McKean SDEs with singular coefficients
Abstract.
The paper investigates existence and uniqueness for a stochastic differential equation (SDE) depending on the law density of the solution, involving a Schwartz distribution. Those equations, known as McKean SDEs, are 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; distributional 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
(1) |
for some given initial condition with density . The terminology McKean refers to the fact that the coefficient of the SDE depends on the law of the solution process itself, while singular reflects 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 the Schwartz distributional nature of the drift, which is encoded in the term .
The problem is -dimensional, in particular the process takes values in , the function is , the term is formally and is a -dimensional Brownian motion, where are two integers. We assume that for some (see below for the definition of Besov spaces ), which means that is a Schwartz distribution and thus the term , as well as its product with , are only formal at this stage. The function is nonlinear.
The term in equation (1) is a special case of a general drift . When is a function, equation (1) was studied by several authors. For example [23] studies existence and uniqueness of the solution under several regularity assumptions on the drift, while [26] requires to be Lipschitz-continuous with respect to the variable , uniformly in time and space, and measurable with respect to time and space. We also mention [2], where the authors obtain existence of the solution when assuming that the drift is a measurable function. For other past contributions see [22].
Different settings of McKean-Vlasov problems have been considered by other authors where the pointwise dependence on the density is replaced by a smoother dependence on the law, typically of Wasserstein type, and the Lipschitz property for the coefficients has been relaxed. From this perspective, the equations are not singular in our sense. For example in [9] the author considers McKean-Vlasov equations with coefficients and which depend on the law of the process in a relatively smooth way, but are Hölder-continuous in time and space. Later on in [15] the authors considered SDEs where both the drift and the diffusion coefficient are of McKean type, with a Wasserstein dependence on the law, and where the drift satisfies a Krylov-Röckner --type dependence. Independently [27] considered in particular SDEs with a McKean drift of the type where is the law of , and is some measurable function and . In [16], the authors study McKean-Vlasov SDEs with drift discontinuous under Wasserstein distance.
In the literature we also find some contributions on (1) with , i.e. when there is no dependence on the law but the drift is a Schwartz distribution. In this case equation (1) becomes an SDE with singular drift. Ordinary SDEs with distributional drift were investigated by several authors, starting from [13, 12, 3, 28] in the one-dimensional case. In the multi-dimensional case it was studied by [11] with being a Schwartz distribution living in a fractional Sobolev space of negative order (up to . Afterwards, [5] extended the study to a smaller negative order (up to ) and formulated the problem as a martingale problem. We also mention [21], 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 [19, 20].
The main analytical tool in the works cited above is an associated singular PDE (either Kolmogorov or Fokker-Planck). In the McKean case, the relevant PDE associated to equation (1) is the nonlinear Fokker-Planck equation
(2) |
where
(3) |
PDEs with similar (ir)regular coefficients were studied in the past, see for example [11, 17] 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 [11], 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 . 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 [25, 24].
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 6.2) and its existence and uniqueness (proved in Theorem 6.5). The second contribution is the study of the singular Fokker-Plank equation (2), in particular we find a unique solution in the sense of Schwartz distributions, see Theorem 3.8 for existence and 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 use the notation to indicate the space of jointly continuous functions with gradient in uniformly continuous in . By a slight abuse of notation we use the same notation for functions which are -valued. When is differentiable, we denote by the matrix given by . When we denote the Hessian matrix of by Hess.
We denote by the space of Schwartz functions on and by the space of Schwartz distributions. For we denote by the Besov space or Hölder-Zygmund space and by its norm, more precisely
where is a partition of unity and denotes the Fourier transform. For more details see for example [1, Section 2.7]. We recall that for one has . If then the space coincides with the classical Hölder space of functions which are -times differentiable and such that the th derivative is -Hölder continuous. For example if the classical -Hölder norm
(4) |
is an equivalent norm in . With an abuse of notation we use to denote (4). For this and for more details see, for example, [29, Chapter 1] or [1, Section 2.7]. Notice that we use the same notation to indicate -valued functions but also or -valued functions. It will be clear from the context which space is needed.
We denote by the space of continuous functions on taking values in , that is . For any given we denote by and the spaces given by
Notice that is an inductive space. We will also use the spaces , recalling that if and only if there exists such that , see Lemma B.2 in Appendix B for a proof of the latter fact.
Similarly, we use the metric space , meaning that if and only if for any we have . Notice that if is continuous and such that then .
Let denote the semigroup generated by on , in particular for all we define , where the kernel is the usual heat kernel
(5) |
It is easy to see that . Moreover we can extend it to by dual pairing (and we denote it with the same notation for simplicity). One has for each and , using the fact that the kernel is symmetric.
Lemma 2.1.
Let be continuous and . The unique (weak) solution of
is given by
(6) |
By weak solution we mean, for every and we have .
Proof.
The fact that (6) is a solution is done by inspection. The uniqueness is a consequence of Fourier transform. ∎
We denote by the usual Gamma function defined as for .
In the whole article the letter or 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 on elements of Besov spaces . These estimates are known as Schauder’s estimates (for a proof we refer to [8, Lemma 2.5], see also [14] for similar results).
Lemma 2.2 (Schauder’s estimates).
Let for some . Then for any there exists a constant such that
(7) |
for all .
Moreover let . For we have
(8) |
Notice that from (8) it readily follows that if for some , then for we have
(9) |
In other words, this means that if then (and in fact it is -Hölder continuous in time). We also recall that Bernstein’s inequalities hold (see [1, Lemma 2.1] and [14, Appendix A.1]), that is for there exists a constant such that
(10) |
for all . 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 and . If then for all we have and
(11) |
The following is an important estimate which allows to define the so called pointwise product between certain distributions and functions, which is based on Bony’s estimates. For details see [4] or [14, Section 2.1]. Let and with and . Then the pointwise product is well-defined as an element of and there exists a constant such that
(12) |
Moreover if and are continuous functions defined on with values in the above Besov spaces, one can easily show that the product is also continuous with values in , and
(13) |
2.3. Assumptions
We now collect the assumptions on the distributional term , the nonlinearity and (see (3)) and on the initial condition that 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 and . In particular .
In the following result we construct a sequence using the heat semigroup and prove certain properties.
Proposition 2.4.
Let as in Assumption 1. Let us define a sequence such that, for any fixed and for all we have
where and is the Gaussian kernel defined in (5).
-
(i)
For each , is globally bounded, together with all its space derivatives.
-
(ii)
For each , is continuous in for all . In particular .
-
(iii)
We have the convergence in .
Proof.
Assumption 2.
Let be Lipschitz and bounded.
Assumption 3.
Let be globally Lipschitz.
We believe that Assumption 3 is unnecessary. Indeed by Assumption 2 one gets that 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 [18, Proposition 3.1] and [26, Theorem 22]. However we assume here to be Lipschitz to improve the readability of the paper.
Assumption 4.
Let .
Assumption 5.
Let be a bounded probability density.
2.4. The singular Martingale Problem
We conclude this section with a short recap of useful results from [21], where the authors consider the Martingale Problem for SDEs of the form
(14) |
where satisfies Assumption 1 (with ) and is a given probability measure. 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 and considering in the SDE in (1).
First of all, let us recall the definition of the operator associated to SDE (14) given in [21]. The operator is defined as
(15) |
where
and . Here and the function is the time-derivative. Note also that is well-defined using (12) and Assumption 1. The Laplacian is intended in the sense of distributions. Notice that the identity functions for any belong to and we have .
Next we give the definition of solution to the martingale problem in [21, Definition 4.3]: a couple is a solution to the martingale problem with distributional drift and initial condition (for shortness, solution of MP with drift and i.c. ) if and only if for every
(16) |
is a local martingale under . The domain is given by
(17) |
where has been defined in (15), and the spaces are defined as where is the closure of compactly supported functions of with respect to the norm of . Finally we recall that if and only if there exists such that , by Remark B.3 part (ii). We say that the martingale problem with drift and i.c. admits uniqueness if, whenever we have two solutions and with , , then the law of under equals the law of under . With this definition at hand, we show in [21, Theorem 4.11] that MP admits existence and uniqueness.
3. Fokker-Planck singular PDE
This section is devoted to the study of the singular Fokker-Planck equation (2), recalled here for ease of reading
After introducing the notions of solution for this PDE (weak and mild, which turns out to be equivalent, see Proposition 3.3), we will show that there exists a unique solution in Theorem 3.8 with Banach’s fixed point theorem.
Below we will need mapping properties of the function when viewed as operator acting on , for some . To this aim, we make a slight abuse of notation and denote by the function when viewed as an operator, that is for we have . We sometimes omit the brackets and write in place of . The result below on is taken from [18], Proposition 3.1 and equation (32).
This mapping property allows us to define weak and mild solutions for the singular Fokker-Planck equation.
Definition 3.2.
Let Assumptions 1, 3 and 4 hold and let .
-
(i)
We say that is a mild solution for the singular Fokker-Planck equation (2) if the integral equation
(18) is satisfied.
-
(ii)
We say that is a weak solution for the singular Fokker-Planck equation (2) if for all and all we have
(19)
Note that the term appearing in both items is well-defined as an element of thanks to (12) and Assumption 1 together with Lemma 3.1.
Proposition 3.3.
Let . The function 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 with . ∎
Let us denote by the solution map for the mild solution of PDE (2), that is for for some we have
Then a mild solution of (2) is a solution of , in other words it is a fixed point of .
We present now an a priori bound for mild solutions, if they exist.
Proposition 3.4.
Proof.
Let
(20) |
for brevity. Using Bernstein’s inequality (10) we get
Then using the definition of from (20), pointwise product property (12) (since ) and Lemma 3.1 we have
(21) |
where we recall that is now a constant that changes from line to line. Now using this, together with Schauder’s estimates (Lemma 2.2 with ) and the fact that , for fixed , one obtains
Now by a generalised Gronwall’s inequality (see Lemma A.1) we have
with and where is the Mittag-Leffler function, see Lemma A.1. Now taking the sup over and using the fact that is increasing we get
This concludes the proof. ∎
We are interested in finding a mild solution of (2) according to Definition 3.2, in the space . Let us denote by and by
(22) |
Then the mild formulation (18) is equivalent to
(23) |
since .
Then a mild solution of (2) is where is a solution of (23), in other words is a fixed point of the map . For any we introduce a family of equivalent norms in given by
Consider then the -ball in of radius , given by
(24) |
Notice that these sets are closed with respect to the topology of , hence they are F-spaces, see [10, Chapter 2.1], with respect to the metric topology of . The -equivalent norm generates the -equivalent metric with respect to the metric of , given by
(25) |
for any . Let and be chosen arbitrarily. The is again an F-space.
In the proofs below we will also use the notation
(26) |
for brevity. In order to show that is a contraction, we first show that it maps balls into balls.
Proposition 3.5.
Proof.
Let , for some to be specified later. Using the definition of , Schauder’s estimate for the semigroup (Lemma 2.2) and the definition of from (26) we have
(28) | ||||
Now we use Bernstein’s inequality (10) to bound
and using again the definition of from (26), the pointwise product property (12) (since ) and Lemma 3.1 we have
(29) |
Now plugging (3) into (28) we get
(30) | ||||
Using the assumption that and choosing we have that (since is a contraction) thus (30) gives
(31) |
where
is positive by Assumption 1 and by . We want to choose and such that , for which it is enough that
(32) | ||||
(33) | ||||
(34) |
provided that the denominator is positive. To do so, we pick large enough so that
(35) |
Then we set
where has been chosen in (35). Then for any , and with this choice of we have indeed that and therefore if then as wanted. ∎
We show below that it is possible to choose large enough such that is a contraction on under , with chosen according to Proposition 3.5.
Lemma 3.6.
Proof.
Let . Using the definition of , of the solution map and of as in (26) we have
(36) |
By Bernstein’s inequality (10), pointwise product property (12), the contraction property of , local Lipschitz property of from Lemma 3.1 and definition of -equivalent metric we get
(37) |
having used in the last line the fact that by choice of and that for any one has
Plugging (3) into (3) and using the Gamma function we get
hence setting
we conclude. ∎
We can now state and prove existence and uniqueness of a mild solution of (2) using the equivalent equation (23).
Proposition 3.7.
Proof.
We show existence and uniqueness of solution of (23) because this is equivalent to existence and uniqueness of a mild solution to (2).
Proof.
Existence. Since by assumption, there exists such that . With such by Proposition 3.7 we know that there exists a (unique) mild solution in .
Uniqueness. Given two solutions there exist such that for . Then choosing we have that for and by uniqueness in from Proposition 3.7 we have that . ∎
Remark 3.9.
Notice that if we suppose that in place of Assumption 4 one gets that a solution exists in .
4. The regularised PDE and its limit
We consider the sequence introduced in Proposition 2.4. When the term is replaced by , with fixed , then we get a smoothed PDE, that is, we get the Fokker-Planck equation
(38) |
where we recall that . For ease of reading, we recall that the mild solution of (38) is given by an element such that
(39) |
Remark 4.1.
At this point we introduce the notation and some useful results on a very similar semilinear PDE studied in [26]. We consider the PDE
(40) |
where is a bounded Borel function. We set
(41) |
Thanks to Assumptions 2 and properties of stated in Proposition 2.4 item (i) we have that the term is uniformly bounded. Below we recall a mild-type solution, introduced in [26], which we call here semigroup solution. We will show that any semigroup solution is also a mild solution in Proposition 4.5.
Definition 4.2.
Notice that this definition is inspired by [26, Definition 6], but we modified it here to include the condition , rather than (the latter as in [26], where moreover the solution is called ‘mild solution’). Indeed integrability of is sufficient for the integrals in the semigroup solution to make sense, because is also bounded and the heat kernel and its derivative are integrable.
The first result we have on (40) is about uniqueness of the semigroup solution in . This result is not included in [26], but we were inspired by proofs therein, in particular by the proof of [26, Lemma 20].
Lemma 4.3.
There exists at most one semigroup solution of (40).
Proof.
First of all we remark that since is the heat kernel then we have two positive constants such that
(43) |
for all , where is a Gaussian probability density.
Let us consider two semigroup solutions of (40). We denote by the semigroup solution map, which is the right-hand side of (4.2). Notice that by Assumption 4, and the function is Lipschitz, uniformly in because is assumed to be Lipschitz in Assumption 3. Using this, together with the bound (43), for fixed , we get
Now, by an application of a fractional Gronwall’s inequality (see Lemma A.1) we conclude that for all , so in particular we have
hence the semigroup solution is unique in . ∎
At this point we want to compare the concept of mild solution and that of semigroup solution. Recall that so in fact PDE (40) is exactly (38). First we state and prove a preparatory lemma, where is vector-valued and will be taken to be for fixed in the following result.
Lemma 4.4.
Let , . Then
(44) |
almost everywhere.
Proof.
We will show that the left-hand side (LHS) and the right-hand side (RHS) are the same object in . Notice that the heat kernel is the same kernel associated to the semigroup , namely if , then with . We now take the Fourier transform in of both sides. The LHS gives
The RHS of (44), on the other hand, gives
Notice that one should be careful that the products appearing above are classical products of an element of (like ) and an element of (like ). ∎
We are now ready to prove that any mild solution is a semigroup solution.
Proposition 4.5.
Any mild solution of (38) is a semigroup solution.
Proof.
Recall that by (41). For to be a semigroup solution it must be an a.e. bounded function that satisfies (4.2). First we notice that, since is a mild solution, there exists such that so the second term on the RHS of expression (4.2) is well-defined. We recall that by Assumption 2, is bounded and by Proposition 2.4 (i) also is bounded hence is also bounded. Moreover by Assumption 4 the initial condition so also the first term on the RHS of expression (4.2) is well-defined.
Now we show that the two terms on the RHS of (39) are equal to the terms on the RHS of (4.2). We start with the initial condition term, which can be written as
since is the kernel of the semigroup . For the second term we use Lemma 4.4 with to get
and so (4.2) becomes (39), i.e. the mild solution is also a semigroup solution. ∎
The next result establishes, in particular, the uniqueness of the solution in and a continuity result with respect to .
Proposition 4.7.
Proof.
Item (i). Let (resp. ) be a solution in to (2) with (resp. ); so there exists such that . We fix . Using Schauder’s estimates and Bernstein’s inequalities, for the difference below we get the bound
(45) |
Now, in order to bound the term inside the integral we use the mapping properties of from Lemma 3.1, the property (12) of the pointwise product, and the fact that and are mild solutions. We get
At this point we use the a priori bound for (resp. for ) found in Proposition 3.4, which depends on and (resp. ) and is increasing with respect to the latter. Thus we get
where is a function increasing in the second variable. Putting this into (4) we get
and by a generalised Gronwall’s inequality (see Lemma A.1) we get
where is again a function increasing in the second variable.
Item (ii). Let be a sequence in . Let us assume that is the unique solution of (2) with by Theorem 3.8. Moreover, by Proposition 3.7, such lives in , where depends only on , hence not on . Let be the unique solution of (2). We apply Item (i) with and to get
(46) |
We have because in , and
because is increasing. Therefore plugging this into (46) we have
where . Thus taking the sup over we get that in if in , which implies the convergence of in because . ∎
5. The regularised SDEs
In this section we consider the regularised version of the McKean SDE introduced in (1), when is replaced by a defined in Proposition 2.4, for fixed . We focus on the SDE
(47) |
where is a given random variable distributed according to . In order to show existence and uniqueness of a solution of (47) and its link to the mild (and semigroup) solution of (38), we make use of Theorems 12 and 13 from [26], as we see below.
Proposition 5.1.
Proof.
We observe that (47) is the special case of equation (1) in [26] when and has a density with respect to the Lebesgue measure. Notice that all assumptions in Theorems 12 and 13 are satisfied. Indeed, the drift is bounded and Lipschitz with respect to because is Lipschitz and bounded by Assumption 2 and is bounded by Proposition 2.4 item (i).
Item (i). We apply the result [26, Theorem 13 point 3]. In fact, the authors forgot to emphasize that the can be chosen to be bounded (contrary to Theorem 13 point 1 where they emphasized it).
Item (ii). We apply the result [26, Theorem 13 point 2].
Item (iii). We apply the result [26, Theorem 12 point 1] to get that is a weak solution of (40). Under [26, Assumption C]111which postulates uniqueness of weak solutions for in the class of measure valued functions, which is true if , see [26, Remark 7]., weak and semigroup solutions are equivalent, see [26, Proposition 16]. ∎
6. Solving the McKean problem
Let Assumptions 1, 2, 3, 4 and 5 be standing assumptions in this section. For ease of reading, we recall the problem at hand, which was illustrated in (1). We want to solve the McKean equation
(48) |
for some given initial condition . The corresponding Fokker-Planck singular equation (already introduced in (2) and recalled here for ease of reading) is
(49) |
where , to which we gave a proper meaning and which we solved in Section 3.
Remark 6.1.
In [23] the authors investigate the propagation of chaos for McKean SDE (48) with smooth coefficients and initial condition, using a system of moderately interacting particles. The corresponding system in our singular framework appears to be
where is a mollifier converging to .
We observe that the above equations can be considered as a -dimensional SDE
with singular drift where
and each . This singular SDE is well-defined using [21] (see also [11]) because since and is Lipschitz and bounded (since both and are Lipschitz and bounded).
We leave the study of this system and its behaviour when to future research.
Definition 6.2.
A solution (in law) of the McKean problem (48) is a triple such that is a probability measure on some measurable space , the function is defined on and belongs to , the couple is a solution to the martingale problem with distributional drift , and is the law density of .
We say that the McKean problem (48) admits uniqueness if, whenever we have two solutions and , then in and the law of under equals the law of under .
Using the tools developed in the previous sections, in Theorem 6.5 we will construct a solution to the McKean problem (48) and show that this solution is unique. We first recall two useful results from [21]. Let us consider a distributional drift that satisfies Assumption 1 with .
The first result concerns convergence in law when the distributional drift is approximated by a sequence of smooth functions . This result is crucial to show existence of the McKean equation.
Proposition 6.3.
(Issoglio-Russo, [21, Theorem 4.16]). Let satisfy Assumption 1. Let be a sequence in converging to in . Let (respectively ) be a solution to the (linear) MP with distributional drift (respectively ). Then the sequence converges in law to . In particular, if is a bounded function (which also belongs to ) and is a (strong) solution of
then converges to in law.
The second result is the fact that the law of the solution to a (linear) martingale problem with distributional drift solves the Fokker-Planck equation in the weak sense. This result is crucial to show uniqueness of the McKean equation.
Proposition 6.4.
We can now state and prove the main result of this paper.
Theorem 6.5.
Proof.
Existence. Let us consider the sequence defined in Proposition 2.4. The corresponding smoothed McKean problem is
(50) |
By Proposition 5.1 part (i) we have a solution of (50) where is bounded and is a (strong) solution of on some fixed probability space , with . By Proposition 5.1 part (iii) we have that is a semigroup solution of (38). On the other hand, we know by Remark 4.1 that a mild solution of the same equation exists. By Proposition 4.5 we know that is a semigroup solution and moreover it is bounded (because it is a mild solution). By uniqueness of semigroup solutions (see Lemma 4.3) we have .
Now we notice that converges to in because of (13), the linearity of the pointwise product, the Lipschitz property of , Lemma 3.1, the convergence by Proposition 2.4 item (iii) and the convergence by Proposition 4.7. By Lemma [21, Lemma 4.14] we know that is also a solution to the MP with distributional drift and initial condition , hence applying Proposition 6.3 we have that in law (as ), and since is the law density of we have that must be the law density of .
Uniqueness. Suppose that we have two solutions of the McKean problem (48), and . By definition we know that is a solution to the (linear) martingale problem with distributional drift . Thus by Proposition 6.4 we have that is a weak solution to the Fokker-Planck equation
which is exactly PDE (49). Item (ii) in Proposition 4.7 guarantees uniqueness of the mild solution of (49) and Proposition 3.3 ensures that weak and mild solutions of the Fokker-Planck equation are equivalent, hence . Note that it is crucial the fact that . This implies that are both solutions of the same (linear) martingale problem with distributional drift , so by uniqueness of the solution of MP (see Section 2.4) we conclude that the law of under equals the law of under . ∎
Appendix A A generalised Gronwall’s inequality
Here we recall a useful generalised Gronwall’s inequality (or fractional Gronwall’s inequality). For a proof see [30, Corollary 2].
Lemma A.1.
Suppose , is a nonnegative function locally integrable on (some ) and nondecreasing on . Let be a nonnegative, nondecreasing continuous function defined on , (constant), and suppose is nonnegative and locally integrable on with
on this interval. Then
where is the Mittag-Leffler function defined by .
Remark A.2.
In [18], the end of the proof of Proposition 4.1 incorrectly uses Gronwall’s lemma. The proper argument should instead cite a generalised Gronwall’s inequality, like the one stated above.
Appendix B Compactness and continuity in inductive spaces
This Appendix is devoted to the proof of a continuity result in inductive spaces. We show in two steps that a function belongs to if and only if it belongs to for some .
The first step is about compactness of sets in inductive spaces .
Lemma B.1.
Let . A set is a compact in if and only if there exists such that and is a compact in .
Proof.
“”. Let be a compact. For any , we know that for some and we pick an arbitrary open neighbourhood in . Thus is an open set of . We have , and since is compact in there exists a finite subcovering . Let . Thus . Next we show that is also a compact in for the chosen . Let be any open covering of in , that is . Each is an open set of thus also of , therefore is also an open covering of , thus there exists a finite covering.
“”. Let be a compact in , for some . The inclusion is obvious. Now let us take an open covering of in , that is , where each is an open set in . Since , then . Finally we notice that since is an open set in , by trace topology we have that is an open set of (because is a closed set of ). Thus we can extract a finite subcovering in , which will be also a finite subcovering of in . ∎
Next we show the continuity result.
Lemma B.2.
Let . Then .
Proof.
The inclusion is obvious.
Next we show the inclusion . Let be continuous. We have to find such that .
Let , which is a compact in since it is the image of the compact via which is continuous. By Lemma B.1 there exists such that is a compact in , in particular, . It remains to show that in when . Since is compact in , there exists a subsequence such that for some , thus . On the other hand, means that in . Thus by uniqueness of the limit we have .
∎
References
- [1] H. Bahouri, J-Y. Chemin, and R. Danchin. Fourier Analysis and Nonlinear Partial Differential Equations. Springer, 2011.
- [2] V. Barbu and M. Röckner. Probabilistic representation for solutions to nonlinear Fokker-Planck equations. SIAM J. Math. Anal., 50(4):4246–4260, 2018.
- [3] R. F. Bass and Z.-Q. Chen. Brownian motion with singular drift. Ann. Probab., 31(2):791–817, 2003.
- [4] J.-M. Bony. Calcul symbolique et propagation des singularites pour les équations aux dérivées partielles non linéaires. Ann. Sci. Ec. Norm. Super., 14:209–246, 1981.
- [5] G. Cannizzaro and K. Chouk. Multidimensional SDEs with singular drift and universal construction of the polymer measure with white noise potential. Ann. Probab., 46(3):1710–1763, 2018.
- [6] R. Carmona and F. Delarue. Probabilistic theory of mean field games with applications I, volume 84 of Probability Theory and Stochastic Modelling. Springer International Publishing, 2018.
- [7] R. Carmona and F. Delarue. Probabilistic theory of mean field games with applications II, volume 84 of Probability Theory and Stochastic Modelling. Springer International Publishing, 2018.
- [8] R. Catellier and K. Chouk. Paracontrolled distributions and the 3-dimensional stochastic quantization equation. Ann. Probab., 46(5):2621–2679, 2018.
- [9] P. E. Chaudru de Raynal. Strong well-posedness of McKean-Vlasov stochastic differential equations with Hölder drift. Stochastic Processes Appl., 130(1):79–107, 2020.
- [10] N. Dunford and J. T. Schwartz. Linear operators. Part I. Wiley Classics Library. John Wiley & Sons Inc., New York, 1988. General theory, With the assistance of William G. Bade and Robert G. Bartle, Reprint of the 1958 original, A Wiley-Interscience Publication.
- [11] F. Flandoli, E. Issoglio, and F. Russo. Multidimensional SDEs with distributional coefficients. T. Am. Math. Soc., 369:1665–1688, 2017.
- [12] F. Flandoli, F. Russo, and J. Wolf. Some SDEs with distributional drift. I. General calculus. Osaka J. Math., 40(2):493–542, 2003.
- [13] F. Flandoli, F. Russo, and J. Wolf. Some SDEs with distributional drift. II. Lyons-Zheng structure, Itô’s formula and semimartingale characterization. Random Oper. Stochastic Equations, 12(2):145–184, 2004.
- [14] M. Gubinelli, P. Imkeller, and N. Perkowski. Paracontrolled distributions and singular PDEs. Forum of Mathematics, Pi, 3:75 pages, 2015.
- [15] X. Huang and F.-Yu. Wang. Distribution dependent SDEs with singular coefficients. Stochastic Processes Appl., 129(11):4747–4770, 2019.
- [16] X. Huang and F.-Yu. Wang. McKean-Vlasov SDEs with drifts discontinuous under Wasserstein distance. Discrete Contin. Dyn. Syst., 41(4):1667–1679, 2021.
- [17] E. Issoglio. Transport equations with fractal noise - existence, uniqueness and regularity of the solution. J. Analysis and its App., 32(1):37–53, 2013.
- [18] E. Issoglio. A non-linear parabolic PDE with a distributional coefficient and its applications to stochastic analysis. J. Differential Equations, 267(10):5976–6003, 2019.
- [19] E. Issoglio and S. Jing. Forward-Backward SDEs with distributional coefficients. Stochastic Processes and their Applications, 130(1):47–78, 2020.
- [20] E. Issoglio and F. Russo. A Feynman–Kac result via Markov BSDEs with generalised drivers. Bernoulli, 26(1):728–766, 2020.
- [21] E. Issoglio and F. Russo. A PDE with drift of negative Besov index and related martingale problem. preprint (in preparation), 2022.
- [22] L. Izydorczyk, N. Oudjane, and F. Russo. McKean Feynman-Kac probabilistic representations of nonlinear partial differential equations. Geometry and Invariance in Stochastic Dynamics. Eds. S. Ugolini et al. Preprint hal-02397045, to appear, 2021.
- [23] B. Jourdain and S. Méléard. Propagation of chaos and fluctuations for a moderate model with smooth initial data. Ann. Inst. H. Poincaré Probab. Statist., 34(6):727–766, 1998.
- [24] A. Le Cavil, N. Oudjane, and F. Russo. Particle system algorithm and chaos propagation related to a non-conservative McKean type stochastic differential equations. Stochastics and Partial Differential Equations: Analysis and Computation., pages 1–37, 2016.
- [25] A. Le Cavil, N. Oudjane, and F. Russo. Probabilistic representation of a class of non-conservative nonlinear partial differential equations. ALEA Lat. Am. J. Probab. Math. Stat, 13(2):1189–1233, 2016.
- [26] J. Lieber, N. Oudjane, and F. Russo. On the well-posedness of a class of McKean Feynman-Kac equations. Markov Process. Relat. Fields, 25(5):821–862, 2019.
- [27] M. Röckner and X. Zhang. Well-posedness of distribution dependent SDEs with singular drifts. Bernoulli, 27(2):1131–1158, 2021.
- [28] F. Russo and G. Trutnau. Some parabolic PDEs whose drift is an irregular random noise in space. Ann. Probab., 35(6):2213–2262, 2007.
- [29] H. Triebel. Bases in Function Spaces, Sampling, Discrepancy, Numerical Integration, volume 11 of EMS Tracts in Mathematics. European Mathematical Society, 2010.
- [30] H. Ye, J. Gao, and Y. Ding. A generalized Gronwall inequality and its application to a fractional differential equation. J. Math. Anal. Appl., 328(2):1075–1081, 2007.