Zero-sum Stochastic Differential Games of Impulse Versus Continuous Control by FBSDEs111This work was supported by the Swedish Energy Agency through grant number 48405-1
Abstract
We consider a stochastic differential game in the context of forward-backward stochastic differential equations, where one player implements an impulse control while the opponent controls the system continuously. Utilizing the notion of “backward semigroups” we first prove the dynamic programming principle (DPP) for a truncated version of the problem in a straightforward manner. Relying on a uniform convergence argument then enables us to show the DPP for the general setting. In particular, this avoids technical constraints imposed in previous works dealing with the same problem. Moreover, our approach allows us to consider impulse costs that depend on the present value of the state process in addition to unbounded coefficients.
Using the dynamic programming principle we deduce that the upper and lower value functions are both solutions (in viscosity sense) to the same Hamilton-Jacobi-Bellman-Isaacs obstacle problem. By showing uniqueness of solutions to this partial differential inequality we conclude that the game has a value.
1 Introduction
The history of differential games is almost as long as the history of modern optimal control theory and traces back to the seminal work by Isaacs [16]. To counter the unrealistic idea that one of the players have to give up their control to the opponent, Elliot and Kalton introduced the notion of strategies defined as non-anticipating maps from the opponents set of controls to the players own controls [10]. Assuming that one player plays a strategy while the opponent plays a classical control, Evans and Souganidis [11] used the theory of viscosity solutions to find a representation of the upper and lower value functions in deterministic differential games as solutions to Hamilton-Jacobi-Bellman-Isaacs (HJBI) equations. Using a discrete time approximation technique, this was later translated to the stochastic setting by Flemming and Souganidis [12]. The natural terminology for these games being zero-sum stochastic differential games (SDGs). Using the theory of backward stochastic differential equations (BSDEs), in particular the notion of backward semigroups, Buckdahn and Li [4] simplified the arguments and further extended the results in [12] to cost functionals defined in terms of BSDEs.
Just as stochastic control was extended to various types of controls in the latter half of the previous century (notably to controls of impulse type in [3]), so has stochastic differential games. Tang and Hou [21] considered the setting of two-player, zero-sum SDGs where both players play switching controls (a particular type of impulse control). Their result was later extended by Djehiche et. al. [7, 8] to incorporate stochastic switching-costs. In the context of general impulse controls, Cosso [5] considered a zero-sum game where both players play impulse controls. By adapting the theory developed in [4], L. Zhang recently extended these results to cost functionals defined by BSDEs [23].
In the present work we will be dealing with SDGs where one player plays an impulse control while the opponent plays a continuous control. This type of game problems have previously be considered by Azimzadeh [1] for linear expectations and when the intervention costs are deterministic and by Bayraktar et. al. [2] when the impulse control is of switching type. We follow the path described above where the cost functional is defined in terms of the solution to a BSDE and introduce the lower value function
and the upper value function
with , where the pair solves the non-standard BSDE
(1.1) |
In the above definitions, (resp. ) and (resp. ) represent the set of impulse (resp. continuous) controls and their corresponding non-anticipative strategies. The generic member of will be denoted by where is the time of the intervention and is the corresponding impulse, taking values in the compact set . Moreover, the impulse cost process is defined as
(1.2) |
where and solves the impulsively and continuously controlled SDE
(1.3) |
for and
(1.4) |
whenever with .
We show that and are both viscosity solutions to the Hamilton-Jacobi-Bellman-Isaacs quasi-variational inequality (HJBI-QVI)
(1.5) |
where and
We then move on to prove that (1.5) admits at most one solution, leading to the main contribution of the paper, namely the conclusion that the game has a value, i.e. that .
As in most previous works on stochastic differential games involving impulse controls, the main technical difficulty we face is showing continuity of the upper and lower value functions in the time variable. In previous works such as [21, 5, 22] continuity is simplified by assuming that the intervention costs do not depend on the state and are non-increasing in time. In [1] the assumption of non-increasing intervention costs is replaced by one where the impulse player commits to, at the start of the game, limit to a fixed number of impulses (where can be chosen arbitrarily large) in addition to assuming that impulses can only be made at rational times.
In the present work we take a completely different approach to the above mentioned articles, where we first show continuity under a truncation and then show that the truncated value functions converge uniformly to the true value functions on compact sets.
The paper is organized as follows. In the next section we give some preliminary definitions and describe the by now well established theory of viscosity solutions to partial differential equations (PDEs) as well as the notion of backward semigroups. Then, in Section 3 we give some preliminary estimates on the solutions to the non-standard BSDE in (1.1). Section 4 is devoted to showing that dynamic programming principles hold for the lower and upper value functions. The proof that the lower and upper value functions are both solutions in viscosity sense to the same HJBI-QVI, that is (1.5), is given in Section 5 while the uniqueness proof is postponed to Section 6.
2 Preliminaries
We let be a complete probability space on which lives a -dimensional Brownian motion . We denote by the augmented natural filtration of .
Throughout, we will use the following notation:
-
•
is the -algebra of -progressively measurable subsets of .
-
•
For , we let be the set of all -valued, -measurable càglàd processes such that and we let be the subset of processes that are continuous.
-
•
We let denote the set of all -valued -measurable processes such that .
-
•
We let be the set of all -stopping times and for each we let be the corresponding subsets of stopping times such that , -a.s.
-
•
We let be the set of all -valued processes where is a compact set.
-
•
We let be the set of all , where is a non-decreasing sequence of -stopping times and is a -measurable r.v. taking values in , such that for all .
-
•
For stopping times we let be the subset of with , -a.s. for . Similarly, we let , be the restriction of to all . When we use the shorthands and .
-
•
For any , we let . Moreover, we introduce and let and .
-
•
We let denote the set of all functions that are of polynomial growth in , i.e. there are constants such that for all .
We also mention that, unless otherwise specified, all inequalities between random variables are to be interpreted in the -a.s. sense.
Definition 2.1.
We introduce that notion of non-anticipative strategies defined as all maps for which whenever , -a.e. on (resp. for which whenever , -a.s.). We denote by (resp. ) the set of non-anticipative strategies.
Moreover, we define the restrictions to an interval denoted (resp. ) as all non-anticipative maps (resp. ).
Definition 2.2.
We will rely heavily on approximation schemes where we limit the number of interventions in the impulse control. To this extent we let for and let be the corresponding set of non-anticipative strategies .
Definition 2.3.
We introduce the concatenation of impulse controls as
and note that for each we have the decomposition .
Similarly, when we let the concatenation of and at be defined as
for all .
Throughout, we make the following assumptions on the parameters in the cost functional where and are fixed constants:
Assumption 2.4.
-
i)
We assume that is Borel measurable, of polynomial growth in , i.e. there is a and a such that
for all , and that there is a constant such that for any , , , and we have
Moreover, we assume that is continuous for all .
-
ii)
The terminal reward satisfies the growth condition
for all , and the following local Lipschitz criterion
-
iii)
The intervention cost is jointly continuous in , bounded from below, i.e.
locally Lipschitz in and locally Hölder continuous in , in particular, we assume that
for some .
-
iv)
For each we have
Remark 2.5.
Moreover, we make the following assumptions on the coefficients of the controlled forward SDE:
Assumption 2.6.
For any , , and we have:
-
i)
The function is jointly continuous and satisfies
and the growth condition
(2.1) for some constants and .
-
ii)
The coefficients and are jointly continuous and satisfy the growth condition
and the Lipschitz continuity
2.1 Viscosity solutions
We define the upper, , and lower, semi-continuous envelope of a function as
Next we introduce the notion of a viscosity solution using the limiting parabolic superjet and subjet of a function (see pp. 9-10 of [6] for a definition):
Definition 2.7.
Let be a locally bounded l.s.c. (resp. u.s.c.) function from to . Then,
-
a)
It is referred to as a viscosity supersolution (resp. subsolution) to (1.5) if:
-
i)
(resp. )
-
ii)
For any and (resp. ) we have
(resp.
-
i)
-
b)
It is referred to as a viscosity solution if it is both a supersolution and a subsolution.
We will sometimes use the following equivalent definition of viscosity supersolutions (resp. subsolutions):
Definition 2.8.
A l.s.c. (resp. u.s.c.) function is a viscosity supersolution (subsolution) to (1.5) if (resp. ) and whenever is such that and has a local maximum (resp. minimum) at , then
Remark 2.9.
denotes the set of real-valued functions that are continuously differentiable up to third order and whose derivatives of order one to three are bounded
2.2 Backward semigroups
For we let and assume that . For all we then define (see [17])
(2.2) |
where is the unique solution333From now on we assume that any referred to uniqueness of solutions to a BSDE is uniqueness in and therefore refrain from referring to the space. to
The so defined family of operators is referred to as the backward semigroup related to the BSDE.
3 Forward- Backward SDEs with impulses
In this section we consider the non-standard BSDE in (1.1). Impulsively controlled BSDEs in the non-Markovian framework were treated in [19], while BSDEs related to switching problems have been treated in [15, 14, 13].
Considering first the forward SDE, we get by repeated use of standard results for SDEs (see e.g. Chapter 5 in [20]) that (1.3)-(1.4) admits a unique solution for any since , -a.s. Now, any solution of (1.1) can be written , where solves the standard BSDE
(3.1) |
By standard results we find that (3.1) admits a unique solution whenever and . By a moment estimate given in the next section we are able to conclude that (1.1) admits a unique solution whenever .
3.1 Estimates for the controlled diffusion process
Proposition 3.1.
For each , there is a such that
(3.2) |
-a.s. for all .
Proof. We use the shorthand . By Assumption 2.6.(i) we get for , using integration by parts, that
We note that if and for some then there is a largest time such that . This means that during the interval interventions will not increase the magnitude . By induction we find that
(3.3) |
for all , where , , for and .
Now, since and coincide on we have
and
Inserted in (3.3) this gives
The Burkholder-Davis-Gundy inequality now gives that for ,
and Grönwall’s lemma gives that for ,
(3.4) |
-a.s., where the constant does not depend on , or and (3.2) follows by letting on both sides and using Fatou’s lemma. The result for general follows by Jensen’s inequality.∎
As mentioned above, inequality (3.2) guarantees existence of a unique solution to the BSDE (1.1). We will also need the following stability property.
Proposition 3.2.
For each and , there is a such that
-a.s. for all , with , and all .
Proof. To simplify notation we let and for . Moreover, we let and set . Define , then if we have , where for any value of ,
When we get for ,
By induction we find that
Now, since
Proposition 3.4 gives that
Similarly,
and we find that
Moreover, we note that for and (with ),
and the Burkholder-Davis-Gundy inequality gives for we have
The Lipschitz conditions on the coefficients combined with Grönwall’s lemma then implies that
Now, since for the result follows by induction.∎
3.2 Estimates for the BSDE
For and we let be the unique solution to the following standard BSDE
(3.5) |
Combining classical results (see e.g. [9]) with Proposition 3.1, we have
(3.6) |
-a.s. for all .
We have the following straightforward generalization of the standard comparison principle:
Lemma 3.3.
(Comparison principle) If satisfies Assumption 2.4, and is defined as but with driver instead of , then if for all , we have , -a.s. for each whenever are such that , -a.s.
Proof. This follows immediately from the standard comparison principle (see Theorem 2.2 in [9]).∎
Using the comparison principle we easily deduce the following moment estimates:
Proposition 3.4.
We have,
(3.7) |
and for each , there is a such that
(3.8) |
-a.s. for all .
Proof. The first statement follows by repeated application of the comparison principle which gives that and using (3.6).
The second statement follows by noting that for fixed , there is a such that
for all .∎
Proposition 3.5.
For each , there is a such that
(3.9) |
-a.s. for all with and all and .
Proof. To simplify notation, we let and and set and . By defining and we have for that
with and . We now introduce the processes and defined as444Throughout, we use the convention that
and
We then have by the Lipschitz continuity of that . Using Ito’s formula we find that
with . Taking expectations on both sides yields
Now,
where we have used Proposition 3.2 to reach the last inequality. Moreover,
Combining the above inequalities, the assertion follows.∎
The above proof immediately gives the following stability result:
Corollary 3.6.
(Stability) If satisfies Assumption 2.4, and is defined as with driver instead of , then there is a such that
-a.s. for all and .
4 Dynamic programming principles
In this section we show that and are jointly continuous (deterministic) functions that satisfy the dynamic programming relations
(4.1) |
and
(4.2) |
for and .
Proposition 4.1.
For every we have and , -a.s.
Proof. This follows by repeating the steps in the proof of Proposition 4.1 in [4].∎
We can thus pick the deterministic versions to represent and . As mentioned in the introduction, the main technical difficult that we encounter appears when trying to show continuity of the upper and lower value functions in the time variable. The reason for this is that the constant in Proposition 3.5 depends on and tends to infinity as tends to infinity. We resolve this issue by first considering the upper and lower value functions under an imposed restriction on the number of interventions in the impulse control. Relying on a uniform convergence result will then give us continuity of and .
4.1 A DPP with limited number of impulses
We introduce the truncated value functions
and
for . Similarly to and we have:
Lemma 4.2.
For every and we have and , -a.s.
Combined with the estimates of the previous section this gives the following estimates:
Proposition 4.3.
For each , there is a such that
(4.3) |
for all . Moreover, there is a such that
for all and .
Proof. Since
we have
for each and some . We also see that the same relation holds for . Taking expectation on both sides and using that and are deterministic, the first inequality follows by Proposition 3.5 since was arbitrary.
The second inequality is an immediate consequence of Proposition 3.4.∎
Turning now to the dynamic programming principles, that will be obtained by applying arguments similar to those in Section 4 of [4], we have:
Proposition 4.4.
For each and any , and we have
(4.4) |
and
(4.5) |
Remark 4.5.
At first glance the DPP for may seem counter-intuitive as, on the right-hand side, could take two different values at time (one under and the other in ) and thus trigger two different reactions from the impulse controller at time . However, by the definition of a non-anticipative strategy, whenever , -a.s. and an arbitrary choice of will not influence the overall value.
Proof. The proof (which is only given for the lower value function as the arguments for are identical) will be carried out over a sequence of lemmata where
Lemma 4.6.
can be chosen to be deterministic.
Proof. Again, this follows by repeating the steps in the proof of Proposition 4.1 in [4].∎
Lemma 4.7.
.
Proof. We begin by picking an arbitrary and note that we can define the restriction, , of to as
We fix and have by a pasting property555We can paste together two controls on sets and by setting and get by uniqueness of solutions to our BSDE that . that there is a such that
Now, given we can define the restriction, , of to as
We let be a partition of such that , then by Proposition 4.3 there is a such that for all and . We pick and have by the same pasting property as above that there is for each and , a such that
Consequently,
with
Using first comparison and then the stability property for BSDEs we find that
where only depends on the coefficients of the BSDE. Now, as this holds for all we conclude that , but was arbitrary and the result follows.∎
The opposite inequality and its proof are classical (see e.g. Proposition 1.10 in [12] and Proposition 3.1 in [21]) and we give the proof only for the sake of completeness.
Lemma 4.8.
.
Proof. We again fix an and let be defined as above. We pick an for each and note that there is a (see [4] Lemma 4.5) such that
for all . Moreover, there is an such that
for all , where is the number of interventions in . Now, each can be uniquely decomposed as with (with interventions) and (with first intervention at ). Then,
with
Since , we conclude that , where does not depend on which in turn was arbitrary and the result follows.∎
4.2 A DPP for the general case
We turn now to the general case where there is no restriction on the number of interventions in the impulse control. Before taking the limit as in and , we need to delimit the set of impulse controls:
Definition 4.9.
For and we let be the set of all such that , -a.s., for all .
Moreover, we let be the subset of all such that for each and ,
-a.s.
Given an we note that the set consists of all controls where it is never (on average) beneficial to abandon and stop intervening on the system for the remainder of the period. Similarly, is the set of strategies where, given that the opponent acts rationally, it will never be beneficial to abandon and stop intervening. The usefulness of the above definitions in our case lies in the fact that they allow us to bound the corresponding solution to (1.1) from below by an expression that does not involve intervention costs. In particular, we have whenever and , that
(4.6) |
for all , and similarly when we have
(4.7) |
for all and .
The following lemma shows that these sets contain all relevant impulse controls and strategies, respectively.
Lemma 4.10.
We have
and
Proof. For any and arbitrary we let
Assumption 2.4.iv implies that and we get that with
and
the set is -negligible.
Moreover, since
it follows that on we have and we conclude that letting we have -a.s. By comparison we thus find that , -a.s. for all . In particular, this gives that and from which we conclude that any is dominated by an element of . Since this holds for any , we have that
proving the first statement.
For the second statement we fix and . We then set and let
Furthermore, we define as and let . By definition we have
(4.8) |
For and we let be the subset of all with on such that
and similarly let be the subset of all with on such that
Then, we can repeat the arguments in Lemma 4.7 to conclude that for all , the sets and are non-empty and comparison implies that
(4.9) |
and
(4.10) |
Moreover, for and with on we have by (4.8), that
and using comparison together with stability implies that
for all . In particular, since was arbitrary, letting and using (4.9) and (4.10) gives that
and we conclude that dominates . On the other hand, by a similar argument we find that
for all and since on we conclude that
and the assertion follows.∎
In particular, we may w.l.o.g. restrict our attention to impulse controls (resp. strategies) in Definition 4.9. The following result relates the number of interventions in these impulse controls and strategies to the magnitude of the initial value and is central in deriving continuity of and .
Lemma 4.11.
Proof. Both statements will follow by a similar argument and we set (resp. ). To simplify notation we let and and get that
Letting
and
we have by the Lipschitz continuity of that . Using Ito’s formula we find that
with . Since the intervention costs are positive, taking the conditional expectation on both sides gives
On the other hand, by (4.6) (resp. (4.7)) we have
Proposition 3.1 then gives
Next, we derive a bound on the -norm of . Applying Ito’s formula to we get
(4.12) |
where is without the first intervention costs. Since the intervention costs are nonnegative, we have
for any . Inserted in (4.12) and using the Lipschitz property of this gives
Now, as , it follows that the stochastic integral is uniformly integrable and thus a martingale. To see this, note that the Burkholder-Davis-Gundy inequality gives
Taking expectations on both sides thus gives
Finally,
and
from which (4.11) follows by choosing sufficiently large.∎
Lemma 4.12.
There is a such that for all we have
for all . In particular, the sequence converges to , uniformly on compact subsets of .
Proof. For each and there is by Lemma 4.10 a such that
(4.13) |
-a.s.
Now, let and for , set
, where we recall that is the truncation of to the first interventions. As on we have, with and , that for all . Letting and , this gives
for some , with . Taking expectation on both sides and using the Cauchy-Schwartz inequality gives
Now, as , Lemma 4.11 implies that
Since was arbitrary we can pick such that
and we find that
from which the desired inequality follows since was arbitrary. In particular, we find that converges uniformly on sets where is bounded.∎
Theorem 4.13.
is continuous and satisfies (4.1)
Proof. Since the sequence is non-decreasing, Lemma 4.12 implies that uniformly on compacts as . Hence, is continuous.
It remains to show that satisfies (4.1). We have by (4.4) and comparison that
and it follows that . On the other hand, for each and any we can repeat the argument in Lemma 4.12 to find that there is a such that
Moreover, for each , let be the unique solution to
while we assume that satisfies
Then by comparison and
(4.14) |
with , where . Since the right-hand side of the above inequality tends to 0 as we conclude by taking the essential supremum over all that there is a such that
for each . We conclude that and since was arbitrary it follows that satisfies (4.1).∎
Lemma 4.14.
There is a constant such that for all we have
for all . In particular, the sequence converges uniformly on compact subsets of .
Proof. For each there is a such that
-a.s. Then, for ,
By arguing as in the proof of Lemma 4.12 the result now follows.∎
Theorem 4.15.
is continuous and satisfies (4.2).
Proof. As above we find that uniformly on compacts and conclude that is continuous.
5 The value functions as viscosity solutions to the HJBI-QVI
Our main motivation for deriving the dynamic programming relations in the previous section is that we wish to use them to prove that the upper and lower value functions are solutions, in viscosity sense, to the Hamilton-Jacobi-Bellman-Isaacs quasi-variational inequality (1.5).
Whenever (resp. ) a simple application of the dynamic programming principle stipulates that it is suboptimal for the impulse controller to intervene on the system at time . One main ingredient when proving that (resp. ) is a viscosity solution to (1.5) is showing that if (resp. ) then, on sufficiently small time intervals, we may (to a sufficient accuracy) assume that the impulse controller does not intervene on the system. As the probability that the state, when starting in at time , leaves any ball with a finite radius containing on a non-empty interval is positive, this results requires a slightly intricate analysis compared to the deterministic setting (something that was pointed out already in [21]). In the following sequence of lemmas we extend the results from [21] to the case when the cost functional is defined in terms of the solution to a BSDE.
The first lemma is given without proof as it follows immediately from the definitions:
Lemma 5.1.
Let be locally bounded functions. is monotone (if pointwise, then ). Moreover, (resp. ) is l.s.c. (resp. u.s.c.).
In addition, rather than relying on the standard DPP from the previous section, formulated at deterministic times, we need the following “weak” DPP:
Lemma 5.2.
Assume that and , then for any we have
(5.1) |
and
(5.2) |
Proof. For we let for all , where is such that666We can repeat the approximation routine from Lemma 4.8 to show that such a strategy exists.
(5.3) |
-a.s., for all , where . Then and there is a such that
On the other hand, the semi-group property of along with (5.3) and comparison gives that
Since was arbitrary the first inequality follows by taking the essential supremum over all .
Concerning the second inequality we have that for each , there is a such that
for all . With (assuming that when does not contain interventions) we let be such that
Applying the continuous control (see Remark 4.5 concerning the value at the point of concatenation) and using the semi-group property of (as above) now leads to the second inequality.∎
Lemma 5.3.
Let be such that then there is a and an such that
for all .
For each we have, with , for that
(5.4) |
-a.s. We introduce the stopping time
(with ) and get that
-a.s. Choosing gives
-a.s. Using this inequality we will show that there is a such that for some and all we have
for all from which the result of this lemma follows by Lemma 5.2. For any , let be the unique solution to
with and let solve
Then, with
(5.5) |
and
(5.6) |
we have by the Lipschitz continuity of that . Then, with
we have
where
and
where we have used the polynomial growth of and together with the fact that on . We can now get rid of and use the martingale property of to find that
whenever is small enough that . Combined, this gives that there is a such that whenever we have . Since and were arbitrary the assertion follows.∎
Lemma 5.4.
Let be such that then there is a and an such that
for all .
Proof. As in the proof of the above lemma, there is a and an such that
We can thus repeat the steps in the previous lemma to conclude that there is a such that
for all and for some . The lemma then follows by applying the second inequality in Lemma 5.2.∎
We now fix , and . Following the standard procedure to go from a DPP to a quasi-variational inequality when dealing with a controlled FBSDE (see e.g. [18]) we introduce the BSDEs
and
(5.7) |
with
Remark 5.5.
In particular, we note that is a viscosity supersolution (subsolution) of (1.5) if , and () on whenever is such that and attains a local minimum (maximum) at .
Note that the only reason that (5.7) is stochastic comes from the fact that is a stochastic control. In regard to Hamiltonian minimization it seems natural to introduce the following ordinary differential equation (ODE)
We have the following auxiliary lemma, that summarize the results in Lemma 5.1 and Lemma 5.3 of [4].
Lemma 5.6.
For every and we have
(5.8) |
Also, we have that
(5.9) |
Proof. The first property follows from the definition of and Ito’s formula applied to . The second result is immediate from the comparison principle of BSDEs.∎
We now give a sequence of lemmata that will help us show that is a viscosity solution to (1.5).
Lemma 5.7.
We have
Proof. Note that
Concerning the first term on the right-hand side we have
For the remaining terms,
and classically we have
Combining the above estimates the desired results follows.∎
Lemma 5.8.
There is a such that
for each and .
Proof. Grönwall’s inequality gives that
and we conclude that .
∎
Lemma 5.9.
Assume that is such that has a local maximum at where . Then, there are constants such that
for all and .
Proof. Since has a local maximum at there are constants and a such that for all . Now, let
and note from the proof of Lemma 5.3 that , -a.s. Assume that and let be the unique solution to
and assume that satisfies
Then, with
where and are given by (5.5)-(5.6). By comparison we have
and the result follows.∎
Theorem 5.10.
is a viscosity solution to (1.5).
Proof. To begin with we clearly have that for all (see Remark 2.5). We first show that is a viscosity supersolution. For this, we fix and assume that is such that has a local maximum at , where .
If we have by the DPP that
On the other hand by Lemma 5.9 we have for sufficiently small that
Now, (5.8) gives
Combined this gives
In particular, by Lemma 5.7 and (5.9) this implies that
Hence, and we conclude by Lemma 5.8 that
and by continuity of it follows that
Assume instead that , then and we conclude that is a viscosity supersolution.
We turn now to the subsolution property. We fix and assume that is such that has a local minimum at , where . If we have by the DPP and Lemma 5.3 that, whenever is sufficiently small,
On the other hand repeating the argument in the proof of Lemma 5.9 gives that
and we get that
i.e. . Now, repeating the above argument we find that
Analogously we get when then and we conclude that is a viscosity subsolution.∎
6 Uniqueness of viscosity solutions to the HJBI-QVI
To be able to conclude that the game has a value, i.e. that , we will now show that (1.5) has at most one solution in the viscosity sense in . We let
(6.1) |
and have that
We will need the following lemma:
Lemma 6.1.
Proof. With we note that, since is a supersolution and , we have so that the terminal condition holds. Moreover, we have
Since is a supersolution, we have
Now, either in which case it follows by (2.1) that or and (2.1) gives that . We conclude that
Next, let be such that has a local maximum of 0 at with . Then and has a local maximum of 0 at . Since is a viscosity supersolution, we have
Consequently,
where the right hand side is non-negative for all and all for some .∎
We have the following results the proof of which we omit since it is classical:
Lemma 6.2.
A locally bounded function is a viscosity supersolution (resp. subsolution) to (1.5) if and only if for every , is a viscosity supersolution (resp. subsolution) to
(6.2) |
Remark 6.3.
We have the following comparison result for viscosity solutions in :
Proposition 6.4.
Let (resp. ) be a supersolution (resp. subsolution) to (1.5). If , then .
Proof. First, we note that we only need to show that the statement holds for solutions to (6.2). We thus assume that (resp. ) is a viscosity supersolution (resp. subsolution) to (6.2).
It is sufficient to show that
for all and any . Then the result follows by taking the limit . Moreover, we know from Lemma 6.1 that there is a such that is a supersolution to (6.2) for each and .
By assumption, , which implies that there are and such that
Hence, for each there is a such that
We search for a contradiction and assume that there is a such that . Then there is a point (the open unit ball of radius centered at 0) such that
We first show that there is at least one point such that
-
a)
and
-
b)
.
We again argue by contradiction and assume that for all . Indeed, as is u.s.c. and is continuous, there is a such that
(6.3) |
Now, set and note that since
it follows that . Moreover, as is a supersolution it satisfies
or
and we conclude from (6.3) that
Hence, and by our assumption it follows that there is a such that
and a corresponding . Now, this process can be repeated indefinitely to find a sequence in such that for any we have
with . Now, as we get a contradiction by letting while noting that is bounded on . We can thus pick a such that a) and b) above holds.
The remainder of the proof is similar to the proof of Proposition 4.1 in [13]. We assume without loss of generality that and define
where
Since is u.s.c. and is l.s.c. there is a (with the closure of ) such that
Now, the inequality gives
Consequently, is bounded (since and are bounded on ) and as . We can, thus, extract subsequences such that as . Since
it follows that
and as the righthand side is dominated by we conclude that
In particular, this gives that which implies that
and
We can extract a subsequence of such that , and
Moreover, since is u.s.c. (see Lemma 5.1) and there is an such that
for all . To simplify notation we will, from now on, denote simply by .
By Theorem 8.3 of [6] there are and such that
and for every ,
where . Now, we have
where is the identity-matrix of suitable dimension and
In particular, since and are bounded, choosing gives that
(6.6) |
By the definition of viscosity supersolutions and subsolutions we have that
for all and
Combined, this gives that
Collecting terms we have that
and since is Lipschitz continuous in and bounded on , we have
where the right-hand side tends to 0 as . Let denote the column of and let denote the column of then by the Lipschitz continuity of and (6.6), we have
and we conclude that
Finally, we have for some that
Repeating the above argument we get that the upper limit of the right-hand side when is bounded by . Put together, this gives that
a contradiction since was arbitrary.∎
References
- [1] P. Azimzadeh. A zero-sum stochastic differential game with impulses,precommitment, and unrestricted cost functions. Appl Math Optim, 79:483–514, 2019.
- [2] E. Bayraktar, A. Cosso, and H. Pham. Robust feedback switching control: dynamic programming and viscosity solutions. SIAM J. Control Optim., 54(5):2594–2628, 2016.
- [3] A. Bensoussan and J.L. Lions. Impulse Control and Quasivariational inequalities. Gauthier-Villars, Montrouge, France, 1984.
- [4] R. Buckdahn and J. Li. Stochastic differential games and viscosity solutions of hamilton-jacobi-bellman-isaacs equations. SIAM J. Control Optim, 47(1):444–475, 2008.
- [5] A. Cosso. Stochastic differential games involving impulse controls and double-obstacle quasi-variational inequalities. SIAM J. Control Optim., 3(51):2102–2131, 2013.
- [6] M. G. Crandall, H. Ishii, and P. L. Lions. User’s guide to viscosity solutions of second order partial differential equations. Bulletin of the American Mathematical Society, 27(1):1–67, 1992.
- [7] B. Djehiche, S. Hamadéne, and M. Morlais. Viscosity solutions of systems of variational inequalities with interconnected bilateral obstacles. Funkcialaj Ekvacioj, 58(1):135–175, 2015.
- [8] B. Djehiche, S. Hamadéne, M.-A. Morlais, and X. Zhao. On the equality of solutions of max-min and min-max systems of variational inequalities with interconnected bilateral obstacles. J. Math. Anal. Appl., 452:148–175, 2017.
- [9] N. El Karoui, S. Peng, and M. C. Quenez. Backward stochastic differential equationsin finance. Math. Finance, 7(1):1–71, 1997.
- [10] R. J. Elliott and N. J. Kalton. The existence of value in differential games. Number 126. Memoirs of the American Mathematical Society, Providence, Rhode Island, 1972.
- [11] L. C. Evans and P. E. Souganidis. Differential games and representation formulasfor solutions of hamilton-jacobi-isaacs equations. Indiana Univ. Math. J., 33:773–797, 1984.
- [12] W. H. Flemming and P. E. Souganidis. On the existence of value functions of two-player, zero-sum stochastic differential games. Indiana Univ. Math. J., 38:293–314, 1989.
- [13] S. Hamadéne and M. A. Morlais. Viscosity solutions of systems of pdes with interconnected obstacles and switching problem. Appl Math Optim., 67:163–196, 2013.
- [14] S. Hamadéne and J. Zhang. Switching problem and related system of reflected backward SDEs. Stochastic Process. Appl., 120(4):403–426, 2010.
- [15] Y. Hu and S. Tang. Multi-dimensional BSDE with oblique reflection and optimal switching. Prob. Theory and Related Fields, 147(1-2):89–121, 2008.
- [16] R. Isaacs. Differential games. A mathematical theory with applications to warfare andpursuit, control and optimization. John Wiley & Sons, Inc., New York-London-Sydney, 1965.
- [17] J. Li and S. Peng. Stochastic optimization theory of backward stochastic differential equations with jumps and viscosity solutions ofhamilton-jacobi-bellman equations. Nonlinear Analysis, 70:1776–1796, 2009.
- [18] J. Li and Q. Wei. Optimal control problems of fully coupled fbsdes and viscosity solutions of hamilton-jacobi-bellman equations. SIAM J. Control Optim, 52(3):1622–1662, 2014.
- [19] M. Perninge. Finite horizon robust impulse control in a non-markovian framework and related systems of reflected bsdes. arXiv:2103.16272, 2021.
- [20] P. Protter. Stochastic Integration and Differential Equations. Springer, Berlin, 2nd edition, 2004.
- [21] S. Tang and Sh. Hou. Switching games of stochastic differential systems. SIAM J. Control Optim., 46(3):900–929, 2007.
- [22] F. Zhang. Stochastic differential games involving impulse controls. ESAIM Control Optim. Calc. Var., 17(3):749–760, 2011.
- [23] L. Zhang. A bsde approach to stochastic differential games involvingimpulse controls and hjbi equation. J Syst Sci Complex, 2021.