Nonzero-Sum Games of Optimal Stopping for Markov Processes
- 455 Downloads
Abstract
Two players are observing a right-continuous and quasi-left-continuous strong Markov process X. We study the optimal stopping problem \(V^{1}_{\sigma }(x)=\sup _{\tau } \mathsf {M}_{x}^{1}(\tau ,\sigma )\) for a given stopping time \(\sigma \) (resp. \(V^{2}_{\tau }(x)=\sup _{\sigma } \mathsf {M}_{x}^{2}(\tau ,\sigma )\) for given \(\tau \)) where \(\mathsf {M}_{x}^{1}(\tau ,\sigma ) = \mathsf {E}_{x} [G_{1}(X_{\tau })I(\tau \le \sigma ) + H_{1}(X_{\sigma })I(\sigma < \tau )]\) with \(G_1,H_1\) being continuous functions satisfying some mild integrability conditions (resp. \(\mathsf {M}_{x}^{2}(\tau ,\sigma ) = \mathsf {E}_{x} [G_{2}(X_{\sigma })I(\sigma < \tau ) + H_{2}(X_{\tau })I(\tau \le \sigma )]\) with \(G_2,H_2\) being continuous functions satisfying some mild integrability conditions). We show that if \(\sigma = \sigma _{D_{2}} = \inf \{t \ge 0: X_t \in D_2\}\) (resp. \(\tau = \tau _{D_{1}} = \inf \{t \ge 0: X_t \in D_1\}\)) where \(D_{2}\) (resp. \(D_1\)) has a regular boundary, then \(V^{1}_{\sigma _{D_{2}}}\) (resp. \(V^{2}_{\tau _{D_{1}}}\)) is finely continuous. If \(D_{2}\) (resp. \(D_1\)) is also (finely) closed then \(\tau _*^{\sigma _{D_2}} = \inf \{t \ge 0: X_{t} \in D_{1}^{\sigma _{D_{2}}}\}\) (resp. \(\sigma _{*}^{\tau _{D_1}} = \inf \{t \ge 0: X_{t} \in D_{2}^{\tau _{D_{1}}}\}\)) where \(D_{1}^{\sigma _{D_{2}}} = \{V^{1}_{\sigma _{D_{2}}} = G_{1}\}\) (resp. \(D_{2}^{\tau _{D_{1}}} = \{V^{2}_{\tau _{D_{1}}} = G_{2}\}\)) is optimal for player one (resp. player two). We then derive a partial superharmonic characterisation for \(V^{1}_{\sigma _{D_2}}\) (resp. \(V^{2}_{\tau _{D_1}}\)) which can be exploited in examples to construct a pair of first entry times that is a Nash equilibrium.
Keywords
Nonzero-sum optimal stopping game Nash equilibrium Markov process Double partial superharmonic characterisation Principle of double smooth fit Principle of double continuous fitMathematics Subject Classification
Primary 91A15 60G40 Secondary 60J25 60G441 Introduction
Optimal stopping games, often referred to as Dynkin games, are extensions of optimal stopping problems. Since the seminal paper of Dynkin [14], optimal stopping games have been studied extensively. Martingale methods for zero-sum games were studied by Kifer [32], Neveu [44], Stettner [55], Lepeltier and Maingueneau [41] and Ohtsubo [45]. The Markovian framework was initially studied by Frid [22], Gusein-Zade [26], Elbakidze [18] and Bismut [5]. Bensoussan and Friedman [2] and Friedman [20, 21] considered zero-sum optimal stopping games for diffusions and developed an analytic approach by relying on variational and quasi-variational inequalities. Ekström and Peskir [16] proved the existence of a value in two-player zero-sum optimal stopping games for right-continuous strong Markov processes and construct a Nash equilibrium point under the additional assumption that the underlying process is quasi-left continuous. Peskir in [51] and [52] extended these results further by deriving a semiharmonic characterisation of the value of the game without assuming that a Nash equilibrium exists a priori. In particular, a necessary and sufficient condition for the existence of a Nash equilibrium is that the value function coincides with the smallest superharmonic and the largest subharmonic function lying between the gain and the loss function which, in the case of absorbed Brownian motion in [0,1], is equivalent to ‘pulling a rope’ between ‘two obstacles’ (that is finding the shortest path between the graphs of two functions). Connections between zero-sum optimal stopping games and singular stochastic control problems were studied in [23, 31] and[4]. Cvitanic and Karatzas [11] showed that backward stochastic differential equations are connected with the value function of a zero-sum Dynkin game. Advances in this direction can be found in [27]. Various authors have also studied zero-sum optimal stopping games with randomised strategies. For further details one can refer to [39] and the references therein. Zero-sum optimal stopping games have been used extensively in the pricing of game contingent claims both in complete and incomplete markets (see for example [15, 17, 24, 25, 30, 33, 35, 36, 38] and [19]).
Literature on nonzero-sum optimal stopping games is mainly concerned with the existence of a Nash equilibrium. Initial studies in discrete time date back to Morimoto [42] wherein a fixed point theorem for monotone mappings is used to derive sufficient conditions for the existence of a Nash equilibrium point. Ohtsubo [46] derived equilibrium values via backward induction and in [47] the same author considers nonzero-sum games in which the lower gain process has a monotone structure, and gives sufficient conditions for a Nash equilibrium point to exist. Shmaya and Solan in [54] proved that every two player nonzero-sum game in discrete time admits an \(\varepsilon \)-equilibrium in randomised stopping times. In continuous time Bensoussan and Friedman [3] showed that, for diffusions, a Nash equilibrium exists if there exists a solution to a system of quasi-variational inequalities. However, the regularity and uniqueness of the solution remain open problems. Nagai [43] studies a nonzero-sum stopping game of symmetric Markov processes. A system of quasi-variational inequalities is introduced in terms of Dirichlet forms and the existence of extremal solutions of a system of quasi-variational inequalities is discussed. Nash equilibrium points of the stopping game are then obtained from these extremal solutions. Cattiaux and Lepeltier [8] study special right-processes, namely Hunt processes in the Ray topology, and they prove existence of a quasi-Markov Nash Equilbrium. The authors follow Nagai’s idea but use probabilistic tools rather than the theory of Dirichlet forms. Huang and Li in [29] prove the existence of a Nash equilibrium point for a class of nonzero-sum noncyclic stopping games using the martingale approach. Laraki and Solan [40] proved that every two-player nonzero-sum Dynkin game in continuous time admits an \(\varepsilon -\)equilibrium in randomised stopping times. Hamadène and Zhang in [28] prove existence of a Nash equilibrium using the martingale approach, for processes with positive jumps. One application of nonzero-sum optimal stopping games is seen in the study of game options in incomplete markets, via the consideration of utility-based arguments (see [34]). Nonzero-sum optimal stopping games have also been used to model the interaction between bondholders and shareholders in the study of convertible bonds, when corporate taxes are included and when the company is allowed to claim default (see [9]).
In this work we consider two player nonzero-sum games of optimal stopping for a general strong Markov process. The aim is to use probabilistic tools to study the optimal stopping problem of player one (resp. player two) when the stopping time of player two (resp. player one) is externally given. Although this work does not deal with the question of existence of mutually best responses (that is the existence of a Nash equilibrium) the results obtained can be exploited further in various examples, to show the existence of a pair of first entry times that will be a Nash equilibrium. Indeed, the results derived here will be used in a separate work (see [1]) to construct Nash equilibrium points for one dimensional regular diffusions and for a certain class of payoff functions.
This paper is organised as follows: In Sect. 2 we introduce the underlying setup and formulate the nonzero-sum optimal stopping game. In Sect. 3 we show that if the strategy chosen by player two (resp. player one) is \(\sigma _{D_2}\) (resp. \(\tau _{D_1}\)), the first entry time into a regular Borel subset \(D_2\) (resp. \(D_1\)) of the state space, then the value function of player one associated with \(\sigma _{D_2}\) (resp. the value function of player two associated with \(\tau _{D_1}\)), which we shall denote by \(V^1_{\sigma _{D_2}}\) (resp. \(V^2_{\tau _{D_1}}\)), is finely continuous. In Section 4 we shall use this regularity property of \(V^1_{\sigma _{D_2}}\) (resp. \(V^2_{\tau _{D_1}}\)) to construct an optimal stopping time for player one (resp. player two). In Sect. 5 we shall use the results obtained in Sects. 3 and 4 to provide a partial superharmonic characterisation for \(V^1_{\sigma _{D_2}}\) (resp. \(V^2_{\tau _{D_1}}\)). More precisely if \({D_2}\) (resp. \(D_1\)) is also a closed or finely closed subset of the state space then \(V^1_{\sigma _{D_2}}\) (resp. \(V^2_{\tau _{D_1}}\)) can be identified with the smallest finely continuous function that is superharmonic in \(D_2^c\) (resp. in \(D_1^c\)) and that majorises the lower payoff function. In Section 6 we shall consider stationary one-dimensional Markov processes and we shall assume that there exists a pair of stopping times \((\tau _{A_*},\sigma _{B_*})\) of the form \(\tau _{A_*}= \inf \{t\ge 0: X_t \le A_*\}\) and \(\sigma _{B_*} = \inf \{t \ge 0 : X_t \ge B_*\}\) where \(A_*< B_*\), that is a Nash equilibrium point. We first show that \(V^{1}_{\sigma _{B_*}}\) (resp. \(V^{1}_{\tau _{A_*}}\)) is continuous at \(A_*\) (resp. at \(B_*\)). Then for the special case of one dimensional regular diffusions we shall use the results obtained in Sect. 5 to show that \(V^{1}_{\sigma _{B_*}}\) (resp. \(V^{2}_{\tau _{A_*}}\)) is also smooth at \(A_*\) (resp. \(B_*\)) provided that the payoff functions are smooth. This is in line with the principle of smooth fit observed in standard optimal stopping problems (see for example [49] for further details).
2 Formulation of the Problem
In this section we shall formulate rigorously the nonzero-sum optimal stopping game. For this we shall first set up the underlying framework. This will be similar to the one presented by Ekström and Peskir (cf. [16, p. 3]). On a given filtered probability space \(\left( \Omega ,\mathcal {F},(\mathcal {F}_t)_{t\ge 0},\mathsf {P}_{x}\right) \) we define a strong Markov process \(X=(X_{t})_{t \ge 0}\) with values in a measurable space \((E,\mathcal {B})\), with E being a locally compact Hausdorff space with a countable base (note that since E has a countable base then it is a Polish space) and \(\mathcal {B}\) the Borel \(\sigma \)-algebra on E. We shall assume that \(\mathsf {P}_{x}\left( X_0=x\right) =1\), that the sample paths of X are right-continuous and that X is quasi-left-continuous (that is \(X_{\rho _{n}} \rightarrow X_{\rho }\) \(\mathsf {P}_{x}\)-a.s. whenever \(\rho _{n}\) and \(\rho \) are stopping times such that \(\rho _{n} \uparrow \rho \) \(\mathsf {P}_{x}\)-a.s.). All stopping times mentioned throughout this text are relative to the filtration \((\mathcal {F}_{t})_{t \ge 0}\) introduced above, which is also assumed to be right-continuous. This means that entry times in open and closed sets are stopping times. Moreover \(\mathcal {F}_0\) is assumed to contain all \(\mathsf {P}_{x}\)-null sets from \(\mathcal {{F}}^{X}_\infty = \sigma \left( X_{t}:t \ge 0\right) \), which further implies that the first entry times to Borel sets are stopping times. We shall also assume that the mapping \(x \mapsto \mathsf {P}_{x}(F)\) is (universally) measurable for each \(F \in \mathcal {F}\) so that the mapping \(x \mapsto \mathsf {E}_{x} [Z]\) is (universally) measurable for each (integrable) random variable Z.
Remark 2.1
Note that a subset F of a Polish space E is said to be universally measurable if it is \(\mu \)-measurable for every finite measure \(\mu \) on \((E,\mathcal {B})\), where \(\mathcal {B}\) is the Borel-\(\sigma \) algebra on E. By \(\mu \)-measurable we mean that F is measurable with respect to the completion of \(\mathcal {B}\) under \(\mu \). If \(\mathcal {B}^*\) is the collection of universally measurable subsets of E then a function \(f:E \rightarrow \mathbb {R}\) is said to be universally measurable if \(f^{-1}(A) \in \mathcal {B}^*\) for all \(A \in \mathcal {B}(\mathbb {R})\), where \(\mathcal {B}(\mathbb {R})\) is the Borel-\(\sigma \) algebra on \(\mathbb {R}\).
Finally we shall assume that \(\Omega \) is the canonical space \(E^{\left[ 0,\infty \right) }\) with \(X_{t}(\omega ) = \omega (t)\) for \(\omega \in \Omega \). In this case the shift operator \(\theta _{t}:\Omega \rightarrow \Omega \) is well defined by \(\theta _{t}(\omega )(s) = \omega (t+s)\) for \(\omega \in \Omega \) and \(t,s \ge 0\).
The game is said to have a solution if there exists a pair of stopping times \((\tau _*,\sigma _{*})\) which is a Nash equilibrium point, that is \(\mathsf {M}_{x}^1(\tau ,\sigma _{*}) \le \mathsf {M}_{x}^1(\tau _*,\sigma _{*})\) and \(\mathsf {M}_{x}^2(\tau _*,\sigma ) \le \mathsf {M}_{x}^2(\tau _*,\sigma _{*})\) for all stopping times \(\tau , \sigma \). This means that none of the players will perform better if they change their strategy independent of each other. In this case \(V^{1}_{\sigma _{*}}(x) = \mathsf {M}_{x}^{1}(\tau _{*},\sigma _{*})\) is the payoff function of player one and \(V^{2}_{\tau _{*}}(x)=\mathsf {M}_{x}^{2}(\tau _{*},\sigma _{*})\) the payoff function of player two in this equilibrium. So \(V^{1}_{\sigma _*}\) and \(V^{2}_{\tau _*}\) can be called the value functions of the game (corresponding to \((\tau _*,\sigma _*))\). In general, as we shall see in Sect. 6, there might be other pairs of stopping times that form a Nash equilibrium point, which can lead to different value functions.
3 Fine continuity property
In this section we show that if the strategy chosen by player two (resp. player one) corresponds to the first entry time into a subset \(D_2\) (resp. \(D_1\)) of E, whose boundary \(\partial D_2\) (resp. \(\partial D_1\)) is regular, then \(V^{1}_{\sigma _{D_2}}\) (resp. \(V^{2}_{\tau _{D_1}})\) is continuous in the fine topology (i.e. finely continuous). For literature on the fine topology one can refer to [6, 10] and [13]. We first define the concept of a finely open set and a regular boundary of a Borel subset of E.
Definition 3.1
An arbitrary set \(B \subseteq E\) is said to be finely open if there exists a Borel set \(A \subseteq B\) such that \(\mathsf {P}_{x}(\rho _{A^{c}} > 0) = 1\) for every \(x \in A\), where \(\rho _{A^{c}} = \inf \{t > 0: X_{t} \in A^{c}\}\) is the first hitting time in \(A^{c}\).
Definition 3.2
The boundary \(\partial D\) of a Borel set \(D \subseteq E\) is said to be regular for D if \(\mathsf {P}_{x}(\rho _{D} = 0) = 1\) for every point \(x \in \partial D\), where \(\rho _{D} = \inf \{t>0:X_{t} \in D \}\).
We now introduce preliminary results which are needed to prove the main theorem of this section.
Lemma 3.3
For any given stopping time \(\sigma \) (resp. \(\tau \)), the mapping \(x \mapsto V^{1}_{\sigma }(x)\), (resp. \(x \mapsto V^{2}_{\tau }(x)\)) is measurable.
Proof
Lemma 3.4
Let D be a Borel subset of E and let \(x \in \partial D\), where \(\partial D\) is a regular boundary for D. Suppose that \((\rho _{n})_{n=1}^{\infty }\) is a sequence of stopping times such that \(\rho _{n} \downarrow 0 \text { } \mathsf {P}_{x}\)-a.s. as \(n \rightarrow \infty \). Set \(\sigma _{\rho _{n}}= \inf \{t \ge \rho _{n} : X_{t} \in D\} \). Then \(\sigma _{\rho _{n}} \downarrow 0 \text { } \mathsf {P}_{x}\text {-a.s.}\) as \(n \rightarrow \infty \).
Proof
Let \(x \in \partial D\). By regularity of \( \partial D\) for any \(\varepsilon > 0\) there exists \(t \in (0,\varepsilon )\) such that \(X_{t} \in D\) \(\mathsf {P}_{x}\)-a.s. Since \(\sigma _{\rho _{n}}\) is a sequence of decreasing stopping times then \(\sigma _{\rho _{n}} \downarrow \beta \text { } \mathsf {P}_{x}\)-a.s. for some stopping time \(\beta \). So suppose for contradiction that \(\beta > 0\). Now \(\rho _{n} \downarrow 0 \text { } \mathsf {P}_{x}\)-a.s. and for each n we have \(\sigma _{\rho _{n}} \ge \beta \text { } \mathsf {P}_{x}\)-a.s. So for any given \(\omega \in \Omega \backslash N\) where \(\mathsf {P}_{x}(N) = 0\) we have that \(X_{t}(\omega ) \notin D\) for all \(t \in (0,\beta (\omega ))\) and this contradicts the fact that \(\partial D\) is regular for D.
The next lemma and theorem, which we shall exploit in this study, provide conditions for fine continuity. The proofs of these results can be found in [13].
Lemma 3.5
Theorem 3.6
Let \(F:E \rightarrow \mathbb {R}\) be a measurable function and suppose that \(K_{1} \subseteq K_{2} \subseteq K_{3} \subseteq ...\) is a nested sequence of compact sets in E. Suppose also that \(\rho _{K_{n}} \downarrow 0 \text { } \mathsf {P}_{x}\)-a.s. as \(n \rightarrow \infty \), where \(\rho _{K_{n}} = \inf \{t \ge 0: X_{t} \in K_{n}\}\). If \(\lim _{n \rightarrow \infty } \mathsf {E}_{x}[F(X_{\rho _{K_{n}}}) ]= F(x)\) then F is finely continuous.
We next state and prove the main result of this section, that is the fine continuity property of \(V^{1}_{\sigma _{D_2}}\) (resp. \(V^{2}_{\tau _{D_1}}\)).
Theorem 3.7
Proof
We will only prove (3.5) as (3.6) follows by symmetry.
We next present an example to show that if \(\partial D\) is not regular for D then \(V^{1}_{\sigma _{D}}\) may not be finely continuous.
Example 3.8
4 Towards a Nash equilibrium
The main result of this section is to show that if \(\sigma _{D_2}\) (resp. \(\tau _{D_1}\)) is externally given as the first entry time in \(D_2\) (resp. \(D_1\)), a set that is either closed or finely closed, and has a regular boundary, then the first entry time \(\tau _{*}^{\sigma _{D_2}} = \inf \{t \ge 0:X_t \in D^{\sigma _{D_2}}_1\}\) (resp. \(\sigma _{*}^{\tau _{D_1}} = \inf \{t \ge 0: X_t \in D^{\tau _{D_1}}_1 \}\)) where \(D^{\sigma _{D_2}}_1 = \{V^{1}_{\sigma _{D_2}} = G_1\}\) (resp. \(D^{\tau _{D_1}}_2 = \{V^{2}_{\tau _{D_1}} = G_2\}\)) solves the optimal stopping problem \(V_{\sigma _{D_{2}}}^{1}(x)=\sup _{\tau } \mathsf {M}_{x}^{1}(\tau ,\sigma _{D_{2}})\) (resp. \(V_{\tau _{D_{1}}}^{2}(x)=\sup _{\sigma } \mathsf {M}_{x}^{2}(\tau _{D_1},\sigma ))\). The proof of this result will be divided into several lemmas and propositions.
Proposition 4.1
Proof
We shall only prove (4.1) as for (4.2) the result follows by symmetry. The proof will be carried out in several steps.
Lemma 4.2
Proof
Since \(\rho _{n}\) is an increasing sequence of stopping times, \(\sigma _{\rho _{n}}\) is increasing and thus \(\sigma _{\rho _{n}} \uparrow \beta \text { }\mathsf {P}_{x}\)-a.s. for some stopping time \(\beta \). We need to prove that \(\beta = \sigma _{\rho } \text { } \mathsf {P}_{x}-a.s.\)
\(\mathbf {2^{\circ }}\) Let \(\Omega _{1}=\{\omega \in \Omega : \beta (\omega ) > \rho (\omega )\}\) and \(\Omega _{2} = \{\omega \in \Omega : \beta (\omega ) = \rho (\omega )\}\). We prove that there exists a set N with \(\mathsf {P}_{x}(N) =0 \) such that \(\beta (\omega ) = \sigma _{\rho }(\omega )\) for every \(\omega \in (\Omega _{1} \cup \Omega _{2}) \backslash {N}\).
\(\mathbf {(i.)}\) Suppose first that \(\mathsf {P}_{x}(\Omega _{1}) > 0\). Since \(\sigma _{\rho _{n}} \uparrow \beta \) \(\mathsf {P}_{x}\)-a.s. then for every \(\omega \in \Omega _{1} \backslash N_{3}\), where \(\mathsf {P}_{x}(N_{3}) = 0\), there exists \(n_{2}(\omega ) \in \mathbb {N}\) such that \(\sigma _{\rho _{n}}(\omega ) > \rho (\omega )\) for all \(n \ge n_{2}(\omega )\). Moreover, since \(\sigma _{\rho }(\omega ) \ge \rho (\omega )\) for every \(\omega \in \Omega _{1} \backslash N_{4}\) where \(\mathsf {P}_{x}(N_{4}) = 0\) it follows that for every \(\omega \in \Omega _{1} \backslash N\), where \(N = N_{3} \cup N_{4}\), there exists \(n_{3}(\omega ) \in \mathbb {N}\) such that \(\sigma _{\rho _{n}}(\omega ) = \sigma _{\rho }(\omega )\) for all \(n \ge n_{3}(\omega )\). From this it follows that \(\beta = \sigma _{\rho }\) \(\mathsf {P}_{x}\)-a.s. on \(\Omega _{1}\).
Proposition 4.3
Let \(D_1,D_{2}\) be either closed or finely closed subsets of E. Suppose also that their respective boundaries \(\partial D_1\) and \(\partial D_{2}\) are regular. Let \(\tau _{D_1}\) and \(\sigma _{D_{2}}\) be the first entry times into \(D_1\) and \(D_2\) respectively. Set \(\tau _{\varepsilon }^{\sigma _{D_{2}}} = \inf \{t \ge 0: X_{t} \in D_{1}^{\sigma _{D_{2}},\varepsilon }\}\) and \(\sigma _{\varepsilon }^{\tau _{D_{1}}} = \inf \{t \ge 0: X_{t} \in D_{2}^{\tau _{D_{1}},\varepsilon }\}\) where \(D_{1}^{\sigma _{D_{2}},\varepsilon } = \{V_{\sigma _{D_{2}}}^{1} \le G_{1} + \varepsilon \}\) and \(D_{2}^{\tau _{D_{1}},\varepsilon } = \{V_{\tau _{D_{1}}}^{2} \le G_{2} + \varepsilon \}\). Then \(\tau _{\varepsilon }^{\sigma _{D_{2}}} \uparrow \tau _{*}^{\sigma _{D_{2}}}\) \(\mathsf {P}_{x}\)-a.s. and \(\sigma _{\varepsilon }^{\tau _{D_{1}}} \uparrow \sigma _{*}^{\tau _{D_{1}}}\) \(\mathsf {P}_{x}\)-a.s., where \(\tau _{*}^{\sigma _{D_{2}}} = \inf \{t \ge 0: X_{t} \in D_{1}^{\sigma _{D_{2}}}\}\) with \(D_{1}^{\sigma _{D_{2}}} = \{V_{\sigma _{D_{2}}}^{1} = G_{1}\}\) and \(\sigma _{*}^{\tau _{D_{1}}} = \inf \{t \ge 0: X_{t} \in D_{2}^{\tau _{D_{1}}}\}\) with \(D_{2}^{\tau _{D_{1}}} = \{V_{\tau _{D_{1}}}^{2} = G_{2}\}\).
Proof
We now state and prove the main result of this section.
Theorem 4.4
Proof
5 Partial superharmonic characterisation
- (i.)
u lies between \(G_{1}\) and \(K_1\), u is identified with \(H_{1}\) in the region \(D_{2} = \{v = G_{2}\}\) and u is the smallest superharmonic function that dominates \(G_{1}\) in the region \(\{v > G_2\}\)
- (ii.)
v lies between \(G_{2}\) and \(K_2\), v is identified with \(H_2\) in the region \(D_{1} = \{u = G_1\}\) and v is the smallest superharmonic function that dominates \(G_2\) in the region \(\{u > G_1\}\).
The double partial superharmonic characterisation of the value functions in a nonzero-sum optimal stopping game for absorbed Brownian motion
Definition 5.1
Let C be a measurable subset of E and \(D = E \backslash C\). A measurable function \(F:E \rightarrow \mathbb {R}\) is said to superharmonic in C if \(\mathsf {E}_{x}[F(X_{\rho \wedge \sigma _{D}})] \le F(x)\) for every stopping time \(\rho \) and for all \(x \in E\), where \(\sigma _{D} = \inf \{t \ge 0 : X_{t} \in D\}\). F is said to be superharmonic if \(\mathsf {E}_{x}[F(X_{\rho })] \le F(x)\) for every stopping time \(\rho \) and for all \(x \in E\).
Lemma 5.2
Proof
Theorem 5.3
Proof
6 The case of stationary one-dimensional Markov processes
In this section we shall assume that the Markov process X takes values in \(\mathbb {R}\) and is such that Law\((X|\mathsf {P}_{x}) = \) Law\((X^{x}|\mathsf {P})\). We shall also assume that there exist points \(A_*\) and \(B_*\) satisfying \(-\infty< A_*< B_*< \infty \) such that (i.) for given \(D_2\) of the form \([B_*,\infty )\), the first entry time \(\tau _{A_*} = \inf \{t \ge 0:X_t \le A_*\}\) (as obtained from Theorem 4.4) is optimal for player one and (ii.) for given \(D_1\) of the form \((-\infty ,A_*]\), the first entry time \(\sigma _{B_*} = \inf \{t \ge 0:X_t \ge B_*\}\) (as obtained from Theorem 4.4) is optimal for player two.
So in this section we will assume the existence of a pair \((\tau _{A_*}, \sigma _{B_*})\) that is a Nash equilibrium.
6.1 The principle of double continuous fit
We shall first state and prove the following lemma.
Lemma 6.1
Let \(\sigma _{B_*}^{A_{*}} = \inf \{t \ge 0: X_t \ge B_*\}\) be the optimal stopping time for player two under \(\mathsf {P}_{A_*}\) and let \(\sigma _{B_*}^{A_{*}+\varepsilon } = \inf \{t \ge 0: X_t \ge B_*\}\) be the optimal stopping time of player two under \(\mathsf {P}_{A_{*} + \varepsilon }\), for given \(\varepsilon > 0\). Then, if condition (6.3) is satisfied we have that \(\sigma _{B_*}^{A_*+ \varepsilon } \uparrow \sigma _{B_*}^{A_*}\) as \(\varepsilon \downarrow 0\). Similarly, if \(\tau _{A_*}^{B_{*}} = \inf \{t \ge 0: X_t \le A_*\}\) is the optimal stopping time for player two under \(\mathsf {P}_{B_*}\) and \(\tau _{A_*}^{B_{*}-\varepsilon } = \inf \{t \ge 0: X_t \le A_*\}\) is the optimal stopping time of player one under \(\mathsf {P}_{B_{*} - \varepsilon }\), for given \(\varepsilon > 0\), then if condition (6.4) is satisfied we have that \(\tau _{A_*}^{B_*- \varepsilon } \uparrow \tau _{A_*}^{B_*}\) as \(\varepsilon \downarrow 0\).
Proof
We shall only prove that \(\sigma _{B_*}^{A_*+ \varepsilon } \uparrow \sigma _{D_2}^{A_*}\) as \(\varepsilon \downarrow 0\). The fact that \(\tau _{A_*}^{B_*- \varepsilon } \uparrow \tau _{A_*}^{B_*}\) as \(\varepsilon \downarrow 0\) can be proved in the same way. Since \(\text {Law}(X|\mathsf {P}_{x}) = \text {Law}(X^x|\mathsf {P})\) we have that \(\sigma _{B_*}^{A_{*}+\varepsilon } \) is equally distributed as \(\hat{\sigma }_{B_*}^{A_{*}+\varepsilon }:= \inf \{t \ge 0: X_t^{A_{*}+\varepsilon } \ge B_*\}\) under \(\mathsf {P}\) whereas \(\sigma _{B_*}^{A_{*}}\) is equally distributed as \(\hat{\sigma }_{B_*}^{A_{*}}:= \inf \{t \ge 0: X_t^{A_*} \ge B_*\}\) under \(\mathsf {P}\). Now \(\hat{\sigma }_{B_*}^{A_*+ \varepsilon } \uparrow \gamma \) as \(\varepsilon \downarrow 0\) for some stopping time \(\gamma \le \hat{\sigma }_{B_*}^{A_*}\). So to prove the result it remains to show that \(\gamma \ge \hat{\sigma }_{B_*}^{A_*}\). By the time-space condition (6.3) we have that \(X_{\hat{\sigma }_{B_*}^{A_*+ \varepsilon }}^{A_*+ \varepsilon } \rightarrow X_{\gamma }^{A_*}\) \(\mathsf {P}\)-a.s. as \(\varepsilon \downarrow 0\). Now since \(X_{\hat{\sigma }_{B_*}^{A_*+ \varepsilon }}^{A_*+ \varepsilon } \ge B_*\) for each \(\varepsilon > 0\) it follows that \(X_{\gamma }^{A_*} \ge B_*\). But this implies that \(\gamma \ge \hat{\sigma }_{B_*}^{A_*}\) and this proves the required result. \(\square \)
Proposition 6.2
Suppose that the payoff functions \(G_{i}, H_{i}\) for \(i=1,2\) are also assumed to be bounded. Then the value functions \(V^{1}_{\sigma _{B_*}}\) and \(V^{2}_{\tau _{A_*}}\) are continuous at \(A_*\) and \(B_*\) respectively.
Proof
Remark 6.3
The assumption of boundedness on \(G_1\) and \(H_1\) in Proposition 6.2 can be relaxed. For example, the result will also hold provided that \(G_1(X^{A_*+ \varepsilon }_{\hat{\tau }_{A_*}^{A_*+ \varepsilon }}) - G_1(X^{A_*}_{\hat{\tau }^{A_*+ \varepsilon }_{A_*}})\) and \(H_1(X^{A_*+ \varepsilon }_{\hat{\sigma }_{B_*}^{A_*+ \varepsilon }}) - H_1(X^{A_*}_{\hat{\sigma }_{B_*}^{A_*}})\) are bounded above by some integrable random variables \(\tilde{Z}_1\) and \(\tilde{Z}_2\) respectively. Similarly the boundedness assumption on \(G_2\) and \(H_2\) can be relaxed.
6.2 The principle of double smooth fit
In this section we will consider the special case when X is a one dimensional regular diffusion process and we shall assume that \(V^{1}_{\sigma _{B_*}}\) and \(V^{2}_{\tau _{A_*}}\) are obtained from the double partial superharmonic characterisation as explained in Sect. 5. More precisely we shall assume that the functions u, v introduced in Theorem 5.3 (i.) coincide with those from Theorem 5.3 (ii.) so that a mutual response is assumed to exist. The aim is to use this characterisation to derive the so-called principle of double smooth fit. This principle is an extension of the principle of smooth fit observed in standard optimal stopping problems (see [49]). We note that in the case of more general strong Markov processes in \(\mathbb {R}\) this principle may break down. As observed in standard optimal stopping problems this may happen for example when the scale function of X is not differentiable (see [50]) or in the case of Poisson process (see [48]). Carr et. al in [7], for example, also showed that this principle breaks down in a CGMY model.
Remark 6.4
Examples of nonzero-sum optimal stopping games for one dimensional regular diffusion processes, for which the optimal stopping regions are of the threshold type are given in [1] and [12]1. In particular the authors therein provide sufficient conditions for existence and uniqueness of Nash equilibria.
Lemma 6.5
Suppose that \(f,g:\mathbb {R}_{+} \rightarrow \mathbb {R}\) are two continuous functions such that \(f(0)=g(0)=0\), \(f(\varepsilon ) > 0\) whenever \(\varepsilon > 0\), and \(g(\delta ) > 0\) whenever \(\delta > 0\). Then for every \(\varepsilon _{n} \downarrow 0\) as \(n \rightarrow \infty \), there exists \(\varepsilon _{n_{k}} \downarrow 0\) and \(\delta _{k} \downarrow 0\) as \(k \rightarrow \infty \) such that \(\lim _{k \rightarrow \infty } \frac{f(\varepsilon _{n_{k}})}{g(\delta _{k})} = 1.\)
Proposition 6.6
Suppose that \(D_{1}\) is of the form \((-\infty ,A_{*}]\) and \(D_{2}\) of the form \([B_{*},\infty )\) for some points \(A_{*},B_{*}\) such that \( A_{*} < B_{*} \). Suppose that \(G_{1}\) is differentiable at \(A_{*}\) and \(G_{2}\) is differentiable at \(B_{*}\). If the scale function S of X is differentiable at \(A_{*}\) and \(B_{*}\), then \(u'(A_{*}) = G'_{1}(A_{*})\) and \(v'(B_{*}) = G'_{2}(B_{*})\).
Proof
Examples for absorbed Brownian motion which show that uniqueness of Nash equilibrium fails in general
Footnotes
- 1.
The second manuscript was available to the author after the first draft of the paper was published on The University of Manchester Website http://www.maths.manchester.ac.uk/our-research/research-groups/statistics-and-its-applications/research-reports/.
Notes
Acknowledgements
The author is grateful to Professor Goran Peskir for introducing the topic of optimal stopping games, for the many fruitful discussions on the subtleties of Markov processes and the principles of smooth and continuous fit in zero-sum games, and for providing insight into the variational approach as a way of observing and understanding the principles of ‘double smooth fit’ and ‘double continuous fit’ in nonzero-sum games.
References
- 1.Attard, N.: Nash equilibrium in nonzero-sum games of optimal stopping for Brownian motion. Research Report No. 2, Probab. Statist. Group Manchester (2015)Google Scholar
- 2.Bensoussan, F., Friedman, A.: Nonlinear variational inequalities and differential games with stopping times. J. Funct. Anal. 16, 305–352 (1974)MathSciNetCrossRefMATHGoogle Scholar
- 3.Bensoussan, F., Friedman, A.: Nonzero-sum stochastic differential games with stopping times and free boundary value problem. Trans. Am. Math. Soc. 213, 275–327 (1977)CrossRefMATHGoogle Scholar
- 4.Boetius, F.: Bounded variation singular stochastic control and Dynkin game. SIAM J. Control Optim. 44, 1289–1321 (2005)MathSciNetCrossRefMATHGoogle Scholar
- 5.Bismut, J.M.: Sur un problème de Dynkin. Z. Wahrsch. Verw. Geb. 39, 31–53 (1977)MathSciNetCrossRefMATHGoogle Scholar
- 6.Blumenthal, R.M., Getoor, R.K.: Markov Processes and Potential Theory. Academic Press, New York (1968)MATHGoogle Scholar
- 7.Carr, P., Geman, H., Madan, D., Yor, M.: The fine structure of returns: an empirical investigation. J. Bus. 75, 305–332 (2002)CrossRefGoogle Scholar
- 8.Cattiaux, P., Lepeltier, J.P.: Existence of a quasi-Markov Nash equilibrium for nonzero-sum Markov stopping games. Stoch. Stoch. Rep. 30, 85–103 (1990)MathSciNetCrossRefMATHGoogle Scholar
- 9.Chen, N., Dai, M., Wan, X.: A nonzero-sum game approach to convertible bonds: tax benefit, bankruptcy cost and early/late calls. Math. Financ. 23(1), 57–93 (2013)MathSciNetCrossRefMATHGoogle Scholar
- 10.Chung, K.L., Walsh, J.B.: Markov Processes, Brownian Motion, and Time Symmetry. Springer, New York (2005)CrossRefMATHGoogle Scholar
- 11.Cvitanic, I., Karatzas, I.: Backward SDEs with reflection and Dynkin games. Ann. Probab. 24, 2024–2056 (1996)MathSciNetCrossRefMATHGoogle Scholar
- 12.De Angelis, T., Ferrari, G., Moriarty, J.: Nash equilibria of threshold type for two-player nonzero-sum games of stopping. (2015) arXiv:1508.03989
- 13.Dynkin, E.: Markov Processes, Volume 1. Academic Press, New York (1965)CrossRefGoogle Scholar
- 14.Dynkin, E.: Game variant of a problem of optimal stopping. Soviet Math. Dokl. 10, 16–19 (1969)MATHGoogle Scholar
- 15.Ekström, E.: Properties of game options. Math. Methods Oper. Res. 63, 221–238 (2006)MathSciNetCrossRefMATHGoogle Scholar
- 16.Ekström, E., Peskir, G.: Optimal stopping games for Markov processes. SIAM J. Control Optim. 47, 684–702 (2008)MathSciNetCrossRefMATHGoogle Scholar
- 17.Ekström, E., Villeneuve, S.: On the value of optimal stopping games. Ann. Appl. Probab. 16, 1576–1596 (2006)MathSciNetCrossRefMATHGoogle Scholar
- 18.Elbakidze, N.V.: Construction of the cost and optimal policies in a game problem of stopping a Markov process. Theory Probab. Appl. 21, 163–168 (1976)MathSciNetCrossRefMATHGoogle Scholar
- 19.Emmerling, T.J.: Perpetual cancellable American call options. Math. Financ. 22, 645–666 (2012)MathSciNetCrossRefMATHGoogle Scholar
- 20.Friedman, A.: Stochastic games and variational inequalities. Arch. Ration. Mech. Anal. 51, 321–341 (1973a)MathSciNetCrossRefMATHGoogle Scholar
- 21.Friedman, A.: Regularity theorem for variational inequalities in unbounded domains and applications to stopping time problems. Arch. Ration. Mech. Anal. 76, 134–160 (1973b)MathSciNetMATHGoogle Scholar
- 22.Frid, E.B.: The optimal stopping rule for a two-person Markov chain with opposing interests. Theory Probab. Appl. 14, 713–716 (1969)CrossRefMATHGoogle Scholar
- 23.Fukushima, M., Taksar, M.: Dynkin games via Dirichlet forms and singular control of one-dimensional diffusions. SIAM J. Control Optim. 41, 682–699 (2002)MathSciNetCrossRefMATHGoogle Scholar
- 24.Gapeev, P.V.: The spread option optimal stopping game. In: Kyprianou, A., et al. (eds.) Exotic Option Pricing and Advanced Lévy Models, pp. 205–293. Wiley, New York (2006)Google Scholar
- 25.Gapeev, P.V., Kühn, C.: Perpetual convertible bonds in jump-diffusion models. Stat. Decis. 23, 15–31 (2005)MathSciNetMATHGoogle Scholar
- 26.Gusein-Zade, S.M.: On a game connected with a Wiener process. Theory Probab. Appl. 14(704), 701 (1969)MathSciNetCrossRefMATHGoogle Scholar
- 27.Hamadène, S., Hassani, M.: BSDEs with two reflecting barriers: the general result. Probab. Theory Related Fields 132(264), 237 (2005)MathSciNetCrossRefMATHGoogle Scholar
- 28.Hamadène, S., Zhang, M.: The continuous time nonzero-sum Dynkin game problem and application in game options. SIAM J. Control Optim. 5, 3659–3669 (2010)MathSciNetCrossRefMATHGoogle Scholar
- 29.Huang, C.-F., Li, L.: Continuous time stopping games with monotone reward structures. Math. Oper. Res. 15, 496–507 (1990)MathSciNetCrossRefMATHGoogle Scholar
- 30.Kallsen, J., Kühn, C.: Pricing derivatives of American and game type in incomplete markets. Financ. Stoch. 9, 261–284 (2004)MathSciNetCrossRefMATHGoogle Scholar
- 31.Karatzas, I., Wang, H.: Connections Between Bounded-Variation Control and Dynkin Games. Volume in Honor of Professor Alain Bensoussan, pp. 353–362. IOS Press, Amsterdam (2001)Google Scholar
- 32.Kifer, Y.: Optimal stopping in games with continuous time. Theory Probab. Appl. 16, 545–550 (1971)MathSciNetCrossRefMATHGoogle Scholar
- 33.Kifer, Y.: Game options. Financ. Stoch. 4, 443–463 (2000)MathSciNetCrossRefMATHGoogle Scholar
- 34.Kühn, C.: Game contingent claims in complete and incomplete markets. J. Math. Econom. 40, 889–902 (2004)MathSciNetCrossRefMATHGoogle Scholar
- 35.Kühn, C., Kyprianou, A.E.: Callable puts as composite exotic options. Math. Financ. 17, 487–502 (2007)MathSciNetCrossRefMATHGoogle Scholar
- 36.Kühn, C., Kyprianou, A.E., Van Schaik, K.: Pricing Israeli options: a pathwise approach. Stochastics 79, 117–137 (2007)MathSciNetCrossRefMATHGoogle Scholar
- 37.Kunita, H.: Stochastic differential equations based on Lévy processes and stochastic flows of diffeomorphisms. In: Rao, M.M. (ed.) Real and Stochastic Analysis: New Perspectives (Trends in Mathematics), pp. 305–373. Birkhäuser, Boston (2004)CrossRefGoogle Scholar
- 38.Kyprianou, A.E.: Some calculations for Israeli options. Financ. Stoch. 8, 73–86 (2004)MathSciNetCrossRefMATHGoogle Scholar
- 39.Laraki, R., Solan, E.: The value of zero-sum stopping games in continuous time. SIAM J. Control Optim. 43, 1913–1922 (2005)MathSciNetCrossRefMATHGoogle Scholar
- 40.Laraki, R., Solan, E.: Equilibrium in two-player nonzero-sum Dynkin games in continuous time. Stochastics 85, 997–1014 (2012)CrossRefMATHGoogle Scholar
- 41.Lepeltier, J.P., Maingueneau, M.A.: Le Jeu de Dynkin en théorie générale sans l’hypothése de Mokobodski. Stochastics 13, 25–44 (1984)MathSciNetCrossRefMATHGoogle Scholar
- 42.Morimoto, Y.: Nonzero-sum discrete parameter stochastic games with stopping times. Probab. Theory Relat. Fields 72, 155–160 (1986)MathSciNetCrossRefMATHGoogle Scholar
- 43.Nagai, H.: Nonzero-sum stopping games of symmetric Markov processes. Probab. Theory Relat. Fields 75, 487–497 (1987)CrossRefMATHGoogle Scholar
- 44.Neveu, J.: Discrete-Parameter Martingales. North-Holland Publishing Company, Amsterdam (1975)MATHGoogle Scholar
- 45.Ohtsubo, Y.: Optimal Stopping in sequential games with or without a constraint of always terminating. Math. Oper. Res. 12, 277–296 (1986)MathSciNetCrossRefMATHGoogle Scholar
- 46.Ohtsubo, Y.: A nonzero-sum extension of Dynkin stopping problem. Math. Oper. Res. 11, 591–607 (1987)MathSciNetCrossRefMATHGoogle Scholar
- 47.Ohtsubo, Y.: On a discrete-time nonzero-sum Dynkin problem with monotonicity. J. Appl. Probab. 28, 466–472 (1991)MathSciNetCrossRefMATHGoogle Scholar
- 48.Peskir, G., Shiryaev, A.N.: Sequential testing problems for Poisson processes. Ann. Stat. 28, 837–859 (2000)MathSciNetCrossRefMATHGoogle Scholar
- 49.Peskir, G., Shiryaev, A.N.: Optimal Stopping and Free-Boundary Problems. Lectures in Mathematics. ETH Zürich, Birkhäuser, Basel (2006)MATHGoogle Scholar
- 50.Peskir, G.: Principle of smooth fit and diffusions with angles. Stochastics 79, 239–302 (2006)MathSciNetGoogle Scholar
- 51.Peskir, G.: Optimal stopping games and Nash equilibrium. Theory Probab. Appl. 53, 558–571 (2008)MathSciNetCrossRefMATHGoogle Scholar
- 52.Peskir, G.: A duality principle for the Legendre transform. J. Convex Anal. 19, 609–630 (2012)MathSciNetMATHGoogle Scholar
- 53.Sharpe, M.: General Theory of Markov Processes. Academic Press, New York (1988)MATHGoogle Scholar
- 54.Shmaya, E., Solan, E.: Two-player nonzero-sum stopping games in discrete time. Ann. Probab. 32, 2733–2764 (2004)MathSciNetCrossRefMATHGoogle Scholar
- 55.Stettner, L.: Zero-sum Markov games with stopping and impuslive strategies. Appl. Math. Optim. 9, 1–24 (1982)MathSciNetCrossRefMATHGoogle Scholar
Copyright information
Open AccessThis article is distributed under the terms of the Creative Commons Attribution 4.0 International License (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, distribution, and reproduction in any medium, provided you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.