Skip to content
BY-NC-ND 4.0 license Open Access Published by De Gruyter Open Access May 10, 2018

Oscillation and non-oscillation of half-linear differential equations with coeffcients determined by functions having mean values

  • Petr Hasil and Michal Veselý EMAIL logo
From the journal Open Mathematics

Abstract

The paper belongs to the qualitative theory of half-linear equations which are located between linear and non-linear equations and, at the same time, between ordinary and partial differential equations. We analyse the oscillation and non-oscillation of second-order half-linear differential equations whose coefficients are given by the products of functions having mean values and power functions. We prove that the studied very general equations are conditionally oscillatory. In addition, we find the critical oscillation constant.

MSC 2010: 34C10; 34C15

1 Introduction

The aim of this paper is to contribute to the rapidly developing theory of conditionally oscillatory equations. The topic of our research belongs to the qualitative theory concerning the oscillatory behaviour of the half-linear differential equation

[R(t)Φ(x)]+S(t)Φ(x)=0,Φ(x)=|x|p1sgn x,p>1,(1)

where coefficients R > 0, S are continuous functions. Function Φ is the so-called one dimensional p-Laplacian which connects half-linear equations with partial differential equations. Of course, some results obtained for equations of type (1) can be transferred or generalized to (elliptic) PDEs (see, e.g., the last section of [1]).

We recall some basic facts about the treated topic, a short historical background, and the motivation of our research. First of all, we point out that one of the biggest disadvantages of the research in the field of half-linear equations is the lack of the additivity of the solution space (it is the reason for the used nomenclature). Nevertheless, Sturm’s separation and comparison theorems remain valid (see, e.g., [2, 3]). Therefore, we can classify half-linear equations as oscillatory and non-oscillatory as well as linear equations. More precisely, Sturm’s separation theorem guarantees that if one non-zero solution is oscillatory (i.e., its zero points tend to infinity), then every solution is oscillatory and Eq. (1) is called oscillatory. There exist important equations whose oscillatory properties can be determined simply by measuring (in some sense) their coefficients. Searching for such equations is based on the study of the so-called conditional oscillation.

On behalf of clarity, we consider the equations of the form

[R(t)Φ(x)]+γC(t)Φ(x)=0,γR.(2)

We say that Eq. (2) is conditionally oscillatory if there exists a constant Γ such that Eq. (2) is oscillatory for γ > Γ and non-oscillatory for γ < Γ. The constant Γ is usually called the critical oscillation constant of Eq. (2). Note that such a critical oscillation constant depends on coefficients R > 0 and C and it is non-negative—this observation comes directly from Sturm’s comparison theorem (see Theorem 2.3 in Section 2 below). The conditionally oscillatory equations are very useful as testing and comparing equations.

The first conditionally oscillatory half-linear equation was found in [4], where the critical oscillation constant Γ = (p − 1)p/pp was revealed for the equation

[Φ(x)]+γtpΦ(x)=0.

Then, motivated by results about linear equations (see [5]), the equation

[R(t)Φ(x)]+γD(t)tpΦ(x)=0(3)

with positive periodic functions R, D was studied in [6] and it turned out that Eq. (3) is conditionally oscillatory as well. Later, it was proved that the critical oscillation constant can be found even in the case of Eq. (3) with coefficients having mean values (see [1]).

As a follow-up of the above mentioned results, a natural question arose, whether it is possible to remove tp from the potential of Eq. (3) and to preserve the conditionally oscillatory behaviour of the considered equation. Concerning this research direction, we mention papers [7, 8, 9] which are, together with paper [1], the main motivations of the results presented here. We should emphasize that, although the first motivation comes from the linear case, our research follows a path in half-linear equations and the linear case remains a special case for p = 2. Hence, our result is new even for linear equations which is demonstrated in Corollary 3.5 at the end of this paper.

To conclude this introductory section, we mention some books and papers that are connected to the treated topic. The theory of half-linear equations is thoroughly described in the already mentioned books [2, 3]. The direction of research which leads to perturbed equations is treated in many papers. We mention at least papers [10, 11, 12, 13, 14, 15]. Half-linear equations are close to non-linear equations, where the p-Laplacian is replaced by more general functions. For results concerning such a type of equations, we refer to [16, 17, 18, 19, 20, 21]. We should not forget to mention the discrete counterparts of results mentioned in this section. The theory of conditionally oscillatory difference equations is not as developed as the continuous one. Nevertheless, some results are already available in the literature (see [22, 23]). Some basic results are known even for dynamic equations on time scales which connect and generalize the continuous and discrete case. For such results, see [24, 25].

The rest of this paper is divided into two sections. The next section contains the description of the used transformation, where we derive the so-called adapted Riccati equation and we state preparatory lemmas. The last section is devoted to our results. We also mention a corollary concerning linear equations (to demonstrate the novelty of the main result and its impact to linear equations) and an illustrative simple example.

2 Preparations

First of all, we recall the definition of mean values for continuous functions.

Definition 2.1

Let a continuous function f : [T, ∞) → ℝ be such that the limit

f¯:=lima1att+af(s)ds

is finite and exists uniformly with respect to t ∈ [T, ∞). The number f is called the mean value of f.

Obviously, the mean value of any β-periodic continuous function f is

f¯=1βττ+βf(s)ds,(4)

where τ is arbitrary.

To prove our results, we use the generalized Riccati technique which is described below. We consider the second-order half-linear differential equation

tαr(t)Φ(x)+tαps(t)Φ(x)=0,Φ(x)=|x|p1sgn x,p>1,(5)

where α ≤ 0 and r, s are continuous functions such that the mean values of functions r1/(1−p) and s exist, the mean value s is positive, and

0<r:=inf{r(t),tR}r+:=sup{r(t),tR}<.(6)

We denote by q the number conjugated with the given number p > 1, i.e.,

p+q=pq.(7)

Immediately, we obtain the inverse function to Φ in the form Φ−1(x) = |x|q − 1sgn x.

The basis of our method is the transformation to the Riccati half-linear equation which can be introduced as follows. We consider a non-zero solution x of Eq. (5) and we define

w(t)=tαr(t)Φx(t)x(t).(8)

Considering Eq. (5), the differentiation of (8) leads to the Riccati half-linear equation

w(t)+tαps(t)+(p1)tαr(t)1q|w(t)|q=0.(9)

The form of Eq. (9) is not sufficient enough for our method. Hence, we apply the transformation ζ(t) = −tpα − 1w(t) which leads to the equation

ζ(t)=1t(pα1)ζ(t)+s(t)+(p1)r1q(t)|ζ(t)|q.(10)

Eq. (10) is called the adapted generalized Riccati equation for the consistency with similar cases in the literature.

Further, in this section, we formulate auxiliary results that will be needed in the following section within the proof of Theorem 3.1 below. We begin with properties of functions having mean values.

Lemma 2.2

Let a continuous function f : [T, ∞) ⊂ (0, ∞) → ℝ have mean value f. For an arbitrarily given a > 0, there exists a constant M(f) > 0 such that

tt+bf(τ)dτ<M(f)(11)

and

tt+bf(τ)τdτ<M(f)t(12)

for all tT, b ∈ [0, a].

Proof

The statement of the lemma follows from the beginning of the proof of [1, Theorem 8] (see directly inequalities (48) and (51) in [1]). □

Next, we recall the well-known Sturm half-linear (also called the Sturm–Picone) comparison theorem.

Theorem 2.3

Let r̂, , ŝ, s̃ be continuous functions satisfying r̂ (t) ≥ (t) > 0 and s̃(t) ≥ ŝ(t) for all sufficiently large t. Consider the pair of equations

r^(t)Φ(x)+s^(t)Φ(x)=0,(13)
r~(t)Φ(x)+s~(t)Φ(x)=0.(14)

  1. If Eq. (14)is non-oscillatory, then Eq. (13)is non-oscillatory.

  2. If Eq. (13)is oscillatory, then Eq. (14)is oscillatory.

Proof

See, e.g., [3, Theorem 1.2.4]. □

Now, we formulate the condition which is equivalent to the non-oscillation of Eq. (5).

Theorem 2.4

Eq. (5)is non-oscillatory if and only if there exists a ∈ ℝ and a solution w: [a, ∞) → ℝ of Eq. (9)satisfying either

w(t)=tταps(τ)dτ+(p1)tταr(τ)1q|w(τ)|qdτ0(15)

or

w(t)=tταps(τ)dτ+(p1)tταr(τ)1q|w(τ)|qdτ0(16)

for all ta.

Proof

See, e.g., [3, Theorems 2.2.4 and 2.2.5], where it suffices to consider that the used divergence of ∫ (τα r(τ))1 − q dτ follows from α ≤ 0 and from (6) and the used convergence of ∫ταps(τ) dτ is proved as (20) in [1]. □

The upcoming lemma describes the connection between the behaviour of solutions of Eq. (5) and the adapted Riccati equation (10).

Lemma 2.5

If Eq. (5)is non-oscillatory, then there exists a solution ζ of the associated adapted generalized Riccati equation(10)such that ζ(t) ≤ 0 for all large t ∈ ℝ.

Proof

We apply Theorem 2.4. From the positivity of s (see also [1]), it is seen that (16) cannot be valid for all large t, i.e., we obtain (15). Hence, there exists a non-negative solution w of Eq. (9) on some interval [T, ∞), i.e., there exists a non-positive solution ζ of Eq. (10) on [T, ∞). □

The last lemma contains the opposite implication to the one in Lemma 2.5.

Lemma 2.6

If there exists a solution ζ of Eq. (10)for all large t ∈ ℝ, then Eq. (5)is non-oscillatory.

Proof

The statement of the lemma follows directly from the half-linear version of the Reid roundabout theorem (see [3, Theorem 1.2.2] and also [1, Lemma 4]). □

3 Results

In this section, we formulate and prove our results. For reader’s convenience, we slightly modify Eq. (5) as follows. We consider the equation

tαrpq(t)Φ(x)+tαps(t)Φ(x)=0,Φ(x)=|x|p1sgn x,p>1,(17)

where t ∈ ℝ is sufficiently large, q is the number conjugated with p (see (7)), α ≤ 0, and r, s are continuous functions having mean values such that (6) is satisfied. Note that s can be non-positive. The only difference between Eq. (5) and Eq. (17) is the power −p/q of r. The reason for this modification is purely technical (it leads to more transparent calculations below) and it does not mean any restriction (consider (6)).

Theorem 3.1

Let us consider Eq. (17).

  1. If

    ppα1ps¯r¯p1>1,(18)

    then Eq. (17)is oscillatory.

  2. If

    ppα1ps¯r¯p1<1,(19)

    then Eq. (17)is non-oscillatory.

Proof

In the both parts of the proof, we will consider such a number a > 2 for which (see Definition 2.1 together with (18) and (19))

ppα1p1att+as(τ)dτ1att+ar(τ)dτp1>1+ε(20)

or

ppα1p1att+as(τ)dτ1att+ar(τ)dτp1<1ε(21)

for all considered t and for some ε ∈ (0, 1). We can rewrite (20) and (21) into the following forms

1att+as(τ)dτ>1+εpα1pp1att+ar(τ)dτ1p(22)

and

1att+as(τ)dτ<1εpα1pp1att+ar(τ)dτ1p.(23)

Using (6), from (22) and (23), we obtain the existence of L > 0, for which

1att+as(τ)dτpα1pp1att+ar(τ)dτ1p>L(24)

and

1att+as(τ)dτpα1pp1att+ar(τ)dτ1p<L(25)

for all considered t. Indeed, one can put

L:=εpα1ppr+1p

in the both cases.

We will consider the associated adapted generalized Riccati equation in the form of Eq. (10) which corresponds Eq. (17), i.e., the equation

ζ(t)=1t(pα1)ζ(t)+s(t)+(p1)r(t)|ζ(t)|q.(26)

At first, we show that any solution ζ of Eq. (26) defined for tT is bounded from below, i.e., we show that there exists K > 1 satisfying

ζ(t)>K,tT.(27)

On the contrary, let us assume that

lim inftζ(t)=(28)

or

lim inftT0ζ(t)=(29)

for some T0 ∈ (T, ∞). For given a and function s, let us consider M(s) from Lemma 2.2. Let P > 0 be an arbitrary number such that

(p1)ryq(pα1)y>M(s),yP.(30)

In particular, considering inftTζ (t) = −∞, the continuity of ζ implies the existence of an interval [t1, t2] such that ζ (t1) ≤ − P, ζ(t) < −P for all t ∈ (t1, t2], and t2t1 ∈ (0, a]. Without loss of generality, we consider T > 2. Using (12) in Lemma 2.2, the form of Eq. (26), and (30), we have

ζ(t2)ζ(t1)=t1t2ζ(τ)dτ=t1t2(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|qτdτ>t1t2M(s)+s(τ)τdτt1t2M(s)τdτt1t2s(τ)τdτ>0M(s)t1M(s)T>M(s).(31)

This inequality proves that (29) cannot be valid for any T0 ∈ ℝ (consider that a > 2 is given). Therefore, there exist arbitrarily long intervals, where ζ(t) ≤ −P. Let I = [t3, t4] be such an interval whose length is at least 2 (i.e., t4t3 ≥ 2) and t4t3a. As in (31) (consider that t3 > 2), we obtain

ζ(t4)ζ(t3)=t3t4ζ(τ)dτ>t3t4M(s)τdτt3t4s(τ)τdτM(s)logt4t3M(s)t3M(s)logt3+2t31t3>0.(32)

Of course, (32) means that ζ(t4) > ζ(t3). In fact, (31) and (32) guarantee that

lim inftζ(t)>,

which contradicts (28). Hence, (27) is valid.

In the both parts of the proof, we will also apply the estimation

tt+as(τ)tdτtt+as(τ)τdτat2M(s)(33)

for all large t and M(s) from Lemma 2.2. We use the mean value theorem of the integral calculus to get this estimation. More precisely, considering t ∈ [t1, t2], where t1 is sufficiently large, since s is integrable and x(t) = t− 1 is monotone for t ∈ [t1, t2], there exists t3 ∈ [t1, t2] such that

t1t2s(τ)τdτ=1t1t1t3s(τ)dτ+1t2t3t2s(τ)dτ.(34)

Immediately, from (34), we obtain (see (11) in Lemma 2.2)

tt+a1t1τs(τ)dτ=tt+as(τ)tdτ1ttt+bs(τ)dτ1t+at+bt+as(τ)dτ=1tt+bt+as(τ)dτ1t+at+bt+as(τ)dτ=at(t+a)t+bt+as(τ)dτat2M(s),(35)

where b ∈ [0, a]. It is seen that (35) gives (33).

Part (I). Evidently (see (18)), s > 0. By contradiction, let us suppose that Eq. (17) is non-oscillatory. From Lemma 2.5, we know that there exists a non-positive solution ζ of Eq. (26) on some interval [T, ∞). For this solution ζ, we introduce the averaging function ζave by

ζave(t):=1att+aζ(τ)dτ,tT.(36)

We know that (see (27))

ζave(t)(K,0],tT.(37)

For t > T (see Eq. (26)), we have

ζ!ave(t)=1att+aζ(τ)dτ=1att+a1τ(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|qdτ.(38)

For t > T, we also have (see (27) and (33))

tt+a1τ(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|qdτ1ttt+a(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|qdτtt+a1t1τ(pα1)K+(p1)r+Kqdτ+tt+as(τ)tdτtt+as(τ)τdτa2t(t+a)(pα1)K+(p1)r+Kq+at2M(s)<Nat2,(39)

where

N:=a(pα1)K+a(p1)r+Kq+M(s).(40)

For t > T, we obtain (see (38), (39), and (40))

ζave(t)1attt+a(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|qNtdτ.(41)

If we put

X(t):=(pα1)pptpatt+ar(τ)dτpq,Y(t):=|ζave(t)|qqtpatt+ar(τ)dτ(42)

for t > T, then we have (see (41))

ζave(t)1attt+a(pα1)ζ(τ)dτ+X(t)+Y(t)Nt2+1attt+as(τ)dτX(t)+1attt+a(p1)r(τ)|ζ(τ)|qdτY(t),t>T.(43)

Taking into account (43), for large t, we will show the inequalities

Nt2L3t,(44)
1attt+a(pα1)ζ(τ)dτ+X(t)+Y(t)0,(45)
1attt+as(τ)dτX(t)Lt,(46)
1attt+a(p1)r(τ)|ζ(τ)|qdτY(t)L3t.(47)

The first inequality (44) is valid for all t ≥ 3N/L. Hence, we can approach to (45). It holds (see (36) and (42))

1attt+a(pα1)ζ(τ)dτ+X(t)+Y(t)=1t(pα1)ζave(t)+(pα1)pppatt+ar(τ)dτpq+|ζave(t)|qqpatt+ar(τ)dτ.(48)

We recall the well-known Young inequality which says that

App+BqqAB0(49)

holds for all non-negative numbers A, B. We take A = (p t X(t))1/p and B = (qtY(t))1/q. Hence,

App=tX(t)=(pα1)pppatt+ar(τ)dτpq,Bqq=tY(t)=|ζave(t)|qqpatt+ar(τ)dτ,

and (see (37))

AB=ptX(t)1pqtY(t)1q=(pα1)patt+ar(τ)dτ1q|ζave(t)|patt+ar(τ)dτ1q=(pα1)ζave(t).

Finally, considering (48) and (49), we have

1attt+a(pα1)ζ(τ)dτ+X(t)+Y(t)=1tApp+BqqAB0,

which proves (45).

Next, (46) is valid. Indeed, we have (see (24) and (42))

1attt+as(τ)dτX(t)=1t1att+as(τ)dτ(pα1)pppatt+ar(τ)dτpq=1t1att+as(τ)dτpα1pp1att+ar(τ)dτ1p>Lt

for all considered t.

To prove (47), we use the form of Eq. (26) together with (6), (12) from Lemma 2.2, and with (27) which immediately give

t+it+jζ(τ)dτt+it+j1τ(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|qdτt+it+j1τ(pα1)K+(p1)r+Kqdτ+t+it+js(τ)τdτat(pα1)K+(p1)r+Kq+M(s)t+iQt(50)

for t > T and i, j ∈ [0, a], ij, where

Q:=a(pα1)K+(p1)r+Kq+M(s).

Hence, we have

ζ(t+j)ζ(t+i)Qt,t>T,i,j[0,a],

which implies (see (36))

ζ(τ)ζave(t)Qt(51)

for all t > T and τ ∈ [t, t+a].

Further, since the function x(t) = |t|q is continuously differentiable on [−K, 0], there exists C > 0 for which

|y|q|z|qC|yz|,y,z[K,0].(52)

Thus, we have (see (6), (7), (27), (37), (42), (51), and (52))

1attt+a(p1)r(τ)|ζ(τ)|qdτY(t)=1t1att+a(p1)r(τ)|ζ(τ)|qdτ|ζave(t)|qqpatt+ar(τ)dτp1attt+a||ζ(τ)|q|ζave(t)|q|r(τ)dτ(p1)r+attt+a||ζ(τ)|q|ζave(t)|q|dτ(p1)r+attt+aC|ζ(τ)ζave(t)|dτ(p1)r+CQt2

for t > T which gives (47) for all sufficiently large t.

Altogether, (43) together with (44), (45), (46), and (47) guarantee

ζave(t)0L3t+LtL3t=L3t(53)

for all large t. From (53) it follows that limt → ∞ζave (t) = ∞. In particular (see (36)), ζ is positive at least in one point which is a contradiction. The proof of part (I) is complete.

Part (II). Without loss of generality, we can assume that s > 0. Indeed, for s ≤ 0, it suffices to replace function s by function s + k for a constant k > 0 such that s + k > 0 and

ppα1ps¯+kr¯p1<1

and to use Theorem 2.3.

Let t0 be a sufficiently large number. We denote (see (6))

Z:=prpα11p.

Let ζ be the solution of the adapted generalized Riccati equation (26) satisfying

ζ(t0)=p(pα1)at0t0+ar(τ)dτ1p2Z,0.(54)

Based on Lemma 2.6, it suffices to prove that this solution exists for all t ∈ [t0, ∞). Let [t0, T) be the maximal interval, where the solution ζ exists. Note that T ∈ (t0, ∞) ∪ {∞}. In fact, considering the continuity and the boundedness from below of ζ (see (27)), it suffices to show that ζ(t) < 0 for all t ∈ [t0, T).

Let us consider an interval J := [t0, t1) such that ζ(t) ∈ (−2 Z, 0) for tJ. For any t2, t3J ∩ [t0, t0 + a], t2t3, we have (see (6) and (12) in Lemma 2.2, cf. (50))

ζ(t3)ζ(t2)=t2t3ζ(τ)dτ=t2t31τ(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|qdτt2t31τ(pα1)2Z+(p1)r+2Zqdτ+t2t3s(τ)τdτ1t0t0t0+a(pα1)2Z+(p1)r+2Zqdτ+M(s)t2Q~t0,(55)

where

Q~:=a(pα1)2Z+(p1)r+2Zq+M(s).

Since t0 is given as a sufficiently large number, from (55), we see that

ζ(t)(2Z,0),t[t0,t0+a].(56)

In addition, (55) means that

|ζ(t0)ζ(τ)|<Q~t0,τ[t0,t0+a].(57)

Similarly as in the first part of the proof, we introduce

ζave(t):=1att+aζ(τ)dτ(58)

for t from a neighbourhood of t0. It holds (see (56), (57), and (58))

ζave(t0)(2Z,0),|ζave(t0)ζ(t0)|<Q~t0.(59)

We have (see Eq. (26), cf. (38))

ζave(t0)=1at0t0+aζ(τ)dτ=1at0t0+a1τ[(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|q]dτ.(60)

In addition, as in the first part of the proof (see (39)), we have

|t0t0+a1τ[(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|q]dτ1t0t0t0+a[(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|q]dτ|<Nat02,(61)

where N is defined in (40) (we can also put K = 2Z).

Hence, from (60) and (61), we obtain (cf. (41))

ζave(t0)1at0t0t0+a[(pα1)ζ(τ)+s(τ)+(p1)r(τ)|ζ(τ)|q+Nt0]dτ.(62)

If we put (cf. (42))

X(t0):=(pα1)pp(pat0t0+ar(τ)dτ)1p,(63)
Y(t0):=(pα1)pq(pat0t0+ar(τ)dτ)1p,(64)

then we have (see (62))

ζave(t0)1at0t0t0+as(τ)X(t0)dτ+1at0t0t0+a(p1)r(τ)|ζ(τ)|qY(t0)dτ+1at0t0t0+a(pα1)ζ(τ)dτ+X(t0)t0+Y(t0)t0+Nt02.(65)

The aim of our process is to prove that ζave(t0)<0. To obtain this inequality, it suffices to show (see (65))

1at0t0+as(τ)X(t0)dτL,(66)
|1at0t0+a(p1)r(τ)|ζ(τ)|qY(t0)dτ|L4,(67)
1at0t0+a(pα1)ζ(τ)dτ+X(t0)+Y(t0)L4,(68)
Nt0L4.(69)

Evidently, (69) is valid, because t0 is sufficiently large. We show that (66) is valid. We have (see (25), (63))

1at0t0+as(τ)X(t0)dτ=1at0t0+as(τ)dτ(pα1)pp(pat0t0+ar(τ)dτ)1p=1at0t0+as(τ)dτ(pα1p)p(1at0t0+ar(τ)dτ)1p<L,

i.e., we obtain (66).

Now we prove (67). Let D > 0 be such that (cf. (52))

||y|q|z|q|D|yz|,y,z[2Z,0].(70)

Using (6), (7), (54), (56), (57), (64), and (70), we have

|1at0t0+a(p1)r(τ)|ζ(τ)|qY(t0)dτ|=|1at0t0+a(p1)r(τ)|ζ(τ)|qdτ(pα1)pq(pat0t0+ar(τ)dτ)1p|=|1at0t0+a(p1)r(τ)|ζ(τ)|qdτpq|ζ(t0)|q(1at0t0+ar(τ)dτ)|p1at0t0+ar(τ)||ζ(τ)|q|ζ(t0)q|dτ(p1)r+at0t0+a||ζ(τ)|q|ζ(t0)q|dτ(p1)r+at0t0+aD|ζ(τ)ζ(t0)|dτ(p1)r+at0t0+aDQ~t0dτ=DQ~(p1)r+t0.(71)

For a sufficiently large number t0, inequality (67) follows from (71).

It remains to prove (68). We have (see (58), (63), and (64))

1at0t0+a(pα1)ζ(τ)dτ+X(t0)+Y(t0)=(pα1)ζave(t0)+(pα1)pp(pat0t0+ar(τ)dτ)1p+(pα1)pq(pat0t0+ar(τ)dτ)1p.(72)

Let us assume that (see (54))

ζave(t0)=ζ(t0)=(p(pα1)at0t0+ar(τ)dτ)1p.(73)

Then, (72) gives

1at0t0+a(pα1)ζ(τ)dτ+X(t0)+Y(t0)=(pα1)(p(pα1)at0t0+ar(τ)dτ)1p+(pα1)pp(pat0t0+ar(τ)dτ)1p+(pα1)pq(pat0t0+ar(τ)dτ)1p=(pα1)p(pat0t0+ar(τ)dτ)1p(1+1p+1q)=0.(74)

Using (59) (consider (73)), one can see that (74) gives (68) for large t0.

Finally, applying (66), (67), (68), and (69) in (65), we have

ζave(t0)1t0(L+L4+L4+L4)=L4t0<0,

i.e., we have (see (58))

ζave(t0)=ζ(t0+a)ζ(t0)a<0,

i.e., ζ (t0 + a) < ζ (t0). Since we can replace t0 by an arbitrary number t > t0 in the process above, we obtain ζ (t + a) < ζ(t) and ζ(τ) < 0 for all τ ∈ [t, t + a] if

ζ(t)=(p(pα1)att+ar(τ)dτ)1p

for some t ∈ (t0, ∞). Considering the fact that a can be arbitrarily large (see Definition 2.1 and (6)) together with (55) and ζave(t)<0, we obtain that ζ (t) < 0 for all t > t0. The proof is complete.∈□

Remark 3.2

Now we describe the connection of Theorem 3.1 and our motivation. We repeat that our basic motivation comes from papers[1, 7, 8, 9]. To the best of our knowledge, the strongest known results about non-perturbed conditionally oscillatory half-linear differential equations are proved just in those articles. In[1, 8, 9], only the caseα = 0 is analysed. Note that, in [8], the considered type of equations differs from Eq. (17). In [7], the general form of Eq. (17)is treated. The process in [7] (and also in [8, 9]) is based on the modified Prüfer transformation. Hence, it is entirely different from the method used in this paper which enables us to cover new types of equations. The coefficients of equations considered in [7] has to be restricted in a certain sense. This restriction is removed in the presented results.

Remark 3.3

Theorem 3.1 does not cover the case when

(ppα1)ps¯r¯p1=1.(75)

It is known (see any of papers [10, 11, 12, 13, 14, 15]) that this case is not generally solvable. More precisely, for

r(t)1,s(t)(pα1p)p,

Eq. (17)is non-oscillatory (see [7]) and, at the same time, there exist continuous functions r, s satisfying(75)for which Eq. (17)is oscillatory (forα = 0, see again papers[10, 11, 12, 13, 14, 15]). We conjecture that Eq. (17)is non-oscillatory in the situation given by(75), where the coefficients r, s are periodic functions (see also(4)). Our conjecture is based on results of[26, 27], but it remains an open problem.

To illustrate Theorem 3.1, we mention the following example of a simple equation whose oscillatory behaviour does not follow from any previously known result (also for p = 2, i.e., in the case of linear equations). In fact, as far as we know, such equations with general α are not studied in the literature.

Example 3.4

Let c, d ≠ 0 be arbitrarily given. We define the functions : [1, ∞) → ℝ by the formula

s(t):={c+4nnd(tn),t[n,n+14n);c+4nnd(n+24nt),t[n+14n,n+34n);c+4nnd(tn44n),t[n+34n,n+44n);c+4nnd(tn4(n1)4n),t[n+4(n1)4n,n+4(n1)+14n);c+4nnd(n+4(n1)+24nt),t[n+4(n1)+14n,n+4(n1)+34n);c+4nnd(tn1),t[n+4(n1)+34n,n+1),

wheren ∈ ℕ. One can easily show that the mean value of this function exists and thats = c. We consider the equation

[tαΦ(x)]+tαps(t)Φ(x)=0,(76)

wherep > 1 andα ≤ 0 are arbitrary. From Theorem 3.1, we know that Eq. (76)is oscillatory forc > (pα − 1)p/ppand non-oscillatory forc < (pα − 1)p/pp.

We repeat that we obtain new results even for linear equations. With regard to the importance of this fact, we formulate the corresponding consequence of Theorem 3.1 as the corollary below.

Corollary 3.5

Let us consider the equation

[xtαr(t)]+s(t)tα+2x=0,(77)

wheret ∈ ℝ is sufficiently large, α ≥ 0, and r, sare continuous functions having mean valuesr, ssuch that(6)is satisfied.

  1. If 4 rs > (1 + α)2, then Eq. (77)is oscillatory.

  2. If 4 rs < (1 + α)2, then Eq. (77)is non-oscillatory.

Proof

The corollary follows directly from Theorem 3.1. □

Acknowledgement

Second author was supported by Czech Science Foundation under Grant GA17-03224S.

References

[1] Hasil P., Mařík R., Veselý M., Conditional oscillation of half-linear differential equations with coefficients having mean values, Abstract Appl. Anal., 2014, article ID 258159, 1–14. 10.1155/2014/258159Search in Google Scholar

[2] Agarwal R.P., Grace A.R., O’Regan D., Oscillation Theory for Second Order Linear, Half-linear, Superlinear and Sublinear Dynamic Equations, Kluwer Academic Publishers, Dordrecht, 2002. 10.1007/978-94-017-2515-6Search in Google Scholar

[3] Došlý O., Řehák P., Half-linear Differential Equations, Elsevier, Amsterdam, 2005.10.1155/JIA.2005.535Search in Google Scholar

[4] Elbert Á., Asymptotic behaviour of autonomous half-linear differential systems on the plane, Studia Sci. Math. Hungar., 1984, 19, No. 2-4, 447–464.Search in Google Scholar

[5] Schmidt K.M., Oscillation of perturbed Hill equation and lower spectrum of radially periodic Schrödinger operators in the plane, Proc. Amer. Math. Soc., 1999, 127, No. 8, 2367–2374.10.1090/S0002-9939-99-05069-8Search in Google Scholar

[6] Hasil P., Conditional oscillation of half-linear differential equations with periodic coefficients, Arch. Math. (Brno), 2008, 44, No. 2, 119–131.Search in Google Scholar

[7] Došlý O., Jaroš J., Veselý M., Generalized Prüfer angle and oscillation of half-linear differential equations, Appl. Math. Lett., 2017, 64, No. 2, 34–41. 10.1016/j.aml.2016.08.004Search in Google Scholar

[8] Hasil P., Veselý M., Oscillation constant for modified Euler type half-linear equations, Electron. J. Diff. Equ., 2015, No. 220, 1–14.10.1186/s13662-015-0544-1Search in Google Scholar

[9] Jaroš J., Veselý M., Conditional oscillation of Euler type half-linear differential equations with unbounded coefficients, Studia Sci. Math. Hungar., 2016, 53, No. 1, 22–41. 10.1556/012.2015.1323Search in Google Scholar

[10] Došlý O., Funková H., Euler type half-linear differential equation with periodic coefficients, Abstract Appl. Anal., 2013, article ID 714263, 1–6. 10.1155/2013/714263Search in Google Scholar

[11] Došlý O., Hasil P., Critical oscillation constant for half-linear differential equations with periodic coefficients, Ann. Mat. Pura Appl., 2011, 190, No. 3, 395–408. 10.1007/s10231-010-0155-0Search in Google Scholar

[12] Došlý O., Veselý M., Oscillation and non-oscillation of Euler type half-linear differential equations, J. Math. Anal. Appl., 2015, 429, 602–621. 10.1016/j.jmaa.2015.04.030Search in Google Scholar

[13] Misir A., Mermerkaya B., Critical oscillation constant for Euler type half-linear differential equation having multi-different periodic coefficients, Int. J. Differ. Equ., 2017, article ID 5042421, 1–8. 10.1155/2017/5042421Search in Google Scholar

[14] Misir A., Mermerkaya B., Critical oscillation constant for half linear differential equations which have different periodic coefficients, Gazi Univ. J. Sci., 2016, 29, No. 1, 79–86.10.1155/2017/5042421Search in Google Scholar

[15] Misir A., Mermerkaya B., Oscillation and nonoscillation of half-linear Euler type differential equations with different periodic coefficients, Open Math., 2017, 15, No. 1, 548–561. 10.1515/math-2017-0046Search in Google Scholar

[16] Došlý O., Yamaoka N., Oscillation constants for second-order ordinary differential equations related to elliptic equations with p-Laplacian, Nonlinear Anal., 2015, 113, 115–136. 10.1016/j.na.2014.09.025Search in Google Scholar

[17] Sugie J., Nonoscillation criteria for second-order nonlinear differential equations with decaying coefficients, Math. Nachr., 2008, 281, No. 11, 1624–1637. 10.1002/mana.200510702Search in Google Scholar

[18] Sugie J., Kita K., Oscillation criteria for second order nonlinear differential equations of Euler type, J. Math. Anal. Appl., 2001, 253, No. 2, 414–439. 10.1006/jmaa.2000.7149Search in Google Scholar

[19] Sugie J., Onitsuka M., A non-oscillation theorem for nonlinear differential equations with p-Laplacian, Proc. Roy. Soc. Edinburgh Sect. A, 2006, 136, No. 3, 633–647. 10.1017/S0308210500005096Search in Google Scholar

[20] Sugie J., Yamaoka N., Growth conditions for oscillation of nonlinear differential equations with p-Laplacian, J. Math. Anal. Appl., 2005, 306, 18–34. 10.1016/j.jmaa.2004.10.009Search in Google Scholar

[21] Sugie J., Yamaoka N., Oscillation of solutions of second-order nonlinear self-adjoint differential equations, J. Math. Anal. Appl., 2004, 291, No. 2, 387–405. 10.1016/j.jmaa.2003.11.027Search in Google Scholar

[22] Hasil P., Veselý M., Oscillation and non-oscillation criteria for linear and half-linear difference equations, J. Math. Anal. Appl., 2017, 452, No. 1, 401–428. 10.1016/j.jmaa.2017.03.012Search in Google Scholar

[23] Hasil P., Veselý M., Oscillation constants for half-linear difference equations with coefficients having mean values, Adv. Differ. Equ., 2015, No. 210, 1–18. 10.1186/s13662-015-0544-1Search in Google Scholar

[24] Hasil P., Vítovec J., Conditional oscillation of half-linear Euler-type dynamic equations on time scales, Electron. J. Qual. Theory Differ. Equ., 2015, No. 6, 1–24.10.14232/ejqtde.2015.1.6Search in Google Scholar

[25] Řehák P., A critical oscillation constant as a variable of time scales for half-linear dynamic equations, Math. Slovaca, 2010, 60 No. 2, 237–256. 10.2478/s12175-010-0009-7Search in Google Scholar

[26] Hasil P., Veselý M., Non-oscillation of half-linear differential equations with periodic coefficients, Electron. J. Qual. Theory Differ. Equ., 2015, No. 1, 1–21.10.14232/ejqtde.2015.1.1Search in Google Scholar

[27] Hasil P., Veselý M., Non-oscillation of perturbed half-linear differential equations with sums of periodic coefficients, Adv. Differ. Equ., 2015, No. 190, 1–17. 10.1186/s13662-015-0533-4Search in Google Scholar

Received: 2017-08-07
Accepted: 2018-04-11
Published Online: 2018-05-10

© 2018 Hasil and Veselý, published by De Gruyter

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 License.

Downloaded on 29.3.2024 from https://www.degruyter.com/document/doi/10.1515/math-2018-0047/html
Scroll to top button