1 Introduction

Let X be a nonempty set and \({\mathbb {R}}_{+}\) be the set of all nonnegative real numbers. A semimetric is a distance function \(d:X\,{ \times }\, X \,{ \rightarrow }\, {\mathbb {R}}_{+}\) that satisfies two axioms: \((d_{1}) \): \(d(x,y)\,{ = }\,0\) if and only if \(x\,{ = }\,y\); \((d_{2}) \): \(d(x,y)\,{ = }\, d(y,x)\) for all \(x,y\,{ \in }\, X\). It is well known that by adding the triangle inequality to the axioms of d it becomes continuous. In 1993, Czerwik [8] investigated a semimetric called b-metric, which satisfies the inequality: \(d(x,y) \le b(d(x,z)+d(z,y))\), where b is a constant in \([1,+\infty )\) and \(x,y,z \,{\in }\, X\). This notion has been studied previously by different authors, for the latest and rather complete bibliography, we refer the reader to the surveys of Berinde and Păcurar [2] and Karapınar [14]. Despite the b-metric is very useful in applications [7, 15, 27], it has a major drawback due to its lack of continuity [26]. In order to overcome this limitation, Kirk and Shahzad proposed a slight modification in the third axiom, see [16, 17].

The suprametric, which was introduced by the author in [3], is a semimetric that fulfill \(d(x,y) \le d(x,z) + d(z,y) + \rho d(x,z)d(z,y)\), where \(\rho \) is a constant in \({\mathbb {R}}_{+}\) and \(x,y,z \,{\in }\, X\). This distance function is very useful to construct projective metrics of Thompson’s type [25], and to prove the existence of solutions to various classes of integral and matrix equations. Very recently, the suprametric has been utilized by Panda et al. [21, 22] to analyze complex valued fractional order neural networks and the existence of a solution of stochastic integral equations.

It is known that the space of p-integrable functions for \(p\in (0,1)\) is a b-metric space, but it is unclear whether it is a suprametric space. The author introduced the b-suprametric [4], which subsume such functional space. Note that the distance function of the b-suprametric is not necessarily continuous [4, Example 2.10], although the continuity is very useful. In order to overcome this drawback here we introduce the strong b-suprametric distance function, a subfamily of the b-suprametric, and shows its continuity.

The objectives of this work are fourfold: (1) To introduce the strong b-suprametric space in which we establish fixed point theorems of Banach and Edelstein types. (2) To prove a variational principle through the Cantor’s intersection theorem, then to derive a Caristi fixed point result via this variational principle. (3) To introduce the strong b-supranormed linear space and to provide the fixed point principles of Brouwer and Schauder in such linear space. (4) To provide new sufficient conditions for the existence of a solution to an integral equation, via a Chebyshev type inequality, where the integral operator involved is not necessarily Lipschitzian with respect to a metric. Then, we show the existence of a unique solution to a third-order boundary value problem.

2 Strong b-Suprametric Spaces

Here and below, the symbols \({\mathbb {R}}\) and \({\mathbb {N}}\) will denote respectively the set of all real numbers and all nonnegative natural numbers. The symbol \(\textrm{cl}(A)\) stands for the closure of a set A. We first need to recall the b-suprametric spaces from [4].

Definition 1.1

Let (Xd) be a semimetric space and \(b\,{\ge }\,1\), \(\rho \,{\ge }\,0\) be two real constants. The function d is called b-suprametric if:

(\(d_{3}\)) \(d(x,y)\le b \,(d(x,z)+d(z,y))+\rho d(x,z)d(z,y)\) for all \(x,y,z\in X\).

A pair (Xd) is called b-suprametric space if X is a nonempty set and d is a b-suprametric.

In the previous definition, if \(\rho \,{ = }\,0\) we obtain the b-metric [8] and if \(b\,{ = }\,1\) we obtain the suprametric [3]. In the sequel, we focus on the following subclass of b-suprametric space.

Definition 1.2

Let (Xd) be a semimetric space and \(b\,{\ge }\,1\), \(\rho \,{\ge }\,0\) be two real constants. The function d is called strong b-suprametric (sb-suprametric space) if:

(\(d'_{3}\)):

\(d(x,y)\le b \, d(x,z)+d(z,y)+\rho d(x,z)d(z,y)\) for all \(x, y, z \in X\).

A pair (Xd) is called sb-suprametric space if X is a nonempty set and d is an sb-suprametric.

Remark 1.3

From \((d_{2})\), it follows that we also have

\(d''_{3}\):

\(d(x,y)\le d(x,z)+b \, d(z,y)+\rho d(x,z)d(z,y)\), for all \(x, y, z \in X\).

Examples 1.4

  • All suprametric spaces of [3] are sb-suprametric spaces.

  • Let \(X=\{1,2,3\}\) and let \(d:X\times X \rightarrow {\mathbb {R}}_{+}\) be a function defined by:

    $$\begin{aligned} d(x,y)=\left\{ \begin{array}{ll} 0,& \quad x=y,\\ 3,& \quad (x,y)\in \big \{(1,2),(2,1)\big \},\\ \frac{1}{2},& \quad \text {otherwise}. \end{array}\right. \end{aligned}$$

    Then (Xd) is an sb-suprametric space with coefficient \(b=\textstyle \frac{3}{2}\) and \(\rho =8\).

  • Let \(X=C_{+}[0,1]\) of continuous nonnegative functions endowed with

    $$\begin{aligned} \delta (x,y)=\sup _{t\in [0,1]}|x(t)-y(t)|(|x(t)-y(t)|+\textstyle \frac{1}{2}),\;\text { for all }x,y\in X. \end{aligned}$$

    Then \((X, \delta )\) is an sb-suprametric space for \(b=\rho =2\).

Proposition 1.5

Let (Xd) be an sb-suprametric space, then for all \(p, q, s, t\,{\in }\,X\)

$$\begin{aligned} \frac{\rho (d(p,q)-d(s,t))^{2}}{(b+\rho d(p,q)+\rho d(s,t))^{2}}\le 2\big (d(p,t)+d(s,q)+\rho \, d(p,t)\, d(s,q)\big ). \end{aligned}$$
(1)

Proof

Let (Xd) be an \(sb\)-suprametric space with \(\rho \,{>}\,0\) (the case \(\rho \,{=}\,0\) is trivial). Then,

$$\begin{aligned} d(p,q)&\le d(p,s)+b\, d(s,q)+\rho \, d(p,s)\, d(s,q)\\&\le d(s,t)+b\, d(p,t)+\rho \, d(s,t) \, d(p,t)+ b \, d(s,q)\\&\quad +\rho \, \big (d(s,t)+b \, d(p,t)+\rho \, d(s,t) \, d(p,t)\big )\, d(s,q)\\&\le d(s,t)+\big (b+\rho \, d(s,t)\big ) \big (d(p,t)+ d(s,q)+\rho \, d(p,t)\, d(s,q)\big ), \end{aligned}$$

which implies

$$\begin{aligned} \frac{d(p,q)-d(s,t)}{(b+\rho \, d(s,t))}\le d(p,t)+ d(s,q)+\rho \, d(p,t)\, d(s,q). \end{aligned}$$
(2)

A similar argument shows that

$$\begin{aligned} \frac{d(s,t)-d(p,q)}{(b+\rho \, d(p,q))}\le d(p,t)+ d(s,q)+\rho \, d(p,t)\, d(s,q). \end{aligned}$$
(3)

Adding (2) to (3), we obtain

$$\begin{aligned} \frac{\rho (d(p,q)-d(s,t))^{2}}{(b+\rho d(p,q))(b+\rho d(s,t))}\le 2\big (d(p,t)+d(s,q)+\rho \, d(p,t)\, d(s,q)\big ), \end{aligned}$$

which implies (1). \(\square \)

Remark 1.6

Let \(u\,{ \ge }\,0\) and \(\rho \,{ > }\,0\). Assume that a sequence \(\{u_{n}\}\subset {\mathbb {R}}_{+}\) satisfies

$$\begin{aligned} \lim _{n\rightarrow \infty } \frac{\rho \,(u_{n}-u)^{2}}{(b+\rho \,u_{n}+\rho \,u)^{2}}=0, \end{aligned}$$

then \(u_{n}\) tends to u as \(n\rightarrow \infty \). Otherwise, if \(u_{n}\) does not tends to u there exists \(\varepsilon >0\) such that for all integer \(k>0\), \(n(k)>k\) and \(|u_{n(k)}-u|>\varepsilon \). Then,

$$\begin{aligned} \frac{\sqrt{\rho }\,\varepsilon }{b+\rho \,u_{n(k)}+\rho \,u} <\frac{\sqrt{\rho }\,|u_{n(k)}-u|}{b+\rho \,u_{n(k)}+\rho \,u}\rightarrow 0\text { as } k\rightarrow \infty , \end{aligned}$$

which implies that \(u_{n(k)}\) tends to infinity as \(k\rightarrow \infty \), and hence

$$\begin{aligned} \lim _{k\rightarrow \infty } \frac{{\rho }\,(u_{n(k)}-u)^{2}}{(b+\rho \,u_{n(k)}+\rho \,u)^{2}}=\frac{1}{\rho }, \end{aligned}$$

yields a contradiction.

Remark 1.7

Let \(\{p_{n}\}\) and \(\{q_{n}\}\) be sequences in X such that \(\lim _{n\rightarrow \infty } d (p_{n}, t) \,{=}\,0\) and \(\lim _{n\rightarrow \infty } d (q_{n}, s)\, {=}\, 0\), then by (1) and Remark 1.6, \( \lim \limits _{n\rightarrow \infty } d (p_{n}, q_{n}) = d (s, t), \) and this means that d is continuous.

Let (Xd) be an sb-suprametric space. An open ball and a closed ball centered at \(a\in X\) and of radius \(r>0\), are respectively given by

$$\begin{aligned} B(a,r)&{:}{=}&\{x\in X: d(a,x)<r\} \text { and } B[a,r]{:}{=}\{x\in X: d(a,x)\le r\}. \end{aligned}$$

Proposition 1.8

Let (Xd) be an sb-suprametric space. Then

  1. (i)

    every open ball is an open set.

  2. (ii)

    every closed ball is a closed set.

Proof

To see (i), let \(r>0\) and \(a\in X\). For \(y\in B(a,r)\) let

$$\begin{aligned} r_{1}&{:}{=}&\frac{r-d(y,a)}{b+\rho d(y,a)}, \end{aligned}$$

then if \(x\in B(y,r_{1})\),

$$\begin{aligned} d(x,a)&\le b \, d(x,y)+ d(y,a)+\rho d(x,y)\, d(y,a)\\&< b \, r_{1}+ d(y,a)+\rho r_{1}\, d(y,a)=r. \end{aligned}$$

Thus, \(B(y,r_{1})\subseteq B(a,r)\), and B(ar) is open.

Now, to see (ii), let \(r\,{ > }\,0\) and \(a\,{ \in }\, X\) and take a sequence \(\{x_{n}\}\) in B[ar] convergent to some x with respect to d. Then

$$\begin{aligned} d(a,x)&\le d(a,x_{n})+ b\,d(x_{n},x)+\rho d(a,x_{n})\, d(x_{n},x)\\&\le r+ (b+\rho r)\, d(x_{n},x), \end{aligned}$$

and as \(n\rightarrow \infty \), we get \(x\in B[a,r]\) which proves that B[ar] is closed. \(\square \)

As a consequence, we obtain the following propositions.

Proposition 1.9

Let (Xd) be an sb-suprametric space. The family of open balls form a base of a topology on X.

Proof

Let \(u\in B(x,\varepsilon ), B(y,\varepsilon ')\) and choose \(r>0\) so that \((b+\rho r)d(x,u)+r<\varepsilon \) and \((b+\rho r)d(y,u)+r<\varepsilon '\). Then by taking a point \(v\in B(u,r)\), we obtain

$$\begin{aligned} d(x,v)\le b d(x,u)+d(u,v)+\rho d(x,u)d(u,v)< (b+\rho r)d(x,u)+r<\varepsilon ,\\ d(y,v)\le b d(y,u)+d(x,u)+\rho d(y,u)d(u,v)<(b+\rho r)d(y,u)+r<\varepsilon ', \end{aligned}$$

which implies that \(B(u,r)\subset B(x,\varepsilon )\cap B(y,\varepsilon ')\). Finally, we conclude by using [10, Lemma I.4.7] and the fact that every \(x\,{ \in }\, X\) is also in \(B(x,\tau )\) for some \(\tau \,{ > }\,0\). \(\square \)

Proposition 1.10

An sb-suprametric space is normal.

Proof

Let (Xd) be an sb-suprametric space. If \(x,y\,{ \in }\, X\) such \(x\,{ \ne }\, y\), then \(U{:}{=}B(x,\frac{d(x,y)}{2})\) and \(V{:}{=}B(y,\frac{d(x,y)}{2\,b+\rho \,d(x,y)})\) are disjoint neighborhoods of x and y respectively. Otherwise, assume that \(U\cap V\ne \varnothing \), so there exists \(z\in U\cap V\). Thus, using that \(d(x,z)<\frac{r}{2}\) and \(d(y,z)<\frac{r}{2\,b+\rho \,r}\) where \(r=d(x,y)\), we obtain

$$\begin{aligned} r=d(x,y)&\le d(x,z)+d(z,y)+\rho \, d(x,z)d(z,y)\\&<\frac{r}{2}+\frac{r}{2\,b+\rho \,r}+\rho \,\frac{r}{2}\,\frac{r}{2\,b+\rho \,r}=r, \end{aligned}$$

a contradiction, so our claim holds. We conclude therefore that X is Hausdorff.

Let now U and V be disjoint closed sets and let

$$\begin{aligned} d(x,U){{:}{=}}\inf \limits _{u\in U}d(x,u) \text { and } d(x,V){{:}{=}}\inf \limits _{v\in V}d(x,v). \end{aligned}$$

Define the sets

$$\begin{aligned} U'{:}{=}\big \{x\in X: d(x,U)<d(x,V)\big \}\text { and } V'{:}{=}\big \{x\in X: d(x,V)<d(x,U)\big \}. \end{aligned}$$

Then \(U'\) and \(V'\) are disjoint neighborhoods of U and V respectively. \(\square \)

Proposition 1.11

In an sb-suprametric space, if a sequence has a limit it is unique.

Proposition 1.12

The strong b-metric space [16] is normal.

Definition 1.13

Let (Xd) be an sb-suprametric space.

  1. (i)

    The sequence \(\{x_{n}\}_{n\in {\mathbb {N}}}\) converges to \(x \in X\) iff \(\lim \limits _{n\rightarrow \infty }d(x_{n},x)\,{ = }\,0\).

  2. (ii)

    The sequence \(\{x_{n}\}_{n\in {\mathbb {N}}}\) is Cauchy iff \(\lim \limits _{n,m\rightarrow \infty }d(x_{n},x_{m})\,{ = }\,0\).

  3. (iii)

    (Xd) is complete iff any Cauchy sequence in X is convergent.

Remark 1.14

Let \(X{=}C_{+}[0,1]\) be the set of continuous functions \(x:[0,1]{\rightarrow } {\mathbb {R}}_{+}\) endowed with \(\delta \) the sb-suprametric of Examples 1.4. The completeness of \((X,\delta )\) follows from that of (Xd) with \(d(x,y){=}\sup \limits _{t\in [0,1]}|x(t){-}y(t)|\) for \(x,y\in X\).

The next lemma is a direct consequence of Definition 1.13.

Lemma 1.15

In an sb-suprametric space, we have:

  1. (i)

    A convergent sequence is a Cauchy sequence.

  2. (ii)

    A Cauchy sequence converges iff it has a convergent subsequence.

  3. (iii)

    A point \(u\in \textrm{cl}(U)\) iff there is a sequence \(\{u_{n}\}\subset U\) converging to u.

Remark 1.16

If a sequence \(\{x_{n}\}_{n\in {\mathbb {N}}}\) is Cauchy in a complete sb-suprametric (Xd), then there exists \(x_{*}\in X\) such that \(\lim \limits _{n\rightarrow \infty }d(x_{n},x_{*})=0\). By \(d'_{3}\) follows that every subsequence \(\{x_{n(k)}\} _{k\in {\mathbb {N}}}\) converges to \(x_{*}\).

Remark 1.17

Let (Xd) be an \(sb\)-suprametric space. By \(d'_{3}\), we have:

$$\begin{aligned} d(x_{0},x_{n}) \le b\max \{1,\rho ^{n}\} \textstyle \sum \limits _{i=1}^{n}{\textbf{e}}_{i}(d_{0},\ldots ,d_{n-1}), \end{aligned}$$

for all \(n \in {\mathbb {N}}\), \(x_{0},\ldots , x_{n} \in X\), where \(d_{i-1}{:}{=}d(x_{i-1},x_{i})\) and \({\textbf{e}}_{i}\) is the ith elementary symmetric polynomial in n variables, that is,

$$\begin{aligned} {\textbf{e}}_{i}(d_{0},\ldots ,d_{n-1})=\textstyle \sum \limits _{0\le j_{1}<j_{2}<\cdots <j_{i}\le n-1}\;\prod \limits _{k=1}^{i}d_{j_{k}}. \end{aligned}$$

It is easy to see that these polynomials possess the following properties:

Proposition 1.18

Let \(n\in {\mathbb {N}}\) and \({\textbf{e}}_{i}(x_{0},\ldots ,x_{n-1})\) be an elementary symmetric polynomial of index \(0\le i\,{\le }\,n\). Then,

  1. (i)

    \(x_{k}\mapsto {\textbf{e}}_{i}(x_{0},\ldots ,x_{k},\ldots ,x_{n-1})\) is a nondecreasing function for \(0\,{\le }\,k\,{\le }\,n\).

  2. (ii)

    \({\textbf{e}}_{i}(ax_{0},\ldots ,ax_{n-1}) = a^{i} {\textbf{e}}_{i}(x_{0},\ldots ,x_{n-1})\) for all \(a\in {\mathbb {R}}_{+}\).

A covering of a set U in X is a family of open sets whose union contains U. A set \(U \subseteq X\) is called compact if and only if every covering of U by open sets in X contains a finite sub-covering. A subset U of a topological space X is sequentially compact, if every sequence of points in U has a subsequence converging to a point of X. A set \(N \subset X\) is called an \(\varepsilon \)-net for a set \(U \subseteq X\) (\(\varepsilon >0\)), if there exists \(x_{\varepsilon }\in N\) for every \(x \in U\) such that \(d(x, x_{\varepsilon }) < \varepsilon \). Next, we provide an extreme value theorem in sb-suprametric spaces.

Theorem 1.19

Let (Xd) be an sb-suprametric space. Let U be a compact subset of X and \(f:U\rightarrow {\mathbb {R}}\) be a continuous function. Then,

  1. (i)

    f is bounded on U,

  2. (ii)

    f attains its supremum and its infimum.

Proof

The proof is similar to that of [18, Chap. 5. Theorem 1] (see also [10, Lemma I.5.8]). \(\square \)

The compactness is discussed in the rest of this section.

Theorem 1.20

For a set U in an sb-suprametric space X to be compact, it is necessary, and in the case of completeness of X, sufficient that there is a finite \(\varepsilon \)-net for the set U for every \(\varepsilon > 0\).

Proof

The proof is similar to that of [18, Chap. 5. Theorem 3], except for the value of the distance between any two points, which does not exceed \(\varepsilon _{n}\left( 1+b+\rho \varepsilon _{n}\right) \). \(\square \)

Corollary 1.21

A subset U of an sb-suprametric space is compact if and only if it is closed and sequentially compact.

Proof

The proof is exactly similar to that of [10, Theorem I.6.13]. \(\square \)

Corollary 1.22

Let (Xd) be a sb-suprametric space and \(U\subseteq X\). If U is compact, then it is bounded.

Proof

Let \(S_{n}=\{x_{1},\ldots ,x_{n}\}\) be a 1-net for U. Let \(a\in X\), \(x\in X\) and \(x_{i}\in S_{n}\) for \(i=1,\ldots ,n\). Then,

$$\begin{aligned} d(x,a)&\le b d(x,x_{i})+d(x_{i},a)+\rho d(x,x_{i})\, d(x_{i},a) \\&\le b \big (d(x,x_{i})+d(x_{i},a)\big )+\textstyle \frac{\rho }{2} \big (d(x,x_{i})+ d(x_{i},a)\big )^{2}\\&\le (b+\textstyle \frac{\rho }{2}) \big (1+\max _{i}d(x_{i},a)\big )^{2}<\infty . \end{aligned}$$

\(\square \)

3 Banach and Edelstein Fixed Point Theorems

We start by presenting a Banach fixed point result in sb-suprametric spaces.

Theorem 2.1

Let (Xd) be a complete sb-suprametric space and \(f:X\rightarrow X\) be a given mapping. Assume that there exists \(c\in [0,b^{-1})\) such that for all \(x,y\in X\),

$$\begin{aligned} d(fx,fy)\le c\, d(x,y). \end{aligned}$$
(4)

Then f has a unique fixed point and \(\big \{f^{n}x\big \}_{n\in {\mathbb {N}}}\) converges to it for all \(x\in X\).

Proof

Assume that \(\rho >0\), since the case \(\rho =0\) is treated in [8] (see also [13]). Let \(x_{0}\in X\) and define the sequence \(\{x_{n}\}\) by \(x_{n} = f^{n}x_{0}\) for all \(n\in {\mathbb {N}}\), where \(f^{n}\) is nth iterates of f. For simplification let us introduce the notation: \(d_{i,j}{:}{=}d(x_{i},x_{j})\), where \(i,j\in {\mathbb {N}}\). Now, from (4), we get

$$\begin{aligned} d_{n,n+1}\le c d_{n-1,n}<d_{n-1,n}. \end{aligned}$$

Hence, \(\{d_{n,n+1}\}\) is decreasing sequence and for all \(k\in {\mathbb {N}}\), we have

$$\begin{aligned} d_{n,n+1}\le c^{n-k}\, d_{k,k+1},\;\text { for all } n>k. \end{aligned}$$
(5)

So \(\lim \limits _{n\rightarrow \infty }d_{n,n+1}=0\), and therefore there exits \(k\in {\mathbb {N}}\) such that for all \(n\ge k\),

$$\begin{aligned} d_{n,n+1}\le 1. \end{aligned}$$
(6)

Next, we shall prove that the sequence \(\{x_{n}\}\) is Cauchy. Using \(d'_{3}\) and (6), and for sufficiently large integers pq such that \(q>p> k\) it follow that

$$\begin{aligned} d_{p,q}\le & d_{p,p+1}+b\,d_{p+1,q}+\rho \,d_{p,p+1}\,d_{p+1,q}\\\le & c^{p-k} d_{k,k+1} +b\, d_{p+1,q}+\rho \, c^{p-k} d_{k+1,k}\, d_{p+1,q}\\\le & c^{p-k} +(b+\rho \, c^{p-k} ) d_{p+1,q}, \end{aligned}$$

where

$$\begin{aligned} d_{p+1,q}\le & d_{p+1,p+2}+b\,d_{p+2,q}+\rho \,d_{p+1,p+2}\,d_{p+2,q}\\\le & c^{p-k+1} d_{k+1,k} +b\,d_{p+2,q}+\rho \, c^{p-k+1} d_{k+1,k} \,d_{p+2,q}\\\le & c^{p-k+1} +(b+\rho \, c^{p-k+1} ) d_{p+2,q}. \end{aligned}$$

By combining the previous inequalities, we obtain

$$\begin{aligned} d_{p,q}\le c^{p-k} +c^{p-k+1} (b+\rho \, c^{p-k} ) +(b+\rho \,c^{p-k}) (b+ \rho \,c^{p-k+1} ) d_{p+2,q}. \end{aligned}$$

Using (6) in all terms of the sum, we obtain by induction

$$\begin{aligned} d_{p,q} \le c^{p-k} \textstyle \sum \limits _{i=0}^{q-p-1}c^{i}\prod \limits _{j=0}^{i-1}(b+\rho \,c^{p-k+j}). \end{aligned}$$

Now, since \(c\in [0,b^{-1})\), then

$$\begin{aligned} d_{p,q} \le c^{p-k} \textstyle \sum \limits _{i=0}^{q-p-1}c^{i}\prod \limits _{j=0}^{i-1}(b+\rho \, c^{j}). \end{aligned}$$

Using d’Alembert’s criterion of convergence of real series, we deduce that \(\sum \limits _{i=0}^{\infty } u_{i}\) converges, where

$$\begin{aligned} u_{i}&{:}{=}&c^{i}\textstyle \prod \limits _{j=0}^{i-1}(b+\rho \,c^{j}). \end{aligned}$$

We conclude \(d_{p,q}\rightarrow 0\) as \(p,q \rightarrow \infty \), so the sequence \(\{x_{n}\}\) is Cauchy. Thus, it follows that \(\{x_{n}\}\) converges to some \(x_{*}\in X\), say, since X is sb-complete, which proves that \(\omega _{f}(x_{0})\) is nonempty. We now shall show that \(x_{*}\) is a fixed point of f. By using (4), we get

$$\begin{aligned} d(fx_{n(k)},fx_{*})\le c\, d(x_{n(k)},x_{*}). \end{aligned}$$

By letting \(k\rightarrow \infty \), we obtain by Proposition 1.11 and Remark 1.16 that \(x_{*}=fx_{*}\). Finally, the uniqueness of the fixed point follows immediately from (4). \(\square \)

Remark 2.2

Theorem 2.1 generalizes [3, Theorem 2.1]. Note also that for the extended suprametric, introduced by Panda et al. [22], an additional continuity assumption was added to obtain the main fixed point theorem.

Proposition 2.3

Let (Xd) be an sb-suprametric space and let \(f:X\rightarrow X\) be a Lipschitz mapping, that is, there is a constant \(\lambda \in [0,\infty )\) such that for all \(x,y\in X\),

$$\begin{aligned} d(fx,fy)\le \lambda \, d(x,y). \end{aligned}$$

Then f is continuous.

Proof

The proof is exactly the same as that of [3, Proposition 1.8]. \(\square \)

Let X be a topological space and \(f:X\rightarrow X\) be a mapping. For \(x_{0}\in X\) the \(\omega \)-limit set is given by

$$\begin{aligned} \omega _{f}(x_{0}){:}{=}\textstyle \bigcap \limits _{n\in {\mathbb {N}}}\textrm{cl}\left( \left\{ f^{k}x_{0}:k\ge n \right\} \right) . \end{aligned}$$

The next result follows immediately from Remark 1.7, Proposition 2.3 and [19, Theorem 1].

Theorem 2.4

Let (Xd) be an sb-suprametric space and let \(f:X\rightarrow X\) be a contractive mapping, that is, for all \(x,y\in X\) with \(x\ne y\),

$$\begin{aligned} d(fx,fy)< d(x,y). \end{aligned}$$

If there exists \(x_{0}\in X\) such that \(\omega _{f}(x_{0})\) is nonempty, then f has a unique fixed point and \(\big \{f^{n}x\big \}_{n\in {\mathbb {N}}}\) converges to this fixed point for all \(x\in X\).

Remark 2.5

Theorem 2.4 generalize [11, Theorem 1] and [3, Theorem 2.3]. In this connection, see also [12, Chapter 1, Theorem 1.2] and [5, Section 6].

4 Ekeland Variational Principle and Caristi Fixed Point Theorem

We first present a Cantor’s intersection theorem in sb-suprametric spaces.

Theorem 3.1

Let (Xd) be a complete sb-suprametric space, and let \(\left\{ C_{n}\right\} _{n\in {\mathbb {N}}} \) be a decreasing nested sequence of nonempty closed sets of X with

$$\begin{aligned} \textrm{diam}(C_{n})\mathrm{{:}{=}}\sup \left\{ d(x,y):x,y\in C_{n}\right\} \rightarrow 0 \text { as }n \rightarrow \infty . \end{aligned}$$

Then \(\bigcap _{n\in {\mathbb {N}}} C_{n}=\{z\}\) for some \(z\in X\).

Proof

Since \(C_{n}\) is nonempty for all n, we take \(z_{n}\) in every \(C_{n}\). We then construct a Cauchy sequence \(\{z_{n}\}\) because \(d(z_{m},z_{n})\le \textrm{diam}(C_{N})\) for all mn greater than some integer N and \(\textrm{diam}(C_{N})\rightarrow 0\) as \(N\rightarrow \infty \). Now, by completeness of (Xd), we deduce that \(\{z_{n}\}\) converges to some z. Next, since \(z_{n}\in C_{N}\) for all \(n\ge N\) and \(C_{N}\) is closed, \(z\in \bigcap _{n\ge {N}} C_{n}\), which implies by the nestedness property that \(z\in \bigcap _{n\in {\mathbb {N}}} C_{n}\). Assume now that there exists \(z'\in \bigcap _{n\in {\mathbb {N}}} C_{n}\) such that \(z'\ne z\), then \(d(z,z')>0\), which implies that there exists \(m\in {\mathbb {N}}\) such that \(\textrm{diam}(C_{n})<d(z,z')\) for all \(n\ge m\). Consequently, \(z'\not \in C_{n}\) for all \(n\ge m\) and therefore \(z'\not \in \bigcap _{n\in {\mathbb {N}}} C_{n}\). This proves that \(\bigcap _{n\in {\mathbb {N}}} C_{n}=\{z\}\). \(\square \)

We next present an Ekeland’s variational principle in the new spaces.

Theorem 3.2

Let (Xd) be a complete sb-suprametric space \((b \,{ > }\,1)\) and let \(\phi : X \rightarrow {\mathbb {R}}\cup \{\pm \infty \}\) be a lower semicontinuous function which is proper and lower bounded. Then, for every \(x_{0} \in X\) and \(\varepsilon > 0\) with

$$\begin{aligned} \phi (x_{0})\le \inf _{x\in X}\phi (x)+\varepsilon , \end{aligned}$$

there exist \(x_{\varepsilon }\in X\) and a sequence \(\{x_{i}\}_{i\in {\mathbb {N}}}\) in X such that:

  1. (i)

    \(\lim _{i\rightarrow \infty }d(x_{i},x_{\varepsilon })=0 \).

  2. (ii)

    \(d(x_{i},x_{\varepsilon })\le {2^{-n}}{\varepsilon }\), for all \(i\in {\mathbb {N}}\).

  3. (iii)

    \(\phi (x_{\varepsilon })+\sum \limits _{i=0}^{\infty }{b^{-i}}d(x_{\varepsilon },x_{i})\le \phi (x_{0})\).

  4. (iv)

    \(\phi (x_{\varepsilon })+\sum \limits _{i=0}^{\infty }{b^{-i}}d(x_{\varepsilon },x_{i})< \phi (x)+\sum \limits _{i=0}^{\infty }{b^{-i}}d(x,x_{i})\) for all \(x\ne x_{\varepsilon }\).

Proof

Let \(x_{0} \in X\) and \(\varepsilon > 0\) and define the set

$$\begin{aligned} C_{0}{:}{=}\big \{x\in X: \phi (x)+d(x,x_{0})\le \phi (x_{0})\big \}. \end{aligned}$$

Clearly, \(C_{0}\) is nonempty and closed since it contains \(x_{0}\), d is continuous and \(\phi \) is lower semi-continuous. Now, for all \(y\in C_{0}\), we have

$$\begin{aligned} d(x,x_{0})\le \phi (x_{0})-\phi (y)\le \phi (x_{0})-\inf _{x\in X}\phi (x)\le \varepsilon . \end{aligned}$$
(7)

Choose \(x_{1}\in C_{0}\) such that

$$\begin{aligned} \phi (x_{1})+d(x_{1},x_{0})\le \inf _{x\in C_{0}}\{\phi (x)+d(x,x_{0})\}+{(2b)^{-1}}{\varepsilon }, \end{aligned}$$

and consider the set

$$\begin{aligned} C_{1}&{:}{=}&\big \{x\in C_{0}: \phi (x)+d(x,x_{0})+b^{-1}d(x,x_{1})\le \phi (x_{1})+d(x_{1},x_{0})\big \}. \end{aligned}$$

By induction, we choose \(x_{n-1}\in C_{n-2}\) \((n\ge 2)\) and consider

$$\begin{aligned} C_{n-1}&{:}{=}&\big \{x\in C_{n-2}: \phi (x)+\textstyle \sum \limits _{i=0}^{n-1}b^{-i}d(x,x_{i})\le \phi (x_{n-1})+\sum \limits _{i=0}^{n-2}b^{-i}d(x_{n-1},x_{i})\big \}. \end{aligned}$$

Then, we choose \(x_{n}\in C_{n-1}\) such that

$$\begin{aligned} \phi (x_{n})+\textstyle \sum \limits _{i=0}^{n-1}b^{-i}d(x_{n},x_{i})\le \inf \limits _{x\in C_{n-1}}\left\{ \phi (x)+\sum \limits _{i=0}^{n-1}b^{-i}d(x_{n-1},x_{i})\right\} +(2b)^{-n}\varepsilon . \end{aligned}$$
(8)

Define again a set

$$\begin{aligned} C_{n}&{:}{=}&\left\{ x\in C_{n-1}: \phi (x)+\textstyle \sum \limits _{i=0}^{n}b^{-i}d(x,x_{i})\le \phi (x_{n})+\sum \limits _{i=0}^{n-1}b^{-i}d(x_{n},x_{i})\right\} . \end{aligned}$$
(9)

Clearly, \(C_{n}\) is nonempty and closed since it contains \(x_{n}\), d is continuous and \(\phi \) is lower semicontinuous. Next, for all \(y\in C_{n}\), we deduce from (8) and (9) that

$$\begin{aligned} b^{-n}d(y,x_{n})&\le \Bigg [ \phi (x_{n})+\textstyle \sum \limits _{i=0}^{n-1}b^{-i}d(x_{n},x_{i})\Bigg ]- \Bigg [\phi (y)+\sum \limits _{i=0}^{n-1}b^{-i}d(y,x_{i}) \Bigg ]\\&\le \Bigg [\phi (x_{n})+\textstyle \sum \limits _{i=0}^{n-1}b^{-i}d(x_{n},x_{i}) \Bigg ]{-}\inf \limits _{x\in C_{n-1}} \Bigg [ \phi (x)+\sum \limits _{i=0}^{n-1}b^{-i}d(x_{n},x_{i}) \Bigg ]\\&\le (2b)^{-n}\varepsilon . \end{aligned}$$

Hence, for all \(y\in C_{n}\), we have

$$\begin{aligned} d(y,x_{n})\le 2^{-n}\varepsilon . \end{aligned}$$
(10)

this implies that (i) holds. Consequently, \(\lim _{n\rightarrow \infty }\textrm{diam}(C_{n})= 0\) and the sequence \(\{C_{n}\}_{n\in {\mathbb {N}}}\) is decreasing nested sequence of nonempty closed sets of X. So, by Theorem 3.1 it follows that \(\bigcap _{n\in {\mathbb {N}}} C_{n}=\{x_{\varepsilon }\}\) for some \(x_{\varepsilon }\in X\). Note that (ii) follows from (7) and (10). Now, since for all \(x\ne x_{\varepsilon }\), \(x\not \in \bigcap _{n\in {\mathbb {N}}} C_{n}\), thus there exists \(m\in {\mathbb {N}}\) such that \(x\not \in C_{m}\), then

$$\begin{aligned} \phi (x_{m})+\textstyle \sum \limits _{i=0}^{m-1}{b^{-i}}d(x_{m},x_{i}) < \phi (x)+\textstyle \sum \limits _{i=0}^{m}{b^{-i}}d(x,x_{i}). \end{aligned}$$

But \(x\not \in C_{m}\) means that \(x\not \in C_{k}\) for all \(k\ge m\), so from the previous inequalities we conclude that for all \(k\ge m\), we have

$$\begin{aligned} \phi (x_{\varepsilon })+\textstyle \sum \limits _{i=0}^{k}{b^{-i}}d(x_{\varepsilon },x_{i})&\le \phi (x_{k})+\textstyle \sum \limits _{i=0}^{k-1}{b^{-i}}d(x_{k},x_{i})\\&\le \phi (x_{m})+\textstyle \sum \limits _{i=0}^{m-1}{b^{-i}}d(x_{m},x_{i})\le \phi (x_{0}). \end{aligned}$$

Consequently (iii) and (iv) hold. \(\square \)

Remark 3.3

Theorem 3.2 generalizes [6, Theorem 2.2].

Corollary 3.4

Let (Xd) be a complete sb-suprametric space \((b \,{ > }\,1)\) and let \(\phi : X \rightarrow {\mathbb {R}}\cup \{\pm \infty \}\) be a lower semi-continuous function which is proper and lower bounded. Then, for every \(\varepsilon > 0\) there exist \(x_{\varepsilon }\in X\) and a sequence \(\{x_{i}\}_{i\in {\mathbb {N}}}\) in X such that:

  1. (i)

    \(\lim _{i\rightarrow \infty }d(x_{i},x_{\varepsilon })=0 \).

  2. (ii)

    \(\phi (x_{\varepsilon })+\textstyle \sum \limits _{i=0}^{\infty }{b^{-i}}d(x_{\varepsilon },x_{i})\le \inf _{x\in X}\phi (x)+\varepsilon \).

  3. (iii)

    \(\phi (x_{\varepsilon })+\textstyle \sum \limits _{i=0}^{\infty }{b^{-i}}d(x_{\varepsilon },x_{i}) \le \phi (x)+\textstyle \sum \limits _{i=0}^{\infty }{b^{-i}}d(x,x_{i})\) for all \(x\in X\).

We next present a fixed point theorem in sb-suprametric spaces.

Theorem 3.5

Let (Xd) be a complete sb-suprametric space \((b\,{ > }\,1)\). Let \(f:X\rightarrow X\) be a mapping for which there exists a proper, lower semicontinuous and lower bounded function \(\phi :X\rightarrow {\mathbb {R}}\cup \{\pm \infty \}\) such that for all \(x\in X\),

$$\begin{aligned} \frac{b^2+\rho }{b-1} d(x, f(x)) \le \phi (x) - \phi (f(x)). \end{aligned}$$
(11)

Then f has a fixed point.

Proof

Assume that for all \(x \in X\), \(f(x)\ne x\). By applying Corollary 3.4, we deduce that for every \(\varepsilon > 0\) there exist \(x_{\varepsilon }\in X\) and a sequence \(\{x_{i}\}_{i\in {\mathbb {N}}}\) in X such that:

$$\begin{aligned} \phi (x_{\varepsilon })+\textstyle \sum \limits _{i=0}^{\infty }{b^{-i}}d(x_{\varepsilon },x_{i}) \le \phi (x)+\textstyle \sum \limits _{i=0}^{\infty }{b^{-i}}d(x,x_{i}), \end{aligned}$$

for all \(x\in X\). By taking \(x=f(x_{\varepsilon })\), where here \(x\ne x_{\varepsilon }\), we get

$$\begin{aligned} \phi (x_{\varepsilon })-\phi (f(x_{\varepsilon }))< \textstyle \sum \limits _{i=0}^{\infty }{b^{-i}}d(f(x_{\varepsilon }),x_{i})-\textstyle \sum \limits _{i=0}^{\infty }{b^{-i}}d(x_{\varepsilon },x_{i}). \end{aligned}$$

Now, by \(d'_{3}\) it follows that

$$\begin{aligned} \phi (x_{\varepsilon })-\phi (f(x_{\varepsilon }))< \textstyle \sum \limits _{i=0}^{\infty }b^{1-i}d(f(x_{\varepsilon }),x_{\varepsilon })+{\rho }\textstyle \sum \limits _{i=0}^{\infty }{b^{-i}}d(f(x_{\varepsilon }),x_{\varepsilon })d(x_{\varepsilon },x_{i}). \end{aligned}$$

Since \(\lim _{i\rightarrow \infty }d(x_{\varepsilon },x_{i})=0\), then there exists an integer \(N>0\) such that for all \(i\ge N\), we have \(d(x_{\varepsilon },x_{i})\le 1\). Hence,

$$\begin{aligned} \phi (x_{\varepsilon })-\phi (f(x_{\varepsilon }))< \Big (\textstyle \frac{b^2+\rho }{b-1} +\max _{0\le i\le N}\big \{b^{-i}d(x_{\varepsilon },x_{i})\big \}\Big )d(f(x_{\varepsilon }),x_{\varepsilon }). \end{aligned}$$

Next, we take \(x=x_{\varepsilon }\) in (11), we obtain

$$\begin{aligned} \frac{b^2+\rho }{b-1} d(x_{\varepsilon }, f(x_{\varepsilon })) \le \phi (x_{\varepsilon }) - \phi (f(x_{\varepsilon })), \end{aligned}$$

and this inequality combined with the previous one yield a contradiction. We conclude that f has a fixed point. \(\square \)

The Caristi’s fixed point theorem in sb-suprametric spaces follows immediately by taking \(\psi =\frac{b-1}{b^2+\rho } \phi \) in the previous theorem.

Corollary 3.6

Let (Xd) be a complete sb-suprametric space \((b\,{ > }\,1)\). Let \(f:X\rightarrow X\) be a mapping for which there exists a proper, lower semicontinuous and lower bounded function \(\psi :X\rightarrow {\mathbb {R}}\cup \{\pm \infty \}\) such that for all \(x\in X\),

$$\begin{aligned} d(x, f(x)) \le \psi (x) - \psi (f(x)). \end{aligned}$$

Then f has a fixed point.

Proposition 3.7

Corollary 3.6 generalizes [17, Theorem 2.14].

5 Strong b-Supranormed Spaces

In this section, we introduce the concept of strong b-supranormed spaces and derive some of its properties.

Definition 4.1

Let X be a nonempty linear space and \(b\,{\ge }\,1\), \(\rho \,{\ge }\,0\) are two real constants. A function \(\Vert \cdot \Vert :X\rightarrow {\mathbb {R}}_{+}\) is called \(b\)-supranorm if the following conditions hold:

(\({n}_{1}\)):

\(\Vert x\Vert =0\) if and only if \(x=0\),

(\({n}_{2}\)):

\(\Vert \lambda x\Vert =|\lambda | \Vert x\Vert \), for all \(x\in X\) and \(\lambda \in {\mathbb {R}}\)

(\({n}_{3}\)):

\(\Vert x+y\Vert \le b(\Vert x\Vert +\Vert y\Vert )+\rho \, \Vert x\Vert \,\Vert y\Vert \) for all \(x,y\in X\).

A pair \((X,\Vert \cdot \Vert )\) is called a b-supranorm space if X is a nonempty set and \(\Vert \cdot \Vert \) is a b-supranorm. The pair \((X,\Vert \cdot \Vert )\) is called a supranorm space if \(b=1\).

Definition 4.2

Let X be a nonempty linear space and \(b\,{\ge }\,1\), \(\rho \,{\ge }\,0\) are two real constants. A function \(\Vert \cdot \Vert :X\rightarrow {\mathbb {R}}_{+}\) is called strong b-supranorm (sb-supranorm) if it satisfies (\({n}_{1}\)), (\({n}_{2}\)) and

(\(n'_{3}\)):

\(\Vert x+y\Vert \le b \,\Vert x\Vert +\Vert y\Vert +\rho \Vert x\Vert \Vert y\Vert \) for all \(x, y\in X\).

A pair \((X,\Vert \cdot \Vert )\) is called a strong b-supranormed (sb-supranormed) linear space if X is a nonempty set and \(\Vert \cdot \Vert \) is a string b-supranorm. The pair \((X,\Vert \cdot \Vert )\) is called a strong supranormed linear space if \(b=1\).

Remark 4.3

Using (\({n}_{2}\)), it follows that

(\(n''_{3}\)):

\(n''_{3}\) \(\Vert x+y\Vert \le \Vert x\Vert +b \,\Vert y\Vert +\rho \Vert x\Vert \Vert y\Vert \) for all \(x, y \in X\).

Examples 4.4

  • Clearly, strong b-normed spaces of [16] are sb-supranormed spaces.

  • If \(\Vert \cdot \Vert \) is an sb-supranorm linear space X, then the function \(d:X\times X\rightarrow {\mathbb {R}}_{+}\) given by \(d(x,y)=\Vert x-y\Vert \) is an sb-suprametric.

  • Consider the set \(X={\mathbb {R}}^{2}\) endowed with a function \(\Vert \cdot \Vert :X \rightarrow {\mathbb {R}}\) defined by

    $$\begin{aligned} \Vert (x,y) \Vert =\left| x-y \right| +\min (|x|,|y|). \end{aligned}$$

    It is not difficult to see that \((X,\Vert \cdot \Vert )\) is an sb-supranormed space for \(b\,{=}\,\rho \,{=}\,2\).

Remark 4.5

Let \((X, \Vert \cdot \Vert )\) be an sb-supranormed linear space. If a sequence \(\{x_{n}\}\) converges simultaneously to x and y, that is, \(\lim \limits _{n\rightarrow \infty }\Vert x_{n}-x\Vert =\lim \limits _{n\rightarrow \infty }\Vert x_{n}-y\Vert =0\), then \(x=y\), and this follows from (\({n}_{1}\)) and (\(n'_{3}\)), since we have

$$\begin{aligned} \Vert x-y\Vert \le \Vert x-x_{n}\Vert +b \Vert x_{n}-y\Vert +\rho \Vert x-x_{n}\Vert \Vert x_{n}-y\Vert . \end{aligned}$$

Moreover, we have the following inequality:

$$\begin{aligned} \textstyle \left\| \sum \limits _{i=0}^{n}x_{i}\right\| \le b\max \{1,\rho ^{n}\} \sum \limits _{i=1}^{n}{\textbf{e}}_{i}(\Vert x_{0}\Vert ,\ldots ,\Vert x_{n}\Vert ), \end{aligned}$$

for all \(n \in {\mathbb {N}}\) and \(x_{0},\ldots , x_{n} \in X\).

Lemma 4.6

Let \((X, \Vert \cdot \Vert )\) be an sb-supranormed linear space. Then, \(\Vert \cdot \Vert \) is a continuous function.

Proof

Assume that \(\rho >0\) and let \(x,y \in X\), then

$$\begin{aligned} \Vert x\Vert =\Vert y+(x-y)\Vert \le \Vert y\Vert +b\,\Vert x-y\Vert +\rho \Vert y\Vert \,\Vert x-y\Vert , \end{aligned}$$

and consequently,

$$\begin{aligned} \frac{\Vert x\Vert -\Vert y\Vert }{b+\rho \Vert y\Vert }\le \Vert x-y\Vert . \end{aligned}$$
(12)

Similarly,

$$\begin{aligned} \frac{\Vert y\Vert -\Vert x\Vert }{b+\rho \Vert x\Vert }\le \Vert x-y\Vert . \end{aligned}$$
(13)

Hence, from (12) and (13), one gets

$$\begin{aligned} \frac{\rho (\Vert x\Vert -\Vert y\Vert )^{2}}{(b+\rho \Vert x\Vert +\rho \Vert y\Vert )^{2}}\le 2\Vert x-y\Vert . \end{aligned}$$

Therefore, if \(\Vert x_{n}-x\Vert \rightarrow 0\) as \(n\rightarrow \infty \), then

$$\begin{aligned} \frac{\rho (\Vert x_{n}\Vert -\Vert x\Vert )^{2}}{(b+\rho \Vert x_{n}\Vert +\rho \Vert x\Vert )^{2}}\le 2\Vert x_{n}-x\Vert \rightarrow 0\;\text { as }\, n\rightarrow \infty , \end{aligned}$$

and using Remark 1.6 it follows that \(\Vert x_{n}\Vert \rightarrow \Vert x\Vert \) as \(n\rightarrow \infty \), which implies that \(\Vert \cdot \Vert \) is a continuous function. \(\square \)

Let \((X_{n}, d)\) be an n-dimensional sb-supranormed linear space and let \(\{u_{1},\ldots , u_{n}\}\) be a base of \(X_{n}\). For every \(x\in X\) there exist unique coefficients \(\alpha _{1},\ldots ,\alpha _{n}\in {\mathbb {R}}\) such that

$$\begin{aligned} x=\textstyle \sum \limits _{i=1}^{n}\alpha _{i} u_{i}, \end{aligned}$$

and define \(\Vert \cdot \Vert _{0}:X \rightarrow {\mathbb {R}}_{+}\) by

$$\begin{aligned} \Vert x\Vert _{0}=\textstyle \sum \limits _{i=1}^{n} {\textbf {e}}_{i}(|\alpha _{1}|,\ldots ,|\alpha _{n}|). \end{aligned}$$

Theorem 4.7

Let \((X_{n}, \Vert \cdot \Vert )\) be an n-dimensional \(sb\)-supranormed linear space. Then, there exists \(\beta > 0\) such that

$$\begin{aligned} \Vert x\Vert \le \beta \Vert x\Vert _{0},\;\text { for all } x \in X. \end{aligned}$$
(14)

Proof

Let \(x\in X_{n}\) and \(\{u_{1},\ldots , u_{n}\}\) be a base of \(X_{n}\). Then, there exist unique coefficients \(\alpha _{1},\ldots ,\alpha _{n}\in {\mathbb {R}}\) such that \(x=\sum _{i=1}^{n}\alpha _{i} u_{i}\). Hence,

$$\begin{aligned} \Vert x\Vert&= \left\Vert \textstyle \sum \limits _{i=1}^{n}\alpha _{i} u_{i}\right\Vert \\&\le b\max \{1,\rho ^{n}\} \textstyle \sum \limits _{i=1}^{n} {\textbf {e}}_{i}(|\alpha _{1}| \, \Vert u_{1} \Vert ,\ldots ,|\alpha _{n}| \, \Vert u_{n} \Vert )\\&\le b\max \{1,\rho ^{n}\}\Big (\max _{_{1\le j_{1}<j_{2}<\cdots <j_{i}\le n}}\textstyle \prod \limits _{k=1}^{i}\Vert u_{j_{k}}\Vert \Big ) \textstyle \sum \limits _{i=1}^{n} {\textbf {e}}_{i}(|\alpha _{1}| ,\ldots ,|\alpha _{n}|)\\&= \beta \; \Vert x\Vert _{0}, \end{aligned}$$

where

$$\begin{aligned} \beta&{:}{=}&b\max \{1,\rho ^{n}\}\Big (\max _{_{1\le j_{1}<j_{2}<\cdots <j_{i}\le n}}\textstyle \prod \limits _{k=1}^{i}\Vert u_{j_{k}}\Vert \Big ). \end{aligned}$$
(15)

\(\square \)

Theorem 4.8

Let \((X_{n}, \Vert \cdot \Vert )\) be an n-dimensional \(sb\)-supranormed linear space. Then, \(\Vert \cdot \Vert \) and \(\Vert \cdot \Vert _{0}\) are equivalent, that is, there exists \(\alpha ,\beta > 0\) such that for all \(x \in X\)

$$\begin{aligned} \alpha \Vert x\Vert _{0}\le \Vert x\Vert \le \beta \Vert x\Vert _{0}. \end{aligned}$$

Proof

Let \(x\in X_{n}\) and \(\{u_{1},\ldots , u_{n}\}\) be a base of \(X_{n}\). Then, there exist unique coefficients \(\alpha _{1},\ldots ,\alpha _{n}\in {\mathbb {R}}\) such that \(x=\sum _{i=1}^{n}\alpha _{i} u_{i}\). Define a set U by

$$\begin{aligned} U&{:}{=}&\{x\in X: \Vert x\Vert _{0}=1\}. \end{aligned}$$

We first show that U is bounded. Let \(\alpha ^{j}_{1},\ldots ,\alpha ^{j}_{n}\in {\mathbb {R}}\) for \(j=1,2\) such that \(x_{j}=\sum _{i=1}^{n}\alpha ^{j}_{i} u_{i}\in U\). Then

$$\begin{aligned} \Vert x_{1}-x_{2}\Vert&=\biggl \Vert \textstyle \sum \limits _{i=1}^{n}(\alpha ^{1}_{i}-\alpha ^{2}_{i}) e_{i}\biggr \Vert \\&\le b\max \{1,\rho ^{n}\} \textstyle \sum \limits _{i=1}^{n} {\textbf {e}}_{i}(|\alpha ^{1}_{1}-\alpha ^{2}_{1}| \, \Vert u_{1} \Vert ,\ldots ,|\alpha ^{1}_{n}-\alpha ^{2}_{n}| \, \Vert u_{n} \Vert )\\&\le \beta \textstyle \sum \limits _{i=1}^{n} {\textbf {e}}_{i}(|\alpha ^{1}_{1}-\alpha ^{2}_{1}| ,\ldots ,|\alpha ^{1}_{n}-\alpha ^{2}_{n}|)\\&\le \beta \textstyle \sum \limits _{i=1}^{n} {\textbf {e}}_{i}(|\alpha ^{1}_{1}|+|\alpha ^{2}_{1}| ,\ldots ,|\alpha ^{1}_{n}|+|\alpha ^{2}_{n}|) \\&\le \beta \, K, \end{aligned}$$

where \(\beta \) is given by (15) and \( K {:}{=}\max \big \{2^{k}n \left( ^{n}_{k}\right) :k=1,\ldots , n\big \} \), which proves that U is bounded.

Define now a function \(\phi :X \rightarrow {\mathbb {R}}_{+}\) by \(\phi (x)=\Vert x\Vert \). It follows by Lemma 4.6 that f is continuous. Note that U is strongly compact, since it is bounded and closed subset of \({\mathbb {R}}^{n}\). Using Theorem 1.19, we deduce that \(\phi \) has an infimum \(\alpha \) in U, which is different from zero because \(\Vert x\Vert _{0}=1\) for every vector \(x\in U\). Hence,

$$\begin{aligned} \alpha =\inf \{\phi (x): x\in U\}=\inf \{\Vert x\Vert : x\in U\}>0. \end{aligned}$$

Thus, from the fact \(\frac{x}{\Vert x\Vert _{0}}\in U\) for all \(x\in X\), it follows that

$$\begin{aligned} \left\| \frac{x}{\Vert x\Vert _{0}}\right\| \ge \alpha >0,\;\text { for all } x \in X, \end{aligned}$$

which implies that

$$\begin{aligned} \alpha \Vert x\Vert _{0} \le \Vert x\Vert ,\;\text { for all } x \in X. \end{aligned}$$
(16)

Finally, combine (14) and (16), we obtain the result. \(\square \)

Remark 4.9

As an immediate consequence of Theorem 4.8, any two sb-supranorms on a finite-dimensional space are equivalent.

Lemma 4.10

Let \((X_{n}, \Vert \cdot \Vert )\) be an n-dimensional \(sb\)-supranormed linear space, let U be a bounded set of X. Then, U is compact.

Proof

Let \(x\in X_{n}\) and \(\{u_{1},\ldots , u_{n}\}\) be a base of \(X_{n}\). Then, there exist unique coefficients \(\alpha _{1},\ldots ,\alpha _{n}\in {\mathbb {R}}\) such that \(x=\sum _{i=1}^{n}\alpha _{i} u_{i}\). Let \({\overline{x}}=(\alpha _{1},\ldots ,\alpha _{n})\in {\mathbb {R}}^{n}\). Define the function \(\phi :U \rightarrow {\mathbb {R}}^{n}\) by \(\phi (x)={\overline{x}}\) for all \(x\in U\) and let \(V=\phi (U)\). Since the function \(\Vert \cdot \Vert _{1}:X\rightarrow {\mathbb {R}}_{+}\) by \(\Vert x\Vert _{1}{:}{=}\Vert \phi (x)\Vert _{n}\) is an sb-supranorm on \(X_{n}\), where \(\Vert \cdot \Vert _{n}\) is an sb-supranorm on \({\mathbb {R}}^{n}\). Hence, according to Remark 4.9, \(\Vert \cdot \Vert _{1}\) is equivalent to the \(sb\)-supranorm \(\Vert \cdot \Vert \). We conclude that there exist \(\alpha ,\beta >0\) such that

$$\begin{aligned} \alpha \Vert {\overline{x}}\Vert _{n} \le \Vert x\Vert \le \beta \Vert {\overline{x}}\Vert _{n}. \end{aligned}$$
(17)

As consequences of (17), U bounded in X if and only if U bounded in \({\mathbb {R}}^{n}\), and a sequence \(\{x_{n}\}\) is convergent in \((X, \Vert \cdot \Vert )\) if and only if the corresponding sequence \(\{{\overline{x}}_{n}\}\) is convergent in \({\mathbb {R}}^{n}\). Consequently, the compactness of U bounded in X follows from the compactness of U bounded in \({\mathbb {R}}^{n}\). \(\square \)

6 Brouwer and Schauder Fixed Point Principles

We first recall the Brouwer fixed point principle in \({\mathbb {R}}^{n}\).

Theorem 5.1

(Brouwer) Let U be a bounded closed convex set of \({\mathbb {R}}^{n}\). If a mapping \(f:U \rightarrow U\) is continuous, then it has a fixed point.

The Brouwer fixed point principle in sb-supranormed space is given next.

Theorem 5.2

Let \((X_{n}, \Vert \cdot \Vert )\) be an n-dimensional sb-supranormed linear space and let U be a bounded closed convex set of \(X_{n}\). If a mapping \(f:U \rightarrow U\) is continuous, then it has a fixed point.

Proof

Let \(x\in X_{n}\) and \(\{u_{1},\ldots , u_{n}\}\) be a base of \(X_{n}\). Then, there exist unique coefficients \(\alpha _{1},\ldots ,\alpha _{n}\in {\mathbb {R}}\) such that \(x=\sum _{i=1}^{n}\alpha _{i} u_{i}\). Let \({\overline{x}}=(\alpha _{1},\ldots ,\alpha _{n})\in {\mathbb {R}}^{n}\). Define the function \(\phi :U \rightarrow {\mathbb {R}}^{n}\) by \(\phi (x)={\overline{x}}\) for all \(x\in U\) and let \(V=\phi (U)\). The function \(\phi :U \rightarrow V\) is bijective, where the mapping \(\phi ^{-1}:V\rightarrow U\) is given by \(\phi ^{-1}({\overline{x}})=x\) for all \({\overline{x}}\in V\). Next, we will prove several claims: \(\square \)

Claim 1

\(\phi :U \rightarrow V\) is an homeomorphism. Indeed \(\phi \) is continuous in U, since by (16), we have

$$\begin{aligned} \Vert \phi (x)-\phi (x_{0})\Vert _{0} =\Vert {\overline{x}}-{\overline{x}}_{0}\Vert _{0} \le \alpha ^{-1}\Vert x-x_{0}\Vert ,\;\text { for all } x,x_{0} \in U, \end{aligned}$$

Similarly, \(\phi ^{-1}\) is continuous in V, because by (14), we have

$$\begin{aligned} \Vert \phi ^{-1}({\overline{x}})-\phi ^{-1}({\overline{x}}_{0})\Vert = \Vert x-x_{0}\Vert \le \beta \Vert {\overline{x}}-{\overline{x}}_{0}\Vert _{0},\;\text { for all } {\overline{x}}, {\overline{x}}_{0} \in V. \end{aligned}$$

Hence, Claim 1 holds.

Claim 2

V is convex. Let \({\overline{x}}=(\alpha _{1},\ldots ,\alpha _{n}), {\overline{y}}=(\beta _{1},\ldots ,\beta _{n})\in V\), where \(\alpha _{i},\beta _{i}\in {\mathbb {R}}\) for \(i=1,\ldots ,n\). For all \(\lambda \in [0,1]\), we obtain by convexity of U that

$$\begin{aligned} \lambda {\overline{x}}+(1-\lambda ){\overline{y}}&=\left( \lambda \alpha _{1}+(1-\lambda )\beta _{1}, \ldots , \lambda \alpha _{n}+(1-\lambda )\beta _{n} \right) \\&= \phi (\lambda {x}+(1-\lambda ){y})\in V, \end{aligned}$$

which implies that Claim 2 holds.

Claim 3

V is bounded. Let \({\overline{x}}, {\overline{y}} \in V\), then by the boundedness of U and Theorem 4.7 it follows that

$$\begin{aligned} \Vert {\overline{x}}-{\overline{y}}\Vert _{0}\le \alpha ^{-1}\Vert x-y\Vert \le \alpha ^{-1} d(U), \end{aligned}$$

where \(d(U){:}{=}\max \{\Vert x-y\Vert : x,y\in U\}\), which proves Claim 3.

Claim 4

V is closed. Let \(x=\sum _{i=1}^{n}\alpha _{i} u_{i}\), \(x_{0}=\sum _{i=1}^{n}\beta _{i} u_{i}\), \({\overline{x}}=(\alpha _{1},\ldots ,\alpha _{n})\in V\), \({\overline{x}}_{0}=(\beta _{1},\ldots ,\beta _{n})\), where \(\alpha _{i},\beta _{i}\in {\mathbb {R}}\) for \(i=1,\ldots , n\). Assume that \(\Vert {\overline{x}}-{\overline{x}}_{0}\Vert _{0}\) tends to zero. Now, by (16), we have

$$\begin{aligned} \Vert {x}-{x}_{0}\Vert \le \beta \Vert {\overline{x}}-{\overline{x}}_{0}\Vert _{0}, \end{aligned}$$

so it follows that by closedness of U that \(x_{0}\in U\), which implies that \({\overline{x}}_{0}\in V\) and this prove the claim.

Claim 5

f has a fixed point. To show this, define the function \(F:V\rightarrow V\) by \(F=\phi {f}\phi ^{-1}\). By Theorem 5.1 and the previous claims, we deduce that there exists \({\overline{x}}\in V\) such that \(F(x)=x\), that is,

$$\begin{aligned} \phi {f}\phi ^{-1}(x)=x, \end{aligned}$$

which is equivalent to \(f(\phi ^{-1}(x))=\phi ^{-1}(x)\), and since \(\phi ^{-1}(x)\in U\), then f has a fixed point in U. \(\square \)

Before establishing the fixed point principle of Schauder type in sb-supranormed spaces, we need to develop some auxiliary results. Let \((E,\Vert \cdot \Vert )\) be an sb-supranormed linear space and \(N {:}{=}\{c_{1}, \ldots , c_{n}\}\) be a finite subset of E. For any fixed \(\varepsilon > 0\), define the set

$$\begin{aligned} (N,{\varepsilon })&{:}{=}&\textstyle \bigcup \limits _{i=1}^{n} B(c_{i},\varepsilon ), \end{aligned}$$

where

$$\begin{aligned} B(c_{i},\varepsilon ){:}{=}\left\{ x\in E: \Vert x-c_{i}\Vert <\varepsilon \right\} ,\quad i=1,\ldots ,n. \end{aligned}$$

Define a mapping \(\mu _{i}:(N,{\varepsilon })\rightarrow {\mathbb {R}}\) by

$$\begin{aligned} \mu _{i}(x){:}{=}\max \big [0, \varepsilon -\Vert x-c_{i}\Vert \big ],\quad i=1,\ldots ,n. \end{aligned}$$

Consider the Schauder projection \(p_{\varepsilon }:(N,{\varepsilon }) \rightarrow \textrm{conv}(N)\) given by

$$\begin{aligned} p_{\varepsilon }(x)=\textstyle \left[ \,\sum \limits _{i=1}^{n}\mu _{i}(x)\right] ^{-1}\sum \limits _{i=1}^{n}\mu _{i}(x)c_{i}, \end{aligned}$$

Note that \(p_{\varepsilon }((N,{\varepsilon }))\subset \textrm{conv}(N)\) as a convex combination of \(\{c_{1}, \ldots , c_{n}\}\). Moreover, if \(x\in (N,{\varepsilon })\), then there exists i such that \(x\in B(C_{i},\varepsilon )\), so \(\sum \nolimits _{i=1}^{n}\mu _{i}(x)\ne 0\), which means that \(p_{\varepsilon }\) is well defined.

Lemma 5.3

Let \((E,\Vert \cdot \Vert )\) be an sb-supranormed linear space, U be a convex subset of E and \(N=\{c_{1},\ldots ,c_{n}\}\subset U\). Then for a sufficiently small \(\varepsilon > 0\), we have

  1. (i)

    \(\Vert x-p_{\varepsilon }(x)\Vert \le n\,b\,\varepsilon \, \max \{1,\rho ^{n}\} \) for all \(x\in (N,{\varepsilon })\),

  2. (ii)

    \(p_{\varepsilon }:(N,{\varepsilon })\rightarrow \textrm{conv}(N)\subset U\) is a continuous compact mapping.

Proof

Let \(\varepsilon \in (0,1]\) be sufficiently small such that for every \(1\,{\le }\, i\, {\le }\, n\) and any x in \((N,\varepsilon )\), \(P_{i}(x)\le P_{1}(x)\), where \(P_{i}(x){:}{=}{\textbf {e}}_{i}(\mu _{1}(x),\ldots ,\mu _{n}(x))\). Then,

$$\begin{aligned}&\Vert x-p_{\varepsilon }(x)\Vert = \textstyle \left[ \,\sum \limits _{i=1}^{n}\mu _{i}(x)\right] ^{-1}\left\| \sum \limits _{i=1}^{n}\mu _{i}(x) (x-c_{i})\right\| \\&\quad \le b \max \{1,\rho ^{n}\} \textstyle (P_{1}(x))^{-1} \sum \limits _{i=1}^{n} {\textbf {e}}_{i}(\mu _{1}(x)\Vert x-c_{1}\Vert ,\ldots ,\mu _{n}(x)\Vert x-c_{n}\Vert )\\&\quad \le b \max \{1,\rho ^{n}\}\textstyle (P_{1}(x))^{-1} \sum \limits _{i=1}^{n} {\textbf {e}}_{i}(\mu _{1}(x)\varepsilon ,\ldots ,\mu _{n}(x)\varepsilon )\\&\quad \le \textstyle b\,\varepsilon \,\max \{1,\rho ^{n}\} (P_{1}(x))^{-1} \sum \limits _{i=1}^{n}P_{i}(x)\\&\quad \le n\,b\,\varepsilon \,\max \{1,\rho ^{n}\}. \end{aligned}$$

Now, since \(p_{\varepsilon }\) is a finite sum of continuous functions and \(\Vert \cdot \Vert \) is continuous according to Lemma 4.6, then \(p_{\varepsilon }\) is continuous. The compactness of \(p_{\varepsilon }\) follows from Lemma 4.10, since its codomain is with finite-dimension. \(\square \)

Lemma 5.4

Let X be a topological space and E be an sb-supranormed linear space. Let U be a convex set of E and \(f :X \rightarrow U\) be a compact mapping. Then for a sufficiently small \(\varepsilon > 0\), there exists a finite set

$$\begin{aligned} N=\{c_{1},\ldots ,c_{n}\}\subset f(X)\subset U, \end{aligned}$$

and a finite-dimensional mapping \(f_{\varepsilon } :X \rightarrow U\) such that:

  1. (i)

    \(\Vert f_{\varepsilon } (x)-f(x)\Vert \le n\, b\, \varepsilon \, \max \{1,\rho ^{n}\} \) for all \(x\in X\),

  2. (ii)

    \(f_{\varepsilon } (X)\subset \textrm{conv}(N) \subset U\).

Proof

(i)::

By Theorem 1.20 and for sufficiently small \(\varepsilon \in (0,1)\) there exists a finite \(\varepsilon \)-net \(\{c_{1},\ldots ,c_{n}\} \subset f(X)\) because f(X) is compact in E. Now, if \(y\in f(X)\), then \(d(y,c_{i})<\varepsilon \) for some \(i\in \{1,\ldots ,n\}\), thus \(y\in B(c_{i},\varepsilon )\), so \(y\in (N,\varepsilon )\) and this proves that \(f(X)\subset (N,\varepsilon )\). Let \(f_{\varepsilon }=p_{\varepsilon }f\). We deduce by Lemma 5.3 that

$$\begin{aligned} \Vert f_{\varepsilon }(x)-f(x)\Vert =\Vert p_{\varepsilon }y-y\Vert \le n\, b\, \varepsilon \, \max \{1,\rho ^{n}\}, \end{aligned}$$

where \(y=f(x)\in (N,\varepsilon )\), for all \(x\in X\).

(ii)::

Let \(y\in f_{\varepsilon }(X)\). Thus, there is \(z=f(x)\in (N,\varepsilon )\) for some \(x\in X\) such that \(y=p_{\varepsilon }(z)\). Consider

$$\begin{aligned} y=p_{\varepsilon }(z)=\textstyle \sum \limits _{i=1}^{n}\lambda _{i} c_{i},\;\; \textstyle \sum \limits _{i=1}^{n}\lambda _{i}=1, \;\;\lambda _{i}\in {\mathbb {R}},\;\; i=1,\ldots ,n. \end{aligned}$$

Thus, \(y\in \textrm{conv}(N)\subset U\), and by convexity of U it follows that \(f_{\varepsilon }\subset \textrm{conv}(N)\subset U\). \(\square \)

Let (Xd) be an sb-suprametric space, U be a nonempty set of X and \(f :U\rightarrow X\) be a given mapping. If for a given \(\varepsilon > 0\), there exists a point \(x \in U \) such that \(d(x,f(x)) < \varepsilon \), then we say that x is an \(\varepsilon \)-fixed point for f.

Theorem 5.5

Let (Xd) be an sb-suprametric space and U be a closed set of X. If a mapping \(f :U \rightarrow X\) is compact, then f has a fixed point if and only if for each \(\varepsilon > 0\) it has an \(\varepsilon \)-fixed point.

Proof

The necessary condition is trivial, so we only show the sufficient condition. Let \(\varepsilon _{n}=\frac{1}{n}\), \(n\in {\mathbb {N}}\). Assume there exists \(u_{n}\in U\) for all \(n\in {\mathbb {N}}\) such that \(u_{n}\) are \(\varepsilon _{n}\)-fixed point, that is,

$$\begin{aligned} d(u_{n},f(u_{n}))< \textstyle \frac{1}{n},\;\text { for all }n\in {\mathbb {N}}. \end{aligned}$$
(18)

The mapping f is compact, so there exists a compact K such that \(f(X)\subseteq K\). Thus, there exists a subsequence \(\{u_{n_{k}}\}\) such that \(f(u_{n_{k}})\) converges to some \(u\in X\) as k tends to infinity. Now, using (18), it follows that for any \(\varepsilon >0\) there exists \(k_{0}\in {\mathbb {N}}\) such that for \(k\ge k_{0}\), we have

$$\begin{aligned} d(u_{n_{k}}, u)&\le b d(u_{n_{k}}, f(u_{n_{k}})) +d(f(u_{n_{k}}), u) + \rho \,d(u_{n_{k}}, f(u_{n_{k}})) d(f(u_{n_{k}}), u) \\&\le \textstyle \frac{b}{n_{k}} + \varepsilon + \frac{\rho \, \varepsilon }{n_{k}} < \varepsilon (b+1+ \varepsilon ), \end{aligned}$$

which implies that \(\{u_{n_{k}}\}\) converges to u in U because U is closed. Observe that \(\{f(u_{n_{k}})\}\) converges to u and by continuity of f it converges also to f(u), which means by Proposition 1.11 that \(u=f(u)\). \(\square \)

Remark 5.6

In Theorem 5.5, if \(f :U \rightarrow U\) is compact, the assumption of closeness of U may be dropped, since the sequence \(f(u_{n_{k}})\) converges to some \(u\in \textrm{cl}\left( f(U)\right) \) which is a subset of U.

Finally, we present a Schauder fixed point principle.

Theorem 5.7

Let \((X, \Vert \cdot \Vert )\) be an sb-supranormed linear space and U be a convex set (not necessarily closed) of X. If a mapping \(f:U \rightarrow U\) is compact, then it has a fixed point.

Proof

It suffice to show that f has an \(\varepsilon \)-fixed point. By Lemma 5.4 it follows that for a sufficiently small \(\varepsilon >0\) there exists \(f_{\varepsilon }: U\rightarrow U\) such that

  1. (i)

    \(\Vert f_{\varepsilon } (x)-f(x)\Vert \le n\, b\, \varepsilon \, \max \{1,\rho ^{n}\} \) for all \(x\in U\),

  2. (ii)

    \(f_{\varepsilon } (U)\subset \textrm{conv}(N) \subset U\).

Since \(\textrm{conv}(N) \subset U\), we get \(f_{\varepsilon } (\textrm{conv}(N))\subset f_{\varepsilon }(U) \subset \textrm{conv}(N)\), which implies that \(f_{\varepsilon }: \textrm{conv}(N)\rightarrow \textrm{conv}(N)\) is well defined. Since \(\textrm{conv}(N)\) is bounded closed convex (see also [24, Propositions C.2 and C5]), we deduce by Theorem 5.2 that there exists \(x_{\varepsilon }\in \textrm{conv}(N) \subset U\) such that \(f_{\varepsilon }x_{\varepsilon }=x_{\varepsilon }\), so by (i), we obtain

$$\begin{aligned} \Vert f(x_{\varepsilon })-x_{\varepsilon }\Vert = \Vert f (x_{\varepsilon })-f_{\varepsilon }(x_{\varepsilon })\Vert \le n\, b\, \varepsilon \, \max \{1,\rho ^{n}\}, \end{aligned}$$

and by letting \({\varepsilon }\) tends to zero together with the continuity of f, we obtain the result by Theorem 5.5 and Remark 5.6. \(\square \)

Remark 5.8

Observe that U is not necessary closed, since Theorem 5.2 is applied to the selfmap \(f_{\varepsilon }\) defined on the closed set \(\textrm{conv}(N)\). Moreover and according to Remark 5.6, Theorem 5.5 can be applied without requiring the closeness of U. This answer the question in [9, Remark 13].

7 Applications

In this section, we study the existence of a unique solution to an integral equation as well as to a boundary value problem, as applications to the fixed point theorem proved in Section 2. We consider the integral equation:

$$\begin{aligned} x(t)=\lambda (t)+\int _{0}^{1} G(t,s)h(s,x(s))\textrm{d}s,\;\; t\in [0,1]. \end{aligned}$$
(19)

The problem of existence of a solution for the integral equation (19) will be discussed under the following assumptions:

(\(a_1\)):

\(\lambda :[0,1]\rightarrow {\mathbb {R}}_{+}\) is a continuous function.

(\(a_2\)):

\(h:[0,1]\,{\times }\,{\mathbb {R}}_{+} {\,\rightarrow \,} {\mathbb {R}}_{+}\) is a continuous function and there exists a continuous function \(u:{\mathbb {R}}^{2}_{+} {\,\rightarrow \,} {\mathbb {R}}_{+}\) such that for all \((s,p,q)\,{\in }\, [0,1]\,{\times }\,{\mathbb {R}}_{+}^{2}\),

$$\begin{aligned} & u(p,p)=0, \end{aligned}$$
(20a)
$$\begin{aligned} & \left| h(s,p)-h(s,q)\right| \le u(p,q), \end{aligned}$$
(20b)
$$\begin{aligned} & u(p,q)^{2}+ {\textstyle \frac{1}{2}}u(p,q) \le |p-q|^{2}+ {\textstyle \frac{1}{2}}|p-q|. \end{aligned}$$
(20c)
(\(a_3\)):

\(G:[0,1]^{2} \rightarrow {\mathbb {R}}_{+}\) is a continuous function such that

$$\begin{aligned} c{:}{=}\max _{s,t\in [0,1]}G(s,t) <1. \end{aligned}$$

Before presenting the main result of this section, we derive an inequality of Chebyshev type. For more details on Chebyshev inequalities, we refer to Chapter IX of [20] and for more recent references, see for instance [1, 23].

Lemma 6.1

Let a and b be real numbers such that \(a\, {< }\,b\), and let \(w(t,\cdot )\) be a nonnegative measurable function for every \(t\,{\in }\,[a,b]\). Let \(x(s)=(x_{1}(s),x_{2}(s),\ldots ,x_{n}(s))\) such that \(\{x_{i}\}_{1\le i \le n}\) are nonnegative functions defined on [ab], and let u be a nonnegative function defined on \([a,b]\times {\mathbb {R}}^{n}_{+}\) such that \(s\mapsto u(s,x(s))\) is integrable with respect to w(ts) for every \(t\in [a,b]\). Then

$$\begin{aligned} \left( \int _{a}^{b} u(s,x(s))w(t,s)\textrm{d}s \right) ^{2}\le \int _{a}^{b} u(s,x(s))^{2}w(t,s)\textrm{d}s \int _{a}^{b} w(t,s)\textrm{d}s,\;t\in [a,b]. \end{aligned}$$

Proof

We have

$$\begin{aligned} 0&\le \int _{a}^{b} \int _{a}^{b} \big (u(r,x(r))-u(s,x(s))\big )^{2}w(t,s)w(t,r)\textrm{d}s\textrm{d}r\\&= \int _{a}^{b} \int _{a}^{b}\big (u(r,x(r))^{2}-2u(r,x(r))u(s,x(s))+u(s,x(s))^{2}\big )w(t,s)w(t,r)\textrm{d}s\textrm{d}r\\&= \int _{a}^{b}\big ( u(r,x(r))^{2}\int _{a}^{b}w(t,s)\textrm{d}s-2u(r,x(r))\int _{a}^{b}u(s,x(s))w(t,s)\textrm{d}s\\ &\quad +\int _{a}^{b}u(s,x(s))^{2}w(t,s)\textrm{d}s\big )w(t,r)\textrm{d}r\\&= 2\int _{a}^{b} u(s,x(s))^{2}w(t,s)\textrm{d}s\int _{a}^{b}w(t,s)\textrm{d}s-2 \left( \int _{a}^{b}u(s,x(s))w(t,s)\textrm{d}s\right) ^{2}. \end{aligned}$$

\(\square \)

Theorem 6.2

Under assumptions (\(a_1\))–(\(a_3\)), the integral equation (19) has a unique solution in \(C_{+}([0, 1])\).

Proof

Let \(X=C_{+}([0, 1])\) be the set of continuous functions \(x:[0, 1]\rightarrow {\mathbb {R}}_{+}\), endowed with the suprametric \(\delta \) of Examples 1.4. First, by Remark 1.14, \((X,\delta )\) is a complete. Consider the operator \(T:X\rightarrow X\) defined by

$$\begin{aligned} Tx(t)=\lambda (t)+\int _{0}^{1} G(t,s)h(s,x(s))\textrm{d}s,\;\; t\in [0,1]. \end{aligned}$$

Observe first that T is well defined. Let \(x,y\in X\), then by using the assumptions (\(a_1\))–(\(a_3\)) and Lemma 6.1, we get

$$\begin{aligned}&|Tx(t)-Ty(t)|(|Tx(t)-Ty(t)|+{\textstyle \frac{1}{2}})\\&\quad = \left| \int _{0}^{1}\! G(t,s)\big (h(s,x(s)){-}h(s,y(s))\big )\textrm{d}s\right| \left( \left| \int _{0}^{1}\! G(t,s)\big (h(s,x(s)){-}h(s,y(s))\big )\textrm{d}s\right| {+}\textstyle \frac{1}{2}\right) \\&\quad \le \int _{0}^{1} G(t,s)\big |h(s,x(s))-h(s,y(s))\big |\textrm{d}s\left( \int _{0}^{1} G(t,s)\big |h(s,x(s))-h(s,y(s))\big |\textrm{d}s{+}{\textstyle \frac{1}{2}}\right) \\&\quad \le \int _{0}^{1} G(t,s)u(x(s),y(s))\textrm{d}s\left( \int _{0}^{1} G(t,s)u(x(s),y(s))\textrm{d}s{+}\textstyle \frac{1}{2}\right) \\&\quad \le \int _{0}^{1} G(t,s)\textrm{d}s\int _{0}^{1} G(t,s)u(x(s),y(s))^{2}\textrm{d}s{+}{\textstyle \frac{1}{2}}\int _{0}^{1} G(t,s)u(x(s),y(s))\textrm{d}s\\&\quad \le \int _{0}^{1} G(t,s)\big (u(x(s),y(s))^{2}+ {\textstyle \frac{1}{2}}u(x(s),y(s))\big )\textrm{d}s\\&\quad \le \int _{0}^{1} G(t,s)|x(s)-y(s)|\big (|x(s)-y(s)|+ {\textstyle \frac{1}{2}}\big )\textrm{d}s\\&\quad \le c\delta (x,y), \end{aligned}$$

and this implies

$$\begin{aligned} \delta (Tx,Ty)\le c\delta (x,y). \end{aligned}$$

By Theorem 2.1, we conclude that the integral equation (19) has a unique solution in X. \(\square \)

Next by Theorem 6.2, we show the existence of a unique solution in \(C_{+}[0,1]\) to the following nonlinear third-order boundary value problem:

$$\begin{aligned}&x'''(t)+{\sqrt{t\,x(t)+1}}(1-e^{-t \, x(t)})=0,\quad t\in [0,1], \end{aligned}$$
(22a)
$$\begin{aligned}&x(0)=x'(1)=0\text { and }x(1)=1. \end{aligned}$$
(22b)

Proposition 6.3

The boundary value problem (22) has a unique solution in \(C_{+}[0,1]\).

Proof

The boundary value problem (22) has a solution \(x\in C_{+}[0,1]\) if and only if the operator \(T:C_{+}[0,1]\rightarrow C_{+}[0,1]\) defined by

$$\begin{aligned} Tx(t)=\int _{0}^{1}G(t,s){\sqrt{ s\,x(s)+1}}(1-e^{-s \, x(s)})\textrm{d}s,\quad t\in [0,1], \end{aligned}$$

has a fixed point in \(C_{+}[0,1]\), where the Green’s function associated to the homogeneous problem \(x'''(t)=0\) that satisfies the boundary condition (22b) is given by

$$\begin{aligned} G(t,s)= \left\{ \begin{array}{ll} {\textstyle \frac{1}{2}}s^2(t-1)^2,& \quad 0 \le s\le t\le 1,\\ {\textstyle \frac{1}{2}}t(s-1)(s(t-2)+t)),& \quad 0 \le t\le s \le 1. \end{array} \right. \end{aligned}$$

Firstly, observe that T is well defined and (\(a_1\)) holds, where \(\lambda =0\). Moreover, it is easy to see that G is continuous and satisfies (\(a_3\)), since we have

$$\begin{aligned} 0\le G(t,s) \le {\textstyle \frac{1}{2}},\text { for all }t,s\in [0,1]. \end{aligned}$$

Consider now the functions \(h:[0,1]\times {\mathbb {R}}_{+}\rightarrow {\mathbb {R}}_{+}\) and \(u:{\mathbb {R}}^{2}_{+}\rightarrow {\mathbb {R}}_{+}\) given by

$$\begin{aligned} h(s,p)=\sqrt{s\,p+1}(1-e^{-s \, p}) \;\text { and }\; u(p,q)=\sqrt{|p-q|+1}(1-e^{-|p-q|}). \end{aligned}$$

In order to use Theorem 6.2 and conclude that T has a unique solution in \(C_{+}[0,1]\), we have to check (\(a_2\)). Note that h and u are continuous and it is not difficult to see that (20b) and (20c) follow from the next lemma. \(\square \)

Lemma 6.4

For all \((p,q,s)\in {\mathbb {R}}^{2}_{+}\times [0,1]\), we have \(A\ge 0\) and \(B\ge 0\), where

$$\begin{aligned} & A=\textstyle \sqrt{|p-q|+1}(1-e^{-|p-q|})-|\sqrt{p+1}(1-e^{-p})-\sqrt{q+1}(1-e^{-q})|,\\ & \quad B= p^2+\textstyle \frac{1}{2}p-(s\,p+1)(1-e^{-s\,p})^2-\frac{1}{2}\sqrt{s\,p+1}(1-e^{-s\,p}). \end{aligned}$$

Proof

Suppose, without loss of generality, that \(p> q\). Then,

$$\begin{aligned} A=\textstyle \sqrt{p-q+1}(1-e^{q-p})-\sqrt{p+1}(1-e^{-p})+\sqrt{q+1}(1-e^{-q}) \end{aligned}$$

Using the mean value theorem twice, it follows that there exists \(c\in (q, p)\) such that

$$\begin{aligned} A=\sqrt{p-q+1}(1-e^{q-p})-\textstyle \frac{1}{2}(p-q)\frac{1+e^{-c}+2ce^{-c}}{ \sqrt{c+1}}, \end{aligned}$$

and also there exists \(c'\in (p-q, p)\) such that

$$\begin{aligned} A=\textstyle \sqrt{q+1}(1-e^{-q})-\frac{1}{2}q\frac{1+e^{-c'}+2c'e^{-c'}}{ \sqrt{c'+1}}. \end{aligned}$$

Now, since the function \(h_{1}:{\mathbb {R}}_{+}\rightarrow {\mathbb {R}}\) given by \(h_{1}(t)=\textstyle \frac{1+e^{-t}+2 t e^{-t}}{ \sqrt{t+1}}\) is decreasing on \({\mathbb {R}}_{+}\), we obtain

$$\begin{aligned} A\ge \textstyle \sqrt{p-q+1}(1-e^{q-p})-\textstyle \frac{1}{2}(p-q)\frac{1+e^{-q}+2 q e^{-q}}{ \sqrt{q+1}}, \end{aligned}$$

and

$$\begin{aligned} A\ge \textstyle \sqrt{q+1}(1-e^{-q})-\frac{1}{2}q\frac{1+e^{-(p-q)}+2(p-q)e^{-(p-q)}}{ \sqrt{p-q+1}}. \end{aligned}$$

Hence, it suffice to know the sign of \(h_{2}(p-q)-h_{1}(q)\) and \(h_{2}(q)-h_{1}(p-q)\), where the function \(h_{2}:{\mathbb {R}}_{+}\rightarrow {\mathbb {R}}\) is given by \(h_{2}(t)=\textstyle {2t^{-1}\sqrt{t+1}(1-e^{-t})}\). It is not difficult to see that \(h_{2}\) is decreasing, so if \(p-q\le q\), \(h_{2}(p-q)-h_{1}(q)\ge h_{2}(p-q)-h_{1}(p-q)\) and if \(p-q>q\), \(h_{2}(q)-h_{1}(p-q)\ge h_{2}(q)-h_{1}(q)\). We conclude from the fact that \(t\mapsto (h_{2}-h_{1})(t)\) is positive that \(A\ge 0\). Finally, we have

$$\begin{aligned} B&\ge p^2+\textstyle \frac{1}{2}p-(s\,p+1)(1-e^{-s\,p})^2-\sqrt{s\,p+1}(1-e^{-s\,p})\\&\ge (s\,p)^2+\textstyle \frac{1}{2}s\,p-(s\,p+1-\sqrt{s\,p+1})(1-e^{-s\,p})\ge 0. \end{aligned}$$

\(\square \)