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

On the L2L^{2} stability of shock waves for finite-entropy solutions of Burgers

Andres A. Contreras Hip  and  Xavier Lamy Department of Mathematical Sciences, New Mexico State University, Las Cruces, New Mexico, USA albertch@nmsu.edu Institut de Mathématiques de Toulouse; UMR 5219, Université de Toulouse; CNRS, UPS IMT, F-31062 Toulouse Cedex 9, France xlamy@math.univ-toulouse.fr
(Date: August 9, 2025)
Abstract.

We prove L2L^{2} stability estimates for entropic shocks among weak, possibly non-entropic, solutions of scalar conservation laws tu+xf(u)=0\partial_{t}u+\partial_{x}f(u)=0 with strictly convex flux function ff. This generalizes previous results by Leger and Vasseur, who proved L2L^{2} stability among entropy solutions. Our main result, the estimate

|u(t,)u0shock(x(t))|2dx|u0u0shock|2+Cμ+([0,t]×),\displaystyle\int_{\mathbb{R}}|u(t,\cdot)-u_{0}^{shock}(\cdot-x(t))|^{2}\,dx\leq\int_{\mathbb{R}}|u_{0}-u_{0}^{shock}|^{2}+C\mu_{+}([0,t]\times\mathbb{R}),

for some Lipschitz shift x(t)x(t), includes an error term accounting for the positive part of the entropy production measure μ=t(u2/2)+xq(u)\mu=\partial_{t}(u^{2}/2)+\partial_{x}q(u), where q(u)=uf(u)q^{\prime}(u)=uf^{\prime}(u). Stability estimates in this general non-entropic setting are of interest in connection with large deviation principles for the hydrodynamic limit of asymmetric interacting particle systems. Our proof adapts the scheme devised by Leger and Vasseur, where one constructs a shift x(t)x(t) which allows to bound from above the time-derivative of the left-hand side. The main difference lies in the fact that our solution u(t,)u(t,\cdot) may present a non-entropic shock at x=x(t)x=x(t) and new bounds are needed in that situation. We also generalize this stability estimate to initial data with bounded variation.

1. Introduction

We consider bounded weak (not necessarily entropy) solutions of Burgers’ equation

tu+xu22=0,\displaystyle\partial_{t}u+\partial_{x}\frac{u^{2}}{2}=0,

or more generally a scalar conservation law

(1.1) tu+xf(u)=0,t0,x,\displaystyle\partial_{t}u+\partial_{x}f(u)=0,\quad t\geq 0,\;x\in\mathbb{R},

with uniformly convex flux f′′α>0f^{\prime\prime}\geq\alpha>0. Let us recall that for any entropy-flux pair (η,q)(\eta,q) i.e. η′′0\eta^{\prime\prime}\geq 0 and q=ηfq^{\prime}=\eta^{\prime}f^{\prime}, the corresponding entropy production of a bounded weak solution uu is the distribution

(1.2) μη=tη(u)+xq(u).\displaystyle\mu_{\eta}=\partial_{t}\eta(u)+\partial_{x}q(u).

In the special case η(t)=t2/2\eta(t)=t^{2}/2 we will drop the subscript η\eta and simply write

(1.3) μ=tu22+xq(u),q(v)=0vtf(t)𝑑t.\displaystyle\mu=\partial_{t}\frac{u^{2}}{2}+\partial_{x}q(u),\qquad q(v)=\int_{0}^{v}tf^{\prime}(t)\,dt.

For smooth solutions the entropy production μη\mu_{\eta} is always zero, but smooth long-time solutions do not exist in general. Entropy solutions are weak solutions whose entropy production is nonpositive, i.e. μη0\mu_{\eta}\leq 0 for all convex entropies η\eta. Kružkov introduced this concept in [15] and showed that for any bounded initial condition u0(x)u_{0}(x) there exists a unique entropy solution.

Finite-entropy solutions

Here in contrast we consider weak solutions whose entropy productions do not necessarily have a sign. Such solutions are not uniquely determined by their initial conditions: they will in general deviate from the unique entropy solution, and the present work addresses the question of estimating this deviation. Our motivation comes from the study of large deviation principles for the hydrodynamic limit of asymmetric interacting particle systems [12, 26], where it is crucial to control how much a general weak solution can deviate from the entropy solution.

To give more details about this issue, we focus on a continuum variant introduced in [22, 3]. There, the question boils down to describing the variational convergence (Γ\Gamma-convergence) of functionals of the form

Eε(u)=1ε[εxux1(tu+xf(u))]2𝑑x𝑑t,\displaystyle E_{\varepsilon}(u)=\frac{1}{\varepsilon}\int\left[\varepsilon\partial_{x}u-\partial_{x}^{-1}(\partial_{t}u+\partial_{x}f(u))\right]^{2}\,dx\,dt,

in the regime ε0+\varepsilon\to 0^{+}. The same problem is considered in [23] (with the motivation of providing a variational point of view on the vanishing viscosity method). Limits u=limuεu=\lim u_{\varepsilon} of sequences of bounded energy Eε(uε)CE_{\varepsilon}(u_{\varepsilon})\leq C are weak solutions of (1.1), but not necessarily entropy solutions. They belong to the wider class that we will call here finite-entropy solutions: bounded weak solutions of (1.1) such that

(1.4) μη is a Radon measure for all convex η,\displaystyle\mu_{\eta}\text{ is a Radon measure for all convex }\eta,

where μη\mu_{\eta} is the entropy production defined in (1.2). The conjectured limiting energy E0(u)E_{0}(u) is the negative part μ([0,T]×)\mu_{-}([0,T]\times\mathbb{R}) of their entropy production, but a proof of this fact is still lacking.

Specifically, the missing part is the upper bound: given a finite-entropy solution uu, can one construct an approximating sequence uεuu_{\varepsilon}\to u in L1L^{1} such that lim supEε(uε)E0(u)\limsup E_{\varepsilon}(u_{\varepsilon})\leq E_{0}(u) ? Very similar questions arise in relation with micromagnetics models (the so-called Aviles-Giga energy), we refer to the introduction of [16] for more details. What makes this question hard is the lack of fine knowledge on finite-entropy solutions. Unlike entropy solutions, they are not necessarily of bounded variation (BV). Only very recently E. Marconi [20, 21] proved that their entropy production is a one-dimensional rectifiable measure. This rectifiability result is a remarkable achievement, but it seems that solving the upper bound problem requires other new ideas.

To the best of our knowledge, only two upper bound constructions are available in the literature, with restrictive assumptions on the finite-entropy solution uu. The first construction in [23] requires uu to be BV, and the approximating sequence is obtained by mollifying uu and using the fine properties of BV functions. The second construction in [3] is based on approximation by vanishing viscosity, which converges in open regions where the entropy production is 0\leq 0. If regions of negative and positive entropy production are not “well separated” this construction breaks down, for want of a good estimate on the distance between uu and entropy solutions when the entropy production changes sign.

In this spirit, the only estimate [16] we are aware of is not homogeneous:

[0,1]t×[1,1]x|uuent|4Cμ+([0,2]t×[2,2]x)γfor some γ(0,1),\displaystyle\int_{[0,1]_{t}\times[-1,1]_{x}}\left|u-u^{ent}\right|^{4}\leq C\,\mu_{+}([0,2]_{t}\times[-2,2]_{x})^{\gamma}\qquad\text{for some }\gamma\in(0,1),

where uentu^{ent} is the entropy solution with initial data u0u_{0} and |u|1|u|\leq 1. If one applies (a rescaled version of) this estimate in small regions where μ+\mu_{+} is small, after summing over all regions the right-hand side may become very large because of the small exponent γ\gamma. As a consequence, this estimate cannot be used to remove the main restriction (that the regions where the entropy production has a constant sign must be well separated) in the approximation scheme of [3]. One would rather need an estimate that is homogeneous, hence amenable to summing rescaled applications of it.

In this work we propose a new approach towards such estimate, beginning with the distance of uu to entropic shocks: if a solution uu starts close to a shock and μ+\mu_{+} is small, then uu remains close to a shock, and this is quantified via a homogeneous estimate. More precisely, our main result takes the form of an L2L^{2} stability estimate for entropic shocks. For entropy solutions this question was adressed in [18, 19] using relative entropy methods. Here we generalize their methods to solutions whose entropy production does not necessarily have a sign. Loosely stated, we prove (Theorem 1.1)

|u(t)shock|2𝑑x|u(0)shock|2𝑑x+C0tμ+(dt,dx),\displaystyle\int|u(t)-\text{shock}|^{2}\,dx\leq\int|u(0)-\text{shock}|^{2}dx+C\,\int_{0}^{t}\!\int\mu_{+}(dt,dx),

where the shock at time tt is a shift of the initial shock. We also provide a generalization to any BVBV initial data (Theorem 1.4).

Strong and very strong traces

As in [19, 13, 14], in order to implement the relative entropy method we need to assume that uu has traces on Lipschitz curves, in a strong enough sense. From [27], it is known that finite-entropy solutions have traces which are reached strongly in L1L^{1}. We call this the strong trace property, precisely defined as follows. A bounded function u:[0,T]×u\colon[0,T]\times\mathbb{R}\to\mathbb{R} satisfies the strong trace property if for any Lipschitz path x:[0,T]x\colon[0,T]\to\mathbb{R} there exist traces tu(t,x(t)±)t\mapsto u(t,x(t)\pm) on each side of x(t)x(t), such that

(1.5) esslimy0+0T|u(t,x(t)±y)u(t,x(t)±)|𝑑t=0.\displaystyle\operatorname*{ess\,lim}_{y\to 0^{+}}\int_{0}^{T}\left|u(t,x(t)\pm y)-u(t,x(t)\pm)\right|\,dt=0.

In [27] entropy solutions are considered, but the proof there uses only a kinetic formulation which is also valid for finite-entropy solutions [8]. The results of [27] also include traces along constant time lines, implying that (for an a.e. representative)

(1.6) [0,T]tu(t,)Lloc1is continuous,\displaystyle[0,T]\ni t\mapsto u(t,\cdot)\in L^{1}_{loc}\quad\text{is continuous,}

whenever uu is a finite-entropy solution of (1.1).

Unfortunately the strong trace property turns out not to be enough for our purposes, and as in [19, 13, 14] we will in fact require an even stronger property. We say that a bounded function u:[0,T]×u\colon[0,T]\times\mathbb{R}\to\mathbb{R} satisfies the very strong trace property if for any Lipschitz path x:[0,T]x\colon[0,T]\to\mathbb{R} there exist traces tu(t,x(t)±)t\mapsto u(t,x(t)\pm) such that (for an a.e. representative of uu)

(1.7) esslimy0+u(t,x(t)±y)=u(t,x(t)±)for a.e. t[0,T].\displaystyle\operatorname*{ess\,lim}_{y\to 0^{+}}u(t,x(t)\pm y)=u(t,x(t)\pm)\qquad\text{for a.e. }t\in[0,T].

By dominated convergence the very strong trace property does imply the strong trace property. Functions uBV([0,T]×)u\in BV([0,T]\times\mathbb{R}) satisfy the very strong trace property, but it is not known whether finite-entropy solutions satisfy it.

Stability of shocks in L2L^{2} for finite-entropy solutions

We are now ready to state our main result, on the L2L^{2} stability of an entropic shock wave ushocku^{shock} with initial datum

(1.8) u0shock(x)\displaystyle u_{0}^{shock}(x) =u𝟏x<0+ur𝟏x>0,u>ur,\displaystyle=u_{\ell}\mathbf{1}_{x<0}+u_{r}\mathbf{1}_{x>0},\qquad u_{\ell}>u_{r},

that is, ushock(t,x)=u0shock(xσt)u^{shock}(t,x)=u_{0}^{shock}(x-\sigma t), with shock speed σ=(f(ur)f(u))/(uru)\sigma=(f(u_{r})-f(u_{\ell}))/(u_{r}-u_{\ell}).

Theorem 1.1.

Let f:f\colon\mathbb{R}\to\mathbb{R} be such that f′′α>0f^{\prime\prime}\geq\alpha>0. Let u:[0,T]×u\colon[0,T]\times\mathbb{R}\to\mathbb{R} be a bounded finite-entropy solution (1.4) of

tu+xf(u)=0,u(0,x)=u0(x).\displaystyle\partial_{t}u+\partial_{x}f(u)=0,\qquad u(0,x)=u_{0}(x).

Assume that uu satisfies the very strong trace property (1.7). Let ushocku^{shock} be the entropic shock wave with initial datum u0shocku_{0}^{shock} (1.8), and set M=supIf′′M=\sup_{I}f^{\prime\prime} and S=supI|f|S=\sup_{I}|f^{\prime}|, where I=[min(ur,infu),max(u,supu)]I=[\min(u_{r},\inf u),\max(u_{\ell},\sup u)].

There exists a Lipschitz path h:[0,T]h\colon[0,T]\to\mathbb{R} such that h(0)=0h(0)=0 and

(1.9) RR|u(t,x)ushock(t,xh(t))|2𝑑x\displaystyle\int_{-R}^{R}\left|u(t,x)-u^{shock}(t,x-h(t))\right|^{2}dx RtSR+tS|u0u0shock|2𝑑x\displaystyle\leq\int_{-R-tS}^{R+tS}\left|u_{0}-u_{0}^{shock}\right|^{2}dx
+CM3α3μ+([0,t]×[RtS,R+tS]),\displaystyle\quad+C\frac{M^{3}}{\alpha^{3}}\,\mu_{+}([0,t]\times[-R-tS,R+tS]),

for all t[0,T]t\in[0,T], all R>0R>0 and some absolute constant C>0C>0, where μ\mu is the entropy production (1.3) associated with η(t)=t2/2\eta(t)=t^{2}/2.

In addition the drift hh is controlled by

(1.10) cαM2(uur)0th(τ)2𝑑τ\displaystyle c\frac{\alpha}{M^{2}}(u_{\ell}-u_{r})\int_{0}^{t}h^{\prime}(\tau)^{2}\,d\tau 2St2St(u0u0shock)2𝑑x\displaystyle\leq\int_{-2St}^{2St}(u_{0}-u_{0}^{shock})^{2}\,dx
+M3α3μ+([0,t]×[2St,2St])\displaystyle\quad+\frac{M^{3}}{\alpha^{3}}\mu_{+}([0,t]\times[-2St,2St])

for some absolute constant c>0c>0 and all t[0,T]t\in[0,T].

Remark 1.2.

We were not able to remove the very strong trace assumption from this statement. In the proof it is used only in Lemma 3.3 to establish that uu admits generalized characteristics: for any x0x_{0}\in\mathbb{R}, there exists a Lipschitz curve x:[0,T]x\colon[0,T]\mapsto\mathbb{R} such that x(0)=x0x(0)=x_{0} and

x(t)=σ(u(t,x(t)),u(t,x(t)+)for a.e. t[0,T],\displaystyle x^{\prime}(t)=\sigma(u(t,x(t)-),u(t,x(t)+)\qquad\text{for a.e. }t\in[0,T],

where σ(u,u+)=(f(u+)f(u))/(u+u)\sigma(u_{-},u_{+})=(f(u_{+})-f(u_{-}))/(u_{+}-u_{-}) when uu+u_{-}\neq u_{+}, and σ(u,u)=f(u)\sigma(u,u)=f^{\prime}(u). Other places where traces are needed require only the strong trace property (1.5), satisfied by finite-entropy solutions.

Remark 1.3.

The necessity of introducing a drift h(t)h(t), and the near-optimality of estimate (1.10) when μ+=0\mu_{+}=0 and u=ur=1u_{\ell}=-u_{r}=1, are proved in [29, Proposition 1.2].

To prove Theorem 1.1 we adapt the relative entropy arguments used in [18, 19, 14] (see also [24, 13, 25, 29]). The relative entropy method was introduced in [6, 10] to study the L2L^{2} stability of smooth solutions among entropy solutions, and later refined in [18, 19] to obtain the L2L^{2} stability (up to a drift) of shock waves (see [1] for LpL^{p} stability estimates up to a drift). This method is also relevant in the study of hydrodynamic limits for fluid equations [28]. The basic idea is that for any constant v0v_{0}, one has an identity of the form

12t(uv0)2=μxq(u;v0).\displaystyle\frac{1}{2}\partial_{t}(u-v_{0})^{2}=\mu-\partial_{x}q(u;v_{0}).

Stability of the constant state v0v_{0} when μ0\mu\leq 0 then follows by integrating over xx\in\mathbb{R}, provided q(u;v0)q(u;v_{0}) is nice enough (e.g. has compact support). In the case of finite-entropy solutions, one also has to take into account the contribution of μ+\mu_{+}. But when studying the stability of a shock, one integrates t(uu)2\partial_{t}(u-u_{\ell})^{2} and t(uur)2\partial_{t}(u-u_{r})^{2} on two complementary half-lines, and boundary terms appear at the junction.

The crucial remark used in [18, 19, 14] is that, if the initial shock is shifted by a well-chosen length x(t)x(t), then the boundary terms combine into a nonpositive contribution. There are two cases to consider, depending on whether or not u(t,)u(t,\cdot) jumps at x(t)x(t). At times tt where it does not jump, the situation is the same for entropy or finite-entropy solutions, and the ideas of [18, 19, 14] apply also in our case. But at times tt where it does jump, an entropy solution can only make a negative jump, while a finite-entropy solution can also make a positive jump. More precisely, denoting by (u,u+)(u_{-},u_{+}) the values of the jump of uu, it is shown in [18, 19] that the dissipation rate D(u,u+;u,ur)D(u_{-},u_{+};u_{\ell},u_{r}) coming from the boundary terms satisfies

D(u,u+;u,ur)0whenever uu+ and uur.\displaystyle D(u_{-},u_{+};u_{\ell},u_{r})\leq 0\qquad\text{whenever }u_{-}\geq u_{+}\text{ and }u_{\ell}\geq u_{r}.

To include finite-entropy solutions, we have to consider also what happens when u<u+u_{-}<u_{+}. One cannot expect the dissipation rate DD to remain 0\leq 0, but what we do show (see Proposition 2.1) is that its positive part is controlled by the entropy cost of the jump, in other words by μ+\mu_{+}. This crucial observation enables us to adapt the techniques of [18, 19, 14] to our situation and to prove the stability estimate (1.9). In fact we prove a sharper upper bound on DD, thanks to which the control (1.10) on the drift h(t)h(t) can then be obtained as in [14].

Stability of entropy solutions with BVBV initial data

As a complement we provide a generalization of Theorem 1.1 where the entropy solution ushocku^{shock} is replaced by any entropy solution with BVBV initial data. This relies, as in [13, 4], on applying the techniques of Theorem 1.1’s proof to obtain estimates between uu and functions with a finite number of shocks. Each shock wave has to be shifted, and all shifts may be different. The estimate (1.10) on the drift of one single shock in Theorem 1.1 is therefore replaced by an estimate on the L1L^{1} distance between the “shifted” function and the actual entropy solution.

Theorem 1.4.

Let f:f\colon\mathbb{R}\to\mathbb{R} be such that f′′α>0f^{\prime\prime}\geq\alpha>0. Let u:[0,T]×u\colon[0,T]\times\mathbb{R}\to\mathbb{R} be a bounded finite-entropy solution (1.4) of

tu+xf(u)=0,u(0,x)=u0(x).\displaystyle\partial_{t}u+\partial_{x}f(u)=0,\qquad u(0,x)=u_{0}(x).

Assume that uu satisfies the very strong trace property (1.7). Let ζ\zeta be an entropy solution of (1.1) with initial datum ζ0LBVloc()\zeta_{0}\in L^{\infty}\cap BV_{loc}(\mathbb{R}), and set M=supIf′′M=\sup_{I}f^{\prime\prime} and S=supI|f|S=\sup_{I}|f^{\prime}|, where I=[min(infζ0,infu),max(supζ0,supu)]I=[\min(\inf\zeta_{0},\inf u),\max(\sup\zeta_{0},\sup u)].

There exists u~LBVloc([0,T]×)Lip([0,T],Lloc1())\tilde{u}\in L^{\infty}\cap BV_{loc}([0,T]\times\mathbb{R})\cap\mathrm{Lip}([0,T],L^{1}_{loc}(\mathbb{R})) such that

(1.11) RR|u(t,x)u~(t,x)|2𝑑x\displaystyle\int_{-R}^{R}\left|u(t,x)-\tilde{u}(t,x)\right|^{2}dx RStR+St|u0ζ0|2𝑑x\displaystyle\leq\int_{-R-St}^{R+St}\left|u_{0}-\zeta_{0}\right|^{2}dx
+CM3α3μ+([0,t]×[RSt,R+St]),\displaystyle\quad+C\frac{M^{3}}{\alpha^{3}}\mu_{+}([0,t]\times[-R-St,R+St]),

and

(1.12) RR|u~(t,x)ζ(t,x)|𝑑x\displaystyle\int_{-R}^{R}\left|\tilde{u}(t,x)-\zeta(t,x)\right|\,dx
CMα12(Dζ0)([RSt,R+St])t\displaystyle\leq C\frac{M}{\alpha^{\frac{1}{2}}}\sqrt{(D\zeta_{0})_{-}([-R-St,R+St])}\sqrt{t}
RStR+St|u0ζ0|2𝑑x+CM3α3μ+([0,t]×[RSt,R+St]),\displaystyle\hskip 30.00005pt\cdot\sqrt{\int_{-R-St}^{R+St}\left|u_{0}-\zeta_{0}\right|^{2}dx+C\frac{M^{3}}{\alpha^{3}}\mu_{+}([0,t]\times[-R-St,R+St])},

for some absolute constant C>0C>0, all t[0,T]t\in[0,T] and all R>0R>0.

Remark 1.5.

In the case ζ0=u0shock\zeta_{0}=u_{0}^{shock}, Theorem 1.4 is a corollary of Theorem 1.1 taking u~(t,x)=u0shock(t,xh(t))\tilde{u}(t,x)=u_{0}^{shock}(t,x-h(t)) so that (1.11) is exactly (1.9), and (1.12) follows from (1.10).

Remark 1.6.

It is well-known that weak differentiability of order s=1/3s=1/3 is critical for finite entropy solutions of (1.1) (see [9, 11, 7]). Somewhat interestingly this critical exponent also comes up in relation with Theorem 1.4: if one wishes to use Theorem 1.4 in order to estimate the distance of uu to the entropy solution starting at u0u_{0} (when u0u_{0} is not BVBV) in terms of μ+\mu_{+}, it seems natural to consider ζ0=u0ρε\zeta_{0}=u_{0}*\rho_{\varepsilon} with ρε(x)=ε1ρ(x/ε)\rho_{\varepsilon}(x)=\varepsilon^{-1}\rho(x/\varepsilon) for some smooth kernel ρ\rho and ε\varepsilon small enough so that |u0u0ρε|2𝑑xμ+(dt,dx)\int|u_{0}-u_{0}*\rho_{\varepsilon}|^{2}\,dx\lesssim\int\mu_{+}(dt,dx). The right-hand side of (1.12) then puts forward the square root of the product

|(u0ρε)|𝑑x|u0u0ρε|2𝑑x.\displaystyle\int|(u_{0}*\rho_{\varepsilon})^{\prime}|\,dx\cdot\int|u_{0}-u_{0}*\rho_{\varepsilon}|^{2}\,dx.

If u0u_{0} enjoys some fractional derivability of order s>0s>0 (e.g. of Besov B2,sB^{s}_{2,\infty} or Sobolev Ws,2W^{s,2} type), the first factor is typically bounded by εs1\varepsilon^{s-1}, the second by ε2s\varepsilon^{2s}, hence this product is bounded by ε3s1\varepsilon^{3s-1}, and the exponent s=1/3s=1/3 is critical.

Outline

The article is organized as follows. In section 2 we prove the new bound on the dissipation rate DD appearing in the relative entropy method. In section 3 we recall and adapt the arguments of [18, 19] to prove Theorem 1.1. In section 4 we prove Theorem 1.4.

Acknowledgements

X.L. is partially supported by ANR project ANR-18-CE40-0023 and COOPINTER project IEA-297303.

2. Upper bound on the dissipation rate DD

We start by setting some notations. We denote by η\eta, qq the entropy-flux pair given by

η(x)=x22,q(x)=0xηf,\displaystyle\eta(x)=\frac{x^{2}}{2},\quad q(x)=\int_{0}^{x}\eta^{\prime}f^{\prime},

and by η(|),q(;)\eta(\cdot|\cdot),q(\cdot;\cdot) the corresponding relative entropy-flux pair

η(x|a)\displaystyle\eta(x|a) =η(x)η(a)η(a)(xa)=(xa)22\displaystyle=\eta(x)-\eta(a)-\eta^{\prime}(a)(x-a)=\frac{{(x-a)}^{2}}{2}
q(x;a)\displaystyle q(x;a) =q(x)q(a)η(a)(f(x)f(a)).\displaystyle=q(x)-q(a)-\eta^{\prime}(a)(f(x)-f(a)).

The propagation speed of a shock (u,u+)(u_{-},u_{+}) is constrained by the Rankine-Hugoniot condition:

(2.1) σ(u,u+)=f(u+)f(u)u+u,\displaystyle\sigma(u_{-},u_{+})=\frac{f(u_{+})-f(u_{-})}{u_{+}-u_{-}},

and by setting σ(u,u)=f(u)\sigma(u,u)=f^{\prime}(u) the function σ\sigma is continuous on 2\mathbb{R}^{2}. Given two shocks (u,u+)(u_{-},u_{+}) and (u,ur)(u_{\ell},u_{r}) we define the dissipation rate

(2.2) D(u,u+;u,ur):=q(u+;ur)q(u;u)σ(u,u+)(η(u+|ur)η(u|u)).\displaystyle D(u_{-},u_{+};u_{\ell},u_{r}):=q(u_{+};u_{r})-q(u_{-};u_{\ell})-\sigma(u_{-},u_{+})\left(\eta(u_{+}|u_{r})-\eta(u_{-}|u_{\ell})\right).

As explained in the introduction, this corresponds to the boundary terms which arise when calculating

tη(u(t,x)|u0shock(xx(t)))𝑑x,\displaystyle\partial_{t}\int\eta\big{(}u(t,x)\big{|}u_{0}^{shock}(x-x(t))\big{)}\,dx,

at times tt where u(t,)u(t,\cdot) has a jump (u,u+)(u_{-},u_{+}) at x=x(t)x=x(t).

Our goal in this section is to compare the dissipation rate DD with the entropy cost of the jump (u,u+)(u_{-},u_{+}), given by

(2.3) E(u,u+)\displaystyle E(u_{-},u_{+}) =q(u+)q(u)σ(u,u+)(η(u+)η(u)).\displaystyle=q(u_{+})-q(u_{-})-\sigma(u_{-},u_{+})(\eta(u_{+})-\eta(u_{-})).

This formula corresponds to the fact that, if a solution uu has a jump (u(t),u+(t))(u_{-}(t),u_{+}(t)) along a curve x(t)x(t), and is smooth everywhere else, then by the BV chain rule (see e.g. [2, § 3.10]) the entropy production μ\mu is given by

μ(A)=𝟏(t,x(t))AE(u(t),u+(t))𝑑t.\displaystyle\mu(A)=\int\mathbf{1}_{(t,x(t))\in A}\,E(u_{-}(t),u_{+}(t))\,dt.

The main result of this section is the following.

Proposition 2.1.

For uuru_{\ell}\geq u_{r} and any u±u_{\pm}\in\mathbb{R} we have

D(u,u+;u,ur)\displaystyle D(u_{-},u_{+};u_{\ell},u_{r}) C1M3α3max(E(u,u+),0)\displaystyle\leq C_{1}\frac{M^{3}}{\alpha^{3}}\max(E(u_{-},u_{+}),0)
C2α(uur)[(uu)2+(uru+)2],\displaystyle\quad-C_{2}\,\alpha\,(u_{\ell}-u_{r})\left[(u_{\ell}-u_{-})^{2}+(u_{r}-u_{+})^{2}\right],

for some absolute constants C1,C2>0C_{1},C_{2}>0 and 0<αM0<\alpha\leq M such that αf′′M\alpha\leq f^{\prime\prime}\leq M on the convex hull of {u,u+,u,ur}\{u_{-},u_{+},u_{\ell},u_{r}\}.

Proof of Proposition 2.1.

The case uu+u_{-}\geq u_{+} can be inferred from the arguments in [14, Lemma 4.1], only the case u<u+u_{-}<u_{+} is really new. For the reader’s convenience we include a proof in both cases. Only the values of ff on the convex hull of {u,u+,u,ur}\{u_{-},u_{+},u_{\ell},u_{r}\} play a role in this inequality, so we assume without loss of generality that αf′′M\alpha\leq f^{\prime\prime}\leq M on \mathbb{R}.

Case 1: u+uu_{+}\geq u_{-}.

We split DD as

(2.4) D(u,u+;u,ur)\displaystyle D(u_{-},u_{+};u_{\ell},u_{r}) =E(u,u+)+F(u,u+;u,ur)+D(u,u;u,ur)\displaystyle=E(u_{-},u_{+})+F_{-}(u_{-},u_{+};u_{\ell},u_{r})+D(u_{-},u_{-};u_{\ell},u_{r})
=E(u,u+)+F+(u,u+;u,ur)+D(u+,u+;u,ur)\displaystyle=E(u_{-},u_{+})+F_{+}(u_{-},u_{+};u_{\ell},u_{r})+D(u_{+},u_{+};u_{\ell},u_{r})

where

F(u,u+;u,ur)\displaystyle F_{-}(u_{-},u_{+};u_{\ell},u_{r}) =D(u,u+;u,ur)E(u,u+)D(u,u;u,ur)\displaystyle=D(u_{-},u_{+};u_{\ell},u_{r})-E(u_{-},u_{+})-D(u_{-},u_{-};u_{\ell},u_{r})
=η(ur)[f(u)f(u+)+σ(u,u+)u+f(u)u]\displaystyle=\eta^{\prime}(u_{r})\left[f(u_{-})-f(u_{+})+\sigma(u_{-},u_{+})u_{+}-f^{\prime}(u_{-})u_{-}\right]
+(σ(u,u+)f(u))[η(ur)η(u)urη(ur)\displaystyle\quad+(\sigma(u_{-},u_{+})-f^{\prime}(u_{-}))\big{[}\eta(u_{r})-\eta(u_{\ell})-u_{r}\eta^{\prime}(u_{r})
+uη(u)η(u)u],\displaystyle\hskip 140.00021pt+u_{\ell}\eta^{\prime}(u_{\ell})-\eta^{\prime}(u_{\ell})u_{-}\big{]},
F+(u,u+;u,ur)\displaystyle F_{+}(u_{-},u_{+};u_{\ell},u_{r}) =D(u,u+;u,ur)E(u,u+)D(u+,u+;u,ur)\displaystyle=D(u_{-},u_{+};u_{\ell},u_{r})-E(u_{-},u_{+})-D(u_{+},u_{+};u_{\ell},u_{r})
=η(u)[f(u)f(u+)σ(u,u+)u+f(u+)u+]\displaystyle=\eta^{\prime}(u_{\ell})\left[f(u_{-})-f(u_{+})-\sigma(u_{-},u_{+})u_{-}+f^{\prime}(u_{+})u_{+}\right]
+(σ(u,u+)f(u+))[η(ur)η(u)urη(ur)\displaystyle\quad+(\sigma(u_{-},u_{+})-f^{\prime}(u_{+}))\big{[}\eta(u_{r})-\eta(u_{\ell})-u_{r}\eta^{\prime}(u_{r})
+uη(u)+η(ur)u+,].\displaystyle\hskip 140.00021pt+u_{\ell}\eta^{\prime}(u_{\ell})+\eta^{\prime}(u_{r})u_{+},\big{]}.

We also define

Δ:=u+u0,A±:=uu±B±:=uru±,\displaystyle\Delta:=u_{+}-u_{-}\geq 0,\quad A_{\pm}:=u_{\ell}-u_{\pm}\geq B_{\pm}:=u_{r}-u_{\pm},

and start by remarking that

(2.5) E(u,u+)\displaystyle E(u_{-},u_{+}) =β12Δ3for some β[α,M],\displaystyle=\frac{\beta}{12}\Delta^{3}\qquad\text{for some }\beta\in[\alpha,M],
(2.6) F±(u,u+;u,ur)\displaystyle F_{\pm}(u_{-},u_{+};u_{\ell},u_{r}) =γ±4Δ(A±2B±2)for some γ±[α,M],\displaystyle=\mp\frac{\gamma_{\pm}}{4}\Delta(A_{\pm}^{2}-B_{\pm}^{2})\qquad\text{for some }\gamma_{\pm}\in[\alpha,M],
(2.7) D(u±,u±;u,ur)\displaystyle D(u_{\pm},u_{\pm};u_{\ell},u_{r}) α6(A±3B±3).\displaystyle\leq-\frac{\alpha}{6}(A_{\pm}^{3}-B_{\pm}^{3}).

Proof of (2.5). Recalling that η(t)=t2/2\eta(t)=t^{2}/2 and q(t)=tf(t)q^{\prime}(t)=tf^{\prime}(t) we have

E(u,u+)\displaystyle E(u_{-},u_{+}) =q(u+)q(u)σ(u,u+)(η(u+)η(u))\displaystyle=q(u_{+})-q(u_{-})-\sigma(u_{-},u_{+})(\eta(u_{+})-\eta(u_{-}))
=uu+tf(t)𝑑tu++u2uu+f(t)𝑑t\displaystyle=\int_{u_{-}}^{u_{+}}tf^{\prime}(t)\,dt-\frac{u_{+}+u_{-}}{2}\int_{u_{-}}^{u_{+}}f^{\prime}(t)\,dt
=14(u+u)211sf(u++u2+su+u2)𝑑s\displaystyle=\frac{1}{4}(u_{+}-u_{-})^{2}\int_{-1}^{1}sf^{\prime}\left(\frac{u_{+}+u_{-}}{2}+s\frac{u_{+}-u_{-}}{2}\right)\,ds
=14(u+u)211s[f(u++u2+su+u2)f(u++u2)]𝑑s\displaystyle=\frac{1}{4}(u_{+}-u_{-})^{2}\int_{-1}^{1}s\left[f^{\prime}\left(\frac{u_{+}+u_{-}}{2}+s\frac{u_{+}-u_{-}}{2}\right)-f^{\prime}\left(\frac{u_{+}+u_{-}}{2}\right)\right]\,ds
=18(u+u)30111s2f′′(u++u2+stu+u2)𝑑s𝑑t.\displaystyle=\frac{1}{8}(u_{+}-u_{-})^{3}\int_{0}^{1}\int_{-1}^{1}s^{2}f^{\prime\prime}\left(\frac{u_{+}+u_{-}}{2}+st\frac{u_{+}-u_{-}}{2}\right)\,ds\,dt.

This last expression implies (2.5) since αf′′M\alpha\leq f^{\prime\prime}\leq M and 0111s2𝑑s𝑑t=2/3\int_{0}^{1}\int_{-1}^{1}s^{2}\,dsdt=2/3.∎

Proof of (2.6). We have

F\displaystyle F_{-} =12(uur)(u+ur2u)1u+uuu+(f(t)f(u))𝑑t\displaystyle=\frac{1}{2}(u_{\ell}-u_{r})(u_{\ell}+u_{r}-2u_{-})\frac{1}{u_{+}-u_{-}}\int_{u_{-}}^{u_{+}}(f^{\prime}(t)-f^{\prime}(u_{-}))\,dt
=12(AB)(A+B)01[f(u+s(u+u))f(u)]𝑑s\displaystyle=\frac{1}{2}(A_{-}-B_{-})(A_{-}+B_{-})\int_{0}^{1}\left[f^{\prime}(u_{-}+s(u_{+}-u_{-}))-f^{\prime}(u_{-})\right]\,ds
=12Δ(A2B2)0101tf′′(u+st(u+u))𝑑s𝑑t,\displaystyle=\frac{1}{2}\Delta(A_{-}^{2}-B_{-}^{2})\int_{0}^{1}\int_{0}^{1}t\,f^{\prime\prime}(u_{-}+st(u_{+}-u_{-}))\,ds\,dt,

which gives (2.6) for FF_{-} since αf′′M\alpha\leq f^{\prime\prime}\leq M and 0101t𝑑s𝑑t=1/2\int_{0}^{1}\int_{0}^{1}t\,dsdt=1/2. Similarly

F+\displaystyle F_{+} =12Δ(A+2B+2)0101tf′′(u+st(u+u))𝑑s𝑑t\displaystyle=-\frac{1}{2}\Delta(A_{+}^{2}-B_{+}^{2})\int_{0}^{1}\int_{0}^{1}t\,f^{\prime\prime}(u_{+}-st(u_{+}-u_{-}))\,ds\,dt

which gives (2.6) for F+F_{+}.∎

Proof of (2.7). We have

D(u,u;u,ur)\displaystyle D(u,u;u_{\ell},u_{r}) =urutf(t)𝑑tuuuf(t)𝑑tururuf(t)𝑑t\displaystyle=\int_{u_{r}}^{u_{\ell}}tf^{\prime}(t)\,dt-u_{\ell}\int_{u}^{u_{\ell}}f^{\prime}(t)\,dt-u_{r}\int_{u_{r}}^{u}f^{\prime}(t)\,dt
+12(uur)(u+ur2u)f(u)\displaystyle\quad+\frac{1}{2}(u_{\ell}-u_{r})(u_{\ell}+u_{r}-2u)f^{\prime}(u)
=uu(tu)f(t)𝑑t+uru(tur)f(t)𝑑t\displaystyle=\int_{u}^{u_{\ell}}(t-u_{\ell})f^{\prime}(t)\,dt+\int_{u_{r}}^{u}(t-u_{r})f^{\prime}(t)\,dt
+(uu(tu)+uru(tur))f(u)\displaystyle\quad+\left(\int_{u}^{u_{\ell}}(t-u_{\ell})+\int_{u_{r}}^{u}(t-u_{r})\right)f^{\prime}(u)
=uu(tu)(f(t)f(u))𝑑t+uru(tur)(f(t)f(u))𝑑t\displaystyle=\int_{u}^{u_{\ell}}(t-u_{\ell})(f^{\prime}(t)-f^{\prime}(u))\,dt+\int_{u_{r}}^{u}(t-u_{r})(f^{\prime}(t)-f^{\prime}(u))\,dt
=uu(tu)(tu)01f′′(u+s(tu))𝑑s𝑑t\displaystyle=\int_{u}^{u_{\ell}}(t-u_{\ell})(t-u)\int_{0}^{1}f^{\prime\prime}(u+s(t-u))\,ds\,dt
+uru(tur)(tu)01f′′(u+s(tu))𝑑s𝑑t.\displaystyle\quad+\int_{u_{r}}^{u}(t-u_{r})(t-u)\int_{0}^{1}f^{\prime\prime}(u+s(t-u))\,ds\,dt.

If u[ur,u]u\in[u_{r},u_{\ell}] we see that f′′αf^{\prime\prime}\geq\alpha implies

D(u,u;u,ur)\displaystyle D(u,u;u_{\ell},u_{r}) α(uu(ut)(tu)𝑑t+uru(tur)(ut)𝑑t)\displaystyle\leq-\alpha\left(\int_{u}^{u_{\ell}}(u_{\ell}-t)(t-u)\,dt+\int_{u_{r}}^{u}(t-u_{r})(u-t)\,dt\right)
=α6(A3B3),\displaystyle=-\frac{\alpha}{6}\left(A^{3}-B^{3}\right),

where A=uuA=u_{\ell}-u and B=uruB=u_{r}-u. If uuru\leq u_{r} we rewrite the above as

D(u,u;u,ur)\displaystyle D(u,u;u_{\ell},u_{r}) =uru(ut)(tu)01f′′(u+s(tu))𝑑s𝑑t\displaystyle=-\int_{u_{r}}^{u_{\ell}}(u_{\ell}-t)(t-u)\int_{0}^{1}f^{\prime\prime}(u+s(t-u))\,ds\,dt
uur(uur)(tu)01f′′(u+s(tu))𝑑s𝑑t,\displaystyle\quad-\int_{u}^{u_{r}}(u_{\ell}-u_{r})(t-u)\int_{0}^{1}f^{\prime\prime}(u+s(t-u))\,ds\,dt,

and if uuu\geq u_{\ell} as

D(u,u;u,ur)\displaystyle D(u,u;u_{\ell},u_{r}) =uu(uur)(ut)01f′′(u+s(tu))𝑑s𝑑t\displaystyle=-\int_{u_{\ell}}^{u}(u_{\ell}-u_{r})(u-t)\int_{0}^{1}f^{\prime\prime}(u+s(t-u))\,ds\,dt
uru(tur)(ut)01f′′(u+s(tu))𝑑s𝑑t,\displaystyle\quad-\int_{u_{r}}^{u_{\ell}}(t-u_{r})(u-t)\int_{0}^{1}f^{\prime\prime}(u+s(t-u))\,ds\,dt,

and in both cases we deduce again that (2.7) is valid.∎

Combining (2.4) with (2.5)-(2.7) we obtain

(2.8) D(u,u+;u,ur)β12Δ3γ±4Δ(A±2B±2)α6(A±3B±3).\displaystyle D(u_{-},u_{+};u_{\ell},u_{r})\leq\frac{\beta}{12}\Delta^{3}\mp\frac{\gamma_{\pm}}{4}\Delta(A_{\pm}^{2}-B_{\pm}^{2})-\frac{\alpha}{6}(A_{\pm}^{3}-B_{\pm}^{3}).

Since Δ0\Delta\geq 0, for any ABA\geq B and γ,λ>0\gamma,\lambda>0, by Young’s inequality ab13a3+23b32ab\leq\frac{1}{3}a^{3}+\frac{2}{3}b^{\frac{3}{2}} (a,b0)a,b\geq 0) we have

γ4Δ|A2B2|γ12λ3Δ3+γ6λ32|A2B2|32.\displaystyle\frac{\gamma}{4}\Delta\left|A^{2}-B^{2}\right|\leq\frac{\gamma}{12\lambda^{3}}\Delta^{3}+\frac{\gamma}{6}\lambda^{\frac{3}{2}}\left|A^{2}-B^{2}\right|^{\frac{3}{2}}.

From

|A2B2|3\displaystyle\left|A^{2}-B^{2}\right|^{3} =(AB)2|A2B2|(A+B)22(AB)2(A2+B2)2\displaystyle=(A-B)^{2}\left|A^{2}-B^{2}\right|(A+B)^{2}\leq 2(A-B)^{2}(A^{2}+B^{2})^{2}
8(AB)2(A2+AB+B2)2=8(A3B3)2,\displaystyle\leq 8(A-B)^{2}(A^{2}+AB+B^{2})^{2}=8(A^{3}-B^{3})^{2},

we deduce

γ4Δ|A2B2|γ12λ3Δ3+2γ3λ32(A3B3),\displaystyle\frac{\gamma}{4}\Delta\left|A^{2}-B^{2}\right|\leq\frac{\gamma}{12\lambda^{3}}\Delta^{3}+{\sqrt{2}}\frac{\gamma}{3}\lambda^{\frac{3}{2}}(A^{3}-B^{3}),

and we see that choosing λ3=α2/(32γ2)\lambda^{3}=\alpha^{2}/(32\gamma^{2}) leads to

γ4Δ|A2B2|83γ3α2Δ3+α12(A3B3).\displaystyle\frac{\gamma}{4}\Delta\left|A^{2}-B^{2}\right|\leq\frac{8}{3}\frac{\gamma^{3}}{\alpha^{2}}\Delta^{3}+\frac{\alpha}{12}(A^{3}-B^{3}).

Plugging this into (2.8) yields

D(u,u+;u,ur)\displaystyle D(u_{-},u_{+};u_{\ell},u_{r}) (1+32γ±3βα2)β12Δ3α12(A±3B±3).\displaystyle\leq\left(1+32\frac{\gamma_{\pm}^{3}}{\beta\alpha^{2}}\right)\frac{\beta}{12}\Delta^{3}-\frac{\alpha}{12}(A_{\pm}^{3}-B_{\pm}^{3}).

Recalling (2.5)-(2.6) this implies

D(u,u+;u,ur)\displaystyle D(u_{-},u_{+};u_{\ell},u_{r}) CM3α3E(u,u+)α12(A±3B±3),\displaystyle\leq C\frac{M^{3}}{\alpha^{3}}E(u_{-},u_{+})-\frac{\alpha}{12}(A_{\pm}^{3}-B_{\pm}^{3}),

for C=33C=33. Remarking that

A3B3\displaystyle A^{3}-B^{3} =(AB)(A2+AB+B2)(AB)A2+B22,\displaystyle=(A-B)(A^{2}+AB+B^{2})\geq(A-B)\frac{A^{2}+B^{2}}{2},
A±B±\displaystyle A_{\pm}-B_{\pm} =uur,\displaystyle=u_{\ell}-u_{r},

we deduce

D(u,u+;u,ur)\displaystyle D(u_{-},u_{+};u_{\ell},u_{r}) CM3α3E(u,u+)\displaystyle\leq C\frac{M^{3}}{\alpha^{3}}E(u_{-},u_{+})
α24(uur)[(uu)2+(uru+)2],\displaystyle\quad-\frac{\alpha}{24}(u_{\ell}-u_{r})\left[(u_{\ell}-u_{-})^{2}+(u_{r}-u_{+})^{2}\right],

which proves Proposition 2.1 when u+uu_{+}\geq u_{-}.

Case 2: u>u+u_{-}>u_{+}.

Following [24] (see also [14, Lemma 3.3]) we write DD as a combination of integrals of the function

g(t)\displaystyle g(t) =f(t)σ(u,u+)t(f(u+)σ(u,u+)u+)\displaystyle=f(t)-\sigma(u_{-},u_{+})t-(f(u_{+})-\sigma(u_{-},u_{+})u_{+})
=f(t)σ(u,u+)t(f(u)σ(u,u+)u),\displaystyle=f(t)-\sigma(u_{-},u_{+})t-(f(u_{-})-\sigma(u_{-},u_{+})u_{-}),

where the second equality follows from the definition of σ\sigma. Specifically we have

D(u,u+;u,ur)\displaystyle D(u_{-},u_{+};u_{\ell},u_{r}) =uru+g(t)𝑑t+uug(t)𝑑t.\displaystyle=-\int_{u_{r}}^{u_{+}}g(t)\,dt+\int_{u_{\ell}}^{u_{-}}g(t)\,dt.

As remarked in [24], since u+uu_{+}\leq u_{-} and uruu_{r}\leq u_{\ell}, this can be written as

(2.9) D=I|g|+J|g|,\displaystyle-D=\int_{I}\left|g\right|+\int_{J}\left|g\right|,

where II and JJ are disjoint intervals such that

IJ=([u+,u][ur,u])([u+,u][ur,u]).\displaystyle I\cup J=([u_{+},u_{-}]\cup[u_{r},u_{\ell}])\setminus([u_{+},u_{-}]\cap[u_{r},u_{\ell}]).

We denote by g0g_{0} the function gg in the special case f(t)=f0(t)=t2/2f(t)=f_{0}(t)=t^{2}/2, i.e.

g0(t)=12(tu+)(tu).\displaystyle g_{0}(t)=\frac{1}{2}(t-u_{+})(t-u_{-}).

Since gαg0g-\alpha g_{0} is a convex function vanishing at u+u_{+} and uu_{-}, we have

g(t)\displaystyle g(t) αg0(t)for t[u+,u],\displaystyle\geq\alpha g_{0}(t)\quad\text{for }t\in\mathbb{R}\setminus[u_{+},u_{-}],
g(t)\displaystyle g(t) αg0(t)for t[u+,u].\displaystyle\leq\alpha g_{0}(t)\quad\text{for }t\in[u_{+},u_{-}].

Therefore |g|α|g0|\left|g\right|\geq\alpha\left|g_{0}\right| on \mathbb{R} and one infers from (2.9) that

DαD0,\displaystyle-D\geq-\alpha D_{0},

where D0D_{0} is the dissipation rate (2.2) for f(t)=f0(t)=t2/2f(t)=f_{0}(t)=t^{2}/2. To conclude the proof of Proposition 2.1 it suffices to check that

(2.10) D0(u,u+;u,ur)c(uur)[(uu)2+(uru+)2],\displaystyle-D_{0}(u_{-},u_{+};u_{\ell},u_{r})\geq c(u_{\ell}-u_{r})\left[(u_{\ell}-u_{-})^{2}+(u_{r}-u_{+})^{2}\right],

for some absolute constant c>0c>0. To this end we compute

u+urg0\displaystyle\int_{u_{+}}^{u_{r}}g_{0} =16(uru+)3+14(u+u)(uru+)2,\displaystyle=\frac{1}{6}(u_{r}-u_{+})^{3}+\frac{1}{4}(u_{+}-u_{-})(u_{r}-u_{+})^{2},
uug0\displaystyle\int_{u_{-}}^{u_{\ell}}g_{0} =16(uu)3+14(uu+)(uu)2,\displaystyle=\frac{1}{6}(u_{\ell}-u_{-})^{3}+\frac{1}{4}(u_{-}-u_{+})(u_{\ell}-u_{-})^{2},

so that, setting H=uuH=u_{\ell}-u_{-} and K=uru+K=u_{r}-u_{+}, we find

D0\displaystyle-D_{0} =14(uu+)(H2+K2)+16(H3K3).\displaystyle=\frac{1}{4}(u_{-}-u_{+})(H^{2}+K^{2})+\frac{1}{6}(H^{3}-K^{3}).

This implies

D0\displaystyle-D_{0} =14(uurH+K)(H2+K2)+16(H3K3)\displaystyle=\frac{1}{4}(u_{\ell}-u_{r}-H+K)(H^{2}+K^{2})+\frac{1}{6}(H^{3}-K^{3})
=14(uur)(H2+K2)112(HK)3\displaystyle=\frac{1}{4}(u_{\ell}-u_{r})(H^{2}+K^{2})-\frac{1}{12}(H-K)^{3}

As HK=uur(uu+)uurH-K=u_{\ell}-u_{r}-(u_{-}-u_{+})\leq u_{\ell}-u_{r} we deduce

D0\displaystyle-D_{0} 14(uur)(H2+K2)112(uur)(HK)2\displaystyle\geq\frac{1}{4}(u_{\ell}-u_{r})(H^{2}+K^{2})-\frac{1}{12}(u_{\ell}-u_{r})(H-K)^{2}
=14(uur)(H2+K213(HK)2)\displaystyle=\frac{1}{4}(u_{\ell}-u_{r})\left(H^{2}+K^{2}-\frac{1}{3}(H-K)^{2}\right)
112(uur)(H2+K2),\displaystyle\geq\frac{1}{12}(u_{\ell}-u_{r})(H^{2}+K^{2}),

proving (2.10) with c=1/12c=1/12. ∎

3. The stability estimate for shocks

We follow [18, 19, 24], where uu is an entropy solution, and explain how their methods adapt to our more general situation. Our goal is to control the increase of

(3.1) F(t)=R+StRStη(u(t,x)|u0shock(xx(t)))𝑑x,\displaystyle F(t)=\int_{-R+St}^{R-St}\eta\big{(}u(t,x)\big{|}u_{0}^{shock}(x-x(t))\big{)}\,dx,

for a well-chosen Lipschitz path x(t)x(t) and RStR\geq St.

First we recall properties of the traces of uu along Lipschitz curves, which require only the strong trace property (and are thus valid for finite-entropy solutions).

Lemma 3.1.

Let uu be a weak bounded solution of (1.1) satisfying the strong trace property (1.5). Let x:[0,T]x\colon[0,T]\to\mathbb{R} be a Lipschitz path, and u(t,x(t)±)u(t,x(t)\pm) the traces of uu along x(t)x(t). Then for almost every t[0,T]t\in[0,T] we have the Rankine-Hugoniot relation

(3.2) f(u(t,x(t)+))f(u(t,x(t))=x(t)(u(t,x(t)+)u(t,x(t))).\displaystyle f(u(t,x(t)+))-f(u(t,x(t)-)=x^{\prime}(t)(u(t,x(t)+)-u(t,x(t)-)).
Proof of Lemma 3.1.

This is proved in [19, Lemma 6]. We sketch the proof for the reader’s convenience. The Rankine Hugoniot relation (3.2) follows from testing the equation

tu+xf(u)=0,\displaystyle\partial_{t}u+\partial_{x}f(u)=0,

against a test function χ\chi of the form

χ(t,ξ)=ψ(t)(Φε(ξx(t))+Φε(x(t)ξ)1),\displaystyle\chi(t,\xi)=\psi(t)\left(\Phi_{\varepsilon}(\xi-x(t))+\Phi_{\varepsilon}(x(t)-\xi)-1\right),
where 𝟏y<εΦε(y)𝟏y<0,\displaystyle\mathbf{1}_{y<-\varepsilon}\leq\Phi_{\varepsilon}(y)\leq\mathbf{1}_{y<0},

The strong trace property (1.5) ensures convergence, as ε0\varepsilon\to 0, to (3.2) tested against ψ(t)\psi(t). ∎

Next we establish a formula for the variations of quantities of the form

tx(t)y(t)η(u(t,x)|v0)𝑑x,\displaystyle t\mapsto\int_{x(t)}^{y(t)}\eta(u(t,x)|v_{0})\,dx,

for some constant v0v_{0}.

Lemma 3.2.

Let uu be a finite-entropy (1.4) solution of (1.1). Let y,z:[0,T]y,z\colon[0,T]\to\mathbb{R} be Lipschitz paths, let 0t1<t2T0\leq t_{1}<t_{2}\leq T and assume that

y(τ)<z(τ)τ(t1,t2).\displaystyle y(\tau)<z(\tau)\qquad\forall\tau\in(t_{1},t_{2}).

For any v0v_{0}\in\mathbb{R}, we have

(3.3) y(t2)z(t2)η(u(t2,ξ)|v0)dξy(t1)z(t1)η(u(t1),ξ)|v0)dξ\displaystyle\int_{y(t_{2})}^{z(t_{2})}\eta(u(t_{2},\xi)|v_{0})\,d\xi-\int_{y(t_{1})}^{z(t_{1})}\eta(u(t_{1}),\xi)|v_{0})\,d\xi
=𝟏t1<τ<t2,y(τ)<ξ<z(τ)μ(dτ,dξ)\displaystyle=\int\mathbf{1}_{t_{1}<\tau<t_{2},\,y(\tau)<\xi<z(\tau)}\;\mu(d\tau,d\xi)
+t1t2[q(u(τ,y(τ)+);v0)y(τ)η(u(τ,y(τ)+)|v0)]𝑑τ\displaystyle\quad+\int_{t_{1}}^{t_{2}}\left[q(u(\tau,y(\tau)+);v_{0})-y^{\prime}(\tau)\eta(u(\tau,y(\tau)+)|v_{0})\right]\,d\tau
t1t2[q(u(τ,z(τ));v0)z(τ)η(u(τ,z(τ))|v0)]𝑑τ.\displaystyle\quad-\int_{t_{1}}^{t_{2}}\left[q(u(\tau,z(\tau)-);v_{0})-z^{\prime}(\tau)\eta(u(\tau,z(\tau)-)|v_{0})\right]\,d\tau.
Proof of Lemma 3.2.

The proof is essentially the same as e.g. [19, Lemma 6], see also [13, Lemma 2.4]. We sketch it here for the reader’s convenience, the only difference being that we keep the terms involving μ\mu. We may assume without loss of generality that y<zy<z in [t1,t2][t_{1},t_{2}] (otherwise consider instead [t1+δ,t2δ][t_{1}+\delta,t_{2}-\delta] and let δ0+\delta\to 0^{+} at the end).

We test the identity

tη(u|u0)+xq(u|u0)=μ,\displaystyle\partial_{t}\eta(u|u_{0})+\partial_{x}q(u|u_{0})=\mu,

against a test function χ\chi of the form

χ(τ,ξ)=ψε(τ)(Φε(y(τ)ξ)Φε(ξz(τ))),\displaystyle\chi(\tau,\xi)=\psi_{\varepsilon}(\tau)\left(\Phi_{\varepsilon}(y(\tau)-\xi)-\Phi_{\varepsilon}(\xi-z(\tau))\right),
where 𝟏t1+ε<τ<t2εψε(τ)𝟏t1<τ<t2,𝟏x<εΦε(x)𝟏x<0.\displaystyle\mathbf{1}_{t_{1}+\varepsilon<\tau<t_{2}-\varepsilon}\leq\psi_{\varepsilon}(\tau)\leq\mathbf{1}_{t_{1}<\tau<t_{2}},\quad\mathbf{1}_{x<-\varepsilon}\leq\Phi_{\varepsilon}(x)\leq\mathbf{1}_{x<0}.

and obtain (3.3) as ε0+\varepsilon\to 0^{+}, thanks to the strong trace property (1.5) and the time-continuity property (1.6). ∎

Using Lemma 3.2 we will obtain a formula for the variations of F(t)F(t) (3.1), and thanks to Lemma 3.1 we will see that at any time tt where u(t,)u(t,\cdot) has a jump (u,u+)(u_{-},u_{+}) at x(t)x(t), the increase of F(t)F(t) is controlled by μ\mu plus the dissipation rate DD, which owing to Proposition 2.1 is in turn controlled by μ+\mu_{+}. Note that so far this is valid for any Lipschitz curve x(t)x(t). However, in order to control the increase of F(t)F(t) at times tt where u(t,)u(t,\cdot) does not jump at x(t)x(t), we need to constrain x(t)x^{\prime}(t). The next lemma gives us a tool to do so. This is the only place where we require the very strong trace property.

Lemma 3.3.

Let uu be a bounded finite-entropy (1.4) solution of (1.1) and assume that uu satisfies the very strong trace property (1.7). Then for any x0x_{0}\in\mathbb{R} there exists a generalized characteristic of uu starting at x0x_{0}, that is, a Lipschitz path x:[0,T]x\colon[0,T]\to\mathbb{R} such that x(0)=x0x(0)=x_{0} and

(3.4) x(t)=σ(u(t,x(t)),u(t,x(t)+)for a.e. t[0,T]\displaystyle x^{\prime}(t)=\sigma(u(t,x(t)-),u(t,x(t)+)\qquad\text{for a.e. }t\in[0,T]

where u(t,x(t)±)u(t,x(t)\pm) denote the traces of uu along x(t)x(t) and σ\sigma is the shock speed (2.1)

Proof of Lemma 3.3.

This is proved e.g. in [19, Proposition 1]. The path xx is obtained as a limit of paths xk(t)x_{k}(t) solving xk(t)=Φk(t,xk(t))x_{k}^{\prime}(t)=\Phi_{k}(t,x_{k}(t)), where Φk\Phi_{k} is a mollification (with respect to the xx variable) of fuf^{\prime}\circ u. The very strong trace property then ensures that xx satisfies

min{f(u(t,x(t)±))}x(t)max{f(u(t,x(t)±))}for a.e. t[0,T],\displaystyle\min\{f^{\prime}(u(t,x(t)\pm))\}\leq x^{\prime}(t)\leq\max\{f^{\prime}(u(t,x(t)\pm))\}\quad\text{for a.e. }t\in[0,T],

so that x(t)=f(u(t,x(t)))x^{\prime}(t)=f^{\prime}(u(t,x(t))) for a.e. tt where there is no jump, i.e. u(t,x(t))=u(t,x(t)+=u(t,x(t))u(t,x(t)-)=u(t,x(t)+=u(t,x(t)), and at jump points (3.4) follows from (3.2). ∎

We are now ready to prove our main result.

Proof of Theorem 1.1.

We apply Lemma 3.3 and let the Lipschitz path x:[0,T]x\colon[0,T]\to\mathbb{R} be a generalized characteristic starting at 0, i.e. x(0)=0x(0)=0 and

(3.5) x(t)=σ(u(t),u+(t))for a.e. t[0,T],\displaystyle x^{\prime}(t)=\sigma(u_{-}(t),u_{+}(t))\quad\text{for a.e. }t\in[0,T],

where u±(t)=u(t,x(t)±)u_{\pm}(t)=u(t,x(t)\pm) denote the traces of uu along x(t)x(t).

Let R>0R>0 and F(t)F(t) defined by (3.1) for all tR/St\leq R/S. We assume without loss of generality that RSTR\geq ST (otherwise replace TT by R/SR/S). Consider the time

t=sup{t[0,T]:R+St<x(t)<RSt}.\displaystyle t_{*}=\sup\left\{t\in[0,T]\colon-R+St<x(t)<R-St\right\}.

By definition of S=supI|f|S=\sup_{I}|f^{\prime}| we know that |x|S\left|x^{\prime}\right|\leq S and deduce

R+St<x(t)<RStt[0,t),\displaystyle-R+St<x(t)<R-St\qquad\forall t\in[0,t_{*}),
and x(t)(,R+St][RSt,)tt.\displaystyle x(t)\in(-\infty,-R+St]\cup[R-St,\infty)\qquad\forall t\geq t_{*}.

For t[0,t]t\in[0,t_{*}] we have

F(t)=R+Stx(t)η(u(t,x)|u)𝑑x+x(t)RStη(u(x,t)|ur)𝑑x.\displaystyle F(t)=\int_{-R+St}^{x(t)}\eta(u(t,x)|u_{\ell})\,dx+\int_{x(t)}^{R-St}\eta(u(x,t)|u_{r})\,dx.

We apply the variation formula (3.3) to compute F(t)F(0)F(t)-F(0). Note that the identity

q(u;v)η(u|v)=2(uv)2vu(tv)f(t)𝑑t=201sf(us+v(1s))𝑑s.\displaystyle\frac{q(u;v)}{\eta(u|v)}=\frac{2}{(u-v)^{2}}\int_{v}^{u}(t-v)f^{\prime}(t)\,dt=2\int_{0}^{1}s\,f^{\prime}(us+v(1-s))\,ds.

and the definition of SS ensure that

(3.6) |q(u;v)|Sη(u|v)u,vI=[min(u,infu),max(ur,supu)].\displaystyle\left|q(u;v)\right|\leq S\eta(u|v)\qquad\forall u,v\in I=[\min(u_{\ell},\inf u),\max(u_{r},\sup u)].

As a consequence, whenever y(t)=Sy^{\prime}(t)=S or z(t)=Sz^{\prime}(t)=-S the corresponding term in the right-hand side of (3.3) gives a nonpositive contribution, and we deduce

F(t)F(0)\displaystyle F(t)-F(0) μ+(BR,tS)\displaystyle\leq\mu_{+}(B_{R,t}^{S})
+0t[q(u+(τ);ur)q(u(τ);u)\displaystyle\quad+\int_{0}^{t}\big{[}q(u_{+}(\tau);u_{r})-q(u_{-}(\tau);u_{\ell})
x(τ)(η(u+(τ)|ur)η(u(τ)|u)]dτ,\displaystyle\hskip 50.00008pt-x^{\prime}(\tau)\left(\eta(u_{+}(\tau)|u_{r})-\eta(u_{-}(\tau)|u_{\ell}\right)\big{]}\,d\tau,

where

BR,tS\displaystyle B_{R,t}^{S} ={(τ,ξ):0<τ<t,R+τS<ξ<RτS}.\displaystyle=\left\{(\tau,\xi)\colon 0<\tau<t,-R+\tau S<\xi<R-\tau S\right\}.

Recalling (3.5) that x=σ(u,u+)x^{\prime}=\sigma(u_{-},u_{+}) a.e. in [0,T][0,T], we recognize the dissipation rate DD (2.2) and rewrite the above as

(3.7) F(t)F(0)\displaystyle F(t)-F(0) μ+(BR,tS)+0tD(u(τ),u+(τ);u,ur)𝑑τ.\displaystyle\leq\mu_{+}(B_{R,t}^{S})+\int_{0}^{t}D(u_{-}(\tau),u_{+}(\tau);u_{\ell},u_{r})\,d\tau.

Since

σ(u,u+)σ(u,ur)=01[f(tu+(1t)u+)f(tu+(1t)ur)]𝑑t.\displaystyle\sigma(u_{-},u_{+})-\sigma(u_{\ell},u_{r})=\int_{0}^{1}\left[f^{\prime}(tu_{-}+(1-t)u_{+})-f^{\prime}(tu_{\ell}+(1-t)u_{r})\right]dt.

and ff^{\prime} is MM-Lipschitz on II we have

|σ(u,u+)σ(u,ur)|M2(|uu|+|uru+|),\displaystyle\left|\sigma(u_{-},u_{+})-\sigma(u_{\ell},u_{r})\right|\leq\frac{M}{2}\left(\left|u_{\ell}-u_{-}\right|+\left|u_{r}-u_{+}\right|\right),

so using the upper bound on DD provided by Proposition 2.1 we deduce from (3.7) that

F(t)F(0)\displaystyle F(t)-F(0) μ+(BR,tS)+C1M3α30tmax(E(τ),0)𝑑τ\displaystyle\leq\mu_{+}(B_{R,t}^{S})+C_{1}\frac{M^{3}}{\alpha^{3}}\int_{0}^{t}\max(E(\tau),0)\,d\tau
C2(uur)αM20t(x(τ)σ)2𝑑τ,\displaystyle\quad-C_{2}(u_{\ell}-u_{r})\frac{\alpha}{M^{2}}\int_{0}^{t}(x^{\prime}(\tau)-\sigma)^{2}\,d\tau,

where E(τ)E(\tau) is the entropy cost of the jump (u(τ),u+(τ))(u_{-}(\tau),u_{+}(\tau)) (2.3). From the characterization [17, 8] of the one-dimensional part of μ\mu we have

μ+{(τ,x(τ))}=max(E(τ),0)dτ,\displaystyle\mu_{+\lfloor\{(\tau,x(\tau))\}}=\max(E(\tau),0)\,d\tau,

and obtain

(3.8) F(t)\displaystyle F(t) F(0)+CM3α3μ+(BR,tS)\displaystyle\leq F(0)+C\frac{M^{3}}{\alpha^{3}}\mu_{+}(B_{R,t}^{S})
C2(uur)αM20t(x(τ)σ)2𝑑τ,t[0,t].\displaystyle\quad-C_{2}(u_{\ell}-u_{r})\frac{\alpha}{M^{2}}\int_{0}^{t}(x^{\prime}(\tau)-\sigma)^{2}\,d\tau,\qquad\forall t\in[0,t_{*}].

Now for t[t,T]t\in[t_{*},T] we have

F(t)=R+StRStη(u(x,t)|v0)𝑑x,\displaystyle F(t)=\int_{-R+St}^{R-St}\eta(u(x,t)|v_{0})\,dx,

where v0=urv_{0}=u_{r} or uu_{\ell}. Therefore, applying (3.3) and remarking again that the terms involving y(t)=Sy^{\prime}(t)=S and z(t)=Sz^{\prime}(t)=-S give nonpositive contributions, we obtain

F(t)F(t)\displaystyle F(t)-F(t_{*}) μ+(BR,tSBR,tS)t[t,T].\displaystyle\leq\mu_{+}(B^{S}_{R,t}\setminus B^{S}_{R,t_{*}})\qquad\forall t\in[t_{*},T].

Combining this with the estimate obtained in [0,t][0,t_{*}] and recalling the definition (3.1) of F(t)F(t) we deduce that

R+tSRtS|u(x,t)u0shock(xx(t))|2𝑑x\displaystyle\int_{-R+tS}^{R-tS}\left|u(x,t)-u_{0}^{shock}(x-x(t))\right|^{2}\,dx RR|u0u0shock|2𝑑x\displaystyle\leq\int_{-R}^{R}\left|u_{0}-u_{0}^{shock}\right|^{2}\,dx
+CM3α3μ+(BR,tS)t[0,T].\displaystyle\quad+C\frac{M^{3}}{\alpha^{3}}\mu_{+}(B_{R,t}^{S})\qquad\forall t\in[0,T].

Provided we set h(t)=x(t)σth(t)=x(t)-\sigma\,t and replace RR by R+StR+St, this implies our main result (1.9) since BR+St,tS[0,t]×[RSt,R+St]B_{R+St,t}^{S}\subset[0,t]\times[-R-St,R+St].

To prove estimate (1.10) on h(t)h(t), simply remark that (3.8) readily implies that

cαM2(uur)0th(τ)2𝑑τRR(u0u0shock)2𝑑x+M3α3μ+(BR,tS)\displaystyle c\frac{\alpha}{M^{2}}(u_{\ell}-u_{r})\int_{0}^{t}h^{\prime}(\tau)^{2}\,d\tau\leq\int_{-R}^{R}(u_{0}-u_{0}^{shock})^{2}\,dx+\frac{M^{3}}{\alpha^{3}}\mu_{+}(B_{R,t}^{S})

for some absolute constant c>0c>0, provided R+St<x(t)<RSt-R+St<x(t)<R-St. Since we know that |x(t)|St\left|x(t)\right|\leq St we may choose R=2StR=2St, and deduce (1.10). ∎

4. The stability estimate for BVBV initial data

This section is dedicated to Theorem 1.4’s proof, following the scheme introduced in [13]. It is based on considering initial conditions ζ0\zeta_{0} with a finite number of entropic shocks at x10<<xN0x_{1}^{0}<\cdots<x_{N}^{0}, of the form

(4.1) ζ0(x)=v00(x)𝟏x<x10+v10(x)𝟏x10<x<x20++vN0(x)𝟏x>xN0,\displaystyle\zeta_{0}(x)=v^{0}_{0}(x)\mathbf{1}_{x<x^{0}_{1}}+v^{0}_{1}(x)\mathbf{1}_{x^{0}_{1}<x<x^{0}_{2}}+\cdots+v^{0}_{N}(x)\mathbf{1}_{x>x^{0}_{N}},
where v00v10vN0 are Lipschitz nondecreasing functions on .\displaystyle\text{where }v^{0}_{0}\geq v^{0}_{1}\geq\cdots\geq v^{0}_{N}\text{ are Lipschitz nondecreasing functions on }\mathbb{R}.

Since ζ0\zeta_{0} depends only on the restrictions of vj0v^{0}_{j} to pairwise disjoint intervals, we may assume that their extensions to \mathbb{R}, in addition to being Lipschitz and nondecreasing, satisfy

infζ0vj0supζ0and0vj10vj0dj0on ,\displaystyle\inf\zeta_{0}\leq v^{0}_{j}\leq\sup\zeta_{0}\quad\text{and}\quad 0\leq v^{0}_{j-1}-v^{0}_{j}\leq d^{0}_{j}\qquad\text{on }\mathbb{R},

for all j{0,,N}j\in\{0,\ldots,N\}, where dj0:=(vj10vj0)(x0j)d^{0}_{j}:=(v^{0}_{j-1}-v^{0}_{j})(x_{0}^{j}) is the schock amplitude of ζ0\zeta_{0} at xj0x_{j}^{0}. Note that with these notations the negative part of Dζ0D\zeta_{0} is given by (Dζ0)=jdj0δxj0(D\zeta_{0})_{-}=\sum_{j}d_{j}^{0}\delta_{x_{j}^{0}}.

The function u~\tilde{u} appearing in Theorem 1.4 is then going to be piecewise equal to the entropy solutions vjv_{j} of (1.1) with initial data vj(0,x)=vj0(x)v_{j}(0,x)=v_{j}^{0}(x). Since ff is convex and the vj0v_{j}^{0} are Lipschitz nondecreasing, vjv_{j} is directly obtained by the method of characteristics and we have that

vj is a Lipschitz solution of (1.1) and xvj0j{0,,N}.\displaystyle v_{j}\text{ is a Lipschitz solution of \eqref{eq:scl} and }\partial_{x}v_{j}\geq 0\qquad\forall j\in\{0,\ldots,N\}.

Moreover the entropy solutions vjv_{j} are ordered as their initial data

supζ0v0v1vNinfζ0,\displaystyle\sup\zeta_{0}\geq v_{0}\geq v_{1}\geq\cdots\geq v_{N}\geq\inf\zeta_{0},

and their differences have the following nonincreasing property: for any time t>0t>0 and R>0R>0, and intervals I=[R,R]I=[-R,R], I0=[RSt,R+St]I_{0}=[-R-St,R+St] we have

(4.2) supI(vj1(t,)vj(t,))supI0(vj10vj0)dj0.\displaystyle\sup_{I}(v_{j-1}(t,\cdot)-v_{j}(t,\cdot))\leq\sup_{I_{0}}(v^{0}_{j-1}-v_{j}^{0})\leq d_{j}^{0}.

This last assertion follows from the method of characteristics: for all xIx\in I there are yzI0y\leq z\in I_{0} such that

vj1(t,x)vj(t,x)=vj10(y)vj0(z)vj10(z)vj0(z).\displaystyle v_{j-1}(t,x)-v_{j}(t,x)=v^{0}_{j-1}(y)-v^{0}_{j}(z)\leq v^{0}_{j-1}(z)-v^{0}_{j}(z).

We construct the function u~\tilde{u} by shifting the shocks between the functions vjv_{j}. To deal with shocks between nondecreasing functions (instead of constant functions as in Theorem 1.1), we need an equivalent of Lemma 3.2 where the constant v0v_{0} is replaced by a Lipschitz solution vv of (1.1) with xv0\partial_{x}v\geq 0. This corresponds to [13, Lemma 2.4] which we transpose here to our setting:

Lemma 4.1.

Let uu be a finite-entropy (1.4) solution of (1.1). Let y,z:[0,T]y,z\colon[0,T]\to\mathbb{R} be Lipschitz paths, let 0t1<t2T0\leq t_{1}<t_{2}\leq T and assume that

y(τ)<z(τ)τ(t1,t2).\displaystyle y(\tau)<z(\tau)\qquad\forall\tau\in(t_{1},t_{2}).

For any Lipschitz solution vv of (1.1) such that xv0\partial_{x}v\geq 0 we have

(4.3) y(t2)z(t2)η(u(t2,ξ)|v(t2,ξ))dξy(t1)z(t1)η(u(t1),ξ)|v(t1,ξ))dξ\displaystyle\int_{y(t_{2})}^{z(t_{2})}\eta(u(t_{2},\xi)|v(t_{2},\xi))\,d\xi-\int_{y(t_{1})}^{z(t_{1})}\eta(u(t_{1}),\xi)|v(t_{1},\xi))\,d\xi
𝟏t1<τ<t2,y(τ)<ξ<z(τ)μ(dτ,dξ)\displaystyle\leq\int\mathbf{1}_{t_{1}<\tau<t_{2},\,y(\tau)<\xi<z(\tau)}\;\mu(d\tau,d\xi)
+t1t2[q(u(τ,y(τ)+);v(τ,y(τ)))y(τ)η(u(τ,y(τ)+)|v(τ,y(τ)))]𝑑τ\displaystyle\quad+\int_{t_{1}}^{t_{2}}\left[q(u(\tau,y(\tau)+);v(\tau,y(\tau)))-y^{\prime}(\tau)\eta(u(\tau,y(\tau)+)|v(\tau,y(\tau)))\right]\,d\tau
t1t2[q(u(τ,z(τ));v(τ,z(τ)))z(τ)η(u(τ,z(τ))|v(τ,z(τ)))]𝑑τ.\displaystyle\quad-\int_{t_{1}}^{t_{2}}\left[q(u(\tau,z(\tau)-);v(\tau,z(\tau)))-z^{\prime}(\tau)\eta(u(\tau,z(\tau)-)|v(\tau,z(\tau)))\right]\,d\tau.
Proof of Lemma 4.1.

Since vv is Lipschitz we may use the chain rule and find

tη(u|v)+xq(u|v)\displaystyle\partial_{t}\eta(u|v)+\partial_{x}q(u|v) =μη′′(v)xv[f(u)f(v)f(v)(uv)]\displaystyle=\mu-\eta^{\prime\prime}(v)\partial_{x}v\left[f(u)-f(v)-f^{\prime}(v)(u-v)\right]
μ,\displaystyle\leq\mu,

because ff is convex and η′′(v)xv0\eta^{\prime\prime}(v)\partial_{x}v\geq 0. Then (4.3) follows as in the proof of Lemma 3.2. ∎

Equipped with Lemma 4.1, the construction and estimates of u~\tilde{u} become very similar to the proof of Theorem 1.4 as it simply consists in shifting the shocks along generalized characteristics of uu. One small technical difficulty is that the curves may merge, or cross the bounds of integration. We start by considering the simplest setting where no merging nor crossing happens:

Proposition 4.2.

Let t1<t2t_{1}<t_{2} and u:[t1,t2]×u\colon[t_{1},t_{2}]\times\mathbb{R}\to\mathbb{R} be a bounded finite-entropy solution (1.4) of (1.1). Let x1,,xN:[t1,t2]x_{1},\ldots,x_{N}\colon[t_{1},t_{2}]\to\mathbb{R} be generalized characteristics of uu, that is,

xj(t)=σ(u(t,xj(t)),u(t,xj(t)+)for a.e. t[t1,t2].\displaystyle x_{j}^{\prime}(t)=\sigma(u(t,x_{j}(t)-),u(t,x_{j}(t)+)\qquad\text{for a.e. }t\in[t_{1},t_{2}].

Assume that R>0R>0 is such that for all τ(t1,t2)\tau\in(t_{1},t_{2}),

RS(t2τ):=x0(τ)<x1(τ)<<xN(τ)<xN+1(τ):=R+S(t2τ).\displaystyle-R-S(t_{2}-\tau):=x_{0}(\tau)<x_{1}(\tau)<\cdots<x_{N}(\tau)<x_{N+1}(\tau):=R+S(t_{2}-\tau).

Then, for any v0vNv_{0}\geq\cdots\geq v_{N} Lipschitz solutions of (1.1) such that xvj0\partial_{x}v_{j}\geq 0 for j=0,,Nj=0,\ldots,N, and u~(t,x)\tilde{u}(t,x) defined by

u~(t,x)=v0(t,x)𝟏x<x1(t)+v1(t,x)𝟏x1(t)<x<x2(t)++vN(t,x)𝟏x>xN(t),\displaystyle\tilde{u}(t,x)=v_{0}(t,x)\mathbf{1}_{x<x_{1}(t)}+v_{1}(t,x)\mathbf{1}_{x_{1}(t)<x<x_{2}(t)}+\cdots+v_{N}(t,x)\mathbf{1}_{x>x_{N}(t)},

we have

(4.4) x0(t2)xN+1(t2)|u(t2,x)u~(t2,x)|2𝑑xx0(t1)xN+1(t1)|u(t1,x)u~(t1,x)|2𝑑x\displaystyle\int_{x_{0}(t_{2})}^{x_{N+1}(t_{2})}\left|u(t_{2},x)-\tilde{u}(t_{2},x)\right|^{2}dx-\int_{x_{0}(t_{1})}^{x_{N+1}(t_{1})}\left|u(t_{1},x)-\tilde{u}(t_{1},x)\right|^{2}dx
Λ(t1,t2):=CM3α3μ+((t1,t2)×(x0(t1),xN+1(t1))),\displaystyle\leq\Lambda(t_{1},t_{2}):=C\frac{M^{3}}{\alpha^{3}}\mu_{+}((t_{1},t_{2})\times(x_{0}(t_{1}),x_{N+1}(t_{1}))),

and, for any entropy solution uentu^{ent} of (1.1),

(4.5) x0(t2)xN+1(t2)|u~(t2,x)uent(t2,x)|𝑑xx0(t1)xN+1(t1)|u~(t1,x)uent(t1,x)|𝑑x\displaystyle\int_{x_{0}(t_{2})}^{x_{N+1}(t_{2})}\left|\tilde{u}(t_{2},x)-u^{ent}(t_{2},x)\right|\,dx-\int_{x_{0}(t_{1})}^{x_{N+1}(t_{1})}\left|\tilde{u}(t_{1},x)-u^{ent}(t_{1},x)\right|\,dx
j=1Nt1t2vj1(t,xj(t))vj(t,xj(t))θj(t)𝑑t,\displaystyle\leq\sum_{j=1}^{N}\int_{t_{1}}^{t_{2}}\sqrt{v_{j-1}(t,x_{j}(t))-v_{j}(t,x_{j}(t))}\,\theta_{j}(t)\,dt,

where the functions θj\theta_{j} satisfy

(4.6) αCj=1Nt1t2θj(t)2𝑑tx0(t1)xN+1(t1)|u(t1,x)u~(t1,x)|2𝑑x+Λ(t1,t2),\displaystyle\frac{\alpha}{C}\sum_{j=1}^{N}\int_{t_{1}}^{t_{2}}\theta_{j}(t)^{2}\,dt\leq\int_{x_{0}(t_{1})}^{x_{N+1}(t_{1})}\left|u(t_{1},x)-\tilde{u}(t_{1},x)\right|^{2}dx+\Lambda(t_{1},t_{2}),

and C>0C>0 is an absolute constant.

Remark 4.3.

In the case of one shock between constant functions, (4.4) corresponds to (1.9) while (4.5) can be inferred from (1.10) and Kružkov’s L1L^{1} stability estimate [5, Theorem 6.2.3].

Remark 4.4.

The right-hand side of (4.5) can be written more compactly as

Ju~|[u~]|12θ|νx|𝑑1,\displaystyle\int_{J_{\tilde{u}}}|[\tilde{u}]|^{\frac{1}{2}}\theta\,|\nu_{x}|\,d\mathcal{H}^{1},

where Ju~=j{(t,xj(t)}t[t1,t2]J_{\tilde{u}}=\bigcup_{j}\{(t,x_{j}(t)\}_{t\in[t_{1},t_{2}]} is the jump set of u~\tilde{u} with normal vector ν=(νt,νx)\nu=(\nu_{t},\nu_{x}), [u~]=u~+u~[\tilde{u}]=\tilde{u}_{+}-\tilde{u}_{-} denotes the jump of u~\tilde{u} along that jump set, and the function θ\theta is defined on Ju~J_{\tilde{u}} by θ(t,xj(t))=θj(t)\theta(t,x_{j}(t))=\theta_{j}(t). Then (4.6) translates into

αCJu~θ2|νx|𝑑1x0(t1)xN+1(t1)|u(t1,x)u~(t1,x)|2𝑑x+Λ(t1,t2).\displaystyle\frac{\alpha}{C}\int_{J_{\tilde{u}}}\theta^{2}\,|\nu_{x}|\,d\mathcal{H}^{1}\leq\int_{x_{0}(t_{1})}^{x_{N+1}(t_{1})}\left|u(t_{1},x)-\tilde{u}(t_{1},x)\right|^{2}dx+\Lambda(t_{1},t_{2}).
Proof of Proposition 4.2.

For all j{0,,N}j\in\{0,\ldots,N\} we apply Lemma 4.1 to the paths y=xjy=x_{j}, z=xj+1z=x_{j+1} and Lipschitz nondecreasing entropy solution vjv_{j}:

xj(t2)xj+1(t2)|u(t2,x)u~(t2,x)|2𝑑xxj(t1)xj+1(t1)|u(t1,x)u~(t1,x)|2𝑑x\displaystyle\int_{x_{j}(t_{2})}^{x_{j+1}(t_{2})}|u(t_{2},x)-\tilde{u}(t_{2},x)|^{2}dx-\int_{x_{j}(t_{1})}^{x_{j+1}(t_{1})}|u(t_{1},x)-\tilde{u}(t_{1},x)|^{2}dx
μ+({t1<τ<t2,xj(t)<ξ<xj+1(t)})\displaystyle\leq\mu_{+}(\{t_{1}<\tau<t_{2},x_{j}(t)<\xi<x_{j+1}(t)\})
+t1t2[q(u(τ,xj(τ)+);v(τ,xj(τ)))\displaystyle\quad+\int_{t_{1}}^{t_{2}}\big{[}q(u(\tau,x_{j}(\tau)+);v(\tau,x_{j}(\tau)))
xj(τ)η(u(τ,xj(τ)+)|v(τ,xj(τ)))]dτ\displaystyle\hskip 50.00008pt-x_{j}^{\prime}(\tau)\eta(u(\tau,x_{j}(\tau)+)|v(\tau,x_{j}(\tau)))\big{]}\,d\tau
t1t2[q(u(τ,xj+1(τ));v(τ,xj+1(τ)))\displaystyle\quad-\int_{t_{1}}^{t_{2}}\big{[}q(u(\tau,x_{j+1}(\tau)-);v(\tau,x_{j+1}(\tau)))
xj+1(τ)η(u(τ,xj+1(τ))|v(τ,xj+1(τ)))]dτ.\displaystyle\hskip 50.00008pt-x_{j+1}^{\prime}(\tau)\eta(u(\tau,x_{j+1}(\tau)-)|v(\tau,x_{j+1}(\tau)))\big{]}\,d\tau.

Summing all these inequalities, we deduce

x0(t2)xN+1(t2)|u(t2,x)u~(t2,x)|2𝑑xx0(t1)xN+1(t1)|u(t1,x)u~(t1,x)|2𝑑x\displaystyle\int_{x_{0}(t_{2})}^{x_{N+1}(t_{2})}\left|u(t_{2},x)-\tilde{u}(t_{2},x)\right|^{2}dx-\int_{x_{0}(t_{1})}^{x_{N+1}(t_{1})}\left|u(t_{1},x)-\tilde{u}(t_{1},x)\right|^{2}dx
μ+((t1,t2)×(x0(t1),xN+1(t1))\displaystyle\leq\mu_{+}((t_{1},t_{2})\times(x_{0}(t_{1}),x_{N+1}(t_{1}))
+j=1Nt1t2D(u(t,xj(τ)),u(τ,xj(τ)+);vj1(τ,xj(τ)),vj(τ,xj(τ)))𝑑τ\displaystyle\quad+\sum_{j=1}^{N}\int_{t_{1}}^{t_{2}}D(u(t,x_{j}(\tau)-),u(\tau,x_{j}(\tau)+);v_{j-1}(\tau,x_{j}(\tau)),v_{j}(\tau,x_{j}(\tau)))\,d\tau

Here we discarded the boundary terms involving the paths x0x_{0} and xN+1x_{N+1} as they give nonpositive contributions thanks to (3.6). Since v0vNv_{0}\geq\cdots\geq v_{N} we may apply Proposition 2.1 to obtain, as for (3.8) in the proof of Theorem 1.1,

x0(t2)xN+1(t2)|u(t2,x)u~(t2,x)|2𝑑xx0(t1)xN+1(t1)|u(t1,x)u~(t1,x)|2𝑑x\displaystyle\int_{x_{0}(t_{2})}^{x_{N+1}(t_{2})}\left|u(t_{2},x)-\tilde{u}(t_{2},x)\right|^{2}dx-\int_{x_{0}(t_{1})}^{x_{N+1}(t_{1})}\left|u(t_{1},x)-\tilde{u}(t_{1},x)\right|^{2}dx
Λ(t1,t2)αCj=1Nt1t2(vj1(τ,xj(τ))vj(τ,xj(τ))sj(τ)2dτ,\displaystyle\leq\Lambda(t_{1},t_{2})-\frac{\alpha}{C}\sum_{j=1}^{N}\int_{t_{1}}^{t_{2}}(v_{j-1}(\tau,x_{j}(\tau))-v_{j}(\tau,x_{j}(\tau))\,s_{j}(\tau)^{2}\,d\tau,

where

(4.7) sj(t)=xj(t)σ(vj1(t,xj(t)),vj(t,xj(t))).\displaystyle s_{j}(t)=x_{j}^{\prime}(t)-\sigma(v_{j-1}(t,x_{j}(t)),v_{j}(t,x_{j}(t))).

This implies in particular (4.4). Note for later use that it also implies

(4.8) αCM2j=1Nt1t2(vj1(τ,xj(τ))vj(τ,xj(τ)))sj(τ)2𝑑τ\displaystyle\frac{\alpha}{CM^{2}}\sum_{j=1}^{N}\int_{t_{1}}^{t_{2}}(v_{j-1}(\tau,x_{j}(\tau))-v_{j}(\tau,x_{j}(\tau)))\,s_{j}(\tau)^{2}\,d\tau
x0(t1)xN+1(t1)|u(t1,x)u~(t1,x)|2𝑑x+Λ(t1,t2).\displaystyle\leq\int_{x_{0}(t_{1})}^{x_{N+1}(t_{1})}\left|u(t_{1},x)-\tilde{u}(t_{1},x)\right|^{2}dx+\Lambda(t_{1},t_{2}).

Now we turn to the proof of (4.5). For any convex entropy η~\tilde{\eta} and associated flux q~\tilde{q}, we compute

μ~η~:=tη~(u~)+xq~(u~)\displaystyle\tilde{\mu}_{\tilde{\eta}}:=\partial_{t}\tilde{\eta}(\tilde{u})+\partial_{x}\tilde{q}(\tilde{u})

Note that u~\tilde{u} need not be a solution of (1.1), but we can still compute this entropy production. Since u~\tilde{u} is a Lipschitz solution of (1.1) outside of the Lipschitz curves xjx_{j}, the BVBV chain rule ensures that μ~η~\tilde{\mu}_{\tilde{\eta}} is concentrated on those jump curves, and

μ~η~\displaystyle\tilde{\mu}_{\tilde{\eta}} =j=1N(t,xj(t)){gj(t)dt},\displaystyle=\sum_{j=1}^{N}(t,x_{j}(t))_{\sharp}\left\{g_{j}(t)\,dt\right\},
gj(t)\displaystyle g_{j}(t) =xj(t)[η~(vj(t,xj(t)))η~(vj1(t,xj(t)))]+q~(vj(t,xj(t)))q~(t,vj1(t))).\displaystyle=-x_{j}^{\prime}(t)\left[\tilde{\eta}(v_{j}(t,x_{j}(t)))-\tilde{\eta}(v_{j-1}(t,x_{j}(t)))\right]+\tilde{q}(v_{j}(t,x_{j}(t)))-\tilde{q}(t,v_{j-1}(t))).

Recalling from the definition of sjs_{j} (4.7) that xj(t)=σ(vj1(t,xj(t)),vj(t,xj(t)))+sj(t)x_{j}^{\prime}(t)=\sigma(v_{j-1}(t,x_{j}(t)),v_{j}(t,x_{j}(t)))+s_{j}(t), we see that

gj(t)\displaystyle g_{j}(t) =Eη~(vj1(t,xj(t)),vj(t,xj(t)))\displaystyle=E_{\tilde{\eta}}(v_{j-1}(t,x_{j}(t)),v_{j}(t,x_{j}(t)))
+sj(t)[η~(vj(t,xj(t)))η~(vj1(t,xj(t)))],\displaystyle\quad+s_{j}(t)\left[\tilde{\eta}(v_{j}(t,x_{j}(t)))-\tilde{\eta}(v_{j-1}(t,x_{j}(t)))\right],

where Eη~(u,u+)E_{\tilde{\eta}}(u_{-},u_{+}) denotes, as in (2.3), the entropy cost of a jump (u,u+)(u_{-},u_{+}) for the entropy η~\tilde{\eta}. Here, since vj1vjv_{j-1}\geq v_{j} we have Eη~(vj1,vj)0E_{\tilde{\eta}}(v_{j-1},v_{j})\leq 0, and therefore

(4.9) μ~η~\displaystyle\tilde{\mu}_{\tilde{\eta}} (sup|η~u~|)λ,\displaystyle\leq\left(\sup|\tilde{\eta}^{\prime}\circ\tilde{u}|\right)\,\lambda,
λ\displaystyle\lambda =j=1N(t,xj(t)){λj(t)dt},\displaystyle=\sum_{j=1}^{N}(t,x_{j}(t))_{\sharp}\left\{\lambda_{j}(t)\,dt\right\},
λj(t)\displaystyle\lambda_{j}(t) =(vj1(t,xj(t))vj(t,xj(t)))|sj(t)|\displaystyle=\left(v_{j-1}(t,x_{j}(t))-v_{j}(t,x_{j}(t))\right)\left|s_{j}(t)\right|

Even though u~\tilde{u} is not a solution, and its entropy production is not nonpositive, we claim that the proof of Kružkov’s L1L^{1} estimate [5, Theorem 6.2.3] can be adapted to obtain

(4.10) t|uentu~|+xQ(uent;u~)λ,\displaystyle\partial_{t}\,|u^{ent}-\tilde{u}|+\partial_{x}\,Q(u^{ent};\tilde{u})\leq\lambda,

where Q(u;v)=sign(uv)(f(u)f(v))Q(u;v)=\mathrm{sign}(u-v)(f(u)-f(v)). We follow [5, Theorem 6.2.3] and use the variable doubling technique: from (4.9) and the fact that uentu^{ent} is an entropy solution we have

(4.11) (t+s)|uent(s,y)u~(t,x)|+(x+y)Q(uent(s,y);u~(t,x))λ(t,x),\displaystyle(\partial_{t}+\partial_{s})|u^{ent}(s,y)-\tilde{u}(t,x)|+(\partial_{x}+\partial_{y})Q(u^{ent}(s,y);\tilde{u}(t,x))\leq\lambda(t,x),

and we test this against

Φε(t,x,s,y)=ψ(t+s2,x+y2)ρε(ts2)ρε(xy2),\displaystyle\Phi_{\varepsilon}(t,x,s,y)=\psi\left(\frac{t+s}{2},\frac{x+y}{2}\right)\rho_{\varepsilon}\left(\frac{t-s}{2}\right)\rho_{\varepsilon}\left(\frac{x-y}{2}\right),

for any nonnegative test function ψ(t,x)\psi(t,x), ρε(t)=ε1ρ(ε1t)\rho_{\varepsilon}(t)=\varepsilon^{-1}\rho(\varepsilon^{1}t) with ρ\rho a smooth nongative function with compact support and unit integral, and small enough ε>0\varepsilon>0. In the proof of [5, Theorem 6.2.3], it is shown that the left-hand side of (4.11) tested against Φε\Phi_{\varepsilon} converges to the left-hand side of (4.10) tested against ψ\psi, as ε0\varepsilon\to 0. To obtain (4.10) we only need to check that the same happens with the right-hand sides. We have

λ(t,x),Φε(t,x,s,y)\displaystyle\langle\lambda(t,x),\Phi_{\varepsilon}(t,x,s,y)\rangle
=ψ(t+s2,x+y2)ρε(ts2)ρε(xy2)λ(dt,dx)𝑑s𝑑y\displaystyle=\int\psi\left(\frac{t+s}{2},\frac{x+y}{2}\right)\rho_{\varepsilon}\left(\frac{t-s}{2}\right)\rho_{\varepsilon}\left(\frac{x-y}{2}\right)\,\lambda(dt,dx)\,dsdy
=ψ(tεs^,xεy^)λ(dt,dx)ρ(s^)ρ(y^)𝑑s^𝑑y^ε0λ,ψ,\displaystyle=\int\psi\left(t-\varepsilon\hat{s},x-\varepsilon\hat{y}\right)\,\lambda(dt,dx)\,\rho(\hat{s})\rho(\hat{y})\,d\hat{s}d\hat{y}\xrightarrow[\varepsilon\to 0]{}\langle\lambda,\psi\rangle,

by dominated convergence, since ψ\psi is bounded and λ\lambda is a finite measure. This proves (4.10). (Direct computations using the BVBV chain rule would also provide a proof.)

Testing (4.10) as in the proof of Lemma 3.2 and discarding nonpositive boundary terms, we deduce that

x0(t2)xN+1(t2)|u~(t2,x)uent(t2,x)|𝑑xx0(t1)xN+1(t2)|u~(t1,x)uent(t1,x)|𝑑x\displaystyle\int_{x_{0}(t_{2})}^{x_{N+1}(t_{2})}\left|\tilde{u}(t_{2},x)-u^{ent}(t_{2},x)\right|\,dx-\int_{x_{0}(t_{1})}^{x_{N+1}(t_{2})}\left|\tilde{u}(t_{1},x)-u^{ent}(t_{1},x)\right|\,dx
𝟏t1<t<t2,x0(t)<x<xN+1(t)λ(dt,dx)\displaystyle\leq\int\mathbf{1}_{t_{1}<t<t_{2},x_{0}(t)<x<x_{N+1}(t)}\lambda(dt,dx)
=j=1Nt1t2vj1(t,xj(t))vj(t,xj(t))θj(t)𝑑t,\displaystyle=\sum_{j=1}^{N}\int_{t_{1}}^{t_{2}}\sqrt{v_{j-1}(t,x_{j}(t))-v_{j}(t,x_{j}(t))}\,\theta_{j}(t)\,dt,

where θj(t)=vj1(t,xj(t))vj(t,xj(t))|sj(t)|\theta_{j}(t)=\sqrt{v_{j-1}(t,x_{j}(t))-v_{j}(t,x_{j}(t))}\,\left|s_{j}(t)\right| satisfies (4.6) thanks to (4.8). ∎

Now we turn to the proof of Theorem 1.4 for initial conditions ζ0\zeta_{0} with a finite number of shocks as in (4.1).

Proof of Theorem 1.4 for ζ0\zeta_{0} as in (4.1).

We let x1,,xN:[0,T]x_{1},\ldots,x_{N}\colon[0,T]\to\mathbb{R} be the generalized characteristics of uu starting at x10,,xN0x^{0}_{1},\ldots,x^{0}_{N}. We let

t=sup{t[0,T]:x1(τ)<<xN(τ)τ[0,t]}>0,\displaystyle t_{*}=\sup\left\{t\in[0,T]\colon x_{1}(\tau)<\cdots<x_{N}(\tau)\;\forall\tau\in[0,t]\right\}>0,

so that the curves xjx_{j} do not intersect on [0,t)[0,t_{*}). If t<Tt_{*}<T, some of the curves intersect at tt_{*}, and for all j{1,,N}j\in\{1,\ldots,N\} such that

xj1(t)<xj(t)=xj+1(t)==x(t)<x+1(t),\displaystyle x_{j-1}(t_{*})<x_{j}(t_{*})=x_{j+1}(t_{*})=\ldots=x_{\ell}(t_{*})<x_{\ell+1}(t_{*}),

for some {j+1,,N}\ell\in\{j+1,\ldots,N\} (with the convention that x0=x_{0}=-\infty and xN+1=+x_{N+1}=+\infty), we modify xj+1,,xx_{j+1},\ldots,x_{\ell} on [t,T][t_{*},T] by setting them all equal to xjx_{j}. In particular after this modification these curves are still generalized characteristics of uu. We repeat that procedure, at most NN times, until we have generalized characteristics x1,,xnx_{1},\ldots,x_{n} starting at x01,,xN0x_{0}^{1},\ldots,x^{0}_{N}, which may intersect, but not cross:

x1x2xN on [0,T].\displaystyle x_{1}\leq x_{2}\leq\ldots\leq x_{N}\quad\text{ on }[0,T].

Then we define u~\tilde{u} on [0,T]×[0,T]\times\mathbb{R} by setting

u~(t,x)=v0(t,x)𝟏x<x1(t)+v1(t,x)𝟏x1(t)<x<x2(t)++vN(t,x)𝟏x>xN(t),\displaystyle\tilde{u}(t,x)=v_{0}(t,x)\mathbf{1}_{x<x_{1}(t)}+v_{1}(t,x)\mathbf{1}_{x_{1}(t)<x<x_{2}(t)}+\cdots+v_{N}(t,x)\mathbf{1}_{x>x_{N}(t)},

where vjv_{j} are the entropy solutions of (1.1) with initial data vj0v^{0}_{j}. Note that some terms of this sum may become zero as tt increases and curves merge.

For any t[0,T]t\in[0,T] and R>0R>0 we set

x0(τ)=RS(tτ),xN+1(τ)=R+S(tτ),\displaystyle x_{0}(\tau)=-R-S(t-\tau),\quad x_{N+1}(\tau)=R+S(t-\tau),
AR,t={(τ,ξ):0<τ<t,x0(τ)<ξ<xN+1(τ)}.\displaystyle A_{R,t}=\left\{(\tau,\xi)\colon 0<\tau<t,\;x_{0}(\tau)<\xi<x_{N+1}(\tau)\right\}.

As τ\tau increases from 0 to tt, some curves x1(τ),,xN(τ)x_{1}(\tau),\cdots,x_{N}(\tau) may merge, or exit the interval [x0(τ),xN+1(τ)][x_{0}(\tau),x_{N+1}(\tau)] (but they can not enter it, as |xj|S|x_{j}^{\prime}|\leq S). We partition [0,T][0,T] into at most NN intervals inside which no merging nor crossing happens. In those subintervals we can apply Proposition 4.2. Concatenating all resulting estimates (4.4), we deduce

RR|u(t2,x)u~(t2,x)|2𝑑xRStR+St|u(0,x)u~(0,x)|2𝑑x\displaystyle\int_{-R}^{R}\left|u(t_{2},x)-\tilde{u}(t_{2},x)\right|^{2}dx-\int_{-R-St}^{R+St}\left|u(0,x)-\tilde{u}(0,x)\right|^{2}dx
Λ(0,t)=CM3α3μ+((0,t)×(RSt,R+St)),\displaystyle\leq\Lambda(0,t)=C\frac{M^{3}}{\alpha^{3}}\mu_{+}((0,t)\times(-R-St,R+St)),

which proves (1.11). Concatenating the estimates (4.5) we obtain a function θ\theta defined on the jump set Ju~J_{\tilde{u}} of u~\tilde{u}, such that for any entropy solution uentu^{ent},

(4.12) RR|u~(t,x)uent(t,x)|𝑑xRStR+St|u~(0,x)uent(0,x)|𝑑x\displaystyle\int_{-R}^{R}\left|\tilde{u}(t,x)-u^{ent}(t,x)\right|\,dx-\int_{-R-St}^{R+St}\left|\tilde{u}(0,x)-u^{ent}(0,x)\right|\,dx
Ju~AR,t|[u~]|12θ|νx|𝑑1,\displaystyle\leq\int_{J_{\tilde{u}}\cap A_{R,t}}|[\tilde{u}]|^{\frac{1}{2}}\theta\,|\nu_{x}|\,d\mathcal{H}^{1},
and αCM2Ju~AR,tθ2|νx|𝑑1RStR+St|u(0,x)u~(0,x)|2𝑑x+Λ(0,t).\displaystyle\frac{\alpha}{CM^{2}}\int_{J_{\tilde{u}}\cap A_{R,t}}\theta^{2}\,|\nu_{x}|\,d\mathcal{H}^{1}\leq\int_{-R-St}^{R+St}\left|u(0,x)-\tilde{u}(0,x)\right|^{2}dx+\Lambda(0,t).

Here we use the notations of Remark 4.4, [u~][\tilde{u}] denotes the jump of u~\tilde{u} along the jump set Ju~J_{\tilde{u}} with normal vector ν=(νt,νx)\nu=(\nu_{t},\nu_{x}).

Then we estimate the right-hand side of (4.12),

Ju~AR,t|[u~]|12θ|νx|𝑑1\displaystyle\int_{J_{\tilde{u}}\cap A_{R,t}}|[\tilde{u}]|^{\frac{1}{2}}\theta\,|\nu_{x}|\,d\mathcal{H}^{1}
(Ju~AR,t|u~||νx|𝑑1)12(Ju~AR,tθ2|νx|𝑑1)12\displaystyle\leq\left(\int_{J_{\tilde{u}}\cap A_{R,t}}|\tilde{u}|\,|\nu_{x}|\,d\mathcal{H}^{1}\right)^{\frac{1}{2}}\left(\int_{J_{\tilde{u}}\cap A_{R,t}}\theta^{2}\,|\nu_{x}|\,d\mathcal{H}^{1}\right)^{\frac{1}{2}}
CMα12(Ju~AR,t|u~||νx|𝑑1)12RStR+St|u(0,x)u~(0,x)|2𝑑x+Λ(0,t).\displaystyle\leq C\frac{M}{\alpha^{\frac{1}{2}}}\left(\int_{J_{\tilde{u}}\cap A_{R,t}}|\tilde{u}|\,|\nu_{x}|\,d\mathcal{H}^{1}\right)^{\frac{1}{2}}\sqrt{\int_{-R-St}^{R+St}\left|u(0,x)-\tilde{u}(0,x)\right|^{2}dx+\Lambda(0,t)}.

Using the nonincreasing property (4.2) of the differences between the functions vjv_{j} we can estimate [u~][\tilde{u}], and obtain, with X0={j:xj0[RSt,R+St]}X_{0}=\{j\colon x^{0}_{j}\in[-R-St,R+St]\},

Ju~AR,t|[u~]||νx|𝑑1\displaystyle\int_{J_{\tilde{u}}\cap A_{R,t}}|[\tilde{u}]|\,|\nu_{x}|\,d\mathcal{H}^{1} 0t[jX0sup[x0(τ),xN+1(τ)](vj1(τ,)vj(τ,))]𝑑τ\displaystyle\leq\int_{0}^{t}\left[\sum_{j\in X_{0}}\sup_{[x_{0}(\tau),x_{N+1}(\tau)]}(v_{j-1}(\tau,\cdot)-v_{j}(\tau,\cdot))\right]\,d\tau
tjX0dj0=t(Dζ0)([RSt,R+St]),\displaystyle\leq t\sum_{j\in X_{0}}d^{0}_{j}=t\cdot(D\zeta_{0})_{-}([-R-St,R+St]),

hence

Ju~AR,t|[u~]|12θ|νx|𝑑1\displaystyle\int_{J_{\tilde{u}}\cap A_{R,t}}|[\tilde{u}]|^{\frac{1}{2}}\theta\,|\nu_{x}|\,d\mathcal{H}^{1} Cα12(Dζ0)([RSt,R+St])t\displaystyle\leq C\alpha^{-\frac{1}{2}}\sqrt{(D\zeta_{0})_{-}([-R-St,R+St])}\sqrt{t}
RStR+St|u(0,x)u~(0,x)|2𝑑x+Λ(0,t).\displaystyle\quad\quad\cdot\sqrt{\int_{-R-St}^{R+St}\left|u(0,x)-\tilde{u}(0,x)\right|^{2}dx+\Lambda(0,t)}.

Combining this with (4.12) and choosing uent=ζu^{ent}=\zeta readily implies (1.12), and concludes the proof of Theorem 1.4 when the initial condition ζ0\zeta_{0} has a finite number of shocks as in (4.1). ∎

Before turning to the proof of Theorem 1.4 for any initial condition ζ0LBVloc()\zeta_{0}\in L^{\infty}\cap BV_{loc}(\mathbb{R}), we gather some estimates on the function u~\tilde{u} that we just constructed:

Lemma 4.5.

When ζ0\zeta_{0} has a finite number of shocks as in (4.1), for any 0s<tT0\leq s<t\leq T and R>0R>0 we have the bounds

(4.13) |Dxu~|([s,t]×[R,R])\displaystyle|D_{x}\tilde{u}|([s,t]\times[-R,R]) 2(ts)(ζ0+(Dζ0)([RSt,R+St])),\displaystyle\leq 2(t-s)\left(\left\|\zeta_{0}\right\|_{\infty}+(D\zeta_{0})_{-}([-R-St,R+St])\right),
(4.14) |Dtu~|([s,t]×[R,R])\displaystyle|D_{t}\tilde{u}|([s,t]\times[-R,R]) S|Dxu~|([s,t]×[R,R]),\displaystyle\leq S|D_{x}\tilde{u}|([s,t]\times[-R,R]),
(4.15) RR|u~(t,x)u~(s,x)|𝑑x\displaystyle\int_{-R}^{R}|\tilde{u}(t,x)-\tilde{u}(s,x)|\,dx 2S(ζ0+(Dζ0)([RSt,R+St]))(ts).\displaystyle\leq 2S\big{(}\left\|\zeta_{0}\right\|_{\infty}+(D\zeta_{0})_{-}([-R-St,R+St])\big{)}(t-s).
Proof of Lemma 4.5.

To obtain (4.13), remark that (Dxu~)(D_{x}\tilde{u})_{-} consists only of shocks which are differences between the vjv_{j}, therefore using the nonincreasing properties (4.2) of such differences and letting X0={j:xj0[RSt,R+St]}X_{0}=\{j\colon x^{0}_{j}\in[-R-St,R+St]\} we have

(Dxu~)([s,t]×[R,R])\displaystyle(D_{x}\tilde{u})_{-}([s,t]\times[-R,R]) =Ju~([s,t]×[R,R])|[u~]||νx|𝑑1\displaystyle=\int_{J_{\tilde{u}}\cap([s,t]\times[-R,R])}|[\tilde{u}]|\,|\nu_{x}|\,d\mathcal{H}^{1}
(ts)jX0dj0=(ts)(Dζ0)([RST,R+ST]).\displaystyle\leq(t-s)\sum_{j\in X_{0}}d^{0}_{j}=(t-s)(D\zeta_{0})_{-}([-R-ST,R+ST]).

This implies (4.13) since |Dxu~|=Dxu~+2(Dxu~)|D_{x}\tilde{u}|=D_{x}\tilde{u}+2(D_{x}\tilde{u})_{-} and

Dxu~([s,t]×[R,R])2(ts)u~2(ts)ζ0.\displaystyle D_{x}\tilde{u}([s,t]\times[-R,R])\leq 2(t-s)\left\|\tilde{u}\right\|_{\infty}\leq 2(t-s)\left\|\zeta_{0}\right\|_{\infty}.

To obtain (4.14) we simply note that outside Ju~J_{\tilde{u}} the function u~\tilde{u} is Lipschitz and satisfies tu~=f(u~)xu~\partial_{t}\tilde{u}=-f^{\prime}(\tilde{u})\partial_{x}\tilde{u}, and for the jump part we take into account that the normal vector satisfies |νt|S|νx||\nu_{t}|\leq S|\nu_{x}|. Finally, using that

RR|u~(t,x)u~(s,x)|𝑑x|Dtu~|([s,t]×[R,R]),\displaystyle\int_{-R}^{R}|\tilde{u}(t,x)-\tilde{u}(s,x)|\,dx\leq|D_{t}\tilde{u}|([s,t]\times[-R,R]),

we obtain (4.15) as a consequence of (4.13)-(4.14). ∎

Remark 4.6.

We also have the Oleinik-type bound Dxu~(t,)1/(αt)D_{x}\tilde{u}(t,\cdot)\leq 1/(\alpha t) for all t(0,T]t\in(0,T], since the functions vjv_{j} satisfy xvj1/(αt)\partial_{x}v_{j}\leq 1/(\alpha t) and all other contributions to Dxu~D_{x}\tilde{u} are negative shocks.

Finally we prove Theorem 1.4 for any initial condition ζ0LBVloc()\zeta_{0}\in L^{\infty}\cap BV_{loc}(\mathbb{R}).

Proof of Theorem 1.4.

We fix ζ0LBVloc()\zeta_{0}\in L^{\infty}\cap BV_{loc}(\mathbb{R}) and approximate it with functions ζε\zeta_{\varepsilon} of the form (4.1) as follows. We have ζ0=ζ1+ζ2\zeta_{0}=\zeta^{1}+\zeta^{2} where ζ1\zeta^{1} is nondecreasing and ζ2\zeta^{2} is nonincreasing, and ζjζ0\|\zeta^{j}\|_{\infty}\leq\left\|\zeta_{0}\right\|_{\infty} for j=1,2j=1,2. We fix a smooth compactly supported nonnegative function ρ\rho with unit integral on \mathbb{R} and define ρε(x)=ε1ρ(ε1x)\rho_{\varepsilon}(x)=\varepsilon^{-1}\rho(\varepsilon^{-1}x). We set

ζε1=ζ1ρε,\displaystyle\zeta^{1}_{\varepsilon}=\zeta^{1}*\rho_{\varepsilon},

so that ζε1\zeta^{1}_{\varepsilon} is Lipschitz nondecreasing and ζε1ζ0\|\zeta^{1}_{\varepsilon}\|_{\infty}\leq\left\|\zeta_{0}\right\|_{\infty}. Taking ε=1/K\varepsilon=1/K for some integer K>0K>0, we let

a=infζ2,b=supζ2,a=a+baK for {0,,K}\displaystyle a=\inf\zeta^{2},\quad b=\sup\zeta^{2},\quad a_{\ell}=a+\ell\frac{b-a}{K}\text{ for }\ell\in\{0,\ldots,K\}

and for an integer K>0K>0 we define

ζε2(x)=a0𝟏ζ2(x)=a0+=1Ka𝟏a1<ζ2(x)a.\displaystyle\zeta^{2}_{\varepsilon}(x)=a_{0}\mathbf{1}_{\zeta^{2}(x)=a_{0}}+\sum_{\ell=1}^{K}a_{\ell}\mathbf{1}_{a_{\ell-1}<\zeta^{2}(x)\leq a_{\ell}}.

Since ζ2\zeta^{2} is nonincreasing, the sets in the above indicator functions are intervals, and we see that

ζε0=ζε1+ζε2,\displaystyle\zeta^{0}_{\varepsilon}=\zeta^{1}_{\varepsilon}+\zeta^{2}_{\varepsilon},

is equal almost everywhere to a function of the form (4.1). Moreover we have

ζε0\displaystyle\zeta^{0}_{\varepsilon} ζ0in Lloc2()\displaystyle\longrightarrow\zeta_{0}\quad\text{in }L^{2}_{loc}(\mathbb{R})
(Dζε0)([R,R])\displaystyle(D\zeta^{0}_{\varepsilon})_{-}([-R,R]) (Dζ0)([R,R])for all R>0,\displaystyle\longrightarrow(D\zeta_{0})_{-}([-R,R])\quad\text{for all }R>0,

as ε0\varepsilon\to 0. Applying Theorem 1.4 to the initial condition ζε0\zeta^{0}_{\varepsilon} we obtain functions u~ε\tilde{u}_{\varepsilon} satisfying (1.11)-(1.12). Thanks to Lemma 4.5 and u~εζ0\|\tilde{u}_{\varepsilon}\|_{\infty}\leq\|\zeta_{0}\|_{\infty} we may extract a subsequence of u~ε\tilde{u}_{\varepsilon} such that u~ε(t,)\tilde{u}_{\varepsilon}(t,\cdot) converges in Lloc2L^{2}_{loc} for every t[0,T]t\in[0,T], and pass to the limit in (1.11)-(1.12). The limit u~\tilde{u} satisfies the bounds of Lemma 4.5. ∎

References

  • [1] Adimurthi, Ghoshal, S. S., and Veerappa Gowda, G. D. LpL^{p} stability for entropy solutions of scalar conservation laws with strict convex flux. J. Differ. Equations 256, 10 (2014), 3395–3416.
  • [2] Ambrosio, L., Fusco, N., and Pallara, D. Functions of bounded variation and free discontinuity problems. Oxford: Clarendon Press, 2000.
  • [3] Bellettini, G., Bertini, L., Mariani, M., and Novaga, M. Γ\Gamma-entropy cost for scalar conservation laws. Arch. Ration. Mech. Anal. 195, 1 (2010), 261–309.
  • [4] Chen, G., Krupa, S. G., and Vasseur, A. F. Uniqueness and weak-BV stability for 2×22\times 2 conservation laws. arXiv:2010.04761.
  • [5] Dafermos, C. Hyperbolic conservation laws in continuum physics, second ed., vol. 325 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 2005.
  • [6] Dafermos, C. M. The second law of thermodynamics and stability. Arch. Rational Mech. Anal. 70, 2 (1979), 167–179.
  • [7] De Lellis, C., and Ignat, R. A regularizing property of the 2D2D-eikonal equation. Comm. Partial Differential Equations 40, 8 (2015), 1543–1557.
  • [8] De Lellis, C., Otto, F., and Westdickenberg, M. Structure of entropy solutions for multi-dimensional scalar conservation laws. Arch. Ration. Mech. Anal. 170, 2 (2003), 137–184.
  • [9] De Lellis, C., and Westdickenberg, M. On the optimality of velocity averaging lemmas. Ann. Inst. H. Poincaré Anal. Non Linéaire 20, 6 (2003), 1075–1085.
  • [10] DiPerna, R. J. Uniqueness of solutions to hyperbolic conservation laws. Indiana Univ. Math. J. 28, 1 (1979), 137–188.
  • [11] Golse, F., and Perthame, B. Optimal regularizing effect for scalar conservation laws. Rev. Mat. Iberoam. 29, 4 (2013), 1477–1504.
  • [12] Kipnis, C., and Landim, C. Scaling limits of interacting particle systems, vol. 320 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999.
  • [13] Krupa, S. G., and Vasseur, A. F. On uniqueness of solutions to conservation laws verifying a single entropy condition. J. Hyperbolic Differ. Equ. 16, 1 (2019), 157–191.
  • [14] Krupa, S. G., and Vasseur, A. F. Stability and uniqueness for piecewise smooth solutions to a nonlocal scalar conservation law with applications to Burgers-Hilbert equation. SIAM J. Math. Anal. 52, 3 (2020), 2491–2530.
  • [15] Kružkov, S. N. First order quasilinear equations with several independent variables. Mat. Sb. (N.S.) 81 (123) (1970), 228–255.
  • [16] Lamy, X., and Otto, F. On the regularity of weak solutions to Burgers’ equation with finite entropy production. Calc. Var. Partial Differential Equations 57, 4 (2018), Paper No. 94, 19.
  • [17] Lecumberry, M. Geometric structure of micromagnetic walls and shock waves in scalar conservation laws. PhD thesis, Université de Nantes, 2004.
  • [18] Leger, N. L2L^{2} stability estimates for shock solutions of scalar conservation laws using the relative entropy method. Arch. Ration. Mech. Anal. 199, 3 (2011), 761–778.
  • [19] Leger, N., and Vasseur, A. Relative entropy and the stability of shocks and contact discontinuities for systems of conservation laws with non-BV perturbations. Arch. Ration. Mech. Anal. 201, 1 (2011), 271–302.
  • [20] Marconi, E. On the structure of weak solutions to scalar conservation laws with finite entropy production. arXiv:1909.07257.
  • [21] Marconi, E. The rectifiability of the entropy defect measure for Burgers equation. arXiv:2004.09932.
  • [22] Mariani, M. Large deviations principles for stochastic scalar conservation laws. Probab. Theory Related Fields 147, 3-4 (2010), 607–648.
  • [23] Poliakovsky, A. On a variational approach to the method of vanishing viscosity for conservation laws. Adv. Math. Sci. Appl. 18, 2 (2008), 429–451.
  • [24] Serre, D., and Vasseur, A. F. L2L^{2}-type contraction for systems of conservation laws. J. Éc. polytech. Math. 1 (2014), 1–28.
  • [25] Serre, D., and Vasseur, A. F. About the relative entropy method for hyperbolic systems of conservation laws. In A panorama of mathematics: pure and applied, vol. 658 of Contemp. Math. Amer. Math. Soc., Providence, RI, 2016, pp. 237–248.
  • [26] Varadhan, S. Large deviations for the asymmetric simple exclusion process. In Stochastic analysis on large scale interacting systems, vol. 39 of Adv. Stud. Pure Math. Math. Soc. Japan, Tokyo, 2004, pp. 1–27.
  • [27] Vasseur, A. Strong traces for solutions of multidimensional scalar conservation laws. Arch. Ration. Mech. Anal. 160, 3 (2001), 181–193.
  • [28] Vasseur, A. F. Recent results on hydrodynamic limits. In Handbook of differential equations: evolutionary equations. Vol. IV, Handb. Differ. Equ. Elsevier/North-Holland, Amsterdam, 2008, pp. 323–376.
  • [29] Vasseur, A. F. Relative entropy and contraction for extremal shocks of conservation laws up to a shift. In Recent advances in partial differential equations and applications, vol. 666 of Contemp. Math. Amer. Math. Soc., Providence, RI, 2016, pp. 385–404.