Robust commitments and the power of partial reputation

Similar documents
Deceptive Advertising with Rational Buyers

Selecting Efficient Correlated Equilibria Through Distributed Learning. Jason R. Marden

Definitions and Proofs

A Rothschild-Stiglitz approach to Bayesian persuasion

A Rothschild-Stiglitz approach to Bayesian persuasion

Slope Fields: Graphing Solutions Without the Solutions

Bayesian Persuasion Online Appendix

A Rothschild-Stiglitz approach to Bayesian persuasion

Ex Post Cheap Talk : Value of Information and Value of Signals

Conservative Belief and Rationality

On the Unique D1 Equilibrium in the Stackelberg Model with Asymmetric Information Janssen, M.C.W.; Maasland, E.

arxiv: v1 [math.oc] 28 Jun 2016

CS364A: Algorithmic Game Theory Lecture #13: Potential Games; A Hierarchy of Equilibria

For general queries, contact

Learning Equilibrium as a Generalization of Learning to Optimize

Near-Potential Games: Geometry and Dynamics

First Prev Next Last Go Back Full Screen Close Quit. Game Theory. Giorgio Fagiolo

On the errors introduced by the naive Bayes independence assumption

Lecture December 2009 Fall 2009 Scribe: R. Ring In this lecture we will talk about

General-sum games. I.e., pretend that the opponent is only trying to hurt you. If Column was trying to hurt Row, Column would play Left, so

On Reputation with Imperfect Monitoring

Lecture 1: Entropy, convexity, and matrix scaling CSE 599S: Entropy optimality, Winter 2016 Instructor: James R. Lee Last updated: January 24, 2016

Selfishness vs Altruism vs Balance

Strategic Abuse and Accuser Credibility

Near-Potential Games: Geometry and Dynamics

Exact and Approximate Equilibria for Optimal Group Network Formation

CS261: A Second Course in Algorithms Lecture #11: Online Learning and the Multiplicative Weights Algorithm

Bayes Correlated Equilibrium and Comparing Information Structures

We set up the basic model of two-sided, one-to-one matching

Correlated Equilibrium in Games with Incomplete Information

Persuading a Pessimist

Discrete Mathematics and Probability Theory Spring 2014 Anant Sahai Note 10

When to Ask for an Update: Timing in Strategic Communication

On the Impossibility of Black-Box Truthfulness Without Priors

Wars of Attrition with Budget Constraints

Models of Reputation with Bayesian Updating

Game Theory. Greg Plaxton Theory in Programming Practice, Spring 2004 Department of Computer Science University of Texas at Austin

ECO421: Reputation. Marcin P ski. March 29, 2018

Towards a General Theory of Non-Cooperative Computation

6.207/14.15: Networks Lecture 10: Introduction to Game Theory 2

Adversarial Classification

6.207/14.15: Networks Lecture 16: Cooperation and Trust in Networks

Uncertainty and Disagreement in Equilibrium Models

Discussion of "Persuasion in Global Games with an Application to Stress Testing" by Nicolas Inostroza and Alessandro Pavan

Area I: Contract Theory Question (Econ 206)

Perfect Bayesian Equilibrium

CS168: The Modern Algorithmic Toolbox Lecture #6: Regularization

Voluntary Leadership in Teams and the Leadership Paradox

EC3224 Autumn Lecture #04 Mixed-Strategy Equilibrium

CS 781 Lecture 9 March 10, 2011 Topics: Local Search and Optimization Metropolis Algorithm Greedy Optimization Hopfield Networks Max Cut Problem Nash

Iterated Strict Dominance in Pure Strategies

Satisfaction Equilibrium: Achieving Cooperation in Incomplete Information Games

Planning With Information States: A Survey Term Project for cs397sml Spring 2002

FORMULATION OF THE LEARNING PROBLEM

When to Ask for an Update: Timing in Strategic Communication. National University of Singapore June 5, 2018

Appendix of Homophily in Peer Groups The Costly Information Case

Other-Regarding Preferences: Theory and Evidence

ECO 199 GAMES OF STRATEGY Spring Term 2004 Precepts Week 7 March Questions GAMES WITH ASYMMETRIC INFORMATION QUESTIONS

Basic Game Theory. Kate Larson. January 7, University of Waterloo. Kate Larson. What is Game Theory? Normal Form Games. Computing Equilibria

Conjectural Variations in Aggregative Games: An Evolutionary Perspective

Entropic Selection of Nash Equilibrium

MATH 521, WEEK 2: Rational and Real Numbers, Ordered Sets, Countable Sets

Weak Robust (Virtual) Implementation

BELIEFS & EVOLUTIONARY GAME THEORY

MARKOV DECISION PROCESSES (MDP) AND REINFORCEMENT LEARNING (RL) Versione originale delle slide fornita dal Prof. Francesco Lo Presti

Sequential Games with Incomplete Information

Theory Field Examination Game Theory (209A) Jan Question 1 (duopoly games with imperfect information)

Econ 618: Correlated Equilibrium

Lecture 4: Lower Bounds (ending); Thompson Sampling

Sequential Equilibrium in Computational Games

Conditional probabilities and graphical models

1 Basic Game Modelling

A Game-Theoretic Analysis of Games with a Purpose

2 Making the Exponential Mechanism Exactly Truthful Without

Blocking Development

Empirical Risk Minimization

The key is that there are two disjoint populations, and everyone in the market is on either one side or the other

Price and Capacity Competition

Crowdsourcing contests

Informational Substitutes and Complements. Bo Waggoner June 18, 2018 University of Pennsylvania

The No-Regret Framework for Online Learning

Price of Stability in Survivable Network Design

Online Learning, Mistake Bounds, Perceptron Algorithm

This chapter covers asymptotic analysis of function growth and big-o notation.

Government 2005: Formal Political Theory I

CMU Noncooperative games 4: Stackelberg games. Teacher: Ariel Procaccia

September Math Course: First Order Derivative

A Note on the Existence of Ratifiable Acts

A New and Robust Subgame Perfect Equilibrium in a model of Triadic Power Relations *

Should all Machine Learning be Bayesian? Should all Bayesian models be non-parametric?

Interior-Point Methods for Linear Optimization

6 Evolution of Networks

Robust Knowledge and Rationality

The Simplex and Policy Iteration Methods are Strongly Polynomial for the Markov Decision Problem with Fixed Discount

σ(a) = a N (x; 0, 1 2 ) dx. σ(a) = Φ(a) =

Players as Serial or Parallel Random Access Machines. Timothy Van Zandt. INSEAD (France)

Evidence with Uncertain Likelihoods

Sender s Small Concern for Credibility and Receiver s Dilemma

Prioritized Sweeping Converges to the Optimal Value Function

Contracts under Asymmetric Information

Transcription:

Robust commitments and the power of partial reputation Vidya Muthukumar, Anant Sahai EECS, UC Berkeley {vidya.muthukumar,sahai}@eecs.berkeley.edu Abstract Agents rarely act in isolation their behavioral history, in particular, is public to others. We seek a non-asymptotic understanding of how a leader agent should shape this history to its maximal advantage, knowing that follower agent(s) will be learning and responding to it. We work within the formal framework of two-player Stackelberg leader-follower games with finite observations of the leader commitment, a setting which can arise in security games, network routing and law enforcement. First, we formally show that for most non-zero-sum game settings the Stackelberg commitment itself is highly non-robust to observational uncertainty. We propose observationrobust, polynomial-time-computable commitment constructions for leaders that approximate the Stackelberg payoff, and also show that these commitment rules approximate the maximum obtainable payoff (which could in general be greater than Stackelberg). Finally, we discuss the potential of modeling partial establishment of reputation, including establishing follower belief in leader commitment. 1 Introduction Consider a selfish, rational agent (designated as the leader ) who is interacting non-cooperatively with another selfish, rational agent (designated as the follower ). If the agents are interacting simultaneously, they will naturally play a ash equilibrium(a) of the two-player game. This is the traditionally studied solution concept in game theory. ow, let s say that the leader has the ability to reveal its strategy in advance in the form of a mixed strategy commitment, and the follower has the ability to observe this commitment and respond to it. The optimal strategy for the leader now corresponds to the Stackelberg equilibrium of the two-player leader-follower game. The Stackelberg solution concept enjoys application in several engineering settings where commitment is the natural framework for the leader, such as security [PPM + 08], network routing [Rou04] and law enforcement [MS17]. The solution concept is interpreted in a broader sense as the ensuing strategy played by a patient leader that wishes to build a reputation by playing against an infinite number of myopic followers [KW82, MR82, FL89, FL92]. Crucially, one can show rigorously that the leader will benefit significantly from commitment power, i.e. its ensuing Stackelberg equilibrium payoff is at least as much as the simultaneous equilibrium payoff [VSZ10]. But which ever interpretation one chooses, the Stackelberg solution concept assumes a very idealized setting (even over and above the assumptions of selfishness and infinite rationality) in which the mixed strategy commitment is exactly revealed to the follower. Further, the follower 100% believes that the leader will actually stick to her commitment. What happens when these assumptions are relaxed? What if the leader could only demonstrate her commitment in a finite number of interactions how would she modify her commitment to maximize payoff, and how much commitment power would she continue to enjoy? The primary purpose of this paper is to provide an answer to these questions. The fundamental thing that changes in a finite-interaction regime is that the follower only observes a part of leader behavioral history and needs to learn about the leader s strategic behavior to the extent that he is able to respond as optimally as possible. By restricting our attention to finitely repeated play, we arrive at a problem setting that is fairly general: these follower agents in general will not know about the preferences of the leader agent. When provided with historical context, it is more likely that they will use it rather than ignore it. A broad umbrella of problems that has received significant attention in the machine learning literature is learning of strategic behavior from samples of play; for example, learning the agent s utility function through inverse 1

reinforcement learning [ZMBD08], learning the agent s level of rationality [WZB13], and inverse game theory [KS15]. While significant progress has been made in this goal of learning strategic behavior, attention has been restricted to the passive learning setting in which the agent is unaware of the presence of the learner, or agnostic to the learner s utility. In many situations, the agent himself will be invested in the outcome of the learning process. In our work, we put ourselves in the shoes of an agent who is shaping her historical context and is aware of the learner s presence as well as preferences, and study her choice of optimal strategy 1. As we will see, the answer will depend both on her utility function itself, as well as what kind of response she is able to elicit from the learner. 1.1 Related work Economists have cared about the so-called reputation effects for a while, but theoretical analysis has been limited to the asymptotics and convergence of an appropriate solution concept to Stackelberg. Reputation effects were first observed in the chain-store paradox, a firm-competition game where an incumbent firm would often deviate from ash equilibrium behavior and play its aggressive Stackelberg (pure, in this case) strategy [Sel78]. Theoretical justification was provided for this particular application [KW82, MR82] by modeling a + 1-player interaction between a leader and multiple followers, and studying the ash equilibrium of an ensuing game as. It was shown that the leader would play its pure Stackelberg strategy 2 in the Bayes-ash equilibrium of this game endowed with a common prior on the leader s payoff structure 3. This model was generalized to such leader-follower ensembles for a general two-player game, and considering the possibility of mixed-strategy reputation, still retaining the asymptotic nature of results [FL89, FL92]. Our analysis is primarily new in that it studies reputation building in a non-asymptotic manner, and plays a predictive role in prescribing how a leader should play to maximally build its reputation in a finite-interaction environment. In particular, if the leading agent is restricted to only < interactions, what would its behavioral strategy be and how much (relative) reputational advantage would it continue to enjoy? The specific case of studying Stackelberg games with observational uncertainty has received relatively more attention, particularly in the security games literature, but the focus has been more algorithmic than analytical. A Stackelberg security game is played between a defender, who places different utility levels on different targets to be protected and accordingly uses her resources to defend some subset of targets; and an attacker, who observes the defender strategy (generally imperfectly) and wishes to attack some of the targets depending on whether he thinks they are left open as well as how much he values those targets. Pita et al [PJT + 10] first proposed a model for the defernder (leader) to account for attacker (follower) observational uncertainty by allowing the follower to anchor [RT97] to a certain extent on its uniform prior. While they showed significant returns from using their model through extensive experiments, they largely circumvented the algorithmic and analytical challenges by not explicitly considering random samples of defender behavior, thus keeping the attacker response deterministic but shifted. A duo of papers [AKK + 12, SAY + 12] later considered a model of full-fledged observational uncertainty with a Bayesian prior and posterior update based on samples of behavior, and proposed heuristic algorithmic techniques to compute the optimum. In fact, they also factored for quantal responses using a bounded rationality model [MP95]. This work showed through simulations that there could be a positive return over and above Stackelberg payoff. In one important piece of analytical work, the problem was also considered for the special case of zero-sum games [BHP14], and it was shown that the Stackelberg commitment itself approximated the optimal payoff. However, this reasoning does not extend to non-zero-sum games as was evidenced by an example in that paper itself, and an analysis of the non-zero-sum case remained very much open. Quantitatively different sources of observational uncertainty have also been considered, such as the possibility of the mixed commitment being fully observed or not observed at all [KCP11]; as well as different ways of handling the uncertainty, eg: showing that for some security games the Stackelberg and 1 It is worth mentioning the recent paradigm of cooperative inverse reinforcement learning [HMRAD16] which studies the problem of agent investment in principal learning where the incentives are not completely aligned, but the setting is cooperative. In contrast, we focus on non-cooperative settings. 2 ote that this is fundamentally different from the mixed strategy Stackelberg solution concept more restrictive, and not necessarily advantageous over ash, but it turns out to be so in the firm competition case. 3 A more nuanced model considered a Bayesian prior over a leader being constrained to play its pure Stackelberg strategy as opposed to unconstrained play. 2

ash equilibria coincide and observation of commitment does not matter [KYK + 11]. Our work differs from these recent developments in that it limits observation in the most natural way (i.e. number of samples of leader behavior), and provides analytical insights for a generic non-zero-sum two-player game. We must also make mention of robust solution concepts that have been developed in the economics literature and are intellectually related to our contributions. The concept of tremblinghand-perfect-equilibrium [Sel75] explicitly studies how robust mixed-strategy equilibria are to slight perturbations in the mixtures themselves, and a similar concept was proposed for Stackelberg [VDH97]. Another solution concept, quantal-response-equilibrium [MP95], studies agents that are boundedly rational, and the Stackelberg equivalent (a leader optimizing its commitment tailored to a boundedly rational follower) was studied empirically for security games [PJT + 10]. This is an orthogonal but equally important source of uncertainty in response. Spiritually important to our paradigm is also the one of games with asymmetric private information; i.e. leading agent possesses private information and follower agent does not know the nature of that information but has a prior distribution on its value. The question then becomes to what extent the leading agent wishes to reveal, or obfuscate, this information either through repeated play [AMS95] or explicit signalling [CS82]. Relatively recent work on Bayesian persuasion [KG11] somewhat connects the two perspectives and shows through analysis and examples the potential advantage of manipulating private information to elicit a favorable response. We can think of the Stackelberg commitment as a special case of private information in a rather different model. With the limited-observation perspective, the question will arise of whether we want this information to be revealed, or obfuscated, to the maximum. 1.2 Our contributions Our main contribution is to understand the extent of reputational advantage when interaction is finite, and prescribe approximately optimal commitment rules for this regime. We work within the formal framework of Stackelberg leader-follower games in which a follower obtains a limited number of observations of the leader commitment. We first show, theoretically and through extensive simulation that in most non-zero-sum games the Stackelberg commitment is highly non-robust to observational uncertainty and therefore a poor choice. ext, we propose robust commitment rules for leaders and show that we can approach the Stackelberg payoff at a specific rate with the number of observations. We do this by making an explicit connection between the problem and best-response-classification, and prove and leverage a new tail bound on discrete distribution learning. Further, we show that a possible advantage for the leader from limited observability is only related to response mismatch (i.e. the follower s response being different than expected), and leverage the results on discrete distribution learning to show that this advantage is only realized in a small-sample regime. In particular, the small-sample regime refers to one where the number of observations is less than an effective dimension of the game 4. The corollary is that we are able to approximate the optimal payoff in many cases through a simple construction which can be obtained in constant time from computation of the Stackelberg commitment (itself a polynomial-time operation [CS06]). We corroborate our theoretical results with extensive simulation on illustrative examples and random ensembles of security games. 2 Problem statement 2.1 Preliminaries We represent a two-player leader-follower game in normal form 5 by the pair of m n matrices (A, B), where A = [ a 1 a 2... a n ] denotes the leader payoff matrix while B = [ b1 b 2... b n ] denotes the follower payoff matrix. We consider a setting of asymmetric private information in which the leader knows about the follower preferences (i.e. it knows A) while the follower does not know about the leader preferences (i.e. it possesses no knowledge of the matrix B) 6. 4 One can think of this as the number of leader pure strategies, but for some regimes eg: security games, it can be significantly smaller. 5 We could have also chosen to express the game in extensive form, but the normal form turns out to be more natural for our analysis. 6 This is an important assumption for the paper, and is in fact used in traditional reputation building frameworks. In future, we will want to better understand situations of repeated interaction where the -level leader and 1-level follower 3

In the infinite limit of play, the well-established effect from the follower s point of view is that the leader has established commitment, or developed a reputation, for playing according to some mixed strategy x m, where m represents the m-dimensional probability simplex. We denote the follower s best response to a mixed strategy x by j (x) [m]. (ote that this is a pure strategy, and an important assumption is that ties are assumed to be broken in favor of the leader [CS06]. We also define best-response regions as the set of leader commitments that would elicit the pure strategy response j from the follower, i.e. R j := {x m : j (x) = j}. Thus, we can define the leader s payoff to be expected with an infinite reputation: Definition 1. A leader with an infinite reputation of playing according to the strategy x should expect payoff f (x) := x, a k. Therefore, the leader s Stackelberg payoff is the solution to the program f := max x m f (x). The argmax of this program is denoted as the Stackelberg commitment x. denote the best response faced in Stackelberg equilibrium by j := j (x ). Further, we It is clear that the Stackelberg commitment is optimal for the leader under two conditions: the leader is 100% known to be committed to a fixed strategy, and the follower knows exactly the leader s committed-to strategy. For a finite number of interactions, neither is true. 2.2 Observational uncertainty with established commitment Even assuming that there is a shared belief in commitment, there is uncertainty. In particular, with a finite number of plays, the follower does not know the exact strategy that the leader has committed to, and only has an estimate. Consider the situation where a leader can only reveal its commitment x through pure strategy plays 7 I 1, I 2,..., I i.i.d x. Let us say that the commitment is known (to both leader and followers) to come from a set of mixed strategies X m. We denote the maximum likelihood estimate of the leader s mixed strategy, as seen by the follower, by X. It is reasonable to expect, under certainty of commitment, that a rational 8 follower would adopt a best-response to X, i.e. play the mixed strategy We can express the expected leader payoff under this learning rule. j ( X ). (1) Definition 2. A leader which will have plays according to the hidden strategy x can expect payoff in the th play of ] f (x) := E [ x, a j ( X ) against a follower that plays according to (1). The maximal payoff a leader can expect is f := max x m f (x) and it acquires this payoff by playing the argmax strategy x. Ideally, we want to understand how close f is to f, and also how close x is to x. An answer to the former question would tell us how observational uncertainty impacts the first-player advantage. An answer to the latter question would shed light on whether the best course of action deviates significantly from Stackelberg commitment. We are also interested in algorithmic techniques for approximately computing the quantity f, as it would involve solving a non-convex optimization problem. are both learning about one another. 7 We intellectually motivated this model by considering followers who do not know about the leader s preferences and learn this through samples of play. Practically, this setting arises in Stackelberg security games [PPM + 08] where it is common knowledge that the defender (leader) has a randomized commitment and is deploying resources according to this commitment in every time step. 8 Rational is in quotes because the follower is not necessarily using expected-utility theory (although there is an expected-utility-maximization interpetation to this estimate if the mixed strategy were uniform drawn from X ). 4

3 Main Results 3.1 Robustness of Stackelberg commitment to observational uncertainty A natural first question is whether the Stackelberg commitment, which is clearly optimal if the game were being played infinitely (or equivalently, if the leader had infinite commitment power), is also suitable for finite play. In particular, we might be interested in evaluating whether we can do better than the baseline Stackelberg performance f. We show through a few baseline examples that the answer can vary. Follower Follower 1 2 Follower 3 1 F. Leader 2 e ns po s re Leader (-1,1) e sp 1 3 (1,-1) (0,0) (-1,1) (0,0) (1,-1) (3,-3) F se = 1 2 se on sp re = =3 F. nse 2 on sp e.r spo (0,0) e =3 (0,1) 2 (-1,1) ns nse 2 1 F. re F (1,0) o sp (3,-3) 1 1 re (1,-1) re 1 F. (0,0) es.r = spo F. re s on p = 2 (a) 2 3 zero-sum game. (b) 2 2 non-zero-sum game. (c) 2 3 non-zero-sum game. Figure 1. Illustration of examples of zero-sum game and non-zero-sum games in the form of normal form tables and ideal leader payoff function f (.). p denotes the probability that the leader will play strategy 1, and fully describes leader mixed commitment for these 2 n games. Actual value Upper bound (adv. over pure Stackelberg) 2 F. = 2 (0,0) = (1,-1) e 1 s on Leader umber of observations Figure 2. Semilog plot of extent of advantage over Stackelberg payoff as a function of in the 2 3 zero-sum game. Example 1. We consider a 2 3 zero-sum game, represented in normal form in Figure 1(a), in which we can express the leader strategy according to the probability it will pick strategy 1, denoted by p, and leader payoff is as follows: f (p; 1) := p if follower best responds with strategy 1 f (p; 2) := 1 p if follower best responds with strategy 2 f (p; 3) := 3 4p if follower best responds with strategy 3 Since the game is zero-sum, the follower responds in a way that is worst-case for the leader. This means that we can express the leader payoff as f (p) = min{f (p; 1), f (p; 2), f (p; 3)}. This leader payoff structure is depicted in Figure 1(a). Therefore, we can express the Stackelberg payoff as f = max f (p) = 1/2, p [0,1] 5

attained at p = 1/2. We wish to evaluate f (p ). It was noted [BHP14] that f (p ) f (p ) by the minimax theorem, but not always clear whether strict inequality would hold (that is, if observational uncertainty gives a strict advantage). For this example, we can actually get a sizeable improvement! To see this, look at the simple example of = 1. Denoting P = 1 j=1 I[Ij = 1], we have f 1(1/2) = Pr[ P 1 = 0].f(1/2; 1) + Pr[ P 1 = 1].f(1/2; 3) = 1/2 1/2 + 1/2 1 = 3/4. The semilog plot in Figure 2 shows that this improvement persists for larger values of, although the extent of improvement decreases rather dramatically, in fact exponentially, with. We can show that f (1/2) f = 1/2 Pr[ P > 2/3] = 1/2 Pr[ P 1/2 > 1/6] exp{ D KL (3/2 1/2)} where D KL (..) denotes the Kullback-Leibler divergence, and the last inequality is due to Sanov s theorem [CK11]. This shows analytically that the advantage does indeed decrease exponentially with. aturally, this is because the stochasticity that enables the more favorable follower response with pure strategy 3 decreases as increases. Example 1 showed us how Stackelberg commitment power could be increased by stochastically eliciting more favorable responses. We now see an example illustrating that the commitment power can disappear completely. Ideal Stackelberg Robust Optimum Robust Expected leader payoff Actual Stackelberg Expected leader payoff umber of observations umber of observations (a) Plot depicting the performance of the sequence of robust commitments {x } 1 and the Stackelberg commitment x as a function of. The benchmark for comparison is idealized Stackelberg payoff f. (b) Plot showing the performance of sequence of robust commitments {x } 1 as compared to the optimum performance f (brute-forced). Figure 3. Example of a 2 2 non-zero-sum game for which observational uncertainty is always undesirable. Example 2. We consider a 2 2 non-zero-sum game, represented in normal form and leader payoff structure in Figure 1(b). Explicitly, the ideal leader payoff function is { p if p 1/2 f (p) = p if p > 1/2. This is essentially the example reproduced in [BHP14], which we repeat for storytelling value. otice that f = 1/2, p = 1/2, but the advantage evaporates with observational uncertainty. For any finite, we have f (1/2) = Pr[ P 1/2](1/2) + Pr[ P > 1/2]( 1/2) = 1/2 1/2 1/2 1/2 = 0. 6

Remarkably, this implies that f f (p ) = 1/2 and so lim f f (p ) 0! This is clearly a very negative result for the robustness of Stackelberg commitment, and as a very pragmatic matter tells us that the idealized Stackelberg commitment p is far from ideal in finite-observation settings. This example shows us a case where stochasticity in follower response is not desired, principally because of the discontinuity in the leader payoff function at p. Example 2 displayed to the fullest the significant disadvantage of observational uncertainty. The game considered was special in that there was no potential for limited-observation gain, while in the game presented in Example 1 there was only potential for limited-observational gain. What could happen in general? Our next and final example provides an illustration. (adv. over pure Stackelberg) Actual value Upper bound Expected leader payoff Ideal Stackelberg Robust Actual Stackelberg Expected leader payoff Optimum Robust umber of observations umber of observations umber of observations (a) Plot of the extent of (dis)advantage (b) Plot depicting the performance of (c) Plot showing the performance over Stackelberg payoff as a function of the sequence of robust commitments of sequence of robust commitments. {x } 1 and the Stackelberg commitment x as a function of. The bench- performance f (brute-forced). {x } 1 as compared to the optimum mark for comparison is idealized Stackelberg payoff f. Figure 4. A more general example of a 2 3 non-zero-sum game in which observational uncertainty could either help or hurt the leader. Example 3. Our final example considers a 2 3 non-zero-sum game, whose normal form and leader payoff structure are depicted in Figure 1(c). The ideal leader payoff function is p if p 1/2 f (p) = 1/2 p if 1/2 < p 5/7 3 4p if p > 5/7. As in the other examples, f = 1/2, p = 1/2. otice that this example captures both positive and negative effects of stochasticity in response. On one hand, follower response 2 is highly undesirable (a la Example 2) but follower response 3 is highly desirable (a la Example 1). What is the net effect? We have f (1/2) = Pr[ P 1/2](1/2) + Pr[1/2 < P < 5/7](0) + Pr[ P 5/7](1) = (1/2)(1/2) + Pr[ P 5/7](1) 1/4 + 1/2 exp{ D KL (5/7 1/2)}. A quick calculation thus tells us that f (p ) <= f if 8, showing that Stackelberg in fact has poor robustness for this example. Intuitively, the probability of the bad stochastic event remains constant while the probability of the good stochastic event decreases exponentially with. Even more damningly, we see that lim f f (p ) lim 1/4 1/2 exp{ D KL (5/7 1/2) = 1/4, again showing that the Stackelberg commitment is far from ideal. We can see the dramatic decay of leader advantage over and above Stackelberg, and ensuing disadvantage even for a very small number of observations, in Figure 3(a). While the three examples detailed above provide differing conclusions, there are some common threads. For one, in all the examples it is the case that committing to the Stackelberg mixture x can result in the follower being agnostic between more than one response. Only one of these responses, the pure strategy j, is desirable for the leader. A very slight misperception in the estimation of the true value x can therefore lead to a different, worse-than-expected response 7

and this misperception happens with a sizeable, non-vanishing probability 9. On the flipside, a different response could also lead to better-than-expected payoff, raising the potential for a gain over and above f. However, these better-than-expected responses cannot share a boundary with the Stackelberg commitment, and we will see that the probability of eliciting them decreases exponentially with. The net effect is that the Stackelberg commitment is, most often, not robust and critically, this is even the case for small amounts of uncertainty. Our first proposition is a formal statement of instability of Stackelberg commitment for a general 2 n game. We denote the leader probability of playing strategy 1 by p [0, 1], and the Stackelberg commitment s probability of playing strategy 1 by p. Furthermore, let φ(t) denote the CDF of the standard normal distribution (0, 1). We are now ready to state the result. Proposition 1. For any 2 n leader-follower game in which p (0, 1) and f (p) discontinuous at p = p, we have f (p ) f C (φ( ) C 12 C ) + exp{ C 2 } (2) where C, C, C are strictly positive constants depending on the parameters of the game. This directly implies the following: 1. For some 0 > 0, we have f (p ) < f for all > 0. 2. We have lim f (p ) < f. The technical ingredients in the proof of Proposition 1 are the Berry-Esseen theorem [?,?], used to show that the detrimental alternate responses on the Stackelberg boundary are non-vanishingly likely and the Hoeffding bound, used to tail bound the probability of potentially beneficial alternate responses not on the boundary. For non-robustness of Stackelberg commitment to hold, the two critical conditions for the game are that there is a discontinuity at the Stackelberg boundary, and that the Stackelberg commitment is mixed. For a zero-sum game, the first condition does not hold and the Stackelberg commitment stays robust as we saw in Example 1. The proposition directly implies that the ideal Stackelberg payoff is only obtained for the exact case of = (when the commitment is perfectly observed), and that for any value of < there is a non-vanishing reduction in payoff. In the simulation section, we will see that this gap is significant in practice. 3.2 Robust commitments achieving close-to-stackelberg performance The surprising message of Proposition 1 is that, in general, the Stackelberg commitment x is undesirable. Colloquially speaking, the commitment x is pushed to the extreme point of the bestresponse-region R j to ensure optimality under idealized conditions; and this is precisely what makes it sub-optimal under uncertainty. What if we could move our commitment a little bit into the interior of the region R j instead, such that we can get a high-probability-guarantee on eliciting the expected response, while staying sufficiently close to the idealized optimum? Our main result quantifies the ensuing tradeoff and shows that we can cleverly construct the commitment to approximate Stackelberg performance. We use the linear system of equations B (j) x c (j) for some matrix B (j) R m m and some vector c (j) R m to describe the best-response-polytope R j. A crucial quantity will also be the dimension of the polytope R j, i.e. We are now ready to define our theorem. dim(r j) := rank(b (j) ). Theorem 1. Let k = rank(b (j ) ) and the best-response polytope R j be measurable in R m 1. k Then, under the condition that >, we can construct commitment x (C(B (j ) )) 1/p = x 9 Making a connecting to the concept of trembling-hand-perfect-equilibrium, which is defined for ash equilibrium [Sel75], one can say that the Stackelberg commitment is, in general, trembling-hand-imperfect. 8

such that ( ) f f (x ) C 1 + p exp{ C 1 2p } (3) for any p < 1/2 and universal constants (C, C ) > 0 and constant C(B (j ) ) > 0 that depends on the geometry of the best-response polytope R j. The proof of Theorem 1, deferred to the appendix, involves some technical steps to achieve as good as possible a scaling in. We need to find a commitment such that the empirical estimate that the follower uses lies within best-response-polytope R j with high probability. To obtain our high-probability-bound, we require, and prove, a more sophisticated variant of Devroye s lemma [Dev83] to carefully tail bound the total variation, or equivalently l 1-norm, between an empirical estimate and the true distribution. To properly leverage the tail bound, we also need to choose the direction of deviation from Stackelberg commitment, i.e. the quantity, carefully. This direction is chosen in a way depending on the localized geometry of the best-response-polytope R j around its extreme point 10 x. The details of the construction are provided in the proof. The caveat of Theorem 1 is that commitment power can be robustly exploited in this way only if there are enough observations of the commitment. One obvious requirement is that the best-response-region R j needs to be Lebesgue-measurable i.e. non-empty in R m 1. Two other crucial quantities dictate how many observations are needed. 1. The number of observations needs to roughly scale as the dimension of the best-response region R j localized around the Stackelberg commitment. This is important in the context of high-dimensional games like security games where the dimension of the game is exponential in certain parameters but the dimension of the best-response regions is actually polynomial. We capitalize on this improvement because the follower effectively needs to be able to learn only its best-response as opposed to the entire mixture of leader strategies. 2. The geometry of the best-response-polytope R j also matters, in a localized sense around the Stackelberg commitment x as well as a globalized sense. Qualitatively, the two desirable properties for robustness are a global property: that the width of the polytope is high, and a local property: that the localized geometry about the Stackelberg commitment x is sufficiently descriptive of the global geometry of the polytope. These notions are formalized in the proof, and are captured by the constant C(B (j ) ) > 0. It is also worth mentioning that our techniques could probably be generalized to convex bodies (as opposed to just polytopes), but the property of convexity is a critical ingredient in the proofs. We mention a couple of special cases of leader-follower games for which the robust commitment constructions of Theorem 1 are not required; in fact, it is simply optimal to play Stackelberg. Remark 1. For games in which the mixed-strategy Stackelberg equilibrium coincides with a pure strategy, the follower s best response is always as expected regardless of the number of observations. There is no tradeoff and it is simply optimal to play Stackelberg even under observational uncertainty. Remark 2. For the zero-sum case, it was observed [BHP14] that a Stackelberg commitment is made assuming that the follower will respond in the worst case. If there is observational uncertainty, the follower can only respond in a way that yields payoff for the leader that is better than expected. This results in an expected payoff greater than or equal to the Stackelberg payoff f, and it simply makes sense to stick with the Stackelberg commitment x. As we have seen, this logic does not hold up for non-zero-sum games because different responses can lead to worse-thanexpected payoff. One way of thinking of this is that the function f (.) is discontinuous in x for a non-zero-sum game, but continuous for the special case of zero-sum. 3.3 The small-sample case As noted, the construction in Theorem 1 provided no guarantees for the case of the number of observations being less than the dimension of the best-response-polytope R j. This can be thought of as the small-sample regime in which the number of samples of a discrete distribution is less than the support of that distribution. Would any commitment work? Our next result takes the form of 10 ote that the Stackelberg commitment always lies at an extreme point owing to it maximizing an appropriate LP [CS06]. 9

a converse and shows that there exists a game for which in the small-sample case, no commitment construction can get us close to Stackelberg advantage in the small-sample case; and furthermore, a strategy that mixes between fewer pure strategies can be optimal as opposed to a fully mixed commitment. Theorem 2. There exists a m m game for which m implies a constant gap to Stackelberg equilibrium for any commitment. That is, for this game (A, B) R m m and any commitment x m, we have f (x) f C for some constant C > 0 that does not depend on. We prove Theorem 2 by explicitly constructing the game, which is a m m security game. We consider this game to be a special case of a response-dominant game, i.e. a game in which the leader prefers the follower to respond with strategy j regardless of her choice of mixed strategy. Formally, this means that a j a l for all l j. The intuition for Theorem 2 is as follows. For a response-dominant game, the performance of a commitment within the Stackelberg best-response-region depends crucially on the probability with which the follower will respond as expected 11. This requires the follower to be able to accurately classify its best-response to the leader strategy, which obviously requires a minimum number of samples. In the small-sample-case, this classification guarantee evaporates regardless of the nature of the commitment, and so the natural thing to do for the leader is to mix between fewer strategies, or even play a pure strategy. For the game construction in Theorem 2, we are able to explicitly calculate the optimum value f, and we observe that the best commitment mixes between m 1 pure strategies, as opposed to m. Effectively, the commitment x ensures a deterministic expected response from the follower for any 1! It is worth noting that for the special case of response-dominant games, it is both in leader s and follower s best interest for the actual best-response to be learned accurately 12. And in the small-sample context, our construction shows that the leader can value certainty in best-response over deviation from the Stackelberg commitment. We have already seen that this alignment in incentives for leader strategy learnability does not hold for a general non-zero-sum game, and this is what creates potential for a gain over and above the Stackelberg advantage: the leader would ideally like to fool the follower into responding differently, and in a way beneficial to itself. ext, we will establish limits on when this is possible. 3.4 Approximating the maximum possible payoff So far, we have considered the limited-observability problem and shown that the Stackelberg commitment x is not a suitable choice. We have constructed robust commitments that come close to idealized Stackelberg payoff f and shown that the guarantee fundamentally depends on the number of observations scaling with the effective dimension of the game. ow, we turn to the question of whether we can approximate f, the actual optimum of the program. ote that since the problem is in general non-convex in x, it is P-hard to exactly compute. Rather than the traditional approach of constructing a polynomial-time-approximation-algorithm, our approach is approximation-theoretic 13. We first show that in the large-sample case, we cannot do much better than the actual Stackelberg payoff f ; informally speaking, our ability to fool the follower into responding strictly-better-than-expected is limited. Combining this with the robust commitment construction of Theorem 1, we obtain an approximation to the optimum payoff. The main result of this section is stated below. Theorem 3. We have f f m + Cn. for some constant C > 0 depending on the parameters of the game (A, B). 11 This was informally discussed in prior work as expected strategies, defined as the set of follower responses that will ensure at least as much payoff to the leader as if the commitment had been fully observed. For response-dominant games, the set of expected strategies is the singleton set {j }. 12 This is much like in the cooperative learning paradigm where both principal and agent would like to increase learnability. 13 In other words, the extent of approximation is measured by the number of samples as opposed to the runtime of an algorithm. This is very much the flavor of previously-obtained results on Stackelberg zero-sum security games [BHP14]. 10

As a corollary the commitment construction defined in Theorem 1 provides a Õ( 1 )-additiveapproximation algorithm to f. The proof of Theorem 3 carefully ties several facts that we have seen formally, as well as alluded to, together. First, a generalization of Proposition 1 tells us that we cannot improve sizeably over Stackelberg by committing to any mixed strategy on the boundary between two or more best-response regions. Interestingly, the tie-breaking tenet is not required to prove such a statement, although it makes writing the proof operationally easier. Second, we show that the improvement gained by a fixed commitment in the interior of any best-response-region decreases exponentially with, simply because the probability of eliciting a better-than-expected response decreases exponentially with. Putting these two facts together, the natural thing to try would be to fix commitments that approach a boundary as increases (much like our robust commitment constructions, but now with a different motive). This should happen fast enough that we maintain a sizeable probability of eliciting a different response for every, while simultaneously ensuring that that different response is actually better-than-expected. We then show that the ensuing gain over and above Stackelberg would have to decrease with according to the rate specified. Intellectually, Theorem 3 tells us that the robust commitments are essentially optimal. The practical benefit that Theorem 3 affords us is that we now have an approximation to the optimum payoff the leader could possibly obtain, which can be computed in constant time after computing the Stackelberg equilibrium, which itself is polynomial time [CS06]. This is because the robust commitment is obtained by first computing Stackelberg equilibrium x, and then deviating away from x in the magnitude and direction specified. We will now study the practical benefits of our analytical insights through extensive simulation. 4 Simulations and experiments We used the Python libraries numpy and scipy for all our simulations. All computations were exact. 4.1 Example 2 2 and 2 3 games First, we return to the non-zero-sum games described in Examples 2 and 3. These were 2 2 and 2 3 games respectively, and the Stackelberg commitment was non-robust for both games. ow, armed with the results in Theorem 1, we can employ our robust commitment constructions and study their performance. To construct our robust commitments, we first computed the Stackelberg commitment using the LP solver in scipy (scipy.optimize.linprog), and then used the construction in Theorem 1. Figures 3(a) and 4(b) compares the expected payoff obtained by our robust commitment construction scheme {x } 1 for different numbers of samples, and for the games described in Examples 2 and 3 respectively. The benchmark with respect to which we measure this expected payoff is the Stackelberg payoff f (obtained by Stackelberg commitment under infinite observability). We also observe a significant gap between the payoffs obtained by these robust commitment constructions and the payoff obtained if we used the Stackelberg commitment x. We showed in theory that there is significant benefit for choosing the commitment to factor in such observational uncertainty, and we can now see it in practice. Furthermore, for the case of 2 leader pure strategies we were able to brute-force the maximum possible obtainable payoff 14 f, and compare the value to the robust commitment payoff. This comparison is particularly valuable for smaller values of, as shown in Figures 3(b) and 4(c). We notice that the values are much closer even than our theory would have predicted, and even for small values of. Thus, our constructions have significant practical benefit as well: we are able to get close to the optimum while drastically reducing the required computation (to just solving n LPs!). Since these examples involved 2 n games, the commitment construction became trivial (i.e. only one direction to move along) next, we test the entire sophistry of the commitment construction for m m security games. And instead of looking at specific examples, we now look at a random ensemble. 14 first we used scipy.optimize.brute with an appropriate grid size to initialize, and then ran a gradient descent at that initialization point. This was feasible for the case of 2 pure strategies. 11

4.2 Random security games Reward Target 1 2 3 4 5 Defender (if protected) Unif[0,1] Unif[0,1] Unif[0,1] Unif[0,1] Unif[0,1] Defender (if unprotected) Unif[-1,0] Unif[-1,0] Unif[-1,0] Unif[-1,0] Unif[-1,0] Attacker (if protected) Unif[-1,0] Unif[-1,0] Unif[-1,0] Unif[-1,0] Unif[-1,0] Attacker (if unprotected) Unif[0,1] Unif[0,1] Unif[0,1] Unif[0,1] Unif[0,1] Figure 5: Illustration of random ensemble of 5 5 security game. Expected leader payoff Ideal Stackelberg Robust Actual Stackelberg (gap to Stackelberg payoff) % gap to Stackelberg payoff umber of observations umber of observations umber of observations (a) Plot of expected defender payoff (b) Log-log plot of the gap between robust commitment payoff and idealized robust commitment payoff and idealized (c) Percentage plot of the gap between when defender uses robust commitments compared to Stackelberg commitment Stackelberg payoff. Stackelberg payoff. as well as idealized Stackelberg payoff. Figure 6. Illustration of performance of robust commitments and Stackelberg commitment in random 5 5 Stackelberg security games for a finite number of observations of defender commitment. Our next set of simulations is inspired by the security games framework. We create a random ensemble of 5 5 security games in which the defender can defend one of 5 targets, and the attacker can attack one of these 5 targets. The defender and attacker rewards are chosen to be uniformly at random between [0, 1], and their penalties are uniform at random between [ 1, 0]. This is essentially the random ensemble that was created in previous empirical work on security games [AKK + 12]. Figure 5 shows the construction of this ensemble in a picture. The purpose of random security games is to show that the properties we observed above unstable Stackelberg commitment, robust commitment payoff approximating the optimum are the norm rather than the exception. Figure 6 illustrates the results for random security games. The performance of the sequence of robust commitments {x } 1, as well as the Stackelberg commitment x is plotted in Figure 6(a) against the benchmark of idealized Stackelberg performance f. Figure 6(b) depicts the rate of convergence of the gap in robust commitment performance to the idealized Stackelberg payoff we can clearly see the O( 1 ) rate of convergence in this plot. Finally, Figure 6(c) plots the percentage gap between robust commitment payoff and idealized Stackelberg payoff as a function of. We can make the following conclusions from these plots: 1. The Stackelberg commitment is extremely non-robust on average. In fact we noticed that this was the case with high probability. This happens because the Stackelberg commitment, although it can vary widely for different games in the random ensemble, is almost surely on a boundary shared with other responses and therefore unstable. 2. The robust commitments are doing much better on average than the original Stackelberg commitment even for very large values of. The stark difference in payoff between the two motivates the construction of the robust commitment, which was as easy to compute as the Stackelberg commitment. 12

It is worth mentioning that unlike in the examples presented before, we were not able to brute-force the optimum feasibly for m > 2 using scipy.optimize solvers; in the near-future we may want to utilize state-of-the-art solvers that have been developed in the security games literature [AKK + 12, PJT + 10] to make the practical comparison. We should expect to draw similar conclusions as we did for the simpler 2 n-game examples. 5 Discussion We constructed robust commitment constructions with several advantages. First, we are able to effectively preserve the Stackelberg payoff by ensuring a high-probability guarantee on the follower responding as expected. An oblique, but significant advantage to our robust commitments is that their guarantees hold even if we removed the pivotal assumption of follower breaking ties in favor of the leader in which we case we will want a guarantee on remaining in the interior of the bestresponse-region. For the large-sample regime, this guarantee will also hold, and as the number of observations grows, our construction naturally converges to the Stackelberg commitment. Second, we established fundamental limits on the ability of the leader to gain over and above Stackelberg payoff. We formally showed that this ability disappears in the large-sample regime, and in a certain sense that our robust commitments are approximately optimal. Our results established a formal connection between leader payoff and follower discrete distribution learning, and in the context of these limits, both players are mutually incentivized to increase learnability under limited samples, even though the setting is non-cooperative which was a rather surprising conclusion. One area that deserves much more scrutiny is the small-sample regime. Here, we should not actually expect the follower to respond directly to the empirical distribution just because some pure strategies are not observed, the follower would not necessarily expect them never to be played! More realistic may be the follower responding to some prior estimate weighted with the side information of observations, or using a more sophisticated estimate of the discrete distribution that provides smaller-sample guarantees [VV16, ADOS16]. It may even be possible that the follower discards observations entirely and the commitment is disregarded in favor of some other generic strategy (eg: minimax or uniform mixtures). The true answer of how we might expect a follower to behave in small-sample-regime might be determined by rigorous experimental analysis in addition to theory, and is an important avenue for future work. Perhaps related to this goal is also a better theoretical understanding of how a follower should respond in the limited-sample case, particularly in the context of recent work on optimally learning discrete mixtures [VV16, ADOS16]. Only once we understand how a follower is expected to behave in this regime can we prescribe commitment rules, and definitively answer whether commitment gains are still possible. Our model also took commitment establishment for granted, i.e. the follower assumed that the leader would indeed be drawing its pure strategies iid from the same mixture in every round. In reality, the belief in commitment needs to be built up over time, and in earlier rounds the follower may not always respond directly to the empirical estimate. He may want to maintain a possibility that the leader will play minimax/ash and respond accordingly. Formally, this framework of reputation building has been explored asymptotically as mentioned before, but a good non-asymptotic understanding does not currently exist. When the iid assumption is removed, a related question is whether the leader could choose to play deterministically, in such a way to increase strategy learnability while maintaining a follower impression of iid commitment. This could be related to recent work on learning from untrusted data which has been proposed for different application domains. Most generally, the non-asymptotic reputation building framework builds a paradigm for a rich set of problems in learning in games. We might want to consider online interaction between an -level agent and a k-level agent 15 While we know how to play a repeated zero-sum game online [FS99], we might want to move beyond the overly pessimistic worst case while playing a non-zero-sum game. If we relax the assumption on the leader knowing the follower preferences, we might want to consider a combined model where both leader and follower solve learning problems through repeated interaction: the leader has to learn about follower utilities, and the follower has to learn the leader commitment if there is any. 15 in our framework, we implicitly considered a 1-level follower agent 13