1: \documentclass[12pt]{article}
2: \usepackage{amsmath,amssymb,amsthm,xspace,picinpar,graphicx,textcomp}
3: \setlength{\textwidth}{6.5 in} \setlength{\textheight}{8.25in}
4: \setlength{\oddsidemargin}{0in} \setlength{\topmargin}{0in}
5: \addtolength{\textheight}{.8in} \addtolength{\voffset}{-.5in}
6: \newcommand{\old}[1]{}
7: \newcommand{\note}[1]{\begin{center}\fbox{\parbox{5in}{\textbf{Note: #1}}}\end{center}}
8: \renewcommand{\th}{th\xspace}
9: \newcommand{\nd}{nd\xspace}
10: \newcommand{\st}{st\xspace}
11: \newcommand{\textem}[1]{\textit{#1\/}}
12: \newtheorem{thm}{Theorem}[subsubsection]
13: \newtheorem{theorem}{Theorem}[subsubsection]
14: \newtheorem{lem}[theorem]{Lemma}
15: \newtheorem{lemma}[theorem]{Lemma}
16: \newtheorem{cor}[theorem]{Corollary}
17: \newtheorem{prop}[theorem]{Proposition}
18: \newtheorem{conj}{Conjecture}
19: \newcommand{\remark}{\medskip\noindent\textbf{Remark:\ }}
20: \newcommand{\remarkend}{\medskip}
21: \newcommand{\R}{{\mathbb R}}
22: \newcommand{\G}{{G}}
23: \newcommand{\E}{{\mathbb E}}
24: \newcommand{\Eng}{{\cal E}}
25: \newcommand{\C}{{\mathbb C}}
26: \newcommand{\Z}{{\mathbb Z}}
27: \newcommand{\M}{{\cal M}}
28: \newcommand{\eps}{\varepsilon}
29: \newcommand{\Cov}{{\rm Cov}}
30: \newcommand{\Var}{{\rm Var}}
31: \renewcommand{\Re}{{\rm{Re}}}
32: \renewcommand{\Im}{{\rm{Im}}}
33: \newcommand{\Xt}{X_{\text{term}}}
34: \def\eref#1{(\ref{#1})}
35:
36: \newcommand{\vs}{\vspace{.15 in}}
37:
38:
39: \newcommand{\PSbox}[3]{\mbox{\rule{0in}{#3}\hspace{#2}\special{psfile=#1}}}
40: \begin{document}
41: \title{Random-Turn Hex and Other Selection Games}
42: \old{
43: \author{Yuval Peres\thanks{U.C. Berkeley. Research supported in part by NSF grants \#DMS-0244479 and \#DMS-0104073.}\ \thanks{Microsoft Research.}
44: \and Oded Schramm\footnotemark[2]
45: \and Scott Sheffield\footnotemark[1] \footnotemark[2]
46: \and David B. Wilson\footnotemark[2]}
47: }
48: \author{Yuval Peres, Oded Schramm, Scott Sheffield, and David B. Wilson}
49: \date{} \maketitle
50:
51: \old{
52: \begin{abstract}
53: The game of Hex has two players who take turns placing
54: stones of their respective colors on the hexagons
55: of a rhombus-shaped hexagonal
56: grid. Black wins by completing a crossing between two opposite
57: edges, while White wins by completing a crossing between the other
58: pair of opposite edges. Although ordinary Hex is famously difficult
59: to analyze, \textit{Random-Turn\/} Hex---in which players toss a coin
60: before each turn to decide who gets to place the next stone---has a
61: simple optimal strategy. It belongs to a general class of
62: random-turn games---called selection games---in which the
63: expected payoff when both players play the random-turn game
64: optimally is the same as when both players play randomly. We also
65: describe the optimal strategy and study the expected length of the
66: game under optimal play for Random-Turn Hex and several other
67: selection games.
68: \end{abstract}
69: }
70:
71: \renewcommand{\thesubsubsection}{\arabic{subsubsection}}
72: \let\section\subsubsection
73: \let\subsection\paragraph
74: \section{INTRODUCTION.}
75: \subsection*{Overview.}
76:
77: \newcommand{\olympcapt}{A game of Hex played at the 5\th Computer
78: Olympiad in London, August 24, 2000 \cite{hexinfo}. Queenbee
79: \cite{queenbee} (black) played against Hexy \cite{hexy} (white).
80: (There is a special rule regarding the first move whose purpose is
81: to offset the advantage of the first player, see~\cite{hexinfo}.)
82: \label{fig:olympiad}}
83: The game of \textem{Hex}, invented independently by Piet Hein in 1942
84: and John Nash in 1948 \cite{nash}, has two players who take turns placing stones
85: of their respective colors on the hexagons of a rhombus-shaped hexagonal grid
86: (see Figure~\ref{fig:olympiad}).
87: A player wins by completing a path connecting the two opposite sides
88: of his or her color.
89: Although it is easy to show that player~I has a winning strategy,
90: it is far from obvious what that strategy is. Hex is usually played
91: on an $11 \times 11$ board (e.g., the commercial version by Parker
92: Brothers$^{\text{\tiny\textregistered}}$ and the Computer Olympiad Hex Tournament both use $11\times
93: 11$ boards),
94: for which the optimal strategy is not yet known. (For a book on practical Hex strategy see \cite{browne}, and for further information on Hex see \cite{stewart}, \cite{gardner}, \cite{hexinfo}, or \cite{gale}.)
95: \begin{figure}[phtb]
96: \centerline{\includegraphics[width=0.5\textwidth]{hex-olympiad.eps}}
97: \caption{\olympcapt}
98: \end{figure}
99:
100: \textem{Random-Turn Hex} is the same as ordinary Hex, except that
101: instead of alternating turns, players toss a coin before each turn
102: to decide who gets to place the next stone.
103: Although ordinary Hex is famously difficult to analyze, the optimal
104: strategy for Random-Turn Hex turns out to be very simple.
105: We introduce random-turn games in part because they are in many cases
106: more tractable than their deterministic analogs; they also exhibit
107: surprising structure and symmetry. In one general class of games
108: called \textit{selection games}, which includes Hex, the probability
109: that player~I wins when both players play \textit{optimally\/} is
110: precisely the probability that player~I wins when both players
111: play \textit{randomly}. Combining this with Smirnov's recent result
112: about percolation on the hexagonal lattice \cite{smirnov},
113: we will see that in a
114: certain ``fine lattice limit'' winning probabilities of Random-Turn
115: Hex are a ``conformal invariant'' of the board shape. In this and
116: other games, the set of moves played during an entire game (when both
117: players play optimally) has an intriguing fractal structure.
118:
119: On a philosophical note, random-turn games are natural models for
120: real-world conflicts, where opposing agents (political parties,
121: lobbyists, businesses, militaries, etc.)\ do not alternate turns.
122: Instead, they continually seek to improve their positions incrementally.
123: The extent to which they succeed is (at least partially)
124: unpredictable and may be modeled using randomness. Random-turn games
125: are also closely related to ``Richman games'' \cite{MR1685133}. In a
126: Richman game, each player offers money to the other player for the
127: right to make the next move, and the player who offers more gets to
128: move. (At the end of the Richman game, the money has no value.)
129: Another class of random-turn games, where the moves of one player may
130: be reversed by the other player, is studied in
131: \cite{MR1685133} and \cite{pssw-tug}.
132:
133: For simplicity, we limit the discussion in this article to two-player,
134: zero-sum, random-turn games with perfect information.
135:
136: \subsection*{Random-turn selection games.}
137: Now we describe a general class of games that includes the famous game of Hex.
138: Let $S$ be an $n$-element set, which will sometimes be called the
139: \textem{board}, and let $f$ be a function from the $2^n$ subsets of $S$ to
140: $\mathbb R$. A \textem{selection game} is played as follows: the
141: first player selects an element of $S$, the second player selects one
142: of the remaining $n-1$ elements, the first player selects one of the
143: remaining $n-2$, and so forth, until all elements have been chosen.
144: Let $S_1$ and $S_2$ signify the sets chosen by the first and second
145: players, respectively. Then player~I receives a payoff of $f(S_1)$
146: and player~II a payoff of $ - f(S_1)$ (selection games are zero-sum).
147: The following are examples of selection games:
148:
149: \newcommand{\spar}[1]{\subsection*{#1.}}
150:
151: \spar{Hex}
152: Here $S$ is the set of hexagons on a rhombus-shaped $L
153: \times L$ hexagonal grid, and $f(S_1)$ is $1$ if $S_1$ contains a
154: left-right crossing, $-1$ otherwise (see Figures~\ref{fig:olympiad}
155: and \ref{fig:hex11}). In this case, once $S_1$ contains a left-right
156: crossing or $S_2$ contains an up-down crossing (which precludes the
157: possibility of $S_1$ having a left-right crossing), the outcome is
158: determined and there is no need to continue the game.
159: \medskip
160:
161: We also consider Hex played on other types of boards.
162: In the general setting, some hexagons are given to the first or second
163: players before the game has begun. One of the reasons for considering
164: such games is that after a number of moves are played in ordinary Hex,
165: the remaining game has this form.
166:
167: \spar{Surround}
168: The famous game of \textit{Go\/} is not a selection game (for one, a player
169: can remove an opponent's pieces), but the game of \textit{Surround}, in
170: which, as in Go, surrounding area is important, is a selection game.
171: In this game $S$ is the set of $n$ hexagons in a hexagonal grid
172: (of any shape). At the end of the game, each hexagon is recolored to
173: be the color of the outermost cluster surrounding it (if there is such
174: a cluster). The payoff $f(S_1)$ is the number of hexagons recolored
175: black minus the number of hexagons recolored white. (Another natural
176: payoff function is $f^*(S_1)=\operatorname{sign}(f(S_1))$.)
177:
178: \spar{Full-Board Tic-Tac-Toe}
179: Here $S$ is the set of spaces in a $3
180: \times 3$ grid, and $f(S_1)$ is the number of horizontal, vertical, or
181: diagonal rows in $S_1$ minus the number of horizontal, vertical, or
182: diagonal rows in $S \backslash S_1$. This is different from ordinary
183: tic-tac-toe in that the game does not end after the first row is
184: completed.
185:
186: \spar{Team Captains}
187: Two team captains are choosing
188: baseball teams from a finite set $S$ of $n$ players for the purpose of
189: playing a single game against each other. The payoff $f(S_1)$ for the
190: first captain is the probability that the players in $S_1$ (together
191: with the first captain) would beat the players in $S_2$ (together with
192: the second captain). The payoff function may be very complicated
193: (depending on which players are skilled at which positions, which players have
194: played together before, which players get along well with which
195: captain, etc.). Because we have not specified the payoff function,
196: this game is as general as the class of selection games.
197:
198: \vs
199:
200: Every selection game has a random-turn variant in which at each turn
201: a fair coin is tossed to decide who moves next.
202:
203: \vs
204:
205: Consider the following questions:
206:
207: \begin{enumerate}
208: \item What can one say about the probability distribution of $S_1$
209: after a typical game of optimally played Random-Turn Surround?
210: \item More generally, in a generic random-turn selection game, how
211: does the probability distribution of the final state depend on the
212: payoff function $f$?
213: \item Less precise: Are the teams chosen in Random-Turn Team Captains
214: ``good teams'' in any objective sense?
215: \end{enumerate}
216:
217: \noindent
218: The answers are surprisingly simple.
219:
220: \section{OPTIMAL STRATEGY.}
221: \label{c.opt}
222:
223: A (pure) \textem{strategy} for a given player
224: in a random-turn selection game is a
225: map $M$ from pairs of disjoint subsets $(T_1, T_2)$ of $S$ to elements
226: of $S$. Here $M(T_1, T_2)$ indicates the element that the player will
227: pick if given a turn at a time in the game at which player~I
228: has thus far picked the elements of $T_1$ and player~II has picked
229: the elements of $T_2$.
230:
231: Denote by $E(T_1, T_2)$ the expected payoff for player~I at this
232: stage in the game, assuming that both players play optimally with the
233: goal of maximizing expected payoff. As is true for all finite
234: perfect-information, two-player games, $E$ is well defined, and one can
235: compute $E$ and the set of possible optimal
236: strategies inductively as follows. First, if $T_1 \cup T_2 = S$, then
237: $E(T_1, T_2) = f(T_1)$. Next, suppose that we have computed $E(T_1, T_2)$
238: whenever $|S \backslash (T_1 \cup T_2)| \leq k$. Then if $|S
239: \backslash (T_1 \cup T_2)|=k+1$, and player~I has the chance to
240: move, player~I will play optimally if and only if she chooses an $s$
241: from $S \backslash (T_1 \cup T_2)$ for which $E(T_1 \cup \{s\}, T_2)$ is
242: maximal. (If she chose any other $s$, this would reduce her expected
243: payoff.) Similarly, player~II plays optimally if and only if she
244: minimizes $E(T_1, T_2 \cup \{s\})$ at each stage.
245:
246: The foregoing analysis also demonstrates a well-known fundamental fact
247: about finite, turn-based, perfect-information games: both players have
248: optimal pure strategies (i.e., strategies that do not require flipping
249: coins), and knowing the other player's strategy does not give a player
250: any advantage when both players play optimally. (This contrasts with
251: the situation in which the players play ``simultaneously,'' as they do
252: in Rock-Paper-Scissors.)
253:
254: \begin{theorem} \label{t.rts}
255: The value of a random-turn selection game is the expectation of
256: $f(T)$ when a set $T$ is selected randomly and uniformly among
257: all subsets of $S$. Moreover, any optimal strategy for one of the
258: players is also an optimal strategy for the other player.
259: \end{theorem}
260:
261: \begin{proof}
262: If player~II plays any optimal strategy, player~I can achieve the
263: expected payoff $\E[f(T)]$ by playing exactly the same strategy
264: (since, when both players play the same strategy, each element will
265: belong to $S_1$ with probability $1/2$, independently). Thus, the
266: value of the game is at least $\E[f(T)]$. However, a symmetric
267: argument applied with the roles of the players interchanged implies
268: that the value is no more than $\E[f(T)]$.
269:
270: Suppose that $M$ is an optimal strategy for the first player. We
271: have seen that when both players use $M$, the expected payoff is
272: $\E[f(T)]=E(\emptyset,\emptyset)$. Since $M$ is optimal for
273: player~I, it follows that when both players use $M$ player~II always
274: plays optimally (otherwise, player~I would gain an advantage, since
275: she is playing optimally). This means that $M(\emptyset,\emptyset)$
276: is an optimal first move for player~II, and therefore every optimal
277: first move for player~I is an optimal first move for player~II. Now
278: note that the game started at any position is equivalent to a
279: selection game. We conclude that every optimal move for one of the
280: players is an optimal move for the other, which completes the proof.
281: \end{proof}
282:
283: If $f$ is identically zero, then all strategies are optimal. However,
284: if $f$ is \textit{generic\/} (meaning that all of the values $f(S_1)$
285: for different subsets $S_1$ of $S$ are linearly independent over
286: $\mathbb Q$), then the preceding argument shows that the optimal
287: choice of $s$ is always unique and that it is the same for both
288: players. We thus have the following result:
289:
290: \begin{prop}\label{p.generic}
291: If $f$ is generic, then there is a unique optimal strategy and it is
292: the same strategy for both players. Moreover, when both players
293: play optimally, the final $S_1$ is equally likely to be any one of
294: the $2^n$ subsets of $S$.
295: \end{prop}
296:
297: Theorem~\ref{t.rts} and Proposition~\ref{p.generic}
298: are in some ways quite surprising. In the
299: baseball team selection, for example, one has to think very hard in
300: order to play the game optimally, knowing that at each stage there
301: is exactly one correct choice and that the adversary can capitalize
302: on any miscalculation. Yet, despite all of that mental effort by the
303: team captains, the final teams look no different than they would
304: look if at each step both captains chose players uniformly at
305: random.
306:
307: Also, for purposes of illustration, suppose that there are only two players who
308: know how to pitch and that a team without a pitcher always loses.
309: In the alternating turn game, a captain can always wait to select a
310: pitcher until just after the other captain selects a pitcher. In
311: the random-turn game, the captains must try to select the pitchers
312: in the opening moves, and there is an even chance the pitchers will
313: end up on the same team.
314:
315: Theorem~\ref{t.rts} and Proposition~\ref{p.generic} generalize to
316: random-turn selection games in which the player to get the next turn
317: is chosen using a biased coin. If player~I gets each turn with
318: probability $p$, independently, then the value of the game is $\E[f(T)]$, where
319: $T$ is a random subset of $S$ for which each element of $S$ is
320: in $T$ with probability $p$, independently.
321: For the corresponding statement of the proposition to hold, the notion
322: of ``generic'' needs to be modified. For example, it suffices to
323: assume that the values of $f$ are linearly independent over $\mathbb Q[p]$.
324: The proofs are essentially the same. We leave it as an exercise to the
325: reader to generalize the proofs further so as to include the following
326: two games:
327:
328: \spar{Restaurant Selection}
329: Two parties (with opposite food preferences) want to select a dinner
330: location. They begin with a map containing $2^n$ distinct points
331: in $\R^2$, indicating restaurant locations. At each step,
332: the player who wins a coin toss may draw a straight line that divides
333: the set of remaining restaurants exactly in half and eliminate all the
334: restaurants on one side of that line. Play continues until one
335: restaurant $z$ remains, at which time player~I receives payoff
336: $f(z)$ and player~II receives $-f(z)$.
337:
338: \spar{Balanced Team Captains}
339: Suppose that the captains wish
340: to have the final teams equal in size (i.e., there are $2 n$ players
341: and we want a guarantee that each team will have exactly $n$ players
342: in the end). Then instead of tossing coins, the captains may shuffle a
343: deck of $2n$ cards (say, with $n$ red cards and $n$ black cards). At
344: each step, a card is turned over and the captain whose color is shown
345: on the card gets to choose the next player.
346:
347: \section{WIN-OR-LOSE SELECTION GAMES.}
348:
349: We say that a game is a \textem{win-or-lose} game
350: if $f(T)$ takes on precisely two values, which we may as well
351: assume to be $-1$ and $1$.
352: If $S_1\subset S$ and $s\in S$, we say
353: that $s$ is \textem{pivotal} for $S_1$ if
354: $f(S_1\cup\{s\})\ne f(S_1\setminus\{s\})$.
355: A selection game is \textem{monotone} if $f$ is monotone; that is,
356: $f(S_1)\ge f(S_2)$ whenever $S_1\supset S_2$.
357: Hex is an example of a monotone, win-or-lose game. For such games,
358: the optimal moves have the following simple description:
359:
360: \begin{lemma}\label{l.piv}
361: In a monotone, win-or-lose, random-turn selection game,
362: a first move $s$ is optimal if and only if $s$ is an element
363: of $S$ that is most likely to be pivotal for a random-uniform
364: subset $T$ of $S$.
365: When the position is $(S_1,S_2)$, the move $s$ in $S\setminus(S_1\cup S_2)$
366: is optimal if and only if $s$ is an element of $S\setminus(S_1\cup S_2)$
367: that is most likely to be pivotal for $S_1\cup T$, where
368: $T$ is a random-uniform subset of $S\setminus(S_1\cup S_2)$.
369: \end{lemma}
370:
371: \noindent
372: The proof of the lemma is straightforward at this point and is
373: left to the reader.
374:
375: For win-or-lose games, such as Hex, the players may stop making moves
376: after the winner has been determined, and it is interesting to
377: calculate how long a random-turn, win-or-lose, selection game will last
378: when both players play optimally. Suppose that the game is a
379: monotone game and that, when there is more than one optimal move, the
380: players break ties in the same way. Then we may take the point of
381: view that the playing of the game is a (possibly randomized) decision
382: procedure for evaluating the payoff function $f$ when the items are
383: randomly allocated. Let $\vec x$ denote the allocation of the items,
384: where $x_i=\pm 1$ according to whether the $i$\th item goes to the
385: first or second player. We may think of the $x_i$ as input bits,
386: and the playing of the game is one way to compute $f(\vec x)$. The
387: number of turns played is the number of bits of $\vec x$ examined
388: before $f(\vec x)$ is computed. We can use certain inequalities from the
389: theory of Boolean functions to bound the average length of play.
390:
391: Let $I_i(f)$ denote the influence of the $i$\th bit on $f$ (i.e., the
392: probability that flipping $x_i$ will change the value of $f(\vec x)$).
393: The following inequality is from O'Donnell and Servedio
394: \cite{odonnell-servedio}:
395: \begin{multline}
396: \sum_i I_i(f) =
397: \E\left[\sum_i f(\vec x) x_i\right] =
398: \E\left[f(\vec x) \sum_i x_i 1_{\text{$x_i$ examined}}\right]
399: \le \text{(by Cauchy-Schwarz)}\\
400: \sqrt{\E[f(\vec x)^2]\,\E\left[\left(\sum_{\text{$i$: $x_i$ examined}} x_i\right)^2 \right]}
401: =
402: \sqrt{ \E\left[\left(\sum_{\text{$i$: $x_i$ examined}} x_i\right)^2 \right]}
403: = \sqrt{\E[ \text{\# bits examined}]}\,.\label{eq:os}
404: \end{multline}
405: The last equality is justified by noting that
406: $\E[x_i\,x_j\, 1_{\text{$x_i$ and $x_j$ both examined}}]=0$
407: when $i\ne j$, which holds since conditioned on $x_i$ being examined
408: before $x_j$, conditioned on the value of $x_i$, and conditioned
409: on $x_j$ being examined, the expected value of $x_j$ is zero.
410: By~\eqref{eq:os} we have $$\E[\text{\# turns}] \geq \left[\sum_i I_i(f)\right]^2.$$
411: We shall shortly apply this bound to the game of Random-Turn Hex.
412:
413: We mention one other inequality from the theory of Boolean functions, this
414: one due to O'Donnell, Saks, Schramm, and Servedio
415: \cite[Theorem 3.1]{odonnell-saks-schramm-servedio}:
416: \begin{equation}\label{eq:osss}
417: \Var[f] \leq \sum_i \Pr[\text{$x_i$ examined}] I_i(f).
418: \end{equation}
419: For the random-turn game this implies that
420: $$ \E[\text{\# turns}] \geq \frac{\Var[f]}{\max_i I_i(f)}.$$
421:
422: \section{RANDOM-TURN HEX.}
423: \label{s.hex}
424:
425: \subsection*{Odds of winning on large boards and under biased play.}
426: In the game of Hex, the propositions discussed earlier imply that the probability that
427: player~I wins is given by the probability that there is a left-right
428: crossing in independent Bernoulli percolation
429: on the sites (i.e., when the sites are independently and randomly colored black or white). One perhaps surprising
430: consequence of the connection to Bernoulli percolation is that, if
431: player~I has a slight edge in the coin toss and wins the coin toss
432: with probability $1/2+\eps$, then for any $\eps>0$, any $\delta>0$,
433: and any $r>0$
434: there is a strategy for player~I
435: that wins with probability at least $1-\delta$ on
436: the $L\times rL$ board, provided that $L$ is sufficiently large.
437:
438: \subsection*{Random-Turn Hex on ordinary-size boards.}
439:
440: Reisch proved in 1981 that determining which player has a winning
441: strategy on an arbitrarily shaped Hex board (with some of the sites
442: already colored in) is PSPACE-complete \cite{reisch}. The online
443: community has, however, made a good deal of progress (much of it
444: unpublished) on solving the problem for smaller boards. Jing Yang
445: \cite{yang} has announced the solution of Hex (and provided associated
446: computer programs) on boards of size up to $9\times 9$. Hex is
447: usually played on an $11 \times 11$ board, for
448: which the optimal strategy is not yet known.
449:
450: What is the situation with Random-Turn Hex? We do not know if the
451: correct move in Random-Turn Hex can be found in polynomial time. On
452: the other hand, for any fixed $\eps$ a computer can sample $O(L^4 \eps^{-2}
453: \log (L^4/\eps))$ percolation configurations (filling
454: in the empty hexagons at random) to estimate which empty site is most
455: likely to be pivotal given the current board configuration. Except
456: with probability $O(\eps/L^2)$, the computer will pick a site that is
457: within $O(\eps/L^2)$ of being optimal. This simple randomized strategy
458: provably beats an optimal opponent $50 - \eps$ percent of the time.
459:
460: \begin{figure}[phtb]
461: \centerline{\includegraphics[width=.48\textwidth]{hex-11.eps}\hfill\includegraphics[width=.48\textwidth]{hex-63.eps}}
462: \caption{Random-Turn Hex on boards of size $11\times 11$ and $63\times 63$ under (near) optimal play.}
463: \label{fig:hex11}
464: \label{fig:hex63}
465: \end{figure}
466:
467: \subsection*{Typical games under optimal play.}
468:
469: What can we say about how long an average game of Random-Turn Hex will
470: last, assuming that both players play optimally? (Here we assume
471: that the game is stopped once a winner is determined.) If the side
472: length of the board is $L$, we wish to know how the expected length of
473: a game grows with $L$ (see Figure~\ref{fig:hex63} for games
474: on a large board). Computer simulations on a variety of board sizes suggest
475: that the exponent is in the range $1.5$--$1.6$. As far as rigorous bounds
476: go, a trivial upper bound is $O(L^2)$. Since the game does not end
477: until a player has found a crossing, the length of the shortest
478: crossing in percolation is a lower bound, and empirically this
479: distance grows as $L^{1.1306\pm0.0003}$ \cite{MR1732786}, where the
480: exponent is known to be strictly larger than $1$. We give a stronger
481: lower bound:
482: \begin{thm}\label{thm:hex-length}
483: Random-Turn Hex under optimal play on an order $L$ board, when the
484: two players break ties in the same manner, takes at least
485: $L^{3/2+o(1)}$ time on average.
486: \end{thm}
487: \begin{proof}
488: To use the O'Donnell-Servedio bound~\eqref{eq:os}, we need to know the influence
489: that the sites have on whether or not there is a percolation crossing
490: (a path of black hexagons connecting the two opposite black sides).
491: The influence $I_i(f)$ is the probability that flipping site $i$ changes
492: whether there is a black crossing or a white crossing. The ``4-arm
493: exponent'' for percolation is $5/4$ \cite{MR1879816}
494: (as predicted earlier in \cite{coniglio}), so
495: $I_i(f)=L^{-5/4+o(1)}$ for sites $i$ ``away from the boundary,'' say in the
496: middle ninth of the region. Thus $\sum_i I_i(f) \geq L^{3/4+o(1)}$,
497: so $\E[ \text{\# turns}] \geq L^{3/2+o(1)}$.
498: \end{proof}
499:
500: \noindent
501: \textbf{Remark.} The function $\sum_i x_i 1_{\text{$x_i$ examined}}$
502: is the number of extra times that player~I wins the coin toss at the
503: time that the game terminates. Naturally this is correlated with the
504: winner $f(\vec x)$ of the game, and for large board sizes this
505: correlation is noticeable. Since the only inequality used in \eqref{eq:os}
506: was the Cauchy-Schwarz inequality, if we know $\sum_i I_i(f)$ and the
507: expected length of the game, we can determine how correlated the
508: winner is with who won most of the coin tosses. Using more detailed
509: knowledge of how $I_i(f)$ behaves near the boundary, one can show that
510: for the standard lozenge-shaped Hex board, $\sum_i I_i(f)=L^{3/4+o(1)}$.
511: If the game lasts for $L^{>1.5}$ steps on average, this would imply
512: that the correlation between the winner of the game and the winner of
513: the majority of the coin tosses before the game is won tends to $0$ as
514: $L\to\infty$.
515:
516: \vs
517:
518: \noindent
519: \textbf{Remark.} The question of determining how many bits need to
520: be examined before one can decide whether or not there is a
521: percolation crossing arose in the context of dynamical percolation,
522: where the sites flip according to independent Poisson clocks
523: \cite{schramm-steif}. Roughly speaking, if few bits need to be
524: examined, then it is easier for there to be exceptional times at which
525: there is an infinite percolating cluster in the plane
526: \cite{schramm-steif}. One possible algorithm for determining whether
527: or not there is a crossing in an order $L$ region would be to follow
528: the black-white interfaces starting at the corners to see how they
529: connect, which exposes $L^{7/4+o(1)}$ hexagons in expectation
530: \cite{MR1879816}. While this algorithm has the best currently
531: provable bound on the number of exposed hexagons, the ``play
532: Random-Turn Hex'' algorithm appears to do better (taking about
533: $L^{\text{$1.5$--$1.6$}}$ time).
534:
535: \vs
536:
537: An optimally played game of Random-Turn Hex on a small board may
538: occasionally have a move that is disconnected from the other played
539: hexagons, as the game in Figure~\ref{fig:discon} shows.
540: But this is very much the exception rather than the rule. For
541: moderate- to large-sized boards it appears that in almost every
542: optimally played game, the set of played hexagons remains a connected
543: set throughout the game (which is in sharp contrast to the usual game
544: of Hex). We do not have an explanation for this phenomenon,
545: nor is it clear to us if it persists as the board size increases beyond
546: the reach of simulations.
547: \begin{figure}[phtb]
548: \centerline{\includegraphics[width=.35\textwidth]{hex-disconnected.eps}}
549: \caption{A rare occurrence---a game of Random-Turn Hex under (near) optimal play with a disconnected play.\label{fig:discon}}
550: \end{figure}
551:
552: \subsection*{Conformal (non)invariance of Random-Turn Hex.}
553:
554: Combined with Smirnov's celebrated recent work in percolation
555: \cite{smirnov}, the connection between Random-Turn Hex and percolation
556: enables us to use Cardy's formula \cite{cardy} to approximate
557: player~I's probability of winning on very large boards of various
558: shapes, such as the $L\times r L$ rectangle. The way this is done is
559: as follows. Suppose that the players play the game on a very large
560: board that, when suitably scaled,
561: approximates some simply connected domain $D$ in the
562: complex plane, where the boundary of $D$ alternates colors at the
563: points $b_1,b_2,b_3,b_4$ with one black side between $b_1$ and $b_2$
564: and the other between $b_3$ and $b_4$. By the Riemann Mapping
565: Theorem, there is an analytic function $\phi$ that bijectively maps
566: the domain $D$ to the upper half plane (i.e., a conformal bijection)
567: such that $\phi(b_1)=0$, $\phi(b_3)=1$, and $\phi(b_4)=\infty$.
568: Smirnov proved that in percolation on boards that are shaped like the
569: domain $D$, the probability of a black percolation crossing is a
570: function of $\phi(b_2)$ plus an error term that goes to zero as the
571: board size (the length of the shortest path connecting opposite sides)
572: goes to infinity. This function of $\phi(b_2)$ is an
573: explicit formula (involving a hypergeometric function) that is known
574: as \textit{Cardy's formula}. These crossing probabilities are known as a
575: ``conformal invariant'' of percolation because they remain invariant
576: under conformal maps (up to error terms that go to zero as the board
577: size goes to infinity). The connection between Random-Turn Hex and
578: percolation tells us that when the two players play optimally, the
579: probability of black winning is (up to an additive $o(1)$ error)
580: conformally invariant, and in particular is given by Cardy's formula.
581:
582: However, the actual game play is \textit{not\/} conformally invariant,
583: and indeed, even the location of the first move is not conformally
584: invariant (see Figure~\ref{fig:pivotal}). The reason for this is not
585: difficult to understand. Recall that the first move is played at the
586: site that is most likely to be pivotal. The probability that a site
587: is pivotal is not scale invariant (for larger boards, the probability
588: that a site is pivotal is smaller). Since a general conformal map
589: scales different parts of the domain by different amounts, the
590: location of the site most likely to be pivotal is not conformally
591: invariant.
592:
593: \begin{figure}[phtb]
594: \centerline{\includegraphics[width=.58\textwidth]{hex-pivotal-lozenge.eps}\hfill\includegraphics[width=.38\textwidth]{hex-pivotal-disk.eps}}
595: \caption{The location of the best first move in Random-Turn Hex is the site that is most likely to be pivotal for a percolation crossing. Shown here are estimates of the probability that a site is critical when the board is the standard Hex board (on the left) and when the board is a disk (on the right), calculated using SLE \cite{schramm-wilson}. For the standard board the best first move is near the center, while for the disk-shaped board the best first move is near one of the four points where the black boundary and white boundary meet.}
596: \label{fig:pivotal}
597: \end{figure}
598:
599: \section{THREE EXAMPLES ON TREES.}
600: \label{s.tree}
601:
602: In this section, we study two different random-turn games based on trees.
603: Although these particular examples can be analyzed rather well, there are
604: other natural games based on trees that we do not know how to analyze.
605: We mention one such example at the end of the section (Recursive Three-Fold
606: Majority).
607:
608: \subsection*{AND-OR trees.}
609:
610: \newcommand{\Tr}{\textsf{T}\xspace}
611: \newcommand{\Fa}{\textsf{F}\xspace}
612: We believe that the expected length of game play in Random-Turn Hex
613: grows like a nonintegral power of the selection set size $|S|$ and that the
614: set of hexagons played has a random fractal structure. The game of
615: Random-Turn AND-OR (see Figure~\ref{fig:and-or})
616: is a selection game for which we can actually prove
617: analogous statements.
618:
619: \begin{figure}[phtb]
620: \centerline{\includegraphics[width=\textwidth]{hex-andor.eps}}
621: \caption{AND-OR tree of depth 4. In Random-Turn AND-OR, the player who wins the coin toss sets the bit at a leaf node. The value of the function determines the winner of the game.}
622: \label{fig:and-or}
623: \end{figure}
624:
625: The selection set $S$ is the set of leaves in a depth $h$ complete
626: binary tree. We number the levels of the tree from $0$ (the root) to
627: $h$ (the leaves). The leaves are treated as binary bits, with labels
628: of \Tr (``True'') or \Fa (``False'') if they are chosen by player~I or
629: II, respectively. The internal nodes of the tree are also treated as
630: binary bits; the players cannot select the internal nodes, instead the
631: label of a bit on the $k$\th level of the tree ($k<h$) is the AND of
632: its two children if $k$ is even and the OR if $k$ is odd. Player~I
633: wins if the root label is \Tr; player~II wins if it is \Fa.
634:
635: When player~I always wins the coin toss with probability $p$ (which
636: need not be $1/2$), it is straightforward to estimate the probability
637: that she wins the game. Let $q_k$ be the probability that a label at
638: the $k$\th level is \Tr. Then $q_k = q_{k+1}^2$ (if $k$ is even) and
639: $q_k = 2 q_{k+1} - q_{k+1}^2$ (if $k$ is odd). When $k$ is even,
640: $q_k = (2 q_{k+2} - q_{k+2}^2)^2$. The map $q \mapsto (2q-q^2)^2$ has a
641: fixed point when $$q^4 - 4q^3 + 4q^2 -q = q(q-1)(q^2-3q+1) = 0$$
642: (i.e., if $q \in \{0,1,(3 \pm \sqrt{5})/2\}$); the fixed points in
643: $[0,1]$ are $0$, $1$, and $(3-\sqrt{5})/2 \approx .382$. If $p$
644: is fixed and $h$ tends to $\infty$ along even integers, then the
645: probability player~I wins tends to $0$ if $p < (3-\sqrt 5)/2$
646: and to $1$ if $p > (3-\sqrt 5)/2$.
647:
648: \begin{theorem}\label{th:andor}
649: Consider an optimally played game of AND-OR on a level $h$ tree with
650: coin-toss probability $p$. If a move is played in some subtree $T$
651: below some vertex $v$, then the succeeding moves are all played in
652: $T$ until the label of $v$ is determined. Moreover, the labels of
653: the level $h-1$ vertices are determined in an order that is an
654: optimally played game on the tree truncated at level $h-1$.
655: \end{theorem}
656: \begin{proof}
657: When a complete set of labels of the leaves (and hence all other
658: vertices) is fixed, we say a vertex on level $k$ is pivotal if it is a
659: pivotal element of the game on the tree truncated at level $k$. A
660: vertex on level $k$ is pivotal if and only if its parent is pivotal
661: \textit{and\/} its sibling is \Tr (if $k$ is odd) or \Fa (if $k$ is
662: even). Thus, a vertex is pivotal if and only if the path from that
663: vertex to the root has the property that every sibling of a vertex
664: along that path is \Tr or \Fa as the level is odd or even. Note that
665: the event ``$v$ is pivotal'' is independent of labels on the subtree
666: below $v$.
667:
668: We assume, inductively, that the statement of the theorem is true for
669: a tree with fewer than $h$ levels and for the first $t$ steps on a
670: level $h$ tree. Before the $(t+1)$\th step is played, there are
671: either $0$ or $1$ partially determined vertices at level $h-1$ (that
672: is, vertices that are undetermined but for which a leaf below them has
673: been played). If there are none, then the leaves that are most likely
674: to be pivotal are children of the level $h-1$ vertices that are most
675: likely to be pivotal, so any optimal $(t+1)$\th step also satisfies
676: the desired condition. Suppose now that there is exactly one
677: partially determined vertex $v$ at level $h-1$, and call its labeled
678: child $x$ and its yet unlabeled child $y$. Let $z$ be any unlabeled
679: leaf different from $y$ that might still affect the outcome of the
680: game. Leaf $x$ must have been labeled at step $t$. We need to show
681: that playing $y$ as the $(t+1)$\th move is better than playing $z$.
682:
683: Let $r_t(u)$ be the conditional probability that a vertex $u$ is
684: pivotal given the labels that have been set prior to move $t$. Let
685: $\bar p=p$ if $h$ is odd and $\bar p=1-p$ if $h$ is even ($\bar p$ is
686: the probability that the value of a leaf node does not determine its
687: parent). Since $x$ was preferred over $z$, we have $r_t(x)\ge
688: r_t(z)$. Now $r_t(z)$ is a nonnegative martingale, which is to say
689: that $\E[r_{t+1}(z)|\text{labels at time $t$}]=r_t(z)\geq 0$. Since
690: $x$ got labeled the way it did with probability $\bar p$, we have
691: $\bar p r_{t+1}(z)\le r_t(z)\le r_t(x)=\bar p r_t(v)=\bar p
692: r_{t+1}(y)$. Thus, we see that at step $t+1$, playing $y$ is at least
693: as good as playing $z$. We need only rule out the case of equality,
694: namely, $r_{t+1}(z)=r_{t+1}(y)$.
695:
696: Let $a$ be the least common ancestor of $y$ and $z$, let $b_1$ be the
697: child of $a$ above $y$, and let $b_2$ be the child of $a$ above $z$.
698: Suppose that $x$ was the first move played in the subtree below $a$.
699: Regardless of which player got leaf $x$, the label of $a$ would not be
700: determined in turn $t$. Thus, $z$ would still be pivotal with
701: positive probability, so the martingale argument would actually give
702: the strict inequality $\bar p r_{t+1}(z)< r_t(z)$ in place of the weak
703: inequality we used earlier. Consequently, we get
704: $r_{t+1}(z)<r_{t+1}(y)$, as required.
705:
706: On the other hand, suppose that $x$ was not the first move played
707: below $a$. Then by induction, the $(t-1)$\th move was played below
708: $b_1$ and, again invoking the inductive hypothesis, we see that
709: $r_t(z)<r_t(x)$. Thus, we get $r_{t+1}(z)<r_{t+1}(y)$ in this case as
710: well.
711: \end{proof}
712:
713: Note that the Theorem~\ref{th:andor} completely characterizes the optimally
714: played games, whether or not the two players break ties in the same
715: way. In fact, for every optimally played game there is an embedding
716: of the tree in the plane for which at each turn the played leaf is the
717: leftmost leaf that at that point has a positive probability to be
718: pivotal.
719:
720: \begin{theorem}
721: If $h$ is even and $p = (3-\sqrt 5)/2$, then the expected
722: length of the game {\rm(}i.e., the expected number of labeled leaves{\rm)}
723: is precisely $$\left(\frac{1+\sqrt 5}{2}\right)^h = (1.6180\ldots)^h.$$
724: \end{theorem}
725: \begin{proof}
726: When all leaves are randomly labeled, a level $k$ vertex is $1$ with
727: probability $p$ if $k$ is even and $1-p$ if $k$ is odd. Under
728: optimal play, once a vertex's label is determined, with
729: probability $p$ it determines the label of its parent. Thus,
730: given that a vertex $v$ is labeled during an optimally played game,
731: the expected number of labeled children of $v$ is $1+1-p$.
732: \end{proof}
733:
734: Since we know the precise expected length of the game, we take a
735: moment to compare this with the lower bounds from the theory of
736: Boolean functions. When the bits $x_i$ are true with probability $p$,
737: rather than taking $x_i=\pm1$ according to whether the bit is true or
738: false, it turns out to be more natural to take
739: $\Tr=\sqrt{(1-p)/p}$ and $\Fa=-\sqrt{p/(1-p)}$. Then $\E[x_i]=0$
740: and $\E[x_i^2]=1$, so the O'Donnell-Servedio bound is still valid.
741: For AND-OR trees with $p=(3-\sqrt5)/2$, the influence of each
742: bit is $(\frac{-1+\sqrt5}{2})^h$, and there are $2^h$ bits, whence the
743: O'Donnell-Servedio bound gives
744: $$\E[\text{\# turns}] \geq
745: (-1+\sqrt{5})^{2 h} = (1.5279\ldots)^h.
746: $$
747:
748: The O'Donnell-Saks-Schramm-Servedio bound~\eqref{eq:osss}
749: generalizes to the
750: $p$-biased case~\cite[Theorem 3.1]{odonnell-saks-schramm-servedio}
751: and yields
752: $$\E[\text{\# turns}] \geq \frac{1}{(\frac{-1+\sqrt5}{2})^h} =
753: \left(\frac{1+\sqrt 5}{2}\right)^h = (1.6180\ldots)^h.$$
754: This lower bound is exactly tight for AND-OR trees.
755:
756:
757: \subsection*{Shannon Switching Game.}
758:
759: We now discuss a game for which we can prove that the set of moves is
760: always connected under optimal play. In some ways, the analysis is
761: similar to the analysis of AND-OR games.
762:
763: The \textit{Shannon Switching Game\/} is played by two players (named Cut and
764: Short) on a graph with two distinguished vertices. When it is Cut's
765: turn, he may delete any unplayed edge of the graph, while Short may render an
766: edge immune to being cut. Cut wins if he manages to disconnect the
767: two distinguished vertices, otherwise Short wins. When the graph is
768: an $(L+1)\times L$ grid, with the vertices of the left side merged
769: into one distinguished vertex and the vertices on the other side
770: merged into another distinguished vertex, then the game is called
771: Bridg-It (also known as Gale after its inventor David Gale). Oliver
772: Gross showed that the first player in Bridg-It has a (simple) winning
773: strategy, and then Lehman \cite{lehman} (see also \cite{mansfield})
774: showed how to solve the general Shannon Switching Game. Here we
775: consider the random-turn version of the Shannon Switching Game.
776:
777: Just as Random-Turn Hex is connected to site percolation on the
778: triangular lattice, where the vertices of the lattice (or, equivalently,
779: faces of the hexaognal lattice) are independently colored black or
780: white with probaiblity $1/2$, Random-Turn Bridg-It is connected to
781: bond percolation on the square lattice, where the edges of the square
782: lattice are independently colored black or white with probability
783: $1/2$. We don't know the optimal strategy for Random-Turn Bridg-It,
784: but as with Random-Turn Hex, we can make a randomized algorithm that
785: plays near optimally. Less is known about bond percolation than site
786: percolation, but it is believed that the crossing probabilities for
787: these two processes are asymptotically the same on
788: ``nice'' domains~\cite{MR1230963}, %Langlands et al
789: so the probability that Cut wins in Random-Turn Bridg-It is
790: well approximated by the probability that a player wins in Random-Turn
791: Hex on a similarly shaped board.
792:
793: Consider the Random-Turn Shannon Switching Game on a tree, where the
794: root is one distinguished vertex and the leaves are (collectively)
795: the other vertex. Thus,
796: Short wins if at the end of the game there is a path from the
797: root to one of the leaves. Under optimal play, the probability that Short wins
798: is just the probability that there is a path from the root to some
799: leaf when each edge is independently deleted with probability $1/2$.
800: For the complete ternary tree with $h$ levels, the probability that
801: Short wins under optimal play converges to $3-\sqrt{5}\approx0.764$
802: for large $h$ (see the proof of Theorem~\ref{thm:ternary}).
803: For the complete binary tree with $h$ levels Short
804: wins with probability $\sim 4/h$, so we define the ``enhanced binary
805: tree'' to be the tree where the root has $\lfloor h \log 2/2\rfloor$
806: children each of whom fathers a complete binary tree with $h-1$
807: levels. Then the Random-Turn Shannon Switching Game on this enhanced
808: binary tree is an approximately fair game when $h$ is large, meaning
809: that Short wins with probability $1/2+o(1)$. But what
810: are the optimal strategies, and how long does an average game last if
811: both players play optimally? We will discuss these issues presently.
812:
813: There is a natural coupling of a game with bond percolation on the tree.
814: The edges played by Short are the open edges, and those played by
815: Cut are the closed edges. The unplayed edges may be considered undecided.
816: We assume that the game terminates when the outcome is decided.
817:
818: In the following result, a subtree below a vertex $v$ in a rooted tree
819: consists of $v$ and all the vertices and edges that are separated
820: from the root by $v$.
821:
822: \begin{theorem}
823: Consider the Random-Turn Shannon Switching Game on a tree of depth
824: $h$ in which each internal node {\rm(}except possibly the root{\rm)} has $b$
825: children and all the leaves are at level $h$. Under optimal play,
826: the set of moves played by Short forms a connected set of edges
827: containing the root, and the set of moves played by Cut are all
828: adjacent to the edges played by Short. At any position, the set of
829: optimal moves {\rm(}for either player{\rm)} consists of all the edges closest
830: to the leaves among the undecided edges adjacent to shorted edges.
831: Consequently, whenever an edge is played in some subtree $T$ below
832: some vertex, the subsequent moves are played in $T$ until all the
833: leaves of $T$ are disconnected from the root by cut edges or until
834: Short wins.
835: \end{theorem}
836:
837: \begin{proof}
838: After Short and Cut have made some moves, define the residual tree to
839: be the graph formed from the original tree by deleting the subtrees
840: disconnected from the root by cut edges and contracting
841: edges that have been shorted.
842: Consider some position of an optimally played game.
843: By induction, we assume that in the residual tree corresponding to
844: the current position
845: the direct descendants of the root are themselves roots of
846: regular $b$-ary subtrees of various depths whose leaves are
847: leaves of the original tree.
848: (The base of the induction is clear.)
849: Suppose that $e=[x,y]$ is an edge in the residual tree, where
850: $x$ is closer to the root than $y$.
851: The probability that $e$ is pivotal for bond percolation
852: from the root to the leaves (i.e., given the status of the
853: other edges, there is a connection from the root to the leaves if $e$
854: is uncut and there is no connection if $e$ is cut)
855: is precisely the probability that $y$ is connected to the leaves
856: (by uncut edges),
857: $x$ is connected to the root, and every open path from the root
858: to the leaves passes through $e$.
859: This is clearly maximized by edges in the residual tree adjacent to the root.
860: Among the edges adjacent to the root, this is maximized by
861: those closest to the leaves, because the probability
862: of having an open path to the root in Bernoulli percolation on
863: the regular $b$-ary tree is monotone decreasing in the depth of the tree.
864: By Lemma~\ref{l.piv}, these correspond to the optimal moves.
865: When one of these edges is shorted or cut, the stated
866: structure of the residual tree is preserved. Thus, the induction step is
867: established. The theorem follows.
868: \end{proof}
869:
870: Next, we consider how long an average optimally played game lasts. In
871: the following two theorems we use the notation (common in computer
872: science) $\Theta(g)$ to denote a quantity that is bounded between $c_1
873: g$ and $c_2 g$, where $c_1$ and $c_2$ are positive universal constants.
874: (The related notation $O(g)$ denotes an expression that is upper
875: bounded by $c_2 g$, but which may or may not be bounded below.)
876: \begin{theorem}
877: Random-Turn Switching on the enhanced binary tree of depth $h$ lasts
878: on average $\Theta(h^2)$ turns under optimal play.
879: \end{theorem}
880: \begin{proof}
881: For the complete binary tree of depth $h$ the expected number of
882: vertices in the percolation component of the root is $h+1$.
883: Accordingly, the expected number
884: of played edges of the enhanced binary
885: tree is $O(h^2)$. For the lower bound, note that for the complete
886: binary tree the expected number of vertices in the percolation
887: component of the root
888: conditional on there being no leaf in this component is $(1+o(1)) h$.
889: (This quantity is easily calculated inductively, for example.) For the
890: enhanced binary tree, there is a good chance that $\Theta(h)$
891: subtrees of the root are explored, so that on average at least
892: $\Theta(h^2)$ moves are played.
893: \end{proof}
894: \begin{theorem}
895: Random-Turn Switching on the complete ternary tree of order $h$
896: lasts on average $\Theta(h)$ turns under optimal play.
897: \label{thm:ternary}
898: \end{theorem}
899: \begin{proof}
900: For the complete ternary tree of depth $h$ let $q_h$ be the
901: probability that Cut wins, let $\mu_h$ be the expected number of
902: explored (played) edges conditional on Cut winning, and let $\nu_h$
903: be the expected number of explored edges conditional on Short winning.
904: We have $q_0=0$, $\mu_0=0$ (say), and $\nu_0=0$.
905: When Cut wins, the top three edges are always explored.
906: Conditional on Cut winning, these top edges are open with probability
907: $q_{h-1}/(1+q_{h-1})$, independently. Consequently, the following recursions hold:
908: \begin{align*}
909: q_{h+1} &= \left(\frac{1+q_h}2\right)^3, \\
910: \mu_{h+1} &= 3 + 3\frac{q_h}{1+q_h} \mu_h\,, \\
911: % \intertext{since the top three edges are always explored, and conditional upon extinction, they are each open with probability $q_h/(1+q_h)$}
912: % \nu_{h+1} &= \frac{1}{1-q_h^3}\left[ (1-q_h^3) \mu_h + (1+q_h+q_h^2) + (q_h+q_h^2+q_h^3)\mu_h\right] \\
913: % &= \mu_h + \frac{1}{1-q_h} + \frac{q_h}{1-q_h}\mu_h.\\
914: \nu_{h+1} &= \frac{1-q_h}{1-q_h^3}\left[ 1+\nu_h\right] + \frac{q_h(1-q_h)}{1-q_h^3}\left[2+\mu_h+\nu_h\right] + \frac{q_h^2(1-q_h)}{1-q_h^3}\left[3+2\mu_h+\nu_h\right] \\
915: &= \nu_h + \frac{1+q_h+q_h^2-3 q_h^3}{1-q_h^3} + \frac{q_h+q_h^2-2 q_h^3}{1-q_h^3} \mu_h\,.
916: \end{align*}
917: We then have $q_h\to \sqrt 5-2$ ($<1/2$), $\mu_h=\Theta(1)$, $\nu_h=\Theta(h)$, and $\E[\text{\# turns}] = \Theta(h)$.
918: % \begin{align*}
919: % \lim_{h\to\infty} q_h &=\sqrt{5}-2 =: q\\
920: % \lim_{h\to\infty} \mu_h &= 3\frac{1+q}{1-2q} = 3+\frac{9}{\sqrt{5}} =: \mu\\
921: % \lim_{h\to\infty} \frac{\nu_h}h &= \frac{1+q \mu}{1-q} -3 q^3\frac{1+\mu}{1-q^3}\\
922: % \lim_{h\to\infty} \frac{\E[\text{\# turns}]}h &= (1-q)\lim_{h\to\infty} \frac{\nu_h}h = \frac{70-54/\sqrt{5}}{19} = \Theta(1). \qedhere
923: % \end{align*}
924: \end{proof}
925:
926: \subsection*{Recursive Majority.}
927:
928: Let $S_h=S$ be a subset of the leaves of the complete ternary tree of
929: depth $h$. Inductively, let $S_j$ be the set of nodes at level $j$
930: such that the majority of the nodes at level $j+1$ under them is in
931: $S_{j+1}$. The payoff function $f(S)$ for \textit{Recursive
932: Three-Fold Majority\/} is $-1$ if $S_0=\emptyset$ and $+1$ if
933: $S_0=\{\text{root}\}$. It seems that this random-turn game cannot be
934: analyzed using the methods we have used for the Random-Turn AND-OR
935: game or the Random-Turn Shannon Switching Game. In particular, we do not know
936: how long the game takes (on average) when played optimally.
937:
938:
939: \section{OPEN PROBLEMS.}
940:
941: We recall here some of the open problems raised earlier in the
942: article. It would be interesting to know the expected length of a
943: game of optimally played Random-Turn Hex and whether or not the true
944: optimal move (not just near-optimal) can be found efficiently. The
945: algorithm ``play Random-Turn Hex'' appears to be an efficient
946: algorithm (in terms of expected number of input bits examined) for
947: determining whether or not there is a percolation crossing, but is there
948: an asymptotically more efficient algorithm? It would also be interesting
949: to know, in optimally played Random-Turn Hex on large boards, whether
950: or not the fraction of disconnected moves is asymptotically zero.
951: Finally, we do not know how long Random-Turn Ternary Recursive
952: Majority takes, or whether or not it is the most efficient algorithm
953: (in terms of expected number of input bits read) for evaluating
954: Recursive Ternary Majority.
955:
956:
957: \subsection*{ACKNOWLEDGMENTS.}
958:
959: We are grateful to Wendelin Werner for suggesting to us that there
960: should be a version of Hex manifesting some conformal invariance in
961: the scaling limit. This was our original motivation for considering
962: the random-turn version of the game.
963:
964: Two of us (Peres and Sheffield) were supported in part by NSF grants \#DMS-0244479 and \#DMS-0104073.
965:
966: \begin{thebibliography}{10}
967:
968: \bibitem{hexinfo}
969: Hex information,
970: V. V. Anshelevich, ed., available at
971: \texttt{http://www.cs.unimaas.nl/ icga/games/hex/}.
972:
973: \bibitem{hexy}
974: V.~V. Anshelevich,
975: The game of Hex: An automatic theorem proving approach to game programming,
976: in \textit{Proceedings of the Seventeenth National Conference on Artificial Intelligence} (2000), pp.~189--194;
977: available at \texttt{http://home.earthlink.net/\char126vanshel/ VAnshelevich-01.pdf}.
978:
979: \bibitem{browne}
980: C.~Browne,
981: \textit{Hex Strategy: Making the Right Connections},
982: AK Peters, Natick, MA, 2000.
983:
984: \bibitem{cardy}
985: J.~L. Cardy,
986: Critical percolation in finite geometries,
987: \textit{J. Phys.\ A} \textbf{25} (1992) L201--L206.
988:
989: \bibitem{coniglio}
990: A.~Coniglio,
991: Fractal structure of {Ising} and {Potts} clusters: Exact results,
992: \textit{Phys.\ Rev.\ Lett.} \textbf{62} (1989) 3054--3057.
993:
994: \bibitem{gale}
995: D.~Gale,
996: The game of {Hex} and the {Brouwer} fixed-point theorem,
997: \textit{Amer.\ Math.\ Monthly}
998: % this \textsc{Monthly}
999: \textbf{86} (1979) 818--827.
1000:
1001: \bibitem{gardner}
1002: M.~Gardner,
1003: The game of {Hex},
1004: in \textit{Hexaflexagons and Other Mathematical Diversions: The First
1005: Scientific American Book of Puzzles and Games}, 1959,
1006: Simon and Schuster, New York, pp.~73--83.
1007:
1008: \bibitem{MR1732786}
1009: P.~Grassberger,
1010: Pair connectedness and shortest-path scaling in critical percolation,
1011: \textit{J. Phys.\ A} \textbf{32} (1999) 6233--6238.
1012:
1013: \bibitem{nash}
1014: H.~W. Kuhn and S.~Nasar, eds.,
1015: \textit{The Essential John Nash},
1016: Princeton University Press, Princeton, 2002.
1017:
1018: \bibitem{MR1230963}
1019: R.~Langlands, P.~Pouliot, and Y.~Saint-Aubin,
1020: Conformal invariance in two-dimensional percolation,
1021: \textit{Bull.\ Amer.\ Math.\ Soc.\ (N.S.)} \textbf{30} (1994) 1--61.
1022:
1023: \bibitem{MR1685133}
1024: A.~J. Lazarus, D.~E. Loeb, J.~G. Propp, W.~R. Stromquist, and D.~H. Ullman,
1025: Combinatorial games under auction play,
1026: \textit{Games Econom.\ Behav.} \textbf{27} (1999) 229--264.
1027:
1028: \bibitem{lehman}
1029: A.~Lehman,
1030: A solution of the {Shannon} switching game,
1031: \textit{J. Soc.\ Indust.\ Appl.\ Math.} \textbf{12} (1964) 687--725.
1032:
1033: \bibitem{mansfield}
1034: R.~Mansfield,
1035: Strategies for the {Shannon} switching game,
1036: \textit{Amer.\ Math.\ Monthly}
1037: % this \textsc{Monthly}
1038: \textbf{103} (1996) 250--252.
1039:
1040: \bibitem{odonnell-saks-schramm-servedio}
1041: R.~O'Donnell, M.~Saks, O.~Schramm, and R.~Servedio,
1042: Every decision tree has an influential variable,
1043: in \textit{Proceedings of the 46th Annual Symposium on Foundations of
1044: Computer Science (FOCS)}, 2005, pp.~31--39;
1045: available at arXiv:cs.CC/0508071.
1046:
1047: \bibitem{odonnell-servedio}
1048: R.~O'Donnell and R.~Servedio,
1049: On decision trees, influences, and learning monotone decision trees,
1050: Technical report CUCS-023-04, Department of
1051: Computer Science, Columbia University, 2004.
1052:
1053: \bibitem{pssw-tug}
1054: Y.~Peres, O.~Schramm, S.~Sheffield, and D.~B. Wilson,
1055: Tug-of-war and the infinity {Laplacian} (2006, preprint).
1056:
1057: \bibitem{reisch}
1058: S.~Reisch,
1059: Hex ist {PSPACE}-vollst\"andig,
1060: \textit{Acta Inform.} \textbf{15} (1981) 167--191.
1061:
1062: \bibitem{queenbee}
1063: J.~van Rijswijck,
1064: \textit{Computer Hex: Are Bees Better than Fruitflies?},
1065: Master's thesis, Department of Computer Science, University of Alberta, 2000.
1066:
1067: \bibitem{schramm-steif}
1068: O.~Schramm and J.~E. Steif,
1069: Quantitative noise sensitivity and exceptional times for percolation
1070: (2005, preprint); available at arXiv:math.PR/0504586.
1071:
1072: \bibitem{schramm-wilson}
1073: O.~Schramm and D.~B. Wilson,
1074: {SLE} coordinate changes,
1075: \textit{New York J. Math.} \textbf{11} (2005) 659--669;
1076: available at arXiv:math.PR/0505368.
1077:
1078: \bibitem{smirnov}
1079: S.~Smirnov,
1080: Critical percolation in the plane. {I.} {Conformal} invariance and
1081: {Cardy's} formula. {II.} {Continuum} scaling limit, 2001,
1082: available at \texttt{http://www.math.kth.se/\char126stas/ papers/percol.ps}.
1083:
1084: \bibitem{MR1879816}
1085: S.~Smirnov and W.~Werner,
1086: Critical exponents for two-dimensional percolation,
1087: \textit{Math.\ Res.\ Lett.} \textbf{8} (2001) 729--744.
1088:
1089: \bibitem{stewart}
1090: I.~Stewart,
1091: Hex marks the spot,
1092: \textit{Sci.\ Amer.} \textbf{283} (September 2000) 100--103.
1093:
1094: \bibitem{yang}
1095: J.~Yang,
1096: \texttt{http://www.ee.umanitoba.ca/\char126jingyang/}.
1097:
1098: \end{thebibliography}
1099:
1100:
1101:
1102: \hrule
1103: \vs \noindent
1104: \textbf{YUVAL PERES} works in probability theory and fractal geometry,
1105: especially on problems involving Brownian motion, probability on
1106: trees, percolation, and projections of fractals. After graduating
1107: from the Hebrew University, Jerusalem, in 1990, he has taught at
1108: Stanford, Yale, and the Hebrew University. He now teaches in the
1109: Statistics and Mathematics Departments at the University of
1110: California, Berkeley.
1111:
1112: \noindent\textit{Department of Statistics; 367 Evans Hall; University of California; Berkeley, CA 94720}%-3860}
1113:
1114: \noindent{\texttt{http://www.stat.berkeley.edu/\char126peres/}}
1115:
1116: \vs \noindent
1117: \textbf{ODED SCHRAMM} was born in December 1961 in Jerusalem, Israel. He studied
1118: at the Hebrew University (BSc and MSc) and completed his Ph.D. at
1119: Princeton University under the direction of William Thurston. After a
1120: two-year postdoc at U.C. San Diego, he joined the Weizmann Institute of
1121: Science and remained there until 1999, at which time he moved to
1122: Redmond, Washington, to join the Theory Group of Microsoft Research. Among
1123: the prizes he has received are the Salem Prize (2001), the Clay Research Award
1124: (2002), the Henri Poincar\'e Prize (2003) and the Loeve Prize (2003).
1125:
1126: \noindent\textit{Microsoft Research; One Microsoft Way; Redmond, WA 98052}
1127:
1128: \noindent{\texttt{http://research.microsoft.com/\char126schramm/}}
1129:
1130: \vs \noindent
1131: \textbf{SCOTT SHEFFIELD} earned his Ph.D. from Stanford in 2003 and his B.A.
1132: and M.A. degrees from Harvard in 1998. He completed postdoctoral
1133: research at Microsoft Research and U.C. Berkeley and is currently an
1134: assistant professor in the Courant Institute at New York University,
1135: where he teaches probability and mathematical finance. He has held
1136: numerous internships and short-term research positions in government
1137: and industry, and his primary research interests are mathematical
1138: physics and probability theory.
1139:
1140: \noindent\textit{Courant Institute; 251 Mercer Street; New York, NY 10012}
1141:
1142: \noindent{\texttt{http://www.cims.nyu.edu/\char126sheff/}}
1143:
1144: \vs\noindent \textbf{DAVID B. WILSON} works in probability theory,
1145: especially in statistical physics and random walks. He studied at MIT
1146: where he earned three S.B. degrees in 1991 and his Ph.D. in 1996 under
1147: the direction of Jim Propp. After postdocs at U.C. Berkeley, DIMACS
1148: at Rutgers, and the Institute for Advanced Study, he joined the Theory
1149: Group of Microsoft Research.
1150:
1151: \noindent\textit{Microsoft Research; One Microsoft Way; Redmond, WA 98052}
1152:
1153: \noindent{\texttt{http://dbwilson.com}}
1154:
1155: \end{document}
1156: