• Aucun résultat trouvé

McKean SDEs with singular coefficients

N/A
N/A
Protected

Academic year: 2022

Partager "McKean SDEs with singular coefficients"

Copied!
27
0
0

Texte intégral

(1)

HAL Id: hal-03306570

https://hal.archives-ouvertes.fr/hal-03306570

Preprint submitted on 29 Jul 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.

McKean SDEs with singular coefficients

Elena Issoglio, Francesco Russo

To cite this version:

Elena Issoglio, Francesco Russo. McKean SDEs with singular coefficients. 2021. �hal-03306570�

(2)

Elena Issoglio and Francesco Russo

Abstract. The paper investigates existence and uniqueness for a sto- chastic differential equation (SDE) with distributional drift depending on the law density of the solution. Those equations are known as McK- ean SDEs. The McKean SDE is interpreted in the sense of a suitable singular martingale problem. A key tool used in the investigation is the study of the corresponding Fokker-Planck equation.

Key words and phrases. Stochastic differential equations; distribu- tional drift; McKean; Martingale problem.

2020 MSC. 60H10; 60H30; 35C99; 35D99; 35K10.

1. Introduction

In this paper we are concerned with the study of singular McKean SDEs of the form

( Xt=X0+Rt

0F(v(s, Xs))b(s, Xs)ds+Wt

v(t,·) is the law density ofXt, (1) for some given initial condition X0 with density v0. The terminology Mc- Kean refers to the fact that the coefficient of the SDE depends on the law of the solution process itself, while singularreflects the fact that one of the coefficients is a Schwartz distribution. The main aim of this paper is to solve the singular McKean problem (1), that is, to define rigorously the meaning of equation (1) and to find a (unique) solution to the equation. The key novelty is theirregularity of the drift, which is encoded in the termb.

The problem is d-dimensional, in particular the process X takes values in X ∈ Rd, the function F is F : R → Rd×n, the term b is formally b : [0, T]×Rd→RnandW is ad-dimensional Brownian motion, wheren, dare two integers. We assume that b(t,·)∈ Cβ(Rn) for some 0< β < 1/2 (see below for the definition of Besov spaces Cβ(Rn)), which means that b(t,·) is a Schwartz distribution and thus the term b(t, Xt), as well as its product withF, are only formal at this stage. The function F is nonlinear.

When b is a function, equation (1) has been recently studied by several authors. For example [19] study existence and uniqueness of the solution un- der some regularity assumptions on the drift, while [22] requires the drift to be of a special form, being Lipschitz-continuous with respect to the variable v, uniformly in time and space, and measurable with respect to space. We

Date: 16th July 2021.

1

(3)

also mention [2], where the authors obtain existence of the solution when as- suming that the drift is a measurable function. For other past contributions see [18].

In the literature we also find some contributions on (1) with F ≡1, i.e.

when there is no dependence on the law v but the drift b is a Schwartz distribution. In this case equation (1) becomes an SDE with singular drift.

Ordinary SDEs with distributional drift were investigated by several au- thors, starting from [10, 9, 3, 23] in the one-dimensional case. In the multi- dimensional case it was studied by [8] with bbeing a Schwartz distribution living in a fractional Sobolev space of negative order (up to −12). After- wards, [5] extended the study to a smaller negative order (up to −23) and formulated the problem as a martingale problem. We also mention [17], where the singular SDE is studied as a martingale problem, with the same setting as in the present paper (in particular the drift belongs to a negative Besov space rather than a fractional Sobolev space). Backwards SDEs with similar singular coefficients have also been studied, see [15, 16].

The main analytical tool in the works cited above is the study of an asso- ciated singular PDE (either Kolmogorov or Fokker-Planck). In the McKean case, the relevant PDE associated to equation (1) is the nonlinear Fokker- Planck equation

tv= 12∆v−div( ˜F(v)b)

v(0) =v0, (2)

where ˜F(v) :=vF(v). PDEs with similar (ir)regular coefficients were stu- died in the past, see for example [8, 13] for the study of singular Kolmogorov equations. One can then use results on existence, uniqueness and continuity of the solution to the PDE (e.g. with respect to the initial condition and the coefficients) to infer results about the stochastic equation. For example in [8], the authors use the singular Kolmogorov PDE to define the meaning of the solution to the SDE and find a unique solution.

Let us remark that the PDEs mentioned above are a classical tool in the study of McKean equations when the dependence on the law density of the process is pointwise, which is the case in the present paper where we have F(v(t, x)). There is, however, a large body of literature that studies McKean equations where the drift depends on the law more regularly, typically it is assumed to be Lipschitz-continuous with respect to the Wasserstein metric.

In this case the McKean equation is treated with different techniques than the ones explained above, in particular it is treated with probabilistic tools.

This is nowadays a well-known approach, for more details see for example the recent books by Carmona and Delarue [6, 7], see also [21, 20].

Our contribution to the literature is twofold. The first and main novel result concerns the notion of solution to the singular McKean equation (1) (introduced in Definition 6.1) and its existence and uniqueness (proved in Theorem 6.4). The second contribution is the study of the singular Fokker- Plank equation (2), in particular we find a unique solutionv∈C([0, T];Cβ+)

(4)

in the sense of Schwartz distributions, see Theorem 3.7 for existence and Proposition 4.7 for uniqueness.

The paper is organised as follows. In Section 2 we introduce the notation and recall some useful results on semigroups and Besov spaces. We also recall briefly some results on the singular martingale problem. In Section 3 we study the singular Fokker-Planck PDE (2). Then we consider a mollified version of the PDE and the SDE in Sections 4 and 5, respectively. Finally in Section 6 we use the mollified PDEs and SDEs and their limits to study (1) and we prove our main theorem of existence and uniqueness of a solution to (1). In Appendix A we recall a useful fractional Gronwall’s inequality. In Appendix B we show a characterization of continuity and compactness in inductive spaces.

2. Setting and useful results

2.1. Notation and definitions. Let us denote byCbuc1,2 :=Cbuc1,2([0, T]×Rd) the space of all C1,2-functions such that the function and its gradient in x are bounded, and the Hessian matrix is uniformly continuous. We also use the notationC0,1:=C0,1([0, T]×Rd) to indicate the space of functions with gradient in x uniformly continuous in (t, x). By a slight abuse of notation we use the same notation Cbuc1,2 and C0,1 for functions which are Rd-valued.

When f :Rd →Rd is differentiable, we denote by ∇f the matrix given by (∇f)i,j = ∂ifj. When f : Rd → R we denote the Hessian matrix of f by Hess(f).

We denote by S = S(Rd) the space of Schwartz functions on Rd and by S = S(Rd) the space of Schwartz distributions. For γ ∈ R we denote by Cγ = Cγ(Rd) the Besov space or H¨older-Zygmund space and by k · kγ

its norm (see for example [1, Section 2.7]). We recall that for γ < γ one has Cγ ⊂ Cγ. If γ ∈ R+ \N then the space coincides with the classical H¨older space of functions which are ⌊γ⌋-times differentiable and such that the⌊γ⌋th derivative is (γ−⌊γ⌋)-H¨older continuous. For example ifγ ∈(0,1) the classical γ-H¨older norm

kfk+ sup

x6=y,|xy|<1

|f(x)−f(y)|

|x−y|γ , (3) is an equivalent norm in Cγ. With an abuse of notation we use kfkγ to denote (3). For this and for more details see, for example, [25, Chapter 1]

or [1, Section 2.7]. Notice that we use the same notation Cγ to indicate R-valued functions but also Rd or Rd×d-valued functions. It will be clear from the context which space is needed.

We denote byCTCγthe space of continuous functions on [0, T] taking values inCγ, that isCTCγ :=C([0, T];Cγ). For any given γ ∈Rwe denote by Cγ+

and Cγ the spaces given by

Cγ+:=∪α>γCα, Cγ :=∩α<γCα.

(5)

Notice thatCγ+is an inductive space. We will also use the spacesCTCγ+:=

C([0, T];Cγ+), with the meaning thatf ∈CTCγ+ if and only if there exists α > γ such that f ∈ CTCα, see Lemma B.2 in Appendix B for a proof of the latter fact.

Similarly, we use the space CTCγ := C([0, T];Cγ), meaning that f ∈ CTCγ if and only if for any α < γ we have f ∈CTCα. Notice that if f is continuous and such that ∇f ∈CTC0+ thenf ∈C0,1.

Let (Pt)t denote the semigroup generated by 12∆ on S, in particular for all φ ∈ S we define (Ptφ)(x) := R

Rdpt(x−y)φ(y)dy, where the kernel p is the usual heat kernel

p(t, z) = 1

(2πt)d/2 exp{−|z|2

t }. (4)

It is easy to see that Pt :S → S. Moreover we can extend it to S by dual pairing (and we denote it with the same notation for simplicity). One has hPtψ, φi=hψ, Ptφifor eachφ∈ S andψ∈ S, using the fact that the kernel is symmetric.

Lemma 2.1. Let g : [0, T]→ S(Rd) be continuous and w0 ∈ S(Rd). The unique (weak) solution of

tw= 12∆w+g w(0) =w0 is given by

Ptw0+ Z t

0

Ptsg(s)ds, t∈[0, T]. (5) By weak solution we mean, for every ϕ ∈ S(Rd) and t ∈ [0, T] we have hw(t), ϕi=hw0, ϕi+Rt

0hw(s),12∆ϕids+Rt

0hg(s), ϕids.

Proof. The fact that (5) is a solution is done by inspection. The uniqueness

is a consequence of Fourier transform.

In the whole article the lettercor C will denote a generic constant which may change from line to line.

2.2. Some useful results. In the sections below, we are interested in the action of Pt on elements of Besov spaces Cγ. These estimates are known as Schauder’s estimates. For a proof we refer to [11], see also [12] for similar results.

Lemma 2.2 (Schauder’s estimates). Let f ∈ Cγ for some γ ∈R. Then for any θ≥0 there exists a constant c such that

kPtfkγ+2θ ≤ctθkfkγ, (6) for all t >0.

Moreover for f ∈ Cγ and for any θ∈(0,1) we have

kPtf−fkγ ≤ctθkfkγ+2θ. (7)

(6)

Notice that from (7) it readily follows that iff ∈ Cγ for some 0 < θ <1, then fort > s >0 we have

kPtf−Psfkγ≤c(t−s)θkfkγ+2θ. (8) In other words, this means that if f ∈ Cγ+θ thenP·f ∈CTCγ (and in fact it isθ-H¨older continuous in time). We also recall that Bernstein’s inequalities hold (see [1, Lemma 2.1] and [12, Appendix A.1]), that is for γ ∈ R there exists a constant c >0 such that

k∇gkγ ≤ckgkγ+1, (9) for allg∈ C1+γ. Using Schauder’s and Bernstein’s inequalities we can easily obtain a useful estimate on the gradient of the semigroup, as we see below.

Lemma 2.3. Let γ ∈R and θ∈(0,1). If g∈ Cγ then for all t >0 we have

∇(Ptg)∈ Cγ+2θ1 and

k∇(Ptg)kγ+2θ1 ≤ctθkgkγ. (10) The following is an important estimate which allows to define the so called pointwise productbetween certain distribution and functions, which is based on Bony’s estimates. For details see [4] or [12, Section 2.1]. Letf ∈ Cα and g ∈ Cβ with α−β > 0 and α, β >0. Then the pointwise product f g is well-defined as an element of Cβ and there exists a constant c > 0 such that

kf gkβ ≤ckfkαkgkβ. (11) Moreover if f and g are continuous functions defined on [0, T] with values in the above Besov spaces, then the product is also continuous with values inCβ, and

kf gkCTC−β ≤ckfkCTCαkgkCTC−β. (12) 2.3. Assumptions. We now collect the assumptions on the distributional termb, the nonlinearity F and ˜F and on the initial conditionv0that will be used later on in order for PDE (2) to be well-defined and for the McKean- Vlasov problem (1) to be solved.

Assumption 1. Let 0< β <1/2 and b∈CTCβ.

Assumption 2. Let (bn)be a sequence of bounded functions in CTCβ that converges to b in CTCβ. Moreover for each n, let bn(t,·) ∈ Cb for all t∈[0, T].

Example 2.4. Assumption 2 is satisfied for instance in the following case.

Let β < β and let b∈ CTCβ. We define the sequence (bn) for any fixed t∈[0, T]and for all n≥1 by

bn(t,·) :=φn∗b(t,·),

where φn(x) := p1/n2(x) and p is the Gaussian kernel defined in (4). If ψ∈ S then φn∗ψ=P1/n2ψ, thus we have bn(t,·) =P1/n2b(t,·).

(7)

(i) We now show that t7→bn(t,·) is continuous in Cβ. For any t, s∈ [0, T] we have

kbn(t,·)−bn(s,·)kβ =kP1/n2b(t,·)−P1/n2b(s,·)kβ

=kP1/n2(b(t,·)−b(s,·))kβ

≤ckb(t,·)−b(s,·)kβ

≤ckb(t,·)−b(s,·)kβ,

having used estimate (6)in Lemma 2.2 (withθ= 0). The conclusion now follows.

(ii) We show bn→b in CTCβ. For t∈[0, T]we have kbn(t,·)−b(t,·)kβ =kP1/n2b(t,·)−b(t,·)kβ

≤c 1

n2 β−β

2

kb(t,·)kβ,

having used (7) in Lemma 2.2. Now we take the sup over t∈[0, T] and we have kbn−bkCTC−β →0 asn→ ∞, since β−β >0.

Assumption 3. Let F be Lipschitz and bounded.

Assumption 4. Let F(z) :=˜ zF(z) be globally Lipschitz.

We believe that Assumption 4 is unnecessary. Indeed by Assumption 3 one gets that ˜F is locally Lipschitz with linear growth. This condition could be sufficient to show that a solution PDE (2) exists, for example using techniques similar to the ones appearing in [14, Proposition 3.1] and [22, Theorem 22]. However we assume here ˜F to be Lipschitz to improve the readability of the paper.

Assumption 5. Let v0∈ Cβ+.

Assumption 6. Let v0 be a bounded probability density.

2.4. The singular Martingale Problem. We conclude this section with a short recap of useful results from [17], where the authors consider the Martingale Problem for SDEs of the form

Xt=X0+ Z t

0

B(s, Xs)ds+Wt, (13) where B satisfies Assumption 1 (with b = B). Notice that this SDE can be considered as the linear counterpart of the McKean-Vlasov problem (1), which can be obtained for example by ‘fixing’ a suitable function v and considering B=F(v)bin the SDE in (1).

First of all, let us recall the definition of the operatorLassociated to SDE (13) given in [17]. The operatorL is defined as

L: DL0 → {S-valued integrable functions}

f 7→ Lf := ˙f+12∆f+∇f B, (14)

(8)

where

DL0 :=CTC(1+β)+∩AC([0, T];S)

and AC stands for “absolutely continuous”. Here f : [0, T]×Rd → R and the function ˙f : [0, T]→ S is defined for anyf ∈ D0L as the unique ˙f such that f(t)−f(0) = Rt

0f˙(s)ds, which always exists because f is absolutely continuous. Note also that∇f Bis well-defined using (11) and Assumption 1. The Laplacian ∆ is intended in the sense of distributions.

Next we give the definition of solution to the martingale problem in [17]:

a couple (X,P) is a solution to the martingale problem with distributional drift B (for shortness, solution of MP) if and only if for every f ∈ DL

f(t, Xt)−f(0, X0)− Z t

0

(Lf)(s, Xs)ds (15) is a local martingale underP, where the domain DL is given by

DL:= {f ∈ D0L: such that ∃g∈CTC0+

such thatf is a weak solution of Lf =g}, (16) and L has been defined in (14). We say that the martingale problem with distributional drift B admits uniquenessif, whenever we have two solutions (X1,P1) and (X2,P2), then the law of X1 under P1 equals the law of X2 under P2. With this definition at hand, we show in [17] that MP admits existence and uniqueness.

3. Fokker-Planck singular PDE

This section is devoted to the study of the singular Fokker-Planck equa- tion (2), recalled here for ease of reading

tv= 12∆v−div( ˜F(v)b) v(0) =v0.

After introducing the notions of solution for this PDE (weak and mild, which turns out to be equivalent), we will show that there exists a solution in Theorem 3.7 with Schaefer’s fixed point theorem. We will show with different techniques (see Section 4, in Proposition 4.7), that such solution is unique.

Below we will need mapping properties of the function ˜F when viewed as operator acting on Cα, for some α ∈ (0,1). To this aim, we make a slight abuse of notation and denote by ˜F the function when viewed as an operator, that is forf ∈ Cα we have ˜F(f) := ˜F(f(·)). We sometimes omit the brackets and write ˜F f in place of ˜F(f). The result below on ˜F is taken from [14], Proposition 3.1 and equation (32).

Lemma 3.1 (Issoglio [14]). Under Assumption 4 and if α∈(0,1) then

• F˜ :Cα → Cα and for all f, g ∈ Cα

kF f˜ −F g˜ kα ≤c(1 +kfk2α+kgk2α)1/2kf−gkα;

(9)

• for all f ∈ Cα then kF f˜ kα≤c(1 +kfkα).

This mapping property allows us to define weak and mild solutions for the singular Fokker-Planck equation.

Definition 3.2. Let Assumptions 1, 4 and 5 hold and let v∈CTCβ+. (i) We say that v is a mild solution for the singular Fokker-Planck

equation (2) if for all t ∈ [0, T] the following integral equation is satisfied

v(t) =Ptv0− Z t

0

Pts[div( ˜F(v(s))b(s))]ds. (17) (ii) We say that v is a weak solution for the singular Fokker-Planck

equation (2) if for allϕ∈ S(Rd) and all t∈[0, T] we have hϕ, v(t)i=hϕ, v0i+

Z t

0 h1

2∆ϕ, v(s)ids+ Z t

0 h∇ϕ,F˜(v)(s)b(s)ids. (18) Note that the termF˜(v)(s)b(s) appearing in both items is well-defined as an element of Cβ thanks to (11) and Assumption 1 together with Lemma 3.1.

Proposition 3.3. Let v ∈ CTCβ+. The function v is a weak solution of PDE (2) if and only if it is a mild solution.

Proof. This is a consequence of Lemma 2.1 withg(s) :=−div( ˜F(v(s))b(s)).

We are interested in finding a mild solution of (2) in the space CTCβ+. To do so we will apply Schafer’s fixed point theorem, following similar ideas as done in [14, Section 4]. To this end, we state and prove a few preparatory results, including a priori estimates and mapping properties of the solution map. Let us denote byJ the solution map for the mild solution of PDE (2), that is for v∈CTCα for someα∈(0,1) we have

Jt(v) :=Ptv0− Z t

0

Pts[div( ˜F(v(s))b(s))]ds.

Then a mild solution of (2) is a solution of v =J(v), in other words it is a fixed point of J. In the proofs below we will also use the notation

Gs(v) := ˜F(v(s))b(s) (19)

for brevity.

We present now an a priori bound for mild solutions, if they exist.

Proposition 3.4. Let Assumptions 1, 4 and 5 hold. Let α ∈(β,1−β). If v∈CTCα is such that v=λJ(v) for some λ∈[0,1], then we have

kvkCTCα ≤K,

where K is a constant depending on kv0kα,kbkCTC−β, T but independent of λ. Moreover K is an increasing function of kbkCTC−β.

(10)

It is obvious that choosing λ= 1 we get a mild solution of (2). Here we state the result for more generalλbecause it will be needed later on.

Proof. Using Bernstein’s inequality (9) we get kdivGs(v)kβ1

d

X

i=1

k ∂

∂xiGs(v)kβ1 ≤c

d

X

i=1

kGs(v)kβ.

Then using the definition of G from (19), pointwise product property (11) (since α−β >0) and Lemma 3.1 we have

kdivGs(v)kβ1 ≤ckF˜(v(s))kαkb(s)kβ ≤c(1 +kv(s)kα)kb(s)kβ. (20) Now using this, together with Schauder’s estimates (Lemma 2.2 with θ :=

α+β+1

2 ) and the fact thatθ <1, for fixed t∈[0, T], one obtains kv(t)kα ≤ kλPtv0kα+

Z t

0 kλPts[divGs(v)]kαds

≤ckv0kα+ Z t

0

c(t−s)α+β+12 kdivGs(v)kβ1ds

≤ckv0kα+c Z t

0

c(t−s)α+β+12 (1 +kv(s)kα)kb(s)kβds

≤ckv0kα+ckbkCTC−β

Z t 0

(t−s)α+β+12 (1 +kv(s)kα)ds

≤ckv0kα+ckbkCTC−βT1−α−β2 +ckbkCTC−β

Z t 0

(t−s)α+β+12 kv(s)kαds.

Now by a generalised Gronwall’s inequality (see Lemma A.1) we have kv(t)kα ≤[ckv0kα+ckbkCTC−βT1−α−β2 ]Eη(ckbkCTC−βΓ(η)tη),

with η = −α+β+12 + 1 = 1α2β > 0 and where Eη is the Mittag-Leffler function, see Lemma A.1. Now taking the sup over t∈[0, T] and using the fact that Eη is increasing we get

kvkCTCα

≤h

ckv0kα+ckbkCTC−βT1−α−β2 i Eη

ckbkCTC−βΓ

1−α−β 2

T1−α−β2

ckv0kα+ckbkCTC−βT

Eη ckbkCTC−βΓ(1)T

=:K.

This concludes the proof.

The next result is about mapping properties of the solution map J.

Proposition 3.5. Let Assumptions 1, 4 and 5 hold. Let us fix α, α, ε such that β < α ≤ α < 1−β, 1−α −β −2ε > 0, ε > 0 and such that v0 ∈ Cα+2ε, which is always possible thanks to Assumptions 1 and 5.

(11)

(i) For any v ∈CTCα we have kJ(v)kCε

TCα ≤ckv0kα+2ε +ckbkCTC−βT1−α′−β−2ε

2 (1 +kvkCTCα) (21) and for any v, u∈CTCα we have

sup

t[0,T]kJt(v)−Jt(u)kα ≤cT1−α′−β2 (1 +kuk2CTCα+kvk2CTCα)1/2 (22)

× kv−ukCTCαkbkCTC−β,

kJt(v)−Jt(u)−(Js(v)−Js(u))kα ≤c(t−s)εT1−α′−β−2ε

2 kbkCTC−β (23)

×(1 +kuk2CTCα+kvk2CTCα)1/2kv−ukCTCα, ∀s < t∈[0, T].

(ii) J :CTCα→CTεCα and it is continuous.

(iii) J :CTCβ+→CTCβ+ and it is continuous.

Proof. Item (iii) is a direct consequence of Item (ii). Item (ii) is a conse- quence of Item (i). In fact, the mapping property follows from (21). As far as continuity is concerned, equations (22) and (23) allow us to bound the norm

kJ(v)−J(u)kCε

TCα = sup

t[0,T]kJt(v)−Jt(u)kα (24) + sup

0s<tT

kJt(v)−Jt(u)−(Js(v)−Js(u))kα

(t−s)ε .

We now show Item (i) in 5 steps. Notice that v0 ∈ Cα because α <

α+ 2ε. Letv∈CTCα.

Step 1. Let0≤t≤T. We show that Jt(v)∈ Cα.

Using the definition ofJ, Schauder’s estimate for the semigroup, estimate (9) and the bound (20) we have

kJt(v)kα ≤ kPtv0kα+ Z t

0 kPts[divGs(v)]kαds

≤ckv0kα +c Z t

0

(t−s)α

+β+1

2 kdivGs(v)kβ1ds

≤ckv0kα +c Z t

0

(t−s)α

+β+1

2 (1 +kv(s)kα)kb(s)kβds

≤ckv0kα +cT1−α′−β2 (1 +kvkCTCα)kbkCTC−β. (25) Step 2. Let0≤s < t≤T. We show

kJt(v)−Js(v)kα ≤c(t−s)εkv0kα+2ε (26) +c(t−s)εT1−α′−β−2ε

2 (1 +kvkCTCα)kbkCTC−β.

(12)

We have

kJt(v)−Js(v)kα ≤k(Pts−I)(Psv0)kα

+k Z s

0

(Pts−I)(Psr[divGr(v))]drkα

+k Z t

s

Ptr[divGr(v)]drkα

=:M1+M2+M3.

ForM1we recall thatv0∈ Cα+2ε and we use Schauder’s estimate (7) (with γ−2θ=α, θ=ε) and continuity of the semigroup to get

M1 ≤c(t−s)εkPsv0kα+2ε ≤c(t−s)εkv0kα+2ε. ForM2 we use (7) as well as (6) and (20) to get

M2 ≤c(t−s)ε Z s

0 k(Psr[divGr(v))]kα+2εdr

≤c(t−s)ε Z s

0

(s−r)α

+β+1+2ε

2 kdivGr(v)kβ1dr

≤c(t−s)εT1−α′−β−

2 (1 +kvkCTCα)kbkCTC−β. ForM3 we use only (6) and (20) to get

M3 ≤c Z t

s

(t−r)α

+β+1

2 kdivGr(v)kβ1dr

≤c Z t

s

(t−r)α

+β+1

2 dr(1 +kvkCTCα)kbkCTC−β

≤c(t−s)1−α′−β2 (1 +kvkCTCα)kbkCTC−β

≤c(t−s)εT1−α′−β−2ε

2 (1 +kvkCTCα)kbkCTC−β.

Putting the three bounds forM1, M2, M3 together we can conclude that (26) holds.

Step 3. We show bound (21).

Using Step 1 and Step 2 we have kJ(v)kCε

TCα = sup

t[0,T]kJt(v)kα + sup

0s<tT

kJt(v)−Js(v)kα

(t−s)ε

≤ckv0kα +ckv0kα+2ε

+cT1−α′−β2 (1 +kvkCTCα)kbkCTC−β

+ 2T1−α′−β−2ε

2 (1 +kvkCTCα)kbkCTC−β

≤ckv0kα+2ε +ckbkCTC−βT1−α′−β−2ε

2 (1 +kvkCTCα), which concludes the proof of (21).

(13)

Step 4. We show bound (22).

By Schauder’s estimates and Lemma 3.1 we get sup

t[0,T]kJt(v)−Jt(u)kα

= sup

t[0,T]k Z t

0

Pts[divGs(v)−divGs(u)]dskα

= sup

t[0,T]k Z t

0

Pts[div( ˜F(v(s))−F˜(u(s))b(s))]dskα

≤ sup

t[0,T]k Z t

0

(t−s)α

+β+1

2 kF˜(v(s))−F(u(s))˜ kαkb(s)kβds

≤ sup

t[0,T]k Z t

0

(t−s)α

+β+1

2 ds(1 +kuk2CTCα+kvk2CTCα)1/2kv−ukCTCαkbkCTC−β

≤cT1−α′−β2 (1 +kuk2CTCα+kvk2CTCα)1/2kv−ukCTCαkbkCTC−β. Step 5. We show bound (23).

We have

kJt(v)−Jt(u)−(Js(v)−Js(u))kα

≤ kPtv0−Ptv0−(Psv0−Psv0)kα

+

Z s

0

(Pts−I)Psr[div([ ˜F(v(r))−F˜(u(r))]b(r))]dr α

+

Z t s

Ptr[div([ ˜F(v(r))−F(u(r))]b(r))]dr˜ α

=:N1+N2+N3.

Notice that N1 = 0. ForN2 and N3 we do similar computations as for M2 and M3 in Step 2, respectively, and using also Lemma 3.1 we get

N2 ≤c(t−s)ε Z s

0

(s−r)α

+β+1+2ε

2 drkbkCTC−β×

×(1 +kuk2CTCα +kvk2CTCα)1/2kv−ukCTCα

≤c(t−s)εT1−α′−β−2ε

2 kbkCTC−β(1 +kuk2CTCα +kvk2CTCα)1/2kv−ukCTCα

and

N3 ≤c(t−s)1−α′−β2 (1 +kuk2CTCα+kvk2CTCα)1/2kv−ukCTCα

≤c(t−s)εT1−α′−β−2ε

2 kbkCTC−β(1 +kuk2CTCα +kvk2CTCα)1/2kv−ukCTCα, where 1α2β >0 by choice of the parameters.

A consequence of Proposition 3.5 is the corollary below.

(14)

Corollary 3.6. Letα, αandε be chosen according to Proposition 3.5. Then for all εsuch that ε ≥ε >0 we have J :CTεCα →CTεCα is continuous.

Proof. We observe that CTCα is continuously embedded in CTεCα for all ε >0. By part (ii) of Proposition 3.5 we then get J :CTεCα → CTεCα and

the mapping is continuous.

We are now ready to show that a solution to (2) exists.

Theorem 3.7. Let Assumptions 1, 4 and 5 hold. Then there exists a mild solution v of (2). Moreover there exists ε > 0 and α > β, only depending on v0, such that v∈CTεCα⊂CTCβ+.

Proof. Let us fix here ε > 0 and α > β such that v0 ∈ Cα+2ε and such that α < 1−β and 1−α−β−2ε > 0, which is always possible thanks to Assumptions 1 and 5. We choose α such thatβ < α≤α and we choose ε such that 0 < ε ≤ ε. Therefore we can apply Corollary 3.6 which tells that J :CTεCα → CTεCα is continuous and thus also J : CTεCα → CTεCα is continuous. We next show that the same operator J is compact. Indeed, by bound (21) and kvkCTCα ≤ kvkCεTCα we have that the image of a ball in CTεCα is a ball in CTεCα, and balls in the latter are precompact sets in CTεCα. The idea is to apply Schaefer’s fixed point theorem, see [24, Theorem 4.3.2]. For this we further need to show that the set

Λ :={v∈CTεCα:v=λJ(v) for someλ∈[0,1]}

is bounded inCTεCα. Notice that (ε, α, α) satisfy the assumptions on (ε, α, α) from Proposition 3.5, in particular 1−α−β−2ε > 0 and v0 ∈ Cα+2ε. If v∈Λ, then by (21) in Proposition 3.5 and by Proposition 3.4 we have

kvkCTεCα =λkJ(v)kCTεCα

≤ kJ(v)kCTεCα

≤ckv0kα+2ε+ckbkCTC−βT1−α−β−2ε2 (1 +kvkCTCα)

≤ckv0kα+2ε+ckbkCTC−βT1−α−β−2ε2 (1 +K)

=:K <∞,

whereKis independent ofv. Thus by Schaefer’s fixed point theorem we can conclude that there exists a fixed pointvofJinCTεCα,v=J(v), and such v is a mild solution of (2) in CTCβ+ sinceCTεCα⊂CTεCβ+⊂CTCβ+. Remark 3.8. We can show that the solution v is more regular if we sup- pose that v0 ∈ C(1β) in place of Assumption 5. In this case with similar arguments we could get that a solution v exists inCTC(1β).

Remark 3.9. We will prove below (with other techniques) that the solution v found in Theorem 3.7 is actually unique in CTCβ+, see Proposition 4.7.

(15)

4. The regularised PDE and its limit

Let Assumptions 1, 2, 3, 4 and 5 hold throughout in this section.

When the term b is replaced by bn satisfying Assumption 2 we get a smoothed PDE, that is, we get the Fokker-Planck equation

tvn= 12∆vn−div( ˜F(vn)bn)

vn(0) =v0, (27)

where we recall that ˜F(vn) =vnF(vn). For ease of reading, we recall that the mild solution of (27) is given by an element vn∈CTCβ+ such that

vn(t) =Ptv0− Z t

0

Pts[div( ˜F(vn(s))bn(s))]ds. (28) Remark 4.1. We observe that, since bn ∈ CTCβ, then all results from Section 3 are still valid, in particular the bound from Proposition 3.5 and the existence result from Theorem 3.7 still apply to (27).

At this point we introduce the notation and some useful results on a very similar semilinear PDE studied in [22]. We consider the PDE

tu(t, x) = 12∆u(t, x)−div(u(t, x)b(t, x, u(t, x)))

u(0,dx) =u0(dx), (29)

whereu0 is a Borel probability measure which admitsv0 as bounded density with respect to the Lebesgue measure. We set

b(t, x, z) :=F(z)bn(t, x). (30) Thanks to Assumptions 3 and 2 we have that the termb(t, x, z) is uniformly bounded.

Definition 4.2. We will call asemigroup solutionof the PDE (29) a func- tion u∈L([0, T]×Rd) that satisfies the integral equation

u(t, x) = Z

Rd

pt(x−y)v0(y)dy +

d

X

j=1

Z t 0

Z

Rd

u(s, y)bj(s, y, u(s, y))∂yjpts(x−y)dyds, (31) where p is the Gaussian heat kernel introduced in (4).

Notice that this definition is inspired by [22, Definition 6], but we mo- dified it here to include the condition u ∈ L([0, T]×Rd), rather than u ∈ L1([0, T]×Rd) (the latter as in [22], where moreover the solution is called ‘mild solution’). Indeed integrability ofuis sufficient for the integrals in the semigroup solution to make sense, becauseb is also bounded and the heat kernel and its derivative are integrable.

The first result we have on (29) is about uniqueness of the semigroup solution in L([0, T]×Rd). This result is not included in [22], but we were inspired by proofs therein, in particular by the proof of [22, Lemma 20].

(16)

Lemma 4.3. There exists at most one semigroup solution of (29).

Proof. First of all we remark that sincept(y) is the heat kernel then we have two positive constants cp, Cp such that

|∂yjpt(y)| ≤ Cp

√tqt(y), (32) for all j = 1, . . . , d, where qt(y) = cpd/2

ecp|y|

2

t is a Gaussian probability density. This can be easily calculated from the explicit form of pt(y), since one gets

yjpt(y) =−yj t pt(y) so that

|∂yjpt(y)|= |yj|

t pt(y) = 1

√t

|√yj| tpt(y).

Now we observe that when |yj|

t ≤ 1 then |yj|

tpt(y) ≤ pt(y) so it can be bounded by qt(y) for suitable constant cp and (32) follows. On the other hand, when |yj|

t > 1, we have that pt(y) goes to zero exponentially with respect to |yj|

t, in particular there exists two positive constants Cp, cp such that

|√yj|

tpt(y) = |√yj| tc 1

td/2ec

|yj|2

t ≤ Cpcp

d/2

ecp|y|

2 t , and (32) follows.

Let us consider two semigroup solutionsu1, u2of (29). We denote by Π(u) the semigroup solution map, which is the right-hand side of (31). Notice that v0 ∈L(Rd) by Assumption 5, and the functionz7→zb(t, x, z) is Lipschitz, uniformly in t, x because ˜F is assumed to be Lipschitz in Assumption 4.

Using this together with the bound (32) for fixed t∈(0, T] we get kΠ(u1)(t,·)−Π(u2)(t,·)k

=

d

X

j=1

Z t

0

Z

Rd

u1(s, y)bj(s, y, u1(s, y))−u2(s, y)bj(s, y, u2(s, y))

·∂yjpts(x−y)dyds

≤C Z t

0

Z

Rd|u1(s, y)−u2(s, y)| 1

√t−sCuqts(x−y)dyds

≤C Z t

0 ku1(s,·)−u2(s,·)k 1

√t−sds· Z

Rd

qts(x−y)dy

≤C Z t

0 ku1(s,·)−u2(s,·)k

√1

t−sds.

Now, by an application of a fractional Gronwall’s inequality (see Lemma A.1) we conclude that ku1(t,·) −u2(t,·)k ≤ 0 for all t ∈ [0, T], so in

(17)

particular we have

ku1−u2kL([0,T]×Rd)= 0,

hence the semigroup solution is unique in L([0, T]×Rd).

At this point we want to compare the concept of mild solution and that of semigroup solution. Recall that b(t, x, z) = F(z)bn(t, x) so in fact PDE (29) is exactly (27). First we state and prove a preparatory lemma, where f is vector-valued and will be taken to be u(t, x)b(t, x, u(t, x)) for fixed tin the following result.

Lemma 4.4. Let f∈L(Rd;Rd), t∈[0, T]. Then

Pt(div f) =−

d

X

j=1

Z

Rd

f(y)∂yjpt(· −y)dy, (33)

almost everywhere.

Proof. We will show that the left-hand side (LHS) and the right-hand side (RHS) are the same object in S. Moreover we know that both sides are functions in L so we conclude. Notice that the heat kernel pt(x) is the same kernel associated to the semigroupPt, namely ifφ∈ S, thenPtφ∈ S withPtφ(x) =R

Rdpt(x−y)φ(y)dy. Let φ∈ S. The LHS gives

hPt(divf), φi=hdiv f, Ptφi

=−

d

X

j=1

hf, ∂yj(Ptφ)i

=−

d

X

j=1

f, ∂yj

Z

Rd

pt(· −x)φ(x)dx

=−

d

X

j=1

f,

Z

Rd

yjpt(· −x)φ(x)dx

.

(18)

The RHS of (33), on the other hand, gives h−

d

X

j=1

Z

Rd

f(y)∂yjpt(· −y)dy, φi=−

d

X

j=1

h Z

Rd

f(y)∂yjpt(· −y)dy, φi

=−

d

X

j=1

Z

Rd

Z

Rd

f(y)∂yjpt(x−y)dy φ(x)dx

=−

d

X

j=1

Z

Rd

f(y) Z

Rd

yjpt(x−y)φ(x)dxdy

=−

d

X

j=1

Z

Rd

f(y) Z

Rd

yjpt(y−x)φ(x)dxdy

=−

d

X

j=1

f,

Z

Rd

yjpt(· −x)φ(x)dx

,

having used the symmetry ofpt(·).

We are now ready to prove that any mild solution is also a semigroup solution.

Proposition 4.5. Any mild solutionvnof (27)is also a semigroup solution.

Proof. Recall thatF(z)bn(t, x) =b(t, x, z) by (30). Forvnto be a semigroup solution it must be an a.e. bounded function that satisfies (31). First we notice that, since vn is a mild solution, there exists α > β such that vn ∈ CTCα ⊂L([0, T]×Rd) so the second term on the RHS of expression (31) is well-defined. We recall that by Assumption 3, F is bounded and by Assumption 2 hence b is bounded. Moreover by Assumption 5 the initial condition v0 ∈ Cβ+⊂L([0, T]×Rd) so also the first term on the RHS of expression (31) is well-defined.

Now we show that the two terms on the RHS of (28) are equal to the terms on the RHS of (31). We start with the initial condition term, which can be written as

(Ptv0)(x) = Z

Rd

pt(x−y)v0(y)dy,

sinceptis the kernel of the semigroupPt. For the second term we use Lemma 4.4 with f=ub to get

Pt(div[u(t)F(u(t))bn(t,·)]) =Pt(div[u(t)b(t, u(t))])

=−

d

X

j=1

Z t 0

Z

Rd

u(s, y)bj(s, y, u(s, y))∂yjpt(· −y)dyds

and so (31) becomes (28), i.e. the mild solution vn is also a semigroup

solution.

Références

Documents relatifs

— Well posedness in the Gevrey Classes of the Cauchy Problem for a Non Strictly Hyperbolic Equation with Coefficients Depending On Time, Ann.. — Two by two strongly hyperbolic systems

The Carleman estimate proven in Section 3 allows to give observability estimates that yield null controllability results for classes of semilinear heat equations... We have

The Carleman estimate (5) proven in the previous section allows to give observability estimates that yield results of controllability to the trajectories for classes of semilinear

Two Results concerning the Small Viscosity Solution of Linear Scalar Conservation Laws with

More precisely, we prove existence and uniqueness of the solution for this problem in two different settings: the case with degenerated enthalpy function and a fixed vectorial

Continuity of the blow-up profile with respect to initial data and to the blow-up point for a semilinear heat equation.. Zaag d,

Recently, the first and third authors and Nakamura [10] introduced a method based on appropriate Carleman estimates to prove a quantitative uniqueness for second-order

Thanks to our result of invertibility of the Jacobian matrix of the flow, we can also prove (Theorem 2 in §4 ) that the process thus defined is a solution to a linear