Abstract

Oscillation criteria obtained by Kusano and Onose (1973) and by Belohorec (1969) are extended to second-order sublinear impulsive differential equations of Emden-Fowler type: 𝑥(𝑡)+𝑝(𝑡)|𝑥(𝜏(𝑡))|𝛼1𝑥(𝜏(𝑡))=0, 𝑡𝜃𝑘; Δ𝑥(𝑡)|𝑡=𝜃𝑘+𝑞𝑘|𝑥(𝜏(𝜃𝑘))|𝛼1𝑥(𝜏(𝜃𝑘))=0; Δ𝑥(𝑡)|𝑡=𝜃𝑘=0, (0<𝛼<1) by considering the cases 𝜏(𝑡)𝑡 and 𝜏(𝑡)=𝑡, respectively. Examples are inserted to show how impulsive perturbations greatly affect the oscillation behavior of the solutions.

1. Introduction

We deal with second-order sublinear impulsive differential equations of the form𝑥||||(𝑡)+𝑝(𝑡)𝑥(𝜏(𝑡))𝛼1𝑥(𝜏(𝑡))=0,𝑡𝜃𝑘,Δ𝑥||(𝑡)𝑡=𝜃𝑘+𝑞𝑘||𝑥𝜏𝜃𝑘||𝛼1𝑥𝜏𝜃𝑘||=0,Δ𝑥(𝑡)𝑡=𝜃𝑘=0,(1.1) where 0<𝛼<1, 𝑡𝑡0, and 𝑘𝑘0 for some 𝑡0+ and 𝑘0, {𝜃𝑘} is a strictly increasing unbounded sequence of positive real numbers, ||Δ𝑧(𝑡)𝑡=𝜃𝜃=𝑧+𝑧(𝜃𝜃),𝑧=lim𝑡𝜃𝑧(𝑡).(1.2)

Let PLC(𝐽,𝑅) denote the set of all real-valued functions 𝑢 defined on 𝐽 such that 𝑢 is continuous for all 𝑡𝐽 except possibly at 𝑡=𝜃𝑘 where 𝑢(𝜃±𝑘) exists and 𝑢(𝜃𝑘)=𝑢(𝜃𝑘).

We assume in the sequel that (a)𝑝PLC([𝑡0,),), (b){𝑞𝑘} is a sequence of real numbers, (c)𝜏𝐶([𝑡0,),+), 𝜏(𝑡)𝑡, lim𝑡𝜏(𝑡)=.

By a solution of (1.1) on an interval 𝐽[𝑡0,), we mean a function 𝑥(𝑡) which is defined on 𝐽 such that 𝑥,𝑥,𝑥PLC(𝐽) and which satisfies (1.1). Because of the requirement Δ𝑥(𝑡)|𝑡=𝜃𝑘=0 every solution of (1.1) is necessarily continuous.

As usual we assume that (1.1) has solutions which are nontrivial for all large 𝑡. Such a solution of (1.1) is called oscillatory if it has no last zero and nonoscillatory otherwise.

In case there is no impulse, (1.1) reduces to Emden-Fowler equation with delay𝑥||||(𝑡)+𝑝(𝑡)𝑥(𝜏(𝑡))𝛼1𝑥(𝜏(𝑡))=0,0<𝛼<1,(1.3) and without delay𝑥+𝑝(𝑡)|𝑥|𝛼1𝑥=0,0<𝛼<1.(1.4)

The problem of oscillation of solutions of (1.3) and (1.4) has been considered by many authors. Kusano and Onose [1] see also [2, 3] proved the following necessary and sufficient condition for oscillation of (1.3).

Theorem 1.1. If 𝑝(𝑡)0, then a necessary and sufficient condition for every solution of (1.3) to be oscillatory is that []𝜏(𝑡)𝛼𝑝(𝑡)𝑑𝑡=.(1.5)

The condition 𝑝(𝑡)0 is required only for the sufficiency part, and no similar criteria is available for 𝑝(𝑡) changing sign, except in the case 𝜏(𝑡)=𝑡. Without imposing a sign condition on 𝑝(𝑡), Belohorec [4] obtained the following sufficient condition for oscillation of (1.4).

Theorem 1.2. If 𝑡𝛽𝑝(𝑡)𝑑𝑡=(1.6) for some 𝛽[0,𝛼], then every solution of (1.4) is oscillatory.

Compared to the large body of papers on oscillation of differential equations, there is only little known about the oscillation of impulsive differential equations; see [57] for equations with delay and [813] for equations without delay. For some applications of such equations, we may refer to [1418]. The books [19, 20] are good sources for a general theory of impulsive differential equations.

The object of this paper is to extend Theorems 1.1 and 1.2 to impulsive differential equations of the form (1.1). The results show that the impulsive perturbations may greatly change the oscillatory behavior of the solutions. A nonoscillatory solution of (1.3) or (1.4) may become oscillatory under impulsive perturbations.

The following two lemmas are crucial in the proof of our main theorems. The first lemma is contained in [21] and the second one is extracted from [22].

Lemma 1.3. If each 𝐴𝑖 is continuous on [𝑎,𝑏], then 𝑏𝑎𝑠𝜃𝑖<𝑏𝐴𝑖(𝑠)𝑑𝑠=𝑎𝜃𝑖<𝑏𝜃𝑖𝑎𝐴𝑖(𝑠)𝑑𝑠.(1.7)

Lemma 1.4. Fix𝐽=[𝑎,𝑏], let 𝑢,𝜆𝐶(𝐽,+), 𝐶(+,+), and 𝑐+, and let {𝜆𝑘} a sequence of positive real numbers. If 𝑢(𝐽)𝐼+ and 𝑢(𝑡)𝑐+𝑡𝑎𝜆(𝑠)(𝑢(𝑠))𝑑𝑠+𝑎<𝜃𝑘<𝑡𝜆𝑘𝑢𝜃𝑘,𝑡𝐽,(1.8) then 𝑢(𝑡)𝐺1𝐺(𝑐)+𝑡𝑎𝜆(𝑠)𝑑𝑠+𝑎<𝜃𝑘<𝑡𝜆𝑘[,𝑡𝑎,𝛽),(1.9) where 𝐺(𝑢)=𝑢𝑢0𝑑𝑥(𝑥),𝑢,𝑢0𝐼,𝛽=sup𝜈𝐽𝐺(𝑐)+𝑡𝑎𝜆(𝑠)𝑑𝑠+𝑎<𝜃𝑘<𝑡𝜆𝑘.𝐺(𝐼),𝑎𝑡𝜈(1.10)

2. The Main Results

We first establish a necessary and sufficient condition for oscillation of solutions of (1.1) when 𝜏(𝑡)𝑡.

Theorem 2.1. If []𝜏(𝑡)𝛼||||𝑝(𝑡)𝑑𝑡+𝜏𝜃𝑘𝛼||𝑞𝑘||<,(2.1) then (1.1) has a solution 𝑥(𝑡) satisfying lim𝑡𝑥(𝑡)𝑡=𝑎0.(2.2)

Proof. Choose 𝑡1max{1,𝑡0}. In view of Lemma 1.3 by integrating (1.1) twice from 𝑡0 to 𝑡, we obtain 𝑡𝑥(𝑡)=𝑥1𝑥𝑡1𝑡𝑡1𝑡1𝜃𝑘<𝑡𝑞𝑘||𝑥𝜏𝜃𝑘||𝛼1𝑥𝜏𝜃𝑘𝑡𝜃𝑘𝑡𝑡1||||(𝑡𝑠)𝑝(𝑠)𝑥(𝜏(𝑠))𝛼1𝑥((𝜏(𝑠)))𝑑𝑠,𝑡𝑡1.(2.3) Set 𝑢(𝑡)=𝑐+𝑡1𝜃𝑘<𝑡||𝑞𝑘||||𝑥𝜏𝜃𝑘||𝛼+𝑡𝑡1||||||||𝑝(𝑠)𝑥(𝜏(𝑠))𝛼𝑑𝑠,𝑡𝑡1,(2.4) where 𝑐=|𝑥(𝑡1)|+|𝑥(𝑡1)|. Then ||||𝑥(𝑡)𝑡𝑢(𝑡),𝑡𝑡1.(2.5) Let 𝑡2𝑡1 be such that 𝜏(𝑡)𝑡1 for all 𝑡𝑡2. Replacing 𝑡 by 𝜏(𝑡) in (2.5) and using the increasing character of 𝑢(𝑡), we see that ||||𝑥(𝜏(𝑡))𝜏(𝑡)𝑢(𝑡),𝑡𝑡2.(2.6) From (2.4), we also see that 𝑢||||||||(𝑡)=𝑝(𝑡)𝑥(𝜏(𝑡))𝛼,𝑡𝜃𝑘,||(2.7)Δ𝑢(𝑡)𝑡=𝜃𝑘=||𝑞𝑘||||𝑥𝜏𝜃𝑘||𝛼(2.8) for 𝑡𝑡2 and 𝜃𝑘𝑡2. Now, in view of (2.6) and (2.8), an integration of (2.7) from 𝑡2 to 𝑡 leads to 𝑢(𝑡)𝑐+𝑡𝑡2||||[]𝑝(𝑠)𝜏(𝑠)𝛼[]𝑢(𝑠)𝛼𝑑𝑠+𝑡2𝜃𝑘<𝑡||𝑞𝑘||𝜏𝜃𝑘𝛼𝑢𝜃𝑘𝛼.(2.9) Applying Lemma 1.4 with (𝑥)=𝑥𝛼||||[],𝜆(𝑠)=𝑝(𝑠)𝜏(𝑠)𝛼,𝜆𝑘=||𝑞𝑘||𝜏𝜃𝑘𝛼,(2.10) we easily see that 𝑢(𝑡)𝐺1𝐺(𝑐)+𝑡𝑡2||||[]𝑝(𝑠)𝜏(𝑠)𝛼𝑑𝑠+𝑡2𝜃𝑘<𝑡||𝑞𝑘||𝜏𝜃𝑘𝛼.(2.11) Since 𝑢𝐺(𝑢)=1𝛼𝑢1𝛼01𝛼1𝛼,𝐺1(𝑢)=(1𝛼)𝑢+𝑢01𝛼1/(1𝛼),(2.12) the inequality (2.11) becomes 𝑐𝑢(𝑡)1𝛼+(1𝛼)𝑡𝑡1||||[]𝑝(𝑠)𝜏(𝑠)𝛼𝑑𝑠+(1𝛼)𝑡1𝜃𝑘<𝑡||𝑞𝑘||𝜏𝜃𝑘𝛼1/(1𝛼),(2.13) from which, on using (2.1), we have 𝑢(𝑡)𝑐1,𝑡𝑡2,(2.14) where 𝑐1=𝑐1𝛼+(1𝛼)𝑡1||||[]𝑝(𝑠)𝜏(𝑠)𝛼𝑑𝑠+(1𝛼)𝑡1𝜃𝑘<||𝑞𝑘||𝜏𝜃𝑘𝛼1/(1𝛼).(2.15) In view of (2.5), (2.6), and (2.14) we see that ||||𝑥(𝑡)𝑐1||||𝑡,𝑥(𝜏(𝑡))𝑐1𝜏(𝑡),𝑡𝑡2.(2.16)
To complete the proof it suffices to show that 𝑥(𝑡) approaches a nonzero limit as 𝑡 tends to . To see this we integrate (1.1) from 𝑡2 to 𝑡 to get𝑥(𝑡)=𝑥𝑡1𝑡𝑡2||||𝑝(𝑠)𝑥(𝜏(𝑠))𝛼1𝑥(𝜏(𝑠))𝑑𝑠𝑡2𝜃𝑘<𝑡𝑞𝑘||𝑥𝜏𝜃𝑘||𝛼1𝑥𝜏𝜃𝑘.(2.17) Employing (2.16) we have 𝑡2||||𝑝(𝑠)𝑥(𝜏(𝑠))𝛼𝑑𝑠𝑐𝛼1𝑡2||||[]𝑝(𝑠)𝜏(𝑠)𝛼𝑑𝑠<,𝑡2𝜃𝑘<||𝑞𝑘𝑥𝜏𝜃𝑘||𝛼𝑐𝛼1𝑡2𝜃𝑘<||𝑞𝑘||𝜏𝜃𝑘𝛼<.(2.18) Therefore, lim𝑡𝑥(𝑡)=𝐿 exists. Clearly, we can make 𝐿0 by requiring that 𝑥𝑡2>𝑐𝛼1𝑡2||||[]𝑝(𝑠)𝜏(𝑠)𝛼𝑑𝑠+𝑡2𝜃𝑘<||𝑞𝑘||𝜏𝜃𝑘𝛼,(2.19) which is always possible by arranging 𝑡2.

Theorem 2.2. Suppose that 𝑝 and {𝑞𝑘} are nonnegative. Then every solution of (1.1) is oscillatory if and only if []𝜏(𝑡)𝛼𝑝(𝑡)𝑑𝑡+𝜏𝜃𝑘𝛼𝑞𝑘=.(2.20)

Proof. Let (2.20) fail to hold. Then, by Theorem 2.1 we see that there is a solution 𝑥(𝑡) which satisfies (2.2). Clearly, such a solution is nonoscillatory. This proves the necessity.
To show the sufficiency, suppose that (2.20) is valid but there is a nonoscillatory solution 𝑥(𝑡) of (1.1). We may assume that 𝑥(𝑡) is eventually positive; the case 𝑥(𝑡) being eventually negative is similar. Clearly, there exists 𝑡1𝑡0 such that 𝑥(𝜏(𝑡))>0 for all 𝑡𝑡1. From (1.1), we have that𝑥(𝑡)0for𝑡𝑡1,𝑡𝜃𝑘.(2.21) Thus, 𝑥(𝑡) is decreasing on every interval not containing 𝑡=𝜃𝑘. From the impulse conditions in (1.1), we also have Δ𝑥(𝜃𝑘)0. Therefore, we deduce that 𝑥(𝑡) is nondecreasing on [𝑡1,).
We may claim that 𝑥(𝑡) is eventually positive. Because if 𝑥(𝑡)<0 eventually, then 𝑥(𝑡) becomes negative for large values of 𝑡. This is a contradiction.
It is now easy to show that𝑥(𝑡)𝑡𝑡1𝑥(𝑡),𝑡𝑡1.(2.22) Therefore, 𝑡𝑥(𝑡)2𝑥(𝑡),𝑡𝑡2=2𝑡1.(2.23) Let 𝑡3𝑡2 be such that 𝜏(𝑡)𝑡2 for 𝑡𝑡3. Using (2.23) and the nonincreasing character of 𝑥(𝑡), we have 𝑥(𝜏(𝑡))𝜏(𝑡)2𝑥(𝑡),𝑡𝑡3,(2.24) and so, by (1.1), 𝑥(𝑡)+2𝛼[]𝑝(𝑡)𝜏(𝑡)𝛼𝑥(𝑡)𝛼0,𝑡𝜃𝑘.(2.25) Dividing (2.25) by [𝑥(𝑡)]𝛼 and integrating from 𝑡3 to 𝑡, we obtain 𝑡3𝜃𝑘<𝑡𝑥𝜃𝑘1𝛼𝑥𝜃𝑘𝑞𝑘𝑥𝜏𝜃𝑘𝛼1𝛼+𝑥(𝑡)1𝛼𝑥𝑡31𝛼+(1𝛼)2𝛼𝑡𝑡3[]𝜏(𝑡)𝛼𝑝(𝑠)𝑑𝑠0(2.26) which clearly implies that 𝑡3𝜃𝑘<𝑡𝑎𝑘+(1𝛼)2𝛼𝑡𝑡3[]𝜏(𝑡)𝛼𝑥𝑝(𝑠)𝑑𝑠𝑡31𝛼,(2.27) where 𝑎𝑘=𝑥𝜃𝑘1𝛼𝑞11𝑘𝑥𝜏𝜃𝑘𝛼𝑥𝜃𝑘1𝛼.(2.28) Since 1(1𝑢)1𝛼(1𝛼)𝑢 for 𝑢(0,) and 0<𝛼<1, by taking 𝑞𝑢=𝑘𝑥𝜏𝜃𝑘𝛼𝑥𝜃𝑘,(2.29) we see from (2.28) that 𝑎𝑘𝑞(1𝛼)𝑘𝑥𝜏𝜃𝑘𝛼𝑥𝜃𝑘𝛼.(2.30) But, (2.24) gives 𝑥𝜏𝜃𝑘𝜏𝜃𝑘2𝑥𝜏𝜃𝑘𝜏𝜃𝑘2𝑥𝜃𝑘,(2.31) and hence 𝑎𝑘(1𝛼)2𝛼𝜏𝜃𝑘𝛼𝑞𝑘.(2.32)
Finally, (2.27) and (2.32) result in𝑡3[]𝜏(𝑡)𝛼𝑝(𝑡)𝑑𝑡+𝑡3<𝜃𝑘<𝜏𝜃𝑘𝛼𝑞𝑘<,(2.33) which contradicts (2.20). The proof is complete.

Example 2.3. Consider the impulsive delay differential equation 𝑥(𝑡)+(𝑡1)2||||𝑥(𝑡1)1/2𝑥(𝑡1)=0,𝑡𝑘,Δ𝑥||(𝑡)𝑡=𝑘+(𝑘1)1||||𝑥(𝑘1)1/2||𝑥(𝑘1)=0,Δ𝑥(𝑡)𝑡=𝑘=0,(2.34) where 𝑡2 and 𝑖2.
We see that 𝜏(𝑡)=𝑡1, 𝛼=1/2, 𝑝(𝑡)=(𝑡1)2, and 𝑞𝑘=(𝑘1)1, 𝜃𝑘=𝑘. Since(𝑡1)3/2𝑑𝑡+(𝑘1)1/2=,(2.35) applying Theorem 2.2 we conclude that every solution of (2.34) is oscillatory.
We note that if the equation is not subject to any impulse condition, then, since(𝑡1)5/2𝑑𝑡<,(2.36) the equation 𝑥(𝑡)+(𝑡1)2||||𝑥(𝑡1)1/2𝑥(𝑡1)=0(2.37) has a nonoscillatory solution by Theorem 1.1.

Let us now consider (1.1) when 𝜏(𝑡)=𝑡. That is,𝑥+𝑝(𝑡)|𝑥|𝛼1𝑥=0,𝑡𝜃𝑘,Δ𝑥||𝑡=𝜃𝑘+𝑞𝑘|𝑥|𝛼1||𝑥=0,Δ𝑥𝑡=𝜃𝑘=0,(2.38) where 0<𝛼<1 and 𝑝𝑞𝑘 are given by (a) and (b).

The following theorem is an extension of Theorem 1.2. Note that no sign condition is imposed on 𝑝(𝑡) and {𝑞𝑘}.

Theorem 2.4. If 𝑡𝛽𝑝(𝑡)𝑑𝑡+𝜃𝛽𝑘𝑞𝑘=(2.39) for some 𝛽[0,𝛼], then every solution of (2.38) is oscillatory.

Proof. Assume on the contrary that (2.38) has a nonoscillatory solution 𝑥(𝑡) such that 𝑥(𝑡)>0 for all 𝑡𝑡0 for some 𝑡00. The proof is similar when 𝑥(𝑡) is eventually negative. We set 𝑡𝑤(𝑡)=1𝑥(𝑡)1𝛼,𝑡𝑡0.(2.40) It is not difficult to see that 𝑤(𝑡)=(𝛼1)𝑡𝛼2[]𝑥(𝑡)1𝛼+(1𝛼)𝑡𝛼1[]𝑥(𝑡)𝛼𝑥(𝑡),𝑡𝜃𝑘,(2.41) and hence ||Δ𝑤𝑡=𝜃𝑘=(1𝛼)𝑞𝑘𝜃𝑘𝛼1.(2.42) From (2.41), we have 𝑡𝛽1𝛼𝑡2𝑤(𝑡)=(1𝛼)𝑡𝛽𝑥(𝑡)𝑥𝛼(𝑡)𝛼(1𝛼)𝑡𝛽2𝑥𝛼1(𝑡)𝑡𝑥(𝑡)𝑥(𝑡)2,(2.43) and so 𝑡𝛽1𝛼𝑡2𝑤(𝑡)(1𝛼)𝑡𝛽𝑝(𝑡),𝑡𝜃𝑘.(2.44) In view of (2.42), by a straightforward integration of (2.44), we have 𝑡𝑡0𝑠𝛽1𝛼𝑠2𝑤(𝑠)𝑑𝑠=𝑠𝛽1𝛼𝑠2𝑤||(𝑠)𝑡𝑡0𝑡0𝜃𝑘<𝑡Δ𝑡𝛽𝛼+1𝑤(||𝑡)𝑡=𝜃𝑘𝑡𝑡0(𝛽1𝛼)𝑠𝛽𝛼𝑤(𝑠)𝑑𝑠=𝑡𝛽𝛼+1𝑤(𝑡)𝑡0𝛽𝛼+1𝑤𝑡0𝑡0𝜃𝑘<𝑡(1𝛼)𝑞𝑘𝜃𝛽𝑘𝑠(𝛽𝛼1)𝛽𝛼𝑤||(𝑠)𝑡𝑡0+(𝛽𝛼)(𝛽𝛼1)𝑡𝑡0𝑠𝛽1𝛼𝑤(𝑠)𝑑𝑠,(2.45) which combined with (2.44) leads to 𝑡𝛽𝛼+1𝑤(𝑡)𝑡0𝛽𝛼+1𝑡𝑤0(𝛽𝛼+1)𝑡0𝛽𝛼𝑤𝑡0+(1𝛼)𝑡0𝜃𝑘<𝑡𝜃𝛽𝑘𝑞𝑘+𝑡𝑡0𝑠𝛽.𝑝(𝑠)𝑑𝑠(2.46) Finally, by using (2.39) in the last inequality, we see that there is a 𝑡1>𝑡0 such that 𝑤(𝑡)𝑡𝛼𝛽1,𝑡𝑡1,(2.47) which, however, implies that 𝑤(𝑡) as 𝑡, a contradiction with 𝑥(𝑡)>0. The proof is complete.

Example 2.5. Consider the impulsive differential equation 𝑥+𝑡7/3|𝑥|1/2𝑥=0,𝑡𝑘,Δ𝑥||𝑡=𝑘+𝑘1/6|𝑥|1/2||𝑥=0,Δ𝑥𝑡=𝑘=0,(2.48) where 𝑡1 and 𝑖1.
We have that 𝑝(𝑡)=𝑡7/3, 𝛼=1/2, and 𝑞𝑘=𝑘1/6, 𝜃𝑘=𝑘. Taking 𝛽=1/3 we see from (2.38) that𝑡2𝑑𝑡+𝑘1/3=.(2.49) Since the conditions of Theorem 2.4 are satisfied, every solution of (2.48) is oscillatory.
Note that if the impulses are absent, then, since𝑡2𝑑𝑡<,(2.50) the equation 𝑥+𝑡7/3|𝑥|1/2𝑥=0(2.51) is oscillatory by Theorem 1.2.

Acknowledgment

This work was partially supported by METU-B AP (project no: 01-01-2011-003).