Home A topological analysis of p(x)-harmonic functionals in one-dimensional nonlocal elliptic equations
Article Open Access

A topological analysis of p(x)-harmonic functionals in one-dimensional nonlocal elliptic equations

  • Christopher S. Goodrich EMAIL logo
Published/Copyright: April 28, 2025

Abstract

We consider a class of one-dimensional elliptic equations possessing a p(x)-harmonic functional as a nonlocal coefficient. As part of our results we treat the model case

M 0 1 u ( x ) p ( x ) d x u ( t ) = λ f t , u ( t ) ,  0 < t < 1

subject to the boundary data

u ( 0 ) = 0 = u ( 1 ) .

In addition, we consider a broader class of problems, of which the model case in a special case, by writing the argument of M as a finite convolution. As part of the analysis, a simple but fundamental lemma in introduced that allows the estimation of u ( x ) p ( x ) in terms of constant exponents; this is the key to circumventing the variable exponent. An unusual array of analytical tools is used, including Sobolev’s inequality. Our results address both existence and nonexistence of solution.

2010 Mathematics Subject Classification: Primary: 33B15; 34B10; 34B18; 42A85; 44A35; Secondary: 26A33; 46E35; 47H30

1 Introduction

Suppose that a and b are L 1 ( 0 , + ) functions. Define the finite convolution of a and b at t ≥ 0, denoted (ab)(t), by

( a b ) ( t ) 0 t a ( t s ) b ( s ) d s .

The finite convolution is an important nonlocal functional, being as it occurs in many frequently studied nonlocal operators, the most well known of which may be the fractional differential and integral operators. For example, if we set a ( t ) 1 Γ ( α ) t α 1 , where 0 < α < 1, then (au)(t) is the α-th order fractional integral of Riemann–Liouville type – see, for example [1], [2], [3], [4], [5], for details.

At the same time, an important functional is the p(x)-harmonic functional, which for u sufficiently regular is defined, in the one-dimensional setting, by

(1.1) u 0 1 u ( x ) p ( x ) d x ,

where p is a sufficiently regular function defined on [0, 1]. In our setting, p and u′ will be continuous on [0, 1]. In the higher dimensional setting, the functional (1.1) takes the form

u Ω D u ( x ) p ( x ) d x ,  Ω R n ,

and it plays a very important role in regularity theory. In part, this is due to minimisers of the functional being related to weak solutions of p(x)-Laplacian equations – see, for example, [6]. Such functionals were introduced by Zhikov [7], subsequently analysed by Coscia and Mingione [8], and since then studied by many authors such as Fey and Foss [9], Ragusa and Tachikawa [10], [11], and Usuba [12]. These problems have applications in electrorheological fluids and thermistors, amongst other areas.

Finally, regarding nonlocal differential equations, one of the most well-studied is the Kirchhoff-type equation, which is a parabolic PDE having the form

u t t M D u L p p Δ u ( x ) = λ f u ( x ) ,  x Ω R n ,

where, clasically, p = 2. When u tt ≡ 0 so that steady-state solutions are sought, and the space dimension is one, then the parabolic PDE reduces to the elliptic one-dimensional nonlocal equation

M u L p p u ( t ) = λ f t , u ( t ) ,  0 < t < 1 .

In this paper we consider a combination of these three themes: finite convolutions, p(x)-harmonic functionals, and Kirchhoff-type nonlocal equations in dimension one. More precisely, we study, for λ > 0 a parameter, both the existence and nonexistence of at least one positive solution to

(1.2) M a | u | p ( ) ( 1 ) u ( t ) = λ f t , u ( t ) ,  0 < t < 1

subject to the boundary data

(1.3) u ( 0 ) = 0 = u ( 1 ) ,

where in (1.2), and, indeed, throughout the remainder of this paper, we use the notation

a | u | p ( ) ( 1 ) 0 1 a ( 1 s ) u ( s ) p ( s ) d s .

Observe that in the model case a(s) ≡ 1, problem (1.2) reduces to the model case

M 0 1 u ( s ) p ( s ) d s u ( t ) = λ f t , u ( t ) ,

which is a one-dimensional Kirchhoff problem with a p(x)-harmonic functional nonlocal coefficient. In case p(s) ≡ p 0 ≥ 1, then the model case further reduces to

M u L p 0 p 0 u ( t ) = λ f t , u ( t ) ,

which is the classical one-dimensional Kirchhoff problem. Thus, problem (1.2) and (1.3) is a natural generalisation. Furthermore, as intimated above, part of the reason for considering the convolutional structure in the nonlocal coefficient is that this permits the study of a wide range of nonlocal coefficients, including fractional-order operators as mentioned above.

One of the difficulties in managing the variable exponent case is that a number of inequalities that we would typically use, amongst these are Sobolev, Jensen, and Hölder, do not interact so nicely with non-constant exponents. In regularity theory, there is a standard strategy for circumventing this difficulty. Indeed, there one typically uses some assumed regularity about the exponent p (often, for example, some degree of Hölder continuity) to approximate quantities involving p(x) by switching from p(x) to either the supremum or infimum of p on a set of very small measure. For typical examples of this strategy, one can look at the arguments in any number of regularity papers involving p(x)-harmonic functionals [6], [8], [10], [11], [12] – see also Diening, et al. [13]. The problem is that in our setting we cannot simply use sets of arbitrarily small measure. Instead we must constantly work on the entire interval on which the differential equation is valid – i.e., 0 < t < 1. This means that we cannot directly import the strategies from regularity theory.

A different strategy, which we utilised in a recent paper [14], is to instead try to utilise a Harnack inequality. In [14] we considered a nonlocal functional coefficient of the form

0 1 u ( s ) p ( s ) d s ,

and to control the variable exponent we calculated, for C > 0 a constant,

0 1 u ( s ) p ( s ) d s C c d u p ( s ) d s , where  ( c , d ) ( 0,1 ) ,

passing from u(t) to ‖u from below by means of a Harnack-type inequality. Then by controlling the size of ‖u we could pass to a constant exponent from below. Once again, this strategy is of no use in the p(x)-harmonic functional setting. The reason is that because the nonlocal functional (in the model case a(t) ≡ 1) has the form

0 1 u ( s ) p ( s ) d s ,

we cannot expect a Harnack-like equality to hold for u′.

So, this is where a completely new idea is required. Indeed, one of the principal contributions of this present work is to develop a way to handle the variable exponent with neither the tools from regularity theory nor the Harnack-inequality approach as in [14]. The resolution is a careful estimation – see Lemma 2.6 – of the variable exponent quantity in terms of a constant exponent quantity, and, crucially, this estimation is global in nature, not merely local as is the custom in regularity theory. Moreover, due to the Dirichlet boundary data (1.3), we also make careful use of a version of Sobolev’s inequality. This was an idea we recently utilised in [15], though some careful adaptations are required in the variable exponent setting.

We have spent most of this introduction explaining the motivation for considering and the difficulties inherent in treating p(x)-harmonic functionals. We wish to conclude this section by highlighting the broader literature on nonlocal differential operators. Most of the existing papers in nonlocal differential equations have a form roughly similar to either the case (and its ODE equivalent)

(1.4) M u L p p Δ u ( x ) = λ f x , u ( x ) ,  x Ω R n ,

or the case (and, again, its ODE equivalent)

(1.5) M D u L p p Δ u ( x ) = λ f x , u ( x ) ,  x Ω R n .

Equation (1.4), in which the nonlocal element depends only on u itself, has been studied in papers by Alves and Covei [16], Corrêa [17], Corrêa, Menezes, and Ferreira [18], do Ó, Lorca, Sánchez, and Ubilla [19], Goodrich [20], Stańczy [21], Wang, Wang, and An [22], Yan and Ma [23], and Yan and Wang [24]. Similarly, equation (1.5), in which the nonlocal element depends on du (or u′), has been studied by Afrouzi, Chung, and Shakeri [25], Ambrosetti and Arcoya [26], Azzouz and Bensedik [27], Boulaaras [28], Boulaaras and Guefaifia [29], Delgado, Morales-Rodrigo, Santos Júnior, and Suárez [30], Graef, Heidarkhani, and Kong [31], Infante [32], [33], and Santos Júnior and Siciliano [34]. Mostly this research has investigated existence of solution (as we do in Section 2), though there are a few works on nonexistence (as we do in Section 3), including recent papers by both Shibata [35], [36], [37] and the author [38], [39], [40]. Additionally, nonlocal differential and integral operators have been studied extensively, and here we direct the interested reader to the following references [41], [42], [43], [44], [45], some of which detail applications to beam deformation, nuclear reactor modelling, and heat transfer problems.

In the analysis of problems such as (1.4) and (1.5) a very common assumption is that M(t) > 0 for all t ≥ 0, and, sometimes in addition, some sort of growth assumptions on M. For example, the results of [17], [18], [19], [21], [22] cannot be applied to the case in which M is either sign-changing or vanishing. Similarly, Ambrosetti and Arcoya [26] require positivity eventually, Santos and Siciliano [34] require positivity on a neighbourhood of zero, and Infante [46] disallows negativity. Moreover, none of these authors’ methods were applied to either the convolutional setting or the p(x)-harmonic setting. By contrast, our recently developed methodology [47], [48], [49], [50], together with Lizama [51], allows a precise localisation of the argument of M, thereby avoiding all manner of assumptions on M; these techniques have been developed further by both Hao and Wang [52] and Song and Hao [53]. Moreover, it is fully compatible with realising the nonlocal element as a finite convolution. The key to this is the judicious construction of a specialised order cone, together with specially constructed open subsets of the order cone – see Section 2 for details.

But none of our previous work allows for p(x)-harmonic functional coefficients. So, in this present work we demonstrate that all of these desirable aspects of our theory carry over to the p(x)-harmonic functional setting. Furthermore, the fundamental lemma we introduce in order to switch between variable and constant exponents (Lemma 2.6) may be useful in other studies related to p(x)-harmonic functionals as it demonstrates a simple but fundamental connection between the variable exponent case and the constant exponent case.

In addition, our results here provide both for an analysis of existence (in Section 2) as well as nonexistence (in Section 3) of solution. The existence results treat the convex regime p(x) > 1, whereas the nonexistence results treat both the convex regime p(x) > 1 as well as the concave regime 0 < p(x) ≤ 1. Moreover, along the way we provide, especially in Section 3, a number of what may be called “norm localisation” results, whereby results are provided that localise, principally, ‖u based on some knowledge of the value of a | u | p ( ) ( 1 ) . Whilst results of this type have appeared in many recent papers on nonlocal differential equations, the ones we provide here are, to the best of the our knowledge, new. Moreover, it is likely that these localisation results will be useful in future studies related to problem (1.2) and (1.3) and its relatives.

2 Existence theory for problem (1.2) and (1.3)

In this section we will begin by detailing both the function spaces in which we work as well as the notation which we will employ throughout. We will then introduce a sequence of lemmata, which will be utilised in the proof of the main existence theorem. A concluding example is presented in order to illustrate the general application of the existence theory.

So, to begin, we will work within the space C 1 [ 0,1 ] equipped with the norm u C 1 max u , u , which subsequently makes the space a Banach space; here we use the notation

u max t [ 0,1 ] u ( t ) ,

for any u C [ 0,1 ] . The notation 1 will henceforth denote the constant function 1 : R { 1 } . In an entirely similar way, the notation 0 will denote the constant function 0 : R { 0 } . In addition, by the notation f [ a , b ] × [ c , d ] M and f [ a , b ] × [ c , d ] m we mean the quantities

f [ a , b ] × [ c , d ] M max ( t , u ) [ a , b ] × [ c , d ] f ( t , u ) f [ a , b ] × [ c , d ] m min ( t , u ) [ a , b ] × [ c , d ] f ( t , u ) ,

respectively, where f is any continuous real-valued function defined on the set [a, b] × [c, d] ⊆ [0, 1] × [0, + ∞).

Since we will prove existence of solution to (1.2) and (1.3) by means of topological fixed point theory, we will study the boundary value problem by studying an equivalent Hammerstein integral operator. As such, we will denote by G : [ 0,1 ] × [ 0,1 ] R the function

(2.1) G ( t , s ) t ( 1 s ) ,  0 t s 1 s ( 1 t ) ,  0 s t 1 ,

which is the Green’s function associated to the operator Lu ≔ − u″ subject to the Dirichlet boundary conditions u(0) = 0 = u(1). Henceforth, by G : [ 0,1 ] R we denote the function

G ( s ) max t [ 0,1 ] G ( t , s ) = s ( 1 s ) .

We next mention the general assumptions imposed on the various functions appearing in the differential equation (1.2).

H1: The functions M : [ 0 , + ) R , f: [0, 1] × [0, + ∞) → [0, + ∞), and a: (0, 1] → [0, + ∞) satisfy the following properties.

  1. Each of M and f is continuous on their respective domains.

  2. Both a L 1 ( 0,1 ] ; [ 0 , + ) and (a1)(1) ≠ 0.

  3. There exist real numbers 0 < ρ 1 < ρ 2 such that M(t) > 0 for t ρ 1 , ρ 2 .

H2: The function p: [0, 1] → (1, + ∞) is continuous, and there exist real numbers p and p + such that

1 < p p ( t ) p + < +

for each t ∈ [0, 1].

As suggested toward the conclusion of Section 1, we will use a positive order cone – namely,

K u C 1 [ 0,1 ] : u ( 0 ) = 0 = u ( 1 ) ,  u 0 , and  ( 1 u ) ( 1 ) C 0 u ,

where the coercivity constant C 0 > 0 in (1u)(1) ≥ C 0u is defined by

C 0 sup s ( 0,1 ) 1 G ( s ) 0 1 G ( t , s ) d t .

In addition, for each ρ ≥ 0 by V ̂ ρ denote the set

V ̂ ρ u K : a | u | p ( ) ( 1 ) < ρ K .

Observe that V ̂ ρ is (relatively) open in K and, moreover, that

V ̂ ρ = u K : a | u | p ( ) ( 1 ) = ρ ,

which is the crucial topological property inasmuch as asserting precise control over the argument of M in (1.3) is concerned. Finally, by T : V ̂ ρ 2 ̄ \ V ̂ ρ 1 C 1 [ 0,1 ] we denote the Hammerstein-type operator

( T u ) ( t ) λ 0 1 M a | u | p ( ) ( 1 ) 1 G ( t , s ) f s , u ( s ) d s .

Then fixed points of T correspond to solutions of (1.2) and (1.3). Note that T is well defined on the domain V ̂ ρ 2 ̄ \ V ̂ ρ 1 on account of condition (H1.3).

Remark 2.1.

Due to the use of Dirichlet boundary conditions and the explicit knowledge of the Green’s function in (2.1), it is easy to compute explicitly the coercivity constant C 0 defined above. Indeed, by, say, [20], Example 2.7], we find by direct computation that

C 0 = sup s ( 0,1 ) 1 G ( s ) 0 1 G ( t , s ) d t = sup s ( 0,1 ) 1 s ( 1 s ) 0 1 G ( t , s ) d t = 1 2 .

In what follows, however, we will generally denote the coercivity constant by explicitly writing C 0. In this way, it will evident where it arises and in what ways it is used in the proofs.

Remark 2.2.

We wish to emphasise, as we mentioned in Section 1, that the hypotheses on the nonlocal coefficient M permit it to be both sign-changing and vanishing. As mentioned in Section 1, this is quite different than most of the existing literature.

We now present a sequence of preliminary lemmata, the first three of which are older results and the remainder of which are original. To this end, we first recall a one-dimensional version of Sobolev’s inequality. The particular statement given here can be found either in Brown, Hinton, and Schwabik [54] or in Talenti [55]. In the sequel, the functions to which we apply Lemma 2.3 will be of class C 1 [ 0,1 ] , which means they will automatically satisfy the regularity hypothesis of absolute continuity in the statement of Lemma 2.3. Also, for notational convenience in the sequel, the function κ : [ 1 , + ) × ( 1 , + ) R will be defined by

(2.2) κ ( p , r ) p 1 + r p 1 r Γ 1 p + 1 r 2 1 + p r 1 p Γ 1 p Γ 1 r , where  1 r + 1 r = 1 .

Lemma 2.3.

Assume that u : [ 0,1 ] R is an absolutely continuous function such that u(0) = 0 = u(1) and

0 1 u ( s ) r d s < + ,

where 1 < r < + ∞. Then for any 1 ≤ p < + ∞ the one-dimensional Sobolev-type inequality

0 1 u ( s ) r d s 1 r 1 κ ( p , r ) 0 1 u ( s ) p d s 1 p

is satisfied, where κ is defined as in (2.2).

We next state the topological fixed point results that we will use. Our results originate from a paper by Yan [56]. The unusual aspect of Yan’s results are that whilst they are in a sense standard fixed point index results, they are nonstandard inasmuch as they do not require the sets in question to be bounded. In particular, in our setting Lemma 2.4 is crucial because whilst V ̂ ρ , for a given ρ ≥ 0, cannot be expected, in general, to satisfy

sup u V ̂ ρ u C 1 < + ,

it does satisfy

sup u V ̂ ρ u < +

as evidenced by Lemma 2.8. And thanks to Yan’s results, the latter is sufficient for our purposes. In addition, we remark that the notation i(T, Ω) in Lemmata 2.4 and 2.5 refers to the fixed point index relative to an operator T on a set Ω K .

Lemma 2.4.

Suppose that Ω C 1 [ 0,1 ] is an open set with

sup x Ω ̄ x < + .

Let T : K Ω ̄ K be continuous such that T K Ω ̄ is relatively compact. If there exists α K \ { 0 } such that

x T x μ α ,

for all x K Ω and μ ≥ 0, then i ( T , K Ω ) = 0 .

Lemma 2.5.

Suppose that Ω C 1 [ 0,1 ] , 0 ∈ Ω, and Ω K is relatively open in K . Let T : K Ω ̄ K be continuous such that T K Ω ̄ is relatively compact. If

T x μ x ,

for all x K Ω and μ ≥ 1, then i ( T , K Ω ) = 1 .

Our first original lemma is a pointwise estimate that allows us to switch from variable exponent growth to constant exponent growth. This lemma states a fundamental pointwise relationship between a function taken to a variable exponent and the same function taken to a constant exponent. As intimated in Section 1, Lemma 2.6 is a sort-of global version of the local variable-to-constant exponent estimates that are used frequently in regularity theory – cf., [6], [8].

Lemma 2.6.

Let f : [ 0,1 ] R be given. Suppose that p: [0, 1] → (1, + ∞) satisfies condition (H2). Then, given any constant q satisfying 1 ≤ q < p , for each t ∈ [0, 1] it holds that

f ( t ) p ( t ) q 2 1 p + q f ( t ) p q 1 .

Proof.

First notice that

(2.3) f ( t ) p q f ( t ) + 1 p q f ( t ) + 1 p ( t ) q .

At the same time, observe, for each 0 ≤ t ≤ 1, that

(2.4) f ( t ) + 1 p ( t ) q 2 p ( t ) q 1 f ( t ) p ( t ) q + 1 2 p + q 1 f ( t ) p ( t ) q + 1 ,

using, pointwise, the inequality

( a + b ) x 2 x 1 a x + b x ,

which holds for any fixed x > 1 by the convexity of tt x . Then putting inequalities (2.3) and (2.4) together we deduce that

(2.5) 2 p + q 1 f ( t ) p ( t ) q + 1 f ( t ) + 1 p ( t ) q f ( t ) p q ,

for each t ∈ [0, 1]. Finally, from inequality (2.5) it follows that

f ( t ) p ( t ) q + 1 2 1 p + q f ( t ) p q

so that

(2.6) f ( t ) p ( t ) q 2 1 p + q f ( t ) p q 1 ,

for each t ∈ [0, 1], as claimed.□

As a consequence of Lemma 2.6 we obtain the following corollary.

Corollary 2.7.

Suppose that the hypotheses of Lemma 2.6 are true. Then under the additional assumption that f C 0 [ 0,1 ] it holds that

0 1 f ( s ) p ( s ) q d t 2 1 p + q 0 1 f ( s ) p q d s 1 = 2 1 p + q f L p q p q 1 .

Proof.

The conclusion of the corollary follows at once by integrating both sides of inequality (2.6) on the interval [0,1], together with the regularity assumptions on f and p as well as the fact that p q > 1 .□

One of the primary consequences of Corollary 2.7, and thus of Lemma 2.6, is that we conclude that for each ρ ≥ 0 the set V ̂ ρ satisfies the boundedness condition

sup u V ̂ ρ u < + .

Thus, much as in [15], we will be able to use Lemmata 2.4 and 2.5 in spite of the fact that

sup u V ̂ ρ u C 1 = +

can occur. The former observation is the content of the next lemma. Furthermore, as an upshot of the next lemma we deduce a localisation of u V ̂ ρ in terms of both ‖u and u C 1 .

Lemma 2.8.

Assume that (H1)–(H2) hold and fix ρ > 0. Assume that

a 1 1 q L 1 ( 0,1 ]

for some number q 1 , p . If u V ̂ ρ , then

u C 0 1 κ 1 , p q 2 p + q 1 ρ 1 q a 1 1 q 1 ( 1 ) q 1 q + 1 q p .

Moreover, if u V ̂ ρ , then both

u C 1 ρ ( a 1 ) ( 1 ) 1 p + 1

and

u C 0 1 κ 1 , p q 2 p + q 1 ρ 1 q a 1 1 q 1 ( 1 ) q 1 q + 1 q p

Finally,

sup u V ̂ ρ u < + .

Proof.

We begin by proving the first statement in the lemma. Let us first observe that since u V ̂ ρ it follows that

ρ > a | u | p ( ) ( 1 ) .

Accordingly, choosing q > 1 sufficiently close to 1 such that q < p , by the reverse Hölder inequality it follows that

(2.7) ρ > a | u | p ( ) ( 1 ) a 1 1 q 1 ( 1 ) 1 q 1 | u | p ( ) q ( 1 ) q .

Here we use the fact that

a 1 1 q 1 ( 1 ) 1 q < +

since, by the assumption in the statement of the lemma,

a 1 1 q L 1 ( 0,1 ] .

Now we estimate from below the second factor appearing on the right-hand side of inequality (2.7). So, on account of Corollary 2.7 note that

(2.8) 1 | u | p ( ) q ( 1 ) q = 0 1 u ( s ) p ( s ) q d s q 2 1 p + q 0 1 u ( s ) p q d s 1 q .

Now, recalling that u(0) = 0 = u(1) since as u V ̂ ρ K , due to Sobolev’s inequality, i.e., Lemma 2.3, we see that

(2.9) 0 1 u ( s ) p q d s 1 κ 1 , p q 0 1 u ( s ) d s p q ,

keeping in mind that

p q > 1 .

Putting estimates (2.8) and (2.9) together we see that

(2.10) 1 | u | p ( ) q ( 1 ) q 2 1 p + q 0 1 u ( s ) p q d s 1 q 2 1 p + q 1 κ 1 , p q 0 1 u ( s ) d s p q 1 q = 2 1 p + q 1 κ 1 , p q 0 1 u ( s ) d s p q 1 q = 2 1 p + q 1 κ 1 , p q ( 1 u ) ( 1 ) p q 1 q 2 1 p + q 1 κ 1 , p q C 0 u p q 1 q ,

where we have used the coercivity condition

( 1 u ) ( 1 ) C 0 u ,

since u K ; note, also, that in (2.10) we have used the fact that u ≡ |u| since u K . Finally, from inequality (2.10) combined with inequality (2.7) it follows that

(2.11) ρ a 1 1 q 1 ( 1 ) 1 q 1 | u | p ( ) q ( 1 ) q a 1 1 q 1 ( 1 ) 1 q 2 1 p + q 1 κ 1 , p q C 0 u p q 1 q .

Therefore, from inequality (2.11) we deduce that

2 1 p + q 1 κ 1 , p q C 0 u p q ρ 1 q a 1 1 q 1 ( 1 ) q 1 q + 1

so that

u C 0 1 κ 1 , p q 2 p + q 1 ρ 1 q a 1 1 q 1 ( 1 ) q 1 q + 1 q p ,

which establishes the first claim in the statement of the lemma. Moreover, taking the supremum on both sides of this inequality over all u V ̂ ρ yields the estimate

sup u V ̂ ρ u < + ,

which proves the third claim in the lemma.

Lastly, we prove the second claim in the lemma. In this case, we note that if u V ̂ ρ , then

(2.12) ρ = a | u | p ( ) ( 1 ) a ( 1 + | u | ) p ( ) ( 1 ) a 1 + u 1 p ( ) ( 1 ) .

Next, using the estimate

1 + u p ( s ) 1 + u C 1 p ( s ) 1 + u C 1 p + ,  s [ 0,1 ] ,

it follows that

(2.13) ρ a 1 + u 1 p ( ) ( 1 ) a 1 + u C 1 1 p + ( 1 ) .

Consequently, putting (2.12) and (2.13) together we conclude that

1 + u C 1 p + ρ ( a 1 ) ( 1 )

so that

u C 1 ρ ( a 1 ) ( 1 ) 1 p + 1 ,

as claimed, and where here we use the fact that (a1)(1) > 0, by assumption. Finally, since the switch from u V ̂ ρ to u V ̂ ρ does not induce any changes in the first part of the proof but for the fact that we initially have the identity

ρ = a | u | p ( ) ( 1 )

instead of an inequality, it follows that the first part of the proof generates the upper bound on ‖u claimed in the second claim of the lemma. And this completes the proof. □

Remark 2.9.

As mentioned elsewhere it may well occur that sup u V ̂ ρ u C 1 = + . But Lemma 2.8 is the key to establishing that the modified fixed point theory of Lemmata 2.4 and 2.5, nonetheless, may be used.

Our final preliminary lemma records the fact, which is well known to specialists in this context, that T maps a solid annular subregion of K into itself. Since this fact has been proved elsewhere in analogous circumstances and since the inclusion of the variable exponent does not materially alter such proofs, we merely record this fact and omit the proof.

Lemma 2.10.

Assume that conditions (H1)–(H2) are true. Then the operator T is completely continuous on V ̂ ρ 2 ̄ \ V ̂ ρ 1 . Moreover,

T V ̂ ρ 2 ̄ \ V ̂ ρ 1 K .

Proof.

Omitted – see, for example, [15], Lemma 2.11].□

With the necessary lemmata dispatched we state and prove our existence theorem. Prior to stating the theorem, we introduce the notation

E ρ 0 , C 0 1 κ 1 , p q 2 p + q 1 ρ 1 q a 1 1 q 1 ( 1 ) q 1 q + 1 q p R ,

for any ρ > 0. This notation will be used in the statement and proof of Theorem 2.11.

Theorem 2.11.

Suppose that each of conditions (H1) and (H2) is true. If, in addition, there exists a number q 1 , p such that a 1 1 q L 1 ( 0,1 ] and the numbers ρ 1 and ρ 2 satisfy, respectively, the inequalities

  1. 2 1 p + a 1 1 q 1 ( 1 ) 1 q λ f [ 0,1 ] × E ρ 1 m M ρ 1 p 0 1 0 1 G ( t , s ) d s d t p ( a 1 ) ( 1 ) > ρ 1 and

  2. 0 1 a ( 1 t ) λ f [ 0,1 ] × E ρ 2 M M ρ 2 t 2 t + 1 2 p ( t ) d t < ρ 2 ,

then problem (1.2) subject to the Dirichlet boundary data (1.3) has at least one positive solution

u 0 V ̂ ρ 2 \ V ̂ ρ 1 ̄ .

Moreover, u 0 satisfies the dual localisation

u 0 C 0 1 κ 1 , p q 2 p + q 1 ρ 2 1 q a 1 1 q 1 ( 1 ) q 1 q + 1 q p

and

u 0 C 1 ρ 1 ( a 1 ) ( 1 ) 1 p + 1 .

Proof.

Let us begin by noting that by Lemma 2.10 we know that T is a completely continuous operator on V ̂ ρ 2 ̄ \ V ̂ ρ 1 that maps into K . For later reference, define the function α: [0, 1] → [0, + ∞) by

α ( t ) t ( 1 t ) .

As was shown in [15], pp. 476–477], the function α is an element of K \ { 0 } . Moreover, we remark that 0 V ̂ ρ 2 since

a | 0 | p ( ) ( 1 ) = ( a 0 ) ( 1 ) = 0 ,

which is necessary for the proper application of the fixed point theory.

We first aim to use Lemma 2.4. On account of Lemma 2.8 we know that the former lemma is properly applicable in this context. Thus, we claim that uTu + μα for each u V ̂ ρ 1 and μ ≥ 0, with the function α as in the previous paragraph. So, for contradiction suppose that there is u V ̂ ρ 1 and μ ≥ 0 such that uTu + μα. In particular, differentiating each side, taking the absolute value of the derivatives, and taking both sides to the p(t)-th power yields

(2.14) u ( t ) p ( t ) = ( T u + μ α ) ( t ) p ( t ) ,

for each t ∈ [0, 1]. Now convolving both sides of (2.14) with the kernel a and evaluating at t = 1 we arrive at

(2.15) ρ 1 = a u p ( ) ( 1 ) = a ( T u + μ α ) p ( ) ( 1 ) ,

using the fact that

u V ̂ ρ 1 a u p ( ) ( 1 ) = ρ 1 .

We next estimate from below the right-hand side of (2.15). First we note that by Lemma 2.6 we can write

(2.16) ( T u + μ e ) ( t ) p ( t ) 2 1 p + ( T u + μ e ) ( t ) p 1 ,

for each t ∈ [0, 1]. Then using (2.16) in (2.15) yields the estimate

(2.17) a ( T u + μ α ) p ( ) ( 1 ) a | ( T u + μ α ) | p 2 1 p + 1 ( 1 ) a 1 1 q 1 ( 1 ) 1 q 1 | ( T u + μ α ) | p q ( 1 ) q 2 1 p + ( a 1 ) ( 1 ) ,

where to obtain the second inequality we have used the reverse Hölder inequality; the number q in (2.17) and henceforth is the number referenced in the statement of the theorem. Next note that the map

t ( T u + μ α ) ( t )

satisfies the boundary data (1.3) since T u K and by direct calculation

( T u + μ α ) ( 0 ) = ( T u ) ( 0 ) + μ α ( 0 ) = 0 = ( T u ) ( 1 ) + μ α ( 1 ) = ( T u + μ α ) ( 1 ) .

In addition, by assumption T u + μ α C 1 [ 0,1 ] . Then applying the Sobolev inequality (Lemma 2.3) we see that

(2.18) 1 | ( T u + μ α ) | p q ( 1 ) q 1 | T u + μ α | ( 1 ) p 1 | T u | ( 1 ) p = ( 1 T u ) ( 1 ) p ,

where we use the fact that |Tu| ≡ Tu. So, all in all, putting inequalities (2.15), (2.17), and (2.18) together we see that

(2.19) ρ 1 = a ( T u + μ α ) p ( ) ( 1 ) a 1 1 q 1 ( 1 ) 1 q 1 | ( T u + μ α ) | p q ( 1 ) q 2 1 p + ( a 1 ) ( 1 ) a 1 1 q 1 ( 1 ) 1 q ( 1 T u ) ( 1 ) p 2 1 p + ( a 1 ) ( 1 ) .

Notice that by Lemma 2.8 since u V ̂ ρ 1 it follows that

0 u ( t ) u C 0 1 κ 1 , p q 2 p + q 1 ρ 1 1 q a 1 1 q 1 ( 1 ) q 1 q + 1 q p

whenever 0 ≤ t ≤ 1. So, we calculate

(2.20) ( 1 T u ) ( 1 ) p = λ 0 1 0 1 M ρ 1 1 G ( t , s ) f s , u ( s ) d s d t p = λ M ρ 1 p 0 1 0 1 G ( t , s ) f s , u ( s ) d s d t p λ f [ 0,1 ] × E ρ 1 m M ρ 1 p 0 1 0 1 G ( t , s ) d s d t p .

So, putting (2.20) into (2.19) yields, by way of condition (1) in the statement of the theorem,

(2.21) ρ 1 2 1 p + a 1 1 q 1 ( 1 ) 1 q λ f [ 0,1 ] × E ρ 1 m M ρ 1 p 0 1 0 1 G ( t , s ) d s d t p ( a 1 ) ( 1 ) > ρ 1 ,

which is a contradiction. Hence, (2.21) together with Lemma 2.3 imply that

(2.22) i T , V ̂ ρ 1 = 0 .

We next aim to utilise Lemma 2.4. So, to this end, suppose for contradiction that Tuμu for some μ ≥ 1 and u V ̂ ρ 2 . Now, notice that since u V ̂ ρ 2 , it follows that

a | u | p ( ) ( 1 ) = ρ 2 .

So, taking the absolute value of the derivative of both sides of μuTu, subsequently exponentiating each side to the p(t)-th power, convolving each side with a and then evaluating the convolutions at t = 1 yields the identity

(2.23) ρ 2 = a | u | p ( ) ( 1 ) a | μ u | p ( ) ( 1 ) = a | ( T u ) | p ( ) ( 1 ) ,

where we have used the calculation

μ p ( t ) μ p 1 ,

for each t ∈ [0, 1], recalling that μ ≥ 1.

We wish to estimate from above the right-hand side of (2.23). So, note that

(2.24) a | ( T u ) | p ( ) ( 1 ) = 0 1 a ( 1 t ) λ 0 1 M ρ 2 1 G t ( t , s ) f s , u ( s ) d s p ( t ) d t 0 1 a ( 1 t ) λ f [ 0,1 ] × E ρ 2 M M ρ 2 p ( t ) 0 1 G t ( t , s ) d s p ( t ) d t = 0 1 a ( 1 t ) λ f [ 0,1 ] × E ρ 2 M M ρ 2 t 2 t + 1 2 p ( t ) d t

using the fact that

0 1 G t ( t , s ) d s = t 2 t + 1 2 .

Then combining (2.23) with (2.24) and recalling condition (2) in the statement of the theorem we see that

(2.25) ρ 2 a | ( T u ) | p ( ) ( 1 ) 0 1 a ( 1 t ) λ f [ 0,1 ] × E ρ 2 M M ρ 2 t 2 t + 1 2 p ( t ) d t < ρ 2 ,

which is a contradiction. Therefore, inequality (2.25) together with an application of Lemma 2.5 implies that

(2.26) i T , V ̂ ρ 2 = 1 .

Finally, combining (2.22), (2.26), and properties of the index i imply that T has a fixed point, say u 0, satisfying

u 0 V ̂ ρ 2 \ V ̂ ρ 1 ̄ .

Furthermore, due to Lemma 2.8, we deduce that u 0 satisfies the dual localisation

u 0 C 0 1 κ 1 , p q 2 p + q 1 ρ 2 1 q a 1 1 q 1 ( 1 ) q 1 q + 1 q p

and

u 0 C 1 ρ 1 ( a 1 ) ( 1 ) 1 p + 1 ,

exactly as in the statement of the theorem. And this completes the proof.□

Recall from Section 1 that the case a1 leads to an important model case, in which the nonlocal data takes the form

M a | u | p ( ) ( 1 ) M 1 | u | p ( ) ( 1 ) = M 0 1 u ( s ) p ( s ) d s ,

where

u 0 1 u ( s ) p ( s ) d s

is a one-dimensional p(x)-harmonic functional. Corollary 2.12 addresses this special case.

Corollary 2.12.

Let a1. Suppose that each of conditions (H1) and (H2) is true. If, in addition, the numbers ρ 1 and ρ 2 satisfy, respectively, the inequalities

  1. 2 1 p + λ f [ c , d ] × E ρ 1 m M ρ 1 p 0 1 0 1 G ( t , s ) d s d t p 1 > ρ 1 and

  2. 0 1 λ f [ 0,1 ] × E ρ 2 M M ρ 2 t 2 t + 1 2 p ( t ) d t < ρ 2 ,

then problem (1.2) subject to the Dirichlet boundary data (1.3) has at least one positive solution

u 0 V ̂ ρ 2 \ V ̂ ρ 1 ̄ .

Moreover, u 0 satisfies the dual localisation

u 0 C 0 1 κ 1 , p q 2 p + q 1 ρ 2 1 q a 1 1 q 1 ( 1 ) q 1 q + 1 q p

and

u 0 C 1 ρ 1 1 p + 1 .

Remark 2.13.

Using the fact that

max t [ 0,1 ] t 2 t + 1 2 = 1 2 ,

it follows that, for example, in condition (2) in the statement of Corollary 2.12 one could replace this condition by the slightly simpler (but stronger) condition

0 1 λ f [ 0,1 ] × E ρ 2 M 2 M ρ 2 p ( t ) d t < ρ 2 ,

which could be easier to calculate since the quantity λ f [ 0,1 ] × E ρ 2 M 2 M ρ 2 is a constant.

We conclude with an example to illustrate the application of the results presented herein.

Example 2.14.

We will consider the case in which a1, as this leads to an important model case and also allows us to utilise the slightly simpler Corollary 2.12. So, we consider the boundary value problem

(2.27) M 0 1 u ( x ) p ( x ) d x u ( t ) = λ f t , u ( t ) ,  t ( 0,1 ) u ( 0 ) = 0 u ( 1 ) = 0 ,

where

p ( x ) 2 + e x cos x

and

M ( t ) 9 + t 2 ,  0 t < 3 ( 3 t ) ( t 9 ) ,  t 3 .

Now, notice that

p 1.82 p ( x ) 3 p +

for 0 ≤ x ≤ 1. Therefore, take

q 1.8

so that 1 < q < p as required. Finally, for ɛ 1 ∈ (0, 3) to be specified momentraily, take

ρ 1 3 + ε 1

and

ρ 2 6 .

With the preceding definitions in hand, we first calculate

(2.28) 2 1 p + λ f [ 0,1 ] × E ρ 1 m M ρ 1 p 0 1 0 1 G ( t , s ) d s d t p 1 = 2 2 λ f [ 0,1 ] × E ρ 1 m 9 + 3 + ε 1 2 1.82 1 12 1.82 1 = 2 2 λ f [ 0,1 ] × E ρ 1 m 6 ε 1 + ε 1 2 1.82 1 12 1.82 1

In addition, we calculate

(2.29) 0 1 λ f [ 0,1 ] × E ρ 2 M M ρ 2 t 2 t + 1 2 p ( t ) d t = 0 1 λ f [ 0,1 ] × E ρ 2 M 9 t 2 t + 1 2 2 + e t cos t d t 0 1 λ f [ 0,1 ] × E ρ 2 M 9 4 + 2 e t cos t d t 1 2 0 1 t 2 t + 1 2 4 + 2 e t cos t d t 1 2 0.085685 0 1 λ f [ 0,1 ] × E ρ 2 M 9 4 + 2 e t cos t d t 1 2 ,

where we have used Hölder’s inequality to simplify the calculation of the integral and replace its exact value with a simpler upper bound.

Now, regarding (2.28) notice that

lim ε 1 0 + 2 2 λ f [ 0,1 ] × E ρ 1 m 6 ε 1 + ε 1 2 1.82 1 12 1.82 1 = + > ρ 1 ,

which means that so long as f [ 0,1 ] × E ρ 1 m 0 , then condition (1) in Corollary 2.12 can be satisfied for any f and λ > 0 so long as ρ 1 is taken sufficiently close to (and greater than) 3. At the same time, (2.29) implies that so long as

(2.30) 0 1 λ f [ 0,1 ] × E ρ 2 M 9 4 + 2 e t cos t d t < 4903.31441 ,

then condition (2) in Corollary 2.12 will be satisfied. In summary, provided that (2.30) is satisfied and f [ 0,1 ] × E 3 + ε 1 m > 0 , then problem (2.27) will have at least one positive solution, u 0, satisfying

u 0 V ̂ 6 ̄ \ V ̂ 3 + ε 1 .

Finally, by the localisations provided in the statement of Corollary 2.12 we know that

u 0 1 2 κ 1 , 1.82 1.8 2 3 1.8 1 6 1 1.8 + 1 1.8 1.82 1.37240

and

u 0 C 1 3 + ε 1 1 3 1 .

Remark 2.15.

Notice that

  1. M(0) < 0; and

  2. lim t→+ M(t) = −∞.

As discussed in Section 1, until recently this is very unusual in the literature on nonlocal elliptic problems. And, in particular, it demonstrates that the good aspects of our theory continue to work even in the setting of p(x)-harmonic functional coefficients.

Remark 2.16.

In principle our results could be extended to radially symmetric solutions of the elliptic PDE

M Ω D u ( x ) p | x | d x Δ u ( x ) = λ f u ( x ) , | x | ,  x Ω R n u ( x ) = 0 ,  x Ω ,

where, for 0 < r < R < + ∞,

Ω x R n : r < | x | < R

is a closed annular region. Such a problem involves a p( x )-harmonic functional coefficient, albeit one in which the exponent p depends at most radially on x – i.e., p | x | in lieu of p( x ). This could represent an interesting, nontrivial avenue for further research.

3 Nonexistence theory for problem (1.2) and (1.3)

In this section we present a nonexistence theory for problem (1.2) and (1.3) both in the concave case 0 < p(x) ≤ 1 and in the convex case 1 < p(x). In order to present the nonexistence results, we will require a sequence of norm localisation results. These results, similar to Lemma 2.8 in Section 2, provide bounds on ‖u under the assumption that u V ̂ ρ for some ρ > 0.

Our first such lemma provides an upper bound on ‖u in case p(x) > 1. Although Lemma 2.8 earlier provided a similar result, here we take a very different approach under the assumption that the kernel a is bounded away from zero – cf., Remark 3.2.

Lemma 3.1.

Assume that conditions (H1)–(H2) hold and that, in addition,

a m inf t ( 0,1 ] a ( t ) > 0 .

If u V ̂ ρ for some ρ > 0, then

u 2 p + 1 + 2 p + 1 a m ρ 1 p .

Proof.

Since u K , it follows that u(0) = 0 = u(1). Let t 0 ∈ [0, 1] be such that

u = u t 0 .

Since u(0) = 0 = u(1) and u is nonnegative, we may assume without loss that 0 < t 0 < 1. Consequently, since p > 1, by Hölder’s inequality we see that

u = u t 0 u ( 0 ) = 0 t 0 u ( s ) d s 0 t 0 u ( s ) d s 0 1 u ( s ) p d s 1 p

so that by Lemma 2.6, choosing q = 1 in the statement of the lemma,

(3.1) u 0 1 u ( s ) p d s 1 p 2 p + 1 + 2 p + 1 0 1 u ( s ) p ( s ) d s 1 p .

Now, since, by assumption,

a m inf t ( 0,1 ] a ( t ) > 0 ,

it follows that

(3.2) 0 1 u ( s ) p ( s ) d s a m 1 0 1 a ( 1 t ) u ( s ) p ( s ) d s = a m 1 a | u | p ( ) ( 1 ) = a m 1 ρ ,

using that u V ̂ ρ . All in all, then, using inequality (3.2) in inequality (3.1) implies that

u 2 p + 1 + 2 p + 1 0 1 u ( s ) p ( s ) d s 1 p 2 p + 1 + 2 p + 1 a m ρ 1 p ,

as desired.□

Remark 3.2.

Note that the localisation provided by Lemma 3.1 is somewhat cleaner than the localisation provided by Lemma 2.8 in Section 2 since we are not using Sobolev’s inequality in the proof of Lemma 3.1. However, we do require that a be bounded away from zero, which makes the result slightly less general. Nonetheless, each of the commonly occurring kernels a1 and a ( t ) = 1 Γ ( α ) t α 1 , 0 < α < 1, does satisfy this requirement. So, it seems to be only a mildly restrictive additional hypothesis.

We next provide a norm localisation result that will be needed in the proof of Lemma 3.4 momentarily. In order to prove this lemma, we assume that u′ is nonincreasing. For the purposes of using Lemma 3.4 in the nonexistence result, this assumption will not present a problem.

Lemma 3.3.

Assume that u K and that u′ is nonincreasing. Then

0 1 u ( s ) d s = 2 u .

Proof.

Let us first assume that u0. Define the measurable sets E +, E , E 0 ⊆ [0, 1] as follows.

E + t [ 0,1 ] : u ( t ) > 0 E t [ 0,1 ] : u ( t ) < 0 E 0 t [ 0,1 ] : u ( t ) = 0

Let t 0E 0. Note that such a point must exist since u C [ 0,1 ] , u(0) = 0 = u(1), and u0. Then, seeing as u K so that u satisfies the Dirichlet boundary data (1.3), it follows that

0 1 u ( s ) d s = E + u ( s ) d s = u t 0 + E u ( s ) d s = u t 0 + E 0 u ( s ) d s = 0 = 2 u t 0 = 2 u ,

as claimed. On the other hand, if u0, then the conclusion of the theorem is trivially true. And this completes the proof of the theorem.□

Our third norm localisation lemma provides a lower bound on ‖u in case 0 < p(x) ≤ 1. For this lemma, and some of the other results in this section, we require the following hypothesis.

H3: The function p: [0, 1] → (0, 1] is continuous, and there exist real numbers p and p + such that

0 < p p ( t ) p + 1

for each t ∈ [0, 1].

Observe that conditions (H2) and (H3) collectively address both qualitative regimes for the exponent function p. In particular, (H2) allows for convex behaviour, whereas (H3) allows for concave behaviour. We also note that, strictly speaking, Lemma 3.4 requires that p + < 1. However, Corollary 3.7 does allow for the case p + = 1.

Lemma 3.4.

Assume that each of conditions (H1) and (H3) holds. In addition, assume that there exists ɛ > 0 such that

  1. (1 + ɛ)p + ≤ 1; and

  2. a L 1 + ε ε ( 0,1 ) .

If u V ̂ ρ , for some ρ > 0, and u′ is nonincreasing, then

u 2 1 + p + p min ρ 1 p a L 1 + ε ε 1 p , ρ 1 p + a L 1 + ε ε 1 p + .

Proof.

Let u V ̂ ρ . Recall from condition (H3) that

0 < p p ( t ) p + 1 .

Let ɛ > 0 be the number from the statement of the lemma – i.e., ɛ > 0 is selected such that both

( 1 + ε ) p + 1

and

a L 1 + ε ε ( 0,1 ) .

Then Hölder’s inequality implies that

(3.3) ρ = 0 1 a ( 1 s ) u ( s ) p ( s ) d s 0 1 a ( s ) 1 + ε ε ε 1 + ε 0 1 u ( s ) ( 1 + ε ) p ( s ) d s 1 1 + ε .

We estimate from above the second factor on the right-hand side of (3.3) by writing

(3.4) 0 1 u ( s ) ( 1 + ε ) p ( s ) d s 1 1 + ε = { s : | u ( s ) | 1 } u ( s ) ( 1 + ε ) p ( s ) d s + { s : | u ( s ) | > 1 } u ( s ) ( 1 + ε ) p ( s ) d s 1 1 + ε { s : | u ( s ) | 1 } u ( s ) ( 1 + ε ) p d s + { s : | u ( s ) | > 1 } u ( s ) ( 1 + ε ) p + d s 1 1 + ε 0 1 u ( s ) ( 1 + ε ) p d s + 0 1 u ( s ) ( 1 + ε ) p + d s 1 1 + ε 0 1 u ( s ) ( 1 + ε ) p d s 1 1 + ε + 0 1 u ( s ) ( 1 + ε ) p + d s 1 1 + ε 0 1 u ( s ) d s p + 0 1 u ( s ) d s p + = 2 p u p + 2 p + u p + 2 p + u p + u p + .

Note that to obtain the third inequality in (3.4) we have used the fact that t t 1 1 + ε is a subadditive map. In addition, to obtain the fourth inequality in (3.4) we have used Jensen’s inequality, keeping in mind that

0 < ( 1 + ε ) p ( 1 + ε ) p + 1 .

Finally, we have used Lemma 3.3 twice in (3.4) to write

0 1 u ( s ) d s = 2 u .

All in all, then, using inequality (3.4) in inequality (3.3) we obtain that

(3.5) ρ = 0 1 a ( 1 s ) u ( s ) p ( s ) d s 0 1 a ( s ) 1 + ε ε ε 1 + ε 0 1 u ( s ) ( 1 + ε ) p ( s ) d s 1 1 + ε a L 1 + ε ε 2 p + u p + u p + .

Now, if ‖u≥ 1, then inequality (3.5) reduces to

(3.6) ρ = 0 1 a ( 1 s ) u ( s ) p ( s ) d s a L 1 + ε ε 2 p + u p + u p + 2 u p + a L 1 + ε ε 2 1 + p + u p + ,

whereas if ‖u < 1, then inequality (3.5) reduces to

(3.7) ρ = 0 1 a ( 1 s ) u ( s ) p ( s ) d s a L 1 + ε ε 2 p + u p + u p + 2 u p a L 1 + ε ε 2 1 + p + u p .

So, altogether from (3.6) and (3.7) we conclude that

(3.8) u min 2 1 + p + p ρ 1 p a L 1 + ε ε 1 p , 2 1 + p + p + ρ 1 p + a L 1 + ε ε 1 p + .

Finally, note that since

1 + p + p 1 + p + p + ,

it follows that

(3.9) 1 + p + p 1 + p + p + < 0 .

And from inequality (3.9) we conclude that

(3.10) 0 < 2 1 + p + p 2 1 + p + p + .

Therefore, from (3.10) we conclude that (3.8) can be rewritten as

u min 2 1 + p + p ρ 1 p a L 1 + ε ε 1 p , 2 1 + p + p + ρ 1 p + a L 1 + ε ε 1 p + 2 1 + p + p min ρ 1 p a L 1 + ε ε 1 p , ρ 1 p + a L 1 + ε ε 1 p + .

And this completes the proof.□

Remark 3.5.

Let’s consider how the Riemann–Liouville fractional derivative kernel, a ( t ) 1 Γ ( α ) t α 1 , 0 < α < 1, interacts with the assumption

a L 1 + ε ε ( 0,1 ) ,

which we required in Lemma 3.4. Note that the improper integral

a L 1 + ε ε 1 + ε ε = 0 1 1 Γ ( α ) t α 1 1 + ε ε d t

converges if and only if

( α 1 ) 1 + ε ε > 1 α > 1 ε 1 + ε = 1 1 + ε .

In other words, a L 1 + ε ε ( 0,1 ) if and only if

α 1 1 + ε , 1 .

Note that the set 1 1 + ε , 1 is nonempty for all ɛ > 0. That is, for any given ɛ > 0, there always exists a nonempty interval of values for α such that t 1 Γ ( α ) t α 1 satisfies the required regularity. Indeed, this is a consequence of the fact that ε 1 1 + ε is decreasing together with the fact that

lim ε 0 + 1 1 + ε = 1 .

In particular, the allowable range of α given ɛ > 0 is illustrated in Figure 1.

Figure 1: 
An illustration of the allowable range of α given ɛ > 0, as described in Remark 3.5. The shaded region represents the allowable range of 0 < α < 1 for a given choice of ɛ > 0.
Figure 1:

An illustration of the allowable range of α given ɛ > 0, as described in Remark 3.5. The shaded region represents the allowable range of 0 < α < 1 for a given choice of ɛ > 0.

Remark 3.6.

If a1, then

a L 1 + ε ε = 1

for all ɛ > 0, and so, the lower bound on ‖u in Lemma 3.4 simplifies to

u 2 1 + p + p min ρ 1 p , ρ 1 p + .

We note that, technically, Lemma 3.4 does not permit p + = 1 since then (1 + ɛ)p + ≤ 1̸ for every ɛ > 0. However, if a L ( 0,1 ) , then as a corollary to Lemma 3.4 we have the following result. And Corollary 3.7 does admit the case p + = 1. Note that a1 is a model case and satisfies a L ( 0,1 ) .

Corollary 3.7.

Assume that each of conditions (H1) and (H3) holds. In addition, assume that a L ( 0,1 ) . If u V ̂ ρ , for some ρ > 0, and u′ is nonincreasing, then

u min 1 2 1 / p a 1 p ρ 1 p , 1 2 1 / p + a 1 p + ρ 1 p + .

Proof.

Since a L ( 0,1 ) and u V ̂ ρ , it follows that

ρ = a | u | p ( ) ( 1 ) a 1 | u | p ( ) ( 1 ) = a 0 1 u ( s ) p ( s ) d s a { s : | u ( s ) | 1 } u ( s ) p d s + { s : | u ( s ) | > 1 } u ( s ) p + d s a u p + u p + .

Thus,

u min 1 2 1 / p a 1 p ρ 1 p , 1 2 1 / p + a 1 p + ρ 1 p + ,

as claimed.□

Remark 3.8.

Note that Corollary 3.7 gives a slightly better lower bound than Lemma 3.4 in case a 1 L ( 0,1 ) since 2 1 p > 2 1 + p + p . Note, furthermore, that Corollary 3.7 does not apply to the fractional derivative kernel a ( t ) = 1 Γ ( α ) t α 1 , 0 < α < 1, since in this case a L ( 0,1 ) .

With our preliminary localisation results dispatched, we now present our nonexistence results. The first of these, namely Theorem 3.9, applies in case 0 < p(x) ≤ 1, whereas the second of these, namely Corollary 3.10, applies in case p(x) > 1. We remark that in each case we assume tacitly that the number λ 0 is finite. Obviously, if it is not, then the nonexistence results are not applicable.

Theorem 3.9.

Assume that each of conditions (H1) and (H3) holds. In addition, assume that there are constants c 0 > 0 and r > 1 such that

f ( t , u ) c 0 u r ,

for each ( t , u ) 1 4 , 3 4 × [ 0 , + ) . Define λ 0 by

λ 0 4 r c 0 max t [ 0,1 ] 1 4 3 4 G ( t , s ) d s 1 sup ρ : M ( ρ ) > 0 M ( ρ ) 2 1 + p + p min ρ 1 p a L 1 + ε ε 1 p , ρ 1 p + a L 1 + ε ε 1 p + 1 r .

Then for each λ > λ 0 problem (1.2) and (1.3) has no positive solution.

Proof.

Suppose that u 0 is a positive solution to problem (1.2) and (1.3). We endeavour to demonstrate that inevitably we are led to a contradiction. Note that since u 00, it follows that there exists ρ 0 > 0 such that

(3.11) u 0 V ̂ ρ 0 .

Indeed, were this not the case, i.e., were u 0 V ̂ 0 , then

(3.12) 0 = a u 0 p ( ) ( 1 ) = 0 1 a ( 1 t ) u 0 ( s ) p ( s ) d s .

Since by assumption a(t) > 0, a.e. t ∈ [0, 1], it would follow from (3.12) that

(3.13) u 0 ( t ) 0 ,

using that u 0 C [ 0,1 ] . But then (3.13) would imply that u 0(t) ≡ 0 in light of the fact that u 0(0) = 0. And this would contradict the assumption that u 0 is a nontrivial solution. Consequently, we conclude that ρ 0 ≠ 0.

Now, let us first suppose that u 0 V ̂ ρ 0 is such that

(3.14) M ρ 0 > 0 .

Then on account of (3.14) it follows that a solution of problem (1.2) and (1.3) may be realised in the operator equation form

(3.15) u 0 T u 0 ,

where T is exactly as in Section 2. Note that due to (3.15) and the definition of T, it follows that u′ ≡ (Tu)′ is nonincreasing on [0,1]. Now, recalling the growth condition

f ( t , u ) c 0 u r

imposed on f in the statement of the theorem, from both (3.11) and (3.15) it follows that

(3.16) u 0 u 0 ( t ) = T u 0 ( t ) = λ 0 1 M ρ 0 1 G ( t , s ) f s , u 0 ( s ) d s λ M ρ 0 1 4 3 4 G ( t , s ) f s , u 0 ( s ) d s λ c 0 M ρ 0 1 4 3 4 G ( t , s ) u 0 ( s ) r d s λ c 0 4 r M ρ 0 u 0 r 1 4 3 4 G ( t , s ) d s

for each t ∈ [0, 1], where we have used the fact (see Erbe and Wang [57]) that

min t 1 4 , 3 4 u 0 ( t ) = min t 1 4 , 3 4 T u 0 ( t ) 1 4 T u 0 = 1 4 u 0 .

Notice that the map

t 1 4 3 4 G ( t , s ) d s

is a continuous function defined on the compact set [0,1]. Consequently, it must attain a maximum value, which is readily seen to be a positive number. Let t 0 ∈ (0, 1) be the value at which it attains its maximum. Then since inequality (3.16) must hold for all t ∈ [0, 1], including at t = t 0, we have that

(3.17) u 0 λ c 0 4 r M ρ 0 u 0 r 1 4 3 4 G t 0 , s d s .

Observe that inequality (3.17) is true if and only if

(3.18) u 1 r λ c 0 4 r M ρ 0 1 4 3 4 G t 0 , s d s .

Now, recalling that 0 < p(x) ≤ 1, Lemma 3.4 implies that

(3.19) u 0 2 1 + p + p min ρ 1 p a L 1 + ε ε 1 p , ρ 1 p + a L 1 + ε ε 1 p + .

Then, recalling that 1 − r < 0, from inequality (3.19) we see that

(3.20) u 1 r 2 1 + p + p min ρ 1 p a L 1 + ε ε 1 p , ρ 1 p + a L 1 + ε ε 1 p + 1 r .

However, at the same time, the definition of λ 0 implies that

(3.21) λ c 0 4 r M ρ 0 1 4 3 4 G t 0 , s d s > c 0 4 r M ρ 0 1 4 3 4 G t 0 , s d s λ 0 c 0 4 r M ρ 0 1 4 3 4 G t 0 , s d s × 4 r c 0 max t [ 0,1 ] 1 4 3 4 G ( t , s ) d s 1 sup ρ : M ( ρ ) > 0 M ( ρ ) 2 1 + p + p min ρ 1 p a L 1 + ε ε 1 p , ρ 1 p + a L 1 + ε ε 1 p + 1 r = λ 0 1 M ρ 0 M ρ 0 2 1 + p + p min ρ 0 1 p a L 1 + ε ε 1 p , ρ 0 1 p + a L 1 + ε ε 1 p + 1 r = 2 1 + p + p min ρ 0 1 p a L 1 + ε ε 1 p , ρ 0 1 p + a L 1 + ε ε 1 p + 1 r

whenever λ > λ 0. Thus, putting inequalities (3.18), (3.20), and (3.21) together we conclude that whenever λ > λ 0 it is the case that

(3.22) 2 1 + p + p min ρ 0 1 p a L 1 + ε ε 1 p , ρ 0 1 p + a L 1 + ε ε 1 p + 1 r u 1 r λ c 0 4 r M ρ 0 1 4 3 4 G t 0 , s d s > 2 1 + p + p min ρ 0 1 p a L 1 + ε ε 1 p , ρ 0 1 p + a L 1 + ε ε 1 p + 1 r .

And since (3.22) is a contradiction, we conclude that if u 0 is a positive solution of (1.2) and (1.3) such that (3.14) holds, then an absurdity is reached, and so, the problem cannot admit a positive solution in this case.

Next we suppose that

(3.23) A ρ 0 = 0 .

This case, unlike the previous one, proceeds in a manner similar to [38], [39], [40]. Letting u 0 represent a fictitious positive solution of (1.2) and (1.3), in this case, by means of the growth condition on f, the differential equation (1.2) degenerates to

0 = λ f t , u 0 ( t ) λ c 0 u 0 ( t ) ,

which implies that u 0(t) ≤ 0, t ∈ [0, 1]. As this contradicts the assumption that u 0 was a positive solution, we conclude that in case (3.23), problem (1.2) and (1.3) cannot have a positive solution.

Finally, we suppose that

(3.24) A ρ 0 < 0 .

This case also proceeds in a manner similar to [38], [39], [40]. In particular, since A ρ 0 0 , a fictitious positive solution u 0 of problem (1.2) and (1.3) once again can be realised as a fixed point of the operator T utilised in Section 2. But then we see at once that

u 0 ( t ) = T u 0 ( t ) = λ 0 1 A ρ 0 1 < 0 G ( t , s ) f s , u 0 ( s ) d s < 0 ,

which again contradicts the assumption that u 0 is a positive solution. Consequently, we conclude that in case (3.24) holds, problem (1.2) and (1.3) cannot have a positive solution.

Now, since the preceding cases are exhaustive, it follows that problem (1.2) and (1.3) cannot have a positive solution whenever λ > λ 0. And this completes the proof of the theorem.□

Our second nonexistence result, which may be considered as a corollary to Theorem 3.9, addresses nonexistence in the case p(x) > 1; i.e., the setting considered in Section 2.

Corollary 3.10.

Assume that conditions (H1)–(H2) hold except that

a m inf t ( 0,1 ] a ( t ) > 0 .

In addition, assume that there are constants c 0 > 0 and 0 < r < 1 such that

f ( t , u ) c 0 u r ,

for each ( t , u ) 1 4 , 3 4 × [ 0 , + ) . Define λ 0 by

λ 0 4 r c 0 max t [ 0,1 ] 1 4 3 4 G ( t , s ) d s 1 sup ρ : M ( ρ ) > 0 M ( ρ ) 2 p + 1 + 2 p + 1 a m ρ 1 r p .

Then for each λ > λ 0 problem (1.2) and (1.3) has no positive solution.

Proof.

Since the majority of the proof is identical to that of Theorem 3.9, we will only state here the part that differs. In particular, the dissimilarity occurs in inequalities (3.19) and (3.20). Since now both 1 − r > 0 and p(x) > 1, from Lemma 3.1 it follows that

u 2 p + 1 + 2 p + 1 a m ρ 1 p

so that

(3.25) u 1 r 2 p + 1 + 2 p + 1 a m ρ 1 r p .

Now, let t 0 have the same meaning as in the proof of Theorem 3.9. Then, imitating (3.21) and (3.22), but with the value of λ 0 as in the statement of the corollary, we conclude that whenever λ > λ 0 it follows that

2 p + 1 + 2 p + 1 a m ρ 1 r p u 1 r λ c 0 4 r M ρ 0 1 4 3 4 G t 0 , s d s > c 0 4 r M ρ 0 1 4 3 4 G t 0 , s d s × 4 r c 0 max t [ 0,1 ] 1 4 3 4 G ( t , s ) d s 1 sup ρ : M ( ρ ) > 0 M ( ρ ) 2 p + 1 + 2 p + 1 a m ρ 1 r p 2 p + 1 + 2 p + 1 a m ρ 1 r p ,

which is a contradiction. Since the remainder of the proof is no different than the corresponding parts of the proof of Theorem 3.9, this completes the proof of the corollary.□

If a1, then the statement of the preceding two results can be simplified slightly. In addition, since

(3.26) max t [ 0,1 ] 1 4 3 4 G ( t , s ) d s = 3 32 ,

both Theorem 3.9 and Corollary 3.10 may be rewritten in light of (3.26). Consequently, we obtain the following corollaries. Note that in the case of Corollary 3.11, we use Corollary 3.7 in light of Remark 3.8.

Corollary 3.11.

Assume that each of conditions (H1) and (H3) holds except that a1. In addition, assume that there are constants c 0 > 0 and r > 1 such that

f ( t , u ) c 0 u r ,

for each ( t , u ) 1 4 , 3 4 × [ 0 , + ) . Define λ 0 by

λ 0 2 5 + 2 r 3 c 0 sup ρ : M ( ρ ) > 0 M ( ρ ) min 1 2 1 / p ρ 1 p , 1 2 1 / 2 + ρ 1 p + 1 r .

Then for each λ > λ 0 problem (1.2) and (1.3) has no positive solution.

Corollary 3.12.

Assume that conditions (H1)–(H2) hold except that a1. In addition, assume that there are constants c 0 > 0 and 0 < r < 1 such that

f ( t , u ) c 0 u r ,

for each ( t , u ) 1 4 , 3 4 × [ 0 , + ) . Define λ 0 by

λ 0 2 5 + 2 r 3 c 0 sup ρ : M ( ρ ) > 0 M ( ρ ) 2 p + 1 + 2 p + 1 ρ 1 r p .

Then for each λ > λ 0 problem (1.2) and (1.3) has no positive solution.

We next provide an example in order to illustrate the application of Corollary 3.11. Both the choice of M, which is the nonlocal coefficient, and the choice of p, which is the exponent function, in Example 3.13 are merely to illustrate the applicability of the nonexistence result. In particular, notice that

lim inf t + M ( t ) = .

As mentioned in Section 1, there remains a relative paucity of results applying to nonpositive nonlocal elements in the current literature. So, we emphasise that our results do apply in this setting – even the nonexistence results.

Example 3.13.

Define the nonlocal coefficient M by

M ( t ) 3 t 3 ,  0 t 1 t 4 3 sin ( π t ) ,  t > 1 .

Let us suppose the nonlinearity f is such that

f ( t , u ) u 2

so that c 0 = 1 and r = 2. Finally, define the exponent function p by

p ( t ) 1 2 + 1 4 cos 2 π t

so that

p 1 4 p ( t ) 3 4 p + .

Now, since M(t) ≤ 0 when 0 ≤ t ≤ 1, the supremum in the statement of Corollary 3.11 will ignore the behaviour of M(ρ) for 0 ≤ ρ ≤ 1. In light of this, we calculate, for ρ ≥ 1,

(3.27) M ( ρ ) min 1 2 1 / p ρ 1 p , 1 2 1 / 2 + ρ 1 p + 1 r = M ( ρ ) min 2 1 4 ρ 4 , 2 3 4 ρ 4 3 1 = ρ 4 3 sin ( π ρ ) 2 3 4 ρ 4 3 = 2 3 4 sin π ρ .

Consequently, on account of (3.27) we calculate

λ 0 = 2 5 + 2 ( 2 ) 3 c 0 sup ρ : M ( ρ ) > 0 M ( ρ ) min 1 2 1 / p ρ 1 p , 1 2 1 / 2 + ρ 1 p + 1 r = 512 3 2 3 4 = 1 3 2 39 4 .

Thus, Corollary 3.11 asserts that whenever

λ > 1 3 2 39 4 287.026

problem (1.2) and (1.3) has no positive solution.

We conclude this section and the paper with one final norm localisation result. This result establishes an upper bound on ‖u provided both that u V ̂ ρ , for some ρ > 0, and that ‖u′‖ ≤ 1. Crucially, whilst this lemma, Lemma 3.14, is in the same vein as Lemma 2.8 from Section 2, this new norm localisation result does not use Sobolev’s inequality, and so, applies in case 0 < p(x) ≤ 1. We have relegated this result to the end of the section since we do not make use of the lemma in the present paper. However, it may be of value in future studies of problem (1.2). And, evidently, it is related to the focus of the first half of this section.

Lemma 3.14.

Assume that each of conditions (H1) and (H3) holds except that

a m inf t ( 0,1 ] a ( t ) > 0 .

If both

  1. u V ̂ ρ for some ρ > 0; and

  2. u′‖ ≤ 1;

then

u 1 + ρ a m .

Proof.

Suppose that u V ̂ ρ . Then

(3.28) 0 1 u ( s ) p d s = { s : | u ( s ) | 1 } u ( s ) p d s + { s : | u ( s ) | > 1 } u ( s ) p d s 1 + 0 1 u ( s ) p ( s ) d s 1 + 1 a m 0 1 a ( 1 s ) u ( s ) p ( s ) d s .

In addition, following the beginning of the proof of Lemma 3.1, we see that

(3.29) u 0 1 u ( s ) d s 0 1 u ( s ) p d s ,

using both the assumption ‖u′‖ ≤ 1 and the assumption p ≤ 1 so that

u ( s ) u ( s ) p ,  s [ 0,1 ] .

Consequently, combining inequalities (3.28) and (3.29) yields

u 0 1 u ( s ) p d s 1 + 1 a m 0 1 a ( 1 s ) u ( s ) p ( s ) d s = 1 + 1 a m a | u | p ( ) ( 1 ) = 1 + ρ a m ,

as desired.□


Corresponding author: Christopher S. Goodrich, School of Mathematics and Statistics, UNSW Sydney, Sydney, NSW 2052, Australia, E-mail: 

I dedicate this paper to the memory of my beloved Maddie Goodrich (16 March 2002–16 March 2020).


Acknowledgments

I gratefully acknowledge the careful reading by the three anonymous referees. I would like to thank, in particular, the handling editor who encouraged me to improve certain aspects of the paper and to make the results more comprehensive.

  1. Research ethics: Not applicable.

  2. Informed consent: Not applicable.

  3. Author contributions: The author has accepted responsibility for the entire content of this manuscript and approved its submission.

  4. Use of Large Language Models, AI and Machine Learning Tools: None declared.

  5. Conflict of interest: The author states no conflict of interest.

  6. Research funding: None declared.

  7. Data availability: None declared.

References

[1] A. Borhanifar, M. A. Ragusa, and S. Valizadeh, “High-order numerical method for two-dimensional Riesz space fractional advection-dispersion equation,” Discrete Contin. Dyn. Syst. Ser. B, vol. 26, no. 10, pp. 5495–5508, 2021. https://doi.org/10.3934/dcdsb.2020355.Search in Google Scholar

[2] K. Q. Lan, “Equivalence of higher order linear Riemann-Liouville fractional differential and integral equations,” Proc. Am. Math. Soc., vol. 148, no. 12, pp. 5225–5234, 2020. https://doi.org/10.1090/proc/15169.Search in Google Scholar

[3] K. Q. Lan, “Compactness of Riemann-Liouville fractional integral operators,” Electron. J. Qual. Theory Differ. Equ., no. 84, p. 15, 2020. https://doi.org/10.14232/ejqtde.2020.1.84.Search in Google Scholar

[4] I. Podlubny, Fractional Differential Equations, New York, Academic Press, 1999.Search in Google Scholar

[5] J. R. L. Webb, “Initial value problems for Caputo fractional equations with singular nonlinearities,” Electr. J. Differ. Equ., vol. 2019, no. 117, p. 32, 2019.Search in Google Scholar

[6] C. S. Goodrich, M. A. Ragusa, and A. Scapellato, “Partial regularity of solutions to p(x)-Laplacian PDEs with discontinuous coefficients,” J. Differ. Equ., vol. 268, no. 9, pp. 5440–5468, 2020.10.1016/j.jde.2019.11.026Search in Google Scholar

[7] V. V. Zhikov, “On some variational problems,” Russ. J. Math. Phys., vol. 5, no. 1, pp. 105–116, 1997.Search in Google Scholar

[8] A. Coscia and G. Mingione, “Hölder continuity of the gradient of p(x)-harmonic mappings,” C. R. Acad. Sci. Paris Ser. I Math. vol. 328, no. 4, pp. 363–368, 1999. https://doi.org/10.1016/s0764-4442(99)80226-2.Search in Google Scholar

[9] K. Fey and M. Foss, “Morrey regularity for almost minimizers of asymptotically convex functionals with nonstandard growth,” Forum Math., vol. 25, no. 5, pp. 887–929, 2013.Search in Google Scholar

[10] M. Ragusa and A. Tachikawa, “On interior regularity of minimizers of p(x)-energy functionals,” Nonlinear Anal., vol. 93, pp. 162–167, 2013. https://doi.org/10.1016/j.na.2013.07.023.Search in Google Scholar

[11] M. Ragusa and A. Tachikawa, “Boundary regularity of minimizers of p(x)-energy functionals,” Ann. Inst. H. Poincaré Anal. Non Linéaire, vol. 33, no. 2, pp. 451–476, 2016. https://doi.org/10.1016/j.anihpc.2014.11.003.Search in Google Scholar

[12] K. Usuba, “Partial regularity of minimizers of p(x)-growth functionals with 1 < p(x) < 2,” Bull. Lond. Math. Soc., vol. 47, no. 3, pp. 455–472, 2015. https://doi.org/10.1112/blms/bdv013.Search in Google Scholar

[13] L. Diening, P. Herjulehto, P. Hästö, and M. Růžička, Lebesgue and Sobolev Spaces with Variable Exponents, Heidelberg, Springer, 2011.10.1007/978-3-642-18363-8Search in Google Scholar

[14] C. S. Goodrich, “p(x)-growth in nonlocal differential equations with convolution coefficients,” Adv. Differ. Equ., vol. 30, nos. 3–4, pp. 115–140, 2025. https://doi.org/10.57262/ade030-0304-115.Search in Google Scholar

[15] C. S. Goodrich, “An application of Sobolev’s inequality to one-dimensional Kirchhoff equations,” J. Differ. Equ., vol. 385, pp. 463–486, 2024. https://doi.org/10.1016/j.jde.2023.12.035.Search in Google Scholar

[16] C. O. Alves and D.-P. Covei, “Existence of solution for a class of nonlocal elliptic problem via sub-supersolution method,” Nonlinear Anal.: Real World Appl., vol. 23, pp. 1–8, 2015. https://doi.org/10.1016/j.nonrwa.2014.11.003.Search in Google Scholar

[17] F. J. S. A. Corrêa, “On positive solutions of nonlocal and nonvariational elliptic problems,” Nonlinear Anal., vol. 59, no. 7, pp. 1147–1155, 2004. https://doi.org/10.1016/s0362-546x(04)00322-0.Search in Google Scholar

[18] F. J. S. A. Corrêa, S. D. B. Menezes, and J. Ferreira, “On a class of problems involving a nonlocal operator,” Appl. Math. Comput., vol. 147, no. 2, pp. 475–489, 2004. https://doi.org/10.1016/s0096-3003(02)00740-3.Search in Google Scholar

[19] J. M. do Ó, S. Lorca, J. Sánchez, and P. Ubilla, “Positive solutions for some nonlocal and nonvariational elliptic systems,” Complex Var. Elliptic Equ., vol. 61, no. 3, pp. 297–314, 2016. https://doi.org/10.1080/17476933.2015.1064404.Search in Google Scholar

[20] C. S. Goodrich, “A topological approach to nonlocal elliptic partial differential equations on an annulus,” Math. Nachr., vol. 294, no. 2, pp. 286–309, 2021. https://doi.org/10.1002/mana.201900204.Search in Google Scholar

[21] R. Stańczy, “Nonlocal elliptic equations,” Nonlinear Anal., vol. 47, no. 5, pp. 3579–3584, 2001. https://doi.org/10.1016/s0362-546x(01)00478-3.Search in Google Scholar

[22] Y. Wang, F. Wang, and Y. An, “Existence and multiplicity of positive solutions for a nonlocal differential equation,” Bound. Value Probl., vol. 2011, p. 5, 2011. https://doi.org/10.1186/1687-2770-2011-5.Search in Google Scholar

[23] B. Yan and T. Ma, “The existence and multiplicity of positive solutions for a class of nonlocal elliptic problems,” Bound. Value Probl., vol. 2016, p. 165, 2016. https://doi.org/10.1186/s13661-016-0670-z.Search in Google Scholar

[24] B. Yan and D. Wang, “The multiplicity of positive solutions for a class of nonlocal elliptic problem,” J. Math. Anal. Appl., vol. 442, no. 1, pp. 72–102, 2016. https://doi.org/10.1016/j.jmaa.2016.04.023.Search in Google Scholar

[25] G. A. Afrouzi, N. T. Chung, and S. Shakeri, “Existence and non-existence results for nonlocal elliptic systems via sub-supersolution method,” Funkcial. Ekvac., vol. 59, no. 3, pp. 303–313, 2016. https://doi.org/10.1619/fesi.59.303.Search in Google Scholar

[26] A. Ambrosetti and D. Arcoya, “Positive solutions of elliptic Kirchhoff equations,” Adv. Nonlinear Stud., vol. 17, no. 1, pp. 3–15, 2017. https://doi.org/10.1515/ans-2016-6004.Search in Google Scholar

[27] N. Azzouz and A. Bensedik, “Existence results for an elliptic equation of Kirchhoff-type with changing sign data,” Funkcial. Ekvac., vol. 55, no. 1, pp. 55–66, 2012. https://doi.org/10.1619/fesi.55.55.Search in Google Scholar

[28] S. Boulaaras, “Existence of positive solutions for a new class of Kirchhoff parabolic systems,” Rocky Mt. J. Math., vol. 50, no. 2, pp. 445–454, 2020. https://doi.org/10.1216/rmj.2020.50.445.Search in Google Scholar

[29] S. Boulaaras and R. Guefaifia, “Existence of positive weak solutions for a class of Kirrchoff elliptic systems with multiple parameters,” Math. Methods Appl. Sci., vol. 41, no. 13, pp. 5203–5210, 2018. https://doi.org/10.1002/mma.5071.Search in Google Scholar

[30] M. Delgado, C. Morales-Rodrigo, J. R. Santos Júnior, and A. Suárez, “Non-local degenerate diffusion coefficients break down the components of positive solution,” Adv. Nonlinear Stud., vol. 20, no. 1, pp. 19–30, 2020.10.1515/ans-2019-2046Search in Google Scholar

[31] J. Graef, S. Heidarkhani, and L. Kong, “A variational approach to a Kirchhoff-type problem involving two parameters,” Results Math., vol. 63, nos. 3–4, pp. 877–889, 2013. https://doi.org/10.1007/s00025-012-0238-x.Search in Google Scholar

[32] G. Infante, “Nonzero positive solutions of nonlocal elliptic systems with functional BCs,” J. Elliptic Parabol. Equ., vol. 5, no. 2, pp. 493–505, 2019. https://doi.org/10.1007/s41808-019-00049-6.Search in Google Scholar

[33] G. Infante, “Eigenvalues of elliptic functional differential systems via a Birkhoff-Kellogg type theorem,” Mathematics, vol. 9, p. 4, 2021. https://doi.org/10.3390/math9010004.Search in Google Scholar

[34] J. R. Santos Júnior and G. Siciliano, “Positive solutions for a Kirchhoff problem with a vanishing nonlocal element,” J. Differ. Equ., vol. 265, no. 5, pp. 2034–2043, 2018.10.1016/j.jde.2018.04.027Search in Google Scholar

[35] T. Shibata, “Global and asymptotic behaviors of bifurcation curves of one-dimensional nonlocal elliptic equations,” J. Math. Anal. Appl., vol. 516, no. 2, p. 12, 2022. https://doi.org/10.1016/j.jmaa.2022.126525.Search in Google Scholar

[36] T. Shibata, “Asymptotic behavior of solution curves of nonlocal one-dimensional elliptic equations,” Bound. Value Probl., p. 15, 2022.10.1186/s13661-022-01644-8Search in Google Scholar

[37] T. Shibata, “Exact solutions and bifurcation curves of nonlocal elliptic equations with convolutional Kirchhoff functions,” Bound. Value Probl., p. 13, 2024.10.1186/s13661-024-01871-1Search in Google Scholar

[38] C. S. Goodrich, “Nonexistence and parameter range estimates for convolution differential equations,” Proc. Am. Math. Soc. Ser. B, vol. 9, pp. 254–265, 2022. https://doi.org/10.1090/bproc/130.Search in Google Scholar

[39] C. S. Goodrich, “A surprising property of nonlocal operators: the deregularising effect of nonlocal elements in convolution differential equations,” Adv. Nonlinear Stud., vol. 24, no. 4, pp. 805–818, 2024. https://doi.org/10.1515/ans-2023-0137.Search in Google Scholar

[40] C. S. Goodrich, “Nonexistence of nontrivial solutions to Kirchhoff-like equations,” Proc. Am. Math. Soc. Ser. B, vol. 11, pp. 304–314, 2024. https://doi.org/10.1090/bproc/224.Search in Google Scholar

[41] A. Cabada, G. Infante, and F. Tojo, “Nonzero solutions of perturbed Hammerstein integral equations with deviated arguments and applications,” Topol. Methods Nonlinear Anal., vol. 47, no. 1, pp. 265–287, 2016.10.12775/TMNA.2016.005Search in Google Scholar

[42] F. Cianciaruso, G. Infante, and P. Pietramala, “Solutions of perturbed Hammerstein integral equations with applications,” Nonlinear Anal.: Real World Appl., vol. 33, pp. 317–347, 2017. https://doi.org/10.1016/j.nonrwa.2016.07.004.Search in Google Scholar

[43] F. Cianciaruso, L. Muglia, and P. Pietramala, “Coupled elliptic systems depending on the gradient with nonlocal BCs in exterior domains,” Bound. Value Probl., p. 17, 2020.10.1186/s13661-020-01384-7Search in Google Scholar

[44] G. Infante and P. Pietramala, “A cantilever equation with nonlinear boundary conditions,” Electron. J. Qual. Theory Differ. Equ., no. 15, p. 14, 2009. https://doi.org/10.14232/ejqtde.2009.4.15.Search in Google Scholar

[45] G. Infante, P. Pietramala, and M. Tenuta, “Existence and localization of positive solutions for a nonlocal BVP arising in chemical reactor theory,” Commun. Nonlinear Sci. Numer. Simul., vol. 19, no. 7, pp. 2245–2251, 2014. https://doi.org/10.1016/j.cnsns.2013.11.009.Search in Google Scholar

[46] G. Infante, “Nontrivial solutions of systems of perturbed Hammerstein integral equations with functional terms,” Mathematics, vol. 9, p. 330, 2021. https://doi.org/10.3390/math9040330.Search in Google Scholar

[47] C. S. Goodrich, “A topological approach to a class of one-dimensional Kirchhoff equations,” Proc. Am. Math. Soc. Ser. B, vol. 8, pp. 158–172, 2021. https://doi.org/10.1090/bproc/84.Search in Google Scholar

[48] C. S. Goodrich, “Nonlocal differential equations with concave coefficients of convolution type,” Nonlinear Anal., vol. 211, 2021, Art. no. 112437. https://doi.org/10.1016/j.na.2021.112437.Search in Google Scholar

[49] C. S. Goodrich, “Nonlocal differential equations with convolution coefficients and applications to fractional calculus,” Adv. Nonlinear Stud., vol. 21, no. 4, pp. 767–787, 2021. https://doi.org/10.1515/ans-2021-2145.Search in Google Scholar

[50] C. S. Goodrich, “Nonlocal differential equations with p-q growth,” Bull. Lond. Math. Soc., vol. 55, no. 3, pp. 1373–1391, 2023. https://doi.org/10.1112/blms.12798.Search in Google Scholar

[51] C. S. Goodrich and C. Lizama, “Existence and monotonicity of nonlocal boundary value problems: the one-dimensional case,” Proc. R. Soc. Edinburgh, Sect. A, vol. 152, no. 1, pp. 1–27, 2022. https://doi.org/10.1017/prm.2020.90.Search in Google Scholar

[52] X. Hao and X. Wang, “Positive solutions of parameter-dependent nonlocal differential equations with convolution coefficients,” Appl. Math. Lett., vol. 153, p. 5, 2024. https://doi.org/10.1016/j.aml.2024.109063.Search in Google Scholar

[53] Q. Song and X. Hao, “Positive solutions for nonlocal differential equations with concave and convex coefficients,” Positivity, vol. 28, no. 5, p. 25, 2024. https://doi.org/10.1007/s11117-024-01086-9.Search in Google Scholar

[54] R. C. Brown, D. B. Hinton, and Š. Schwabik, “Applications of a one-dimensional Sobolev inequality to eigenvalue problems,” Differ. Integral Equ., vol. 9, no. 3, pp. 481–498, 1996. https://doi.org/10.57262/die/1367969967.Search in Google Scholar

[55] G. Talenti, “Best constant in Sobolev inequality,” Ann. Mat. Pura Appl., vol. 110, no. 4, pp. 353–372, 1976. https://doi.org/10.1007/bf02418013.Search in Google Scholar

[56] B. Yan, “Multiple positive solutions for singular boundary-value problems with derivative dependence on finite and infinite intervals,” Electr. J. Differ. Equ., no. 74, p. 25, 2006.Search in Google Scholar

[57] L. H. Erbe and H. Wang, “On the existence of positive solutions of ordinary differential equations,” Proc. Am. Math. Soc., vol. 120, no. 3, pp. 743–748, 1994. https://doi.org/10.1090/s0002-9939-1994-1204373-9.Search in Google Scholar

Received: 2024-11-17
Accepted: 2025-04-11
Published Online: 2025-04-28

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 12.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2023-0185/html
Scroll to top button