1: \documentclass[11pt]{article}
2: \usepackage{fullpage, times}
3: \bibliographystyle{alpha}
4:
5: % for the notations in this paper
6: \newcommand{\defeq}{\stackrel{\Delta}{=}}
7: \newcommand{\PAR}{\mbox{{\sf PAR}}}
8: \newcommand{\GT}{\mbox{{\sf GT}}}
9: \newcommand{\Tr}{\mbox{{\rm Tr} }}
10: \newcommand{\ANN}{\mbox{{\sf ANN}}}
11: \newcommand{\ellone}[1]{\| #1 \|_1}
12: \newcommand{\E}{\mathop{\rm {\bf E}}}
13: \newcommand{\R}{{\mathbb R}}
14: \newcommand{\cH}{{\cal H}}
15: \newcommand{\cK}{{\cal K}}
16: \newcommand{\cI}{{\cal I}}
17: \newcommand{\cM}{{\cal M}}
18: \newcommand{\cP}{{\cal P}}
19: \newcommand{\tP}{\tilde{\cal P}}
20: \newcommand{\cQ}{{\cal Q}}
21: \newcommand{\cX}{{\cal X}}
22: \newcommand{\cY}{{\cal Y}}
23: \newcommand{\cZ}{{\cal Z}}
24: \newcommand{\ket}[1]{| #1 \rangle}
25: \newcommand{\ketbra}[1]{| #1 \rangle \langle #1 |}
26: \newcommand{\braket}[2]{\langle #1 | #2 \rangle}
27: \newcommand{\bzero}{{\mathbf 0}}
28:
29: % theorem-like environments in this paper
30: \newtheorem{definition}{Definition} % Specify Definition
31: \newtheorem{corollary}{Corollary} % Specify Corollary
32: \newtheorem{theorem}{Theorem} % Specify Theorem
33: \newtheorem{result}{Result} % Specify Result
34: \newtheorem{fact}{Fact} % Specify Fact
35: \newtheorem{lemma}{Lemma} % Specify Lemma
36: \newenvironment{proof}{\noindent{\bf Proof:}}{\qed} % Specify proof
37: \newcommand{\qed}{\hfill{$\rule{6pt}{6pt}$}} %Box at end of proof
38:
39: % Start of the paper
40: \title{Lower bounds for predecessor searching
41: in the cell probe model\footnote{
42: This paper is the journal version of the conference papers
43: \cite{sen:quantcell} and \cite{sen:pred}.
44: }
45: }
46:
47: \author{
48: Pranab Sen\thanks{
49: Department of Combinatorics and Optimization,
50: University of Waterloo,
51: Waterloo, Ontario N2L 3G1, Canada.
52: Email: {\sf p2sen@iqc.ca}.
53: Work done while the author was a graduate student at TIFR, Mumbai,
54: India visiting UC Berkeley and DIMACS under a Sarojini Damodaran
55: International Fellowship grant, and a
56: postdoctoral researcher at LRI, Orsay, France supported by
57: the EU 5th framework program QAIP IST-1999-11234 and
58: by CNRS/STIC 01N80/0502 and 01N80/0607 grants.
59: }
60: \and
61: S~.Venkatesh\thanks{
62: Algorithms and Complexity Group, Max-Planck Institut f\"{u}r
63: Informatik, Stuhlsatzenhausweg 85, 66123 Saarbr\"{u}ken, Germany.
64: Email: {\sf srini@mpi-sb.mpg.de}. Work done while the author
65: was a postdoctoral researcher at IAS, Princeton, USA supported
66: by NSF grant CCR-9987845 and a joint IAS-DIMACS post doctoral
67: fellowship.
68: }
69: }
70:
71: \date{}
72:
73: \begin{document}
74:
75: \maketitle
76: \thispagestyle{empty}
77:
78: \begin{abstract}
79: We consider a fundamental problem in data structures,
80: {\em static predecessor searching}: Given a subset $S$ of
81: size $n$ from the universe
82: $[m]$, store $S$ so that queries of the form ``What
83: is the predecessor of $x$ in
84: $S$?'' can be answered efficiently. We study this problem in
85: the cell probe
86: model introduced by Yao~\cite{yao:tablesort}. Recently,
87: Beame and Fich~\cite{beame:pred} obtained optimal bounds on the
88: number of probes
89: needed by any deterministic query scheme if the associated storage
90: scheme uses only $n^{O(1)}$ cells of word size $(\log m)^{O(1)}$ bits.
91:
92: We give a new lower bound proof for this problem that matches
93: the bounds of Beame and Fich. Our lower bound proof has
94: the following advantages: it
95: works for randomised query schemes too, while
96: Beame and Fich's proof works for deterministic query schemes only.
97: In addition, it is simpler than
98: Beame and Fich's proof.
99: In fact, our lower bound for predecessor searching
100: extends to the `quantum address-only' query schemes that we
101: define in this paper. In these query schemes, quantum parallelism
102: is allowed
103: only over the `address lines' of the queries. These query schemes
104: subsume classical randomised query schemes,
105: and include many quantum query algorithms
106: like Grover's algorithm~\cite{grover:search}.
107:
108: We prove our lower bound using the round elimination approach of
109: Miltersen, Nisan, Safra and Wigderson~\cite{miltersen:roundelim}.
110: Using tools from information theory, we prove a strong
111: round elimination lemma for communication complexity that enables
112: us to
113: obtain a tight lower bound for the predecessor problem.
114: Our strong round elimination lemma also extends to quantum
115: communication complexity.
116: We also use our
117: round elimination lemma to obtain a rounds versus communication
118: tradeoff for the `greater-than' problem, improving on the
119: tradeoff in \cite{miltersen:roundelim}.
120: We believe that
121: our round elimination lemma is of independent interest and should have
122: other applications.
123: \end{abstract}
124:
125: %\noindent {\bf Keywords:} Predecessor searching, cell probe
126: %model, data structures, communication complexity,
127: %information theory, lower bounds.
128:
129: \section{Introduction}
130: \subsection{The problem and the model}
131: A static data structure problem
132: consists of a set of data $D$, a set of queries $Q$, a set of
133: answers $A$, and a function
134: $f:D \times Q \rightarrow A$. The aim is to store the data efficiently
135: and succinctly,
136: so that any query can be answered with only a few probes to the data
137: structure.
138: {\em Static predecessor searching}
139: is a well studied problem in data
140: structure design (see e.g.~\cite{boas:priority, boas:order,
141: willard:loglog, andersson:lin, ajtai:pred, miltersen:union,
142: miltersen:roundelim, beame:pred}).
143: Data structures for answering predecessor queries can be used to
144: construct data structures to
145: answer other queries like rank (finding the number of elements in $S$
146: that are smaller than or equal to $x$) and nearest neighbour
147: (finding an
148: element in $S$ closest to $x$) efficiently. This motivates the need to
149: design efficient data structures that support predecessor queries.
150:
151: Let $[m]$ denote the set of integers $\{0, \ldots, m-1\}$.
152: \begin{definition}[Static predecessor searching]
153: In the problem of {\em $(m,n)$-static predecessor searching},
154: we are given a subset $S$ of size $n$ from the universe
155: $[m]$. Our goal is to store the set $S$ succinctly
156: so that
157: queries of the form ``What is the predecessor of $x$ in $S$?'' for
158: $x \in [m]$ can be answered with a few probes to the data structure.
159: If $x$ has no predecessor
160: in $S$, that is, $x$ is smaller than
161: every element in $S$, then return a default value, say, $-1$.
162: \end{definition}
163:
164: In this paper, we study the static predecessor searching problem in
165: Yao's {\em cell probe} model~\cite{yao:tablesort}.
166: The cell
167: probe model is a natural and general model for proving upper and lower
168: bound results in data structures.
169: \begin{definition}[The cell probe model]
170: An $(s,w,t)$ {\em cell probe scheme} for a
171: static data structure problem $f:D \times Q \rightarrow A$
172: has two components:
173: a {\em storage scheme} and a {\em query scheme}. The storage scheme
174: stores the data $d \in D$ as a table $T[d]$ of $s$
175: cells, each cell of word size $w$ bits.
176: The storage scheme is classical deterministic.
177: Given a query $q \in Q$, the query scheme computes $f(d, q)$ by
178: making at most $t$ probes to $T[d]$, where each probe reads one cell
179: at a time, and the probes can be adaptive.
180: In a deterministic cell probe scheme the query scheme is
181: classical deterministic,
182: in a randomised cell probe scheme it is classical randomised, and
183: in an address-only quantum cell probe scheme it is bounded error
184: address-only quantum (defined in Section~\ref{subsec:quantcell}).
185: \end{definition}
186: Since in the cell probe model we only charge a
187: scheme for the number of probes made to memory cells and for the
188: total number of cells of storage used, and all internal computation
189: is for free,
190: lower bounds proved in the cell probe model hold in all reasonable
191: data structure models (e.g. the unit cost RAM with the same word
192: size) and give us insight into the intrinsic difficulty
193: of the problem.
194:
195: The goal is to design cell probe schemes for $(m, n)$-static
196: predecessor searching using
197: small space i.e. $s = n^{O(1)}$ and $w = O(\log m)$, and at the
198: same time making a small number of probes $t$ in the worst case.
199:
200: \subsection{Previous work}
201: We start by describing the sequence of results that lead to the
202: currently
203: best known upper bounds for the $(m, n)$-static predecessor problem.
204: For a long time, the best upper bound known for the
205: predecessor problem
206: was due to the data structures of van Emde
207: Boas et al.~\cite{boas:priority, boas:order},
208: and the data structures of
209: Fredman and Willard~\cite{fredman:fusion}.
210: In their papers, van Emde
211: Boas et al.~\cite{boas:priority, boas:order}
212: gave a $(\Omega(m), O(\log m), O(\log \log m))$ deterministic
213: cell probe solution for predecessor.
214: The main drawback of their solution is that the number of cells
215: used is very large. Later, Willard~\cite{willard:loglog}
216: reduced the number of cells used to $O(n)$.
217: Building on the work of van Emde Boas et al. and Willard, Fredman and
218: Willard~\cite{fredman:fusion}, and
219: Andersson~\cite{andersson:lin} designed
220: $(O(n), O(\log m), O(\sqrt{\log n}))$
221: deterministic cell probe schemes for predecessor.
222: Recently, Beame and
223: Fich~\cite{beame:pred} improved on these upper
224: bounds and showed a
225: $\left(O\left(\frac{n^2 \log n}{\log \log n}\right),
226: O(\log m),t\right)$,
227: where
228: $t = \min\left\{O\left(\frac{\log\log m}{\log\log\log m}\right),
229: O\left(\sqrt{\frac{\log n}{\log \log n}}\,\right)\right\}$,
230: deterministic cell probe scheme for predecessor.
231:
232: The first lower bound for the $(m, n)$-static predecessor
233: problem was proved by Ajtai~\cite{ajtai:pred}, who showed that
234: no $(n^{O(1)}, O(\log m), t)$ deterministic cell probe scheme for
235: predecessor can have constant number of probes $t$.
236: Miltersen~\cite{miltersen:union} observed that there
237: is a close connection
238: between the cell probe complexity of a data structure problem and
239: the communication complexity of a related communication game, and
240: used this to improve Ajtai's lower bound to
241: $\Omega(\sqrt{\log\log m}\,)$ probes.
242: Recently, building on Ajtai's and Miltersen's work,
243: Beame and Fich~\cite{beame:pred}
244: showed that their data structure described above is indeed optimal
245: in the following sense: any
246: $(n^{O(1)}, 2^{(\log m)^{1 - \Omega(1)}}, t)$ deterministic cell
247: probe scheme for predecessor must satisfy
248: $t = \Omega\left(\frac{\log\log m}{\log\log\log m}\right)$
249: as a function of $m$, and any
250: $(n^{O(1)}, (\log m)^{O(1)}, t)$ deterministic cell
251: probe scheme for predecessor must satisfy
252: $t = \Omega\left(\sqrt{\frac{\log n}{\log \log n}}\,\right)$
253: as a function of $n$. Similar lower bounds
254: were proved by Xiao~\cite{xiao:cell}. We would like to
255: stress here that all the above lower bound proofs
256: are complicated with many
257: technical details. Also, they hold for deterministic cell
258: probe schemes only.
259:
260: The result of Beame and Fich gives rise to the following two
261: questions: does their lower bound hold for randomised query
262: schemes as well? It has been observed recently that randomisation
263: enormously helps in the case of membership
264: queries~\cite{buhrman:bitvectors} and
265: approximate nearest neighbour
266: queries~\cite{liu:ann, kushilevitz:ann, indyk:ann, chakrabarti:ann},
267: and one might believe that it
268: could help answer predecessor queries quickly as well. Secondly,
269: is it possible to give a simple proof of the lower bound result
270: of Beame and Fich?
271:
272: A partial answer to both of the above questions was given by
273: Miltersen, Nisan, Safra and Wigderson~\cite{miltersen:roundelim}.
274: In their paper, they
275: proved a general {\em round elimination} lemma for communication
276: complexity.
277: Using the connection between cell probe
278: complexity of data structures and communication complexity, and
279: their round elimination lemma, they showed
280: the following lower bound for the
281: predecessor problem: any $(n^{O(1)}, (\log m)^{O(1)}, t)$ randomised
282: cell probe scheme for $(m, n)$-static predecessor must
283: satisfy $t = \Omega(\sqrt{\log\log m}\,)$
284: as a function of $m$, and $t = \Omega((\log n)^{1/3})$
285: as a function of $n$. Though the
286: lower bound proved by \cite{miltersen:roundelim} is
287: weaker than that of
288: \cite{beame:pred}, their approach had two advantages: their lower
289: bound holds for randomised query schemes too,
290: and the proof is much simpler.
291: In their paper, Miltersen et al. ask if their round elimination
292: based approach can be strengthened
293: to obtain the lower bound of Beame and Fich.
294:
295: \subsection{Our Results}
296: We answer the question posed by Miltersen, Nisan, Safra and Wigderson.
297: Our main result in this paper shows that the lower bound of Beame and
298: Fich holds for address-only quantum cell probe schemes (and hence,
299: for randomised cell probe schemes) as well.
300:
301: \subsubsection{The Predecessor Problem}
302: \begin{result}
303: Suppose there is a $(n^{O(1)}, (\log m)^{O(1)}, t)$ randomised
304: cell probe
305: scheme for the $(m, n)$-static predecessor problem with error
306: probability less than $1/3$.
307: Then,
308: \begin{enumerate}
309: \item[(a)]
310: $t = \Omega \left( \frac{\log \log m}{\log \log \log m} \right)$
311: as a function of $m$;
312: \item[(b)]
313: $t = \Omega \left( \sqrt{\frac{\log n}{\log \log n}} \, \right)$
314: as a function of $n$.
315: \end{enumerate}
316: The same lower bound also holds for address-only quantum cell
317: probe schemes for static predecessor searching.
318: \end{result}
319:
320: We prove our lower bound for predecessor searching
321: by combining the approach in \cite{miltersen:roundelim}
322: with a new round elimination lemma for communication complexity.
323: Our round elimination lemma is a strengthening of the one proved
324: in ~\cite{miltersen:roundelim},
325: and we believe it is of independent interest.
326:
327: \subsubsection{An improved round elimination lemma}
328: In this paper, all communication protocols are two-party.
329: The error probability of a randomised or quantum communication
330: protocol is defined as the maximum error of the
331: protocol for any input.
332: For a general introduction to (classical)
333: communication complexity, see the
334: book by Kushilevitz and Nisan~\cite{kushilevitz:cc}.
335:
336: Let $f: \cX \times \cY \rightarrow \cZ$ be any
337: communication problem. Let
338: us denote by $f^{(n),A}$ a new communication game in which
339: Alice is given $x_1,x_2,\ldots,x_n \in \cX$, and Bob
340: is given $y \in \cY$, $i \in [n]$ and also
341: copies of $x_1, x_2, \ldots, x_{i-1}$.
342: Their task is to compute $f(x_i,y)$. $f^{(n), B}$ is defined
343: similarly. Intuitively,
344: if Alice starts the communication and her first message
345: is much smaller than $n$ bits, then she is unlikely to send
346: much useful information about $x_i$
347: to Bob as she is unaware of $i$. So it should be possible to
348: eliminate the first message of Alice, giving rise to a protocol
349: where Bob starts, with one less round of communication, and having
350: similar message complexity and error probability.
351: The round elimination lemma captures
352: this intuition.
353:
354: \begin{definition}
355: A $[t;l_1,l_2,\ldots,l_t]^A$ ($[t;l_1,l_2,\ldots,l_t]^B$)
356: communication protocol is one
357: where Alice (Bob) starts the communication, the
358: $i$th message is $l_i$ bits long, and the communication goes on
359: for $t$ rounds.
360: \end{definition}
361:
362: \begin{result}
363: Suppose the communication game $f^{(n),A}$ has a
364: $[t;l_1,l_2,\ldots,l_t]^A$
365: public coin randomised protocol with error less than
366: $\delta$. Then, $f$ has a $[t-1;l_2,\ldots,l_t]^B$ public coin
367: randomised protocol
368: with error less than
369: $\epsilon \defeq \delta + (1/2)(2l_1\ln 2/n)^{1/2}$.
370: A similar result holds for public coin quantum protocols (defined
371: in Section~\ref{subsec:quantcomm}) too.
372: \end{result}
373:
374: The proof of this lemma uses tools from information theory.
375: In particular,
376: we use the {\em average encoding} theorem of
377: Klauck, Nayak, Ta-Shma and Zuckerman~\cite{klauck:interaction}.
378: Intuitively, this theorem
379: says that if the mutual information between a random variable and its
380: randomised encoding is small, then the probability distributions
381: on code words for various values of the random variable
382: are indeed close to the average probability distribution on
383: code words.
384:
385: \subsubsection{Applications to other problems}
386: We prove our lower bound result for predecessor by actually proving
387: a lower bound for the {\em rank parity problem}. In the rank parity
388: problem, we need to store a subset $S$ of the universe $[m]$ so that
389: given a query element $x \in [m]$, we can output whether the number
390: of elements in $S$ less than or equal to $x$ is even or odd.
391: Lower bounds for rank parity imply similar lower bounds for some other
392: data structure problems like {\em point separation}~\cite{beame:pred}
393: and {\em two-dimensional reporting
394: range query}~\cite{miltersen:roundelim}. For
395: details of the reduction from rank parity to the above problems,
396: see the respective papers cited above.
397:
398: Independently, the round elimination lemma has applications
399: to problems
400: in communication complexity. For example, let us consider
401: communication protocols for the `greater-than' problem $\GT_n$
402: in which Alice and Bob are given bit strings $x$ and $y$ respectively
403: of length $n$ each, and
404: the goal is to find out if $x > y$ or not (treating
405: $x, y$ as integers between $0$ and $2^n - 1$).
406: Miltersen, Nisan, Safra and Wigderson~\cite{miltersen:roundelim},
407: and Smirnov~\cite{smirnov:shannon} have
408: studied rounds versus communication tradeoffs for $\GT_n$.
409: Miltersen et al. show an
410: $\Omega\left(n^{1/t} 2^{-O(t)}\right)$ lower bound for $t$-round
411: bounded error public coin randomised protocols for $\GT_n$.
412: Using our stronger round elimination lemma, we improve Miltersen
413: et al.'s result.
414: \begin{result}
415: The bounded error public coin randomised $t$-round
416: communication complexity of $\GT_n$
417: is lower bounded by $\Omega(n^{1/t}t^{-2})$.
418: For bounded error quantum protocols with input-independent
419: prior entanglement for
420: $\GT_n$, we have a lower bound of $\Omega(n^{1/t}t^{-1})$.
421: \end{result}
422:
423: \paragraph{Remark:} The lower
424: bound for quantum protocols is better because, by definition, a
425: quantum protocol always sends messages whose length is independent
426: of the input.
427:
428: There exists a bounded error classical randomised protocol for
429: $GT_n$ using $t$ rounds of communication and having a complexity of
430: $O(n^{1/t} \log n)$. Hence, for a constant number of rounds, our
431: lower bound matches the upper bound to within
432: logarithmic factors. For one round quantum protocols, our result
433: implies
434: an $\Omega(n)$ lower bound for $GT_n$ (which is optimal to within
435: constant factors), improving upon the
436: previous $\Omega(n/\log n)$ lower bound of
437: Klauck~\cite{klauck:ccsurvey}. No rounds versus communication tradeoff
438: for this problem, for more than one round,
439: was known earlier in the quantum setting.
440: If the number of rounds is unbounded, then there is a private
441: coin classical
442: randomised protocol for $GT_n$ using $O(\log n)$ rounds of
443: communication and having a complexity of
444: $O(\log n)$~\cite{nisan:threshold}. An
445: $\Omega(\log n)$ lower bound for the bounded error
446: quantum communication
447: complexity of $GT_n$ (irrespective of the number of rounds)
448: follows from Kremer's
449: result~\cite{kremer:quantcc} that the bounded error quantum
450: communication complexity of a function is lower bounded (up to
451: constant factors) by the
452: logarithm of the one round (classical)
453: deterministic communication complexity.
454:
455: \subsection{Our techniques}
456: The starting point of our work is the paper of Miltersen, Nisan, Safra
457: and Wigderson~\cite{miltersen:roundelim} showing lower
458: bounds for randomised
459: cell probe schemes for predecessor.
460: The crux of Miltersen et al.'s lower bound is the following
461: round elimination lemma for communication complexity.
462:
463: \bigskip
464:
465: \noindent {\bf Fact (Round elimination lemma,
466: \cite{miltersen:roundelim})}
467: {\em
468: Let $f:\cX \times \cY \rightarrow \cZ$ be a function.
469: Let $\epsilon, \delta > 0$ be real numbers. Suppose that
470: $\delta \leq \epsilon^2 (100 \ln (8/\epsilon))^{-1}$.
471: Suppose the communication game $f^{(n),A}$ has a
472: $[t;l_1,\ldots,l_t]^A$ public coin randomised protocol
473: with error less than $\delta$. Also suppose that
474: $n \geq 20 (l_1 \ln 2 + \ln 5) \epsilon^{-1}$.
475: Then there is a $[t-1;l_2,\ldots,l_t]^B$ public coin
476: randomised protocol for $f$ with error less than
477: $\epsilon$.
478: }
479:
480: \bigskip
481:
482: \noindent Note that in the round elimination lemma
483: of \cite{miltersen:roundelim}, the
484: dependence between $\delta$ and $\epsilon$ is quadratic.
485: In their paper, Miltersen et al. ask if their
486: round elimination based
487: approach can be strengthened to obtain Beame
488: and Fich's~\cite{beame:pred} lower bound.
489:
490: Our first observation
491: is that if we can prove a stronger round elimination lemma in
492: which $\delta$ and $\epsilon$ are related by a {\em small additive
493: term}, and the additive term is upper
494: bounded by $(\frac{l_1}{n})^{\Omega(1)}$,
495: then we can obtain the lower bound of Beame and Fich for
496: randomised cell probe schemes solving the predecessor problem.
497:
498: Our next observation is that Klauck, Nayak, Ta-Shma and
499: Zuckerman~\cite{klauck:interaction} have studied rounds
500: versus quantum communication tradeoffs
501: for the `tree pointer chasing' problem
502: using tools from quantum information theory.
503: In fact, their
504: quantum lower bound for the `tree pointer chasing' problem is
505: better than its previously known classical
506: lower bound~\cite{miltersen:roundelim}!
507: An important ingredient of their quantum lower bound was a
508: quantum information-theoretic result called
509: {\em average encoding} theorem. This result says informally
510: that when messages carry very
511: little information about the input, the average message is
512: essentially
513: as good as the individual messages. This result gives us a new way
514: of attacking the round elimination problem. The information-theoretic
515: round reduction
516: arguments in \cite{klauck:interaction} are
517: average-case (under the uniform
518: distribution on the inputs) arguments, and do not immediately
519: give a worst-case result like the round elimination lemma.
520: The information-theoretic arguments
521: have to be combined with
522: Yao's minimax lemma~\cite{yao:minimax} (also used in the
523: proof of the round elimination lemma in
524: \cite{miltersen:roundelim}) to prove
525: the strong round elimination lemma of this paper.
526: The information-theoretic approach
527: brings out more clearly the
528: intuition behind round elimination,
529: as opposed to the ad hoc combinatorial
530: proof in \cite{miltersen:roundelim}.
531: We believe that this strong round
532: elimination lemma is an important technical contribution
533: of this paper.
534:
535: \subsection{Organisation of the paper}
536: We start with some preliminaries in the next
537: section. Assuming the average encoding theorem, we prove an
538: intermediate result in Section~\ref{sec:rndreduce}
539: which allows us to reduce the number of rounds
540: of a communication protocol if the first message does not convey
541: much information about the sender's input.
542: Proofs of the classical and quantum versions of
543: the average encoding
544: theorem can be found in the appendix for completeness.
545: Using the intermediate result, we
546: prove our strong round elimination lemma in
547: Section~\ref{sec:roundelim}.
548: Sections~\ref{sec:rndreduce} and \ref{sec:roundelim}
549: each have two subsections: the first
550: one treats the classical version of the results and
551: the second one treats the quantum version.
552: Using the strong round elimination lemma, we prove the optimal
553: lower bound for the predecessor problem in Section~\ref{sec:predlb}.
554: The rounds versus communication tradeoff for the `greater-than'
555: problem is sketched in Section~\ref{sec:gt}.
556: We finally conclude
557: mentioning some open problems in Section~\ref{sec:conclusion}.
558:
559: \section{Preliminaries}
560:
561: \subsection{The address-only quantum cell probe model}
562: \label{subsec:quantcell}
563: A quantum {\em $(s,w,t)$ cell probe scheme} for a
564: static data structure
565: problem $f: D \times Q \rightarrow A$
566: has two components: a classical deterministic
567: {\em storage scheme} that
568: stores the data $d \in D$ in a table $T[d]$ using
569: $s$ cells each containing $w$ bits, and a quantum {\em query scheme}
570: that answers queries by `quantumly probing a cell at a time'
571: $t$ times.
572: Formally speaking, the table $T[d]$ is made available to the
573: query algorithm in the form of an oracle unitary transform $O_d$. To
574: define $O_d$ formally, we represent the basis states of the query
575: algorithm as
576: $\ket{j,b,z}$, where $j \in [s-1]$ is a binary string of
577: length $\log s$, $b$ is a binary string of length $w$, and
578: $z$ is a binary string of
579: some fixed length. Here,
580: $j$ denotes the address of a cell in the table $T[d]$,
581: $b$ denotes the qubits which will hold the contents of a cell
582: and $z$ stands for the rest of the qubits (`work qubits')
583: in the query algorithm.
584: $O_d$ maps $\ket{j,b,z}$ to $\ket{j, b \oplus T[d]_j, z}$, where
585: $T[d]_j$ is a bit string of length $w$ and
586: denotes the contents of the $j$th cell in $T[d]$.
587: A quantum query scheme with $t$ probes is just a sequence of
588: unitary transformations
589: \begin{displaymath}
590: U_{0} \rightarrow O_d \rightarrow U_{1} \rightarrow O_d
591: \rightarrow \ldots U_{t-1}
592: \rightarrow O_d \rightarrow U_{t}
593: \end{displaymath}
594: where $U_{j}$'s are arbitrary unitary transformations that do not
595: depend on $d$ ($U_{j}$'s represent the internal computations
596: of the query algorithm). For a query $q \in Q$, the
597: computation starts in a computational basis state
598: $\ket{q}\ket{0}$, where we
599: assume that the ancilla qubits are initially in the basis state
600: $\ket{0}$. Then we apply in succession,
601: the operators $U_0, O_d, U_1, \ldots, U_{t-1}, O_d, U_t$,
602: and measure the final state.
603: The answer consists of the values on some of the
604: output wires of the circuit.
605: We say that the scheme has worst case error probability less than
606: $\epsilon$ if
607: the answer is equal to $f(d,q)$, for
608: every $(d,q) \in D \times Q$, with
609: probability greater than $1 - \epsilon$.
610: The term `bounded error quantum scheme' means that
611: $\epsilon = 1/3$.
612:
613: We now formally define the {\em address-only quantum cell
614: probe model}.
615: Here the storage scheme is classical deterministic as before,
616: but the query scheme is restricted to be `address-only quantum'.
617: This means that the state vector before a query to the oracle $O_d$
618: is always a {\em tensor product} of a state vector
619: on the address and work qubits (the $(j,z)$ part in
620: $(j,b,z)$ above), and a state
621: vector on the data qubits (the $b$ part in
622: $(j,b,z)$ above). The state vector on the data qubits before a
623: query to the oracle $O_d$ is {\em independent of the
624: query element $q$ and the data $d$} but
625: can vary with the probe number. Intuitively, we are only
626: making use of quantum parallelism over the address lines of
627: a query. This
628: mode of querying a table subsumes classical (deterministic or
629: randomised) querying, and also many
630: non-trivial
631: quantum algorithms like Grover's algorithm~\cite{grover:search},
632: Farhi et al.'s algorithm~\cite{farhi:search},
633: H{\o}yer et al.'s algorithm~\cite{hoyer:search} etc.
634: satisfy the `address-only' condition.
635: For classical (deterministic or randomised)
636: querying, the state vector on the data qubits is
637: $\ket{0}$, independent of the probe number.
638: For Grover's algorithm and Farhi et al.'s algorithm,
639: the state vector on the data qubit is
640: $(\ket{0} - \ket{1}) / \sqrt{2}$, independent of the probe number.
641: For H{\o}yer et al.'s algorithm,
642: the state vector on the data qubit is
643: $\ket{0}$ for some probe numbers, and
644: $(\ket{0} - \ket{1}) / \sqrt{2}$ for the other probe numbers.
645:
646: \subsection{Quantum communication protocols}
647: \label{subsec:quantcomm}
648: In this paper, we adopt the `interacting unitary quantum circuits'
649: definition of quantum
650: communication protocols of Yao~\cite{yao:quantcc}.
651: Thus, Alice and Bob send a certain number of fixed length
652: messages to each other, and the number and length of these messages
653: is independent of their inputs.
654: If Alice's and Bob's inputs are in computational basis
655: states, the global state of all the qubits of Alice and Bob is pure
656: at all times during the execution of the protocol.
657: Measurements are not allowed during the execution of
658: the protocol. At the end of the protocol, the last recipient of
659: a message make a von Neumann measurement in the computational basis
660: of certain qubits (the `answer qubits')
661: in her possession in order to determine the answer
662: of the protocol. The choice of `answer qubits' is independent of
663: Alice's and Bob's inputs.
664:
665: We require that Alice and Bob
666: make a secure copy of their inputs before beginning the quantum
667: communication protocol.
668: This is possible since the inputs to Alice and Bob are in
669: computational basis states e.g. CNOT gates can be used for this
670: purpose. Thus, without loss of generality,
671: the input qubits of Alice and Bob are never sent
672: as messages, their state remains unchanged throughout the protocol,
673: and they are never
674: measured i.e. some work qubits are measured to determine the result of
675: the protocol.
676: We call such quantum protocols {\em secure}
677: and will assume
678: henceforth that all our quantum protocols are secure.
679:
680: We now define the concept of a {\em safe}
681: quantum communication
682: protocol, which will be used in the statement of the
683: quantum round elimination lemma.
684: \begin{definition}[Safe quantum protocol]
685: A $[t;c;l_1,\ldots,l_t]^A$ ($[t;c;l_1,\ldots,l_t]^B$) safe quantum
686: communication
687: protocol is a secure quantum protocol where Alice (Bob) starts
688: the communication, the first message is $l_1 + c$ qubits long,
689: the $i$th message, for $i \ge 2$, is $l_i$ qubits long, and
690: the communication goes on for $t$ rounds. We think of the
691: first message as having two parts: the `main part' which is
692: $l_1$ qubits long, and the `safe overhead part' which is $c$ qubits
693: long. The density matrix of the `safe overhead' is independent of the
694: inputs to Alice and Bob.
695: \end{definition}
696:
697: \paragraph{Remarks:} \ \\
698: 1. \ \ The safe overhead is nothing but a way to send
699: input independent
700: prior entanglement from Alice to Bob. The reason we use
701: this notation is that we will later define safe public coin
702: quantum protocols where there will be two kinds of input
703: independent prior entanglement, the first being the one provided
704: by the safe overhead and the second being the one proved by the
705: public coin. \\
706: 2. \ \ The reason for defining the concept of a safe overhead,
707: intuitively speaking, is as follows.
708: The communication games arising from data structure problems
709: often have an asymmetry between the message lengths of Alice and Bob.
710: This asymmetry is crucial to prove lower bounds on the number
711: of rounds of communication.
712: In the previous quantum round reduction arguments (e.g. those
713: of Klauck et al.~\cite{klauck:interaction}), the complexity
714: of the first message in the protocol increases quickly as the
715: number of rounds is reduced and the asymmetry gets lost.
716: This leads to a problem where the
717: first message soon gets big enough to potentially
718: convey substantial information about the input of one player to the
719: other, destroying any hope of proving strong lower bounds on the
720: number of rounds.
721: The concept of a safe protocol allows us to get around this problem.
722: We show through a careful
723: quantum information theoretic analysis of the round reduction process,
724: that in a safe protocol,
725: though the complexity of the first message increases a lot, this
726: increase is confined to the safe overhead and so, the
727: information content does not increase much. This is the key property
728: which allows us to prove a round elimination lemma for safe
729: quantum protocols.
730:
731: In this paper we will deal with quantum protocols with {\em public
732: coins}. Intuitively, a public coin quantum protocol is
733: a probability distribution over finitely many ({\em coinless})
734: quantum protocols. We shall
735: henceforth call the standard definition of a quantum protocol
736: without prior entanglement as {\em coinless}. Our definition
737: is similar to the classical scenario, where a randomised protocol
738: with public coins is a probability distribution over finitely
739: many deterministic
740: protocols. We note however, that our definition of a
741: public coin quantum
742: protocol is {\em not} the same as that of a
743: quantum protocol with prior
744: entanglement, which has been studied previously
745: (see e.g.~\cite{cleve:ip}). Our definition is weaker, in that it
746: does not allow the unitary transformations of Alice and Bob to
747: alter the `public coin'.
748: \begin{definition}[Public coin quantum protocol]
749: In a quantum protocol with a public coin, there is, before the
750: start of the protocol, a
751: quantum state called a {\em public coin}, of the form
752: $\sum_c \sqrt{p_c} \ket{c}_A \ket{c}_B$, where the
753: subscripts denote ownership of qubits by Alice and Bob,
754: $p_c$ are finitely many non-negative real
755: numbers and $\sum_c p_c = 1$.
756: Alice and Bob make (entangled) copies of their respective
757: halves of the
758: public coin using {\sc CNOT} gates before commencing the
759: protocol. The unitary transformations of Alice and Bob during the
760: protocol do not touch the public coin.
761: The public coin is never measured, nor is it ever sent as a message.
762: \end{definition}
763: Hence, one can
764: think of the public coin quantum protocol to be a probability
765: distribution, with probability $p_c$, over finitely many
766: coinless quantum protocols
767: indexed by the coin basis states $\ket{c}$. A {\em safe public coin}
768: quantum protocol is similarly defined as
769: a probability distribution over finitely many safe
770: coinless quantum protocols.
771:
772: \paragraph{Remarks:} \ \\
773: 1.\ \ We need to define public coin quantum protocols in order to
774: make use of the harder direction of Yao's minimax
775: lemma~\cite{yao:minimax}. The minimax lemma
776: is the main tool which allows us to convert `average case'
777: round reduction arguments to `worst case' arguments. We need `worst
778: case' round reduction arguments in proving lower
779: bounds for the rounds
780: complexity of communication games arising from data
781: structure problems. This is because many of these lower
782: bound proofs
783: use some notion of ``self-reducibility'', arising from the
784: original data
785: structure problem, which fails to hold in the `average case' but holds
786: for the `worst case'. The quantum round reduction arguments of
787: Klauck et al.~\cite{klauck:interaction}
788: are `average case' arguments, and this is
789: one of the reasons why they do not suffice to prove lower bounds
790: for the rounds complexity of communication games arising from data
791: structure problems. \\
792: 2.\ \ For $[t;c;l_1,\ldots,l_t]^A$ safe quantum protocols computing
793: a function $f$, Yao's minimax lemma says that
794: the infimum (worst-case) error of a public coin protocol is equal
795: to the supremum over all input probability distributions of
796: the infimum distributional error of a coinless protocol.
797:
798: \subsection{Predecessor searching and communication complexity}
799: \label{subsec:mnswpredlb}
800: We first describe the connection between the address-only
801: quantum cell probe
802: complexity of a static data structure problem and the quantum
803: communication complexity of an associated communication game.
804: Let $f: D \times Q \rightarrow A$ be a static data structure problem.
805: Consider a two-party
806: communication problem where Alice is given a query $q \in Q$,
807: Bob is given a datum $d \in D$, and they have to communicate and
808: find out the answer $f(d,q)$. We have the following lemma, which
809: is a quantum analogue of a lemma of Miltersen~\cite{miltersen:union}
810: relating cell probe complexity to communication complexity in
811: the classical setting.
812: \begin{lemma}
813: \label{lem:queryandcc}
814: Suppose we have an $(s,w,t)$ quantum cell
815: probe solution to a static data structure problem
816: $f: D \times Q \rightarrow A$.
817: Then we have a
818: $(2t, 0, \log s + w, \log s + w)^A$ safe coinless quantum
819: protocol for the corresponding
820: communication problem. If the query scheme is address-only,
821: we also have a $(2t, 0, \log s, \log s + w)^A$ safe coinless
822: quantum protocol for the corresponding communication problem.
823: The error probability of the communication protocol is the same as
824: that of the cell probe scheme.
825: \end{lemma}
826: \begin{proof}
827: Given a quantum $(s,w,t)$ cell
828: probe solution to the static data structure problem $f$,
829: we can get a
830: $(2t, 0, \log s + w, \log s + w)^A$ safe coinless quantum
831: protocol for the corresponding
832: communication problem by just
833: simulating the cell probe solution. If in addition,
834: the query scheme is address-only, the messages from Alice to
835: Bob need consist only of the `address' part. This can be
836: seen as follows. Let the state vector of the data qubits before
837: the $i$th query be $\ket{\theta_i}$. $\ket{\theta_i}$ is independent
838: of the query element and the stored data.
839: Bob keeps $t$ special ancilla registers in states
840: $\ket{\theta_i}, 1 \leq i \leq t$ at the start of the protocol $P$.
841: These special ancilla registers are in tensor with the rest of the
842: qubits of Alice and Bob at the start of $P$.
843: Protocol $P$ simulates the cell probe solution, but with the following
844: modification. To simulate the $i$th query of the cell probe
845: solution, Alice prepares her `address' and `data' qubits as in the
846: query scheme, but sends the `address' qubits only. Bob treats those
847: `address' qubits together with $\ket{\theta_i}$ in the $i$th
848: special ancilla register as Alice's query, and performs the oracle
849: table transformation on them. He then sends these qubits (both
850: the `address' as well as the $i$th special register qubits) to Alice.
851: Alice exchanges the contents of the $i$th special register with
852: her `data' qubits (i.e. exchanges the basis states), and proceeds
853: with the simulation of the query scheme. This gives us a
854: $(2t, 0, \log s, \log s + w)^A$ safe coinless
855: quantum protocol with the same error probability as
856: that of the cell probe query scheme.
857: \end{proof}
858:
859: \paragraph{Remark:}
860: In many natural data structure problems $\log s$ is much smaller than
861: $w$ and thus, in the address-only quantum case, we get a
862: $(2t, 0, \log s, O(w))^A$ safe protocol. In the classical
863: setting of \cite{miltersen:union}, one gets a
864: $(2t, 0, \log s, w)^A$ protocol.
865: This asymmetry in message lengths is crucial
866: in proving non-trivial lower bounds on $t$.
867: The concept of a safe quantum protocol helps us in exploiting
868: this asymmetry.
869:
870: We now recall some facts about the connection between
871: cell probe schemes for predecessor and communication
872: complexity of rank parity
873: from \cite{miltersen:roundelim}. We give proof sketches of these
874: facts for completeness.
875:
876: \begin{definition}
877: A $(t,a,b)^A$ ($(t,a,b)^B$) classical communication
878: protocol is a
879: $[t;l_1,\ldots,l_t]^A$ ($[t;l_1,\ldots,l_t]^B$)
880: classical protocol, where assuming
881: Alice (Bob) starts, $l_i = a$ for $i$ odd and $l_i = b$ for
882: $i$ even ($l_i = b$ for $i$ odd and $l_i = a$ for $i$ even).
883: A $(t,c,a,b)^A$ ($(t,c,a,b)^B$) quantum communication
884: protocol is a
885: $[t;c;l_1,\ldots,l_t]^A$ ($[t;c;l_1,\ldots,l_t]^B$)
886: safe quantum protocol, where assuming
887: Alice (Bob) starts, $l_i = a$ for $i$ odd and $l_i = b$ for
888: $i$ even ($l_i = b$ for $i$ odd and $l_i = a$ for $i$ even).
889: \end{definition}
890:
891: \begin{definition}[Rank parity]
892: In the {\em rank parity} communication
893: game $\PAR_{p,q}$,
894: Alice is given a bit string $x$ of
895: length $p$, Bob is given a set $S$ of bit strings of length $p$,
896: $|S| \leq q$, and they have to communicate and decide
897: whether the rank of $x$ in $S$ (treating the bit
898: strings as integers) is odd or even.
899: By the rank of $x$ in $S$, we mean the cardinality of the
900: set $\{y \in S \mid y \leq x \}$.
901: In the communication game $\PAR^{(k),A}_{p,q}$,
902: Alice is given $k$ bit strings $x_1,\ldots,x_k$ each of
903: length $p$, Bob is given a set $S$ of bit strings of length $p$,
904: $|S| \leq q$, an index $i \in [k]$,
905: and copies of $x_1,\ldots,x_{i-1}$;
906: they have to communicate and decide
907: whether the rank of $x_i$ in $S$ is odd or even.
908: In the communication game $\PAR^{(k),B}_{p,q}$,
909: Alice is given a bit string $x$ of
910: length $p$ and an index $i \in [k]$, Bob is given $k$ sets
911: $S_1,\ldots,S_k$ of bit strings of length $p$,
912: $|S_j| \leq q, 1 \leq j \leq k$;
913: they have to communicate and decide
914: whether the rank of $x$ in $S_i$ is odd or even.
915: \end{definition}
916:
917: \begin{fact}
918: \label{fact:pred2rank}
919: Let $m$ be a positive integer such that $m$ is a power of $2$.
920: Suppose that there is a $(n^{O(1)}, (\log m)^{O(1)}, t)$
921: randomised (address-only quantum) cell
922: probe scheme for the $(m, n)$-static predecessor problem.
923: Then the rank parity communication game $\PAR_{\log m, n}$ has a
924: $\left(2t+O(1), O(\log n), (\log m)^{O(1)} \right)^A$
925: private coin randomised (safe coinless quantum) protocol.
926: The error probability of the communication protocol is the same as
927: that of the cell probe scheme.
928: \end{fact}
929: \begin{proof}
930: Consider the static rank parity data structure problem where
931: the storage scheme has to store a set $S \subseteq [m]$, $|S| \leq n$,
932: and the query scheme, given a query $x \in [m]$, has to decide whether
933: the rank of $x$ in $S$ is odd or even.
934: Fredman, Koml\'{o}s and Szemer\'{e}di~\cite{fredman:hashing}
935: have shown the existence of two-level
936: perfect hash tables containing, for each member $y$ of the
937: stored subset $S$, $y$'s rank in $S$, and
938: using $O(n)$ cells of word size $O(\log m)$ and requiring
939: only $O(1)$ deterministic cell probes.
940: Combining a $(n^{O(1)}, (\log m)^{O(1)}, t)$
941: cell probe solution to the static predecessor problem with such
942: a perfect hash table gives us a
943: $(n^{O(1)} + O(n), \max((\log m)^{O(1)}, O(\log m)), t + O(1))$
944: cell probe solution to the static rank parity problem.
945: The error probability of the cell probe scheme for the rank
946: parity problem is the same as the error probability of the cell
947: probe scheme for the predecessor problem.
948: Converting the cell probe scheme for the rank parity problem
949: into a communication protocol by \cite{miltersen:union}
950: (by Lemma~\ref{fact:pred2rank}), we get a
951: $(2t+O(1), O(\log n), (\log m)^{O(1)})^A$ private coin
952: randomised (safe coinless quantum) protocol for the rank
953: parity communication game $\PAR_{\log m, n}$.
954: The error probability of the communication protocol is the same as
955: that of the cell probe scheme for the predecessor problem.
956: \end{proof}
957:
958: \begin{fact}
959: \label{fact:rankred1}
960: Let $k, p$ be positive integers such that $k \mid p$.
961: A communication protocol with Alice starting for $\PAR_{p,q}$
962: gives us a communication protocol with
963: Alice starting for $\PAR^{(k),A}_{p/k,q}$ with the
964: same message complexity, number of rounds
965: and error probability.
966: \end{fact}
967: \begin{proof}
968: Consider the problem $\PAR^{(k),A}_{p/k,q}$.
969: Alice, who is given $x_1,\ldots,x_k$, computes the concatenation
970: $\widehat{x} \defeq x_1 \cdot x_2 \cdots x_k$. Bob, who is given
971: $S$, $i$ and $x_1,\ldots,x_{i-1}$, computes
972: \begin{displaymath}
973: \widehat{S} \defeq
974: \left\{x_1 \cdot x_2 \cdots x_{i-1}
975: \cdot y \cdot 0^{p(1 - i/k)}: y \in S
976: \right\}.
977: \end{displaymath}
978: After this, Alice and Bob run the protocol for $\PAR_{p,q}$ on inputs
979: $\widehat{x}$, $\widehat{S}$ to solve the
980: problem $\PAR^{(k),A}_{p/k,q}$.
981: \end{proof}
982:
983: \begin{fact}
984: \label{fact:rankred2}
985: Let $k, q$ be positive integers such that
986: $k \mid q$ and $k$ is a power of $2$.
987: A communication protocol with Bob starting for $\PAR_{p,q}$
988: gives us a communication protocol with Bob starting for
989: $\PAR^{(k),B}_{p - \log k - 1,q/k}$ with the same message complexity,
990: number of rounds and error probability.
991: \end{fact}
992: \begin{proof}
993: Consider the problem $\PAR^{(k),B}_{p - \log k - 1,q/k}$.
994: Alice, who is given $x$ and $i$,
995: computes the concatenation $\hat{x} \defeq (i-1) \cdot 0 \cdot x$.
996: Bob, who is given $S_1,\ldots,S_k$, computes the sets
997: $S'_1,\ldots,S'_k$ where
998: \begin{displaymath}
999: S'_j \defeq
1000: \left\{ \begin{array}{l}
1001: \left\{(j-1) \cdot 0 \cdot y: y \in S_j \right\}\\
1002: ~~~~~~~~~~~~~~~~~~~~ \mbox{if $|S_j|$ is even}, \\
1003: \left\{(j-1) \cdot 0 \cdot y: y \in S_j \right\}
1004: \bigcup \left\{(j-1) \cdot 1^{p - \log k} \right\}\\
1005: ~~~~~~~~~~~~~~~~~~~~ \mbox{if $|S_j|$ is odd}.
1006: \end{array}
1007: \right.
1008: \end{displaymath}
1009: Above, the integers $(i-1), (j-1)$ are to be thought of as
1010: bit strings of length $\log k$.
1011: Bob also computes $\hat{S} \defeq \bigcup_{j=1}^k S'_j$.
1012: Alice and Bob then run the protocol for $\PAR_{p,q}$ on inputs
1013: $\hat{x}$, $\hat{S}$ to solve the problem
1014: $\PAR^{(k),B}_{p - \log k - 1,q/k}$.
1015: \end{proof}
1016:
1017: \subsection{Some classical information theoretic facts}
1018: \label{sec:classicalinfoprelim}
1019: In this subsection, we discuss some classical
1020: information theoretic facts
1021: which will be used in the proof of our improved classical
1022: round elimination lemma.
1023: For a good account of classical information theory, see
1024: the book by Cover and Thomas~\cite{cover:infotheory}.
1025:
1026: In this paper, all random variables have finite range and all
1027: sample spaces have finite cardinality.
1028: Let $X,Y,Z$ be random variables with some joint distribution.
1029: The {\em Shannon entropy} of $X$ is defined as
1030: $H(X) \defeq -\sum_x \Pr[X=x] \log \Pr[X=x]$.
1031: The {\em mutual information} of $X$ and $Y$ is defined
1032: as $I(X:Y) \defeq H(X) + H(Y) - H(XY)$. If the range of $X$
1033: has cardinality at most $d$, $I(X:Y) \leq \log d$.
1034: $I(X:Y) = 0$ iff $X$ and $Y$ are independent.
1035: Let $I((X:Y) \mid Z = z)$ denote the mutual information of
1036: $X$ and $Y$ conditioned on the event $Z=z$.
1037:
1038: The next fact follows easily from the definitions.
1039: \begin{fact}
1040: \label{fact:classicalaveraging}
1041: Let $X,Y,Z$ be random variables with some joint distribution.
1042: Then,
1043: \begin{enumerate}
1044: \item[(a)] $I(XY:Z) = I(X:Z) + I(Y:ZX) - I(X:Y)$.
1045: In particular, if $X, Y$ are independent
1046: $I(XY:Z) = I(X:Z) + I(Y:ZX)$.
1047: \item[(b)] $\displaystyle
1048: I(Y:ZX) = I(X:Y) + \E_{X} [I((Y:Z) \mid X=x)]$, where
1049: is expectation is over the marginal distribution of $X$.
1050: \end{enumerate}
1051: \end{fact}
1052:
1053: We use total variation distance to quantify the distance between
1054: two probability distributions.
1055: \begin{definition}[Total variation distance]
1056: Let $P,Q$ be probability distributions on the same
1057: sample space $\Omega$.
1058: The {\em total variation distance} (also known as the
1059: $\ell_1$-distance) between
1060: $P$ and $Q$, denoted by $\ellone{P - Q}$, is defined as
1061: $\displaystyle \ellone{P - Q} \defeq \sum_{x \in \Omega} |P(x)-Q(x)|$.
1062: \end{definition}
1063:
1064: We will need the {\em average encoding} theorem of
1065: Klauck, Nayak, Ta-Shma and Zuckerman~\cite{klauck:interaction}.
1066: Klauck et al. actually prove a quantum version of this
1067: theorem in their paper, but we will use the classical version
1068: in the proof of our round elimination lemma.
1069: Intuitively speaking, the theorem says that if the mutual information
1070: between a random variable and its randomised encoding is small,
1071: then the probability distributions on the code words
1072: for various values of the random variable
1073: are close to the average probability distribution on the code words.
1074: \begin{fact}[Average encoding theorem, \cite{klauck:interaction}]
1075: \label{fact:averageencodingclassical}
1076: Let $X, M$ be correlated
1077: random variables. Let $p_x$ denote the (marginal) probability
1078: that $X=x$, and $\Pi^x$ denote the conditional distribution of $M$
1079: given that $X=x$.
1080: Let $\Pi$ denote the (marginal) probability distribution of $M$ i.e.
1081: $\Pi = \sum_x p_x \Pi^x$.
1082: Then,
1083: \begin{displaymath}
1084: \sum_x p_x \ellone{\Pi^x - \Pi} \leq \sqrt{(2 \ln 2) I(X:M)}.
1085: \end{displaymath}
1086: \end{fact}
1087: A self-contained classical
1088: proof, without using quantum information theory,
1089: of this fact can be found in the appendix for completeness.
1090:
1091: \subsection{Some quantum information theoretic facts}
1092: \label{sec:quantinfoprelim}
1093: In this subsection, we discuss some quantum
1094: information theoretic facts
1095: which will be used in the proof of our quantum
1096: round elimination lemma.
1097: For a good account of quantum information theory, see
1098: the book by Nielsen and Chuang~\cite{nielsen:quant}.
1099:
1100: In this paper, all quantum systems, Hilbert spaces and superoperators
1101: are finite dimensional.
1102: Let $X,Y$ be quantum systems with some joint density matrix
1103: $\rho_{XY}$.
1104: If $\rho_X$ is the reduced density matrix of $X$,
1105: the {\em von Neumann entropy} of $X$ is defined as
1106: $S(X) \defeq -\Tr \rho_X \log \rho_X$.
1107: The {\em mutual information} of $X$ and $Y$ is defined
1108: as $I(X:Y) \defeq S(X) + S(Y) - S(XY)$. If $X$
1109: is at most $d$-dimensional, $I(X:Y) \leq 2 \log d$.
1110: $I(X:Y) = 0$ iff $X$ and $Y$ are independent i.e.
1111: $\rho_{XY} = \rho_X \otimes \rho_Y$.
1112: Suppose $X,Y,Z$ are quantum systems with some joint density matrix
1113: where $Z$ is a classical random variable i.e. the reduced density
1114: matrix of $Z$ is diagonal in the computational basis.
1115: Let $I((X:Y) \mid Z = z)$ denote the mutual information of
1116: $X$ and $Y$ conditioned on the event $Z=z$.
1117:
1118: The next fact follows easily from the definitions.
1119: \begin{fact}
1120: \label{fact:quantaveraging}
1121: Let $X,Y,Z,W$ be quantum systems with some joint density matrix,
1122: where $X$ and $Y$ are classical random variables.
1123: Then,
1124: \begin{enumerate}
1125: \item[(a)] $I(XY:Z) = I(X:Z) + I(Y:ZX) - I(X:Y)$.
1126: In particular, if $X, Y$ are independent
1127: $I(XY:Z) = I(X:Z) + I(Y:ZX)$.
1128: \item[(b)] $\displaystyle
1129: I(Y:ZX) = I(X:Y) + \E_{X} [I((Y:Z) \mid X=x)]$, where
1130: is expectation is over the marginal distribution of $X$.
1131: \item[(c)] Suppose $W$ is independent of $X$ and $Z$ is supported
1132: on $m$ qubits. Then, $I(X:ZW) \leq 2m$.
1133: \end{enumerate}
1134: \end{fact}
1135:
1136: \paragraph{Remarks:} \ \\
1137: 1. \ \ Fact~\ref{fact:quantaveraging}(c) is the key observation
1138: allowing us to
1139: ``ignore'' the size of the ``safe'' overhead $W$
1140: in quantum round elimination applications.
1141: In these applications, the complexity of the first message
1142: in the protocol increases quickly, but the blow up is confined to
1143: the ``safe'' overhead. Earlier round reduction arguments were unable
1144: to handle this large blow up in the complexity of the first message.\\
1145: 2. \ \ In Fact~\ref{fact:quantaveraging}(c), if $Z$ is a
1146: classical random variable, we get the improved
1147: inequality $I(X:ZW) \leq m$.
1148:
1149: We use trace distance to quantify the distance between
1150: two density matrices.
1151: \begin{definition}[Trace distance]
1152: Let $\rho,\sigma$ be density matrices in the same Hilbert space.
1153: The {\em trace distance} between
1154: $\rho$ and $\sigma$, denoted by $\ellone{\rho - \sigma}$,
1155: is defined as
1156: $\ellone{\rho - \sigma} \defeq
1157: \Tr \sqrt{(\rho - \sigma)^\dagger (\rho - \sigma)}$.
1158: \end{definition}
1159: If $\rho$ is a density matrix in a Hilbert space $\cH$ and
1160: $\cM$ is a general measurement i.e. a POVM on $\cH$, let
1161: $\cM \rho$ denote the probability distribution on
1162: the (classical) outcomes of $\cM$ got by performing measurement
1163: $\cM$ on $\rho$.
1164: The importance of the trace distance as a metric on density
1165: matrices stems from the following fundamental fact
1166: (see e.g.~\cite{aharonov:mixed}).
1167: \begin{fact}
1168: \label{fact:l1tracedist}
1169: Let $\rho_1,\rho_2$ be two density matrices in the same Hilbert
1170: space $\cH$. Let $\cM$ be a POVM on $\cH$. Then,
1171: $\ellone{\cM\rho_1 - \cM\rho_2} \leq \ellone{\rho_1 - \rho_2}$.
1172: \end{fact}
1173:
1174: Let $\cH, \cK$ be disjoint Hilbert spaces such that
1175: $\dim(\cK) \geq \dim(\cH)$. Let $\rho$ be a density matrix in
1176: $\cH$ and $\ket{\psi}$ be a pure state in $\cH \otimes \cK$.
1177: $\ket{\psi}$ is said to be a {\em purification} of $\rho$
1178: if $\Tr_{\cK} \ketbra{\psi} = \rho$. We will require the following
1179: basic fact about two purifications of the same density matrix.
1180: \begin{fact}
1181: \label{fact:localtransition}
1182: Let $\cH, \cK$ be disjoint Hilbert spaces such that
1183: $\dim(\cK) \geq \dim(\cH)$. Let $\rho$ be a density matrix in
1184: $\cH$ and $\ket{\psi}, \ket{\phi}$ be two purifications of
1185: $\rho$ in $\cH \otimes \cK$. Then there is a local
1186: unitary transformation $U$ on $\cK$ such that
1187: $\ket{\psi} = (I \otimes U) \ket{\phi}$, where $I$ is the
1188: identity operator on $\cH$.
1189: \end{fact}
1190:
1191: We now state an improved version of the quantum
1192: {\em average encoding} theorem of
1193: Klauck, Nayak, Ta-Shma and Zuckerman~\cite{klauck:interaction}.
1194: This improved
1195: version follows from a direct connection between fidelity and
1196: relative entropy described in \cite{dacunha:BgeqS}, and was
1197: also independently observed by Klauck (private communication).
1198: Intuitively speaking, the theorem says that if the mutual information
1199: between a random variable and its quantum encoding is small,
1200: then given any purifications of the quantum code words
1201: for various values of the random variable,
1202: one can find purifications of the average code word that are close
1203: to the respective purifications of the quantum code words.
1204: \begin{fact}
1205: \label{fact:averageencodingquant}
1206: Let $X, M$ be quantum systems with some joint density matrix.
1207: Let $X$ be a classical random variable
1208: and $p_x$ denote the (marginal) probability
1209: that $X=x$. Let $\cH$ denote the Hilbert space of $M$
1210: and $\rho^x$ denote the conditional density matrix
1211: of $M$ given that $X=x$.
1212: Let $\rho$ denote the reduced density matrix of $M$. Note that
1213: $\rho = \sum_x p_x \rho^x$.
1214: Let $\cK$ denote a disjoint
1215: Hilbert space such that $\dim(\cK) \geq \dim(\cH)$.
1216: Let $\ket{\psi^x}$ be purifications of $\rho^x$ in
1217: $\cH \otimes \cK$. Then there exist purifications $\ket{\phi^x}$
1218: of $\rho$ in $\cH \otimes \cK$ such that
1219: \begin{displaymath}
1220: \sum_x p_x \ellone{\ketbra{\psi^x} - \ketbra{\phi^x}} \leq
1221: \sqrt{(4 \ln 2) I(X:M)}.
1222: \end{displaymath}
1223: \end{fact}
1224: A proof of this fact can be found in the appendix
1225: for completeness.
1226:
1227: \section{Reducing the number of rounds}
1228: \label{sec:rndreduce}
1229: In this section, we prove an intermediate result
1230: which will be required to prove our strong round elimination lemmas.
1231: In Section~\ref{subsec:classicalrndreduce} we prove the
1232: intermediate result in its classical version
1233: (Lemma~\ref{lem:classicalroundreduce}),
1234: whereas in Section~\ref{subsec:quantrndreduce} we prove the
1235: intermediate result in its quantum version
1236: (Lemma~\ref{lem:quantroundreduce}).
1237: The proof of Lemma~\ref{lem:classicalroundreduce}
1238: is similar to the proof of Lemma~4.4 in
1239: \cite{klauck:interaction} (see also \cite{maneva:interaction}),
1240: but much simpler since we are in the classical setting.
1241: The proof of Lemma~\ref{lem:quantroundreduce}
1242: is a slight refinement of the proof of Lemma~4.4 in
1243: ~\cite{klauck:interaction}, and is included here for completeness.
1244: Intuitively speaking, the intermediate result
1245: says that if the first message
1246: in a communication protocol
1247: carries little information about the sender's
1248: input, under some probability
1249: distribution on Alice's and Bob's inputs,
1250: then it can be eliminated, giving rise
1251: to a protocol where the other player starts, with one less
1252: round of communication, smaller message complexity, and with similar
1253: average error probability with respect to the same probability
1254: distribution on Alice's and Bob's inputs.
1255:
1256: Consider a communication protocol $\cP$ computing
1257: a function $f: \cX \times \cY \rightarrow \cZ$.
1258: For an input $(x,y) \in \cX \times \cY$, we define the error
1259: $\epsilon^{\cP}_{x,y}$ of
1260: $\cP$ on $(x,y)$ to be the probability
1261: that the result of $\cP$ on input $(x,y)$ is not
1262: equal to $f(x,y)$. For
1263: a protocol $\cP$, given a probability distribution
1264: $D$ on $\cX \times \cY$, we define the
1265: average error
1266: $\epsilon^{\cP}_D$ of $\cP$ with respect to $D$ as the expectation
1267: over $D$ of the error of $\cP$ on inputs $(x,y) \in \cX \times \cY$.
1268: We define $\epsilon^\cP$ to be the maximum error of $\cP$ on inputs
1269: $(x,y) \in \cX \times \cY$ i.e.
1270: $\epsilon^\cP$ is the error of protocol $\cP$.
1271:
1272: \subsection{Classical}
1273: \label{subsec:classicalrndreduce}
1274: \begin{lemma}
1275: \label{lem:classicalroundreduce}
1276: Suppose $f: \cX \times \cY \rightarrow \cZ$ is a function. Let $D$ be
1277: a probability distribution on $\cX \times \cY$, and
1278: $\cP$ be a $[t;l_1,\ldots,l_t]^A$ private coin
1279: randomised protocol for $f$.
1280: Let $X$ stand for the classical random variable denoting
1281: Alice's input,
1282: $M$ stand for the random variable denoting the
1283: first message of Alice in $\cP$, and
1284: $I(X:M)$ denote the mutual information between $X$ and $M$ when
1285: the inputs to $\cP$ are distributed according to $D$.
1286: Then there exists a $[t-1;l_2,\ldots,l_t]^B$
1287: deterministic protocol $\cQ$ for $f$, such that
1288: $\epsilon^\cQ_D \leq \epsilon^\cP_D + (1/2) ((2 \ln 2) I(X:M))^{1/2}$.
1289: \end{lemma}
1290: \begin{proof}
1291: We first give an overview of the idea of the proof before
1292: getting down to the details. The proof proceeds in stages.
1293:
1294: \paragraph{Idea of Stage 1:} Starting from protocol $\cP$, we
1295: construct a
1296: $[t;l_1,\ldots,l_t]^A$ private coin protocol $\cP'$
1297: where the first message
1298: is independent of Alice's input, and $\epsilon^{\cP'}_D
1299: \leq \epsilon^{\cP}_D + (1/2) ((2 \ln 2) I(X:M))^{1/2}$.
1300: The important idea here is to first generate Alice's
1301: message using a new private coin without `looking' at her input,
1302: and after that,
1303: to adjust Alice's old private coin in a suitable manner so as to
1304: be consistent with her message and input.
1305:
1306: \paragraph{Idea of Stage 2:} Since the first message of
1307: $\cP'$ is independent
1308: of Alice's input, Bob can generate it himself.
1309: Doing this and setting coin tosses appropriately gives us a
1310: $[t-1;l_2,\ldots,l_t]^B$ deterministic protocol
1311: $\cQ$ for $f$ such that
1312: $\epsilon^\cQ_D \leq \epsilon^{\cP'}_D
1313: \leq \epsilon^\cP_D + (1/2) ((2 \ln 2) I(X:M))^{1/2}$.
1314:
1315: \bigskip
1316:
1317: We now give the details of the proof.
1318: Let $\Pi^x$ be the probability distribution of the first message
1319: $M$ of protocol $\cP$ when Alice's input $X=x$.
1320: Define $\Pi \defeq \sum_x d_x \Pi^x$, where
1321: $d_x$ is the marginal probability of $X=x$ under distribution $D$.
1322: $\Pi$ is the probability distribution of the average first
1323: message of $\cP$ under distribution $D$.
1324: For $x \in \cX$ and an instance $m$ of the first message of Alice, let
1325: $q^{xm}_r$ denote the conditional probability that the private
1326: coin toss of Alice results in $r$, given that Alice's input is $x$
1327: and her first message in $\cP$ is $m$.
1328: If in $\cP$ message $m$ cannot occur when Alice's
1329: input is $x$, then we
1330: define $q^{xm}_r \defeq 0$.
1331: Let $\pi_m^x$
1332: denote the probability that the first message of
1333: Alice in $\cP$ is $m$,
1334: given that her input is $x$. Let $\pi_m$ denote the
1335: probability that the first message of Alice in $\cP$ is $m$, when
1336: Alice's and Bob's inputs are distributed according to $D$.
1337: Then, $\pi_m = \sum_x d_x \pi_m^x$.
1338:
1339: \paragraph{Stage 1: }
1340: We construct a $[t;l_1,\ldots,l_t]^A$ private coin
1341: randomised protocol $\cP'$ for $f$ with average error under
1342: distribution $D$
1343: $\epsilon^{\cP'}_D \leq \epsilon^\cP_D +
1344: (1/2) ((2 \ln 2) I(X:M))^{1/2}$,
1345: and where the probability distribution of
1346: the first message is independent of the input to Alice.
1347: We now describe the protocol $\cP'$.
1348: Suppose Alice is given
1349: $x \in \cX$ and Bob is given $y \in \cY$. Alice tosses a fresh private
1350: coin to pick $m$ with probability $\pi_m$ and set
1351: her old private coin to $r$ with probability $q^{xm}_r$.
1352: After this, Alice and Bob behave as in protocol $\cP$ (henceforth,
1353: Alice ignores the new private coin which she had tossed to generate
1354: her first message $m$).
1355: Hence in $\cP'$ the probability distribution of the first message
1356: is independent of Alice's input.
1357:
1358: Let us now compare the situations in protocols
1359: $\cP$ and $\cP'$ when Alice's input is $x$, Bob's input is $y$,
1360: Alice has finished tossing her private coins (both old and new),
1361: but no communication has taken place as yet.
1362: In protocol $\cP$, the probability that Alice's private coin toss
1363: results in $r$ is $\sum_m \pi_m^x q^{xm}_r$.
1364: In protocol $\cP'$, the probability that Alice's old private coin
1365: is set to $r$ is $\sum_m \pi_m q^{xm}_r$.
1366: Thus, the total variation distance between the probability
1367: distributions on Alice's old private coin is
1368: \begin{eqnarray*}
1369: \sum_r \left|\sum_m q^{xm}_r (\pi_m^x - \pi_m)\right|
1370: & \leq & \sum_r \sum_m q^{xm}_r \left| \pi_m^x - \pi_m\right| \\
1371: & = & \sum_m \left(\left| \pi_m^x - \pi_m \right|
1372: \sum_r q^{xm}_r
1373: \right) \\
1374: & = & \sum_m \left| \pi_m^x - \pi_m \right| \\
1375: & = & \ellone{\Pi^x - \Pi}.
1376: \end{eqnarray*}
1377: Hence, the error probability of $\cP'$ on input $x,y$
1378: $\epsilon^{\cP'}_{x,y} \leq \epsilon^\cP_{x,y} +
1379: (1/2) \ellone{\Pi^x - \Pi}$.
1380: Let $d_{xy}$ be the probability that $(X,Y)=(x,y)$ under distribution
1381: $D$. Then, the average error of $\cP'$ under distribution $D$
1382: \begin{eqnarray*}
1383: \epsilon^{\cP'}_D
1384: & = & \sum_{x,y} d_{xy} \epsilon^{\cP'}_{x,y} \\
1385: & \leq & \sum_{x,y} d_{xy} \left(\epsilon^\cP_{x,y} + \frac{1}{2}
1386: \ellone{\Pi^x - \Pi}\right) \\
1387: & = & \epsilon^\cP_D +
1388: \frac{1}{2} \sum_x d_x \ellone{\Pi^x - \Pi} \\
1389: & \leq & \epsilon^\cP_D + \frac{1}{2} ((2 \ln 2) I(X:M))^{1/2}.
1390: \end{eqnarray*}
1391: The last inequality follows from
1392: Fact~\ref{fact:averageencodingclassical}.
1393:
1394: \paragraph{Stage 2:}
1395: We now construct our desired $[t-1;l_2,\ldots,l_t]^B$
1396: deterministic protocol $\cQ$ for $f$ with
1397: $\epsilon^\cQ_D \leq \epsilon^{\cP'}_D$.
1398: Suppose all the coin tosses
1399: of Alice and Bob in $\cP'$ are done publicly before any communication
1400: takes place. Now there is no need for the first message from
1401: Alice to Bob, because Bob can reconstruct the message by looking
1402: at the public coin tosses.
1403: This gives us a $[t-1;l_2,\ldots,l_t]^B$ public coin
1404: protocol $\cQ'$ such that
1405: $\epsilon^{\cQ'}_{x,y} = \epsilon^{\cP'}_{x,y}$ for
1406: every $(x,y) \in \cX \times \cY$. By setting the public coin tosses of
1407: $\cQ'$ to an appropriate value, we get a $[t-1;l_2,\ldots,l_t]^B$
1408: deterministic protocol
1409: $\cQ$ such that
1410: $\epsilon^\cQ_D \leq \epsilon^{\cQ'}_D = \epsilon^{\cP'}_D
1411: \leq \epsilon^\cP_D + (1/2) ((2 \ln 2) I(X:M))^{1/2}$.
1412:
1413: This completes the proof of Lemma~\ref{lem:classicalroundreduce}.
1414: \end{proof}
1415:
1416: \subsection{Quantum}
1417: \label{subsec:quantrndreduce}
1418: \begin{lemma}
1419: \label{lem:quantroundreduce}
1420: Suppose $f: \cX \times \cY \rightarrow \cZ$ is a function. Let $D$ be
1421: a probability distribution on $\cX \times \cY$, and
1422: $\cP$ be a $[t;c;l_1,\ldots,l_t]^A$ safe coinless
1423: quantum protocol for $f$.
1424: Let $X$ stand for the classical random variable denoting
1425: Alice's input, $M$ denote the
1426: first message of Alice in $\cP$, and
1427: $I(X:M)$ denote the mutual information between $X$ and $M$ when
1428: the inputs to $\cP$ are distributed according to $D$.
1429: Then there exists a $[t-1;c+l_1;l_2,\ldots,l_t]^B$
1430: safe coinless quantum protocol $\cQ$ for $f$, such that
1431: $\epsilon^\cQ_D \leq \epsilon^\cP_D + (1/2) ((4 \ln 2) I(X:M))^{1/2}$.
1432: \end{lemma}
1433: \begin{proof}
1434: We first give an overview of the plan of the proof, before
1435: getting down to the details. The proof proceeds in stages.
1436: Stage~1 of the quantum proof corresponds to
1437: Stage~1 of the classical proof, and
1438: Stages~2A and 2B of the quantum proof together correspond to
1439: Stage~2 of the classical proof.
1440:
1441: \paragraph{Idea of Stage 1:}
1442: Starting from protocol $\cP$, we construct a
1443: $[t;c;l_1,\ldots,l_t]^A$ safe coinless quantum protocol $\cP'$
1444: where the first message
1445: is independent of Alice's input, and
1446: $\epsilon^{\cP'}_D \leq
1447: \epsilon^{\cP}_D + (1/2)((4 \ln 2) I(X:M))^{1/2}$.
1448: The important idea here is to generate a purification $\ket{\phi^x}$
1449: of the average first
1450: message $\rho$ of Alice in $\cP$ that is close to
1451: the purification $\ket{\psi^x}$ of the message $\rho^x$ of Alice
1452: in $\cP$ when her input is $x$. The existence of such a purification
1453: $\ket{\phi^x}$ is guaranteed by Fact~\ref{fact:averageencodingquant}.
1454:
1455: \paragraph{Idea of Stage 2A:}
1456: Since the first message of $\cP'$ is independent of
1457: Alice's input (in fact its density matrix is $\rho$),
1458: Bob can generate it himself. This suffices if $\cP'$ is a one-round
1459: protocol, and completes the proof of Lemma~\ref{lem:quantroundreduce}
1460: for such protocols.
1461: But if $\cP'$ has more than one round, it is also necessary for Bob
1462: to achieve the correct entanglement between
1463: Alice's work qubits and the first message i.e. it is necessary that
1464: the joint state of Alice's work qubits and the qubits of the first
1465: message be $\ket{\phi^x}$.
1466: Bob achieves this by first sending a safe message of $l_1+c$ qubits.
1467: If Alice's input is $x$, she then applies a unitary transformation
1468: $V_x$ on her work qubits in order to make
1469: the joint state of her work qubits and the qubits of the first
1470: message $\ket{\phi^x}$.
1471: The existence of such a $V_x$ follows from
1472: Fact~\ref{fact:localtransition}.
1473: Doing all this gives us a
1474: $[t+1;c+l_1;0,0,l_2,\ldots,l_t]^B$ safe coinless quantum
1475: protocol $\cQ'$ for $f$ such that
1476: $\epsilon^{\cQ'}_{x,y} = \epsilon^{\cP'}_{x,y}$ for
1477: every $(x,y) \in \cX \times \cY$.
1478:
1479: \paragraph{Idea of Stage 2B:}
1480: Since the first message of Alice in $\cQ'$ is zero qubits
1481: long, Bob can concatenate his first two messages, giving
1482: us a $[t-1;c+l_1;l_2,\ldots,l_t]^B$ safe coinless quantum
1483: protocol $\cQ$ for $f$ such that
1484: $\epsilon^{\cQ}_{x,y} = \epsilon^{\cQ'}_{x,y}$ for
1485: every $(x,y) \in \cX \times \cY$. The technical reason
1486: behind this is that unitary transformations on disjoint
1487: sets of qubits commute.
1488: Note that
1489: $\epsilon^{\cQ}_D = \epsilon^{\cQ'}_D = \epsilon^{\cP'}_D
1490: \leq \epsilon^{\cP}_D + (1/2)((4 \ln 2) I(X:M))^{1/2}$.
1491:
1492: We now give the details of the proof.
1493: Let $\rho^x$ be the density matrix of the first message $M$ of
1494: protocol $\cP$ when Alice's input $X=x$.
1495: Let $A$ and $B$
1496: denote Alice's work qubits excluding the qubits of $M$
1497: and Bob's work qubits respectively.
1498: Let $\ket{\psi^x}_{AM}$ be the joint pure state of $AM$
1499: in $\cP$ when Alice's input $X=x$, she has finished preparing
1500: her first message, but no communication has taken place as yet.
1501: Without loss of generality, the number of
1502: qubits in $A$ is at least the number of qubits in $M$.
1503: Define $\rho \defeq \sum_x p_x \rho^x$, where
1504: $p_x$ is the marginal probability of $X=x$ under distribution $D$.
1505: $\rho$ is the density
1506: matrix of the average first message of $\cP$ under distribution $D$.
1507:
1508: \paragraph{Stage 1:}
1509: We construct a $[t;c;l_1,\ldots,l_t]^A$ safe coinless quantum
1510: protocol $\cP'$ for $f$ with average error under distribution $D$
1511: $\epsilon^{\cP'}_D \leq \epsilon^{\cP}_D +
1512: (1/2)((4 \ln 2) I(X:M))^{1/2}$,
1513: and where the density matrix of
1514: the first message is independent of the input $x$ to Alice.
1515: We now describe the protocol $\cP'$.
1516: Suppose Alice is given
1517: $x \in \cX$ and Bob is given $y \in \cY$.
1518: The qubits of $AM$ are initialised to zero. Alice applies
1519: a unitary transformation $U'_x$ on $AM$
1520: in order to prepare a purification $\ket{\phi^x}$
1521: of $\rho$. She then sends the qubits of $M$ as her first message
1522: to Bob. After this, Alice and Bob behave as in protocol $\cP$.
1523: Hence in $\cP'$ the density matrix of the first message
1524: is independent of Alice's input.
1525:
1526: Let us now compare the situation in protocols
1527: $\cP$ and $\cP'$ when Alice's input is $x$, Bob's input is $y$,
1528: Alice has finished preparing the state $\ket{\phi^x}$, but no
1529: communication has taken place as yet.
1530: In protocol $\cP$, the state of $A M B$ at this point in time
1531: is $\ket{\psi^x}_{AM} \ket{\bzero}_B$.
1532: In protocol $\cP'$, the state of $A M B$ at this point in time
1533: is $\ket{\phi^x}_{AM} \ket{\bzero}_B$.
1534: Hence,
1535: $\epsilon^{\cP'}_{x,y} \leq \epsilon^{\cP}_{x,y} +
1536: (1/2)\ellone{\ketbra{\psi^x} - \ketbra{\phi^x}}$.
1537: Let $q_{xy}$ denote
1538: the probability that $(X,Y)=(x,y)$ under distribution
1539: $D$. Then, the average error of $\cP'$ under distribution $D$
1540: \begin{eqnarray*}
1541: \epsilon^{\cP'}_D
1542: & = & \sum_{x,y} q_{xy} \epsilon^{\cP'}_{x,y} \\
1543: & \leq & \sum_{x,y} q_{xy}
1544: \left(\epsilon^{\cP}_{x,y} + \frac{1}{2}
1545: \ellone{\ketbra{\psi^x} - \ketbra{\phi^x}}
1546: \right) \\
1547: & = & \epsilon^{\cP}_D + \frac{1}{2}
1548: \sum_x p_x \ellone{\ketbra{\psi^x} - \ketbra{\phi^x}} \\
1549: & \leq & \epsilon^{\cP}_D + \frac{1}{2} ((4 \ln 2) I(X:M))^{1/2}
1550: \end{eqnarray*}
1551: The last inequality follows by taking $\ket{\phi^x}$ to be the
1552: purifications promised by Fact~\ref{fact:averageencodingquant}.
1553:
1554: \paragraph{Stage 2A:}
1555: We now construct a $[t+1;c+l_1;0,0,l_2,\ldots,l_t]^B$ safe
1556: coinless quantum protocol $\cQ'$ for $f$ with
1557: $\epsilon^{\cQ'}_{x,y} = \epsilon^{\cP'}_{x,y}$, for all
1558: $(x,y) \in \cX \times \cY$.
1559: Suppose
1560: Alice is given $x \in \cX$ and Bob is given $y \in \cY$.
1561: Let $A_1$ denote
1562: all the qubits of $A$, except the last $l_1+c$ qubits.
1563: Let $A_2$ denote the last $l_1+c$ qubits of $A$. Thus, $A = A_1 A_2$.
1564: In protocol $\cQ'$, Alice initially starts with work qubits $A_1$
1565: only, and Bob initially starts with work qubits $B M A_2$.
1566: The work qubits of Alice and Bob are initialised to zero.
1567: Bob commences protocol $\cQ'$ by constructing
1568: a canonical purification $\ket{\eta}_{M A_2}$ of $\rho$, where
1569: the reduced density matrix of $M$ is $\rho$.
1570: Bob then sends $A_2$ to Alice.
1571: The density matrix of $A_2$ is independent
1572: of the inputs $x,y$ (in fact, if $\ket{\eta}_{M A_2}$
1573: is the Schmidt purification then the reduced
1574: density matrix of $A_2$ is also $\rho$).
1575: After receiving $A_2$, Alice applies a unitary transformation
1576: $V_x$ on $A$ so that the state vector of $A M$ becomes
1577: $\ket{\phi^x}_{AM}$. The existence of such
1578: a $V_x$ follows from Fact~\ref{fact:localtransition}.
1579: The global state of Alice's and Bob's qubits at this point
1580: in protocol $\cQ'$ is the same as the global state of
1581: Alice's and Bob's qubits at the point in protocol $\cP'$ just
1582: after Alice has sent her first message to Bob.
1583: Bob now treats $M$ as if
1584: it were the first message of Alice in $\cP'$, and proceeds to compute
1585: his response $N$ (the qubits of $N$ are a subset of the qubits
1586: of $M B$) of length $l_2$. Bob sends $N$ to Alice and
1587: after this protocol $\cQ'$ proceeds as protocol $\cP'$. In $\cQ'$
1588: Bob starts the communication, the communication
1589: goes on for $t+1$ rounds, the first message of Bob of length
1590: $l_1+c$ viz. $A_2$ is a safe message, and the first
1591: message of Alice is zero qubits long.
1592:
1593: \paragraph{Stage 2B:}
1594: We finally construct our desired $[t-1;c+l_1;l_2,\ldots,l_t]^B$ safe
1595: coinless quantum protocol $\cQ$ for $f$ with
1596: $\epsilon^{\cQ}_{x,y} = \epsilon^{\cQ'}_{x,y}$, for all
1597: $(x,y) \in \cX \times \cY$.
1598: In protocol $\cQ$ Bob, after doing the same computations as
1599: in $\cQ'$, first sends as a single
1600: message the $(l_1+c)+l_2$ qubits $A_2 N$.
1601: After receiving $A_2 N$,
1602: Alice applies $V_x$ on $A$ followed by her appropriate
1603: unitary transformation
1604: on $A N$ viz. the unitary transformation of Alice in $\cQ'$ on
1605: the qubits of $A N$ after she has received the first two
1606: messages of Bob.
1607: The global state of all the qubits of
1608: Alice and Bob at this point in protocol $\cQ$
1609: is the same as the global state of all the qubits of Alice
1610: and Bob at the point in protocol $\cQ'$ just after Alice has finished
1611: generating her second message but before she has sent it to Bob.
1612: This is because
1613: unitary transformations on disjoint sets of qubits commute.
1614: After this, protocol $\cQ$ proceeds as protocol $\cQ'$.
1615: In protocol $\cQ$ Bob starts the communication, the
1616: communication
1617: goes on for $t-1$ rounds, and the first message of Bob of length
1618: $(l_1+c)+l_2$ viz. $A_2 N$ contains a safe overhead viz. $A_2$ of
1619: $l_1+c$ qubits.
1620:
1621: This completes the proof of Lemma~\ref{lem:quantroundreduce}.
1622: \end{proof}
1623:
1624: \paragraph{Remark:}
1625: The proof of Lemma~\ref{lem:quantroundreduce} requires the global
1626: state of all the qubits of Alice and Bob to be pure at all times
1627: during the execution of protocol $\cP$, when Alice's and Bob's
1628: inputs are in computational basis states. This is because we use
1629: the machinery of purifications in the proof. The proof also ensures
1630: that the purity property holds for the final protocol $\cQ$.
1631:
1632: \section{The round elimination lemma}
1633: \label{sec:roundelim}
1634: We now prove the classical and quantum versions of
1635: our strong round elimination lemma in
1636: Sections~\ref{subsec:classicalroundelim} and
1637: \ref{subsec:quantroundelim} respectively. The round elimination
1638: lemma is stated for public coin protocols
1639: only. Since a public coin classical randomised protocol can
1640: be converted to a private coin classical randomised
1641: protocol at the expense of an
1642: additive increase in the communication complexity by at most
1643: logarithm of the total bit size of the inputs~\cite{newman:private},
1644: we also get a similar round elimination lemma for private
1645: coin classical randomised protocols. A similar statement about
1646: round elimination can
1647: be made for safe coinless quantum protocols; for such protocols
1648: the safe overhead increases by an additional additive term
1649: that is at most
1650: logarithm of the total bit size of the inputs.
1651: But since the statement of the round elimination lemma is cleanest
1652: for public coin protocols, we give it below
1653: for such protocols only.
1654:
1655: \subsection{Classical}
1656: \label{subsec:classicalroundelim}
1657: \begin{lemma}[Round elimination lemma, classical version]
1658: \label{lem:classicalroundelim}
1659: Suppose $f:\cX \times \cY \rightarrow \cZ$ is a function.
1660: Suppose the communication game $f^{(n),A}$ has a
1661: $[t;l_1,\ldots,l_t]^A$ public coin randomised protocol
1662: with error less than $\delta$.
1663: Then there is a $[t-1;l_2,\ldots,l_t]^B$ public coin
1664: randomised protocol for $f$ with error less than
1665: $\epsilon \defeq \delta + (1/2) (2 l_1 \ln 2 /n)^{1/2}$.
1666: \end{lemma}
1667: \begin{proof}
1668: Suppose the given protocol for $f^{(n),A}$ has error
1669: $\tilde{\delta} < \delta$. Define
1670: $\tilde{\epsilon} \defeq \tilde{\delta} +
1671: (1/2) (2 l_1 \ln 2 /n)^{1/2}$.
1672: To prove the round elimination lemma
1673: it suffices to give,
1674: by the harder direction of Yao's minimax lemma~\cite{yao:minimax},
1675: for any probability distribution $D$ on $\cX \times \cY$, a
1676: $[t-1;l_2,\ldots,l_t]^B$ deterministic
1677: protocol $\cP$ for $f$ with
1678: $\epsilon^\cP_D \leq \tilde{\epsilon} < \epsilon$.
1679: To this end, we will first construct a probability distribution
1680: $D^\ast$ on $\cX^n \times [n] \times \cY$ as follows:
1681: Choose $i \in [n]$ uniformly at random. Choose independently, for
1682: each $j \in [n]$,
1683: $(x_j,y_j) \in \cX \times \cY$ according to distribution $D$. Set
1684: $y=y_i$ and throw away $y_j, j \neq i$.
1685: By the easier direction of Yao's minimax lemma, we get a
1686: $[t;l_1,\ldots,l_t]^A$ deterministic protocol $\cP^\ast$
1687: for $f^{(n),A}$ with
1688: $\epsilon^{\cP^\ast}_{D^\ast} \leq \tilde{\delta} < \delta$.
1689: In $\cP^\ast$, Alice gets $x_1,\ldots,x_n \in \cX$, Bob gets
1690: $i \in [n]$, $y \in \cY$ and copies of $x_1,\ldots,x_{i-1}$.
1691: We shall construct the desired protocol $\cP$
1692: from the protocol $\cP^\ast$.
1693:
1694: In $\cP^\ast$, let Alice's and Bob's inputs be distributed according
1695: to $D^\ast$. Let the input to Alice be denoted by
1696: the random variable $X \defeq X_1 \cdots X_n$, where $X_i$ is the
1697: random variable corresponding to the $i$th input to Alice.
1698: Let the random variables $Y, \cI$ correspond to the inputs $y, i$
1699: respectively of Bob.
1700: Let $M$ denote the random variable corresponding to
1701: the first message of Alice in $\cP^\ast$.
1702: Define probability distribution $D^\ast_{i;x_1,\ldots,x_{i-1}}$ on
1703: $\cX^n \times [n] \times \cY$ to be the distribution $D^\ast$
1704: conditioned on $\cI=i$ and $X_1,\ldots,X_{i-1}=x_1,\ldots,x_{i-1}$.
1705: Define probability distribution $D^\ast_{i;y;x_1,\ldots,x_i}$ on
1706: $\cX^n \times [n] \times \cY$ to be the distribution $D^\ast$
1707: conditioned on $\cI=i$, $Y=y$ and $X_1,\ldots,X_i=x_1,\ldots,x_i$.
1708: Let $\epsilon^{\cP^\ast}_{D^\ast;i;x_1,\ldots,x_{i-1}}$ denote
1709: the average
1710: error of $\cP^\ast$ under distribution
1711: $D^\ast_{i;x_1,\ldots,x_{i-1}}$.
1712: Using Fact~\ref{fact:classicalaveraging} and the fact that
1713: under distribution $D^\ast$
1714: $X_1,\ldots,X_n$ are independent random variables, we get that
1715: \begin{equation}
1716: \label{eq:info1classical}
1717: \begin{array}{lrl}
1718: \E_{i,X}\, [I((X_i:M) \mid X_1,\ldots,X_{i-1}=x_1,\ldots,x_{i-1})]
1719: & = & \displaystyle \E_i\, [I (X_i:M X_1 \cdots X_{i-1})] \\
1720: & = & \frac{I(X:M)}{n} \\
1721: & \leq & \frac{l_1}{n}.
1722: \end{array}
1723: \end{equation}
1724: Also,
1725: \begin{equation}
1726: \label{eq:error1classical}
1727: \tilde{\delta} \geq \epsilon^{\cP^\ast}_{D^\ast} =
1728: \E_{i,X}\, \left[\epsilon^{\cP^\ast}_{D^\ast;i;x_1,\ldots,x_{i-1}}
1729: \right].
1730: \end{equation}
1731: Above, the expectations are under distribution $D^\ast$ and the
1732: mutual informations are for protocol $\cP^\ast$ with its inputs
1733: distributed according to $D^\ast$.
1734:
1735: For any $i \in [n]$, $x_1, \ldots, x_{i-1} \in \cX$, let us now define
1736: the $[t;l_1,\ldots,l_t]^A$
1737: private coin randomised protocol $\cP'_{i;x_1,\ldots,x_{i-1}}$
1738: for the function $f$ in terms of protocol $\cP^\ast$ as follows:
1739: Alice is given
1740: $x \in \cX$ and Bob is given $y \in \cY$. Bob sets
1741: $\cI=i$, and both Alice and Bob set
1742: $X_1, \ldots, X_{i-1} = x_1, \ldots, x_{i-1}$.
1743: Alice tosses a fresh private coin
1744: to choose $X_{i+1},\ldots,X_n \in \cX$, where each
1745: $X_j, i+1 \leq j \leq n$ is chosen independently according to the
1746: marginal distribution on $\cX$ induced by $D$.
1747: Alice sets $X_i=x$ and Bob sets
1748: $Y=y$. They then run protocol $\cP^\ast$ on these inputs.
1749: The probability that $\cP'_{i;x_1,\ldots,x_{i-1}}$
1750: makes an error for an input $(x,y)$,
1751: $\epsilon^{\cP'_{i;x_1,\ldots,x_{i-1}}}_{x,y}$,
1752: is the average probability of
1753: error of $\cP^\ast$ under distribution $D^\ast_{i;y;x_1,\ldots,x_i}$.
1754: Hence, the average probability of error of
1755: $\cP'_{i;x_1,\ldots,x_{i-1}}$
1756: under distribution $D$,
1757: \begin{equation}
1758: \label{eq:error2classical}
1759: \epsilon^{\cP'_{i;x_1,\ldots,x_{i-1}}}_D =
1760: \epsilon^{\cP^\ast}_{D^\ast;i;x_1,\ldots,x_{i-1}}.
1761: \end{equation}
1762: Let $M'$ denote the random variable corresponding to Alice's
1763: first message
1764: and $X'$ denote the random variable $X_i$, when Alice's and Bob's
1765: inputs are distributed according
1766: to $D$ in $\cP'_{i;x_1,\ldots,x_{i-1}}$.
1767: Then
1768: \begin{equation}
1769: \label{eq:info2classical}
1770: I((X_i:M) \mid X_1,\ldots,X_{i-1}=x_1,\ldots,x_{i-1}) = I (X':M'),
1771: \end{equation}
1772: where the left hand side refers to
1773: the mutual information in protocol $\cP^\ast$ when its inputs are
1774: distributed according to $D^\ast$ and
1775: the right hand side refers to the mutual information in
1776: protocol $\cP'_{i;x_1,\ldots,x_{i-1}}$ when its inputs are distributed
1777: according to $D$.
1778:
1779: Using Lemma~\ref{lem:classicalroundreduce} and equations
1780: (\ref{eq:error2classical}) and (\ref{eq:info2classical}), we get a
1781: $[t-1;l_2,\ldots,l_t]^B$
1782: deterministic protocol $\cP_{i;x_1,\ldots,x_{i-1}}$ for
1783: $f$ with
1784: \begin{equation}
1785: \label{eq:average1classical}
1786: \begin{array}{lcl}
1787: \epsilon^{\cP_{i;x_1,\ldots,x_{i-1}}}_D
1788: & \leq & \epsilon^{\cP'_{i;x_1,\ldots,x_{i-1}}}_D +
1789: \frac{1}{2} ((2 \ln 2) I(X':M'))^{1/2} \\
1790: & = & \epsilon^{\cP^\ast}_{D^\ast;i;x_1,\ldots,x_{i-1}} +
1791: \frac{1}{2} ((2 \ln 2)
1792: I((X_i:M) \mid X_1,\ldots, X_{i-1} =
1793: x_1,\ldots,x_{i-1}))^{1/2}.
1794: \end{array}
1795: \end{equation}
1796: We have that
1797: (note that the expectations below are under distribution $D^\ast$ and
1798: the mutual informations are for protocol $\cP^\ast$ with its
1799: inputs distributed according to $D^\ast$)
1800: \begin{eqnarray*}
1801: \E_{i,X} \left[\epsilon^{\cP_{i;x_1,\ldots,x_{i-1}}}_D\right]
1802: & \leq & \E_{i,X} \left[\epsilon^{\cP^\ast}_{D^\ast;i;
1803: x_1,\ldots,x_{i-1}}
1804: \right] + \\
1805: & & \frac{1}{2} \E_{i,X}\, \left[
1806: ((2 \ln 2) I((X_i:M) \mid X_1,\ldots,X_{i-1} =
1807: x_1,\ldots,x_{i-1}))^{1/2}
1808: \right] \\
1809: & \leq & \E_{i,X} \left[\epsilon^{\cP^\ast}_{D^\ast;i;
1810: x_1,\ldots,x_{i-1}}
1811: \right] + \\
1812: & & \frac{1}{2} \left(
1813: (2 \ln 2) \E_{i,X}
1814: [I((X_i:M) \mid X_1,\ldots, X_{i-1}=
1815: x_1,\ldots,x_{i-1})]
1816: \right)^{1/2} \\
1817: & \leq & \tilde{\delta} + \frac{1}{2}
1818: \left(\frac{2 l_1 \ln 2}{n} \right)^{1/2} \\
1819: & = & \tilde{\epsilon}.
1820: \end{eqnarray*}
1821: The first inequality follows from (\ref{eq:average1classical}),
1822: the second
1823: inequality follows from the concavity of the square root function
1824: and the last inequality from (\ref{eq:info1classical}) and
1825: (\ref{eq:error1classical}).
1826:
1827: Thus, we can immediately
1828: see that there exist $i \in [n]$ and
1829: $x_1, \ldots, x_{i-1} \in \cX$ such that
1830: $\epsilon^{\cP_{i;x_1,\ldots,x_{i-1}}}_D \leq \tilde{\epsilon}$.
1831: Let $\cP \defeq \cP_{i;x_1,\ldots,x_{i-1}}$. $\cP$ is
1832: our desired $[t-1;l_2,\ldots,l_t]^B$
1833: deterministic protocol for $f$ with
1834: $\epsilon^\cP_D \leq \tilde{\epsilon}$, thus completing the proof
1835: of the round elimination lemma.
1836: \end{proof}
1837:
1838: \subsection{Quantum}
1839: \label{subsec:quantroundelim}
1840: \begin{lemma}[Round elimination lemma, quantum version]
1841: \label{lem:quantroundelim}
1842: Suppose $f:\cX \times \cY \rightarrow \cZ$ is a function.
1843: Suppose the communication game $f^{(n),A}$ has a
1844: $[t;c;l_1,\ldots,l_t]^A$ safe public coin quantum protocol
1845: with error less than $\delta$.
1846: Then there is a $[t-1;c+l_1;l_2,\ldots,l_t]^B$ safe public coin
1847: quantum protocol for $f$ with error less than
1848: $\epsilon \defeq \delta + (1/2) (8 l_1 \ln 2 /n)^{1/2}$.
1849: \end{lemma}
1850: \begin{proof} {\bf (Sketch)}
1851: The proof is very similar to the proof
1852: of Lemma~\ref{lem:classicalroundelim}.
1853: We just point out some important things below. Note
1854: that in the quantum setting, the upper bound in
1855: (\ref{eq:info1classical}) is
1856: $\frac{2 l_1}{n}$ by Fact~\ref{fact:quantaveraging}(c).
1857: Also, in the definition of protocol $\cP'_{i;x_1,\ldots,x_{i-1}}$,
1858: instead of feeding probabilistic mixtures for the inputs
1859: $X_{i+1}, \ldots, X_n$, Alice feeds appropriate pure states
1860: (pure states that would give the correct probabilistic mixture
1861: were they to be measured in the computational basis).
1862: Equations~\ref{eq:error2classical} and \ref{eq:info2classical}
1863: continue to hold because protocol $\cP^\ast$ is secure.
1864: Since the global
1865: state of all the qubits of Alice and Bob is pure at all times
1866: during the execution of protocol $\cP'_{i;x_1,\ldots,x_{i-1}}$
1867: when Alice's and Bob's
1868: inputs are in computational basis states,
1869: Lemma~\ref{lem:quantroundreduce} can now be used to get the
1870: $[t-1;c+l_1;l_2,\ldots,l_t]^B$ safe coinless quantum protocol
1871: $\cP_{i;x_1,\ldots,x_{i-1}}$ for $f$.
1872: \end{proof}
1873:
1874: \section{Optimal lower bounds for predecessor}
1875: \label{sec:predlb}
1876: In this section, we prove our (optimal) lower bounds on the query
1877: complexity of static
1878: predecessor searching in the
1879: cell probe model with randomised or address-only quantum query
1880: schemes.
1881:
1882: \begin{theorem}
1883: \label{thm:classicalpredlb}
1884: Suppose there is a $(n^{O(1)}, (\log m)^{O(1)}, t)$ randomised
1885: cell probe
1886: scheme for the $(m, n)$-static predecessor problem with error
1887: probability less than $1/3$.
1888: Then,
1889: \begin{enumerate}
1890: \item[(a)]
1891: $t = \Omega \left( \frac{\log \log m}{\log \log \log m} \right)$
1892: as a function of $m$;
1893: \item[(b)]
1894: $t = \Omega \left( \sqrt{\frac{\log n}{\log \log n}} \, \right)$
1895: as a function of $n$.
1896: \end{enumerate}
1897: The same lower bound also holds for address-only quantum cell
1898: probe schemes for static predecessor searching.
1899: \end{theorem}
1900: \begin{proof}
1901: The proof is similar to the proof of the lower bound for predecessor
1902: in \cite{miltersen:roundelim},
1903: but with different parameters, and using our stronger round
1904: elimination lemma in its classical version
1905: (Lemma~\ref{lem:classicalroundelim}).
1906:
1907: By Fact~\ref{fact:pred2rank}, it suffices to consider
1908: communication protocols for the
1909: rank parity communication game $\PAR_{\log m, n}$.
1910: Let $n = 2^{(\log \log m)^2 / \log \log \log m}$.
1911: Let $c_1 \defeq (2 \ln 2) 6^2$.
1912: For any given constants $c_2, c_3 \geq 1$, define
1913: \begin{displaymath}
1914: a \defeq c_2 \log n ~~~~~~~~~~ b \defeq (\log m)^{c_3}
1915: \end{displaymath}
1916: \begin{displaymath}
1917: t \defeq \frac{\log \log m}{(c_1 + c_2 + c_3) \log \log \log m}.
1918: \end{displaymath}
1919: We will show that $\PAR_{\log m, n}$ does not have
1920: $(2t,a,b)^A$ public coin randomised communication
1921: protocols with error less than $1/3$,
1922: thus proving both the desired lower bounds for the
1923: predecessor problem.
1924:
1925: Given a $(2t,a,b)^A$ public coin protocol for
1926: $\PAR_{\log m, n}$ with error probability at most $\delta$,
1927: we can get a $(2t,a,b)^A$ public coin protocol for
1928: $\PAR^{(c_1 a t^2),A}_{\frac{\log m}{c_1 a t^2}, n}$
1929: with error probability at most $\delta$
1930: by Fact~\ref{fact:rankred1}. Using
1931: Lemma~\ref{lem:classicalroundelim}, we get a
1932: $(2t-1,a,b)^B$ public coin protocol for
1933: $\PAR_{\frac{\log m}{c_1 a t^2}, n}$,
1934: but the error probability increases to at most $\delta + (12t)^{-1}$.
1935: By Fact~\ref{fact:rankred2}, we get
1936: a $(2t-1,a,b)^B$ public coin protocol for
1937: $\PAR^{(c_1 b t^2),B}_{\frac{\log m}{c_1 a t^2} -
1938: \log (c_1 b t^2) - 1, \frac{n}{c_1 b t^2}}$
1939: with error probability at most $\delta + (12t)^{-1}$.
1940: From the given values of the parameters, we see that
1941: $\frac{\log m}{(2 c_1 a t^2)^t} \geq \log (c_1 b t^2) + 1$.
1942: This implies that we also have a $(2t-1,a,b)^B$ public
1943: coin protocol for
1944: $\PAR^{(c_1 b t^2),B}_{\frac{\log m}{2 c_1 a t^2},
1945: \frac{n}{c_1 b t^2}}$
1946: with error probability at most $\delta + (12t)^{-1}$.
1947: Using Lemma~\ref{lem:classicalroundelim} again, we get a
1948: $(2t-2,a,b)^A$ public coin protocol for
1949: $\PAR_{\frac{\log m}{2 c_1 a t^2}, \frac{n}{c_1 b t^2}}$,
1950: but the error probability increases to at most
1951: $\delta + 2 (12t)^{-1}$.
1952:
1953: We do the above steps repeatedly. We start off with a
1954: $(2t,a,b)^A$ public coin protocol for
1955: $\PAR_{\log m, n}$ with error probability less than $1/3$.
1956: After applying the above steps
1957: $i$ times, we get a
1958: $(2t-2i,a,b)^A$ public coin protocol for
1959: $\PAR_{\frac{\log m}{(2 c_1 a t^2)^i}, \frac{n}{(c_1 b t^2)^i}}$
1960: with error probability less than $1/3 + 2i (12t)^{-1}$.
1961:
1962: By applying the above steps $t$ times, we finally get a
1963: $(0,a,b)^A$ public coin randomised protocol for the problem
1964: $\PAR_{\frac{\log m}{(2 c_1 a t^2)^t}, \frac{n}{(c_1 b t^2)^t}}$
1965: with error probability less than $1/3 + 2t (12t)^{-1} = 1/2$.
1966: From the given values of the parameters, we see that
1967: $\frac{\log m}{(2 c_1 a t^2)^t} \geq (\log m)^{\Omega(1)}$ and
1968: $\frac{n}{(c_1 b t^2)^t} \geq n^{\Omega(1)}$.
1969: Thus, we get a zero round protocol for a rank parity problem on a
1970: non-trivial domain with error probability less than $1/2$, which
1971: is a contradiction.
1972:
1973: In the above proof, we are tacitly ignoring ``rounding off'' problems.
1974: We remark that this does not affect the correctness of the proof.
1975:
1976: Finally we observe that by using Lemma~\ref{lem:quantroundreduce},
1977: one can prove that the same lower bound holds for address-only
1978: quantum cell probe schemes for static predecessor.
1979: \end{proof}
1980:
1981: \section{The `greater-than' problem}
1982: \label{sec:gt}
1983: We illustrate another application of the round elimination
1984: lemma to communication complexity by proving
1985: improved rounds versus communication tradeoffs for the
1986: `greater-than' problem.
1987: \begin{theorem}
1988: The bounded error public coin randomised $t$-round
1989: communication complexity of $\GT_n$
1990: is lower bounded by $\Omega(n^{1/t}t^{-2})$.
1991: For bounded error quantum protocols with input-independent
1992: prior entanglement for
1993: $\GT_n$, we have a lower bound of $\Omega(n^{1/t}t^{-1})$.
1994: \end{theorem}
1995: \begin{proof}
1996: We recall the following reduction from $\GT_{n/k}^{(k),A}$ to
1997: $\GT_n$ (see \cite{miltersen:roundelim}): In $\GT_{n/k}^{(k),A}$,
1998: Alice is given
1999: $x_1,\ldots,x_k \in \{0,1\}^{n/k}$, Bob is given $i \in [k]$,
2000: $y \in \{0,1\}^{n/k}$, and copies of $x_1,\ldots,x_{i-1}$, and
2001: they have to communicate and decide if $x_i > y$. To reduce
2002: $\GT_{n/k}^{(k),A}$ to $\GT_n$, Alice constructs
2003: $\hat{x} \in \{0,1\}^n$ by concatenating $x_1,\ldots,x_k$,
2004: Bob constructs $\hat{y} \in \{0,1\}^n$ by concatenating
2005: $x_1,\ldots,x_{i-1},y,1^{n(1 - i/k)}$. It is
2006: easy to see that $\hat{x} > \hat{y}$ iff $x_i > y$.
2007:
2008: Suppose there is a $t$-round bounded error public
2009: coin randomised protocol for $\GT_n$ with communication
2010: complexity $c$.
2011: We can think of the protocol as a $(t,c,c)^A$ public coin randomised
2012: protocol with error probability less than $1/3$.
2013: Suppose $n \geq k^t$, where $k \defeq (2 \ln 2) (3 t)^2 c$.
2014: Applying the self-reduction and Lemma~\ref{lem:classicalroundelim}
2015: alternately for $t$ stages gives us a zero round protocol
2016: for the `greater-than' problem on a non-trivial domain
2017: with error probability less than $1/2$,
2018: which is a contradiction.
2019: Thus, $c = \Omega(n^{1/t} t^{-2})$.
2020:
2021: In the above proof, we are tacitly ignoring ``rounding off'' problems.
2022: We remark that this does not affect the correctness of the proof.
2023:
2024: Finally, we observe that for bounded error quantum protocols with
2025: input-independent prior entanglement for $\GT_n$, one can
2026: improve the lower bound to $\Omega(n^{1/t} t^{-1})$ by exploiting
2027: the fact that by definition, a quantum protocol sends fixed length
2028: messages independent of the input.
2029: \end{proof}
2030:
2031: \section{Conclusion and open problems}
2032: \label{sec:conclusion}
2033: In this paper, we proved a lower bound for the randomised and
2034: address-only quantum query
2035: complexity of a cell probe scheme for the static predecessor
2036: searching problem. Our lower bound matches the deterministic
2037: cell probe upper bound of Beame and Fich.
2038: We proved our lower bound by proving a strong
2039: round elimination lemma in communication
2040: complexity. Our round elimination lemma improves on the round
2041: elimination lemma of Miltersen, Nisan, Safra and Wigderson, and is
2042: crucial to proving our optimal lower bound for predecessor searching.
2043: Our strong round elimination lemma also gives us improved
2044: rounds versus communication
2045: tradeoffs for the `greater-than' problem.
2046: We believe that
2047: our round elimination lemma is of independent interest and should have
2048: other applications.
2049: In fact recently, Chakrabarti and Regev~\cite{chakrabarti:ann}
2050: have proved an optimal lower bound for randomised cell probe schemes
2051: for the {\em approximate nearest neighbour} searching problem
2052: on the Hamming cube $\{0, 1\}^d$. Their proof uses the classical
2053: version of our round
2054: elimination lemma and combines it with a {\em message switching}
2055: argument, drawing on the message compression
2056: ideas of Jain, Radhakrishnan and Sen~\cite{jain:compress},
2057: for classical communication protocols that
2058: further exploits the
2059: asymmetry in the message lengths of Alice of Bob.
2060:
2061: We believe
2062: that our work brings out an interesting fact. Sometimes,
2063: in order to prove lower bound results, it helps to work in a more
2064: general model of computation. Of course, this
2065: makes the task of proving
2066: lower bounds harder, but also we now
2067: have more tools and techniques at
2068: our disposal. This sometimes enables us to attack the problem
2069: in a clearer fashion, without letting irrelevant details about the
2070: restricted model distract us.
2071: In our case, our attempt to prove a lower bound result in the more
2072: general address-only quantum cell probe model led us
2073: to make better use of powerful tools from information theory,
2074: which finally enabled us to prove optimal lower bounds for predecessor
2075: in the randomised cell probe model! Also, the
2076: information-theoretic approach gives us a simpler
2077: and clearer lower bound proof as compared to previous lower bound
2078: proofs for predecessor.
2079:
2080: The lower bound for predecessor searching for quantum cell probe
2081: schemes works only if the query scheme
2082: is {\em address-only}. If the query scheme is not address-only, the
2083: asymmetry in the message lengths of Alice and Bob in the
2084: corresponding quantum communication protocol breaks down.
2085: For a general quantum query scheme, it is an open problem
2086: to prove non-trivial lower bounds for static data structure problems.
2087:
2088: The {\em message switching} idea of Chakrabarti
2089: and Regev~\cite{chakrabarti:ann}
2090: works for classical communication protocols only. Thus, their lower
2091: bound for approximate nearest neighbour searching
2092: on the Hamming cube
2093: holds for classical cell probe schemes only. Our round elimination
2094: lemma alone does not seem to be able to fully exploit the asymmetry
2095: in the message lengths of Alice and Bob in a communication protocol
2096: for this problem. Proving a non-trivial lower bound for this
2097: problem in the address-only quantum cell probe model remains an
2098: open problem.
2099:
2100: \section*{Acknowledgements}
2101: We thank Rahul Jain, Hartmut
2102: Klauck and Peter Bro Miltersen for helpful discussions and feedback,
2103: and Amit Chakrabarti for sending us a copy of \cite{chakrabarti:ann}.
2104: We also thank Ashwin Nayak for helpful discussions and for
2105: pointing out the reference \cite{maneva:interaction},
2106: and Jaikumar Radhakrishnan for reading an early draft of
2107: \cite{sen:quantcell} and for enlightening discussions.
2108:
2109: \bibliography{pred}
2110:
2111: \appendix
2112:
2113: \section{The average encoding theorem}
2114:
2115: \subsection{Classical}
2116: In this subsection, we give a self-contained classical proof,
2117: without using quantum information theory, of
2118: Fact~\ref{fact:averageencodingclassical}.
2119: We first recall the definition of a classical information-theoretic
2120: quantity called {\em relative entropy}, also known as
2121: {\em Kullback-Leibler divergence}.
2122: \begin{definition}[Relative entropy]
2123: Let $P$ and $Q$ be probability distributions on the same
2124: sample space $\Omega$. The {\em relative entropy} between
2125: $P$ and $Q$ is defined as
2126: \begin{displaymath}
2127: S(P \| Q) \defeq
2128: \sum_{x \in \Omega} P(x) \log \left(\frac{P(x)}{Q(x)}\right).
2129: \end{displaymath}
2130: \end{definition}
2131:
2132: We now require a non-trivial fact from classical information
2133: theory, which upper bounds the total variation distance of a pair
2134: of probability distributions in terms of their relative entropy.
2135: A proof of the fact can be found in
2136: \cite[Lemma~12.6.1]{cover:infotheory}.
2137: \begin{fact}
2138: \label{thm:divell1}
2139: Let $P$ and $Q$ be probability distributions on the same finite
2140: sample space $\Omega$. Then,
2141: \[
2142: \ellone{P - Q} \leq \sqrt{(2 \ln 2) S(P \| Q)}.
2143: \]
2144: \end{fact}
2145:
2146: We can now prove
2147: Fact~\ref{fact:averageencodingclassical}.
2148:
2149: \ \\
2150: {\bf Fact~\ref{fact:averageencodingclassical} (Average encoding
2151: theorem, classical version)}
2152: {\it
2153: Let $X, M$ be correlated
2154: random variables. Let $p_x$ denote the (marginal) probability
2155: that $X=x$, and $\Pi^x$ denote the conditional distribution of $M$
2156: given that $X=x$.
2157: Let $\Pi$ denote the (marginal) probability distribution of $M$ i.e.
2158: $\Pi = \sum_x p_x \Pi^x$.
2159: Then,
2160: \begin{displaymath}
2161: \sum_x p_x \ellone{\Pi^x - \Pi} \leq \sqrt{(2 \ln 2) I(X:M)}.
2162: \end{displaymath}
2163: \/}
2164: \begin{proof}
2165: Let $\cX$, $\cM$ be the finite ranges of random
2166: variables $X$, $M$ respectively.
2167: We define two probability distributions $P$, $Q$ on
2168: $\cX \times \cM$.
2169: In distribution $P$, the probability of
2170: $(x,m) \in \cX \times \cM$ is
2171: $p_x \cdot \pi_m^x$, where $\pi_m^x$ is the conditional probability
2172: that $M = m$ given that $X = x$. In distribution $Q$, the
2173: probability of $(x,m) \in \cX \times \cM$ is $p_x \cdot \pi_m$, where
2174: $\pi_m$ is the (marginal) probability that $M=m$
2175: i.e. $\pi_m = \sum_x p_x \pi_m^x$.
2176:
2177: It is easy to check that
2178: $S(P \| Q) = I(X:M)$ and
2179: $\ellone{P - Q} = \sum_x p_x \ellone{\Pi_x - \Pi}$.
2180: The result now follows by applying Fact~\ref{thm:divell1} to $P$
2181: and $Q$.
2182: \end{proof}
2183:
2184: \subsection{Quantum}
2185: In this subsection, we give a proof of
2186: Fact~\ref{fact:averageencodingquant}.
2187: We first recall some basic definitions and facts from quantum
2188: information theory.
2189: Let $\rho$ and $\sigma$ be density matrices in the
2190: same finite dimensional Hilbert space $\cH$.
2191: The {\em fidelity} (also called Uhlmann's transition probability
2192: or the Bhattacharya coefficient) of $\rho$ and $\sigma$ is defined as
2193: $B(\rho, \sigma) \defeq \ellone{\sqrt{\rho} \sqrt{\sigma}}$.
2194: The {\em von Neumann relative entropy} between
2195: $\rho$ and $\sigma$ is defined as
2196: $S(\rho \| \sigma) \defeq \Tr (\rho (\log \rho - \log \sigma))$.
2197:
2198: Jozsa~\cite{jozsa:fidelity} gave an elementary proof for finite
2199: dimensional Hilbert spaces of the following basic and remarkable
2200: property about fidelity.
2201: \begin{fact}
2202: \label{fact:jozsa}
2203: Let $\rho, \sigma$ be density matrices in the same
2204: Hilbert space $\cH$. Let $\cK$ be a disjoint
2205: Hilbert space such that $\dim (\cK) \geq \dim (\cH)$. Then
2206: for any purification $\ket{\psi}$ of $\rho$ in $\cH \otimes \cK$,
2207: there exists a purification $\ket{\phi}$ of $\sigma$
2208: in $\cH \otimes \cK$ such that
2209: $B(\rho, \sigma) = |\braket{\psi}{\phi}|$.
2210: \end{fact}
2211:
2212: For two probability distributions $P, Q$ on the same sample space
2213: $\Omega$, their fidelity is defined as
2214: $B(P, Q) \defeq \sum_{x \in \Omega} \sqrt{P(x) Q(x)}$.
2215: We will need the following result about
2216: fidelity proved by Fuchs and Caves~\cite{fuchs:fidelity}.
2217: \begin{fact}
2218: \label{fact:fuchscaves}
2219: Let $\rho, \sigma$ be density matrices in the same
2220: Hilbert space $\cH$. Then
2221: $\displaystyle B(\rho, \sigma)=\inf_{\cM} B(\cM \rho, \cM \sigma)$,
2222: where $\cM$ ranges over POVM's on $\cH$.
2223: In fact, the infimum above can be attained by a complete
2224: von Neumann measurement on $\cH$.
2225: \end{fact}
2226:
2227: The following fundamental fact (see e.g~\cite{nielsen:quant})
2228: states that the relative entropy can only decrease on performing
2229: a measurement.
2230: \begin{fact}[Monotonicity of relative entropy]
2231: \label{fact:monorelentropy}
2232: Let $\rho, \sigma$ be density matrices in the same finite
2233: dimensional Hilbert space $\cH$. Let $\cM$ be a
2234: POVM on $\cH$. Then,
2235: $S(\cM \rho \| \cM \sigma) \leq S(\rho \| \sigma)$.
2236: \end{fact}
2237:
2238: We require the following explicit expression for the
2239: trace distance of two pure states (see e.g.~\cite{nielsen:quant}).
2240: \begin{fact}
2241: \label{fact:trdistpure}
2242: For pure states $\ket{\psi}$ and $\ket{\phi}$ in the same Hilbert
2243: space,
2244: $\ellone{\ketbra{\psi} - \ketbra{\phi}} =
2245: 2 \sqrt{1 - |\braket{\psi}{\phi}|^2}$.
2246: \end{fact}
2247:
2248: The following information-theoretic fact follows easily from the
2249: definitions.
2250: \begin{fact}
2251: \label{fact:inforelentropy}
2252: Let $X, M$ be quantum systems with some joint density matrix, where
2253: $X$ is a classical random variable. Let
2254: $p_x$ denote the (marginal) probability that $X=x$
2255: and $\rho^x$ denote the conditional density matrix
2256: of $M$ given that $X=x$.
2257: Let $\rho$ denote the reduced density matrix of $M$. Note that
2258: $\rho = \sum_x p_x \rho^x$.
2259: Then, $I(X:M) = \sum_x p_x S(\rho^x \| \rho)$.
2260: \end{fact}
2261:
2262: We now recall the following direct connection between relative
2263: entropy and fidelity observed in the classical setting by
2264: \cite{dacunha:BgeqS}, and in the quantum setting by Klauck (private
2265: communication).
2266: \begin{fact}
2267: \label{fact:BleqS}
2268: Let $\rho$ and $\sigma$ be two density matrices in the same
2269: Hilbert space. Then,
2270: \begin{displaymath}
2271: 1 - B(\rho , \sigma) \leq \frac{(\ln 2) S(\rho \| \sigma)}{2}.
2272: \end{displaymath}
2273: \end{fact}
2274: \begin{proof}
2275: Let $\cM$ be the complete von Neumann measurement that
2276: achieves the infimum
2277: in Fact~\ref{fact:fuchscaves}. Let $\Omega$ denote the set of
2278: possible (classical) outcomes of $\cM$.
2279: Define probability distributions $P \defeq \cM \rho$ and
2280: $Q \defeq \cM \sigma$.
2281: From Fact~\ref{fact:monorelentropy} and concavity of the $\log$
2282: function it follows that
2283: \begin{displaymath}
2284: \begin{array}{c}
2285: \displaystyle
2286: -\frac{S(\rho \| \sigma)}{2} \leq -\frac{S(P \| Q)}{2}
2287: = \sum_{x \in \Omega} P(x) \log \sqrt{\frac{Q(x)}{P(x)}} \\
2288: \displaystyle
2289: \leq \log \sum_{x \in \Omega} \sqrt{Q(x) P(x)}
2290: = \log B(P, Q) = \log B(\rho,\sigma).
2291: \end{array}
2292: \end{displaymath}
2293: Thus,
2294: $B(\rho , \sigma)
2295: \geq 2^{- S(\rho \| \sigma) / 2}
2296: = \exp (- (\ln 2) S(\rho \| \sigma) / 2)
2297: \geq 1 - ((\ln 2) S(\rho \| \sigma) / 2)$.
2298: \end{proof}
2299:
2300: We can now prove
2301: Fact~\ref{fact:averageencodingquant}.
2302:
2303: \ \\
2304: {\bf Fact~\ref{fact:averageencodingquant} (Average encoding
2305: theorem, quantum version)}
2306: {\it
2307: Let $X, M$ be quantum systems with some joint density matrix.
2308: Let $X$ be a classical random variable
2309: and $p_x$ denote the (marginal) probability
2310: that $X=x$. Let $\cH$ denote the Hilbert space of $M$
2311: and $\rho^x$ denote the conditional density matrix
2312: of $M$ given that $X=x$.
2313: Let $\rho$ denote the reduced density matrix of $M$. Note that
2314: $\rho = \sum_x p_x \rho^x$.
2315: Let $\cK$ denote a disjoint
2316: Hilbert space such that $\dim(\cK) \geq \dim(\cH)$.
2317: Let $\ket{\psi^x}$ be purifications of $\rho^x$ in
2318: $\cH \otimes \cK$. Then there exist purifications $\ket{\phi^x}$
2319: of $\rho$ in $\cH \otimes \cK$ such that
2320: \begin{displaymath}
2321: \sum_x p_x \ellone{\ketbra{\psi^x} - \ketbra{\phi^x}} \leq
2322: \sqrt{(4 \ln 2) I(X:M)}.
2323: \end{displaymath}
2324: \/} \\
2325: {\bf Proof:}
2326: Using the concavity of the square root function,
2327: Facts~\ref{fact:inforelentropy},
2328: \ref{fact:jozsa}, \ref{fact:trdistpure}, \ref{fact:BleqS},
2329: and the fact that fidelity is always at most $1$, we get
2330: \begin{displaymath}
2331: \begin{array}{c}
2332: \displaystyle
2333: \sqrt{I(X:M)}
2334: = \sqrt{\sum_x p_x S(\rho^x \| \rho)}
2335: \geq \sum_x p_x \sqrt{S(\rho^x \| \rho)}
2336: \geq \sum_x p_x \sqrt{\frac{2 (1 - B(\rho^x, \rho))}{\ln 2}} \\
2337: \displaystyle
2338: \geq \sum_x p_x \sqrt{\frac{1 - (B(\rho^x, \rho))^2}{\ln 2}}
2339: = \sum_x p_x \sqrt{\frac{1 - |\braket{\psi^x}{\phi^x}|^2}
2340: {\ln 2}}
2341: = \sum_x p_x \frac{\ellone{\ketbra{\psi^x} - \ketbra{\phi^x}}}
2342: {\sqrt{4 \ln 2}}.
2343: \end{array}
2344: \end{displaymath}
2345: \qed
2346:
2347: \end{document}
2348:
2349: