1 Introduction

Recently, the study of the periodic solutions of prescribed mean curvature equations has become very active; see [15] and the references therein. Meanwhile, the Liénard equation, Liénard system, and p-Laplacian equations are also studied by many people; see for example [68]. These kinds of equations have wide applications in many fields, such as physics, mechanics, and engineering.

In [9] Wang studied the following prescribed mean curvature Rayleigh equation with one deviating argument:

( φ p ( x ( t ) 1 + x 2 ( t ) ) ) +f ( t , x ( t ) ) +g ( t , x ( t τ ( t ) ) ) =e(t)
(1.1)

under the assumptions:

f ( t , u ) a | u | r , ( t , u ) R 2 and g ( t , u ) e ( t ) m 1 | u | m 2 , t R , x d .
(1.2)

By using the theory of coincidence degree, the author obtained some sufficient conditions for the existence and uniqueness of periodic solution in the case of τ(t)=0.

Stimulated by [9], we study the periodic solutions for a kind of prescribed mean curvature Liénard p-Laplacian equation with two deviating arguments in the form

( φ p ( x ( t ) 1 + x 2 ( t ) ) ) +f ( x ( t ) ) x (t)+g ( x ( t τ ( t ) ) ) +h ( x ( t γ ( t ) ) ) =e(t),
(1.3)

where φ p (s)= | s | p 2 s, p2, τ, γ are continuous functions with period T, τ (t)<1, γ (t)<1, and 0 T e(t)dt=0, f,g,h C 1 (R,R).

To the best of our knowledge, there are few results on this topic. The purpose of this paper is to establish a criterion to guarantee the existence of T-periodic solution. Our methods and results are different from the corresponding ones of [9]. So our results are essentially new.

2 Preliminaries

Let X and Y be real Banach spaces and L:D(L)XY be a Fredholm operator with index zero, that is, X=KerL X 1 and Y=ImL Y 1 . Furthermore, let P:XKerL and Q:Y Y 1 be the continuous projectors. Clearly, KerL(D(L) X 1 )={0}, thus the restriction L P =L | D ( L ) X 1 is invertible. Denote by K the inverse of L P .

Let Ω be an open bounded subset of X with D(L)ΩΦ. A map N: Ω ¯ Y is called L-compact in Ω ¯ if QN( Ω ¯ ) is bounded and the operator K(IQ)N: Ω ¯ X is compact. The following, Mawhin’s continuation theorem, is well known.

Lemma 2.1 (Gaines and Mawhin [10])

Suppose that X and Y are two Banach spaces, and L:D(L)XY is a Fredholm operator with index zero. Furthermore, ΩX is an open bounded subset and N: Ω ¯ Y is L-compact in Ω ¯ . If all the following conditions hold:

  1. (1)

    LxλNx, xΩD(L), λ(0,1);

  2. (2)

    NxImL, xΩKerL;

  3. (3)

    deg{JQN,ΩKerL,0}0, where J:ImQKerL is an isomorphism.

Then the equation Lx=Nx has a solution in Ω ¯ D(L).

To use Mawhin’s continuation theorem to study (1.3), we firstly transform (1.3) into the following form:

{ x 1 ( t ) = φ q ( x 2 ( t ) ) 1 φ q 2 ( x 2 ( t ) ) , x 2 ( t ) = f ( x 1 ( t ) ) φ q ( x 2 ( t ) ) 1 φ q 2 ( x 2 ( t ) ) g ( x 1 ( t τ ( t ) ) ) h ( x 1 ( t γ ( t ) ) ) + e ( t ) .
(2.1)

Denote by ϕ(x)= φ q ( x ) 1 φ q 2 ( x ) , φ q (s)= | s | q 2 s, 1/p+1/q=1, then prescribed mean curvature p-Laplacian operator φ p ( x 1 + x 2 ) is marked by ϕ 1 (x). Clearly, if x(t)= ( x 1 ( t ) , x 2 ( t ) ) is a T-periodic solution to (2.1), then x 1 (t) must be a T-periodic solution to (1.3).

Set C T 1 ={x:xC(R, R 2 ),x(t+T)x(t)} with the norm | x | 0 = max t [ 0 , T ] |x(t)|, X=Y={x(t)= ( x 1 ( t ) , x 2 ( t ) ) C(R, R 2 ):x(t+T)x(t)} with the norm x=max{ | x 1 | 0 , | x 2 | 0 }, | x | 2 = ( 0 T x 2 ( t ) d t ) 1 / 2 . Then X and Y are Banach spaces. Let

L:D(L)XY,Lx= x = ( x 1 x 2 ) ,
(2.2)
N : X Y , where  N x = ( ϕ ( x 2 ( t ) ) f ( x 1 ( t ) ) ϕ ( x 2 ( t ) ) g ( x 1 ( t τ ( t ) ) ) h ( x 1 ( t γ ( t ) ) ) + e ( t ) ) .
(2.3)

It is easy to see that KerL= R 2 , ImL={x:xY, 0 T x(s)ds=0}. So L is a Fredholm operator with index zero. Define projectors P:XKerL and Q:YImQ by

Px= 1 T 0 T x(s)ds,Qy= 1 T 0 T y(s)ds,

and let K be the inverse of L | Ker P D ( L ) . Obviously KerL=ImQ= R 2 ,

[ K y ] 1 (t)= 0 T k(t,s)y(s)ds,
(2.4)

where

k(t,s)= { s T , 0 s < t T , s T T , 0 t s T .

From (2.3) and (2.4), one can easily see that N is L-compact on Ω, where Ω is an open bounded subset of X. The following lemma is useful to estimate a priori bounds of periodic solutions of (2.1).

Lemma 2.2 (Liu et al. [11])

Suppose that x(t) C T 1 , and that there is a point t R such that x( t )=0, then

0 T |x(t) | 2 dt T 2 π 2 0 T | x (t) | 2 dt.

Lemma 2.3 ([12])

Let τ C T 1 , | τ | 0 <1. If x(t) C T 1 then

0 T |x ( t τ ( t ) ) | 2 dt τ 0 0 T |x(t) | 2 dt,

where τ 0 = 1 1 | τ | 0 , γ(t) satisfies the above inequality and γ 0 = 1 1 | γ | 0 .

At the end of this section, we list the basic assumptions which will be used in Section 3.

(H1) (g( x 1 )g( x 2 ))( x 1 x 2 )>0, x 1 , x 2 R and x 1 x 2 , (h( x 1 )h( x 2 ))( x 1 x 2 )>0, x 1 , x 2 R and x 1 x 2 .

(H2) There is a constant d>0 such that x(g(x)+h(x))>0 if and only if |x|>d.

(H3) There exist nonnegative constants k 1 , k 2 , c 1 , c 2 such that

| g ( x ) | k 1 | x | + c 1 , | x | > d ; | h ( x ) | k 2 | x | + c 2 , | x | > d .

(H4) There is a constant σ>0 such that |f(s)|σ for all sR.

3 Main results and the proof

In this section we state the main results and give its proof.

Theorem 3.1 Suppose the assumptions (H1)-(H4) hold. Then (1.3) has at least one T-periodic solution provided T π ( k 1 τ 0 + k 2 γ 0 )<σ.

Proof Let Ω 1 ={xX:Lx=λNx,λ(0,1)}, suppose x(t) Ω 1 , then

{ x 1 ( t ) = λ φ q ( x 2 ( t ) ) 1 φ q 2 ( x 2 ( t ) ) λ ϕ ( x 2 ( t ) ) , x 2 ( t ) = λ f ( x 1 ( t ) ) φ q ( x 2 ( t ) ) 1 φ q 2 ( x 2 ( t ) ) λ g ( x 1 ( t τ ( t ) ) ) λ h ( x 1 ( t γ ( t ) ) ) + λ e ( t ) .
(3.1)

Integrating both sides of the second equation of (3.1) from 0 to T, we have

0 T ( g ( x 1 ( t τ ( t ) ) ) + h ( x 1 ( t γ ( t ) ) ) ) dt=0,

which implies that there exists a point ξ[0,T] such that

g ( x 1 ( ξ τ ( ξ ) ) ) +h ( x 1 ( ξ γ ( ξ ) ) ) =0.
(3.2)

Now we claim that there must exist a point t 0 R such that

| x 1 ( t 0 )|d.
(3.3)

In fact, let ξτ(ξ)= t 1 , ξγ(ξ)= t 2 , then from (3.2), g( x 1 ( t 1 ))+h( x 1 ( t 2 ))=0. If x 1 ( t 1 )= x 1 ( t 2 ) holds, (3.3) is clearly true. Now assume x 1 ( t 1 )< x 1 ( t 2 ).

Set F(x)g(x)+h(x), according to (H1), one has

F ( x 1 ( t 1 ) ) = g ( x 1 ( t 1 ) ) + h ( x 1 ( t 1 ) ) < g ( x 1 ( t 1 ) ) + h ( x 1 ( t 2 ) ) < g ( x 1 ( t 2 ) ) + h ( x 1 ( t 2 ) ) = F ( x 1 ( t 2 ) ) ,

namely F( x 1 ( t 1 ))<0<F( x 1 ( t 2 )), then there must exist a t 0 such that F( x 1 ( t 0 ))=0. That is, g( x 1 ( t 0 ))+h( x 1 ( t 0 ))=0. According to (H2), (3.3) holds. If x 1 ( t 1 )> x 1 ( t 2 ) holds, by a similar method, we can show that (3.3) is also true.

Let t ˜ =nT+ t 0 , where t ˜ [0,T] and n is an integer. Then

| x 1 (t)|=| x 1 ( t ˜ )+ t ˜ t x 1 (s)ds|d+ 0 T | x 1 (s)|dsfor all s[0,T].

Thus we get

| x 1 | 0 = max t [ 0 , T ] | x 1 (t)|d+ T | x 1 | 2 .
(3.4)

On the other hand, multiplying both sides of the second equation of (3.1) by x 1 (t) and integrating over [0,T], we have

0 T ( ϕ 1 ( x 1 ( t ) λ ) ) x 1 ( t ) d t + 0 T f ( x 1 ( t ) ) ( x 1 ( t ) ) 2 d t + λ 0 T g ( x 1 ( t τ ( t ) ) ) x 1 ( t ) d t + λ 0 T h ( x 1 ( t γ ( t ) ) ) x 1 ( t ) d t = λ 0 T e ( t ) x 1 ( t ) d t ,
(3.5)

where ϕ 1 ( x 1 ( t ) λ )= φ p ( x 1 ( t ) λ 1 + ( x 1 ( t ) λ ) 2 ). Furthermore, in view of (H4), we have

| 0 T f ( x 1 ( t ) ) ( x 1 ( t ) ) 2 d t | = 0 T | f ( x 1 ( t ) ) | | x 1 ( t ) | 2 d t | 0 T ( ϕ 1 ( x 1 ( t ) λ ) ) x 1 ( t ) d t | + λ | 0 T g ( x 1 ( t τ ( t ) ) ) x 1 ( t ) d t | + λ | 0 T h ( x 1 ( t γ ( t ) ) ) x 1 ( t ) d t | + λ | 0 T e ( t ) x 1 ( t ) d t | .
(3.6)

Denote w(t)= ϕ 1 ( x 1 (t)/λ), then

0 T ( ϕ 1 ( x 1 ( t ) / λ ) ) x 1 (t)dt=λ 0 T ϕ ( ω ( t ) ) dω(t)=0.

Together with (3.6) and the fact λ(0,1) yields

σ 0 T | x 1 ( t ) | 2 d t | 0 T g ( x 1 ( t τ ( t ) ) ) x 1 ( t ) d t | + | 0 T h ( x 1 ( t γ ( t ) ) ) x 1 ( t ) d t | + | 0 T e ( t ) x 1 ( t ) d t | .

Set

E 1 = { t [ 0 , 1 ] , | x 1 ( t τ ( t ) ) | d } , E 2 = { t [ 0 , 1 ] , | x 1 ( t τ ( t ) ) | > d } , E 3 = { t [ 0 , 1 ] , | x 1 ( t γ ( t ) ) | d } , E 4 = { t [ 0 , 1 ] , | x 1 ( t γ ( t ) ) | > d } .

In view of Lemma 2.3, we obtain

σ | x 1 | 2 2 E 1 | g ( x 1 ( t τ ( t ) ) ) | | x 1 ( t ) | d t + E 2 | g ( x 1 ( t τ ( t ) ) ) | | x 1 ( t ) | d t + E 3 | h ( x 1 ( t γ ( t ) ) ) | | x 1 ( t ) | d t + E 4 | h ( x 1 ( t γ ( t ) ) ) | | x 1 ( t ) | d t + 0 T | e ( t ) | | x 1 ( t ) | d t g d 0 T | x 1 ( t ) | d t + h d 0 T | x 1 ( t ) | d t + k 1 0 T | x 1 ( t τ ( t ) ) | | x 1 ( t ) | d t + k 2 0 T | x 1 ( t γ ( t ) ) | | x 1 ( t ) | d t + c 1 0 T | x 1 ( t ) | d t + c 2 0 T | x 1 ( t ) | d t + | e | 0 T | x 1 ( t ) | d t ( g d + h d + c 1 + c 2 + | e | ) T | x 1 | 2 + k 1 τ 0 | x 1 | 2 | x 1 | 2 + k 2 γ 0 | x 1 | 2 | x 1 | 2 ,
(3.7)

where τ 0 , γ 0 are defined by Lemma 2.3, and

g d = max | x | d |g(x)|, h d = max | x | d |h(x)|, | e | = max t [ 0 , T ] |e(t)|.

From (3.3), there exists a point t 0 R such that | x 1 ( t 0 )|d. Let ϖ(t)= x 1 (t) x 1 ( t 0 ), we have ϖ(t+T)=ϖ(t), ϖ (t)= x 1 (t), and ϖ( t 0 )=0; then, by Lemma 2.2,

| ϖ | 2 T π | ϖ | 2 = T π | x 1 | 2 .

By the Minkowski inequality, we get

| x 1 | 2 = ( 0 T | x 1 ( t ) | 2 d t ) 1 2 = ( 0 T | ϖ ( t ) + x 1 ( t 0 ) | 2 d t ) 1 2 ( 0 T | x 1 ( t 0 ) | 2 d t ) 1 2 + ( 0 T | ϖ ( t ) | 2 d t ) 1 2 .
(3.8)

So

| x 1 | 2 T d+ | ϖ | 2 T d+ T π | x 1 | 2 .

Combining (3.7) and (3.8), we obtain

σ | x 1 | 2 2 ( g d + h d + c 1 + c 2 + | e | ) T | x 1 | 2 + ( k 1 τ 0 + k 2 γ 0 ) | x 1 | 2 ( T d + T π | x 1 | 2 ) = ( g d + h d + c 1 + c 2 + | e | ) T | x 1 | 2 + ( k 1 τ 0 + k 2 γ 0 ) T d | x 1 | 2 + T π ( k 1 τ 0 + k 2 γ 0 ) | x 1 | 2 2 .
(3.9)

Since T π ( k 1 τ 0 + k 2 γ 0 )<σ, there is a constant M 0 >0 independent of λ such that | x 1 | 2 M 0 , i.e., | x 1 | 0 d+ T | x 1 | 2 d+ T M 0 := M 1 .

By the first equation of (3.1), we have

0 T λ φ q ( x 2 ( t ) ) 1 φ q 2 ( x 2 ( t ) ) dt=0,

together with φ q (0)=0, which implies that there is a constant η[0,T] such that x 2 (η)=0. Hence

| x 2 | 0 0 η | x 2 (s)|ds 0 T | x 2 (s)|ds.
(3.10)

By the two equations of (3.1) and Hölder’s inequality, we obtain

0 T | x 2 ( s ) | d s 0 T | f ( x 1 ( t ) ) x 1 ( t ) | d t + λ 0 T | g ( x 1 ( t τ ( t ) ) ) | d t + λ 0 T | h ( x 1 ( t γ ( t ) ) ) | d t + λ 0 T | e ( t ) | d t 0 T | f ( x 1 ( t ) ) | | x 1 ( t ) | d t + g M 1 T + h M 1 T + | e | T < | f | 0 M 0 T + ( g M 1 + h M 1 + | e | ) T ,
(3.11)

where | f | 0 = max | x 1 | M 1 |f( x 1 )|, g M 1 = max | x 1 | M 1 |g( x 1 )|, h M 1 = max | x 1 | M 1 |h( x 1 )|. By (3.10), (3.11), | x 2 | 0 ( g M 1 + h M 1 + | e | )T+ | f | 0 M 0 T +d=: M 2 .

Let Ω 2 ={xKerL:NxImL}. If x Ω 2 , then QNx=0. In view of 0 T e(t)dt=0, we have

{ φ q ( x 2 ) 1 φ q 2 ( x 2 ) = 0 , 0 T g ( x 1 ) d t + 0 T h ( x 1 ) d t = 0 .
(3.12)

So x 2 =0 M 2 . Together with (H2) yields

| x 1 |d M 2 .

Let Ω={x:x=( x 1 , x 2 )X, | x 1 | 0 < M 1 +1, | x 2 | 0 < M 2 +1}. Then Ω( Ω 1 Ω 2 ) and Ω is a bounded open set of X. So (1) and (2) of Lemma 2.1 are satisfied.

In the next step we show that condition (3) of Lemma 2.1 holds. Define a linear isomorphism J:ImQKerL by J( x 1 , x 2 )=( x 2 , x 1 ), and let

H(v,μ):=μv+ 1 μ T JQNv,(v,μ)Ω×[0,1].

The direct computation and (H2) show that for (x,μ)(ΩKerL)×[0,1],

x H(x,μ)=μ ( x 1 2 + x 2 2 ) + 1 μ T ( g ( x 1 ) + h ( x 1 ) ) x 1 + 1 μ T | x 2 | q 2 x 2 2 1 φ q 2 ( x 2 ) >0.

Thus, x H(x,μ)0 for (x,μ)ΩKerL×[0,1], which implies

deg { J Q N , Ω Ker L , 0 } = deg { H ( x , 0 ) , Ω Ker L , 0 } = deg { H ( x , 1 ) , Ω Ker L , 0 } = deg { I , Ω Ker L , 0 } 0 .

Condition (3) of Lemma 2.1 holds. By Lemma 2.1, the equation Lx=Nx has a solution. This completes the proof of Theorem 3.1. □

We can use a similar method to conclude the following result, the details are omitted.

Theorem 3.2 Suppose (H3), (H4), and the following assumptions hold:

( H 1 ) (g( x 1 )g( x 2 ))( x 1 x 2 )<0, x 1 , x 2 R, and x 1 x 2 , (h( x 1 )h( x 2 ))( x 1 x 2 )<0, x 1 , x 2 R, and x 1 x 2 .

( H 2 ) There is a constant d>0 such that x(g(x)+h(x))<0 if and only if |x|>d.

Then (1.3) has at least one T-periodic solution if T π ( k 1 τ 0 + k 2 γ 0 )<σ.

4 Uniqueness results

It is difficult to establish the uniqueness of the T-periodic solution of (1.3), but in the special case we obtain the uniqueness of (1.3).

Theorem 4.1 Assume ( H 1 ) holds, and τ(t)= ε 1 , γ(t)= ε 2 ( ε 1 , ε 2 are sufficiently small constants). Then (1.3) has at most one T-periodic solution.

Proof Denote

I(x)= 0 x f(w)dw,y(t)= φ p ( x ( t ) 1 + x 2 ( t ) ) +I ( x ( t ) ) .

Then (1.3) can be converted into the following form:

{ x ( t ) = φ q ( y ( t ) I ( x ( t ) ) ) 1 φ q 2 ( y ( t ) I ( x ( t ) ) ) , y ( t ) = g ( x ( t ε 1 ) ) h ( x ( t ε 2 ) ) + e ( t ) .
(4.1)

Suppose that x i (t) (i=1,2) are two T-periodic solutions of (1.3). Then from (4.1), we have

{ x i ( t ) = φ q ( y i ( t ) I ( x i ( t ) ) ) 1 φ q 2 ( y i ( t ) I ( x i ( t ) ) ) , y i ( t ) = g ( x i ( t ε 1 ) ) h ( x i ( t ε 2 ) ) + e ( t ) , i = 1 , 2 .
(4.2)

Let u(t)= x 1 (t) x 2 (t), v(t)= y 1 (t) y 2 (t), which together with (4.2) yields

{ u ( t ) = φ q ( y 1 ( t ) I ( x 1 ( t ) ) ) 1 φ q 2 ( y 1 ( t ) I ( x 1 ( t ) ) ) φ q ( y 2 ( t ) I ( x 2 ( t ) ) ) 1 φ q 2 ( y 2 ( t ) I ( x 2 ( t ) ) ) , v ( t ) = ( g ( x 1 ( t ε 1 ) ) g ( x 2 ( t ε 1 ) ) ) v ( t ) = ( h ( x 1 ( t ε 2 ) ) h ( x 2 ( t ε 2 ) ) ) .
(4.3)

Now, we claim that

v(t)0,tR.

We argue by contradiction, in view of v C 2 [0,T] and v(t+T)=v(t), tR, we obtain max t R v(t)>0.

Then there must exist t R such that

{ v ( t ) = max t [ 0 , T ] v ( t ) = max t R v ( t ) > 0 , v ( t ) = ( g ( x 1 ( t ε 1 ) ) g ( x 2 ( t ε 1 ) ) ) v ( t ) = ( h ( x 1 ( t ε 2 ) ) h ( x 2 ( t ε 2 ) ) ) = 0 , v ( t ) = ( g ( x 1 ( t ε 1 ) ) x 1 ( t ε 1 ) g ( x 2 ( t ε 1 ) ) x 2 ( t ε 1 ) ) v ( t ) = ( h ( x 1 ( t ε 2 ) ) x 1 ( t ε 2 ) h ( x 2 ( t ε 2 ) ) x 2 ( t ε 2 ) ) 0 .
(4.4)

Noticing ( H 1 ), that is, g (x)<0, h (x)<0, ε 1 , ε 2 sufficiently small, it follows from the second equation of (4.4) that

x 1 ( t ε 1 ) x 2 ( t ε 1 ) , x 1 ( t ε 2 ) x 2 ( t ε 2 ) .

Then from the third equation of (4.4), we get

v ( t ) = g ( x 1 ( t ε 1 ) ) ( x 1 ( t ε 1 ) x 2 ( t ε 1 ) ) h ( x 1 ( t ε 2 ) ) ( x 1 ( t ε 2 ) x 2 ( t ε 2 ) ) = g ( x 1 ( t ε 1 ) ) ( φ q ( y 1 ( t ε 1 ) I ( x 1 ( t ε 1 ) ) ) 1 φ q 2 ( y 1 ( t ε 1 ) I ( x 1 ( t ε 1 ) ) ) φ q ( y 2 ( t ε 1 ) I ( x 2 ( t ε 1 ) ) ) 1 φ q 2 ( y 2 ( t ε 1 ) I ( x 2 ( t ε 1 ) ) ) ) h ( x 1 ( t ε 2 ) ) ( φ q ( y 1 ( t ε 2 ) I ( x 1 ( t ε 1 ) ) ) 1 φ q 2 ( y 1 ( t ε 2 ) I ( x 1 ( t ε 2 ) ) ) φ q ( y 2 ( t ε 2 ) I ( x 2 ( t ε 2 ) ) ) 1 φ q 2 ( y 2 ( t ε 2 ) I ( x 2 ( t ε 2 ) ) ) ) .

In view of g ( x 1 ( t ε 1 ))<0, h ( x 1 ( t ε 2 ))<0, v( t )= y 1 ( t ) y 2 ( t )>0, x 1 x 2 , φ q (x) are increasing on x, and ε 1 , ε 2 are sufficiently small, thus we have

v ( t ) >0.

This contradicts the third equation of (4.4), which implies that

v(t)= y 1 (t) y 2 (t)0,tR.

By using a similar argument, we can also show that y 2 (t) y 1 (t)0.

Hence, we obtain

y 1 (t) y 2 (t),tR.

Then from (4.3) and g (x)<0, we have x 1 (t ε 1 ) x 2 (t ε 1 ), x 1 (t ε 2 ) x 2 (t ε 2 ), tR. That is

x 1 (t) x 2 (t).

Therefore, (1.3) has at most one T-periodic solution. The proof of Theorem 4.1 is now complete. □

From Theorem 3.2 and Theorem 4.1, we see that (1.3) has an unique T-periodic solution when τ(t)= ε 1 , γ(t)= ε 2 .

5 An example

In this section we give an example to illustrate the application of Theorem 3.1.

Example 5.1 Consider the prescribed mean curvature Liénard equation with two deviating arguments:

( φ 4 ( x ( t ) 1 + x 2 ( t ) ) ) + f ( x ( t ) ) x ( t ) + g ( x ( t cos 3 π t 4 π ) ) + h ( x ( t sin 3 π t 4 π ) ) = sin 3 π t ,
(5.1)

where f(x)=σ+ x 2 , σ>5, τ(t)= cos 3 π t 4 π , γ(t)= sin 3 π t 4 π , e(t)=sin3πt, g(x)= 85 16 x+6, h(x)=5x+2. So τ 0 = γ 0 =4. Choosing T= 2 3 , d=10, k 1 = 85 16 , k 2 =5, then ( k 1 τ 0 + k 2 γ 0 ) T π 4.38<5<σ. It is obvious that (H1), (H2), (H3), and (H4) hold. Then it follows from Theorem 3.1 that (5.1) has at least one 2 3 -periodic solution.