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

A LWR model with constraints at moving interfaces

Abraham Sylla1
Abstract

We propose a mathematical framework to the study of scalar conservation laws with moving interfaces. This framework is developed on a LWR model with constraint on the flux along these moving interfaces. Existence is proved by means of a finite volume scheme. The originality lies in the local modification of the mesh and in the treatment of the crossing points of the trajectories.

11footnotetext: [email protected]
Institut Denis Poisson, CNRS UMR 7013, Université de Tours, Université d’Orléans
Parc de Grandmont, 37200 Tours cedex, France
ORCID number: 0000-0003-1784-4878

2020 AMS Subject Classification: 35L65, 76A30, 65M08.

Keywords: Hyperbolic scalar conservation laws, moving interfaces, flux constraints, finite volume scheme.

1 Introduction

Being given a regular concave flux f𝐂2([0,1],+)f\in\mathbf{C}^{2}([0,1],\mathbb{R}^{+}) verifying

f(0)=f(1)=0,!ρ¯]0,1[,for a.e.ρ]0,1[,f(ρ)(ρ¯ρ)>0,f(0)=f(1)=0,\quad\exists!\;\overline{\rho}\in\mathopen{]}0,1\mathclose{[},\;\text{for a.e.}\;\rho\in\mathopen{]}0,1\mathclose{[},\quad f^{\prime}(\rho)(\overline{\rho}-\rho)>0, (1.1)

and a finite family of trajectories (yi)i[[1;J]](y_{i})_{i\in[\![1;J]\!]} and constraints (qi)i[[1;J]](q_{i})_{i\in[\![1;J]\!]} defined on ]si,Ti[\mathopen{]}s_{i},T_{i}\mathclose{[} (0si<Ti0\leq s_{i}<T_{i}), we tackle the following problem:

{tρ(x,t)+x(f(ρ(x,t)))=0(x,t)×]0,+[:=Ωρ(x,0)=ρo(x)xi[[1;J]],(f(ρ)y˙i(t)ρ)|x=yi(t)qi(t)t]si,Ti[.\left\{\begin{array}[]{rlr}\partial_{t}\rho(x,t)+\partial_{x}\left(f(\rho(x,t))\right)&=0&(x,t)\in\mathbb{R}\times\mathopen{]}0,+\infty\mathclose{[}:=\Omega\\[5.0pt] \rho(x,0)&=\rho_{o}(x)&x\in\mathbb{R}\\[5.0pt] \forall i\in[\![1;J]\!],\quad\left.\left(f(\rho)-\dot{y}_{i}(t)\rho\right)\right|_{x=y_{i}(t)}&\leq q_{i}(t)&t\in\mathopen{]}s_{i},T_{i}\mathclose{[}.\end{array}\right. (1.2)

Systems of the type (1.2) have naturally arisen in the recent years. Let us give a non-exhaustive review on how our Problem (1.2) relates to the existing literature.

  • The authors of [12, 15] considered a model very similar to (1.2). In their framework, (yi)i(y_{i})_{i} represented the trajectories of autonomous vehicles, and the authors aimed at modeling the regulation impact on a few autonomous vehicles on the traffic flow. In the same framework but with different applications in mind, the model of [20] accounts for the boundedness of traffic acceleration. Note that in each of these models, the trajectories of the moving interfaces (yi)i(y_{i})_{i} were not given a priori, but rather obtained as solutions to an ODE involving the density of traffic, a mechanism reminiscent of [2, 9, 21] for instance. Let us also mention the work of [16] where the authors studied a different model for the situation of several moving bottlenecks.

  • The numerical aspect of (1.2) was treated in [7] (for one trajectory) and [10] (for multiple trajectories), where the authors modeled the moving bottlenecks created by buses on a road.

  • In a class of problems close to (1.2), i.e. without constraint on the flux, but still with coupling interfaces/density, the authors of [14] described the interaction between a platoon of vehicles and the surrounding traffic flow on a highway.

  • Problem (1.2) can be seen as a conservation law with discontinuous flux and special treatments at the interfaces. In that directions, the authors of [18, 4, 1, 6, 23] studied such problems but with the classical vanishing viscosity coupling at the interfaces.

In several of these works [15, 20], the existence issue is tackled using the wave-front tracking procedure which is very sensible to the details of the model. On the other hand, when numerical schemes are considered, see [10, 7], the numerical analysis is usually left out.

The contribution of this paper is to provide a robust mathematical setting both in the theoretical and numerical aspects of (1.2). The proof of uniqueness is based upon a combination of Kruzhkov classical method of doubling variables and the theory of dissipative germs in the framework of discontinuous flux [3], and it is analogous to the one of [4]. To prove existence, we build a finite volume scheme with a grid that adapts locally to the trajectories (yi)i(y_{i})_{i} and to their crossing points, but remains a simple Cartesian grid away from the interfaces. Our work can serve as a basis for constructing solutions to more involved models, e.g. via the splitting approach. As an example of application, we can point out the variant of our recent work [21] with multiple slow vehicles involved; this is a mildly non-local analogue of the problem considered numerically in [10].

As the fundamental ingredient of the well-posedness proof and numerical approximation of (1.2), we will first tackle the one trajectory/one constraint problem:

{tρ+x(f(ρ))=0ρ(,0)=ρo(f(ρ)y˙(t)ρ)|x=y(t)q(t)t>0,\left\{\begin{aligned} \partial_{t}\rho+\partial_{x}\left(f(\rho)\right)&=0\\[5.0pt] \rho(\cdot,0)&=\rho_{o}\\[5.0pt] \left.\left(f(\rho)-\dot{y}(t)\rho\right)\right|_{x=y(t)}&\leq q(t)&t>0,\end{aligned}\right. (1.3)

with y𝐖𝐥𝐨𝐜1,(]0,+[,)y\in\mathbf{W}_{\mathbf{loc}}^{1,\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}) and q𝐋𝐥𝐨𝐜(]0,+[,)q\in\mathbf{L}_{\mathbf{loc}}^{\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}). Models in the class of (1.3) have been greatly investigated in the past few decades. Motivated by the modeling of tollgates and traffic lights for instance, the authors of [8] considered (1.3) with the trivial trajectory y0y\equiv 0 and proved a well-posedness result in the 𝐁𝐕\mathbf{BV} framework (i.e. with both qq and ρo\rho_{o} with bounded variation, locally). The authors of [2] then extended the well-posedness in the 𝐋\mathbf{L}^{\infty} framework and also constructed a convergent numerical scheme. More recently, in [9, 11, 21], the authors studied a variant of (1.3) in which ρ\rho and y˙\dot{y} were coupled via an ODE. The coupling was thought to model the influence of a slow vehicle, traveling at speed y˙\dot{y}, on road traffic.

The reduction of (1.2) to localized problem (1.3) requires the construction of a finite volume scheme in the original coordinates (x,t)(x,t), while the treatment of (1.3) in the literature is most often based upon the rectification of the interface via a variable change, see [9, 11, 21]. For (1.2), this approach leads to a cumbersome and singular construction, see [4]. In our well-posedness analysis and approximation of (1.3), having in mind (1.2), we will not change the coordinate system.

Let us detail how the paper is organized. Sections 2-3 are devoted to Problem (1.3). We start by giving two definitions of solutions. One, most frequently used in traffic dynamics (see [8, 5]), is composed of classical Kruzhkov entropy inequalities with reminder term taking into account the constraint and of a weak formulation for the constraint, see Definition 2.1. The second definition emanates from the theory of conservation laws with dissipative interface coupling (see [3, 1]). It consists of Kruzhkov entropy inequalities with test functions that vanish along the interface {x=y(t)}\{x=y(t)\} and of an explicit treatment of the traces of the solution along the interface, see Definition 2.4. Before tackling the well-posedness issue, we prove that these two definitions are equivalent, see Propositions 2.6-2.6, similarly to what the authors of [2] did. Uniqueness follows from the stability obtained in Section 2, see Theorem 2.13. In Section 3, we construct a finite volume scheme for (1.3) and prove of its convergence. In the construction, we do not rectify the trajectory, but instead we locally modify the mesh to mold the trajectory. Moreover, we fully make use of techniques and results put forward by the author of [22] to derive localized 𝐁𝐕\mathbf{BV} estimates away from the interface, essential to obtain strong compactness for the approximate solutions created by the scheme, see Corollary 3.7. This is a way to highlight the generality of the compactness technique of [22].

In Section 4, we get back to the original problem (1.2). Our strategy is to assemble the study of (1.2) from several local studies of (1.3) with the help of a partition of unity argument. This concerns, in particular, the convergence of finite volume approximation of (1.2) which is addressed via a localization argument. However, the scheme needs to be defined globally, which makes it impossible to use the rectification strategy as soon as the interfaces have crossing points, see [4] for a singular rectification strategy.

2 Uniqueness and stability for the single trajectory problem

The content of this section is not original in the sense that it is a rigorous adaptation and assembling of existing techniques reminiscent of [24, 19, 8, 2, 3].

2.1 Equivalent definitions of solutions

Throughout the paper, for all ss\in\mathbb{R}, we denote by

ρ[0,1],Fs(ρ):=f(ρ)sρanda,b[0,1],Φs(a,b):=sgn(ab)(Fs(a)Fs(b)),\forall\rho\in[0,1],\;F_{s}(\rho):=f(\rho)-s\rho\quad\text{and}\quad\forall a,b\in[0,1],\;\Phi_{s}(a,b):=\mathop{\rm sgn}(a-b)(F_{s}(a)-F_{s}(b)),

the normal flux through {x=xo+st}\{x=x_{o}+st\} (xox_{o}\in\mathbb{R}) and its entropy flux associated with the Kruzhkov entropy ρ|ρκ|\rho\mapsto|\rho-\kappa|, for all κ[0,1]\kappa\in[0,1], see [19]. Let us also denote by Γ\Gamma the trajectory/interface:

Γ:={(y(t),t):t[0,+[}.\Gamma:=\{(y(t),t)\;:\;t\in[0,+\infty[\}.
Definition 2.1.

Let ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]). We say that ρ𝐋(Ω,[0,1])\rho\in\mathbf{L}^{\infty}(\Omega,[0,1]) is an admissible entropy solution to (1.3) if

(i) for all test functions φ𝐂𝐜(Ω¯,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}^{+}) and κ[0,1]\kappa\in[0,1], the following entropy inequalities are verified:

0+\displaystyle\int_{0}^{+\infty}\int_{\mathbb{R}} (|ρκ|tφ+Φ(ρ,κ)xφ)dxdt+|ρo(x)κ|φ(x,0)dx\displaystyle\biggl{(}|\rho-\kappa|\partial_{t}\varphi+\Phi(\rho,\kappa)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{o}(x)-\kappa|\varphi(x,0)\mathinner{\mathrm{d}{x}} (2.1)
+0+y˙(t)(κ,q(t))φ(y(t),t)dt0,\displaystyle+\int_{0}^{+\infty}\mathcal{R}_{\dot{y}(t)}(\kappa,q(t))\varphi(y(t),t)\mathinner{\mathrm{d}{t}}\geq 0,

where

y˙(t)(κ,q(t)):=2(Fy˙(t)(κ)min{Fy˙(t)(κ),q(t)});\mathcal{R}_{\dot{y}(t)}(\kappa,q(t)):=2\left(F_{\dot{y}(t)}(\kappa)-\min\left\{F_{\dot{y}(t)}(\kappa),q(t)\right\}\right);

(ii) for all test functions φ𝐂𝐜(Ω,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\Omega,\mathbb{R}^{+}) the following constraint inequalities are verified:

Ω+(ρtφ+f(ρ)xφ)dxdt0+q(t)φ(y(t),t)dt,-\iint_{\Omega^{+}}\biggl{(}\rho\partial_{t}\varphi+f(\rho)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\leq\int_{0}^{+\infty}q(t)\varphi(y(t),t)\mathinner{\mathrm{d}{t}}, (2.2)

where Ω+:={(x,t)Ω:x>y(t)}\displaystyle{\Omega^{+}:=\left\{(x,t)\in\Omega\;:\;x>y(t)\right\}}.

Remark 2.1.

Taking κ=0\kappa=0, then κ=1\kappa=1 in (2.1), from the condition ρ(x,t)[0,1]\rho(x,t)\in[0,1] a.e. we deduce that any admissible weak solution to Problem (1.3) is also a distributional solution to the conservation law tρ+xf(ρ)=0\partial_{t}\rho+\partial_{x}f(\rho)=0. If ρ\rho is a regular enough solution, then for all test functions φ𝐂𝐜(Ω,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\Omega,\mathbb{R}^{+}), we have

0\displaystyle 0 =Ω+div(x,t)(f(ρ)ρ)φdxdt\displaystyle=\iint_{\Omega^{+}}\text{div}_{(x,t)}\begin{pmatrix}f(\rho)\\ \rho\end{pmatrix}\varphi\ \mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
=Ω+(f(ρ)φρφ)(1y˙(t))dtΩ+(f(ρ)ρ)x,tφdxdt\displaystyle=\int_{\partial\Omega^{+}}\begin{pmatrix}f(\rho)\varphi\\ \rho\varphi\end{pmatrix}\cdot\begin{pmatrix}-1\\ \dot{y}(t)\end{pmatrix}\mathinner{\mathrm{d}{t}}-\iint_{\Omega^{+}}\begin{pmatrix}f(\rho)\\ \rho\end{pmatrix}\cdot\nabla_{x,t}\varphi\ \mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
=0+((f(ρ)y˙(t)ρ)|x=y(t))φ(y(t),t)dtΩ+(ρtφ+f(ρ)xφ)dxdt.\displaystyle=-\int_{0}^{+\infty}\biggl{(}\left(f(\rho)-\dot{y}(t)\rho\right)_{|x=y(t)}\biggr{)}\varphi(y(t),t)\mathinner{\mathrm{d}{t}}-\iint_{\Omega^{+}}\biggl{(}\rho\partial_{t}\varphi+f(\rho)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}.

Moreover, if ρ\rho satisfies the flux inequality of (1.3) a.e. on ]0,+[\mathopen{]}0,+\infty\mathclose{[}, then the previous computations lead to

Ω+(ρtφ+f(ρ)xφ)dxdt0+q(t)φ(y(t),t)dt;-\iint_{\Omega^{+}}\biggl{(}\rho\partial_{t}\varphi+f(\rho)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\leq\int_{0}^{+\infty}q(t)\varphi(y(t),t)\mathinner{\mathrm{d}{t}};

this is where inequalities (2.2) come from. Note how they make sense irrespective of the regularity of ρ\rho. Integrating on Ω:={(x,t)Ω:x<y(t)}\displaystyle{\Omega^{-}:=\left\{(x,t)\in\Omega\;:\;x<y(t)\right\}} would lead to similar and equivalent inequalities.

Definition 2.1 is well suited for passage to the limit of a.e. convergent sequences of exact or approximate solutions. However, we cannot derive uniqueness by the standard arguments like in the classical case of Kruzhkov. Using an equivalent notion of solution, which we adapt from [3], based on explicit treatment of traces of ρ\rho on Γ\Gamma, we rather combine the arguments of [19] and [24]. In this definition a couple plays a major role, the one which realizes the equality in the flux constraint in (1.3). More precisely, fix first s0s\geq 0. By (1.1) and concavity of ff, for all q[0,maxFs)q\in[0,\max F_{s}), the equation Fs(ρ)=qF_{s}(\rho)=q admits exactly two solutions in [0,1][0,1], see Figure 1, left. The same way, if s0s\leq 0, then for all q[s˙,maxFs)q\in[-\dot{s},\max F_{s}), the equation still admits two solutions in [0,1][0,1]. The couple formed by these two solutions, denoted by (ρ^s(q),ρwidechecks(q))\left(\widehat{\rho}_{s}(q),\widecheck{\rho}_{s}(q)\right) in Definition 2.2 below, will serve both in the proof of uniqueness and existence.

Refer to caption
Figure 1: Illustration of Assumption (2.3)

Following the previous discussion, in the sequel, we will assume that qq verifies the following assumption:

for a.e.t>0,q(t)[0,maxFy˙(t)[ify˙(t)0andq(t)[y˙(t),maxFy˙(t)[ify˙(t)<0.\text{for a.e.}\;t>0,\quad q(t)\in[0,\max F_{\dot{y}(t)}\mathclose{[}\;\text{if}\;\dot{y}(t)\geq 0\;\;\text{and}\;\;q(t)\in[-\dot{y}(t),\max F_{\dot{y}(t)}\mathclose{[}\;\text{if}\;\dot{y}(t)<0. (2.3)

In particular, remark that

for a.e.t>0,y˙(t)+q(t)0.\text{for a.e.}\;t>0,\quad\dot{y}(t)+q(t)\geq 0. (2.4)
Definition 2.2.

Let s+s\in\mathbb{R}^{+} and q[0,maxFs[q\in[0,\max F_{s}\mathclose{[}, or ss\in\mathbb{R}^{-} and q[s,maxFs[q\in[-s,\max F_{s}\mathclose{[}. The admissibility germ for the conservation law in (1.3) associated with the constraint Fs(ρ)|x=stq\displaystyle{F_{s}(\rho)_{|x=st}\leq q} is the subset 𝒢s(q)[0,1]2\mathcal{G}_{s}(q)\subset[0,1]^{2} defined as the union:

𝒢s(q):=(ρ^s(q),ρwidechecks(q))𝒢s1(q){(κ,κ):Fs(κ)q}𝒢s2(q){(kl,kr):kl<krandFs(kl)=Fs(kr)q}𝒢s3(q),\mathcal{G}_{s}(q):=\underbrace{\left(\widehat{\rho}_{s}(q),\widecheck{\rho}_{s}(q)\right)}_{\mathcal{G}^{1}_{s}(q)}\bigcup\underbrace{\{(\kappa,\kappa)\;:\;F_{s}(\kappa)\leq q\}}_{\mathcal{G}^{2}_{s}(q)}\bigcup\underbrace{\left\{(k_{l},k_{r})\;:\;k_{l}<k_{r}\;\text{and}\;F_{s}(k_{l})=F_{s}(k_{r})\leq q\right\}}_{\mathcal{G}^{3}_{s}(q)},

where, due to the bell-shaped profile of FsF_{s}, the couple (ρ^s(q),ρwidechecks(q))\left(\widehat{\rho}_{s}(q),\widecheck{\rho}_{s}(q)\right) is uniquely defined by the conditions

Fs(ρ^s(q))=Fs(ρwidechecks(q))=qandρ^s(q)>ρwidechecks(q).F_{s}(\widehat{\rho}_{s}(q))=F_{s}(\widecheck{\rho}_{s}(q))=q\quad\text{and}\quad\widehat{\rho}_{s}(q)>\widecheck{\rho}_{s}(q).
Lemma 2.3.

For all s+s\in\mathbb{R}^{+} and q[0,maxFs[q\in[0,\max F_{s}\mathclose{[}, and for all ss\in\mathbb{R}^{-} and q[s,maxFs[q\in[-s,\max F_{s}\mathclose{[}, the admissibility germ 𝒢s(q)\mathcal{G}_{s}(q) is 𝐋1\mathbf{L}^{1}-dissipative in the sense that:

(i) for all (kl,kr)𝒢s(q)(k_{l},k_{r})\in\mathcal{G}_{s}(q), Fs(kl)=Fs(kr)F_{s}(k_{l})=F_{s}(k_{r}) (Rankine-Hugoniot condition);

(ii) for all (kl,kr),(cl,cr)𝒢s(q)(k_{l},k_{r}),(c_{l},c_{r})\in\mathcal{G}_{s}(q),

Φs(kl,cl)Φs(kr,cr).\Phi_{s}(k_{l},c_{l})\geq\Phi_{s}(k_{r},c_{r}). (2.5)

Proof.  The point (i) is obvious from the definition. Let us prove the dissipative feature (2.5). The following table summarizes which values can take the difference Δ=Φs(kl,cl)Φs(kr,cr)\Delta=\Phi_{s}(k_{l},c_{l})-\Phi_{s}(k_{r},c_{r}), depending on in which parts of the germ the couples (kl,kr),(cl,cr)𝒢s(q)(k_{l},k_{r}),(c_{l},c_{r})\in\mathcal{G}_{s}(q) belong to.

(cl,cr)(c_{l},c_{r}) (kl,kr)(k_{l},k_{r}) 𝒢s1(q)\in\mathcal{G}_{s}^{1}(q) 𝒢s2(q)\in\mathcal{G}_{s}^{2}(q) 𝒢s3(q)\in\mathcal{G}_{s}^{3}(q)
𝒢s1(q)\in\mathcal{G}_{s}^{1}(q) 0 0 0 or 2(qFs(kl))2(q-F_{s}(k_{l}))
𝒢s2(q)\in\mathcal{G}_{s}^{2}(q) 0 0 0 or 2|Fs(c)Fs(kl)|2|F_{s}(c)-F_{s}(k_{l})|
𝒢s3(q)\in\mathcal{G}_{s}^{3}(q) 0 or 2(qFs(cl))2(q-F_{s}(c_{l})) 0 or 2|Fs(cl)Fs(k)|2|F_{s}(c_{l})-F_{s}(k)| 0 or 2|Fs(cl)Fs(kl)|2|F_{s}(c_{l})-F_{s}(k_{l})|

Having in mind the definition of 𝒢s3(q)\mathcal{G}^{3}_{s}(q), we can conclude that Δ0\Delta\geq 0.   \square

Definition 2.4.

Let ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]). We say that ρ𝐋(Ω,[0,1])\rho\in\mathbf{L}^{\infty}(\Omega,[0,1]) is a 𝒢y˙(q)\mathcal{G}_{\dot{y}}(q)-entropy solution to (1.3) if:

(i) for all test functions φ𝐂𝐜(Ω¯\Γ,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega}\backslash\Gamma,\mathbb{R}^{+}) and κ[0,1]\kappa\in[0,1], the following entropy inequalities are verified:

0+(|ρκ|tφ+Φ(ρ,κ)xφ)dxdt+|ρo(x)κ|φ(x,0)dx0;\int_{0}^{+\infty}\int_{\mathbb{R}}\biggl{(}|\rho-\kappa|\partial_{t}\varphi+\Phi(\rho,\kappa)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{o}(x)-\kappa|\varphi(x,0)\mathinner{\mathrm{d}{x}}\geq 0; (2.6)

(ii) for a.e. t>0t>0,

(ρ(y(t),t),ρ(y(t)+,t))𝒢y˙(t)(q(t)).(\rho(y(t)-,t),\rho(y(t)+,t))\in\mathcal{G}_{\dot{y}(t)}(q(t)). (2.7)
Remark 2.2.

Condition (2.7) is to be understood in the sense of strong traces along Γ\Gamma. An important fact we stress is that it is not restrictive to assume that entropy solutions, i.e. bounded functions verifying (2.6), admit strong traces. Usually, it is ensured provided a nondegeneracy assumption on the flux function:

for any nonempty interval]a,b[]0,1[,f 1]a,b[is not constant.\text{for any nonempty interval}\;\mathopen{]}a,b\mathclose{[}\subset\mathopen{]}0,1\mathclose{[},\quad f\;\mathbf{1}_{\mathopen{]}a,b\mathclose{[}}\;\text{is not constant.} (2.8)

In the context of traffic flow, however, we sometimes consider fluxes which do not verify (2.8). Such fluxes, which have linear parts, usually model constant traffic velocity for small densities. In those situations, and when y0y\equiv 0, one can prove that under a mild assumption on the constraint, if the initial datum has bounded variation, then solutions to (1.3) are in 𝐋(]0,T[,𝐁𝐕())\mathbf{L}^{\infty}(\mathopen{]}0,T\mathclose{[},\mathbf{BV}(\mathbb{R})), and traces are then to be understood in the sense of 𝐁𝐕()\mathbf{BV}(\mathbb{R}) functions, see [21, Theorem 3.2]. Also note that the germ formalism can be adapted to the situations where the flux is degenerate and no variation bound is assumed, see [3, Remarks 2.2, 2.3].

We now prove that Definitions 2.1 and 2.4 are equivalent.

Proposition 2.5.

Any admissible entropy solution to (1.3) is a 𝒢y˙(q)\mathcal{G}_{\dot{y}}(q)-entropy solution.

Proof.  Fix ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]) and let ρ𝐋(Ω,[0,1])\rho\in\mathbf{L}^{\infty}(\Omega,[0,1]) be an admissible entropy solution to (1.3). Let φ𝐂𝐜(Ω¯,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}^{+}) and κ[0,1]\kappa\in[0,1]. If φ\varphi vanishes along Γ\Gamma, then (2.1) becomes (2.6). Moreover, it is known that the Rankine-Hugoniot condition is contained in (2.1). Combining it with (2.2) gives us:

for a.e.t>0,Fy˙(t)(ρ(y(t),t))=Fy˙(t)(ρ(y(t)+,t))q(t).\text{for a.e.}\;t>0,\quad F_{\dot{y}(t)}(\rho(y(t)-,t))=F_{\dot{y}(t)}(\rho(y(t)+,t))\leq q(t). (2.9)

Let us show that for a.e. t>0t>0, (ρ(y(t),t),ρ(y(t)+,t))𝒢y˙(t)(q(t))(\rho(y(t)-,t),\rho(y(t)+,t))\in\mathcal{G}_{\dot{y}(t)}(q(t)).

Case 1: ρ(y(t),t)ρ(y(t)+,t)\rho(y(t)-,t)\leq\rho(y(t)+,t). Condition (2.9) implies that (ρ(y(t),t),ρ(y(t)+,t))𝒢y˙(t)2(q(t))𝒢y˙(t)3(q(t))(\rho(y(t)-,t),\rho(y(t)+,t))\in\mathcal{G}^{2}_{\dot{y}(t)}(q(t))\cup\mathcal{G}^{3}_{\dot{y}(t)}(q(t)).

Case 2: ρ(y(t),t)>ρ(y(t)+,t)\rho(y(t)-,t)>\rho(y(t)+,t). Suppose now that φ𝐂𝐜(Ω,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\Omega,\mathbb{R}^{+}) and fix nn\in\mathbb{N}^{*}. By a standard approximation argument, we can apply (2.1) with the Lipschitz test function ξnφ\xi_{n}\varphi, where ξn\xi_{n} is the cut-off function:

ξn(x,t)={1if|xy(t)|<1n2n|xy(t)|if1n|xy(t)|2n0if|xy(t)|>2n.\xi_{n}(x,t)=\left\{\begin{array}[]{ccc}1&\text{if}&\displaystyle{|x-y(t)|<\frac{1}{n}}\\[5.0pt] 2-n|x-y(t)|&\text{if}&\displaystyle{\frac{1}{n}\leq|x-y(t)|\leq\frac{2}{n}}\\[5.0pt] 0&\text{if}&\displaystyle{|x-y(t)|>\frac{2}{n}}.\end{array}\right.

This yields:

0+|ρκ|(ξntφ+ny˙(t)sgn(xy(t))𝟏{1n<|xy(t)|<2n}φ)dxdt\displaystyle\int_{0}^{+\infty}\int_{\mathbb{R}}|\rho-\kappa|\left(\xi_{n}\partial_{t}\varphi+n\dot{y}(t)\mathop{\rm sgn}(x-y(t))\mathbf{1}_{\left\{\frac{1}{n}<|x-y(t)|<\frac{2}{n}\right\}}\varphi\right)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
+0+Φ(ρ,κ)(ξnxφnsgn(xy(t)𝟏{1n<|xy(t)|<2n}φ)dxdt\displaystyle+\int_{0}^{+\infty}\int_{\mathbb{R}}\Phi(\rho,\kappa)\left(\xi_{n}\partial_{x}\varphi-n\mathop{\rm sgn}(x-y(t)\mathbf{1}_{\left\{\frac{1}{n}<|x-y(t)|<\frac{2}{n}\right\}}\varphi\right)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
+0+y˙(t)(κ,q(t))φ(y(t),t)dt0.\displaystyle+\int_{0}^{+\infty}\mathcal{R}_{\dot{y}(t)}(\kappa,q(t))\varphi(y(t),t)\mathinner{\mathrm{d}{t}}\geq 0.

Taking the limit when n+n\to+\infty, we obtain:

0+(Φy˙(t)(ρ(y(t),t),κ)Φy˙(t)(ρ(y(t)+,t),κ)+y˙(t)(κ,q(t)))φ(y(t),t)dt0,\int_{0}^{+\infty}\biggl{(}\Phi_{\dot{y}(t)}\left(\rho(y(t)-,t),\kappa\right)-\Phi_{\dot{y}(t)}\left(\rho(y(t)+,t),\kappa\right)+\mathcal{R}_{\dot{y}(t)}(\kappa,q(t))\biggr{)}\varphi(y(t),t)\mathinner{\mathrm{d}{t}}\geq 0,

which implies that for a.e. t>0t>0 and for all κ[0,1]\kappa\in[0,1],

Φy˙(t)(ρ(y(t),t),κ)Φy˙(t)(ρ(y(t)+,t),κ)+y˙(t)(κ,q(t))0.\Phi_{\dot{y}(t)}\left(\rho(y(t)-,t),\kappa\right)-\Phi_{\dot{y}(t)}\left(\rho(y(t)+,t),\kappa\right)+\mathcal{R}_{\dot{y}(t)}(\kappa,q(t))\geq 0.

Taking in particular κ=argmax(Fy˙(t))\kappa=\text{argmax}(F_{\dot{y}(t)}), we get:

Φy˙(t)(ρ(y(t),t),κ)Φy˙(t)(ρ(y(t)+,t),κ)+2(Fy˙(t)(κ)q(t))0.\Phi_{\dot{y}(t)}\left(\rho(y(t)-,t),\kappa\right)-\Phi_{\dot{y}(t)}\left(\rho(y(t)+,t),\kappa\right)+2(F_{\dot{y}(t)}(\kappa)-q(t))\geq 0. (2.10)

Since ρ(y(t),t)>ρ(y(t)+,t)\rho(y(t)-,t)>\rho(y(t)+,t), (2.10) leads to Fy˙(t)(ρ(y(t),t))q(t)\displaystyle{F_{\dot{y}(t)}(\rho(y(t)-,t))\geq q(t)}, which combined with (2.9), implies Fy˙(t)(ρ(y(t),t))=Fy˙(t)(ρ(y(t)+,t))=q(t)\displaystyle{F_{\dot{y}(t)}(\rho(y(t)-,t))=F_{\dot{y}(t)}(\rho(y(t)+,t))=q(t)}. We deduce that (ρ(y(t),t),ρ(y(t)+,t))𝒢y˙(t)1(q(t))(\rho(y(t)-,t),\rho(y(t)+,t))\in\mathcal{G}^{1}_{\dot{y}(t)}(q(t)), which completes the proof.   \square

Proposition 2.6.

Any 𝒢y˙(q)\mathcal{G}_{\dot{y}}(q)-entropy solution to (1.3) is an admissible entropy solution.

Proof.  Fix ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]) and let ρ𝐋(Ω,[0,1])\rho\in\mathbf{L}^{\infty}(\Omega,[0,1]) be a 𝒢y˙(q)\mathcal{G}_{\dot{y}}(q)-entropy solution to (1.3). Let φ𝐂𝐜(Ω¯,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}^{+}) and κ[0,1]\kappa\in[0,1]. We still denote by ξn\xi_{n} the cut-off function from the last proof. We write φ=(1ξn)φ+ξnφ\displaystyle{\varphi=(1-\xi_{n})\varphi+\xi_{n}\varphi}. Since ϕn=(1ξn)φ\phi_{n}=(1-\xi_{n})\varphi vanishes along Γ\Gamma, we have

𝐈\displaystyle\mathbf{I} =0+(|ρκ|tφ+Φ(ρ,κ)xφ)dxdt+|ρo(x)κ|φ(x,0)dx\displaystyle=\int_{0}^{+\infty}\int_{\mathbb{R}}\biggl{(}|\rho-\kappa|\partial_{t}\varphi+\Phi(\rho,\kappa)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{o}(x)-\kappa|\varphi(x,0)\mathinner{\mathrm{d}{x}}
+0+y˙(t)(κ,q(t))φ(y(t),t)dt\displaystyle+\int_{0}^{+\infty}\mathcal{R}_{\dot{y}(t)}(\kappa,q(t))\varphi(y(t),t)\mathinner{\mathrm{d}{t}}
=0+(|ρκ|tϕn+Φ(ρ,κ)xϕn)dxdt+|ρo(x)κ|ϕn(x,0)dx0\displaystyle=\underbrace{\int_{0}^{+\infty}\int_{\mathbb{R}}\biggl{(}|\rho-\kappa|\partial_{t}\phi_{n}+\Phi(\rho,\kappa)\partial_{x}\phi_{n}\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{o}(x)-\kappa|\phi_{n}(x,0)\mathinner{\mathrm{d}{x}}}_{\geq 0}
+0+(|ρκ|t(ξnφ)+Φ(ρ,κ)x(ξnφ))dxdt+|ρo(x)κ|ξn(x,0)φ(x,0)dx\displaystyle+\int_{0}^{+\infty}\int_{\mathbb{R}}\biggl{(}|\rho-\kappa|\partial_{t}(\xi_{n}\varphi)+\Phi(\rho,\kappa)\partial_{x}(\xi_{n}\varphi)\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{o}(x)-\kappa|\xi_{n}(x,0)\varphi(x,0)\mathinner{\mathrm{d}{x}}
+0+y˙(t)(κ,q(t))φ(y(t),t)dt\displaystyle+\int_{0}^{+\infty}\mathcal{R}_{\dot{y}(t)}(\kappa,q(t))\varphi(y(t),t)\mathinner{\mathrm{d}{t}}
0+|ρκ|(ξntφ+ny˙(t)sgn(xy(t))𝟏{1n<|xy(t)|<2n}φ)dxdt\displaystyle\geq\int_{0}^{+\infty}\int_{\mathbb{R}}|\rho-\kappa|\left(\xi_{n}\partial_{t}\varphi+n\dot{y}(t)\mathop{\rm sgn}(x-y(t))\mathbf{1}_{\left\{\frac{1}{n}<|x-y(t)|<\frac{2}{n}\right\}}\varphi\right)\ \mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
+0+Φ(ρ,κ)(ξnxφnsgn(xy(t)𝟏{1n<|xy(t)|<2n}φ)dxdt\displaystyle+\int_{0}^{+\infty}\int_{\mathbb{R}}\Phi(\rho,\kappa)\left(\xi_{n}\partial_{x}\varphi-n\mathop{\rm sgn}(x-y(t)\mathbf{1}_{\left\{\frac{1}{n}<|x-y(t)|<\frac{2}{n}\right\}}\varphi\right)\ \mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
+|ρo(x)κ|ξn(x,0)φ(x,0)dx+0+y˙(t)(κ,q(t))φ(y(t),t)dt.\displaystyle+\int_{\mathbb{R}}|\rho_{o}(x)-\kappa|\xi_{n}(x,0)\varphi(x,0)\mathinner{\mathrm{d}{x}}+\int_{0}^{+\infty}\mathcal{R}_{\dot{y}(t)}(\kappa,q(t))\varphi(y(t),t)\mathinner{\mathrm{d}{t}}.

Taking the limit when n+n\to+\infty, we obtain:

𝐈0+(Φy˙(t)(ρ(y(t),t),κ)Φy˙(t)(ρ(y(t)+,t),κ)+y˙(t)(κ,q(t))Δ(t,κ))φ(y(t),t)dt.\mathbf{I}\geq\int_{0}^{+\infty}\biggl{(}\underbrace{\Phi_{\dot{y}(t)}\left(\rho(y(t)-,t),\kappa\right)-\Phi_{\dot{y}(t)}\left(\rho(y(t)+,t),\kappa\right)+\mathcal{R}_{\dot{y}(t)}(\kappa,q(t))}_{\Delta(t,\kappa)}\biggr{)}\varphi(y(t),t)\mathinner{\mathrm{d}{t}}.

To conclude, we are going to prove that for a.e. t>0t>0 and for all κ[0,1]\kappa\in[0,1], Δ(t,κ)0\Delta(t,\kappa)\geq 0. Remember that by assumption, for a.e. t>0t>0, (ρ(y(t),t),ρ(y(t)+,t))𝒢y˙(t)(q(t))(\rho(y(t)-,t),\rho(y(t)+,t))\in\mathcal{G}_{\dot{y}(t)}(q(t)). The following table, in which we dropped the y˙(t)/q(t)\dot{y}(t)/q(t)-indexing, summarizes which values can take the difference Δ(t,κ)\Delta(t,\kappa) according to the position of κ\kappa with respect to the couple (ρ(y(t),t),ρ(y(t)+,t))(\rho(y(t)-,t),\rho(y(t)+,t)), which is simply denoted by (ρl,ρr)(\rho_{l},\rho_{r}). Note that the case marked by ×\mathbf{\times} does not happen.

κ\kappa (ρl,ρr)(\rho_{l},\rho_{r}) 𝒢1\in\mathcal{G}^{1} 𝒢2\in\mathcal{G}^{2} 𝒢3\in\mathcal{G}^{3}
κ<min{ρl,ρr}\kappa<\min\{\rho_{l},\rho_{r}\} 0 (κ,q(t))\mathcal{R}(\kappa,q(t)) 0
κ>max{ρl,ρr}\kappa>\max\{\rho_{l},\rho_{r}\} 0 (κ,q(t))\mathcal{R}(\kappa,q(t)) 0
κ\kappa between ρl\rho_{l} and ρr\rho_{r} 0 ×\mathbf{\times} 2(F(κ)F(ρl))+(κ,q(t))2(F(\kappa)-F(\rho_{l}))+\mathcal{R}(\kappa,q(t))

Clearly, Δ(t,κ)0\Delta(t,\kappa)\geq 0, which proves that 𝐈0\mathbf{I}\geq 0, hence ρ\rho satisfies (2.1). Moreover, by assumption, for a.e. t>0t>0, (ρ(y(t),t),ρ(y(t)+,t))𝒢y˙(t)(q(t))(\rho(y(t)-,t),\rho(y(t)+,t))\in\mathcal{G}_{\dot{y}(t)}(q(t)). This implies, in particular, that ρ\rho satisfies the flux constraint inequality (f(ρ)y˙(t)ρ)|x=y(t)q(t)\displaystyle{(f(\rho)-\dot{y}(t)\rho)_{|x=y(t)}\leq q(t)} in the a.e. sense. By Remark 2.1, ρ\rho satisfies (2.2) as well i.e. ρ\rho is an admissible entropy solution to (1.3).   \square

2.2 Uniqueness of 𝒢\mathcal{G}-entropy solutions

We now prove uniqueness using Definition 2.4.

Lemma 2.7 (Kato inequality).

Fix ρo,σo𝐋(,[0,1])\rho_{o},\sigma_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]), y𝐖𝐥𝐨𝐜1,(]0,+[,)y\in\mathbf{W}_{\mathbf{loc}}^{1,\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}), q𝐋𝐥𝐨𝐜(]0,+[,)q\in\mathbf{L}_{\mathbf{loc}}^{\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}). We denote by ρ\rho a 𝒢y˙(q)\mathcal{G}_{\dot{y}}(q)-entropy solution to (1.3). The same way, let σ\sigma be 𝒢y˙(r)\mathcal{G}_{\dot{y}}(r)-entropy solution to Problem (1.3) with initial datum σo\sigma_{o}. We suppose that q,rq,r satisfy (2.3). Then for all test functions φ𝐂𝐜(Ω¯,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}^{+}), we have

0+(|ρσ|tφ+Φ(ρ,σ)xφ)dxdt+|ρo(x)σo(x)|φ(x,0)dx\displaystyle\int_{0}^{+\infty}\int_{\mathbb{R}}\biggl{(}|\rho-\sigma|\partial_{t}\varphi+\Phi(\rho,\sigma)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{o}(x)-\sigma_{o}(x)|\varphi(x,0)\mathinner{\mathrm{d}{x}} (2.11)
+0+(Φy˙(t)(ρ(y(t)+,t),σ(y(t)+,t))Φy˙(t)(ρ(y(t),t),σ(y(t),t)))φ(y(t),t)dt0.\displaystyle+\int_{0}^{+\infty}\biggl{(}\Phi_{\dot{y}(t)}\left(\rho(y(t)+,t),\sigma(y(t)+,t)\right)-\Phi_{\dot{y}(t)}\left(\rho(y(t)-,t),\sigma(y(t)-,t)\right)\biggr{)}\varphi(y(t),t)\mathinner{\mathrm{d}{t}}\geq 0.

Proof.  Take ϕ=ϕ(x,t,χ,τ)𝐂𝐜(Ω¯2,+)\phi=\phi(x,t,\chi,\tau)\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega}^{2},\mathbb{R}^{+}) with support contained in the set (Ω¯\Γ)2\displaystyle{\left(\overline{\Omega}\backslash\Gamma\right)^{2}}. The classical method of doubling variables leads us to:

|ρ(x,t)σ(χ,τ)|(tϕ+τϕ)+Φ(ρ(x,t),σ(χ,τ))(xϕ+χϕ)dxdtdχdτ\displaystyle\iiiint|\rho(x,t)-\sigma(\chi,\tau)|(\partial_{t}\phi+\partial_{\tau}\phi)+\Phi(\rho(x,t),\sigma(\chi,\tau))(\partial_{x}\phi+\partial_{\chi}\phi)\ \mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\mathinner{\mathrm{d}{\chi}}\mathinner{\mathrm{d}{\tau}} (2.12)
+|ρo(x)σ(χ,τ)|ϕ(x,0,χ,τ)dxdχdτ+|ρ(x,t)σo(χ)|ϕ(x,t,χ,0)dxdtdχ0.\displaystyle+\iiint|\rho_{o}(x)-\sigma(\chi,\tau)|\phi(x,0,\chi,\tau)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{\chi}}\mathinner{\mathrm{d}{\tau}}+\iiint|\rho(x,t)-\sigma_{o}(\chi)|\phi(x,t,\chi,0)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\mathinner{\mathrm{d}{\chi}}\geq 0.

Again, a standard approximation argument allows us to apply (2.12) with the Lipschitz function

ϕn(x,t,χ,τ)=γn(x,t)φ(x+χ2,t+τ2)δn(xχ2)δn(tτ2),\phi_{n}(x,t,\chi,\tau)=\gamma_{n}(x,t)\varphi\left(\frac{x+\chi}{2},\frac{t+\tau}{2}\right)\delta_{n}\left(\frac{x-\chi}{2}\right)\delta_{n}\left(\frac{t-\tau}{2}\right),

where φ=φ(X,T)𝐂𝐜(Ω¯,+)\varphi=\varphi(X,T)\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}^{+}), (δn)n(\delta_{n})_{n} is a smooth approximation of the Dirac mass at the origin, and

γn(x,t)={0if|xy(t)|<1nn(|xy(t)|1n)if1n|xy(t)|2n1if|xy(t)|>2n.\gamma_{n}(x,t)=\left\{\begin{array}[]{ccc}0&\text{if}&\displaystyle{|x-y(t)|<\frac{1}{n}}\\[5.0pt] \displaystyle{n\left(|x-y(t)|-\frac{1}{n}\right)}&\text{if}&\displaystyle{\frac{1}{n}\leq|x-y(t)|\leq\frac{2}{n}}\\[5.0pt] 1&\text{if}&\displaystyle{|x-y(t)|>\frac{2}{n}}.\end{array}\right.

Using the fact that for a.e. t>0t>0,

tϕn+τϕn\displaystyle\partial_{t}\phi_{n}+\partial_{\tau}\phi_{n} =ny˙(t)sgn(xy(t))𝟏{1n<|xy(t)|<2n}φ(x+χ2,t+τ2)δn(xχ2)δn(tτ2)\displaystyle=-n\dot{y}(t)\mathop{\rm sgn}(x-y(t))\mathbf{1}_{\left\{\frac{1}{n}<|x-y(t)|<\frac{2}{n}\right\}}\varphi\left(\frac{x+\chi}{2},\frac{t+\tau}{2}\right)\delta_{n}\left(\frac{x-\chi}{2}\right)\delta_{n}\left(\frac{t-\tau}{2}\right)
+γn(x,t)Tφ(x+χ2,t+τ2)δn(xχ2)δn(tτ2)\displaystyle+\gamma_{n}(x,t)\partial_{T}\varphi\left(\frac{x+\chi}{2},\frac{t+\tau}{2}\right)\delta_{n}\left(\frac{x-\chi}{2}\right)\delta_{n}\left(\frac{t-\tau}{2}\right)
xϕn+χϕn\displaystyle\partial_{x}\phi_{n}+\partial_{\chi}\phi_{n} =nsgn(xy(t))𝟏{1n<|xy(t)|<2n}φ(x+χ2,t+τ2)δn(xχ2)δn(tτ2)\displaystyle=n\mathop{\rm sgn}(x-y(t))\mathbf{1}_{\left\{\frac{1}{n}<|x-y(t)|<\frac{2}{n}\right\}}\varphi\left(\frac{x+\chi}{2},\frac{t+\tau}{2}\right)\delta_{n}\left(\frac{x-\chi}{2}\right)\delta_{n}\left(\frac{t-\tau}{2}\right)
+γn(x,t)Xφ(x+χ2,t+τ2)δn(xχ2)δn(tτ2),\displaystyle+\gamma_{n}(x,t)\partial_{X}\varphi\left(\frac{x+\chi}{2},\frac{t+\tau}{2}\right)\delta_{n}\left(\frac{x-\chi}{2}\right)\delta_{n}\left(\frac{t-\tau}{2}\right),

we obtain:

|ρ(x,t)σ(χ,τ)|(tϕn+τϕn)dxdtdχdτ\displaystyle\iiiint|\rho(x,t)-\sigma(\chi,\tau)|(\partial_{t}\phi_{n}+\partial_{\tau}\phi_{n})\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\mathinner{\mathrm{d}{\chi}}\mathinner{\mathrm{d}{\tau}}
n+0+y˙(t)(|ρ(y(t)+,t)σ(y(t)+,t)||ρ(y(t),t)σ(y(t),t)|)φ(y(t),t)dt\displaystyle{\ \underset{n\to+\infty}{\longrightarrow}\ }-\int_{0}^{+\infty}\dot{y}(t)\biggl{(}\left|\rho(y(t)+,t)-\sigma(y(t)+,t)\right|-\left|\rho(y(t)-,t)-\sigma(y(t)-,t)\right|\biggr{)}\varphi(y(t),t)\mathinner{\mathrm{d}{t}}
+0+|ρ(x,t)σ(x,t)|Tφ(x,t)dxdt,\displaystyle+\int_{0}^{+\infty}\int_{\mathbb{R}}|\rho(x,t)-\sigma(x,t)|\partial_{T}\varphi(x,t)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}},

and

Φ(ρ(x,t),σ(χ,τ))(xϕn+χϕn)dxdtdχdτ\displaystyle\iiiint\Phi(\rho(x,t),\sigma(\chi,\tau))(\partial_{x}\phi_{n}+\partial_{\chi}\phi_{n})\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\mathinner{\mathrm{d}{\chi}}\mathinner{\mathrm{d}{\tau}}
n+0+(Φ(y(t)+,t),σ(y(t)+,t)Φ(ρ(y(t),t),σ(y(t),t)))φ(y(t),t)dt\displaystyle{\ \underset{n\to+\infty}{\longrightarrow}\ }\int_{0}^{+\infty}\biggl{(}\Phi(y(t)+,t),\sigma(y(t)+,t)-\Phi(\rho(y(t)-,t),\sigma(y(t)-,t))\biggr{)}\varphi(y(t),t)\mathinner{\mathrm{d}{t}}
+0+Φ(ρ(x,t),σ(x,t))Xφ(x,t)dxdt.\displaystyle+\int_{0}^{+\infty}\int_{\mathbb{R}}\Phi(\rho(x,t),\sigma(x,t))\partial_{X}\varphi(x,t)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}.

Finally, since

|ρo(x)σ(χ,τ)|ϕn(x,0,χ,τ)dxdχdτand|ρ(x,t)σo(χ)|ϕn(x,t,χ,0)dxdχdt\displaystyle\iiint|\rho_{o}(x)-\sigma(\chi,\tau)|\phi_{n}(x,0,\chi,\tau)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{\chi}}\mathinner{\mathrm{d}{\tau}}\;\;\text{and}\;\;\iiint|\rho(x,t)-\sigma_{o}(\chi)|\phi_{n}(x,t,\chi,0)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{\chi}}\mathinner{\mathrm{d}{t}}
both converge to12|ρo(x)σo(x)|φ(x,0)dx,\displaystyle\text{both converge to}\;\frac{1}{2}\int_{\mathbb{R}}|\rho_{o}(x)-\sigma_{o}(x)|\varphi(x,0)\mathinner{\mathrm{d}{x}},

we get (2.11) by assembling the above ingredients together.   \square

Theorem 2.8.

Fix ρo,σo𝐋(,[0,1])\rho_{o},\sigma_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]), y𝐖𝐥𝐨𝐜1,(]0,+[,)y\in\mathbf{W}_{\mathbf{loc}}^{1,\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}), q𝐋𝐥𝐨𝐜(]0,+[,)q\in\mathbf{L}_{\mathbf{loc}}^{\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}). We denote by ρ\rho a 𝒢y˙(q)\mathcal{G}_{\dot{y}}(q)-entropy solution to (1.3). The same way, let σ\sigma be 𝒢y˙(r)\mathcal{G}_{\dot{y}}(r)-entropy solution to Problem (1.3) with initial datum σo\sigma_{o}. We suppose that q,rq,r satisfy (2.3). Then for all T>0T>0, we have

ρ(T)σ(T)𝐋1()ρoσo𝐋1()+20T|q(t)r(t)|dt.\|\rho(T)-\sigma(T)\|_{\mathbf{L}^{1}(\mathbb{R})}\leq\|\rho_{o}-\sigma_{o}\|_{\mathbf{L}^{1}(\mathbb{R})}+2\int_{0}^{T}|q(t)-r(t)|\mathinner{\mathrm{d}{t}}. (2.13)

In particular, Problem (1.3) admits at most one solution.

Proof.  Fix T>0T>0, Ry𝐋(]0,T[)R\geq\|y\|_{\mathbf{L}^{\infty}(\mathopen{]}0,T\mathclose{[})} and set L:=f𝐋+y˙𝐋(]0,T[)L:=\|f^{\prime}\|_{\mathbf{L}^{\infty}}+\|\dot{y}\|_{\mathbf{L}^{\infty}(\mathopen{]}0,T\mathclose{[})}. Consider for all nn\in\mathbb{N}^{*} the function:

φn(x,t):=14(1ξn(tT))(1ξn(|x|R+L(tT))),\varphi_{n}(x,t):=\frac{1}{4}\left(1-\xi_{n}(t-T)\right)\left(1-\xi_{n}\left(|x|-R+L(t-T)\right)\right),

where (ξn)n(\xi_{n})_{n} is a smooth approximation of the sign function. The sequence (φn)n(\varphi_{n})_{n} is a smooth approximation of the characteristic function of the trapezoid

𝒯:={(x,t)Ω¯:t[0,T]and|x|RL(tT)}{(y(t),t):t[0,T]}.\mathcal{T}:=\left\{(x,t)\in\overline{\Omega}\;:\;t\in[0,T]\;\text{and}\;|x|\leq R-L(t-T)\right\}\supset\left\{(y(t),t)\;:\;t\in[0,T]\right\}.

Let us apply Kato inequality (2.11) with (φn)n(\varphi_{n})_{n}. For all nn\in\mathbb{N}, we have

0+|ρσ|tφndxdt\displaystyle\int_{0}^{+\infty}\int_{\mathbb{R}}|\rho-\sigma|\partial_{t}\varphi_{n}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}} =140+|ρσ|ξn(tT)(1ξn(|x|R+L(tT)))dxdt\displaystyle=-\frac{1}{4}\int_{0}^{+\infty}\int_{\mathbb{R}}|\rho-\sigma|\xi_{n}^{\prime}(t-T)\left(1-\xi_{n}\left(|x|-R+L(t-T)\right)\right)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
L40+|ρσ|(1ξn(tT))ξn(|x|R+L(tT))dxdt\displaystyle-\frac{L}{4}\int_{0}^{+\infty}\int_{\mathbb{R}}|\rho-\sigma|\left(1-\xi_{n}(t-T)\right)\xi_{n}^{\prime}\left(|x|-R+L(t-T)\right)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
n+|x|R|ρ(x,T)σ(x,T)|dxL0T|x|=RL(tT)|ρσ|dxdt.\displaystyle{\ \underset{n\to+\infty}{\longrightarrow}\ }-\int_{|x|\leq R}|\rho(x,T)-\sigma(x,T)|\mathinner{\mathrm{d}{x}}-L\int_{0}^{T}\int_{|x|=R-L(t-T)}|\rho-\sigma|\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}.

Then,

0+Φ(ρ,σ)xφndxdt\displaystyle\int_{0}^{+\infty}\int_{\mathbb{R}}\Phi(\rho,\sigma)\partial_{x}\varphi_{n}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}} =140+Φ(ρ,σ)(1ξn(tT))sgn(x)ξn(|x|R+L(tT))dxdt\displaystyle=-\frac{1}{4}\int_{0}^{+\infty}\int_{\mathbb{R}}\Phi(\rho,\sigma)\left(1-\xi_{n}(t-T)\right)\mathop{\rm sgn}(x)\xi_{n}^{\prime}\left(|x|-R+L(t-T)\right)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
n+0T|x|=RL(tT)Φ(ρ,σ)sgn(x)dxdt.\displaystyle{\ \underset{n\to+\infty}{\longrightarrow}\ }-\int_{0}^{T}\int_{|x|=R-L(t-T)}\Phi(\rho,\sigma)\mathop{\rm sgn}(x)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}.

Finally, we have

|ρo(x)σo(x)|φn(x,0)dxn+|x|R+LT|ρo(x)σo(x)|dx\int_{\mathbb{R}}|\rho_{o}(x)-\sigma_{o}(x)|\varphi_{n}(x,0)\mathinner{\mathrm{d}{x}}{\ \underset{n\to+\infty}{\longrightarrow}\ }\int_{|x|\leq R+LT}|\rho_{o}(x)-\sigma_{o}(x)|\mathinner{\mathrm{d}{x}}

Remark also that the choices of RR and LL imply that for all t>0t>0,

φn(y(t),t)n+ 1.\varphi_{n}(y(t),t){\ \underset{n\to+\infty}{\longrightarrow}\ }1.

Assembling the previous limits together, we get:

|x|R|ρ(x,T)σ(x,T)|dx+|x|R+LT|ρo(x)σo(x)|dx\displaystyle-\int_{|x|\leq R}|\rho(x,T)-\sigma(x,T)|\mathinner{\mathrm{d}{x}}+\int_{|x|\leq R+LT}|\rho_{o}(x)-\sigma_{o}(x)|\mathinner{\mathrm{d}{x}}
0T|x|=RL(tT)(L|ρσ|+Φ(ρ,σ)sgn(x))dxdt\displaystyle-\int_{0}^{T}\int_{|x|=R-L(t-T)}\left(L|\rho-\sigma|+\Phi(\rho,\sigma)\mathop{\rm sgn}(x)\right)\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
+0T(Φy˙(t)(ρ(y(t)+,t),σ(y(t)+,t))Φy˙(t)(ρ(y(t),t),σ(y(t),t)))dt0.\displaystyle+\int_{0}^{T}\biggl{(}\Phi_{\dot{y}(t)}\left(\rho(y(t)+,t),\sigma(y(t)+,t)\right)-\Phi_{\dot{y}(t)}\left(\rho(y(t)-,t),\sigma(y(t)-,t)\right)\biggr{)}\mathinner{\mathrm{d}{t}}\geq 0.

Note that for all ρ,σ[0,1]\rho,\sigma\in[0,1] and for all xx\in\mathbb{R},

L|ρσ|+Φ(ρ,σ)sgn(x)L|ρσ||f(ρ)f(σ)|(Lf𝐋)|ρσ|0.L|\rho-\sigma|+\Phi(\rho,\sigma)\mathop{\rm sgn}(x)\geq L|\rho-\sigma|-|f(\rho)-f(\sigma)|\geq(L-\|f^{\prime}\|_{\mathbf{L}^{\infty}})|\rho-\sigma|\geq 0.

Consequently, we have shown that

|x|R|ρ(x,T)σ(x,T)|dx\displaystyle\int_{|x|\leq R}|\rho(x,T)-\sigma(x,T)|\mathinner{\mathrm{d}{x}} |x|R+LT|ρo(x)σo(x)|dx\displaystyle\leq\int_{|x|\leq R+LT}|\rho_{o}(x)-\sigma_{o}(x)|\mathinner{\mathrm{d}{x}}
+0T(Φy˙(t)(ρ(y(t)+,t),σ(y(t)+,t))Φy˙(t)(ρ(y(t),t),σ(y(t),t))Δ(t))dt.\displaystyle+\int_{0}^{T}\biggl{(}\underbrace{\Phi_{\dot{y}(t)}\left(\rho(y(t)+,t),\sigma(y(t)+,t)\right)-\Phi_{\dot{y}(t)}\left(\rho(y(t)-,t),\sigma(y(t)-,t)\right)}_{\Delta(t)}\biggr{)}\mathinner{\mathrm{d}{t}}.

What is left to do is to take the limit when R+R\to+\infty and to estimate the last two terms of the right-hand side of the previous inequality. The following table, in which we dropped the tt-indexing, summarizes which values can take the difference Δ(t)\Delta(t) according to which parts of their respective germs the couples (ρ(y(t),t),ρ(y(t)+,t))(\rho(y(t)-,t),\rho(y(t)+,t)) and (σ(y(t),t),σ(y(t)+,t))(\sigma(y(t)-,t),\sigma(y(t)+,t)), respectively denoted by (ρl,ρr)(\rho_{l},\rho_{r}) and (σl,σr)(\sigma_{l},\sigma_{r}) belong to.

(σl,σr)(\sigma_{l},\sigma_{r}) (ρl,ρr)(\rho_{l},\rho_{r}) 𝒢y˙1(q)\in\mathcal{G}^{1}_{\dot{y}}(q) 𝒢y˙2(q)\in\mathcal{G}^{2}_{\dot{y}}(q) 𝒢y˙3(q)\in\mathcal{G}^{3}_{\dot{y}}(q)
𝒢y˙1(r)\in\mathcal{G}^{1}_{\dot{y}}(r) 2(qr)2(q-r) 0 or 2(Fy˙(ρl)r)2(F_{\dot{y}}(\rho_{l})-r) 2(Fy˙(ρl)r)2(F_{\dot{y}}(\rho_{l})-r)
𝒢y˙2(r)\in\mathcal{G}^{2}_{\dot{y}}(r) 0 0 0\leq 0
𝒢y˙3(r)\in\mathcal{G}^{3}_{\dot{y}}(r) 2(Fy˙(σl)q)2(F_{\dot{y}}(\sigma_{l})-q) 0\leq 0 0\leq 0

We clearly see the bound Δ(t)2|q(t)r(t)|\displaystyle{\Delta(t)\leq 2|q(t)-r(t)|}, which leads us to (2.13), which clearly implies uniqueness. This concludes the proof.   \square

3 Existence for the single trajectory problem

We build a simple finite volume scheme and prove its convergence to an admissible entropy solution to (1.3). From now on, we denote by

ab:=max{a,b}andab:=min{a,b}.a\vee b:=\max\{a,b\}\quad\text{and}\quad a\wedge b:=\min\{a,b\}.

Fix ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]) and y𝐖𝐥𝐨𝐜1,(]0,+[,)y\in\mathbf{W}_{\mathbf{loc}}^{1,\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}).

3.1 Adapted mesh and definition of the scheme

We start by defining the sequence of approximate slopes:

n,sn=1Δttntn+1y˙(t)dt;t0,sΔ(t)=nsn𝟏[tn,tn+1[(t),\forall n\in\mathbb{N},\;s^{n}=\frac{1}{\Delta t}\int_{t^{n}}^{t^{n+1}}\dot{y}(t)\mathinner{\mathrm{d}{t}};\quad\forall t\geq 0,\;s_{\Delta}(t)=\sum_{n\in\mathbb{N}}s^{n}\mathbf{1}_{[t^{n},t^{n+1}\mathclose{[}}(t),

and the sequence of approximate trajectories:

t0,yΔ(t)=y(0)+0tsΔ(τ)dτ;n,yn=yΔ(tn).\forall t\geq 0,\;y_{\Delta}(t)=y(0)+\int_{0}^{t}s_{\Delta}(\tau)\mathinner{\mathrm{d}{\tau}};\quad\forall n\in\mathbb{N},\;y^{n}=y_{\Delta}(t^{n}).

Since (sΔ)Δ(s_{\Delta})_{\Delta} converges y˙\dot{y} in 𝐋𝐥𝐨𝐜1(]0,+[,)\mathbf{L}_{\mathbf{loc}}^{1}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}), (yΔ)Δ(y_{\Delta})_{\Delta} converges to yy in 𝐋𝐥𝐨𝐜(]0,+[,)\mathbf{L}_{\mathbf{loc}}^{\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}).

The same way, we define (qΔ)Δ(q_{\Delta})_{\Delta}, the sequence of approximate constraints:

qΔ(t)=nqn𝟏[tn,tn+1[(t);qn=1Δttntn+1q(t)dt,q_{\Delta}(t)=\sum_{n\in\mathbb{N}}q^{n}\mathbf{1}_{[t^{n},t^{n+1}\mathclose{[}}(t);\quad q^{n}=\frac{1}{\Delta t}\int_{t^{n}}^{t^{n+1}}q(t)\mathinner{\mathrm{d}{t}},

which converges to qq in 𝐋𝐥𝐨𝐜1(]0,+[,)\mathbf{L}_{\mathbf{loc}}^{1}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}).

Remark 3.1.

With our choices, from (2.4), we deduce that

n,sn+qn=1Δttntn+1(y˙(t)+q(t))dt0.\forall n\in\mathbb{N},\quad s^{n}+q^{n}=\frac{1}{\Delta t}\int_{t^{n}}^{t^{n+1}}\left(\dot{y}(t)+q(t)\right)\mathinner{\mathrm{d}{t}}\geq 0. (3.1)

This fact will come in handy in the proof of stability for the scheme.

Fix now T>0T>0 and a spatial mesh size Δx>0\Delta x>0 with λ=Δt/Δx\lambda=\Delta t/\Delta x fixed, verifying the CFL condition

2(f𝐋+y˙𝐋(]0,T[):=L)λ1.2\left(\underbrace{\|f^{\prime}\|_{\mathbf{L}^{\infty}}+\|\dot{y}\|_{\mathbf{L}^{\infty}(\mathopen{]}0,T\mathclose{[})}}_{:=L}\right)\lambda\leq 1. (3.2)

For all nn\in\mathbb{N}, there exists a unique index jnj_{n}\in\mathbb{Z} such that yn]xjn,xjn+1[y^{n}\in\mathopen{]}x_{j_{n}},x_{j_{n}+1}\mathclose{[}, see Figure 2. Introduce the sequence (χjn)j(\chi_{j}^{n})_{j\in\mathbb{Z}} defined by

χjn={xjifjjn1ynifj=jnxj+1ifjjn+1.\chi_{j}^{n}=\left\{\begin{array}[]{ccl}x_{j}&\text{if}&j\leq j_{n}-1\\[5.0pt] y^{n}&\text{if}&j=j_{n}\\[5.0pt] x_{j+1}&\text{if}&j\geq j_{n}+1.\end{array}\right.

We define the cell grids:

Ω¯=nj𝒫j+1/2n,\overline{\Omega}=\bigcup_{n\in\mathbb{N}}\bigcup_{j\in\mathbb{Z}}\mathcal{P}_{j+1/2}^{n},

where for all nn\in\mathbb{N} and jj\in\mathbb{Z}, 𝒫j+1/2n\mathcal{P}_{j+1/2}^{n} is the rectangle ]χjn,χj+1n[×[tn,tn+1[\mathopen{]}\chi_{j}^{n},\chi_{j+1}^{n}\mathclose{[}\times[t^{n},t^{n+1}\mathclose{[} if jjn2j\leq j_{n}-2, one of the parallelograms represented in Figure 2 if j{jn1,jn}j\in\{j_{n}-1,j_{n}\} and the rectangle ]χj+1n,χj+2n[×[tn,tn+1[\mathopen{]}\chi_{j+1}^{n},\chi_{j+2}^{n}\mathclose{[}\times[t^{n},t^{n+1}\mathclose{[} if jjn+1j\geq j_{n}+1.

Refer to caption
Figure 2: Illustration of the modification to the mesh.

We start by discretizing the initial datum ρo\rho_{o} with (ρj+1/20)j\left(\rho_{j+1/2}^{0}\right)_{j} where for all jj\in\mathbb{Z}, ρj+1/20\displaystyle{\rho_{j+1/2}^{0}} is its mean value on the cell ]χj0,χj+10[\mathopen{]}\chi_{j}^{0},\chi_{j+1}^{0}\mathclose{[}. Clearly, for this choice, we have:

ρj+1/20[0,1]andρΔ0=jρj+1/20𝟏]χj0,χj+10[Δx0ρoin𝐋𝐥𝐨𝐜1().\rho_{j+1/2}^{0}\in[0,1]\quad\text{and}\quad\rho_{\Delta}^{0}=\sum_{j\in\mathbb{Z}}\rho_{j+1/2}^{0}\mathbf{1}_{\mathopen{]}\chi_{j}^{0},\chi_{j+1}^{0}\mathclose{[}}{\ \underset{\Delta x\to 0}{\longrightarrow}\ }\rho_{o}\;\text{in}\;\mathbf{L}_{\mathbf{loc}}^{1}(\mathbb{R}).

Let us denote by 𝐄𝐎=𝐄𝐎(a,b)\mathbf{EO}=\mathbf{EO}(a,b) the Engquist-Osher numerical flux associated with ff and for all ss\in\mathbb{R}, 𝐆𝐨𝐝s=𝐆𝐨𝐝s(u,v)\mathbf{God}^{s}=\mathbf{God}^{s}(u,v) be the Godunov flux associated with ρf(ρ)sρ\rho\mapsto f(\rho)-s\rho.

Fix nn\in\mathbb{N}. To simplify the reading, we introduce the notations:

j,fjn:=𝐄𝐎(ρj1/2n,ρj+1/2n)andfintn:=𝐆𝐨𝐝sn(ρjn1/2n,ρjn+1/2n)qn.\forall j\in\mathbb{Z},\quad f_{j}^{n}:=\mathbf{EO}\left(\rho_{j-1/2}^{n},\rho_{j+1/2}^{n}\right)\quad\text{and}\quad f_{int}^{n}:=\mathbf{God}^{s^{n}}\left(\rho_{j_{n}-1/2}^{n},\rho_{j_{n}+1/2}^{n}\right)\wedge q^{n}. (3.3)

We now proceed to the definition of the scheme. It comes from a discretization of the conservation law written in each volume control 𝒫j+1/2n\mathcal{P}_{j+1/2}^{n} (nn\in\mathbb{N}, jj\in\mathbb{Z}). Away from the trajectory/constraint, it is the standard 33-point marching formula and when j{jn1,jn}j\in\{j_{n}-1,j_{n}\}, we have to deal with both the constraint and the interface which is not vertical. Three cases have to be considered when describing the marching formula of the scheme, but we really give the details for only one of them.

Case 1: jn+1=jn+1j_{n+1}=j_{n}+1. This means that the line joining (yn,tn)(y^{n},t^{n}) and (yn+1,tn+1)(y^{n+1},t^{n+1}) crosses the line x=xjn+1x=x_{j_{n}+1}, see Figure 2. If j{jn1,jn}j\notin\{j_{n}-1,j_{n}\}, the conservation written in the rectangle 𝒫j+1/2n\mathcal{P}_{j+1/2}^{n} is given by the standard equation:

(ρj+1/2n+1ρj+1/2n)Δx+(fj+1nfjn)Δt=0.\left(\rho_{j+1/2}^{n+1}-\rho_{j+1/2}^{n}\right)\Delta x+(f_{j+1}^{n}-f_{j}^{n})\Delta t=0. (3.4)

From the conservation in the cell 𝒫jn1/2n\mathcal{P}_{j_{n}-1/2}^{n}, we set:

ρjn+11/2n+1(yn+1χjn+12n+1)ρjn1/2n(ynχjn1n)+(fintnfjn1n)Δt=0.\rho_{j_{n+1}-1/2}^{n+1}\left(y^{n+1}-\chi_{j_{n+1}-2}^{n+1}\right)-\rho_{j_{n}-1/2}^{n}\left(y^{n}-\chi_{j_{n}-1}^{n}\right)+(f_{int}^{n}-f_{j_{n}-1}^{n})\Delta t=0. (3.5)

This formula corresponds to the choice of putting the same value for ρΔ\rho_{\Delta} on ]χjn+12n+1,χjn+11n+1[\mathopen{]}\chi_{j_{n+1}-2}^{n+1},\chi_{j_{n+1}-1}^{n+1}\mathclose{[} and on ]χjn+11n+1,yn+1[\mathopen{]}\chi_{j_{n+1}-1}^{n+1},y^{n+1}\mathclose{[} at time t=tn+1t=t^{n+1}, i.e. ρjn+13/2n+1=ρjn+11/2n+1\rho_{j_{n+1}-3/2}^{n+1}=\rho_{j_{n+1}-1/2}^{n+1}. In the cell 𝒫jn+1/2n\mathcal{P}_{j_{n}+1/2}^{n}, the conservation takes the form:

ρjn+1+1/2n+1(χjn+1+1n+1yn+1)ρjn+1/2n(χjn+1nyn)ρjn+3/2nΔx+(fjn+2nfintn)Δt=0.\rho_{j_{n+1}+1/2}^{n+1}\left(\chi_{j_{n+1}+1}^{n+1}-y^{n+1}\right)-\rho_{j_{n}+1/2}^{n}\left(\chi_{j_{n}+1}^{n}-y^{n}\right)-\rho_{j_{n}+3/2}^{n}\Delta x+(f_{j_{n}+2}^{n}-f_{int}^{n})\Delta t=0. (3.6)

Let us introduce the two functions

Hjn1n(u,v,w):=v(ynχjn1n)(𝐆𝐨𝐝sn(v,w)qn𝐄𝐎(u,v))Δtyn+1χjn+12n+1H_{j_{n}-1}^{n}(u,v,w):=\frac{v(y^{n}-\chi_{j_{n}-1}^{n})-\left(\mathbf{God}^{s^{n}}(v,w)\wedge q^{n}-\mathbf{EO}(u,v)\right)\Delta t}{y^{n+1}-\chi_{j_{n+1}-2}^{n+1}}

and

Hjnn(u,v,w,z):=v(χjn+1nyn)+wΔx(𝐄𝐎(w,z)𝐆𝐨𝐝sn(u,v)qn)Δtχjn+1+1n+1yn+1,H_{j_{n}}^{n}(u,v,w,z):=\frac{v(\chi_{j_{n}+1}^{n}-y^{n})+w\Delta x-\left(\mathbf{EO}(w,z)-\mathbf{God}^{s^{n}}(u,v)\wedge q^{n}\right)\Delta t}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}},

so that

{ρjn+11/2n+1=Hjn1n(ρjn3/2n,ρjn1/2n,ρjn+1/2n)ρjn+1+1/2n+1=Hjnn(ρjn1/2n,ρjn+1/2n,ρjn+3/2n,ρjn+5/2n).\left\{\begin{aligned} \rho_{j_{n+1}-1/2}^{n+1}&=H_{j_{n}-1}^{n}(\rho_{j_{n}-3/2}^{n},\rho_{j_{n}-1/2}^{n},\rho_{j_{n}+1/2}^{n})\\[5.0pt] \rho_{j_{n+1}+1/2}^{n+1}&=H_{j_{n}}^{n}(\rho_{j_{n}-1/2}^{n},\rho_{j_{n}+1/2}^{n},\rho_{j_{n}+3/2}^{n},\rho_{j_{n}+5/2}^{n}).\end{aligned}\right. (3.7)

The key point in the proofs of the next section (stability and discrete entropy inequalities) is that the functions Hjn1nH_{j_{n}-1}^{n} and HjnnH_{j_{n}}^{n} are nondecreasing with respect to their arguments, therefore the modification in (3.3) did not affect the monotonicity of the resulting scheme (3.4) – (3.6).

Finally, the approximate solution ρΔ\rho_{\Delta} is defined almost everywhere on Ω¯\overline{\Omega}:

ρΔ=n(jjnρj+1/2n𝟏𝒫j+1/2n+jjn+1ρj+3/2n𝟏𝒫j+1/2n).\rho_{\Delta}=\sum_{n\in\mathbb{N}}\left(\sum_{j\leq j_{n}}\rho_{j+1/2}^{n}\mathbf{1}_{\mathcal{P}_{j+1/2}^{n}}+\sum_{j\geq j_{n}+1}\rho_{j+3/2}^{n}\mathbf{1}_{\mathcal{P}_{j+1/2}^{n}}\right).

The other cases (jn+1=jnj_{n+1}=j_{n} or jn+1=jn1j_{n+1}=j_{n}-1) follow from similar geometric considerations. Note that in the context of traffic dynamics, yy would be the trajectory of a stationary or a forward moving obstacle and therefore, we should have y˙0\dot{y}\geq 0. This implies that for all nn\in\mathbb{N}, either jn+1=jnj_{n+1}=j_{n} or jn+1=jn+1j_{n+1}=j_{n}+1. This is why we will focus on the case presented in Figure 2.

3.2 Stability and discrete entropy inequalities

Proposition 3.1 (𝐋\mathbf{L}^{\infty} stability).

Under the CFL condition (3.2), the scheme (3.4) – (3.6) is stable:

n,j,ρj+1/2n[0,1].\forall n\in\mathbb{N},\;\forall j\in\mathbb{Z},\quad\rho_{j+1/2}^{n}\in[0,1]. (3.8)

Proof.Monotonicity. Fix nn\in\mathbb{N}. Clearly, the expression (3.4) allows us to express ρn+1\rho^{n+1} as a function of three values of ρn\rho^{n} in a nondrecreasing way, see the [13, Chapter 5] for instance. We now verify that the functions Hjn1nH_{j_{n}-1}^{n} and HjnnH_{j_{n}}^{n} are also nondecreasing. Let us detail the proof for HjnnH_{j_{n}}^{n}. Recall that HjnnH_{j_{n}}^{n} is Lipschitz continuous by construction, therefore we can study its monotonicity in terms of its a.e. derivatives. Making use of both the CFL condition (3.2) and of the monotonicity of 𝐄𝐎\mathbf{EO} and 𝐆𝐨𝐝sn\mathbf{God}^{s^{n}}, for a.e. u,v,w,z[0,1]u,v,w,z\in[0,1], we have

1Hjnn(u,v,w,z)\displaystyle\partial_{1}H_{j_{n}}^{n}(u,v,w,z) =12Δtχjn+1+1n+1yn+1𝐆𝐨𝐝sna(u,v)(1sgn(𝐆𝐨𝐝sn(u,v)qn))0,\displaystyle=\frac{1}{2}\frac{\Delta t}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}\frac{\partial\mathbf{God}^{s^{n}}}{\partial a}(u,v)(1-\mathop{\rm sgn}(\mathbf{God}^{s^{n}}(u,v)-q^{n}))\geq 0,
2Hjnn(u,v,w,z)\displaystyle\partial_{2}H_{j_{n}}^{n}(u,v,w,z) =χjn+1nynχjn+1+1n+1yn+1+Δtχjn+1+1n+1yn+1𝐆𝐨𝐝snb(u,v)(1sgn(𝐆𝐨𝐝sn(u,v)qn))2\displaystyle=\frac{\chi_{j_{n}+1}^{n}-y^{n}}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}+\frac{\Delta t}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}\frac{\partial\mathbf{God}^{s^{n}}}{\partial b}(u,v)\frac{(1-\mathop{\rm sgn}(\mathbf{God}^{s^{n}}(u,v)-q^{n}))}{2}
χjn+1n(yn+LΔt)χjn+1+1n+1yn+1χjn+1n(yn+Δx2)χjn+1+1n+1yn+10,\displaystyle\geq\frac{\chi_{j_{n}+1}^{n}-(y^{n}+L\Delta t)}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}\geq\frac{\chi_{j_{n}+1}^{n}-\left(y^{n}+\frac{\Delta x}{2}\right)}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}\geq 0,
3Hjnn(u,v,w,z)\displaystyle\partial_{3}H_{j_{n}}^{n}(u,v,w,z) =Δxχjn+1+1n+1yn+1Δtχjn+1+1n+1yn+1𝐄𝐎a(w,z)\displaystyle=\frac{\Delta x}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}-\frac{\Delta t}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}\frac{\partial\mathbf{EO}}{\partial a}(w,z)
ΔxLΔtχjn+1+1n+1yn+1ΔxΔx/2χjn+1+1n+1yn+10,\displaystyle\geq\frac{\Delta x-L\Delta t}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}\geq\frac{\Delta x-\Delta x/2}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}\geq 0,
4Hjnn(u,v,w,z)\displaystyle\partial_{4}H_{j_{n}}^{n}(u,v,w,z) =Δtχjn+1+1n+1yn+1𝐄𝐎b(w,z)0,\displaystyle=-\frac{\Delta t}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}\frac{\partial\mathbf{EO}}{\partial b}(w,z)\geq 0,

proving the monotonicity of HjnnH_{j_{n}}^{n}. Similar computations show that Hjn1nH_{j_{n}-1}^{n} is nondecreasing with respect to its arguments as well.
Stability. We now turn to the proof of (3.8), which is done by induction on nn. If n=0n=0, it is verified by definition of (ρj+1/20)j\left(\rho_{j+1/2}^{0}\right)_{j}. Suppose now that (3.8) holds for some integer n0n\geq 0 and let us show that it still holds for n+1n+1. Remark that 0 and 11 are stationary solutions to the scheme. It is obviously true in the case (3.4). The definitions of Hjn1nH_{j_{n}-1}^{n} and HjnnH_{j_{n}}^{n} do not change this fact. For instance, Hjn1n(0,0,0)=0H_{j_{n}-1}^{n}(0,0,0)=0 since qn0q^{n}\geq 0 and because of (3.1), we also have:

Hjn1n(1,1,1)=(ynχjn1n)((sn)qn)Δtyn+1χjn+12n+1=(ynχjn1n)+snΔtyn+1χjn+12n+1=1.H_{j_{n}-1}^{n}(1,1,1)=\frac{(y^{n}-\chi_{j_{n}-1}^{n})-\left((-s^{n})\wedge q^{n}\right)\Delta t}{y^{n+1}-\chi_{j_{n+1}-2}^{n+1}}=\frac{(y^{n}-\chi_{j_{n}-1}^{n})+s^{n}\Delta t}{y^{n+1}-\chi_{j_{n+1}-2}^{n+1}}=1.

Similar computations would ensure that it holds also for HjnnH_{j_{n}}^{n}. Using now the monotonicity of Hjn1nH_{j_{n}-1}^{n} for instance, we deduce that

0=Hjn1n(0,0,0)\displaystyle 0=H_{j_{n}-1}^{n}(0,0,0) Hjn1n(ρjn3/2n,ρjn1/2n,ρjn+1/2n)\displaystyle\leq H_{j_{n}-1}^{n}(\rho_{j_{n}-3/2}^{n},\rho_{j_{n}-1/2}^{n},\rho_{j_{n}+1/2}^{n})
=ρjn+11/2n+1\displaystyle=\rho_{j_{n+1}-1/2}^{n+1}
=Hjn1n(ρjn3/2n,ρjn1/2n,ρjn+1/2n)Hjn1n(1,1,1)=1,\displaystyle=H_{j_{n}-1}^{n}(\rho_{j_{n}-3/2}^{n},\rho_{j_{n}-1/2}^{n},\rho_{j_{n}+1/2}^{n})\leq H_{j_{n}-1}^{n}(1,1,1)=1,

which concludes the induction argument. The remaining cases follow from similar computations.   \square

Corollary 3.2 (Discrete entropy inequalities).

Fix nn\in\mathbb{N}, j\{jn+12}j\in\mathbb{Z}\backslash\{j_{n+1}-2\} and κ[0,1]\kappa\in[0,1]. Then the numerical scheme (3.4) – (3.6) fulfills the following discrete entropy inequalities:

|ρj+1/2n+1κ|(χj+1n+1χjn+1){|ρj+1/2nκ|(χj+1nχjn)(Φj+1nΦjn)Δtifj{jn+11,jn+1}|ρjn+11/2n+1κ|Δx+|ρjn1/2nκ|(χjnnχjn1n)(ΦintnΦjn1n)Δt+12sn(κ,qn)Δtifj=jn+11|ρjn+1/2nκ|(χjn+1nχjnn)+|ρjn+3/2nκ|Δx(Φjn+2nΦintn)Δt+12sn(κ,qn)Δtifj=jn+1,\displaystyle|\rho_{j+1/2}^{n+1}-\kappa|(\chi_{j+1}^{n+1}-\chi_{j}^{n+1})\leq\left\{\begin{array}[]{lll}&|\rho_{j+1/2}^{n}-\kappa|(\chi_{j+1}^{n}-\chi_{j}^{n})-\left(\Phi_{j+1}^{n}-\Phi_{j}^{n}\right)\Delta t&\text{if}\;j\notin\{j_{n+1}-1,j_{n+1}\}\\[10.0pt] &-|\rho_{j_{n+1}-1/2}^{n+1}-\kappa|\Delta x+|\rho_{j_{n}-1/2}^{n}-\kappa|(\chi_{j_{n}}^{n}-\chi_{j_{n}-1}^{n})&\\[5.0pt] &-\left(\Phi_{int}^{n}-\Phi_{j_{n}-1}^{n}\right)\Delta t+\frac{1}{2}\mathcal{R}_{s^{n}}(\kappa,q^{n})\Delta t&\text{if}\;j=j_{n+1}-1\\[10.0pt] &|\rho_{j_{n}+1/2}^{n}-\kappa|(\chi_{j_{n}+1}^{n}-\chi_{j_{n}}^{n})+|\rho_{j_{n}+3/2}^{n}-\kappa|\Delta x\\[5.0pt] &-\left(\Phi_{j_{n}+2}^{n}-\Phi_{int}^{n}\right)\Delta t+\frac{1}{2}\mathcal{R}_{s^{n}}(\kappa,q^{n})\Delta t&\text{if}\;j=j_{n+1},\end{array}\right. (3.9)

where Φjn\Phi_{j}^{n} and Φintn\Phi_{int}^{n} denote the numerical entropy fluxes:

Φjn:=𝐄𝐎(ρj1/2nκ,ρj+1/2nκ)𝐄𝐎(ρj1/2nκ,ρj+1/2nκ);\displaystyle\Phi_{j}^{n}:=\mathbf{EO}(\rho_{j-1/2}^{n}\vee\kappa,\rho_{j+1/2}^{n}\vee\kappa)-\mathbf{EO}(\rho_{j-1/2}^{n}\wedge\kappa,\rho_{j+1/2}^{n}\wedge\kappa);
Φintn:=min{𝐆𝐨𝐝sn(ρjn1/2nκ,ρjn+1/2nκ),qn}min{𝐆𝐨𝐝sn(ρjn1/2nκ,ρjn+1/2nκ),qn}\displaystyle\Phi_{int}^{n}:=\min\{\mathbf{God}^{s^{n}}(\rho_{j_{n}-1/2}^{n}\vee\kappa,\rho_{j_{n}+1/2}^{n}\vee\kappa),q^{n}\}-\min\{\mathbf{God}^{s^{n}}(\rho_{j_{n}-1/2}^{n}\wedge\kappa,\rho_{j_{n}+1/2}^{n}\wedge\kappa),q^{n}\}

Proof.  This result is mostly a consequence of the scheme monotonicity. When the interface/constraint does not enter the calculations i.e. when j{jn+11,jn+1}j\notin\{j_{n+1}-1,j_{n+1}\}, the proof follows [13, Lemma 5.4]. The key point is not only the monotonicity, but also the fact that in the classical case, all the constants states κ[0,1]\kappa\in[0,1] are stationary solutions of the scheme. This observation does not hold when the constraint enters the calculations. Suppose for example that j=jn+1j=j_{n+1} (which corresponds to the function HjnnH_{j_{n}}^{n}). Here, we have

Hjnn(κ,κ,κ,κ)\displaystyle H_{j_{n}}^{n}(\kappa,\kappa,\kappa,\kappa) =κ(χjn+1nyn)+κΔx(f(κ)(f(κ)snκ)qn)Δtχjn+1+1n+1yn+1\displaystyle=\frac{\kappa(\chi_{j_{n}+1}^{n}-y^{n})+\kappa\Delta x-\left(f(\kappa)-(f(\kappa)-s^{n}\kappa)\wedge q^{n}\right)\Delta t}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}
=(χjn+2nynsnΔt)κχjn+1+1n+1yn+1Δt2(χjn+1+1n+1yn+1)sn(κ,qn)\displaystyle=\frac{(\chi_{j_{n}+2}^{n}-y^{n}-s^{n}\Delta t)\kappa}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}-\frac{\Delta t}{2(\chi_{j_{n+1}+1}^{n+1}-y^{n+1})}\mathcal{R}_{s^{n}}(\kappa,q^{n})
=κΔt2(χjn+1+1n+1yn+1)sn(κ,qn),\displaystyle=\kappa-\frac{\Delta t}{2(\chi_{j_{n+1}+1}^{n+1}-y^{n+1})}\mathcal{R}_{s^{n}}(\kappa,q^{n}),

and it implies:

Hjnn(ρjn1/2nκ,ρjn+1/2nκ,ρjn+3/2nκ,ρjn+5/2nκ)\displaystyle H_{j_{n}}^{n}(\rho_{j_{n}-1/2}^{n}\wedge\kappa,\rho_{j_{n}+1/2}^{n}\wedge\kappa,\rho_{j_{n}+3/2}^{n}\wedge\kappa,\rho_{j_{n}+5/2}^{n}\wedge\kappa)
ρjn+1+1/2n+1κ,ρjn+1+1/2n+1κ\displaystyle\leq\rho_{j_{n+1}+1/2}^{n+1}\wedge\kappa,\ \rho_{j_{n+1}+1/2}^{n+1}\vee\kappa
Hjnn(ρjn1/2nκ,ρjn+1/2nκ,ρjn+3/2nκ,ρjn+5/2nκ)+Δt2(χjn+1+1n+1yn+1)sn(κ,qn).\displaystyle\leq H_{j_{n}}^{n}(\rho_{j_{n}-1/2}^{n}\vee\kappa,\rho_{j_{n}+1/2}^{n}\vee\kappa,\rho_{j_{n}+3/2}^{n}\vee\kappa,\rho_{j_{n}+5/2}^{n}\vee\kappa)+\frac{\Delta t}{2(\chi_{j_{n+1}+1}^{n+1}-y^{n+1})}\mathcal{R}_{s^{n}}(\kappa,q^{n}).

We deduce:

|ρjn+1+1/2n+1κ|\displaystyle|\rho_{j_{n+1}+1/2}^{n+1}-\kappa| =ρjn+1+1/2n+1κρjn+1+1/2n+1κ\displaystyle=\rho_{j_{n+1}+1/2}^{n+1}\vee\kappa-\rho_{j_{n+1}+1/2}^{n+1}\wedge\kappa
Hjnn(ρjn1/2nκ,ρjn+1/2nκ,ρjn+3/2nκ,ρjn+5/2nκ)\displaystyle\leq H_{j_{n}}^{n}(\rho_{j_{n}-1/2}^{n}\vee\kappa,\rho_{j_{n}+1/2}^{n}\vee\kappa,\rho_{j_{n}+3/2}^{n}\vee\kappa,\rho_{j_{n}+5/2}^{n}\vee\kappa)
Hjnn(ρjn1/2nκ,ρjn+1/2nκ,ρjn+3/2nκ,ρjn+5/2nκ)+Δt2(χjn+1+1n+1yn+1)sn(κ,qn)\displaystyle-H_{j_{n}}^{n}(\rho_{j_{n}-1/2}^{n}\wedge\kappa,\rho_{j_{n}+1/2}^{n}\wedge\kappa,\rho_{j_{n}+3/2}^{n}\wedge\kappa,\rho_{j_{n}+5/2}^{n}\wedge\kappa)+\frac{\Delta t}{2(\chi_{j_{n+1}+1}^{n+1}-y^{n+1})}\mathcal{R}_{s^{n}}(\kappa,q^{n})
=χjn+1nynχjn+1+1n+1yn+1|ρjn+1/2nκ|+Δxχjn+1+1n+1yn+1|ρjn+3/2nκ|\displaystyle=\frac{\chi_{j_{n}+1}^{n}-y^{n}}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}|\rho_{j_{n}+1/2}^{n}-\kappa|+\frac{\Delta x}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}|\rho_{j_{n}+3/2}^{n}-\kappa|
Δtχjn+1+1n+1yn+1(Φjn+2nΦintn)+Δt2(χjn+1+1n+1yn+1)sn(κ,qn),\displaystyle-\frac{\Delta t}{\chi_{j_{n+1}+1}^{n+1}-y^{n+1}}\left(\Phi_{j_{n}+2}^{n}-\Phi_{int}^{n}\right)+\frac{\Delta t}{2(\chi_{j_{n+1}+1}^{n+1}-y^{n+1})}\mathcal{R}_{s^{n}}(\kappa,q^{n}),

which is exactly (3.9) in the case j=jn+1j=j_{n+1}. The obtaining of (3.9) in the case j=jn+11j=j_{n+1}-1 is similar, so we omit the details of the proof for this case.   \square

3.3 Continuous inequalities for the approximate solution

The next step of the reasoning is to derive analogous inequalities to (2.1)-(2.2), verified by the approximate solution ρΔ\rho_{\Delta}, starting from (3.9) and (3.4) – (3.6).

In this section, we fix a test function φ𝐂𝐜(Ω¯,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}^{+}) and set:

n,j,φj+1/2n:=1χj+1nχjnχjnχj+1nφ(x,tn)dx=χjnχj+1nφ(x,tn)dx.\forall n\in\mathbb{N},\;\forall j\in\mathbb{Z},\quad\varphi_{j+1/2}^{n}:=\frac{1}{\chi_{j+1}^{n}-\chi_{j}^{n}}\int_{\chi_{j}^{n}}^{\chi_{j+1}^{n}}\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}=\fint_{\chi_{j}^{n}}^{\chi_{j+1}^{n}}\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}.

We start by deriving continuous entropy inequalities verified by ρΔ\rho_{\Delta}. Define the approximate entropy flux:

ΦΔ(ρΔ,κ):=n(jjnΦjn𝟏𝒫j+1/2n+jjn+1Φj+1n𝟏𝒫j+1/2n).\Phi_{\Delta}(\rho_{\Delta},\kappa):=\sum_{n\in\mathbb{N}}\left(\sum_{j\leq j_{n}}\Phi_{j}^{n}\mathbf{1}_{\mathcal{P}_{j+1/2}^{n}}+\sum_{j\geq j_{n}+1}\Phi_{j+1}^{n}\mathbf{1}_{\mathcal{P}_{j+1/2}^{n}}\right).
Proposition 3.3 (Approximate entropy inequalities).

Fix nn\in\mathbb{N} and κ[0,1]\kappa\in[0,1]. Then we have

tntn+1(|ρΔκ|tφ+ΦΔ(ρΔ,κ)xφ)dxdt\displaystyle\int_{t^{n}}^{t^{n+1}}\int_{\mathbb{R}}\biggl{(}|\rho_{\Delta}-\kappa|\partial_{t}\varphi+\Phi_{\Delta}\left(\rho_{\Delta},\kappa\right)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}} (3.10)
+|ρΔ(x,tn)κ|φ(x,tn)dx|ρΔ(x,tn+1)κ|φ(x,tn+1)dx\displaystyle+\int_{\mathbb{R}}|\rho_{\Delta}(x,t^{n})-\kappa|\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}-\int_{\mathbb{R}}|\rho_{\Delta}(x,t^{n+1})-\kappa|\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}}
+tntn+1sΔ(t)(κ,qΔ(t))φ(yΔ(t),t)dtO(Δx2)+O(ΔxΔt)+O(Δt2).\displaystyle+\int_{t^{n}}^{t^{n+1}}\mathcal{R}_{s_{\Delta}(t)}(\kappa,q_{\Delta}(t))\varphi(y_{\Delta}(t),t)\mathinner{\mathrm{d}{t}}\geq\mathop{}\mathopen{}O\mathopen{}\left(\Delta x^{2}\right)+\mathop{}\mathopen{}O\mathopen{}\left(\Delta x\Delta t\right)+\mathop{}\mathopen{}O\mathopen{}\left(\Delta t^{2}\right).

Proof.  For all j\{jn+12}j\in\mathbb{Z}\backslash\{j_{n+1}-2\}, we multiply the discrete entropy inequalities (3.9) by φj+1/2n+1\varphi_{j+1/2}^{n+1} and take the sum to obtain:

jjn+12|ρj+1/2n+1κ|φj+1/2n+1(χj+1n+1χjn+1)\displaystyle\sum_{j\neq j_{n+1}-2}\left|\rho_{j+1/2}^{n+1}-\kappa\right|\varphi_{j+1/2}^{n+1}(\chi_{j+1}^{n+1}-\chi_{j}^{n+1})
j{jn+12,jn+11,jn+1}(|ρj+1/2nκ|(χj+1nχjn)(Φj+1nΦjn)Δt)φj+1/2n+1\displaystyle\leq\sum_{j\notin\{j_{n+1}-2,j_{n+1}-1,j_{n+1}\}}\left(\left|\rho_{j+1/2}^{n}-\kappa\right|(\chi_{j+1}^{n}-\chi_{j}^{n})-(\Phi_{j+1}^{n}-\Phi_{j}^{n})\Delta t\right)\varphi_{j+1/2}^{n+1}
+|ρjn1/2nκ|φjn+11/2n+1(χjnnχjn1n)|ρjn+11/2n+1κ|φjn+11/2n+1Δx(ΦintnΦjn1n)φjn+11/2n+1Δt\displaystyle+|\rho_{j_{n}-1/2}^{n}-\kappa|\varphi_{j_{n+1}-1/2}^{n+1}(\chi_{j_{n}}^{n}-\chi_{j_{n}-1}^{n})-|\rho_{j_{n+1}-1/2}^{n+1}-\kappa|\varphi_{j_{n+1}-1/2}^{n+1}\Delta x-\left(\Phi_{int}^{n}-\Phi_{j_{n}-1}^{n}\right)\varphi_{j_{n+1}-1/2}^{n+1}\Delta t
+|ρjn+1/2nκ|φjn+1+1/2n+1(χjn+1nχjnn)+|ρjn+3/2nκ|φjn+1+1/2n+1Δx(Φjn+2nΦintn)φjn+1+1/2n+1Δt\displaystyle+|\rho_{j_{n}+1/2}^{n}-\kappa|\varphi_{j_{n+1}+1/2}^{n+1}(\chi_{j_{n}+1}^{n}-\chi_{j_{n}}^{n})+|\rho_{j_{n}+3/2}^{n}-\kappa|\varphi_{j_{n+1}+1/2}^{n+1}\Delta x-\left(\Phi_{j_{n}+2}^{n}-\Phi_{int}^{n}\right)\varphi_{j_{n+1}+1/2}^{n+1}\Delta t
+12sn(κ,qn)(φjn+11/2n+1+φjn+1+1/2n+1)Δt.\displaystyle+\frac{1}{2}\mathcal{R}_{s^{n}}(\kappa,q^{n})(\varphi_{j_{n+1}-1/2}^{n+1}+\varphi_{j_{n+1}+1/2}^{n+1})\Delta t.

This inequality can be rewritten as

j|ρj+1/2n+1κ|φj+1/2n+1(χj+1n+1χjn+1)j|ρj+1/2nκ|φj+1/2n+1(χj+1nχjn)\displaystyle\sum_{j\in\mathbb{Z}}\left|\rho_{j+1/2}^{n+1}-\kappa\right|\varphi_{j+1/2}^{n+1}(\chi_{j+1}^{n+1}-\chi_{j}^{n+1})-\sum_{j\in\mathbb{Z}}\left|\rho_{j+1/2}^{n}-\kappa\right|\varphi_{j+1/2}^{n+1}(\chi_{j+1}^{n}-\chi_{j}^{n})
|ρjn+11/2n+1κ|(φjn+11/2n+1φjn+13/2n+1)Δxε1+|ρjn1/2nκ|(φjn+11/2n+1φjn+13/2n+1)(χjnnχjn1n)ε2\displaystyle\leq-\underbrace{\left|\rho_{j_{n+1}-1/2}^{n+1}-\kappa\right|\left(\varphi_{j_{n+1}-1/2}^{n+1}-\varphi_{j_{n+1}-3/2}^{n+1}\right)\Delta x}_{\mathrm{\varepsilon}_{1}}+\underbrace{\left|\rho_{j_{n}-1/2}^{n}-\kappa\right|\left(\varphi_{j_{n+1}-1/2}^{n+1}-\varphi_{j_{n+1}-3/2}^{n+1}\right)(\chi_{j_{n}}^{n}-\chi_{j_{n}-1}^{n})}_{\mathrm{\varepsilon}_{2}}
+|ρjn+1/2nκ|(φjn+1+1/2n+1φjn+11/2n+1)(χjn+1nχjnn)ε3\displaystyle+\underbrace{\left|\rho_{j_{n}+1/2}^{n}-\kappa\right|\left(\varphi_{j_{n+1}+1/2}^{n+1}-\varphi_{j_{n+1}-1/2}^{n+1}\right)(\chi_{j_{n}+1}^{n}-\chi_{j_{n}}^{n})}_{\mathrm{\varepsilon}_{3}}
j{jn+12,jn+11,jn+1}(Φj+1nΦjn)φj+1/2n+1Δt(ΦintnΦjn1n)φjn+11/2n+1Δt\displaystyle-\sum_{j\notin\{j_{n+1}-2,j_{n+1}-1,j_{n+1}\}}(\Phi_{j+1}^{n}-\Phi_{j}^{n})\varphi_{j+1/2}^{n+1}\Delta t-\left(\Phi_{int}^{n}-\Phi_{j_{n}-1}^{n}\right)\varphi_{j_{n+1}-1/2}^{n+1}\Delta t
(Φjn+2nΦintn)φjn+1+1/2n+1Δt+12sn(κ,qn)(φjn+11/2n+1+φjn+1+1/2n+1)Δt,\displaystyle-\left(\Phi_{j_{n}+2}^{n}-\Phi_{int}^{n}\right)\varphi_{j_{n+1}+1/2}^{n+1}\Delta t+\frac{1}{2}\mathcal{R}_{s^{n}}(\kappa,q^{n})(\varphi_{j_{n+1}-1/2}^{n+1}+\varphi_{j_{n+1}+1/2}^{n+1})\Delta t,

with for all i[[1;3]]i\in[\![1;3]\!], |εi|8xφ𝐋Δx2|\mathrm{\varepsilon}_{i}|\leq 8\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x^{2}. We now proceed to the Abel’s transformation and reorganize the terms of the inequality. This leads us to:

j|ρj+1/2n+1κ|φj+1/2n+1(χj+1n+1χjn+1)j|ρj+1/2nκ|φj+1/2n(χj+1nχjn)A\displaystyle\underbrace{\sum_{j\in\mathbb{Z}}\left|\rho_{j+1/2}^{n+1}-\kappa\right|\varphi_{j+1/2}^{n+1}(\chi_{j+1}^{n+1}-\chi_{j}^{n+1})-\sum_{j\in\mathbb{Z}}\left|\rho_{j+1/2}^{n}-\kappa\right|\varphi_{j+1/2}^{n}(\chi_{j+1}^{n}-\chi_{j}^{n})}_{A}
j|ρj+1/2nκ|(φj+1/2n+1φj+1/2n)(χj+1nχjn)B+j{jn+12,jn+11}Φjn(φj+1/2n+1φj1/2n+1)ΔtC\displaystyle-\underbrace{\sum_{j\in\mathbb{Z}}\left|\rho_{j+1/2}^{n}-\kappa\right|\left(\varphi_{j+1/2}^{n+1}-\varphi_{j+1/2}^{n}\right)(\chi_{j+1}^{n}-\chi_{j}^{n})}_{B}+\underbrace{\sum_{j\notin\{j_{n+1}-2,j_{n+1}-1\}}\Phi_{j}^{n}\left(\varphi_{j+1/2}^{n+1}-\varphi_{j-1/2}^{n+1}\right)\Delta t}_{C}
12sn(κ,qn)(φjn+11/2n+1+φjn+1+1/2n+1)ΔtD+i=15εi,\displaystyle\leq\underbrace{\frac{1}{2}\mathcal{R}_{s^{n}}(\kappa,q^{n})(\varphi_{j_{n+1}-1/2}^{n+1}+\varphi_{j_{n+1}+1/2}^{n+1})\Delta t}_{D}+\sum_{i=1}^{5}\mathrm{\varepsilon}_{i},

with for all i[[4;5]]i\in[\![4;5]\!], |εi|4f𝐋xφ𝐋ΔxΔt|\mathrm{\varepsilon}_{i}|\leq 4\|f\|_{\mathbf{L}^{\infty}}\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x\Delta t. We immediately see that

A=|ρΔ(x,tn+1)κ|φ(x,tn+1)dx|ρΔ(x,tn)κ|φ(x,tn)dx.A=\int_{\mathbb{R}}\left|\rho_{\Delta}(x,t^{n+1})-\kappa\right|\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}}-\int_{\mathbb{R}}\left|\rho_{\Delta}(x,t^{n})-\kappa\right|\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}.

We conclude this proof by estimating the remaining terms of the inequality.

Estimating BB. First, note that

B\displaystyle B =jjn2𝒫j+1/2n|ρΔκ|tφdxdt+jjn+1𝒫j+1/2n|ρΔκ|tφdxdt\displaystyle=\sum_{j\leq j_{n}-2}\iint_{\mathcal{P}_{j+1/2}^{n}}\left|\rho_{\Delta}-\kappa\right|\partial_{t}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\sum_{j\geq j_{n}+1}\iint_{\mathcal{P}_{j+1/2}^{n}}\left|\rho_{\Delta}-\kappa\right|\partial_{t}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
+|ρjn1/2nκ|(χjn1n+1χjn+1n+1φ(x,tn+1)dxχjn1nynφ(x,tn)dx)(ynχjn1n)B1\displaystyle+\underbrace{\left|\rho_{j_{n}-1/2}^{n}-\kappa\right|\biggl{(}\fint_{\chi_{j_{n}-1}^{n+1}}^{\chi_{j_{n}+1}^{n+1}}\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}}-\fint_{\chi_{j_{n}-1}^{n}}^{y^{n}}\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}\biggr{)}(y^{n}-\chi_{j_{n}-1}^{n})}_{B_{1}}
+|ρjn+1/2nκ|(χjnn+1yn+1φ(x,tn+1)dxynχjn+1nφ(x,tn)dx)(χjn+1nyn)B2\displaystyle+\underbrace{\left|\rho_{j_{n}+1/2}^{n}-\kappa\right|\biggl{(}\fint_{\chi_{j_{n}}^{n+1}}^{y^{n+1}}\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}}-\fint_{y^{n}}^{\chi_{j_{n}+1}^{n}}\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}\biggr{)}(\chi_{j_{n}+1}^{n}-y^{n})}_{B_{2}}
+|ρjn+3/2nκ|(yn+1χjn+2n+1φ(x,tn+1)dxχjn+1nχjn+2nφ(x,tn)dx)ΔxB3.\displaystyle+\underbrace{\left|\rho_{j_{n}+3/2}^{n}-\kappa\right|\biggl{(}\fint_{y^{n+1}}^{\chi_{j_{n}+2}^{n+1}}\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}}-\fint_{\chi_{j_{n}+1}^{n}}^{\chi_{j_{n}+2}^{n}}\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}\biggr{)}\Delta x}_{B_{3}}.

Since

𝒫jn1/2n|ρΔκ|tφdxdt\displaystyle\iint_{\mathcal{P}_{j_{n}-1/2}^{n}}\left|\rho_{\Delta}-\kappa\right|\partial_{t}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}
=|ρjn1/2nκ|(χjn1n+1yn+1φ(x,tn+1)dxχjn1nynφ(x,tn)dxsntntn+1φ(yn+sn(ttn),t)dt)\displaystyle=\left|\rho_{j_{n}-1/2}^{n}-\kappa\right|\biggl{(}\int_{\chi_{j_{n}-1}^{n+1}}^{y^{n+1}}\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}}-\int_{\chi_{j_{n}-1}^{n}}^{y^{n}}\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}-s^{n}\int_{t^{n}}^{t^{n+1}}\varphi(y^{n}+s^{n}(t-t^{n}),t)\mathinner{\mathrm{d}{t}}\biggr{)}
=|ρjn1/2nκ|(yn+1χjn1n+1ynχjn1nχjn1n+1yn+1φ(x,tn+1)dxχjn1nynφ(x,tn)dx.\displaystyle=\left|\rho_{j_{n}-1/2}^{n}-\kappa\right|\biggl{(}\frac{y^{n+1}-\chi_{j_{n}-1}^{n+1}}{y^{n}-\chi_{j_{n}-1}^{n}}\fint_{\chi_{j_{n}-1}^{n+1}}^{y^{n+1}}\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}}-\fint_{\chi_{j_{n}-1}^{n}}^{y^{n}}\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}\biggr{.}
.+ynyn+1ynχjn1ntntn+1φ(yn+sn(ttn),t)dt)(ynχjn1n),\displaystyle\biggl{.}+\frac{y^{n}-y^{n+1}}{y^{n}-\chi_{j_{n}-1}^{n}}\fint_{t^{n}}^{t^{n+1}}\varphi(y^{n}+s^{n}(t-t^{n}),t)\mathinner{\mathrm{d}{t}}\biggr{)}(y^{n}-\chi_{j_{n}-1}^{n}),

we deduce the bound:

|B1𝒫jn1/2n|ρΔκ|tφdxdt|\displaystyle\left|B_{1}-\iint_{\mathcal{P}_{j_{n}-1/2}^{n}}\left|\rho_{\Delta}-\kappa\right|\partial_{t}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\right|
=|ρjn1/2nκ|(yn+1yn)|χjn1n+1yn+1φ(x,tn+1)dxtntn+1φ(yn+sn(ttn),t)dt|\displaystyle=\left|\rho_{j_{n}-1/2}^{n}-\kappa\right|(y^{n+1}-y^{n})\left|\fint_{\chi_{j_{n}-1}^{n+1}}^{y^{n+1}}\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}}-\fint_{t^{n}}^{t^{n+1}}\varphi(y^{n}+s^{n}(t-t^{n}),t)\mathinner{\mathrm{d}{t}}\right|
y˙𝐋(3xφ𝐋Δx+tφ𝐋Δt+2y˙𝐋xφ𝐋Δt)Δt.\displaystyle\leq\|\dot{y}\|_{\mathbf{L}^{\infty}}\biggl{(}3\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x+\|\partial_{t}\varphi\|_{\mathbf{L}^{\infty}}\Delta t+2\|\dot{y}\|_{\mathbf{L}^{\infty}}\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta t\biggr{)}\Delta t.

The same way, we would derive the estimation:

|B2+B3𝒫jn+1/2n|ρΔκ|tφdxdt|\displaystyle\left|B_{2}+B_{3}-\iint_{\mathcal{P}_{j_{n}+1/2}^{n}}\left|\rho_{\Delta}-\kappa\right|\partial_{t}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\right|
6xφ𝐋Δx2+y˙𝐋(2xφ𝐋Δx+tφ𝐋Δt+2y˙𝐋xφ𝐋Δt)Δt.\displaystyle\leq 6\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x^{2}+\|\dot{y}\|_{\mathbf{L}^{\infty}}\biggl{(}2\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x+\|\partial_{t}\varphi\|_{\mathbf{L}^{\infty}}\Delta t+2\|\dot{y}\|_{\mathbf{L}^{\infty}}\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta t\biggr{)}\Delta t.

Estimating CC. We write:

C\displaystyle C =λj{jn+12,jn+11,jn+1}χjnχj+1nxΔxxΦjnxφ(y,tn+1)dydx+Φjn+1n(φjn+1+1/2n+1φjn+11/2n+1)Δtε6\displaystyle=\lambda\sum_{j\notin\{j_{n+1}-2,j_{n+1}-1,j_{n+1}\}}\int_{\chi_{j}^{n}}^{\chi_{j+1}^{n}}\int_{x-\Delta x}^{x}\Phi_{j}^{n}\partial_{x}\varphi(y,t^{n+1})\mathinner{\mathrm{d}{y}}\mathinner{\mathrm{d}{x}}+\underbrace{\Phi_{j_{n+1}}^{n}\left(\varphi_{j_{n+1}+1/2}^{n+1}-\varphi_{j_{n+1}-1/2}^{n+1}\right)\Delta t}_{\mathrm{\varepsilon}_{6}}
=tntn+1ΦΔ(ρΔ,κ)xφdxdt+ε6jn+12jjn+11𝒫j+1/2nΦΔ(ρΔ,κ)xφdxdtε7\displaystyle=\int_{t^{n}}^{t^{n+1}}\int_{\mathbb{R}}\Phi_{\Delta}(\rho_{\Delta},\kappa)\partial_{x}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\mathrm{\varepsilon}_{6}-\underbrace{\sum_{j_{n+1}-2\leq j\leq j_{n+1}-1}\iint_{\mathcal{P}_{j+1/2}^{n}}\Phi_{\Delta}(\rho_{\Delta},\kappa)\partial_{x}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}}_{\mathrm{\varepsilon}_{7}}
+j{jn+12,jn+11,jn+1}(λχjnχj+1nxΔxxΦjnxφ(y,tn+1)dydx)tntn+1ΦΔ(ρΔ,κ)xφdxdtε8,\displaystyle+\underbrace{\sum_{j\notin\{j_{n+1}-2,j_{n+1}-1,j_{n+1}\}}\biggl{(}\lambda\int_{\chi_{j}^{n}}^{\chi_{j+1}^{n}}\int_{x-\Delta x}^{x}\Phi_{j}^{n}\partial_{x}\varphi(y,t^{n+1})\mathinner{\mathrm{d}{y}}\mathinner{\mathrm{d}{x}}\biggr{)}-\int_{t^{n}}^{t^{n+1}}\int_{\mathbb{R}}\Phi_{\Delta}(\rho_{\Delta},\kappa)\partial_{x}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}}_{\mathrm{\varepsilon}_{8}},

with

|ε6|+|ε7|8f𝐋xφ𝐋ΔxΔt,|\mathrm{\varepsilon}_{6}|+|\mathrm{\varepsilon}_{7}|\leq 8\|f\|_{\mathbf{L}^{\infty}}\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x\Delta t,

and

|ε8|f𝐋(4xx2φ𝐋(+,𝐋1)Δx+tx2φ𝐋(+,𝐋1)Δt)Δt.|\mathrm{\varepsilon}_{8}|\leq\|f\|_{\mathbf{L}^{\infty}}\biggl{(}4\|\partial_{xx}^{2}\varphi\|_{\mathbf{L}^{\infty}(\mathbb{R}^{+},\mathbf{L}^{1})}\Delta x+\|\partial_{tx}^{2}\varphi\|_{\mathbf{L}^{\infty}(\mathbb{R}^{+},\mathbf{L}^{1})}\Delta t\biggr{)}\Delta t.

Estimating DD. Finally, we have

D\displaystyle D =sn(κ,qn)φ(yn+1,tn+1)Δt+1yn+1χjn+11χjn+11n+1yn+1(φ(x,tn+1)φ(yn+1,tn+1))Δtε9\displaystyle=\mathcal{R}_{s^{n}}(\kappa,q^{n})\varphi(y^{n+1},t^{n+1})\Delta t+\underbrace{\frac{1}{y^{n+1}-\chi_{j_{n+1}-1}}\int_{\chi_{j_{n+1}-1}^{n+1}}^{y^{n+1}}(\varphi(x,t^{n+1})-\varphi(y^{n+1},t^{n+1}))\Delta t}_{\mathrm{\varepsilon}_{9}}
+1χjn+1+1yn+1yn+1χjn+1+1n+1(φ(x,tn+1)φ(yn+1,tn+1))Δtε10\displaystyle+\underbrace{\frac{1}{\chi_{j_{n+1}+1}-y^{n+1}}\int_{y^{n+1}}^{\chi_{j_{n+1}+1}^{n+1}}(\varphi(x,t^{n+1})-\varphi(y^{n+1},t^{n+1}))\Delta t}_{\mathrm{\varepsilon}_{10}}
=tntn+1sΔ(t)(κ,qΔ(t))φ(yΔ(t),t)dt+ε9+ε10+tntn+1sΔ(t)(κ,qΔ(t))(φ(yn+1,tn+1)φ(yΔ(t),t))dtε11,\displaystyle=\int_{t^{n}}^{t^{n+1}}\mathcal{R}_{s_{\Delta}(t)}(\kappa,q_{\Delta}(t))\varphi(y_{\Delta}(t),t)\mathinner{\mathrm{d}{t}}+\mathrm{\varepsilon}_{9}+\mathrm{\varepsilon}_{10}+\underbrace{\int_{t^{n}}^{t^{n+1}}\mathcal{R}_{s_{\Delta}(t)}(\kappa,q_{\Delta}(t))(\varphi(y^{n+1},t^{n+1})-\varphi(y_{\Delta}(t),t))\mathinner{\mathrm{d}{t}}}_{\mathrm{\varepsilon}_{11}},

with

|ε9|+|ε10|+|ε11|2f𝐋(2xφ𝐋Δx+y˙𝐋xφ𝐋Δt+tφ𝐋Δt)Δt.|\mathrm{\varepsilon}_{9}|+|\mathrm{\varepsilon}_{10}|+|\mathrm{\varepsilon}_{11}|\leq 2\|f\|_{\mathbf{L}^{\infty}}\biggl{(}2\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x+\|\dot{y}\|_{\mathbf{L}^{\infty}}\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta t+\|\partial_{t}\varphi\|_{\mathbf{L}^{\infty}}\Delta t\biggr{)}\Delta t.

\square

Note that if φ\varphi is supported in time in [0,T][0,T], with T[tN,tN+1[T\in[t^{N},t^{N+1}\mathclose{[}, then by summing (3.10) over n[[0;N+1]]n\in[\![0;N+1]\!], we obtain (recall that λ\lambda is fixed):

0T\displaystyle\int_{0}^{T}\int_{\mathbb{R}} (|ρΔκ|tφ+ΦΔ(ρΔ,κ)xφ)dxdt+|ρΔ0κ|φ(x,0)dx\displaystyle\biggl{(}|\rho_{\Delta}-\kappa|\partial_{t}\varphi+\Phi_{\Delta}\left(\rho_{\Delta},\kappa\right)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{\Delta}^{0}-\kappa|\varphi(x,0)\mathinner{\mathrm{d}{x}} (3.11)
+0TsΔ(t)(κ,qΔ(t))φ(yΔ(t),t)dtO(Δx)+O(Δt).\displaystyle+\int_{0}^{T}\mathcal{R}_{s_{\Delta}(t)}(\kappa,q_{\Delta}(t))\varphi(y_{\Delta}(t),t)\mathinner{\mathrm{d}{t}}\geq\mathop{}\mathopen{}O\mathopen{}\left(\Delta x\right)+\mathop{}\mathopen{}O\mathopen{}\left(\Delta t\right).

We now turn to the proof of an approximate version of (2.2). Let us define the approximate flux function:

𝐅Δ(ρΔ):=n(jjnfjn𝟏𝒫j+1/2n+jjn+1fj+1n𝟏𝒫j+1/2n).\mathbf{F}_{\Delta}\left(\rho_{\Delta}\right):=\sum_{n\in\mathbb{N}}\left(\sum_{j\leq j_{n}}f_{j}^{n}\mathbf{1}_{\mathcal{P}_{j+1/2}^{n}}+\sum_{j\geq j_{n}+1}f_{j+1}^{n}\mathbf{1}_{\mathcal{P}_{j+1/2}^{n}}\right).
Proposition 3.4 (Approximate constraint inequalities).

Fix nn\in\mathbb{N} and κ[0,1]\kappa\in[0,1]. Then we have

yn+ρΔ(x,tn)φ(x,tn)dxyn+1+ρΔ(x,tn+1)φ(x,tn+1)dx\displaystyle\int_{y^{n}}^{+\infty}\rho_{\Delta}(x,t^{n})\varphi(x,t^{n})\mathinner{\mathrm{d}{x}}-\int_{y^{n+1}}^{+\infty}\rho_{\Delta}(x,t^{n+1})\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}} (3.12)
tntn+1(ρΔtφ+𝐅Δ(ρΔ)xφ)dxdttntn+1qΔ(t)φ(yΔ(t),t)dt+O(Δx2)+O(ΔxΔt)+O(Δt2).\displaystyle\begin{array}[]{cl}\displaystyle{-\int_{t^{n}}^{t^{n+1}}\int_{\mathbb{R}}\biggl{(}\rho_{\Delta}\partial_{t}\varphi+\mathbf{F}_{\Delta}\left(\rho_{\Delta}\right)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}}&\leq\displaystyle{\int_{t^{n}}^{t^{n+1}}q_{\Delta}(t)\varphi(y_{\Delta}(t),t)\mathinner{\mathrm{d}{t}}}\\[10.0pt] &+\mathop{}\mathopen{}O\mathopen{}\left(\Delta x^{2}\right)+\mathop{}\mathopen{}O\mathopen{}\left(\Delta x\Delta t\right)+\mathop{}\mathopen{}O\mathopen{}\left(\Delta t^{2}\right).\end{array}

Proof.  Following the steps of the proof of Proposition 3.3, we first multiply the scheme (3.4)-(3.6) by φj+1/2n+1\varphi_{j+1/2}^{n+1}, sum over jjn+1j\geq j_{n+1} and then apply the summation by parts procedure. This time, we obtain:

jjn+1ρj+1/2n+1φj+1/2n+1(χj+1n+1χjn+1)jjnρj+1/2nφj+1/2n(χj+1nχjn)A\displaystyle\underbrace{\sum_{j\geq j_{n+1}}\rho_{j+1/2}^{n+1}\varphi_{j+1/2}^{n+1}(\chi_{j+1}^{n+1}-\chi_{j}^{n+1})-\sum_{j\geq j_{n}}\rho_{j+1/2}^{n}\varphi_{j+1/2}^{n}(\chi_{j+1}^{n}-\chi_{j}^{n})}_{A}
jjnρj+1/2n(φj+1/2n+1φj+1/2n)(χj+1nχjn)B+jjn+2fjn(φj+1/2n+1φj1/2n+1)ΔtCqnφjn+1+1/2n+1ΔtD+ε,\displaystyle-\underbrace{\sum_{j\geq j_{n}}\rho_{j+1/2}^{n}\left(\varphi_{j+1/2}^{n+1}-\varphi_{j+1/2}^{n}\right)(\chi_{j+1}^{n}-\chi_{j}^{n})}_{B}+\underbrace{\sum_{j\geq j_{n}+2}f_{j}^{n}\left(\varphi_{j+1/2}^{n+1}-\varphi_{j-1/2}^{n+1}\right)\Delta t}_{C}\leq\underbrace{q^{n}\varphi_{j_{n+1}+1/2}^{n+1}\Delta t}_{D}+\mathrm{\varepsilon},

with ε8xφ𝐋Δx2\mathrm{\varepsilon}\leq 8\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x^{2}. Clearly,

A=yn+1+ρΔ(x,tn+1)φ(x,tn+1)dxyn+ρΔ(x,tn)φ(x,tn)dx,A=\int_{y^{n+1}}^{+\infty}\rho_{\Delta}(x,t^{n+1})\varphi(x,t^{n+1})\mathinner{\mathrm{d}{x}}-\int_{y^{n}}^{+\infty}\rho_{\Delta}(x,t^{n})\varphi(x,t^{n})\mathinner{\mathrm{d}{x}},

and estimate (3.12) follows from the bounds:

|Btntn+1ρΔtφdxdt|\displaystyle\left|B-\int_{t^{n}}^{t^{n+1}}\int_{\mathbb{R}}\rho_{\Delta}\partial_{t}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\right|
(3xφ𝐋Δx+tφ𝐋Δt)Δt+y˙𝐋(2xφ𝐋Δx+2y˙𝐋xφ𝐋Δt+tφ𝐋Δt)Δt\displaystyle\leq(3\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x+\|\partial_{t}\varphi\|_{\mathbf{L}^{\infty}}\Delta t)\Delta t+\|\dot{y}\|_{\mathbf{L}^{\infty}}\biggl{(}2\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x+2\|\dot{y}\|_{\mathbf{L}^{\infty}}\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta t+\|\partial_{t}\varphi\|_{\mathbf{L}^{\infty}}\Delta t\biggr{)}\Delta t
|Ctntn+1𝐅Δ(ρΔ)xφdxdt|f𝐋(6xφ𝐋+4xx2φ𝐋(+,𝐋1)+tx2φ𝐋(+,𝐋1))ΔxΔt\displaystyle\left|C-\int_{t^{n}}^{t^{n+1}}\int_{\mathbb{R}}\mathbf{F}_{\Delta}\left(\rho_{\Delta}\right)\partial_{x}\varphi\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\right|\leq\|f\|_{\mathbf{L}^{\infty}}\biggl{(}6\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}+4\|\partial_{xx}^{2}\varphi\|_{\mathbf{L}^{\infty}(\mathbb{R}^{+},\mathbf{L}^{1})}+\|\partial_{tx}^{2}\varphi\|_{\mathbf{L}^{\infty}(\mathbb{R}^{+},\mathbf{L}^{1})}\biggr{)}\Delta x\Delta t
|Dtntn+1qΔ(t)φ(yΔ(t),t)dt|q𝐋(2xφ𝐋Δx+tφ𝐋Δt+y˙𝐋xφ𝐋Δt)Δt.\displaystyle\left|D-\int_{t^{n}}^{t^{n+1}}q_{\Delta}(t)\varphi(y_{\Delta}(t),t)\mathinner{\mathrm{d}{t}}\right|\leq\|q\|_{\mathbf{L}^{\infty}}\biggl{(}2\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta x+\|\partial_{t}\varphi\|_{\mathbf{L}^{\infty}}\Delta t+\|\dot{y}\|_{\mathbf{L}^{\infty}}\|\partial_{x}\varphi\|_{\mathbf{L}^{\infty}}\Delta t\biggr{)}\Delta t.

\square

If φ\varphi is supported in time in (0,T)(0,T), with T[tN,tN+1[T\in[t^{N},t^{N+1}\mathclose{[}, then by summing (3.10) over n[[0;N+1]]n\in[\![0;N+1]\!], we obtain:

0T(ρΔtφ+𝐅Δ(ρΔ)xφ)dxdt0TqΔ(t)φ(yΔ(t),t)dt+O(Δx)+O(Δt).\displaystyle-\int_{0}^{T}\int_{\mathbb{R}}\biggl{(}\rho_{\Delta}\partial_{t}\varphi+\mathbf{F}_{\Delta}\left(\rho_{\Delta}\right)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\leq\int_{0}^{T}q_{\Delta}(t)\varphi(y_{\Delta}(t),t)\mathinner{\mathrm{d}{t}}+\mathop{}\mathopen{}O\mathopen{}\left(\Delta x\right)+\mathop{}\mathopen{}O\mathopen{}\left(\Delta t\right). (3.13)

3.4 Compactness and convergence

The remaining part of the reasoning consists in obtaining sufficient compactness for the sequence (ρΔ)Δ(\rho_{\Delta})_{\Delta} in order to pass to the limit in (3.11)-(3.13). To doing so, we adapt techniques and results put forward by Towers in [22]. With this in mind, we suppose in this section that the flux function, still bell-shaped, is such that

μ>0,ρ[0,1],f′′(ρ)μ.\exists\mu>0,\;\forall\rho\in[0,1],\quad f^{\prime\prime}(\rho)\leq-\mu. (3.14)

We denote for all nn\in\mathbb{N} and jj\in\mathbb{Z},

Djn:=max{ρj1/2nρj+1/2n,0}.D_{j}^{n}:=\max\left\{\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n},0\right\}.

We will also use the notation

n,^n+1=\{jn+12,jn+11,jn+1,jn+1+1}.\forall n\in\mathbb{N},\quad\widehat{\mathbb{Z}}_{n+1}=\mathbb{Z}\backslash\{j_{n+1}-2,j_{n+1}-1,j_{n+1},j_{n+1}+1\}.

In [22], the author dealt with a discontinuous in both time and space flux and the specific “vanishing viscosity” coupling at the interface. The discontinuity in space was localized along the curve {x=0}\{x=0\}. Here, we deal with a smooth flux, but we have a flux constraint along the curve {x=y(t)}\{x=y(t)\}. The applicability of the technique of [22] for our case with moving interface and flux-constrained interface coupling relies on the fact that one can derive a bound on Djn+1D_{j}^{n+1} as long as the interface does not enter the calculations for Djn+1D_{j}^{n+1} i.e. as long as j^n+1j\in\widehat{\mathbb{Z}}_{n+1} in the case jn+1=jn+1j_{n+1}=j_{n}+1.

Lemma 3.5.

Let nn\in\mathbb{N}, j^n+1j\in\widehat{\mathbb{Z}}_{n+1}, a:=μΔt4Δx\displaystyle{a:=\mu\frac{\Delta t}{4\Delta x}} and ψ(x):=xax2\displaystyle{\psi(x):=x-ax^{2}}. Then

Djn+1ψ(max{Dj1n,Djn,Dj+1n}).D_{j}^{n+1}\leq\psi\left(\max\left\{D_{j-1}^{n},D_{j}^{n},D_{j+1}^{n}\right\}\right). (3.15)

Proof.  For the sake of completeness, the proof, largely inspired by [22], can be found in Appendix A.   \square

Remark 3.2.

Fix nn\in\mathbb{N} and j^n+1j\in\widehat{\mathbb{Z}}_{n+1}. Remark that if Djn>0D_{j}^{n}>0, then we can write that for some ν(j){j1,j,j+1}\nu(j)\in\{j-1,j,j+1\}, we have

Djn+1\displaystyle D_{j}^{n+1} Dν(j)na(Dν(j)n)2\displaystyle\leq D_{\nu(j)}^{n}-a\left(D_{\nu(j)}^{n}\right)^{2}
=Dν(j)n(1aDν(j)n)=Dν(j)n1a2(Dν(j)n)21+aDν(j)nDν(j)n1+aDν(j)n=11Dν(j)n+a.\displaystyle=D_{\nu(j)}^{n}\left(1-aD_{\nu(j)}^{n}\right)=D_{\nu(j)}^{n}\frac{1-a^{2}\left(D_{\nu(j)}^{n}\right)^{2}}{1+aD_{\nu(j)}^{n}}\leq\frac{D_{\nu(j)}^{n}}{1+aD_{\nu(j)}^{n}}=\frac{1}{\frac{1}{D_{\nu(j)}^{n}}+a}.
Corollary 3.6.

Let nn\in\mathbb{N}. Then the scheme (3.4) – (3.6) verifies the following one-sided Lipschitz condition:

Djn+1{1(n+1)aifjjn+13n1((jn+12)j)aifjn+13njjn+131(j(jn+1+1))aifjn+1+2jjn+1+2+n1(n+1)aifjjn+1+2+n.D_{j}^{n+1}\leq\left\{\begin{array}[]{cll}\displaystyle{\frac{1}{(n+1)a}}&\text{if}&j\leq j_{n+1}-3-n\\[10.0pt] \displaystyle{\frac{1}{((j_{n+1}-2)-j)a}}&\text{if}&j_{n+1}-3-n\leq j\leq j_{n+1}-3\\[15.0pt] \displaystyle{\frac{1}{(j-(j_{n+1}+1))a}}&\text{if}&j_{n+1}+2\leq j\leq j_{n+1}+2+n\\[10.0pt] \displaystyle{\frac{1}{(n+1)a}}&\text{if}&j\geq j_{n+1}+2+n.\end{array}\right. (3.16)
Refer to caption
Figure 3: Illustration of the OSL bound (3.16).

Proof.  Fix nn\in\mathbb{N}. We only prove (3.16) in the cases jjn+1+2j\geq j_{n+1}+2. The reasoning for the cases jj03j\leq j_{0}-3 is very similar. Let us first prove by induction on kk\in\mathbb{N}^{*} that

k,j,min{n+1,j(jn+1+1)}kDjn+11ka.\forall k\in\mathbb{N}^{*},\;\forall j\in\mathbb{Z},\quad\min\{n+1,j-(j_{n+1}+1)\}\geq k\implies D_{j}^{n+1}\leq\frac{1}{ka}. (3.17)

Inequality (3.17) holds if k=1k=1. Indeed, if k=1k=1, then jjn+1+2j\geq j_{n+1}+2 i.e. j^n+1j\in\widehat{\mathbb{Z}}_{n+1}. By (3.15),

νj{j1,j,j+1},Djn+1Dνjna(Dνjn)2.\exists\nu_{j}\in\{j-1,j,j+1\},\quad D_{j}^{n+1}\leq D_{\nu_{j}}^{n}-a\left(D_{\nu_{j}}^{n}\right)^{2}.

If Dνjn=0D_{\nu_{j}}^{n}=0, then Djn+1=01/aD_{j}^{n+1}=0\leq 1/a. Otherwise, we can write:

Djn+111Dνjn+a1a=1ka.D_{j}^{n+1}\leq\frac{1}{\frac{1}{D_{\nu_{j}}^{n}}+a}\leq\frac{1}{a}=\frac{1}{ka}.

Now, let us assume that (3.17) holds for some integer kk\in\mathbb{N}^{*} and suppose that min{n+1,j(jn+1+1)}k+1\min\{n+1,j-(j_{n+1}+1)\}\geq k+1. Again, by (3.15),

νj{j1,j,j+1},Djn+1Dνjna(Dνjn)2.\exists\nu_{j}\in\{j-1,j,j+1\},\quad D_{j}^{n+1}\leq D_{\nu_{j}}^{n}-a\left(D_{\nu_{j}}^{n}\right)^{2}.

Since

nkandνj(jn+1)(j1)(jn+1+1)=j(jn+1+1)1k,n\geq k\quad\text{and}\quad\nu_{j}-(j_{n}+1)\geq(j-1)-(j_{n+1}+1)=j-(j_{n+1}+1)-1\geq k,

we deduce that min{n,j(jn+1)}k\min\{n,j-(j_{n}+1)\}\geq k, hence, using the induction property:

Djn+111Dνjn+a1(k+1)a,D_{j}^{n+1}\leq\frac{1}{\frac{1}{D_{\nu_{j}}^{n}}+a}\leq\frac{1}{(k+1)a},

which concludes the induction argument. Estimates (3.16) in the cases jjn+1+2j\geq j_{n+1}+2 follow for suitable choices of kk in (3.17).   \square

Corollary 3.7 (Localized 𝐁𝐕\mathbf{BV} estimates).

Fix 0<ε<X0<\mathrm{\varepsilon}<X and suppose that 3Δxε3\Delta x\leq\mathrm{\varepsilon} and that tn+1ε2L\displaystyle{t^{n+1}\geq\frac{\mathrm{\varepsilon}}{2L}}. Then there exists a constant Λ=Λ(1ε,X)\displaystyle{\Lambda=\Lambda\left(\frac{1}{\mathrm{\varepsilon}},X\right)}, nondecreasing with respect to its arguments such that

𝐓𝐕(ρΔ(tn+1) 1]yn+1+ε,yn+1+X[)Λ,\mathbf{TV}\left(\rho_{\Delta}(t^{n+1})\;\mathbf{1}_{\mathopen{]}y^{n+1}+\mathrm{\varepsilon},y^{n+1}+X\mathclose{[}}\right)\leq\Lambda, (3.18)

and

yn+1+εyn+1+X|ρΔ(x,tn+2)ρΔ(x,tn+1))|dx2Δx+L(2Λ+1)Δt.\int_{y^{n+1}+\mathrm{\varepsilon}}^{y^{n+1}+X}\left|\rho_{\Delta}(x,t^{n+2})-\rho_{\Delta}(x,t^{n+1)})\right|\mathinner{\mathrm{d}{x}}\leq 2\Delta x+L\left(2\Lambda+1\right)\Delta t. (3.19)

Note that we have the same bounds for the quantities:

𝐓𝐕(ρΔ(tn+1) 1]yn+1X,yn+1ε[)andyn+1Xyn+1ε|ρΔ(x,tn+2)ρΔ(x,tn+1))|dx.\mathbf{TV}\left(\rho_{\Delta}(t^{n+1})\;\mathbf{1}_{\mathopen{]}y^{n+1}-X,y^{n+1}-\mathrm{\varepsilon}\mathclose{[}}\right)\;\;\text{and}\;\;\int_{y^{n+1}-X}^{y^{n+1}-\mathrm{\varepsilon}}\left|\rho_{\Delta}(x,t^{n+2})-\rho_{\Delta}(x,t^{n+1)})\right|\mathinner{\mathrm{d}{x}}.

Proof.  Let kn+1,Jn+1k_{n+1},J_{n+1}\in\mathbb{Z} such that yn+1+ε]χkn+1n+1,χkn+1n+1+Δx[y^{n+1}+\mathrm{\varepsilon}\in\mathopen{]}\chi_{k_{n+1}}^{n+1},\chi_{k_{n+1}}^{n+1}+\Delta x\mathclose{[} and yn+1+X]χJn+1n+1,χJn+1n+1+Δx[y^{n+1}+X\in\mathopen{]}\chi_{J_{n+1}}^{n+1},\chi_{J_{n+1}}^{n+1}+\Delta x\mathclose{[}. We have:

𝐓𝐕(ρΔ(tn+1) 1]yn+1+ε,yn+1+X[)\displaystyle\mathbf{TV}\left(\rho_{\Delta}(t^{n+1})\;\mathbf{1}_{\mathopen{]}y^{n+1}+\mathrm{\varepsilon},y^{n+1}+X\mathclose{[}}\right) =j=kn+1+1Jn+1|ρj+1/2n+1ρj1/2n+1|\displaystyle=\sum_{j=k_{n+1}+1}^{J_{n+1}}|\rho_{j+1/2}^{n+1}-\rho_{j-1/2}^{n+1}|
=2j=kn+1+1Jn+1Djn+1j=kn+1+1Jn+1(ρj+1/2n+1ρj1/2n+1)\displaystyle=2\sum_{j=k_{n+1}+1}^{J_{n+1}}D_{j}^{n+1}-\sum_{j=k_{n+1}+1}^{J_{n+1}}(\rho_{j+1/2}^{n+1}-\rho_{j-1/2}^{n+1})
=2j=kn+1+1Jn+1Djn+1(ρJn+11/2n+1ρkn+1+1/2n+1)1+2j=kn+1+1Jn+1Djn+1.\displaystyle=2\sum_{j=k_{n+1}+1}^{J_{n+1}}D_{j}^{n+1}-(\rho_{J_{n+1}-1/2}^{n+1}-\rho_{k_{n+1}+1/2}^{n+1})\leq 1+2\sum_{j=k_{n+1}+1}^{J_{n+1}}D_{j}^{n+1}.

Now, for all jkn+1+1j\geq k_{n+1}+1, we have

j(jn+1+1)(kn+1+1)(jn+1+1))ΔxΔx\displaystyle j-(j_{n+1}+1)\geq\frac{(k_{n+1}+1)-(j_{n+1}+1))\Delta x}{\Delta x} =(χkn+1n+1+Δx)χjn+1n+1xΔ\displaystyle=\frac{(\chi_{k_{n+1}}^{n+1}+\Delta x)-\chi_{j_{n+1}}^{n+1}}{{}^{\Delta}x}
(yn+1+ε)(yn+1+2Δx)Δx=εΔx21.\displaystyle\geq\frac{(y^{n+1}+\mathrm{\varepsilon})-(y^{n+1}+2\Delta x)}{\Delta x}=\frac{\mathrm{\varepsilon}}{\Delta x}-2\geq 1.

Lemma 3.16 ensures that

𝐓𝐕(ρΔ(tn+1) 1]yn+1+ε,yn+1+X[)1+2aj=kn+1+1Jn+11min{n+1,j(jn+1+1)}.\mathbf{TV}\left(\rho_{\Delta}(t^{n+1})\;\mathbf{1}_{\mathopen{]}y^{n+1}+\mathrm{\varepsilon},y^{n+1}+X\mathclose{[}}\right)\leq 1+\frac{2}{a}\sum_{j=k_{n+1}+1}^{J_{n+1}}\frac{1}{\min\{n+1,j-(j_{n+1}+1)\}}.

However, we also have:

n+1=tn+1Δtε2LΔtεΔx=(yn+1+ε)yn+1Δxχkn+1n+1(χjn+1n+1+Δx)Δx=kn+1(jn+1+1).n+1=\frac{t^{n+1}}{\Delta t}\geq\frac{\mathrm{\varepsilon}}{2L\Delta t}\geq\frac{\mathrm{\varepsilon}}{\Delta x}=\frac{(y^{n+1}+\mathrm{\varepsilon})-y^{n+1}}{\Delta x}\geq\frac{\chi_{k_{n+1}}^{n+1}-(\chi_{j_{n+1}}^{n+1}+\Delta x)}{\Delta x}=k_{n+1}-(j_{n+1}+1).

We deduce that for all j[[kn+1+1;Jn+1]]j\in[\![k_{n+1}+1;J_{n+1}]\!], we have min{n+1,j(jn+1+1)}kn+1(jn+1+1)\min\{n+1,j-(j_{n+1}+1)\}\geq k_{n+1}-(j_{n+1}+1). Therefore,

j=kn+1+1Jn+1|ρj+1/2n+1ρj1/2n+1|\displaystyle\sum_{j=k_{n+1}+1}^{J_{n+1}}|\rho_{j+1/2}^{n+1}-\rho_{j-1/2}^{n+1}| 1+2a×(Jn+1kn+1kn+1(jn+1+1))\displaystyle\leq 1+\frac{2}{a}\times\left(\frac{J_{n+1}-k_{n+1}}{k_{n+1}-(j_{n+1}+1)}\right)
1+2a×(Xε+Δxε2Δx)\displaystyle\leq 1+\frac{2}{a}\times\left(\frac{X-\mathrm{\varepsilon}+\Delta x}{\mathrm{\varepsilon}-2\Delta x}\right)
Λ,Λ:=1+6Xaε,\displaystyle\leq\Lambda,\quad\Lambda:=1+\frac{6X}{a\mathrm{\varepsilon}},

which is exactly (3.18). Then,

yn+1+εyn+1+X|ρΔ(x,tn+2)ρΔ(x,tn+1))|dx\displaystyle\int_{y^{n+1}+\mathrm{\varepsilon}}^{y^{n+1}+X}\left|\rho_{\Delta}(x,t^{n+2})-\rho_{\Delta}(x,t^{n+1)})\right|\mathinner{\mathrm{d}{x}}
2Δx+j=kn+1+1Jn+1|ρj+1/2n+2ρj+1/2n+1|Δx\displaystyle\leq 2\Delta x+\sum_{j=k_{n+1}+1}^{J_{n+1}}|\rho_{j+1/2}^{n+2}-\rho_{j+1/2}^{n+1}|\Delta x
2Δx+f𝐋(j=kn+1+1Jn+1|ρj+3/2n+1ρj+1/2n+1|+j=kn+1+1Jn+1|ρj+1/2n+1ρj1/2n+1|)Δt\displaystyle\leq 2\Delta x+\|f^{\prime}\|_{\mathbf{L}^{\infty}}\left(\sum_{j=k_{n+1}+1}^{J_{n+1}}|\rho_{j+3/2}^{n+1}-\rho_{j+1/2}^{n+1}|+\sum_{j=k_{n+1}+1}^{J_{n+1}}|\rho_{j+1/2}^{n+1}-\rho_{j-1/2}^{n+1}|\right)\Delta t
2Δx+L(2Λ+1)Δt,\displaystyle\leq 2\Delta x+L\left(2\Lambda+1\right)\Delta t,

concluding the proof.   \square

Theorem 3.8.

Fix ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]), y𝐖𝐥𝐨𝐜1,(]0,+[,)y\in\mathbf{W}_{\mathbf{loc}}^{1,\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}) and q𝐋𝐥𝐨𝐜(]0,+[,)q\in\mathbf{L}_{\mathbf{loc}}^{\infty}(\mathopen{]}0,+\infty\mathclose{[},\mathbb{R}). Suppose that f𝐂2([0,1],+)f\in\mathbf{C}^{2}([0,1],\mathbb{R}^{+}) satisfies (1.1)-(3.14) and that yy is nondecreasing. Then as Δ0\Delta\to 0 while satisfying the CFL condition (3.2), (ρΔ)Δ(\rho_{\Delta})_{\Delta} converges a.e. on Ω\Omega to the admissible entropy solution to (1.3).

Proof.  Fix nn\in\mathbb{N}^{*}. The uniform convergence of (yΔ)Δ(y_{\Delta})_{\Delta} to yy, coupled with the 𝐁𝐕\mathbf{BV} bounds (3.18)-(3.19) and the uniform 𝐋\mathbf{L}^{\infty} bound (3.8) provide (up to a subsequence) a.e. convergence for the sequence (ρΔ)Δ(\rho_{\Delta})_{\Delta} in any rectangular bounded domains of the open subset

On:={(x,t)Ω:|xy(t)|>1/n},O_{n}:=\{(x,t)\in\Omega\;:\;|x-y(t)|>1/n\},

see [17, Appendix A]. The a.e. convergence on any compact subsets of Ωn\Omega_{n} follows by a classical covering argument. Then a diagonal procedure provides the a.e. convergence on any compact subsets of O:={(x,t)Ω:xy(t)}O:=\{(x,t)\in\Omega\;:\;x\neq y(t)\}. A further extraction yields the a.e. convergence on Ω\Omega.

Equipped with the convergence of (ρΔ)Δ(\rho_{\Delta})_{\Delta} to ρ\rho, we let Δ0\Delta\to 0 in (3.11) and (3.13) to establish that ρ\rho is an admissible entropy solution to (1.3). By uniqueness, the whole sequence converges to ρ\rho, which proves the theorem.   \square

Corollary 3.9.

Fix ρo𝐋(;[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R};[0,1]), y𝐖𝐥𝐨𝐜1,(]0,+[),y˙0y\in\mathbf{W}_{\mathbf{loc}}^{1,\infty}(\mathopen{]}0,+\infty\mathclose{[}),\dot{y}\geq 0 and q𝐋𝐥𝐨𝐜(]0,+[),q0q\in\mathbf{L}_{\mathbf{loc}}^{\infty}(\mathopen{]}0,+\infty\mathclose{[}),q\geq 0. Suppose that f𝐂2([0,1])f\in\mathbf{C}^{2}([0,1]) satisfies (1.1)-(3.14). Then Problem (1.3) admits a unique admissible entropy solution.

Proof.  Existence comes from Theorem 3.8 while uniqueness was established by Theorem 2.8.   \square

4 Well-posedness for the multiple trajectory problem

We now get back to the original problem (1.2). Let us detail the organization of this section. First, we construct a partition of the unity to reduce the study of (1.2) to an assembling of several local studies of (1.3), see Section 4.1. Using the definition based on germs, analogous to Definition 2.4, we will prove a stability estimate, leading to uniqueness, see Theorem 4.3. Then in Section 4.3, we construct a finite volume scheme in which we fully use the precise study of Section 3. A special treatment of the crossing points is described, see Section 4.3.1.

Let us recall that we are given a finite (or more generally locally finite) family of trajectories and constraints (yi,qi)i[[1;J]](y_{i},q_{i})_{i\in[\![1;J]\!]} defined on ]si,Ti[\mathopen{]}s_{i},T_{i}\mathclose{[} (0si<Ti0\leq s_{i}<T_{i}). Introduce the notations:

i[[1;J]],Γi:={(yi(t),t):t[si,Ti]}.\forall i\in[\![1;J]\!],\quad\Gamma_{i}:=\{(y_{i}(t),t)\;:\;t\in[s_{i},T_{i}]\}.

We suppose that for all i[[1;J]]i\in[\![1;J]\!], yi𝐖1,(]si,Ti[,)y_{i}\in\mathbf{W}^{1,\infty}(\mathopen{]}s_{i},T_{i}\mathclose{[},\mathbb{R}) and qi𝐋(]si,Ti[,+)q_{i}\in\mathbf{L}^{\infty}(\mathopen{]}s_{i},T_{i}\mathclose{[},\mathbb{R}^{+}). This notation means that what can be seen as crossing points between interfaces will be considered as endpoints of the interfaces; for instance, given two crossing lines, we split them into four interfaces having a common endpoint. We denote by (𝒞m)m[[1;M]](\mathcal{C}_{m})_{m\in[\![1;M]\!]} the set of all endpoints of the interfaces Γi\Gamma_{i}, i[[1;J]]i\in[\![1;J]\!].

4.1 Reduction to a single interface

Fix φ𝐂𝐜(Ω¯\m=1M𝒞m,)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega}\backslash\cup_{m=1}^{M}\mathcal{C}_{m},\mathbb{R}). Let us denote by KK the compact support of φ\varphi.

Step 1. For all i[[1;J]]i\in[\![1;J]\!], KΓiK\cap\Gamma_{i} is a compact subset (maybe empty) of Ω¯\overline{\Omega}, and the family (KΓi)i(K\cap\Gamma_{i})_{i} is pairwise disjoint. By compactness,

δ>0,i,j[[1;J]],ijdist(KΓi,KΓj)2δ.\exists\delta>0,\;\forall i,j\in[\![1;J]\!],\quad i\neq j\implies\text{dist}(K\cap\Gamma_{i},K\cap\Gamma_{j})\geq 2\delta.

Step 2. For all i[[1;J]]i\in[\![1;J]\!], set

Ωi:=(x,t)KΓiB((x,t),δ),\Omega_{i}:=\bigcup_{(x,t)\in K\cap\Gamma_{i}}\textbf{B}((x,t),\delta),

where B((x,t),δ)\textbf{B}((x,t),\delta) denotes the 2\mathbb{R}^{2}-euclidean open ball centered on (x,t)(x,t) and of radius δ\delta. Clearly, Ωi\Omega_{i} is an open subset of Ω¯\overline{\Omega} containing Γi\Gamma_{i}. Moreover, the family (Ωi)i(\Omega_{i})_{i} is pairwise disjoint. Indeed, suppose instead that for some i,j[[1;J]]i,j\in[\![1;J]\!] (iji\neq j), we have ΩiΩj\Omega_{i}\cap\Omega_{j}\neq\emptyset, and fix (x,t)ΩiΩj(x,t)\in\Omega_{i}\cap\Omega_{j}. By definition, there exists (xi,ti)KΓi(x_{i},t_{i})\in K\cap\Gamma_{i} and (xj,tj)KΓj(x_{j},t_{j})\in K\cap\Gamma_{j} such that

(x,t)B((xi,ti),δ)B((xj,tj),δ).(x,t)\in\textbf{B}((x_{i},t_{i}),\delta)\cap\textbf{B}((x_{j},t_{j}),\delta).

Using the triangle inequality, we deduce that

dist(KΓi,KΓj)dist((xi,ti),(xj,tj))dist((xi,ti),(x,t))+dist((x,t),(xj,tj))<2δ,\text{dist}(K\cap\Gamma_{i},K\cap\Gamma_{j})\leq\text{dist}((x_{i},t_{i}),(x_{j},t_{j}))\leq\text{dist}((x_{i},t_{i}),(x,t))+\text{dist}((x,t),(x_{j},t_{j}))<2\delta,

yielding the contradiction.

Step 3. Define the open subset (finite intersection of open subsets):

Ωo:={(x,t)Ω¯:i[[1;J]],dist((x,t),KΓi)δ2}.\Omega_{o}:=\left\{(x,t)\in\overline{\Omega}\;:\;\forall i\in[\![1;J]\!],\;\text{dist}((x,t),K\cap\Gamma_{i})\geq\frac{\delta}{2}\right\}.

The family (Ωi)i[[0;J]](\Omega_{i})_{i\in[\![0;J]\!]} is an open covering of ×+\mathbb{R}\times\mathbb{R}^{+}. Consequently, there exists a partition of the unity (θi)i[[0;J]](\theta_{i})_{i\in[\![0;J]\!]} associated with this covering:

i[[0;J]],θi𝐂𝐜(Ωi,+);(x,t)×+,i=0Jθi(x,t)=1.\forall i\in[\![0;J]\!],\;\theta_{i}\in\mathbf{C}_{\mathbf{c}}^{\infty}(\Omega_{i},\mathbb{R}^{+});\quad\forall(x,t)\in\mathbb{R}\times\mathbb{R}^{+},\;\sum_{i=0}^{J}\theta_{i}(x,t)=1.

Step 4. We write the function φ\varphi in the following manner:

φ=i=0J(φθi)=φo+i=1Jφi.\varphi=\sum_{i=0}^{J}(\varphi\theta_{i})=\varphi_{o}+\sum_{i=1}^{J}\varphi_{i}. (4.1)

Note that:

  1. 1.

    φo\varphi_{o} vanishes along all the interfaces;

  2. 2.

    for all i[[1;J]]i\in[\![1;J]\!], φi\varphi_{i} vanishes along all the interfaces but Γi\Gamma_{i}.

4.2 Definition of solutions and uniqueness

Following Section 2 and Definition 2.4, we give the following definition of solution.

Definition 4.1.

Let ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]). We say that ρ𝐋(Ω,[0,1])\rho\in\mathbf{L}^{\infty}(\Omega,[0,1]) is a 𝒢\mathcal{G}-entropy solution to (1.2) if:

(i) for all test functions φ𝐂𝐜(Ω¯\i=1JΓi,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega}\backslash\cup_{i=1}^{J}\Gamma_{i},\mathbb{R}^{+}) and κ[0,1]\kappa\in[0,1], the following entropy inequalities are verified:

0+(|ρκ|tφ+Φ(ρ,κ)xφ)dxdt+|ρo(x)κ|φ(x,0)dx0;\int_{0}^{+\infty}\int_{\mathbb{R}}\biggl{(}|\rho-\kappa|\partial_{t}\varphi+\Phi(\rho,\kappa)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{o}(x)-\kappa|\varphi(x,0)\mathinner{\mathrm{d}{x}}\geq 0; (4.2)

(ii) for all i[[1;J]]i\in[\![1;J]\!] and for a.e. t]si,Ti[t\in\mathopen{]}s_{i},T_{i}\mathclose{[},

(ρ(yi(t),t),ρ(yi(t)+,t))𝒢y˙i(t)(qi(t)),(\rho(y_{i}(t)-,t),\rho(y_{i}(t)+,t))\in\mathcal{G}_{\dot{y}_{i}(t)}(q_{i}(t)), (4.3)

where the admissibility germ 𝒢y˙i(qi)\mathcal{G}_{\dot{y}_{i}}(q_{i}) is defined in Definition 2.2.

Lemma 4.2 (Kato inequality).

Fix ρo,σo𝐋(,[0,1])\rho_{o},\sigma_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]). Let (qi)i[[1;J]](q_{i})_{i\in[\![1;J]\!]} and (qi)i[[1;J]](\overset{\sim}{q}_{i})_{i\in[\![1;J]\!]} be two family of constraints, where for all i[[1;J]]i\in[\![1;J]\!], qiq_{i}, qi𝐋(]si,Ti[,)\overset{\sim}{q}_{i}\in\mathbf{L}^{\infty}(\mathopen{]}s_{i},T_{i}\mathclose{[},\mathbb{R}). We denote by ρ\rho (resp. σ\sigma) a 𝒢\mathcal{G}-entropy solution to Problem (1.2) corresponding to initial datum ρo\rho_{o} (resp. σo\sigma_{o}) and constraints (qi)i[[1;J]](q_{i})_{i\in[\![1;J]\!]} (resp. (qi)i[[1;J]](\overset{\sim}{q}_{i})_{i\in[\![1;J]\!]}). Then for all test functions φ𝐂𝐜(Ω¯,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}^{+}), we have

0+(|ρσ|tφ+Φ(ρ,σ)xφ)dxdt+|ρo(x)σo(x)|φ(x,0)dx\displaystyle\int_{0}^{+\infty}\int_{\mathbb{R}}\biggl{(}|\rho-\sigma|\partial_{t}\varphi+\Phi(\rho,\sigma)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{o}(x)-\sigma_{o}(x)|\varphi(x,0)\mathinner{\mathrm{d}{x}} (4.4)
+i=1JsiTi(Φy˙i(t)(ρ(yi(t)+,t),σ(yi(t)+,t))Φy˙i(t)(ρ(yi(t),t),σ(yi(t),t)))φ(yi(t),t)dt0.\displaystyle+\sum_{i=1}^{J}\int_{s_{i}}^{T_{i}}\biggl{(}\Phi_{\dot{y}_{i}(t)}\left(\rho(y_{i}(t)+,t),\sigma(y_{i}(t)+,t)\right)-\Phi_{\dot{y}_{i}(t)}\left(\rho(y_{i}(t)-,t),\sigma(y_{i}(t)-,t)\right)\biggr{)}\varphi(y_{i}(t),t)\mathinner{\mathrm{d}{t}}\geq 0.

Proof.  We split the reasoning in two steps.

Step 1. Suppose first that φ𝐂𝐜(Ω¯\m=1M𝒞m,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega}\backslash\cup_{m=1}^{M}\mathcal{C}_{m},\mathbb{R}^{+}). In this case, we write φ\varphi using the partition of unity (4.1). Fix i[[1;J]]i\in[\![1;J]\!]. Following the computations of Lemma 2.7, we obtain:

Ωi(|ρσ|tφi+Φ(ρ,σ)xφi)dxdt+{x:(x,0)Ωi}|ρo(x)σo(x)|φi(x,0)dx\displaystyle\iint_{\Omega_{i}}\biggl{(}|\rho-\sigma|\partial_{t}\varphi_{i}+\Phi(\rho,\sigma)\partial_{x}\varphi_{i}\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\{x\in\mathbb{R}\;:\;(x,0)\in\Omega_{i}\}}|\rho_{o}(x)-\sigma_{o}(x)|\varphi_{i}(x,0)\mathinner{\mathrm{d}{x}} (4.5)
+siTi(Φy˙i(t)(ρ(yi(t)+,t),σ(yi(t)+,t))Φy˙i(t)(ρ(yi(t),t),σ(yi(t),t)))φi(yi(t),t)dt0.\displaystyle+\int_{s_{i}}^{T_{i}}\biggl{(}\Phi_{\dot{y}_{i}(t)}\left(\rho(y_{i}(t)+,t),\sigma(y_{i}(t)+,t)\right)-\Phi_{\dot{y}_{i}(t)}\left(\rho(y_{i}(t)-,t),\sigma(y_{i}(t)-,t)\right)\biggr{)}\varphi_{i}(y_{i}(t),t)\mathinner{\mathrm{d}{t}}\geq 0.

Now, since φo\varphi_{o} vanishes along all the interfaces, standard computations lead to

Ωo(|ρσ|tφo+Φ(ρ,σ)xφo)dxdt+{x:(x,0)Ωo}|ρo(x)σo(x)|φo(x,0)dx0.\iint_{\Omega_{o}}\biggl{(}|\rho-\sigma|\partial_{t}\varphi_{o}+\Phi(\rho,\sigma)\partial_{x}\varphi_{o}\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\{x\in\mathbb{R}\;:\;(x,0)\in\Omega_{o}\}}|\rho_{o}(x)-\sigma_{o}(x)|\varphi_{o}(x,0)\mathinner{\mathrm{d}{x}}\geq 0. (4.6)

We now sum (4.5) (i[[1;J]]i\in[\![1;J]\!]) and (4.6) to obtain (4.4). This inequality is analogous to (2.11).

Step 2. Consider now φ𝐂𝐜(Ω¯,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}^{+}). Fix nn\in\mathbb{N}^{*}. From the first step, a classical approximation argument allows us to apply (4.4) with the Lipschitz test function

ψn(x,t)=(m=1Mδm,n(x,t))φ(x,t),\psi_{n}(x,t)=\left(\sum_{m=1}^{M}\delta_{m,n}(x,t)\right)\varphi(x,t),

where for all m[[1;M]]m\in[\![1;M]\!],

δm,n(x,t)={0ifdist1((x,t),𝒞m)<1nn(dist1((x,t),𝒞m)1n)if1ndist1((x,t),𝒞m)2n1ifdist1((x,t),𝒞m)>2n,\delta_{m,n}(x,t)=\left\{\begin{array}[]{ccc}0&\text{if}&\displaystyle{\text{dist}_{1}((x,t),\mathcal{C}_{m})<\frac{1}{n}}\\[5.0pt] \displaystyle{n\left(\text{dist}_{1}((x,t),\mathcal{C}_{m})-\frac{1}{n}\right)}&\text{if}&\displaystyle{\frac{1}{n}\leq\text{dist}_{1}((x,t),\mathcal{C}_{m})\leq\frac{2}{n}}\\[5.0pt] 1&\text{if}&\displaystyle{\text{dist}_{1}((x,t),\mathcal{C}_{m})>\frac{2}{n}},\end{array}\right.

where, by analogy with the proof of Lemma 2.7, dist1\text{dist}_{1} denotes the 2\mathbb{R}^{2} distance associated with the norm 1\|\cdot\|_{1}. We let n+n\to+\infty, keeping in mind that:

(m=1Mδm,n)φφ𝐋1(Ω,)n+ 0;m[[1;M]],δm,n𝐋1(Ω,2)=O(1n).\left\|\left(\sum_{m=1}^{M}\delta_{m,n}\right)\varphi-\varphi\right\|_{\mathbf{L}^{1}(\Omega,\mathbb{R})}{\ \underset{n\to+\infty}{\longrightarrow}\ }0;\quad\forall m\in[\![1;M]\!],\;\left\|\nabla\delta_{m,n}\right\|_{\mathbf{L}^{1}(\Omega,\mathbb{R}^{2})}=\mathop{}\mathopen{}O\mathopen{}\left(\frac{1}{n}\right).

Straightforward computations lead to (4.4) with φ𝐂𝐜(Ω¯,)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}), concluding the proof.   \square

Theorem 4.3.

Fix ρo,σo𝐋(,[0,1])\rho_{o},\sigma_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]). Let (qi)i[[1;J]](q_{i})_{i\in[\![1;J]\!]} and (qi)i[[1;J]](\overset{\sim}{q}_{i})_{i\in[\![1;J]\!]} be two family of constraints, where for all i[[1;J]]i\in[\![1;J]\!], qiq_{i}, qi𝐋(]si,Ti[,)\overset{\sim}{q}_{i}\in\mathbf{L}^{\infty}(\mathopen{]}s_{i},T_{i}\mathclose{[},\mathbb{R}). We denote by ρ\rho (resp. σ\sigma) a 𝒢\mathcal{G}-entropy solution to Problem (1.2) corresponding to initial datum ρo\rho_{o} (resp. σo\sigma_{o}) and constraints (qi)i[[1;J]](q_{i})_{i\in[\![1;J]\!]} (resp. (qi)i[[1;J]](\overset{\sim}{q}_{i})_{i\in[\![1;J]\!]}). Then for all T>0T>0, we have

ρ(T)σ(T)𝐋1()ρoσo𝐋1()+i=1J2siTi|qi(t)qi(t)|dt.\|\rho(T)-\sigma(T)\|_{\mathbf{L}^{1}(\mathbb{R})}\leq\|\rho_{o}-\sigma_{o}\|_{\mathbf{L}^{1}(\mathbb{R})}+\sum_{i=1}^{J}2\int_{s_{i}}^{T_{i}}\left|q_{i}(t)-\overset{\sim}{q}_{i}(t)\right|\mathinner{\mathrm{d}{t}}. (4.7)

In particular, Problem (1.2) admits at most one 𝒢\mathcal{G}-entropy solution.

Proof.  Estimate (4.7) follows from Kato inequality (4.4) with a suitable choice of test function and in light of the inequality:

i[[1;J]],for a.e.t]si,Ti[,\displaystyle\forall i\in[\![1;J]\!],\;\text{for a.e.}\;t\in\mathopen{]}s_{i},T_{i}\mathclose{[},
Φy˙i(t)(ρ(yi(t)+,t),σ(yi(t)+,t))Φy˙i(t)(ρ(yi(t),t),σ(yi(t),t))2|qi(t)qi(t)|,\displaystyle\Phi_{\dot{y}_{i}(t)}\left(\rho(y_{i}(t)+,t),\sigma(y_{i}(t)+,t)\right)-\Phi_{\dot{y}_{i}(t)}\left(\rho(y_{i}(t)-,t),\sigma(y_{i}(t)-,t)\right)\leq 2|q_{i}(t)-\overset{\sim}{q}_{i}(t)|,

see Theorem 2.8 and its proof.   \square

4.3 Proof of existence

Following the reasoning of Sections 2-3, we introduce a second definition of solutions, more suitable to prove existence.

Definition 4.4.

Let ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]). We say that ρ𝐋(Ω,[0,1])\rho\in\mathbf{L}^{\infty}(\Omega,[0,1]) is an admissible entropy solution to (1.2) if

(i) for all test functions φ𝐂𝐜(Ω¯,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega},\mathbb{R}^{+}) and κ[0,1]\kappa\in[0,1], the following entropy inequalities are verified:

0+\displaystyle\int_{0}^{+\infty}\int_{\mathbb{R}} (|ρκ|tφ+Φ(ρ,κ)xφ)dxdt+|ρo(x)κ|φ(x,0)dx\displaystyle\biggl{(}|\rho-\kappa|\partial_{t}\varphi+\Phi(\rho,\kappa)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{o}(x)-\kappa|\varphi(x,0)\mathinner{\mathrm{d}{x}} (4.8)
+i=1JsiTiy˙i(t)(κ,qi(t))φ(yi(t),t)dt0,\displaystyle+\sum_{i=1}^{J}\int_{s_{i}}^{T_{i}}\mathcal{R}_{\dot{y}_{i}(t)}(\kappa,q_{i}(t))\varphi(y_{i}(t),t)\mathinner{\mathrm{d}{t}}\geq 0,

where y˙i(κ,qi)\mathcal{R}_{\dot{y}_{i}}(\kappa,q_{i}) is defined in Definition 2.1 ;

(ii) for all test functions φ𝐂𝐜(Ω\m=1M𝒞m,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\Omega\backslash\cup_{m=1}^{M}\mathcal{C}_{m},\mathbb{R}^{+}), written under the form (4.1), the following constraint inequalities are verified for all i[[1;J]]i\in[\![1;J]\!]:

Ωi+(ρtφ+f(ρ)xφ)dxdtsiTiqi(t)φi(yi(t),t)dt,-\iint_{\Omega_{i}^{+}}\biggl{(}\rho\partial_{t}\varphi+f(\rho)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\leq\int_{s_{i}}^{T_{i}}q_{i}(t)\varphi_{i}(y_{i}(t),t)\mathinner{\mathrm{d}{t}}, (4.9)

where Ωi+:={(x,t)Ωi:x>yi(t)}\displaystyle{\Omega_{i}^{+}:=\left\{(x,t)\in\Omega_{i}\;:\;x>y_{i}(t)\right\}}.

Proposition 4.5.

Definition 4.1 and Definition 4.4 are equivalent. Moreover, in Definition 4.4 (i), it is equivalent that (4.8) holds with φ𝐂𝐜(Ω¯\m=1M𝒞m,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega}\backslash\cup_{m=1}^{M}\mathcal{C}_{m},\mathbb{R}^{+}).

Proof.  The proof of the equivalence of Definitions 4.1 and 4.4 is a straightforward adaptation of the proofs of Propositions 2.5-2.6. The last part of the statement follows using the same approximation argument described at the end of the proof of Lemma 4.2.   \square

Let us now turn to the proof of existence for admissible entropy solutions of (1.2). We make use of the precise study of Section 3 in the case of a single trajectory and build a finite volume scheme. We keep the notations of Section 3 when there is no ambiguity.

4.3.1 Construction of the mesh, definition of the scheme

For the sake of clarity, suppose that we only have two trajectories/constraints (yi,qi)(y_{i},q_{i}), i{1,2}i\in\{1,2\}, defined on [0,τ][0,\tau], which cross at time τ\tau. We denote by 𝒞\mathcal{C} this crossing point. Suppose also that this crossing point results in two additional trajectories/constraints (yi,qi)(y_{i},q_{i}), i{3,4}i\in\{3,4\}, defined on [τ,T][\tau,T], and which do not cross, as represented in Figure 4.

Let us fully make explicit the steps of the reasoning leading to the construction of our scheme in that situation. Suppose that λ=Δt/Δx\lambda=\Delta t/\Delta x is fixed and verifies the CFL condition

2(f𝐋+maxi[[1;4]]y˙i𝐋(]0,T[):=L)λ1.2\left(\underbrace{\|f^{\prime}\|_{\mathbf{L}^{\infty}}+\max_{i\in[\![1;4]\!]}\|\dot{y}_{i}\|_{\mathbf{L}^{\infty}(\mathopen{]}0,T\mathclose{[})}}_{:=L}\right)\lambda\leq 1. (4.10)

Set NN\in\mathbb{N} such that τ[tN,tN+1[\tau\in[t^{N},t^{N+1}\mathclose{[}. We divide the discussion in four parts.

Part 1. Introduce the number

N1:=inf{n:|yΔ1(tn)yΔ2(tn)|4Δx}.N_{1}:=\inf\left\{n\in\mathbb{N}\;:\;|y_{\Delta}^{1}(t^{n})-y_{\Delta}^{2}(t^{n})|\leq 4\Delta x\right\}.

The definition of N1N_{1} ensures that for all n[[0;N11]]n\in[\![0;N_{1}-1]\!], we can independently modify the mesh near the two trajectories yΔ1y_{\Delta}^{1} and yΔ2y_{\Delta}^{2}, as presented in Figure 5.

Consequently, we can simply define the approximate solution ρΔ\rho_{\Delta} on ×[0,tN11]\mathbb{R}\times[0,t^{N_{1}-1}] as the finite volume approximation of a conservation law, with initial datum ρo\rho_{o}, with flux constraints on two non-interacting trajectories, using the recipe of Section 3 for each trajectory/constraint.

Refer to caption
Figure 4: Illustration of the configuration.

Part 2. Fix now n[[N1;N]]n\in[\![N_{1};N]\!]. In these time intervals, since the two trajectories are too close to each other, one cannot modify the mesh in the neighborhood of one of them without affecting the other. However, the scheme has to be defined globally, so we proceed as described below.

  • First, introduce the mean trajectory and the new constraint:

    t[0,τ],y12(t):=y1(t)+y2(t)2;q12(t):=min{q1(t),q2(t)},\forall t\in[0,\tau],\quad y_{12}(t):=\frac{y_{1}(t)+y_{2}(t)}{2};\quad q_{12}(t):=\min\{q_{1}(t),q_{2}(t)\},

    represented in purple in Figure 5, before the crossing point (in red). The choice of taking the minimal level of constraint in the definition of q12q_{12} stems from the nature of the constrained problem; see however Remark 4.1 below.

  • Then, define ρΔ\rho_{\Delta} on ×[tN1,tN]\mathbb{R}\times[t^{N_{1}},t^{N}] as the finite volume approximation of the one trajectory/one constraint problem:

    {tρ+x(f(ρ))=0ρ(,tN1)=ρΔ(,tN11)(f(ρ)y˙12(t)ρ)|x=y12(t)q12(t)t]tN1,tN[,\left\{\begin{aligned} \partial_{t}\rho+\partial_{x}\left(f(\rho)\right)&=0\\ \rho(\cdot,t^{N_{1}})&=\rho_{\Delta}(\cdot,t^{N_{1}-1})\\ \left.\left(f(\rho)-\dot{y}_{12}(t)\rho\right)\right|_{x=y_{12}(t)}&\leq q_{12}(t)&t\in\mathopen{]}t^{N_{1}},t^{N}\mathclose{[},\end{aligned}\right.

    using exactly the recipe of Section 3.1.

Part 3. Introduce the number:

N2:=inf{n:n>Nand|yΔ3(tn)yΔ4(tn)|4Δx}.N_{2}:=\inf\left\{n\in\mathbb{N}\;:\;n>N\;\;\text{and}\;\;|y_{\Delta}^{3}(t^{n})-y_{\Delta}^{4}(t^{n})|\geq 4\Delta x\right\}.

For n[[N;N2]]n\in[\![N;N_{2}]\!], we are in the same situation as Part 2. We proceed to the same construction, mutatis mutandis.

  • As in Part 2, define the mean trajectory and the new constraint:

    t[τ,T],y34(t):=y3(t)+y4(t)2;q34(t):=min{q3(t),q4(t)},\forall t\in[\tau,T],\quad y_{34}(t):=\frac{y_{3}(t)+y_{4}(t)}{2};\quad q_{34}(t):=\min\{q_{3}(t),q_{4}(t)\},

    represented in purple in Figure 5, after the crossing point.

  • Define ρΔ\rho_{\Delta} on ×[tN,tN2]\mathbb{R}\times[t^{N},t^{N_{2}}] as the finite volume approximation of the one trajectory/one constraint problem:

    {tρ+x(f(ρ))=0ρ(,tN)=ρΔ(,tN)(f(ρ)y˙34(t)ρ)|x=y34(t)q34(t)t]tN,tN2[.\left\{\begin{aligned} \partial_{t}\rho+\partial_{x}\left(f(\rho)\right)&=0\\ \rho(\cdot,t^{N})&=\rho_{\Delta}(\cdot,t^{N})\\ \left.\left(f(\rho)-\dot{y}_{34}(t)\rho\right)\right|_{x=y_{34}(t)}&\leq q_{34}(t)&t\in\mathopen{]}t^{N},t^{N_{2}}\mathclose{[}.\end{aligned}\right.
Refer to caption
Figure 5: Illustration of the local modifications of the mesh.

Part 4. Finally, ρΔ\rho_{\Delta} is defined on ×[tN2,T]\mathbb{R}\times[t^{N_{2}},T] like in Part 1 with y3,q3,ρΔ(,tN2)y_{3},q_{3},\rho_{\Delta}(\cdot,t^{N_{2}}) (resp. y4,q4y_{4},q_{4}) playing the role of y1,q1,ρoy_{1},q_{1},\rho_{o} (resp. of y2,q2y_{2},q_{2}).

Remark 4.1.

Let us stress out that the details of the treatment done in Parts 2-3 do not play any significant role in the convergence proof below thanks to the choice of test functions vanishing at neighborhood of the crossing points, see Proposition 4.5. Consequently, taking the mean trajectory and the minimum of the constraint is merely an example aiming at preserving some consistency while keeping the scheme simple to understand and implement.

The general case of a finite number of interfaces (locally finite number can be easily included) is treated in the same way, leading to a pattern with the uniform rectangular mesh adapted to each of the interfaces Γi\Gamma_{i}, i[[1;J]]i\in[\![1;J]\!] except for small (in terms of the number of impacted mesh cells) neighborhoods of the crossing points 𝒞m\mathcal{C}_{m}, m[[1;M]]m\in[\![1;M]\!].

4.3.2 Proof of convergence

Theorem 4.6.

Fix T>0T>0, f𝐂2([0,1],+)f\in\mathbf{C}^{2}([0,1],\mathbb{R}^{+}) satisfying (1.1)-(3.14) and ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]). Let (yi,qi)i[[1;J]](y_{i},q_{i})_{i\in[\![1;J]\!]} be a finite family of trajectories and constraints defined on ]si,Ti[\mathopen{]}s_{i},T_{i}\mathclose{[} (0si<Ti0\leq s_{i}<T_{i}). We suppose that for all i[[1;J]]i\in[\![1;J]\!], yi𝐖1,(]si,Ti[,)y_{i}\in\mathbf{W}^{1,\infty}(\mathopen{]}s_{i},T_{i}\mathclose{[},\mathbb{R}) and qi𝐋(]si,Ti[,+)q_{i}\in\mathbf{L}^{\infty}(\mathopen{]}s_{i},T_{i}\mathclose{[},\mathbb{R}^{+}). Suppose also that the interfaces (Γi)i(\Gamma_{i})_{i} defined by the trajectories (yi)i(y_{i})_{i} have a finite number of crossing points. Then as Δ0\Delta\to 0 while satisfying the CFL condition

2(f𝐋+maxi[[0;J]]y˙i𝐋(]0,T[):=L)λ1,2\left(\underbrace{\|f^{\prime}\|_{\mathbf{L}^{\infty}}+\max_{i\in[\![0;J]\!]}\|\dot{y}_{i}\|_{\mathbf{L}^{\infty}(\mathopen{]}0,T\mathclose{[})}}_{:=L}\right)\lambda\leq 1,

the sequence (ρΔ)Δ(\rho_{\Delta})_{\Delta} constructed by the procedure of Section 4.3.1 converges a.e. on Ω\Omega to the admissible entropy solution to (1.2).

Proof.  We make use of the fact that in Definition 4.4, we only need to consider test functions that vanish at a neighborhood of the crossing points (this is the key observation leading to Remark 4.1 here above).

(i) Proof of the entropy inequalities. Fix φ𝐂𝐜(Ω¯\m=1M𝒞m,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\overline{\Omega}\backslash\cup_{m=1}^{M}\mathcal{C}_{m},\mathbb{R}^{+}), written as φ=φo+i=1Jφi\varphi=\varphi_{o}+\sum_{i=1}^{J}\varphi_{i}, using the appropriate partition of unity, see Section 4.1. Since φo\varphi_{o} vanishes along all the interfaces, ρΔ\rho_{\Delta} verifies inequality (3.11) with 0\mathcal{R}\equiv 0 on the domain Ωo\Omega_{o} and with test function φo\varphi_{o}. Indeed, for a sufficiently small Δx>0\Delta x>0, the scheme we constructed in the previous section reduces to a standard finite volume in Ωo\Omega_{o}. Fix now i[[1;J]]i\in[\![1;J]\!]. Since φi\varphi_{i} vanishes along all the interfaces but Γi\Gamma_{i}, ρΔ\rho_{\Delta} verifies inequality (3.11) with reminder term sΔi(κ,qΔi)\mathcal{R}_{s_{\Delta}^{i}}(\kappa,q_{\Delta}^{i}) along the trajectory yΔiy_{\Delta}^{i} on the domain Ωi\Omega_{i} and with test function φi\varphi_{i}, due to the analysis of Section 3; indeed, in the support of the test function, our scheme for the multi-interface problem reduces to the scheme for the single-interface problem. By summing these previous inequalities, we obtain an approximate version of (4.8) verified by ρΔ\rho_{\Delta}:

0+\displaystyle\int_{0}^{+\infty}\int_{\mathbb{R}} (|ρΔκ|tφ+ΦΔ(ρΔ,κ)xφ)dxdt+|ρΔ0(x)κ|φ(x,0)dx\displaystyle\biggl{(}|\rho_{\Delta}-\kappa|\partial_{t}\varphi+\Phi_{\Delta}(\rho_{\Delta},\kappa)\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}+\int_{\mathbb{R}}|\rho_{\Delta}^{0}(x)-\kappa|\varphi(x,0)\mathinner{\mathrm{d}{x}} (4.11)
+i=1JsiTisΔi(t)(κ,qΔi(t))φ(yΔi(t),t)dtO(Δx)+O(Δt).\displaystyle+\sum_{i=1}^{J}\int_{s_{i}}^{T_{i}}\mathcal{R}_{s_{\Delta}^{i}(t)}(\kappa,q_{\Delta}^{i}(t))\varphi(y_{\Delta}^{i}(t),t)\mathinner{\mathrm{d}{t}}\geq\mathop{}\mathopen{}O\mathopen{}\left(\Delta x\right)+\mathop{}\mathopen{}O\mathopen{}\left(\Delta t\right).

(ii) Proof of the weak constraint inequalities. Let φ𝐂𝐜(Ω\m=1M𝒞m,+)\varphi\in\mathbf{C}_{\mathbf{c}}^{\infty}(\Omega\backslash\cup_{m=1}^{M}\mathcal{C}_{m},\mathbb{R}^{+}), written under the form (4.1). Fix i[[1;J]]i\in[\![1;J]\!]. Since φi\varphi_{i} vanishes along all the interfaces but Γi\Gamma_{i}, for a sufficiently small Δx\Delta x, ρΔ\rho_{\Delta} verifies inequality (3.13) with constraint qΔiq_{\Delta}^{i} along the trajectory yΔiy_{\Delta}^{i} on the domain Ωi+\Omega_{i}^{+} and with test function φi\varphi_{i}. We obtain an approximate version of (4.12) verified by ρΔ\rho_{\Delta}:

Ωi+(ρΔtφ+𝐅Δ(ρΔ)xφ)dxdtsiTiqΔi(t)φi(yΔi(t),t)dt+O(Δx)+O(Δt).-\iint_{\Omega_{i}^{+}}\biggl{(}\rho_{\Delta}\partial_{t}\varphi+\mathbf{F}_{\Delta}(\rho_{\Delta})\partial_{x}\varphi\biggr{)}\mathinner{\mathrm{d}{x}}\mathinner{\mathrm{d}{t}}\leq\int_{s_{i}}^{T_{i}}q_{\Delta}^{i}(t)\varphi_{i}(y_{\Delta}^{i}(t),t)\mathinner{\mathrm{d}{t}}+\mathop{}\mathopen{}O\mathopen{}\left(\Delta x\right)+\mathop{}\mathopen{}O\mathopen{}\left(\Delta t\right). (4.12)

(iii) Compactness and convergence. Compactness of the sequence (ρΔ)Δ(\rho_{\Delta})_{\Delta} follows directly from the study of Section 3.4 where we derived local 𝐁𝐕\mathbf{BV} bounds for (ρΔ)Δ(\rho_{\Delta})_{\Delta} under the assumption (3.14). Indeed, these local bounds lead to compactness in the domain complementary to the interfaces, we only use the fact that the interfaces together with the crossing points form a closed subset of Ω\Omega with zero Lebesgue measure. Once the a.e. convergence (up to a subsequence) on Ω\Omega to some ρ𝐋(Ω,[0,1])\rho\in\mathbf{L}^{\infty}(\Omega,[0,1]) obtained, we simply pass to the limit in (4.11)-(4.12). This proves that ρ\rho is an admissible solution to (1.2). By the uniqueness of Theorem 4.3, the whole sequence converges to ρ\rho. This concludes the proof.   \square

Corollary 4.7.

Fix T>0T>0, f𝐂2([0,1],+)f\in\mathbf{C}^{2}([0,1],\mathbb{R}^{+}) satisfying (1.1)-(3.14) and ρo𝐋(,[0,1])\rho_{o}\in\mathbf{L}^{\infty}(\mathbb{R},[0,1]). Let (yi,qi)i[[1;J]](y_{i},q_{i})_{i\in[\![1;J]\!]} be a finite family of trajectories and constraints defined on ]si,Ti[\mathopen{]}s_{i},T_{i}\mathclose{[} (0si<Ti0\leq s_{i}<T_{i}). We suppose that for all i[[1;J]]i\in[\![1;J]\!], yi𝐖1,(]si,Ti[,)y_{i}\in\mathbf{W}^{1,\infty}(\mathopen{]}s_{i},T_{i}\mathclose{[},\mathbb{R}) and qi𝐋(]si,Ti[,+)q_{i}\in\mathbf{L}^{\infty}(\mathopen{]}s_{i},T_{i}\mathclose{[},\mathbb{R}^{+}). Suppose also that the interfaces (Γi)i(\Gamma_{i})_{i} defined by the trajectories (yi)i(y_{i})_{i} have a finite number of crossing points. Then Problem (1.2) admits a unique admissible entropy solution.

Proof.  Existence comes from Theorem 4.6 while uniqueness is established by Theorem 4.3.   \square

5 Numerical experiment with crossing trajectories

In this section, we perform a numerical test to illustrate the scheme analyzed in Section 3 and Section 4.3. We take the GNL flux f(ρ)=ρ(1ρ)f(\rho)=\rho(1-\rho).

We model the following situation. A vehicle breaks down on a road and reduces by half the surrounding traffic flow, which initial state is given by ρo=0.8×𝟏[1,3]\displaystyle{\rho_{o}=0.8\times\mathbf{1}_{[1,3]}}. At some point, a tow truck comes to move the immobile vehicle. We summarized this situation in Figure 6. Notice the time interval in which q30.1q_{3}\equiv 0.1. This corresponds to the time needed for the tow truck to move the vehicle. Remark also that the value of the constraint on this time interval is smaller than the one when only the broken down vehicle was reducing the traffic flow.

Refer to caption
Figure 6: A tow truck comes moving an immobile vehicle.

The evolution of the numerical solution is represented in Figure 7. Let us comment on the profile of the numerical solution.

  • At first (0t5.800\leq t\leq 5.80), the solution is composed of traveling waves separated by a stationary non-classical shock located at the immobile vehicle position.

  • When the tow truck catches up with the vehicle (6.30t8.06.30\leq t\leq 8.0), the profile of the numerical solution is the same, but the greater value of the constraint in this time interval changes the magnitude of the non-classical shock; at this point the combined presence of both the tow truck and the immobile vehicle clogs the traffic flow even more.

  • Finally, once the tow truck starts again (t>8.0t>8.0), the traffic congestion is reduced.

Notice at time t=7.44t=7.44 the small artifact (circled in red in Figure 7) created by Parts 2-3 in the construction of the approximate solution and reproduced by the scheme. This highlights the fact that even if the treatment of the crossing points brings inconsistencies or artifacts to the numerical solution, these undesired effects are not amplified by the scheme, and become negligible when one refines the mesh.

Refer to caption
Figure 7: The numerical solution at different fixed times; for an animated evolution of the solution, follow: https://www.abrahamsylla.com/numerical-simulations

Appendix A Proof of the OSL bound

We prove in this appendix Lemma 3.5. All the notations are taken from Sections 3.1 and 3.4. The proof is a simple rewriting of the proof of [22, Lemma 4.2].

It will be convenient to write the Engquist-Osher flux under the form:

a,b[0,1],𝐄𝐎(a,b)=(f(aρ¯)f(ρ¯)2)q+(a)+(f(bρ¯)f(ρ¯)2)q(b),\forall a,b\in[0,1],\quad\mathbf{EO}(a,b)=\underbrace{\left(f(a\wedge\overline{\rho})-\frac{f(\overline{\rho})}{2}\right)}_{q_{+}(a)}+\underbrace{\left(f(b\vee\overline{\rho})-\frac{f(\overline{\rho})}{2}\right)}_{q_{-}(b)},

so that for all nn\in\mathbb{N}, when j^n+1j\in\widehat{\mathbb{Z}}_{n+1}, the scheme (3.4) can be rewritten as:

ρj+1/2n+1=ρj+1/2nλ(q+(ρj+1/2n)+q(ρj+3/2n)q+(ρj1/2n)q(ρj+1/2n)).\rho_{j+1/2}^{n+1}=\rho_{j+1/2}^{n}-\lambda\biggl{(}q_{+}\left(\rho_{j+1/2}^{n}\right)+q_{-}\left(\rho_{j+3/2}^{n}\right)-q_{+}\left(\rho_{j-1/2}^{n}\right)-q_{-}\left(\rho_{j+1/2}^{n}\right)\biggr{)}. (A.1)
Lemma A.1.

For all nn\in\mathbb{N} and jj\in\mathbb{Z}, we have

ρj1/2nρj+1/2n1λμandDjn1λμ.\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}\leq\frac{1}{\lambda\mu}\quad\text{and}\quad D_{j}^{n}\leq\frac{1}{\lambda\mu}. (A.2)

Proof.  Indeed, using first the uniform convexity of ff and then the CFL condition (3.2), we can write:

(ρj1/2nρj+1/2n)μρj+1/2nρj1/2nf′′(u)du2f𝐋ΔxΔt,\left(\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}\right)\mu\leq-\int_{\rho_{j+1/2}^{n}}^{\rho_{j-1/2}^{n}}f^{\prime\prime}(u)\mathinner{\mathrm{d}{u}}\leq 2\|f^{\prime}\|_{\mathbf{L}^{\infty}}\leq\frac{\Delta x}{\Delta t},

from which we deduce (A.2).   \square

Lemma A.2.

Let nn\in\mathbb{N}, j^n+1j\in\widehat{\mathbb{Z}}_{n+1}, a=λμ4\displaystyle{a=\frac{\lambda\mu}{4}} and ψ(x)=xax2\displaystyle{\psi(x)=x-ax^{2}}. Then

Djn+1ψ(max{Dj1n,Djn,Dj+1n}).D_{j}^{n+1}\leq\psi\left(\max\left\{D_{j-1}^{n},D_{j}^{n},D_{j+1}^{n}\right\}\right). (A.3)

Proof.  We divide the proof in three steps.

Step 1: The function ψ\psi is nonnegative on [0,1/a][0,1/a] and nondecreasing on [0,1/(2a)][0,1/(2a)]. Note that by (A.2), max{Dj1n,Djn,Dj+1n}1/(4a)\displaystyle{\max\left\{D_{j-1}^{n},D_{j}^{n},D_{j+1}^{n}\right\}\leq 1/(4a)}, which will allow us to use the monotonicity of ψ\psi.

Step 2. We assume that

ρj+1/2nρj+3/2n0andρj3/2nρj1/2n0,\rho_{j+1/2}^{n}-\rho_{j+3/2}^{n}\geq 0\quad\text{and}\quad\rho_{j-3/2}^{n}-\rho_{j-1/2}^{n}\geq 0, (A.4)

and we are going to prove that (A.3) holds. Using the uniform concavity assumption of ff, we can write that

a,b[0,1],q+(b)q+(a)(bρ¯aρ¯)f(aρ¯)μ2(bρ¯aρ¯)2.\forall a,b\in[0,1],\quad q_{+}(b)-q_{+}(a)\leq(b\wedge\overline{\rho}-a\wedge\overline{\rho})f^{\prime}(a\wedge\overline{\rho})-\frac{\mu}{2}(b\wedge\overline{\rho}-a\wedge\overline{\rho})^{2}. (A.5)

A similar inequality holds for qq_{-} as well. Using (A.1), we obtain:

ρj1/2n+1ρj+1/2n+1\displaystyle\rho_{j-1/2}^{n+1}-\rho_{j+1/2}^{n+1} =ρj1/2nρj+1/2n\displaystyle=\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n} (A.6)
λ(q+(ρj1/2n)q+(ρj3/2n)q+(ρj+1/2n)+q+(ρj1/2n))\displaystyle-\lambda\left(q_{+}\left(\rho_{j-1/2}^{n}\right)-q_{+}\left(\rho_{j-3/2}^{n}\right)-q_{+}\left(\rho_{j+1/2}^{n}\right)+q_{+}\left(\rho_{j-1/2}^{n}\right)\right)
λ(q(ρj+1/2n)q(ρj1/2n)q(ρj+3/2n)+q(ρj+1/2n))\displaystyle-\lambda\left(q_{-}\left(\rho_{j+1/2}^{n}\right)-q_{-}\left(\rho_{j-1/2}^{n}\right)-q_{-}\left(\rho_{j+3/2}^{n}\right)+q_{-}\left(\rho_{j+1/2}^{n}\right)\right)
=ρj1/2nρj+1/2n\displaystyle=\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}
+λ{(q+(ρj+1/2n)q+(ρj1/2n))+(q+(ρj3/2n)q+(ρj1/2n))\displaystyle+\lambda\left\{\left(q_{+}\left(\rho_{j+1/2}^{n}\right)-q_{+}\left(\rho_{j-1/2}^{n}\right)\right)+\left(q_{+}\left(\rho_{j-3/2}^{n}\right)-q_{+}\left(\rho_{j-1/2}^{n}\right)\right)\right.
+(q(ρj+3/2n)q(ρj+1/2n))+(q(ρj1/2n)q(ρj+1/2n))}\displaystyle\left.+\left(q_{-}\left(\rho_{j+3/2}^{n}\right)-q_{-}\left(\rho_{j+1/2}^{n}\right)\right)+\left(q_{-}\left(\rho_{j-1/2}^{n}\right)-q_{-}\left(\rho_{j+1/2}^{n}\right)\right)\right\}
ρj1/2nρj+1/2n\displaystyle\leq\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}
+λ(ρj+1/2nρ¯ρj1/2nρ¯)f(ρj1/2nρ¯)λμ2(ρj+1/2nρ¯ρj1/2nρ¯)2\displaystyle+\lambda\left(\rho_{j+1/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right)f^{\prime}(\rho_{j-1/2}^{n}\wedge\overline{\rho})-\frac{\lambda\mu}{2}\left(\rho_{j+1/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right)^{2}
+λ(ρj3/2nρ¯ρj1/2nρ¯)f(ρj1/2nρ¯)λμ2(ρj3/2nρ¯ρj1/2nρ¯)2\displaystyle+\lambda\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right)f^{\prime}(\rho_{j-1/2}^{n}\wedge\overline{\rho})-\frac{\lambda\mu}{2}\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right)^{2}
+λ(ρj+3/2nρ¯ρj+1/2nρ¯)f(ρj+1/2nρ¯)λμ2(ρj+3/2nρ¯ρj+1/2nρ¯)2\displaystyle+\lambda\left(\rho_{j+3/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)f^{\prime}(\rho_{j+1/2}^{n}\vee\overline{\rho})-\frac{\lambda\mu}{2}\left(\rho_{j+3/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)^{2}
+λ(ρj1/2nρ¯ρj+1/2nρ¯)f(ρj+1/2nρ¯)λμ2(ρj1/2nρ¯ρj+1/2nρ¯)2,\displaystyle+\lambda\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)f^{\prime}(\rho_{j+1/2}^{n}\vee\overline{\rho})-\frac{\lambda\mu}{2}\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)^{2},

where the last inequality comes from using (A.5). The proof now reduces to four cases, depending on the ordering of ρ¯\overline{\rho}, ρj1/2n\rho_{j-1/2}^{n} and ρj1/2n\rho_{j-1/2}^{n}.

Case 1: ρ¯ρj1/2n,ρj+1/2n\overline{\rho}\geq\rho_{j-1/2}^{n},\ \rho_{j+1/2}^{n}. Under assumption (A.4), we have ρ¯ρj+3/2n\overline{\rho}\geq\rho_{j+3/2}^{n} as well. Inequality (A.6) becomes:

ρj1/2n+1ρj+1/2n+1\displaystyle\rho_{j-1/2}^{n+1}-\rho_{j+1/2}^{n+1} (1λf(ρj1/2n))(ρj1/2nρj+1/2n)+λf(ρj1/2n)(ρj3/2nρ¯ρj1/2n)\displaystyle\leq\left(1-\lambda f^{\prime}(\rho_{j-1/2}^{n})\right)\left(\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}\right)+\lambda f^{\prime}(\rho_{j-1/2}^{n})\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\right) (A.7)
λμ2((ρj1/2nρj+1/2n)2+(ρj3/2nρ¯ρj1/2n)2)\displaystyle-\frac{\lambda\mu}{2}\left(\left(\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}\right)^{2}+\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\right)^{2}\right)
(1λf(ρj1/2n))(ρj1/2nρj+1/2n)+λf(ρj1/2n)(ρj3/2nρ¯ρj1/2n)\displaystyle\leq\left(1-\lambda f^{\prime}(\rho_{j-1/2}^{n})\right)\left(\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}\right)+\lambda f^{\prime}(\rho_{j-1/2}^{n})\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\right)
λμ4((ρj1/2nρj+1/2n)2+(ρj3/2nρ¯ρj1/2n)2)\displaystyle-\frac{\lambda\mu}{4}\left(\left(\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}\right)^{2}+\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\right)^{2}\right)
(1λf(ρj1/2n))(ρj1/2nρj+1/2n)+λf(ρj1/2n(ρj3/2nρ¯ρj1/2n)\displaystyle\leq\left(1-\lambda f^{\prime}(\rho_{j-1/2}^{n})\right)\left(\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}\right)+\lambda f^{\prime}(\rho_{j-1/2}^{n}\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\right)
λμ4max{ρj1/2nρj+1/2n,ρj3/2nρ¯ρj1/2n}2,\displaystyle-\frac{\lambda\mu}{4}\max\left\{\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n},\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\right\}^{2},

where the last inequality comes from the bound: a2+b2max{a,b}2\displaystyle{a^{2}+b^{2}\geq\max\{a,b\}^{2}}. The CFL condition (3.2) ensures that the two first terms of the right-hand side of the last inequality are a convex combination of (ρj1/2nρj+1/2n)\displaystyle{\left(\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}\right)} and (ρj3/2nρ¯ρj1/2n)\displaystyle{\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\right)}. Consequently, inequality (A.7) then becomes

ρj1/2n+1ρj+1/2n+1ψ(max{ρj1/2nρj+1/2n,ρj3/2nρ¯ρj1/2n}).\rho_{j-1/2}^{n+1}-\rho_{j+1/2}^{n+1}\leq\psi\left(\max\left\{\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n},\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\right\}\right).

Since ρj3/2nρ¯ρj1/2nρj3/2nρj1/2n\displaystyle{\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\leq\rho_{j-3/2}^{n}-\rho_{j-1/2}^{n}}, the monotonicity of ψ\psi ensures that

ρj1/2n+1ρj+1/2n+1\displaystyle\rho_{j-1/2}^{n+1}-\rho_{j+1/2}^{n+1} ψ(max{ρj1/2nρj+1/2n,ρj3/2nρj1/2n})\displaystyle\leq\psi\left(\max\left\{\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n},\rho_{j-3/2}^{n}-\rho_{j-1/2}^{n}\right\}\right)
ψ(max{Dj1n,Djn})\displaystyle\leq\psi\left(\max\left\{D_{j-1}^{n},D_{j}^{n}\right\}\right)
ψ(max{Dj1n,Djn,Dj+1n}).\displaystyle\leq\psi\left(\max\left\{D_{j-1}^{n},D_{j}^{n},D_{j+1}^{n}\right\}\right).

Since the right-hand side of this inequality is nonnegative, we can replace its left-hand side by Djn+1D_{j}^{n+1}, which concludes the proof in this case.

Case 2: ρ¯ρj1/2n,ρj+1/2n\overline{\rho}\leq\rho_{j-1/2}^{n},\ \rho_{j+1/2}^{n}. The proof of in this case similar to the last one so we omit the details.

Case 3: ρj+1/2nρ¯ρj1/2n\rho_{j+1/2}^{n}\leq\overline{\rho}\leq\rho_{j-1/2}^{n}. Under Assumption (A.4), we have the following ordering:

ρj+3/2nρj+1/2nρ¯ρj1/2nρj3/2n.\rho_{j+3/2}^{n}\leq\rho_{j+1/2}^{n}\leq\overline{\rho}\leq\rho_{j-1/2}^{n}\leq\rho_{j-3/2}^{n}.

Inequality (A.6) becomes

ρj1/2n+1ρj+1/2n+1\displaystyle\rho_{j-1/2}^{n+1}-\rho_{j+1/2}^{n+1} ρj1/2nρj+1/2nλμ2((ρj1/2nρ¯)2+(ρ¯ρj+1/2n)2)\displaystyle\leq\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}-\frac{\lambda\mu}{2}\left((\rho_{j-1/2}^{n}-\overline{\rho})^{2}+(\overline{\rho}-\rho_{j+1/2}^{n})^{2}\right)
ρj1/2nρj+1/2nλμ4(ρj1/2nρj+1/2n)2,\displaystyle\leq\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}-\frac{\lambda\mu}{4}(\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n})^{2},

where we used the inequality 2(a2+b2)(a+b)22(a^{2}+b^{2})\geq(a+b)^{2}. From here, we can conclude as in Case 1.

Case 4: ρj1/2nρ¯ρj+1/2n\rho_{j-1/2}^{n}\leq\overline{\rho}\leq\rho_{j+1/2}^{n}. Using the decomposition

ρj1/2nρj+1/2n=(ρj1/2nρ¯ρj+1/2nρ¯)+(ρj1/2nρ¯ρj+1/2nρ¯),\rho_{j-1/2}^{n}-\rho_{j+1/2}^{n}=(\rho_{j-1/2}^{n}\wedge\overline{\rho}-\rho_{j+1/2}^{n}\wedge\overline{\rho})+(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}),

inequality (A.6) becomes

ρj1/2n+1ρj+1/2n+1\displaystyle\rho_{j-1/2}^{n+1}-\rho_{j+1/2}^{n+1} (1λf(ρj1/2n))(ρj1/2nρ¯ρj+1/2nρ¯)+λf(ρj1/2n)(ρj3/2nρ¯ρj1/2nρ¯)\displaystyle\leq\left(1-\lambda f^{\prime}(\rho_{j-1/2}^{n})\right)\left(\rho_{j-1/2}^{n}\wedge\overline{\rho}-\rho_{j+1/2}^{n}\wedge\overline{\rho}\right)+\lambda f^{\prime}(\rho_{j-1/2}^{n})\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right) (A.8)
+(1+λf(ρj+1/2n))(ρj1/2nρ¯ρj+1/2nρ¯)λf(ρj+1/2n)(ρj+1/2nρ¯ρj+3/2nρ¯)\displaystyle+\left(1+\lambda f^{\prime}(\rho_{j+1/2}^{n})\right)\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)-\lambda f^{\prime}(\rho_{j+1/2}^{n})\left(\rho_{j+1/2}^{n}\vee\overline{\rho}-\rho_{j+3/2}^{n}\vee\overline{\rho}\right)
λμ2{(ρj1/2nρ¯ρj+1/2nρ¯)2+(ρj3/2nρ¯ρj1/2nρ¯)2\displaystyle-\frac{\lambda\mu}{2}\left\{\left(\rho_{j-1/2}^{n}\wedge\overline{\rho}-\rho_{j+1/2}^{n}\wedge\overline{\rho}\right)^{2}+\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right)^{2}\right.
+(ρj1/2nρ¯ρj+1/2nρ¯)2+(ρj+1/2nρ¯ρj+3/2nρ¯)2}\displaystyle\left.+\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)^{2}+\left(\rho_{j+1/2}^{n}\vee\overline{\rho}-\rho_{j+3/2}^{n}\vee\overline{\rho}\right)^{2}\right\}
(1λf(ρj1/2n))(ρj1/2nρ¯ρj+1/2nρ¯)+λf(ρj1/2n)(ρj3/2nρ¯ρj1/2nρ¯)\displaystyle\leq\left(1-\lambda f^{\prime}(\rho_{j-1/2}^{n})\right)\left(\rho_{j-1/2}^{n}\wedge\overline{\rho}-\rho_{j+1/2}^{n}\wedge\overline{\rho}\right)+\lambda f^{\prime}(\rho_{j-1/2}^{n})\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right)
+(1+λf(ρj+1/2n))(ρj1/2nρ¯ρj+1/2nρ¯)λf(ρj+1/2n)(ρj+1/2nρ¯ρj+3/2nρ¯)\displaystyle+\left(1+\lambda f^{\prime}(\rho_{j+1/2}^{n})\right)\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)-\lambda f^{\prime}(\rho_{j+1/2}^{n})\left(\rho_{j+1/2}^{n}\vee\overline{\rho}-\rho_{j+3/2}^{n}\vee\overline{\rho}\right)
λμ2{(ρj1/2nρ¯ρj+1/2nρ¯)2+(ρj1/2nρ¯ρj+1/2nρ¯)2}.\displaystyle-\frac{\lambda\mu}{2}\left\{\left(\rho_{j-1/2}^{n}\wedge\overline{\rho}-\rho_{j+1/2}^{n}\wedge\overline{\rho}\right)^{2}+\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)^{2}\right\}.

The CFL condition (3.2) and the ordering ρj+1/2nρ¯ρj1/2n\rho_{j+1/2}^{n}\leq\overline{\rho}\leq\rho_{j-1/2}^{n} result in

(1λf(ρj1/2n))(ρj1/2nρ¯ρj+1/2nρ¯)0and(1+λf(ρj+1/2n))(ρj1/2nρ¯ρj+1/2nρ¯)0\left(1-\lambda f^{\prime}(\rho_{j-1/2}^{n})\right)\left(\rho_{j-1/2}^{n}\wedge\overline{\rho}-\rho_{j+1/2}^{n}\wedge\overline{\rho}\right)\leq 0\quad\text{and}\quad\left(1+\lambda f^{\prime}(\rho_{j+1/2}^{n})\right)\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)\leq 0

so we can replace (A.8) by

ρj1/2n+1ρj+1/2n+1\displaystyle\rho_{j-1/2}^{n+1}-\rho_{j+1/2}^{n+1} λf(ρj1/2n)(ρj3/2nρ¯ρj1/2nρ¯)λf(ρj+1/2n)(ρj+1/2nρ¯ρj+3/2nρ¯)\displaystyle\leq\lambda f^{\prime}(\rho_{j-1/2}^{n})\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right)-\lambda f^{\prime}(\rho_{j+1/2}^{n})\left(\rho_{j+1/2}^{n}\vee\overline{\rho}-\rho_{j+3/2}^{n}\vee\overline{\rho}\right)
λμ2{(ρj1/2nρ¯ρj+1/2nρ¯)2+(ρj1/2nρ¯ρj+1/2nρ¯)2}\displaystyle-\frac{\lambda\mu}{2}\left\{\left(\rho_{j-1/2}^{n}\wedge\overline{\rho}-\rho_{j+1/2}^{n}\wedge\overline{\rho}\right)^{2}+\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)^{2}\right\}
12((ρj3/2nρ¯ρj1/2nρ¯)+(ρj+1/2nρ¯ρj+3/2nρ¯))\displaystyle\leq\frac{1}{2}\left(\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right)+\left(\rho_{j+1/2}^{n}\vee\overline{\rho}-\rho_{j+3/2}^{n}\vee\overline{\rho}\right)\right)
λμ4{(ρj1/2nρ¯ρj+1/2nρ¯)2+(ρj1/2nρ¯ρj+1/2nρ¯)2}\displaystyle-\frac{\lambda\mu}{4}\left\{\left(\rho_{j-1/2}^{n}\wedge\overline{\rho}-\rho_{j+1/2}^{n}\wedge\overline{\rho}\right)^{2}+\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)^{2}\right\}
ψ(max{(ρj3/2nρ¯ρj1/2nρ¯),(ρj1/2nρ¯ρj+1/2nρ¯)}),\displaystyle\leq\psi\left(\max\left\{\left(\rho_{j-3/2}^{n}\wedge\overline{\rho}-\rho_{j-1/2}^{n}\wedge\overline{\rho}\right),\left(\rho_{j-1/2}^{n}\vee\overline{\rho}-\rho_{j+1/2}^{n}\vee\overline{\rho}\right)\right\}\right),

and we exploit the monotonicity of ψ\psi to conclude.

Step 3: We no longer assume (A.4), and we get back to the general case. Let us introduce

uj3/2n=ρj3/2nρj1/2n,uj1/2n=ρj1/2n,uj+1/2n=ρj+1/2n,uj+3/2n=ρj+3/2nρj1/2n,u_{j-3/2}^{n}=\rho_{j-3/2}^{n}\vee\rho_{j-1/2}^{n},\;u_{j-1/2}^{n}=\rho_{j-1/2}^{n},\;u_{j+1/2}^{n}=\rho_{j+1/2}^{n},\;u_{j+3/2}^{n}=\rho_{j+3/2}^{n}\wedge\rho_{j-1/2}^{n},

and

uj1/2n+1=H(uj3/2n,uj1/2n,uj+1/2n);uj+1/2n+1=H(uj1/2n,uj+1/2n,uj+3/2n).u_{j-1/2}^{n+1}=H(u_{j-3/2}^{n},u_{j-1/2}^{n},u_{j+1/2}^{n});\quad u_{j+1/2}^{n+1}=H(u_{j-1/2}^{n},u_{j+1/2}^{n},u_{j+3/2}^{n}).

Using the monotonicity of HH, we get:

ρj1/2n+1ρj+1/2n+1\displaystyle\rho_{j-1/2}^{n+1}-\rho_{j+1/2}^{n+1} =H(ρj3/2n,ρj1/2n,ρj+1/2n)H(ρj1/2n,ρj+1/2n,ρj+3/2n)\displaystyle=H(\rho_{j-3/2}^{n},\rho_{j-1/2}^{n},\rho_{j+1/2}^{n})-H(\rho_{j-1/2}^{n},\rho_{j+1/2}^{n},\rho_{j+3/2}^{n})
H(uj3/2n,uj1/2n,uj+1/2n)H(uj1/2n,uj+1/2n,uj+3/2n)=uj1/2n+1uj+1/2n+1.\displaystyle\leq H(u_{j-3/2}^{n},u_{j-1/2}^{n},u_{j+1/2}^{n})-H(u_{j-1/2}^{n},u_{j+1/2}^{n},u_{j+3/2}^{n})=u_{j-1/2}^{n+1}-u_{j+1/2}^{n+1}.

Since uj+1/2nuj+3/2n0u_{j+1/2}^{n}-u_{j+3/2}^{n}\geq 0 and uj3/2nuj1/2n0u_{j-3/2}^{n}-u_{j-1/2}^{n}\geq 0, Step 2 ensures that

Djn+1ψ(max{Dj1n,Djn,Dj+1n}),Djn=max{uj1/2nuj+1/2n,0}.{\overset{\sim}{D}}_{j}^{n+1}\leq\psi\left(\max\left\{{\overset{\sim}{D}}_{j-1}^{n},{\overset{\sim}{D}}_{j}^{n},{\overset{\sim}{D}}_{j+1}^{n}\right\}\right),\quad{\overset{\sim}{D}}_{j}^{n}=\max\left\{u_{j-1/2}^{n}-u_{j+1/2}^{n},0\right\}.

Clearly,

Dj1nDj1n,Djn=Djn,Dj+1nDj+1n.{\overset{\sim}{D}}_{j-1}^{n}\leq D_{j-1}^{n},\quad{\overset{\sim}{D}}_{j}^{n}=D_{j}^{n},\quad{\overset{\sim}{D}}_{j+1}^{n}\leq D_{j+1}^{n}.

Using the monotonicity of ψ\psi, we get:

ρj1/2n+1ρj+1/2n+1uj1/2n+1uj+1/2n+1ψ(max{Dj1n,Djn,Dj+1n}),\rho_{j-1/2}^{n+1}-\rho_{j+1/2}^{n+1}\leq u_{j-1/2}^{n+1}-u_{j+1/2}^{n+1}\leq\psi\left(\max\left\{D_{j-1}^{n},D_{j}^{n},D_{j+1}^{n}\right\}\right),

concluding the proof.   \square

References

  • [1] B. Andreianov. New approaches to describing admissibility of solutions of scalar conservation laws with discontinuous flux. ESAIM: Proceedings and Surveys, 50:40–65, 2015.
  • [2] B. Andreianov, P. Goatin, and N. Seguin. Finite volume schemes for locally constrained conservation laws. Numerische Mathematik, 115(4):609–645, 2010.
  • [3] B. Andreianov, K. Karlsen, and N. H. Risebro. A theory of L1\text{L}^{1}-dissipative solvers for scalar conservation laws with discontinuous flux. Archive for Rational Mechanics and Analysis, 201(1):27–86, 2011.
  • [4] B. Andreianov and D. Mitrović. Entropy conditions for scalar conservation laws with discontinuous flux revisited. Ann. Inst. H. Poincaré Anal. Non Linéaire, 32(6):1307–1335, 2015.
  • [5] B. Andreianov and A. Sylla. A macroscopic model to reproduce self-organization at bottlenecks. In International Conference on Finite Volumes for Complex Applications, pages 243–254. Springer, 2020.
  • [6] A. Bressan, G. Guerra, and W. Shen. Vanishing viscosity solutions for conservation laws with regulated flux. Journal of Differential Equations, 266(1):312–351, 2019.
  • [7] C. Chalons, M. L. Delle Monache, and P. Goatin. A conservative scheme for non-classical solutions to a strongly coupled pde-ode problem. Interfaces and Free Boundaries, 19(4):553–570, 2018.
  • [8] R. M. Colombo and P. Goatin. A well posed conservation law with a variable unilateral constraint. Journal of Differential Equations, 234(2):654–675, 2007.
  • [9] M. L. Delle Monache and P. Goatin. Scalar conservation laws with moving constraints arising in traffic flow modeling: an existence result. Journal of Differential Equations, 257(11):4015––4029, 2014.
  • [10] M. L. Delle Monache and P. Goatin. A numerical scheme for moving bottlenecks in traffic flow. Bulletin of the Brazilian Mathematical Society, New Series, 47(2):605–617, 2016.
  • [11] M. L. Delle Monache and P. Goatin. Stability estimates for scalar conservation laws with moving flux constraints. Networks and Heterogeneous Media, 12(2):245–258, 2017.
  • [12] M. L. Delle Monache, T. Liard, B. Piccoli, R. Stern, and D. Work. Traffic reconstruction using autonomous vehicles. SIAM Journal on Applied Mathematics, 79(5):1748–1767, 2019.
  • [13] R. Eymard, T. Gallouët, and R. Herbin. Finite Volume Methods, volume VII of Handbook of Numerical Analysis. North-Holland, Amsterdam, 2000.
  • [14] A. Ferrara, P. Goatin, and G. Piacentini. A macroscopic model for platooning in highway traffic. SIAM Journal on Applied Mathematics, 80(1):639–656, 2020.
  • [15] M. Garavello, P. Goatin, T. Liard, and B. Piccoli. A multiscale model for traffic regulation via autonomous vehicles. Journal of Differential Equations, 269(7):6088–6124, 2020.
  • [16] I. Gasser, C. Lattanzio, and A. Maurizi. Vehicular traffic flow dynamics on a bus route. Multiscale Modeling & Simulation, 11(3):925–942, 2013.
  • [17] H. Holden and N. H. Risebro. Front Tracking for Hyperbolic Conservation Laws, volume 152 of Applied Mathematical Sciences. Springer-Verlag New York, 2002.
  • [18] K. H. Karlsen and J. D. Towers. Convergence of the lax-friedrichs scheme and stability for conservation laws with a discontinuous space-time dependent flux. Chinese Annals of Mathematics, 25(03):287–318, 2004.
  • [19] S. N. Kruzhkov. First order quasilinear equations with several independent variables. Mathematics of the USSR-Sbornik, 81(123):228–255, 1970.
  • [20] N. Laurent-Brouty, G. Costeseque, and P. Goatin. A macroscopic traffic flow model accounting for bounded acceleration. SIAM Journal on Applied Mathematics, 81(1):173–189, 2021.
  • [21] A. Sylla. Influence of a slow moving vehicle on traffic: Well-posedness and approximation for a mildly nonlocal model. Networks & Heterogeneous Media, 16(2):221–256, 2021.
  • [22] J. D. Towers. Convergence via OSLC of the Godunov scheme for a scalar conservation law with time and space flux discontinuities. Numerische Mathematik, 139(4):939–969, 2018.
  • [23] J. D. Towers. An existence result for conservation laws having BV spatial flux heterogeneities-without concavity. Journal of Differential Equations, 269(7):5754–5764, 2020.
  • [24] A. I. Vol’pert. The spaces BV and quasilinear equations. Matematicheskii Sbornik, 115(2):255–302, 1967.