[[!meta title="Game Theory: a Critical Introduction"]] ## Index * Difference between meteorological and traffic-jam-style predictions: 18. * Frequent assumption (but not always) on game theory that individuals knows the rules of the game; that they even known their inner motives: 28. * Individualism, separation of structure and choice, 31. * MAD, 86; 88. * Elimination of non-credible threats, 88. # Excerpts ## Intro What is game theory: In many respects this enthusiasm is not difficult to understand. Game theory was probably born with the publication of The Theory of Games and Economic Behaviour by John von Neumann and Oskar Morgenstern (first published in 1944 with second and third editions in 1947 and 1953). They defined a game as any interaction between agents that is governed by a set of rules specifying the possible moves for each participant and a set of outcomes for each possible combination of moves. How it can help: If game theory does make a further substantial contribution, then we believe that it is a negative one. The contribution comes through demonstrating the limits of a particular form of individualism in social science: one based exclusively on the model of persons as preference satisfiers. This model is often regarded as the direct heir of David Hume’s (the 18th century philosopher) conceptualisation of human reasoning and motivation. It is principally associated with what is known today as rational choice theory, or with the (neoclassical) economic approach to social life (see Downs, 1957, and Becker, 1976). Our main conclusion on this theme (which we will develop through the book) can be rephrased accordingly: we believe that game theory reveals the limits of ‘rational choice’ and of the (neoclassical) economic approach to life. In other words, game theory does not actually deliver Jon Elster’s ‘solid microfoundations’ for all social science; and this tells us something about the inadequacy of its chosen ‘microfoundations’. Assumptions: three key assumptions: agents are instrumentally rational (section 1.2.1); they have common knowledge of this rationality (section 1.2.2); and they know the rules of the game (section 1.2.3). These assumptions set out where game theory stands on the big questions of the sort ‘who am I, what am I doing here and how can I know about either?’. The first and third are ontological. 1 They establish what game theory takes as the material of social science: in particular, what it takes to be the essence of individuals and their relation in society. The second raises epistemological issues 2 (and in some games it is not essential for the analysis). It is concerned with what can be inferred about the beliefs which people will hold about how games will be played when they have common knowledge of their rationality. Instrumental rationality (_Homo economicus_): We spend more time discussing these assumptions than is perhaps usual in texts on game theory because we believe that the assumptions are both controversial and problematic, in their own terms, when cast as general propositions concerning interactions between individuals. This is one respect in which this is a critical introduction. The discussions of instrumental rationality and common knowledge of instrumental rationality (sections 1.2.1 and 1.2.2), in particular, are indispensable for anyone interested in game theory. In comparison section 1.2.3 will appeal more to those who are concerned with where game theory fits in to the wider debates within social [...] Individuals who are instrumentally rational have preferences over various ‘things’, e.g. bread over toast, toast and honey over bread and butter, rock over classical music, etc., and they are deemed rational because they select actions which will best satisfy those preferences. One of the virtues of this model is that very little needs to be assumed about a person’s preferences. Rationality is cast in a means-end framework with the task of selecting the most appropriate means for achieving certain ends (i.e. preference satisfaction); and for this purpose, preferences (or ‘ends’) must be coherent in only a weak sense that we must be able to talk about satisfying them more or less. Technically we must have a ‘preference ordering’ because it is only when preferences are ordered that we will be able to begin to make judgements about how different actions satisfy our preferences in different degrees. [...] Thus it appears a promisingly general model of action. For instance, it could apply to any type of player of games and not just individuals. So long as the State or the working class or the police have a consistent set of objectives/ preferences, then we could assume that it (or they) too act instrumentally so as to achieve those ends. Likewise it does not matter what ends a person pursues: they can be selfish, weird, altruistic or whatever; so long as they consistently motivate then people can still act so as to satisfy them best. An agent is "rational" in this conext when they have preference ordering" and if "they select the action that maximizes those preferences: Readers familiar with neoclassical Homo economicus will need no further introduction. This is the model found in standard introductory texts, where preferences are represented by indifference curves (or utility functions) and agents are assumed rational because they select the action which attains the highest feasible indifference curve (maximises utility). For readers who have not come across these standard texts or who have forgotten them, it is worth explaining that preferences are sometimes represented mathematically by a utility function. As a result, acting instrumentally to satisfy best one’s preferences becomes the equivalent of utility maximising behaviour. Reason and slavery: Even when we accept the Kantian argument, it is plain that reason’s guidance is liable to depend on characteristics of time and place. For example, consider the objective of ‘owning another person’. This obviously does not pass the test of the categorical imperative since all persons could not all own a person. Does this mean then we should reject slave-holding? At first glance, the answer seems to be obvious: of course, it does! But notice it will only do this if slaves are considered people. Of course we consider slaves people and this is in part why we abhor slavery, but ancient Greece did not consider slaves as people and so ancient Greeks would not have been disturbed in their practice of slavery by an application of the categorical imperative. Reason dependent on culture: Wittgenstein suggests that if you want to know why people act in the way that they do, then ultimately you are often forced in a somewhat circular fashion to say that such actions are part of the practices of the society in which those persons find themselves. In other words, it is the fact that people behave in a particular way in society which supplies the reason for the individual person to act: or, if you like, actions often supply their own reasons. This is shorthand description rather than explanation of Wittgenstein’s argument, but it serves to make the connection to an influential body of psychological theory which makes a rather similar point. Cognitive dissonance and free market proponents: Festinger’s (1957) cognitive dissonance theory proposes a model where reason works to ‘rationalise’ action rather than guide it. The point is that we often seem to have no reason for acting the way that we do. For instance, we may recognise one reason for acting in a particular way, but we can equally recognise the pull of a reason for acting in a contrary fashion. Alternatively, we may simply see no reason for acting one way rather than another. In such circumstances, Festinger suggests that we experience psychological distress. It comes from the dissonance between our self-image as individuals who are authors of our own action and our manifest lack of reason for acting. It is like a crisis of self-respect and we seek to remove it by creating reasons. In short we often rationalise our actions ex post rather than reason ex ante to take them as the instrumental model suggests. [...] Research has shown that people seek out and read advertisements for the brand of car they have just bought. Indeed, to return us to economics, it is precisely this insight which has been at the heart of one of the Austrian and other critiques of the central planning system when it is argued that planning can never substitute for the market because it presupposes information regarding preferences which is in part created in markets when consumers choose. Infinite regress of the economics of information acquisiton (i.e learning, eg. from a secret service): Actually most game theorists seem to agree on one aspect of the problem of belief formation in the social world: how to update beliefs in the presence of new information. They assume agents will use Bayes’s rule. This is explained in Box 1.6. We note there some difficulties with transplanting a technique from the natural sciences to the social world which are related to the observation we have just made. We focus here on a slightly different problem. Bayes provides a rule for updating, but where do the original (prior) expectations come from? Or to put the question in a different way: in the absence of evidence, how do agents form probability assessments governing events like the behaviour of others? There are two approaches in the economics literature. One responds by suggesting that people do not just passively have expectations. They do not just wait for information to fall from trees. Instead they make a conscious decision over how much information to look for. Of course, one must have started from somewhere, but this is less important than the fact that the acquisition of information will have transformed these original ‘prejudices’. The crucial question, on this account, then becomes: what determines the amount of effort agents put into looking for information? This is deceptively easy to answer in a manner consistent with instrumental rationality. The instrumentally rational agent will keep on acquiring information to the point where the last bit of search effort costs her or him in utility terms the same amount as the amount of utility he or she expects to get from the information gained by this last bit of effort. The reason is simple. As long as a little bit more effort is likely to give the agent more utility than it costs, then it will be adding to the sum of utilities which the agent is seeking to maximise. [...] This looks promising and entirely consistent with the definition of instrumentally rational behaviour. But it begs the question of how the agent knows how to evaluate the potential utility gains from a bit more information _prior to gaining that information_. Perhaps he or she has formulated expectations of the value of a little bit more information and can act on that. But then the problem has been elevated to a higher level rather than solved. How did he or she acquire that expectation about the value of information? ‘By acquiring information about the value of information up to the point where the marginal benefits of this (second-order) information were equal to the costs’, is the obvious answer. But the moment it is offered, we have the beginnings of an infinite regress as we ask the same question of how the agent knows the value of this second-order information. To prevent this infinite regress, we must be guided by something _in addition_ to instrumental calculation. But this means that the paradigm of instrumentally rational choices is incomplete. The only alternative would be to assume that the individual _knows_ the benefits that he or she can expect on average from a little more search (i.e. the expected marginal benefits) because he or she knows the full information set. But then there is no problem of how much information to acquire because the person knows everything! Infinite recursion of the common knowledge (CKR): If you want to form an expectation about what somebody does, what could be more natural than to model what determines their behaviour and then use the model to predict what they will do in the circumstances that interest you? You could assume the person is an idiot or a robot or whatever, but most of the time you will be playing games with people who are instrumentally rational like yourself and so it will make sense to model your opponent as instrumentally rational. This is the idea that is built into the analysis of games to cover how players form expectations. We assume that there is common knowledge of rationality held by the players. It is at once both a simple and complex approach to the problem of expectation formation. The complication arises because with common knowledge of rationality I know that you are instrumentally rational and since you are rational and know that I am rational you will also know that I know that you are rational and since I know that you are rational and that you know that I am rational I will also know that you know that I know that you are rational and so on…. This is what common knowledge of rationality means. [...] It is difficult to pin down because common knowledge of X (whatever X may be) cannot be converted into a finite phrase beginning with ‘I know…’. The best one can do is to say that if Jack and Jill have common knowledge of X then ‘Jack knows that Jill knows that Jack knows …that Jill knows that Jack knows…X’—an infinite sentence. The idea reminds one of what happens when a camera is pointing to a television screen that conveys the image recorded by the very same camera: an infinite self-reflection. Put in this way, what looked a promising assumption suddenly actually seems capable of leading you anywhere. [...] The problem of expectation formation spins hopelessly out of control. Nevertheless game theorists typically assume CKR and many of them, and certainly most people who apply game theory in economics and other disciplines Uniformity: Consistent Alignment of Beliefs (CAB), another weak assumption based on Harsanyi doctrine requiring equal information; followed by a comparison with Socract dialectics: Put informally, the notion of _consistent alignment of beliefs_ (CAB) means that no instrumentally rational person can expect another similarly rational person who has the same infor mation to develop different thought processes. Or, alternatively, that no rational person expects to be surprised by another rational person. The point is that if the other person’s thought is genuinely moving along rational lines, then since you know the person is rational and you are also rational then your thoughts about what your rational opponent might be doing will take you on the same lines as his or her own thoughts. The same thing applies to others provided they respect _your_ thoughts. So your beliefs about what your opponents will do are consistently aligned in the sense that if you actually knew their plans, you would not want to change your beliefs; and if they knew your plans they would not want to change the beliefs they hold about you and which support their own planned actions. Note that this does not mean that everything can be deterministically predicted. Reason reflecting on itself: These observations are only designed to signal possible trouble ahead and we shall examine this issue in greater detail in Chapters 2 and 3. We conclude the discussion now with a pointer to wider philosophical currents. Many decades before the appearance of game theor y, the Ger man philosophers G.F.W.Hegel and Immanuel Kant had already considered the notion of the self-conscious reflection of human reasoning on itself. Their main question was: can our reasoning faculty turn on itself and, if it can, what can it infer? Reason can certainly help persons develop ways of cultivating the land and, therefore, escape the tyranny of hunger. But can it understand how it, itself, works? In game theory we are not exactly concerned with this issue but the question of what follows from common knowledge of rationality has a similar sort of reflexive structure. When reason knowingly encounters itself in a game, does this tell us anything about what reason should expect of itself? What is revealing about the comparison between game theory and thinkers like Kant and Hegel is that, unlike them, game theory offers something settled in the form of CAB. What is a source of delight, puzzlement and uncertainty for the German philosophers is treated as a problem solved by game theory. For instance, Hegel sees reason reflecting on reason as it reflects on itself as part of the restlessness which drives human history. This means that for him there are no answers to the question of what reason demands of reason in other people outside of human history. Instead history offers a changing set of answers. Likewise Kant supplies a weak answer to the question. Rather than giving substantial advice, reason supplies a negative constraint which any principle of knowledge must satisfy if it is to be shared by a community of rational people: any rational principle of thought must be capable of being followed by all. O’Neill (1989) puts the point in the following way: [Kant] denies not only that we have access to transcendent meta- physical truths, such as the claims of rational theology, but also that reason has intrinsic or transcendent vindication, or is given in consciousness. He does not deify reason. The only route by which we can vindicate certain ways of thinking and acting, and claim that those ways have authority, is by considering how we must discipline our thinking if we are to think or act at all. This disciplining leads us not to algorithms of reason, but to certain constraints on all thinking, communication and interaction among any plurality. In particular we are led to the principle of rejecting thought, act or communication that is guided by principles that others cannot adopt. (O’Neill p. 27) Summary: To summarise, game theory is avowedly Humean in orientation. [...] The second [aspect] is that game theorists seem to assume _too much_ on behalf of reason [even more than Hume did]. Giddens, Wittgenstein language games and the "organic or holistic view of the relation between action and structure" (pages 30-31): The question is ontological and it connects directly with the earlier discussion of instrumental rationality. Just as instrumental rationality is not the only ontological view of what is the essence of human rationality, there is more than one ontological view regarding the essence of social interaction. Game theory works with one view of social interaction, which meshes well with the instrumental account of human rationality; but equally there are other views (inspired by Kant, Hegel, Marx, Wittgenstein) which in turn require different models of (rational) action. State (pages 32-33): Perhaps the most famous example of this type of institutional creation comes from the early English philosopher Thomas Hobbes who suggested in Leviathan that, out of fear of each other, individuals would contract with each other to form a State. In short, they would accept the absolute power of a sovereign because the sovereign’s ability to enforce contracts enables each individual to transcend the dog- eat-dog world of the state of nature, where no one could trust anyone and life was ‘short, nasty and brutish’. Thus, the key individualist move is to draw attention to the way that structures not only constrain; they also enable (at least those who are in a position to create them). It is the fact that they enable which persuades individuals consciously (as in State formation) or unconsciously (in the case of those which are generated spontaneously) to build them. To bring out this point and see how it connects with the earlier discussion of the relation between action and structure it may be helpful to contrast Hobbes with Rousseau. Hobbes has the State emerging from a contract between individuals because it serves the interests of those individuals. Rousseau also talked of a social contract between individuals, but he did not speak this individualist language. For him, the political (democratic) process was not a mere means of ser ving persons’ interests by satisfying their preferences. It was also a process which changed people’s preferences. People were socialised, if you like, and democracy helped to create a new human being, more tolerant, less selfish, better educated and capable of cherishing the new values of the era of Enlightenment. By contrast, Hobbes’ men and women were the same people before and after the contract which created the State. 4 Game theory as justification of individualism (pages 32-33), which reminds the discussion made by Dany-Robert Dufour in La Cite Perverse; it is also noted that the State is considered a "collective action agency": Where do structures come from when they are separate from actions? An ambitious response which distinguishes methodological individualists of all types is that the structures are merely the deposits of previous interactions (potentially understood, of course, as games). This answer may seem to threaten an infinite regress in the sense that the structures of the previous interaction must also be explained and so on. But, the individualist will want to claim that ultimately all social str uctures spring from interactions between some set of asocial individuals; this is why it is ‘individualist’. [...] Returning to game theory’s potential contribution, we can see that, in so far as individuals are modelled as Humean agents, game theory is well placed to help assess the claims of methodological individualists. After all, game theory purports to analyse social interaction between individuals who, as Hume argued, have passions and a reason to serve them. Thus game theory should enable us to examine the claim that, beginning from a situation with no institutions (or structures), the self-interested behaviour of these instrumentally rational agents will either bring about institutions or fuel their evolution. An examination of the explanatory power of game theory in such settings is one way of testing the individualist claims. In fact, as we shall see in subsequent chapters, the recurring difficulty [...] Suppose we take the methodological individualist route and see institutions as the deposits of previous interactions between individuals. Individualists are not bound to find that the institutions which emerge in this way are fair or just. Indeed, in practice, many institutions reflect the fact that they were created by one group of people and then imposed on other groups. All that any methodological individualist is committed to is being able to find the origin of institutions in the acts of individuals qua individuals. The political theory of liberal individualism goes a stage further and tries to pass judgement on the legitimacy of particular institutions. Institutions in this view are to be regarded as legitimate in so far as all individuals who are governed by them would have broadly ‘agreed’ to their creation. Naturally, much will turn on how ‘agreement’ is to be judged because people in desperate situations will often ‘agree’ to the most desperate of outcomes. Thus there are disputes over what constitutes the appropriate reference point (the equivalent to Hobbes’s state of nature) for judging whether people would have agreed to such and such an arrangement. We set aside a host of further problems which emerge the moment one steps outside liberal individualist premises and casts doubt over whether people’s preferences have been autonomously chosen. Game theory has little to contribute to this aspect of the dispute. However, it does make two significant contributions to the discussions in liberal individualism with respect to how we might judge ‘agreement’. Prisioner's dilemma and the hobbesian argument for the creation of a State (pages 36-37): resolution would require a higher State in the next upper level of recursion: Finally there is the prisoners’ dilemma game (to which we have dedicated the whole of Chapter 5 and much of Chapter 6). Recall the time when there were still two superpowers each of which would like to dominate the other, if possible. They each faced a choice between arming and disarming. When both arm or both disarm, neither is able to dominate the other. Since arming is costly, when both decide to arm this is plainly worse than when both decide to disarm. However, since we have assumed each would like to dominate the other, it is possible that the best outcome for each party is when that party arms and the other disarms since although this is costly it allows the arming side to dominate the other. These preferences are reflected in the ‘arbitrary’ utility pay-offs depicted in Figure 1.4. Game theory makes a rather stark prediction in this game: both players will arm (the reasons will be given later). It is a paradoxical result because each does what is in their own interest and yet their actions are collectively self- defeating in the sense that mutual armament is plainly worse than the alternative of mutual disarmament which was available to them (pay-off 1 for utility pay-offs depicted in Figure 1.4. Game theory makes a rather stark prediction in this game: both players will arm (the reasons will be given later). It is a paradoxical result because each does what is in their own interest and yet their actions are collectively self- defeating in the sense that mutual armament is plainly worse than the alternative of mutual disarmament which was available to them (pay-off 1 for each rather than 2). The existence of this type of interaction together with the inference that both will arm has provided one of the strongest arguments for the creation of a State. This is, in effect, Thomas Hobbes’s argument in Leviathan. And since our players here are themselves States, both countries should agree to submit to the authority of a higher State which will enforce an agreement to disar m (an argument for a strong, independent, United Nations?). Too much trust in that type of instrumental rationality might lead to lower outcomes in some games: The term rationalisable has been used to describe such strategies because a player can defend his or her choice (i.e. rationalise it) on the basis of beliefs about the beliefs of the opponent which are not inconsistent with the game’s data. However, to pull this off, we need ‘more’ commonly known rationality than in the simpler games in Figures 2.1 and 2.3. Looking at Figure 2.4 we see that outcome (100, 90) is much more inviting than the rationalisable outcome (1, 1). It is the deepening confidence in each other’s instrumental rationality (fifth-order CKR, to be precise) which leads our players to (1, 1). In summary notation, the rationalisable strategies R2, C2 are supported by the following train of thinking (which reflects the six steps described earlier): -- 48 Nash-equilibrium: self-confirming strategy: A set of rationalisable strategies (one for each player) are in a Nash equilibrium if their implementation confirms the expectations of each player about the other’s choice. Put differently, Nash strategies are the only rationalisable ones which, if implemented, confirm the expectations on which they were based. This is why they are often referred to as self-confirming strategies or why it can be said that this equilibrium concept requires that players’ beliefs are consistently aligned (CAB). -- 53 Arguments against CAB: In the same spirit, it is sometimes argued (borrowing a line from John von Neumann and Oskar Morgenstern) that the objective of any analysis of games is the equivalent of writing a book on how to play games; and the minimum condition which any piece of advice on how to play a game must satisfy is simple: the advice must remain good advice once the book has been published. In other words, it could not really be good advice if people would not want to follow it once the advice was widely known. On this test, only (R2, C2) pass, since when the R player follows the book’s advice, the C player would want to follow it as well, and vice versa. The same cannot be said of the other rationalisable strategies. For instance, suppose (R1, C1) was recommended: then R would not want to follow the advice when C is expected to follow it by selecting C1 and likewise, if R was expected to follow the advice, C would not want to. Both versions of the argument with respect to what mutual rationality entails seem plausible. Yet, there is something odd here. Does respect for each other’s rationality lead each person to believe that neither will make a mistake in a game? Anyone who has talked to good chess players (perhaps the masters of strategic thinking) will testify that rational persons pitted against equally rational opponents (whose rationality they respect) do not immediately assume that their opposition will never make errors. On the contrary, the point in chess is to engender such errors! Are chess players irrational then? One is inclined to answer no, but why? And what is the difference as -- 57 Limits conceptualizing reason as an algorithm ("Humean approach to reason is algorithmic"): Harsanyi doctrine seems to depend on a powerfully algorithmic and controversial view of reason. Reason on this account (at least in an important part) is akin to a set of rules of inference which can be used in moving from evidence to expectations. That is why people using reason (because they are using the same algorithms) should come to the same conclusion. However, there is genuine puzzlement over whether such an algorithmic view of reason can apply to all circumstances. Can any finite set of rules contain rules for their own application to all possible circumstances? The answer seems to be no, since under some sufficiently detailed level of description there will be a question of whether the rule applies to this event and so we shall need rules for applying the rules for applying the rules. And as there is no limit to the detail of the description of events, we shall need rules for applying the rules for applying the rules, and so on to infinity. In other words, every set of rules will require creative interpretation in some circumstances and so in these cases it is perfectly possible for two individuals who share the same rules to hold divergent expectations. This puts a familiar observation from John Maynard Keynes and Frank Knight regarding genuine uncertainty in a slightly different way, but nevertheless it yields the same conclusion. There will be circumstances under which individuals are unable to decide rationally what probability assessment to attach to events because the events are uncertain and so it should not be surprising to find that they disagree. Likewise, the admiration for entrepreneurship found among economists of the Austrian school depends on the existence of uncertainty. Entrepreneurship is highly valued precisely because, as a result of uncertainty, people can hold different expectations regarding the future. In this context, the entrepreneurs are those who back their judgement against that of others and succeed. In other words, there would be no job for entrepreneurs if we all held common expectations in a world ruled by CAB! A similar conclusion regarding ineliminable uncertainty is shared by social theorists who have been influenced by the philosophy of Kant. They deny that reason should be understood algorithmically or that it always supplies answers as to what to do. For Kantians reason supplies a critique of itself which is the source of negative restraints on what we can believe rather than positive instructions as to what we should believe. Thus the categorical imperative (see section 1.2.1), which according to Kant ought to determine many of our significant choices, is a sieve for beliefs and it rarely singles out one belief. Instead, there are often many which pass the test and so there is plenty of room for disagreement over what beliefs to hold. Perhaps somewhat surprisingly though, a part of Kant’s argument might lend support to the Nash equilibrium concept. In particular Kant thought that rational agents should only hold beliefs which are capable of being universalised. This idea, taken by itself, might prove a powerful ally of Nash. [...] Of course, a full Kantian perspective is likely to demand rather more than this and it is not typically adopted by game theorists. Indeed such a defence of Nash would undo much of the foundations of game theory: for the categorical imperative would even recommend choosing dominated strategies if this is the type of behaviour that each wished everyone adopted. Such thoughts sit uncomfortably with the Humean foundations of game theory and we will not dwell on them for now. Instead, since the spirit of the Humean approach to reason is algorithmic, we shall continue discussing the difficulties with the Harsanyi—Aumann defence of Nash. -- 58-60 "Irrational" plays which might intend to send a message to other players: Indeed why should one assume in this way that players cannot (or should not) try to make statements about themselves through patterning their ‘trembles? The question becomes particularly sharp once it is recalled that, on the conventional account, players must expect that there is always some chance of a tremble. Trembles in this sense are part of normal behaviour, and the critics argue that agents may well attempt to use them as a medium for signalling something to each other. Of course, players will not do so if they believe that their chosen pattern is going to be ignored by others. But that is the point: why assume that this is what they will believe from the beginning, especially when agents can see that the generally accepted use of trembles as signals might secure a better outcome for both players [...]? Note that this is not an argument against backward induction per se: it is an argument against assuming CKR while working out beliefs via backward induction (i.e. a criticism of Nash backward induction). When agents consider patterning their ‘trembles’, they project forward about future behaviour given that there are trembles now or in the past. What makes it ambiguous whether they should do this, or stick to Nash backward induction instead, is that there is no uniquely rational way of playing games like Figures 3.5 or 3.6 (unlike the race to 20 game in which there is). In this light, the subgame perfect Nash equilibrium offers one of many possible scenarios of how rational agents will behave. -- 93 Why not expand this affirmation so _any_ move to signal some intention? ## Misc Page 101: Hence these refinements (e.g. proper equilibria), likethe Nash equilibrium project itself, seem to have to appeal to somethingother than the traditional assumptions of game theory regarding rationalaction in a social context. Page 102: regarding the relation betweenconvention following and instrumental rationality. The worry here takes usback to the discussion of section 1.2.3 where for instance it was suggested thatconventions might best be understood in the way suggested by Wittgenstein orHegel. In short, the acceptance of convention may actually require a radicalreassessment of the ontological foundations of game theory. Page 102: actually require a radicalreassessment of the ontological foundations of game theory. Page 103: Why not give up on the Nash concept altogether? This ‘giving up’ might takeon one of two forms. Firstly, game theory could appeal to the concept ofrationalisable strategies (recall section 2.4 of Chapter 2) which seemuncontentiously to flow from the assumptions of instrumental rationalityand CKR. The difficulty with such a move is that it concedes that gametheory is unable to say much about many games (e.g. Figures 2.6, 2.12, etc.).Naturally, modesty of this sort might be entirely appropriate for gametheory, although it will diminish its claims as a solid foundation for socialscience. Page 104: Unlike the instrumentally rational model, for Hegelians and Marxists actionbased on preferences feeds back to affect preferences, and so on, in an everunfolding chain. (See Box 3.1 for a rather feeble attempt to blend desires andbeliefs.) Likewise some social psychologists might argue that the key to actionlies less with preferences and more with the cognitive processes used bypeople; and consequently we should address ourselves to understanding theseprocesses. Page 105: 105 Page 106: Quite simply, the significant social processes which write history cannot beunderstood through the lens of instrumental rationality. This destines gametheory to a footnote in some future text on the history of social theory. Welet the reader decide.3 Page 108: Thirdly, the sociology of the discipline may provide further clues. Twoconditions would seem to be essential for the modern development of adiscipline within the academy. Firstly the discipline must be intellectuallydistinguishable from other disciplines. Secondly, there must be some barriersto the amateur pursuit of the discipline. (A third condition which goes withoutsaying is that the discipline must be able to claim that what it does ispotentially worth while.) The first condition reduces the competition fromwithin the academy which might come from other disciplines (to do thisworthwhile thing) and the second ensures that there is no effectivecompetition from outside the academy. In this context, the rational choicemodel has served economics very well. It is the distinguishing intellectualfeature of economics as a discipline and it is amenable to such formalisationthat it keeps most amateurs well at bay. Thus it is plausible to argue that thesuccess of economics as a discipline within the social sciences has been closelyrelated to its championing of the rational choice model. Page 108: kind of amnesia or lobotomy which thediscipline seems to have suffered regarding most things philosophical duringthe postwar period. Page 108: It isoften more plausible to think of the academy as a battleground betweendisciplines rather than between ideas and the disciplines which have goodsurvival features (like the barriers to entry identified above) Page 109: explanations willonly prosper in so far as they are both superior and they are not institutionallyundermined by the rise of neoclassical economics and the demise ofsociology. It is not necessary to see these things conspiratorially to see thepoint of this argument. All academics have fought their corner in battles overresources and they always use the special qualities of their discipline asammunition in one way or another. Thus one might explain in functionalist termsthe mystifying attachment of economics and game theory to Nash. Page 110: We have no special reason to prioritise one strand of our proposedexplanation. Yet, there is more than a hint of irony in the last suggestionbecause Jon Elster has often championed game theory and its use of the Nashequilibrium concept as an alternative to functional arguments in social science.Well, if the use of Nash by game theorists is itself to be explainedfunctionally, then… Page 111: Liberal theorists often explain the State with reference to state of nature. Forinstance, within the Hobbesian tradition there is a stark choice between astate of nature in which a war of all against all prevails and a peacefulsociety where the peace is enforced by a State which acts in the interest ofall. The legitimacy of the State derives from the fact that people who wouldotherwise live in Hobbes’s state of nature (in which life is ‘brutish, nasty andshort’) can clearly see the advantages of creating a State. Even if a State had Page 111: not surfaced historically for all sorts of other reasons, it would have to beinvented.Such a hypothesised ‘invention’ would require a cooperative act of comingtogether to create a State whose purpose will be to secure rights over life andproperty. Nevertheless, even if all this were common knowledge, it wouldnot guarantee that the State will be created. There is a tricky further issuewhich must be resolved. The people must agree to the precise property rightswhich the State will defend and this is tricky because there are typically avariety of possible property rights and the manner in which the benefits ofpeace will be distributed depends on the precise property rights which areselected (see Box 4.1).In other words, the common interest in peace cannot be the onlyelement in the liberal explanation of the State, as any well-defined andpoliced property rights will secure the peace. The missing element is anaccount of how a particular set of property rights are selected and thiswould seem to require an analysis of how people resolve conflicts ofinterest. This is where bargaining theory promises to make an importantcontribution to the liberal theory of the State because it is concernedprecisely with interactions of this sort. Page 112: State creation in Hobbes’s world provides one example (which especiallyinterests us because it suggests that bargaining theory may throw light onsome of the claims of liberal political theory with respect to the State), butthere are many others. Page 113: The creation of the institutions for enforcing agreements (like the State)which are presumed by cooperative game theory requires as we have seenthat agents first solve the bargaining problem non-cooperatively. Page 113: Indeed for this reason, and following thepractice of most game theorists, we have so far discussed the non-cooperative play of games ‘as if ’ there was no communication, therebyimplicitly treating any communication which does take place in the absenceof an enforcement agency as so much ‘cheap talk’ Page 113: In cooperative games agents cantalk to each other and make agreements which are binding on later play. Innon-cooperative games, no agreements are binding. Players can say whateverthey like, but there is no external agency which will enforce that they dowhat they have said they will do. Page 114: Thus it will have shown not just what sort of State rational agents mightagree to create, but also how rational agents might solve a host of bargainingproblems in social life. Unfortunately we have reasons to doubt therobustness of this analysis and it is not difficult to see our grounds forscepticism. If bargaining games resemble the hawk-dove game and thediscussion in Chapter 2 is right to point to the existence of multipleequilibria in this game under the standard assumptions of game theory, thenhow does bargaining theory suddenly manage to generate a uniqueequilibrium? Page 114: 114face value, the striking result of the non-cooperative analysis of thebargaining problem is that it yields the same solution to the bargainingproblem as the axiomatic approach. If this result is robust, then it seems thatgame theory will have done an extraordinary service by showing thatbargaining problems have unique solutions (whichever route is preferred).Thus it will have shown not just what sort of State rational agents mightagree to create, but also how rational agents might solve a host of bargainingproblems in social life. Unfortunately we have reasons to doubt therobustness of this analysis and it is not difficult to see our grounds forscepticism. If bargaining games resemble the hawk-dove game and thediscussion in Chapter 2 is right to point to the existence of multipleequilibria in this game under the standard assumptions of game theory, thenhow does bargaining theory suddenly manage to generate a uniqueequilibrium? Page 116: A threat or promise which, if carried out, costs more tothe agent who issued it than if it is not carried out, iscalled an incredible threat or promise. Page 138: However, this failure topredict should be welcomed by John Rawls and Robert Nozick as it providesan opening to their contrasting views of what counts as justice betweenrational agents. Page 138: If the Nash solution were unique, then game theory would have answeredan important question at the heart of liberal theory over the type of Statewhich rational agents might agree to create. In addition, it would have solveda question in moral philosophy over what justice might demand in this and avariety of social interactions. After all, how to divide the benefits from socialcooperation seems at first sight to involve a tricky question in moralphilosophy concerning what is just, but if rational agents will only ever agreeon the Nash division then there is only one outcome for rational agents.Whether we want to think of this as just seems optional. But if we do or ifwe think that justice is involved, then we will know, and for onceunambiguously, what justice apparently demands between instrumentallyrational agents.Unfortunately, though, it seems we cannot draw these inferences becausethe Nash solution is not the unique outcome. Accepting this conclusion, weare concerned in this section with what bargaining theory then contributes tothe liberal project of examining the State as if it were the result of rational Page 140: behind the veil of ignorance. Page 142: Torture: Another example in moral philosophy is revealed by the problem oftorture for utilitarians. For instance, a utilitarian calculation focuses onoutcomes by summing the individual utilities found in society. In so doing itdoes not enquire about the fairness or otherwise of the processes responsiblefor generating those utilities with the result that it could sanction torture whenthe utility gain of the torturer exceeds the loss of the person being tortured.Yet most people would feel uncomfortable with a society which sanctionedtorture on these grounds because it unfairly transgresses the ‘rights’ of thetortured. Page 143: Granted that society (andthe State) are not the result of some living-room negotiation, what kind of“axioms” would have generated the social outcomes which we observe in agiven society?’ That is, even if we reject the preceding fictions (i.e. of the Stateas a massive resolution of an n-person bargaining game, or of the veil ofignorance) as theoretically and politically misleading, we may still pinpointcertain axioms which would have generated the observed income distributions(or distributions of opportunities, social roles, property rights, etc.) as a resultof an (utterly) hypothetical bargaining game. Page 143: Roemer (1988) considers a problem faced by an international agencycharged with distributing some resources with the aim of improving health(say lowering infant mortality rates). How should the authority distributethose resources? This is a particularly tricky issue because different countriesin the world doubtless subscribe to some very different principles which theywould regard as relevant to this problem; and so agreement on a particularrule seems unlikely. Nevertheless, he suggests that we approach the problemby considering the following constraints (axioms) which we might want toapply to the decision rule because they might be the object of significantagreement. Page 144: rule which allocates resources in such a way as to raise the country with thelowest infant survival rate to that of the second lowest, and then if the budgethas not been exhausted, it allocates resources to these two countries until theyreach the survival rate of the third lowest country, and so on until the budgetis exhausted. Page 147: It is tempting to think that the problem only arises here because theprisoners cannot communicate with one another. If they could get togetherthey would quickly see that the best for both comes from ‘not confessing’.But as we saw in the previous chapter, communication is not all that isneeded. Each still faces the choice of whether to hold to an agreement that Page 148: The recognition ofthis predicament helps explain why individuals might rationally submit to theauthority of a State, which can enforce an agreement for ‘peace’. Theyvoluntarily relinquish some of their freedom that they enjoy in the(hypothesised) state of nature to the State because it unlocks the prisoners’dilemma. (It should be added perhaps that this is not to be taken as a literalaccount of how all States or enforcement agencies arise. The point of theargument is to demonstrate the conditions under which a State or enforcementagency would enjoy legitimacy among a population even though it restrictedindividual freedoms.) Page 148: their normal business with the result that they prosper and enjoy a more‘commodious’ living (as Hobbes phrased it), choosing strategy ‘peace’ is like‘not confessing’ above; when everyone behaves in this manner it is much betterthan when they all choose ‘war’ (’confess’). However, and in spite of wideranging recognition that peace is better than war, the same prisoners’ dilemmaproblem surfaces and leads to war. Page 148: While Hobbes thought that the authority of the State should be absolute soas to discourage any cheating on ‘peace’, he also thought the scope of itsinterventions in this regard would be quite minimal. In contrast much of themodern fascination with the prisoners’ dilemma stems from the fact that theprisoners’ dilemma seems to be a ubiquitous feature of social life. Forinstance, it plausibly lies at the heart of many problems which groups Page 148: they have struck over ‘not confessing’. Is it in the interest of either party tokeep to such an agreement? No, a quick inspection reveals that the bestaction in terms of pay-offs is still to ‘confess’. As Thomas Hobbes remarkedin Leviathan when studying a similar problem, ‘covenants struck without thesword are but words’. The prisoners may trumpet the virtue of ‘notconfessing’ but if they are only motivated instrumentally by the pay-offs,then it is only so much hot air because each will ‘confess’ when the timecomes for a decision. Page 148: What seems to be required to avoid this outcome is a mechanism whichallows for joint or collective decision making, thus ensuring that both actuallydo ‘not confess’. In other words, there is a need for a mechanism for enforcingan agreement—Hobbes’s ‘sword’, if you like. And it is this recognition whichlies at the heart of a traditional liberal argument dating back to Hobbes for thecreation of the State which is seen as the ultimate enforcement agency.(Notice, however, that such an argument applies equally to some otherinstitutions which have the capacity to enforce agreements, for example theMafia.) In Hobbes’s story, each individual in the state of nature can behavepeacefully or in a war-like fashion. Since peace allows everyone to go about Page 149: it is notuncommon to find the dilemma treated as the essential model of social life Page 149: The following four sectionsand the next chapter, on repeated games, discuss some of the developmentsin the social science literature which have been concerned with how thedilemma might be unlocked without the services of the State. In otherwords, the later sections focus on the question of whether the widespreadnature of this type of interaction necessarily points to the (legitimate inliberal terms) creation of an activist State. Are there other solutions whichcan be implemented without involving the State or any public institution?Since the scope of the State’s activities has become one of the mostcontested issues in contemporary politics, it will come as no surprise todiscover that the discussions around alternative solutions to the dilemmahave assumed a central importance in recent political (and especially inliberal and neoliberal) theory. Page 149: It arises as a problem of trust in every elemental economic exchangebecause it is rare for the delivery of a good to be perfectly synchronised withthe payment for it and this affords the opportunity to cheat on the Page 150: These are two-person examples of the dilemma, but it is probably the ‘n-person’ version of the dilemma (usually called the free rider problem) which hasattracted most attention. It creates a collective action problem among groupsof individuals. Again the examples are legion. Page 151: The instrumentally rational individual will recognise that the best action is‘do not attach’ (i.e. defection) whatever the others do. This means that in apopulation of like-minded individuals, all will decide similarly with the resultthat each individual gains 2 utils. This is plainly an inferior outcome for allbecause everyone could have attached the device and if they all had done soeach would have enjoyed 3 utils.In these circumstances the individuals in this economy might agree to theState enforcing attachment of the device. Alternatively, it is easy to see howanother popular intervention by the State would also do the trick. The Statecould tax each individual who did not attach the device a sum equivalent to 2utils and this would turn ‘attach’ (C) into the dominant strategy. Page 151: There is nothinglike the State which can enforce contracts within the household to keep akitchen clean, but interestingly within a family household one oftenobserves the exercise of patriarchal or paternal power instead. Of course,the potential difficulty with such an arrangement is that the patriarch mayrule in a partial manner with the result that the kitchen is clean but with nohelp from the hands of the patriarch! The role of the State has in suchcases been captured, so to speak, by an interested party determined bygender. Then gender becomes the determinant of who bears the burdenand who has the more privileged role. Social power which ‘solves’prisoners’ dilemmas can be thus exercised without the direct involvementof the State (even though the State often enshrines such power in its owninstitutions). Page 152: Hence the prisoners’ dilemma/free rider might plausibly lie atthe distinction which is widely attributed to Marx in the discussion of classconsciousness between a class ‘of itself’ and ‘for itself’ (see Elster, 1986b). Onsuch a view a class transforms itself into a ‘class for itself’, or a society avoidsdeficient demand, by unlocking the dilemma. Page 153: Adam Smith’s account of how the self-interest of sellers combines with thepresence of many sellers to frustrate their designs and to keep prices lowmight also fit this model of interaction. If you are the seller choosing from thetwo row strategies C and D, then imagine that C and D translate into ‘charge ahigh price’ and ‘charge a low price’ respectively. Figure 5.2 could reflect yourpreference ordering as high prices for all might be better than low prices forall and charging a low price when all others charge a high might be the bestoption because you scoop market share. Presumably the same applies to yourcompetitors. Thus even though all sellers would be happier with a high level ofprices, their joint interest is subverted because each acting individually quiterationally charges a low price. It is as if an invisible hand was at work onbehalf of the consumers. Page 155: This is perhaps the most radical departure from the conventionalinstrumental understanding of what is entailed by rationality because, whileaccepting the pay-offs, it suggests that agents should act in a different wayupon them. The notion of rationality is no longer understood in the means—end framework as the selection of the means most likely to satisfy given ends. Page 155: thus enabling‘rationality’ to solve the problem when there are sufficient numbers ofKantian agents. Page 155: For instance, we mighthave wrongly assumed earlier that there is no honour among thieves becauseacting honourably could be connected to acting rationally in some fullaccount of rationality in which case the dilemma might be unlocked withoutthe intervention of the State (or some such agency). This general idea oflinking a richer notion of rational agency with the spontaneous solution ofthe dilemma has been variously pursued in the social science literature andthis section and the following three consider four of the more prominentsuggestions. Page 155: The first connects rationality with morality and Kant provides a readyreference. His practical reason demands that we should undertake thoseactions which when generalised yield the best outcomes. It does not matterwhether others perform the same calculation and actually undertake thesame action as you. The morality is deontological and it is rational for theagent to be guided by a categorical imperative (see Chapter 1). Consequently,in the free rider problem, the application of the categorical imperative willinstruct Kantian agents to follow the cooperative action Page 156: Similarly partisans in occupied Europe during the Second World War riskedtheir lives even when it was not clear that it was instrumentally rational toconfront the Nazis. In such cases, it seems people act on a sense of what isright. Page 156: Likewise, Hardin (1982) suggests thatthe existence of environmental and other voluntary organisations usuallyentails overcoming a free rider problem and in the USA this may beexplained in part by an American commitment to a form ofcontractarianism whereby ‘people play fair if enough others do’ Page 156: Instead, rationality is conceived more as an expression of what is possible: ithas become an end in its own right. This is not only radical, it is alsocontroversial. Deontological moral philosophy is controversial for the obviousreason that it is not concerned with the actual consequences of an action, aswell as for the move to connect it with rationality. (Nevertheless, O’Neill(1989) presents a recent argument and provides an extended discussion of thismoral psychology and how it might be applied.)Kant’s morality may seem rather demanding for these reasons, but thereare weaker or vaguer types of moral motivation which also seem capableof unlocking the prisoners’ dilemma. For example, a general altruisticconcern for the welfare of others may provide a sufficient reason forpeople not to defect on the cooperative arrangement. Page 157: Another departure from the strict instrumental model of rational action comeswhen individuals make decisions in a context of norms and these norms arecapable of overriding considerations of what is instrumentally rational. Page 157: On the other hand, given the well-known difficultiesassociated with any coherent system of ethics (like utilitarianism), it seemsquite likely that a person’s ethical concerns will not be captured by a well-behaved set of preferences (see for instance Sen (1970) on the problems ofbeing a Paretian Liberal). Indeed rational agents may well base their actions onreasons which are external to their preferences. Page 157: Of course, there is a tricky issue concerning whether these rather weaker orvaguer moral motivations (like altruism, acting on what is fair or what is right)mark a deep breach with the instrumental model of action. It might be arguedthat such ethical concerns can be represented in this model by introducing theconcept of ethical preferences. Thus the influence of ethical preferencestransforms the pay-offs in the game. Page 158: Disputeswithin Aboriginal society are neither perceived as simply between twoindividuals nor subject to some established community tribunal. It is for thisreason that the resolution of a major conflict will involve a significant amountof negotiation between the parties. Yet the informal laws which govern thecontents of the negotiations are well entrenched in the tribal culture. Forexample, it is not uncommon for family members of the perpetrator to beasked to accept ‘punishment’ if the individual offender is in prison andtherefore unavailable. Page 159: First World War. This was a war of unprecedentedcarnage both at the beginning and the end. Yet during a middle period, non-aggression between the two opposing trenches emerged spontaneously in theform of a ‘live and let live’ norm. Christmas fraternisation is one well-knownexample, but the ‘live and let live’ norm was applied much more widely.Snipers would not shoot during meal times and so both sides could go abouttheir business ‘talking and laughing’ at these hours. Artillery was predictablyused both at certain times and at certain locations. So both sides couldappear to demonstrate aggression by venturing out at certain times and tocertain locations, knowing that the shells would fall predictably close to, butnot on, their chosen route. Likewise, it was not considered ‘etiquette’ to fireon working parties who had been sent out to repair a position or collect thedead and so on. Page 159: For instance, it is sometimes argued that thenorms of Confucian societies enable those economies to solve the prisoners’dilemma/free rider problems within companies without costly contracting andmonitoring activity and that this explains, in part, the economic success ofthose economies (see Hargreaves Heap, 1991, Casson, 1991, North, 1991).Akerlof ’s (1983) discussion of loyalty filters, where he explains the relativesuccess of Quaker groups in North America by their respect for the norm ofhonesty, is another example— Page 160: Wittgenstein of Philosophical Investigations1 is an obvious source for this viewbecause he would deny that the meaning of something like a person’sinterests or desires can be divorced from a social setting; and this is a usefulopportunity to take that argument further. The attribution of meaningrequires language rules and it is impossible to have a private language. Thereis a long argument around the possibility or otherwise of private languagesand it may be worth pursuing the point in a slightly different way by askinghow agents have knowledge of what action will satisfy the condition ofbeing instrumentally rational. Any claim to knowledge involves a firstunquestioned premise: I know this because I accept x. Otherwise an infiniteregress is inevitable: I accept x because I accept y and I accept ybecause…and so on. Accordingly, if each person’s knowledge of what isrational is to be accessible to one another, then they must share the samefirst premises. It was Wittgenstein’s point that people must share somepractices if they are to attach meaning to words and so avoid the problem ofinfinite redescription which comes with any attempt to specify the rules forapplying the rules of a language. Page 161: There is another similarity and difference which might also be usefullymarked. To make it very crudely one might draw an analogy between thedifficulty which Wittgenstein encounters over knowledge claims and a similardifficulty which Simon (1982) addresses. (Herbert Simon is well known ineconomics for his claim that agents are procedurally rational, or boundedlyrational, because they do not have the computing capacity to work out whatis the best to do in complex settings.) To be sure, Wittgenstein finds theproblem in an infinite regress of first principles while Simon finds thedifficulty in the finite computing capacity of the brain. Nevertheless, both Page 161: discussion of the Harsanyi doctrine because a similar claim seems to underpinthat doctrine. Namely that all rational individuals must come to the sameconclusion when faced by the same evidence. Wittgenstein would agree to theextent that some such shared basis of interpretation must be present ifcommunication is to be possible. But he would deny that all societies andpeoples will share the same basis for interpretations. The source of the sharingfor Wittgenstein is not some universal ‘rationality’, as it is for Harsanyi; ratherit is the practices of the community in which the people live, and these willvary considerably across time and space. Page 162: let us make the view inspired by Wittgensteinvery concrete. The suggestion is that what is instrumentally rational is notwell defined unless one appeals to the prevailing norms of behaviour. Thismay seem a little strange in the context of a prisoners’ dilemma where thedemands of instrumental rationality seem plain for all to see: defect! But,in reply, those radically inspired by Wittgenstein would complain that thenorms have already been at work in the definition of the matrix and itspay-offs because it is rare for any social setting to throw up unvarnishedpay-offs. A social setting requires interpretation before the pay-offs can beassigned and norms are implicated in those interpretations. (See forexample Polanyi (1945) who argues, in his celebrated discussion of the riseof industrial society, that the incentives of the market system are onlyeffective when the norms of society place value on private materialadvance.) Page 162: The last reflection on rationality comes from David Gauthier. He remainsfirmly in the instrumental camp and ambitiously argues that its dictates havebeen wrongly understood in the prisoners’ dilemma game. Instrumental rationalitydemands cooperation and not defection! To make his argument he distinguishesbetween two sorts of maximisers: a straightforward maximiser (SM) and aconstrained maximiser (CM). A straightforward maximiser defects (D)following the same logic that we have used so far. The constrained maximiseruses a conditional strategy of cooperating (C) with fellow constrainedmaximisers and defecting with straightforward maximisers. He then asks:which disposition (straightforward or constrained) should an instrumentallyrational person choose to have? (The decision can be usefully compared with asimilar one confronting Ulysses in connection with listening to the Sirens, Page 164: The point is that if instrumental rationality is what motivates the CM inthe prisoners’ dilemma, then a CM must want to defect Page 164: In other words, being a CM may be better than beingan SM, but the best strategy of all is to label yourself a CM and then cheaton the deal. And, of course, when people do this, we are back in a worldwhere everyone defects. Page 164: Surely, this line of argument goes,it pays not to ‘zap’ a fellow CM because your reputation as a CM is therebypreserved and this enables you to interact more fruitfully with fellow CMs inthe future. Should you zap a fellow CM now, then everyone will know that youare a rogue and so in your future interactions, you will be treated as an SM. Inshort, in a repeated setting, it pays to forgo the short run gain from defectingbecause this ensures the benefits of cooperation over the long run. Thusinstrumental calculation can make true CM behaviour the best course ofaction. Page 165: Moreover, it achieved aremarkable degree of cooperation. Page 165: each program.Tit-for-Tat, submitted by Anatol Rapoport, won the tournament. Theprogram starts with a cooperative move and then does whatever theopponent did on the previous move. It was, as Axelrod points out, not onlythe simplest program, it was also the best! Page 165: dilemma can be defeated without the intervention of a collective agency likethe State—that is, provided the interaction is repeated sufficiently often tomake the long term benefits outweigh the short gains. Page 166: ‘Is the Prisoners’ dilemma all of sociology?’Of course, it is not, he answers. Nevertheless, it has fascinated social scientistsand proved extremely difficult to unlock in one-shot plays of the game—atleast, without the creation of a coercive agency like the State which is capableof enforcing a collective action or without the introduction of norms or somesuitable form of moral motivation on the part of the individuals playing thegame. Of course, many interactions are repeated and so this stark conclusionmay be modified by the discussion of the next chapter. Page 167: Perhaps somewhat surprisingly, mutualdefection remains the only Nash equilibrium. The following two sectionsdiscuss, respectively, indefinitely repeated prisoners’ dilemma and therelated free rider games. We show (section 6.4) that mutual cooperation isa possible Nash equilibrium outcome in these games provided there is a‘sufficient’ degree of uncertainty over when the repetition will cease.There are some significant implications here both for liberal politicaltheory and for the explanatory power of game theory. We notice that thisresult means that mutual cooperation might be achieved without theintervention of a collective agency like the State and/or withoutappealing to some expanded notion of rational agency Page 168: the absence of a theory of equilibriumselection. Page 170: Firstly, it provides a theoretical warrant for the belief that cooperation in theprisoners’ dilemma can be rationally sustained without the intervention ofsome collective agency like the State, provided there is sufficient (to be definedlater) doubt over when the repeated game will end. Thus the presence of aprisoners’ dilemma interaction does not necessarily entail either a poor socialoutcome or the institutions of formal collective decision making. The thirdalternative is for players to adopt a tit-for-tat strategy rationally.1 If they adoptthis third alternative the socially inferior outcome of mutual defection will beavoided without the interfering presence of the State or some other formal(coercive) institution. Page 171: Equally, it is probable that both prisonersin the original example may think twice about ‘confessing’ because each knowsthat they are likely to encounter one another again (if not in prison, at leastoutside) and so there are likely to be opportunities for exacting ‘punishment’ ata later date. Page 171: Folk theorem Page 172: This is an extremely important result for the social sciences because itmeans that there are always multiple Nash equilibria in such indefinitelyrepeated games. Hence, even if Nash is accepted as the appropriateequilibrium concept for games with individuals who are instrumentally rationaland who have common knowledge of that rationality, it will not explain howindividuals select their strategies because there are many strategy pairs whichform Nash equilibria in these repeated games. Of course, we have encounteredthis problem in some one-shot games before, but the importance of this resultis that it means the problem is always there in indefinitely repeated games.Even worse, it is amplified by repetition. In other words, game theory needs tobe supplemented by a theory of equilibrium selection if it is to explain actionin these indefinitely repeated games, especially if it is to explain howcooperation actually arises spontaneously in indefinitely repeated prisoners’dilemma games. Page 175: Now consider a tit-for-tat strategy in this group which works in thefollowing way. The strategy partitions the group into those who are in ‘goodstanding’ and those who are in ‘no standing’ based on whether the individualcontributed to the collective fund in the last time period. Those in ‘goodstanding’ are eligible for the receipt of help from the group if they fall ‘ill’ thistime period, whereas those who are in ‘no standing’ are not eligible for help.Thus tit-for-tat specifies cooperation and puts you in ‘good standing’ for thereceipt of a benefit if you fall ‘ill’ (alternatively, to connect with the earlierdiscussion, one might think of cooperating as securing a ‘reputation’ whichputs one in ‘good standing’ Page 175: Notice your decision now will determine whether you are in ‘good standing’from now until the next opportunity that you get to make this decision (whichwill be the next period if you do not fall ‘ill’ or the period after that if you fall‘ill’). So we focus on the returns from your choice now until you next get theopportunity to choose. Page 176: Who needs the State? Page 176: Here we pick up threads of the Hobbesianargument for the State and see what the result holds for this argument. At firstglance, the argument for the State seems to be weakened because it appearsthat a group can overcome the free rider problem without recourse to theState for contract enforcement. So long as the group can punish free riders byexcluding them from the benefits of cooperation (as for instance the Pygmiespunished Cephu—see Chapter 5), then there is the possibility of ‘spontaneous’public good provision through the generalisation of the tit-for-tat strategy.Having noted this, nevertheless, the point seems almost immediately to beblunted since the difference between a Hobbesian State which enforcescollective agreements and the generalised tit-for-tat arrangement is notaltogether clear and so in proving one we are hardly undermining the other.After all, the State merely codifies and implements the policies of‘punishment’ on behalf of others in a very public way (with the rituals ofpolice stations, courts and the like). But, is this any different from the golfclub which excludes a member from the greens when the dues have not beenpaid or the Pygmies’ behaviour towards Cephu? Or the gang which excludespeople who have not contributed ‘booty’ to the common fund? Page 176: Box 6.2 Page 178: contract—that the creation of the State by the individual also helps shape asuperior individual.) Hayek, however, prefers the ‘English tradition’ because hedoubts (a) that the formation of the State is part of a process which liberates(and moulds) the social agent and (b) that there is the knowledge to informsome central design so that it can perform the task of resolving free ridingbetter than spontaneously generated solutions (like tit-for-tat). In other words,reason should know its limits and this is what informs Hayek’s support forEnglish pragmatism and its suspicion of the State.Of course there is a big ‘if in Hayek’s argument. Although Beirut stillmanaged to function without a grand design, most of its citizens prayed forone. In short, the spontaneous solution is not always the best. Indeed, as wehave seen, the cooperative solution is just one among many Nash equilibria inrepeated games, so in the absence of some machinery of collective decisionmaking, there seems no guarantee it will be selected. Against this, however, itis sometimes argued that evolution will favour practices which generate thecooperative outcome since societies that achieve cooperation in these gameswill prosper as compared with those which are locked in mutual defection.This is the cue for a discussion of evolutionary game theory and we shall leavefurther discussion of the State until we turn to evolutionary game theory Page 178: Instead the result seems important because it demythologises the State.Firstly the State qua State (that is, the State with its police force, its courts andthe like) is not required to intrude into every social interaction which suffersfrom a free rider problem. There are many practices and institutions which aresurrogates for the State in this regard. Indeed, the Mafia has plausiblydisplaced the State in certain areas precisely because it provides the services ofa State. Likewise, during the long civil war years inhabitants of Beirutsomehow still managed to maintain services which required the overcoming offree rider problems.Secondly since something like the State as contract enforcer might well arise‘spontaneously’ through the playing of free rider games repeatedly, it need notrequire any grand design. There need be no constitutional conventions. In thisway the result counts strongly for what Hayek (1962) refers to as the Englishas opposed to the European continental Enlightenment tradition. The latterstresses the power of reason to construct institutions that overcome problemslike those of the free rider. (It also often presupposes—recall Rousseau’s social Page 178: different if you pay the State in the form of taxes or the Mafia in the form oftribute? Page 185: example comes from strategicdecisions by the legislature when the Executive is trying to push throughParliament a series of bills that the latter is unsympathetic towards. Page 185: President proposes legislation. The Congress is notin sympathy with the proposal and must decide whether to make amendments.If it decides to make an amendment, then the President must decide whetherto fight the amendment or acquiesce. Looking at the President’s pay-offs it isobvious that, even though he or she prefers that the Congress does not amendthe legislation, if it does, he or she would not want to fight Page 186: the Folk theorem ensures that aninfinity of war/acquiescence patterns are compatible with instrumentalrationality. Nevertheless, the duration of such games is usually finite andsometimes their length is definite—e.g. US Presidents have a fixed term andincumbents have only a fixed number of local markets that they wish todefend. What happens then? Would it make sense for the President or theincumbent to put on a show of strength early on (e.g. by fighting the Congressor unleashing a price war) in order to create a reputation for belligerence thatwould make the Congress and the entrant think that, in future rounds, theywill end up with pay-off -1/2 if they dare them? Page 186: In the finitely repeated version of the game Nash backward inductionargues against this conclusion. Just as in the case of the prisoners’ dilemmain the previous subsection, it suggests that, since there will be no fighting atthe last play of the game, the reputation of the President/incumbent willunravel to the first stage and no fighting will occur (rationally). Theconclusion changes again once we drop CKR (or allow for different types ofplayers). Page 190: Of course, there may be actions that can be takenoutside the game and which have a similar effect on the beliefs of others. Such‘signalling’ behaviour is considered briefly in this section to round out thediscussion of reputations. It is of potential relevance not only to repeated, butalso to one-shot games. Page 192: when the game isrepeated and there is a unique Nash equilibrium things change. The Nashequilibrium is attractive because as time goes by and agents adjust theirexpectations of what others will do in the light of experience, then they willseem naturally drawn to the Nash equilibrium because it is the only restingplace for beliefs. Any other set of beliefs will upset itself. Page 192: Nevertheless, there is still no guarantee that a Nash equilibrium willsurface even if it exists and it is unique. Page 193: The strength of the Nash equilibrium is that forward looking agents mayrealise that (R2, C2) is the only outcome that does not engender such thoughts.We just saw that adaptive (or backward looking) expectations will not do thetrick. If, however, after having been around the pay-off matrix a few timesplayers ask themselves the question ‘How can we reach a stable outcome?’,they may very well conclude that the only such outcome is the Nashequilibrium (R2, C2).But why would they want to ask such a question? What is so wrong withinstability (and disequilibrium) after all? Indeed in the case of Figure 2.6 ourplayers have an incentive to avoid a stable outcome (observe that on averagethe cycle which takes them from one extremity of the pay-off matrix toanother yields a much higher pay-off than the Nash equilibrium result). If, onthe other hand, pay-offs were as in Figure 6.4 below, they would be stronglymotivated to reach the Nash equilibrium. Page 193: It is easy to see that this type of adaptivelearning will never lead the players to the Nash equilibrium outcome (R2, C2).Instead, they will be oscillating between outcomes (R1, C1), (R1, C3), (R3, C1)and (R3, C3).Can they break away from this never ending cycle and hit the Nashequilibrium? They can provided they converge onto a common forwardlooking train of thought. For Page 194: Thus we conclude that whether repetition makes the Nashequilibrium more or less likely when it is unique must depend on thecontingencies of how people learn and the precise pay-offs from non-Nashbehaviour. Page 194: Broadly put, this is one and the same problem. It is a problem withspecifying how agents come to hold beliefs which are extraneous to the game(in the sense that they cannot be generated endogenously through theapplication of the assumptions of instrumental rationality and commonknowledge of instrumental rationality) Page 195: the insights of evolutionary game theory arecrucial material for many political and philosophical debates, especially thosearound the State. Page 195: The argument for suchan agency turns on the general problem of equilibrium selection and on theparticular difficulty of overcoming the prisoners’ dilemma. When there aremultiple equilibria, the State can, through suitable action on its own part, guidethe outcomes towards one equilibrium rather than another. Thus the problemof equilibrium selection is solved by bringing it within the ambit of consciouspolitical decision making. Likewise, with the prisoners’ dilemma/ free riderproblem, the State can provide the services of enforcement. Alternativelywhen the game is repeated sufficiently and the issue again becomes one ofequilibrium selection, then the State can guide the outcomes towards thecooperative Nash equilibrium. Page 195: intransigent Right’ Page 196: —that is, the idea that you can turn social outcomes intomatters of social choice through the intervention of a collective action agencylike the State. The positive argument against ‘political rationalism’, as the quoteabove suggests, turns on the idea that these interventions are not evennecessary. The failure to intervene does not spell chaos, chronic indecision,fluctuations and outcomes in which everyone is worse off than they couldhave been. Instead, a ‘spontaneous order’ will be thrown up as a result ofevolutionary processes. Page 196: Likewise, there are problems of ‘political failure’ that subvert the ideal ofdemocratic decision making and which can match the market failures that theState is attempting to rectify. For example, Buchanan and Wagner (1977) andTullock (1965) argue that special interests are bound to skew ‘democraticdecisions’ towards excessively large bureaucracies and high governmentexpenditures. Furthermore there are difficulties, especially after the Arrowimpossibility theorem, with making sense of the very idea of something likethe ‘will of the people’ in whose name the State might be acting (see Arrow,1951, Riker, 1982, Hayek, 1962, and Buchanan, 1954).1These, so to speak, are a shorthand list of the negative arguments comingfrom the political right against ‘political rationalism’ or ‘socialconstructivism’ Page 196: Forinstance, there are problems of inadequate knowledge which can mean thateven the best intentioned and executed political decision generates unintendedand undesirable consequences. Indeed this has always been an importanttheme in Austrian economics, featuring strongly in the 1920s debate over thepossibility of socialist planning as well as contemporary doubts over thewisdom of more minor forms of State intervention. Page 196: Hayek (1962) himself tracesthe battlelines in the dispute back to the beginning of Enlightenmentthinking:Hayek distinguished two intellectual lines of thought about freedom, ofradically opposite upshot. The first was an empiricist, essentially Britishtradition descending from Hume, Smith and Ferguson, and seconded byBurke and Tucker, which understood political development as aninvoluntary process of gradual institutional improvement, comparable tothe workings of a market economy or the evolution of common law. Thesecond was a rationalist, typically French lineage descending fromDescartes through Condorcet to Comte, with a horde of modernsuccessors, which saw social institutions as fit for premeditatedconstruction, in the spirit of polytechnic engineering. The former lineled to real liberty; the latter inevitably destroyed it. Page 197: evolutionary stable strategies Page 197: In particular, wesuggest that the evolutionary approach can help elucidate the idea that poweris mobilised through institutions and conventions. We conclude the chapterwith a summing-up of where the issue of equilibrium selection and the debateover the State stands after the contribution of the evolutionary approach. Page 197: The basic idea behind this equilibrium concept is that an ESS is a strategywhich when used among some population cannot be ‘invaded’ by anotherstrategy because it cannot be bested. So when a population uses a strategy I,‘mutants’ using any other strategy J cannot get a toehold and expand amongthat population. Page 197: This is why evolutionary game theory assumes significance in the debateover an active State. It should help assess the claims of ‘spontaneous order’made by those in the British corner and so advance one of the central debatesin Enlightenment political thinking. Page 198: This is, if youlike, a version of Hobbes’s nightmare where there are no property rightsand everyone you come across will potentially claim your goods. Page 202: Secondly, and more specifically, there is the result that although the symmetricalplay of this game yields a unique equilibrium, it becomes unstable the momentrole playing begins and some players start to recognise asymmetry. Sincecreative agents seem likely to experiment with different ways of playing thegame, it would be surprising if there was never some deviation based on anasymmetry. Indeed it would be more than surprising because there is muchevidence to support the idea that people look for ‘extraneous’ reasons whichmight explain what are in fact purely random types of behaviour (see theadjacent box on winning streaks).Formally, this leaves us with the old problem of how the solution to thegame comes about. However, evolutionary game theory does at least point usin the direction of an answer. The phase diagram in Figure 7.2 reveals that theselection of an equilibrium depends critically on the initial set of beliefs Page 202: once animperfect form of rationality is posited. In other words, it is not beingdeduced as an implication of the common knowledge of rationalityassumption which has been the traditional approach of mainstream gametheory. Page 203: Thirdly, it can be noted that the selection of one ESS rather than anotherembodies a convention Page 203: To put these observationsrather less blandly, since rationality on this account is only responsible for thegeneral impulse towards mimicking profitable behaviour, the history of thegame depends in part on what are the idiosyncratic and unpredictable (non-rational, one might say, as opposed to irrational) features of individual beliefsand learning. Page 204: Fourthly, the selection of one equilibrium rather than another potentiallymatters rather deeply. In effect in the hawk—dove game over contestedproperty, what happens in the course of moving to one of the ESSs is theestablishment of a form of property rights. Either those playing role R get theproperty and role C players concede this right, or those playing role C get theproperty and role R players concede this right. This is interesting not onlybecause it contains the kernel of a possible explanation of property rights (onwhich we shall say more later) but also because the probability of playing roleR or role C is unlikely to be distributed uniformly over the population. Indeed,this distribution will depend on whatever is the source of the distinction usedto assign people to roles. Page 204: The question, then, of how a source of differentiation gets establishedbecomes rather important. Page 204: Thus the behaviour at one of theseESSs is conventionally determined and, to repeat the earlier point, we can plotthe emergence of a particular convention with the use of this phase diagram.It will depend both on the presumption that agents learn from experience(the rational component of the explanation) and on the particularidiosyncratic (and non-rational) features of initial beliefs and precise learningrules. Page 206: After all, perhaps the presence of these conventions can only beaccounted for by a move towards a Wittgensteinian ontology, in which casemainstream game theory’s foundations look decidedly wobbly. To prevent thisdrift a more robust response is required.The alternative response is to deny that the appeal to shared prominenceor salience involves either an infinite regress or an acknowledgement thatindividuals are necessarily ontologically social Page 206: There is a further and deeper problem with the concept of salience basedon analogy because the attribution of terms like ‘possession’ plainly begs thequestion by presupposing the existence of some sort of property rights in thepast. In other words, people already share a convention in the past and this isbeing used to explain a closely related convention in the present. Thus we havenot got to the bottom of the question concerning how people come to holdconventions in the first place.3 Page 206: So, of course, we cannot hope to explainhow they actually achieve a new coordination without appealing to thosebackground conventions. In this sense, it would be foolish for socialscientists (and game theorists, in particular) to ignore the social context inwhich individuals play new games. Page 208: This conclusion reinforces the earlier result that the course of historydepends in part on what seem from the instrumental account of rationalbehaviour to be non-rational (and perhaps idiosyncratic) and thereforefeatures of human beliefs and action which are difficult to predict Page 209: mechanically. One can interpret this in the spirit of methodologicalindividualism at the expense of conceding that individuals are, in this regard,importantly unpredictable. On the one hand, this does not look good for theexplanatory claims of the theory. On the other hand, to render theindividuals predictable, it seems that they must be given a shared history andthis will only raise the methodological concern again of whether we canaccount for this sharing satisfactorily without a changed ontology. Insummary, if individuals are afforded a shared history, then social context is‘behind’ no one and ‘in’ everyone and then the question is whether it is agood idea to analyse behaviour by assuming (as methodological individualistsdo) the separability of context and action.4 Page 213: The underlying point here is that discrimination may be evolutionarystable if the dominated cannot find ways of challenging the social conventionthat supports their subjugation. This conclusion is not necessarily rightbecause there are other potential sources of change. The insight that we preferto draw is that individual attempts to buck an established convention areunlikely to succeed, whereas the same is not true when individuals takecollective action. Page 213: Stasis, status quo: Thus the introduction of a convention will benefit the average person, butif you happen to be so placed with respect to the convention that you onlyplay the dominant role with a probability of less than 1/3, then you would bebetter off without the convention. This result may seem puzzling at first: whydo the people who play a dominant role less than 1/3 of the time not revert tothe symmetric play of the game and so undermine the convention? The answeris that even though the individual would be better off if everyone quit theconvention, it does not make sense to do so individually. After all, aconvention will tell your opponent to play either H or D, and then instruct youto play D or H respectively; and you can do no better than follow thisconvention since the best reply to H remains D and likewise the best reply toD is H. It is just tough luck if you happen to get the D instruction all thetime!We take the force of this individual calculation to be a powerful contributorto the status quo and it might seem to reveal that evolutionary processes yieldto stasis. Page 213: Conventions, inequality and revolt Page 214: To summarise, we should expect a convention to emerge even though itmay not suit everyone, or indeed even if it short-changes the majority. It maybe discriminatory, inequitable, non-rational, indeed thoroughly disagreeable, yetsome such convention is likely to arise whenever a social interaction like hawk-dove is repeated. Which convention emerges will depend on the sharedsalience of extraneous features of the interaction, initial beliefs and the waythat people learn. Page 214: Standstill: A potential weakness of evolutionary game theory has just becomeapparent. Once the bandwagon has come to a standstill, and one conventionhas been selected, the theory cannot account for a potential subversion of theestablished convention. Such an account would require, as we argued in theprevious paragraph, an understanding of political (that is, collective) actionbased on a more active form of human agency than the one provided byinstrumental rationality. Can evolutionary game theory go as far? Page 219: Recall the idea of a trembling hand in section 2.7.1 and suppose thatplayers make mistakes sometimes. In particular, when they intend tocooperate they occasionally execute the decision wrongly and they defect. Inthese circumstances, playing t punishes you for the mistake endlessly becauseit means that your opponent defects next round in response to your mistakendefection. If in the next period you cooperate, you are bound to get zapped.If you follow your t-strategy next time, then you will be defecting while youropponent will be cooperating and a frustrating sequence of alternatingdefections and cooperations will ensue. One way out of this bind is toamend t to t’ whereby, if you defect by mistake, then you cooperate twiceafterwards: the first time as a gesture of acknowledging your mistake and thesecond in order to coordinate your cooperative behaviour with that of youropponent. In other words, the amended tit-for-tat instructs you to cooperatein response to a defection which has been provoked by an earlier mistakendefection on your part. Page 219: Eventhough strategy C would do equally well as a reply to t’, if your opponentmade the mistake (last period) then you know that your opponent willcooperate in the next two rounds no matter what you do this period. Thusyour best response in this round is to defect Page 221: Conventions as covert social power Page 221: even more covert power that comes from being able to mould the preferencesand the beliefs of others so that a conflict of interest is not even latentlypresent. Page 221: with the interests of another.It is common in discussions of power to distinguish between the overt andthe covert exercise of power. Thus, for instance, Lukes (1974) distinguishesthree dimensions of power. There is the power that is exercised in the politicalor the economic arena where individuals, or firms, institutions, etc., are able tosecure decisions which favour their interests over others quite overtly. This isthe overt exercise of power along the first dimension. In addition, there is themore covert power that comes from keeping certain items off the politicalagenda. Some things simply do not get discussed in the political arena and inthis way the status quo persists. Yet the status quo advantages some rather thanothers and so this privileging of the status quo by keeping certain issues offthe political agenda is the second dimension of power. Finally, there is the Page 222: The figure of Spartacus captured imaginations over theages, not so much because of his military antics, but because he personifiedthe possibility of liberating the slaves from the beliefs which sustained theirsubjugation. Page 222: this is the power which works through the mind and which dependsfor its influence on the involvement or agreement of large numbers of thepopulation (again connecting with the earlier observation about the force ofcollective action). Page 222: State were consciously to select a convention in these circumstances thenwe might observe the kind of political haggling associated with the overtexercise of power. Naturally when a convention emerges spontaneously, we donot observe this because there is no arena for the haggling to occur, yet theemergence of a convention is no less decisive than a conscious politicalresolution in resolving the conflict of interest.6Evolutionary game theory also helps reveal the part played by beliefs,especially the beliefs of the subordinate group, in securing the power of thedominant group (a point, for example, which is central to Gramsci’s notion ofhegemony and Hart’s contention that the power of the law requires voluntarycooperation). Page 224: Theannexing of virtue can happen as a result of well-recognised patterns ofcognition. Page 224: Of course, like all theories of cognitive dissonance removal,this story begs the question of whether the adjustment of beliefs can do thetrick once one knows that the beliefs have been adjusted for the purpose.Nevertheless, there seem to be plenty of examples of dissonance removal Page 225: Our final illustration of how evolutionary game theory might help sharpenour understanding of debates around power in the social sciences relates tothe question of how gender and race power distributions are constitutedand persist. The persistence of these power imbalances is a puzzle to some. Page 227: Once a convention isestablished in this game, a set of property relations are also established.Hence the convention could encode a set of class relations for this gamebecause it will, in effect, indicate who owns what and some may end upowning rather a lot when others own scarcely anything. However, as wehave seen a convention of this sort will only emerge once the game isplayed asymmetrically and this requires an appeal to some piece ofextraneous information like sex or age or race, etc. In short, the creationof private property relations from the repeated play of these gamesdepends on the use of some other asymmetry and so it is actuallyimpossible to imagine a situation of pure class relations, as they couldnever emerge from an evolutionary historical process. Or to put thisslightly differently: asymmetries always go in twos!This understanding of the relation has further interesting implications.For instance, an attack on gender stratification is in part an attack on classstratification and vice versa. Page 227: Likewise, however, it would be wrong toimagine that the attack on either if successful would spell the end of theother. Page 227: On this account of powerthrough the working of convention, the ideological battle aimed atpersuading people not to think of themselves as subordinate is half thebattle because these beliefs are part of the way that power is mobilised. Page 228: . The feedback mechanism, however, ispresent in this analysis and it arises because there is ‘learning’. It is theassumption that people shift towards practices which secure better outcomes(without knowing quite why the practice works for the best) which is thefeedback mechanism responsible for selecting the practices. Thus in the debateover functional explanation, the analysis of evolutionary games lends supportto van Parijs’s (1982) argument that ‘learning’ might supply the generalfeedback mechanism for the social sciences which will license functionalexplanations in exactly the same way as natural selection does in the biologicalsciences. Page 228: effect, the explanation of gender and racial inequalities using thisevolutionary model is an example of functional argument. Page 228: The differencebetween men and women or between whites and blacks has no merit inthe sense that it does not explain why the differentiation persists. Thedifferentiation has the unintended consequence of helping the populationto coordinate its decision making in settings where there are benefitsfrom coordination. It is this function of helping the population to selectan equilibrium in a situation which would otherwise suffer from theconfusion of multiple equilibria which explains the persistence of thedifferentiation. Page 229: So far, however, the difference between the two camps (H&EVGT andMarx) is purely based on value judgements: one argues that illusory moralsare good for all, the other that they are not. In this sense, both canprofitably make use of the analysis in evolutionary game theory. Indeed, aswe have already implied in section 7.3.4, a radical political project grounded Page 229: On the side of H&EVGT, Hume thinks that suchillusions play a positive role (in providing the ‘cement’ which keeps societytogether) in relation to the common good. So do neo-Humeans (like Sugden)who are, of course, less confident that invocation of the ‘common good’ is agood idea (as we mentioned in section 7.6.2) but who are still happy to seeconventions (because of the order they bring) become entrenched in sociallife even if this is achieved with the help of a few moral ‘illusions’. On theother side, however, Marx insists that moral illusions are never a good idea(indeed he dislikes all illusions). Especially since, as he sees it, their socialfunction is to help some dreadful conventions survive (recall how in section7.3.4 we showed that disagreeable conventions may become stable even ifthey are detrimental to the majority). Marx believed that we can Page 229: which sound quite like observations that Marxists might make: theimportance of taking collective action if one wants to change a convention;how power can be covertly exercised; how beliefs (particularly moral beliefs)may become endogenous to the conventions we follow; how propertyrelations might develop functionally; and so on. Page 229: Indeedmost of the ideas developed on the basis of H&EVGT in the precedingpages would find Marx in agreement. Page 229: People may think that their beliefson such matters go beyond material values (i.e. self-interest, which in ourcontext means pay-offs); that they respond to certain universal ideals aboutwhat is ‘good’ and ‘right’, when all along their moral beliefs are a direct(even if unpredictable) repercussion of material conditions and interests. Page 230: An analysis of hawk—dove games, along the lines of H&EVGT, helpsexplain the evolution of property rights in primitive societies. Once theserights are in place and social production is under way, each group in society(e.g. the owners of productive means, or those who do not own tools, land,machines, etc.) develops its own interest. And since (as H&EVGT concurs)conventions evolve in response to such interests, it is not surprising thatdifferent conventions are generated within different social groups in responseto the different interests. The result is conflicting sets of conventions which Page 230: Finally, the established (stable) conventions acquire moral weight and even leadpeople to believe in something called the common good—which is most likelyanother illusion Page 230: In summary, H&EVGTbegins with a behavioural theory based on the individual interest and eventuallylands on its agreeable by-product: the species interest. There is nothing inbetween the two types of interest. By contrast, Marx posits another type ofinterest in between: class interest.Marx’s argument is that humans are very different from other speciesbecause we produce commodities in an organised way before distributingthem. Whereas other species share the fruits of nature (hawk—dove games aretherefore ‘naturally’ pertinent in their state of nature), humans have developedcomplex social mechanisms for producing goods. Naturally, the norms ofdistribution come to depend on the structure of these productive mechanisms.They involve a division of labour and lead to social divisions (classes). Whichclass a person belongs to depends on his or her location (relative to others)within the process of production. The moment collective production (as in thecase of Cephu and his tribe in Chapter 5) gave its place to a separationbetween those who owned the tools of production and those who workedthose tools, then groups with significantly different (and often contradictory)interests developed. Page 230: in collective action is as compatible with evolutionary game theory as is theneo-Humeanism of Sugden (1986, 1989). But is there something more inMarx than a left wing interpretation of evolutionary game theory? We thinkthere is. Page 231: lead to conflicting morals. Each set of morals becomes an ideology.9 Which setof morals (or ideology) prevails at any given time? Marx thinks that, inevitably,the social class which is dominant in the sphere of production and distributionwill also be the one whose set of conventions and morals (i.e. whose ideology)will come to dominate over society as a whole.To sum up Marx’s argument so far, prevailing moral beliefs are illusoryproducts of a social selection process where the driving force is not somesubjective individual interest but objective class interest rooted in thetechnology and relations of production. Although there are many conflictingnorms and morals, at any particular time the morality of the ruling class isuniquely evolutionary stable. The mélange of legislation, moral codes, norms,etc., reflects this dominant ideology.But is there a fundamental difference between the method of H&EVGTand Marx? Or is it just a matter of introducing classes in the analysiswithout changing the method? Page 231: So, how would Marx respond to evolutionary game theory if he werearound today? He would, we think, be very interested in some of the radicalconclusions in this chapter. However, he would also speak derisively of thematerialism of H&EVGT Marx habitually poured scorn on those (e.g.Spinoza and Feuerbach) who transplanted models from the natural sciencesto the social sciences with little or no modification to allow for the fact thathuman beings are very different to atoms, planets and molecules.12 Wemention this because at the heart of H&EVGT lies a simple Darwinianmechanism (witness that there is no analytical difference between the modelsin the biology of John Maynard Smith and the models in this chapter). Marxwould probably claim that the theory is not sufficiently evolutionary because(a) its mechanism comes to a standstill once a stable convention has evolved,and (b) of its reliance on instrumental rationality which reduces humanactions to passive reflex responses to some (meta-physical) self-interest. Page 232: Especially in hisphilosophical (as opposed to economic) works, Marx argued strongly for anevolutionary (or more precisely historical) theory of society with a modelof human agency which retains human activity as a positive (creative) forceat its core. In addition, Marx often spoke out against mechanism; againstmodels borrowed directly from the natural sciences (astronomy andbiology are two examples that he warned against). It is helpful to preservesuch an aversion since humans are ontologically different to atoms andgenes. Of course Marx himself has been accused of mechanism and,indeed, in the modern (primarily Anglo-Saxon) social theory literature he istaken to be an exemplar of 19th century mechanism. Nevertheless hewould deny this, pointing to the dialectical method he borrowed fromHegel and which (he would claim) allowed him to have a scientific, yetnon-mechanistic, outlook. Do we believe him? As authors we disagree here.SHH does not, while YV does. Page 232: Of course there is always the answer that self-interest feeds into moral beliefsand then moral beliefs feed back into self-interest and alter people’s desires.And so on. But that would be too circular for Marx. It would not explainwhere the process started and where it is going. By contrast, his version ofmaterialism (which he labelled historical materialism) starts from thetechnology of production and the corresponding social organisation. Thelatter entails social classes which in turn imbue people with interests; peopleact on those interests and, mostly without knowing it, they shape theconventions of social life which then give rise to morals. The process,however, is grounded on the technology of production at the beginning of thechain. And as this changes (through technological innovations) it provides theimpetus for the destabilisation of the (temporarily) evolutionary stableconventions at the other end of the chain. Page 232: Ifmorals are socially manufactured, then so is self-interest. Page 233: Perhaps our disagreement needs to be understood in terms of thelack of a shared history in relation to these debates—one of us embarkingfrom an Anglo-Saxon, the other from a (south) European, tradition. It was,after all, one of our important points in earlier chapters that game theoristsshould not expect a convergence of beliefs unless agents have a sharedhistory! Page 234: most of the population. This would seem to provide ammunition for the socialconstructivists, but of course it depends on them believing that collectiveaction agencies like the State will have sufficient information to distinguish thesuperior outcomes. Perhaps all that can be said on this matter is that, if youreally believe that evolutionary forces will do the best that is possible, then it isbeyond dispute that these forces have thrown up people who are predisposedto take collective action. Thus it might be argued that our evolutionarysuperiority as a species derives in part precisely from the fact that we are pro-active through collective action agencies rather than reactive as we would beunder a simple evolutionary scheme. Page 234: Turning to another dispute, that between social constructivism and spontaneousorder within liberal political theory, two clarifications have occurred. The first isthat there can be no presumption that a spontaneous order will deliveroutcomes which make everyone better off, or even outcomes which favour Page 234: Thesetheoretical moves will threaten to dissolve the distinction between action andstructure which lies at the heart of the game theoretical depiction of social lifebecause it will mean that the structure begins to supply reasons for action andnot just constraints upon action. On the optimistic side, this might be seen asjust another example of how discussions around game theory help to dissolvesome of the binary oppositions which have plagued some debates in socialscience—just as it helped dissolve the opposition between gender and classearlier in this chapter. However, our concern here is not to point to requiredchanges in ontology of a particular sort. The point is that some change isnecessary, and that it is likely to threaten the basic approach of game theory tosocial life. Page 234: Secondly, on the difficult cases where equilibrium selection involveschoices over whose interests are to be favoured (i.e. it is not a matter ofselecting the equilibrium which is better for everyone), then it is notobvious that a collective action agency like the State is any better placed tomake this decision than a process of spontaneous order. This may come asa surprise, since we have spent most of our time here focusing on theindeterminacy of evolutionary games when agents are only weaklyinstrumentally rational. Page 235: In other words the very debate within liberal political theory over socialconstructivism versus spontaneous order is itself unable to come to aresolution precisely because its shared ontological foundations are inadequatefor the task of social explanation. In short, we conclude that not only willgame theory have to embrace some expanded form of individual agency, if itis to be capable of explaining many social interactions, but also that this isnecessary if it is to be useful to the liberal debate over the scope of theState. Page 237: sabotage Page 238: What it does mean is thatour interpretation of results must be cautious and that, ultimately,laboratory experiments may only be telling us how people behave inlaboratories. Page 241: becausethere are some players who are unconditionally cooperative or ‘altruistic’ in theway that they play this game and, secondly, because whether someone iscooperative or not seems to be determined by one’s background, rather thanby how clever (or rational) he or she is (see adjacent box on the curse ofeconomics). In this sense, the evidence seems to point to a falsification of theassumption of instrumentally rational action based on the pay-offs Page 242: divisions of an army are stationed on two hill-tops overlooking a valley inwhich an enemy division can be clearly seen. It is known that if both divisionsattack simultaneously they will capture the enemy with none, or very little, lossof life. However, there were no prior plans to launch such an attack, as it wasnot anticipated that the enemy would be spotted in that location. How will thetwo divisions coordinate their attack (we assume that they must maintain visualand radio silence)? Neither commanding officer will launch an attack unless heis sure that the other will attack at the same time. Thus a classic coordinationproblem emerges.Imagine now that a messenger can be sent but that it will take him about anhour to convey the message. However, it is also possible that he will be caughtby the enemy in the meantime. If everything goes smoothly and the messengergets safely from one hill-top to another, is this enough for a coordinated attackto be launched? Suppose the message sent by the first commanding officer tothe second read: ‘Let’s attack at dawn!’ Will the second officer attack at dawn?No, unless he is confident that the first commanding officer (who sent the Page 242: message) knows that the message has been received. So, the secondcommanding officer sends the messenger back to the first with the message:‘Message received. Dawn it is!’ Will the second officer attack now? Not untilhe knows that the messenger has delivered his message. Paradoxically, noamount of messages will do the trick since confirmation of receipt of the lastmessage will be necessary regardless of how many messages have been alreadyreceived. Page 242: We see that in a coordination game like the above, even a very highdegree of common knowledge of the plan to attack at dawn is not enough toguarantee coordination (see Box 8.3 for an example of how different degreesof common knowledge can be engendered in the laboratory). What is needed(at least in theory) is a consistent alignment of beliefs (CAB) about the plan.1And yet this does not exclude the possibility that the two commandingofficers will both attack at dawn with very high probability. How successfullythey coordinate will, however, depend on more than a high degree ofcommon knowledge. Indeed the latter may even be un-necessary providedthe time of the attack is carefully chosen. The classic early experiments byThomas Schelling on behaviour in coordination games have confirmed this— Page 246: Thus in experiments, Pareto superiority does not seem to be a generalcriterion which players use to select between Nash equilibria (see also Chapter7). In conclusion, so far it seems that the way people actually play these gamesis neither directly controlled by the strategic aspects of the game (i.e. thelocation of the best response marks (+) and (-) in the matrix) nor by the size ofthe return from coordinating on non-Nash outcomes such as (R3, C3): it is aso-far-unexplained mixture of the two factors that decides. Page 251: To phrase this conclusion slightly differently, but in a way which connectswith the results in the next section, bargaining is a ‘complex socialphenomenon’ where people take cues from aspects of their social life whichgame theory typically overlooks. Thus players seem to base their behaviouron aspects of the social interaction which game theory typically treats asextraneous; and when players share these extraneous reference points such Page 258: What we have here is an evolution ofsocial roles. Players with the R label develop a different attitude towardsreflective cooperation to those players with the C role in spite of the fact that theRs and the Cs are the same people. In other words, the signal which causes theobserved pattern of cooperation seems to be emitted by the label R or C. Thisreminds us of the discussion in Chapter 7 about the capacity of sex, race andother extraneous features to pin down a convention on which the structure ofdiscrimination is grounded. Page 258: Experimentation with game theory is good, clean fun. Can it be more thanthat? Can it offer a way out of the obtuse debates on CKR, CAB, NEMS,Nash backward induction, out-of-equilibrium behaviour, etc.? The answerdepends on how we interpret the results. And as interpretation leaves plentyof room for controversy, we should not expect the data from the laboratoryunequivocally to settle any disputes. Our suspicion is that experiments are togame theory what the latter is to liberal individualism: a brilliant means ofcodifying its problems and of creating a taxonomy of time-honoureddebates.There are, however, important benefits from experimenting. Watchingpeople play games reminds us of their inherent unpredictability, their sense offairness, their complex motivation—of all those things that we tend to forgetwhen we model humans as bundles of preferences moving around some pay- Page 258: radical breakwith the exclusive reliance of instrumental rationality is also necessary. Page 260: At root we suspect that the major problem is the one that the experimentsin the last chapter isolate: namely, that people appear to be more complexlymotivated than game theory’s instrumental model allows and that a part ofthat greater complexity comes from their social location.We do not regard this as a negative conclusion. Quite the contrary, it standsas a challenge to the type of methodological individualism which has had afree rein in the development of game theory. Page 260: Along the way to this conclusion, we hope also that you have had fun.Prisoners’ dilemmas and centipedes are great party tricks. They are easy todemonstrate and they are amenable to solutions which are paradoxical enoughto stimulate controversy and, with one leap of the liberal imagination, theaudience can be astounded by the thought that the fabric of society (even theexistence of the State) reduces to these seemingly trivial games—Fun andGames, as the title of Binmore’s (1992) text on game theory neatly puts it. Butthere is a serious side to all this. Game theory is, indeed, well placed toexamine the arguments in liberal political theory over the origin and the scopeof agencies for social choice like the State. In this context, the problems whichwe have identified with game theory resurface as timely warnings of thedifficulties any society is liable to face if it thinks of itself only in terms ofliberal individualism. Page 260: The ambitious claim that game theory will provide a unified foundation for allsocial science seems misplaced to