Tug-of-war and the infinity Laplacian

By Yuval Peres, Oded Schramm, Scott Sheffield, and David B. Wilson

Abstract

We prove that every bounded Lipschitz function on a subset of a length space admits a tautest extension to , i.e., a unique Lipschitz extension for which for all open . This was previously known only for bounded domains in , in which case is infinity harmonic; that is, a viscosity solution to , where

We also prove the first general uniqueness results for on bounded subsets of (when is uniformly continuous and bounded away from ) and analogous results for bounded length spaces. The proofs rely on a new game-theoretic description of . Let be the value of the following two-player zero-sum game, called tug-of-war: fix . At the turn, the players toss a coin and the winner chooses an with . The game ends when , and player I’s payoff is . We show that . Even for bounded domains in , the game theoretic description of infinity harmonic functions yields new intuition and estimates; for instance, we prove power law bounds for infinity harmonic functions in the unit disk with boundary values supported in a -neighborhood of a Cantor set on the unit circle.

1. Introduction and preliminaries

1.1. Overview

We consider a class of zero-sum two-player stochastic games called tug-of-war and use them to prove that every bounded real-valued Lipschitz function on a subset of a length space admits a unique absolutely minimal (AM) extension to , i.e., a unique Lipschitz extension for which for all open . We present an example that shows this is not generally true when is merely Lipschitz and positive. (Recall that a metric space is a length space if for all , the distance is the infimum of the lengths of continuous paths in that connect to . Length spaces are more general than geodesic spaces, where the infima need to be achieved.)

When is the closure of a bounded domain and is its boundary, a Lipschitz extension of is AM if and only it is infinity harmonic in the interior of ; i.e., it is a viscosity solution (defined below) to , where is the so-called infinity Laplacian

(informally, this is the second derivative of in the direction of the gradient of ). Aronsson proved this equivalence for smooth in 1967, and Jensen proved the general statement in 1993 Reference 1Reference 13. Our analysis of tug-of-war also shows that in this setting has a unique viscosity solution (extending ) when is continuous and or , but not necessarily when assumes values of both signs. We note that in the study of the homogenous equation , the normalizing factor in (Equation 1.1) is sometimes omitted; however, it is important to include it in the non-homogenous equation. Observe that with the normalization, coincides with the ordinary Laplacian in the one-dimensional case.

Unlike the ordinary Laplacian or the -Laplacian for , the infinity Laplacian can be defined on any length space with no additional structure (such as a measure or a canonical Markov semigroup)—that is, we will see that viscosity solutions to are well defined in this generality. We will establish the above stated uniqueness of solutions to in the setting of length spaces.

Originally, we were motivated not by the infinity Laplacian but by random turn Hex Reference 22 and its generalizations, which led us to consider the tug-of-war game. As we later learned, tug-of-war games have been considered by Lazarus, Loeb, Propp and Ullman in Reference 16 (see also Reference 15).

Tug-of-war on a metric space is very natural and conceivably applicable (like differential game theory) to economic and political modeling.

The intuition provided by thinking of strategies for tug-of-war yields new results even in the classical setting of domains in . For instance, in Section 4 we show that if is infinity harmonic in the unit disk and its boundary values are in and supported on a -neighborhood of the ternary Cantor set on the unit circle, then for some .

Before precisely stating our main results, we need several definitions.

1.2. Random turn games and values

We consider two-player, zero-sum random-turn games, which are defined by the following parameters: a set of states of the game, two directed transition graphs with vertex set , a non-empty set of terminal states (a.k.a. absorbing states), a terminal payoff function , a running payoff function , and an initial state .

The game play is as follows: a token is initially placed at position . At the step of the game, a fair coin is tossed, and the player who wins the toss may move the token to any for which is a directed edge in her transition graph. The game ends the first time , and player I’s payoff is . Player I seeks to maximize this payoff, and since the game is zero-sum, player II seeks to minimize it.

We will use the term tug-of-war (on the graph with edges ) to describe the game in which (i.e., players have identical move options) and is undirected (i.e., all moves are reversible). Generally, our results pertain only to the undirected setting. Occasionally, we will also mention some counterexamples showing that the corresponding results do not hold in the directed case.

In the most conventional version of tug-of-war on a graph, is a union of “target sets” and , there is no running payoff (), and is identically on and identically on . Players then try to “tug” the game token to their respective targets (and away from their opponent’s targets), and the game ends when a target is reached.

A strategy for a player is a way of choosing the player’s next move as a function of all previously played moves and all previous coin tosses. It is a map from the set of partially played games to moves (or in the case of a random strategy, a probability distribution on moves). Normally, one would think of a good strategy as being Markovian, i.e., as a map from the current state to the next move, but it is useful to allow more general strategies that take into account the history.

Given two strategies , let and be the expected total payoff (including the running payoffs received) at the termination of the game, if the game terminates with probability one and this expectation exists in ; otherwise, let and .

The value of the game for player I is defined as . The value for player II is . We use the expressions and to denote the values for players I and II, respectively, as a function of the starting state of the game.

Note that if player I cannot force the game to end almost surely, then , and if player II cannot force the game to end almost surely, then . Clearly, . When , we say that the game has a value, given by .

Our definition of value for player I penalizes player I severely for not forcing the game to terminate with probability one, awarding in this case.

(As an alternative definition, one could define , and hence player I’s value, by assigning payoffs to all of the non-terminating sequences . If the payoff function for the non-terminating games is a zero-sum Borel-measurable function of the infinite sequence, then player I’s value is equal to player II’s value in great generality Reference 17; see also Reference 20 for more on stochastic games. The existence of a value by our strong definition implies the existence and equality of the values defined by these alternative definitions.)

Considering the two possibilities for the first coin toss yields the following lemma, a variant of which appears in Reference 16.

Lemma 1.1.

The function satisfies the equation

for every non-terminal state for which the right-hand side is well defined, and when the right-hand side is of the form . The analogous statement holds for , except that when the right-hand side of Equation 1.2 is of the form .

When , the operator

is called the (discrete) infinity Laplacian. A function is infinity harmonic if Equation 1.2 holds and at all non-terminal . When is finite, this is equivalent to . However, it will be convenient to adopt the convention that is infinity harmonic at if (resp. ) and the right-hand side in Equation 1.2 is also (resp. ). Similarly, it will be convenient to say is a solution to at if (resp. ) and the right-hand side in Equation 1.2 is also (resp. ).

In a tug-of-war game, it is natural to guess that the value exists and is the unique solution to

and also that (at least when is locally finite) player I’s optimal strategy will be to always move to the vertex that maximizes and that player II’s optimal strategy will be to always move to the vertex that minimizes . This is easy to prove when is undirected and finite and is everywhere positive or everywhere negative. Subtleties arise in more general cases ( infinite, directed, unbounded, having values of both signs, etc.).

Our first theorem addresses the question of the existence of a value.

Theorem 1.2.

A tug-of-war game with parameters has a value whenever the following hold:

(1)

Either everywhere or .

(2)

.

(3)

is undirected.

Counterexamples exist when any one of the three criteria is removed. In Section 5 (a section devoted to counterexamples) we give an example of a tug-of-war game without a value, where is undirected, , and the running payoff satisfies but .

The case where , is bounded, and is locally finite was proved earlier in Reference 15. That paper discusses an (essentially non-random) game in which the two players bid for the right to choose the next move. That game, called the Richman game, has the same value as tug-of-war with , where takes the values and . Additionally, a simple and efficient algorithm for calculating the value when and is finite is presented there.

1.3. Tug-of-war on a metric space

Now consider the special case where is a metric space, , and Lipschitz functions and are given. Let be the edge-set in which if and only if , and let be the value (if it exists) of the game played on with terminal payoff and running payoff normalized to be .

In other words, is the value of the following two-player zero-sum game, called -tug-of-war: fix . At the turn, the players toss a coin and the winner chooses an with . The game ends when , and player I’s payoff is .

When the limit exists pointwise, we call the continuum value (or just “value”) of the quintuple . We define the continuum value for player I (or II) analogously.

The reader may wonder why we have chosen not to put an edge in between and when exactly. This choice has some technical implications. Specifically, we will compare the -game with the -game. If are such that , then in a length space it does not follow that there is a such that and . However, it does follow if you replace the weak inequalities with strong inequalities throughout.

We prove the following:

Theorem 1.3.

Suppose is a length space, is non-empty, is bounded below and has an extension to a uniformly continuous function on , and either satisfies or all three of the following hold: , is uniformly continuous, and has finite diameter. Then the continuum value exists and is a uniformly continuous function extending . Furthermore, as . If is Lipschitz, then so is . If and are Lipschitz, then .

The above condition that extends to a uniformly continuous function on is equivalent to having uniformly continuous on and “Lipschitz on large scales,” as we prove in Lemma 3.9 below.

We will see in Section 5.1 that this fails in general when but . When assumes values of both signs, it fails even when is a closed disk in , is its boundary and . In Section 5.3 we show by means of an example that in such circumstances it may happen that and moreover, .

1.4. Absolutely minimal Lipschitz extensions

Given a metric space , a subset and a function , we write

and . Thus is Lipschitz iff . Given , we say that is a minimal extension of if and for all .

It is well known that for any metric space , any Lipschitz on a subset of admits a minimal extension. The largest and smallest minimal extensions (introduced by McShane Reference 18 and Whitney Reference 24 in the 1930’s) are respectively

We say is an absolutely minimal (AM) extension of if and for every open set . We say that is AM on if it is defined on and is an AM extension of its restriction to . AM extensions were first introduced by Aronsson in 1967 Reference 1 and have applications in engineering and image processing (see Reference 4 for a recent survey).

We prove the following:

Theorem 1.4.

Let be a length space and let be Lipschitz, where . If , then the continuum value function described in Theorem 1.3 (with ) is an AM extension of . If is also bounded, then is the unique AM extension of .

We present in the counterexample section, Section 5, an example in which is Lipschitz, non-negative, and unbounded, and although the continuum value is an AM extension, it is not the only AM extension.

Prior to our work, the existence of AM extensions in the above settings was known only for separable length spaces Reference 14 (see also Reference 19). The uniqueness in Theorem 1.4 was known only in the case that is the closure of a bounded domain and . (To deduce this case from Theorem 1.4, one needs to replace by the smallest closed ball containing , say.) Three uniqueness proofs in this setting have been published, by Jensen Reference 13, by Barles and Busca Reference 6, and by Aronsson, Crandall, and Juutinen Reference 4. The third proof generalizes from the Euclidean norm to uniformly convex norms.

Our proof applies to more general spaces because it invokes no outside theorems from analysis (which assume existence of a local Euclidean geometry, a measure, a notion of twice differentiability, etc.), and relies only on the structure of as a length space.

As noted in Reference 4, AM extensions do not generally exist on metric spaces that are not length spaces. (For example, if is the -shaped region with the Euclidean metric, and , then no non-constant has an AM extension. Indeed, suppose that is an AM extension of . Let , and . Then, considering , it follows that . Likewise, . Now taking , we see that . Hence . By symmetry, . Since is assumed to be non-constant, , and hence . Then , which contradicts .)

One property that makes length spaces special is the fact that the Lipschitz norm is determined locally. More precisely, if is closed, then either or for every

The definition of AM was inspired by the notion that if is the “tautest possible” Lipschitz extension of , it should be tautest possible on any open , given the values of on and ignoring the rest of the metric space. Without locality, the rest of the metric space cannot be ignored (since long-distance effects may change the global Lipschitz constant), and the definition of AM is less natural. Another important property of length spaces is the fact that the graph distance metric on scaled by approximates the original metric, namely, it is within of .

1.5. Infinity Laplacian on

The continuum version of the infinity Laplacian is defined for functions on domains by

This is the same as , where is the Hessian of and . Informally, is the second derivative of in the direction of the gradient of . If , then is undefined; however, we adopt the convention that if the second derivative of happens to be the same in every direction (i.e., the matrix is times the identity), then , which is the second derivative in any direction. (As mentioned above, some texts on infinity harmonic functions define without the normalizing factor . When discussing viscosity solutions to , the two definitions are equivalent. The fact that the normalized version is sometimes undefined when does not matter because it is always well defined at when is a cone function, i.e., when has the form for and with , and viscosity solutions can be defined via comparison with cones; see Section 1.6.) As in the discrete setting, is infinity harmonic if .

While discrete infinity harmonic functions are a recent concept, introduced in finite-difference schemes for approximating continuous infinity harmonic functions Reference 21, related notions of value for stochastic games are of course much older. The continuous infinity Laplacian first appeared in the work of Aronsson Reference 1 and has been very thoroughly studied Reference 4. Key motivations for studying this operator are the following:

(1)

AM extensions: Aronsson proved that extensions on domains (of functions on ) are infinity harmonic if and only if they are AM.

(2)

-harmonic functions: As noted by Aronsson Reference 1, the infinity Laplacian is the formal limit, as of the (properly normalized) -Laplacians. Recall that -harmonic functions, i.e., minimizers of subject to boundary conditions, solve the Euler-Lagrange equation

which can be rewritten

where is the ordinary Laplacian. Dividing by , we see that (at least when ) -harmonic functions satisfy , where ; the second term vanishes in the large limit. It is not too hard to see that as tends to infinity, the Lipschitz norm of any limit of the -harmonic functions extending will be . So it is natural to guess (and was proved in Reference 7) that as tends to infinity the -harmonic extensions of converge to a limit that is both absolutely minimal and a viscosity solution to .

In the above setting, Aronsson also proved that there always exists an AM extension, and that in the planar case , there exists at most one infinity harmonic extension; however infinity harmonic extensions do not always exist Reference 2.

To define the infinity Laplacian in the non- setting requires us to consider weak solutions; the right notion here is that of viscosity solution, as introduced by Crandall and Lions (1983) Reference 11. Start by observing that if and are functions, , and in a neighborhood of , then has a local minimum at , whence (if both sides of this inequality are defined). This comparison principle (which has analogs for more general degenerate elliptic PDEs Reference 5) suggests that if is not , in order to define we want to compare it to functions for which is defined. Let be the set of real valued functions defined and in a neighborhood of for which has been defined; that is, either , or and the limit exists.

Definition.

Let be a domain in and let be continuous. Set

Thus satisfies in a domain , iff every such that has a local minimum at some satisfies . In this case is called a viscosity subsolution of . Note that if , then wherever .

Similarly, let

and call a viscosity supersolution of iff in .

Finally, is a viscosity solution of if in (i.e., is both a supersolution and a subsolution).

Here is a little caveat. At present, we do not know how to show that in the viscosity sense determines . For example, if is Lipschitz, and are continuous, and holds for (in the viscosity sense), how does one prove that ?

The following result of Jensen (alluded to above) is now well known Reference 1Reference 13Reference 4: if is a domain in and is continuous, then for every bounded open set (i.e., is AM) if and only if is a viscosity solution to in .

Let , where is closed, and is a length space. If , one can define the -harmonic measure of from as the infimum of over all functions that are Lipschitz on , AM in and satisfy on . This quantity will be denoted by . In Section 4 we prove

Theorem 1.5.

Let be the unit ball in , , let , let be the center of the ball, and for each let be a spherical cap of radius (of dimension ). Then

where are absolute constants (which do not depend on ).

Numerical calculations Reference 21 had suggested that in the setting of the theorem tends to as , but this was only recently proved Reference 12, and the proof did not yield any quantitative information on the rate of decay. In contrast to our other theorems in the paper, the proof of this theorem does not use tug-of-war. The primary tool is the comparison of a specific AM function in with decay discovered by Aronsson Reference 3.

1.6. Quadratic comparison on length spaces

To motivate the next definition, observe that for continuous functions , the inequality reduces to convexity. The definition of convexity requiring a function to lie below its chords has an analog, comparison with cones, which characterizes infinity harmonic functions. Call the function a cone based at . For an open , say that a continuous satisfies comparison with cones from above on if for every open for every , and for every cone based at such that the inequality holds on , the same inequality is valid throughout . Comparison with cones from below is defined similarly using the inequality .

Jensen Reference 13 proved that viscosity solutions to for domains in satisfy comparison with cones (from above and below), and Crandall, Evans, and Gariepy Reference 9 proved that a function on is absolutely minimal in a bounded domain if and only if it satisfies comparison with cones in .

Champion and De Pascale Reference 8 adapted this definition to length spaces, where cones are replaced by functions of the form , where . Their precise definition is as follows. Let be an open subset of a length-space and let be continuous. Then is said to satisfy comparison with distance functions from above on if for every open , for every , for every and for every , if holds on , then it also holds in . The function is said to satisfy comparison with distance functions from below if satisfies comparison with distance functions from above. Finally, satisfies comparison with distance functions if it satisfies comparison with distance functions from above and from below.

The following result from Reference 8 will be used in the proof of Theorem 1.4:

Lemma 1.6 (Reference 8).

Let be an open subset of a length space. A continuous satisfies comparison with distance functions in if and only if it is AM in .

To study the inhomogenous equation , because will in general have a non-zero second derivative (in its gradient direction), it is natural to extend these definitions to comparison with functions that have a quadratic term.

Definitions.

Let with and let be a length space.

Let . We call the function a quadratic distance function (centered at ).

We say that a quadratic distance function is -increasing (in distance from ) on an open set if either (1) and for every , we have , or (2) and and . Similarly, we say that a quadratic distance function is -decreasing on if is -increasing on .

If is a continuous function defined on an open set in a length space , we say that satisfies -quadratic comparison on if the following two conditions hold:

(1)

-quadratic comparison from above: For every open and -increasing quadratic distance function on with quadratic term , the inequality on implies on .

(2)

-quadratic comparison from below: For every open and -decreasing quadratic distance function on with quadratic term , the inequality on implies on .

The following theorem is proved in Section 6.

Theorem 1.7.

Let be a real-valued continuous function on a bounded domain in , and suppose that is a continuous function on . Then satisfies -quadratic comparison on if and only if is a viscosity solution to in .

This equivalence motivates the study of functions satisfying quadratic comparison. Note that satisfying in the viscosity sense depends only on the local behavior of near . We may use Theorem 1.7 to extend the definition of to length spaces, saying that on an open subset of a length space if and only if every has a neighborhood on which satisfies -quadratic comparison. We warn the reader, however, that even for length spaces contained within , there can be solutions to that do not satisfy comparison with distance functions (or -quadratic comparison, for that matter): for example, and . The point here is that when we take and compare with some function on , the “appropriate” notion of the boundary is the boundary in , not in .

The continuum value of the tug-of-war game sometimes gives a construction of a function satisfying -quadratic comparison. We prove:

Theorem 1.8.

Suppose is a length space, is non-empty, is uniformly continuous and bounded, and either satisfies or all three of the following hold: , is uniformly continuous, and has finite diameter. Then the continuum value is the unique continuous function satisfying -quadratic comparison on and on . Moreover, if is continuous, satisfies on and -quadratic comparison from below on , then 1 throughout .

Putting these last two theorems together, we obtain

Corollary 1.9.

Suppose is a bounded open set, is uniformly continuous, and satisfies either or and is uniformly continuous. Then there is a unique continuous function that is a viscosity solution to on and satisfies on . This unique solution is the continuum value of tug-of-war on .

It is easy to verify that indeed satisfies the assumptions in Theorem 1.8. In order to deduce the corollary, we may take the length space as a ball in which contains and extend to , say. Alternatively, we may consider with its intrinsic metric and lift to the completion of .

We present in Section 5 an example showing that the corollary may fail if is permitted to take values of both signs. Specifically, the example describes two functions defined in the closed unit disk in and having boundary values identically zero on the unit circle such that with some Lipschitz function we have for (in the viscosity sense), while .

The plan of the paper is as follows. Section 2 discusses the discrete tug-of-war on graphs and proves Theorem 1.2, and Section 3 deals with tug-of-war on length spaces and proves Theorems 1.3, 1.4 and 1.8. Section 4 is devoted to estimates of -harmonic measure. In Section 5, we present a few counterexamples showing that some of the assumptions in the theorems we prove are necessary. Section 6 is devoted to the proof of Theorem 1.7. Section 7 presents some heuristic arguments describing what the limiting trajectories of some -tug-of-war games on domains in may look like and states a question regarding the length of the game. We conclude with additional open problems in Section 8.

2. Discrete game value existence

2.1. Tug-of-war on graphs without running payoffs

In this section, we will generally assume that is undirected and connected and that .

Though we will not use this fact, it is interesting to point out that in the case where is finite, there is a simple algorithm from Reference 15 which calculates the value of the game and proceeds as follows. Assuming the value is already calculated at some set of vertices, find a path , , with interior vertices and endpoints which maximizes and set for . Repeat this as long as .

Recall that is the value function for player I.

Lemma 2.1.

Suppose that and . Then is the smallest -harmonic function bounded from below on that extends . More generally, if is an -harmonic function which is bounded from below on and on , then on .

Similarly, if is bounded from above on , then is the largest -harmonic function bounded from above on that extends .

Proof.

Player I could always try to move closer to some specific point . Since in almost every infinite sequence of fair coin tosses there will be a time when the number of tails exceeds the number of heads by , this ensures that the game terminates a.s., and we have . Suppose that on and is -harmonic on . Given , consider an arbitrary strategy for player I and let II play a strategy that at step (if II wins the coin toss) II selects a state where is within of its infimum among the states to which II could move. We will show that the expected payoff for player I is at most . We may assume the game terminates a.s. at a time . Let denote the random sequence of states encountered in the game. Since is -harmonic, the sequence is a supermartingale. Optional sampling and Fatou’s lemma imply that . Thus . By Lemma 1.1, this completes the proof.

Now, we prove the first part of Theorem 1.2. When the graph is locally finite and is finite, this was proven in Reference 15, Thm. 17.

Theorem 2.2.

Suppose that is connected, and . If is bounded below (or above) on and , then , so the game has a value.

Before we prove this, we discuss several counterexamples that occur when the conditions of the theorem are not met. First, this theorem can fail if is directed. A trivial counterexample is when is finite and there is no directed path from the initial state to .

If is infinite and is directed, then there are counterexamples to Theorem 2.2, even when every vertex lies on a directed path towards a terminal state. For example, suppose and , with . If consists of directed edges of the form and , then II may play so that with positive probability the game never terminates, and hence the value for player I is by definition .

Even in the undirected case, a game may not have a value if is not bounded either from above or below. The reader may check that if is the integer lattice and the terminal states are the -axis with , then the players’ value functions are given by and . (Roughly speaking, this is because, in order to force the game to end in a finite amount of time, player I has to “give up” opportunities to move to the right.) Observe also that in this case any linear function which agrees with on the -axis is -harmonic.

As a final remark before proving this theorem, let us consider the obvious strategy for the two players, namely for player I to always maximize on her turn and for II to always minimize on her turn. Even when is (undirected and) locally finite and the payoff function satisfies , these obvious strategies need not be optimal. Consider, e.g., the game shown in Figure 1, where is given by , , and consists of edges of the form and . The terminal states are on the left and right edges of the square, and the payoff is on the left and on the right. Clearly the function (the Euclidean distance from the right edge of the square) is infinity harmonic, and by Corollary 2.3 below, we have . The obvious strategy for player I is to always move left, and for II it is to always move right. Suppose however that player I always pulls left and player II always pulls up. It is easy to check that the probability that the game ever terminates when starting from is at most (this function is a supermartingale under the corresponding Markov chain). Therefore the game continues forever with positive probability, resulting in a payoff of to player I. Thus, a near-optimal strategy for player I must be able to force the game to end, and must be prepared to make moves which do not maximize . (This is a well-known phenomenon, not particular to tug-of-war.)

Proof of Theorem 2.2.

If is bounded above but not below, then we may exchange the roles of players  and and negate to reduce to the case where is bounded from below. Since always holds, we just need to show that . Since player I could always pull towards a point in and thereby ensure that the game terminates, we have . Let and write . Let be the sequence of positions of the game. For ease of exposition, we begin by assuming that is locally finite (so that the suprema and infima in the definition of the -Laplacian definition are achieved) and that ; later we will remove these assumptions.

To motivate the following argument, we make a few observations. In order to prove that , we need to show that player II can guarantee that the game terminates while also making sure that the expected payoff is not much larger than . These are two different goals, and it is not a priori clear how to combine them. To resolve this difficulty, observe that is non-decreasing in if players I and II adopt the strategies of maximizing (respectively, minimizing) at every step. As we will later see, this implies that the game terminates a.s. under these strategies. On the other hand, if player I deviates from this strategy and thereby reduces , then perhaps player II can spend some turns playing suboptimally with respect to in order to increase . Let . For let and , which is the last position in up to time . We will shortly describe a strategy for II based on the idea of backtracking to when not in . If , we may define a backtracking move from as any move to a neighbor of that is closer to than in the subgraph spanned by the vertices . Here, “closer” refers to the graph metric of . When II plays the backtracking strategy, she backtracks whenever not in and plays to a neighbor minimizing when in .

Now consider the game evolving under any strategy for player I and the backtracking strategy for II. Let be the distance from to in the subgraph . Set

It is clear that , because there is a path of length from to in and the change in across any edge in this path is less than , by the definition of . It is easy to verify that is a supermartingale, as follows. If , and player I plays, then , while if II gets the turn, then . If and II plays, then . If and player I does not play into , then . The last case to consider is that and player I plays into a vertex in . In such a situation,

Thus, indeed, is a supermartingale (bounded from below). Let denote the first time a terminal state is reached (so if the game does not terminate). By the martingale convergence theorem, the limit exists. But when player II plays we have . Therefore, the game must terminate with probability . The expected outcome of the game thus played is at most . Consequently, , which completes the proof in the case where is locally finite and .

Next, what if is not locally finite, so that suprema and infima might not be achieved? In this case, we fix a small , and use the same strategy as above, except that if and II gets the turn, she moves to a neighbor at which is at most larger than its infimum value among neighbors of . In this case, is a supermartingale, and hence the expected payoff is at most . Since this can be done for any , we again have that .

Finally, suppose that . Let , and let player II pull toward until the first time a vertex with or is reached. After that, II continues as above. Since , this completes the proof.

Corollary 2.3.

If is connected, and , then is the unique bounded -harmonic function agreeing with on .

Proof.

This is an immediate consequence of Lemma 2.1, the remark that follows it, and Theorem 2.2.

If , and , then is an example of an (unbounded) -harmonic function that is different from .

2.2. Tug-of-war on graphs with running payoffs

Suppose now that . Then the analog of Theorem 2.2 does not hold without additional assumptions. For a simple counterexample, suppose that is a triangle with self-loops at its vertices (i.e., a player may opt to remain in the same position), that the vertex is a terminal vertex with final payoff , and the running payoff is given by and . Then the function given by , , is a solution to , provided . The reader may check that is the smallest of these functions and is the largest. The gap of between and appears because a player would have to give up a move (sacrificing one) in order to force the game to end. This is analogous to the example given in Section 2.1. Both players are earning payoffs in the interior of the game, and moving to a terminal vertex costs a player a turn.

One way around this is to assume that is either uniformly positive or uniformly negative, as in the following analog of Theorem 2.2. We now prove the second half of Theorem 1.2:

Theorem 2.4.

Suppose that is connected and . Assume that is bounded from below and . Then . If, additionally, and are bounded from above, then any bounded solution to with the given boundary conditions is equal to .

Proof.

By considering a strategy for player I that always pulls toward a specific terminal state , we see that . By Lemma 1.1, on . Let be any solution to on that is bounded from below on and has the given boundary values on .

Claim.

. In proving this, we may assume without loss of generality that is connected, where is the graph . Then if at some vertex in , we also have throughout , in which case the Claim is obvious. Thus, assume that is finite on . Fix and let II use the strategy that at step , if the current state is and II wins the coin toss, selects a state with Then for any strategy chosen by player I, the sequence is a supermartingale bounded from below, which must converge a.s. to a finite limit. Since , this also forces the game to terminate a.s. Let denote the termination time. Then

Thus this strategy for II shows that . Since is arbitrary, this verifies the claim. In particular, in , so .

Now suppose that and , and is a bounded solution to with the given boundary values. By the claim above, . On the other hand, player I can play to maximize or nearly maximize in every move. Under such a strategy, she guarantees that by turn the expected payoff is at least , where is if the game has terminated by time , and otherwise. If the expected number of moves played is infinite, the expected payoff is infinite. Otherwise, , since is bounded. Thus, in any case.

3. Continuum value of tug-of-war on a length space

3.1. Preliminaries and outline

In this section, we will prove Theorem 1.3, Theorem 1.8 and Theorem 1.4. Throughout this section, we assume denotes a tug-of-war game, i.e., is a length space with distance function , is a non-empty set of terminal states, is the final payoff function, and is the running payoff function. We let denote the game state at time in -tug-of-war.

It is natural to ask for a continuous-time version of tug-of-war on a length space. Precisely and rigorously defining such a game (which would presumably involve replacing coin tosses with white noise, making sense of what a continuum no-look-ahead strategy means, etc.) is a technical challenge we will not undertake in this paper (though we include some discussion of the small- limiting trajectory of -tug-of-war in the finite-dimensional Euclidean case in Section 7). But we can make sense of the continuum game’s value function by showing that the value function for the -step tug-of-war game converges as .

The value functions do not satisfy any nice monotonicity properties as . In the next subsection we define two modified versions of tug-of-war whose values closely approximate , and which do satisfy a monotonicity property along sequences of the form , allowing us to conclude that exists. Then we show that any such limit is a bounded from below viscosity solution to , and that any viscosity solution bounded from below is an upper bound on such a limit, so that any two such limits must be equal, which will allow us to prove that the continuum limit exists.

Because the players can move the game state almost as far as , either player can ensure that is “almost a supermartingale” up until the time that . When doing calculations it is more convenient to instead work with a related metric defined by

Since is the graph distance scaled by , it is in fact a metric, and either player may choose to make a supermartingale up until the time .

3.2. II-favored tug-of-war and dyadic limits

We define a game called II-favored -tug-of-war that is designed to give a lower bound on player I’s expected payoff. It is related to ordinary -step tug-of-war, but II is given additional options, and player I’s running payoffs are slightly smaller. At the  step, player I chooses a point in and a coin is tossed. If player I wins the coin toss, the game position moves to a point, of player II’s choice, in . (If , this means simply moving to .) If II wins, then the game position moves to a point in of II’s choice. The game ends at the first time for which . Player I’s payoff is then if the game never terminates, and otherwise it is

where is the point that player I targets on the  turn, and is defined to be zero if . We let be the value for player I for this game. Given a strategy for player II in the ordinary -game, player II can easily mimic this strategy in the II-favored -game and do at least as well, so .

Let be the value for player II of I-favored -tug-of-war, defined analogously but with the roles of player I and player II reversed (i.e., at each move, II selects the target less than units away, instead of player I, etc., and the in the running payoff term in equation Equation 3.1 is replaced with a , and games that never terminate have payoff ). For any we have

Lemma 3.1.

For any ,

Proof.

We have already noted that . We will prove that ; the inequality follows by symmetry. Consider a strategy for a II-favored -tug-of-war. We define a strategy for player I for the II-favored -game that mimics as follows. Whenever strategy would choose a target point , player I “aims” for for one “round,” which we define to be the time until one of the players has won the coin toss two more times than the other player. By “aiming for we mean that player I picks a target point that, in the metric , is units closer to than the current point. With probability , player I gets two surplus moves before II, and then the game position reaches (or a point in ) before the game position exits . If player II gets two surplus moves before player I, then the game position will be in . (See Figure 2.)

The expected number of moves in this round of the II-favored -game is ; and the running payoff at each move is an times the infimum over a ball of radius that is a subset of (as opposed to times the infimum over the whole ball). Hence strategy guarantees for the II-favored -game an expected total payoff that is at least as large as what guarantees for the II-favored -game.

Thus converges along dyadic sequences: exists. A priori the subsequential limit could depend upon the choice of the dyadic sequence, i.e., the initial .

The same argument can be used to show that for positive integers .

3.3. Comparing favored and ordinary tug-of-war

We continue with a preliminary bound on how far apart and can be. Let denote the Lipschitz constant of with respect to the restriction of the metric to . Since , the Lipschitz constant of with respect to upper bounds .

Lemma 3.2.

Let . Suppose that and either

(1)

everywhere,  or

(2)

is bounded above and has finite diameter.

Then for each and ,

Such an expected payoff is guaranteed for player I if she adopts a pull towards strategy, which at each move attempts to reduce . Similarly a pull towards strategy for II gives

Proof.

Let player I use a pull towards strategy. Let be the time at which is reached, which will be finite a.s. The distance is a supermartingale, except possibly at the last step, where player II may have moved the game state to a terminal point up to a distance of from the target, even if player I wins the coin toss. Thus , whence

If , then this implies . If and has finite diameter, then the expected number of steps before the game terminates is at most the expected time that a simple random walk on the interval of integers (with a self-loop added at the right endpoint) takes to reach when started at , i.e., at most . Hence

This gives the desired lower bound on . The symmetric argument gives the upper bound for .

Next, we show that the lower bound on is a good lower bound.

Lemma 3.3.

Suppose is Lipschitz, and either

(1)

everywhere, or

(2)

is uniformly continuous, has finite diameter and .

Then as . If is also Lipschitz, then .

Note that since is assumed to be a length space, the assumptions imply that is constant and .

Proof.

In order to prove that is not much larger than , consider a strategy for player I in ordinary -tug-of-war, which achieves an expected payoff of at least against any strategy for player II. We shortly describe a modified strategy for player I playing the II-favored game, which does almost as well as does in the ordinary game. To motivate , observe that a turn in the II-favored -tug-of-war can alternatively be described as follows. Suppose that the position at the end of the previous turn is . First, player I gets to make a move to an arbitrary point satisfying . Then a coin is tossed. If player II wins the toss, she gets to make two steps from , each of distance less than . Otherwise, II gets to move to an arbitrary point in , but only if the latter set is non-empty. This completes the turn. The strategy is based on the idea that after a win in the coin toss by II, player I may use her move to reverse one of the two steps executed by II (provided has not been reached).

As strategy is playing the II-favored game against player II, it keeps track of a virtual ordinary game. At the outset, the II-favored game is in state , as is the virtual game. As long as the virtual and the favored game have not ended, each turn in the favored game corresponds to a turn in the virtual game, and the virtual game uses the same coin tosses as the favored game. In each such turn , the target for player I in the favored game is the current state in the ordinary game. If player I wins the coin toss, then the new game state in the favored game is his current target , which is the state of the virtual game , and the new state of the virtual game is chosen according to strategy applied to the history of the virtual game. If player II wins the toss and chooses the new state of the favored game to be , where necessarily , then in the virtual game the virtual player II chooses the new state as some point satisfying and . Induction shows that as long as both games are running, and thus the described moves are all legal.

If at some time the virtual game has terminated, but the favored game has not, we let player I continue playing the favored game by always pulling towards the final state of the virtual game. If the favored game has terminated, for the sake of comparison, we continue the virtual game, but this time let player II pull towards the final state of the favored game and let player I continue using strategy .

Let be the time at which the favored game has ended, and let be the time at which the ordinary virtual game has ended. By Lemma 3.2, if , then the conditioned expectation of the remaining running payoffs and final payoff to player I in the favored game after time , given what happened up to time , is at least (here the implicit constant may depend on , and ). Likewise, from Lemma 3.1 and the second inequality from Lemma 3.2 show that if , then the conditioned expectation of the remaining running and final payoffs in the virtual game is at most .

Set . Then if is Lipschitz and if is uniformly continuous. At each time , since , the running payoff in the virtual game and in the favored game differ by at most . Lemmas 3.1 and 3.2 show that there is a constant , which may depend on and , but not on , such that . Thus, in the case where , since guarantees a payoff of at least , we have . Assume that player II plays the II-favored game (up to time ) using a strategy such that the expected payoff to player I who uses is at most . Then, in the case where , we will have , again. There is a strategy for player II in the ordinary game, which corresponds to the play of player II in the virtual game, when player I uses and and player II uses in the favored game. (The description of the virtual game defines for some game histories, and we may take an arbitrary extension of this partial strategy to all possible game histories.) The above shows that the expected payoff for player I in the ordinary game when player I uses and player II uses differs from the expected payoff for player I in the favored game when player I uses and player II uses by at most . Thus . Since , the proof is now complete.

Hence under the assumption of Lemma 3.3, .

3.4. Dyadic limits satisfy quadratic comparison

We start by showing that almost satisfies -quadratic comparison from above.

Lemma 3.4.

Let , let be an open subset of and write . Suppose that is a quadratic distance function that is -increasing on , where satisfies

Also suppose that or . If the value function for player I in -tug-of-war satisfies on , then on .

Proof.

Fix some . Consider the strategy for player II that from a state at distance from pulls to state (if ) or else moves to reduce the distance to by “almost” units, enough to ensure that . If , then , whence with . In this case, if II wins the toss, then , whence . Thus for all , regardless of what strategy player I adopts,

Setting , we conclude that is a supermartingale.

Suppose that player I uses a strategy with expected payoff larger than . (If there is no such strategy, the assertion of the lemma is obvious.) Then a.s. We claim that

Clearly this holds if is replaced by . To pass to the limit as , consider two cases:

If , then since is -increasing on , it is also bounded from below on . Consequently, is a supermartingale bounded from below, so Equation 3.3 holds.

If , then , by Equation 3.2. By assumption therefore , which implies . If , we get , and hence Equation 3.3 holds. On the other hand, if , then dominated convergence gives Equation 3.3.

Since , we deduce that

where runs over all possible strategies for player I with expected payoff larger than . Since was arbitrary, the proof is now complete.

In order for to satisfy -quadratic comparison (from above), we would like to know that if on the boundary of an open set, then this (almost) holds in a neighborhood of the boundary, so that we can apply the above lemma. To do this we prove a uniform Lipschitz lemma:

Lemma 3.5 (Uniform Lipschitz).

Suppose that is Lipschitz, and either

(1)

everywhere,  or

(2)

is bounded from above and has finite diameter.

Then for each , and are Lipschitz on w.r.t. the metric with the Lipschitz constant depending only on and .

Proof.

By symmetry, it suffices to prove this for . Set

Let be distinct. If , then . If and , Lemmas 3.2 and 3.1 give

Now suppose . Set , on and . Then, clearly, the value of for the game where is replaced by and is replaced by is the same as for the original game. By Equation 3.4, we have . Consequently, Equation 3.4 gives

which completes the proof.

Lemma 3.6.

Suppose is Lipschitz and , and either (1) identically or else (2) , is uniformly continuous, and has finite diameter. Then the subsequential limit satisfies -quadratic comparison on .

Proof.

By Theorems 2.2 and 2.4, , so from Lemma 3.3 we have as . Thus .

Note that the hypotheses imply that is bounded. Consider an open and an -increasing quadratic distance function on with quadratic term , such that on . We must show that on . Since , we have by Lemma 3.1 that converges uniformly to . So for any , if is large enough, then on . Note also that is necessarily uniformly continuous on . (If or , this is clear. Otherwise, and . However, implies that , since is -increasing on .) Hence, by the uniform Lipschitz lemma, Lemma 3.5, on for all sufficiently large , where we use the notation of Lemma 3.4. By that lemma on all of . Letting and shows that satisfies -quadratic comparison from above.

To prove quadratic comparison from below, note that the only assumptions which are not symmetric under exchanging the roles of the players are and . However, we only used these assumptions to prove . Consequently, comparison from below follows by symmetry.

3.5. Convergence

Lemma 3.7.

Suppose that is continuous and satisfies -quadratic comparison from below on , and that is locally bounded from below. Let . Then in II-favored -tug-of-war, when player I (using any strategy) targets point on step , player II may play to make a supermartingale, where

and .

Proof.

Let be the point that player I has targeted at time . Assume that . We define the following:

(1)

,

(2)

(the infimum value of that II can guarantee if II wins the coin toss),

(3)

, and

(4)

(, , , and ).

Player II can play so that , i.e., so that . We will show that whenever we have , and then it will follow that is a supermartingale. There are two cases to check, depending on whether or :

Suppose . Note that . If , the inequality on follows from and the definition of . Assume therefore that . Then . Thus is decreasing on . Let be the point where attains its minimum in . Then is decreasing on . Set . We have , and for we have . Thus, for . Since satisfies -quadratic comparison from below, and , we get for . In particular, for one has .

Now suppose . The function is a lower bound for on , and hence applying -quadratic comparison in with gives . Therefore, , which together with implies that is decreasing on . By applying -quadratic comparison on we see that for we have .

Lemma 3.8.

Suppose that is Lipschitz, and either

(1)

everywhere,  or

(2)

is bounded from above and has finite diameter.

Also suppose that is continuous, satisfies -quadratic comparison from below on , on , and . Then for all .

Proof.

The idea is for player II to make the defined in Lemma 3.7 a supermartingale, but we need to pick a stopping time such that , while is unlikely to be much larger than . Let and let be defined as in Lemma 3.7; that is, . Set , and let . We first show

In the case that , the supermartingale is bounded from below, so we can choose . Player I is compelled to ensure . Conditional on the game up to time , player I cannot guarantee a conditional expected payoff better than . Since , we get Equation 3.5, as desired.

In the case , we let , where . Then . Note that follows from , and Lemma 3.5. Suppose player II makes a supermartingale up until time . Given play until time , player II may make sure that the conditional expected payoff to player I is at most

Taking expectation and separating into cases in which or not, we get

Since , this gives

The first term above is , independent of . Player I is compelled to play a strategy that ensures is finite a.s., since otherwise the payoff is . With such a strategy, as , and since is bounded from below and , the last summand tends to as . Thus, we get Equation 3.5 in this case as well.

Next, we show that . Let , and set , where , , and . Let . Then for . Since satisfies comparison from below and is -decreasing , we get on . Thus, if , then . In conjunction with Lemma 3.5, this implies that . Choosing and taking to therefore gives in Equation 3.5 . However, the inequality from Lemma 3.1 implies , completing the proof.

Proof of Theorem 1.3.

First, suppose that is Lipschitz. Let . Let and . We know that these limits exist from Lemma 3.1. Lemma 3.3 tells us that as . Since the assumptions of that lemma are player-symmetric, we likewise get . From Theorems 2.2 and 2.4 (possibly with the roles of the players interchanged) we know that , and hence the above gives . By Lemma 3.1, and , and so we conclude that and . Note that the assumptions imply that . Therefore, from Lemma 3.5 and we conclude that is Lipschitz. Precisely the same argument gives if is also assumed to be Lipschitz, and similar estimates also hold for . Note that the assumptions imply that is bounded if . Lemma 3.6 (applied possibly with the roles of the players reversed) tells us that satisfies -quadratic comparison on . Clearly . (If identically, then , while if , we may use the fact that is Lipschitz.) Thus, Lemma 3.8 implies that . Consequently, . By symmetry , and hence . This completes the proof in the case where is Lipschitz.

Using the result of Lemma 3.9 below, we know that for every there is a Lipschitz such that . Then the Lipschitz case applies to the functions in place of . Since the game value for is bounded between the corresponding value with and , and the latter two values differ by , the result easily follows.

Lemma 3.9.

Let be a length space, and let be defined on a non-empty subset . The following conditions are equivalent:

(1)

is uniformly continuous and .

(2)

extends to a uniformly continuous function on .

(3)

There is a sequence of Lipschitz functions on tending to in .

Proof.

We start by assuming (1) and proving (2). Let and . We now show that . Let , and let satisfy . Such a exists because is uniformly continuous. Condition (1) implies that . For , we have

which proves that . Two other immediate properties of which we will use are that holds for and when and . It follows that is subadditive: for .

Now for set . Then on . We now prove

for . Indeed, let . Then

Consequently, subadditivity and monotonicity of gives

Taking the supremum over all then implies Equation 3.6. Therefore, is uniformly continuous and (2) holds.

We now assume (2) and prove (3). Let be a uniformly continuous extension of to . It clearly suffices to approximate by Lipschitz functions on in . For let . The same argument which was used above to prove Equation 3.6 now shows that . Clearly, . Let for . Since is a length space, is subadditive. Let and . Then

Taking the supremum over all and using the subadditivity of therefore gives

Therefore, once . Since and , this proves (3).

The passage from (3) to (1) is standard, and therefore omitted. This concludes the proof.

Proof of Theorem 1.8.

Lemma 3.9 tells us that extends to a uniformly continuous function on and that it can be approximated in by Lipschitz functions. We know from Theorem 1.3 that the continuum value exists and is uniformly continuous. Suppose first that is Lipschitz. Lemma 3.6 (applied possibly with the roles of the players reversed) says that satisfies -quadratic comparison on . If is not Lipschitz, we may deduce the same result by approximating from below by Lipschitz functions and observing that a monotone non-decreasing -limit of functions satisfying -quadratic comparison from above also satisfies -quadratic comparison from above, and making the symmetric argument for comparison from below. Now suppose that is as in the second part of the theorem. This clearly implies . Now Lemma 3.8 gives (again, we may need to first approximate by Lipschitz functions). The uniqueness follows directly.

Proof of Theorem 1.4.

If and , then from Lemma 3.2 it follows that . Let be open and define for . It is clear that the continuum value of is the same as , since any player may first play a strategy that is appropriate for the game, and once is hit, start playing a strategy that is appropriate for the original game. The above argument shows that if and . In particular, , which implies ; that is, is AM in .

Now suppose that is an AM extension of and . Lemma 1.6 tells us that satisfies comparison with distance functions. Observe that the proof of Lemma 3.8 shows that in the case we may replace the hypothesis that satisfies -quadratic comparison from below on by the hypothesis that satisfies comparison with distance functions from below (since only comparisons with distance functions are used in this case). Thus, . Similarly, , which implies the required uniqueness statement and completes the proof.

4. Harmonic measure for

Here, we present a few estimates of the -harmonic measure . Before proving Theorem 1.5, we consider the -harmonic of porous sets. Recall that a set in a metric space is -porous if for every every ball of radius contains a ball of radius that is disjoint from . An example of a porous set is the ternary Cantor set in . We start with a general lemma in the setting of length spaces.

Lemma 4.1.

Suppose is a length space and on the terminal states , where is continuous. Let . Suppose that for some integer and positive constants and , for every with , there is a sequence of points such that , , and

for . Then the AM extension of satisfies

Proof.

We will obtain bounds on that are independent of (as long as ), and these yield bounds on . Since , we need only give a good strategy for player II to obtain an upper bound on . The idea is for player II to always have a “plan” for reaching a terminal state at which is , while staying away from the terminal states at which is non-zero. A plan consists of a sequence of points , where is the game state when the plan was formed, , and . As soon as the game state reaches , player II starts pulling towards . If player I is lucky and gets many moves, player II may have to give up on the plan and form a new plan. When tugging towards , we suppose that player II gives up and forms a new plan as soon as , and otherwise plays to ensure that is a supermartingale. While tugging towards , the plan will be aborted at that stage with probability at most , so with probability at least the plan is never aborted and succeeds in reaching .

Suppose player II aborts the plan at time while tugging towards , and forms a new plan starting at . Then , so

so the game state remains far from . Since player II can always find a short plan (length at most ) when the distance from is at least , player I gets to with probability at most , which yields the desired upper bound.

The reader may wish to check that the lemma implies in the setting of Theorem 1.5 (i.e., when is the unit ball in , , and is a spherical cap of radius .

We are now ready to state and prove an upper bound on the -harmonic measure of neighborhoods of porous sets.

Theorem 4.2.

Let , , be the closed unit ball and let be its boundary, the unit sphere. Let and let . Let be an -porous subset of , and let be the closure of the -neighborhood of . Then

Of course, in the above, is -porous as a subset of . (Every subset of is -porous as a subset of .)

Proof.

The plan is to use the lemma, of course. Let . Let , and suppose that . Let be a closest point to on . Inside there is a point such that . (See Figure 3.) Therefore, . We define the sequence inductively, as follows. If , then we take to be any closest point to on . Otherwise, let be the point on the line segment from to whose distance from is .

It can be checked that after steps the sequence hits and that the assumptions of the lemma hold with this , with and with some , independent of . The theorem now easily follows from the lemma.

We now proceed to study the -harmonic measure of spherical caps.

Proof of Theorem 1.5.

It turns out to be more convenient to work with in place of . By an obvious comparison argument, it is sufficient to estimate , where is the AM function in with boundary values

The function is invariant under rotations of preserving . Therefore, is also AM in . Thus, we henceforth restrict ourselves to the case , with no loss of generality.

Aronsson Reference 3 constructed a family of viscosity solutions to in that are separable in polar coordinates: (for ). We are interested in the solution, which may be written as

Observe that when this solution is non-negative, and when , say, where is some fixed constant. Then , where

Since is monotone in , and satisfies , it follows that .

We now show that the bound is tight. It is easy to see, using comparison with a cone centered at , say, that , where is a constant that does not depend on . Comparison with a cone centered at any point in the unit disk shows that if . Using such estimates, it is easy to see that there is a constant such that on the disk of radius centered at , provided that , say. Now consider the function

By the choice of the constant , we have on , since in . If denote the polar coordinates centered at , the center of , and denote the standard polar coordinates centered at , then when , and we choose . Also, is bounded from below by a constant times when and . Since when , it follows that on the unit circle

Therefore on the boundary of the unit circle, as well as on . Consequently, in the complement of in the unit disk. This implies that there is some such that for all sufficiently small . The required estimate now follows by several applications of comparison with cones (the number of which depends on ), similar to the above argument estimating a lower bound for on .

One could also try to use Aronsson’s other solutions to bound -harmonic measure of certain sets in other domains. Also, Aronsson has some explicit solutions for , which may serve a similar purpose.

5. Counterexamples

5.1. Tug-of-war games with positive payoffs and no value

Here we give the promised counterexample showing that the hypothesis in Theorem 1.2 cannot be relaxed to . As we later point out, a similar construction works in the continuum setting of length spaces.

The comb game is tug-of-war with running payoffs on an infinite graph shaped like a comb, as shown in Figure 4. The comb is defined by an infinite sequence of positive integers , and has states (vertices) . The edges of the comb are of the form and , giving the graph the shape of a comb, where the tooth of the comb has length . The running payoff is if and zero otherwise. (Later we will consider a variation where everywhere.) The terminal states are the states of the form , and the terminal payoff is zero on all terminal states.

Lemma 5.1.

For any comb (choice of the sequence ), we have .

Proof.

First we argue that . Denote by the accumulated running payoff. Fix a large and equip player I with the following strategy: at all times for which , player I pulls down if , left if and , and right if ; if , then player I pulls toward the closest terminal state. Define the termination time and also the stopping time . Fix any strategy for player II and observe that is a non-negative supermartingale, which must converge a.s.; therefore, from any initial state, a.s.; therefore, the stopping time is almost surely finite, whence a.s. as well.

Consider the process

Then is a submartingale. Note that if , then and , while if , then and . In any case, . Thus by optional stopping,

Next, to show that , suppose that player II adopts the strategy of always pulling up, and player I knows this and seeks to maximize her payoff. Because player II always pulls up, is a positive supermartingale, so a.s. converges to , and by optional stopping,

This shows that by always pulling up, player II can force player I’s expected payoff from to be no more than .

Remark 5.2.

Note that the strategy for player II of always pulling up does not necessarily force the game to end with probability one; this strategy always gives an upper bound on , but it only yields an upper bound on if , when the Borel-Cantelli Lemma ensures termination.

Suppose that the teeth of the comb are long, i.e., , and that player I plays the strategy of always pulling down if and always pulling right if . If player II plays the strategy of always pulling up, then we may calculate the probability that the game terminates when started in state . Either the state goes to the terminal state of the tooth or goes to the base of the next tooth, and the latter probability is . When the teeth of the comb are long, the game lasts forever with probability . If player II needs to ensure that the game terminates, she will need to be prepared to sometimes pull left instead of up, and this necessity will be costly for player II.

Lemma 5.3.

For every there is a suitable comb (choice of the sequence ) such that .

Proof.

Let player I play the strategy of pulling down when and pulling to the right when . We will estimate from below the expected payoff when player II plays any strategy which guarantees that the game terminates in finite time a.s. against the above strategy for player I. Let denote the probability that the game terminates at the terminal state . Then . For every let and denote the expected number of game transitions from to the left, upwards and to the right, respectively (that is, to , to and to , respectively).

First, observe that

because every time that the game state arrives at , with conditional probability at most , the game terminates at before returning to . Next, observe that

because the number of transitions from to can exceed the number of transitions from to by at most one. (More precisely, we have , but this will not be needed.) Finally, note that

because player I always pulls right at . (We set .)

An easy induction shows that the above relations imply

The expected payoff satisfies

We plug in the above lower bound on to obtain

Since , the lemma follows if we choose with sufficiently large.

Note that when player II always pulls up, for every vertex in the comb there is a finite upper bound on the expected number of visits to , which holds regardless of the strategy used by player I. Since in the proof of Lemma 5.1 player II always pulls up, it follows that the value for player I is at most even when is replaced by , where and . This modification will certainly not decrease the value for player II. Therefore, in Theorem 1.2, the assumption cannot be replaced by the assumption , even if throughout .

Note that we may convert the discrete comb graph to a continuous length space by adding line segments corresponding to edges in the comb. It is then easy to define the corresponding continuous and conclude that in Theorem 1.3 the assumption cannot be relaxed to . The details are left to the reader.

Remark 5.4.

The dependence of on the growth of is somewhat surprising. If grows slowly enough so that , then , because (as remarked above) the Borel-Cantelli lemma implies that the strategy of always pulling up is guaranteed to terminate, and thus the proof of Lemma 5.1 applies. We have seen that for some sequences of polynomial growth, can be arbitrarily large. However, if grows rapidly enough so that , then . This is immediate from the following more general statement: for every , we have To prove this, suppose player II adopts the strategy of always pulling left when and and up otherwise. Let denote the payoff accumulated for player I at points in up to time . Then , because

is a positive supermartingale with . Then we claim that

Fix . Starting at , the expected number of visits to before returning to is at most 1 (by comparison to simple random walk). The expected number of visits to is at most , since each time , there is a chance of at least of terminating at without returning to . Thus the expected accumulated payoff at is at most ; summing over proves the claim.

5.2. Positive Lipschitz function with multiple AM extensions

Here we show that uniqueness in Theorem 1.4 may fail if is unbounded; that is, we give an (here ) for which is Lipschitz and positive and the continuum value is not the only AM extension of .

Let be the rooted ternary tree, where each node has three direct descendants and every node but the root has one parent. Let be the corresponding length space, where we glue in a line segment of length for every edge in . For every node in , we label the three edges leading to descendants of by , and . (One may interpret as the set of finite stacks of cards, where each card has one of the three labels , , and .)

We define a function on the nodes of by induction on the distance from the root. For any vertex let be the number of edges on the simple path from the root to whose label is not . Set . Next, if is the parent of and the edge from to is labeled , then , respectively. Finally, if the edge from to is labeled , let , say. This defines on the vertices of . We define it on by linear interpolation along the edges.

Let denote the set of nodes consisting of the root and all vertices such that the edge from to its parent is not labeled . For each let be some large integer, whose value will be later specified, and let be the vertex at distance away from along the (unique) infinite simple path starting at which contains only edges labeled . Let and let be the set of all nodes such that . Set .

We first claim that is AM on . Indeed, let be an open subset of . If does not contain any tree node, then , because interpolates linearly inside the edges. Suppose now that is a node. Let be the infinite path starting at always going away from the root which uses only edges labeled . For each positive integer let be the infinite path starting at always going away from the root whose first edges are labeled and the rest are labeled . Observe that meets and meets . Consequently, and . If and , then by the construction of . Thus . Since , this proves that is AM on .

Next, we consider the discrete tug-of-war game on the graph , where and is the restriction of to . Let , the starting position of the game, be the third vertex on the infinite simple path from the root whose edges are all labeled . We claim that the value for player I satisfies . Before proving this we explain the idea: at each step, player I has one or two moves that increase the value of by (either moving away from the root along an edge with label — i.e., “adding a card to the deck” — or moving towards the root along an edge labeled — i.e., “removing a card from the top of the deck”), and similarly, player II has one or two moves that decrease the value of by . Player I can thus make a submartingale by always making moves of this type.

This does not force the game to terminate, however. In order to end the game favorably by reaching a point in , player I must add a sequence of cards labeled to the top of the deck. If player II adopts the strategy of always choosing the edge labeled (“adding a card to the deck”), then (provided that increases rapidly enough) the expected number of suboptimal moves that player I must make (by moving along an edge with label ) to reach any particular element of is large enough to significantly decrease the total expected payoff for player I.

We proceed to prove that . Let be a vertex in , which is in the same connected component of as . Let be the simple path from to . We now abbreviate , and .

Let player II use the naive strategy of always trying to move from the current state along the edge labeled going away from the root. For let be the first time such that , and if no such exists let . (As before is the game position at time .) Let for and for . Then is clearly a supermartingale, regardless of the strategy used by player I. Since the game terminates at time if , this implies that for

Consequently,

Therefore

Since , we can make this smaller than any required positive number by choosing sufficiently large. We may therefore choose the function so that

Since on , this implies .

Let be the linear interpolation of to the edges. By Lemma 1.1 is discrete -harmonic. It follows that in the II-favored -tug-of-war game, for every and , player II can play to make a supermartingale. Consequently, Lemma 3.3 implies that the value of the continuum game is bounded by on the vertices. Since the continuum value is AM on , by Theorem 1.4, this proves our claim that the assumption is necessary for uniqueness to follow in the setting of that theorem.

5.3. Smooth on a disk with no game value

The following example is a continuum analog of the triangle example at the beginning of Section 2.2.

Let be a closed disk in , and let be its boundary. We now show that for some smooth function , the value for player I and the value for player II differ in -tug-of-war with on . Not only are the values different, but the difference does not shrink to zero as . There is a lot of freedom in the choice of the function , but an essential property, at least for the proof, is that is anti-symmetric about a line of symmetry of .

It will be convenient to identify with , and use complex numbers to denote points in . Let be a function such that:

(1)

inside the disk ,

(2)

in the right half plane ,

(3)

in , and

(4)

is anti-symmetric about the imaginary line.

Let be large, let , and on . Consider small. We want to show that for an appropriate choice of , then

Indeed, suppose that this is not the case. Let be small, and suppose that . Symmetry gives . The assumption therefore tells us that . In particular, on the imaginary axis. We now abbreviate . Clearly on and on . By considering strategies which pull towards the imaginary axis and then using whatever strategy gives a value in , it is easy to see that (as usual, means that there is some universal constant such that ) and that in , say. (See, e.g., the proof of Lemma 3.5.) It is then easy to see that there is a constant , which does not depend on and (as long as is sufficiently small and , say), such that on the disk and on the disk . Since is nearly anti-symmetric, it is enough to prove the first claim. Indeed, a strategy for player I which demonstrates this is one in which she pulls towards until she accumulates a payoff of . If successful, she then aims towards the boundary , but whether successful or not, whenever the game position comes within distance of the imaginary axis, she changes strategy and adopts an arbitrary strategy that yields a payoff . We assume that and are sufficiently small so that the term is much smaller than .

It now easily follows from Theorem 1.5 that . Consequently, there is some constant such that on , provided that is sufficiently small. Set , and let be the set of points such that

Since on and on , it follows that every path in the graph from to must intersect . (We are assuming .) Let denote the union of and all connected components of that are disjoint from .

We now claim that there is some (which does not depend on ) such that and therefore . Indeed, if satisfy and , say, then the strategies for either player of pulling towards and only giving up when the current position is within distance of show that

This proves the existence of such an .

We now let player II play a strategy very similar to the backtracking strategy she used in the proof of Theorem 2.2, which we presently describe in detail. Let denote the sequence of positions of the game. Assume that . For each , let . Note that if , then . By Lemma 1.1 inside the function satisfies . If and player II gets the turn, she moves to some that satisfies . While if and player II gets the turn, she moves to any neighbor of that is closer (in the graph metric) to than in the subgraph of spanned by the vertices .

Now consider the game evolving under any strategy for player I and the above strategy for player II. Let be the graph distance from to in . Set

As in the proof of Theorem 2.2, it is easy to verify that and that is a supermartingale.

Let be the time in which the game stops; that is, . Assume that a.s. Suppose that we could apply the optional stopping time theorem. Then we would have

Thus,

But , because the Euclidean distance from to is at least . Since is at most the supremum of expected payoffs under the current strategy for player II and an arbitrary strategy guaranteeing for player I, we obtain . This contradicts our previous conclusion that , because we may choose large.

How do we justify the application of the optional stopping time theorem? We slightly modify the strategy for player II. So far, our analysis utilized one advantage to player II in the calculation of ; that is, that it is player I’s responsibility to make sure that . Now we have to use the other advantage to player II, which is that player II gets to choose her strategy after knowing what player I’s strategy is. Let be the first time such that with the strategy which player I is using and the above strategy for player II, the game terminates by time with probability at least . The new strategy for player II is to play the above strategy until time , and if the game lasts longer, to use an arbitrary strategy which would guarantee a conditioned expected future payoff of . We may certainly apply the optional stopping time theorem at time . This does give a contradiction as above, because , and the contradiction proves our claim Equation 5.1.

5.4. Lipschitz on unit disk with multiple solutions to

Here we construct the counterexample showing that if we omit the assumption , in Corollary 1.9, then it may fail. In the example, is the unit disk in , is its boundary, , and is Lipschitz and take values of both signs. The example is motivated by and similar to the example of Section 5.3. However, since the assumptions of Corollary 1.9 do not hold (obviously), we need to construct the solutions to not by using tug-of-war, but by other means. In fact, we will use a smoothing of Aronsson’s function Equation 4.1.

The following analytic-geometric lemma will replace the use of the backtracking strategy for player II in Section 5.3. The following notation will be needed. For and let

Lemma 5.5.

Let be a length space, and let be bounded Lipschitz real-valued functions on a non-empty subset . Let and be the corresponding AM extensions to . Fix , and let . Suppose that on , that and

Then in .

We need a few simple observations before we begin the proof. Suppose that is an open connected subset of a length space , that is non-empty, and that is Lipschitz and bounded. Let denote the set of all points such that there is a finite length path where and . Set , and for two points let be the infimum length of paths joining and such that is finite. Note that is a length space. Also, since on , the restriction of to is Lipschitz. Consequently, this restriction of has an AM extension with respect to . We claim that this is also an AM extension of with respect to .

Observe that for any Lipschitz function where is open, we have

(This can be verified by considering for every pair of distinct points the restriction of to a nearly shortest path joining and in .) Hence, to show that is AM with respect to it is enough to prove that

holds for arbitrary open . Observe that for we have , where refers to . Since is AM with respect to , we have for all . Finally, because on . This proves Equation 5.3 and thereby shows that is AM on with respect to . In particular, we conclude that

for the AM extension of to .

Proof of Lemma 5.5.

First, it is clear that is a closed set. Let

and let be the AM extension of to . We claim that is also an AM extension of to . Since uniqueness of AM extensions holds in this setting, this will imply and complete the proof, because on .

Let be a connected component of , and let denote the corresponding metric on . Note that , since for . By Equation 5.2 and the assumption , it follows that . Thus, we get for , which implies for .

Now let be open. In order to prove that is AM, we need to establish Equation 5.3. Without loss of generality, we assume that is connected. If , then Equation 5.3 certainly holds, since is AM in .

Suppose now that . Then , by the definition of . Since is AM, by Equation 5.4 with in place of , in place of and in place of , there is a sequence of pairs in and a sequence of paths from to , respectively, such that ,

and is finite. For all sufficiently large , . Hence . Let be the first point on that is in , and let be the last point on that is in . Note that as , because , Equation 5.5 holds and . (If at some significant proportion of the length of the function does not change in speed very close to , it will not have enough distance to catch up.) Similarly, . Since and and on , it follows that and similarly . Thus, , proving that , as needed.

Our example is based on smoothing two different modifications of Aronsson’s -harmonic function from Equation 4.1. For let be the set of points in such that . Define the inner and outer radii of : and , and note that , while for every . Now fix some large . Let denote restricted to the annulus . Let , , denote the AM function on whose boundary values are equal to on the inner circle and are on the outer circle, respectively. Lemma 5.5 implies that on (provided that was chosen sufficiently large). Shortly, we will show that there is a function defined on the unit disk that agrees with in and such that (in the viscosity sense), where is a Lipschitz function. Assuming this for the moment, we can then define in and inside the disk . Then and are both on the unit circle, and both satisfy , providing the required example. It thus remains to prove

Lemma 5.6.

For every there are Lipschitz functions such that agrees with on and holds in the viscosity sense.

Proof.

Let

be the angle factor of in Equation 4.1. First, observe that and . Next, note that when . We now verify that is in a neighborhood of . Write

The first two factors are real-analytic near . Setting , which is real-analytic near , allows us to rewrite the last factor as

which is also real-analytic near . We conclude that is real-analytic at every .

It is instructive to see that is -harmonic in . Verifying away from the real line can be done by differentiation. At points where , we have . At such points , and indeed there. However,

near . This implies that there is no function that satisfies and in a neighborhood of . Thus, .

We return to the proof of the lemma, and first establish that the lemma holds when we only require that be continuous, instead of Lipschitz. In this case, the function may be written as

for appropriately chosen functions and . The functions and will be , will satisfy in a neighborhood of and , for all . It follows that is Lipschitz and that is away from the -axis and in a neighborhood of . It remains to check the behavior of at points on and close to the -axis. Based on Equation 5.6, we may estimate for close to and :

where is some polynomial. Let satisfy . Choose as a function that is for , is for , and is strictly monotone in . Choose as a function that is in and is for . When , , and hence is by Equation 5.7. In the range , we use Equation 5.8 to conclude that the limit as of exists.

We now argue that there is a continuous function such that in the viscosity sense. Suppose that is a function defined in a small open set containing the point such that the minimum of in occurs at . We will now perturb . For sufficiently small the function is a function defined in and is the unique minimum of in . If is chosen much smaller, then the function will still satisfy that the infimum of is attained in . Moreover, that infimum cannot be attained on the -axis, because is zero on the -axis, is not parallel to the -axis, and exists. Since and are arbitrarily small, we conclude that as tends to through a point not on the -axis. Thus, . Similarly, . This proves , where is the continuous extension of off the -axis to the whole plane.

If we want to be Lipschitz, we need to eliminate the term in Equation 5.8. Define

Then near we have

Set

where and are to be chosen soon. A calculation shows that near we have

Of course, when and are chosen so that , all the terms on the right hand side vanish. Our goal is to choose these functions so that the following holds:

(1)

for we have ,

(2)

when is sufficiently small, we have ,

(3)

the term vanishes identically,

(4)

we don’t have a blow up due to , and finally

(5)

the functions are , say.

This is not hard. Fix such that . For , we choose the so that . In particular, in this range. In the range we maintain . The function is chosen so that throughout we have , while does not vanish in . This is possible, because for and at . For every , takes the unique value for which the term vanishes. Since and do not vanish in the interval, it is clear that such a choice for is possible and is provided that and are also . Throughout we take . This allows us to simplify the expression for in this range, to obtain

Now the denominators cannot vanish before . Thus, takes care to make the term vanish while evolves to become zero throughout . This completes the proof.

6. Viscosity solutions and quadratic comparison in

In this section we prove Theorem 1.7, which states that in bounded domains in Euclidean space , is a viscosity solution of iff satisfies -quadratic comparison. The following lemma will be useful in that proof. Let denote the Hessian matrix .

Lemma 6.1.

Let be any real-valued function which is in a neighborhood of and satisfies . Then for every , there exist quadratic distance functions and such that:

(1)

; .

(2)

for . Also, as quadratic forms, strictly dominates , which in turn strictly dominates .

(3)

Consequently, at all points in a neighborhood of .

(4)

and are centered at and , respectively, with for .

(5)

In a neighborhood of , the function is -decreasing in the distance from and is -increasing in the distance from .

Proof.

We begin with preliminary calculations about the quadratic distance function , centered at the origin. Fix some . Let be the unit vector in the -direction and be any unit vector orthogonal to . Write .

Then direct calculations show:

(1)

.

(2)

.

(3)

.

(4)

.

(5)

whenever .

Of course, it is enough to complete this calculation in dimension two when and and to deduce the general case by symmetry.

We now construct the described in the lemma (the construction of is similar). We will take for some small value of (specified below) and write , where we first choose to be an arbitrary real with the property that , we then choose so that and we then choose so that . We must now show that the thus constructed satisfies the requirements of the lemma.

We can compute explicitly in terms of , , , , and using the relation . As tends to zero, tends to .

The description of given above implies that if and is a unit vector orthogonal to , and , then

(1)

.

(2)

.

(3)

.

Note that tends to as . Let . We claim that by choosing sufficiently small we can make sure that dominates as a quadratic form (or any other fixed quadratic form satisfying , for that matter). Let and . If is non-zero, we may write , where and is orthogonal to . Then

On the other hand . Since for some constant , the domination follows from . This constructs with the required properties. A similar construction applies to .

Proof of Theorem 1.7.

Let and suppose that satisfies -quadratic comparison in a neighborhood of . Let be a real-valued function defined in a neighborhood of . Suppose that and has a local minimum at . The above lemma implies that for every there is a quadratic distance function such that for all in some neighborhood of , and is -increasing in a neighborhood of . Since has a strict local minimum at , it follows by -quadratic comparison that arbitrarily close to there are points for which . By continuity of , this implies . That is, . Since was arbitrary, this also implies .

Now we remove the assumption that and assume instead that as . If , we may take , and the same argument as above gives . Now suppose and let . In this case, we have to modify the argument, because is not -increasing in a neighborhood of . Recall that has a local minimum at and has a strict local minimum at . Therefore, has a strict local minimum at . Let be sufficiently small so that on . By continuity, for every with sufficiently small for every . Fix such an satisfying , and let be a point in where attains its minimum. Since

it follows that . On the other hand, , because

Therefore, . The first case we have considered therefore gives . Since was arbitrary, continuity gives , as required. Thus, we conclude that . By symmetry, , and hence is a viscosity solution as claimed.

This completes the first half of the proposition. For the converse, suppose that is a viscosity solution of . Let be open. Suppose that is a -increasing quadratic distance function in , on and there is some such that . For all sufficiently small we still have on . Let be a point in where attains its minimum. Note that , since . Consequently, . Thus, . Therefore, satisfies -quadratic comparison on , and the proof is complete.

7. Limiting trajectory

7.1. A general heuristic

For every , when both players play optimally in -tug-of-war, the sequence of points visited is random. Do the laws of these random sequences, properly normalized, converge in some sense to the law of a random continuous path as tends to zero?

We give a complete answer in only a couple of simple cases. However, we can more generally compute the limiting trajectory when is in a domain contained in and the players move to maximize/minimize instead of ; we suspect but cannot generally prove that the limiting behavior will be the same when the players use .

Consider a point in the domain at which . If is small enough, then throughout the closed ball , so the extrema of on the closed ball lie on the surface of the ball. Then at any such extremum , by the Lagrange multipliers theorem, for some real . Since is , we have , where . We define , , and . Then , where . If we solve this with small , we find

Then , and thus , and so . Hence,

When is and has non-zero gradient in a domain, this suggests the SDE

where and is equal to minus its projection onto (so that and are always orthogonal). Now Itô’s formula implies that, as expected, is a martingale. In particular, when is infinity harmonic, is a martingale.

All of the above analysis applies only when players make moves to optimize instead of , as they would do if they were playing optimally. This difference is what makes the calculation of the limiting trajectory a heuristic, except in a few special cases as described in the next subsection.

7.2. Special cases

The above analysis does apply to optimally played games in a couple of simple cases for which . Let and let be the complement of a bounded set. Then the following are infinity harmonic functions which satisfy both and :

(1)

is the distance from to a fixed convex set whose -neighborhood is contained in .

(2)

is the argument of on a -neighborhood of , and defined arbitrarily elsewhere (here we assume that the -neighborhood of does not contain a simple closed curve surrounding , so that the argument can be defined there).

(Due to boundary errors the above need not hold when is a bounded domain and is its boundary.) In the first case, is simply a Brownian motion along a (straight) gradient flow line of . In the second case, is a diffusion with drift in the direction and diffusion of constant magnitude orthogonal to .

Crandall and Evans Reference 10 have explored the following question in some detail, and it has recently been answered affirmatively by Savin Reference 23 when : is every -harmonic function on a domain in everywhere differentiable? This question can be rewritten as a question about the amount of variation of the optimal direction of the first move (as a function of the starting point) in -tug-of-war. An affirmative answer might be a step towards a more complete analysis of the limiting game trajectories when and when is not smooth, since it would at least ensure that is well defined everywhere that the gradient is non-zero.

8. Additional open problems

(1)

If is open and bounded, Lipschitz, and Lipschitz, is there a unique viscosity solution for with boundary values given by for generic ? Here, generic could mean in the sense of Baire category, or it could mean almost every, or perhaps this could be true except for a countable set of .

(2)

Does Theorem 1.8 continue to hold if and are merely continuous instead of uniformly continuous? Can the requirement be replaced with (or ) when has finite diameter? When solving on bounded domains in , can the condition that be continuous be replaced with a natural weaker condition (e.g., semicontinuity or piecewise continuity)?

(3)

In Section 5.2 we gave a triple with positive and Lipschitz for which the continuum value of tug-of-war is not the unique AM extension of .

Is there an example where is the closure of a connected open subset of and is its boundary? In particular, let be the set of points in above the graph of the function , let be the boundary of in , and let on . Is the only AM extension of to ? What is the value of the corresponding game? (Note added in revision: Changyou Wang and Yifeng Yu (personal communication) have recently shown that is the only AM extension of to . The proof turned out to be rather simple.)

(4)

Suppose that is Lipschitz and are continuous on an open set and that as well as , both in the viscosity sense. Does it follow that ? (Note added in revision: Yifeng Yu Reference 25 proved this in the case .)

Acknowledgements

We thank Alan Hammond and Gábor Pete for useful discussions and the referee for corrections to an earlier version.

Figures

Figure 1.

A graph for which the obvious tug-of-war strategy of maximizing/minimizing does poorly.

Graphic without alt text
Figure 2.

At the end of the round, player I reaches the target (or a point in ) with probability at least , and otherwise the state still remains within . During the first step of the round, when player I has target , the running payoff is the infimum of over , and during any step of the round the infimum is over a subset of .

Graphic without alt text
Figure 3.

The terminal states are the unit sphere, and the support of is , the -neighborhood of a porous set on the unit sphere. From a starting point , player II finds a point on the sphere that is near the closest point on the sphere but far from . Player II then tugs towards along a sequence of points, and then when it gets much closer to the sphere than to , it tugs straight to the sphere.

Graphic without alt text
Figure 4.

A comb.

Graphic without alt text

Mathematical Fragments

Equation (1.1)
Lemma 1.1.

The function satisfies the equation

for every non-terminal state for which the right-hand side is well defined, and when the right-hand side is of the form . The analogous statement holds for , except that when the right-hand side of 1.2 is of the form .

Theorem 1.2.

A tug-of-war game with parameters has a value whenever the following hold:

(1)

Either everywhere or .

(2)

.

(3)

is undirected.

Theorem 1.3.

Suppose is a length space, is non-empty, is bounded below and has an extension to a uniformly continuous function on , and either satisfies or all three of the following hold: , is uniformly continuous, and has finite diameter. Then the continuum value exists and is a uniformly continuous function extending . Furthermore, as . If is Lipschitz, then so is . If and are Lipschitz, then .

Theorem 1.4.

Let be a length space and let be Lipschitz, where . If , then the continuum value function described in Theorem 1.3 (with ) is an AM extension of . If is also bounded, then is the unique AM extension of .

Theorem 1.5.

Let be the unit ball in , , let , let be the center of the ball, and for each let be a spherical cap of radius (of dimension ). Then

where are absolute constants (which do not depend on ).

Lemma 1.6 (Reference 8).

Let be an open subset of a length space. A continuous satisfies comparison with distance functions in if and only if it is AM in .

Theorem 1.7.

Let be a real-valued continuous function on a bounded domain in , and suppose that is a continuous function on . Then satisfies -quadratic comparison on if and only if is a viscosity solution to in .

Theorem 1.8.

Suppose is a length space, is non-empty, is uniformly continuous and bounded, and either satisfies or all three of the following hold: , is uniformly continuous, and has finite diameter. Then the continuum value is the unique continuous function satisfying -quadratic comparison on and on . Moreover, if is continuous, satisfies on and -quadratic comparison from below on , then 1 throughout .

Corollary 1.9.

Suppose is a bounded open set, is uniformly continuous, and satisfies either or and is uniformly continuous. Then there is a unique continuous function that is a viscosity solution to on and satisfies on . This unique solution is the continuum value of tug-of-war on .

Lemma 2.1.

Suppose that and . Then is the smallest -harmonic function bounded from below on that extends . More generally, if is an -harmonic function which is bounded from below on and on , then on .

Theorem 2.2.

Suppose that is connected, and . If is bounded below (or above) on and , then , so the game has a value.

Corollary 2.3.

If is connected, and , then is the unique bounded -harmonic function agreeing with on .

Theorem 2.4.

Suppose that is connected and . Assume that is bounded from below and . Then . If, additionally, and are bounded from above, then any bounded solution to with the given boundary conditions is equal to .

Equation (3.1)
Lemma 3.1.

For any ,

Lemma 3.2.

Let . Suppose that and either

(1)

everywhere,  or

(2)

is bounded above and has finite diameter.

Then for each and ,

Such an expected payoff is guaranteed for player I if she adopts a pull towards strategy, which at each move attempts to reduce . Similarly a pull towards strategy for II gives

Lemma 3.3.

Suppose is Lipschitz, and either

(1)

everywhere, or

(2)

is uniformly continuous, has finite diameter and .

Then as . If is also Lipschitz, then .

Lemma 3.4.

Let , let be an open subset of and write . Suppose that is a quadratic distance function that is -increasing on , where satisfies

Also suppose that or . If the value function for player I in -tug-of-war satisfies on , then on .

Equation (3.3)
Lemma 3.5 (Uniform Lipschitz).

Suppose that is Lipschitz, and either

(1)

everywhere,  or

(2)

is bounded from above and has finite diameter.

Then for each , and are Lipschitz on w.r.t. the metric with the Lipschitz constant depending only on and .

Equation (3.4)
Lemma 3.6.

Suppose is Lipschitz and , and either (1) identically or else (2) , is uniformly continuous, and has finite diameter. Then the subsequential limit satisfies -quadratic comparison on .

Lemma 3.7.

Suppose that is continuous and satisfies -quadratic comparison from below on , and that is locally bounded from below. Let . Then in II-favored -tug-of-war, when player I (using any strategy) targets point on step , player II may play to make a supermartingale, where

and .

Lemma 3.8.

Suppose that is Lipschitz, and either

(1)

everywhere,  or

(2)

is bounded from above and has finite diameter.

Also suppose that is continuous, satisfies -quadratic comparison from below on , on , and . Then for all .

Equation (3.5)
Lemma 3.9.

Let be a length space, and let be defined on a non-empty subset . The following conditions are equivalent:

(1)

is uniformly continuous and .

(2)

extends to a uniformly continuous function on .

(3)

There is a sequence of Lipschitz functions on tending to in .

Equation (3.6)
Equation (4.1)
Lemma 5.1.

For any comb (choice of the sequence ), we have .

Equation (5.1)
Lemma 5.5.

Let be a length space, and let be bounded Lipschitz real-valued functions on a non-empty subset . Let and be the corresponding AM extensions to . Fix , and let . Suppose that on , that and

Then in .

Equation (5.3)
Equation (5.4)
Equation (5.5)
Equation (5.6)
Equation (5.7)
Equation (5.8)

References

Reference [1]
Gunnar Aronsson, Extension of functions satisfying Lipschitz conditions, Ark. Mat. 6 (1967), 551–561. MR 0217665 (36:754)
Reference [2]
—, On the partial differential equation , Ark. Mat. 7 (1968), 395–425. MR 0237962 (38:239)
Reference [3]
—, Construction of singular solutions to the -harmonic equation and its limit equation for , Manuscripta Math. 56 (1986), no. 2, 135–158. MR 850366 (87j:35070)
Reference [4]
Gunnar Aronsson, Michael G. Crandall, and Petri Juutinen, A tour of the theory of absolutely minimizing functions, Bull. Amer. Math. Soc. (N.S.) 41 (2004), no. 4, 439–505 (electronic). MR 2083637 (2005k:35159)
Reference [5]
M. Bardi, M. G. Crandall, L. C. Evans, H. M. Soner, and P. E. Souganidis, Viscosity solutions and applications, Lecture Notes in Mathematics, vol. 1660, Springer-Verlag, Berlin, 1997, Lectures given at the 2nd C.I.M.E. Session held in Montecatini Terme, June 12–20, 1995, edited by I. Capuzzo Dolcetta and P. L. Lions, Fondazione C.I.M.E. [C.I.M.E. Foundation]. MR 1462698 (98b:35005)
Reference [6]
G. Barles and Jérôme Busca, Existence and comparison results for fully nonlinear degenerate elliptic equations without zeroth-order term, Comm. Partial Differential Equations 26 (2001), no. 11-12, 2323–2337. MR 1876420 (2002k:35078)
Reference [7]
T. Bhattacharya, E. DiBenedetto, and J. Manfredi, Limits as of and related extremal problems, Rend. Sem. Mat. Univ. Politec. Torino 1989 (1991). Special issue on topics in nonlinear PDEs, 15–68. MR 1155453 (93a:35049)
Reference [8]
Thierry Champion and Luigi De Pascale, Principles of comparison with distance functions for absolute minimizers, J. Convex Anal. 14 (2007), no. 3, 515–541. MR 2341302
Reference [9]
M. G. Crandall, L. C. Evans, and R. F. Gariepy, Optimal Lipschitz extensions and the infinity Laplacian, Calc. Var. Partial Differential Equations 13 (2001), no. 2, 123–139. MR 1861094 (2002h:49048)
Reference [10]
Michael G. Crandall and L. C. Evans, A remark on infinity harmonic functions, Proceedings of the USA-Chile Workshop on Nonlinear Analysis (Viña del Mar-Valparaiso, 2000) (San Marcos, TX), Electron. J. Differ. Equ. Conf., vol. 6, Southwest Texas State Univ., 2001, pp. 123–129 (electronic). MR 1804769 (2001j:35076)
Reference [11]
Michael G. Crandall and Pierre-Louis Lions, Viscosity solutions of Hamilton-Jacobi equations, Trans. Amer. Math. Soc. 277 (1983), no. 1, 1–42. MR 690039 (85g:35029)
Reference [12]
Lawrence C. Evans and Yifeng Yu, Various properties of solutions of the infinity-Laplacian equation, Comm. Partial Differential Equations 30 (2005), no. 7-9, 1401–1428. MR 2180310 (2006g:35062)
Reference [13]
Robert Jensen, Uniqueness of Lipschitz extensions: minimizing the sup norm of the gradient, Arch. Rational Mech. Anal. 123 (1993), no. 1, 51–74. MR 1218686 (94g:35063)
Reference [14]
Petri Juutinen, Absolutely minimizing Lipschitz extensions on a metric space, Ann. Acad. Sci. Fenn. Math. 27 (2002), no. 1, 57–67. MR 1884349 (2002m:54020)
Reference [15]
Andrew J. Lazarus, Daniel E. Loeb, James G. Propp, Walter R. Stromquist, and Daniel H. Ullman, Combinatorial games under auction play, Games Econom. Behav. 27 (1999), no. 2, 229–264. MR 1685133 (2001f:91023)
Reference [16]
Andrew J. Lazarus, Daniel E. Loeb, James G. Propp, and Daniel Ullman, Richman games, Games of No Chance (Richard J. Nowakowski, ed.), MSRI Publications, vol. 29, Cambridge Univ. Press, Cambridge, 1996, pp. 439–449. MR 1427981 (97j:90101)
Reference [17]
Donald A. Martin, The determinacy of Blackwell games, J. Symbolic Logic 63 (1998), no. 4, 1565–1581. MR 1665779 (2000d:03118)
Reference [18]
E. J. McShane, Extension of range of functions, Bull. Amer. Math. Soc. 40 (1934), no. 12, 837–842. MR 1562984
Reference [19]
V. A. Milman, Absolutely minimal extensions of functions on metric spaces, Mat. Sb. 190 (1999), no. 6, 83–110. MR 1719573 (2000j:41041)
Reference [20]
Abraham Neyman and Sylvain Sorin (eds.), Stochastic games and applications, NATO Science Series C: Mathematical and Physical Sciences, vol. 570, Dordrecht, Kluwer Academic Publishers, 2003. MR 2032421 (2004h:91004)
Reference [21]
Adam M. Oberman, A convergent difference scheme for the infinity Laplacian: construction of absolutely minimizing Lipschitz extensions, Math. Comp. 74 (2005), no. 251, 1217–1230 (electronic). MR 2137000 (2006h:65165)
Reference [22]
Yuval Peres, Oded Schramm, Scott Sheffield, and David B. Wilson, Random-turn Hex and other selection games, Amer. Math. Monthly 114 (2007), no. 5, 373–387. MR 2309980 (2008a:91039)
Reference [23]
Ovidiu Savin, regularity for infinity harmonic functions in two dimensions, Arch. Ration. Mech. Anal. 176 (2005), no. 3, 351–361. MR 2185662 (2006i:35108)
Reference [24]
Hassler Whitney, Analytic extensions of differentiable functions defined in closed sets, Trans. Amer. Math. Soc. 36 (1934), no. 1, 63–89. MR 1501735
Reference [25]
Yifeng Yu, Uniqueness of values of Aronsson operators and applications to “tug-of-war” game theory, 2007, http://www.ma.utexas.edu/yifengyu/uofa2d.pdf.

Article Information

MSC 2000
Primary: 91A15 (Stochastic games), 91A24 (Positional games), 35J70 (Elliptic partial differential equations of degenerate type), 54E35 (Metric spaces, metrizability), 49N70 (Differential games)
Keywords
  • Infinity Laplacian
  • absolutely minimal Lipschitz extension
  • tug-of-war
Author Information
Yuval Peres
Microsoft Research, One Microsoft Way, Redmond, Washington 98052, and Department of Statistics, 367 Evans Hall, University of California, Berkeley, California 94720
Homepage
MathSciNet
Oded Schramm
Microsoft Research, One Microsoft Way, Redmond, Washington 98052
Homepage
Scott Sheffield
Microsoft Research, One Microsoft Way, Redmond, Washington 98052, and Department of Statistics, 367 Evans Hall, University of California, Berkeley, California 94720
Address at time of publication: Courant Institute, 251 Mercer Street, New York, New York 10012
Homepage
David B. Wilson
Microsoft Research, One Microsoft Way, Redmond, Washington 98052
Homepage
Additional Notes

Research of the first and third authors was supported in part by NSF grants DMS-0244479 and DMS-0104073.

Journal Information
Journal of the American Mathematical Society, Volume 22, Issue 1, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on and published on .
Copyright Information
Copyright 2008 by the authors. This paper or any part thereof may be reproduced for non-commercial purposes.
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/S0894-0347-08-00606-1
  • MathSciNet Review: 2449057
  • Show rawAMSref \bib{2449057}{article}{ author={Peres, Yuval}, author={Schramm, Oded}, author={Sheffield, Scott}, author={Wilson, David}, title={Tug-of-war and the infinity Laplacian}, journal={J. Amer. Math. Soc.}, volume={22}, number={1}, date={2009-01}, pages={167-210}, issn={0894-0347}, review={2449057}, doi={10.1090/S0894-0347-08-00606-1}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.