10.02.2014 Views

Price dynamics on a stock market with asymmetric information - CMUP

Price dynamics on a stock market with asymmetric information - CMUP

Price dynamics on a stock market with asymmetric information - CMUP

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

C<strong>on</strong>tents lists available at ScienceDirect<br />

Games and Ec<strong>on</strong>omic Behavior<br />

www.elsevier.com/locate/geb<br />

<str<strong>on</strong>g>Price</str<strong>on</strong>g> <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> <strong>on</strong> a <strong>stock</strong> <strong>market</strong> <strong>with</strong> <strong>asymmetric</strong> informati<strong>on</strong><br />

Bernard De Meyer 1<br />

PSE-CES, Université Paris 1, 112 Bd de l’Hôpital, 75013 Paris, France<br />

article info abstract<br />

Article history:<br />

Received 19 February 2007<br />

Available <strong>on</strong>line 6 February 2010<br />

JEL classificati<strong>on</strong>:<br />

G14<br />

C72<br />

C73<br />

D44<br />

Keywords:<br />

Repeated games<br />

Incomplete informati<strong>on</strong><br />

<str<strong>on</strong>g>Price</str<strong>on</strong>g> <str<strong>on</strong>g>dynamics</str<strong>on</strong>g><br />

Martingales of maximal variati<strong>on</strong><br />

When two <strong>asymmetric</strong>ally informed risk-neutral agents repeatedly exchange a risky<br />

asset for numéraire, they are essentially playing an n-times repeated zero-sum game of<br />

incomplete informati<strong>on</strong>. In this setting, the price L q at period q can be defined as the<br />

expected liquidati<strong>on</strong> value of the risky asset given players’ past moves. This paper indicates<br />

that the asymptotics of this price process at equilibrium, as n goes to ∞, iscompletely<br />

independent of the “natural” trading mechanism used at each round: it c<strong>on</strong>verges, as n<br />

increases, to a C<strong>on</strong>tinuous Martingale of Maximal Variati<strong>on</strong>. This martingale class thus<br />

provides natural <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> that could be used in financial ec<strong>on</strong>ometrics. It c<strong>on</strong>tains in<br />

particular Black and Scholes’ <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>. We also prove here a mathematical theorem <strong>on</strong><br />

the asymptotics of martingales of maximal M-variati<strong>on</strong>, extending Mertens and Zamir’s<br />

paper <strong>on</strong> the maximal L 1 -variati<strong>on</strong> of a bounded martingale.<br />

© 2010 Elsevier Inc. All rights reserved.<br />

1. Introducti<strong>on</strong><br />

One fundamental problem in financial analysis is to accurately identify the <strong>stock</strong> price <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>: different <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> for<br />

the underlying asset will lead to different pricing formulae for derivatives.<br />

Financial ec<strong>on</strong>ometrics does not completely solve this problem: Statistical methods can calibrate the parameters of a<br />

model, finding in a general class of possible <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>, the <strong>on</strong>e that best fits the historical data. But still, assumpti<strong>on</strong>s have<br />

to be made regarding the class of possible <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>. Most of the classes used in practice (Bachelier’s <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>, Black and<br />

Scholes <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>, diffusi<strong>on</strong> models, stochastic volatility models, GARCH-models, etc.) are chosen by a kind of rule of thumb,<br />

<strong>with</strong> no real ec<strong>on</strong>omic justificati<strong>on</strong>. The randomness of the prices is often c<strong>on</strong>ceived as completely exogenous. The first<br />

sentences in Bachelier’s (1900) thesis illustrate quite well this kind of explanati<strong>on</strong>: “The influences that determine the price<br />

variati<strong>on</strong>s <strong>on</strong> the <strong>stock</strong> <strong>market</strong> are uncountable. Past, present or even future expected events, having often nothing to do<br />

<strong>with</strong> the <strong>stock</strong> <strong>market</strong>, have repercussi<strong>on</strong> <strong>on</strong> the prices.”<br />

In this paper however, we suggest that part of the randomness in the <strong>stock</strong> price <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> is endogenous: it is introduced<br />

by the agents in order to maximize their profit. This idea was already present in De Meyer and Moussa-Saley (2003),<br />

where the Brownian term in the price <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> was explained endogenously. Instituti<strong>on</strong>al investors clearly have better access<br />

to informati<strong>on</strong> <strong>on</strong> the <strong>market</strong> than the private <strong>on</strong>es: they are better skilled to analyze the flow of informati<strong>on</strong> and in<br />

some cases they are even part of the board of directors of the firms of which they are trading the shares. So, instituti<strong>on</strong>al<br />

investors are better informed and this informati<strong>on</strong>al advantage is known publicly. As a c<strong>on</strong>sequence, each of their moves<br />

<strong>on</strong> the <strong>market</strong>s is analyzed by the other agents to extract its informati<strong>on</strong>al c<strong>on</strong>tent. If informed agents act naively, making<br />

E-mail address: demeyer@univ-paris1.fr.<br />

1 The main ideas of this paper were developed during my stay at the Cowles Foundati<strong>on</strong> for research in Ec<strong>on</strong>omics at Yale University. I am very thankfull<br />

to all the members of the Cowles foundati<strong>on</strong> for their hospitality. I would like to thank John Geanakoplos and Pradeep Dubey for fruitful discussi<strong>on</strong>s.<br />

0899-8256/$ – see fr<strong>on</strong>t matter © 2010 Elsevier Inc. All rights reserved.<br />

doi:10.1016/j.geb.2010.01.011


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 43<br />

moves that depend deterministically <strong>on</strong> their informati<strong>on</strong>, they will completely reveal this informati<strong>on</strong> to the other agents,<br />

and doing so, they will lose their strategic advantage for the future. The <strong>on</strong>ly way to benefit from the informati<strong>on</strong> <strong>with</strong>out<br />

revealing it too fast is to introduce noise <strong>on</strong> their moves: this is tantamount to selecting random moves via lotteries that<br />

depend <strong>on</strong> their informati<strong>on</strong>. The main idea in De Meyer and Moussa-Saley (2003) is that the noise introduced by the<br />

informed agents in the day-to-day transacti<strong>on</strong>s will generate a Brownian moti<strong>on</strong>.<br />

The central result of this paper is that this kind of argument leads to a particular class of price <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>, hereafter<br />

referred to as C<strong>on</strong>tinuous Martingales of Maximal Variati<strong>on</strong> (CMMV), which is quite robust: CMMV will appear in any<br />

repeated exchange between two risk neutral <strong>asymmetric</strong>ally-informed players, independently of the “natural” trading mechanism<br />

used in the exchanges.<br />

The structure of the paper is as follows: in the next secti<strong>on</strong>, we define CMMV precisely. As suggested by our main result,<br />

this class of <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> is a natural candidate that could be used in financial ec<strong>on</strong>ometrics as the class of possible <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>.<br />

As will be seen, this class of <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> is a subclass of local volatility models that c<strong>on</strong>tains as particular cases Bachelier’s<br />

<str<strong>on</strong>g>dynamics</str<strong>on</strong>g> as well as Black and Scholes’ <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>.<br />

In Secti<strong>on</strong> 3, we introduce the game Γ n to model the repeated exchanges between two risk neutral <strong>asymmetric</strong>ally<br />

informed players in the most general way: Player 1 initially receives a private message c<strong>on</strong>cerning the liquidati<strong>on</strong> value L<br />

of the risky asset traded. During n c<strong>on</strong>secutive rounds, the players exchange the risky asset against a numéraire, using a<br />

general trading mechanism 〈I, J, T 〉 which is simply a game <strong>with</strong> respective acti<strong>on</strong> spaces I and J for players 1 and 2, and<br />

whose outcome T (i, j) is a transfer vector representing the quantities of risky asset and numéraire exchanged when acti<strong>on</strong>s<br />

are (i, j). At each round, acti<strong>on</strong>s are chosen simultaneously and are then publicly announced. The players aim to maximize<br />

the liquidati<strong>on</strong> value of their final portfolio.<br />

The game Γ n is equivalent to a zero sum repeated game <strong>with</strong> <strong>on</strong>e sided informati<strong>on</strong> à la Aumann–Maschler. In Secti<strong>on</strong> 4,<br />

we define the c<strong>on</strong>cepts of strategy, value and optimal strategy for Γ n . Since players are risk neutral, the natural noti<strong>on</strong> of<br />

price L q of the risky asset at period q is defined as the c<strong>on</strong>diti<strong>on</strong>al expected liquidati<strong>on</strong> value given player 2’s previous<br />

observati<strong>on</strong>s.<br />

The trading mechanism introduced above should satisfy some properties in order to represent real exchanges <strong>on</strong> the<br />

<strong>stock</strong> <strong>market</strong>. In Secti<strong>on</strong> 5, we introduce 5 axioms that must be satisfied by a “natural” trading mechanism. Let us describe<br />

them very briefly here:<br />

(H1) The game Γ n has a value V n whatever the distributi<strong>on</strong> of the liquidati<strong>on</strong> value is.<br />

(H2) is a c<strong>on</strong>tinuity assumpti<strong>on</strong> of V 1 as a functi<strong>on</strong> of the law of the liquidati<strong>on</strong> value.<br />

(H3) stipulates that the mechanism should be invariant <strong>with</strong> respect to the scale of numéraire: If two players use this<br />

mechanism to exchange a risky asset R against the dollar or against the cent, the same transacti<strong>on</strong>s in value will be<br />

observed in both cases. More specifically, the quantity of risky asset exchanged will be the same, but the counterpart<br />

in cents will be the counterpart in dollars multiplied by 100. (H3) is thus a 1-homogeneity property of the value.<br />

(H4) is an invariance axiom <strong>with</strong> respect to the riskless part of the risky asset: If two players use the trading mechanism<br />

to exchange <strong>with</strong> the dollar as numéraire an asset R ′ c<strong>on</strong>sisting of <strong>on</strong>e share of asset R and a $100 bill, then the<br />

transacti<strong>on</strong> observed will be the same in value as if they were exchanging R for the dollar. In other words, the<br />

quantity x of R- and R ′ -shares exchanged will be the same in both cases, but the x bills of $100 exchanged <strong>with</strong>in the<br />

R ′ -shares will be paid back in dollars, that is, if y and y ′ denotes the counterpart in numéraire when exchanging R and<br />

R ′ respectively, then y ′ = y + 100x. The value of the game must thus remain unchanged if <strong>on</strong>e shifts the liquidati<strong>on</strong><br />

value by a c<strong>on</strong>stant amount.<br />

(H5) There exists a situati<strong>on</strong> in which player 1 can take a strictly positive profit from his private informati<strong>on</strong>: he is strictly<br />

better off <strong>with</strong> his message than <strong>with</strong>out. This axiom is <strong>on</strong> the <strong>on</strong>e hand completely natural to model the <strong>stock</strong> <strong>market</strong>:<br />

it seems indeed comm<strong>on</strong>sense that private informati<strong>on</strong> has a strictly positive value <strong>on</strong> the <strong>market</strong>. Otherwise, there<br />

would clearly be no need for insider trading regulati<strong>on</strong>, since no <strong>on</strong>e would have incentive to make such trades.<br />

On the other hand, however, this axiom is in a way unnatural. This game is zero sum and has a positive value. So<br />

why should the uninformed player participate in a game where he is loosing m<strong>on</strong>ey? This is a particular case of Milgrom–<br />

Stokey’s No Trade Theorem. Some agents <strong>on</strong> the <strong>market</strong> are in fact forced to trade: for instance, a <strong>market</strong> maker facing<br />

a more informed trader. Since the bid and the ask posted by a <strong>market</strong> maker is a commitment to buy or sell at these<br />

prices any quantity of shares up to a prefixed limit, the <strong>on</strong>ly way for the <strong>market</strong> maker to avoid trading would be to post<br />

a very large bid-ask spread. Most <strong>market</strong> regulati<strong>on</strong>s however impose explicit limit <strong>on</strong> <strong>market</strong> makers’ bid-ask spread, thus<br />

steering past the No Trade paradox.<br />

At the end of Secti<strong>on</strong> 5, we state the main result of the paper which is Theorem 1. It indicates that if the trading mechanism<br />

is natural in the above sense, if the price process (L q ) q=0,...,n at equilibrium in Γ n is represented by the c<strong>on</strong>tinuous<br />

time process (Π n t ) t∈[0,1], <strong>with</strong>Π n t := L q <strong>on</strong> the time interval [q/n,(q + 1)/n[, thenΠ n c<strong>on</strong>verges in law to a particular<br />

CMMV Π μ depending just <strong>on</strong> the law μ of the liquidati<strong>on</strong> value of R. The limit is thus completely independent of the<br />

natural trading mechanism c<strong>on</strong>sidered, showing in this way the robustness of the CMMV class of <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>. This paper<br />

differs from the existing literature <strong>on</strong> trading <strong>with</strong> <strong>asymmetric</strong>al informati<strong>on</strong> (see e.g. Kyle, 1985) by the fact that the price<br />

randomness is essentially c<strong>on</strong>sidered as endogenous. In Kyle’s paper however, to get rid of the above menti<strong>on</strong>ed No Trade


44 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

paradox, noise traders have to be introduced, and the resulting <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> will thus crucially depend <strong>on</strong> the hypotheses made<br />

<strong>on</strong> this exogenous source of randomness.<br />

We will provide explicit examples of trading mechanism satisfying (H1) to (H5) is Secti<strong>on</strong> 6, and we will compare our<br />

results <strong>with</strong> De Meyer and Moussa-Saley (2003). In that paper a particular trading mechanism is analyzed for which optimal<br />

strategies can be explicitly computed and the c<strong>on</strong>vergence to the CMMV Π μ can be proved directly. The result of this paper<br />

however is much more general and applies to games <strong>with</strong> abstract trading mechanisms. In the absence of closed-form<br />

formulae, a more subtle and abstract line of analysis has to be followed.<br />

The proof of Theorem 1 is presented in Secti<strong>on</strong> 7. By linking more or less his moves to his initial message, player 1 can<br />

chose the rate of revelati<strong>on</strong> of his private informati<strong>on</strong> and in this way, he can c<strong>on</strong>trol the price process (L q ) q=1,...,n , which<br />

is precisely the c<strong>on</strong>diti<strong>on</strong>al expectati<strong>on</strong> of L given the past moves. We then express the maximal amount player 1 can<br />

guarantee in Γ n <strong>with</strong> a given revelati<strong>on</strong> martingale (L q ) q=1,...,n . We prove in particular that this amount is the V 1 -variati<strong>on</strong><br />

of the martingale (L q ) q=1,...,n : for a functi<strong>on</strong> M, suchasV 1 , that maps probability measures (the laws of the liquidati<strong>on</strong><br />

value for V 1 ) toR, the M-variati<strong>on</strong> of the martingale (L q ) q=1,...,n is defined as E[ ∑ n−1<br />

q=0 M([L q+1 − L q |(L s ) sq ])], where<br />

[L q+1 − L q |(L s ) sq ] denotes the c<strong>on</strong>diti<strong>on</strong>al law of the increment L q+1 − L q given the past at time q. The informed player is<br />

thus facing a martingale optimizati<strong>on</strong> problem: the price martingale (L q ) at equilibrium must maximize the V 1 -variati<strong>on</strong>.<br />

Theorem 1 then follows at <strong>on</strong>ce from Theorem 5 which is a mathematical result proved in the sec<strong>on</strong>d part of the paper.<br />

It is an interesting result that generalizes both Mertens and Zamir (1977) and De Meyer (1998) <strong>on</strong> the maximal L 1 - and<br />

L p -variati<strong>on</strong> of a bounded martingale: It states that the c<strong>on</strong>tinuous representati<strong>on</strong> (Πt n) t∈[0,1] of a martingale (L n q ) q=0,...,n<br />

(i.e. Πt<br />

n = L n q if t ∈[q/n,(q + 1)/n[) <strong>with</strong> final distributi<strong>on</strong> μ that maximizes the M-variati<strong>on</strong> c<strong>on</strong>verges in law, as n →∞,<br />

to the CMMV Π μ , provided M satisfies a homogeneity and a c<strong>on</strong>tinuity property. This result justifies the terminology<br />

C<strong>on</strong>tinuous Martingale of Maximal Variati<strong>on</strong> adopted in this paper. Since the limit Π μ is independent of M, which is a<br />

completely general functi<strong>on</strong> that could even have no relati<strong>on</strong> <strong>with</strong> a game, this result underlines even more the robustness<br />

of the CMMV class. The proof of this result is based <strong>on</strong> three ingredients: duality, the central limit theorem and Skorokhod’s<br />

embedding techniques. We refer the reader to Secti<strong>on</strong> 8 for more details.<br />

It is generally fairly difficult to prove that <strong>on</strong>e particular mechanism satisfies to (H1): Games of any length must have a<br />

value. In Appendix A, we prove that it is often sufficient to prove that the <strong>on</strong>e shot game has a value.<br />

2. C<strong>on</strong>tinuous martingales of maximal variati<strong>on</strong><br />

Let be the set of probability distributi<strong>on</strong>s <strong>on</strong> (R, B R ), where B R is the Borel tribe <strong>on</strong> R. In the following, the<br />

probability distributi<strong>on</strong> of a random variable X will be denoted [X]. Ifμ ∈ , we will use both notati<strong>on</strong>s X ∼ μ or<br />

[X] =μ to indicate that the random variable X is μ-distributed. We also write ‖μ‖ L p for ‖X‖ L p where X ∼ μ, and we<br />

set p := {μ ∈ : ‖μ‖ L p < ∞}.<br />

A c<strong>on</strong>tinuous martingale of maximal variati<strong>on</strong> is defined as a martingale Π <strong>on</strong> time interval [0, 1] that can be written as: ∀t: Π t =<br />

f (B t , t) where B is a standard Brownian moti<strong>on</strong> and f : R ×[0, 1]→R is increasing in the first variable.<br />

Notice that we can recover the whole functi<strong>on</strong> f (x, t) from its terminal values g(x) := f (x, 1): Indeed, using the Markov<br />

property of the Brownian moti<strong>on</strong>, we get:<br />

f (B t , t) = Π t = [ ] [ ] [ (<br />

E Π 1 |(B s ) st = E g(B1 )|B t = E g Bt + (B 1 − B t ) ) ]<br />

|B t .<br />

Since, B 1 − B t ∼ N (0,(1 − t)) is independent of B t ,weget<strong>with</strong>h t denoting the density functi<strong>on</strong> of N (0,(1 − t)):<br />

f (x, t) =<br />

∫ ∞<br />

−∞<br />

g(x − y)h t (y) dy = g ∗ h t (x),<br />

∗ representing the c<strong>on</strong>voluti<strong>on</strong> product. Due to the smoothing property of the normal kernel, the functi<strong>on</strong> f is thus C ∞ <strong>on</strong><br />

R ×[0, 1[. We may then also apply Itô’s formula to the process Π t and we get:<br />

(<br />

dΠ t = ∂ x f (B t , t) dB t + ∂ t f (B t , t) + 1 )<br />

2 ∂2 x,x f (B t, t) dt.<br />

For Π to be a martingale, the drift term must vanish and f must thus satisfy the heat equati<strong>on</strong> ∂ t f (x, t) + 1 2 ∂2 x,x f (x, t) = 0.<br />

One useful property of CMMV is that for a given μ ∈ 1 , there exists a unique CMMV denoted hereafter Π μ such that<br />

Π μ 1<br />

∼ μ. Indeed, as is well known, there exists a unique (up to a null set) increasing functi<strong>on</strong> f μ : R → R that f μ (Z) ∼ μ<br />

if Z ∼ N (0, 1). The functi<strong>on</strong> f μ can be expressed in terms of the cumulative distributi<strong>on</strong> functi<strong>on</strong>s F μ and F N of μ and<br />

N (0, 1) by the following formula: f μ (x) = Fμ −1(F<br />

N (x)), where Fμ −1(y)<br />

:= inf{s: F μ(s)>y}.<br />

Since Π μ 1<br />

= f (B 1, 1) ∼ μ and f (x, 1) is by hypothesis increasing in x, it must be the case that f (x, 1) = f μ (x), as<br />

B 1 ∼ N (0, 1). As menti<strong>on</strong>ed above, we then get f (x, t) = ∫ ∞<br />

−∞ f μ(x − y)h t (y) dy which is clearly increasing in x, since so is<br />

f μ :ifx > x ′ : f μ (x − y) f μ (x ′ − y). Multiplying both sides of this inequality by h t (y)>0, we get indeed after integrati<strong>on</strong><br />

f (x, t) f (x ′ , t).<br />

We c<strong>on</strong>clude this secti<strong>on</strong> by showing that the CMMV class is a class of local volatility models for the price process that<br />

c<strong>on</strong>tains as particular cases Bachelier’s Dynamics as well as Black and Scholes’ <strong>on</strong>e. Indeed, if μ is not a Dirac measure, f μ


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 45<br />

is not a c<strong>on</strong>stant functi<strong>on</strong>. Therefore, the strict inequality f μ (x − y)> f μ (x ′ − y) will hold in the previous argument for a set<br />

of y of positive Lebesgue measure. As a result, for t < 1, f (x, t)> f (x ′ , t), whenever x > x ′ . The strictly increasing functi<strong>on</strong><br />

x → f (x, t) has thus an inverse φ t and the relati<strong>on</strong> Π t = f (B t , t) yields B t = φ t (Π t ). Our above formula for dΠ t thus<br />

becomes: dΠ t = ∂ x f (B t , t) dB t = a(Π t , t) dB t , where a(y, t) := ∂ x f (φ t (y), t). WealsohavedΠ t = a(Π t , t) dB t <strong>with</strong> a(y, t) = 0<br />

when μ is a Dirac measure. A model of actualized price <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> under the risk-neutral probability measure satisfying this<br />

diffusi<strong>on</strong> equati<strong>on</strong> is referred to as a local volatility model in finance, a being the volatility functi<strong>on</strong> (e.g. Dupire, 1997).<br />

Since a must satisfy specific properties for the corresp<strong>on</strong>ding process to be a CMMV, the CMMV class is indeed a subclass<br />

of local volatility models.<br />

Under the risk-neutral probability, Bachelier’s <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> and Black and Scholes’ <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> for the actualized price process<br />

are given by the formulae dΠ t = σ dB t and dΠ t = σ Π t dB t , where σ > 0 is the volatility parameter. As is well known, in<br />

both case we get Π t = f (B t , t), <strong>with</strong> f defined respectively as f (x, t) = σ x + c and f (x, t) = ce σ x− σ 2<br />

2 t . These two functi<strong>on</strong>s<br />

are increasing in x.<br />

3. The game Γ n (μ)<br />

In the game Γ n analyzed in this paper, two players are repeatedly trading a risky asset R against a numéraire N.<br />

We first describe the informati<strong>on</strong> asymmetry: At the beginning of the game, player 1 (P1) receives a private message m<br />

c<strong>on</strong>cerning the risky asset. This message is randomly chosen by nature <strong>with</strong> a given probability distributi<strong>on</strong> ν. P2doesnot<br />

receive this message. He just knows that P1 has been informed and he also knows the probability distributi<strong>on</strong> ν.<br />

The message m will be publicly revealed at a future date T , say at the next shareholder meeting. The price L of R <strong>on</strong><br />

the <strong>market</strong> at that date is called the liquidati<strong>on</strong> value of R. It will depend <strong>on</strong> m and L is thus a functi<strong>on</strong> L(.) of m. The<br />

liquidati<strong>on</strong> value of N is independent of m and is fixed to be 1. We assume that both players know how to interpret the<br />

message m and they therefore know the functi<strong>on</strong> L(.).<br />

Let μ denote the probability distributi<strong>on</strong> of L(m) when m ∼ ν. We will assume in this paper that μ ∈ 1 .AsL(m) is the<br />

<strong>on</strong>ly relevant informati<strong>on</strong> in the message m, we may assume that P1 is <strong>on</strong>ly informed of L(m).<br />

So, the initial stage in Γ n (μ) is simply the following: Nature picks L at random <strong>with</strong> probability μ. P1isinformedofL,<br />

while P2 is not. Both players know μ.<br />

We next c<strong>on</strong>sider n rounds of exchange before the revelati<strong>on</strong> date T . To model these repeated exchanges in the most<br />

general way, we introduce the noti<strong>on</strong> of a trading mechanism. Such a mechanism is defined as a triple 〈(I, I), ( J, J ), T 〉,<br />

where I and J are the respective acti<strong>on</strong> sets of P1 and P2, endowed <strong>with</strong> σ -algebras I, J , and where T : I × J → R 2<br />

is the transfer functi<strong>on</strong>, assumed to be measurable from (I ⊗ J ) to the Borel σ -algebra B R 2 .Iftheplayersplay(i, j),<br />

T (i, j) = (A ij , B ij ) represents the transfer from P2 to P1: A ij and B ij are the respective numbers of R and N shares that P1<br />

receives from P2. (Typically <strong>on</strong>e is positive and the other negative.)<br />

So, in Γ n (μ), attradingperiodq (q = 1,...,n), the players select an acti<strong>on</strong> pair (i q , j q ) independently of each other,<br />

based <strong>on</strong> their prior observati<strong>on</strong>s and private informati<strong>on</strong>. Acti<strong>on</strong>s are then made public and the portfolios are then in-<br />

) represent P1 and P2’s portfolios at the end of round q, we have thus:<br />

cremented. If y q = (yq R, yN q ) and z q = (zq R, zN q<br />

y q = y q−1 + T (i q , j q ) and z q = z q−1 − T (i q , j q ). We assume that both players have sufficiently large portfolios, or have<br />

short-selling capacities, so as to h<strong>on</strong>or the trade T (i q , j q ).<br />

We will give examples of trading mechanisms in Secti<strong>on</strong> 6. The exchange at period q could result from a bargaining<br />

game or from an aucti<strong>on</strong> procedure, and acti<strong>on</strong>s in these setups are the strategies used by the players in these mechanisms.<br />

Another example of interest is a <strong>market</strong> game: Players’ acti<strong>on</strong>s i and j are demand functi<strong>on</strong>s for asset R inagivenclassE. In<br />

this case, I = J = E is a set of strictly decreasing functi<strong>on</strong>s i from R to R satisfying lim p→−∞ i(p)>0andlim p→∞ i(p)


46 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

4. Strategies, value, equilibria and price process in Γ n (μ)<br />

Let us first define strategies in Γ n (μ). AmixedstrategyforP2inΓ 1 (μ) is a probability distributi<strong>on</strong> τ <strong>on</strong> ( J, J ). However,<br />

since A ij and B ij are a priori unbounded, we have to restrict a little bit this definiti<strong>on</strong>. Let ( J) be the set of probability<br />

distributi<strong>on</strong>s τ <strong>on</strong> ( J, J ) such that,<br />

∫<br />

∫<br />

∀i ∈ I: |A ij | dτ ( j) −∞,<br />

which implies that g n (μ, σ , τ ) is well defined in R ∪{∞}.LetSn adm be the set of admissible strategies for P1.<br />

Astrategyσ ∈ Sn<br />

adm guarantees α to P1 if, ∀τ ∈ Tn adm : g n (μ, σ , τ ) α.


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 47<br />

The maximum amount P1 can guarantee in Γ n (μ) is<br />

V n (μ) :=<br />

sup inf<br />

σ ∈Sn<br />

adm τ ∈Tn<br />

adm<br />

g n (μ,σ,τ).<br />

Astrategyσ is optimal if it guarantees V n (μ).<br />

It is always true that V n (μ) V n (μ). When equality holds, the game Γ n (μ) is said to have a value V n (μ) := V n (μ) =<br />

V n (μ).<br />

If the game has a value, and if σ ∗ and τ ∗ are optimal strategies, then (σ ∗ , τ ∗ ) is a Nash equilibrium of the game.<br />

C<strong>on</strong>versely, if (σ ∗ , τ ∗ ) is a Nash equilibrium of the game, then the game has a value and σ ∗ and τ ∗ are optimal strategies.<br />

In the “<strong>market</strong> game” example of trading mechanism described in the previous secti<strong>on</strong>, where both players submit a<br />

demand functi<strong>on</strong>, a natural noti<strong>on</strong> of price at period q would be the <strong>market</strong> clearing price. However, <strong>with</strong> general trading<br />

mechanisms, acti<strong>on</strong>s by the players are not necessarily price related and we have to define what we mean by the price<br />

L q of the risky asset R at period q. One possible definiti<strong>on</strong> of L q could be based <strong>on</strong> the quotient − B iq, jq<br />

, that is the<br />

A iq, jq<br />

counterpart in numéraire paid per R-shares acquired. But this definiti<strong>on</strong> is not completely satisfying: <strong>on</strong> <strong>on</strong>e hand, it would<br />

lead to technical problems in case A iq , j q<br />

= 0. On the other hand, both B iq , j q<br />

and A iq , j q<br />

are random quantities. At stage 1<br />

for instance, the counterpart to the lottery yielding A i1 , j 1<br />

units of R is a lottery yielding −B i1 , j 1<br />

of N. So, should the price<br />

prevailing at the first exchange be defined as the quotient − E[B i 1 , j 1<br />

]<br />

E[A i1 , j 1<br />

] or as E[− B i 1 , j 1<br />

]? These two noti<strong>on</strong>s do not coincide in<br />

A i1 , j 1<br />

general.<br />

To avoid these difficulties, we prefer in this paper to define the price L q as the price at which the risk-neutral player<br />

P2 would agree to trade <strong>with</strong> another uninformed player: L q = E[L|i 1 ,...,i q , j 1 ,..., j q ]. This definiti<strong>on</strong> implies in particular<br />

that the price process is a martingale. It will also be c<strong>on</strong>venient in this paper to represent this discrete time price process<br />

(L q ) q=0,...,n by a c<strong>on</strong>tinuous time process (Πt n) t∈[0,1], where the time t represents the proporti<strong>on</strong> of already elapsed rounds:<br />

Πt n := L [[nt]] , where [[x]] denotes the largest integer below x.<br />

5. Natural exchange mechanism<br />

The noti<strong>on</strong> of a trading mechanism involved in the definiti<strong>on</strong> of Γ n (μ) is completely general and some mechanisms could<br />

be unrealistic to represent actual exchanges <strong>on</strong> the <strong>stock</strong> <strong>market</strong>, for instance a dictatorial mechanism where P1 selects the<br />

transfer vectors, regardless to P2’s acti<strong>on</strong>.<br />

Rather than focusing <strong>on</strong> an explicit trading mechanism representing <strong>on</strong>e particular <strong>market</strong>, we isolate in this secti<strong>on</strong> five<br />

axioms (H1) to (H5) that should be met by any natural modelizati<strong>on</strong> of an exchange. These are expressed in terms of the<br />

value of the <strong>on</strong>e shot game. We also give some additi<strong>on</strong>al c<strong>on</strong>diti<strong>on</strong>s (H1 ′ )to(H4 ′ ) that imply the corresp<strong>on</strong>ding (H) axiom<br />

and that are easier to check.<br />

We are c<strong>on</strong>cerned in this paper <strong>with</strong> qualitative properties of the price process at equilibrium in Γ n (μ) when such<br />

equilibria exist. Even if equilibria could fail to exist in Γ n (ν), forν ≠ μ, we however have to assume that the value exists.<br />

This is the c<strong>on</strong>tent of the two versi<strong>on</strong>s of axiom (H1) for k = 2or∞:<br />

(H1- k )Existenceofthevalue:The game Γ n (μ) has a value V n (μ) for all distributi<strong>on</strong> μ ∈ k and all n ∈ N.<br />

As we will see, axioms (H3) and (H4) hereafter will imply in particular that acti<strong>on</strong> spaces I and J c<strong>on</strong>tain infinitely many<br />

acti<strong>on</strong>s. Even when I and J are subsets of R m , the transfer functi<strong>on</strong> T will typically have some disc<strong>on</strong>tinuities in order to<br />

satisfy (H5). So proving that a particular mechanism satisfies (H1) is in general fairly difficult. The problem is essentially to<br />

prove that:<br />

(H1 ′ ): ∀μ ∈ ∞ , the game Γ 1 (μ) has a value.<br />

Indeed, we prove in Appendix A of this paper that (H1 ′ ) joint to our following c<strong>on</strong>tinuity assumpti<strong>on</strong> (H2) implies<br />

(H1- ∞ ). We also prove that (H1 ′ ) and (H2 ′ )imply(H1- 2 ).<br />

The sec<strong>on</strong>d axiom is a c<strong>on</strong>tinuity assumpti<strong>on</strong> of V 1 (μ) as a functi<strong>on</strong> of μ: if two assets R and R ′ have nearly the same<br />

the liquidati<strong>on</strong> values L and L ′ , i.e. there is a join distributi<strong>on</strong> of (L, L ′ ) where, for some p ∈[1, 2[, ‖L − L ′ ‖ p is small, then<br />

it is natural to expect that exchanging R or R ′ will lead to similar profits for P1: V 1 ([L]) and V 1 ([L ′ ]) should be close to<br />

each other. Axiom (H2) is in fact a Lipschitz c<strong>on</strong>tinuity assumpti<strong>on</strong>:<br />

(H2) C<strong>on</strong>tinuity: ∃p ∈[1, 2[, ∃A ∈ R: ∀L, L ′ ∈ L 2 :<br />

∣ V 1<br />

(<br />

[L]<br />

)<br />

− V 1<br />

([<br />

L<br />

′ ])∣ ∣ A · ∥∥ L − L<br />

′ ∥ ∥<br />

L p .<br />

The Wasserstein distance W p (μ, μ ′ ) of order p between μ and μ ′ ∈ 2 is defined as W p (μ, μ ′ ) := inf{‖X − X ′ ‖ L p :<br />

X ∼ μ, X ′ ∼ μ ′ }, and therefore (H2) is simply a Lipschitz c<strong>on</strong>tinuity assumpti<strong>on</strong> in terms of W p (μ, μ ′ ): ∃p ∈[1, 2[,<br />

∃A ∈ R: ∀μ, μ ′ ∈ 2 : |V 1 (μ) − V 1 (μ ′ )| A · W p (μ, μ ′ ).<br />

Notice that a mechanism T satisfying (H2 ′ )willclearlysatisfy(H2)<strong>with</strong>p = 1:<br />

(H2 ′ )Boundedexchanges:∀i, j: |A i, j | A.


48 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

Indeed, if nature initially selects jointly the liquidati<strong>on</strong> values (L, L ′ ) of two asset R and R ′ and informs P1 of its joint<br />

choice, the additi<strong>on</strong>al informati<strong>on</strong> c<strong>on</strong>tained in L ′ is useless to play in Γ 1 ([L]), and similarly, the additi<strong>on</strong>al informati<strong>on</strong><br />

c<strong>on</strong>tained in L is useless to play in Γ 1 ([L ′ ]). Viewing thus a strategy of P1 as a map from (L, L ′ ) to (I), we may c<strong>on</strong>sider<br />

that strategy spaces are identical in Γ 1 ([L]) and Γ 1 ([L ′ ]). If both players use the same strategies (σ , τ ) in these games, the<br />

payoffs are respectively E[A i, j L + B i, j ] and E[A i, j L ′ + B i, j ]. The difference of P1’s payoffs in both games E[A i, j (L − L ′ )] is<br />

then bounded by A‖L − L ′ ‖ 1 and, as announced, (H2) is thus satisfied.<br />

For the next two axioms, we c<strong>on</strong>sider a <strong>market</strong> as a system <strong>with</strong> which agents agree <strong>on</strong> trades of some asset in counterpart<br />

for some numéraire. The same system and the trading mechanism representing it could be used to trade different<br />

assets and numéraires.<br />

In axiom (H3), we c<strong>on</strong>sider the effect of a change of numéraire: we compare two situati<strong>on</strong>s where the same risky asset is<br />

exchanged for numéraire N 1 and N 2 respectively, <strong>with</strong> <strong>on</strong>e unit of N 1 being worth α > 0unitsofN 2 .IfL is the liquidati<strong>on</strong><br />

value of R in terms of N 1 , the liquidati<strong>on</strong> value of R in terms of N 2 will then be αL and we are thus comparing Γ 1 ([L])<br />

<strong>with</strong> Γ 1 ([αL]). Quite intuitively, we expect to observe same value trades in both situati<strong>on</strong>s and the value of both game will<br />

just differ by the scaling factor α. This is our hypothesis (H3):<br />

(H3): Invariance <strong>with</strong> respect <strong>with</strong> the numéraire scale:<br />

( ) ( )<br />

∀α > 0: ∀L ∈ L 2 : V 1 [αL] = αV 1 [L] .<br />

We could see this invariance property of the value V 1 as a c<strong>on</strong>sequence of an invariance property of the trading mechanism<br />

itself: when comparing Γ 1 ([L]) <strong>with</strong> Γ 1 ([αL]), we expect that the number of exchanged R-shares will be the same<br />

in both games, but the number of N 2 -shares given in counterpart in Γ 1 ([αL]) will just be the product of α and the N 1 -<br />

counterpart in Γ 1 ([L]).<br />

Clearly, the players will not use the same acti<strong>on</strong>s when trading <strong>with</strong> N 1 or N 2 , but to each acti<strong>on</strong> in the N 1 -trading<br />

game corresp<strong>on</strong>ds an acti<strong>on</strong> in the N 2 -game <strong>with</strong> similar effect. These corresp<strong>on</strong>dences between acti<strong>on</strong>s are represented by<br />

the translati<strong>on</strong> rules in the following axiom:<br />

(H3 ′ ): For all α > 0, there are <strong>on</strong>e-to-<strong>on</strong>e translati<strong>on</strong> rules ψ 1 : I → Iandψ 2 : J → J <strong>with</strong> the property that ∀i, j: A ψ1 (i),ψ 2 ( j) =<br />

A i, j and B ψ1 (i),ψ 2 ( j) = α · B i, j .<br />

If P1 plays ψ 1 (i) in Γ 1 ([αL]) whenever he would play i in Γ 1 ([L]), then his payoff in Γ 1 ([αL]) against an acti<strong>on</strong> j =<br />

ψ 2 ( j ′ ) of P2 will be:<br />

E[A ψ1 (i),ψ 2 ( j ′ )αL + B ψ1 (i),ψ 2 ( j ′ )]=αE[A i, j ′ L + B i, j ′].<br />

Therefore P1 can guarantee the same amount, up to a factor α in Γ 1 ([L]) and Γ 1 ([αL]). A similar argument holds for P2,<br />

and therefore our hypothesis (H3 ′ ) implies (H3).<br />

With the next axiom, we analyze the effect of shifting the liquidati<strong>on</strong> value by a c<strong>on</strong>stant quantity β: we compare the<br />

game Γ 1 ([L]), where a risky asset R <strong>with</strong> liquidati<strong>on</strong> value L is traded for numéraire N, <strong>with</strong>thegameΓ 1 ([L + β]) where<br />

a risky asset R ′ is traded for N, the asset R ′ c<strong>on</strong>sisting of <strong>on</strong>e unit of R and β units of N. We expect to observe similar<br />

trades in both games: the same number x of R shares will be exchanged in both cases, but the x.β units of N included in<br />

the x units of R ′ in the sec<strong>on</strong>d game are immediately paid back in N. In both games, the players will not use the same<br />

acti<strong>on</strong>s, however, their acti<strong>on</strong>s in the first game can be translated into acti<strong>on</strong>s in the sec<strong>on</strong>d <strong>on</strong>e. This leads to the following<br />

c<strong>on</strong>diti<strong>on</strong>, in terms of the trading mechanism:<br />

(H4 ′ ): For all β ∈ R, there exist <strong>on</strong>e-to-<strong>on</strong>e translati<strong>on</strong> rules φ 1 : I → Iandφ 2 : J → J that map Pi’s acti<strong>on</strong>s in Γ 1 ([L]) to Pi’s<br />

acti<strong>on</strong>s in Γ 1 ([L + β]) <strong>with</strong> the property that ∀i, j: A φ1 (i),φ 2 ( j) = A i, j and B φ1 (i),φ 2 ( j) = B i, j − β · A i, j .<br />

As we will see, (H3 ′ ) and (H4 ′ ) are quite natural when “price related” mechanisms are c<strong>on</strong>cerned, that is mechanisms<br />

where acti<strong>on</strong>s are prices or demand curves as the two first examples presented in the next secti<strong>on</strong>. For these mechanisms,<br />

the translati<strong>on</strong> rules ψ i and φ i appearing in (H3 ′ ) and (H4 ′ ) are respectively a rescaling by a factor α or a shift by a c<strong>on</strong>stant<br />

β of the corresp<strong>on</strong>ding price or demand curve.<br />

Since the trades in Γ 1 ([L]) and Γ 1 ([L + β]) will have the same value whenever the players use their translated strategies,<br />

it follows from (H4 ′ ) that:<br />

(H4): Invariance <strong>with</strong> respect to the risk-less part of the risky asset:<br />

( ) ( )<br />

∀β ∈ R: V 1 [L + β] = V 1 [L] .<br />

Using successively (H4), (H3) and (H2) we infer that<br />

∀β ∈ R:<br />

V 1<br />

(<br />

[β]<br />

)<br />

= V 1<br />

(<br />

[0 + β]<br />

)<br />

= V 1<br />

(<br />

[0]<br />

)<br />

= lim<br />

α→0<br />

V 1<br />

(<br />

[α]<br />

)<br />

= lim<br />

α→0<br />

αV 1 (1) = 0.<br />

In other words, no benefit can be made of exchanging a riskless asset for a numéraire. Notice that this last property would<br />

be automatically satisfied by a symmetrical trading mechanism (I = J and T (i, j) =−T ( j, i)). Indeed, in this case, the <strong>on</strong>ly<br />

asymmetry in Γ 1 (μ) is the initial message sent to P1. When the liquidati<strong>on</strong> value of R is a c<strong>on</strong>stant β, thegameisthusa<br />

completely symmetric zero sum game (<strong>with</strong> anti-symmetric payoff matrix) and its value V 1 ([β]) must then be 0.


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 49<br />

In the game Γ 1 ([L]), P1 could decide to ignore his private informati<strong>on</strong> L. Since the players are risk-neutral, the risky<br />

asset R could then be replaced by a risk-less asset <strong>with</strong> c<strong>on</strong>stant liquidati<strong>on</strong> value β := E[L]: <strong>with</strong>out his informati<strong>on</strong>, P1<br />

could then guarantee V 1 ([β]) = 0. So, using his informati<strong>on</strong> in Γ 1 ([L]), P1 can guarantee a positive amount: V 1 ([L]) 0.<br />

Our next axiom is that there exists a situati<strong>on</strong> in which P1 can make a strict profit out of his informati<strong>on</strong>:<br />

(H5): Positive value of informati<strong>on</strong>: ∃L ∈ L 2 : V 1 ([L])>0.<br />

As already explained in the introducti<strong>on</strong>, this last axiom is paradoxical: <strong>on</strong> <strong>on</strong>e hand, the fact that private informati<strong>on</strong><br />

has a strictly positive value <strong>on</strong> the <strong>market</strong> is so obvious that a model <strong>with</strong>out (H5) would be completely unrealistic. Rating<br />

firms can sell informati<strong>on</strong> at a positive price because it has a positive value. Insider trading regulati<strong>on</strong> would not exist if<br />

private informati<strong>on</strong> had no value. On the other hand, (H5) is <strong>on</strong>ly possible if the uninformed player is forced to play the<br />

game: he would be better off if he could avoid trading. As previously menti<strong>on</strong>ed, we argue that some agents, for instance<br />

<strong>market</strong> makers, are in fact forced to trade. Notice also that most <strong>market</strong> models <strong>with</strong> <strong>asymmetric</strong> informati<strong>on</strong> avoid this no<br />

trade paradox by introducing noise traders, that is traders that are forced to trade for exogenous liquidity reas<strong>on</strong>s.<br />

A trading mechanism will be referred to as k-natural if it satisfies the previous five axioms (H1- k ), (H2),...,(H5). We<br />

are now ready to state the main result of the paper:<br />

Theorem 1. For k ∈{2, ∞} and μ in k , c<strong>on</strong>sider a sequence {(σ n , τ n )} n∈N of equilibria in the repeated exchange games Γ n (μ)<br />

indexed by the length n of the game. Let L n be the price process at equilibrium in Γ n (μ):<br />

L n q := E π(μ,σ n ,τ n )[L|i 1 ,...,i q , j 1 ,..., j q ].<br />

:= L n [[nt]] c<strong>on</strong>verges in finite-<br />

Then, if the trading mechanism is k-natural, the c<strong>on</strong>tinuous time representati<strong>on</strong> Π n of L n defined as Πt<br />

n<br />

dimensi<strong>on</strong>al distributi<strong>on</strong> to the c<strong>on</strong>tinuous martingale of maximal variati<strong>on</strong> Π μ .<br />

The asymptotics of the price process is thus completely independent of the natural trading mechanism. Before proving<br />

this theorem in Secti<strong>on</strong> 7, we provide in the next secti<strong>on</strong> some examples of natural trading mechanism.<br />

6. Examples of natural trading mechanism<br />

6.1. Market maker game<br />

As a first example, let us c<strong>on</strong>sider the game between two <strong>asymmetric</strong>ally informed <strong>market</strong> makers. At each period, they<br />

post simultaneously a price i and j ∈ R which are a commitment to sell or buy <strong>on</strong>e unit of the risky asset at this price. If<br />

i ≠ j, an arbitrageur will see a possibility of arbitrage: he will buy <strong>on</strong>e share of R at the lowest price and sell it immediately<br />

at the highest <strong>on</strong>e. This game, referred to hereafter as the <strong>market</strong> maker game <strong>with</strong> arbitrageur is not an exchange between<br />

two individuals and does not bel<strong>on</strong>g to the class of zero-sum games analyzed in this paper. It can however be approximated<br />

by the following trading mechanism: when i > j, <strong>on</strong>e share of R goes from P2 to P1 as in the <strong>market</strong> maker <strong>with</strong> arbitrageur<br />

game, but we now assume that both players are exchanging R at a comm<strong>on</strong> price f (i, j) in numéraire, where f is a functi<strong>on</strong><br />

satisfying min(i, j) f (i, j) max(i, j), and ∀α > 0, ∀β: f (αi +β,α j +β) = α f (i, j)+β. Symmetric transfers happen when<br />

i < j, and there is no transfer if i = j. So, formally, the trading mechanism is represented by the transfer functi<strong>on</strong><br />

⎧<br />

⎨ (1, − f (i, j)) if i > j,<br />

T (i, j) := (0, 0) if i = j,<br />

⎩<br />

(−1, f (i, j)) if i < j.<br />

In De Meyer and Moussa-Saley (2003), the particular case f (i, j) = max(i, j) was c<strong>on</strong>sidered. Similar results would<br />

hold for other functi<strong>on</strong> f as f (i, j) = λ max(i, j) + (1 − λ) min(i, j), <strong>with</strong>λ ∈[0, 1]. In that paper, the game Γ n (μ) was<br />

analyzed for distributi<strong>on</strong> μ c<strong>on</strong>centrated <strong>on</strong> the two points 0, 1. The analysis was extended to general distributi<strong>on</strong>s μ ∈ 2<br />

in De Meyer and Moussa-Saley (2002). We prove in these papers that Γ n (μ) have a value and both players have optimal<br />

strategies. We further have a closed form formula for the value: V n (μ) = max E[LS n ], where the maximum is taken over<br />

the joint distributi<strong>on</strong>s of (L, S n ) satisfying L ∼ μ and S n = ∑ n<br />

q=1 U q is a sum of n independent random variables U q that<br />

are uniformly distributed <strong>on</strong> [−1, 1]. This is a M<strong>on</strong>ge–Kantorovich mass transportati<strong>on</strong> problem and its optimal soluti<strong>on</strong><br />

(L, S n ) is characterized by the fact that L can be expressed as the unique increasing functi<strong>on</strong> gμ n (S n) satisfying gμ n (S n) ∼ μ.<br />

It follows from this formula that (H1- 2 )–(H5) are satisfied by the mechanism. Based <strong>on</strong> these explicit formulae, we can<br />

also prove that the price process at equilibrium is given by L q = E[gμ n (S n)|U 1 ,...,U q ]. Using the central limit theorem,<br />

the asymptotics of this price process can then be derived, and we prove in De Meyer and Moussa-Saley (2003, 2002) that<br />

the above Theorem 1 holds for this precise trading mechanism. The proof of Theorem 1 presented in the present paper is<br />

however much more general as it applies to arbitrary natural trading mechanism where no closed form formulae are known<br />

for V n (μ).<br />

The particular mechanism dealt <strong>with</strong> in these papers is “price related” in the sense that the players’ acti<strong>on</strong>s are the<br />

proposed prices for the exchange. In this setting, a natural noti<strong>on</strong> of the price process for R could be the sequence of P1’s


50 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

posted prices. We prove in De Meyer and Moussa-Saley (2003) that the price i q posted at period q will be very close to L q ,<br />

and the posted price process will have the same asymptotic as (L q ) q=1,...,n .<br />

De Meyer and Marino (2005a, 2005b), Domansky (2007) analyze the same trading mechanism as in De Meyer and<br />

Moussa-Saley (2003), but <strong>market</strong> makers have to post prices <strong>with</strong>in a discrete grid. In this case, however, the price process<br />

fails to c<strong>on</strong>verge to a CMMV. Theorem 1 does not apply in this setting since neither (H3) nor (H4) are satisfied by these<br />

discretized mechanism: the grid should be both shift- and scale-invariant, which is impossible.<br />

6.2. Market game<br />

The sec<strong>on</strong>d class of examples of natural mechanisms we will present are the following <strong>market</strong> games: Players’ acti<strong>on</strong>s<br />

i and j are demand functi<strong>on</strong>s for asset R in terms of numéraire N. Inthiscase,I and J are sets of strictly decreasing<br />

functi<strong>on</strong>s i from R to R satisfying lim p→−∞ i(p)>0 and lim p→∞ i(p) 0 and all β ∈ R, ifi bel<strong>on</strong>gs to I then so does the functi<strong>on</strong><br />

i α,β : p → i α,β (p) := i(αp + β). In the same way, j ∈ J will imply j α,β ∈ J ,foralla > 0 and β ∈ R.<br />

Axioms (H3) and (H4) will then be satisfied by this mechanism: we just have to prove that (H3 ′ ) and (H4 ′ )hold.For<br />

α > 0, define the translati<strong>on</strong> maps ψ 1 (i) := i α −1 ,0 and ψ 2( j) := j α −1 ,0 .Sincep∗ (i α −1 ,0 , j α −1 ,0 ) = αp∗ (i, j), wegetindeed<br />

A ψ1 (i),ψ 2 ( j) = i α −1 ,0 (p∗ (i α −1 ,0 , j α −1 ,0 )) = i(p∗ (i, j)) = A i, j and B ψ1 (i),ψ 2 ( j) =−p ∗ (i α −1 ,0 , j α −1 ,0 )i α −1 ,0 (p∗ (i α −1 ,0 , j α −1 ,0 )) =<br />

αB i, j .(H3 ′ )isthusproved.<br />

To prove (H4 ′ ), define, for β ∈ R, the translati<strong>on</strong> maps φ 1 (i) := i 1,−β and φ 2 ( j) := j 1,−β .Thenp ∗ (i 1,−β , j 1,−β ) = p ∗ (i, j)+<br />

β and (H4 ′ ) follows:<br />

and<br />

A φ1 (i),φ 2 ( j) = i 1,−β<br />

(<br />

p ∗ (i 1,−β , j 1,−β ) ) = i ( p ∗ (i, j) ) = A i, j<br />

(<br />

B φ1 (i),φ 2 ( j) =−i 1,−β p ∗ (i 1,−β , j 1,−β ) ) p ∗ (i 1,−β , j 1,−β ( )<br />

=−i p ∗ (i, j) )( p ∗ (i, j) + β )<br />

= B i, j − A i, j β.<br />

We next discuss axiom (H5) in the <strong>market</strong> game model. This axiom precludes P2 from no trading and the zero demand<br />

functi<strong>on</strong> could therefore not bel<strong>on</strong>g to J nor be approximated by an element of J in order to (H5) to hold. In fact, this<br />

c<strong>on</strong>diti<strong>on</strong> will imply that any functi<strong>on</strong> j ∈ J is disc<strong>on</strong>tinuous at the point γ := sup{p: j(p)>0}. At this threshold price γ ,<br />

P2 passes from a strictly buying positi<strong>on</strong> (lim < p →γ<br />

j(p) >0) to a strictly selling <strong>on</strong>e (lim > p →γ<br />

j(p) 0 and i 0,q is the inelastic demand functi<strong>on</strong> at price q. Anacti<strong>on</strong>i of P1 can thus<br />

be identified <strong>with</strong> the corresp<strong>on</strong>ding pair (q, α). Let J be the set of functi<strong>on</strong>s j r for some r ∈ R, where j r (p) := 1ifp < r,<br />

j r (p) := −1 ifp > r and j r (p) := 0ifp = r. The <strong>market</strong> clearing price in this setup will be p ∗ (α, q, r) = q − α if q − α > r,<br />

p ∗ (α, q, r) = q + α if q + α < r, and there will be no clearing price if q − α r q + α. It is c<strong>on</strong>venient to represent a pair<br />

(α, q) by a pair (x, y), <strong>with</strong>x := q − α y =: q + α. With these notati<strong>on</strong>s, the trading mechanism is thus<br />

⎧<br />

⎨ (1, x) if r < x,<br />

T (x, y, r) = (−1, y) if r > y,<br />

⎩<br />

(0, 0) if x r y.<br />

This game could then also be viewed as an exchange between an informed <strong>market</strong> maker (P1) posting a bid x and an ask<br />

y and an investor who is forced to trade: he can just decide the price r at which he switches from a buying positi<strong>on</strong> to a<br />

selling <strong>on</strong>e.<br />

We now prove that this mechanism satisfies (H1 ′ ). Let μ ∈ ∞ and let K be a compact interval of positive length such<br />

that μ(K ) = 1. A pure strategy for P1 in Γ 1 (μ) is a pair of functi<strong>on</strong>s (X, Y ) that maps the liquidati<strong>on</strong> value L to the acti<strong>on</strong><br />

X(L), Y (L). IfP2playsapurestrategyr, thepayoffisE μ [g(L, X(L), Y (L), r)], where, <strong>with</strong> χ(x) := 1 {x


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 51<br />

P2 has an optimal strategy τ ∗ , as it results from Propositi<strong>on</strong> 1.17 in Mertens et al. (1994): K is compact, and due to our<br />

assumpti<strong>on</strong> ∀L: X(L) L Y (L), the coefficients of χ in (3) are positive and g(X, Y , r) is thus lower semic<strong>on</strong>tinuous in r.<br />

With this mechanism, there is at most <strong>on</strong>e share of R exchanged, and thus (H2 ′ ) holds. With Theorem 24, we thus<br />

c<strong>on</strong>clude that (H1- 2 ) holds and axiom (H2) is implied by (H2 ′ ). Since I and J are closed for shift and dilatati<strong>on</strong>, (H3)<br />

and (H4) will hold. Finally, to prove that (H5) holds, c<strong>on</strong>sider the measure μ that puts a weight 1/2 <strong>on</strong> both 0 and 1. The<br />

following strategy (X, Y ) guarantees 1/8 to P1 in Γ(μ)<br />

{<br />

3/4 if L = 1,<br />

X(L) :=<br />

0 if L = 0<br />

{ 1/4 if L = 0,<br />

and Y (L) :=<br />

1 if L = 1,<br />

and V 1 (μ) 1/8 > 0. The mechanism is thus 2-natural.<br />

6.3. Can<strong>on</strong>ical games<br />

We c<strong>on</strong>clude this secti<strong>on</strong> <strong>with</strong> a technical remark to show the class of games we are analyzing in this paper is far<br />

from empty: We provide a general way to generate natural trading mechanisms. If a functi<strong>on</strong> V : 2 → R is the value of a<br />

game <strong>with</strong> <strong>on</strong>e sided informati<strong>on</strong> where the state of nature space is R, it is well known that V must be both c<strong>on</strong>cave and<br />

Blackwell increasing. This last property means that if (Y 1 , Y 2 ) is a random vector such that E[Y 2 |Y 1 ]=Y 1 ,thenV ([Y 2 ]) <br />

V ([Y 1 ]). We argue here that any functi<strong>on</strong> V <strong>with</strong> these properties and satisfying further (H2) to (H5) can be implemented<br />

as the value of a game <strong>with</strong> a natural trading mechanism. V (μ) := (E μ [|L − E μ [L]| p ]) 1/p ,1 p < 2, is an example of such<br />

a functi<strong>on</strong>. A c<strong>on</strong>cave functi<strong>on</strong> V is the infimum of the linear functi<strong>on</strong>als by which it is dominated. In this setting, the<br />

c<strong>on</strong>tinuous linear functi<strong>on</strong>als of measures are maps of the form μ → E μ [φ(L)], where φ is a c<strong>on</strong>tinuous functi<strong>on</strong>. Therefore<br />

if Φ is the set of c<strong>on</strong>tinuous functi<strong>on</strong>s φ such that ∀μ: E μ [φ(L)] V (μ), wewillhave∀μ: V (μ) = inf φ∈Φ E μ [φ(L)]. The<br />

Blackwell m<strong>on</strong>ot<strong>on</strong>icity indicates that <strong>on</strong>ly c<strong>on</strong>vex functi<strong>on</strong>s φ in Φ are to be c<strong>on</strong>sidered: in other words, if Φ vex is the set<br />

of c<strong>on</strong>vex φ in Φ then ∀μ: V (μ) = inf φ∈Φ vex E μ [φ(L)]. A proof of this property can be found in De Meyer et al. (2009).<br />

By a density argument, the set Φ 1,vex of c<strong>on</strong>tinuously differentiable functi<strong>on</strong>s φ in Φ vex will have the same property:<br />

∀μ: V (μ) = inf<br />

φ∈Φ 1,vex E μ<br />

[<br />

φ(L)<br />

]<br />

. (4)<br />

C<strong>on</strong>sider then the following trading mechanism: I = R, J = Φ 1,vex and ∀i ∈ I,φ∈ Φ 1,vex , T (i,φ) is defined as: T (i,φ):=<br />

(φ ′ (i), φ(i) − iφ ′ (i)).<br />

With this trading mechanism, both players can guarantee V (μ) in Γ 1 (μ). IndeedifP2playsφ in Γ 1 (μ), forallpure<br />

strategy i(L), the payoff will be E μ [φ ′ (i(L))(L − i(L)) + φ(i(L))] which is less than E μ [φ(L)], as a c<strong>on</strong>sequence of the<br />

c<strong>on</strong>vexity of φ. Minimizing this in φ, P2 can the guarantee V (μ) in Γ 1 (μ), as it follows from (4).<br />

P1 can guarantee the same amount <strong>with</strong> the pure strategy i ∗ (L) := L. Indeed,ifP2playsφ the payoff is then E μ [φ(L)] <br />

V (μ).<br />

Therefore V (μ) is the value of Γ 1 (μ) and thus V 1 = V will satisfy (H1 ′ ) and thus (H1- ∞ ), as it follows from Theorem<br />

24. According to our assumpti<strong>on</strong> <strong>on</strong> V the mechanism will also satisfy (H2) to (H5): It thus is an ∞-natural mechanism.<br />

In the above game, the strategy space J could be reduced to any subset F of φ 1,vex such that ∀μ: V (μ) =<br />

inf φ∈F E μ [φ(L)]. As an applicati<strong>on</strong> of this remark, let f be a strictly positive c<strong>on</strong>tinuously differentiable c<strong>on</strong>vex functi<strong>on</strong><br />

such that f (x)/(1 +|x| p ) is bounded <strong>on</strong> R for some p ∈[1, 2[. IfF denotes then the set of φ of the form as φ(L) = α f ( L−β<br />

α )<br />

for some α > 0 and β ∈ R. Then the functi<strong>on</strong> V (μ) := inf α>0,β E μ [α f ( L−β<br />

α )] is the value of the game Γ 1(μ) <strong>with</strong> the following<br />

trading mechanism: In this case a functi<strong>on</strong> φ ∈ E can be identified <strong>with</strong> the corresp<strong>on</strong>ding pair (α,β), and thus<br />

J = (R + × R), I = R and T (i, α,β)= ( f ′ ( i−β i−β<br />

α ), α f ( α ) − if′ ( i−β<br />

α )). The resulting mechanism will be ∞-natural.<br />

7. P1’s martingale optimizati<strong>on</strong> problem<br />

In this secti<strong>on</strong> we show that the choice of a strategy for player 1 turns out to be a choice of the optimal rate of revelati<strong>on</strong><br />

of his private informati<strong>on</strong>. If, at period q, P1 uses a strategy that depends <strong>on</strong> his informati<strong>on</strong> L, P2willmakeaBayesian<br />

reevaluati<strong>on</strong> of his beliefs about L. Therefore, the price L q , which is the expectati<strong>on</strong> of L <strong>with</strong> respect to P2’s believe after<br />

stage q, will be different of L q−1 . The amount of informati<strong>on</strong> revealed can thus be represented by the price process. The<br />

optimal rate of revelati<strong>on</strong> for P1 will be that for which the price process L is optimal in the maximizati<strong>on</strong> problem (6) here<br />

below.<br />

If Y is a random variable <strong>on</strong> a probability space (Ω, A, P) and if H ⊂ A is a σ -algebra, the probability distributi<strong>on</strong> of<br />

Y will be denoted [Y ], and the c<strong>on</strong>diti<strong>on</strong>al distributi<strong>on</strong> of Y given H will be [Y |H]. WewillalsowriteΓ n [Y ] and V n [Y ]


52 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

instead of Γ n ([Y ]) and V n ([Y ]). InparticularV 1 [Y |H] is an H-measurable random variable. 2 Let W n (μ) be the set of pairs<br />

(F, X) where F := (F q ) q=0,...,n+1 is a filtrati<strong>on</strong> <strong>on</strong> a probability space, and X = (X q ) q=0,...,n+1 is an F -martingale whose<br />

n + 1-th value X n+1 is μ-distributed. For (F, X) ∈ W n (μ), wedefineV n (F, X) as:<br />

[ n−1<br />

]<br />

∑<br />

V n (F, X) := E V 1 [X q+1 |F q ] . (5)<br />

q=0<br />

Let us also define V n (μ) as<br />

V n (μ) := sup { V n (F, X): (F, X) ∈ W n (μ) } . (6)<br />

Lemma 2. For all μ ∈ 2 : V n (μ) V n (μ).<br />

Proof. Given (F, X) ∈ W n (μ) <strong>on</strong> a probability space (Ω, A, P), we have to prove that P1 can guarantee V n (F, X) in Γ n (μ).<br />

At the initial stage of Γ n (μ), nature selects a μ distributed random variable L and informs P1 of its choice. We can clearly<br />

assume that nature uses the probability space (Ω, A, P) as lottery and sets L = X n+1 ,sinceX n+1 ∼ μ. We can also assume<br />

that P1 observes the whole space (Ω, A, P). He can therefore adopt the following strategy in Γ n [X n+1 ]:atstageq<br />

he plays an ɛ-optimal strategy σ ∗ q in Γ 1[X q+1 |F q ]. The probability distributi<strong>on</strong> σ ∗ q used by P1 to select i q is thus measurable<br />

<strong>with</strong> respect to σ (F q , X q+1 ) ⊂ F q+1 and i q is chosen independently of X n+1 c<strong>on</strong>diti<strong>on</strong>ally to σ (F q , X q+1 ). Therefore<br />

E[X n+1 |σ (F q , X q+1 , i q )]=E[X n+1 |σ (F q , X q+1 )]=X q+1 ,sinceX is an F -martingale. The payoff at stage q is thus:<br />

E[A iq ,τ q<br />

X n+1 + B iq ,τ q<br />

]=E[A iq ,τ q<br />

X q+1 + B iq ,τ q<br />

]=E [ E[A iq ,τ q<br />

X q+1 + B iq ,τ q<br />

|F q ] ] .<br />

Since the move of P1 is ɛ-optimal in Γ 1 [X q+1 |F q ],hegetsatleastV 1 [X q+1 |F q ]−ɛ c<strong>on</strong>diti<strong>on</strong>ally to F q . The result follows<br />

then easily since ɛ > 0isarbitrary. ✷<br />

Lemma 3. For all μ ∈ 2 : V n (μ) V n (μ).<br />

Proof. We will prove that P1 will not be able to guarantee a higher payoff than V n (μ). Indeed,letσ be a strategy of P1 in<br />

Γ n (μ). Toreplytoσ , P2 may adopt the following strategy: since he knows σ 1 , he may compute the distributi<strong>on</strong> of L 1 :=<br />

E[L|i 1 ].Heplaysthenanɛ-optimal strategy τ 1 in Γ 1 [L 1 ].Atperiodq, he computes [L q |H q−1 ],<strong>with</strong>H q := σ (i 1 , j 1 ...i q , j q ),<br />

where L q := E[L|i q , H q−1 ], and plays an ɛ-optimal strategy τ q in Γ 1 [L q |H q−1 ]. Clearly, we also have L q = E[L|H q ],since<br />

c<strong>on</strong>diti<strong>on</strong>ally to H q−1 ,themove j q is independent of L. Therefore, <strong>with</strong> H n+1 := σ (L, H n ), L n+1 := L, H 0 := {∅,(I × J) n ×R},<br />

L 0 := E[L], L := (L q ) q=0,...,n+1 and H := (H q ) q=0,...,n+1 ,thepair(H, L) bel<strong>on</strong>gs to W n (μ).<br />

With that reply P1’s c<strong>on</strong>diti<strong>on</strong>al payoff at period q, givenH q−1 is the at most V 1 [L q |H q−1 ]+ɛ and the overall payoff in<br />

Γ n (μ) is then less than V n (H, L) + nɛ V n (μ) + nɛ. ✷<br />

Theorem 4. If the mechanism satisfies (H1- k ) (k ∈{2, ∞}), then for all μ ∈ k : V n (μ) = V n (μ). Furthermore,ifσ , τ are optimal<br />

strategies in Γ n (μ), ifL q := E π(μ,σ ,τ ) [L|H q ],whereH q := σ (i 1 , j 1 ...i q , j q ),andL := (L q ) q=0,...,n+1 ,then(H, L) solves the<br />

maximizati<strong>on</strong> problem (6).<br />

Proof. The first claim follows the two previous lemmas and the fact that the value exists as stated in (H1- k ).<br />

Let us next assume that the players are playing a pair (σ , τ ) of optimal strategies in Γ n (μ). LetX q denote the expectati<strong>on</strong>,<br />

c<strong>on</strong>diti<strong>on</strong>al to H q ,ofthesumofthen − q stage payoffs following stage q.<br />

Let us then first observe that, for all q, X q must be at least V n−q [L|H q ]. Otherwise P1 could deviate from stage q + 1<br />

<strong>on</strong> to an ɛ-optimal strategy in Γ n−q [L|H q ], obtaining thus a higher payoff against τ than <strong>with</strong> σ , which is impossible since<br />

(σ , τ ) is an equilibrium of the game.<br />

At period q, P2 may compute v q−1 := V 1 [L q |H q−1 ] and u q−1 := E[A iq , j q<br />

L q + B iq , j q<br />

|H q−1 ].Ontheevent{v q−1 < u q−1 },P2<br />

could then deviate at stage q <strong>with</strong> an ɛ-optimal strategy in Γ 1 [L q |H q−1 ], bringing the expected payoff of that stage to less<br />

than v q−1 + ɛ. Provided ɛ is small enough, this is strictly less than the payoff u q−1 P1 would obtain <strong>with</strong> τ .Thisdeviati<strong>on</strong><br />

will not change the behavior of P1 at period q, and the distributi<strong>on</strong> [L|H q ] remains thus unchanged.<br />

2<br />

The set of probability measures <strong>on</strong> R may be endowed <strong>with</strong> the weak*-topology: the weakest topology such that φ g : μ → E μ [g] is c<strong>on</strong>tinuous,<br />

for all c<strong>on</strong>tinuous bounded functi<strong>on</strong> g : R → R. IfY is a random variable <strong>on</strong> a probability space (Ω, A, P) and if H ⊂ A is a σ -algebra, the c<strong>on</strong>diti<strong>on</strong>al<br />

distributi<strong>on</strong> [Y |H] can then be seen as a measurable map from (Ω, H) to (, B ) where B denotes the Borel σ -algebra <strong>on</strong> corresp<strong>on</strong>ding to the<br />

weak*-topology. Let r 2 be the set of μ ∈ such that ‖μ‖ L 2 r. 2 r is a closed subset of . (Indeed2 r = ⋂ n φ−1([0,<br />

gn r2 ]), whereg n (x) := min(x 2 , n).) We<br />

next prove that the restricti<strong>on</strong> of V 1 to 2 r is c<strong>on</strong>tinuous: If {μ n } n∈N ⊂ 2 r c<strong>on</strong>verges weakly to μ, then, according to Skorokhod representati<strong>on</strong> theorem,<br />

there exists a sequence X n of random variables <strong>with</strong> X n ∼ μ n that c<strong>on</strong>verges a.s. to X ∼ μ. Since‖X n − X‖ L 2 2r, we c<strong>on</strong>clude that |X n − X| p is a<br />

uniformly integrable sequence (p < 2), andthus‖X n − X‖ L p → 0, implying <strong>with</strong> (H2) that V 1 [X n ]→V 1 [X]. SoV 1 is indeed c<strong>on</strong>tinuous <strong>on</strong> 2 r . Therefore<br />

V 1 : 2 → R is measurable <strong>on</strong> the trace B 2 of B <strong>on</strong> 2 .Letnext[Y ] be in 2 . As the compositi<strong>on</strong> of the measurable maps [Y |H]:(Ω, H) → ( 2 , B 2 )<br />

and V 1 : ( 2 , B 2 ) → (R, B R ), V 1 [Y |H], isthusH-measurable, as announced.


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 53<br />

If, after stage q, P2 follows <strong>with</strong> an ɛ-optimal strategy in Γ n−q [L|H q ],thepayoffY q of P1 in the n − q last stages will<br />

be less than or equal to V n−q [L|H q ]+ɛ X q + ɛ. Therefore, π (μ,σ ,τ ) [v q−1 < u q−1 ]=0, since otherwise, P2 would have a<br />

profitable deviati<strong>on</strong>: v q−1 + ɛ + Y q v q−1 + X q + 2ɛ u q−1 + X q ,ifɛ is small enough.<br />

So for all q, E[v q−1 ] E[u q−1 ]. Summing up all these inequalities, we get V n (H, L) g n (μ, σ , τ ) = V n (μ) = V n (μ), and<br />

the sec<strong>on</strong>d asserti<strong>on</strong> is proved. ✷<br />

8. The asymptotic behavior of martingales of maximal variati<strong>on</strong><br />

In this secti<strong>on</strong>, we define the M-variati<strong>on</strong> of a martingale and we analyze in Theorem 5 the limit behavior of the<br />

martingales maximizing this M-variati<strong>on</strong>. As we will see at the end of this secti<strong>on</strong>, Theorem 1 is a simple corollary of<br />

Theorem 4 and Theorem 5.<br />

Let us start <strong>with</strong> some notati<strong>on</strong>s. L 2 0 will denote hereafter the set of random variables X ∈ L2 such that E[X]=0, and<br />

2 0 be the set of measure μ <strong>on</strong> R such that X ∼ μ implies X ∈ L2 0 . For a functi<strong>on</strong> M : 2 0<br />

→ R, and a random variable<br />

X in L 2 0 ,wewillwriteM[X] instead of M([X]). W n(μ) was defined in Secti<strong>on</strong> 7 as the set of pairs (F, X) where F :=<br />

(F q ) q=0,...,n+1 is a filtrati<strong>on</strong> <strong>on</strong> a probability space, and X = (X q ) q=0,...,n+1 is an F -martingale X whose n + 1-th value X n+1<br />

is μ-distributed. Observe that if μ ∈ 2 and (F, X) ∈ W n (μ), then[X q+1 − X q |F q ]∈ 2 0 , and we may therefore define the<br />

M-variati<strong>on</strong> Vn M (F, X) as<br />

[ n−1<br />

]<br />

∑<br />

Vn M (F, X) := E M[X q+1 − X q |F q ] . (7)<br />

q=0<br />

Since we <strong>on</strong>ly will deal in this paper <strong>with</strong> functi<strong>on</strong>s M that are Lipschitz in L p -norm for p < 2, we refer to footnote 2 <strong>on</strong><br />

page 52 for a proof of the measurability of M[X q+1 − X q |F q ].LetusnextdefineV n M (μ) as<br />

V M n (μ) := sup{ V M n (F, X): (F, X) ∈ W n(μ) } . (8)<br />

For μ ∈ 2 , the functi<strong>on</strong> f μ was defined in Secti<strong>on</strong> 2 as the unique increasing functi<strong>on</strong> such that f μ (Z) ∼ μ when<br />

Z ∼ N (0, 1). The CMMV Π μ was also defined there as Π μ t<br />

:= E[ f μ (B 1 )|(B s ) st ], where B is a standard Brownian moti<strong>on</strong>.<br />

With these definiti<strong>on</strong>s, we are ready to state the main result of this secti<strong>on</strong>:<br />

Theorem 5. If M satisfies:<br />

(i) For all random variable X ∈ L 2 0 , ∀α 0: M[α X]=αM[X].<br />

(ii) There exist p ∈[1, 2[ and A ∈ R such that for all X, Y ∈ L 2 0 :<br />

∣<br />

∣ M[X]−M[Y ] A‖X − Y ‖ L p .<br />

Then for all μ ∈ 2 ,<br />

Vn M lim √ (μ) = ρ · [ ] E f μ (Z)Z ,<br />

n→∞ n<br />

where Z ∼ N (0, 1) and ρ := sup{M(ν): ν ∈ 2 0 , ‖ν‖ L 2 1}.<br />

Furthermore, if ρ > 0 and if, for all n, (F n , X n ) ∈ W n (μ) satisfies Vn M(Fn , X n ) = Vn M (μ), then the c<strong>on</strong>tinuous time representati<strong>on</strong><br />

Π n of X n defined as Πt n := X[[n·t]] n c<strong>on</strong>verges in finite-dimensi<strong>on</strong>al distributi<strong>on</strong> to the CMMV Πμ .<br />

This theorem justifies our terminology when referring to Π μ as the c<strong>on</strong>tinuous martingale of maximal variati<strong>on</strong> corresp<strong>on</strong>ding<br />

to μ.<br />

With M[X]:=‖X‖ L 1 , V M n (F, X) is just the L1 -variati<strong>on</strong> of the martingale X and we recover here Mertens and Zamir’s<br />

(1977) result <strong>on</strong> the maximal variati<strong>on</strong> of a bounded martingale, taking μ such that μ({1}) = s, μ({0}) = 1 − s. With<br />

M[X]:=‖X‖ L p ,<strong>with</strong>p ∈[1, 2[ we also recover the results of De Meyer (1998). The proof presented here is in fact inspired<br />

by this last paper.<br />

The previous theorem implies also Theorem 1.<br />

Proof of Theorem 1. Indeed, (H4) indicates that for all variable X: V 1 [X] =V 1 [(X − E[X])] and thus also, if F is a σ -<br />

algebra: V 1 [X|F]=V 1 [(X − E[X|F])|F]. Therefore, if (F, X) ∈ W n (μ), the functi<strong>on</strong> V n (F, X) defined in (5) is just equal<br />

to:<br />

V n (F, X) = V V 1<br />

n (F, X).


54 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

Since (H3) indicates that V 1 satisfies the first c<strong>on</strong>diti<strong>on</strong> of Theorem 5, (H2) indicates that V 1 fulfills the sec<strong>on</strong>d <strong>on</strong>e and<br />

V<br />

(H5) indicates that ρ > 0, Theorem 1 follows then from Theorem 4. As a byproduct, we also get that lim n (μ)<br />

n→∞ √ n<br />

=<br />

ρ · E[ f μ (Z)Z]. ✷<br />

Notice that the techniques presented in this paper could also be applied to analyze general repeated zero-sum game, and<br />

Mertens and Zamir’s (1976) result is in fact an easy c<strong>on</strong>sequence of Theorem 5 and of an adapted versi<strong>on</strong> of Theorem 4.<br />

The remaining secti<strong>on</strong>s of this paper are devoted to the proof of Theorem 5 that relies <strong>on</strong> duality techniques, a central<br />

limit theorem for martingales and a Skorokhod embedding of martingales in the Brownian filtrati<strong>on</strong>. With these techniques,<br />

the problem of maximizing the M-variati<strong>on</strong> of martingales will become a problem of maximizing a covariati<strong>on</strong> as introduced<br />

in the next secti<strong>on</strong>.<br />

We provide an upper bound for M in Secti<strong>on</strong> 10 that leads to an upper bound for Vn<br />

M in the next <strong>on</strong>e. We will then<br />

c<strong>on</strong>clude in Secti<strong>on</strong> 13 that ρ · E[ f μ (Z)Z] dominates the lim sup of V n M √ (μ) , using a central limit theorem for martingales<br />

n<br />

presented in Secti<strong>on</strong> 12.<br />

In Secti<strong>on</strong> 14, we prove that ρ · E[ f μ (Z)Z] is the lim inf of V n M √ (μ) , and in Secti<strong>on</strong> 15, we prove the c<strong>on</strong>vergence of the<br />

n<br />

Π n to Π μ .<br />

Notice that the case ρ = 0 is trivial in Theorem 5, since then M[μ] 0forallμ in 2 0 , and thus the c<strong>on</strong>stant martingale<br />

X q = E[X n+1 ],forallq n and X n+1 ∼ μ will be optimal in the maximizati<strong>on</strong> problem (8), and so Vn M (μ) = 0. In the<br />

sequel, we therefore assume ρ > 0.<br />

As a remark, observe that the hypothesis p < 2 in (ii) of Theorem 5 could not be weakened in p 2. A counterexample<br />

of this is given at the end of Secti<strong>on</strong> 13.<br />

9. The maximal covariati<strong>on</strong><br />

In this secti<strong>on</strong>, we solve an auxiliary optimizati<strong>on</strong> problem that will be central in our argument: it explains where the<br />

expressi<strong>on</strong> E[ f μ (Z)Z] appearing in Theorem 5 comes from. It is a classical M<strong>on</strong>ge–Kantorovich mass transportati<strong>on</strong> problem<br />

and claim (1) in the next theorem is well known. However, claim (2) is original and a proof is needed.<br />

In this secti<strong>on</strong> all random variables will be <strong>on</strong> a probability space (Ω, A, P), and Z will denote an N (0, 1) random<br />

variable.<br />

Theorem 6. For μ ∈ 1+ ,letusdefineα(μ) := sup{E[XZ]: X ∼ μ} then<br />

(1) α(μ) = E[ f μ (Z)Z].<br />

(2) If {X n } n∈N is a sequence of μ-distributed random variables such that E[X n Z] c<strong>on</strong>verges to α(μ) then X n c<strong>on</strong>verges in L 1 -norm to<br />

f μ (Z).<br />

Proof. Let X be a μ-distributed random variable. For a real number x, we set x + := max{x, 0} and x − := (−x) + . Since<br />

μ ∈ 1+ , X + Z and X − Z are in L 1 and <strong>with</strong> Fubini–T<strong>on</strong>elli theorem<br />

[<br />

E X + ] [ ∫∞<br />

] ∫ ∞<br />

Z = E 1 {cX} · Zdc = E[1 {cX} Z] dc.<br />

0<br />

0<br />

Similarly, E[X − · Z]= ∫ ∞<br />

0<br />

E[1 {c


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 55<br />

E [( f μ (Z) ) − · Z<br />

]<br />

− E<br />

[<br />

X− · Z ] =<br />

∫ ∞<br />

0<br />

E [ (1 {c


56 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

10. An upper bound for M<br />

On a probability space (Ω, A, P), for q 1 and r > 0, let B q r (Ω, A, P) be the set B q r (Ω, A, P) := {X ∈ L 2 (Ω, A, P)|<br />

‖X‖ L q r}. LetnextB ∗ (Ω, A, P) be defined as:<br />

B ∗ (Ω, A, P) := B 2 p′<br />

ρ (Ω, A, P) ∩ B2A (Ω, A, P),<br />

<strong>with</strong> A, ρ and p as in Theorem 5 and p ′ such that 1 p + 1 p ′ = 1. Let us finally define, for X ∈ L 2 (Ω, A, P):<br />

B(X) := sup { E[XY]: Y ∈ B ∗ (Ω, A, P) } . (9)<br />

Due to Jensen’s inequality, for all r 1, ‖E[Y |X]‖ L r ‖Y ‖ L r . Therefore, if Y ∈ B ∗ (Ω, A, P) then E[Y |X] ∈B ∗ (Ω, A, P).<br />

So, we infer that B(X) = sup{E[Xf(X)]: f (X) ∈ B ∗ (Ω, A, P)}, and therefore B(X) just depends <strong>on</strong> the distributi<strong>on</strong> [X]. In<br />

other words, if [X]=[X ′ ],thenB(X) = B(X ′ ),evenifX and X ′ are defined <strong>on</strong> different probability spaces. We will therefore<br />

abuse the notati<strong>on</strong>s and write B[X] instead of B(X).<br />

Lemma 7.<br />

(1) For all X ∈ L 2 0 (Ω, A, P): M[X] ρ ·‖X‖ L 2 .<br />

(2) If B(Ω, A, P) := {X ∈ L 2 (Ω, A, P): B[X] 1},then<br />

B(Ω, A, P) ⊂ ( c<strong>on</strong>v B 2 1 (Ω, A, P) ∪ B p 1<br />

(Ω, A, P) ) .<br />

ρ 2A<br />

(3) For all X ∈ L 2 0<br />

(Ω, A, P): B[X] M[X].<br />

Proof. Claim (1) is an obvious c<strong>on</strong>sequence of the definiti<strong>on</strong> of ρ as sup{M[X]: X ∈ L 2 0 , ‖X‖ L2 1} and of the 1-<br />

homogeneity of M.<br />

We next prove claim (2): Let C denote c<strong>on</strong>v(B 2 1<br />

(Ω, A, P) ∪ B p 1<br />

(Ω, A, P)). SinceB 2 1<br />

(Ω, A, P) and B p 1<br />

(Ω, A, P) are<br />

ρ 2A<br />

ρ 2A<br />

bounded c<strong>on</strong>vex closed sets in L 2 -norm (p < 2), they are weakly compact. C is therefore closed. Therefore, if Z ∈ L 2 (Ω, A, P)<br />

does not bel<strong>on</strong>g to C, wecanseparate{Z} and C in L 2 (Ω, A, P) by a separating vector Y : E[Y Z] > α := sup{E[Y X]: X ∈ C}.<br />

In particular α sup{E[Y X]: X ∈ B 2 1<br />

}= ρ 1 ·‖Y ‖ L 2 , and α sup{E[Y X]: X ∈ B p 1<br />

}= 1<br />

2A ·‖Y ‖ L p′ . This indicates that Y ′ :=<br />

ρ 2A<br />

Y<br />

α ∈ B∗ (Ω, A, P). Therefore B[Z] E[Y ′ Z] > 1 and so Z /∈ B(Ω, A, P). So the complementary of C is included in the<br />

complementary of B(Ω, A, P), orequivalently:B(Ω, A, P) ⊂ C.<br />

To prove claim (3) observe that both M and B are 1-homogeneous <strong>on</strong> L 2 0<br />

that, for all X ∈ L 2 0<br />

(Ω, A, P). Therefore, we just have to prove<br />

(Ω, A, P), M[X] 1 whenever B[X] 1. But if B[X] 1, then X ∈ B(Ω, A, P). So, by the previous claim<br />

X ∈ C. SinceC is the c<strong>on</strong>vex hull of two c<strong>on</strong>vex sets, we get that X = λY + λ ′ Y ′ ,<strong>with</strong>λ,λ ′ 0, λ + λ ′ = 1, Y ∈ B 2 1<br />

and<br />

ρ<br />

Y ′ ∈ B p 1<br />

.SinceE[X]=0, we also have X = λ(Y − E[Y ]) + λ ′ (Y ′ − E[Y ′ ]). Due to property (ii) in Theorem 5, we get:<br />

2A<br />

M[X] M [ λ ( Y − E[Y ] )] + A ∥ ∥λ ′( Y ′ − E [ Y ′])∥ ∥<br />

L p .<br />

Since ‖(Y − E[Y ])‖ L 2 ‖Y ‖ L 2 ρ 1 , it follows from claim (1) that the first term is bounded by λ. The sec<strong>on</strong>d <strong>on</strong>e is bounded<br />

by λ ′ since ‖Y ′ − E[Y ′ ]‖ L p ‖Y ′ ‖ L p +‖E[Y ′ ]‖ L p 2‖Y ′ ‖ L p 1 .ThusM[X] 1 and the lemma is proved.<br />

A ✷<br />

11. An upper bound for V M n (μ)<br />

For (F, X) ∈ W n (μ), Vn M(F, X) was defined in (7). The term M[X q+1 − X q |F q ] involved there is then dominated by<br />

B[X q+1 − X q |F q ], and we will next focus <strong>on</strong> the B variati<strong>on</strong> Vn B (F, X) that dominates the M-variati<strong>on</strong>.<br />

The next lemma presents E[B[X q+1 − X q |F q ]] as the result of an optimizati<strong>on</strong> problem.<br />

Let F 1 ⊂ F 2 be two σ -algebras <strong>on</strong> a probability space (Ω, A, P). LetB ∗ (F 2 |F 1 ) denote the set of Y ∈ L 2 (F 2 ) such that<br />

E[Y 2 |F 1 ] ρ 2 and E[|Y | p′ |F 1 ] (2A) p′ .LetB(ρ,C) ∗ (F 2|F 1 ) denote the set of Y ∈ L 2 (F 2 ) such that<br />

(1) E[Y |F 1 ]=0;<br />

(2) E[Y 2 |F 1 ]=ρ 2 ;<br />

(3) E[|Y | p′ |F 1 ] C p′ .<br />

Lemma 8.<br />

(1) For all X ∈ L 2 0 (F 2):<br />

E [ B[X|F 1 ] ] = sup { E[XY]|Y ∈ B ∗ (F 2 |F 1 ) } .


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 57<br />

(2) If there exists in L 2 (Ω, F 2 , P) a random variable U that is independent of σ (F 1 , X) and taking the values 1 and −1 <strong>with</strong> probability<br />

1/2 then<br />

E [ B[X|F 1 ] ] sup { E[XY]|Y ∈ B ∗ (ρ,4A+ρ) (F 2|F 1 ) } .<br />

Proof. If X ∈ L 2 (F 2 ) and Y ∈ B ∗ (F 2 |F 1 ) then E[XY]=E[E[XY|F 1 ]]. Since c<strong>on</strong>diti<strong>on</strong>ally to F 1 , Y bel<strong>on</strong>gs to B ∗ ,weget<br />

E[XY|F 1 ] B[X|F 1 ], and thus E[B[X|F 1 ]] sup{E[XY]|Y ∈ B ∗ (F 2 |F 1 )}.<br />

To prove the reverse inequality, we just have to prove that, ∀ɛ > 0, there exists a measurable map φ : 2 × R → R such<br />

that, ∀μ ∈ 2 ,ifX ∼ μ, thenE[φ(μ, X) 2 ] ρ 2 , E[|φ(μ, X)| p′ ] (2A) p′ and B(μ) − ɛ E[Xφ(μ, X)]. Indeed, the random<br />

variable Y := φ([X|F 1 ], X) bel<strong>on</strong>gs then to B ∗ (F 2 |F 1 ) and E[XY] E[B[X|F 1 ]] − ɛ.<br />

We now prove the existence of such a measurable map. As argued just after Eq. (9), B(μ) = sup{E μ [Xf(X)]:<br />

E μ [ f (X) 2 ] ρ 2 ; E μ [| f (X)| p ] (2A) p }. Since there exists a countable set { f n } n∈N of c<strong>on</strong>tinuous functi<strong>on</strong>s, <strong>with</strong> f 0 ≡ 0, that<br />

is dense in L 2 (R, B R , μ), forallμ ∈ 2 ,wealsohaveB(μ) = sup n g n (μ), where g n (μ) := E μ [Xf n (X)] if E μ [ f n (X) 2 ] ρ 2<br />

and E μ [| f n (X)| p ] (2A) p , and g n (μ) := 0 otherwise. The maps μ → g n (μ) are measurable <strong>with</strong> respect to the Borel<br />

σ -algebra B 2 associated <strong>with</strong> the weak topology <strong>on</strong> 2 . Indeed, the maps μ → E μ [Xf n (X)], μ → E μ [ f n (X) 2 ] and<br />

μ → E μ [| f n (X)| p ] are weakly c<strong>on</strong>tinuous and thus measurable <strong>with</strong> respect to B 2 . We infer therefore that μ → B(μ)<br />

and thus μ → B(μ) − g n (μ) are also B 2 -measurable. For ɛ > 0, we may then define N(μ) as the smallest integer n such<br />

that B(μ) − g n (μ) ɛ. Themapφ : (μ, x) → f N(μ) (x) will therefore be measurable.<br />

It has also the required properties: ∀μ ∈ 2 ,ifX ∼ μ, thenE[φ(μ, X) 2 ] ρ 2 , E[|φ(μ, X)| p′ ] (2A) p′ and B(μ) − ɛ <br />

E[Xφ(μ, X)].<br />

Indeed, for all μ ∈ 2 ,eitherB(μ) ɛ, which implies N(μ) = 0 and thus E μ [ f N(μ) (X)X]=0 B(μ) − ɛ, since f 0 ≡ 0.<br />

We also have in this case E μ [ f N(μ) 2 (X)]=0 ρ2 and E μ [| f N(μ) (X)| p ]=0 (2A) p .<br />

Or, B(μ) >ɛ, and thus g N(μ) (μ) >0, which indicates that E μ [ f N(μ) 2 (X)] ρ2 , E μ [| f N(μ) (X)| p ] (2A) p and also<br />

E μ [ f N(μ) (X)X]=g N(μ) (μ) B(μ) − ɛ, as it results from the definiti<strong>on</strong> of g n .<br />

We next turn to claim (2): Just observe that if Y ∈ B ∗ (F 2 |F 1 ),thenY ′ := E[Y |X] also bel<strong>on</strong>gs to B ∗ (F 2 |F 1 ) and has<br />

thus the property that E[XY]=E[XY ′ ]. Now, c<strong>on</strong>sider Y ′′ := Y ′ − E[Y ′ |F 1 ].SinceX ∈ L 2 0 (F 2|F 1 ),wehaveE[XE[Y ′ |F 1 ]] = 0.<br />

Therefore E[XY]=E[XY ′′ ]. Now, observe that<br />

θ 2 := E [( Y ′′) 2<br />

|F1<br />

]<br />

= E<br />

[(<br />

Y<br />

′ ) 2<br />

|F1<br />

]<br />

−<br />

(<br />

E<br />

[<br />

Y ′ |F 1<br />

]) 2<br />

E<br />

[<br />

Y 2 |F 1<br />

]<br />

ρ 2 .<br />

Finally, let Y ′′′ be defined as Y ′′′ := Y ′′ + √ ρ 2 − θ 2 U ,sinceU is independent of σ (F 1 , X), and since Y ′′ is measurable <strong>on</strong><br />

this σ -algebra, we get obviously<br />

E[XY]=E [ XY ′′′] , E [ Y ′′′ |F 1<br />

]<br />

= 0 and E<br />

[(<br />

Y<br />

′′′ ) 2<br />

|F1<br />

]<br />

= ρ 2 .<br />

Observing that (E[(Y ′′′ ) p′ |F 1 ]) 1 p ′<br />

is just a c<strong>on</strong>diti<strong>on</strong>al L p′ -norm, we get<br />

(<br />

E<br />

[∣ ∣Y ′′′ ∣ ∣ p′ |F 1<br />

]) 1<br />

p ′ ( E [∣ ∣ Y<br />

′′ ∣ ∣ p′ |F 1<br />

]) 1<br />

p ′ + ρ<br />

( E [∣ ∣ Y<br />

′ ∣ ∣ p′ |F 1<br />

]) 1<br />

p ′ + ( E [∣ ∣ E<br />

[<br />

Y ′ |F 1<br />

]∣ ∣<br />

p ′ |F 1<br />

]) 1<br />

p ′ + ρ<br />

2 ( E [∣ ∣ Y<br />

′ ∣ ∣ p′ |F 1<br />

]) 1<br />

p ′ + ρ<br />

2 ( E [ |Y | p′ |F 1<br />

]) 1<br />

p ′ + ρ<br />

4A + ρ.<br />

Therefore, for all Y ∈ B ∗ (F 2 |F 1 ),thereisaY ′′′ ∈ B(ρ,4A+ρ) ∗ (F 2|F 1 ) such that E[XY]=E[XY ′′′ ], and claim (2) then follows<br />

from claim (1). ✷<br />

LetusnowusethislemmatocomputeVn M(F, X) for a pair (F, X) ∈ W n(μ) defined <strong>on</strong> a probability space (Ω, A, P).<br />

Let us first enlarge this space obtaining a new <strong>on</strong>e (Ω ′ , A ′ , P ′ ) where A may be seen as a sub-σ -algebra of A ′ , P ′ and<br />

P coincide <strong>on</strong> A and where there is a system of n independent random variables (U q ) q=1,...,n , independent of A, <strong>with</strong><br />

P ′ (U q = 1) = P ′ (U q =−1) = 1/2. C<strong>on</strong>sider next the filtrati<strong>on</strong> F ′ defined by F ′ q := σ (F q, U k , k q). X is then also a martingale<br />

<strong>on</strong> F ′ and [X q+1 − X q |F q ]=[X q+1 − X q |F ′ q ]. Therefore<br />

Vn M (F, X) V n B (F, X) = V (<br />

n<br />

B F ′ , ) X .<br />

We will denote B(ρ,4A+ρ) ∗ (F ′ ) the set of F ′ - adapted processes Y such that for all q = 1,...,n: Y q ∈ B(ρ,4A+ρ) ∗ (F ′ q |F ′ q−1 ).<br />

Then, since X q − X q−1 ∈ L 2 0 (F ′ q |F ′ q−1 ), we may apply claim (2) of last lemma to get


58 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

V B n<br />

(<br />

F ′ , X ) =<br />

n∑ [ [ E B X q − X q−1 |F ′ ]]<br />

q−1<br />

q=1<br />

n∑<br />

<br />

[ ]<br />

sup E (X q − X q−1 ) · Y q .<br />

Y ∈B(ρ,4A+ρ) ∗ (F ′ )<br />

q=1<br />

Since X is an F ′ -martingale and E[Y q |F ′ q−1 ]=0, we get<br />

E [ (X q − X q−1 ) · Y q<br />

]<br />

= E[Xq · Y q ]=E [ E [ X n+1 |F ′ q]<br />

· Yq<br />

]<br />

= E[Xn+1 · Y q ],<br />

and therefore<br />

V M n<br />

(<br />

F ′ , ) X sup E<br />

Y ∈B(ρ,4A+ρ) ∗ (F ′ )<br />

[<br />

X n+1 ·<br />

n∑<br />

Y q<br />

]. (10)<br />

q=1<br />

For a given Y ∈ B(ρ,4A+ρ) ∗ (F ′ ),letS q be defined as S 0 := 0, S q := S q−1 + Y q . Observe then that S is an F ′ -martingale.<br />

We will denote S (ρ,4A+ρ) (F ′ ) the set of F ′ -martingale S whose increments S q+1 − S q bel<strong>on</strong>g to B(ρ,4A+ρ) ∗ (F ′ q |F ′ q−1 ),for<br />

all q, and such that S 0 = 0. The last formula becomes then<br />

(<br />

Vn<br />

M F ′ , ) X sup<br />

S∈S(ρ,4A+ρ) ∗ (F ′ )<br />

E[X n+1 · S n ]. (11)<br />

Let us make two comments <strong>on</strong> the last formula: the quantity Vn M(F ′ , X) depends <strong>on</strong> the laws [X q+1 − X q |F q ] which<br />

are intimately related to the filtrati<strong>on</strong> F . The bound we found in the last formula just depends <strong>on</strong> the laws [X q+1 −<br />

X q |X 1 ,...,X q ]. Therefore, if we create a martingale ˜X <strong>on</strong> another filtrati<strong>on</strong> G <strong>with</strong> the same law as X — we call this<br />

procedure the embedding of X in the filtrati<strong>on</strong> G —, the right-hand side of last inequality can equivalently be evaluated<br />

<strong>on</strong> ˜X.<br />

The sec<strong>on</strong>d comment is that we will have to deal <strong>with</strong> V n M(F ′ ,X)<br />

√<br />

S<br />

, and will then have to evaluate E[X n n+1 · √n n<br />

], for<br />

S ∈ S(ρ,4A+ρ) ∗ (F ′ ). Since the increments of S have a c<strong>on</strong>diti<strong>on</strong>al variance equal to ρ 2 S<br />

, √n n<br />

will be approximatively normally<br />

distributed, due to a central limit theorem for martingales. We need however precise bounds for this approximati<strong>on</strong>. These<br />

bounds are provided in the next secti<strong>on</strong> which is in fact the crux point of the argument. We embed there both martingales<br />

√S<br />

and X in a Brownian filtrati<strong>on</strong>.<br />

n<br />

12. The embedding in the Brownian filtrati<strong>on</strong><br />

Let B be a Brownian moti<strong>on</strong> <strong>on</strong> a probability space (Ω 0 , A 0 , P 0 ) and let G be the natural filtrati<strong>on</strong> of B. Skorokhod<br />

posed the following questi<strong>on</strong>: Given a probability distributi<strong>on</strong> μ ∈ p′ ,isthereaG-stopping time θ such that B θ ∼ μ? To<br />

avoid trivial uninteresting soluti<strong>on</strong>s to this problem, <strong>on</strong>e further requires that E[θ p′<br />

2 ] < ∞. It is well known that Skorokhod’s<br />

problem has a soluti<strong>on</strong> for all μ ∈ p′<br />

0 (see for instance Azéma and Yor, 1979) and we will denote θ μ <strong>on</strong>e of these soluti<strong>on</strong>s.<br />

We also will need the following fact: For all p ′ > 1, there exist two n<strong>on</strong>-negative c<strong>on</strong>stants c p ′ and C p ′ , called the<br />

Burkholder–Davis–Gundy c<strong>on</strong>stants (see Burkholder, 1973), such that, for all G-stopping times τ τ ′ :<br />

[<br />

E τ p′ ]<br />

2 < ∞ ⇒ cp ′ · [( p′<br />

E τ − τ ′) 2<br />

[ ]<br />

|G τ ′]<br />

E |Bτ − B τ ′| p′ |G τ ′ C p ′ · [( p′<br />

E τ − τ ′) 2<br />

|G τ ′]<br />

.<br />

In the particular case p ′ = 2, we have c 2 = C 2 = 1.<br />

Lemma 9. Let R = (R q ) q:=0,...,n be a martingale <strong>with</strong> R n ∈ L p′<br />

0 , then there is an increasing sequence of stopping times {τ q} q:=0,...,n<br />

such that E[τ p′<br />

2<br />

n ] < ∞ and such that both processes R and ˆR have the same distributi<strong>on</strong> where ˆRq := B τq .<br />

Proof. Just take τ 0 := θ [R0 ] so that [ ˆR0 ]=[B τ0 ]=[R 0 ].Onceτ q is defined, define τ q+1 as follows: B ′ t := B τ q +t − B τq is another<br />

Brownian moti<strong>on</strong> <strong>on</strong> its natural filtrati<strong>on</strong> G ′ .Forall(r 0 ,...,r q ) ∈ R q+1 define ˜θ(r 0 ,...,r q ) := θ ′ [R q+1 −R q |R 0 =r 0 ;...;R q =r q ] ,<br />

where θ μ ′ is a soluti<strong>on</strong> of μ-Skorokhod’s problem for the Brownian moti<strong>on</strong> B′ . This mapping can be chosen measurable<br />

from R q+1 to (Ω 0 , A 0 ), and define τ q+1 := τ q + ˜θ(ˆR0 ,..., ˆRq ). Then ˆRq+1 − ˆRq = B τq+1 − B τq = B ′˜θ(ˆR0 ,..., ˆRq ) . Therefore<br />

[ ˆRq+1 − ˆRq | ˆR0 ,..., ˆRq ]=[R q+1 − R q |R 0 ,...,R q ]. We then c<strong>on</strong>clude that R and ˆR have the same laws. Next, since<br />

c p · E[ ˜θ(ˆR0 ,..., ˆRq ) p′<br />

2 |G τq ] E[( ˆRq+1 − ˆRq ) p′ |G τq ], we c<strong>on</strong>clude by inducti<strong>on</strong> that E[τ p′<br />

2<br />

n ] < ∞. ✷<br />

We are now ready to start the embedding procedure. Let us c<strong>on</strong>sider (F ′ , X) ∈ W n (μ) and S ∈ S(ρ,4A+ρ) ∗ (F ′ ),asinthe<br />

last secti<strong>on</strong>.


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 59<br />

Let R denote R :=<br />

S<br />

ρ √ . To embed both R and X, we will have to slightly perturb the above procedure: For ɛ > 0<br />

n<br />

we define τ 0 = 0, τ 1 := ɛ and ˆR0 := 0. There exists a functi<strong>on</strong> f 0 such that [ f 0 ( √ ɛ Z)]=[X 0 ] if Z ∼ N (0, 1) and we set<br />

2<br />

ˆX 0 := f 0 (B τ 1<br />

− B τ0 ). The random vectors (R 0 , X 0 ) and ( ˆR0 , ˆX0 ) will thus have the same distributi<strong>on</strong>.<br />

2<br />

For q = 0,...,n − 1, we then define τ q+1 , τ q+<br />

3 , ˆRq+1 recursively as follows: Let G ′ be the natural σ -algebra of B ′ t :=<br />

2<br />

B t+τq+ 1<br />

2<br />

− B τq+ . Define as above<br />

1<br />

2<br />

˜θ(r 0 ,...,r q , x 0 ,...,x q ) := θ ′ [R q+1 −R q |R 0 =r 0 ,...,R q =r q ,X 0 =x 0 ,...,X q =x q ] ,<br />

and then set τ q+1 := τ q+<br />

1 + ˜θ(ˆR0 ,..., ˆRq , ˆX0 ,..., ˆXq ), τ<br />

2<br />

q+<br />

3<br />

2<br />

:= τ q+1 + ɛ and finally ˆRq+1 := ˆRq + B τq+1 − B τq+ . It follows<br />

1<br />

2<br />

from this definiti<strong>on</strong> that ˆRq+1 − ˆRq has the same law c<strong>on</strong>diti<strong>on</strong>ally to ( ˆRs , ˆXs ) sq as R q+1 − R q c<strong>on</strong>diti<strong>on</strong>ally to (R s , X s ) sq ,<br />

and thus (( ˆRs ) sq+1 ,(ˆXs ) sq ) and ((R s ) sq+1 ,(X s ) sq ) have the same law. There exists a functi<strong>on</strong> f q+1 : R q+1 ×R q ×R → R<br />

such that, ∀((r s ) sq+1 ,(x s ) sq ) ∈ R q+1 × R q ,<strong>with</strong>Z ∼ N (0, 1):<br />

[<br />

fq+1<br />

(<br />

(rs ) sq+1 ,(x s ) sq , √ ɛ Z )] = [ X q+1 |(R s = r s ) sq+1 ,(X s = x s ) sq<br />

]<br />

.<br />

We then set ˆXq+1 := f q+1 (( ˆRs ) sq+1 ,(ˆXs ) sq , B τq+ − B<br />

3 τq+1 ), and it follows that (R s , X s ) sq+1 and ( ˆRs , ˆXs ) sq+1 are equally<br />

2<br />

distributed.<br />

It is c<strong>on</strong>venient to define also τ n+1 as τ n+1 = τ n+<br />

1 := τ n + ɛ and ˆXn+1 := f n+1 (( ˆRs ) sn ,(ˆXs ) sn , B τn+ − B<br />

1 τn ), where<br />

2<br />

2<br />

f n+1 is such that, for all (r s ), (x s ):<br />

[ (<br />

fq+1 (rs ) sn ,(x s ) sn , √ ɛ )] Z = [ ]<br />

X n+1 |(R s = r s ) sn ,(X s = x s ) sn ,<br />

whenever Z ∼ N (0, 1).<br />

The resulting process ( ˆR, ˆX) has the same distributi<strong>on</strong> as (R, X). It follows from the above definiti<strong>on</strong> that ( ˆRq , ˆXq ) is<br />

G τq -measurable and the law ( ˆRq , ˆXq ) c<strong>on</strong>diti<strong>on</strong>ally to G τq−1 is just the law of ( ˆRq , ˆXq ) c<strong>on</strong>diti<strong>on</strong>ally to ( ˆRs , ˆXs ) s


60 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

Therefore, E[τ [[nt]] ]=[[nt]](ɛ + 1 ), as announced.<br />

n<br />

We next prove claim (2). Since E[τ q − τ q−1 |G τq−1 ]=E[τ q − τ q−1 ],weget<br />

where<br />

[[nt]]<br />

∑( τ [[nt]] − E[τ [[nt]] ]= (τq − τ q−1 ) − E[τ q − τ q−1 ] ) = Q [[nt]] ,<br />

Q s :=<br />

q=1<br />

s∑ (<br />

(τq − τ q−1 ) − E[τ q − τ q−1 |G τq−1 ] ) =<br />

q=1<br />

s∑<br />

(<br />

q=1<br />

τ q − τ q−<br />

1<br />

2<br />

− 1 n<br />

The process Q = (Q s ) s=0,...,n is clearly a G τs -martingale starting at 0. Since p ∈[1, 2[ and 1 p + 1 p ′ = 1, we get p ′ > 2, and so,<br />

˜p := min(p′ ,4)<br />

2<br />

∈]1, 2]. Therefore<br />

∥<br />

∥τ [[nt]] − E[τ [[nt]] ] ∥ ∥<br />

L 1 =‖Q [[nt]] ‖ L 1 ‖Q n ‖ L 1 ‖Q n ‖ L ˜p . (13)<br />

We claim next that<br />

[<br />

E |Q n | ˜p] ∑n−1<br />

2 2− ˜p [ E |Q k+1 − Q k | ˜p] . (14)<br />

k=0<br />

This follows at <strong>on</strong>ce from a recursive use of the relati<strong>on</strong>:<br />

E [ |x + Y | ˜p] |x| ˜p + 2 2− ˜p E [ |Y | ˜p] ,<br />

that holds for all x in R, whenever Y is a centered random variable: Indeed,<br />

|x + Y | ˜p −|x| ˜p = Y<br />

∫ 1<br />

Thus, since E[Y ]=0, we get<br />

[<br />

E [ |x + Y | ˜p] −|x| ˜p = E<br />

0<br />

˜p|x + sY|˜p−1 sgn(x + sY)ds.<br />

Y<br />

∫ 1<br />

0<br />

)<br />

.<br />

˜p ( |x + sY|˜p−1 sgn(x + sY) −|x| ˜p−1 sgn(x) ) ]<br />

ds .<br />

Since ˜p 2, straightforward computati<strong>on</strong> indicates that, for fixed a, the functi<strong>on</strong> g(x) := ||x+a| ˜p−1 sgn(x+a)−|x| ˜p−1 sgn(x)|<br />

reaches its maximum at x =−a/2, implying g(x) 2 2− ˜p |a| ˜p−1 .<br />

So, E[|x + Y | ˜p ]−|x| ˜p E[|Y | ∫ 1<br />

0 22− ˜p ˜p|sY|˜p−1 ds]=2 2− ˜p E[|Y | ˜p ], as announced and inequality (14) follows.<br />

Next ‖Q k+1 − Q k ‖ L ˜p =‖τ k+1 − τ k+<br />

1<br />

2<br />

τ k+<br />

1 ‖<br />

2 L ˜p , and thus<br />

− 1 n ‖ L ˜p ‖τ k+1 − τ k+<br />

1<br />

2<br />

‖Q k+1 − Q k ‖ L ˜p 2‖τ k+1 − τ k+<br />

1 ‖<br />

2 L ˜p 2‖τ k+1 − τ k+<br />

1 ‖ p ′ .<br />

2 L 2<br />

‖ L ˜p + 1 n . Since 1 n = E[τ k+1 − τ k+<br />

1<br />

2<br />

], wealsohave 1 n ‖τ k+1 −<br />

Finally, ˆRk+1 − ˆRk = B τk+1 − B τk+ 1<br />

2<br />

. Therefore, we get <strong>with</strong> Burkholder–Davis–Gundy inequality, and since S ∈<br />

S ∗ (ρ,4A+ρ) (F ′ ):<br />

[<br />

E (τ k+1 − τ k+<br />

1 ) p′ ] 1 [ ]<br />

2 E | ˆRk+1 − ˆRk | p′ 1 [ ]<br />

= E |S<br />

2 c p ′<br />

c p ′ρ p′ n p′ k+1 − S k | p′ (4A + ρ) p′<br />

.<br />

2<br />

c p ′ρ p′ n p′<br />

2<br />

So: E[|Q k+1 − Q k | ˜p ] 2 ˜p (4A + ρ) 2 ˜p (c p ′) − 2 ˜p<br />

p ′ ρ −2 ˜p n − ˜p , and, <strong>with</strong> (14), we c<strong>on</strong>clude<br />

E [ |Q n | ˜p] 2 2 (4A + ρ) 2 ˜p (c p ′) − 2 ˜p<br />

p ′ ρ −2 ˜p n 1− ˜p .<br />

Therefore, <strong>with</strong> (13), we get:<br />

∥<br />

∥τ [[nt]] − E[τ [[nt]] ] ∥ ( ) L 1 2 2˜p<br />

4A + ρ 2<br />

−1<br />

n<br />

1˜p<br />

(c p ′) 1 p ′ ρ<br />

and claim (2) is proved.<br />

We next prove claim (3): let τ denote E[τ [[nt]] ].Then<br />

‖B τ[[nt]] − B t ‖ L 2 ‖B τ[[nt]] − B τ ‖ L 2 +‖B τ − B t ‖ L 2 = √ ‖τ [[nt]] − τ ‖ L 1 + √ |τ − t|.<br />

The first term is bounded by claim (2), and the sec<strong>on</strong>d <strong>on</strong>e by claim (1).<br />


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 61<br />

13. An upper bound for lim sup V M n (μ)/√ n<br />

Let us c<strong>on</strong>sider (F ′ , X) ∈ W n (μ). According to (11), we have:<br />

Vn M(F ′ [<br />

, X)<br />

S n<br />

√ sup E X n+1 · √<br />

].<br />

n S∈S(ρ,4A+ρ) ∗ (F ′ )<br />

n<br />

For S ∈ S ∗ (ρ,4A+ρ) (F ′ ),letusdefineR n :=<br />

S n<br />

ρ √ n<br />

secti<strong>on</strong>, for ɛ > 0.<br />

Then E[X n+1 · Sn √ n<br />

]=ρ · E[X n+1 · R n ]=ρ · E[ ˆXn+1 · ˆRn ].<br />

and let us embed (X, R) in the Brownian filtrati<strong>on</strong>, as d<strong>on</strong>e in the last<br />

Claim (3) for t = 1 in Lemma 10 yields ‖B 1 − B τn ‖ L 2 κ · n 1<br />

p ′ − 1 ∧4 2<br />

+ √ ɛn. With (12), we get then ‖ ˆRn − B 1 ‖ L 2 <br />

κ · n 1<br />

p ′ − 1 ∧4 2<br />

+ 2 √ ɛ · n.<br />

Therefore, since ˆXn+1 ∼ μ and B 1 ∼ N (0, 1),<br />

E[ ˆXn+1 · ˆRn ] E[ ˆXn+1 · [ B 1 ]+E ˆXn+1 · ( ˆRn − B 1 ) ]<br />

E[ ˆXn+1 · B 1 ]+‖ˆXn+1 ‖ L 2 ·‖ˆRn − B 1 ‖ L 2<br />

α(μ) +‖μ‖ L 2 · (κ · n 1<br />

p ′ ∧4 − 1 2<br />

+ 2 √ ɛ · n ) ,<br />

where α(μ) = E[ f μ (Z)Z] was defined in Theorem 6. Since this holds for all ɛ > 0 and all S ∈ S(ρ,4A+ρ) ∗ (F ′ ), we c<strong>on</strong>clude<br />

that for all (F ′ , X) ∈ W n (μ):<br />

V M n (F ′ , X)<br />

√ n<br />

ρ · α(μ) + ρ ·‖μ‖ L 2 · κ · n 1<br />

p ′ ∧4 − 1 2<br />

.<br />

Since p ′ > 2, and since the c<strong>on</strong>stant κ in Lemma 10 is independent of n, wethushaveproved:<br />

Theorem 11. Under the hypotheses of Theorem 5,<br />

lim sup<br />

n→∞<br />

V M n (μ) √ n<br />

ρ · E [ f μ (Z)Z ] .<br />

We will prove in the next secti<strong>on</strong> that ρ · α(μ) is the limit of V n M √ (μ) as n increases to ∞. To c<strong>on</strong>clude this secti<strong>on</strong>, we<br />

n<br />

give here an example to illustrate that p must be strictly less than 2 in hypothesis (ii) of Theorem 5 in order to get the<br />

satisfies hypothesis (i) of Theorem 5, and would also satisfy hypothesis (ii) <strong>with</strong><br />

result: Clearly, the functi<strong>on</strong> M[μ]:=‖μ‖ L 2<br />

A = 1ifp = 2 was allowed. For this M, ρ = 1. Let then μ be the probability<br />

√<br />

that assigns a weight 1/2 to+1 and −1. Let X n<br />

be the unique martingale of length n + 1 such that ∀q = 0,...,n, |Xq n|= q<br />

n and such that Xn n+1 := Xn n . In other words, if, for<br />

√<br />

√ √<br />

q < n, Xq n = q<br />

n ,thenXn q+1 jumps to q+1<br />

<strong>with</strong> probability γ or −<br />

n<br />

q+1<br />

n<br />

<strong>with</strong> probability 1 − γ , where γ := 1 2 (1 + √<br />

q<br />

q+1 ),<br />

√<br />

and symmetric jumps are made if Xq n =− q<br />

n . An easy computati<strong>on</strong> shows that E[(Xn q+1 − Xn q )2 |X n 1 ,...,Xn q ]= 1 , and thus,<br />

n<br />

if F n denotes the natural filtrati<strong>on</strong> of X n , we getVn M(Fn , X n ) = √ n. Since for all pair (F ′ , X) ∈ W n (μ), we can write<br />

as in (11): V M n (F ′ , X) = E[X n+1 S n ], where S n = ∑ n<br />

k=1 Y k,<strong>with</strong>E[Y 2 k ]=1wegetE[S2 n<br />

inequality, it comes Vn M(F ′ √<br />

, X) ‖X n+1 ‖ L 2√ n =‖μ‖L 2√ n = n. Therefore:<br />

√ (<br />

n V<br />

M<br />

n (μ) Vn<br />

M F n , X n) = √ n,<br />

]=n, and due to Cauchy–Schwarz<br />

and thus<br />

Vn M lim √ (μ) = 1 > ρ · [ ] E f μ (Z)Z = [ E |Z| ] √<br />

2<br />

=<br />

n→∞ n π ,<br />

since f μ (Z) = 1 {Z0} − 1 {Z


62 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

Using Lemma 9, we can c<strong>on</strong>struct, for each n, an increasing sequence (τq n) q=0,...,n of stopping times <strong>on</strong> the Brownian<br />

filtrati<strong>on</strong> G such that Yq n := √ n · (B τ<br />

n<br />

q<br />

− B τ<br />

n ) is an i.i.d. sequence <strong>with</strong> [Y n q−1 q ]=[Y ]. Observe in particular that τ q n − τ n q−1 is<br />

also an i.i.d. sequence. We also set τ n n+1 := τ n n ∨ 1.<br />

The argument of Lemma 10 can be applied to this sequence of stopping times, replacing ρ by 1, p ′ by 4, ɛ by 0 and<br />

4A + ρ by ‖Y ‖ L 4 . We obtain in this way:<br />

Lemma 12.<br />

(1) For all t ∈[0, 1]: E[τ[[nt]] n [[nt]]<br />

]=<br />

n .<br />

(2) ‖τ[[nt]] n − E[τ [[nt]] n ]‖ L 2 γ 2 · n − 1 ‖Y ‖<br />

2 ,whereγ := L 4<br />

(3) ∀u ∈[0, 1], ∀q [[nu]]: p(τq n < u) γ 4<br />

n( q .<br />

n −u)2<br />

(4) ‖B 1 − B τ<br />

n<br />

n<br />

‖ L 2 γ · n − 1 4 .<br />

.<br />

(c 4 ) 1 4<br />

Proof. By c<strong>on</strong>structi<strong>on</strong> of the sequence τ n q , θn q := τ n q − τ n q−1 is an i.i.d. sequence of random variables and 1 = E[(Y n q )2 ]=<br />

n · E[(B τ<br />

n<br />

q<br />

− B τ<br />

n )2 ]=n · E[θ n q−1 q ]. Burkholder–Davis–Gundy inequality indicates that<br />

c 4 · var [ θ n q<br />

]<br />

c4 · E [( θ n q<br />

) 2 ]<br />

E<br />

[<br />

(Bτ n<br />

q<br />

− B τ<br />

n<br />

q−1 )4] = E [ Y 4] /n 2 .<br />

Therefore, since τ[[nt]] n = ∑ [[nt]]<br />

q=1 θn,wegetE[τ q [[nt]] n ]=[[nt]]/n and<br />

∥<br />

∥τ [[nt]] n − E[ τ[[nt]]]∥ n ∥<br />

2<br />

= ( )<br />

L 2 var τ[[nt]]<br />

n E[Y 4 ]·[[nt]]<br />

. E[Y 4 ]<br />

c 4 · n 2 c 4 · n .<br />

Claim (3) is just a <strong>on</strong>e-sided Chebichev inequality: Indeed, <strong>with</strong> t := q/n, claims (1) and (2) indicate that E[τq n]=q/n<br />

and var(τq n) γ 4 /n. Therefore: p(τq n < u) = p(τ q n − E[τ q n] < u − q/n) var(τ q n)<br />

(q/n−u) . 2<br />

We finally c<strong>on</strong>clude ‖B τ<br />

n<br />

n<br />

− B 1 ‖ 2 = E[|τ n L 2 n − 1|] ‖Y √ ‖2 L 4<br />

c4·n<br />

, and the lemma is proved. ✷<br />

We define next Fq n := G τq<br />

n and Xq n := E[ f μ(B 1 )|Fq n],forq = 0,...,n + 1. Since τ n n+1 1, we have Xn n+1 = f μ(B 1 ), and<br />

due to the definiti<strong>on</strong> of f μ ,wehaveX n n+1 ∼ μ. Therefore (Fn , X n ) ∈ W n (μ).<br />

We will have to compute Vn M(Fn , X n ). To do so, it is c<strong>on</strong>venient to introduce an approximati<strong>on</strong> ˜Xn of X n . As explained<br />

in Secti<strong>on</strong> 2, due to the Markov property of the Brownian moti<strong>on</strong>, Π μ t := E[ f μ (B 1 )|G t ]= f (B t , t) where f (x, t) := E[ f μ (x +<br />

√<br />

1 − t · Z)] <strong>with</strong> Z ∼ N (0, 1). As a c<strong>on</strong>voluti<strong>on</strong> <strong>with</strong> a normal density, f is twice c<strong>on</strong>tinuously differentiable <strong>on</strong> R ×[0, 1[,<br />

and it further satisfies the heat equati<strong>on</strong>, so that Π μ t = f (0, 0) + ∫ t<br />

0 r s dB s ,<strong>with</strong>r s = 0fors 1 and r s = ∂ ∂x f (B s, s) for s < 1.<br />

Let us observe here that f (x, t) is increasing in x at fixed t since so is f μ (x), andthusr s 0foralls. Observe also that<br />

r s is c<strong>on</strong>tinuous <strong>on</strong> [0, 1[ and that Xq n = f (0, 0) + ∫ τq<br />

n<br />

0 r s dB s . We will then define ˜Xn by:<br />

τ<br />

∫<br />

q<br />

n<br />

˜X q n = f (0, 0) + r n s dB s, (15)<br />

0<br />

where r n := T n (r) is the image of the process r by the map T n we now define. Let H 2 be the linear space of G-progressively<br />

measurable processes a such that<br />

[ ∫∞<br />

]<br />

‖a‖ 2 :=<br />

H 2 E a 2 s ds < ∞.<br />

0<br />

Let also H 2 [0,1] denote the set of a ∈ H2 such that a s = 0, for all s 1. For a ∈ H 2 [0,1] ,wedefineT n(a) as the simple process<br />

n−1<br />

∑<br />

T n (a) t :=<br />

q=0<br />

n · E<br />

[ q+1<br />

∫n<br />

q<br />

n<br />

]<br />

a s ds|G τ<br />

n<br />

q<br />

· 1 [τ<br />

n<br />

q ,τ n q+1 [ (t).<br />

Lemma 13. T n is a linear mapping from H 2 [0,1] to H2 ,and<br />

∀a ∈ H 2 [0,1] : ∥<br />

∥Tn (a) ∥ ∥<br />

H 2 ‖a‖ H 2.


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 63<br />

Proof. As a simple process, T n (a) is progressively measurable and<br />

∥ Tn (a) ∥ [ n−1<br />

( [ q+1 ]) ∑<br />

∫n<br />

2 2 =<br />

H 2 E n · E a s ds|G τ<br />

n<br />

q<br />

· (τ n q+1 − τ ) ]<br />

q<br />

n .<br />

q=0<br />

q<br />

n<br />

Since Y n q+1 := √ n · (B τ<br />

n − B q+1<br />

τq n) satisfies [Y n q+1 |G τq n ]=[Y ], weget<br />

E [ τ n q+1 − τ n q |G τ n q<br />

Furthermore, <strong>with</strong> Jensens inequality:<br />

(<br />

E<br />

[ q+1<br />

∫n<br />

] [<br />

= E (Bτ n − ] B E[Y 2 ]<br />

q+1<br />

τq n)2 |G τ<br />

n<br />

q<br />

= = 1 n n .<br />

a s ds|G τ<br />

n<br />

q<br />

]) 2<br />

E<br />

[( q+1<br />

∫n<br />

) 2 ]<br />

a s ds |G τ<br />

n<br />

q<br />

,<br />

q<br />

n<br />

q<br />

n<br />

and by Cauchy–Schwarz inequality: ( ∫ q+1<br />

n<br />

q<br />

n<br />

∥ Tn (a) ∥ [ n−1<br />

2 <br />

H 2 E<br />

∑<br />

q=0<br />

E<br />

[ q+1<br />

∫n<br />

q<br />

n<br />

a s ds) 2 ∫ q+1<br />

n<br />

q<br />

n<br />

]]<br />

a 2 s ds|G τq<br />

n =‖a‖ 2 ,<br />

H 2<br />

a 2 s ds · 1<br />

n . Therefore<br />

and the lemma is proved.<br />

✷<br />

Lemma 14. ∀a ∈ H 2 [0,1] :lim n→∞ ‖T n (a) − a‖ H<br />

2 = 0.<br />

Proof. As it follows from the last lemma, the linear maps W n defined by W n (a) := T n (a) − a form an equi-c<strong>on</strong>tinuous<br />

sequence of linear mappings. Therefore, we just have to prove the result for elementary processes a of the form: a s :=<br />

ψ u · 1 [u,v[ , where u < v < 1 and ψ u ∈ L ∞ (G u ). Indeed, these elementary processes engender a dense subspace of H[0,1] 2 .If<br />

ψ t := E[ψ u |G t ], the process ψ is a martingale <strong>on</strong> the Brownian filtrati<strong>on</strong> and, as such, has c<strong>on</strong>tinuous sample paths. It is<br />

further uniformly integrable since ψ u ∈ L ∞ (G u ), and <strong>with</strong> the stopping theorem, we c<strong>on</strong>clude that E[ψ u |G τ<br />

n<br />

q<br />

]=ψ τ<br />

n<br />

q<br />

.Next:<br />

T n (a) = ψ τ<br />

n<br />

[[nu]] · ([[nu]]<br />

− ) nu · 1 [τ<br />

n<br />

[[nu]]<br />

,τ n [[nu]]+1 [ + ψ τ<br />

n<br />

[[nv]] · (nv<br />

−[[nv]] ) [[nv]]−1<br />

∑<br />

· 1 [τ<br />

n<br />

[[nv]]<br />

,τ n [[nv]]+1 [ + ψ τ<br />

n<br />

q · 1 [τ<br />

n<br />

q ,τ n q+1 [ .<br />

q=[[nu]]<br />

The H 2 norm of the two first terms goes to 0 <strong>with</strong> n since ‖ψ‖ L ∞ < ∞ and E[τ n q+1 − τ q n]=1/n.<br />

For all positive numbers x, x ′ , y, y ′ such that x y and x ′ y ′ ∫ ∞<br />

, it is easy to check that 0 (1 [x,y[ − 1 [x ′ ,y ′ [) 2 dt =|x − x ′ |+<br />

|y − y ′ | if [x, y[∩[x ′ , y ′ ∫ ∞<br />

[≠∅.Otherwise, 0 (1 [x,y[ − 1 [x ′ ,y ′ [) 2 dt = y − x + y ′ − x ′ .So,<strong>with</strong>D n := {τ[[nu]] n v}∪{τ [[nv]] n u},<br />

we get: ‖a − ψ u · 1 [τ<br />

n<br />

[[nu]]<br />

,τ[[nv]] n [ ‖ 2 =‖ψ<br />

H 2 u · 1 [u,v[ − ψ u · 1 [τ<br />

n<br />

[[nu]]<br />

,τ[[nv]] n [ ‖ 2 which is also equal to E[1<br />

H 2<br />

D<br />

c<br />

n<br />

ψu 2(|u − τ [[nu]] n |+|v −<br />

τ[[nv]] n |)] +E[1 D n<br />

ψu 2(τ [[nv]] n − τ [[nu]] n + v − u)]. The first expectati<strong>on</strong> c<strong>on</strong>verges to 0 as n increases, since ‖ψ u‖ L ∞ < ∞ and<br />

both ‖v − τ[[nv]] n ‖ L 1 and ‖u − τ [[nu]] n ‖ L 1 c<strong>on</strong>verge to 0, according to Lemma 12. The same lemma indicates that ‖τ [[nt]] n ‖ L 2 <br />

1 + γ 2 and ψu 2(τ [[nv]] n − τ [[nu]] n + v − u) is thus bounded in L2 . With Cauchy–Schwarz inequality, we c<strong>on</strong>clude then that<br />

E[1 Dn ψu 2(τ [[nv]] n − τ [[nu]] n + v − u)] c<strong>on</strong>verges to 0 as n →∞,sincep(D n) = p(τ[[nu]] n v) + p(τ [[nv]] n u) also c<strong>on</strong>verges to 0:<br />

τ[[nu]] n c<strong>on</strong>verges in L1 to u < v, and τ[[nv]] n to v > u. Therefore ‖a − ψ u · 1 [τ<br />

n<br />

[[nu]]<br />

,τ[[nv]] n [ ‖ 2 c<strong>on</strong>verges to 0 and, to prove the<br />

H 2<br />

lemma, it just remains to prove that η n c<strong>on</strong>verges to 0, where<br />

η n :=<br />

∥ ψ u · 1 [τ<br />

n<br />

[[nu]]<br />

,τ[[nv]] n [ −<br />

[[nv]]−1<br />

∑<br />

q=[[nu]]<br />

ψ τ<br />

n<br />

q · 1 [τ<br />

n<br />

q ,τ n q+1 [ ∥ ∥∥∥∥<br />

2<br />

H 2 .<br />

Now<br />

η n =<br />

∥<br />

∑<br />

[[nv]]−1<br />

q=[[nu]]<br />

(ψ u − ψ τ<br />

n<br />

q<br />

) · 1 [τ<br />

n<br />

q ,τ n q+1 [ ∥ ∥∥∥∥<br />

2<br />

[[nv]]−1<br />

∑<br />

=<br />

H 2 q=[[nu]]<br />

[ E (ψ u − ψ τ<br />

n<br />

q<br />

) 2 · (τ n q+1 − τ )]<br />

q<br />

n .


64 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

It results from the definiti<strong>on</strong> of ψ t that ψ t = ψ u if t u. Therefore, we infer that: (ψ u − ψ τ<br />

n<br />

q<br />

) 2 4‖ψ u ‖ 2 ∞ 1 τ n q


Claim (2) follows then from claim (1).<br />

B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 65<br />

= √ √<br />

n · A ·<br />

[∣ ∣<br />

E ∣ X<br />

n<br />

n − ˜Xn n 2]<br />

.<br />

We will next compute V M n (Fn , ˜Xn ). Defining λ n q as: λn q<br />

✷<br />

q+1<br />

n<br />

:= n · E[∫ q<br />

n<br />

r s ds|G τ<br />

n<br />

q<br />

],wehaver n t := ∑ n−1<br />

q=0 λn q · 1 [τ n q ,τ n q+1 [ (t). Since<br />

r is a positive process, we clearly have λ n q 0. Next, ˜Xn<br />

q+1 − ˜Xn q = λ n q · (B τ n − B q+1<br />

τq n) = an q · Y n q+1 , where an q := λn q<br />

√ . a n n q is thus<br />

positive and Fq n -measurable. Therefore, since M[X] is 1-homogeneous in X according to assumpti<strong>on</strong> (i) in Theorem 5, since<br />

[Y n q+1 |F q]=[Y ], and since E[Y 2 ]=1, E[Y ]=0, we get:<br />

V M n<br />

(<br />

F n , ) [ ∑n−1<br />

˜Xn = [ E M ˜Xn<br />

q+1 − ] ]<br />

˜Xn q |Fq<br />

n<br />

q=0<br />

[ ∑n−1<br />

= [ E M a n q · ] ]<br />

Y n<br />

q+1 |Fn q<br />

q=0<br />

[ n−1<br />

]<br />

∑<br />

= E a n q · M[Y ]<br />

q=0<br />

= M[Y ]·E<br />

= M[Y ]·E<br />

[ ∑n−1<br />

]<br />

a n q · (Y ) n 2<br />

q+1<br />

q=0<br />

[( ∑n−1<br />

) ( n−1<br />

)] ∑<br />

a n q · Y n q+1<br />

· Y n q+1<br />

q=0<br />

q=0<br />

Since E[B 2 τn<br />

n ]=1, we also have<br />

V M n (Fn , ˜Xn )<br />

M[Y ]·√n<br />

= √ n · M[Y ]·E [ ˜Xn n · B τ<br />

n<br />

n<br />

]<br />

.<br />

E[ Xn n · ] ∥ B τn<br />

n − ˜Xn n − Xn∥<br />

n ∥<br />

L 2<br />

= [ E X n n+1 · ] ∥ B τn<br />

n − ˜Xn n − Xn∥<br />

n ∥<br />

L 2<br />

[ E X n n+1 · ] ∥ B 1 − ∥X n ∥<br />

n+1 L 2 ·‖B τ<br />

n<br />

n<br />

− B 1 ‖ L 2 − ∥ ∥<br />

˜Xn n − Xn<br />

n<br />

= [ ]<br />

E f μ (B 1 ) · B 1 −‖μ‖L 2 ·‖B τ<br />

n<br />

n<br />

− B 1 ‖ L 2 − ∥ ˜Xn n − Xn<br />

n ∥<br />

L 2<br />

= α(μ) −‖μ‖ L 2 ·‖B τ<br />

n<br />

n<br />

− B 1 ‖ L 2 − ∥ ˜Xn n − Xn<br />

n ∥<br />

L 2.<br />

With claim (4) in Lemma 12 and claim (1) in Lemma 15, we c<strong>on</strong>clude then that:<br />

lim inf<br />

n→∞<br />

V M n (Fn , X n )<br />

√ n<br />

= lim inf<br />

n→∞<br />

V M n (Fn , ˜Xn )<br />

√ n<br />

M[Y ]·α(μ).<br />

Since V M n (μ) V M n (Fn , X n ), we thus have proved that for all Y ∈ L 4 <strong>with</strong> E[Y ]=0andE[Y 2 ]=1:<br />

lim inf<br />

n→∞<br />

V M n (μ) √ n<br />

M[Y ]·α(μ).<br />

Since ˜D := {Ỹ ∈ L 4 : E[Ỹ ]=0 and E[Ỹ 2 ] 1} is dense for the L 2 -norm in D := {Ỹ ∈ L 2 : E[Ỹ ]=0 and E[Ỹ 2 ] 1}, and since<br />

M is c<strong>on</strong>tinuous for the L p -norm and thus for the L 2 -norm, we infer that there exists a sequence {Ỹ n } n∈N ⊂ ˜D such that<br />

lim<br />

n→∞ M[Ỹ n ]=ρ := { sup M[Ỹ ]: Ỹ ∈ D } > 0.<br />

We may further assume that M[Ỹ n ] > 0, so that, since M is 1-homogeneous, we have that M[Y n ] M[Ỹ n ], where Y n =<br />

Ỹ n<br />

‖Ỹ n ‖ L 2<br />

.SinceY n ∈ L 4 satisfies E[Y n ]=0 and E[Yn 2 ]=1, we thus have proved that<br />

lim inf<br />

n→∞<br />

V M n (μ) √ n<br />

With Theorem 11, we get then<br />

lim M[Y ]·α(μ) = ρ · α(μ).<br />

n→∞<br />

∥<br />

L 2


66 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

Theorem 16. Under the hypotheses of Theorem 5,<br />

Vn M lim √ (μ) = ρ · α(μ).<br />

n→∞ n<br />

The first part of Theorem 5 is thus proved. The sec<strong>on</strong>d part will be proved in the next secti<strong>on</strong>.<br />

15. C<strong>on</strong>vergence to the c<strong>on</strong>tinuous martingale of maximal variati<strong>on</strong><br />

Let B be a standard <strong>on</strong>e-dimensi<strong>on</strong>al Brownian moti<strong>on</strong> <strong>on</strong> a filtered probability space (Ω, A, P,(G t ) t0 ).Ifμ ∈ 1+ ,<br />

the martingale Π μ t<br />

:= E[ f μ (B 1 )|G t ] is referred to in this paper as the c<strong>on</strong>tinuous martingales of maximal variati<strong>on</strong> of final<br />

distributi<strong>on</strong> μ. This terminology is justified by the next result that clearly implies the sec<strong>on</strong>d part of Theorem 5.<br />

If (F, X) ∈ W n (μ), we define the c<strong>on</strong>tinuous time representati<strong>on</strong> ˜X of X as the process ( ˜Xt ) t∈[0,1] <strong>with</strong> ˜Xt := X [[nt]] ,<br />

where [[a]] is the greatest integer less or equal to a.<br />

Theorem 17. Assume that M satisfies the hypotheses (i) and (ii) of Theorem 5, thatρ > 0, thatμ ∈ 2 and that {(F n , X n )} n∈N is a<br />

sequence of martingales <strong>with</strong> for all n (F n , X n ) ∈ M n (μ), that asymptotically maximizes the M-variati<strong>on</strong>, i.e.:<br />

Vn M lim<br />

(Fn , X n )<br />

√ = ρ · α(μ).<br />

n→∞ n<br />

Then ˜Xn c<strong>on</strong>verges in finite-dimensi<strong>on</strong>al distributi<strong>on</strong> to Π μ : For all finite set J ⊂[0, 1], ( ˜Xn t<br />

) t∈ J c<strong>on</strong>verges in law to (Π μ t<br />

) t∈ J .<br />

Proof. Let {(F n , X n )} n∈N be an asymptotically maximizing sequence. Without loss of generality, we may assume that F n<br />

c<strong>on</strong>tains an adapted system (U q ) q=0,...,n of independent uniform random variables, independent of X n (otherwise F n could<br />

be widened). Therefore, <strong>with</strong> (11), there exists S n ∈ S(ρ,4A+ρ) ∗ (Fn ) such that Vn M(Fn , X n ) − 1 E[X n n+1 · Sn n ], and thus<br />

E[X n n+1 · lim<br />

Sn n ]<br />

n→∞ ρ · √n = α(μ).<br />

As in Secti<strong>on</strong> 12, for ɛ n > 0 to be determined later, we may embed (X n , R n ) in the Brownian filtrati<strong>on</strong> G, where R n :=<br />

obtaining thus an increasing sequence (τq n) q=0,...,n+1 and a pair ( ˆXn , ˆRn ) of ˆF n martingales, where ˆF q n := G τq<br />

n such that<br />

(X n , R n ) and ( ˆXn , ˆRn ) are equally distributed. We then have<br />

[ ] [<br />

E ˆXn<br />

n+1 B 1 E ˆXn ˆR n n+1 n]<br />

−‖μ‖L 2 · ∥∥ ∥<br />

B1 − ˆRn ∥L n 2.<br />

Since E[ ˆXn ˆR n n+1 n ]=E[Xn n+1 Rn n ], the first term in the right-hand side of the last inequality c<strong>on</strong>verges to α(μ). Next, according<br />

to (12) and claim (3) in Lemma 10 <strong>with</strong> t = 1:<br />

∥<br />

∥ B1 − ˆRn ∥L n 2 ‖B 1 − B τ<br />

n<br />

n<br />

‖ L 2 + ∥ ∥ Bτ n<br />

n<br />

− ˆRn ∥L n 2<br />

κ · n 1<br />

p ′ ∧4 − 1 2<br />

+ 2 √ ɛ n n.<br />

So if ɛ n is chosen so as to ensure nɛ n → 0asn →∞,weget<br />

E[ ˆXn<br />

n+1 · lim<br />

B 1]<br />

n→∞ ρ · √n = α(μ).<br />

Since B 1 ∼ N (0, 1) and ˆXn<br />

n+1 ∼ μ, we may then apply claim (2) in Theorem 6 to infer that ˆXn<br />

n+1 c<strong>on</strong>verges in L1 -norm to<br />

f μ (B 1 ) = Π μ 1 .<br />

Next observe that ‖ ˆXn [[nt]]<br />

− Π μ t<br />

‖ L 1 ‖ ˆXn [[nt]]<br />

− Π μ τ<br />

‖<br />

[[nt]]<br />

n L 1 +‖Π μ τ<br />

− Π μ<br />

[[nt]] n t<br />

‖ L 1 .But<br />

∥ ˆXn [[nt]] − Π μ ∥<br />

τ[[nt]]<br />

n<br />

∥<br />

L 1 = ∥ ∥ E<br />

[<br />

ˆXn<br />

n+1 − Πμ 1 |G τ n [[nt]]<br />

]∥<br />

∥L<br />

1 ∥ ˆXn<br />

n+1 − Πμ ∥<br />

1 L 1.<br />

On the other hand, <strong>with</strong> claims (1) and (2) in Lemma 10 and our choice of ɛ n we infer that τ n [[nt]] → t in L1 .SinceΠ μ is<br />

uniformly integrable and, as a martingale <strong>on</strong> the Brownian filtrati<strong>on</strong>, it has c<strong>on</strong>tinuous sample paths, we then c<strong>on</strong>clude that<br />

‖Π μ τ n [[nt]]<br />

− Π μ t ‖ L 1 → 0asn increases. Therefore ˆXn [[nt]] c<strong>on</strong>verges to Π μ t in L 1 .<br />

This implies in particular the c<strong>on</strong>vergence in finite distributi<strong>on</strong> of the process ( ˆXn [[nt]]<br />

) t0 to Π μ , and this process has<br />

same distributi<strong>on</strong> as ˜Xn . ✷<br />

ρ·√n Sn ,


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 67<br />

16. C<strong>on</strong>clusi<strong>on</strong><br />

We c<strong>on</strong>clude this paper <strong>with</strong> a list of open problems and some possible extensi<strong>on</strong>s of the model.<br />

(1) It is assumed in the descripti<strong>on</strong> of Γ n that acti<strong>on</strong>s are observed at each round. This is however an unrealistic hypothesis<br />

in the case of a <strong>market</strong> game, where the complete individual demand functi<strong>on</strong>s remain typically private. Only a sample<br />

of this demand functi<strong>on</strong> will be revealed during the tât<strong>on</strong>nement process leading to the <strong>market</strong> clearing price. What<br />

is certainly observed by P2 in an exchange game is the trade. This raises the questi<strong>on</strong> whether similar <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> will<br />

appear in a game where <strong>on</strong>ly transfers are observed at each round. This however is a game <strong>with</strong> imperfect m<strong>on</strong>itoring<br />

and is thus more difficult to analyze.<br />

(2) It is quite reductive to represent the <strong>market</strong> by a two player game. It is interesting to note at this respect that, in Γ n ,<br />

P1 has no need to observe P2’s acti<strong>on</strong>s to play optimally, and thus, at each stage, P2 is essentially maximizing his stage<br />

payoff, since his acti<strong>on</strong> will not influence P1’s future behavior. Therefore, we could replace P2 by a successi<strong>on</strong> of players,<br />

<strong>on</strong>e per stage, playing against a single informed P1, and we would obtain the same equilibria.<br />

Our result relies crucially <strong>on</strong> the min max approach and the noti<strong>on</strong> of value that characterizes zero-sum games. Dealing<br />

<strong>with</strong> more general model where N-players interact at each round is more difficult as the noti<strong>on</strong> of optimal strategies<br />

has to be replaced by that of Nash equilibrium which has much less structure. There could in particular exist multiple<br />

equilibria <strong>with</strong> different payoffs.<br />

(3) One criticism of the model c<strong>on</strong>cerns hypothesis (H5) and the fact that it implies that P2 is forced to trade (see the<br />

discussi<strong>on</strong> of (H5) is Secti<strong>on</strong> 5). The <strong>on</strong>ly way to avoid this no trade paradox is to c<strong>on</strong>sider risk proclivity for P2. This<br />

however turns also to be a n<strong>on</strong>-zero sum game. Instead of c<strong>on</strong>sidering <strong>on</strong>e single risk-seeking player 2, we are currently<br />

analyzing a game where P1 faces a successi<strong>on</strong> of risk-seeking P2s and we expect the same price <str<strong>on</strong>g>dynamics</str<strong>on</strong>g> to appear.<br />

Indeed, in this setting, the <strong>on</strong>e shot game has a single equilibrium for all μ and P1’s payoff at equilibrium is also<br />

afuncti<strong>on</strong>M of the law of L q+1 − L q . P1 thus maximizes the M variati<strong>on</strong>. However, due to the risk proclivity, this<br />

functi<strong>on</strong> M is not 1-homogeneous any more, but it is locally 1-homogeneous around 0. Apparently this is enough to get<br />

our asymptotic results: Martingales <strong>with</strong> maximal variati<strong>on</strong> c<strong>on</strong>verge to c<strong>on</strong>tinuous processes: as n increases, the size<br />

of the increments goes to 0 and <strong>on</strong>ly the local behavior of M around zero seems to matter.<br />

(4) As menti<strong>on</strong>ed above, analyzing n<strong>on</strong>-zero sum games is much more difficult. As a first step in that directi<strong>on</strong>, we are<br />

currently analyzing the <strong>market</strong> maker game <strong>with</strong> arbitrageur model menti<strong>on</strong>ed is Secti<strong>on</strong> 6.1. This game is n<strong>on</strong>-zero<br />

sum and we find in this model a sequence of equilibria in which the price process c<strong>on</strong>verges to c<strong>on</strong>tinuous martingales<br />

close to the CMMV class.<br />

(5) Since CMMV is a quite robust class of <str<strong>on</strong>g>dynamics</str<strong>on</strong>g>, it seems natural to use it in financial ec<strong>on</strong>ometrics. As a particular<br />

local volatility model, it could be used to price derivatives, taking into account the volatility smile, as Dupire’s (1997)<br />

method. We are currently working <strong>on</strong> a pricing method <strong>with</strong> volatility smile using CMMV. As compared to Dupire’s <strong>on</strong>e,<br />

less informati<strong>on</strong> is needed <strong>on</strong> the volatility manifold to calibrate the model accurately.<br />

Appendix A<br />

In this appendix we aim to prove that (H1 ′ ) joint (H2) implies (H1- ∞ ) as stated in Theorem 24. The trading mechanism<br />

c<strong>on</strong>sidered in this secti<strong>on</strong> is thus assumed to satisfy both (H1 ′ ) and (H2). We are c<strong>on</strong>cerned in this secti<strong>on</strong> <strong>with</strong> measures<br />

in ∞ having thus a compact support. Let K =[K, K ] be a compact interval. All measures μ c<strong>on</strong>sidered in this secti<strong>on</strong> are<br />

in (K ), the set of probability distributi<strong>on</strong> over K . In Lemma 2, we proved that P1 can guarantee V n (μ) in Γ n (μ). Wewill<br />

prove that player 2 can guarantee the same amount, proving thus that V n (μ) is the value of Γ n (μ), <strong>with</strong>out assuming that<br />

this value exists. With the next lemma, we analyze the c<strong>on</strong>tinuity of V n (μ).<br />

Lemma 18.<br />

(1) If V 1 satisfies (H2) in L p -norm <strong>with</strong> the Lipschitz c<strong>on</strong>stant A, then for all random variables L 1 , L 2 , for all n:<br />

∣ Vn<br />

(<br />

[L1 ] ) − V n<br />

(<br />

[L2 ] )∣ ∣ nA‖L1 − L 2 ‖ L p .<br />

(2) In particular V n is c<strong>on</strong>tinuous for the weak topology <strong>on</strong> (K ).<br />

(3) V n (μ) is further c<strong>on</strong>cave in μ.<br />

(4) Let Φ n denote the set of c<strong>on</strong>tinuous functi<strong>on</strong>s φ <strong>on</strong> K satisfying for all ν ∈ (K ), E ν [φ(L)] V n (ν), then<br />

∀μ ∈ (K):<br />

V n (μ) = inf<br />

φ∈Φ n<br />

E μ<br />

[<br />

φ(L)<br />

]<br />

.<br />

(5) ∀ɛ > 0, there exists a finite subset Φ ′ ∈ Φ n such that<br />

∀μ ∈ (K):<br />

[ ]<br />

min E φ∈Φ ′ μ φ(L) Vn (μ) + ɛ.


68 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

Proof. (1) Let (L 1 , L 2 ) be a random vector <strong>with</strong> marginals μ 1 , μ 2 in (K ). Let(F, X) ∈ W n (μ 1 ) be such that V n (F, X) <br />

V n (μ 1 ) − ɛ. LetthenY n+1 be a random variable <strong>on</strong> the same probability space as X and such that the random vectors<br />

(X n+1 , Y n+1 ) and (L 1 , L 2 ) are equally distributed. Set then Y q := E[Y n+1 |F q ].Then(F, Y ) ∈ W n (μ 2 ). It follows from Jensen’s<br />

inequality:<br />

[∣<br />

E ∣ V 1 [X q+1 |F q ]−V 1 [Y q+1 |F q ] ∣ ] A · [( [∣ ∣<br />

E E ∣ Xq+1 − Y q+1 p ]) 1 ]<br />

|Fq<br />

p<br />

A · (E [ E [ |X q+1 − Y q+1 | p |F q<br />

]]) 1<br />

p<br />

= A ·‖X q+1 − Y q+1 ‖ L p<br />

A ·‖X n+1 − Y n+1 ‖ L p<br />

= A ·‖L 1 − L 2 ‖ L p .<br />

Therefore |V n (F, X) − V n (F, Y )| nA‖Y n+1 − X n+1 ‖ L p , and thus V n (μ 1 ) − ɛ V n (F, X) V n (F, Y ) + nA‖L 1 − L 2 ‖ L p <br />

V n (μ 2 ) + nA‖L 1 − L 2 ‖ L p .Sinceɛ > 0isarbitrary,wegetV n ([L 1 ]) − V n ([L 2 ]) nA‖L 1 − L 2 ‖ L p . Interchanging L 1 and L 2 ,we<br />

get claim (1).<br />

(2) Next, let μ m be weakly c<strong>on</strong>vergent in (K ) to μ. According to Skorokhod’s representati<strong>on</strong> theorem, there exists a<br />

sequence X m of μ m -distributed random variables that c<strong>on</strong>verges a.s. to a μ-distributed limit X. Since all variables are K -<br />

valued, we c<strong>on</strong>clude <strong>with</strong> Lebesgue dominated c<strong>on</strong>vergence theorem that X m c<strong>on</strong>verges to X in L p -norm. Claim (1) implies<br />

then that V n (μ m ) = V n ([X m ]) c<strong>on</strong>verges to V n ([X]) = V n (μ), and V n is thus weakly c<strong>on</strong>tinuous as announced.<br />

(3) If μ = λ 1 μ 1 + λ 2 μ 2 ,<strong>with</strong>λ i 0, λ 1 + λ 2 = 1, let, for ɛ > 0, (F i , X i ) ∈ W n (μ i ) be such that V n (F i , X i ) V n (μ i ) − ɛ.<br />

Assume that X 1 and X 2 are <strong>on</strong> two independent probability spaces. On the product space, we can then c<strong>on</strong>sider X :=<br />

1 A X 1 + (1 − 1 A )X 2 , where A is an event of probability λ 1 independent of X 1 , X 2 .ThensetF q := σ (A, Fq 1, F q 2 ). It follows<br />

that (F, X) ∈ W n (μ). Therefore V n (μ) V n (F, X) = ∑ i λ iV n (F i , X i ) ∑ i λ iV n (μ i ) − ɛ. Letting ɛ go to 0, we get the<br />

announced c<strong>on</strong>cavity.<br />

(4) This claim follows at <strong>on</strong>ce from the fact that a c<strong>on</strong>tinuous c<strong>on</strong>cave functi<strong>on</strong> f <strong>on</strong> a Banach space is the infimum of<br />

the set of c<strong>on</strong>tinuous linear functi<strong>on</strong>als that dominate f . In this case, V n is a functi<strong>on</strong> <strong>on</strong> the closed subset (K ) of the<br />

Banach space of bounded measures <strong>on</strong> K and C(K ) is the dual of this space.<br />

(5) For ɛ > 0 and φ ∈ Φ n ,defineC φ as {μ ∈ (K )|E μ [φ(L)]−V n (μ) 0, let Φ ′ as in claim (5) of Lemma 18. Let φ 1 ,...,φ R be<br />

an enumerati<strong>on</strong> of Φ ′ . Then, if r ∗ (ω) denotes the smallest r that minimizes E[φ r (X)|F](ω), r ∗ (ω) is F measurable and we<br />

clearly get <strong>with</strong> φ ω := φ r∗ (ω) : E[φ ω (X(ω))]=E[min r {E[φ r (X)|F]}] E[V n ([X|F])]+ɛ. Letting ɛ go to 0, we get claim (1).<br />

(2) Let F i the set corresp<strong>on</strong>ding to F i in claim (1). Then F 1 ⊂ F 2 and thus inf φ. ∈F 2<br />

E[φ ω (X(ω))] inf φ. ∈F 1<br />

E[φ ω (X(ω))],<br />

and the result follows. ✷<br />

Throughout this paper, V n (μ), as defined in (6), has been c<strong>on</strong>sidered as a problem of maximizati<strong>on</strong> of V n (F, X) over a<br />

martingale space W n (μ). The next lemma will allow us to view V n (μ) as a maximizati<strong>on</strong> problem over a measure space.<br />

Let n<br />

mart be the subset of ρ ∈ (K n+1 ) such that if (L 1 ,...,L n+1 ) is ρ-distributed then ∀q: E ρ [L q+1 |L q ]=L q , where L q<br />

is a notati<strong>on</strong> for (L 1 ,...,L q ).Thesetofρ ∈ (K n+1 ) that further satisfy L n+1 ∼ μ will be denoted n<br />

mart (μ). This last set<br />

is thus the set of laws of martingales in W n (μ). Clearly,ρ ∈ (K n+1 ) bel<strong>on</strong>gs to n<br />

mart if and <strong>on</strong>ly if for all c<strong>on</strong>tinuous<br />

functi<strong>on</strong> f : E ρ [ f (L q )(L q+1 − L q )]=0, and it bel<strong>on</strong>gs to n<br />

mart (μ) if furthermore E ρ [ f (L n+1 )]=E μ [ f (L)]. Since all these<br />

c<strong>on</strong>diti<strong>on</strong>s are linear c<strong>on</strong>tinuous in ρ, n<br />

mart and n mart (μ) are closed c<strong>on</strong>vex subsets of the weakly compact space (K n+1 ).<br />

Lemma 20.<br />

(1) ∀(F, X) ∈ W n (μ): V n (X, F) ∑ n<br />

q=1 E[V 1[X q |X


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 69<br />

(3) ∀μ ∈ (K ): V n (μ) = sup ρ∈ mart<br />

1 (μ) V 1([L 1 ] ρ ) + E ρ [V n−1 ([L 2 |L 1 ] ρ )].<br />

(4) The map ρ → E ρ [V n ([L 2 |L 1 ] ρ )] is c<strong>on</strong>cave in ρ and weakly upper semic<strong>on</strong>tinuous <strong>on</strong> mart<br />

1 .<br />

Proof. (1) For (F, X) ∈ W n (μ), setF ′ q := σ (X q). SinceX q is F q -measurable, we get F ′ q ⊂ F q. Therefore, <strong>with</strong> claim (2)<br />

Lemma 19,<br />

n∑<br />

V n (X, F) =<br />

[ E V 1 [X q |F q−1 ] ] n∑<br />

<br />

[ [<br />

E V 1 Xq |F q−1]] ′ ,<br />

q=1<br />

q=1<br />

and claim (1) is proved.<br />

(2) If ρ denotes the law of X, then ρ ∈ n<br />

mart (μ) and the coordinate process (L n+1 ) <strong>on</strong> the probability space<br />

(K n+1 , B K n+1, ρ) has the same law as X. IfG q := σ (L q ), we get thus ∑ n<br />

q=1 E[V 1[X q |F ′ q−1 ]] = ∑ n<br />

q=1 E ρ[V 1 [L q |G q−1 ]] =<br />

∑<br />

V n (G, L) V n (μ), and thus V n (μ) = n<br />

sup ρ∈ mart<br />

n (μ) q=1 E ρ[V 1 [L q |L 1 is a martingale of length n, <strong>with</strong> final distributi<strong>on</strong> [L n+1 |L 1 ]. Therefore, for all<br />

ρ : E ρ [ ∑ n<br />

q=2 V 1[L q |L q ]|L 1 ] V n−1 ([L n+1 |L 1 ]). Thus,V n (μ) sup ρ∈ mart<br />

n (μ) V 1([L 1 ] ρ ) + E ρ [V n−1 ([L n+1 |L 1 ])]. C<strong>on</strong>versely,<br />

let ρ be ɛ-optimal in the right-hand side of this formula. Let then ρ L1 denote the law of an ɛ-optimal martingale in this<br />

in V n−1 ([L n+1 |L 1 ]), then selecting L 1 <strong>with</strong> ρ and L >1 <strong>with</strong> ρ L1 gives a martingale that satisfies ∑ n<br />

q=1 E[V 1[L q |L


70 B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71<br />

compact, we c<strong>on</strong>clude <strong>with</strong> Propositi<strong>on</strong> 1.8 in Mertens et al. (1994) that inf and sup comute in the previous formula and<br />

thus:<br />

[<br />

max E μ φτ (L) − φ(L) ] .<br />

μ∈(K )<br />

0 inf<br />

τ ∈T K<br />

If τ in and ɛ-optimal strategy in this inf sup, we get ∀μ: ɛ E μ [φ τ (L) − φ(L)], and in particular ∀L: ɛ φ τ (L) − φ(L).<br />

We thus have proved that ∀φ ∈ Φ 1 , ∀ɛ > 0, there exists τ ∈ T K such that φ + ɛ φ τ .<br />

Let E be a countable dense subset in the separable space (C(K ), ‖.‖ ∞ ).Forall f ∈ E ∩ Φ 1 , ∀n ∈ N, letτ f ,n ∈ T K be such<br />

that f + 1/n φ τ f ,n<br />

. The countable set T 1 of {τ f ,n | f ∈ E ∩ Φ 1 ,n ∈ N} will have the required property. Indeed, if φ ∈ Φ 1 ,for<br />

all ɛ > 0, there exists f ∈ E such that φ + ɛ/2 f φ. Inparticular f ∈ Φ 1 .Ifn 2/ɛ, thenφ τ f ,n<br />

f + 1/n φ + ɛ. ✷<br />

We next prove recursively a similar property for Γ n :<br />

Lemma 23. There exists a countable set T n of P2’s strategies in Γ n such that, if φ ∈ Φ n then, for all ɛ > 0, there exists a strategy τ in<br />

T n such that ∀L ∈ K : φ(L) + ɛ φτ n(L),whereφn τ was defined in (2).<br />

Proof. According to the previous lemma, the result holds for n = 1. Assume next it holds for n − 1. We argue that it will<br />

also hold for n. Indeed,φ ∈ Φ n implies that ∀μ ∈ (K ): E μ [φ(L)] V n (μ). Therefore 0 sup μ V n (μ) − E μ [φ(L)]. According<br />

to claim (3) in Lemma 20, we get thus<br />

0 sup<br />

μ<br />

sup<br />

ρ∈ mart<br />

1<br />

(μ)<br />

V 1<br />

(<br />

[L1 ] ρ<br />

)<br />

+ Eρ<br />

[<br />

Vn−1<br />

(<br />

[L2 |L 1 ] ρ<br />

)]<br />

− Eρ<br />

[<br />

φ(L2 ) ] .<br />

Since mart<br />

1<br />

= ⋃ μ mart 1 (μ), we get <strong>with</strong> Lemma 21:<br />

0 sup<br />

ρ∈ mart<br />

1<br />

[<br />

inf E ρ φτ (L 1 ) ] [ ( )] [<br />

+ E ρ Vn−1 [L2 |L 1 ] ρ − Eρ φ(L2 ) ] .<br />

τ ∈T K<br />

The payoff in this sup inf is finite since all the φ τ and φ are c<strong>on</strong>tinuous <strong>on</strong> K and thus bounded. It is further c<strong>on</strong>cave<br />

weakly u.s.c. in ρ as it results from claim (4) in Lemma 20. mart<br />

1 is weakly compact as a closed subset of (K 2 ).Onthe<br />

other hand the map τ → E ρ [φ τ (L 1 )] is c<strong>on</strong>vex. We c<strong>on</strong>clude <strong>with</strong> Propositi<strong>on</strong> 1.8 in Mertens et al. (1994) that inf and sup<br />

comute in the previous formula and thus: 0 inf τ ∈TK sup ρ∈ mart E<br />

1 ρ [φ τ (L 1 )]+E ρ [V n−1 ([L 2 |L 1 ] ρ )]−E ρ [φ(L 2 )].<br />

Let then τ ∗ ∈ T K be an ɛ/2-optimal strategy in this inf sup. We get thus for all ρ ∈ mart<br />

1 : ɛ/2 E ρ [φ τ ∗(L 1 )]+<br />

E ρ [V n−1 ([L 2 |L 1 ] ρ )]−E ρ [φ(L 2 )]. Forallμ ∈ (K ), ifρ is the law of a vector (L 1 , L 2 ) <strong>with</strong> L 2 ∼ μ and L 1 = E μ [L 2 ],then<br />

ρ ∈ mart<br />

1 and the last formula yields: ɛ/2 φ τ ∗(E μ [L 2 ])] +V n−1 (μ) − φ(E μ [L 2 ]) and in particular, as it results from<br />

the definiti<strong>on</strong> of φ τ ∗ , for all acti<strong>on</strong> i of P1: ɛ/2 A iτ ∗ E μ [L 2 ]+B iτ ∗ + V n−1 (μ) − φ(E μ [L 2 ]). In other words, the functi<strong>on</strong><br />

φ i (L) := ɛ/2 + φ(L) − A i,τ ∗ L − B iτ ∗ satisfies ∀μ ∈ (K ): E μ [φ i (L)] V n−1 (μ) and thus bel<strong>on</strong>gs to Φ n−1 . According to our<br />

hypothesis that the lemma holds for n − 1, there must be a strategy τ in T n−1 such that ɛ/2 + φ i φτ<br />

n−1 .Letτ (i) denote<br />

the first such τ in a given enumerati<strong>on</strong> of T n−1 . The functi<strong>on</strong> τ (i) will then be measurable in i.<br />

C<strong>on</strong>sider then the following strategy τ of P2 in Γ n : at the first stage he plays τ 1 := τ ∗ and starting from the sec<strong>on</strong>d<br />

stage <strong>on</strong>, he plays according to τ (i 1 ) if P1’s first move was i 1 .<br />

For this strategy,<br />

( n∑<br />

) ( n∑<br />

)<br />

φτ n (L) = sup A i1 τ ∗ L + B i 1 τ ∗ + L A iq ,τ q−1 (i 1 ) + B iq ,τ q−1 (i 1 )<br />

i 1 ,...i n q=2<br />

q=2<br />

= sup A i1 τ ∗ L + B i 1 τ ∗ + φn−1 τ (i 1 ) (L)<br />

i 1<br />

sup A i1 τ ∗ L + B i 1 τ ∗ + φ i 1<br />

(L) + ɛ/2<br />

i 1<br />

= φ τ ∗(L) + ɛ.<br />

We thus have proved that for all φ ∈ Φ n , ∀ɛ > 0, there exists a strategy τ such that ɛ + φ φ n τ . The c<strong>on</strong>structi<strong>on</strong> of T n<br />

is then similar to that of T 1 in the previous lemma. ✷<br />

Theorem 24.<br />

(1) If the trading mechanism satisfies (H1 ′ ) and (H2) then for all μ ∈ (∞), V n (μ) is the value of Γ n (μ). The mechanism satisfies<br />

thus to (H1- ∞ ).<br />

(2) If the mechanism further satisfies to (H2 ′ ), then the above asserti<strong>on</strong> holds for all μ ∈ 2 ,and(H1- 2 ) is satisfied.


B. De Meyer / Games and Ec<strong>on</strong>omic Behavior 69 (2010) 42–71 71<br />

Proof. (1) Any μ ∈ ∞ has a compact support, and the previous results apply. In particular, <strong>with</strong> claim (4) in Lemma 18,<br />

for all ɛ > 0, there exists φ ∈ Φ n such that ɛ + V n (μ) E μ [φ]. According to Lemma 23, P2 has a strategy τ in Γ n (μ) such<br />

that φ + ɛ φτ n . The maximal amount P1 can get if P2 uses this strategy is E μ[φτ n(L)] E μ[φ(L)]+ɛ V n (μ) + 2ɛ. Since<br />

ɛ > 0 is arbitrary, we c<strong>on</strong>clude that P2 can guarantee V n (μ) in Γ n (μ). The fact that P1 can guarantee the same amount was<br />

proved in Lemma 2.<br />

(2) If ∀i, j: |A i, j | A, asstatedin(H2 ′ ), then for all admissible strategy τ in Γ n , the functi<strong>on</strong> φτ n (L) defined in (2) will<br />

be Lipschitz in L <strong>with</strong> c<strong>on</strong>stant nA.Letthenμ ∈ 2 and let L be a μ-distributed random variable. There exists a sequence<br />

L m of random variables in L ∞ that c<strong>on</strong>verges to L in L 1 .Letthenτ m be a 1/m-optimal strategy of P2 in Γ n ([Y m ]). Then,<br />

since E[φτ n m<br />

(L)] E[φτ n m<br />

(L m )]+nA‖L − L m ‖ L 1 V n ([L m ]) + 1/m + nA‖L − L m ‖ L 1 V n ([L]) + 2nA‖Y − Y m ‖ L 1 + 1/m. Since<br />

the right-hand side c<strong>on</strong>verges to V n ([Y ]) as m →∞, we infer that P2 can guarantee V n (μ) in Γ(μ), forallμ ∈ 2 . ✷<br />

References<br />

Azéma, J., Yor, M., 1979. Une soluti<strong>on</strong> simple au problème de Skorokhod. In: Séminaire de Probabilités XIII. In: Lecture Notes in Math., vol. 721, pp. 90–115.<br />

Bachelier, L., 1900. Théorie de la spéculati<strong>on</strong>. In: Annales scientifiques de l’Ecole Normale Supérieure. Gauthier–Villars, Paris, pp. 3–17, 21–86.<br />

Burkholder, D.L., 1973. Distributi<strong>on</strong> functi<strong>on</strong> inequalities for martingales. Ann. Probab. 1, 19–42.<br />

De Meyer, B., 1998. The maximal variati<strong>on</strong> of a bounded martingale and the central limit theorem. Ann. Inst. H. Poincaré B 34, 49–59.<br />

De Meyer, B., Marino, A., 2005a. C<strong>on</strong>tinuous versus discrete Market games. Cowles Foundati<strong>on</strong> Discussi<strong>on</strong> Paper 1535. Yale University.<br />

De Meyer, B., Marino, A., 2005b. Duality and optimal strategies in the finitely repeated zero-sum games <strong>with</strong> incomplete informati<strong>on</strong> <strong>on</strong> both sides. Cahier<br />

de la mais<strong>on</strong> des sciences éc<strong>on</strong>omiques 2005.27. Université Panthé<strong>on</strong>-Sorb<strong>on</strong>ne, Paris.<br />

De Meyer, B., Lehrer, E., Rosenberg, D., 2009. Evaluating informati<strong>on</strong> in zero-sum games <strong>with</strong> incomplete informati<strong>on</strong> <strong>on</strong> both sides. Documents de travail<br />

du Centre d’Ec<strong>on</strong>omie de la Sorb<strong>on</strong>ne 2009.35. Université Panthé<strong>on</strong>-Sorb<strong>on</strong>ne, Paris.<br />

De Meyer, B., Moussa-Saley, H., 2002. A Model of Games <strong>with</strong> a C<strong>on</strong>tinuum of States of Nature. Prépublicati<strong>on</strong> de l’Institut Elie Cartan, Nancy.<br />

De Meyer, B., Moussa-Saley, H., 2003. On the strategic origin of Brownian moti<strong>on</strong> in finance. Int. J. Game Theory 31, 285–319.<br />

Domansky, V., 2007. Repeated games <strong>with</strong> <strong>asymmetric</strong> informati<strong>on</strong> and random price fluctuati<strong>on</strong>s at finance <strong>market</strong>s. Int. J. Game Theory 36, 241–257.<br />

Dupire, B., 1997. Pricing and hedging <strong>with</strong> smiles. In: Dempster, M.A.H., Pliska, S.R. (Eds.), Mathematics of Derivative Securities. Cambridge University Press,<br />

pp. 103–111.<br />

Kyle, A.S., 1985. C<strong>on</strong>tinuous aucti<strong>on</strong>s and insider trading. Ec<strong>on</strong>ometrica 53, 1315–1336.<br />

Mertens, J.-F. Sorin, S. Zamir, S., 1994. Repeated games. Core Discussi<strong>on</strong> Papers 9420, 9421, 9422. Université Catholique de Louvain, Louvain-la-Neuve.<br />

Mertens, J.-F., Zamir, S., 1976. The normal distributi<strong>on</strong> and repeated games. Int. J. Game Theory 5, 187–197.<br />

Mertens, J.-F., Zamir, S., 1977. The maximal variati<strong>on</strong> of a bounded martingale. Israel J. Math. 27, 252–276.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!