This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Scalar conservation laws with white noise initial data

Mehdi Ouaki Department of Statistics #3860
451 Evans Hall
University of California at Berkeley
Berkeley, CA 94720-3860
USA
mouaki@berkeley.edu
Abstract.

The statistical description of the scalar conservation law of the form ρt=H(ρ)x\rho_{t}=H(\rho)_{x} with H:H:\mathbb{R}\rightarrow\mathbb{R} a smooth convex function has been an object of interest when the initial profile ρ(,0)\rho(\cdot,0) is random. The special case when H(ρ)=ρ22H(\rho)=\frac{\rho^{2}}{2} (Burgers equation) has in particular received extensive interest in the past and is now understood for various random initial conditions. We prove in this paper a conjecture on the profile of the solution at any time t>0t>0 for a general class of Hamiltonians HH and show that it is a stationary piecewise-smooth Feller process. Along the way, we study the excursion process of the two-sided linear Brownian motion WW below any strictly convex function ϕ\phi with superlinear growth and derive a generalized Chernoff distribution of the random variable argmaxz(W(z)ϕ(z))\text{argmax}_{z\in\mathbb{R}}(W(z)-\phi(z)). Finally, when ρ(,0)\rho(\cdot,0) is a white noise derived from an abrupt Lévy process, we show that the structure of shocks of the solution is a.s discrete at any fixed time t>0t>0 under some mild assumptions on HH.

Key words and phrases:
scalar conservation law, white noise, random initial data, path decomposition, Chernoff distribution, abrupt process
2010 Mathematics Subject Classification:
60G51, 60J65, 60J60, 60J75, 35L65

1. Introduction

We are interested in the following conservation law problem

(1.1) {ρt=(H(ρ))x, for t>0,xρ(x,0)=ξ(x),x\left\{\begin{array}[]{ll}\rho_{t}=(H(\rho))_{x}~{}~{},\text{ for }~{}~{}t>0,x\in\mathbb{R}\\ \rho(x,0)=\xi(x)~{}~{},~{}~{}x\in\mathbb{R}\end{array}\right.

where HH is a C2C^{2} strictly convex function with superlinear growth at infinity and ξ\xi is a white noise. A question of interest is to describe the law of the process ρ(,t)\rho(\cdot,t) at any given time t>0t>0.

1.1. Background

There is a straightforward link between the scalar conservation law and the Hamilton-Jacobi PDE. Indeed, if one defines

u(x,t)=xρ(y,t)𝑑y\displaystyle u(x,t)=\int_{-\infty}^{x}\rho(y,t)dy

and the potential

U0(x)=xξ(y)𝑑y\displaystyle U_{0}(x)=\int_{-\infty}^{x}\xi(y)dy

then uu solves the PDE

(1.2) ut=H(ux)u_{t}=H(u_{x})

and is determined by the Hopf-Lax formula (see [6][Theorem 4, Chapter 3.3])

(1.3) u(x,t)=supy(U0(y)tL(yxt))u(x,t)=\sup_{y\in\mathbb{R}}\left(U_{0}(y)-tL\left(\frac{y-x}{t}\right)\right)

where LL is the Legendre transform of HH defined as

L(q)=maxp(qpH(p))\displaystyle L(q)=\max\limits_{p\in\mathbb{R}}\left(qp-H(p)\right)

The rightmost maximizer y(x,t)y(x,t) in the equation (1.3) is called the backward Lagrangian, and is directly linked to the entropy solution ρ\rho of the scalar conservation law (1.1) by the Lax-Oleinik formula (see [6][Theorem 1, Chapter 3.4])

ρ(x,t)=(H)1(y(x,t)xt)=L(y(x,t)xt)\rho(x,t)=(H^{\prime})^{-1}\left(\frac{y(x,t)-x}{t}\right)=L^{\prime}\left(\frac{y(x,t)-x}{t}\right)

The reader may be familiar with this other form of the Hamilton-Jacobi PDE

(1.4) ut+H(ux)=0u_{t}+H(u_{x})=0

If we denote by uu a solution of (1.4), then it is easy to see that u~(x,t):=u(x,t)\tilde{u}(x,t):=-u(x,t) verifies u~t=H~(u~x)\tilde{u}_{t}=\tilde{H}(\tilde{u}_{x}) for the Hamiltonian H~(ρ)=H(ρ)\tilde{H}(\rho)=H(-\rho). We will thus only restrict ourselves to the version of the scalar conservation law in (1.1).

When the Hamiltonian HH takes the simple form H(ρ)=ρ22H(\rho)=\frac{\rho^{2}}{2}, the scalar conservation law (1.1) is called Burgers equation and is written ρt=ρρx\rho_{t}=\rho\rho_{x}. The Lax-Oleinik formula simplifies to

(1.5) ρ(x,t)=y(x,t)xt\rho(x,t)=\frac{y(x,t)-x}{t}

The Burgers equation has seen an extensive interest when the initial data ρ(,0)\rho(\cdot,0) is random in the context of Burgers turbulence. We will present thereby the most relevant results in this area.

1.2. Burgers equation when ρ(,0)\rho(\cdot,0) is a Brownian white noise

This is the case when the initial potential U0U_{0} is expressed as

(1.6) U0(x)=σB(x),xU_{0}(x)=\sigma B(x),~{}~{}x\in\mathbb{R}

where σ>0\sigma>0 is a diffusion factor and BB is a two-sided standard linear Brownian motion. In a remarkable paper [9] with the aim of studying the global behavior of isotonic estimators, Groeneboom completely determined the statistics of the process

(V(a):=sup{x:B(x)(xa)2 is maximal},a)\displaystyle\left(V(a):=\sup\left\{x\in\mathbb{R}:B(x)-(x-a)^{2}\text{ is maximal}\right\},a\in\mathbb{R}\right)

He showed that this process is pure-jump with jump kernels expressed in terms of Airy functions. By the Hopf-Lax formula and (1.5), this process is related to the solution of the Burgers equation with Brownian white noise initial data.

More precisely, let ρσ(x,t)\rho_{\sigma}(x,t) be the entropy solution of Burgers equation when the initial potential is determined by (1.6). Since in the Burgers case the Hamiltonian enjoys the same scaling as the Brownian motion. It follows that for every t>0t>0, the process (ρσ(x,t),x)(\rho_{\sigma}(x,t),x\in\mathbb{R}) has the same law as (σ23t13ρ1(x((σt)23,1),x)(\sigma^{\frac{2}{3}}t^{-\frac{1}{3}}\rho_{1}(x((\sigma t)^{-\frac{2}{3}},1),x\in\mathbb{R}). The following theorem gives a precise description of the law of the entropy solution at time t=1t=1.

Theorem 1.1 (Groeneboom 89, [9]).

The process (ρ12(x,1),x)(\rho_{\frac{1}{\sqrt{2}}}(x,1),x\in\mathbb{R}) is a stationary piecewise-linear Markov process with generator 𝒜\mathcal{A} acting on a test function φCc()\varphi\in C_{c}^{\infty}(\mathbb{R}) as

𝒜φ(y)=φ(y)+y(φ(z)φ(y))n(y,z)𝑑z\mathcal{A}\varphi(y)=-\varphi^{\prime}(y)+\int_{y}^{\infty}(\varphi(z)-\varphi(y))n(y,z)dz

The jump density nn is given by the formula

n(y,z)=J(z)J(y)K(zy),z>y\displaystyle n(y,z)=\frac{J(z)}{J(y)}K(z-y)~{}~{},~{}~{}z>y

where JJ and ZZ are positive functions defined on the line and positive half-line respectively, whose Laplace transforms

j(q)=eqyJ(y)𝑑y,k(q)=0eqyK(y)𝑑y\displaystyle j(q)=\int_{-\infty}^{\infty}e^{qy}J(y)dy,~{}~{}~{}~{}~{}k(q)=\int_{0}^{\infty}e^{-qy}K(y)dy

are meromorphic functions on \mathbb{C} given by

j(q)=1Ai(q),k(q)=2d2dq2logAi(q)\displaystyle j(q)=\frac{1}{\mathrm{Ai}(q)},~{}~{}~{}~{}k(q)=-2\frac{d^{2}}{dq^{2}}\mathrm{log}\mathrm{Ai}(q)

where Ai\mathrm{Ai} denotes the first Airy function.

Remark 1.2.

For general t>0t>0, the process (ρ12(x,t),x)(\rho_{\frac{1}{\sqrt{2}}}(x,t),x\in\mathbb{R}) is also a stationary piecewise-linear Markov process with generator

𝒜tφ(y)=1tφ(y)+yt13n(yt13,zt13)(φ(z)φ(y))𝑑z\mathcal{A}^{t}\varphi(y)=-\frac{1}{t}\varphi^{\prime}(y)+\int_{y}^{\infty}t^{-\frac{1}{3}}n(yt^{\frac{1}{3}},zt^{\frac{1}{3}})(\varphi(z)-\varphi(y))dz

In particular, the linear pieces have slope 1t\displaystyle-\frac{1}{t}.

1.3. Burgers equation when ρ(,0)\rho(\cdot,0) is a spectrally negative Lévy process

A Lévy process (Xt)t(X_{t})_{t\in\mathbb{R}} is a process with stationary independent increments and such that X0=0X_{0}=0. By spectrally negative Lévy process, we mean a process that has only downward jumps. For the Burgers equation, Bertoin in [4] proved a remarkable closure theorem for this class of initial data. We quote here his result.

Theorem 1.3 (Bertoin 98, [4]).

Consider Burgers equation of the form ρt+ρρx=0\rho_{t}+\rho\rho_{x}=0 with initial data ξ(x)\xi(x) which is a spectrally negative Lévy process for x0x\geq 0 and ξ(x)=0\xi(x)=0 for x<0x<0. Assume that the expected value of ξ(1)\xi(1) is positive. Then for each fixed t>0t>0, the backward Lagrangian y(x,t)y(x,t) has the property that (y(x,t)y(0,t))x0(y(x,t)-y(0,t))_{x\geq 0} is independent of y(0,t)y(0,t) and is in the parameter xx a subordinator, i.e. a nondecreasing Lévy process. Its distribution is that of the first passage process

xinf{z0:tξ(z)+z>x}\displaystyle x\mapsto\inf\{z\geq 0:t\xi(z)+z>x\}

Furthermore, if we denote by ψ(s)\psi(s) and Θ(t,s)\Theta(t,s) (s0)(s\geq 0) respectively the Laplace exponents of ξ(x)\xi(x) and y(x,t)y(x,0)y(x,t)-y(x,0),

𝔼[exp(sξ(x))]=exp(xψ(s))\displaystyle\mathbb{E}[\mathrm{exp}(s\xi(x))]=\mathrm{exp}(x\psi(s))
𝔼[exp(s(y(x,t)y(0,t)))]=exp(xΘ(t,s))\displaystyle\mathbb{E}[\mathrm{exp}(s(y(x,t)-y(0,t)))]=\mathrm{exp}(x\Theta(t,s))

then we have the functional identity

ψ(tΘ(t,s))+Θ(t,s)=s\displaystyle\psi(t\Theta(t,s))+\Theta(t,s)=s

Moreover, the process (ρ(x,t)ρ(0,t))x0(\rho(x,t)-\rho(0,t))_{x\geq 0} is a Lévy process, and its Laplace exponent ψ(t,q)\psi(t,q) verifies the Burgers equation

(1.7) ψt+ψψq=0\psi_{t}+\psi\psi_{q}=0
Remark 1.4.

This theorem is remarkable in the sense that it provides an infinite-dimensional, nonlinear dynamical system which preserves the independence and homogeneity properties of its random initial configuration. Moreover, it was observed in [14] that the evolution according to Burgers equation of the Laplace exponents in (1.7) corresponds to a Smoluchowski coagulation equation [21] with additive rate which determines the jump statistics. This connection is simply due to the Lévy-Khintchine representation of Laplace exponents.

1.4. Scalar conservation law with general Hamiltonian HH

A natural question that arises is if the previous phenomenon (the entropy solution at later times having a simple form that can be explicitly described) is intrinsic to the Burgers equation or if the same holds for scalar conservation laws with general Hamiltonians HH. In an attempt to answer this question, Menon and Srinivasan in [15] proved that when the initial condition ξ\xi is a spectrally positive strong Markov process, then the entropy solution of (1.1) at later times remains Markov and spectrally positive. However, it is not as clear whether the Feller property is preserved through time. The following conjecture was stated in that paper, together with different heuristic but convincing ways to see why that must be true.

Conjecture 1.5.

If the initial data ξ\xi of the scalar conservation law in (1.1) is either

  1. (1)

    A white noise derived from a spectrally positive Lévy process.

  2. (2)

    A stationary spectrally positive Feller process with bounded variation.

then the solution ρ(,t)\rho(\cdot,t) for any fixed time t>0t>0 is a stationary spectrally positive Feller process with bounded variation. Moreover, its jump kernel and drift verify an integro-differential equation.

Remark 1.6.

By a result of Courrège (see [3][Theorem 3.5.3]), the generator 𝒜\mathcal{A} of any spectrally positive Feller process with bounded variation takes the form

𝒜φ(y)=b(y)φ(y)+y(φ(z)φ(y))n(y,dz)\displaystyle\mathcal{A}\varphi(y)=b(y)\varphi^{\prime}(y)+\int_{y}^{\infty}(\varphi(z)-\varphi(y))n(y,dz)

given that Cc()𝒟(𝒜)C^{\infty}_{c}(\mathbb{R})\subset\mathcal{D}(\mathcal{A}) (Cc()C^{\infty}_{c}(\mathbb{R}) is the space of infinitely differentiable functions with compact support and 𝒟(𝒜)\mathcal{D}(\mathcal{A}) is the domain of the generator 𝒜\mathcal{A}). Moreover the kernel nn verifies the integrability condition : y(1|yz|2)n(y,dz)<\int_{y}^{\infty}(1\wedge|y-z|^{2})n(y,dz)<\infty.

A variant111Under some mild conditions on the Hamiltonian HH, and a slight modification of the nature of the initial data. of the second part of this conjecture when the initial data is a piecewise-deterministic spectrally positive Feller process was recently proved by Kaspar and Rezakhanlou in [12] and [13]. We give below an explicit exposition of their result together with the exact form of the integro-differential equation verified by the drift and the jump kernel. This equation (1.8) was formally derived by Menon and Srinivasan in [15] and shown to be equivalent to the following Lax equation

t𝒜=[𝒜,]=𝒜𝒜\partial_{t}\mathcal{A}=[\mathcal{A},\mathcal{B}]=\mathcal{A}\mathcal{B}-\mathcal{B}\mathcal{A}

where 𝒜t\mathcal{A}^{t} is the generator of xρ(x,t)x\mapsto\rho(x,t) and t\mathcal{B}^{t} is the generator of tρ(x,t)t\mapsto\rho(x,t). We give explicit formulas for these generators below in the statement of Theorem 1.8.

Notation 1.7.

We write 1\mathcal{M}_{1} for the set of probability measures on the real line, and

[H]y,z=H(y)H(z)yz for yz\displaystyle[H]_{y,z}=\frac{H(y)-H(z)}{y-z}\text{ for }y\neq z
Theorem 1.8 (Kaspar and Rezakhanlou 20, [13]).

Assume that the initial data ρ0=ρ0(x)\rho^{0}=\rho^{0}(x) is zero of x<0x<0 and is a Markov process for x0x\geq 0 that starts at ρ0(0)=0\rho^{0}(0)=0. More precisely, its infinitesimal generator 𝒜0\mathcal{A}^{0} has the form

𝒜0φ(ρ)=b0(ρ)φ(ρ)+ρ(φ(ρ+)φ(ρ))f0(ρ,ρ+)𝑑ρ+\displaystyle\mathcal{A}^{0}\varphi(\rho_{-})=b^{0}(\rho_{-})\varphi^{\prime}(\rho_{-})+\int_{\rho_{-}}^{\infty}(\varphi(\rho_{+})-\varphi(\rho_{-}))f^{0}(\rho_{-},\rho_{+})d\rho_{+}

Furthermore, assume that

  1. (1)

    The rate kernel f0(p,p+)f^{0}(p{-},p_{+}) is C1C^{1} and is supported on

    {(p,p+):Ppp+P+}\displaystyle\{(p_{-},p_{+}):P_{-}\leq p_{-}\leq p_{+}\leq P_{+}\}

    for some constants P±P_{\pm}.

  2. (2)

    The Hamiltonian function H:[P,P+]H:[P_{-},P_{+}]\to\mathbb{R} is C2C^{2}, convex, has positive right-derivative at p=Pp=P_{-} and finite left-derivative at p=P+p=P_{+}.

  3. (3)

    The initial drift b0b^{0} is C1C^{1} and satisfies b00b^{0}\leq 0 with b0(ρ)=0b^{0}(\rho)=0 whenever ρ[P,P+]\rho\notin[P_{-},P_{+}].

Then for each fixed t>0t>0, the process xρ(x,t)x\mapsto\rho(x,t) (where ρ\rho is a solution of (1.1)) has x=0x=0 marginal given by 0(dρ0,t)\ell^{0}(d\rho_{0},t) where 0:[0,)1\ell^{0}:[0,\infty)\to\mathcal{M}_{1} is the unique function such that 0(dρ,0)=δ0(dρ)\ell^{0}(d\rho,0)=\delta_{0}(d\rho) and

d0(dρ,t)dt=(t0(,t))(dρ,t)\displaystyle\frac{d\ell^{0}(d\rho,t)}{dt}=(\mathcal{B}^{t*}\ell^{0}(\cdot,t))(d\rho,t)

where t\mathcal{B}^{t*} is the adjoint operator of t\mathcal{B}^{t}, that acts on measures with

tφ(ρ)=H(ρ)b(ρ,t)φ(ρ)ρ[H]ρ,ρ+(φ(ρ+)φ(ρ))f(ρ,ρ+,t)𝑑ρ+\displaystyle\mathcal{B}^{t}\varphi(\rho_{-})=-H^{\prime}(\rho_{-})b(\rho_{-},t)\varphi^{\prime}(\rho_{-})-\int_{\rho_{-}}^{\infty}[H]_{\rho_{-},\rho_{+}}(\varphi(\rho_{+})-\varphi(\rho_{-}))f(\rho_{-},\rho_{+},t)d\rho_{+}

for any test function φ\varphi. Moreover the process xρ(x,t)x\mapsto\rho(x,t) evolves for 0<x<0<x<\infty according to a Markov process with generator 𝒜t\mathcal{A}^{t} given by

𝒜tφ(ρ)=b(ρ,t)φ(ρ)+ρ(φ(ρ+)φ(ρ))f(ρ,ρ+,t)𝑑ρ+\displaystyle\mathcal{A}^{t}\varphi(\rho_{-})=b(\rho_{-},t)\varphi^{\prime}(\rho_{-})+\int_{\rho_{-}}^{\infty}(\varphi(\rho_{+})-\varphi(\rho_{-}))f(\rho_{-},\rho_{+},t)d\rho_{+}

Here bb and ff are obtained from their initial conditions

b(ρ,0)=b0(ρ),f(ρ,ρ+,0)=f0(ρ,ρ+)\displaystyle b(\rho,0)=b^{0}(\rho),~{}~{}~{}~{}f(\rho_{-},\rho_{+},0)=f^{0}(\rho_{-},\rho_{+})

bb solves the ODE with parameter

tb(ρ,t)=H′′(ρ)b(ρ,t)2\partial_{t}b(\rho,t)=H^{\prime\prime}(\rho)b(\rho,t)^{2}

and ff solves the following Boltzmann-like kinetic equation

(1.8) tf(ρ,ρ,t)=Q(f,f)+C(f)+ρ(fVρ(ρ,ρ+,t))+ρ+(fVρ+(ρ,ρ+,t))\displaystyle\partial_{t}f(\rho_{-},\rho_{-},t)=Q(f,f)+C(f)+\partial_{\rho_{-}}(fV_{\rho_{-}}(\rho_{-},\rho_{+},t))+\partial_{\rho_{+}}(fV_{\rho_{+}}(\rho_{-},\rho_{+},t))

where the velocities VρV_{\rho_{-}} and Vρ+V_{\rho_{+}} are given by

Vρ(ρ,ρ+,t)=([H]ρ,ρ+H(ρ))b(ρ,t)\displaystyle V_{\rho_{-}}(\rho_{-},\rho_{+},t)=([H]_{\rho_{-},\rho_{+}}-H^{\prime}(\rho_{-}))b(\rho_{-},t)
Vρ+(ρ,ρ+,t)=([H]ρ,ρ+H(ρ+))b(ρ+,t)\displaystyle V_{\rho_{+}}(\rho_{-},\rho_{+},t)=([H]_{\rho_{-},\rho_{+}}-H^{\prime}(\rho_{+}))b(\rho_{+},t)

the coagulation-like collision kernel QQ is

Q(f,f)(ρ,ρ+,t)=ρρ+([H]ρ,ρ+[H]ρ,ρ)f(ρ,ρ,t)f(ρ,ρ+,t)𝑑ρ\displaystyle Q(f,f)(\rho_{-},\rho_{+},t)=\int_{\rho_{-}}^{\rho_{+}}([H]_{\rho_{*},\rho_{+}}-[H]_{\rho_{-},\rho_{*}})f(\rho_{-},\rho_{*},t)f(\rho_{*},\rho_{+},t)d\rho_{*}
ρ+([H]ρ,ρ+[H]ρ+,ρ)f(ρ,ρ+,t)f(ρ+,ρ,t)𝑑ρ\displaystyle-\int_{\rho_{+}}^{\infty}([H]_{\rho_{-},\rho_{+}}-[H]_{\rho_{+},\rho_{*}})f(\rho_{-},\rho_{+},t)f(\rho_{+},\rho_{*},t)d\rho_{*}
ρ([H]ρ,ρ[H]ρ,ρ+)f(ρ,ρ+,t)f(ρ,ρ,t)𝑑ρ\displaystyle-\int_{\rho_{-}}^{\infty}([H]_{\rho_{-},\rho_{*}}-[H]_{\rho_{-},\rho_{+}})f(\rho_{-},\rho_{+},t)f(\rho_{-},\rho_{*},t)d\rho_{*}

and the linear operator CC is given by

C(f)(ρ,ρ+)=f(ρ,ρ+)(b(ρ,t)H′′(ρ)([H]ρ,ρ+H(ρ))ρb(ρ,t))\displaystyle C(f)(\rho_{-},\rho{+})=f(\rho_{-},\rho_{+})(b(\rho_{-},t)H^{\prime\prime}(\rho_{-})-([H]_{\rho_{-},\rho_{+}}-H^{\prime}(\rho_{-}))\partial_{\rho_{-}}b(\rho_{-},t))

The purpose of this paper is to prove the first part of the conjecture when the initial data ξ\xi is a Brownian white noise and thus extend the results of Groeneboom [9] in the Burgers case. We show that at any fixed time t>0t>0, the solution ρ(,t)\rho(\cdot,t) is a stationary piecewise-smooth Feller process and we give an explicit description of its generator. This result proves the complete integrability of scalar conservation laws for this class of initial data and moves away from the unnatural emphasis on Burgers equation. Our method as will be seen by the reader can be extended when the white noise is derived from a spectrally positive Lévy process with non-zero Brownian exponent. Our shortcoming in this case will be not having explicit formulas for the jump kernel. We also show that the structure of shocks of Burgers turbulence holds for the general scalar conservation law under the assumption of rough initial data.

Since the entropy solution is expressed via the Lax-Oleinik formula. It is natural to study the law of the process Ψϕ\Psi^{\phi} defined as

(1.9) Ψϕ(x)=sup{y:U0(y)ϕ(yx)=maxz(U0(z)ϕ(zx))},x\Psi^{\phi}(x)=\sup\left\{y\in\mathbb{R}:U_{0}(y)-\phi(y-x)=\max\limits_{z\in\mathbb{R}}\left(U_{0}(z)-\phi(z-x)\right)\right\},~{}~{}x\in\mathbb{R}

where U0U_{0} is a spectrally positive Lévy process and ϕ\phi is a C2C^{2} strictly convex function with superlinear growth, such that U0(y)=o(ϕ(y))U_{0}(y)=o(\phi(y))222We write f=o(g)f=o(g) if limfg=0\lim\frac{f}{g}=0 and f=O(g)f=O(g) if fg\frac{f}{g} is bounded. for |y||y|\to\infty. The relationship between the process Ψϕ\Psi^{\phi} and the entropy solution ρ(,t)\rho(\cdot,t) of (1.1) is the following

ρ(x,t)=L(ΨtL(t)(x)xt)\displaystyle\rho(x,t)=L^{\prime}\left(\frac{\Psi^{tL(\frac{\cdot}{t})}(x)-x}{t}\right)

Our paper is organized as follows

  1. (1)

    In Section 2, we give some preliminary results on the process Ψϕ\Psi^{\phi} when U0U_{0} is a spectrally positive Lévy process such as its Markovian property.

  2. (2)

    In Section 3, we will focus on the case where U0U_{0} is a two-sided Brownian motion and show that the process Ψϕ\Psi^{\phi} is pure jump, following similar ideas used by Groeneboom in [9]. The main ingredient being the path decomposition of Markov processes when they reach their ultimate maximum. This result implies that the Brownian motion U0U_{0} has excursions below the sequence of convex functions (xϕ(xxn))n(x\mapsto\phi(x-x_{n}))_{n\in\mathbb{N}} where (xn)n(x_{n})_{n\in\mathbb{N}} are the jump times of the process Ψϕ\Psi^{\phi} (which is a discrete set by a result of Section 5). However, the justification of many manipulations used in [9] rely on the regularity and asymptotic properties of Airy functions at infinity, as those arise naturally in the expressions of transition densities used throughout the study of the Brownian motion with parabolic drift. Unfortunately, those special functions are intrinsic to this special case as we will explain later, and one do not have similar expressions in the general case.

  3. (3)

    In Section 4, we circumvent this difficulty by using a more analytic approach to prove the smoothness and integrability of the densities that were used in Section 3. Moreover, via Girsanov theorem we manage to express explicitly the jump kernel of the process Ψϕ\Psi^{\phi} in terms of the distribution of Brownian excursion areas. Along the way, we find the joint density of the maximum and its location of the process (W(z)ϕ(z))z(W(z)-\phi(z))_{z\in\mathbb{R}} where WW is a two-sided Brownian motion. In particular, the density of argmaxz(W(z)ϕ(z))\text{argmax}_{z\in\mathbb{R}}(W(z)-\phi(z)) enjoys a simple expression similar to Chernoff distribution for the parabolic drift.

  4. (4)

    Finally, in Section 5 we give a sufficient condition on the Lévy process U0U_{0} for the process Ψϕ\Psi^{\phi} to have discrete range (with the convention that a set is discrete if it is countable with no accumulation points). As a consequence, this implies that the structure of shocks of the entropy solution ρ(,t)\rho(\cdot,t) is discrete for any time t>0t>0 when the initial data belongs to the large class of abrupt Lévy processes introduced by Vigon in [20], this result generalizes the findings of Bertoin [5] and Abramson [1] when U0U_{0} is spectrally positive.

We give here our main results

Theorem 1.9.

Suppose that the initial potential U0U_{0} is a two-sided Brownian motion and let ρ\rho be the solution of the scalar conservation law ρt=(H(ρ))x\rho_{t}=(H(\rho))_{x}. Then for every fixed t>0t>0, the process xρ(x,t)x\mapsto\rho(x,t) is a stationary piecewise-smooth Feller process. Its generator is given by

𝒜tφ(ρ)=φ(ρ)tH′′(ρ)+ρ(φ(ρ+)φ(ρ))n(ρ,ρ+,t)𝑑ρ+\displaystyle\mathcal{A}^{t}\varphi(\rho_{-})=-\frac{\varphi^{\prime}(\rho_{-})}{tH^{\prime\prime}(\rho_{-})}+\int_{\rho_{-}}^{\infty}(\varphi(\rho_{+})-\varphi(\rho_{-}))n(\rho_{-},\rho_{+},t)d\rho_{+}

for any test function φCc()\varphi\in C_{c}^{\infty}(\mathbb{R}), where

(1.10) n(ρ,ρ+,t)=(ρ+ρ)K(ρ,ρ+,t)2πt(H(ρ+)H(ρ))3ρ++ρ+H′′(ρ)K(ρ+,ρ,t)2πt(H(ρ)H(ρ+))3𝑑ρρ+ρH′′(ρ)K(ρ,ρ,t)2πt(H(ρ)H(ρ))3𝑑ρn(\rho_{-},\rho_{+},t)=\frac{(\rho_{+}-\rho_{-})K(\rho_{-},\rho_{+},t)}{\sqrt{2\pi t(H^{\prime}(\rho_{+})-H^{\prime}(\rho_{-}))^{3}}}\frac{\rho_{+}+\int_{\rho_{+}}^{\infty}\frac{H^{\prime\prime}(\rho)-K(\rho_{+},\rho,t)}{\sqrt{2\pi t(H^{\prime}(\rho)-H^{\prime}(\rho_{+}))^{3}}}d\rho}{\rho_{-}+\int_{\rho_{-}}^{\infty}\frac{H^{\prime\prime}(\rho)-K(\rho_{-},\rho,t)}{\sqrt{2\pi t(H^{\prime}(\rho)-H^{\prime}(\rho_{-}))^{3}}}d\rho}

for ρ<ρ+\rho_{-}<\rho_{+}, and

K(ρ,ρ+,t)=H′′(ρ+)exp(t2ρρ+ρ2H′′(ρ)dρ)×\displaystyle K(\rho_{-},\rho_{+},t)=H^{\prime\prime}(\rho_{+})\text{exp}\left(-\frac{t}{2}\int_{\rho_{-}}^{\rho_{+}}\rho_{*}^{2}H^{\prime\prime}(\rho_{*})d\rho_{*}\right)\times
𝔼[exp(ρρ+e(tH(ρ))𝑑ρ)]\displaystyle\mathbb{E}\left[\text{exp}\left(-\int_{\rho_{-}}^{\rho_{+}}\textbf{e}(tH^{\prime}(\rho_{*}))d\rho_{*}\right)\right]

where e is a Brownian excursion on the interval [tH(ρ),tH(ρ+)][tH^{\prime}(\rho_{-}),tH^{\prime}(\rho_{+})].

Remark 1.10.
  1. (1)

    The profile of the solution at any fixed time t>0t>0 is a concatenation of smooth pieces that evolve as solutions of ODEs with vector field (or drift) b(ρ,t):=1tH′′(ρ)b(\rho,t):=-\frac{1}{tH^{\prime\prime}(\rho)} and are interrupted by stochastic upward jumps distributed via the jump kernel n(,,t)n(\cdot,\cdot,t). We prove in Section 5 that in the Brownian white noise case, under mild assumptions on the Hamiltonian HH, the set of jump times is discrete, i.e. : there are only a finite number of jumps on any given compact interval.

  2. (2)

    For any ϵ>0\epsilon>0, the profile of xρ(x,ϵ)x\mapsto\rho(x,\epsilon) is a piecewise-deterministic Markov process and belongs to the class of initial data considered in the second part of the Conjecture 1.5. A consequence of this observation would be that the kernel (ρ,ρ+,t)n(ρ,ρ+,t)(\rho_{-},\rho_{+},t)\mapsto n(\rho_{-},\rho_{+},t) in the expression (1.10) verifies the kinetic equation (1.8). However, Theorem 1.8 only considers a variant of the original statement of the conjecture as it forces the initial data to be flat on the negative real-line (whereas here we deal with a stationary process) and restricts the range of ρ0\rho^{0} on a compact interval [P,P+][P_{-},P_{+}]. These technical modifications arise from the very challenging proof of existence and uniqueness of a classical solution to (1.8) under general assumptions. Verifying that the kernel nn in the Brownian white noise case is a solution to the kinetic equation (1.8) from the explicit expression (1.10) seems also inaccessible at the present due to the complicated term involving the Brownian excursion. This verification was done for the Burgers case by Menon and Srinivasan in [15][Section 6] through many non-trivial calculations, but relied extensively on the connection with Airy functions and an associated Painlevé property.

The following result is a consequence of our study of the process Ψϕ\Psi^{\phi}. It gives an explicit formula for the density of the random variable argmaxω(W(ω)ϕ(ω))\text{argmax}_{\omega\in\mathbb{R}}(W(\omega)-\phi(\omega)) where WW is a two-sided Brownian motion. From results of Section 4, we also have access to the joint distribution of

(argmaxω(W(ω)ϕ(ω)),maxω(W(ω)ϕ(ω)))\displaystyle(\text{argmax}_{\omega\in\mathbb{R}}(W(\omega)-\phi(\omega)),\max_{\omega\in\mathbb{R}}(W(\omega)-\phi(\omega)))

but we omit it here because the expression is quite large.

Theorem 1.11.

Let ωM\omega_{M} be the location of the maximum of the process (S(ω)=W(ω)ϕ(ω))ω(S(\omega)=W(\omega)-\phi(\omega))_{\omega\in\mathbb{R}} where WW is a two-sided Brownian motion, its density is equal to

[ωMdt]dt=12fϕ(t)fϕ()(t)\displaystyle\frac{\mathbb{P}[\omega_{M}\in dt]}{dt}=\frac{1}{2}f^{\phi}(t)f^{\phi(-\cdot)}(-t)

for any tt\in\mathbb{R}, and where

fϕ(t)=ϕ(t)+01pϕ(t,u)2πu3𝑑u\displaystyle f^{\phi}(t)=\phi^{\prime}(t)+\int_{0}^{\infty}\frac{1-p^{\phi}(t,u)}{\sqrt{2\pi u^{3}}}du

with

pϕ(t,u)=exp(12tt+uϕ(z)2𝑑z)𝔼[exp(tt+uϕ′′(z)e(z)𝑑z)] for u>0\displaystyle p^{\phi}(t,u)=\text{exp}\left(-\frac{1}{2}\int_{t}^{t+u}\phi^{\prime}(z)^{2}dz\right)\mathbb{E}\left[\text{exp}\left(-\int_{t}^{t+u}\phi^{\prime\prime}(z)\textbf{e}(z)dz\right)\right]\text{ for }u>0

where e is a Brownian excursion on [t,t+u][t,t+u].

Remark 1.12.

In the parabolic drift case (Chernoff distribution), the term ϕ′′\phi^{\prime\prime} is constant and the Laplace transform of a standard Brownian excursion area is known to be expressed via Airy functions. We will develop on the connection between the formulas found by Groeneboom in [9] and ours at the end of Section 4. Also, we refer the reader to the survey [11] for a more detailed exposition on the distribution and Laplace transform of various Brownian paths areas.

x1x_{1}x2x_{2}Drift b(,t)b(\cdot,t)xxρ(,t)\rho(\cdot,t)
Figure 1. The typical profile of the entropy solution at a given time t>0t>0.

We define now a class of rough Lévy processes called abrupt that were introduced by Vigon in [20].

Definition 1.13.

A Lévy process (Xt)t(X_{t})_{t\in\mathbb{R}} is said to be abrupt if its paths have unbounded variation and almost surely for all local maxima mm of XX we have

lim infh01h(XmhXm)=+ and lim suph01h(Xm+hXm)=\displaystyle\liminf\limits_{h\downarrow 0}\frac{1}{h}(X_{m-h}-X_{m-})=+\infty\text{ and }\limsup\limits_{h\downarrow 0}\frac{1}{h}(X_{m+h}-X_{m})=-\infty
Remark 1.14.

A Lévy process XX with paths of unbounded variation is abrupt if and only if

01t1[Xt[at,bt]]𝑑t<,a<b\displaystyle\int_{0}^{1}t^{-1}\mathbb{P}\left[X_{t}\in[at,bt]\right]dt<\infty,~{}~{}\forall a<b

Examples of abrupt Lévy processes include stable processes with index α(1,2]\alpha\in(1,2] and any process with non-zero Brownian exponent.

Our last main result determines the structure of shocks of the scalar conservation law when the initial data is a white noise derived from an abrupt Lévy process.

Theorem 1.15.

Assume that the Lévy process U0U_{0} is spectrally positive, abrupt and is such that U0(y)=O(|y|)U_{0}(y)=O(|y|) for |y||y|\rightarrow\infty, then the set

t={y:y=y(x,t) or y=y(x,t) for some x}\displaystyle\mathcal{L}^{t}=\{y\in\mathbb{R}:y=y(x,t)\text{ or }y=y(x-,t)\text{ for some }x\in\mathbb{R}\}

is almost surely discrete for any fixed time t>0t>0. We say then that the structure of shocks of the entropy solution ρ(,t)\rho(\cdot,t) is discrete.

Remark 1.16.

From a point of view of hydrodynamic turbulence, a discontinuity of the entropy solution ρ(,t)\rho(\cdot,t) at position xx means the presence of a cluster of particles at this location at time tt. Those clusters interact with each other via inelastic shocks, and the cluster at location xx and at time tt contains all the particles that were initially located in [y(x,t),y(x,t))[y(x-,t),y(x,t)). Our result shows that at any given time t>0t>0, the set of clusters is discrete. When the initial data is a Lévy white noise, we can picture that there are infinitely many particles initially scattered everywhere with i.i.d velocities. Therefore, when we assume that this initial profile is rough (as it is the case when the potential U0U_{0} is an abrupt Lévy process), this turbulence forces all the particles to aggregate in heavy disjoint lumps instantaneously for any time t>0t>0.

Acknowledgement

I would like to thank my advisor Fraydoun Rezakhanlou for many fruitful discussions.

2. Preliminaries

Notation 2.1.

We will use the notation argmax+f\text{argmax}^{+}f to denote the rightmost maximizer of a function ff (i.e. : the last time at which a function ff reaches its maximum).

Menon and Srinivisan proved in their paper [15] a closure theorem for white noise initial data for the scalar conservation law solutions. They showed that if initially the potential U0U_{0} is spectrally positive with independent increments then ρ(,t)\rho(\cdot,t) is a spectrally positive Markov process for any fixed t>0t>0. The proof of this statement follows from standard use of path decomposition of strong Markov processes at their ultimate maximum. The same holds for our process Ψϕ\Psi^{\phi}. Precisely, we have the following theorem for which we give the proof for the sake of completeness.

Theorem 2.2.

Assume that U0U_{0} is a spectrally positive Lévy process, then the process Ψϕ\Psi^{\phi} is a non-decreasing Markov process. Moreover for any aa\in\mathbb{R}, the process Ψϕ(.+a)a\Psi^{\phi}(.+a)-a has the same distribution as Ψϕ\Psi^{\phi}.

Proof.

For x1x2x_{1}\leq x_{2} and yΨϕ(x1)y\leq\Psi^{\phi}(x_{1}), we have that

U0(Ψϕ(x1))U0(y)\displaystyle U_{0}(\Psi^{\phi}(x_{1}))-U_{0}(y) ϕ(Ψϕ(x1)x1)ϕ(yx1)\displaystyle\geq\phi(\Psi^{\phi}(x_{1})-x_{1})-\phi(y-x_{1})
ϕ(Ψϕ(x1)x2)ϕ(yx2)\displaystyle\geq\phi(\Psi^{\phi}(x_{1})-x_{2})-\phi(y-x_{2})

By the convexity of ϕ\phi, and hence Ψϕ(x1)Ψϕ(x2)\Psi^{\phi}(x_{1})\leq\Psi^{\phi}(x_{2}). Also, by definition Ψϕ\Psi^{\phi} is a càdlàg process (right continuous with left hand limits). Take h>0h>0, then

(2.1) Ψϕ(x+h)=Ψϕ(x)+argmaxy0+(U0(y+Ψϕ(x))ϕ(y+Ψϕ(x)(x+h)))\Psi^{\phi}(x+h)=\Psi^{\phi}(x)+\\ \text{argmax}^{+}_{y\geq 0}\left(U_{0}(y+\Psi^{\phi}(x))-\phi(y+\Psi^{\phi}(x)-(x+h))\right)

The process Ux(y):=U0(y)ϕ(yx)U^{x}(y):=U_{0}(y)-\phi(y-x) is clearly Markov. By Millar’s theorem of path decomposition of Markov processes when they reach their ultimate maximum (see [16]), the process (Ux(y+Ψϕ(x))y0(U^{x}(y+\Psi^{\phi}(x))_{y\geq 0} is independent of (Ux(y))yΨϕ(x)(U^{x}(y))_{y\leq\Psi^{\phi}(x)} given (Ψϕ(x),Ux(Ψϕ(x)))(\Psi^{\phi}(x),U^{x}(\Psi^{\phi}(x))) (because of the upward jumps of U0U_{0}, the maximum is attained at the right hand limit). Moreover, because of the independence of the increments of UxU^{x}, the process (Ux(y+Ψϕ(x))Ux(Ψϕ(x)))y0(U^{x}(y+\Psi^{\phi}(x))-U^{x}(\Psi^{\phi}(x)))_{y\geq 0} is independent of (Ux(y))yΨϕ(x)(U^{x}(y))_{y\leq\Psi^{\phi}(x)} given Ψϕ(x)\Psi^{\phi}(x). Now it suffices to see that (Ψϕ(y))yx(\Psi^{\phi}(y))_{y\leq x} only depends on the pre-maximum process (Ux(y))yΨϕ(x)(U^{x}(y))_{y\leq\Psi^{\phi}(x)} because of the monotonicity of Ψϕ\Psi^{\phi}, this fact alongside the equation (2.1) gives the Markov property of the process Ψϕ\Psi^{\phi}. The last statement follows easily from the stationarity of increments of U0U_{0}. ∎

Remark 2.3.

Notice that except in the last statement, the stationarity of increments was not used in the proof of the Markovian property of the process Ψϕ\Psi^{\phi}.

3. The process Ψϕ\Psi^{\phi} in the Brownian case

In this section, we assume that W:=U0W:=U_{0} is a two-sided Brownian motion. We proved in the previous section that the process Ψϕ\Psi^{\phi} is Markov and enjoys a space-time shifted stationarity property. Hence, we shall only determine its transition function at time zero and consequently the form of its generator at this time. In this section we will differentiate and switch the order of integrals and differentiations without any justification, as Section 4 is devoted to take care of all those technicalities.

Notation 3.1.

In the sequel, we will deal with functions of the form f(s,x,t,y)f(s,x,t,y) where tt and ss play the role of temporal variables, and xx and yy that of spatial variables. Without confusion, the notation xf(s,x,t,y)\partial_{x}f(s,x,t,y) (resp. yf(s,x,t,y)\partial_{y}f(s,x,t,y)) refer to the partial derivative of ff with respect to the second variable (resp. fourth variable).

We state here the first result regarding the transition function of the process Ψϕ\Psi^{\phi}.

Theorem 3.2.

Let h>0h>0 and ω1<ω2\omega_{1}<\omega_{2} be two real numbers. Then we have that

[Ψϕ(h)dω2|Ψϕ(0)=ω1]=[argmaxωω1+Xh(ω)dω2]\displaystyle\mathbb{P}[\Psi^{\phi}(h)\in d\omega_{2}|\Psi^{\phi}(0)=\omega_{1}]=\mathbb{P}[\text{argmax}^{+}_{\omega\geq\omega_{1}}X^{h}(\omega)\in d\omega_{2}]

where Xh(ω):=S(ω)+rh(ω)X^{h}(\omega):=S^{\downarrow}(\omega)+r^{h}(\omega) and

  • (S(ω))ωω1(S^{\downarrow}(\omega))_{\omega\geq\omega_{1}} is the Markov process (S(ω):=W(ω)ϕ(ω))ωω1(S(\omega):=W(\omega)-\phi(\omega))_{\omega\geq\omega_{1}} started at zero and Doob-conditioned to stay negative (i.e to hit -\infty before 0). Precisely, its transition function is given by

    (3.1) [S(t)dy|S(s)=x]=[τ0=|S(t)=y][τ0=|S(s)=x]f(s,x,t,y)dy\mathbb{P}[S^{\downarrow}(t)\in dy|S^{\downarrow}(s)=x]=\frac{\mathbb{P}[\tau_{0}=\infty|S(t)=y]}{\mathbb{P}[\tau_{0}=\infty|S(s)=x]}f(s,x,t,y)dy

    for t>s>ω1t>s>\omega_{1} and x,y<0x,y<0, and where τ0\tau_{0} is the first hitting time of zero of the process SS. The function ff is the transition density of the process SS killed at zero, at time tt and state yy, formally defined as

    [S(t)dy,maxsutS(u)<0|S(s)=x]=f(s,x,t,y)dy\displaystyle\mathbb{P}[S(t)\in dy,\max\limits_{s\leq u\leq t}S(u)<0|S(s)=x]=f(s,x,t,y)dy

    Moreover, the entrance law of SS^{\downarrow} is given by

    (3.2) [S(t)dy]=[τ0=|S(t)=y]x[τ0=|S(s)=x]|x=0xf(ω1,0,t,y)dy\mathbb{P}[S^{\downarrow}(t)\in dy]=\frac{\mathbb{P}[\tau_{0}=\infty|S(t)=y]}{\partial_{x}\mathbb{P}[\tau_{0}=\infty|S(s)=x]_{|x=0}}\partial_{x}f(\omega_{1},0,t,y)dy
  • The function rhr^{h} is defined as rh(ω)=ϕ(ω)ϕ(ωh)+cr^{h}(\omega)=\phi(\omega)-\phi(\omega-h)+c where cc is a constant such that rh(ω1)=0r^{h}(\omega_{1})=0.

Proof.

We have that

[Ψϕ(h)dω2|Ψϕ(0)=ω1]\displaystyle\mathbb{P}[\Psi^{\phi}(h)\in d\omega_{2}|\Psi^{\phi}(0)=\omega_{1}] =[argmaxωω1+(W(ω)ϕ(ωh))dω2|Ψϕ(0)=ω1]\displaystyle=\mathbb{P}[\text{argmax}^{+}_{\omega\geq\omega_{1}}(W(\omega)-\phi(\omega-h))\in d\omega_{2}|\Psi^{\phi}(0)=\omega_{1}]
=[argmaxωω1+(S(ω)S(ω1)+rh(ω))dω2\displaystyle=\mathbb{P}[\text{argmax}^{+}_{\omega\geq\omega_{1}}(S(\omega)-S(\omega_{1})+r^{h}(\omega))\in d\omega_{2}
|argmax+S(ω)=ω1]\displaystyle|\text{argmax}^{+}S(\omega)=\omega_{1}]

Now, using Millar path decomposition of Markov processes when they reach their ultimate maximum, the expression of the transition densities of the post-maximum process in [16][Equation 9] on the process SS, and the spatial homogeneity of the Brownian motion (and thus of SS), we get (3.1). To get the entrance law it suffices to send ss to ω1\omega_{1} and xx to zero. ∎

Let us now introduce some notation to keep our formulas compact.

Notation 3.3.

Denote by

J(s,x)=[τ0=|S(s)=x]=[S(u)<0 for all us|S(s)=x],x<0\displaystyle J(s,x)=\mathbb{P}[\tau_{0}=\infty|S(s)=x]=\mathbb{P}[S(u)<0\text{ for all }u\geq s|S(s)=x],~{}~{}x<0

and define

j(s,x)=xJ(s,x),s,x<0\displaystyle j(s,x)=\frac{\partial}{\partial x}J(s,x),~{}~{}s\in\mathbb{R},~{}~{}x<0
j(s)=limx0xJ(s,x),s\displaystyle j(s)=\lim_{x\uparrow 0}\frac{\partial}{\partial x}J(s,x),~{}~{}s\in\mathbb{R}

Also denote

Φ(s,x,ω)=[τ0dω|S(s)=x]dω,s<ω,x\displaystyle\Phi(s,x,\omega)=\frac{\mathbb{P}[\tau_{0}\in d\omega|S(s)=x]}{d\omega},~{}~{}s<\omega,~{}~{}x\in\mathbb{R}

Furthermore, let S~\tilde{S} be the process defined as (S~(ω):=W(ω)ϕ(ω))ω(\tilde{S}(\omega):=W(\omega)-\phi(-\omega))_{\omega\in\mathbb{R}}. We define f~\tilde{f} and Φ~\tilde{\Phi} analogously.

With this notation, the entrance law of the process SS^{\downarrow} is expressed as

(3.3) [S(t)dy]=J(t,y)j(s)xf(ω1,0,t,y)dy,t>ω1 and y<0\mathbb{P}[S^{\downarrow}(t)\in dy]=\frac{J(t,y)}{j(s)}\partial_{x}f(\omega_{1},0,t,y)dy,~{}~{}t>\omega_{1}\text{ and }y<0

The next result will allow us to recover the transition function of the process Ψϕ\Psi^{\phi}.

Theorem 3.4.

Let ω1<ω2\omega_{1}<\omega_{2} and x(0,rh(ω2))x^{*}\in(0,r^{h}(\omega_{2})). Define ω(ω1,ω2)\omega^{*}\in(\omega_{1},\omega_{2}) to be the unique point such that rh(ω)=xr_{h}(\omega^{*})=x^{*} (such a time exists because of the strict convexity of ϕ\phi that makes rhr_{h} strictly increasing). Then we have that

[argmaxωω1+Xh(ω)dω2,maxωω1Xh(ω)dx]dω2dx=\displaystyle\frac{\mathbb{P}[\text{argmax}^{+}_{\omega\geq\omega_{1}}X^{h}(\omega)\in d\omega_{2},\max\limits_{\omega\geq\omega_{1}}X^{h}(\omega)\in dx^{*}]}{d\omega_{2}dx^{*}}=
2xj(ω2h)j(ω1)Φ(ωh,yx,ω2h)Φ~(ω,yx,ω1)𝑑y\displaystyle 2\int_{-\infty}^{x_{*}}\frac{j(\omega_{2}-h)}{j(\omega_{1})}\Phi(\omega^{*}-h,y-x^{*},\omega_{2}-h)\tilde{\Phi}(-\omega^{*},y-x^{*},-\omega_{1})dy

Before proving this theorem, we will state a lemma that links the joint distribution of the maximum of the diffusion SS and its location with the functionals ff and JJ .

Lemma 3.5.

Let MM and ωM\omega_{M} be respectively the maximum of the process (S(ω))ωs(S(\omega))_{\omega\geq s} and its location, we have then that

(3.4) [ωMdt,Mdz|S(s)=x]dtdz=12j(t)yf(s,xz,t,0)=j(t)Φ(s,xz,t)\frac{\mathbb{P}[\omega_{M}\in dt,M\in dz|S(s)=x]}{dtdz}=\frac{1}{2}j(t)\partial_{y}f(s,x-z,t,0)=-j(t)\Phi(s,x-z,t)
Proof.

We have by the Markov property that

[ωM>t,Mdz|S(s)=x]\displaystyle\mathbb{P}[\omega_{M}>t,M\in dz|S(s)=x] =[maxsutS(u)<z,maxutS(u)dz|S(s)=x]\displaystyle=\mathbb{P}[\max\limits_{s\leq u\leq t}S(u)<z,\max\limits_{u\geq t}S(u)\in dz|S(s)=x]
=zf(s,xz,t,yz)[maxutS(u)dz|S(t)=y]𝑑y\displaystyle=\int_{-\infty}^{z}f(s,x-z,t,y-z)\mathbb{P}[\max_{u\geq t}S(u)\in dz|S(t)=y]dy

Now we see that

[maxutS(u)dz|S(t)=y]=J(t,yzdz)J(t,yz)=j(t,yz)dz\mathbb{P}[\max_{u\geq t}S(u)\in dz|S(t)=y]=J(t,y-z-dz)-J(t,y-z)=-j(t,y-z)dz

Hence

[ωM>t,Mdz|S(s)=x]=0f(s,xz,t,y)j(t,y)𝑑y𝑑z\mathbb{P}[\omega_{M}>t,M\in dz|S(s)=x]=-\int_{-\infty}^{0}f(s,x-z,t,y)j(t,y)dydz

Thus

(3.5) [ωMdt,Mdz]dzdt=t0f(s,xz,t,y)j(t,y)𝑑y\frac{\mathbb{P}[\omega_{M}\in dt,M\in dz]}{dzdt}=\frac{\partial}{\partial t}\int_{-\infty}^{0}f(s,x-z,t,y)j(t,y)dy

Now, by Kolmogorov forward and backward equations on the diffusion SS we have that

tf=ϕ(t)yf+12y2f\partial_{t}f=\phi^{\prime}(t)\partial_{y}f+\frac{1}{2}\partial^{2}_{y}f

and

tj=ϕ(t)yj12y2j\partial_{t}j=\phi^{\prime}(t)\partial_{y}j-\frac{1}{2}\partial^{2}_{y}j

By interchanging the time partial derivative and the integral sign in (3.5), we find by integration by parts

t0f(s,xz,t,y)j(t,y)𝑑y=ϕ(t)[fj]0+12[jyffyj]0\displaystyle\frac{\partial}{\partial t}\int_{-\infty}^{0}f(s,x-z,t,y)j(t,y)dy=\phi^{\prime}(t)\left[fj\right]_{-\infty}^{0}+\frac{1}{2}\left[j\partial_{y}f-f\partial_{y}j\right]_{-\infty}^{0}

Now it suffices to see that ff vanishes at both zero and infinity, from which the first equality follows. For the second equality, it suffices to see that

[τ0>t|S(s)=x]=0f(s,x,t,y)𝑑y\displaystyle\mathbb{P}[\tau_{0}>t|S(s)=x]=\int_{-\infty}^{0}f(s,x,t,y)dy

Differentiating with respect to time and using the Kolmogorov forward equation in the same fashion as was done before gives the result. ∎

Remark 3.6.

All these differentiations and integrations by parts are justified by the fact that ff and jj are sufficiently smooth and integrable away from {t=s}\{t=s\}. This fact will be proved in the next section.

Proof of Theorem 3.4.

We have that

[argmax+ωω1Xh(ω)dω2,maxωω1Xh(ω)dx]=\displaystyle\mathbb{P}[\text{argmax}^{+}_{\omega\geq\omega_{1}}X^{h}(\omega)\in d\omega_{2},\max\limits_{\omega\geq\omega_{1}}X^{h}(\omega)\in dx^{*}]=
x[Xh(ω)dy,argmax+ωω1Xh(ω)dω2,maxωω1Xh(ω)dx]\displaystyle\int_{-\infty}^{x^{*}}\mathbb{P}[X^{h}(\omega^{*})\in dy,\text{argmax}^{+}_{\omega\geq\omega_{1}}X^{h}(\omega)\in d\omega_{2},\max\limits_{\omega\geq\omega_{1}}X^{h}(\omega)\in dx^{*}]

Because for ω[ω1,ω)\omega\in[\omega_{1},\omega^{*}), we have that Xh(ω)rh(ω)<xX^{h}(\omega)\leq r_{h}(\omega)<x^{*}, then by the Markov property we get that

[Xh(ω)dy,argmax+ωω1Xh(ω)dω2,maxωω1Xh(ω)dx]=\displaystyle\mathbb{P}[X^{h}(\omega^{*})\in dy,\text{argmax}^{+}_{\omega\geq\omega_{1}}X^{h}(\omega)\in d\omega_{2},\max\limits_{\omega\geq\omega_{1}}X^{h}(\omega)\in dx^{*}]=
[Xh(ω)dy][argmax+ωωXh(ω)dω2,maxωωXh(ω)dx|Xh(ω)=y]\displaystyle\mathbb{P}[X^{h}(\omega^{*})\in dy]\mathbb{P}[\text{argmax}^{+}_{\omega\geq\omega^{*}}X^{h}(\omega)\in d\omega_{2},\max\limits_{\omega\geq\omega^{*}}X^{h}(\omega)\in dx^{*}|X^{h}(\omega^{*})=y]

Let us focus first on the second term of this product. The law of the Markov process XhX^{h} is that of the process S+rhS+r^{h} conditioned to stay below rhr_{h}. However, when XhX^{h} starts from the state y<xy<x^{*} at time ω\omega^{*}, the event we condition on has positive probability and hence it is just the naive conditioning. Thus, we can write

[argmax+ωωXh(ω)dω2,maxωωXh(ω)dx|Xh(ω)=y]=\displaystyle\mathbb{P}\left[\text{argmax}^{+}_{\omega\geq\omega^{*}}X^{h}(\omega)\in d\omega_{2},\max\limits_{\omega\geq\omega^{*}}X^{h}(\omega)\in dx^{*}|X^{h}(\omega^{*})=y\right]=
[argmax+ωωS(ω)+rh(ω)dω2,maxωωS(ω)+rh(ω)dx\displaystyle\mathbb{P}\left[\text{argmax}^{+}_{\omega\geq\omega^{*}}S(\omega)+r^{h}(\omega)\in d\omega_{2},\max\limits_{\omega\geq\omega^{*}}S(\omega)+r^{h}(\omega)\in dx^{*}\right.
|S(ω)=yx,S(ω)0 for all ωω]\displaystyle\left.|S(\omega^{*})=y-x^{*},S(\omega)\leq 0\text{ for all }\omega\geq\omega^{*}\right]

This probability is equal to the ratio of this probability

1=[argmax+ωωS(ω)+rh(ω)dω2,maxωωS(ω)+rh(ω)dx,\displaystyle\mathbb{P}_{1}=\mathbb{P}\left[\text{argmax}^{+}_{\omega\geq\omega^{*}}S(\omega)+r^{h}(\omega)\in d\omega_{2},\max\limits_{\omega\geq\omega^{*}}S(\omega)+r^{h}(\omega)\in dx^{*},\right.
S(ω)0 for all ωω|S(ω)=yx]\displaystyle\left.S(\omega)\leq 0\text{ for all }\omega\geq\omega^{*}|S(\omega^{*})=y-x^{*}\right]

over the probability

2=[S(ω)0 for all ωω|S(ω)=yx]\displaystyle\mathbb{P}_{2}=\mathbb{P}\left[S(\omega)\leq 0\text{ for all }\omega\geq\omega^{*}|S(\omega^{*})=y-x^{*}\right]

For the first probability 1\mathbb{P}_{1}, notice that on the event that {maxωωS(ω)+rh(ω)dx}\{\max\limits_{\omega\geq\omega^{*}}S(\omega)+r^{h}(\omega)\in dx^{*}\}, we always have that S(ω)0 for all ωωS(\omega)\leq 0\text{ for all }\omega\geq\omega^{*}, because rh(ω)xr^{h}(\omega)\geq x^{*} for ωω\omega\geq\omega^{*}. Thus

1=[argmax+ωωS(ω)+rh(ω)dω2,maxωωS(ω)+rh(ω)dx|S(ω)=yx]\displaystyle\mathbb{P}_{1}=\mathbb{P}\left[\text{argmax}^{+}_{\omega\geq\omega^{*}}S(\omega)+r^{h}(\omega)\in d\omega_{2},\max\limits_{\omega\geq\omega^{*}}S(\omega)+r^{h}(\omega)\in dx^{*}\right.\left.|S(\omega^{*})=y-x^{*}\right]

Now we have that

S(ω)+rh(ω)=W(ω)ϕ(ωh)+c,ωω\displaystyle S(\omega)+r^{h}(\omega)=W(\omega)-\phi(\omega-h)+c~{}~{},~{}~{}\omega\geq\omega^{*}

Hence

(S(ω)+rh(ω)|S(ω)=yx)ωω=d(S(ωh)|S(ωh)=y)ωω\displaystyle(S(\omega)+r^{h}(\omega)|S(\omega^{*})=y-x^{*})_{\omega\geq\omega^{*}}\overset{\mathrm{d}}{=}(S(\omega-h)|S(\omega^{*}-h)=y)_{\omega\geq\omega^{*}}

Thus by using Lemma 3.5 for s=ωhs=\omega^{*}-h and x=yxx=y-x^{*}, we get that

1=j(ω2h)Φ(ωh,yx,ω2h)dω2dx\displaystyle\mathbb{P}_{1}=-j(\omega_{2}-h)\Phi(\omega^{*}-h,y-x^{*},\omega_{2}-h)d\omega_{2}dx^{*}

Therefore

(3.6) 12=j(ω2h)Φ(ωh,yx,ω2h)J(ω,yx)dω2dx\frac{\mathbb{P}_{1}}{\mathbb{P}_{2}}=-\frac{j(\omega_{2}-h)\Phi(\omega^{*}-h,y-x^{*},\omega_{2}-h)}{J(\omega^{*},y-x^{*})}d\omega_{2}dx^{*}

Finally for the first term [Xh(ω)dy]\mathbb{P}[X^{h}(\omega^{*})\in dy], we have that

[Xh(ω)dy]\displaystyle\mathbb{P}[X^{h}(\omega^{*})\in dy] =[S(ω)d(yrh(ω))]\displaystyle=\mathbb{P}[S^{\downarrow}(\omega^{*})\in d(y-r^{h}(\omega^{*}))]
=[S(ω)d(yx)]\displaystyle=\mathbb{P}[S^{\downarrow}(\omega^{*})\in d(y-x^{*})]
=J(ω,yx)j(ω1)xf(ω1,0,ω,yx)dy\displaystyle=\frac{J(\omega^{*},y-x^{*})}{j(\omega_{1})}\partial_{x}f(\omega_{1},0,\omega^{*},y-x^{*})dy

Now it is not hard to see that we have the following equality

(3.7) f~(s,x,t,y)=f(t,y,s,x)\tilde{f}(s,x,t,y)=f(-t,y,-s,x)

This is true because both those functions verify the same PDE with the same boundary and growth conditions, by combining the backward and forward Kolmogorov equations. Hence

xf(s,x,t,y)=yf~(t,y,s,x)\partial_{x}f(s,x,t,y)=\partial_{y}\tilde{f}(-t,y,-s,x)

Hence, by Lemma 3.5

xf(ω1,0,ω,yx)\displaystyle\partial_{x}f(\omega_{1},0,\omega^{*},y-x^{*}) =yf~(ω,yx,ω1,0)\displaystyle=\partial_{y}\tilde{f}(-\omega^{*},y-x^{*},-\omega_{1},0)
=2Φ~(ω,yx,ω1)\displaystyle=-2\tilde{\Phi}(-\omega^{*},y-x^{*},-\omega_{1})

Thus

(3.8) [Xh(ω)dy|Xh(ω1)=0]=2J(ω,yx)j(ω1)Φ~(ω,yx,ω1)dy\mathbb{P}[X^{h}(\omega^{*})\in dy|X^{h}(\omega_{1})=0]=-2\frac{J(\omega^{*},y-x^{*})}{j(\omega_{1})}\tilde{\Phi}(-\omega^{*},y-x^{*},-\omega_{1})dy

Multiplying equations (3.6) and (3.8) and integrating with respect to yy on (,x)(-\infty,x^{*}) gives the result. ∎

ω\omegayyrhr^{h}XhX^{h}ω2\omega_{2}ω\omega^{*}xx^{*}ω1\omega_{1}
Figure 2. Path decomposition of XhX^{h} at its maximum

We are ready now to state the main result of this section.

Theorem 3.7.

The transition function of the process Ψϕ\Psi^{\phi} is given by

[Ψϕ(h)dω2|Ψϕ(0)=ω1]=2j(ω2h)j(ω1)×\displaystyle\mathbb{P}[\Psi^{\phi}(h)\in d\omega_{2}|\Psi^{\phi}(0)=\omega_{1}]=2\frac{j(\omega_{2}-h)}{j(\omega_{1})}\times
ω1ω20(rh)(ω)Φ(ωh,y,ω2h)Φ~(ω,y,ω1)dydω\displaystyle\int_{\omega_{1}}^{\omega_{2}}\int_{\infty}^{0}(r^{h})^{\prime}(\omega)\Phi(\omega-h,y,\omega_{2}-h)\tilde{\Phi}(-\omega,y,-\omega_{1})dyd\omega

Moreover, the process Ψϕ\Psi^{\phi} is pure-jump and its generator at zero is given by its action on any test function φCc()\varphi\in C_{c}^{\infty}(\mathbb{R})

𝒜ϕφ(y)=y(φ(z)φ(y))nϕ(y,z)dz\displaystyle\mathcal{A}^{\phi}\varphi(y)=\int_{y}^{\infty}(\varphi(z)-\varphi(y))n^{\phi}(y,z)dz

where

nϕ(y,z)=2j(z)j(y)yz0ϕ(ω)Φ(ω,x,z)Φ~(ω,x,y)dxdω=:j(z)j(y)Kϕ(y,z)\displaystyle n^{\phi}(y,z)=2\frac{j(z)}{j(y)}\int_{y}^{z}\int_{-\infty}^{0}\phi^{\prime\prime}(\omega)\Phi(\omega,x,z)\tilde{\Phi}(-\omega,x,-y)dxd\omega=:\frac{j(z)}{j(y)}K^{\phi}(y,z)
Proof.

By integrating the formula in Theorem 3.4 with respect to xx^{*} between and 0 and rh(ω2)r^{h}(\omega_{2}) (as XhX^{h} is pointwise at most rhr^{h}), we get that

[argmax+ωω1Xh(ω)dω2]=2j(ω2h)j(ω1)0rh(ω2)xΦ(ωh,yx,ω2h)×\displaystyle\mathbb{P}[\text{argmax}^{+}_{\omega\geq\omega_{1}}X^{h}(\omega)\in d\omega_{2}]=2\frac{j(\omega_{2}-h)}{j(\omega_{1})}\int_{0}^{r^{h}(\omega_{2})}\int_{-\infty}^{x^{*}}\Phi(\omega^{*}-h,y-x^{*},\omega_{2}-h)\times
Φ~(ω,yx,ω1)dydx\displaystyle\tilde{\Phi}(-\omega^{*},y-x^{*},-\omega_{1})dydx^{*}

Now it suffices to do the change of variables y=yxy^{\prime}=y-x^{*} and ω=(rh)1(x)\omega=(r^{h})^{-1}(x^{*}) to get the transition density. As for the generator part, it suffices to do the following Taylor expansion for h0h\rightarrow 0

(rh)(ω)=ϕ(ω)h+O(h2)\displaystyle(r^{h})^{\prime}(\omega)=\phi^{\prime\prime}(\omega)h+O(h^{2})

Remark 3.8.

In the next section, we will greatly simplify this expression of the generator by giving explicit formulas of KϕK^{\phi} and jj in Proposition 4.7 and Theorem 4.8 respectively.

4. Regularity of the transition functions and explicit formulas

The goal of this section is to prove the regularity of the transition density f(s,x,t,y)f(s,x,t,y) away from the line {t=s}\{t=s\}, so that we can justify all the operations we did in the previous section and to deduce along the way explicit formulas for the jump kernel of the process Ψϕ\Psi^{\phi}.

Processes such as the three-dimensional Bessel process, the three-dimensional Bessel bridges, and the Brownian motion killed at zero will be mentioned in some of the results of this section. We refer the unfamiliar reader to [17][Chapters 3,6,11] for basic facts about these processes.

The following proposition gives a closed formula for the density ff.

Proposition 4.1.

Let x,y<0x,y<0 and t>st>s, the density ff is given by the formula

f(s,x,t,y)=G(s,x,t,y)exp(ϕ(t)y+ϕ(s)x12st(ϕ(u))2du)×\displaystyle f(s,x,t,y)=G(s,x,t,y)\text{exp}\left(-\phi^{\prime}(t)y+\phi^{\prime}(s)x-\frac{1}{2}\int_{s}^{t}(\phi^{\prime}(u))^{2}du\right)\times
𝔼[exp(stϕ(u)B(u)du)|B(s)=x,B(t)=y]\displaystyle\mathbb{E}\left[\text{exp}\left(-\int_{s}^{t}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=-x,B(t)=-y\right]

where BB is a three-dimensional Bessel process, and GG is the transition density function of the Brownian motion killed at zero, given explicitly by

G(s,x,t,y)=12π(ts)(e(xy)22(ts)e(x+y)22(ts))G(s,x,t,y)=\frac{1}{\sqrt{2\pi(t-s)}}\left(e^{-\frac{(x-y)^{2}}{2(t-s)}}-e^{-\frac{(x+y)^{2}}{2(t-s)}}\right)
Proof.

The process SS can be expressed as

S(t)=W(t)ϕ(t)=W(t)stϕ(u)duϕ(s)S(t)=W(t)-\phi(t)=W(t)-\int_{s}^{t}\phi^{\prime}(u)du-\phi(s)

Thus by Girsanov theorem, SS is a Brownian motion under the measure \mathbb{Q} with Radon-Nikodym derivative given by

dd|t=exp(stϕ(u)dWu12st(ϕ(u))2du)\frac{d\mathbb{Q}}{d\mathbb{P}}_{|\mathcal{F}_{t}}=\text{exp}\left(\int_{s}^{t}\phi^{\prime}(u)dW_{u}-\frac{1}{2}\int_{s}^{t}(\phi^{\prime}(u))^{2}du\right)

where t:=σ{W(u):sut}\mathcal{F}_{t}:=\sigma\{W(u):s\leq u\leq t\} is the canonical filtration of WW. Thus for any function FF we have that

𝔼[F(S(t))1{maxsutS(u)<0}|S(s)=x]=𝔼[Z(t)F(W(t))1{maxsutW(u)<0}|W(s)=x]\mathbb{E}[F(S(t))\mathbbold{1}_{\{\max\limits_{s\leq u\leq t}S(u)<0\}}|S(s)=x]=\mathbb{E}[Z(t)F(W(t))\mathbbold{1}_{\{\max\limits_{s\leq u\leq t}W(u)<0\}}|W(s)=x]

where

Z(t):=exp(stϕ(u)dWu12stϕ(u)2du)Z(t):=\exp\left(-\int_{s}^{t}\phi^{\prime}(u)dW_{u}-\frac{1}{2}\int_{s}^{t}\phi^{\prime}(u)^{2}du\right)

In particular for F=Fϵ:=12ϵ1[yϵ,y+ϵ]F=F_{\epsilon}:=\frac{1}{2\epsilon}\mathbbold{1}_{[y-\epsilon,y+\epsilon]}, we have that

f(s,x,t,y)\displaystyle f(s,x,t,y) =limϵ0𝔼[Fϵ(S(t))1{maxsutS(u)<0}|S(s)=x]\displaystyle=\lim_{\epsilon\rightarrow 0}\mathbb{E}[F_{\epsilon}(S(t))\mathbbold{1}_{\{\max\limits_{s\leq u\leq t}S(u)<0\}}|S(s)=x]
=limϵ012ϵyϵy+ϵ𝔼[Z(t)1{W(t)dz,maxsutW(u)<0}|W(s)=x]\displaystyle=\lim_{\epsilon\rightarrow 0}\frac{1}{2\epsilon}\int_{y-\epsilon}^{y+\epsilon}\mathbb{E}[Z(t)\mathbbold{1}_{\{W(t)\in dz,\max\limits_{s\leq u\leq t}W(u)<0\}}|W(s)=x]

Now if we denote by WW^{\partial} the Brownian motion killed at zero whose law is defined as

𝔼[F(W(t))|W(s)=x]=𝔼[F(W(t))1{maxsutW(u)<0}|W(s)=x]\displaystyle\mathbb{E}[F(W^{\partial}(t))|W^{\partial}(s)=x]=\mathbb{E}[F(W(t))\mathbbold{1}_{\{\max_{s\leq u\leq t}W(u)<0\}}|W(s)=x]

Thus

f(s,x,t,y)\displaystyle f(s,x,t,y) =limϵ012ϵyϵy+ϵ𝔼[Z(t)|W(t)=y,W(s)=x]pts(x,z)dz\displaystyle=\lim_{\epsilon\rightarrow 0}\frac{1}{2\epsilon}\int_{y-\epsilon}^{y+\epsilon}\mathbb{E}[Z^{\partial}(t)|W^{\partial}(t)=y,W^{\partial}(s)=x]p_{t-s}^{\partial}(x,z)dz
=pts(x,y)𝔼[Z(t)|W(t)=y,W(s)=x]\displaystyle=p_{t-s}^{\partial}(x,y)\mathbb{E}[Z^{\partial}(t)|W^{\partial}(t)=y,W^{\partial}(s)=x]

where ZZ^{\partial} is the same as ZZ with WW replaced by WW^{\partial}, and pt(x,y)p_{t}^{\partial}(x,y) is the transition density function of the process WW^{\partial}. However it is a well-known fact that pts(x,y)=G(s,x,t,y)p_{t-s}^{\partial}(x,y)=G(s,x,t,y), and the law of the Brownian motion killed at zero between ss and tt conditioned on its extreme values is the law of the reflection of a three-dimensional Bessel bridge between (s,x)(s,-x) and (t,y)(t,-y) (as our killed Brownian motion stays negative and the Bessel bridges are by definition positive). Finally, by using an integration by parts we have that

d(B(u)ϕ(u))=ϕ(u)dB(u)+ϕ(u)B(u)dud(B(u)\phi^{\prime}(u))=\phi^{\prime}(u)dB(u)+\phi^{\prime\prime}(u)B(u)du

Integrating between ss and tt, we get the desired result. ∎

Remark 4.2.

From the last proposition, one can readily see that for fixed ss and xx

0f(s,x,t,y)C(t)eA(t)y2 for all y\displaystyle 0\leq f(s,x,t,y)\leq C(t)e^{-A(t)y^{2}}\text{ for all }y

where CC and AA are locally bounded, and AA is locally bounded from below by a positive constant.

Let us now prove that ff is smooth. First of all, one can extend ff to the positive line as well by defining

f(s,x,t,y)=f(s,x,t,y),y>0\displaystyle f(s,x,t,y)=-f(s,x,t,-y),~{}~{}y>0

Then ff verifies in the distribution sense the following PDE (Kolmogorov forward equation)

(4.1) tf122yf=ϕ(t)yf on (t,y)(s,+)×\partial_{t}f-\frac{1}{2}\partial^{2}_{y}f=\phi^{\prime}(t)\partial_{y}f\text{ on }(t,y)\in(s,+\infty)\times\mathbb{R}

and with boundary conditions f(s,x,s,)=δxδxf(s,x,s,\cdot)=\delta_{x}-\delta_{-x}, and obviously f(s,x,t,0)=0f(s,x,t,0)=0. Now, it is well-known that the function GG that we defined in Proposition 4.1 verifies the heat equation

tG12y2G=0\displaystyle\partial_{t}G-\frac{1}{2}\partial_{y}^{2}G=0

with the same boundary conditions as ff. Moreover, if one defines the function G^\hat{G} as

G^(s,x,t,y)=12π(ts)e(xy)22(ts)\displaystyle\hat{G}(s,x,t,y)=\frac{1}{\sqrt{2\pi(t-s)}}e^{-\frac{(x-y)^{2}}{2(t-s)}}

it is also a solution for the heat equation but with boundary condition G^(s,x,s,)=δx\hat{G}(s,x,s,\cdot)=\delta_{x}. Thus, in order to study the regularity properties of the solution to (4.1), one might use Duhamel’s principle to get a representation formula for ff. More precisely, we shall prove the following theorem

Theorem 4.3.

Fix s,xs,x\in\mathbb{R}. There exists a function hC([s,+),L1()L())h\in C([s,+\infty),L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R})) (where here L1()L()L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R}) is the space of continuous functions on the real line that are uniformly bounded and absolutely integrable), such that

h(t,y)=stϕ(u)G^(u,z,t,y)zG(s,x,u,z)dzdu\displaystyle h(t,y)=\int_{s}^{t}\int_{\mathbb{R}}\phi^{\prime}(u)\hat{G}(u,z,t,y)\partial_{z}G(s,x,u,z)dzdu
stϕ(u)zG^(u,z,t,y)h(u,z)dzdu\displaystyle-\int_{s}^{t}\int_{\mathbb{R}}\phi^{\prime}(u)\partial_{z}\hat{G}(u,z,t,y)h(u,z)dzdu

Furthermore, hh is smooth.

Proof.

Let us fix T>sT>s. Define the functional ΞT\Xi^{T} from 𝒞T:=C([s,T],L1()L())\mathcal{C}_{T}:=C([s,T],L^{1}(\mathbb{R})\cap L^{\infty}(\mathbb{R})) into itself equipped with the norm

||h||𝒞T:=supstT||h(t)||L+||h(t)||L1||h||_{\mathcal{C}_{T}}:=\sup_{s\leq t\leq T}||h(t)||_{L^{\infty}}+||h(t)||_{L^{1}}

by

ΞT[h](t,y)=stϕ(u)G^(u,z,t,y)zG(s,x,u,z)dzdu\displaystyle\Xi^{T}[h](t,y)=\int_{s}^{t}\int_{\mathbb{R}}\phi^{\prime}(u)\hat{G}(u,z,t,y)\partial_{z}G(s,x,u,z)dzdu
stϕ(u)zG^(u,z,t,y)h(u,z)dzdu\displaystyle-\int_{s}^{t}\int_{\mathbb{R}}\phi^{\prime}(u)\partial_{z}\hat{G}(u,z,t,y)h(u,z)dzdu

It is clear that ΞT\Xi^{T} sends 𝒞T\mathcal{C}_{T} to itself due to the growth rate of the Green functions GG and G^\hat{G} at infinity in space. Moreover we have that for any two functions hh and h~\tilde{h} in 𝒞T\mathcal{C}_{T}

||ΞT[h](t,.)ΞT[h~](t,.)||L1st|ϕ(u)|du|h(u,z)h~(u,z)|dz|zG^(u,z,t,y)|dy\displaystyle||\Xi^{T}[h](t,.)-\Xi^{T}[\tilde{h}](t,.)||_{L^{1}}\leq\int_{s}^{t}|\phi^{\prime}(u)|du\int_{\mathbb{R}}|h(u,z)-\tilde{h}(u,z)|dz\int_{\mathbb{R}}|\partial_{z}\hat{G}(u,z,t,y)|dy

Now we see that

zG^(u,z,t,y)=yz2π(tu)3e(yz)22(tu)\displaystyle\partial_{z}\hat{G}(u,z,t,y)=\frac{y-z}{\sqrt{2\pi(t-u)^{3}}}e^{-\frac{(y-z)^{2}}{2(t-u)}}

Hence

|zG^(u,z,t,y)|dy22π(tu)30ωeω22(tu)dω=22π(tu)\displaystyle\int_{\mathbb{R}}|\partial_{z}\hat{G}(u,z,t,y)|dy\leq\frac{2}{\sqrt{2\pi(t-u)^{3}}}\int_{0}^{\infty}\omega e^{-\frac{\omega^{2}}{2(t-u)}}d\omega=\frac{2}{\sqrt{2\pi(t-u)}}

Thus

||ΞT[h](t,.)ΞT[h~](t,.)||L14Tssupu[s,T]|ϕ(u)|π||hh~||𝒞T\displaystyle||\Xi^{T}[h](t,.)-\Xi^{T}[\tilde{h}](t,.)||_{L^{1}}\leq\frac{4\sqrt{T-s}\sup_{u\in[s,T]}|\phi^{\prime}(u)|}{\sqrt{\pi}}||h-\tilde{h}||_{\mathcal{C}_{T}}

A similar bound is found for the LL^{\infty} norm. Thus, for TT close enough to ss, the operator ΞT\Xi^{T} becomes a contraction, and thus by Picard theorem, it admits a unique fixed point.
Now define

T=sup{Ts:h𝒞T such that ΞT[h]=h}\displaystyle T^{*}=\sup\{T\geq s:\exists h\in\mathcal{C}_{T}\text{ such that }\Xi^{T}[h]=h\}

Suppose that T<T^{*}<\infty, then it is easy to see by Gronwall inequality that for any sequence (tm)m(t_{m})_{m\in\mathbb{N}} such that tmTt_{m}\uparrow T^{*}, the sequence (h(tm,))m(h(t_{m},\cdot))_{m\in\mathbb{N}} is Cauchy in L()L1()L^{\infty}(\mathbb{R})\cap L^{1}(\mathbb{R}) and thus converge strongly to a unique limit that we denote h(T,)h(T^{*},\cdot). This extension thus belongs to 𝒞T\mathcal{C}_{T^{*}}. However, for small ϵ>0\epsilon>0, one can further extend the fixed point hh to 𝒞T+ϵ\mathcal{C}_{T^{*}+\epsilon} by the same contraction argument. This contradicts the definition of TT^{*}, and thus T=T^{*}=\infty from which follow the existence of a global solution. The smoothness of hh follows readily from that of the Green function G^\hat{G} and the dominated convergence theorem. ∎

We are now ready to prove the following result

Theorem 4.4.

The function fGf-G is everywhere smooth in the variables (t,y)(t,y), in particular the function ff is smooth away from {t=s}\{t=s\}.

Proof.

Define the function qq by

q(s,x,t,y)=h(t,y)+G(s,x,t,y)\displaystyle q(s,x,t,y)=h(t,y)+G(s,x,t,y)

where hh is the global solution from Theorem 4.3. By integration by parts we have that

q(s,x,t,y)=G(s,x,t,y)+stϕ(u)G^(u,z,t,y)zG(s,x,u,z)dzdu+\displaystyle q(s,x,t,y)=G(s,x,t,y)+\int_{s}^{t}\int_{\mathbb{R}}\phi^{\prime}(u)\hat{G}(u,z,t,y)\partial_{z}G(s,x,u,z)dzdu+
stϕ(u)G^(u,z,t,y)zh(u,z)dudz\displaystyle\int_{s}^{t}\int_{\mathbb{R}}\phi^{\prime}(u)\hat{G}(u,z,t,y)\partial_{z}h(u,z)dudz
=G(s,x,t,y)+sϕ(u)1{t(u,+)}G^(u,z,t,y)zq(u,z)dudz\displaystyle=G(s,x,t,y)+\int_{s}^{\infty}\int_{\mathbb{R}}\phi^{\prime}(u)\mathbbold{1}_{\{t\in(u,+\infty)\}}\hat{G}(u,z,t,y)\partial_{z}q(u,z)dudz

Now it suffices to see that

(t122y)(1t(u,+)G^(u,z,t,y))=δ0(tu)G^(u,z,u,y)=δ0(tu)δ0(yz)\displaystyle(\partial_{t}-\frac{1}{2}\partial^{2}_{y})(\mathbbold{1}_{t\in(u,+\infty)}\hat{G}(u,z,t,y))=\delta_{0}(t-u)\hat{G}(u,z,u,y)=\delta_{0}(t-u)\delta_{0}(y-z)

and thus the function qq verifies the PDE (4.1) with the boundary conditions q(s,x,s,.)=δxδxq(s,x,s,.)=\delta_{x}-\delta_{-x}. The result now would follow if we can prove that f=qf=q. Consider the function v:=fqv:=f-q, it verifies the PDE (4.1) with vanishing initial condition. The growth condition of vv at infinity in space ensures that vv can be viewed as a tempered distribution. By taking the Fourier transform in space in the PDE (4.1) we get that

tv^(t,k)=(12k2+iϕ(t)k)v^(t,k)\displaystyle\partial_{t}\hat{v}(t,k)=\left(-\frac{1}{2}k^{2}+i\phi^{\prime}(t)k\right)\hat{v}(t,k)

Thus

t(v^(t,k)e12k2tiϕ(t)k)=0\displaystyle\partial_{t}(\hat{v}(t,k)e^{\frac{1}{2}k^{2}t-i\phi(t)k})=0

which means that the distribution v^(t,k)e12k2tiϕ(t)k\hat{v}(t,k)e^{\frac{1}{2}k^{2}t-i\phi(t)k} is constant along the time variable tt. Moreover, we also have that

limtsv(t,)=0\displaystyle\lim_{t\rightarrow s}v(t,\cdot)=0

in the tempered distribution sense. Indeed for any φ\varphi in the Schwartz space 𝒮()\mathcal{S}(\mathbb{R}), if we denote by SS^{\partial} is the diffusion SS killed at zero we have that

limtsφ(y)v(t,y)dy=limts[(𝔼[φ(S(t))|S(s)=x]𝔼[φ(W(t))|W(s)=x])\displaystyle\lim_{t\to s}\int_{\mathbb{R}}\varphi(y)v(t,y)dy=\lim_{t\to s}\left[\left(\mathbb{E}[\varphi(S^{\partial}(t))|S^{\partial}(s)=x]-\mathbb{E}[\varphi(W^{\partial}(t))|W^{\partial}(s)=x]\right)\right.
(𝔼[φ(S(t))|S(s)=x]𝔼[φ(W(t))|W(s)=x])\displaystyle\left.-\left(\mathbb{E}[\varphi(-S^{\partial}(t))|S^{\partial}(s)=x]-\mathbb{E}[\varphi(W^{\partial}(t))|W^{\partial}(s)=-x]\right)\right.
φ(y)h(t,y)dy]=0\displaystyle\left.-\int_{\mathbb{R}}\varphi(y)h(t,y)dy\right]=0

as h(s,)=0h(s,\cdot)=0 and by using the dominated convergence theorem. Thus by continuity of the Fourier transform, one deduces that vv is zero everywhere, and hence q=fq=f as desired. ∎

Let us introduce now a function that is going to play a fundamental role in our calculations. Define gg by

(4.2) g(s,x,t,y)=G(s,x,t,y)𝔼[exp(stϕ(u)B(u)du)|B(s)=x,B(t)=y]g(s,x,t,y)=G(s,x,t,y)\mathbb{E}\left[\text{exp}\left(-\int_{s}^{t}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=-x,B(t)=-y\right]

for x,y<0x,y<0 and tst\geq s, where BB is a three-dimensional Bessel process. Because ff is smooth away from {t=s}\{t=s\}, the same holds for gg. We have then the following lemma.

Lemma 4.5.

The function gg verifies the following PDE

(4.3) tg=122yg+ϕ(t)yg\partial_{t}g=\frac{1}{2}\partial^{2}_{y}g+\phi^{\prime\prime}(t)yg

for (t,y)(s,+)×(,0)(t,y)\in(s,+\infty)\times(-\infty,0).

Proof.

We can replace the Bessel process BB by the Brownian motion killed at zero WW^{\partial} in the expression of gg in (4.2) for the same reasons we gave earlier. Now let φCc((s,+)×(,0))\varphi\in C_{c}^{\infty}((s,+\infty)\times(-\infty,0)) be a test function. We apply Ito formula to the following semi-martingale

Y(t)=φ(t,W(t))exp(stϕ(u)W(u)du)\displaystyle Y(t)=\varphi(t,W(t))\text{exp}\left(\int_{s}^{t}\phi^{\prime\prime}(u)W(u)du\right)

where WW is a Brownian motion started at xx. We get then

dY(t)=yφ(t,W(t))exp(stϕ(u)W(u)du)dW(t)+\displaystyle dY(t)=\partial_{y}\varphi(t,W(t))\text{exp}\left(\int_{s}^{t}\phi^{\prime\prime}(u)W(u)du\right)dW(t)+
(tφ(t,W(t))+122yφ(t,W(t))+φ(t,W(t))ϕ(t)W(t))exp(stϕ(u)W(u)du)dt\displaystyle\left(\partial_{t}\varphi(t,W(t))+\frac{1}{2}\partial^{2}_{y}\varphi(t,W(t))+\varphi(t,W(t))\phi^{\prime\prime}(t)W(t)\right)\text{exp}\left(\int_{s}^{t}\phi^{\prime\prime}(u)W(u)du\right)dt

We integrate between ss and tτ0t\wedge\tau_{0} (where τ0\tau_{0} is the first hitting time of zero of WW). As the first term is a bounded local martingale (and hence a true martingale), by taking the expectation we get that

𝔼[φ(tτ0,W(tτ0))]=𝔼[stτ0(tφ(u,W(u))+122yφ(u,W(u))\displaystyle\mathbb{E}\left[\varphi(t\wedge\tau_{0},W(t\wedge\tau_{0}))]=\mathbb{E}[\int_{s}^{t\wedge\tau_{0}}\left(\partial_{t}\varphi(u,W(u))+\frac{1}{2}\partial^{2}_{y}\varphi(u,W(u))\right.\right.
+φ(u,W(u))ϕ(u)W(u))exp(suϕ(ω)W(ω)dω)du]\displaystyle\left.\left.+\varphi(u,W(u))\phi^{\prime\prime}(u)W(u)\right)\text{exp}\left(\int_{s}^{u}\phi^{\prime\prime}(\omega)W(\omega)d\omega\right)du\right]

Therefore

𝔼[φ(tτ0,W(tτ0))]=𝔼[st1{maxszuW(z)<0}(tφ(u,W(u))+122yφ(u,W(u))\displaystyle\mathbb{E}\left[\varphi(t\wedge\tau_{0},W(t\wedge\tau_{0}))]=\mathbb{E}[\int_{s}^{t}\mathbbold{1}_{\{\max\limits_{s\leq z\leq u}W(z)<0\}}\left(\partial_{t}\varphi(u,W(u))+\frac{1}{2}\partial^{2}_{y}\varphi(u,W(u))\right.\right.
+φ(u,W(u))ϕ(u)W(u))exp(suϕ(ω)W(ω)dω)du]\displaystyle\left.\left.+\varphi(u,W(u))\phi^{\prime\prime}(u)W(u)\right)\text{exp}\left(\int_{s}^{u}\phi^{\prime\prime}(\omega)W(\omega)d\omega\right)du\right]
=st𝔼[(tφ(u,W(u))+122yφ(u,W(u))\displaystyle=\int_{s}^{t}\mathbb{E}\left[\left(\partial_{t}\varphi(u,W^{\partial}(u))+\frac{1}{2}\partial^{2}_{y}\varphi(u,W^{\partial}(u))\right.\right.
+φ(u,W(u))ϕ(u)W(u))exp(suϕ(ω)W(ω)dω)]du\displaystyle\left.\left.+\varphi(u,W^{\partial}(u))\phi^{\prime\prime}(u)W^{\partial}(u)\right)\text{exp}\left(\int_{s}^{u}\phi^{\prime\prime}(\omega)W^{\partial}(\omega)d\omega\right)\right]du

By sending tt\rightarrow\infty and conditioning on the value of W(u)W^{\partial}(u), we get

s0(tφ(u,y)+122yφ(u,y)+ϕ(u)yφ(u,y))g(u,y)dydu=0\displaystyle\int_{s}^{\infty}\int_{-\infty}^{0}\left(\partial_{t}\varphi(u,y)+\frac{1}{2}\partial^{2}_{y}\varphi(u,y)+\phi^{\prime\prime}(u)y\varphi(u,y)\right)g(u,y)dydu=0

Thus we get the PDE in the distribution sense, but also in the classical sense because gg is smooth on the interior of its domain. ∎

We give now an explicit formula for the functional Φ\Phi that was introduced in the previous section.

Proposition 4.6.

The function Φ\Phi can be expressed as

Φ(s,x,t)=x2π(ts)3ex22(ts)exp(ϕ(s)x12st(ϕ(u))2du)×\displaystyle\Phi(s,x,t)=\frac{-x}{\sqrt{2\pi(t-s)^{3}}}e^{-\frac{x^{2}}{2(t-s)}}\text{exp}\left(\phi^{\prime}(s)x-\frac{1}{2}\int_{s}^{t}(\phi^{\prime}(u))^{2}du\right)\times
𝔼(s,x)(t,0)[exp(stϕ(u)Bbr(u)du)]\displaystyle\mathbb{E}^{(s,-x)\rightarrow(t,0)}\left[\text{exp}\left(-\int_{s}^{t}\phi^{\prime\prime}(u)B^{br}(u)du\right)\right]

for s<ts<t and x<0x<0. BbrB^{br} here is a three-dimensional Bessel bridge from (s,x)(s,-x) to (t,0)(t,0).

Proof.

From Lemma 3.5, we have that

Φ(s,x,t)=12yf(s,x,t,0)\displaystyle\Phi(s,x,t)=-\frac{1}{2}\partial_{y}f(s,x,t,0)

Since

f(s,x,t,y)=exp(ϕ(t)y+ϕ(s)x12st(ϕ(u))2du)g(s,x,t,y)\displaystyle f(s,x,t,y)=\text{exp}\left(-\phi^{\prime}(t)y+\phi^{\prime}(s)x-\frac{1}{2}\int_{s}^{t}(\phi^{\prime}(u))^{2}du\right)g(s,x,t,y)

and

yg(s,x,t,0)=limy0yG(s,x,t,y)𝔼[exp(stϕ(u)B(u)du)|B(s)=x,B(t)=y]+limy0G(s,x,t,y)y𝔼[exp(stϕ(u)B(u)du)|B(s)=x,B(t)=y]\displaystyle\begin{split}\partial_{y}g(s,x,t,0)=\lim_{y\uparrow 0}\partial_{y}G(s,x,t,y)\mathbb{E}\left[\text{exp}\left(-\int_{s}^{t}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=-x,B(t)=-y\right]+\\ \lim_{y\uparrow 0}G(s,x,t,y)\partial_{y}\mathbb{E}\left[\text{exp}\left(-\int_{s}^{t}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=-x,B(t)=-y\right]\end{split}

it suffices to prove that

limy0y𝔼[exp(stϕ(u)B(u)du)|B(s)=x,B(t)=y]<\displaystyle\lim_{y\uparrow 0}\partial_{y}\mathbb{E}\left[\text{exp}\left(-\int_{s}^{t}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=-x,B(t)=-y\right]<\infty

as G(s,x,t,0)=0G(s,x,t,0)=0. We have by Hopital’s rule applied twice

limy0y𝔼[exp(stϕ(u)B(u)du)|B(s)=x,B(t)=y]\displaystyle\lim_{y\uparrow 0}\partial_{y}\mathbb{E}\left[\text{exp}\left(-\int_{s}^{t}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=-x,B(t)=-y\right] =limy0(yg)G(yG)gG2\displaystyle=\lim_{y\uparrow 0}\frac{(\partial_{y}g)G-(\partial_{y}G)g}{G^{2}}
=limy0(2yg)G(y2G)g2GyG\displaystyle=\lim_{y\uparrow 0}\frac{(\partial^{2}_{y}g)G-(\partial_{y}^{2}G)g}{2G\partial_{y}G}
=limy0y2g2yGlimy0(y2G)g2GyG\displaystyle=\lim_{y\uparrow 0}\frac{\partial_{y}^{2}g}{2\partial_{y}G}-\lim_{y\uparrow 0}\frac{(\partial_{y}^{2}G)g}{2G\partial_{y}G}
=limy0(y2G)g2GyG\displaystyle=-\lim_{y\uparrow 0}\frac{(\partial_{y}^{2}G)g}{2G\partial_{y}G}
=limy0(y3G)g+(y2G)yg2(yG)2+2Gy2G\displaystyle=-\lim_{y\uparrow 0}\frac{(\partial_{y}^{3}G)g+(\partial_{y}^{2}G)\partial_{y}g}{2(\partial_{y}G)^{2}+2G\partial_{y}^{2}G}
limy0y𝔼[exp(stϕ(u)B(u)du)|B(s)=x,B(t)=y]\displaystyle\lim_{y\uparrow 0}\partial_{y}\mathbb{E}\left[\text{exp}\left(-\int_{s}^{t}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=-x,B(t)=-y\right] =0\displaystyle=0

In the fourth line we used the fact that limy0y2g=0\lim_{y\uparrow 0}\partial_{y}^{2}g=0. This follows from the PDE (4.3) verified by gg and the fact that g(s,x,t,0)=0g(s,x,t,0)=0. Moreover because limy0yG0\lim_{y\uparrow 0}\partial_{y}G\neq 0, we can conclude that the limit is equal to zero in the penultimate equality.

To finish the proof, we refer to the fact that the weak limit of the law of the three-dimensional Bessel process conditioned to end at yy when yy goes to zero is that of the corresponding three-dimensional Bessel bridge, and thus the result follows from the expression of the Green function GG. ∎

We are ready to give an explicit formula of the kernel KϕK^{\phi}.

Proposition 4.7.

The kernel KϕK^{\phi} has the following expression

Kϕ(y,z)=ϕ(z)ϕ(y)2π(zy)3exp(12yz(ϕ(u))2du)𝔼[exp(yzϕ(u)e(u)du)]\displaystyle K^{\phi}(y,z)=\frac{\phi^{\prime}(z)-\phi^{\prime}(y)}{\sqrt{2\pi(z-y)^{3}}}\text{exp}\left(-\frac{1}{2}\int_{y}^{z}(\phi^{\prime}(u))^{2}du\right)\mathbb{E}\left[\text{exp}\left(-\int_{y}^{z}\phi^{\prime\prime}(u)\textbf{e}(u)du\right)\right]

for yzy\leq z, where (e(u),yuz)(\textbf{e}(u),y\leq u\leq z) is a Brownian excursion on [y,z][y,z].

Proof.

Recall that KϕK^{\phi} is given by

Kϕ(y,z)=2yz0ϕ(ω)Φ(ω,x,z)Φ~(ω,x,y)dxdω\displaystyle K^{\phi}(y,z)=2\int_{y}^{z}\int_{0}^{\infty}\phi^{\prime\prime}(\omega)\Phi(\omega,-x,z)\tilde{\Phi}(-\omega,-x,-y)dxd\omega

Remember that Φ~\tilde{\Phi} is the same as Φ\Phi with the function ϕ\phi replaced by ϕ()\phi(-\cdot). Hence

Φ(ω,x,z)Φ~(ω,x,y)=x22π(zω)3(ωy)3ex22(zω)x22(ωy)×\displaystyle\Phi(\omega,-x,z)\tilde{\Phi}(-\omega,-x,-y)=\frac{x^{2}}{2\pi\sqrt{(z-\omega)^{3}(\omega-y)^{3}}}e^{-\frac{x^{2}}{2(z-\omega)}-\frac{x^{2}}{2(\omega-y)}}\times
exp(12ωz(ϕ(u))2du12ωy(ϕ(u))2du)×\displaystyle\text{exp}\left(-\frac{1}{2}\int_{\omega}^{z}(\phi^{\prime}(u))^{2}du-\frac{1}{2}\int_{-\omega}^{-y}(\phi^{\prime}(-u))^{2}du\right)\times
𝔼(ω,x)(z,0)[exp(ωzϕ(u)Bbr(u)du)]×\displaystyle\mathbb{E}^{(\omega,x)\rightarrow(z,0)}\left[\text{exp}\left(-\int_{\omega}^{z}\phi^{\prime\prime}(u)B^{br}(u)du\right)\right]\times
𝔼(ω,x)(y,0)[exp(ωyϕ(u)Bbr(u)du)]\displaystyle\mathbb{E}^{(-\omega,x)\rightarrow(-y,0)}\left[\text{exp}\left(-\int_{-\omega}^{-y}\phi^{\prime\prime}(-u)B^{br}(u)du\right)\right]

Consider now a Brownian excursion e on [y,z][y,z], conditionally on its value at ω[y,z]\omega\in[y,z], the two paths (e(u),yuω)(\textbf{e}(u),~{}~{}y\leq u\leq\omega) and (e(u),ωuz)(\textbf{e}(u),~{}~{}\omega\leq u\leq z) are independent, and each path has the distribution of a three-dimensional Bessel bridge. Furthermore, because of the Brownian scaling we have that

(4.4) (e(u),yuz)=d(yzestd(uyzy),yuz)(\textbf{e}(u),y\leq u\leq z)\stackrel{{\scriptstyle d}}{{=}}(\sqrt{y-z}\textbf{e}^{\text{std}}\left(\frac{u-y}{z-y}\right),y\leq u\leq z)

where (estd(u),0u1)(\textbf{e}^{\text{std}}(u),0\leq u\leq 1) is a standard Brownian excursion. Thus, using the fact that

[estd(t)dx]=2x22πt3(1t)3ex22t(1t)dx\displaystyle\mathbb{P}\left[\textbf{e}^{\text{std}}(t)\in dx\right]=\frac{2x^{2}}{\sqrt{2\pi t^{3}(1-t)^{3}}}e^{-\frac{x^{2}}{2t(1-t)}}dx

then it follows that for ω[y,z]\omega\in[y,z]

[e(ω)dx]=2x2(zy)32π(zω)3(ωy)3ex22(zω)x22(ωy)dx\displaystyle\mathbb{P}\left[\textbf{e}(\omega)\in dx\right]=\frac{2x^{2}\sqrt{(z-y)^{3}}}{\sqrt{2\pi(z-\omega)^{3}(\omega-y)^{3}}}e^{-\frac{x^{2}}{2(z-\omega)}-\frac{x^{2}}{2(\omega-y)}}dx

Thus by the time-reversal property of the three-dimensional Bessel bridges we have that

Φ(ω,x,z)Φ~(ω,x,y)=12π(zy)3𝔼[exp(yzϕ(u)e(u)du)|e(ω)=x]×\displaystyle\Phi(\omega,-x,z)\tilde{\Phi}(-\omega,-x,-y)=\frac{1}{\sqrt{2\pi(z-y)^{3}}}\mathbb{E}[\text{exp}\left(-\int_{y}^{z}\phi^{\prime\prime}(u)\textbf{e}(u)du\right)|\textbf{e}(\omega)=x]\times
[e(ω)dx]dx\displaystyle\frac{\mathbb{P}[\textbf{e}(\omega)\in dx]}{dx}

By integrating with respect to xx and ω\omega we get the desired result. ∎

The next theorem gives a closed formula for the function jj.

Theorem 4.8.

Let ss\in\mathbb{R}, define the function lsl^{s} on (0,)(0,\infty) by

ls(u)=exp(12ss+uϕ(z)2dz)𝔼[exp(ss+uϕ(z)e(z)dz)],u>0\displaystyle l^{s}(u)=\text{exp}\left(-\frac{1}{2}\int_{s}^{s+u}\phi^{\prime}(z)^{2}dz\right)\mathbb{E}\left[\text{exp}\left(-\int_{s}^{s+u}\phi^{\prime\prime}(z)\textbf{e}(z)dz\right)\right],~{}~{}u>0

where e is a Brownian excursion on [s,s+u][s,s+u]. Then

j(s)=ϕ(s)+0ls(u)12πu3du\displaystyle j(s)=-\phi^{\prime}(s)+\int_{0}^{\infty}\frac{l^{s}(u)-1}{\sqrt{2\pi u^{3}}}du
Proof.

The function JJ is defined as

J(s,x)\displaystyle J(s,x) =[S(ω)<0 for all ωs|S(s)=x]\displaystyle=\mathbb{P}[S(\omega)<0\text{ for all }\omega\geq s|S(s)=x]
=limt[S(ω)<0 for all sωt|S(s)=x]\displaystyle=\lim_{t\rightarrow\infty}\mathbb{P}[S(\omega)<0\text{ for all }s\leq\omega\leq t|S(s)=x]
=limt0f(s,x,t,y)dy\displaystyle=\lim_{t\rightarrow\infty}\int_{-\infty}^{0}f(s,x,t,y)dy
=eϕ(s)xlimte12st(ϕ(u))2du0eϕ(t)yg(s,x,t,y)dy\displaystyle=e^{\phi^{\prime}(s)x}\lim_{t\rightarrow\infty}e^{-\frac{1}{2}\int_{s}^{t}(\phi^{\prime}(u))^{2}du}\int_{-\infty}^{0}e^{-\phi^{\prime}(t)y}g(s,x,t,y)dy
=eϕ(s)xlimte12st(ϕ(u))2du0eϕ(t)ym(s,x,t,y)dy\displaystyle=e^{\phi^{\prime}(s)x}\lim_{t\rightarrow\infty}e^{-\frac{1}{2}\int_{s}^{t}(\phi^{\prime}(u))^{2}du}\int_{0}^{\infty}e^{\phi^{\prime}(t)y}m(s,x,t,y)dy

where the function mm is defined as

m(s,x,t,y)=g(s,x,t,y)\displaystyle m(s,x,t,y)=g(s,x,t,-y)

It verifies the following PDE

(4.5) tm=122yymϕ(t)ym\partial_{t}m=\frac{1}{2}\partial^{2}_{yy}m-\phi^{\prime\prime}(t)ym

Because of the asymptotic behavior of gg in space at infinity, we can define for every λ\lambda\in\mathbb{R} the Laplace transform

m^(s,x,t,λ)=0eλym(s,x,t,y)dy\displaystyle\hat{m}(s,x,t,\lambda)=\int_{0}^{\infty}e^{\lambda y}m(s,x,t,y)dy

From the representation formula of the function hh (and thus that of gg) in the statement of Theorem 4.3 and the fast decay of the Green functions GG and G^\hat{G} in space, we can interchange the order of differentiation and integration for the Laplace transform m^\hat{m}, hence

tm^\displaystyle\partial_{t}\hat{m} =0eλytm(s,x,t,y)dy\displaystyle=\int_{0}^{\infty}e^{\lambda y}\partial_{t}m(s,x,t,y)dy
=0eλy(122yym(s,x,t,y)ϕ(t)ym(s,x,t,y))dy\displaystyle=\int_{0}^{\infty}e^{\lambda y}\left(\frac{1}{2}\partial^{2}_{yy}m(s,x,t,y)-\phi^{\prime\prime}(t)ym(s,x,t,y)\right)dy
=12[eλyym(s,x,t,y)]0+12λ2m^(s,x,t,λ)ϕ(t)λm^(s,x,t,λ)\displaystyle=\frac{1}{2}\left[e^{\lambda y}\partial_{y}m(s,x,t,y)\right]_{0}^{\infty}+\frac{1}{2}\lambda^{2}\hat{m}(s,x,t,\lambda)-\phi^{\prime\prime}(t)\partial_{\lambda}\hat{m}(s,x,t,\lambda)
=12λ2m^(s,x,t,λ)ϕ(t)λm^(s,x,t,λ)12ym(s,x,t,0)\displaystyle=\frac{1}{2}\lambda^{2}\hat{m}(s,x,t,\lambda)-\phi^{\prime\prime}(t)\partial_{\lambda}\hat{m}(s,x,t,\lambda)-\frac{1}{2}\partial_{y}m(s,x,t,0)

by integration by parts and using the fact that m(s,x,t,0)=0m(s,x,t,0)=0. From the expression of gg we deduce that

ym(s,x,t,0)=yg(s,x,t,0)=2x2π(ts)3ex22(ts)×\displaystyle\partial_{y}m(s,x,t,0)=-\partial_{y}g(s,x,t,0)=\frac{-2x}{\sqrt{2\pi(t-s)^{3}}}e^{-\frac{x^{2}}{2(t-s)}}\times
𝔼[exp(stϕ(u)B(u)du)|B(s)=x,B(t)=0]\displaystyle\mathbb{E}\left[\text{exp}\left(-\int_{s}^{t}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=-x,B(t)=0\right]
=2Φ(s,x,t)exp(ϕ(s)x+12st(ϕ(u))2du)=:2Υ(t)\displaystyle=2\Phi(s,x,t)\text{exp}\left(-\phi^{\prime}(s)x+\frac{1}{2}\int_{s}^{t}(\phi^{\prime}(u))^{2}du\right)=:-2\Upsilon(t)

Since xx and ss are fixed for now, we will often omit them when writing out expressions where they do not vary. Thus, the PDE verified by m^\hat{m} takes the form

tm^+ϕ(t)λm^12λ2m^Υ(t)=0\displaystyle\partial_{t}\hat{m}+\phi^{\prime\prime}(t)\partial_{\lambda}\hat{m}-\frac{1}{2}\lambda^{2}\hat{m}-\Upsilon(t)=0

This is a first order non-linear PDE that can be solved by the method of characteristics. If we denote the variables by x1:=tx_{1}:=t and x2:=λx_{2}:=\lambda and the value of the function z=m^(x1,x2)z=\hat{m}(x_{1},x_{2}), the characteristic ODEs take the form

{x1˙(u)=1x2˙(u)=ϕ(x1(u))z˙(u)=12x22(u)z(u)+Υ(x1(u))\left\{\begin{array}[]{ll}\dot{x_{1}}(u)=1\\ \dot{x_{2}}(u)=\phi^{\prime\prime}(x_{1}(u))\\ \dot{z}(u)=\frac{1}{2}x_{2}^{2}(u)z(u)+\Upsilon(x_{1}(u))\end{array}\right.

We choose the initial conditions such that x1(u)=ux_{1}(u)=u and x2(u)=ϕ(u)+(λϕ(t))x_{2}(u)=\phi^{\prime}(u)+(\lambda-\phi^{\prime}(t)) for usu\geq s. Hence

z˙(u)=12(ϕ(u)+λϕ(t))2z(u)+Υ(u)\dot{z}(u)=\frac{1}{2}(\phi^{\prime}(u)+\lambda-\phi^{\prime}(t))^{2}z(u)+\Upsilon(u)

Introduce the function vλv^{\lambda} defined by

vλ(u)=exp(12su(ϕ(z)+λϕ(t))2dz)v^{\lambda}(u)=\text{exp}\left(-\frac{1}{2}\int_{s}^{u}(\phi^{\prime}(z)+\lambda-\phi^{\prime}(t))^{2}dz\right)

Then it is clear that

(vλz˙)(u)=vλ(u)Υ(u)(\dot{v^{\lambda}z})(u)=v^{\lambda}(u)\Upsilon(u)

In order to avoid the singularity at {t=s}\{t=s\}, we integrate thus between s+ϵs+\epsilon and tt for ϵ>0\epsilon>0 to get that

vλ(t)z(t)vλ(s+ϵ)z(s+ϵ)=s+ϵtvλ(u)Υ(u)duv^{\lambda}(t)z(t)-v^{\lambda}(s+\epsilon)z(s+\epsilon)=\int_{s+\epsilon}^{t}v^{\lambda}(u)\Upsilon(u)du

which is equivalent to

m^(s,x,t,λ)vλ(t)m^(s,x,s+ϵ,ϕ(s+ϵ)+λϕ(t))vλ(s+ϵ)=s+ϵtvλ(u)Υ(u)du\displaystyle\hat{m}(s,x,t,\lambda)v^{\lambda}(t)-\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon)+\lambda-\phi^{\prime}(t))v^{\lambda}(s+\epsilon)=\int_{s+\epsilon}^{t}v^{\lambda}(u)\Upsilon(u)du

By taking λ=ϕ(t)\lambda=\phi^{\prime}(t), we get

(4.6) m^(s,x,t,ϕ(t))e12stϕ(u)2dum^(s,x,s+ϵ,ϕ(s+ϵ))e12ss+ϵϕ(u)2du=s+ϵte12suϕ(ω)2dωΥ(u)du\begin{split}\hat{m}(s,x,t,\phi^{\prime}(t))e^{-\frac{1}{2}\int_{s}^{t}\phi^{\prime}(u)^{2}du}-\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon))e^{-\frac{1}{2}\int_{s}^{s+\epsilon}\phi^{\prime}(u)^{2}du}\\ =\int_{s+\epsilon}^{t}e^{-\frac{1}{2}\int_{s}^{u}\phi^{\prime}(\omega)^{2}d\omega}\Upsilon(u)du\end{split}

As J(s,x)=eϕ(s)xlimte12st(ϕ(u))2dum^(s,x,t,ϕ(t))J(s,x)=e^{\phi^{\prime}(s)x}\lim\limits_{t\rightarrow\infty}e^{-\frac{1}{2}\int_{s}^{t}(\phi^{\prime}(u))^{2}du}\hat{m}(s,x,t,\phi^{\prime}(t)). By sending tt to \infty in the expression (4.6), we have

J(s,x)=eϕ(s)x[m^(s,x,s+ϵ,ϕ(s+ϵ))e12ss+ϵϕ(u)2du+\displaystyle J(s,x)=e^{\phi^{\prime}(s)x}\left[\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon))e^{-\frac{1}{2}\int_{s}^{s+\epsilon}\phi^{\prime}(u)^{2}du}\right.+
s+ϵe12suϕ(ω)2dωΥ(s,x,u)du]\displaystyle\left.\int_{s+\epsilon}^{\infty}e^{-\frac{1}{2}\int_{s}^{u}\phi^{\prime}(\omega)^{2}d\omega}\Upsilon(s,x,u)du\right]

It follows that

(4.7) j(s):=limx0xJ(s,x)=e12ss+ϵϕ(u)2dulimx0xm^(s,x,s+ϵ,ϕ(s+ϵ))+s+ϵe12suϕ(ω)2dωlimx0xΥ(s,x,u)du\displaystyle\begin{split}j(s):=\lim_{x\uparrow 0}\frac{\partial}{\partial x}J(s,x)=e^{-\frac{1}{2}\int_{s}^{s+\epsilon}\phi^{\prime}(u)^{2}du}\lim_{x\uparrow 0}\frac{\partial}{\partial x}\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon))+\\ \int_{s+\epsilon}^{\infty}e^{-\frac{1}{2}\int_{s}^{u}\phi^{\prime}(\omega)^{2}d\omega}\lim_{x\uparrow 0}\frac{\partial}{\partial x}\Upsilon(s,x,u)du\end{split}

since m(s,0,s+ϵ,)=0m(s,0,s+\epsilon,\cdot)=0, and we can interchange differentiation and the integral sign in the second term because we are away from the singularity line {t=s}\{t=s\}. Now, we have that

m^(s,x,s+ϵ,ϕ(s+ϵ))=0eϕ(s+ϵ)ym(s,x,s+ϵ,y)dy\displaystyle\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon))=\int_{0}^{\infty}e^{\phi^{\prime}(s+\epsilon)y}m(s,x,s+\epsilon,y)dy

It is clear that mm is smooth in the parameters (s,x)(s,x) as well. Our analysis of regularity of the function f(s,x,t,y)f(s,x,t,y) consisted of using the Kolmogorov forward equation where the parameters were tt and yy, but similarly the Kolmogorov backward equation that holds for the parameters ss and xx, we see that the solution enjoys the same smoothness and integrability properties away from the line {s=t}\{s=t\} (it is formally just the adjoint problem). Hence we can differentiate inside the integral sign to get

limx0xm^(s,x,s+ϵ,ϕ(s+ϵ))\displaystyle\lim_{x\uparrow 0}\frac{\partial}{\partial x}\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon)) =0eϕ(s+ϵ)ylimx0xm(s,x,s+ϵ,y)dy\displaystyle=\int_{0}^{\infty}e^{\phi^{\prime}(s+\epsilon)y}\lim_{x\uparrow 0}\frac{\partial}{\partial x}m(s,x,s+\epsilon,y)dy

since we have that

limx0xm(s,x,s+ϵ,y)=2y2πϵ3ey22ϵ×\displaystyle\lim_{x\uparrow 0}\frac{\partial}{\partial x}m(s,x,s+\epsilon,y)=-\frac{2y}{\sqrt{2\pi\epsilon^{3}}}e^{-\frac{y^{2}}{2\epsilon}}\times
𝔼[exp(ss+ϵϕ(u)B(u)du)|B(s)=0,B(s+ϵ)=y]\displaystyle\mathbb{E}\left[\text{exp}\left(-\int_{s}^{s+\epsilon}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=0,B(s+\epsilon)=y\right]

Thus

limx0xm^(s,x,s+ϵ,ϕ(s+ϵ))=0eϕ(s+ϵ)yy22ϵ2y2πϵ3×\displaystyle\lim_{x\uparrow 0}\frac{\partial}{\partial x}\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon))=-\int_{0}^{\infty}e^{\phi^{\prime}(s+\epsilon)y-\frac{y^{2}}{2\epsilon}}\frac{2y}{\sqrt{2\pi\epsilon^{3}}}\times
𝔼[exp(ss+ϵϕ(u)B(u)du)|B(s)=0,B(s+ϵ)=y]dy\displaystyle\mathbb{E}\left[\text{exp}\left(-\int_{s}^{s+\epsilon}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=0,B(s+\epsilon)=y\right]dy

However, the density of a three-dimensional Bessel process is given by

(4.8) [B(s+ϵ)dy|B(s)=0]=2y22πϵ3ey22ϵdy\mathbb{P}[B(s+\epsilon)\in dy|B(s)=0]=\frac{2y^{2}}{\sqrt{2\pi\epsilon^{3}}}e^{-\frac{y^{2}}{2\epsilon}}dy

Hence

limx0xm^(s,x,s+ϵ,ϕ(s+ϵ))=\displaystyle\lim_{x\uparrow 0}\frac{\partial}{\partial x}\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon))=
𝔼[1B(s+ϵ)exp(ϕ(s+ϵ)B(s+ϵ)ss+ϵϕ(u)B(u)du)|B(s)=0]\displaystyle-\mathbb{E}\left[\frac{1}{B(s+\epsilon)}\text{exp}\left(\phi^{\prime}(s+\epsilon)B(s+\epsilon)-\int_{s}^{s+\epsilon}\phi^{\prime\prime}(u)B(u)du\right)|B(s)=0\right]
=𝔼[1B(ϵ)exp(ϕ(s+ϵ)B(ϵ)0ϵϕ(u+s)B(u)du)|B(0)=0]\displaystyle=-\mathbb{E}\left[\frac{1}{B(\epsilon)}\text{exp}\left(\phi^{\prime}(s+\epsilon)B(\epsilon)-\int_{0}^{\epsilon}\phi^{\prime\prime}(u+s)B(u)du\right)|B(0)=0\right]

However by Brownian scaling, we know that

(B(u),u0)=d(ϵB(uϵ),u0)\displaystyle(B(u),u\geq 0)\overset{\mathrm{d}}{=}(\sqrt{\epsilon}B\left(\frac{u}{\epsilon}\right),u\geq 0)

Hence

limx0xm^(s,x,s+ϵ,ϕ(s+ϵ))\displaystyle\lim_{x\uparrow 0}\frac{\partial}{\partial x}\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon)) =𝔼[1ϵB(1)exp(ϕ(s+ϵ)ϵB(1)\displaystyle=-\mathbb{E}\left[\frac{1}{\sqrt{\epsilon}B(1)}\text{exp}\left(\phi^{\prime}(s+\epsilon)\sqrt{\epsilon}B(1)-\right.\right.
ϵ301ϕ(ϵu+s)B(u)du)|B(0)=0]\displaystyle\left.\left.\sqrt{\epsilon^{3}}\int_{0}^{1}\phi^{\prime\prime}(\epsilon u+s)B(u)du\right)|B(0)=0\right]
=1ϵ𝔼[1B(1)]+ϕ(s)+O(ϵ)\displaystyle=-\frac{1}{\sqrt{\epsilon}}\mathbb{E}\left[\frac{1}{B(1)}\right]+\phi^{\prime}(s)+O(\sqrt{\epsilon})

It follows then that

(4.9) limx0xm^(s,x,s+ϵ,ϕ(s+ϵ))=22πϵ+ϕ(s)+O(ϵ)\lim_{x\uparrow 0}\frac{\partial}{\partial x}\hat{m}(s,x,s+\epsilon,\phi^{\prime}(s+\epsilon))=-\frac{2}{\sqrt{2\pi\epsilon}}+\phi^{\prime}(s)+O(\sqrt{\epsilon})

for ϵ\epsilon small. The expectation of the inverse of B(1)B(1) is computed using the density given in (4.8). Now, on the other hand for the second term in (4.7), we have

limx0xΥ(s,x,u)\displaystyle\lim_{x\uparrow 0}\frac{\partial}{\partial x}\Upsilon(s,x,u) =xΦ(s,0,u)exp(12suϕ(ω)2dω)\displaystyle=-\partial_{x}\Phi(s,0,u)\text{exp}\left(\frac{1}{2}\int_{s}^{u}\phi^{\prime}(\omega)^{2}d\omega\right)
=12π(us)3𝔼[exp(suϕ(z)e(z)dz)]\displaystyle=\frac{1}{\sqrt{2\pi(u-s)^{3}}}\mathbb{E}\left[\text{exp}\left(-\int_{s}^{u}\phi^{\prime\prime}(z)\textbf{e}(z)dz\right)\right]

Hence

(4.10) s+ϵe12suϕ(ω)2dωlimx0xΥ(s,x,u)du=ϵls(u)2πu3du\int_{s+\epsilon}^{\infty}e^{-\frac{1}{2}\int_{s}^{u}\phi^{\prime}(\omega)^{2}d\omega}\lim_{x\uparrow 0}\frac{\partial}{\partial x}\Upsilon(s,x,u)du=\int_{\epsilon}^{\infty}\frac{l^{s}(u)}{\sqrt{2\pi u^{3}}}du

and thus, from combining (4.7), (4.9) and (4.10) we get

j(s)=ϵls(u)2πu3du22πϵϕ(s)+O(ϵ)\displaystyle j(s)=\int_{\epsilon}^{\infty}\frac{l^{s}(u)}{\sqrt{2\pi u^{3}}}du-\frac{2}{\sqrt{2\pi\epsilon}}-\phi^{\prime}(s)+O(\sqrt{\epsilon})

Finally, see that

ϵdu2πu3=22πϵ\displaystyle\int_{\epsilon}^{\infty}\frac{du}{\sqrt{2\pi u^{3}}}=\frac{2}{\sqrt{2\pi\epsilon}}

and then send ϵ\epsilon to zero to finish the proof. ∎

Remark 4.9.

When ϕ\phi is parabolic (ϕ(y)=y2\phi(y)=y^{2}), the term ϕ\phi^{\prime\prime} in the PDE (4.5) of mm becomes a constant and thus it takes the simple form

tm=122yym2ym\displaystyle\partial_{t}m=\frac{1}{2}\partial^{2}_{yy}m-2ym

By taking the Fourier transform in time we get

12(m^(τ,y))=(iτ+2y)m^(τ,y)\displaystyle\frac{1}{2}(\hat{m}(\tau,y))^{\prime\prime}=(i\tau+2y)\hat{m}(\tau,y)

This is a Sturm-Liouville equation. Its solution can be expressed in terms of Airy functions, from which follows all the analytical descriptions that Groeneboom found in [9]. It is clear that when ϕ\phi^{\prime\prime} is not constant, this method fails which makes the study more delicate as one doesn’t have any asymptotic or regularity properties of the function mm, which was a crucial part in the analysis of Groeneboom. For those reasons, we had to take advantage of the space Laplace transform.

As a consequence of the explicit formula of jj and Φ\Phi, we are able to provide the joint distribution of the maximum of the process (W(ω)ϕ(ω))ωs(W(\omega)-\phi(\omega))_{\omega\geq s} and its location. This is given by the expression of Φ\Phi and jj and using Lemma 3.5. However, the formula is involving many terms, in particular the Bessel bridge area. On the other hand, the density of the location of the maximum takes a simpler formula. This is a generalization of Chernoff distribution, where the parabolic drift is replaced by any strictly convex drift ϕ\phi.

Theorem 4.10.

Let ωM\omega_{M} be the location of the unique maximum of the process (S(ω)=W(ω)ϕ(ω))ω(S(\omega)=W(\omega)-\phi(\omega))_{\omega\in\mathbb{R}}, its density is equal to

[ωMdt]dt=12j(t)j~(t)\displaystyle\frac{\mathbb{P}[\omega_{M}\in dt]}{dt}=\frac{1}{2}j(t)\tilde{j}(-t)

where j~\tilde{j} is the analogue of jj for the process S~(ω):=W(ω)ϕ(ω)\tilde{S}(\omega):=W(\omega)-\phi(-\omega).

Proof.

We will prove the equality for t0t\geq 0, the case t0t\leq 0 is completely identical. From Lemma 3.5 with s=0s=0 and any x>zx>z

[argmaxω0S(ω)dt,maxω0S(ω)dz|S(0)=x]dtdz=12j(t)yf(0,xz,t,0)\displaystyle\frac{\mathbb{P}[\text{argmax}_{\omega\geq 0}S(\omega)\in dt,\max\limits_{\omega\geq 0}S(\omega)\in dz|S(0)=x]}{dtdz}=\frac{1}{2}j(t)\partial_{y}f(0,x-z,t,0)

Hence

[ωMdt|S(0)=x]=x+[argmaxω0S(ω)dt,maxω0S(ω)dz,\displaystyle\mathbb{P}[\omega_{M}\in dt|S(0)=x]=\int_{x}^{+\infty}\mathbb{P}[\text{argmax}_{\omega\geq 0}S(\omega)\in dt,\max\limits_{\omega\geq 0}S(\omega)\in dz,
maxω0S(ω)<z|S(0)=x]\displaystyle\max\limits_{\omega\leq 0}S(\omega)<z|S(0)=x]
=x+12j(t)yf(0,xz,t,0)[S(ω)<z for all ω0|S(0)=x]dzdt\displaystyle=\int_{x}^{+\infty}\frac{1}{2}j(t)\partial_{y}f(0,x-z,t,0)\mathbb{P}[S(\omega)<z\text{ for all }\omega\leq 0|S(0)=x]dzdt

by independence of the paths (S(ω),ω0)(S(\omega),\omega\leq 0) and (S(ω),ω0)(S(\omega),\omega\geq 0). However by time reversal of the Brownian motion we have

[S(ω)<z for all ω0|S(0)=x]\displaystyle\mathbb{P}[S(\omega)<z\text{ for all }\omega\leq 0|S(0)=x] =[S~(ω)<z for all ω0|S~(0)=x]\displaystyle=\mathbb{P}[\tilde{S}(\omega)<z\text{ for all }\omega\geq 0|\tilde{S}(0)=x]
=[S~(ω)<0 for all ω0|S~(0)=xz]\displaystyle=\mathbb{P}[\tilde{S}(\omega)<0\text{ for all }\omega\geq 0|\tilde{S}(0)=x-z]
=J~(0,xz)\displaystyle=\tilde{J}(0,x-z)

Thus

[ωMdt|S(0)=x]dt=012j(t)yf(0,z,t,0)J~(0,z)dz\displaystyle\frac{\mathbb{P}[\omega_{M}\in dt|S(0)=x]}{dt}=\int_{-\infty}^{0}\frac{1}{2}j(t)\partial_{y}f(0,z,t,0)\tilde{J}(0,z)dz

Notice that the right hand-side is independent of xx, so we can drop the conditional probability in the left hand-side. Moreover by (3.7), we have

(4.11) yf(0,z,t,0)=xf~(t,0,0,z)\partial_{y}f(0,z,t,0)=\partial_{x}\tilde{f}(-t,0,0,z)

Using the expression of the entrance law of the process S~\tilde{S}^{\downarrow} in (3.3), we have

(4.12) [S~(0)dz|S~(t)=0]=J~(0,z)j~(t)xf~(t,0,0,z)dz\mathbb{P}[\tilde{S}^{\downarrow}(0)\in dz|\tilde{S}^{\downarrow}(-t)=0]=\frac{\tilde{J}(0,z)}{\tilde{j}(-t)}\partial_{x}\tilde{f}(-t,0,0,z)dz

Hence combining (4.11) and (4.12) we get

0yf(0,z,t,0)J~(0,z)dz=j~(t)0[S~(0)dz|S~(t)=0]=j~(t)\displaystyle\int_{0}^{\infty}\partial_{y}f(0,z,t,0)\tilde{J}(0,z)dz=\tilde{j}(-t)\int_{-\infty}^{0}\mathbb{P}[\tilde{S}^{\downarrow}(0)\in dz|\tilde{S}^{\downarrow}(-t)=0]=\tilde{j}(-t)

which completes the proof.

Remark 4.11.

This last theorem is exactly Theorem 1.11 by noticing that fϕ(t)=j(t)f^{\phi}(t)=-j(t) and fϕ()(t)=j~(t)f^{\phi(-\cdot)}(-t)=-\tilde{j}(-t).

Remark 4.12.

From [9] results in the parabolic drift case, the Chernoff distribution can be expressed as

[argmaxz(W(z)z2)dt]dt=12k(t)k(t)\displaystyle\frac{\mathbb{P}[\text{argmax}_{z\in\mathbb{R}}(W(z)-z^{2})\in dt]}{dt}=\frac{1}{2}k(t)k(-t)

where k(t)=e23t3g(t)k(t)=e^{\frac{2}{3}t^{3}}g(t) and gg has the Fourier transform given by

g^(τ):=eitτg(t)dt=213Ai(i213τ)\displaystyle\hat{g}(\tau):=\int_{-\infty}^{\infty}e^{it\tau}g(t)dt=\frac{2^{\frac{1}{3}}}{\mathrm{Ai}(i2^{-\frac{1}{3}}\tau)}

This expression is not clear from the formula we provided in Theorem 1.11. We will prove thus in the following proposition that those two indeed coincide.

Proposition 4.13.

For any tt\in\mathbb{R} we have

2t+012πu3(1e23((u+t)3t3)𝔼[exp(20ue(z)dz)])du=\displaystyle 2t+\int_{0}^{\infty}\frac{1}{\sqrt{2\pi u^{3}}}\left(1-e^{-\frac{2}{3}((u+t)^{3}-t^{3})}\mathbb{E}\left[\text{exp}\left(-2\int_{0}^{u}\textbf{e}(z)dz\right)\right]\right)du=
e23t32πeitvg^(v)dv\displaystyle\frac{e^{\frac{2}{3}t^{3}}}{2\pi}\int_{-\infty}^{\infty}e^{-itv}\hat{g}(v)dv
Proof.

From equation (1.6) in [10] 333There is a typo in the published paper, the term 4234^{\frac{2}{3}} in the denominator should be there instead of 4134^{\frac{1}{3}}. , we have that

(4.13) 12πv=Ai(iξ413x)Ai(iξ)u=0eiuv23((u+t)3t3)dudv=e2txe23t3423v=eitvAi(iξ)Bi(iξ413x)Ai(iξ413x)Bi(iξ)Ai(iξ)dv\displaystyle\begin{split}\frac{1}{2\pi}\int_{v=-\infty}^{\infty}\frac{\mathrm{Ai}(i\xi-4^{\frac{1}{3}}x)}{\mathrm{Ai}(i\xi)}\int_{u=0}^{\infty}e^{iuv-\frac{2}{3}((u+t)^{3}-t^{3})}dudv=e^{-2tx}\\ -\frac{e^{\frac{2}{3}t^{3}}}{4^{\frac{2}{3}}}\int_{v=-\infty}^{\infty}e^{-itv}\frac{\mathrm{Ai}(i\xi)\mathrm{Bi}(i\xi-4^{\frac{1}{3}}x)-\mathrm{Ai}(i\xi-4^{\frac{1}{3}}x)\mathrm{Bi}(i\xi)}{\mathrm{Ai}(i\xi)}dv\end{split}

where ξ=213v\xi=2^{-\frac{1}{3}}v, and Bi\mathrm{Bi} is the second Airy function. By differentiating both sides with respect to xx and sending xx to zero, we get

(4.14) e23t3413πv=eitvAi(iξ)dv=2t+limx0x12πv=Ai(iξ413x)Ai(iξ)u=0eiuv23((u+t)3t3)dudv\frac{e^{\frac{2}{3}t^{3}}}{4^{\frac{1}{3}}\pi}\int_{v=-\infty}^{\infty}\frac{e^{-itv}}{\mathrm{Ai}(i\xi)}dv=\\ 2t+\lim_{x\uparrow 0}\frac{\partial}{\partial x}\frac{1}{2\pi}\int_{v=-\infty}^{\infty}\frac{\mathrm{Ai}(i\xi-4^{\frac{1}{3}}x)}{\mathrm{Ai}(i\xi)}\int_{u=0}^{\infty}e^{iuv-\frac{2}{3}((u+t)^{3}-t^{3})}dudv

as the Wronskian of the Airy functions Ai\mathrm{Ai} and Bi\mathrm{Bi} is constant and equal to 1π\frac{1}{\pi}. In the right-hand side of (4.13), we cannot differentiate inside the integral sign because it becomes divergent. However for fixed x<0x<0, the integrand is absolutely integrable and thus we can use Fubini theorem. Now from [11][Equation 384, Page 141] we have that

0eλs𝔼[exp(20sB(u)du)|B(s)=x]x2πs3ex22sds=Ai(213λ413x)Ai(213λ)\displaystyle-\int_{0}^{\infty}e^{-\lambda s}\mathbb{E}\left[\text{exp}\left(-2\int_{0}^{s}B(u)du\right)|B(s)=-x\right]\frac{x}{\sqrt{2\pi s^{3}}}e^{-\frac{x^{2}}{2s}}ds=\frac{\mathrm{Ai}(2^{-\frac{1}{3}}\lambda-4^{\frac{1}{3}}x)}{\mathrm{Ai}(2^{-\frac{1}{3}}\lambda)}

where BB is as usual a three-dimensional Bessel process. Thus, by inverse Laplace transform we have

𝔼[exp(20uB(z)dz)|B(u)=x]x2πu3ex22u=12πeiuvAi(iξ413x)Ai(iξ)dv\displaystyle-\mathbb{E}\left[\text{exp}\left(-2\int_{0}^{u}B(z)dz\right)|B(u)=-x\right]\frac{x}{\sqrt{2\pi u^{3}}}e^{-\frac{x^{2}}{2u}}=\frac{1}{2\pi}\int_{-\infty}^{\infty}e^{iuv}\frac{\mathrm{Ai}(i\xi-4^{\frac{1}{3}}x)}{\mathrm{Ai}(i\xi)}dv

Hence the integral in the RHS of (4.14) is equal to

(4.15) 0e23((u+t)3t3)x2πu3ex22u𝔼[exp(20uB(z)dz)|B(u)=x]du-\int_{0}^{\infty}e^{-\frac{2}{3}((u+t)^{3}-t^{3})}\frac{x}{\sqrt{2\pi u^{3}}}e^{-\frac{x^{2}}{2u}}\mathbb{E}\left[\text{exp}\left(-2\int_{0}^{u}B(z)dz\right)|B(u)=-x\right]du

By splitting this integral on (0,ϵ)(0,\epsilon) and (ϵ,)(\epsilon,\infty), we can interchange the integral and the differentiation for the integral on (ϵ,)(\epsilon,\infty), and so we get after sending xx to zero

(4.16) ϵe23((u+t)3t3)12πu3𝔼[exp(20ue(z)dz)]du-\int_{\epsilon}^{\infty}e^{-\frac{2}{3}((u+t)^{3}-t^{3})}\frac{1}{\sqrt{2\pi u^{3}}}\mathbb{E}\left[\text{exp}\left(-2\int_{0}^{u}\textbf{e}(z)dz\right)\right]du

where e is as usual a Brownian excursion on the corresponding interval. As for the first term (the integral on (0,ϵ)(0,\epsilon)), by the change of variable y=xuy=\frac{x}{\sqrt{u}} (dy=x2u3dudy=-\frac{x}{2\sqrt{u^{3}}}du), it is equal to

0ϵe23((u+t)3t3)x2πu3ex22u𝔼[exp(20uB(z)dz)|B(u)=x]du\displaystyle-\int_{0}^{\epsilon}e^{-\frac{2}{3}((u+t)^{3}-t^{3})}\frac{x}{\sqrt{2\pi u^{3}}}e^{-\frac{x^{2}}{2u}}\mathbb{E}\left[\text{exp}\left(-2\int_{0}^{u}B(z)dz\right)|B(u)=-x\right]du
=xϵe23((x2y2+t)3t3)22πey22𝔼[exp(2x3y301B(z)dz)|B(1)=y]dy\displaystyle=\int_{-\infty}^{\frac{x}{\sqrt{\epsilon}}}e^{-\frac{2}{3}((\frac{x^{2}}{y^{2}}+t)^{3}-t^{3})}\frac{2}{\sqrt{2\pi}}e^{-\frac{y^{2}}{2}}\mathbb{E}\left[\text{exp}\left(-2\frac{x^{3}}{y^{3}}\int_{0}^{1}B(z)dz\right)|B(1)=-y\right]dy

by Brownian scaling on the Bessel process BB. Differentiating with respect to xx, we get by Leibniz rule

(4.17) 22πϵe23((ϵ+t)3t3)ex22ϵ𝔼[exp(2ϵ301B(z)dz)|B(1)=xϵ]+Fϵ(x)\frac{2}{\sqrt{2\pi\epsilon}}e^{-\frac{2}{3}((\epsilon+t)^{3}-t^{3})}e^{-\frac{x^{2}}{2\epsilon}}\mathbb{E}\left[\text{exp}\left(-2\sqrt{\epsilon^{3}}\int_{0}^{1}B(z)dz\right)|B(1)=-\frac{x}{\sqrt{\epsilon}}\right]+F^{\epsilon}(x)

where FϵF^{\epsilon} is equal to

Fϵ(x)=22πxϵ(4x5y68tx3y44t2xy26x2y3)ey22e23((x2y2+t)3t3)×\displaystyle F^{\epsilon}(x)=\frac{2}{\sqrt{2\pi}}\int_{-\infty}^{\frac{x}{\sqrt{\epsilon}}}\left(-4\frac{x^{5}}{y^{6}}-8t\frac{x^{3}}{y^{4}}-4t^{2}\frac{x}{y^{2}}-6\frac{x^{2}}{y^{3}}\right)e^{-\frac{y^{2}}{2}}e^{-\frac{2}{3}((\frac{x^{2}}{y^{2}}+t)^{3}-t^{3})}\times
𝔼[exp(2x3y301B(z)dz)|B(1)=y]dy\displaystyle\mathbb{E}\left[\text{exp}\left(-2\frac{x^{3}}{y^{3}}\int_{0}^{1}B(z)dz\right)|B(1)=-y\right]dy

However we have that for xx small enough (such that |xϵ|=xϵ1|\frac{x}{\sqrt{\epsilon}}|=-\frac{x}{\sqrt{\epsilon}}\leq 1)

|xϵxy2ey22e23((x2y2+t)3t3)𝔼[exp(2x3y301B(z)dz)|B(1)=y]dy|\displaystyle\left|\int_{-\infty}^{\frac{x}{\sqrt{\epsilon}}}\frac{x}{y^{2}}e^{-\frac{y^{2}}{2}}e^{-\frac{2}{3}((\frac{x^{2}}{y^{2}}+t)^{3}-t^{3})}\mathbb{E}\left[\text{exp}\left(-2\frac{x^{3}}{y^{3}}\int_{0}^{1}B(z)dz\right)|B(1)=-y\right]dy\right|\leq
|x|xϵey22y2dy|x|(1ϵx+1ey22dy)\displaystyle|x|\int_{-\frac{x}{\sqrt{\epsilon}}}^{\infty}\frac{e^{-\frac{y^{2}}{2}}}{y^{2}}dy\leq|x|(1-\frac{\sqrt{\epsilon}}{x}+\int_{1}^{\infty}e^{-\frac{y^{2}}{2}}dy)

so

lim supx0|xϵxy2ey22e23((x2y2+t)3t3)𝔼[exp(2x3y301B(z)dz)|B(1)=y]dy|ϵ\displaystyle\limsup\limits_{x\uparrow 0}\left|\int_{-\infty}^{\frac{x}{\sqrt{\epsilon}}}\frac{x}{y^{2}}e^{-\frac{y^{2}}{2}}e^{-\frac{2}{3}((\frac{x^{2}}{y^{2}}+t)^{3}-t^{3})}\mathbb{E}\left[\text{exp}\left(-2\frac{x^{3}}{y^{3}}\int_{0}^{1}B(z)dz\right)|B(1)=-y\right]dy\right|\leq\sqrt{\epsilon}

Similarly with the other terms we find that there is a constant C>0C>0 (that depends on tt) such that

lim supx0|Fϵ(x)|Cϵ\displaystyle\limsup\limits_{x\uparrow 0}|F^{\epsilon}(x)|\leq C\sqrt{\epsilon}

Hence, by combining (4.16) and (4.17), the limit of the derivative of the expression in (4.15) when xx goes to zero is equal to

ϵe23((u+t)3t3)12πu3𝔼[exp(20ue(z)dz)]du+\displaystyle-\int_{\epsilon}^{\infty}e^{-\frac{2}{3}((u+t)^{3}-t^{3})}\frac{1}{\sqrt{2\pi u^{3}}}\mathbb{E}\left[\text{exp}\left(-2\int_{0}^{u}\textbf{e}(z)dz\right)\right]du+
22πϵe23((ϵ+t)3t3)𝔼[exp(2ϵ301e(z)dz)]+lim supx0Fϵ(x)\displaystyle\frac{2}{\sqrt{2\pi\epsilon}}e^{-\frac{2}{3}((\epsilon+t)^{3}-t^{3})}\mathbb{E}\left[\text{exp}\left(-2\sqrt{\epsilon^{3}}\int_{0}^{1}\textbf{e}(z)dz\right)\right]+\limsup\limits_{x\uparrow 0}F^{\epsilon}(x)

Now it suffices to see that

22πϵe23((ϵ+t)3t3)𝔼[exp(2ϵ301e(z)dz)]=22πϵ+O(ϵ)\displaystyle\frac{2}{\sqrt{2\pi\epsilon}}e^{-\frac{2}{3}((\epsilon+t)^{3}-t^{3})}\mathbb{E}\left[\text{exp}\left(-2\sqrt{\epsilon^{3}}\int_{0}^{1}\textbf{e}(z)dz\right)\right]=\frac{2}{\sqrt{2\pi\epsilon}}+O(\sqrt{\epsilon})
=ϵ12πu3du+O(ϵ)\displaystyle=\int_{\epsilon}^{\infty}\frac{1}{\sqrt{2\pi u^{3}}}du+O(\sqrt{\epsilon})

By sending ϵ\epsilon to zero we get the desired result. ∎

We are now ready to prove the Theorem 1.9.

Proof of Theorem 1.9.

Recall that our solution is expressed as

ρ(x,t)=L(y(x,t)xt)=L(ΨtL(t)(x)xt)\displaystyle\rho(x,t)=L^{\prime}\left(\frac{y(x,t)-x}{t}\right)=L^{\prime}\left(\frac{\Psi^{tL(\frac{\cdot}{t})}(x)-x}{t}\right)

Hence, ρ\rho is stationary by Theorem 2.2, and so it is a time-homogenous Markov process, its generator is determined by

𝒜tφ(y)\displaystyle\mathcal{A}^{t}\varphi(y) =limh0𝔼[φ(ρ(h,t))φ(ρ)|ρ(0,t)=ρ]h\displaystyle=\lim_{h\rightarrow 0}\frac{\mathbb{E}[\varphi(\rho(h,t))-\varphi(\rho_{-})|\rho(0,t)=\rho_{-}]}{h}
=limh0𝔼[φ(L(ΨtL(.t)(h)ht))φ(ρ)|ΨtL(t)(0)=tH(ρ)]h\displaystyle=\lim_{h\rightarrow 0}\frac{\mathbb{E}[\varphi(L^{\prime}(\frac{\Psi^{tL(\frac{.}{t})}(h)-h}{t}))-\varphi(\rho_{-})|\Psi^{tL(\frac{\cdot}{t})}(0)=tH^{\prime}(\rho_{-})]}{h}
=1tL(H(ρ))φ(ρ)+𝒜tL(t)φ(L(t))(tH(ρ))\displaystyle=-\frac{1}{t}L^{\prime\prime}(H^{\prime}(\rho_{-}))\varphi^{\prime}(\rho_{-})+\mathcal{A}^{tL(\frac{\cdot}{t})}\varphi(L^{\prime}(\frac{\cdot}{t}))(tH^{\prime}(\rho_{-}))
=φ(ρ))tH(ρ)+𝒜tL(.t)φ(L(t))(tH(ρ))\displaystyle=-\frac{\varphi^{\prime}(\rho_{-}))}{tH^{\prime\prime}(\rho_{-})}+\mathcal{A}^{tL(\frac{.}{t})}\varphi(L^{\prime}(\frac{\cdot}{t}))(tH^{\prime}(\rho_{-}))
=φ(ρ)tH(ρ)+ρ(φ(ρ+)φ(ρ))n(ρ,ρ+,t)dρ+\displaystyle=-\frac{\varphi^{\prime}(\rho_{-})}{tH^{\prime\prime}(\rho_{-})}+\int_{\rho_{-}}^{\infty}(\varphi(\rho_{+})-\varphi(\rho_{-}))n(\rho_{-},\rho_{+},t)d\rho_{+}

where

n(ρ,ρ+,t)=tH(ρ+)jtL(t)(tH(ρ+))jtL(t)(tH(ρ))KtL(t)(tH(ρ),tH(ρ+))\displaystyle n(\rho_{-},\rho_{+},t)=tH^{\prime\prime}(\rho_{+})\frac{j^{tL(\frac{\cdot}{t})}(tH^{\prime}(\rho_{+}))}{j^{tL(\frac{\cdot}{t})}(tH^{\prime}(\rho_{-}))}K^{tL(\frac{\cdot}{t})}(tH^{\prime}(\rho_{-}),tH^{\prime}(\rho_{+}))

By a change of variables we have

KtL(t)(tH(ρ),tH(ρ+))=ρ+ρ2πt3(H(ρ+)H(ρ))3×\displaystyle K^{tL(\frac{\cdot}{t})}(tH^{\prime}(\rho_{-}),tH^{\prime}(\rho_{+}))=\frac{\rho_{+}-\rho_{-}}{\sqrt{2\pi t^{3}(H^{\prime}(\rho_{+})-H^{\prime}(\rho_{-}))^{3}}}\times
exp(t2ρρ+(ρ)2H(ρ)dρ)𝔼[exp(ρρ+e(tH(ρ))dρ)]\displaystyle\text{exp}\left(-\frac{t}{2}\int_{\rho_{-}}^{\rho_{+}}(\rho_{*})^{2}H^{\prime\prime}(\rho_{*})d\rho_{*}\right)\mathbb{E}\left[\text{exp}\left(-\int_{\rho_{-}}^{\rho_{+}}\textbf{e}(tH^{\prime}(\rho_{*}))d\rho_{*}\right)\right]

Similarly

jtL(t)(tH(ρ))=ρ+ρ1p(ρ,ρ,t)2πt(H(ρ)H(ρ))3H(ρ)dρ\displaystyle-j^{tL(\frac{\cdot}{t})}(tH^{\prime}(\rho_{-}))=\rho_{-}+\int_{\rho_{-}}^{\infty}\frac{1-p(\rho_{-},\rho,t)}{\sqrt{2\pi t(H^{\prime}(\rho)-H^{\prime}(\rho_{-}))^{3}}}H^{\prime\prime}(\rho)d\rho

where

p(ρ,ρ,t)=exp(t2ρρ(ρ)2H(ρ)dρ)𝔼[exp(ρρe(tH(ρ))dρ)]\displaystyle p(\rho_{-},\rho,t)=\text{exp}\left(-\frac{t}{2}\int_{\rho_{-}}^{\rho}(\rho_{*})^{2}H^{\prime\prime}(\rho_{*})d\rho_{*}\right)\mathbb{E}\left[\text{exp}\left(-\int_{\rho_{-}}^{\rho}\textbf{e}(tH^{\prime}(\rho_{*}))d\rho_{*}\right)\right]

The theorem then follows by appropriately defining the kernel KK. ∎

Remark 4.14.

While our main study focused on the case where the initial potential is a two-sided Brownian motion. It is not hard to see that we can extend the result about the profile of the entropy solution when the potential is a spectrally positive Lévy process with non-zero Brownian exponent. The main ingredients that were used were respectively the path decomposition of Markov processes at their ultimate maximum and the regularity properties of the transition function ff. Both these facts hold true in the Lévy case when the initial potential U0U_{0} has a non-zero Brownian exponent, as the only difference in the Kolmogorov forward equation is an added integral operator accounting for the jumps of the Lévy process. A similar approach will lead to the same smoothness property away from the singularity line {t=s}\{t=s\} (the presence of the heat operator t122y\partial_{t}-\frac{1}{2}\partial^{2}_{y} is key to have parabolic smoothing), which will allow all the operations in the second section to be valid. Moreover, one should be able to extract similar expression for the jump kernel nn by using the Girsanov theorem version for Lévy processes. We chose in this paper to only discuss the Brownian motion case because it gives a general idea on how things work and also because it simplifies greatly the computations. One would expect to have similar formulas where the equivalent of the Brownian excursion will be the Lévy bridge informally defined as a Lévy process conditioned to stay positive and to start and end at zero. Those bridges are discussed in [19].

5. Structure of shocks of the entropy solution

A priori, from the involved expression of the generator in Theorem 1.9, one cannot easily claim whether if the structure of shocks of the solution ρ\rho is discrete or not. Indeed, this amounts to checking if the following integrability condition on the jump kernel nn holds

λ(ρ)=ρn(ρ,ρ+,t)dρ+< for all ρ\displaystyle\lambda(\rho_{-})=\int_{\rho_{-}}^{\infty}n(\rho_{-},\rho_{+},t)d\rho_{+}<\infty\text{ for all }\rho_{-}\in\mathbb{R}

However, using the recent theory of Lipschitz minorants of Lévy processes developed in [2] and [7], and following some of the arguments from the study of the structure of shocks in Burgers equation of [1], it turns out that when the initial potential is an abrupt spectrally positive Lévy process, one can prove that the set of jump times of the solution ρ\rho is discrete.

As we did with Theorem 1.9, we will prove a general statement for the process Ψϕ\Psi^{\phi} from which Theorem 1.15 will follow. We state thus the following theorem

Theorem 5.1.

Assume that U0U_{0} is an abrupt spectrally positive Lévy process and ϕ\phi is a strictly convex function such that lim|y||ϕ(y)|=+\lim_{|y|\to\infty}|\phi^{\prime}(y)|=+\infty and lim|y|+U0(y)ϕ(y)=0\displaystyle\lim_{|y|\rightarrow+\infty}\frac{U_{0}(y)}{\phi(y)}=0 almost surely, then the range of Ψϕ\Psi^{\phi} is a.s discrete.

Proof.

From Theorem 2.2, we know that for every nn\in\mathbb{Z}

(Ψϕ(x+n)n)x=d(Ψϕ(x))x\displaystyle(\Psi^{\phi}(x+n)-n)_{x\in\mathbb{R}}\overset{\mathrm{d}}{=}(\Psi^{\phi}(x))_{x\in\mathbb{R}}

hence it suffices to prove that the set range(Ψϕ)[0,1]\text{range}(\Psi^{\phi})\cap[0,1] is a.s discrete. Moreover, we can restrict the process Ψϕ\Psi^{\phi} on [M,M][-M,M]. Indeed, we claim that the probability of the event

AM:={there exists a such that |a|M and Ψϕ(a)[0,1]}\displaystyle A_{M}:=\{\text{there exists }a\text{ such that }|a|\geq M\text{ and }\Psi^{\phi}(a)\in[0,1]\}

goes to zero as MM goes to infinity. To show this claim, assume that there exists a sequence (an)n(a_{n})_{n\in\mathbb{N}} such that λn:=Ψϕ(an)[0,1]\lambda_{n}:=\Psi^{\phi}(a_{n})\in[0,1] and |an||a_{n}|\to\infty. By definition we have that

(5.1) U0(λn)ϕ(λnan)U0(y)ϕ(yan) for all yU_{0}(\lambda_{n})-\phi(\lambda_{n}-a_{n})\geq U_{0}(y)-\phi(y-a_{n})\text{ for all }y

Up to taking subsequences, we have either that ana_{n}\to\infty or ana_{n}\to-\infty. If ana_{n}\to\infty, take y=an1y=a_{n}-1 in (5.1), then

(5.2) U0(λn)ϕ(λnan)U0(an1)ϕ(1)U_{0}(\lambda_{n})-\phi(\lambda_{n}-a_{n})\geq U_{0}(a_{n}-1)-\phi(-1)

As ϕ\phi^{\prime} is strictly increasing, we must have limyϕ(y)=\lim_{y\rightarrow-\infty}\phi^{\prime}(y)=-\infty, and thus ϕ\phi is decreasing for yy\rightarrow-\infty. Hence from (5.2) and the fact that λn1\lambda_{n}\leq 1, we get

(5.3) U0(λn)U0(an1)ϕ(λnan)ϕ(1)ϕ(1an)ϕ(1)U_{0}(\lambda_{n})-U_{0}(a_{n}-1)\geq\phi(\lambda_{n}-a_{n})-\phi(-1)\geq\phi(1-a_{n})-\phi(-1)

for nn large enough. However, because (U0(y))y(U_{0}(y))_{y\in\mathbb{R}} has the same distribution as (U0((y)))y(-U_{0}((-y)-))_{y\in\mathbb{R}}, then almost surely limnU0(an1)ϕ(1an)=0\lim_{n\rightarrow\infty}\frac{U_{0}(a_{n}-1)}{\phi(1-a_{n})}=0, which is a contradiction with (5.3). The case ana_{n}\rightarrow-\infty is similar by taking y=any=a_{n} in (5.1), proving thus our claim.

Define now the event BMB_{M} as

BM={Card(range(Ψϕ|[M,M]|)[0,1])=}\displaystyle B_{M}=\left\{\text{Card}\left(\text{range}(\Psi^{\phi}_{|[-M,M]|})\cap[0,1]\right)=\infty\right\}

It suffices to prove that limM[BM]=0\lim\limits_{M\rightarrow\infty}\mathbb{P}\left[B_{M}\right]=0.

Suppose initially that 𝔼[|U0(1)|]<\mathbb{E}[|U_{0}(1)|]<\infty and let CM:=supt[2M,2M]|ϕ(t)|C_{M}:=\sup\limits_{t\in[-2M,2M]}|\phi^{\prime}(t)|. Because of our assumption on ϕ\phi, then for MM large enough we have that 𝔼[|U0(1)|]<CM\mathbb{E}[|U_{0}(1)|]<C_{M}. For any a[M,M]a\in[-M,M] such that λa:=Ψϕ(a)[0,1]\lambda_{a}:=\Psi^{\phi}(a)\in[0,1], we have for all t[M,M]t\in[-M,M]

(5.4) U0(t)U0(λa)ϕ(ta)ϕ(λaa)CM|tλa|U_{0}(t)-U_{0}(\lambda_{a})\leq\phi(t-a)-\phi(\lambda_{a}-a)\leq C_{M}|t-\lambda_{a}|

For α>0\alpha>0 such that 𝔼[|U0(1)|]<α\mathbb{E}[|U_{0}(1)|]<\alpha, let us consider now the process Lα0L^{\alpha}_{0} that is the α\alpha-Lipschitz majorant of U0U_{0}, defined formally as

Lα0(y)=supz{U0(z)α|zy|}\displaystyle L^{\alpha}_{0}(y)=\sup_{z\in\mathbb{R}}\left\{U_{0}(z)-\alpha|z-y|\right\}

We refer the reader to the two papers [2] and [7] for a detailed study of the Lipschitz minorant of a Lévy process. Consider GαtG^{\alpha}_{t} (resp. DαtD^{\alpha}_{t}) to be the last contact point before tt (resp. the first contact point after tt) of Lα0L^{\alpha}_{0} with U0U_{0}, i.e.

Gαt=sup{y<t:L0α(y)=U0(y)} and Dαt=inf{y>t:L0α(y)=U0(y)}\displaystyle G^{\alpha}_{t}=\sup\left\{y<t:L_{0}^{\alpha}(y)=U_{0}(y)\right\}\text{ and }D^{\alpha}_{t}=\inf\left\{y>t:L_{0}^{\alpha}(y)=U_{0}(y)\right\}

for any tt\in\mathbb{R}. Moreover, let 𝒵α\mathcal{Z}_{\alpha} be the contact set of Lα0L^{\alpha}_{0} and U0U_{0} defined as

𝒵α:={y:Lα0(y)=U0(y)}\displaystyle\mathcal{Z}_{\alpha}:=\{y\in\mathbb{R}:L^{\alpha}_{0}(y)=U_{0}(y)\}

Then on the event {GCM0,DCM1[M,M]}\{G^{C_{M}}_{0},D^{C_{M}}_{1}\in[-M,M]\}, from the inequality (5.4), we have

U0(GCM0)U0(λa)CM(λaGCM0) and U0(DCM1)U0(λa)CM(DCM1λa)\displaystyle U_{0}(G^{C_{M}}_{0})-U_{0}(\lambda_{a})\leq C_{M}(\lambda_{a}-G^{C_{M}}_{0})\text{ and }U_{0}(D^{C_{M}}_{1})-U_{0}(\lambda_{a})\leq C_{M}(D^{C_{M}}_{1}-\lambda_{a})

Hence for tMt\geq M, we have

U0(t)U0(λa)\displaystyle U_{0}(t)-U_{0}(\lambda_{a}) U0(DCM1)+CM(tDCM1)U0(λa)\displaystyle\leq U_{0}(D^{C_{M}}_{1})+C_{M}(t-D^{C_{M}}_{1})-U_{0}(\lambda_{a})
CM(DCM1λa)+CM(tDCM1)=CM|tλa|\displaystyle\leq C_{M}(D^{C_{M}}_{1}-\lambda_{a})+C_{M}(t-D^{C_{M}}_{1})=C_{M}|t-\lambda_{a}|

Similarly for tMt\leq-M we get the same result. Together with (5.4), we deduce that for any a[M,M]a\in[-M,M] such that λa:=Ψϕ(a)[0,1]\lambda_{a}:=\Psi^{\phi}(a)\in[0,1], λa\lambda_{a} is in the contact set 𝒵CM\mathcal{Z}_{C_{M}}. However when U0U_{0} is abrupt, we know from [2][See proof of Proposition 6.1] that this set is discrete, and hence 𝒵CM[0,1]\mathcal{Z}_{C_{M}}\cap[0,1] is finite. Thus

(5.5) [BM][GCM0M]+[DCM1M]\mathbb{P}[B_{M}]\leq\mathbb{P}[G^{C_{M}}_{0}\leq-M]+\mathbb{P}[D^{C_{M}}_{1}\geq M]

Now it is not hard to see that for α<α\alpha<\alpha^{\prime}, we have that 𝒵α𝒵α\mathcal{Z}_{\alpha}\subset\mathcal{Z}_{\alpha^{\prime}}. Hence, for MM large enough we have

(5.6) D1CMD1β,GCM0Gβ0D_{1}^{C_{M}}\leq D_{1}^{\beta},~{}~{}G^{C_{M}}_{0}\geq G^{\beta}_{0}

where β=𝔼[|U0(1)|]+1\beta=\mathbb{E}[|U_{0}(1)|]+1 is independent of MM. However, from [2][Theorem 2.6] we know that the set 𝒵β\mathcal{Z}_{\beta} is stationary and regenerative (see [8] for the precise definition of stationary regenerative sets), thus the random variables Dβ11D^{\beta}_{1}-1 and Gβ0-G^{\beta}_{0} have the same distribution as Dβ0D^{\beta}_{0}. Moreover from [2][Equation (4.7)], we have that

[Dβ0Gβ0dx]=xΛβ(dx)+xΛβ(dx)\displaystyle\mathbb{P}[D^{\beta}_{0}-G^{\beta}_{0}\in dx]=\frac{x\Lambda^{\beta}(dx)}{\int_{\mathbb{R}+}x\Lambda^{\beta}(dx)}

where Λβ\Lambda^{\beta} is the Lévy measure of the subordinator associated with the contact set 𝒵β\mathcal{Z}_{\beta} (the stationarity of 𝒵β\mathcal{Z}_{\beta} ensuring that +xΛβ(dx)<\int_{\mathbb{R}+}x\Lambda^{\beta}(dx)<\infty). It follows thus from (5.6) that the right-hand side of (5.5) goes to zero when MM\rightarrow\infty, from which we get the desired result that the range of Ψϕ\Psi^{\phi} is discrete when 𝔼[|U0(1)|]<\mathbb{E}[|U_{0}(1)|]<\infty.

Now, if 𝔼[|U0(1)|]=\mathbb{E}[|U_{0}(1)|]=\infty, consider for any NN\in\mathbb{N} the truncated process U0NU_{0}^{N}, that is the process U0U_{0} started at zero and with its jumps of size greater than NN removed. It is formally defined as :

(5.7) U0N(y)={U0(y)0zy(U0(z)U0(z))1{U0(z)U0(z)N} if y0U0(y)+yz0(U0(z)U0(z))1{U0(z)U0(z)N} if y0U_{0}^{N}(y)=\left\{\begin{array}[]{ll}U_{0}(y)-\sum_{0\leq z\leq y}(U_{0}(z)-U_{0}(z-))\mathbbold{1}_{\{U_{0}(z)-U_{0}(z-)\geq N\}}\text{ if }~{}~{}y\geq 0\\ U_{0}(y)+\sum_{y\leq z\leq 0}(U_{0}(z)-U_{0}(z-))\mathbbold{1}_{\{U_{0}(z)-U_{0}(z-)\geq N\}}\text{ if }~{}~{}y\leq 0\end{array}\right.

We have that 𝔼[|U0N(1)|]<\mathbb{E}[|U_{0}^{N}(1)|]<\infty as any Lévy process with uniformly bounded jumps has finite moments of any order (see [18][Lemma 8.2]). Hence, if we denote by ΨϕN\Psi^{\phi}_{N} the process Ψϕ\Psi^{\phi} where we replace U0U_{0} by U0NU_{0}^{N}. By what we proved previously, we have that almost surely, the set range(ΨϕN)[0,1]\text{range}(\Psi^{\phi}_{N})\cap[0,1] is finite for every NN\in\mathbb{N} (as the finiteness of the moment of order 11 of U0N(1)U_{0}^{N}(1) ensures by the law of large numbers that U0N(y)=o(ϕ(y))U_{0}^{N}(y)=o(\phi(y))). By the arguments provided before, it suffices to prove that range(Ψϕ|[M,M])[0,1]\text{range}(\Psi^{\phi}_{|[-M,M]})\cap[0,1] is finite for every M0M\geq 0. Now, for N1N\geq 1 and y0y\geq 0 we have that

|U0N(y)|\displaystyle|U_{0}^{N}(y)| |U0(y)|+0zy(U0(z)U0(z))1{U0(z)U0(z)N}\displaystyle\leq|U_{0}(y)|+\sum_{0\leq z\leq y}(U_{0}(z)-U_{0}(z-))\mathbbold{1}_{\{U_{0}(z)-U_{0}(z-)\geq N\}}
|U0(y)|+|U0(y)U01(y)|2|U0(y)|+|U01(y)|\displaystyle\leq|U_{0}(y)|+|U_{0}(y)-U_{0}^{1}(y)|\leq 2|U_{0}(y)|+|U_{0}^{1}(y)|

and similarly for y0y\leq 0. Thus almost surely

lim|y|supN1|U0N(y)|ϕ(y±M)=0\displaystyle\lim_{|y|\to\infty}\sup_{N\geq 1}\frac{|U_{0}^{N}(y)|}{\phi(y\pm M)}=0

as by the law of large numbers U01(y)=O(|y|)U_{0}^{1}(y)=O(|y|). Let K1>0K_{1}>0 such that for all |y|K1|y|\leq K_{1}, we have almost surely

supN1|U0N(y)|ϕ(y±M)12\displaystyle\sup_{N\geq 1}\frac{|U_{0}^{N}(y)|}{\phi(y\pm M)}\leq\frac{1}{2}

Let K2>0K_{2}>0 such that yϕ(y)y\mapsto\phi(y) is increasing on [K2,+)[K_{2},+\infty) and decreasing on (,K2)(-\infty,-K_{2}), then for |y|max(K1,K2)+M|y|\geq\max(K_{1},K_{2})+M and a[M,M]a\in[-M,M], we have

U0N(y)ϕ(ya)\displaystyle U_{0}^{N}(y)-\phi(y-a) U0N(y)ϕ(y±M)\displaystyle\leq U_{0}^{N}(y)-\phi(y\pm M)
12ϕ(y±M)y+\displaystyle\leq-\frac{1}{2}\phi(y\pm M)\underset{y\to+\infty}{\rightarrow}-\infty

Hence there exists K>1K>1 large enough such that

(5.8) sup|y|Ksupa[M,M]supN1(U0N(y)ϕ(ya))B:=infλ[0,1],a[M,M](U0(λ)ϕ(λa))\sup_{|y|\geq K}\sup_{a\in[-M,M]}\sup_{N\geq 1}\left(U_{0}^{N}(y)-\phi(y-a)\right)\leq B:=\inf_{\lambda\in[0,1],a\in[-M,M]}\left(U_{0}(\lambda)-\phi(\lambda-a)\right)

Now, the largest jump size of the process U0U_{0} on any compact interval [R,R][-R,R] is almost surely finite, because

[y[R,R],U0(y)U0(y)N]=1e2RΠ([N,+))N0\displaystyle\mathbb{P}[\exists y\in[-R,R],U_{0}(y)-U_{0}(y-)\geq N]=1-e^{-2R\Pi([N,+\infty))}\underset{N\to\infty}{\rightarrow}0

where Π\Pi is the Lévy measure of U0U_{0}. Hence there exists a random N~\tilde{N} such that U0N~(y)=U0(y)U_{0}^{\tilde{N}}(y)=U_{0}(y) on [K,K][-K,K], and thus if range(Ψϕ|[M,M])[0,1]\text{range}(\Psi^{\phi}_{|[-M,M]})\cap[0,1] is infinite, then there exists infinitely many λa[0,1]\lambda_{a}\in[0,1] such that

U0(λa)ϕ(λaa)U0(y)ϕ(ya) for all y\displaystyle U_{0}(\lambda_{a})-\phi(\lambda_{a}-a)\geq U_{0}(y)-\phi(y-a)\text{ for all }y

which in light of (5.8) implies that

U0N~(λa)ϕ(λaa)U0N~(y)ϕ(ya) for all y\displaystyle U_{0}^{\tilde{N}}(\lambda_{a})-\phi(\lambda_{a}-a)\geq U_{0}^{\tilde{N}}(y)-\phi(y-a)\text{ for all }y

and this is a contradiction with the fact that range(ΨϕN~)[0,1]\text{range}(\Psi^{\phi}_{\tilde{N}})\cap[0,1] is finite, thus completing the proof. ∎

Finally, we are left to prove Theorem 1.15

Proof of Theorem 1.15.

In light of Theorem 5.1, it suffices to check that for any t>0t>0 we have

lim|x||L(xt)|=+\displaystyle\lim_{|x|\to\infty}\left|L^{\prime}\left(\frac{x}{t}\right)\right|=+\infty

However, due to the convexity of LL, the function LL^{\prime} is increasing and thus the limits,

l+:=limxL(x) and l:=limxL(x)l^{+}:=\lim_{x\rightarrow\infty}L^{\prime}(x)\text{ and }l^{-}:=\lim_{x\rightarrow-\infty}L^{\prime}(x)

exist. However, due to the superlinear growth of HH (and thus of LL), it must be that l+=l^{+}=\infty and l=l^{-}=-\infty, which gives the desired result. ∎

Remark 5.2.

The class of abrupt Lévy processes mentioned in Theorem 1.15 is quite large. Indeed, it contains any linear combination of Brownian motion with linear drift and stable Lévy processes with index α(1,2)\alpha\in(1,2) with its negative jumps removed.

References

  • [1] J. Abramson. Structure of shocks in Burgers turbulence with Lévy noise initial data. J. Stat. Phys., 152(3):541–568, 2013.
  • [2] J. Abramson and S. N. Evans. Lipschitz minorants of Brownian motion and Lévy processes. Probab. Theory Relat. Fields, 158(3-4):809–857, 2014.
  • [3] D. Applebaum. Lévy processes and stochastic calculus, volume 93 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2004.
  • [4] J. Bertoin. The inviscid burgers equation with brownian initial velocity. Communications in Mathematical Physics, 193(2):397, 1998.
  • [5] J. Bertoin. Structure of shocks in burgers turbulence with stable noise initial data. Communications in Mathematical Physics, 203(3):729, 1999.
  • [6] L. C. Evans. Partial differential equations. American Mathematical Society, Providence, R.I., 2010.
  • [7] S. N. Evans and M. Ouaki. Excursions away from the Lipschitz minorant of a Lévy process. arXiv e-prints, page arXiv:1905.07038, May 2019.
  • [8] P. J. Fitzsimmons and M. Taksar. Stationary regenerative sets and subordinators. Ann. Probab., 16(3):1299–1305, 1988.
  • [9] P. Groeneboom. Brownian motion with a parabolic drift and airy functions. Probability theory and related fields, 81(1):79 – 109, 1989.
  • [10] P. Groeneboom, S. Lalley, and N. Temme. Chernoff’s distribution and differential equations of parabolic and Airy type. J. Math. Anal. Appl., 423(2):1804–1824, 2015.
  • [11] S. Janson. Brownian excursion area, Wright’s constants in graph enumeration, and other Brownian areas. Probab. Surv., 4:80–145, 2007.
  • [12] D. C. Kaspar and F. Rezakhanlou. Scalar conservation laws with monotone pure-jump Markov initial conditions. Probab. Theory Relat. Fields, 165(3-4):867–899, 2016.
  • [13] D. C. Kaspar and F. Rezakhanlou. Kinetic statistics of scalar conservation laws with piecewise-deterministic markov process data. Archive for Rational Mechanics and Analysis, 237(1):259, 2020.
  • [14] G. Menon and R. L. Pego. Universality classes in Burgers turbulence. Comm. Math. Phys., 273(1):177–202, 2007.
  • [15] G. Menon and R. Srinivasan. Kinetic theory and lax equations for shock clustering and burgers turbulence. Journal of Statistical Physics, 140(6):1 – 29, 2010.
  • [16] M. P. W. A path decomposition for markov processes. The Annals of Probability, 6(2):345, 1978.
  • [17] D. Revuz and M. Yor. Continuous martingales and Brownian motion, volume 293 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1991.
  • [18] R. L. Schilling. An introduction to Lévy and Feller processes. In From Lévy-type processes to parabolic SPDEs, Adv. Courses Math. CRM Barcelona, pages 1–126. Birkhäuser/Springer, Cham, 2016.
  • [19] G. Uribe Bravo. Bridges of Lévy processes conditioned to stay positive. Bernoulli, 20(1):190–206, 02 2014.
  • [20] V. Vigon. Abrupt Lévy processes. Stochastic Process. Appl., 103(1):155–168, 2003.
  • [21] M. Von Smoluchowski. Drei Vortrage uber Diffusion. Brownsche Bewegung und Koagulation von Kolloidteilchen. Z. Phys., 17:557–585, 1916.