Abstract

Using the Kato theorem for abstract differential equations, the local well-posedness of the solution for a nonlinear dissipative Camassa-Holm equation is established in space 𝐶([0,𝑇),𝐻𝑠(𝑅))𝐶1([0,𝑇),𝐻𝑠1(𝑅)) with 𝑠>3/2. In addition, a sufficient condition for the existence of weak solutions of the equation in lower order Sobolev space 𝐻𝑠(𝑅) with 1𝑠3/2 is developed.

1. Introduction

Camassa and Holm [1] used the Hamiltonian method to derive a completely integrable wave equation𝑢𝑡𝑢𝑥𝑥𝑡+2𝑘𝑢𝑥+3𝑢𝑢𝑥=2𝑢𝑥𝑢𝑥𝑥+𝑢𝑢𝑥𝑥𝑥,(1.1) by retaining two terms that are usually neglected in the small amplitude, shallow water limit. Its alternative derivation as a model for water waves can be found in Constantin and Lannes [2] and Johnson [3]. Equation (1.1) also models wave current interaction [4], while Dai [5] derived it as a model in elasticity (see Constantin and Strauss [6]). Moreover, it was pointed out in Lakshmanan [7] that the Camassa-Holm equation (1.1) could be relevant to the modeling of tsunami waves (see Constantin and Johnson [8]).

In fact, a huge amount of work has been carried out to investigate the dynamic properties of (1.1). For 𝑘=0, (1.1) has traveling wave solutions of the form 𝑐e|𝑥𝑐𝑡|, called peakons, which capture the main feature of the exact traveling wave solutions of greatest height of the governing equations (see [911]). For 𝑘>0, its solitary waves are stable solitons [6, 11]. It was shown in [1214] that the inverse spectral or scattering approach was a powerful tool to handle Camassa-Holm equation. Equation (1.1) is a completely integrable infinite-dimensional Hamiltonian system (in the sense that for a large class of initial data, the flow is equivalent to a linear flow at constant speed [15]). It should be emphasized that (1.1) gives rise to geodesic flow of a certain invariant metric on the Bott-Virasoro group (see [16, 17]), and this geometric illustration leads to a proof that the Least Action Principle holds. It is worthwhile to mention that Xin and Zhang [18] proved that the global existence of the weak solution in the energy space 𝐻1(𝑅) without any sign conditions on the initial value, and the uniqueness of this weak solution is obtained under some conditions on the solution [19]. Coclite et al. [20] extended the analysis presented in [18, 19] and obtained many useful dynamic properties to other equations (also see [2124]). Li and Olver [25] established the local well-posedness in the Sobolev space 𝐻𝑠(𝑅) with 𝑠>3/2 for (1.1) and gave conditions on the initial data that lead to finite time blowup of certain solutions. It was shown in Constantin and Escher [26] that the blowup occurs in the form of breaking waves, namely, the solution remains bounded but its slope becomes unbounded in finite time. After wave breaking, the solution can be continued uniquely either as a global conservative weak solution [21] or a global dissipative solution [22]. For peakons, these possibilities are explicitly illustrated in the paper [27]. For other methods to handle the problems relating to various dynamic properties of the Camassa-Holm equation and other shallow water models, the reader is referred to [10, 2832] and the references therein.

In this paper, motivated by the work in [25, 33], we study the following generalized Camassa-Holm equation𝑢𝑡𝑢𝑡𝑥𝑥+2𝑘𝑢𝑥+𝑎𝑢𝑚𝑢𝑥=2𝑢𝑥𝑢𝑥𝑥+𝑢𝑢𝑥𝑥𝑥+𝛽𝜕𝑥𝑢𝑥𝑁,(1.2) where 𝑚1 and 𝑁1 are natural numbers, and 𝑎, 𝑘, and 𝛽 are arbitrary constants. Obviously, (1.2) reduces to (1.1) if we set 𝑎=3, 𝑚=1, and 𝛽=0. Actually, Wu and Yin [34] consider a nonlinearly dissipative Camassa-Holm equation which includes a nonlinearly dissipative term 𝐿(𝑢), where 𝐿 is a differential operator or a quasidifferential operator. Therefore, we can regard the term 𝛽𝜕𝑥[(𝑢𝑥)𝑁] as a nonlinearly dissipative term for the dissipative Camassa-Holm equation (1.2).

Due to the term 𝛽𝜕𝑥[(𝑢𝑥)𝑁] in (1.2), the conservation laws in previous works [10, 25] for (1.1) lose their powers to obtain some bounded estimates of the solution for (1.2). A new conservation law different from those presented in [10, 25] will be established to prove the local existence and uniqueness of the solution to (2.3) subject to initial value 𝑢0(𝑥)𝐻𝑠(𝑅) with 𝑠>3/2. We should address that all the generalized versions of the Camassa-Holm equation in previous works (see [17, 25, 34]) do not involve the nonlinear term 𝜕𝑥[(𝑢𝑥)𝑁]. Lai and Wu [33] only studied a generalized Camassa-Holm equation in the case where 𝛽0 and 𝑁 is an odd number. Namely, (1.2) with 𝛽<0 and arbitrary positive integer 𝑁 was not investigated in [33].

The main tasks of this paper are two-fold. Firstly, by using the Kato theorem for abstract differential equations, we establish the local existence and uniqueness of solutions for (1.2) with any 𝛽 and arbitrary positive integer 𝑁 in space 𝐶([0,𝑇),𝐻𝑠𝐶(𝑅))1([0,𝑇),𝐻𝑠1(𝑅)) with 𝑠>3/2. Secondly, it is shown that the existence of weak solutions in lower order Sobolev space 𝐻𝑠(𝑅) with 1𝑠3/2. The ideas of proving the second result come from those presented in Li and Olver [25].

2. Main Results

Firstly, we give some notation.

The space of all infinitely differentiable functions 𝜙(𝑡,𝑥) with compact support in [0,+)×𝑅 is denoted by 𝐶0. 𝐿𝑝=𝐿𝑝(𝑅)(1𝑝<+) is the space of all measurable functions such that 𝑝𝐿𝑝=𝑅|(𝑡,𝑥)|𝑝𝑑𝑥<. We define 𝐿=𝐿(𝑅) with the standard norm 𝐿=inf𝑚(𝑒)=0sup𝑥𝑅𝑒|(𝑡,𝑥)|. For any real number 𝑠, 𝐻𝑠=𝐻𝑠(𝑅) denotes the Sobolev space with the norm defined by𝐻𝑠=𝑅||𝜉||1+2𝑠||||||(𝑡,𝜉)2𝑑𝜉1/2<,(2.1) where (𝑡,𝜉)=𝑅𝑒𝑖𝑥𝜉(𝑡,𝑥)𝑑𝑥.

For 𝑇>0 and nonnegative number 𝑠, 𝐶([0,𝑇);𝐻𝑠(𝑅)) denotes the Frechet space of all continuous 𝐻𝑠-valued functions on [0,𝑇). We set Λ=(1𝜕2𝑥)1/2.

In order to study the existence of solutions for (1.2), we consider its Cauchy problem in the form𝑢𝑡𝑢𝑡𝑥𝑥=2𝑘𝑢𝑥𝑎𝑢𝑚+1𝑚+1𝑥+2𝑢𝑥𝑢𝑥𝑥+𝑢𝑢𝑥𝑥𝑥+𝛽𝜕𝑥𝑢𝑥𝑁=𝑘𝑢𝑥𝑎𝑢𝑚+1𝑚+1𝑥+12𝜕3𝑥𝑢212𝜕𝑥𝑢2𝑥+𝛽𝜕𝑥𝑢𝑥𝑁,𝑢(0,𝑥)=𝑢0(𝑥),(2.2) which is equivalent to𝑢𝑡+𝑢𝑢𝑥=Λ2𝑎𝑘𝑢𝑢𝑚+1𝑚+1𝑥+Λ2𝑢𝑢𝑥12Λ2𝜕𝑥𝑢2𝑥+𝛽Λ2𝜕𝑥𝑢𝑥𝑁,𝑢(0,𝑥)=𝑢0(𝑥).(2.3)

Now, we state our main results.

Theorem 2.1. Let 𝑢0(𝑥)𝐻𝑠(𝑅) with 𝑠>3/2. Then problem (2.2) or problem (2.3) has a unique solution 𝑢(𝑡,𝑥)𝐶([0,𝑇);𝐻𝑠𝐶(𝑅))1([0,𝑇);𝐻𝑠1(𝑅)) where 𝑇>0 depends on 𝑢0𝐻𝑠(𝑅).

Theorem 2.2. Suppose that 𝑢0(𝑥)𝐻𝑠 with 1𝑠3/2 and 𝑢0𝑥𝐿<. Then there exists a 𝑇>0 such that (1.2) subject to initial value 𝑢0(𝑥) has a weak solution 𝑢(𝑡,𝑥)𝐿2([0,𝑇],𝐻𝑠) in the sense of distribution and 𝑢𝑥𝐿([0,𝑇]×𝑅).

3. Local Well-Posedness

We consider the abstract quasilinear evolution equation𝑑𝑣𝑑𝑡+𝐴(𝑣)𝑣=𝑓(𝑣),𝑡0,𝑣(0)=𝑣0.(3.1) Let 𝑋 and 𝑌 be Hilbert spaces such that 𝑌 is continuously and densely embedded in 𝑋, and let 𝑄𝑌𝑋 be a topological isomorphism. Let 𝐿(𝑌,𝑋) be the space of all bounded linear operators from 𝑌 to 𝑋. If 𝑋=𝑌, we denote this space by 𝐿(𝑋). We state the following conditions in which 𝜌1,𝜌2,𝜌3, and 𝜌4 are constants depending on max{𝑦𝑌,𝑧𝑌}.(i)𝐴(𝑦)𝐿(𝑌,𝑋) for 𝑦𝑋 with (𝐴(𝑦)𝐴(𝑧))𝑤𝑋𝜌1𝑦𝑧𝑋𝑤𝑌,𝑦,𝑧,𝑤𝑌,(3.2) and 𝐴(𝑦)𝐺(𝑋,1,𝛽) (i.e., 𝐴(𝑦) is quasi-𝑚-accretive), uniformly on bounded sets in 𝑌.(ii)𝑄𝐴(𝑦)𝑄1=𝐴(𝑦)+𝐵(𝑦), where 𝐵(𝑦)𝐿(𝑋) is bounded, uniformly on bounded sets in 𝑌. Moreover, (𝐵(𝑦)𝐵(𝑧))𝑤𝑋𝜌2𝑦𝑧𝑌𝑤𝑋,𝑦,𝑧𝑌,𝑤𝑋.(3.3)(iii)𝑓𝑌𝑌 extends to a map from 𝑋 into 𝑋 is bounded on bounded sets in 𝑌, and satisfies𝑓(𝑦)𝑓(𝑧)𝑌𝜌3𝑦𝑧𝑌,𝑦,𝑧𝑌,𝑓(𝑦)𝑓(𝑧)𝑋𝜌4𝑦𝑧𝑋,𝑦,𝑧𝑌.(3.4)

Kato Theorem (see [35])
Assume that (i), (ii), and (iii) hold. If 𝑣0𝑌, there is a maximal 𝑇>0 depending only on 𝑣0𝑌, and a unique solution 𝑣 to problem (3.1) such that 𝑣=𝑣,𝑣0[𝐶𝐶(0,𝑇);𝑌)1([0,𝑇);𝑋).(3.5) Moreover, the map 𝑣0𝑣(,𝑣0) is a continuous map from 𝑌 to the space [𝐶𝐶(0,𝑇);𝑌)1([0,𝑇);𝑋).(3.6)

For problem (2.3), we set 𝐴(𝑢)=𝑢𝜕𝑥, 𝑌=𝐻𝑠(𝑅), 𝑋=𝐻𝑠1(𝑅), Λ=(1𝜕2𝑥)1/2, 𝑓(𝑢)=Λ2𝑎𝑘𝑢𝑢𝑚+1𝑚+1𝑥+Λ2𝑢𝑢𝑥12Λ2𝜕𝑥𝑢2𝑥+𝛽Λ2𝜕𝑥𝑢𝑥𝑁,(3.7) and 𝑄=Λ. In order to prove Theorem 2.1, we only need to check that 𝐴(𝑢) and 𝑓(𝑢) satisfy assumptions (i)–(iii).

Lemma 3.1. The operator 𝐴(𝑢)=𝑢𝜕𝑥 with 𝑢𝐻𝑠(𝑅), 𝑠>3/2 belongs to 𝐺(𝐻𝑠1,1,𝛽).

Lemma 3.2. Let 𝐴(𝑢)=𝑢𝜕𝑥 with 𝑢𝐻𝑠 and 𝑠>3/2. Then 𝐴(𝑢)𝐿(𝐻𝑠,𝐻𝑠1) for all 𝑢𝐻𝑠. Moreover, (𝐴(𝑢)𝐴(𝑧))𝑤𝐻𝑠1𝜌1𝑢𝑧𝐻𝑠1𝑤𝐻𝑠,𝑢,𝑧,𝑤𝐻𝑠(𝑅).(3.8)

Lemma 3.3. For 𝑠>3/2, 𝑢,𝑧𝐻𝑠 and 𝑤𝐻𝑠1, it holds that 𝐵(𝑢)=[Λ,𝑢𝜕𝑥]Λ1𝐿(𝐻𝑠1) for 𝑢𝐻𝑠 and (𝐵(𝑢)𝐵(𝑧))𝑤𝐻𝑠1𝜌2𝑢𝑧𝐻𝑠𝑤𝐻𝑠1.(3.9)

Proofs of the above Lemmas 3.13.3 can be found in [29] or [31].

Lemma 3.4 (see [35]). Let 𝑟 and 𝑞 be real numbers such that 𝑟<𝑞𝑟. Then 𝑢𝑣𝐻𝑞𝑐𝑢𝐻𝑟𝑣𝐻𝑞1,if𝑟>2,𝑢𝑣𝐻𝑟+𝑞1/2𝑐𝑢𝐻𝑟𝑣𝐻𝑞1,if𝑟<2.(3.10)

Lemma 3.5. Let 𝑢,𝑧𝐻𝑠 with 𝑠>3/2, then 𝑓(𝑢) is bounded on bounded sets in 𝐻𝑠 and satisfies 𝑓(𝑢)𝑓(𝑧)𝐻𝑠𝜌3𝑢𝑧𝐻𝑠,(3.11)𝑓(𝑢)𝑓(𝑧)𝐻𝑠1𝜌4𝑢𝑧𝐻𝑠1.(3.12)

Proof. Using the algebra property of the space 𝐻𝑠0 with 𝑠0>1/2, we have 𝑓(𝑢)𝑓(𝑧)𝐻𝑠Λ𝑐2𝑎𝑘𝑢𝑢𝑚+1𝑚+1𝑥𝑎𝑘𝑧𝑧𝑚+1𝑚+1𝑥𝐻𝑠+Λ2𝑢𝑢𝑥𝑧𝑧𝑥𝐻𝑠+Λ2𝜕𝑥𝑢2𝑥𝑧2𝑥𝐻𝑠+Λ2𝜕𝑥𝑢𝑥𝑁Λ2𝜕𝑥𝑧𝑥𝑁𝐻𝑠𝑐𝑢𝑧𝐻𝑠1+𝑢𝑚+1𝑧𝑚+1𝐻𝑠1+𝑢𝑢𝑥𝑧𝑧𝑥𝐻𝑠1+𝑢2𝑥𝑧2𝑥𝐻𝑠1+(𝑢𝑥)𝑁(𝑧𝑥)𝑁𝐻𝑠1𝑐𝑢𝑧𝐻𝑠1+𝑚𝑗=0𝑢𝐻𝑚𝑗𝑠𝑧𝑗𝐻𝑠+𝑢𝐻𝑠+𝑧𝐻𝑠+𝑁1𝑗=0𝑢𝑥𝐻𝑁𝑗𝑠1𝑧𝑥𝑗𝐻𝑠1𝜌3𝑢𝑧𝐻𝑠,(3.13) from which we obtain (3.11).
Applying Lemma 3.4, 𝑢𝑢𝑥=(1/2)(𝑢2)𝑥, 𝑠>3/2,𝑢𝐿𝑐𝑢𝐻𝑠1 and 𝑢𝑥𝐿𝑐𝑢𝐻𝑠, we get 𝑓(𝑢)𝑓(𝑧)𝐻𝑠1𝑐𝑢𝑧𝐻𝑠2+𝑢𝑚+1𝑧𝑚+1𝐻𝑠2+𝑢2𝑧2𝐻𝑠2+(𝑢𝑥𝑧𝑥)(𝑢𝑥+𝑧𝑥)𝐻𝑠2+(𝑢𝑥𝑧𝑥)𝑁1𝑗=0𝑢𝑥𝑁1𝑗𝑧𝑗𝑥𝐻𝑠2𝑐𝑢𝑧𝐻𝑠11+𝑚𝑗=0𝑢𝐻𝑚𝑗𝑠1𝑧𝑗𝐻𝑠1+𝑢𝐻𝑠1+𝑧𝐻𝑠1+𝑢𝐻𝑠+𝑧𝐻𝑠+𝑁1𝑗=0𝑢𝑥𝐻𝑁𝑗𝑠1𝑧𝑥𝑗𝐻𝑠1𝜌4𝑢𝑧𝐻𝑠1,(3.14) which completes the proof of (3.12).

Proof of Theorem 2.1. Using the Kato Theorem, Lemmas 3.13.3, and 3.5, we know that system (2.2) or problem (2.3) has a unique solution [𝑢(𝑡,𝑥)𝐶(0,𝑇);𝐻𝑠𝐶(𝑅))1[0,𝑇);𝐻𝑠1(𝑅).(3.15)

4. Existence of Weak Solutions

For 𝑠2, using the first equation of system (2.2) derives𝑑𝑑𝑡𝑅𝑢2+𝑢2𝑥+2𝛽𝑡0𝑢𝑥𝑁+1𝑑𝜏𝑑𝑥=0,(4.1) from which we have the conservation law𝑅𝑢2+𝑢2𝑥+2𝛽𝑡0𝑢𝑥𝑁+1𝑑𝜏𝑑𝑥=𝑅𝑢20+𝑢20𝑥𝑑𝑥.(4.2)

Lemma 4.1 (Kato and Ponce [36]). If 𝑟>0, then 𝐻𝑟𝐿 is an algebra. Moreover, 𝑢𝑣𝑟𝑐𝑢𝐿𝑣𝑟+𝑢𝑟𝑣𝐿,(4.3) where 𝑐 is a constant depending only on 𝑟.

Lemma 4.2 (Kato and Ponce [36]). Let 𝑟>0. If 𝑢𝐻𝑟𝑊1, and 𝑣𝐻𝑟1𝐿, then Λ𝑟𝑣,𝑢𝐿2𝜕𝑐𝑥𝑢𝐿Λ𝑟1𝑣𝐿2+Λ𝑟𝑢𝐿2𝑣𝐿.(4.4)

Lemma 4.3. Let 𝑠2 and the function 𝑢(𝑡,𝑥) is a solution of problem (2.2) and the initial data 𝑢0(𝑥)𝐻𝑠(𝑅). Then the following inequality holds 𝑢𝐿𝑢𝐻1𝑢0𝐻1𝑒|𝛽|𝑡0𝑢𝑥𝐿𝑁1𝑑𝜏.(4.5)
For 𝑞(0,𝑠1], there is a constant 𝑐, which only depends on 𝑚, 𝑁, 𝑘, 𝑎, and 𝛽, such that 𝑅Λ𝑞+1𝑢2𝑑𝑥𝑅Λ𝑞+1𝑢02𝑑𝑥+𝑐𝑡0𝑢𝑥𝐿𝑢2𝐻𝑞+11+𝑢𝐿𝑚1𝑑𝜏+𝑐𝑡0𝑢2𝐻𝑞+1𝑢𝑥𝐿𝑁1𝑑𝜏.(4.6)
For 𝑞[0,𝑠1], there is a constant 𝑐, which only depends on 𝑚,𝑁,𝑘,𝑎, and 𝛽, such that 𝑢𝑡𝐻𝑞𝑐𝑢𝐻𝑞+11+1+𝑢𝐿𝑚1𝑢𝐻1+𝑢𝑥𝐿𝑁1.(4.7)

Proof. Using 𝑢2𝐻1=𝑅(𝑢2+𝑢2𝑥)𝑑𝑥 and (4.2) derives (4.5).
Using 𝜕2𝑥=Λ2+1 and the Parseval equality gives rise to 𝑅Λ𝑞𝑢Λ𝑞𝜕3𝑥𝑢2𝑑𝑥=2𝑅Λ𝑞+1𝑢Λ𝑞+1𝑢𝑢𝑥𝑑𝑥+2𝑅(Λ𝑞𝑢)Λ𝑞𝑢𝑢𝑥𝑑𝑥.(4.8)
For 𝑞(0,𝑠1], applying (Λ𝑞𝑢)Λ𝑞 to both sides of the first equation of system (2.3) and integrating with respect to 𝑥 by parts, we have the identity 12𝑑𝑑𝑡𝑅(Λ𝑞𝑢)2+Λ𝑞𝑢𝑥2𝑑𝑥=𝑎𝑅(Λ𝑞𝑢)Λ𝑞𝑢𝑚𝑢𝑥𝑑𝑥𝑅Λ𝑞+1𝑢Λ𝑞+1𝑢𝑢𝑥1𝑑𝑥+2𝑅Λ𝑞𝑢𝑥Λ𝑞𝑢2𝑥+𝑑𝑥𝑅(Λ𝑞𝑢)Λ𝑞𝑢𝑢𝑥𝑑𝑥𝛽𝑅Λ𝑞𝑢𝑥Λ𝑞𝑢𝑥𝑁𝑑𝑥.(4.9) We will estimate the terms on the right-hand side of (4.9) separately. For the first term, by using the Cauchy-Schwartz inequality and Lemmas 4.1 and 4.2, we have 𝑅(Λ𝑞𝑢)Λ𝑞𝑢𝑚𝑢𝑥𝑑𝑥=𝑅(Λ𝑞Λ𝑢)𝑞𝑢𝑚𝑢𝑥𝑢𝑚Λ𝑞𝑢𝑥𝑑𝑥+𝑅(Λ𝑞𝑢)𝑢𝑚Λ𝑞𝑢𝑥𝑑𝑥𝑐𝑢𝐻𝑞𝑚𝑢𝐿𝑚1𝑢𝑥𝐿𝑢𝐻𝑞+𝑢𝑥𝐿𝑢𝐿𝑚1𝑢𝐻𝑞+12𝑢𝐿𝑚1𝑢𝑥𝐿Λ𝑞𝑢2𝐿2𝑐𝑢2𝐻𝑞𝑢𝐿𝑚1𝑢𝑥𝐿.(4.10) Using the above estimate to the second term yields 𝑅Λ𝑞+1𝑢Λ𝑞+1𝑢𝑢𝑥𝑑𝑥𝑐𝑢2𝐻𝑞+1𝑢𝑥𝐿.(4.11) For the third term, using the Cauchy-Schwartz inequality and Lemma 4.1, we obtain 𝑅Λ𝑞𝑢𝑥Λ𝑞𝑢2𝑥Λ𝑑𝑥𝑞𝑢𝑥𝐿2Λ𝑞𝑢2𝑥𝐿2𝑐𝑢𝐻𝑞+1𝑢𝑥𝐿𝑢𝑥𝐻𝑞+𝑢𝑥𝐿𝑢𝑥𝐻𝑞𝑐𝑢2𝐻𝑞+1𝑢𝑥𝐿.(4.12)
For the last term in (4.9), using Lemma 4.1 repeatedly results in ||||𝑅Λ𝑞𝑢𝑥Λ𝑞𝑢𝑥𝑁||||𝑢𝑑𝑥𝑥𝐻𝑞𝑢𝑁𝑥𝐻𝑞𝑐𝑢2𝐻𝑞+1𝑢𝑥𝐿𝑁1.(4.13) It follows from (4.9) to (4.13) that there exists a constant 𝑐 depending only on 𝑚,𝑁 and the coefficients of (1.2) such that 12𝑑𝑑𝑡𝑅(Λ𝑞𝑢)2+Λ𝑞𝑢𝑥2𝑢𝑑𝑥𝑐𝑥𝐿𝑢2𝐻𝑞+11+𝑢𝐿𝑚1+𝑐𝑢2𝐻𝑞+1𝑢𝑥𝐿𝑁1.(4.14) Integrating both sides of the above inequality with respect to 𝑡 results in inequality (4.6).
To estimate the norm of 𝑢𝑡, we apply the operator (1𝜕2𝑥)1 to both sides of the first equation of system (2.3) to obtain the equation 𝑢𝑡=1𝜕2𝑥12𝑘𝑢𝑥+𝜕𝑥𝑎𝑢𝑚+1𝑚+1+12𝜕2𝑥𝑢212𝑢2𝑥+𝛽𝜕𝑥𝑢𝑥𝑁.(4.15) Applying (Λ𝑞𝑢𝑡)Λ𝑞 to both sides of (4.15) for 𝑞(0,𝑠1] gives rise to 𝑅Λ𝑞𝑢𝑡2𝑑𝑥=𝑅Λ𝑞𝑢𝑡Λ𝑞2𝜕𝑥𝑎2𝑘𝑢𝑢𝑚+1𝑚+1+12𝜕2𝑥𝑢212𝑢2𝑥+𝛽𝜕𝑥𝑢𝑥𝑁𝑑𝜏.(4.16) For the right-hand side of (4.16), we have 𝑅Λ𝑞𝑢𝑡Λ𝑞22𝑘𝑢𝑥𝑢𝑑𝑥𝑐𝑡𝐻𝑞𝑢𝐻𝑞,𝑅Λ𝑞𝑢𝑡1𝜕2𝑥1Λ𝑞𝜕𝑥𝑎𝑢𝑚+1𝑚+112𝑢2𝑥𝑢𝑑𝑥𝑐𝑡𝐻𝑞𝑅1+𝜉2𝑞1×𝑅𝑎𝑢𝑚+1𝑚(1𝜉𝜂)̂𝑢(𝜂)2𝑢𝑥(𝜉𝜂)𝑢𝑥(𝜂)𝑑𝜂21/2𝑢𝑐𝑡𝐻𝑞𝑢𝐻1𝑢𝐻𝑞+11+𝑢𝐿𝑚1.(4.17) Since Λ𝑞𝑢𝑡1𝜕2𝑥1Λ𝑞𝜕2𝑥𝑢𝑢𝑥Λ𝑑𝑥=𝑞𝑢𝑡Λ𝑞𝑢𝑢𝑥Λ𝑑𝑥+𝑞𝑢𝑡1𝜕2𝑥1Λ𝑞𝑢𝑢𝑥𝑑𝑥,(4.18) using Lemma 4.1, 𝑢𝑢𝑥𝐻𝑞𝑐(𝑢2)𝑥𝐻𝑞𝑐𝑢𝐿𝑢𝐻𝑞+1 and 𝑢𝐿𝑢𝐻1, we have Λ𝑞𝑢𝑡Λ𝑞𝑢𝑢𝑥𝑢𝑑𝑥𝑐𝑡𝐻𝑞𝑢𝑢𝑥𝐻𝑞𝑢𝑐𝑡𝐻𝑞𝑢𝐻1𝑢𝐻𝑞+1,Λ𝑞𝑢𝑡1𝜕2𝑥1Λ𝑞𝑢𝑢𝑥𝑢𝑑𝑥𝑐𝑡𝐻𝑞𝑢𝐻1𝑢𝐻𝑞+1.(4.19) Using the Cauchy-Schwartz inequality and Lemma 4.1 yields ||||𝑅Λ𝑞𝑢𝑡1𝜕2𝑥1Λ𝑞𝜕𝑥𝑢𝑁𝑥||||𝑢𝑑𝑥𝑐𝑡𝐻𝑞𝑢𝑥𝐿𝑁1𝑢𝐻𝑞+1.(4.20) Applying (4.17)–(4.20) into (4.16) yields the inequality 𝑢𝑡𝐻𝑞𝑐𝑢𝐻𝑞+11+1+𝑢𝐿𝑚1𝑢𝐻1+𝑢𝑥𝐿𝑁1.(4.21) This completes the proof of Lemma 4.3.

Defining𝑒𝜙(𝑥)=1/(𝑥21),|𝑥|<1,0,|𝑥|1,(4.22) and setting 𝜙𝜀(𝑥)=𝜀1/4𝜙(𝜀1/4𝑥) with 0<𝜀<1/4 and 𝑢𝜀0=𝜙𝜀𝑢0, we know that 𝑢𝜀0𝐶 for any 𝑢0𝐻𝑠(𝑅) and 𝑠>0.

It follows from Theorem 2.1 that for each 𝜀 the Cauchy problem𝑢𝑡𝑢𝑡𝑥𝑥=𝜕𝑥𝑎2𝑘𝑢𝑢𝑚+1𝑚+1+12𝜕3𝑥𝑢212𝜕𝑥𝑢2𝑥+𝛽𝜕𝑥𝑢𝑥𝑁,𝑢(0,𝑥)=𝑢𝜀0(𝑥),𝑥𝑅,(4.23) has a unique solution 𝑢𝜀(𝑡,𝑥)𝐶([0,𝑇);𝐻).

Lemma 4.4. Under the assumptions of problem (4.23), the following estimates hold for any 𝜀 with 0<𝜀<1/4 and 𝑠>0𝑢𝜀0𝑥𝐿𝑐1𝑢0𝑥𝐿,𝑢𝜀0𝐻𝑞𝑐1𝑢,if𝑞𝑠,𝜀0𝐻𝑞𝑐1𝜀(𝑠𝑞)/4𝑢,if𝑞>𝑠,𝜀0𝑢0𝐻𝑞𝑐1𝜀(𝑠𝑞)/4𝑢,if𝑞𝑠,𝜀0𝑢0𝐻𝑠=𝑜(1),(4.24) where 𝑐1 is a constant independent of 𝜀.

The proof of this Lemma can be found in Lai and Wu [33].

Lemma 4.5. If 𝑢0(𝑥)𝐻𝑠(𝑅) with 𝑠[1,3/2] such that 𝑢0𝑥𝐿<. Let 𝑢𝜀0 be defined as in system (4.23). Then there exist two positive constants 𝑇 and 𝑐, which are independent of 𝜀, such that the solution 𝑢𝜀 of problem (4.23) satisfies 𝑢𝜀𝑥𝐿𝑐 for any 𝑡[0,𝑇).

Proof. Using notation 𝑢=𝑢𝜀 and differentiating both sides of the first equation of problem (4.23) or (4.15) with respect to 𝑥 give rise to 𝑢𝑡𝑥+12𝜕2𝑥𝑢212𝑢2𝑥𝑎=2𝑘𝑢+𝑢𝑚+1𝑚+112𝑢2𝛽𝑢𝑁𝑥Λ2𝑎2𝑘𝑢+𝑢𝑚+1𝑚+112𝑢2+12𝑢2𝑥𝛽𝑢𝑁𝑥.(4.25) Letting 𝑝>0 be an integer and multiplying the above equation by (𝑢𝑥)2𝑝+1 and then integrating the resulting equation with respect to 𝑥 yield the equality 1𝑑2𝑝+2𝑑𝑡𝑅𝑢𝑥2𝑝+2𝑝𝑑𝑥+2𝑝+2𝑅𝑢𝑥2𝑝+3=𝑑𝑥𝑅𝑢𝑥2𝑝+1𝑎2𝑘𝑢+𝑢𝑚+1𝑚+112𝑢2𝛽𝑢𝑁𝑥𝑑𝑥𝑅𝑢𝑥2𝑝+1Λ2𝑎2𝑘𝑢+𝑢𝑚+1𝑚+1𝑢22+12𝑢2𝑥𝛽𝑢𝑁𝑥𝑑𝑥.(4.26) Applying the Hölder's inequality yields 1𝑑2𝑝+2𝑑𝑡𝑅𝑢𝑥2𝑝+2||||𝑑𝑥2𝑘𝑅|𝑢|2𝑝+2𝑑𝑥1/(2𝑝+2)+𝑎𝑚+1𝑅||𝑢𝑚+1||2𝑝+2𝑑𝑥1/(2𝑝+2)+12𝑅||𝑢2||2𝑝+2𝑑𝑥1/(2𝑝+2)+𝛽𝑅||𝑢𝑁𝑥||2𝑝+2𝑑𝑥1/(2𝑝+2)+𝑅||𝐺||2𝑝+2𝑑𝑥1/(2𝑝+2)𝑅||𝑢𝑥||2𝑝+2𝑑𝑥(2𝑝+1)/(2𝑝+2)+𝑝𝑢2𝑝+2𝑥𝐿𝑅||𝑢𝑥||2𝑝+2𝑑𝑥,(4.27) or 𝑑𝑑𝑡𝑅𝑢𝑥2𝑝+2𝑑𝑥1/(2𝑝+2)||||2𝑘𝑅|𝑢|2𝑝+2𝑑𝑥1/(2𝑝+2)+𝑎𝑚+1𝑅||𝑢𝑚+1||2𝑝+2𝑑𝑥1/(2𝑝+2)+12𝑅||𝑢2||2𝑝+2𝑑𝑥1/(2𝑝+2)+𝛽𝑅||𝑢𝑁𝑥||2𝑝+2𝑑𝑥1/(2𝑝+2)+𝑅||𝐺||2𝑝+2𝑑𝑥1/(2𝑝+2)+𝑝𝑢2𝑝+2𝑥𝐿𝑅||𝑢𝑥||2𝑝+2𝑑𝑥1/(2𝑝+2),(4.28) where 𝐺=Λ2𝑎2𝑘𝑢+𝑢𝑚+1𝑚+1𝑢22+12𝑢2𝑥𝛽𝑢𝑁𝑥.(4.29) Since 𝑓𝐿𝑝𝑓𝐿 as 𝑝 for any 𝑓𝐿𝐿2, integrating both sides of the inequality (4.28) with respect to 𝑡 and taking the limit as 𝑝 result in the estimate 𝑢𝑥𝐿𝑢0𝑥𝐿+𝑡0𝑐𝑢𝐿+𝑢2𝐿+𝑢𝑚+1𝐿𝑢+𝛽𝑥𝑁𝐿+𝐺𝐿+12𝑢𝑥2𝐿𝑑𝜏.(4.30) Using the algebra property of 𝐻𝑠0(𝑅) with 𝑠0>1/2 yields (𝑢𝜀𝐻(1/2)+ means that there exists a sufficiently small 𝛿>0 such that 𝑢𝜀(1/2)+=𝑢𝜀𝐻1/2+𝛿)𝐺𝐿𝑐𝐺𝐻(1/2)+Λ𝑐2𝑎2𝑘𝑢+𝑢𝑚+1𝑚+1𝑢22+12𝑢2𝑥𝛽𝑢𝑁𝑥𝐻(1/2)+𝑐𝑢𝐻1+𝑢2𝐻1+𝑢𝐻𝑚+11+Λ2(𝑢2𝑥)𝐻(1/2)++Λ2(𝑢𝑁𝑥)𝐻(1/2)+𝑐𝑢𝐻1+𝑢2𝐻1+𝑢𝐻𝑚+11+𝑢2𝑥𝐻0+𝑢𝑁𝑥𝐻0𝑐𝑢𝐻1+𝑢2𝐻1+𝑢𝐻𝑚+11+𝑢𝑥𝐿𝑢𝐻1+𝑢𝑥𝐿𝑁1𝑢𝐻1𝑐𝑒𝑐𝑡0𝑢𝑥𝐿𝑁1𝑑𝜏𝑢1+𝑥𝐿+𝑢𝑥𝐿𝑁1,(4.31) in which Lemma 4.3 is used. Therefore, we get 𝑡0𝐺𝐿𝑑𝜏𝑐𝑡0𝑒𝑐𝜏0𝑢𝑥𝐿𝑁1𝑑𝜉𝑢1+𝑥𝐿+𝑢𝑥𝐿𝑁1𝑑𝜏.(4.32)
From (4.30) and (4.32), one has 𝑢𝑥𝐿𝑢0𝑥𝐿+𝑐𝑡0𝑢𝑥2𝐿+𝑢𝑥𝑁𝐿+𝑒𝑐𝑡0𝑢𝑥𝐿𝑁1𝑑𝜏+𝑒𝑐𝜏0𝑢𝑥𝐿𝑁1𝑑𝜉𝑢1+𝑥𝐿+𝑢𝑥𝐿𝑁1𝑑𝜏.(4.33)
From Lemma 4.4, it follows from the contraction mapping principle that there is a 𝑇>0 such that the equation 𝑊𝐿=𝑢0𝑥𝐿+𝑐𝑡0𝑊2𝐿+𝑊𝑁𝐿+𝑒𝑐𝑡0𝑊𝐿𝑁1𝑑𝜏+𝑒𝑐𝜏0𝑊𝐿𝑁1𝑑𝜉1+𝑊𝐿+𝑊𝐿𝑁1𝑑𝜏(4.34) has a unique solution 𝑊𝐶[0,𝑇]. Using the Theorem presented at page 51 in [25] or Theorem 2 in Section 1.1 presented in [37] yields that there are constants 𝑇>0 and 𝑐>0 independent of 𝜀 such that 𝑢𝑥𝐿𝑊(𝑡) for arbitrary 𝑡[0,𝑇], which leads to the conclusion of Lemma 4.5.
Using Lemmas 4.3 and 4.5, notation 𝑢𝜀=𝑢 and Gronwall's inequality results in the inequalities 𝑢𝜀𝐻𝑞𝐶𝑇𝑒𝐶𝑇,𝑢𝜀𝑡𝐻𝑟𝐶𝑇𝑒𝐶𝑇,(4.35) where 𝑞(0,𝑠], 𝑟(0,𝑠1] and 𝐶𝑇 depends on 𝑇. It follows from Aubin's compactness theorem that there is a subsequence of {𝑢𝜀}, denoted by {𝑢𝜀𝑛}, such that {𝑢𝜀𝑛} and their temporal derivatives {𝑢𝜀𝑛𝑡} are weakly convergent to a function 𝑢(𝑡,𝑥) and its derivative 𝑢𝑡 in 𝐿2([0,𝑇],𝐻𝑠) and 𝐿2([0,𝑇],𝐻𝑠1), respectively. Moreover, for any real number 𝑅1>0, {𝑢𝜀𝑛} is convergent to the function 𝑢 strongly in the space 𝐿2([0,𝑇],𝐻𝑞(𝑅1,𝑅1)) for 𝑞[0,𝑠) and {𝑢𝜀𝑛𝑡} converges to 𝑢𝑡 strongly in the space 𝐿2([0,𝑇],𝐻𝑟(𝑅1,𝑅1)) for 𝑟[0,𝑠1]. Thus, we can prove the existence of a weak solution to (2.2).

Proof of Theorem 2.2. From Lemma 4.5, we know that {𝑢𝜀𝑛𝑥}(𝜀𝑛0) is bounded in the space 𝐿. Thus, the sequences {𝑢𝜀𝑛} and {𝑢𝜀𝑛𝑥} are weakly convergent to 𝑢 and 𝑢𝑥 in 𝐿2[0,𝑇],𝐻𝑟(𝑅,𝑅) for any 𝑟[0,𝑠1), respectively. Therefore, 𝑢 satisfies the equation 𝑇0𝑅𝑢𝑔𝑡𝑔𝑥𝑥𝑡𝑑𝑥𝑑𝑡=𝑇0𝑅𝑎2𝑘𝑢+𝑢𝑚+1𝑚+1+12𝑢2𝑥𝑔𝑥12𝑢2𝑔𝑥𝑥𝑥𝑢𝛽𝑥𝑁𝑔𝑥𝑑𝑥𝑑𝑡,(4.36) with 𝑢(0,𝑥)=𝑢0(𝑥) and 𝑔𝐶0. Since 𝑋=𝐿1([0,𝑇]×𝑅) is a separable Banach space and {𝑢𝜀𝑛𝑥} is a bounded sequence in the dual space 𝑋=𝐿([0,𝑇]×𝑅) of 𝑋, there exists a subsequence of {𝑢𝜀𝑛𝑥}, still denoted by {𝑢𝜀𝑛𝑥}, weakly star convergent to a function 𝑣 in 𝐿([0,𝑇]×𝑅). It derives from the {𝑢𝜀𝑛𝑥} weakly convergent to 𝑢𝑥 in 𝐿2([0,𝑇]×𝑅) that 𝑢𝑥=𝑣 almost everywhere. Thus, we obtain 𝑢𝑥𝐿([0,𝑇]×𝑅).

Acknowledgments

The author is very grateful to the reviewers for their helpful and valuable comments, which have led to a meaningful improvement of the paper. This work is supported by the Key Project of Chinese Ministry of Education (109140).