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

The shock formation and optimal regularities of the resulting shock curves for 1-D scalar conservation laws

Yin Huicheng1,∗,  Zhu Lu2, 111Yin Huicheng (huicheng@nju.edu.cn, 05407@njnu.edu.cn) and Zhu Lu (zhulu@hhu.edu.cn) are supported by the NSFC (No.11571177, No.11731007, No.12001162).
1. School of Mathematical Sciences and Institute of Mathematical Sciences,
Nanjing Normal University, Nanjing, 210023, China.
2. College of Science, Hohai University, Nanjing, 210098, China
Abstract

The study on the shock formation and the regularities of the resulting shock surfaces for hyperbolic conservation laws is a basic problem in the nonlinear partial differential equations. In this paper, we are concerned with the shock formation and the optimal regularities of the resulting shock curves for the 1-D conservation law tu+xf(u)=0\partial_{t}u+\partial_{x}f(u)=0 with the smooth initial data u(0,x)=u0(x)u(0,x)=u_{0}(x). If u0(x)C1()u_{0}(x)\in C^{1}(\mathbb{R}) and f(u)C2()f(u)\in C^{2}(\mathbb{R}), it is well-known that the solution uu will blow up on the time T=1ming(x)T^{*}=-\frac{1}{\min{g^{\prime}(x)}} when ming(x)<0\min{g^{\prime}(x)}<0 holds for g(x)=f(u0(x))g(x)=f^{\prime}(u_{0}(x)). Let x0x_{0} be a local minimum point of g(x)g^{\prime}(x) such that g(x0)=ming(x)<0g^{\prime}(x_{0})=\min{g^{\prime}(x)}<0 and g′′(x0)=0g^{\prime\prime}(x_{0})=0, g(3)(x0)>0g^{(3)}(x_{0})>0 (which is called the generic nondegenerate condition), then by Theorem 2 of [11], a weak entropy solution uu together with the shock curve x=φ(t)C2[T,T+ε)x=\varphi(t)\in C^{2}[T^{*},T^{*}+\varepsilon) starting from the blowup point (T,x=x0+g(x0)T)(T^{*},x^{*}=x_{0}+g(x_{0})T^{*}) can be locally constructed. When the generic nondegenerate condition is violated, namely, when x0x_{0} is a local minimum point of g(x)g^{\prime}(x) such that g′′(x0)=g(3)(x0)==g(2k0)(x0)=0g^{\prime\prime}(x_{0})=g^{(3)}(x_{0})=...=g^{(2k_{0})}(x_{0})=0 but g(2k0+1)(x0)>0g^{(2k_{0}+1)}(x_{0})>0 for some k0k_{0}\in\mathbb{N} with k02k_{0}\geq 2; or g(k)(x0)=0g^{(k)}(x_{0})=0 for any kk\in\mathbb{N} and k2k\geq 2, we will study the shock formation and the optimal regularity of the shock curve x=φ(t)x=\varphi(t), meanwhile, some precise descriptions on the behaviors of uu near the blowup point (T,x)(T^{*},x^{*}) are given. Our main aims are to show that: around the blowup point, the shock really appears whether the initial data are degenerate with finite orders or with infinite orders; the optimal regularities of the shock solution and the resulting shock curve have the explicit relations with the degenerate degrees of the initial data.

Keywords: Shock formation, shock curve, entropy condition, Rankine-Hugoniot condition, hyperbolic conservation law

Mathematical Subject Classification 2000: 35L05, 35L72

1 Main result

The study on the blowup and shock formation of smooth solutions to the hyperbolic conservation laws is a basic problem in the nonlinear partial differential equations, which has made much progress for the multi-dimensional cases in recent years (see [2]-[3], [6]-[9], [12]-[15]). In the present paper, we are concerned with the shock formation and the optimal regularities of the resulting shock curves for the 1-D conservation law

{tu+xf(u)=0,(t,x)+×,u(0,x)=u0(x),x,\left\{\begin{aligned} &\partial_{t}u+\partial_{x}f(u)=0,\ (t,x)\in{\mathbb{R}}_{+}\times{\mathbb{R}},\\ &u(0,x)=u_{0}(x),\ x\in{\mathbb{R}},\end{aligned}\right. (1.1)

where f(u)C2()f(u)\in C^{2}(\mathbb{R}) and u0(x)C1()u_{0}(x)\in C^{1}(\mathbb{R}). It is well-known that the C1C^{1} solution uu of (1.1) will blow up at the time T=1ming(x)T^{*}=-\frac{1}{\min{g^{\prime}(x)}} with g(x)=f(u0(x))g(x)=f^{\prime}(u_{0}(x)) and minxg(x)<0\min_{x\in\mathbb{R}}{g^{\prime}(x)}<0. If we further assume g(x)L()Cp()g(x)\in L^{\infty}(\mathbb{R})\cap C^{p}(\mathbb{R}) with p4p\geq 4, and let x0x_{0} be a local minimum point of g(x)g^{\prime}(x) such that

g(x0)=minxg(x)<0,g′′(x0)=0,g(3)(x0)>0,\displaystyle g^{\prime}(x_{0})=\min_{x\in{\mathbb{R}}}{g^{\prime}(x)}<0,\quad g^{\prime\prime}(x_{0})=0,\quad g^{(3)}(x_{0})>0, (1.2)

which is called the generic nondegenerate condition in [1], then by Theorem 2 of [11], a weak entropy solution uu of (1.1) together with the shock curve x=φ(t)x=\varphi(t) starting from the blowup point (T,x=x0+g(x0)T)(T^{*},x^{*}=x_{0}+g(x_{0})T^{*}) can be locally obtained as follows:

(i)

φ(t)Cp(T,T+ε)Cp2[T,T+ε).\displaystyle\varphi(t)\in C^{p}(T^{*},T^{*}+\varepsilon)\cap C^{\frac{p}{2}}[T^{*},T^{*}+\varepsilon). (1.3)

(ii) In some part of the neighbourhood of (T,x)(T^{*},x^{*}) near x=φ(t)x=\varphi(t), for tTt\geq T^{*} and xφ(t)x\not=\varphi(t),

{|u(t,x)u(T,x)|C((tT)3+(xx)2)16,|tu(t,x)|C((tT)3+(xx)2)16,|xu(t,x)|C((tT)3+(xx)2)13,|x2u(t,x)|C((tT)3+(xx)2)56.\left\{\begin{aligned} &|u(t,x)-u(T^{*},x^{*})|\leq C((t-T^{*})^{3}+(x-x^{*})^{2})^{\frac{1}{6}},\\ &|\partial_{t}u(t,x)|\leq\frac{C}{((t-T^{*})^{3}+(x-x^{*})^{2})^{\frac{1}{6}}},\\ &|\partial_{x}u(t,x)|\leq\frac{C}{((t-T^{*})^{3}+(x-x^{*})^{2})^{\frac{1}{3}}},\\ &|\partial_{x}^{2}u(t,x)|\leq\frac{C}{((t-T^{*})^{3}+(x-x^{*})^{2})^{\frac{5}{6}}}.\end{aligned}\right. (1.4)

When the generic nondegenerate condition (1.2) is violated, namely, if x0x_{0} is a local minimum point of g(x)g^{\prime}(x) such that

{g(x)L()C2k+2()for k with k2,g(x0)=minxg(x)<0,g′′(x0)=g(3)(x0)==g(2k)(x0)=0,g(2k+1)(x0)>0,\left\{\begin{aligned} &g(x)\in L^{\infty}(\mathbb{R})\cap C^{2k+2}(\mathbb{R})\quad\text{for $k\in\mathbb{N}$ with $k\geq 2$,}\\ &g^{\prime}(x_{0})=\min_{x\in{\mathbb{R}}}{g^{\prime}(x)}<0,\quad g^{\prime\prime}(x_{0})=g^{(3)}(x_{0})=...=g^{(2k)}(x_{0})=0,\quad g^{(2k+1)}(x_{0})>0,\end{aligned}\right. (1.5)

or

{g(x)L()C(),g(x0)=minxg(x)<0,g(k)(x0)=0for any k and k2,\left\{\begin{aligned} &g(x)\in L^{\infty}(\mathbb{R})\cap C^{\infty}(\mathbb{R}),\\ &g^{\prime}(x_{0})=\min_{x\in{\mathbb{R}}}{g^{\prime}(x)}<0,\quad g^{(k)}(x_{0})=0\quad\text{for any $k\in\mathbb{N}$ and $k\geq 2$},\end{aligned}\right. (1.6)

we will study the shock formation and the optimal regularity of the resulting shock x=φ(t)x=\varphi(t) from the blowup point (T,x)(T^{*},x^{*}), meanwhile, some precise descriptions on the behaviors of the solution uu around the blowup point (T,x)(T^{*},x^{*}) (rather than only some part near the shock curve) will be given.

Without loss of generality and for convenience, we set x0=0x_{0}=0 in (1.5) and (1.6). In addition, under condition (1.5), near x0=0x_{0}=0 we assume

g(x)=x+x2k+1+r(x),\displaystyle g(x)=-x+x^{2k+1}+r(x), (1.7)

where r(x)C2k+2r(x)\in C^{2k+2} satisfies that r(j)(x)=O(x2kj+2)r^{(j)}(x)=O(x^{2k-j+2}) for 0j2k+20\leq j\leq 2k+2; under condition (1.6), we choose a class of initial data

g(x)=x+e|x|p(xp+r0(x)),\displaystyle{g(x)=-x+e^{-|x|^{-p}}\left(\frac{x}{p}+r_{0}(x)\right),} (1.8)

where p>0p>0 and r0(x)CLr_{0}(x)\in C^{\infty}\cap L^{\infty} with

r0(j)(x)={O(x2j),j=0,1,2,O(1),j3{r^{(j)}_{0}(x)=\left\{\begin{array}[]{ll}O(x^{2-j}),&j=0,1,2,\\ O(1),&j\geq 3\end{array}\right.} (1.9)

for xx near 0.

Starting from the blowup point (1,0)(1,0) of (1.1), let the formed shock curve Γ\Gamma be denoted by x=φ(t)x=\varphi(t) if the shock really appears. On the left hand side and right hand side of Γ\Gamma for t1t\geq 1, the weak entropy solution uu is represented by uu_{-} and u+u_{+} respectively (see Figure 1 below).

Refer to caption
Figure 1: Shock formation

It follows from the Rankine-Hugoniot condition and entropy condition on Γ\Gamma that

φ(t)[u](t,φ(t))=[f(u)](t,φ(t)),\displaystyle\varphi^{\prime}(t)[u](t,\varphi(t))=[f(u)](t,\varphi(t)), (1.10)

where [u](t,φ(t))=u+(t,φ(t))u(t,φ(t))[u](t,\varphi(t))=u_{+}(t,\varphi(t))-u_{-}(t,\varphi(t)) is the jump of uu across Γ\Gamma, and

f(u+(t,φ(t)))<φ(t)<f(u(t,φ(t))).\displaystyle f^{\prime}(u_{+}(t,\varphi(t)))<\varphi^{\prime}(t)<f^{\prime}(u_{-}(t,\varphi(t))). (1.11)

Our main results are

Theorem 1.1.

Under assumption (1.7), there exists a unique solution uC1((0,1)×)C([0,1]×)u\in C^{1}((0,1)\times{\mathbb{R}})\cap C([0,1]\times{\mathbb{R}}) to problem (1.1) together with (1.10)-(1.11) for t1t\geq 1. Furthermore,

(1) φ(t)Ck+1k[1,1+ε)\varphi(t)\in C^{\frac{k+1}{k}}[1,1+\varepsilon) and uC1((1,1+ε)×){x=φ(t)})u\in C^{1}((1,1+\varepsilon)\times{\mathbb{R}})\setminus\{x=\varphi(t)\}) for some ε>0\varepsilon>0.

(2) near the blowup point (1,0)(1,0), the behaviors of uu and its derivatives are as follows

|u(t,x)u(1,0)|\displaystyle|u(t,x)-u(1,0)| =\displaystyle= O(|t1|12k+|x|12k+1),\displaystyle O(|t-1|^{\frac{1}{2k}}+|x|^{\frac{1}{2k+1}}), (1.12)
|tu(t,x)|\displaystyle|\partial_{t}u(t,x)| =\displaystyle= O((|t1|12k+|x|12k+1)(2k1)),\displaystyle O((|t-1|^{\frac{1}{2k}}+|x|^{\frac{1}{2k+1}})^{-(2k-1)}), (1.13)
|xu(t,x)|\displaystyle|\partial_{x}u(t,x)| =\displaystyle= O((|t1|12k+|x|12k+1)2k).\displaystyle O((|t-1|^{\frac{1}{2k}}+|x|^{\frac{1}{2k+1}})^{-2k}). (1.14)
Theorem 1.2.

Under assumption (1.8), there exists a unique solution uC1((0,1)×)C([0,1]×)u\in C^{1}((0,1)\times{\mathbb{R}})\cap C([0,1]\times{\mathbb{R}}) to problem (1.1) together with (1.10)-(1.11) for t1t\geq 1. Furthermore,

(i) φ(t)C1[1,1+ε]\varphi(t)\in C^{1}[1,1+\varepsilon] and uC1(((1,1+ε)×){x=φ(t)})u\in C^{1}(((1,1+\varepsilon)\times{\mathbb{R}})\setminus\{x=\varphi(t)\}) for some ε>0\varepsilon>0. In addition, φ(t)=O((t1)|ln(t1)|2p)\varphi(t)=O((t-1)|\ln(t-1)|^{-\frac{2}{p}}) near t=1t=1 and for t>1t>1.

(ii) near the blowup point (1,0)(1,0), the behaviors of uu and its derivatives are as follows

|u(t,x)u(1,0)|\displaystyle|u(t,x)-u(1,0)| =\displaystyle= O(|ln|t1||1p+|ln|x||1p),\displaystyle O(|\ln|t-1||^{-\frac{1}{p}}+|\ln|x||^{-\frac{1}{p}}), (1.15)
|tu(t,x)|\displaystyle|\partial_{t}u(t,x)| =\displaystyle= O(|t1|1|ln|t1||11p+|x|1|ln|x||11p),\displaystyle O(|t-1|^{-1}|\ln|t-1||^{-1-\frac{1}{p}}+|x|^{-1}|\ln|x||^{-1-\frac{1}{p}}), (1.16)
|xu(t,x)|\displaystyle|\partial_{x}u(t,x)| =\displaystyle= O(|t1|1|ln|t1||1+|x|1|ln|x||1).\displaystyle O(|t-1|^{-1}|\ln|t-1||^{-1}+|x|^{-1}|\ln|x||^{-1}). (1.17)
Remark 1.1.

If we take k=1k=1, then Theorem 1.1 coincides with the result in Theorem 2 of [11]. In addition, the author in [11] only shows the behaviors of uu in some part of the neighbourhood of the blowup point (1,0)(1,0), which corresponds to the smallness of the variable |λ|=|x|(t1)32|\lambda|=\frac{|x|}{(t-1)^{\frac{3}{2}}} for t>1t>1. This can be referred to the proof of Lemma 2.1 in [11], where |λ||\lambda| is required to be small.

Remark 1.2.

The regularities of φ(t)\varphi(t) in Theorem 1.1 and Theorem 1.2 are optimal. One can see Remark 3.1 and Remark 4.1 below.

Remark 1.3.

Under the generic nondegenerate assumption of the initial data, for the 1-D 2×22\times 2 pp- system of polytropic gases, the authors in [10]-[11] and [5] obtain the formation and construction of the shock wave starting from the blowup point under some variant conditions; for the 1-D 3×33\times 3 strictly hyperbolic conservation laws with the small initial data or the 3-D full compressible Euler equations with symmetric structure and small perturbation, the authors in [4], [16] and [7] also get the formation and construction of the resulting shock waves, respectively.

In order to prove Theorem 1.1-1.2, our focus is to solve the singular and nonlinear ordinary differential equation (1.10) as in [11]. Note that the equation (1.10) is equivalent to φ(t)=G(t,φ(t))01f(θu+(t,φ(t))+(1θ)u(t,φ(t)))𝑑θ\varphi^{\prime}(t)=G(t,\varphi(t))\triangleq\int_{0}^{1}f^{\prime}(\theta u_{+}(t,\varphi(t))+(1-\theta)u_{-}(t,\varphi(t)))d\theta, where the function G(t,φ)G(t,\varphi) is not Lipschtzian with respect to variable φ\varphi since the first order derivative of u±(t,x)u_{\pm}(t,x) with respect to the variable xx admits the strong singularities (see (1.13) and (1.17)). To get the uniqueness and regularity of (φ(t),u±(t,x))(\varphi(t),u_{\pm}(t,x)), we require to carefully analyze the behavior and regularity of solution uu near the blowup point (1,0)(1,0). Due to the more degenerate conditions (1.5) and (1.6), we shall introduce some different transformations of (t,x)(t,x) from that in [11] (for examples, see (2.9), (2.20), (2.36) and so on). By involved computation, the behaviors of solution uu around the point (1,0)(1,0) are derived and then the optimal regularities of φ(t)\varphi(t) are also established. From our results, we have known two basic facts for problem (1.1): (1) Around the blowup point, the shock really appears whether the initial data are degenerate with finite orders or with infinite orders. (2) The optimal regularities of the shock solution and the resulting shock curve have explicit relations with the degenerate degrees of the initial data.

Our paper is organized as follows. In Section 2, we give some basic analysis on the characteristics envelope of equation (1.1) near (1,0)(1,0), meanwhile, the detailed behaviors of the characteristics near (1,0)(1,0) are established. The proofs of Theorem 1.1 and Theorem 1.2 are given in Section 3 and Section 4 respectively.

2 Some preliminary

For problem (1.1), we define the characteristics x=x(t,y)x=x(t,y) starting from the point (0,y)(0,y) as follows

{dx(t,y)dt=f(u(t,x(t,y))),x(0,y)=y.\linespread{1.2}\begin{cases}&\displaystyle\frac{dx(t,y)}{dt}=f^{\prime}(u(t,x(t,y))),\\ &x(0,y)=y.\end{cases} (2.1)

Then along this characteristics we have

u(t,x(t,y))u0(y).u(t,x(t,y))\equiv u_{0}(y). (2.2)

This means that the characteristics x(t,y)x(t,y) is straight and

x(t,y)=y+tg(y).x(t,y)=y+tg(y). (2.3)

For any fixed t>0t>0, in order to solve y=y(t,x)y=y(t,x) in (2.3) such that the solution uu in (2.2) can be obtained, it is necessary to let

xy(t,y)=1+tg(y)>0.\frac{\partial x}{\partial y}(t,y)=1+tg^{\prime}(y)>0.

By assumption (1.7) or (1.8), we have that near x=0x=0,

(i) for 0t<10\leq t<1, xy(t,y)>0\frac{\partial x}{\partial y}(t,y)>0;

(ii) xy(1,y)0\frac{\partial x}{\partial y}(1,y)\geq 0, and only at y=0y=0, xy(1,y)=0\frac{\partial x}{\partial y}(1,y)=0.

Thus for 0t10\leq t\leq 1, one can get a function y=y(t,x)y=y(t,x) satisfying (2.3) such that the solution to (1.1) is

u(t,x)=u0(y(t,x)).u(t,x)=u_{0}(y(t,x)). (2.4)

On the other hand, one can compute that for 0t<10\leq t<1,

{yt=g(y)1+tg(y),yx=11+tg(y).\left\{\begin{array}[]{l}\frac{\partial y}{\partial t}=-\frac{g(y)}{1+tg^{\prime}(y)},\\ \frac{\partial y}{\partial x}=\frac{1}{1+tg^{\prime}(y)}.\end{array}\right. (2.5)

This means that as (t,x)(t,x) tends to (1,0)(1-,0), then y(t,x)0y(t,x)\rightarrow 0 and |xy(t,x)|+|\partial_{x}y(t,x)|\rightarrow+\infty.


Let ε>0\varepsilon>0 be a sufficiently small constant. Under assumption (1.2), it is easy to check that for 1<t<1+ε1<t<1+\varepsilon and yy near 0, there exist two roots of yx(t,y)=0\partial_{y}x(t,y)=0 with respect to the variable yy, which are denoted by η(t)\eta_{-}(t) and η+(t)\eta_{+}(t) with η(t)<η+(t)\eta_{-}(t)<\eta_{+}(t). Set x±(t)=x(t,η±(t))x_{\pm}(t)=x(t,\eta_{\pm}(t)), we then have

  • for x<x+(t)x<x_{+}(t) (x>x(t)x>x_{-}(t) resp.) and equality (2.3), there exists a unique root denoted by y(t,x)y_{-}(t,x) (y+(t,x)y_{+}(t,x) resp.).

  • for x=x+(t)x=x_{+}(t) (x=x(t)x=x_{-}(t) resp.) and equality (2.3), there exist two roots denoted by y(t,x)<η+(t)y_{-}(t,x)<\eta_{+}(t) (η(t)<y+(t,x)\eta_{-}(t)<y_{+}(t,x) resp.).

  • for x+(t)<x<x(t)x_{+}(t)<x<x_{-}(t) and equality (2.3), there exist three roots denoted by y(t,x)<y0(t,x)<y+(t,x)y_{-}(t,x)<y_{0}(t,x)<y_{+}(t,x).

Set

Ω\displaystyle\Omega_{-} =\displaystyle= {(t,x):1<t<1+ε,x<x(t)}\displaystyle\{(t,x):1<t<1+\varepsilon,\ x<x_{-}(t)\}
Ω+\displaystyle\Omega_{+} =\displaystyle= {(t,x):1<t<1+ε,x>x+(t)}\displaystyle\{(t,x):1<t<1+\varepsilon,\ x>x_{+}(t)\}
Ω0\displaystyle\Omega_{0} =\displaystyle= {(t,x):1<t<1+ε,x+(t)<x<x(t)}.\displaystyle\{(t,x):1<t<1+\varepsilon,\ x_{+}(t)<x<x_{-}(t)\}.

Under (1.7), we derive some properties of η±(t)\eta_{\pm}(t) and x±(t)x_{\pm}(t) near the blowup point (1,0)(1,0).

Lemma 2.1.

There exists an ε>0\varepsilon>0 sufficiently small such that
(1) η±(t)C2k+1(1,1+ε)C12k[1,1+ε)\eta_{\pm}(t)\in C^{2k+1}(1,1+\varepsilon)\cap C^{\frac{1}{2k}}[1,1+\varepsilon) admit the following expansion

η±(t)=±(2k+1)12k(t1)12kg(2k+2)(0)2k(2k)!(2k+1)2k+1k(t1)1k+o((t1)1k);\eta_{\pm}(t)=\pm(2k+1)^{-\frac{1}{2k}}(t-1)^{\frac{1}{2k}}-\frac{g^{(2k+2)}(0)}{2k(2k)!}(2k+1)^{-\frac{2k+1}{k}}(t-1)^{\frac{1}{k}}+o((t-1)^{\frac{1}{k}}); (2.6)

(2) x±(t)=x(t,η±(t))C2k+1(1,1+ε)C2k+12k[1,1+ε)x_{\pm}(t)=x(t,\eta_{\pm}(t))\in C^{2k+1}(1,1+\varepsilon)\cap C^{\frac{2k+1}{2k}}[1,1+\varepsilon) are the envelopes of the characteristic lines (2.3) which form a cusp at (1,0)(1,0), meanwhile,

x±(t)=2k(2k+1)2k+12k(t1)2k+12k+g(2k+2)(0)(2k+2)!(2k+1)k+1k(t1)k+1k+o((t1)k+1k).x_{\pm}(t)=\mp 2k(2k+1)^{-\frac{2k+1}{2k}}(t-1)^{\frac{2k+1}{2k}}+\frac{g^{(2k+2)}(0)}{(2k+2)!}(2k+1)^{-\frac{k+1}{k}}(t-1)^{\frac{k+1}{k}}+o((t-1)^{\frac{k+1}{k}}). (2.7)
Proof.

(1) Note that η±(t)\eta_{\pm}(t) are the solutions of

1+tg(y)=(t1)+(2k+1)ty2k+tr(y)=0.1+tg^{\prime}(y)=-(t-1)+(2k+1)ty^{2k}+tr^{\prime}(y)=0. (2.8)

This immediately yields η±C2k+1(1,1+ε]\eta_{\pm}\in C^{2k+1}(1,1+\varepsilon] by the implicit function theorem. For t1+t\to 1+, set

s=(t1)12k,z=ys.s=(t-1)^{\frac{1}{2k}},\ z=\frac{y}{s}. (2.9)

Then (2.8) becomes

F(s,z)(1+s2k)[(2k+1)z2k+s2kr(sz)]1=0.F(s,z)\triangleq(1+s^{2k})[(2k+1)z^{2k}+s^{-2k}r^{\prime}(sz)]-1=0. (2.10)

Since r(sz)=O(s2k+1)r^{\prime}(sz)=O(s^{2k+1}) for ss near 0, F(0,z±0)=0F(0,z^{0}_{\pm})=0 holds for z±0=±(2k+1)12kz^{0}_{\pm}=\pm(2k+1)^{-\frac{1}{2k}}. By direct computation, we have that

sF(s,z)\displaystyle\partial_{s}F(s,z) =\displaystyle= 2k(2k+1)s2k1z2k2ks2k1r(sz)+(s2k+1)zr′′(sz),\displaystyle 2k(2k+1)s^{2k-1}z^{2k}-2ks^{-2k-1}r^{\prime}(sz)+(s^{-2k}+1)zr^{\prime\prime}(sz), (2.11)
zF(s,z)\displaystyle\partial_{z}F(s,z) =\displaystyle= (1+s2k)[2k(2k+1)z2k1+s2k+1r′′(sz)].\displaystyle(1+s^{2k})[2k(2k+1)z^{2k-1}+s^{-2k+1}r^{\prime\prime}(sz)]. (2.12)

Together with r(sz)=g(2k+2)(0)(2k+2)!(sz)2k+2+o(s2k+2)r(sz)=\frac{g^{(2k+2)}(0)}{(2k+2)!}(sz)^{2k+2}+o(s^{2k+2}), this yields

sF(0,z±0)\displaystyle\partial_{s}F(0,z^{0}_{\pm}) =\displaystyle= ±g(2k+2)(0)(2k+1)!(2k+1)2k+12k,\displaystyle\pm\frac{g^{(2k+2)}(0)}{(2k+1)!}(2k+1)^{-\frac{2k+1}{2k}}, (2.13)
zF(0,z±0)\displaystyle\partial_{z}F(0,z^{0}_{\pm}) =\displaystyle= ±2k(2k+1)12k0.\displaystyle\pm 2k(2k+1)^{\frac{1}{2k}}\neq 0. (2.14)

By the implicit function theorem, for small ε>0\varepsilon>0 there exist

z=z±(s)C2k+1[0,ε]z=z_{\pm}(s)\in C^{2k+1}[0,\varepsilon] (2.15)

such that F(s,z±(s))=0F(s,z_{\pm}(s))=0 and

z±(s)=z±0g(2k+2)(0)2k(2k)!(2k+1)2k+1ks+o(s).z_{\pm}(s)=z^{0}_{\pm}-\frac{g^{(2k+2)}(0)}{2k(2k)!}(2k+1)^{-\frac{2k+1}{k}}s+o(s). (2.16)

Therefore, (2.6) is shown and then η±(t)C12k[1,1+ε]\eta_{\pm}(t)\in C^{\frac{1}{2k}}[1,1+\varepsilon].


(2) By (1.7) and (2.3), we have

x±(t)=x(t,η±(t))=(t1)η±(t)+tη±2k+1(t)+tr(η±(t)).x_{\pm}(t)=x(t,\eta_{\pm}(t))=-(t-1)\eta_{\pm}(t)+t\eta^{2k+1}_{\pm}(t)+tr(\eta_{\pm}(t)).

Together with (2.6), this yields x±(t)C2k+1(1,1+ε]C2k+12k[1,1+ε]x_{\pm}(t)\in C^{2k+1}(1,1+\varepsilon]\cap C^{\frac{2k+1}{2k}}[1,1+\varepsilon] and the expansion (2.7). In addition, due to yx(t,η±(t))=0\frac{\partial}{\partial y}x(t,\eta_{\pm}(t))=0, then for t[1,1+ε]t\in[1,1+\varepsilon],

ddtx±(t)=tx(t,η±(t))=g(η±(t)).\frac{d}{dt}x_{\pm}(t)=\frac{\partial}{\partial t}x(t,\eta_{\pm}(t))=g(\eta_{\pm}(t)). (2.17)

This means that the tangent direction of x=x±(t)x=x_{\pm}(t) coincides with the characteristic speed of (2.3) at (t,x±(t))(t,x_{\pm}(t)). Consequently, the proof of (2) is finished. ∎

Under (1.8)-(1.9), we have

Lemma 2.2.

For η±(t),x±(t),y±(t,x)\eta_{\pm}(t),\ x_{\pm}(t),\ y_{\pm}(t,x) and y0(t,x)y_{0}(t,x), we can deduce the following properties for small ε>0\varepsilon>0:
(1) η±(t)C(1,1+ε]C[1,1+ε]\eta_{\pm}(t)\in C^{\infty}(1,1+\varepsilon]\cap C[1,1+\varepsilon] with η±(τ)=±|ln(t1)|1p+O(ln|ln(t1))||ln(t1)|)\eta_{\pm}(\tau)=\pm|\ln(t-1)|^{-\frac{1}{p}}+O(\frac{\ln|\ln(t-1))|}{|\ln(t-1)|}).
(2) x±(t)=x(t,η±(t))C(1,1+ε]C1[1,1+ε]x_{\pm}(t)=x(t,\eta_{\pm}(t))\in C^{\infty}(1,1+\varepsilon]\cap C^{1}[1,1+\varepsilon] are the envelopes of the characteristic lines and form a cusp at (1,0)(1,0). Moreover we have the expansion x±(t)=(t1)|ln(t1)|1p+O((t1)ln|ln(t1)||ln(t1)|)x_{\pm}(t)=\mp(t-1)|\ln(t-1)|^{-\frac{1}{p}}+O(\frac{(t-1)\ln|\ln(t-1)|}{|\ln(t-1)|}).
(3) For any t(1,1+ε]t\in(1,1+\varepsilon], y(t,)y_{-}(t,\cdot) is an increasing function from (,x(t)](-\infty,x_{-}(t)] onto (,η(t)](-\infty,\eta_{-}(t)]; y0(t,)y_{0}(t,\cdot) is a decreasing function from [x+(t),x(t)][x_{+}(t),x_{-}(t)] onto [η(t),η+(t)][\eta_{-}(t),\eta_{+}(t)]; y+(t,)y_{+}(t,\cdot) is an increasing function from [x+(t),+)[x_{+}(t),+\infty) onto [η+(t),+)[\eta_{+}(t),+\infty). Moreover, ym(t,x)C(Ωm)C(Ω¯m)y_{m}(t,x)\in C^{\infty}(\Omega_{m})\cap C(\bar{\Omega}_{m}), where m=,+,0m=-,+,0.

Proof.

(1) Set τ=t1\tau=t-1 for t1t\geq 1. Note that η±(t)\eta_{\pm}(t) are defined for small τ>0\tau>0 and are the solutions of the following equation

1+tg(y)=(τ+1|y|pe|y|p)+τ|y|pe|y|p+(τ+1)e|y|pr1(y)=0,1+tg^{\prime}(y)=\left(-\tau+\frac{1}{|y|^{p}}e^{-|y|^{-p}}\right)+\frac{\tau}{|y|^{p}}e^{-|y|^{-p}}+(\tau+1)e^{-|y|^{-p}}r_{1}(y)=0, (2.18)

where r1(y)=1p+yr0(y)|y|p+2+r0(y)=O(|y|min{p+1,0})r_{1}(y)=\frac{1}{p}+\frac{yr_{0}(y)}{|y|^{p+2}}+r_{0}^{\prime}(y)=O(|y|^{\min\{-p+1,0\}}). Denote ω=|lnτ|1\omega=|\ln\tau|^{-1} and z=|y|p(ω1lnω)z=|y|^{-p}-(\omega^{-1}-\ln\omega), then (2.18) becomes

F(τ,z)1+(1+ωzωlnω)ez+ωez(τ(z+ω1lnω)+(τ+1)r1(y))=0.F(\tau,z)\triangleq-1+\left(1+\omega z-\omega\ln\omega\right)e^{-z}+\omega e^{-z}\left(\tau\left(z+\omega^{-1}-\ln\omega\right)+(\tau+1)r_{1}(y)\right)=0. (2.19)

Obviously, F(0,0)=0F(0,0)=0. In addition, by direct computation, we have that for small |y||y|,

r1(y)=p+1|y|p+2r0(y)+yr0(y)|y|p+2+r0′′(y)=O(|y|p)r^{\prime}_{1}(y)=-\frac{p+1}{|y|^{p+2}}r_{0}(y)+\frac{yr^{\prime}_{0}(y)}{|y|^{p+2}}+r^{\prime\prime}_{0}(y)=O(|y|^{-p})

and

yz=y|y|pp.\frac{\partial y}{\partial z}=-\frac{y|y|^{p}}{p}.

Thus one can check that

Fz=ωez(1+zωωlnω)ezωez(τ(z+ω1lnω1)+(τ+1)(r1(y)r1(y)yz)).\frac{\partial F}{\partial z}=\omega e^{-z}-\left(1+z\omega-\omega\ln\omega\right)e^{-z}-\omega e^{-z}\left(\tau\left(z+\omega^{-1}-\ln\omega-1\right)+(\tau+1)(r_{1}(y)-r^{\prime}_{1}(y)\frac{\partial y}{\partial z})\right).

This yields Fz(0,0)=1\frac{\partial F}{\partial z}(0,0)=-1 by |y|ω1p|y|\lesssim\omega^{\frac{1}{p}}. Since F(τ,z)F(\tau,z) is continuous and has the continuous partial derivative Fz\frac{\partial F}{\partial z} near (0,0)(0,0), it follows from the implicit function theorem that there exists a continuous function z=z(τ)z=z(\tau) near z=0z=0 to satisfy F(τ,z)=0F(\tau,z)=0. This deduces η±(t)=±(|lnτ|+ln|lnτ|+z(τ))1p=±|lnτ|1p+O(ln|lnτ||lnτ|)\eta_{\pm}(t)=\pm(|\ln\tau|+\ln|\ln\tau|+z(\tau))^{-\frac{1}{p}}=\pm|\ln\tau|^{-\frac{1}{p}}+O(\frac{\ln|\ln\tau|}{|\ln\tau|}). On the other hand, for τ>0\tau>0, we have xy(t,y)0\frac{\partial x}{\partial y}(t,y)\neq 0 and gCg^{\prime}\in C^{\infty}. Then by the implicit function theorem, η±(t)C(1,1+ε]\eta_{\pm}(t)\in C^{\infty}(1,1+\varepsilon] hold.

(2) By (1.8) and (2.3), we have that for small τ>0\tau>0,

x±(t)=τη±(t)+t(η±(t)pe|η±(t)|p+r0(η±(t)))=τ[η±(t)+t(η±(t)p+o(η±(t)))e|O(ln|lnτ||lnτ|)|],x_{\pm}(t)=-\tau\eta_{\pm}(t)+t(\frac{\eta_{\pm}(t)}{p}e^{-|\eta_{\pm}(t)|^{-p}}+r_{0}(\eta_{\pm}(t)))=\tau[-\eta_{\pm}(t)+t(\frac{\eta_{\pm}(t)}{p}+o(\eta_{\pm}(t)))e^{-|O(\frac{\ln|\ln\tau|}{|\ln\tau|})|}],

which derives x±(t)C(1,1+ε]C[1,1+ε]x_{\pm}(t)\in C^{\infty}(1,1+\varepsilon]\cap C[1,1+\varepsilon]. On the other hand, it holds that for t(1,1+ε]t\in(1,1+\varepsilon],

ddtx±(t)\displaystyle\frac{d}{dt}x_{\pm}(t) =\displaystyle= tx(t,η±(t))+yx(t,η±(t))ddtη±(t)\displaystyle\frac{\partial}{\partial t}x(t,\eta_{\pm}(t))+\frac{\partial}{\partial y}x(t,\eta_{\pm}(t))\frac{d}{dt}\eta_{\pm}(t)
=\displaystyle= g(η±(t)),\displaystyle g(\eta_{\pm}(t)),

which means that the tangent direction of x=x±(t)x=x_{\pm}(t) is same as the characteristic speed of (2.3) at (t,x±(t))(t,x_{\pm}(t)). In addition, at the point (1,0)(1,0), one has

x±(1)\displaystyle x^{\prime}_{\pm}(1) =\displaystyle= limt1+x±(t,η±(t))0t1\displaystyle\lim_{t\rightarrow 1_{+}}\frac{x_{\pm}(t,\eta_{\pm}(t))-0}{t-1}
=\displaystyle= limt1+[η±(t)+t(η±(t)p+o(η±(t)))eO(ln|lnτ||lnτ|)]\displaystyle\lim_{t\rightarrow 1_{+}}[-\eta_{\pm}(t)+t(\frac{\eta_{\pm}(t)}{p}+o(\eta_{\pm}(t)))e^{O(\frac{\ln|\ln\tau|}{|\ln\tau|})}]
=\displaystyle= 0\displaystyle 0
=\displaystyle= g(0).\displaystyle g(0).

Hence we finish the proof of (2).

(3) For any fixed t(1,1+ε]t\in(1,1+\varepsilon], due to

yx(t,y){>0, for y(,η(t))(η+(t),+),=0, for y=η±(t),<0, for y(η(t),η+(t)),\frac{\partial}{\partial y}x(t,y)\left\{\begin{array}[]{ll}>0,&\text{ for }y\in(-\infty,\eta_{-}(t))\cup(\eta_{+}(t),+\infty),\\ =0,&\text{ for }y=\eta_{\pm}(t),\\ <0,&\text{ for }y\in(\eta_{-}(t),\eta_{+}(t)),\end{array}\right.

then by the inverse function theorem, ym(t,)y_{m}(t,\cdot) with m=,+,0m=-,+,0 are well defined and satisfy the corresponding monotonicity. Moreover, y+(t,)C(x+(t),+)C[x+(t),+)y_{+}(t,\cdot)\in C^{\infty}(x_{+}(t),+\infty)\cap C[x_{+}(t),+\infty), y0(t,)C(x+(t),x(t))C[x+(t),x(t)]y_{0}(t,\cdot)\in C^{\infty}(x_{+}(t),x_{-}(t))\cap C[x_{+}(t),x_{-}(t)] and y(t,)C(,x(t))C(,x(t)]y_{-}(t,\cdot)\in C^{\infty}(-\infty,x_{-}(t))\cap C(-\infty,x_{-}(t)].

On the other hand, because of yx(t,x)0\partial_{y}x(t,x)\neq 0 for (t,x){x=x±(t)}(t,x)\notin\{x=x_{\pm}(t)\}, thus it follows from the implicit function theorem that ym(t,x)C(Ωm)y_{m}(t,x)\in C^{\infty}(\Omega_{m}), m=,+,0m=-,+,0. For the continuity of ym(t,x)y_{m}(t,x) in Ω¯m\bar{\Omega}_{m}, we take y+(t,x)y_{+}(t,x) as an example. By x+(t)C1([1,1+ε])x_{+}(t)\in C^{1}([1,1+\varepsilon]), we then get

|y+(t¯,x+(t¯))y+(t,x)|=|y+(t¯,x+(t¯))y+(t,x+(t))|+|y+(t,x+(t))y+(t,x)|0|y_{+}(\bar{t},x_{+}(\bar{t}))-y_{+}(t,x)|=|y_{+}(\bar{t},x_{+}(\bar{t}))-y_{+}(t,x_{+}(t))|+|y_{+}(t,x_{+}(t))-y_{+}(t,x)|\rightarrow 0

as (t,x)(t¯,x+(t¯))(t,x)\rightarrow(\bar{t},x_{+}(\bar{t})) for t¯[1,1+ε]\bar{t}\in[1,1+\varepsilon] and (t,x)Ω+(t,x)\in\Omega_{+}. Thus y+(t,x)C(Ω+)C(Ω¯+)y_{+}(t,x)\in C^{\infty}(\Omega_{+})\cap C(\bar{\Omega}_{+}) holds.

To study the formation of shock wave and the regularity of the resulting shock x=φ(t)x=\varphi(t) to equation (1.1), it is required to study the properties of y±(t,x)y_{\pm}(t,x) for (t,x)(t,x) in the cusp domain Ω0\Omega_{0}. Under assumption (1.5), motivated by [11], we take the following change of the variables

τ=t1,s=τ12k,μ=ys,λ=xs2k+1,\tau=t-1,\ s=\tau^{\frac{1}{2k}},\ \mu=\frac{y}{s},\ \lambda=\frac{x}{s^{2k+1}}, (2.20)

and will establish the behavior of y±(t,x)y_{\pm}(t,x) near (1,0)(1,0) in some sub-domain of Ω0\Omega_{0}.

Lemma 2.3.

For small ε>0\varepsilon>0, under assumption (1.7), there exists some constant δ>0\delta>0 such that for (s,λ){0sε,|λ|δ}(s,\lambda)\in\{0\leq s\leq\varepsilon,\ |\lambda|\leq\delta\}, (s,λ)sjy±(t,x)(s,\lambda)\rightarrow s^{j}y_{\pm}(t,x) are of Cj+2C^{j+2} for j=1,0,1,,2kj=-1,0,1,\ldots,2k and y±(t,x)y_{\pm}(t,x) admit the following expansions

y+(t,x)\displaystyle y_{+}(t,x) =\displaystyle= s(1+λ2kg(2k+2)(0)2k(2k+2)!s)+O(s3+sλ2),\displaystyle s(1+\frac{\lambda}{2k}-\frac{g^{(2k+2)}(0)}{2k(2k+2)!}s)+O(s^{3}+s\lambda^{2}), (2.21)
y(t,x)\displaystyle y_{-}(t,x) =\displaystyle= s(1+λ2kg(2k+2)(0)2k(2k+2)!s)+O(s3+sλ2).\displaystyle s(-1+\frac{\lambda}{2k}-\frac{g^{(2k+2)}(0)}{2k(2k+2)!}s)+O(s^{3}+s\lambda^{2}). (2.22)
Proof.

Let

h(y)r(y)y2k+1=01(1θ)2k(2k)!g(2k+1)(θy)𝑑θ1.h(y)\triangleq\frac{r(y)}{y^{2k+1}}=\int_{0}^{1}\frac{(1-\theta)^{2k}}{(2k)!}g^{(2k+1)}(\theta y)d\theta-1. (2.23)

Then h(y)Cp2k1h(y)\in C^{p-2k-1} and h(0)=0h(0)=0. Furthermore,

yh(y)=y(2k)!01(1θ)2kθg(2k+2)(θy)𝑑θ=1(2k)!01g(2k+1)(θy)(1θ)2k1[1(2k+1)θ]𝑑θ.yh^{\prime}(y)=\frac{y}{(2k)!}\int_{0}^{1}(1-\theta)^{2k}\theta g^{(2k+2)}(\theta y)d\theta=-\frac{1}{(2k)!}\int_{0}^{1}g^{(2k+1)}(\theta y)(1-\theta)^{2k-1}[1-(2k+1)\theta]~{}d\theta. (2.24)

This derives yh(y)C1yh^{\prime}(y)\in C^{1}. Similarly, yjh(j)(y)C1y^{j}h^{(j)}(y)\in C^{1} holds for j=2,,2k+1j=2,\ldots,2k+1. Divided by s2ks^{2k}, (2.3) becomes

G(s,λ,μ)μ+(1+s2k)μ2k+1+(1+s2k)μ2k+1h(sμ)λ=0.G(s,\lambda,\mu)\triangleq-\mu+(1+s^{2k})\mu^{2k+1}+(1+s^{2k})\mu^{2k+1}h(s\mu)-\lambda=0. (2.25)

For s=λ=0s=\lambda=0, by G(0,0,μ)=μ+μ2k+1=μ(1μ2k)=0G(0,0,\mu)=-\mu+\mu^{2k+1}=-\mu(1-\mu^{2k})=0 we get the roots μ±0=±1\mu^{0}_{\pm}=\pm 1 and μc0=0\mu^{0}_{c}=0. Note that

μG(s,λ,μ)=1+(2k+1)(1+s2k)μ2k+(1+s2k)((2k+1)μ2kh(sμ)+μ2k+1sh(sμ)).\partial_{\mu}G(s,\lambda,\mu)=-1+(2k+1)(1+s^{2k})\mu^{2k}+(1+s^{2k})((2k+1)\mu^{2k}h(s\mu)+\mu^{2k+1}sh^{\prime}(s\mu)). (2.26)

Then

μG(0,0,±1)=2k0.\partial_{\mu}G(0,0,\pm 1)=2k\neq 0. (2.27)

By the implicit function theorem, there exist functions μ=μ±(s,λ)C1\mu=\mu_{\pm}(s,\lambda)\in C^{1} near (s,λ)=(0,0)(s,\lambda)=(0,0) such that

G(s,λ,μ±(s,λ))=0,μ±(0,0)=±1,G(s,\lambda,\mu_{\pm}(s,\lambda))=0,\ \mu_{\pm}(0,0)=\pm 1, (2.28)

and then s1y±C1s^{-1}y_{\pm}\in C^{1}. On the other hand, due to

sG(s,λ,μ)\displaystyle\partial_{s}G(s,\lambda,\mu) =\displaystyle= 2ks2k1μ2k+2ks2k1μ2k+1h(sμ)+(1+s2k)μ2k+2h(sμ),\displaystyle 2ks^{2k-1}\mu^{2k}+2ks^{2k-1}\mu^{2k+1}h(s\mu)+(1+s^{2k})\mu^{2k+2}h^{\prime}(s\mu), (2.29)
λG(s,λ,μ)\displaystyle\partial_{\lambda}G(s,\lambda,\mu) =\displaystyle= 1,\displaystyle-1, (2.30)

then

sG(0,0,±1)\displaystyle\partial_{s}G(0,0,\pm 1) =\displaystyle= h(0)=g(2k+2)(0)(2k)!01(1θ)2kθ𝑑θ=g(2k+2)(0)(2k+2)!\displaystyle h^{\prime}(0)=\frac{g^{(2k+2)}(0)}{(2k)!}\int_{0}^{1}(1-\theta)^{2k}\theta d\theta=\frac{g^{(2k+2)}(0)}{(2k+2)!} (2.31)
λG(0,0,±1)\displaystyle\partial_{\lambda}G(0,0,\pm 1) =\displaystyle= 1.\displaystyle-1. (2.32)

It follows from (2.27), (2.31) and (2.32) that

sμ±(0,0)=g(2k+2)(0)2k(2k+2)!,λμ±(0,0)=12k.\partial_{s}\mu_{\pm}(0,0)=-\frac{g^{(2k+2)}(0)}{2k(2k+2)!},\quad\partial_{\lambda}\mu_{\pm}(0,0)=\frac{1}{2k}. (2.33)

Consequently, the expansions (2.21) and (2.22) are shown.

We next prove (s,λ)y±C2(s,\lambda)\rightarrow y_{\pm}\in C^{2}. By

ssG(s,λ,μ±)+μG(s,λ,μ±)(ssμ±(s,λ))=0s\partial_{s}G(s,\lambda,\mu_{\pm})+\partial_{\mu}G(s,\lambda,\mu_{\pm})(s\partial_{s}\mu_{\pm}(s,\lambda))=0 (2.34)

and

ssG(s,λ,μ±(s,λ))=2ks2kμ±2k+2ks2kμ±2k+1h(sμ±)+(1+s2k)μ±2k+1(sμ±h(sμ±))C1,s\partial_{s}G(s,\lambda,\mu_{\pm}(s,\lambda))=2ks^{2k}\mu_{\pm}^{2k}+2ks^{2k}\mu_{\pm}^{2k+1}h(s\mu_{\pm})+(1+s^{2k})\mu_{\pm}^{2k+1}(s\mu_{\pm}h^{\prime}(s\mu_{\pm}))\in C^{1}, (2.35)

we then have ssμ±(s,λ)C1s\partial_{s}\mu_{\pm}(s,\lambda)\in C^{1}. Thus from sy±=μ±+ssμ±\partial_{s}y_{\pm}=\mu_{\pm}+s\partial_{s}\mu_{\pm}, one can see that (s,λ)y±(s,\lambda)\to y_{\pm} is of C1C^{1}. In addition, by λy±=sλμ±(s,λ)\partial_{\lambda}y_{\pm}=s\partial_{\lambda}\mu_{\pm}(s,\lambda) and similar computation, we can get y±C2y_{\pm}\in C^{2} with respect to ss and λ\lambda.

Note that

sjsjG(s,λ,μ±(s,λ))\displaystyle s^{j}\partial^{j}_{s}G(s,\lambda,\mu_{\pm}(s,\lambda)) =\displaystyle= 𝒢(s,λ,ssμ±,,sj1sj1μ±,h(sμ±),(sμ±)h(sμ±),,(sμ±)jh(j)(sμ±))\displaystyle{\mathcal{G}}(s,\lambda,s\partial_{s}\mu_{\pm},\ldots,s^{j-1}\partial^{j-1}_{s}\mu_{\pm},h(s\mu_{\pm}),(s\mu_{\pm})h^{\prime}(s\mu_{\pm}),\ldots,(s\mu_{\pm})^{j}h^{(j)}(s\mu_{\pm}))
+μG(s,λ,μ±)(sjsjμ±(s,λ)),j=2,3,,2k+1,\displaystyle+\partial_{\mu}G(s,\lambda,\mu_{\pm})(s^{j}\partial^{j}_{s}\mu_{\pm}(s,\lambda)),\ j=2,3,\ldots,2k+1,

where 𝒢{\mathcal{G}} is a polynomial with respect to its arguments. Then sjsjμ±(s,λ)C1s^{j}\partial^{j}_{s}\mu_{\pm}(s,\lambda)\in C^{1} for j=2,3,,2k+1j=2,3,\ldots,2k+1 by induction. Similarly, sjλmsjmμ±(s,λ)C1s^{j}\partial^{m}_{\lambda}\partial^{j-m}_{s}\mu_{\pm}(s,\lambda)\in C^{1} for 1mj2k+11\leq m\leq j\leq 2k+1. Consequently, the proof of (s,λ)sjy±(t,x)Cj+2(s,\lambda)\rightarrow s^{j}y_{\pm}(t,x)\in C^{j+2} for j=1,0,1,,2kj=-1,0,1,\ldots,2k is completed. ∎

Under assumption (1.8), we now study the asymptotic behavior of y±(t,x)y_{\pm}(t,x) near (1,0)(1,0). In this case, we take the following change of the variables

τ=t1,s=|lnτ|1p,λ=xsτ,μ=ys.\tau=t-1,\ s=|\ln\tau|^{-\frac{1}{p}},\ \lambda=\frac{x}{s\tau},\ \mu=\frac{y}{s}. (2.36)

Then we obtain

Lemma 2.4.

Under assumption (1.8), there exist some small constants ε\varepsilon, δ>0\delta>0 such that for (t,x)Ω0{1t1+ε,δsτ<x<δsτ}(t,x)\in\Omega_{0}\triangleq\{1\leq t\leq 1+\varepsilon,\ -\delta s\tau<x<\delta s\tau\}, y±C1+py_{\pm}\in C^{1+p} hold for the variables ss and λ\lambda. Furthermore,

y+(t,x)\displaystyle y_{+}(t,x) =\displaystyle= s(1+lnppsp+spλp)+O(smin{p+2,2p+1}+sp+1|λ|2),\displaystyle s\left(1+\frac{\ln p}{p}s^{p}+\frac{s^{p}\lambda}{p}\right)+O\left(s^{\min\{p+2,2p+1\}}+s^{p+1}|\lambda|^{2}\right), (2.37)
y(t,x)\displaystyle y_{-}(t,x) =\displaystyle= s(1lnppsp+spλp)+O(smin{p+2,2p+1}+sp+1|λ|2).\displaystyle s\left(-1-\frac{\ln p}{p}s^{p}+\frac{s^{p}\lambda}{p}\right)+O\left(s^{\min\{p+2,2p+1\}}+s^{p+1}|\lambda|^{2}\right). (2.38)
Proof.

By divided by sτs\tau, (2.3) can be written as follows

G(s,λ,μ)μ(1+1pesp(1|μ|p))+μpesp|μ|p+esp+1sesp|μ|pr0(sμ)λ=0.G(s,\lambda,\mu)\triangleq\mu(-1+\frac{1}{p}e^{s^{-p}(1-|\mu|^{-p})})+\frac{\mu}{p}e^{-s^{-p}|\mu|^{-p}}+\frac{e^{s^{-p}}+1}{s}e^{-s^{-p}|\mu|^{-p}}r_{0}(s\mu)-\lambda=0. (2.39)

Below we assume μ0\mu\neq 0. Without loss of generality, one can assume μ>0\mu>0 (corresponding to the case of y+(t,x)y_{+}(t,x)). We divide the proof of Lemma 2.4 into the two cases of p1p\geq 1 and p(0,1)p\in(0,1).


Case 1. p1p\geq 1.

Set ζ=sp(1μp)lnp\zeta=s^{-p}(1-\mu^{-p})-\ln p and μ=(1sp(ζ+lnp))1p\mu=\left(1-s^{p}(\zeta+\ln p)\right)^{-\frac{1}{p}}. Then (2.39) becomes

F1(s,λ,ζ)G(s,λ,μ)=μ(1+eζ)+μeζsp+peζ(1+esp)sr0(sμ)λ.F_{1}(s,\lambda,\zeta)\triangleq G(s,\lambda,\mu)=\mu(-1+e^{\zeta})+\mu e^{\zeta-s^{-p}}+\frac{pe^{\zeta}(1+e^{-s^{-p}})}{s}r_{0}(s\mu)-\lambda. (2.40)

Obviously F1(0,0,0)=0F_{1}(0,0,0)=0 and lims0+,ζ0μ=1\displaystyle\lim_{s\to 0+,\zeta\to 0}\mu=1. Note that

sμ=sp1(ζ+lnp)(1sp(ζ+lnp))1p1,ζμ=spp(1sp(ζ+lnp))1p1.\partial_{s}\mu=s^{p-1}(\zeta+\ln p)\left(1-s^{p}(\zeta+\ln p)\right)^{-\frac{1}{p}-1},\quad\partial_{\zeta}\mu=\frac{s^{p}}{p}\left(1-s^{p}(\zeta+\ln p)\right)^{-\frac{1}{p}-1}.

Thanks to p1p\geq 1, we have that sμ\partial_{s}\mu is bounded. On the other hand,

sF1\displaystyle\partial_{s}F_{1} =\displaystyle= psp1μeζsp+(1+eζ+eζsp)sμ\displaystyle ps^{-p-1}\mu e^{\zeta-s^{-p}}+\left(-1+e^{\zeta}+e^{\zeta-s^{-p}}\right)\partial_{s}\mu
+peζ(esp(psp1)1)s2r0(sμ)+peζ(1+esp)sr0(sμ)(μ+ssμ),\displaystyle+\frac{pe^{\zeta}\left(e^{-s^{-p}}(ps^{-p}-1)-1\right)}{s^{2}}r_{0}(s\mu)+\frac{pe^{\zeta}(1+e^{-s^{-p}})}{s}r^{\prime}_{0}(s\mu)\left(\mu+s\partial_{s}\mu\right),
λF1\displaystyle\partial_{\lambda}F_{1} =\displaystyle= 1,\displaystyle-1,
ζF1\displaystyle\partial_{\zeta}F_{1} =\displaystyle= μ(eζ+eζsp)+(1+eζ+eζsp)ζμ+peζ(1+esp)s(r0(sμ)+sr0(sμ)ζμ).\displaystyle\mu\left(e^{\zeta}+e^{\zeta-s^{-p}}\right)+\left(-1+e^{\zeta}+e^{\zeta-s^{-p}}\right)\partial_{\zeta}\mu+\frac{pe^{\zeta}(1+e^{-s^{-p}})}{s}\left(r_{0}(s\mu)+sr^{\prime}_{0}(s\mu)\partial_{\zeta}\mu\right).

This derives F1C1F_{1}\in C^{1} and

sF1(0,0,0)=p2r0′′(0),λF1(0,0,0)=1,ζF1(0,0,0)=1.\partial_{s}F_{1}(0,0,0)=\frac{p}{2}r^{\prime\prime}_{0}(0),\quad\partial_{\lambda}F_{1}(0,0,0)=-1,\quad\partial_{\zeta}F_{1}(0,0,0)=1. (2.41)

Thus by the implicit function theorem, one can obtain that there exists a unique function ζ(s,λ)C1\zeta(s,\lambda)\in C^{1} satisfying F1(s,λ,ζ(s,λ))=0F_{1}(s,\lambda,\zeta(s,\lambda))=0 and admitting the following expansion

ζ(s,λ)=p2r0′′(0)s+λ+O(s2+λ2).\zeta(s,\lambda)=\frac{p}{2}r^{\prime\prime}_{0}(0)s+\lambda+O\left(s^{2}+\lambda^{2}\right). (2.42)

At this time, we get

μ(s,λ)\displaystyle\mu(s,\lambda) =\displaystyle= (1sp(ζ+lnp))1p\displaystyle\left(1-s^{p}(\zeta+\ln p)\right)^{-\frac{1}{p}} (2.43)
=\displaystyle= (1sp(lnp+p2r0′′(0)s+λ+O(s2+λ2))1p\displaystyle\left(1-s^{p}(\ln p+\frac{p}{2}r^{\prime\prime}_{0}(0)s+\lambda+O\left(s^{2}+\lambda^{2}\right)\right)^{-\frac{1}{p}}
=\displaystyle= 1+lnppsp+spλp+O(sp+1+sp|λ|2)\displaystyle 1+\frac{\ln p}{p}s^{p}+\frac{s^{p}\lambda}{p}+O\left(s^{p+1}+s^{p}|\lambda|^{2}\right)

and slμCl+ps^{l}\mu\in C^{l+p}, l=0,1l=0,1.

If we consider the case of μ<0\mu<0, then by the same method, one can obtain the expansion

μ(s,λ)=1lnppsp+spλp+O(sp+1+sp|λ|2).\mu(s,\lambda)=-1-\frac{\ln p}{p}s^{p}+\frac{s^{p}\lambda}{p}+O\left(s^{p+1}+s^{p}|\lambda|^{2}\right). (2.44)

Case 2. 0<p<10<p<1.

Set ω=sp\omega=s^{p}, ζ=ω1(1μp)lnp\zeta=\omega^{-1}(1-\mu^{-p})-\ln p and μ=(1ω(ζ+lnp))1p\mu=\left(1-\omega(\zeta+\ln p)\right)^{-\frac{1}{p}}. Then (2.39) becomes

F2(ω,λ,ζ)G(s,λ,μ)=μ(1+eζ)+μeζω1+peζ(1+eω1)ω1pr0(ω1pμ)λ.F_{2}(\omega,\lambda,\zeta)\triangleq G(s,\lambda,\mu)=\mu(-1+e^{\zeta})+\mu e^{\zeta-\omega^{-1}}+\frac{pe^{\zeta}(1+e^{-\omega^{-1}})}{\omega^{\frac{1}{p}}}r_{0}(\omega^{\frac{1}{p}}\mu)-\lambda. (2.45)

Obviously F2(0,0,0)=0F_{2}(0,0,0)=0 and limω0+,ζ0μ=1\displaystyle\lim_{\omega\to 0+,\ \zeta\to 0}\mu=1. Note that

ωμ=ζ+lnpp(1ω(ζ+lnp))1p1,ζμ=ωp(1ω(ζ+lnp))1p1.\partial_{\omega}\mu=\frac{\zeta+\ln p}{p}\left(1-\omega(\zeta+\ln p)\right)^{-\frac{1}{p}-1},\quad\partial_{\zeta}\mu=\frac{\omega}{p}\left(1-\omega(\zeta+\ln p)\right)^{-\frac{1}{p}-1}.

On the other hand,

ωF2\displaystyle\partial_{\omega}F_{2} =\displaystyle= ω2μeζω1+(1+eζ+eζω1)ωμ\displaystyle\omega^{-2}\mu e^{\zeta-\omega^{-1}}+\left(-1+e^{\zeta}+e^{\zeta-\omega^{-1}}\right)\partial_{\omega}\mu
+eζ(eω1(pω11)1)ω1p+1r0(ω1pμ)+peζ(1+eω1)ω1pr0(ω1pμ)ω1p1(1pμ+ωωμ),\displaystyle+\frac{e^{\zeta}\left(e^{-\omega^{-1}}(p\omega^{-1}-1)-1\right)}{\omega^{\frac{1}{p}+1}}r_{0}(\omega^{\frac{1}{p}}\mu)+\frac{pe^{\zeta}(1+e^{-\omega^{-1}})}{\omega^{\frac{1}{p}}}r^{\prime}_{0}(\omega^{\frac{1}{p}}\mu)\cdot\omega^{\frac{1}{p}-1}\left(\frac{1}{p}\mu+\omega\partial_{\omega}\mu\right),
λF2\displaystyle\partial_{\lambda}F_{2} =\displaystyle= 1,\displaystyle-1,
ζF2\displaystyle\partial_{\zeta}F_{2} =\displaystyle= μ(eζ+eζω1)+(1+eζ+eζω1)ζμ+peζ(1+eω1)ω1p(r0(ω1pμ)+ω1pr0(ω1pμ)ζμ).\displaystyle\mu\left(e^{\zeta}+e^{\zeta-\omega^{-1}}\right)+\left(-1+e^{\zeta}+e^{\zeta-\omega^{-1}}\right)\partial_{\zeta}\mu+\frac{pe^{\zeta}(1+e^{-\omega^{-1}})}{\omega^{\frac{1}{p}}}\left(r_{0}(\omega^{\frac{1}{p}}\mu)+\omega^{\frac{1}{p}}r^{\prime}_{0}(\omega^{\frac{1}{p}}\mu)\partial_{\zeta}\mu\right).

Then F2C1F_{2}\in C^{1} and

ωF2(0,0,0)=0,λF2(0,0,0)=1,ζF2(0,0,0)=1.\partial_{\omega}F_{2}(0,0,0)=0,\quad\partial_{\lambda}F_{2}(0,0,0)=-1,\quad\partial_{\zeta}F_{2}(0,0,0)=1. (2.46)

Thus by the implicit function theorem, one can obtain that there exists a unique function ζ(ω,λ)C1\zeta(\omega,\lambda)\in C^{1} satisfying F2(ω,λ,ζ(ω,λ))=0F_{2}(\omega,\lambda,\zeta(\omega,\lambda))=0 and admitting such an expansion

ζ(ω,λ)=λ+O(ω2+λ2).\zeta(\omega,\lambda)=\lambda+O\left(\omega^{2}+\lambda^{2}\right). (2.47)

This yields

μ(s,λ)\displaystyle\mu(s,\lambda) =\displaystyle= (1sp(ζ(sp,λ)+lnp))1p\displaystyle\left(1-s^{p}\left(\zeta(s^{p},\lambda)+\ln p\right)\right)^{-\frac{1}{p}} (2.48)
=\displaystyle= (1sp(lnp+λ+O(s2p+λ2))1p\displaystyle\left(1-s^{p}(\ln p+\lambda+O\left(s^{2p}+\lambda^{2}\right)\right)^{-\frac{1}{p}}
=\displaystyle= 1+lnppsp+spλp+O(s2p+sp|λ|2)\displaystyle 1+\frac{\ln p}{p}s^{p}+\frac{s^{p}\lambda}{p}+O\left(s^{2p}+s^{p}|\lambda|^{2}\right)

and slμCl+ps^{l}\mu\in C^{l+p} for l=0,1l=0,1.

If we consider the case of μ<0\mu<0, then by the same method, one can obtain the expansion

μ(s,λ)=1lnppsp+spλp+O(s2p+sp|λ|2).\mu(s,\lambda)=-1-\frac{\ln p}{p}s^{p}+\frac{s^{p}\lambda}{p}+O\left(s^{2p}+s^{p}|\lambda|^{2}\right). (2.49)

Consequently, we complete the proof of (2.37) and (2.38).

3 Proof of Theorem 1.1

We now construct the shock curve x=φ(t)x=\varphi(t) of (1.1) in Ω0\Omega_{0}. It follows from the Rankine-Hugoniot condition that

{φ(t)=f(u0(y+(t,φ(t))))f(u0(y(t,φ(t))))u0(y+(t,φ(t)))u0(y(t,φ(t))),φ(1)=0.\left\{\begin{array}[]{l}\varphi^{\prime}(t)=\frac{f(u_{0}(y_{+}(t,\varphi(t))))-f(u_{0}(y_{-}(t,\varphi(t))))}{u_{0}(y_{+}(t,\varphi(t)))-u_{0}(y_{-}(t,\varphi(t)))},\\ \varphi(1)=0.\end{array}\right. (3.1)

This, together with the mean-value theorem, yields

g(y+(t,φ(t)))<φ(t)<g(y(t,φ(t))),g(y_{+}(t,\varphi(t)))<\varphi^{\prime}(t)<g(y_{-}(t,\varphi(t))), (3.2)

which means that the entropy condition on x=φ(t)x=\varphi(t) holds. Denote

a(x,y){f(u0(x))f(u0(y))u0(x)u0(y), if xy,g(x), if x=y.a(x,y)\triangleq\left\{\begin{array}[]{ll}\frac{f(u_{0}(x))-f(u_{0}(y))}{u_{0}(x)-u_{0}(y)},&\text{ if }x\neq y,\\ g(x),&\text{ if }x=y.\end{array}\right. (3.3)

Under assumption (1.7), it is easy to verify a(x,y)C2k+2(2)a(x,y)\in C^{2k+2}({\mathbb{R}}^{2}) with a(x,y)=12(x+y)+b(x,y)a(x,y)=-\frac{1}{2}(x+y)+b(x,y), where b(x,y)=O(x2+y2)C2k+2b(x,y)=O(x^{2}+y^{2})\in C^{2k+2}.


Lemma 3.1.

Under assumption (1.7), for (3.1) and small ε>0\varepsilon>0, there exists a solution x=φ(t)x=\varphi(t) on [1,1+ε)[1,1+\varepsilon) to satisfy
(1) φ(t)\varphi(t) is a C2C^{2} function on s[0,ε)s\in[0,\varepsilon), where s=(t1)12ks=(t-1)^{\frac{1}{2k}};
(2) φ(t)Ck+1k[1,1+ε)\varphi(t)\in C^{\frac{k+1}{k}}[1,1+\varepsilon).

Proof.

(1) Set λ(s)=φ(t)s2k+1\lambda(s)=\frac{\varphi(t)}{s^{2k+1}}. Then (3.1) becomes

{sλ(s)+(2k+2)λ(s)=sd(s,λ(s)),λ(0)=0,\left\{\begin{array}[]{l}s\lambda^{\prime}(s)+(2k+2)\lambda(s)=sd(s,\lambda(s)),\\ \lambda(0)=0,\end{array}\right. (3.4)

where

d(s,λ)=k(μ+(s,λ)+μ(s,λ)λ)s+2ks2b(sμ+(s,λ),sμ(s,λ)).d(s,\lambda)=-\frac{k(\mu_{+}(s,\lambda)+\mu_{-}(s,\lambda)-\lambda)}{s}+\frac{2k}{s^{2}}b(s\mu_{+}(s,\lambda),s\mu_{-}(s,\lambda)). (3.5)

By the proof procedure of Lemma 2.3, we have sjd(s,λ)Cjs^{j}d(s,\lambda)\in C^{j} for j=0,1,2j=0,1,2. Then by the same analysis in [11], there exists a unique solution λ(s)C1[0,ε)\lambda(s)\in C^{1}[0,\varepsilon) to (3.4) and further sλ(s)C1s\lambda^{\prime}(s)\in C^{1}. Due to (sλ(s))=sλ(s)+λ(s)(s\lambda(s))^{\prime}=s\lambda^{\prime}(s)+\lambda(s), then sλ(s)C2s\lambda(s)\in C^{2}, and s2λ(s)=s2d(2k+1)sλC2s^{2}\lambda^{\prime}(s)=s^{2}d-(2k+1)s\lambda\in C^{2}. Therefore, sφ(t)=12k[s2λ(s)+(2k+1)sλ(s)]s\rightarrow\varphi^{\prime}(t)=\frac{1}{2k}[s^{2}\lambda^{\prime}(s)+(2k+1)s\lambda(s)] is a C2C^{2} function.

(2) Let s0+s\rightarrow 0+ in (3.4), we have

λ(0)=d(0,0)2k+3=g(2k+2)(0)(2k+3)!+lims0+b(sμ+,sμ)s2=O(1).\lambda^{\prime}(0)=\frac{d(0,0)}{2k+3}=\frac{g^{(2k+2)}(0)}{(2k+3)!}+\lim_{s\rightarrow 0_{+}}\frac{b(s\mu_{+},s\mu_{-})}{s^{2}}=O(1).

Therefore λ(s)=O(s)\lambda(s)=O(s) and then φ(t)=O(s2k+2)=O((t1)k+1k)Ck+1k[1,1+ε)\varphi(t)=O(s^{2k+2})=O((t-1)^{\frac{k+1}{k}})\in C^{\frac{k+1}{k}}[1,1+\varepsilon).

Remark 3.1. The regularity of φ(t)\varphi(t) in Lemma 3.1 is optimal. Indeed, we consider the following problem

{tu+x(12u2)=0,u(0,x)=x+x2k+1+|x|2k+2+ε,ε>0.\left\{\begin{array}[]{ll}&\partial_{t}u+\partial_{x}(\frac{1}{2}u^{2})=0,\\ &\displaystyle u(0,x)=-x+x^{2k+1}+|x|^{2k+2+\varepsilon},\ \varepsilon>0.\end{array}\right.

In this case, we have g(x)=x+x2k+1+x2k+2g(x)=-x+x^{2k+1}+x^{2k+2} and

y+(t,φ(t))\displaystyle y_{+}(t,\varphi(t)) =\displaystyle= (t1)12k(1+φ(t)2k(t1)2k+12k(t1)12k2k)+o((t1)1k)\displaystyle(t-1)^{\frac{1}{2k}}(1+\frac{\varphi(t)}{2k(t-1)^{\frac{2k+1}{2k}}}-\frac{(t-1)^{\frac{1}{2k}}}{2k})+o((t-1)^{\frac{1}{k}})
y(t,φ(t))\displaystyle y_{-}(t,\varphi(t)) =\displaystyle= (t1)12k(1+φ(t)2k(t1)2k+12k(t1)12k2k)+o((t1)1k).\displaystyle(t-1)^{\frac{1}{2k}}(-1+\frac{\varphi(t)}{2k(t-1)^{\frac{2k+1}{2k}}}-\frac{(t-1)^{\frac{1}{2k}}}{2k})+o((t-1)^{\frac{1}{k}}). (3.6)

It follows from Rankine-Hugoniot condition that

φ(t)\displaystyle\varphi^{\prime}(t) =\displaystyle= y(t,φ(t))+y+(t,φ(t))2\displaystyle-\frac{y_{-}(t,\varphi(t))+y_{+}(t,\varphi(t))}{2}
=\displaystyle= φ(t)k(t1)(t1)1kk+o((t1)1k).\displaystyle\frac{\varphi(t)}{k(t-1)}-\frac{(t-1)^{\frac{1}{k}}}{k}+o((t-1)^{\frac{1}{k}}).

This derives φ(t)Ck+1k[1,1+ε)\varphi(t)\in C^{\frac{k+1}{k}}[1,1+\varepsilon) which is optimal.

Lemma 3.2.

Under assumption (1.7), for any c(2k(2k+1)12k+1,+)c\in(-\frac{2k}{(2k+1)^{\frac{1}{2k}+1}},+\infty), there exist ε=ε(c)>0\varepsilon=\varepsilon(c)>0 and δ=δ(c)>0\delta=\delta(c)>0 such that for (s,λ){0<s<ε,cδ<λ<c+δ}(s,\lambda)\in\{0<s<\varepsilon,c-\delta<\lambda<c+\delta\}, (s,λ)y+(t,x)(s,\lambda)\rightarrow y_{+}(t,x) has the expansion

y+(t,x)=s(μcμc2k+2g(2k+2)(0)(1+μc2k)(2k+2)!s+λc1+μc2k)+O(s3+s(λc)2),y_{+}(t,x)=s\left(\mu_{c}-\frac{\mu_{c}^{2k+2}g^{(2k+2)}(0)}{(-1+\mu_{c}^{2k})(2k+2)!}s+\frac{\lambda-c}{-1+\mu_{c}^{2k}}\right)+O(s^{3}+s(\lambda-c)^{2}), (3.7)

and for (s,λ){0<s<ε,cδ<λ<c+δ}(s,\lambda)\in\{0<s<\varepsilon,-c-\delta<\lambda<-c+\delta\}, (s,λ)y(t,x)(s,\lambda)\rightarrow y_{-}(t,x) has the expansion

y(t,x)=s(μcμc2k+2g(2k+2)(0)(1+μc2k)(2k+2)!s+λ+c1+μc2k)+O(s3+s(λ+c)2),y_{-}(t,x)=s\left(-\mu_{c}-\frac{\mu_{c}^{2k+2}g^{(2k+2)}(0)}{(-1+\mu_{c}^{2k})(2k+2)!}s+\frac{\lambda+c}{-1+\mu_{c}^{2k}}\right)+O(s^{3}+s(\lambda+c)^{2}), (3.8)

where μc\mu_{c} is the unique solution in (1(2k+1)12k,+)(\frac{1}{(2k+1)^{\frac{1}{2k}}},+\infty) of the equation

G(0,c,μ)=μ+μ2k+1c=0.G(0,c,\mu)=-\mu+\mu^{2k+1}-c=0. (3.9)
Proof.

By (2.26), (2.29) and (2.30), we have that

μG(0,±c,±μc)\displaystyle\partial_{\mu}G(0,\pm c,\pm\mu_{c}) =\displaystyle= 1+(2k+1)μc2k>0,\displaystyle-1+(2k+1)\mu_{c}^{2k}>0, (3.10)
sG(0,±c,±μc)\displaystyle\partial_{s}G(0,\pm c,\pm\mu_{c}) =\displaystyle= μc2k+2h(0)=μc2k+2g(2k+2)(0)(2k+2)!,\displaystyle\mu_{c}^{2k+2}h^{\prime}(0)=\frac{\mu_{c}^{2k+2}g^{(2k+2)}(0)}{(2k+2)!}, (3.11)
λG(0,±c,±μc)\displaystyle\partial_{\lambda}G(0,\pm c,\pm\mu_{c}) =\displaystyle= 1.\displaystyle-1. (3.12)

Then by the implicit function theorem one has that there is a unique function μ±(s,λ)\mu_{\pm}(s,\lambda) near (0,±c)(0,\pm c) satisfying G(s,λ,μ±(s,λ))0G(s,\lambda,\mu_{\pm}(s,\lambda))\equiv 0 and admitting the expansion

μ±(s,λ)=±μc11+μc2k(μc2k+2g(2k+2)(0)(2k+2)!s(λc))+O(s2+(λc)2).\mu_{\pm}(s,\lambda)=\pm\mu_{c}-\frac{1}{-1+\mu_{c}^{2k}}\left(\frac{\mu_{c}^{2k+2}g^{(2k+2)}(0)}{(2k+2)!}s-(\lambda-c)\right)+O(s^{2}+(\lambda-c)^{2}). (3.13)

Thus (3.7) and (3.8) are proved. ∎

To study the asymptotic behavior of y±y_{\pm} near the xx-axis, we now take the following transform

ξ=x12k+1,η=t1ξ2k,ν=yξ.\xi=x^{\frac{1}{2k+1}},\ \eta=\frac{t-1}{\xi^{2k}},\ \nu=\frac{y}{\xi}. (3.14)

Under assumption (1.6), by divided ξ2k+1\xi^{2k+1}, (2.3) then becomes

H(η,ξ,ν)ην+(1+ξ2kη)ν2k+1+(1+ξ2kη)ν2k+1h(ξν)1=0.H(\eta,\xi,\nu)\triangleq-\eta\nu+(1+\xi^{2k}\eta)\nu^{2k+1}+(1+\xi^{2k}\eta)\nu^{2k+1}h(\xi\nu)-1=0. (3.15)

We now have

Lemma 3.3.

Under assumption (1.6), for small δ>0\delta>0, there exists some small constant ε>0\varepsilon>0 such that for (η,ξ){|η|ε, 0<ξ<δ}(\eta,\xi)\in\{|\eta|\leq\varepsilon,\ 0<\xi<\delta\}, we can get the expansion of y+(t,x)y_{+}(t,x) on (ξ,η)(\xi,\eta)

y+(t,x)=ξ(1+η2k+1g(2k+2)(0)(2k+2)!ξ)+O(η2ξ+ξ3),y_{+}(t,x)=\xi\left(1+\frac{\eta}{2k+1}-\frac{g^{(2k+2)}(0)}{(2k+2)!}\xi\right)+O(\eta^{2}\xi+\xi^{3}), (3.16)

and for (η,ξ){|η|ε,δ<ξ<0}(\eta,\xi)\in\{|\eta|\leq\varepsilon,\ -\delta<\xi<0\}, we can get

y(t,x)=ξ(1η2k+1g(2k+2)(0)(2k+2)!ξ)+O(η2ξ+ξ3).y_{-}(t,x)=\xi\left(1-\frac{\eta}{2k+1}-\frac{g^{(2k+2)}(0)}{(2k+2)!}\xi\right)+O(\eta^{2}\xi+\xi^{3}). (3.17)
Proof.

It follows from direct computation that H(0,0,1)=0H(0,0,1)=0 and

νH\displaystyle\partial_{\nu}H =\displaystyle= η+(2k+1)(1+ξ2kη)ν2k+(2k+1)(1+ξ2kη)ν2kh(ξν)+(1+ξ2kη)ν2k+1ξh(ξν),\displaystyle-\eta+(2k+1)(1+\xi^{2k}\eta)\nu^{2k}+(2k+1)(1+\xi^{2k}\eta)\nu^{2k}h(\xi\nu)+(1+\xi^{2k}\eta)\nu^{2k+1}\xi h^{\prime}(\xi\nu), (3.18)
ηH\displaystyle\partial_{\eta}H =\displaystyle= ν+ξ2kν2k+1+ξ2kν2k+1h(ξν),\displaystyle-\nu+\xi^{2k}\nu^{2k+1}+\xi^{2k}\nu^{2k+1}h(\xi\nu), (3.19)
ξH\displaystyle\partial_{\xi}H =\displaystyle= 2kξ2k1ην2k+1+2kξ2k1ην2k+1h(ξη)+(1+ξ2kη)ν2k+1νh(ξν).\displaystyle 2k\xi^{2k-1}\eta\nu^{2k+1}+2k\xi^{2k-1}\eta\nu^{2k+1}h(\xi\eta)+(1+\xi^{2k}\eta)\nu^{2k+1}\nu h^{\prime}(\xi\nu). (3.20)

Then

νH(0,0,1)\displaystyle\partial_{\nu}H(0,0,1) =\displaystyle= 2k+1,\displaystyle 2k+1, (3.21)
ηH(0,0,1)\displaystyle\partial_{\eta}H(0,0,1) =\displaystyle= 1,\displaystyle-1, (3.22)
ξH(0,0,1)\displaystyle\partial_{\xi}H(0,0,1) =\displaystyle= h(0)=g(2k+2)(0)(2k+2)!.\displaystyle h^{\prime}(0)=\frac{g^{(2k+2)}(0)}{(2k+2)!}. (3.23)

By the implicit function theorem, there exists a unique solution ν=ν(η,ξ)\nu=\nu(\eta,\xi) of (3.15) near (η,ξ)=(0,0)(\eta,\xi)=(0,0) such that

ν(η,ξ)=1+η2k+1g(2k+2)(0)(2k+2)!ξ+O(η2+ξ2).\nu(\eta,\xi)=1+\frac{\eta}{2k+1}-\frac{g^{(2k+2)}(0)}{(2k+2)!}\xi+O(\eta^{2}+\xi^{2}). (3.24)

Therefore, we obtain (3.16) for ξ>0\xi>0. Analogously, (3.17) holds for ξ<0\xi<0. ∎

Next we consider the asymptotic behavior of y(t,x)y(t,x) near the blowup point (1,0)(1,0) in the domain {(t,x):t<1}\{(t,x):\ t<1\}. Without confusions, we still use the same notation as for t>1t>1. Note that through each point in {t<1}\{t<1\}, there exists a unique characteristic line. By taking the following transform similar to (2.20)

τ=1t,s=τ12k,μ=ys,λ=xs2k+1,\tau=1-t,\ s=\tau^{\frac{1}{2k}},\ \mu=\frac{y}{s},\ \lambda=\frac{x}{s^{2k+1}}, (3.25)

and then by divided s2k+1s^{2k+1} on two sides of (2.3), then (2.3) becomes

G(s,λ,μ)μ+(1s2k)μ2k+1+(1s2k)μ2k+1h(sμ)λ=0.G(s,\lambda,\mu)\triangleq\mu+(1-s^{2k})\mu^{2k+1}+(1-s^{2k})\mu^{2k+1}h(s\mu)-\lambda=0. (3.26)
Lemma 3.4.

For each cc\in{\mathbb{R}}, there exist ε=ε(c),δ=δ(c)>0\varepsilon=\varepsilon(c),\delta=\delta(c)>0 such that for (s,λ){0<s<ε,cδ<λ<c+δ}(s,\lambda)\in\{0<s<\varepsilon,c-\delta<\lambda<c+\delta\}, (s,λ)y(t,x)(s,\lambda)\rightarrow y(t,x) has the expansion

y(t,x)=s(μcμc2k+2g(2k+2)(0)(1+(2k+1)μc2k)(2k+2)!sλc1+(2k+1)μc2k)+O(s3+s(λc)2),y(t,x)=s\left(\mu_{c}-\frac{\mu_{c}^{2k+2}g^{(2k+2)}(0)}{(1+(2k+1)\mu_{c}^{2k})(2k+2)!}s-\frac{\lambda-c}{1+(2k+1)\mu_{c}^{2k}}\right)+O(s^{3}+s(\lambda-c)^{2}), (3.27)

where μc\mu_{c} is the unique solution of the equation

G(0,c,μ)=μ+μ2k+1c=0.G(0,c,\mu)=\mu+\mu^{2k+1}-c=0. (3.28)
Proof.

It follows from direct computation that

μG(s,λ,μ)\displaystyle\partial_{\mu}G(s,\lambda,\mu) =\displaystyle= 1+(2k+1)(1s2k)μ2k+(2k+1)(1s2k)μ2k+1sh(sμ),\displaystyle 1+(2k+1)(1-s^{2k})\mu^{2k}+(2k+1)(1-s^{2k})\mu^{2k+1}sh^{\prime}(s\mu), (3.29)
sG(s,λ,μ)\displaystyle\partial_{s}G(s,\lambda,\mu) =\displaystyle= 2ks2k1μ2k+12ks2k1μ2k+1h(sμ)+(1s2k)μ2k+2h(sμ),\displaystyle-2ks^{2k-1}\mu^{2k+1}-2ks^{2k-1}\mu^{2k+1}h(s\mu)+(1-s^{2k})\mu^{2k+2}h^{\prime}(s\mu), (3.30)
λG(s,λ,μ)\displaystyle\partial_{\lambda}G(s,\lambda,\mu) =\displaystyle= 1.\displaystyle-1. (3.31)

Then

μG(0,c,μc)\displaystyle\partial_{\mu}G(0,c,\mu_{c}) =\displaystyle= 1+(2k+1)μc2k>0,\displaystyle 1+(2k+1)\mu_{c}^{2k}>0, (3.32)
sG(0,c,μc)\displaystyle\partial_{s}G(0,c,\mu_{c}) =\displaystyle= μc2k+2h(0)=μc2k+2g(2k+2)(0)(2k+2)!,\displaystyle\mu_{c}^{2k+2}h^{\prime}(0)=\frac{\mu_{c}^{2k+2}g^{(2k+2)}(0)}{(2k+2)!}, (3.33)
λG(0,c,μc)\displaystyle\partial_{\lambda}G(0,c,\mu_{c}) =\displaystyle= 1.\displaystyle-1. (3.34)

By the implicit function theorem, there exists a μ(s,λ)\mu(s,\lambda) near (c,μc)(c,\mu_{c}) satisfying

μ(s,λ)=μc11+(2k+1)μc2k(μc2k+2g(2k+2)(0)(2k+2)!s(λc))+O(s2+(λc)2),\mu(s,\lambda)=\mu_{c}-\frac{1}{1+(2k+1)\mu_{c}^{2k}}\left(\frac{\mu_{c}^{2k+2}g^{(2k+2)}(0)}{(2k+2)!}s-(\lambda-c)\right)+O(s^{2}+(\lambda-c)^{2}), (3.35)

from which we can deduce (3.27). ∎


We start to prove Theorem 1.1.

Proof of Theorem 1.1:

(1) By Lemma 3.1, φ(t)Ck+1k[1,1+ε)\varphi(t)\in C^{\frac{k+1}{k}}[1,1+\varepsilon) and uC1((1,1+ε)×){x=φ(t)})u\in C^{1}((1,1+\varepsilon)\times{\mathbb{R}})\setminus\{x=\varphi(t)\}) for small ε>0\varepsilon>0 have been shown.

(2) Let δ,ε>0\delta,\ \varepsilon>0 be the constants obtained in Lemma 3.3 and denote

Ωx,+\displaystyle\Omega_{x,+} =\displaystyle= B{(t,x): 0<x<δ2k+1,|t1|<εx2k2k+1},\displaystyle B\cap\{(t,x):\ 0<x<\delta^{2k+1},\ |t-1|<\varepsilon x^{\frac{2k}{2k+1}}\}, (3.36)
Ωx,\displaystyle\Omega_{x,-} =\displaystyle= B{(t,x):δ2k+1<x<0,|t1|<ε(x)2k2k+1},\displaystyle B\cap\{(t,x):\ -\delta^{2k+1}<x<0,\ |t-1|<\varepsilon(-x)^{\frac{2k}{2k+1}}\}, (3.37)
Ω0\displaystyle\Omega_{0} =\displaystyle= B{(t,x):t<1,|x|12k+1<2ε12k(1t)12k}.\displaystyle B\cap\{(t,x):\ t<1,\ |x|^{\frac{1}{2k+1}}<\frac{2}{\varepsilon^{\frac{1}{2k}}}(1-t)^{\frac{1}{2k}}\}. (3.38)

Let c0(0,2k(2k+1)1+12k)c_{0}\in(0,\frac{2k}{(2k+1)^{1+\frac{1}{2k}}}) be some fixed constant and denote

Ωt,+\displaystyle\Omega_{t,+} =\displaystyle= B{(t,x):c0(t1)12k<x12k+1<2ε12k(t1)12k},\displaystyle B\cap\{(t,x):\ -c_{0}(t-1)^{\frac{1}{2k}}<x^{\frac{1}{2k+1}}<\frac{2}{\varepsilon^{\frac{1}{2k}}}(t-1)^{\frac{1}{2k}}\}, (3.39)
Ωt,\displaystyle\Omega_{t,-} =\displaystyle= B{(t,x):2ε12k(t1)12k<x12k+1<c0(t1)12k}.\displaystyle B\cap\{(t,x):\ -\frac{2}{\varepsilon^{\frac{1}{2k}}}(t-1)^{\frac{1}{2k}}<x^{\frac{1}{2k+1}}<c_{0}(t-1)^{\frac{1}{2k}}\}. (3.40)

It is easy to see that for (t,x)Ωx,+Ωt,+(t,x)\in\Omega_{x,+}\cup\Omega_{t,+}, u(t,x)=u0(y+(t,x))u(t,x)=u_{0}(y_{+}(t,x)); for (t,x)Ωx,Ωt,(t,x)\in\Omega_{x,-}\cup\Omega_{t,-}, u(t,x)=u0(y(t,x))u(t,x)=u_{0}(y_{-}(t,x)); for (t,x)Ω0(t,x)\in\Omega_{0}, u(t,x)=u0(y(t,x))u(t,x)=u_{0}(y(t,x)). By Heine-Borel property of compactness, there exist {cj,±,δj,±=δj,±(cj,±),εj,±=εj,±(cj,±)}j=1n\{c_{j,\pm},\delta_{j,\pm}=\delta_{j,\pm}(c_{j,\pm}),\varepsilon_{j,\pm}=\varepsilon_{j,\pm}(c_{j,\pm})\}_{j=1}^{n} and {cj,0,δj,0=δj,0(cj,0),εj,0=εj,0(cj,0)}j=1n\{c_{j,0},\delta_{j,0}=\delta_{j,0}(c_{j,0}),\varepsilon_{j,0}=\varepsilon_{j,0}(c_{j,0})\}_{j=1}^{n} such that

Ωt,+j=1nΩt,+j,Ωt,j=1nΩt,j,Ω0j=1nΩ0j,\Omega_{t,+}\subset\cup_{j=1}^{n}\Omega_{t,+}^{j},\ \Omega_{t,-}\subset\cup_{j=1}^{n}\Omega_{t,-}^{j},\ \Omega_{0}\subset\cup_{j=1}^{n}\Omega_{0}^{j}, (3.41)

where

Ωt,+j\displaystyle\Omega_{t,+}^{j} =\displaystyle= {(t,x): 0<(t1)12k<εj,+,cj,+δj,+<x(t1)2k+12k<cj,++δj,+},\displaystyle\{(t,x):\ 0<(t-1)^{\frac{1}{2k}}<\varepsilon_{j,+},\ c_{j,+}-\delta_{j,+}<\frac{x}{(t-1)^{\frac{2k+1}{2k}}}<c_{j,+}+\delta_{j,+}\}, (3.42)
Ωt,j\displaystyle\Omega_{t,-}^{j} =\displaystyle= {(t,x): 0<(t1)12k<εj,,cj,δj,<x(t1)2k+12k<cj,+δj,},\displaystyle\{(t,x):\ 0<(t-1)^{\frac{1}{2k}}<\varepsilon_{j,-},\ c_{j,-}-\delta_{j,-}<\frac{x}{(t-1)^{\frac{2k+1}{2k}}}<c_{j,-}+\delta_{j,-}\}, (3.43)
Ω0j\displaystyle\Omega_{0}^{j} =\displaystyle= {(t,x): 0<(1t)12k<εj,0,cj,+δj,+<x(1t)2k+12k<cj,++δj,+},\displaystyle\{(t,x):\ 0<(1-t)^{\frac{1}{2k}}<\varepsilon_{j,0},\ c_{j,+}-\delta_{j,+}<\frac{x}{(1-t)^{\frac{2k+1}{2k}}}<c_{j,+}+\delta_{j,+}\}, (3.44)

and these domains satisfy the corresponding properties in Lemma 3.2 and 3.4.

Set B={(t,x): 0<(t1)2+x2<ρ}B=\{(t,x):\ 0<\sqrt{(t-1)^{2}+x^{2}}<\rho\}, and choose ρ>0\rho>0 sufficiently small such that

B=Ωx,+Ωx,Ωt,+Ωt,Ω0.B=\Omega_{x,+}\cup\Omega_{x,-}\cup\Omega_{t,+}\cup\Omega_{t,-}\cup\Omega_{0}. (3.45)

We now establish the behaviors of uu and its derivatives near (1,0)(1,0). It suffices to only consider this in the domains Ωx,+\Omega_{x,+}, Ωt,+j\Omega_{t,+}^{j} and Ω0j\Omega_{0}^{j} since the other cases can be treated analogously.

For (t,x)Ωx,+(t,x)\in\Omega_{x,+}, we have

|u(t,x)u(1,0)|=|u0(y+(t,x))||y+(t,x)|x12k+1;|u(t,x)-u(1,0)|=|u_{0}(y_{+}(t,x))|\lesssim|y_{+}(t,x)|\lesssim x^{\frac{1}{2k+1}}; (3.46)

for (t,x)Ωt,+j(t,x)\in\Omega_{t,+}^{j},

|u(t,x)u(1,0)|=|u0(y+(t,x))||y+(t,x)|(t1)12k;|u(t,x)-u(1,0)|=|u_{0}(y_{+}(t,x))|\lesssim|y_{+}(t,x)|\lesssim(t-1)^{\frac{1}{2k}}; (3.47)

and for (t,x)Ω0j(t,x)\in\Omega_{0}^{j},

|u(t,x)u(1,0)|=|u0(y(t,x))||y(t,x)|(1t)12k.|u(t,x)-u(1,0)|=|u_{0}(y(t,x))|\lesssim|y(t,x)|\lesssim(1-t)^{\frac{1}{2k}}. (3.48)

Therefore (1.12) is obtained.

Let’s turn to prove the estimates (1.13) and (1.14). Note that

{ux=u0(y(t,x))yx(t,x)=u0(y(t,x))1+tg(y(t,x)),ut=u0(y(t,x))yt(t,x)=u0(y(t,x))g(y(t,x))1+tg(y(t,x)).\left\{\begin{array}[]{ll}&\frac{\partial u}{\partial x}=u^{\prime}_{0}(y(t,x))\frac{\partial y}{\partial x}(t,x)=\frac{u^{\prime}_{0}(y(t,x))}{1+tg^{\prime}(y(t,x))},\\ &\frac{\partial u}{\partial t}=u^{\prime}_{0}(y(t,x))\frac{\partial y}{\partial t}(t,x)=-\frac{u^{\prime}_{0}(y(t,x))g(y(t,x))}{1+tg^{\prime}(y(t,x))}.\end{array}\right. (3.49)

For (t,x)Ωt,+j(t,x)\in\Omega_{t,+}^{j}, by (3.7) in Lemma 3.2 we have

1+tg(y+(t,x))\displaystyle 1+tg^{\prime}(y_{+}(t,x)) =\displaystyle= s2k+(2k+1)(1+s2k)y+2k(t,x)+O(y+2k+1(t,x))\displaystyle-s^{2k}+(2k+1)(1+s^{2k})y_{+}^{2k}(t,x)+O(y_{+}^{2k+1}(t,x)) (3.50)
=\displaystyle= (1+(2k+1)μcj,+2k)s2k+O(s2k+1+s2k|λcj,+|)\displaystyle(-1+(2k+1)\mu_{c_{j,+}}^{2k})s^{2k}+O(s^{2k+1}+s^{2k}|\lambda-c_{j,+}|)
\displaystyle\gtrsim s2k\displaystyle s^{2k}
=\displaystyle= (t1)\displaystyle(t-1)
\displaystyle\gtrsim |t1|+|x|2k2k+1,\displaystyle|t-1|+|x|^{\frac{2k}{2k+1}},

where s=t1s=t-1, and the fact of 1+(2k+1)μcj,+2k>0-1+(2k+1)\mu_{c_{j,+}}^{2k}>0 has been used.

For (t,x)Ωx,+(t,x)\in\Omega_{x,+}, by (3.16) in Lemma 3.3 we have

1+tg(y(t,x))\displaystyle 1+tg^{\prime}(y(t,x)) =\displaystyle= ηξ2k+(2k+1)(1+ηξ2k)y2k(t,x)+O(y2k+1(t,x))\displaystyle-\eta\xi^{2k}+(2k+1)(1+\eta\xi^{2k})y^{2k}(t,x)+O(y^{2k+1}(t,x)) (3.51)
=\displaystyle= ξ2k+O(ξ2k+1+ηξ2k)\displaystyle\xi^{2k}+O(\xi^{2k+1}+\eta\xi^{2k})
\displaystyle\gtrsim ξ2k\displaystyle\xi^{2k}
=\displaystyle= x2k2k+1\displaystyle x^{\frac{2k}{2k+1}}
\displaystyle\gtrsim |t1|+|x|2k2k+1,\displaystyle|t-1|+|x|^{\frac{2k}{2k+1}},

where ξ=x12k+1\xi=x^{\frac{1}{2k+1}}.

For (t,x)Ω0,+j(t,x)\in\Omega_{0,+}^{j}, by (3.27) in Lemma 3.4 we arrive at

1+tg(y(t,x))\displaystyle 1+tg^{\prime}(y(t,x)) =\displaystyle= s2k+(2k+1)(1+s2k)y2k(t,x)+O(y2k+1(t,x))\displaystyle s^{2k}+(2k+1)(1+s^{2k})y^{2k}(t,x)+O(y^{2k+1}(t,x)) (3.52)
=\displaystyle= (1+(2k+1)μcj,02k)s2k+O(s2k+1+s2k|λcj,0|)\displaystyle(1+(2k+1)\mu_{c_{j,0}}^{2k})s^{2k}+O(s^{2k+1}+s^{2k}|\lambda-c_{j,0}|)
\displaystyle\gtrsim s2k\displaystyle s^{2k}
=\displaystyle= (1t)\displaystyle(1-t)
\displaystyle\gtrsim |t1|+|x|2k2k+1,\displaystyle|t-1|+|x|^{\frac{2k}{2k+1}},

where s=1ts=1-t, and the fact of 1+(2k+1)μcj,02k>01+(2k+1)\mu_{c_{j,0}}^{2k}>0 has been used.

Therefore, 1+tg(y(t,x))|t1|+|x|2k2k+11+tg^{\prime}(y(t,x))\gtrsim|t-1|+|x|^{\frac{2k}{2k+1}} holds for (t,x)B(t,x)\in B. In light of (3.49) and the fact of g(y(t,x))y(t,x)g(y(t,x))\sim y(t,x) in BB, (1.13)-(1.14) hold and thus the proof of Theorem 1.1 is completed.

\square

4 Proof of Theorem 1.2

By the characteristics method, we can define u±(t,x)=u0(y±(t,x))u_{\pm}(t,x)=u_{0}(y_{\pm}(t,x)). By (1.10), the shock curve x=φ(t)x=\varphi(t) satisfies

{φ(t)=f(u0(y+(t,φ(t))))f(u0(y(t,φ(t))))u0(y+(t,φ(t)))u0(y(t,φ(t))),φ(1)=0.\linespread{1.2}\begin{cases}&\varphi^{\prime}(t)=\frac{f(u_{0}(y_{+}(t,\varphi(t))))-f(u_{0}(y_{-}(t,\varphi(t))))}{u_{0}(y_{+}(t,\varphi(t)))-u_{0}(y_{-}(t,\varphi(t)))},\\ &\varphi(1)=0.\end{cases} (4.1)

Denote

a(x,y){f(u0(x))f(u0(y))u0(x)u0(y), if xy,g(x), if x=y.a(x,y)\triangleq\linespread{1.2}\begin{cases}&\displaystyle\frac{f(u_{0}(x))-f(u_{0}(y))}{u_{0}(x)-u_{0}(y)},\text{ if }x\neq y,\\ &g(x),\text{ if }x=y.\end{cases} (4.2)

By (1.8), we have a(x,y)a(x,y) in C(2)C^{\infty}({\mathbb{R}}^{2}) and

a(x,y)=12(x+y)+b(x,y),a(x,y)=-\frac{1}{2}(x+y)+b(x,y), (4.3)

where b(x,y)=b(y,x)b(x,y)=b(y,x) and b(x,y)=O(x2+y2)Cb(x,y)=O(x^{2}+y^{2})\in C^{\infty}.

We now study the regularity of the shock wave x=φ(t)x=\varphi(t) as a function of s=|ln(t1)|1ps=|\ln(t-1)|^{-\frac{1}{p}}.

Lemma 4.1.

Under assumption (1.8), for (4.1) and small ε>0\varepsilon>0, there exists a solution x=φ(t)x=\varphi(t) on [1,1+ε)[1,1+\varepsilon) such that
(1) sφ(t)s\rightarrow\varphi(t) is of C1C^{1} on [0,ε)[0,\varepsilon);
(2) x=φ(t)x=\varphi(t) is of C1C^{1} on [1,1+ε][1,1+\varepsilon] with the behavior φ(t)=O(s2τ)\varphi(t)=O(s^{2}\tau).

Proof.

(1) Set λ(s)=φ(t)sτ\lambda(s)=\frac{\varphi(t)}{s\tau}. Then

dφdt(t)=s(s1+ppdλ(s)ds+(spp+1)λ(s)).\frac{d\varphi}{dt}(t)=s\left(\frac{s^{1+p}}{p}\frac{d\lambda(s)}{ds}+(\frac{s^{p}}{p}+1)\lambda(s)\right).

Substituting this into (4.1) yields

s1+ppdλ(s)ds+(spp+1)λ(s)\displaystyle\frac{s^{1+p}}{p}\frac{d\lambda(s)}{ds}+(\frac{s^{p}}{p}+1)\lambda(s) (4.4)
=\displaystyle= 1sa(sμ+(s,λ(s)),sμ(s,λ(s)))\displaystyle\frac{1}{s}a(s\mu_{+}(s,\lambda(s)),s\mu_{-}(s,\lambda(s)))
=\displaystyle= 12(μ+(s,λ(s))+μ(s,λ(s)))+1sb(sμ+(s,λ(s)),sμ(s,λ(s)))\displaystyle-\frac{1}{2}\left(\mu_{+}(s,\lambda(s))+\mu_{-}(s,\lambda(s))\right)+\frac{1}{s}b(s\mu_{+}(s,\lambda(s)),s\mu_{-}(s,\lambda(s)))
=\displaystyle= spλp+O(s+s2p+1|λ|+sp+1|λ|2).\displaystyle-\frac{s^{p}\lambda}{p}+O\left(s+s^{2p+1}|\lambda|+s^{p+1}|\lambda|^{2}\right). (4.5)

By (2.37) and (2.38), we have

d(s,λ)1sa(sμ+(s,λ),sμ(s,λ))+sppλ=O(s+s2p+1|λ|+sp+1|λ|2).d(s,\lambda)\triangleq\frac{1}{s}a(s\mu_{+}(s,\lambda),s\mu_{-}(s,\lambda))+\frac{s^{p}}{p}\lambda=O\left(s+s^{2p+1}|\lambda|+s^{p+1}|\lambda|^{2}\right). (4.6)

Moreover, sld(s,λ)Cl+ps^{l}d(s,\lambda)\in C^{l+p} holds for l=0,1l=0,1, which is derived by Lemma 2.4 and

d(sd(s,λ(s)))ds=O(1+s2p|λ|+s2p+1|λ|+sp|λ|2+sp+2|λ|2).\frac{d(sd(s,\lambda(s)))}{ds}=O(1+s^{2p}|\lambda|+s^{2p+1}|\lambda^{\prime}|+s^{p}|\lambda|^{2}+s^{p+2}|\lambda^{\prime}|^{2}). (4.7)

In addition, (4.1) in (s,λ)(s,\lambda) can be written as

{s1+ppdλ(s)ds+(2psp+1)λ(s)=d(s,λ(s)),λ(0)=0.\linespread{1.2}\begin{cases}&\frac{s^{1+p}}{p}\frac{d\lambda(s)}{ds}+(\frac{2}{p}s^{p}+1)\lambda(s)=d(s,\lambda(s)),\\ &\lambda(0)=0.\end{cases} (4.8)

This yields

λ(s)=ps20sω1pespωpd(ω,λ(ω))𝑑ω.\lambda(s)=ps^{-2}\int_{0}^{s}\omega^{1-p}e^{s^{-p}-\omega^{-p}}d(\omega,\lambda(\omega))d\omega. (4.9)

It follows from direct computation that

|λ(s)|\displaystyle|\lambda(s)| \displaystyle\leq s20ss2espω1peωp|d(ω,λ(ω))|dω)\displaystyle s^{-2}\int_{0}^{s}s^{2}e^{s^{-p}}\omega^{-1-p}e^{-\omega^{-p}}|d(\omega,\lambda(\omega))|d\omega)
\displaystyle\lesssim (s+s2p+1λL[0,s]+sp+1λL[0,s]2)0sespω1peωp𝑑ω\displaystyle(s+s^{2p+1}\|\lambda\|_{L^{\infty}[0,s]}+s^{p+1}\|\lambda\|^{2}_{L^{\infty}[0,s]})\int_{0}^{s}e^{s^{-p}}\omega^{-1-p}e^{-\omega^{-p}}d\omega
\displaystyle\lesssim s+s2p+1λL[0,s]+sp+1λL[0,s]2.\displaystyle s+s^{2p+1}\|\lambda\|_{L^{\infty}[0,s]}+s^{p+1}\|\lambda\|^{2}_{L^{\infty}[0,s]}.

Thus λL[0,s]Cs\|\lambda\|_{L^{\infty}[0,s]}\leq Cs for s(0,ε]s\in(0,\varepsilon] and small ε>0\varepsilon>0. By the analogous computation, we can apply the contraction mapping theorem to prove that there exists a continuous solution λ\lambda to the integral equation (4.9). From (4.9), we have

λ(s)\displaystyle\lambda^{\prime}(s) =\displaystyle= ps1pd(s,λ(s))2ps30sω1pespωpd(ω,λ(ω))𝑑ωpsp3esp0sω2d(ω,λ(ω))𝑑eωp\displaystyle ps^{-1-p}d(s,\lambda(s))-2ps^{-3}\int_{0}^{s}\omega^{1-p}e^{s^{-p}-\omega^{-p}}d(\omega,\lambda(\omega))d\omega-ps^{-p-3}e^{s^{-p}}\int_{0}^{s}\omega^{2}d(\omega,\lambda(\omega))de^{-\omega^{-p}}
=\displaystyle= 2ps30sω1pespωpd(ω,λ(ω))𝑑ω+psp3esp0seωp(ωd(ω,λ(ω))+ωd(ωd(ω,λ(ω)))dω)𝑑ω.\displaystyle-2ps^{-3}\int_{0}^{s}\omega^{1-p}e^{s^{-p}-\omega^{-p}}d(\omega,\lambda(\omega))d\omega+ps^{-p-3}e^{s^{-p}}\int_{0}^{s}e^{-\omega^{-p}}\big{(}\omega d(\omega,\lambda(\omega))+\omega\frac{d(\omega d(\omega,\lambda(\omega)))}{d\omega}\big{)}d\omega.

This derives

|λ(s)|\displaystyle|\lambda^{\prime}(s)| \displaystyle\lesssim s30ss2ω1pespωp|d(ω,λ(ω))|𝑑ω+sp30ss2+pω1pespωp|d(ω,λ(ω))|𝑑ω\displaystyle s^{-3}\int_{0}^{s}s^{2}\omega^{-1-p}e^{s^{-p}-\omega^{-p}}|d(\omega,\lambda(\omega))|d\omega+s^{-p-3}\int_{0}^{s}s^{2+p}\omega^{-1-p}e^{s^{-p}-\omega^{-p}}|d(\omega,\lambda(\omega))|d\omega
+sp3esp0ss2+pω1peωp|d(ωd(ω,λ(ω)))dω|𝑑ω\displaystyle+s^{-p-3}e^{s^{-p}}\int_{0}^{s}s^{2+p}\omega^{-1-p}e^{-\omega^{-p}}|\frac{d(\omega d(\omega,\lambda(\omega)))}{d\omega}|d\omega
\displaystyle\lesssim 1+λL[0,s]s+s1esp0sω1peωp(s+s2λL[0,s]+λL[0,s]2)𝑑ω\displaystyle 1+\frac{\|\lambda\|_{L^{\infty}[0,s]}}{s}+s^{-1}e^{s^{-p}}\int_{0}^{s}\omega^{-1-p}e^{-\omega^{-p}}(s+s^{2}\|\lambda^{\prime}\|_{L^{\infty}[0,s]}+\|\lambda\|^{2}_{L^{\infty}[0,s]})d\omega
\displaystyle\lesssim 1+λL[0,s]s+sλL[0,s],\displaystyle 1+\frac{\|\lambda\|_{L^{\infty}[0,s]}}{s}+s\|\lambda^{\prime}\|_{L^{\infty}[0,s]},

and then λ(s)C[0,ε]\lambda^{\prime}(s)\in C[0,\varepsilon] can be shown. In addition, by φ(t)=sτλ(s)\varphi(t)=s\tau\lambda(s) and s=|lnτ|1ps=|\ln\tau|^{-\frac{1}{p}}, then φ(t)=O(s2τ)\varphi(t)=O(s^{2}\tau) holds.

Remark 4.1. The regularity of φ(t)\varphi(t) in Lemma 4.1 is optimal. Indeed, we consider Burgers’ equation

{ut+x(12u2)=0,u(0,x)=x+1pe|x|p(x+x2),p>0.\left\{\begin{array}[]{ll}&\displaystyle\frac{\partial u}{\partial t}+\frac{\partial}{\partial x}(\frac{1}{2}u^{2})=0,\\ &\displaystyle u(0,x)=-x+\frac{1}{p}e^{-|x|^{-p}}\left(x+x^{2}\right),\ p>0.\end{array}\right.

In this case, g(x)=u0(x)=x+1pe|x|p(x+x2)g(x)=u_{0}(x)=-x+\frac{1}{p}e^{-|x|^{-p}}\left(x+x^{2}\right). On the other hand, (2.39) can be written as

F(s,λ,ζ)G(s,λ,μ)=μ(1+eζ)+μeζsp+sμ2eζ(1+esp)λ,F(s,\lambda,\zeta)\triangleq G(s,\lambda,\mu)=\mu(-1+e^{\zeta})+\mu e^{\zeta-s^{-p}}+s\mu^{2}e^{\zeta}(1+e^{-s^{-p}})-\lambda,

which derives Fs|s=λ=ζ=0=1\frac{\partial F}{\partial s}|_{s=\lambda=\zeta=0}=1. So we have that

y+(t,x)\displaystyle y_{+}(t,x) =\displaystyle= |ln(t1)|1p(1+lnpp|ln(t1)|1)+φ(t)p(t1)|ln(t1)|+1p|ln(t1)|12p+o(|ln(t1)|12p),\displaystyle|\ln(t-1)|^{-\frac{1}{p}}(1+\frac{\ln p}{p}|\ln(t-1)|^{-1})+\frac{\varphi(t)}{p(t-1)|\ln(t-1)|}+\frac{1}{p}|\ln(t-1)|^{-1-\frac{2}{p}}+o(|\ln(t-1)|^{-1-\frac{2}{p}}),
y(t,x)\displaystyle y_{-}(t,x) =\displaystyle= |ln(t1)|1p(1+lnpp|ln(t1)|1)+φ(t)p(t1)|ln(t1)|+1p|ln(t1)|12p+o(|ln(t1)|12p).\displaystyle-|\ln(t-1)|^{-\frac{1}{p}}(1+\frac{\ln p}{p}|\ln(t-1)|^{-1})+\frac{\varphi(t)}{p(t-1)|\ln(t-1)|}+\frac{1}{p}|\ln(t-1)|^{-1-\frac{2}{p}}+o(|\ln(t-1)|^{-1-\frac{2}{p}}).

It follows from Rankine-Hugoniot condition that

φ(t)\displaystyle\varphi^{\prime}(t) =\displaystyle= y(t,φ(t))+y+(t,φ(t))2\displaystyle-\frac{y_{-}(t,\varphi(t))+y_{+}(t,\varphi(t))}{2}
=\displaystyle= φ(t)p(t1)|ln(t1)|1p|ln(t1)|12p+o(|ln(t1)|12p).\displaystyle-\frac{\varphi(t)}{p(t-1)|\ln(t-1)|}-\frac{1}{p}|\ln(t-1)|^{-1-\frac{2}{p}}+o(|\ln(t-1)|^{-1-\frac{2}{p}}).

This means that φ(t)=O((t1)|ln(t1)|2p)\varphi(t)=O((t-1)|\ln(t-1)|^{-\frac{2}{p}}) is optimal.

Lemma 4.2.

Under assumption (1.7), for any c(1,+)c\in(-1,+\infty), there exist ε=ε(c),δ=δ(c)>0\varepsilon=\varepsilon(c),\delta=\delta(c)>0 such that for (s,λ){0<s<ε,cδ<λ<c+δ}(s,\lambda)\in\{0<s<\varepsilon,c-\delta<\lambda<c+\delta\}, (s,λ)y+(t,x)(s,\lambda)\rightarrow y_{+}(t,x) has the expansion

y+(t,x)=s(1+ln(c+1)+lnppsp+sp(λc)p(c+1))+Oc(smin{p+2,2p+1}+sp+1|λc|2),y_{+}(t,x)=s\left(1+\frac{\ln(c+1)+\ln p}{p}s^{p}+\frac{s^{p}(\lambda-c)}{p(c+1)}\right)+O_{c}(s^{\min\{p+2,2p+1\}}+s^{p+1}|\lambda-c|^{2}), (4.10)

and for (s,λ){0<s<ε,cδ<λ<c+δ}(s,\lambda)\in\{0<s<\varepsilon,-c-\delta<\lambda<-c+\delta\}, (s,λ)y(t,x)(s,\lambda)\rightarrow y_{-}(t,x) has the expansion

y(t,x)=s(1ln(c+1)+lnppsp+sp(λ+c)p(c+1))+Oc(smin{p+2,2p+1}+sp+1|λ+c|2).y_{-}(t,x)=s\left(-1-\frac{\ln(c+1)+\ln p}{p}s^{p}+\frac{s^{p}(\lambda+c)}{p(c+1)}\right)+O_{c}(s^{\min\{p+2,2p+1\}}+s^{p+1}|\lambda+c|^{2}). (4.11)
Proof.

Similarly to Lemma 2.4, we first consider the case of p1p\geq 1. By taking λ=c>1\lambda=c>-1 and s=0s=0 in (2.40), we have the solution ζc=ln(c+1)\zeta_{c}=\ln(c+1) and μc=1\mu_{c}=1. Furthermore, direct computation yields

sF1(0,ζc,c)\displaystyle\partial_{s}F_{1}(0,\zeta_{c},c) =\displaystyle= c(ln(c+1)+lnp)δp1+p(c+1)2r0′′(0),\displaystyle c\left(\ln(c+1)+\ln p\right)\delta_{p}^{1}+\frac{p(c+1)}{2}r^{\prime\prime}_{0}(0), (4.12)
ζF1(0,ζc,c)\displaystyle\partial_{\zeta}F_{1}(0,\zeta_{c},c) =\displaystyle= c+1,\displaystyle c+1, (4.13)
λF1(0,ζc,c)\displaystyle\partial_{\lambda}F_{1}(0,\zeta_{c},c) =\displaystyle= 1,\displaystyle-1, (4.14)

where δp1={1,p=1,0,p>1.\delta_{p}^{1}=\left\{\begin{array}[]{cc}1,&\ p=1,\\ 0,&\ p>1.\end{array}\right. By the implicit function theorem, we have

ζ(s,λ)=ln(c+1)(c(ln(c+1)+lnp)δp1+p(c+1)2r0′′(0))s+λcc+1+Oc(s2+|λc|2),\zeta(s,\lambda)=\ln(c+1)-\left(c\left(\ln(c+1)+\ln p\right)\delta_{p}^{1}+\frac{p(c+1)}{2}r^{\prime\prime}_{0}(0)\right)s+\frac{\lambda-c}{c+1}+O_{c}(s^{2}+|\lambda-c|^{2}), (4.15)

and then

μ+(s,λ)=(1sp(ζ+lnp))1p=1+ln(c+1)+lnppsp+sp(λc)p(c+1)+Oc(sp+1+sp|λc|2).\mu_{+}(s,\lambda)=(1-s^{p}(\zeta+\ln p))^{-\frac{1}{p}}=1+\frac{\ln(c+1)+\ln p}{p}s^{p}+\frac{s^{p}(\lambda-c)}{p(c+1)}+O_{c}(s^{p+1}+s^{p}|\lambda-c|^{2}). (4.16)

from which (4.10) follows.

For p(0,1]p\in(0,1], recalling ω=sp\omega=s^{p} and then taking ω=0\omega=0, λ=c\lambda=c and ζ=ζc=ln(c+1)\zeta=\zeta_{c}=\ln(c+1) in (2.45), one can arrive at

ωF2(0,ζc,c)\displaystyle\partial_{\omega}F_{2}(0,\zeta_{c},c) =\displaystyle= c(ln(c+1)+lnp)p,\displaystyle\frac{c\left(\ln(c+1)+\ln p\right)}{p}, (4.17)
ζF2(0,ζc,c)\displaystyle\partial_{\zeta}F_{2}(0,\zeta_{c},c) =\displaystyle= c+1,\displaystyle c+1, (4.18)
λF2(0,ζc,c)\displaystyle\partial_{\lambda}F_{2}(0,\zeta_{c},c) =\displaystyle= 1,\displaystyle-1, (4.19)

and then by the implicit function theorem we have

ζ(s,λ)=ζ~(ω,λ)=ln(c+1)+c(ln(c+1)+lnp)pω+λcc+1+Oc(s2+|λc|2).\zeta(s,\lambda)=\tilde{\zeta}(\omega,\lambda)=\ln(c+1)+\frac{c\left(\ln(c+1)+\ln p\right)}{p}\omega+\frac{\lambda-c}{c+1}+O_{c}(s^{2}+|\lambda-c|^{2}). (4.20)

Thus

μ+(s,λ)=μ~+(ω,λ)=(1ω(ζ+lnp))1p=1+ln(c+1)+lnppω+ω(λc)p(c+1)+Oc(ω2+ω|λc|2)\mu_{+}(s,\lambda)=\tilde{\mu}_{+}(\omega,\lambda)=(1-\omega(\zeta+\ln p))^{-\frac{1}{p}}=1+\frac{\ln(c+1)+\ln p}{p}\omega+\frac{\omega(\lambda-c)}{p(c+1)}+O_{c}(\omega^{2}+\omega|\lambda-c|^{2}) (4.21)

and (4.10) can be obtained.

On the other hand for λ=c\lambda=-c and μc=1\mu_{-c}=-1, (4.11) can be proved by the same method. ∎


Next we consider the behavior of y(t,x)y(t,x) near xx-axis. Note that for y>0y>0,

x=ξeξpx=\xi e^{-\xi^{-p}} (4.22)

is a monotonically increasing function of ξ\xi from [0,+)[0,+\infty) to [0,+)[0,+\infty). Then there exists a unique inverse function h(x)h(x) of (4.22) satisfying that for x>0x>0 sufficiently small,

h(x)=|lnx|1p+O(|lnx|11pln|lnx|).h(x)=|\ln x|^{-\frac{1}{p}}+O(|\ln x|^{-1-\frac{1}{p}}\ln|\ln x|). (4.23)

Define

ξ={h(|x|),x>0,h(|x|),x<0,ν=yξ,η=(t1)h(|x|)|x|.\xi=\left\{\begin{array}[]{ll}h(|x|),&x>0,\\ -h(|x|),&x<0,\end{array}\right.\nu=\frac{y}{\xi},\ \eta=\frac{(t-1)h(|x|)}{|x|}. (4.24)
Lemma 4.3.

Under assumption (1.7), there exist some constants ε\varepsilon, δ>0\delta>0 small enough such that for (η,ξ){0<ξ<δ,ε<η<ε}(\eta,\xi)\in\{0<\xi<\delta,-\varepsilon<\eta<\varepsilon\}, (η,ξ)y+(t,x)(\eta,\xi)\rightarrow y_{+}(t,x) has the expansion

y+(t,x)=ξ(1+lnppξp+1pξpη)+O(ξmin{p+2,2p+1}+ξp+1η2),y_{+}(t,x)=\xi\left(1+\frac{\ln p}{p}\xi^{p}+\frac{1}{p}\xi^{p}\eta\right)+O(\xi^{\min\{p+2,2p+1\}}+\xi^{p+1}\eta^{2}), (4.25)

and for (η,ξ){δ<ξ<0,ε<η<ε}(\eta,\xi)\in\{-\delta<\xi<0,-\varepsilon<\eta<\varepsilon\}, (η,ξ)y+(t,x)(\eta,\xi)\rightarrow y_{+}(t,x) has the expansion

y(t,x)=ξ(1+lnpp(ξ)p+1p(ξ)pη)+O((ξ)min{p+2,2p+1}+(ξ)p+1η2).y_{-}(t,x)=\xi\left(1+\frac{\ln p}{p}(-\xi)^{p}+\frac{1}{p}(-\xi)^{p}\eta\right)+O((-\xi)^{\min\{p+2,2p+1\}}+(-\xi)^{p+1}\eta^{2}). (4.26)
Proof.

We only consider the case of x>0x>0 and then y=y+(t,x)>0y=y_{+}(t,x)>0 since the other case can be treated analogously. By x=ξeξpx=\xi e^{-\xi^{-p}}, y=ξνy=\xi\nu, t1=ηeξpt-1=\eta e^{-\xi^{-p}} and (1.8), (2.3) becomes

H(η,ξ,ν)ην+νpeξpνp(η+eξp)+eξpνp(η+eξp)ξr0(ξν)1=0.H(\eta,\xi,\nu)\triangleq-\eta\nu+\frac{\nu}{p}e^{-\xi^{-p}\nu^{-p}}\left(\eta+e^{\xi^{-p}}\right)+\frac{e^{-\xi^{-p}\nu^{-p}}\left(\eta+e^{\xi^{-p}}\right)}{\xi}r_{0}(\xi\nu)-1=0. (4.27)

Similarly to Lemma 2.4, we divide the proof procedure into two cases of p1p\geq 1 and 0<p<10<p<1. Firstly we consider p1p\geq 1. Set θ=ξp(1νp)lnp\theta=\xi^{-p}(1-\nu^{-p})-\ln p and ν=(1ξp(θ+lnp))1p\nu=\left(1-\xi^{p}(\theta+\ln p)\right)^{-\frac{1}{p}}. Then (4.27) becomes

J1(η,ξ,θ)H(η,ξ,ν)=ην+(νeθ1)+ηνeθξp+peθ(ηeξp+1)ξr0(ξν).J_{1}(\eta,\xi,\theta)\triangleq H(\eta,\xi,\nu)=-\eta\nu+\left(\nu e^{\theta}-1\right)+\eta\nu e^{\theta-\xi^{-p}}+\frac{pe^{\theta}\left(\eta e^{-\xi^{-p}}+1\right)}{\xi}r_{0}(\xi\nu). (4.28)

Note J1(0,0,0)=0J_{1}(0,0,0)=0, and

ξν=ξp1(θ+lnp)(1ξp(θ+lnp))1p1,θν=1pξp(1ξp(θ+lnp))1p1\partial_{\xi}\nu=\xi^{p-1}(\theta+\ln p)(1-\xi^{p}(\theta+\ln p))^{-\frac{1}{p}-1},\ \partial_{\theta}\nu=\frac{1}{p}\xi^{p}(1-\xi^{p}(\theta+\ln p))^{-\frac{1}{p}-1} (4.29)

are bounded near ξ=0\xi=0 and θ=0\theta=0 by p1p\geq 1. Due to

ξJ1\displaystyle\partial_{\xi}J_{1} =\displaystyle= (eθη)ξν+ηeθξp(ξν+pνξp1)\displaystyle(e^{\theta}-\eta)\partial_{\xi}\nu+\eta e^{\theta-\xi^{-p}}(\partial_{\xi}\nu+p\nu\xi^{-p-1}) (4.31)
peθξ2(ηeξp(1pξp)+1)r0(ξν)+peθ(ηeξp+1)ξr0(ξν)(ν+ξξν),\displaystyle-\frac{pe^{\theta}}{\xi^{2}}(\eta e^{-\xi^{-p}}(1-p\xi^{-p})+1)r_{0}(\xi\nu)+\frac{pe^{\theta}\left(\eta e^{-\xi^{-p}}+1\right)}{\xi}r^{\prime}_{0}(\xi\nu)(\nu+\xi\partial_{\xi}\nu),
θJ1\displaystyle\partial_{\theta}J_{1} =\displaystyle= eθν+(eθη)θν+ηeθξp(ν+θν)+peθ(ηeξp+1)ξ(r0(ξν)+ξr0(ξν)θν),\displaystyle e^{\theta}\nu+(e^{\theta}-\eta)\partial_{\theta}\nu+\eta e^{\theta-\xi^{-p}}(\nu+\partial_{\theta}\nu)+\frac{pe^{\theta}\left(\eta e^{-\xi^{-p}}+1\right)}{\xi}(r_{0}(\xi\nu)+\xi r^{\prime}_{0}(\xi\nu)\partial_{\theta}\nu), (4.32)
ηJ1\displaystyle\partial_{\eta}J_{1} =\displaystyle= ν+eθξp(ν+pξr0(ξν)),\displaystyle-\nu+e^{\theta-\xi^{-p}}(\nu+\frac{p}{\xi}r_{0}(\xi\nu)), (4.33)

we then obtain

ξJ1(0,0,0)=δ1plnp+p2r0′′(0),θJ1(0,0,0)=1,ηJ1(0,0,0)=1.\partial_{\xi}J_{1}(0,0,0)=\delta_{1}^{p}\ln p+\frac{p}{2}r^{\prime\prime}_{0}(0),\ \partial_{\theta}J_{1}(0,0,0)=1,\ \partial_{\eta}J_{1}(0,0,0)=-1. (4.34)

Thus by the implicit function theorem, one can deduce that there exists a unique function θ=θ(η,ξ)\theta=\theta(\eta,\xi) near (η,ξ)=(0,0)(\eta,\xi)=(0,0) satisfying

θ(η,ξ)=(δ1plnp+p2r0′′(0))ξ+η+O(ξ2+η2).\theta(\eta,\xi)=-\left(\delta_{1}^{p}\ln p+\frac{p}{2}r^{\prime\prime}_{0}(0)\right)\xi+\eta+O(\xi^{2}+\eta^{2}). (4.35)

Recalling ν=(1ξp(θ+lnp))1p\nu=\left(1-\xi^{p}(\theta+\ln p)\right)^{-\frac{1}{p}}, we then have

ν(η,ξ)=1+lnppξp+1pξpη+O(ξp+1+ξpη2).\nu(\eta,\xi)=1+\frac{\ln p}{p}\xi^{p}+\frac{1}{p}\xi^{p}\eta+O\left(\xi^{p+1}+\xi^{p}\eta^{2}\right). (4.36)

For p(0,1)p\in(0,1), set ς=ξp\varsigma=\xi^{p} and then ν=(1ς(θ+lnp))1p\nu=(1-\varsigma(\theta+\ln p))^{-\frac{1}{p}}. In this case, (4.27) becomes

J2(η,ς,θ)H(η,ξ,ν)=ην+(νeθ1)+ηνeθς1+peθ(ηeς1+1)ς1pr0(ς1pν).J_{2}(\eta,\varsigma,\theta)\triangleq H(\eta,\xi,\nu)=-\eta\nu+\left(\nu e^{\theta}-1\right)+\eta\nu e^{\theta-\varsigma^{-1}}+\frac{pe^{\theta}\left(\eta e^{-\varsigma^{-1}}+1\right)}{\varsigma^{\frac{1}{p}}}r_{0}(\varsigma^{\frac{1}{p}}\nu). (4.37)

By J2(0,0,0)=0J_{2}(0,0,0)=0 and

ςν=1p(θ+lnp)(1ς(θ+lnp))1p1,θν=ςp(1ς(θ+lnp))1p1,\partial_{\varsigma}\nu=\frac{1}{p}(\theta+\ln p)(1-\varsigma(\theta+\ln p))^{-\frac{1}{p}-1},\ \partial_{\theta}\nu=\frac{\varsigma}{p}(1-\varsigma(\theta+\ln p))^{-\frac{1}{p}-1}, (4.38)

under assumption (1.8), we have that J2(η,ξ,θ)C1pJ_{2}(\eta,\xi,\theta)\in C^{\frac{1}{p}} and

ξJ2\displaystyle\partial_{\xi}J_{2} =\displaystyle= (eθη)ςν+ηeθς1(ςν+νς2)\displaystyle(e^{\theta}-\eta)\partial_{\varsigma}\nu+\eta e^{\theta-\varsigma^{-1}}(\partial_{\varsigma}\nu+\nu\varsigma^{-2}) (4.40)
eθς1p+1(ηeς1(1pς1)+1)r0(ς1pν)+peθ(ηeς1+1)ς1pr0(ς1pν)(1pς1p1ν+ς1pςν),\displaystyle-\frac{e^{\theta}}{\varsigma^{\frac{1}{p}+1}}(\eta e^{-\varsigma^{-1}}(1-p\varsigma^{-1})+1)r_{0}(\varsigma^{\frac{1}{p}}\nu)+\frac{pe^{\theta}\left(\eta e^{-\varsigma^{-1}}+1\right)}{\varsigma^{\frac{1}{p}}}r^{\prime}_{0}(\varsigma^{\frac{1}{p}}\nu)(\frac{1}{p}\varsigma^{\frac{1}{p}-1}\nu+\varsigma^{\frac{1}{p}}\partial_{\varsigma}\nu),
θJ2\displaystyle\partial_{\theta}J_{2} =\displaystyle= eθν+(eθη)θν+ηeθς1(ν+θν)+peθ(ηeς1+1)ς1p(r0(ς1pν)+ς1pr0(ς1pν)θν),\displaystyle e^{\theta}\nu+(e^{\theta}-\eta)\partial_{\theta}\nu+\eta e^{\theta-\varsigma^{-1}}(\nu+\partial_{\theta}\nu)+\frac{pe^{\theta}\left(\eta e^{-\varsigma^{-1}}+1\right)}{\varsigma^{\frac{1}{p}}}(r_{0}(\varsigma^{\frac{1}{p}}\nu)+\varsigma^{\frac{1}{p}}r^{\prime}_{0}(\varsigma^{\frac{1}{p}}\nu)\partial_{\theta}\nu), (4.41)
ηJ2\displaystyle\partial_{\eta}J_{2} =\displaystyle= ν+eθς1(ν+pς1pr0(ς1pν)).\displaystyle-\nu+e^{\theta-\varsigma^{-1}}(\nu+\frac{p}{\varsigma^{\frac{1}{p}}}r_{0}(\varsigma^{\frac{1}{p}}\nu)). (4.42)

This yields

ςJ2(0,0,0)=lnpp,θJ2(0,0,0)=1,ηJ2(0,0,0)=1.\partial_{\varsigma}J_{2}(0,0,0)=\frac{\ln p}{p},\ \partial_{\theta}J_{2}(0,0,0)=1,\ \partial_{\eta}J_{2}(0,0,0)=-1. (4.43)

By the implicit function theorem, we know that there exists a unique solution θ=θ(η,ς)\theta=\theta(\eta,\varsigma) for (η,ς)(\eta,\varsigma) near (0,0)(0,0), which satisfies θ(0,0)=0\theta(0,0)=0 and

θ(η,ς)=lnppς+η+O(ς2+η2).\theta(\eta,\varsigma)=-\frac{\ln p}{p}\varsigma+\eta+O\left(\varsigma^{2}+\eta^{2}\right). (4.44)

By ξ=ς1p\xi=\varsigma^{\frac{1}{p}}, then

ν(η,ξ)\displaystyle\nu(\eta,\xi) =\displaystyle= (1ς(θ+lnp))1p\displaystyle(1-\varsigma(\theta+\ln p))^{-\frac{1}{p}} (4.45)
=\displaystyle= (1ςlnpςη+O(ς2+η2))1p\displaystyle\left(1-\varsigma\ln p-\varsigma\eta+O\left(\varsigma^{2}+\eta^{2}\right)\right)^{-\frac{1}{p}}
=\displaystyle= 1+lnppξp+1pξpη+O(ξ2p+ξpη2).\displaystyle 1+\frac{\ln p}{p}\xi^{p}+\frac{1}{p}\xi^{p}\eta+O(\xi^{2p}+\xi^{p}\eta^{2}).

Therefore we finish the proof of (4.25).

For x<0x<0, we can transform (2.3) to H(η,ξ,ν)=0H(\eta,-\xi,-\nu)=0. Analogously, we can obtain (4.26) about y(t,x)y_{-}(t,x) and x<0x<0. ∎


Next we consider the behavior of y(t,x)y(t,x) for t<1t<1 near (1,0)(1,0). Without of confusion, we still denote

τ=1t,s=|lnτ|1p,λ=xsτ,μ=ys,\tau=1-t,\ s=|\ln\tau|^{-\frac{1}{p}},\ \lambda=\frac{x}{s\tau},\ \mu=\frac{y}{s}, (4.46)

as in the case of t>1t>1. By divided sτs\tau, (2.3) then becomes

G(s,λ,μ)μ(1+1pesp(1|μ|p))μpesp|μ|p+e|μ|psp(esp1)sr0(sμ)λ=0.G(s,\lambda,\mu)\triangleq\mu(1+\frac{1}{p}e^{s^{-p}(1-|\mu|^{-p})})-\frac{\mu}{p}e^{-s^{-p}|\mu|^{-p}}+\frac{e^{-|\mu|^{-p}s^{-p}}(e^{s^{-p}}-1)}{s}r_{0}(s\mu)-\lambda=0. (4.47)
Lemma 4.4.

Under assumption (1.7), we have that

(1) for any c>1c>1, there exist ε=ε(c),δ=δ(c)>0\varepsilon=\varepsilon(c),\delta=\delta(c)>0 such that for (s,λ){0<s<ε,1<cδ<λ<c+δ}(s,\lambda)\in\{0<s<\varepsilon,1<c-\delta<\lambda<c+\delta\}, (s,λ)y(t,x)(s,\lambda)\rightarrow y(t,x) has the expansion

y(t,x)=s(1+ln(c1)+lnppsp+sp(λc)p(c1))+O(smin{p+2,2p+1}+sp+1|λc|2).y(t,x)=s\left(1+\frac{\ln(c-1)+\ln p}{p}s^{p}+\frac{s^{p}(\lambda-c)}{p(c-1)}\right)+O(s^{\min\{p+2,2p+1\}}+s^{p+1}|\lambda-c|^{2}). (4.48)

(2) for any 0c<10\leq c<1, there exist ε=ε(c),δ=δ(c)>0\varepsilon=\varepsilon(c),\delta=\delta(c)>0 such that for (s,λ){0<s<ε,cδ<λ<c+δ<1}(s,\lambda)\in\{0<s<\varepsilon,c-\delta<\lambda<c+\delta<1\}, (s,λ)y(t,x)(s,\lambda)\rightarrow y(t,x) has the expansion

y(t,x)=s(c+(λc))+O(s3+s|λc|2).y(t,x)=s\left(c+(\lambda-c)\right)+O(s^{3}+s|\lambda-c|^{2}). (4.49)

(3) for any c<1c<-1, there exist ε=ε(c),δ=δ(c)>0\varepsilon=\varepsilon(c),\delta=\delta(c)>0 such that for (s,λ){0<s<ε,cδ<λ<c+δ<1}(s,\lambda)\in\{0<s<\varepsilon,c-\delta<\lambda<c+\delta<-1\}, (s,λ)y(t,x)(s,\lambda)\rightarrow y(t,x) has the expansion

y(t,x)=s(1+ln(c+1)+lnppspsp(λc)p(c+1))+O(smin{p+2,2p+1}+sp+1|λc|2).y(t,x)=-s\left(1+\frac{\ln(-c+1)+\ln p}{p}s^{p}-\frac{s^{p}(\lambda-c)}{p(c+1)}\right)+O(s^{\min\{p+2,2p+1\}}+s^{p+1}|\lambda-c|^{2}). (4.50)

(4) for any 1<c<0-1<c<0, there exist ε=ε(c),δ=δ(c)>0\varepsilon=\varepsilon(c),\delta=\delta(c)>0 such that for (s,λ){0<s<ε,1<cδ<λ<c+δ<1}(s,\lambda)\in\{0<s<\varepsilon,-1<c-\delta<\lambda<c+\delta<1\}, (s,λ)y(t,x)(s,\lambda)\rightarrow y(t,x) has the expansion

y(t,x)=s(c+(λc))+O(s3+s|λc|2).y(t,x)=s\left(c+(\lambda-c)\right)+O(s^{3}+s|\lambda-c|^{2}). (4.51)
Proof.

We only prove the cases (1), (2) for c0c\geq 0 since the cases (3), (4) can be obtained by the same way.


(1) If c>1c>1, it is similar to the proof of Lemma 2.4 that we adopt the variable transformation

ζ=sp(1μp)lnp,\zeta=s^{-p}(1-\mu^{-p})-\ln p, (4.52)

and then μ=(1sp(ζ+lnp))1p\mu=(1-s^{p}(\zeta+\ln p))^{-\frac{1}{p}}.

At first, we assume p1p\geq 1. Then (4.47) becomes

F1(s,λ,ζ)μ(1+eζ)μeζsp+peζ(1esp)sr0(sμ)λ.F_{1}(s,\lambda,\zeta)\triangleq\mu(1+e^{\zeta})-\mu e^{\zeta-s^{-p}}+\frac{pe^{\zeta}(1-e^{-s^{-p}})}{s}r_{0}(s\mu)-\lambda. (4.53)

It is easy to see that for c>1c>1, F1(0,c,ln(c1))=0F_{1}(0,c,\ln(c-1))=0 and

sμ=sp1(ζ+lnp)(1sp(ζ+lnp))1p1,ζμ=spp(1sp(ζ+lnp))1p1.\partial_{s}\mu=s^{p-1}(\zeta+\ln p)(1-s^{p}(\zeta+\ln p))^{-\frac{1}{p}-1},\ \partial_{\zeta}\mu=\frac{s^{p}}{p}(1-s^{p}(\zeta+\ln p))^{-\frac{1}{p}-1}.

Then

sF1(s,λ,ζ)\displaystyle\partial_{s}F_{1}(s,\lambda,\zeta) =\displaystyle= (1+eζeζsp)sμ+psp1μeζsp\displaystyle\left(1+e^{\zeta}-e^{\zeta-s^{-p}}\right)\partial_{s}\mu+ps^{-p-1}\mu e^{\zeta-s^{-p}}
+peζs2(esp(1psp)1)r0(sμ)+peζ(1esp)sr0(sμ)(μ+ssμ),\displaystyle+\frac{pe^{\zeta}}{s^{2}}\left(e^{-s^{-p}}(1-ps^{-p})-1\right)r_{0}(s\mu)+\frac{pe^{\zeta}(1-e^{-s^{-p}})}{s}r^{\prime}_{0}(s\mu)\left(\mu+s\partial_{s}\mu\right),
λF1(s,λ,ζ)\displaystyle\partial_{\lambda}F_{1}(s,\lambda,\zeta) =\displaystyle= 1,\displaystyle-1,
ζF1(s,λ,ζ)\displaystyle\partial_{\zeta}F_{1}(s,\lambda,\zeta) =\displaystyle= (1+eζeζsp)ζμ+μeζ(1esp)+peζ(1esp)s(r0(sμ)+r0(sμ)ζμ).\displaystyle\left(1+e^{\zeta}-e^{\zeta-s^{-p}}\right)\partial_{\zeta}\mu+\mu e^{\zeta}(1-e^{-s^{-p}})+\frac{pe^{\zeta}(1-e^{-s^{-p}})}{s}(r_{0}(s\mu)+r^{\prime}_{0}(s\mu)\partial_{\zeta}\mu).

This yields

sF1(0,c,ln(c1))=c(lnp+ln(c1))δ1p+p(c1)2r0(0),λF1(0,c,ln(c1))=1,\displaystyle\partial_{s}F_{1}(0,c,\ln(c-1))=c\left(\ln p+\ln(c-1)\right)\delta_{1}^{p}+\frac{p(c-1)}{2}r^{\prime\prime}_{0}(0),\ \partial_{\lambda}F_{1}(0,c,\ln(c-1))=-1,
ζF1(0,c,ln(c1))=c1.\displaystyle\partial_{\zeta}F_{1}(0,c,\ln(c-1))=c-1. (4.54)

By the implicit function theorem, there exists a unique solution ζ(s,λ)\zeta(s,\lambda) satisfying ζ(0,c)=ln(c1)\zeta(0,c)=\ln(c-1) and ζ(s,λ)\zeta(s,\lambda) is CC^{\infty} near (0,c)(0,c) with

ζ=ln(c1)c(lnp+ln(c1))δ1p+p(c1)2r0(0)c1s+λcc1+O(s+λ2).\zeta=\ln(c-1)-\frac{c\left(\ln p+\ln(c-1)\right)\delta_{1}^{p}+\frac{p(c-1)}{2}r^{\prime\prime}_{0}(0)}{c-1}s+\frac{\lambda-c}{c-1}+O(s+\lambda^{2}). (4.55)

Thus

μ\displaystyle\mu =\displaystyle= (1sp(ζ+lnp))1p\displaystyle(1-s^{p}(\zeta+\ln p))^{-\frac{1}{p}} (4.56)
=\displaystyle= 1+ln(c1)+lnppsp+sp(λc)p(c1)+O(sp+1+sp(λc)2).\displaystyle 1+\frac{\ln(c-1)+\ln p}{p}s^{p}+\frac{s^{p}(\lambda-c)}{p(c-1)}+O(s^{p+1}+s^{p}(\lambda-c)^{2}).

Secondly, we consider the case of p(0,1)p\in(0,1). Let ω=sp\omega=s^{p} and then μ=(1ω(ζ+lnp))1p\mu=\left(1-\omega(\zeta+\ln p)\right)^{-\frac{1}{p}}. In this case, (4.53) becomes

F2(s,λ,ω)μ(1+eζ)μeζω1+peζ(1eω1)ω1pr0(ω1pμ)λ.F_{2}(s,\lambda,\omega)\triangleq\mu(1+e^{\zeta})-\mu e^{\zeta-\omega^{-1}}+\frac{pe^{\zeta}(1-e^{-\omega^{-1}})}{\omega^{\frac{1}{p}}}r_{0}(\omega^{\frac{1}{p}}\mu)-\lambda. (4.57)

It is obvious that for c>1c>1, F2(0,c,ln(c1))=0F_{2}(0,c,\ln(c-1))=0. By direct computation, we have that

ωμ=ζ+lnpp(1ω(ζ+lnp))1p1,ζμ=ωp(1ω(ζ+lnp))1p1\partial_{\omega}\mu=\frac{\zeta+\ln p}{p}(1-\omega(\zeta+\ln p))^{-\frac{1}{p}-1},\ \partial_{\zeta}\mu=\frac{\omega}{p}(1-\omega(\zeta+\ln p))^{-\frac{1}{p}-1}

and

ωF2(ω,λ,ζ)\displaystyle\partial_{\omega}F_{2}(\omega,\lambda,\zeta) =\displaystyle= (1+eζeζω1)ωμ+s2μeζω1\displaystyle\left(1+e^{\zeta}-e^{\zeta-\omega^{-1}}\right)\partial_{\omega}\mu+s^{-2}\mu e^{\zeta-\omega^{-1}}
+eζω1p+1(eω1(1pω1)1)r0(ω1pμ)+eζ(1eω1)ω1pr0(ω1pμ)(ω1p1μ+ω1pωμ),\displaystyle+\frac{e^{\zeta}}{\omega^{\frac{1}{p}+1}}\left(e^{-\omega^{-1}}(1-p\omega^{-1})-1\right)r_{0}(\omega^{\frac{1}{p}}\mu)+\frac{e^{\zeta}(1-e^{-\omega^{-1}})}{\omega^{\frac{1}{p}}}r^{\prime}_{0}(\omega^{\frac{1}{p}}\mu)\left(\omega^{\frac{1}{p}-1}\mu+\omega^{\frac{1}{p}}\partial_{\omega}\mu\right),
λF2(ω,λ,ζ)\displaystyle\partial_{\lambda}F_{2}(\omega,\lambda,\zeta) =\displaystyle= 1,\displaystyle-1,
ζF2(ω,λ,ζ)\displaystyle\partial_{\zeta}F_{2}(\omega,\lambda,\zeta) =\displaystyle= (1+eζeζω1)ζμ+μeζ(1eω1)+peζ(1eω1)ω1p(r0(ω1pμ)+r0(ω1pμ)ζμ).\displaystyle\left(1+e^{\zeta}-e^{\zeta-\omega^{-1}}\right)\partial_{\zeta}\mu+\mu e^{\zeta}(1-e^{-\omega^{-1}})+\frac{pe^{\zeta}(1-e^{-\omega^{-1}})}{\omega^{\frac{1}{p}}}(r_{0}(\omega^{\frac{1}{p}}\mu)+r^{\prime}_{0}(\omega^{\frac{1}{p}}\mu)\partial_{\zeta}\mu).

This yields

sF1(0,c,ln(c1))=c(lnp+ln(c1))p,λF1(0,c,ln(c1))=1,ζF1(0,c,ln(c1))=c1.\partial_{s}F_{1}(0,c,\ln(c-1))=\frac{c(\ln p+\ln(c-1))}{p},\ \partial_{\lambda}F_{1}(0,c,\ln(c-1))=-1,\ \partial_{\zeta}F_{1}(0,c,\ln(c-1))=c-1. (4.58)

By the implicit function theorem, there exists a unique solution ζ(s,λ)\zeta(s,\lambda) satisfying ζ(0,c)=ln(c1)\zeta(0,c)=\ln(c-1) and F2(s,λ,ζ(s,λ))=0F_{2}(s,\lambda,\zeta(s,\lambda))=0 with

ζ=ln(c1)c(lnp+ln(c1))p(c1)s+λcc1+O(s+λ2).\zeta=\ln(c-1)-\frac{c\left(\ln p+\ln(c-1)\right)}{p(c-1)}s+\frac{\lambda-c}{c-1}+O(s+\lambda^{2}). (4.59)

Thus

μ\displaystyle\mu =\displaystyle= (1sp(ζ+lnp))1p\displaystyle(1-s^{p}(\zeta+\ln p))^{-\frac{1}{p}} (4.60)
=\displaystyle= 1+ln(c1)+lnppsp+sp(λc)p(c1)+O(sp+1+sp(λc)2).\displaystyle 1+\frac{\ln(c-1)+\ln p}{p}s^{p}+\frac{s^{p}(\lambda-c)}{p(c-1)}+O(s^{p+1}+s^{p}(\lambda-c)^{2}).

Together with (4.56) and (4.60), this derives (4.48).


(2) For 0<c<10<c<1, define

G(s,λ,μ)μ(1+1pesp(1|μ|p))μpesp|μ|p+e|μ|psp(esp1)sr0(sμ)λ=0.G(s,\lambda,\mu)\triangleq\mu(1+\frac{1}{p}e^{s^{-p}(1-|\mu|^{-p})})-\frac{\mu}{p}e^{-s^{-p}|\mu|^{-p}}+\frac{e^{-|\mu|^{-p}s^{-p}}(e^{s^{-p}}-1)}{s}r_{0}(s\mu)-\lambda=0. (4.61)

It is clear that G(0,c,c)=0G(0,c,c)=0 and

sG(s,λ,μ)\displaystyle\partial_{s}G(s,\lambda,\mu) =\displaystyle= sp1(μ+pr0(sμ)s)(esp(1μp)(μp1)espμpμp)\displaystyle s^{-p-1}\left(\mu+\frac{pr_{0}(s\mu)}{s}\right)\left(e^{s^{-p}(1-\mu^{-p})}(\mu^{-p}-1)-e^{-s^{-p}\mu^{-p}}\mu^{-p}\right)
+sμr0(sμ)r0(sμ)s2(esp(1μp)espμp),\displaystyle+\frac{s\mu r^{\prime}_{0}(s\mu)-r_{0}(s\mu)}{s^{2}}\left(e^{s^{-p}(1-\mu^{-p})}-e^{-s^{-p}\mu^{-p}}\right),
λG(s,λ,μ)\displaystyle\partial_{\lambda}G(s,\lambda,\mu) =\displaystyle= 1,\displaystyle-1,
μG(s,λ,μ)\displaystyle\partial_{\mu}G(s,\lambda,\mu) =\displaystyle= 1+spμp1(μ+pr0(sμ)s)(esp(1μp)(1μp)espμpμp)\displaystyle 1+s^{-p}\mu^{-p-1}\left(\mu+\frac{pr_{0}(s\mu)}{s}\right)\left(e^{s^{-p}(1-\mu^{-p})}(1-\mu^{-p})-e^{-s^{-p}\mu^{-p}}\mu^{-p}\right)
+(1p+r0(sμ))(esp(1μp)espμp).\displaystyle+\left(\frac{1}{p}+r^{\prime}_{0}(s\mu)\right)\left(e^{s^{-p}(1-\mu^{-p})}-e^{-s^{-p}\mu^{-p}}\right).

Then in light of c<1c<1, it follows that

sG(0,c,c)\displaystyle\partial_{s}G(0,c,c) =\displaystyle= 0,\displaystyle 0,
λG(0,c,c)\displaystyle\partial_{\lambda}G(0,c,c) =\displaystyle= 1,\displaystyle-1,
μG(0,c,c)\displaystyle\partial_{\mu}G(0,c,c) =\displaystyle= 1.\displaystyle 1.

Thus by the implicit function theorem, there exists a unique solution μ=μ(s,λ)\mu=\mu(s,\lambda) satisfying that μ(0,c)=c\mu(0,c)=c and

μ(s,λ)=c+(λc)+Oc(s2+|λc|2).\mu(s,\lambda)=c+(\lambda-c)+O_{c}(s^{2}+|\lambda-c|^{2}). (4.62)

Therefore, by y=sμy=s\mu, we finish the proof of (4.49).


Proof of Theorem 1.2:

(1) It can be obtained by Lemma 4.1.

(2) Since we don’t get the behavior of y(t,x)y(t,x) for c=±1c=\pm 1 in Lemma 4.4, we have to choose the domain Ω,t,+0\Omega_{-,t,+}^{0} and Ω,t,0\Omega_{-,t,-}^{0} as follows

Ωt,,+0\displaystyle\Omega_{t,-,+}^{0} =\displaystyle= {(t,x): 0<s<ε0, 1δ0<xsτ<1+δ0},\displaystyle\{(t,x):\ 0<s<\varepsilon_{0},\ 1-\delta_{0}<\frac{x}{s\tau}<1+\delta_{0}\}, (4.63)
Ωt,,0\displaystyle\Omega_{t,-,-}^{0} =\displaystyle= {(t,x): 0<s<ε0,1δ0<xsτ<1+δ0},\displaystyle\{(t,x):\ 0<s<\varepsilon_{0},\ -1-\delta_{0}<\frac{x}{s\tau}<-1+\delta_{0}\}, (4.64)

where τ=1t\tau=1-t, s=|lnτ|1ps=|\ln\tau|^{-\frac{1}{p}} and ε0\varepsilon_{0}, δ0>0\delta_{0}>0 sufficiently small.

We only consider the behavior in Ω,t,+0\Omega_{-,t,+}^{0} since the treatment in Ω,t,0\Omega_{-,t,-}^{0} is similar. By monotonicity of y(t,)y(t,\cdot) for each fixed t[0,1]t\in[0,1], we know that for (t,x)Ω,t,+0(t,x)\in\Omega_{-,t,+}^{0},

y(t,(1δ0)sτ)y(t,x)y(t,(1+δ0)sτ).y(t,(1-\delta_{0})s\tau)\leq y(t,x)\leq y(t,(1+\delta_{0})s\tau). (4.65)

Let’s firstly turn to consider y(t,(1+δ0)sτ)y(t,(1+\delta_{0})s\tau). Note that μ=y(t,(1+δ0)sτ)s\mu=\frac{y(t,(1+\delta_{0})s\tau)}{s} satisfies

1+δ0=μ(1+1pesp(1μp))μpespμp+esp1sespμpr0(sμ).1+\delta_{0}=\mu\left(1+\frac{1}{p}e^{s^{-p}(1-\mu^{-p})}\right)-\frac{\mu}{p}e^{-s^{-p}\mu^{-p}}+\frac{e^{s^{-p}}-1}{s}e^{-s^{-p}\mu^{-p}}r_{0}(s\mu). (4.66)

For p1p\geq 1, we let

ζ=sp(1μp)lnp,\zeta=s^{-p}(1-\mu^{-p})-\ln p, (4.67)

and then μ=(1sp(ζ+lnp))1p\mu=\left(1-s^{p}(\zeta+\ln p)\right)^{-\frac{1}{p}}. In this case, (4.66) becomes

J1(s,ζ)μ(1+eζ)μeζsp+peζ(1esp)sr0(sμ)(1+δ0)=0.J_{1}(s,\zeta)\triangleq\mu\left(1+e^{\zeta}\right)-\mu e^{\zeta-s^{-p}}+\frac{pe^{\zeta}(1-e^{-s^{-p}})}{s}r_{0}(s\mu)-(1+\delta_{0})=0. (4.68)

Note that if s0+s\to 0+, then ζ(lnδ0)+\zeta\to(\ln{\delta_{0}})+ and μ1+\mu\to 1+. In addition,

sμ=sp1(ζ+lnp)(1sp(ζ+lnp))1p1,ζμ=spp(1sp(ζ+lnp))1p1,\partial_{s}\mu=s^{p-1}(\zeta+\ln p)\left(1-s^{p}(\zeta+\ln p)\right)^{-\frac{1}{p}-1},\ \partial_{\zeta}\mu=\frac{s^{p}}{p}\left(1-s^{p}(\zeta+\ln p)\right)^{-\frac{1}{p}-1}, (4.69)

and then by taking s0+s\to 0+, we have

sμ(0,lnδ0)=δ1p(lnδ0+lnp),ζμ(0,lnδ0)=0.\partial_{s}\mu(0,\ln\delta_{0})=\delta_{1}^{p}\left(\ln\delta_{0}+\ln p\right),\ \partial_{\zeta}\mu(0,\ln\delta_{0})=0. (4.70)

On the other hand, it follows from direct computation that

sJ1(s,ζ)\displaystyle\partial_{s}J_{1}(s,\zeta) =\displaystyle= (1+eζeζsp)sμpsp1μeζsp\displaystyle(1+e^{\zeta}-e^{\zeta-s^{-p}})\partial_{s}\mu-ps^{-p-1}\mu e^{\zeta-s^{-p}}
peζ(esp(psp1)+1)s2r0(sμ)+peζ(esp1)sr0(sμ)(μ+ssμ),\displaystyle-\frac{pe^{\zeta}\left(e^{-s^{-p}}(ps^{-p}-1)+1\right)}{s^{2}}r_{0}(s\mu)+\frac{pe^{\zeta}(e^{-s^{-p}}-1)}{s}r^{\prime}_{0}(s\mu)\left(\mu+s\partial_{s}\mu\right),
ζJ1(s,ζ)\displaystyle\partial_{\zeta}J_{1}(s,\zeta) =\displaystyle= μeζ(1esp)+(1+eζeζsp)ζμ+peζ(esp1)s(r0(sμ)+sζμr0(sμ)).\displaystyle\mu e^{\zeta}(1-e^{-s^{-p}})+(1+e^{\zeta}-e^{\zeta-s^{-p}})\partial_{\zeta}\mu+\frac{pe^{\zeta}(e^{-s^{-p}}-1)}{s}\left(r_{0}(s\mu)+s\partial_{\zeta}\mu r^{\prime}_{0}(s\mu)\right).

Together with (1.8), this yields

sJ1(0,lnδ0)=(1+δ0)(lnδ0+lnp)+pδ02r0(0),ζJ1(0,lnδ0)=1.\partial_{s}J_{1}(0,\ln\delta_{0})=(1+\delta_{0})(\ln\delta_{0}+\ln p)+\frac{p\delta_{0}}{2}r^{\prime\prime}_{0}(0),\ \partial_{\zeta}J_{1}(0,\ln\delta_{0})=1. (4.71)

Thus by the implicit function theorem, there exists a function ζ(s)\zeta(s) satisfying J(s,ζ(s))=0J(s,\zeta(s))=0 such that

ζ(s)=lnδ0((1+δ0)(lnδ0+lnp)+pδ02r0(0))s+Oδ0(s2),s(0,ε0],\zeta(s)=\ln\delta_{0}-\left((1+\delta_{0})(\ln\delta_{0}+\ln p)+\frac{p\delta_{0}}{2}r^{\prime\prime}_{0}(0)\right)s+O_{\delta_{0}}(s^{2}),\quad s\in(0,\varepsilon_{0}], (4.72)

where ε0=ε0(δ0)>0\varepsilon_{0}=\varepsilon_{0}(\delta_{0})>0 is small enough. Therefore

μ(s)=(1sp(ζ+lnp))1p=1+lnδ0+lnppsp+Oδ0(sp+1),s(0,ε0].\mu(s)=\left(1-s^{p}(\zeta+\ln p)\right)^{-\frac{1}{p}}=1+\frac{\ln\delta_{0}+\ln p}{p}s^{p}+O_{\delta_{0}}(s^{p+1}),\quad s\in(0,\varepsilon_{0}]. (4.73)

For p(0,1)p\in(0,1), we take ω=sp\omega=s^{p} and then (4.68) becomes

J2(ω,ζ)μ(1+eζ)μeζω1+peζ(1eω1)ω1pr0(ω1pμ)(1+δ0)=0,J_{2}(\omega,\zeta)\triangleq\mu\left(1+e^{\zeta}\right)-\mu e^{\zeta-\omega^{-1}}+\frac{pe^{\zeta}(1-e^{-\omega^{-1}})}{\omega^{\frac{1}{p}}}r_{0}(\omega^{\frac{1}{p}}\mu)-(1+\delta_{0})=0, (4.74)

where μ=(1ω(ζ+lnp))1p\mu=\left(1-\omega(\zeta+\ln p)\right)^{-\frac{1}{p}}. By direct computation, one has

ωJ2(ω,ζ)\displaystyle\partial_{\omega}J_{2}(\omega,\zeta) =\displaystyle= (1+eζeζω1)ωμω2μeζω1\displaystyle(1+e^{\zeta}-e^{\zeta-\omega^{-1}})\partial_{\omega}\mu-\omega^{-2}\mu e^{\zeta-\omega^{-1}}
eζ(eω1(pω11)+1)ω1p+1r0(ω1pμ)+peζ(eω11)ω1pr0(ω1pμ)(1pω1p1μ+ωωμ),\displaystyle-\frac{e^{\zeta}\left(e^{-\omega^{-1}}(p\omega^{-1}-1)+1\right)}{\omega^{\frac{1}{p}+1}}r_{0}(\omega^{\frac{1}{p}}\mu)+\frac{pe^{\zeta}(e^{-\omega^{-1}}-1)}{\omega^{\frac{1}{p}}}r^{\prime}_{0}(\omega^{\frac{1}{p}}\mu)\left(\frac{1}{p}\omega^{\frac{1}{p}-1}\mu+\omega\partial_{\omega}\mu\right),
ζJ2(ω,ζ)\displaystyle\partial_{\zeta}J_{2}(\omega,\zeta) =\displaystyle= μeζ(1eω1)+(1+eζeζω1)ζμ+peζ(eω11)ω1p(r0(ω1pμ)+ω1pζμr0(ω1pμ)),\displaystyle\mu e^{\zeta}(1-e^{-\omega^{-1}})+(1+e^{\zeta}-e^{\zeta-\omega^{-1}})\partial_{\zeta}\mu+\frac{pe^{\zeta}(e^{-\omega^{-1}}-1)}{\omega^{\frac{1}{p}}}\left(r_{0}(\omega^{\frac{1}{p}}\mu)+\omega^{\frac{1}{p}}\partial_{\zeta}\mu r^{\prime}_{0}(\omega^{\frac{1}{p}}\mu)\right),

and

ωμ=1p(ζ+lnp)(1ω(ζ+lnp))1p1,ζμ=1pθ(1ω(ζ+lnp))1p1.\partial_{\omega}\mu=\frac{1}{p}(\zeta+\ln p)\left(1-\omega(\zeta+\ln p)\right)^{-\frac{1}{p}-1},\ \partial_{\zeta}\mu=\frac{1}{p}\theta\left(1-\omega(\zeta+\ln p)\right)^{-\frac{1}{p}-1}. (4.75)

This yields

ωμ(0,lnδ0)=lnδ0+lnpp,ζμ(0,lnδ0)=0,\partial_{\omega}\mu(0,\ln\delta_{0})=\frac{\ln\delta_{0}+\ln p}{p},\ \partial_{\zeta}\mu(0,\ln\delta_{0})=0, (4.76)

and then

ωJ2(0,lnδ0)=(1+δ0)(lnδ0+lnp)p,ζJ2(0,lnδ0)=lnδ0.\partial_{\omega}J_{2}(0,\ln\delta_{0})=\frac{(1+\delta_{0})(\ln\delta_{0}+\ln p)}{p},\ \partial_{\zeta}J_{2}(0,\ln\delta_{0})=\ln\delta_{0}. (4.77)

So we can obtain

ζ(ω)=lnδ0(1+δ0)(lnδ0+lnp)plnδ0ω+Oδ0(ω2),\zeta(\omega)=\ln\delta_{0}-\frac{(1+\delta_{0})(\ln\delta_{0}+\ln p)}{p\ln\delta_{0}}\omega+O_{\delta_{0}}(\omega^{2}), (4.78)

and then

μ=(1ω(ζ+lnp))1p=1+lnδ0+lnppsp+Oδ0(s2p).\mu=\left(1-\omega(\zeta+\ln p)\right)^{-\frac{1}{p}}=1+\frac{\ln\delta_{0}+\ln p}{p}s^{p}+O_{\delta_{0}}(s^{2p}). (4.79)

By (4.73) and (4.79), we have that for p>0p>0 and 0<s<ε00<s<\varepsilon_{0},

μ(s)=1+lnδ0+lnppsp+Oδ0(smin{2p,p+1}),\mu(s)=1+\frac{\ln\delta_{0}+\ln p}{p}s^{p}+O_{\delta_{0}}(s^{\min\{2p,p+1\}}), (4.80)

where ε0>0\varepsilon_{0}>0 is a small constant depending on δ0\delta_{0}. It follows that for s[0,ε0)s\in[0,\varepsilon_{0}),

y(t,(1+δ0)sτ)=sμ(s)=s+lnδ0+lnppsp+1+Oδ0(smin{2p+1,p+2}).y(t,(1+\delta_{0})s\tau)=s\mu(s)=s+\frac{\ln\delta_{0}+\ln p}{p}s^{p+1}+O_{\delta_{0}}(s^{\min\{2p+1,p+2\}}). (4.81)

Secondly, we consider the behavior of y(t,(1δ0)sτ)y(t,(1-\delta_{0})s\tau). For x=(1δ0)sτx=(1-\delta_{0})s\tau, μ=y(t,(1δ0)sτ)s\mu=\frac{y(t,(1-\delta_{0})s\tau)}{s} satisfies

L(s,μ)μ(1+1pesp(1μp))μpespμp+esp1sespμpr0(sμ)(1δ0)=0.L(s,\mu)\triangleq\mu\left(1+\frac{1}{p}e^{s^{-p}(1-\mu^{-p})}\right)-\frac{\mu}{p}e^{-s^{-p}\mu^{-p}}+\frac{e^{s^{-p}}-1}{s}e^{-s^{-p}\mu^{-p}}r_{0}(s\mu)-(1-\delta_{0})=0. (4.82)

It is easy to know μ=1δ0\mu=1-\delta_{0} as s0+s\to 0_{+}. Furthermore, by direct computation, one has

sL(s,μ)\displaystyle\partial_{s}L(s,\mu) =\displaystyle= sp1(μ+psr0(sμ))(esp(1μp)(1μp)espμpμp)\displaystyle s^{-p-1}\left(\mu+\frac{p}{s}r_{0}(s\mu)\right)\left(e^{s^{-p}(1-\mu^{-p})}(1-\mu^{-p})-e^{-s^{-p}\mu^{-p}}\mu^{-p}\right)
+esp(1μp)espμps2(r0(sμ)+sμr0(sμ)),\displaystyle+\frac{e^{s^{-p}(1-\mu^{-p})}-e^{-s^{-p}\mu^{-p}}}{s^{2}}\left(-r_{0}(s\mu)+s\mu r^{\prime}_{0}(s\mu)\right),
μL(s,μ)\displaystyle\partial_{\mu}L(s,\mu) =\displaystyle= (1+1pesp(1μp)1pespμp)+spμp(esp1)espμp\displaystyle\left(1+\frac{1}{p}e^{s^{-p}(1-\mu^{-p})}-\frac{1}{p}e^{-s^{-p}\mu^{-p}}\right)+s^{-p}\mu^{-p}\left(e^{s^{-p}}-1\right)e^{-s^{-p}\mu^{-p}}
+(esp1)espμp(psp1μp1r0(sμ)+r0(sμ)).\displaystyle+\left(e^{s^{-p}}-1\right)e^{-s^{-p}\mu^{-p}}\left(ps^{-p-1}\mu^{-p-1}r_{0}(s\mu)+r^{\prime}_{0}(s\mu)\right).

By assumption (1.8), we have

sL(0,1δ0)=0,μL(0,1δ0)=1.\partial_{s}L(0,1-\delta_{0})=0,\quad\partial_{\mu}L(0,1-\delta_{0})=1. (4.83)

Then there exists a function

μ(s)=1δ0+Oδ0(s2)\mu(s)=1-\delta_{0}+O_{\delta_{0}}(s^{2}) (4.84)

such that L(s,μ(s))=0L(s,\mu(s))=0 for s>0s>0 sufficiently small and dependent on δ0\delta_{0}. It follows that for s[0,ε0)s\in[0,\varepsilon_{0}),

y(t,(1δ0)sτ)=sμ(s)=(1δ0)s+Oδ0(s3).y(t,(1-\delta_{0})s\tau)=s\mu(s)=(1-\delta_{0})s+O_{\delta_{0}}(s^{3}). (4.85)

Thus for small fixed δ0>0\delta_{0}>0, we have that for s[0,ε0)s\in[0,\varepsilon_{0}) and x((1δ0)sτ,(1+δ0)sτ)x\in((1-\delta_{0})s\tau,(1+\delta_{0})s\tau),

12sy(t,x)32s.\frac{1}{2}s\leq y(t,x)\leq\frac{3}{2}s. (4.86)

Recalling u(t,x)=u0(y(t,x))u(t,x)=u_{0}(y(t,x)), we have that

|u(t,x)u(1,0)||y(t,x)|s=|ln(1t)|1p,|u(t,x)-u(1,0)|\lesssim|y(t,x)|\lesssim s=|\ln(1-t)|^{-\frac{1}{p}}, (4.87)

and by

1+tg(y(t,x))=(1t)+(1pe|y|p+|y|pe|y|p+r0(y))|ln(1t)|(1t),1+tg^{\prime}(y(t,x))=(1-t)+\left(\frac{1}{p}e^{-|y|^{-p}}+|y|^{-p}e^{-|y|^{-p}}+r^{\prime}_{0}(y)\right)\gtrsim|\ln(1-t)|(1-t), (4.88)

we have that for (t,x)Ωt,,+0(t,x)\in\Omega_{t,-,+}^{0},

ux(t,x)\displaystyle\frac{\partial u}{\partial x}(t,x) =\displaystyle= u0(y(t,x))1+tg(y(t,x))|ln(1t)|1(1t)1,\displaystyle\frac{u^{\prime}_{0}(y(t,x))}{1+tg^{\prime}(y(t,x))}\lesssim|\ln(1-t)|^{-1}(1-t)^{-1}, (4.89)
ut(t,x)\displaystyle\frac{\partial u}{\partial t}(t,x) =\displaystyle= u0(y(t,x))g(y(t,x))1+tg(y(t,x))|ln(1t)|11p(1t)1.\displaystyle-\frac{u^{\prime}_{0}(y(t,x))g(y(t,x))}{1+tg^{\prime}(y(t,x))}\lesssim|\ln(1-t)|^{-1-\frac{1}{p}}(1-t)^{-1}. (4.90)

We decompose the neighbourhood BB of (1,0)(1,0) into Ωx,+,Ωx,,Ωjt,+,Ωjt,,m\Omega_{x,+},\ \Omega_{x,-},\ \Omega^{j}_{t,+},\ \Omega^{j}_{t,-,m} for j=1,2,,Nj=1,2,\ldots,N and Ωjt,,+,Ωjt,,\Omega^{j}_{t,-,+},\ \Omega^{j}_{t,-,-} for j=0,1,2,,Nj=0,1,2,\ldots,N as follows (see Figure 2 below).

Refer to caption
Figure 2: Decomposition of BB
Ωx,+\displaystyle\Omega_{x,+} =\displaystyle= {(t,x):x>0, 0<ξ<δ,ε<η<ε},\displaystyle\{(t,x):\ x>0,\ 0<\xi<\delta,\ -\varepsilon<\eta<\varepsilon\}, (4.91)
Ωx,\displaystyle\Omega_{x,-} =\displaystyle= {(t,x):x<0, 0<ξ<δ,ε<η<ε},\displaystyle\{(t,x):\ x<0,\ 0<\xi<\delta,\ -\varepsilon<\eta<\varepsilon\}, (4.92)

where (ξ,η)(\xi,\eta) are defined in (4.24). Taking the suitable constants c1,+<c2,+<<cN1,+<cN,+c_{1,+}<c_{2,+}<\ldots<c_{N-1,+}<c_{N,+}, {εj,+}j=1N\{\varepsilon_{j,+}\}_{j=1}^{N} and {δj,+}j=1N\{\delta_{j,+}\}_{j=1}^{N}, and setting

Ωjt,+={(t,x): 0<s<εj,+,cj,+δj,+<λ<cj,++δj,+},s=|ln(t1)|1p,λ=xs(t1),\Omega^{j}_{t,+}=\{(t,x):\ 0<s<\varepsilon_{j,+},\ c_{j,+}-\delta_{j,+}<\lambda<c_{j,+}+\delta_{j,+}\},s=|\ln(t-1)|^{-\frac{1}{p}},\ \lambda=\frac{x}{s(t-1)}, (4.93)

where B{t0}(j=1NΩjt,+)Ωx,+Ωx,+B\cap\{t\geq 0\}\subset\left(\cup_{j=1}^{N}\Omega^{j}_{t,+}\right)\cup\Omega_{x,+}\cup\Omega_{x,+}. By choosing small δ0>0\delta_{0}>0, we can define Ω0t,,+\Omega^{0}_{t,-,+} and Ω0t,,\Omega^{0}_{t,-,-} as in (4.63) and (4.65). Meanwhile, taking some suitable constants {c1,,m}j=1N\{c_{1,-,m}\}_{j=1}^{N}, {εj,,m}j=1N\{\varepsilon_{j,-,m}\}_{j=1}^{N} and {δj,,m}j=1N\{\delta_{j,-,m}\}_{j=1}^{N} where m=+,0,m=+,0,-, and setting

Ωjt,,m={(t,x): 0<s<εj,,m,cj,,mδj,,m<λ<cj,,m+δj,,m},s=|ln(1t)|1p,λ=xs(1t),\Omega^{j}_{t,-,m}=\{(t,x):\ 0<s<\varepsilon_{j,-,m},\ c_{j,-,m}-\delta_{j,-,m}<\lambda<c_{j,-,m}+\delta_{j,-,m}\},s=|\ln(1-t)|^{-\frac{1}{p}},\ \lambda=\frac{x}{s(1-t)}, (4.94)

where cN,,<<c1,,<1<c1,,0<c1,,0<<cN,,0<1<c1,,+<<cN,,+c_{N,-,-}<\ldots<c_{1,-,-}<-1<c_{1,-,0}<c_{1,-,0}<\ldots<c_{N,-,0}<1<c_{1,-,+}<\ldots<c_{N,-,+}, such that B{t0}(j=0NΩjt,,+)(j=1NΩjt,,0)(j=0NΩjt,,)Ωx,+Ωx,+B\cap\{t\leq 0\}\subset\left(\cup_{j=0}^{N}\Omega^{j}_{t,-,+}\right)\cup\left(\cup_{j=1}^{N}\Omega^{j}_{t,-,0}\right)\cup\left(\cup_{j=0}^{N}\Omega^{j}_{t,-,-}\right)\cup\Omega_{x,+}\cup\Omega_{x,+} holds. Note that in Ωx,±\Omega_{x,\pm}, y(t,x)ξy(t,x)\sim\xi; and in others, y(t,x)sy(t,x)\sim s. Then similarly to (4.87), we can obtain (1.15). On the other hand, in Ωx,±\Omega_{x,\pm}, 1+tg(y(t,x))|x||ln|x||1+tg^{\prime}(y(t,x))\gtrsim|x||\ln|x||; and in others, 1+tg(y(t,x))|t1||ln|t1||1+tg^{\prime}(y(t,x))\gtrsim|t-1||\ln|t-1|| as in (4.88). Then similarly to the proof for (4.89) and (4.90), we can establish (1.16) and (1.17).

\square

References

  • [1] S. Alinhac, Blowup of small data solutions for a class of quasilinear wave equations in two space dimensions. Ann. of Math. 149 (1999), no. 1, 97-127.
  • [2] T. Buckmaster, S. Shkoller, V. Vicol, Formation of shocks for 2D isentropic compressible Euler, arXiv: 1907.03784v2, 8 Jul 2019.
  • [3] T. Buckmaster, S. Shkoller, V. Vicol, Formation of point shocks for 3D compressible Euler, arXiv: 1912.04429v2, 22 Jun 2020.
  • [4] Chen Shuxing, Xin Zhouping, Yin Huicheng, Formation and construction of shock wave for quasilinear hyperbolic system and its application to inviscid compressible flow. The Institute of Mathematical Sciences at CUHK, 2010, Research Reports: 2000-10(069)
  • [5] Chen Shuxing, Dong Liming, Formation of shock for the p-system with general smooth initial data. Sci. in China, Ser. A, 44 (2001), no. 9, 1139-1147.
  • [6] D. Christodoulou, The formation of shocks in 3-dimensional fluids. EMS Monographs in Mathematics. European Mathematical Society (EMS), Zürich, 2007.
  • [7] D. Christodoulou, A. Lisibach, Shock development in spherical symmetry. Ann. PDE 2 (2016), no. 1, Art. 3, 246 pp.
  • [8] D. Christodoulou, Miao Shuang, Compressible flow and Euler’s equations, Surveys of Modern Mathematics, 9. International Press, Somerville, MA; Higher Education Press, Beijing, 2014.
  • [9] G. Holzegel, S. Klainerman, J. Speck, Wong Willie Wai-Yeung, Small-data shock formation in solutions to 3D quasilinear wave equations: an overview. J. Hyperbolic Differ. Equ. 13 (2016), no. 1, 1-105.
  • [10] Kong Dexing, Formation and propagation of singularities for 2×22\times 2 quasilinear hyperbolic systems. Trans. Amer. Math. Soc. 354 (2002), no. 8, 3155-3179.
  • [11] M.P. Lebaud, Description de la formation d’un choc dans le pp-système. J. Math. Pures Appl. (9) 73 (1994), no. 6, 523-565.
  • [12] J. Luk, J. Speck, Shock formation in solutions to the 2D compressible Euler equations in the presence of non-zero vorticity. Invent. Math. 214 (2018), no. 1, 1-169.
  • [13] F. Merle, P. Raphaël, I. Rodniaski, J. Szefel, On the implosion of a three dimensional compressible fluid, arXiv: 1912.11009v2, 13 Jun 2020.
  • [14] Miao Shuang, Yu Pin, On the formation of shocks for quasilinear wave equations. Invent. Math. 207 (2017), no. 2, 697-831.
  • [15] J. Speck, Shock formation for 2D quasilinear wave systems featuring multiple speeds: blowup for the fastest wave, with non-trivial interactions up to the singularity. Ann. PDE 4 (2018), no. 1, Art. 6, 131 pp.
  • [16] Yin Huicheng, Formation and construction of a shock wave for 3-D compressible Euler equations with the spherical initial data. Nagoya Math. J. 175 (2004), 125-164.