Path-dependent Hamilton-Jacobi equations
with super-quadratic growth in the gradient
and the vanishing viscosity method
Abstract.
The non-exponential Schilder-type theorem in Backhoff-Veraguas, Lacker and Tangpi [Ann. Appl. Probab., 30 (2020), pp. 1321–1367] is expressed as a convergence result for path-dependent partial differential equations with appropriate notions of generalized solutions. This entails a non-Markovian counterpart to the vanishing viscosity method.
We show uniqueness of maximal subsolutions for path-dependent viscous Hamilton-Jacobi equations related to convex super-quadratic backward stochastic differential equations.
We establish well-posedness for the Hamilton-Jacobi-Bellman equation associated to a Bolza problem of the calculus of variations with path-dependent terminal cost. In particular, uniqueness among lower semi-continuous solutions holds and state constraints are admitted.
Key words and phrases:
Large deviations, vanishing viscosity method, path-dependent partial differential equations, minimax solutions, Dini solutions, calculus of variations, optimal control, state constraints, backward stochastic differential equations, nonsmooth analysis2010 Mathematics Subject Classification:
60H30; 60F10, 35F21, 35K10, 49J221. Introduction
Backhoff-Veraguas, Lacker and Tangpi [BLT] derived a non-exponential Schilder-type theorem, which they used to obtain new limit theorems for backward stochastic differential equations (BSDEs) and, in the Markovian case, for the corresponding partial differential equations (PDEs). They posed the question whether it is possible to have a corresponding PDE result in the non-Markovian case as well. Purpose of our work is to provide an answer to this question.
We establish well-posedness for (second order) path-dependent viscous Hamilton-Jacobi equations and for (first-order) path-dependent Hamilton-Jacobi-Bellman (HJB) equations with possibly super-quadratic growth in the gradient. Together with a modification of the Schilder type theorem in [BLT], we obtain a non-Markovian vanishing viscosity result for path-dependent PDEs and thereby address the mentioned open problem in [BLT].
The notions of solutions for our path-dependent PDEs are in the spirit of contingent solutions for PDEs (see, e.g., [Frankowska89AMO]), also known as Dini solutions (see, e.g., [BardiCapuzzoDolcetta] and [Vinter]) or minimax solutions (see, e.g., [SubbotinBook]).
In the context of contingent or Dini solutions for first-order standard PDEs related to Bolza problems, [DM-F00] and also [PQ02] are very close to our approach. More recent works in this direction are [Misztela14], [BernisBettiol18COCV] and [BernisBettiol20chapter]. Regarding the possible use of viscosity solution techniques, we refer the reader to the remarks on p. 1202 in [BettiolVinter17SICON]. In particular, fast growth in the gradient, discontinuity of the Lagrangian, and extended real-valued lower semi-continuous terminal data (to allow right-end point constraints in optimal control problems) cause non-trivial issues. For example, solutions of HJB equations can be expected then to be only lower semi-continuous.
In the second-order case, the only works we are aware of that use contingent-type solutions are [SubbotinaEtAl85] and [Subbotina06], where Isaacs equations corresponding to Markovian stochastic differential games with drift control and bounded control spaces are studied. However, similar constructions (stochastic or Gaussian derivatives) are also used in [Haussmann92MOR] and [Haussmann94SICON].
In the context of first-order path-dependent PDEs, [KaiseEtAl18PPDE] is most relevant. It deals with a calculus of variations problem involving a path-dependent terminal cost and the related path-dependent HJB equation. The setting is very close to our problem (DOC) below. The main difference is that in [KaiseEtAl18PPDE] the terminal cost is required to be Lipschitz continuous, which leads to Lipschitz continuity of the corresponding value function and also makes it possible in [KaiseEtAl18PPDE] to develop a viscosity solution theory. In our work, we require only continuity resp. lower semi-continuity of the terminal cost, which is one of the reasons why we establish a Dini resp. minimax solution theory. For the current state of the art for first-order path-dependent PDEs and for further relevant works, see [GLP21AMO] and the references therein.
In the literature [EKTZ11, K14, ETZ_I, ETZ_II, PhamZhang14SICON, EkrenZhang16PUQR, RTZ14overview, Ren16AAP, Ren17Stoch, Ekren17SPA, RTZ17, cosso18, RTZ20SICON, RR20SIMA] on viscosity solutions of second-order path-dependent PDEs, the Hamiltonian is required to grow at most linearly in the gradient (the same condition is also needed in [CossoRusso19Osaka], where a notion of strong-viscosity solutions is used). Overcoming this restriction for any notion of generalized solutions has been a longstanding open problem. By proving wellposedness of maximal (Dini) subsolutions for a class of second-order path-dependent PDEs with quadratic and even super-quadratic growth in the gradient, we establish first results related to this problem.
Non-Markovian large deviation problems and their connections to path-dependent PDEs are also studied in [MRTZ16]. In contrast to our work, in [MRTZ16] only the (limiting) rate function is characterized as a solution of a (first-order) path-dependent PDE. Moreover, the terminal condition is required to be Lipschitz continuous whereas we need only continuity.
2. Setup
2.1. Notation
Let . The canonical process on is denoted by , i.e., for each . Let be the (raw) filtration generated by . Given a probability measure on , denote by the -completion of the right-limit of .
We equip with the supremum norm and with the pseudo-metric defined by
Continuity and semi-continuity of functions defined on (resp. ) are to be understood with respect to (resp. ). Note that semi-continuous functions on are -progressive. From now on, we write l.s.c. (resp. u.s.c.) instead of lower semi-continuous (resp. upper semi-continuous).
With slight abuse of notation, we also use the notation to express the sup-norm for functions belonging to other function spaces.
We often identify vectors with constant functions, e.g., given a map , a vector , and a path , we write instead of .
Given , denote by be the unique probability measure on such that is a -dimensional standard -Wiener process starting at and that . We write for the corresponding expected value. Moreover, and .
As space of controls, the set of all bounded -progressive processes from to is used (whereas in [BLT] the controls are -progressive).
We denote by the effective domain of an extended real-valued function.
2.2. Data
Let and be measurable functions. We use the following hypotheses for .
(H1) The function satisfies the Tonelli-Nagumo condition, i.e., there is a function bounded from below with as such that on . Moreover, is l.s.c., proper, and convex for every .
(H2) .
These hypotheses are nearly identical with the corresponding condition (TI) for the Lagrangian in [BLT] (where it is denoted by ). In some of our main results, we will, in addition to (H1) and (H2), also assume that is continuous and finite-valued. In those cases, (TI) is satisfied as pointed out in [BLT].
2.3. The optimal control problems and HJB equations
Let . The value for our stochastic optimal control problem with initial data is given by
where and is a continuous process on defined by . The terminal value problem involving the corresponding HJB equation is
(TVP ) |
Remark 2.1.
If (H1) and (H2) hold and is bounded, then .
The value for our deterministic optimal control problem with initial data is given by
where
Here, , , is the Sobolev space of all that have a weak derivative . The terminal value problem involving the corresponding HJB equation is
(TVP) |
3. Notions of solutions of path-dependent HJB equations
We call a function non-anticipating if for every . Note that whenever a function on is l.s.c. or u.s.c. (with respect to ), then it is automatically non-anticipating.
3.1. Dini solutions
Given a non-anticipating function , we define the lower and upper Dini derivative
at points in direction . Here, in unison with the process in Subsection 2.3, .
The following (path-dependent) notion of Dini semi-solutions is motivated by the notion of (contingent) solutions used in Theorem 4.1 of [DM-F00] for HJB equations related to Bolza problems. Our notion is also related to the infinitesimal version of minimax solutions for path-dependent Isaacs equations used in [Lukoyanov01JAMM].
Definition 3.1.
Let .
(i) We call a Dini supersolution of (TVP) if is l.s.c., , and, for every with ,
(3.1) |
Example 3.2.
Let , , where , and be defined by . Then satisfies
Note that, for each with and each , we have for which the constant perturbation involving is responsible. (To obtain a finite value for our Dini derivative, we would need to permit the non-constant perturbation .) Thus is never satisfied. Hence, in our example there does not exist a Dini supersolution of (TVP). This justifies the need for an appropriate weaker notion of solution (see Subsection 3.2).
Given a non-anticipating function , we define the upper stochastic Dini derivative
at points in direction (cf. [SubbotinaEtAl85, Subbotina06, Haussmann92MOR, Haussmann94SICON]).
The following notion of subsolutions for second order path-dependent PDEs is motivated by the minimax solutions used in [SubbotinaEtAl85, Subbotina06] in a Markovian framework.
Definition 3.3.
Remark 3.4.
In our specific setting, the use of Dini type semiderivatives such as those introduced in this section suffices. This motivates us to call our generalized solutions Dini solutions. For more general data, path-dependent counterparts of contingent derivatives such as the Clio derivatives in [AubinHaddad02PPDE] need to be utilized. Corresponding generalized solutions would be called contingent solutions.
3.2. Minimax solutions
Here, we introduce a weaker notion of solution, which is an adjustment of the infinitesimal notion of minimax solutions in [BK18JFA]. It is also motivated by the notion of (l.s.c. contingent) solutions used in Theorem 5.1 of [DM-F00]. The problem in Example 3.2, which partially motivated this weaker notion, is overcome by allowing non-constant perturbations (see also the notion of contingent solutions in [Carja12SICON] that are defined via contingent derivatives with function-valued directions).
3.3. Consistency with classical solutions
First, we provide the definitions for path derivatives. The first-order ones are due to Kim [KimBook] and the second-order ones are due to Dupire [dupirefunctional]. Our presentation follows [ETZ_I] and [Lukoyanov03] .
Definition 3.6.
Let .
(i) We write if and if there are functions and called first-order path derivatives such that, for every , every and every , we have
(ii) We write if with corresponding first-order path derivatives and and if there is a function called second-order path derivative such that, for every , every probability measure on such that is a -Itô-semimartingale after time with bounded characteristics and with , and every , we have
Here, is the quadratic variation of and, given matrices , , is the trace of .
Remark 3.7.
If , then its first-order path derivatives are uniquely determined. If, in addition, , then its second-order path-derivative is uniquely determined as well. We refer to Section 2.3 of [ETZ_I] for more details.
Definition 3.8.
Let .
(i) We call a classical subsolution (resp. classical supersolution, classical solution) of (TVP) if , (resp. , ), and, for every ,
(ii) We call a classical subsolution (resp. classical supersolution, classical solution) of (TVP ) if (resp. , ) and, for every ,
Proposition 3.9 (Consistency of Dini solutions with classical solutions).
Assume that is continuous.
Proposition 3.10 (Partial consistency of l.s.c. minimax solutions with classical solutions).
Assume that is continuous and real-valued. Let .
Remark 3.11.
Proof of Proposition 3.10.
Part (b) follows immediately from Definition 3.5 (i) and Definition 3.8 (i). It remains to prove part (a). To this end, fix and assume that is an l.s.c. minimax subsolution of (TVP). Fix . Fix . Then, for any with a continuous derivative that satisfies , we have
Since , is continuous, and was arbitrary in , we have
i.e., is classical subsolution of (TVP). ∎
4. Main results
Theorem 4.1.
Assume (H1) with for some and (H2). Let be continuous and bounded. Then converges to uniformly on compacta and is continuous. Moreover, is the unique l.s.c. minimax solution of (TVP) that is bounded from below. If, in addition, is continuous and finite-valued, then we have the following:
(i) For each , the function is the unique bounded maximal Dini subsolution of (TVP )
(ii) The function is the unique bounded maximal Dini subsolution of (TVP).
Proof.
Remark 4.2.
Well-posedness of (TVP ) requires to be only u.s.c. and bounded (see Theorem 7.2). The corresponding result in the Markovian case treated in [BLT] is of similar strength (well-posedness holds for maximal viscosity supersolutions of the corresponding viscous Hamilton-Jacobi equations, which is due to [DrapeauMainberger16EJP]).
Theorem 4.3.
Assume (H1).
(a) Let be l.s.c., proper, and bounded from below. Then the value function is the unique l.s.c. minimax solution of (TVP) that is bounded from below.
(b) Assume (H2). Let be continuous and finite-valued. Let be u.s.c. and bounded. Then is the unique maximal bounded Dini subsolution of (TVP).
Proof.
See Section 8. ∎
5. Proof of the convergence result
Consider the semicontinuous envelopes and defined by
for every . Here, is the open -neighborhood of in .
First, we establish an auxiliary result.
Lemma 5.1.
Assume (H1). Let in as . Consider a probability space . Let be a sequence in that converges weakly to some . Then .
Proof.
We follow the lines of the proof of the corresponding deterministic closure theorem 10.8.ii in [Cesari]. First, note that, by lower semi-continuity and convexity of , for every , we have (more details can be found the proof of Lemma A.1 of [BLT]). Next, fix . Since is bounded from below, there is a independent from such that, for all , we have and thus . Again, as is bounded from below, either the right-hand side of the previous inequality converges to as or otherwise the left-hand side equals . This concludes the proof as was chosen arbitrarily. ∎
The following two statements adapt Theorem 2.2 in [BLT] to our slightly more general setting.
Lemma 5.2.
Assume (H1) with for some and (H2). Let be l.s.c. and bounded from below. Then .
Proof.
Let . It suffices to consider the case . Let be a sequence in that converges to in and that satisfies . Let be a sequence in such that each belongs to and is an -minimizer of with initial data . Then there exists a subsequence of with
and, for all ,
Since and are bounded from below, we can assume that
for some , with (cf. Theorem 11.1.i and its proof in [Cesari]). As and for all , one can proceed nearly exactly as in the proof of Lemma A.1 in [BLT] to show that the probability measures are tight. Let us point out the differences to [BLT]. Thanks to our additional requirement that the function from Hypothesis (H1) satisfies for some , we can estimate instead of (cf. with the first displayed equation in the proof of Lemma A.1 in [BLT]) and thus we can invoke Lemma 2 in [Zheng85] to obtain tightness (cf. also with the proof of Lemma 3.13 in [TanTouzi13]) instead of using the Aldous tightness criterion (Theorem 16.11 in [Kallenberg2nd]). Let us also note that the sequence is weakly convergent because for each ,
as , i.e., a sequence of copies of converges in distribution to the constant . Consequently, the probability measures are tight. Thus, by Skorohod’s representation theorem, there exists a probability space with a sequence of -valued random variables that satisfies for each and that converges (after passing to a subsequence), -a.s., to some random variable . Next, define a sequence of -valued processes on by . Again as in the proof of Lemma A.1 in [BLT], one can deduce that is equiabsolutely integrable and thus has a subsequential weak limit in that we denote by and that satisfies, by Lemma 5.1,
as well as , -a.s., for every . Moreover, , -a.s. Hence, together with being l.s.c. and =, -a.s., we have
(5.1) |
To conclude the proof, it suffices to note that there exists some such that the right-hand side (5.1) is greater than or equal to ) (cf. also Remark 2.6 of [HaussmannLepeltier90SICON]). ∎
Lemma 5.3.
Assume (H1) and (H2). Let be u.s.c. and bounded. Then .
Proof.
It suffices to follow the arguments of the first paragraph of the proof of Theorem 2.2 in [BLT] and make the obvious adjustments. For the convenience of the reader, we quickly go over it. Fix and with . Given , define for and for . Next, consider a sequence that converges to in and satisfies . Then
as is u.s.c. as well as bounded and , which follows from , (H2), and the convexity of (cf. the first displayed equation after (38) in [BLT]). Finally, letting concludes the proof as in [BLT]. ∎
Remark 5.4 (No Lavrentiev phenomenon).
Assume (H1) and (H2). Let be continuous and bounded. Then
where
This result follows from the proofs of Lemmata 5.2 and 5.3 but with each , , replaced by the unique probability measure under which a.s. For a more direct proof, it suffices to slightly modify the proof of Proposition 4.1 in [ButtazzoBelloni95] (this result is due to [DeArcangelis89]), where the case is treated.
6. Connections to BSDEs
We present parts of the theory of (convex) superquadratic BSDEs from [DrapeauEtAl13AOP] that are relevant for our work. Fix , , , and a -measurable random variable . Consider the BSDE
(6.1) |
Definition 6.1.
A pair is a supersolution of (6.1) if
-
•
is a càdlàg and -adapted process,
-
•
is an -predictable process with
-
•
, , is an -supermartingale, and,
-
•
for every , with , we have, -a.s.,
Proposition 6.2.
Assume (H1) and (H2). Let be bounded. Then (6.1) has a unique minimal supersolution and we write
(6.2) |
Proof.
By the proof of Theorem 3.4 in [DrapeauEtAl16AIHP] and by Theorem A.1 in [DrapeauEtAl16AIHP], the set of supersolutions of (6.1) is non-empty. Thus we can apply Theorem 4.17 in [DrapeauEtAl13AOP], which concludes the proof. ∎
From now on, let in this section for the sake of simplicity. Fix . Recall that . Given , define by . We also identify a control with the function , , where . Recall that , , and that .
For the next statement, we borrow the following constructions from [BLT]. Given , , with , a path is defined by
(6.3) |
(note that in [BLT] is used instead of ) and a path is defined by
Moreover, we use the function defined by
(6.4) |
Also set , , and .
Lemma 6.3.
Assume (H1) and (H2). Let be bounded. Then
(6.5) |
The calculations in our proof are essentially the same as those in the proof of Lemma 5.1 in [BLT]. Although of similar nature, the corresponding statements are different. This permits a more elementary proof in our case.
Proof of Lemma 6.3.
One should keep in mind, that , -a.s. We only consider the case as otherwise (6.5) is clearly satisfied.
The following statement together with Theorem 4.1 provide a non-Markovian non-linear Feynman-Kac formula that connects maximal Dini subsolutions of path-dependent PDEs with minimal supersolutions of convex superquadratic BSDEs, for which we use the notation (6.2).
Theorem 6.4.
Assume (H1) and (H2). Let be continuous and finite-valued. Let be bounded. Fix and . Then . Moreover, for -a.e. and every , we have .
Proof.
We prove the theorem only for the case and .
We will employ expressions , being a measurable function from to that satisfies (H1) and (H2) in place of , from Section 2 of [BLT], which are defined by
where is the set of all probability measures on that are absolutely continuous with respect to and is the unique -progressive process with , -a.s., that satisfies
(Section 2 of [BLT]).
First note that the right-hand side of (6.5) equals according to (BBD) in [BLT]. Moreover, by a straight-forward adjustment of the proof of Lemma 5.1 in [BLT], for every and -a.e. . In this context, note that Lemma 5.1 and Lemma A.2, both in [BLT], formally require continuity of , which, however, is not necessary (cf. the corresponding material in [DrapeauEtAl16AIHP]). Using Lemma 6.3, we can deduce that , -a.s., for every . ∎
7. The second order HJB equations
Lemma 7.1.
Let . Assume (H1). Let be continuous in . A bounded u.s.c. function is a Dini subsolution of (TVP ) if and only if and, for every , , and ,
(7.1) |
Proof.
We adapt the proof of Theorem V.3 in [Subbotina06], which is situated in a Markovian context with bounded control spaces, to our non-Markovian setting with the unbounded control space .
(i) Let be a Dini subsolution of (TVP ). Fix . Put and . Assume that there are and , and such that
(7.2) |
Consider the set of all -stopping times that satisfy , -a.s., and
(7.3) |
We equip with an order defined by if and only if , -a.s. Now, let be a totally ordered non-empty subset of . Note that there exists a sequence in such that as . Since is totally ordered and is linear, is increasing and converges to . We have because
thanks to (7.3) as well as to being u.s.c and bounded. Hence, by Zorn’s lemma, has a maximal element . We show that , -a.s. To this end, assume that . Define a set-valued function on by
The sets are non-empty because they contain and they are compact because, for any sequence in that converges to some , we have
as is u.s.c. and bounded. Moreover, for every sequence that belongs to and that converges to some with respect to the metric and hence also with respect to , every sequence in with has a subsequential limit that belongs to because, for every subsequential limit of (there is at least one), we have, by possibly passing to a subsequence,
Hence, considered as a map from equipped here with the metric into the set of all compact non-empty subsets of is u.s.c. (Theorem 2.2 on p. 31 in [Kisielewicz91DI]) and thus measurable (Theorem 2.1 on p. 29 in [Kisielewicz91DI]), i.e., all sets of the form , being a closed subset of , belong to (see p. 41 in [Kisielewicz91DI]). Consequently, by Theorem 3.13 on p. 49 in [Kisielewicz91DI], there exists a -measurable selector of that satisfies for every . Therefore, the map
is -measurable. Also note that, since, by (3.3) and the continuity of , every set contains a strictly positive element, we have the inequality and hence . Thus, by Lemma V.3 in [Subbotina06], is an -stopping time with on . Finally, since
and since, for each , thanks to and thanks to the definition of ,
we have
where the last inequality follows from (7.3) with . We can conclude that . Since and , we have but not , -a.s., i.e., we have a contradiction to being the maximal element in . Hence, , -a.s. However, this in turn contradicts (7.2), which concludes the proof of this direction.
(ii) Showing the remaining direction is straight-forward. ∎
Theorem 7.2.
Proof.
(i) Existence and regularity: First, we show that is bounded and u.s.c. Boundedness of , (H1), and (H2) yield boundedness of . To establish upper semi-continuity, we will use defined by (6.3) and the notation (6.4). For this reason, we assume that and but this assumption is not restrictive. We continue by fixing a pair and considering a sequence in that converges to and satisfies . We distinguish between two cases.
Case 1: . Fix an such that . We can assume, without loss of generality, that for all . Let . Then the map (see (6.4))
from to has a real upper bound due to being continuous and being bounded. Note that , , is continuous for every with . Thus, by Lemma 6.3,
Since was arbitrary, we can deduce after again invoking Lemma 6.3 that .
Case 2: . Then, proceeding similarly as in Case 1 but using the constant control , we have
We can conclude that is u.s.c.
Next, we establish the subsolution property via the BSDE connection in Theorem 6.4. To this end, we use the notation (6.2) and fix , , and . Note first that, by Proposition 3.6 (1) in [DrapeauEtAl13AOP],
(7.4) |
Since is bounded and, by Theorem 6.4, , -a.s., we can apply Theorem 3.4 in [DrapeauEtAl16AIHP] (see also Lemma A.2 in [BLT]) together with (3) in [BLT] to deduce that, with ,
Thus, by (7.4) and Theorem 6.4, satisfies (7.1) in place of . Hence, by Lemma 7.1, is a Dini subsolution of (TVP ).
(ii) Uniqueness: Let be a bounded Dini subsolution of (TVP ). Without loss of generality, let . Fix and . It suffices to show that . By (a slight modification) of (36) and (BBD), both in [BLT], there exists an with , where and each is -measurable, such that . Here, for any , , , and is the unique solution of , , with initial condition . Next, note that and on , -a.s. Thus, by Lemma 7.1,
Also note that, for -a.e. , we have, with , and on , -a.s., and thus, by Lemma 7.1,
Therefore
Repeating this procedure and noting that yields
which concludes the proof. ∎
8. The first order HJB equation
Lemma 8.1.
Assume (H1). An l.s.c. function bounded from below is a minimax supersolution of (TVP) if and only if and, for every and , there exists an such that
(8.1) |
Proof.
(i) Let be a minimax supersolution of (TVP). Assume that there exist a with and a such that, for all , . Then there is an such that
(8.2) |
for all because has a minimizer in (this can be shown exactly as the corresponding statement for the non-path-dependent counterpart Theorem 11.1.i in Section 11.1 of [Cesari] and its extension to unbounded domains, which is treated in Section 11.2 of [Cesari]). Next, consider the (non-void) set of all for which holds. Let . This supremum is attained. To see this, consider a sequence in with . Since and is bounded from below, one can show as in Sections 11.1 and 11.2 of [Cesari] that there is a subsequence such that converges to some in , i.e., . Lower semi-continuity of together with Theorem 10.8.ii in [Cesari] yield
Thus and, by (8.2), . Hence, by (3.4), there is a and an such that and . Thus , which is a contradiction to the maximality of .
(ii) Fix with . Suppose that, for every for every , there exists an such that (8.1) holds with replaced by . Then one can proceed similarly as in the proof of Lemma 3.6 in [BK18JFA] to show that there exists a sequence in and an increasing sequence of finite subsets of whose union is dense in such that, for each and every , (8.1) holds with replaced by . Thus , where is a lower bound of . Note that we are in a similar situation as in part (i) of this proof and thus it can be shown in the same way that there is a subsequence of that converges to some in . Now, fix and a sequence in with for each and with as . By Theorem 10.8.ii in [Cesari] and lower semi-continuity of ,
i.e., there is an such that, for every , (8.1) holds. From this point, (3.4) follows easily. ∎
Lemma 8.2.
Assume (H1). An l.s.c. function is an l.s.c. minimax subsolution of (TVP) if and only if and, for every , , and , we have
(8.3) |
Proof.
(i) Let be an l.s.c. minimax subsolution of (TVP). For the sake of a contradiction, assume that there exist a , a , an , and an such that
(8.4) |
as well as and for all . Put
We show that this infimum is attained. To this end, consider a sequence in with and . Then
i.e., is a minimum and . Moreover, by (8.4), . Finally, by (3.5), there is a such that . Hence, , which is a contradiction to the minimality of .
(ii) Showing the remaining direction is straight-forward. ∎
Proof of Theorem 4.3.
(a) Establishing the lower semi-continuity of is quite standard. It is very similar to the proof of Lemma 5.2 and actually slightly easier as no probability is involved (cf. also Proposition 3.1 in [DM-F00] for the non-path-dependent case). To deduce that is an l.s.c. minimax solution, it suffices to apply the dynamic programming principle with the existence of a minimizer for (cf. Theorem 11.1.i in [Cesari]) together with Lemmata 8.1 and 8.2. Finally, we can apply Lemmata 8.1 and 8.2 again to obtain a comparison principle between l.s.c. minimax subsolutions and minimax supersolutions, from which uniqueness follows.
(b) Taking Remark 5.4 into account, one can see that the proof follows essentially from the content of Section 7. The measures need to be replaced by the Dirac measures under which a.s. for each and one should note that the domains of the controls are different ( here vs. in Section 7). Moreover, the BSDE argument in part (i) of the proof of Theorem 7.2 needs to be replaced by the deterministic dynamic programming principle. ∎
9. Conclusion
The main contributions of this paper are a non-Markovian vanishing viscosity result for path-dependent PDEs (PPDEs) that corresponds to the non-exponential Schilder theorem in [BLT], well-posedness for new notions of generalized solutions of PPDEs that can have quadratic or even super-quadratic growth in the gradient, and a non-Markovian Feynman-Kac formula for convex superquadratic BSDEs.
We want to emphasize that, here, control-theoretic methods, or equivalently (in our case), results from the theory of large deviations have been applied to obtain stability results for PPDEs (corresponding PDEs results have been obtained in [BLT] in the Markovian case). Of great interest would be research that investigates the opposite direction, i.e., to establish stronger results (in particular, suitable stability results) for PPDEs with (only) quadratic growth in the gradient in order to derive non-Markovian large deviation results similarly as it has been successfully done in the Markovian case via PDEs.