• Aucun résultat trouvé

CRANDALL-LIONS VISCOSITY SOLUTIONS FOR PATH-DEPENDENT PDES: THE CASE OF HEAT EQUATION

N/A
N/A
Protected

Academic year: 2021

Partager "CRANDALL-LIONS VISCOSITY SOLUTIONS FOR PATH-DEPENDENT PDES: THE CASE OF HEAT EQUATION"

Copied!
56
0
0

Texte intégral

(1)

HAL Id: hal-02383626

https://hal.archives-ouvertes.fr/hal-02383626v3

Preprint submitted on 26 Jan 2021

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of

sci-entific research documents, whether they are

pub-lished or not. The documents may come from

teaching and research institutions in France or

abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est

destinée au dépôt et à la diffusion de documents

scientifiques de niveau recherche, publiés ou non,

émanant des établissements d’enseignement et de

recherche français ou étrangers, des laboratoires

publics ou privés.

PATH-DEPENDENT PDES: THE CASE OF HEAT

EQUATION

Andrea Cosso, Francesco Russo

To cite this version:

Andrea Cosso, Francesco Russo.

CRANDALL-LIONS VISCOSITY SOLUTIONS FOR

(2)

Crandall-Lions Viscosity Solutions for

Path-Dependent PDEs: the Case of Heat

Equation

Andrea Cosso∗ Francesco Russo

We address our interest to the development of a theory of viscosity solutions `a la Crandall-Lions for path-dependent partial differential equations (PDEs), namely PDEs in the space of continuous paths C([0, T ]; Rd

). Path-dependent PDEs can play a central role in the study of certain classes of optimal control problems, as for instance optimal control problems with delay. Typically, they do not admit a smooth solution satisfying the corresponding HJB equation in a classical sense, it is therefore natural to search for a weaker notion of solution. While other notions of generalized solution have been proposed in the literature, the extension of the Crandall-Lions framework to the path-dependent setting is still an open problem. The question of uniqueness of the solutions, which is the most delicate issue, will be based on early ideas from the theory of viscosity solutions and a suitable variant of Ekeland’s variational principle. This latter is based on the construction of a smooth gauge-type function, where smooth is meant in the horizontal/vertical (rather than Fr´echet) sense. In order to make the presentation more readable, we address the path-dependent heat equation, which in particular simplifies the smoothing of its natural “candidate” solution. Finally, concerning the existence part, we provide a functional Itˆo formula under general assumptions, extending earlier results in the literature. MSC 2010 subject classifications:35D40, 35R15, 60H30.

Keywords:Path-dependent partial differential equations, viscosity solutions, functional Itˆo for-mula.

1. Introduction

Path-dependent heat equation refers to the second-order partial differential equation in the space of continuous paths

     −∂H t v(t, x) − 1 2tr∂ V xxv(t, x)  = 0, (t, x) ∈ [0, T ) × C([0, T ]; Rd), v(T, x) = ξ(x), x∈ C([0, T ]; Rd). (1.1)

Here C([0, T ]; Rd) denotes the Banach space of continuous paths x : [0, T ] → Rdequipped

with the supremum norm kxk∞:= supt∈[0,T ]|x(t)|, with |·| denoting the Euclidean norm

University of Bologna, Dipartimento di Matematica, Piazza di Porta S. Donato, 5, 40126 Bologna,

Italy, andrea.cosso at unibo.it

ENSTA Paris, Institut Polytechnique de Paris, Unit´e de Math´ematiques Appliqu´ees, 828, boulevard

des Mar´echaux, 91120 Palaiseau, France, francesco.russo at ensta-paris.fr

(3)

on Rd. The terminal condition ξ : C([0, T ]; Rd) → R is assumed to be continuous and

bounded. We refer to equation (1.1) as path-dependent heat equation. Similarly as for the usual heat equation, it admits the following Feynman-Kac representation formula in terms of the d-dimensional Brownian motion W = (Ws)s∈[0,T ]:

v(t, x) = Eξ(Wt,x), ∀ (t, x) ∈ [0, T ] × C([0, T ]; Rd), (1.2) where Wt,xs := ( x(s), s ≤ t, x(t) + Ws− Wt, s > t.

In the case of the classical heat equation ξ only depends on the terminal value Wt,xT . The peculiarity of equation (1.1) is the presence of the so-called functional or path-wise derivatives ∂H

t v, ∂xxV v, where ∂tHv is known as horizontal derivative, while ∂xxV v is

the matrix of second-order vertical derivatives. Those derivatives appeared in [56, 57] (under the name of coinvariant derivatives) as building block of the so-called i-smooth analysis, and independently in [1], where they were denoted Clio derivatives; later, they were rediscovered by [31] (from which we borrow terminology and definitions), which adopted a slightly different definition based on the space of c`adl`ag paths and in addi-tion developed a related stochastic calculus, known as funcaddi-tional Itˆo calculus, including in particular the so-called functional Itˆo formula. Differently from the classical Fr´echet derivative on C([0, T ]; Rd), the distinguished features of the pathwise derivatives are their

finite-dimensional nature and the property of being non-anticipative, which follow from the interpretation of t in x(t) as time variable. This means that v(t, x) only depends on the values of the path x up to time t; moreover, the horizontal and vertical derivatives at time t are computed keeping the past values frozen, while only the present value of the path (that is x(t)) can vary. The related functional Itˆo calculus was rigorously in-vestigated in [12, 13, 14]. [15, 18] also gave a contribution in this direction, exploring the relation between pathwise derivatives and Banach space stochastic calculus, built on an appropriate notion of Fr´echet type derivative and firstly conceived in [26], see also [27,28,29,25].

Partial differential equations in the space of continuous paths (also known as func-tional or Clio or path-dependent partial differential equations) are mostly motivated by optimal control problems of deterministic and stochastic systems with delay (or path-dependence) in the state variable. Such control systems arise in many fields, as for instance optimal advertising theory [49, 50], chemical engineering [46], financial man-agement [39, 74], economic growth theory [2], mean field game theory [5], biomedicine [47,84], systemic risk [10]. The underlying deterministic or stochastic controlled differen-tial equations with delay can be studied in two ways: first using a direct approach (see for instance [52, 86,56, 53, 58]), second by lifting them into a suitable infinite-dimensional framework, leading to evolution equations in Hilbert (as in [11,24,42]) or Banach spaces (as in [71,72, 26]). The latter methodology turned out to be preferable to address gen-eral optimal control problems with delay (see for instance [92, 51, 49, 40, 41,45, 38]), although such an infinite-dimensional reformulation may require some additional artifi-cial assumptions to be imposed on the original control problem. On the other hand, the

(4)

direct approach was adopted for special problems where the Hamilton-Jacobi-Bellman equation reduces to a finite-dimensional differential equation, as in [36, 59]. This ap-proach can now regain relevance thanks to a well-grounded theory of path-dependent partial differential equations. To this regard, the path-dependent heat equation repre-sents the primary test for such a theory, it indeed requires the main building blocks of the methodology, without overloading the proofs with additional technicalities.

Path-dependent partial differential equations represent a quite recent area of research. Typically, they do not admit a smooth solution satisfying the equation in a classical sense, mainly because of the awkward nature of the underlying space C([0, T ]; Rd). This

happens also for the path-dependent heat equation, which in particular does not have the smoothing effect characterizing the classical heat equation, except in some specific cases (as shown in [26, 30]) with ξ belonging to the class of so-called cylinder or tame functions (therefore depending specifically on a finite number of integrals with respect to the path) or ξ being smoothly Fr´echet differentiable. It is indeed quite easy, relying on the probabilistic representation formula (1.2), to see that the function v is not smooth (in the horizontal/vertical sense mentioned above) for terminal conditions of the form

ξ(x) = sup

0≤t≤T

x(t), ξ(x) = x(t0),

for some fixed t0∈ (0, T ). For a detailed analysis of the first case above we refer to Section

3.2 in [17], see also Remark 3.8 in [18]. Concerning the second case, see for instance Example 11.1.3 of [93]. It is however worth mentioning that some positive results on smooth solutions were obtained in [18, 78]. We also refer to Chapter 9 of [26] and [30], where smooth solutions were investigated using a Fr´echet type derivative formulation.

It is therefore natural to search for a weaker notion of solution, as the notion of vis-cosity solution, commonly used in the standard finite-dimensional case. The theory of viscosity solutions, firstly introduced in [21,22] for first-order equations in finite dimen-sion and later extended to the second-order case in [61, 62, 63], provides a well-suited framework guaranteeing the desired existence, uniqueness, and stability properties (for a comprehensive account see [20]). The extension of such a theory to equations in infinite dimension was initiated by [23, 64,65, 66, 87,90]. One of the structural assumption is that the state space has to be a Hilbert space or, slightly more general, certain Banach space with smooth norm, not including for instance the Banach space C([0, T ]; Rd)

(no-tice however that in this paper we do not directly generalize those results to C([0, T ]; Rd),

as we adopt horizontal/vertical, rather than Fr´echet, derivatives on C([0, T ]; Rd)).

First-order path-dependent partial differential equations were deeply investigated in [70] using a viscosity type notion of solution, which differs from the Crandall-Lions def-inition as the maximum/minimum condition is formulated on the subset of absolutely continuous paths. Such a modification does not affect existence in the first-order case, however it is particularly convenient for uniqueness, which is indeed established under general conditions. Other notions of generalized solution designed for first-order equations were adopted in [1] as well as in [67,68, 69], where the minimax framework introduced in [88,89] was implemented. We also mention [4], where such a minimax approach was extended to first-order path-dependent Hamilton-Jacobi-Bellman equations in infinite

(5)

di-mension. Concerning the second-order case, a first attempt to extend the Crandall-Lions framework to the path-dependent case was carried out in [76], even though a technical condition on the semi-jets was imposed, namely condition (16) in [76], which narrows down the applicability of such a result. In the literature, this was perceived as an almost insurmountable obstacle, so that the Crandall-Lions definition was not further investi-gated, while other notions of generalized solution were devised, see [33,77,91,19,60,3,9]. We mention in particular the framework designed in [33] and further investigated in [34, 35,81, 82, 83,16], where the notion of sub/supersolution adopted differs from the Crandall-Lions definition as the tangency condition is not pointwise but in the sense of expectation with respect to an appropriate class of probability measures. On the other hand, in [19] we introduced the so-called strong-viscosity solution, which is quite similar to the notion of good solution for partial differential equations in finite dimension, that in turn is known to be equivalent to the definition of Lp-viscosity solution, see for

in-stance [54]. We also mention [3], where the authors deal with semilinear path-dependent equations and propose the notion of decoupled mild solution, formulated in terms of gen-eralized transition semigroups; such a notion also adapts to path-dependent equations with integro-differential terms.

In the present paper we adopt the natural generalization of the well-known definition of viscosity solution `a la Crandall-Lions given in terms of test functions and, under this notion, we establish existence and uniqueness for the path-dependent heat equation (1.1). The uniqueness property is derived, as usual, from the comparison theorem. The proof of this latter, which is the most delicate issue, is known to be quite involved even in the classical finite-dimensional case (see for instance [20]), and in its latest form is based on Ishii’s lemma. Here we follow instead an earlier approach (see for instance Theorem II.1 in [63] or Theorem IV.1 in [64]), which in principle can be applied to any path-dependent equation admitting a “candidate” solution v, for which a probabilistic representation formula holds. This is the case for equation (1.1), where the candidate solution is given by formula (1.2), but it is also the case for Kolmogorov type equations or, more generally, for Hamilton-Jacobi-Bellman equations. This latter is the class of equations studied in [63] and [64], whose methodology in a nutshell can be described as follows. Let u (resp. w) be a viscosity subsolution (resp. supersolution) of the same path-dependent equation. The desired inequality u ≤ w follows if we compare both u and w to the “candidate” solution v, that is if we prove the two inequalities u ≤ v and v ≤ w. Let us consider for instance the first inequality u ≤ v. In the non-path-dependent and finite-dimensional case (as in [63]), this is proved proceeding as follows: firstly, performing a smoothing of v through its probabilistic representation formula; secondly, taking a local maximum of u − vn (here it is used the local compactness of the finite-dimensional underlying

space), with vn being a smooth approximation of v; finally, the inequality u ≤ vn is

proved proceeding as in the so-called partial comparison theorem (comparison between a viscosity subsolution/supersolution and a smooth supersolution/subsolution), namely exploiting the viscosity subsolution property of u with vnplaying the role of test function.

In [64], where such a methodology was extended to the infinite-dimensional case, the existence of a maximum of u − vn is achieved relying on Ekeland’s variational principle,

(6)

In this paper we generalize the methodology sketched above to the path-dependent case. There are however at least two crucial mathematical issues required by such a proof, still not at disposal in the path-dependent framework.

Firstly, given a candidate solution v, it is not a priori obvious how to perform a smooth approximation of v itself starting from its probabilistic representation formula. Here we exploit the results proved in [18] (Theorem 3.5) and [19] (Theorem 3.12), which are re-ported and adapted to the present framework in Appendix D(Lemma D.1and Lemma D.2, respectively). Notice that such results apply to the case of the path-dependent heat equation (1.1), where there is only the terminal condition ξ in the probabilistic repre-sentation formula (1.2) for v. More general results are at disposal in [18] and [19], which cover the case of semilinear path-dependent partial differential equations, characterized by the presence of four coefficients b, σ, F , ξ (see, in particular, Theorem 3.16 in [19] for more details). However, when those other coefficients appear in the path-dependent partial differential equation, we need more information on the sequence {vn}n

approxi-mating v. For instance, we also have to estimate the derivatives of vnin order to proceed

as in [63] or [64]. Since such results are still not at disposal in the path-dependent setting, in order to make the paper more readable and not excessively lengthy, here we address the case of the path-dependent heat equation.

Secondly, concerning the existence of a maximum of u − vn, we rely on a generalized

version of Ekeland’s variational principle for which we need a smooth gauge-type function with bounded derivatives, as explained below. Our equation is in fact formulated on the non-locally compact space [0, T ] × C([0, T ]; Rd) endowed with the pseudometric

d∞ (t, x), (t′, x′) := |t − t′| + kx(· ∧ t) − x′(· ∧ t′)k∞.

Recall that Ekeland’s variational principle, in its original form, applied to ([0, T ] × C([0, T ]; Rd), d

∞) states that a perturbation u(·, ·) − vn(·, ·) − δd∞((·, ·), (¯t, ¯x)) of u(·, ·) −

vn(·, ·) has a strict global maximum, with the perturbation being expressed in terms of

the distance d∞ (the point (¯t, ¯x) is fixed). As the map (t, x) 7→ d∞((t, x), (¯t, ¯x)) is not

smooth, it cannot be a test function. In order to have a smooth map instead of d∞,

we need a smooth variational principle on [0, T ] × C([0, T ]; Rd). To this end, the starting

point is a generalization of the so-called Borwein-Preiss smooth variant of Ekeland’s vari-ational principle (see for instance [8]), which works when d∞ is replaced by a so-called

gauge-type function (see Definition 3.1). For the proof of the comparison theorem, we have to construct a gauge-type function which is also smooth and with bounded deriva-tives, recalling that smooth in the present context means in the horizontal/vertical (rather than in the Fr´echet) sense. In Section3 such a gauge-type function is built through a smoothing of d∞itself (more precisely, of the part concerning the supremum norm). This

latter smoothing is performed by convolution, firstly in the vertical direction, that is in the direction of the map 1[t,T ] (Lemma 3.1), then in the horizontal direction (Lemma

3.2), the ordering of smoothings being crucial. Notice in particular that the supremum norm is already smooth in the horizontal direction; however, after the vertical smoothing, we lose in general the horizontal regularity because of the presence of the term 1[t,T ];

(7)

gauge-type function with bounded derivatives corresponds to the function ρ∞defined in

(3.8).

Regarding existence, we prove that the candidate solution v in (1.2) solves in the vis-cosity sense equation (1.1). We proceed essentially as in the classical non-path-dependent case, relying as usual on Itˆo’s formula, which in the present context corresponds to the functional Itˆo formula. Such a formula was firstly stated in [31] and then rigorously proved in [12, 13], see also [14,43,18,60,73]. In the present paper we provide a functional Itˆo formula under general assumptions (Theorem2.2). In particular, we do not require any boundedness assumption on the functional u : [0, T ] × C([0, T ]; Rd) → R, thus improving

(when the semimartingale process is continuous) the results stated in [12,13].

The paper is organized as follows. Section 2 is devoted to pathwise derivatives and functional Itˆo calculus. In particular, there is the functional Itˆo formula (Theorem2.2) whose complete proof is reported in AppendixA. In Section3we prove the smooth vari-ational principle on [0, T ] × C([0, T ]; Rd), constructing the smooth gauge-type function

with bounded derivatives. In Section4we provide the (path-dependent) Crandall-Lions definition of viscosity solution for a general path-dependent partial differential equa-tion. We then study in detail the path-dependent heat equaequa-tion. In particular, we prove existence showing that the so-called candidate solution v solves in the viscosity sense the path-dependent heat equation (Theorem 4.1). We conclude Section 4 proving the comparison theorem (Theorem4.2) and uniqueness (Corollary4.1).

2. Pathwise derivatives and functional Itˆ

o calculus

In the present section we define the pathwise derivatives and prove the functional Itˆo formula under general assumptions.

2.1. Maps on c`

adl`

ag paths

Given T > 0 and d ∈ N∗, we denote by D([0, T ]; Rd) the set of c`adl`ag functions

ˆ

x: [0, T ] → Rd. We denote by ˆx(t) the value of ˆx at t ∈ [0, T ]. We also denote by 0

the function ˆx: [0, T ] → Rd identically equal to zero. We consider on D([0, T ]; Rd) the

supremum norm k·k∞, namely kˆxk∞:= supt∈[0,T ]|ˆx(t)|, where |·| denotes the Euclidean

norm on Rd (we use the same symbol | · | to denote the Euclidean norm on Rk, for any

k ∈ N). We refer to Chapter V in [79] and to Section 15 of Chapter 3 in [6] for a study of the set of c`adl`ag functions endowed with the uniform metric and a comparison with the Skorokhod space.

We set ˆΛ := [0, T ] × D([0, T ]; Rd) and define ˆd

∞: ˆΛ × ˆΛ → [0, ∞) as ˆ d∞ (t, ˆx), (t′, ˆx′)  := |t − t′| + ˆx(· ∧ t) − ˆx′(· ∧ t′) .

Notice that ˆd∞is a pseudometric on ˆΛ, that is ˆd∞is not a true metric because one may

(8)

true metric space ( ˆΛ∗, ˆd∗

∞), called the metric space induced by the pseudometric space

( ˆΛ, ˆd∞), by means of the equivalence relation which follows from the vanishing of the

pseudometric. We also observe that ( ˆΛ, ˆd∞) is a complete pseudometric space. Finally,

we denote by B( ˆΛ) the Borel σ-algebra on ˆΛ induced by ˆd∞.

Definition 2.1. A map (or functional) ˆu : ˆΛ → R is said to be non-anticipative (on ˆ

Λ) if it satisfies

ˆ

u(t, ˆx) = ˆu(t, ˆx(· ∧ t)), for all (t, ˆx) ∈ ˆΛ.

Remark 2.1. (i) The property of being non-anticipative is crucial and automatically true if the map ˆu : ˆΛ → R is continuous with respect to ˆd∞.

(ii) More generally, it holds that whenever ˆu : ˆΛ → R is Borel measurable, namely ˆu is measurable with respect to B( ˆΛ), then ˆu is non-anticipative on ˆΛ. As a matter of fact, notice that every open subset B of ˆΛ, endowed with ˆd∞, satisfies the following property: if

(t, ˆx) ∈ B then (t, ˆx(·∧t)) ∈ B (this follows from the fact that ˆd∞((t, ˆx), (t, ˆx(·∧t))) = 0).

As a consequence, by a monotone class argument, the same property holds true for every Borel subset of ˆΛ. Now, let ˆu : ˆΛ → R be Borel measurable. For every (t, ˆx) ∈ ˆΛ, denote

Bu(t,ˆˆ x) := (s, ˆy) ∈ ˆΛ : ˆu(s, ˆy) = ˆu(t, ˆx) .

Notice that Bu(t,ˆˆ x)∈ B( ˆΛ) and since (t, ˆx) ∈ Bu(t,ˆˆ x)we deduce that (t, ˆx(·∧t)) ∈ Bˆu(t,ˆx).

This means that ˆu(t, ˆx(· ∧ t)) = ˆu(t, ˆx), namely the map ˆu is non-anticipative.

Definition 2.2. We denote by C( ˆΛ) the set of maps ˆu : ˆΛ → R which are continuous on ˆΛ with respect to ˆd∞.

Definition 2.3 (Pathwise derivatives). Let ˆu : ˆΛ → R be non-anticipative.

(i) Given (t, ˆx) ∈ ˆΛ, with t < T , the horizontal derivative of ˆu at (t, ˆx) (if the corresponding limit exists) is defined as

∂tHu(t, ˆˆ x) := lim δ→0+

ˆ

u(t + δ, ˆx(· ∧ t)) − ˆu(t, ˆx)

δ .

At t = T the horizontal derivative is defined as ∂H

t u(T, ˆˆ x) := lim t→T−∂

H t u(t, ˆˆ x).

(ii) Given (t, ˆx) ∈ ˆΛ, the vertical derivatives of first and second-order of ˆu at (t, ˆx) (if the corresponding limits exist) are defined as

∂xViu(t, ˆˆ x) := limh→0

ˆ

u(t, ˆx+ h ei1[t,T ]) − ˆu(t, ˆx)

(9)

∂V xixju(t, ˆˆ x) := ∂ V xj(∂ V xiu)(t, ˆˆ x),

where e1, . . . , ed is the standard orthonormal basis of Rd.

Finally, we denote ∂V xu = (∂ˆ xV1u, . . . , ∂ˆ V xdu) and ∂ˆ V xxu = (∂ˆ xVixju)ˆ i,j=1,...,d.

Definition 2.4. We denote by C1,2( ˆΛ) the set of ˆu ∈ C( ˆΛ) such that ∂H

t u, ∂ˆ xVu, ∂ˆ xxV uˆ

exist everywhere on ˆΛ and are continuous.

For later use, we also introduce the following set of maps on c`adl`ag paths.

Definition 2.5. We denote by C0,2( ˆΛ) the set of ˆu ∈ C( ˆΛ) such that ∂V

xu, ∂ˆ xxV u existˆ

everywhere on ˆΛ and are continuous.

We can finally state the functional Itˆo formula for maps on c`adl`ag paths, whose proof is reported in AppendixA.

Theorem 2.1. Let ˆu ∈ C1,2( ˆΛ). Then, for every d-dimensional continuous semi-martingale X = (Xt)t∈[0,T ], where X = (X1, . . . , Xd), defined on some filtered

prob-ability space (Ω, F, (Ft)t∈[0,T ], P), with (Ft)t∈[0,T ]satisfying the usual conditions, the

fol-lowing functional Itˆo formulaholds:

ˆ u(t, X) = ˆu(0, X) + Z t 0 ∂tHu(s, X) ds +ˆ 1 2 d X i,j=1 Z t 0 ∂xVixju(s, X) d[Xˆ i, Xj] s + d X i=1 Z t 0 ∂xViu(s, X) dXˆ i

s, for all 0 ≤ t ≤ T, P-a.s.

Proof. See AppendixA.

2.2. Maps on continuous paths

Let C([0, T ]; Rd) denote the set of continuous functions x : [0, T ] → Rd. Notice that

C([0, T ]; Rd) is a subset of D([0, T ]; Rd). We set Λ := [0, T ] × C([0, T ]; Rd) and denote

d∞ the restriction of ˆd∞ to Λ × Λ. Then, d∞ is a pseudometric on Λ and (Λ, d∞) is a

complete pseudometric space. We denote by B(Λ) the Borel σ-algebra on Λ induced by d∞.

Definition 2.6. Let ˆu : ˆΛ → R be non-anticipative and consider u: Λ → R. We say that ˆu is consistent with u if

u(t, x) = ˆu(t, x), for all (t, x) ∈ Λ.

(10)

The following consistency property is crucial as it implies that, given u admitting two maps ˆu1 and ˆu2, both being consistent with u, their pathwise derivatives coincide on

continuous paths (see also Remark2.2).

Lemma 2.1. If ˆu1, ˆu2∈ C1,2( ˆΛ) satisfy ˆ u1(t, x) = ˆu2(t, x), ∀ (t, x) ∈ Λ, then ∂tHuˆ1(t, x) = ∂tHuˆ2(t, x), ∂V xuˆ1(t, x) = ∂xVuˆ2(t, x), ∂xxV uˆ1(t, x) = ∂xxV uˆ2(t, x), for all (t, x) ∈ Λ. Proof. See AppendixB.

Thanks to Lemma2.1we can now give the following definition (see also Remark2.2).

Definition 2.7. Let u : Λ → R. We say that u ∈ C1,2(Λ) if there exists ˆu : ˆΛ → R consistent with u and satisfying ˆu ∈ C1,2( ˆΛ). Moreover, we define

∂H t u(t, x) := ∂tHu(t, x),ˆ ∂xVu(t, x) := ∂ V xu(t, x),ˆ ∂xxV u(t, x) := ∂xxV u(t, x),ˆ for all (t, x) ∈ Λ.

Remark 2.2. Notice that, by Lemma 2.1, if u ∈ C1,2(Λ) then the definition of the pathwise derivatives of u is independent of the map ˆu ∈ C1,2( ˆΛ) consistent with u. Theorem 2.2. Let u ∈ C1,2(Λ). Then, for every d-dimensional continuous semi-martingale X = (Xt)t∈[0,T ], where X = (X1, . . . , Xd), defined on some filtered

prob-ability space (Ω, F, (Ft)t∈[0,T ], P), with (Ft)t∈[0,T ]satisfying the usual conditions, the

fol-lowing functional Itˆo formulaholds:

u(t, X) = u(0, X) + Z t 0 ∂tHu(s, X) ds + 1 2 d X i,j=1 Z t 0 ∂xVixju(s, X) d[X i, Xj] s (2.1) + d X i=1 Z t 0 ∂Vxiu(s, X) dX i

(11)

Proof. Since u ∈ C1,2(Λ), by Definition 2.7 there exists a map ˆu : ˆΛ → R consistent with u and satisfying ˆu ∈ C1,2( ˆΛ). Then, by Theorem2.1, the following functional Itˆo formula holds: ˆ u(t, X) = ˆu(0, X) + Z t 0 ∂tHu(s, X) ds +ˆ 1 2 d X i,j=1 Z t 0 ∂xVixju(s, X) d[Xˆ i, Xj] s + d X i=1 Z t 0 ∂xViˆu(s, X) dX i

s, for all 0 ≤ t ≤ T, P-a.s.

The claim follows identifying the pathwise derivatives of ˆu with those of u.

3. Smooth variational principle on Λ

The goal of the present section is the proof of a smooth variational principle on Λ, which plays a crucial role in the proof of the comparison theorem (Theorem4.2). To this end, we begin recalling a generalization of the so-called Borwein-Preiss smooth variant ([7]) of Ekeland’s variational principle ([32]), corresponding to Theorem3.1below. We state it for the case of real-valued (rather than R ∪ {+∞}-valued as in [8]) maps on Λ. We firstly recall the definition of gauge-type function for the specific set Λ.

Definition 3.1. We say that Ψ : Λ×Λ → [0, +∞) is a gauge-type function provided that the properties below hold:

a) Ψ is continuous on Λ × Λ;

b) Ψ((t, x), (t, x)) = 0, for every (t, x) ∈ Λ;

c) for every ε > 0 there exists η > 0 such that, for all (t′, x), (t′′, x′′) ∈ Λ, the inequality

Ψ((t′, x), (t′′, x′′)) ≤ η implies d

∞((t′, x′), (t′′, x′′)) < ε.

Theorem 3.1. Let G : Λ → R be an upper semicontinuous map, bounded from above. Suppose that Ψ : Λ × Λ → [0, +∞) is a gauge-type function (according to Definition 3.1) and {δn}n≥0 is a sequence of strictly positive real numbers. For every ε > 0, let

(t0, x0) ∈ Λ such that

sup G − ε ≤ G(t0, x0).

Then, there exists a sequence {(tn, xn)}n≥1 ⊂ Λ which converges to some (¯t, ¯x) ∈ Λ

satisfying the following properties.

i) Ψ((¯t, ¯x), (tn, xn)) ≤ 2δ0, for every n ≥ 0.

ii) G(t0, x0) ≤ G(¯t, ¯x) −Pn=0+∞δnΨ((¯t, ¯x), (tn, xn)).

iii) For every (t, x) 6= (¯t, ¯x), G(t, x) − +∞ X n=0 δnΨ (t, x), (tn, xn) < G(¯t, ¯x) − +∞ X n=0 δnΨ (¯t, ¯x), (tn, xn).

(12)

Proof. Theorem3.1follows trivially from Theorem 2.5.2 in [8], the only difference being that the latter result is stated on complete metric spaces, while here Λ is a complete pseudometric space.

The main ingredient of Theorem3.1is the gauge-type function Ψ. In the proof of the comparison theorem we need such a gauge-type function to be also smooth as a map of its first pair, namely (t, x) 7→ Ψ((t, x), (t0, x0)), and with bounded derivatives. The

most important example of gauge-type function is the pseudometric d∞ itself, which

unfortunately is not smooth enough. The major contribution of the present section is the construction of such a smooth gauge-type function with bounded derivatives, which corresponds to the function ρ∞ in (3.8). In order to do it, we perform a smoothing of

the pseudometric d∞ itself (more precisely of the part concerning the supremum norm),

first in the vertical direction, and then in the horizontal direction. In particular, the next result concerns the smoothing in the vertical direction. The precise form of the mollifier ζ in (3.1) is used to get explicit bounds on ˆκ(t0,x0)

∞ and its derivatives.

Lemma 3.1. Let ζ : Rd→ R be the probability density function of the standard normal

multivariate distribution ζ(z) := 1 (2π)d2 e−12|z| 2 , ∀ z ∈ Rd. (3.1)

For every fixed (t0, x0) ∈ Λ, define the map ˆκ(t∞0,x0): ˆΛ → R as

ˆ κ(t0,x0) ∞ (t, ˆx) := Z Rd ˆx(· ∧ t) − x0(· ∧ t0) − z 1[t,T ] ζ(z) dz − Z Rd|z| ζ(z) dz, (3.2)

for all (t, ˆx) ∈ ˆΛ. Moreover, let κ(t0,x0)

∞ : Λ → R be given by

κ(t0,x0)

∞ (t, x) := ˆκ(t∞0,x0)(t, x),

for every (t, x) ∈ Λ. Then, the following properties hold.

1) For every (t, ˆx) ∈ ˆΛ, the vertical derivatives of first and second-order of ˆκ(t0,x0)

∞ at (t, ˆx) (namely ∂V xiκˆ (t0,x0) ∞ (t, ˆx) and ∂xVixjˆκ (t0,x0)

∞ (t, ˆx), for every i, j = 1, . . . , d) exist.

2) For every i, j = 1, . . . , d, ∂V xiˆκ

(t0,x0)

∞ is bounded by the constant 1 and ∂Vxixjκˆ

(t0,x0)

∞ is

bounded by the constantq2 π.

3) ˆκ(t0,x0)

∞ ≥ −Cζ and κ∞(t0,x0)(t, x) ≥ kx(· ∧ t) − x0(· ∧ t0)k∞− Cζ, for every (t, x) ∈ Λ,

with Cζ := Z Rd|z|ζ(z) dz = √ 2Γ d 2 + 1 2  Γ d 2  > 0, (3.3) where Γ(·) is the Gamma function.

(13)

4) For every fixed d, there exists some constant αd> 0 such that

αd kx(· ∧ t) − x0(· ∧ t0)k∞d+1∧ kx(· ∧ t) − x0(· ∧ t0)k∞ (3.4)

≤ κ(t0,x0)

∞ (t, x) ≤ kx(· ∧ t) − x0(· ∧ t0)k∞,

for all (t, x) ∈ Λ. In particular, it holds that κ(t0,x0)

∞ ≥ 0.

Proof. See AppendixC, SectionC.1.

We now address the problem of smoothing the map ˆκ(t0,x0)

∞ in the horizontal direction.

This is required by the fact that the presence of 1[t,T ] in the definition of ˆκ(t∞0,x0) is an

obstruction to horizontal regularity, therefore a further convolution in the time variable t is needed. The latter convolution also provides the continuity on ˆΛ (notice that the map (t, ˆx) 7→ ˆκ(t0,x0)

∞ (t, ˆx) is not continuous on ˆΛ, see Remark3.1).

We perform such a horizontal smoothing to ˆκ(t0,x0)

∞ /(1 + Cζ + ˆκ(t∞0,x0)). We apply it

to such a map (rather than to ˆκ(t0,x0)

∞ directly) in order to have bounded derivatives (see

item 3 of Lemma 3.2). Moreover, we consider 1 + Cζ + ˆκ∞(t0,x0) (instead of 1 + ˆκ(t∞0,x0))

in order to have a denominator greater than or equal to 1 (this follows from inequality ˆ

κ(t0,x0)

∞ ≥ −Cζ, see item 3 of Lemma3.1). The precise form of the mollifier η in (3.5) is

used to get explicit bounds on ˆχ(t0,x0)

∞ and its derivatives.

Remark 3.1 ([48]). Notice that the map (t, ˆx) 7→ ˆκ(t0,x0)

∞ (t, ˆx) is not continuous on ˆΛ.

As a matter of fact, consider the following example. Take d = 1, T = 2, t0= 0, x0≡ 0,

t = 1, ˆx= 1[1,2]. Then, it holds that

ˆ κ(t0,x0) ∞ (t, ˆx) = Z R ˆx(· ∧ t) − x0(· ∧ t0) − z 1[t,T ] ζ(z) dz − Z R |z| ζ(z) dz = Z R |1 − z| ζ(z) dz − Z R |z| ζ(z) dz. Now, take δ ∈ (0, 1), then

ˆ κ(t0,x0) ∞ (t + δ, ˆx) = Z R xˆ(· ∧ (t + δ)) − x0(· ∧ t0) − z 1[t+δ,T ] ∞ζ(z) dz − Z R|z| ζ(z) dz = Z R max1, |1 − z| ζ(z) dz − Z R|z| ζ(z) dz. In conclusion, we have ˆκ(t0,x0) ∞ (t + δ, ˆx) − ˆκ(t∞0,x0)(t, ˆx) = Z R n max1, |1 − z| − |1 − z|oζ(z) dz = Z 2 0 1 − |1 − z|  ζ(z) dz =: ε∗ > 0,

(14)

where ε∗is a constant independent of δ. This proves that |ˆκ(t∞0,x0)(t+δ, ˆx)−ˆκ(t∞0,x0)(t, ˆx)| 6→

0 as δ → 0+and shows that ˆκ(t0,x0)

∞ is not continuous on ˆΛ.

Lemma 3.2. Let η : R → R be given by

η(s) := s e−s, ∀ s ∈ R. (3.5) For every fixed (t0, x0) ∈ Λ, let ˆκ(t∞0,x0)be the map defined in Lemma 3.1and define the

map ˆχ(t0,x0) ∞ : ˆΛ → R as ˆ χ(t0,x0) ∞ (t, ˆx) := Z +∞ 0 ˆ κ(t0,x0) ∞ (t + s) ∧ T, ˆx(· ∧ t) 1 + Cζ+ ˆκ(t∞0,x0) (t + s) ∧ T, ˆx(· ∧ t) η(s) ds,

for all (t, ˆx) ∈ ˆΛ, with Cζ as in (3.3), where we recall that 1 + Cζ+ ˆκ(t∞0,x0)≥ 1 (see item

3 of Lemma3.1). Moreover, let χ(t0,x0)

∞ : Λ → R be given by

χ(t0,x0)

∞ (t, x) := ˆχ(t∞0,x0)(t, x), ∀ (t, x) ∈ Λ. (3.6)

Then, the following properties hold.

1) For every (t, ˆx) ∈ ˆΛ, the horizontal and vertical derivatives of first and second-order of ˆχ(t0,x0) ∞ at (t, ˆx) (namely ∂tHχˆ (t0,x0) ∞ (t, ˆx), ∂Vxiχˆ (t0,x0) ∞ (t, ˆx) and ∂xVixjχˆ (t0,x0) ∞ (t, ˆx),

for every i, j = 1, . . . , d) exist. 2) ˆχ(t0,x0)

∞ ∈ C1,2( ˆΛ) and the map (t0, x0), (t, ˆx) 7→ ˆχ(t∞0,x0)(t, ˆx) is continuous on

Λ × ˆΛ.

3) The horizontal derivative of ˆχ(t0,x0)

∞ is bounded by the constant 2e; the first-order

verti-cal derivatives of ˆχ(t0,x0)

∞ are bounded by the constant 1 + Cζ; the second-order vertical

derivatives of ˆχ(t0,x0)

∞ are bounded by the constant (1 + Cζ)

q 2 π+ 2  . 4) For every (t, x) ∈ Λ, αdkx(· ∧ t) − x0(· ∧ t0)k d+1 ∞ ∧ kx(· ∧ t) − x0(· ∧ t0)k∞ 1 + Cζ+ kx(· ∧ t) − x0(· ∧ t0)k∞ (3.7) ≤ χ(t0,x0) ∞ (t, x) ≤ kx(· ∧ t) − x0(· ∧ t0)k∞∧ 1,

with the same constant αd as in (3.4). In particular, it holds that χ(t∞0,x0)≥ 0.

Proof. See AppendixC, SectionC.2.

In conclusion, by Lemma3.2it follows that the map ρ∞: Λ × Λ → [0, +∞) given by

ρ∞ (t, x), (t0, x0) = |t − t0|2+ χ∞(t0,x0)(t, x), ∀ (t, x), (t0, x0) ∈ Λ, (3.8)

with χ∞as in (3.6), is a gauge-type function, which is also smooth as a map of the first

pair, namely (t, x) 7→ ρ∞((t, x), (t0, x0)), and with bounded derivatives.

We now apply Theorem 3.1 to the smooth gauge-type function ρ∞ with bounded

(15)

Theorem 3.2 (Smooth variational principle on Λ). Let δ > 0 and G : Λ → R be an upper semicontinuous map, bounded from above. For every ε > 0, let (t0, x0) ∈ Λ satisfy

sup G − ε ≤ G(t0, x0).

Then, there exists a sequence {(tn, xn)}n≥1 ⊂ Λ which converges to some (¯t, ¯x) ∈ Λ

fulfilling the properties below.

i) ρ∞((¯t, ¯x), (tn, xn)) ≤ 2δ, for every n ≥ 0.

ii) G(t0, x0) ≤ G(¯t, ¯x) − δϕε(t, x), where the map ϕε: Λ → [0, +∞) is defined as

ϕε(t, x) := +∞ X n=0 1 2nρ∞ (t, x), (tn, xn), ∀ (t, x) ∈ Λ.

iii) For every (t, x) 6= (¯t, ¯x), G(t, x) − δ ϕε(t, x) < G(¯t, ¯x) − δ ϕε(¯t, ¯x).

Finally, the map ϕε satisfies the following properties.

1) ϕε∈ C1,2(Λ) and is bounded.

2) ∂H

t ϕε is bounded by the constant 2 2 T +2e.

3) For every i, j = 1, . . . , d, ∂V

xiϕε is bounded by the constant 2 (1 + Cζ) and ∂

V xixjϕε is

bounded by the constant 2 (1 + Cζ)

q

2 π+ 2

 .

Proof. Items i)-ii)-iii) follow directly from Theorem3.1, while items 1)-2)-3) follow easily from items 2)-3)-4) of Lemma3.2.

4. Crandall-Lions (path-dependent) viscosity

solutions

4.1. Viscosity solutions

In the present section we consider the second-order path-dependent partial differential equation

  

∂H

t u(t, x) = F t, x, u(t, x), ∂Vxu(t, x), ∂xxV u(t, x), (t, x) ∈ [0, T ) × C([0, T ]; Rd),

u(T, x) = ξ(x), x∈ C([0, T ]; Rd), (4.1)

with F : [0, T ] × C([0, T ]; Rd) × Rd× S(d) → R and ξ : C([0, T ]; Rd) → R, where S(d) is

the set of symmetric d × d matrices.

Definition 4.1. We denote by C1,2pol(Λ) the set of ϕ ∈ C1,2(Λ) such that ϕ, ∂H t ϕ,

∂V

(16)

Definition 4.2. We say that an upper semicontinuous map u : Λ → R is a (path-dependent) viscosity subsolution of equation (4.1) if the following holds.

• u(T , x)≤ ξ(x), for all x ∈ C([0, T ]; Rd);

• for any (t, x)∈ [0, T ) × C([0, T ]; Rd) and ϕ ∈ C1,2pol(Λ), satisfying

(u − ϕ)(t, x) = sup (t′ ,x′ )∈Λ(u − ϕ)(t ′, x), we have −∂tHϕ(t, x) + F t, x, u(t, x), ∂xVϕ(t, x), ∂ V xxϕ(t, x)  ≤ 0.

We say that a lower semicontinuous map u : Λ → R is a (path-dependent) viscosity supersolutionof equation (4.1) if:

• u(T , x)≥ ξ(x), for all x ∈ C([0, T ]; Rd);

• for any (t, x)∈ [0, T ) × C([0, T ]; Rd) and ϕ ∈ C1,2pol(Λ), satisfying:

(u − ϕ)(t, x) = inf (t′,x)∈Λ(u − ϕ)(t ′, x), we have −∂tHϕ(t, x) + F t, x, u(t, x), ∂xVϕ(t, x), ∂ V xxϕ(t, x)  ≥ 0.

We say that a continuous map u : Λ → R is a (path-dependent) viscosity solution of equation (4.1) if u is both a (path-dependent) viscosity subsolution and a (path-dependent) viscosity supersolution of (4.1).

4.2. Path-dependent heat equation

In the present section we focus on the path-dependent heat equation, namely when F (t, x, r, p, M ) = −1 2tr[M ]      ∂H t u(t, x) + 1 2tr∂ V xxu(t, x)  = 0, (t, x) ∈ [0, T ) × C([0, T ]; Rd), u(T, x) = ξ(x), x∈ C([0, T ]; Rd). (4.2)

In the sequel we denote

Lu(t, x) := ∂tHu(t, x) +

1 2tr[∂

V

xxu(t, x)]. (4.3)

On the terminal condition ξ, we impose the assumption

(17)

4.2.1. Existence

The “candidate solution” to equation (4.2) is v(t, x) := Eξ(Wt,x),

for all (t, x) ∈ Λ, (4.4) where W = (Ws)s∈[0,T ] is a d-dimensional Brownian motion on some complete

proba-bility space (Ω, F, P), and the stochastic process Wt,x= (Wt,xs )s∈[0,T ]is given by

Wt,xs :=

(

x(s), s ≤ t,

x(t) + Ws− Wt, s > t.

(4.5)

Remark 4.1. The boundedness of ξ in Assumption (A) will be used in the proof of (the comparison) Theorem4.2. On the other hand, the proof that the function v in (4.4) is continuous and is a viscosity solution of equation (4.2) (see the proof of Theorem4.1) holds under weaker growth condition on ξ (for instance, ξ having polynomial growth).

Theorem 4.1. Under Assumption (A), the function v in (4.4) is continuous and bounded. Moreover, v is a (path-dependent) viscosity solution of equation (4.2).

Proof. Step I. Continuity of v. Given (t, x), (t′, x) ∈ Λ, with t ≤ t, from (4.5) we

have Wt,xs − Wt ′ ,x′ s =      x(s) − x′(s), s ≤ t, x(t) − x′(s) + W s− Wt, t < s ≤ t′, x(t) − x′(t) + W t′− Wt, s > t′. Hence sup s∈[0,T ] Wt,xs − Wts′,x′ ≤ kx(· ∧ t) − x′(· ∧ t′)k∞+ sup s∈[t,t′] Ws− Wt ≤ kx(· ∧ t) − x′(· ∧ t′)k∞+ d X i=1 sup s∈[t,t′] Wsi− Wti ,

where W = (W1, . . . , Wd) and the second inequality follows from the fact the Euclidean

norm on Rdis estimated by the 1-norm. By the reflection principle, sup

s∈[t,t′]|Wi

s− Wti|

has the same law as |Wi

t′− Wti|, therefore Eh sup s∈[0,T ] Wt,xs − Wt ′ ,x′ s i ≤ kx(· ∧ t) − x′(· ∧ t)k ∞+ d X i=1 E|Wti′− Wti|  = kx(· ∧ t) − x′(· ∧ t′)k∞+ d r 2 πp|t − t ′|.

Then, since ξ is bounded and continuous, the continuity of v follows from the above estimate together with the Lebesgue dominated convergence theorem.

(18)

Step II.v is a viscosity solution of equation (4.2). For every t ∈ [0, T ], let Ft= (Ft

s)s∈[t,T ]

be the completion of the filtration generated by (Ws− Wt)s∈[t,T ]. Now, fix (t, x) ∈ Λ

and t′∈ [t, T ]. We first prove that

v(t, x) = Ev t′, Wt,x. (4.6)

To this end, we begin noticing that by (4.5) we have

Wt,x· = x(· ∧ t) + W·∨t− Wt. (4.7)

Therefore

v(t, x) = E[ξ(x(· ∧ t) + W·∨t− Wt)]. (4.8)

Now, notice that, by (4.7),

Wt ′ ,Wt,x · = W t,x ·∧t′+ W·∨t′ − Wt′ = Wt,x · .

This proves the flow property Wt,x· = Wt

,Wt,x

· . Then, by the freezing lemma for

con-ditional expectation and formula (4.8), we obtain

v(t, x) = Eξ(Wt,x ) = Eξ Wt′ ,Wt,x = Eξ Wt,x ·∧t′+ W·∨t′− Wt′  = EEξ Wt,x ·∧t′+ W·∨t′ − Wt′  Ftt′  = Ev t′, Wt,x ·∧t′.

Finally, recalling that v is non-anticipative we deduce that v(t′, Wt,x

·∧t′) = v(t

, Wt,x),

which concludes the proof of formula (4.6).

Let us now prove that v is a viscosity solution of equation (4.2). We only prove the viscosity subsolution property, as the supersolution property can be proved in a similar way. We proceed along the same lines as in the proof of the subsolution property in Theorem 3.66 of [38]. Let (t, x) ∈ [0, T ) × C([0, T ]; Rd) and ϕ ∈ C1,2

pol(Λ), satisfying:

(v − ϕ)(t, x) = sup

(t′,x)∈Λ(v − ϕ)(t

, x).

We suppose that (v − ϕ)(t, x) = 0 (if this is not the case, we replace ϕ by ψ(·, ·) := ϕ(·, ·) + v(t, x) − ϕ(t, x)). Take

ϕ(t, x) = v(t, x) = Ev t + ε, Wt,x

≤ Eϕ t + ε, Wt,x,

(4.9)

where the latter inequality follows from the fact that sup(v − ϕ) = 0, so that v ≤ ϕ on Λ. Notice that the last expectation in (4.9) is finite, as ϕ has polynomial growth. Now, by the functional Itˆo formula (2.1), we have

ϕ(t + ε, Wt,x) = ϕ(t, x) + Z t+ε t Lϕ(s, W t,x) ds + d X i=1 Z t+ε t ∂xViϕ(s, W t,x) dWi s,

(19)

where L was defined in (4.3). Since ∂V

xiϕ has polynomial growth, the corresponding

stochastic integral is a martingale. Then, plugging the above formula into (4.9) and dividing by ε, we find − E 1ε Z t+ε t Lϕ(s, W t,x) ds  ≤ 0. Letting ε → 0+, we conclude that

−Lϕ(t, x) ≤ 0, which proves the viscosity subsolution property.

4.2.2. Comparison theorem and uniqueness

Theorem 4.2. Suppose that Assumption (A) holds. Let u, w : Λ → R be respectively upper and lower semicontinuous, satisfying

sup u < +∞, inf w > −∞.

Suppose that u (resp. w) is a (path-dependent) viscosity subsolution (resp. supersolution) of equation (4.2). Then u ≤ w on Λ.

Proof. The proof consists in showing that u ≤ v and v ≤ w on Λ (with v given by (4.4)), from which we immediately deduce the claim. In what follows, we only report the proof of the inequality u ≤ v, as the other inequality (that is v ≤ w) can be deduced from the first one replacing u, v, ξ with −w, −v, −ξ, respectively.

We proceed by contradiction and assume that sup(u − v) > 0. Then, there exists (t0, x0) ∈ Λ such that

(u − v)(t0, x0) > 0.

Notice that t0< T , since u(T, ·) ≤ ξ(·) = v(T, ·). We split the rest of the proof into five

steps.

Step I. Let {ξN}N be the sequence given by LemmaD.2. Since ξ is bounded, we have that ξN is bounded uniformly with respect to N . Now, denote

vN(t, x) := EξN(Wt,x), for all (t, x) ∈ Λ.

Then, vN is bounded uniformly with respect to N . Moreover, by Lemma D.1it follows

that, for every N , vN ∈ C1,2(Λ) and is a classical (smooth) solution of equation (4.2)

with terminal condition ξN. Finally, recalling from Lemma D.2 that {ξN}N converges

pointwise to ξ as N → +∞, it follows from the Lebesgue dominated convergence theorem that {vN}N converges pointwise to v as N → +∞. Then, we notice that there exists

N0∈ N such that

(u − vN0)(t0, x0) > 0. (4.10)

We also suppose that (possibly enlarging N0)

|ξ(x0) − ξN0(x0)| ≤

1

(20)

Step II. For every λ > 0, we set(t, x) := eλtu(t, x), ξλ(x) := eλTξ(x), vλ N0(t, x) := e λtv N0(t, x), ξ λ N0(x) := e λTξ N0(x).

for all (t, x) ∈ Λ. Notice that uλ is a (path-dependent) viscosity subsolution of the

path-dependent partial differential equation    ∂H t uλ(t, x) + 12tr[∂xVuλ(t, x)] = λ uλ(t, x), (t, x) ∈ [0, T ) × C([0, T ]; Rd), uλ(T, x) = ξλ(x), x∈ C([0, T ]; Rd). (4.12) Similarly, vλ

N0 is a classical (smooth) solution of equation (4.12) with ξ

λreplaced by ξλ N0.

We finally notice that by (4.10) we have

(uλ− vNλ0)(t0, x0) > 0. So, in particular, sup(uλ− vNλ0) − ε = (u λ − vλN0)(t0, x0) ≤ sup(u λ − vλN0), (4.13) where ε := sup(uλ− vλ N0) − (u λ− vλ N0)(t0, x0).

Step III.Notice that uλ− vλ

N0 is upper semicontinuous and bounded from above. Then,

by (4.13) and the smooth variational principle (Theorem 3.2) with G = uλ− vλ N0, we

deduce that for every δ > 0 there exists a sequence {(tn, xn)}n≥1 ⊂ Λ converging to

some (¯t, ¯x) ∈ Λ (possibly depending on ε, δ, λ, N0) such that the following holds.

i) ρ∞((tn, xn), (¯t, ¯x)) ≤ 2δ, for every n ≥ 0, where ρ∞ is the smooth gauge-type

function with bounded derivatives defined by (3.8). ii) (uλ− vλ N0)(t0, x0) ≤ u λ− (vλ N0+ δϕε)(¯t, ¯x), where ϕε(t, x) := +∞ X n=0 1 2nρ∞ (t, x), (tn, xn)  ∀ (t, x) ∈ Λ. iii) It holds that

− (vλ

N0+ δϕε)(¯t, ¯x) = sup

(t,x)∈Λ

− (vλ

N0+ δϕε)(t, x). (4.14)

We also recall from Theorem3.2that ϕεsatisfies the following properties.

1) ϕε∈ C1,2(Λ) and is bounded.

2) ∂tHϕε(t, x)

≤ 2 2 T +2e, for every (t, x) ∈ [0, T ) × C([0, T ]; Rd). 3) For every i, j = 1, . . . , d, ∂V

xiϕε is bounded by the constant 2 (1 + Cζ) and ∂

V xixjϕε is

bounded by the constant 2 (1 + Cζ)

q

2 π+ 2

 .

(21)

In particular, ϕε∈ C1,2pol(Λ).

Step IV. We prove below that ¯t < T . As a matter of fact, by item ii) of Step III we have

uλ− (vλN0+ δϕε)(¯t, ¯x) ≥ (u

λ

− vNλ0)(t0, x0). (4.15)

On the other hand, if ¯t = T we obtain

uλ−(vNλ0+δϕε)(¯t, ¯x) = e

λT(ξ(¯x

)−ξN0(¯x))−δϕε(T, ¯x) ≤ e

λT(ξ(¯x

)−ξN0(¯x)), (4.16)

where the latter inequality comes from the fact that ϕε≥ 0. Hence, by (4.15) and (4.16)

we get

eλt0(u − v

N0)(t0, x0) ≤ e

λT(ξ(¯x

) − ξN0(¯x)).

Letting ε → 0, it follows from item i) above with n = 0 and (3.7) that d∞((¯t, ¯x), (t0, x0)) →

0. Therefore, letting ε → 0 in the previous inequality, we obtain eλt0(u − v N0)(t0, x0) ≤ e λT(ξ(x 0) − ξN0(x0)). By (4.11), we end up with eλt0 1 2e

λT. Letting λ → 0, we find a contradiction.

Step V. Here again λ > 0 is fixed. By (4.14) and the definition of viscosity subsolution of (4.12) applied to uλ at the point (¯t, ¯x) with test function vλ

N0+ δϕε, we obtain

−L(vλ

N0+ δϕε)(¯t, ¯x) + λ u

λt, ¯x) ≤ 0.

Recalling that vλ

N0 is a classical (smooth) solution of equation (4.12) with ξ

λreplaced by

ξλ

N0, we find

λ (uλ− vλN0)(¯t, ¯x) ≤ δ Lϕε(¯t, ¯x).

By item ii) in Step III (namely (4.15)), subtracting from both sides the quantity λδϕε(¯t, ¯x), we obtain

λ (uλ− vλN0)(t0, x0) ≤ λ u

λ

− (vNλ0+ δϕε)(¯t, ¯x) ≤ δ Lϕε(¯t, ¯x) − λ δ ϕε(¯t, ¯x).

Recalling that ϕε≥ 0, we see that

λ (uλ− vNλ0)(t0, x0) ≤ λ u

λ

− (vNλ0+ δϕε)(¯t, ¯x) ≤ δ Lϕε(¯t, ¯x).

From items 2) and 3) above, it follows that Lϕε(¯t, ¯x) is bounded by a constant (not

depending on ε, δ, λ). Therefore, letting δ → 0+, taking into account the notations of

Step II, we have

λ eλt0(u − v

N0)(t0, x0) = λ (u

λ

− vNλ0)(t0, x0) ≤ 0,

which gives a contradiction to (4.10).

As a direct consequence of the comparison theorem (Theorem 4.2), we obtain the following uniqueness result.

(22)

Corollary 4.1. Under Assumption (A), the function v in (4.4) is the unique (path-dependent) viscosity solution of equation (4.2), where uniqueness holds in the class of all continuous and bounded functions from Λ to R.

Proof. By Theorem 4.1 we know that v is continuous and bounded, moreover it is a (path-dependent) viscosity solution of equation (4.2).

Now, let u : Λ → R be a continuous and bounded function such that u is a (path-dependent) viscosity solution of equation (4.2). Then, in particular, u (resp. v) is a (path-dependent) viscosity subsolution (resp. supersolution) of equation (4.2). As a con-sequence, by the comparison theorem (Theorem4.2) we deduce that u ≤ v on Λ. Chang-ing the roles of u and v we get the opposite inequality, from which we conclude that u ≡ v.

Appendix A: Functional Itˆ

o’s formula

We start with a definition and a technical result.

Definition A.1. We denote by C1,0( ˆΛ) the set of ˆu ∈ C( ˆΛ) such that ∂H

t u existsˆ

everywhere on ˆΛ and is continuous.

Lemma A.1. Let ˆu ∈ C1,0( ˆΛ). Then, for every (t, ˆx) ∈ [0, T ) × D([0, T ]; Rd), the map

φ : [0, T − t] → R, defined as

φ(a) := ˆu(t + a, ˆx(· ∧ t)), ∀ a ∈ [0, T − t], is in C1([0, T − t]) and

φ′(a) = ∂tHu(t + a, ˆˆ x(· ∧ t)), ∀ a ∈ [0, T − t].

Proof. Let a ∈ [0, T − t). We have, for any δ ∈ (0, T − t − a], φ(a + δ) − φ(a) δ = ˆ u(t + a + δ, ˆx(· ∧ t)) − ˆu(t + a, ˆx(· ∧ t)) δ δ→0−→+∂ H t u(t + a, ˆˆ x(· ∧ t)).

This shows that φ is right-differentiable on [0, T − t) and that such a right-derivative is continuous on [0, T − t). Then, it follows for instance from Corollary 1.2, Chapter 2, in [75] that φ ∈ C1([0, T − t)). Finally, at a = T − t, we have

lim a→(T −t)−φ ′(a) = lim a→(T −t)−∂ H t u(t + a, ˆˆ x(· ∧ t)) = ∂tHu(T, ˆˆ x(· ∧ t)).

This implies that φ ∈ C1([0, T − t]) and concludes the proof. We now report the proof of Theorem2.1.

(23)

Proof of Theorem 2.1. Fix t ∈ (0, T ]. For every ˆx∈ D([0, T ]; Rd), denote by ˆx(·∧t−) the path ˆ x(s ∧ t−) := ( ˆ x(s), 0 ≤ s < t, ˆ x(t−), t ≤ s ≤ T. (A.1) When t = 0, we set ˆx(· ∧ 0−) := ˆx(· ∧ 0). We split the rest of the proof into several steps. Step 1. Piecewise constant approximation of X. For every n ∈ N, consider the dyadic partition {tn

0, . . . , tnℓ, . . . , tn2n} of [0, t], with tn = ℓ

2nt, for every ℓ = 0, . . . , 2n. Then, we

consider the piecewise constant approximation of X

Xns = 2n−1 X ℓ=0 Xtn ℓ 1[tnℓ,tnℓ+1)(s) + Xt1[t,T ](s), ∀ s ∈ [0, T ]. (A.2)

Now, recalling the definition of ˆx(· ∧ t−) in (A.1), we notice that ˆ u(t, Xn·∧t−) − ˆu(0, X) = 2n−1 X ℓ=0  ˆ u tnℓ+1, Xn·∧tn ℓ+1− − ˆu t n ℓ, Xn·∧tn ℓ−  , (A.3)

where we have used the telescoping property and the equality ˆu(0, X) = ˆu(0, Xn). For every ℓ = 0, . . . , 2n− 1, we have ˆ u tnℓ+1, Xn·∧tn ℓ+1− − ˆu t n ℓ, Xn·∧tn ℓ−  = Iℓn+ Jℓn, where Iℓn := ˆu tnℓ+1, Xn·∧tn ℓ+1− − ˆu t n ℓ, Xn·∧tn ℓ+1−  = ˆu tnℓ+1, Xn·∧tn ℓ − ˆu t n ℓ, Xn·∧tn ℓ, Jℓn := ˆu tnℓ, Xn·∧tn ℓ+1− − ˆu t n ℓ, Xn·∧tn ℓ−.

Notice that when ℓ = 0 we have Jn

ℓ = 0. Moreover, when ℓ = 1, . . . , 2n− 1, Jℓn can be

written as Jℓn = ˆu tnℓ, Xn·∧tn ℓ − ˆu t n ℓ, Xn·∧tn ℓ−1.

Step 2.Definitions of In, Jn,1, Jn,2, Jn,3, Jn,4. By LemmaA.1we have

Iℓn = Z tnℓ+1 tn ℓ ∂tHu s, Xˆ n·∧tn ℓ ds = Z tnℓ+1 tn ℓ ∂tHu s, Xˆ n·∧s ds. Then, we define In := 2n−1 X ℓ=0 Iℓn = 2n−1 X ℓ=0 Z tnℓ+1 tn ℓ ∂tHu s, Xˆ n·∧s ds = Z t 0 ∂tHu s, Xˆ n·∧s ds. (A.4)

On the other hand, when ℓ = 1, . . . , 2n− 1 the term Jn

ℓ can be written as Jn ℓ = J n,1 ℓ + J n,2 ℓ + J n,3 ℓ + J n,4 ℓ ,

(24)

where (notice that Xn = (Xn,1, . . . , Xn,d)) Jn,1= d X i=1 ∂xViu tˆ n ℓ, Xn·∧tn ℓ−1  Xtn,in ℓ − X n,i tn ℓ−1  = d X i=1 ∂xViu tˆ n ℓ, Xn·∧tn ℓ−1  Xtin ℓ − X i tn ℓ−1, Jn,2=1 2 d X i,j=1 ∂xVixjˆu t n ℓ, X·∧tn ℓ−1  Xtin ℓ − X i tn ℓ−1  Xtjn ℓ − X j tn ℓ−1, Jn,3=1 2 d X i,j=1  ∂V xixju tˆ n ℓ, Xn·∧tn ℓ−1 − ∂ V xixju tˆ n ℓ, X·∧tn ℓ−1  Xi tn ℓ − X i tn ℓ−1  Xtjn ℓ − X j tn ℓ−1, Jn,4=1 2 d X i,j=1 Z 1 0  ∂xVixju tˆ n ℓ, Xn·∧tn ℓ−1+ a Xt n ℓ − Xtnℓ−11[tnℓ,T ]  − ∂V xixju tˆ n ℓ, Xn·∧tn ℓ−1  Xi tn ℓ − X i tn ℓ−1  Xtjn ℓ − X j tn ℓ−1 da. Then, we define Jn,k := 2n−1 X ℓ=1 Jn,k, (A.5) for every k = 1, 2, 3, 4. As a consequence, by (A.3), (A.4), (A.5), we obtain

ˆ

u(t, Xn·∧t−) − ˆu(0, X) = In+ Jn,1+ Jn,2+ Jn,3+ Jn,4. (A.6)

Notice that Jn,1 = 2n−1 X ℓ=1 d X i=1 ∂xViu tˆ n ℓ, Xn·∧tn ℓ−1  Xtin ℓ − X i tn ℓ−1  = d X i=1 Z t 0 Zsn,idXsi, (A.7) where Zsn,i := 2n−1 X ℓ=1 ∂xViu tˆ n ℓ, X n ·∧s1[tn ℓ−1,tnℓ)(s), ∀ s ∈ [0, t].

Step 3.Convergence. Before studying the convergence of the various terms in (A.6), we make some observations. Recall that for a.e. ω ∈ Ω the trajectory X(ω) is continuous on [0, T ], so that it admits a modulus of continuity ρX(ω): [0, +∞) → [0, +∞). Then, it

holds that ˆ

d∞ (t, Xn·∧t−(ω)), (t, X·∧t(ω)) = X·∧t−n (ω) − X·∧t(ω) ≤ ρX(ω)(2−n), (A.8)

for a.e. ω ∈ Ω. Similarly, we have ˆ d∞ (s, Xn·∧s(ω)), (s, X·∧s(ω)) = Xn·∧s(ω) − X·∧s(ω) ≤ ρX(ω)(2−n), (A.9)

for a.e. ω ∈ Ω and for every s ∈ [0, t]. Moreover, for every ℓ = 1, . . . , 2n− 1 and tn ℓ−1 ≤ s ≤ tn ℓ, it holds that ˆ d∞ (tnℓ, X n ·∧s(ω)), (s, X·∧s(ω)) = |tnℓ − s| + Xn·∧s(ω) − X·∧s(ω)

(25)

≤ 2−n+ ρ

X(ω)(2−n), (A.10)

for a.e. ω ∈ Ω. Now, for every x ∈ C([0, T ]; Rd) and n ∈ N, set (as in (A.2))

xn(s) = 2n−1 X ℓ=0 x(tnℓ) 1[tn ℓ,tnℓ+1)(s) + x(t) 1[t,T ](s), ∀ s ∈ [0, T ].

For such an x, define

K1(x) := (s, ˆx) ∈ ˆΛ : ˆx = xn(· ∧ s), for some n ∈ N , (A.11)

K2(x) := (r, ˆx) ∈ ˆΛ : r, ˆx = tnℓ, xn(· ∧ s), for some n ∈ N,

ℓ = 1, . . . , 2n− 1, s ∈ [tnℓ−1, tnℓ) , (A.12)

K3(x) := (r, ˆx) ∈ ˆΛ : r, ˆx = tnℓ, xn(· ∧ tnℓ−1) + a(x(tnℓ) − x(tnℓ−1))1[tn ℓ,T ](·),

for some n ∈ N, ℓ = 1, . . . , 2n− 1, a ∈ [0, 1] . (A.13) Notice that K2(x) ⊂ K3(x) (observe that xn(·∧s) = xn(·∧tnℓ−1), whenever s ∈ [tnℓ−1, tnℓ),

and consider the pairs in K3(x) with a = 0). It is easy to see that K1(x) and K3(x) (and

a fortiori K2(x)) are relatively compact subsets of ( ˆΛ, ˆd∞). Now, we study separately

the convergence of the various terms appearing in (A.6).

Substep 3.1. Convergence of ˆu(t, Xn·∧t−) − ˆu(0, X). Since ˆu is continuous on ( ˆΛ, ˆd∞)

and (A.8) holds, we deduce that

ˆ

u(t, Xn·∧t−) − ˆu(0, X) P-a.s.−→

n→∞ u(t, Xˆ ·∧t) − ˆu(0, X).

Substep 3.2. Convergence of In. Recalling (A.11), we see that for a.e. ω ∈ Ω the set

K1(X(ω)) is a relatively compact subset of ˆΛ. Since ∂tHu is continuous on ( ˆˆ Λ, ˆd∞), for

such an ω we see that ∂H

t u(s, Xˆ n·∧s(ω)) is uniformly bounded with respect to s and n. In

addition, using again the continuity of ∂H

t ˆu and (A.9), we deduce that for a.e. ω ∈ Ω the

sequence of maps s 7→ ∂H

t u(s, Xˆ n·∧s(ω)) converges pointwise to s 7→ ∂tHu(s, Xˆ ·∧s(ω)). In

conclusion, we can apply Lebesgue’s dominated convergence theorem, which yields

In = Z t 0 ∂tHu s, Xˆ n·∧s ds n→∞P-a.s.−→ Z t 0 ∂tHu s, Xˆ ·∧s ds.

Substep 3.3. Convergence of Jn,1. In order to study the convergence of the sum of

stochastic integrals in (A.7), we write each Xi as Vi+ Mi, where Vi is a bounded

variation process and Miis a continuous local martingale, so that the stochastic integral

with respect to Xi can be written as the sum of two integrals with respect to Vi and

Mi, respectively.

Convergence ofRt

0Z n,i

s dVsi. Recalling (A.12), we see that for a.e. ω ∈ Ω the set K2(X(ω))

(26)

continuity of ∂V

xiu and also (A.10), we see that we can apply Lebesgue’s dominatedˆ

convergence theorem, from which we get Z t 0 Zsn,idVsi n→∞P-a.s.−→ Z t 0 ∂xViu s, Xˆ ·∧s dV i s. Convergence ofRt 0Z n,i

s dMsi. Let hMii denote the quadratic variation of Mi. Reasoning

as in the proof of the convergence of Rt

0Z n,i

s dVsi, we obtain by Lebesgue’s dominated

convergence theorem Z t 0 Zsn,i− ˆu s, X·∧s 2 dhMiis P-a.s.−→ n→∞ 0.

By Proposition 2.26, Chapter 3, in [55], we deduce the convergence in probability (even u.c.p.) Z t 0 Zsn,idMsi P −→ n→∞ Z t 0 ∂xViu s, Xˆ ·∧s dM i s.

Substep 3.4. Convergence of Jn,2. For every n ∈ N and i, j = 1, . . . , d, consider the

process hXi, Xjin defined as hXi, Xjins := 2n X ℓ=1 Xtin ℓ − X i tn ℓ−1  Xtjn ℓ − X j tn ℓ−1 1[t n ℓ−1,T ](s), ∀ s ∈ [0, T ]. Notice that Jn,2 = 1 2 d X i,j=1 Z [tn ℓ−1,tnℓ) ∂xVixju s, Xˆ ·∧s dhX i, Xj ins and, therefore, Jn,2 = 1 2 d X i,j=1 Z [0,(1−2−n)t) ∂xVixju s, Xˆ ·∧s dhX i, Xj ins.

From Theorem 23 in Section II.6 of [80], it is easy to deduce that

sup s∈[0,t] hXi, Xjins − hXi, Xjis P −→ n→∞ 0,

where hXi, Xji is the covariation of Xiand Xj. Then, up to a subsequence, there exists a

P-null set N such that, whenever ω /∈ N, hXi, Xjins(ω) → hXi, Xjis(ω), for all s ∈ [0, t]. For such an ω, this implies that the measure dhXi, Xjin(ω) converges weakly to the

measure dhXi, Xjin(ω). Since ∂V

xixju is continuous, we deduce (up to a subsequence)ˆ

Jn,2 n→∞P-a.s.−→ 12 d X i,j=1 Z t 0 ∂xVixju s, Xˆ ·∧s dhX i, Xj is.

(27)

Substep 3.5. Convergence of Jn,3. We have (we set hXiin := hXi, Xiin, for every i = 1, . . . , d) |Jn,3| ≤ 1 4 d X i,j=1 sup ℓ=1,...,2n−1 ∂ V xixju tˆ n ℓ, Xn·∧tn ℓ−1 − ∂ V xixju tˆ n ℓ, X·∧tn ℓ−1  hX iin T + hXjinT.

Recalling (A.12), we see that for a.e. ω ∈ Ω the set K2(X(ω)) is a relatively compact

subset of ˆΛ. Since ∂V

xixju is continuous on ( ˆˆ Λ, ˆd∞), for a.e. ω ∈ Ω there exists a

(non-decreasing) modulus of continuity ρω

i,j: [0, +∞) → [0, +∞) such that

|Jn,3(ω)| ≤ 1 4 d X i,j=1 sup ℓ=1,...,2n−1 ρω i,j Xn·∧tn ℓ−1(ω) − X·∧t n ℓ−1(ω) ∞  hXiin T(ω) + hXjinT(ω).

From Substep 3.4 we know that, up to a subsequence, hXiin

T(ω) converges to hXiiT(ω),

for a.e. ω ∈ Ω and for every i = 1, . . . , d. On the other hand, by (A.9) we have that supℓ=1,...,2n−1ρωi,j(kXn·∧tn

ℓ−1(ω) − X·∧t n

ℓ−1(ω)k∞) ≤ ρ

ω

i,j(ρX(ω)(2−n)), for a.e. ω ∈ Ω.

Hence, up to a subsequence,

Jn,3 n→∞P-a.s.−→ 0.

Substep 3.6. Convergence of Jn,4. We have (as in Substep 3.5 we set hXiin :=

hXi, Xiin, for every i = 1, . . . , d) |Jn,4| ≤ 14 d X i,j=1 hXiinT + hXjinT  Z 1 0 sup ℓ=1,...,2n−1 ∂ V xixjˆu t n ℓ, Xn·∧tn ℓ−1  − ∂xVixju tˆ n ℓ, Xn·∧tn ℓ−1+ a Xt n ℓ−Xtnℓ−11[tnℓ,T ]  da. Recalling (A.13), we see that for a.e. ω ∈ Ω the set K3(X(ω)) is a relatively compact

subset of ˆΛ. Since ∂V

xixju is continuous on ( ˆˆ Λ, ˆd∞), for a.e. ω ∈ Ω there exists a

(non-decreasing) modulus of continuity ¯ρω

i,j: [0, +∞) → [0, +∞) such that

|Jn,4(ω)| ≤ 1 4 d X i,j=1 hXiin T(ω) + hXjinT(ω)  Z 1 0 sup ℓ=1,...,2n−1 ¯ ρω i,j a Xtn ℓ(ω) − Xtnℓ−1(ω)  da.

We know from Substep 3.4 that, up to a subsequence, hXiin

T(ω) converges to hXiiT(ω),

for a.e. ω ∈ Ω and for every i = 1, . . . , d. On the other hand, supℓ=1,...,2n−1ρ¯ωi,j(a|Xtn ℓ(ω)−

Xtn

ℓ−1(ω)|) ≤ ¯ρ

ω

i,j(ρX(ω)(2−n)), for a.e. ω ∈ Ω, where we recall that ρX(ω)is the modulus

of continuity of the trajectory X(ω). Hence, up to a subsequence,

(28)

Appendix B: Consistency

Proof of Lemma 2.1. The claim concerning the horizontal derivatives follows directly from their definition (Definition2.3-(i)).

It remains to prove the claim concerning the vertical derivatives. To this end, let X= (Xt)t∈[0,T ], with X = (X1, . . . , Xd), be a d-dimensional continuous semimartingale

on some filtered probability space (Ω, F, (Ft)t∈[0,T ], P), where (Ft)t∈[0,T ] satisfies the

usual conditions. Since ˆu1, ˆu2∈ C1,2( ˆΛ), by Theorem2.1the functional Itˆo formulae

ˆ u1(t, X) = ˆu1(0, X) + Z t 0 ∂H t uˆ1(s, X) ds + 1 2 d X i,j=1 Z t 0 ∂V xixjuˆ1(s, X) d[X i, Xj] s + d X i=1 Z t 0 ∂xViuˆ1(s, X) dX i

s, for all 0 ≤ t ≤ T, P-a.s.

and ˆ u2(t, X) = ˆu2(0, X) + Z t 0 ∂tHuˆ2(s, X) ds +1 2 d X i,j=1 Z t 0 ∂xVixjuˆ2(s, X) d[X i, Xj] s + d X i=1 Z t 0 ∂V xiuˆ2(s, X) dX i

s, for all 0 ≤ t ≤ T, P-a.s.

hold. Recalling that ˆu1 and ˆu2, together with their horizontal derivatives, coincide on

continuous paths, identifying bounded variation and local martingale parts in the above formulae, (up to a P-null set), for every i, j = 1, . . . , d and any t ∈ [0, T ], we have

∂xViuˆ1(t, X) = ∂ V xiuˆ2(t, X), ∂ V xixjuˆ1(t, X) = ∂ V xixjuˆ2(t, X). (B.1)

Now, recall that C([0, T ]; Rd) is endowed with the uniform topology and let B(C([0, T ]; Rd))

denote the Borel σ-algebra on C([0, T ]; Rd). We also recall that given a probability

measure µ : B(C([0, T ]; Rd)) → [0, 1], the support of µ is the smallest closed set S µ ⊂

C([0, T ]; Rd) such that µ(S

µ) = 1. Consider a continuous semimartingale X = (Xs)s∈[0,T ]

whose law has support equal to C([0, T ]; Rd). An example of such an X is given by

Xs:= η + Ws, s ∈ [0, T ], where W = (Ws)s∈[0,T ]is a d-dimensional Brownian motion,

while η : Ω → Rdis independent of W and has a standard normal multivariate distribu-tion. Then, for every fixed t ∈ [0, T ], using equalities (B.1) with such a semimartingale X, and exploiting the continuity of ∂V

xuˆ1, ∂xVuˆ2, ∂xxV uˆ1, ∂xxV uˆ2, we obtain

∂xVuˆ1(t, x) = ∂xVuˆ2(t, x), ∂xxV ˆu1(t, x) = ∂xxV uˆ2(t, x),

for every x ∈ C([0, T ]; Rd). Since the above equalities hold for every t ∈ [0, T ], the claim

(29)

Appendix C: Smooth variational principle on Λ

C.1. Proof of Lemma

3.1

Proof of Lemma 3.1. We split the proof into several steps. Step I. Proof of item 1). We first notice that ˆκ(t0,x0)

∞ is a non-anticipative map. Now,

let (t, ˆx) ∈ ˆΛ, h ∈ R\{0}, and i = 1, . . . , d (recall that e1, . . . , ed denotes the standard

orthonormal basis of Rd), then we have

ˆ κ(t0,x0) ∞ (t, ˆx+ h ei1[t,T ]) − ˆκ(t∞0,x0)(t, ˆx) h = Z Rd ˆx(· ∧ t) − x0(· ∧ t0) − z 1[t,T ] ζ(z + h ei) − ζ(z) h dz h→0 −→ Z Rd ˆx(· ∧ t) − x0(· ∧ t0) − z 1[t,T ] ∂ziζ(z) dz,

where ∂ziζ(z) denotes the partial derivative of ζ in the ei-direction at the point z, which

is given by −ziζ(z). This proves that ˆκ(t∞0,x0) admits first-order vertical derivatives at

every (t, ˆx) ∈ ˆΛ. In a similar way we can prove that ˆκ(t0,x0)

∞ also admits second-order

vertical derivatives at every (t, ˆx) ∈ ˆΛ.

Step II. Proof of item 2). We begin noticing that |ˆκ(t0,x0) ∞ (t, ˆx+ h ei1[t,T ]) − ˆκ∞(t0,x0)(t, ˆx)| = Z Rd ˆx(· ∧ t) − x0(· ∧ t0) − (z − h ei) 1[t,T ] ζ(z) dz − Z Rd ˆx(· ∧ t) − x0(· ∧ t0) − z 1[t,T ] ζ(z) dz ≤ |h ei| = |h|, where we have used the fact that R

Rdζ(z) dz = 1. It is then easy to see that, for every

i = 1, . . . , d, ∂V xiˆκ

(t0,x0)

∞ is bounded by the constant 1. Proceeding along the same lines

as for ˆκ(t0,x0) ∞ , we deduce that |∂xViκˆ (t0,x0) ∞ (t, ˆx+ h ej1[t,T ]) − ∂xViˆκ (t0,x0) ∞ (t, ˆx)| is bounded by |h|R Rd|∂ziζ(z)| dz = |h| R Rd|zi| ζ(z) dz = |h| q 2

π. This allows to prove that, for every

i, j = 1, . . . , d, ∂V xixjˆκ (t0,x0) ∞ is bounded by q 2 π.

Step III. Proof of item 3). We begin noting that

ˆ κ(t0,x0) ∞ (t, ˆx) = Z Rd xˆ(· ∧ t) − x0(· ∧ t0) − z 1[t,T ] ∞ζ(z) dz − Z Rd|z| ζ(z) dz ≥ −C ζ,

with Cζ :=RRd|z| ζ(z) dz, which proves the first part of item 3).

Now, let (t, x) ∈ Λ. Using the fact that ζ is a radial function, we have (when d = 1, (−∞, 0] × R0and [0, +∞) × R0 stand for (−∞, 0] and [0, +∞), respectively)

Z Rd x(· ∧ t) − x0(· ∧ t0) − z1[t,T ] ∞ζ(z) dz

(30)

= Z [0,+∞)×Rd−1 x(· ∧ t) − x0(· ∧ t0) − z1[t,T ] ζ(z) dz + Z (−∞,0]×Rd−1 x(· ∧ t) − x0(· ∧ t0) − z1[t,T ] ∞ζ(z) dz = Z [0,+∞)×Rd−1 x(· ∧ t) − x0(· ∧ t0) − z1[t,T ] ζ(z) dz + Z [0,+∞)×Rd−1 x(· ∧ t) − x0(· ∧ t0) + z1[t,T ] ∞ζ(z) dz. (C.1)

Now, we observe that, for every z ∈ Rd, we have

x(· ∧ t) − x0(· ∧ t0) − z1[t,T ] ∞ = maxn x(· ∧ t) − x0(· ∧ t ∧ t0) ∞, maxt≤s≤T x(t) − x0(s ∧ t0) − z o (C.2)

and similarly for kx(· ∧ t) − x0(· ∧ t0) + z1[t,T ]k∞. Moreover, by the elementary inequality

|x − z| + |x + z| ≥ 2|x|, valid for every x, z ∈ Rd, we have

max t≤s≤T x(t) − x0(s ∧ t0) − z + max t≤s≤T x(t) − x0(s ∧ t0) + z ≥ max t≤s≤T n x(t) − x0(s ∧ t0) − z + x(t) − x0(s ∧ t0) + z o ≥ 2 max t≤s≤T x(t) − x0(s ∧ t0) .

Then, using the elementary fact that if a + b ≥ 2c it holds that max{ℓ, a} + max{ℓ, b} ≥ 2 max{ℓ, c}, for every a, b, c, ℓ ∈ R, we find

maxn x(· ∧ t) − x0(· ∧ t ∧ t0) , max t≤s≤T x(t) − x0(s ∧ t0) − z o + maxn x(· ∧ t) − x0(· ∧ t ∧ t0) ∞, maxt≤s≤T x(t) − x0(s ∧ t0) + z o ≥ 2 maxn x(· ∧ t) − x0(· ∧ t ∧ t0) ∞, maxt≤s≤T x(t) − x0(s ∧ t0) o .

By (C.2) it follows that the last quantity coincides with 2 kx(·∧t)−x0(·∧t0)k∞. Therefore,

by (C.1) we obtain (also recalling that Cζ :=RRd|z| ζ(z) dz)

κ(t0,x0) ∞ (t, x) + Cζ = Z Rd x(· ∧ t) − x0(· ∧ t0) − z1[t,T ] ζ(z) dz ≥ 2 Z [0,+∞)×Rd−1 x(· ∧ t) − x0(· ∧ t0) ζ(z) dz = x(· ∧ t) − x0(· ∧ t0) ,

which concludes the proof of item 3). Finally, the explicit expression of the constant Cζ,

reported in (3.3), follows from the fact that Cζ = µχ2(d),1

2, where µχ2(d),12 denotes the

Références

Documents relatifs

Once provided these preliminaries, we can introduce the PDE on [0, T] × H associated with the original PPDE on Θ , recall the notion of viscosity solution in Hilbert spaces, and

Viscosity Solutions of Fully Nonlinear Elliptic Path Dependent Partial Differential Equations.. Zhenjie Ren ∗ October

En outre, la méthode de Perron est non seulement utile pour prouver l’existence de solutions de viscosité, mais a également des applications dans de divers contextes, par exemple,

(ii) the uniqueness may be easier under our definition because our notion of viscosity solution is constrained by a bigger set of test functions; indeed recently comparison results

First of all, Bonnet's theorem tells us that for a compact manifold with positive sectional curvature, the fundamental group must be finite and this was later generalized by

BAOUENDI.- Solvability of partial differential equations in the traces of analytic solutions of the heat equation,- Technical report, The University of Kansan, Lawrence,

The principal local result, The Local Ex- tension Theorem (Appendix 6), is also the main step in all the appli- cations of the Convex Integration technique to solving open and

We prove existence and uniqueness of Crandall-Lions viscosity solutions of Hamilton- Jacobi-Bellman equations in the space of continuous paths, associated to the optimal control