Controlling conservation laws I: entropy–entropy flux
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.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.
(1) |
where , is a unknown vector function, is a given initial condition, is a “conservative” differential operator, is a dissipative differential (diffusion) operator and is a diffusion constant. For simplicity of presentation, we assume that is a compact convex set, and the PDE (1) has periodic boundary conditions. E.g., , where is a –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.

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 , such that
Assumption 2: Suppose that there exists a “symmetric nonnegative definite” operator , such that
Following [22], under assumption 1 and 2, we demonstrate that
forms a Lyapunov functional for PDE (1). In other words, along PDE (1), the time derivative of is non-positive:
Based on the Lyapunov functional , we design optimal control problems of PDE (1). In detail: let ,
where is a flux function and are both symmetric non-negative definite matrix functions. Given a suitable potential functional and a terminal functional , consider a variational problem
(2a) | |||
where the infimum is taken among variables , , and satisfying | |||
(2b) |
By Proposition 8, we derive a critical point system of the optimal control problem (2). Define a Lagrange multiplier function . Then , and satisfies a pair of PDEs:
(3a) | |||
with both initial and terminal time boundary conditions | |||
(3b) |
In above formulations, is the first variation operator w.r.t. variable . 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)
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 , which can solve the initial value conservation laws. And [6] also enforces 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
Here , are the first and the second derivatives w.r.t. , and the unknown variable is . In this case, the conservative and dissipative operators satisfy
Define a function by
Then assumption 1 is satisfied since
Denote a “symmetric nonnegative definite operator” by
Assumption 2 is satisfied since
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
(4a) | |||
such that | |||
(4b) |
Here, the minimizer system of variational problem (4) satisfies a pair of PDEs: and
Velocity control. The unknown variable in Burgers’ equation describes the velocity filed of the fluid over time. In the variational problem (4), we design a potential field to control the evolution of velocity field . Here, the background dynamics of 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
Here , are divergence, Laplacian operators w.r.t , respectively, and the unknown variable is . In this case, the conservative and the dissipative operators satisfy
Define a function by
Then assumption 1 is satisfied since
where . Denote a “symmetric nonnegative definite operator” by
Assumption 2 is satisfied since
where the second equality is from the integration by parts formula and the third equality holds by the fact that , i.e., .
Formulations: Variational problem (2) forms an optimal control problem for traffic flows. In detail, consider
(5a) | |||
such that | |||
and | |||
(5b) |
Here, the minimizer system of variational problem (4) satisfies a pair of PDEs. When , there exists a scalar function , such that and
Position control. The unknown variable in traffic flows represents the density function of cars (particles) in a given spatial domain. Here, the background dynamics of 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
In this case, equation (1) satisfies
(6) |
where is a scalar function, is a flux vector function, and is a semi-positive definite matrix function. If , equation (6) is a scalar conservation law equation
(7) |
Definition 1 (Entropy-entropy flux pair condition [22]).
We call an entropy-entropy flux pair for the conservation law (7) if there exists a convex function , and , such that
Remark 1.
We remark that the entropy-entropy flux condition is trivial for scalar conservation laws. Every differentiable convex function 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
For assumption 1,
where we apply the fact that and in the last equality.
For assumption 2,
where we apply in the last equality. We observe a fact that , since is nonnegative definite and . Hence we know that
This implies that 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 . Here, the metric space connects Lyapunov functionals with dissipative operators through gradient descent flows.
Definition 2 (Entropy-entropy flux-metric condition).
We call an entropy-entropy flux pair-metric for equation (6) if there exists a convex function , and , such that
Under the entropy-entropy flux-metric condition, assumption 2 implies a metric operator below. Define the space of function by
The tangent space of at point is defined by
Denote an elliptic operator by
We are ready to define the metric inner product for the dissipation operator.
Definition 3 (Metric).
The inner product is given below.
where satisfies
From now on, we call the metric space. We next review and present the gradient decent flow in metric space . 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 , the gradient flow of in satisfies
If , then the above gradient flow satisfies
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 is defined by
for any . Let , the above definition forms
Since the above equality holds for any smooth function , we let
Hence the gradient decent flow satisfies
If , then
∎
Remark 2.
An example of metric is the Wasserstein-2 metric, which has been widely studied in optimal transport literature [1, 36]. In other words, let , , then . In this case, the heat equation is the gradient flow of entropy functional 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, . 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.
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 by a general energy functional .
Definition 5 (Flux–gradient flow).
Equations in formulation (9) provide a class of conservative–dissipative dynamics.
Proposition 6 (Entropy-entropy flux-production).
Proof.
Remark 3.
We remark that if and , then the functional
is known as the Fisher information functional. A known fact is that the dissipation of Lyapunov/entropy functional 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 , , consider a variational problem | |||
(12a) | |||
where the infimum is taken among variables , , and satisfying | |||
(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 . If the control variable for all , , 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 , , , , , and is a fixed function with , then problem (12) forms
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 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 , such that
and
(13) |
Here ′ represents the derivative w.r.t. variable . The initial and terminal time conditions satisfy
Proof.
We first rewrite the variables in variational formula (12) by , where
Here we denote . In this case, the variational problem (12) forms
(14) |
Denote the Lagrange multiplier of problem (14) by . Consider the following saddle point problem
In the above formula, we have
where we use the integration by parts formula w.r.t. in the second equality. In other words,
We next derive the critical point for the above saddle point problem. In other words, consider
where , , , are first variational derivatives w.r.t. functions , , , , respectively. We thus derive the pair of PDEs (13) in . ∎
We next present the Hamiltonian formalism for the PDE system (13).
Proposition 9 (Hamiltonian flows in metric space).
Proposition 10 (Functional Hamilton-Jacobi equations of conservation laws).
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,
and
Clearly, is conserved since
This finishes the proof of Proposition 9.
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
where is the probability density function. It satisfies PDE (6), where
In this case, a function pair , where is a convex function and , satisfies the entropy-entropy flux condition. In other words,
In particular, if , then the variational problem (12) satisfies
where the infimum is taken among variables satisfying
Here the minimizer system is given below. There exists a function , such that and
The above optimal control problem and critical point system has been widely studied in the optimal transport (, ), Schrödinger bridge problems (, ) and potential mean field games. There is a Hamiltonian formalism for the PDE system . In other words,
where the Hamiltonian functional satisfies
Example 2 (Controlling scalar conservation laws).
Consider
where and .
-
(i)
Let . Then variational problem (12) satisfies
where the infimum is taken among variables satisfying
Here the minimizer system is given below. There exists a function , such that and
I.e.,
where the Hamiltonian functional satisfies
-
(ii)
Let . Then variational problem (12) satisfies
where the infimum is taken among variables satisfying
Here the minimizer system is given below. There exists a function , such that and
I.e.,
where the Hamiltonian functional satisfies
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
where is a finite or infinite dimensional Hilbert space, and are convex functions and is a linear operator between Hilbert spaces. This problem can be rewritten in the saddle-point problem form
where , which is the convex conjugate of . The algorithm updates , by taking proximal gradient descent, proximal gradient ascent steps alternatively. At the -th iteration, the algorithm updates as follows
Here , are stepsizes, which have to satisfy in order to guarantee the convergence of the algorithm. When the operator is nonlinear, we use the extension of PDHG algorithm [13]. The idea is to use the linear approximation of :
Here, the extension of the PDHG scheme is as follows
(16) | ||||
When is some unbounded linear operator, for instance , the operator norm 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 () 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
We rewrite the varational problem as follows:
(17) |
where
(18) | ||||
and is the traffic flux function. We denote the indicator function by
We also choose norm for updating and norm for , specifically
We summarize the algorithm to solve problem (17) as follows.
Algorithm 1:PDHG for the conservation law control system |
While Maximal number of iteration |
; |
; |
; |
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
(19) |
We review the Lax–Friedrichs scheme. Denote a discretization in time and space, , then the update follows
(20) |
The last term contains diffusion coefficient from the original equation and as coefficient for artificial viscosity with . Notice that the monotone scheme gives entropy solutions. Here the definition of monotone scheme is given below.
Definition 11.
For , a scheme
is called a monotone scheme if 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:
As we want the scheme works when , the restriction on and space–time stepsizes can be simplified as follows:
The first inequality suggests the artificial viscosity we need to add. The second one impose a strong restriction on the stepsize in time when .
5.2.2. Discreitization of the control problem
We consider the control problem of scalar conservation law defined in , where is a given constant. We apply the periodic boundary condition on the spatial domain. Given , we have , . For , define
where and . Note here . Denote
The first conservation law equation adapted from the Lax–Friedrichs scheme is as follows:
(21) |
where , . We choose . 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:
where
By taking the first order derivative of , we automatically get the implicit finite difference scheme for the dual equation of that is backward in time. The positive and negative parts of are split, which help enhance the monotonicity of the discrete Hamiltonian.
(22) |
for , .
5.2.3. Solve conservation laws via primal-dual algorithms
When , for some constant , the variational problem (17) becomes classical conservation laws with initial data. In this case, no control will be enforced on the density function . Therefore, the density movement is only determined by the flux term 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 is point-wise operation for each . 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 .
where . The entropy solution to this problem at satisfies
In the following examples, we use the same spatial time domain with 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 .

5.5. Example 2.
We consider the traffic flow equation
where .

The entropy solution to this problem at time is
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 waiting at the traffic light at . At the time , 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 is closer to the red light (), moves faster. The density at position doesn’t change until time .
5.6. Example 3
We again consider the traffic flow equation with . The final cost functional . We set . In this case, the density will tend to form a ‘similar’ distribution as . When and , this final cost functional is equivalent to imposes the constraint that . We also compare the result from the control of conservation law with a mean-field game problem, i.e., :
(23) |
As shown in Figure 4, we set
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.


5.7. Example 4
Consider the traffic flow equation with . The final cost functional . The initial density and final cost function are as follows
The term penalizes the density for getting too concentrate. Figure 6 shows the effect of the term , where the density is more spreading in space for the case than the case and . We can also see from the profile at the terminal time. In Figure 7 (middle), case has 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 is preserved with error of order of .


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.