Hostname: page-component-788cddb947-w95db Total loading time: 0 Render date: 2024-10-19T18:33:23.960Z Has data issue: true hasContentIssue false

A unified characterization of convolution coefficients in nonlocal differential equations

Published online by Cambridge University Press:  18 September 2024

Christopher S. Goodrich*
Affiliation:
School of Mathematics and Statistics, UNSW Sydney, Sydney, NSW 2052 Australia (c.goodrich@unsw.edu.au)
*
*Corresponding author
Rights & Permissions [Opens in a new window]

Abstract

In loving memory of my beloved miniature dachshund Maddie (16 March 2002 – 16 March 2020). We consider nonlocal differential equations with convolution coefficients of the form

\[{-}M\Big(\big(a*(g\circ |u|)\big)(1)\Big)u''(t)=\lambda f\big(t,u(t)\big),\quad t\in(0,1), \]
in the case in which $g$ can satisfy very generalized growth conditions; in addition, $M$ is allowed to be both sign-changing and vanishing. Existence of at least one positive solution to this equation equipped with boundary data is considered. We demonstrate that the nonlocal coefficient $M$ allows the forcing term $f$ to be free of almost all assumptions other than continuity.

Type
Research Article
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press on behalf of The Royal Society of Edinburgh

1. Introduction

For two $L^1((0,\,+\infty ))$ functions, $a$ and $b$, let $(a*b)(t)$ denote the finite convolution of $a$ and $b$ at some $t\ge 0$ – i.e.,

\[ (a*b)(t):=\int_0^ta(t-s)b(s)\ {\rm d}s,\quad t\ge0. \]

In this paper we consider the following convolution-type differential equation, where $\lambda >0$ is a parameter.

(1.1)\begin{equation} {-}M\Big(\big(a*(g\circ |u|)\big)(1)\Big)u''(t)=\lambda f\big(t,u(t)\big),\quad t\in(0,1) \end{equation}

As will be further clarified in § 2, we assume that $M$ is continuous and possibly both sign-changing and vanishing. Moreover, $a\in L^1((0,\,1))$, which is assumed to be a.e. positive, allows for various nonlocal operators to be captured by the convolutional formulation. For example, if we set

\[ a(t):=\frac{1}{\Gamma(\alpha)}t^{\alpha-1},\quad t>0, \]

where $0<\alpha <1$, then $(a*(g\circ u))(1)$ is the $\alpha$-th order Riemann–Liouville fractional integral of $g\circ u$ at $t=1$. Fractional integrals and derivatives are a well studied class of nonlocal operators – see, for example, [Reference Abbas and Ragusa1, Reference Borhanifar, Ragusa and Valizadeh7, Reference Goodrich19, Reference Goodrich and Lizama32, Reference Goodrich and Lizama33, Reference Lan45, Reference Lan46, Reference Webb53] for some of the research in this area, together with the monographs [Reference Goodrich and Peterson35, Reference Podlubny47].

Our primary contribution herein is to prove the existence of at least one positive solution to (1.1) when subjected to boundary data and, in particular, to do so whilst making the following contributions.

  • We require only that $g$ satisfy the growth bound

    (1.2)\begin{equation} 0\le\xi_1(u)\le g(u)\le\xi_2(u),\quad u\ge0, \end{equation}
    where both $\xi _1$ and $\xi _2$ are strictly increasing functions. Figure 1 illustrates a possible configuration of $\xi _1$, $\xi _2$, and $g$ satisfying (1.2). Notice that this includes the model case, in which $g(u):=u^p$ for some $p>0$. However, this assumption is more general than other recent assumptions. For example, it has been assumed previously [Reference Goodrich29, Reference Goodrich31] that $g$ satisfies $p$-$q$ growth, i.e., $c_1u^p\le g(u)\le c_2+c_3u^q$, which is clearly a special case of the above with $\xi _1(u):=c_1u^p$ and $\xi _2(u):=c_2+c_3u^q$. Similarly, it has been assumed previously [Reference Goodrich24] that $g$ is bounded by convex (or concave) functions. Once again, this is obviously a special case in which one further requires the convexity or concavity of the $\xi _i$ functions. So, the generalization investigated here really gets to the heart of the matter inasmuch as what assumptions on $g$ are necessary – i.e., it is sufficient to assume that $g$ is merely bounded above and below by strictly increasing functions. All other assumptions (e.g., convexity, concavity, particular growth regimes such as polynomial growth) are superfluous.
  • We characterize in a numerically precise way the fact that the forcing term $f$ satisfies essentially no restriction other than continuity provided that $M$ assumes both very small and very large positive values. Whilst this observation is not new per se, for it can be recovered in a general sense even from our original work with Kirchhoff equations, cf., [Reference Goodrich22, Theorem 2.6], in this work we provide a more precise characterization of this phenomenon by utilizing a different growth assumption on $f$. Essentially, we demonstrate that if $M$ is very large at some point and close to zero at another point, then nearly the only assumption needed of $f$ is that it is continuous. This phenomenon is unusual in the theory of boundary value problems (cf., Erbe and Wang [Reference Erbe and Wang18]), though it has been characterized in the context of boundary value problems with nonlocal boundary conditions [Reference Goodrich20, Reference Goodrich21]. In any case, we demonstrate that this phenomenon exists even under the more general condition imposed on the function $g$.

Figure 1. Illustration of the admissible region for the graph of $g$ satisfying condition (1.2).

Let us mention that an important model case of (1.1) occurs when $a(t)\equiv 1$ and $g(u):=u^p$ for $p\ge 1$. In this case, equation (1.1) reduces to

(1.3)\begin{equation} {-}M\Big(\Vert u\Vert_{L^p}^{p}\Big)u''(t)=\lambda f\big(t,u(t)\big). \end{equation}

In case we instead set $\displaystyle a(t)=\frac {1}{\Gamma (\alpha )}t^{\alpha -1}$, $0<\alpha <1$, as mentioned above, then (1.1) reduces to

\[{-}M\Big(\big(I_{0^+}^{\alpha}(g\circ|u|)\big)(1)\Big)u''(t)=\lambda f\big(t,u(t)\big). \]

where by $(I_{0^+}^{\alpha }u)(t)$ we denote the $\alpha$-th order Riemann–Liouville fractional integral of $u$ at $t$.

More generally, nonlocal equations of the form (1.3), or its relatives, have been well studied in recent years. Two model cases seem to have attracted the most attention. One is (1.3) and its PDE equivalent

(1.4)\begin{equation} {-}M\Big(\Vert u\Vert_{L^p}^{p}\Big)\Delta u(\boldsymbol{x})=\lambda f\big(\boldsymbol{x},u(\boldsymbol{x})\big),\quad x\in\Omega\subset\mathbb{R}^{n}, \end{equation}

whereas the other is

(1.5)\begin{equation} {-}M\Big(\Vert u'\Vert_{L^p}^{p}\Big)u''(t)=\lambda f\big(t,u(t)\big),\quad 0< t<1 \end{equation}

and its PDE equivalent

(1.6)\begin{equation} {-}M\Big(\Vert Du\Vert_{L^p}^{p}\Big)\Delta u(\boldsymbol{x})=\lambda f\big(\boldsymbol{x},u(\boldsymbol{x})\big),\quad x\in\Omega\subset\mathbb{R}^{n}. \end{equation}

Each of (1.3)–(1.6) has its origins in the steady-state version of the Kirchhoff-type wave PDE

\[ u_{tt}-M\Big(\Vert Du\Vert_{L^p}^{p}\Big)\Delta u(\boldsymbol{x})=\lambda f\big(\boldsymbol{x},u(\boldsymbol{x})\big),\quad x\in\Omega\subset\mathbb{R}^{n}. \]

Regarding equations of the type (1.3)–(1.4) some recent contributions include papers by Alves and Covei [Reference Alves and Covei3], Corrêa [Reference Corrêa14], Corrêa, Menezes, and Ferreira [Reference Corrêa, Menezes and Ferreira15], do Ó, Lorca, Sánchez, and Ubilla [Reference do Ó, Lorca, Sánchez and Ubilla17], Goodrich [Reference Goodrich22], Stańczy [Reference Stańczy51], Wang, Wang, and An [Reference Wang, Wang and An52], Yan and Ma [Reference Yan and Ma54], and Yan and Wang [Reference Yan and Wang55]. On the other hand, regarding equations of the type (1.5)–(1.6) some recent contributions include papers by Afrouzi, Chung, and Shakeri [Reference Afrouzi, Chung and Shakeri2], Ambrosetti and Arcoya [Reference Ambrosetti and Arcoya4], Azzouz and Bensedik [Reference Azzouz and Bensedik5], Boulaaras [Reference Boulaaras8], Boulaaras and Guefaifia [Reference Boulaaras and Guefaifia9], Chung [Reference Chung12], Delgado, Morales-Rodrigo, Santos Júnior, and Suárez [Reference Delgado, Morales-Rodrigo, Santos Júnior and Suárez16], Graef, Heidarkhani, and Kong [Reference Graef, Heidarkhani and Kong36], Infante [Reference Infante38, Reference Infante39], and Santos Júnior and Siciliano [Reference Santos Júnior and Siciliano48]. In addition to Kirchhoff-like nonlocal differential operators, there is, from a functional analytic viewpoint, a very closely related literature on differential equations equipped with nonlocal boundary operators – see, for example, the papers by Infante, et al. [Reference Biagi, Calamai and Infante6, Reference Cabada, Infante and Tojo10, Reference Cabada, Infante and Tojo11, Reference Cianciaruso, Infante and Pietramala13, Reference Infante37, Reference Infante and Pietramala40Reference Infante, Pietramala and Tenuta44] and Yang [Reference Yang56, Reference Yang57], which in addition to the associated mathematical theory, demonstrate applications to the deformation of a beam under a load, the thermodynamics of a heated filament, and nuclear reactor theory. Additionally, Shibata [Reference Shibata49, Reference Shibata50], along with Goodrich [Reference Goodrich28], has provided some nonexistence results for nonlocal ODEs and nonlocal radially symmetric PDEs of the types mentioned above.

Recently, in the setting of both nonlocal ODEs and nonlocal radially symmetric PDEs, we have developed [Reference Goodrich23, Reference Goodrich25Reference Goodrich27], together with Lizama [Reference Goodrich and Lizama34], a very general methodology for making minimal assumptions of $M$. The methodology utilizes specialized order cones together with topological fixed point theory. An advantage of this methodology is that we are able to make minimal assumptions on the coefficient function $M$. For example, in the study of positive solutions of nonlocal differential equations it is almost always assumed that the nonlocal coefficient $M$ satisfies one of the following three conditions.

  1. (1) $M(t)>0$ for all $t\ge 0$ – see, for example, [Reference Corrêa14, Reference Corrêa, Menezes and Ferreira15, Reference do Ó, Lorca, Sánchez and Ubilla17, Reference Stańczy51, Reference Wang, Wang and An52]

  2. (2) $M(t)$ can only vanish at $0$ or ‘at $+\infty$’ - - see, for example, [Reference Ambrosetti and Arcoya4]

  3. (3) $M(t)>0$ on a neighbourhood of zero – see, for example, [Reference Santos Júnior and Siciliano48]

One can see why such assumptions would be made since if $M(t)=0$, then the differential equation degenerates. Since we are able, by means of our theory, to precisely localize the argument of $M$, i.e., $(a*(g\circ |u|))(1)$, we can avoid making such sweeping assumptions – cf., remark 2.9. Indeed, instead of having to assume that $M(t)>0$ on a pre-specified subset of the real line, our theory simply requires $M(t)$ to be positive somewhere. This is quite different than (1)–(3) above – even than (3), which is the least restrictive of the lot.

So, here we continue the development of this theory by clarifying the generality of the function $g$ and also focussing on the interaction between the behaviour of $M$ and the assumptions required of $f$. And, in particular, we demonstrate that the good aspects of our theory continue to work properly even under the more general assumptions on $g$ utilized herein – not only the minimal assumptions required of $M$, but, furthermore, how $M$ itself can obviate the usual assumptions on $f$.

2. Main result

Throughout this section we denote by $\Vert \cdot \Vert _{\infty }$ the usual maximum norm on $[0,\,1]$, and we will always work within the Banach space $\mathscr {C}([0,\,1])$ equipped with this norm. In addition, we will let $\boldsymbol {1}$ denote the constant function $\boldsymbol {1} : \ \mathbb {R}\rightarrow \{1\}$. Similarly, by $\boldsymbol {0}$ we will denote the constant function $\boldsymbol {0}\ : \mathbb {R}\rightarrow \{0\}$. We will also use the notation

\[ (a*\boldsymbol{1})(c,d):=\int_c^da(1-s)\,{\rm d}s \]

for any $0\le c< d\le 1$.

We next list the assumptions imposed on the various functions appearing in (1.1). In addition, since our approach to studying (1.1) will be via studying the fixed points of an associated Hammerstein integral operator, we will equip (1.1) with boundary data via a Green's function, which we henceforth denote by $G$. The properties of $G$ are listed in condition (H2) below. Observe that condition (H1.2) implies that $(a*\boldsymbol {1})(1)>0$, a fact that will be used in the sequel without explicit mention.

  1. H1: The functions $M\ : \ [0,\,+\infty )\rightarrow \mathbb {R}$, $f\ : \ [0,\,1]\times [0,\,+\infty )\rightarrow [0,\,+\infty )$, $g\ : \ [0,\,+\infty )\rightarrow [0,\,+\infty )$, and $a\ : \ (0,\,1]\rightarrow [0,\,+\infty )$ satisfy the following properties.

    1. (1) Each of $M$, $f$, and $g$ is continuous.

    2. (2) $a\in L^1((0,\,1];[0,\,+\infty ))$ is a.e. positive.

    3. (3) There exist numbers $0<\rho _1<\rho _2$ such that $M(t)>0$ for $t\in [\rho _1,\,\rho _2]$.

    4. (4) There exist strictly increasing continuous functions $\xi _1$, $\xi _2\ : \ [0,\,+\infty )\rightarrow [0,\,+\infty )$ such that

      \begin{equation} \xi_1(u)\le g(u)\le\xi_2(u),\quad u\ge0.\nonumber \end{equation}

  2. H2: The continuous function $G\ : \ [0,\,1]\times [0,\,1]\rightarrow [0,\,+\infty )$ satisfies each of the following.

    1. (1) There exist numbers $0\le c< d\le 1$ and a constant $\eta _0:=\eta _0(c,\,d)\in (0,\,1]$ such that

      \begin{equation} \min_{t\in[c,d]}G(t,s)\ge\eta_0\mathscr{G}(s),\quad s\in[0,1],\nonumber \end{equation}
      where $\mathscr {G}\ : [0,\,1]\rightarrow [0,\,+\infty )$ denotes the function $\displaystyle \mathscr {G}(s):=\max _{t\in [0,1]}G(t,\,s)$.
    2. (2) With $\eta _0$, $c$, and $d$ as in (H2.1), and both $\rho _1$ and $\rho _2$ as in (H1.3), there exist constants $c_1>0$, $c_2\ge 0$, and $c_3>0$ such that

      \begin{align*} & f(t,u)\ge c_1u,\quad (t,u)\in[c,d]\\& \quad \times\left[\eta_0\xi_2^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(1)}\right),\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(c,d)}\right)\right] \end{align*}
      and that
      \[ f(t,u)\le c_2+c_3u,\quad (t,u)\in[0,1]\times\left[0,\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_2}{(a*\boldsymbol{1})(c,d)}\right)\right]. \]

We will study problem (1.1), equipped with suitable boundary data, by means of the operator $T\ : \mathscr {C}([0,\,1])\rightarrow \mathscr {C}([0,\,1])$ defined by

\[ (Tu)(t):=\lambda\int_0^1\left(M\Big(\big(a*(g\circ |u|)\big)(1)\Big)\right)^{{-}1}G(t,s)f\big(s,u(s)\big)\,{\rm d}s. \]

It will be convenient to restrict the domain of $T$ to specialized sets, which allow us to provide precise control over the argument of $M$. Indeed, this is the strategy that permits us to avoid wide ranging assumptions on $M$ such as the uniform positivity of $M$. In particular, we will work within the order cone

\[ \mathscr{K}:=\left\{u\in\mathscr{C}\big([0,1]\big)\ : \ u\ge0\text{ and }\min_{t\in[c,d]}u(t)\ge\eta_0\Vert u\Vert_{\infty}\right\}. \]

Furthermore, for any $\rho \ge 0$, define the set $\widehat {V}_{\rho }\subseteq \mathscr {K}$ by

\[ \widehat{V}_{\rho}:=\Big\{u\in\mathscr{K}\ : \ \big(a*(g\circ |u|)\big)(1)<\rho\Big\}. \]

Observe that $\widehat {V}_{\rho }$ is (relatively) open in $\mathscr {K}$. Crucially, we note that

\[ \partial\widehat{V}_{\rho}:=\Big\{u\in\mathscr{K}\ : \ \big(a*(g\circ |u|)\big)(1)=\rho\Big\}, \]

which gives us very precise control over the argument of $M$. Since whenever $T$ is restricted to a subset of $\mathscr {K}$ it holds that $u\equiv |u|$, henceforth we will omit the absolute value when performing calculations with $T$.

We begin by providing a result that localizes $u$ in either the case $u\in \widehat {V}_{\rho }$ or $u\in \partial \widehat {V}_{\rho }$ for some $\rho >0$. This lemma will be used repeatedly in the sequel. It also establishes that the $\widehat {V}_{\rho }$ set is bounded, with respect to $\Vert \cdot \Vert _{\infty }$, for each $\rho \ge 0$ – a necessary condition for the application of the topological fixed theorem that we employ later.

Lemma 2.1 Suppose that conditions (H1)–(H2) are satisfied. Then for any $\rho >0$ such that

\begin{align*} \xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right)>0, \end{align*}

whenever $u\in \partial \widehat {V}_{\rho }$, it follows that

\begin{align*} \xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right)<\Vert u\Vert_{\infty}<\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right). \end{align*}

In addition, for any $\rho >0$, whenever $u\in \widehat {V}_{\rho }$, it follows that

\begin{align*} \Vert u\Vert_{\infty}<\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right). \end{align*}

Proof. Let us first suppose that $u\in \partial \widehat {V}_{\rho }$ for some $\rho >0$. Then, on the one hand, we calculate

(2.1)\begin{equation} \begin{aligned} \rho=\big(a*(g\circ u)\big)(1) < \big(a*(\xi_2\circ u)\big)(1) & \le\big(a*(\xi_2\circ\Vert u\Vert_{\infty})\boldsymbol{1}\big)(1)\\ & =\xi_2\big(\Vert u\Vert_{\infty}\big)(a*\boldsymbol{1})(1). \end{aligned} \end{equation}

Then using the fact that $\xi _2$ is strictly increasing, it follows from (2.1) that

(2.2)\begin{equation} \Vert u\Vert_{\infty}>\xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right). \end{equation}

On the other hand, we calculate

(2.3)\begin{equation} \begin{aligned} \rho & =\big(a*(g\circ u)\big)(1)>\big(a*(\xi_1\circ u)\big)(1)\\ & \ge\int_c^da(1-s)\xi_1\big(u(s)\big)\ {\rm d}s\\ & \ge\int_c^da(1-s)\xi_1\big(\eta_0\Vert u\Vert_{\infty}\big)\ {\rm d}s\\ & =\xi_1\big(\eta_0\Vert u\Vert_{\infty}\big)(a*\boldsymbol{1})(c,d) \end{aligned} \end{equation}

Then using the fact that $\xi _1$ is strictly increasing, it follows from (2.3) that

(2.4)\begin{equation} \Vert u\Vert_{\infty}<\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right). \end{equation}

And so from both (2.2) and (2.4) we obtain, for any $\rho >0$, the localization estimate

\[ u\in\partial\widehat{V}_{\rho}\Longrightarrow\xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right)<\Vert u\Vert_{\infty}<\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right). \]

Next assume that $u\in \widehat {V}_{\rho }$ for some $\rho >0$. Then, by means of the preceding calculations, we see that the localization

\[ u\in\widehat{V}_{\rho}\Longrightarrow\Vert u\Vert_{\infty}<\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right). \]

holds. And this completes the proof.

Remark 2.2 We wish to emphasize at this juncture that even though the proof of lemma 2.1 is similar to the related results [Reference Goodrich24, Lemma 2.3], [Reference Goodrich26, Lemma 2.4], and [Reference Goodrich31, Lemma 2.8], it, nonetheless, encompasses far greater generality. Indeed, there is no requirement that either $\xi _1$ or $\xi _2$ satisfy any particular type of growth (e.g., polynomial), and there is no requirement that either function satisfy any convexity or concavity assumption. In addition, even in the model case in which $g(u)=u^p$, here the cases $0< p<1$ and $p\ge 1$ are treated in a unified fashion. And this is not something that has been accomplished before, to the best of our knowledge.

Next we prove a technical lemma regarding how large $M(t)$ needs to be in order for a certain inequality to be satisfied. This result will be used in the existence theorem later. Note that in both the statement of lemma 2.3 as well as the sequel we use the following notation:

\[ \overline{G_{[a,b]}}:=\max_{t\in[0,1]}\int_a^bG(t,s)\ {\rm d}s, \]

for any $0\le a< b\le 1$.

Lemma 2.3 Fix $\rho >0$. Assume that each of conditions (H1) and (H2) is satisfied and that $\displaystyle \xi _2^{-1}(\frac {\rho }{(a*\boldsymbol {1})(1)})>0$. If $u\in \partial \widehat {V}_{\rho }$ and

\[ M(\rho)\ge\lambda\overline{G_{[0,1]}}\left(c_2+\frac{c_3}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right)\right)\left(\xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right)\right)^{{-}1}, \]

then

\[ \lambda\big(c_2+c_3\Vert u\Vert_{\infty}\big)\big(M(\rho)\big)^{{-}1}\overline{G_{[0,1]}}\le\Vert u\Vert_{\infty}. \]

Proof. First recall that if $u\in \partial \widehat {V}_{\rho }$, then from lemma 2.1 it follows that

(2.5)\begin{equation} 0<\xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right)<\Vert u\Vert_{\infty}<\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right). \end{equation}

Note that

(2.6)\begin{equation} \lambda\big(c_2+c_3\Vert u\Vert_{\infty}\big)\big(M(\rho)\big)^{{-}1}\overline{G_{[0,1]}}\le\Vert u\Vert_{\infty} \end{equation}

if and only if

(2.7)\begin{equation} M(\rho)\ge\frac{\lambda\left(c_2+c_3\Vert u\Vert_{\infty}\right)\overline{G_{[0,1]}}}{\Vert u\Vert_{\infty}}. \end{equation}

Now, using (2.5) note that

(2.8)\begin{align} & \frac{\lambda\left(c_2+c_3\Vert u\Vert_{\infty}\right)\overline{G_{[0,1]}}}{\Vert u\Vert_{\infty}}\nonumber\\ & \quad<\left(\xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right)\right)^{{-}1}\lambda\left(c_2+c_3\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right)\right)\overline{G_{[0,1]}}\nonumber\\ & \quad =\lambda\overline{G_{[0,1]}}\left(c_2+\frac{c_3}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right)\right)\left(\xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right)\right)^{{-}1}. \end{align}

Then, upon combining (2.8) with (2.6)–(2.7) we see that if

\[ M(\rho)\ge\lambda\overline{G_{[0,1]}}\left(c_2+\frac{c_3}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right)\right)\left(\xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right)\right)^{{-}1}, \]

then

\[ \lambda\big(c_2+c_3\Vert u\Vert_{\infty}\big)\big(M(\rho)\big)^{{-}1}\overline{G_{[0,1]}}\le\Vert u\Vert_{\infty}, \]

which completes the proof.

Lemma 2.4 Assume that conditions (H1)–(H2) hold. Then $T\ : \overline {\widehat {V}_{\rho _2}}\setminus \widehat {V}_{\rho _1}\rightarrow \mathscr {K}$ is completely continuous and, in particular,

\[ T\left(\overline{\widehat{V}_{\rho_2}}\setminus\widehat{V}_{\rho_1}\right)\subseteq\mathscr{K}. \]

Proof. The proof is similar to part of the proof of [Reference Goodrich20, Theorem 3.1], for example. Therefore, we omit the proof.

We finalize our preliminary lemmata with the following result, known as the Guo-Krasnosel'skiĭ theorem – see, for example, [Reference Zeidler58]. This will be the topological fixed point theorem that we utilize in our existence theorem.

Lemma 2.5 Let $\mathscr {B}$ be a Banach space and let $\mathscr {K}\subseteq \mathscr {B}$ be a cone. Assume that $\Omega _1$ and $\Omega _2$ are bounded open sets contained in $\mathscr {B}$ such that $\boldsymbol {0}\in \Omega _1$ and $\overline {\Omega }_1\subseteq \Omega _2$. Assume, further, that $T\ : \mathscr {K}\cap (\overline {\Omega }_2\setminus \Omega _1)\rightarrow \mathscr {K}$ is a completely continuous operator. If either

  1. (1) $\Vert Ty\Vert \le \Vert y\Vert$ for $y\in \mathscr {K}\cap \partial \Omega _1$ and $\Vert Ty\Vert \ge \Vert y\Vert$ for $y\in \mathscr {K}\cap \partial \Omega _2$; or

  2. (2) $\Vert Ty\Vert \ge \Vert y\Vert$ for $y\in \mathscr {K}\cap \partial \Omega _1$ and $\Vert Ty\Vert \le \Vert y\Vert$ for $y\in \mathscr {K}\cap \partial \Omega _2$;

then $T$ has at least one fixed point in $\mathscr {K}\cap (\overline {\Omega }_2\setminus \Omega _1)$.

We now present our existence theorem.

Theorem 2.6 Assume that each of conditions (H1) and (H2) holds. In addition, assume both that

(2.9)\begin{equation} \frac{c_1\lambda\eta_0\overline{G_{[c,d]}}}{M\left(\rho_1\right)}\ge1 \end{equation}

and that

(2.10)\begin{equation} M(\rho_2)\ge\lambda\overline{G_{[0,1]}}\left(c_2+\frac{c_3}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_2}{(a*\boldsymbol{1})(c,d)}\right)\right)\left(\xi_2^{{-}1}\left(\frac{\rho_2}{(a*\boldsymbol{1})(1)}\right)\right)^{{-}1}. \end{equation}

If

\[ g(0)<\frac{\rho_1}{(a*\boldsymbol{1})(1)}, \]

then (1.1) equipped with the boundary data inherited from the Green's function $G$ has at least one positive solution, say $u_0$, such that

\[ u_0\in\overline{\widehat{V}_{\rho_2}}\setminus\widehat{V}_{\rho_1}. \]

Moreover, $u_0$ satisfies the localization

\[ \xi_2^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(1)}\right)<\Vert u_0\Vert_{\infty}<\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_2}{(a*\boldsymbol{1})(c,d)}\right). \]

Proof. First of all, by lemma 2.4 we note that

\[ T\left(\overline{\widehat{V}_{\rho_2}}\setminus\widehat{V}_{\rho_1}\right)\subseteq\mathscr{K}, \]

where $T$ is completely continuous on its domain. Furthermore, we note that

\[ \boldsymbol{0}\in\widehat{V}_{\rho_1} \]

because

\[ \big(a*(g\circ\boldsymbol{0})\boldsymbol{1}\big)(1)=g(0)(a*\boldsymbol{1})(1)<\rho_1 \]

by assumption. And, in addition, we see that

\[ \overline{\widehat{V}_{\rho_1}}\subseteq\widehat{V}_{\rho_2} \]

owing both to the definition of $\widehat {V}_{\rho }$ and to the fact that $\rho _1<\rho _2$. Finally, by lemma 2.1 the sets $\widehat {V}_{\rho _i}$, $i\in \{1,\,2\}$, are bounded. So, each of the technical conditions in lemma 2.5 is satisfied.

We first demonstrate that for each $u\in \partial \widehat {V}_{\rho _1}$ it follows that

(2.11)\begin{equation} \Vert Tu\Vert_{\infty}\ge\Vert u\Vert_{\infty}; \end{equation}

that is, $T$ is a cone expansion on $\partial \widehat {V}_{\rho _1}$. To this end, first note that since $u\in \partial \widehat {V}_{\rho _1}$ it follows that

(2.12)\begin{equation} \big(a*(g\circ u)\big)(1)=\rho_1. \end{equation}

Next recall that $f$ satisfies the growth estimate

(2.13)\begin{align} f(t,u)\ge c_1u,\quad (t,u)\in[c,d]\times\left[\eta_0\xi_2^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(1)}\right),\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(c,d)}\right)\right]. \end{align}

Note that whenever $u\in \partial \widehat {V}_{\rho _1}$, it follows both that

\[ u(t)\ge\eta_0\Vert u\Vert_{\infty}\ge\eta_0\xi_2^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(1)}\right),\quad t\in[a,b] \]

and that

\[ u(t)\le\Vert u\Vert_{\infty}\le\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(c,d)}\right),\quad t\in[0,1]. \]

Consequently, the growth estimate (2.13) is satisfied for any $u\in \partial \widehat {V}_{\rho _1}$. Then it follows from a combination of both (2.12) and (2.13) that, for each $t\in [0,\,1]$,

(2.14)\begin{equation} \begin{aligned} (Tu)(t) & =\lambda\int_0^1\big(M\left(\rho_1\right)\big)^{{-}1}G(t,s)f\big(s,u(s)\big)\ {\rm d}s\\ & \ge\frac{\lambda}{M\left(\rho_1\right)}\int_0^1G(t,s)c_1u(s)\ {\rm d}s\\ & \ge\frac{c_1\lambda}{M\left(\rho_1\right)}\int_{c}^{d}G(t,s)\eta_0\Vert u\Vert_{\infty}\ {\rm d}s\\ & \ge\left(\frac{c_1\lambda\eta_0}{M\left(\rho_1\right)}\int_c^dG(t,s)\ {\rm d}s\right)\Vert u\Vert_{\infty}. \end{aligned} \end{equation}

Now taking the maximum over $t\in [0,\,1]$ on both sides of (2.14) yields

(2.15)\begin{equation} \Vert Tu\Vert_{\infty}\ge\left(\frac{c_1\lambda\eta_0}{M\left(\rho_1\right)}\max_{t\in[0,1]}\int_c^dG(t,s)\ {\rm d}s\right)\Vert u\Vert_{\infty}=\frac{c_1\lambda\eta_0\overline{G_{[c,d]}}}{M\left(\rho_1\right)}\Vert u\Vert_{\infty}. \end{equation}

Finally, using assumption (2.9) in the statement of the theorem, we conclude from inequality (2.15) that

(2.16)\begin{equation} \Vert Tu\Vert_{\infty}\ge\underbrace{\frac{c_1\lambda\eta_0\overline{G_{[a,b]}}}{M\left(\rho_1\right)}}_{\ge1}\Vert u\Vert_{\infty}\ge\Vert u\Vert_{\infty}. \end{equation}

So, inequality (2.16) implies that $T$ is a cone expansion on $\partial \widehat {V}_{\rho _1}$ – that is, inequality (2.11) is satisfied.

We next demonstrate that for each $u\in \partial \widehat {V}_{\rho _2}$ it follows that

(2.17)\begin{equation} \Vert Tu\Vert_{\infty}\le\Vert u\Vert_{\infty}; \end{equation}

that is, $T$ is a cone compression on $\partial \widehat {V}_{\rho _2}$. To this end, first note that since $u\in \partial \widehat {V}_{\rho _2}$ it follows that

(2.18)\begin{equation} \big(a*(g\circ u)\big)(1)=\rho_2. \end{equation}

Next recall that $f$ satisfies the growth estimate

(2.19)\begin{equation} f(t,u)\le c_2+c_3u,\quad (t,u)\in[0,1]\times\left[0,\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_2}{(a*\boldsymbol{1})(c,d)}\right)\right]. \end{equation}

Similar to the first part of the proof, whenever $u\in \partial \widehat {V}_{\rho _2}$ it follows that

(2.20)\begin{equation} u(t)\le\Vert u\Vert_{\infty}\le\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_2}{(a*\boldsymbol{1})(c,d)}\right),\quad t\in[0,1], \end{equation}

where we, once again, have used lemma 2.1; in other words, condition (H2) implies that for each $u\in \partial \widehat {V}_{\rho _2}$ it follows that $f(t,\,u(t))\le c_2+c_3u(t)$, $t\in [0,\,1]$. Then from (2.18)–(2.20), for each $t\in [0,\,1]$, we deduce that

(2.21)\begin{equation} \begin{aligned} (Tu)(t) & =\lambda\int_0^1\big(M\left(\rho_2\right)\big)^{{-}1}G(t,s)f\big(s,u(s)\big)\ {\rm d}s\\ & \le\frac{\lambda}{M\left(\rho_2\right)}\int_0^1G(t,s)\big(c_2+c_3u(s)\big)\ {\rm d}s\\ & \le\frac{\lambda}{M\left(\rho_2\right)}\int_0^1G(t,s)\big(c_2+c_3\Vert u\Vert_{\infty}\big)\ {\rm d}s. \end{aligned} \end{equation}

Taking the maximum over $t\in [0,\,1]$ on both sides of inequality (2.21) yields

(2.22)\begin{equation} \Vert Tu\Vert_{\infty}\le\frac{\lambda}{M\left(\rho_2\right)}\max_{t\in[0,1]}\int_0^1G(t,s)\big(c_2+c_3\Vert u\Vert_{\infty}\big)\ {\rm d}s=\frac{\lambda \overline{G_{[0,1]}}}{M\left(\rho_2\right)}\big(c_2+c_3\Vert u\Vert_{\infty}\big). \end{equation}

Finally, an application of lemma 2.3 to inequality (2.22), keeping in mind assumption (2.10), implies that

(2.23)\begin{equation} \Vert Tu\Vert_{\infty}\le\frac{\lambda \overline{G_{[0,1]}}}{M\left(\rho_2\right)}\big(c_2+c_3\Vert u\Vert_{\infty}\big)\le\Vert u\Vert_{\infty}. \end{equation}

Thus, (2.23) implies the desired inequality (2.17).

All in all, then, by lemma 2.5 we deduce the existence of

\[ u_0\in\overline{\widehat{V}_{\rho_2}}\setminus\widehat{V}_{\rho_1} \]

such that $Tu_0\equiv u_0$. And this function $u_0$ is, therefore, a positive solution of (1.1) equipped with the boundary data inherited from $G$. Finally, the conclusion of lemma 2.1 implies the localization

\[ \xi_2^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(1)}\right)<\Vert u_0\Vert_{\infty}<\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_2}{(a*\boldsymbol{1})(c,d)}\right). \]

And this completes the proof.

Remark 2.7 Let us consider what conditions (2.9)–(2.10) imply regarding the constants $c_1$, $c_2$, and $c_3$ appearing in the growth condition (H2.2) imposed on the forcing function $f$ in (1.1). Figure 2 provides an idealized drawing of the lower and upper bounding functions for the graph of $f$. In the drawing, the numbers $\alpha _1$, $\alpha _2$, and $\alpha _3$ are defined by

\begin{equation} \alpha_1:=\eta_0\xi_2^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(1)}\right)\nonumber \end{equation}

and

\begin{equation} \alpha_i:=\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_i}{(a*\boldsymbol{1})(c,d)}\right),\quad i\in\{2,3\}.\nonumber \end{equation}

That is, the $\alpha _i$'s are the bounds on the $u$ variable for which $(t,\,u)\mapsto f(t,\,u)$ satisfies the various growth restrictions in condition (H2.2).

Figure 2. The light shaded region shows where the graph of $f$ can live. The dark shaded region is the set $E_0$ referenced in remark 2.7. The numbers $\alpha _i$, $i\in \{1,\,2,\,3\}$, are defined in remark 2.7.

Note first that (2.9) is equivalent to

\[ c_1\ge\frac{M\left(\rho_1\right)}{\lambda\eta_0\overline{G_{[c,d]}}} \]

so that as $M(\rho _1)\to 0^+$, it follows that the lower bound on $c_1$ tends to $0$. More precisely and as in figure 2, define $E_0\subset [0,\,+\infty )$ by

\begin{align*} E_0 & :=\left\{(u,v)\in\mathbb{R}^2\ : \ \eta_0\xi_2^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(1)}\right)\right.\nonumber\\ & \left.\qquad\ \le u\le\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_1}{(a*\boldsymbol{1})(c,d)}\right),\quad 0\le v\le c_1u\right\}. \end{align*}

Then, denoting by $m(E_0)$ the Lebesgue measure of the set $E_0$, we see that

\[ \lim_{M\left(\rho_1\right)\to0^+}m\left(E_0\right)=0. \]

In other words, as $M(\rho _1)$ tends to zero, the restriction $f(t,\,u)\ge c_1u$ is obviated.

On the other hand, (2.10) is equivalent to

\[ M(\rho_2)\ge\lambda\eta_0\overline{G_{[0,1]}}\left(c_2+\frac{c_3}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(c,d)}\right)\right)\left(\xi_2^{{-}1}\left(\frac{\rho}{(a*\boldsymbol{1})(1)}\right)\right)^{{-}1}. \]

So, in a similar way, this implies that if there exists $\rho _2$ such that $M(\rho _2)\gg 1$, then $c_2$ and $c_3$ can be very large, thus implying that the upper bound on $f$ is very mild in this case. Indeed, in terms of the drawing in figure 2, both the slope and the $y$-intercept of the line $u\mapsto c_2+c_3u$ will tend to $+\infty$ as $M(\rho _2)\to +\infty$, which means that the upper bound on $f$ becomes less and less restrictive. Thus, in a simplified sense, we see that if $M$ is alternatively very large somewhere and very close to zero somewhere, then the restrictions on $f$ are obviated, and so, there are then essentially no restrictions on $f$ other than continuity (and nonnegativity).

We conclude with an example in order to clarify the application of theorem 2.6. The example will demonstrate how the nonlocal coefficient $M$ can eliminate nearly all restrictions other than continuity and nonnegativity from $f$.

Example 2.8 Let

\[ g(t):=t+e^t\sin^2{t} \]

and

\[ M(t):=\begin{cases} -3+t,\quad & 0\le t\le 3\\ (t-3)(5-t),\quad & 3\le t\le 5\\ - (t-5)^2,\quad & t\ge5\end{cases}. \]

In addition, set $a\equiv \boldsymbol {1}$ and

\[ G(t,s):=\begin{cases} t(1-s),\quad & 0\le t\le s\le 1\\ s(1-t),\quad & 0\le s\le t\le 1\end{cases}. \]

Then $G$ equips (1.1) with Dirichlet boundary conditions so that we are considering the problem

(2.24)\begin{equation} \begin{aligned} -M\big(\boldsymbol{1}*(g\circ |u|)(1)\big)u''(t) & =\lambda f\big(t,u(t)\big),\quad 0< t<1\\ u(0) & =0\\ u(1) & =0. \end{aligned} \end{equation}

Now, one can show that

\[ \xi_1(t):=t\le g(t)\le e^{2t}=:\xi_2(t). \]

Observe that $\xi _1^{-1}(t)=t$ and $\xi _2^{-1}(t)=\ln {\sqrt {t}}$. In addition, for the Green's function $G$ it is known (see Erbe and Wang [Reference Erbe and Wang18], for example) that one may choose $\displaystyle c:=\frac {1}{4}$, $\displaystyle d:=\frac {3}{4}$, and $\displaystyle \eta _0=\frac {1}{4}$. Then we calculate

\[ \overline{G_{[c,d]}}=\max_{t\in[0,1]}\int_{\frac{1}{4}}^{\frac{3}{4}}G(t,s)\ {\rm d}s=\frac{15}{32} \]

and

\[ \overline{G_{[0,1]}}=\max_{t\in[0,1]}\int_0^1G(t,s)\ {\rm d}s=\frac{1}{2}. \]

In addition, since $a\equiv \boldsymbol {1}$, it follows that

\[ (a*\boldsymbol{1})(1)=(\boldsymbol{1}*\boldsymbol{1})(1)=1 \]

and that

\[ (a*\boldsymbol{1})(c,d)=(\boldsymbol{1}*\boldsymbol{1})\left(\frac{1}{4},\frac{3}{4}\right):=\int_{\frac{1}{4}}^{\frac{3}{4}}\ {\rm d}s=\frac{1}{2}. \]

Finally, set, for $0<\varepsilon _1<1$,

\[ \rho_1:=3+\varepsilon_1\nonumber \]

and

\[ \rho_2:=4. \]

Note that

\[ g(0)=0<\frac{\rho_1}{(\boldsymbol{1}*\boldsymbol{1})(1)}=\rho_1. \]

Then condition (2.9) is satisfied provided that

(2.25)\begin{equation} \varepsilon_1\left(2-\varepsilon_1\right)=M\left(3+\varepsilon_1\right)\le\frac{15}{128}c_1\lambda, \end{equation}

whereas condition (2.10) is satisfied provided that

\begin{align*} & 1=M(4)=M\left(\rho_2\right)\\ & \quad\ge\lambda\overline{G_{[0,1]}}\left(c_2+\frac{c_3}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_2}{(a*\boldsymbol{1})(c,d)}\right)\right)\left(\xi_2^{{-}1}\left(\frac{\rho_2}{(a*\boldsymbol{1})(1)}\right)\right)^{{-}1}\\ & \quad=\frac{1}{2}\lambda\big(c_2+32c_3\big)\frac{1}{\ln{2}}\\ & \quad=\frac{1}{\ln{4}}\lambda\big(c_2+32c_3\big), \end{align*}

which is equivalent to

(2.26)\begin{equation} c_2+32c_3\le\frac{\ln{4}}{\lambda}. \end{equation}

Now, since

\[ \lim_{\varepsilon_1\to0^+}\varepsilon_1\left(2-\varepsilon_1\right)=0, \]

it follows that inequality (2.25) can be satisfied for any $\lambda$ and $c_1$ provided that $\varepsilon _1$ is chosen sufficiently close to $0$. So, given any forcing term $f$ satisfying both (H1)–(H2) and inequality (2.26), there exists $\varepsilon _1>0$ sufficiently small such that by theorem 2.6 problem (2.24) admits a positive solution, say

\[ u_0\in\overline{\widehat{V}_{4}}\setminus\widehat{V}_{3+\varepsilon_1}, \]

where $u_0$ satisfies the localization

\[ \ln{\sqrt{3+\varepsilon_1}}=\xi_2^{{-}1}\left(\frac{\rho_1}{(\boldsymbol{1}*\boldsymbol{1})(1)}\right)<\Vert u_0\Vert_{\infty}<\frac{1}{\eta_0}\xi_1^{{-}1}\left(\frac{\rho_2}{(\boldsymbol{1}*\boldsymbol{1})\left(\frac{1}{4},\frac{3}{4}\right)}\right)=32. \]

Finally, observe since

\[ \lim_{\lambda\to0^+}\frac{\ln{4}}{\lambda}={+}\infty, \]

it follows from condition (2.25) that any $f\in \mathscr {C}([0,\,+\infty );[0,\,+\infty ))$ there exists $\lambda _0>0$ sufficiently small such that for each $\lambda \in (0,\,\lambda _0)$ problem (2.24) admits a positive solution. In other words, there is no growth restriction on $f$.

Remark 2.9 Note that the function $g$ in example 2.8

  1. (1) alternates between concave and convex;

  2. (2) alternates between increasing and decreasing; and

  3. (3) does not satisfy

    \[ g(u)\le c_2+c_3u^q, \]
    for any $1\le q<+\infty$, $c_2\ge 0$, and $c_3>0$, seeing as $g$ grows exponentially.

This is seen by the graph of $g$, which is provided in figure 3. Observation (1) implies that the results of [Reference Goodrich24, Reference Goodrich30] cannot be applied. Observation (2) also implies that the results of [Reference Goodrich24, Reference Goodrich30] cannot be applied. And observation (3) implies that the results of [Reference Goodrich22, Reference Goodrich25Reference Goodrich27, Reference Goodrich29, Reference Goodrich31] cannot be applied. Moreover, and as discussed in § 1, other earlier results in the ODEs setting, such as [Reference Corrêa14, Reference Corrêa, Menezes and Ferreira15, Reference do Ó, Lorca, Sánchez and Ubilla17, Reference Stańczy51, Reference Wang, Wang and An52], cannot be applied due both to the sign-changing and vanishing nature of the nonlocal coefficient $M$ as defined in (2.24); in particular, both $\displaystyle \lim _{t\to \infty }M(t)=-\infty$ and $M(0)<0$ in contrast to the restrictions imposed (albeit in the PDEs setting) in [Reference Ambrosetti and Arcoya4] and [Reference Santos Júnior and Siciliano48], respectively. Therefore, the results presented herein are genuinely more broadly applicable than those previously reported in the literature.

Figure 3. Illustration of the graphs of $\xi _1$, $\xi _2$, and $g$ in example 2.8. The shaded region is the area bounded between the graphs of $\xi _1$ and $\xi _2$ – i.e., the admissible region for the graph of $g$.

References

Abbas, M. I. and Ragusa, M. A.. On the hybrid fractional differential equations with fractional proportional derivatives of a function with respect to a certain function. Symmetry 13 (2021), 264.Google Scholar
Afrouzi, G. A., Chung, N. T. and Shakeri, S.. Existence and non-existence results for nonlocal elliptic systems via sub-supersolution method. Funkcial. Ekvac. 59 (2016), 303313.Google Scholar
Alves, C. O. and Covei, D.-P.. Existence of solution for a class of nonlocal elliptic problem via sub-supersolution method. Nonlinear Anal. Real World Appl. 23 (2015), 18.Google Scholar
Ambrosetti, A. and Arcoya, D.. Positive solutions of elliptic Kirchhoff equations. Adv. Nonlinear Stud. 17 (2017), 315.Google Scholar
Azzouz, N. and Bensedik, A.. Existence results for an elliptic equation of Kirchhoff-type with changing sign data. Funkcial. Ekvac. 55 (2012), 5566.Google Scholar
Biagi, S., Calamai, A. and Infante, G.. Nonzero positive solutions of elliptic systems with gradient dependence and functional BCs. Adv. Nonlinear Stud. 20 (2020), 911931.Google Scholar
Borhanifar, A., Ragusa, M. A. and Valizadeh, S.. High-order numerical method for two-dimensional Riesz space fractional advection-dispersion equation. Discrete Cont. Dyn. Syst. Series B 26 (2021), 54955508.Google Scholar
Boulaaras, S.. Existence of positive solutions for a new class of Kirchhoff parabolic systems. Rocky Mountain J. Math. 50 (2020), 445454.Google Scholar
Boulaaras, S. and Guefaifia, R.. Existence of positive weak solutions for a class of Kirrchoff elliptic systems with multiple parameters. Math. Methods Appl. Sci. 41 (2018), 52035210.Google Scholar
Cabada, A., Infante, G. and Tojo, F.. Nonzero solutions of perturbed Hammerstein integral equations with deviated arguments and applications. Topol. Methods Nonlinear Anal. 47 (2016), 265287.Google Scholar
Cabada, A., Infante, G. and Tojo, F. A. F.. Nonlinear perturbed integral equations related to nonlocal boundary value problems. Fixed Point Theory 19 (2018), 6592.Google Scholar
Chung, N. T.. Existence of positive solutions for a class of Kirchhoff type systems involving critical exponents. Filomat 33 (2019), 267280.Google Scholar
Cianciaruso, F., Infante, G. and Pietramala, P.. Solutions of perturbed Hammerstein integral equations with applications. Nonlinear Anal. Real World Appl. 33 (2017), 317347.Google Scholar
Corrêa, F. J. S. A.. On positive solutions of nonlocal and nonvariational elliptic problems. Nonlinear Anal. 59 (2004), 11471155.Google Scholar
Corrêa, F. J. S. A., Menezes, S. D. B. and Ferreira, J.. On a class of problems involving a nonlocal operator. Appl. Math. Comput. 147 (2004), 475489.Google Scholar
Delgado, M., Morales-Rodrigo, C., Santos Júnior, J. R. and Suárez, A.. Non-local degenerate diffusion coefficients break down the components of positive solution. Adv. Nonlinear Stud. 20 (2020), 1930.Google Scholar
do Ó, J. M., Lorca, S., Sánchez, J. and Ubilla, P.. Positive solutions for some nonlocal and nonvariational elliptic systems. Complex Var. Elliptic Equ. 61 (2016), 297314.Google Scholar
Erbe, L. H. and Wang, H.. On the existence of positive solutions of ordinary differential equations. Proc. Amer. Math. Soc. 120 (1994), 743748.Google Scholar
Goodrich, C. S.. Existence of a positive solution to a class of fractional differential equations. Appl. Math. Lett. 23 (2010), 10501055.Google Scholar
Goodrich, C. S.. New Harnack inequalities and existence theorems for radially symmetric solutions of elliptic PDEs with sign changing or vanishing Green's function. J. Differ. Equ. 264 (2018), 236262.Google Scholar
Goodrich, C. S.. Radially symmetric solutions of elliptic PDEs with uniformly negative weight. Ann. Mat. Pura Appl. (4) 197 (2018), 15851611.Google Scholar
Goodrich, C. S.. A topological approach to nonlocal elliptic partial differential equations on an annulus. Math. Nachr. 294 (2021), 286309.Google Scholar
Goodrich, C. S.. A topological approach to a class of one-dimensional Kirchhoff equations. Proc. Amer. Math. Soc. Ser. B 8 (2021), 158172.Google Scholar
Goodrich, C. S.. Nonlocal differential equations with concave coefficients of convolution type. Nonlinear Anal. 211 (2021), 112437.Google Scholar
Goodrich, C. S.. Differential equations with multiple sign changing convolution coefficients. Internat. J. Math. 32 (2021), 2150057.Google Scholar
Goodrich, C. S.. Nonlocal differential equations with convolution coefficients and applications to fractional calculus. Adv. Nonlinear Stud. 21 (2021), 767787.Google Scholar
Goodrich, C. S.. A one-dimensional Kirchhoff equation with generalized convolution coefficients. J. Fixed Point Theory Appl. 23 (2021), 73.Google Scholar
Goodrich, C. S.. Nonexistence and parameter range estimates for convolution differential equations. Proc. Amer. Math. Soc. Ser. B 9 (2022), 254265.Google Scholar
Goodrich, C. S.. Nonlocal differential equations with $p$-$q$ growth. Bull. Lond. Math. Soc. 55 (2023), 13731391.Google Scholar
Goodrich, C. S.. Nonlocal differential equations with convex convolution coefficients. J. Fixed Point Theory Appl. 25 (2023), 4.Google Scholar
Goodrich, C. S.. An application of Sobolev's inequality to one-dimensional Kirchhoff equations. J. Differ. Equ. 385 (2024), 463486.Google Scholar
Goodrich, C. S. and Lizama, C.. A transference principle for nonlocal operators using a convolutional approach: Fractional monotonicity and convexity. Israel J. Math. 236 (2020), 533589.Google Scholar
Goodrich, C. S. and Lizama, C.. Positivity, monotonicity, and convexity for convolution operators. Discrete Contin. Dyn. Syst. Series A. 40 (2020), 49614983.Google Scholar
Goodrich, C. S. and Lizama, C.. Existence and monotonicity of nonlocal boundary value problems: the one-dimensional case. Proc. Roy. Soc. Edinburgh Sect. A 152 (2022), 127.Google Scholar
Goodrich, C. S. and Peterson, A. C.. Discrete Fractional Calculus (Cham, Springer International Publishing, 2015).Google Scholar
Graef, J., Heidarkhani, S. and Kong, L.. A variational approach to a Kirchhoff-type problem involving two parameters. Results. Math. 63 (2013), 877889.Google Scholar
Infante, G.. Positive solutions of some nonlinear BVPs involving singularities and integral BCs. Discrete Contin. Dyn. Syst. Ser. S 1 (2008), 99106.Google Scholar
Infante, G.. Nonzero positive solutions of nonlocal elliptic systems with functional BCs. J. Elliptic Parabol. Equ. 5 (2019), 493505.Google Scholar
Infante, G.. Eigenvalues of elliptic functional differential systems via a Birkhoff–Kellogg type theorem. Mathematics 9 (2021), 4.Google Scholar
Infante, G. and Pietramala, P.. A cantilever equation with nonlinear boundary conditions. Electron. J. Qual. Theory Differ. Equ. (2009), 14.Google Scholar
Infante, G. and Pietramala, P.. A third order boundary value problem subject to nonlinear boundary conditions. Math. Bohem. 135 (2010), 113121.Google Scholar
Infante, G. and Pietramala, P.. Multiple nonnegative solutions of systems with coupled nonlinear boundary conditions. Math. Methods Appl. Sci. 37 (2014), 20802090.Google Scholar
Infante, G. and Pietramala, P.. Nonzero radial solutions for a class of elliptic systems with nonlocal BCs on annular domains. NoDEA Nonlinear Differ. Equ. Appl. 22 (2015), 9791003.Google Scholar
Infante, G., Pietramala, P. and Tenuta, M.. Existence and localization of positive solutions for a nonlocal BVP arising in chemical reactor theory. Commun. Nonlinear Sci. Numer. Simul., 19 (2014), 22452251.Google Scholar
Lan, K. Q.. Equivalence of higher order linear Riemann–Liouville fractional differential and integral equations. Proc. Amer. Math. Soc. 148 (2020), 52255234.Google Scholar
Lan, K. Q.. Compactness of Riemann–Liouville fractional integral operators. Electron. J. Qual. Theory Differ. Equ. 84 (2020), 15.Google Scholar
Podlubny, I.. Fractional Differential Equations (New York, Academic Press, 1999).Google Scholar
Santos Júnior, J. R. and Siciliano, G.. Positive solutions for a Kirchhoff problem with a vanishing nonlocal element. J. Differ.l Equ. 265 (2018), 20342043.Google Scholar
Shibata, T.. Global and asymptotic behaviors of bifurcation curves of one-dimensional nonlocal elliptic equations. J. Math. Anal. Appl. 516 (2022), 126525.Google Scholar
Shibata, T.. Asymptotic behavior of solution curves of nonlocal one-dimensional elliptic equations. Bound. Value Probl. (2022), 63.Google Scholar
Stańczy, R.. Nonlocal elliptic equations. Nonlinear Anal. 47 (2001), 35793584.Google Scholar
Wang, Y., Wang, F. and An, Y.. Existence and multiplicity of positive solutions for a nonlocal differential equation. Bound. Value Probl. (2011), 5.Google Scholar
Webb, J. R. L.. Initial value problems for Caputo fractional equations with singular nonlinearities. Electron. J. Differ. Equ. (2019), 117.Google Scholar
Yan, B. and Ma, T.. The existence and multiplicity of positive solutions for a class of nonlocal elliptic problems. Bound. Value Probl. (2016), 165.Google Scholar
Yan, B. and Wang, D.. The multiplicity of positive solutions for a class of nonlocal elliptic problem. J. Math. Anal. Appl. 442 (2016), 72102.Google Scholar
Yang, Z.. Positive solutions to a system of second-order nonlocal boundary value problems. Nonlinear Anal. 62 (2005), 12511265.Google Scholar
Yang, Z.. Positive solutions of a second-order integral boundary value problem. J. Math. Anal. Appl. 321 (2006), 751765.Google Scholar
Zeidler, E.. Nonlinear Functional Analysis and Its Applications I: Fixed-Point Theorems (New York, Springer, 1986).Google Scholar
Figure 0

Figure 1. Illustration of the admissible region for the graph of $g$ satisfying condition (1.2).

Figure 1

Figure 2. The light shaded region shows where the graph of $f$ can live. The dark shaded region is the set $E_0$ referenced in remark 2.7. The numbers $\alpha _i$, $i\in \{1,\,2,\,3\}$, are defined in remark 2.7.

Figure 2

Figure 3. Illustration of the graphs of $\xi _1$, $\xi _2$, and $g$ in example 2.8. The shaded region is the area bounded between the graphs of $\xi _1$ and $\xi _2$ – i.e., the admissible region for the graph of $g$.