Abstract

This paper is concerned with time-periodic solution of the weakly dissipative Camassa-Holm equation with a periodic boundary condition. The existence and uniqueness of a time periodic solution is presented.

1. Introduction

The Camassa-Holm equationπ‘’π‘‘βˆ’π‘’π‘‘π‘₯π‘₯+3𝑒𝑒π‘₯=2𝑒π‘₯𝑒π‘₯π‘₯+𝑒𝑒π‘₯π‘₯π‘₯,𝑑>0,π‘₯βˆˆβ„,(1.1) modeling the unidirectional propagation of shallow water waves over a flat bottom, where 𝑒(𝑑,π‘₯) represents the fluid’s free surface above a flat bottom (or equivalently, the fluid velocity at time 𝑑β‰₯0 and in the spatial π‘₯ direction).

Since the equation was derived physically by Camassa and Holm [1, 2], many researchers have paid extensive attention to it. The Camassa-Holm equation is also a model for the propagation of axially symmetric waves in hyperelastic rods [3, 4]. It has a bi-Hamiltonian structure [5, 6] and is completely integrable [1, 2, 7–11]. It is a reexpression of geodesic flow on the diffeomorphism group of the circle [12] and on the Virasoro group [13]. Its solitary waves are peaked [7], and they are orbitally stable and interact like solitons [14–16]. The peakons capture a characteristic of the traveling waves of greatest height-exact traveling solutions of the governing equations for water waves with a peak at their crest [17–19].

The Cauchy problem of the Camassa-Holm equation has been extensively studied. It has been shown that this equation is locally well posed [20–25] for initial data 𝑒0βˆˆπ»π‘ (ℝ) with 𝑠>3/2. Moreover, it has global strong solutions modeling permanent waves [20, 24–27] but also blow-up solutions modeling wave breaking [20–28]. On the other hand, it has global weak solutions with initial data 𝑒0∈𝐻1 [29–35]. Moreover, the initial-boundary value problem for the Camassa-Holm equation on the half-line and on a finite interval was discussed in [36, 37]. It is observed that if 𝑒 is the solution of the Camassa-Holm equation with the initial data 𝑒0 in 𝐻1(ℝ), we have for all 𝑑>0,‖𝑒(𝑑,β‹…)β€–πΏβˆžβ‰€βˆš22‖𝑒(𝑑,β‹…)‖𝐻1β‰€βˆš22‖‖𝑒0β€–β€–(β‹…)𝐻1.(1.2)

It is worth pointing out that the advantage of the Camassa-Holm equation in comparison with the KdV equation lies in the fact that the Camassa-Holm equation has peaked solitons and models wave breaking [2, 20, 21].

In general, it is difficult to avoid energy dissipation mechanisms in a real world. Ott and Sudan [38] investigated how the KdV equation was modified by the presence of dissipation and the effect of such dissipation on the solitary solution of the KdV equation. Ghidaglia [39] investigated the long-time behavior of solutions to the weakly dissipative KdV equation as a finite-dimensional dynamical system.

The Camassa-Holm equation with dissipative term isπ‘’π‘‘βˆ’π‘’π‘‘π‘₯π‘₯+3𝑒𝑒π‘₯βˆ’2𝑒π‘₯𝑒π‘₯π‘₯βˆ’π‘’π‘’π‘₯π‘₯π‘₯+𝐿(𝑒)=𝑓(𝑑,π‘₯),𝑑>0,π‘₯βˆˆβ„,(1.3) where 𝑓(𝑑,π‘₯) is the forcing term, 𝐿(𝑒) is a dissipative term, 𝐿 can be a differential operator or a quasi-differential operator according to different physical situations.

With 𝑓=0 and 𝐿(𝑒)=𝛾(1βˆ’πœ•2π‘₯)𝑒, (1.3) becomes weakly dissipative Camassa-Holm equationπ‘’π‘‘βˆ’π‘’π‘‘π‘₯π‘₯+3𝑒𝑒π‘₯ξ€·+π›Ύπ‘’βˆ’π‘’π‘₯π‘₯ξ€Έ=2𝑒π‘₯𝑒π‘₯π‘₯+𝑒𝑒π‘₯π‘₯π‘₯,𝑑>0,π‘₯βˆˆβ„,(1.4) where 𝛾>0 is a constant.

The local well-posedness, global existence, and blow-up phenomena of the Cauchy problem of (1.4) on the line [40] and on the circle [41] were studied. A new global existence result and a new blow-up result for strong solutions to this equation with certain profiles are presented recently [42]. We found that the behaviors of (1.4) are similar to the Camassa-Holm equation in a finite interval of time, such as the local well-posedness and the blow-up phenomena, and that there are considerable differences between (1.4) and the Camassa-Holm equation in their long-time behaviors. The global solutions of (1.4) decay to zero as time goes to infinity provided the potential 𝑦0=(1βˆ’πœ•2π‘₯)𝑒0 is of one sign (see [40, 41]). This long-time behavior is an important feature that the Camassa-Holm equation does not possess. It is well known that the Camassa-Holm equation has peaked traveling wave solutions. But the fact that any global solution of (1.4) decays to zero means that there are no traveling wave solutions of (1.4).

Another difference between (1.4) and the Camassa-Holm equation is that (1.4) does not have the following conservation laws𝐼1=ξ€œπ‘†π‘’π‘‘π‘₯,𝐼2=ξ€œπ‘†ξ€·π‘’2+𝑒2π‘₯𝑑π‘₯,(1.5) which play an important role in the study of the Camassa-Holm equation.

Equation (1.4) has the same blow-up rate as the Camassa-Holm equation does when the blow-up occurs [41]. This fact shows that the blow-up rate of the Camassa-Holm equation is not affected by the weakly dissipative term, but the occurrence of blow-up of (1.4) is affected by the dissipative parameter [40, 41].

In the paper, we would like to consider the following weakly dissipative Camassa-Holm equationπ‘’π‘‘βˆ’π‘’π‘‘π‘₯π‘₯+3𝑒𝑒π‘₯βˆ’2𝑒π‘₯𝑒π‘₯π‘₯βˆ’π‘’π‘’π‘₯π‘₯π‘₯ξ€·+π›Ύπ‘’βˆ’π‘’π‘₯π‘₯ξ€Έ=𝑓(𝑑,π‘₯),𝑑>0,π‘₯βˆˆβ„,(1.6)𝑒(𝑑,π‘₯+𝐿)=𝑒(𝑑,π‘₯),𝑑>0,π‘₯βˆˆβ„,(1.7)𝑒(𝑑+πœ”,π‘₯)=𝑒(𝑑,π‘₯),𝑑>0,π‘₯βˆˆβ„,(1.8) where 𝛾(1βˆ’πœ•2π‘₯)𝑒 is the weakly dissipative term, 𝛾>0 is a constant, and the forcing term 𝑓 is πœ”-periodic in time 𝑑 and 𝐿-periodic in spatial π‘₯. Without loss of generality, we assume further βˆ«Ξ©π‘“(𝑑,π‘₯)𝑑π‘₯=0, where Ξ©=[0,𝐿]. When system is periodically dependent on time 𝑑, we want to know whether there exists time-periodic solution with the same period for the system. In many nonlinear evolution equations, the study of time-periodic solution has attracted considerable interest (e.g., [43–45]). In this paper, we will prove that (1.6)–(1.8) have a solution by using the Galerkin method [46], and Leray-Schauder fixed point theorem [44].

Our paper is organized as follows. In Section 2, we give some notations and definition of some space used in this paper. In Section 3, we prove the existence of the approximate solution and give uniform a priori estimates needed where we prove the convergence of a sequence of the approximate solution. Section 4 is devoted to the study of the existence and uniqueness of time-periodic solution for (1.6)–(1.8).

2. Preliminaries

Before starting our work, it is appropriate to introduce some notations and inequalities that will be used in the paper.

Let 𝑋 be a Banach space, we denote by πΆπ‘˜(πœ”;𝑋) the set of πœ”-periodic 𝑋-valued measurable functions on ℝ1 with continuous derivatives up to order π‘˜. The norm in the space πΆπ‘˜(πœ”;𝑋) is β€–π‘’β€–πΆπ‘˜(πœ”;𝑋)=sup0β‰€π‘‘β‰€πœ”{βˆ‘π‘˜π‘–=0‖𝐷𝑖𝑑𝑒‖𝑋}.

For 1β‰€π‘β‰€βˆž, the space 𝐿𝑝(πœ”;𝑋) is the set of πœ”-periodic 𝑋-valued measurable functions on ℝ such that ‖𝑒‖𝐿𝑝(πœ”;𝑋)=⎧βŽͺ⎨βŽͺβŽ©ξ‚΅ξ€œπœ”0‖𝑒‖𝑝𝑋𝑑𝑑1/𝑝<∞,1≀𝑝<∞sup0β‰€π‘‘β‰€πœ”β€–π‘’β€–π‘‹<∞,𝑝=∞.(2.1)

The space π‘Šπ‘˜,𝑝(πœ”;𝑋) denote the set of functions which belong to 𝐿𝑝(πœ”;𝑋) together with their derivatives up to order π‘˜, and we write π‘Šπ‘˜,2(πœ”;𝑋)=π»π‘˜(πœ”;𝑋) in particular when 𝑋 is a Hilbert space.

𝐿𝑝(Ξ©) and π»π‘š(Ξ©) are classical Sobolev spaces. For simplicity, we write ‖⋅‖𝐿𝑝(Ξ©) by ‖⋅‖𝑝 as 𝑝≠2 and ‖⋅‖𝐿2(Ξ©) by β€–β‹…β€–.

The following inequalities (see [47]) will be used in the proofs laterβ€–π‘’β€–βˆžβ‰€π‘˜1‖𝑒‖𝐻1.‖‖𝐷(2.2)π‘—π‘’β€–β€–π‘β‰€π‘˜2β€–π‘’β€–πœƒπ»π‘šβ€–π‘’β€–1βˆ’πœƒ,(2.3) where 𝐷𝑗𝑒=(πœ•π‘—π‘’)/(πœ•π‘₯𝑗), 1/𝑝=𝑗+πœƒ(1/2βˆ’π‘š)+(1βˆ’πœƒ)(1/2) as 0≀𝑗<π‘š, 𝑗/π‘šβ‰€πœƒβ‰€1.β€–π‘’β€–β‰€π‘˜3‖‖𝑒π‘₯β€–β€–,ξ€œΞ©π‘’(π‘₯)𝑑π‘₯=0.(2.4)

3. A Priori Estimates

In this section, we first prove that (1.6)–(1.8) have a sequence of approximate solutions {𝑒𝑛}βˆžπ‘›=1, then give a prior, estimates about {𝑒𝑛}βˆžπ‘›=1.

We denote the unbounded linear operator 𝐴𝑒=βˆ’π‘’π‘₯π‘₯ on 𝑋=𝐿2∫∩{π‘’βˆ£π‘’(π‘₯+𝐿)=𝑒(π‘₯),Ω𝑒𝑑π‘₯=0}, then the set of all linearly independent eigenvectors {𝑀𝑗}βˆžπ‘—=0 of 𝐴, that is, 𝐴𝑀𝑗=πœ†π‘—π‘€π‘—, with 0<πœ†1β‰€πœ†2β‰€β‹―β‰€πœ†π‘—β†’βˆž, is an orthonormal basis of 𝐿2(Ξ©). For any 𝑛 and a group of function {π‘Žπ‘—π‘›(𝑑)}𝑛𝑗=1, where π‘Žπ‘—π‘›(𝑑)(𝑗=1,2,…,𝑛)∈𝐢1(πœ”;ℝ), the function 𝑒𝑛=βˆ‘π‘›π‘—=1π‘Žπ‘—π‘›(𝑑)π‘€π‘—βˆˆπΆ1(πœ”;𝐻𝑛) is called an approximate solution to (1.6)–(1.8) if it satisfies the equation as follows:ξ€·π‘’π‘›π‘‘βˆ’π‘’π‘›π‘₯π‘₯𝑑𝑒+π›Ύπ‘›βˆ’π‘’π‘›π‘₯π‘₯ξ€Έ,𝑀𝑗=𝑁𝑒𝑛+𝑓,𝑀𝑗,𝑗=1,…,𝑛,(3.1) where 𝑁𝑒𝑛=βˆ’3𝑒𝑛𝑒𝑛π‘₯+2𝑒𝑛π‘₯𝑒𝑛π‘₯π‘₯+𝑒𝑛𝑒𝑛π‘₯π‘₯π‘₯ and 𝐻𝑛=span{𝑀1,𝑀2,…,𝑀𝑛}. By the classical theory of ordinary differential equations, for any fixed π‘£π‘›βˆ‘(𝑑)=𝑛𝑗=1𝑏𝑗𝑛(𝑑)π‘€π‘—βˆˆπΆ1(πœ”;𝐻𝑛), the equation (π‘’π‘›π‘‘βˆ’π‘’π‘›π‘₯π‘₯𝑑+𝛾(π‘’π‘›βˆ’π‘’π‘›π‘₯π‘₯),𝑀𝑗)=(𝑁𝑣𝑛+𝑓,𝑀𝑗), 𝑗=1,…,𝑛 has a unique πœ”-periodic solution 𝑒𝑛 and the mapping πΉβˆΆπ‘£π‘›β†’π‘’π‘› is continuous and compact in 𝐢1(πœ”;𝐻𝑛). Hence by Leray-Schauder fixed point theorem, we want to prove the existence of an approximate solution only to show sup0β‰€π‘‘β‰€πœ”β€–π‘’π‘›β€–2≀𝑐 for all possible solution of (3.1) replaced by πœ†π‘π‘’π‘›(0β‰€πœ†β‰€1) instead of nonlinear term 𝑁𝑒𝑛, where 𝑐 is a constant which only depends on 𝐿, πœ€, πœ”, 𝛾, and 𝑓.

Lemma 3.1. If π‘“βˆˆπΆ1(πœ”;π»βˆ’1(Ξ©)), then sup0β‰€π‘‘β‰€πœ”ξ‚€β€–β€–π‘’π‘›β€–β€–2+‖‖𝑒𝑛π‘₯β€–β€–2≀𝑐1,(3.2) where 𝑐1 is a constant which only depends on 𝐿, πœ”, πœ€, 𝛾, π‘˜3, and 𝑓, 𝑀=sup0β‰€π‘‘β‰€πœ”{‖𝑓(𝑑,π‘₯)β€–2π»βˆ’1(Ξ©)} and 𝑑1=min{2𝛾,2π›Ύβˆ’πœ€}>0.

Proof. Multiplying (3.1) by π‘Žπ‘—π‘›(𝑑) and summing up over 𝑗 from 1 to 𝑛, we obtain ξ€·π‘’π‘›π‘‘βˆ’π‘’π‘›π‘₯π‘₯𝑑𝑒+π›Ύπ‘›βˆ’π‘’π‘›π‘₯π‘₯ξ€Έ,𝑒𝑛=𝑁𝑒𝑛+𝑓,𝑒𝑛.(3.3)
Then, we can get 12𝑑‖‖𝑒𝑑𝑑𝑛‖‖2+‖‖𝑒𝑛π‘₯β€–β€–2‖‖𝑒+𝛾𝑛‖‖2+‖‖𝑒𝑛π‘₯β€–β€–2=𝑁𝑒𝑛+𝑓,𝑒𝑛.(3.4)
Notice that βˆ«βˆ’3Ω𝑒2𝑛𝑒𝑛π‘₯𝑑π‘₯=0, 2βˆ«Ξ©π‘’π‘›π‘’π‘›π‘₯𝑒𝑛π‘₯π‘₯βˆ«π‘‘π‘₯+Ω𝑒2𝑛𝑒𝑛π‘₯π‘₯π‘₯𝑑π‘₯=0.
From Young’s inequality, we have βˆ«Ξ©π‘“π‘’π‘›π‘‘π‘₯≀(πœ€/2)‖𝑒𝑛π‘₯β€–2+π‘˜23𝑀/2πœ€, where πœ€>0 is a constant.
According to the above relations, we can derive from (3.4) that 𝑑‖‖𝑒𝑑𝑑𝑛‖‖2+‖‖𝑒𝑛π‘₯β€–β€–2+𝑑1‖‖𝑒𝑛‖‖2+‖‖𝑒𝑛π‘₯β€–β€–2ξ‚β‰€π‘˜23π‘€πœ€,(3.5) where 𝑑1=min{2𝛾,2π›Ύβˆ’πœ€}>0.
Considering the time periodicity of 𝑒𝑛 and integrating (3.5) over [0,πœ”], we get 𝑑1ξ€œπœ”0‖‖𝑒𝑛‖‖2+‖‖𝑒𝑛π‘₯β€–β€–2ξ‚π‘‘π‘‘β‰€πœ”π‘˜23π‘€πœ€.(3.6)
Hence, there exists π‘‘βˆ—βˆˆ[0,πœ”) such that ‖𝑒𝑛(π‘‘βˆ—)β€–2+‖𝑒𝑛π‘₯(π‘‘βˆ—)β€–2β‰€π‘˜23𝑀/𝑑1πœ€.
From (3.5), we have (𝑑/𝑑𝑑)(‖𝑒𝑛‖2+‖𝑒𝑛π‘₯β€–2)β‰€π‘˜23𝑀/πœ€.
Integrating the above inequality with respect to 𝑑 from π‘‘βˆ— to π‘‘βˆˆ[π‘‘βˆ—,π‘‘βˆ—+πœ”], we deduce that ‖‖𝑒𝑛‖‖(𝑑)2+‖‖𝑒𝑛π‘₯β€–β€–(𝑑)2≀‖‖𝑒𝑛(π‘‘βˆ—)β€–β€–2+‖‖𝑒𝑛π‘₯(π‘‘βˆ—)β€–β€–2+πœ”π‘˜23π‘€πœ€β‰€ξ‚΅1𝑑1ξ‚Άπ‘˜+πœ”23π‘€πœ€.(3.7)
Hence, we infer sup0β‰€π‘‘β‰€πœ”ξ‚€β€–β€–π‘’π‘›β€–β€–2+‖‖𝑒𝑛π‘₯β€–β€–2≀1𝑑1ξ‚Άπ‘˜+πœ”23π‘€πœ€β‰œπ‘1,(3.8) which concludes our proof.

From Lemma 3.1 and Leray-Schauder fixed point theorem, (3.1) has solution {𝑒𝑛}βˆžπ‘›=1, which is also a sequence of approximate solutions of (1.6)–(1.8). In order to obtain the convergence of sequence {𝑒𝑛}βˆžπ‘›=1, we need to give a priori estimates for the high-order derivers of {𝑒𝑛}βˆžπ‘›=1.

Lemma 3.2. If π‘“βˆˆπΆ1(πœ”;π»βˆ’1(Ξ©)), then sup0β‰€π‘‘β‰€πœ”ξ‚€β€–β€–π‘’π‘›π‘₯β€–β€–2+‖‖𝑒𝑛π‘₯π‘₯β€–β€–2≀𝑐2,(3.9) where 𝑐2 is a constant which only depends on 𝐿, πœ”, πœ€, 𝛾, πœ†π‘›, π‘˜1, π‘˜2, π‘˜3, and 𝑓, 𝑀=sup0β‰€π‘‘β‰€πœ”{‖𝑓(𝑑,π‘₯)β€–2π»βˆ’1(Ξ©)} and 𝑑2=min{2π›Ύβˆ’(13/2)πœ€πœ†π‘›,2π›Ύβˆ’(21/2)πœ€}>0.

Proof. Multiplying (3.1) by βˆ’πœ†π‘—π‘Žπ‘—π‘›(𝑑) and summing up over 𝑗 from 1 to 𝑛, we have ξ€·π‘’π‘›π‘‘βˆ’π‘’π‘›π‘₯π‘₯𝑑𝑒+π›Ύπ‘›βˆ’π‘’π‘›π‘₯π‘₯ξ€Έ,𝑒𝑛π‘₯π‘₯ξ€Έ=𝑁𝑒𝑛+𝑓,𝑒𝑛π‘₯π‘₯ξ€Έ.(3.10)
The above equation yields βˆ’12𝑑‖‖𝑒𝑑𝑑𝑛π‘₯β€–β€–2+‖‖𝑒𝑛π‘₯π‘₯β€–β€–2ξ‚ξ‚€β€–β€–π‘’βˆ’π›Ύπ‘›π‘₯β€–β€–2+‖‖𝑒𝑛π‘₯π‘₯β€–β€–2=𝑁𝑒𝑛+𝑓,𝑒𝑛π‘₯π‘₯ξ€Έ.(3.11)
From Young’s inequality, we have ||||ξ€œΞ©π‘“π‘’π‘›π‘₯π‘₯||||‖‖𝑒𝑑π‘₯β‰€πœ€π‘›π‘₯π‘₯π‘₯β€–β€–2+π‘˜23𝑀4πœ€,(3.12) where πœ€>0 is a constant.
From (2.2), (3.8), and Young’s inequality, we can deduce that ||||ξ€œΞ©π‘’π‘›π‘’π‘›π‘₯𝑒𝑛π‘₯π‘₯||||≀‖‖𝑒𝑑π‘₯π‘›β€–β€–βˆžξ€œΞ©||𝑒𝑛π‘₯𝑒𝑛π‘₯π‘₯||𝑑π‘₯β‰€π‘˜1‖‖𝑒𝑛‖‖𝐻1ξ€œΞ©||𝑒𝑛π‘₯𝑒𝑛π‘₯π‘₯||𝑑π‘₯β‰€π‘˜1𝑐11/2ξƒ©πœ€π‘˜1𝑐11/2‖‖𝑒𝑛π‘₯π‘₯β€–β€–2+π‘˜1𝑐11/2‖‖𝑒4πœ€π‘›π‘₯β€–β€–2ξƒͺβ€–β€–π‘’β‰€πœ€π‘›π‘₯π‘₯β€–β€–2+π‘˜21𝑐21.4πœ€(3.13)
From (2.3), (3.8), Cauchy-Schwarz inequality, Young’s inequality, and Lemma 3.1, we get ||||ξ€œΞ©π‘’π‘›π‘’π‘›π‘₯π‘₯𝑒𝑛π‘₯π‘₯π‘₯||||=||||βˆ’1𝑑π‘₯2ξ€œΞ©π‘’π‘›π‘₯𝑒2𝑛π‘₯π‘₯||||≀1𝑑π‘₯2‖‖𝑒𝑛π‘₯‖‖‖‖𝑒𝑛π‘₯π‘₯β€–β€–24≀12𝑐11/2π‘˜22‖‖𝑒𝑛‖‖1/2‖‖𝑒𝑛‖‖𝐻3/23≀34πœ€β€–β€–π‘’π‘›β€–β€–2𝐻3+𝑐31π‘˜8264πœ€3≀34πœ€β€–β€–π‘’π‘›π‘₯π‘₯β€–β€–2+34πœ€β€–β€–π‘’π‘›π‘₯π‘₯π‘₯β€–β€–2+34πœ€π‘1+𝑐31π‘˜8264πœ€3,‖‖𝑒(3.14)𝑛π‘₯π‘₯π‘₯β€–β€–2=ξ€œΞ©|||||(𝑛𝑗=1π‘Žπ‘—π‘›(𝑑)𝑀𝑗)π‘₯π‘₯π‘₯|||||2ξ€œπ‘‘π‘₯=Ξ©|||||(𝑛𝑗=1πœ†π‘—π‘Žπ‘—π‘›(𝑑)𝑀𝑗)π‘₯|||||2𝑑π‘₯β‰€πœ†π‘›β€–β€–π‘’π‘›π‘₯β€–β€–2.(3.15)
Taking (3.11)–(3.15) into account, we can infer that 𝑑‖‖𝑒𝑑𝑑𝑛π‘₯β€–β€–2+‖‖𝑒𝑛π‘₯π‘₯β€–β€–2+𝑑2‖‖𝑒𝑛π‘₯β€–β€–2+‖‖𝑒𝑛π‘₯π‘₯β€–β€–2ξ‚β‰€π‘˜23𝑀+2πœ€3π‘˜21𝑐21+2πœ€9πœ€π‘12+3𝑐31π‘˜8232πœ€3,(3.16) where 𝑑2=min{2π›Ύβˆ’(13/2)πœ€πœ†π‘›,2π›Ύβˆ’(21/2)πœ€}>0.
Integrating (3.16) about 𝑑 from 0 to πœ” and considering the time periodicity of 𝑒𝑛, we get 𝑑2ξ€œπœ”0‖‖𝑒𝑛π‘₯β€–β€–2+‖‖𝑒𝑛π‘₯π‘₯β€–β€–2ξ‚ξƒ©π‘˜π‘‘π‘‘β‰€23𝑀+2πœ€3π‘˜21𝑐21+2πœ€9πœ€π‘12+3𝑐31π‘˜8232πœ€3ξƒͺπœ”.(3.17)
Hence, there exists π‘‘βˆ—βˆˆ[0,πœ”) such that ‖‖𝑒𝑛π‘₯(π‘‘βˆ—)β€–β€–2+‖‖𝑒𝑛π‘₯π‘₯(π‘‘βˆ—)β€–β€–2≀1𝑑2ξƒ©π‘˜23𝑀+2πœ€3π‘˜21𝑐21+2πœ€9πœ€π‘12+3𝑐31π‘˜8232πœ€3ξƒͺ.(3.18)
From (3.16), we have ‖‖𝑒𝑛π‘₯β€–β€–(𝑑)2+‖‖𝑒𝑛π‘₯π‘₯β€–β€–(𝑑)2≀‖‖𝑒𝑛π‘₯(π‘‘βˆ—)β€–β€–2+‖‖𝑒𝑛π‘₯π‘₯(π‘‘βˆ—)β€–β€–2+ξƒ©π‘˜23𝑀+2πœ€3π‘˜21𝑐21+2πœ€9πœ€π‘12+3𝑐31π‘˜8232πœ€3ξƒͺπœ”.(3.19)
Then we can obtain sup0β‰€π‘‘β‰€πœ”ξ‚€β€–β€–π‘’π‘›π‘₯β€–β€–2+‖‖𝑒𝑛π‘₯π‘₯β€–β€–2≀1𝑑2ξ‚Άξƒ©π‘˜+πœ”23𝑀+2πœ€3π‘˜21𝑐21+2πœ€9πœ€π‘12+3𝑐31π‘˜8232πœ€3ξƒͺβ‰œπ‘2,(3.20) which concludes our proof.

In the following, we continue to establish a priori estimates for high-order derivers of the approximate solution {𝑒𝑛}βˆžπ‘›=1 by an inductive argument.

Lemma 3.3. For any π‘˜β‰₯0, if π‘“βˆˆπΆ1(πœ”;π»π‘˜βˆ’1(Ξ©)), then sup0β‰€π‘‘β‰€πœ”ξ‚€β€–β€–π·π‘˜+1𝑒𝑛‖‖2+β€–β€–π·π‘˜+2𝑒𝑛‖‖2≀𝑐,(3.21) where 𝑐 is a constant which only depends on 𝐿, πœ”, πœ€, 𝛾, πœ†π‘›, π‘˜, π‘˜1, π‘˜2, π‘˜3, 𝑓 and 𝑑3={2π›Ύβˆ’16πœ€πœ†π‘›,2π›Ύβˆ’14πœ€}>0.

Proof. By Lemma 3.2, we know the conclusion of Lemma 3.3 holds for π‘˜=0.
Assume that for π‘˜β‰€π‘šβˆ’1(π‘šβ‰₯2) the conclusion of Lemma 3.3 holds, we want to prove that the same statement holds for π‘˜=π‘š also.
Multiplying (3.1) by (βˆ’1)π‘š+1πœ†π‘—π‘š+1π‘Žπ‘—π‘›(𝑑) and summing up over 𝑗 from 1 to 𝑛, we have (βˆ’1)π‘š+112π‘‘ξ‚€β€–β€–π·π‘‘π‘‘π‘š+1𝑒𝑛‖‖2+β€–β€–π·π‘š+2𝑒𝑛‖‖2+(βˆ’1)π‘š+1π›Ύξ‚€β€–β€–π·π‘š+1𝑒𝑛‖‖2+β€–β€–π·π‘š+2𝑒𝑛‖‖2=𝑁𝑒𝑛+𝑓,𝐷2(π‘š+1)𝑒𝑛.(3.22)
Follow the same methods discussed in Lemma 3.2, we have ||||ξ€œΞ©π‘“π·2(π‘š+1)𝑒𝑛||||=||||ξ€œπ‘‘π‘₯Ξ©π·π‘šβˆ’1π‘“π·π‘š+3𝑒𝑛||||‖‖𝐷𝑑π‘₯β‰€πœ€π‘š+3𝑒𝑛‖‖2+1‖‖𝐷4πœ€π‘šβˆ’1𝑓‖‖2.(3.23)
From the conclusion of Lemma 3.3 for π‘˜β‰€π‘šβˆ’1, (2.2), (2.4) and Young’s inequality, we can deduce that ||||ξ€œΞ©π‘’π‘›π‘’π‘›π‘₯𝐷2(π‘š+1)𝑒𝑛||||=|||||ξ€œπ‘‘π‘₯Ξ©ξƒ©π‘š+1𝑖=0πΆπ‘–π‘š+1π·π‘–π‘’π‘›π·π‘š+1βˆ’π‘–π‘’π‘›π‘₯ξƒͺπ·π‘š+1𝑒𝑛|||||β‰€ξ€œπ‘‘π‘₯Ξ©||π‘’π‘›π·π‘š+2π‘’π‘›π·π‘š+1𝑒𝑛||+ξ€œπ‘‘π‘₯Ξ©|||||ξƒ©π‘š+1𝑖=1πΆπ‘–π‘š+1π·π‘–π‘’π‘›π·π‘š+1βˆ’π‘–π‘’π‘›π‘₯ξƒͺπ·π‘š+1𝑒𝑛|||||‖‖𝐷𝑑π‘₯β‰€πœ€π‘š+2𝑒𝑛‖‖2ξ€·+π‘πœ€,π‘˜1,π‘˜3ξ€Έβ€–β€–π·π‘š+1𝑒𝑛‖‖2ξ€·+π‘π‘š,π‘˜1ξ€Έβ€–β€–π·β‰€πœ€π‘š+2𝑒𝑛‖‖2ξ€·+π‘πœ€,π‘˜1,π‘˜3ξ€Έ.,π‘š(3.24)
Similarly, we can also deduce that ||||ξ€œΞ©π‘’π‘›π‘’π‘›π‘₯π‘₯π‘₯𝐷2(π‘š+1)𝑒𝑛||||=|||||ξ€œπ‘‘π‘₯Ξ©ξƒ©π‘šξ“π‘–=0πΆπ‘–π‘šπ·π‘–π‘’π‘›π·π‘šβˆ’π‘–π‘’π‘›π‘₯π‘₯π‘₯ξƒͺπ·π‘š+2𝑒𝑛|||||β‰€ξ€œπ‘‘π‘₯Ξ©||π‘’π‘›π·π‘š+3π‘’π‘›π·π‘š+2𝑒𝑛||ξ€œπ‘‘π‘₯+Ξ©|||𝐢1π‘šπ·π‘’π‘›ξ€·π·π‘š+2𝑒𝑛2|||+ξ€œπ‘‘π‘₯Ξ©||𝐢2π‘šπ·2π‘’π‘›π·π‘š+1π‘’π‘›π·π‘š+2𝑒𝑛||+ξ€œπ‘‘π‘₯Ξ©|||||ξƒ©π‘šξ“π‘–=3πΆπ‘–π‘šπ·π‘–π‘’π‘›π·π‘š+3βˆ’π‘–π‘’π‘›ξƒͺπ·π‘š+2𝑒𝑛|||||≀‖‖𝑒𝑑π‘₯π‘›β€–β€–βˆžξ€œΞ©||π·π‘š+3π‘’π‘›π·π‘š+2𝑒𝑛||‖‖𝑑π‘₯+π‘šπ·π‘’π‘›β€–β€–βˆžξ€œΞ©||π·π‘š+2𝑒𝑛||2𝑑π‘₯+𝐢2π‘šβ€–β€–π·2π‘’π‘›β€–β€–βˆžξ€œΞ©||π·π‘š+1π‘’π‘›π·π‘š+2𝑒𝑛||+𝑑π‘₯π‘šξ“π‘–=3πΆπ‘–π‘šβ€–β€–π·π‘–π‘’π‘›β€–β€–βˆžβ€–β€–π·π‘š+3βˆ’π‘–π‘’π‘›β€–β€–βˆžξ€œΞ©||π·π‘š+2𝑒𝑛||ξ€·π‘˜π‘‘π‘₯≀𝑐1,π‘˜3ξ€Έξ‚΅ξ€œ,π‘šΞ©||π·π‘š+3π‘’π‘›π·π‘š+2𝑒𝑛||ξ€œπ‘‘π‘₯+Ξ©||π·π‘š+2𝑒𝑛||2+ξ€œπ‘‘π‘₯Ξ©||π·π‘š+1π‘’π‘›π·π‘š+2𝑒𝑛||ξ€œπ‘‘π‘₯+Ξ©||π·π‘š+2𝑒𝑛||ξ‚Ά.𝑑π‘₯(3.25)
From the conclusion of Lemma 3.3 for π‘˜β‰€π‘šβˆ’1, Young’s inequality and (2.3), we have π‘ξ€·π‘˜1,π‘˜3ξ€Έξ€œ,π‘šΞ©||π·π‘š+3π‘’π‘›π·π‘š+2𝑒𝑛||‖‖𝐷𝑑π‘₯β‰€πœ€π‘š+3𝑒𝑛‖‖2ξ€·π‘˜+𝑐1,π‘˜3‖‖𝐷,π‘š,πœ€π‘š+2𝑒𝑛‖‖2β€–β€–π·β‰€πœ€π‘š+3𝑒𝑛‖‖2ξ€·π‘˜+𝑐1,π‘˜2,π‘˜3×‖‖𝑒,π‘š,πœ€π‘›β€–β€–π»2(π‘š+2)/π‘š+3π‘š+3‖‖𝑒𝑛‖‖2(1βˆ’(π‘š+2)/(π‘š+3))β€–β€–π·β‰€πœ€π‘š+3𝑒𝑛‖‖2‖‖𝑒+πœ€π‘›β€–β€–2π»π‘š+3ξ€·π‘˜+𝑐1,π‘˜2,π‘˜3‖‖𝑒,π‘š,πœ€π‘›β€–β€–2‖‖𝐷≀2πœ€π‘š+3𝑒𝑛‖‖2‖‖𝐷+πœ€π‘š+2𝑒𝑛‖‖2ξ€·π‘˜+𝑐1,π‘˜2,π‘˜3ξ€Έ,π‘ξ€·π‘˜,π‘š,πœ€1,π‘˜3ξ€Έξ€œ,π‘šΞ©|||ξ€·π·π‘š+2𝑒𝑛2|||ξ€·π‘˜π‘‘π‘₯=𝑐1,π‘˜3‖‖𝐷,π‘šπ‘š+2𝑒𝑛‖‖2ξ€·π‘˜β‰€π‘1,π‘˜2,π‘˜3‖‖𝑒,π‘šπ‘›β€–β€–π»2(π‘š+2)/(π‘š+3)π‘š+3‖‖𝑒𝑛‖‖2(1βˆ’(π‘š+2)/(π‘š+3))β€–β€–π·β‰€πœ€π‘š+3𝑒𝑛‖‖2‖‖𝐷+πœ€π‘š+2𝑒𝑛‖‖2ξ€·π‘˜+𝑐1,π‘˜2,π‘˜3ξ€Έ,π‘ξ€·π‘˜,π‘š,πœ€1,π‘˜3ξ€Έξ€œ,π‘šΞ©||π·π‘š+1π‘’π‘›π·π‘š+2𝑒𝑛||‖‖𝐷𝑑π‘₯β‰€πœ€π‘š+2𝑒𝑛‖‖2ξ€·π‘˜+𝑐1,π‘˜3‖‖𝐷,π‘š,πœ€π‘š+1𝑒𝑛‖‖2β€–β€–π·β‰€πœ€π‘š+2𝑒𝑛‖‖2ξ€·π‘˜+𝑐1,π‘˜3ξ€Έ,π‘ξ€·π‘˜,π‘š,πœ€1,π‘˜3ξ€Έξ€œ,π‘šΞ©||π·π‘š+2𝑒𝑛||‖‖𝐷𝑑π‘₯β‰€πœ€π‘š+2𝑒𝑛‖‖2ξ€·π‘˜+𝑐1,π‘˜3ξ€Έ.,π‘š,πœ€,𝐿(3.26)
Combining (3.25) and the above inequality, we can get ||||ξ€œΞ©π‘’π‘›π‘’π‘›π‘₯π‘₯π‘₯𝐷2(π‘š+1)𝑒𝑛||||‖‖𝐷𝑑π‘₯≀3πœ€π‘š+3𝑒𝑛‖‖2‖‖𝐷+4πœ€π‘š+2𝑒𝑛‖‖2ξ€·π‘˜+𝑐1,π‘˜2,π‘˜3ξ€Έ,π‘š,πœ€,𝐿.(3.27)
Similarly, ||||ξ€œΞ©π‘’π‘›π‘₯𝑒𝑛π‘₯π‘₯𝐷2(π‘š+1)𝑒𝑛||||β‰€ξ€œπ‘‘π‘₯Ξ©||𝑒𝑛π‘₯π·π‘š+1π‘’π‘›π·π‘š+3𝑒𝑛||+ξ€œπ‘‘π‘₯Ξ©|||||ξƒ©π‘šβˆ’1𝑖=1πΆπ‘–π‘šβˆ’1𝐷𝑖+1π‘’π‘›π·π‘š+1βˆ’π‘–π‘’π‘›ξƒͺπ·π‘š+3𝑒𝑛|||||≀‖‖𝑒𝑑π‘₯𝑛π‘₯β€–β€–βˆžξ€œΞ©||π·π‘š+1π‘’π‘›π·π‘š+3𝑒𝑛||+𝑑π‘₯π‘šβˆ’1𝑖=1πΆπ‘–π‘šβˆ’1‖‖𝐷𝑖+1π‘’π‘›β€–β€–βˆžβ€–β€–π·π‘š+1βˆ’π‘–π‘’π‘›β€–β€–βˆžξ€œΞ©||π·π‘š+3𝑒𝑛||‖‖𝐷𝑑π‘₯≀2πœ€π‘š+3𝑒𝑛‖‖2ξ€·+π‘π‘š,π‘˜1ξ€Έ.,πœ€,𝐿(3.28)
Taking (3.22)–(3.24) and (3.27)-(3.28) into account, we can deduce that π‘‘ξ‚€β€–β€–π·π‘‘π‘‘π‘š+1𝑒𝑛‖‖2+β€–β€–π·π‘š+2𝑒𝑛‖‖2‖‖𝐷+2π›Ύπ‘š+1𝑒𝑛‖‖2+β€–β€–π·π‘š+2𝑒𝑛‖‖2‖‖𝐷≀16πœ€π‘š+3𝑒𝑛‖‖2‖‖𝐷+14πœ€π‘š+2𝑒𝑛‖‖2ξ€·π‘˜+𝑐1,π‘˜2,π‘˜3ξ€Έ.,π‘š,πœ€,𝑓,𝐿(3.29)
From the above relation, we can infer π‘‘ξ‚€β€–β€–π·π‘‘π‘‘π‘š+1𝑒𝑛‖‖2+β€–β€–π·π‘š+2𝑒𝑛‖‖2+𝑑3ξ‚€β€–β€–π·π‘š+1𝑒𝑛‖‖2+β€–β€–π·π‘š+2𝑒𝑛‖‖2ξ‚ξ€·π‘˜β‰€π‘1,π‘˜2,π‘˜3ξ€Έ,π‘š,πœ€,𝑓,𝐿,(3.30) where 𝑑3={2π›Ύβˆ’16πœ€πœ†π‘›,2π›Ύβˆ’14πœ€}>0.
Integrating (3.30) about 𝑑 from 0 to πœ”, there exists π‘‘βˆ—βˆˆ[0,πœ”) such that β€–β€–π·π‘š+1𝑒𝑛(π‘‘βˆ—)β€–β€–2+β€–β€–π·π‘š+2𝑒𝑛(π‘‘βˆ—)β€–β€–2β‰€π‘ξ€·π‘˜1,π‘˜2,π‘˜3ξ€Έ,π‘š,πœ€,𝑓,𝐿𝑑3.(3.31)
From (3.30), we have π‘‘ξ‚€β€–β€–π·π‘‘π‘‘π‘š+1𝑒𝑛‖‖2+β€–β€–π·π‘š+2𝑒𝑛‖‖2ξ‚ξ€·π‘˜β‰€π‘1,π‘˜2,π‘˜3ξ€Έ,π‘š,πœ€,𝑓,𝐿.(3.32)
Integrating the above inequality from π‘‘βˆ— to π‘‘βˆˆ[π‘‘βˆ—,π‘‘βˆ—+πœ”] and with (3.31), we can easily obtain sup0β‰€π‘‘β‰€πœ”ξ‚€β€–β€–π·π‘š+1𝑒𝑛‖‖2+β€–β€–π·π‘š+2𝑒𝑛‖‖2≀1𝑑3ξ‚Άπ‘ξ€·π‘˜+πœ”1,π‘˜2,π‘˜3ξ€Έ,π‘š,πœ€,𝑓,πΏβ‰œπ‘.(3.33)
The proof is completed.

Lemma 3.4. For any π‘˜β‰₯0, if π‘“βˆˆπΆ1(πœ”;π»π‘˜+1(Ξ©)), then sup0β‰€π‘‘β‰€πœ”ξ‚€β€–β€–π·π‘˜π‘’π‘›π‘‘β€–β€–2+β€–β€–π·π‘˜+1𝑒𝑛𝑑‖‖2≀𝑐(3.34) where 𝑐 is a constant which only depends on 𝐿, πœ”, πœ€, 𝛾, πœ†π‘›, π‘˜, π‘˜1, π‘˜2, π‘˜3, and 𝑓.

Proof. We first prove the conclusion of Lemma 3.4 holds for π‘˜=0. Multiplying (3.1) by π‘Žξ…žπ‘—π‘›(𝑑) and summing up over 𝑗 from 1 to 𝑛, we have ‖‖𝑒𝑛𝑑‖‖2+‖‖𝑒𝑛π‘₯𝑑‖‖2=𝑁𝑒𝑛𝑒+π‘“βˆ’π›Ύπ‘›βˆ’π‘’π‘›π‘₯π‘₯ξ€Έ,𝑒𝑛𝑑.(3.35)
By Lemma 3.3, if π‘“βˆˆπΆ1(πœ”;𝐻1(Ξ©)), then we have ‖𝑒𝑛‖2𝐻4≀𝑐. Hence, ||𝑁𝑒𝑛𝑒+π‘“βˆ’π›Ύπ‘›βˆ’π‘’π‘›π‘₯π‘₯ξ€Έ,𝑒𝑛𝑑||≀‖‖𝑁𝑒𝑛𝑒+π‘“βˆ’π›Ύπ‘›βˆ’π‘’π‘›π‘₯π‘₯‖‖‖‖𝑒𝑛𝑑‖‖‖‖𝑒≀𝑐𝑛𝑑‖‖.(3.36)
Therefore, from (3.35) and (3.36), it is easy to know that sup0β‰€π‘‘β‰€πœ”ξ‚€β€–β€–π‘’π‘›π‘‘β€–β€–2+‖‖𝑒𝑛π‘₯𝑑‖‖2≀𝑐.(3.37)
Assume that the conclusion of Lemma 3.4 holds for π‘˜β‰€π‘š(π‘šβ‰₯1), we want to prove that the conclusion of Lemma 3.4 also holds for π‘˜=π‘š+1.
Multiplying (3.1) by (βˆ’1)π‘š+1πœ†π‘—π‘š+1π‘Žξ…žπ‘—π‘›(𝑑) and summing up over 𝑗 from 1 to 𝑛, we have (βˆ’1)π‘š+1ξ‚€β€–β€–π·π‘š+1𝑒𝑛𝑑‖‖2+β€–β€–π·π‘š+2𝑒𝑛𝑑‖‖2=𝑁𝑒𝑛𝑒+π‘“βˆ’π›Ύπ‘›βˆ’π‘’π‘›π‘₯π‘₯ξ€Έ,𝐷2(π‘š+1)𝑒𝑛𝑑.(3.38)
By Lemma 3.3, if π‘“βˆˆπΆ1(πœ”;π»π‘š+2(Ξ©)), then β€–π·π‘˜π‘’π‘›β€–2≀𝑐 for π‘˜β‰€π‘š+5. Hence, ||𝑁𝑒𝑛𝑒+π‘“βˆ’π›Ύπ‘›βˆ’π‘’π‘›π‘₯π‘₯ξ€Έ,𝐷2(π‘š+1)𝑒𝑛𝑑||β‰€β€–β€–π·π‘š+1𝑁𝑒𝑛𝑒+π‘“βˆ’π›Ύπ‘›βˆ’π‘’π‘›π‘₯π‘₯β€–β€–β€–β€–π·ξ€Έξ€»π‘š+1π‘’π‘›π‘‘β€–β€–β€–β€–π·β‰€π‘π‘š+1𝑒𝑛𝑑‖‖.(3.39)
Taking (3.38) and (3.39) into account, it follows sup0β‰€π‘‘β‰€πœ”ξ‚€β€–β€–π·π‘š+1𝑒𝑛𝑑‖‖2+β€–β€–π·π‘š+2𝑒𝑛𝑑‖‖2≀𝑐.(3.40)
This completes the proof of Lemma 3.4 by an inductive argument.

4. Existence and Uniqueness of Time-Periodic Solution

We have proved that (1.6)–(1.8) have a sequence of approximate solutions {𝑒𝑛}βˆžπ‘›=1. In this section, we want to prove that the sequence converges and the limit is a solution of (1.6)–(1.8).

By Lemmas 3.1–3.4 and standard compactness arguments, we conclude that there is a subsequence which we denote also by {𝑒𝑛} such that for any 𝐾β‰₯0, if π‘“βˆˆπΆ1(πœ”;π»π‘˜+1(Ξ©)), we have𝑒𝑛(𝑑)βŸΆπ‘’(𝑑),weaklyβˆ—inπΏβˆžξ€·πœ”;π»π‘˜+4ξ€Έ,𝑒(Ξ©)𝑛(𝑑)βŸΆπ‘’(𝑑),stronglyinπΏβˆžξ€·πœ”;π»π‘˜+3ξ€Έ,𝑒(Ξ©)𝑛𝑑(𝑑)βŸΆπ‘’π‘‘(𝑑),weaklyβˆ—inπΏβˆžξ€·πœ”;π»π‘˜+1ξ€Έ,𝑒(Ξ©)𝑛𝑑(𝑑)βŸΆπ‘’π‘‘(𝑑),stronglyinπΏβˆžξ€·πœ”;π»π‘˜(ξ€Έ.Ξ©)(4.1)

From the above lemmas, we know that the nonlinear terms are well defined‖‖𝑒𝑛𝑒𝑛π‘₯βˆ’π‘’π‘’π‘₯‖‖≀‖‖𝑒𝑛𝑒𝑛π‘₯βˆ’π‘’π‘₯ξ€Έβ€–β€–+‖‖𝑒π‘₯ξ€·π‘’π‘›ξ€Έβ€–β€–β‰€β€–β€–π‘’βˆ’π‘’π‘›β€–β€–βˆžβ€–β€–π‘’π‘›π‘₯βˆ’π‘’π‘₯β€–β€–+‖‖𝑒π‘₯β€–β€–βˆžβ€–β€–π‘’π‘›β€–β€–βˆ’π‘’βŸΆ0,(4.2) as π‘›β†’βˆž, uniformly in 𝑑,‖‖𝑒𝑛π‘₯𝑒𝑛π‘₯π‘₯βˆ’π‘’π‘₯𝑒π‘₯π‘₯‖‖≀‖‖𝑒𝑛π‘₯𝑒𝑛π‘₯π‘₯βˆ’π‘’π‘₯π‘₯ξ€Έβ€–β€–+‖‖𝑒π‘₯π‘₯𝑒𝑛π‘₯βˆ’π‘’π‘₯‖‖≀‖‖𝑒𝑛π‘₯β€–β€–βˆžβ€–β€–π‘’π‘›π‘₯π‘₯βˆ’π‘’π‘₯π‘₯β€–β€–+‖‖𝑒π‘₯π‘₯β€–β€–βˆžβ€–β€–π‘’π‘›π‘₯βˆ’π‘’π‘₯β€–β€–βŸΆ0,(4.3) as π‘›β†’βˆž, uniformly in 𝑑,‖‖𝑒𝑛𝑒𝑛π‘₯π‘₯π‘₯βˆ’π‘’π‘’π‘₯π‘₯π‘₯‖‖≀‖‖𝑒𝑛𝑒𝑛π‘₯π‘₯π‘₯βˆ’π‘’π‘₯π‘₯π‘₯ξ€Έβ€–β€–+‖‖𝑒π‘₯π‘₯π‘₯ξ€·π‘’π‘›ξ€Έβ€–β€–β‰€β€–β€–π‘’βˆ’π‘’π‘›β€–β€–βˆžβ€–β€–π‘’π‘›π‘₯π‘₯π‘₯βˆ’π‘’π‘₯π‘₯π‘₯β€–β€–+β€–β€–π‘’π‘›β€–β€–βˆ’π‘’βˆžβ€–β€–π‘’π‘₯π‘₯π‘₯β€–β€–β‰€β€–β€–π‘’π‘›β€–β€–βˆžβ€–β€–π‘’π‘›π‘₯π‘₯π‘₯βˆ’π‘’π‘₯π‘₯π‘₯β€–β€–+π‘˜1β€–β€–π‘’π‘›β€–β€–βˆ’π‘’π»1‖‖𝑒π‘₯π‘₯π‘₯β€–β€–βŸΆ0,(4.4) as π‘›β†’βˆž, uniformly in 𝑑.

Consequently, it follows thatξ€·π‘’π‘‘βˆ’π‘’π‘₯π‘₯𝑑+π›Ύπ‘’βˆ’π‘’π‘₯π‘₯ξ€Έξ€Έ,πœ‚=(𝑁𝑒+𝑓,πœ‚),πœ‚βˆˆπΏ2per.(4.5)

Thanks to the estimates obtained in the previous section, we haveπ‘’π‘‘βˆ’π‘’π‘₯π‘₯𝑑+π›Ύπ‘’βˆ’π‘’π‘₯π‘₯ξ€Έ=𝑁𝑒+𝑓,(4.6) a.e. on ℝ1Γ—Ξ©.

So we obtain that the existence of time periodic solution for (1.6)–(1.8), which is the following theorem.

Theorem 4.1. Given π‘“βˆˆπΆ1(πœ”;π»π‘˜+1(Ξ©)),β€‰β€‰π‘˜β‰₯0, there exists a time periodic solution 𝑒(𝑑,π‘₯) to (1.6)–(1.8), such that 𝑒(𝑑,π‘₯)∈𝐿∞(πœ”;π»π‘˜+4(Ξ©))βˆ©π‘Š1,∞(πœ”;π»π‘˜(Ξ©)).

Under the assumption of Theorem 4.1, we are unable to prove the uniqueness of the solution for (1.6)–(1.8). But if we assume that 𝑀 is sufficiently small, then the result can be obtained.

Theorem 4.2. Suppose that the assumption in Theorem 4.1 holds. If 𝑀 is sufficiently small, then the time periodic solution of (1.6)–(1.8) in Theorem 4.1 is unique.

Proof. Let 𝑒 and 𝑒 be any two time periodic solutions of (1.6)–(1.8). With 𝑣=π‘’βˆ’π‘’, we can get from (1.6) that π‘£π‘‘βˆ’π‘£π‘₯π‘₯𝑑+π›Ύπ‘£βˆ’π‘£π‘₯π‘₯ξ€Έ=π‘π‘’βˆ’π‘π‘’.(4.7)
Taking the inner product of (4.7) with 𝑣, we have 12𝑑𝑑𝑑‖𝑣‖2+‖‖𝑣π‘₯β€–β€–2+𝛾‖𝑣‖2+‖‖𝑣π‘₯β€–β€–2=ξ€·π‘π‘’βˆ’π‘ξ€Έπ‘’,𝑣.(4.8)
Since, ||ξ€·βˆ’3𝑒𝑒π‘₯+3𝑒𝑒π‘₯ξ€Έ||≀||ξ€·,π‘£βˆ’3𝑒𝑣π‘₯ξ€Έ||+||ξ€·,π‘£βˆ’3𝑒π‘₯ξ€Έ||≀3𝑣,𝑣2β€–π‘’β€–βˆžξ‚€β€–π‘£β€–2+‖‖𝑣π‘₯β€–β€–2‖‖+3𝑒π‘₯β€–β€–βˆžβ€–π‘£β€–2≀3π‘˜12𝑐11/2+3π‘˜1𝑐21/2‖𝑣‖2+3π‘˜12𝑐11/2‖‖𝑣π‘₯β€–β€–2,||ξ€·(4.9)2𝑒π‘₯𝑒π‘₯π‘₯βˆ’2𝑒π‘₯𝑒π‘₯π‘₯ξ€Έ||≀||ξ€·,𝑣2𝑒π‘₯𝑣π‘₯π‘₯ξ€Έ||+||ξ€·2,𝑣𝑒π‘₯π‘₯𝑣π‘₯ξ€Έ||≀‖‖𝑒,𝑣π‘₯π‘₯β€–β€–βˆžξ‚€β€–π‘£β€–2+‖‖𝑣π‘₯β€–β€–2‖‖𝑒+2π‘₯β€–β€–βˆžβ€–β€–π‘£π‘₯β€–β€–2+‖‖𝑒π‘₯π‘₯β€–β€–βˆžξ‚€β€–π‘£β€–2+‖‖𝑣π‘₯β€–β€–2≀2π‘˜1𝑐2ξ€Έ+𝑐1/2‖𝑣‖2+2π‘˜1𝑐2ξ€Έ+𝑐1/2+2π‘˜1𝑐21/2‖‖𝑣π‘₯β€–β€–2,||ξ€·(4.10)𝑒𝑒π‘₯π‘₯π‘₯βˆ’π‘’π‘’π‘₯π‘₯π‘₯ξ€Έ||≀||ξ€·,𝑣𝑒𝑣π‘₯π‘₯π‘₯ξ€Έ||+||ξ€·,𝑣𝑒π‘₯π‘₯π‘₯ξ€Έ||β‰€ξ€œπ‘£,𝑣Ω||𝑒π‘₯π‘₯𝑣𝑣π‘₯||3𝑑π‘₯+2ξ€œΞ©||𝑒π‘₯𝑣2π‘₯||ξ€œπ‘‘π‘₯+2Ξ©||𝑒π‘₯π‘₯𝑣𝑣π‘₯||≀1𝑑π‘₯2‖‖𝑒π‘₯π‘₯β€–β€–βˆžξ‚€β€–π‘£β€–2+‖‖𝑣π‘₯β€–β€–2+32‖‖𝑒π‘₯β€–β€–βˆžβ€–β€–π‘£π‘₯β€–β€–2+‖‖𝑒π‘₯π‘₯β€–β€–βˆžξ‚€β€–π‘£β€–2+‖‖𝑣π‘₯β€–β€–2≀3π‘˜12𝑐2ξ€Έ+𝑐1/2‖𝑣‖2+ξ‚Έ3π‘˜12𝑐2ξ€Έ+𝑐1/2+3π‘˜12𝑐21/2‖‖𝑣π‘₯β€–β€–2.(4.11)
Hence, if 𝑀 is sufficient small such that 2𝛾β‰₯3π‘˜1𝑐11/2+6π‘˜1𝑐21/2+7π‘˜1(𝑐2+𝑐)1/2, 2𝛾β‰₯3π‘˜1𝑐11/2+7π‘˜1𝑐21/2+7π‘˜1(𝑐2+𝑐)1/2, then it follows from (4.8)–(4.11), we get 𝑑𝑑𝑑‖𝑣‖2+‖‖𝑣π‘₯β€–β€–2+πœŒβ€–π‘£β€–2+‖‖𝑣π‘₯β€–β€–2≀0,(4.12) where 𝜌β‰₯0 is suitable constant.
Applying Gronwall’s inequality, we derive that ‖𝑣(𝑑)β€–2+‖‖𝑣π‘₯β€–β€–(𝑑)2≀‖𝑣(0)β€–2+‖‖𝑣π‘₯β€–β€–(0)2ξ‚π‘’βˆ’πœŒπ‘‘,forany𝑑β‰₯0.(4.13)
Since 𝑣 is πœ”-periodic in 𝑑, then for any positive integer π‘š we have ‖𝑣(𝑑)β€–2+‖‖𝑣π‘₯β€–β€–(𝑑)2=‖𝑣(𝑑+π‘šπœ”)β€–2+‖‖𝑣π‘₯β€–β€–(𝑑+π‘šπœ”)2.(4.14)
Then we can infer that ‖𝑣(𝑑)β€–2+‖‖𝑣π‘₯β€–β€–(𝑑)2≀‖𝑣(0)β€–2+‖‖𝑣π‘₯β€–β€–(0)2ξ‚π‘’βˆ’πœŒ(𝑑+π‘šπœ”).(4.15)
It follows from 𝑣(0)=𝑣π‘₯(0)=0 that 𝑒(𝑑,π‘₯)=𝑒(𝑑,π‘₯), which completes the proof of Theorem 4.2.