Abstract

We study the following second-order periodic system: where has a singularity and . Under some assumptions on the and , by Moser's twist theorem we obtain the existence of quasiperiodic solutions and boundedness of all the solutions.

1. Introduction and Main Result

In the early 1960s, Littlewood [1] asked whether or not the solutions of the Duffing-type equations are bounded for all time, that is, whether there are resonances that might cause the amplitude of the oscillations to increase without bound. Littlewood suggested studying the following two cases: (i)superlinear case: as ,(ii)sublinear case: and as .

The first positive result of boundedness of solutions in the superlinear case (i) was due to Morris [2]. By means of KAM theorem, Morris proved that every solution of the second-order system (1.1) is bounded if and is piecewise continuous and periodic. This result relies on the fact that the nonlinearity can guarantee the twist condition of KAM theorem. Later, several authors (see [35]) improved the Morris’s result and obtained similar results for a large class of superlinear function .

In 1999, the first result in the sublinear case was proved by Küpper and You [6] in the study of where and . The authors transform (1.2) into a perturbation of an integrable Hamiltonian system and then prove that the Poincaré map of the transformed system is close to a so-called twist map. So, the Moser’s twist theorem guarantees the boundedness of all solutions of (1.2). The general sublinear case was considered by Liu [7] under certain reasonable conditions.

The Littlewood problem for singular potentials is known to be challenging, and there are only very a few results. Recently, Capietto et al. [8] studied with is a -periodic function and , where and is a positive integer. Under the Lazer-Leach assumption that they prove the boundedness of solutions and the existence of quasiperiodic solution by Moser’ twist theorem. It is the first time that the equation of the boundedness of all solution is treated in case of a singular potential.

In this paper, We consider the following sublinearly growing potential: where .

Our main result is the following theorem.

Theorem 1.1. If is 1-periodic continuous, then all the solutions of (1.5) are bounded.

The idea for proving the boundedness of solutions of (1.5) is as follows. By means of transformation theory, (1.5) is, outside of a large disc in the -plane, transformed into a perturbation of an integrable Hamiltonian system. Then, Poincaré map of the transformed system is close to a so-called twist map in . The Moser’s twist theorem [9] guarantees the existence of arbitrarily large invariant curves diffeomorphic to circles and surrounding the origin in the -plane. Every such curves is the base of a time-periodic and flow-invariant cylinder in the extended phase space , which confines the solutions in the interior and which leads to a bound of these solutions.

The paper is organized as follows. In Sections 2.1 and 2.2, we give action-angle variables and some estimates which is useful for our proof. In Section 2.3, we will give an asymptotic expression of the Poincaré map and prove the main result by Moser’s twist theorem [9].

2. Proof of Theorem

2.1. Action-Angle Variables and Some Estimates

Without loss of generality and for brevity of arguments, we assume that the average value of vanishes; that is, . Hence the function is also 1-periodic in and is in .

System (1.5) is equivalent to the planar Hamiltonian system where Hamiltonian is .

In order to introduce action and angle variables, we first consider the auxiliary autonomous system which is integrable with the Hamiltonian The closed curves are just the integral curves of (2.2). Denote by the time period of the integral curve : and by the area enclosed by the closed curve . Let Then, .

It is easy to see that Denote Then,

The following estimates on the functions , , and and , , and are crucial for this paper. We first estimate and . Since is the area enclosed by the closed curve and axis when , we can easily prove that

Let , then we get Since , we have

We now give the estimates on the function and .

Lemma 2.1. One has where , . Note that here and below, one always uses , , or to indicate some constants.

Proof. Now, we estimate the first inequality. We choose as the new variable of integration, then we have Since and , we have . By direct computation, we have then we get When and sufficient large, there exits such that , so we have Since , we have then Observing that there is such that when and , we have By (2.15)–(2.18), we have , .
The proof of the second inequality is similar to the first one, so we only give the brief proof.
We choose as the new variable of integration, so we have By direct computation, we have By (2.20), we can easily get where .
By the similar way in estimating , we get which means that
Thus, we complete the proof of Lemma 2.1.

Remark 2.2. It follows from (2.10) and Lemma 2.1 that Thus, the time period is dominated by when is sufficiently large. By , we know is dominated by when is sufficiently large.

Remark 2.3. It also follows from the definition of , , and Remark 2.2 that In particular,

Remark 2.4. Note that is the inverse function of . By Remark 2.3, we have

We now carry out the standard reduction to the action-angle variables. For this purpose, we define the generating function , where is the part of the the closed curve connecting the point on the -axis and point .

We define the well-known map by It is well-known that the map is symplectic, since From the above discussion, we can easily get In the new variables , the system (2.1) becomes where where is the inverse function of and .

In order to estimate , we need the estimate on the functions and . For this purpose, we first give some definitions which are very similar to those in [4].

Define a function in terms of and by and a linear differential operator acting on functions of , according to where is a smooth function, and we denote .

The following equality (its proof can be found in [4]) is crucial for the proof of the following lemmas:

Before giving the estimates on and , we now prove some lemmas which will be used frequently in the following proof.

Lemma 2.5. Suppose that there is a constant such that , then one can find a constant and such that, for ,

Proof. We now prove (2.37). Let then By direct calculation, Since we have By , there is such that that is which means that That is, Thus, we complete the proof (2.37).
Now, we prove (2.38). Let Then, we have By direct computation, we have Since and , it follows that for so for , we have which means that By the definition of , we have Thus, we complete the proof of (2.38) and Lemma 2.5.

By Lemma 2.5, we have the following Lemma which is important to our estimation.

Lemma 2.6. One can find a constant such that, for , where .

Proof. When , we have By the definitions of and , we have and . By (2.37), we obtain Suppose that , we have We now proven that when , By direct computation, we have where denotes linear combination of functions with integer coefficients and .
Since , we have . By (2.37), we obtain By direct computation, we get By assumption (2.59), we have So, we have By (2.62) and (2.65), we have Thus, we have proved (2.55).
The inequality (2.56) can be proved by (2.38), and the process of proof is similar to that of (2.55), so we omit it.
Thus, we have proved Lemma 2.6.

Now, we give the estimates of and .

Lemma 2.7. For sufficient large and , the following estimates hold:

Proof. We now prove the first inequality. It is sufficient to prove that Case . Differentiating (2.30) by and noting , we have Now, we choose as the new variable of integration, so Observing that and simplifying, we have By (2.69)–(2.72), we have Since , by Lemma 2.6, we have We observe that where , so , which means that By (2.74) and (2.76), we have .
We suppose that where . We will prove .
For this purpose, we firstly estimate We differentiate in (2.73) and using (2.73), then we obtain Since , we have By (2.37), we have By (2.78)–(2.80), we have We suppose that when, We will prove that when, By direct computation, we have where and .
By the assume (2.77), we have By (2.55), we get By (2.85), (2.86) and noting the fact that we obtain By assumption (2.82), we have By (2.89) and the fact that , we have So, by (2.88) and (2.90), we have proved (2.83).
By (2.56), we have By (2.83) and (2.91), we have By the assumption (2.77), the fact that , and noting that where , we have By (2.92) and (2.94), we have which means
We now prove Since we have we have proved (2.83), so we have which means that The proof of Lemma 2.7 is complete.

Remark 2.8. Lemma 2.7 also holds when . Since the idea and the process of the proof is more easily than that of Lemma 2.7, we omit the details.

Now, we give the estimate of .

Lemma 2.9.

Proof. This lemma can be proved easily form the definition of and .

2.2. New Action and Angle Variables

Now, we are concerned with the Hamiltonian system (2.32) with Hamiltonian function given by (2.33). Note that This means that if one can solve form (2.32) as a function of ( and as parameters), then is also a Hamiltonian system with Hamiltonian function , and now, the action, angle, and time variables are , , and .

Form Remarks 2.3 and 2.4, we have Hence, by the implicit function theorem, there is a function such that By Lemma 2.9, we have So, there is a function with such that Let Then, From Remark 2.3, we have known the estimate of , so we need to give the estimate of . For this propose, we need firstly the following Lemma on the estimate of .

Lemma 2.10. The function possesses the following estimates: for .

Proof. From (2.108) and (2.110), it follows that When , by Lemma 2.9, we have When , we first denote By Remark 2.3, we observe that is increasing and By Lemma 2.9 and (2.115), we have So, By direct computation, we have By (2.118), we can easily get When , one may get where , . It is easy to verify for . This complete the proof.

Now, we give the estimates of .

Lemma 2.11. The function possesses the following estimates: for .

Proof. When . By Remark 2.3 and , we have By Lemma 2.10, we know Since it is easy to get that .
When . By direct computation, we have where , . Now, we need to estimate the first term of the integrand. The following equality is important: where , , , , , , . Assume that there are members: in which equal to 1. Noting that By the above discussions, we have By Lemma 2.10, we have known then we have

2.3. Proof of the Main Result

Up to now, we have given an equivalent form of (1.5), that is, the system (2.32), which is expressed in the action and angle variables . In this section, we first introduce some transformations such that in the transformed system, the perturbation terms of (2.32) depending on the new angle variable are very small if the new action variable is sufficiently large and then prove, by Moser’s twist theorem, the statement of Theorem 1.1.

Lemma 2.12. There is a canonical transformation of the form where the functions and are 1-periodic in and satisty uniformly for such that under this mapping, the system and the Hamiltonian function in (2.110) is changed into the form where with Moreover, the new perturbation possesses the estimate

Proof. We will look for the required transformation by a generating function in the following way: where the function will be given later. Under , the transformed system of (2.104) is of the form where By Taylor’s formula, one can write where We choose Then, is of the form (2.135).
We now show that satisfies (2.137). From Remark 2.3 and Lemma 2.11, it follows that for and . In particular if . So we can solve the second equation of (2.138) for , where the function satisfies Set Then, the canonical transformation is of the form (2.132). Moreover, similar to the proof of [5, Lemma  2], we can verify that for and as . Let It is not difficult to prove that for . Note that , we have .
Hence, we have The proof of Lemma 2.12 is complete.
For , we denote by the domain

Lemma 2.13. The Poincaré mapping of (2.134) has the intersection property on ; that is, if is an embedded circle in homotopic to a circle const. in , then .

Proof. The proof can be found in [5].
Definite a diffeomorphism Then, the system (2.134) under the transformation becomes where with defined through the transformation .
Now, we estimate and . Since we have then Moreover, we have When , we have so for all , where .
Since and are sufficiently small as , all solution of (2.155) exist for when the initial values are sufficiently large. Hence, the Poincaré map associated to (2.155) is well defined on as . In fact, by integrating (2.155) from to , we see that has the form where and satisfy the same estimates as those of and ; that is, where .
Since satisfies all the assumptions of Moser’s twist theorem [9], from which we conclude that for any satisfying There is an invariant curve of which is conjugated to pure rotation of the circle with rotation number . Tracing back to the system (2.32), gives rise to an invariant closed curve of the Poincaré map of (2.32) with rotation number which surrounds and is arbitrarily far away form the origin. Hence, all solutions of (1.5) are bounded. This completes the proof of the Theorem.

Acknowledgments

The authors would like to thank the referees for many helpful suggsetions. The work was supported by the National Natural Science Foundation of China (11071038) and the Natural Science Foundation of Jiangsu Province (BK2010420).