Abstract

Existence of positive solutions for advanced equations with several terms ̇𝑥(𝑡)+𝑚𝑘=1𝑎𝑘(𝑡)𝑥(𝑘(𝑡))=0,𝑘(𝑡)𝑡 is investigated in the following three cases: (a) all coefficients 𝑎𝑘 are positive; (b) all coefficients 𝑎𝑘 are negative; (c) there is an equal number of positive and negative coefficients. Results on asymptotics of nonoscillatory solutions are also presented.

1. Introduction

This paper deals with nonoscillation properties of scalar advanced differential equations. Advanced differential equations appear in several applications, especially as mathematical models in economics; an advanced term may, for example, reflect the dependency on anticipated capital stock [1, 2].

It is not quite clear how to formulate an initial value problem for such equations, and existence and uniqueness of solutions becomes a complicated issue. To study oscillation, we need to assume that there exists a solution of such equation on the halfline. In the beginning of 1980s, sufficient oscillation conditions for first-order linear advanced equations with constant coefficients and deviations of arguments were obtained in [3] and for nonlinear equations in [4]. Later oscillation properties were studied for other advanced and mixed differential equations (see the monograph [5], the papers [612] and references therein). Overall, these publications mostly deal with sufficient oscillation conditions; there are only few results [7, 9, 12] on existence of positive solutions for equations with several advanced terms and variable coefficients, and the general nonoscillation theory is not complete even for first-order linear equations with variable advanced arguments and variable coefficients of the same sign. The present paper partially fills up this gap. We obtain several nonoscillation results for advanced equations using the generalized characteristic inequality [13]. The main method of this paper is based on fixed point theory; thus, we also state the existence of a solution in certain cases.

In the linear case, the best studied models with advanced arguments were the equations of the types ̇𝑥(𝑡)𝑎(𝑡)𝑥((𝑡))+𝑏(𝑡)𝑥(𝑡)=0,̇𝑥(𝑡)𝑎(𝑡)𝑥(𝑡)+𝑏(𝑡)𝑥(𝑔(𝑡))=0,(1.1) where 𝑎(𝑡)0, 𝑏(𝑡)0, (𝑡)𝑡, and 𝑔(𝑡)𝑡.

Let us note that oscillation of higher order linear and nonlinear equations with advanced and mixed arguments was also extensively investigated, starting with [14]; see also the recent papers [1519] and references therein.

For equations with an advanced argument, the results obtained in [20, 21] can be reformulated as Theorems A–C below.

Theorem A (see [20]). If 𝑎, 𝑏, and are equicontinuous on [0,), 𝑎(𝑡)0, 𝑏(𝑡)0, (𝑡)𝑡, and limsup𝑡[(𝑡)𝑡]<, then the advanced equation ̇𝑥(𝑡)+𝑎(𝑡)𝑥((𝑡))+𝑏(𝑡)𝑥(𝑡)=0(1.2) has a nonoscillatory solution.

In the present paper, we extend Theorem A to the case of several deviating arguments and coefficients (Theorem 2.10).

Theorem B (see [20]). If 𝑎, 𝑏, and are equicontinuous on [0,), 𝑎(𝑡)0, 𝑏(𝑡)0, (𝑡)𝑡, limsup𝑡[(𝑡)𝑡]<, and limsup𝑡𝑡(𝑡)𝑎(𝑠)exp𝑠(𝑠)𝑏1(𝜏)𝑑𝜏𝑑𝑠<𝑒,(1.3) then the advanced equation ̇𝑥(𝑡)𝑎(𝑡)𝑥((𝑡))𝑏(𝑡)𝑥(𝑡)=0(1.4) has a nonoscillatory solution.

Corollary 2.3 of the present paper extends Theorem B to the case of several coefficients 𝑎𝑘0 and advanced arguments 𝑘 (generally, 𝑏(𝑡)0); if max𝑘𝑘𝑡𝑚(𝑡)𝑖=1𝑎𝑖1(𝑠)𝑑𝑠𝑒,(1.5) then the equation ̇𝑥(𝑡)+𝑚𝑘=1𝑎𝑘(𝑡)𝑥𝑘(𝑡)=0(1.6)

has an eventually positive solution. To the best of our knowledge, only the opposite inequality (with min𝑘𝑘(𝑡) rather than max𝑘𝑘(𝑡) in the upper bound) was known as a sufficient oscillation condition. Coefficients and advanced arguments are also assumed to be of a more general type than in [20]. Comparison to equations with constant arguments deviations, and coefficients (Corollary 2.8) is also outlined.

For advanced equations with coefficients of different sign, the following result is known.

Theorem C (see [21]). If 0𝑎(𝑡)𝑏(𝑡) and (𝑡)𝑡, then there exists a nonoscillatory solution of the equation ̇𝑥(𝑡)𝑎(𝑡)𝑥((𝑡))+𝑏(𝑡)𝑥(𝑡)=0.(1.7)

This result is generalized in Theorem 2.13 to the case of several positive and negative terms and several advanced arguments; moreover, positive terms can also be advanced as far as the advance is not greater than in the corresponding negative terms.

We also study advanced equations with positive and negative coefficients in the case when positive terms dominate rather than negative ones; some sufficient nonoscillation conditions are presented in Theorem 2.15; these results are later applied to the equation with constant advances and coefficients. Let us note that analysis of nonoscillation properties of the mixed equation with a positive advanced term ̇𝑥(𝑡)+𝑎(𝑡)𝑥((𝑡))𝑏(𝑡)𝑥(𝑔(𝑡))=0,(𝑡)𝑡,𝑔(𝑡)𝑡,𝑎(𝑡)0,𝑏(𝑡)0(1.8)

was also more complicated compared to other cases of mixed equations with positive and negative coefficients [21].

In nonoscillation theory, results on asymptotic properties of nonoscillatory solutions are rather important; for example, for equations with several delays and positive coefficients, all nonoscillatory solutions tend to zero if the integral of the sum of coefficients diverges; under the same condition for negative coefficients, all solutions tend to infinity. In Theorems 2.6 and 2.11, the asymptotic properties of nonoscillatory solutions for advanced equations with coefficients of the same sign are studied.

The paper is organized as follows. Section 2 contains main results on the existence of nonoscillatory solutions to advanced equations and on asymptotics of these solutions: first for equations with coefficients of the same sign, then for equations with both positive and negative coefficients. Section 3 involves some comments and open problems.

2. Main Results

Consider first the equatioṅ𝑥(𝑡)𝑚𝑘=1𝑎𝑘(𝑡)𝑥𝑘(𝑡)=0,(2.1) under the following conditions:(a1)𝑎𝑘(𝑡)0, 𝑘=1,,𝑚, are Lebesgue measurable functions locally essentially bounded for 𝑡0,(a2)𝑘[0,) are Lebesgue measurable functions, 𝑘(𝑡)𝑡, 𝑘=1,,𝑚.

Definition 2.1. A locally absolutely continuous function 𝑥[𝑡0,)𝑅 is called a solution of problem (2.1) if it satisfies (2.1) for almost all 𝑡[𝑡0,).

The same definition will be used for all further advanced equations.

Theorem 2.2. Suppose that the inequality 𝑢(𝑡)𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢(𝑠)𝑑𝑠,𝑡𝑡0(2.2) has a nonnegative solution which is integrable on each interval [𝑡0,𝑏], then (2.1) has a positive solution for 𝑡𝑡0.

Proof. Let 𝑢0(𝑡) be a nonnegative solution of inequality (2.2). Denote 𝑢𝑛+1(𝑡)=𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢𝑛(𝑠)𝑑𝑠,𝑛=0,1,,(2.3) then 𝑢1(𝑡)=𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢0(𝑠)𝑑𝑠𝑢0(𝑡).(2.4) By induction we have 0𝑢𝑛+1(𝑡)𝑢𝑛(𝑡)𝑢0(𝑡). Hence, there exists a pointwise limit 𝑢(𝑡)=lim𝑛𝑢𝑛(𝑡). By the Lebesgue convergence theorem, we have 𝑢(𝑡)=𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡).𝑢(𝑠)𝑑𝑠(2.5) Then, the function 𝑡𝑥(𝑡)=𝑥0exp𝑡𝑡0𝑡𝑢(𝑠)𝑑𝑠forany𝑥0>0(2.6) is a positive solution of (2.1).

Corollary 2.3. If max𝑘𝑘𝑡𝑚(𝑡)𝑖=1𝑎𝑖1(𝑠)𝑑𝑠𝑒,𝑡𝑡0,(2.7) then (2.1) has a positive solution for 𝑡𝑡0.

Proof. Let 𝑢0(𝑡)=𝑒𝑚𝑘=1𝑎𝑘(𝑡), then 𝑢0 satisfies (2.2) at any point 𝑡 where 𝑚𝑘=1𝑎𝑘(𝑡)=0. In the case when 𝑚𝑘=1𝑎𝑘(𝑡)0, inequality (2.7) implies 𝑢0(𝑡)𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢0𝑢(𝑠)𝑑𝑠0(𝑡)𝑚𝑘=1𝑎𝑘(𝑡)expmax𝑘𝑘𝑡(𝑡)𝑢0(=𝑒𝑠)𝑑𝑠𝑚𝑘=1𝑎𝑘(𝑡)𝑚𝑘=1𝑎𝑘(𝑒𝑡)expmax𝑘𝑘𝑡(𝑡)𝑚𝑖=1𝑎𝑖𝑒(𝑠)𝑑𝑠𝑚𝑘=1𝑎𝑘(𝑡)𝑚𝑘=1𝑎𝑘(𝑡)𝑒=1.(2.8) Hence, 𝑢0(𝑡) is a positive solution of inequality (2.2). By Theorem 2.2, (2.1) has a positive solution for 𝑡𝑡0.

Corollary 2.4. If there exists 𝜎>0 such that 𝑘(𝑡)𝑡𝜎 and 0𝑚𝑘=1𝑎𝑘(𝑠)𝑑𝑠<, then (2.1) has an eventually positive solution.

Corollary 2.5. If there exists 𝜎>0 such that 𝑘(𝑡)𝑡𝜎 and lim𝑡𝑎𝑘(𝑡)=0, then (2.1) has an eventually positive solution.

Proof. Under the conditions of either Corollary 2.4 or Corollary 2.5, obviously there exists 𝑡00 such that (2.7) is satisfied.

Theorem 2.6. Let 𝑚𝑘=1𝑎𝑘(𝑠)𝑑𝑠= and 𝑥 be an eventually positive solution of (2.1), then lim𝑡𝑥(𝑡)=.

Proof. Suppose that 𝑥(𝑡)>0 for 𝑡𝑡1, then ̇𝑥(𝑡)0 for 𝑡𝑡1 and ̇𝑥(𝑡)𝑚𝑘=1𝑎𝑘𝑡(𝑡)𝑥1,𝑡𝑡1,(2.9) which implies 𝑡𝑥(𝑡)𝑥1𝑡𝑡1𝑚𝑘=1𝑎𝑘(𝑠)𝑑𝑠.(2.10) Thus, lim𝑡𝑥(𝑡)=.

Consider together with (2.1) the following equation:̇𝑥(𝑡)𝑚𝑘=1𝑏𝑘𝑔(𝑡)𝑥𝑘(𝑡)=0,(2.11) for 𝑡𝑡0. We assume that for (2.11) conditions (a1)-(a2) also hold.

Theorem 2.7. Suppose that 𝑡𝑔𝑘(𝑡)𝑘(𝑡), 0𝑏𝑘(𝑡)𝑎𝑘(𝑡), 𝑡𝑡0, and the conditions of Theorem 2.2 hold, then (2.11) has a positive solution for 𝑡𝑡0.

Proof. Let 𝑢0(𝑡)0 be a solution of inequality (2.2) for 𝑡𝑡0, then this function is also a solution of this inequality if 𝑎𝑘(𝑡) and 𝑘(𝑡) are replaced by 𝑏𝑘(𝑡) and 𝑔𝑘(𝑡). The reference to Theorem 2.2 completes the proof.

Corollary 2.8. Suppose that there exist 𝑎𝑘>0 and 𝜎𝑘>0 such that 0𝑎𝑘(𝑡)𝑎𝑘, 𝑡𝑘(𝑡)𝑡+𝜎𝑘, 𝑡𝑡0, and the inequality 𝜆𝑚𝑘=1𝑎𝑘𝑒𝜆𝜎𝑘(2.12) has a solution 𝜆0, then (2.1) has a positive solution for 𝑡𝑡0.

Proof. Consider the equation with constant parameters ̇𝑥(𝑡)𝑚𝑘=1𝑎𝑘𝑥𝑡+𝜎𝑘=0.(2.13) Since the function 𝑢(𝑡)𝜆 is a solution of inequality (2.2) corresponding to (2.13), by Theorem 2.2, (2.13) has a positive solution. Theorem 2.7 implies this corollary.

Corollary 2.9. Suppose that 0𝑎𝑘(𝑡)𝑎𝑘, 𝑡𝑘(𝑡)𝑡+𝜎 for 𝑡𝑡0, and 𝑚𝑘=1𝑎𝑘1,𝑒𝜎(2.14) then (2.1) has a positive solution for 𝑡𝑡0.

Proof. Since 𝑚𝑘=1𝑎𝑘1/𝑒𝜎, the number 𝜆=1/𝜎 is a positive solution of the inequality 𝜆𝑚𝑘=1𝑎𝑘𝑒𝜆𝜎,(2.15) which completes the proof.

Consider now the equation with positive coefficientṡ𝑥(𝑡)+𝑚𝑘=1𝑎𝑘(𝑡)𝑥𝑘(𝑡)=0.(2.16)

Theorem 2.10. Suppose that 𝑎𝑘(𝑡)0 are continuous functions bounded on [𝑡0,) and 𝑘 are equicontinuous functions on [𝑡0,) satisfying 0𝑘(𝑡)𝑡𝛿, then (2.16) has a nonoscillatory solution.

Proof. In the space 𝐶[𝑡0,) of continuous functions on [𝑡0,), consider the set 𝑀=𝑢0𝑢𝑚𝑘=1𝑎𝑘,(𝑡)(2.17) and the operator (𝐻𝑢)(𝑡)=𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡).𝑢(𝑠)𝑑𝑠(2.18) If 𝑢𝑀, then 𝐻𝑢𝑀.
For the integral operator (𝑇𝑢)(𝑡)=𝑘𝑡(𝑡)𝑢(𝑠)𝑑𝑠,(2.19) we will demonstrate that 𝑇𝑀 is a compact set in the space 𝐶[𝑡0,). If 𝑢𝑀, then (𝑇𝑢)(𝑡)𝐶[𝑡0,)sup𝑡𝑡0𝑡𝑡+𝛿||||𝑢(𝑠)𝑑𝑠sup𝑡𝑡0𝑚𝑘=1𝑎𝑘(𝑡)𝛿<.(2.20) Hence, the functions in the set 𝑇𝑀 are uniformly bounded in the space 𝐶[𝑡0,).
Functions 𝑘 are equicontinuous on [𝑡0,), so for any 𝜀>0, there exists a 𝜎0>0 such that for |𝑡𝑠|<𝜎0, the inequality ||𝑘(𝑡)𝑘||<𝜀(𝑠)2sup𝑡𝑡0𝑚𝑘=1𝑎𝑘(𝑡)1,𝑘=1,,𝑚(2.21) holds. From the relation 𝑘(𝑡0)𝑡0𝑘𝑡(𝑡)=𝑡𝑡0+𝑘(𝑡0)𝑡𝑘(𝑡0)𝑡𝑘(𝑡)𝑘𝑡0=𝑡𝑡0𝑘(𝑡)𝑘𝑡0,(2.22) we have for |𝑡𝑡0|<min{𝜎0,𝜀/2sup𝑡𝑡0𝑚𝑘=1𝑎𝑘(𝑡)} and 𝑢𝑀 the estimate ||𝑡(𝑇𝑢)(𝑡)(𝑇𝑢)0||=||||𝑘(𝑡)t𝑢(𝑠)𝑘(𝑡0)𝑡0||||𝑢(𝑠)𝑑𝑠𝑡𝑡0||||𝑢(𝑠)𝑑𝑠+𝑘(𝑡)𝑘𝑡0||||||𝑢(𝑠)𝑑𝑠𝑡𝑡0||sup𝑡𝑡0𝑚𝑘=1𝑎𝑘||(𝑡)+𝑘(𝑡)𝑘𝑡0||sup𝑡𝑡0𝑚𝑘=1𝑎𝑘<𝜀(𝑡)2+𝜀2=𝜀.(2.23) Hence, the set 𝑇𝑀 contains functions which are uniformly bounded and equicontinuous on [𝑡0,), so it is compact in the space 𝐶[𝑡0,); thus, the set 𝐻𝑀 is also compact in 𝐶[𝑡0,).
By the Schauder fixed point theorem, there exists a continuous function 𝑢𝑀 such that 𝑢=𝐻𝑢, then the function 𝑥(𝑡)=exp𝑡𝑡0𝑢(𝑠)𝑑𝑠(2.24) is a bounded positive solution of (2.16). Moreover, since 𝑢 is nonnegative, this solution is nonincreasing on [𝑡0,).

Theorem 2.11. Suppose that the conditions of Theorem 2.10 hold, 𝑡0𝑚𝑘=1𝑎𝑘(𝑠)𝑑𝑠=,(2.25) and 𝑥 is a nonoscillatory solution of (2.16), then lim𝑡𝑥(𝑡)=0.

Proof. Let 𝑥(𝑡)>0 for 𝑡𝑡0, then ̇𝑥(𝑡)0 for 𝑡𝑡0. Hence, 𝑥(𝑡) is nonincreasing and thus has a finite limit. If lim𝑡𝑥(𝑡)=𝑑>0, then 𝑥(𝑡)>𝑑 for any 𝑡, and thus ̇𝑥(𝑡)𝑑𝑚𝑘=1𝑎𝑘(𝑡) which implies lim𝑡𝑥(𝑡)=. This contradicts to the assumption that 𝑥(𝑡) is positive, and therefore lim𝑡𝑥(𝑡)=0.

Let us note that we cannot guarantee any (exponential or polynomial) rate of convergence to zero even for constant coefficients 𝑎𝑘, as the following example demonstrates.

Example 2.12. Consider the equation ̇𝑥(𝑡)+𝑥((𝑡))=0, where (𝑡)=𝑡𝑡ln𝑡, 𝑡3, 𝑥(3)=1/ln3. Then, 𝑥(𝑡)=1/(ln𝑡) is the solution which tends to zero slower than 𝑡𝑟 for any 𝑟>0.

Consider now the advanced equation with positive and negative coefficientṡ𝑥(𝑡)𝑚𝑘=1𝑎𝑘(𝑡)𝑥𝑘(𝑡)𝑏𝑘𝑔(𝑡)𝑥𝑘(𝑡)=0,𝑡0.(2.26)

Theorem 2.13. Suppose that 𝑎𝑘(𝑡) and 𝑏𝑘(𝑡) are Lebesgue measurable locally essentially bounded functions, 𝑎𝑘(𝑡)𝑏𝑘(𝑡)0, 𝑘(𝑡) and 𝑔𝑘(𝑡) are Lebesgue measurable functions, 𝑘(𝑡)𝑔𝑘(𝑡)𝑡, and inequality (2.2) has a nonnegative solution, then (2.26) has a nonoscillatory solution; moreover, it has a positive nondecreasing and a negative nonincreasing solutions.

Proof. Let 𝑢0 be a nonnegative solution of (2.2) and denote 𝑢𝑛+1(𝑡)=𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢𝑛(𝑠)𝑑𝑠𝑏𝑘(𝑡)exp𝑔𝑘𝑡(𝑡)𝑢𝑛(𝑠)𝑑𝑠,𝑡𝑡0,𝑛0.(2.27) We have 𝑢00, and by (2.2), 𝑢0𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢0(𝑠)𝑑𝑠𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢0(𝑠)𝑑𝑠𝑏𝑘(𝑡)exp𝑔𝑘𝑡(𝑡)𝑢0(𝑠)𝑑𝑠=𝑢1(𝑡).(2.28) Since 𝑎𝑘(𝑡)𝑏𝑘(𝑡)0 and 𝑡𝑔𝑘(𝑡)𝑘(𝑡), then 𝑢1(𝑡)0.
Next, let us assume that 0𝑢𝑛(𝑡)𝑢𝑛1(𝑡). The assumptions of the theorem imply 𝑢𝑛+10. Let us demonstrate that 𝑢𝑛+1(𝑡)𝑢𝑛(𝑡). This inequality has the form 𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢𝑛(𝑠)𝑑𝑠𝑏𝑘(𝑡)exp𝑔𝑘𝑡(𝑡)𝑢𝑛(𝑠)𝑑𝑠𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢𝑛1(𝑠)𝑑𝑠𝑏𝑘(𝑡)exp𝑔𝑘𝑡(𝑡)𝑢𝑛1,(𝑠)𝑑𝑠(2.29) which is equivalent to 𝑚𝑘=1exp𝑘𝑡(𝑡)𝑢𝑛𝑎(𝑠)𝑑𝑠𝑘(𝑡)𝑏𝑘(𝑡)exp𝑘𝑔(𝑡)𝑘(𝑡)𝑢𝑛(𝑠)𝑑𝑠𝑚𝑘=1exp𝑘𝑡(𝑡)𝑢𝑛1𝑎(𝑠)𝑑𝑠𝑘(𝑡)𝑏𝑘(𝑡)exp𝑘𝑔(𝑡)𝑘(𝑡)𝑢𝑛1.(𝑠)𝑑𝑠(2.30) This inequality is evident for any 0𝑢𝑛(𝑡)𝑢𝑛1(𝑡), 𝑎𝑘(𝑡)0, and 𝑏𝑘0; thus, we have 𝑢𝑛+1(𝑡)𝑢𝑛(𝑡).
By the Lebesgue convergence theorem, there is a pointwise limit 𝑢(𝑡)=lim𝑛𝑢𝑛(𝑡) satisfying 𝑢(𝑡)=𝑚𝑘=1𝑎𝑘(𝑡)exp𝑘𝑡(𝑡)𝑢(𝑠)𝑑𝑠𝑏𝑘(𝑡)exp𝑔𝑘𝑡(𝑡)𝑢(𝑠)𝑑𝑠,𝑡𝑡0,(2.31)𝑢(𝑡)0, 𝑡𝑡0. Then, the function 𝑡𝑥(𝑡)=𝑥0exp𝑡𝑡0𝑢(𝑠)𝑑𝑠,𝑡𝑡0(2.32) is a positive nondecreasing solution of (2.26) for any 𝑥(𝑡0)>0 and is a negative nonincreasing solution of (2.26) for any 𝑥(𝑡0)<0.

Corollary 2.14. Let 𝑎𝑘(𝑡) and 𝑏𝑘(𝑡) be Lebesgue measurable locally essentially bounded functions satisfying 𝑎𝑘(𝑡)𝑏𝑘(𝑡)0, and let 𝑘(𝑡) and 𝑔𝑘(𝑡) be Lebesgue measurable functions, where 𝑘(𝑡)𝑔𝑘(𝑡)𝑡. Assume in addition that inequality (2.7) holds. Then, (2.26) has a nonoscillatory solution.

Consider now the equation with constant deviations of advanced argumentṡ𝑥(𝑡)𝑚𝑘=1𝑎𝑘(𝑡)𝑥𝑡+𝜏𝑘𝑏𝑘(𝑡)𝑥𝑡+𝜎𝑘=0,(2.33) where 𝑎𝑘,𝑏𝑘 are continuous functions, 𝜏𝑘0, 𝜎𝑘0.

Denote 𝐴𝑘=sup𝑡𝑡0𝑎𝑘(𝑡), 𝑎𝑘=inf𝑡𝑡0𝑎𝑘(𝑡), 𝐵𝑘=sup𝑡𝑡0𝑏𝑘(𝑡), 𝑏𝑘=inf𝑡𝑡0𝑏𝑘(𝑡).

Theorem 2.15. Suppose that 𝑎𝑘0, 𝑏𝑘0, 𝐴𝑘<, and 𝐵𝑘<.
If there exists a number 𝜆0<0 such that 𝑚𝑘=1𝑎𝑘𝑒𝜆0𝜏𝑘𝐵𝑘𝜆0,(2.34)𝑚𝑘=1𝐴𝑘𝑏𝑘𝑒𝜆0𝜎𝑘0,(2.35) then (2.33) has a nonoscillatory solution; moreover, it has a positive nonincreasing and a negative nondecreasing solutions.

Proof. In the space 𝐶[𝑡0,), consider the set 𝑀={𝑢𝜆0𝑢0} and the operator (𝐻𝑢)(𝑡)=𝑚𝑘=1𝑎𝑘(𝑡)exp𝑡+𝜏𝑘𝑡𝑢(𝑠)𝑑𝑠𝑏𝑘(𝑡)exp𝑡+𝜎𝑘𝑡.𝑢(𝑠)𝑑𝑠(2.36) For 𝑢𝑀, we have from (2.34) and (2.35) (𝐻𝑢)(𝑡)𝑚𝑘=1𝐴𝑘𝑏𝑘𝑒𝜆0𝜎𝑘0,(𝐻𝑢)(𝑡)𝑚𝑘=1𝑎𝑘𝑒𝜆0𝜏𝑘𝐵𝑘𝜆0.(2.37) Hence, 𝐻𝑀𝑀.
Consider the integral operator (𝑇𝑢)(𝑡)=𝑡𝑡+𝛿𝑢(𝑠)𝑑𝑠,𝛿>0.(2.38) We will show that 𝑇𝑀 is a compact set in the space 𝐶[𝑡0,). For 𝑢𝑀, we have (𝑇𝑢)(𝑡)𝐶[𝑡0,)sup𝑡𝑡0𝑡𝑡+𝛿||||||𝜆𝑢(𝑠)𝑑𝑠0||𝛿.(2.39) Hence, the functions in the set 𝑇𝑀 are uniformly bounded in the space 𝐶[𝑡0,).
The equality 𝑡0𝑡+𝛿0𝑡𝑡+𝛿=𝑡𝑡0+𝑡0𝑡+𝛿𝑡0𝑡+𝛿𝑡𝑡+𝛿0+𝛿=𝑡𝑡0𝑡𝑡+𝛿0+𝛿 implies ||𝑡(𝑇𝑢)(𝑡)(𝑇𝑢)0||=||||𝑡𝑡+𝛿𝑢(𝑠)𝑡0𝑡+𝛿0||||𝑢(𝑠)𝑑𝑠𝑡𝑡0||||𝑢(𝑠)𝑑𝑠+𝑡𝑡+𝛿0+𝛿||||||𝜆𝑢(𝑠)𝑑𝑠20||||𝑡𝑡0||.(2.40) Hence, the set 𝑇𝑀 and so the set 𝐻𝑀 are compact in the space 𝐶[𝑡0,).
By the Schauder fixed point theorem, there exists a continuous function 𝑢 satisfying 𝜆0𝑢0 such that 𝑢=𝐻𝑢; thus, the function 𝑡𝑥(𝑡)=𝑥0exp𝑡𝑡0𝑢(𝑠)𝑑𝑠,𝑡𝑡0(2.41) is a positive nonincreasing solution of (2.33) for any 𝑥(𝑡0)>0 and is a negative nondecreasing solution of (2.26) for any 𝑥(𝑡0)<0.

Let us remark that (2.35) for any 𝜆0<0 implies 𝑚𝑘=1(𝐴𝑘𝑏𝑘)<0.

Corollary 2.16. Let 𝑚𝑘=1(𝐴𝑘𝑏𝑘)<0, 𝑚𝑘=1𝐴𝑘>0, and for 𝜆0=ln𝑚𝑘=1𝐴𝑘/𝑚𝑘=1𝑏𝑘max𝑘𝜎k,(2.42) the inequality 𝑚𝑘=1𝑎𝑘𝑒𝜆0𝜏𝑘𝐵𝑘𝜆0(2.43) holds, then (2.33) has a bounded positive solution.

Proof. The negative number 𝜆0 defined in (2.42) is a solution of both (2.34) and (2.35); by definition, it satisfies (2.35), and (2.43) implies (2.34).

Example 2.17. Consider the equation with constant advances and coefficients ̇𝑥(𝑡)𝑎𝑥(𝑡+𝑟)+𝑏𝑥(𝑡+𝑑)=0,(2.44) where 0<𝑎<𝑏, 𝑑>0, 𝑟0. Then, 𝜆0=(1/𝑑)ln(𝑎/𝑏) is the minimal value of 𝜆 for which inequality (2.35) holds; for (2.44), it has the form 𝑎𝑏𝑒𝜆𝑑0.
Inequality (2.34) for (2.44) can be rewritten as 𝑓(𝜆)=𝑎𝑒𝜆𝑟𝑏𝜆0,(2.45) where the function 𝑓(𝑥) decreases on (,ln(𝑎𝑟)/𝑟] if 𝜏>0 and for any negative 𝑥 if 𝑟=0; besides, 𝑓(0)<0. Thus, if 𝑓(𝜆1)<0 for some 𝜆1<0, then 𝑓(𝜆)<0 for any 𝜆[𝜆1,0). Hence, the inequality 𝑓𝜆0𝑎=𝑎𝑏𝑟/𝑑1𝑏𝑑𝑎ln𝑏0(2.46) is necessary and sufficient for the conditions of Theorem 2.15 to be satisfied for (2.44).
Figure 1 demonstrates possible values of advances 𝑑 and 𝑟, such that Corollary 2.16 implies the existence of a positive solution in the case 1=𝑎<𝑏=2. Then, (2.46) has the form 0.5𝑟/𝑑2(ln2)/𝑑, which is possible only for 𝑑>0.5ln20.347 and for these values is equivalent to 𝑟𝑑ln(2ln2/𝑑).ln2(2.47)

3. Comments and Open Problems

In this paper, we have developed nonoscillation theory for advanced equations with variable coefficients and advances. Most previous nonoscillation results deal with either oscillation or constant deviations of arguments. Among all cited papers, only [8] has a nonoscillation condition (Theorem 2.11) for a partial case of (2.1) (with 𝑘(𝑡)=𝑡+𝜏𝑘), which in this case coincides with Corollary 2.4. The comparison of results of the present paper with the previous results of the authors was discussed in the introduction.

Finally, let us state some open problems and topics for research.(1)Prove or disprove: if (2.1), with 𝑎𝑘(𝑡)0, has a nonoscillatory solution, then (2.26) with positive and negative coefficients also has a nonoscillatory solution.

As the first step in this direction, prove or disprove that if (𝑡)𝑡 and the equation ̇𝑥(𝑡)𝑎+(𝑡)𝑥((𝑡))=0(3.1)

has a nonoscillatory solution, then the equation ̇𝑥(𝑡)𝑎(𝑡)𝑥((𝑡))=0(3.2)

also has a nonoscillatory solution, where 𝑎+(𝑡)=max{𝑎(𝑡),0}.

If these conjectures are valid, obtain comparison results for advanced equations.(2)Deduce nonoscillation conditions for (2.1) with oscillatory coefficients. Oscillation results for an equation with a constant advance and an oscillatory coefficient were recently obtained in [22].(3)Consider advanced equations with positive and negative coefficients when the numbers of positive and negative terms do not coincide.(4)Study existence and/or uniqueness problem for the initial value problem or boundary value problems for advanced differential equations.

Acknowledgments

L. Berezansky was partially supported by the Israeli Ministry of Absorption. E. Braverman was partially supported by NSERC.