learning in network contexts: experimental results from …parkes/cs286r/spring06/papers/... ·...

44
Ž . Games and Economic Behavior 35, 80123 2001 doi:10.1006game.2000.0835, available online at http:www.idealibrary.com on Learning in Network Contexts: Experimental Results from Simulations Amy Greenwald 1,2 Department of Computer Science, Brown Uni ersity, Box 1910, Pro idence, Rhode Island 02912 E-mail: [email protected] Eric J. Friedman Department of Economics, Rutgers Uni ersity, New Brunswick, New Jersey 08903 E-mail: [email protected] and Scott Shenker 3 Xerox Palo Alto Research Center, 3333 Coyote Hill Road, Palo Alto, California 94304 E-mail: [email protected] Received January 21, 1999 This paper describes the results of simulation experiments performed on a suite of learning algorithms. We focus on games in network contexts. These are contexts Ž. Ž. in which 1 agents have very limited information about the game and 2 play can be extremely asynchronous. There are many proposed learning algorithms in the literature. We choose a small sampling of such algorithms and use numerical simulation to explore the nature of asymptotic play. In particular, we explore the extent to which the asymptotic play depends on three factors: limited informa- tion, asynchronous play, and the degree of responsiveness of the learning algo- rithm. Journal of Economic Literature Classification Numbers: C63, C72. 2001 Academic Press 1 This work was partially supported by an American Association of University Women Dissertation Fellowship. 2 The authors thank Dean Foster for extensive discussions and insightful suggestions. 3 Present address: International Computer Science Institute, 1947 Center Street, Suite 600, Berkeley, CA 94704. 80 0899-825601 $35.00 Copyright 2001 by Academic Press All rights of reproduction in any form reserved.

Upload: others

Post on 03-Sep-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

Ž .Games and Economic Behavior 35, 80�123 2001doi:10.1006�game.2000.0835, available online at http:��www.idealibrary.com on

Learning in Network Contexts:Experimental Results from Simulations

Amy Greenwald1,2

Department of Computer Science, Brown Uni�ersity, Box 1910, Pro�idence,Rhode Island 02912

E-mail: [email protected]

Eric J. Friedman

Department of Economics, Rutgers Uni�ersity, New Brunswick, New Jersey 08903E-mail: [email protected]

and

Scott Shenker 3

Xerox Palo Alto Research Center, 3333 Coyote Hill Road, Palo Alto, California 94304E-mail: [email protected]

Received January 21, 1999

This paper describes the results of simulation experiments performed on a suiteof learning algorithms. We focus on games in network contexts. These are contexts

Ž . Ž .in which 1 agents have very limited information about the game and 2 play canbe extremely asynchronous. There are many proposed learning algorithms in theliterature. We choose a small sampling of such algorithms and use numericalsimulation to explore the nature of asymptotic play. In particular, we explorethe extent to which the asymptotic play depends on three factors: limited informa-tion, asynchronous play, and the degree of responsiveness of the learning algo-rithm. Journal of Economic Literature Classification Numbers: C63, C72. � 2001

Academic Press

1 This work was partially supported by an American Association of University WomenDissertation Fellowship.

2 The authors thank Dean Foster for extensive discussions and insightful suggestions.3 Present address: International Computer Science Institute, 1947 Center Street, Suite 600,

Berkeley, CA 94704.

800899-8256�01 $35.00Copyright � 2001 by Academic PressAll rights of reproduction in any form reserved.

Page 2: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 81

1. INTRODUCTION

While much of classical game theory is based on the assumption ofcommon knowledge, there are many contexts in which this assumptiondoes not apply. Correspondingly, in recent years there has been increasedinterest in the process by which a set of initially naive agents learn throughrepeated play of a game. The central question concerns the nature ofasymptotic play; what set of strategies do the agents learn to play in the

Ž Ž .long-time limit? The review by Fudenberg and Levine 1998 provides an.overview of the literature.

In this paper we focus our attention on learning that occurs in what wecall a network context. In network contexts, agents interact through thecommon use of a resource, such as a communication link or a shareddatabase, which is accessed over a network. The interactions of Internetcongestion control algorithms where agents share network bandwidth, as

Ž .described in Shenker 1995 , is perhaps the most studied example of arepeated game in a network context. As the Internet continues to grow,and more resources are shared by remote users, we expect the networkcontext to become increasingly common. As discussed in Friedman and

Ž .Shenker 1997 , the network context differs from the traditional game-theoretic context in four important ways.

I. First, agents have very limited a priori information. In general,agents are not aware of the underlying characteristics of the sharedresource. In other words, they do not know the payoff structure of thegame; they know neither their own payoff function, nor that of the otherplayers. In addition, agents are not explicitly aware of the existence ofother players. While agents are aware that there may be other agentssimultaneously using the shared resource, they do not have any way ofdirectly observing their presence and, consequently, they know nothingabout the number or the characteristics of their opponents. In particular,this implies that players cannot observe the actions of other players, astandard assumption in many classical models of learning in economics.

II. Second, the payoff function and the agent population are subjectto change over time. Shared resources like network links and computerservers periodically crash and often experience other unpredictable changesin their capabilities, such as upgrades or route changes. In addition, usersof these shared resources come and go quite frequently. Thus, when anagent detects a change in his or her payoff while keeping his or her ownstrategy fixed, the agent cannot tell whether this change is due to changingstrategies of the other players, changes in the players themselves, orvariations in the characteristics of the shared resource.

Page 3: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER82

III. Third, in many, but not all, cases, learning is actually carried outby a computer algorithm, not by a human user. For instance, congestion

Ž .control algorithms e.g., TCP embedded in a computer’s operating systemcontrol the sharing of a network link. Similarly, automated algorithms cancontrol the retry behavior for query submission to a database. Thus, thelearning that takes place in these contexts is explicitly laid out in the formof a well-defined algorithm. Moreover, these algorithms are intended to bequite general in nature and do not depend on the detailed specifics of aparticular situation. This means that Bayesian learning algorithms areinappropriate, because the initial beliefs depend on the specific context. Inany event, the complexity of prior probabilities is such that it is notpossible to use Bayesian updating in any realistic setting.

IV. Fourth, in network contexts, games are often played in an asyn-chronous fashion. There need not be any notion of definable ‘‘rounds ofplay’’; users can update their strategies at any time. Moreover, the rates atwhich agents update their strategies can vary widely, although in generalthese rates are determined by circumstances and are usually not a strategicvariable. For example, automated agents can learn at different rates,depending on their processor speeds and the nature of their learningalgorithms. In addition, due to the geographic dispersion of users of theInternet, there can be varying communication delays to a shared resource,which can lead to updating rates that differ by several orders of magnitude.4

Thus, asynchrony does not arise from Stackelbergian-type strategic manip-ulation but merely from properties of communication and computation.Agents closer to the shared resource, or those who have faster processorsor smarter algorithms, have the potential to learn more rapidly andeffectively.

We focus on contexts that have these four properties: low informationcontent, nonstationary payoffs, automated learning, and asynchrony. Weare interested in what happens when a set of automated agents play agame repeatedly in such a context, and we investigate this behaviorempirically. We consider a small sampling of learning algorithms, some ofwhich have been well studied in the literature; for each algorithm wenumerically simulate a set of agents using that algorithm and we observethe set of strategies played in the long-time regime. Our simulationexperiments can be seen as a natural counterpart to human economic

Ž .experiments; in particular, Chen 1997 investigates some issues closely

4 Ž .As discussed in Blackwell 1956 , standard control theoretic results imply that thefrequency at which strategies are updated should not be greater than the inverse of theround-trip communication delay to the shared resource; otherwise, instability may result.

Page 4: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 83

related to those considered here using human subjects rather than auto-mated learning algorithms.5

We concentrate on the extent to which the asymptotic play depends onthe amount of information available to the agents, the degree of respon-siveness of the learning algorithm, and the level of asynchrony of play. Ofparticular interest is the extent to which the asymptotic play is containedin the various solution concepts such as Nash equilibria, the set of serially

Ž �.undominated strategies D , and the less traditional concepts of seriallyŽ �.unoverwhelmed strategies O and serially Stackelberg-undominated

Ž �.strategies S which are discussed below. Our findings suggest that theasymptotic play of games in network contexts can be quite different fromthat of standard contexts, where play is typically contained in the seriallyundominated strategy set.

This paper has six sections. Section 2 presents the learning algorithmsused in this study and discusses the relevant solution concepts. Section 3contains the simulation results for the sample games considered: severalsimple two-player games, a class of externality games, and the congestiongame. In Section 4, we compare these results with the results of simula-tions in non-network contexts. Section 5 describes related work, and weconclude in Section 6 with a brief discussion of our results. Finally,Appendix A describes the suite of learning algorithms that we simulate insome detail.

2. BACKGROUND

2.1. Learning Algorithms

The literature is replete with learning algorithms, but not all of them areapplicable in network contexts. Because knowledge of the payoff structureand the other agents is extremely limited, games in a network context are,from a single agent’s perspective, most naturally modeled as games against

Ž .nature in which each strategy has some random and possibly time-varyingpayoff about which the agent has no a priori knowledge. Consequently, in

Ž .contrast with belief-based approaches to learning e.g., Bayesian updatingadopted in much of the literature, learning algorithms for network con-texts typically utilize simple updating schemes that do not rely on anydetailed assumptions about the structure of the game. Instead, thesealgorithms employ ‘‘trial-and-error’’ experimentation in an attempt toidentify optimal strategies.

5 ŽAlthough our present focus is solely on automated agents, experimental evidence seeŽ . Ž . Ž ..Banos 1968 , Bernheim 1984 , and Chen 1997 suggests that our results are of relevance in

describing human�human and human�machine interactions as well.

Page 5: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER84

The learning algorithms which we simulate are distinguished first of allby their varying degrees of experimentation; we will, for convenience,

� .denote this level of experimentation by the parameter � � 0, 1 . In staticenvironments, where the payoff structure and the set and characteristics ofthe other agents is fixed, it may be reasonable to decrease the level ofexperimentation as time progresses, with experimentation ceasing in the

Ž . 6infinite-time limit i.e., � � 0 as t � � . Many learning algorithms pro-posed in the literature have this property. In network contexts, however,the environment is not static; the underlying payoff structure and thepopulation of agents are subject to change at any time without explicitnotification. As a result, agents should be prepared to respond to changingconditions at all times and should do so in a bounded amount of time. Thisrequires that a nonzero level of experimentation be maintained in thelong-time limit and that future play be more heavily influenced by payoffsobtained in the recent, rather than the distant, past. This second point can

Ž � Žbe achieved via a parameter � � 0, 1 which dictates the rate and typi-.cally inverse accuracy of learning. We call the ability to respond to

changes in the environment in bounded time responsi�eness and posit thatthis property is fundamental to learning in network contexts. As we shallsee, responsiveness has important implications for the resulting asymptoticplay.

The learning algorithms we discuss also differ in the particular criteriathat they are designed to satisfy. Perhaps the simplest criterion is that,when playing a static game-against-nature, the algorithm rapidly learns to

Ž . 7play with high probability the strategy with the highest average payoff.When combined with responsiveness, and a certain monotonicity property,this leads to the class of so-called reasonable learning algorithms intro-

Ž .duced in Friedman and Shenker 1997 . One example of such an algorithmis the stage learning algorithm. Stage learners partition the game intostages, which consist of 1�� rounds of a game. At each round of play, astage learner chooses its strategy at random based on the probabilities, orweights, it has assigned to each of its strategies. These weights are updatedupon termination of each stage, with weight 1 � � assigned to the purestrategy that obtained the highest average payoffs during the previous

Ž .stage, and weight �� n � 1 assigned to all other strategies. Anotherexample of a reasonable learning algorithm, the so-called responsi�e learn-

6 ŽThis is apparent in decision problems such as classic bandit problems Borgers and Sarin,.1995; Borodin and El-Yaniv, 1998 . It is less transparent, however, in games with multiple

players and strategies.7 The formal definition of probabilistic converge in finite time is described in Blackwell

Ž .1956 . In this paper we do not formally define convergence, but take a more pragmaticapproach which is appropriate for simulations. That is, we say that play has converged whenthe numerical properties are unchanged by additional iterations as evidenced by simulations.

Page 6: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 85

Ž .ing automata introduced in Friedman and Shenker 1996 , is a responsiveŽversion of simple learning automata see, for example, Narendra and

Ž ..Thathachar 1974 . This algorithm updates weights after every round ofplay using quite a different method. Another reasonable learning algo-

Ž .rithm for certain choices of parameters is defined in Erev and RothŽ .1996 and has been used to model human behavior in game-theoreticexperiments.

A second criterion, which is a worst-case measure of performance,involves the concept of regret. Intuitively, a sequence of plays is optimal ifthere is no regret for playing a given strategy sequence rather than playinganother possible sequence of strategies. Regret comes in two forms:external and internal. A sequence of plays is said to exhibit no externalregret if the difference between the cumulative payoffs that are achievedby the learner and those that could be achieved by any other pure strategyis insignificant. The no internal regret optimality criterion is a refinementof the no external regret criterion where the difference between theperformance of a learner’s strategies and any remapped sequence of thosestrategies is insignificant. By remapped we mean that there is a mapping fof the strategy space into itself such that for every occurrence of a given

Ž .strategy s in the original sequence the mapped strategy f s appears in theremapped sequence of strategies. The learning procedures described in

Ž . Ž .Foster and Vohra 1997 and Hart and Mas-Colell 1997 satisfy theproperty of no internal regret. Early no external regret algorithms were

Ž . Ž . Ž .discovered by Blackwell 1956 , Hannan 1957 , Banos 1968 , and MegiddoŽ . Ž .1980 ; recently, no external regret algorithms appeared in Cover 1991 ,

Ž . Ž .Freund and Schapire 1996 , and Auer et al. 1995 .We investigate six learning algorithms: the reasonable learners discussed

Ž Ž . Ž ..above see Friedman and Shenker 1996; 1997 and Erev and Roth 1996 ,Ž Ž .two based on external regret see Auer et al. 1995 and Foster and Vohra

Ž .. Ž Ž ..1993 , and one based on internal regret see Hart and Mas-Colell 1997 .Some of these algorithms were initially proposed for quite differentsettings, in which responsiveness is not necessary and the information level

Ž .is significantly higher e.g., agents know their own payoff function . Wehave extended these learning algorithms for use in network contexts.8 Wecall the versions designed for low-information settings nai�e and thosedesigned for higher information contexts informed. We also consider bothresponsi�e and nonresponsi�e variants. Lastly, each agent has a time-scaleparameter A that determines the rate at which it updates its strategies. A

Ž .player updates its strategy i.e., runs its learning algorithm only every Arounds and treats the average payoff during those A rounds as its actual

8 Detailed descriptions of these algorithms, in both their original and modified forms, arecontained in Appendix A.

Page 7: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER86

payoff. We are interested in how the asymptotic play depends on whetheragents are responsive, whether they are informed, and on the degree of

Ž .asynchrony differences in the A values among the agents.

2.2. Solution Concepts

To describe the asymptotic play, we introduce several solution concepts.We begin with some notation. We restrict our attention to finite games.

� 4Let NN � 1, . . . , N be a finite set of players, where N � NN is the numberof players. The finite set of strategies available to player i � NN is denotedby S , with element s � S . The set of pure strategy profiles is thei i iCartesian product S � Ł S . In addition, let S � Ł S with elementi i �i j� i j

Ž .s � S , and write s � s , s � S. The payoff function � : S � � for�i �i i �i iplayer i is a real-valued function on S.

Recall that strategy s � S is strictly dominated for player i if therei i� Ž . Ž � .exists some strategy s � S such that � s , s � � s , s for alli i i i �i i i �i

s � S . Let D� denote the serially undominated strategy set, i.e., the�i �iset of strategies that remains after the iterated elimination of strictly

Ž .dominated strategies. Milgrom and Roberts 1991 show that the asymp-totic play of a set of adapti�e learners�learners that eventually learn toplay only undominated strategies�eventually lies within D�. In addition,

Ž .it is shown in Friedman and Shenker 1996 that certain responsivelearners playing synchronously also converge to D�. The set D� is widelyconsidered to be an upper bound in terms of solution concepts; that is, it iscommonly held that the appropriate solution concept that arises vialearning through repeated play is a subset of the serially undominated set.9

This may indeed be true in standard game-theoretic contexts.Ž .In Friedman and Shenker 1996, 1997 , however, it is shown that in

network contexts, where there is potentially asynchrony and responsivelearning, play can asymptotically remain outside the serially undominatedset. A more appropriate solution concept for such settings is based on theconcept of o�erwhelmed strategies. We say that a strategy s � S is strictlyi i

� Ž .overwhelmed if there exists some strategy s � S such that � s , s �i i i i �iŽ � � . � �� s , s for all s , s � S . Let O denote the set of strategies thati i �i �i �i �i

remains after the iterated elimination of strictly overwhelmed strategies. ItŽ .is shown in Friedman and Shenker 1997 that the asymptotic play of a set

of reasonable learners lies within O�, regardless of the level of asynchrony.However, it is conjectured that O� is not a precise solution concept, onlyan upper bound.

9 Note that this also holds for one-shot games with common knowledge, as the set D�

Ž .contains all the rationalizable strategies Auer et al., 1995; Cover, 1991 .

Page 8: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 87

A refinement of O� called S� is introduced and formally defined inŽ . 10Friedman and Shenker 1997 . Because it is rather cumbersome, we do

not present the precise definition of S�. Intuitively, the set S� extends theset D� by allowing for the possibility that play is asynchronous rather thansynchronous as is standard in repeated game theory. More formally, the

� Ž .computation of S is as follows: pick a specific nonstrict ordering of theplayers and construct the extensive form game arising from players movingaccording to that order; now compute the set of actions which survive theiterated deletion of strictly dominated strategies in the new game; finally,take the union of these actions for all orderings. Since the ordering isnonstrict, asynchronicity defined in this way incorporates synchronicity,from which it follows that D� � S� � O�.

Ž .Another result of interest, due to Foster and Vohra 1997 , is that a setof no internal regret learners converges to a correlated equilibrium. Notethat the support of a set of correlated equilibria is a subset of D�; in otherwords, correlated equilibria do not assign positive probabilities to strate-gies outside D�, but neither do they necessarily converge to Nash equilib-ria. In contrast, the asymptotic play of a set of no external regret learners

� Ž . 11need not remain inside D Greenwald, 1999 .In the remainder of this paper, we present the result of simulations of

the six learning algorithms on various games. We ask, in particular,whether the asymptotic play converges within the sets D�, S�, or O�.Recall that the term convergence is used informally, both because ofexperimentation, which precludes true convergence, and our interest inachieving results in finite time. We are interested in which of theseconcepts, if any, represents an appropriate solution concept for games innetwork contexts.

3. SIMULATIONS IN NETWORK CONTEXTS

ŽWe consider three sets of games: simple games two players, few. Ž .strategies , the congestion game two players, many strategies , and an

Ž .externality game many players, two strategies . The simulations wereconducted with varying degrees of asynchrony, ranging from synchronous

Žplay to extreme asynchrony with one player acting as the leader i.e., wevary the value of A from 1 to 10,000 for the leading player and we set

10 Ž .In Blackwell 1956 , a class of ‘‘Stackelberg’’ solution concepts is built from variousŽ .primitives such as undominated strategies or correlated equilibria ; however, for the games

studied in this paper most of the ‘‘Stackelberg’’ solution concepts coincide.11 Note that this remark pertains only to the no external regret criterion but says nothing

about the convergence properties of specific algorithms which are defined to satisfy thiscriterion, such as those considered in this study.

Page 9: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER88

.A � 1 for all other players . The degree of responsiveness is determined byparameters � and � ; for each game, we describe the particular parametersettings.

3.1. Simple Two-Player Games

This subsection presents the results of simulations of four simple two-player games with either two or three strategies per player. The row playeris taken to be the leader. The responsive parameter � was set to 0.01 forall algorithms, while the degree of experimentation � was set to 0.025 forthe reasonable learning algorithms and 0.05 for the no regret algorithms.12

In addition, the no regret algorithms depend on tuning parameters; for themixing method, � � 1.5, for multiplicative updating, � � 1, and for the nointernal regret algorithm, � 2.13 Unless otherwise stated, the simulationsdescribed in this paper were run for 108 iterations; although play generallyconverged in far fewer iterations, this rather lengthy simulation timeeliminated the transient effects of initial conditions in the final long-runempirical frequency calculations. Initially, all strategies were assignedequal weights.

3.1.1. Game D

The game depicted in Fig. 1 is referred to herein as Game D since it isD-solvable, but it is not S-solvable or O-solvable: i.e., D� � S� � O�.More specifically, the set D� is a singleton that contains only the strategy

FIG. 1. Game D.

12 The choice of parameter values reflects the usual trade-off between exploration andexploitation. Increasing the rate of responsiveness � and the rate of experimentation � leadsto increased error in stationary environments but increased accuracy in nonstationaryenvironments. In our experience, the results are fairly robust to small changes in parameterssettings, although we have not formally measured this robustness.

13 These parameters represent the learning rates in the no regret algorithms. Slowerlearning rates correspond to higher degrees of accuracy in stationary environments; innonstationary environments, faster learning rates induce more responsive behavior.

Page 10: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 89

Ž .profile T , L , which is the unique Nash equilibrium. On the other hand,� � � 4 � 4 Ž .S � O � T , B � L, R . Note that B, R is a Stackelberg equilibrium

in which the row player is the leader.The graph depicted in Fig. 2 describes the overall results of simulations

of Game D, assuming responsive learning in a naive setting. In particular,Ž .Figure 2 a plots the percentage of time in which the Nash equilibrium

solution arises as the degree of asynchrony varies. Asynchrony of 100, forexample, implies that the column player is learning 100 times as fast as therow player; thus, the row player is viewed as the leader and the columnplayer the follower. Note that when play is synchronous, all the algorithmsconverge to the unique Nash solution. In the presence of sufficientasynchrony, however, play does not converge to the Nash solution for anyof the algorithms studied. Instead, play converges to the Stackelberg

Ž . �equilibrium, as depicted in Fig. 2 b . These results demonstrate that Ddoes not always contain the asymptotic play. Note that these results arerobust; in particular, the results are unchanged even when the game isstudied with ‘‘noisy’’ payoffs � , where � � � � , for small � 0.ˆ ˆi i i

The transition from Nash equilibrium to Stackelberg equilibrium de-picted in Fig. 2 is rather abrupt. This observation prompted us to conductfurther simulations at a series of intermediate values to more preciselydetermine the impact of asynchrony. For the reasonable learning algo-rithms, the transition between equilibria takes place when A falls between100 and 1000; for the no regret algorithms, this transition takes place whenA lies between 10 and 100. Figure 3 depicts the details of these transitionsin the respective ranges of asynchrony for the two sets of algorithms. Onlyreasonable learning algorithms ever clearly exhibit out-of-equilibrium be-havior; the no regret algorithms transition directly from one equilibrium tothe other.

Ž .Recall that B, R is the Stackelberg equilibrium in Game D. Figure 4plots the changing weights over time of strategy B for player 1 and strategyR for player 2 for the no external regret algorithm due to Foster andVohra. The individual plays are also plotted; marks at 100 signify play ofStackelberg strategies, and marks at 0 signify play of Nash strategies. Forcomparison purposes, the synchronous case, where play converges to Nashequilibrium, as well as the asynchronous case with A � 100, where playconverges to Stackelberg equilibrium, are depicted. Play converges in theformer case after roughly 500 iterations, while it converges in the latter

Žcase after only about 300 iterations for the given choices of parameters �. Ž . Ž .and � . In Fig. 4 b , the leader player 1 slowly learns to play the

Ž .Stackelberg solution, and because the follower player 2 is responsive, hisweights follow the leader’s plays. This behavior is representative of all thelearning algorithms considered.

Page 11: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER90

Ž .FIG. 2. Con�ergence to equilibria in Game D. a plots the percentage of time in which NashŽ .equilibrium arises as the degree of asynchrony varies, while b plots the percentage of time

in which Stackelberg equilibrium arises.

Page 12: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 91

Ž .FIG. 3. Detail of con�ergence to Stackelberg equilibrium in Game D. a Reasonable learningŽ .algorithms. b No regret learning algorithms.

Page 13: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER92

Ž .FIG. 4. Con�ergence to equilibria in Game D for Foster ’s and Vohra’s algorithm. a plots theweights of the Stackelberg equilibrium strategies over time when A � 1; notice that play

Ž .converges to Nash equilibrium. b plots these weights when A � 100; notice that playconverges to Stackelberg equilibrium.

Page 14: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 93

FIG. 5. Game O.

3.1.2. Game O

The next game that is studied in this section is depicted in Fig. 5 and isŽ .referred to as Game O, since it is O-solvable. In this game, T , L is the

�Ž .4 � � �unique Nash equilibrium and T , L � D � S � O . Simulations of allthe algorithms, for levels of asynchrony ranging from 1 to 10,000, show thatNash equilibrium is played over 95% of the time. In particular, play doesnot diverge from the Nash equilibrium solution in this O-solvable game, asit did in Game D, regardless of the degree of asynchrony. It has beenestablished that, for reasonable learners, O� is an upper bound on thesolution concept. The same result holds for the other algorithms consid-ered, but this observation is far short of a proof that the O� solutionconcept applies to them as well. The next game addresses the question ofwhether the O� solution concept might in fact be too large a set.

3.1.3. Prisoners’ Dilemma

This section presents the results of simulations of the repeated Prison-Ž . Ž . Žers’ Dilemma see Fig. 6 . In this game, D, D is the unique Nash and

. �Ž .4 � � � �Stackelberg equilibrium, and D, D � D � S � O , since O is theentire game. The Prisoner’s Dilemma provides a simple test of the conjec-ture that the outcome of responsive learning in network contexts is

FIG. 6. Prisoners’ Dilemma.

Page 15: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER94

FIG. 7. Game S.

described by the S� solution concept rather than the the larger solutionset O�.

Simulations of all the algorithms, for levels of asynchrony ranging from 1to 10,000, show that in this game the Nash equilibrium is played over 95%

Žof the time. Since play does not diverge significantly from the Nash and. �Stackelberg equilibrium, the asymptotic play is not spread throughout O ;

instead, it is confined to S�.

3.1.4. Game S

The last simple two-player game that is studied is a game in which theplayers have three strategies. The game is depicted in Fig. 7 and is referred

� �Ž .4 � � 4 � 4 �to as Game S. In Game S, D � T , L , S � T , L � B, R , and O isthe entire game; thus, D� � S� � O�. The results of simulations of GameS resemble the results of simulations of Game D.14 Figure 8 shows that thelearning algorithms do not converge to the Nash equilibrium solution ofthis game when there is asynchrony. Instead, play converges to the Stackel-berg equilibrium, as in Game D. This game provides a second test of theconjecture that the outcome of responsive learning in network settings is astrict subset of the O� solution.

3.2. Externality Games

To test whether similar results apply to games with more than twoplayers, we experiment with externality games. An externality game, as

Ž .defined in Friedman 1996 , is one in which each agent can choose eitherŽ .to participate or not to participate in some joint venture and where the

14 Note that in these simulations, the reasonable learning algorithms utilized � � 0.1667.

Page 16: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 95

Ž .FIG. 8. Con�ergence to equilibria in Game S. a plots the percentage of time in which NashŽ .equilibrium arises as the degree of asynchrony varies, while b plots the percentage of time

in which Stackelberg equilibrium arises.

Page 17: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER96

payoffs obtained by a given player depend only on whether that playerparticipates and on the total number of participating players. In thissection, we study a related class of games which are D-solvable, andmoreover, for certain choices of the parameters, the games in this class areS-solvable and O-solvable as well.

Ž .The class of games which we consider class EG is a discretization ofŽ .the nonatomic games discussed in Friedman 1998 . The set of players

� 4NN � 0, . . . , N � 1 , with i � NN. The players have two possible strategies,namely 0 and 1, where 1 corresponds to participation, and 0 corresponds to

Ž .nonparticipation. The number of participants, therefore, is given by � s �Ý s , where s denotes the strategic choice of player i. Payoffs arei� NN i idetermined as follows. The value to player i of participation is � � �, andi

Ž .the cost of participation is C � , where C is a nondecreasing function ofi iŽ .the externality. Thus, if player i participates, then s � 1 and � 1, s �i i �i

Ž Ž ..� � C � s . Otherwise, if player i does not participate, then s � 0, andi i iŽ . Ž . � .� s � �� 1, s , for � � 0, 1 . Intuitively, � measures the extent toi i �i

which players can opt out.Note that the parameter � does not affect the standard strategic

elements of a given game in this class, such as best-replies or dominatedstrategies. In particular, if the game is D-solvable for � � 0 then it is

� .D-solvable for all � � 0, 1 . Similarly, varying � does not change the setof Nash equilibria. Moreover, it is straightforward to show that when

Ž� � 0, if the game is D-solvable, then it must also be O-solvable and.therefore, also S-solvable . In contrast, for � sufficiently close to 1, the

Ž .game is not S-solvable and therefore, not O-solvable . Thus, by varying �we can create a class of games which are D-solvable but not necessarilyS-solvable and not O-solvable.15

Ž � 4.In our simulations, we consider eight players i.e., NN � 0, . . . , 7 , andŽ .we set � � i and C � � �� , for � �. In the first set of simulations,i i

we choose � 1.9; we call this Game EG . This game is D-solvable and1.9therefore has a unique Nash equilibrium. Moreover, this implies that for� � 0, this game must also be O-solvable; however, for � sufficiently close

Ž .to 1, it is neither S-solvable nor O-solvable see Appendix B . Morespecifically, when � � 6�11 � 0.54, Game EG has a Stackelberg equi-1.9librium with player 2 as the leader, which differs from the Nash equilib-

� . Ž .rium: the Nash equilibrium for all � � 0, 1 is s � 0, 0, 0, 1, 1, 1, 1, 1 ,Žwhile the Stackelberg equilibrium with player 2 as the leader and � �

. Ž .6�11 is s � 0, 0, 1, 0, 1, 1, 1, 1 .

15 Proofs of these claims appear in Appendix B for the class of games considered.

Page 18: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 97

TABLE IGame EG with � � 0.002, A � 50001.9

Algorithm � � 0 � � 0.5 � � 0.6 � � 0.9

Foster and Vohra NE NE SE SEFreund and Schapire NE NE SE SEHart and Mas-Colell NE NE SE SE

Simulations of Game EG were conducted using the naive, responsive1.9variants of the no regret learning algorithms.16 The convergence results in

Ž .the asynchronous case for A � 5000 are listed in Table I, for � � 0.0217 Žand � � 0.002. Simulations of all algorithms show rapid convergence in

. Ž .the empirical sense described earlier to the Nash equilibrium NE for allvalues of � in the synchronous case and to the Stackelberg equilibriumŽ .SE when � � 0.6 and � � 0.9 in the asynchronous case. In particular, inGame EG , for certain choices of the parameters, the asymptotic play is1.9not contained in the set D�.

Another game which we simulated in the class EG is Game EG , which2.1is identical to Game EG , except that � 2.1. Like Game EG this1.9 1.9game is D-solvable. This implies that for � � 0, this game must also beO-solvable; however, for � sufficiently close to 1, this game is not O-solva-

Ž .ble see Appendix B . Lastly, unlike Game EG , Game EG , is S-solva-1.9 2.1Ž .ble see Appendix B . This game was simulated assuming the same

parameter values as the previous set of simulations, as well as someadditional choices for � . Selected results of these simulations appear inTable II. Notice that regardless of the values of � and �, and even in the

TABLE II� 4Game EG with � � 0.01, 0.005, 0.002, 0.001 , A � 10,0002.1

Algorithm � � 0 � � 0.5 � � 0.6 � � 0.9

Foster and Vohra NE NE NE NEFreund and Schapire NE NE NE NEHart and Mas-Colell NE NE NE NE

16 The payoffs were translated by N� in simulations of responsive learning automata andthe algorithm due to Roth and Erev in order to avoid negative payoffs.

17 In the algorithm due to Foster and Vohra, � � 1,000; in the algorithm due to Freundand Schapire, � � 1; finally, in the algorithm due to Hart and Mas-Colell, � 5.

Page 19: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER98

presence of extreme asynchrony, play converges to Nash equilibrium.Thus, as D� � S� � O�, this game provides further evidence for theconjecture that the asymptotic play of learning in network contexts iscontained in S�.

3.3. The Congestion Game

So far we have considered games with rather small strategy spaces. Inthis section, we experiment with a larger strategy space, using an examplethat arises in computer networks. Consider several agents simultaneouslysharing a network link, where each agent controls the rate at which he orshe is transmitting data. If the sum of the transmission rates is greaterthan the total link capacity, then the link becomes congested and theagents’ packets experience high delays and high loss rates. The transmis-sion rates are controlled by each agent’s congestion control algorithm,which varies the rates in response to the level of congestion detected.

One can model the interaction of congestion control algorithms as acost-sharing game where the cost to be shared is the congestion experi-enced. That is, we can model this as a game where the strategies are the

Ž .transmission rates r and the outcomes are the pairs r , c , where c is thei i i i

congestion experienced as function of the strategy profile r. The alloca-Ž .tions must obey the sum rule Ý r � F Ý c where F is a constrainti i i i

Žfunction i.e., the total congestion experienced must be a function of the.total load .

Most current Internet routers use a FIFO packet scheduling algorithm,which results in congestion that is proportional to the transmission rate:

� � Ž . Ž .i.e., c � r �Ý r F Ý c ; this is called the average cost pricing ACPi i j j j jmechanism. In contrast, the fair queuing packet scheduling algorithm canbe modeled as leading to allocations such that c is independent of r asi j

Žlong as r r this condition, plus anonymity, uniquely specifies thej i. Ž Ž .allocations . This is called the serial mechanism see Shenker 1995 for a

.detailed description .Ž .Chen 1997 investigates the two-player congestion game with the fol-

Ž . Ž . Ž .lowing properties: 1 linear utilities U r , c � � r � c , 2 quadratici i i i i iŽ . 2 Ž . � 4congestion F x � x , and 3 a discrete strategy space S � 1, 2, . . . , 12 .i

For the choice of parameters � � 16.1 and � � 20.1, the game defined1 2by the ACP mechanism is D-solvable, but it is not S-solvable or O-solva-

Ž .ble. The unique Nash equilibrium is 4, 8 and the Stackelberg equilibriumŽ .with player 2 leading is 2, 12 . In contrast, the game defined by the serial

Ž .mechanism is O-solvable, with the unique Nash equilibrium at 4, 6 .We conducted simulations of both the serial and the ACP mechanism

using the naive, responsive variants of the no regret algorithms with degree

Page 20: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 99

of experimentation � � 0.02.18,19 In our simulations of the serial mecha-nism, the learning algorithms concentrate their play around the Nashequilibrium. In the ACP mechanism, when play is synchronous, the asymp-totic behavior again centers around the Nash equilibrium. Given sufficient

Ž .asynchrony e.g., A � 5000, when � � 0.002 , however, play converges tothe Stackelberg equilibrium.

3.4. Discussion

Our results thus far indicate that when responsive learning algorithmsplay repeated games against one another, play can persist outside the setof serially undominated strategies, given sufficient asynchrony. It comes asno surprise that asynchrony can lead to play outside Nash or correlatedequilibria. It is known, for example, that Stackelbergian behavior arises in

Ž .games among ‘‘patient players’’ Fudenberg and Levine, 1989 , or whenŽ .there exists the ability to make commitments Rosenthal, 1991 or the

Ž .capacity to establish reputations Watson, 1993 . In these prior analyses,the Stackelberg leaders are seen as manipulating the system. The asyn-chrony that we discuss here comes from a quite different source: theunderlying speed of communication and computation of the agents.

In our examples, the trajectory of play is always either largely inside theserially undominated set or, with sufficient asynchrony, it converges to theStackelberg equilibrium. The transition from one behavior to the other isgenerally quite sudden, although some of the reasonable learning algo-

Ž .rithms raise exception see Fig. 3 . Based on work by Foster and VohraŽ .1997 , we conjecture that more complicated behavior, with probabilitiesspread over a wider set of strategy profiles, can arise in more complexgames but always remains within the set S�.

4. SIMULATIONS IN NON-NETWORK CONTEXTS

Network contexts differ from standard learning contexts in three impor-tant ways: asynchronous play, limited information, and responsive learning.In the previous sections we have looked at how varying asynchrony affectsthe convergence properties of learners. In this section, we briefly considerthe remaining two properties of learning in network contexts.

First we augment the information structure by considering contexts inwhich learners are informed, where such learners know the payoffs that

18 In the algorithm due to Foster and Vohra, � � 5,000; in the algorithm due to Freundand Schapire, � � 1; finally, in the algorithm due to Hart and Mas-Colell, � 2,000.

19 The payoffs were translated to � in simulations of responsive learning automata andthe algorithm due to Roth and Erev in order to avoid negative payoffs.

Page 21: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER100

would have occurred had they chosen an alternative action. This typicallyŽ . Ž .arises when players i know the payoff matrix and ii can observe the

actions of the other players. Not surprisingly, in this setting play outsideD� does not arise. Unfortunately, in network contexts, informed learning isnot an option; the basic structure of the Internet is such that learning isinherently uninformed.

Our second consideration, namely responsiveness, is not inevitable, butŽ .instead reflects a common and appropriate design choice on the Internet.

We find the behavior of naive and nonresponsive learners, in the presenceof asynchrony, to be complex and do not claim to understand suchasymptotic behavior; for example, we do not know whether play alwaysconverges to D�. We demonstrate some of this complexity through simula-tions of the Shapley game, a classic example for which fictitious play doesnot converge. Irrespective of these complexities, nonresponsive learning isnot viable in network contexts due to the nonstationarity of the payofffunctions.

The results of this section show that the seemingly obvious conjecturethat asynchrony alone leads to Stackelberg behavior is not true in general.In our simulations, this conjecture holds only when we consider naive andresponsive learning algorithms, as are relevant for network contexts. If thealgorithms are informed, or nonresponsive, we do not observe Stackelber-gian behavior.

4.1. Informed Learning

Recall that the simulations that lead to asymptotic play outside D�

utilize the responsive and naive variants of the set of learning algorithms.Ž .In our simulations of Game D see Fig. 1 , responsive but informed

learning does not lead to play outside D�, even in the presence of extremeasynchrony, for enhanced versions of reasonable learning that utilize fullpayoff information and the original no regret algorithms. Specifically, forlevels of asynchrony between 1 and 10,000, simulations of all responsiveand informed algorithms show that Nash equilibrium is played over 95% ofthe time.20

Intuitively this occurs because the set of informed learning algorithmscompares the current payoff with the potential payoffs of the otherpossible strategies, assuming that the other agents keep their strategies fixed.The key to the Stackelberg solution is that the leader evaluates his or herpayoff in light of the probable responses of other agents. Naive learners,

20 The simulations of Game D discussed in this section and the next depend on the sameset of parameter values as in Section 3.1; specifically, � � 0.01 for all algorithms, while� � 0.025 for the reasonable learning algorithms and � � 0.05 for the no regret algorithms.

Page 22: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 101

when learning at a slow rate, do this implicitly; that is, they only receivetheir payoffs after the other players respond to their play. The informedlearning algorithms which we consider do not take this reaction intoaccount.

If informed and responsive learning algorithms do indeed converge toD� in general, this might be seen as an argument to consider onlyinformed learning algorithms. However, in network contexts this is not anoption; the information about other payoffs is not available and so we areforced to use naive learning algorithms. However, agents do have a choiceas whether to use responsive or nonresponsive algorithms, a subject towhich we now turn.

4.2. Nonresponsi�e Learning

We now consider, as a benchmark, nonresponsive learners. SimulationsŽ .of Game D see Fig. 1 using nonresponsive algorithms for levels of

asynchrony between 1 and 10,000 show that the Nash equilibrium is playedover 99% of the time for the set of informed algorithms and over 95% ofthe time for the naive set. Figure 9 depicts the weight over time of strategyB for player 1 and strategy R for player 2 in simulations of the no externalregret learning algorithm due to Foster and Vohra with level of asyn-chrony 100.

The behavior of informed and nonresponsive learners is similar to thatof informed and responsive learners, in that they learn to eliminate

� Ž Ž ..dominated strategies, yielding convergence to D see Fig. 9 a . Thisseems reasonable given that the informed, nonresponsive algorithms whichwe simulate are approximately adaptive learners in the sense of Milgrom

Ž . �and Roberts 1991 , who prove that such adaptive learners converge to D .Ž Ž ..In the simulation of naive, nonresponsive learners see Fig. 9 b , play

initiates at the Stackelberg equilibrium, until time almost 3000, when theleader experiments with the Nash equilibrium strategy. Since he or she is

Ž .followed shortly thereafter albeit slowly by the follower, he or she neverŽhas an incentive to return to the Stackelberg solution. Eventually near

.time 8000 , the follower’s strategy converges to that of Nash as well.While in our simulations of relatively simple games the asymptotic play

of nonresponsive learning algorithms is confined to D�, we have no proofthat nonresponsive, naive learning algorithms in general remain inside D�;in fact, the authors are divided as to whether this result holds in thepresence of asynchrony. This subject warrants further study. From theperspective of network contexts, however, the analysis of such questions ismoot, since nonresponsive learners are unsatisfactory for a much simplerreason: their performance is suboptimal in nonstationary settings.

Page 23: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER102

FIG. 9. Asynchronous, nonresponsi�e learning in Game D �ia Foster ’s and Vohra’s algo-Ž .rithm. a plots the Stackelberg equilibrium strategy weights in the informed case; notice that

Ž .play quickly converges to Nash equilibrium. b plots these weights in the naive case; noticethat play again converges to Nash equilibrium.

Page 24: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 103

4.2.1. Nonstationary En�ironments

Consider a simple, one-player, two-strategy game where the payoffsinitially are 1 for strategy A and 0 for strategy B. We simulate a nonsta-tionary version of this game where the payoffs are reversed every 5000rounds.

Ž . Ž .Figures 10 a and 11 a plot the cumulative percentage of time spentplaying the optimal strategy in simulations of sample reasonable learningalgorithms.21 All the reasonable learning algorithms�namely stage learn-ing, responsive learning automata, and the algorithm due to Roth andErev�spend over 90% of their time at the current optimal strategy in thesimulated quasi-static environment. In addition, the resulting fluctuations

Ž .in the weight of strategy A in this game are depicted in Figs. 10 b andŽ .11 b ; observe that the weight of strategy A changes with the state of the

environment.In contrast to the reasonable learning algorithms, the nonresponsive, no

Ž .regret algorithms both the naive and informed versions perform poorly innonstationary environments. Figure 12 plots the cumulative percentage oftime spent playing the Nash equilibrium for Freund’s and Schapire’s noexternal regret algorithm, in both its responsive and nonresponsive forms.22

Note that the nonresponsive version of the algorithm spends only about50% of its time playing the currently optimal strategy. This behavior isrepresentative of all the no regret algorithms studied. This is because thenonresponsive no regret algorithms fixate on one strategy early on�theone that is initially optimal�and are unable to adjust to future changes inenvironmental conditions.

Thus, the criterion of no regret, while perhaps appropriate for learningin static environments, is apparently not sufficient for learning nonstation-ary payoff functions. Since network contexts are typically nonstationaryand since this nonstationarity can be detected only via experimentationŽ .one cannot observe the change in the structure of the game directly , thelearning algorithms employed should be responsive.

4.2.2. Shapley Game

In this section, we compare the behavior of responsive and nonrespon-Ž .sive learners in the Shapley game see Fig. 13 , a well-known game in

which fictitious play, an informed and nonresponsive algorithm, does notconverge. In this game, fictitious play results in cycling through the cells

21 The algorithmic parameters for the reasonable learning algorithms were chosen asfollows: � � � � 0.01.

22 For simulation purposes, in the algorithm due to Freund and Schapire, � � 1, and in theresponsive case, � was set equal to 0.01.

Page 25: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER104

Ž .FIG. 10. Responsi�e LA in quasi-static en�ironment. a Percentage time at Nash equilib-Ž .rium. b Weight of strategy A.

Page 26: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 105

Ž .FIG. 11. Roth’s and Ere� ’s algorithm in quasi-static en�ironment. a Percentage time atŽ .Nash equilibrium. b Weight of strategy A.

Page 27: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER106

FIG. 12. Hart’s and Mas-Colell’s algorithm in quasi-static en�ironment. Responsive vs.Ž . Ž .nonresponsive learning. a Informed version. b Naive version.

Page 28: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 107

FIG. 13. Shapley game.

with 1’s in them, namely cells 1, 2, 5, 6, 7, and 9 in Fig. 13, withŽ .ever-increasing lengths of play in each of these cells Shapley, 1953 . One

is led to conjecture that this fascinating behavior arises because of theclear-cut choices made by fictitious play�with probability 1, the strategywith the highest expected payoff is played, leading to abrupt transitions inthe trajectory of play.

Surprisingly, in our simulations,23 we observe behavior similar to that offictitious play for the nonresponsive learning algorithms24 �both informedŽ Ž . Ž .. Ž Ž . Ž ..see Figs. 14 a and 15 a and naive see Fig. 14 b and 15 b �eventhough these algorithms do not in general induce discrete changes. Figure14 plots the cumulative percentage of time that player 1 plays each of thethree strategies; although not depicted here, the behavior patterns ofplayer 2 are identical. In addition, Fig. 15 depicts the joint empiricalfrequencies of the various strategy profiles after 106 iterations, where thex-axis is labeled 1, . . . , 9 corresponding to the cells in Fig. 13. Via this jointperspective, we see that both informed and naive nonresponsive learnersspend very little time playing in cells without a 1 in them. Specifically, inthe informed case, the likelihood of play in cells 3, 4, and 8 approaches 0,and in the naive case this likelihood approaches ��N.

In contrast, the responsive algorithms, while they do display the samecycling behavior, the duration of play in each cell does not continue togrow. Instead the responsive algorithms spend equal amounts of time in

Ž . Žeach of the distinguished cells. This is depicted in Fig. 16 a the informed. Ž . Ž .case and Fig. 16 b the naive case , which plot the cumulative percentage

23 The graphs present the results of simulations of the algorithm due to Freund andSchapire.

24 The only exception is the algorithm of Hart and Mas-Colell which is known to convergeto correlated equilibrium, and in fact converges to the mixed strategy Nash equilibrium in theShapley game.

Page 29: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER108

FIG. 14. Nonresponsi�e learning in the Shapley game. The cumulative percentage of timeŽ .player 1 plays each of his or her three strategies assuming nonresponsive learning. a

Ž .Informed, nonresponsive. b Naive, nonresponsive.

Page 30: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 109

FIG. 15. Joint empirical frequencies of strategy profiles in the Shapley game assuming nonre-Ž .sponsi�e learning. The x-axis is labeled 1, . . . , 9, which corresponds to the cells in Fig. 13. a

Ž .Informed, nonresponsive. b Naive, nonresponsive.

Page 31: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER110

FIG. 16. Responsi�e learning in Shapley game. Cumulative percentage of time player 1 playsŽ .each of his or her three strategies assuming responsive learning. a Informed, responsive.

Ž .b Naive, responsive.

Page 32: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 111

of time player 1 spends playing each of the three strategies. Notice thatthese graphs converge to 33% for all strategies. In addition, Fig. 17 depictsthe joint empirical frequencies of the various strategy profiles after 106

iterations. One interesting feature of the set of responsive learning algo-rithms is that their empirical frequencies converge to that of the fair andPareto optimal correlated equilibrium in the Shapley game; in particular,both players have expected average payoff of 1�2. This follows from thebounded memory of these algorithms.

5. RELATED WORK

There is a vast body of economics literature on learning throughrepeated play of games, and we make no attempt here to provide adetailed review; for a comprehensive discussion, see the review by Fuden-

Ž .berg and Levine 1998 . There is also substantial interest within theartificial intelligence community in the area of multiagent learning; see therecent special issue of Machine Learning on this topic. In this section, weplace our work in the context of the varying approaches to learning takenby economics and artificial intelligence researchers.

5.1. Rele�ance to Economics

The work on learning in economics falls roughly into two camps. The‘‘high-rationality’’ approach involves learning algorithms that aim to pre-dict their opponents’ strategies and then optimize with respect to those

Žpredictions. Prediction methods can be Bayesian as in Kalai and LehrerŽ .. Ž Ž .. Ž1993 , calibrated as in Foster and Vohra 1997 , or consistent as in

Ž .Fudenberg and Levine 1995b, 1998 , just to name a few examples. Typi-cally, the asymptotic play of high-rationality learning is either a correlatedor Nash equilibrium. Since these algorithms depend on knowledge of theunderlying structure of the game, however, they are not applicable innetwork contexts, which are of interest in this study.

In contrast, the ‘‘low-rationality’’ approaches to learning are concernedwith situations similar to that which concerns us here, where agents haveno information about the game other than the payoffs they receive.

Ž .Examples of such work include Roth and Erev 1995 , Erev and RothŽ . Ž . Ž .1996 , Borgers and Sarin 1995 , Mookerji and Sopher 1996 , and Van

Ž .Huyck et al. 1996 ; most of these algorithms as they were initiallyproposed are not responsive, but as we show in Appendix A, they can bemade responsive via slight modifications. The focus of these papers istypically on matching the results of human experiments, whereas we focusinstead on the nature of the asymptotic play. Chen however, performed

Page 33: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER112

FIG. 17. Joint empirical frequencies of strategy profiles in the Shapley game assuming respon-si�e learning. The x-axis is labeled 1, . . . , 9, which corresponds to the cells in Fig. 13.Ž . Ž .a Informed, responsive. b Naive, responsive.

Page 34: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 113

experiments on the congestion game discussed in Section 3.3 in which shecompared synchronous and asynchronous play as well as learning in full

Žinformation one-shot games zipper design, where play is repeated, but is.against different opponents versus naive learning in repeated games.

5.2. Rele�ance to Artificial Intelligence

In the context of artificial intelligence research, the present study isrelated to work in machine learning, where out sample of algorithms wouldbe classified as reinforcement learning schemes. Unlike machine learning

Ž Ž ..in multiagent systems see Stone and Veloso 1997 �systems designed tocoordinate the interactions of albeit independently motivated but centrallydesigned agents�we are interested in multiagent learning in networkcontexts, where agent design is outside the realm of system design. Recallthat network contexts are characterized by naivete regarding payoff infor-´mation, nonstationary environments, and asynchronous interactions.

ŽTraditionally, reinforcement learning algorithms see textbook by SuttonŽ . Ž ..and Barto 1998 and survey by Kaelbling et al. 1996 were designed for

single agent learning in stationary environments where payoffs are deter-mined by state. Like learning in network contexts, reinforcement learningis naive in the sense that full payoff information is not assumed to beavailable to the learner; however, since reinforcement learning algorithmsare not designed for use in nonstationary environments, they aim tomaximize the discounted sum of expected future returns, rather thanexhibit myopic behavior as do the algorithms considered here. The theoret-ical bounds that are known for reinforcement learning algorithms such as

Ž .Q-learning Watkins, 1989 do not apply in the nonstationary settings thatare typical of network contexts.

Recently, reinforcement learning researchers have expanded their inter-ests to multiagent learning. Such learning, however, can be tricky simplybecause as agents learn, their actions change, which implies that theirimpact on the environment changes, often leading to nonstationarities. Forsimplicity, most studies in multiagent learning focus on settings where

Žpayoffs are either negatively correlated, as in zero-sum games see, forŽ ..example, Littman 1994 , or positively correlated, as in coordination

Ž Ž ..games see, for example, Shoham and Tennenholtz 1993 . Two notableŽ .exceptions include Wellman and Hu 1998 , who conducted a theoretical

investigation of multiagent learning in market interactions, and SandholmŽ .and Crites 1995 , who conducted empirical studies of multiagent rein-

forcement learning in the Prisoners’ Dilemma. Similarly, this work consid-ers multiagent learning in positive sum games.

The results described here on convergence to Stackelberg equilibria,rather than traditional game-theoretic solutions such as D�, were largely a

Page 35: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER114

result of the asynchronous nature of play. Likewise, Huberman andŽ .Glance 1993 simulate the evolutionary dynamics of both synchronous and

asynchronous updating in the repeated Prisoners’ Dilemma and find thatrather different behaviors arise. Perhaps of closer interest to the presentstudy, empirical investigations of asynchronous Internet interactions were

Ž .reported by Lukose and Huberman 1997 , but their studies concern timecorrelations arising from such behavior rather than properties of conver-gence. The convergence of asynchronous machine learning algorithms,however, where players’ moves are determined by discrete random pro-cesses, has been investigated recently in e-commerce pricing models by

Ž .Greenwald and Kephart 2000 .Finally, regarding theoretical computer science, the algorithms which we

refer to as satisfying no regret optimality criteria are also often describedŽ .as achieving reasonable competitive ratios. Borodin and El-Yaniv 1998

present a thorough discussion of the competitive analysis of on-line algo-rithms.

6. CONCLUSION

This paper presents the results of experiments conducted using sixlearning algorithms, embodying three distinct notions of optimality: one

Ž .average-case performance measure reasonable learning and two worst-Ž .case performance measures no external and no internal regret learning .

In the suite of relatively simple games which were examined here, all thealgorithms exhibited qualitatively similar behavior. Thus, it seems that innetwork contexts, the key property is not which type of optimality isachieved but, rather, responsiveness.

In low-information settings, where learners are necessarily naive andthus cannot detect changes in the structure of the game directly, algo-rithms should be able to respond to environmental changes in boundedtime. It is shown in this paper that when such naive and responsivelearning algorithms operate in asynchronous environments, the asymptoticplay need not lie within D�. The question of whether play outside D�

arises for naive but nonresponsive players in asynchronous environmentsremains open but presumably would only arise in games with more playersand larger strategy sets than have been studied in this paper.

It has been established previously that for reasonable learning algo-rithms the asymptotic play is contained within O�, and it was furtherconjectured that such play is contained within the smaller set S�. Ourexperimental results are consistent with this conjecture. While theseresults are suggestive, they are in no way conclusive, and so we are left

Page 36: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 115

with the open question of what the appropriate solution concept is fornaive and responsive learning algorithms in asynchronous settings.

APPENDIX A: LEARNING ALGORITHMS

For completeness, this appendix describes the learning algorithms whichwere simulated in this study: responsive learning automata, due to Fried-

Ž .man and Shenker 1996 , responsive learning via additive updating, due toŽ .Erev and Roth 1996 , multiplicative updating, due to Freund and Schapire

Ž . Ž .1996 , the mixing method, due to Foster and Vohra 1993 , and finally,Ž .learning via no internal regret, due originally to Foster and Vohra 1997

Ž .and simplified by Hart and Mas-Colell 1997 . The algorithms are pre-sented in several varieties, depending upon whether they are applicable ininformed or naive settings and whether or not they are responsive. Ininformed settings, the counterfactual is known�in other words, playersare informed of the potential payoff which they might have achieved hadthey employed any of their other strategies�while in naive settings, theonly information pertaining to the payoff structure that is available at agiven time is the payoff of the strategy that is in fact employed at thattime. Informally, an algorithm is responsive if it weights recent informationmore heavily than past information and has a nonzero lower bound on thelevel of experimentation.25

A.1. Notation

The algorithms in this appendix are described from the point of view ofthe single player engaging in a game against nature. Note that this point ofview gives rise to notation which deviates from that of the main paper. The

� 4player chooses strategic actions from a strategy set N � 1, . . . , n , forn � �. Let i, j � N. The payoff awarded at time t by employing strategy iis denoted � t and, in general, the payoff function at time t is given by � t:i

� �N � a, b , for a, b � �. Note that payoffs are assumed to be bounded.For all the types of learning considered in this paper, the algorithm thatdetermines strategic choices is described in terms of a time-dependent

t Ž t t t . tvector of probabilities, namely p � p , . . . , p , . . . , p , where p denotes1 i n ithe probability of playing strategy i at time t.

A.2. Responsi�e Learning Automata

Ž .Responsive learning automata RLAs are introduced in Friedman andŽ .Shenker 1996 . This algorithm is designed for network contexts; it is

25 Ž .The formal definition of responsiveness is given in Blackwell 1956 .

Page 37: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER116

applicable in naive settings and it is responsive, as the name suggests. It isknown that RLAs are reasonable learners, and therefore, RLAs converge

� Ž .to O Friedman and Shenker, 1997 ; moreover, in the case of syn-� Ž .chronous play, RLAs converge to D Friedman and Shenker, 1996 .

ŽRLAs are a responsive extension of a simple learning automaton for aŽ .survey of the literature, see Narendra and Thathachar 1974 . This simple

learning automata is constructed by letting p0 � 1�n and using the updateirule when strategy i is played at time t,

pt1 � pt �� t 1 � pt 1Ž .Ž .i i i i

pt1 � pt 1 � �� t � j � i � N , 2Ž .Ž .j j i

where � is a parameter that controls the trade-off between learningŽ . Ž .rapidly for � close to 1 and accuracy for � close to 0 . This learning

automaton is not responsive. RLAs achieve responsiveness simply byimposing the requirement that the probability of any strategy never fall

Ž .below �� n � 1 according to the update rules

pt1 � pt �� t at pt 3Ž .Ýi i i j jj�i

pt1 � pt 1 � �� tat � j � i � N , 4Ž .Ž .j j i j

where

pt � �� n � 1Ž .jta � min 1, .j t t½ 5�� pi j

The following two sections describe alternative learning mechanismsŽ .based on additive updating due to Erev and Roth 1996 and Foster and

Ž .Vohra 1997 .

A.3. Simple Additi�e Updating

A second example of learning via additive updating is the naive learningŽ .algorithm of Erev and Roth 1996 which is reasonable, for certain choices

of the parameters. The updating scheme is as follows: pt � qt�Ý qt. Ifi i j� N jstrategy i is played at time t then the q ’s are updated asi

q t1 � 1 � � qt � t 1 � � 5Ž . Ž . Ž .i i i

q t1 � 1 � � qt � t �� n � 1 � j � i � N , 6Ž . Ž . Ž .Ž .j j i

where � behaves as in the RLA. It is straightforward, although quitetedious, to show that this algorithm is reasonable.

Page 38: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 117

The following sections describe additive and multiplicative learning rulesthat satisfy the no external regret optimality criterion.

A.4. Additi�e Updating Re�isited

This section presents an additive updating rule due to Foster and VohraŽ .1993 , known as the mixing method, which exhibits no external regret. Themixing method was originally designed for use in informed settings andconsequently updates weights according to the cumulative payoffs achievedby all strategies, including the payoffs that would have been obtained bystrategies which were not played. The cumulative payoff obtained at time t

Ž t. t t xby strategy i notation � is computed as follows: � � Ý � .i i x�0 iIn the case of two strategies, say A and B, the mixing method updates

the weight of strategy i as

1 � � t � � tŽ .A Bt1w � min max , 0 , 1 , 7Ž .A t t½ 5½ 52 2 � �Ž .A B

where the truncations ensure that the probability assigned to strategy A isŽ .between 0 and 1. It is shown in Foster and Vohra 1993 that the optimal

'value of � is T , where T is the number of iterations at which point themixing method exhibits no external regret. In general, when the number ofstrategies n � 2, the algorithm utilizes pairwise mixing of strategies via Eq.Ž .7 , followed by further mixing of the mixtures.

The mixing method can be modified for use in naive settings by utilizingan estimate of cumulative payoffs that depends only on the payoffsobtained by the strategies that are actually employed and the weights

t Ž t t. t tassociated with those strategies. In particular, let � � I �p � where Iˆ i i i i iis the indicator function: I t � 1 if strategy i is played at time t, and I t � 0i iotherwise. In other words, � t is equal to 0 if strategy i is not employed atˆ itime t; otherwise, � t is the payoff achieved by strategy i at time t scaledˆ iby the likelihood of playing strategy i. Now estimated cumulative payoffsŽ t. t t xnotation p are given by � � Ý � . The naive variant of the mixingˆ ˆ ˆi i x�0 i

t t Ž .method uses � in place of � in Eq. 7 . This update procedure yields ai iset of weights which must then be adjusted for use in naive settings, inorder to ensure that the space of possible payoffs be adequately explored.This is achieved by imposing an artificial lower bound on the probability

t Ž . twith which strategies are played. In particular, let p � 1 � � p ��n,i iand compute � t in terms of pt.ˆ ˆi i

Finally, the mixing method can be modified to achieve responsiveness byutilizing exponential smoothing. In the responsive variant of this algorithm� t is substituted for either � t or � t, depending on whether the setting is˜ ˆi i iinformed or naive. In particular, for 0 � � � 1, in informed settings and

Page 39: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER118

t1 Ž . t t1 t1 Žnaive settings, respectively, � � 1 � � � � and � � 1 �˜ ˜ ˜i i i i. t t1� � � .˜ ˆi i

A.5. Multiplicati�e Updating

Ž .This section describes an algorithm due to Freund and Schapire 1996that achieves no external regret in informed settings via multiplicativeupdating. The development of the variants of the multiplicative updatingalgorithm is analogous to the development of additive updating.

The multiplicative update rule is computed in terms of the cumulativeŽ t.payoffs achieved by all strategies namely � including the surmisedi

payoffs of those strategies which are not played. In particular, the weightassigned to strategy i at time t 1, for � � 0, is given by

� ti1 �Ž .

t1p � . 8Ž .ti � jÝ 1 �Ž .j� N

Ž .The multiplicative updating rule given in Eq. 8 can be modified in amanner identical to the mixing method, using � t and pt to becomeˆ ˆi iapplicable in naive settings, and using � t to achieve responsiveness. Ainaive variant of this multiplicative updating algorithm which achieves no

Ž .external regret appears in Auer et al. 1995 .

A.6. No Internal Regret

Ž .This section describes an algorithm due to Foster and Vohra 1997 thatachieves no internal regret in informed environments and a simple imple-

Ž .mentation due to Hart and Mas-Colell 1997 . In addition, the appropriatenaive and responsive modifications are presented. Note that learning via

Žno internal regret algorithms converges to correlated equilibrium Foster.and Vohra, 1995; Hart and Mas-Colell, 1997 and therefore converges

inside the set D�.Regret can be interpreted as a feeling of remorse over something that

happened as a result of one’s own actions. Formally, the regret felt byplayer i at time t is formulated as the difference between the payoffsobtained via strategy i and the payoffs that could have been achieved hadstrategy j been played whenever i had been played in the past: i.e.,

t t� t t � t t xR � I � � � . Cumulative regret is given by CR � Ý Ri� j i j i i� j x�0 i� jt Ž t . and internal regret is defined as IR � CR , where X �i� j i� j

� 4max X, 0 .

Page 40: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 119

Consider the case of a two-strategy game, with strategies A and B. Thecomponents of the weight vector, namely pt1 and pt1, are updated viaA Bthe following formulae, which reflect cumulative feelings of regret:

IRt IRtB � A A� Bt1 t1p � and p � . 9Ž .A Bt t t tIR IR IR IRA� B B � A A� B B � A

If there is significant regret for having played strategy B rather thanstrategy A, then the algorithm updates weights such that the probability ofplaying strategy A is increased. This algorithm is due to Foster and VohraŽ .1997 . In general, if strategy i is played at time t,

1t1 t t1 t1p � IR and p � 1 � p 10Ž .Ýj i� j i j j�i

Ž .for � 0. This algorithm is due to Hart and Mas-Colell 1997 .In a naive setting, it is necessary to compute an estimate of internal

regret. Recall that the regret at time t for having played strategy i rathert t� t t �than playing strategy j is given by R � I � � � . An estimatedi� j i j i

ˆ t ˆ t t t t� �measure of expected regret R is given by R � I � � � . Noteˆ ˆi� j i� j i j ithat the expected value of this estimated measure of regret is the actual

ˆ tmeasure of regret. Now estimated cumulative internal regret is IR �i� jt ˆ x ˆ tŽ . Ž .Ý R . Finally, weights are updated as in Eq. 10 , with IR usedx�0 i� j i� j

in place of IRt .i� jLike the additive and multiplicative updating algorithms, the no internal

regret learning algorithm can be made responsive via exponential smooth-˜ t1 ˜ t t1 t1Ž .ing of regret. In the informed case, R � 1 � � R 1 R , andi� j i� j i i� j

˜ t1 ˜ t t1ˆ t1Ž .in the naive case R � 1 � � R 1 R . Finally, it suffices toi� j i� j i i� j˜ t ˜ t Ž .use IR � R as a measure internal regret in the responsive case,i� j i� j

˜ twhere R is given by the relevant one of the two previous expressions,i� jdepending on whether the setting is informed or naive.

This concludes the discussion of the learning algorithms that wereselected for simulation in this study. While this appendix is primarily asurvey of the literature, it also includes extensions to the algorithms forapplicability in network contexts. In future work, we hope to extend thescope of our simulations by considering additional related algorithms.

APPENDIX B: RESULTS ON THE CLASS EG

In this appendix, we discuss the claims that are made in Section 3.2 thatpertain to Games EG and EG . First of all, we note that for � � 0,1.9 2.1both of these games are D-solvable. This can be shown via the general

Page 41: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER120

Ž .technique of analyzing the best-reply Cournot dynamics of the relevantŽ .class of games as in Friedman 1998 ; however, it can also be shown

directly for these simple examples.Consider Game EG for a set of eight players. For player 7, even when1.9

� � 8, � � 7 � 8�1.9 � 0. Thus, participation dominates nonparticipa-7tion for player 7, and similarly for players 5 and 6, implying that � 3.Now, given that players 5, 6, and 7 eliminate their dominated strategies,for players 0, 1, and 2, nonparticipation dominates participation. Forexample, if player 2 participates, then � � 2 � 4�1.9 � 0. Thus, given2that players 0, 1, and 2 eliminate participation because it is dominated,continuing this line of reasoning, we find that nonparticipation is domi-nated for players 3 and 4. Finally, the unique outcome via the iterated

Ž .elimination of dominated strategies is s � 0, 0, 0, 1, 1, 1, 1, 1 . By an analo-gous argument, we note that Game EG is also D-solvable with outcome2.1

Ž .s � 0, 0, 0, 1, 1, 1, 1, 1 .The above argument also shows that Games EG and EG are1.9 2.1

Ž .D-solvable, for � � 0, 1 , since dominated strategies are unchanged forthese values of �. Moreover, these games are also O-solvable, for � � 0,since in this case one can show that any dominated strategy is alsooverwhelmed using the monotonicity properties of the payoff functions.However, for � close to 1, neither game is O-solvable. In particular,participation sometimes yields a higher payoff than nonparticipation, butsometimes this effect is reversed, depending on the strategies of the otherplayers. For example, consider player 7 in Game EG , when � � 0.9.1.9Participation with seven other players yields payoffs of � � 2.79, but7nonparticipation with no other participants yields � � 5.83, while nonpar-7ticipation with seven participants yields � � 2.51. In fact, no strategies7are overwhelmed either game for � close to 1.

While both Game EG and Game EG are D-solvable, and neither1.9 2.1are O-solvable, these games differ in terms of S-solvability. In particular,Game EG is S-solvable, while Game EG is not. Consider first Game2.1 1.9EG ; in particular, consider the possible payoffs of player 2 as the leader.1.9If player 2 participates, then player 3 does not, which results in payoffs of

Ž .� 1 � 2 � 5�1.9; on the other hand, if player 2 opts out, then player 32Ž . Ž .does in fact participate, which yields payoffs of � 0 � � 2 � 6�1.9 .2

Consequently, player 2 participates when � � 6�11, since this implies thatŽ . Ž .� 1 � � 0 . Therefore, there exists a Stackelberg equilibrium in Game2 2

EG with player 2 as leader which differs from the Nash equilibrium.1.9Now consider Game EG . In this game, regardless of whether or not2.1player 2 participates, player 3 participates, since � � 3 � 6�2.1 � 0.3

Ž . ŽConsequently, player 2 only participates when � 1 � 2 � 6�2.1 � � 22. Ž .� 6�2.1 � � 0 which cannot be satisfied for any � � 1. A more de-2

tailed argument can be used to show that this game is indeed S-solvable.

Page 42: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 121

REFERENCES

Ž .Auer, P., Cesa-Bianchi, N., Freund, Y., and Schapire, R. 1995 . ‘‘Gambling in a RiggedCasino: The Adversarial Multi-armed Bandit Problem,’’ in Proceedings of the 36th AnnualSymposium on Foundations of Computer Science, pp. 322�331. New York: Assoc. Comput.Mach.

Ž .Banos, A. 1968 . ‘‘On Pseudo Games,’’ Ann. Math. Statist. 39, 1932�1945.Ž .Bernheim, B. D. 1984 . ‘‘Rationalizable Strategic Behavior,’’ Econometrica 52, No. 4,

1007�1028.Ž .Blackwell, D. 1956 . ‘‘An Analog of the Minimax Theorem for Vector Payoffs,’’ Pacific J.

Math. 6, 1�8.Ž .Borgers, T., and Sarin, R. 1995 . ‘‘Learning through Reinforcement and Replicator Dynam-

ics,’’ Mimeo.Ž .Borodin, A., and El-Yaniv, R. 1998 . Online Computation and Competiti�e Analysis. Cam-

bridge, UK: Cambridge Univ. Press.Ž .Chen, Y. 1997 . ‘‘Asynchronicity and Learning in Cost-Sharing Mechanisms,’’ Mimeo.Ž .Cover, T. 1991 . ‘‘Universal Portfolios,’’ Math. Finance 1, No. 1, 1�29.

Ž .Erev, I., and Roth, A. 1996 . ‘‘On the Need for Low Rationality Cognitive Game Theory:Reinforcement Learning in Experimental Games with Unique Mixed Strategy Equilibria,’’Mimeo.

Ž .Foster, D., and Vohra, R. 1993 . ‘‘A Randomization Rule for Selecting Forecasts,’’ Oper. Res.41, No. 4, 704�709.

Ž .Foster, D., and Vohra, R. 1995 . ‘‘Calibrated Learning and Correlated Equilibrium,’’ Preprint.Ž .Foster, D., and Vohra, R. 1997 . ‘‘Regret in the On-Line Decision Problem,’’ Games Econ.

Beha� . 21, 40�55.Ž .Freund, Y., and Schapire, R. 1996 . ‘‘Game Theory, On-Line Prediction, and Boosting,’’ in

Proceedings of the 9th Annual Conference on Computational Learning Theory, pp. 325�332.New York: Assoc. Comput. Math.

Ž .Friedman, E. J. 1996 . ‘‘Dynamics and Rationality in Ordered Externality Games,’’ GamesEcon. Beha� . 16, 65�76.

Ž .Friedman, E. 1998 . ‘‘Learnability in a Class of Non-Atomic Games Arising on the Internet,’’Mimeo.

Ž .Friedman, E., and Shenker, S. 1996 . ‘‘Synchronous and Asynchronous Learning by Respon-sive Learning Automata,’’ Mimeo.

Ž .Friedman, E., and Shenker, S. 1997 . ‘‘Learning and Implementation on the Internet,’’Mimeo.

Ž .Fudenberg, D., and Levine, D. 1989 . ‘‘Reputation and Equilibrium Selection in Games witha Patient Player,’’ Econometrica 57, 759�778.

Ž .Fudenberg, D., and Levine, D. 1995a . ‘‘Conditional Universal Consistency,’’ Mimeo.Ž .Fudenberg, D., and Levine, D. 1995b . ‘‘Consistency and Cautious Fictitious Play,’’ J. Econ.

Dynam. Control 19, 1065�1089.Ž .Fudenberg, D., and Levine, D. 1998 . The Theory of Learning in Games. Cambridge, MA:

MIT Press.Ž .Gittens, J. 1989 . Multi-armed Bandit Allocation Indices. New York: Wiley.

Ž .Greenwald, A. 1999 . ‘‘Learning to Play Network Games,’’ Ph.D. thesis, Courant Institute ofMathematical Sciences, New York University, New York.

Page 43: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

GREENWALD, FRIEDMAN, AND SHENKER122

Ž .Greenwald, A. R., and Kephart, J. O. 2000 . ‘‘Shopbots and Pricebots,’’ in Lecture Notes onArtificial Intelligence: Proceedings of the IJCAI Workshop on Agent-Mediated ElectronicCommerce. Berlin�New York: Springer-Verlag.

Ž .Hannan, J. 1957 . ‘‘Approximation to Bayes Risk in Repeated Plays,’’ in Contributions to theŽ .Theory of Games M. Dresher, A. W. Tucker, and P. Wolfe, Eds. , Vol. 3, pp. 97�139.

Princeton, NJ: Princeton Univ. Press.Ž .Hart, S., and Mas-Colell, A. 1997 . ‘‘A Simple Adaptive Procedure Leading to Correlated

Equilibrium,’’ Technical Report, Center for Rationality and Interactive Decision Theory.Ž .Huberman, B. A., and Glance, N. S. 1993 . ‘‘Evolutionary Games and Computer Simulations,’’

Proc. Nat. Acad. Sci. U.S. A. 90, 7716�7718.Ž .Van Huyck, J., Battalio, R., and Rankin, F. 1996 . ‘‘Selection Dynamics and Adaptive

Behavior without Much Information,’’ Mimeo.Ž .Kaelbling, L., Littman, M., and Moore, A. 1996 . ‘‘Reinforcement Learning: A Survey,’’

Artificial Intelligence Res. 4, 237�285.Ž .Kalai, E., and Lehrer, E. 1993 . ‘‘Rational Learning Leads to Nash Equilibrium,’’ Economet-

rica 61, 1019�1045.Ž .Littman, M. 1994 . ‘‘Markov Games as a Framework for Multi-agent Reinforcement Learn-

ing,’’ in Proceedings of Ele�enth International Conference on Machine Learning, pp. 157�163.San Mateo, CA: Morgan Kaufmann.

Ž .Lukose, R., and Huberman, B. 1997 . ‘‘A Methodology for Managing Risk in ElectronicTransactions over the Internet,’’ presented at the Third International Conference onComputational Economics.

Ž .Megiddo, N. 1986 . ‘‘On Repeated Games with Incomplete Information Played by Non-Bayesian Players,’’ Int. J. Game Theory 9, 157�167.

Ž .Milgrom, P., and Roberts, J. 1991 . ‘‘Adaptive and Sophisticated Learning in Normal FormGames,’’ Games Econ. Beha� . 3, 82�100.

Ž .Mookherjee, D., and Sopher, B. 1996 . ‘‘Learning and Decision Costs in ExperimentalConstant Sum Games,’’ Games Econ. Beha� . 19, 97�132.

Ž .Narendra, K., and Thathachar, M. A. L. 1974 . ‘‘Learning Automata: A Survey,’’ IEEETrans. Systems Man Cybernet. SMC-4, No. 4, 323�334.

Ž .Pearce, D. 1984 . ‘‘Rationalizable Strategic Behavior and the Problem of Perfection,’’Econometrica 52, 1029�1050.

Ž .Rosenthal, R. 1991 . ‘‘A Note on the Robustness of Equilibria with Respect to CommitmentOpportunities,’’ Games and Econ. Beha� . 3, 237�243.

Ž .Roth, A., and Erev, I. 1995 . ‘‘Learning in Extensive Form Games: Experimental Data andSimple Dynamic Models in the Intermediate Term,’’ Games Econ. Beha� . 8, 164�212.

Ž .Sandholm, T., and Crites, R. 1995 . ‘‘Multiagent Reinforcement Learning in the IteratedPrisoners’ Dilemma,’’ in Special Issue on the Prisoners’ Dilemma, Biosystems 37, 147�166.

Ž .Shapley, L. S. 1953 . ‘‘A Value for n-Person Games,’’ in Contributions to the Theory of GamesŽ .H. Kuhn and A. Tucker, Eds. , Vol. II, pp. 307�317. Princeton, NJ: Princeton Univ. Press.

Ž .Shenker, S. 1995 . ‘‘Making Greed Work in Networks: A Game-Theoretic Analysis of SwitchService Disciplines,’’ IEEE�ACM Trans. Networking 3, 819�831.

Ž .Shoham, Y., and Tennenholtz, M. 1993 . ‘‘Co-learning and the Evolution of Social Activity,’’Mimeo.

Ž .Stone, P., and Veloso, M. 1997 . ‘‘Multiagent Systems: A Survey from a Machine LearningPerspective,’’ Technical Report 193, Carnegie Mellon University.

Page 44: Learning in Network Contexts: Experimental Results from …parkes/cs286r/spring06/papers/... · 2006. 1. 25. · LEARNING IN NETWORK CONTEXTS 81 1. INTRODUCTION While much of classical

LEARNING IN NETWORK CONTEXTS 123

Ž .Sutton, R., and Barto, A. 1998 . Reinforcement Learning: An Introduction. Cambridge, MA:MIT Press.

Ž .Watkins, C. 1989 . Learning from Delayed Rewards. Ph.D. thesis, Cambridge University,Cambridge, UK.

Ž .Watson, J. 1993 . ‘‘A Reputation Refinement without Equilibrium,’’ Econometrica 61,199�206.

Ž .Wellman, M., and Hu, J. 1998 . ‘‘Conjectural Equilibrium in Multi-agent Learning,’’ inSpecial Issue on Multi-agent Learning, Mach. Learning 33, 179�200.