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

Controlling conservation laws I: entropy–entropy flux

Wuchen Li [email protected] Department of Mathematics, University of South Carolina, Columbia Siting Liu [email protected] Department of Mathematics, University of California, Los Angeles  and  Stanley Osher [email protected] Department of Mathematics, University of California, Los Angeles
Abstract.

We study a class of variational problems for regularized conservation laws with Lax’s entropy-entropy flux pairs. We first introduce a modified optimal transport space based on conservation laws with diffusion. Using this space, we demonstrate that conservation laws with diffusion are “flux–gradient flows”. We next construct variational problems for these flows, for which we derive dual PDE systems for regularized conservation laws. Several examples, including traffic flow and Burgers’ equation, are presented. Incorporating both primal-dual algorithms and monotone schemes, we successfully compute the control of conservation laws.

Key words and phrases:
Conservation laws; Entropy–entropy flux pairs; Optimal transport; Mean-field games; Hamiltonian flows; Monotone schemes; Primal-dual algorithms.
W. Li thanks the start-up funding from the University of South Carolina, and DEPSCoR Research Collaboration white paper “Transport information geometric modeling of complex systems ”. W. Li, S. Liu and S. Osher thank the funding from AFOSR MURI FA9550-18-1-0502 and ONR grants: N00014-18-1-2527, N00014-20-1-2093, and N00014-20-1-2787.

1. Introduction

Regularized conservation laws111For simplicity, we omit the regularized throughout the paper. are essential classes of dynamics in physics, materials sciences, and mathematical modeling, with applications to inverse problems and AI sampling problems [16, 22, 35]. Examples of conservation law equations include traffic flows, Burgers’ equation, and compressible Navier-Stokes equations, etc.

Consider a system of initial value PDEs below.

tu(t,x)+(u(t,x))=β𝒞(u(t,x)),u(0,x)=u0,\partial_{t}u(t,x)+\mathcal{B}(u(t,x))=\beta\mathcal{C}(u(t,x)),\quad u(0,x)=u_{0}, (1)

where xΩnx\in\Omega\subset\mathbb{R}^{n}, u:[0,)×Ωdu\colon[0,\infty)\times\Omega\rightarrow\mathbb{R}^{d} is a unknown vector function, u0u_{0} is a given initial condition, :C(Ω;d)C(Ω;d)\mathcal{B}\colon C^{\infty}(\Omega;\mathbb{R}^{d})\rightarrow C^{\infty}(\Omega;\mathbb{R}^{d}) is a “conservative” differential operator, 𝒞:C(Ω;d)C(Ω;d)\mathcal{C}\colon C^{\infty}(\Omega;\mathbb{R}^{d})\rightarrow C^{\infty}(\Omega;\mathbb{R}^{d}) is a dissipative differential (diffusion) operator and β0\beta\geq 0 is a diffusion constant. For simplicity of presentation, we assume that Ω\Omega is a compact convex set, and the PDE (1) has periodic boundary conditions. E.g., Ω=𝕋n\Omega=\mathbb{T}^{n}, where 𝕋n\mathbb{T}^{n} is a nn–dimensional torus.

In this paper, we introduce variational problems related to the PDE (1). They generalize mean-field information dynamics [26, 27]; see Figure 1.

Refer to caption
Figure 1. We study variational problems of conservation laws in entropy-entropy flux pairs induced metric spaces.

Here we shall design suitable modified optimal transport spaces for PDE (1), namely mean-field information metric spaces. In these metric spaces, we demonstrate that the PDE (1) has a dissipative variational structure. And we name PDE (1) flux–gradient flows. We then design a control problem of PDE (1) in metric space. By finding the critical point of control problem, we derive a dual PDE system for equation (1). The primal-dual pair of PDE system satisfies a Hamiltonian flow in the metric space, where the Hamiltonian functional depends on the flux function and “kinetic energy”. Several numerical examples, including traffic flow and Burgers’ equation, are presented.

The main results are sketched below.

Assumption 1: Suppose that there exists a function G:dG\colon\mathbb{R}^{d}\rightarrow\mathbb{R}, such that

ΩG(u)(u)𝑑x=0.\int_{\Omega}G^{\prime}(u)\cdot\mathcal{B}(u)dx=0.

Assumption 2: Suppose that there exists a “symmetric nonnegative definite” operator L𝒞(u):C(Ω)C(Ω)L_{\mathcal{C}}(u)\colon C^{\infty}(\Omega)\rightarrow C^{\infty}(\Omega), such that

ΩG(u)𝒞(u)𝑑x=Ω(G(u),L𝒞(u)G(u))𝑑x0.\int_{\Omega}G^{\prime}(u)\cdot\mathcal{C}(u)dx=-\int_{\Omega}(G^{\prime}(u),L_{\mathcal{C}}(u)G^{\prime}(u))dx\leq 0.

Following [22], under assumption 1 and 2, we demonstrate that

𝒢(u(t,))=ΩG(u(t,x))𝑑x,\mathcal{G}(u(t,\cdot))=\int_{\Omega}G(u(t,x))dx,

forms a Lyapunov functional for PDE (1). In other words, along PDE (1), the time derivative of 𝒢(u)\mathcal{G}(u) is non-positive:

ddt𝒢(u(t,))=ΩG(u)tudx=ΩG(u)((u)+β𝒞(u))𝑑x=βΩG(u)𝒞(u)𝑑x=βΩ(G(u),L𝒞(u)G(u))𝑑x0.\begin{split}\frac{d}{dt}\mathcal{G}(u(t,\cdot))=&\int_{\Omega}G^{\prime}(u)\partial_{t}udx=\int_{\Omega}G^{\prime}(u)\cdot(-\mathcal{B}(u)+\beta\mathcal{C}(u))dx\\ =&\beta\int_{\Omega}G^{\prime}(u)\cdot\mathcal{C}(u)dx=-\beta\int_{\Omega}(G^{\prime}(u),L_{\mathcal{C}}(u)G^{\prime}(u))dx\leq 0.\end{split}

Based on the Lyapunov functional 𝒢\mathcal{G}, we design optimal control problems of PDE (1). In detail: let d=1d=1,

(u)=f(u),𝒞(u)=(A(u)u),𝒞(u):=(A(u)G′′(u)1),\mathcal{B}(u)=\nabla\cdot f(u),\quad\mathcal{C}(u)=\nabla\cdot(A(u)\nabla u),\quad\mathcal{L}_{\mathcal{C}}(u):=-\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla),

where f:nf\colon\mathbb{R}\rightarrow\mathbb{R}^{n} is a flux function and A,AG′′1:Ωn×nA,AG^{\prime\prime-1}\colon\Omega\rightarrow\mathbb{R}^{n\times n} are both symmetric non-negative definite matrix functions. Given a suitable potential functional :C(Ω;)\mathcal{F}\colon C^{\infty}(\Omega;\mathbb{R})\rightarrow\mathbb{R} and a terminal functional :C(Ω;)\mathcal{H}\colon C^{\infty}(\Omega;\mathbb{R})\rightarrow\mathbb{R}, consider a variational problem

infu,v,u101[Ω12(v,A(u)G′′(u)1v)𝑑x(u)]𝑑t+(u1),\inf_{u,v,u_{1}}~{}\int_{0}^{1}\Big{[}\int_{\Omega}\frac{1}{2}\big{(}v,A(u)G^{\prime\prime}(u)^{-1}v\big{)}dx-\mathcal{F}(u)\Big{]}dt+\mathcal{H}(u_{1}), (2a)
where the infimum is taken among variables v:[0,1]×Ωnv\colon[0,1]\times\Omega\rightarrow\mathbb{R}^{n}, u:[0,1]×Ωu\colon[0,1]\times\Omega\rightarrow\mathbb{R}, and u1:Ωu_{1}\colon\Omega\rightarrow\mathbb{R} satisfying
tu+f(u)+(A(u)G′′(u)1v)=β(A(u)u),u(0,x)=u0(x).\begin{split}\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}v)=\beta\nabla\cdot(A(u)\nabla u),\quad u(0,x)=u_{0}(x).\end{split} (2b)

By Proposition 8, we derive a critical point system of the optimal control problem (2). Define a Lagrange multiplier function Φ:[0,1]×Ω\Phi\colon[0,1]\times\Omega\rightarrow\mathbb{R}. Then v=Φv=\nabla\Phi, and (u,Φ)(u,\Phi) satisfies a pair of PDEs:

{tu+(u)L𝒞(u)(Φ)β𝒞(u)=0,tΦ+δδuΩ(12(Φ,L𝒞(u)Φ)((u),Φ)+β(Φ,𝒞(u)))𝑑x+δδu(u)=0,\left\{\begin{aligned} &\partial_{t}u+\mathcal{B}(u)-L_{\mathcal{C}}(u)(\Phi)-\beta\mathcal{C}(u)=0,\\ &\partial_{t}\Phi+\frac{\delta}{\delta u}\int_{\Omega}\Big{(}\frac{1}{2}(\Phi,L_{\mathcal{C}}(u)\Phi)-(\mathcal{B}(u),\Phi)+\beta(\Phi,\mathcal{C}(u))\Big{)}dx+\frac{\delta}{\delta u}\mathcal{F}(u)=0,\end{aligned}\right. (3a)
with both initial and terminal time boundary conditions
u(0,x)=u0(x),δδu1(x)(u1)+Φ(1,x)=0.u(0,x)=u_{0}(x),\quad\frac{\delta}{\delta u_{1}(x)}\mathcal{H}(u_{1})+\Phi(1,x)=0. (3b)

In above formulations, δδu\frac{\delta}{\delta u} is the L2L^{2} first variation operator w.r.t. variable uu. Here we aim to study the variational problem (2) and its critical point system (3) in the following two aspects.

  • (i)

    Modeling: We design and control models for conservation law dynamics. A typical example is that we control the kinetic energy of the density under traffic flow.

  • (ii)

    Computation: By choosing some special potential and terminal energies, the minimizer of the mean–field information control problem (2) is consistent with the initial value problem (1). This leads to stable and convergent primal-dual algorithms for initial value conservation laws.

In the literature, there are joint studies towards dynamics in mean-field information metric spaces; see [1, 2, 4, 6, 18, 32, 34, 35, 36] and many references therein. Several types of dynamics in these metric spaces have been studied in recent decades. First, gradient flows have been systematically studied in [1, 9, 14]. Next, Hamiltonian flows with generalizations to differential games have been investigated by [8, 21, 19]. A particular type of Hamiltonian flows, namely Schrödinger bridge systems, and their mean-field generalizations, are widely studied in [3, 11, 15, 12, 24, 23]. Compared to previous works, we focus on conservative-dissipative equations, which are non–gradient flow systems. In particular, we connect the conservation law equations with generalized optimal transport variational structures. Concretely, we study equation (1) associated with a least-square type control problem in the designed metric space. We remark that the control problem in (2) connects with, but is different from, the ones in [5, 6]. In detail, [6] designed variational problems in term of entropy functional (u)=𝒢(u)=ΩG(u)𝑑x-\mathcal{F}(u)=\mathcal{G}(u)=\int_{\Omega}G(u)dx, which can solve the initial value conservation laws. And [6] also enforces v=0v=0 in the control problem (2). In contrast, we control “kinetic energy” in the modified optimal transport space generated by the entropy–entropy flux pairs. Following this approach, the PDE pair (3) can be used to control conservation laws.

The paper is organized below. In section 2, we provide two motivating examples of variational problem (2). In section 3 and 4, we present the main result of this paper. We first present the variational formulation of equation (1) in modified optimal transport space. We next derive a variational problem and a pair of PDEs for conservation laws. In section 5, we design primal-dual algorithms to numerically solve the optimal control problems of conservation laws.

2. Motivation

In this section, we provide two examples of conservative-dissipative equations (1). In these examples, we demonstrate variational problems (2) and PDE pairs (3) with their physical and modeling explanations.

2.1. Viscous Burgers’ equation

Consider a one dimensional viscous Burgers’ equation

tu(t,x)+xu(t,x)22=βxxu(t,x).\partial_{t}u(t,x)+\partial_{x}\frac{u(t,x)^{2}}{2}=\beta\partial_{xx}u(t,x).

Here x\partial_{x}, xx\partial_{xx} are the first and the second derivatives w.r.t. xx, and the unknown variable is u:+×Ωu\colon\mathbb{R}_{+}\times\Omega\rightarrow\mathbb{R}. In this case, the conservative and dissipative operators satisfy

(u)=x(u22),𝒞(u)=xxu.\mathcal{B}(u)=\partial_{x}(\frac{u^{2}}{2}),\qquad\mathcal{C}(u)=\partial_{xx}u.

Define a function G:11G\colon\mathbb{R}^{1}\rightarrow\mathbb{R}^{1} by

G(u)=u22,G(u)=u.G(u)=\frac{u^{2}}{2},\quad G^{\prime}(u)=u.

Then assumption 1 is satisfied since

ΩG(u)(u)𝑑x=Ωx(u22)udx=Ωu2xudx=Ωx(u33)dx=0.\int_{\Omega}G^{\prime}(u)\cdot\mathcal{B}(u)dx=\int_{\Omega}\partial_{x}(\frac{u^{2}}{2})\cdot udx=\int_{\Omega}u^{2}\cdot\partial_{x}udx=\int_{\Omega}\partial_{x}(\frac{u^{3}}{3})dx=0.

Denote a “symmetric nonnegative definite operator” by

L𝒞(u)=xx.L_{\mathcal{C}}(u)=-\partial_{xx}.

Assumption 2 is satisfied since

ΩG(u)𝒞(u)𝑑x=Ωuxxudx=Ω|xu|2𝑑x,\int_{\Omega}G^{\prime}(u)\cdot\mathcal{C}(u)dx=\int_{\Omega}u\cdot\partial_{xx}udx=-\int_{\Omega}|\partial_{x}u|^{2}dx,

where the second equality is from the integration by parts formula.

Formulations: Variational problem (2) forms an optimal control problem for viscous Burgers’ equation. Consider

infu,v,u101[Ω12|v(t,x)|2𝑑x(u)]𝑑t+(u1),\inf_{u,v,u_{1}}~{}\int_{0}^{1}\Big{[}\int_{\Omega}\frac{1}{2}|v(t,x)|^{2}dx-\mathcal{F}(u)\Big{]}dt+\mathcal{H}(u_{1}), (4a)
such that
tu(t,x)+x(u(t,x)22)+xv(t,x)=βxxu(t,x),u(0,x)=u0(x).\partial_{t}u(t,x)+\partial_{x}(\frac{u(t,x)^{2}}{2})+\partial_{x}v(t,x)=\beta\partial_{xx}u(t,x),\quad u(0,x)=u_{0}(x). (4b)

Here, the minimizer system of variational problem (4) satisfies a pair of PDEs: v(t,x)=Φ(t,x)v(t,x)=\nabla\Phi(t,x) and

{tu(t,x)+x(u(t,x)22)+xxΦ(t,x)=βxxu(t,x),tΦ(t,x)+(u(t,x),xΦ(t,x))+δδu(u)(t,x)=βxxΦ(t,x).\left\{\begin{aligned} &\partial_{t}u(t,x)+\partial_{x}(\frac{u(t,x)^{2}}{2})+\partial_{xx}\Phi(t,x)=\beta\partial_{xx}u(t,x),\\ &\partial_{t}\Phi(t,x)+(u(t,x),\partial_{x}\Phi(t,x))+\frac{\delta}{\delta u}\mathcal{F}(u)(t,x)=-\beta\partial_{xx}\Phi(t,x).\end{aligned}\right.

Velocity control. The unknown variable uu in Burgers’ equation describes the velocity filed of the fluid over time. In the variational problem (4), we design a potential field Φ\Phi to control the evolution of velocity field uu. Here, the background dynamics of uu is the classical initial value viscous Burgers’ equation. The designed variational problem is to control a velocity field under suitable running and terminal costs.

2.2. Traffic flow equation

Consider a one dimensional traffic flow equation

tu(t,x)+(u(t,x)(1u(t,x)))=βΔu(t,x).\partial_{t}u(t,x)+\nabla\cdot\big{(}u(t,x)(1-u(t,x))\big{)}=\beta\Delta u(t,x).

Here \nabla\cdot, Δ\Delta are divergence, Laplacian operators w.r.t xx, respectively, and the unknown variable is u:+×Ω+u\colon\mathbb{R}_{+}\times\Omega\rightarrow\mathbb{R}_{+}. In this case, the conservative and the dissipative operators satisfy

(u)=(u(1u)),𝒞(u)=Δu.\mathcal{B}(u)=\nabla\cdot(u(1-u)),\qquad\mathcal{C}(u)=\Delta u.

Define a function G:+1+1G\colon\mathbb{R}_{+}^{1}\rightarrow\mathbb{R}_{+}^{1} by

G(u)=uloguu,G(u)=logu.G(u)=u\log u-u,\quad G^{\prime}(u)=\log u.

Then assumption 1 is satisfied since

ΩG(u)(u)𝑑x=Ωlogu(u(1u))𝑑x=ΩΨ(u)𝑑x=0,\int_{\Omega}G^{\prime}(u)\cdot\mathcal{B}(u)dx=\int_{\Omega}\log u\cdot\nabla\cdot(u(1-u))dx=\int_{\Omega}\nabla\cdot\Psi(u)dx=0,

where Ψ(u)=0u(12z)logzdz\Psi(u)=\int_{0}^{u}(1-2z)\log zdz. Denote a “symmetric nonnegative definite operator” by

L𝒞(u)=(u).L_{\mathcal{C}}(u)=-\nabla\cdot(u\nabla).

Assumption 2 is satisfied since

ΩG(u)𝒞(u)𝑑x=ΩloguΔudx=Ω(logu,u)𝑑x=Ωlogu2u𝑑x=Ω(logu,(ulogu))𝑑x,\begin{split}\int_{\Omega}G^{\prime}(u)\cdot\mathcal{C}(u)dx=&\int_{\Omega}\log u\cdot\Delta udx\\ =&\int_{\Omega}(\nabla\log u,\nabla u)dx\\ =&-\int_{\Omega}\|\nabla\log u\|^{2}udx\\ =&-\int_{\Omega}\big{(}\log u,-\nabla\cdot(u\nabla\log u)\big{)}dx,\end{split}

where the second equality is from the integration by parts formula and the third equality holds by the fact that uu=logu\frac{\nabla u}{u}=\nabla\log u, i.e., u=ulogu\nabla u=u\nabla\log u.

Formulations: Variational problem (2) forms an optimal control problem for traffic flows. In detail, consider

infu,v,v101[Ω12v(t,x)2u(t,x)𝑑x(u)]𝑑t+(u1),\inf_{u,v,v_{1}}~{}\int_{0}^{1}\Big{[}\int_{\Omega}\frac{1}{2}\|v(t,x)\|^{2}u(t,x)dx-\mathcal{F}(u)\Big{]}dt+\mathcal{H}(u_{1}), (5a)
such that
0u(t,x)1,for all t[0,1],0\leq u(t,x)\leq 1,\quad\textrm{for all $t\in[0,1]$},
and
tu(t,x)+(u(t,x)(1u(t,x)))+(u(t,x)v(t,x))=βΔu(t,x),u(0,x)=u0(x).\partial_{t}u(t,x)+\nabla\cdot\big{(}u(t,x)(1-u(t,x))\big{)}+\nabla\cdot\big{(}u(t,x)v(t,x)\big{)}=\beta\Delta u(t,x),\quad u(0,x)=u_{0}(x). (5b)

Here, the minimizer system of variational problem (4) satisfies a pair of PDEs. When u(0,1)u\in(0,1), there exists a scalar function Φ\Phi, such that v(t,x)=Φ(t,x)v(t,x)=\nabla\Phi(t,x) and

{tu(t,x)+(u(t,x)(1u(t,x)))+(u(t,x)Φ(t,x))=βΔu(t,x),tΦ(t,x)+(12u(t,x),Φ(t,x))+12Φ(t,x)2+δδu(t,x)(u)=βΔΦ(t,x).\left\{\begin{aligned} &\partial_{t}u(t,x)+\nabla\cdot\big{(}u(t,x)(1-u(t,x))\big{)}+\nabla\cdot\big{(}u(t,x)\nabla\Phi(t,x)\big{)}=\beta\Delta u(t,x),\\ &\partial_{t}\Phi(t,x)+\big{(}1-2u(t,x),\nabla\Phi(t,x)\big{)}+\frac{1}{2}\|\nabla\Phi(t,x)\|^{2}+\frac{\delta}{\delta u(t,x)}\mathcal{F}(u)=-\beta\Delta\Phi(t,x).\end{aligned}\right.

Position control. The unknown variable uu in traffic flows represents the density function of cars (particles) in a given spatial domain. Here, the background dynamics of uu is the classical traffic flow. The control variable is the velocity for enforcing each car’s velocity in addition to its background traffic flow dynamics. The goal is to control the “total enforced kinetic energy” of all cars, in which individual cars can determine their velocities through both noises and traffic flow interactions.

We shall demonstrate that variational problems (2) can be formulated in term of general conservation law equations associated with entropy-entropy flux pairs.

3. Entropy-entropy flux and flux–gradient flows

In this section, we first recall Lax’s entropy-entropy flux pairs for conservation law equations; see [16, 22]. Here the entropy-entropy flux pair is used to construct a Lyapunov functional for PDE (1). Using this Lyapunov functional with the dissipative operator, we next review and formulate both metric spaces and gradient flows; see [1, 9, 32]. Combining these facts with flux functions, we name PDE (1) flux–gradient flows in metric spaces. In later on sections, we shall demonstrate that the flux–gradient flow formulation is useful in designing control problems of conservation laws.

3.1. Entropy-Entropy flux pairs and Lyapunov functionals

Consider

u:+×Ω1,(u)=f(u),𝒞(u)=(A(u)u).u\colon\mathbb{R}_{+}\times\Omega\rightarrow\mathbb{R}^{1},\quad\mathcal{B}(u)=\nabla\cdot f(u),\quad\mathcal{C}(u)=\nabla\cdot(A(u)\nabla u).

In this case, equation (1) satisfies

tu(t,x)+f(u(t,x))=β(A(u)u),\partial_{t}u(t,x)+\nabla\cdot f(u(t,x))=\beta\nabla\cdot(A(u)\nabla u), (6)

where u:nu\colon\mathbb{R}^{n}\rightarrow\mathbb{R} is a scalar function, f=(f1,,fn)f=(f_{1},\cdots,f_{n}) is a flux vector function, and A=(Aij)1i,jnn×nA=(A_{ij})_{1\leq i,j\leq n}\in\mathbb{R}^{n\times n} is a semi-positive definite matrix function. If β=0\beta=0, equation (6) is a scalar conservation law equation

tu(t,x)+f(u(t,x))=0.\partial_{t}u(t,x)+\nabla\cdot f(u(t,x))=0. (7)
Definition 1 (Entropy-entropy flux pair condition [22]).

We call (G,Ψ)(G,\Psi) an entropy-entropy flux pair for the conservation law (7) if there exists a convex function G:G\colon\mathbb{R}\rightarrow\mathbb{R}, and Ψ:n\Psi\colon\mathbb{R}\rightarrow\mathbb{R}^{n}, such that

Ψ(u)=f(u)G(u).\Psi^{\prime}(u)=f^{\prime}(u)G^{\prime}(u).
Remark 1.

We remark that the entropy-entropy flux condition is trivial for scalar conservation laws. Every differentiable convex function GG satisfies this condition. For simplicity of presentation, we only focus on the scalar case, and leave the study of conservation law systems in future work.

In fact, Lax’s entropy–entropy flux pair introduces a class of functionals, which can be used as Lyapunov functionals for PDE (1). Denote

𝒢(u)=ΩG(u)𝑑x.\mathcal{G}(u)=\int_{\Omega}G(u)dx.

For assumption 1,

ΩG(u)(u)𝑑x=ΩG(u)f(u)𝑑x=ΩG(u)(f(u),u)𝑑x=Ω(Ψ(u),u)𝑑x=ΩΨ(u)𝑑x=0,\begin{split}\int_{\Omega}G^{\prime}(u)\cdot\mathcal{B}(u)dx=&\int_{\Omega}G^{\prime}(u)\nabla\cdot f(u)dx\\ =&\int_{\Omega}G^{\prime}(u)(f^{\prime}(u),\nabla u)dx\\ =&\int_{\Omega}(\Psi^{\prime}(u),\nabla u)dx\\ =&\int_{\Omega}\nabla\cdot\Psi(u)dx=0,\end{split}

where we apply the fact that f(u)G(u)=Ψ(u)f^{\prime}(u)G^{\prime}(u)=\Psi^{\prime}(u) and ΩΨ(u)𝑑x=0\int_{\Omega}\nabla\cdot\Psi(u)dx=0 in the last equality.

For assumption 2,

ΩG(u)𝒞(u)𝑑x=ΩG(u)(A(u)u)𝑑x=Ω(G(u),A(u)u)𝑑x=Ω(G(u),A(u)G′′(u)1G(u))𝑑x,\begin{split}\int_{\Omega}G^{\prime}(u)\cdot\mathcal{C}(u)dx=&\int_{\Omega}G^{\prime}(u)\nabla\cdot(A(u)\nabla u)dx\\ =&-\int_{\Omega}\big{(}\nabla G^{\prime}(u),A(u)\nabla u\big{)}dx\\ =&-\int_{\Omega}\big{(}\nabla G^{\prime}(u),A(u)G^{\prime\prime}(u)^{-1}\nabla G^{\prime}(u)\big{)}dx,\end{split}

where we apply G(u)=G′′(u)u\nabla G^{\prime}(u)=G^{\prime\prime}(u)\nabla u in the last equality. We observe a fact that A(u)G′′(u)10A(u)G^{\prime\prime}(u)^{-1}\succeq 0, since AA is nonnegative definite and G′′(u)>0G^{\prime\prime}(u)>0. Hence we know that

t𝒢(u)=ΩG(u)tudx=ΩG(u)(u)𝑑x+βΩG(u)𝒞(u)𝑑x=βΩ(G(u),A(u)G′′(u)1G(u))𝑑x0.\begin{split}\partial_{t}\mathcal{G}(u)=&\int_{\Omega}G^{\prime}(u)\cdot\partial_{t}udx\\ =&-\int_{\Omega}G^{\prime}(u)\cdot\mathcal{B}(u)dx+\beta\int_{\Omega}G^{\prime}(u)\cdot\mathcal{C}(u)dx\\ =&-\beta\int_{\Omega}\big{(}\nabla G^{\prime}(u),A(u)G^{\prime\prime}(u)^{-1}\nabla G^{\prime}(u)\big{)}dx\leq 0.\end{split}

This implies that 𝒢(u)\mathcal{G}(u) is a Lyapunov functional for PDE (6).

3.2. Metric spaces and gradient flows

We next provide a condition to define a metric space for the unknown variable uu. Here, the metric space connects Lyapunov functionals with dissipative operators through gradient descent flows.

Definition 2 (Entropy-entropy flux-metric condition).

We call (G,Ψ)(G,\Psi) an entropy-entropy flux pair-metric for equation (6) if there exists a convex function G:G\colon\mathbb{R}\rightarrow\mathbb{R}, and Ψ:n\Psi\colon\mathbb{R}\rightarrow\mathbb{R}^{n}, such that

Ψ(u)=f(u)G(u),A(u)G′′(u)1 is a semi positive definite symmetric matrix function.\Psi^{\prime}(u)=f^{\prime}(u)G^{\prime}(u),\quad\textrm{$A(u)G^{\prime\prime}(u)^{-1}$ is a semi positive definite symmetric matrix function}.

Under the entropy-entropy flux-metric condition, assumption 2 implies a metric operator below. Define the space of function uu by

={uC(Ω):Ωu(x)𝑑x=constant}.\mathcal{M}=\Big{\{}u\in C^{\infty}(\Omega)\colon\int_{\Omega}u(x)dx=\mathrm{constant}\Big{\}}.

The tangent space of (u)\mathcal{M}(u) at point uu is defined by

Tu={σC(Ω):Ωσ(x)𝑑x=0}.T_{u}\mathcal{M}=\Big{\{}\sigma\in C^{\infty}(\Omega)\colon\int_{\Omega}\sigma(x)dx=0\Big{\}}.

Denote an elliptic operator L𝒞:C(Ω)C()L_{\mathcal{C}}\colon C^{\infty}(\Omega)\rightarrow C^{\infty}(\mathbb{R}) by

L𝒞(u)=(A(u)G′′(u)1).L_{\mathcal{C}}(u)=-\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla).

We are ready to define the metric inner product for the dissipation operator.

Definition 3 (Metric).

The inner product 𝐠(u):Tu×Tu\mathbf{g}(u)\colon{T_{u}}\mathcal{M}\times{T_{u}}\mathcal{M}\rightarrow\mathbb{R} is given below.

𝐠(u)(σ1,σ2)=Ω(Φ1,L𝒞(u)Φ2)𝑑x=ΩΦ1(A(u)G′′(u)1Φ2)𝑑x=Ω(Φ1,A(u)G′′(u)1Φ2)𝑑x=Ωσ1Φ2𝑑x=Ωσ2Φ1𝑑x,\begin{split}\mathbf{g}(u)(\sigma_{1},\sigma_{2})=&\int_{\Omega}(\Phi_{1},L_{\mathcal{C}}(u)\Phi_{2})dx\\ =&-\int_{\Omega}\Phi_{1}\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\Phi_{2})dx\\ =&\int_{\Omega}(\nabla\Phi_{1},A(u)G^{\prime\prime}(u)^{-1}\nabla\Phi_{2})dx\\ =&\int_{\Omega}\sigma_{1}\Phi_{2}dx=\int_{\Omega}\sigma_{2}\Phi_{1}dx,\end{split}

where ΦiC(Ω)\Phi_{i}\in C^{\infty}(\Omega) satisfies

σi=(A(u)G′′(u)1Φi),i=1,2.\sigma_{i}=-\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\Phi_{i}),\quad i=1,2.

From now on, we call (,𝐠)(\mathcal{M},\mathbf{g}) the metric space. We next review and present the gradient decent flow in metric space (,𝐠)(\mathcal{M},\mathbf{g}). Here the dissipative part of PDE (1) comes from the gradient flow of the proposed Lyapunov (entropy) functional.

Proposition 4 (Gradient flow).

Given an energy functional :\mathcal{E}\colon\mathcal{M}\rightarrow\mathbb{R}, the gradient flow of \mathcal{F} in (,𝐠)(\mathcal{M},\mathbf{g}) satisfies

tu=(A(u)G′′(u)1δδu(u)).\partial_{t}u=\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u}\mathcal{E}(u)).

If (u)=𝒢(u)=ΩG(u)𝑑x\mathcal{E}(u)=\mathcal{G}(u)=\int_{\Omega}G(u)dx, then the above gradient flow satisfies

tu=(A(u)u).\partial_{t}u=\nabla\cdot(A(u)\nabla u).
Proof.

The derivation of gradient flow follows from the definition. For the completeness of this paper, we still present the proof here. The gradient operator gradTu\mathrm{grad}\mathcal{E}\in T_{u}\mathcal{M} is defined by

𝐠(u)(grad(u),σ)=Ω(δδu(u),σ)𝑑x,\mathbf{g}(u)\big{(}\mathrm{grad}\mathcal{E}(u),\sigma\big{)}=\int_{\Omega}(\frac{\delta}{\delta u}\mathcal{E}(u),\sigma)dx,

for any σTu\sigma\in T_{u}\mathcal{M}. Let σ=L𝒞(u)(Φ)=(A(u)G′′(u)1Φ)\sigma=L_{\mathcal{C}}(u)(\Phi)=-\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\Phi), the above definition forms

𝐠(u)(grad(u),σ)=Ω(grad(u),Φ)𝑑x=Ω(δδu(u),L𝒞(u)(Φ))𝑑x=Ω(L𝒞(u)(δδu(u)),Φ)𝑑x.\begin{split}\mathbf{g}(u)\big{(}\mathrm{grad}\mathcal{E}(u),\sigma\big{)}=&\int_{\Omega}(\mathrm{grad}\mathcal{E}(u),\Phi)dx\\ =&\int_{\Omega}(\frac{\delta}{\delta u}\mathcal{E}(u),L_{\mathcal{C}}(u)(\Phi))dx\\ =&\int_{\Omega}(L_{\mathcal{C}}(u)(\frac{\delta}{\delta u}\mathcal{E}(u)),\Phi)dx.\end{split}

Since the above equality holds for any smooth function Φ\Phi, we let

grad(u)=L𝒞(u)(δδu(u))=(A(u)G′′(u)1δδu(u)).\mathrm{grad}\mathcal{E}(u)=L_{\mathcal{C}}(u)\big{(}\frac{\delta}{\delta u}\mathcal{E}(u)\big{)}=-\nabla\cdot\Big{(}A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u}\mathcal{E}(u)\Big{)}.

Hence the gradient decent flow satisfies

tu=grad(u)=(A(u)G′′(u)1δδu(u)).\partial_{t}u=-\mathrm{grad}\mathcal{E}(u)=\nabla\cdot\Big{(}A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u}\mathcal{E}(u)\Big{)}.

If (u)=𝒢(u)\mathcal{E}(u)=\mathcal{G}(u), then

tu=(A(u)G′′(u)1δδu𝒢(u))=(A(u)G′′(u)1G(u))=(A(u)G′′(u)1G′′(u)u)=(A(u)u).\begin{split}\partial_{t}u=&\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u}\mathcal{G}(u))\\ =&\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla G^{\prime}(u))\\ =&\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}G^{\prime\prime}(u)\nabla u)\\ =&\nabla\cdot(A(u)\nabla u).\end{split}

Remark 2.

An example of metric gg is the Wasserstein-2 metric, which has been widely studied in optimal transport literature [1, 36]. In other words, let G(u)=uloguuG(u)=u\log u-u, A(u)=𝕀A(u)=\mathbb{I}, then L𝒞(u)=(u)L_{\mathcal{C}}(u)=-\nabla\cdot(u\nabla). In this case, the heat equation is the gradient flow of entropy functional Ωuloguudx\int_{\Omega}u\log u-udx in Wasserstein-2 space. Recently, generalized optimal transport metrics and nonlinear diffusions have been widely studied in [9, 14, 31].

3.3. Flux–gradient flows

In summary, we illustrate the relation among entropy-entropy flux pairs, gradient flows and PDE (6). On the one hand, the entropy-entropy flux pairs introduce a Lyapunov functional, along which the entropy-entropy flux flow is non positive. In addition, dGdt+Ψ0\frac{dG}{dt}+\nabla\cdot\Psi\leq 0. This entropy condition [22] picks out the unique physical weak solution for inviscid conservation law (7). On the other hand, both Lyapunov functional and dissipative operator define a metric space, under which the dissipative operator forms the gradient flow. These facts imply a formulation for PDE (6). We call it flux–gradient flows.

In detail, equation (6) can be written below.

tu+f(u)=β(A(u)G′′(u)1δδu𝒢(u)),\partial_{t}u+\nabla\cdot f(u)=\beta\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u}\mathcal{G}(u)), (8)

where the flux function ff satisfies

Ωf(u)δδu𝒢(u)𝑑x=0.\int_{\Omega}f(u)\cdot\nabla\frac{\delta}{\delta u}\mathcal{G}(u)dx=0.

The above formulation of the PDE (6) is a gradient flow equation added with the flux function. We can consider its general formulation. We keep the metric operator invariant and replace the Lyapunov functional 𝒢(u)\mathcal{G}(u) by a general energy functional (u)\mathcal{E}(u).

Definition 5 (Flux–gradient flow).

Given an energy functional :\mathcal{E}\colon\mathcal{M}\rightarrow\mathbb{R}, consider a class of PDE

tu+(f1(x,u))=β(A(u)G′′(u)1δδu(u)),\partial_{t}u+\nabla\cdot(f_{1}(x,u))=\beta\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u}\mathcal{E}(u)), (9)

where f1:Ω×1nf_{1}\colon\Omega\times\mathbb{R}^{1}\rightarrow\mathbb{R}^{n} is a flux function satisfying

Ωf1(x,u)δδu(x)(u)𝑑x=0.\int_{\Omega}f_{1}(x,u)\cdot\nabla\frac{\delta}{\delta u(x)}\mathcal{E}(u)dx=0. (10)

If (u)=𝒢(u)=ΩG(u)𝑑x\mathcal{E}(u)=\mathcal{G}(u)=\int_{\Omega}G(u)dx and f1(x,u)=f(u)f_{1}(x,u)=f(u), then equation (9) forms PDE (6).

Equations in formulation (9) provide a class of conservative–dissipative dynamics.

Proposition 6 (Entropy-entropy flux-production).

Energy functional (u)\mathcal{E}(u) is a Lyapunov functional for PDE (9). In other words, the following dissipation result holds. Suppose u(t,x)u(t,x) is the solution of equation (9), then

ddt(u(t,))=β(u(t,))0,\frac{d}{dt}\mathcal{E}(u(t,\cdot))=-\beta\mathcal{I}_{\mathcal{E}}(u(t,\cdot))\leq 0,

where the functional :+\mathcal{I}_{\mathcal{E}}\colon\mathcal{M}\rightarrow\mathbb{R}_{+} is defined by

(u)=Ω(δδu(u),A(u)G′′(u)1δδu(u))𝑑x.\mathcal{I}_{\mathcal{E}}(u)=\int_{\Omega}\big{(}\nabla\frac{\delta}{\delta u}\mathcal{E}(u),A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u}\mathcal{E}(u)\big{)}dx. (11)
Proof.

The proof follows directly from Definition 5. In other words,

ddt(u)=Ωδδu(u)tudx=Ω[δδu(u)f1(x,u)+βδδu(u)(A(u)G′′(u)1δδu(u))]𝑑x=Ω(f1(x,u),δδu(u))𝑑xβΩ(δδu(u),A(u)G′′(u)1δδu(u))𝑑x=β(u),\begin{split}\frac{d}{dt}\mathcal{E}(u)=&\int_{\Omega}\frac{\delta}{\delta u}\mathcal{E}(u)\cdot\partial_{t}udx\\ =&\int_{\Omega}\Big{[}-\frac{\delta}{\delta u}\mathcal{E}(u)\nabla\cdot f_{1}(x,u)+\beta\frac{\delta}{\delta u}{\mathcal{E}(u)}\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u}\mathcal{E}(u))\Big{]}dx\\ =&\int_{\Omega}(f_{1}(x,u),\nabla\frac{\delta}{\delta u}\mathcal{E}(u))dx-\beta\int_{\Omega}\big{(}\nabla\frac{\delta}{\delta u}\mathcal{E}(u),A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u}\mathcal{E}(u)\big{)}dx\\ =&-\beta\mathcal{I}_{\mathcal{E}}(u),\end{split}

where the third equality applies the integration by parts formula and the last equality uses condition (10). ∎

Remark 3.

We remark that if (u)=𝒢(u)=Ω(uloguu)𝑑x\mathcal{E}(u)=\mathcal{G}(u)=\int_{\Omega}(u\log u-u)dx and A(u)=𝕀A(u)=\mathbb{I}, then the functional

(u)=Ωlogu2u𝑑x=Ωu(x)2u(x)𝑑x,\mathcal{I}_{\mathcal{E}}(u)=\int_{\Omega}\|\nabla\log u\|^{2}udx=\int_{\Omega}\frac{\|\nabla u(x)\|^{2}}{u(x)}dx,

is known as the Fisher information functional. A known fact is that the dissipation of Lyapunov/entropy functional 𝒢(u)=Ωuloguudx\mathcal{G}(u)=\int_{\Omega}u\log u-udx along the heat flow equals to the negative Fisher information functional. This fact is called the de Bruijn equality. Here, the de-Bruijn type equalities hold naturally in conservation laws with entropy-entropy flux pairs. In future works, we shall study the dynamical effect of flux function in metric spaces; see related techniques developed in [26].

Remark 4.

We remark that equation (9) are generalized variational formulations for flux–gradient flows in metric spaces. They have potential applications in designing Markov-Chain-Monte-Carlo algorithms, and deriving the neural network variational algorithms for conservation laws; see [28]. We leave the detailed studies of these areas in future works.

4. Controlling Conservation laws

In this section, we present the main results of this paper. We study the variational problems for conservation laws (6). From now on, we assume that the entropy-entropy flux-metric condition in Definition 2 holds.

We first design an optimal control problem over flux–gradient flows in a metric space.

Definition 7 (Optimal control of conservation laws).
Given smooth functionals \mathcal{F}, :\mathcal{H}\colon\mathcal{M}\rightarrow\mathbb{R}, consider a variational problem
infu,v,u101[Ω12(v,A(u)G′′(u)1v)𝑑x(u)]𝑑t+(u1),\inf_{u,v,u_{1}}~{}\int_{0}^{1}\Big{[}\int_{\Omega}\frac{1}{2}\big{(}v,A(u)G^{\prime\prime}(u)^{-1}v\big{)}dx-\mathcal{F}(u)\Big{]}dt+\mathcal{H}(u_{1}), (12a)
where the infimum is taken among variables v:[0,1]×Ωnv\colon[0,1]\times\Omega\rightarrow\mathbb{R}^{n}, u:[0,1]×Ωu\colon[0,1]\times\Omega\rightarrow\mathbb{R}, and u1:Ωu_{1}\colon\Omega\rightarrow\mathbb{R} satisfying
tu+f(u)+(A(u)G′′(u)1v)=β(A(u)u),u(0,x)=u0(x).\begin{split}\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}v)=\beta\nabla\cdot(A(u)\nabla u),\quad u(0,x)=u_{0}(x).\end{split} (12b)

Here the equation (12b) is a control dynamic for the conservation law (6). The objective functional is an enforced “kinetic energy” minus a “potential energy” in the metric space (,𝐠)(\mathcal{M},\mathbf{g}). If the control variable v(t,x)=0v(t,x)=0 for all t=[0,1]t=[0,1], xΩx\in\Omega, then dynamics (12b) becomes the original conservation law equation (6).

Remark 5.

We notice that variational problem (12) is a generalized dynamical optimal transport problem. In other words, if u:[0,1]×Ω+u\colon[0,1]\times\Omega\rightarrow\mathbb{R}_{+}, A(u)=𝕀A(u)=\mathbb{I}, G(u)=uloguuG(u)=u\log u-u, β=0\beta=0, f=0f=0, (u)=0\mathcal{F}(u)=0 and u1u_{1} is a fixed function with Ωu0𝑑x=Ωu1𝑑x\int_{\Omega}u_{0}dx=\int_{\Omega}u_{1}dx, then problem (12) forms

infu,v{01Ω12v2udxdt:tu+(uv)=0,u(0,x)=u0(x),u(1,x)=u1(x)}.\inf_{u,v}~{}\Big{\{}\int_{0}^{1}\int_{\Omega}\frac{1}{2}\|v\|^{2}udxdt\colon\partial_{t}u+\nabla\cdot(uv)=0,\quad u(0,x)=u_{0}(x),\quad u(1,x)=u_{1}(x)\Big{\}}.

The above minimization is known as Benamou-Brenier’s formula [4] studied in classical optimal transport problems. Here the variational problem (12) generalizes the dynamical optimal transport, which contains the inverse of the Hessian operator of entropy functionals to model the “kinetic energy”. In particular, we formulate the conservation laws in the constraint set.

We next derive critical point systems of variational problem (12). They are Hamiltonian flows in (,𝐠)(\mathcal{M},\mathbf{g}) associated with conservation laws.

Proposition 8 (Hamiltonian flows of conservation laws).

The critical point system of variational problem (12) is given below. There exists a function Φ:[0,1]×Ω\Phi\colon[0,1]\times\Omega\rightarrow\mathbb{R}, such that

v(t,x)=Φ(t,x),v(t,x)=\nabla\Phi(t,x),

and

{tu+f(u)+(A(u)G′′(u)1Φ)=β(A(u)u),tΦ+(Φ,f(u))+12(Φ,(A(u)G′′(u)1)Φ)+δδu(u)=β(A(u)Φ)+β(Φ,A(u)u).\left\{\begin{aligned} &\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\Phi)=\beta\nabla\cdot(A(u)\nabla u),\\ &\partial_{t}\Phi+(\nabla\Phi,f^{\prime}(u))+\frac{1}{2}(\nabla\Phi,(A(u)G^{\prime\prime}(u)^{-1})^{\prime}\nabla\Phi)+\frac{\delta}{\delta u}\mathcal{F}(u)=-\beta\nabla\cdot(A(u)\nabla\Phi)+\beta(\nabla\Phi,A^{\prime}(u)\nabla u).\end{aligned}\right. (13)

Here represents the derivative w.r.t. variable uu. The initial and terminal time conditions satisfy

u(0,x)=u0(x),δδu1(u1)+Φ(1,x)=0.u(0,x)=u_{0}(x),\quad\frac{\delta}{\delta u_{1}}\mathcal{H}(u_{1})+\Phi(1,x)=0.
Proof.

We first rewrite the variables (u,v)(u,v) in variational formula (12) by (u,m)(u,m), where

m(t,x)=A(u(t,x))G′′(u(t,x))1v(t,x)=V(u(t,x))v(t,x).m(t,x)=A(u(t,x))G^{\prime\prime}(u(t,x))^{-1}v(t,x)=V(u(t,x))v(t,x).

Here we denote V(u(t,x))=A(u(t,x))G′′(u(t,x))1V(u(t,x))=A(u(t,x))G^{\prime\prime}(u(t,x))^{-1}. In this case, the variational problem (12) forms

infm,u,u1{01[Ω12(m,V(u)1m)dx(u)]dt+(u1):tu+f(u)+m=β(A(u)u),fixed u0.}.\begin{split}&\inf_{m,u,u_{1}}\Big{\{}\int_{0}^{1}\Big{[}\int_{\Omega}\frac{1}{2}(m,V(u)^{-1}m)dx-\mathcal{F}(u)\Big{]}dt+\mathcal{H}(u_{1})\colon\\ &\qquad\quad\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot m=\beta\nabla\cdot(A(u)\nabla u),\quad\textrm{fixed $u_{0}$.}\Big{\}}.\end{split} (14)

Denote the Lagrange multiplier of problem (14) by Φ:[0,1]×Ω\Phi\colon[0,1]\times\Omega\rightarrow\mathbb{R}. Consider the following saddle point problem

infm,u,u1supΦ(m,u,u1,Φ).\begin{split}\inf_{m,u,u_{1}}\sup_{\Phi}\quad\mathcal{L}(m,u,u_{1},\Phi).\end{split}

In the above formula, we have

(m,u,u1,Φ)=01Ω[12(m,V(u)1m)+Φ(tu+f(u)+mβ(A(u)u))]𝑑x𝑑t01(u)𝑑t+(u1)=01Ω[12(m,V(u)1m)+Φ(f(u)+mβ(A(u)u))]𝑑x𝑑t+Ω(Φ(1,x)u1(x)Φ(0,x)u0(x))𝑑x01ΩtΦudxdt01(u)𝑑t+(u1),\begin{split}\mathcal{L}(m,u,u_{1},\Phi)=&\int_{0}^{1}\int_{\Omega}\Big{[}\frac{1}{2}(m,V(u)^{-1}m)+\Phi\Big{(}\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot m-\beta\nabla\cdot(A(u)\nabla u)\Big{)}\Big{]}dxdt\\ &\quad-\int_{0}^{1}\mathcal{F}(u)dt+\mathcal{H}(u_{1})\\ =&\int_{0}^{1}\int_{\Omega}\Big{[}\frac{1}{2}(m,V(u)^{-1}m)+\Phi\Big{(}\nabla\cdot f(u)+\nabla\cdot m-\beta\nabla\cdot(A(u)\nabla u)\Big{)}\Big{]}dxdt\\ &+\int_{\Omega}\big{(}\Phi(1,x)u_{1}(x)-\Phi(0,x)u_{0}(x)\big{)}dx-\int_{0}^{1}\int_{\Omega}\partial_{t}\Phi udxdt-\int_{0}^{1}\mathcal{F}(u)dt+\mathcal{H}(u_{1}),\end{split}

where we use the integration by parts formula w.r.t. tt in the second equality. In other words,

01ΩΦtudxdt=ΩΦ(1,x)u(1,x)𝑑xΩΦ(0,x)u(0,x)𝑑x01ΩtΦudxdt.\int_{0}^{1}\int_{\Omega}\Phi\partial_{t}udxdt=\int_{\Omega}\Phi(1,x)u(1,x)dx-\int_{\Omega}\Phi(0,x)u(0,x)dx-\int_{0}^{1}\int_{\Omega}\partial_{t}\Phi udxdt.

We next derive the critical point for the above saddle point problem. In other words, consider

{δδm=0δδu=0δδu1=0δδΦ=0{V(u)1m=Φ,12(m,V(u)1V(u)V(u)1m)δδutΦ(Φ,f(u))β(A(u)Φ)+β(Φ,A(u)u)=0,Φ1+δδu1(u1)=0,tu+f(u)+mβ(A(u)u)=0,\left\{\begin{split}&\frac{\delta}{\delta m}\mathcal{L}=0\\ &\frac{\delta}{\delta u}\mathcal{L}=0\\ &\frac{\delta}{\delta u_{1}}\mathcal{L}=0\\ &\frac{\delta}{\delta\Phi}\mathcal{L}=0\end{split}\right.\quad\Rightarrow\quad\left\{\begin{split}&V(u)^{-1}m=\nabla\Phi,\\ &-\frac{1}{2}(m,V(u)^{-1}V(u)^{\prime}V(u)^{-1}m)-\frac{\delta}{\delta u}\mathcal{F}-\partial_{t}\Phi-(\nabla\Phi,f^{\prime}(u))\\ &\hskip 105.2751pt-\beta\nabla\cdot(A(u)\nabla\Phi)+\beta(\nabla\Phi,A^{\prime}(u)\nabla u)=0,\\ &\Phi_{1}+\frac{\delta}{\delta u_{1}}\mathcal{H}(u_{1})=0,\\ &\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot m-\beta\nabla\cdot(A(u)\nabla u)=0,\end{split}\right.

where δδm\frac{\delta}{\delta m}, δδu\frac{\delta}{\delta u}, δδu1\frac{\delta}{\delta u_{1}}, δδΦ\frac{\delta}{\delta\Phi} are L2L^{2} first variational derivatives w.r.t. functions mm, uu, u1u_{1}, Φ\Phi, respectively. We thus derive the pair of PDEs (13) in (Ω)\mathcal{M}(\Omega). ∎

We next present the Hamiltonian formalism for the PDE system (13).

Proposition 9 (Hamiltonian flows in metric space).

PDE system (13) has the following Hamiltonian flow formulation.

tu=δδΦ𝒢(u,Φ),tΦ=δδu𝒢(u,Φ),\partial_{t}u=\frac{\delta}{\delta\Phi}\mathcal{H}_{\mathcal{G}}(u,\Phi),\quad\partial_{t}\Phi=-\frac{\delta}{\delta u}\mathcal{H}_{\mathcal{G}}(u,\Phi),

where we define the Hamiltonian functional 𝒢:×C(Ω)\mathcal{H}_{\mathcal{G}}\colon\mathcal{M}\times C^{\infty}(\Omega)\rightarrow\mathbb{R} by

𝒢(u,Φ)=Ω[12(Φ,A(u)G′′(u)1Φ)+(Φ,f(u))β(Φ,A(u)u)]𝑑x+(u).\mathcal{H}_{\mathcal{G}}(u,\Phi)=\int_{\Omega}\Big{[}\frac{1}{2}(\nabla\Phi,A(u)G^{\prime\prime}(u)^{-1}\nabla\Phi)+(\nabla\Phi,f(u))-\beta(\nabla\Phi,A(u)\nabla u)\Big{]}dx+\mathcal{F}(u). (15)

In other words, the Hamiltonian functional 𝒢(u,Φ)\mathcal{H}_{\mathcal{G}}(u,\Phi) is conserved along dynamics (13). Suppose (u,Φ)(u,\Phi) are solutions of equation (13), then

ddt𝒢(u,Φ)=0.\frac{d}{dt}\mathcal{H}_{\mathcal{G}}(u,\Phi)=0.
Proposition 10 (Functional Hamilton-Jacobi equations of conservation laws).

The Hamilton-Jacobi equation in (,𝐠)(\mathcal{M},\mathbf{g}) for equation (13) satisfies

t𝒰(t,u)+Ω[12(δδu(x)𝒰(t,u),A(u)G′′(u)1δδu(x)𝒰(t,u))+(δδu(x)𝒰(t,u),f(u))β(δδu(x)𝒰(t,u),A(u)u)]dx+(u)=0,\begin{split}&\partial_{t}\mathcal{U}(t,u)+\int_{\Omega}\Big{[}\frac{1}{2}\big{(}\nabla\frac{\delta}{\delta u(x)}\mathcal{U}(t,u),A(u)G^{\prime\prime}(u)^{-1}\nabla\frac{\delta}{\delta u(x)}\mathcal{U}(t,u)\big{)}+\big{(}\nabla\frac{\delta}{\delta u(x)}\mathcal{U}(t,u),f(u)\big{)}\\ &\hskip 170.71652pt-\beta(\nabla\frac{\delta}{\delta u(x)}\mathcal{U}(t,u),A(u)\nabla u)\Big{]}dx+\mathcal{F}(u)=0,\end{split}

where 𝒰:[0,1]×L2(Ω)\mathcal{U}\colon[0,1]\times L^{2}(\Omega)\rightarrow\mathbb{R} is a value functional.

Proof of Proposition 9 and Proposition 10.

The proof follows from the definition of Hamiltonian dynamics. We can check it directly by using the first variation operators. In other words,

δδΦ𝒢(u,Φ)=(A(u)G′′(u)1Φ)f(u)+β(A(u)u),\frac{\delta}{\delta\Phi}\mathcal{H}_{\mathcal{G}}(u,\Phi)=-\nabla\cdot(A(u)G^{\prime\prime}(u)^{-1}\nabla\Phi)-\nabla\cdot f(u)+\beta\nabla\cdot(A(u)\nabla u),

and

δδu𝒢(u,Φ)=12(Φ,(A(u)G′′(u)1)Φ)+(Φ,f(u))+δδu(u)β(Φ,A(u)u)+β(A(u)Φ).\begin{split}\frac{\delta}{\delta u}\mathcal{H}_{\mathcal{G}}(u,\Phi)=&\frac{1}{2}(\nabla\Phi,(A(u)G^{\prime\prime}(u)^{-1})^{\prime}\nabla\Phi)+(\nabla\Phi,f^{\prime}(u))+\frac{\delta}{\delta u}\mathcal{F}(u)\\ &-\beta(\nabla\Phi,A^{\prime}(u)\nabla u)+\beta\nabla\cdot(A(u)\nabla\Phi).\end{split}

Clearly, 𝒢(u,Φ)\mathcal{H}_{\mathcal{G}}(u,\Phi) is conserved since

ddt𝒢(u,Φ)=Ω(δδu𝒢tu+δδΦ𝒢tΦ)𝑑x=Ω(δδu𝒢δδΦ𝒢δδΦ𝒢δδu𝒢)𝑑x=0.\begin{split}\frac{d}{dt}\mathcal{H}_{\mathcal{G}}(u,\Phi)=&\int_{\Omega}\Big{(}\frac{\delta}{\delta u}\mathcal{H}_{\mathcal{G}}\cdot\partial_{t}u+\frac{\delta}{\delta\Phi}\mathcal{H}_{\mathcal{G}}\cdot\partial_{t}\Phi\Big{)}dx\\ =&\int_{\Omega}\Big{(}\frac{\delta}{\delta u}\mathcal{H}_{\mathcal{G}}\cdot\frac{\delta}{\delta\Phi}\mathcal{H}_{\mathcal{G}}-\frac{\delta}{\delta\Phi}\mathcal{H}_{\mathcal{G}}\cdot\frac{\delta}{\delta u}\mathcal{H}_{\mathcal{G}}\Big{)}dx\\ =&0.\end{split}

This finishes the proof of Proposition 9.

We next use the Hamiltonian functional to formulate the Hamilton-Jacobi equation in (,𝐠)(\mathcal{M},\mathbf{g}). Define a functional by 𝒰:+×\mathcal{U}\colon\mathbb{R}_{+}\times\mathcal{M}\rightarrow\mathbb{R}. Let Φ(t,x)=δδu(x)𝒰(t,u)\Phi(t,x)=\frac{\delta}{\delta u(x)}\mathcal{U}(t,u) in Hamiltonian flow (13). We obtain the Hamilton-Jacobi equation in metric space (,𝐠)(\mathcal{M},\mathbf{g}), where

t𝒰(t,u)+𝒢(u,δδu𝒰(t,u))=0.\partial_{t}\mathcal{U}(t,u)+\mathcal{H}_{\mathcal{G}}(u,\frac{\delta}{\delta u}\mathcal{U}(t,u))=0.

This finishes the proof of Proposition 10. ∎

Remark 6 (Fisher information regularizations).

We remark that the Fisher information functional is also connected with control problems of conservation laws. See similar studies in [21]. We leave the study about information functional regularizations of conservation laws in a sequential work.

4.1. Examples

In this subsection, we list several examples of control problems for scalar conservation laws.

Example 1 (Controlling heat equations).

Consider the heat equation

tu=βΔu,\partial_{t}u=\beta\Delta u,

where u:[0,+)×Ω+u\colon[0,+\infty)\times\Omega\rightarrow\mathbb{R}_{+} is the probability density function. It satisfies PDE (6), where

u+1,f(u)=0,A(u)=𝕀.u\in\mathbb{R}_{+}^{1},\quad f(u)=0,\quad A(u)=\mathbb{I}.

In this case, a function pair (G,Ψ)(G,\Psi), where GG is a convex function and Ψ=0\Psi=0, satisfies the entropy-entropy flux condition. In other words,

Ψ(u)=f(u)G(u)=0.\Psi^{\prime}(u)=f^{\prime}(u)G^{\prime}(u)=0.

In particular, if G(u)=uloguuG(u)=u\log u-u, then the variational problem (12) satisfies

infu,v,u101[Ω12|v|2u𝑑x(u)]𝑑t+(u1),\inf_{u,v,u_{1}}~{}\int_{0}^{1}\Big{[}\int_{\Omega}\frac{1}{2}|v|^{2}udx-\mathcal{F}(u)\Big{]}dt+\mathcal{H}(u_{1}),

where the infimum is taken among variables u,v,u1u,v,u_{1} satisfying

tu+(uv)=βΔu,u(0,x)=u0(x).\partial_{t}u+\nabla\cdot(uv)=\beta\Delta u,\quad u(0,x)=u_{0}(x).

Here the minimizer system is given below. There exists a function Φ\Phi, such that v=Φv=\nabla\Phi and

{tu+(uΦ)=βΔu,tΦ+12Φ2+δδu(u)=βΔΦ.\left\{\begin{aligned} &\partial_{t}u+\nabla\cdot(u\nabla\Phi)=\beta\Delta u,\\ &\partial_{t}\Phi+\frac{1}{2}\|\nabla\Phi\|^{2}+\frac{\delta}{\delta u}\mathcal{F}(u)=-\beta\Delta\Phi.\end{aligned}\right.

The above optimal control problem and critical point system has been widely studied in the optimal transport (=0\mathcal{F}=0, β=0\beta=0), Schrödinger bridge problems (=0\mathcal{F}=0, β>0\beta>0) and potential mean field games. There is a Hamiltonian formalism for the PDE system (u,Φ)(u,\Phi). In other words,

tu=δδΦ𝒢(u,Φ),tΦ=δδu𝒢(u,Φ),\partial_{t}u=\frac{\delta}{\delta\Phi}\mathcal{H}_{\mathcal{G}}(u,\Phi),\quad\partial_{t}\Phi=-\frac{\delta}{\delta u}\mathcal{H}_{\mathcal{G}}(u,\Phi),

where the Hamiltonian functional satisfies

𝒢(u,Φ)=Ω[12Φ2uβ(Φ,u)]𝑑x+(u).\mathcal{H}_{\mathcal{G}}(u,\Phi)=\int_{\Omega}\Big{[}\frac{1}{2}\|\nabla\Phi\|^{2}u-\beta(\nabla\Phi,\nabla u)\Big{]}dx+\mathcal{F}(u).
Example 2 (Controlling scalar conservation laws).

Consider

tu+f(u)=βΔu,\partial_{t}u+\nabla\cdot f(u)=\beta\Delta u,

where u:[0,+)×Ωu\colon[0,+\infty)\times\Omega\rightarrow\mathbb{R} and A=𝕀A=\mathbb{I}.

  • (i)

    Let G(u)=uloguuG(u)=u\log u-u. Then variational problem (12) satisfies

    infu,v,u101[Ω12|v|2u𝑑x(u)]𝑑t+(u1),\inf_{u,v,u_{1}}~{}\int_{0}^{1}\Big{[}\int_{\Omega}\frac{1}{2}|v|^{2}udx-\mathcal{F}(u)\Big{]}dt+\mathcal{H}(u_{1}),

    where the infimum is taken among variables u,v,u1u,v,u_{1} satisfying

    tu+f(u)+(uv)=βΔu,u(0,x)=u0(x).\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot(uv)=\beta\Delta u,\quad u(0,x)=u_{0}(x).

    Here the minimizer system is given below. There exists a function Φ\Phi, such that v=Φv=\nabla\Phi and

    {tu+f(u)+(uΦ)=βΔu,tΦ+(Φ,f(u))+12Φ2+δδu(u)=βΔΦ.\left\{\begin{aligned} &\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot(u\nabla\Phi)=\beta\Delta u,\\ &\partial_{t}\Phi+(\nabla\Phi,f^{\prime}(u))+\frac{1}{2}\|\nabla\Phi\|^{2}+\frac{\delta}{\delta u}\mathcal{F}(u)=-\beta\Delta\Phi.\end{aligned}\right.

    I.e.,

    tu=δδΦ𝒢(u,Φ),tΦ=δδu𝒢(u,Φ),\partial_{t}u=\frac{\delta}{\delta\Phi}\mathcal{H}_{\mathcal{G}}(u,\Phi),\quad\partial_{t}\Phi=-\frac{\delta}{\delta u}\mathcal{H}_{\mathcal{G}}(u,\Phi),

    where the Hamiltonian functional satisfies

    𝒢(u,Φ)=Ω[12Φ2u+(Φ,f(u))β(Φ,u)]𝑑x+(u).\mathcal{H}_{\mathcal{G}}(u,\Phi)=\int_{\Omega}\Big{[}\frac{1}{2}\|\nabla\Phi\|^{2}u+(\nabla\Phi,f(u))-\beta(\nabla\Phi,\nabla u)\Big{]}dx+\mathcal{F}(u).
  • (ii)

    Let G(u)=u22G(u)=\frac{u^{2}}{2}. Then variational problem (12) satisfies

    infu,v,u101[Ω12|v|2𝑑x(u)]𝑑t+(u1),\inf_{u,v,u_{1}}~{}\int_{0}^{1}\Big{[}\int_{\Omega}\frac{1}{2}|v|^{2}dx-\mathcal{F}(u)\Big{]}dt+\mathcal{H}(u_{1}),

    where the infimum is taken among variables u,v,u1u,v,u_{1} satisfying

    tu+f(u)+v=βΔu,u(0,x)=u0(x).\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot v=\beta\Delta u,\quad u(0,x)=u_{0}(x).

    Here the minimizer system is given below. There exists a function Φ\Phi, such that v=Φv=\nabla\Phi and

    {tu+f(u)+(Φ)=βΔu,tΦ+(Φ,f(u))+δδu(u)=βΔΦ.\left\{\begin{aligned} &\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot(\nabla\Phi)=\beta\Delta u,\\ &\partial_{t}\Phi+(\nabla\Phi,f^{\prime}(u))+\frac{\delta}{\delta u}\mathcal{F}(u)=-\beta\Delta\Phi.\end{aligned}\right.

    I.e.,

    tu=δδΦ𝒢(u,Φ),tΦ=δδu𝒢(u,Φ),\partial_{t}u=\frac{\delta}{\delta\Phi}\mathcal{H}_{\mathcal{G}}(u,\Phi),\quad\partial_{t}\Phi=-\frac{\delta}{\delta u}\mathcal{H}_{\mathcal{G}}(u,\Phi),

    where the Hamiltonian functional satisfies

    𝒢(u,Φ)=Ω[12Φ2+(Φ,f(u))β(Φ,u)]𝑑x+(u).\mathcal{H}_{\mathcal{G}}(u,\Phi)=\int_{\Omega}\Big{[}\frac{1}{2}\|\nabla\Phi\|^{2}+(\nabla\Phi,f(u))-\beta(\nabla\Phi,\nabla u)\Big{]}dx+\mathcal{F}(u).

5. Numerical methods and examples

In this section, we first review classical primal-dual hybrid gradient algorithms (PDHG) and their extensions. We next apply this approach to solve the variational problem defined in equation (12) subject to the constraint involving conservation laws. We design a finite difference scheme to discretize conservation laws and solve the variational problem on grids. Several numerical examples, including Burgers’ equation and traffic flow, are presented.

5.1. PDHG algorithm and its extension

The PDHG algorithm [10] solves the following constrained convex optimization problem

minzh(Kz)+g(z),\displaystyle\min_{z}h(Kz)+g(z),

where 𝒵\mathcal{Z} is a finite or infinite dimensional Hilbert space, hh and gg are convex functions and K:𝒵K:\mathcal{Z}\rightarrow\mathcal{H} is a linear operator between Hilbert spaces. This problem can be rewritten in the saddle-point problem form

minzmaxpKz,pL2+g(z)h(p),\displaystyle\min_{z}\max_{p}\langle Kz,p\rangle_{L^{2}}+g(z)-h^{*}(p),

where h(p)=supzKz,pL2h(z)h^{*}(p)=\sup_{z}\langle Kz,p\rangle_{L^{2}}-h(z), which is the convex conjugate of hh. The algorithm updates zz, pp by taking proximal gradient descent, proximal gradient ascent steps alternatively. At the nn-th iteration, the algorithm updates as follows

zn+1\displaystyle z^{n+1} =argminzKz,p¯nL2+g(z)+12τzznL22,\displaystyle=\text{arg}\min_{z}\langle Kz,\bar{p}^{n}\rangle_{L^{2}}+g(z)+\frac{1}{2\tau}\|z-z^{n}\|^{2}_{L^{2}},
pn+1\displaystyle p^{n+1} =argmaxpKzn+1,pL2h(p)12σppnL22,\displaystyle=\text{arg}\max_{p}\langle Kz^{n+1},p\rangle_{L^{2}}-h^{*}(p)-\frac{1}{2\sigma}\|p-p^{n}\|^{2}_{L^{2}},
p¯n+1\displaystyle\bar{p}^{n+1} =2pn+1pn.\displaystyle=2p^{n+1}-p^{n}.

Here τ\tau, σ\sigma are stepsizes, which have to satisfy στKTK<1\sigma\tau\|K^{T}K\|<1 in order to guarantee the convergence of the algorithm. When the operator KK is nonlinear, we use the extension of PDHG algorithm [13]. The idea is to use the linear approximation of KK:

K(z)K(z¯)+K(z¯)(zz¯).\displaystyle K(z)\approx K(\bar{z})+\nabla K(\bar{z})(z-\bar{z}).

Here, the extension of the PDHG scheme is as follows

zn+1\displaystyle z^{n+1} =argminzz,[K(zn)]Tp¯nL2+g(x)+12τzznL22,\displaystyle=\text{arg}\min_{z}\langle z,[\nabla K(z^{n})]^{T}\bar{p}^{n}\rangle_{L^{2}}+g(x)+\frac{1}{2\tau}\|z-z^{n}\|^{2}_{L^{2}}, (16)
pn+1\displaystyle p^{n+1} =argmaxpK(zn+1),pL2h(p)12σppnL22,\displaystyle=\text{arg}\max_{p}\langle K(z^{n+1}),p\rangle_{L^{2}}-h^{*}(p)-\frac{1}{2\sigma}\|p-p^{n}\|^{2}_{L^{2}},
p¯n+1\displaystyle\bar{p}^{n+1} =2pn+1pn.\displaystyle=2p^{n+1}-p^{n}.

When KK is some unbounded linear operator, for instance K=K=\nabla, the operator norm K\|K\| can increase when we refine the grid size. Consequently, the algorithm may converge slowly due to small stepsizes. We apply a generalization of PDHG, namely the General-proximal Primal-Dual Hybrid Gradient (G-prox PDHG) method from [20]. We choose a proper norm (L2,H1,L^{2},H^{1},...) for the proximal step to allow larger stepsizes.

We apply the nonlinear G-prox PDHG algorithm to solve the variational problem (12), in particular its equivalent format (14). For illustration purpose, we use the traffic flow variational problem (5) as an example and give details on the algorithm. Set

z\displaystyle z =(u,m),\displaystyle=(u,m),
p\displaystyle p =Φ,\displaystyle=\Phi,
K((u,m))\displaystyle K\left((u,m)\right) =tu+f(u)+mβΔu,\displaystyle=\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot m-\beta\Delta u,
g((u,m))\displaystyle g\left((u,m)\right) =01(Ωm22u+𝟏[0,1](u)dx(u))𝑑t,\displaystyle=\int_{0}^{1}\left(\int_{\Omega}\frac{\|m\|^{2}}{2u}+\mathbf{1}_{[0,1]}(u)dx-\mathcal{F}(u)\right)dt,
h(Kz)\displaystyle h(Kz) ={0ifKz=0+else.\displaystyle=\begin{cases}0\quad\text{if}\;Kz=0\\ +\infty\quad\text{else}\end{cases}.

We rewrite the varational problem as follows:

infu,msupΦ(u,m,Φ),withu(0,x)=u0(x),Φ(1,x)=δδu(1,x)(u),\inf_{u,m}\sup_{\Phi}~{}~{}\mathcal{L}(u,m,\Phi),\quad\text{with}\;u(0,x)=u_{0}(x),~{}\Phi(1,x)=-\frac{\delta}{\delta u(1,x)}\mathcal{H}(u), (17)

where

(u,m,Φ)=\displaystyle{}\mathcal{L}(u,m,\Phi)= 01(Ωm22u+𝟏[0,1](u)dx(u))𝑑t\displaystyle\int_{0}^{1}\left(\int_{\Omega}\frac{\|m\|^{2}}{2u}+\mathbf{1}_{[0,1]}(u)dx-\mathcal{F}(u)\right)dt (18)
+01ΩΦ(tu+f(u)+mβΔu)𝑑x𝑑t,\displaystyle+\int_{0}^{1}\int_{\Omega}\Phi\left(\partial_{t}u+\nabla\cdot f(u)+\nabla\cdot m-\beta\Delta u\right)dxdt,

and f(u)=u(1u)f(u)=u(1-u) is the traffic flux function. We denote the indicator function by

𝟏A(x)={+xA0xA.\mathbf{1}_{A}(x)=\begin{cases}+\infty\quad x\notin A\\ 0\quad\quad\;x\in A.\end{cases}

We also choose L2L^{2} norm for updating (u,m)(u,m) and H1H^{1} norm for Φ\Phi, specifically

vL22=01Ωv2𝑑x𝑑t,vH122=vL22+tvL22.\displaystyle\|v\|^{2}_{L^{2}}=\int_{0}^{1}\int_{\Omega}v^{2}dxdt,\quad\|v\|^{2}_{H_{1}^{2}}=\|\nabla v\|^{2}_{L^{2}}+\|\partial_{t}v\|^{2}_{L^{2}}.

We summarize the algorithm to solve problem (17) as follows.

 
Algorithm 1:PDHG for the conservation law control system
 
While k<k< Maximal number of iteration
(u(k+1),m(k+1))\left(u^{(k+1)},m^{(k+1)}\right) =argminu(u,m,Φ¯(k))+12τuu(k)L22+12τmm(k)L22=\text{argmin}_{u}~{}~{}\mathcal{L}(u,m,\bar{\Phi}^{(k)})+\frac{1}{2\tau}\|u-u^{(k)}\|^{2}_{L^{2}}+\frac{1}{2\tau}\|m-m^{(k)}\|^{2}_{L^{2}};
Φ(k+1)\Phi^{(k+1)} =argmaxΦ(u(k+1),m(k+1),Φ)12σΦΦ(k)H122=\text{argmax}_{\Phi}~{}~{}\mathcal{L}(u^{(k+1)},m^{(k+1)},\Phi)-\frac{1}{2\sigma}\|\Phi-\Phi^{(k)}\|^{2}_{H_{1}^{2}};
Φ¯(k+1)\bar{\Phi}^{(k+1)} =2Φ(k+1)Φ(k)=2\Phi^{(k+1)}-\Phi^{(k)};
 

5.2. Finite Difference Discretization

In this subsection, we review basic numerical concepts of conservation law. More details can be found in [25]. Then we design the finite difference discretization for the control of conservation laws.

5.2.1. Lax–Friedrichs scheme for the conservation law

Consider a nonlinear scalar conservation law

{tu+xf(u)=βxxu,u(0,x)=u0(x).\begin{cases}\partial_{t}u+\partial_{x}f(u)=\beta\partial_{xx}u,\\ u(0,x)=u_{0}(x).\end{cases} (19)

We review the Lax–Friedrichs scheme. Denote a discretization Δt,Δx{\Delta t},{\Delta x} in time and space, ujk=u(kΔt,jΔx)u_{j}^{k}=u(k\Delta t,j\Delta x), then the update follows

ujk+1=ujkΔt2Δx(f(uj+1k)f(uj1k))+(β+cΔx)Δt(Δx)2(uj+1k2ujk+uj1k).u_{j}^{k+1}=u_{j}^{k}-\frac{\Delta t}{2\Delta x}\left(f(u_{j+1}^{k})-f(u_{j-1}^{k})\right)+(\beta+c\Delta x)\frac{\Delta t}{(\Delta x)^{2}}\left(u_{j+1}^{k}-2u_{j}^{k}+u_{j-1}^{k}\right). (20)

The last term contains diffusion coefficient β\beta from the original equation and cΔxc\Delta x as coefficient for artificial viscosity with c>0c>0. Notice that the monotone scheme gives entropy solutions. Here the definition of monotone scheme is given below.

Definition 11.

For p,qp,q\in\mathbb{N}, a scheme

ujk+1=G(ujp1k,,uj+qk)\displaystyle u_{j}^{k+1}=G(u_{j-p-1}^{k},...,u_{j+q}^{k})

is called a monotone scheme if GG is a monotonically nondecreasing function of each argument.

In order to guarantee that the scheme (20) is monotone, the following inequalities have to be satisfied:

12(β+cΔx)Δt(Δx)20,\displaystyle 1-2(\beta+c\Delta x)\frac{\Delta t}{(\Delta x)^{2}}\geq 0,
Δt2Δx|f(u)|+(β+cΔx)Δt(Δx)20.\displaystyle-\frac{\Delta t}{2\Delta x}|f^{\prime}(u)|+(\beta+c\Delta x)\frac{\Delta t}{(\Delta x)^{2}}\geq 0.

As we want the scheme works when β0\beta\rightarrow 0, the restriction on cc and space–time stepsizes can be simplified as follows:

c12|f(u)|,\displaystyle c\geq\frac{1}{2}|f^{\prime}(u)|,
(Δx)22(β+cΔx)Δt.\displaystyle{(\Delta x)^{2}}\geq 2(\beta+c\Delta x)\Delta t.

The first inequality suggests the artificial viscosity we need to add. The second one impose a strong restriction on the stepsize in time when β>0\beta>0.

5.2.2. Discreitization of the control problem

We consider the control problem of scalar conservation law defined in [0,b]×[0,1][0,b]\times[0,1], where bb is a given constant. We apply the periodic boundary condition on the spatial domain. Given Nx,Nt>0N_{x},N_{t}>0, we have Δx=bNx\Delta x=\frac{b}{N_{x}}, Δt=1Nt\Delta t=\frac{1}{N_{t}}. For xi=iΔx,tl=lΔtx_{i}=i\Delta x,t_{l}=l\Delta t, define

uil=u(tl,xi)\displaystyle u_{i}^{l}=u(t_{l},x_{i}) 1iNx,0lNt,\displaystyle\quad 1\leq i\leq N_{x},0\leq l\leq N_{t},
m1,il=(mx1(tl,xi))+\displaystyle m_{1,i}^{l}=\left(m_{x_{1}}(t_{l},x_{i})\right)^{+} 1iNx,0lNt1,\displaystyle\quad 1\leq i\leq N_{x},0\leq l\leq N_{t}-1,
m2,il=(mx1(tl,xi))\displaystyle m_{2,i}^{l}=-\left(m_{x_{1}}(t_{l},x_{i})\right)^{-} 1iNx,0lNt1,\displaystyle\quad 1\leq i\leq N_{x},0\leq l\leq N_{t}-1,
Φil=Φ(tl,xi)\displaystyle\Phi_{i}^{l}=\Phi(t_{l},x_{i}) 1iNx,0lNt,\displaystyle\quad 1\leq i\leq N_{x},0\leq l\leq N_{t},
ΦiNt=δδu(1,xi)(uiNt)\displaystyle\Phi_{i}^{Nt}=-\frac{\delta}{\delta u(1,x_{i})}\mathcal{H}(u_{i}^{N_{t}}) 1iNx,\displaystyle\quad 1\leq i\leq N_{x},

where u+:=max(u,0)u^{+}:=\max(u,0) and u=u+uu^{-}=u^{+}-u. Note here m1,il+,m2,ilm_{1,i}^{l}\in\mathbb{R}_{+},m_{2,i}^{l}\in\mathbb{R}_{-}. Denote

(Du)i:=ui+1uiΔx\displaystyle\left(Du\right)_{i}:=\frac{u_{i+1}-u_{i}}{\Delta x}
[Du]i:=((Du)i,(Du)i1)\displaystyle[Du]_{i}:=\left(\left(Du\right)_{i},\left(Du\right)_{i-1}\right)
[Du]^i=((Du)i+,(Du)i1)\displaystyle\widehat{[Du]}_{i}=\left(\left(Du\right)^{+}_{i},-\left(Du\right)^{-}_{i-1}\right)
Lap(u)i=ui+12ui+ui1(Δx)2\displaystyle Lap(u)_{i}=\frac{u_{i+1}-2u_{i}+u_{i-1}}{(\Delta x)^{2}}

The first conservation law equation adapted from the Lax–Friedrichs scheme is as follows:

1Δt(uil+1uil)+12Δx(f(ui+1l+1)f(ui1l+1))+(Dm)1,i1l+(Dm)2,il=(β+cΔx)Lap(u)il+1,\frac{1}{\Delta t}\left(u_{i}^{l+1}-u_{i}^{l}\right)+\frac{1}{2\Delta x}\left(f(u_{i+1}^{l+1})-f(u_{i-1}^{l+1})\right)+\left(Dm\right)^{l}_{1,i-1}+\left(Dm\right)^{l}_{2,i}=(\beta+c\Delta x)Lap(u)^{l+1}_{i}, (21)

where 1iNx1\leq i\leq N_{x}, 0lNt10\leq l\leq N_{t}-1. We choose c12maxu|f(u)|c\geq\frac{1}{2}\max_{u}|f^{\prime}(u)|. Unlike the Lax–Friedrichs scheme in explicit form, we use an implicit discretization in time to encode the feedback control structure. Another benefit of the implicit scheme is that it allows a larger mesh size in time.

Following the discretization of the conservation law, the discrete saddle point problem has the following form:

minu,mmaxΦL(u,m,Φ),\displaystyle\min_{u,m}\max_{\Phi}L(u,m,\Phi),

where

L(u,m,Φ)=\displaystyle L(u,m,\Phi)= ΔxΔt1iNx1lNt(m1,il1)2+(m2,il1)22uilΔt1lNt(ul)\displaystyle\Delta x\Delta t\sum_{\begin{subarray}{c}1\leq i\leq N_{x}\\ 1\leq l\leq N_{t}\end{subarray}}\frac{(m_{1,i}^{l-1})^{2}+(m_{2,i}^{l-1})^{2}}{2u_{i}^{l}}-\Delta t\sum_{1\leq l\leq N_{t}}\mathcal{F}(u^{l})
+Δx1iNx(uiNt)+1iNx1lNt𝟏[0,1](uil)\displaystyle+\Delta x\sum_{1\leq i\leq N_{x}}\mathcal{H}(u_{i}^{N_{t}})+\sum_{\begin{subarray}{c}1\leq i\leq N_{x}\\ 1\leq l\leq N_{t}\end{subarray}}\mathbf{1}_{[0,1]}(u_{i}^{l})
+ΔxΔt1iNx0lNt1Φil(1Δt(uil+1uil)+12Δx(f(ui+1l+1)f(ui1l+1))\displaystyle+\Delta x\Delta t\sum_{\begin{subarray}{c}1\leq i\leq N_{x}\\ 0\leq l\leq N_{t}-1\end{subarray}}\Phi_{i}^{l}\Big{(}\frac{1}{\Delta t}(u_{i}^{l+1}-u_{i}^{l})+\frac{1}{2\Delta x}(f(u_{i+1}^{l+1})-f(u_{i-1}^{l+1}))
+(Dm)1,i1l+(Dm)2,il(β+cΔx)Lap(u)il+1).\displaystyle\hskip 71.13188pt+(Dm)^{l}_{1,i-1}+(Dm)^{l}_{2,i}-(\beta+c\Delta x)Lap(u)^{l+1}_{i}\Big{)}.

By taking the first order derivative of uilu_{i}^{l}, we automatically get the implicit finite difference scheme for the dual equation of Φ\Phi that is backward in time. The positive and negative parts of (Dϕ)il(D\phi)_{i}^{l} are split, which help enhance the monotonicity of the discrete Hamiltonian.

1Δt(Φil+1Φil)+(Φi+1lΦi1l)2Δx(f(uil))+12[DΦ^]il2+δ(uil)δu=(β+cΔx)Lap(Φ)il,\frac{1}{\Delta t}\left(\Phi_{i}^{l+1}-\Phi_{i}^{l}\right)+\frac{(\Phi_{i+1}^{l}-\Phi_{i-1}^{l})}{2\Delta x}\left(f^{\prime}(u_{i}^{l})\right)+\frac{1}{2}\|[\widehat{D\Phi}]_{i}^{l}\|^{2}+\frac{\delta\mathcal{F}(u_{i}^{l})}{\delta u}=-(\beta+c\Delta x)Lap(\Phi)^{l}_{i}, (22)

for 1iNx1\leq i\leq N_{x}, 1lNt1\leq l\leq N_{t}.

We remark that the above discretizations, when f(u)=0f(u)=0, reduce to the finite difference scheme for the mean-field game system proposed in [7]. The discrete form of the Hamiltonian functional (15) at t=tlt=t_{l} takes the form

H𝒢(u,Φ)=1iNx(12[DΦ^]il2uil+(Φi+1lΦi1l)2Δx(f(uil))+(β+cΔx)uilLap(Φ)il)(ul).\displaystyle H_{\mathcal{G}}(u,\Phi)=\sum_{\begin{subarray}{c}1\end{subarray}\leq i\leq N_{x}}\left(\frac{1}{2}\|[\widehat{D\Phi}]_{i}^{l}\|^{2}u_{i}^{l}+\frac{(\Phi_{i+1}^{l}-\Phi_{i-1}^{l})}{2\Delta x}\left(f(u_{i}^{l})\right)+(\beta+c\Delta x)u_{i}^{l}Lap(\Phi)^{l}_{i}\right)-\mathcal{F}(u^{l}).

We shall verify the conservation of Hamiltonian functional with numerical examples.

5.2.3. Solve conservation laws via primal-dual algorithms

When =0\mathcal{F}=0, =c\mathcal{H}=c for some constant cc, the variational problem (17) becomes classical conservation laws with initial data. In this case, no control will be enforced on the density function uu. Therefore, the density movement is only determined by the flux term (f(u))x.\left(f(u)\right)_{x}. This means that the problem is reduced to initial value conservation laws. In this scenario, our approach proposed in Algorithm 1 provides an alternative way to solve the nonlinear conservation law with implicit discretization in time. In the implementation, we observed that this method allows a larger mesh size in time, thanks to the primal–dual variational structure. The method is also highly parallelizable as the proximal gradient descent step for (u,m)(u,m) is point-wise operation for each (l,i)(l,i). We demonstrate this part with two examples in the next section.

5.3. Numerical examples

In this subsection, we present numerical examples for control of conservation laws in one dimensional space.

5.4. Example 1.

We consider the Burgers’ equation on (t,x)[0,1]×[0,4](t,x)\in[0,1]\times[0,4].

tu+xf(u)=0,u(0,x)={12x30else,\displaystyle\partial_{t}u+\partial_{x}f(u)=0,\quad u(0,x)=\begin{cases}1\quad&2\leq x\leq 3\\ 0\quad\quad&\text{else}\end{cases},

where f(u)=12u2,=0,=0,β=0,k=0.5f(u)=\frac{1}{2}u^{2},\mathcal{F}=0,\mathcal{H}=0,\beta=0,k=0.5. The entropy solution to this problem at t=1t=1 satisfies

u(1,x)={13x3.5(x2)2<x30else.u(1,x)=\begin{cases}1\quad&3\leq x\leq 3.5\\ (x-2)\quad&2<x\leq 3\\ 0\quad\quad&\text{else}\end{cases}.

In the following examples, we use the same spatial time domain with Nt=50,Nx=100N_{t}=50,N_{x}=100 for the finite difference scheme. We solve the Burgers’ equation using two approaches: solve the control problem; use the forward Lax–Friedrichs scheme. Figure 2 shows the solutions to the Burgers’ equation. We can see that both numerical solutions are consistent with the exact entropy solution despite some numerical diffusions. From two plots on the right, there are clear formations of rarefaction wave and shock. We have verified that the shock travels at speed v=12v=\frac{1}{2}.

Refer to caption
Figure 2. Numerical results for the Burgers’ equation. Left: a comparison with the exact solution at t=1t=1; middle: the numerical solution via solving the control problem; right: the numerical solution to the conservation law using Lax–Friedrichs scheme.

5.5. Example 2.

We consider the traffic flow equation

tu+xf(u)=0,u(0,x)={0.8,1x20else,\displaystyle\partial_{t}u+\partial_{x}f(u)=0,\quad u(0,x)=\begin{cases}0.8,\quad&1\leq x\leq 2\\ 0\quad\quad&\text{else}\end{cases},

where f(u)=12u(1u),=0,=0,β=0,k=0.5f(u)=\frac{1}{2}u(1-u),\mathcal{F}=0,\mathcal{H}=0,\beta=0,k=0.5.

Refer to caption
Figure 3. Numerical results for the traffic flow equation. Left: a comparison with the exact solution at t=1t=1; middle: the numerical solution via solving the control problem; right: the numerical solution to the conservation law using Lax–Friedrichs scheme.

The entropy solution to this problem at time t=1t=1 is

u(1,x)={11.2x1.412(3x)1.4<x30else.u(1,x)=\begin{cases}1\quad&1.2\leq x\leq 1.4\\ \frac{1}{2}(3-x)\quad&1.4<x\leq 3\\ 0\quad\quad&\text{else}\end{cases}.

We can see from Figure 3 (left) that the solution to the control problem gives the same solution to the conservation law, which matches the entropy solution. In the two plots on the right, we observe that both shocks and rarefaction waves are formed. This traffic flow example describes a group of cars u0u_{0} waiting at the traffic light at x=2x=2. At the time t=0t=0, the red light turns green. But this group of cars doesn’t move at a uniform constant speed. Instead, the car, whose originally position at time t=0t=0 is closer to the red light (x=0x=0), moves faster. The density at position x<0x<0 doesn’t change until time t=12(3x)t=\frac{1}{2}(3-x).

5.6. Example 3

We again consider the traffic flow equation with f(u)=u(1u),β=0.1,=0,c=0.5f(u)=u(1-u),\beta=0.1,\mathcal{F}=0,c=0.5. The final cost functional (u(1,))=μΩu(1,x)log(u(1,x)u1)𝑑x,μ>0\mathcal{H}(u(1,\cdot))=\mu\int_{\Omega}u(1,x)\log(\frac{u(1,x)}{u_{1}})dx,\mu>0. We set μ=1\mu=1. In this case, the density u(1,)u(1,\cdot) will tend to form a ‘similar’ distribution as u1u_{1}. When Ωu1𝑑x=Ωu0𝑑x\int_{\Omega}u_{1}dx=\int_{\Omega}u_{0}dx and μ+\mu\rightarrow+\infty, this final cost functional is equivalent to imposes the constraint that u(1,)=u1u(1,\cdot)=u_{1}. We also compare the result from the control of conservation law with a mean-field game problem, i.e., f=0f=0:

{tu+(uΦ)=βΔu,tΦ+12Φ2+δδu(u)=βΔΦ,u(0,x)=u0(x),Φ(1,x)=δδu1(u(1,)).\begin{cases}\partial_{t}u+\nabla\cdot(u\nabla\Phi)=\beta\Delta u,\\ \partial_{t}\Phi+\frac{1}{2}\|\nabla\Phi\|^{2}+\frac{\delta}{\delta u}\mathcal{F}(u)=-\beta\Delta\Phi,\\ u(0,x)=u_{0}(x),~{}\Phi(1,x)=-\frac{\delta}{\delta u_{1}}\mathcal{H}(u(1,\cdot)).\end{cases} (23)

As shown in Figure 4, we set

u0\displaystyle u_{0} =0.001+0.9e10(x2)2,\displaystyle=0.001+0.9e^{-10(x-2)^{2}},
u1\displaystyle u_{1} =0.001+0.45e10(x1)2+0.45e10(x3)2.\displaystyle=0.001+0.45e^{-10(x-1)^{2}}+0.45e^{-10(x-3)^{2}}.

The solutions are presented in Figure 5. We can see that in the mean-field game setup, there is an even split of the density. While for the control of conservation law, more density travels towards the location of the right-side Gaussian distribution. This demonstrates the difference between solutions in mean-field games and the ones in control of conservation laws.

Refer to caption
Figure 4. From left to right: initial configurations of u0=0.001+0.9e10(x2)2u_{0}=0.001+0.9e^{-10(x-2)^{2}}, u1=0.001+0.45e10(x1)2+0.45e10(x3)2u_{1}=0.001+0.45e^{-10(x-1)^{2}}+0.45e^{-10(x-3)^{2}}, solution u(1,x)u(1,x) for the control of conservation law, solution u(1,x)u(1,x) for the mean-field game problem.
Refer to caption
Figure 5. Left: solution u(t,x)u(t,x) for the problem of controlling the conservation law; right: solution u(t,x)u(t,x) for the mean-field game problem.

5.7. Example 4

Consider the traffic flow equation with f(u)=u(1u),β=103,=αΩulog(u)𝑑x,α0f(u)=u(1-u),\beta=10^{-3},\mathcal{F}=-\alpha\int_{\Omega}u\log(u)dx,\alpha\geq 0. The final cost functional (u(1,))=Ωu(1,x)g(x)𝑑x\mathcal{H}(u(1,\cdot))=\int_{\Omega}u(1,x)g(x)dx. The initial density and final cost function are as follows

u0(x)={0.40.5x1.5103else,g(x)=0.1sin(2πx).\displaystyle u_{0}(x)=\begin{cases}0.4\quad&0.5\leq x\leq 1.5\\ 10^{-3}\quad&\text{else}\\ \end{cases},\quad\quad g(x)=-0.1\sin(2\pi x).

The term =αΩulog(u)𝑑x,α0\mathcal{F}=-\alpha\int_{\Omega}u\log(u)dx,\alpha\geq 0 penalizes the density for getting too concentrate. Figure 6 shows the effect of the term αulog(u)\alpha u\log(u), where the density is more spreading in space for the case α=1\alpha=1 than the case α=0.5\alpha=0.5 and α=0\alpha=0. We can also see from the uu profile at the terminal time. In Figure 7 (middle), α=0\alpha=0 case has u(1,)u(1,\cdot) has the most concentrated densities than others. We also numerically verify that the Hamiltonian functional is preserved over time in this example. In Figure 7 (right), the numerical Hamiltonian H𝒢(u,Φ){H}_{\mathcal{G}}(u,\Phi) is preserved with error of order of O(Δx)O(\Delta x).

Refer to caption
Figure 6. Solution u(t,x)u(t,x) for the problem of controlling the conservation law. From left to right: α=0,0.5,1\alpha=0,0.5,1.
Refer to caption
Figure 7. Left: boundary conditions for the control problems u0u_{0}. Middle: solution u(1,x)u(1,x) for the problem of controlling the conservation law. Right: the numerical Hamiltonian H𝒢(u,Φ){H}_{\mathcal{G}}(u,\Phi).

References

  • [1] L. Ambrosio, N. Gigli and G. Savaré. Gradient Flows in Metric Spaces and in the Space of Probability Measures. Lectures in Mathematics ETH Zurich, Birkhauser Verlag, Basel, 2nd ed. 2008.
  • [2] A. Arnold, P. Markowich, G. Toscani, and A. Unterreiter. On convex Sobolev inequalities and the rate of convergence to equilibrium for Fokker-Planck type equations. Communications in Partial Differential Equations, 26 (1), 2000.
  • [3] J. Backhoff, G. Conforti, I. Gentil, and C. Leonard. The mean field Schrödinger problem: ergodic behavior, entropy estimates and functional inequalities. Probability Theory and Related Fields, 2020.
  • [4] J.D. Benamou and Y. Brenier. A Computational Fluid Mechanics Solution to the Monge-Kantorovich Mass Transfer Problem. Numerische Mathematik, 84(3):3750–393, 2000.
  • [5] Y. Brenier. The Initial Value Problem for the Euler Equations of Incompressible Fluids Viewed as a Concave Maximization Problem. Commun. Math. Phys. 364, 579–-605, 2018.
  • [6] Y. Brenier and I. Moyano. Relaxed solutions for incompressible inviscid flows: A variational and gravitational approximation to the initial value problem. arXiv:2104.14803, 2021.
  • [7] L. Briceño-Arias, D. Kalise, Z. Kobeissi, M. Laurière, A. M. González, and F. J. Silva. On the implementation of a primal-dual algorithm for second order time-dependent mean field games with local couplings. ESAIM: Proceedings and Surveys, 65:330–348, 2019.
  • [8] P. Cardaliaguet, F. Delarue, J. Lasry, and P. Lions. The master equation and the convergence problem in mean field games. arXiv:1509.02505, 2015.
  • [9] J. A. Carrillo, S. Lisini, G. Savare and D. Slepcev. Nonlinear mobility continuity equations and generalized displacement convexity. Journal of Functional Analysis 258(4):1273-1309, 2009.
  • [10] A. Chambolle and T. Pock. A first-order primal-dual algorithm for convex problems with applications to imaging. J. Math. Imaging Vision., 40(1):120-145, 2011.
  • [11] Y. Chen, T. Georgiou, and M. Pavon. On the Relation Between Optimal Transport and Schrodinger Bridges: A Stochastic Control Viewpoint. J Optim Theory Appl, 169:671Ð691, 2016.
  • [12] A. Chiarini, G. Conforti and L. Tamanini. Schrodinger Problem for Lattice Gases: A Heuristic Point of View. Geometric Science of Information, 2021.
  • [13] C. Clason and T. Valkonen. Primal-dual extragradient methods for nonlinear nonsmooth pde-constrained optimization. SIAM Journal on Optimization, 27(3):1314–1339, 2017.
  • [14] J. Dolbeault, B. Nazaret, and G. Savare A new class of transport distances. Calculus of Variations and Partial Differential Equations, (2):193–231, 2009.
  • [15] M.D. Donsker, and S.R.S. Varadhan. Large deviations from a hydrodynamic scaling limit. Commun. Pure Appl. Math, 42(3), 243–270, 1989.
  • [16] L.C. Evans. A Survey of Entropy Methods for Partial Differential Equations. Bull. Amer. Math. Soc., 41, 409-438, 2004.
  • [17] W. Gangbo, W. Li, S. Osher, and M. Puthawala. Unnormalized Optimal Transport. Journal of Computational Physics, 2019.
  • [18] W. Gangbo, T. Nguyen, and A. Tudorascu. Hamilton-Jacobi equations in the Wasserstein space. Meth. Appl. Anal. Vol. 15 no 2, 155-184, 2008.
  • [19] M. Huang, R.P. Malhame, and P. Caines. Large population stochastic dynamic games: closed-loop Mckean-Vlasov systems and the Nash certainty equivalence principle. Communications in Information &\& Systems, 6(3):221-252, 2006.
  • [20] M. Jacobs, F. Léger, W. Li, and S. Osher. Solving Large-Scale Optimization Problems with a Convergence Rate Independent of Grid Size. SIAM J. Numer. Anal., 57(3), 1100–1123, 2019.
  • [21] J. M. Lasry, and P. L. Lions, Mean field games. Japanese Journal of Mathematics, 2, 229-260, 2007.
  • [22] P. Lax. Shock Waves and Entropy. Contributions to Nonlinear Functional Analysis, Proceedings of a Symposium Conducted by the Mathematics Research Center, the University of Wisconsin–Madison, April 12–14, 1971.
  • [23] F. Leger and W. Li. Hopf–Cole transformation via generalized Schrödinger bridge problem. Journal of Differential Equations, Volume 274, 788–827, 2021.
  • [24] C. Leonard. A survey of the Schrodinger problem and some of its connections with optimal transport. DCDS, 34(4):1533–1574, 2014.
  • [25] R. J. LeVeque and R. J. Leveque. Numerical methods for conservation laws, volume 132. Springer, 1992.
  • [26] W. Li. Transport information geometry: Riemannian calculus on probability simplex. Information Geometry, 2021.
  • [27] W. Li. Hessian metric via transport information geometry. Journal of Mathematical Physics, 62, 033301, 2021.
  • [28] W. Li, and S. Osher. Constrained dynamical optimal transport and its Lagrangian formulation. arXiv:1807.00937, 2018.
  • [29] W. Li, and L. Ying. Hessian transport gradient flows. Research in the Mathematical Sciences, 6, 34, 2019.
  • [30] S. Lisini, D. Matthes, and G. Savaréa. Cahn–Hilliard and thin film equations with nonlinear mobility as gradient flows in weighted-Wasserstein metrics. Journal of differential equations, (253): 814-850, 2012.
  • [31] A. Mielke. A gradient structure for reaction diffusion systems and for energy-drift-diffusion systems. Nonlinearity, 4(4):1329, 2011.
  • [32] A. Mielke. Free energy, free entropy, and a gradient structure for thermoplasticity. In “K. Weinberg and A. Pandolfi (eds): Innovative Numerical Approaches for Multi-Field and Multi-Scale Problems”. Lecture Notes in Appl. Comp. Mechanics Vol. 81, pages 135-160, Springer 2016.
  • [33] A. Mielke, D. R. M. Renger, and M. A. Peletier. A generalization of Onsagers reciprocity relations to gradient flows with nonlinear mobility. J. Non-Equil. Thermodyn., 41(2), 141-149, 2016.
  • [34] F. Otto. The geometry of dissipative evolution equations: the porous medium equation. Communications in Partial Differential Equations, 26(1-2):101–174, 2001.
  • [35] C. Villani. Hypocoercivity. Mem. Amer. Math. Soc., no. 950, iv+141, 202, 2009.
  • [36] C. Villani. Optimal Transport: Old and New. Number 338 in Grundlehren Der Mathematischen Wissenschaften. Springer, Berlin, 2009.