Game settings in which each player (or one of the players) by his/her actions can choose a convenient time for stopping the entire game are due to numerous applications, primarily in economics; see, e.g., [9, 15, 17]. The game setting in which the player (players) controls the termination of the process is usually associated with the fundamental paper by Dynkin [2]. In such games, the dynamics is usually assumed to be nondeterministic and can be given by a stochastic differential equation [13, 14], a Lévy–Feller process [12], or a Markov chain with continuous [20] or discrete [7] time.

In the present paper, the dynamics of the game is given by a deterministic conflict-controlled system, and the distribution of the stopping time is not considered to be known to the players in advance as, for example, in [6, 19] and is not considered an absolutely continuous random variable, albeit depending on the players’ actions as in [8]. Each player can try to stop the game only at some discrete times by assigning a conditional stopping probability (from subsets of the compact set \( [0;1] \) known in advance). The sets of time points at which players can initiate the end of the game do not overlap and are also known in advance. The case in which the game will not be stopped is not excluded (see condition (1.4)). The cost functional is a random variable and depends both on the time when the game ends and the state of the system at that time as well as on the player who initiated the early stopping of the game; the task of the players is to optimize the expected value of this random variable. To this end, each player, based on the information he/she has about the realized trajectory of the system, makes decisions both about the conditional probability of the end of the game and about his/her own control of the conflict-controlled system. Here the strategies are actually assumed to be random processes that depend in a nonanticipating way on the realized trajectory of the system. In particular, each player has the right to use both classical deterministic strategies—piecewise continuous programming strategies or piecewise constant positional strategies—and their arbitrary distributions, including strategies with a strategic guide.

The control scheme with a guide was proposed by Krasovskii and Subbotin (see [4]) as a control method resistant to small information noise. A little later, its stochastic version was proposed in [3]. In [5], the stochastic guide was used in the framework of the vanishing slackness method for the Hamilton–Jacobi equations. In [10, 11], a stochastic guide was constructed based on a continuous-time Markov game. In [8], a version of such a stochastic guide was used to construct approximately optimal strategies in a differential game on a finite interval the early stopping of which was a Poisson process with intensity determined by the players.

Assuming the saddle point condition in the small game and the insignificance of infinity (see (1.4)), we show the existence of a game value and propose a procedure based on a stochastic guide that realizes the game value with arbitrary accuracy. The corresponding procedure does not require any knowledge (accumulation of knowledge) about the opponent’s actions or memory about the entire realized trajectory. As part of this procedure, the player changes his/her control \( \bar {u} \) at some times \( t_k \) subject to a Poisson process using the rule

$$ \bar {u}(t)\equiv {u}^{\text {to}\thinspace z(t_k)}\big (t_k,y(t_k)\big )\qquad \forall t\in [t_k;t_{k+1}); $$

here \( y(t_k) \) and \( z(t_k) \) are the positions of the original game and the model game realized at that time, and \( {u}^{\text {to}\thinspace z(t_k)} \) is the control that maximally shifts the position of the original system to the point \( z(t_k) \). The process \( z(\cdot ) \) will be given by some conflict-controlled continuous-time Markov chain in which the fictitious opponent player moves the position of the chain \( z(t) \) as far as possible from the latest available observation of the position of the original system, position \( y (t_k) \), and the fictitious ally player plays optimally. Note that his/her optimal strategy can be constructed exactly by solving a finite system of equations (3.2).

The article is organized as follows. In the first section, we formalize the original differential game and state the main result—the existence of a saddle point and approximately optimal strategies realizing it for each player. In the second section, the necessary constants are selected for each accuracy parameter and a set of states is specified for the model based on a Markov game; in the same section, we state temporary assumptions for the times and probabilities of the end of the game that are convenient for proofs. The third section describes a continuous-time Markov game and its optimal strategies and proves a number of estimates for the divergence of trajectories and values of the functional in this game. In the fourth section, a stochastic guide is constructed and the existence of a game value is proved under the temporary assumptions. In the last section, this result is extended to the general case.

1. ORIGINAL DIFFERENTIAL GAME

Consider the following conflict-controlled system operating in \( \mathbb {R}^d \):

$$ \begin {gathered} \frac {d}{dt}y(t)=f\big (y(t),u(t),v(t)\big ),\quad y(0)=x_*\in \mathbb {R}^d,\\ t\geq 0,\quad u(t)\in {{\mathbb {U}}},\quad v(t)\in {{\mathbb {V}}}. \end {gathered} $$
(1.1)

In what follows, we assume that \( {\mathbb {U}} \) and \( {\mathbb {V}} \) are metric compact sets and the function \( f\!:\!\mathbb {R}^{d}\!\times \! {{\mathbb {U}}}\!\times \! {{\mathbb {V}}}\!\to \!\mathbb {R}^{d} \) is \( L \)-Lipschitz in \( x \) for some constant \( L>1 \) and is bounded by the same constant. We will also require that the saddle point condition in the small game (the Isaacs condition) be satisfied: for all \( x,w\in \mathbb {R}^d \) one has

$$ \min _{u\in {{\mathbb {U}}}}\max _{v\in {{\mathbb {V}}}}\big \langle w,f(x,u,v)\big \rangle =\max _{v\in {{\mathbb {V}}}}\min _{u\in {{\mathbb {U}}}}\big \langle w,f(x,u,v)\big \rangle . $$
(1.2)

In the setting we are considering, the game can end on the initiative of any of the players, and the end time \( \theta _{\min } \) of the game is nondeterministic and determined by the following procedure. Players have infinitely increasing sets of times \( (t^I_k)_{k\in \mathbb {N}} \), \( (t^{II}_k)_{k\in \mathbb {N}} \) and sequences of intervals \( [\phi ^-_k;\phi ^+_k] \) and \( [\psi ^-_k;\psi ^+_k] \) nested in \( [0;1] \). We will assume that the sets of time points are disjoint. Each of the players at their own times ( \( t^{I}_k \) or \( t^{II}_k \), respectively), if the game has not yet been stopped, assigns an element of the corresponding interval, \( \phi _k \) or \( \psi _k \), respectively, as the conditional probability of stopping the game at this time on his/her initiative. In view of the above, the time \( \theta _{\min } \) of the end of the game has the distribution

$$ \begin{aligned} {\mathbb {P}}(\theta _{\min }=t^{I}_k)&=\phi _k\cdot \prod _{i<k}(1-\phi _i)\cdot \prod _{t^{II}_k< t^{I}_k}(1-\psi _i), \\ {\mathbb {P}}(\theta _{\min }=t^{II}_k)&=\psi _k\cdot \prod _{i<k}(1-\psi _i)\cdot \prod _{t^{I}_k< t^{II}_k}(1-\phi _i).\end{aligned} $$

Thus, except for \( \phi _i \) and \( \psi _i \) the players do not have any other instruments of influence upon the time of the termination of the game. In particular, for the same \( \phi _i \) and \( \psi _i \), any probabilities of events about the time of the termination of the game and about its initiating player do not depend on the realized trajectory; for example, if any \( \phi ^-_k \) or \( \psi ^-_k \) are equal to one, then the game always ends at this time; if \( \phi ^+_k \) or \( \psi ^+_k \) are zero, then the game cannot end at this time.

Let us also denote by \( \theta _1 \) a random element equal to the game end time if it were initiated by the first player and equal to \( +\infty \) otherwise; similarly, by \( \theta _2 \) we denote the game end time if it were initiated by the second player and assume \( \theta _2 \) to be equal to \( +\infty \) otherwise. Now we also have \( \theta _{\min }=\min (\theta _1,\theta _2) \) and

$$ \begin {aligned} {\mathbb {P}}(\theta _{1}\geq t^{I}_k)&=\prod _{i<k}(1-\phi _i),\\ {\mathbb {P}}(\theta _{2}\geq t^{II}_k)&=\prod _{i<k}(1-\psi _i). \end {aligned} $$
(1.3)

1.1. Strategies

We will assume that, in addition to public information about the dynamics of system (1.1), the initial position of the game, the possible times \( \theta ^k_{I} \) and \( \theta ^k_{II} \) of the end of the game, and conditional probabilities of the game stopping at these times, the players also have information about the realized trajectory of system (1.1). They have the right to submit both their control to system (1.1) and the conditional probabilities of the game stopping, which depend on the realized trajectory in a nonanticipatory way. Finally, we will allow the corresponding rules to be nondeterministic.

Recall that càdlàg functions are functions continuous on the right and having limit on the left; the set of such functions from the set \( A \) into the set \( B \) will be called Skorokhod’s space (see [1]).

By \( \mathfrak {D}^I \) we denote the set of all measurable mappings \( U \) of \( D(\mathbb {R}_+,\mathbb {R}^d) \) into the space of probability measures over \( D(\mathbb {R}_+,\mathbb {U}) \) such that for any time \( t>0 \) there exists a time \( t^{\prime }>t \) such that for all \( x,x^{\prime }\in D(\mathbb {R}_+,\mathbb {R}^d) \) the equality of the restrictions \( x|_{[0;t)}=x^{\prime }|_{[0;t)} \) implies the equality of the induced probabilities \( U(x) \) and \( U(x^{\prime }) \) on \( D([0;t^{\prime }],\mathbb {U}) \). In a similar way, we define the set \( \mathfrak {D}^{II} \). The elements of the sets \( \mathfrak {D}^{I} \) and \( \mathfrak {D}^{II} \) will be called admissible strategies.

Let us discuss what admissible strategies might be, for example, for the first player. These may turn out to be constant controls (elements of \( \mathbb {U} \)); such controls do not depend on time or on the realized trajectory. We can consider programming controls (controls from \( B(\mathbb {R}_+,\mathbb {U}) \) continuous on the right and having a limit on the left); such controls depend only on time but in no way depend on the already realized trajectory \( y \). Within the framework of the classical formalization of the theory of differential games, positional controls of the following form are usually considered: \( U_t=u(t_i,y(t_i)) \) for each \( t \) in the interval \( [t_i;t_{i+1}) \); moreover, both the mapping \( u:\mathbb {R}_+\times \mathbb {R}^d\to \mathbb {U} \) and the infinitely increasing sequence of times \( t_i \) can be chosen arbitrarily. Note that positional strategies already use information about the realized trajectory; moreover, for each time \( t\in [t_i;t_{i+1}) \), the control is known at least up to the time \( t_{i+1} \) and positional strategies are thus also valid. One can use the control with a guide: the motion of the model \( Z \) is constructed in a nonpredictive way along the real-time trajectory, after which the control is selected by the rule \( U_t=u(t_i,Z(t_i)) \) for all \( t\in [t_i;t_{ i+1}) \); the resulting strategy is also admissible. All of the strategies listed above were deterministic, but if at each time \( t_i \) we choose not just one element from \( \mathbb {U} \), but some distribution of them, then the corresponding rule will specify, generally speaking, a nondeterministic admissible strategy. One way to specify the appropriate distributions is to make the model motion of \( Z \) a random process consistent with some filtering containing the original one and then take a marginal distribution on \( \mathbb {R}^d \) due to this process.

For the stochastic basis, we take Skorokhod’s space \( D(\mathbb {R}_+,\mathbb {R}^d) \) (see [1]) with the Borel sigma-algebra \( \mathcal {B}(D(\mathbb {R}_+,\mathbb {R}^d)) \) and the canonical filtration—the flow of the sigma-algebras \( \mathcal {F}_{t}\!\stackrel {\triangle }{=}\!\sigma (y|_{[0;t]}) \), \( t\!\geq \! 0 \). Let us show that each pair of admissible strategies \( (U,V)\in \mathfrak {D}^I\times \mathfrak {D}^{II} \) determines the unique probability \( {\mathbb {P}}_{U,V} \) on this stochastic basis, thereby determining a random process \( \hat {Y} \) on the solutions of system (1.1).

Note first that at the initial time the process is given, since the position of the process \( \hat {Y} \) is known, and hence the control distributions corresponding to it at this time are also known. Then for some positive \( t^{\prime } \) the induced probabilities \( U(x)\otimes V(x) \) over \( D([0;t^{\prime }],\mathbb {U})\times D([0;t^{\prime }] ,\mathbb {V}) \) are known, since each pair of elements from \( D([0;t^{\prime }],\mathbb {U})\times D([0;t^{\prime }],\mathbb {U}) \) uniquely recovers some solution of system (1.1), the restriction of \( P_{U,V} \) over \( D([0;t^{\prime }],\mathbb {R^d}) \) is uniquely recovered as the measure image by virtue of this mapping, and hence so is the process \( \hat {Y} \) on the interval \( [0;t^{\prime }] \).

Therefore, the set of intervals containing the time \( 0 \) and contained in the semiaxis on which the process \( \hat {Y} \) is defined and uniquely determined is nonempty. If this set contains the semiaxis, then everything has been shown. Assume it does not. Then we find the largest element in our set—the interval \( \mathcal {T} \). Owing to the continuity of the trajectories of system (1.1) and hence of \( \hat {Y}|_{\mathcal {T}} \), we can uniquely reconstruct the trajectories of \( \hat {Y} \) on the closure of \( \mathcal {T} \)—some interval \( [0;t] \). Then there exists a positive time \( t^{\prime \prime } \) such that for each realization of the process \( \hat {Y} \) on \( [0;t] \), the distribution of controls on \( [0;t^{\prime \prime }) \) is known. Knowing these controls for each such implementation, it is possible to construct the distribution of trajectories of system (1.1) up to the time \( t^{\prime \prime } \); thereby the process \( \hat {Y} \) is uniquely reconstructed already on the interval \( [0;t^{\prime \prime }) \) containing both \( \mathcal {T} \) and its closure \( [0;t] \). However, this contradicts the choice of \( \mathcal {T} \) as the largest interval on which the process \( \hat {Y} \) is defined and uniquely determined. Therefore, this process is uniquely determined on the entire semiaxis.

1.2. Stopping Times

Apart from controls of system (1.1), the players also control the probabilities \( \phi _i,\psi _i \) of the termination of the game. To describe the possibilities of the first player admissible with such a choice, we demand that each \( \phi _k \) were an \( \mathcal {F}_{\theta ^I_k}\otimes \mathcal {F}^{\mathbb {U}}_{\theta ^I_k} \)-measurable random variable (on \( {D}(\mathbb {R}_+,\mathbb {R}^m)\times {D}(\mathbb {R}_+,\mathbb {U}) \)) ranging in \( [\phi ^-_k;\phi ^+_k] \), in particular, depending in a nonanticipating way on a pair (trajectory, control) in \( {D}(\mathbb {R}_+,\mathbb {R}^m)\times {D}(\mathbb {R}_+,\mathbb {U}) \).

In particular, they will turn out to be random variables with values in \( [\phi ^-_k;\phi ^+_k] \) that are independent of the trajectory being realized, as analogs of programming controls. Admissible sequences will also be arbitrary probabilities that satisfy the same probability rules and depend on the state of the process at the time \( t^I_{k} \) (or earlier). Finally, nothing prevents us from specifying the probabilities satisfying these rules in terms of the trajectories of some other random process \( Z \) adapted to the filtration containing the original one and responsible for the strategy \( U \).

Note that rule (1.3) associates each such sequence \( \phi _k \) with a distribution (probability) \( \theta _1 \) on \( \{\theta ^I_k\thinspace |\thinspace i\in \mathbb {N}\}\cup \{+\infty \} \). Moreover, in this case, we can assume that there exists some standard probability space \( \Omega ^I \) (for example, the instance \( ([0;1],\mathcal {B}([0;1]),\lambda ) \) was used for this purpose in [8]), and for each element in \( {D}(\mathbb {R}_+,\mathbb {R}^m)\times {D}(\mathbb {R}_+,\mathbb {U}) \) its value \( \theta _1 \) is the distribution of some random variable given in this space and denoted by the same letter for simplicity. We will call such random variables (on \( \Omega ^I \)) depending on \( {D}(\mathbb {R}_+,\mathbb {R}^m)\times {D}(\mathbb {R}_+,\mathbb {U}) \) and ranging in \( \{\theta ^I_k\thinspace |\thinspace i\in \mathbb {N}\}\cup \{+\infty \} \) the admissible times for the first player to exit. Let us place all such random variables into a set \( \mathfrak {Q}^{I} \). Similarly, for the second player we fix a standard probability space \( \Omega ^{II} \) and put each sequence of \( \mathcal {F}_{\theta ^I_k}\otimes \mathcal {F}^{\mathbb {V}}_{\theta ^I_k} \)-measurable random variables with values in \( [\phi ^-_k;\phi ^+_k] \) in correspondence with a random variable \( \theta _2 \) given on \( \Omega ^{II} \) whose distribution obeys (1.3). Such random variables (on \( \Omega ^{II} \)) depending on \( {D}(\mathbb {R}_+,\mathbb {R}^m) \) and ranging in \( \{\theta ^I_k\thinspace |\thinspace i\in \mathbb {N}\}\cup \{+\infty \} \) will be called the admissible times for the second player to exit. These times form a set \( \mathfrak {Q}^{II} \).

Now on the probability space \( \Omega \stackrel {\triangle }{=} \Omega ^I\times \Omega ^{II} \) we can correctly define a random variable \( \theta _{\min }=\min (\theta _1,\theta _2) \) depending on \( {D}(\mathbb {R}_+,\mathbb {R}^m) \times {D}(\mathbb {R}_+,\mathbb {U}) \times {D}(\mathbb {R}_+,\mathbb {V}) \). Moreover, any pair of admissible strategies \( (U,V) \) also uniquely determines the probability \( P_{\Omega ,U,V} \) on \( \Omega ^I\times \Omega ^{II}\times {D}(\mathbb {R}_+,\mathbb {R}^m) \).

1.3. Player’s Objectives

Now let us define the objectives of the players. Let the functions \( \sigma _1:\mathbb {R}_+\times \mathbb {R}^d\to [-1;1] \) and \( \sigma _2:\mathbb {R}_+\times \mathbb {R}^d\to [-1;1] \) be given. We will also assume, increasing \( L \) if needed, that they are \( L \)-Lipschitz in \( x,t \). The number \( \sigma _1(\theta _1,y(\theta _1)) \) will be the payment of the first player to the second if the first player initiated the end of the game at time \( \theta _1 \). The number \( \sigma _2(\theta _2,y(\theta _2)) \) will become the payment of the first player to the second if the second player initiated the end of the game at time \( \theta _2 \).

Now in the case of stopping the game, the objective function is defined at the final time. Since there is, possibly zero, probability that the game will never be stopped, we require at least one of three conditions to be met:

  1. 1.

    The objective functions \( \sigma \) and the controlled system (1.1) are such that \( \sigma _1(t,y(t)) \) and \( \sigma _2(t,y(t)) \) tend to a common limit as \( t\to \infty \) with the convergence being uniform over all possible trajectories \( y \) of system (1.1). In other words, for any positive \( \varepsilon \) there exists a positive integer \( N \) such that for all \( t>N \) and all trajectories \( y \) of system (1.1) one has

    $$ \Big |\sigma _1\big (t,y(t)\big )-\sigma _2\big (t,y(t)\big )\Big |+ \Big |\sigma _1\big (N,y(N)\big )-\sigma _1\big (t,y(t)\big )\Big |<{\varepsilon }. $$
  2. 2.

    At least one of the tuples \( (\phi ^-_k)_{k\in \mathbb {N}} \) and \( (\psi ^-_k)_{k\in \mathbb {N}} \) contains one.

  3. 3.

    At least one of the series \( \sum \limits _{k\in \mathbb {N}}\phi ^-_k \) and \( \sum \limits _{k\in \mathbb {N}}\psi ^-_k \) is diverging.

Both the second and the third conditions guarantee that with probability \( 1 \), regardless of the choice of admissible strategies \( U,V \) and terminal times \( \theta _1,\theta _2 \) of players, the game ends in finite time. Moreover, by virtue of the validity of any of the three conditions above, there exists a sufficiently large positive integer \( N_{\varepsilon } \) such that for each \( i\in \{1,2\} \), for any admissible strategies \( {U,V} \) and terminal times \( \theta _1,\theta _2 \) of the players, the end time of the game \( \theta _{\min }=\min (\theta _1,\theta _2) \) satisfies

$$ \begin {aligned} &{\mathbb {P}}_{\Omega ,U,V}(\theta _{\min }\geq N_{\varepsilon })\sup _{t\geq N_{\varepsilon }}\mathbb {E}_{U,V}\Big (\sigma _i\big (\theta _{\min },y(\theta _{\min })\big )\thinspace \big |\thinspace \theta _{\min }\geq t\Big )-{\varepsilon }\\ &\qquad \qquad {}<{\mathbb {P}}_{\Omega ,U,V}(\theta _{\min }\geq N_{\varepsilon })\thinspace \mathbb {E}_{\Omega ,U,V} \big (J(y,\theta _1,\theta _2)\thinspace \big |\thinspace \theta _{\min }\geq N_{\varepsilon }\big ) \\ &\qquad \qquad {}<{\mathbb {P}}_{\Omega ,U,V}(\theta _{\min }\geq N_{\varepsilon })\inf _{t\geq N_{\varepsilon }}\mathbb {E}_{\Omega ,U,V}\Big (\sigma _i\big (\theta _{\min },y(\theta _{\min })\big )\thinspace \big |\thinspace \theta _{\min }\geq t\Big )+{\varepsilon }. \end {aligned} $$
(1.4)

Indeed, if the first condition in this formula is satisfied, then all expectations differ little from each other, the second condition nullifies the probabilities in this formula with sufficiently large \( N_{\varepsilon } \), and the third condition allows at least one of the expressions

$$ -\sum _{t^{II}_k<N_{\varepsilon }}\ln (1-\phi ^-_k),\quad \text {or}\quad -\sum _{t^{II}_k<N_{\varepsilon }}\ln (1-\psi ^-_k) $$

to be made arbitrarily large, thus making the probabilities in this formula tend to zero.

Now, if the game has not been stopped in finite time and the probability of such an event is positive, it is correct to write

$$ \sigma _1\big (+\infty ,y(+\infty )\big )=\sigma _2\big (+\infty ,y(+\infty )\big ), $$

and hence the payment of the first player to the second is also defined in this case, \( \sigma _1(\theta _1,y(\theta _1))=\sigma _2(\theta _2,y(\theta _2)) \). However, then, subject to any of the three conditions above, the payment of the first player to the second is equal almost everywhere to

$$ J(y,\theta _1,\theta _2)= \begin {cases} \sigma _1(\theta _1,y(\theta _1)),& \theta _1=\theta _{\min } \\ \sigma _2(\theta _2,y(\theta _2)),& \theta _2=\theta _{\min }. \end {cases} $$

1.4. Game Value

Assume that the players have chosen some admissible strategies \( U \) and \( V \). As has been shown above, one can unambiguously recover the distribution of the process \( \hat {Y} \) on solutions of (1.1); thus, on the set of all possible solutions of (1.1) one can unambiguously reconstruct the probability \( {\mathbb {P}}_{U,V} \); in turn, this defines the corresponding expectation \( \mathbb {E}_{U,V} \). Now, with each choice of admissible terminal times \( \theta _1,\theta _2 \) one can consider the distribution of the random variable \( J(y,\theta _1,\theta _2) \). We assume that by their actions the players try to optimize its expectation, the number

$$ \mathbb {E}_{\Omega ,U,V} J(y,\theta _1,\theta _2). $$
(1.5)

Depending on which of the players is informationally discriminated, we obtain the following two game values:

$$ \begin{aligned} \mathcal {V}^+\thinspace &{}\stackrel {\triangle }{=}\thinspace \inf _{U\in \mathfrak {D}^I,\thinspace \theta _1\in \mathfrak {Q}^I}\sup _{ V\in \mathfrak {D}^{II},\thinspace \theta _2\in \mathfrak {Q}^{II}} \mathbb {E}\thinspace _{\Omega ,U,V} J(y,\theta _1,\theta _2),\\ \mathcal {V}^-\thinspace &{}\stackrel {\triangle }{=}\thinspace \sup _{ V\in \mathfrak {D}^{II}, \theta _2\in \mathfrak {Q}^{II}}\inf _{U\in \mathfrak {D}^I, \theta _1\in \mathfrak {Q}^I} \mathbb {E}\thinspace _{\Omega ,U,V} J(y,\theta _1,\theta _2).\end{aligned} $$

The main purpose of this paper is to prove the following assertion.

Theorem 1.1.

Under condition (1.4), one has the equality \( \mathcal {V}^-= \mathcal {V}^+ \). Moreover, for each positive \( \varepsilon \) there exist players’ admissible strategies \( U^{\varepsilon }\in \mathfrak {D}^I \) and \( V^{\varepsilon }\in \mathfrak {D}^{II} \) and terminal times \( \theta ^{\varepsilon }_1\in \mathfrak {Q}^{I} \), \( \theta ^{\varepsilon }_2\in \mathfrak {Q}^{II} \) such that

$$ -{\varepsilon }+\sup _{V\in \mathfrak {D}^{II}, \theta _2\in \mathfrak {Q}^{II}} \mathbb {E}_{\Omega ,U^{\varepsilon },V} J(y,\theta ^{\varepsilon }_1,\theta _2) \leq \mathcal {V}^+=\mathcal {V}^-\leq {\varepsilon }+\inf _{U\in \mathfrak {D}^{I}, \theta _1\in \mathfrak {Q}^{I}} \mathbb {E}_{\Omega ,U,V^{\varepsilon }} J(y,\theta _1,\theta ^{\varepsilon }_2). $$

Let us describe the scheme of proof. Since \( \mathcal {V}^-\leq \mathcal {V}^+ \), from the symmetry of the players, for each positive \( \varepsilon \) it is required to describe necessary admissible strategies \( U_{\varepsilon } \) only for the first player and a terminal time \( \theta ^{\varepsilon }_1 \) satisfying

$$ -{\varepsilon }+\sup _{V\in \mathfrak {D}^{II}, \theta _2\in \mathfrak {Q}^{II}} \mathbb {E}_{\Omega ,U^{\varepsilon },V} J(y,\theta ^{\varepsilon }_1,\theta _2) \leq \mathcal {V}^+. $$
(1.6)

To this end, we first make temporary assumptions about admissible terminal times of the players convenient for the proof and introduce the necessary constants, in particular, by passing from \( t^{I}_k,t^{II}_k \) to \( t_n=nh \); then we consider a continuous-time Markov game as a model game, indicate a series of estimates for it, and construct optimal strategies in this game. On the basis of these optimal strategies, we will further describe the behavior of the guide model and, based on it, the construction of the strategy for the original game, show the mean-square convergence of the corresponding trajectories, and hence the values, on average, of the objective function. In the last section, we remove the temporary assumptions on admissible terminal times for the players.

2. TEMPORARY ASSUMPTIONS AND INTRODUCTION OF CONSTANTS

Let us take some positive number \( {\varepsilon }<\min (1/16L\pi ,6/\sqrt {d}) \), and, together with it, also some positive integer \( N_{{\varepsilon }/12} \) (see (1.4)). Then one can also find some subinterval \( [T_-;T_+)\subset [N_{{\varepsilon }/12};+\infty ) \) that does not contain elements of unboundedly increasing sequences \( (t^{I}_k)_{k\in \mathbb {N}} \) and \( (t^{II}_k)_{k\in \mathbb {N}} \). Since there are finitely many elements of such sequences in \( [0;T_+) \), there exists a positive \( h<\frac {T_+-T_-}{2} \) such that, for all positive integers \( k \) and \( l \), the inequality \( t^{I}_k, t^{II}_l\leq T_+ \) implies that

$$ t^{I}_k>2h, \quad t^{II}_l>2h,\quad |t^{I}_k-t^{II}_l|>2h,\quad t^{I}_k/h\notin \mathbb {Z},\quad t^{II}_l/h\notin \mathbb {Z}. $$

In this case, decreasing \( h \) if necessary, if \( t^{I}_k\leq T_+ \) and \( \phi ^-_{k}>1-h \), then one can also guarantee that \( \phi ^-_{k}=1 \). Moreover, the same property can also be ensured for pairs \( (t^{II}_k,\psi ^+_{k}) \).

Take the smallest element of \( h\mathbb {Z} \) greater than \( T_- \) for \( T \). Now \( T\in [T_-;T_+)\cap h\mathbb {Z} \), and the interval \( (T-h;T+h) \) does not contain elements of the sequences \( (t^{I}_k)_{k\ in\mathbb {N}} \) and \( (t^{II}_k)_{k\in \mathbb {N}} \).

Now let us make the following temporary assumptions about \( T \) and \( h \) as well as the sequences \( (t^{I}_k)_{k\in \mathbb {N}} \), \( (t^{II}_k)_{k\in \mathbb {N}} \) and the corresponding intervals \( [\phi ^-_k;\phi ^+_k] \), \( [\psi ^-_k;\psi ^+_k] \):

  1. 1.

    For all \( k\in \mathbb {N} \), \( t^I_k\leq T \) implies \( \phi ^+_k<1-h \) and \( t^{II}_k\leq T \) implies \( \psi ^+_k<1-h \).

  2. 2.

    All elements of sequences \( (t^{I}_k)_{k\in \mathbb {N}} \) and \( (t^{II}_k)_{k\in \mathbb {N}} \) not exceeding \( T \) lie in \( h\mathbb {N} \).

The first assumption guarantees that the probability of not stopping the game by time \( T \) is separated from zero by a constant independent of the actions of the players. The second assumption means that all possible times of the end of the game up to time \( T \) inclusive lie in some lattice.

In addition to the temporary assumptions above, taking \( h \) equal to \( h/k \) if necessary for sufficiently large \( k\in \mathbb {N} \), we assume here and in the sequel that the following inequality is guaranteed everywhere:

$$ h<\min \left ( \frac {{\varepsilon }^2 e^{-{8L}T}}{288L^2d},\thinspace \frac {{\varepsilon }}{12T(1+16^{1/3}T^{-2/3})^2},\thinspace \frac {{\varepsilon }}{6(L+4)} \right ); $$

since \( L>1 \) and \( {\varepsilon }<\min (1/16\pi L,6/\sqrt {d})<192(\sqrt [3]{2}-1) \), we see that this also guarantees that

$$ 6h(L+4)<{\varepsilon },\quad h<\frac {1}{32\pi L(L+1)}<\frac {1}{L(2L/3+5)},\quad Lh\sqrt {d}<1, $$
(2.1)
$$ 3h< 32\pi h(hL\sqrt {d}+L^2)<1,\quad Le^{{4L}T}\sqrt {2{d}}\sqrt {h}<{\varepsilon }/12, $$
(2.2)
$$ h<\sqrt [3]{32T}-\sqrt [3]{16T},\quad hT(1+16^{1/3}T^{-2/3})^2 <{\varepsilon }/12. $$
(2.3)

By construction, for any positive integer \( n \) there exists at most one element of sequences \( (t^I_k)_{k\in \mathbb {N}} \) and \( (t^{II}_k)_{k\in \mathbb {N}} \) on the interval \( [nh;nh+h) \). If there exists some \( t^I_k \) on this interval, then we assign \( \Phi _n\stackrel {\triangle }{=}[\phi ^-_k/h;\phi ^+_k/h] \); otherwise, we take \( \Phi _n\stackrel {\triangle }{=} \{0\} \). If this interval also contains some \( t^{II}_k \), then we assign \( \Psi _n\stackrel {\triangle }{=} [\psi ^-_k;\psi ^+_k] \), otherwise, we set \( \Psi _n\stackrel {\triangle }{=}\{0\} \).

Note that, owing to the boundedness of the dynamics of system (1.1), there exists a compact set \( K_<\subset \mathbb {R}^d \) beyond which no solution \( y \) of Eq. (1.1) with the initial condition \( y(0)=x_* \) can go. Extending this compact set up to \( K_> \) if necessary, we can assume that any motion in (1.1) at any time in \( [0;T] \) is separated from the boundaries of \( K_{>} \) by at least a distance \( L \).

Fix some infinitely smooth monotone nonincreasing scalar function \( a:\mathbb {R}\to [0;1] \) so that \( a(0)=1 \), \( a(L)=0 \), and its Lipschitz constant does not exceed \( 1 \). Let us also introduce

$$ \hat {f}(x,{u},{v})\stackrel {\triangle }{=} a(\mathrm {dist}\thinspace (x;K_<))f(x,{u},{v}) $$

for all \( (x,u,v)\in \mathbb {R}^{d}\times \mathbb {U}\times \mathbb {V} \). Remaining continuous everywhere, this function is equal to zero outside \( {K}_> \). Now, following \( f \), the function \( \hat {f}|_{K_{>}\times {{\mathbb {U}}}\times {{\mathbb {V}}}} \) is Lipschitz in \( x \) with the constant \( 2L \) and its norm is also at most \( L \).

Only now we finally take \( \mathcal {Z}\stackrel {\triangle }{=} h((\mathbb {N}\cup \{0\})\times \mathbb {Z}^{d+1}) \subset \mathbb {R}^{d+2} \) and \( \mathcal {Z}_<\stackrel {\triangle }{=}\mathcal {Z} \cap (\mathbb {R}\times \mathbb {K}_> \times ((-\infty ;-1]\cup [1;+\infty )\cup \{0\}) \subset \mathbb {R}^{d+2} \).

Denote the basis in \( \mathbb {R}^{d+2} \) by \( (\pi _0,\pi _1,\dots ,\pi _d,\pi _{d+1}) \). For brevity, we write the coordinates of a vector \( z\in \mathbb {R}^{d+2} \) in the form

$$ (\pi _0z,\pi _1z,\dots ,\pi _d z,\pi _{d+1}z). $$

In this case, the zero coordinate of each point will correspond to time, and the last one is intended to fix the game stopping time. Since the coordinates \( (\pi _1z,\dots ,\pi _d z) \) track the trajectory of the original system, we will assume that \( \mathbb {R}^{d} \) coincides with the subset

$$ \big \{z\in \mathbb {R}^{d+2}\thinspace |\thinspace \pi _{0}z=\pi _{d+1}z=0\big \}\subset \mathbb {R}^{d+2}; $$

consequently, we can subtract elements of \( \mathbb {R}^{d} \) and \( \mathbb {R}^{d+2} \) from each other to obtain an element of \( \mathbb {R}^{d+2} \). Moreover, any element \( z\in \mathbb {R}^{d+2} \) is projected onto \( \mathbb {R}^{d} \) according to the rule

$$ z\mapsto z-(\pi _0 z)\pi _0-(\pi _{d+1} z)\pi _{d+1}. $$

Then we can apply the inner product and norm on \( \mathbb {R}^d \) to vectors in \( \mathbb {R}^{d+2} \), denoting them by \( \langle \cdot ,\cdot \rangle _d \) and \( \|\cdot \|_d \), respectively.

3. MODEL GAME

For a guide we will use the trajectories of a specially selected model game—a conflict-controlled continuous-time Markov chain with the phase space \( \mathcal {Z}_< \).

3.1. Randomized Strategies in the Markov Game

For the players’ strategies we take time-invariant randomized strategies.

By a time-invariant randomized strategy \( \bar \mu \) of the first player we understand a pair \( ({\mu },\phi _{\mu }) \) of mappings that take each \( w=(nh,x,s)\in \mathcal {Z}_< \) to some probability measure \( {\mu }[w] \) on \( \mathbb {U} \) and some number \( {\phi }_{\mu }[w]\in \Phi _n \). By a randomized strategy \( \bar \nu \) of the second player we mean a pair \( ({\nu },\psi _{\nu }) \) of mappings that take each \( w=(nh,x,s)\in \mathcal {Z}_< \) to some probability measure \( {\nu }[w] \) on \( {\mathbb {V}} \) and some number \( {\psi }_{\nu }[w]\in \Psi _n \). Denote the set of all time-invariant strategies of the first and second players by \( \check {\mathbb {U}}_\varsigma \) and \( \check {\mathbb {V}}_\varsigma \). By \( {\mathbb {U}}_\varsigma \) and \( {\mathbb {V}}_\varsigma \) we denote their projections—the sets of mappings \( w\mapsto {\mu }[w] \) and \( w\mapsto {\nu }[w] \), respectively.

Define a mapping \( \check {f}:\mathcal {Z}_<\times {\mathbb {U}}_\varsigma \times {\mathbb {V}}_\varsigma \to \mathbb {R}^{d} \) by the following rule: for every pair of strategies \( ({\mu },{\nu })\in {\mathbb {U}}_\varsigma \times {\mathbb {V}}_\varsigma \) and a point \( w=(t,x,s)\in \mathcal {Z}_< \),

$$ \check {f}(w,\mu ,\nu )\stackrel {\triangle }{=}1_{\{0\}}(s)\int _{\mathbb {U}}\int _{{V}}\hat {f}(x,u,{v})\thinspace \mu [w](du)\nu [w](dv). $$

For each \( i\in [1\thinspace :\thinspace d] \), we define its projections \( \pi ^+_i\check {f} \), \( \pi ^-_i\check {f} \) by the rules

$$ \begin{aligned} \pi ^+_i\check {f}(w,\mu ,\nu )\thinspace &{}\stackrel {\triangle }{=}\thinspace 1_{\{0\}}(s)\int _{\mathbb {U}}\int _{{V}}\max \big (0,\pi _i\hat {f}(x,u,{v})\big )\thinspace \mu [w](du)\nu [w](dv) ,\\ \pi ^-_i\check {f}(w,\mu ,\nu )\thinspace &{}\stackrel {\triangle }{=}\thinspace 1_{\{0\}}(s)\int _{\mathbb {U}}\int _{{V}}\min \big (0,\pi _i\hat {f}(x,u,{v})\big )\thinspace \mu [w](du)\nu [w](dv).\end{aligned} $$

We will also need time-dependent strategies—mappings in \( D(\mathbb {R}_+,\check {\mathbb {U}}_\varsigma ) \) and \( D(\mathbb {R}_+,\check {\mathbb {V}}_\varsigma ) \). Denote their sets by \( \check {\mathbb {U}}_\varpi \) and \( \check {\mathbb {V}}_\varpi \), respectively. The mappings \( \check {f} \), \( \pi ^+_i\check {f} \), and \( \pi ^-_i\check {f} \) are continued to these sets using the same formulas.

3.2. Dynamics of the Markov Game

For all points \( w=(t,x,s)\in \mathcal {Z}_< \) and time-invariant randomized strategies \( \bar \mu ,\bar \nu \) of the players, we define the Lévy measure \( \check \eta (w,\bar \mu ,\bar \nu ;\cdot ) \) by setting \( \check \eta (w,\bar {\mu },\bar {\nu };A) \) for each subset \( A\subset \mathcal {Z}_< \) to be equal to

$$ \begin{aligned} &{}\frac {1}{h}\delta _{\pi _{0}}(A)+\frac {1}{h} \sum _{i=1}^{d}\Big [\pi _i^+ \check {f}(w,\mu ,\nu )\delta _{h\pi _i}(A)+\pi _i^-\check {f}(w,\mu ,\nu )\delta _{-h\pi _i}(A)\Big ]\\ &\qquad \qquad \qquad \qquad {}+ \frac {1_{\{0\}}(\pi _{d+1} w)}{h}\left [\frac {{\phi }_\nu [w]}{1-{\phi }_\nu [w]}\delta _{(s+1)\pi _{d+1}}(A)+ \frac {{\psi }_\nu [w]}{1-{\psi }_\nu [w]}\delta _{-(s+1)\pi _{d+1}}(A)\right ].\end{aligned} $$

Note that with fixed strategies \( \mu \) and \( \nu \) the mappings

$$ \begin{aligned} w&{}\mapsto \check {f}(w,\mu ,\nu ), \\ w&{}\mapsto \big (\pi ^+_i\check {f}(w,\mu ,\nu ), \pi ^-_i\check {f}(w,\mu ,\nu )\big )\end{aligned} $$

inherit the \( 2L \)-Lipschitz continuity in \( x \) and the boundedness of the norm by the constant \( L \) from the function \( f \). We have also thereby shown the following inequalities: for all \( w\in \mathcal {Z}_< \),

$$ \begin {gathered} \left \|\thinspace \thinspace \int _{\mathbb {R}^{d+2}} \sum _{i=1}^{d}\pi _iy\thinspace \eta (w,\bar {\mu },\bar {\nu };dy)\right \| \leq L,\\ \sum _{i=1}^{d}\Big |\pi _i^+ \check {f}(w,\mu ,\nu )+\pi _i^-\check {f}(w,\mu ,\nu )\Big |\leq L\sqrt {d}. \end {gathered} $$
(3.1)

For all randomized strategies \( \bar \mu ,\bar \nu \) for such a measure \( \check {\eta } \), we consider the continuous-time Markov chain corresponding to the Kolmogorov matrix \( (\bar {Q}_{wy}(\bar \mu ,\bar \nu ))_{w,y\in \mathcal {Z}_<} \) introduced by the rule

$$ \begin {cases} \displaystyle \frac {1}{h}, & y=w+h\pi _0 \\[.7em] \displaystyle \frac {1}{h}\pi ^+_i\check {f}(w,\mu ,\nu ), & y=w+h\pi _i,\quad i\in [1:d] \\[.7em] \displaystyle \frac {1}{h}\pi ^-_i\check {f}(w,\mu ,\nu ), & y=w-h\pi _i,\quad i\in [1:d] \\[.7em] \displaystyle \frac {1_{\{0\}}(s)}{h}\cdot \frac {{\phi }_\nu [w]}{1-{\phi }_\nu [w]}, & y=w+(1+\pi _0w)\pi _{d+1} \\[.7em] \displaystyle \frac {1_{\{0\}}(s)}{h}\cdot \frac {{\psi }_\nu [w]}{1-{\psi }_\nu [w]}, & y=w-(1+\pi _0w)\pi _{d+1} \\[.7em] -\displaystyle \frac {1}{h} -\frac {1}{h}\sum _{i=1}^{d}\big (\pi ^+_i\check {f}(w,\mu ,\nu )+\pi ^-_i\check {f}(w,\mu ,\nu )\big )& \\[.7em] \qquad \qquad {}-\displaystyle \frac {1_{\{0\}}(s)}{h}\left (\frac {{\phi }_\nu [w]}{1-{\phi }_\nu [w]}+\frac {{\psi }_\nu [w]}{1-{\psi }_\nu [w]}\right ), & y=w \\ 0, & \text {otherwise}. \end {cases} $$

Remark 3.1.

By the construction of the Markov chain, each of its stopping times

$$ \tau _k\stackrel {\triangle }{=}\big \{t\thinspace |\thinspace \pi _0\check {Y}(t)=kh\big \} $$

is finite almost everywhere; moreover, being the sum of random variables collectively independent and distributed exponentially with parameter \( 1/h \), each of them does not depend in any way on the players’ actions, in particular, on the past and/or future values of \( \mu \), \( \nu \), \( \phi \), and \( \psi \).

Remark 3.2.

Note also that, by the construction of the Markov chain, the probability of jumping along the last coordinate (in \( \pi _{d+1}\check {Y} \)) between the stopping times \( \tau _{k-1} \) and \( \tau _k \) for intensities \( (\phi ,\psi ) \) coincides with the probability of ending the original game at time \( t_k=kh \) for the same intensities of the players.

3.3. Game Value and Optimal Strategies in the Markov Game

Since the matrices \( \bar {Q}^h_{wy} \), as well as the lengths of the jumps, are uniformly bounded, all the assumptions in [16, Remark 4.2(b)] are satisfied. Then, as shown in [16, Proposition 3.1(a)], for each pair of time-depending randomized strategies \( (\bar {\mu },\bar {\nu })\in \check {\mathbb {U}}_\varpi \times \check {\mathbb {V}}_\varpi \) and initial conditions \( z_0 \) (at time \( 0 \)) there exists a process \( (\check {Y}(t))_{t\geq 0} \) generated by them and hence also the probability \( \check {{\mathbb {P}}}_{\bar {\mu },\bar {\nu }}[z_0] \) on all possible càdlàg mappings of \( \mathbb {R}_+ \) into \( \mathcal {Z}_< \).

Now we introduce the current payment of this Markov game. For every \( w=(t,x,s)\in \mathbb {R}^{d+1}\times ([-1-T;-1] \cup \{0\}\cup [1;1+T]) \), we define

$$ \check {\sigma }(w)= \begin {cases} \frac {1}{2}\big (\sigma _1(T,x)+\sigma _2(T,x)\big ),& s=0\\ \sigma _1(s-1,x),& s\geq 1\\ \sigma _2(-s-1,x),& s\leq -1. \end {cases} $$

Just as \( \sigma _i \), this function is bounded by one and is \( L \)-Lipschitz. For each initial position \( z_0\in \mathcal {Z}_< \) at time \( 0 \), the players can ensure one of the values depending on which one of them is informationally discriminated,

$$ \begin{aligned} \check {\mathcal {V}}^+(z_0)&=\inf _{\bar {\mu }\in \check {\mathbb {U}}_\varpi }\sup _{\bar {\nu }\in \check {\mathbb {V}}_\varpi }\check {\mathbb {E}}_{\bar {\mu },\bar {\nu }}[z_0]\int _{0}^\infty he^{-ht} \check {\sigma }\big (z(t)\big )\thinspace dt,\\ \check {\mathcal {V}}^-(z_0)&=\sup _{\bar {\nu }\in \check {\mathbb {V}}_\varpi }\inf _{\bar {\mu }\in \check {\mathbb {U}}_\varpi }\check {\mathbb {E}}_{\bar {\mu },\bar {\nu }}[z_0]\int _{0}^\infty he^{-ht} \check {\sigma }\big (z(t)\big )\thinspace dt.\end{aligned} $$

Now, according to [16, Theorem 5.1] and [21, Theorem 2], the system of equations (see, [16, (5.4)] and [21, (11)])

$$ \begin {aligned} \inf _{\bar \mu [w]}\sup _{\bar \nu [w]}\sum _{y\in \mathcal {Z}_<}\bar {Q}_{wy}(\bar \mu ,\bar \nu )\check {\mathcal {V}}(y)&= h\big (\check {\mathcal {V}}(w)-e^{-h\pi _0w} \check {\sigma }(w)\big ) \\ &=\sup _{\bar \nu [w]}\inf _{\bar \mu [w]} \sum _{y\in \mathcal {Z}_<}\bar {Q}_{wy}(\bar \mu ,\bar \nu )\check {\mathcal {V}}(y), \quad w\in \mathcal {Z}_<, \end {aligned} $$
(3.2)

has a unique solution by virtue of the saddle point condition in the small game, and this solution coincides with \( \check {\mathcal {V}}\stackrel {\triangle }{=}\check {\mathcal {V}}^-=\check {\mathcal {V}}^+ \). Moreover, any time-invariant strategies \( \bar {\mu }^\mathrm {opt} \) and \( \bar {\nu }^\mathrm {opt} \) of the players that realize a saddle point in this system for all \( w\in \mathcal {Z}_< \) are optimal strategies in this problem (see, e.g., [21, (6)] and [16, (5.5)–(5.7)]). Let us choose some \( \bar {\mu }^\mathrm {opt} \) and \( \bar {\nu }^\mathrm {opt} \) with such a property.

3.4. Estimates for the Objective Functional in the Markov Game

Note first that, just as in [8, (4.4)], one can show that the inequality

$$ \check {\mathbb {E}}\max _{t\in [0;T+r],t/h\in \mathbb {N}}\big |\pi _0\check {Y}(t)-t\big |^2\leq 4\check {\mathbb {E}}\big |\pi _0\check {Y}(T+r)-T-r\big |^2=8h(T+r) $$

holds for all positive integer \( r/h \). Now, by virtue of the Markov inequality, the probability of the maximum value of \( |\pi _0\check {Y}(t)-t|^2 \) being greater than \( r^2 \) on this interval is at most \( 8h(T+r)/r^2 \). Then the probability of the stopping time \( \tau _{T/h} \) corresponding to time \( T \) to be greater than \( T+r \) is at most \( 8h(T+r)/r^2 \).

Consider an arbitrary trajectory \( \check {Y} \) of the Markov game. Note that by construction, beginning from the stopping time \( \tau _{T/h} \), this trajectory changes only in the zero coordinate; in particular, \( \check {\sigma }(\check {Y}(t)) =\check {\sigma }(\check {Y}(\tau _{T/h})) \) for all \( t\geq \tau _{T/h} \). By virtue of the boundedness of \( \check {\sigma } \), for all trajectories \( \check {Y} \) it follows from \( \tau _{T/h}\leq T+r \) that

$$ \left |\thinspace \int _{0}^\infty he^{-ht} \check {\sigma }\big (\check {Y}(t)\big )\thinspace dt- \check {\sigma }\big (\check {Y}(\tau _{T/h})\big )\right |\leq h \int _{0}^{\tau _{T/h}}e^{-hT}\thinspace dt\leq h(T+r), $$

and, in the case of \( \tau _{(T+r)/h}>T+r \), this modulus is estimated from above by the number \( 2 \). Since the probability of the event \( \tau _{T/h}>T+r \), as noted above, is at most \( 8h(T+r)/r^2 \), we have shown that

$$ \check {\mathbb {E}}\left |\thinspace \int _{0}^\infty he^{-ht} \check {\sigma }\big (\check {Y}(t)\big )\thinspace dt- \check {\sigma }\big (\check {Y}(\tau _{T/h})\big )\right | \leq h(T+r)(1+16/r^2). $$
(3.3)

The right-hand side of (3.3) decreases with respect to \( r \) for \( r\leq \sqrt [3]{32T} \). Take a positive integer in the interval \( [\sqrt [3]{16T}/h;\sqrt [3]{16T}/h+1) \) for \( r/h \); by virtue of (2.3), it is ensured that \( r\leq \sqrt [3]{32T} \). Now, estimating the right-hand side in (3.3) by

$$ h(T+r)(1+16/r^2)\leq hT\left (1+\sqrt [3]{16T}/T\right )^2\stackrel {(2.3))}{\leq }{\varepsilon }/12, $$

we obtain an estimate independent of the actions of the players,

$$ \check {\mathbb {E}} \left |\thinspace \int _{0}^\infty he^{-ht} \check {\sigma }\big (\check {Y}(t)\big )\thinspace dt- \check {\sigma }\big (\check {Y}(\tau _{T/h})\big )\right | \leq {\varepsilon }/12. $$
(3.4)

3.5. Estimates on a Trajectory of the Markov Game

Fix some pair of randomized strategies \( \bar {\mu }=(\mu ,\phi _\mu ), \) \( \bar {\nu }=(\nu ,\psi _\nu ) \); hence the distribution \( \check {{\mathbb {P}}}\stackrel {\triangle }{=} \check {{\mathbb {P}}}_{\bar {\mu },\bar {\nu }}[z_0] \) and the Lévy measure \( {\eta }(z;\cdot )\stackrel {\triangle }{=}\check {\eta } (z,\bar {\mu },\bar {\nu };\cdot ) \). Such a Lévy measure is associated with the Lévy–Khintchine generator (see, e.g., [18, (5.1)] and [22, (2.14)]) that takes each function \( g\in C^2_c(\mathcal {Z}_<) \) to a mapping \( x\mapsto \Lambda g(x) \) according to the rule

$$ \check {\Lambda } g(x)=\int _{\mathbb {X}} \big [g(x+y)-g(x)\big ]\eta (x;dy)\qquad \forall x\in \mathcal {Z}_<. $$
(3.5)

In this case, one has Dynkin’s formula [22, Proposition 2.3]

$$ \check {\mathbb {E}} \left (g\big (\hat {Y}(t^{\prime \prime })\big )-\int _{t^{\prime }}^{t^{\prime \prime }} \Lambda g\big (\check {Y}(s)\big )\thinspace ds\thinspace \Bigg |\thinspace \check {Y}(t^{\prime })\right )= g\big (\check {Y}(t^{\prime })\big ) $$
(3.6)

for all \( g\in C^2_c(\mathcal {Z}_<) \) and at all times \( t^{\prime } \) and \( t^{\prime \prime } \) ( \( t^{\prime \prime }\geq t^{\prime } \)). Then the following process is a martingale:

$$ g\big (\check {Y}(t^{\prime \prime })\big )-\int _{t^{\prime }}^{t^{\prime \prime }} \Lambda g\big (\check {Y}(s)\big )\thinspace ds. $$

Consider the behavior of the Markov chain if \( t^{\prime }=0 \) and the initial state \( w^{\prime }\stackrel {\triangle }{=}\check {Y}(0) \) is known. Assume that \( M\stackrel {\triangle }{=}\sqrt {hL\sqrt {d}+L^2} \). Applying the function

$$ w\mapsto \|w-w^{\prime }\|^2_d\stackrel {\triangle }{=}\sum _{i=1}^d (\pi _iw-\pi _iw^{\prime })^2\qquad \forall w\in \mathbb {R}^{d+2} $$

to the above martingale as \( g \), we can find from the estimates (3.1) (see [8, (4.14)]) that

$$ \check {\mathbb {E}}\thinspace \big \|\check {Y}(t)-\check {Y}(0)\big \|^2_d\leq hL\sqrt {d}t+\frac {4LM}{3}({e^{t}-1})^{3/2}. $$

Now, since \( e^{r}-1\leq re^r \) for \( r\geq 0 \), we have \( ({e^{r}-1})^{3/2}\leq \sqrt {r}(e^{3r/2}-e^{r/2}) \) and

$$ \check {\mathbb {E}}\thinspace \big \|\check {Y}(t)-\check {Y}(0)\big \|^2_d\leq hL\sqrt {d}t+\frac {4LM\sqrt {t}}{3}(e^{3t/2}-e^{t/2}). $$

In particular, for the stopping time \( \tau =\inf \{t\geq 0\thinspace |\thinspace \pi _0 \check {Y}(t)\neq \pi _0 \check {Y}(0)\} \), by virtue of its density being exponential with indicator \( 1/h \) and in view of the equality \( \int \nolimits _0^\infty \sqrt {r}e^{-\nu r}dr=\sqrt {\frac {\pi }{4\nu ^3}} \), we have

$$ \begin{aligned} \check {\mathbb {E}}\thinspace \big \|\check {Y}(\tau )-\check {Y}(0)\big \|^2_d &\leq hL\sqrt {d}\thinspace \check {\mathbb {E}} \tau +\frac {4LM}{3}\thinspace \check {\mathbb {E}}\int _{0}^{\infty }\sqrt {s}(e^{3s/2}-e^{s/2}) e^{-s/h} \thinspace ds \\ &= h^2L\sqrt {d}+\frac {2LM\sqrt {\pi }}{3}\big ({(1/h-3/2)^{-3/2}}-{(1/h-1/2)^{-3/2}}\big ) \\ &\leq h^2L\sqrt {d}+{LM}\sqrt {\pi }(1/h-3/2)^{-5/2}\leq h^2L\left (\sqrt {d}+1\right );\end{aligned} $$

at the last step, we have also used the condition \( 3h<2^5M^2\pi h<1 \) (see (2.2)). In a similar way, the equality \( \int \nolimits _0^\infty {r}^{3/2}e^{-\nu r}dr=\frac {3}{4}\sqrt {\frac {\pi }{\nu ^5}} \) also implies

$$ \begin{aligned} \check {\mathbb {E}} \int _{0}^{\tau } \big \|\check {Y}(t)-\check {Y}(0)\big \|^2_d\thinspace dt &\leq \int _{0}^{\infty } \int _{0}^{s}\left [hL\sqrt {d}t+\frac {4LM\sqrt {t}}{3}(e^{3t/2}-e^{t/2}) \right ]\thinspace dt\thinspace e^{-s/h} \thinspace ds \\ &\leq hL\sqrt {d}\int _{0}^{\infty } \frac {s^2}{2} e^{-s/h}\thinspace ds+\frac {4LM}{3} \int _{0}^{\infty }s^{3/2}(e^{3s/2}-e^{s/2}) e^{-s/h} \thinspace dt\thinspace ds \\ &\leq h^3L\sqrt {d}+LM\sqrt {\pi } \big ((1/h-3/2)^{-5/2}-(1/h-1/2)^{-5/2}\big ) \\ &\leq h^3L\sqrt {d}+\frac {5LM\sqrt {\pi }}{2}(1/h-3/2)^{-7/2} \leq h^3L\left (\sqrt {d}+5\right ).\end{aligned} $$

Since the strategy is time-invariant and the considered process is strongly Markov, the same estimates hold if we consider an arbitrary stopping time \( \tau ^{\prime } \) instead of the initial time, take any stopping time \( \tau ^{\prime \prime }\geq \tau ^{\prime } \) not exceeding

$$ \inf \big \{t\geq 0\thinspace |\thinspace \pi _0 \check {Y}(t)\neq \pi _0 \check {Y}(\tau ^{\prime }),t\geq \tau ^{\prime }\big \} $$

for \( \tau \), and finally, replace the expectation by the conditional expectation at the stopping time \( \tau ^{\prime } \). Thus,

$$ \begin {aligned} \check {\mathbb {E}}_{\tau ^{\prime }} \big \|\check {Y}(\tau ^{\prime \prime })-\check {Y}(\tau ^{\prime })\big \|^2_d&\leq h^2L\left (\sqrt {d}+1\right ), \\ \check {\mathbb {E}}_{\tau ^{\prime }} \int _{\tau ^{\prime }}^{\tau ^{\prime \prime }} \big \|\check {Y}(t)-\check {Y}(\tau ^{\prime })\big \|^2_d\thinspace dt&\leq h^3L\left (\sqrt {d}+5\right ). \end {aligned} $$
(3.7)

4. GUIDE SCHEME AND DOUBLE GAME

4.1. Aiming

When constructing the scheme with a guide, we need to be able to aim in a given direction (and deviate from it). First, we introduce functions necessary for this in a deterministic system. To this end, note that, owing to (1.2), for any vectors \( x,z\in \mathbb {R}^{d} \) there exist controls \( u^{\text {to}\thinspace z}(x)\in \mathbb {U} \) and \( v^{\text {from}\thinspace z}(x)\in \mathbb {V} \) such that

$$ \begin{aligned} \min _{u\in {{\mathbb {U}}}}\max _{v\in {{\mathbb {V}}}}\big \langle x-z,\hat {f}(x,u,v)\big \rangle &=\max _{v\in {{\mathbb {V}}}}\Big \langle x-z,\hat {f}\big (t,x,u^{\text {to}\thinspace z}(x),v\big )\Big \rangle \\ &= \max _{v\in {{\mathbb {V}}}}\min _{u\in {{\mathbb {U}}}}\big \langle x-z,\hat {f}(x,u,v)\big \rangle = \min _{u\in {{\mathbb {U}}}}\Big \langle x-z,\hat {f}\big (x,u,v^{\text {from}\thinspace z}(x)\big )\Big \rangle .\end{aligned} $$

For convenience, we also assume that \( u^{\text {to}\thinspace w^{\prime }}\equiv u^{\text {to}\thinspace z^{\prime }} \) for all \( w^{\prime }=(t,z^{\prime },s)\in \mathbb {R}^{m+2} \). Likewise, for each \( x^{\prime }\in \mathbb {R}^{d} \) we equip the second player with a time-invariant randomized strategy \( \bar {\nu }^{\text {from}\thinspace x^{\prime }} \),

$$ \mathcal {Z}_<\ni w \stackrel {\triangle }{=} (t,z,s)\mapsto \bar {\nu }^{\text {from}\thinspace x^{\prime }}(w)\stackrel {\triangle }{=}\big (v^{\text {from}\thinspace x^{\prime }}(z),\psi ^\mathrm {opt}[w]\big ). $$

In this case, for all \( x,z\in \mathbb {R}^d \) and \( w=(\tau ,z,0)\in \mathbb {R}^{d+2} \), by virtue of the \( 2L \)-Lipschitz continuity of \( \hat {f} \) and \( \check {f} \), we have

$$ \begin {aligned} &{}\Big \langle x-w,{f}(x,u^{\text {to}\thinspace w}(x),v\big )- \check {f}\big (w,{\mu }^\mathrm {opt}(w),{\nu }^{\text {from}\thinspace x}(w)\big )\Big \rangle _d \\ &\qquad \qquad {}\leq \max _{{u}}\Big \langle x-w,{f}(x,u,v)-{f}\big (x,u,{v}^{\text {from}\thinspace x}(z)\big )\Big \rangle _d \\ &\qquad \qquad \qquad \qquad {}+\max _{\bar {\mu }}\Big \langle x-w, \check {f}\big (w,\delta _{u^{\text {to}\thinspace w}(x)},{\nu }^{\text {from}\thinspace x}(w)\big )-\check {f}\big (w,\bar {\mu },{\nu }^{\text {from}\thinspace x}(w)\big )\Big \rangle _d \\ &\qquad \qquad \qquad \qquad \qquad \qquad {}+2L\|x-w\|_d^2\leq 2L\|x-w\|_d^2. \end {aligned} $$
(4.1)

Fix some \( x^{\prime }\in \mathbb {R}^d \) and \( w^{\prime }\in \mathbb {R}^{d+2} \). Consider the mapping \( R_{x^{\prime },w^{\prime }}:\mathbb {R}^{d+2}\times \mathbb {R}^{d+2}\to \mathbb {R} \) independent of the zero and last coordinates of the arguments and given by the following rule: for all \( x\in \mathcal {Z}_< \),

$$ \hat {R}_{x^{\prime },w^{\prime }}(x)=\sum _{i=1}^d(\pi _ix^{\prime }-\pi _iw^{\prime })(\pi _ix-\pi _ix^{\prime })=\langle x^{\prime }-w^{\prime },x-x^{\prime }\rangle _d; $$
(4.2)

for the argument of this function we can also use the elements of \( \mathbb {R}^d \) by virtue of the embedding \( \mathbb {R}^d\subset \mathbb {R}^{d+2} \). By a straightforward calculation of the generator, using the \( 2L \)-Lipschitz continuity of \( \check {f} \), we conclude (see [8, (5.2)]) that in the case of \( x\in \mathbb {R}^{d} \) and \( w\in \mathbb {R}^{d+2} \) we have

$$ \begin {aligned} \check {\Lambda }[\bar \mu ^\mathrm {opt},\bar {\nu }^{\text {from}\thinspace x^{\prime }}] {R}_{w^{\prime },x^{\prime }}(w)&= -\big \langle x^{\prime }-w^{\prime },\check {f}(w,\mu ^\mathrm {opt},{\nu }^{\text {from}\thinspace x^{\prime }})\big \rangle \\ &\leq -\big \langle x^{\prime }-w^{\prime },\check {f}(w^{\prime },\mu ^\mathrm {opt},{\nu }^{\text {from}\thinspace x^{\prime }})\big \rangle _d +2L\|x^{\prime }-w^{\prime }\|_d\cdot \|w-w^{\prime }\|_d \\ &\leq -\big \langle x^{\prime }-w^{\prime },\check {f}(w^{\prime },\mu ^\mathrm {opt},{\nu }^{\text {from}\thinspace x^{\prime }})\big \rangle _d +L\|x^{\prime }-w^{\prime }\|^2_d+L\|w-w^{\prime }\|^2_d. \end {aligned} $$
(4.3)

Such a representation will prove useful when substituting the generator into Dynkin’s formula (3.6).

Considering now the random process in the original game, we can verify (see [8, (5.4)]) that for an arbitrary strategy of the second player \( v \), an analog of Dynkin’s formula holds for the generator

$$ K_<\ni x \mapsto \hat {\Lambda }\big [{u}^{\text {to}\thinspace w^{\prime }}(x^{\prime }),{v}\big ] {R}_{x^{\prime },w^{\prime }}(x)=\big \langle x^{\prime }-w^{\prime }, f(x,{u}^{\text {to}\thinspace w^{\prime }}(x^{\prime }),{v})\big \rangle _d; $$
(4.4)

namely, for any stopping times \( \tau ^{\prime } \) and \( \tau ^{\prime \prime } \) ( \( \tau ^{\prime }\leq \tau ^{\prime \prime } \)) one has

$$ {R}_{x^{\prime },w^{\prime }}\big (y(\tau ^{\prime })\big )=\mathbb {E}_{\tau ^{\prime }}\left [{R}_{x^{\prime },w^{\prime }}\big (y(\tau ^{\prime \prime })\big )-\int _{\tau ^{\prime }}^{\tau ^{\prime \prime }}\hat {\Lambda }\big [{u}^{\text {to}\thinspace w^{\prime }}(x^{\prime }),{v}\big ] {R}_{x^{\prime },w^{\prime }}\big (y(t)\big )\thinspace dt\right ]. $$
(4.5)

In particular, applying the \( 2L \)-Lipschitz continuity of \( f \) to (4.4), evaluating the product by half the sum of squares, and using (4.1), we obtain

$$ \begin {aligned} &{}\hat {\Lambda }\big [\bar {u}^{\text {to}\thinspace w^{\prime }}(x^{\prime }),\bar {v}(t)\big ] {R}_{x^{\prime },w^{\prime }}(x) \\ &\qquad \qquad {}\leq \Big \langle x^{\prime }-w^{\prime },f\big (x^{\prime },{u}^{\text {to}\thinspace w^{\prime }}(x^{\prime }),{v}[t]\big )\Big \rangle _d+2L\|x-x^{\prime }\|_d\cdot \|x^{\prime }-w^{\prime }\|_d \\ &\qquad \qquad {}\stackrel {(4.1)}{\leq } \big \langle x^{\prime }-w^{\prime },\check {f}(w^{\prime },\mu ^\mathrm {opt},{v}^{\text {from}\thinspace x^{\prime }})\big \rangle _d+ 3L\|x^{\prime }-w^{\prime }\|_d^2+L\|x-x^{\prime }\|^2_d. \end {aligned} $$
(4.6)

4.2. Construction of the Guide for the First Player

Now we will construct a càdlàg random process in the phase space \( K_<\times \mathcal {Z}_< \); we will construct this process up to the end of the game. Here we assume that some admissible strategy \( V \) and terminal time \( \theta _2 \) are chosen by the second player but are not known to the first one. For each choice of \( V \) on trajectories in the phase space \( K_<\times \mathcal {Z}_< \), we will obtain its own probability distribution; fixing the choice of \( V \), we also fix the distribution \( \tilde {{\mathbb {P} }} \) and the corresponding expectation \( \widetilde {\mathbb {E}} \). The projection \( \hat {Y} \) of this distribution with trajectories with values \( K_<\times \mathcal {Z}_< \) onto the first component will give the distribution \( {\mathbb {P}}_{U,V} \) of trajectories \( y \) of the original game, and indeed, as we will see later, it will be generated by some admissible strategy \( U \) of the first player. The projection \( \check {Y} \) of this distribution onto the second component will be responsible for the distribution of the guide and will be constructed by the first player based on the distributions \( \check {{\mathbb {P}}} \) generated by the time-invariant strategies \( (\mu ^\mathrm {opt},{\nu }^{\text {from}\thinspace y(t_k)}) \).

Thus, suppose that at predetermined times \( (t_k=kh)_{k\in \mathbb {N}} \) the first player reads the position \( y(\cdot ) \) of the real game; we can also assume that \( t_0=0 \). Let us determine the stopping times \( \tau _k\stackrel {\triangle }{=}\min \{t\thinspace |\thinspace \pi _0\check {Y}(t)=kh\} \) for the random process \( \check {Y} \). As noted in Remark 3.1, by the construction of the Markov chain, each of these stopping times is almost everywhere finite; moreover, being the sum of independent random variables distributed exponentially with parameter \( 1/h \), each of these times does not depend in any way on the actions of the players. Moreover, their entire sequence can be considered generated even before the start of the game. In this way, we already know the zero coordinate of the model, \( \pi _0\check {Y}(t)\stackrel {\triangle }{=}|\{i\in \mathbb {N}\cup \{0\}\thinspace | \thinspace \tau _i\leq t\}|h \).

To generate the first to \( d \)th model coordinates up to the stopping time \( \check {\theta }_{\min } \), we will use the projection of the same measure \( \check {\eta } \) but only onto the first \( d \) coordinates,

$$ \eta (w,\bar {\mu },\bar {\nu };A)\stackrel {\triangle }{=}\frac {1}{h} \sum _{i=1}^{d}\big [\pi _i^+ \check {f}(w,\mu ,\nu )\delta _{h\pi _i}+\pi _i^-\check {f}(w,\mu ,\nu )\delta _{-h\pi _i}\big ]; $$

this corresponds to the Kolmogorov matrix \( ({Q}_{wy}(\bar \mu ,\bar \nu ))_{w,y\in \mathcal {Z}_<} \) introduced by the rule

$$ \begin {cases} \pi ^+_i\check {f}(w,\mu ,\nu )/h, & y=w+h\pi _i,i\in [1:d] \\ \pi ^-_i\check {f}(w,\mu ,\nu )/h, & y=w-h\pi _i,i\in [1:d] \\[.3em] -\displaystyle \sum _{i=1}^{d}\big (\pi ^+_i\check {f}(w,\mu ,\nu )+\pi ^-_i\check {f}(w,\mu ,\nu )\big )/h, & y=w, \\ 0, & \textrm {otherwise}. \end {cases} $$

Finally, for the first player, on each interval \( [t_{k-1};t_k) \) we describe the control procedure for the initial conflict-controlled system if the game has not yet been stopped at time \( t_{k} \) and he/she has already constructed the random process \( {Y} \) until the stopping time \( \tau _k \).

At time \( t_0=0 \), the player initializes a random process \( Y \) by assigning \( \hat {Y}(0)=y(0)=x_* \) and \( \check {Y}(0)=w_0 \), where \( w_0 \) is the point of the lattice \( h\mathbb {R}^{d+2} \) closest to \( (0,y(0),0) \), and assigns the control \( {u}^{\text {to}\thinspace {w_0}} \) on the interval \( [0;h) \). Then the first player calculates the trajectory \( \check {Y} \) of the process with the Lévy measure \( \eta \) due to the pair of strategies \( (\mu ^\mathrm {opt},{\nu }^{\text {from}\thinspace y( 0)}) \) until the stopping time \( \tau _1 \), after which he/she assigns \( w_1=\check {Y}(\tau _1) \). Since, according to the condition, it is impossible to terminate the game at the initial time, the only possible value for the first (and for the second) player for \( \phi _0 \) (and for \( \psi _0 \)) is zero.

The first player does not know how the second player is controlled on the interval \( [0;h) \), however, if the game has not yet ended, then at time \( t_1=h \) the first player, knowing the position \( y(t_1)=\hat {Y}( t_1) \), assigns \( {u}^{\text {to}\thinspace {w_1}} \) as his/her control in the original game on the interval \( [h;2h) \), and for the model \( \check {Y} \), assigns time-invariant strategies \( (\mu ^\mathrm {opt},{\nu }^{\text {from}\thinspace y(t_1)}) \) for both players on the interval \( [\tau _1;\tau _2) \). He/she will also take \( \phi _1=\phi ^\mathrm {opt}(\check {Y}(\tau _1))\in \Phi _1 \) as the probability of stopping of the original game initiated by him/her.

Thus, suppose that we have constructed the control in the original game and the trajectory of the Markov process up to the time \( t_k \) and the stopping time \( \tau _{k} \), respectively. Let us show how the first player is controlled until the time \( t_{k+1} \) and the stopping time \( \tau _{k+1} \) in the case where the original game has not yet ended on the interval \( [t_{k-1};t_{k}) \) and the trajectory of the original game is known until the time \( t_{k} \) inclusive.

At time \( t_k=kh \), knowing the positions \( y(t_k)=\hat {Y}(t_k) \) and \( w_k\stackrel {\triangle }{=} \check {Y}(\tau _k) \), the first player assigns the function \( {u}^{\text {to}\thinspace {w_k}} \) as his/her control in the original game on the interval \( [t_k;t_{k+1}) \) and assigns time-invariant strategies \( (\mu ^\mathrm {opt},{\nu }^{\text {from}\thinspace y(t_k)}) \) for both players for the model on the interval \( [ \tau _k;\tau _{k+1}) \). He/she will also take \( \phi _k=\phi ^\mathrm {opt}({Y}(\tau _k)) \). As before, he/she will assume that only the termination of the original game during \( [t_k;t_{k+1}) \) means the termination of the model game; thus, he/she still does not need information about the probability of termination initiated by the second player neither for calculating the Markov chain up to stopping time \( \tau _{k+1} \) nor for the trajectory of the original game up to time \( t_{k+1} \). The trajectories up to time \( t_{k+1} \) and the stopping time \( \tau _{k+1} \) are constructed, and hence the induction hypothesis is proved.

Thus, we have described both the construction of the trajectory \( y(\cdot )=\hat {Y}(\cdot ) \) of the original game up to the time of its end \( \theta _{\min } \) and the construction of all, except for the last one, coordinates of the model \( \check {Y} \) until time \( \check {\theta }_{\min } \). Let us do this for the last coordinate, too.

No matter how the original game develops, it can be finished ahead of time only at one of the times \( t^I_n,t^{II}_n \), and on each interval \( [t_k;t_{k+1}) \) there is at most one such time. Then we can assume that the original game ended at the time \( \theta _{\min } \) in \( [t_k;t_{k+1}) \) exactly when the model game ends on the interval \( [\tau _k;\tau _{ k+1}) \), that is, one of the events \( \check \theta _i\in [\tau _k;\tau _{k+1}) \) will be realized. By virtue of Remark 3.2, the same probabilities of termination for each player in these games imply that the probability of this event coincides with the probability of transfer along the last coordinate in the Markov game. Then on each time interval \( [0;\tau _k) \) we have the opportunity to reconstruct the last coordinate of the model: \( \pi _{d+1}\check {Y}(t)\stackrel {\triangle }{=} 0 \) until the time \( \check {\theta }_{\min } \) of stopping of the model game, \( \pi _{d+1}\check {Y}(t)\stackrel {\triangle }{=} 1+\check {\theta }_{\min } \) for \( t\geq \check {\theta }_{\min }=\check {\theta }_1 \), and \( \pi _{d+1}\check {Y}(t)\stackrel {\triangle }{=} -1-\check {\theta }_{\min } \) for \( t\geq \check {\theta }_{\min }=\check {\theta }_2 \). On the other hand, by prescribing the probabilities \( \phi _k=\phi ^\mathrm {opt}(\check {Y}(\tau _k)) \) the first player has determined not only the stopping time \( \check {\theta }_1 \) of the Markov game but, by virtue of the same remark, has also prescribed the terminal time \( {\theta }_1 \) in the original game.

Thus, for each choice of strategy and the time of termination by the second player on the trajectories of the process \( (\hat {Y},\check {Y}) \) in the phase space \( K_<\times \mathcal {Z}_< \) we have constructed the probability distribution \( \tilde {{\mathbb {P}}} \) and the corresponding expectation \( \widetilde {\mathbb {E}} \). Disintegrating it first over \( \bar {\phi }^\mathrm {opt}(\check {Y}|_{[0;\tau _{k-1}]}) \) and then over \( y|_{[ 0;t_{k-1}]}=\hat {Y}|_{[0;t_{k-1}]} \) and \( u^{\text {to} y()}|_{[0;t_{ k-1}]} \), we obtain the distribution of the first player’s controls up to time \( t_{k+1} \), and so the strategy presented in this section is admissible. In a similar way, each probability \( \phi _k \) of termination by the first (second) player is determined as a marginal probability with the already known, realized part of the trajectory and control and is also admissible.

Note that the indicated procedure for constructing an admissible strategy \( U \) and a terminal time \( \theta _1 \) did not use time assumptions anywhere. Indeed, we were not interested anywhere at what exact time in the interval \( [t_k=kh;t_{k+1}=kh+h) \) the original game could end; we were only interested in whether it ended on this interval. Further, the probability \( \phi ^\mathrm {opt}({Y}(\tau _k)) \) given by us certainly falls into the desired compact set by the construction of \( \Phi _k \), and since we have used only the information available at time \( t_k \), which means that it is certainly present at any time in the interval \( [t_k;t_{k+1}) \), the resulting stopping time \( \theta _1 \) is admissible. Finally, this procedure used the information about whether the game stopped only at the end of each interval \( [t_k;t_{k+1}) \). In particular, it did not prohibit the second player to decide on his/her next probability to end the game during the interval \( [t_k;t_{k+1}) \) only at the time \( t_{k+1} \). Thus, this procedure allows not only all admissible stopping times for the second player but also all stopping times \( \theta _2 \) of the second player for which, with all \( n\in \mathbb {N} \) and \( s\in [0; 1) \), the probabilities of the events \( \{\theta _2<nh+sh\} \) are only \( \mathcal {F}_{nh} \)-measurable. Let us fix a stopping time \( \theta _2 \) with this property.

4.3. Divergence of Trajectories of Original and Markov Games

Let us introduce stopping times \( \theta _{\min }=\min (\theta _1,\theta _2) \) and \( \check {\theta }_{\min }=\min (\check {\theta }_1,\check {\theta }_2) \) in the original and Markov games, respectively.

Fix some \( k\in \mathbb {N} \) and, together with it, times \( t_{k-1}=(k-1)h,t_k=kh\leq T \) from the partition. Now we can also determine the stopping times

$$ \begin {aligned} t^{\prime }&{}\stackrel {\triangle }{=}\min (t_{k-1},{\theta }_{\min }),&\quad t^{\prime \prime }&{}\stackrel {\triangle }{=}\min (t_{k},{\theta }_{\min }), \\ \tau ^{\prime }&{}\stackrel {\triangle }{=}\min (\tau _{k-1},\check {\theta }_{\min }),&\quad \tau ^{\prime \prime }&{}\stackrel {\triangle }{=}\min (\tau _{k},\check {\theta }_{\min }). \end {aligned} $$

By \( x^{\prime } \) we denote the position \( y(t^{\prime }) \) and by \( w^{\prime } \), a point with the coordinates \( (\pi _1\check {Y}(\tau ^{\prime }),\dots ,\pi _d\check {Y}(\tau ^{\prime })) \). In particular, for all \( t \) in \( \tau ^{\prime }\leq t<\tau ^{\prime \prime } \) it follows that \( \pi _0\check {Y}(t)=(k-1)h \); moreover, the estimates (3.7) hold.

Let us disintegrate the probability \( \tilde {{\mathbb {P}}} \) over \( t^{\prime } \) and \( \tau ^{\prime } \). We denote the conditional expectation with respect to the sigma-algebra \( \mathcal {F}_{t^{\prime }}\otimes \check {\mathcal {F}}_{\tau ^{\prime }} \) by \( \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }} \); note that the pair \( (w^{\prime },x^{\prime }) \) is \( \mathcal {F}_{t^{\prime }}\otimes \check {\mathcal {F}}_{\tau ^{\prime }} \)-measurable.

For all \( w,x\in \mathbb {R}^{d+2} \), the definitions of \( R_{x^{\prime },w^{\prime }}=\langle \cdot - x^{\prime },x^{\prime }-w^{\prime }\rangle _d \) and \( R_{w^{\prime },x^{\prime }} \) and the equality

$$ \|w-x\|^2_d-\|w-x+x^{\prime }-w^{\prime }\|^2_d= 2R_{x^{\prime },w^{\prime }}(x)+2R_{w^{\prime },x^{\prime }}(w)+\|x^{\prime }-w^{\prime }\|^2_d $$

imply the inequalities

$$ \frac {\|x-w\|^2_d-\|x^{\prime }-w^{\prime }\|^2_d}{2} \leq \|x-x^{\prime }\|^2_d+\|w-w^{\prime }\|^2_d+R_{x^{\prime },w^{\prime }}(x)+R_{w^{\prime },x^{\prime }}(w) $$

and

$$ \begin {aligned} &{}\widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}\big \|y(t^{\prime \prime })-\check {Y}(\tau ^{\prime \prime })\big \|_d^2-\|x^{\prime }-w^{\prime }\|_d^2 \leq 2\thinspace \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}\big \|\check {Y}(\tau ^{\prime \prime })-w^{\prime }\big \|^2_d \\ &\qquad \qquad {}+2\thinspace \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}\big \|y(t^{\prime \prime })-x^{\prime }\big \|^2_d +2\thinspace \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}R_{x^{\prime },w^{\prime }}\big (y(t^{\prime \prime })\big )+2\thinspace \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}R_{w^{\prime },x^{\prime }}\big (\check {Y}(\tau ^{\prime \prime })\big ). \end {aligned} $$
(4.7)

The first term on the right-hand side has already been estimated in (3.7),

$$ 2\thinspace \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }} \big \|\check {Y}(t^{\prime \prime })-w^{\prime }\big \|^2_d=2 \check {\mathbb {E}}_{\tau ^{\prime }} \big \|\check {Y}(\tau ^{\prime \prime })-w^{\prime }\big \|^2_d\leq 2h^2L\left (\sqrt {d}+1\right ). $$
(4.8)

The second term on the right-hand side in (4.7) is easy to estimate by virtue of the boundedness of the dynamics of (1.1),

$$ 2\thinspace \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }} \big \|y(t^{\prime \prime })-x^{\prime }\big \|^2_d\leq 2\thinspace L^2\widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}(t^{\prime \prime }-t^{\prime })^2\leq 2L^2h^2. $$
(4.9)

For the third term on the right-hand side in (4.7), using Dynkin’s formula (4.5), from (4.6), \( \|y(s)-x^{\prime }\|^2_d\leq L(s-t^{\prime }) \), and \( t^{\prime \prime }-t^{\prime }\leq h \) we conclude that

$$ \begin {aligned} 2\thinspace \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }} R_{x^{\prime },w^{\prime }}\big (y(t^{\prime \prime })\big )&=2 \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }} \Big [R_{x^{\prime },w^{\prime }}\big (y(t^{\prime \prime })\big )-R_{x^{\prime },w^{\prime }}\big (y(t^{\prime })\big )\Big ]\\ &\leq 2h\widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}\big \langle x^{\prime }-w^{\prime },\check {f}(x^{\prime },\mu ^\mathrm {opt},{v}^{\text {from}\thinspace x^{\prime }})\big \rangle _d \\ &\qquad {}+{6L}\|x^{\prime }-w^{\prime }\|^2_d+4L{\mathbb {E}}_{t^{\prime }}\int _{t^{\prime }}^{t^{\prime \prime }}\big \|y(s)-x^{\prime }\big \|^2_d\thinspace ds \\ &\leq 2\big \langle x^{\prime }-w^{\prime },\check {f}(x^{\prime },\mu ^\mathrm {opt},{v}^{\text {from}\thinspace x^{\prime }})\big \rangle _dh+{6L}\|x^{\prime }-w^{\prime }\|^2_d h+4L^3h^3/3. \end {aligned} $$
(4.10)

By virtue of Dynkin’s formula (3.6), the estimates (4.3) and (3.7), and

$$ \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}(\tau ^{\prime \prime }-\tau ^{\prime })\leq \check {\mathbb {E}}_{\tau ^{\prime }}(\tau _k-\tau _{k-1})=h, $$

the last term in (4.7) can be estimated via

$$ \begin {aligned} &{}2\thinspace \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}R_{w^{\prime },x^{\prime }}\big (\check {Y}(\tau ^{\prime \prime })\big )\leq 2\thinspace \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }} \int _{\tau ^{\prime }}^{\tau ^{\prime \prime }}\hat {\Lambda }\big [{u}^{\text {to}\thinspace w^{\prime }}(x^{\prime }),{v}(s)\big ] {R}_{x^{\prime },w^{\prime }}\big (\check {Y}(s)\big )\thinspace ds \\ &\qquad \qquad {}\leq 2\Big (L\|x^{\prime }-w^{\prime }\|^2_d-\big \langle x^{\prime }-w^{\prime },\check {f}(x^{\prime },\mu ^\mathrm {opt},{\nu }^{\text {from}\thinspace x^{\prime }})\big \rangle _d \Big ) \check {\mathbb {E}}_{\tau ^{\prime }}(\tau ^{\prime \prime }-\tau ^{\prime }) \\ &\qquad \qquad \qquad {}+2L\int _{\tau ^{\prime }}^{\tau ^{\prime \prime }}\check {\mathbb {E}}_{\tau ^{\prime }} \big \|\check {Y}(s)-w^{\prime }\big \|^2_d\thinspace ds \\ &\qquad \qquad {}\stackrel {(3.7)}{\leq } {2L}\|x^{\prime }-w^{\prime }\|^2_dh-2\big \langle x^{\prime }-w^{\prime },\check {f}(x^{\prime },\mu ^\mathrm {opt},{\nu }^{\text {from}\thinspace x^{\prime }})\big \rangle _d h+2h^3L^2\left (\sqrt {d}+5\right ). \end {aligned} $$
(4.11)

Substituting the estimates (4.9)–(4.11) into (4.7) and using the inequalities \( L\sqrt {d}h<1 \) and \( Lh(2L/3+5)<1 \) (see (2.1)), we conclude that

$$ \begin{aligned} \widetilde {\mathbb {E}}_{t^{\prime },\tau ^{\prime }}\big \|y(t^{\prime \prime })-\check {Y}(\tau ^{\prime \prime })\big \|_d^2&\leq (1+{8L}h)\|x^{\prime }-w^{\prime }\|^2_d +2h^2L\left (\sqrt {d}+1+2L^2h/3+L\sqrt {d}h+5Lh\right ) \\ &{}\stackrel {(2.1)}{\leq }\big \|y(t^{\prime })-\check {Y}(\tau ^{\prime })\big \|_d^2(1+{8L}h)+ 6h^2L\left (1+\sqrt {d}\right ).\end{aligned} $$

For brevity, for all positive integer \( i \) we introduce the stopping times \( t^\theta _{i}\stackrel {\triangle }{=}\min (t_{i},{\theta }_{\min }) \) and \( \tau ^\theta _{i}\stackrel {\triangle }{=}\min (\tau _{i},\check {\theta }_{\min }) \). Now for \( i\leq T/h \) the above entails

$$ \widetilde {\mathbb {E}}_{t^\theta _{i-1},\tau ^\theta _{i-1}}\big \|y(t^\theta _{i})-\check {Y}(\tau ^\theta _{i})\big \|_d^2 \leq \big \|y(t^\theta _{i-1})-\check {Y}(\tau ^\theta _{i-1})\big \|_d^2\thinspace e^{{8L}h}+ 6Lh^2\left (1+\sqrt {d}\right ). $$
(4.12)

Iterating (4.12), for all values \( i \) less than \( k\stackrel {\triangle }{=} T/h \) we also obtain

$$ \begin{aligned} &{}\widetilde {\mathbb {E}}_{t^\theta _{k-1},\tau ^\theta _{k-1}}\big \|y(t^\theta _{k})-\check {Y}(\tau ^\theta _{k})\big \|_d^2 \\ &\qquad {}\leq \widetilde {\mathbb {E}}_{t^\theta _{k-2},\tau ^\theta _{k-2}}\big \|y(t^\theta _{k-1})-\check {Y}(\tau ^\theta _{k-1})\big \|_d^2\thinspace e^{2\cdot {8L}h}+ 6Lh^2\left (1+\sqrt {d}\right )(1+e^{{8L}h}) \\ &\qquad {}\leq \widetilde {\mathbb {E}}\big \|y(t_{0})-\check {Y}(\tau _{0})\big \|_d^2\thinspace e^{{8kL}h}+ 6Lh^2\left (1+\sqrt {d}\right )\sum _{i=0}^{k-1}e^{{8iL}h} \\ &\qquad {}\leq \left (\frac {hd}{2}+ \frac {6Lh^2\left (1+\sqrt {d}\right )}{{8L}h}\right )e^{8LT}\leq {2hd}e^{{8L}T}.\end{aligned} $$

Here, apart from \( \|y(t_{0})-\check {Y}(\tau _{0})\|_d^2=\|x_0-x_*\|^2\leq hd/2 \), at the last step we also used the inequality \( 1+\sqrt {d}\leq 2d \). Thus, it has been shown that

$$ \widetilde {\mathbb {E}}\Big \|y\big (\min (T,{\theta }_{\min })\big )-\check {Y}\big (\min (\tau _{T/h},\check {\theta }_{\min })\big )\Big \|_d\leq e^{{4L}T}\sqrt {2{d}}\sqrt {h}. $$
(4.13)

4.4. Estimating the Difference of Payments

Now our task is to estimate the difference of \( J(y,\theta _1,\theta _2) \) and \( \check {\sigma }(\check {Y}(\tau _{T/h})) \). To this end, we consider three cases: \( \theta _{\min }=\theta _1\leq T \), \( \theta _{\min }=\theta _2\leq T \), and \( \theta _{\min }> T \).

In the first case, we have \( \theta _{\min }=\theta _1\leq T \); then \( J(y,\theta _1,\theta _2)=\sigma _1(y(\theta _{\min })) \) and \( \check {\theta }_{\min }\leq \tau _{T/h} \); since

$$ \check {\sigma }\big (\check {Y}(T/h)\big )=\check {\sigma }\big (\check {Y}(\theta _{\min })\big )= \sigma _1\big (\pi _0\check {Y}(\theta _{\min }),\dots , \pi _d\check {Y}(\theta _{\min })\big ), $$

we have

$$ \Big |J(y,\theta _1,\theta _2)-\check {\sigma }\big (\check {Y}(\tau _{T/h})\big )\Big |=\Big |\sigma _1\big (y(\theta _{\min })\big )-\sigma _1\big (\pi _1\check {Y}(\theta _{\min }),\dots ,\pi _d\check {Y}(\theta _{\min })\big )\Big |. $$

Now, by virtue of the \( L \)-Lipschitz continuity of \( \sigma _1 \), from (4.13) we obtain

$$ \widetilde {\mathbb {E}}\Big (\big |J(y,\theta _1,\theta _2)-\check {\sigma }(\check {Y}(\tau _{T/h}))\big |\thinspace \bigg |\thinspace \theta _{\min }=\theta _1\leq T\Big )\leq Le^{{4L}T}\sqrt {2d}\sqrt {h}. $$

In the second case, a similar estimate follows from the \( L \)-Lipschitz continuity of \( \sigma _2 \).

For the third case of \( \theta _{\min }> T \), note that, by virtue of the choice of \( T>N_{{\varepsilon }/12} \), condition (1.4) for \( \sigma _i \) implies, now for \( \check {\sigma } \), the inequalities

$$ \begin{aligned} &\widetilde {{\mathbb {P}}}(\theta _{\min }\geq T)\thinspace \widetilde {\mathbb {E}}\Big (\check {\sigma }\big (T,y(T),0\big )\thinspace \bigg |\thinspace \theta _{\min }\geq T\Big )-{\varepsilon }/12 \\ &\qquad \qquad \qquad {}<\widetilde {{\mathbb {P}}}(\theta _{\min }\geq T)\thinspace \widetilde {\mathbb {E}} \Big (J\big (y,\theta _1,\theta _2\big )\thinspace \bigg |\thinspace \theta _{\min }\geq T\Big ) \\ &\qquad \qquad \qquad {}<\widetilde {{\mathbb {P}}}(\theta _{\min }\geq T)\thinspace \widetilde {\mathbb {E}}\Big (\check {\sigma }\big (T,y(T),0\big )\thinspace \bigg |\thinspace \theta _{\min }\geq T\Big )+{\varepsilon }/12.\end{aligned} $$

Now from the \( L \)-Lipschitz continuity of \( \sigma _i \) we have

$$ \begin{aligned} \widetilde {{\mathbb {P}}}(\theta _{\min }\geq T)\widetilde {\mathbb {E}} \big (|J(y,\theta _1,\theta _2)-\check {\sigma }(\check {Y}(\tau _{T/h}))|\thinspace \big |\thinspace \theta _{\min }\geq T\big )<{\varepsilon }/12+Le^{{4L}T}\sqrt {2{d}}\sqrt {h} \stackrel {(2.2)}{\leq } {\varepsilon }/6.\end{aligned} $$

Combining all these cases, we obtain

$$ \begin{aligned} \widetilde {\mathbb {E}} \Big |J(y,\theta _1,\theta _2)-\check {\sigma }\big (\check {Y}(\tau _{T/h})\big )\Big | \leq {\varepsilon }/6+Le^{{4L}T}\sqrt {2{d}}\sqrt {h}\stackrel {(2.2)}{\leq }{\varepsilon }/4.\end{aligned} $$

It remains to note that now (3.4) implies

$$ \check {\mathbb {E}} \left |\thinspace \int _{0}^\infty he^{-ht} \check {\sigma }\big (\check {Y}(t)\big )\thinspace dt- J(y,\theta _1,\theta _2)\right |\leq {\varepsilon }/4+{\varepsilon }/12={\varepsilon }/3. $$

This estimate is independent of the strategy chosen by the second player and the strategies \( \mu ^{opt},\phi ^{opt} \) chosen by the first player entail the expected value \( \widetilde {\mathbb {E}}\int \nolimits _{0}^\infty he^{-ht} \check {\sigma }(\check {Y}(t))\thinspace dt \) no more than \( \check {\mathcal {V}}(0,z_0,0) \) regardless of the strategy of the second player, therefore we have shown that

$$ \widetilde {\mathbb {E}} J(y,\theta _1,\theta _2)\geq \check {\mathcal {V}}(0,z_0,0)-{\varepsilon }/3. $$

Note that \( J(y,\theta _1,\theta _2) \) has been determined in the original game; then

$$ {\mathbb {E}}_{\Omega ,U,V} J(y,\theta _1,\theta _2)\geq \check {\mathcal {V}}(0,z_0,0)-{\varepsilon }/3. $$

Since the admissible strategy \( V \) and the terminal time \( \theta _2 \) of the second player were arbitrary, it has been shown that

$$ {\mathcal {V}}_-\geq \check {\mathcal {V}}(0,z_0,0)-{\varepsilon }/3. $$

Considering the guide for the second player based on the same Markov game, we obtain the symmetric estimate \( {\mathcal {V}}_+\leq \check {\mathcal {V}}(0,z_0,0)+{\varepsilon }/3 \), based on which,

$$ |{\mathcal {V}}_–{\mathcal {V}}_+|\leq 2{\varepsilon }/3; $$
(4.14)

now the pair \( (U,\theta _1) \) constructed in Sec. 4.2 is \( 2{\varepsilon }/3 \)-optimal in the upper game; in particular, (1.6) also holds for it. Thus, the theorem is proved if the temporary assumptions are also satisfied in addition to condition (1.4).

5. GETTING RID OF TEMPORARY ASSUMPTIONS

Now we will show how to reduce the original game to a game with these assumptions. To this end we need the following assertion.

Lemma 1.

Suppose that players 1 and 2, in addition to the sets of admissible strategies, \( \mathfrak {U} \) and \( \mathfrak {V} \) , respectively, have two classes of terminal times, \( \mathfrak {Q}^{I}_{\flat } \) , \( \mathfrak {Q}^{I}_{\sharp } \) and \( \mathfrak {Q}^{II}_{\flat } \) , \( \mathfrak {Q}^{II}_{\sharp } \) , respectively. Suppose that we have the mappings

$$ \begin{aligned} \mathfrak {Q}^{I}_{\sharp }\ni \theta _1&{}\mapsto \theta ^\flat _1\in \mathfrak {Q}^{I}_{\flat }, \\ \mathfrak {Q}^{II}_{\flat }\ni \theta _2&\mapsto \theta ^\sharp _2\in \mathfrak {Q}^{II}_{\sharp }\end{aligned} $$

and nonnegative numbers \( r^{\prime } \) and \( r^{\prime \prime } \) such that with any choice of terminal times \( \theta _1\in \mathfrak {Q}^{I}_{\sharp } \) , \( \theta _2\in \mathfrak {Q}^{II}_{\flat } \) and with any choice of the strategies of the players, the probabilities of the events \( |\theta _1-\theta ^\flat _1|>r^{\prime } \) and \( |\theta _2-\theta ^\sharp _2|>r^{\prime } \) are at most \( r^{\prime \prime } \) .

Then the inequality \( |\mathcal {V}^{\flat }_–\mathcal {V}^{\sharp }_-|\leq 4r^{\prime \prime }+Lr^{\prime } \) holds, here \( \mathcal {V}^{\sharp }_- \) is the lower game value where the terminal times from the classes \( \mathfrak {Q}^{I}_{\sharp } \) and \( \mathfrak {Q}^{II}_{\sharp } \) are admissible and \( \mathcal {V}^{\flat }_- \) is the lower game value where the terminal times from the classes \( \mathfrak {Q}^{I}_{\flat } \) and \( \mathfrak {Q}^{II}_{\flat } \) are admissible.

Proof. First, note that by assumption, with any choice of admissible strategies \( U \) and \( V \) and admissible terminal times \( \theta _1\!\in \!\mathfrak {Q}^{I}_{\sharp } \) and \( \theta _2\!\in \!\mathfrak {Q}^{II}_{\flat } \), the probability of the event \( |\min ({\theta }_{1},{\theta }^{\sharp }_2)-\min (\hat {\theta }^\flat _{1},\hat {\theta }_{2})|\!>\!r^{\prime } \) is at most \( 2r^{\prime \prime } \). Since \( J \) is bounded in absolute value by \( 1 \), it follows that the expectation generated by the strategies \( U \) and \( V \) for \( |J(y,{\theta }_{1},\hat {\theta }^{\sharp }_2) -J(y,{\theta }^\flat _{1},\hat {\theta }_{2})| \) is at most \( 4r^{\prime \prime }+Lr^{\prime } \).

Let an admissible strategy \( U \) with terminal time \( \theta _1\in \mathfrak {Q}^{I}_{\sharp } \) be an \( s \)-optimal pair in the lower game with \( \mathfrak {Q}^{I}_{\sharp } \) and \( \mathfrak {Q}^{II}_{\sharp } \), and assume that there exists

$$ \mathbb {E}_{\Omega ,U,V} J(y,{\theta }_1,{\theta }_2)\leq {\mathcal {V}}^{\sharp }_-+s $$

for every admissible strategy \( V \) and admissible terminal time \( {\theta }_2\in \mathfrak {Q}^{II}_{\sharp } \). Then for every admissible strategy \( V \) and admissible terminal time \( {\theta }_2\in \mathfrak {Q}^{II}_{\flat } \) we have

$$ \mathbb {E}_{\Omega ,U,V} J(y,{\theta }^\flat _{1},{\theta }_2)-4r^{\prime \prime }-Lr^{\prime }\leq \mathbb {E}_{\Omega ,U,V} J(y,{\theta }_{1},{\theta }^\sharp _2)\leq {\mathcal {V}}^{\sharp }_-+s; $$

thus, since \( \theta ^{\flat }_1\in \mathfrak {Q}^{I}_{\flat } \), it has been shown that \( {\mathcal {V}}^{\flat }_-\leq {\mathcal {V}}^{\sharp }_-+4r^{\prime \prime }+Lr^{\prime }+s \). By virtue of the arbitrariness of the positive \( s \), the inequality \( \mathcal {V}^{\flat }_–\mathcal {V}^{\sharp }_-\leq 4r^{\prime \prime }+Lr^{\prime } \) is also proved. The symmetry \( (\flat ,1)\leftrightarrow (\sharp ,2) \) implies the opposite inequality. The proof of the lemma is complete. \( \quad \blacksquare \)

5.1. Getting Rid of the First Assumption

It was shown earlier that under the temporary assumptions, (4.14) holds and (1.6) is valid for some \( 2{\varepsilon }/3 \)-optimal \( U \) and \( \theta _1 \). Let us show that if there is a pair \( (T,h) \) for the original game that satisfies only the second temporary assumption, then the pair \( (U,\theta _1) \) constructed for it in Sec. 4.2 is \( 2{\varepsilon }/ 3+4h \)-optimal in the upper game; in particular, (1.6) holds, and moreover,

$$ {\mathcal {V}}_-\geq {\mathcal {V}}_+-2{\varepsilon }/3-4h\stackrel {(2.1)}{\geq }{\mathcal {V}}_+-{\varepsilon }. $$
(5.1)

Consider a modified game—a game with the same dynamics, the same objective function, and the same sequences \( t^{I}_k \) and \( t^{I}_k \) but satisfying

$$ \begin {aligned} \hat {\phi }^-_k&{}\stackrel {\triangle }{=} \min (1-h,{\phi }^-_k),&\quad \hat {\phi }^+_k&{}\stackrel {\triangle }{=} \min (1-h,{\phi }^+_k),\\ \hat {\psi }^-_l&{}\stackrel {\triangle }{=} \min (1-h,{\psi }^-_l),&\quad \hat {\psi }^+_k&{}\stackrel {\triangle }{=} \min (1-h,{\psi }^+_l) \end {aligned} $$

in the case of \( {t}^{I}_k\leq T \), \( {t}^{II}_l\leq T \) and \( \hat {\phi }^-_k\stackrel {\triangle }{=}\phi ^-_k \), \( \hat {\phi }^+_k\stackrel {\triangle }{=}\phi ^+_k \), \( \hat {\psi }^-_k\stackrel {\triangle }{=}\psi ^-_k \), and \( \hat {\psi }^+_k\stackrel {\triangle }{=}\psi ^+_k \) otherwise. This game satisfies the first assumption. By \( \mathfrak {Q}^{I}_{\sharp } \) and \( \mathfrak {Q}^{II}_{\sharp } \) we denote the classes of terminal times of the first and second players, respectively, admissible in the modified game. We also adopt \( \mathfrak {Q}^{I}_{\flat }\stackrel {\triangle }{=}\mathfrak {Q}^{I} \) and \( \mathfrak {Q}^{II}_{\flat }\stackrel {\triangle }{=}\mathfrak {Q}^{II} \).

Now consider an arbitrary terminal time \( \theta _2\in \mathfrak {Q}^{II}=\mathfrak {Q}^{II}_{\flat } \) of the second player admissible in the original game. It is generated by \( \mathcal {F}_{t^{II}_{k+1}} \)-measurable random variables \( \psi [k] \) ranging in \( [\psi ^-_k;\psi ^+_k] \) on \( D(\mathbb {R}_+,\mathbb {R}^d) \). Then for all \( k\in \mathbb {N} \) we take \( \hat {\psi }[k]\stackrel {\triangle }{=} \min (1-h,{\psi }[k]) \). This guarantees the values of the random variables \( \hat {\psi }[k] \) in \( [\hat {\psi }^-_k;\hat {\psi }^+_k] \) with \( |{\psi }[k]-\hat {\psi }[k]|\leq h \). Now the rules \( \hat {\psi }[k] \) restore the terminal time \( {\theta }^{\sharp }_2\in \mathfrak {Q}^{II}_{\sharp } \) now admissible in the modified game.

Now consider a terminal time \( \hat {\theta }_1\in \mathfrak {Q}^{I}_{\sharp } \) of the first player admissible in the modified game and the corresponding sequence of random variables \( \hat {\phi }[k] \). Note that the value of \( \hat {\phi }[k] \) outside the interval \( [\phi _k^-;\phi _k^+] \) is possible only for \( \hat {t}^I_{k+1}<T \) and either \( 1-h\leq \phi _k^- \) or \( h\geq \phi _k^- \); based on this, by the choice of \( h \), we have \( \phi _k^-=\phi _k^+=1 \) and \( \hat {\phi }_k^-=1-h \). Then we take \( {\phi }[k]=\hat {\phi }[k] \) for \( \hat {\phi }[k]\in [\phi _k^-;\phi _k^+] \) and \( {\phi }[k]=\phi _k^+ \) otherwise; this guarantees \( |\hat {\phi }[k]-{\phi }[k]|\leq h \). These rules reconstruct the terminal time \( \hat {\theta }^\flat _1\in \mathfrak {Q}^{I}_{\flat } \) admissible in the original game for the first player.

Due to the same dynamics, the sets \( \mathfrak {D}^I \) and \( \mathfrak {D}^{II} \) are the sets of all admissible strategies both in the original and in the modified game. In particular, any choice of an admissible pair \( (U,V)\in \mathfrak {D}^I\times \mathfrak {D}^{II} \) defines the same probability on all possible trajectories in each of the games. By construction, the probabilities of the events \( {\theta }_{1}\neq {\theta }^\flat _{1} \) and \( {\theta }_{2}\neq {\theta }^{\sharp }_2 \) following the events \( |{\psi }[k]-\hat {\psi }[k]|\leq h \) are at most \( h \).

All conditions in the lemma are satisfied for \( r^{\prime \prime }=h \) and \( r^{\prime }=0 \), hence the lower values of the original and modified games differ by no more than \( 4h \); now if the original game satisfies the second assumption, then the modified game satisfies both and its lower value is (4.14); so (5.1) is proved for the lower value of the original game.

Moreover, if the original game satisfies the second assumption, then the strategy \( U \) and terminal time \( \theta _1 \) specified in Sec. 4.2 coincide with those constructed by the same procedure in the modified game. In particular, they are admissible and \( 2{\varepsilon }/3+4h \)-optimal for the upper modified game. Therefore, in the original problem they are \( 2{\varepsilon }/3+8h \)-optimal; in particular, they satisfy (1.6). Thus, (5.1) and (1.6) are proved if the original game satisfies the second assumption.

5.2. Getting Rid of the Second Assumption

We first need two unary operations on \( \mathbb {R}_+ \): \( \lceil \cdot \rceil \) and \( \lfloor \cdot \rfloor \). For each \( R\in [0;T/h] \), let \( \lceil R\rceil \) denote the least element of the lattice \( h\mathbb {N} \) greater than \( R \); \( \lfloor R\rfloor \) designates the largest element of \( h\mathbb {N} \) lattice not greater than \( R \). For every \( R\notin [0;T/h] \), set \( \lceil R\rceil =\lfloor R\rfloor =R \). Moreover, since any subinterval of the interval \( [0;T] \) of length \( 2h \) does not contain more than one element of the sequences \( (t^{I}_k)_{k\in \mathbb {N}} \) and \( (t^ {II}_k)_{k\in \mathbb {N}} \), it follows that there are no such elements on \( [T-h;T+h] \) either; then it follows from \( 0<|\lceil {t}\rceil -t| \), \( \lfloor {t}\rfloor -t<h \) that these unary operations preserve the order between the elements \( {t}^{I}_k \) and \( {t}^{II}_l \); moreover, the order is not violated if the operation is applied to only one of the elements.

Now introduce two auxiliary games with the same dynamics, the same objective function, and the same compact sets \( [{\phi }^-_k;{\phi }^+_k] \) and \( [{\psi }^-_k;{\psi }^+_k] \) but differing from the original game by possible terminal times. In the \( \lceil \rfloor \)-game, these sequences will be formed by the elements \( \lceil t^{I}_k\rceil \) for the first player and the elements \( \lfloor t^{II}_k\rfloor \) for the second one. In the \( \lfloor \rceil \)-game, this will be the other way round: \( \lfloor t^{I}_k\rfloor \) for the first player and \( \lceil t^{I}_k\rceil \) for the second one. Denote the corresponding classes of terminal times admissible for the players by \( \lceil \mathcal {Q}^{I}\rceil \), \( \lfloor \mathcal {Q}^{II}\rfloor \) and \( \lfloor \mathcal {Q}^I\rfloor \), \( \lceil \mathcal {Q}^{II}\rceil \) for the \( \lceil \rfloor \)-game and the \( \lfloor \rceil \)-game. Note that in this case the unary operations \( \lceil \cdot \rceil \) and \( \lfloor \cdot \rfloor \) send \( \lfloor \mathcal {Q}^I\rfloor \) and \( \lceil \mathcal {Q}^{I}\rceil \), as well as \( \lfloor \mathcal {Q}^{II}\rfloor \) and \( \lceil \mathcal {Q}^{II}\rceil \), to each other.

Now consider an arbitrary terminal time \( \theta _i \) of one of the players admissible in the original game; in particular, for every \( t \) the events \( \theta _i\leq t \) are \( \mathcal {F}_t \)-measurable. Then at the terminal time \( \lceil \theta _i \rceil \) for every \( t \) the events \( \lceil \theta _i \rceil \leq t \) are \( \mathcal {F}_t \)-measurable as well. Therefore, the terminal time \( \lceil \theta _1 \rceil \) will be admissible in the \( \lceil \rfloor \)-game and the time \( \lceil \theta _2 \rceil \), in the \( \lfloor \rceil \)-game as the times having suitable sets of values. The same reasoning leads to the same conclusion if \( \theta _1 \) is taken to be a terminal time allowed in the \( \lfloor \rceil \)-game and \( \theta _2 \) is a terminal time allowed in the \( \lceil \rfloor \)-game.

Again, owing to the same dynamics, the sets of admissible strategies are common to all the games under consideration, and each such admissible strategy creates the same distribution for all games on the trajectories \( y \). By construction, the probabilities of events \( |{\theta }_{i}-\lceil {\theta }_{1}\rceil |>h \) and \( |{\theta }_{i}-\lfloor {\theta } _{1}\rfloor |>h \) are zero. By virtue of Lemma 1, for \( r^{\prime }=h \) and \( r^{\prime \prime }=0 \) the lower values of the \( \lceil \rfloor \)-game and the \( \lfloor \rceil \)-game satisfy the inequality \( |\mathcal {V}_- ^{\lceil \rfloor }-\mathcal {V}_-^{\lfloor \rceil }|\leq Lh \). Now, using the proved embeddings for terminal times in the original game, we have

$$ \begin{aligned} \mathcal {V}_-&=\sup _{ V\in \mathfrak {D}^{II}, \theta _2\in \mathfrak {Q}^{II}}\inf _{U\in \mathfrak {D}^I, \theta _1\in \mathfrak {Q}^I} \mathbb {E}_{\Omega , U,V} J(y,\theta _1,\theta _2) \\ &\leq Lh+ \sup _{ V\in \mathfrak {D}^{II}, \theta _2\in \mathfrak {Q}^{II}}\inf _{U\in \mathfrak {D}^I, \theta _1\in \mathfrak {Q}^I} \mathbb {E}_{\Omega , U,V} J\big (y,\lceil \theta _1\rceil ,\lfloor \theta _2\rfloor \big ) \\ &\leq Lh+ \sup _{ V\in \mathfrak {D}^{II}, \theta ^{\prime }_2\in \lfloor \mathfrak {Q}^{II}\rfloor }\inf _{U\in \mathfrak {D}^I, \theta _1\in \lceil \mathfrak {Q}^I\rceil } \mathbb {E}_{\Omega , U,V} J(y,\theta ^{\prime }_1,\theta ^{\prime }_2)=Lh+\mathcal {V}^{\lceil \rfloor }_-.\end{aligned} $$

In a similar way, we can prove the inequality \( \mathcal {V}_-\geq Lh-\mathcal {V}^{\lfloor \rceil }_- \). Thus, we have proved that

$$ |\mathcal {V}_-^{\lfloor \rceil }-\mathcal {V}_-|\leq 2Lh. $$

Since the second assumption holds for the constructed \( \lceil \rfloor \)-game, it follows that the strategy \( U \) and the terminal time \( \theta ^{\lceil \rfloor }_1 \) constructed for it in Sec. 4.2 satisfy (1.6). In addition, (5.1) holds for this game, whence for the lower value \( {\mathcal {V}}_- \) of the original game it follows that

$$ {\mathcal {V}}_-\geq {\mathcal {V}}_+-2{\varepsilon }/3-2h(L+2)\stackrel {(2.1)}{\geq } {\mathcal {V}}_+-{\varepsilon }. $$

It remains to note that the strategy \( U \) does not change when going to the original game, and the terminal time \( \theta _1 \) constructed in Sec. 4.2 for the original game is associated with \( \theta ^{\lceil \rfloor }_1 \) by the relation \( \theta _1=\lceil \theta ^{\lceil \rfloor }_1\rceil \). Now the \( 2{\varepsilon }/3+4h \)-optimality of the pair \( (U,\theta ^{\lceil \rfloor }_1) \) for the upper \( \lceil \rfloor \)-game guarantees the \( 2{\varepsilon }/3+2h (L+4) \)-optimality of \( (U,\theta _1) \) for the original game. By virtue of (2.1), its \( \varepsilon \)-optimality has also been proved; i.e., (1.6) is also shown to hold true. Thus, the strategy \( U \) and the terminal time \( \theta _1 \) constructed in Sec. 4.2 for this game satisfy (1.6) already without any assumptions on the original game other than property (1.4).

The proof of the theorem is complete. \( \quad \blacksquare \)