Publication Cover
Stochastics
An International Journal of Probability and Stochastic Processes
Latest Articles
243
Views
0
CrossRef citations to date
0
Altmetric
Research Article

The Donsker delta function and local time for McKean–Vlasov processes and applications

& ORCID Icon
Received 28 Feb 2023, Accepted 16 Nov 2023, Published online: 06 Dec 2023

Abstract

The purpose of this paper is to establish a stochastic differential equation for the Donsker delta measure of the solution of a McKean–Vlasov (mean-field) stochastic differential equation.

If the Donsker delta measure is absolutely continuous with respect to Lebesgue measure, then its Radon–Nikodym derivative is called the Donsker delta function. In that case it can be proved that the local time of such a process is simply the integral with respect to time of the Donsker delta function. Therefore we also get an equation for the local time of such a process.

For some particular McKean–Vlasov processes, we find explicit expressions for their Donsker delta functions and hence for their local times.

MSC 2020:

1. Introduction

The Donsker delta function of a random variable or a stochastic process arises in many studies, including quantum mechanical particles on a circle [Citation7], financial markets with insider trading as in [Citation10] and in [Citation3] for financial markets with singular drift. It has also been used as a tool to determine explicit formulae for replicating portfolios in complete and incomplete markets [Citation9].

Moreover, the Donsker delta function is also of interest because it can be regarded as a time derivative of the local time. Therefore, explicit expressions for the Donsker delta function lead to explicit formulae of the local time.

For example, if we let B be a Brownian motion defined on a filtered probability space (Ω,F,F={Ft}t0,P), then the Donsker delta function δB(t)(x) of a Brownian motion B at the point x can be regarded as the time derivative of the local time Lt(x) of B. More precisely, we have Lt(x)=0tδB(s)(x)ds.Such an integral exists as an element of the Hida space (S) of stochastic distributions, see Section 2.3.

In [Citation6], the authors use white noise theory to obtain an explicit solution formula for a general stochastic differential equation (SDE), and this is used to find an expression for the Donsker delta function for the solution of an SDE. Subsequently this was also extended to SDEs driven by Lévy noise in [Citation8].

The main result of the current paper is that the Donsker delta measure of a McKean–Vlasov process (see below) always satisfies a certain Fokker–Planck type SPDE in the sense of distributions. Moreover, we use this to find explicit formulae for the Donsker delta functions for McKean–Vlasov processes, and hence their local times, in specific cases.

Let X(t)=XtR be the solution of a McKean–Vlasov SDE, i.e. a mean-field stochastic differential equation, of the form (using matrix notation), dX(t)=α(t,X(t),μtX)dt+β(t,X(t),μtX)dB(t);X(0)=ZR.We call X a McKean–Vlasov process.

Here the σ-algebra F={Ft}t0 denotes the filtration generated by Z and B(), Z is a random variable which is independent of the σ-algebra generated by B() and such that E[|Z|2]<.

Definition 1.1

Define μtX(dx)=μtX(dx,ω) to be regular conditional distribution of X(t) given Ft generated by the Brownian motion B. This means that μtX(dx,ω) is a Borel probability measure on R for all t[0,T],ωΩ and Rg(x)μtX(dx,ω)=E[g(X(t))|Ft](ω),for all functions g such that E[|g(X(t))|]<.

Since we consider only a one-dimensional Brownian motion B(t)R, we will show that the regular conditional distribution of X(t) given the filtration {Ft}t0 can be identified with the Donsker delta measure in the sense of distribution. See details in Section 3.1

2. Preliminaries

In this section, we review some basic notions and results that will be used throughout this work.

2.1. Radon measures

A Radon measure on Rd is a Borel measure which is finite on compact sets, outer regular on all Borel sets and inner regular on all open sets. In particular, all Borel probability measures on Rd are Radon measures.

In the following, we let

  • M0 be the set of deterministic Radon measures.

  • C0(Rd) be the uniform closure of the space Cc(Rd) of continuous functions with compact support.

If we equip M0 with the total variation norm ||μ||:=|μ|(Rd), then M0 becomes a Banach space, and it is the dual of C0(Rd). See Chapter 7 in Folland [Citation5] for more information.

If μM0 is a finite measure, we define (1) μ^(y):=F[μ](y):=Rdeixyμ(dx);yRd(1) to be the Fourier transform of μ at y.

In particular, if μ(dx) is absolutely continuous with respect to Lebesgue measure dx with Radon–Nikodym derivative m(x)=μ(dx)dx, so that μ(dx)=m(x)dx with mL1(Rd), we define the Fourier transform of m at y, denoted by m^(y) or F[m](y), by F[m](y)=m^(y)=Rdeixym(x)dx;yRd.We let M denote the set of all random measures μ(dx,ω);ωΩ such that μ(dx,ω)M0 for each given ωΩ.

2.2. The Schwartz space of tempered distributions

We recall now some notions from white noise analysis.

  • S=S(Rd) be the Schwartz space of rapidly decreasing smooth real functions on Rd. It is a Fréchet space with respect to the family of seminorms: fk,α:=supxRd{(1+|x|k)|αf(x)|},where k=0,1,, α=(α1,,αd) is a multi-index with αj=0,1, (j=1,,d) and αf:=|α|fx1α1xdαdfor|α|=α1++αd.

  • S=S(Rd) is the space of tempered distributions. It is the dual of S.

2.3. The Hida space (S) of stochastic distributions

We restrict ourselves to the white noise probability space (Ω=S,F=B,P), where B is the Borel σ-algebra and the probability P is the probability measure on S defined in virtue of the Bochner–Minlos–Sazonov theorem.

Let J denote the set of all finite multi-indices α=(α1,α2,,αm), m=1,2,, of non-negative integers αi. (2) (2N)α=j=1m(2j)αj=(21)α1(22)α2(23)α3(2m)αm.(2) If α=(α1,α2,)J, we put (3) Hα(ω):=j=1mhαj(θj(ω))=hα1(θ1)hα2(θ2)hαm(θm),ωΩ.(3) The family {Hα}αJ constitutes an orthogonal basis of L2(P).

  • ((S)k)kR is the Hilbert space consisting of all f=αJcαHαL2(P) such that fk2:=αJα!cα2(2N)αk<, for numbers cαR.

  • The space (S)=kR(S)k equipped with the projective topology is the Hida space of stochastic test functions.

  • ((S)k)kR is the Hilbert space consisting of all formal sums F=αJcαHα equipped with the norm Fk2:=αJα!cα2(2N)αk<.

  • The space (S)=kR(S)k equipped with the inductive topology is the Hida space of stochastic distributions. It can be regarded as the dual of (S).

2.4. The Donsker delta function

We now recall some basic definitions:

Definition 2.1

Let Y:ΩR be a random variable which also belongs to the Hida space (S) of stochastic distributions. Then a continuous function (4) δY():R(S)(4) is called a Donsker delta function of Y if it has the property that (5) Rg(y)δY(y)dy=g(Y)a.s.(5) for all (measurable) g:RR such that the integral converges in (S).

The Donsker delta function is related to the regular conditional distribution. The connection is the following: the regular conditional distribution with respect to the σ-algebra F of a given real random variable Y, denoted by μY(dy)=μY(dy,ω);ωΩ, is defined by the following properties:

  • For any Borel set ΛR, μY(Λ,) is a version of E[1YΛ|F].

  • For each fixed ωΩ, μY(dy,ω) is a probability measure on the Borel subsets of R.

It is well known that such a regular conditional distribution always exists. See, e.g. [Citation4, p. 79].

From the required properties of μY(dy,ω), we get the following formula: (6) Rf(y)μY(dy,ω)=E[f(Y)|F].(6)

Definition 2.2

We call μY(dy,ω) the Donsker delta measure of the random variable Y and denote it by δY(dy,ω).

Comparing this with the definition of the Donsker delta function, we obtain the following representation of the regular conditional distribution:

Lemma 2.3

Suppose μY(dy,ω) is absolutely continuous with respect to Lebesgue measure dy on R and that Y is measurable with respect to F. Then the Donsker delta function of Y, δY(y,ω), is the Radon–Nikodym derivative of μY(dy,ω) with respect to Lebesgue measure dy, i.e. (7) δY(y,ω)=μY(dy,ω)dy.(7)

We will prove in Theorem 3.3 that the Donsker delta function can be regarded as a stochastic distribution in S, satisfying a Fokker–Planck type SPDE in the sense of distributions. It can also be represented as an element of the Hida stochastic distribution space (S), and as such it can in some cases be expressed explicitly in terms of Wick calculus. For example, if Y(t)=B(t), we have (8) δB(t)(x)=(2πt)12exp((B(t)x)22t)(S),(8) where ⋄ denotes Wick multiplication and exp denotes Wick exponential. Note that even though the Donsker delta function can only be represented as a distribution, its conditional expectation can be a real-valued stochastic process. For example, for t<T we have (9) E[δB(T)(x)|Ft]=(2π(Tt))12exp[(B(t)x)22(Tt)].(9) For more examples, we refer to e.g. [Citation1] or [Citation9].

3. The Donsker delta equation for McKean–Vlasov processes

3.1. The general multidimensional Fokker–Planck equation

To explain the background for this section, let us recall the general multidimensional situation studied in [Citation2], where X(t)Rd is a McKean–Vlasov diffusion, of the form (using matrix notation), (10) dX(t)=b(t,X(t),μtX)dt+σ(t,X(t),μtX)dB(t),X(0)=Z,(10) where B is a multi-dimensional Brownian motion.

Here Z is a random variable which is independent of the σ-algebra generated by B() and such that E[|Z|2]<.Define the σ-algebra F={Ft}t0 to be the filtration generated by Z and B().

Let M denote the set of all Borel measures on Rd. We assume that the coefficients b(t,x,μ):[0,T]×Rd×MRd and σ(t,x,μ):[0,T]×Rd×MRd are bounded and F-predictable processes for all x,μ, and that b and σ are continuous with respect to t and x for all μ.

One can check that under some assumptions, such as Lipschitz and linear growth conditions, there exists a unique solution of Equation (Equation10).

Definition 3.1

Fix one of the Brownian motions, say B1=B1(t,ω), with filtration {Ft(1)}t0. We define μtX=μtX(dx,ω) to be regular conditional distribution of X(t) given Ft(1). This means that μtX(ω,dx) is a Borel probability measure on Rd for all t[0,T],ωΩ and (11) Rdg(x)μtX(dx,ω)=E[g(X(t))|Ft(1)](ω)(11) for all functions g such that E[|g(X(t))|]<.

The following version of the stochastic Fokker–Planck integro-differential equation for the conditional law for McKean–Vlasov jump diffusions was proved by Agram and Øksendal [Citation2]. For simplicity we consider only the case without jumps here.

Theorem 3.2

Conditional stochastic Fokker–Planck equation [Citation2]

Let X(t) be as in (Equation10) with d2 and let μtX:=μtX(dx,ω) be the regular conditional distribution of X(t) given Ft(1).

Then for a.a. ωΩ the conditional law μtXS and it satisfies the following SPDE (in the sense of distributions): (12) dμtX=A0μtXdt+A1μtXdB1(t),μ0=L(X(0)).(12) Here A0,A1 are the integro-differential operator and the differential operator which are given respectively by (13) A0μ=j=1dDj[bjμ]+12n,j=1dDn,j[(σσT)n,jμ](13) and (14) A1μ=j=1dDj[b1,jμ].(14) In the above Dj,Dn,j denote xj and 2xnxj respectively, in the sense of distributions, and Dj[bjμ]=xj[bj(t,x,μ)μtX(dx)]|μ=μtX, and similarly with the other terms.

3.2. The Fokker–Planck equation for the Donsker measure

In [Citation2], the theorem above was proved under the assumption that d2. However, the proof also works if d = 1 and Ft(1)=Ft. Note that in this case, since X(t) is Ft-measurable, the identity (Equation11) states that (15) Rdg(x)μtX(dx,ω)=g(X(t))(15) for all functions g such that Rd|g(x)|μtX(dx,ω)<.

In particular, if we choose d = 1 in the above we get that the conditional law coincides with the Donsker measure, i.e. (16) μtX(x,ω)=δX(t)(dx,ω).(16) Therefore we get the following Fokker–Planck equation for the Donsker measure:

Theorem 3.3

Assume that X(t) is as in (Equation10), but with d = 1.

Then the Donsker delta measure μtX=δX(t)(dx,ω) satisfies the following equation (in the sense of distribution): (17) dμtX={D[b(t,x,μ)μtX]|μ=μtX+12D2[σ2(t,x,μ)μtX]|μ=μtX}dtD[σ(t,x,μ)μtX]|μ=μtXdB(t);μ0=L(X(0)),(17) where D=x and D2=2x2.

4. Local time

In this section, we first recall the definition of local time of a stochastic process Y():

Definition 4.1

The local time Lt(y) of Y() at the point y and at time t is defined by Lt(y)=limϵ012ϵλ({s[0,t];Y(s)(yϵ,y+ϵ)}),where λ denotes Lebesgue measure on R and the limit is in L2(P).

In the white noise context, the local time can be represented as the integral of the Donsker delta function. More precisely, we have the following result:

Theorem 4.2

The local time Lt(x) of X at the point x and the time t is given by (18) Lt(x)=0tδX(s)(x)ds,(18) where the integration takes place in (S) (or in S for each ω).

Proof.

For completeness, we give the proof.

By definition of the local time and the Donsker delta function, we have Lt(z)=limϵ00tχ(zϵ,z+ϵ)(Y(s))ds=limϵ00t(Rχ(zϵ,z+ϵ)(y)δY(s)(y)dy)ds=limϵ0Rχ(zϵ,z+ϵ)(y)(0tδY(s)(y)ds)dy=0tδY(s)(z)ds,because the function yδY(s)(y) is continuous in (S) (and in S).

Remark 4.3

Note that even though we in general can only say that δX(t)(x)(S), Lt(x) usually exists as a real-valued stochastic process.

5. Explicit solutions

In this section, we find explicitly the Donsker delta function for some particular McKean–Vlasov processes and accordingly their local time.

Suppose that μtX is absolutely continuous, i.e. (19) μtX(dx)=mX(t,x)dx.(19) Then (Equation10) gets the form (20) dX(t)=b(t,X(t),mtX)dt+σ(t,X(t),mtX)dB(t);X(0)=Z,(20) where mtX(x)=mX(t,x) and (Equation17) becomes a stochastic partial differential equation (SPDE), as follows:

Theorem 5.1

Suppose (Equation19) holds. Then the Donsker delta function mX(t,x)=δX(t)(x) is the solution in (S) of the following SPDE: (21) dtmX(t,x)={x[b(t,x,m)mX(t,x)]+122x2[σ2(t,x,m)mX(t,x)]}dt(21) (22) x[σ(t,x,m)mX(t,x)]dB(t);t0,m(0,x)=xL(X(0)).(22)

5.1. Brownian motion

Consider the special case when X(t)=B(t);B(0)=Z. Then b = 0 and σ=1 and Equation (Equation17) becomes (23) mXt(t,x)=122mXx2(t,x)+mXx(t,x)B˙(t);t0,(23) (24) m(0,x)=xL(X(0)).(24) We can easily verify by Wick calculus that a solution in (S) of Equation (Equation23) is (25) δB(t)(x)=(2πt)12exp((B(t)x)22t),(25) which is in agreement with (Equation8). The details are as follows:

Try mX(t,x)=12πtexp[(xB(t))22t].Then mXt(t,x)=12t3/22πexp[(xB(t))22t]+12πtexp[(xB(t))22t](xB(t)t)B˙(t)+12πtexp[(xB(t))22t](xB(t))22t2,and mXx=12πtexp[(xB(t))22t](xB(t)t),and 2mXx2(t,x)=12πtexp[(xB(t))22t](xB(t)t)2+12πtexp[(xB(t))22t](1t).Collecting the terms we see that mX(t,x)=12πtexp((xB(t))22t),satisfies the Fokker–Planck equation (Equation23) for the conditional law of B(t).

From white noise theory, we know that

  • E[XY]=E[X]E[Y]

  • E[expY]:=E[n=01n!Yn]=n=01n!E[Yn]=n=01n!E[Y]n=exp(E[Y])

for all random variables X, Y with a finite expectation (independent or not). From this we see that (26) E[δB(t)(x)]=12πtexp(E[(xB(t))22t])=12πtexp((xE[Z])22t).(26) In particular, if X(0)=Z=z (constant) R a.e., then (27) E[δB(t)(x)]=12πtexp((xz)22t),(27) which has a singularity at x = z.

5.2. Coefficients not depending on x

The next result shows that, under some conditions, the Donsker delta function can be an ordinary function if the initial value X(0) has a density.

Theorem 5.2

Assume that X(t) is the solution of the following McKean–Vlasov equation: (28) dX(t)=α(t,μtX)dt+β(t,μtX)dB(t);X(0)=Z,(28) where the coefficients α(t,x,μ)=α(t,μ) and β(t,x,μ)=β(t,μ) do not depend on x. Suppose that X(0)=Z is a random variable (independent of B) with density (29) h(z)=zL(Z)(z);zR.(29)

(1)

Define (30) Yt(x)=h(x0tα(s,μsX)ds0tβ(s,μsX)dB(s))=h(K(t,x)),(30) where K(t,x)=x0tα(s,μsX)ds0tβ(s,μsX)dB(s).Then Y(t,x) is the Donsker delta function of X(t).

(2)

The solution X(t) of (Equation28) is given by (31) X(t)=RxYt(x)dx.(31)

Proof.

  1. We show that Y(t,x) satisfies Equation (Equation21).

    By the Ito formula, we have (32) dtY(t,x)=h(K(t,x))dtK(t,x)+12h(K(t,x))β2(t)dt={α(t,μtX)h(K(t,x))+12β2(t,μtX)h(K(t,x))}dtβ(t,μtX)h(K(t,x))dB(t).(32) Since (33) h(K(t,x))=ddzh(z)z=K(t,x)=xY(t,x),(33) we see that Equation (Equation32) can be written as (34) dtY(t,x)=[α(t,μtX)xY(t,x)+12β2(t,μtX)2x2Y(t,x)]dtβ(t,μtX)xY(t,x)dB(t),(34) which is the same as Equation (Equation21).

    Since Y(0,x)=h(x)=m(0,x), we conclude by uniqueness that Y(t,x)=m(t,x) for all t.

  2. This follows from the definition of the Donsker delta function.

5.2.1. Constant coefficients

As a special case of the case above, suppose that (35) dX(t)=αdt+βdB(t),X(0)=Z,(35) where α and β are constants. Then by Theorem 5.2, the Donsker delta function is (36) δX(t)(x)=h(xαtβB(t)).(36)

5.3. Mean-field geometric Brownian motion

Suppose that X(t) is a McKean–Vlasov process of the form (37) dXt=α(t,μtX)Xtdt+β(t,μtX)XtdBt;X0=Z>0.(37) We call this a mean-field geometric Brownian motion. For such processes, we have:

Theorem 5.3

(i)

The Donsker delta function mtX(x) for the mean-field geometric Brownian motion X(t) is (38) mtX(x)=δXt(x)=1xH(lnx0tα(s,μsX)ds0tβ(s,μsX)dB(s)),(38) where (39) H(z)=zL(lnZ)(z);zR.(39)

(ii)

The solution X(t) of the mean-field geometric Brownian motion equation (Equation37) can be written as (40) X(t)=0H(lnx0tα(s,μsX)ds0tβ(s,μsX)dB(s))dx=ReuH(u0tα(s,μsX)ds0tβ(s,μsX)dB(s))du.(40)

Proof.

  1. The corresponding Fokker–Planck equation for the Donsker delta function

    mt=mtX(x)=δX(t)(x) is (41) dmt(x)={x[α(t,m)xmt(x)]+12β2(t,m)2x2[x2mt(x)]}dtβ(t,m)x[xmt(x)]dBt={α(t,m)mt(x)α(t,m)xmt(x)+12β2(t,m)[2mt(x)+4xmt(x)+x2mt(x)]}dtβ(t,m)[mt(x)+xmt(x)]dBt;m0(x)=xL(Z)(x).(41) This is a stochastic partial differential equation in mt(x). It seems difficult to find directly an explicit solution of this equation. However, we can find the solution mt(x)=δX(t)(x) by proceeding as follows:

    The solution of (Equation37) is Xt=Zexp(0tβ(s,ms)dB(s)+0t{α(s,ms)12β2(s,ms)}ds)=exp(Yt),where Yt=lnZ+0tβ(s,ms)dB(s)+0t{α(s,ms)12β2(s,ms)}ds.By Theorem 5.2, we know that δYt(x)=H(x0tα(s,μs)ds0tβ(s,μs)dB(s)),where H(z)=zL(lnZ)(z).By definition  we have R+g(y)δYt(y)dy=g(Yt).With g(y)=exp(y), this gives Rexp(y)δYt(y)dy=exp(Yt)=Xt.Hence, substituting exp(y)=x, Xt=Rexp(y)δYt(y)dy=R+xδYt(ln(x))dxx.From this we deduce that (42) mt(x)=δXt(x)=δYt(ln(x))x=1xH(lnx0tα(s,μs)ds0tβ(s,μs)dB(s))(42) is the Donsker delta function of Xt.

  2. This part follows by the definition of the Donsker delta function.

5.4. An example related to the Burgers equation

Suppose the McKean–Vlasov equation has the form (43) dX(t)=αm(t,X(t))dt+βdB(t);X(0)=Z,(43) where m(t,x)=xμX(t,x)=xL(X(t))(x)L2([0,T]×R).

Then the corresponding FP equation for the Donsker function m(t,x) is (44) dm(t,x)={αx(m2(t,x))+12β22x2m(t,x)}dtβxm(t,x)dB(t),(44) (45) m(0,x)=h(x)=xZ(x).(45) This is a stochastic Burgers equation. It is well known that by using the Cole–Hopf transformation the equation can be transformed into the classical heat equation. The details are as follows: if we introduce a new function ψ=ψ(t,x) such that (46) m:=ψx:=xψ,(46) then we see that the Burgers equation (Equation44) becomes the following equation in ψ: (47) (ψx)t=2αψx(ψx)x+12β2(ψx)xxβψxxB˙(t).(47) Integrating with respect to x this gives (48) ψt=α(ψx)2+12β2ψxxβψxB˙(t).(48) Now define the function φ=φ(t,x) by (49) ψ=γlnφ,(49) for some constant γ. Then in terms of φ the above equation gets the form (50) γφtφ=αγ2(φxφ)2+12γβ2(φxφ)xβγφxφB˙(t)=αγ2(φxφ)2+12γβ2(φφxx(φx)2φ2)xβγφxφB˙(t)=αγ2(φxφ)2+12γβ2(φxxφ)12γβ2(φxφ)2βγφxφB˙(t).(50) This simplifies to (51) φt=(γα+12β2)(φxφ)2+12β2φxxβφxB˙(t).(51) If we choose (52) γ=β22α,(52) the equation for φ reduces to the (linear) stochastic heat equation (53) φt=12β2φxxβφxB˙(t),(53) (54) φ(0,x)=k(x)(tobedetermined),(54) or, using Ito differential notation, (55) dφ(t,x)=12β2φxx(t,x)dtβφx(t,x)dB(t);t0,(55) (56) φ(0,x)=k(x).(56) To find an expression for the solution of (Equation53), define an auxiliary process R(t)=R(x)(t) by (57) R(t)=xβB(t)+βB~(t);t0,(57) where B~ is an auxiliary Brownian motion with law P~ and independent of B. Then by the Feynman–Kac formula (58) φ(t,x):=E~[k(R(x)(t))]=E~[k(xβB(t)+βB~(t))],(58) where E~ denotes expectation with respect to P~ and k(z)=φ(0,z), solves Equation (Equation53). Going back to m, we get (59) m(t,x)=ψx(t,x)=γxlnφ(t,x))=γφx(t,x)φ(t,x)=γE~[kx(xβB(t)+βB~(t))]E~[k(xβB(t)+βB~(t))].(59) In particular, setting t = 0 we get (60) h(x):=m(0,x)=γE~[kx(x)]E~[k(x)]=γkx(x)k(x),(60) from which we deduce that (61) k(x)=exp(1γ0xh(y)dy).(61) We summarize what we have proved as follows:

Theorem 5.4

(1)

The Donsker delta function m(t,x)=δX(t)(x) for the solution X(t) of the McKean–Vlasov equation (Equation43) is given by (62) m(t,x)=γφx(t,x)φ(t,x)=γE~[kx(xβB(t)+βB~(t))]E~[k(xβB(t)+βB~(t))],(62) where (63) k(x)=exp(1γ0xm(0,y)dy)=exp(1γ0xL(Z)(y)dy)(63) and (64) γ=β22α.(64)

(2)

The solution X(t) of (Equation43) is given by (65) X(t)=Rxm(t,x)dx(65) with m(t,x) as in part 1.

5.5. A solution approach based on Laplace and Fourier transforms

Consider the Fokker–Planck equation, with μ=μtX, (66) dμt={D[αμ]+12D2[β2μ]}dtD[βμ]dB(t);μ0=δx0,(66) for the McKean–Vlasov equation (Equation35). If α,β are constants, this becomes (67) dμt={αD[μ]+12β2D2[μ]}dtβD[μ]dB(t);μ0=δx0.(67) If dμt=m(t,x)dx, the equation can be written as (68) tm(t,x)=αxm(t,x)+12β22x2m(t,x)+βxm(t,x)B˚(t).(68) Let (69) f~(s)=Lf(s)=0sestf(t)dtdenotetheLaplacetransform(69) and (70) f^(y)=Ff(y)=Reixyf(x)dxdenotetheFouriertransform.(70) Then (71) L(tf(t))(s)=s(Lf)(s)f(0)(71) and (72) L(exp(bt))(s)=1sb(72) and (73) F[Dnw](y)=(iy)nF[w](y).(73) Hence, applying the Laplace and Fourier transform to (Equation68), we get sm~^(s,y)m^(0,x0)=iαym~^(s,y)+12β2(iy)2m~^(s,y)+βiy(m^(.,y)B˚(.))(s)or m^~(s,y)[s+iαy+12β2y2]=m^(0,x0)+βiy(m^(.,y)B˚(.)~)(s)or (74) m^~(s,y)=m^(0,x0)s+iαy+12β2y2+βiys+iαy+12β2y2(m^(.,y)B˚(.)~)(s)=m^(0,x0)s+iαy+12β2y2+βiyL(e(iαy12β2y2)t(s)m^(.,y)B˚(.)~)(s).(74) Put g(t)=e(iαy12β2y2)t and h(t)=m^(t,y)B˚(t)).

Taking inverse Laplace transform, we get m^(t,y)=m^(0,x0)exp((iαy12β2y2)t)+βiyL1(LgLh)(t,y)=m^(0,x0)exp((iαy12β2y2)t)+βiy(gh)(t,y),where (gh)(t)=0tg(s)h(ts)ds.Recall that (75) Reay22bydy=πaeb2a;a>0.(75) Hence F1(g)=12πReiαyt12β2y2t+iyxdy=12πRe12β2ty22yi(12αt12x)dy=12ππ12β2texp(i2(12αt+12x)212β2t)=12πβ2texp((αt+x)22β2t)=12πβ2texp(x22β2tαxβ2)exp(α2t2β2)=:k(t,x).Therefore g(t,x)=F[k(t,.)](y)] and (Equation75) can be written as m^(t,y)=m^(0,x0)F[k](t,y)+βiy0tF[k(ts,.)](y)F[m(s,y)B˚(s))]ds.Taking inverse Fourier transform we get, with k=ddxk(t,x) m(t,x)=m(0,x0)k)(t,x)+F1[β0tF[k(ts,)](y)F[m(s,y)B˚(s))]](t,x)=Rδx0(xy)k(t,y)dy+β0t(Rk(ts,xy)m(s,y)B˚(s)dsdy=k(t,xx0)+βR(0tk(ts,xy)m(s,y)dB(s))dy.We have proved the following:

Theorem 5.5

Suppose α and β are constants and that the Donsker delta measure is absolutely continuous with respect to Lebesgue measure. Then the Donsker delta function m(t,x)=δX(t)(x) of the corresponding McKean–Vlasov process is a solution in (S) of the following stochastic Volterra equation: m(t,x)=k(t,xx0)+βR(0tk(ts,xy)m(s,y)dB(s))dy,where k(t,z)=12πβ2texp(z22β2tαzβ2)exp(α2t2β2);k(u,z)=ddzk(u,z).

Remark 5.6

If α=0,β=1, we get k(t,z)=12πtexp(z22t)k(u,z)=zu12πuexp(z22u).For comparison, recall that the density of Brownian motion at t, x (when starting at x0) is p(t,x)=12πtexp((xx0)22t).

Acknowledgments

We are grateful to Frank Proske for helpful comments.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

Work supported by the Swedish Research Council grant (2020-04697).

References

  • K. Aase, B. Øksendal, and J. Ubøe, Using the Donsker delta function to compute hedging strategies, Potential Anal. 14(4) (2001), pp. 351–374.
  • N. Agram and B. Øksendal, Stochastic Fokker–Planck PIDE for conditional McKean–Vlasov jump diffusions and applications to optimal control. To appear in SICON. arXiv preprint arXiv:2110.02193v3, 2021.
  • N. Agram and B. Øksendal, A financial market with singular drift and no arbitrage, Math. Financ. Econ. 15(3) (2021), pp. 477–500.
  • B.L. Breiman, Probability, Addison-Wesley, 1968.
  • G.B. Folland, Real Analysis: Modern Techniques and Their Applications, Wiley, 1984.
  • A. Lanconelli and F. Proske, On explicit strong solution of Itô-SDE's and the Donsker delta function, Infin. Dimens. Anal. Quantum Probab. Relat. Top. 7(03) (2004), pp. 437–447.
  • A. Lascheck, P. Leukert, L. Streit, and W. Westerkamp, More about Donsker's delta function. arXiv preprint math-ph/0303011, 2003.
  • T. Meyer-Brandis and F. Proske, On the existence and explicit representability of strong solutions of Lévy noise driven SDEs with irregular coefficients, Commun. Math. Sci. 4(1) (2006), pp. 129–154.
  • G. Di Nunno, B. Øksendal, and F. Proske, (Eds.) Malliavin Calculus for Lévy Processes with Applications to Finance, Springer Berlin Heidelberg, 2009.
  • B. Øksendal and E. Engen Røse, A white noise approach to insider trading, in Let Us Use White Noise, 2017, pp. 191–203.