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

Regularity Structure of conservative solutions to the Hunter-Saxton equation

Yu Gao Department of Applied Mathematics, The Hong Kong Polytechnic University, Hung Hom, Kowloon, Hong Kong mathyu.gao@polyu.edu.hk Hao Liu Department of Mathematics, The University of Hong Kong, Pokfulam, Hong Kong u3005509@connect.hku.hk  and  Tak Kwong Wong Department of Mathematics, The University of Hong Kong, Pokfulam, Hong Kong. takkwong@maths.hku.hk
Abstract.

In this paper we characterize the regularity structure, as well as show the global-in-time existence and uniqueness, of (energy) conservative solutions to the Hunter-Saxton equation by using the method of characteristics. The major difference between the current work and previous results is that we are able to characterize the singularities of energy measure and their nature in a very precise manner. In particular, we show that singularities, whose temporal and spatial locations are also explicitly given in this work, may only appear at at most countably many times, and are completely determined by the absolutely continuous part of initial energy measure. Our mathematical analysis is based on using the method of characteristics in a generalized framework that consists of the evolutions of solution to the Hunter-Saxton equation and the energy measure. This method also provides a clear description of the semi-group property for the solution and energy measure for all times.

Key words and phrases:
Keywrods: formulation of singularity, well-posedness, integrable system, decomposition of energy measure, semi-group property

1. Introduction

In this paper, we study the regularity structure of (energy) conservative solutions to the Hunter-Saxton equation on the whole real line \mathbb{R}: for xx, tt\in\mathbb{R},

(1.1) ut+uux=12xuy2(y,t)dy.\displaystyle u_{t}+uu_{x}=\frac{1}{2}\int_{-\infty}^{x}u_{y}^{2}(y,t)\mathop{}\!\mathrm{d}y.

Here, utu_{t} and uxu_{x} represent the time and spatial derivative of uu, respectively. The equation (1.1) is an integrable equation that was first proposed by Hunter and Saxton [13] to study a nonlinear instability in the director field of a nematic liquid crystal. When uu is a classical solution to (1.1), differentiating (1.1) with respect to the spatial variable xx yields

(1.2) (ut+uux)x=12ux2,\displaystyle(u_{t}+uu_{x})_{x}=\frac{1}{2}u_{x}^{2},

and hence, the following energy conservation law holds:

(1.3) (ux2)t+(uux2)x=0.\displaystyle(u_{x}^{2})_{t}+(uu_{x}^{2})_{x}=0.

According to equation (1.3), it is nature to find the global-in-time conservative solution uu that satisfies ux(,t)L2=u¯xL2\|u_{x}(\cdot,t)\|_{L^{2}}=\|\bar{u}_{x}\|_{L^{2}} for any tt\in\mathbb{R}, provided that the initial datum u|t=0=u¯u|_{t=0}=\bar{u} satisfies u¯xL2()\bar{u}_{x}\in L^{2}(\mathbb{R}). However, similar to the inviscid Burgers equation, using the characteristics method, one can verify the following fact: if the initial datum u¯\bar{u} satisfies infxu¯x(x)<0\inf_{x\in\mathbb{R}}\bar{u}_{x}(x)<0, then we have ux(,t)u_{x}(\cdot,t)\to-\infty as tT0:=2/infxu¯x(x).t\to T_{0}:=-2/\inf_{x\in\mathbb{R}}{\bar{u}_{x}(x)}. At the blow-up time, the function uu will lose the information of energy conservation temporarily. This can be seen from the following simple example. Consider the solution with initial datum (see also in [15, 7])

u¯(x)={0,x0,x,0<x<1,1,x1.\bar{u}(x)=\left\{\begin{aligned} &0,\quad x\leq 0,\\ &-x,\quad 0<x<1,\\ &-1,\quad x\geq 1.\end{aligned}\right.

The conservative solution uu is explicitly given by

u(x,t)={0,x0,2x2t,0<x(1t2)2,1+t2,x(1t2)2,ux(x,t)={22t,0<x(1t2)2,0,otherwise.u(x,t)=\left\{\begin{aligned} &0,\quad x\leq 0,\\ &-\frac{2x}{2-t},\quad 0<x\leq\left(1-\frac{t}{2}\right)^{2},\\ &-1+\frac{t}{2},\quad x\geq\left(1-\frac{t}{2}\right)^{2},\end{aligned}\right.\quad u_{x}(x,t)=\left\{\begin{aligned} &-\frac{2}{2-t},\quad 0<x\leq\left(1-\frac{t}{2}\right)^{2},\\ &0,\quad~{}\textrm{otherwise.}\end{aligned}\right.

We have u(x,2)0u(x,2)\equiv 0 and ux2(,t)δu_{x}^{2}(\cdot,t)\to\delta in the sense of distributions as t2t\to 2, where δ\delta is the Dirac delta mass at the origin x=0x=0. Notice that ux(,t)L2=1\|u_{x}(\cdot,t)\|_{L^{2}}=1 for any t2t\neq 2 and ux(,2)L2=0\|u_{x}(\cdot,2)\|_{L^{2}}=0. Therefore, if we only study the solution uu, the energy ux(,t)L2\|u_{x}(\cdot,t)\|_{L^{2}} is not conserved only at t=2t=2; this viewpoint, which causes a temporary/momentary disappearance of energy, is very non-physical. It is worth noting that at the blow-up time t=2t=2, the energy density is converted into the singular measure δ\delta, and ux2(,2)0u_{x}^{2}(\cdot,2)\equiv 0 corresponds to the absolutely continuous part of the measure δ\delta with respect to the Lebesgue measure \mathcal{L}. To describe the energy conservation of weak solutions, we use a nonnegative Radon measure μ(t)\mu(t), which can be seen as the energy measure, to replace ux2(,t)u_{x}^{2}(\cdot,t) in equations (1.1) and (1.3), and study the Hunter-Saxton equation in the following generalized framework:

Generalized Framework:

(1.4) ut+uux=12xdμ(t),\displaystyle u_{t}+uu_{x}=\frac{1}{2}\int_{-\infty}^{x}\mathop{}\!\mathrm{d}\mu(t),
(1.5) μt+(uμ)x=0,\displaystyle\mu_{t}+(u\mu)_{x}=0,
(1.6) dμac(t)=ux2(,t)dx.\displaystyle\mathop{}\!\mathrm{d}\mu_{ac}(t)=u_{x}^{2}(\cdot,t)\mathop{}\!\mathrm{d}x.

Let us comment on the notations in (1.4)-(1.6) as follows. In (1.6), the measure μac(t)\mu_{ac}(t) stands for the absolutely continuous part of the energy measure μ(t)\mu(t) with respect to the Lebesgue measure. In (1.4), the integral xdμ(t)\int_{-\infty}^{x}\mathop{}\!\mathrm{d}\mu(t) is not well defined when μ(t)\mu(t) contains pure point measures. However, we are going to show that there are only at most countably many time tt such that the energy measure μ(t)\mu(t) has pure point measures (see Theorem 2.1 (iii) for details), and hence, the ambiguity of this integral will not affect the definition of weak (energy conservative) solutions below. Therefore, we will keep this integral notation in (1.4). Equation (1.5) is viewed in the sense of distributions. We then study the conservative solutions to the generalized framework (1.4)-(1.6). First of all, let us define the following space for conservative solutions:

Definition 1.1.

Let 𝒟\mathcal{D} be the set of pairs (u,μ)(u,\mu) satisfying

  1. (i)

    uCb()u\in C_{b}(\mathbb{R}), uxL2()u_{x}\in L^{2}(\mathbb{R});

  2. (ii)

    μ+()\mu\in\mathcal{M}_{+}(\mathbb{R});

  3. (iii)

    dμac=ux2dx\mathop{}\!\mathrm{d}\mu_{ac}=u_{x}^{2}\mathop{}\!\mathrm{d}x, where μac\mu_{ac} is the absolutely continuous part of measure μ\mu with respect to the Lebesgue measure \mathcal{L}.

Here, +()\mathcal{M}_{+}(\mathbb{R}) stands for the set of finite positive Radon measure endowed with the weak topology.

Following the physical meaning of the energy measure μ(t)\mu(t), we require μ(t)+()\mu(t)\in\mathcal{M}_{+}(\mathbb{R}), so that the total energy is finite; however, technically speaking, our mathematical analysis in this paper only requires that the positive measure μ(t)((,x])<\mu(t)((-\infty,x])<\infty for all xx\in\mathbb{R}. There are several different definitions for conservative solutions to the Hunter-Saxton equation; see [6, Definition 4.4] for instance. In this paper, we will use the following definition of conservative solutions to the Hunter-Saxton equation:

Definition 1.2 (Conservative solutions).

Let (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D} be a given initial datum. The pair (u(t),μ(t))(u(t),\mu(t)) is said to be a global-in-time conservative solution to the generalized framework (1.4)-(1.6) with the initial datum (u¯,μ¯)(\bar{u},\bar{\mu}), if the pair (u(t),μ(t))(u(t),\mu(t)) satisfies the following:

  1. (i)

    uC(;Cb())Cloc1/2(×)u\in C(\mathbb{R};C_{b}(\mathbb{R}))\cap C^{1/2}_{loc}(\mathbb{R}\times\mathbb{R}), utLloc2(×)u_{t}\in L_{loc}^{2}(\mathbb{R}\times\mathbb{R}), ux(,t)L2()u_{x}(\cdot,t)\in L^{2}(\mathbb{R}) for all tt\in\mathbb{R}, and μC(;+())\mu\in C(\mathbb{R};\mathcal{M}_{+}(\mathbb{R}));

  2. (ii)

    (u(,0),μ(0))=(u¯,μ¯)(u(\cdot,0),\mu(0))=(\bar{u},\bar{\mu}), and dμ(t)=ux2(x,t)dx\mathop{}\!\mathrm{d}\mu(t)={u}_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x for a.e. tt\in\mathbb{R};

  3. (iii)

    the equation

    (1.7) uϕtϕ(uux12F)dxdt=0\displaystyle\int_{\mathbb{R}}\int_{\mathbb{R}}u\phi_{t}-\phi\left(uu_{x}-\frac{1}{2}F\right)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=0

    holds for all ϕCc(×)\phi\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{R}); the function F(x,t)F(x,t) is defined by F(x,t):=xdμ(t)F(x,t):=\int_{-\infty}^{x}\mathop{}\!\mathrm{d}\mu(t);

  4. (iv)

    conservation of energy:

    (1.8) (ϕt+uϕx)dμ(t)dt=0\int_{\mathbb{R}}\int_{\mathbb{R}}\left(\phi_{t}+u\phi_{x}\right)\mathop{}\!\mathrm{d}\mu(t)\mathop{}\!\mathrm{d}t=0

    for every ϕCc(×)\phi\in C^{\infty}_{c}(\mathbb{R}\times\mathbb{R});

  5. (v)

    equation (1.6) holds for all tt\in\mathbb{R}.

The last condition is equivalent to say that (u(t),μ(t))𝒟(u(t),\mu(t))\in\mathcal{D} for all tt\in\mathbb{R}.

It follows from the regularity requirement (i) in Definition 1.2 that the condition (iii) in Definition 1.2 (namely Identity (1.7) holds for all test functions) is equivalent to the following condition:

  1. (iii)’

    Equation (1.4) holds in the Lloc2(×)L^{2}_{loc}(\mathbb{R}\times\mathbb{R}) sense.

The condition (iii)’ is equivalent to [7, Definition 1.1(ii)], when the flux is given by u2/2u^{2}/2.

Remark 1.1 (Energy conservation).

Indeed, any weak solution (u,μ)(u,\mu) to the Hunter-Saxton equation under the generalized framework (1.4)-(1.6) in the sense of Definition 1.2 satisfies the energy conservation μ(t)()=μ¯()\mu(t)(\mathbb{R})=\bar{\mu}(\mathbb{R}) for tt\in\mathbb{R}. This is why we call it a conservative solution. To verify this energy conservation, we now consider any arbitrary time t>0t>0 (the case for t<0t<0 will be similar). We choose non-negative smooth functions χϵ(s)\chi_{\epsilon}(s) and gR(x)g_{R}(x) for ϵ>0\epsilon>0 and R>0R>0, where χϵ(s)=0\chi_{\epsilon}(s)=0 for sϵs\leq-\epsilon and st+ϵs\geq t+\epsilon. χϵ(s)=1\chi_{\epsilon}(s)=1 for s(0,t)s\in(0,t), χϵ(s)0\chi_{\epsilon}^{\prime}(s)\geq 0 for s(ϵ,0)s\in(-\epsilon,0) and χϵ(s)0\chi_{\epsilon}^{\prime}(s)\leq 0 for s(t,t+ϵ)s\in(t,t+\epsilon). The function gRg_{R} satisfies gR(x)=1g_{R}(x)=1 for |x|R|x|\leq R, gR(x)=0g_{R}(x)=0 for |x|2R|x|\geq 2R and |gR(x)|2R\left|g^{\prime}_{R}(x)\right|\leq\frac{2}{R}. Finally, let ϕ(x,t)=χϵ(t)gR(x)\phi(x,t)=\chi_{\epsilon}(t)g_{R}(x), and substituting this ϕ\phi into equation (1.8), we obtain

χϵ(s)gR(x)dμ(s)ds+u(x,s)χϵ(s)gR(x)dμ(s)ds=0.\int_{\mathbb{R}}\int_{\mathbb{R}}\chi^{\prime}_{\epsilon}(s)g_{R}(x)\mathop{}\!\mathrm{d}\mu(s)\mathop{}\!\mathrm{d}s+\int_{\mathbb{R}}\int_{\mathbb{R}}u(x,s)\chi_{\epsilon}(s)g^{\prime}_{R}(x)\mathop{}\!\mathrm{d}\mu(s)\mathop{}\!\mathrm{d}s=0.

Passing to the limit as RR\to\infty, and using uu is bounded, we have

ϵ0χϵ(s)μ(s)()ds+tt+ϵχϵ(s)μ(s)()ds=0.\int_{-\epsilon}^{0}\chi^{\prime}_{\epsilon}(s)\mu(s)(\mathbb{R})\mathop{}\!\mathrm{d}s+\int_{t}^{t+\epsilon}\chi^{\prime}_{\epsilon}(s)\mu(s)(\mathbb{R})\mathop{}\!\mathrm{d}s=0.

Since μC(;+())\mu\in C(\mathbb{R};\mathcal{M}_{+}(\mathbb{R})),

|ϵ0χϵ(s)μ(s)()dsμ¯()|=\displaystyle\left|\int_{-\epsilon}^{0}\chi^{\prime}_{\epsilon}(s)\mu(s)(\mathbb{R})\mathop{}\!\mathrm{d}s-\bar{\mu}(\mathbb{R})\right|= |ϵ0χϵ(s)μ(s)()dsϵ0χϵ(s)μ¯()ds|\displaystyle\left|\int_{-\epsilon}^{0}\chi^{\prime}_{\epsilon}(s)\mu(s)(\mathbb{R})\mathop{}\!\mathrm{d}s-\int_{-\epsilon}^{0}\chi^{\prime}_{\epsilon}(s)\bar{\mu}(\mathbb{R})\mathop{}\!\mathrm{d}s\right|
\displaystyle\leq ϵ0χϵ(s)|μ(s)()μ¯()|ds0asϵ0+.\displaystyle\int_{-\epsilon}^{0}\chi^{\prime}_{\epsilon}(s)\left|\mu(s)(\mathbb{R})-\bar{\mu}(\mathbb{R})\right|\mathop{}\!\mathrm{d}s\to 0~{}\textrm{as}~{}\epsilon\to 0^{+}.

This shows that ϵ0χϵ(s)μ(s)()dsμ¯()\int_{-\epsilon}^{0}\chi^{\prime}_{\epsilon}(s)\mu(s)(\mathbb{R})\mathop{}\!\mathrm{d}s\to\bar{\mu}(\mathbb{R}) as ϵ0+\epsilon\to 0^{+}. On the other hand, one can also verify that tt+ϵχϵ(s)μ(s)()ds\int_{t}^{t+\epsilon}\chi^{\prime}_{\epsilon}(s)\mu(s)(\mathbb{R})\mathop{}\!\mathrm{d}s \to μ(t)()-\mu(t)(\mathbb{R}) as ϵ0+\epsilon\to 0^{+} in a similar manner, and hence, we finally obtain μ(t)()=μ¯()\mu(t)(\mathbb{R})=\bar{\mu}(\mathbb{R}).

Remark 1.2.

It follows from Remark 1.1 and Equation (1.6) that we also have uxL(;L2())u_{x}\in L^{\infty}(\mathbb{R};L^{2}(\mathbb{R})).

Remark 1.3.

It follows μC(;+())\mu\in C(\mathbb{R};\mathcal{M}_{+}(\mathbb{R})) and (ii) in the Definition 1.2 that F(,t)F(\cdot,t) is well defined for a.e. tt\in\mathbb{R}, and FLloc(×)F\in L^{\infty}_{loc}(\mathbb{R}\times\mathbb{R}). Moreover, according to Remark 1.2, we also know FL(×)F\in L^{\infty}(\mathbb{R}\times\mathbb{R}).

Under the above definition, we have the following results:

Theorem 1.1.

Let (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D} be given. Then there exists a unique global-in-time conservative solution (in the sense of Definition 1.2) (u(t),μ(t))(u(t),\mu(t)) to the generalized framework (1.4)-(1.6) with (u¯,μ¯)(\bar{u},\bar{\mu}) as its initial datum. Let μ(t)=μac(t)+μpp(t)+μsc(t)\mu(t)=\mu_{ac}(t)+\mu_{pp}(t)+\mu_{sc}(t), where μac(t)\mu_{ac}(t), μpp(t)\mu_{pp}(t), and μsc(t)\mu_{sc}(t) are the absolutely continuous part, pure point part and the singular continuous part of μ\mu respectively. The following statements also hold:

  1. (i)

    (Energy conservation) we have μC(;+())\mu\in C(\mathbb{R};\mathcal{M}_{+}(\mathbb{R})) and

    (1.9) μ(t)()=μ¯(),for all t.\displaystyle\mu(t)(\mathbb{R})=\bar{\mu}(\mathbb{R}),\quad\mbox{for all }~{}t\in\mathbb{R}.
  2. (ii)

    (Formation of singularity, and its temporal and spatial locations) For any t0t\neq 0, μpp(t)\mu_{pp}(t) and μsc(t)\mu_{sc}(t) are determined by the absolutely continuous part μ¯ac\bar{\mu}_{ac}, namely determined by u¯x\bar{u}_{x}. More precisely, for any t0t\neq 0, we define

    (1.10) AtE={x:u¯x(x)=2t}.\displaystyle A_{t}^{E}=\left\{x:~{}~{}\bar{u}_{x}(x)=-\frac{2}{t}\right\}.

    If (AtE)0\mathcal{L}(A_{t}^{E})\neq 0, then μpp(t)+μsc(t)0\mu_{pp}(t)+\mu_{sc}(t)\neq 0 (i.e., μpp(t)+μsc(t)\mu_{pp}(t)+\mu_{sc}(t) is not a zero measure) and

    (1.11) supp(μpp(t)+μsc(t)){x+u¯(x)t+t24μ¯((,x)):xAtE}.\displaystyle\mathrm{supp}\Big{(}\mu_{pp}(t)+\mu_{sc}(t)\Big{)}\subset\left\{x+\bar{u}(x)t+\frac{t^{2}}{4}\bar{\mu}((-\infty,x)):~{}x\in A_{t}^{E}\right\}.

    All the intervals with positive length in AtEA_{t}^{E} will generate the pure point part μpp(t)\mu_{pp}(t), and the rest of AtEA_{t}^{E} will generate the singular continuous part μsc(t)\mu_{sc}(t).

  3. (iii)

    (Countably many singular times) There are at most countably many times tt\in\mathbb{R}, such that either the pure point part or the singular continuous part of μ(t)\mu(t) is not identically equal to zero, namely both the sets

    (1.12) Tp:={t:μpp(t)0}andTs:={t:μsc(t)0}\displaystyle T_{p}:=\{t:~{}\mu_{pp}(t)\neq 0\}\quad\mbox{and}\quad T_{s}:=\{t:~{}\mu_{sc}(t)\neq 0\}

    are countable.

  4. (iv)

    (Regularity) For all time tt\in\mathbb{R}, the function u(x,t)u(x,t) is globally absolutely continuous with respect to the spatial variable xx and

    (1.13) dμac(t)=ux2(x,t)dx.\displaystyle\mathop{}\!\mathrm{d}\mu_{ac}(t)=u_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x.

    Furthermore,

    (1.14) uC(;Cb())Cloc1/2(×),uxL(;L2()),andutLloc2(×).\displaystyle u\in C(\mathbb{R};C_{b}(\mathbb{R}))\cap C^{1/2}_{loc}(\mathbb{R}\times\mathbb{R}),~{}~{}u_{x}\in L^{\infty}(\mathbb{R};L^{2}(\mathbb{R})),~{}~{}\mbox{and}~{}~{}u_{t}\in L_{loc}^{2}(\mathbb{R}\times\mathbb{R}).
  5. (v)

    (Asymptotic behaviour) If u¯():=limxu¯(x)\bar{u}(-\infty):=\lim_{x\to-\infty}\bar{u}(x) exists, then we have

    (1.15) limxu(x,t)=u¯().\displaystyle\lim_{x\to-\infty}u(x,t)=\bar{u}(-\infty).

    On the other hand, if u¯(+):=limx+u¯(x)\bar{u}(+\infty):=\lim_{x\to+\infty}\bar{u}(x) exists, then we also have

    (1.16) limx+u(x,t)=u¯(+)+12μ¯()t.\displaystyle\lim_{x\to+\infty}u(x,t)=\bar{u}(+\infty)+\frac{1}{2}\bar{\mu}(\mathbb{R})t.

Let us start with the absolutely continuous initial data (as described below) to illustrate our main strategies to prove the above results. Consider an initial datum (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D}. When the initial energy measure μ¯\bar{\mu} is absolutely continuous with respect to Lebesgue measure \mathcal{L} (i.e., μ¯\bar{\mu}\ll\mathcal{L} and dμ¯=u¯x2dx\mathop{}\!\mathrm{d}\bar{\mu}=\bar{u}_{x}^{2}\mathop{}\!\mathrm{d}x), we can apply the usual Lagrangian coordinates (ξ,t)(\xi,t) to define the flow map via tX(ξ,t)=u(X(ξ,t),t)\partial_{t}X(\xi,t)=u(X(\xi,t),t) and X(ξ,0)=ξX(\xi,0)=\xi, and obtain the following global characteristics (see also (2.26) below):

(1.17) X(ξ,t)=ξ+u¯(ξ)t+t24(,ξ)u¯x2(y)dy=ξ+u¯(ξ)t+t24μ¯((,ξ)).\displaystyle X(\xi,t)=\xi+\bar{u}(\xi)t+\frac{t^{2}}{4}\int_{(-\infty,\xi)}\bar{u}_{x}^{2}(y)\mathop{}\!\mathrm{d}y=\xi+\bar{u}(\xi)t+\frac{t^{2}}{4}\bar{\mu}((-\infty,\xi)).

Since Xξ(ξ,t)=[1+t2u¯x(ξ)]20X_{\xi}(\xi,t)=[1+\frac{t}{2}\bar{u}_{x}(\xi)]^{2}\geq 0, X(,t)X(\cdot,t) is an increasing function for any fixed tt. The solution (u,μ)(u,\mu) can be recovered by u(X(ξ,t),t)=tX(ξ,t)u(X(\xi,t),t)=\partial_{t}X(\xi,t) and μ(t)=X(,t)#(u¯x2dx)\mu(t)=X(\cdot,t)\#(\bar{u}_{x}^{2}\mathop{}\!\mathrm{d}x). Then, we can apply an elementary result for push-forward measures (see Lemma A.2 for instance) to decompose the measure μ(t)\mu(t) into its absolutely continuous part, pure point part, and singular continuous part by using the derivative Xξ(,t)X_{\xi}(\cdot,t). More precisely, the singular part of μ(t)\mu(t) is determined by {ξ:Xξ(ξ,t)=0}\{\xi:~{}~{}X_{\xi}(\xi,t)=0\}, or equivalently {x:u¯x(x)=2t}\{x:~{}~{}\bar{u}_{x}(x)=-\frac{2}{t}\}; in particular, the intervals in {x:u¯x(x)=2t}\{x:~{}~{}\bar{u}_{x}(x)=-\frac{2}{t}\} generate the pure point part of μ(t)\mu(t), and the rest points in {x:u¯x(x)=2t}\{x:~{}~{}\bar{u}_{x}(x)=-\frac{2}{t}\} corresponds to the singular continuous part of μ(t)\mu(t). The absolutely continuous part of μ(t)\mu(t) is given by {ξ:Xξ(ξ,t)>0}\{\xi:~{}~{}X_{\xi}(\xi,t)>0\}.

The Lagrangian coordinates (ξ,t)(\xi,t) are applicable to all absolutely continuous initial data (i.e. the initial energy measure μ¯\bar{\mu} is absolutely continuous with respect to Lebesgue measure). For such an initial datum (u¯,μ¯)(\bar{u},\bar{\mu}), since there is no point mass at any particular point ξ\xi, the cumulative energy distribution in (1.17) satisfies

(1.18) μ¯((,ξ))=μ¯((,ξ])=(,ξ)u¯x2(y)dy.\displaystyle\bar{\mu}((-\infty,\xi))=\bar{\mu}((-\infty,\xi])=\int_{(-\infty,\xi)}\bar{u}_{x}^{2}(y)\mathop{}\!\mathrm{d}y.

However, if the initial energy measure μ¯\bar{\mu} contains any pure point part, then the above relation is not true and it is impossible to obtain global flow map XX in the Lagrangian coordinates (ξ,t)(\xi,t). To overcome the difficulty brought by the singular parts of μ¯\bar{\mu}, we will apply another set of variables used in [6, 8, 3]. We “flatten” the singular part of μ¯\bar{\mu} with the help of an increasing Lipschitz function x¯:\bar{x}:\mathbb{R}\to\mathbb{R} defined by

x¯(α)+μ¯((,x¯(α)))αx¯(α)+μ¯((,x¯(α)]),α.\bar{x}(\alpha)+\bar{\mu}((-\infty,\bar{x}(\alpha)))\leq\alpha\leq\bar{x}(\alpha)+\bar{\mu}((-\infty,\bar{x}(\alpha)]),\quad\alpha\in\mathbb{R}.

Then, the function f(α):=1x¯(α)0f(\alpha):=1-\bar{x}^{\prime}(\alpha)\geq 0 will play the role of the energy density in the α\alpha-variable; see Remark 2.2 (i) for the explanation. We actually have μ¯=x¯#(fdα)\bar{\mu}=\bar{x}\#(f\mathop{}\!\mathrm{d}\alpha) in this case; see Proposition 2.1 for further details. Since x¯\bar{x} is increasing and fL1()f\in L^{1}(\mathbb{R}), the absolutely continuous part of the push-forward measure μ¯=x¯#(fdα)\bar{\mu}=\bar{x}\#(f\mathop{}\!\mathrm{d}\alpha) corresponds to B0L={α:x¯(α)>0}B_{0}^{L}=\{\alpha:~{}~{}\bar{x}^{\prime}(\alpha)>0\}, and the singular part of μ¯\bar{\mu} is determined by A0L={α:x¯(α)=0}A_{0}^{L}=\{\alpha:~{}~{}\bar{x}^{\prime}(\alpha)=0\}; see Lemma A.2 for instance. Moreover, similar to X(ξ,t)X(\xi,t), we can also define the global characteristics y(α,t)y(\alpha,t) in the α\alpha-variable depending only on the initial datum with an explicit form (2.12), and y(α,t)y(\alpha,t) is increasing with respect to α\alpha at any particular time tt\in\mathbb{R}. Then, the global conservative solution (u,μ)(u,\mu) to the generalized framework (1.4)-(1.6) will be recovered via u(y(α,t),t)=ty(α,t)u(y(\alpha,t),t)=\partial_{t}y(\alpha,t) and μ(t)=y(,t)#(fdα)\mu(t)=y(\cdot,t)\#(f\mathop{}\!\mathrm{d}\alpha). We can also obtain the following important property (see Proposition 2.2):

yα(α,t)={x¯(α)[1+t2u¯x(x¯(α))]2,αB0L,t24,αA0L.y_{\alpha}(\alpha,t)=\left\{\begin{aligned} &\bar{x}^{\prime}(\alpha)\left[1+\frac{t}{2}\bar{u}_{x}(\bar{x}(\alpha))\right]^{2},\quad\alpha\in B_{0}^{L},\\ &\frac{t^{2}}{4},\quad\alpha\in A_{0}^{L}.\end{aligned}\right.

The singular part of μ(t)\mu(t) comes from the set {α:yα(α,t)=0}\{\alpha:~{}~{}y_{\alpha}(\alpha,t)=0\} in the α\alpha-coordinate, or equivalently, the set {x:u¯x(x)=2t}\{x:~{}~{}\bar{u}_{x}(x)=-\frac{2}{t}\} in the Eulerian coordinates. Since A0LA_{0}^{L} corresponds to the singular part of the initial datum μ¯\bar{\mu}, the above formula implies that it will never create the singular part of μ(t)\mu(t) for any t0t\neq 0. All the properties in Theorem 1.1 and the structure of conservative solutions are obtained with the help of the characteristics y(α,t)y(\alpha,t). In a nutshell, the existence and regularity structure of conservative solutions can be shown by using the method of characteristics. The uniqueness of conservative solutions follows from the existence and uniqueness of characteristics for conservative solutions in the sense of Definition 1.2 (see Theorem 3.2). We also note that the singular continuous part of μ(t)\mu(t) is inevitable even u¯\bar{u} is a smooth function; see Example A.1 for instance.

For the flow map X(ξ,t)X(\xi,t), we obviously have the semi-group property X(,t)=X(,ts)X(,s)X(\cdot,t)=X(\cdot,t-s)\circ X(\cdot,s). A similar relation between the above two types of characteristics X(ξ,t)X(\xi,t) and y(α,t)y(\alpha,t) still holds. Indeed, we have the following theorem:

Theorem 1.2 (Semi-group property).

Let (u,μ)(u,\mu) be a solution to the generalized framework (1.4)-(1.6) given by Theorem 1.1 with initial datum (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D}. Consider a time ss\in\mathbb{R} such that μ(s)\mu(s) is absolutely continuous with respect to the Lebesgue measure. Set u~(x)=u(x,s)\tilde{u}(x)=u(x,s), and X(ξ,t)X(\xi,t) is defined by (1.17) with u¯\bar{u} replaced by u~\tilde{u} (and μ¯((,x))\bar{\mu}((-\infty,x)) is replaced by (,x)u~x2(x)dx\int_{(-\infty,x)}\tilde{u}_{x}^{2}(x)\mathop{}\!\mathrm{d}x). Let y(α,t)y(\alpha,t) be the characteristics (in the α\alpha-coordinate) corresponding to (u¯,μ¯)(\bar{u},\bar{\mu}), see (2.12) for its definition. Then, we have

u~Cb(),u~xL2(),u~xL2=μ¯().\displaystyle\tilde{u}\in C_{b}(\mathbb{R}),\quad\tilde{u}_{x}\in L^{2}(\mathbb{R}),\quad\|\tilde{u}_{x}\|_{L^{2}}=\bar{\mu}(\mathbb{R}).

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

y(,t)=X(,ts)y(,s),μ(t)=X(,ts)#(u~x2dx),\displaystyle y(\cdot,t)=X(\cdot,t-s)\circ y(\cdot,s),\quad\mu(t)=X(\cdot,t-s)\#(\tilde{u}_{x}^{2}\mathop{}\!\mathrm{d}x),

and

u(x,t)=tX(ξ,ts)=u~(ξ)+(ts)2(,ξ)u~x2(y)dy for x=X(ξ,ts).\displaystyle u(x,t)=\frac{\partial}{\partial t}X(\xi,t-s)=\tilde{u}(\xi)+\frac{(t-s)}{2}\int_{(-\infty,\xi)}\tilde{u}_{x}^{2}(y)\mathop{}\!\mathrm{d}y~{}\textrm{ for }~{}x=X(\xi,t-s).

Global characteristics similar to y(α,t)y(\alpha,t) were also used in [6, 8] to construct the semi-group property of conservative solutions to the Hunter-Saxton equation, and obtain the Lipschitz metrics for stability. In [3], similar variables were used to show the uniqueness of conservative solutions to the Camassa-Holm equation via the characteristics method. There is another different set of variables defined by

k¯(η)=inf{x:μ¯((,x])η},\bar{k}(\eta)=\inf\{x:~{}~{}\bar{\mu}((-\infty,x])\geq\eta\},

provided that a Radon measure μ¯+()\bar{\mu}\in\mathcal{M}_{+}(\mathbb{R}) is given. Global characteristics k(η,t)k(\eta,t) can also be defined globally via initial datum (u¯,μ¯)(\bar{u},\bar{\mu}); see [7, 8] for instance. In [7], Bressan, Zhang and Zheng used this kind of characteristics to study the following more general model:

ut+g(u)x=120xg′′(u)ux2dx.u_{t}+g(u)_{x}=\frac{1}{2}\int_{0}^{x}g^{\prime\prime}(u)u_{x}^{2}\mathop{}\!\mathrm{d}x.

Here, the flux gg is a function with a Lipschitz continuous second-order derivative such that g′′(0)>0g^{\prime\prime}(0)>0. When g(u)=u2/2g(u)=u^{2}/2, the above equation becomes the Hunter-Saxton equation. They obtained global existence and uniqueness of conservative solutions on the half real line by using the above characteristics for compactly supported initial measure μ¯\bar{\mu}. This kind of characteristics was also used in [8] to study the stability of conservative solutions to the Hunter-Saxton equation. With the above map k¯\bar{k}, one has μ¯=k¯#|[0,μ¯()].\bar{\mu}=\bar{k}\#\mathcal{L}|_{[0,\bar{\mu}(\mathbb{R})]}. Comparing with x¯\bar{x} (which is globally Lipschitz) that we use in this paper, the function k¯\bar{k}, however, has potential jump discontinuities.

Except for conservative solutions, the Hunter-Saxton equation also has a class of weak solutions called dissipative solutions which dissipate the energy. Hunter and Zheng [14, 15] established the global existence of both dissipative and conservative weak solutions to (1.2) with initial data u¯xBV(0,)\bar{u}_{x}\in BV(0,\infty). Then, Zhang and Zheng [19, 20] studied the global existence of dissipative weak solutions to (1.2) with nonnegative initial data u¯xLp()\bar{u}_{x}\in L^{p}(\mathbb{R}) for any p2p\geq 2. Later in [21], they obtained global solutions (in both the dissipative and conservative cases) to the Hunter-Saxton equation on the half-line with any initial data u¯xL2()\bar{u}_{x}\in L^{2}(\mathbb{R}) by the theory of Young measures. In [5], Bressan and Constantin rewrote the equation in terms of new variables, and obtained global existence and uniqueness of dissipative solutions. The uniqueness of dissipative solutions were also studied by the characteristics methods in [9, 10, 11].

The rest of this paper is organized as follows. In Section 2, we are going to introduce global characteristics and construct global conservative solutions. The structure of these solutions will be studied in details. In Section 3, we will show that the construction based on the method of characteristics provides the global solution in the sense of Definition 1.2. The uniqueness of conservative solutions will be obtained via the characteristics method as well. In Appendix A, we will present some useful facts from real analysis, and construct an absolutely continuous initial datum by Cantor fat set; this initial datum will generate singular continuous part within a finite time.

2. Structure of conservative solutions via characteristics

In this section, we will introduce the characteristics to construct global conservative solutions and study their structure.

2.1. Global characteristics

To illustrate the idea, we first consider a smooth solution uu to the Hunter-Saxton equation (1.1). For any β\beta\in\mathbb{R}, we define x(β,t)x(\beta,t) by

(2.1) x(β,t)+(,x(β,t))ux2(y,t)dy=β,t.\displaystyle x(\beta,t)+\int_{(-\infty,x(\beta,t))}u_{x}^{2}(y,t)\mathop{}\!\mathrm{d}y=\beta,~{}~{}t\in\mathbb{R}.

Differentiating the above identity with respect to tt and β\beta respectively, we obtain

(2.2) tx(β,t)=(uux2)(x(β,t),t)1+ux2(x(β,t),t),βx(β,t)=11+ux2(x(β,t),t),\partial_{t}x(\beta,t)=\frac{(uu_{x}^{2})(x(\beta,t),t)}{1+u_{x}^{2}(x(\beta,t),t)},\qquad\partial_{\beta}x(\beta,t)=\frac{1}{1+u_{x}^{2}(x(\beta,t),t)},

where we used (1.3) for computing tx(β,t)\partial_{t}x(\beta,t). Then, for fixed α\alpha\in\mathbb{R}, we define a function β(t)\beta(t) by solving the following ordinary differential equation (ODE):

(2.3) β(t)=u(x(β(t),t),t),β(0)=α.\displaystyle\beta^{\prime}(t)=u(x(\beta(t),t),t),\quad\beta(0)=\alpha\in\mathbb{R}.

It follows from the chain rule, (2.2) and (2.3) that

(2.4) ddtx(β(t),t)=(tx+βxβ)(β(t),t)=u(x(β(t),t),t)=β(t).\displaystyle\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}x(\beta(t),t)=(\partial_{t}x+\partial_{\beta}x\cdot\beta^{\prime})(\beta(t),t)=u(x(\beta(t),t),t)=\beta^{\prime}(t).

Differentiating (2.3) with respect to the time tt again, we obtain

β′′(t)=\displaystyle\beta^{\prime\prime}(t)= tu(x(β(t),t),t)+xu(x(β(t),t),t)ddtx(β(t),t)=(ut+uux)(x(β(t),t),t)\displaystyle\partial_{t}u(x(\beta(t),t),t)+\partial_{x}u(x(\beta(t),t),t)\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}x(\beta(t),t)=(u_{t}+uu_{x})(x(\beta(t),t),t)
=\displaystyle= 12(,x(β(t),t))ux2(y,t)dy=12[β(t)x(β(t),t)],\displaystyle\frac{1}{2}\int_{(-\infty,x(\beta(t),t))}u_{x}^{2}(y,t)\mathop{}\!\mathrm{d}y=\frac{1}{2}[\beta(t)-x(\beta(t),t)],

where the last equality comes from (2.1). Taking one more time derivative and using (2.4), we have

(2.5) β′′′(t)=0.\displaystyle\beta^{\prime\prime\prime}(t)=0.

Denote x¯(α)=x(α,0)\bar{x}(\alpha)=x(\alpha,0), which is uniquely determined by

x¯(α)+(,x¯(α))ux2(y,0)dy=α.\displaystyle\bar{x}(\alpha)+\int_{(-\infty,\bar{x}(\alpha))}u_{x}^{2}(y,0)\mathop{}\!\mathrm{d}y=\alpha.

Then, we have the following initial data for the ODE (2.5):

(2.6) β(0)=α,β(0)=u(x¯(α),0),β′′(0)=12(αx¯(α))=12(,x¯(α))ux2(y,0)dy.\displaystyle\beta(0)=\alpha,\quad\beta^{\prime}(0)=u(\bar{x}(\alpha),0),\quad\beta^{\prime\prime}(0)=\frac{1}{2}(\alpha-\bar{x}(\alpha))=\frac{1}{2}\int_{(-\infty,\bar{x}(\alpha))}u_{x}^{2}(y,0)\mathop{}\!\mathrm{d}y.

Therefore, solving (2.5) and (2.6) yields

(2.7) β(t)=α+u(x¯(α),0)t+t24(αx¯(α)),\displaystyle\beta(t)=\alpha+u(\bar{x}(\alpha),0)t+\frac{t^{2}}{4}(\alpha-\bar{x}(\alpha)),

and hence, using (2.4), we also have

(2.8) x(β(t),t)=x¯(α)+u(x¯(α),0)t+t24(αx¯(α)).\displaystyle x(\beta(t),t)=\bar{x}(\alpha)+u(\bar{x}(\alpha),0)t+\frac{t^{2}}{4}(\alpha-\bar{x}(\alpha)).

Therefore, to obtain global formulas for β(t)\beta(t) and x(β(t),t)x(\beta(t),t), we only need the information of initial datum u(x,0)u(x,0). The above idea can be generalized to non-smooth111Here, a non-smooth pair (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D} means that the initial data μ¯\bar{\mu} is not absolutely continuous with respect to the Lebesgue measure \mathcal{L}. initial data in 𝒟\mathcal{D} as described below.

Consider an initial datum (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D}. For any α\alpha\in\mathbb{R}, we can also define the function x¯\bar{x} via

(2.9) x¯(α)+μ¯((,x¯(α)))αx¯(α)+μ¯((,x¯(α)]).\displaystyle\bar{x}(\alpha)+\bar{\mu}((-\infty,\bar{x}(\alpha)))\leq\alpha\leq\bar{x}(\alpha)+\bar{\mu}((-\infty,\bar{x}(\alpha)]).

If follows directly from the above definition of x¯\bar{x} that x¯(α)α\bar{x}(\alpha)\leq\alpha for all α\alpha\in\mathbb{R}. As an analogy to (2.7) and (2.8), we also have global-in-time β(t)\beta(t) and x(β(t),t)x(\beta(t),t) for non-smooth initial datum in 𝒟\mathcal{D} as follows:

(2.10) β(t)=α+u¯(x¯(α))t+t24(αx¯(α)),\displaystyle\beta(t)=\alpha+\bar{u}(\bar{x}(\alpha))t+\frac{t^{2}}{4}(\alpha-\bar{x}(\alpha)),

and

(2.11) x(β(t),t)=x¯(α)+u¯(x¯(α))t+t24(αx¯(α)).\displaystyle x(\beta(t),t)=\bar{x}(\alpha)+\bar{u}(\bar{x}(\alpha))t+\frac{t^{2}}{4}(\alpha-\bar{x}(\alpha)).

Hence, we can think of x(β(t),t)x(\beta(t),t) as a function of α\alpha and tt, then we define y(α,t)y(\alpha,t) as follows:

(2.12) y(α,t)=x¯(α)+u¯(x¯(α))t+t24(αx¯(α)).\displaystyle y(\alpha,t)=\bar{x}(\alpha)+\bar{u}(\bar{x}(\alpha))t+\frac{t^{2}}{4}(\alpha-\bar{x}(\alpha)).

Now, let us first show some properties for initial data in 𝒟\mathcal{D} as follows.

Proposition 2.1.

Let (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D} and μ¯=μ¯ac+μ¯pp+μ¯sc,\bar{\mu}=\bar{\mu}_{ac}+\bar{\mu}_{pp}+\bar{\mu}_{sc}, where μ¯ac\bar{\mu}_{ac}, μ¯pp\bar{\mu}_{pp}, and μ¯sc\bar{\mu}_{sc} are the absolutely continuous part, pure point part and the singular continuous part of μ¯\bar{\mu} respectively. Define a map x¯:\bar{x}:\mathbb{R}\to\mathbb{R} by (2.9). Then the following statements hold:

  1. (i)

    The function x¯\bar{x} is Lipschitz continuous with Lipschitz constant bounded by 11.

  2. (ii)

    Definition (2.9) is the same as

    (2.13) x¯(α)=sup{x:x+μ¯((,x))α},\displaystyle\bar{x}(\alpha)=\sup\left\{x:~{}x+\bar{\mu}((-\infty,x))\leq\alpha\right\},

    and

    (2.14) x¯(α)=inf{x:x+μ¯((,x])α}.\displaystyle\bar{x}(\alpha)=\inf\left\{x:~{}x+\bar{\mu}((-\infty,x])\geq\alpha\right\}.
  3. (iii)

    For α\alpha\in\mathbb{R}, we define

    (2.15) f(α)=1x¯(α).\displaystyle f(\alpha)=1-\bar{x}^{\prime}(\alpha).

    Then

    (2.16) x¯#(fdα)=μ¯,fL1=μ¯().\displaystyle\bar{x}\#(f\mathop{}\!\mathrm{d}\alpha)=\bar{\mu},\qquad\|f\|_{L^{1}}=\bar{\mu}(\mathbb{R}).
  4. (iv)

    Let

    A0L,pp={α:x¯(α)=0,z1(x¯(α))<z2(x¯(α))},A_{0}^{L,pp}=\{\alpha:~{}~{}\bar{x}^{\prime}(\alpha)=0,~{}~{}z_{1}(\bar{x}(\alpha))<z_{2}(\bar{x}(\alpha))\},

    and

    A0L,sc={α:x¯(α)=0,z1(x¯(α))=z2(x¯(α))},B0L={α:x¯(α)>0},\displaystyle A_{0}^{L,sc}=\{\alpha:~{}~{}\bar{x}^{\prime}(\alpha)=0,~{}~{}z_{1}(\bar{x}(\alpha))=z_{2}(\bar{x}(\alpha))\},\quad B_{0}^{L}=\{\alpha:~{}~{}\bar{x}^{\prime}(\alpha)>0\},

    where z1(x)=inf{α:x¯(α)=x}z_{1}(x)=\inf\{\alpha:~{}~{}\bar{x}(\alpha)=x\} and z2(x)=sup{α:x¯(α)=x}.z_{2}(x)=\sup\{\alpha:~{}~{}\bar{x}(\alpha)=x\}. Then, we have

    (2.17) μ¯pp=x¯#(f|A0L,ppdα),μ¯sc=x¯#(f|A0L,scdα),μ¯ac=x¯#(f|B0Ldα).\displaystyle\bar{\mu}_{pp}=\bar{x}\#(f|_{A_{0}^{L,pp}}\mathop{}\!\mathrm{d}\alpha),\quad\bar{\mu}_{sc}=\bar{x}\#(f|_{A_{0}^{L,sc}}\mathop{}\!\mathrm{d}\alpha),\quad\bar{\mu}_{ac}=\bar{x}\#(f|_{B_{0}^{L}}\mathop{}\!\mathrm{d}\alpha).

    Moreover,

    (2.18) u¯x2(x¯(α))x¯(α)=f(α),αB0L.\displaystyle\bar{u}_{x}^{2}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)=f(\alpha),\quad\alpha\in B_{0}^{L}.
Remark 2.1.
  1. (i)

    In this paper, the superscript LL on a set ALA^{L} means that we consider the set in the Lagrangian coordinates, while the superscript EE on a set AEA^{E} means that we consider the set in the Eulerian coordinates.

  2. (ii)

    It is worth noting that A0L,ppA_{0}^{L,pp} is defined in the point-wise sense, however, A0L,scA_{0}^{L,sc} and B0LB_{0}^{L} are defined in the a.e. sense. Moreover, when we say some identities hold for αB0L\alpha\in B_{0}^{L} or αA0L,sc\alpha\in A^{L,sc}_{0} (such as (2.18), (2.21) or (2.22) below), we always mean that the identities hold for a.e. αB0L\alpha\in B_{0}^{L} or αA0L,sc\alpha\in A^{L,sc}_{0}. Since this will not affect the analysis, we will omit the emphasize of a.e. in this paper for convenience. Similar treatment also applies to AtL,ppA_{t}^{L,pp}, AtL,scA_{t}^{L,sc}, and BtLB_{t}^{L}, which will be defined in Theorem 2.1 below.

Proof.

Since xx+μ¯((,x])x\mapsto x+\bar{\mu}((-\infty,x]) is strictly increasing with possible jumps, x¯:\bar{x}:\mathbb{R}\to\mathbb{R} is a non-decreasing function.

(i) Let α1,α2\alpha_{1},\alpha_{2}\in\mathbb{R} and α1<α2\alpha_{1}<\alpha_{2}. According to (2.9), we have

0x¯(α2)x¯(α1)\displaystyle 0\leq\bar{x}(\alpha_{2})-\bar{x}(\alpha_{1})\leq α2μ¯((,x¯(α2)))α1+μ¯((,x¯(α1)])\displaystyle\alpha_{2}-\bar{\mu}((-\infty,\bar{x}(\alpha_{2})))-\alpha_{1}+\bar{\mu}((-\infty,\bar{x}(\alpha_{1})])
\displaystyle\leq α2α1μ¯((x¯(α1),x¯(α2)))α2α1.\displaystyle\alpha_{2}-\alpha_{1}-\bar{\mu}((\bar{x}(\alpha_{1}),\bar{x}(\alpha_{2})))\leq\alpha_{2}-\alpha_{1}.

Hence, x¯\bar{x} is Lipschitz continuous with Lipschitz constant bounded by 11.

(ii) First, we show that definition (2.9) is equivalent to definition (2.13). Fix α\alpha\in\mathbb{R}. Let x1x_{1} and x2x_{2} satisfy

x1+μ¯((,x1))αx1+μ¯((,x1]),x_{1}+\bar{\mu}((-\infty,x_{1}))\leq\alpha\leq x_{1}+\bar{\mu}((-\infty,x_{1}]),

and

x2=sup{x|x+μ¯((,x))α}.x_{2}=\sup\left\{x|x+\bar{\mu}((-\infty,x))\leq\alpha\right\}.

Obviously, we have x2x1x_{2}\geq x_{1} and x2+μ¯((,x2))αx_{2}+\bar{\mu}((-\infty,x_{2}))\leq\alpha. If x1<x2x_{1}<x_{2}, then

αx1+μ¯((,x1])<x2+μ¯((,x2)),\alpha\leq x_{1}+\bar{\mu}((-\infty,x_{1}])<x_{2}+\bar{\mu}((-\infty,x_{2})),

which gives a contradiction.

Next, we show that definition (2.9) is equivalent to definition (2.14). Fix α\alpha\in\mathbb{R}. Let x1x_{1} and x3x_{3} satisfy

x1+μ¯((,x1))αx1+μ¯((,x1]),x_{1}+\bar{\mu}((-\infty,x_{1}))\leq\alpha\leq x_{1}+\bar{\mu}((-\infty,x_{1}]),

and

x3=inf{x|x+μ¯((,x])α}.x_{3}=\inf\left\{x|x+\bar{\mu}((-\infty,x])\geq\alpha\right\}.

Obviously, we have x3x1x_{3}\leq x_{1} and x3+μ¯((,x3])αx_{3}+\bar{\mu}((-\infty,x_{3}])\geq\alpha. If x3<x1x_{3}<x_{1}, then

αx1+μ¯((,x1))>x3+μ¯((,x3]),\alpha\geq x_{1}+\bar{\mu}((-\infty,x_{1}))>x_{3}+\bar{\mu}((-\infty,x_{3}]),

which gives a contradiction.

(iii) It suffices to show

μ¯((a,b))=x¯1((a,b))fdα,\bar{\mu}((a,b))=\int_{\bar{x}^{-1}((a,b))}f\mathop{}\!\mathrm{d}\alpha,

for any open interval (a,b)(a,b)\subset\mathbb{R}. Let

α1:=inf{α:x¯(α)>a},α2:=sup{α:x¯(α)<b}.\alpha_{1}:=\inf\{\alpha:~{}~{}\bar{x}(\alpha)>a\},\quad\alpha_{2}:=\sup\{\alpha:~{}~{}\bar{x}(\alpha)<b\}.

Since x¯\bar{x} is continuous, we have x¯(α1)=a\bar{x}(\alpha_{1})=a and x¯(α2)=b\bar{x}(\alpha_{2})=b. Hence, we have x¯1((a,b))=(α1,α2)\bar{x}^{-1}((a,b))=(\alpha_{1},\alpha_{2}),

a+μ¯((,a))α1a+μ¯((,a]),a+\bar{\mu}((-\infty,a))\leq\alpha_{1}\leq a+\bar{\mu}((-\infty,a]),

and

b+μ¯((,b))α2b+μ¯((,b]).b+\bar{\mu}((-\infty,b))\leq\alpha_{2}\leq b+\bar{\mu}((-\infty,b]).

We claim that

(2.19) α1=a+μ¯((,a]),α2=b+μ¯((,b)).\displaystyle\alpha_{1}=a+\bar{\mu}((-\infty,a]),\quad\alpha_{2}=b+\bar{\mu}((-\infty,b)).

We only prove the first one in (2.19) and the other one can be obtained similarly. If

a+μ¯((,a))α1<a+μ¯((,a])=:α~1,a+\bar{\mu}((-\infty,a))\leq\alpha_{1}<a+\bar{\mu}((-\infty,a])=:\tilde{\alpha}_{1},

then

x¯(α)=a,α[α1,α~1],\bar{x}(\alpha)=a,\quad\alpha\in[\alpha_{1},\tilde{\alpha}_{1}],

and

x¯(α)>a,α>α~1.\bar{x}(\alpha)>a,\quad\alpha>\tilde{\alpha}_{1}.

This implies a contradiction:

inf{α:x¯(α)>a}=α~1>α1.\inf\{\alpha:~{}~{}\bar{x}(\alpha)>a\}=\tilde{\alpha}_{1}>\alpha_{1}.

Hence, the claim (2.19) holds. We have

x¯1(a,b)fdα=α1α2[1x¯(α)]dα=\displaystyle\int_{\bar{x}^{-1}(a,b)}f\mathop{}\!\mathrm{d}\alpha=\int_{\alpha_{1}}^{\alpha_{2}}[1-\bar{x}^{\prime}(\alpha)]\mathop{}\!\mathrm{d}\alpha= α2α1x¯(α2)+x¯(α1)\displaystyle\alpha_{2}-\alpha_{1}-\bar{x}(\alpha_{2})+\bar{x}(\alpha_{1})
=\displaystyle= μ¯((,b))μ¯((,a])=μ¯((a,b)).\displaystyle\bar{\mu}((-\infty,b))-\bar{\mu}((-\infty,a])=\bar{\mu}((a,b)).

Next, we prove fL1=μ¯()\|f\|_{L^{1}}=\bar{\mu}(\mathbb{R}). By definition (2.9) of x¯(α)\bar{x}(\alpha), we have

μ¯((,x¯(α)))αx¯(α)μ¯((,x¯(α)]),α.\bar{\mu}((-\infty,\bar{x}(\alpha)))\leq\alpha-\bar{x}(\alpha)\leq\bar{\mu}((-\infty,\bar{x}(\alpha)]),\quad\alpha\in\mathbb{R}.

It is easy to see that x¯(α)±\bar{x}(\alpha)\to\pm\infty as α±\alpha\to\pm\infty. Therefore,

(2.20) limα[αx¯(α)]=0,limα+[αx¯(α)]=μ¯().\displaystyle\lim_{\alpha\to-\infty}[\alpha-\bar{x}(\alpha)]=0,\quad\lim_{\alpha\to+\infty}[\alpha-\bar{x}(\alpha)]=\bar{\mu}(\mathbb{R}).

This implies fL1=μ¯()\|f\|_{L^{1}}=\bar{\mu}(\mathbb{R}).

(iv) Relations (2.17) directly follow from Lemma A.2. Furthermore, since x¯#(f|B0Ldα)=μ¯ac\bar{x}\#(f|_{B_{0}^{L}}\mathop{}\!\mathrm{d}\alpha)=\bar{\mu}_{ac}, we actually have, for any test function φCb()\varphi\in C_{b}(\mathbb{R}),

φ(x)dμ¯ac=φ(x¯(α))f|B0L(α)dα.\int_{\mathbb{R}}\varphi(x)\mathop{}\!\mathrm{d}\bar{\mu}_{ac}=\int_{\mathbb{R}}\varphi(\bar{x}(\alpha))f|_{B_{0}^{L}}(\alpha)\mathop{}\!\mathrm{d}\alpha.

On the other hand, since (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D}, we have dμ¯ac=u¯x2dx\mathop{}\!\mathrm{d}\bar{\mu}_{ac}=\bar{u}_{x}^{2}\mathop{}\!\mathrm{d}x, and hence,

φ(x)dμ¯ac=φ(x)u¯x2(x)dx=φ(x¯(α))u¯x2(x¯(α))x¯(α)dα.\int_{\mathbb{R}}\varphi(x)\mathop{}\!\mathrm{d}\bar{\mu}_{ac}=\int_{\mathbb{R}}\varphi(x)\bar{u}_{x}^{2}(x)\mathop{}\!\mathrm{d}x=\int_{\mathbb{R}}\varphi(\bar{x}(\alpha))\bar{u}_{x}^{2}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)\mathop{}\!\mathrm{d}\alpha.

Combining the above two identities, we obtain (2.18). ∎

From Proposition 2.1, we have

(2.21) f(α)={u¯x2(x¯(α))x¯(α),αB0L,1,αA0L=A0L,ppA0L,sc.f(\alpha)=\left\{\begin{aligned} &\bar{u}_{x}^{2}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha),\quad\alpha\in B_{0}^{L},\\ &1,\quad\alpha\in A_{0}^{L}=A_{0}^{L,pp}\cup A_{0}^{L,sc}.\end{aligned}\right.

This means we use the increasing function x¯(α)\bar{x}(\alpha) to transform the singular part of μ¯\bar{\mu} to some constant part of f(α)f(\alpha). For example, if mδxm\delta_{x} is a pure point part of μ¯\bar{\mu}, we have z1(x)<z2(x)z_{1}(x)<z_{2}(x) and x¯(α)=x\bar{x}(\alpha)=x for α[z1(x),z2(x)]\alpha\in[z_{1}(x),z_{2}(x)]. Moreover, we also have f(α)=1f(\alpha)=1 for α[z1(x),z2(x)]\alpha\in[z_{1}(x),z_{2}(x)] and

m=μ¯({x})=x¯1({x})f(α)dα=[z1(x),z2(x)]dα=z2(x)z1(x).m=\bar{\mu}(\{x\})=\int_{\bar{x}^{-1}(\{x\})}f(\alpha)\mathop{}\!\mathrm{d}\alpha=\int_{[z_{1}(x),z_{2}(x)]}\mathop{}\!\mathrm{d}\alpha=z_{2}(x)-z_{1}(x).

Hence, the function x¯\bar{x} makes the singular part of μ¯\bar{\mu} supported at one point xx with mass mm uniformly distributed on the interval [z1(x),z2(x)][z_{1}(x),z_{2}(x)] with length mm.

We have the following important results about global characteristics y(α,t)y(\alpha,t):

Proposition 2.2.

Let (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D}. Let x¯:\bar{x}:\mathbb{R}\to\mathbb{R} be defined by (2.9) and y(α,t)y(\alpha,t) be defined by (2.12) for α,t\alpha,~{}t\in\mathbb{R}. Then, for any tt\in\mathbb{R}, y(,t)y(\cdot,t) is increasing and we have

(2.22) yα(α,t)={x¯(α)[1+t2u¯x(x¯(α))]2,αB0L,t24,αA0L.y_{\alpha}(\alpha,t)=\left\{\begin{aligned} &\bar{x}^{\prime}(\alpha)\left[1+\frac{t}{2}\bar{u}_{x}(\bar{x}(\alpha))\right]^{2},\quad\alpha\in B_{0}^{L},\\ &\frac{t^{2}}{4},\quad\alpha\in A_{0}^{L}.\end{aligned}\right.

Here, the sets A0LA_{0}^{L} and B0LB_{0}^{L} are the same as in (2.21).

Proof.

Differentiating (2.12) with respect to α\alpha yields

yα(α,t)=x¯(α)+u¯x(x¯(α))x¯(α)t+t24f(α).y_{\alpha}(\alpha,t)=\bar{x}^{\prime}(\alpha)+\bar{u}_{x}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)t+\frac{t^{2}}{4}f(\alpha).

Due to (2.21) and x¯(α)=0\bar{x}^{\prime}(\alpha)=0 for αA0L\alpha\in A_{0}^{L}, we have (2.22). ∎

Let us end this subsection with some remarks about the flow map, provided that the initial data μ¯\bar{\mu} is absolutely continuous with respect to the Lebesgue measure.

Remark 2.2.

Consider some initial datum (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D} such that dμ¯=u¯x2dx\mathop{}\!\mathrm{d}\bar{\mu}=\bar{u}_{x}^{2}\mathop{}\!\mathrm{d}x.

  1. (i)

    The cumulative energy distribution is defined by

    (2.23) F¯(x)=μ¯((,x))=μ¯((,x])=(,x)u¯x2(y)dy.\displaystyle\bar{F}(x)=\bar{\mu}((-\infty,x))=\bar{\mu}((-\infty,x])=\int_{(-\infty,x)}\bar{u}_{x}^{2}(y)\mathop{}\!\mathrm{d}y.

    Obviously,

    (2.24) F¯()=0 and F¯(+)=u¯xL22.\displaystyle\bar{F}(-\infty)=0~{}\textrm{ and }~{}\bar{F}(+\infty)=\|\bar{u}_{x}\|_{L^{2}}^{2}.

    In the α\alpha-variable, the function ff given by (2.15) plays the role of the energy density in the α\alpha-variable, and the cumulative energy distribution is given by

    (,α)f(α)dα=αx¯(α)=F¯(x¯(α)).\int_{(-\infty,\alpha)}f(\alpha)\mathop{}\!\mathrm{d}\alpha=\alpha-\bar{x}(\alpha)=\bar{F}(\bar{x}(\alpha)).
  2. (ii)

    Since μ¯\bar{\mu} is absolutely continuous, there is no need to use the variable α\alpha to define the cumulative energy distribution F¯\bar{F}. Indeed, we could also use the usual flow map to define the characteristics, which is defined by

    (2.25) {tX(ξ,t)=u(X(ξ,t),t),ξ,t>0,X(ξ,0)=ξ.\left\{\begin{aligned} &\frac{\partial}{\partial t}X(\xi,t)=u(X(\xi,t),t),~{}~{}\xi\in\mathbb{R},~{}~{}t>0,\\ &X(\xi,0)=\xi.\end{aligned}\right.

    Taking time derivative again and using the energy conservation condition, we could formally derive the following global trajectories by similar method to (2.7) or (2.8):

    (2.26) X(ξ,t)=ξ+u¯(ξ)t+t24F¯(ξ).\displaystyle X(\xi,t)=\xi+\bar{u}(\xi)t+\frac{t^{2}}{4}\bar{F}(\xi).

    We will discuss about the relation between y(α,t)y(\alpha,t) (given by (2.12)) and X(ξ,t)X(\xi,t) in the next subsection.

2.2. Structure of solutions

In this subsection and in particularly, Theorem 2.1, we will first define uu and μ\mu via (2.27) and (2.28) for a given initial datum (u¯,μ¯)(\bar{u},\bar{\mu}) as well as study their properties; we state and prove our main results for the structure of (u,μ)({u},{\mu}) using the variable α\alpha, which covers Theorem 1.1 and Theorem 1.2, except that, we will postpone to show that they are a global-in-time conservative solution to the generalized framework (1.4)-(1.6) in the sense of Definition 1.2 with initial datum (u¯,μ¯)(\bar{u},\bar{\mu}) in Section 3, see Theorem 3.1 .

Theorem 2.1.

Let (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D}, x¯(α)\bar{x}(\alpha), and f(α)f(\alpha) be the same as in Proposition 2.1. Let y(α,t)y(\alpha,t) be defined by (2.12), and z1(x,t)=inf{α:y(α,t)=x},z2(x,t)=sup{α:y(α,t)=x}z_{1}(x,t)=\inf\{\alpha:~{}~{}y(\alpha,t)=x\},~{}~{}z_{2}(x,t)=\sup\{\alpha:~{}~{}y(\alpha,t)=x\} be two pseudo-inverses of y(,t)y(\cdot,t) for a fixed tt\in\mathbb{R}. For any fixed tt\in\mathbb{R}, let

AtL,pp={α:yα(α,t)=0,z1(y(α,t),t)<z2(y(α,t),t)},A_{t}^{L,pp}=\{\alpha:~{}~{}y_{\alpha}(\alpha,t)=0,~{}~{}z_{1}(y(\alpha,t),t)<z_{2}(y(\alpha,t),t)\},
AtL,sc={α:yα(α,t)=0,z1(y(α,t),t)=z2(y(α,t),t)},andBtL={α:yα(α,t)>0}.\displaystyle A_{t}^{L,sc}=\{\alpha:~{}~{}y_{\alpha}(\alpha,t)=0,~{}~{}z_{1}(y(\alpha,t),t)=z_{2}(y(\alpha,t),t)\},\quad\mbox{and}\quad B_{t}^{L}=\{\alpha:~{}~{}y_{\alpha}(\alpha,t)>0\}.

For any tt\in\mathbb{R}, define

(2.27) u(x,t)=ty(α,t)=u¯(x¯(α))+t2(αx¯(α)) for x=y(α,t),\displaystyle u(x,t)=\frac{\partial}{\partial t}y(\alpha,t)=\bar{u}(\bar{x}(\alpha))+\frac{t}{2}(\alpha-\bar{x}(\alpha))~{}\textrm{ for }~{}x=y(\alpha,t),

and

(2.28) μ(t)=y(,t)#(fdα).\mu(t)=y(\cdot,t)\#(f\mathop{}\!\mathrm{d}\alpha).

Let μ(t)=μac(t)+μpp(t)+μsc(t)\mu(t)=\mu_{ac}(t)+\mu_{pp}(t)+\mu_{sc}(t), where μac(t)\mu_{ac}(t), μpp(t)\mu_{pp}(t), and μsc(t)\mu_{sc}(t) are the absolutely continuous part, pure point part and the singular continuous part of μ(t){\mu}(t) respectively. Then we have

  1. \bullet

    Properties of μ(t)\mu(t):

  2. (i)

    Energy conservation: we have μ(t)C(;+())\mu(t)\in C(\mathbb{R};\mathcal{M}_{+}(\mathbb{R})) and

    (2.29) μ(t)()=μ¯(),t.\displaystyle\mu(t)(\mathbb{R})=\bar{\mu}(\mathbb{R}),\quad t\in\mathbb{R}.
  3. (ii)

    For the structure of μ(t)\mu(t), we have

    (2.30) μpp(t)=y(,t)#(f|AtL,ppdα),μsc(t)=y(,t)#(f|AtL,scdα),μac(t)=y(,t)#(f|BtLdα).\displaystyle\mu_{pp}(t)=y(\cdot,t)\#(f|_{A_{t}^{L,pp}}\mathop{}\!\mathrm{d}\alpha),\quad\mu_{sc}(t)=y(\cdot,t)\#(f|_{A_{t}^{L,sc}}\mathop{}\!\mathrm{d}\alpha),\quad\mu_{ac}(t)=y(\cdot,t)\#(f|_{B_{t}^{L}}\mathop{}\!\mathrm{d}\alpha).

    Moreover, for any t0t\neq 0, we have

    (2.31) AtL,ppAtL,sc={αB0L:u¯x(x¯(α))=2t}:=AtL,\displaystyle A_{t}^{L,pp}\cup A_{t}^{L,sc}=\left\{\alpha\in B_{0}^{L}:~{}~{}\bar{u}_{x}(\bar{x}(\alpha))=-\frac{2}{t}\right\}:=A^{L}_{t},

    where B0LB_{0}^{L} was defined in Proposition 2.1 (iv). This implies that for any t0t\neq 0, μpp(t)\mu_{pp}(t) and μsc(t)\mu_{sc}(t) are determined by the absolutely continuous part of μ¯ac\bar{\mu}_{ac}, i.e., u¯x\bar{u}_{x}. Moreover,

    supp(μpp(t)+μsc(t)){x+u¯(x)t+t24μ¯((,x)):xAtE},\mathrm{supp}\Big{(}\mu_{pp}(t)+\mu_{sc}(t)\Big{)}\subset\Big{\{}x+\bar{u}(x)t+\frac{t^{2}}{4}\bar{\mu}((-\infty,x)):~{}x\in A_{t}^{E}\Big{\}},

    where AtEA_{t}^{E} is defined by (1.10).

  4. (iii)

    There are at most countably many time tt\in\mathbb{R} such that either the pure point part or the singular continuous part of μ(t)\mu(t) is not zero; in other words, both the sets

    (2.32) Tp:={t:μpp(t)0}andTs:={t:μsc(t)0}\displaystyle T_{p}:=\{t:~{}\mu_{pp}(t)\neq 0\}\quad\mbox{and}\quad T_{s}:=\{t:~{}\mu_{sc}(t)\neq 0\}

    are countable.

  5. \bullet

    Properties of u(x,t)u(x,t):

  6. (iv)

    For all time tt\in\mathbb{R}, the function u(,t)u(\cdot,t) is globally absolutely continuous and

    (2.33) dμac(t)=ux2(x,t)dx.\displaystyle\mathop{}\!\mathrm{d}\mu_{ac}(t)=u_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x.

    Moreover,

    (2.34) uC(;Cb())Cloc1/2(×),uxL(;L2()),utLloc2(×).\displaystyle u\in C(\mathbb{R};C_{b}(\mathbb{R}))\cap C^{1/2}_{loc}(\mathbb{R}\times\mathbb{R}),~{}~{}u_{x}\in L^{\infty}(\mathbb{R};L^{2}(\mathbb{R})),~{}~{}u_{t}\in L_{loc}^{2}(\mathbb{R}\times\mathbb{R}).
  7. (v)

    If u¯():=limxu¯(x)\bar{u}(-\infty):=\lim_{x\to-\infty}\bar{u}(x) exists, then we have

    (2.35) limxu(x,t)=u¯().\displaystyle\lim_{x\to-\infty}u(x,t)=\bar{u}(-\infty).

    On the other hand, if u¯(+):=limx+u¯(x)\bar{u}(+\infty):=\lim_{x\to+\infty}\bar{u}(x) exists, then we also have

    (2.36) limx+u(x,t)=u¯(+)+12μ¯()t.\displaystyle\lim_{x\to+\infty}u(x,t)=\bar{u}(+\infty)+\frac{1}{2}\bar{\mu}(\mathbb{R})t.
  8. \bullet

    Relations with the absolutely continuous part:

  9. (vi)

    Consider a time ss\in\mathbb{R} such that μ(s)\mu(s) is absolutely continuous with respect to the Lebesgue measure. Let u~(x)=u(x,s)\tilde{u}(x)=u(x,s), and X(ξ,t)X(\xi,t) and F~\tilde{F} be defined by (2.26) and (2.23) corresponding to u~\tilde{u}. Then, we have

    (2.37) u~Cb(),u~xL2(),u~xL2=μ¯().\displaystyle\tilde{u}\in C_{b}(\mathbb{R}),\quad\tilde{u}_{x}\in L^{2}(\mathbb{R}),\quad\|\tilde{u}_{x}\|_{L^{2}}=\bar{\mu}(\mathbb{R}).

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

    (2.38) y(,t)=X(,ts)y(,s),\displaystyle y(\cdot,t)=X(\cdot,t-s)\circ y(\cdot,s),
    (2.39) μ(t)=X(,ts)#(u~x2dx),\displaystyle\mu(t)=X(\cdot,t-s)\#(\tilde{u}_{x}^{2}\mathop{}\!\mathrm{d}x),

    and

    (2.40) u(x,t)=tX(ξ,ts)=u~(ξ)+(ts)2F~(ξ),for x=X(ξ,ts).\displaystyle u(x,t)=\frac{\partial}{\partial t}X(\xi,t-s)=\tilde{u}(\xi)+\frac{(t-s)}{2}\tilde{F}(\xi),\quad\textrm{for $x=X(\xi,t-s)$}.
Proof.

(i) For the continuity of measure μ(t)\mu(t), we take any bounded continuous function ϕ(x)\phi(x) and using Lebesgue dominated convergence theorem to obtain

limstϕ(x)dμ(s)=limstϕ(y(α,s))f(α)dα=ϕ(y(α,t))f(α)dα=ϕ(x)dμ(t).\lim_{s\to t}\int_{\mathbb{R}}\phi(x)\mathop{}\!\mathrm{d}\mu(s)=\lim_{s\to t}\int_{\mathbb{R}}\phi(y(\alpha,s))f(\alpha)\mathop{}\!\mathrm{d}\alpha=\int_{\mathbb{R}}\phi(y(\alpha,t))f(\alpha)\mathop{}\!\mathrm{d}\alpha=\int_{\mathbb{R}}\phi(x)\mathop{}\!\mathrm{d}\mu(t).

The conservation of total variation of μ(t)\mu(t) (i.e. (2.29)) follows directly from (2.16).

(ii) From Lemma A.2, we have decomposition (2.30).

From (2.22), we know that the singular part of μ(t)\mu(t) comes from the set (2.31). Hence, all the singular parts of μ(t)\mu(t) for t0t\neq 0 are determined by the function u¯x\bar{u}_{x}, or equivalently, the absolutely continuous part of μ¯\bar{\mu}. More precisely, the singular parts come from AtEA_{t}^{E} defined by (1.10), and

supp(μpp(t)+μsc(t))y(AtL,t)={x+u¯(x)t+t24μ¯((,x)):xAtE}.\mathrm{supp}\Big{(}\mu_{pp}(t)+\mu_{sc}(t)\Big{)}\subset y(A_{t}^{L},t)=\Big{\{}x+\bar{u}(x)t+\frac{t^{2}}{4}\bar{\mu}((-\infty,x)):~{}x\in A_{t}^{E}\Big{\}}.

(iii) Since AtL,ppA^{L,pp}_{t} consists of closed intervals with positive length, we can always pick a rational number from each interval. Due to u¯x(x¯(α))=2t\bar{u}_{x}(\bar{x}(\alpha))=-\frac{2}{t} for all αAtL,pp\alpha\in A^{L,pp}_{t}, for different tTpt\in T_{p}, the chosen rational numbers are different because the closed intervals at different times are different. Therefore, using the chosen rational numbers as indices of these intervals, we know that there are at most countable numbers in TpT_{p}.

Let tTst\in T_{s}. Since μsc(t)0\mu_{sc}(t)\neq 0, it follows from (2.30) that (AtL,sc)>0\mathcal{L}(A_{t}^{L,sc})>0. Furthermore, for t1,t2Tst_{1},t_{2}\in T_{s} with t1t2t_{1}\neq t_{2}, we have At1L,scAt2L,sc=A_{t_{1}}^{L,sc}\cap A_{t_{2}}^{L,sc}=\emptyset, since u¯x(x¯(α))=2t\bar{u}_{x}(\bar{x}(\alpha))=-\frac{2}{t} for all αAtL,sc\alpha\in A^{L,sc}_{t}. We obtain the desired result by applying Lemma A.1 (i).

(iv) \bullet Proof of the local absolutely continuity of u(,t)u(\cdot,t) for tt\in\mathbb{R}.

We first prove that u(,t)u(\cdot,t) is locally absolutely continuous. Consider mm non-overlapping intervals {(xk,yk)}k=1m\{(x_{k},y_{k})\}_{k=1}^{m} contained in (R,R)(-R,R) for some R>0R>0. Let xk=y(αk,t)x_{k}=y(\alpha_{k},t) and yk=y(βk,t)y_{k}=y(\beta_{k},t). Using (2.27) and the fundamental theorem of calculus, we have

k=1m|u(yk,t)u(xk,t)|=\displaystyle\sum_{k=1}^{m}|u(y_{k},t)-u(x_{k},t)|= k=1m|u¯(x¯(βk))+t2(βkx¯(βk))u¯(x¯(αk))t2(αkx¯(αk))|\displaystyle\sum_{k=1}^{m}\left|\bar{u}(\bar{x}(\beta_{k}))+\frac{t}{2}(\beta_{k}-\bar{x}(\beta_{k}))-\bar{u}(\bar{x}(\alpha_{k}))-\frac{t}{2}(\alpha_{k}-\bar{x}(\alpha_{k}))\right|
=\displaystyle= k=1m|[αk,βk]u¯x(x¯(α))x¯(α)+t2f(α)dα|.\displaystyle\sum_{k=1}^{m}\left|\int_{[\alpha_{k},\beta_{k}]}\bar{u}_{x}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)+\frac{t}{2}f(\alpha)\mathop{}\!\mathrm{d}\alpha\right|.

Notice that on AtLA_{t}^{L} given by (2.31), equality (2.18) holds and we have

(2.41) u¯x(x¯(α))x¯(α)+t2f(α)=u¯x(x¯(α))x¯(α)[1+t2u¯x(x¯(α))]=0.\displaystyle\bar{u}_{x}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)+\frac{t}{2}f(\alpha)=\bar{u}_{x}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)\left[1+\frac{t}{2}\bar{u}_{x}(\bar{x}(\alpha))\right]=0.

Define DL=k=1m[αk,βk]AtLD^{L}=\cup_{k=1}^{m}[\alpha_{k},\beta_{k}]\setminus A_{t}^{L}. Then, we have

(2.42) k=1m|u(yk,t)u(xk,t)|DL|u¯x(x¯(α))x¯(α)+t2f(α)|dα.\displaystyle\sum_{k=1}^{m}|u(y_{k},t)-u(x_{k},t)|\leq\int_{D^{L}}\left|\bar{u}_{x}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)+\frac{t}{2}f(\alpha)\right|\mathop{}\!\mathrm{d}\alpha.

Next, we will prove that for any η>0,\eta>0, there exists δ>0\delta>0 such that if k=1m(ykxk)<δ\sum_{k=1}^{m}(y_{k}-x_{k})<\delta, then

(DL)<η.\mathcal{L}(D^{L})<\eta.

This will prove the absolute continuity of u(,t)u(\cdot,t) due to the absolute continuity of the Lebesgue integral on the right hand side of (2.42). For this purpose, we define

DE=k=1m(xk,yk)y(AtL,t).D^{E}=\cup_{k=1}^{m}(x_{k},y_{k})\setminus y(A_{t}^{L},t).

Then z1(x,t)=z2(x,t)z_{1}(x,t)=z_{2}(x,t) on DED^{E}. Moreover, z1x(,t)z_{1x}(\cdot,t) exists on DED^{E} and

0<z1x(x,t)=1/yα(α,t)<,xDE,x=y(α,t).0<z_{1x}(x,t)=1/y_{\alpha}(\alpha,t)<\infty,\quad x\in D^{E},\quad x=y(\alpha,t).

Let

g(x)={z1x(x,t),x\y(AtL,t)0,xy(AtL,t).g(x)=\left\{\begin{aligned} &z_{1x}(x,t),~{}~{}x\in\mathbb{R}\backslash y(A_{t}^{L},t)\\ &0,~{}~{}x\in y(A_{t}^{L},t).\end{aligned}\right.

Then, obviously gLloc1()g\in L^{1}_{loc}(\mathbb{R}), since (a,b)g(x)dx(a,b)z1x(x,t)dxz1(b,t)z1(a,t)<\int_{(a,b)}g(x)\mathop{}\!\mathrm{d}x\leq\int_{(a,b)}z_{1x}(x,t)\mathop{}\!\mathrm{d}x\leq z_{1}(b,t)-z_{1}(a,t)<\infty for any b>ab>a. We have the following relation:

(DL)=(z1(DE,t))DEz1x(x,t)dx=k=1m[xk,yk]g(x)dx,\displaystyle\mathcal{L}(D^{L})=\mathcal{L}(z_{1}(D^{E},t))\leq\int_{D^{E}}z_{1x}(x,t)\mathop{}\!\mathrm{d}x=\sum_{k=1}^{m}\int_{[x_{k},y_{k}]}g(x)\mathop{}\!\mathrm{d}x,

where we used [1, Eq. (4.41)] for the inequality. Since gL1(K)g\in L^{1}(K) for any compact set KK\subset\mathbb{R}, if k=1m(ykxk)K\sum_{k=1}^{m}(y_{k}-x_{k})\subseteq K is sufficiently small (depending on the compact set KK), then the measure (DL)\mathcal{L}(D^{L}) will be smaller than η\eta. This shows that u(,t)u(\cdot,t) is locally absolutely continuous. Moreover, since ux(,t)L2()u_{x}(\cdot,t)\in L^{2}(\mathbb{R}) (which will be shown below), we can apply Lemma A.3 (with pp =2) to conclude that u(,t)u(\cdot,t) is globally absolutely continuous.

\bullet Proof of (2.33).

Since u(,t)u(\cdot,t) is absolutely continuous, differentiating u(y(α,t),t)=u¯(x¯(α))+t2(αx¯(α))u(y(\alpha,t),t)=\bar{u}(\bar{x}(\alpha))+\frac{t}{2}(\alpha-\bar{x}(\alpha)) with respect to α\alpha, and then taking a square, we obtain from (2.21) that

ux2(y(α,t),t)yα2(α,t)=[u¯x(x¯(α))x¯(α)]2[1+t2u¯x(x¯(α))]2for αB0L.u^{2}_{x}(y(\alpha,t),t)y^{2}_{\alpha}(\alpha,t)=[\bar{u}_{x}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)]^{2}\left[1+\frac{t}{2}\bar{u}_{x}(\bar{x}(\alpha))\right]^{2}\textrm{for $\alpha\in B_{0}^{L}$}.

Due to (2.22) and (2.21), we have

ux2(y(α,t),t)yα(α,t)=u¯x2(x¯(α))x¯(α)=f(α),αBtLB0L.\displaystyle u^{2}_{x}(y(\alpha,t),t)y_{\alpha}(\alpha,t)=\bar{u}_{x}^{2}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)=f(\alpha),~{}~{}\alpha\in B^{L}_{t}\cap B^{L}_{0}.

On the other hand, since x¯(α)=0\bar{x}^{\prime}(\alpha)=0 on A0LA^{L}_{0}, we have

ux(y(α,t),t)yα(α,t)=u¯x(x¯(α))x¯(α)+t2(1x¯(α))=t2for αA0L.u_{x}(y(\alpha,t),t)y_{\alpha}(\alpha,t)=\bar{u}_{x}(\bar{x}(\alpha))\bar{x}^{\prime}(\alpha)+\frac{t}{2}(1-\bar{x}^{\prime}(\alpha))=\frac{t}{2}~{}\textrm{for $\alpha\in A^{L}_{0}$}.

Using (2.22) again, we conclude that ux(y(α,t),t)=2tu_{x}(y(\alpha,t),t)=\frac{2}{t} on A0LBtLA^{L}_{0}\subset B^{L}_{t}. Hence, on A0LA^{L}_{0},

ux2(y(α,t),t)yα(α,t)=4t2t24=1=f(α).u_{x}^{2}(y(\alpha,t),t)y_{\alpha}(\alpha,t)=\frac{4}{t^{2}}\frac{t^{2}}{4}=1=f(\alpha).

In summary, we have

(2.43) ux2(y(α,t),t)yα(α,t)=f(α),αBtL.\displaystyle u^{2}_{x}(y(\alpha,t),t)y_{\alpha}(\alpha,t)=f(\alpha),~{}~{}\alpha\in B^{L}_{t}.

Since μac(t)=y(,t)#(f|BtLdα)\mu_{ac}(t)=y(\cdot,t)\#(f|_{B^{L}_{t}}\mathop{}\!\mathrm{d}\alpha) and (y(BtL,t))=0\mathcal{L}(\mathbb{R}\setminus y(B_{t}^{L},t))=0 (see Lemma A.2), relation (2.33) holds. \bullet Proof of (2.34).

From the above proof, we have uC(;Cb())u\in C(\mathbb{R};C_{b}(\mathbb{R})) and uxL(;L2())u_{x}\in L^{\infty}(\mathbb{R};L^{2}(\mathbb{R})). From the definition of uu, we also have utLloc2(×)u_{t}\in L_{loc}^{2}(\mathbb{R}\times\mathbb{R}). To prove (2.34), it suffice to show uCloc1/2(×)u\in C_{loc}^{1/2}(\mathbb{R}\times\mathbb{R}). Since we have, for any xyx\neq y and tst\neq s,

|u(x,t)u(y,s)|(|xy|2+|ts|2)14\displaystyle\frac{|u(x,t)-u(y,s)|}{{(|x-y|^{2}+|t-s|^{2})}^{\frac{1}{4}}} C|u(x,t)u(y,s)||xy|12+|ts|12\displaystyle\leq C\frac{|u(x,t)-u(y,s)|}{|x-y|^{\frac{1}{2}}+|t-s|^{\frac{1}{2}}}
C|u(x,t)u(y,t)||xy|12+C|u(y,t)u(y,s)||ts|12\displaystyle\leq C\frac{|u(x,t)-u(y,t)|}{|x-y|^{\frac{1}{2}}}+C\frac{|u(y,t)-u(y,s)|}{|t-s|^{\frac{1}{2}}}

for some constant C>0C>0, it is enough to show that uu is locally Hölder continuous with order 12\frac{1}{2} in spatial and temporal variable respectively. Since u(,t)Hloc1()u(\cdot,t)\in H^{1}_{loc}(\mathbb{R}) for all time tt\in\mathbb{R}, we have u(,t)Cloc1/2()u(\cdot,t)\in C^{1/2}_{loc}(\mathbb{R}). To show the local temporal Hölder continuity, we choose an α\alpha such that y(α,t)=xy(\alpha,t)=x for a given xx\in\mathbb{R}. Then using the spatial Hölder continuity and the definition of uu, we have

|u(x,t)u(x,s)|\displaystyle|u(x,t)-u(x,s)| |u(y(α,t),t)u(y(α,s),s)|+|u(y(α,s),s)u(y(α,t),s)|\displaystyle\leq|u(y(\alpha,t),t)-u(y(\alpha,s),s)|+|u(y(\alpha,s),s)-u(y(\alpha,t),s)|
|αx¯(α)|2|ts|+C|y(α,s)y(α,t)|12C|ts|12.\displaystyle\leq\frac{|\alpha-\bar{x}(\alpha)|}{2}|t-s|+C|y(\alpha,s)-y(\alpha,t)|^{\frac{1}{2}}\leq C|t-s|^{\frac{1}{2}}.

(v) This follows from (2.20) and definition (2.27).

(vi) Notice that (2.37) follows directly from (i) and (iv).

For this particular ss, μ(s)=y(,s)#(fdα)\mu(s)=y(\cdot,s)\#(f\mathop{}\!\mathrm{d}\alpha) is absolutely continuous, we have yα(α,s)>0y_{\alpha}(\alpha,s)>0 for a.e. α\alpha\in\mathbb{R}. Due to (2.22), this is equivalent to u¯x(x)2s\bar{u}_{x}(x)\neq-\frac{2}{s} for a.e. xx\in\mathbb{R}. From the definitions, we have

(2.44) y(α,s)=x¯(α)+u¯(x¯(α))s+s24(αx¯(α)),α,\displaystyle y(\alpha,s)=\bar{x}(\alpha)+\bar{u}(\bar{x}(\alpha))s+\frac{s^{2}}{4}(\alpha-\bar{x}(\alpha)),~{}~{}\alpha\in\mathbb{R},

and

(2.45) u~(x)=u(x,s)=u¯(x¯(α))+s2(αx¯(α)) for x=y(α,s).\displaystyle\tilde{u}(x)=u(x,s)=\bar{u}(\bar{x}(\alpha))+\frac{s}{2}(\alpha-\bar{x}(\alpha))~{}\textrm{ for }~{}x=y(\alpha,s).

By (ii) and (iii), we have u~x2dx=y(,s)#(fdα)\tilde{u}_{x}^{2}\mathop{}\!\mathrm{d}x=y(\cdot,s)\#(f\mathop{}\!\mathrm{d}\alpha). Hence,

(2.46) F~(y(α,s))=(,y(α,s))u~x2(x)dx=(,α)f(α)dα=αx¯(α).\displaystyle\tilde{F}(y(\alpha,s))=\int_{(-\infty,y(\alpha,s))}\tilde{u}_{x}^{2}(x)\mathop{}\!\mathrm{d}x=\int_{(-\infty,\alpha)}f(\alpha)\mathop{}\!\mathrm{d}\alpha=\alpha-\bar{x}(\alpha).

By (2.26), we have

X(ξ,ts)=ξ+u~(ξ)(ts)+(ts)24F~(ξ),ξ.X(\xi,t-s)=\xi+\tilde{u}(\xi)(t-s)+\frac{(t-s)^{2}}{4}\tilde{F}(\xi),~{}~{}\xi\in\mathbb{R}.

Hence, combining (2.44), (2.45), and (2.46) yields (2.38): for any α\alpha, tt\in\mathbb{R},

X(y(α,s),ts)\displaystyle X(y(\alpha,s),t-s) =y(α,s)+u~(y(α,s))(ts)+(ts)24F~(y(α,s))\displaystyle=y(\alpha,s)+\tilde{u}(y(\alpha,s))(t-s)+\frac{(t-s)^{2}}{4}\tilde{F}(y(\alpha,s))
=x¯(α)+u¯(x¯(α))s+s24(αx¯(α))+[u¯(x¯(α))+s2(αx¯(α))](ts)\displaystyle=\bar{x}(\alpha)+\bar{u}(\bar{x}(\alpha))s+\frac{s^{2}}{4}(\alpha-\bar{x}(\alpha))+\left[\bar{u}(\bar{x}(\alpha))+\frac{s}{2}(\alpha-\bar{x}(\alpha))\right](t-s)
+(ts)24(αx¯(α))\displaystyle\quad+\frac{(t-s)^{2}}{4}(\alpha-\bar{x}(\alpha))
=x¯(α)+u¯(x¯(α))t+t24(αx¯(α))=y(α,t).\displaystyle=\bar{x}(\alpha)+\bar{u}(\bar{x}(\alpha))t+\frac{t^{2}}{4}(\alpha-\bar{x}(\alpha))=y(\alpha,t).

For (2.39), we have

μ(t)=y(,t)#(fdα)=[X(,ts)y(,s)]#(fdα)=X(,ts)#[y(,s)#(fdα)]=X(,ts)#(u~x2dα).\mu(t)=y(\cdot,t)\#(f\mathop{}\!\mathrm{d}\alpha)=[X(\cdot,t-s)\circ y(\cdot,s)]\#(f\mathop{}\!\mathrm{d}\alpha)=X(\cdot,t-s)\#[y(\cdot,s)\#(f\mathop{}\!\mathrm{d}\alpha)]=X(\cdot,t-s)\#(\tilde{u}_{x}^{2}\mathop{}\!\mathrm{d}\alpha).

Next, we show (2.40). For x=X(ξ,ts)x=X(\xi,t-s), consider α\alpha satisfying ξ=y(α,s)\xi=y(\alpha,s) and then

x=X(y(α,s),ts)=y(α,t).x=X(y(\alpha,s),t-s)=y(\alpha,t).

We have

u~(ξ)+(ts)2F~(ξ)\displaystyle\tilde{u}(\xi)+\frac{(t-s)}{2}\tilde{F}(\xi) =u¯(x¯(α))+s2(αx¯(α))+(ts)2(αx¯(α))\displaystyle=\bar{u}(\bar{x}(\alpha))+\frac{s}{2}(\alpha-\bar{x}(\alpha))+\frac{(t-s)}{2}(\alpha-\bar{x}(\alpha))
=u¯(x¯(α))+t2(αx¯(α))=u(x,t).\displaystyle=\bar{u}(\bar{x}(\alpha))+\frac{t}{2}(\alpha-\bar{x}(\alpha))=u(x,t).

Let us end this section by some remarks.

Remark 2.3.
  1. (i)

    (Relation with regular initial data) For a non-smooth initial datum (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D}, let (u,μ)(u,\mu) be a solution to the generalized framework (1.4)-(1.6) defined by (2.27) and (2.28). Then (u,μ)(u,\mu) can also be constructed by a regular function u~(x)\tilde{u}(x) as given in Theorem 2.1 (vi). More precisely, such a regular function u~(x):=u(x,s)\tilde{u}(x):=u(x,s) is obtained by evaluating uu at a time ss such that μ\mu is absolutely continuous. This time ss always exists due to Theorem 2.1 (iii). Start from u~\tilde{u} and we can use the traditional flow map X(ξ,t)X(\xi,t) in the Lagrangian coordinates to construct solution (v,ν)(v,\nu). Then (u,μ)(u,\mu) will be recovered by some time shifting of (v,ν)(v,\nu).

  2. (ii)

    (Separation of singular part and absolutely continuous part of μ(t)\mu(t)) In Proposition 2.1 (iv), we used x¯(α)\bar{x}(\alpha) to identify the singular parts (corresponding to A0L=A0L,ppA0L,scA_{0}^{L}=A_{0}^{L,pp}\cup A_{0}^{L,sc}) and absolutely continuous part of μ¯\bar{\mu} (corresponding to B0LB_{0}^{L}). Since the singular part of μ¯\bar{\mu} comes from A0LA_{0}^{L}, we can conclude from Theorem 2.1 (ii) (or (2.31)) that this singular part will never create singular parts of μ(t)\mu(t) again for any other tt, and all the singular parts of μ(t)\mu(t) (t0t\neq 0) are generated by the absolutely continuous part of μ¯\bar{\mu}, i.e. u¯x\bar{u}_{x}. We can also see this from the fact that yα(α,t)=t24>0y_{\alpha}(\alpha,t)=\frac{t^{2}}{4}>0 for αA0L\alpha\in A_{0}^{L} (see (2.22)).

  3. (iii)

    (Calculation of mass for singular part of μ(t)\mu(t)) Due to f(α)=1x¯(α)f(\alpha)=1-\bar{x}^{\prime}(\alpha), we deduce from (2.18) that x¯(α)=t2t2+4\bar{x}^{\prime}(\alpha)=\frac{t^{2}}{t^{2}+4} and f(α)=4t2+4f(\alpha)=\frac{4}{t^{2}+4} on AtLA^{L}_{t}. We consider a point x0y(AtL,pp,t)x_{0}\in y(A_{t}^{L,pp},t) and α1:=z1(x0,t)<z2(x0,t)=:α2\alpha_{1}:=z_{1}(x_{0},t)<z_{2}(x_{0},t)=:\alpha_{2}. Then, u¯x(x)=2t\bar{u}_{x}(x)=-\frac{2}{t} for all x[x1,x2]:=[x¯(α1),x¯(α2)]x\in[x_{1},x_{2}]:=[\bar{x}(\alpha_{1}),\bar{x}(\alpha_{2})]. The mass concentrated at xx is calculated by

    μ(t)({x0})=[α1,α2]f(α)dα=4t2+4(α2α1)=4t2(x2x1).\displaystyle\mu(t)(\{x_{0}\})=\int_{[\alpha_{1},\alpha_{2}]}f(\alpha)\mathop{}\!\mathrm{d}\alpha=\frac{4}{t^{2}+4}(\alpha_{2}-\alpha_{1})=\frac{4}{t^{2}}(x_{2}-x_{1}).

    For the singular continuous part, define AtE,sc=x¯(AtL,sc)A_{t}^{E,sc}=\bar{x}(A_{t}^{L,sc}) and then u¯x(x)=2t\bar{u}_{x}(x)=-\frac{2}{t} for xAtE,scx\in A_{t}^{E,sc}. We have

    μsc(t)()=4t2(AtE,sc).\displaystyle\mu_{sc}(t)(\mathbb{R})=\frac{4}{t^{2}}\mathcal{L}(A_{t}^{E,sc}).

    See Example A.1 for a detailed calculation.

  4. (iv)

    (Luzin N property) The absolutely continuous result stated in Theorem 2.1 (iv) is more straightforward when tTsTpt\notin T_{s}\cup T_{p}. This is because y(,t)y(\cdot,t) is strictly increasing when tTsTpt\notin T_{s}\cup T_{p}, and there exists a unique absolutely continuous inverse of y(,t)y(\cdot,t) by (ii) in Lemma A.1. Using this inverse, we can directly prove that u(,t)u(\cdot,t) is absolutely continuous on any bounded interval. Moreover, we have

    ux2(x,t)dx=μac(t)()=μ(t)()=μ¯(),for all tTsTp.\int_{\mathbb{R}}u_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x=\mu_{ac}(t)(\mathbb{R})=\mu(t)(\mathbb{R})=\bar{\mu}(\mathbb{R}),\quad\mbox{for all }t\notin T_{s}\cup T_{p}.

    However, for tTsTpt\in T_{s}\cup T_{p}, the inverse of y(,t)y(\cdot,t) does not satisfy the Luzin N property on the set AtL,ppAtL,scA_{t}^{L,pp}\cup A_{t}^{L,sc} (see Example A.1 for instance) and it is not absolutely continuous. In this case, the function u(,t)u(\cdot,t) is still absolutely continuous, and we have

    ux2(x,t)dx=μac(t)()<μ(t)()=μ¯(),for all tTsTp.\int_{\mathbb{R}}u_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x=\mu_{ac}(t)(\mathbb{R})<\mu(t)(\mathbb{R})=\bar{\mu}(\mathbb{R}),\quad\mbox{for all }t\in T_{s}\cup T_{p}.

    The above analysis also shows that usually uxC(;L2())u_{x}\notin C(\mathbb{R};L^{2}(\mathbb{R})).

3. Existence and uniqueness of conservative solutions

In this section, we are going to show that (u,μ)(u,\mu) defined by (2.27)-(2.28) is a conservative solution to the system (1.4)-(1.6). We will also show the uniqueness of conservative solutions via characteristics method.

3.1. Existence

We have the following existence theorem:

Theorem 3.1 (Existence).

Let (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D} be an initial datum. Let uu be defined by (2.27) and μ\mu be defined by (2.28). Then, (u(t),μ(t))(u(t),\mu(t)) is a global-in-time conservative solution to the generalized framework (1.4)-(1.6) in the sense of Definition 1.2 with initial datum (u¯,μ¯)(\bar{u},\bar{\mu}). Moreover, the function uu and energy measure μ\mu satisfy all the properties in Theorem 2.1.

Proof.

It follows from Theorem 2.1, Equation (2.27) and (2.28) that (u(t),μ(t))(u(t),\mu(t)) satisfy properties (i), (ii) and (v) of Definition 1.2. We are going to use the change of variables y(α,t)y(\alpha,t) to prove (iii) and (iv) (i.e., (1.7) and (1.8)) of Definition 1.2. Since utu_{t} has enough regularity and y(α,t)y(\alpha,t) is an absolutely continuous function of α\alpha, we have, for ϕCc(×)\phi\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{R}),

uϕtdxdt=utϕdxdt=ut(y(α,t),t)ϕ(y(α,t),t)yα(α,t)dαdt\displaystyle\int_{\mathbb{R}}\int_{\mathbb{R}}u\phi_{t}\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=-\int_{\mathbb{R}}\int_{\mathbb{R}}u_{t}\phi\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t=-\int_{\mathbb{R}}\int_{\mathbb{R}}u_{t}(y(\alpha,t),t)\phi(y(\alpha,t),t)y_{\alpha}(\alpha,t)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t
=\displaystyle= ddtu(y(α,t),t)ϕ(y(α,t),t)yα(α,t)dαdt+(uux)(y(α,t),t)ϕ(y(α,t),t)yα(α,t)dαdt\displaystyle-\int_{\mathbb{R}}\int_{\mathbb{R}}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}u(y(\alpha,t),t)\phi(y(\alpha,t),t)y_{\alpha}(\alpha,t)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t+\int_{\mathbb{R}}\int_{\mathbb{R}}(uu_{x})(y(\alpha,t),t)\phi(y(\alpha,t),t)y_{\alpha}(\alpha,t)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t
=\displaystyle= ddtu(y(α,t),t)ϕ(y(α,t),t)yα(α,t)dαdt+(uux)(x,t)ϕ(x,t)dxdt.\displaystyle-\int_{\mathbb{R}}\int_{\mathbb{R}}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}u(y(\alpha,t),t)\phi(y(\alpha,t),t)y_{\alpha}(\alpha,t)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t+\int_{\mathbb{R}}\int_{\mathbb{R}}(uu_{x})(x,t)\phi(x,t)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t.

Comparing with equation (1.7), we are left to show

(3.1) ddtu(y(α,t),t)ϕ(y(α,t),t)yα(α,t)dαdt\displaystyle\int_{\mathbb{R}}\int_{\mathbb{R}}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}u(y(\alpha,t),t)\phi(y(\alpha,t),t)y_{\alpha}(\alpha,t)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t =12ϕ(x,t)F(x,t)dxdt\displaystyle=\frac{1}{2}\int_{\mathbb{R}}\int_{\mathbb{R}}\phi(x,t)F(x,t)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}t
=12ϕ(y(α,t),t)F(y(α,t),t)yα(α,t)dαdt.\displaystyle=\frac{1}{2}\int_{\mathbb{R}}\int_{\mathbb{R}}\phi(y(\alpha,t),t)F(y(\alpha,t),t)y_{\alpha}(\alpha,t)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t.

Notice that from the definition of uu (see (2.27)), we have ddtu(y(α,t),t)=12(αx¯(α))\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}u(y(\alpha,t),t)=\frac{1}{2}(\alpha-\bar{x}(\alpha)) and the integral on the left hand side of (3.1) becomes

ddtu(y(α,t),t)ϕ(y(α,t),t)yα(α,t)dαdt=12(αx¯(α))ϕ(y(α,t),t)yα(α,t)dαdt.\int_{\mathbb{R}}\int_{\mathbb{R}}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}u(y(\alpha,t),t)\phi(y(\alpha,t),t)y_{\alpha}(\alpha,t)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t=\int_{\mathbb{R}}\int_{\mathbb{R}}\frac{1}{2}(\alpha-\bar{x}(\alpha))\phi(y(\alpha,t),t)y_{\alpha}(\alpha,t)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t.

On the other hand, for all but countably many time tt, μ(t)\mu(t) is absolutely continuous. For these time tt, the absolutely continuity of μ(t)\mu(t) and (2.28) imply

F(y(α,t),t)=(,y(α,t))dμ(t)=(,α)f(α)dα=αx¯(α).\displaystyle F(y(\alpha,t),t)=\int_{(-\infty,y(\alpha,t))}\mathop{}\!\mathrm{d}\mu(t)=\int_{(-\infty,\alpha)}f(\alpha)\mathop{}\!\mathrm{d}\alpha=\alpha-\bar{x}(\alpha).

This proves (3.1), and hence, equation (1.7) follows.

Next, we prove (1.8). Using the definition (2.28) of μ(t)\mu(t) again, we obtain, for ϕCc(×)\phi\in C_{c}^{\infty}(\mathbb{R}\times\mathbb{R}),

(3.2) (ϕt+uϕx)dμ(t)dt\displaystyle\int_{\mathbb{R}}\int_{\mathbb{R}}(\phi_{t}+u\phi_{x})\mathop{}\!\mathrm{d}\mu(t)\mathop{}\!\mathrm{d}t =[ϕt(y(α,t),t)+u(y(α,t),t)ϕx(y(α,t),t)]f(α)dαdt\displaystyle=\int_{\mathbb{R}}\int_{\mathbb{R}}\Big{[}\phi_{t}(y(\alpha,t),t)+u(y(\alpha,t),t)\phi_{x}(y(\alpha,t),t)\Big{]}f(\alpha)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t
=ddtϕ(y(α,t),t)f(α)dαdt\displaystyle=\int_{\mathbb{R}}\int_{\mathbb{R}}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\phi(y(\alpha,t),t)f(\alpha)\mathop{}\!\mathrm{d}\alpha\mathop{}\!\mathrm{d}t
=ddt(ϕ(y(α,t),t)f(α)dα)dt=0.\displaystyle=\int_{\mathbb{R}}\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}\left(\int_{\mathbb{R}}\phi(y(\alpha,t),t)f(\alpha)\mathop{}\!\mathrm{d}\alpha\right)\mathop{}\!\mathrm{d}t=0.

Remark 3.1.

It is obvious to see that (1.8) still holds for all bounded smooth ϕ(x,t)\phi(x,t) with bounded derivatives supported on (,)×(M,M)(-\infty,\infty)\times(-M,M) for any M>0M>0 from the proof of (3.2). This property will be used in the proof of Lemma 3.1.

3.2. Uniqueness

We will show the uniqueness of conservative solutions via the characteristics method. We use the methods in [2, 3, 4] with some improvements.

Lemma 3.1.

Let (v,ν)(v,\nu) be a conservative solution to the generalized framework (1.4)-(1.6) in the sense of Definition 1.2. Consider the time tt and τ\tau such that ν\nu is absolutely continuous. Then, for any fixed yy\in\mathbb{R} and ϵ0>0\epsilon_{0}>0,

(3.3) (,y+a(tτ))vx2(x,t)dx(,y)vx2(x,τ)dx(,y+a+(tτ))vx2(x,t)dx,\displaystyle\int_{(-\infty,y+a_{-}(t-\tau))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x\leq\int_{(-\infty,y)}v_{x}^{2}(x,\tau)\mathop{}\!\mathrm{d}x\leq\int_{(-\infty,y+a_{+}(t-\tau))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x,

provided that tτ>0t-\tau>0 is small enough (depending on vv, yy and ϵ0\epsilon_{0}), where a±:=v(y,τ)±ϵ0a_{\pm}:=v(y,\tau)\pm\epsilon_{0}. Moreover, for any T>0T>0 and any Tτ<tT-T\leq\tau<t\leq T,

(3.4) (,yCT(tτ))vx2(x,t)dx(,y)vx2(x,τ)dx(,y+CT(tτ))vx2(x,t)dx,\displaystyle\int_{(-\infty,y-C_{T}(t-\tau))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x\leq\int_{(-\infty,y)}v_{x}^{2}(x,\tau)\mathop{}\!\mathrm{d}x\leq\int_{(-\infty,y+C_{T}(t-\tau))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x,

for all CTC_{T} satisfying vCb(×[T,T])CT\|v\|_{C_{b}(\mathbb{R}\times[-T,T])}\leq C_{T}.

Proof.

We are going to construct some test functions and use (1.8) to prove this lemma.

For any fixed ϵ>0\epsilon>0 sufficiently small, consider the following two non-negative smooth functions:

(3.5) θϵ(x)={1,xy,0,xy+ϵ,\theta_{\epsilon}(x)=\left\{\begin{aligned} &1,\quad x\leq y,\\ &0,\quad x\geq y+\epsilon,\end{aligned}\right.

and χϵ(s)=1\chi_{\epsilon}(s)=1 for s[τ,t]s\in[\tau,t] and suppχϵ(τϵ,t+ϵ)\operatorname{supp}\chi_{\epsilon}\subset(\tau-\epsilon,t+\epsilon). Assume that

θϵ0,χϵ(s)>0 for s(τϵ,τ),χϵ(s)<0 for s(t,t+ϵ).\theta_{\epsilon}^{\prime}\leq 0,\quad\chi^{\prime}_{\epsilon}(s)>0~{}\textrm{ for }~{}s\in(\tau-\epsilon,\tau),\quad\chi^{\prime}_{\epsilon}(s)<0~{}\textrm{ for }~{}s\in(t,t+\epsilon).

Let ϕϵ±(x,s)=θϵ(xa±(sτ))χϵ(s)Cb(×[0,))\phi^{\pm}_{\epsilon}(x,s)=\theta_{\epsilon}(x-a_{\pm}(s-\tau))\chi_{\epsilon}(s)\in C^{\infty}_{b}(\mathbb{R}\times[0,\infty)). Then,

(ϕϵt±+vϕϵx±)(x,s)=χϵ(s)θϵ(xa±(sτ))+(va±)χϵ(s)θϵ(xa±(sτ)).\displaystyle(\phi^{\pm}_{\epsilon t}+v\phi^{\pm}_{\epsilon x})(x,s)=\chi^{\prime}_{\epsilon}(s)\theta_{\epsilon}(x-a_{\pm}(s-\tau))+(v-a_{\pm})\chi_{\epsilon}(s)\theta^{\prime}_{\epsilon}(x-a_{\pm}(s-\tau)).

Since the conservative solution vv satisfies (1.8) and ϕϵ±\phi^{\pm}_{\epsilon} is supported on (,)×[τϵ,t+ϵ](-\infty,\infty)\times[\tau-\epsilon,t+\epsilon], we can take ϕϵ±\phi^{\pm}_{\epsilon} as a test function in (1.8) (cf. Remark 3.1), and obtain

(3.6) 0=(τϵτ+tt+ϵ)y+a±(sτ)+ϵχϵ(s)θϵ(xa±(sτ))vx2(x,s)dxds+τϵt+ϵy+a±(sτ)y+a±(sτ)+ϵ(v(x,s)a±)χϵ(s)θϵ(xa±(sτ))vx2(x,s)dxds.0=\left(\int_{\tau-\epsilon}^{\tau}+\int_{t}^{t+\epsilon}\right)\int_{-\infty}^{y+a_{\pm}(s-\tau)+\epsilon}\chi^{\prime}_{\epsilon}(s)\theta_{\epsilon}(x-a_{\pm}(s-\tau))v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}s\\ +\int_{\tau-\epsilon}^{t+\epsilon}\int_{y+a_{\pm}(s-\tau)}^{y+a_{\pm}(s-\tau)+\epsilon}(v{(x,s)}-a_{\pm})\chi_{\epsilon}(s)\theta^{\prime}_{\epsilon}(x-a_{\pm}(s-\tau))v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}s.

Consider the case for a+a_{+}. When tτt-\tau and ϵ\epsilon are sufficiently small, by the continuity of vv, we have |v(x,s)v(y,τ)|<ϵ0|v(x,s)-v(y,\tau)|<\epsilon_{0} for s(τϵ,t+ϵ)s\in(\tau-\epsilon,t+\epsilon) and x(y+a+(sτ),y+a+(sτ)+ϵ)x\in(y+a_{+}(s-\tau),y+a_{+}(s-\tau)+\epsilon), which implies v(x,s)a+=v(x,s)v(y,τ)ϵ00v(x,s)-a_{+}=v(x,s)-v(y,\tau)-\epsilon_{0}\leq 0. Hence, the second term in (3.6) is positive due to θϵ0\theta^{\prime}_{\epsilon}\leq 0 and χϵ0\chi_{\epsilon}\geq 0. Therefore,

(3.7) (τϵτ+tt+ϵ)y+a+(sτ)+ϵχϵ(s)θϵ(xa+(sτ))vx2(x,s)dxds0.\displaystyle\left(\int_{\tau-\epsilon}^{\tau}+\int_{t}^{t+\epsilon}\right)\int_{-\infty}^{y+a_{+}(s-\tau)+\epsilon}\chi^{\prime}_{\epsilon}(s)\theta_{\epsilon}(x-a_{+}(s-\tau))v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}s\leq 0.

By the definition of θϵ\theta_{\epsilon} and χϵ\chi_{\epsilon}, we have

τϵτy+a+(sτ)+ϵχϵ(s)θϵ(xa+(sτ))vx2(x,s)dxds\displaystyle\int_{\tau-\epsilon}^{\tau}\int_{-\infty}^{y+a_{+}(s-\tau)+\epsilon}\chi^{\prime}_{\epsilon}(s)\theta_{\epsilon}(x-a_{+}(s-\tau))v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}s\geq τϵτχϵ(s)y+a+(sτ)vx2(x,s)dxds\displaystyle\int_{\tau-\epsilon}^{\tau}\chi^{\prime}_{\epsilon}(s)\int_{-\infty}^{y+a_{+}(s-\tau)}v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}s
\displaystyle\geq infs[τϵ,τ]0y+a+(sτ)vx2(x,s)dx,\displaystyle\inf_{s\in[\tau-\epsilon,\tau]^{0}}\int_{-\infty}^{y+a_{+}(s-\tau)}v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x,

and similarly,

tt+ϵy+a+(sτ)+ϵχϵ(s)θϵ(xa+(sτ))vx2(x,s)dxdssups[t,t+ϵ]0y+a+(sτ)+ϵvx2(x,s)dx.\int_{t}^{t+\epsilon}\int_{-\infty}^{y+a_{+}(s-\tau)+\epsilon}\chi^{\prime}_{\epsilon}(s)\theta_{\epsilon}(x-a_{+}(s-\tau))v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}s\geq-\sup_{s\in[t,t+\epsilon]^{0}}\int_{-\infty}^{y+a_{+}(s-\tau)+\epsilon}v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x.

Here, we use [τϵ,τ]0[\tau-\epsilon,\tau]^{0} and [t,t+ϵ]0[t,t+\epsilon]^{0} to stand for those times in [τϵ,τ][\tau-\epsilon,\tau] and [t,t+ϵ][t,t+\epsilon] such that ν\nu is absolutely continuous. As a measure, ν\nu is continuous in time, so using [12, Theorem 1.40], we have

limϵ0+τϵτy+a+(sτ)+ϵχϵ(s)θϵ(xa+(sτ))vx2(x,s)dxdsyvx2(x,τ)dx,\lim_{\epsilon\to 0^{+}}\int_{\tau-\epsilon}^{\tau}\int_{-\infty}^{y+a_{+}(s-\tau)+\epsilon}\chi^{\prime}_{\epsilon}(s)\theta_{\epsilon}(x-a_{+}(s-\tau))v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}s\geq\int_{-\infty}^{y}v_{x}^{2}(x,\tau)\mathop{}\!\mathrm{d}x,

and

limϵ0+tt+ϵy+a+(sτ)+ϵχϵ(s)θϵ(xa+(sτ))vx2(x,s)dxdsy+a+(tτ)vx2(x,t)dx.\lim_{\epsilon\to 0^{+}}\int_{t}^{t+\epsilon}\int_{-\infty}^{y+a_{+}(s-\tau)+\epsilon}\chi^{\prime}_{\epsilon}(s)\theta_{\epsilon}(x-a_{+}(s-\tau))v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}s\geq-\int_{-\infty}^{y+a_{+}(t-\tau)}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x.

Passing to the limit as ϵ0+\epsilon\to 0^{+} in (3.7) and using the above two inequalities, we obtain the second inequality in (3.3).

Consider the case for aa_{-}. When tτt-\tau and ϵ\epsilon are sufficiently small, by the continuity of vv, we have |v(x,s)v(y,τ)|<ϵ0|v(x,s)-v(y,\tau)|<\epsilon_{0} for s(τϵ,t+ϵ)s\in(\tau-\epsilon,t+\epsilon) and x(y+a(sτ),y+a(sτ)+ϵ)x\in(y+a_{-}(s-\tau),y+a_{-}(s-\tau)+\epsilon), which implies v(x,s)a=v(x,s)v(y,τ)+ϵ00v(x,s)-a_{-}=v(x,s)-v(y,\tau)+\epsilon_{0}\geq 0. Hence, the second term in (3.6) is negative due to θϵ0\theta^{\prime}_{\epsilon}\leq 0 and χϵ0\chi_{\epsilon}\geq 0. Therefore,

(3.8) (τϵτ+tt+ϵ)y+a(sτ)+ϵχϵ(s)θϵ(xa(sτ))vx2(x,s)dxds0.\displaystyle\left(\int_{\tau-\epsilon}^{\tau}+\int_{t}^{t+\epsilon}\right)\int_{-\infty}^{y+a_{-}(s-\tau)+\epsilon}\chi^{\prime}_{\epsilon}(s)\theta_{\epsilon}(x-a_{-}(s-\tau))v_{x}^{2}(x,s)\mathop{}\!\mathrm{d}x\mathop{}\!\mathrm{d}s\geq 0.

Similarly to the case for a+a_{+}, passing to the limit as ϵ0+\epsilon\to 0^{+} in (3.8), we obtain the first inequality in (3.3).

One can show (3.4) by applying a similar argument as in the above proof of (3.3) by using a=CTa_{-}=-C_{T} and a+=CTa_{+}=C_{T} instead. It is worth noting that CTvCT-C_{T}\leq v\leq C_{T}, so the proof will not need the smallness condition for tτt-\tau because we do not need to apply the continuity of vv as in the above proof of (3.3).

Next, we apply the above lemma to prove the following uniqueness theorem.

Theorem 3.2 (Uniqueness of characteristics and conservative solutions).

Let (v,ν)(v,\nu) be a conservative solution to the generalized framework (1.4)-(1.6) in the sense of Definition 1.2 with initial datum (u¯,μ¯)𝒟(\bar{u},\bar{\mu})\in\mathcal{D}. Then, there exists a unique characteristic y1(α,t)y_{1}(\alpha,t) satisfying

(3.9) ty1(α,t)=v(y1(α,t),t),y1(α,0)=x¯(α),\displaystyle\frac{\partial}{\partial t}y_{1}(\alpha,t)=v(y_{1}(\alpha,t),t),\quad y_{1}(\alpha,0)=\bar{x}(\alpha),

and

(3.10) ν(t)((,y1(α,t)))αx¯(α)ν(t)((,y1(α,t)]),\nu(t)((-\infty,y_{1}(\alpha,t)))\leq\alpha-\bar{x}(\alpha)\leq\nu(t)((-\infty,y_{1}(\alpha,t)]),

for any α\alpha\in\mathbb{R} and a.e tt\in\mathbb{R}, where x¯(α)\bar{x}(\alpha) is defined by (2.9). The uniqueness of characteristics and conservative solutions follows, i.e., (v,ν)=(u,μ)(v,\nu)=(u,\mu), where (u,μ)(u,\mu) is defined by (2.27) and (2.28)\eqref{eq:measuresolutionmu}.

Proof.

We will separate the proof into three steps.

Step 1. We use a similar definition for x1(β,t)x_{1}(\beta,t) to that of x¯(α)\bar{x}(\alpha): for any β\beta\in\mathbb{R} and tt\in\mathbb{R},

(3.11) x1(β,t)+ν(t)((,x1(β,t)))βx1(β,t)+ν(t)((,x1(β,t)]),\displaystyle x_{1}(\beta,t)+\nu(t)((-\infty,x_{1}(\beta,t)))\leq\beta\leq x_{1}(\beta,t)+\nu(t)((-\infty,x_{1}(\beta,t)]),

which is actually the inverse function of the sum of identity and cumulative energy distribution: x+ν(t)((,x])x+\nu(t)((-\infty,x]). For a.e. tt\in\mathbb{R}, we have dν(t)=vx2(x,t)dx\mathop{}\!\mathrm{d}\nu(t)=v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x, so at these times, Definition (3.11) is equivalent to

(3.12) x1(β,t)+(,x1(β,t))vx2(y,t)dy=β,\displaystyle x_{1}(\beta,t)+\int_{(-\infty,x_{1}(\beta,t))}v_{x}^{2}(y,t)\mathop{}\!\mathrm{d}y=\beta,

which implies

(3.13) |x1β(β,t)|=11+vx2(x1(β,t),t)1,β, a.e. t\displaystyle|x_{1\beta}(\beta,t)|=\frac{1}{1+v_{x}^{2}(x_{1}(\beta,t),t)}\leq 1,\quad\forall\beta\in\mathbb{R},\textrm{ a.e. }t\in\mathbb{R}

and

(3.14) |βv(x1(β,t),t)|=|vx(x1(β,t),t)1+vx2(x1(β,t),t)|12,β, a.e. t.\displaystyle|\partial_{\beta}v(x_{1}(\beta,t),t)|=\left|\frac{v_{x}(x_{1}(\beta,t),t)}{1+v_{x}^{2}(x_{1}(\beta,t),t)}\right|\leq\frac{1}{2},\quad\forall\beta\in\mathbb{R},\textrm{ a.e. }t\in\mathbb{R}.

Next, we prove tx1(β,t)t\mapsto x_{1}(\beta,t) is Lipschitz. First, consider τ,t[T,T]\tau,t\in[-T,T] and τ<t\tau<t such that ν\nu is absolutely continuous at τ,t\tau,t. Let y=x1(β,τ)y=x_{1}(\beta,\tau). It follows from the first inequality of (3.4) that

yCT(tτ)+(,yCT(tτ))vx2(x,t)dx\displaystyle y-C_{T}(t-\tau)+\int_{(-\infty,y-C_{T}(t-\tau))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x x1(β,τ)+(,x1(β,τ))vx2(x,τ)dx=β\displaystyle\leq x_{1}(\beta,\tau)+\int_{(-\infty,x_{1}(\beta,\tau))}v_{x}^{2}(x,\tau)\mathop{}\!\mathrm{d}x=\beta
=x1(β,t)+(,x1(β,t))vx2(x,t)dx,\displaystyle=x_{1}(\beta,t)+\int_{(-\infty,x_{1}(\beta,t))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x,

which implies yCT(tτ)x1(β,t)y-C_{T}(t-\tau)\leq x_{1}(\beta,t), i.e. x1(β,τ)x1(β,t)CT(tτ)x_{1}(\beta,\tau)-x_{1}(\beta,t)\leq C_{T}(t-\tau). Similarly, it follows from the second inequality of (3.4) that

y+CT(tτ)+(,y+CT(tτ))vx2(x,t)dx\displaystyle y+C_{T}(t-\tau)+\int_{(-\infty,y+C_{T}(t-\tau))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x x1(β,τ)+(,x1(β,τ))vx2(x,τ)dx=β\displaystyle\geq x_{1}(\beta,\tau)+\int_{(-\infty,x_{1}(\beta,\tau))}v_{x}^{2}(x,\tau)\mathop{}\!\mathrm{d}x=\beta
=x1(β,t)+(,x1(β,t))vx2(x,t)dx,\displaystyle=x_{1}(\beta,t)+\int_{(-\infty,x_{1}(\beta,t))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x,

which implies y+CT(tτ)x1(β,t)y+C_{T}(t-\tau)\geq x_{1}(\beta,t), i.e. x1(β,t)x1(β,τ)CT(tτ)x_{1}(\beta,t)-x_{1}(\beta,\tau)\leq C_{T}(t-\tau). Combining the above two results yields

|x1(β,t)x1(β,τ)|CT|tτ|,|x_{1}(\beta,t)-x_{1}(\beta,\tau)|\leq C_{T}|t-\tau|,

for any tt, τ\tau such that ν\nu is absolutely continuous. With the above results, when ν\nu is not absolutely continuous at tt and/or τ\tau, we only need to show the continuity of map tx1(β,t)t\mapsto x_{1}(\beta,t). Actually, it follows from Definition (3.11) that

(3.15) x1(β,t)+ν(t)((,x1(β,t)))βx1(β,s)+ν(s)((,x1(β,s)]),\displaystyle x_{1}(\beta,t)+\nu(t)((-\infty,x_{1}(\beta,t)))\leq\beta\leq x_{1}(\beta,s)+\nu(s)((-\infty,x_{1}(\beta,s)]),

which implies

x1(β,s)x1(β,t)ν(t)((,x1(β,t)))ν(s)((,x1(β,s)])x_{1}(\beta,s)-x_{1}(\beta,t)\geq\nu(t)((-\infty,x_{1}(\beta,t)))-\nu(s)((-\infty,x_{1}(\beta,s)])

for any ss, tt\in\mathbb{R}. Fix a tt\in\mathbb{R}. We prove the continuity of x1(β,)x_{1}(\beta,\cdot) at tt by a contradiction argument. Seeking for a contradiction, we assume that there exists a sequence skts_{k}\to t as kk\to\infty but limkx1(β,sk)=A<x1(β,t)\lim_{k\to\infty}x_{1}(\beta,s_{k})=A<x_{1}(\beta,t). Due to Remark 1.1, we have

0>\displaystyle 0> limkx1(β,sk)x1(β,t)ν(t)((,x1(β,t)))lim supkν(sk)((,x1(β,sk)])\displaystyle\lim_{k\to\infty}x_{1}(\beta,s_{k})-x_{1}(\beta,t)\geq\nu(t)((-\infty,x_{1}(\beta,t)))-\limsup_{k\to\infty}\nu(s_{k})((-\infty,x_{1}(\beta,s_{k})])
=\displaystyle= ν(t)((,x1(β,t)))lim supk[ν(sk)()ν(sk)((x1(β,sk),+))]\displaystyle\nu(t)((-\infty,x_{1}(\beta,t)))-\limsup_{k\to\infty}[\nu(s_{k})(\mathbb{R})-\nu(s_{k})((x_{1}(\beta,s_{k}),+\infty))]
=\displaystyle= ν(t)((,x1(β,t)))μ¯()+lim infkν(sk)((x1(β,sk),+)).\displaystyle\nu(t)((-\infty,x_{1}(\beta,t)))-\bar{\mu}(\mathbb{R})+\liminf_{k\to\infty}\nu(s_{k})((x_{1}(\beta,s_{k}),+\infty)).

For kk big enough, we have x1(β,sk)<x1(β,t)+A2<x1(β,t)x_{1}(\beta,s_{k})<\frac{x_{1}(\beta,t)+A}{2}<x_{1}(\beta,t), and combining [12, Theorem 1.40] yields a contradiction:

0>\displaystyle 0> ν(t)((,x1(β,t)))μ¯()+lim infkν(sk)((x1(β,t)+A2,+))\displaystyle\nu(t)((-\infty,x_{1}(\beta,t)))-\bar{\mu}(\mathbb{R})+\liminf_{k\to\infty}\nu(s_{k})\left(\left(\frac{x_{1}(\beta,t)+A}{2},+\infty\right)\right)
\displaystyle\geq ν(t)((,x1(β,t)))μ¯()+ν(t)((x1(β,t)+A2,+))0,\displaystyle\nu(t)((-\infty,x_{1}(\beta,t)))-\bar{\mu}(\mathbb{R})+\nu(t)\left(\left(\frac{x_{1}(\beta,t)+A}{2},+\infty\right)\right)\geq 0,

which is a contradiction. On the other hand, if there is a sequence s~kt\tilde{s}_{k}\to t as kk\to\infty but limkx1(β,s~k)=B>x1(β,t)\lim_{k\to\infty}x_{1}(\beta,\tilde{s}_{k})=B>x_{1}(\beta,t), we can also obtain a contradiction by a similar argument. Combining the above arguments, we know x1(β,t)x_{1}(\beta,t) is continuous with respect to time tt.

Step 2. In this step, we are going to prove (3.9) and (3.10).

Consider the integral equation:

(3.16) β1(t)=α+0tv(x1(β1(s),s),s)ds.\beta_{1}(t)=\alpha+\int_{0}^{t}v(x_{1}(\beta_{1}(s),s),s)\mathop{}\!\mathrm{d}s.

For tt\in\mathbb{R}, due to (3.14) and Banach fixed point theorem, there always exists a unique global solution β1(t)\beta_{1}(t) to (3.16). Define

(3.17) y1(α,t)=x1(β1(t),t),α,t.\displaystyle y_{1}(\alpha,t)=x_{1}(\beta_{1}(t),t),\quad\alpha\in\mathbb{R},~{}~{}t\in\mathbb{R}.

Then, combining (2.9) and (3.11) yields the initial datum: y1(α,0)=x¯(α)y_{1}(\alpha,0)=\bar{x}(\alpha). From (3.11), (3.16), and (3.17), the following relation holds: for a.e. tt\in\mathbb{R},

(3.18) y1(α,t)+ν(t)((,y1(α,t)))α+0tv(y1(α,s),s)dsy1(α,t)+ν(t)((,y1(α,t)]).\displaystyle y_{1}(\alpha,t)+\nu(t)((-\infty,y_{1}(\alpha,t)))\leq\alpha+\int_{0}^{t}v(y_{1}(\alpha,s),s)\mathop{}\!\mathrm{d}s\leq y_{1}(\alpha,t)+\nu(t)((-\infty,y_{1}(\alpha,t)]).

Fix α\alpha\in\mathbb{R}. From Step 1, we know that y1(α,t)y_{1}(\alpha,t) is differentiable for a.e. t>0t>0. In the following, we only consider those tt\in\mathbb{R} such that y1(α,)y_{1}(\alpha,\cdot) is differentiable and ν(t)\nu(t) is absolutely continuous with dν(t)=vx2(x,t)dx\mathop{}\!\mathrm{d}\nu(t)=v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x. We will prove (3.9) by a contradiction argument. Seeking for a contradiction, we assume that there exists a τ\tau\in\mathbb{R} such that dν(τ)=vx2(x,τ)dx\mathop{}\!\mathrm{d}\nu(\tau)=v_{x}^{2}(x,\tau)\mathop{}\!\mathrm{d}x and y1(α,τ)y_{1}(\alpha,\tau) is differentiable but ty1(α,τ)v(y1(α,τ),τ)\partial_{t}y_{1}(\alpha,\tau)\neq v(y_{1}(\alpha,\tau),\tau). Then, there exists some ϵ0>0\epsilon_{0}>0 such that either one of the following two cases holds:

  1. Case 1:
    (3.19) ty1(α,τ)v(y1(α,τ),τ)2ϵ0.\displaystyle\partial_{t}y_{1}(\alpha,\tau)\leq v(y_{1}(\alpha,\tau),\tau)-2\epsilon_{0}.
  2. Case 2:
    (3.20) ty1(α,τ)v(y1(α,τ),τ)+2ϵ0.\displaystyle\partial_{t}y_{1}(\alpha,\tau)\geq v(y_{1}(\alpha,\tau),\tau)+2\epsilon_{0}.

Next, we derive a contradiction from Case 1. Let y=y1(α,τ)y=y_{1}(\alpha,\tau) and a=v(y,τ)ϵ0a_{-}=v(y,\tau)-\epsilon_{0}. Then, the first inequality in (3.3) holds for tτ>0t-\tau>0 small enough. Due to (3.19), we have

y1(α,t)<y1(α,τ)+[v(y1(α,τ),τ)ϵ0](tτ)=y+a(tτ)y_{1}(\alpha,t)<y_{1}(\alpha,\tau)+[v(y_{1}(\alpha,\tau),\tau)-\epsilon_{0}](t-\tau)=y+a_{-}(t-\tau)

for tτt-\tau small enough. Combining the first inequality in (3.3) yields

(3.21) β1(t)\displaystyle\beta_{1}(t) =x1(β1(t),t)+(,x1(β1(t),t))vx2(x,t)dx=y1(α,t)+(,y1(α,t))vx2(x,t)dx\displaystyle=x_{1}(\beta_{1}(t),t)+\int_{(-\infty,x_{1}(\beta_{1}(t),t))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x=y_{1}(\alpha,t)+\int_{(-\infty,y_{1}(\alpha,t))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x
<y+a(tτ)+(,y+a(tτ))vx2(x,t)dx\displaystyle<y+a_{-}(t-\tau)+\int_{(-\infty,y+a_{-}(t-\tau))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x
y1(α,τ)+[v(y1(α,τ),τ)ϵ0](tτ)+(,y1(α,τ))vx2(x,τ)dx\displaystyle\leq y_{1}(\alpha,\tau)+[v(y_{1}(\alpha,\tau),\tau)-\epsilon_{0}](t-\tau)+\int_{(-\infty,y_{1}(\alpha,\tau))}v_{x}^{2}(x,\tau)\mathop{}\!\mathrm{d}x
=β1(τ)+[v(y1(α,τ),τ)ϵ0](tτ).\displaystyle=\beta_{1}(\tau)+[v(y_{1}(\alpha,\tau),\tau)-\epsilon_{0}](t-\tau).

Therefore,

ϵ0<v(y1(α,τ),τ)β1(t)β1(τ)tτ=v(x1(β1(τ),τ),τ)β1(t)β1(τ)tτ.\epsilon_{0}<v(y_{1}(\alpha,\tau),\tau)-\frac{\beta_{1}(t)-\beta_{1}(\tau)}{t-\tau}=v(x_{1}(\beta_{1}(\tau),\tau),\tau)-\frac{\beta_{1}(t)-\beta_{1}(\tau)}{t-\tau}.

Passing to the limit as tτt\to\tau in the above inequality, and using (3.16), we obtain ϵ00\epsilon_{0}\leq 0, which is a contradiction.

For Case 2, we can use the second inequality in (3.3) to derive a contradiction in a similar manner.

From (3.9), we have y1(α,t)=x¯(α)+0tv(y1(α,s),s)dsy_{1}(\alpha,t)=\bar{x}(\alpha)+\int_{0}^{t}v(y_{1}(\alpha,s),s)\mathop{}\!\mathrm{d}s, and hence, (3.10) follows directly from (3.18).

Step 3. Uniqueness. Notice that vv satisfies the Hunter-Saxton equation (1.1) in Lloc2(×)L_{loc}^{2}(\mathbb{R}\times\mathbb{R}), and hence, vv satisfies (1.1) classically almost everywhere in ×\mathbb{R}\times\mathbb{R}. Since we have already proved that there exists a characteristics function y1(α,t)y_{1}(\alpha,t) satisfying (3.9) and (3.10), we also have

2t2y1(α,t)=ddtv(y1(α,t),t)=(vt+vvx)(y1(α,t),t)=12(,y1(α,t))vx2(x,t)dx=12(αx¯(α)).\displaystyle\frac{\partial^{2}}{\partial t^{2}}y_{1}(\alpha,t)=\frac{\mathop{}\!\mathrm{d}}{\mathop{}\!\mathrm{d}t}v(y_{1}(\alpha,t),t)=(v_{t}+vv_{x})(y_{1}(\alpha,t),t)=\frac{1}{2}\int_{(-\infty,y_{1}(\alpha,t))}v_{x}^{2}(x,t)\mathop{}\!\mathrm{d}x=\frac{1}{2}\left(\alpha-\bar{x}(\alpha)\right).

Since

ty1(α,0)=u¯(x¯(α)),\frac{\partial}{\partial t}y_{1}(\alpha,0)=\bar{u}(\bar{x}(\alpha)),

we have

y1(α,t)=x¯(α)+u¯(x¯(α))t+t24(αx¯(α))=y(α,t),y_{1}(\alpha,t)=\bar{x}(\alpha)+\bar{u}(\bar{x}(\alpha))t+\frac{t^{2}}{4}(\alpha-\bar{x}(\alpha))=y(\alpha,t),

where y(α,t)y(\alpha,t) is defined by (2.12). By Definition (2.27) of uu, we have

(3.22) v(y(α,t),t)=ty(α,t)=u¯(x¯(α))+t2(αx¯(α))=u(y(α,t),t).\displaystyle v(y(\alpha,t),t)=\frac{\partial}{\partial t}y(\alpha,t)=\bar{u}(\bar{x}(\alpha))+\frac{t}{2}(\alpha-\bar{x}(\alpha))=u(y(\alpha,t),t).

Acknowledgements Y. Gao is supported by the Start-up fund from The Hong Kong Polytechnic University. T. K. Wong is partially supported by the HKU Seed Fund for Basic Research under the project code 201702159009, the Start-up Allowance for Croucher Award Recipients, and Hong Kong General Research Fund (GRF) grant “Solving Generic Mean Field Type Problems: Interplay between Partial Differential Equations and Stochastic Analysis” with project number 17306420.

Appendix A Some useful facts from real analysis

In this appendix, we state and prove three useful lemmas from real analysis. All of them are fundamental and somewhat classical. For readers’ convenience, we also provide their proofs here.

Lemma A.1.

The following two statements holds:

  1. (i)

    The real line \mathbb{R} cannot be written as the union of uncountably many disjoint subsets with positive measures.

  2. (ii)

    Let X:X:\mathbb{R}\to\mathbb{R} be an absolutely continuous function satisfying Xξ(ξ)>0X_{\xi}(\xi)>0 for a.e. ξ\xi\in\mathbb{R}. Then, there exists a unique absolutely continuous inverse of XX.

Proof.

(i) We prove this by a contradiction argument. Let Λ\Lambda be an uncountable index set and {Aα}αΛ\{A_{\alpha}\}_{\alpha\in\Lambda} be a family of uncountable disjoint subsets of \mathbb{R} with (Aα)>0\mathcal{L}(A_{\alpha})>0 such that

=αΛAα.\mathbb{R}=\cup_{\alpha\in\Lambda}A_{\alpha}.

Because (Aα)>0\mathcal{L}(A_{\alpha})>0, there must be some integer nn such that (Aα[n,n+1])>0\mathcal{L}(A_{\alpha}\cap[n,n+1])>0. Since Λ\Lambda is an uncountable set, there must be some integer n0n_{0} such that

Λ0={αΛ:(Aα[n0,n0+1])>0}\Lambda_{0}=\{\alpha\in\Lambda:~{}~{}\mathcal{L}(A_{\alpha}\cap[n_{0},n_{0}+1])>0\}

is an uncountable set. Then there exists a positive integer m>0m>0 such that the set

Λm={αΛ:(Aα[n0,n0+1])>1m}\Lambda_{m}=\left\{\alpha\in\Lambda:~{}~{}\mathcal{L}(A_{\alpha}\cap[n_{0},n_{0}+1])>\frac{1}{m}\right\}

is uncountable. Since the sets (Aα[n0,n0+1])\mathcal{L}(A_{\alpha}\cap[n_{0},n_{0}+1]) for αΛm\alpha\in\Lambda_{m} are disjoint with each other, this contradicts with ([n0,n0+1])=1\mathcal{L}([n_{0},n_{0}+1])=1.

(ii) We prove this on arbitrary interval [a,b][a,b]. Clearly XX is continuous and strictly increasing on [a,b][a,b] and hence, it has a continuous and strictly increasing inverse ZZ. Let

A={ξ[a,b]:0<Xξ(ξ)<},B=[0,1]A,A=\{\xi\in[a,b]:~{}~{}0<X_{\xi}(\xi)<\infty\},\quad B=[0,1]\setminus A,

and

C=X(A),D=X(B).C=X(A),\quad D=X(B).

We have (B)=0\mathcal{L}(B)=0 and CD=[X(a),X(b)]C\cup D=[X(a),X(b)]. Since XX is absolutely continuous, it satisfies Luzin N property and (X(B))=(D)=0\mathcal{L}(X(B))=\mathcal{L}(D)=0.

We know that ZZ has a finite positive derivative at each point of CC. Let ECE\subset C satisfying (E)=0\mathcal{L}(E)=0. Then

E=n=1En,E=\cup_{n=1}^{\infty}E_{n},

where En={xE:0<Zx(x)n}E_{n}=\{x\in E:~{}~{}0<Z_{x}(x)\leq n\}. Then, we have (En)=0\mathcal{L}(E_{n})=0 and

(Z(E))n=1(Z(En))n=1n(En)=0,\mathcal{L}(Z(E))\leq\sum_{n=1}^{\infty}\mathcal{L}(Z(E_{n}))\leq\sum_{n=1}^{\infty}n\mathcal{L}(E_{n})=0,

where we have used (En)n(En)\mathcal{L}(E_{n})\leq n\mathcal{L}(E_{n}); see [16, Lemma 6.3]. Hence, the inverse ZZ on [X(a),X(b)][X(a),X(b)] is a continuous function of bounded variation satisfying the Luzin N property, which means that ZZ is absolutely continuous.

We have the following lemma for push-forward measures:

Lemma A.2.

Let X:X:\mathbb{R}\to\mathbb{R} be a continuous increasing surjective function. Define two pseudo-inverse functions of XX by

Z1(x)=inf{ξ:X(ξ)=x},Z2(x)=sup{ξ:X(ξ)=x}.Z_{1}(x)=\inf\{\xi:~{}~{}X(\xi)=x\},\quad Z_{2}(x)=\sup\{\xi:~{}~{}X(\xi)=x\}.

Define

App={ξ:Xξ(ξ)=0,Z1(X(ξ))<Z2(X(ξ))},Asc={ξ:Xξ(ξ)=0,Z1(X(ξ))=Z2(X(ξ))},A^{pp}=\{\xi:~{}~{}X_{\xi}(\xi)=0,~{}~{}Z_{1}(X(\xi))<Z_{2}(X(\xi))\},\quad A^{sc}=\{\xi:~{}~{}X_{\xi}(\xi)=0,~{}~{}Z_{1}(X(\xi))=Z_{2}(X(\xi))\},

and

B={ξ:Xξ(ξ)>0}.B=\{\xi:~{}~{}X_{\xi}(\xi)>0\}.

Here, AppA^{pp} is defined in point-wise sense, and AscA^{sc} and BB are defined in a.e. sense.

Let g0L1()g\geq 0\in L^{1}(\mathbb{R}) . Consider the measure

μ=X#(gdξ).\mu=X\#(g\mathop{}\!\mathrm{d}\xi).

Let g1=g1Bg_{1}=g\cdot 1_{B}, g2=g1Appg_{2}=g\cdot 1_{A^{pp}} and g3=g1Ascg_{3}=g\cdot 1_{A^{sc}}, here 1App1_{A^{pp}}, 1Asc1_{A^{sc}} and 1B1_{B} are characteristic functions on AppA^{pp}, AscA^{sc} and BB respectively. Then the following statements hold:

  1. (i)

    (X(AppAsc))=0\mathcal{L}(X(A^{pp}\cup A^{sc}))=0, where \mathcal{L} is the Lebesgue measure.

  2. (ii)

    The absolutely continuous part of μ\mu is given by X#(g1dξ)X\#(g_{1}\mathop{}\!\mathrm{d}\xi).

  3. (iii)

    The set X(App)X(A^{pp}) is a countable set, and the pure point part of μ\mu is given by X#(g2dξ)X\#(g_{2}\mathop{}\!\mathrm{d}\xi).

  4. (iv)

    The singular continuous part of μ\mu is given by X#(g3dξ)X\#(g_{3}\mathop{}\!\mathrm{d}\xi).

Proof.

(i) It follows from the change of variable formula (see [17, Theorem 4] for instance) that

(A.1) 1X(AppAsc)(x)dx=1AppAsc(ξ)Xξ(ξ)dξ=0.\int_{\mathbb{R}}1_{X(A^{pp}\cup A^{sc})}(x)\mathop{}\!\mathrm{d}x=\int_{\mathbb{R}}1_{A^{pp}\cup A^{sc}}(\xi)X_{\xi}(\xi)\mathop{}\!\mathrm{d}\xi=0.

(ii) We prove that X#(g1dξ)X\#(g_{1}\mathop{}\!\mathrm{d}\xi)\ll\mathcal{L}. Let EE\subset\mathbb{R} satisfy (E)=0\mathcal{L}(E)=0. Then

[X#(g1dξ)](E)=X1(E)g1(ξ)dξ=BX1(E)g(ξ)dξ.[X\#(g_{1}\mathop{}\!\mathrm{d}\xi)](E)=\int_{X^{-1}(E)}g_{1}(\xi)\mathop{}\!\mathrm{d}\xi=\int_{B\cap X^{-1}(E)}g(\xi)\mathop{}\!\mathrm{d}\xi.

Hence, we only need to show (BX1(E))=(Z1(X(B)E))=0\mathcal{L}(B\cap X^{-1}(E))=\mathcal{L}(Z_{1}(X(B)\cap E))=0. This means Z1Z_{1} has Luzin N property on X(B)X(B). Set B~=X(B)\tilde{B}=X(B). Since

B~E=n=1(B~nE),where B~n:={xB~:0<Z1x(x)n},\tilde{B}\cap E=\cup_{n=1}^{\infty}(\tilde{B}_{n}\cap E),\quad\mbox{where }\tilde{B}_{n}:=\left\{x\in\tilde{B}:~{}~{}0<Z_{1x}(x)\leq n\right\},

and

(Z1(B~nE))n(B~nE)=0,\mathcal{L}(Z_{1}(\tilde{B}_{n}\cap E))\leq n\mathcal{L}(\tilde{B}_{n}\cap E)=0,

we have

(Z1(B~E))n=1(Z1(B~nE))=0.\mathcal{L}(Z_{1}(\tilde{B}\cap E))\leq\sum_{n=1}^{\infty}\mathcal{L}(Z_{1}(\tilde{B}_{n}\cap E))=0.

Hence, X#(g1dξ)X\#(g_{1}\mathop{}\!\mathrm{d}\xi)\ll\mathcal{L}. Due to (i), we have μX#(g1dξ)\mu-X\#(g_{1}\mathop{}\!\mathrm{d}\xi)\perp\mathcal{L}, and hence, X#(g1dξ)X\#(g_{1}\mathop{}\!\mathrm{d}\xi) is the absolutely continuous part of μ\mu.

(iii) Suppose AppA^{pp}\neq\emptyset. Let xx, yX(App)y\in X(A^{pp}) and xyx\neq y. By the definition of X(App)X(A^{pp}), we have

Z1(x)<Z2(x),Z1(y)<Z2(y).Z_{1}(x)<Z_{2}(x),\quad Z_{1}(y)<Z_{2}(y).

For each xX(App)x\in X(A^{pp}), we choose a rational number in [Z1(x),Z2(x)][Z_{1}(x),Z_{2}(x)] to stand for it. Due to the increasing nature of XX, we have [Z1(x),Z2(x)][Z1(y),Z2(y)]=[Z_{1}(x),Z_{2}(x)]\cap[Z_{1}(y),Z_{2}(y)]=\emptyset, which implies different points in X(App)X(A^{pp}) correspond to different rational numbers. Therefore, the set X(App)X(A^{pp}) is at most countable.

For ξApp\xi\in A^{pp}, denote x=X(ξ)x=X(\xi). Then, we have z1(x)<z2(x)z_{1}(x)<z_{2}(x) and

μ({x})=[X#(gdξ)]({x})=X1({x})g(ξ)dξ=[z1(x),z2(x)]g(ξ)dξ0.\mu(\{x\})=[X\#(g\mathop{}\!\mathrm{d}\xi)](\{x\})=\int_{X^{-1}(\{x\})}g(\xi)\mathop{}\!\mathrm{d}\xi=\int_{[z_{1}(x),z_{2}(x)]}g(\xi)\mathop{}\!\mathrm{d}\xi\geq 0.

Therefore, there is a pure point measure of μ\mu at xx if the above value is strictly positive. Moreover, there is no other pure point parts as for ξAscB\xi\in A^{sc}\cup B, Z1(X(ξ))=Z2(X(ξ))Z_{1}(X(\xi))=Z_{2}(X(\xi)) and the above value must be 0.

(iv) It is direct to see from the definition that μ=X#(g1dξ)+X#(g2dξ)+X#(g3dξ)\mu=X\#(g_{1}\mathop{}\!\mathrm{d}\xi)+X\#(g_{2}\mathop{}\!\mathrm{d}\xi)+X\#(g_{3}\mathop{}\!\mathrm{d}\xi). We have proven in (i) and (ii) that X#(g1dξ)X\#(g_{1}\mathop{}\!\mathrm{d}\xi) is the the absolutely continuous part of μ\mu and X#(g2dξ)X\#(g_{2}\mathop{}\!\mathrm{d}\xi) is the pure point part of μ\mu, hence, by Lebesgue decomposition theorem, the remaining part X#(g3dξ)X\#(g_{3}\mathop{}\!\mathrm{d}\xi) is the singular continuous part of μ\mu. ∎

The last lemma shows that a locally absolutely continuous (i.e., absolutely continuous on any closed and bounded interval of \mathbb{R}) function has to be globally absolutely continuous if its derivative is in Lp()L^{p}(\mathbb{R}) for some p[1,]p\in[1,\infty].

Lemma A.3.

Let 1p1\leq p\leq\infty. Then any locally absolutely continuous function uu on \mathbb{R} with its derivative uxLp()u_{x}\in L^{p}(\mathbb{R}) is globally absolutely continuous.

Proof.

If p=1p=1, it follows from the absolute continuity of Lebesgue integral that uu must be globally absolutely continuous. If p=p=\infty, then uu is globally Lipschitz continuous, and hence, globally absolutely continuous.

Let 1<p<1<p<\infty. For any ϵ>0\epsilon>0, we are going to find some δ>0\delta>0, such that for any disjoint open intervals {(ai,bi)}i=1k\{(a_{i},b_{i})\}_{i=1}^{k} with i=1k|biai|<δ\sum_{i=1}^{k}|b_{i}-a_{i}|<\delta, we have i=1k|u(bi)u(ai)|<ϵ\sum_{i=1}^{k}|u(b_{i})-u(a_{i})|<\epsilon. For this purpose, we define En:={x:n|ux(x)|<n+1}E_{n}:=\{x:n\leq|u_{x}(x)|<n+1\} for all integer n0n\geq 0. Let n0n_{0} be a large enough integer such that n0p12ϵuxLppn_{0}^{p-1}\geq\frac{2}{\epsilon}\|u_{x}\|^{p}_{L^{p}}. Then

i=1k|u(bi)u(ai)|\displaystyle\sum_{i=1}^{k}|u(b_{i})-u(a_{i})| =i=1k|aibiux(x)dx|i=1kaibi|ux(x)|dx\displaystyle=\sum_{i=1}^{k}\left|\int_{a_{i}}^{b_{i}}u_{x}(x)\mathop{}\!\mathrm{d}x\right|\leq\sum_{i=1}^{k}\int_{a_{i}}^{b_{i}}|u_{x}(x)|\mathop{}\!\mathrm{d}x
=n=0n01(i=1k(ai,bi))En|ux(x)|dx+n=n0(i=1k(ai,bi))En|ux(x)|dx\displaystyle=\sum_{n=0}^{n_{0}-1}\int_{\left(\cup_{i=1}^{k}(a_{i},b_{i})\right)\cap E_{n}}|u_{x}(x)|\mathop{}\!\mathrm{d}x+\sum_{n=n_{0}}^{\infty}\int_{\left(\cup_{i=1}^{k}(a_{i},b_{i})\right)\cap E_{n}}|u_{x}(x)|\mathop{}\!\mathrm{d}x
n0δ+n=n0(i=1k(ai,bi))En|ux(x)|dx.\displaystyle\leq n_{0}\delta+\sum_{n=n_{0}}^{\infty}\int_{\left(\cup_{i=1}^{k}(a_{i},b_{i})\right)\cap E_{n}}|u_{x}(x)|\mathop{}\!\mathrm{d}x.

Moreover,

n=n0(i=1k(ai,bi))En|ux(x)|dx\displaystyle\sum_{n=n_{0}}^{\infty}\int_{\left(\cup_{i=1}^{k}(a_{i},b_{i})\right)\cap E_{n}}|u_{x}(x)|\mathop{}\!\mathrm{d}x n=n0En|ux(x)|dxn=n01np1En|ux(x)|pdx\displaystyle\leq\sum_{n=n_{0}}^{\infty}\int_{E_{n}}|u_{x}(x)|\mathop{}\!\mathrm{d}x\leq\sum_{n=n_{0}}^{\infty}\frac{1}{n^{p-1}}\int_{E_{n}}|u_{x}(x)|^{p}\mathop{}\!\mathrm{d}x
1n0p1uxLppϵ2.\displaystyle\leq\frac{1}{n_{0}^{p-1}}\|u_{x}\|^{p}_{L^{p}}\leq\frac{\epsilon}{2}.

Now we choose δ<ϵ2n0\delta<\frac{\epsilon}{2n_{0}}, combining the above two inequalities we have

i=1k|u(bi)u(ai)|<ϵ2+ϵ2=ϵ.\sum_{i=1}^{k}|u(b_{i})-u(a_{i})|<\frac{\epsilon}{2}+\frac{\epsilon}{2}=\epsilon.

In the following example, we are going to use a fat Cantor set to create an initial datum (u¯,μ¯)(\bar{u},\bar{\mu}) such that the singular continuous part of the corresponding energy measure μ(t)\mu(t) is nonzero at t=2t=2.

Example A.1 (Fat Cantor set for AtL,scA_{t}^{L,sc}).

\bullet Fat Cantor set: The fat Cantor set that we use is defined as follows: we consider the set [0,1][0,1], in the first step, we remove the middle open interval with length 14\frac{1}{4} (i.e. [38,58][\frac{3}{8},\frac{5}{8}]) from [0,1][0,1]. At nn-th step, we remove the open sub-intervals of length 14n\frac{1}{4^{n}} from the middle of each of the 2n12^{n-1} remaining intervals. Continuing this procedure, the fat Cantor set EE is defined as the points that are never removed. We have the following well-known results: (i) set EE is closed; (ii) set EE is nowhere dense in [0,1][0,1]; (iii) (E)=12\mathcal{L}(E)=\frac{1}{2}; and (iv) all the endpoints of the removed intervals are dense in set EE.

\bullet A function vanishing exactly on EE: From Now, we use the above fat Cantor set to define a continuous function hh. We choose a sequence of smooth non-negative function hn(x)h_{n}(x) such that each hn(x)h_{n}(x) is positive inside each open interval removed at the nn-th step and hn(x)=0h_{n}(x)=0 outside the intervals removed at the nn-th step. We also assume that supx[0,1]hn(x)=1\sup_{x\in[0,1]}h_{n}(x)=1. Then we define

h(x)=n=1hn(x)n2.h(x)=\sum_{n=1}^{\infty}\frac{h_{n}(x)}{n^{2}}.

As the uniform limit of a sequence of continuous functions, the function hh is continuous. Since hnh_{n}’s are supported in disjoint sets and 0hn10\leq h_{n}\leq 1 for all positive integer nn, we also have 0hπ260\leq h\leq\frac{\pi^{2}}{6}. Since h=0h=0 on the endpoints of all the removed intervals, combing the property (iv) above yields h(x)=0h(x)=0 for all xEx\in E. Moreover, h(x)>0h(x)>0 for x[0,1]\Ex\in[0,1]\backslash E.

\bullet Initial datum (u¯,μ¯)(\bar{u},\bar{\mu}): Finally, our target initial datum in this example is defined by

u¯x(x)={h12(x)1,x[0,1],0,otherwise,\bar{u}_{x}(x)=\left\{\begin{aligned} &h^{\frac{1}{2}}(x)-1,~{}~{}x\in[0,1],\\ &0,~{}~{}\textrm{otherwise},\end{aligned}\right.

and

u¯(x)=(,x)u¯x(η)dη,dμ¯=u¯x2dx.\bar{u}(x)=\int_{(-\infty,x)}\bar{u}_{x}(\eta)\mathop{}\!\mathrm{d}\eta,\quad\mathop{}\!\mathrm{d}\bar{\mu}=\bar{u}_{x}^{2}\mathop{}\!\mathrm{d}x.

Hence, the function u¯Cb()\bar{u}\in C_{b}(\mathbb{R}) and u¯xL2()\bar{u}_{x}\in L^{2}(\mathbb{R}). Since μ¯\bar{\mu} is absolutely continuous, we know that x¯(α)\bar{x}(\alpha) (defined by (2.9)) is strictly increasing and (B0L)=0\mathcal{L}(\mathbb{R}\setminus B_{0}^{L})=0 for B0LB_{0}^{L} given in Proposition 2.1 (iv). From (2.22), we have the derivative of characteristic y(α,t)y(\alpha,t) at t=2t=2:

yα(α,2)=x¯(α)[1+u¯x(x¯(α))]2=x¯(α)h(x¯(α)),αB0Land x¯(α)[0,1].y_{\alpha}(\alpha,2)=\bar{x}^{\prime}(\alpha)\left[1+\bar{u}_{x}(\bar{x}(\alpha))\right]^{2}=\bar{x}^{\prime}(\alpha)h(\bar{x}(\alpha)),\quad\alpha\in B_{0}^{L}~{}\textrm{and $\bar{x}(\alpha)\in[0,1]$}.

and

yα(α,2)=x¯(α)[1+u¯x(x¯(α))]2=x¯(α)>0,αB0Land x¯(α)[0,1].y_{\alpha}(\alpha,2)=\bar{x}^{\prime}(\alpha)\left[1+\bar{u}_{x}(\bar{x}(\alpha))\right]^{2}=\bar{x}^{\prime}(\alpha)>0,\quad\alpha\in B_{0}^{L}~{}\textrm{and $\bar{x}(\alpha)\notin[0,1]$}.

It follows from Remark 2.3 (iii) that x¯(α)=t2t2+4\bar{x}^{\prime}(\alpha)=\frac{t^{2}}{t^{2}+4} and f(α)=4t2+4f(\alpha)=\frac{4}{t^{2}+4} on AtL,scA_{t}^{L,sc}, so A2L,sc={α:yα(α,2)=0}=x¯1(E)A_{2}^{L,sc}=\{\alpha:~{}~{}y_{\alpha}(\alpha,2)=0\}=\bar{x}^{-1}(E) has a positive measure. Hence, the inverse of y(,2)y(\cdot,2) does not have Luzin N property since (y(A2L,sc,2))=0\mathcal{L}(y(A_{2}^{L,sc},2))=0 and (A2L,sc)=(x¯1(E))=2(E)=1>0\mathcal{L}(A_{2}^{L,sc})=\mathcal{L}(\bar{x}^{-1}(E))=2\mathcal{L}(E)=1>0. Moreover,

μsc(2)()=A2L,scf(α)dα=12(x¯1(E))=(E)=12.\mu_{sc}(2)(\mathbb{R})=\int_{A_{2}^{L,sc}}f(\alpha)\mathop{}\!\mathrm{d}\alpha=\frac{1}{2}\mathcal{L}(\bar{x}^{-1}(E))=\mathcal{L}(E)=\frac{1}{2}.

In the above construction, u¯x(x)\bar{u}_{x}(x) is continuous, and 1+u¯x(x)1+\bar{u}_{x}(x) vanishes on EE. We remark that one can have a smooth function u¯\bar{u} such that 1+u¯x(x)1+\bar{u}_{x}(x) vanishes exactly at the closed set EE by Whitney’s theorem [18, Theorem I]. Obviously, the construction above for time t=2t=2 can be adapted to any time t0t\neq 0.

References

  • [1] J. J. Benedetto and W. Czaja. Integration and modern analysis. Springer Science & Business Media, 2010.
  • [2] A. Bressan. Uniqueness of conservative solutions for nonlinear wave equations via characteristics. Bull. Braz. Math. Soc., 47(1):157–169, 2016.
  • [3] A. Bressan, G. Chen, and Q. Zhang. Uniqueness of conservative solutions to the Camassa-Holm equation via characteristics. Discrete Contin. Dyn. Syst., 35(1):25.
  • [4] A. Bressan, G. Chen, and Q. Zhang. Unique conservative solutions to a variational wave equation. Arch. Ration. Mech. Anal., 217(3):1069–1101, 2015.
  • [5] A. Bressan and A. Constantin. Global solutions of the Hunter–Saxton equation. SIAM J. Math. Anal., 37(3):996–1026, 2005.
  • [6] A. Bressan, H. Holden, and X. Raynaud. Lipschitz metric for the Hunter–Saxton equation. J. Math. Pures Appl., 94(1):68–92, 2010.
  • [7] A. Bressan, P. Zhang, and Y. Zheng. Asymptotic variational wave equations. Arch. Ration. Mech. Anal., 183(1):163–185, 2007.
  • [8] J. A. Carrillo, K. Grunert, and H. Holden. A Lipschitz metric for the Hunter–Saxton equation. Comm. Partial Differential Equations, 44(4):309–334, 2019.
  • [9] T. Cieślak and G. Jamróz. Maximal dissipation in Hunter–Saxton equation for bounded energy initial data. Adv. Math., 290:590–613, 2016.
  • [10] C. M. Dafermos. Generalized characteristics and the Hunter–Saxton equation. J. Hyperbolic Differ. Equ., 8(01):159–168, 2011.
  • [11] C. M. Dafermos. Maximal dissipation in equations of evolution. J. Differential Equations, 252(1):567–587, 2012.
  • [12] L. C. Evans and R. F. Gariepy. Measure theory and fine properties of functions. CRC press, 2015.
  • [13] J. K. Hunter and R. Saxton. Dynamics of director fields. SIAM J. Appl. Math., 51(6):1498–1521, 1991.
  • [14] J. K. Hunter and Y. Zheng. On a nonlinear hyperbolic variational equation: I. Global existence of weak solutions. Arch. Ration. Mech. Anal., 129(4):305–353, 1995.
  • [15] J. K. Hunter and Y. Zheng. On a nonlinear hyperbolic variational equation: Ii. the zero-viscosity and dispersion limits. Arch. Ration. Mech. Anal., 129(4):355–383, 1995.
  • [16] S. Saks. Theory of the integral. 1937.
  • [17] J. Serrin and D. E. Varberg. A general chain rule for derivatives and the change of variables formula for the lebesgue integral. Amer. Math. Monthly, 76(5):514–520, 1969.
  • [18] H. Whitney. Analytic extensions of differentiable functions defined in closed sets. Trans. Amer. Math. Soc., 36(1):63–89, 1934.
  • [19] P. Zhang and Y. Zheng. On oscillations of an asymptotic equation of a nonlinear variational wave equation. Asymptot. Anal., 18(3-4):307–327, 1998.
  • [20] P. Zhang and Y. Zheng. On the existence and uniqueness of solutions to an asymptotic equation of a variational wave equation. Acta Math. Sin., 15(1):115–129, 1999.
  • [21] P. Zhang and Y. Zheng. Existence and uniqueness of solutions of an asymptotic equation arising from a variational wave equation with general data. Arch. Ration. Mech. Anal., 155(1):49–83, 2000.