortnow.
1609: \newblock {\em Complexity-Theoretic Aspects of Interactive Proof Systems}.
1610: \newblock PhD thesis, Massachusetts Institute of Technology, 1989.
1611: 
1612: \bibitem[FRS94]{FoRoSi94}
1613: Lance Fortnow, John Rompel, and Michael Sipser.
1614: \newblock On the power of multi-prover interactive proof systems.
1615: \newblock {\em Theoretical Computer Science}, 134(21):545--557, 1994.
1616: 
1617: \bibitem[FV02]{FeiVer02}
1618: Uriel Feige and Oleg Verbitsky.
1619: \newblock Error reduction by parallel repetition -- a negative result.
1620: \newblock {\em Combinatorica}, 22(4):461--478, 2002.
1621: 
1622: \bibitem[HR06]{HolRen06}
1623: Thomas Holenstein and Renato Renner.
1624: \newblock On the randomness of independent experiments, 2006.
1625: \newblock (cs.IT/0608007).
1626: 
1627: \bibitem[LS95]{LapSha95}
1628: Dror Lapidot and Adi Shamir.
1629: \newblock A one-round, two-prover, zero-knowledge protocol for {NP}.
1630: \newblock {\em Combinatorica}, 15(2):204--214, 1995.
1631: 
1632: \bibitem[NC00]{NieChu00}
1633: Michael~A. Nielsen and Isaac~L. Chuang.
1634: \newblock {\em Quantum Computation and Quantum Information}.
1635: \newblock Cambridge University Press, first edition, 2000.
1636: \newblock ISBN 0-521-63503-9.
1637: 
1638: \bibitem[PR94]{PopRoh94}
1639: Sandu Popescu and Daniel Rohrlich.
1640: \newblock Nonlocality as an axiom for quantum theory.
1641: \newblock {\em Foundations of Physics}, 24(3):379--385, March 1994.
1642: \newblock (quant-ph/9508009).
1643: 
1644: \bibitem[PRW97]{PaRaWi97}
1645: Itzhak Parnafes, Ran Raz, and Avi Wigderson.
1646: \newblock Direct product results and the {GCD} problem, in old and new
1647:   communication models.
1648: \newblock In {\em Proceedings of the Twenty-Ninth Annual ACM Symposium on
1649:   Theory of Computing}, pages 363--372, 1997.
1650: 
1651: \bibitem[Raz98]{Raz98}
1652: Ran Raz.
1653: \newblock A parallel reptition theorem.
1654: \newblock {\em {SIAM} Journal on Computing}, 27(3):763--803, 1998.
1655: 
1656: \bibitem[Ver94]{Verbit94}
1657: Oleg Verbitsky.
1658: \newblock Towards the parallel repetition conjecture.
1659: \newblock In {\em Structure in Complexity Theory Conference}, pages 304--307,
1660:   1994.
1661: 
1662: \bibitem[Wat02]{Watrou02}
1663: John Watrous.
1664: \newblock A note on parallel repetition of two-prover quantum-interactive
1665:   proofs, 2002.
1666: \newblock Manuscript.
1667: 
1668: \end{thebibliography}
1669: 
1670: }
1671: 
1672: \appendix
1673: \section{Non-triviality}
1674: \label{app:nontriviality}
1675: \paragraph{Local case}
1676: We quickly reproduce a slight modification\footnote{Fortnow also lets
1677:   the referee choose~$x=y=1$ with some probability, in which case the
1678:   players cannot win the game.} of Fortnow's example \cite{Fortno89}
1679: which shows that in general~$v(\mkG^2) > (v(\mkG))^2$.  The same
1680: variation was also considered by Feige and Lov\'asz \cite{FeiLov92}.
1681: 
1682: The game we describe is over bits (i.e., all the queries and all the
1683: responses are bits).  We
1684: set~$\Pd_{XY}(0,0):=\Pd_{XY}(0,1):=\Pd_{XY}(1,0):=\frac13$, and define
1685: \begin{align}Q(x,y,a,b) := \bigl((x \lor a) \neq (y \lor b)\bigr).
1686: \end{align}
1687: This can be described in words: Alice and Bob receive a bit, and at
1688: least one of these bits is~$0$.  If both players receive~$0$,
1689: exactly one player must respond with $1$.  If one of the players
1690: receives~$1$, the other must respond with~$0$.  
1691: 
1692: We first show that for this game~$v=\frac{2}{3}$.  Clearly, $v \geq
1693: \frac23$ (e.g.,~both players always answer~$0$).  To
1694: show~$v\leq\frac23$ we check all deterministic strategies.  If both
1695: players reply~$0$ on query~$0$, this fails in case~$x=y=0$ (and thus
1696: with probability~$\frac13$).  If one player, w.l.o.g.~Alice,
1697: answers~$0$ with~$1$ the players fail in case~$x=0$ and~$y=1$.
1698: 
1699: If this game is repeated twice in parallel, setting~$(a_1,a_2) :=
1700: (x_2,x_1)$,~$(b_1,b_2) := (y_2,y_1)$ also wins with
1701: probability~$\frac23$.  One can check this as follows: for every fixed
1702: query~$(x_1,y_1)$ answering with~$(x_2,y_2)$ wins the first subgame
1703: with probability~$\frac23$.  Moreover, with this strategy
1704: $Q(x_1,y_1,a_1,b_1) \equiv
1705: Q(x_2,y_2,a_2,b_2)$ which implies the claim.  
1706: 
1707: \paragraph{No-signaling case}
1708: We now show that for the above game 
1709: \begin{align}\label{eq:41}
1710:   v(\mkG)= v(\mkG^2)=\vns(\mkG)=\vns(\mkG^2).
1711: \end{align}
1712: Previously, it was known that quantum strategies do not help Alice and
1713: Bob to win this game \cite{Watrou02} (in both the single instance case
1714: and where two parallel instances are used).  
1715: 
1716: To show~(\ref{eq:41}) it is sufficient to show that that $v(\mkG) =
1717: \vns(\mkG)$ (since $\vns(\mkG^2)\leq \vns(\mkG)$ and $\vns(\mkG^2)
1718: \geq v(\mkG^2) = v(\mkG)$ are already known).  There are two ways to
1719: see that $v(\mkG) = \vns(\mkG)$.  First, one can notice that the joint
1720: probability of Alice's and Bob's reply only matters if~$x=y=0$; i.e.,
1721: only for one query.  In such a case one can always get a local
1722: strategy which is as good as a given no-signaling strategy.
1723: Alternatively, let $p$ be the probability that Alice replies~$0$ on
1724: query~$0$ and $q$ be the probability that Bob replies~$0$ on
1725: query~$0$.  In this case, the players win with probability at most~$p$
1726: on query~$(x,y)=(0,1)$, with probability at most~$q$ on query~$(1,0)$,
1727: and with probability at most $(1-p)+(1-q)$ on query~$(0,0)$, which
1728: gives an overall winnig probability of at most $\frac23$.
1729: 
1730: \section{A Lemma on Relative Entropy}\label{app:relentrlemma}
1731: 
1732: This following lemma is well known, but we do not know of a standard
1733: reference containing a proof of it.
1734: \begin{lemma}
1735:   Let $\Pd_{U^k} = \Pd_{U_1}\dots\Pd_{U_k}$ and
1736:   $\Pd_{V^k}$ be distributions over the same set.  Then,
1737:   \begin{align*}
1738:     \sum_{j=1}^k D(\Pd_{V_j}\|\Pd_{U_j}) \leq
1739:     D(\Pd_{V^k}\|\Pd_{U^k}).
1740:   \end{align*}
1741: \end{lemma}
1742: \begin{proof}
1743:   We prove the
1744:   bipartite case; the general case follows by induction.
1745:   \begin{align*}
1746:     D(\Pd_{V_1 V_2}&\|\Pd_{U_1}\Pd_{U_2}) \\
1747:     &=
1748:     \sum_{(u_1,u_2)}
1749:     \Pd_{V_1 V_2}(u_1,u_2)\log\Bigl(
1750:     \frac{\Pd_{V_1 V_2}(u_1,u_2)}
1751:     {\Pd_{U_1}(u_1)\Pd_{U_2}(u_2)}\Bigr)\displaybreak[3]\\
1752:     &=
1753:     \sum_{(u_1,u_2)}
1754:     \Pd_{V_1 V_2}(u_1,u_2)\log\Bigl(
1755:     \frac{\Pd_{V_1}(u_1)}
1756:     {\Pd_{U_1}(u_1)}\Bigr)+
1757:     \\ &\qquad\qquad+
1758:     \sum_{(u_1,u_2)}\!\!
1759:     \Pd_{V_1 V_2}(u_1,u_2)\log\Bigl(
1760:     \frac{\Pd_{V_2|V_1=u_1}(u_2)}
1761:     {\Pd_{U_2}(u_2)}\Bigr)\displaybreak[3]\\
1762:     &=
1763:     D(\Pd_{V_1}\|\Pd_{U_1})
1764:     +
1765:     \sum_{(u_1,u_2)}
1766:     \Pd_{V_1 V_2}(u_1,u_2)\log\Bigl(
1767:     \frac{\Pd_{V_2}(u_2)}{\Pd_{U_2}(u_2)}\cdot
1768:     \frac{\Pd_{V_1 V_2}(u_1,u_2)}
1769:     {\Pd_{V_1}(u_1)\Pd_{V_2}(u_2)}
1770:     \Bigr)\\
1771:     &=D(\Pd_{V_1}\|\Pd_{U_1})+D(\Pd_{V_2}\|\Pd_{U_2})
1772:     +\!\!\sum_{(u_1,u_2)}\!\!
1773:     \Pd_{V_1 V_2}(u_1,u_2)\log\Bigl(
1774:     \frac{\Pd_{V_1 V_2}(u_1,u_2)}
1775:     {\Pd_{V_1}(u_1)\Pd_{V_2}(u_2)}\Bigr)\\
1776:     &\geq D(\Pd_{V_1}\|\Pd_{U_1})+D(\Pd_{V_2}\|\Pd_{U_2}),
1777:   \end{align*}
1778:   where the last inequality follows from the log-sum inequality (see
1779:   \cite[Theorem 2.7.1]{CovTho91}).
1780: \end{proof}
1781: 
1782: \section{Converse of Lemma~\ref{lem:fractionalCovEq}}
1783: \label{app:fractEquivalence}
1784: 
1785: In this appendix we show that Lemma~\ref{lem:fractionalCovEq} can be
1786: strengthened to get an ``if and only if'' condition.
1787: \begin{lemma}\label{lem:converse}
1788:   Let a conditional distribution $\Pd_{Z|AB}$ be given.  If, for all
1789:   product distributions~$\Pd_{AB}=\Pd_{A}\Pd_{B}$ the Markov condition
1790:   $A\leftrightarrow Z \leftrightarrow B$ is satisfied, then  there
1791:   exists functions $f(a,z): \cA\times\cZ \rightarrow [0,1]$
1792:   and~$g(b,z): \cB\times\cZ\rightarrow[0,1]$ such that
1793:   \begin{align}
1794:     \Pd_{Z|A=a B=b} = f(a,z)\cdot g(b,z).
1795:   \end{align}
1796: \end{lemma}
1797: \begin{proof}
1798:   Fix an arbitrary~$z$ throughout the proof, and consider
1799:   arbitrary elements $a,a'\in\cA$ and $b,b' \in \cB$.  We
1800:   set~$\Pd_{A}(a) = \Pd_{A}(a') = \frac12$ and $\Pd_{B}(b) =
1801:   \Pd_{B}(b') = \frac12$.  The Markov condition implies
1802:   \begin{align*}
1803:     \Pd_{A|Z=z B=b}(a) = \Pd_{A|Z=z B=b'}(a)
1804:   \end{align*}
1805:   which is equivalent to 
1806:   \begin{align*}
1807:     \frac{\Pd_{ABZ}(a,b,z)}{\Pd_{ABZ}(a,b,z)+\Pd_{ABZ}(a',b,z)} =
1808:     \frac{\Pd_{ABZ}(a,b',z)}{\Pd_{ABZ}(a,b',z)+\Pd_{ABZ}(a',b',z)}
1809:   \end{align*}
1810:   or (because of our choice of~$\Pd_{AB}$)
1811:   \begin{align*}
1812:     \frac{\Pd_{Z|A=a,B=b}(z)}{\Pd_{Z|A=a,B=b}(z)+\Pd_{Z|A=a',B=b}(z)} =
1813:     \frac{\Pd_{Z|A=a,B=b'}(z)}{\Pd_{Z|A=a,B=b'}(z)+\Pd_{Z|A=a',B=b'}(z)}.
1814:   \end{align*}
1815:   Analogously one gets (by swapping the roles of~$a$ and~$a'$)
1816:   \begin{align*}
1817:     \frac{\Pd_{Z|A=a',B=b}(z)}{\Pd_{Z|A=a,B=b}(z)+\Pd_{Z|A=a',B=b}(z)} =
1818:     \frac{\Pd_{Z|A=a',B=b'}(z)}{\Pd_{Z|A=a,B=b'}(z)+\Pd_{Z|A=a',B=b'}(z)}.
1819:   \end{align*}
1820:   Together, this implies
1821:   \begin{align}\label{eq:16}
1822:     \Pd_{Z|A=a,B=b}(z) \Pd_{Z|A=a',B=b'}(z) = 
1823:     \Pd_{Z|A=a,B=b'}(z) \Pd_{Z|A=a',B=b}(z).
1824:   \end{align}
1825:   
1826:   Fix now~$z$ and let~$f(\cdot,z)$ and~$g(\cdot,z)$ be functions onto~$[0,1]$
1827:   which satisfy
1828:   \begin{align}\label{eq:40}
1829:     f(a,z)g(b,z) \geq \Pd_{Z|A=a,B=b}(z)
1830:   \end{align}
1831:   for all~$(a,b)$ and for which the number of pairs $(a,b)$ for which
1832:   $f(a,z)g(b,z) > \Pd_{Z|A=aB=b}(z)$ is minimal (such functions exist
1833:   since $f(a,z) = g(b,z) = 1$ satisfy~(\ref{eq:40})).  We assume this
1834:   number is non-zero and obtain a contradiction.  For this,
1835:   let~$(a_1,b_1)$ be a pair for which $f(a_1,z)g(b_1,z) > 0$ and for
1836:   which the quotient $\Pd_{Z|A=a_1,B=b_1}(z)/f(a_1,z)g(b_1,z) < 1$ is
1837:   minimal.
1838: 
1839:   We define
1840:   \begin{align*}
1841:     f'(a,z) &:= \begin{cases}
1842:       f(a,z) & \text{if $a\neq a_1$,}\\
1843:       \frac{\Pd_{Z|A=a,B=b_1}(z)}{g(b_1,z)} & \text{if $a = a_1$}
1844:     \end{cases}\\
1845:     \intertext{and}
1846:     g'(b,z) &:= \begin{cases}
1847:       g(b,z) & \text{if $b\neq b_1$,}\\
1848:       \frac{\Pd_{Z|A=a_1,B=b}(z)}{f(a_1,z)} & \text{if $b=b_1$}.
1849:     \end{cases}
1850:   \end{align*}
1851:   We note that $f'$ and~$g'$ cannot take values larger than~$1$.  For
1852:   example, $f'(a_1,z)>1$ implies $\Pd_{Z|A=a_1,B=b_1}(z)>g(b_1,z) \geq
1853:   f(a_1,z) g(b_1,z)$, which contradicts~(\ref{eq:40}).  We further
1854:   claim that either the pair $(f',g)$ or $(f,g')$ still
1855:   satisfies~(\ref{eq:40}).  Otherwise, there are values~$a_2$
1856:   and~$b_2$ such that
1857:   \begin{align*}
1858:     f'(a_1,z)g(b_2,z) &= \frac{\Pd_{Z|A=a_1,B=b_1}(z)}{g(b_1,z)}
1859:     g(b_2,z) > \Pd_{Z|A=a_1 B=b_2}(z) \\
1860:     \intertext{and}
1861:     f(a_2,z)g'(b_1,z) &= f(a_2,z) \frac{\Pd_{Z|A=a_1,B=b_1}(z)}{f(a_1,z)}
1862:      > \Pd_{Z|A=a_2 B=b_1}(z),
1863:   \end{align*}
1864:   which implies
1865:   \begin{align*}
1866:     \frac{\Pd_{Z|A=a_1,B=b_1}(z)}{g(b_1,z)}\frac{\Pd_{Z|A=a_1,B=b_1}(z)}{f(a_1,z)}
1867:         g(b_2,z)f(a_2,z) > \Pd_{Z|A=a_2 B=b_1}(z) \Pd_{Z|A=a_1 B=b_2}(z),
1868:   \end{align*}
1869:   and using~(\ref{eq:16}) 
1870:   \begin{align*}
1871:     \frac{\Pd_{Z|A=a_1,B=b_1}(z)}{f(a_1,z)g(b_1,z)}
1872:      >\frac{\Pd_{Z|A=a_2,B=b_2}(z)}{g(b_2,z)f(a_2,z)}
1873:   \end{align*}
1874:   contradicting the way we chose $(a_1,b_1)$.  Thus, either $(f',g)$
1875:   or~$(f,g')$ still satisfies (\ref{eq:40}) and since the
1876:   respective version of (\ref{eq:40}) is satisfied with equality for
1877:   at least one more pair $(a,b)$ (namely for $(a_1,b_1)$) than for
1878:   which $(f,g)$ satisfies it, we get a contradiction.
1879: \end{proof}
1880: \end{document}
1881: 
1882: %%% Local Variables: 
1883: %%% mode: latex
1884: %%% TeX-master: t
1885: %%% End: 
1886: