Abstract:
This paper gives a survey of recent investigations on nonlinear Fokker–Planck–Kolmogorov equations of elliptic and parabolic types and contains a number of new results. We discuss in detail the problems of existence and uniqueness of solutions, various estimates of solutions, connections with linear equations, and the convergence of solutions of parabolic equations to stationary solutions.
Bibliography: 116 items.
The classical linear stationary Fokker–Planck–Kolmogorov equation with respect to a probability density ϱ has the form
Δϱ−div(ϱb)=0,
where Δ is the Laplace operator and b is a smooth vector field on Rd, called the drift. A nonlinear equation arises if the drift can depend on the solution ϱ, for example, if in the one-dimensional case b=ϱ. The nonlinearity makes this equation non-trivial even on the real line, where a linear equation is solved in the explicit form in an elementary way. A more general nonlinear equation with a diffusion matrix A=(aij), which can also depend on the solution, has the form
Throughout, the measure ϱtdx is identified with its density ϱt with respect to Lebesgue measure. Such equations are often solved with respect to measures, rather than with respect to functions; the precise definitions are given below. In some problems it is important to indicate separately the dependence of coefficients on the values of the density ϱ(x) at points in the elliptic case (or on ϱt(x) and also on t and the measure ϱt in the parabolic case). A typical example is the drift coefficient b of the form
A great number of papers is devoted to the investigation of nonlinear Fokker–Planck–Kolmogorov equations; this is connected with numerous applications in physics and biology, social and economic areas. Even a simple search in the MathSciNet database of the American Mathematical Society for the keywords ‘Fokker-Planck’ produces thousands of papers; moreover, this does not include a large number of papers of the physical and applied nature. Constructions and discussions of models on the basis of Fokker–Planck–Kolmogorov equations can be found in [62], [63], [65], [77], [78], [99], and [100]. For example, in biological models for a swarm (see [99]) a population density ϱ satisfies the one-dimensional (in x) stationary equation with unit diffusion coefficient and drift coefficient b of the form
In [79] another equation popular in biology was considered:
∂tϱ−Δϱ+∇⋅(ϱ∇(−Δ)−1ϱ)=0,ϱ(0,x)=ϱ0(x).
In socio-economic models (see [65]) and in traffic models (see [112]) one considers linear and nonlinear Fokker–Planck–Kolmogorov equations on the half-line or an interval with diffusion coefficients degenerate on the boundary. Such equations are reduced by a change of coordinates to the form indicated, with coefficients growing strongly at infinity.
In recent years there were intensive studies (see [48], [87], and [109]) of equations of the form
∂tϱt=div[f(ϱt)∇(H′(ϱt)+V+W∗ϱt)],
where
W∗ϱt(x)=∫RdW(x−y)ϱt(y)dy.
In such equations local and non-local nonlinearities are also combined, that is, the equation contains both nonlinear terms depending on the value of ϱt at the point x and terms depending on ϱt by means of the convolution with a kernel ∇W.
Note also the recent papers [3], [7]–[12], which study equations of the form
∂tϱt=Δβ(ϱt)+div(ϱtb(ϱt)∇Φ)
and prove the existence and uniqueness of solutions and their convergence to the stationary solution in L1 as t→+∞, and discuss probabilistic interpretations of solutions. The expression Δβ(ϱt) and a special form of the drift coefficient bring substantial specific features. Probabilistic representations of solutions to linear and nonlinear Fokker–Planck–Kolmogorov equations were also considered in [13] and [30].
The problems of the existence, uniqueness, and convergence of probability solutions (that is, non-negative solutions whose integral over Rd equals one) to the stationary solution for the equation with a non-local nonlinearity of the form
∫RdK(x,y)ϱt(y)dy
were studied in [1], [64], and [93], and some recent results and a survey of literature can be found in [27]–[29], [59], [96], and [115]. In [28], [29], [38], and [115] a nonlinear equation is written as a perturbation of the linear equation obtained after substituting the stationary solution into the coefficients. Then, assuming some smallness of the nonlinear part, one proves convergence to the stationary solution as t→+∞. This approach enables one to justify the convergence of solutions of the parabolic equation to solutions of the stationary equation in the case of irregular and strongly growing coefficients. In [2] the equation of the form
(∂tf+v∇x)f(t,x,v)=(∫Rdf(t,x,v)dv)βLFPf(t,x,v)
was studied, where
LFP=∇v⋅(∇v+v).
The book [4] develops functional methods of investigation of nonlinear elliptic and parabolic equations, in particular, by employing nonlinear semigroups. A Trotter-type formula for nonlinear Fokker–Planck–Kolmogorov equations was obtained in [5]. In [81] solutions to some nonlinear equations were represented as gradient flows for suitable functionals. Considerable attention has been paid to Vlasov-type equations (see [54], [37], [83], [84], and [114]).
In most papers on nonlinear Fokker–Planck–Kolmogorov equations a very important role is played by specific features of the equations under consideration. It is hardly possible to create a general theory covering all types of equations, however, there are papers of a fairly general nature. For example, in [110] and [23] the existence of solutions is established with the aid of Schauder’s fixed point theorem under broad assumptions. The books [76] and [80] study equations of a rather general form, including nonlinear Fokker–Planck–Kolmogorov equations.
Linear and nonlinear Fokker–Planck–Kolmogorov equation have interesting connections with other directions of research, including mean field games (see [41], [40], [42], and [106]) and Monge–Kantorovich optimal transportation problems (see [16] and [20]). For stochastic versions of such equations, see, for example, [6], [51], and [66]. Investigation of nonlinear Fokker–Planck–Kolmogorov equations on infinite-dimensional spaces was initiated in [17] and continued in [88] and [53], but so far these problems remain poorly studied. At the end of § 6 we mention an interesting special class of infinite-dimensional Fokker–Planck–Kolmogorov equations on spaces of measures, which arose in connection with finite-dimensional stochastic Itô equations whose coefficients can depend on distributions of solutions.
We survey investigations of the last decade on nonlinear Fokker–Planck–Kolmogorov equations of elliptic and parabolic types and present some new results. We discuss problems of the existence and uniqueness of solutions, various estimates for solutions, and the convergence of solutions of parabolic equations to stationary solutions. In § 3 we briefly recall some basic facts about linear equations, since the nonlinear equations under consideration look as linear equations with coefficients depending on solutions. In addition, in § 4 we present some estimates for various distances between solutions to linear equations, obtained by different methods. These estimates play an important role in the study of nonlinear equations. In § 5 we begin the discussion of nonlinear equations from the elliptic case. Here we present results on the existence of solutions obtained with the aid of two types of fixed point theorems: the Schauder–Tychonoff theorem and its multivalued version, the Kakutani–Ky Fan theorem, and also with the aid of the contracting mapping theorem, which employs the estimates from § 4 and also gives conditions for the uniqueness of solutions. The last section, § 6, is concerned with nonlinear parabolic equations and contains results on the existence and uniqueness of solutions and on the convergence of solutions of parabolic equations to solutions of stationary equations. In all sections we give illustrative examples. The vast bibliography in this survey includes only the works most close to the problems we consider and covers a small part of all papers in this very intensively developing area. Of course, under reasonable restrictions on the size of the bibliography it was not possible to present all authors working in this field. A simple search on the basis of the sources included in the bibliography and the papers citing these works increases many times the number of papers and the number of authors on this subject.
2. Notation and terminology
The standard inner product on Rn is denoted by ⟨x,y⟩ and the Euclidean norm by |x|. The operator norm of an operator A on Rn is denoted by ‖. For a Lebesgue measurable set E in \mathbb{R}^n the symbol L^p(E) denotes the standard Banach space of equivalence classes of functions whose absolute value is integrable over E to power p\geqslant 1.
For an open set U\subset \mathbb{R}^n, let W^{p,k}(U), where p\geqslant 1 and k\in\mathbb{N}, be the Sobolev space of functions f in L^p(U) having generalized derivatives \partial_{x_{i_1}}\cdots \partial_{x_{i_m}}f\in L^p(U), m=1,\dots,k. This space is Banach with the norm
Let W^{p,k}_{\rm loc} denote the set of all functions on \mathbb{R}^n whose restriction to every ball U belongs to W^{p,k}(U).
The symbol C_0^\infty(U) denotes the set of all infinitely differentiable functions with compact support in U.
A Borel measure on a topological space X is a bounded (possibly, signed) measure on the \sigma-algebra \mathcal{B}(X) of Borel sets. Such a measure \mu is decomposed into the difference \mu=\mu^+-\mu^- of its positive and negative parts, the sum |\mu|=\mu^+ +\mu^- of which is called the total variation of the measure \mu. The variation, or variation norm, of the measure \mu is defined by the equality \|\mu\|=|\mu|(X). Below we mainly deal with Borel measures on \mathbb{R}^d and \mathbb{R}^d\times [0,T].
The set of all Borel probability measures on \mathbb{R}^d, that is, measures \mu\geqslant 0 such that \mu(\mathbb{R}^d)=1, is denoted by \mathcal{P}(\mathbb{R}^d). We also use the larger set \mathcal{SP}(\mathbb{R}^d) of subprobability Borel measures \mu\geqslant 0 such that \mu(\mathbb{R}^d)\leqslant 1.
The space of all bounded Borel measures on \mathbb{R}^d is equipped with the Kantorovich–Rubinshtein norm
which makes the sets \mathcal{SP}(\mathbb{R}^d) and \mathcal{P}(\mathbb{R}^d) (but not the whole space of signed measures) complete separable metric spaces (see [15]). The topology on \mathcal{SP}(\mathbb{R}^d) and \mathcal{P}(\mathbb{R}^d) generated by this metric is the weak topology, for which the convergence of a sequence of measures \mu_j to a measure \mu is the convergence of the integrals
In the case where \mu_j,\mu\in \mathcal{P}(\mathbb{R}^d) this convergence is equivalent to the convergence of the integrals of functions from C_0^\infty(\mathbb{R}^d).
Let \mathcal{P}_k(\mathbb{R}^d) denote the subspace of \mathcal{P}(\mathbb{R}^d) consisting of the measures with finite kth moment (that is, such that the function |x|^k is integrable). This space is equipped with the Kantorovich k-metric W_k defined by the formula
where the infimum is taken over all measures \sigma\in \mathcal{P}(\mathbb{R}^{d}\times \mathbb{R}^{d}) with projections \mu and \nu onto the factors.
3. Linear Fokker–Planck–Kolmogorov equations
The linear elliptic (stationary) Fokker–Planck–Kolmogorov equation on the Euclidean space \mathbb{R}^d arises if on \mathbb{R}^d we are given a mapping
called the drift coefficient or drift, for which the elements a^{ij} and b^i are Borel functions. These objects generate a second-order elliptic operator
with the operator L_{A,b} if the functions a^{ij} and b^i are integrable on compact sets with respect to the measure |\mu| and the following identity is fulfilled:
In the theory of partial differential equations such equations are called double divergence form equations. In the sense of distributions identity (3.2) can be written as
where summation over repeated indices is meant and a^{ij}\mu and b^{i}\mu are distributions, being locally finite measures. However, one should bear in mind that in the case of non-smooth coefficients the operator L_{A,b}^* is not defined at all distributions.
With the aid of this identity the equation can also be defined for measures on domains or on Riemannian manifolds, moreover, one can admit measures finite on compact sets, but for the purposes of this survey we confine ourselves to globally bounded measures. Detailed surveys of the properties of solutions are given in [22] and [23] (see also the recent survey [31]), we only note here that in case of nonnegative solutions the measure (\det A)^{1/d}\mu possesses a density with respect to Lebesgue measure, so that for a non-degenerate diffusion matrix the solution itself has a density \varrho. If the matrix A is non-degenerate, its elements belong to the local Sobolev class W^{p,1}_{\rm loc} for some p>d, and the function |b^i|^p is locally integrable with respect to Lebesgue measure or the measure |\mu|, then \mu has a continuous density in the same Sobolev class, and in the case of a positive measure \mu this density has no zeros. For solutions with densities in Sobolev classes and coefficients a^{ij} from Sobolev classes (3.1) reduces to a divergence-form equation with respect to the density:
In this case a solution can fail to have second Sobolev derivatives, but if they exist and the functions a^{ij} have second Sobolev derivatives, while the functions b^i have first Sobolev derivatives, then this equation becomes a classical elliptic equation. However, not every classical equation can be obtained from an equation of the form (3.1).
One can also consider Fokker–Planck–Kolmogorov equations with a potential term c, which is a Borel function, locally integrable with respect to the solution \mu, which now satisfies the equation
understood similarly to (3.2) for the operator L_{A,b,c}f=L_{A,b}f+c f.
Various conditions are known for the existence and uniqueness of probability solutions. Most convenient ones in applications employ Lyapunov functions.
Theorem 3.1. Let V\in C^2(\mathbb{R}^d) be a quasicompact function, so that \mathbb{R}^d is the union of the compact sets \{V\leqslant c_k\} for some increasing sequence of numbers c_k, and
\begin{equation*}
L_{A,b}V(x)\leqslant -1 \quad \textit{if}\ \ |x|\geqslant R
\end{equation*}
\notag
for some R>0. Suppose that A is continuous and either b is also continuous, or b is locally bounded and \det A>0. Then there exists a probability solution to equation (3.1).
It is unclear whether there is always a probability solution in the case where a Lyapunov function exists if A and A^{-1} are locally bounded and the drift b is locally integrable to power p>d with respect to Lebesgue measure.
Let us also mention the following result on the uniqueness of solutions.
Theorem 3.2. If a^{ij}\in W^{p,1}_{\rm loc} for some p>d, the continuous version of the mapping A (which exists by the Sobolev embedding theorem) is non-degenerate and b^i\in L^{p}_{\rm loc}, then a probability solution \mu is unique, provided that the functions
belong to L^1(\mu). In particular, this is true if the coefficients have an at most linear growth.
In terms of Lyapunov functions the following condition holds: there are no different probability solutions if there exists a function V\in C^2(\mathbb{R}^d) such that V(x)\to +\infty as |x|\to\infty and the estimate L_{A,b}V\leqslant C+CV holds for some C>0.
Without additional conditions on the drift b even when it is smooth the equation with the unit diffusion matrix can have different probability solutions for d>1.
The linear parabolic Fokker–Planck–Kolmogorov equation is also associated with a pair of mappings, the diffusion coefficient A(x,t)=(a^{ij}(x,t))_{i,j\leqslant d} and the drift coefficient b(x,t)=(b^i(x,t))_{i\leqslant d}, which are now defined on \mathbb{R}^d\times [0,T] for some T>0 and take values in the space of symmetric non-negative definite operators and in \mathbb{R}^d, respectively. As in the elliptic case, we assume that they are Borel measurable.
Solutions to parabolic equations will be considered in the class of (possibly, signed) measures of the form \mu(dx\, dt)=\mu_t(dx)\, dt for some family of Borel measures on \mathbb{R}^d depending on t Borel measurably, that is, the functions t\mapsto \mu_t(B) must be Borel measurable for Borel sets B, the function t\mapsto \|\mu_t\| must be Lebesgue integrable on [0,T], and the expression indicated means that for any bounded Borel function f on \mathbb{R}^d\times [0,T] we have the equality
if the functions a^{ij} and b^i are integrable with respect to |\mu| on compact sets in \mathbb{R}^d\times (0,T) and the following equality holds for every function \varphi\in C_0^\infty (\mathbb{R}^d\times (0,T)):
This is actually the same equation as in the elliptic case, but with an elliptic operator with two arguments which is completely degenerate in t. An additional difference is the form of the measure. Equation (3.4) can also be written in the form
without the index t for the measure, when it is meant that \mu has the form \mu(dx\, dt)=\mu_t(dx)\, dt indicated above. This agrees with the convention that in the case where a density \varrho(x,t) exists the equation is written as \partial_t\varrho=L_{A,b}^* \varrho. The measure \mu is often identified with the family of measures (\mu_t)_{t\in (0,T)}.
The Cauchy problem for the Fokker–Planck–Kolmogorov equation with the operator L_{A,b} is posed by adding an initial condition, that is, an initial measure \nu. Here, however, several different (inequivalent in the general case) technical formulations are possible. We define a solution to the Cauchy problem
with initial measure \nu as a family of Borel measures (\mu_t)_{t\in [0,T]} on the space \mathbb{R}^d, Borel measurably depending on t, such that \mu_0=\nu, the measure |\mu_t|\, dt on \mathbb{R}^d\times [0,T] is finite and for every function \varphi\in C_0^\infty(\mathbb{R}^d) the equality
The measure \mu(dx\, dt)=\mu_t(dx)\, dt on \mathbb{R}^d\times [0,T] is also called a solution. A probability solution is a solution in which almost all the \mu_t are probability measures.
Note that any solution to the Cauchy problem also satisfies the Fokker–Planck–Kolmogorov equation (3.4). This is first verified for functions
with smooth functions \varphi_1 and \varphi_2 with compact support, by multiplying (3.7) by the function \varphi_2'(t), integrating over [0,T], and then integrating by parts on the right.
There is a Borel set E\subset [0,T] of Lebesgue measure zero such that (3.7) is true for all t\not\in E and all \varphi\in C_0^\infty (\mathbb{R}). Indeed, there exists a countable system of functions \varphi_n\in C_0^\infty (\mathbb{R}^d) with the following property: for every function \varphi\in C_0^\infty (\mathbb{R}^d) there is a subsequence \{\varphi_{n_j}\} such that the functions \varphi_{n_j} have support in some common ball containing the support of \varphi and converge uniformly to \varphi so that their first and second derivatives also converge uniformly. The set E of points t for which (3.7) is violated for some function \varphi=\varphi_n is Borel and has measure zero. Let t\not\in E and \varphi\in C_0^\infty (\mathbb{R}^d). Let us take the corresponding subsequence \{\varphi_{n_j}\} with the properties indicated above. Then the functions L_{A,b}\varphi_{n_j} converge to L_{A,b}\varphi in L^1 with respect to the measure |\mu_t|\, dt, and the integrals of the functions \varphi_{n_j} against the measures \mu_t and \nu converge to the integrals of \varphi. Hence equality (3.7), valid for \varphi_n, remains true for \varphi.
In this connection the question arises: why do not we redefine the solution at the points of the set E of measure zero in order to obtain the equality at all points? As the next proposition shows, this is in principle possible, but an important nuance appears: if all measures \mu_t of the solution we wish to redefine are probability measures (as desired in many problems), then this property can be lost after the redefinition. There are examples showing that even in the case of smooth coefficients a probability solution (\mu_t)_{t\geqslant 0} cannot be redefined in such a way that all measures \mu_t become probability measures and (3.7) is true for all t.
Proposition 3.3. Suppose that we are given a family of bounded Borel measures \mu_t on \mathbb{R}^d, t\in [0,T], Borel measurably depending on t, such that
Then there exist bounded measures \widetilde{\mu}_t Borel measurably depending on t, for which \sup_t\|\widetilde{\mu}_t\|\leqslant M, \widetilde{\mu}_t=\mu_t for almost all t, and the above equality is fulfilled for all t\in [0,T].
Proof. Let us take the Borel set E of Lebesgue measure zero indicated above and for t\in E redefine the measure \mu_t as follows. Fix the open ball B_k of radius k\in\mathbb{N} with centre at the origin and on the space of infinitely differentiable functions with support in B_k consider the linear functional
By Riesz’s theorem there exists a Borel measure \sigma_{t,k} on B_k for which the functional L_{t,k} is the integral against the measure \sigma_{t,k}; moreover, \|\sigma_{t,k}\|\leqslant M. The restriction of the measure \sigma_{t,k+1} to B_k coincides with \sigma_{t,k+1} since both measures assign equal integrals to smooth functions with support in B_k. Therefore, we have a measure \widetilde{\mu}_t on \mathbb{R}^d with restrictions \sigma_{t,k} to B_k and variation at most M. For points in [0,T]\setminus E we retain the original measures \mu_t. The family obtained satisfies our integral identity for all t. This yields its Borel measurability in t, since the right-hand side of the integral identity is continuous in t. \Box
In the case of subprobability solutions there is a useful similar interpretation of solutions as elements of the space L^0([0,T],\mathcal{SP}(\mathbb{R}^d)) of equivalence classes of Lebesgue measurable mappings from [0,T] to the space \mathcal{SP}(\mathbb{R}^d) of subprobability Borel measures on \mathbb{R}^d equipped with the Kantorovich–Rubinshtein metric d_{\rm KR}, making it a complete separable metric space. Thus, also L^0([0,T],\mathcal{SP}(\mathbb{R}^d)) becomes a complete separable metric space with the metric of convergence in measure
where some representatives of equivalence classes are used (the integral does not depend on their choice).
For the existence and uniqueness of solutions to the Cauchy problem broad sufficient condition are also known, expressed in terms of estimates for coefficients and their global integrability with respect to solutions or in terms of Lyapunov functions (see [23], Chaps. 6 and 9). Only in the one-dimensional case, in the recent paper [21] the uniqueness of probability solutions was established without such conditions (but for drifts independent of time). In addition, there are results about the properties of solution densities (see [23], Chaps. 7 and 8). In particular, the very existence of a density is ensured by the condition \det A(x,t)>0, as in the elliptic case.
Parabolic equations are solvable under much broader conditions than elliptic ones. For example, for A={\rm I} and b=0 the elliptic equation has no non-zero solutions, while the parabolic equation has the standard Gaussian solution. There is an inverse implication: under some restrictions on A and b the existence of a probability solution to the elliptic equation L_{A,b}^*\mu=0 implies the existence of a subprobability (but not always a probability) solution to the Cauchy problem with an arbitrary initial probability distribution.
4. Estimates of distances between solutions to linear equations
In the study of nonlinear Fokker–Planck–Kolmogorov equations it is useful to have estimates for various distances between solutions to linear elliptic and parabolic equations with distinct coefficients. This subject deserves a separate survey; here we only present several results from [25] and [26] used below. First we give a number of estimates from [26] (see also [18] and [19]) for elliptic equations. Suppose that two measures
are probability solutions to the equations L_{A_{\mu}, b_{\mu}}^{*}\mu=0 and L_{A_{\sigma},b_{\sigma}}^{*}\sigma=0 the coefficients of which satisfy the following conditions: the elements of the diffusion matrices belong to the Sobolev class W^{p, 1}_{\rm loc} for some p>d, the continuous versions of these matrices are positive definite, and the components of the drifts belong to L^p_{\rm loc}.
We denote by W^{p,1}(\mu) the weighted Sobolev class obtained by the completion of the space C_0^\infty(\mathbb{R}^d) with respect to the Sobolev norm \|f\|_{p,1,\mu} which differs from the standard norm by using the measure \mu in place of Lebesgue measure. As the density of \mu is continuous and positive, functions in this class do not differ locally from functions in W^{p,1}_{\rm loc}. Hence W^{p,1}(\mu) consists of functions of class W^{p,1}_{\rm loc} with finite norm \|\cdot\|_{p,1,\mu}.
Theorem 4.1. Suppose that |A_{\mu}^{-1/2}\Phi|\in L^2(\sigma) and at least one of the following conditions is fulfilled:
(i) the functions (1+|x|)^{-2}a^{ij}_{\mu} and (1+|x|)^{-1}|b_{\mu}| belong to L^1(\mu);
(ii) there exists a function V\in C^2(\mathbb{R}^d) such that \lim_{|x|\to\infty}V(x)=+\infty, V\geqslant 0,
If A_{\mu}\geqslant \alpha\cdot {\rm I}, then it follows from this estimate that \sqrt{v}\in W^{2,1}(\mu).
Given a number C_{\rm S} and a Borel measurable matrix-valued mapping A, we say that a probability measure \mu satisfies the logarithmic Sobolev inequality with constant C_{\rm S} and matrix A if
for every function f\in C_0^\infty(\mathbb{R}^d). Under our assumptions this inequality extends to all functions f\in W^{2,1}(\mu) if A is bounded.
Corollary 4.2. If, in addition to the hypotheses of Theorem 4.1, it is known that the measure \mu satisfies the logarithmic Sobolev inequality with constant C_{\rm S} and matrix A_{\mu}, then the following assertions are true.
is fulfilled. Then the measure \mu satisfies the logarithmic Sobolev inequality with constant 2/\kappa (see [23], Theorem 5.6.36) and the estimates in (i)–(iii) above hold with this constant in place of C_{\rm S}. Moreover, the integral in the right-hand side of these inequalities equals \|b_{\mu}-b_{\sigma}\|_{L^2(\sigma)}^2 in this case.
Given a number C_{\rm P} and a Borel measurable matrix-valued mapping A, we say that a probability measure \mu satisfies Poincaré’s inequality with constant C_{\rm P} and matrix A if
Corollary 4.4. If, in addition to the hypotheses of Theorem 4.1, it is known that the measure \mu satisfies Poincaré’s inequality with constant C_{\rm P} and matrix A_{\mu}, then the following inequalities are valid:
Another method of obtaining estimates for the total variation distance between solutions consists in using solutions to Poisson’s equation L_{\mu}u=\psi, where the integral of \psi against the measure \mu vanishes. Let us give an example of a result proved in this way.
Theorem 4.5. Suppose that there exist a positive function V\in C^2(\mathbb{R}^d) with the property \lim_{|x|\to \infty}V(x)=+\infty, a positive number \gamma, and a ball Q of radius R with centre at the origin such that
where C depends on d, \gamma, R, \|a^{ij}\|_{W^{p, 1}(Q_1)}, \|b^i\|_{L^p(Q_1)}, \sup_{x\in Q_1}\|A(x)^{-1}\|, and also on the minimum of the function V on Q_1 and the maxima of the function V and of the absolute values of its first and second derivatives on Q_1, where Q_1 is the ball of radius R+1 with centre at the origin.
Suppose that |b_{\mu}(x)|+|b_{\sigma}(x)|\leqslant C_0(1+|x|)^m for some numbers m\geqslant 1 and C_0>0. If the measure \sigma has a finite moment of order 2m+2, then the hypotheses of Theorem 4.5 are fulfilled and the estimate
In [60] an estimate for the quadratic Kantorovich distance W_2 between stationary distributions was obtained under the assumption that the coefficients of the first equation belong to the Sobolev class with weight \varrho_\mu, the measures \mu= \varrho_\mu\, dx and \nu=\varrho_\nu\, dx are mutually absolutely continuous, and solutions of the Cauchy problem for the first operator converge exponentially rapidly to the stationary distribution \varrho_\mu in the metric W_2.
In [25] an estimate for distances between solutions to parabolic Fokker–Planck–Kolmogorov equations was obtained. Consider a time-dependent elliptic operator
where A(x,t)=(a^{ij}(x,t))_{i,j\leqslant d} is a positive definite symmetric matrix with Borel elements and b(x,t)=(b^i(x,t))_{i=1}^d \colon \mathbb{R}^d\times [0,T]\to \mathbb{R}^d is a Borel mapping; moreover, b is locally bounded, that is, for every ball U\subset \mathbb{R}^d there is a number B=B(U)\geqslant 0 such that
where \nu\in \mathcal{P}(\mathbb{R}^d). As above, we consider solutions of the form \mu(dx\,dt)=\mu_t(dx)\,dt on \mathbb{R}^d\times [0,T] given by families of probability measures (\mu_t)_{t\in[0,T]} on \mathbb{R}^d, Borel measurably depending on t.
Suppose now that there are two solutions \mu=(\mu_t)_{t\in [0, T]} and \sigma=(\sigma_t)_{t\in [0, T]} to the Cauchy problem (4.2) with coefficients A_{\mu}, b_{\mu} and A_{\sigma}, b_{\sigma}, respectively, and a common initial condition \nu. So far we are speaking of linear equations, hence the indices \mu and \sigma do not mean any dependence of coefficients on solutions, but only label equations.
The distance between \mu_t and \sigma_t can be estimated in terms of the L^2(\sigma)-norm of the vector field A_\mu^{-1/2}\Phi. In the case of equal diffusion matrices \Phi=b_\sigma-b_\mu.
Recall that for two measures \mu_1 and \mu_2 in \mathcal{P}(\mathbb{R}^d) such that \mu_1=w\cdot \mu_2 the entropy H(\mu_1\,|\,\mu_2) is defined by the formula
\begin{equation*}
H(\mu_1\,|\,\mu_2)=\int w\log w \, d\mu_2,
\end{equation*}
\notag
provided that w\log w \in L^1(\mu_2). If \mu_1 and \mu_2 are given by positive densities \varrho_1 and \varrho_2 such that \varrho_1\log (\varrho_1/\varrho_2)\in L^1(\mathbb{R}^d), then H(\mu_1\,|\,\mu_2) is the integral of \varrho_1\log(\varrho_1/\varrho_2).
Theorem 4.7. Suppose that |A_{\mu}^{-1/2}\Phi|\in L^2(\mathbb{R}^d\times[0,T],\sigma) and at least one of the following two conditions is fulfilled:
Finally, for \varphi=1 these estimates are fulfilled with 1 in place of 1+\log \alpha(t).
Estimates of distances between solutions to Fokker–Planck–Kolmogorov equations with dissipative drift coefficients were obtained in [89] with the aid of expressions generalizing the classical Kantorovich metric, and estimates of distances between solutions to equations with partially degenerate diffusion matrices, when degeneracy occurs only in part of the variables, were obtained in [91]. Note that such equations arise in the study of the system of stochastic Langevin equations.
There are many papers on estimates of the rate of convergence of solutions of linear parabolic Fokker–Planck–Kolmogorov equations to stationary solutions as t\to+\infty (see, for example, [35], [58], [59], [75], and also the subsequent works citing these papers).
5. Nonlinear elliptic equations
Nonlinear elliptic Fokker–Planck–Kolmogorov equations differ from linear ones by a dependence of coefficients on solutions. A typical example of such a dependence is supplied by expressions of the form
The interest in nonlinear elliptic Fokker–Plank–Kolmogorov equations is explained in the first place by the fact that solutions of nonlinear parabolic Fokker–Plank– Kolmogorov equations approach to solutions of such equations as t\to+\infty. The key role in obtaining sufficient conditions for the existence and uniqueness of solutions is played by estimates of solutions of linear equations with the aid of the method of Lyapunov functions, by estimates of distances between solutions of two different linear equations, and by sufficient conditions for the existence of continuous solution densities with respect to Lebesgue measure. Let us proceed to precise formulations.
The set \mathcal{P}_{\alpha,V}(\mathbb{R}^d) is non-empty for sufficiently large \alpha and is convex, metrizable, and compact in the weak topology. The set of all probability measures with respect to which the function V is integrable is denoted by \mathcal{P}_V(\mathbb{R}^d). We say that a sequence of measures \mu_n\in \mathcal{P}_V(\mathbb{R}^d) converges V-weakly to a measure \mu\in\mathcal{P}_V(\mathbb{R}^d) if
for all continuous functions f such that \lim_{|x|\to\infty} f(x)/V(x)=0.
Let us verify that weak convergence of a sequence of measures \mu_n\in \mathcal{P}_{\alpha,V}(\mathbb{R}^d) to a measure \mu\in\mathcal{P}_{\alpha,V}(\mathbb{R}^d) yields V-weak convergence of \mu_n to \mu. Let f be a continuous function, and let \lim_{|x|\to\infty} f(x)/V(x)=0. Set
follows since the continuous function f-f_N vanishes outside some ball.
Suppose that for every measure \mu\in\mathcal{P}_V(\mathbb{R}^d) and all i,j\leqslant d we are given Borel functions a^{ij}(\,\cdot\,,\mu) and b^i(\,\cdot\,,\mu) such that the matrix A(x,\mu)=(a^{ij}(x,\mu))_{1\leqslant i,j\leqslant d} is symmetric and non-negative definite. Set
if \mu is a solution to the linear equation with operator L_{\mu}.
Theorem 5.1. Suppose that the following conditions are fulfilled:
(i) for every ball B and every \alpha>0 the functions a^{ij}(\,\cdot\,,\mu) and b^i(\,\cdot\,,\mu) are bounded on B uniformly in \mu\in \mathcal{P}_{\alpha, V}(\mathbb{R}^d) and are equicontinuous on B uniformly in \mu\in \mathcal{P}_{\alpha, V}(\mathbb{R}^d);
(ii) if measures \mu_n\in \mathcal{P}_V(\mathbb{R}^d) converge V-weakly to a measure \mu\in\mathcal{P}_V(\mathbb{R}^d), then for all x we have
for all x\in\mathbb{R}^d and \mu\in\mathcal{P}_{V}(\mathbb{R}^d).
Then there exists a solution \mu\in\mathcal{P}_V(\mathbb{R}^d) to the equation L_{\mu}^{*}\mu=0.
Proof. Fix \alpha \geqslant C/\Lambda such that the metrizable compact set \mathcal{P}_{\alpha,V}(\mathbb{R}^d) is non-empty. According to [23], Corollary 2.4.4, for every measure \sigma\in\mathcal{P}_{\alpha, V}(\mathbb{R}^d) there exists a solution \mu to the linear equation L_{\sigma}^{*}\mu=0. Since
is true. Denote by \Phi(\sigma) the set of all solutions \mu\in\mathcal{P}_{\alpha, V}(\mathbb{R}^d) to the linear Fokker– Planck– Kolmogorov equation L_{\sigma}^{*}\mu=0. This is a non-empty convex subset of the convex metrizable compact set \mathcal{P}_{\alpha,V}(\mathbb{R}^d). Let us verify that the graph of the multivalued mapping \Phi is closed. Let the measures \sigma_n\in \mathcal{P}_{\alpha, V}(\mathbb{R}^d) converge weakly to a measure \sigma\in\mathcal{P}_{\alpha,V}(\mathbb{R}^d) and the measures \mu_n\in \Phi(\sigma_n) converge weakly to a measure \mu\in \mathcal{P}_{\alpha, V}(\mathbb{R}^d). Then the measures \sigma_n converge V-weakly to \sigma. Let us verify that L_{\sigma}^{*}\mu=0. Let \varphi\in C_0^{\infty}(\mathbb{R}^d). We have the equality
Let B be a ball containing the support of \varphi. The sequences a^{ij}(\,\cdot\,,\sigma_n) and b^i(\,\cdot\,,\sigma_n) converge uniformly on B to the continuous functions a^{ij}(\,\cdot\,,\sigma) and b^i(\,\cdot\,, \sigma), respectively. We have
where the first term tends to zero as n\to\infty by the continuity of the coefficients and the weak convergence of \mu_n to \mu. The second term tends to zero as n\to\infty due to the uniform convergence on B of the coefficients of the operators L_{\sigma_n} to the coefficients of the operator L_{\sigma}. Letting n\to\infty, we obtain
Thus, the graph of the mapping \Phi is closed, hence we can apply the Kakutani–Ky Fan theorem (see [34], Theorem 1.2.12). This theorem, which is a multivalued version of the Schauder–Tychonoff theorem, asserts that if a multivalued mapping \Phi from a convex compact set K in a locally convex space to the set of non-empty convex compact subsets of K has a closed graph, then there exists a point k such that k\in \Phi(k). Thus, there exists \mu\in\Phi(\mu). \Box
Note that the factor 1-\delta is needed in condition (iii) to make the term (1-\delta)C vanishing for \delta=1.
Let us give a typical example, where the hypotheses of Theorem 5.1 are fulfilled.
Example 5.2. Suppose that the functions q^{ij} are continuous and bounded on \mathbb{R}^d\times \mathbb{R}^d and the matrices Q(x,y)=\bigl(q^{ij}(x,y)\bigr)_{i,j\leqslant d} are symmetric and non-negative definite. Suppose also that a vector field \beta=(\beta^i)_{i\leqslant d} is continuous on \mathbb{R}^d, a vector field K=(K^i)_{i\leqslant d} is continuous on \mathbb{R}^d\times\mathbb{R}^d, and there exist positive numbers C_1, C_2, C_3, and C_4 such that C_4<C_2 and for all x and y the estimates
that is, condition (iii) is fulfilled for \delta=0. Clearly, (ii) is also fulfilled. Let us verify condition (i). Let B be a ball in \mathbb{R}^d, and let
Let us verify the equicontinuity of the coefficients. It suffices to show that for every continuous function f on \mathbb{R}^d\times\mathbb{R}^d satisfying the condition \sup_{x\in B}|f(x,y)|\leqslant C(B)F(y) on every ball B, where F\geqslant 0 and \lim_{|x|\to\infty} F(x)/V(x)=0, the functions
It remains to observe that the function f is uniformly continuous on the compact set B\times\{y\colon V(y)\leqslant R\}. Thus, all conditions in Theorem 5.1 are fulfilled, and there exists a solution \mu\in\mathcal{P}_V(\mathbb{R}^d) to the nonlinear equation L_{\mu}^{*}\mu=0.
Let us consider an example illustrating condition (iii) in Theorem 5.1.
Example 5.3. Suppose that d=1 and \varepsilon,\gamma\in\mathbb{R}; let
Our equation has the form (b(x,\mu)\mu)'=0, so that b(x,\mu)\mu=\operatorname{const}. Since |x|\in L^1(\mu), we have b\in L^1(\mu) and b(x,\mu)\mu=0. Integrating the equality b(x,\mu)\mu=0 we obtain
The unique solution for \varepsilon\ne 1 is \delta_{\gamma/(1-\varepsilon)}.
We show that, whenever |\varepsilon|<1, for the function V(x)=|x|^2/2 one can find numbers \delta\in [0, 1], C>0, and \Lambda>0 for which condition (iii) in Theorem 5.1 is fulfilled. Let us take \beta\in (0,1) such that |\varepsilon|<1-\beta/2. Then
Thus, with the aid of condition (iii) we have managed to indicate precisely one of the intervals of values of \varepsilon for which a solution exists. Recall that for \varepsilon=1 there is no solution. Let us show that for \varepsilon>1 there is no suitable function V. Suppose that there is a function V\in C^2(\mathbb{R}) satisfying the following conditions: V\geqslant 0, \lim_{|x|\to\infty}V(x)=+\infty, and for some numbers \delta\in [0,1], C>0, and \Lambda>0 the inequality
is fulfilled for all x and all probability measures with finite integrals of the functions |x| and V(x). Substituting \mu=\delta_x into this inequality we obtain
Due to the inequalities 1-\delta\geqslant 0 and \Lambda>0, this leads to a contradiction, because the expression (1-\delta)(C-\Lambda V(x_n)) is negative for sufficiently large n. Recall that for \varepsilon>1 the equation has solutions. Moreover, for \gamma=0 and \varepsilon=1 every measure \delta_a is a solution, but similarly to what has been done above, one can show that also in this case there is no suitable function V. Below we discuss how to modify the hypotheses of Theorem 5.1 for a finer analysis of the existence of solutions.
Let W(x)=\sqrt{1+V(x)}. Suppose, in addition to the hypotheses of Theorem 5.1, that for every measure \mu\in\mathcal{P}_V(\mathbb{R}^d) there exists a positive number C_{\mu} such that for all x\in\mathbb{R}^d we have
holds. It is clear that the set I_0^W is a linear space and contains 1.
Proposition 5.4. Suppose that for some function \psi\in I_0^W there exists a function h\in C(\mathbb{R}^d) such that \sup_x|h(x)|/V(x)<\infty, and for every measure \mu\in \mathcal{P}_V(\mathbb{R}^d) there exist numbers C_1(\mu)\ne 0 and C_2(\mu) satisfying the equality
Let \mu\in \mathcal{P}_V(\mathbb{R}^d) be a solution to the linear equation L_{\sigma}^{*}\mu=0 for some measure \sigma\in \mathcal{P}_V(\mathbb{R}^d). Then
(ii) for every ball B and every \alpha>0, the functions a^{ij}(\,\cdot\,,\mu) and b^i(\,\cdot\,,\mu) on B are uniformly bounded and equicontinuous on B uniformly in \mu\in \mathcal{P}_{\alpha,V}(\mathbb{R}^d) satisfying the equality \mu\big|_{\mathcal{H}}=\nu\big|_{\mathcal{H}};
(iii) if measures \mu_n\in \mathcal{P}_V(\mathbb{R}^d) converge V-weakly to a measure \mu\in\mathcal{P}_V(\mathbb{R}^d), then
for all x\in\mathbb{R}^d and all \mu\in\mathcal{P}_{V}(\mathbb{R}^d) such that \mu\big|_{\mathcal{H}}=\nu\big|_{\mathcal{H}}.
Then there exists a solution \mu to the equation L_{\mu}^{*}\mu=0 such that \mu\big|_{\mathcal{H}}=\nu\big|_{\mathcal{H}}.
Proof. The justification is similar to the proof of Theorem 5.1. We just observe that one has to consider only measures \mu\in \mathcal{P}_{\alpha,V}(\mathbb{R}^d) such that \mu\big|_{\mathcal{H}}=\nu\big|_{\mathcal{H}} and take \alpha greater than C/\Lambda and \|V\|_{L^1(V)}. In addition, if measures \mu_n\in \mathcal{P}_{\alpha,V}(\mathbb{R}^d) converge V-weakly to \mu\in\mathcal{P}_{\alpha, V}(\mathbb{R}^d) and \mu_n\big|_{\mathcal{H}}=\nu\big|_{\mathcal{H}}, then we have \mu\big|_{\mathcal{H}}=\nu\big|_{\mathcal{H}}. \Box
Thus, condition (iv) in Theorem 5.5 is fulfilled for C=a^2/2, \Lambda=1, and \delta=0. This agrees fully with the fact that every measure \mu=\delta_a satisfies the equation L^{*}_{\mu}\mu=0.
and the function K is continuous. Set V(x)=|x|^2/2 and W(x)=\sqrt{1+V(x)}. Observe that the function \psi(x)=\langle v, x\rangle satisfies the equality
Therefore, \psi\in I_0^W and the function x\mapsto \langle v,x\rangle belongs to \mathcal{H}. Let Q\in \mathbb{R}. We consider measures \mu\in\mathcal{P}_V(\mathbb{R}^d) satisfying the condition
holds for \delta=0, \Lambda=q, and C=d+\bigl(|Q\theta|+\sup_{u, v}|K(u, v)|\bigr)^2/(2q). Thus, the hypotheses of the theorem are fulfilled, hence for every Q there exists a solution \mu to the equation L_{\mu}^{*}\mu=0 that satisfies the condition indicated.
In the case where the matrix A is non-degenerate, any solution has a density with respect to Lebesgue measure. Moreover, if the matrix A satisfies the Dini mean oscillation condition, then the solution density has a continuous positive version (see [32], [55], and [57]). These observations enable us in the case of a non-degenerate and sufficiently regular matrix A to construct solutions under weaker restrictions on the drift coefficient and the dependence of coefficients on solutions.
Let \mathcal{CL}_V(\mathbb{R}^d) denote the set of continuous functions \varrho such that
Let \mathcal{CP}_V(\mathbb{R}^d) denote the subset in \mathcal{CL}_V(\mathbb{R}^d) consisting of probability densities. By \mathcal{CP}_{\alpha, V}(\mathbb{R}^d) we denote the subset in \mathcal{CP}_V(\mathbb{R}^d) consisting of densities \varrho satisfying the estimate
We say that a measurable function f on a ball B satisfies the Dini mean oscillation condition with modulus of continuity w_B if there exists a continuous increasing function w_B on [0,+\infty) such that w_B(0)=0,
It is known that f has a continuous version (see [73]), which we deal with below.
Suppose that for every function \varrho\in\mathcal{CP}_V(\mathbb{R}^d) and all i,j\leqslant d we are given Borel functions a^{ij}(\,\cdot\,,\varrho) and b^i(\,\cdot\,,\varrho) on \mathbb{R}^d, where the matrices A(x,\varrho)=\bigl(a^{ij}(x,\varrho)\bigr)_{i,j\leqslant d} are symmetric and non-negative definite. Set
if the measure \mu=\varrho\,dx is a solution to the linear equation with operator L_{\varrho}.
Theorem 5.8. Suppose that the following conditions are fulfilled:
(i) for every ball B and every \alpha>0 one can find a number \lambda_{B,\alpha}>0 and a function w_{B,\alpha} such that for all \varrho\in\mathcal{CP}_{\alpha,V}(\mathbb{R}^d) the functions a^{ij}(\,\cdot\,,\varrho) are continuous on B, satisfy the Dini mean oscillation condition with the modulus of continuity w_{B,\alpha} on B, and for all x\in B the estimate
(iii) if some functions \varrho_n\in \mathcal{CP}_V(\mathbb{R}^d) converge to \varrho\in \mathcal{CP}_V(\mathbb{R}^d) locally uniformly and \|\varrho_n-\varrho\|_{W}\to 0, then the functions a^{ij}(\,\cdot\,, \varrho_n) converge pointwise to the function a^{ij}(\,\cdot\,, \varrho_n) and the functions b^i(\,\cdot\,, \varrho_n) converge to the function b^i(\,\cdot\,,\varrho) in L^1(B) for every ball B;
(iv) there are numbers \delta\in [0,1], \Lambda>0, and C>0 such that
Then there exists a unique solution to the equation L_{\varrho}^{*}\varrho=0 and this solution is a continuous positive function.
Proof. Let \alpha>0 and \sigma\in\mathcal{CP}_{V}(\mathbb{R}^d). According to [29], Theorems 4.2 and 4.7, there exists a unique solution \varrho to the equation L_{\sigma}^{*}\varrho=0. In addition, by [29], Theorem 3.5, for every ball B one can find a number C(B,\alpha) and a modulus of continuity \omega_{B,\alpha} (a strictly increasing function on [0,+\infty) vanishing at zero), which do not depend on \sigma, such that
Let us consider the set X_{\alpha} of functions \varrho\in\mathcal{CP}_{\alpha, V}(\mathbb{R}^d) such that for every ball B, whenever x,y\in B, we have
The set X_{\alpha} is non-empty for sufficiently large \alpha. Every sequence \varrho_n\in X_{\alpha} contains a subsequence \varrho_{n_j} such that it converges uniformly on any ball to some non-negative continuous function \varrho, \varrho\,dx is a probability measure, and the measures \varrho_{n_j}\,dx converge V-weakly to \varrho\,dx. Therefore, \|\varrho_{n_j}-\varrho\|_W\to 0. Thus, the set X_{\alpha} is convex and compact in the space \mathcal{CL}_V(\mathbb{R}^d) with the metric generated by the norm \|\cdot\|_W.
Fix a number \alpha>0 such that the set X_{\alpha} is non-empty and \alpha>C/\Lambda. For every function \sigma\in X_{\alpha} we denote by \Phi(\sigma) the unique solution \varrho to the equation L^{*}_{\sigma}\varrho=0. Let us verify the continuity of the mapping \Phi. Suppose that \sigma_n\in X_{\alpha}, \varrho_n=\Phi(\sigma_n), \sigma\in X_{\alpha}, and \|\sigma-\sigma_n\|_W\to 0. Every subsequence \{\varrho_{n_k}\} contains a further subsequence \{\varrho_{n_{k_j}}\} that converges locally uniformly and in the norm \|\cdot\|_W to \varrho\in X_{\alpha}. Moreover, we can assume that \{\sigma_{n_{j_k}}\} converges locally uniformly to \sigma. Let us verify that \varrho=\Phi(\sigma). It will follow from this that the whole original sequence \{\varrho_n\} converges to \varrho. For every function \varphi\in C_0^{\infty}(\mathbb{R}^d) we have
Let B be a ball containing the support of \varphi. We observe that the sequence of functions a^{ij}(\,\cdot\,,\sigma_{n_{j_k}}) converges to a^{ij}(\,\cdot\,, \sigma) uniformly on B, and the functions \varrho_{n_{j_k}} converge to \varrho in L^1(B). Therefore,
Since the functions b^{i}(\,\cdot\,,\sigma_{n_{j_k}}) converge to b^i(\,\cdot\,,\sigma) in L^1(B) and the sequence \{\varrho_{n_{j_k}}\} is uniformly bounded on B and converges uniformly on B to \varrho, we have
Therefore, \varrho=\Phi(\sigma). Since the mapping \Phi from X_{\alpha} to X_{\alpha} is continuous and X_{\alpha} is convex and compact with respect to the norm \|\cdot\|_W, by Schauder’s theorem (see [34], Theorem 1.12.8) there exists \varrho\in X_{\alpha} such that \varrho=\Phi(\varrho).
Consider an example showing that one cannot expect the uniqueness of solutions under the assumptions of Theorem 5.8.
Observe that c=\|\varrho_1-\varrho_2\|_{L^1(\mathbb{R})}>0 and for every \varrho\in L^1(\mathbb{R}) one has
\begin{equation*}
\|\varrho-\varrho_2\|_{L^1(\mathbb{R})}+ \|\varrho_1-\varrho\|_{L^1(\mathbb{R})}\geqslant c.
\end{equation*}
\notag
Set a(x,\varrho)=1, b(x,\varrho)=-c^{-1}\bigl(\|\varrho- \varrho_2\|_{L^1(\mathbb{R})}+ 4\|\varrho_1-\varrho\|_{L^1(\mathbb{R})}\bigr)x, W(x)=1 and V(x)=|x|^2/2. Since
However, the equation L_{\varrho}^{*}\varrho=0 has two solutions \varrho_1 and \varrho_2. We draw attention to the fact that in this example the diffusion coefficient is one and the drift coefficient is smooth in x and Lipschitz in \varrho with respect to the weighted norm \|\cdot\|_W. Therefore, uniqueness is connected not only with the regularity of coefficients.
The key role in obtaining sufficient conditions for uniqueness is played by estimates for the distances between solutions. Let us give an example of sufficient conditions for uniqueness obtained in this way.
Theorem 5.10. Let A(x,\mu)=\mathrm{I} and V(x)=|x|^2. Suppose that for every measure \mu\in\mathcal{P}_V(\mathbb{R}^d) we are given a function b(\,\cdot\,,\mu) bounded on every ball and satisfying the condition
Then \mu=\sigma. In particular, if \kappa(\mu)=\kappa does not depend on \mu and Q(x)=Q does not depend on x, then for Q<\kappa the equation L_{\mu}^{*}\mu=0 has at most one solution.
Therefore, W_2(\mu,\sigma)\leqslant \kappa(\mu)^{-1}\|Q\|_{L^2(\sigma)}W_2(\mu,\sigma). If \|Q\|_{L^2(\sigma)}<\kappa(\mu), then we obtain W_2(\mu,\sigma)=0. \Box
Consider an example showing that the conditions in Theorem 5.10 are sharp.
Thus, the assumptions of the theorem are fulfilled for \kappa=1 and Q=\varepsilon. If \varepsilon<1, then the solution is unique. It is readily verified that for \varepsilon=1 every measure \mu given by a density (2\pi)^{-1/2}e^{-(x-a)^2/2}, a\in\mathbb{R}, is a solution.
We draw the reader’s attention to the fact that the condition Q<\kappa in Theorem 5.10 and the condition
in the existence theorem mean a certain smalleness of the nonlinear (in \mu) part of the coefficients. The same can be seen from the examples presented above. If we consider a nonlinear equation as a perturbation of a linear equation, then it is natural to expect that under small perturbations existence and uniqueness theorems valid for linear equations remain in forth. Let us mention the following result from [28].
Let \mathcal{P}_k^a(\mathbb{R}^d) denote the subset of L^1(\mathbb{R}^d) consisting of almost everywhere non-negative functions \varrho such that
Suppose that for every function \varrho\in\mathcal{P}_k^a(\mathbb{R}^d) (as usual, we do not mean a single function, but rather the whole equivalence class with respect to equality almost everywhere) and every \varepsilon\in[0,1] we are given Borel functions a^{ij}_{\varepsilon}(\,\cdot\,,\varrho) and b^i_{\varepsilon}(\,\cdot\,,\varrho). The corresponding operator is denoted by L_{\varrho,\varepsilon}. For all functions f such that (1+|x|^k)|f|\in L^1(\mathbb{R}^d) we set
Theorem 5.12. Suppose that the following conditions are fulfilled:
(i) the matrices A_{\varepsilon}(x, \varrho)= \bigl(a^{ij}_{\varepsilon}(x, \varrho)\bigr)_{i,j\leqslant d} are symmetric, non-negative definite, and there exist positive numbers \lambda and \theta such that for all \varrho\in\mathcal{P}_k^a(\mathbb{R}^d) and \varepsilon\in[0,1] one has
(ii) there exist numbers \beta_0>0, \beta_1>0, \beta_2>0, \beta_3>0, and m\geqslant 0 such that for all \varrho\in\mathcal{P}_k^a(\mathbb{R}^d) and \varepsilon\in[0,1] one has
Then there exists \varepsilon_0\in (0, 1] such that for every \varepsilon\in[0, \varepsilon_0] a solution \varrho_{\varepsilon}\in\mathcal{P}_k^a(\mathbb{R}^d) to the equation L_{\varrho, \varepsilon}^{*}\varrho=0 exists and is unique.
The proof makes use of estimates for the distances between solutions and a priori estimates for the logarithmic gradient of the solution.
Let consider an example where the hypotheses of Theorem 5.12 are fulfilled.
Then the hypotheses of Theorem are fulfilled for k=1.
The parameter \varepsilon in the hypotheses of Theorem 5.12 is responsible for the size of the nonlinear part. For \varepsilon=0 condition (iii) yields that the coefficients do not depend on \varrho and the equation is linear. Correspondingly, for the values of the parameter \varepsilon close to zero the nonlinear equation does not differ much from the linear equation and inherits the existence and uniqueness theorems valid for it. Consider a frequently encountered form of the dependence of coefficients on the parameter, when
Suppose that for some \varepsilon>0 there exists a probability solution \varrho_{\varepsilon} to the nonlinear equation with coefficients a^{ij}_{\varepsilon}(x,\varrho) and b^i_{\varepsilon}(x,\varrho). Then \sigma=\varepsilon\varrho_{\varepsilon} is a solution to the equation with the coefficients a^{ij}(x,\varrho) and b^i(x,\varrho) independent of \varepsilon. However, \sigma is not a probability density and the integral of \sigma equals \varepsilon. Thus, in place of the problem of the construction of a probability solution for small values of the parameter \varepsilon one can solve the problem of constructing a non-negative solution with integral equal to \varepsilon. In some physical problems exactly this setting is natural; for example, this is the case for solutions of the Bose–Einstein equation.
We have discussed above conditions on the coefficients admitting only a non-local dependence on solutions, a typical example of which is the expression of the form
However, many popular physical, biological, and economic models (see [48], [65], and [109]) lead to Fokker–Planck–Kolmogorov equations with local nonlinearity, when the coefficients depend on the values of the solution at points.
Let us mention a result from [33] on the existence and uniqueness of non-negative solutions with a prescribed value of the integral for a nonlinear equation with local and non-local nonlinearities.
For a fixed number k\geqslant 1 we denote by \mathcal{CL}_k^{+}(\mathbb{R}^d) the set of non-negative continuous functions \varrho satisfying the condition
As above, a function \varrho \in \mathcal{CL}_k^{+}(\mathbb{R}^d) is called a solution to the nonlinear equation L^{*}_{\varrho}\varrho=0 if \varrho is a solution to the linear equation with the operator L_{\varrho}.
For any function g in the linear span of \mathcal{CL}_k^{+}(\mathbb{R}^d) we set
(ii) there exist positive numbers \beta_1, \beta_2, and r_0 such that for every function g\in \mathcal{CL}_k^{+}(\mathbb{R}^d) such that \|g\|_k\leqslant r_0 and all x\in\mathbb{R}^d and u\in [0,r_0] one has
\begin{equation*}
\langle x, b(x, u, g)\rangle \leqslant \beta_1-\beta_2|x|^2;
\end{equation*}
\notag
(iii) for every r>0 there exists a number C_r>0 such that the inequality
\begin{equation*}
|b(x, u, g)-b(x, v, h)|\leqslant C_r(1+|x|^m)\bigl(|u-v|+\|g-h\|_{k}\bigr)
\end{equation*}
\notag
holds for all x\in\mathbb{R}^d, u,v\in[0,r], and all g, h\in \mathcal{CL}_k^{+}(\mathbb{R}^d) such that \|g\|_k\leqslant r and \|h\|_k\leqslant r.
Then there exists M_0>0 such that for every M \in (0, M_0) the stationary equation L^{*}_{\varrho}\varrho=0 has a unique solution \varrho\in \mathcal{CL}_k^{+}(\mathbb{R}^d) satisfying the conditions
The drift b has the form b(x,u,\varrho)=-x-xu, that is, this equation is a particular case of the equation from the previous example. The hypotheses of Theorem 5.14 are fulfilled, hence for sufficiently small positive M there exists a positive continuous solution the integral of which equals M. However, in this case we can fully investigate the problem of the existence of solutions. It is readily verified that the functions
are solutions. We show that there are no other solutions integrable over \mathbb{R}^d. Let \varrho be a positive continuous integrable solution. Since for k\geqslant 2 we have
Since for \varrho(x)>1 the estimate \varrho(x)/(1+\varrho(x))\geqslant 1/2 holds, it suffices to establish integrability on the set \{x\colon \varrho(x)\leqslant 1\}. Moreover, it suffices to prove integrability on this set for the function \varrho(x)\log^2\varrho(x). Observe that on this set we have the estimate
Substituting in place of \psi the function \psi_N(x)=\zeta(x/N), where \zeta\in C_0^{\infty}(\mathbb{R}^d) and \zeta(x)=1 for |x|\leqslant 1, we obtain
is finite, and therefore there is no stationary solution with integral over \mathbb{R}^d greater than M_1. In the one-dimensional and two-dimensional cases this integral diverges, hence for every M>0 there is a positive solution with integral equal to M.
It was proved in [39], [49], and [47] that in the one-dimensional and two-dimensional cases, for every M>0 there exists a global in time solution \varrho to the Cauchy problem for the equation
such that the integral of the function x\mapsto \varrho(x,t) over \mathbb{R}^d equals M for every t, and it was also shown that, as t\to+\infty, this solution converges to the stationary solution with integral equal to M. In [111] sufficient conditions were obtained for the non-existence of a global in time solution in the three-dimensional case. Finally, in [46] and [69] approximation of equation (5.1) by more regular equations was considered in the multidimensional case, and it was shown that the solutions converge to the curve in the space of measures that is the sum of the Dirac measure at zero and the measure given by the density with respect to Lebesgue measure such that this density satisfies equation (5.1) away from the origin.
6. Nonlinear parabolic equations
Nonlinear parabolic Fokker–Planck–Kolmogorov equations, like nonlinear elliptic equations, differ from linear ones by the dependence of their coefficients on solutions. The nature of the dependence of coefficients is determined by the applied problem that leads to the equation. An important particular case of nonlinear Fokker–Planck–Kolmogorov equations are the famous Vlasov equations, arising, for example, in the study of the behaviour of a large number of interrelated particles. Let us consider the system of differential equations
\begin{equation*}
\dot{x}^i_t=\frac{1}{N}\sum_{j=1}^NK(x^i_t, x^j_t), \qquad x^i_t\in\mathbb{R}^d,\quad 1\leqslant i\leqslant N.
\end{equation*}
\notag
The distribution of the points x^1_t,\dots,x^N_t in \mathbb{R}^d can be described with the aid of the measures
Under broad conditions on the function K one can show (see Dobrushin’s celebrated paper [54]) that if the initial distributions \mu_0^N converge weakly to some measure \nu, then the solutions \mu^N converge weakly to the solution \mu of the Cauchy problem for the same Vlasov equation with initial condition \nu. Note that in [54] the Kantorovich–Rubinshtein metric was used for the first time to prove the solvability of the nonlinear Vlasov equation with the aid of the contracting mapping theorem. In that paper the drift coefficient has the form
where a and B have bounded continuous first order derivatives and B is also bounded.
Now consider the system of stochastic equations
\begin{equation*}
dx^i_t=\frac{1}{N}\sum_{j=1}^NK(x^i_t,x^j_t)\,dt+\sqrt{2}\,dw_t^i, \qquad 1\leqslant i\leqslant N,
\end{equation*}
\notag
where the w_t^i are independent d-dimensional Wiener processes. Suppose that random variables x^1_0,\dots,x^N_0 are independent and have distribution \nu. Under broad conditions on K, for every t the sequence of empiric measures \mu_t^N=\dfrac{1}{N}\displaystyle\sum_{i=1}^N\delta_{x^i_t(\omega)} converges in probability, as N\to\infty, to a (non-random) probability measure \mu_t, where the measure \mu=\mu_t\,dt is a solution to the Cauchy problem for the nonlinear Fokker–Planck–Kolmogorov equation
with the initial condition \mu_0=\nu. Moreover, as N\to\infty, the processes x_t^1,\dots,x_t^N converge to N independent processes y_t^1,\dots,y_t^N satisfying the stochastic Mckean–Vlasov equations
where the w_t^i are independent d-dimensional Wiener processes, \mu_t is the distribution of the random variable y_t^i, and the random variables y_0^i are independent and have distribution \nu. This phenomenon is called ‘propagation of chaos’ (see [77], [93], and [94]).
Thus, a typical example of the dependence of coefficients on solutions, as already noted in the elliptic case, is delivered by the expression
The most important questions in the theory of nonlinear Fokker–Plank–Kolmogorov equations are the existence and uniqueness of solutions to the Cauchy problem and their convergence to the stationary solution as t\to+\infty. In what follows we discuss solutions in the classes of probability measures or non-negative measures; however, equations for signed measures are also of great interest.
Let us proceed to precise formulations and examples. Let T>0 and let V\in C^2(\mathbb{R}^d) be a function such that
For every \tau\in(0,T] let \mathcal{M}_{\tau}(V) denote the set of finite non-negative Borel measures \mu on \mathbb{R}^d\times [0,\tau] of the form \mu=\mu_t\, dt, where \mu_t\in \mathcal{P}(\mathbb{R}^d), for which the mapping t\mapsto \mu_t is continuous in the weak topology and
We say that a sequence of measures \mu^n=\mu^n_t\,dt in \mathcal{M}_{\tau}(V) converges V-weakly to a measure \mu=\mu_t\,dt in \mathcal{M}_{\tau}(V) if for all t\in [0,\tau] the sequence of measures \mu_t^n converges V-weakly to \mu.
Let C^{+}[0,\tau] be the set of non-negative continuous functions on the interval [0,\tau].
For \tau\in(0,T] and a function \alpha\in C^{+}[0,\tau] we denote by M_{\tau,\alpha}(V) the set of measures \mu=\mu_t\,dt\in \mathcal{M}_{\tau}(V) such that for all t\in[0,\tau] we have
As we noted in the elliptic case, if \mu_n=\mu^n_t\, dt\in M_{\tau,\alpha}(V) are measures such that for every t the measures \mu_t^n converge weakly to measures \mu_t and \mu=\mu_t\, dt, then one has the V-convergence of the measures \mu_n to \mu.
Below we assume that each measure \mu=\mu_t\,dt\in \mathcal{M}_{\tau}(V) is extended to [0,T]\times\mathbb{R}^d by the equality \mu_t=\mu_{\tau} if t>\tau. If \mu=\mu_t\,dt\in \mathcal{M}_{\tau,\alpha}(V), then after this extension the new measure \widetilde{\mu} belongs to \mathcal{M}_{\tau,\widetilde{\alpha}}(V), where \widetilde{\alpha}(t)=\alpha(t) for t\in[0,\tau] and \widetilde{\alpha}(t)=\alpha(\tau) for t>\tau. If a sequence of measures \mu^n in \mathcal{M}_{\tau,\alpha}(V) converges V-weakly to a measure \mu\in\mathcal{M}_{\tau,\alpha}(V), then after this extension the sequence of measures \widetilde{\mu}^n converges V-weakly to the measure \widetilde{\mu}.
We define similarly the classes \mathcal{SM}_{\tau}(V) and \mathcal{SM}_{\tau,\alpha}(V) consisting of the measures \mu=\mu_t\,dt, where each \mu_t is subprobability measure on \mathbb{R}^d.
Suppose that for all \mu\in\mathcal{M}_{T}(V) and all i, j\leqslant d we are given Borel functions a^{ij}(\,\cdot\,{,}\,\cdot\, ,\mu) and b^{i}(\,\cdot\, {,} \,\cdot\,, \mu) on \mathbb{R}^d\times[0,T] such that the matrix A(x,t,\mu)=(a^{ij}(x,t,\mu))_{i,j\leqslant d} is symmetric and non-negative definite. Set
Let \nu\in \mathcal{P}(\mathbb{R}^d). A measure \mu=\mu_t\, dt\in\mathcal{M}_{\tau}(V) on \mathbb{R}^d\times[0,\tau] is called a solution to the Cauchy problem for the nonlinear Fokker–Planck–Kolmogorov equation
if \mu is a solution to the Cauchy problem for the linear equation with the operator L_{\mu}. As in the linear case, for the measure \mu=\mu_t\, dt this expression is identical to the shortened version
The next theorem generalizes some results from [92] (see also [23], Chap. 6). An analogous assertion for the transport equation was obtained earlier in [24].
Theorem 6.1. Suppose that the following conditions are fulfilled:
(i) for each ball B\subset\mathbb{R}^d and each function \alpha\in C^{+}[0, T] the functions a^{ij}(\,\cdot\,, t, \mu) and b^i(\,\cdot\,, t, \mu) are bounded on B uniformly in \mu\in \mathcal{M}_{T,\alpha}(V) and t\in[0,T], and for almost all t\in[0,T] they are equicontinuous on B uniformly in \mu\in \mathcal{M}_{T,\alpha}(V);
(ii) if a sequence of measures \mu^n\in \mathcal{M}_{T,\alpha}(V) converges V-weakly to some measure \mu\in \mathcal{M}_{T,\alpha}(V), then for almost every t\in[0,T], for all x\in B we have
(iii) there are mappings \Lambda_1 and \Lambda_2 associating to any function \alpha\in C^{+}[0,T] functions \Lambda_1[\alpha] and \Lambda_2[\alpha] in C^{+}[0,T] such that for all \alpha\in C^{+}[0,T] and all \mu\in\mathcal{M}_{T,\alpha}(V) we have
Then for every probability measure \nu such that V\in L^1(\nu) there exists \tau in (0,T] such that there is a measure \mu=\mu_t\,dt\in\mathcal{M}_{\tau}(V) on \mathbb{R}^d\times[0,\tau] that is a solution to the Cauchy problem \partial_t\mu_t=L_{\mu}^{*}\mu_t, \mu_0=\nu. Moreover, if
where g_1 and g_2 are positive non-decreasing continuous functions on [0,+\infty), then as \tau one can take an arbitrary number not greater than T and smaller than \tau^{*}, where
Proof. Let \sigma=\sigma_t\,dt\in\mathcal{M}_{\tau,\alpha}(V). According to [23], Theorems 6.7.3 and 7.1.1, the Cauchy problem \partial_t\mu_t=L_{\sigma}^{*}\mu_t, \mu_0=\nu has a solution \mu=\mu_t\,dt\in\mathcal{M}_{\tau}(V) for which
and take \tau\in(0,T] sufficiently small so that Q(t)\leqslant 1 and R(t)\leqslant 2 for t\in[0,\tau]. When \Lambda_1[\alpha](t)=g_1(\alpha(t)) and \Lambda_2[\alpha](t)=g_2(\alpha(t)), we fix a number \tau satisfying the conditions
\begin{equation*}
0<\tau\leqslant T \quad\text{and}\quad \tau<\int_{u_0}^{+\infty}\frac{du}{g_1(u)+ug_2(u)}\,,
\end{equation*}
\notag
and find the function \alpha by solving on [0,\tau] the differential equation
It is clear that the integral of V against the measure \mu_t does not exceed \alpha(t). Choosing \tau and \alpha in this way, we obtain that the measure \sigma\in\mathcal{M}_{\tau, \alpha}(V) corresponds to the solution \mu in \mathcal{M}_{\tau, \alpha}(V). Let \varphi\in C_0^{\infty}(\mathbb{R}^d). Since for all t,s\in[0,\tau] we have
and by condition (i) the coefficients of operator L_{\sigma} on the support of \varphi are bounded uniformly in \sigma\in\mathcal{M}_{\tau,\alpha}(V) and t\in[0,\tau], we obtain
The set \mathcal{K} is convex and non-empty. We show that it is compact in the space of finite non-negative measures on \mathbb{R}^d\times[0,\tau] (which is metrizable by the Kantorovich–Rubinshtein metric). Let \sigma^n\in\mathcal{K}. Since \|V\|_{L^1(\sigma_t^n)}\leqslant\alpha(t), for each fixed t we can find a weakly convergent subsequence of \{\sigma_t^n\}. Using the diagonal procedure we extract a subsequence \{n_k\} such that the measures \sigma_r^{n_k} converge weakly to some probability measures \sigma_r for all rational numbers r\in [0,\tau]. Let t\in [0,T] and \varepsilon>0. For every function \varphi\in C_0^{\infty}(\mathbb{R}^d) there exists a rational number r\in[0,T] such that C(\varphi)|t-r|<\varepsilon. Let N be such that
is Cauchy. This yields the weak convergence of the sequence \{\sigma_t^{n_k}\} to some probability measure \sigma_t. Observe that \|V\|_{L^1(\sigma_t)}\leqslant\alpha(t) and the mapping t\mapsto\sigma_t is continuous, because
Thus, we have obtained a measure \sigma=\sigma_t\,dt\in\mathcal{M}_{\tau, \alpha}(V) to which the subsequence \{\sigma^{n_k}\} converges weakly.
Consider the multivalued mapping \Phi that with every measure \sigma\in\mathcal{K} associates the subset \Phi(\sigma)\subset\mathcal{K} consisting of the solutions to the Cauchy problem \partial_t\mu=L_{\sigma}^{*}\mu, \mu_0=\nu. It is clear that \Phi(\sigma) is non-empty and convex. Let us verify that the graph of \Phi is closed. Suppose that measures \sigma^n\in\mathcal{K} converge weakly to \sigma\in\mathcal{K} and measures \mu^n\in\mathcal{K} converge weakly to \mu. We show that \mu\in\Phi(\sigma). Passing to a subsequence \{n_k\} we can assume that for every t the measures \sigma^{n_k}_t and \mu^{n_k}_t converge V-weakly to \sigma_t and \mu_t, respectively. Then by conditions (i) and (ii), on every ball, for almost all t the sequences of functions a^{ij}(\,\cdot\,, t, \sigma^{n_k}), b^i(\,\cdot\, , t, \sigma^{n_k}) converge uniformly to the continuous functions a^{ij}(\,\cdot\,, t, \sigma) and b^i(\,\cdot\,, t, \sigma). This observation and the weak convergence of \{\mu^{n_k}_t\} to \mu_t enable us to justify the equality
Thus, the measure \mu=\mu_t\,dt belongs to \Phi(\sigma). It also follows that \Phi(\sigma) is closed. By the Kakutani–Ky Fan theorem there exists a measure \mu\in\mathcal{K} belonging to \Phi(\mu), that is, \mu is a solution to the Cauchy problem \partial_t\mu=L^{*}_{\mu}\mu, \mu_0=\nu. By the definition of the set \mathcal{K} the mapping t\mapsto\mu_t is continuous. \Box
Let us give a typical example when Theorem 6.1 is applicable.
where q^{ij} and h^i are continuous functions, the matrix Q(x,y)=(q^{ij}(x,y))_{i,j\leqslant d} is symmetric, and 0\leqslant Q(x,y)\leqslant \theta\cdot {\rm I} for some number \theta and all x and y. Suppose that the vector field h=(h^i)_{i\leqslant d} satisfies the estimates
Thus, the hypotheses of Theorem 6.1 are fulfilled for the constant functions g_1=g_2=C, hence a solution to the Cauchy problem exists on the interval [0,\tau] for every \tau>0 since the integral of the function 1/(C+Cu) diverges.
With the aid of the superposition principle (see, for example, [30]), from the existence of a solution to the Cauchy problem for the Fokker–Planck–Kolmogorov equation one can derive the existence of a solution to the nonlinear martingale problem. Recall that the superposition principle for a solution (\mu_t)_{t\in [0,T]} to the Cauchy problem for the Fokker–Planck–Kolmogorov equation with the operator L gives (under broad assumptions on L) a Borel probability measure P on the path space C([0,T],\mathbb{R}^d) such that \mu_t is the image P\circ e_t^{-1} of the measure P under the mapping e_t(\omega)=\omega(t), t\in [0,T], and for all \varphi\in C_0^{\infty}(\mathbb{R}^d) the process
is a martingale with respect to P and the filtration \mathcal{F}_t=\sigma\bigl(\omega(s), s\leqslant t\bigr). The measure P is called a solution to the martingale problem with operator L. Thus, the superposition principle means a representation of the solution to the Cauchy problem for the Fokker–Planck–Kolmogorov equation in the form of the one-point distributions of a solution to the martingale problem.
Theorem 6.3. Suppose that for every measure \mu\in\mathcal{M}_T(V) there exists C(\mu)>0 such that
Let \tau\in(0,T], and let \nu be a probability measure such that V\in L^1(\nu). Suppose that a measure \mu=\mu_t\,dt\in\mathcal{M}_{\tau}(V) is a solution to the Cauchy problem \partial_t\mu_t=L_{\mu}^{*}\mu_t, \mu_0=\nu.
Then there exists a probability measure P on C([0,T],\mathbb{R}^d) that is a solution to the martingale problem for the operator L_{\mu} such that \mu_t=P\circ e_t^{-1}.
Proof. We observe that the functions \|A(\,\cdot\,{,}\,\cdot\,,\mu)\| and |b(\,\cdot\,{,}\,\cdot\,,\mu)| are integrable against the measure \mu on \mathbb{R}^d\times[0,\tau] and apply the superposition principle for the linear equation with the operator L_{\mu}, which is possible under the conditions indicated in the theorem (see [30]). \Box
The existence and uniqueness of solutions to the nonlinear martingale problem and the corresponding stochastic equation were studied in [64], [95], and [96].
Now consider an example where a solution does not exist on every interval [0,\tau].
Let \nu=\delta_a, where a>0. Then f(t) is defined on [0,a^{-1}), and the Cauchy problem for the equation \partial_t\mu_t=L_{\mu}^{*}\mu_t with initial condition \nu has no solutions on [0,\tau] for \tau\geqslant a^{-1}, which agrees with the estimate \tau<u_0^{-1/2} obtained above since here we have u_0=a^2.
Let us give a sufficient condition for the absence of a global solution. The problem of the ‘blow-up’ of solutions to equations of the form under consideration has a vast literature; see [97], [98], [14], [43], and [85]. The following result was obtained in [92].
Theorem 6.5. Suppose that conditions (i) and (ii) in Theorem 6.1 are fulfilled and there is a continuous increasing positive function G on [0,+\infty) such that
Also let |\sqrt{A(x,t,\mu)}\,\nabla V(x)|^2\leqslant C_1+C_2V(x) for some positive numbers C_1 and C_2. Suppose that u_{0}=\|V\|_{L^1(\nu)}>0 and in case (a) the function 1/(uG(u)) is integrable on [u_0,+\infty) and its integral equals \tau_1<T, while in case (b) the function 1/G(u) is integrable on [u_0,+\infty) and its integral equals \tau_1<T. Then the Cauchy problem \partial_t\mu_t=L_{\mu}^{*}\mu_t, \mu_0=\nu has no solution in \mathcal{M}_T(V) on the interval [0,\tau] for \tau\geqslant \tau_1.
In [68] some integral conditions involving Lyapunov functions were considered. Let us state for example the following assertion, which is analogous to a result in [68] for the stochastic equation, proved under some additional restrictions on the growth of coefficients, but for more general Lyapunov functions. The existence result for the stochastic equation implies the solvability of the Cauchy problem for the nonlinear Fokker–Planck–Kolmogorov equation (but the converse result can be obtained from the superposition principle). Let a^{ij}(x,t,\mu) and b^i(x,t,\mu) be defined for all measures \mu\in\mathcal{SM}_T(V).
Theorem 6.6. Let conditions (i) and (ii) in Theorem 6.1 be fulfilled, but for the classes \mathcal{SM}_{T,\alpha}(V) in place of \mathcal{M}_{T,\alpha}(V). Suppose that there exist positive numbers C_1 and C_2 such that for all measures \mu\in\mathcal{SM}_{T,\alpha}(V) with compact support in \mathbb{R}^d\times[0,T] we have
Then, for every probability measure \nu such that V\in L^1(\nu) there exists a solution \mu=\mu_t\,dt\in\mathcal{M}_T(V) to the Cauchy problem \partial_t\mu_t=L_{\mu}^{*}\mu_t, \mu_0=\nu.
Proof. Suppose that \varphi\in C_0^{\infty}(\mathbb{R}^d), 0\leqslant\varphi\leqslant 1, and \varphi(x)=1 if |x|<1, \nu^n is a probability measure with support in the ball |x|<n, the measures \nu^n converge weakly to \nu, and the norms \|V\|_{L^1(\nu^n)} are uniformly bounded by some number C_{\nu}. Set \varphi_n(x)=\varphi(x/n) and L_{\mu,n}=\varphi_nL_{\varphi_n\mu}. By Theorem 6.1 there exists a solution \mu^n\in\mathcal{M}_T(V) to the Cauchy problem \partial_t\mu_t=L_{\mu,n}^{*}\mu_t, \mu_0=\nu^n. Observe that \varphi_n\mu^n \in \mathcal{SM}_{T,\alpha}(V), and so
It follows from our conditions on coefficients that in this case equality (3.7) remains valid for all functions \varphi\in C^2(\mathbb{R}^d) with compact support, but since the function \varphi_n vanishes outside a ball, for \mu=\mu^n this equality is also true for \varphi=V. Therefore,
Since \varphi_n(x)a^{ij}(x,t,\varphi_n\mu^n) and \varphi_n(x)b^{i}(x,t,\varphi_n\mu^n) are bounded on every ball B\subset \mathbb{R}^d uniformly in t\in[0,T] and n, for every function \psi\in C_0^{\infty}(\mathbb{R}^d) we obtain the estimate
for some constant C(\psi) independent of t, s, and n. Repeating the reasoning from the proof of Theorem 6.1, we find a subsequence \{n_k\} such that the sequence \{\mu_t^{n_k}\} converges weakly to some probability measure \mu_t; moreover, \|V\|_{L^1(\mu_t)}\leqslant\alpha(t) and the mapping t\mapsto\mu_t is continuous. Thus, we have obtained the measure \mu=\mu_t\,dt\in\mathcal{M}_{T,\alpha}(V). It is clear that the measures \mu^{n_k} converge V-weakly to \mu, and the analogous assertion is true for the sequence \{\varphi_{n_k}\mu^{n_k}\}. For every function \psi\in C_0^{\infty}(\mathbb{R}^d) we have
By the hypotheses of the theorem the sequences a^{ij}(x,t,\varphi_n\mu^n) and b^{i}(x,t,\varphi_n\mu^n) converge uniformly on every ball to functions a^{ij}(x,t,\mu) and b^{i}(x,t,\mu) continuous in x. This enables us to pass to the limit as n\to\infty and obtain the equality
In this situation Theorem 6.1 enables us to construct a solution only for \tau<u_0^{-2}, where u_0=\|V\|_{L^1(\nu)}. The condition from the theorem has the form
However, by Theorem 6.6 a solution exists for all \tau>0.
One also considers Lyapunov functions depending on the measure \mu. In this case an important role is played by differential calculus on the space of probability measures and generalizations of Itôs formula to the case of functions depending on the distributions of random process (see [41], Chap. 5, and [68]). Applications of differential calculus on the space of probability measures to the justification of the existence and uniqueness of solutions to nonlinear martingale problems and the corresponding nonlinear Fokker–Planck–Kolmogorov equations were discussed in [50]. The main novelty of that work consists in conditions on the dependence of the diffusion coefficient on the solution, because in place of the classical Lipschitz condition the existence and regularity of a linear derivative of A with respect to the measure are assumed.
In the case where the matrix A is non-degenerate and sufficiently regular, a solution possesses a continuous density with respect to Lebesgue measure, and it is more natural to consider nonlinear equations in the space of densities rather than measures. Let us give an example of such an assertion.
We denote by \mathcal{L}_{\tau}(V) and \mathcal{L}_{\tau,\alpha}(V) the subsets of \mathcal{M}_{\tau}(V) and \mathcal{M}_{\tau,\alpha}(V), respectively, consisting of the measures \mu=\mu_t\,dt such that \mu_t(dx)=\varrho(x,t)\,dx for every t\in(0,\tau]. Note that we do not assume the absolute continuity of the measure \mu_0.
Let B be an open ball in \mathbb{R}^d. Let B_r(z) denote the intersection B\cap B(z,r), where B(z,r) is the open ball of radius r with centre z. Set
We say that a measurable function f on B\times(0, T) satisfies the Dini mean oscillation condition in x with modulus of continuity w_B if there exists a continuous increasing function w_B on [0,+\infty) such that w_B(0)=0 and
Now assume that the coefficients of the operator L_{\mu} are defined only on the set \mathcal{L}_T(V). By a solution to the Cauchy problem \partial_t\mu=L^{*}_{\mu}\mu, \mu_0=\nu we mean a measure \mu\in\mathcal{L}_T(V) that is a solution to the Cauchy problem for the linear equation with operator L_{\mu} and initial condition \nu.
Theorem 6.8. Suppose that the following conditions are fulfilled:
(i) for every ball B and every function \alpha\in C^{+}[0,T] one can find a positive number \lambda_{B,\alpha} and function w_{B,\alpha} such that for all \mu\in\mathcal{L}_{T,\alpha}(V) and all t\in [0,T] the functions a^{ij}(\,\cdot\,,t,\mu) are continuous on B, satisfy the Dini mean oscillation condition with modulus of continuity w_{B,\alpha} on B\times(0,T), and for all x\in B and t\in[0,T] one has
(ii) for every ball B\subset\mathbb{R}^d and every function \alpha\in C^{+}[0, T] the functions b^i(\,\cdot\,,t,\mu) are bounded on B uniformly in \mu\in \mathcal{L}_{T,\alpha}(V) and t\in[0,T], and for almost all t\in[0,T] they are equicontinuous on B uniformly in \mu\in \mathcal{L}_{T,\alpha}(V);
(iii) if measures \mu^n=\varrho^n(x,t)\,dx\,dt and \mu=\varrho(x,t)\,dx\,dt belong to \mathcal{L}_{T,\alpha}(V) and for almost every t\in (0,T] the sequence \varrho^n(\,\cdot\,,t) converges to \varrho(\,\cdot\,,t) in L^1(\mathbb{R}^d), then for almost every t\in[0,T], for all x\in\mathbb{R}^d we have
Then for every probability measure \nu satisfying the condition V\!\in L^1(\nu) there exists a solution \mu\in\mathcal{L}_T(V) to the Cauchy problem \partial_t\mu_t=L_{\mu}^{*}\mu_t, \mu_0=\nu.
Proof. Let \omega\in C_0^{\infty}(\mathbb{R}^d) satisfy
that is, (\omega_n*\mu)(x,t)\,dx\,dt\in\mathcal{L}_{T,\widetilde{\alpha}}. Moreover, if some measures \mu^k\in\mathcal{M}_{T,\alpha}(V) converge V-weakly to \mu\in\mathcal{M}_{T, \alpha}(V), then for every t\in[0,T] the probability densities x\mapsto (\omega_n*\mu^k)(x,t) converge pointwise to the probability density x\mapsto (\omega_n*\mu)(x,t), which implies their convergence in L^1(\mathbb{R}^d).
Set L_{\mu, n}=L_{(\omega_n*\mu)\,dx\,dt}. According to the observations made above, the coefficients of the operator L_{(\omega_n*\mu)\,dx\,dt} satisfy all conditions in Theorem 6.1 for \Lambda_1=\Lambda_2= C_2. Therefore, on \mathbb{R}^d\times[0,T] there exists a solution \mu^n\in\mathcal{M}_T(V) to the Cauchy problem \partial_t\mu_t=L_{\mu, n}^{*}\mu_t, \mu_0=\nu. Observe that \mu^n satisfies the inequality
that is, \mu^n\in\mathcal{M}_{T, \alpha}(V) for all n. Hence \omega_n*\mu^n\in\mathcal{M}_{T,\widetilde{\alpha}}(V), where \widetilde{\alpha}=C_1+C_1\alpha. Moreover, the solution \mu^n has a continuous density \varrho^n with respect to Lebesgue measure, and according to [56], Theorem 1.4, for every ball B and every interval J\subset(0,T) one can find a number C(B,J,\widetilde{\alpha}) and a continuous increasing function \omega_{B,J,\widetilde{\alpha}} independent of n such that for all x,y\in B and t,s\in J one has
Passing to a subsequence, we can assume that \{\varrho^n\} converges to some continuous function \varrho on \mathbb{R}^d\times(0,T) uniformly on every cylinder B\times J. Since for t>0 the functions \varrho^n(\,\cdot\,,t) are probability densities and the integrals of V(x)\varrho^n(x,t) over \mathbb{R}^d are uniformly bounded, the function \varrho(\,\cdot\,,t) is a probability density and the sequence \varrho^n(\,\cdot\,,t) converges to it in L^1(\mathbb{R}^d). In addition, for every t>0 the sequence (\omega_n*\mu^n)(\,\cdot\,, t) converges to \varrho(\,\cdot\,, t) in L^1(\mathbb{R}^d) because
Thus, the measure \mu=\varrho(x,t)\,dx\,dt is the required solution. \Box
Note that in the proof of Theorem 6.8 we actually approximated the nonlinear equation for the density \varrho by a nonlinear equation for the measure \mu, by using the substitution of the expression \omega_n*\mu in place of \varrho into the coefficients of the equation. This approach has successfully been applied to the approximation of nonlinear Fokker–Planck–Kolmogorov equations with local nonlinearity by equations with a nonlinearity of nonlocal type (see [52]). One cannot expect the uniqueness of solutions in Theorem 6.8.
and a(x,0,\mu)=1+c^{-1}\|\mu_0-\delta_0\|. Since a(x,t,\varrho_1\,dx\,dt)=1 and a(x,t,\varrho_2\,dx\,dt)=2 for t>0, the measures \varrho_1\,dx\,dt and \varrho_2\,dx\,dt are solutions to the Cauchy problem
that is, in this example the coefficient a is non-degenerate, constant in x, and Lipschitz with respect to \mu.
Other examples of non-uniqueness can be found in [90]. One method for deriving sufficient conditions for uniqueness consists in applying a suitable estimate for distances between solutions to linear equations. Let us mention an assertion from [108].
Theorem 6.10. Let V(x)=1+|x|^{2m}, m\geqslant 1. Suppose that the following conditions are fulfilled:
(i) for every measure \mu\in\mathcal{L}_T(V) there exist positive numbers \lambda(\mu) and \theta(\mu) such that for all x, y\in\mathbb{R}^d and all t\in[0, T] one has
hold for almost all (x,t)\in\mathbb{R}^d\times[0,T].
Then for every probability measure \nu=\varrho_0\,dx satisfying the conditions \varrho_0\log\varrho_0\in L^1(\mathbb{R}^d) and V\varrho_0\in L^1(\mathbb{R}^d), the Cauchy problem \partial_t\mu_t=L_{\mu}^{*}\mu_t, \mu_0=\nu can have at most one solution \mu\in \mathcal{L}_T(V).
The key role in justification is played by the estimate from Theorem 4.7. Various sufficient conditions for uniqueness are presented in [90].
An important problem is the study of the behaviour of solutions \mu=\mu_t\,dt to the parabolic equation as t\to\infty and, in particular, obtaining sufficient conditions for convergence to the stationary solution. A considerable number of papers are devoted to this question, among which we single out [36], [44], [45], [67], [82], [86], and [107]. For example, in [36] solutions \mu_t\in \mathcal{P}_2(\mathbb{R}^n) to the nonlinear equation
were obtained. Convergence to the stationary solution in the total variation distance in place of the Kantorovich metric was investigated in [38], [28], [29], and [12]. Let us present a result from [28].
Let A(x)=(a^{ij}(x))_{i,j\leqslant d} be a symmetric positive definite matrix such that there exist positive numbers \theta and \lambda for which
Let V\geqslant 1. Recall that \mathcal{P}_V(\mathbb{R}^d) is the set of probability measures \mu on \mathbb{R}^d for which V\in L^1(\mu), and \mathcal{P}_{\alpha, V}(\mathbb{R}^d) is the subset of \mathcal{P}_V(\mathbb{R}^d) consisting of measures \mu for which \|V\|_{L^1(\mu)}\leqslant\alpha. Suppose that for every \varepsilon\in[0, 1] and every measure \mu\in\mathcal{P}_V(\mathbb{R}^d) we are given a Borel vector field b_{\varepsilon}(\,\cdot\,,\mu) on \mathbb{R}^d. Set
Theorem 6.11. Suppose that the following conditions are fulfilled:
(i) for every \alpha>0 there exist positive numbers N_1(\alpha) and N_2(\alpha) such that for every \varepsilon\in[0, 1], all \mu\in\mathcal{P}_{\alpha,V}(\mathbb{R}^d), and all x\in\mathbb{R}^d one has
Then one can find \varepsilon_0\in (0, 1] such that for every \varepsilon\in[0, \varepsilon_0] there exists a solution \mu^{\varepsilon}\in\mathcal{P}_V(\mathbb{R}^d) to the stationary equation L^{*}_{\mu,\varepsilon}\mu=0, and for every probability measure \nu\in\mathcal{P}_V(\mathbb{R}^d) on \mathbb{R}^d\times[0,+\infty) there exists a unique solution \mu_t^{\varepsilon}\,dt with \mu_t^{\varepsilon}\in\mathcal{P}_V(\mathbb{R}^d) to the Cauchy problem \partial_t\mu_t=L_{\mu_t,\varepsilon}^{*}\mu_t, \mu_0=\nu, moreover,
A similar result on convergence to the stationary solution in the case where the diffusion matrix depends on the solution was obtained in [29]. The existence of a global solution to a parabolic equation and its convergence to the stationary solution in the case where the equation contains nonlinear terms of local and non-local type were studied in [33].
In [70]–[72], [74], and [116], nonlinear Fokker–Plank–Kolmogorov equations on infinite-dimensional spaces were considered. Let E=C([-1,0],\mathbb{R}^d) be the space of continuous mappings with the standard \sup-norm, and let the functions b^i and a^{ij} on E\times \mathcal{P}(\mathbb{R}^d) be Borel measurable. For a fixed measure \mu\in\mathcal{P}(\mathbb{R}^d) the operator L_{\mu} on C_0^\infty(\mathbb{R}^d) with values in the space of Borel functions on E is defined by the formula
Then a Borel mapping t\mapsto \mu_t\in \mathcal{P}(E) with values in the space of probability measures on E is called a solution to the nonlinear Cauchy problem \partial_{t}\mu(t)=L_{\mu_t}^*\mu_t with initial condition \mu_0, where \mu(t) is the image of the measure \mu_t under the projection v\mapsto v(0) of E onto \mathbb{R}^d, if for all f\in C_0^\infty(\mathbb{R}^d) the function L_{\mu_t}f is integrable against the measure \mu_t\, dt on the sets E\times [0,T] and
In the papers cited above one can find conditions for the solvability of such equations and some properties of solutions. In [113] the author studied the differentiability of nonlinear functionals on the space of measures with respect to a solution to a parabolic Fokker–Planck–Kolmogorov equation.
For the selection of solutions with the flow property in the case of the non-unique solvability of a nonlinear equation, see [103].
We do not discuss equations on domains. In this connection see, for example, [102], where the authors considered the equation with nonlinearity in the right-hand side, of the form
with a singular convolution kernel K was considered. The stochastic approximation of the Navier–Stokes–Vlasov–Fokker–Planck system was studied in [61].
A linearization of nonlinear parabolic Fokker–Planck–Kolmogorov equations was discussed in [105] and [104].
The authors thank the referee for useful comments.
Bibliography
1.
N. U. Ahmed and Xinhong Ding, “On invariant measures of nonlinear Markov processes”, J. Appl. Math. Stochastic Anal., 6:4 (1993), 385–406
2.
F. Anceschi and Yuzhe Zhu, “On a spatially inhomogeneous nonlinear Fokker–Planck equation: Cauchy problem and diffusion asymptotics”, Anal. PDE, 17:2 (2024), 379–420
3.
V. Barbu, “Generalized solutions to nonlinear Fokker–Planck equations”, J. Differential Equations, 261:4 (2016), 2446–2471
4.
V. Barbu, Semigroup approach to nonlinear diffusion equations, World Sci. Publ., Hackensack, NJ, 2022, vii+212 pp.
5.
V. Barbu, “The Trotter product formula for nonlinear Fokker–Planck flows”, J. Differential Equations, 345 (2023), 314–333
6.
V. Barbu and M. Röckner, “Nonlinear Fokker–Planck equations driven by Gaussian linear multiplicative noise”, J. Differential Equations, 265:10 (2018), 4993–5030
7.
V. Barbu and M. Röckner, “Probabilistic representation for solutions to nonlinear Fokker–Planck equations”, SIAM J. Math. Anal., 50:4 (2018), 4246–4260
8.
V. Barbu and M. Röckner, “Solutions for nonlinear Fokker–Planck equations with measures as initial data and McKean–Vlasov
equations”, J. Funct. Anal., 280:7 (2021), 108926, 35 pp.
9.
V. Barbu and M. Röckner, “Uniqueness for nonlinear Fokker–Planck equations and weak uniqueness for McKean–Vlasov SDEs”, Stoch. Partial Differ. Equ. Anal. Comput., 9:3 (2021), 702–713; correction: 11:1 (2023), 426–431
10.
V. Barbu and M. Röckner, “The invariance principle for nonlinear Fokker–Planck equations”, J. Differential Equations, 315 (2022), 200–221
11.
V. Barbu and M. Röckner, “Uniqueness for nonlinear Fokker–Planck equations and for McKean–Vlasov SDEs: the degenerate case”, J. Funct. Anal., 285:4 (2023), 109980, 37 pp.
12.
V. Barbu and M. Röckner, “The evolution to equilibrium of solutions to nonlinear Fokker–Planck equation”, Indiana Univ. Math. J., 72:1 (2023), 89–131
13.
Ya. I. Belopol'skaya, “Systems of nonlinear backward and forward Kolmogorov equations: generalized solutions”, Theory Probab. Appl., 66:1 (2021), 15–43
14.
A. L. Bertozzi, J. A. Carrillo, and T. Laurent, “Blow-up in multidimensional aggregation equations with mildly singular interaction kernels”, Nonlinearity, 22:3 (2009), 683–710
15.
V. I. Bogachev, Weak convergence of measures, Math. Surveys Monogr., 234, Amer. Math. Soc., Providence, RI, 2018, xii+286 pp.
16.
V. I. Bogachev, “Kantorovich problem of optimal transportation of measures: new directions of research”, Russian Math. Surveys, 77:5 (2022), 769–817
17.
V. I. Bogachev, G. Da Prato, M. Röckner, and S. V. Shaposhnikov, “Nonlinear evolution equations for measures on infinite dimensional spaces”, Stochastic partial differential equations and applications, Quad. Mat., 25, Dept. Math., Seconda Univ. Napoli, Caserta, 2010, 51–64
18.
V. I. Bogachev, A. I. Kirillov, and S. V. Shaposhnikov, “The Kantorovich and variation distances between invariant measures
of diffusions and nonlinear stationary Fokker–Planck–Kolmogorov equations”, Math. Notes, 96:5 (2014), 855–863
19.
V. I. Bogachev, A. I. Kirillov, and S. V. Shaposhnikov, “Distances between stationary distributions of diffusions and solvability of nonlinear
Fokker–Planck–Kolmogorov equations”, Theory Probab. Appl., 62:1 (2018), 12–34
20.
V. I. Bogachev, A. V. Kolesnikov, and S. V.Shaposhnikov, Monge and Kantorovich problems of optimal transportation, Regulyarnaya i Khaotichaskaya Dinanika, Institute for Computer Studies, Moscow–Izhevsk, 2023, 664 pp. (Russian)
21.
V. I. Bogachev, T. I. Krasovitskii, and S. V. Shaposhnikov, “On uniqueness of probability solutions of the Fokker–Planck–Kolmogorov equation”, Sb. Math., 212:6 (2021), 745–781
22.
V. I. Bogachev, N. V. Krylov, and M. Röckner, “Elliptic and parabolic equations for measures”, Russian Math. Surveys, 64:6 (2009), 973–1078
23.
V. I. Bogachev, N. V. Krylov, M. Röckner, and S. V. Shaposhnikov, Fokker–Planck–Kolmogorov equations, Math. Surveys Monogr., 207, Amer. Math. Soc., Providence, RI, 2015, xii+479 pp.
24.
V. I. Bogachev, M. Röckner, and S. V. Shaposhnikov, “Nonlinear evolution and transport equations for measures”, Dokl. Math., 80:3 (2009), 785–789
25.
V. I. Bogachev, M. Röckner, and S. V. Shaposhnikov, “Distances between transition probabilities of diffusions and applications to nonlinear
Fokker–Planck–Kolmogorov equations”, J. Funct. Anal., 271:5 (2016), 1262–1300
26.
V. I. Bogachev, M. Röckner, and S. V. Shaposhnikov, “The Poisson equation and estimates for distances between stationary distributions of diffusions”, J. Math. Sci. (N.Y.), 232:3 (2018), 254–282
27.
V. I. Bogachev, M. Röckner, and S. V. Shaposhnikov, “Convergence to stationary measures in nonlinear Fokker–Planck–Kolmogorov equations”, Dokl. Math., 98:2 (2018), 452–457
28.
V. I. Bogachev, M. Röckner, and S. V. Shaposhnikov, “Convergence in variation of solutions of nonlinear Fokker–Planck–Kolmogorov equations to stationary
measures”, J. Funct. Anal., 276:12 (2019), 3681–3713
29.
V. I. Bogachev, M. Röckner, and S. V. Shaposhnikov, “On convergence to stationary distributions for solutions of nonlinear Fokker–Planck–Kolmogorov
equations”, J. Math. Sci. (N. Y.), 242:1 (2019), 69–84
30.
V. I. Bogachev, M. Röckner, and S. V. Shaposhnikov, “On the Ambrosio–Figalli–Trevisan superposition principle for probability solutions to
Fokker–Planck–Kolmogorov equations”, J. Dynam. Differential Equations, 33:2 (2021), 715–739
31.
V. I. Bogachev, M. Röckner, and S. V. Shaposhnikov, “Kolmogorov problems on equations for stationary and transition probabilities of diffusion processes”, Theory Probab. Appl., 68:3 (2023), 342–369
32.
V. I. Bogachev, M. Röckner, and S. V. Shaposhnikov, “Zvonkin's transform and the regularity of solutions to double divergence form elliptic equations”, Comm. Partial Differential Equations, 48:1 (2023), 119–149
33.
V. I. Bogachev, D. I. Salakhov, and S. V. Shaposhnikov, “The Fokker–Planck–Kolmogorov equation with nonlinear terms of local and nonlocal type”, St. Petersburg Math. J., 35:5 (2024), 749–767
34.
V. I. Bogachev and O. G. Smolyanov, Topological vector spaces and their applications, Springer Monogr. Math., Springer, Cham, 2017, x+456 pp.
35.
F. Bolley, I. Gentil, and A. Guillin, “Convergence to equilibrium in Wasserstein distance for Fokker–Planck equations”, J. Funct. Anal., 263:8 (2012), 2430–2457
36.
F. Bolley, I. Gentil, and A. Guillin, “Uniform convergence to equilibrium for granular media”, Arch. Ration. Mech. Anal., 208:2 (2013), 429–445
37.
F. Bouchut, “Existence and uniqueness of a global smooth solution for the Vlasov–Poisson–Fokker–Planck
system in three dimensions”, J. Funct. Anal., 111:1 (1993), 239–258
38.
O. A. Butkovsky, “On ergodic properties of nonlinear Markov chains and stochastic McKean–Vlasov equations”, Theory Probab. Appl., 58:4 (2014), 661–674
39.
J. A. Cañizo, J. A. Carrillo, P. Laurençot, and J. Rosado, “The Fokker–Planck equation for bosons in 2D: well-posedness and asymptotic behavior”, Nonlinear Anal., 137 (2016), 291–305
40.
P. Cardaliaguet, F. Delarue, J.-M. Lasry, and P.-L. Lions, The master equation and the convergence problem in mean field games, Ann. of Math. Stud., 201, Princeton Univ. Press, Princeton, NJ, 2019, x+212 pp.
41.
R. Carmona and F. Delarue, Probabilistic theory of mean field games with applications, v. I, Probab. Theory Stoch. Model., 83, Mean field FBSDEs, control, and games, Springer, Cham, 2018, xxv+713 pp. ; v. II, 84, Mean field games with common noise and master equations, xxiv+697 pp.
42.
J. A. Carrillo, A. Clini, and S. Solem, “The mean field limit of stochastic differential equation systems modeling grid cells”, SIAM J. Math. Anal., 55:4 (2023), 3602–3634
43.
J. A. Carrillo, M. DiFrancesco, A. Figalli, T. Laurent, and D. Slepčev, “Global-in-time weak measure solutions and finite-time aggregation for non-local interaction equations”, Duke Math. J., 156:2 (2011), 229–271
44.
J. A. Carrillo, R. Duan, and A. Moussa, “Global classical solutions close to equilibrium to the Vlasov–Fokker–Planck–Euler system”, Kinet. Relat. Models, 4:1 (2011), 227–258
45.
J. A. Carrillo, D. Gómez-Castro, and J. L. Vázquez, “Infinite-time concentration in aggregation–diffusion equations with a given potential”, J. Math. Pures Appl. (9), 157 (2022), 346–398
46.
J. A. Carrillo, K. Hopf, and J. L. Rodrigo, “On the singularity formation and relaxation to equilibrium in 1D Fokker–Planck model with superlinear
drift”, Adv. Math., 360 (2020), 106883, 66 pp.
47.
J. A. Carrillo, P. Laurençot, and J. Rosado, “Fermi–Dirac–Fokker–Planck equation: well-posedness & long-time asymptotics”, J. Differential Equations, 247:8 (2009), 2209–2234
48.
J. A. Carrillo, S. Lisini, G. Savaré, and D. Slepčev, “Nonlinear mobility continuity equations and generalized displacement convexity”, J. Funct. Anal., 258:4 (2010), 1273–1309
49.
J. A. Carrillo, J. Rosado, and F. Salvarani, “1D nonlinear Fokker–Planck equations for fermions and bosons”, Appl. Math. Lett., 21:2 (2008), 148–154
50.
P.-E. Chaudru de Raynal and N. Frikha, “Well-posedness for some non-linear SDEs and related PDE on the Wasserstein space”, J. Math. Pures Appl. (9), 159 (2022), 1–167
51.
M. Coghi and B. Gess, “Stochastic nonlinear Fokker–Planck equations”, Nonlinear Anal., 187 (2019), 259–278
52.
M. Colombo, G. Crippa, M. Graffe, and L. V. Spinolo, “Recent results on the singular local limit for nonlocal conservation laws”, Hyperbolic problems: theory, numerics, applications, AIMS Ser. Appl. Math., 10, Am. Inst. Math. Sci. (AIMS), Springfield, MO, 2020, 369–376
53.
M. Dieckmann, “A restricted superposition principle for (non-)linear Fokker–Planck–Kolmogorov
equations on Hilbert spaces”, J. Evol. Equ., 22:2 (2022), 55, 28 pp.
54.
R. L. Dobrushin, “Vlasov equations”, Funct. Anal. Appl., 13:2 (1979), 115–123
55.
Hongjie Dong, L. Escauriaza, and Seick Kim, “On C^1, C^2, and weak type-(1,1) estimates for linear elliptic operators. II”, Math. Ann., 370:1-2 (2018), 447–489
56.
Hongjie Dong, L. Escauriaza, and Seick Kim, “On C^{1/2,1}, C^{1,2}, and C^{0,0} estimates for linear parabolic operators”, J. Evol. Equ., 21:4 (2021), 4641–4702
57.
Hongjie Dong and Seick Kim, “On C^1, C^2, and weak type-(1, 1) estimates for linear elliptic operators”, Comm. Partial Differential Equations, 42:3 (2017), 417–435
58.
A. Eberle, “Reflection couplings and contraction rates for diffusions”, Probab. Theory Related Fields, 166:3-4 (2016), 851–886
59.
A. Eberle, A. Guillin, and R. Zimmer, “Quantitative Harris-type theorems for diffusions and McKean–Vlasov processes”, Trans. Amer. Math. Soc., 371:10 (2019), 7135–7173
60.
M. Fathi and M. Mikulincer, “Stability estimates for invariant measures of diffusion processes, with applications to stability of
moment measures and Stein kernels”, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 23:3 (2022), 1417–1445
61.
F. Flandoli, M. Leocata, and C. Ricci, “The Navier–Stokes–Vlasov–Fokker–Planck system as a scaling limit of particles in a fluid”, J. Math. Fluid Mech., 23:2 (2021), 40, 39 pp.
62.
T. D. Frank, Nonlinear Fokker–Planck equations. Fundamentals and applications, Springer Ser. Synergetics, Springer-Verlag, Berlin, 2005, xii+404 pp.
63.
T. D. Frank, “Linear and nonlinear Fokker–Planck equations”, Synergetics, Encycl. Complex. Syst. Sci., Springer, New York, 2020, 149–182
64.
T. Funaki, “A certain class of diffusion processes associated with nonlinear parabolic equations”, Z. Wahrsch. Verw. Gebiete, 67:3 (1984), 331–348
65.
G. Furioli, A. Pulvirenti, E. Terraneo, and G. Toscani, “Fokker–Planck equations in the modeling of socio-economic phenomena”, Math. Models Methods Appl. Sci., 27:1 (2017), 115–158
66.
S. Grube, “Strong solutions to McKean–Vlasov SDEs with coefficients of Nemytskii type: the time-dependent case”, J. Evol. Equ., 24:2 (2024), 37, 14 pp.
67.
A. Guillin, P. Le Bris, and P. Monmarché, “Convergence rates for the Vlasov–Fokker–Planck equation and uniform in time propagation of
chaos in non convex cases”, Electron. J. Probab., 27 (2022), 124, 44 pp.
68.
W. R. P. Hammersley, D. Šiška, and Ł. Szpruch, “McKean–Vlasov SDEs under measure dependent Lyapunov conditions”, Ann. Inst. Henri Poincaré Probab. Stat., 57:2 (2021), 1032–1057
69.
K. Hopf, “Singularities in L^1-supercritical Fokker–Planck equations: a qualitative analysis”, Ann. Inst. H. Poincaré C Anal. Non Linéaire, 41:2 (2024), 357–403
Xing Huang, M. Röckner, and Feng-Yu Wang, “Nonlinear Fokker–Planck equations for probability measures on path space and path-distribution dependent
SDEs”, Discrete Contin. Dyn. Syst., 39:6 (2019), 3017–3035
72.
Xing Huang and Feng-Yu Wang, “Singular McKean–Vlasov (reflecting) SDEs with distribution dependent noise”, J. Math. Anal. Appl., 514:1 (2022), 126301, 21 pp.
73.
Sukjung Hwang and Seick Kim, “Green's function for second order elliptic equations in non-divergence form”, Potential Anal., 52:1 (2020), 27–39
74.
E. Issoglio and F. Russo, “McKean SDEs with singular coefficients”, Ann. Inst. Henri Poincaré Probab. Stat., 59:3 (2023), 1530–1548
75.
Min Ji, Zhongwei Shen, and Yingfei Yi, “Convergence to equilibrium in Fokker–Planck equations”, J. Dynam. Differential Equations, 31:3 (2019), 1591–1615
76.
A. Jüngel, Entropy methods for diffusive partial differential equations, SpringerBriefs Math., Springer, Cham, 2016, viii+139 pp.
77.
M. Kac, “Foundations of kinetic theory”, Proceedings of the third Berkeley symposium on mathematical statistics and probability, 1954–1955, v. 3, Univ. California Press, Berkeley–Los Angeles, CA, 1956, 171–197
78.
E. F. Keller and L. A. Segel, “Initiation of slime mold aggregation viewed as an instability”, J. Theoret. Biol., 26:3 (1970), 399–415
79.
A. Kiselev, F. Nazarov, L. Ryzhik, and Yao Yao, “Chemotaxis and reactions in biology”, J. Eur. Math. Soc. (JEMS), 25:7 (2023), 2641–2696
80.
V. Kolokoltsov, Differential equations on measures and functional spaces, Birkhäuser Adv. Texts Basler Lehrbücher, Birkhäuser/Springer, Cham, 2019, xvi+525 pp.
81.
S. Kondratyev and D. Vorotnikov, “Nonlinear Fokker–Planck equations with reaction as gradient flows of the free energy”, J. Funct. Anal., 278:2 (2020), 108310, 40 pp.
82.
A. A. Kon'kov, “Stabilization of solutions of the nonlinear Fokker–Planck equation”, J. Math. Sci. (N. Y.), 197:3 (2014), 358–366
83.
V. V. Kozlov, “The generalized Vlasov kinetic equation”, Russian Math. Surveys, 63:4 (2008), 691–726
84.
V. V. Kozlov, “The Vlasov kinetic equation, dynamics of continuum and turbulence”, Regul. Chaotic Dyn., 16:6 (2011), 602–622
85.
Hailiang Li and G. Toscani, “Long-time asymptotics of kinetic models of granular flows”, Arch. Ration. Mech. Anal., 172:3 (2004), 407–428
86.
Jie Liao, Qianrong Wang, and Xiongfeng Yang, “Global existence and decay rates of the solutions near Maxwellian for non-linear Fokker–Planck equations”, J. Stat. Phys., 173:1 (2018), 222–241
87.
S. Lisini and A. Marigonda, “On a class of modified Wasserstein distances induced by concave mobility functions defined on bounded
intervals”, Manuscripta Math., 133:1-2 (2010), 197–224
88.
O. A. Manita, “Nonlinear Fokker–Planck–Kolmogorov equations in Hilbert spaces”, J. Math. Sci. (N. Y.), 216:1 (2016), 120–135
89.
O. A. Manita, “Estimates for transportation costs along solutions to Fokker–Planck–Kolmogorov equations
with dissipative drifts”, Atti Accad. Naz. Lincei Rend. Lincei Mat. Appl., 28:3 (2017), 601–618
90.
O. A. Manita, M. S. Romanov, and S. V. Shaposhnikov, “On uniqueness of solutions to nonlinear Fokker–Planck–Kolmogorov equations”, Nonlinear Anal., 128 (2015), 199–226
91.
O. A. Manita, M. S. Romanov, and S. V. Shaposhnikov, “Estimates of distances between solutions of Fokker–Planck–Kolmogorov equations with partially
degenerate diffusion matrices”, Theory Stoch. Process., 23:2 (2018), 41–54
92.
O. A. Manita and S. V. Shaposhnikov, “Nonlinear parabolic equations for measures”, St. Petersburg Math. J., 25:1 (2014), 43–62
93.
H. P. McKean, Jr., “A class of Markov processes associated with nonlinear parabolic equations”, Proc. Nat. Acad. Sci. U.S.A., 56:6 (1966), 1907–1911
94.
H. P. McKean, Jr., “Propagation of chaos for a class of non-linear parabolic equations”, Stochastic differential equations, Lecture Series in Differential Equations, Session 7, Air Force Office of Scientific Research, Office of Aerospace Research, United States Air Force, Arlington, VA, 1967, 41–57
95.
S. Mehri and W. Stannat, “Weak solutions to Vlasov–McKean equations under Lyapunov-type conditions”, Stoch. Dyn., 19:6 (2019), 1950042, 23 pp.
96.
Y. Mishura and A. Veretennikov, “Existence and uniqueness theorems for solutions of McKean–Vlasov stochastic equations”, Theory Probab. Math. Statist., 103 (2020), 59–101
97.
È. Mitidieri and S. I. Pokhozhaev, “A priori estimates and blow-up of solutions to nonlinear partial differential equations and inequalities”, Proc. Steklov Inst. Math., 234 (2001), 1–362
98.
È. Mitidieri and S. I. Pokhozhaev, “Liouville theorems for some classes of nonlinear nonlocal problems”, Proc. Steklov Inst. Math., 248:1 (2005), 158–178
99.
A. Mogilner and L. Edelstein-Keshet, “A non-local model for a swarm”, J. Math. Biol., 38:6 (1999), 534–570
100.
A. Okubo and S. A. Levin, Diffusion and ecological problems: modern perspectives, Interdiscip. Appl. Math., 14, 2nd ed., Springer-Verlag, New York, 2001, xx+467 pp.
101.
C. Olivera, A. Richard, and M. Tomašević, “Quantitative particle approximation of nonlinear Fokker–Planck equations with singular kernel”, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 24:2 (2023), 691–749
102.
R. Precup and P. Rubbioni, “Stationary solutions of Fokker–Planck equations with nonlinear reaction terms in bounded domains”, Potential Anal., 57:2 (2022), 181–199
103.
M. Rehmeier, “Flow selections for (nonlinear) Fokker–Planck–Kolmogorov equations”, J. Differential Equations, 328 (2022), 105–132
104.
M. Rehmeier, “Linearization and a superposition principle for deterministic and stochastic nonlinear
Fokker–Planck–Kolmogorov equations”, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 24:3 (2023), 1705–1739
105.
Panpan Ren, M. Röckner, and Feng-Yu Wang, “Linearization of nonlinear Fokker–Planck equations and applications”, J. Differential Equations, 322 (2022), 1–37
106.
Zhenjie Ren, Xiaolu Tan, N. Touzi, and Junjian Yang, “Entropic optimal planning for path-dependent mean field games”, SIAM J. Control Optim., 61:3 (2023), 1415–1437
107.
K. Schuh, “Global contractivity for Langevin dynamics with distribution-dependent forces and uniform in time
propagation of chaos”, Ann. Inst. Henri Poincaré Probab. Stat., 60:2 (2024), 753–789
108.
S. V. Shaposhnikov, “Nonlinear Fokker–Planck–Kolmogorov equations for measures”, Stochastic partial differential equations and related fields, Springer Proc. Math. Stat., 229, Springer, Cham, 2018, 367–379
109.
Zheng Sun, J. A. Carrillo, and Chi-Wang Shu, “A discontinuous Galerkin method for nonlinear parabolic equations and gradient flow problems with
interaction potentials”, J. Comput. Phys., 352:4 (2018), 76–104
110.
L. G. Tonoyan, “Nonlinear elliptic equations for measures”, Dokl. Math., 84:1 (2011), 558–561
111.
G. Toscani, “Finite time blow up in Kaniadakis–Quarati model of Bose–Einstein particles”, Comm. Partial Differential Equations, 37:1 (2012), 77–87
112.
A. Tosin and M. Zanella, “Kinetic-controlled hydrodynamics for traffic models with driver-assist vehicles”, Multiscale Model. Simul., 17:2 (2019), 716–749
113.
Alvin Tse, “Higher order regularity of nonlinear Fokker–Planck PDEs with respect to the measure component”, J. Math. Pures Appl. (9), 150 (2021), 134–180
114.
V. Vedenyapin, A. Sinitsyn, and E. Dulov, Kinetic Boltzmann, Vlasov and related equations, Elsevier, Inc., Amsterdam, 2011, xvi+304 pp.
115.
A. Yu. Veretennikov, “On ergodic measures for McKean–Vlasov stochastic equations”, Monte Carlo and quasi-Monte Carlo methods 2004, Springer-Verlag, Berlin, 2006, 471–486