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

Convergence of Solutions of the Porous Medium Equation with Reactions

Bendong Lou  and Maolin Zhou
B. Lou: Mathematics and Science College, Shanghai Normal University, Shanghai, 200234, China. Email: [email protected]
M. Zhou: Chern Institute of Mathematics, Nankai University, Tianjin, 300071, China. Email: [email protected]
This research was partially supported by the National Key Research and Development Program of China (2021YFA1002400) and the NSF of China (No. 12071299).

Abstract. Consider the Cauchy problem of one dimensional porous medium equation (PME) with reactions. We first prove a general convergence result, that is, any bounded global solution starting at a nonnegative compactly supported initial data converges as tt\to\infty to a nonnegative zero of the reaction term or a ground state stationary solution. Based on it, we give out a complete classification on the asymptotic behaviors of the solutions for PME with monostable, bistable and combustion types of nonlinearities.

Key words and phrases: Porous medium equation with reactions (RPME); Cauchy problem; free boundary; asymptotic behavior; spreading and vanishing.

2020 Mathematics Subject Classification: 35K65, 35K57,35B40,35K15.

1. Introduction

Consider the following reaction diffusion equation (abbreviated as RDE in the following)

(1.1) ut=uxx+f(u),x,t>0.u_{t}=u_{xx}+f(u),\quad x\in{\mathbb{R}},\ t>0.

In the last decades, many authors gave systemic qualitative study for this equation. The reaction terms include the following three typical types:

(fM)  monostable case,       (fB)  bistable case,      (fC)  combustion case.

In the monostable case, we assume f=f(u)C1([0,))f=f(u)\in C^{1}([0,\infty)) and

(1.2) f(0)=f(1)=0,f(0)>0,f(1)<0,(1u)f(u)>0for u>0,u1.f(0)=f(1)=0,\quad f^{\prime}(0)>0,\quad f^{\prime}(1)<0,\quad(1-u)f(u)>0\ \mbox{for }u>0,u\not=1.

The well known example is the logistic term, that is, the Verhulst law: f(u)=u(1u)f(u)=u(1-u). In the bistable case, we assume f=f(u)C1([0,))f=f(u)\in C^{1}([0,\infty)) and

(1.3) f(0)=f(θ)=f(1)=0,f(u){<0in (0,θ),>0in (θ,1),<0in (1,),01f(s)𝑑s>0,f(0)=f(\theta)=f(1)=0,\quad f(u)\left\{\begin{array}[]{l}<0\ \ \mbox{in }(0,\theta),\\ >0\ \ \mbox{in }(\theta,1),\\ <0\ \ \mbox{in }(1,\infty),\end{array}\right.\quad\int_{0}^{1}f(s)ds>0,

for some θ(0,1)\theta\in(0,1), f(0)<0f^{\prime}(0)<0, f(1)<0f^{\prime}(1)<0. A typical example is f(u)=u(uθ)(1u)f(u)=u(u-\theta)(1-u) with θ(0,12)\theta\in(0,\frac{1}{2}). In the combustion case, we assume there exists θ(0,1)\theta\in(0,1) such that fC1([θ,])f\in C^{1}([\theta,\infty]) and

(1.4) {f(u)=0in [0,θ],f(u)>0in (θ,1),f(1)<0,f(u)<0in (1,),there exists a small δ>0 such that f(u)>0 for u(θ,θ+δ].\left\{\begin{array}[]{l}f(u)=0\ \ \mbox{in }[0,\theta],\quad f(u)>0\ \mbox{in }(\theta,1),\quad f^{\prime}(1)<0,\quad f(u)<0\ \mbox{in }(1,\infty),\\ \mbox{there exists a small }\delta>0\mbox{ such that }f^{\prime}(u)>0\mbox{ for }u\in(\theta,\theta+\delta].\end{array}\right.

In 1937, Fisher [17] and Kolmogorov et al. [22] studied the RDE with logistic reaction in their pioneer works. In particular, they specified the traveling wave solutions. In 1975 and 1978, Aronson and Weinberger [6, 7] studied the asymptotic behavior for the solutions of the Cauchy problem of RDE. Among others, in the monostable case, they proved the hair-trigger effect, which says that spreading happens (that is, u1u\to 1 as tt\to\infty, which is also called persistence or propagation phenomenon) for any positive solution of the monostable equation; in the bistable case, they gave sufficient conditions for spreading and that for vanishing (that is, u0u\to 0 as tt\to\infty, which is also called extinction phenomenon). In 1977, Fife and McLeod [16] also studied the bistable equation, proved the existence and stability of the traveling wave solution. In 2006, Zlatoš [32] gave further systemic study on the asymptotic behavior for the solutions of bistable and combustion equations. More precisely, he considered the Cauchy problem of (1.1) with initial data u(x,0)=χ[L,L](x)u(x,0)=\chi_{[-L,L]}(x) (which denotes the characteristic function over [L,L][-L,L]), and proved a trichotomy result: there exists a sharp value L>0L_{*}>0 such that when L>LL>L_{*} (resp. L<LL<L_{*}), spreading (resp. vanishing) happens for the solution. In addition, the solution develops to a transition one if and only if L=LL=L_{*}, which is a ground state solution in the bistable case and the largest zero θ(0,1)\theta\in(0,1) of ff in the combustion case. In 2010, Du and Matano [13] extended these results to the Cauchy problem with general initial data like u(x,0)=σϕ(x)u(x,0)=\sigma\phi(x) for positive number σ\sigma and continuous, nonnegative, compactly supported function ϕ\phi. For bistable and combustion equations, they also proved the trichotomy and sharp transition results as in [32]. In 2015, Du and Lou [11] proved similar results for the nonlinear Stefan problem.

In this paper, we study the porous medium equation with reactions:

(CP) {ut=(um)xx+f(u),x,t>0,u(x,0)=u0(x),x,\mbox{(CP)}\hskip 113.81102pt\left\{\begin{array}[]{ll}u_{t}=(u^{m})_{xx}+f(u),&\ \ x\in{\mathbb{R}},\ t>0,\\ u(x,0)=u_{0}(x),&\ \ x\in{\mathbb{R}},\end{array}\right.\hskip 142.26378pt

where m>1m>1 is a constant. This problem can be used to model population dynamics with diffusion flux depending on the population density, the combustion, propagation of interfaces and nerve impulse propagation phenomena in porous media, as well as the propagation of intergalactic civilizations in the field of astronomy (cf. [20, 21, 24, 27] etc.). For simplicity, we use PME/RPME to refer to the porous media equation in (CP) without/with a reaction.

The well-posedness of the problem (CP) was studied in [2, 3, 26, 30] etc.. For the qualitative property, the monostable equation was considered widely. For example, in 1979, Aronson [3] considered the problem with logistic reaction. He studied the existence of traveling wave solutions and compared the results with that in RDE. It was shown that (see also [19, 25, 29]) the monostable RPME has a traveling wave solution u(x,t)=U~(xct)u(x,t)=\widetilde{U}(x-ct) if and only if m1m\geq 1 and only for the speed cc(m)c\geq c_{*}(m). For bistable and combustion RPMEs, however, there is only one traveling wave solution U~(xct)\widetilde{U}(x-c_{*}t) (cf. [19]). Recently, Du, Quirós and Zhou [14] considered the (NN-dimensional) RPME with Fisher-KPP type of reaction. In particular, they presented a precise estimate for the spreading speed of the free boundary with a logarithmic correction. On the other hand, the qualitative properties for the RPMEs with bistable or combustion reaction were not as clear as the monostable case. In 1982, Aronson, Crandall and Peletier [5] considered the bistable RPME in a bounded interval. Besides the well-posedness, they specified the stationary solutions which can be ω\omega-limits of the solutions. Recently, Gárriz [18] considered the problem (CP) with reactions of the monostable, bistable and combustion types. He gave sufficient conditions ensuring the spreading or vanishing, and used the traveling wave solution to characterize the spreading solutions starting at compactly supported initial data.

Our aim in this paper is to give out a complete classification for the asymptotic behaviors of the solutions of (CP). We will specify its stationary solutions; prove a general convergence result, that is, any nonnegative bounded global solution converges as tt\to\infty to a stationary one. Then, based on the general result, we prove the hair-trigger effect for monostable PME and spreading-transition-vanishing trichotomy results on the asymptotic behavior for solutions of bistable or combustion RPMEs. In some sense, this paper can be regarded as a RPME version of the important works [11, 13, 32] for RDEs.

Our basic assumption on ff is

(F) f(u)C2([0,)) with Lipschitz number K,f(0)=0 and f(u)<0 for u>1.\mbox{(F)}\hskip 14.22636ptf(u)\in C^{2}([0,\infty))\mbox{ with Lipschitz number }K,\ \ f(0)=0\mbox{ and }f(u)<0\mbox{ for }u>1.\hskip 71.13188pt

The C2C^{2} smoothness is mainly used to give the a priori estimates, including the lower bound for vxxv_{xx} (where vv is the pressure, see Appendix for details). In the study of asymptotic behavior, we actually only need fC1f\in C^{1}. The assumption f(u)<0f(u)<0 for u>1u>1 is natural in population dynamics, which means that the population has a finite capacity. If one only studies the well-posedness, this requirement can be weakened or omitted. The initial data are chosen from the following set (see Figure 1)

(I)u0𝔛:={ψC()|there exist b>0 and a finite number of open intervals in [b,b]such that ψ(x)>0 in these intervals, and ψ(x)=0 otherwise}.{\rm(I)}\hskip 14.22636ptu_{0}\in\mathfrak{X}:=\left\{\psi\in C({\mathbb{R}})\left|\begin{array}[]{l}\mbox{there exist }b>0\mbox{ and a finite number of open intervals in }[-b,b]\\ \mbox{such that }\psi(x)>0\mbox{ in these intervals, and }\psi(x)=0\mbox{ otherwise}\end{array}\right.\right\}.\ \ \ \

It is also possible to consider more general initial data, for example, u0C()u_{0}\in C({\mathbb{R}}) with compact support, or, u0L1()L()u_{0}\in L^{1}({\mathbb{R}})\cap L^{\infty}({\mathbb{R}}). For simplicity and clarity of presentation, especially, for clarity of the free boundaries, we consider in this paper only the initial data in 𝔛\mathfrak{X}. In this case, the solution has a left-most free boundary l(t)l(t), a right-most free boundary r(t)r(t), and finite number of interior free boundaries in (l(t),r(t))(l(t),r(t)).

Refer to caption
Figure 1. An example of the initial data.

For any T>0T>0, denote QT:=×(0,T)Q_{T}:={\mathbb{R}}\times(0,T). A function u(x,t)C(QT)L(QT)u(x,t)\in C(Q_{T})\cap L^{\infty}(Q_{T}) is called a very weak solution of (CP) if for any φCc(QT)\varphi\in C^{\infty}_{c}(Q_{T}), there holds

(1.5) u(x,T)φ(x,T)𝑑x=u0(x)φ(x,0)𝑑x+QTf(u)φ𝑑x𝑑t+QT[uφt+umφxx]𝑑x𝑑t.\int_{{\mathbb{R}}}u(x,T)\varphi(x,T)dx=\int_{{\mathbb{R}}}u_{0}(x)\varphi(x,0)dx+\iint_{Q_{T}}f(u)\varphi dxdt+\iint_{Q_{T}}[u\varphi_{t}+u^{m}\varphi_{xx}]dxdt.

As an extension of the definition, if uu satisfies (1.5) with inequality “\geq” (resp. “\leq”), instead of equality, for every test function φ0\varphi\geq 0, then uu is called a very weak supersolution/very weak upper solution (resp. very weak subsolution/very weak lower solution) of (CP) (cf. [30, Chapter 5]).

From the references [5, 25, 26, 30] etc. we know that under the assumption (F) and (I), the problem (CP) has a unique very weak solution u(x,t)C(QT)L(QT)u(x,t)\in C(Q_{T})\cap L^{\infty}(Q_{T}) for any T>0T>0, u(x,t)0u(x,t)\geq 0 in QTQ_{T}, and the support of u(,t)u(\cdot,t), denoted by spt[u(,t)]{\rm spt}[u(\cdot,t)], is contained in [l(t),r(t)][l(t),r(t)] for all t>0t>0. Moreover, we will show below that both l(t)-l(t) and r(t)r(t) are non-decreasing Lipschitz functions, as in PME, and so the following limits exist:

(1.6) l:=limtl(t),r:=limtr(t).l^{\infty}:=\lim\limits_{t\to\infty}l(t),\quad r^{\infty}:=\lim\limits_{t\to\infty}r(t).

With the global existence in hand, it is possible to study the asymptotic behavior for the solutions. For RDEs, the equations are uniform parabolic and the strong maximum principle is applicable. So, any solution is a classical one and the convergence of uu to its ω\omega-limit is taken in the topology of Cloc2,1()C^{2,1}_{loc}({\mathbb{R}}). For our RPME, however, we have only very weak solutions with CαC^{\alpha} bounds in any compact domain. A nonnegative solution is not necessarily to be positive and classical. Hence the convergence of a solution to its limit is first considered in the topology of Cloc()C_{loc}({\mathbb{R}}) or Lloc()L^{\infty}_{loc}({\mathbb{R}}). In addition, if u(x,tn)w(x)>0u(x,t_{n})\to w(x)>0 as tnt_{n}\to\infty in some domain JJ, then u>0u>0 in JJ by the positivity persistence and so it is a classical solution. Thus, the convergence u(x,tn)w(x)u(x,t_{n})\to w(x) holds in C2C^{2} topology in any compact subdomain of JJ. In summary, if no other specification, throughout this paper we use the following definition for ω\omega-limits of uu:

(1.7) limnu(x,tn)=w(x) means {u(x,tn)w(x) in Cloc() topology;u(x,tn)w(x) in Cloc2(J) topology if w(x)>0 in J.\lim\limits_{n\to\infty}u(x,t_{n})=w(x)\mbox{ means }\left\{\begin{array}[]{l}u(x,t_{n})\to w(x)\mbox{ in }C_{loc}({\mathbb{R}})\mbox{ topology};\\ u(x,t_{n})\to w(x)\mbox{ in }C^{2}_{loc}(J)\mbox{ topology if }w(x)>0\mbox{ in }J.\end{array}\right.

In some cases, our problem may have a ground state solution, which means one of the following two types of nonnegative stationary solutions of (CP):

\bullet Type I ground state solution (see Figure 2): U0(x)C2()U_{0}(x)\in C^{2}({\mathbb{R}}) is an even stationary solution of (CP) with

U0(x)>0>U0(x) for x>0,U0(+)[0,1);U_{0}(x)>0>U^{\prime}_{0}(x)\mbox{ for }x>0,\quad U_{0}(+\infty)\in[0,1);
Refer to caption
Refer to caption
Figure 2. Type I ground state solution.

\bullet Type II ground state solution (see Figure 3): there exist a positive integer kk, zi(i=1,2,,k)z_{i}\in{\mathbb{R}}\ (i=1,2,\cdots,k) and L>0L>0 with zi+2Lzi+1(i=1,2,,k1)z_{i}+2L\leq z_{i+1}\ (i=1,2,\cdots,k-1) such that

(1.8) 𝒰(x)=U(xz1)+U(xz2)++U(xzk),x,\mathcal{U}(x)=U(x-z_{1})+U(x-z_{2})+\cdots+U(x-z_{k}),\quad x\in{\mathbb{R}},

where U(x)C()C2(\{±L})U(x)\in C({\mathbb{R}})\cap C^{2}({\mathbb{R}}\backslash\{\pm L\}) is an even stationary solution of the problem (CP) with

(1.9) U(x)>0>U(x) for x(0,L),U(x)=0 for xL,(Um1)(L)=0.U(x)>0>U^{\prime}(x)\mbox{ for }x\in(0,L),\quad U(x)=0\mbox{ for }x\geq L,\quad(U^{m-1})^{\prime}(L)=0.
Refer to caption
Figure 3. Type II ground state solution.

A Type I ground state solution is positive on the whole {\mathbb{R}}, which is the analogue of that in bistable RDEs (cf. [13, 32]). For our RPME, this kind of solution exists if, in particular, ff is a bistable nonlinearity with

(1.10) f(u)=λuα(1+o(1)) as u0+0,f(u)=-\lambda u^{\alpha}(1+o(1))\mbox{\ \ as\ \ }u\to 0+0,

for some λ>0\lambda>0 and αm\alpha\geq m (see Lemma 6.1). The component UU in Type II ground state solution has a compact support, which is different from Type I. Such a solution exists when (1.10) holds for 0<α<m0<\alpha<m (see Lemma 6.1). The last equality in (1.9) is used to indicate that U(x)U(x) is a stationary solution not only of the RPME but also of the Cauchy problem (CP). In fact, a necessary and sufficient condition ensuring a stationary solution u¯\bar{u} of RPME with compact support [b,b][-b,b] is also a stationary solution of the Cauchy problem (CP) is that the waiting times at its boundaries ±b\pm b are infinite, which requires that (u¯m1)(±b)=0(\bar{u}^{m-1})(\pm b)=0 (see Lemma 2.3 below).

Our first main theorem is a general convergence result, which says that any nonnegative bounded global solution of (CP) converges to a nonnegative zero of ff or a ground state solution:

Theorem 1.1 (General convergence theorem).

Assume (F) and (I). Let u(x,t)u(x,t) be a bounded, nonnegative, time-global solution of (CP). Then u(,t)u(\cdot,t) converges as tt\to\infty, in the sense of (1.7), to a stationary solution of (CP), which is one of the following types:

  1. (i)

    a nonnegative zero of ff;

  2. (ii)

    U0(xz0)U_{0}(x-z_{0}) for some z0[b,b]z_{0}\in[-b,b] and some Type I ground state solution U0U_{0};

  3. (iii)

    𝒰(x)\mathcal{U}(x) as in (1.8) for k1k\geq 1 points z1,z2,,zk[b,b]z_{1},z_{2},\cdots,z_{k}\in[-b,b]. In addition, l,rl^{\infty},\ r^{\infty} are bounded in this case.

This theorem is a fundamental result and can be used conveniently to study various RPME. For example, when the reaction term in RPME is a typical monostable, bistable or combustion one, it is easy to clarify all of the nonnegative stationary solutions. Hence, using the general convergence result in the previous theorem, we can give a rather complete analysis on the asymptotic behavior for the solutions of (CP) with these reactions.

To obtain the above convergence result, our main tool is the so-called zero number argument in the version of porous medium equation, which can be used to describe the intersection points between two solutions. The main difficulty to establish the argument is brought by the existence of free boundaries and the degeneracy of diffusion here, which is a significant difference compared with the classical heat equation. In particular, the intersection number may increase. For the convenience of the readers, we present the intersection number properties as a theorem in the below. Let u1,u2u_{1},u_{2} be two solutions of (CP) with initial data in 𝔛\mathfrak{X}, then they are zero outside of their supports. It is a little confusing to define the intersection number between u1u_{1} and u2u_{2} in these places. Instead, we denote 𝒵0(t)\mathcal{Z}_{0}(t) as the number of their positive intersection points, that is,

(1.11) 𝒵0(t):=#{xu1(x,t)=u2(x,t)>0},t0,\mathcal{Z}_{0}(t):=\#\{x\in{\mathbb{R}}\mid u_{1}(x,t)=u_{2}(x,t)>0\},\qquad t\geq 0,

where #S\#S denotes the number of the elements in a set SS.

Theorem 1.2 (Intersection number properties).

Assume (F). Let u1,u2u_{1},u_{2} be bounded, nonnegative, time-global solutions of (CP) with initial data in 𝔛\mathfrak{X}. If 𝒵0(0)<\mathcal{Z}_{0}(0)<\infty, then 𝒵0(t)<\mathcal{Z}_{0}(t)<\infty for all t>0t>0, and there exists a time T0T\geq 0 such that

𝒵0(t) decreases for t>T.\mathcal{Z}_{0}(t)\mbox{ decreases for }t>T.

More specifically, there exist {tj}0jk\{t_{j}\}_{0\leq j\leq k} with 1k1\leq k\in\mathbb{N} and 0=t0<t1<<T:=tk1<tk=+0=t_{0}<t_{1}<\cdots<T:=t_{k-1}<t_{k}=+\infty such that

  1. (i)

    𝒵0(t)\mathcal{Z}_{0}(t) decreases if tj<t<tj+1(j=0,1,,k1)t_{j}<t<t_{j+1}\ (j=0,1,\cdots,k-1);

  2. (ii)

    𝒵0(t)\mathcal{Z}_{0}(t) strictly increases at t=tj(j=1,,k1)t=t_{j}\ (j=1,\cdots,k-1) in the following sense:

    limttj+0𝒵0(t)>limttj0𝒵0(t).\lim_{t\rightarrow t_{j}+0}\mathcal{Z}_{0}(t)>\lim_{t\rightarrow t_{j}-0}\mathcal{Z}_{0}(t).

The last property (that is, 𝒵0(t)\mathcal{Z}_{0}(t) increases at tjt_{j}) happens in particular when a component where u1>0u_{1}>0 meets a component where u2>0u_{2}>0 as their boundaries approaching to each other (see more in Lemma 3.5). As we have mentioned above, this is significant different from the classical zero number diminishing properties. Such a phenomena happens only in PMEs where a solution has free boundaries but not in RDEs.

In order to describe the threshold of transition solutions, as in [11, 13, 32], we introduce a one-parameter family of initial data u0=ϕσ(x)u_{0}=\phi_{\sigma}(x) satisfying the following conditions:

  • (Φ1\Phi_{1})

    ϕσ𝔛\phi_{\sigma}\in\mathfrak{X} for every σ>0\sigma>0, and the map σϕσ\sigma\mapsto\phi_{\sigma} is continuous from (0,)(0,\infty) to L()L^{\infty}({\mathbb{R}});

  • (Φ2\Phi_{2})

    if 0<σ1<σ20<\sigma_{1}<\sigma_{2}, then ϕσ1,ϕσ2\phi_{\sigma_{1}}\leq,\not\equiv\phi_{\sigma_{2}};

  • (Φ3\Phi_{3})

    limσ0ϕσL()=0\lim\limits_{\sigma\to 0}\|\phi_{\sigma}\|_{L^{\infty}({\mathbb{R}})}=0.

Theorem 1.3.

Assume fC2f\in C^{2} is a monostable reaction term satisfying (1.2). Let uσ(x,t)u_{\sigma}(x,t) be the very weak solution of (CP) with u0=ϕσ(x)u_{0}=\phi_{\sigma}(x) satisfying (Φ1)(\Phi_{1})-(Φ3)(\Phi_{3}). Then the hair-trigger effect holds: for all σ>0\sigma>0, we have

uσ(,t)1 as t, in the topology of Lloc().u_{\sigma}(\cdot,t)\to 1\mbox{\ \ as\ \ }t\to\infty,\ \ \mbox{ in the topology of }L^{\infty}_{loc}({\mathbb{R}}).
Remark 1.4.

Theorem 1.3 is a direct application of Theorem 1.1, since the structure of the ω\omega-limit set is rather simple in this case (see details in Section 4). Through a different proof, [18] obtained a deeper result on the hair-trigger effect for more general monostable RPME in NN-dimensional space, where ff satisfies lim infu0+0f(u)um+2/N>0\liminf\limits_{u\rightarrow 0+0}\frac{f(u)}{u^{m+2/N}}>0.

In the combustion case, we have the following result.

Theorem 1.5.

Assume fC2f\in C^{2} is a combustion reaction term satisfying (1.4). Let uσ(x,t)u_{\sigma}(x,t) be the very weak solution of (CP) with u0=ϕσu_{0}=\phi_{\sigma} satisfying (Φ1)(\Phi_{1})-(Φ3)(\Phi_{3}). Then there exist 0<σ0<\sigma_{*}\leq\infty such that the following trichotomy result holds:

  • (i)

    spreading happens for σ>σ\sigma>\sigma_{*}:

    uσ(,t)1 as t, in the topology of Lloc().u_{\sigma}(\cdot,t)\to 1\mbox{\ \ as\ \ }t\to\infty,\ \ \mbox{ in the topology of }L^{\infty}_{loc}({\mathbb{R}}).
  • (ii)

    vanishing happens for 0<σ<σ0<\sigma<\sigma_{*}:

    uσ(,t)0 as t, in the topology of L().u_{\sigma}(\cdot,t)\to 0\mbox{\ \ as\ \ }t\to\infty,\ \ \mbox{ in the topology of }L^{\infty}({\mathbb{R}}).
  • (iii)

    transition case for σ=σ\sigma=\sigma_{*}:

    uσ(,t)θ as t, in the topology of Lloc().u_{\sigma}(\cdot,t)\to\theta\mbox{\ \ as\ \ }t\to\infty,\ \ \mbox{ in the topology of }L^{\infty}_{loc}({\mathbb{R}}).

    Moreover for large tt, uσu_{\sigma} has exactly two free boundaries l(t)<r(t)l(t)<r(t) satisfying

    (1.12) l(t),r(t)=2y0t[1+o(1)] as t,-l(t),r(t)=2y_{0}\sqrt{t}\;[1+o(1)]\mbox{\ \ as\ \ }t\to\infty,

    for some y0(0,θm12)y_{0}\in(0,\theta^{\frac{m-1}{2}}).

In the bistable case, if we further assume that f(0)<0f^{\prime}(0)<0, then the problem (CP) has Type II ground state solutions, and so we have the following trichotomy result.

Theorem 1.6.

Assume fC2f\in C^{2} is a bistable reaction term satisfying (1.3) and f(0)<0f^{\prime}(0)<0. Let uσ(x,t)u_{\sigma}(x,t) be the very weak solution of (CP) with u0=ϕσu_{0}=\phi_{\sigma} satisfying (Φ1)(\Phi_{1})-(Φ3)(\Phi_{3}). Then there exist 0<σσ0<\sigma_{*}\leq\sigma^{*}\leq\infty such that the following trichotomy holds:

  • (i)

    spreading happens for σ>σ\sigma>\sigma^{*}:

    uσ(,t)1 as t, in the topology of Lloc().u_{\sigma}(\cdot,t)\to 1\mbox{\ \ as\ \ }t\to\infty,\ \ \mbox{ in the topology of }L^{\infty}_{loc}({\mathbb{R}}).
  • (ii)

    vanishing happens for 0<σ<σ0<\sigma<\sigma_{*}:

    uσ(,t)0 as t, in the topology of L().u_{\sigma}(\cdot,t)\to 0\mbox{\ \ as\ \ }t\to\infty,\ \ \mbox{ in the topology of }L^{\infty}({\mathbb{R}}).
  • (iii)

    transition case for σσσ\sigma_{*}\leq\sigma\leq\sigma^{*}:

    uσ(,t)𝒰(x) as t, in the topology of L(),u_{\sigma}(\cdot,t)\to\mathcal{U}(x)\mbox{\ \ as\ \ }t\to\infty,\ \ \mbox{ in the topology of }L^{\infty}({\mathbb{R}}),

    where 𝒰(x)\mathcal{U}(x) is a Type II ground state solution defined as (1.8) for k(1)k\;(\geq 1) points z1,z2,,zk[b,b]z_{1},z_{2},\cdots,z_{k}\in[-b,b].

Remark 1.7.
  • (1)

    Under some additional conditions, we can show the sharpness of the transition solution: σ=σ\sigma_{*}=\sigma^{*} (see details in Lemma 6.3).

  • (2)

    The behaviour of a transition solution in Theorem 1.6 (iii) is quite different from that of the bistable RDE, where the transition solution converges to a Type I ground state solution which is positive on the whole {\mathbb{R}}.

  • (3)

    When spreading happens, the estimate on the propagation of the free boundaries has been shown in [18].

The main features of the paper are as follows:

  • (1)

    we give out a complete classification for the asymptotic behaviors of the solutions of (CP) with general nonlinear term;

  • (2)

    we establish the zero number argument for RPME with solid proof.

This paper is arranged as follows. In Section 2 we first present some results on the well-posedness and the a priori estimates, which are not entirely new and some details are postponed to Appendix. We also present sufficient conditions ensuring the waiting time of a free boundary to be 0, finite or infinite, specify the relationship between the stationary solutions of the RPME and those of the Cauchy problem, and derive the monotonicity and the Darcy law for the free boundaries. In Section 3, we study the monotonicity of the solution outside of the initial support, show the intersection number properties, and then use it, as well as the classical zero number argument, to prove the general convergence result. In Section 4, we consider the RPME with monostable reaction term and prove the hair-trigger effect. In Section 5 we consider the RPME with combustion reaction term: we present all possible nonnegative stationary solutions, prove a trichotomy result with a sharp transition for the asymptotic behavior. We also present a precise estimate for the propagation speed of the free boundaries of the transition solutions. In Section 6, we consider the RPME with bistable reaction term: we present all possible nonnegative stationary solutions, specify Type I and Type II ground state solutions, prove a trichotomy result with transition for the asymptotic behavior. In Section 7, we give a sufficient condition for the complete vanishing phenomena. Finally, for the convenience of the readers, we present in Appendix (Section 8) the a priori estimates, well-posedness and other important properties for PME with general reaction terms. They are not entirely new but not well collected in the literature.

We collect some notations and symbols which will be used in this paper.

  • For ψC()\psi\in C({\mathbb{R}}), we use spt[ψ()]{\rm spt}[\psi(\cdot)] to denote its support;

  • v(x)b(xx00)v(x)\to b\ (x\to x_{0}-0), or, limxx00v(x)=b\lim\limits_{x\to x_{0}-0}v(x)=b means the left limit limx<x0,xx0v(x)=b\lim\limits_{x<x_{0},\ x\to x_{0}}v(x)=b. The right one is denoted similarly;

  • Dxv(x0)=limxx00v(x)v(x0)xx0D^{-}_{x}v(x_{0})=\lim\limits_{x\to x_{0}-0}\frac{v(x)-v(x_{0})}{x-x_{0}} denotes the left derivative of vv at x0x_{0}. The right one Dx+v(x0)D^{+}_{x}v(x_{0}) is similar;

  • 0<1x10<1-x\ll 1 means that x<1x<1 and 1x1-x is sufficiently small;

  • a(x)b(x)a(x)\sim b(x) as xx0±0x\to x_{0}\pm 0 means that a(x)=b(x)(1+o(1))a(x)=b(x)(1+o(1)) as xx0±0x\to x_{0}\pm 0.

2. Preliminaries

In this section we first collect some basic results on the a priori estimates and the well-posedness. Then we specify the waiting time, stationary solutions, the regularity and the Darcy law of the free boundaries, which turn out not direct consequences of the corresponding results for PME.

2.1. Well-posedness and a priori estimates

As we have mentioned above, the existence and uniqueness of (CP) have been studied a lot by many authors (see, for example, [5, 25, 26, 30] etc.). In addition, by the assumption f(u)<0f(u)<0 for u>1u>1 in (F) and by the comparison principle, it is easily to obtain

(2.1) 0u(x,t)M0:=max{1,u0L()},x,t>0.0\leq u(x,t)\leq M^{\prime}_{0}:=\max\{1,\|u_{0}\|_{L^{\infty}({\mathbb{R}})}\},\quad x\in{\mathbb{R}},\ t>0.

The persistence of positivity, the finite spreading speed and the a priori estimates were also studied widely, especially for the PME without reactions/sources or RPMEs with polynomial or logistic reactions. For our equation with a general reaction term, however, these properties seem not conveniently collected in literature, though they are not entirely new. We collect some of them here and present some detailed proof in the appendix.

1. Positivity persistence. If u(x,t0)ε>0u(x,t_{0})\geq\varepsilon>0 in a neighborhood of x0x_{0}\in{\mathbb{R}}, then u(x0,t)>0u(x_{0},t)>0 for all tt0t\geq t_{0} (see Proposition 8.2). Thus, uu is classical near the line {(x0,t)t>t0})\{(x_{0},t)\mid t>t_{0}\}).

2. Finite propagation speed. The solution has a left-most free boundary l(t)l(t) and a right-most free boundary r(t)r(t) (also called interfaces) which propagate in finite speed: spt[u(,t)][l(t),r(t)][s¯(t),s¯(t)]{\rm spt}[u(\cdot,t)]\subset[l(t),r(t)]\subset[-\bar{s}(t),\bar{s}(t)] with

(2.2) s¯(t):=O(1)(t+1)1/2eK(m1)(t+1)2,t>0,\bar{s}(t):=O(1)(t+1)^{1/2}e^{\frac{K(m-1)(t+1)}{2}},\quad t>0,

(see Proposition 8.4).

In a gas flow problem through a porous medium, uu denotes the density of the gas, and

(2.3) v(x,t):=mm1[u(x,t)]m1v(x,t):=\frac{m}{m-1}[u(x,t)]^{m-1}

represents the pressure of the gas. In many cases, it is convenient to consider vv rather than uu. Using vv, (CP) is converted into the following problem

(pCP){vt=(m1)vvxx+vx2+g(v),x,t>0,v(x,0)=v0(x):=mm1u0m1(x),x,{\rm(pCP)}\hskip 85.35826pt\left\{\begin{array}[]{ll}v_{t}=(m-1)vv_{xx}+v_{x}^{2}+g(v),&x\in{\mathbb{R}},\ t>0,\\ \displaystyle v(x,0)=v_{0}(x):=\frac{m}{m-1}u_{0}^{m-1}(x),&x\in{\mathbb{R}},\end{array}\right.\hskip 142.26378pt

with

(2.4) g(v):=m((m1)vm)m2m1f(((m1)vm)1m1).g(v):=m\Big{(}\frac{(m-1)v}{m}\Big{)}^{\frac{m-2}{m-1}}f\left(\Big{(}\frac{(m-1)v}{m}\Big{)}^{\frac{1}{m-1}}\right).

For simplicity, we call the equation in (pCP) as pRPME. Since ff is Lipschitz with number KK we have

(2.5) |g(v)|K(m1)v,v0.|g(v)|\leq K(m-1)v,\quad v\geq 0.

Since uu is bounded as in (2.1), we see that vv is also bounded:

(2.6) 0v(x,t)M0:=max{mm1,v0L},x,t>0.0\leq v(x,t)\leq M_{0}:=\max\left\{\frac{m}{m-1},\|v_{0}\|_{L^{\infty}}\right\},\quad x\in{\mathbb{R}},\ t>0.

3. A priori estimates. We have a priori estimates for vx,vxxv_{x},v_{xx} and vtv_{t} as follows. If v>0v>0 in R:=(a,b)×(0,T]R:=(a,b)\times(0,T], then for any 0<δ<ba2,τ<T0<\delta<\frac{b-a}{2},\ \tau<T there holds:

(2.7) |vx(x,t)|M1(m,f,δ,τ),(x,t)[a+δ,bδ]×[τ,T].|v_{x}(x,t)|\leq M_{1}(m,f,\delta,\tau),\quad(x,t)\in[a+\delta,b-\delta]\times[\tau,T].

There exist τ0=τ0(m,v0)\tau_{0}=\tau_{0}(m,v_{0}) and C=C(m,v0)C=C(m,v_{0}) such that

vxx(x,t)C(t+τ0),x,t(0,T],v_{xx}(x,t)\geq-C(t+\tau_{0}),\quad x\in{\mathbb{R}},\ t\in(0,T],

in the sense of distributions in ×(0,T]{\mathbb{R}}\times(0,T]. For any τ>0\tau>0 there holds

(2.8) C3tC4vt(x,t)C1t+C2,x,tτ,-C_{3}t-C_{4}\leq v_{t}(x,t)\leq C_{1}t+C_{2},\quad x\in{\mathbb{R}},\ t\geq\tau,

where CiC_{i} depends on m,fm,f and v0v_{0}, while C2C_{2} depends also on vt(x,τ)v_{t}(x,\tau).

In addition, we have the following locally uniform Hölder estimates (cf. [10] and [30, Theorem 7.18]): for any τ>0\tau>0 and any compact domain D×(τ,)D\subset{\mathbb{R}}\times(\tau,\infty), there are positive constants CC and α\alpha, both depending only on M0,mM^{\prime}_{0},m and τ\tau, such that

(2.9) vCα(D)C.\|v\|_{C^{\alpha}(D)}\leq C.

Note that the Hölder estimates in [10, 30] are given for uu, but it is easy to convert to that for vv.

2.2. Waiting times of the free boundaries and compactly supported stationary solutions

In PME and RPMEs, the degeneracy of the equation not only causes finite propagation and free boundaries. It may even occur that a free boundary is stationary for a while if the mass distribution near the border of the support is very small.

Let x0x_{0}\in{\mathbb{R}} be a zero of u0(x)u_{0}(x). Define the waiting time at x0x_{0} by

t(x0):=inf{t>0u(x0,t)>0}.t^{*}(x_{0}):=\inf\{t>0\mid u(x_{0},t)>0\}.

For the one dimensional PME, it is known that t(x0)t^{*}(x_{0}) can be 0 or a positive number, depending on the decay rate of u0u_{0} near x0x_{0} (cf. [30, §15.3], [31]). For our equation with a source, the situation is more complicated since positive reactions can promote the propagation while the negative reactions will restrain it. We will show the following results.

  • (1).

    t(x0)=0t^{*}(x_{0})=0 when the pressure v0v_{0} near x0x_{0} is bounded from below by a linear function;

  • (2).

    t(x0)>0t^{*}(x_{0})>0 when the pressure v0v_{0} lies below a quadratic function near x0x_{0}. In addition t(x0)<t^{*}(x_{0})<\infty if the reaction term f0f\geq 0 near 0, or, if ff is negative but v0v_{0} lies above another quadratic function with big coefficient.

  • (3).

    t(x0)=t^{*}(x_{0})=\infty when the pressure v0v_{0} is less than a quadratic function near x0x_{0} and f(u)Kuf(u)\leq-K^{\prime}u for some large K>0K^{\prime}>0 and 0u10\leq u\ll 1.

For simplicity, we assume 0 is a zero of v0v_{0} and consider its waiting time.

1. The case t(0)=0t^{*}(0)=0.

Lemma 2.1.

Assume (F) and (I). Assume v0(0)=0v_{0}(0)=0 and

(2.10) v0(x)ρx,0xr0,v_{0}(x)\geq\rho x,\quad 0\leq x\leq r_{0},

for some ρ>0,r0>0\rho>0,\ r_{0}>0, then t(0)=0t^{*}(0)=0.

Proof.

Set

a:=πr0,h:=ρr0π,μ:=K(m1)+mha2,b(t):=ha2μ(eμt1),a:=\frac{\pi}{r_{0}},\quad h:=\frac{\rho r_{0}}{\pi},\quad\mu:=K(m-1)+mha^{2},\quad b(t):=\frac{ha^{2}}{\mu}(e^{-\mu t}-1),

and, for t0t\geq 0, define

v¯(x,t):={heμtsin(axb(t)),b(t)axb(t)+πa,0, otherwise.\underline{v}(x,t):=\left\{\begin{array}[]{ll}he^{-\mu t}\sin(ax-b(t)),&\ \ \ \displaystyle\frac{b(t)}{a}\leq x\leq\frac{b(t)+\pi}{a},\\ 0,&\ \ \ \mbox{ otherwise}.\end{array}\right.

By a direct calculation we have

v¯(x,0)v0(x),x,\underline{v}(x,0)\leq v_{0}(x),\quad x\in{\mathbb{R}},

and, with z:=axb(t)z:=ax-b(t),

𝒩(v¯)\displaystyle\mathcal{N}(\underline{v}) :=\displaystyle:= v¯t(m1)v¯v¯xxv¯x2g(v¯)\displaystyle\underline{v}_{t}-(m-1)\underline{v}\;\underline{v}_{xx}-\underline{v}_{x}^{2}-g(\underline{v})
\displaystyle\leq v¯t(m1)v¯v¯xxv¯x2+K(m1)v¯\displaystyle\underline{v}_{t}-(m-1)\underline{v}\;\underline{v}_{xx}-\underline{v}_{x}^{2}+K(m-1)\underline{v}
=\displaystyle= heμt[μsinz+(m1)a2heμtsin2z+K(m1)sinz\displaystyle he^{-\mu t}[-\mu\sin z+(m-1)a^{2}he^{-\mu t}\sin^{2}z+K(m-1)\sin z
b(t)cosza2heμtcos2z]\displaystyle-b^{\prime}(t)\cos z-a^{2}he^{-\mu t}\cos^{2}z]
\displaystyle\leq a2h2eμt[sinz+eμtcoszeμtcos2z]\displaystyle a^{2}h^{2}e^{-\mu t}[-\sin z+e^{-\mu t}\cos z-e^{-\mu t}\cos^{2}z]
\displaystyle\leq a2h2e2μt[sinz+coszcos2z]\displaystyle a^{2}h^{2}e^{-2\mu t}[-\sin z+\cos z-\cos^{2}z]
=\displaystyle= 12a2h2e2μt[(1cosz)2+2sinzsin2z]\displaystyle-\frac{1}{2}a^{2}h^{2}e^{-2\mu t}[(1-\cos z)^{2}+2\sin z-\sin^{2}z]
\displaystyle\leq 0,b(t)axb(t)+πa,t0.\displaystyle 0,\qquad\frac{b(t)}{a}\leq x\leq\frac{b(t)+\pi}{a},\ t\geq 0.

Therefore, v¯(x,t)\underline{v}(x,t) is a subsolution, and so by the comparison result (see, for example, [30, Theorem 6.5]) we have v(0,t)v¯(0,t)>0v(0,t)\geq\underline{v}(0,t)>0 for small t>0t>0. Then t(0)=0t^{*}(0)=0 since b(t)<0b^{\prime}(t)<0 for all t0t\geq 0. ∎

2. The case t(0)(0,)t^{*}(0)\in(0,\infty).

Lemma 2.2.

Assume (F) and (I). Assume v0(0)=0v_{0}(0)=0,

(2.11) v0(x)A2x2,0xr0,v_{0}(x)\geq A_{2}x^{2},\quad 0\leq x\leq r_{0},

for some r0>0r_{0}>0 and some A2A_{2} satisfying

(2.12) A2>A2(m,K):=K(m1)2σmm+12σ,σ:=(mm+1)m+1.A_{2}>A_{2}(m,K):=\frac{K(m-1)}{2\sigma^{\frac{m}{m+1}}-2\sigma},\quad\sigma:=\Big{(}\frac{m}{m+1}\Big{)}^{m+1}.

and

(2.13) v0(x)A1x2,x0,v_{0}(x)\leq A_{1}x^{2},\quad x\geq 0,

for some A1>A2A_{1}>A_{2}. Then t(0)(0,)t^{*}(0)\in(0,\infty).

Proof.

First we show that t(0)>0t^{*}(0)>0. Set

a:=(m1)K2(m+1)A1+(m1)K,T1:=a(m1)K,C1:=1a2(m+1),a:=\frac{(m-1)K}{2(m+1)A_{1}+(m-1)K},\quad T_{1}:=\frac{a}{(m-1)K},\quad C_{1}:=\frac{1-a}{2(m+1)},

and define

v¯(x,t):=C1x2T1t,x, 0t<T1.\bar{v}(x,t):=\frac{C_{1}x^{2}}{T_{1}-t},\quad x\in{\mathbb{R}},\ 0\leq t<T_{1}.

Then v¯(x,0)=A1x2v0(x)\bar{v}(x,0)=A_{1}x^{2}\geq v_{0}(x), and

𝒩(v¯)\displaystyle\mathcal{N}(\bar{v}) :=\displaystyle:= v¯t(m1)v¯v¯xxv¯x2g(v¯)\displaystyle\bar{v}_{t}-(m-1)\bar{v}\bar{v}_{xx}-\bar{v}_{x}^{2}-g(\bar{v})
\displaystyle\geq v¯t(m1)v¯v¯xxv¯x2K(m1)v¯\displaystyle\bar{v}_{t}-(m-1)\bar{v}\bar{v}_{xx}-\bar{v}_{x}^{2}-K(m-1)\bar{v}
=\displaystyle= C1x2(T1t)2[12(m+1)C1K(m1)(T1t)]\displaystyle\frac{C_{1}x^{2}}{(T_{1}-t)^{2}}[1-2(m+1)C_{1}-K(m-1)(T_{1}-t)]
\displaystyle\geq 0,x,t[0,T1).\displaystyle 0,\quad x\in{\mathbb{R}},\ t\in[0,T_{1}).

By comparison we have v(x,t)v¯(x,t)v(x,t)\leq\bar{v}(x,t) in t[0,T1)t\in[0,T_{1}). Hence, the waiting time of vv at 0 is not smaller than T1=T1(m,A1,K)T_{1}=T_{1}(m,A_{1},K).

Next we prove t(0)<t^{*}(0)<\infty. Set, for some δ(0,1)\delta\in(0,1),

τ:=σK(m21),C2:=δr022(m+1)τ12(m+1),x1:=[C2(1+2(m+1)A2τ)A2τmm+1]1/2.\tau:=\frac{\sigma}{K(m^{2}-1)},\quad C_{2}:=\frac{\delta r_{0}^{2}}{2(m+1)\tau^{\frac{1}{2(m+1)}}},\quad x_{1}:=\left[\frac{C_{2}(1+2(m+1)A_{2}\tau)}{A_{2}\tau^{\frac{m}{m+1}}}\right]^{1/2}.

Then by the choice of A2A_{2} we have

x12C2=1A2τmm+1+2(m+1)ττmm+1<2σmm+1σK(m1)τmm+1+2(m+1)ττmm+1=2[K(m21)]mm+1K(m1).\frac{x_{1}^{2}}{C_{2}}=\frac{1}{A_{2}\tau^{\frac{m}{m+1}}}+\frac{2(m+1)\tau}{\tau^{\frac{m}{m+1}}}\\ <2\frac{\sigma^{\frac{m}{m+1}}-\sigma}{K(m-1)\tau^{\frac{m}{m+1}}}+2\frac{(m+1)\tau}{\tau^{\frac{m}{m+1}}}\\ =2\frac{[K(m^{2}-1)]^{\frac{m}{m+1}}}{K(m-1)}.

Define

v¯(x,t):={C2(t+τ)mm+1(xx1)22(m+1)(t+τ),x[l¯(t),r¯(t)], 0t<T2,0,x<l¯(t) or x>r¯(t), 0t<T2,\underline{v}(x,t):=\left\{\begin{array}[]{ll}\displaystyle C_{2}(t+\tau)^{-\frac{m}{m+1}}-\frac{(x-x_{1})^{2}}{2(m+1)(t+\tau)},&x\in[\underline{l}(t),\underline{r}(t)],\ 0\leq t<T_{2},\\ 0,&x<\underline{l}(t)\mbox{ or }x>\underline{r}(t),\ 0\leq t<T_{2},\end{array}\right.

where

l¯(t):=x12(m+1)C2(t+τ)12(m+1),r¯(t):=x1+2(m+1)C2(t+τ)12(m+1)\underline{l}(t):=x_{1}-\sqrt{2(m+1)C_{2}}(t+\tau)^{\frac{1}{2(m+1)}},\quad\underline{r}(t):=x_{1}+\sqrt{2(m+1)C_{2}}(t+\tau)^{\frac{1}{2(m+1)}}

and

T2:=1σK(m21).T_{2}:=\frac{1-\sigma}{K(m^{2}-1)}.

We now show that v¯(x,t)\underline{v}(x,t) is a subsolution in 0t<T20\leq t<T_{2}. In fact, by our construction it is easy to verify that

A2x2v¯(x,0),x[l¯(0),r¯(0)][0,r0]A_{2}x^{2}\geq\underline{v}(x,0),\quad x\in[\underline{l}(0),\underline{r}(0)]\subset[0,r_{0}]

provided δ(0,1)\delta\in(0,1) is small, and the equality holds at

x=x^:=x12(m+1)τA2+1.x=\hat{x}:=\frac{x_{1}}{2(m+1)\tau A_{2}+1}.

In addition, for x[l¯(t),r¯(t)], 0tT2x\in[\underline{l}(t),\underline{r}(t)],\ 0\leq t\leq T_{2}, by a direct calculation we have

𝒩v¯\displaystyle\mathcal{N}\underline{v} :=\displaystyle:= v¯t(m1)v¯v¯xxv¯x2g(v¯)\displaystyle\underline{v}_{t}-(m-1)\underline{v}\underline{v}_{xx}-\underline{v}_{x}^{2}-g(\underline{v})
\displaystyle\leq v¯t(m1)v¯v¯xxv¯x2+K(m1)v¯\displaystyle\underline{v}_{t}-(m-1)\underline{v}\underline{v}_{xx}-\underline{v}_{x}^{2}+K(m-1)\underline{v}
\displaystyle\leq C2(t+τ)2m+1m+1[1m+1+K(m1)(t+τ)]0.\displaystyle\frac{C_{2}}{(t+\tau)^{\frac{2m+1}{m+1}}}\Big{[}\frac{-1}{m+1}+K(m-1)(t+\tau)\Big{]}\leq 0.

This implies that v¯\underline{v} is a subsolution at leat in the time interval [0,T2][0,T_{2}].

By the choice of the above parameters, we have

l¯(0)2C2=(1+2(m+1)A2τA2τmm+1)1/2[2(m+1)τ1m+1]1/2>0,\frac{\underline{l}(0)}{\sqrt{2C_{2}}}=\left(\frac{1+2(m+1)A_{2}\tau}{A_{2}\tau^{\frac{m}{m+1}}}\right)^{1/2}-\left[2(m+1)\tau^{\frac{1}{m+1}}\right]^{1/2}>0,

and

l¯(T2)2C2\displaystyle\frac{\underline{l}(T_{2})}{\sqrt{2C_{2}}} <\displaystyle< ([K(m21)]mm+1K(m1))1/2[(m+1)(T2+τ)1m+1]1/2\displaystyle\left(\frac{[K(m^{2}-1)]^{\frac{m}{m+1}}}{K(m-1)}\right)^{1/2}-\left[(m+1)(T_{2}+\tau)^{\frac{1}{m+1}}\right]^{1/2}
=\displaystyle= ([K(m21)]mm+1K(m1))1/2(m+1[K(m21)]1m+1)1/2=0.\displaystyle\left(\frac{[K(m^{2}-1)]^{\frac{m}{m+1}}}{K(m-1)}\right)^{1/2}-\left(\frac{m+1}{[K(m^{2}-1)]^{\frac{1}{m+1}}}\right)^{1/2}=0.

This means that before t=T2t=T_{2}, the left free boundary l¯(t)\underline{l}(t) of v¯(x,t)\underline{v}(x,t) moves leftward from a positive point to a negative one. Then, v¯(0,t)\underline{v}(0,t) becomes positive at T2T_{2}, and so does v(0,t)v(0,t). This proves our lemma. ∎

The case t(0)>0t^{*}(0)>0 still includes two subcases: t(0)(0,)t^{*}(0)\in(0,\infty) and t(0)=t^{*}(0)=\infty. In the PME, it is known that only the former subcase is possible, which is called the hole-filling phenomena (see, for example, [30, Theorem 15.15]). For our RPME with a reaction term, however, the latter is also possible.

3. The case where t(0)=t^{*}(0)=\infty and compactly supported stationary solutions. To discuss the possibility of infinite waiting time, we first specify the compactly supported stationary solutions, including its sufficient and necessary conditions, examples and uniqueness. We remark that the following two concepts should be distinguished:

compactly supported stationary solutions of the equation pRPME,  and
compactly supported stationary solutions of the Cauchy problem
(pCP).

The latter means all compactly supported functions v¯\bar{v} such that the solution v(x,t;v¯)v(x,t;\bar{v}) of (pCP) with initial data v0=v¯v_{0}=\bar{v} satisfies v(x,t;v¯)v¯v(x,t;\bar{v})\equiv\bar{v} for all xx\in{\mathbb{R}} and t0t\geq 0.

Lemma 2.3.

Assume v¯\bar{v} is a continuous nonnegative stationary solution of pRPME with support [l,r][l,r] and v¯(x)>0\bar{v}(x)>0 in (l,r)(l,r). Then

  1. (i).

    v¯\bar{v} is also a stationary solution of (pCP) if and only if t(l),t(r)τt^{*}(l),\ t^{*}(r)\geq\tau for some τ>0\tau>0;

  2. (ii).

    a sufficient condition for (i) is v¯(x)A(xl)2\bar{v}(x)\leq A(x-l)^{2} and v¯(x)A(xr)2\bar{v}(x)\leq A(x-r)^{2} in (l,r)(l,r) for some A>0A>0; a necessary condition for (i) is v¯(l)=v¯(r)=0\bar{v}^{\prime}(l)=\bar{v}^{\prime}(r)=0;

  3. (iii).

    v¯\bar{v} is not a stationary solution of (pCP) if v¯(l)>0\bar{v}^{\prime}(l)>0 or v¯(r)<0\bar{v}^{\prime}(r)<0;

  4. (iv).

    in any case, v¯\bar{v} is a time-independent (also called a stationary) very weak subsolution of (pCP), and v(x,t;v¯)v¯(x)v(x,t;\bar{v})\geq\bar{v}(x) for x,t>0x\in{\mathbb{R}},\ t>0. The inequality is strict on [l,r][l,r] when v¯\bar{v} is not a stationary solution of (pCP).

Proof.

(i). If t(l),t(r)τt^{*}(l),\ t^{*}(r)\geq\tau, then for all t[0,τ]t\in[0,\tau] we have

v(x,t;v¯)=v¯(x),l(t)=l,r(t)=r.v(x,t;\bar{v})=\bar{v}\ \ (x\in{\mathbb{R}}),\quad l(t)=l,\quad r(t)=r.

Therefore, these equalities hold for all t0t\geq 0, and so v¯\bar{v} is a stationary solution of (pCP).

(ii). The sufficient condition follows from Lemma 2.2, and the necessary condition follows from Lemma 2.1.

(iii). The conclusion is a consequence of (ii).

(iv). Denote the corresponding density functions of v¯(x)\bar{v}(x) and v(x,t;v¯)v(x,t;\bar{v}) by u¯(x)\bar{u}(x) and u(x,t;u¯)u(x,t;\bar{u}), respectively. For any T>0T>0 and any test function φCc(QT)\varphi\in C^{\infty}_{c}(Q_{T}) (QT:=×(0,T)Q_{T}:={\mathbb{R}}\times(0,T)) with φ0\varphi\geq 0,

QT[u¯φt+u¯mφxx]𝑑x𝑑t\displaystyle\iint_{Q_{T}}[\bar{u}\varphi_{t}+\bar{u}^{m}\varphi_{xx}]dxdt =\displaystyle= u¯(x)[φ(x,T)φ(x,0)]𝑑x+0Tlr(u¯m)xx(x)φ(x,t)𝑑x𝑑t\displaystyle\int_{{\mathbb{R}}}\bar{u}(x)[\varphi(x,T)-\varphi(x,0)]dx+\int_{0}^{T}\int_{l}^{r}(\bar{u}^{m})_{xx}(x)\varphi(x,t)dxdt
+0T[(u¯m)x(l)φ(l,t)(u¯m)x(r)φ(r,t)]𝑑t\displaystyle+\int_{0}^{T}[(\bar{u}^{m})_{x}(l)\varphi(l,t)-(\bar{u}^{m})_{x}(r)\varphi(r,t)]dt
\displaystyle\geq u¯(x)[φ(x,T)φ(x,0)]𝑑x+0Tlr(u¯m)xx(x)φ(x,t)𝑑x𝑑t\displaystyle\int_{{\mathbb{R}}}\bar{u}(x)[\varphi(x,T)-\varphi(x,0)]dx+\int_{0}^{T}\int_{l}^{r}(\bar{u}^{m})_{xx}(x)\varphi(x,t)dxdt
=\displaystyle= u¯(x)[φ(x,T)φ(x,0)]𝑑xQTf(u¯(x))φ(x,t)𝑑x𝑑t.\displaystyle\int_{{\mathbb{R}}}\bar{u}(x)[\varphi(x,T)-\varphi(x,0)]dx-\iint_{Q_{T}}f(\bar{u}(x))\varphi(x,t)dxdt.

So u¯\bar{u} is a time-independent very weak solution of (CP), and then u(x,t;u¯)u¯(x)u(x,t;\bar{u})\geq\bar{u}(x) follows from the comparison result (cf. [30, Theorem 6.5]).

To show the strict inequality, we note that, in the domain (l,r)(l,r), both v¯\bar{v} and vv are positive, and so the classical strong maximum principle is applied. On the other hand, by the conclusion in (i), the waiting times at ll and rr are zero since v¯\bar{v} is not a stationary solution of (pCP). So v(l,t;v¯)>0=v¯(l)v(l,t;\bar{v})>0=\bar{v}(l) and v(r,t;v¯)>0=v¯(r)v(r,t;\bar{v})>0=\bar{v}(r). This proves the strict inequality on [l,r][l,r]. ∎

Due to the strong maximum principle, the Cauchy problem of a RDE does not have compactly supported stationary solutions. For (pCP), however, such solutions are possible since the strong maximum principle is not true on the free boundaries. In fact, we will show in Section 5 (see details in Lemma 6.1) that, if ff is a bistable nonlinearity with

f(u)=λuα(1+o(1)) as u0+0,f(u)=-\lambda u^{\alpha}(1+o(1))\mbox{\ \ as\ \ }u\to 0+0,

for some λ>0\lambda>0 and 1α<m1\leq\alpha<m, then the bistable pRPME has nonnegative compactly supported stationary solution v¯\bar{v}, which is a ground state solution satisfying

v¯(x)>0 in (L,L),v¯(x)=0 for |x|L,\bar{v}(x)>0\mbox{ in }(-L,L),\quad\bar{v}(x)=0\mbox{ for }|x|\geq L,

and

v¯(x)A(x±L)2(m1)mα as xL±0.\bar{v}(x)\sim A(x\pm L)^{\frac{2(m-1)}{m-\alpha}}\mbox{\ \ as\ \ }x\to\mp L\pm 0.

Since 2(m1)mα2\frac{2(m-1)}{m-\alpha}\geq 2, by Lemma 2.3 (ii) we know that v¯\bar{v} is also a stationary solution of (pCP).

Comparing the ZKB solution of PME (named after Zel’dovich, Kompaneets and Barenblatt) with the solution of (pCP) for monostable and combustion reactions, it is easy to see that compactly supported stationary solutions do not exist in such problems. But they can exist for (pCP) with bistable reactions (see Lemma 6.1). We now show that it is unique when it exists.

Lemma 2.4.

Assume (F) and VV is a compactly supported stationary solution of (pCP) with

V(x)>0 in (L,L),V(x)=0 for |x|L,V(x)>0\mbox{ in }(-L,L),\quad V(x)=0\mbox{ for }|x|\geq L,

for some L>0L>0. Then

(2.14) V(x)=V(x) and V(x)<0 in (0,L),V(0)=V(±L)=0.\quad V(x)=V(-x)\mbox{ and }V^{\prime}(x)<0\mbox{ in }(0,L),\quad V^{\prime}(0)=V^{\prime}(\pm L)=0.

(So, VV is a Type II ground state solution of (pCP).) Moreover, such compactly supported stationary solution is unique.

Proof.

By the equation we know that VV is symmetric with respect to each positive critical points, so we have the symmetry and monotonicity in (2.14) and V(0)=0V^{\prime}(0)=0. The equality V(±L)=0V^{\prime}(\pm L)=0 follow from Lemma 2.3.

We now prove the uniqueness of such compactly supported stationary solution. Set

(2.15) w:=(m1mV)1m1V.w:=\Big{(}\frac{m-1}{m}V\Big{)}^{\frac{1}{m-1}}V^{\prime}.

Then w(±L)=0w(\pm L)=0. Hence, on the (V,w)(V,w)-phase plane (cf. Subsections 4.1 and 5.1 for more details), V(x)V(x) corresponds to a homoclinic orbit starting and ending at (0,0)(0,0). Using the equation in (pCP) we have

w=f((m1mV)1m1).w^{\prime}=-f\left(\Big{(}\frac{m-1}{m}V\Big{)}^{\frac{1}{m-1}}\right).

Multiplying this equation by (2.15) and integrating it over [x,0][x,0] for x[L,0)x\in[-L,0), we have

w2(x)=F(V(x))F(V(0))w^{2}(x)=F(V(x))-F(V(0))

with

F(v):=20vf((m1mr)1m1)(m1mr)1m1𝑑r.F(v):=-2\int_{0}^{v}f\left(\Big{(}\frac{m-1}{m}r\Big{)}^{\frac{1}{m-1}}\right)\cdot\Big{(}\frac{m-1}{m}r\Big{)}^{\frac{1}{m-1}}dr.

Taking x=Lx=-L we have w2(L)=0=F(0)F(V(0))w^{2}(-L)=0=F(0)-F(V(0)), and so F(V(0))=0F(V(0))=0. For x(L,0)x\in(-L,0) we have V(x)>0V^{\prime}(x)>0 and w(x)>0w(x)>0. Thus,

F(v)>F(V(0))=0,v(0,V(0)).F(v)>F(V(0))=0,\quad v\in(0,V(0)).

(This happens only if f(u)<0f(u)<0 for 0<u10<u\ll 1.) Therefore,

(2.16) Θ1:=V(0)\Theta_{1}:=V(0)

is the smallest positive zero of F()F(\cdot), and there exists Θ(0,Θ1)\Theta\in(0,\Theta_{1}) such that

(2.17) f((m1mΘ)1m1)=g(Θ)=0 and f((m1mv)1m1)>0 for v(Θ,Θ1).f\left(\Big{(}\frac{m-1}{m}\Theta\Big{)}^{\frac{1}{m-1}}\right)=g(\Theta)=0\mbox{ and }f\left(\Big{(}\frac{m-1}{m}v\Big{)}^{\frac{1}{m-1}}\right)>0\mbox{ for }v\in(\Theta,\Theta_{1}).

Consequently, for x<0x<0, V(x)V(x) is uniquely determined by the problem

(m1mV)1m1V=F(V(x)) for x<0,V(0)=Θ1,V(0)=0.\Big{(}\frac{m-1}{m}V\Big{)}^{\frac{1}{m-1}}V^{\prime}=\sqrt{F(V(x))}\mbox{\ for\ }x<0,\qquad V(0)=\Theta_{1},\quad V^{\prime}(0)=0.

This proves the uniqueness of compactly supported stationary solution of (pCP), that is, the Type II ground state solution with one connected support. ∎

We have given sufficient and necessary conditions, and uniqueness for compactly supported stationary solutions of the Cauchy problem (pCP). Clearly, they correspond to the analogues of the original problem (CP). Using them we finally give some examples of initial data, which are not necessarily to be stationary solutions, such that the waiting times on their boundaries are infinite. For example, for a ground state solution VV of the bistable pRPME in the previous lemma, we choose v0(x)C()v_{0}(x)\in C({\mathbb{R}}) such that

v0(x)V(x) in ,v0(x)>0 in (L,L).v_{0}(x)\leq V(x)\mbox{ in }{\mathbb{R}},\quad v_{0}(x)>0\mbox{ in }(-L,L).

Then v0v_{0} lies below VV and has the same initial boundaries ±L\pm L as VV. Since the boundaries l(t)l(t) and r(t)r(t) of v(x,t;v0)v(x,t;v_{0}) are monotone in time and since vv always lies below VV by comparison, we have t(±L)=t^{*}({\pm L})=\infty.

2.3. Lipschitz continuity, Darcy law and strict monotonicity

We now address the Lipschitz continuity, the Darcy law and the monotonicity for each free boundary. For definiteness, in this subsection, we work with the right free boundary x=r(t)x=r(t), that is,

v(x,t)>0 for r(t)1x<r(t),v(x,t)=0 for xr(t).v(x,t)>0\mbox{ for }r(t)-1\ll x<r(t),\quad v(x,t)=0\mbox{ for }x\geq r(t).

We use one-sided partial derivatives in space and time:

D±r(t):=r(t±0),Dx±v(x,t)=vx(x±0,t).D^{\pm}r(t):=r^{\prime}(t\pm 0),\quad D^{\pm}_{x}v(x,t)=v_{x}(x\pm 0,t).

Also we use r(t)r^{\prime}(t) to denote the derivative of rr when D+r(t)=Dr(t)D^{+}r(t)=D^{-}r(t).

First we recall the related results on PME.

Lemma 2.5 ([30, Chapter 15]).

Let vv be the global solution of (pCP) with g0g\equiv 0, r(t)r(t) be its right free boundary with waiting time t0t^{*}\geq 0. Then, for any t0>tt_{0}>t^{*}, there exists a small ε>0\varepsilon>0 such that

  1. (i).

    rC((t0ε,t0+ε))r\in C^{\infty}((t_{0}-\varepsilon,t_{0}+\varepsilon)), vC(𝒩ε)v\in C^{\infty}(\mathcal{N}_{\varepsilon}) for

    𝒩ε={(x,t)xr(t) and |(xx0,tt0)|ε},x0:=r(t0);\mathcal{N}_{\varepsilon}=\{(x,t)\mid x\leq r(t)\mbox{ and }|(x-x_{0},t-t_{0})|\leq\varepsilon\},\quad x_{0}:=r(t_{0});
  2. (ii).

    there holds

    r(t0)=Dxv(x0,t0)>0,r′′(t)=mr(t)Dxxv(x0,t0);r^{\prime}(t_{0})=-D^{-}_{x}v(x_{0},t_{0})>0,\quad r^{\prime\prime}(t)=mr^{\prime}(t)D^{-}_{xx}v(x_{0},t_{0});
  3. (iii).

    for any positive integer jj, there exists CjC_{j} depending on (x0,t0),m,j,ε,v(x_{0},t_{0}),m,j,\varepsilon,v such that

    |jxjv(x,t)|Cj,(x,t)𝒩ε.\left|\frac{\partial^{j}}{\partial x^{j}}v(x,t)\right|\leq C_{j},\quad(x,t)\in\mathcal{N}_{\varepsilon}.

We will use this lemma to prove the important properties for the free boundaries, that is, the C1C^{1} regularity and the Darcy law. Since we have no semiconvexity for the free boundary r(t)r(t) of RPME as for the PME (cf. [30, §15.4.1]), we have to deal with the problem in a more complicated way.

We first show the left side continuity of vxv_{x}:

Lemma 2.6.

Let vv be the solution of (pCP), r(t)r(t) be its right free boundary. Then

(2.18) Dxv(r(t),t)=limxr(t)0vx(x,t),t>0.D^{-}_{x}v(r(t),t)=\lim\limits_{x\to r(t)-0}v_{x}(x,t),\quad t>0.
Proof.

As we mentioned before, vxxv_{xx} is bounded from below: vxxC1tC2v_{xx}\geq-C_{1}t-C_{2} for some C1,C2>0C_{1},C_{2}>0. Hence

v~(x,t):=v(x,t)+12(C1t+C2)x2\tilde{v}(x,t):=v(x,t)+\frac{1}{2}(C_{1}t+C_{2})x^{2}

is a continuous and convex function of xx for t>0t>0. So, for every x1,t1>0x_{1}\in{\mathbb{R}},\ t_{1}>0, the function v~(,t1)\tilde{v}(\cdot,t_{1}) admits one-sided derivatives at x1x_{1}: Dx+v~(x1,t1)D^{+}_{x}\tilde{v}(x_{1},t_{1}) and Dxv~(x1,t1)D^{-}_{x}\tilde{v}(x_{1},t_{1}) with Dxv~(x1,t1)Dx+v~(x1,t1)D^{-}_{x}\tilde{v}(x_{1},t_{1})\leq D^{+}_{x}\tilde{v}(x_{1},t_{1}), and the functions Dx±v~(,t1)D^{\pm}_{x}\tilde{v}(\cdot,t_{1}) are non-decreasing ones. Consequently, Dx±v(x1,t1)D^{\pm}_{x}v(x_{1},t_{1}) exist with Dxv(x1,t1)Dx+v(x1,t1)D^{-}_{x}v(x_{1},t_{1})\leq D^{+}_{x}v(x_{1},t_{1}) and Dx±v(x,t1)+(C1t1+C2)xD^{\pm}_{x}v(x,t_{1})+(C_{1}t_{1}+C_{2})x are non-decreasing in xx. Note that, for small ε>0\varepsilon>0, when xI1:=[r(t1)ε,r(t1))x\in I_{1}:=[r(t_{1})-\varepsilon,r(t_{1})), v(x,t1)v(x,t_{1}) is smooth and so

vx(x,t1)+(C1t1+C2)xDx±v(r(t1),t1)+(C1t1+C2)r(t1).v_{x}(x,t_{1})+(C_{1}t_{1}+C_{2})x\leq D^{\pm}_{x}v(r(t_{1}),t_{1})+(C_{1}t_{1}+C_{2})r(t_{1}).

The right hand side is bounded since Dxv(r(t1),t1)0=Dx+v(r(t1),t1)D^{-}_{x}v(r(t_{1}),t_{1})\leq 0=D^{+}_{x}v(r(t_{1}),t_{1}). So k1:=limxr(t1)0vx(x,t1)k_{1}:=\lim\limits_{x\to r(t_{1})-0}v_{x}(x,t_{1}) exists and k1Dxv(r(t1),t1)k_{1}\leq D^{-}_{x}v(r(t_{1}),t_{1}). The conclusion (2.18) is proved if we can show the equality. By contradiction, we assume that k1<Dxv(r(t1),t1)0k_{1}<D^{-}_{x}v(r(t_{1}),t_{1})\leq 0. Then there exists k2k_{2} lying between them and

vx(x,t1)<k2,xI1.v_{x}(x,t_{1})<k_{2},\qquad x\in I_{1}.

For any x2,x3x_{2},x_{3} satisfying r(t1)εx3<x2<r(t1)r(t_{1})-\varepsilon\ll x_{3}<x_{2}<r(t_{1}) we have

v(x3,t1)v(x2,t1)x3x2<k2.\frac{v(x_{3},t_{1})-v(x_{2},t_{1})}{x_{3}-x_{2}}<k_{2}.

Taking the limit as x2r(t1)0x_{2}\to r(t_{1})-0 and using the continuity of vv we obtain

v(x3,t1)v(r(t1),t1)x3r(t1)k2.\frac{v(x_{3},t_{1})-v(r(t_{1}),t_{1})}{x_{3}-r(t_{1})}\leq k_{2}.

This implies that Dxv(r(t1),t1)k2D^{-}_{x}v(r(t_{1}),t_{1})\leq k_{2}, a contradiction. ∎

Theorem 2.7.

Let vv be the solution of (pCP), r(t)r(t) be its right free boundary with waiting time t[0,)t^{*}\in[0,\infty). Then rC1((t,))r\in C^{1}((t^{*},\infty)) and the Darcy law holds:

(2.19) r(t)=Dxv(r(t),t)0,t>t.r^{\prime}(t)=-D^{-}_{x}v(r(t),t)\geq 0,\quad t>t^{*}.

Moreover, r(t)0r^{\prime}(t)\not\equiv 0 in any interval in (t,)(t^{*},\infty).

Proof.

We prove the Darcy law at any given time t0>tt_{0}>t^{*}. Denote x0:=r(t0)x_{0}:=r(t_{0}) and β:=Dxv(x0,t0)\beta:=-D^{-}_{x}v(x_{0},t_{0}).

Step 1. Assume β>0\beta>0 and to show D+r(t0)=βD^{+}r(t_{0})=\beta. Consider the following problem

{v¯t=(m1)v¯v¯xx+v¯x2+K(m1)v¯,x,t>0,v¯(x,0)=v¯0(x),x,\left\{\begin{array}[]{ll}\bar{v}_{t}=(m-1)\bar{v}\bar{v}_{xx}+\bar{v}_{x}^{2}+K(m-1)\bar{v},&x\in{\mathbb{R}},\ t>0,\\ \bar{v}(x,0)=\bar{v}_{0}(x),&x\in{\mathbb{R}},\end{array}\right.

where v¯0(x)\bar{v}_{0}(x) satisfies

v¯0(x00)=β,v¯0(x)v(x,t0) for x,\bar{v}^{\prime}_{0}(x_{0}-0)=-\beta,\quad\bar{v}_{0}(x)\geq v(x,t_{0})\mbox{ for }x\in{\mathbb{R}},

and v¯0(x)\bar{v}_{0}(x) is sufficiently smooth so that the right free boundary r¯(t)\bar{r}(t) of v¯(,t)\bar{v}(\cdot,t) is smooth in t0t\geq 0. Then v¯\bar{v} is a supersolution of (pCP), and so

r(t+t0)r¯(t),t0.r(t+t_{0})\leq\bar{r}(t),\quad t\geq 0.

To estimate r¯(t)\bar{r}(t) precisely, we change the time variable from tt to ss by

s=S(t):=eK(m1)t1K(m1)t=T(s):=ln(1+K(m1)s)K(m1),s=S(t):=\frac{e^{K(m-1)t}-1}{K(m-1)}\Leftrightarrow t=T(s):=\frac{\ln(1+K(m-1)s)}{K(m-1)},

and define

V(x,s):=v¯(x,T(s))1+K(m1)s,V(x,s):=\frac{\bar{v}(x,T(s))}{1+K(m-1)s},

then we have

(2.20) {Vs=(m1)VVxx+Vx2,x,s>0,V(x,0)=v¯0(x),x.\left\{\begin{array}[]{ll}V_{s}=(m-1)VV_{xx}+V_{x}^{2},&x\in{\mathbb{R}},\ s>0,\\ V(x,0)=\bar{v}_{0}(x),&x\in{\mathbb{R}}.\end{array}\right.

By Lemma 2.5, for some small ε0>0\varepsilon_{0}>0, the right free boundary x=R(s)=r¯(T(s))x=R(s)=\bar{r}(T(s)) of V(,s)V(\cdot,s) satisfies

R(s)=DxV(R(s),s),|R′′(s)|C¯1,0sS(ε0),R^{\prime}(s)=-D^{-}_{x}V(R(s),s),\quad|R^{\prime\prime}(s)|\leq\bar{C}_{1},\quad 0\leq s\leq S(\varepsilon_{0}),

or, equivalently,

r¯(t)=R(s)dsdt=Dxv¯(r¯(t),t),|r¯′′(t)|C¯,0tε0.\bar{r}^{\prime}(t)=R^{\prime}(s)\frac{ds}{dt}=-D^{-}_{x}\bar{v}(\bar{r}(t),t),\quad|\bar{r}^{\prime\prime}(t)|\leq\bar{C},\quad 0\leq t\leq\varepsilon_{0}.

Consequently, we have upper estimate for r(t+t0)r(t+t_{0}):

(2.21) r(t+t0)r¯(t)r¯(0)+r¯(0)t+C¯t2x0+βt+C¯t2,0tε0.r(t+t_{0})\leq\bar{r}(t)\leq\bar{r}(0)+\bar{r}^{\prime}(0)t+\bar{C}t^{2}\leq x_{0}+\beta t+\bar{C}t^{2},\quad 0\leq t\leq\varepsilon_{0}.

Next we consider the problem

{v¯t=(m1)v¯v¯xx+v¯x2K(m1)v¯,x,t>0,v¯(x,0)=v¯0(x),x,\left\{\begin{array}[]{ll}\underline{v}_{t}=(m-1)\underline{v}\underline{v}_{xx}+\underline{v}_{x}^{2}-K(m-1)\underline{v},&x\in{\mathbb{R}},\ t>0,\\ \underline{v}(x,0)=\underline{v}_{0}(x),&x\in{\mathbb{R}},\end{array}\right.

where v¯0(x)\underline{v}_{0}(x) satisfies

v¯0(x00)=β,v¯0(x)v(x,t0) for x,\underline{v}^{\prime}_{0}(x_{0}-0)=-\beta,\quad\underline{v}_{0}(x)\leq v(x,t_{0})\mbox{ for }x\in{\mathbb{R}},

and v¯0(x)\underline{v}_{0}(x) is sufficiently smooth so that the right free boundary r¯(t)\underline{r}(t) of v¯(,t)\underline{v}(\cdot,t) is smooth in t0t\geq 0. Then v¯\bar{v} is a subsolution of (pCP) and, as above,

(2.22) r(t+t0)r¯(t)x0+βtC¯t2,0tε0,r(t+t_{0})\geq\underline{r}(t)\geq x_{0}+\beta t-\underline{C}t^{2},\quad 0\leq t\leq\varepsilon_{0},

for some positive C¯>0\underline{C}>0. Combining with (2.21) we have D+r(t0)=βD^{+}r(t_{0})=\beta.

Step 2. To show D+r(t)0D^{+}r(t)\not\equiv 0 in any interval in (t,)(t^{*},\infty) with positive measure. Note that this conclusion implies that r(t)r(t) is strictly increasing in t>tt>t^{*}. By the positivity persistence, r(t)r(t) is increasing in t>tt>t^{*}. Assume by contradiction that D+r(t)=0D^{+}r(t)=0 for t[t1,t2](t,)t\in[t_{1},t_{2}]\subset(t^{*},\infty). We will derive

(2.23) D+r(t)=0,t[0,t2],D^{+}r(t)=0,\quad t\in[0,t_{2}],

which, then, contradicts the definition of tt^{*} and the fact t2>tt_{2}>t^{*}.

To show (2.23), we suppose, on the contrary, that t3:=inf{t~D+r(t)0 in [t~,t2]}(0,t1]t_{3}:=\inf\{\tilde{t}\mid D^{+}r(t)\equiv 0\mbox{ in }[\tilde{t},t_{2}]\}\in(0,t_{1}], and, by the continuity, there exists τ(t,t3)\tau\in(t^{*},t_{3}) such that

(2.24) D+r(τ)>0 and r(t3)r(τ)<Δ:=Pπμ(1eμ(t2t1)),D^{+}r(\tau)>0\mbox{\ \ and\ \ }r(t_{3})-r(\tau)<\Delta:=\frac{P\pi}{\mu}(1-e^{-\mu(t_{2}-t_{1})}),

where P-P is the lower bound of vxxv_{xx}: vxx>Pv_{xx}>-P for x,t[t,t2]x\in{\mathbb{R}},\ t\in[t^{*},t_{2}], and μ:=K(m1)+mPπ\mu:=K(m-1)+mP\pi. By Step 1 we have

0<2ρ:=D+r(τ)=Dxv(r(τ),τ).0<2\rho:=D^{+}r(\tau)=-D^{-}_{x}v(r(\tau),\tau).

Hence, for ε:=ρP\varepsilon:=\frac{\rho}{P} we have

Dxv(x,τ)ρ and v(x,τ)ρ(xr(τ)),r(τ)εxr(τ).-D^{-}_{x}v(x,\tau)\geq\rho\mbox{\ \ and\ \ }v(x,\tau)\geq-\rho(x-r(\tau)),\quad r(\tau)-\varepsilon\leq x\leq r(\tau).

Denote

a:=Pπρ,h:=ρ2Pπ,b(t):=Pπμ(1eμ(tτ))a:=\frac{P\pi}{\rho},\quad h:=\frac{\rho^{2}}{P\pi},\quad b(t):=\frac{P\pi}{\mu}(1-e^{-\mu(t-\tau)})

and using the subsolution v¯\underline{v} as in Lemma 2.1 we conclude that

(2.25) r(t)r(τ)+b(t),t>τ.r(t)\geq r(\tau)+b(t),\quad t>\tau.

In particular, at t=t3t=t_{3} we have

r(t3)=r(t2)r(τ)+b(t2)r(τ)+Δ,r(t_{3})=r(t_{2})\geq r(\tau)+b(t_{2})\geq r(\tau)+\Delta,

contradict (2.24). This proves (2.23), and then the conclusion in this step is proved.

Step 3. Assume β>0\beta>0 and to show Dr(t0)=βD^{-}r(t_{0})=\beta. We assume that, for some increasing sequence {tn}\{t_{n}\} tending to t0t_{0},

(2.26) β:=limnr(tn)r(t0)tnt0,D+r(tn)=βn:=Dxv(r(tn),tn)>0.\beta^{\prime}:=\lim\limits_{n\to\infty}\frac{r(t_{n})-r(t_{0})}{t_{n}-t_{0}},\quad D^{+}r(t_{n})=\beta_{n}:=-D^{-}_{x}v(r(t_{n}),t_{n})>0.

(The latter holds by Step 2.) If we can show that, for each of such time sequence {tn}\{t_{n}\}, the limit β=β\beta^{\prime}=\beta. Then we obtain the conclusion Dr(t0)=βD^{-}r(t_{0})=\beta.

By contradiction, we assume β>β\beta^{\prime}>\beta. The reversed case is proved similarly. Then there exist a small δ>0\delta>0 and a large n0n_{0} such that

(2.27) r(tn)r(t0)tnt0β+2δ,nn0.\frac{r(t_{n})-r(t_{0})}{t_{n}-t_{0}}\geq\beta+2\delta,\quad n\geq n_{0}.

For any given nn0n\geq n_{0}, using the estimates (2.21) and (2.22) at the time tnt_{n} instead of t0t_{0} we see that, for some ε1(0,ε0)\varepsilon_{1}\in(0,\varepsilon_{0}) and some positive numbers C¯,C¯\underline{C}^{\prime},\ \bar{C}^{\prime}, there holds

(2.28) r(tn)+βntC¯t2r(t+tn)r(tn)+βnt+C¯t2,0tε1.r(t_{n})+\beta_{n}t-\underline{C}^{\prime}t^{2}\leq r(t+t_{n})\leq r(t_{n})+\beta_{n}t+\bar{C}^{\prime}t^{2},\quad 0\leq t\leq\varepsilon_{1}.

Taking t=t0tnt=t_{0}-t_{n} in the second inequality we have

r(t0)r(tn)βn(t0tn)+C¯(t0tn)2.r(t_{0})-r(t_{n})\leq\beta_{n}(t_{0}-t_{n})+\bar{C}^{\prime}(t_{0}-t_{n})^{2}.

Together with (2.27) we see that, when nn is sufficiently large,

βnβ+δ.\beta_{n}\geq\beta+\delta.

Taking nn large such that ϵ:=t0tnε12\epsilon:=t_{0}-t_{n}\leq\frac{\varepsilon_{1}}{2} and taking t=2ϵt=2\epsilon in (2.28) we have

r(t0+ϵ)\displaystyle r(t_{0}+\epsilon) \displaystyle\geq r(tn)+2βnϵ+O(ϵ2)\displaystyle r(t_{n})+2\beta_{n}\epsilon+O(\epsilon^{2})
\displaystyle\geq r(tn)+βnϵ+C¯ϵ2+(β+δ)ϵ+O(ϵ2)\displaystyle r(t_{n})+\beta_{n}\epsilon+\bar{C}^{\prime}\epsilon^{2}+(\beta+\delta)\epsilon+O(\epsilon^{2})
\displaystyle\geq r(t0)+(β+δ)ϵ+O(ϵ2).\displaystyle r(t_{0})+(\beta+\delta)\epsilon+O(\epsilon^{2}).

Together with (2.21) we have

O(ϵ2)δϵ+O(ϵ2).O(\epsilon^{2})\geq\delta\epsilon+O(\epsilon^{2}).

This is a contradiction when ϵ\epsilon is small, or, equivalently nn is sufficiently large.

Step 4. Existence and continuity of r(t)r^{\prime}(t) at t0t_{0} when β=r(t0)>0\beta=r^{\prime}(t_{0})>0. Combining Step 1 and Step 3 we have

r(t0)=D+r(t0)=Dr(t0)=β, when β=Dxv(x0,t0)>0.r^{\prime}(t_{0})=D^{+}r(t_{0})=D^{-}r(t_{0})=\beta,\quad\mbox{ when }\beta=-D^{-}_{x}v(x_{0},t_{0})>0.

The proof in Step 3 also implies that, for any subsequence {ni}\{n_{i}\} of {n}\{n\}, βniβ+δ\beta_{n_{i}}\geq\beta+\delta is impossible, and so lim supnβnβ\limsup_{n\to\infty}\beta_{n}\leq\beta. In a similar way as in Step 3, one can show that βniβδ\beta_{n_{i}}\leq\beta-\delta is impossible, and so lim infnβnβ\liminf_{n\to\infty}\beta_{n}\geq\beta. Therefore, r(t)r^{\prime}(t) is continuous at t0t_{0} from the left side. On the other hand, if the time sequence {tn}\{t_{n}\} in (2.26) is a decreasing one tending to t0t_{0} from right side, then one can prove as above that r(tn)β(n)r^{\prime}(t_{n})\to\beta\ (n\to\infty). This means that r(t)r^{\prime}(t) is continuous at t0t_{0} from the right side. Combining the above conclusions together we get rC1r\in C^{1} at any time t0t_{0} where r(t0)>0r^{\prime}(t_{0})>0.

Step 5. The case β:=Dxv(x0,t0)=0\beta:=-D^{-}_{x}v(x_{0},t_{0})=0. For any small ε>0\varepsilon>0, we define

v~0(x):=v(x,t0)+ε(x0x),x,\tilde{v}_{0}(x):=v(x,t_{0})+\varepsilon(x_{0}-x),\quad x\in{\mathbb{R}},

and consider the equation of vv for tt0t\geq t_{0}, with initial data v(x,t0)v(x,t_{0}) replaced by v~0(x)\tilde{v}_{0}(x). Then by Step 1 we have D+r(t0)Dxv~0(x0)=εD^{+}r(t_{0})\leq-D^{-}_{x}\tilde{v}_{0}(x_{0})=\varepsilon. Since ε>0\varepsilon>0 is arbitrary we actually have D+r(t0)=0D^{+}r(t_{0})=0. Now as the proof in Step 3, with β>0\beta>0 being replaced by β=0\beta=0, one can prove in a similar way that Dr(t0)=0D^{-}r(t_{0})=0, and so r(t0)=0r^{\prime}(t_{0})=0. This implies that the Darcy law (2.19) remains hold even if β=0\beta=0. Finally, the continuity of r(t)r^{\prime}(t) at t0t_{0} where r(t0)=0r^{\prime}(t_{0})=0 can be shown as in Step 4.

This completes the proof of the theorem. ∎

Remark 2.8.

From this theorem we know that, for t>tt>t^{*}, r(t)0r^{\prime}(t)\geq 0 and it is not identical zero in any time interval. Though we do not have the strong conclusion r(t)>0r^{\prime}(t)>0 as in PME (cf. [30, Corollary 15.23]), the current properties are enough to continue the study for the asymptotic behavior of the solutions. For example, they are enough in the application of the zero number argument.

3. Zero Number Argument and General Convergence Result

In this section we present the intersection number properties, prove that the solution is strictly monotone outside its initial support, and then prove the general convergence result.

3.1. Zero number and intersection number properties

A useful tool we will use in the qualitative study is the so-called zero number argument. For the convenience of the readers, we first prepare some basic results in this area. Consider

(3.1) ηt=a(x,t)ηxx+b(x,t)ηx+c(x,t)η in E0:={(x,t)b1(t)<x<b2(t),t(t1,t2)},\eta_{t}=a(x,t)\eta_{xx}+b(x,t)\eta_{x}+c(x,t)\eta\quad\mbox{ in }E_{0}:=\{(x,t)\mid b_{1}(t)<x<b_{2}(t),\ t\in(t_{1},t_{2})\},

where b1b_{1} and b2b_{2} are continuous functions in (t1,t2)(t_{1},t_{2}). For each t(t1,t2)t\in(t_{1},t_{2}), denote by

𝒵(t):=#{xJ¯(t)η(,t)=0}\mathcal{Z}(t):=\#\{x\in\bar{J}(t)\mid\eta(\cdot,t)=0\}

the number of zeroes of η(,t)\eta(\cdot,t) in the interval J¯(t):=[b1(t),b2(t)]\bar{J}(t):=[b_{1}(t),b_{2}(t)]. A point x0J¯(t)x_{0}\in\bar{J}(t) is called a multiple zero (or degenerate zero) of η(,t)\eta(\cdot,t) if η(x0,t)=ηx(x0,t)=0\eta(x_{0},t)=\eta_{x}(x_{0},t)=0. In 1988, Angenent [1] proved a zero number diminishing property, and in 1998, the conditions in [1] were weakened by Chen [9] for solutions with lower regularity. One of their results can be summarized as the following:

Proposition 3.1 ([1, 9], zero number diminishing properties).

Assume the coefficients in (3.1) satisfies

(3.2) a,a1,at,ax,b,cL.a,a^{-1},a_{t},a_{x},b,c\in L^{\infty}.

Let η\eta be a nontrivial Wp,loc2,1W^{2,1}_{p,loc} solution of (3.1). Further assume that, for i=1,2i=1,2, either

(3.3) η(bi(t),t)0,t(t1,t2),\eta(b_{i}(t),t)\not=0,\quad t\in(t_{1},t_{2}),

or,

(3.4) bi(t)C1 and η(bi(t),t)0,t(t1,t2),b_{i}(t)\in C^{1}\mbox{ and }\eta(b_{i}(t),t)\equiv 0,\quad t\in(t_{1},t_{2}),

or,

(3.5) bi(t)C1 and ηx(bi(t),t)0,t(t1,t2).b_{i}(t)\in C^{1}\mbox{ and }\eta_{x}(b_{i}(t),t)\equiv 0,\quad t\in(t_{1},t_{2}).

Then

  • (i)

    𝒵(t)\mathcal{Z}(t) is finite and decreasing in t(t1,t2)t\in(t_{1},t_{2});

  • (ii)

    if s(t1,t2)s\in(t_{1},t_{2}) and x0J¯(s)x_{0}\in\bar{J}(s) is a multiple zero of η(,s)\eta(\cdot,s), then 𝒵(s1)>𝒵(s2)\mathcal{Z}(s_{1})>\mathcal{Z}(s_{2}) for all s1,s2s_{1},s_{2} satisfying t1<s1<s<s2<t2t_{1}<s_{1}<s<s_{2}<t_{2}.

Note that, in the original results in [1, 9], the problem was considered in fixed intervals, that is, bi(t)b_{i}(t) are constants. Using our assumptions (3.3)-(3.5), for any time interval with small length, we can straighten the domain boundaries so that the zero number diminishing properties remain hold. The boundedness assumption of a1a^{-1} in (3.2) is clearly a difficulty in using the zero number argument for RPMEs. Nevertheless, under some additional conditions, we can prove a diminishing property for the number of intersection points between two solutions of (pCP), despite the degeneracy of the equation. The analogue results for RDEs was proved in [12, Lemma 2.4], but the proof for our current result is more complicated due to the degeneracy.

For i=1,2i=1,2, let viv_{i} be the solution of (pCP) with initial data in

{ψC()ψ(x)>0 in (li(0),ri(0)),ψ(x)=0 for x(li(0),ri(0))}.\{\psi\in C({\mathbb{R}})\mid\psi(x)>0\mbox{ in }(l_{i}(0),r_{i}(0)),\quad\psi(x)=0\mbox{ for }x\not\in(l_{i}(0),r_{i}(0))\}.

Denote li(t)l_{i}(t) and ri(t)r_{i}(t) the left and right free boundaries of viv_{i}, respectively. Write

l(t):=max{l1(t),l2(t)},r(t):=min{r1(t),r2(t)},t0.l(t):=\max\{l_{1}(t),l_{2}(t)\},\quad r(t):=\min\{r_{1}(t),r_{2}(t)\},\qquad t\geq 0.

Then in the case where l(t)<r(t)l(t)<r(t), the number of the positive intersection points between v1(,t)v_{1}(\cdot,t) and v2(,t)v_{2}(\cdot,t) in {\mathbb{R}}, which is denoted by 𝒵0(t)\mathcal{Z}_{0}(t) as in (1.11), is the same as the number of intersection points in J0(t):=(l(t),r(t))J_{0}(t):=(l(t),r(t)).

Theorem 3.2.

Assume (F), v1,v2v_{1},v_{2} are given as above. Assume that 𝒵0(0)\mathcal{Z}_{0}(0) is finite. Then there exists a time τ[0,+]\tau\in[0,+\infty] , such that

  1. (i)

    𝒵0(t)\mathcal{Z}_{0}(t) decreases for t(0,τ)t\in(0,\tau) and for t(τ,)t\in(\tau,\infty);

  2. (ii)

    limtτ+0𝒵0(t)=1\lim\limits_{t\rightarrow\tau+0}\mathcal{Z}_{0}(t)=1 and limtτ0𝒵0(t)=0\lim\limits_{t\rightarrow\tau-0}\mathcal{Z}_{0}(t)=0 if τ(0,)\tau\in(0,\infty).

The proof of this theorem consists of the following two lemmas.

Lemma 3.3.

Assume (F), v1,v2v_{1},v_{2} are given as above. Further assume that l(0)<r(0)l(0)<r(0) and one of the following hypotheses holds:

  1. (a)

    l1(0)l2(0),r1(0)r2(0)l_{1}(0)\not=l_{2}(0),\ r_{1}(0)\not=r_{2}(0);

  2. (b)

    𝒵0(0)<\mathcal{Z}_{0}(0)<\infty.

Then

  1. (i).

    𝒵0(t)\mathcal{Z}_{0}(t) is decreasing in t>0t>0;

  2. (ii).

    𝒵0(t)\mathcal{Z}_{0}(t) is strictly decreasing when tt passes through ss, if v1(,s)v2(,s)v_{1}(\cdot,s)-v_{2}(\cdot,s) has a degenerate zero in J0(s)J_{0}(s), or, if one interior zero moves to the boundary l(t)l(t) or r(t)r(t) as ts0t\to s-0.

Proof.

We first prove the conclusions under the assumption (a). The main complexity is caused by the possibility that an interior intersection point moves up to the domain boundaries. Without loss of generality, we assume

l1(t)<l2(t) for small t0,r1(t)<r2(t) for all t0.l_{1}(t)<l_{2}(t)\mbox{ for small }t\geq 0,\quad r_{1}(t)<r_{2}(t)\mbox{ for all }t\geq 0.

Then the left boundaries l1(t)l_{1}(t) and l2(t)l_{2}(t) may meet after some time, but the right boundaries will not.

At the early stage, say t[0,t1]t\in[0,t_{1}] for some t1>0t_{1}>0, l(t)=l2(t)>l1(t)l(t)=l_{2}(t)>l_{1}(t), and η(x,t):=v1(x,t)v2(x,t)\eta(x,t):=v_{1}(x,t)-v_{2}(x,t) satisfies a linear equation

(3.6) ηt=a(x,t)ηxx+b(x,t)ηx+c(x,t)η,xJ0(t):=(l(t),r(t)),t(0,t1],\eta_{t}=a(x,t)\eta_{xx}+b(x,t)\eta_{x}+c(x,t)\eta,\quad x\in J_{0}(t):=(l(t),r(t)),\ t\in(0,t_{1}],

with

a(x,t):=(m1)v1(x,t),b(x,t):=v1x(x,t)+v2x(x,t),a(x,t):=(m-1)v_{1}(x,t),\quad b(x,t):=v_{1x}(x,t)+v_{2x}(x,t),

and

c(x,t):=(m1)v2xx(x,t)+{g(v1(x,t))g(v2(x,t))v1(x,t)v2(x,t),v1(x,t)v2(x,t),0,v1(x,t)=v2(x,t).c(x,t):=(m-1)v_{2xx}(x,t)+\left\{\begin{array}[]{ll}\frac{g(v_{1}(x,t))-g(v_{2}(x,t))}{v_{1}(x,t)-v_{2}(x,t)},&v_{1}(x,t)\not=v_{2}(x,t),\\ 0,&v_{1}(x,t)=v_{2}(x,t).\end{array}\right.

By the continuity of v1,v2v_{1},v_{2},

η(l(t),t)=v1(l2(t),t)ρ,η(r(t),t)=v2(r1(t),t)ρ,t[0,t1],\eta(l(t),t)=v_{1}(l_{2}(t),t)\geq\rho,\quad\eta(r(t),t)=-v_{2}(r_{1}(t),t)\leq-\rho,\quad t\in[0,t_{1}],

for some small ρ>0\rho>0. Take two smooth curves l^(t)\hat{l}(t) and r^(t)\hat{r}(t) such that

l(t)<l^(t)l(t)+1,r(t)1r^(t)<r(t),l(t)<\hat{l}(t)\ll l(t)+1,\quad r(t)-1\ll\hat{r}(t)<r(t),
(3.7) η(x,t)0,x[l(t),l^(t)][r^(t),r(t)],t[0,t1],\eta(x,t)\not=0,\quad x\in[l(t),\hat{l}(t)]\cup[\hat{r}(t),r(t)],\ t\in[0,t_{1}],

and, for some ρ1>0\rho_{1}>0,

v1(x,t),v2(x,t)ρ1 in E1:={(x,t)xJ^(t):=[l^(t),r^(t)],t[0,t1]}.v_{1}(x,t),\ v_{2}(x,t)\geq\rho_{1}\mbox{\ \ in \ \ }E_{1}:=\{(x,t)\mid x\in\hat{J}(t):=[\hat{l}(t),\hat{r}(t)],\ t\in[0,t_{1}]\}.

Then both v1v_{1} and v2v_{2} are classical in E1E_{1}, we can use the classical zero number diminishing property in E1E_{1} to conclude that 𝒵J^(t)(t)\mathcal{Z}_{\hat{J}(t)}(t), which denotes the zero number of η(,t)\eta(\cdot,t) in the interval J^(t)\hat{J}(t), is finite, decreasing in tt, and strictly decreasing when tt passes through a moment ss when η(,s)\eta(\cdot,s) has a degenerate zero. Note that, the same conclusions hold true for 𝒵0(t)\mathcal{Z}_{0}(t) since it equals to 𝒵J^(t)(t)\mathcal{Z}_{\hat{J}(t)}(t) by (3.7).

If l2(t)l_{2}(t) will never approach l1(t)l_{1}(t) for all t>0t>0, then the conclusions are proved.

We now consider the case: there is a moment T>t1T>t_{1} such that l2(t)l_{2}(t) moves leftward and meets l1(t)l_{1}(t) as tT0t\to T-0, and suppose that TT is the first one of such moments. Due to the diminishing property for 𝒵0(t)\mathcal{Z}_{0}(t) in the time interval (0,T)(0,T) we see that there are at most finite moments when η(,t)\eta(\cdot,t) has degenerate zeros (only in J0(t)J_{0}(t)). Thus, there exists t2(t1,T)t_{2}\in(t_{1},T) such that η(,t)\eta(\cdot,t) has no degenerate zeros in J0(t)J_{0}(t) in the time interval (t2,T)(t_{2},T), that is, in this time interval η(,t)\eta(\cdot,t) has fixed finite number of non-degenerate zeros. For simplicity and without loss of generality, in the rest of the proof we assume that η(,t)\eta(\cdot,t) has no interior degenerate zeros in J0(t)J_{0}(t) for all t(t2,)t\in(t_{2},\infty). So we can assume the null curves of η\eta in E2:={(x,t)xJ0(t),t2<t<T}E_{2}:=\{(x,t)\mid x\in J_{0}(t),\ t_{2}<t<T\} are

γ1(t)<γ2(t)<<γk(t),t(t2,T),\gamma_{1}(t)<\gamma_{2}(t)<\cdots<\gamma_{k}(t),\quad t\in(t_{2},T),

for some positive integer kk. Using [15, Theorem 2] as in the proof of Lemma 2.4 in [12] one can show that

xj:=limtT0γj(t),j=1,2,,k,x_{j}:=\lim\limits_{t\to T-0}\gamma_{j}(t),\quad j=1,2,\cdots,k,

exist. We divide the situation into two cases.

Case 1. l(T)<x1l(T)<x_{1}, that is, l2(t)l_{2}(t) moves leftward to catch up with l1(t)l_{1}(t) as tT0t\to T-0, while the left-most null curve γ1(t)\gamma_{1}(t) remains on the right of l(t)l(t) till t=Tt=T. In this case, by using the comparison principle for very weak solutions in the domain

E3:={(x,t)<x<γ1(t),t(t2,T)}E_{3}:=\{(x,t)\mid-\infty<x<\gamma_{1}(t),\ t\in(t_{2},T)\}

we have

η(x,T)>0 for l(T)<x<x1,η(l(T),T)=η(x1,T)=0.\eta(x,T)>0\mbox{ for }l(T)<x<x_{1},\quad\eta(l(T),T)=\eta(x_{1},T)=0.

As a consequence we have 𝒵0(T)=k\mathcal{Z}_{0}(T)=k. The boundary zero l(T)l(T) of η(,T)\eta(\cdot,T) is not included in 𝒵0(T)\mathcal{Z}_{0}(T). Recalling the Darcy law we have the additional fact Dx+η(l(T),T)=0D^{+}_{x}\eta(l(T),T)=0. We remark that, this will lead to a contradiction with the Hopf lemma immediately in the Stefan problems for RDEs (cf. [11, 12]). However, in the current problem, the fact Dx+η(l(T),T)=0D^{+}_{x}\eta(l(T),T)=0 does not contradict the Hopf lemma, since the Hopf lemma is no longer necessarily to be true at (l(T),T)(l(T),T) where the equation (3.6) is degenerate.

Next, we consider the time period from the time TT to t3(T,]t_{3}\in(T,\infty], where t3t_{3} is the smallest moment when γ1(t3)\gamma_{1}(t_{3}) meets l(t3)l(t_{3}). Using the comparison principle in the domain

E4:={(x,t)<x<γ1(t),t[T,t3)},E_{4}:=\{(x,t)\mid-\infty<x<\gamma_{1}(t),\ t\in[T,t_{3})\},

we conclude that η(x,t)0\eta(x,t)\geq 0 in this domain. This implies that, for t[T,t3)t\in[T,t_{3}), l1(t)l2(t)l_{1}(t)\leq l_{2}(t), but the equality can hold from time to time. However, no matter l1(t)<l2(t)l_{1}(t)<l_{2}(t) or l1(t)=l2(t)l_{1}(t)=l_{2}(t), we have v1,v2>0v_{1},v_{2}>0 for x(l(t),γ1(t)]x\in(l(t),\gamma_{1}(t)], and so the strong maximum principle holds in

E5:={(x,t)l(t)<xγ1(t),Tt<t3},E_{5}:=\{(x,t)\mid l(t)<x\leq\gamma_{1}(t),\ T\leq t<t_{3}\},

which implies that η(x,t)>0\eta(x,t)>0 in E5E_{5}. As a consequence, 𝒵0(t)=k\mathcal{Z}_{0}(t)=k for t[T,t3)t\in[T,t_{3}).

At the moment t3t_{3}, γ1(t3)\gamma_{1}(t_{3}) meets l(t3)l(t_{3}). The following analysis for tt3t\geq t_{3} is similar as what we will do in Case 2.

Case 2. γ1(t)\gamma_{1}(t) tends to l(t)l(t) at tT0t\to T-0, and so l1(T)=l2(T)=γ1(T)l_{1}(T)=l_{2}(T)=\gamma_{1}(T) (see Figure 4). Recall that we assumed before that η(,t)\eta(\cdot,t) no longer has interior degenerate zeros. However, this does not exclude that possibility that γ1(T)=γ2(T)==γj0(T)=l(T)\gamma_{1}(T)=\gamma_{2}(T)=\cdots=\gamma_{j_{0}}(T)=l(T), that is, j0j_{0}-null curves tend to the point (l(T),T)(l(T),T) at the same time as tT0t\to T-0. Since the discussion is similar, in what follows, we only consider the case

l1(T)=l2(T)=l(T)=γ1(T)<γ2(T)<<γk(T).l_{1}(T)=l_{2}(T)=l(T)=\gamma_{1}(T)<\gamma_{2}(T)<\cdots<\gamma_{k}(T).

Then we have 𝒵0(T)=k1\mathcal{Z}_{0}(T)=k-1, and

η(x,T)<0,l(T)<x<γ2(T).\eta(x,T)<0,\quad l(T)<x<\gamma_{2}(T).
Refer to caption
Figure 4. Vanishing of zero at t=Tt=T.

As above we next consider the time period from TT to t4(T,]t_{4}\in(T,\infty], where t4t_{4} is the smallest moment when γ2(t4)\gamma_{2}(t_{4}) meets l(t4)l(t_{4}). Using the comparison principle for very weak solutions in the domain

E6:={(x,t)<x<γ2(t),t[T,t4)},E_{6}:=\{(x,t)\mid-\infty<x<\gamma_{2}(t),\ t\in[T,t_{4})\},

we conclude that η(x,t)0\eta(x,t)\leq 0 in this domain. This implies that, for t[T,t4)t\in[T,t_{4}), l2(t)l1(t)l_{2}(t)\leq l_{1}(t), but the equality can hold from time to time. However, no matter l2(t)<l1(t)l_{2}(t)<l_{1}(t) or l2(t)=l1(t)l_{2}(t)=l_{1}(t), we have v1,v2>0v_{1},v_{2}>0 for x(l(t),γ2(t)]x\in(l(t),\gamma_{2}(t)], and so the strong maximum principle holds in

E7:={(x,t)l(t)<xγ2(t),Tt<t4},E_{7}:=\{(x,t)\mid l(t)<x\leq\gamma_{2}(t),\ T\leq t<t_{4}\},

which implies that η(x,t)<0\eta(x,t)<0 in E7E_{7}. As a consequence, 𝒵0(t)=k1\mathcal{Z}_{0}(t)=k-1 for t[T,t4)t\in[T,t_{4}).

In the case where t4=t_{4}=\infty, there is nothing left to prove. In the case where t4<t_{4}<\infty, we can analyze the situation for tt4t\geq t_{4} as in the current case.

As a conclusion, we see that, in case no interior degenerate zeros appear, the zero number of η(,t)\eta(\cdot,t) in the open interval J0(t)J_{0}(t) is decreasing. It is strictly decreasing when some interior null curves touch the boundaries.

Next we consider the assumption (b). Without loss of generality, we assume that l1(0)=l2(0)l_{1}(0)=l_{2}(0), that is, the left free boundaries of v1v_{1} and v2v_{2} glue together at the beginning. Since 𝒵0(t)<\mathcal{Z}_{0}(t)<\infty, there exists x~>l1(0)=l2(0)\tilde{x}>l_{1}(0)=l_{2}(0), such that v1(x,0)v_{1}(x,0) and v2(x,0)v_{2}(x,0) has no intersection points in (l1(0),x~)(l_{1}(0),\tilde{x}). This is a situation like that in Case 1 at time TT. The rest discussion is then similar as above.

This completes the proof of the lemma. ∎

Remark 3.4.

In the above proof, as well as in other related places of the paper, we use the maximum principle for η\eta. One may worry about the unboundedness of the coefficient cc in (3.6) at the points where v2=0v_{2}=0. We remark that, though we write v1v2v_{1}-v_{2} in the form of η\eta for convenience, we actually compare the very weak solutions v1v_{1} and v2v_{2}. As in [30, Theorem 5.5] the comparison principle between v1,v2v_{1},v_{2} holds safely. The main reason is that each of them can be approximated by a sequence of decreasing classical solutions.

Lemma 3.5.

Assume (F), v1,v2v_{1},v_{2} are given as above. Further assume that l(0)r(0)l(0)\geq r(0). Then there exists a time t1[0,]t_{1}\in[0,\infty] such that

{𝒵0(t)=0,tt1,𝒵0(t)=1,t1<tt1+1,𝒵0(t) decreases,t>t1.\begin{cases}\mathcal{Z}_{0}(t)=0,&t\leq t_{1},\\ \mathcal{Z}_{0}(t)=1,&t_{1}<t\ll t_{1}+1,\\ \mathcal{Z}_{0}(t)\mbox{ decreases},&t>t_{1}.\end{cases}
Proof.

By the assumption l(0)r(0)l(0)\geq r(0), we see that the supports of two initial data v1(x,0)v_{1}(x,0) and v2(x,0)v_{2}(x,0) do not have overlap. Without loss of generality, we assume v1v_{1} lies on the left of v2v_{2} in the beginning time, that is r1(0)l2(0)r_{1}(0)\leq l_{2}(0). Define t1:=sup{t0r1(t)l2(t)}[0,]t_{1}:=\sup\{t\geq 0\mid r_{1}(t)\leq l_{2}(t)\}\in[0,\infty], then 𝒵0(t)=0\mathcal{Z}_{0}(t)=0 when tt1t\leq t_{1} by the definition of 𝒵0(t)\mathcal{Z}_{0}(t).

Refer to caption
Figure 5. Generation of zero at t=t1t=t_{1}.

When t1<t_{1}<\infty, r1(t)r_{1}(t) will surpass l2(t)l_{2}(t) at t=t1t=t_{1} (see Figure 5). According to the monotonicity of viv_{i} near the free boundaries, we obtain 𝒵0(t)=1\mathcal{Z}_{0}(t)=1 for t1<tt1+1t_{1}<t\ll t_{1}+1. After t1t_{1}, the problem can be treated like that in Lemma 3.3. Then the lemma follows. ∎

Proof of Theorem 1.2. The proof is almost the same as that of Theorem 3.2. We only note that the support of initial data may not be singly connected, but consists of finite intervals. Similar to t1t_{1} in Lemma 3.5, we define tjt_{j} as the jj-th time for two components (in one of them v1>0v_{1}>0 and in the other v2>0v_{2}>0) to meet. It is easy to see that #{tj}\#\{t_{j}\} is at most finite up to the choice of initial data. However, more complex than the case in Lemma 3.5, 𝒵0(t)\mathcal{Z}_{0}(t) may not increase strictly at t=tjt=t_{j}, since there may exist other interior degenerate zeros of v1v2v_{1}-v_{2} which vanishes at the same time. To exclude such cases, we select a subset of {tj}\{t_{j}\} such that 𝒵0(t)\mathcal{Z}_{0}(t) strictly increases at these times. This proves Theorem 1.2. ∎

3.2. Monotonicity outside of the initial support

Assume the initial data u0u_{0} of (CP) belongs to

(3.8) 𝔛1:={ψC()ψ(x)>0 for x(b,b), and ψ(x)=0 for |x|b}.\mathfrak{X}_{1}:=\{\psi\in C({\mathbb{R}})\mid\psi(x)>0\mbox{ for }x\in(-b,b),\mbox{ and }\psi(x)=0\mbox{ for }|x|\geq b\}.

Then the solution has exactly two free boundaries: l(t)<r(t)l(t)<r(t) with l(0)=bl(0)=-b and r(0)=br(0)=b. Moreover, l(t)l(t) is strictly decreasing after the waiting time t1(b)t^{*}_{1}(-b), r(t)r(t) is strictly increasing after the waiting time t2(b)t^{*}_{2}(b), and spt[u(,t)]=[l(t),r(t)]\mbox{spt}[u(\cdot,t)]=[l(t),r(t)]. We now prove a monotonicity property for u(,t)u(\cdot,t) outside of [b,b][-b,b].

Lemma 3.6.

Assume (F) and u0𝔛1u_{0}\in\mathfrak{X}_{1}. Then

  1. (i).

    for any given x¯>b\bar{x}>b, if r(t¯)=x¯r(\bar{t})=\bar{x} for some t¯>t(b)\bar{t}>t^{*}(b), then vx(x¯,t)<0v_{x}(\bar{x},t)<0 for t>t¯t>\bar{t}. Similarly, for any given x¯<b\bar{x}^{\prime}<-b, if l(t¯)=x¯l(\bar{t}^{\prime})=\bar{x}^{\prime} for some t¯>t(b)\bar{t}^{\prime}>t^{*}(-b), then vx(x¯,t)>0v_{x}(\bar{x}^{\prime},t)>0 for t>t¯t>\bar{t}^{\prime};

  2. (ii).

    2br(t)+l(t)2b-2b\leq r(t)+l(t)\leq 2b for all t>0t>0.

Proof.

(i). We only work on x¯>b\bar{x}>b since the case x¯<b\bar{x}^{\prime}<-b is studied similarly. Given x¯>b\bar{x}>b, either r(t)x¯r(t)\leq\bar{x} for all t>0t>0, or, by Theorem 2.7, there exists a unique t¯>t(b)\bar{t}>t^{*}(b) such that r(t¯)=x¯>br(\bar{t})=\bar{x}>b and r(t)>x¯r(t)>\bar{x} for t>t¯t>\bar{t}. We consider only the latter case and will use the so-called Aleksandrov’s reflection principle. Set

η(x,t):=v(x,t)v(2x¯t),(x,t)E1:={(x,t)<xx¯,tt¯}.\eta(x,t):=v(x,t)-v(2\bar{x}-t),\quad(x,t)\in E_{1}:=\{(x,t)\mid-\infty<x\leq\bar{x},\ t\geq\bar{t}\}.

Then, for some functions c1=c1(vx)c_{1}=c_{1}(v_{x}) and c2=c2(m,g,v,vxx)c_{2}=c_{2}(m,g^{\prime},v,v_{xx}), η\eta satisfies

{ηt=(m1)vηxx+c1ηx+c2η,(x,t)E1,η(x¯,t)=0,tt¯,η(x,t¯)0,xx¯.\left\{\begin{array}[]{ll}\eta_{t}=(m-1)v\eta_{xx}+c_{1}\eta_{x}+c_{2}\eta,&(x,t)\in E_{1},\\ \eta(\bar{x},t)=0,&t\geq\bar{t},\\ \eta(x,\bar{t})\geq 0,&x\leq\bar{x}.\end{array}\right.

Using the maximum principle (see Remark 3.4 if one worries about the unboundedness of c2c_{2}) we have

(3.9) η(x,t)0 in E1,ηx(x¯,t)0 for tt¯.\eta(x,t)\geq 0\mbox{ in }E_{1},\quad\eta_{x}(\bar{x},t)\leq 0\mbox{ for }t\geq\bar{t}.

Furthermore, v(x¯,t)>0v(\bar{x},t)>0 for t>t¯t>\bar{t} by the positivity persistence, and so both vv and η\eta are classical near the line {x=x¯,t>t¯}\{x=\bar{x},\ t>\bar{t}\}. This implies that the Hopf lemma is applicable and so

ηx(x¯,t)=2vx(x¯,t)<0 for t>t¯.\eta_{x}(\bar{x},t)=2v_{x}(\bar{x},t)<0\mbox{ for }t>\bar{t}.

(ii). The first inequality in (3.9) also implies that

l(t)2x¯r(t),t>t¯,l(t)\leq 2\bar{x}-r(t),\quad t>\bar{t},

that is,

l(t)+r(t)2x¯,t>t¯.l(t)+r(t)\leq 2\bar{x},\quad t>\bar{t}.

Finally, since x¯b>0\bar{x}-b>0 can be chosen as small as possible, so does t¯t(b)\bar{t}-t^{*}(b), we actually have r(t)+l(t)2br(t)+l(t)\leq 2b for t>0t>0. In a similar way one can show that r(t)+l(t)2br(t)+l(t)\geq-2b for all t>0t>0. ∎

In the above lemma we consider initial data with exactly one connected support. Now we consider the case u0𝔛u_{0}\in\mathfrak{X}. More precisely, assume

(3.10) u0{ψC()|there exist bl1<r1l2<r2ln<rnb such that ψ(x)>0 in (lj,rj) for j=1,,n, and ψ(x)=0 otherwise}.u_{0}\in\left\{\psi\in C({\mathbb{R}})\left|\begin{array}[]{l}\mbox{there exist }-b\leq l_{1}<r_{1}\leq l_{2}<r_{2}\leq\cdots\leq l_{n}<r_{n}\leq b\mbox{ such that }\\ \psi(x)>0\mbox{ in }(l_{j},r_{j})\mbox{ for }j=1,\cdots,n,\mbox{ and }\psi(x)=0\mbox{ otherwise}\end{array}\right.\right\}.

Note that this set is exactly the same as 𝔛\mathfrak{X}. Here we explicitly give the name of each boundary (see Figure 1). As we will see below that, the purpose for choosing u0u_{0} in this set is mainly for the clarity of the statement for the free boundaries. Our approach remains valid for general continuous and compactly supported initial data with infinite many free boundaries.

For each i=1,2,,ni=1,2,\cdots,n, denote by li(t)l_{i}(t) (resp. ri(t)r_{i}(t)) the free boundary of the solution starting at lil_{i} (resp. rir_{i}). Given j{1,2,,n1}j\in\{1,2,\cdots,n-1\}, the boundary point rjr_{j} either keeps stationary for all time (that is, t(rj)=t^{*}(r_{j})=\infty), or it moves rightward after the finite waiting time. In the latter case, we define

Tj:=sup{Tthere exists xj[rj,lj+1] such that u(xj,t)=0 for 0tT}.T_{j}:=\sup\{T\mid\mbox{there exists }x_{j}\in[r_{j},l_{j+1}]\mbox{ such that }u(x_{j},t)=0\mbox{ for }0\leq t\leq T\}.

In case Tj=T_{j}=\infty, the free boundaries rj(t)r_{j}(t) will never meet lj+1(t)l_{j+1}(t), or, rj=lj+1r_{j}=l_{j+1} and both of them have infinite waiting times. In case Tj<T_{j}<\infty, these two free boundaries meet in some finite time and then this free boundary vanish after TjT_{j}. (Note that, the following case will not happen: they meet together in finite time, and this zero of uu will never vanish. In fact, by Step 2 in the proof of Theorem 2.7, any free boundary will never stop once it begins to move.) Hence, for t>Tjt>T_{j}, the intervals (lj(t),rj(t))(l_{j}(t),r_{j}(t)) and (lj+1(t),rj+1(t))(l_{j+1}(t),r_{j+1}(t)) merge into one: (lj(t),rj+1(t))(l_{j}(t),r_{j+1}(t)). We remark that, it may merge other intervals at the same time, or at some time after TjT_{j}. Whatever happens, there exists a large TT such that, for tTt\geq T, there exist fixed number of free boundaries:

(3.11) l1(t)ln1(t)<rn2(t)ln3(t)<rn4(t)ln2k1(t)<rn2k(t)rn(t),l_{1}(t)\equiv l_{n_{1}}(t)<r_{n_{2}}(t)\leq l_{n_{3}}(t)<r_{n_{4}}(t)\leq\cdots\leq l_{n_{2k-1}}(t)<r_{n_{2k}}(t)\equiv r_{n}(t),

for n1,,n2k{1,2,,n}n_{1},\cdots,n_{2k}\in\{1,2,\cdots,n\} satisfying

n1n2<n3n4<<n2k1n2k,n_{1}\leq n_{2}<n_{3}\leq n_{4}<\cdots<n_{2k-1}\leq n_{2k},

such that v(x,t)>0v(x,t)>0 in (ln2i1(t),rn2i(t))(i=1,2,,k)(l_{n_{2i-1}}(t),r_{n_{2i}}(t))\ (i=1,2,\cdots,k) and v=0v=0 otherwise. Moreover, by (3.10) and the monotonicity of li(t),ri(t)l_{i}(t),r_{i}(t) we have

(3.12) b<rn2(t),ln2k1(t)<b,t0,-b<r_{n_{2}}(t),\quad l_{n_{2k-1}}(t)<b,\quad t\geq 0,

Using the previous Lemma 3.6 in each of these intervals we have the following result.

Corollary 3.7.

Assume (F) and (I). Then there exists a large TT such that, for tTt\geq T, u(,t)u(\cdot,t) has 2k2k free boundaries as in (3.11) and (3.12). Moreover, for 1jk1\leq j\leq k, u(,t)u(\cdot,t) is strictly increasing in [l2j1(t),l2j1(0))[l_{2j-1}(t),l_{2j-1}(0)), and strictly decreasing in (r2j(0),r2j(t)](r_{2j}(0),r_{2j}(t)]. In particular, u(,t)u(\cdot,t) is strictly increasing in [l1(t),l1(0))[l_{1}(t),l_{1}(0)) and strictly decreasing in (rn(0),rn(t)](r_{n}(0),r_{n}(t)].

3.3. General convergence

Proof of Theorem 1.1. We prove the results by using the pressure vv instead of uu.

Step 1. To show the ω\omega-limit set is non-empty. By the locally uniform Hölder bound in (2.9), there exist C>0,α1(0,1)C>0,\ \alpha_{1}\in(0,1), both depend on v0C\|v_{0}\|_{C} and mm, such that, for any M>0M>0 and any increasing time sequence {tn}\{t_{n}\}, there holds,

v(x,tn+t)Cα1([M,M]×[1,1])C.\|v(x,t_{n}+t)\|_{C^{\alpha_{1}}([-M,M]\times[-1,1])}\leq C.

Hence, for any α(0,α1)\alpha\in(0,\alpha_{1}), there is a subsequence of {tn}\{t_{n}\}, denoted again by {tn}\{t_{n}\}, such that

v(x,tn+t)wM(x,t)Cα([M,M]×[1,1])0 as n.\|v(x,t_{n}+t)-w_{M}(x,t)\|_{C^{\alpha}([-M,M]\times[-1,1])}\to 0\mbox{\ \ as \ \ }n\to\infty.

Using Cantor’s diagonal argument, there exist a subsequence of {tn}\{t_{n}\}, denoted again by {tn}\{t_{n}\} and a function w(x,t)Cα(×)w(x,t)\in C^{\alpha}({\mathbb{R}}\times{\mathbb{R}}) such that

v(x,tn+t)w(x,t) as n,in the topology of Clocα(2).v(x,t_{n}+t)\to w(x,t)\mbox{\ \ as\ \ }n\to\infty,\quad\mbox{in the topology of }C^{\alpha}_{loc}({\mathbb{R}}^{2}).

In addition, if w(x,t)>0w(x,t)>0 in a domain E2E\subset{\mathbb{R}}^{2}, then for any compact subset DED\subset E, there exists small ρ>0\rho>0 such that

w(x,t),v(x,tn+t)ρ>0,(x,t)D,n1.w(x,t),\ v(x,t_{n}+t)\geq\rho>0,\quad(x,t)\in D,\ n\gg 1.

Then, v(x,tn+t)v(x,t_{n}+t) is classical in DD, and so, for any β1(0,1)\beta_{1}\in(0,1), v(x,tn+t)C2+β1,1+β1/2(D)C\|v(x,t_{n}+t)\|_{C^{2+\beta_{1},1+\beta_{1}/2}(D)}\leq C for any large nn and some CC independent of nn. This implies that a subsequence of {v(x,tn+t)}\{v(x,t_{n}+t)\} converges in C2+β,1+β/2(D)(0<β<β1)C^{2+\beta,1+\beta/2}(D)\ (0<\beta<\beta_{1}) to the limit w(x,t)w(x,t). Thus, w(x,t)w(x,t) is a classical solution of (pCP) in the domain where w(x,t)>0w(x,t)>0, and it is a very weak solution of (pCP) for (x,t)2(x,t)\in{\mathbb{R}}^{2}. Consequently, the ω\omega-limit set of vv in the topology of (1.7) is non-empty, compact, connected and invariant.

Step 2. Limits of the free boundaries. By Corollary 3.7, there exists a large TT such that v(,t)v(\cdot,t) has exactly 2k2k free boundaries as in (3.11) and (3.12) for all tTt\geq T. For each i=1,2,,ki=1,2,\cdots,k, since ln2i1(t)l_{n_{2i-1}}(t) is decreasing and rn2i(t)r_{n_{2i}}(t) is increasing, the following limits exist:

(3.13) l:=limtln1(t),ln2i1:=limtln2i1(t),rn2i:=limtrn2i(t),r:=limtrn2k(t).l^{\infty}:=\lim\limits_{t\to\infty}l_{n_{1}}(t),\quad l^{\infty}_{n_{2i-1}}:=\lim\limits_{t\to\infty}l_{n_{2i-1}}(t),\quad r^{\infty}_{n_{2i}}:=\lim\limits_{t\to\infty}r_{n_{2i}}(t),\quad r^{\infty}:=\lim\limits_{t\to\infty}r_{n_{2k}}(t).

We have the following possibilities:

Case 1. k2k\geq 2. In this case all the free boundaries lie in [b,b][-b,b] except for the left most one ln1(t)l1(t)l_{n_{1}}(t)\equiv l_{1}(t) and the right most one rn2k(t)rn(t)r_{n_{2k}}(t)\equiv r_{n}(t). Together with Lemma 3.6 (ii) we have

l1(t)2brn2(t)3b,rn(t)2bln2k1(t)3b.l_{1}(t)\geq-2b-r_{n_{2}}(t)\geq-3b,\quad r_{n}(t)\leq 2b-l_{n_{2k-1}}(t)\leq 3b.

Consequently, all the limits in (3.13) are in [3b,3b][-3b,3b].

Case 2. k=1k=1 and l,r<-l^{\infty},r^{\infty}<\infty. In this case v(,t)v(\cdot,t) has exactly one connected compact support [ln1(t),rn2(t)]=[l1(t),rn(t)][l_{n_{1}}(t),r_{n_{2}}(t)]=[l_{1}(t),r_{n}(t)] which tends to the bounded interval [l,r][l^{\infty},r^{\infty}] as tt\to\infty. Note that, even in this case, the support of the ω\omega-limit w(x,t1)w(x,t_{1}) may have several separated intervals as in the ground state solution in (1.8), since the convergence vwv\to w is taken in Clocα()C^{\alpha}_{loc}({\mathbb{R}}) topology, the positivity of vv does not guarantee the positivity of its limit ww.

Case 3. k=1k=1 and l,r=-l^{\infty},r^{\infty}=\infty. In this case v(,t)v(\cdot,t) has exactly one connected compact support [ln1(t),rn2(t)]=[l1(t),rn(t)][l_{n_{1}}(t),r_{n_{2}}(t)]=[l_{1}(t),r_{n}(t)] which tends to {\mathbb{R}} as tt\to\infty.

Step 3. Quasi-convergence and Type II ground state solutions in the case l,r<-l^{\infty},r^{\infty}<\infty. This includes Cases 1 and 2 in Step 2. We first use the Lyapunov functional to prove the quasi-convergence result, that is, any ω\omega-limit is a stationary solution of (pCP). Define

(3.14) E[v(,t)]:=l1(t)rn(t)[m12v2m1vx2G(v)]𝑑x,tT,E[v(\cdot,t)]:=\int_{l_{1}(t)}^{r_{n}(t)}\left[\frac{m-1}{2}v^{\frac{2}{m-1}}v_{x}^{2}-G(v)\right]dx,\quad t\geq T,

with

G(v):=0vg(r)r3mm1𝑑r,G(v):=\int_{0}^{v}g(r)r^{\frac{3-m}{m-1}}dr,

which is convergent by |g(v)|K(m1)|v||g(v)|\leq K(m-1)|v|. When vv is the solution of (pCP), a direct calculation shows that

ddtE[v(,t)]=l1(t)rn(t)v3mm1vt2𝑑x0.\frac{d}{dt}E[v(\cdot,t)]=-\int_{l_{1}(t)}^{r_{n}(t)}v^{\frac{3-m}{m-1}}v_{t}^{2}dx\leq 0.

So, E[v(,t)]E[v(\cdot,t)] is a Lyapunov functional. It is bounded from below by our assumption ll1(t)<rn(t)rl^{\infty}\leq l_{1}(t)<r_{n}(t)\leq r^{\infty}. Hence, we can use the standard argument to show that any ω\omega-limit of v(,t)v(\cdot,t) is a stationary solution of the equation pRPME.

We now explain that any ω\omega-limit is also a stationary solution of the Cauchy problem (pCP). Assume, w(x,t)w(x,t) is the limit function obtained in Step 1. For each τ1\tau_{1}\in{\mathbb{R}}, we claim that the solution v(x,t;w(,τ1))v(x,t;w(\cdot,\tau_{1})) of (pCP) with v0(x)=w(x,τ1)v_{0}(x)=w(x,\tau_{1}) is a stationary one: v(x,t;w(,τ1))w(x,τ1)v(x,t;w(\cdot,\tau_{1}))\equiv w(x,\tau_{1}). Since w(x,τ1)w(x,\tau_{1}) has boundaries l,ln2i1,rn2i,rl^{\infty},l^{\infty}_{n_{2i-1}},r^{\infty}_{n_{2i}},r^{\infty} as in (3.13), by Lemma 2.3 we only need to show that the waiting time of each of these boundaries is not zero. Without loss of generality, we only prove t(l)>0t^{*}(l^{\infty})>0. By contradiction, assume the left boundary lw(t)l_{w}(t) of v(x,t;w(,τ1))v(x,t;w(\cdot,\tau_{1})) satisfies lw(τ2)<ll_{w}(\tau_{2})<l^{\infty} for some τ2>0\tau_{2}>0. Since w(x,τ1+τ2)w(x,\tau_{1}+\tau_{2}) is also an ω\omega-limit of vv:

v(x,tn+τ1+τ2)w(x,τ1+τ2)v(x,τ2;w(,τ1)),n.v(x,t_{n}+\tau_{1}+\tau_{2})\to w(x,\tau_{1}+\tau_{2})\equiv v(x,\tau_{2};w(\cdot,\tau_{1})),\quad n\to\infty.

(Note that w(x,τ1+t)w(x,\tau_{1}+t) solves the (pCP) with initial data w(x,τ1)w(x,\tau_{1}) at t=0t=0. It is nothing but v(x,t;w(,τ1))v(x,t;w(\cdot,\tau_{1})) by the uniqueness of the very weak solution of (pCP).) Thus, we have

l1(tn+τ1+τ2)lw(τ2)<l,n,l_{1}(t_{n}+\tau_{1}+\tau_{2})\to l_{w}(\tau_{2})<l^{\infty},\quad n\to\infty,

contradicts the definition of ll^{\infty}. Therefore, w(x,τ1)w(x,\tau_{1}) is a stationary solution of (pCP).

We continue to specify the possible shapes of the ω\omega-limits of vv. Denote by w(x)w(x) one of them. By the monotonicity in Lemma 3.6 and Corollary 3.7, we see that either w(x)0w(x)\equiv 0, or, it has maximums in [b,b][-b,b]. In the latter case, we claim that ww does not have positive minimums. Otherwise, by the equation pRPME, ww is symmetric with respect to each maximum point and each minimum point as long as it keeps positive, and so ww must be a positive periodic function, which is impossible by the monotonicity outside of [b,b][-b,b]. Therefore, the only possible choice for ww is that it has only maximum points in [b,b][-b,b] but has no positive minimum points, and so it is a Type II ground state solution of (pCP). More precisely, for some

bz1<z2<<zkb and L>0,-b\leq z_{1}<z_{2}<\cdots<z_{k}\leq b\mbox{\ \ and\ \ }L>0,

with zi+2Lzi+1(i=1,2,,k1)z_{i}+2L\leq z_{i+1}\ (i=1,2,\cdots,k-1), there holds

(3.15) w(x)=𝒱(x):=V(xz1)+V(xz2)++V(xzk),x,w(x)=\mathcal{V}(x):=V(x-z_{1})+V(x-z_{2})+\cdots+V(x-z_{k}),\quad x\in{\mathbb{R}},

where V(x)V(x) is the unique nonnegative compactly supported stationary solution of (pCP) as in Lemma 2.4.

Step 4. Convergence in case l,r<-l^{\infty},r^{\infty}<\infty. In the previous step we know that any ω\omega-limit w(x)w(x) of v(x,t)v(x,t) is either 0 or a Type II ground state solution as 𝒱\mathcal{V}. In the former case we actually have the convergence result:

v(,t)0 as t,v(\cdot,t)\to 0\mbox{\ \ as\ \ }t\to\infty,

since the ω\omega-limit set of vv is connected. In the latter case, we now improve the conclusion in Step 3 to a convergence one: v(x,t)𝒱(x)(t)v(x,t)\to\mathcal{V}(x)\ (t\to\infty), that is, the shifts {z1,z2,,zk}\{z_{1},z_{2},\cdots,z_{k}\} in (3.15) is unique.

We first prove the following claim.

Claim 1. There exists T1>0T_{1}>0 such that v(,t)v(\cdot,t) has exactly kk maximums larger than Θ\Theta for tT1t\geq T_{1}, where Θ\Theta is the largest positive zero of gg in (0,Θ1)(0,\Theta_{1}) as in (2.17).

First, we show that v(,t)Θv(\cdot,t)-\Theta has exactly 2k2k non-degenerate zeros for large time. Using the standard zero number argument for v(,t)Θv(\cdot,t)-\Theta we know that, for some T2>0T_{2}>0, v(,t)Θv(\cdot,t)-\Theta has fixed even number of non-degenerate zeros when t>T2t>T_{2}. Since 𝒱\mathcal{V} defined by (3.15) is one ω\omega-limit of vv, we see that the fixed number must be 2k2k. Denote them by

y1(t)<y2(t)<<y2k1(t)<y2k(t).y_{1}(t)<y_{2}(t)<\cdots<y_{2k-1}(t)<y_{2k}(t).

Next, we show that vv has exactly one maximum zi(t)z_{i}(t) in (y2i1(t),y2i(t))(y_{2i-1}(t),y_{2i}(t)) for all large tt and each i=1,2,,ki=1,2,\cdots,k. We only prove the conclusion in [y1(t),y2(t)][y_{1}(t),y_{2}(t)]. In this interval ζ(x,t):=vx(x,t)\zeta(x,t):=v_{x}(x,t) satisfies

(3.16) {ζt=(m1)vζxx+(m+1)ζζx+g(v)ζ,y1(t)<x<y2(t),t>T2,ζ(y1(t),t)>0>ζ(y2(t),t),t>T2.\left\{\begin{array}[]{ll}\zeta_{t}=(m-1)v\zeta_{xx}+(m+1)\zeta\zeta_{x}+g^{\prime}(v)\zeta,&y_{1}(t)<x<y_{2}(t),t>T_{2},\\ \zeta(y_{1}(t),t)>0>\zeta(y_{2}(t),t),&t>T_{2}.\end{array}\right.

Since v(x,t)>Θv(x,t)>\Theta for y1(t)<x<y2(t),t>T2y_{1}(t)<x<y_{2}(t),t>T_{2}, the equation is a uniform parabolic one, and so we can use the classical zero number argument to conclude that, for some T3>T2T_{3}>T_{2}, ζ(,t)\zeta(\cdot,t) has fixed number of zeros in (y1(t),y2(t))(y_{1}(t),y_{2}(t)). By our assumption, for some increasing time sequence {tn}\{t_{n}\}, we have

[y1(tn),y2(tn)](L+z1,L+z1),[y_{1}(t_{n}),y_{2}(t_{n})]\subset(-L+z_{1},L+z_{1}),

and v(x,tn)𝒱(x)(n)v(x,t_{n})\to\mathcal{V}(x)\ (n\to\infty) in the topology Cloc2,1((L+z1,L+z1))C^{2,1}_{loc}((-L+z_{1},L+z_{1})). This implies that v(x,tn)v(x,t_{n}) has exactly one maximum point z1(tn)(y1(tn),y2(tn))z_{1}(t_{n})\in(y_{1}(t_{n}),y_{2}(t_{n})) for large nn. This prove Claim 1.

By v(x,tn)𝒱(x)v(x,t_{n})\to\mathcal{V}(x) again, for some large T4>T3T_{4}>T_{3}, v(x,T4)>Θv(x,T_{4})>\Theta near z1z_{1}. For simplicity, when tT4t\geq T_{4}, denote by l(t),r(t)l(t),\ r(t) the two free boundaries of the support of v(,t)v(\cdot,t) containing z1z_{1}. They are nothing but l1(t)ln1(t)l_{1}(t)\equiv l_{n_{1}}(t) and rn2(t)r_{n_{2}}(t) in (3.11). Then l(t)-l(t) and r(t)r(t) are non-decreasing continuous functions.

We now use Claim 1 to prove the convergence result, that is, the shift z1z_{1} is unique. By contradiction we assume that vv has another ω\omega-limit point

𝒱ˇ(x):=V(xzˇ1)+V(xzˇ2)++V(xzˇk),\check{\mathcal{V}}(x):=V(x-\check{z}_{1})+V(x-\check{z}_{2})+\cdots+V(x-\check{z}_{k}),

and, without loss of generality, assume z1<zˇ1z_{1}<\check{z}_{1}. Choose x0,x1,x2[z1,zˇ1]x_{0},x_{1},x_{2}\in[z_{1},\check{z}_{1}] such that

x0l(T4)+r(T4)2,x0x1=x2x0>0,x_{0}\not=\frac{l(T_{4})+r(T_{4})}{2},\quad x_{0}-x_{1}=x_{2}-x_{0}>0,

that is, x0x_{0} is the center of [x1,x2][z1,zˇ1][x_{1},x_{2}]\subset[z_{1},\check{z}_{1}] but not that of [l(T4),r(T4)][l(T_{4}),r(T_{4})]. Then, by the connectedness of the ω\omega-limit set ω(v)\omega(v), each V(xxi)(i=0,1,2)V(x-x_{i})\ (i=0,1,2) is a part of an ω\omega-limit of vv. Define

l~(t):=2x0r(t),r~(t):=2x0l(t),v~(x,t):=v(2x0x,t) for x,tT4.\tilde{l}(t):=2x_{0}-r(t),\quad\tilde{r}(t):=2x_{0}-l(t),\quad\tilde{v}(x,t):=v(2x_{0}-x,t)\mbox{\ \ for\ \ }x\in{\mathbb{R}},\ t\geq T_{4}.

Then l~(t)\tilde{l}(t) and r~(t)\tilde{r}(t) are free boundaries of v~(,t)\tilde{v}(\cdot,t), v~(x,t)>0\tilde{v}(x,t)>0 for x(l~(t),r~(t)),tT4x\in(\tilde{l}(t),\tilde{r}(t)),\ t\geq T_{4}, and

v~t=(m1)v~v~xx+v~x2+g(v~),x,tT4.\tilde{v}_{t}=(m-1)\tilde{v}\tilde{v}_{xx}+\tilde{v}_{x}^{2}+g(\tilde{v}),\quad x\in{\mathbb{R}},t\geq T_{4}.

It is easily seen that l~(t)<r(t),r~(t)>l(t)\tilde{l}(t)<r(t),\ \tilde{r}(t)>l(t), and so v(,t)v(\cdot,t) and v~(,t)\tilde{v}(\cdot,t) have non-empty common domain [l^(t),r^(t)][\hat{l}(t),\hat{r}(t)] with

l^(t):=max{l(t),l~(t)}<x0<r^(t):=min{r(t),r~(t)},tT4.\hat{l}(t):=\max\{l(t),\tilde{l}(t)\}\ <x_{0}<\ \hat{r}(t):=\min\{r(t),\tilde{r}(t)\},\quad t\geq T_{4}.

By the choice of x0x_{0} we see that

l(T4)l~(T4)=2x0r(T4),r(T4)r~(T4)=2x0l(T4).l(T_{4})\not=\tilde{l}(T_{4})=2x_{0}-r(T_{4}),\quad r(T_{4})\not=\tilde{r}(T_{4})=2x_{0}-l(T_{4}).

Hence, we can use Lemma 3.3 which is valid in open intervals to conclude that the number of the zeros of η(,t):=v(,t)v~(,t)\eta(\cdot,t):=v(\cdot,t)-\tilde{v}(\cdot,t) in the open interval J0(t):=(l^(t),r^(t))J_{0}(t):=(\hat{l}(t),\hat{r}(t)) is finite, decreasing in t>T4t>T_{4}, and strictly decreasing when η(,t)\eta(\cdot,t) has degenerate zeros in J0(t)J_{0}(t). Consequently, for some large time T5>T4T_{5}>T_{4},

(3.17) x0 is not a degenerate zero of η(,t):=v(,t)v~(,t) for t>T5.x_{0}\mbox{ is not a degenerate zero of }\eta(\cdot,t):=v(\cdot,t)-\tilde{v}(\cdot,t)\mbox{ for }t>T_{5}.

On the other hand, by the previous assumption, there exist two time sequences {sn(1)}\{s^{(1)}_{n}\} and {sn(2)}\{s^{(2)}_{n}\} with

(3.18) <sn(1)<sn(2)<sn+1(1)<sn+1(2)<,n=1,2,\cdots<s^{(1)}_{n}<s^{(2)}_{n}<s^{(1)}_{n+1}<s^{(2)}_{n+1}<\cdots,\quad n=1,2,\cdots

such that, for i=1,2i=1,2,

v(x,sn(i))V(xxi)L([L+xi,L+xi])0 as n.\|v(x,s^{(i)}_{n})-V(x-x_{i})\|_{L^{\infty}([-L+x_{i},L+x_{i}])}\to 0\mbox{\ \ as\ \ }n\to\infty.

Therefore, for all large nn we have

z1(sn(i))xi,i=1,2.z_{1}(s^{(i)}_{n})\approx x_{i},\quad i=1,2.

By the continuity of z1(t)z_{1}(t), there exists sn(sn(1),sn(2))s_{n}\in(s^{(1)}_{n},s^{(2)}_{n}) such that z1(sn)=x0z_{1}(s_{n})=x_{0}. Thus, for large nn we have

η(x0,sn)=v(x0,sn)v~(x0,sn)=0,ηx(x0,sn)=vx(x0,sn)v~x(x0,sn)=0,\eta(x_{0},s_{n})=v(x_{0},s_{n})-\tilde{v}(x_{0},s_{n})=0,\quad\eta_{x}(x_{0},s_{n})=v_{x}(x_{0},s_{n})-\tilde{v}_{x}(x_{0},s_{n})=0,

contradicting (3.17). This proves the uniqueness of z1z_{1}, and so the convergence in this step is proved.

Step 5. To show that any ω\omega-limit is a stationary solution of pRPME in case l,r=-l^{\infty},r^{\infty}=\infty. This is Case 3 in Step 2. For large tt, say tT6t\geq T_{6}, v(,t)v(\cdot,t) has exactly one connected support. Denote, for simplicity, its left and right free boundaries by l(t)l(t) and r(t)r(t), respectively. Due to the unboundedness of l(t)l(t) and r(t)r(t), the Lyapunov functional as in (3.14) is no longer a suitable tool to derive the quasi-convergence result. Instead, we will use the zero number argument to continue our argument.

Using the limit function w(x,t)w(x,t) in Step 1, we want to show that, for each tt\in{\mathbb{R}}, w(x,t)w(x,t) is a stationary solution of (pCP). It is sufficient to prove that w(x,0)w(x,0) is so. When w(x,0)0w(x,0)\equiv 0 in xx\in{\mathbb{R}}, there is nothing left to proof. So, we assume without loss of generality that w(0,0)>0w(0,0)>0. We now construct directly a positive stationary solution as the following:

(3.19) (m1)v¯v¯xx+v¯x2+g(v¯)=0,v¯(0)=w(0,0)>0,v¯(0)=wx(0,0).(m-1)\bar{v}\bar{v}_{xx}+\bar{v}^{2}_{x}+g(\bar{v})=0,\quad\bar{v}(0)=w(0,0)>0,\ \ \bar{v}^{\prime}(0)=w_{x}(0,0).

As long as v¯>0\bar{v}>0 the equation is a non-degenerate second order ordinary differential equation, and so the existence and uniqueness of the solution of (3.19) follow from the standard theory. Noticing from the equation we see that v¯\bar{v} is symmetric with respect to any positive critical point. Hence, we have the following possibilities:

  1. (a).

    v¯(x)>0\bar{v}(x)>0 for all xx\in{\mathbb{R}};

  2. (b).

    v¯(x)>0\bar{v}(x)>0 and v¯(x)>0\bar{v}^{\prime}(x)>0 in (l0,)(l_{0},\infty) for some l0(,0)l_{0}\in(-\infty,0), and v¯(l0)=0\bar{v}(l_{0})=0;

  3. (c).

    v¯(x)>0\bar{v}(x)>0 and v¯(x)<0\bar{v}^{\prime}(x)<0 in (,r0)(-\infty,r_{0}) for some r0(0,)r_{0}\in(0,\infty), and v¯(r0)=0\bar{v}(r_{0})=0;

  4. (d).

    v¯(x)>0\bar{v}(x)>0 for x(l0,r0)x\in(l_{0},r_{0}) for some l0<0<r0l_{0}<0<r_{0}, v¯(l0)=v¯(r0)=0\bar{v}(l_{0})=\bar{v}(r_{0})=0 and it has a unique maximum point z0:=l0+r02z_{0}:=\frac{l_{0}+r_{0}}{2}.

In this step we will prove w(x,0)v¯(x)w(x,0)\equiv\bar{v}(x) in spt[v¯]{\rm spt}[\bar{v}].

(a). Set

η(x,t):=v(x,t)v¯(x),x,tT6.\eta(x,t):=v(x,t)-\bar{v}(x),\quad x\in{\mathbb{R}},\ t\geq T_{6}.

In the current case (a) we have

(3.20) η(l(t),t)<0,η(r(t),t)<0,tT6,\eta(l(t),t)<0,\quad\eta(r(t),t)<0,\quad t\geq T_{6},

and

(3.21) ηt=a(x,t)ηxx+b(x,t)ηx+c(x,t)η,xJ0(t):=(l(t),r(t)),tT6,\eta_{t}=a(x,t)\eta_{xx}+b(x,t)\eta_{x}+c(x,t)\eta,\quad x\in J_{0}(t):=(l(t),r(t)),\ t\geq T_{6},

with

a(x,t):=(m1)v(x,t),b(x,t):=vx(x,t)+v¯x(x),a(x,t):=(m-1)v(x,t),\quad b(x,t):=v_{x}(x,t)+\bar{v}_{x}(x),

and

c(x,t):=(m1)v¯xx(x)+{g(v(x,t))g(v¯(x))v(x,t)v¯(x),vv¯,0,v=v¯.c(x,t):=(m-1)\bar{v}_{xx}(x)+\left\{\begin{array}[]{ll}\frac{g(v(x,t))-g(\bar{v}(x))}{v(x,t)-\bar{v}(x)},&v\not=\bar{v},\\ 0,&v=\bar{v}.\end{array}\right.

We will use Proposition 3.1 to study the zero number of η(,t)\eta(\cdot,t) in any bounded time interval (t1,t2)(T6,)(t_{1},t_{2})\subset(T_{6},\infty). For this purpose, we should require that a,b,ca,b,c satisfy the assumption (3.2). This is obviously not true in the domain

E1:={(x,t)xJ0(t),t1<t<t2},E_{1}:=\{(x,t)\mid x\in J_{0}(t),\ t_{1}<t<t_{2}\},

since a1=[(m1)v]1a^{-1}=[(m-1)v]^{-1} is not bounded near the boundaries of E1E_{1}. This difficulty will be solved below by cutting J0(t)J_{0}(t) a little bit near its boundaries. Since l(t)l(t) and r(t)r(t) are continuous and bounded functions in [t1,t2][t_{1},t_{2}], there exists ρ1>0\rho_{1}>0 small such that

η(l(t),t)=v¯(l(t))ρ1,η(r(t),t)=v¯(r(t))ρ1,t[t1,t2].\eta({l}(t),t)=-\bar{v}(l(t))\leq-\rho_{1},\quad\eta({r}(t),t)=-\bar{v}(r(t))\leq-\rho_{1},\quad t\in[t_{1},t_{2}].

By the uniform continuity of vv, there exists ε1>0\varepsilon_{1}>0 small such that

η(x,t)<0,l(t)xl(t)+ε1 or r(t)+ε1xr(t),t[t1,t2]\eta(x,t)<0,\quad l(t)\leq x\leq l(t)+\varepsilon_{1}\mbox{ or }r(t)+\varepsilon_{1}\leq x\leq r(t),\ t\in[t_{1},t_{2}]

and, for some small ρ2>0\rho_{2}>0, there holds

v(x,t)ρ2 in E2:={(x,t)l(t)+ε1xr(t)ε1,t1tt2}.v(x,t)\geq\rho_{2}\mbox{ in }E_{2}:=\{(x,t)\mid l(t)+\varepsilon_{1}\leq x\leq r(t)-\varepsilon_{1},\ t_{1}\leq t\leq t_{2}\}.

Now we see that a,b,ca,b,c satisfy the conditions in (3.2) in the domain E2E_{2}. Hence the classical zero number diminishing properties as in Proposition 3.1 hold. Since ε1>0,t1T6>0\varepsilon_{1}>0,t_{1}-T_{6}>0 can be chosen as small as possible and t2t_{2} can be chosen as large as possible, we actually have the same zero number diminishing properties in E1E_{1} with (t1,t2)=(T6,)(t_{1},t_{2})=(T_{6},\infty). As a consequence, η\eta has only degenerate zeros (not on the boundaries l(t)l(t) and r(t)r(t)) in finite time, that is, there exists T7>T6T_{7}>T_{6} such that η(,t)\eta(\cdot,t) not longer has degenerate zeros for tT7t\geq T_{7}. Thus, for any t[1,1]t\in[-1,1] and all large nn, η(x,tn+t)=v(x,tn+t)v¯(x)\eta(x,t_{n}+t)=v(x,t_{n}+t)-\bar{v}(x) has no degenerate zeros in J0(tn+t)J_{0}(t_{n}+t). This implies that (see, for example [13, Lemma 2.6]) its limit w(x,t)v¯(x)w(x,t)-\bar{v}(x) either is 0 identically, or has no degenerate zeros in the spatial domain where w(,t)>0w(\cdot,t)>0. The latter, however, contradicts the initial conditions in (3.19). This proves that w(x,0)v¯(x)w(x,0)\equiv\bar{v}(x) in spt[v¯]{\rm spt}[\bar{v}].

(b). Now we assume v¯\bar{v} is a stationary solution as in case (b). Since l(t)(t)l(t)\to-\infty\ (t\to\infty), there exists T8>T6T_{8}>T_{6} such that l(t)<l0l(t)<l_{0} for tT8t\geq T_{8}. Hence, we should consider η\eta in

E3:={(x,t)l0<x<r(t),tT8}.E_{3}:=\{(x,t)\mid l_{0}<x<r(t),\ t\geq T_{8}\}.

On its left boundary {x=l0,t>T8}\{x=l_{0},\ t>T_{8}\}, we have η>0\eta>0. By continuity, for any bounded time interval (t1,t2)(T8,)(t_{1},t_{2})\subset(T_{8},\infty), there exists ε2>0\varepsilon_{2}>0 and ρ3=ρ3(t1,t2,l0)>0\rho_{3}=\rho_{3}(t_{1},t_{2},l_{0})>0 such that

η(x,t)>ρ3,l0xl0+ε2,t(t1,t2).\eta(x,t)>\rho_{3},\quad l_{0}\leq x\leq l_{0}+\varepsilon_{2},\ t\in(t_{1},t_{2}).

Furthermore, the term v¯xx(l0+ε2)\bar{v}_{xx}(l_{0}+\varepsilon_{2}) in the coefficient cc of (3.21) is also bounded, though v¯xx(l0)\bar{v}_{xx}(l_{0}) maybe not. Thus, we can use the zero number diminishing properties in the domain {(x,t)l0+ε2<x<r(t),t1<t<t2}\{(x,t)\mid l_{0}+\varepsilon_{2}<x<r(t),t_{1}<t<t_{2}\}. By the arbitrariness of ε2\varepsilon_{2} and (t1,t2)(t_{1},t_{2}), we see that the zero number diminishing properties hold actually in the domain E3E_{3}. The rest proof is similar as that in (a). This proves the quasi-convergence in case (b).

The proof for w(x,0)v¯(xspt[v¯])w(x,0)\equiv\bar{v}\ (x\in{\rm spt}[\bar{v}]) in cases (c) and (d) are similar as that in case (b).

Step 6. To show that any ω\omega-limit of vv is a positive stationary solution of (pCP) in the case l,r=-l^{\infty},r^{\infty}=\infty. From the previous step we see that, in the case l,r=-l^{\infty},r^{\infty}=\infty, any ω\omega-limit of vv must be a stationary solution of the equation pRPME as in (a)-(d). By the monotonicity in Corollary 3.7, the cases (b) and (c) are impossible. We now prove that the case (d) is also impossible.

1). By the monotonicity of vv and ww outside of [b,b][-b,b], we see that the maximum point z0:=l0+r02z_{0}:=\frac{l_{0}+r_{0}}{2} of v¯(x)w(x,0)|spt[v¯]\bar{v}(x)\equiv w(x,0)|_{{\rm spt}[\bar{v}]} lies in [b,b][-b,b], and so, with :=r0l02\ell:=\frac{r_{0}-l_{0}}{2}, we have

bl0<z0<r0b+.-b-\ell\leq l_{0}<z_{0}<r_{0}\leq b+\ell.

2). We show that v(3b2,t)<v¯(z0)v(-3b-2\ell,t)<\bar{v}(z_{0}) for all t>0t>0. Assume by contradiction that the reversed inequality holds at some time t1>0t_{1}>0. Then by the monotonicity of vv in (l(t1),b)(l(t_{1}),-b) we have

v(x,t1)v¯(x+z0+3b+),x[3b2,3b].v(x,t_{1})\geq\bar{v}(x+z_{0}+3b+\ell),\quad x\in[-3b-2\ell,-3b].

Since v¯(x+z0+3b+)\bar{v}(x+z_{0}+3b+\ell) is a time-independent weak subsolution of the problem (pCP), by Lemma 2.3 we have

v(x,t)v¯(x+z0+3b+),x,tt1,v(x,t)\geq\bar{v}(x+z_{0}+3b+\ell),\quad x\in{\mathbb{R}},\ t\geq t_{1},

and so

w(x,0)v¯(x+z0+3b+),x.w(x,0)\geq\bar{v}(x+z_{0}+3b+\ell),\quad x\in{\mathbb{R}}.

By the monotonicity of w(x,0)w(x,0) in [3b2,b][-3b-2\ell,-b] we have

(3.22) w(x,0)v¯(z0),x[3b,b].w(x,0)\geq\bar{v}(z_{0}),\quad x\in[-3b-\ell,-b].

On the other hand, w(x,0)=0w(x,0)=0 at l0[b,b+]l_{0}\in[-b-\ell,b+\ell]. So w(x,0)w(x,0) has at least one maximum point x1[b,l0)x_{1}\in[-b,l_{0}). Without loss of generality, assume x1x_{1} is the smallest one of such points. Then, as a stationary solution of pRPME, w(x,0)w(x,0) is symmetric with respect to x1x_{1}. This implies by (3.22) that

w(x,0)v¯(z0),x[3b,2x1+3b+][b,b+],w(x,0)\geq\bar{v}(z_{0}),\quad x\in[-3b-\ell,2x_{1}+3b+\ell]\supset[-b-\ell,b+\ell],

contradicts 1).

3). If v¯(x)w(x,0)|spt[v¯]\bar{v}(x)\equiv w(x,0)|_{{\rm spt}[\bar{v}]} is not only a stationary solution of the equation pRPME, but also a stationary solution of (pCP), then we can derive a contradiction as the following. Choose tt^{\prime} large such that l(t)<3b4l(t^{\prime})<-3b-4\ell, and choose LL^{\prime} such that

L+l0<l(t)<L+r0-L^{\prime}+l_{0}<l(t^{\prime})<-L^{\prime}+r_{0}

and v¯(x+L)\bar{v}(x+L^{\prime}) has exactly one intersection point with v(x,t)v(x,t^{\prime}). Then using the intersection number properties Lemma 3.3 and using the fact in 2) we conclude that l(t)l(t) will never pass across the left boundary L+l0-L^{\prime}+l_{0} of v¯(x+L)\bar{v}(x+L^{\prime}). This contradicts the assumption l=l^{\infty}=-\infty.

4). We then consider the case where v¯(x)w(x,0)|spt[v¯]\bar{v}(x)\equiv w(x,0)|_{{\rm spt}[\bar{v}]} is just a stationary solution of the equation pRPME, but not a stationary one of (pCP). Denote by v(x,t;v¯)v(x,t;\bar{v}) the solution of (pCP) with initial data v¯\bar{v}, and denote its free boundaries by l¯(t)\bar{l}(t) and r¯(t)\bar{r}(t). By Lemma 2.3, the waiting times t(l0)=t(r0)=0t^{*}(l_{0})=t^{*}(r_{0})=0. Hence, l¯(t)-\bar{l}(t) and r¯(t)\bar{r}(t) are strictly increasing in t>0t>0, and

v(x,t;v¯)|[l¯(t),r¯(t)]w(x,t)|[l¯(t),r¯(t)],0<t1.v(x,t;\bar{v})|_{[\bar{l}(t),\bar{r}(t)]}\equiv w(x,t)|_{[\bar{l}(t),\bar{r}(t)]},\quad 0<t\ll 1.

Since v¯\bar{v} is a (stationary) subsolution of (pCP), by Lemma 2.3 we have v(x,t;v¯)>v¯(x)v(x,t;\bar{v})>\bar{v}(x) on the support of v¯\bar{v}, and so v(x,t;v¯)v(x,t;\bar{v}) is strictly increasing in tt. Therefore, for any small τ>0\tau>0, we have

l¯(2τ)<l¯(τ)<l¯(0)=l0<r0=r¯(0)<r¯(τ)<r¯(2τ),\bar{l}(2\tau)<\bar{l}(\tau)<\bar{l}(0)=l_{0}<r_{0}=\bar{r}(0)<\bar{r}(\tau)<\bar{r}(2\tau),

and

w(x,0)<w(x,τ) for x[l0,r0],w(x,τ)<w(x,2τ) for x[l¯(τ),r¯(τ)].w(x,0)<w(x,\tau)\mbox{ for }x\in[l_{0},r_{0}],\qquad w(x,\tau)<w(x,2\tau)\mbox{ for }x\in[\bar{l}(\tau),\bar{r}(\tau)].

By Step 1, v(x,tn+2τ)w(x,2τ)v(x,t_{n}+2\tau)\to w(x,2\tau) as nn\to\infty, and so for some large n0n_{0} we have

v(x,tn0+2τ)w(x,τ),x[l¯(τ),r¯(τ)].v(x,t_{n_{0}}+2\tau)\geq w(x,\tau),\quad x\in[\bar{l}(\tau),\bar{r}(\tau)].

From the previous step we know that w(x,τ)w(x,\tau) is a stationary solution of pRPME in [l¯(τ),r¯(τ)][\bar{l}(\tau),\bar{r}(\tau)]. Hence, w(x,τ)|[l¯(τ),r¯(τ)]w(x,\tau)|_{[\bar{l}(\tau),\bar{r}(\tau)]} is a time-independent weak subsolution of (pCP). By comparison we have

v(x,t)w(x,τ),x[l¯(τ),r¯(τ)],t>tn0+2τ.v(x,t)\geq w(x,\tau),\quad x\in[\bar{l}(\tau),\bar{r}(\tau)],\ t>t_{n_{0}}+2\tau.

This contradicts the fact that w(x,0)w(x,0) is an ω\omega-limit of v(x,t)v(x,t).

Consequently, the case (d) is also impossible. In other words, the only possible case is (a) in Step 5. We then have the following claim.

Claim 2. When l,r=-l^{\infty},r^{\infty}=\infty, any ω\omega-limit w(x)w(x) of v(x,t;u0)v(x,t;u_{0}) is a stationary solution of pRPME which is positive in {\mathbb{R}}, and so is a stationary solution of the Cauchy problem (pCP).

Step 7. Convergence in case l,r=-l^{\infty},r^{\infty}=\infty. By Claim 2, the convergence v(x,tn;u0)w(x)v(x,t_{n};u_{0})\to w(x) in Step 1 holds actually in Cloc2()C^{2}_{loc}({\mathbb{R}}) topology. Therefore, the solution v(,t)v(\cdot,t) is a classical one in (ln1(t),r2n(t))(l_{n_{1}}(t),r_{2n}(t)). Then, using the classical zero number diminishing properties and using similar arguments as proving [11, Theorem 1.1], [12, Theorem 1.1] (see also [13, Theorem 1.1]) we conclude that v(,t)v(\cdot,t) converges as tt\to\infty to either a nonnegative zero of gg, or

v(,t)V0(z0)>0,v(\cdot,t)\to V_{0}(\cdot-z_{0})>0,

where z0[b,b]z_{0}\in[-b,b] and V0V_{0} is an evenly decreasing positive stationary solution of pRPME, that is,

U0(x):=(m1mV0(x))1m1>0,xU_{0}(x):=\left(\frac{m-1}{m}V_{0}(x)\right)^{\frac{1}{m-1}}>0,\quad x\in{\mathbb{R}}

is a Type I ground state solution of (CP) as given in Section 1.

This completes the proof of Theorem 1.1. \Box

4. Monostable RPME

In this section we study the hair-trigger effect for the solutions of (CP) or (pCP) with monostable type of reaction.

4.1. Stationary solutions

A stationary solution u=q(x)u=q(x) of (CP) satisfies

(4.1) (qm)′′+f(q)=0,xJ.(q^{m})^{\prime\prime}+f(q)=0,\quad x\in J\subset{\mathbb{R}}.

We will use the phase portrait to present all the nonnegative stationary solutions that we are interested in. The equation (4.1) can be rewritten as a pair of first order equations

(4.2) {p=(qm)=mqm1q,p=f(q),\left\{\begin{array}[]{l}p=(q^{m})^{\prime}=mq^{m-1}q^{\prime},\\ p^{\prime}=-f(q),\end{array}\right.

or the first order equation

(4.3) pdp=mqm1f(q)dq.pdp=-mq^{m-1}f(q)dq.

Its first integral is

(4.4) p2=C2m0qrm1f(r)𝑑r,p^{2}=C-2m\int_{0}^{q}r^{m-1}f(r)dr,

for suitable choice of CC. We consider only the trajectories in the region {q0}\{q\geq 0\} which correspond to nonnegative solutions.

Case A. The points (0,0)(0,0) and (1,0)(1,0) are singular ones on the (q,p)(q,p)-phase plane for the system (4.2). Hence, they correspond to constant stationary solutions:

Uq(x)q,x,U_{q}(x)\equiv q,\quad x\in{\mathbb{R}},

for q=0, 1q=0,\ 1. We will show below that U1U_{1} is the only possible choice for the ω\omega-limit of the solution uu of (CP).

Case B. When C=C0:=2m01rm1f(r)𝑑rC=C_{0}:=2m\int_{0}^{1}r^{m-1}f(r)dr, the trajectories on the phase plane are two curves, whose functions are

p=P0(q):=(2mq1rm1f(r)𝑑r)1/2 and p=P0(q),0q<1.p=P_{0}(q):=\left(2m\int_{q}^{1}r^{m-1}f(r)dr\right)^{1/2}\mbox{\ \ and\ \ }p=-P_{0}(q),\quad 0\leq q<1.

They connect regular points Q±:=(0,±C0)Q_{\pm}:=(0,\pm\sqrt{C_{0}}) to the singular point (1,0)(1,0). Denote the corresponding stationary solution by U±(x)U^{\pm}(x). Then

U+(l0)=0,U+(x)>0 for x>l0, and U+(x)1 as x+;U^{+}(l_{0})=0,\quad U^{+}(x)>0\mbox{ for }x>l_{0},\mbox{\ \ and\ \ }U^{+}(x)\nearrow 1\mbox{ as }x\to+\infty;
U(r0)=0,U(x)>0 for x<r0, and U(x)1 as x,U^{-}(r_{0})=0,\quad U^{-}(x)>0\mbox{ for }x<r_{0},\mbox{\ \ and\ \ }U^{-}(x)\nearrow 1\mbox{ as }x\to-\infty,

where l0l_{0} is the left boundary of U+U^{+} and r0r_{0} is the right boundary of UU^{-}.

Case C. When C>C0C>C_{0}, the trajectories start at (0,±C)(0,\pm\sqrt{C}), go beyond the line q=1q=1 and so the the corresponding stationary solutions are unbounded ones, which will not be used below.

Case D. When 0<C<C00<C<C_{0}, the trajectory p2+2m0qrm1f(r)𝑑r=Cp^{2}+2m\int_{0}^{q}r^{m-1}f(r)dr=C is a connected curve, symmetric with respect to the qq-axis, and connecting the points Q+:=(0,C)Q_{+}:=(0,\sqrt{C}), Q0:=(q0,0)Q_{0}:=(q_{0},0) and Q:=(0,C)Q_{-}:=(0,-\sqrt{C}) for a unique q0(0,1)q_{0}\in(0,1) which is defined as the following

C=2m0q0rm1f(r)𝑑r.C=2m\int_{0}^{q_{0}}r^{m-1}f(r)dr.

This trajectory corresponds to a stationary solution Uq0(x)U_{q_{0}}(x), under the normalized condition Uq0(0)=0U^{\prime}_{q_{0}}(0)=0, it satisfies

(4.5) Uq0(0)=q0,Uq0(±L0)=0 for some L0>0,Uq0(x)<0,Uq0(x)=Uq0(x) for x(0,L0).U_{q_{0}}(0)=q_{0},\quad U_{q_{0}}(\pm L_{0})=0\mbox{ for some }L_{0}>0,\quad U^{\prime}_{q_{0}}(x)<0,\ \ U_{q_{0}}(x)=U_{q_{0}}(-x)\mbox{ for }x\in(0,L_{0}).

We now study the order of the solution UU in Cases B (in this case, it is U+U^{+}), C or D near its left boundary. Denote the left boundary by l0l_{0}. By (4.4) we have

(Um)=(C2m0Urm1f(r)𝑑r)1/2C,xl0+0.(U^{m})^{\prime}=\left(C-2m\int_{0}^{U}r^{m-1}f(r)dr\right)^{1/2}\sim\sqrt{C},\quad x\to l_{0}+0.

So,

(4.6) Um(x)C(xl0),xl0+0.U^{m}(x)\sim\sqrt{C}(x-l_{0}),\quad x\to l_{0}+0.

If we use V:=mm1Um1V:=\frac{m}{m-1}U^{m-1} to denote the corresponding stationary solution of pRPME we have

(4.7) V(x)=mm1Cm12m(xl0)m1m,xl0+0.V(x)=\frac{m}{m-1}C^{\frac{m-1}{2m}}(x-l_{0})^{\frac{m-1}{m}},\quad x\to l_{0}+0.
Remark 4.1.

We point out that the solutions U±U^{\pm} in Case B and Uq0U_{q_{0}} in Case D are just stationary solutions of RPME but not stationary solutions of the Cauchy problem (CP). In fact, the expression in (4.7) and Lemma 2.1 imply that, when we choose V(x)V(x) as an initial data of the problem (pCP), the waiting time of its boundaries are 0. In particular, in the bistable RPME, Uq0U_{q_{0}} is not a Type II ground state solution.

We continue to specify the relationship between the support width 2L02L_{0} of Uq0U_{q_{0}} in Case D and its height q0q_{0}.

Lemma 4.2.

When ff is a monostable reaction term, there holds L00L_{0}\to 0 as q00+0q_{0}\to 0+0. In addition, if ff satisfies f(ρu)>ρmf(u)f(\rho u)>\rho^{m}f(u) for ρ,u(0,1)\rho,u\in(0,1), then L0L_{0} is strictly increasing in q0(0,1)q_{0}\in(0,1).

Proof.

By f(0)>0f^{\prime}(0)>0, there exists δ>0\delta>0 small such that

(4.8) f(u)π24ε2um,u[0,δ).f(u)\geq\frac{\pi^{2}}{4\varepsilon^{2}}u^{m},\quad u\in[0,\delta).

For any q0(0,δ)q_{0}\in(0,\delta), by (4.4) we have

p2=2mqq0rm1f(r)𝑑rπ24ε2(q02mq2m).p^{2}=2m\int_{q}^{q_{0}}r^{m-1}f(r)dr\geq\frac{\pi^{2}}{4\varepsilon^{2}}(q_{0}^{2m}-q^{2m}).

Therefore, on x[L0,0]x\in[-L_{0},0] we have

p=(qm)π2εq0m[1(qq0)2m]1/2.p=(q^{m})^{\prime}\geq\frac{\pi}{2\varepsilon}q_{0}^{m}\Big{[}1-\Big{(}\frac{q}{q_{0}}\Big{)}^{2m}\Big{]}^{1/2}.

Since q(L0)=0,q(0)=q0q(-L_{0})=0,q(0)=q_{0}, by integrating the above inequality over [L0,0][-L_{0},0] we have

π2εL001dr1r2=π2,{\frac{\pi}{2\varepsilon}}L_{0}\leq\int_{0}^{1}\frac{dr}{\sqrt{1-r^{2}}}=\frac{\pi}{2},

that is, L0εL_{0}\leq\varepsilon.

Denote v~:=um\tilde{v}:=u^{m} and f~(s):=f(s1m)\tilde{f}(s):=f(s^{\frac{1}{m}}), then the stationary problem

(um)′′+f(u)=0,u(x)>0 in (L0,L0),u(±L0)=0,(u^{m})^{\prime\prime}+f(u)=0,\ \ u(x)>0\mbox{ in }(-L_{0},L_{0}),\qquad u(\pm L_{0})=0,

is equivalent to

v~′′+f~(v~)=0,v~(x)>0 in (L0,L0),v~(±L0)=0.\tilde{v}^{\prime\prime}+\tilde{f}(\tilde{v})=0,\ \ \tilde{v}(x)>0\mbox{ in }(-L_{0},L_{0}),\qquad\tilde{v}(\pm L_{0})=0.

By our assumption f~\tilde{f} is a Fisher-KPP type of nonlinearity: f~(s)/s\tilde{f}(s)/s is decreasing in s(0,1)s\in(0,1). Hence v~(0)\tilde{v}(0) is strictly increasing in L0L_{0}. This proves the lemma. ∎

4.2. Hair-trigger effect

Now we can use the general convergence theorem to prove the hair-trigger effect easily.

Proof of Theorem 1.3. By Lemma 4.2, there exist sufficiently small q0>0q_{0}>0, small L0>0L_{0}>0 and some x0x_{0}\in{\mathbb{R}} such that

(4.9) u0(x)Uq0(xx0),x[L0+x0,L0+x0].u_{0}(x)\geq U_{q_{0}}(x-x_{0}),\quad x\in[-L_{0}+x_{0},L_{0}+x_{0}].

Since Uq0(xx0)U_{q_{0}}(x-x_{0}) is a time-independent very weak subsolution of (CP), by the comparison result in Lemma 2.3 (iv) we have

u(x,t)Uq0(xx0),x,t>0.u(x,t)\geq U_{q_{0}}(x-x_{0}),\quad x\in{\mathbb{R}},\ t>0.

In the monostable PME, the only stationary solution of (CP) larger than Uq0U_{q_{0}} is 11. So, by using the general convergence theorem we conclude that u(x,t)1u(x,t)\to 1 in Lloc()L^{\infty}_{loc}({\mathbb{R}}) topology. This proves the hair-trigger effect. \Box

Remark 4.3.

From the above proofs we see that (4.8) holds if limu0+0f(u)um=\lim\limits_{u\to 0+0}\frac{f(u)}{u^{m}}=\infty. So, the hair-trigger effect for the monostable RPME actually holds under this condition rather than the stronger one f(0)>0f^{\prime}(0)>0.

5. Combustion RPME

In this section we study the asymptotic behavior for the solutions of (CP) or (pCP) with combustion type of reaction: ff satisfies (fC).

5.1. Stationary solutions

On the nonnegative stationary solutions of the equation in (CP), we first have Case A - Case D types as in the monostable RPME. The difference is that: in Case A the solutions are UqqU_{q}\equiv q for q[0,θ]{1}q\in[0,\theta]\cup\{1\} in the combustion case (However, only U0,UθU_{0},U_{\theta} and U1U_{1} are possible choices for the ω\omega-limits of the solution uu of (CP).); in Case D, the compactly supported stationary solutions Uq0U_{q_{0}} exists only for q0(θ,1)q_{0}\in(\theta,1). The decay rates in (4.6) and (4.7) remain valid (in fact, they are true for x[l0,l0+C1/2θm]x\in[l_{0},l_{0}+C^{-1/2}\theta^{m}] due to f(u)0f(u)\equiv 0 in [0,θ][0,\theta]). As we mentioned in Remark 4.1, each Uq0U_{q_{0}} is just a stationary solution of the equation but not of the Cauchy problem (CP).

5.2. Asymptotic behavior of the solutions

Since the current problem has no Type I and Type II ground state solutions, by the general convergence results in Theorem 1.1 we know that u(,t)u(\cdot,t) converges as tt\to\infty to some constant solution in Case A.

Now we present a sufficient condition for the spreading phenomena: u(,t)1u(\cdot,t)\to 1 as tt\to\infty.

Lemma 5.1.

Assume ff is a combustion reaction, u0𝔛u_{0}\in\mathfrak{X} satisfies

u0(x)Uq0(x),x,u_{0}(x)\geq U_{q_{0}}(x),\quad x\in{\mathbb{R}},

for some stationary solution Uq0U_{q_{0}} in Case D. Then spreading happens for the solution uu of (CP).

Proof.

By the comparison result, the solution u(x,t)u(x,t) of (CP) remains above Uq0(x)U_{q_{0}}(x). Hence any ω\omega-limit of uu is also above Uq0(x)U_{q_{0}}(x). The only possible choice for such stationary solutions in Case A is nothing but U11U_{1}\equiv 1. Moreover, due to the finite propagation speed and the properties in Lemma 3.6, we know that spt[u(,t)]=[l(t),r(t)]{\rm spt}[u(\cdot,t)]=[l(t),r(t)] for large tt, hence the convergence u1u\to 1 holds in the sense of (1.7). ∎

Next we present some sufficient conditions for the vanishing phenomena: u(,t)0u(\cdot,t)\to 0 as tt\to\infty.

Lemma 5.2.

Assume ff is a combustion reaction, u0𝔛u_{0}\in\mathfrak{X} satisfies u0(x)θu_{0}(x)\leq\theta. Then vanishing happens in the topologies of L()L^{\infty}({\mathbb{R}}) for the solution uu of (CP).

Proof.

The proof is simple since uu actually solves the PME when u(x,t)θu(x,t)\leq\theta, and so we can take a ZKB solution as a supersolution to suppress uu to 0. ∎

Remark 5.3.

For any q(0,θ)q\in(0,\theta), UqqU_{q}\equiv q will not be an ω\omega-limit. In fact, if u(x,tn)q(n)u(x,t_{n})\to q\ (n\to\infty) in the Cloc()C_{loc}({\mathbb{R}}) topology, then, in the bounded interval [b,b][-b,b], u(x,T)<θu(x,T)<\theta for some large TT. This inequality also holds in [l(T),r(T)][l(T),r(T)] by the monotonicity outside of [b,b][-b,b]. Hence vanishing happens for uu by the above lemma, a contradiction.

To prove the sharpness of the transition: uθu\to\theta, we need the following lemmas.

Lemma 5.4.

Assume (F). For i=1,2i=1,2, assume ui0C()u_{i0}\in C({\mathbb{R}}) satisfies

ui0(x)>0 in (li(0),ri(0)),ui0(x)=0 for xli(0) and xri(0),u_{i0}(x)>0\mbox{ in }(l_{i}(0),r_{i}(0)),\quad u_{i0}(x)=0\mbox{ for }x\leq l_{i}(0)\mbox{ and }x\geq r_{i}(0),

and

u10(x)u20(x),x.u_{10}(x)\leq u_{20}(x),\quad x\in{\mathbb{R}}.

Let uiu_{i} be the solution of (CP) with initial data ui0u_{i0} and let li(t),ri(t)l_{i}(t),\ r_{i}(t) be its left and right free boundaries. Then there exist T0,τ0T\geq 0,\ \tau\geq 0 and ε0>0\varepsilon_{0}>0 such that

(5.1) u1(x,t)u2(x+ε,t+τ),x,tT,ε[0,ε0] or ε[ε0,0],u_{1}(x,t)\leq u_{2}(x+\varepsilon,t+\tau),\quad x\in{\mathbb{R}},\ t\geq T,\ \varepsilon\in[0,\varepsilon_{0}]\mbox{ or }\varepsilon\in[-\varepsilon_{0},0],

provided one of the following conditions holds:

  1. (i).

    r2(t)l2(t)r1(t)l1(t)r_{2}(t)-l_{2}(t)\not\equiv r_{1}(t)-l_{1}(t) for all t0t\geq 0;

  2. (ii).

    t(l2(0))t^{*}(l_{2}(0)) or t(r2(0))t^{*}(r_{2}(0)) is finite, and r2(t)l2(t)=r1(t)l1(t)r_{2}(t)-l_{2}(t)=r_{1}(t)-l_{1}(t) for all t0t\geq 0.

Proof.

By comparison we have

u1(x,t)u2(x,t) for x and t0,l2(t)l1(t)<r1(t)r2(t) for t0.u_{1}(x,t)\leq u_{2}(x,t)\mbox{ for }x\in{\mathbb{R}}\mbox{ and }t\geq 0,\quad l_{2}(t)\leq l_{1}(t)<r_{1}(t)\leq r_{2}(t)\mbox{ for }t\geq 0.

(i). In case r2(T0)l2(T0)>r1(T0)l1(T0)r_{2}(T_{0})-l_{2}(T_{0})>r_{1}(T_{0})-l_{1}(T_{0}) for some T00T_{0}\geq 0, we assume without loss of generality that l2(T0)<l1(T0)l_{2}(T_{0})<l_{1}(T_{0}), and by continuity assume T0>0T_{0}>0. Then using the strong maximum principle in {(x,t)l1(t)x<r1(t),T01tT0}\{(x,t)\mid l_{1}(t)\leq x<r_{1}(t),\ T_{0}-1\ll t\leq T_{0}\} we have

u1(x,T0)<u2(x,T0) for x[l1(T0),r1(T0)).u_{1}(x,T_{0})<u_{2}(x,T_{0})\mbox{ for }x\in[l_{1}(T_{0}),r_{1}(T_{0})).

So, there exists a small ε0>0\varepsilon_{0}>0 such that

u1(x,T0)<u2(x+ε,T0) for x[l1(T0),r1(T0)],ε[ε0,0).u_{1}(x,T_{0})<u_{2}(x+\varepsilon,T_{0})\mbox{ for }x\in[l_{1}(T_{0}),r_{1}(T_{0})],\quad\varepsilon\in[-\varepsilon_{0},0).

The conclusion then follows from the comparison principle for T=T0T=T_{0} and τ=0\tau=0.

(ii). Assume without loss of generality that t(l2(0))<t^{*}(l_{2}(0))<\infty. Given T1>t(l2(0))T_{1}>t^{*}(l_{2}(0)), then both u1(,t)u_{1}(\cdot,t) and u2(,t)u_{2}(\cdot,t) are strictly increasing in J1:=[x1,l1(0)]J_{1}:=[x_{1},l_{1}(0)], where x1:=l2(T1)=l1(T1)x_{1}:=l_{2}(T_{1})=l_{1}(T_{1}). By the monotonicity of l2(t)l_{2}(t) after the waiting time we see that l2(t)l_{2}(t) is strictly decreasing in tT1t\geq T_{1}. Hence, for any small t>0t>0 we have

l2(T1+t)<x1=l2(T1) and 0<u2(x1,T1+t)=u1(x2(t),T1)<u1(l1(0),T1),l_{2}(T_{1}+t)<x_{1}=l_{2}(T_{1})\mbox{\ \ and\ \ }0<u_{2}(x_{1},T_{1}+t)=u_{1}(x_{2}(t),T_{1})<u_{1}(l_{1}(0),T_{1}),

for some x2(t)[x1,l1(0)]x_{2}(t)\in[x_{1},l_{1}(0)]. Choose τ>0\tau>0 small, and denote

E1:={(x,t)l2(T1+t)xx1,0tτ} and E2:={(x,t)x1xx2(t),0tτ}.E_{1}:=\{(x,t)\mid l_{2}(T_{1}+t)\leq x\leq x_{1},0\leq t\leq\tau\}\mbox{\ \ and\ \ }E_{2}:=\{(x,t)\mid x_{1}\leq x\leq x_{2}(t),0\leq t\leq\tau\}.

By the monotonicity of u1(,T1)u_{1}(\cdot,T_{1}) and u2(,T1+τ)u_{2}(\cdot,T_{1}+\tau) in [x1,x2(τ)][x_{1},x_{2}(\tau)], we have

u1(x,T1)<u1(x2(τ),T1)=u2(x1,T1+τ)<u2(x,T1+τ),x[x1,x2(τ)].u_{1}(x,T_{1})<u_{1}(x_{2}(\tau),T_{1})=u_{2}(x_{1},T_{1}+\tau)<u_{2}(x,T_{1}+\tau),\quad x\in[x_{1},x_{2}(\tau)].

Thus there exists ε0>0\varepsilon_{0}>0 small such that

(5.2) u1(x,T1)<u2(x+ε,T1+τ),x[x1,x2(τ)],ε[ε0,0].u_{1}(x,T_{1})<u_{2}(x+\varepsilon,T_{1}+\tau),\quad x\in[x_{1},x_{2}(\tau)],\ \varepsilon\in[-\varepsilon_{0},0].

On the other hand, when τ>0\tau>0 and ε0\varepsilon_{0} are sufficiently small we have

u1(x,T1)<u2(x+ε,T1+τ),x[x2(τ),r1(T1)],ε[ε0,0].u_{1}(x,T_{1})<u_{2}(x+\varepsilon,T_{1}+\tau),\quad x\in[x_{2}(\tau),r_{1}(T_{1})],\ \varepsilon\in[-\varepsilon_{0},0].

Combining with (5.2) we have

u1(x,T1)<u2(x+ε,T1+τ),x[l1(T1),r1(T1)],ε[ε0,0].u_{1}(x,T_{1})<u_{2}(x+\varepsilon,T_{1}+\tau),\quad x\in[l_{1}(T_{1}),r_{1}(T_{1})],\ \varepsilon\in[-\varepsilon_{0},0].

Then the conclusion follows from the comparison principle. ∎

Lemma 5.5.

Under the assumptions of the previous lemma, if we further assume that u1(,t)θu_{1}(\cdot,t)\to\theta as tt\to\infty, then u2(,t)↛θu_{2}(\cdot,t)\not\to\theta as tt\to\infty.

Proof.

Since u1θu_{1}\to\theta, there exists TT+1T^{\prime}\geq T+1 for TT in (5.1) such that

(5.3) u1(x,t)<θ+δ2,x,tT1,u_{1}(x,t)<\theta+\frac{\delta}{2},\quad x\in{\mathbb{R}},\ t\geq T^{\prime}-1,

where δ\delta is given in (1.4). Now, for λ(0,1)\lambda\in(0,1), we define

wλ(x,t):=λ2u1(λmx,λ2t),x,tT.w^{\lambda}(x,t):=\lambda^{-2}\>\!u_{1}(\lambda^{m}x,\lambda^{2}t),\quad x\in{\mathbb{R}},\ t\geq T^{\prime}.

We choose λ0(0,1)\lambda_{0}\in(0,1) close enough to 11 so that, for every λ[λ0,1)\lambda\in[\lambda_{0},1),

(5.4) wλ(x,t)θ+δ,x,tT,w^{\lambda}(x,t)\leq\theta+\delta,\quad x\in{\mathbb{R}},\ t\geq T^{\prime},

and, by (5.1),

(5.5) wλ(x,T)u2(x+ε1,T),x[λml1(T),λmr1(T)]w^{\lambda}(x,T^{\prime})\leq u_{2}(x+\varepsilon_{1},T^{\prime}),\quad x\in[\lambda^{-m}l_{1}(T^{\prime}),\lambda^{-m}r_{1}(T^{\prime})]

for some ε1\varepsilon_{1}. Observe that wλw^{\lambda} satisfies the equation

wtλ=(wλ)xxm+f(λ2wλ),x,t>0.w^{\lambda}_{t}=(w^{\lambda})^{m}_{xx}+f(\lambda^{2}w^{\lambda}),\quad x\in{\mathbb{R}},\ t>0.

By (5.4) and the assumption in (fC) we have f(λ2wλ)f(wλ)f(\lambda^{2}w^{\lambda})\leq f(w^{\lambda}). (Here is the only place we use the assumption f(u)>0f^{\prime}(u)>0 in (θ,θ+δ](\theta,\theta+\delta].) Therefore in view of (5.5), we find that wλw^{\lambda} is a subsolution of (CP) for tTt\geq T^{\prime}. It follows that u2(x+ε1,t)wλ(x,t)u_{2}(x+\varepsilon_{1},t)\geq w^{\lambda}(x,t) for tTt\geq T^{\prime}. If u2θ(t)u_{2}\to\theta\ (t\to\infty), then by taking limit as tt\to\infty we conclude that θλ2θ\theta\geq\lambda^{-2}\theta, a contradiction. ∎


Proof of Theorem 1.5. From above lemmas we know that u(,t)u(\cdot,t) converges as tt\to\infty to 0, or θ\theta, or 11.

For the initial data ϕσ𝔛\phi_{\sigma}\in\mathfrak{X}, when σ>0\sigma>0 is small we have by (Φ3)(\Phi_{3}) that ϕσθ\phi_{\sigma}\leq\theta. Hence, vanishing happens for uσu_{\sigma} by Lemma 5.2. Denote

Σ0:={σ>0vanishing happens for uσ}.\Sigma_{0}:=\{\sigma>0\mid\mbox{vanishing happens for }u_{\sigma}\}.

Then Σ0\Sigma_{0} is not empty, and, by comparison, it is an interval. Since ϕσ\phi_{\sigma} depends on σ\sigma continuously, so does the solution uσu_{\sigma} (cf. [5]). Then, Σ0\Sigma_{0} is an open interval (0,σ)(0,\sigma_{*}) for some σ(0,]\sigma_{*}\in(0,\infty].

In case σ=\sigma_{*}=\infty (this is the complete vanishing case, as stated in Theorem 7.1), there is nothing left to prove.

The left case is σ(0,)\sigma_{*}\in(0,\infty). (This happens in particular when the condition in Lemma 5.1 holds.) In this case, the solution uσu_{\sigma_{*}} with critical initial data ϕσ\phi_{\sigma_{*}} converges to θ\theta or 11. We will show that uσu_{\sigma_{*}} is a transition solution. Otherwise, spreading happens for uσ1u_{\sigma_{*}}\to 1, that is,

σΣ1:={σσspreading happens for uσ}.\sigma_{*}\in\Sigma_{1}:=\{\sigma\geq\sigma_{*}\mid\mbox{spreading happens for }u_{\sigma}\}.

Then, for any q0(θ,1)q_{0}\in(\theta,1) and the corresponding stationary solution Uq0U_{q_{0}} in Case D in Subsection 5.1, there exists a sufficiently large TT such that

uσ(x,T)>Uq0(x),xspt[Uq0].u_{\sigma_{*}}(x,T)>U_{q_{0}}(x),\quad x\in{\rm spt}[U_{q_{0}}].

By the continuous dependence of uu on the initial data, for any σ\sigma satisfying 0<σσ10<\sigma_{*}-\sigma\ll 1, we also have

uσ(x,T)>Uq0(x),xspt[Uq0].u_{\sigma}(x,T)>U_{q_{0}}(x),\quad x\in{\rm spt}[U_{q_{0}}].

Consequently, the ω\omega-limits of uσu_{\sigma} are bigger than Uq0U_{q_{0}}, but these ω\omega-limits are nothing but 11, that is, spreading happens for uσu_{\sigma}. This implies that Σ1\Sigma_{1} is an open set and it contains σ\sigma_{*}, contradicts the definition of σ\sigma_{*}. Therefore, uσu_{\sigma_{*}} is a transition solution, that is, u(x,t)θu(x,t)\to\theta as tt\to\infty, and so both of its left and right boundaries tend to infinity. It follows from Lemmas 5.4 and 5.5 that uσ↛θu_{\sigma}\not\to\theta for any σ>σ\sigma>\sigma_{*}. Consequently, Σ1=(σ,)\Sigma_{1}=(\sigma_{*},\infty).

To finish the proof of Theorem 1.5, we only need to prove the asymptotic speeds of the free boundaries in (1.12) for the transition solution, which follows from Proposition 5.7 below.

This completes the proof of Theorem 1.5. \Box

At the end of this section we study the asymptotic speeds for the boundaries of the transition solution of (CP). In the current combustion case, the equation is PME ut=(um)xxu_{t}=(u^{m})_{xx} when uθu\leq\theta. So, it is convenient to compare the solutions of this equation, especially, the selfsimilar solutions, with the solution of (CP). For this purpose, we need a special selfsimilar solution which solves the following problem:

(5.6) {vt=(m1)vvxx+vx2,0<x<ρ(t),t>0,,v(0,t)=Θ:=mm1θm1,t>0,v(ρ(t),t)=0,t>0,ρ(t)=vx(ρ(t),t),t>0.\left\{\begin{array}[]{ll}v_{t}=(m-1)vv_{xx}+v_{x}^{2},&0<x<\rho(t),\ t>0,,\\ v(0,t)=\Theta:=\frac{m}{m-1}\theta^{m-1},&t>0,\\ v(\rho(t),t)=0,&t>0,\\ \rho^{\prime}(t)=-v_{x}(\rho(t),t),&t>0.\end{array}\right.
Lemma 5.6.

There exists y0y_{0} satisfying

(5.7) 0<y0<θm120<y_{0}<\theta^{\frac{m-1}{2}}

such that, with ρ(t)=2y0t\rho(t)=2y_{0}\sqrt{t}, the problem (5.6) has a selfsimilar solution v(x,t)=V(x2t)v(x,t)=V(\frac{x}{2\sqrt{t}}).

Proof.

The proof is divided into two steps.

Step 1. An auxiliary problem. We first consider the following initial value problem

(5.8) {ξ′′(y)=2y(ξ1m)(y)=2ymξ1mmξ(y),y>0,ξ(0)=θm,ξ(0)=2θm+12.\left\{\begin{array}[]{l}\displaystyle\xi^{\prime\prime}(y)=-2y\left(\xi^{\frac{1}{m}}\right)^{\prime}(y)=-2\frac{y}{m}\xi^{\frac{1-m}{m}}\xi^{\prime}(y),\quad y>0,\\ \displaystyle\xi(0)=\theta^{m},\ \xi^{\prime}(0)=-2\theta^{\frac{m+1}{2}}.\end{array}\right.

It is easily seen that, for any h(0,θm)h\in(0,\theta^{m}), the constant function ξh\xi\equiv h is a solution of the equation. Hence, it follows from ξ(0)<0\xi^{\prime}(0)<0 that ξ(y)<0\xi^{\prime}(y)<0 as long as ξ>0\xi>0. In other words, there exists y0(0,]y_{0}\in(0,\infty] such that

ξ(y)<0,ξ(y)>0,0y<y0.\xi^{\prime}(y)<0,\quad\xi(y)>0,\quad 0\leq y<y_{0}.

We now show that y0<y_{0}<\infty and so ξ(y0)=0\xi(y_{0})=0. In fact, for any y~,y(0,y0)\tilde{y},y\in(0,y_{0}) with y~<y\tilde{y}<y we have

ξ′′(s)<2y~mξ1mm(s)ξ(s),0<s<y~.\xi^{\prime\prime}(s)<-2\frac{\tilde{y}}{m}\xi^{\frac{1-m}{m}}(s)\xi^{\prime}(s),\quad 0<s<\tilde{y}.

Integrating it over [0,y~][0,\tilde{y}] we have

ξ(y~)<2θm+12+2y~(θξ1m(y~))<2θm+12+2θy~.\xi^{\prime}(\tilde{y})<-2\theta^{\frac{m+1}{2}}+2\tilde{y}\left(\theta-\xi^{\frac{1}{m}}(\tilde{y})\right)<-2\theta^{\frac{m+1}{2}}+2\theta\tilde{y}.

Integrating it again over [0,y][0,y] we have

ξ(y)<ξ~(y):=θm2θm+12y+θy2.\xi(y)<\tilde{\xi}(y):=\theta^{m}-2\theta^{\frac{m+1}{2}}y+\theta y^{2}.

Since ξ~(y)\tilde{\xi}(y) has a unique zero y:=θm12y_{*}:=\theta^{\frac{m-1}{2}}, we conclude that

(5.9) y0<y.y_{0}<y_{*}.

Step 2. Selfsimilar solution. A direct calculation shows that

u(x,t)=ξ1m(x2t)u(x,t)=\xi^{\frac{1}{m}}\left(\frac{x}{2\sqrt{t}}\right)

is a selfsimilar solution of

{ut=(um)xx,0<x<ρ(t),t>0,,u(0,t)=θ,u(ρ(t),t)=0,t>0,\left\{\begin{array}[]{l}u_{t}=(u^{m})_{xx},\quad 0<x<\rho(t),\ t>0,,\\ u(0,t)=\theta,\ \ u(\rho(t),t)=0,\quad t>0,\end{array}\right.

with ρ(t):=2y0t\rho(t):=2y_{0}\sqrt{t}. Therefore,

v(x,t)=V(x2t):=mm1ξm1m(x2t)v(x,t)=V\left(\frac{x}{2\sqrt{t}}\right):=\frac{m}{m-1}\xi^{\frac{m-1}{m}}\left(\frac{x}{2\sqrt{t}}\right)

is a selfsimilar solution of the first equation in (5.6). It also satisfies the boundary conditions in (5.6) at x=0x=0 and x=ρ(t)x=\rho(t). Moreover, the interface x=ρ(t)x=\rho(t) satisfies the Darcy law (cf. [30, Theorem 15.19]):

ρ(t)=vx(ρ(t),t).\rho^{\prime}(t)=-v_{x}(\rho(t),t).

In other words, we have V(y0)=2y0V^{\prime}(y_{0})=-2y_{0}. This proves the lemma. ∎

This lemma implies that V(x2t)V\left(\frac{x}{2\sqrt{t}}\right) is not only a solution of the equation in (5.6), but also a solution of the following Cauchy-Dirichlet problem

(5.10) {vt=(m1)vvxx+vx2,x>0,t>0,,v(0,t)=Θ:=mm1θm1,t>0,\left\{\begin{array}[]{ll}v_{t}=(m-1)vv_{xx}+v_{x}^{2},&x>0,\ t>0,,\\ v(0,t)=\Theta:=\frac{m}{m-1}\theta^{m-1},&t>0,\end{array}\right.

with one free boundary ρ(t)\rho(t).

Proposition 5.7.

Assume ff satisfies (fC), uu is a transition solution of (CP). Then its left and right free boundaries l(t)l(t) and r(t)r(t) satisfy the asymptotic speed as in (1.12).

Proof.

The conclusion can be proved in a similar way as the authors did for the reaction diffusion equation in [12, Proposition 4.2]. Note that there are several key points in the proof: (1) using the zero number argument to study the number of intersection points between two solutions. Here we see that the argument remains valid for our problem by Lemma 3.3 and by the Darcy law; (2) using a selfsimilar solution satisfying the Darcy law (which is called the Stefan boundary condition in [12]) to construct lower and supersolutions. Here we only need to use the selfsimilar solution V(x2t)V(\frac{x}{2\sqrt{t}}) obtained in the previous lemma to replace the selfsimilar solution Φ(t,x)\Phi(t,x) in [12]; (3) using the limit θ(t)/r(t)0(t)\theta(t)/r(t)\to 0\ (t\to\infty) as an extra condition, where θ(t)\theta(t) is the θ\theta-level set of u(,t)u(\cdot,t). This limit remains true for our problem and the proof is the same as that in Subsection 4.3 in [12]. ∎

6. Bistable RPME

In this section we study the asymptotic behavior for the solutions of (CP) or (pCP) with bistable type of reaction: ff satisfies (fB).

6.1. Stationary solutions

On the nonnegative stationary solutions of the equation in (CP), we first have Case A - Case D types as in the monostable RPME. The difference is that: in Case A the solutions are UqU_{q} for q=0,θ,1q=0,\theta,1 in the current case; in Case D, the compactly supported stationary solutions Uq0U_{q_{0}} exists for q0(θ1,1)q_{0}\in(\theta_{1},1), where θ1\theta_{1} is defined below by (6.1). The decay rates in (4.6) and (4.7) remain valid. As we mentioned in Remark 4.1, each Uq0U_{q_{0}} is just a stationary solution of the equation but not of the Cauchy problem (CP).

Besides these stationary solutions, for the bistable RPME, there are two other kinds of stationary solutions: periodic solutions and the ground state solution.

Case E. In the (q,p)(q,p)-phase plane, besides (0,0)(0,0) and (1,0)(1,0), we have another singular point (θ,0)(\theta,0), which is a center. There are infinitely many closed trajectories surrounding (θ,0)(\theta,0). Each of them corresponds to a positive stationary periodic solution Uper(x)U_{per}(x). By the monotonicity result in Corollary 3.7, none of them can be a candidate of the ω\omega-limit for the solution of (CP) with u0𝔛u_{0}\in\mathfrak{X}.

Case F. In the phase plane, there is a homoclinic orbit which starts and ends at (0,0)(0,0), crossing the point (θ1,0)(\theta_{1},0), with θ1(θ,1)\theta_{1}\in(\theta,1) given by

(6.1) 0θ1rm1f(r)𝑑r=0.\int_{0}^{\theta_{1}}r^{m-1}f(r)dr=0.

The function of this homoclinic orbit is

p2=2m0qrm1f(r)𝑑r,0<qθ1.p^{2}=-2m\int_{0}^{q}r^{m-1}f(r)dr,\quad 0<q\leq\theta_{1}.

Denote the corresponding solution by U(x)U(x), and assume it satisfies the normalized condition

U(0)=θ1,U(0)=0.U(0)=\theta_{1},\quad U^{\prime}(0)=0.

Assume the largest interval where UU remains positive is (L,L)(-L,L) for some L(0,]L\in(0,\infty]. We can give a formula for LL as the following. Denote η:=Um\eta:=U^{m}, then

(6.2) dηdx=p(η):=(2m0η1/mrm1f(r)𝑑r)1/2,x(L,0].\frac{d\eta}{dx}=p(\eta):=\left(-2m\int_{0}^{\eta^{1/m}}r^{m-1}f(r)dr\right)^{1/2},\quad x\in(-L,0].

Therefore,

(6.3) L:=0θ1mdηp(η)=0θ1m(2m0η1/mrm1f(r)𝑑r)1/2𝑑η.L:=\int_{0}^{\theta_{1}^{m}}\frac{d\eta}{p(\eta)}=\int_{0}^{\theta_{1}^{m}}\left(-2m\int_{0}^{\eta^{1/m}}r^{m-1}f(r)dr\right)^{-1/2}d\eta.

Under the assumptions in (fB) we have f(0)<0f^{\prime}(0)<0 and so f(q)f(0)qf(q)\sim f^{\prime}(0)q as q0+0q\to 0+0. We will see below that L<L<\infty in this case. However, LL might be \infty if ff is a bistable nonlinearity but without the assumption f(0)<0f^{\prime}(0)<0.

Lemma 6.1.

Assume fC([0,))C1((0,))f\in C([0,\infty))\cap C^{1}((0,\infty)) is a bistable nonlinearity satisfying (1.3). Further assume that

(6.4) f(q)=λqα(1+o(1)) as q0+0,f(q)=-\lambda q^{\alpha}(1+o(1))\mbox{\ \ as\ \ }q\to 0+0,

for some λ>0\lambda>0 and α>0\alpha>0. Then L=L=\infty when αm\alpha\geq m; L<L<\infty when 0<α<m0<\alpha<m.

Proof.

Substituting (6.4) into the formula p(η)p(\eta) in (6.2) we have

(6.5) p(η)(2mλm+α)1/2ηm+α2m as η0+0.p(\eta)\sim\left(\frac{2m\lambda}{m+\alpha}\right)^{1/2}\eta^{\frac{m+\alpha}{2m}}\mbox{\ \ as\ \ }\eta\to 0+0.

Case 1. αm\alpha\geq m. When 0<δ10<\delta\ll 1, we have

(6.6) L>0δ(m+α3mλ)1/2ηm+α2m𝑑η=.L>\int_{0}^{\delta}\left(\frac{m+\alpha}{3m\lambda}\right)^{1/2}\eta^{-\frac{m+\alpha}{2m}}d\eta=\infty.

Thus, the solution U(x)>0U(x)>0 in {\mathbb{R}}, it is a Type I ground state solution, as that in the bistable RDEs.

Case 2. 0<α<m0<\alpha<m. When 0<δ10<\delta\ll 1, we have

(6.7) L<δθ1mdηp(η)+0δ(m+αmλ)1/2ηm+α2m𝑑η<.L<\int_{\delta}^{\theta_{1}^{m}}\frac{d\eta}{p(\eta)}+\int_{0}^{\delta}\left(\frac{m+\alpha}{m\lambda}\right)^{1/2}\eta^{-\frac{m+\alpha}{2m}}d\eta<\infty.

By (6.5) and (6.2) we have

dηdx(2mλm+α)1/2ηm+α2m as xL+0,\frac{d\eta}{dx}\sim\left(\frac{2m\lambda}{m+\alpha}\right)^{1/2}\eta^{\frac{m+\alpha}{2m}}\mbox{\ \ as\ \ }x\to-L+0,

and so

2mmα(m+α2mλ)1/2(ηmα2m)1, as xL+0.\frac{2m}{m-\alpha}\Big{(}\frac{m+\alpha}{2m\lambda}\Big{)}^{1/2}\Big{(}\eta^{\frac{m-\alpha}{2m}}\Big{)}^{\prime}\sim 1,\mbox{\ \ as\ \ }x\to-L+0.

Hence,

(6.8) U(x)=η1/m(x)A1(α,λ)(x+L)2mα as xL+0,U(x)=\eta^{1/m}(x)\sim A_{1}(\alpha,\lambda)(x+L)^{\frac{2}{m-\alpha}}\mbox{\ \ as\ \ }x\to-L+0,

with

A1(α,λ):=(mα2m)2mα(2mλm+α)1mα.A_{1}(\alpha,\lambda):=\left(\frac{m-\alpha}{2m}\right)^{\frac{2}{m-\alpha}}\left(\frac{2m\lambda}{m+\alpha}\right)^{\frac{1}{m-\alpha}}.

From (6.8) we see that, at the boundaries ±L\pm L, UU is continuous for all m>αm>\alpha, Lipschitz when α<m2+α\alpha<m\leq 2+\alpha and C1C^{1} when α<m<2+α\alpha<m<2+\alpha.

If we denote by V(x):=mm1Um1V(x):=\frac{m}{m-1}U^{m-1} the corresponding pressure of UU, then we have

(6.9) {VC([L,L]),V(0)=Θ1:=mm1θ1m1,V(±L)=0,V(x)>0,V(x)<0,V(x)=V(x) for x(0,L),V(x)A(x+L)2(m1)mα as xL+0,with A:=mm1A1m1.\left\{\begin{array}[]{l}V\in C([-L,L]),\quad V(0)=\Theta_{1}:=\frac{m}{m-1}\theta_{1}^{m-1},\quad V(\pm L)=0,\\ V(x)>0,\ \ V^{\prime}(x)<0,\ \ V(x)=V(-x)\mbox{ for }x\in(0,L),\\ V(x)\sim A(x+L)^{\frac{2(m-1)}{m-\alpha}}\mbox{\ \ as\ \ }x\to-L+0,\quad\mbox{with }A:=\frac{m}{m-1}A_{1}^{m-1}.\end{array}\right.

This proves the lemma. ∎

Remark 6.2.

In the special case when ff is a bistable nonlinearity with f(0)<0f^{\prime}(0)<0 as in (fB), we have α=1\alpha=1 and λ=f(0)\lambda=-f^{\prime}(0), and so VV in (6.9) satisfies:

V(x)A(x+L)2 as xL+0,with A:=(m1)f(0)2(m+1).V(x)\sim A(x+L)^{2}\mbox{\ \ as\ \ }x\to-L+0,\quad\mbox{with }A:=\frac{-(m-1)f^{\prime}(0)}{2(m+1)}.

Thus V(L+0)=0V^{\prime}(-L+0)=0, and so VV or U=(m1mV)1m1U=(\frac{m-1}{m}V)^{\frac{1}{m-1}} is a Type II ground state solution. By Lemma 2.3, V(x)V(x) is a stationary solution not only for the bistable RPME, but also for the Cauchy problem (pCP). This is different from the compactly supported solutions in Case D, where they are only stationary solutions of the RPME but not of (pCP). Furthermore, the combinations of such solutions, like

(6.10) 𝒱(x):=V(xz1)+V(xz2),with z2z12L,\mathcal{V}(x):=V(x-z_{1})+V(x-z_{2}),\quad\mbox{with\ \ }z_{2}-z_{1}\geq 2L,

are also Type II ground state stationary solutions of the problem (pCP).

6.2. Asymptotic behavior

Proof of Theorem 1.6: We prove the trichotomy result in Theorem 1.6. By f(0)<0f^{\prime}(0)<0, it follows from Remark 6.2 that, the problem has only Type II ground state solution as in (6.9) or (6.10). By the general convergence result Theorem 1.1 we know that uu converges as tt\to\infty to 0,θ,10,\theta,1, or a Type II ground state solution.

The constant solution θ\theta can be excluded easily from the candidates. In fact, for any positive stationary periodic solution Uper(x)U_{per}(x) of the equation (see Case E in Subsection 6.1), by the classical zero number argument we see that the zero number of uσ(,t)Uper()u_{\sigma}(\cdot,t)-U_{per}(\cdot) is finite and decreasing in t>0t>0. However, if uσθu_{\sigma}\to\theta for some σ>0\sigma>0, then the above zero number tends to that of θUper()\theta-U_{per}(\cdot), which is infinite. So we derive a contradiction.

For the initial data ϕσ𝔛\phi_{\sigma}\in\mathfrak{X}, when σ>0\sigma>0 is small we have by (Φ3\Phi_{3}) that ϕσθ\phi_{\sigma}\leq\theta. Hence, a similar argument as in Lemma 5.2 implies that vanishing happens for uσu_{\sigma}. Denote

Σ0:={σ>0vanishing happens for uσ}.\Sigma_{0}:=\{\sigma>0\mid\mbox{vanishing happens for }u_{\sigma}\}.

Then Σ0\Sigma_{0} is an open interval (0,σ)(0,\sigma_{*}) for some σ(0,]\sigma_{*}\in(0,\infty], as in the combustion equation.

In case σ=\sigma_{*}=\infty (this is the complete vanishing case, as stated in Theorem 7.1), there is nothing left to prove.

In case σ\sigma_{*} is a positive number, we can show that uσu_{\sigma_{*}} is a transition solution, that is, it converges to some Type II ground state solution 𝒰\mathcal{U}. Otherwise, it tends to 11. Then the set

Σ1:={σ>0uσ1 as t}\Sigma_{1}:=\{\sigma>0\mid u_{\sigma}\to 1\mbox{ as }t\to\infty\}

is not empty. Since this set is open as proved in the combustion case and since σΣ1\sigma_{*}\in\Sigma_{1} by our assumption, we see that σΣ1\sigma\in\Sigma_{1} for any σ\sigma with 0<σσ10<\sigma_{*}-\sigma\ll 1. This contradicts the definition of σ\sigma_{*}. In what follows we assume u(x,t)u(x,t) converges to a Type II ground state solution 𝒰(x)\mathcal{U}(x) as in (1.8).

If Σ1\Sigma_{1} is empty, then uσu_{\sigma} is a transition solution for all σσ\sigma\geq\sigma_{*} (this is the case σ(0,)\sigma_{*}\in(0,\infty) and σ=\sigma^{*}=\infty in Theorem 1.6), and the proof is also completed. In case Σ1\Sigma_{1} is a non-empty open set, by the comparison principle, it is actually an interval (σ,)(\sigma^{*},\infty) for some σσ\sigma^{*}\geq\sigma_{*}. For each σ[σ,σ]\sigma\in[\sigma_{*},\sigma^{*}], uσu_{\sigma} is a transition solution.

This completes the proof for Theorem 1.6. \Box

On the sharpness of the transition solution: σ=σ\sigma_{*}=\sigma^{*}, we give some sufficient conditions.

Lemma 6.3.

Let uσu_{\sigma} be the solution of (CP) with bistable ff and with initial data ϕσ\phi_{\sigma} satisfying (Φ1)(\Phi_{1})-(Φ3)(\Phi_{3}). Assume uσu_{\sigma_{*}} is a transition solution, then, for any σ>σ\sigma>\sigma_{*}, uσu_{\sigma} is no longer a transition one, if one of the following conditions holds:

  1. (i).

    (Φ2)(\Phi_{2}) is strengthened as: ϕσ(x)ϕσ(x+ε)\phi_{\sigma_{*}}(x)\leq\phi_{\sigma}(x+\varepsilon) when |ε||\varepsilon| is small;

  2. (ii).

    there is i{1,2,,k}i\in\{1,2,\cdots,k\} such that ri(t)li(t)ri(t)li(t)r_{i}(t)-l_{i}(t)\not\equiv r_{*i}(t)-l_{*i}(t), where li(t),ri(t)l_{i}(t),r_{i}(t) are the left and right boundaries of the ii-th connected component of the the support spt[uσ(,t)]{\rm spt}[u_{\sigma}(\cdot,t)], li(t),ri(t)l_{*i}(t),r_{*i}(t) are the analogues of spt[uσ(,t)]{\rm spt}[u_{\sigma_{*}}(\cdot,t)];

  3. (iii).

    li(t)li(t),ri(t)ri(t)l_{i}(t)\equiv l_{*i}(t),\ r_{i}(t)\equiv r_{*i}(t) for all i={1,2,,k}i=\{1,2,\cdots,k\}, and one of the waiting times of these boundaries is finite.

Proof.

In a similar way as proving Lemma 5.4 we see that any of the three conditions (i)-(iii) implies that, for some T0,τ0T\geq 0,\ \tau\geq 0 and ε0>0\varepsilon_{0}>0, there holds

uσ(x,t)uσ(x+ε,t+τ),x[li(t),ri(t)],tT,ε[0,ε0] or ε[ε0,0].u_{\sigma_{*}}(x,t)\leq u_{\sigma}(x+\varepsilon,t+\tau),\quad x\in[l_{*i}(t),r_{*i}(t)],\ t\geq T,\ \varepsilon\in[0,\varepsilon_{0}]\mbox{ or }\varepsilon\in[-\varepsilon_{0},0].

In case uσu_{\sigma} is also a transition solution, by taking limit as tt\to\infty in this inequality we conclude that

U(xz)U(x+εzσ),x[l,r],ε[0,ε0] or ε[ε0,0],U(x-z_{*})\leq U(x+\varepsilon-z_{\sigma}),\quad x\in[l^{\infty}_{*},r^{\infty}_{*}],\ \varepsilon\in[0,\varepsilon_{0}]\mbox{ or }\varepsilon\in[-\varepsilon_{0},0],

for some z,zσ[b,b]z_{*},z_{\sigma}\in[-b,b], where l=limtli(t)l^{\infty}_{*}=\lim\limits_{t\to\infty}l_{*i}(t), r=limtri(t)r^{\infty}_{*}=\lim\limits_{t\to\infty}r_{*i}(t). This is impossible since UU is a Type II ground state solution. ∎

Remark 6.4.

The only situation not included in (ii) and (iii) in the previous lemma is

(6.11) {li(t)li(t)li(0),ri(t)ri(t)ri(0) for all i={1,2,,k},and all the waiting times of these boundaries are infinite.\left\{\begin{array}[]{l}l_{i}(t)\equiv l_{*i}(t)\equiv l_{i}(0),\ r_{i}(t)\equiv r_{*i}(t)\equiv r_{i}(0)\mbox{\ \ for all }i=\{1,2,\cdots,k\},\\ \mbox{and all the waiting times of these boundaries are infinite}.\end{array}\right.

We guess that, even in this case, the transition solution is also unique. The instability of Type II ground state solution should be the essential reason.

Remark 6.5.

In case the initial data ϕσ\phi_{\sigma} has a unique maximum point, so does u(,t)u(\cdot,t). Then the limit of the transition solution is a Type II ground state solution U(xz)U(x-z) for some z[b,b]z\in[-b,b]. Even in this case, the transition solution may have a very wide support but just converges to a positive limit in a bounded interval. For example, assume u(x,t;σ(b2x2))u(x,t;\sigma(b^{2}-x^{2})) is a transition solution for some σ>0\sigma>0, and b>Lb>L, then

u(x,t)>0 for x(l(t),r(t)),t>0,u(x,t)>0\mbox{ for }x\in(l(t),r(t)),\ t>0,
u(x,t)u(±L,t),x(l(t),L][L,r(t)),t>0,u(x,t)\leq u(\pm L,t),\quad x\in(l(t),-L]\cup[L,r(t)),\ t>0,

and

u(±L,t)0,u(,t)U()L([l(t),r(t)])0 as t.u(\pm L,t)\to 0,\quad\|u(\cdot,t)-U(\cdot)\|_{L^{\infty}([l(t),r(t)])}\to 0\mbox{\ \ as\ \ }t\to\infty.

7. Complete Vanishing Phenomena

Note that, Theorem 1.3 gives the hair-trigger effect for (CP) with monostable reactions (see also Remark 1.4 and 4.3). In the bistable and combustion cases, however, Theorems 1.5 and 1.6 show that vanishing happens for the solutions uσu_{\sigma} when σ<σ\sigma<\sigma_{*}, where σ(0,]\sigma_{*}\in(0,\infty]. A natural question is that: is it really possible that σ=\sigma_{*}=\infty? In other words, is there b>0b>0 such that, for any initial data u0u_{0} with support in [b,b][-b,b], vanishing always happens for uu, no matter how large u0L\|u_{0}\|_{L^{\infty}} is? We call such a result as a complete vanishing phenomena. Of course, this phenomena seems difficult in most cases. However, it does happen for some problems. Recently, Li and Lou [23] proved that such a phenomena really happens in bistable and combustion RDEs provided the nonlinearity f(u)f(u) decreases sufficiently fast as uu\to\infty. Also, it was shown in [11] that this phenomena occurs in the free boundary problems for RDEs with bistable, combustion or even monostable reactions. We now show that this phenomena happens for (CP), provided f(u)f(u) decreases fast as uu\to\infty.

Theorem 7.1.

Assume fC2f\in C^{2} is a bistable or combustion reaction term. Assume also that

(F) lim infu+f(u)up0>0,\mbox{\rm(F)${}_{\infty}$}\hskip 170.71652pt\liminf\limits_{u\to+\infty}\frac{-f(u)}{u^{p_{0}}}>0,\hskip 184.9429pt

for some p0>mp_{0}>m. Then there exists a small b>0b>0 such that vanishing always happens for the solution of (CP) provided spt[u0][b,b]{\rm spt}[u_{0}]\subset[-b,b].

Proof.

Recall that, when ff is a bistable or combustion nonlinearity, there exists K>0K>0 such that

f(u)Ku,u0.f(u)\leq Ku,\quad u\geq 0.

Using the pressure notation vv and

g(v):=m((m1)vm)m2m1f(((m1)vm)1m1),g(v):=m\left(\frac{(m-1)v}{m}\right)^{\frac{m-2}{m-1}}f\left(\Big{(}\frac{(m-1)v}{m}\Big{)}^{\frac{1}{m-1}}\right),

we see that

(7.1) g(v)K(m1)v for v0.g(v)\leq K(m-1)v\mbox{ for }v\geq 0.

Denote

p1:=m2+p0m1,p:=p0+3m42(m1),α:=2p2.p_{1}:=\frac{m-2+p_{0}}{m-1},\quad p:=\frac{p_{0}+3m-4}{2(m-1)},\quad\alpha:=\frac{2}{p-2}.

Then p1p=p2=p0m2(m1)=2αp_{1}-p=p-2=\frac{p_{0}-m}{2(m-1)}=\frac{2}{\alpha}. By (F) we have

lim infv+g(v)vp1>0.\liminf\limits_{v\to+\infty}\frac{-g(v)}{v^{p_{1}}}>0.

So, there exists L0>0L_{0}>0 and M>1M>1 such that

(7.2) g(v)L0vp1pvpLvp,vM,g(v)\leq-L_{0}v^{p_{1}-p}v^{p}\leq-Lv^{p},\qquad v\geq M,

where L:=L0Mp1pL:=L_{0}M^{p_{1}-p}.

Recall that θ\theta is a positive zero of ff in the bistable and combustion equations. Choose M>1M>1 larger if necessary (to be determined below) and denote M0:=((m1)Mm)1m1M_{0}:=\Big{(}\frac{(m-1)M}{m}\Big{)}^{\frac{1}{m-1}}, then we can choose s0>0s_{0}>0, with 0<1K(m1)s010<1-K(m-1)s_{0}\ll 1 when M0M_{0} is sufficiently large, such that

(7.3) θ(1K(m1)s02K)1m+1(1+(m1)K(m21)s02)1m1=2M0s01m+1>1.\theta\frac{\left(\frac{1-K(m-1)s_{0}}{2K}\right)^{\frac{1}{m+1}}}{\left(1+\frac{(m-1)-K(m^{2}-1)s_{0}}{2}\right)^{\frac{1}{m-1}}}=2M_{0}s_{0}^{\frac{1}{m+1}}>1.

Define b1=C1(m)s01m+1b_{1}=C_{1}(m)s_{0}^{\frac{1}{m+1}} by

(7.4) (m1)b122m(m+1)s02m+1=112m1.\frac{(m-1)b^{2}_{1}}{2m(m+1)s_{0}^{\frac{2}{m+1}}}=1-\frac{1}{2^{m-1}}.

Then, we can choose M>1M>1 sufficiently large such that

(7.5) b:=M1αmin{b13,1},b:=M^{-\frac{1}{\alpha}}\leq\min\left\{\frac{b_{1}}{3},1\right\},
(7.6) L:=L0Mp1p3cb1M+2p[(m1)α(α+1)+α2],L:=L_{0}M^{p_{1}-p}\geq\frac{3c}{b_{1}M}+2^{p}[(m-1)\alpha(\alpha+1)+\alpha^{2}],

where c>0c>0 is defined by

c:=21αb1αα1[2(m1)α(α+1)+2α2+K(m1)].c:=2^{\frac{1}{\alpha}}b^{-1-\alpha}\alpha^{-1}[2(m-1)\alpha(\alpha+1)+2\alpha^{2}+K(m-1)].

Denote C:=(2M0)m1s0m1m+1>1C:=(2M_{0})^{m-1}s_{0}^{\frac{m-1}{m+1}}>1, then

(7.7) (C(m1)b122m(m+1)s02m+1)1m112C1m1=M0s01m+1.\left(C-\frac{(m-1)b^{2}_{1}}{2m(m+1)s_{0}^{\frac{2}{m+1}}}\right)^{\frac{1}{m-1}}\geq\frac{1}{2}C^{\frac{1}{m-1}}=M_{0}s_{0}^{\frac{1}{m+1}}.

We will show that the complete vanishing phenomena happens if spt[u(x,0)][b,b]{\rm spt}[u(x,0)]\subset[-b,b].

Step 1. We first consider the stage when the solution is suppressed by a spatial homogeneous upper solution down to the range vMv\leq M. Consider the equation

v1t=Lv12.v_{1t}=-Lv_{1}^{2}.

Then v1=(Lt)1v_{1}=(Lt)^{-1} is monotonically decreasing from ++\infty, and till it reaches MM at time t=t1:=(LM)1t=t_{1}:=(LM)^{-1}, it is a supersolution of (pCP) due to (7.2) and p>2p>2, no matter how large u0L\|u_{0}\|_{L^{\infty}} is. So,

v(x,t)v1(t1)=M,tt1.v(x,t)\leq v_{1}(t_{1})=M,\quad t\geq t_{1}.

This inequality gives an estimate for vv from above. We also need estimates on the right and left boundaries. We state only the case on the right side since the left side is studied similarly. For c,bc,b given above, we define

v2(x,t):=(xctb)αbα,b+ct<x2b+ct,t0.v_{2}(x,t):=(x-ct-b)^{-\alpha}-b^{-\alpha},\quad b+ct<x\leq 2b+ct,\ t\geq 0.

We now show that v2v_{2} is also a supersolution of (pCP) in the domain

E:={(x,t)b+ct<x2b+ct,t0}.E:=\{(x,t)\mid b+ct<x\leq 2b+ct,\ t\geq 0\}.

1). First, on the interval J1:=[b+ct+21αb,2b+ct]J_{1}:=[b+ct+2^{-\frac{1}{\alpha}}b,2b+ct], we have

21αbH:=xctbb,0v2bα,2^{-\frac{1}{\alpha}}b\leq H:=x-ct-b\leq b,\quad 0\leq v_{2}\leq b^{-\alpha},

and so by the first inequality in (7.1) and the choice of cc we have

𝒩v2\displaystyle\mathcal{N}v_{2} :=\displaystyle:= v2t(m1)v2v2xxv2x2g(v2)\displaystyle v_{2t}-(m-1)v_{2}v_{2xx}-v_{2x}^{2}-g(v_{2})
\displaystyle\geq αcHα1(m1)α(α+1)v2Hα2α2H2α2K(m1)v2\displaystyle\alpha cH^{-\alpha-1}-(m-1)\alpha(\alpha+1)v_{2}H^{-\alpha-2}-\alpha^{2}H^{-2\alpha-2}-K(m-1)v_{2}
\displaystyle\geq Hα2[αcH(m1)α(α+1)Hαα2Hα]K(m1)bα\displaystyle H^{-\alpha-2}\left[\alpha cH-(m-1)\alpha(\alpha+1)H^{-\alpha}-\alpha^{2}H^{-\alpha}\right]-K(m-1)b^{-\alpha}
\displaystyle\geq bα2[αc21αb2(m1)α(α+1)bα2α2bα]K(m1)bα\displaystyle b^{-\alpha-2}\left[\alpha c2^{-\frac{1}{\alpha}}b-2(m-1)\alpha(\alpha+1)b^{-\alpha}-2\alpha^{2}b^{-\alpha}\right]-K(m-1)b^{-\alpha}
\displaystyle\geq b2α2[αc21αb1+α2(m1)α(α+1)2α2K(m1)]0.\displaystyle b^{-2\alpha-2}\left[\alpha c2^{-\frac{1}{\alpha}}b^{1+\alpha}-2(m-1)\alpha(\alpha+1)-2\alpha^{2}-K(m-1)\right]\geq 0.

2). Next, in the interval J2:=(b+ct,b+ct+21αb]J_{2}:=(b+ct,b+ct+2^{-\frac{1}{\alpha}}b] we have

0<H:=xctb21αb,v2bαM,0<H:=x-ct-b\leq 2^{-\frac{1}{\alpha}}b,\quad v_{2}\geq b^{-\alpha}\geq M,

and by the choice of LL we have

𝒩v2\displaystyle\mathcal{N}v_{2} \displaystyle\geq Hα2[αcH(m1)α(α+1)Hαα2Hα]+Lv2p\displaystyle H^{-\alpha-2}\left[\alpha cH-(m-1)\alpha(\alpha+1)H^{-\alpha}-\alpha^{2}H^{-\alpha}\right]+Lv_{2}^{p}
\displaystyle\geq L(Hαbα)p[(m1)α(α+1)+α2]H2α2\displaystyle L(H^{-\alpha}-b^{-\alpha})^{p}-[(m-1)\alpha(\alpha+1)+\alpha^{2}]H^{-2\alpha-2}
\displaystyle\geq 2pLHαp[(m1)α(α+1)+α2]H2α2\displaystyle 2^{-p}LH^{-\alpha p}-[(m-1)\alpha(\alpha+1)+\alpha^{2}]H^{-2\alpha-2}
\displaystyle\geq H2α2[2pL(m1)α(α+1)α2]0.\displaystyle H^{-2\alpha-2}\left[2^{-p}L-(m-1)\alpha(\alpha+1)-\alpha^{2}\right]\geq 0.

3). On the free boundary r2(t):=2b+ctr_{2}(t):=2b+ct of v2v_{2}, by the choice of cc, we easily obtain

v2x(2b+ct,t)=αbα1c,t0.-v_{2x}(2b+ct,t)=\alpha b^{-\alpha-1}\leq c,\quad t\geq 0.

Therefore, v2v_{2} is a supersolution of (pCP) in the domain EE. As a consequence, at t1t_{1}, the right free boundary r(t)r(t) of vv satisfies

r(t1)r2(t1)=2b+ct12b+cLMb1,r(t_{1})\leq r_{2}(t_{1})=2b+ct_{1}\leq 2b+\frac{c}{LM}\leq b_{1},

by the choice of b,cb,c and LL. Similarly, the left free boundary l(t)l(t) of vv satisfies l(t1)b1l(t_{1})\geq-b_{1}.

Using the notation of uu we conclude from above that

(7.8) u(x,t1)u¯(x):=M0χ[b1,b1].u(x,t_{1})\leq\bar{u}(x):=M_{0}\cdot\chi_{[-b_{1},b_{1}]}.

Step 2. Next we use the ZKB solution as another upper solution to suppress uu further such that it goes down below θ\theta after some time. More precisely, consider

{u3t=(u3m)xx+Ku3,x,t>0,u3(x,0)=u¯(x),x.\left\{\begin{array}[]{ll}u_{3t}=(u^{m}_{3})_{xx}+Ku_{3},&x\in{\mathbb{R}},\ t>0,\\ u_{3}(x,0)=\bar{u}(x),&x\in{\mathbb{R}}.\end{array}\right.

Then u3u_{3} is a upper solution of (CP) and

u(x,t+t1)u3(x,t),x,t>0.u(x,t+t_{1})\leq u_{3}(x,t),\quad x\in{\mathbb{R}},\ t>0.

Change the time variable from tt to ss by

s:=eK(m1)t1K(m1)t=T(s):=ln(1+K(m1)s)K(m1),s:=\frac{e^{K(m-1)t}-1}{K(m-1)}\Leftrightarrow t=T(s):=\frac{\ln(1+K(m-1)s)}{K(m-1)},

and define

w(x,s):=u3(x,T(s))[1+K(m1)s]1m1,w(x,s):=u_{3}(x,T(s))[1+K(m-1)s]^{-\frac{1}{m-1}},

then we have

(7.9) {ws=(wm)xx,x,s>0,w(x,0)=u¯(x),x.\left\{\begin{array}[]{ll}w_{s}=(w^{m})_{xx},&x\in{\mathbb{R}},\ s>0,\\ w(x,0)=\bar{u}(x),&x\in{\mathbb{R}}.\end{array}\right.

To estimate w(x,s)w(x,s), we consider the ZKB solution

w1(x,s):=s1m+1(C(m1)x22m(m+1)s2m+1)+1m1,x,ss0,w_{1}(x,s):=s^{-\frac{1}{m+1}}\left(C-\frac{(m-1)x^{2}}{2m(m+1)s^{\frac{2}{m+1}}}\right)^{\frac{1}{m-1}}_{+},\quad x\in{\mathbb{R}},\ s\geq s_{0},

with s0s_{0} and CC given above. By the choice of b1,Cb_{1},C and s0s_{0}, for x[b1,b1]x\in[-b_{1},b_{1}] we have

w1(x,s0)\displaystyle w_{1}(x,s_{0}) \displaystyle\geq w1(±b1,s0)=s01m+1(C(m1)b122m(m+1)s02m+1)1m1\displaystyle w_{1}(\pm b_{1},s_{0})=s_{0}^{-\frac{1}{m+1}}\left(C-\frac{(m-1)b^{2}_{1}}{2m(m+1)s_{0}^{\frac{2}{m+1}}}\right)^{\frac{1}{m-1}}
\displaystyle\geq 12s01m+1C1m1>M0u¯(x).\displaystyle\frac{1}{2}s_{0}^{-\frac{1}{m+1}}C^{\frac{1}{m-1}}>M_{0}\geq\bar{u}(x).

Therefore, w1w_{1} is a upper solution of (7.9), and so

w(x,s)w1(x,s+s0),x,s>0.w(x,s)\leq w_{1}(x,s+s_{0}),\quad x\in{\mathbb{R}},\ s>0.

In particular, at x=0x=0 and s=s1:=1K(m+1)s02Ks=s_{1}:=\frac{1-K(m+1)s_{0}}{2K} we have

w(0,s1)w1(0,s1+s0)=(s1+s0)1m+1C1m1.w(0,s_{1})\leq w_{1}(0,s_{1}+s_{0})=(s_{1}+s_{0})^{-\frac{1}{m+1}}C^{\frac{1}{m-1}}.

Using the notation uu we have

u(x,T(s1)+t1)\displaystyle u(x,T(s_{1})+t_{1}) \displaystyle\leq u3(x,T(s1))=w(x,s1)[1+K(m1)s1]1m1\displaystyle u_{3}(x,T(s_{1}))=w(x,s_{1})[1+K(m-1)s_{1}]^{\frac{1}{m-1}}
\displaystyle\leq (s1+s0)1m+1C1m1[1+K(m1)s1]1m1θ,\displaystyle(s_{1}+s_{0})^{-\frac{1}{m+1}}C^{\frac{1}{m-1}}[1+K(m-1)s_{1}]^{\frac{1}{m-1}}\leq\theta,

by (7.3) and (7.7).

Step 3. By the sufficient condition for vanishing in bistable and combustion RPMEs (cf. Lemma 5.2), we conclude that u0u\to 0 as tt\to\infty.

This proves the complete vanishing phenomena in Theorem 7.1. ∎

References

  • [1] S. B. Angenent, The zero set of a solution of a parabolic equation, J. Reine Angew. Math., 390 (1988), 79-96.
  • [2] D. G. Aronson, Regularity properties of flows through porous media, SIAM J. Appl. Math., 17 (1969), 461-467.
  • [3] D. G. Aronson, Density-dependent interaction-diffusion systems, Dynamics and modelling of reactive systems (Proc. Adv. Sem., Math. Res. Center, Univ. Wisconsin, Madison, Wis., 1979), pp. 161-176, Publ. Math. Res. Center Univ. Wisconsin, 44, Academic Press, New York-London, 1980.
  • [4] D. G. Aronson and P. Bénilan, Régularité des solutions de l’équation des milieux poreux dans n{\mathbb{R}}^{n}, C. R. Acad. Sci. Paris Ser. A-B, 288 (1979), 103-105.
  • [5] D. G. Aronson, M. G. Crandall and L. A. Peletier, Stabilization of solutions of a degenerate nonlinear diffusion problem, Nonlinear Anal. TMA, 6 (1982), 1001-1022.
  • [6] D. G. Aronson and H. F. Weinberger, Nonlinear diffusion in population genetics, conbustion, and nerve pulse propagation, In: Partial Differential Equations and Related Topics, Lecture Notes in Math. 446, Springer, Berlin, 5-49, 1975.
  • [7] D. G. Aronson and H. F. Weinberger, Multidimensional nonlinear diffusion arising in population genetics, Adv. Math., 30 (1978), 33-76.
  • [8] P. Bénilan, Evolution Equations and Accretive Operators, Lecture Notes, Univ. Kentucky, Manuscript, 1981.
  • [9] X. Y. Chen, A strong unique continuation theorem for parabolic equations, Math. Ann., 311 (1998), 603-630.
  • [10] E. DiBenedetto and A. Friedman, Hölder estimates for nonlinear degenerate parabolic systems, J. Reine Angew. Math., 357 (1985), 1-22. Addendum to: Hölder estimates for nonlinear degenerate parabolic systems, J. Reine Angew. Math., 363, 217-220.
  • [11] Y. Du and B. Lou, Spreading and vanishing in nonlinear diffusion problems with free boundaries, J. Eur. Math. Soc., 17 (2015), 2673-2724.
  • [12] Y. Du, B. Lou and M. Zhou, Nonlinear diffusion problems with free boundaries: Convergence, transition speed and zero number arguments, SIAM J. Math. Anal., 47 (2015), 3555-3584.
  • [13] Y. Du and H. Matano, Convergence and sharp thresholds for propagation in nonlinear diffusion problems, J. Eur. Math. Soc., 12 (2010), 279-312.
  • [14] Y. Du, F. Quirós and M. Zhou, Logarithmic corrections in Fisher-KPP problems for the porous medium equations, J. Math. Pures Appl., 136 (2020), 415-455.
  • [15] F. J. Fernandez, Unique continuation for parabolic operators. II, Comm. Partial Differential Equations, 28 (2003), 1597-1604.
  • [16] P. C. Fife and J. B. McLeod, The approach of solutions of nonlinear diffusion equations to travelling front solutions, Arch. Ration. Mech. Anal., 65 (1977), 335-361.
  • [17] R. A. Fisher, The wave of advance of advantageous genes, Ann. Eugenics, 7 (1937), 335-369.
  • [18] A. Gárriz, Propagation of solutions of the porous medium equation with reaction and their travelling wave behaviour, Nonl. Anal., 195 (2020), 111736, pp. 23.
  • [19] B. H. Gilding and R. Kersner, Travelling waves in nonlinear diffusion-convection reaction, in: Progress in Nonlinear Differential Equations and their Applications, Vol. 60, Birkhäuser Verlag, Basel, ISBN: 3-7643-7071-8.
  • [20] W. S. C. Gurney and R. M. Nisbet, A note on non-linear population transport, J. Theor. Biol., 56 (1976), 249-251.
  • [21] M. E. Gurtin and R. C. MacCamy, On the diffusion of biological populations, Math. Biosci., 33 (1977), 35-49.
  • [22] A. N. Kolmogorov, I. G. Petrovski and N. S. Piskunov, A study of the diffusion equation with increase in the amount of substance, and its application to a biological problem, Bull. Moscow Univ., Math. Mech., 1 (1937), 1-25.
  • [23] Q. Li and B. Lou, Vanishing phenomena in fast decreasing generalized bistable equations, J. Math. Anal. Appl., 500 (2021), 125096.
  • [24] W. Newman and C. Sagan, Galactic civilizations: population dynamics and interstellar diffusion, Icarus, 46 (1981), 293-327.
  • [25] A. de Pablo and J. L. Vázquez, Travelling waves and finite propagation in a reaction-diffusion equation, J. Differential Equations, 91 (1991), 19-61.
  • [26] P. E. Sacks, The initial and boundary value problem for a class of degenerate parabolic equations, Comm. Partial Differential Equations, 8 (1983), 693-733.
  • [27] F. Sanchez-Garduno and P.K. Maini, Existence and uniqueness of a sharp travelling wave in degenerate non-linear diffusion Fisher-KPP equations, J. Math. Biol., 33 (1994), 163-192.
  • [28] O. A. Oleĭnik, A. S. Kalašinkov and Y.-L. Čžou, The Cauchy problem and boundary problems for equations of the type of non-stationary filtration (Russian), Izv. Akad. Nauk SSSR. Ser. Mat., 22 (1958), 667-704.
  • [29] J. A. Sherratt, On the form of smooth-front travelling waves in a reaction-diffusion equation with degenerate nonlinear diffusion, Math. Model. Nat. Phenom., 5 (2010), 64-79.
  • [30] J. L. Vázquez, The Porous Medium Equation Mathematical Theory, Clarendon Press, Oxford, 2007.
  • [31] Z. Wu, J. Zhao, J. Yin and H. Li, Nonlinear Diffusion Equations, World Scientific Publishing Co., Inc., River Edge, NJ, 2001. xviii+502 pp.
  • [32] A. Zlatoš, Sharp transition between extinction and propagation of reaction, J. Amer. Math. Soc., 19 (2006), 251-263.

8. Appendix: Well-posedness and a priori Estimates

For our RPME with general reaction terms, it seems that the a priori estimates, well-posedness and other important properties as that in PME are not well collected in literature, though they are not entirely new. We present some details here for the convenience of the readers.

8.1. Well-posedness

The existence of very weak solution of (CP) has been well studied by many authors (cf. [5, 25, 26, 30, 31]).

Proposition 8.1.

Assume (I) and ff is Lipschitz continuous with f(0)=0f(0)=0. Then the problem (CP) has a unique very weak solution u(x,t)C(QT)L(QT)u(x,t)\in C(Q_{T})\cap L^{\infty}(Q_{T}) for any T>0T>0, and

0u(x,t)M0:=max{1,u0L} in QT=×(0,T).0\leq u(x,t)\leq M^{\prime}_{0}:=\max\{1,\|u_{0}\|_{L^{\infty}}\}\quad\mbox{ in }Q_{T}={\mathbb{R}}\times(0,T).

Moreover, there are two free boundaries l(t)<r(t)l(t)<r(t) such that spt(u(,t))[l(t),r(t)]{\rm spt}(u(\cdot,t))\subset[l(t),r(t)] for all t>0t>0.

We can not obtain u(x,t)>0u(x,t)>0 as in RDEs due to the degeneracy of the equation at u=0u=0. Nevertheless, the solution uu solves the equation in the classical sense by the standard parabolic theory in a neighborhood of every point (x,t)(x,t) at which the solution is positive (see also [28]).

8.2. Persistence of positivity and finite propagation speed

We now show the following positivity persistence property as in PME.

Proposition 8.2 (Persistence of positivity).

Assume u(x0,t0)>0u(x_{0},t_{0})>0 for some x0,t0>0x_{0}\in{\mathbb{R}},\ t_{0}>0. Then u(x0,t)>0u(x_{0},t)>0 for all tt0t\geq t_{0}. ((Thus, uu is classical near the line {(x0,t)t>t0}).\{(x_{0},t)\mid t>t_{0}\}).

Proof.

Since t0>0t_{0}>0 we have u(x,t0)C()u(x,t_{0})\in C({\mathbb{R}}), and so, by our assumption there exist ε>0,δ>0\varepsilon>0,\ \delta>0 such that

u(x,t0)ε for xJ:=[x0δ,x0+δ].u(x,t_{0})\geq\varepsilon\mbox{ for }x\in J:=[x_{0}-\delta,x_{0}+\delta].

We use KK to denote the Lipschitz constant for ff. Denote

β:=1m1,μ:=K+π24δ2εm1,\beta:=\frac{1}{m-1},\qquad\mu:=K+\frac{\pi^{2}}{4\delta^{2}}\varepsilon^{m-1},

and define

(8.1) u¯(x,t):={εeμt(sinπ(xx0+δ)2δ)1m,xJ,t0,0,x\J,t0.\underline{u}(x,t):=\left\{\begin{array}[]{ll}\displaystyle\varepsilon e^{-\mu t}\left(\sin\frac{\pi(x-x_{0}+\delta)}{2\delta}\right)^{\frac{1}{m}},&x\in J,\ t\geq 0,\\ 0,&x\in{\mathbb{R}}\backslash J,\ t\geq 0.\end{array}\right.

It is easily seen that

u¯(x,0)εu(x,t0),xJ,\underline{u}(x,0)\leq\varepsilon\leq u(x,t_{0}),\quad x\in J,

and

u¯t(u¯m)xxf(u¯)u¯t(u¯m)xx+Ku¯0,x,t0.\underline{u}_{t}-(\underline{u}^{m})_{xx}-f(\underline{u})\leq\underline{u}_{t}-(\underline{u}^{m})_{xx}+K\underline{u}\leq 0,\quad x\in{\mathbb{R}},\ t\geq 0.

By the comparison principle we have

u(x,t+t0)u¯(x,t)>0,|xx0|<δ,t>0.u(x,t+t_{0})\geq\underline{u}(x,t)>0,\quad|x-x_{0}|<\delta,\ t>0.

This proves the proposition. ∎

Remark 8.3.

A positive lower bound can also be given by a subsolution with the form of ZKB solution:

(8.2) u¯1(x,t):=1(t+1)βeK(t+1)(C(xx0)2e(m1)K(t+1)2m)+β,x,t0,\underline{u}_{1}(x,t):=\frac{1}{(t+1)^{\beta}e^{K(t+1)}}\left(C-\frac{(x-x_{0})^{2}e^{(m-1)K(t+1)}}{2m}\right)^{\beta}_{+},\quad x\in{\mathbb{R}},\ t\geq 0,

with w+:=max{w(x,t),0}w_{+}:=\max\{w(x,t),0\} and C>0C>0 satisfying

2mCe(m1)Kδ2,CβεeK.2mCe^{-(m-1)K}\leq\delta^{2},\quad C^{\beta}\leq\varepsilon e^{K}.

Using this subsolution, we can conclude that

u(x,t)u¯1(x,t)>0 if |xx0|<s¯(t):=(2mC)1/2e(m1)K(t+1)2,t0.u(x,t)\geq\underline{u}_{1}(x,t)>0\mbox{ if }|x-x_{0}|<\underline{s}(t):=(2mC)^{1/2}e^{-\frac{(m-1)K(t+1)}{2}},\ t\geq 0.

Note that, using this estimate we have u>0u>0 in a shrinking domain (x0s¯(t),x0+s¯(t))(x_{0}-\underline{s}(t),x_{0}+\underline{s}(t)), while using u¯\underline{u} we have u>0u>0 in a fixed domain (x0δ,x0+δ)(x_{0}-\delta,x_{0}+\delta).

When u00u_{0}\geq 0 is compactly supported, besides the upper bound in Proposition 8.1, we also have the finite propagation speed as in PME.

Proposition 8.4.

Assume (I) and ff is Lipschitz continuous with f(0)=0f(0)=0. Then the support of u(,t)u(\cdot,t) lies in [s¯(t),s¯(t)][-\bar{s}(t),\bar{s}(t)] with

(8.3) s¯(t):=[C1(t+1)]1/2e(m1)K(t+1)2,t>0,\bar{s}(t):=[C_{1}(t+1)]^{1/2}e^{\frac{(m-1)K(t+1)}{2}},\quad t>0,

for some C1C_{1} depending only on m,Km,K and u0L\|u_{0}\|_{L^{\infty}}, and so l(t),r(t)s¯(t)-l(t),r(t)\leq\bar{s}(t).

Proof.

We construct a supersolution of the ZKB form. Set A:=(4mβ)βA:=(4m\beta)^{-\beta}, choose C1>0C_{1}>0 such that

AeK(C1b2e(m1)K)βu0L,Ae^{K}\left(C_{1}-\frac{b^{2}}{e^{(m-1)K}}\right)^{\beta}\geq\|u_{0}\|_{L^{\infty}},

where b>0b>0 is the number in (I), and define

u¯(x,t):=AeK(t+1)(C1x2(t+1)e(m1)K(t+1))+β,x,t0.\bar{u}(x,t):=Ae^{K(t+1)}\left(C_{1}-\frac{x^{2}}{(t+1)e^{(m-1)K(t+1)}}\right)^{\beta}_{+},\quad x\in{\mathbb{R}},\ t\geq 0.

By the choose of C1C_{1} we see that u¯(x,0)u0(x)\bar{u}(x,0)\geq u_{0}(x), and by a direct calculation we have

u¯t(u¯m)xx+Ku¯(u¯m)xx+f(u¯),|x|s¯(t):=[C1(t+1)]1/2e(m1)K(t+1)2,t0.\bar{u}_{t}\geq(\bar{u}^{m})_{xx}+K\bar{u}\geq(\bar{u}^{m})_{xx}+f(\bar{u}),\quad|x|\leq\bar{s}(t):=[C_{1}(t+1)]^{1/2}e^{\frac{(m-1)K(t+1)}{2}},\ t\geq 0.

Hence, u¯\bar{u} is a supersolution of (CP), and the conclusion follows from the comparison. ∎

Remark 8.5.

Our construction for the subsolution in (8.2) and the supersolution in this proof is inspired by the ZKB solution, but with obvious difference due to the presence of the linear term KuKu.

8.3. A priori estimate for the pressure

As we have mentioned in Section 2, the pressure v(x,t):=mm1[u(x,t)]m1v(x,t):=\frac{m}{m-1}[u(x,t)]^{m-1} solves

(pCP){vt=(m1)vvxx+vx2+g(v),x,t>0,v(x,0)=v0(x):=mm1u0m1(x),x,{\rm(pCP)}\hskip 85.35826pt\left\{\begin{array}[]{ll}v_{t}=(m-1)vv_{xx}+v_{x}^{2}+g(v),&x\in{\mathbb{R}},\ t>0,\\ \displaystyle v(x,0)=v_{0}(x):=\frac{m}{m-1}u_{0}^{m-1}(x),&x\in{\mathbb{R}},\end{array}\right.\hskip 142.26378pt

with

g(v):=m((m1)vm)m2m1f(((m1)vm)1m1)g(v):=m\Big{(}\frac{(m-1)v}{m}\Big{)}^{\frac{m-2}{m-1}}f\left(\Big{(}\frac{(m-1)v}{m}\Big{)}^{\frac{1}{m-1}}\right)

satisfying

(8.4) |g(v)|K(m1)v,v0,|g(v)|\leq K(m-1)v,\quad v\geq 0,

by (F). Since uu is bounded as in Proposition 8.1, we see that vv is also bounded:

(8.5) 0v(x,t)M0:=max{mm1,v0L},x,t>0.0\leq v(x,t)\leq M_{0}:=\max\left\{\frac{m}{m-1},\|v_{0}\|_{L^{\infty}}\right\},\quad x\in{\mathbb{R}},\ t>0.

Hence,

(8.6) |g(v)|G0(m,K,v0):=K(m1)M0,0vM0.|g(v)|\leq G_{0}(m,K,v_{0}):=K(m-1)M_{0},\quad 0\leq v\leq M_{0}.

By the above propositions we have the following result.

Proposition 8.6.

Assume (I) and ff is Lipschitz continuous with f(0)=0f(0)=0. Then the problem (pCP) has a unique nonnegative solution v(x,t)C(QT)L(QT)v(x,t)\in C(Q_{T})\cap L^{\infty}(Q_{T}) for any T>0T>0.

Moreover, if v(x0,t0)>0v(x_{0},t_{0})>0 for some x0x_{0}\in{\mathbb{R}} and t0>0t_{0}>0, then v(x0,t)>0v(x_{0},t)>0 for all tt0t\geq t_{0}, and so vv is a classical solution near the line {(x0,t)tt0}\{(x_{0},t)\mid t\geq t_{0}\}.

The support of v(,t)v(\cdot,t) lies in [s¯(t),s¯(t)][-\bar{s}(t),\bar{s}(t)] with s¯(t)\bar{s}(t) given by (8.3).

In what follows, we continue to present the a priori estimates for vx,vxxv_{x},\ v_{xx} and vtv_{t} under further restrictions on ff: fC2f\in C^{2}. In this case we have

(8.7) |g(v)|G1(m,K,v0)and|g′′(v)|G2(m,K,v0),0vM0.|g^{\prime}(v)|\leq G_{1}(m,K,v_{0})\quad\mbox{and}\quad|g^{\prime\prime}(v)|\leq G_{2}(m,K,v_{0}),\quad 0\leq v\leq M_{0}.

In a similar way as Aronson [2] (with obvious modification caused by the reaction term g(v)g(v)) we have

Lemma 8.7 ([2], uniform estimate for vxv_{x}).

Assume (I) and fC1([0,))f\in C^{1}([0,\infty)) with f(0)=0f(0)=0. Let vv be a smooth positive classical solution of (pCP) in R:=(a,b)×(0,T]R:=(a,b)\times(0,T] for some a,b,Ta,b,T\in{\mathbb{R}} with b>a,T>0b>a,\ T>0. Then for any 0<δ<ba2,τ<T0<\delta<\frac{b-a}{2},\ \tau<T there holds:

(8.8) |vx(x,t)|M1(m,M0,G0,G1,δ,τ),(x,t)[a+δ,bδ]×[τ,T].|v_{x}(x,t)|\leq M_{1}(m,M_{0},G_{0},G_{1},\delta,\tau),\quad(x,t)\in[a+\delta,b-\delta]\times[\tau,T].

Moreover, if τ=0\tau=0 and

M10:=max[a,b]|v0(x)|<,M_{1}^{0}:=\max\limits_{[a,b]}|v^{\prime}_{0}(x)|<\infty,

then (8.8) holds in (a+δ,bδ)×(0,T](a+\delta,b-\delta)\times(0,T] for M1=M1(m,M0,G0,G1,δ,M10)M_{1}=M_{1}(m,M_{0},G_{0},G_{1},\delta,M_{1}^{0}).

Note that the a priori bound M1M_{1} of vxv_{x} obtained in this lemma is uniform in TT due to the fact that the bound of vv is so. The result holds for classical solutions. For the very weak solution in C(QT)L(QT)C(Q_{T})\cap L^{\infty}(Q_{T}), however, one can show the following Lipschitz continuity for vv and Hölder continuity for uu (which can be regarded as the limit of a sequence of classical solutions).

Proposition 8.8 ([2], Lipschitz continuity for vv and Hölder continuity for uu).

Assume (I) and fC1([0,))f\in C^{1}([0,\infty)) with f(0)=0f(0)=0. If we further assume that u0mu^{m}_{0} is Lipschitz continuous, then for T>τ>0T>\tau>0,

(8.9) |v(x,t)v(y,t)|C1(m,τ,u0L)|xy|,x,y,t[τ,T],|v(x,t)-v(y,t)|\leq C_{1}(m,\tau,\|u_{0}\|_{L^{\infty}})\cdot|x-y|,\quad x,y\in{\mathbb{R}},\ t\in[\tau,T],
(8.10) |u(x,t)u(y,t)|C2(m,τ,u0L)|xy|ν,x,y,t[τ,T],ν:=min{1,1m1}|u(x,t)-u(y,t)|\leq C_{2}(m,\tau,\|u_{0}\|_{L^{\infty}})\cdot|x-y|^{\nu},\quad x,y\in{\mathbb{R}},\ t\in[\tau,T],\ \nu:=\min\Big{\{}1,\frac{1}{m-1}\Big{\}}

and umx(x,t)\frac{\partial u^{m}}{\partial x}(x,t) exists and is continuous in xx\in{\mathbb{R}}, with umx(x,t)=0\frac{\partial u^{m}}{\partial x}(x,t)=0 if u(x,t)=0u(x,t)=0;

Proof.

The proof follows from the previous lemma and the fact: the very weak solution uu is the pointwise limit of a decreasing sequence of positive classical functions. (For the PME, this was shown in [28]. For the current problem, the proof is similar). ∎

Next we consider the lower bound of vxxv_{xx}. For the PME, Aronson and Bénilan [4] gave a lower bound for vxxv_{xx} in 1979:

(8.11) vxx1(m+1)t,t>0.v_{xx}\geq-\frac{1}{(m+1)t},\quad t>0.

This inequality, usually known as the Aronson-Bénilan estimate, is understood in the sense of distributions. It is optimal in the sense that equality is actually attained by the source-type or ZKB solutions. It is a significant novelty of the Cauchy problem, and is used so often in the theory of nonnegative solutions in the whole space. For our problem (pCP), we will also give a lower bound for vxxv_{xx}, but with quite different order from (8.11).

Proposition 8.9.

Assume (F), (I), and v0C2v_{0}\in C^{2} a.e. in {\mathbb{R}}, with

0v0(x)M00,M10:=sup|v0(x)|<,v0′′(x)M20>,a.e.x.0\leq v_{0}(x)\leq M_{0}^{0},\quad M_{1}^{0}:=\sup\limits_{{\mathbb{R}}}|v^{\prime}_{0}(x)|<\infty,\quad v^{\prime\prime}_{0}(x)\geq-M_{2}^{0}>-\infty,\quad a.e.\ x\in{\mathbb{R}}.

Let vC(QT)L(QT)v\in C(Q_{T})\cap L^{\infty}(Q_{T}) for any T>0T>0 be the solution of (pCP). Then

vxx(x,t)C(t+τ),x,t(0,T],v_{xx}(x,t)\geq-C(t+\tau),\quad x\in{\mathbb{R}},\ t\in(0,T],

for some τ=τ(m,M00,M10,M20,G0,G1,G2)\tau=\tau(m,M_{0}^{0},M_{1}^{0},M_{2}^{0},G_{0},G_{1},G_{2}) and C=C(m,M00,M10,G0,G1,G2)C=C(m,M_{0}^{0},M_{1}^{0},G_{0},G_{1},G_{2}), here the inequality holds in the sense of distributions in ×(0,T]{\mathbb{R}}\times(0,T].

Proof.

We follow the idea in Aronson and Bénilan [4] (see also [30, §9.3]).

1). First we assume v>0v>0 in QT:=×(0,T]Q_{T}:={\mathbb{R}}\times(0,T]. In this case, vv is a classical solution of (pCP). We write the equation satisfied by η:=vxx\eta:=v_{xx} by differentiating the equation of vv twice:

ηt=(η)+g′′vx2,\eta_{t}=\mathcal{L}(\eta)+g^{\prime\prime}v_{x}^{2},

with

(η):=(m1)vηxx+2mvxηx+(m+1)η2+g(v)η.\mathcal{L}(\eta):=(m-1)v\eta_{xx}+2mv_{x}\eta_{x}+(m+1)\eta^{2}+g^{\prime}(v)\eta.

Note that here we use the C2C^{2} smoothness assumption for ff. By Lemma 8.7 we have

|vx(x,t)|M1(m,M00,M10,G0,G1),x,t>0.|v_{x}(x,t)|\leq M_{1}(m,M_{0}^{0},M_{1}^{0},G_{0},G_{1}),\quad x\in{\mathbb{R}},\ t>0.

Hence,

ηt(η)G2(M1+1)2.\eta_{t}\geq\mathcal{L}(\eta)-G_{2}(M_{1}+1)^{2}.

Now we construct a subsolution of this operator. Define

(8.12) τ1:=G1+(m+1)M20(m+1)G2(M1+1)2.\tau_{1}:=\frac{G_{1}+(m+1)M_{2}^{0}}{(m+1)G_{2}(M_{1}+1)^{2}}.

Then ζ(x,t):=G2(M1+1)2(t+τ1)\zeta(x,t):=-G_{2}(M_{1}+1)^{2}(t+\tau_{1}) satisfies

ζt(ζ)+G2(M1+1)2=G2(M1+1)2(t+τ1)[G1(m+1)G2(M1+1)2(t+τ1)]0.\zeta_{t}-\mathcal{L}(\zeta)+G_{2}(M_{1}+1)^{2}=G_{2}(M_{1}+1)^{2}(t+\tau_{1})[G_{1}-(m+1)G_{2}(M_{1}+1)^{2}(t+\tau_{1})]\leq 0.

In addition,

ζ(x,0)=G2(M1+1)2τ1<M20v0′′(x)=η(x,0+),x.\zeta(x,0)=-G_{2}(M_{1}+1)^{2}\tau_{1}<-M_{2}^{0}\leq v^{\prime\prime}_{0}(x)=\eta(x,0+),\quad x\in{\mathbb{R}}.

By comparison we have

vxx(x,t)=η(x,t)ζ(x,t)=G2(M1+1)2(t+τ1),x,t>0.v_{xx}(x,t)=\eta(x,t)\geq\zeta(x,t)=-G_{2}(M_{1}+1)^{2}(t+\tau_{1}),\quad x\in{\mathbb{R}},\ t>0.

2). We now construct a family of approximate positive solutions to approach the nonnegative general solution. Set

u0ε(x):=u0(x)+ε,x, 0<ε1.u_{0\varepsilon}(x):=u_{0}(x)+\varepsilon,\quad x\in{\mathbb{R}},\ 0<\varepsilon\ll 1.

Then the corresponding pressure is v0ε:=mm1u0εm1v_{0\varepsilon}:=\frac{m}{m-1}u_{0\varepsilon}^{m-1}, which satisfies

0<εv0ε(x)M00+1,|v0ε(x)|M10,v0ε′′(x)M20,x.0<\varepsilon\leq v_{0\varepsilon}(x)\leq M_{0}^{0}+1,\quad|v^{\prime}_{0\varepsilon}(x)|\leq M_{1}^{0},\quad v^{\prime\prime}_{0\varepsilon}(x)\geq-M_{2}^{0},\quad x\in{\mathbb{R}}.

According to the standard theory, the problem (CP) with initial data u0εu_{0\varepsilon} has a unique solution uε(x,t)u_{\varepsilon}(x,t). In a similar way as in the previous step, we see that the corresponding pressure vεv_{\varepsilon} satisfies

(8.13) 0<vε(x,t)max{mm1,M00+1},x,t>0, 0<ε1,0<v_{\varepsilon}(x,t)\leq\max\Big{\{}\frac{m}{m-1},M_{0}^{0}+1\Big{\}},\quad x\in{\mathbb{R}},\ t>0,\ 0<\varepsilon\ll 1,

vεxv_{\varepsilon x} satisfies

|vεx(x,t)|M1(m,M00,M10,G0,G1),x,t>0, 0<ε1,|v_{\varepsilon x}(x,t)|\leq M^{\prime}_{1}(m,M_{0}^{0},M_{1}^{0},G_{0},G_{1}),\quad x\in{\mathbb{R}},\ t>0,\ 0<\varepsilon\ll 1,

and vεxxv_{\varepsilon xx} satisfies

(8.14) vεxx(x,t)C(t+τ),x,t(0,T], 0<ε1,v_{\varepsilon xx}(x,t)\geq-C(t+\tau),\quad x\in{\mathbb{R}},\ t\in(0,T],\ 0<\varepsilon\ll 1,

for

C=C(m,M00,M10,G0,G1,G2) and τ=τ(m,M00,M10,M20,G0,G1,G2).C=C(m,M_{0}^{0},M_{1}^{0},G_{0},G_{1},G_{2})\ \ \mbox{ and }\ \ \tau=\tau(m,M_{0}^{0},M_{1}^{0},M_{2}^{0},G_{0},G_{1},G_{2}).

Since vεv_{\varepsilon} is strictly decreasing in ε\varepsilon and vεvv_{\varepsilon}\geq v. There exists v^v0\hat{v}\geq v\geq 0 such that

limε0vε(x,t)=v^(x,t) in the topology of Cloc(QT).\lim\limits_{\varepsilon\to 0}v_{\varepsilon}(x,t)=\hat{v}(x,t)\mbox{ in the topology of }C_{loc}(Q_{T}).

(In particular, v^C(QT)L(QT)\hat{v}\in C(Q_{T})\cap L^{\infty}(Q_{T})). Furthermore, in the area D:={(x,t)QTv^(x,t)>0}D:=\{(x,t)\in Q_{T}\mid\hat{v}(x,t)>0\}, by standard interior parabolic estimates we see that the convergence holds also in the topology Cloc2,1(D)C^{2,1}_{loc}(D). Using the Lebesgue theorem to take limit as ε0\varepsilon\to 0 in

[vεφxx+C(t+τ)φ]𝑑x𝑑t=[vεxx+C(t+τ)]φ𝑑x𝑑t0\iint[v_{\varepsilon}\varphi_{xx}+C(t+\tau)\varphi]dxdt=\iint[v_{\varepsilon xx}+C(t+\tau)]\varphi dxdt\geq 0

for every φCc(QT),φ0\varphi\in C^{\infty}_{c}(Q_{T}),\ \varphi\geq 0, we obtain

[v^φxx+C(t+τ)φ]𝑑x𝑑t0.\iint[\hat{v}\varphi_{xx}+C(t+\tau)\varphi]dxdt\geq 0.

This means that (8.14) holds for v^\hat{v} in the distribution sense.

To complete our proof we only need to show that v^v\hat{v}\equiv v. Denote by

u^=(m1mv^)1m1C(QT)L(QT)\hat{u}=\Big{(}\frac{m-1}{m}\hat{v}\Big{)}^{\frac{1}{m-1}}\in C(Q_{T})\cap L^{\infty}(Q_{T})

the original density variable corresponding to the pressure v^\hat{v}. Recalling that uεu_{\varepsilon} is a classical solution of (CP) with initial data u0εu_{0\varepsilon} we have

uε(x,T)φ(x,T)𝑑x=u0ε(x)φ(x,0)𝑑x+QTf(uε)φ𝑑x𝑑t+QT[uεφt+uεmφxx]𝑑x𝑑t.\int_{{\mathbb{R}}}u_{\varepsilon}(x,T)\varphi(x,T)dx=\int_{{\mathbb{R}}}u_{0\varepsilon}(x)\varphi(x,0)dx+\iint_{Q_{T}}f(u_{\varepsilon})\varphi dxdt+\iint_{Q_{T}}[u_{\varepsilon}\varphi_{t}+u^{m}_{\varepsilon}\varphi_{xx}]dxdt.

Taking limit as ε0\varepsilon\to 0 we have

u^(x,T)φ(x,T)𝑑x=u0(x)φ(x,0)𝑑x+QTf(u^)φ𝑑x𝑑t+QT[u^φt+umφxx]𝑑x𝑑t.\int_{{\mathbb{R}}}\hat{u}(x,T)\varphi(x,T)dx=\int_{{\mathbb{R}}}u_{0}(x)\varphi(x,0)dx+\iint_{Q_{T}}f(\hat{u})\varphi dxdt+\iint_{Q_{T}}[\hat{u}\varphi_{t}+u^{m}\varphi_{xx}]dxdt.

This means that u^\hat{u} is a very weak solution of (CP) with u^(x,0)=u0\hat{u}(x,0)=u_{0}. By the uniqueness, u^\hat{u} must be uu, and so v^v\hat{v}\equiv v. ∎

Finally, we show that vv is also Lipschitz in time.

Lemma 8.10.

Assume the hypotheses in Proposition 8.9 hold. Then for any τ>0\tau>0 there holds

(8.15) vt(x,t)C1t+C2,x,tτ,v_{t}(x,t)\leq C_{1}t+C_{2},\quad x\in{\mathbb{R}},\ t\geq\tau,

where C1C_{1} depends on m,M10,G0m,M^{0}_{1},G_{0} and G1G_{1}, C2C_{2} depends on m,G0,G1,vt(x,τ)m,G_{0},G_{1},v_{t}(x,\tau), and

(8.16) vtC3tC4,x,t>0,v_{t}\geq-C_{3}t-C_{4},\quad x\in{\mathbb{R}},\ t>0,

where C3C_{3} and C4C_{4} depend on m,M00,M10,G0,G1,G2m,M_{0}^{0},M_{1}^{0},G_{0},G_{1},G_{2}.

Proof.

By approximation, we may assume that vv is positive and smooth. In this case we have

0<v1+v0L,|vx(x,t)|M1(m,M10,G0,G1),x,t>0,0<v\leq 1+\|v_{0}\|_{L^{\infty}},\quad|v_{x}(x,t)|\leq M_{1}(m,M_{1}^{0},G_{0},G_{1}),\quad x\in{\mathbb{R}},\ t>0,

as above. We consider the function

P(x,t):=vt+(m1)vx2=(m1)vvxx+mvx2+g(v).P(x,t):=v_{t}+(m-1)v_{x}^{2}=(m-1)vv_{xx}+mv_{x}^{2}+g(v).

Then

Px=(m1)vvxxx+(3m1)vxvxx+g(v)vx,Pxx=vtxx+2(m1)(vxx2+vxvxxx),Pt=(m1)vvtxx+(m1)vtvxx+2m(m+1)vx2vxx+2m(m1)vvxvxx+2mvxg(v)vx+g(v)vt.\begin{array}[]{l}P_{x}=(m-1)vv_{xxx}+(3m-1)v_{x}v_{xx}+g^{\prime}(v)v_{x},\\ P_{xx}=v_{txx}+2(m-1)(v_{xx}^{2}+v_{x}v_{xxx}),\\ P_{t}=(m-1)vv_{txx}+(m-1)v_{t}v_{xx}+2m(m+1)v_{x}^{2}v_{xx}+2m(m-1)vv_{x}v_{xx}\\ \ \ \ \ \ \ \ +2mv_{x}g^{\prime}(v)v_{x}+g^{\prime}(v)v_{t}.\end{array}

Hence we have

(P)\displaystyle\mathcal{L}(P) :=\displaystyle:= Pt(m1)vPxx2vxPx\displaystyle P_{t}-(m-1)vP_{xx}-2v_{x}P_{x}
=\displaystyle= g~(x,t,P):=1v[(Pg)2+A1(Pg)+A2]+A3,\displaystyle\widetilde{g}(x,t,P):=\frac{1}{v}[-(P-g)^{2}+A_{1}(P-g)+A_{2}]+A_{3},

with

A1:=(4m1)vx2+vg,A2:=m(3m1)vx4,A3:=2mvxg(v)vx+g(v)(g(m1)vx2).\begin{array}[]{l}A_{1}:=(4m-1)v_{x}^{2}+vg^{\prime},\qquad A_{2}:=-m(3m-1)v_{x}^{4},\\ A_{3}:=2mv_{x}g^{\prime}(v)v_{x}+g^{\prime}(v)(g-(m-1)v_{x}^{2}).\end{array}

By fC2f\in C^{2} and the bounds of v,vxv,\ v_{x} we see that

|Ai(x,t)|A¯i,x,tτ,i=1,2,3.|A_{i}(x,t)|\leq\bar{A}_{i},\quad x\in{\mathbb{R}},\ t\geq\tau,\ i=1,2,3.

Choose

C1:=A¯3,C2:=G0+A¯1+A¯12+4A¯22+(m1)M12+supxvt(x,τ),C_{1}:=\bar{A}_{3},\quad C_{2}:=G_{0}+\frac{\bar{A}_{1}+\sqrt{\bar{A}_{1}^{2}+4\bar{A}_{2}}}{2}+(m-1)M_{1}^{2}+\sup\limits_{x\in{\mathbb{R}}}v_{t}(x,\tau),

then

(C1t+C2g)2+A1(C1t+C2g)+A20-(C_{1}t+C_{2}-g)^{2}+A_{1}(C_{1}t+C_{2}-g)+A_{2}\leq 0

and so

(C1t+C2)g~(x,t,C1t+C2)0\mathcal{L}(C_{1}t+C_{2})-\widetilde{g}(x,t,C_{1}t+C_{2})\geq 0

and

(C1t+C2)|t=τ>C2>P(x,τ).(C_{1}t+C_{2})|_{t=\tau}>C_{2}>P(x,\tau).

Hence C1t+C2C_{1}t+C_{2} is a supersolution of (P)=g~(x,t,P)\mathcal{L}(P)=\widetilde{g}(x,t,P) in tτt\geq\tau, and so the (8.15) follows from the maximum principle. Note that the auxiliary function PP was used for PME in [8] and [30, Chapter 15], but the supersolution they used is C/t-C/t, different from ours.

The lower bound for vtv_{t} is an immediate consequence of Proposition 8.9:

vt=(m1)vvxx+vx2+g(v)C(m1)(t+τ)(1+v0L)G0.v_{t}=(m-1)vv_{xx}+v_{x}^{2}+g(v)\geq-C(m-1)(t+\tau)(1+\|v_{0}\|_{L^{\infty}})-G_{0}.

This proves the Lipschitz continuity for vv in time. ∎