OMA – Vol 4 – Issue 2 (2020) – PISRT https://old.pisrt.org Tue, 08 Jun 2021 17:56:26 +0000 en-US hourly 1 https://wordpress.org/?v=6.7 Some applications of second-order differential subordination for a class of analytic function defined by the lambda operator https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/some-applications-of-second-order-differential-subordination-for-a-class-of-analytic-function-defined-by-the-lambda-operator/ Sun, 27 Dec 2020 15:16:44 +0000 https://old.pisrt.org/?p=4847
OMA-Vol. 4 (2020), Issue 2, pp. 170 - 177 Open Access Full-Text PDF
B. Venkateswarlu, P. Thirupathi Reddy, S. Sridevi, Sujatha
Abstract: In this paper, we introduce a new class of analytic functions by using the lambda operator and obtain some subordination results.
]]>

Open Journal of Mathematical Analysis

Some applications of second-order differential subordination for a class of analytic function defined by the lambda operator

B. Venkateswarlu\(^1\), P. Thirupathi Reddy, S. Sridevi, Sujatha
Department of Mathematics, GSS, GITAM University, Doddaballapur- 562 163, Bengaluru Rural, Karnataka, India.; (B.V & S.S & S)
Department of Mathematics, Kakatiya Univeristy, Warangal- 506 009, Telangana, India.; (P.T.R)
\(^{1}\)Corresponding Author: bvlmaths@gmail.com

Abstract

In this paper, we introduce a new class of analytic functions by using the lambda operator and obtain some subordination results.

Keywords:

Analytic, convex, subordination, symmetric.

1. Introduction

Let \( \mathbb{C} \) be complex plane and let \(\mathbb{ U } = \{z: z \in \mathbb{C} ~\text{and}~ |z| < 1\} = \mathbb{ U } \setminus \{0\}\) be an open unit disc in \( \mathbb{C} .\) Also let \(H(\mathbb{ U } )\) be a class of analytic functions in \(\mathbb{ U } .\) For \(n \in \mathbb{N}= \{1, 2, 3, \cdots , \} \) and \(a \in \mathbb{C} ,\) let \(H[a, n]\) be a subclass of \(H(\mathbb{ U } )\) formed by the functions of the form \[ f(z) = z+ a_n z^n + a_{n+1} z^{n+1} + \cdots \] with \(H_0 \equiv H[0, 1]\) and \(H \equiv H[1, 1].\) Suppose that \(A_n\) is a class of all analytic functions of the form

\begin{equation} \label{1.1} f(z) = z +\sum \limits _{k=n+1}^{\infty}a_{n}z^{n} \end{equation}
(1)
in the open unit disk \(\mathbb{ U } \) with \(A_1 = A.\) A function \(f \in H(\mathbb{ U } )\) is univalent if it is a one-to-one function in \(\mathbb{ U } .\) By \(S\), we denote a subclass of \(A\) formed by functions univalent in \(\mathbb{ U } .\) If a function \(f \in A\) maps \(\mathbb{ U } \) onto a convex domain and \(f\) is univalent, then \(f\) is called a convex function. By \[ K= \left\{ f\in A: \Re \left\{1+ \frac{zf''(z)}{f'(z)}\right\}> 0, \ \ z \in \mathbb{ U } \right\}, \] we denote a class of all convex functions defined in \(\mathbb{ U } \) and normalized by \(f(0) = 0\) and \(f'(0) = 1.\)

Let \(f\) and \(F\) be elements of \(H(\mathbb{ U } ).\) A function \(f\) is said to be subordinate to \(F\), if there exists a Schwartz function \(w\) analytic in \(\mathbb{ U } \) with \( w (0) = 0 ~~ \text{ and } ~~ |w(z) | < 1, \ \ z \in \mathbb{ U } , \) such that \(f(z) = F(w(z)).\) In this case, we write \( f(z) \prec F(z) ~~~ \text{ or } ~~ f \prec F.\) Furthermore, if the function \(F\) is univalent in \(\mathbb{ U } ,\) then we get the following equivalence [1,2]:

\[f(z) \prec F(z) \Leftrightarrow f(0) = F(0) ~~~ \text{ and}~~ f(\mathbb{ U } ) \prec F(\mathbb{ U } ). \] The method of differential subordinations (also known as the method of admissible functions) was first introduced by Miller and Mocanu in 1978 [3], and the development of the theory was originated in 1981 [4]. All details can be found in the book by Miller and Mocanu [2]. In recent years, numerous authors studied the properties of differential subordinations (see [5,6,7,8], etc.).

Let \( \Psi : \mathbb{C}^3 \times \mathbb{ U } \rightarrow \mathbb{C} \) and let \(h\) be univalent in \(\mathbb{ U } .\) If \(p\) is analytic in \(\mathbb{ U } \) and satisfies the second-order differential subordination:

\begin{equation} \label{1.2} \Psi \left( p(z), zp'(z), zp''(z); z \right) \prec h(z), \end{equation}
(2)
then \(p\) is called the solution of differential subordination. The univalent function \(q\) is called a dominant of the solution of the differential subordination or, simply, a dominant if \( p \prec q\) for all \(p\) satisfying (2). The dominant \(q_1\) satisfying \(q_1 \prec q\) for all dominants \(q\) of (2) is called the best dominant of (2).

Let us recall lambda function [9] defined by:

\[ \lambda (z, s ) = \sum \limits _{k =2} ^{\infty} \frac{z^k }{ (2k +1) ^k }\] where \(z \in \mathbb{ U } , s \in \mathbb{C},\) when \(|z| < 1, \Re (s) > 1, \) when \(|z| = 1\) and let \(\lambda ^{(-1)} (z, s ) \) be defined such that \[ \lambda (z, s ) * \lambda ^{(-1)} (z, s ) = \frac{1}{(1- z) ^{ \mu +1}}, ~ \mu > -1 .\] We now define \(\left( z \lambda ^{(-1)} (z, s ) \right)\) as: \begin{align*} (z \lambda (z,s)) * \left( z \lambda ^{(-1)} (z, s ) \right) = \frac{z}{(1-z)^{\mu+1}} = z+ \sum \limits _{k =2} ^{\infty} \frac{ ( \mu +1)_{k -1}}{ (k -1) !} z^k , \mu > -1 \end{align*} and obtain the linear operator \( \mathcal{I}_{\mu}^{s} f(z)= \left( z \lambda ^{(-1)} (z, s ) \right) * f(z),\) where \(f \in A, z \in \mathbb{ U } \) and \( \left( z \lambda ^{(-1)} (z, s ) \right) = z+ \sum \limits _{k =2} ^{\infty} \frac{ ( \mu +1)_{k -1} (2k -1) ^s}{ (k -1) !} z^k .\) A simple computation gives us
\begin{align} \label{1.3a} \mathcal{I}_{\mu}^{s} f(z) & =z+ \sum \limits _{k =2} ^{\infty} L(k, \mu, s) a_k z^k , \end{align}
(3)
where
\begin{align}\label{1.4a} L(k, \mu, s) & = \frac{ ( \mu +1)_{k -1} (2k -1) ^s}{ (k -1) !}, \end{align}
(4)
where \(( \mu)_{k } \) is the Pochhammer symbol defined in terms of the Gamma function by: \[( \mu)_{k } = \frac{\Gamma ( \mu+k )}{\Gamma (\mu)}=\left\{ \begin{array}{ll} 1, & \hbox{ if \(k =0 \);} \\ \mu ( \mu +1) \cdots (\mu +k -1), & \hbox{ if \( k \in \mathbb{N} \)}. \end{array} \right . .\]

Definition 1. Let \(\mathfrak{ L} _{\mu , s } ( \varrho ) \) be a class of function \(f \in A\) satisfying the inequality \[ \mathfrak { \Re } \left( \mathcal{I}_ {\mu}^{s}f(z) \right) \geq \varrho ,\] where \( z\in \mathbb{ U }, \ \ 0 \leq \varrho < 1 \) and \(\mathcal{I}_ {\mu}^{s}f(z) \) is the Lambda operator.

Lemma 1. let \(h\) be a convex function with \(h(0)=a \) and let \(\gamma\in \mathbb { C^* }:= \mathbb { C } \setminus \{ 0 \} \) be a complex number with \( \mathfrak { \Re \{ \gamma \} }\geq 0. \) If \( p \in H[a, n] \) and

\begin{equation} \label{1.5} p(z)+\frac{1}{\gamma}zp'(z) \prec h(z), \end{equation}
(5)
then \( p(z) \prec q(z) \prec h(z), \) where \( q(z) = \frac{\gamma}{nz^\frac{\gamma}{n}}\int \limits_{0}^{z}t^\frac{\gamma}{n-1}h(t)dt, ~~ z\in \mathbb{ U }.\) The function \(q\) is convex and is the best dominant for subordination (5).

Lemma 2. [10] Let \( \mathfrak { \Re \{ \mu \} }> 0, ~~n\in \mathbb { N } \) and \( w= \frac{n^2 + |\mu|^2-|n^2-\mu^2|}{4n \mathfrak { \Re \{ \mu \} }} . \) Also, let \(h\) be an analytic function in \(\mathbb { U } \) with \( h(0)=1.\) Suppose that \( \mathfrak { \Re }\left \{ 1+\frac{zh''(z)}{h'(z)} \right \} > -w. \) If \( p(z) = 1+p_nz^n+p_{n+1}z^{n+1}+ \cdots \) is analytic in \(\mathbb { U } \) and

\begin{equation} \label{1.6} p(z)+\frac{1}{\mu}zp'(z) \prec h(z), \end{equation}
(6)
then \( p(z) \prec q(z), \) where \(q\) is a solution of the differential equation \( q(z)+ \frac{n}{\mu}zq'(z) = h(z), ~~q(0)=1, \) given by \( q(z) = \frac{\mu}{nz^\frac{\mu}{n}}\int \limits_{0}^{z}t^\frac{\mu}{n-1}h(t) dt, ~~z\in \mathbb { U } . \) Moreover, \(q\) is the best dominant for the differential subordination (6).

Lemma 3. [11] Let \(r\) be a convex function in \(\mathbb { U } \) and let \( h(z)=r(z)+n\varrho zr'(z), ~~ z\in \mathbb { U } , \) where \(\varrho > 0 ~~ and ~~ n \in \mathbb { N }.\) If \( p(z)= r(0)+p_nz^n+p_{n+1}z^{n+1}+ \cdots , ~~z\in \mathbb { U } , \) is holomorphic in \( \mathbb { U } \) and \( p(z)+\varrho zp'(z)\prec h(z), ~~z\in \mathbb { U } , \) then \( p(z)\prec r(z) \) and this result is sharp.

In the present paper, we use the subordination results from [10] to prove our main results.

2. Main results

Theorem 1. The set \( \mathfrak{ L} _{\mu , s } ( \varrho ) \) is convex.

Proof. Let \( f_j(z)=z+\sum\limits_{k=2}^{\infty}a_{k,j}z^k, ~~ z\in \mathbb { U } , ~~ j= 1, \cdots , m \) be in the class \( \mathfrak{ L} _{\mu , s } ( \varrho ). \) Then, by Definition 1, we get

\begin{equation} \label{2.1} \mathfrak{ \Re } \left \{ (\mathcal{I}_ {\mu}^{s}f(z))' \right \} = \mathfrak{ \Re } \left \{ 1+\sum\limits_{k=2}^{\infty}L(k, \mu, s)a_{k,j}kz^{k-1} \right \} > \varrho . \end{equation}
(7)
For any positive numbers \( \varsigma _1, \varsigma _2, \varsigma _3, \cdots ,\varsigma _m \) such that \( \sum\limits_{j=1}^{m}\varsigma _j=1 , \) it is necessary to show that the function \(h(z)=\sum\limits_{j=1}^{m}\varsigma _jf_j(z)\) is an element of \( \mathfrak{ L} _{\mu , s } ( \varrho ) \), i.e.,
\begin{equation} \label{2.2} \mathfrak{ \Re } \left \{ (\mathcal{I}_ {\mu}^{s}h(z))' \right \} > \varrho . \end{equation}
(8)
Thus, we have
\begin{equation} \label{2.3} \mathcal{I}_ {\mu}^{s} h(z) = z+\sum\limits_{k=2}^{\infty}L(k,\mu,s) \left \{ \sum\limits_{j=1}^{m} \varsigma _ja_{k,j} \right \} z^k. \end{equation}
(9)
If we differentiate (9) with respect to \(z\), then we obtain \[ (\mathcal{I}_ \mu ^s h(z))' = 1+\sum\limits_{k=2}^{\infty}kL(k,\mu,s) \left \{ \sum\limits_{j=1}^{m} \varsigma _ja_{k,j} \right \} z^{k-1}. \] Thus by using (8), we have \begin{align*} \mathfrak{ \Re } \left \{ (\mathcal{I}_ \mu ^s h(z))' \right \} = 1+\sum\limits_{j=1}^{m}\varsigma _j \mathfrak{ \Re } \left \{ \sum\limits_{k=2}^{\infty}kL(k,\mu,s)a_{k,j}z^{k-1} \right \} > 1+\sum\limits_{j=1}^{m}\varsigma _j(\varrho -1) = \varrho . \end{align*} Hence, inequality (7) is true and we arrive at the desired result.

Theorem 2. Let \(q\) be convex function in \(\mathbb {U} \) with \(q(0)=1\) and \( h(z)=q(z) + \frac{1}{\gamma+1}zq'(z), ~~ z\in \mathbb { U } , \) where \(\gamma\) is a complex number with \(\mathfrak{ \Re } \{ {\gamma} \} > -1\). If \(f\in \mathfrak{ L} _{\mu , s } ( \varrho ) \) and \( \aleph =\Upsilon_\gamma f,\) where

\begin{equation} \label{2.4} \aleph (z) = \Upsilon_\gamma f(z) = \frac{\gamma+1}{z^\gamma}\int\limits_0^zt^{\gamma-1}f(t)dt, \end{equation}
(10)
then
\begin{equation} \label{2.5} (\mathcal{I}_ \mu ^s f(z))' \prec h(z) \end{equation}
(11)
implies that \( (\mathcal{I}_ \mu ^s \aleph (z))' \prec q(z) \) and this result is sharp.

Proof. In view of equality (10), we can write

\begin{equation} \label{2.6} z^\gamma \aleph (z)=(\gamma+1)\int\limits_0^zt^{\gamma-1}f(t)dt. \end{equation}
(12)
Differentiating (12) with respect to \(z,\) we obtain \( (\gamma) \aleph (z)+z \aleph ~'(z)=(\gamma+1)f(z). \) Further, by applying the operator \(\mathcal{I}_ \mu ^s\) to the last equation, we get
\begin{equation} \label{2.7} (\gamma)\mathcal{I}_ \mu ^s \aleph(z)+z(\mathcal{I}_ \mu ^s \aleph (z))'=(\gamma+1)\mathcal{I}_ \mu ^s f(z). \end{equation}
(13)
If we differentiate (13) with respect to \(z\), then we find
\begin{equation} \label{2.8} (\mathcal{I}_ \mu ^s \aleph(z))'+\frac{1}{\gamma+1}z(\mathcal{I}_ \mu ^s f(z))''=(\mathcal{I}_ \mu ^s f(z))'. \end{equation}
(14)
By using the differential subordination given by (11) in equality (14), we obtain
\begin{equation} \label{2.9} (\mathcal{I}_ \mu ^s \aleph(z))'+\frac{1}{\gamma+1}z(\mathcal{I}_ \mu ^s f(z))'' \prec h(z). \end{equation}
(15)
We define
\begin{equation} \label{2.10} p(z)=(\mathcal{I}_ \mu ^s \aleph(z))' . \end{equation}
(16)
Hence, as a result of simple computations, we get \begin{align*} p(z) &= \left \{ z+ \sum\limits_{k=2}^{\infty}L(k,\mu,s)\frac{\gamma+1}{\gamma+k}a_kz^k \right \}' = 1+ p_1z+p_2z^2+ \cdots ,~~p\in H[1,1]. \end{align*} By using (16) in subordination (15), we obtain \begin{equation*} p(z)+\frac{1}{\gamma+1}zp'(z)\prec h(z)=q(z)+\frac{1}{\gamma+1}zq'(z),~~ z\in \mathbb{ U }. \end{equation*} If we use Lemma 2, then we write \( p(z)\prec q(z). \) Thus, we obtained the desired result and \(q\) is the best dominant.

Example 1. If we choose \( \gamma=i+1\) and \(q(z)= \frac{1+z}{1-z}, \) in Theorem 2, then we get \( h(z)=\frac{(i+2)-((i+2)z+2)z}{(i+2)(1-z)^2}. \) If \( f \in \mathfrak{ L} _{\mu , s } ( \varrho ) \) and \(\aleph\) is given as \( \aleph(z)= \Upsilon_if(z)=\frac{i+2}{z^{i+1}}\int\limits_0^zt^if(t)dt, \) then, by virtue of Theorem 2, we find \( (\mathcal{I}_ {\mu}^{s}f(z))'\prec h(z) = \frac{(i+2)-((i+2)z+2)z}{(i+2)(1-z)^2},\) implies \((\mathcal{I}_ {\mu}^{s}f(z))' \prec \frac{1+z}{1-z}.\)

Theorem 3. Let \( \mathfrak{ \Re }{ \left \{ \gamma \right \} } > -1 \) and \( w=\frac{1+|\gamma+1|^2-|\gamma^2+2\gamma|}{4\mathfrak{ \Re }{ \left \{ \gamma+1 \right \}}}. \) Suppose that \(h\) is an analytic function in \(\mathbb{ U } \) with \(h(0)=1\) and that \( \mathfrak { \Re }\left \{ 1+\frac{zh''(z)}{h'(z)} \right \} > -w. \) If \( f \in \mathfrak{ L} _{\mu , s } ( \varrho ) \) and \( \aleph = \Upsilon_\mu ^s f, \) where \(\aleph\) is defined by (10), then

\begin{equation} \label{2.11} (\mathcal{I}_ \mu ^s f(z))' \prec h(z) \end{equation}
(17)
implies that \( (\mathcal{I}_ \mu ^s \aleph(z))' \prec q(z), \) where \(q\) is the solution of the differential equation \( h(z) = q(z) + \frac{1}{\gamma+1}zq'(z), ~~q(0)=1, \) given by \( q(z) = \frac{\gamma+1}{z^{\gamma+1}}\int\limits_0^zt^\gamma f(t)dt. \) Moreover, \(q\) is the best dominant for subordination (17).

Proof. If we choose \( n=1 \) and \( \mu={\gamma+1} \) in Lemma 1, then the proof is obtained by means of the proof of Theorem 3.

Theorem 4. Let

\begin{equation} \label{2.12} h(z)=\frac{1+(2\varrho - 1)z}{1+z}, ~~ 0\leq \varrho < 1 \end{equation}
(18)
be convex in \( \mathbb { U } \) with \( h(0)=1 .\) If \( f\in A \) and verifies the differential subordination \( (\mathcal{I}_ \mu ^s f(z))' \prec h(z), \) then \( (\mathcal{I}_ \mu ^s \aleph (z))' \prec q(z)=(2\varrho -1)+\frac{2(1-\varrho )(\gamma+1) \tau(\gamma)}{z^{\gamma+1}}, \) where \(\tau\) is given by the formula
\begin{equation} \label{2.13} \tau(\gamma)=\int\limits_0^z\frac{t^\gamma}{t+1}dt \end{equation}
(19)
and \(\aleph\) is given by equation (10). The function \(q\) is convex and is the best dominant.

Proof. If \( h(z)=\frac{1+(2\varrho - 1)z}{1+z}, ~~ 0\leq \varrho < 1, \) then \( h\) is convex and, in view of Theorem 3, we can write \( (\mathcal{I}_ \mu ^s \aleph(z))' \prec q(z). \) Now, by using Lemma 1, we get \begin{align*} q(z) = \frac{\gamma+1}{z^{\gamma+1}}\int\limits_0^zt^\gamma h(t)dt = \frac{\gamma+1}{z^{\gamma+1}}\int\limits_0^zt^\gamma \left \{ \frac{1+(2\varrho -1)t}{1+t} \right \} dt = (2\varrho -1)+\frac{2(1-\varrho )(\gamma+1)}{z^{\gamma+1}}\tau(\gamma), \end{align*} where \(\tau\) is given by (19). Hence, we obtain \begin{equation*} (\mathcal{I}_ \mu ^s \aleph (z))' \prec q(z)=(2\varrho -1)+\frac{2(1-\varrho )(\gamma+1) \tau(\gamma)}{z^{\gamma+1}}. \end{equation*} The function \( q\) is convex. Moreover, it is the best dominant. Hence the theorem is proved.

Theorem 5. If \( 0 \leq \varrho < 1, 0 \leq \mu < 1, \delta \geq 0, \mathfrak{ \Re } \{ {\gamma} \} > -1, \) and \(\aleph= \Upsilon_\gamma f \) is defined by (10), then \( \Upsilon_\gamma( \mathfrak{ L} _{\mu , s } ( \varrho )) \subset \mathfrak{ L} _{\mu , s } ( \rho ), \) where

\begin{equation} \label{2.14} \rho= \min \limits _{ |z|=1 }\mathfrak { \Re } \{ {q(z)} \} = \rho(\gamma,\varrho )=(2\varrho -1)+2(1-\varrho )(\gamma+1)\tau(\gamma) \end{equation}
(20)
and \(\tau\) is given by (19).

Proof. Assume that \(h\) is given by equation (18), \(f \in \mathfrak{ L} _{\mu , s } ( \varrho ), \) and \( \aleph = \Upsilon_\gamma f \) is defined by (10). Then \(h\) is convex and, by Theorem 3, we deduce

\begin{equation} \label{2.15} (\mathcal{I}_ \mu ^s \aleph (z))' \prec q(z)=(2\varrho -1)+\frac{2(1-\varrho )(\gamma+1) \tau(\gamma)}{z^{\gamma+1}}, \end{equation}
(21)
where \(\tau\) is given by (19). Since \( q \) is convex, \( q( \mathbb{ U }) \) is symmetric about the real axis, and \( \mathfrak{ \Re } \{ {\gamma} \} > -1,\) we find \begin{align*} \mathfrak{ \Re } \left \{ (\mathcal{I}_ \mu ^s \aleph(z))' \right \} \geq \min \limits _{|z|=1} \mathfrak{ \Re } \{ {q(z)} \} = \mathfrak{ \Re } \{ q(1) \}= \rho(\gamma,\varrho ) = (2\varrho -1)+2(1-\varrho )(\gamma+1)(1-\varrho )\tau(\gamma). \end{align*} It follows from inequality (21) that \( \Upsilon_\gamma( \mathfrak{ L} _{\mu , s } ( \varrho )) \subset \mathfrak{ L} _{\mu , s } ( \rho ), \) where \(\rho\) is given by (20). Hence the theorem is proved.

Theorem 6. Let \(q\) be a convex function with \( q(0) = 1 \) and \(h\) be a function such that \( h(z)= q(z) + zq'(z), ~~ z\in \mathbb{ U }. \) If \( f \in A,\) then the subordination

\begin{equation} \label{2.16} (\mathcal{I}_ \mu ^s f(z))' \prec h(z) \end{equation}
(22)
implies that \( \frac{\mathcal{I}_ \mu ^s f(z)}{z} \prec q(z), \) and the result is sharp.

Proof. Let

\begin{equation} \label{2.17} p(z)=\frac{\mathcal{I}_ \mu ^s f(z)}{z}. \end{equation}
(23)
Differentiating (23), we find \( (\mathcal{I}_ \mu ^s f(z))' = p(z)+zp'(z). \) We now compute \( p(z) \). This gives
\begin{align} p(z) =\frac{\mathcal{I}_ \mu ^s f(z)}{z} = \frac{z+\sum\limits_{k=2}^{\infty}L(k,\mu,s)a_kz^k}{z} =1+p_1z+p_2z^2+ \cdots , ~~p \in H[1,1]. \end{align}
(24)
By using (24) in subordination (22), we find \( p(z)+zp'(z)\prec h(z)=q(z)+zq'(z). \) Hence, by applying Lemma 3, we conclude that \( p(z) \prec q(z) \) i.e., \( \frac{\mathcal{I}_ \mu ^s f(z)}{z} \prec q(z). \) This result is sharp and \(q\) is the best dominant. Hence the theorem is proved.

Example 2. If we take \( \mu = 0 \) and \( s = 1 \) in equality (4) and \( q(z) = \frac{1}{1-z} \) in Theorem 5, then \( h(z)=\frac{1}{(1-z)^2} \) and

\begin{equation} \label{2.19} I_0^1 f(z)=z+\sum\limits_{k=2}^{\infty}\frac{(2k-1)}{(k-1)!}a_kz^k. \end{equation}
(25)
Differentiating (25) with respect to \(z\), we get \begin{align*} (I_0^1f(z))' = 1+ \sum\limits_{k=2}^{\infty}\frac{(2k-1)}{(k-1)!}a_kz^{k-1} = 1+p_1z+p_2z^2+ \cdots , ~~ p \in H[1,1]. \end{align*} By using Theorem 5, we find \( (I_0^1f(z))'\prec h(z) = \frac{1}{(1-z)^2}. \) This yields \( \frac{I_0^1f(z)}{z} \prec q(z) = \frac{1}{1-z}. \)

Theorem 7. Let \( h(z)=\frac{1+(2\varrho -1)z}{1+z}, ~~ z\in \mathbb{ U } \) be convex in \( \mathbb { U } \) with \( h(0)=1\) and \( 0 \leq \varrho < 1.\) If \( f \in A \) satisfies the differential subordination

\begin{equation} \label{2.20} (\mathcal{I}_ \mu ^s f(z))' \prec h(z), \end{equation}
(26)
then \( \frac{\mathcal{I}_ \mu ^s f(z)}{z} \prec q(z)= (2\varrho -1)+\frac{2(1-\varrho )ln(1+z)}{z}. \) The function \(q\) is convex and, in addition, it is the best dominant.

Proof. Let

\begin{equation} \label{2.21} p(z)=\frac{\mathcal{I}_ \mu ^s f(z)}{z}= 1+p_1z+p_2z^2+ \cdots , ~~ p \in H[1,1]. \end{equation}
(27)
Differentiating (27), we find
\begin{equation} \label{2.22} (\mathcal{I}_ \mu ^s f(z))' = p(z) + zp'(z). \end{equation}
(28)
In view of (28), the differential subordination (26) becomes \( (\mathcal{I}_ \mu ^s f(z))' \prec h(z) = \frac{1+(2\varrho -1)z}{1+z},\) and by using Lemma 1, we deduce \( p(z)\prec q(z)=\frac{1}{z}\int h(t)dt = (2\varrho -1)+\frac{2(1-\varrho )ln(1+z)}{z}. \) Now, by virtue of relation (27) we obtained the desired result.

Corollary 1. If \( f \in \mathfrak{ L} _{\mu , s } ( \varrho ), \) then \( \mathfrak { \Re } \left( \frac{\mathcal{I}_ \mu ^s f(z)}{z} \right) > (2\varrho -1)+2(1-\varrho )ln(2). \)

Proof. If \( f \in \mathfrak { L} _{\mu , s } ( \varrho ), \) then it follows from Definition 1 that \( \mathfrak { \Re } \left \{ (\mathcal{I}_ \mu ^s f(z))' \right \} > \varrho , ~~ z \in \mathbb { U }, \) which is equivalent to \( (\mathcal{I}_ \mu ^s f(z))' \prec h(z) = \frac{1+(2\varrho -1)z}{1+z}. \) Now, by using Theorem 7, we obtain \begin{equation*} \frac{\mathcal{I}_ \mu ^s f(z)}{z} \prec q(z) = (2\varrho -1)+\frac{2(1-\varrho )ln(1+z)}{z}. \end{equation*} Since \(q\) is convex and \(q( \mathbb { U } ) \) is symmetric about the real axis, we conclude that \begin{equation*} \mathfrak { \Re } \left( \frac{\mathcal{I}_ \mu ^s f(z)}{z} \right) > \mathfrak { \Re } (q(1)) = (2\varrho -1)+2(1-\varrho )ln(2). \end{equation*}

Theorem 8. Let \(q\) be a convex function such that \(q(0)=1\) and \(h\) be the function given by the formula \( h(z)=q(z)+zq'(z), ~~z \in \mathbb { U }. \) If \( f \in A \) and verifies the differential subordination

\begin{equation} \label{2.23} \left \{ \frac{z \mathcal{I}_ \mu ^s f(z)}{\mathcal{I}_ \mu ^s \aleph(z)} \right \}' \prec h(z), ~~ z \in \mathbb{ U }, \end{equation}
(29)
then \( \frac{\mathcal{I}_ \mu ^s f(z)}{\mathcal{I}_ \mu ^s \aleph(z)} \prec q(z), ~~ z \in \mathbb { U }, \) and this result is sharp.

Proof. For function \( f \in A, \) given by Equation (1), we get \begin{equation*} \mathcal{I}_ \mu ^s \aleph (z) = z + \sum\limits_{k=2}^{\infty}L(k,\mu,s)\frac{\gamma+1}{k+\gamma}a_kb_kz^k, ~~z\in \mathbb { U }. \end{equation*} We now consider the function \begin{align*} p(z)=\frac{\mathcal{I}_ \mu ^s f(z)}{\mathcal{I}_ \mu ^s \aleph (z)} = \frac{z+\sum\limits_{k=2}^{\infty}L(k,\mu,s )a_kb_kz^k}{z+\sum\limits_{k=2}^{\infty}L(k,\mu,s)\frac{\gamma+1}{k+\gamma}a_kb_kz^k} = \frac{1+\sum\limits_{k=2}^{\infty}L(k,\mu,s)a_kb_kz^{k-1}}{1+\sum\limits_{k=2}^{\infty}L(k,\mu,s)\frac{\gamma+1}{k+\gamma}a_kb_kz^{k-1}}. \end{align*} In this case, we get \begin{equation*} (p(z))'=\frac{(\mathcal{I}_ \mu ^s f(z))'}{\mathcal{I}_ \mu ^s \aleph(z)} - p(z) \frac{(\mathcal{I}_ \mu ^s \aleph(z))'}{\mathcal{I}_ \mu ^s \aleph(z)}. \end{equation*} Then

\begin{equation} \label{2.24} p(z)+zp'(z)= \left \{ \frac{z \mathcal{I}_ \mu ^s f(z)}{\mathcal{I}_ \mu ^s \aleph(z)} \right \}' , ~~ z \in \mathbb{ U }. \end{equation}
(30)
By using relation (30) in inequality (29), we obtain \( p(z)+zp'(z)\prec h(z)=q(z)+zq'(z) \) and, by virtue of Lemma 3, \( p(z)\prec q(z), \) i.e., \( \frac{\mathcal{I}_ \mu ^s f(z)}{\mathcal{I}_ \mu ^s \aleph(z)} \prec q(z). \) Hence the theorem is proved.

Acknowledgments

The authors warmly thank the referees for the careful reading of the paper and their comments.

Author Contributions

All authors contributed equally to the writing of this paper. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bulboaca, T. (2005). Differential subordinations and superordinations: Recent results. Casa Cartii de Stiinta. [Google Scholor]
  2. Miller, S. S., & Mocanu, P. T. (2000). Differential Subordinations: Theory and Applications. CRC Press. [Google Scholor]
  3. Miller, S. S., & Mocanu, P. T. (1978). Second order differential inequalities in the complex plane. Journal of Mathematical Analysis and Applications, 65(2), 289-305. [Google Scholor]
  4. Miller, S. S., & Mocanu, P. T. (1981). Differential subordinations and univalent functions. The Michigan Mathematical Journal, 28(2), 157-172. [Google Scholor]
  5. Akgül, A. (2017). On second-order differential subordinations for a class of analytic functions defined by convolution. Journal of Nonlinear Sciences and Application, 10, 954-963.[Google Scholor]
  6. Lupas, A. A. (2012). Certain differential subordinations using Salagean and Ruscheweyh operators. Acta Universitatis Apulensis, 29, 125-129. [Google Scholor]
  7. Bulut, S. (2014). Some applications of second-order differential subordination on a class of analytic functions defined by Komatu integral operator. International Scholarly Research Notices, 2014, Artical ID 606235. [Google Scholor]
  8. Oros, G. I., & Oros, G. (2008). On a class of univalent functions defined by a generalized Salagean operator. Complex Variables and Elliptic Equations, 53(9), 869-877. [Google Scholor]
  9. Spanier, J., & Oldham, K. B. (1987). An Atlas of Functions. New York: Hemisphere publishing corporation. [Google Scholor]
  10. Hallenbeck, D. J., & Ruscheweyh, S. (1975). Subordination by convex functions. Proceedings of the American Mathematical Society, 52(1), 191-195.[Google Scholor]
  11. Oros, G., & Oros, G. I. (2003). A class of holomorphic functions II. Libertas Mathematica, 23, 65-68. [Google Scholor]
  12. Salagean, G. S. (1983). Subclasses of univalent functions. In Complex Analysis—Fifth Romanian-Finnish Seminar (pp. 362-372). Springer, Berlin, Heidelberg. [Google Scholor]
]]>
On properties of inner product type integral transformers https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/on-properties-of-inner-product-type-integral-transformers/ Wed, 23 Dec 2020 14:35:40 +0000 https://old.pisrt.org/?p=4820
OMA-Vol. 4 (2020), Issue 2, pp. 160 - 169 Open Access Full-Text PDF
Benard Okelo
Abstract: In this paper, we give characterizations of certain properties of inner product type integral transformers. We first consider unitarily invariant norms and operator valued functions. We then give results on norm inequalities for inner product type integral transformers in terms of Landau inequality, Grüss inequality. Lastly, we explore some of the applications in quantum theory.
]]>

Open Journal of Mathematical Analysis

On properties of inner product type integral transformers

Benard Okelo
Department of Pure and Applied Mathematics, Jaramogi Oginga Odinga University of Science and Technology, Box 210-40601, Bondo-Kenya.; bnyaare@yahoo.com

Abstract

In this paper, we give characterizations of certain properties of inner product type integral transformers. We first consider unitarily invariant norms and operator valued functions. We then give results on norm inequalities for inner product type integral transformers in terms of Landau inequality, Grüss inequality. Lastly, we explore some of the applications in quantum theory.

Keywords:

Norm inequality, unitarily invariant norm, operator valued function, norm ideal, inner product type integral transformer.

1. Introduction

Let \(\mathcal{H}\) be an infinite dimensional complex Hilbert space and \(\mathcal{B(H)}\) be the algebra of all bounded linear operators on \(\mathcal{H}.\) In this paper, we discuss various types of norm inequalities for inner product type integral transformers in terms of Landau type inequality, Grüss type inequality and Cauchy-Schwarz type inequality. We shall also consider the applications in quantum theory. We begin by the following definition:

Definition 1. Grüss inequality states that if \(f\) and \(g\) are integrable real functions on \([a,b]\) such that \(C\leq f(x)\le D\) and \(E\leq g(x)\le F\) hold for some real constants \(C,D,E,F\) and for all \(x\in[a,b]\), then

\begin{equation} \left|\frac1{b-a}\int_a^bf(x)g(x)dx-\frac1{(b-a)^2}\int_a^b f(x)dx\int_a^b g(x)dx\right| \leq\frac14(D-C)(F-E).\label{grisovaca} \end{equation}
(1)
Inequality (1) is very interesting to many researchers and it has been considered in many studies whereby conditions on functions are varied to give different estimates (see [1] and references therein). More on this inequality (and the classical one [2]) are discussed in the sequel.

Next, we discuss a very important definition of inner product type integral \((i.p.t.i)\) transformer which is key to our study.

Definition 2. Consider weakly \(\mu^*\)-measurable operator valued \((o.v)\) functions \(A, B:\Omega\rightarrow \mathcal{B(H)}\) and for all \(X\in \mathcal{B(H)}\). Let the function \(t\rightarrow A_t X B_t\) be also weakly \(\mu^*\)-measurable. If these functions are Gel'fand integrable for all \(X\in \mathcal{B(H)}\), then the inner product type linear transformation \(X\to\int_\Omega A_t X B_t dt\) is called an inner product type integral \((i.p.t.i)\) transformer on \(\mathcal{B(H)}\) and denoted by \(\int_\Omega A_t \otimes B_t dt\) or \({\mathcal I}_{A,B}\).

Remark 1. If \(\mu\) is the counting measure on \(\mathbb N\) then such transformers are known as elementary operators whose certain properties have been studied in details (see [3] and the references therein).

2. Preliminaries

In this section, we consider a special type of norms called the unitarily invariant norm. We give its description in details which will be useful in the sequel. Let \(\mathcal{C}_\infty(\mathcal{H})\) denote the space of all compact linear operators acting on a separable, complex Hilbert space \(\mathcal{H}\). Each symmetric gauge function \(\Phi,\) denoted by \((s.g.)\), on sequences gives rise to a unitarily invariant \((u.i)\) norm on operators defined by \(\left\|X\right\|_\Phi=\Phi(\{s_n(X)\}_{n=1}^\infty)\) with \(s_1(X)\ge s_2(X)\ge...\) being the singular values of \(X\), i.e., the eigenvalues of \(|X|=(X^*X)^\frac12.\) We denote any such norm by the symbol \(\left|\left|\left|\cdot\right|\right|\right|\), which is therefore defined on a naturally associated norm ideal \(\mathcal{C}_{\left|\left|\left|\cdot\right|\right|\right|}(\mathcal{H})\) of \(\mathcal{C}_\infty(\mathcal{H})\) and satisfies the invariance property \( |\|UXV|\|=|\|X|\|\) for all \(X\in\mathcal{C}_{\left|\left|\left|\cdot\right|\right|\right|}(\mathcal{H})\) and for all unitary operators \(U,V\in \mathcal{B(H)}\). One of the well known among \(u.i.\) norms are the Schatten \(p\)-norms defined for \(1\le p< \infty\) as \(\|X\|_p=\sqrt[p]{\,\sum_{n=1}^\infty s_n^p(X)}\), while \(\|X\|_\infty =\|X\|=s_1(X)\) coincides with the operator norm \(\|X\|\). Minimal and maximal \(u.i.\) norm are among Schatten norms, i.e., \(\|X\|_\infty\le|\|X\||\le\|X\|_1\) for all \(X\in\mathcal{C}_1(\mathcal{H})\) (see inequality (IV.38) [4]). For \(f,g\in\mathcal{H}\), we will denote by \(g^*\otimes f\) one dimensional operator \((g^*\otimes f)h=\langle h,g\rangle f\) for all \(h\in\mathcal{H}\) and it is known that the linear span of \(\{g^*\otimes f\,|\, f,g\in \mathcal{H}\}\) is dense in each of \(\mathcal{C}_p(\mathcal{H})\) for \(1\le p\le\infty\). Schatten \(p\)-norms are also classical examples of \(p\)-reconvexized norms. Namely, any \(u.i.\) norm \(\|.\|_\Phi\) could be \(p\)-reconvexized for any \(p\ge1\) by setting \(\|A\|_{\Phi^{(p)}} = \| |A|^p\|_{\Phi}^{\frac1p}\) for all \(A\in \mathcal{B(H)}\) such that \(|A|^p\in \Phi(\mathcal{H})\). For the proof of the triangle inequality and other properties of these norms, see [2] and for the characterization of the dual norm for \(p\)-reconvexized, see Theorem 2.1 [2].

The set \(\mathcal{C}_{|||\cdot|||}=\{A \in \mathcal{K}(\mathcal{H}) : \left\vert \left\vert \left\vert A \right\vert \right\vert \right\vert < \infty \}\) is a closed self-adjoint ideal \(\mathcal{J}\) of \(\mathcal{B}( \mathcal{H})\) containing finite rank operators. It enjoys the following properties. First, for all \(A,B\in \mathcal{B(H)}\) and \(X \in \mathcal{J}\), \( \left\vert \left\vert \left\vert AXB\right\vert \right\vert \right\vert \leq \left\vert \left\vert A\right\vert \right\vert \ \left\vert \left\vert \left\vert X\right\vert \right\vert \right\vert \ \left\vert \left\vert B\right\vert \right\vert\,. \) Secondly, if \(X\) is a rank one operator, then \( \left\vert \left\vert \left\vert X\right\vert \right\vert \right\vert =\|X\|\,. \) The Ky Fan norm as an example of unitarily invariant norms is defined by \(\| A\| _{(k)}=\sum_{j=1}^{k}s_{j}(A)\) for \(k=1,2,\ldots\). The Ky Fan dominance Theorem [5] states that \(\| A\| _{(k)}\leq \| B\| _{(k)}\,\,(k=1,2,\ldots )\) if and only if \(|||A||| \leq |||B|||\) for all unitarily invariant norms \(|||\cdot|||\), see [6] for more information on unitarily invariant norms. The inequalities involving unitarily invariant norms have been of special interest (see [5] and the references therein).

Lemma 1. Let \(\mathcal{T}\) and \(\mathcal{S}\) be linear mappings defined on \(\mathcal{C}_\infty(\mathcal{H}).\) If \(\|\mathcal{T}X\|\le\|\mathcal{S}X\|\mbox{ for all }X\in \mathcal{C}_\infty(\mathcal{H}), \;\|\mathcal{T}X\|_1\le\|\mathcal{S}X\|_1\mbox{ for all }X\in \mathcal{C}_\infty(\mathcal{H})\), then \( \mathcal{T}X\le\mathcal{S}X\) for all unitarily invariant norms.

Proof. The norms \(\|\cdot\|\) and \(\|\cdot\|_1\) are dual to each other in the sense that \(\|X\|=\sup_{\|Y\|_1=1}|tr(XY)|\) and \( \|X\|_1=\sup_{\|Y\|=1}|tr(XY)|.\) Hence \(\|\mathcal{T}^*X\|\le\|\mathcal{S}^*X\|\) and \(\|\mathcal{T}^*X\|_1\le\|\mathcal{S}^*X\|_1\). Consider the Ky Fan norm \(\|\cdot\|_{(k)}\). Its dual norm is \(\|\cdot\|_{(k)}^\sharp=\max\{\|\cdot\|,(1/k)\|\cdot\|_1\}\). Thus, by duality, \(\|\mathcal{T}X\|_{(k)}\le\|\mathcal{S}X\|_{(k)}\) and the result follows by Ky Fan dominance property [6].

An operator \(A\in \mathcal{B(H)}\) is called \(G_{1}\) operator if the growth condition

\[ \left\Vert (z-A)^{-1}\right\Vert =\frac{1}{{dist}(z,\sigma (A))} \] holds for all \(z\) not in the spectrum \(\sigma (A)\) of \(A\). Here \({dist}(z,\sigma (A))\) denotes the distance between \(z\) and \(\sigma (A)\). It is known that hyponormal (in particular, normal) operators are \( G_{1}\) operators [4].

Let \(A, B\in \mathcal{B(H)}\) and let \(f\) be a function which is analytic on an open neighborhood \( \Omega \) of \(\sigma (A)\) in the complex plane. Then \(f(A)\) denotes the operator defined on \(\mathcal{H}\) by \( f(A)=\frac{1}{2\pi i}\int\limits_{C}f(z)(z-A)^{-1}dz, \label{4} \) called the Riesz-Dunford integral, where \(C\) is a positively oriented simple closed rectifiable contour surrounding \(\sigma (A)\) in \(\Omega \) (see [2] and the references therein). The spectral mapping theorem asserts that \(\sigma (f(A))=f(\sigma (A))\). Throughout this paper, \(\mathbb{D}=\{z\in\mathbb{C}:\left\vert z\right\vert < 1\}\) denotes the unit disk, \(\partial\mathbb{D}\) stands for the boundary of \(\mathbb{D}\) and \(d_{A}={dist}(\partial\mathbb{D},\sigma (A))\). In addition, we adopt the notation \(\mathfrak{H}=\{f: \mathbb{D}\to \mathbb{C}: f \mbox{   is analytic}, \Re(f)>0 \mbox{   and  } f(0)=1\}.\)

In this work, we present some upper bounds for \(|||f(A)Xg(B)\pm X|||\), where \(A, B\) are \(G_{1}\) operators, \(|||\cdot|||\) is a unitarily invariant norm and \(f, g\in \mathfrak{H}\). Further, we find some new upper bounds for the the Schatten \(2\)-norm of \(f(A)X\pm Xg(B)\). Up-to this juncture, we find some upper estimates for \(|||f(A)Xg(B)+ X|||\) in terms of \(|||\,|AXB|+|X|\,|||\) and \(|||f(A)Xg(B)- X|||\) in terms of \(|||\,|AX|+|XB|\,|||\), where \(A, B\) are \(G_{1}\) operators and \(f, g\in \mathcal{H}\).

Proposition 1. If \(A,B\in \mathcal{B(H)}\) are \(G_{1}\) operators with \(\sigma (A)\cup \sigma (B)\subset\mathbb{D}\) and \(f, g \in \mathcal{H}\), then for every \(X\in \mathcal{B(H)}\) and for every unitarily invariant norm \(\left\vert \left\vert \left\vert\cdot \right\vert \right\vert \right\vert \), the inequality \( \left\vert \left\vert \left\vert f(A)Xg(B)+X\right\vert \right\vert \right\vert \leq \frac{2\sqrt{2}}{d_{A}d_{B}} \left\vert\left\vert \left\vert\,|AXB|+|X|\,\right\vert \right\vert \right\vert \label{5} \) holds.

Proof. From the Herglotz representation Theorem [1], it follows that \(f\in \mathcal{H}\) can be represented as

\begin{eqnarray} f(z)=\int\limits_{0}^{2\pi }\frac{e^{i\alpha}+z}{e^{i\alpha}-z}d\mu(\alpha)+i\Im f(0)=\int\limits_{0}^{2\pi }\frac{e^{i\alpha}+z}{e^{i\alpha}-z}d\mu(\alpha), \label{6} \end{eqnarray}
(2)
where \(\mu \) is a positive Borel measure on the interval \([0,2\pi ]\) with finite total mass \(\int\limits_{0}^{2\pi }d\mu(\alpha)=f(0)=1\). Similarly \(g(z)=\int\limits_{0}^{2\pi }\frac{e^{i\alpha}+z}{e^{i\alpha}-z}d\nu(\alpha)\) for some positive Borel measure \(\nu\) on the interval \([0,2\pi ]\) with finite total mass \(1\). We have \begin{eqnarray*} f(A)Xg(B)+X=\int\limits_{0}^{2\pi }\int\limits_{0}^{2\pi }\left[ \left( e^{i\alpha}-A\right)^{-1} \left( e^{i\alpha}+A\right)X\left( e^{i\beta}+B\right) \left( e^{i\beta}-B\right)^{-1}+X\right] d\mu(\alpha)d\nu(\beta). \end{eqnarray*} By some computation, we have \begin{eqnarray} \label{s1} \left\vert \left\vert \left\vert f(A)Xg(B)+X\right\vert \right\vert \right\vert \nonumber\leq \int\limits_{0}^{2\pi }\int\limits_{0}^{2\pi }2\left\Vert \left( e^{i\alpha }-A\right)^{-1}\right\Vert \left\vert \left\vert \left\vert AXB+ e^{i\alpha}Xe^{i\beta}\right\vert \right\vert \right\vert \left\Vert \left( e^{i\alpha }-B\right)^{-1}\right\Vert d\mu(\alpha)d\nu(\beta).\ \end{eqnarray} Since \(A\) and \(B\) are \(G_{1}\) operators, we deduce that
\begin{eqnarray} \label{s2} \left\vert \left\vert \left( e^{i\alpha}-A\right)^{-1}\right\vert \right\vert =\frac{1}{{dist}(e^{i\alpha},\sigma (A))}\leq \frac{1}{{dist}(\partial\mathbb{D},\sigma (A))}=\frac{1}{d_{A}}, \end{eqnarray}
(3)
and similarly \( \left\vert \left\vert \left( e^{i\beta}-B\right)^{-1}\right\vert \right\vert \leq \frac{1}{d_{B}}. \) Now, for every positive operators \(C, D\), every non-negative operator monotone function \(h(t)\) on \([0,\infty)\) and every unitarily invariant norm \(|||\cdot|||\), we have \(|||h(A+B)||| \leq |||h(A)+h(B)|||\). Now, from the Ky Fan dominance theorem, we infer that
\begin{eqnarray} \label{s4} \left\vert\left\vert \left\vert AXB+ e^{i\alpha}Xe^{i\beta} \right\vert \right\vert \right\vert \leq \sqrt{2} \left\vert\left\vert \left\vert\, |AXB|+|X| \, \right\vert \right\vert \right\vert. \end{eqnarray}
(4)
Therefore, it follows from inequality (3) and Equation (4) that \[ \left\vert \left\vert \left\vert f(A)Xg(B)+X\right\vert \right\vert \right\vert \leq \frac{2\sqrt{2}}{d_{A}d_{B}} \left\vert\left\vert \left\vert\,|AXB|+|X|\,\right\vert \right\vert \right\vert, \] which completes the proof.

Theorem 1. Let \(f, g\in \mathcal{H}\) and \(A\in\mathcal{B(H)}\) be a \(G_{1}\) operator with \(\sigma (A)\subset\mathbb{D}\). The inequality \( \left\vert \left\vert \left\vert f(A)Xg(A^*)+X\right\vert \right\vert \right\vert \leq \frac{2}{d_{A}^2} \left\vert\left\vert \left\vert\ A|X|A^*+|X|\ \right\vert \right\vert \right\vert \) holds for every normal operator \(X\in\mathcal{B(H)}\) commuting with \(A\) and for every unitarily invariant norm \(\left\vert \left\vert \left\vert\cdot \right\vert \right\vert \right\vert \).

Proof. Let \(X\) and \(AXB\) be normal. Since \(||| C+D |||\leq |||\,|C|+|D|\,|||\) for any normal operators \(C\) and \(D\), the constant \(\sqrt{2}\) can be reduced to \(1\) in Equation (4). Now from Fuglede-Putnam theorem, if \(A\in \mathcal{B(H)}\) is an operator, \(X\in {\mathcal(B)}({\mathcal(H)})\) is normal and \(AX=XA\), then \(AX^*=X^*A\). Thus if \(X\) is a normal operator commuting with a \(G_{1}\) operator \(A\), then \(AXA^*\) is normal, \(|AXA^*|=A|X|A^*\) and \(A^*\) is a \(G_1\) operator with \(d_{A^*}=d_A\). By Proposition 1 the proof is complete.

Next, letting \(A=B\) in Proposition 1, we obtain the following result.

Corollary 1. Let \(f, g\in \mathcal{H}\) and \(A\in \mathcal{B(H)}\) be a \(G_{1}\) operator with \(\sigma (A)\subset\mathbb{D}\). Then \( \left\vert \left\vert \left\vert f(A)Xg(A)-X\right\vert \right\vert \right\vert \leq \frac{2\sqrt{2}}{d_{A}^2} \left\vert\left\vert \left\vert \,|AX|+|XA|\,\right\vert \right\vert \right\vert \) for every \(X\in\mathcal{B}(\mathcal{H})\) and for every unitarily invariant norm \(\left\vert \left\vert \left\vert \cdot \right\vert \right\vert \right\vert \).

Setting \(X=I\) in Proposition 1 again, we obtain the following result.

Corollary 2. Let \(f, g\in \mathfrak{H}\) and \(A,B\in\mathbb{M}_n\) be \(G_{1}\) matrices such that \(\sigma (A)\cup \sigma (B)\subset\mathbb{D}\). Then \( \left\vert \left\vert \left\vert f(A)g(B)+I\right\vert \right\vert \right\vert \leq \frac{2\sqrt{2}}{d_{A}d_{B}} \left\vert\left\vert \left\vert\,|AB|+I\,\right\vert \right\vert \right\vert \) for every unitarily invariant norm \(\left\vert \left\vert \left\vert \cdot \right\vert \right\vert \right\vert. \)

Corollary 3. If \(A\in \mathcal{B}(\mathcal{H})\) is self-adjoint and \(f\) is a continuous complex function on \(\sigma(A)\), then \(f(UAU^*)=Uf(A)U^*\) for all unitaries \(U\).

Proof. By the Stone-Weierstrass theorem, there is a sequence \((p_n)\) of polynomials uniformly converging to \(f\) on \(\sigma(A)\). Hence, \[f(UAU^*)=\lim_np_n(UAU^*)=U(\lim_np_n(A))U^*=Uf(A)U^*.\] We note that \(\sigma(UAU^*)=\sigma(A)\).

We conclude this section by presenting some inequalities involving the Hilbert-Schmidt norm \(\|\cdot\|_2.\)

Theorem 2. Let \(A,B\in\mathbb{M}_n\) be Hermitian matrices satisfying \(\sigma(A)\cup \sigma(B)\subset \mathbb{D}\) and let \(f, g\in \mathfrak{H}\). Then \( \|f(A)X\pm Xg(B)\|_2\leq \left\|\frac{X+|A|X}{d_A}+\frac{X+X|B|}{d_B}\right\|_2. \)

Proof. Let \(A=UD(\nu_j)U^*\) and \(B=VD(\mu_k)V^*\) be the spectral decomposition of \(A\) and \(B\) and let \(Y=U^*XV:=[y_{jk}].\) Noting that \(|e^{i\alpha}-\lambda_j|\geq d_A\) and \(|e^{i\beta}-\mu_k|\geq d_B,\) we have from [7] that \begin{align*} \|f(A)X\pm Xg(B)\|_2^2&=\sum_{j,k}|f(\lambda_j)\pm g(\mu_k)|^2|y_{jk}|^2\\&\leq\sum_{j,k}\left(\frac{1+|\lambda_j|}{d_A}+\frac{1+|\mu_k|}{d_B}\right)^2|y_{jk}|^2\\ &=\left\|\frac{X+|A|X}{d_A}+\frac{X+X|B|}{d_B}\right\|_2^2, \end{align*} which completes the proof.

3. Operators in function spaces

In this section, we present some results on operator valued functions. From [5], if \((\Omega,\mathcal{M,}\mu)\) is a measure space, for a \(\sigma\)-finite measure \(\mu\) on \(\mathcal{M}\), the mapping \(\mathcal{A}:\Omega\rightarrow \mathcal{B(H)}\) will be called \([\mu]\) weakly\(^{*}\)-measurable if a scalar valued function \(t \rightarrow tr (A_{t} Y)\) is measurable for any \(Y\in\mathcal{C}_{1}(\mathcal{H})\). Moreover, if all these functions are in \(L^{1}(\Omega, \mu)\), then since \(\mathcal{B(H)}\) is the dual space of \(\mathcal{C}_{1}(\mathcal{H})\), for any \(E\in \mathcal{M}\), we have the unique operator \(I_{E}\in \mathcal{B(H)}\), called the Gel'fand or weak \(^*\)-integral of \(\mathcal{A}\) over \(E\), such that
\begin{equation} tr(\mathcal{I}_{E} Y)=\int_E tr(A_{t} Y)dt \;\;\;\text{ for all}\;\;\; Y\in \mathcal{C_{1}(H)}. \label{geljfandovintegral} \end{equation}
(5)
We denote it by \(\int_EA_{t}d\mu(t)\) or \(\int_E A d\mu.\) We consider the following important aspect.

Proposition 2. \(A:\Omega \rightarrow \mathcal{B(H)}\) is \([\mu]\) if and only if scalar valued functions \(t \rightarrow \langle A_{t} f,f\rangle\) are \([\mu]\) measurable (resp. integrable) for every \(f\in \mathcal{H}\).

Proof. Every one dimensional operator \(f^{*} \otimes f\) is in \(C_{1}(H)\) and \(tr(A_{t}( f^{*} \otimes f))=tr(f^{*} \otimes A_{t} f)=\left< A_{t} f,f\right>,\) so that \([\mu]\) weak \(^*\)-measurability (resp. \([\mu]\) weak \(^*\)-integrability) of \(A\) directly implies measurability (resp. integrability) of \(\left< A_{t} f,f\right>\) for any \(f\in \mathcal{H}\). The converse follows immediately from [4] and this completes the proof.

We note that in view of Proposition 2, the Equation (5) of Gel'fand integral for \(o.v.\) functions can be reformulated as follows [2]:

Proposition 3. If \(\left< A f,f\right>\in L^1(E,\mu)\) for all \(f\in \mathcal{H}\), for some \(E\in \mathcal{M}\) and a \(\mathcal{B(H)}\)-valued function \(A\) on \(E\), then the mapping \(f\rightarrow\int_E \left< A_{t} f,f\right>d\mu(t)\) represents a quadratic form of bounded operator \(\int_E A dm\) or \(\int_E A_t d\mu(t)\), satisfying the following \( \left< \left(\int_E A_t d\mu (t)\right) f,g\right>= \int_E \left< At f,g\right>\,d\mu (t), for\;\; all\;\; f,g\in \mathcal{H}.\)

Proof. It suffices to show that for all \(E\in \mathcal{M},\) \( \Phi_E(f,g)=\int_E \left< A_{t} f,g\right>\,d\mu (t),\) for all \(f, g\in \mathcal{H}\), defines a bounded sesquilinear functional \(\Phi\) on \(\mathcal{H}\). Indeed, by [1], we have \( | \Phi_E(f,g)| \le \int_E|\left< A_{t} f,g\right>|\,d\mu (t) \le \| A_{t} f,g\|_{L^1} \le M \|f\|\|g\| \) for all \(f,g\in \mathcal{H}\) since integration is a contractive functional on \(L^{1}(\Omega ,\mu)\). This completes the proof.

Remark 2. It is known from [1] that for a \([\mu]\) \(A:\Omega \rightarrow \mathcal{B(H)}\) we have that \(A^*A\) is Gel'fand integrable if and only if \( \int_\Omega \|A_t f\|^2d\mu (t)< \infty,\) for all \(f\in \mathcal{H}\). Moreover, for a \([\mu]\) function \(A:\Omega \rightarrow \mathcal{B(H)}\). Let us consider a linear transformation \(\vec{A}:D_{\vec{A}}\rightarrow L^{2}(\Omega,\mu , \mathcal{H})\), with the domain \(D_{\vec{A}}=\{ f\in \mathcal{H} \,| \, \int_\Omega \|A_t f\|^2 d\mu (t)< \infty\}\), defined by \( ({\vec{A}}f)(t)=A_t f .\) and all \(f\in D_{\vec{A}}.\)

In the next section, we devote our efforts to results on inner product type integral transformers in terms of Landau, Cauchy-Schwarz and Grüss type norm inequalities.

4. Norm inequalities

In this section, we consider various types of norm inequalities for inner product type integral transformers discussed in [1,2,4,7]. From [1], a sufficient condition is provided when \(A^*\) and \(B\) from Definition 2 are both in \(L^2_G(\Omega,d\mu, \mathcal{B(H)}).\) If each of families \((A_t)_{t\in\Omega}\) and \((B_t)_{t\in\Omega}\) consists of commuting normal operators, then by Theorem 3.2 [1], the \(i.p.t.i\) transformer \(\int_\Omega A_t \otimes B_t d\mu(t)\) leaves every \(u.i.\) norm ideal \(\mathcal{C}_{|\|\cdot|\|}(\mathcal{H})\) invariant and the following Cauchy-Schwarz inequality holds:
\begin{equation} \left|\left\| \int_\Omega A_t X B_t d\mu(t) \right|\right\|\le \left|\left\| \sqrt{\int_\Omega A_t^* A_t } d\mu(t) \sqrt{\int_\Omega B_t^*B_t d\mu(t)}\right|\right\|, \end{equation}
(6)
for all \(X\in \mathcal{C}_{|\|\cdot|\|}(\mathcal{H})\). Normality and commutativity condition can be dropped for Schatten \(p\)-norms as shown in Theorem 3.3 [1]. In Theorem 3.1 [2], a formula for the exact norm of the \(i.p.t.i\) transformer \(\int_\Omega A_t \otimes B_td\mu(t)\) acting on \(\mathcal{C}_2(\mathcal{H})\) is found. In Theorem 2.1 [2], the exact norm of the \(i.p.t.i\) transformer \(\int_\Omega A_t^* \otimes A_t d\mu(t)\) is given for two specific cases:
\begin{equation} \left\| \int_\Omega A_t^*\otimes A_t d\mu(t) \right\|_{B(H)\to\mathcal{C}_{\Phi}(\mathcal{H})}= \left\| \int_\Omega A_t^*A_t d\mu(t) \right\|_{\mathcal{C}_\Phi(\mathcal{H})}, \label{bhubiloshta} \end{equation}
(7)
\begin{equation} \left\| \int_\Omega A_t^*\otimes A_t d\mu(t) \right\|_{\mathcal{C}_{\Phi}(\mathcal{H})\to\mathcal{C}_1(\mathcal{H})}= \left\| \int_\Omega A_t A_t^* d\mu(t) \right \|_{\mathcal{C}_{\Phi_*}(\mathcal{H})}, \nonumber%\label{biloshtaunuklearne} \end{equation} where \(\Phi_*\) stands for a \(s.g.\) function related to the dual space \((\mathcal{C}_{\Phi}(\mathcal{H}))^*\). The norm appearing in (7) and its associated space \(L_G^2(\Omega,d\mu,\mathcal{B(H)},\mathcal{C}_\Phi(\mathcal{H}))\) present only a special case of norming a field \(A=(A_t)_{t\in\Omega}\). A much wider class of norms \( \| \cdot\|_{\Phi,\Psi}\) and their associated spaces \(L_G^2(\Omega,d\mu,\mathcal{B(H)},\mathcal{C}_\Phi(\mathcal{H}))\) are given by [2]:
\begin{equation} \| A\|_{\Phi,\Psi}= \left\|\int_\Omega A_t^*\otimes A_t d\mu(t) \right\|_{B(\mathcal{C}_\Phi(\mathcal{H}),\mathcal{C}_\Psi(\mathcal{H}))}^\frac12 \end{equation}
(8)
for an arbitrary pair of \(s.g.\) functions \(\Phi\) and \(\Psi\). For the proof of completeness of the space \(L_G^2(\Omega,d\mu,\mathcal{C}_\Phi(\mathcal{H}), \mathcal{C}_\Psi(\mathcal{H}))\), see Theorem 2.2 [2]. Before going into the details of this section lets consider the following Proposition which will be useful in the sequel [7]. We give its proof for completion.

Proposition 4. Let \(\mu\) be a probability measure on \(\Omega\), then for every field \((\mathcal{A}_t)_{t\in\Omega}\) in \(L^2(\Omega,\mu,\mathcal{B}(\mathcal{H}))\), for all \(B\in\mathcal{B}(\mathcal{H})\), for all unitarily invariant norms \(|\|\cdot|\|\) and for all \(\theta>0\),

\begin{eqnarray} \int_\Omega\left|\mathcal{A}_t-B\right|^2 d\mu(t) &=& \int_\Omega\left|\mathcal{A}_t-\int_\Omega A_t d\mu(t)\right|^2 d\mu(t) +\left| \int_\Omega A_t d\mu(t)-B\right|^2\label{nulto} \end{eqnarray}
(9)
\begin{eqnarray} &\ge& \int_\Omega\left|\mathcal{A}_t-\int_\Omega A_t d\mu(t)\right|^2 d\mu(t) =\int_\Omega|\mathcal{A}_t|^2 d\mu(t)-\left|\int_\Omega A_t d\mu(t)\right|^2, \label{prvo} \end{eqnarray}
(10)
\begin{eqnarray} \min_{B\in\mathcal{B}(\mathcal{H})}\left|\left\|\left|\int_\Omega\left|\mathcal{A}_t-B\right|^2 d\mu(t)|\right|^\theta\right|\right\| &=& \left|\left\| \left|\int_\Omega\left|\mathcal{A}_t-\int_\Omega A_t d\mu(t)\right|^2 d\mu(t)\right|^\theta \right\|\right| \end{eqnarray}
(11)
\begin{eqnarray} &=& \left|\int_\Omega|\mathcal{A}_t|^2 d\mu(t)- \left|\int_\Omega A_t d\mu(t)\right|^2\||^\theta|\right\|. \label{drugo} \end{eqnarray}
(12)
Thus, the considered minimum is always obtained for \(B=\int_\Omega A_t d\mu(t)\).

Proof. The expression (9) is trivial and the inequality (10) follows from (9), while identity (10) is just a a special case of Lemma 2.1 [1] applied for \(k=1\) and \(\delta_1=\Omega\).

As \(0\le A\le B\) for \(A,B\in \mathcal{C_{\infty}(H)}\) implies \( s_n^\theta(A)\le s_n^\theta(B)\) for all \(n\in \mathbb{N}\), as well as \(|\| A^\theta|\|\le |\| B^\theta|\|,\) then (12) follows.

Recall that, for a pair of random real variables \((Y,Z)\), its coefficient of correlation

\[\rho_{Y,Z}=\frac{| E(YZ)-E(Y)E(Z)|}{\sigma(Y)\sigma(Z)}= \frac{| E(YZ)-E(Y)E(Z)|}{ \sqrt{E(Y^2)-E^2(Y)} \sqrt{E(Z^2)-E^2(Z)}}\] always satisfies \(|\rho_{Y,Z}|\le 1.\) For square integrable functions \(f\) and \(g\) on \([0,1]\) and \(D(f,g)=\int_0^1f(t)g(t)\,d t- \int_0^1f(t)\,d t\int_0^1g(t)\,d t.\) Landau proved that \( | D(f,g)|\le \sqrt{D(f,f)D(g,g)}.\)

The following result represents a generalization of Landau inequality in \(u.i.\) norm ideals [2] for Gel'fand integrals of \(o.v.\) functions with relative simplicity of its formulation.

Theorem 3. If \(\mu\) is a probability measure on \(\Omega\). Let both fields \((A_t)_{t\in\Omega}\) and \((B_t)_{t\in\Omega}\) be in \(L^2(\Omega,\mu,\mathcal{B(H)})\) consisting of commuting normal operators and consider \[\sqrt{\,\int_\Omega|A_{t}|^2 -\left|\int_\Omega A_{t} d\mu(t)\right|^2}X \sqrt{\,\int_\Omega| B_{t}|^2 d\mu(t)-\left|\int_\Omega B_{t} d\mu(t)\right|^2},\] for some \(X\in B(H)\). Then \[\int_\Omega A_tX B_t d\mu(t)-\int_\Omega A_{t} dt X\!\!\int_\Omega B_{t} d\mu(t) \in C_{|\|.|\|}(H).\]

Proof. First, we have the following Korkine type identity for \(i.p.t.i\) transformers

\begin{eqnarray} \int_\Omega A_t X\ B_t d\mu(t)\!-\!\int_\Omega A_{t} d\mu(t) X\! \! \int_\Omega B_{t} d\mu(t) &=&\int_\Omega d\mu(s)\int_\Omega A_t X B_t d\mu(t)\!-\!\int_\Omega\!\int_\Omega A_t X B_s\,d\mu(s)d\mu(t)\nonumber\\ &=&\dfrac12\int_{\Omega^2}(A_s- A_t)X( B_s- B_t)d(\mu\times\mu)(s,t).\label{21} \end{eqnarray}
(13)
In this representation, we have \((A_s-A_t)_{(s,t)\in\Omega^2}\) and \((B_s-B_t)_{(s,t)\in\Omega^2}\) to be in \(L^2(\Omega^2,\mu\times\mu, \mathcal{B(H)})\) because by an application of the identity (13),
\begin{eqnarray} \dfrac12\int_{\Omega^2}\left| A_s- A_t\right|^2d(\mu\times\mu)(s,t)=\int_\Omega| A_{t}|^2 d\mu(t)- \left|\int_\Omega A_{t} d\mu(t)\right|^2 =\int_\Omega\left| A_t-\int_\Omega A_{t} d\mu(t)\right|^2 d\mu(t) \in B(H).\label{23} \end{eqnarray}
(14)
Both families \((A_s-A_t)_{(s,t)\in\Omega^2}\) and \((B_s-B_t)_{(s,t)\in\Omega^2}\) consist of commuting normal operators and by Theorem 3.2 [1] \[\dfrac12\int_{\Omega^2}(A_s-A_t)X(B_s-B_t)d(\mu\times\mu)(s,t)\in \mathcal{C}_{|\|\cdot| |\|}(\mathcal{H})\] due to identity (14), and so the conclusion (13) follows.

Lemma 2. Let \(\mu\) (resp. \(\nu\)) be a probability measure on \(\Omega\) (resp. \(\mho\)). Further, let both families \(\{A_s,C_t\}_{(s,t)\in\Omega\times\mho}\) and \(\{B_s, D_t\}_{(s,t)\in\Omega\times\mho}\) consist of commuting normal operators and let \begin{equation}\sqrt{\,\int_\Omega|A_s|^2d\mu(s) \int_\mho|C_t|^2d\nu(t)-\left|\int_\Omega A_sd\mu(s)\int_\mho C_td\nu(t)\right|^2} X \sqrt{\,\int_\Omega| B_s|^2d\mu(s) \int_\mho|D_t|^2d\nu(t)-\left|\int_\Omega B_sd\mu(s)\int_\mho D_td\nu(t)\right|^2} \end{equation} be in \(\mathcal{C}_{|\|\cdot| |\|}(\mathcal{H})\) for some \(X\in \mathcal{B(H)}\). Then \begin{eqnarray*}\int_\Omega \int_\mho A_s C_tX B_s D_t\,d\mu(s)\,d\nu(t) -\int_\Omega A_s \,d\mu(s)\int_\mho C_t\,d\nu(t) X\!\!\int_\Omega B_s \,d\mu(s) \int_\mho D_t\,d\nu(t) \in\mathcal{C}_{|\|\cdot| |\|}(\mathcal{H}).\end{eqnarray*}

Proof. Apply Theorem 3 to the probability measure \(\mu\times\nu\) on \(\Omega\times\mho\) and families \((A_s C_t)_{(s,t)\in\Omega\times\mho}\) and \(( B_s D_t)_{(s,t)\in\Omega\times\mho}\) of normal commuting operators in \(L_G^2(\Omega\times\mho,d\mu\times\nu,\mathcal{B(H)}).\) The rest follows trivially.

Next, we consider Landau type inequality for \(i.p.t.i\) transformers in Schatten ideals for the Schatten \(p\)-norms.

Proposition 5. Let \(\mu\) be a probability measure on \(\Omega\) and \((A_t)_{t\in\Omega}\) and \((B_t)_{t\in \Omega}\) be \(\mu\)-weak\({}^*\) measurable families of bounded Hilbert space operators such that \(\int_\Omega\left(\|A_t f\|^2+\|A_t^* f\|^2+\| B_t f\|^2+\|B_t^* f\|^2\right)d\mu(t)< \infty\) for all \(f\in \mathcal{H}\) and let \(p,q,r\ge1\) such that \(\dfrac1p=\dfrac1{2q}+\dfrac1{2r}\,\). Then for all \(X\in \mathcal{C}_p(\mathcal{H})\),

\begin{eqnarray} \label{grussp}&&\notag \left\|\int_\Omega A_t X B_t d\mu(t)- \int_\Omega A_{t} d\mu(t) X \int_\Omega B_{t} d\mu(t)\right\|_p\\ &&\leqslant\kern-4pt\left\| \left(\int_\Omega\left|\left(\int_\Omega\left|A_t^*-\int_\Omega A_{t}^* d\mu(t) \right|^2 d\mu(t) \right)^{\frac{q-1}2} \left(A_t-\int_\Omega A_{t} d\mu(t)\right)\right|^2 d\mu(t)\right)^{\frac1{2q}}\right\|\nonumber\\&& X\left\|\left(\int_\Omega\left|\left(\int_\Omega\left| B_t-\int_\Omega B_{t}d\mu(t) \right|^2 d\mu(t) \right)^{\frac{r-1}2} \left( B_t^*-\int_\Omega B_{t}^* d\mu(t)\right)\right|^2 d\mu(t)\right)^{\frac1{2r}}\right\|_p. \end{eqnarray}
(15)

Proof. According to identity (14), applying Theorem 3.3 [1] to families \((\mathcal{A}_s-\mathcal{A}_t)_{(s,t)\in\Omega^2}\) and \((\mathcal{B}_s-\mathcal{B}_t)_{(s,t)\in\Omega^2}\) gives

\begin{eqnarray} \label{pnorm} && \left\|\int_\Omega A_t X B_t d\mu(t)-\int_\Omega\mathcal{A}_td\mu(t) X\int_\Omega\mathcal{B}_td\mu(t)\right\|_p =\left\|\dfrac12\int_{\Omega^2}(A_s-A_t)X(B_s-B_t)d(\mu\times\mu)(s,t)\right\|_p \nonumber\\ &&\le\left\|\left(\dfrac12\int_{\Omega^2}(\mathcal{A}_s^*-\mathcal{A}_t^*) \left(\dfrac12\int_{\Omega^2}|\mathcal{A}_s^*-\mathcal{A}_t^*|^2(\mu\times\mu)(s,t)\right)^{q-1}\kern-13.1pt (\mathcal{A}_s-\mathcal{A}_t)d(\mu\times\mu)(s,t)\right)^{\frac1{2q}}\right\|\nonumber\\ &&\times\left\|\left(\dfrac12\int_{\Omega^2}(\mathcal{B}_s-\mathcal{B}_t) \Bigl(\dfrac12\int_{\Omega^2}|\mathcal{B}_s-\mathcal{B}_t|^2(\mu\times\mu)(s,t)\Bigr)^{r-1}\kern-8pt (\mathcal{B}_s^*-\mathcal{B}_t^*)d(\mu\times\mu)(s,t)\right)^{\frac1{2r}}\kern-2pt\right\|_p.\end{eqnarray}
(16)
By applying identity (14) once again, the last expression in (16) becomes \[\bigg\|\biggl(\dfrac12\int_{\Omega^2}(\mathcal{A}_s-\mathcal{A}_t)^* \left(\int_\Omega\left|\mathcal{A}_t^*-\int_\Omega \mathcal{A}^* _t d\mu(t)\right|^2d\mu(t)\right)^{q-1} (\mathcal{A}_s-\mathcal{A}_t)d(\mu\times\mu)(s,t)\biggr)^{\frac1{2q}}\] \[\biggl(\dfrac12\int_{\Omega^2}(\mathcal{B}_s-\mathcal{B}_t) \Bigl(\int_\Omega\left|\mathcal{B}_s-\int_\Omega \mathcal{B} _t d\mu(t)\right|^2d\mu(s)\Bigr)^{r-1}\kern-5pt(\mathcal{B}_s-\mathcal{B}_t)^*d(\mu\times\mu)(s,t)\biggr)^{\frac1{2r}}\bigg\|_p.\label{odvizraz} \] Denoting \(\Bigl(\int_\Omega\left|A_s^*-\int_\Omega\mathcal{A}^*d\mu\right|^2d\mu(s)\Bigr)^{\frac{p-1}2}\) (resp. \(\Bigl(\int_\Omega\left|B_s-\int_\Omega\mathcal{B}d\mu\right|^2d\mu(s)\Bigr)^{\frac{r-1}2}\)) by \(Y\) (resp. \(Z\)), then the expression (16) becomes
\begin{eqnarray} \label{saYZ} \biggl\|\left(\dfrac12\int_{\Omega^2}\left|Y A_s-Y A_t\right|^2d(\mu\times\mu)(s,t)\right)^{\frac1{2q}}. \left(\dfrac12\int_{\Omega^2}\left|Z B_s^*-Z B_t^*\right|^2d(\mu\times\mu)(s,t)\right)^{\frac1{2r}}\biggr\|_p. \end{eqnarray}
(17)
Again applying identity (14) to families \((Y A_t)_{t\in\Omega}\) and \((Z B_t^*)_{t\in\Omega}\), (17) becomes \[\left\|\left(\int_\Omega\left|Y\mathcal{A}_t-\int_\Omega Y\mathcal{A}_t d\mu(t)\right|^2d\mu(t)\right)^{\frac1{2q}}\kern-4pt . \left(\int_\Omega\left|Z\mathcal{B}_t^*-\int_\Omega Z\mathcal{B}_t^* d\mu(t)\right|^2d\mu(t)\right)^{\frac1{2r}}\right\|_p,\] which obviously equals to the righthand side of (15).

The next result [1] is a special case of an abstract Hölder inequality presented in Theorem 3.1.(e) [1] for Cauchy-Schwarz inequality for \(o.v.\) functions in \(u.i.\) norm ideals.

Proposition 6. Let \(\mu\) be a measure on \(\Omega\). Further, let \((A_t)_{t\in\Omega}\) and \((B_t)_{t\in \Omega}\) be \(\mu\)-weak\({}^*\) measurable in \(\mathcal{B(H)}\) such that \(|\int_\Omega|A_t|^2 d\mu(t)|^\theta\) and \(|\int_\Omega|B_t|^2 d\mu(t)|^\theta\) are in \(\mathcal{C}_{\||.|\|}\mathcal{H}\) for some \(\theta>0\) and for \(u.i.\) norm. Then we have \[ \|||\int_\Omega A_t^* B_t d\mu(t)\|||^\theta \|||\le \|||\int_\Omega A_t^* A_t d\mu(t)\|||^\theta \|||^\frac12 \|||\int_\Omega B_t^* B_t d\mu(t)\|||^\theta \|||^\frac12. \]

Proof. Take \(\Phi\) to be a \(s.g.\) function that generates \(u.i.\) norm \(\||\cdot\||\), \(\Phi_1=\Phi\), \(\Phi_2=\Phi_3=\Phi^{(2)}\) (2-reconvexization of \(\Phi\)), \(\alpha=2\theta\) and \(X=I\), and then apply Theorem 3.1 [1], we get our desired result.

Now, we give another generalization of Landau inequality for Gel'fand integrals of \(o.v.\) functions in \(u.i.\) norm ideals.

Theorem 4. If \(\mu\) is a probability measure on \(\Omega\), \(\theta>0\) and \((A_t)_{t\in\Omega}\) and \((B_t)_{t\in \Omega}\) are as in Proposition 6, \(\mu\)-weak\({}^*\) measurable families of bounded Hilbert space operators such that \(\|||\int_\Omega|A_t|^2d\mu(t)\|||^\theta\) and \(\|||\int_\Omega|B_t|^2d\mu(t)\|||^\theta\) are in \(\mathcal{C}_{\||.|\|}\mathcal{H}\) for some \(\theta>0\) and for some \(u.i.\) norm \(\||\cdot\||\) we have \begin{eqnarray*} && \left\|\left|\int_\Omega A_t^* B_td\mu(t) -\int_\Omega A_t^*d\mu(t)\int_\Omega B_td\mu(t) \|||^\theta \right\|\right|^2\\ &&\le \||\int_\Omega \||| A_t \|||^2d\mu(t)- \|||\int_\Omega A_td\mu(t) \|||^2 \|||^\theta \||| \|||\int_\Omega \||| B_t \|||^2d\mu(t)- \|||\int_\Omega B_td\mu(t) \|||^2 \|||^\theta \||.\end{eqnarray*}

Proof. It suffices to invoke Proposition 6 to \(o.v.\) families \((A_s-A_t)_{(s,t)\in\Omega^2}\) and \((B_s-B_t)_{(s,t)\in\Omega^2}\) and use identity [7] to proof this result.

Now, we consider some interesting quantities that relate to norm inequalities. For bounded set of operators \(A=(\mathcal{A}_t)_{t\in\Omega}\), we see that the radius of the smallest disk that essentially contains its range is

\[r_\infty(A)=\inf_{A\in \mathcal{B(H)}}ess \sup_{t\in\Omega}\| A_t-A\|= \inf_{A\in \mathcal{B(H)}}\| A_t-A\|_\infty=\min_{A\in \mathcal{B(H)}}\| A_t-A\|_\infty.\] From the triangle inequality, we have \(\bigl|\|\mathcal{A}_t-A'\|-\|\mathcal{A}_t-A\|\bigr|\leq\|A'-A\|\), so the mapping \(A\to ess \sup_{t\in\Omega}\|A_t-A\|\) is nonnegative and continuous on \(\mathcal{B(H)}\). Since \((\mathcal{A}_t)_{t\in\Omega}\) is bounded field of operators, we also have \(\| A_{t}-A\|\to\infty\) when \(\|A\|\to\infty\), so this mapping attains minimum [5], and it actually attains at some \(A_0\in \mathcal{B(H)}\), which represents a center of the disk considered [6]. Any such field of operators is of finite diameter, therefore, we have that \(r_\infty(A)=ess \sup_{s,t\in\Omega}\| A_s-A_t\|,\) with the simple inequalities given as \(r_\infty(A)\le diam_\infty(A)\le 2r_\infty(A)\) relating those quantities. For such fields of operators we can now state the following stronger version of Grüss inequality [2].

Lemma 3. Let \(\mu\) be a \(\sigma\)-finite measure on \(\Omega\) and let \(A=(\mathcal{A}_t)_{t\in\Omega}\) and \(B=(\mathcal{B}_t)_{t\in\Omega}\) be \([\mu]\) a.e. bounded fields of operators. Then, for all \(X\in \mathcal{C_{|\|.|\|}(H)}\), \( \sup_{\mu(\delta)>0}\||\frac1{\mu(\delta)}\int_\delta\mathcal{A}_tX\mathcal{B}_t d\mu(t) - \frac1{\mu(\delta)}\int_\delta\mathcal{A}_t d\mu(t) \,X \frac1{\mu(\delta)}\int_\delta\mathcal{B}_t d\mu(t) |\|\le \min_{i} \mathcal{P}_{i}\cdot\|| X\||.\) (Here \(\sup\) is taken over all measurable sets \(\delta\subseteq\Omega\) such that \(0< \mu(\delta)< \infty\)).

Lemma 3 has the following immediate implication when \((\mathcal{A}_t)_{t\in\Omega}\) and \((\mathcal{B}_t)_{t\in\Omega}\) are bounded fields of self-adjoint operators.

Theorem 5. If \(\mu\) is a probability measure on \(\Omega\) and \(C,D,E,F\) be bounded self-adjoint operators. Also, let \((\mathcal{A}_t)_{t\in\Omega}\) and \((\mathcal{B}_t)_{t\in\Omega}\) be bounded self-adjoint fields satisfying \(C\le\mathcal{A}_t\le D\) and \(E\le\mathcal{B}_t\le F\) for all \(t\in\Omega\). Then for all \(X\in \mathcal{C_{|\|.|\|}(H)}\), we have

\begin{equation} \left\|\left|\int_\Omega\mathcal{A}_tX\mathcal{B}_t d\mu(t)- \int_\Omega\mathcal{A}_td\mu(t) \,X \int_\Omega\mathcal{B}_t d\mu(t)\right|\right\| \le\dfrac{\|D-C\|\cdot\|F-E\|}4\cdot \|| X|\|. \label{oshtrina} \end{equation}
(18)

Proof. As \(\frac{C-D}2\le\mathcal{A}_t-\frac{C+D}2\le\frac{D-C}2\) for every \(t\in\Omega\), then \begin{eqnarray*} ess \sup_{t\in\Omega}\| \mathcal{A}_t-\frac{C+D}2\|= ess \sup_{t\in\Omega}\sup_{\| f\|=1}\||\langle\mathcal{A}_t-\frac{C+D}2 \| f,f\rangle|\le \sup_{\| f\|=1}|\langle\frac{D-C}2 f,f\rangle|= \frac{\| D-C\|}2, \end{eqnarray*} which implies \(r_\infty(A)\le \frac{\| D-C\|}2,\) and similarly \(r_\infty(B)\le\frac{\| F-E\|}2.\) Thus (18) follows directly.

In case of \(\mathcal{H}=\mathbb{C}\) and \(\mu\) being the normalized Lebesgue measure on \([a,b]\) (i.e. \(d\,\mu(t)=\frac{dt}{b-a}\)), then (1) follows from Theorem 5. This special case also confirms the sharpness of the constant \(\frac14\) in the inequality (18).

Lastly, we consider, the Grüss type inequality for elementary operators in the example below.

Example 1. Let \(A_1, \ldots, A_n\), \(B_1, \ldots, B_n\), \(C, D, E\) and \(F\) be bounded linear self-adjoint operators acting on a Hilbert space \(\mathcal{H}\) such that \(C\le A_i\le D\) and \(E\le B_i\le F\) for all \(i=1,2,\cdots,n\), then for arbitrary \(X\in\mathcal{C}_{\||.|\|}\mathcal{H}\), we have \begin{eqnarray} \nonumber%\label{mainc} \|| \frac1n\sum_{i=1}^n A_i XB_i-\frac1{n^2}\sum_{i=1}^nA_i\, X\sum_{i=1}^nB_i\|| \leq \frac{\|D-C\|\|F-E\|}4 \|| X\||. \end{eqnarray} Indeed, it is sufficient to prove that the elementary operator is normally represented and that Grüss type inequality holds for it [3].

In the next section, we dedicate our effort to the applications of this study in other fields. We consider quantum theory in particular, whereby, we describe the application in quantum chemistry and quantum mechanics.

5. Applications in quantum theory

Norm inequalities and other properties of \(i.p.t.i\) transformers have various applications in other fields. We discuss the applications in quantum theory involving two cases [3]. The first case is in quantum chemistry, whereby, we consider the Hamiltonian which is a bounded, self-adjoint operator on some infinite-dimensional Hilbert space which governs a quantum chemical system. The Hamiltonian helps in estimation of ground state energies of chemical systems via subsystems.

The quantum mechanics deals with commutator approximation. The discussions of approximation by commutators \(AX-XA\) or by generalized commutator \(AX-XB\) originates from quantum theory. For instance, the Heisenberg uncertainly principle may be mathematically deduced as saying that there exists a pair \(A,X\) of linear operators and a non-zero scalar \(\alpha\) for which \(AX - XA = \alpha I\). A natural question immediately arises: How close can \(AX - XA\) be to the identity? In [3], it is discussed that if \(A\) is normal, then, for all \(X \in B(H)\), \(||I - (AX - XA)|| \geq ||I||.\) In the inequality here, the zero commutator is a commutator approximate in \(B(H)\).

Acknowledgments

The author is grateful to the referees for the useful comments

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Jocic, D. R. (2005). Cauchy-Schwarz norm inequalities for weak*-integrals of operator valued functions. Journal of Function Analysis, 218, 318-346.[Google Scholor]
  2. Jocic, D. R. (2009). Interpolation norms between row and column spaces and the norm problem for elementary operators. Linear Algebra Applications, 430, 2961-2974. [Google Scholor]
  3. Okelo, N. B, Agure, J. O., & Oleche, P. O. (2013). Various notions of orthogonality in normed spaces. Acta Mathematica Scientia, 33, 1387-1397. [Google Scholor]
  4. Jocic, D. R. (1999). The Cauchy-Schwarz norm inequality for elementary operators in Schatten ideals. Journal of London Mathematical Society, 60, 925-934.[Google Scholor]
  5. Conway, J. B. (1990). A course in Functional Analysis, Second edition. Springer-Verlag, New York. [Google Scholor]
  6. Kreyzig, E. (1978). Introductory Functional Analysis with Applications. John Wiley and sons, New York.[Google Scholor]
  7. Li, X., Mohapatra, R. N., & Rodriguez, R. S. (2002). Grüss-type inequalities. Journal of Mathematical Analysis Applications, 267, 434-443. [Google Scholor]
]]>
Boundary value problems for a class of stochastic nonlinear fractional order differential equations https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/boundary-value-problems-for-a-class-of-stochastic-nonlinear-fractional-order-differential-equations/ Mon, 21 Dec 2020 13:54:39 +0000 https://old.pisrt.org/?p=4812
OMA-Vol. 4 (2020), Issue 2, pp. 152 - 159 Open Access Full-Text PDF
McSylvester Ejighikeme Omaba, Louis O. Omenyi
Abstract: Consider a class of two-point Boundary Value Problems (BVP) for a stochastic nonlinear fractional order differential equation \(D^\alpha u(t)=\lambda\sqrt{I^\beta[\sigma^2(t,u(t))]}\dot{w}(t),\,\,00\) is a level of the noise term, \(\sigma:[0,1]\times\mathbb{R}\rightarrow\mathbb{R}\) is continuous, \(\dot{w}(t)\) is a generalized derivative of Wiener process (Gaussian white noise), \(D^\alpha\) is the Riemann-Liouville fractional differential operator of order \(\alpha\in (3,4)\) and \(I^\beta,\,\,\beta>0\) is a fractional integral operator. We formulate the solution of the equation via a stochastic Volterra-type equation and investigate its existence and uniqueness under some precise linearity conditions using contraction fixed point theorem. A case of the above BVP for a stochastic nonlinear second order differential equation for \(\alpha=2\) and \(\beta=0\) with \(u(0)=u(1)=0\) is also studied.
]]>

Open Journal of Mathematical Analysis

Boundary value problems for a class of stochastic nonlinear fractional order differential equations

McSylvester Ejighikeme Omaba\(^1\), Louis O. Omenyi
Department of Mathematics, College of Science, University of Hafr Al Batin, P. O Box 1803 Hafr Al Batin 31991, KSA.; (M.E.O)
Department of Mathematics/Computer Science/Statistics/Informatics, Alex Ekwueme Federal University, Ndufu-Alike, Ikwo, Nigeria.; (L.O.O)
\(^{1}\)Corresponding Author: mcomaba@uhb.edu.sa

Abstract

Consider a class of two-point Boundary Value Problems (BVP) for a stochastic nonlinear fractional order differential equation \(D^\alpha u(t)=\lambda\sqrt{I^\beta[\sigma^2(t,u(t))]}\dot{w}(t),\,\,0<t<1\) with boundary conditions \(u(0)=0,\,\,u'(0)=u'(1)=0,\) where \(\lambda>0\) is a level of the noise term, \(\sigma:[0,1]\times\mathbb{R}\rightarrow\mathbb{R}\) is continuous, \(\dot{w}(t)\) is a generalized derivative of Wiener process (Gaussian white noise), \(D^\alpha\) is the Riemann-Liouville fractional differential operator of order \(\alpha\in (3,4)\) and \(I^\beta,\,\,\beta>0\) is a fractional integral operator. We formulate the solution of the equation via a stochastic Volterra-type equation and investigate its existence and uniqueness under some precise linearity conditions using contraction fixed point theorem. A case of the above BVP for a stochastic nonlinear second order differential equation for \(\alpha=2\) and \(\beta=0\) with \(u(0)=u(1)=0\) is also studied.

Keywords:

Boundary value problems, existence and uniqueness result, fractional integral, Riemann-Liouville fractional derivative, stochastic volterra-type equation.

1. Introduction

Fractional derivatives and fractional order differential equations have received an increasingly high interest and attention due to its physical and modeling applications in science, engineering and mathematics[1,2,3,4]. On the other hand, researches have shown that two-point boundary value problems have found their applications in theoretical physics, applied mathematics, optimization and control theory and engineering [5]. The use of second-order Boundary Value Problem (BVP) arises in several areas of engineering and applied sciences such as celestial mechanics, circuit theory, astrophysics, chemical kinetics, and biology [6].

They are particularly encountered in the study of following natural phenomena: in the study of surface-tension-induced flows of a liquid metal or semiconductor [7,8], used in modeling biological materials (elastic and hyperelastic materials) [9]. In [10], authors studied the existence and uniqueness of a nontrivial solution of a two-point boundary value problems. Stochastic nonlinear fractional order differential equation and its associated BVPs, on the other hand, will undoubtedly give more realistic models for the above natural occurrences.

Motivated by the above applications and the results of the papers [2,3,4], we study the white noise pertubation of a nonlinear BVP and consider the following BVP for a stochastic nonlinear fractional differential equation

\begin{eqnarray} \label{eqn-fract} \left \{ \begin{array}{lll} D^\alpha u(t)=\lambda\sqrt{I^\beta[\sigma^2(t,u(t))]}\dot{w}(t),\,\,0< t< 1\\ u(0)=0,\,\,u'(0)=u'(1)=0, \end{array}\right. \end{eqnarray}
(1)
where \(\lambda>0\) is a noise level parameter, \(\sigma:[0,1]\times\mathbb{R}\rightarrow\mathbb{R}\) is continuous, \(\dot{w}(t)\) is a Gaussian white noise, \(D^\alpha=D^\alpha_{0_+}\) is the Riemann-Liouville (R-L) fractional derivative of order \(\alpha\in (3, 4)\) and \(I^\beta\) is the fractional integral of order \(\beta>0\). Existence and uniqueness results are very crucial in the study of BVP since they are not well behaved like the initial value problems, therefore, we aim to establish the existence and uniqueness of solution to Equation (1). To the best of our knowledge, the above model has not been studied before and we therefore, try to make sense of the solution to the above equation as follows:

Definition 1. We say that \(\{u(t)\}_{0\leq t\leq 1}\) is a mild solution to Equation (1) if \(a. s\), the following is satisfied \begin{eqnarray*} u(t)=\lambda\int_0^1 G(t,s)\sqrt{I^\beta[\sigma^2(s,u(s))]}\dot{w}(s)ds=\lambda\int_0^1 G(t,s)\sqrt{I^\beta[\sigma^2(s,u(s))]}dw(s), \end{eqnarray*} where \(G(t,s)\) is as defined in (3).

If \(\{u(t)\}_{0< t< 1}\) satisfies the additional condition \(\sup_{0\leq t\leq 1}{\mathbb E}|u(t)|^2< \infty,\) then we say that \(\{u(t)\}_{0\leq t\leq 1}\) is a random field solution to Equation (1).

Definition 2. The R-L fractional integral of order \(\beta>0\) of a given continuous function \(f:(0,\infty)\rightarrow\mathbb{R}\) is defined by \[I^\beta f(t)=\frac{1}{\Gamma(\beta)}\int_0^t(t-s)^{\beta-1}f(s)ds,\] provided the right integral exists. Denote \(I^0 f(t)=f(t).\)

Definition 3. The R-L fractional derivative of order \(\alpha>0\) of a given continuous function \(f:(0,\infty)\rightarrow\mathbb{R}\) is defined by \[D^\alpha f(t)=\frac{1}{\Gamma(n-\alpha)}\bigg(\frac{d}{dt}\bigg)^n\int_0^t\frac{f(s)}{(t-s)^{\alpha-n+1}}ds,\] where \(n=[\alpha]+1\), provided the right integral converges.

Many authors studied different boundary value problems of nonlinear fractional order differential equations, see [1,11,12,13,14] and their references. Xu in [3] considered the following boundary value problem of a nonlinear fractional differential equation:

Lemma 1 ([3]). Given \(h\in C[0,1]\) and \(3< \alpha\leq 4\), the unique solution of \begin{equation*} \begin{cases} D^\alpha u(t)=h(t),\,\,0< t< 1\\ u(0)=u(1)=u'(0)=u'(1)=0, \end{cases}\end{equation*} is \(u(t)=\displaystyle\int_0^1 G(t,s)h(s)ds\) where

\begin{eqnarray} \label{Green-1} G(t,s)=\left \{ \begin{array}{lll} \frac{(t-s)^{\alpha-1}+(1-s)^{\alpha-2}t^{\alpha-2}[(s-t)+(\alpha-2)(1-t)s]}{\Gamma(\alpha)},& 0\leq s\leq t\leq1\\ \frac{t^{\alpha-2}(1-s)^{\alpha-2}[(s-t)+(\alpha-2)(1-t)s]}{\Gamma(\alpha)},&0\leq t\leq s\leq1. \end{array}\right. \end{eqnarray}
(2)

Lemma 2 ([4]). The Green function \(G(t,s)\) in (2) has the following properties

  • (1) \(G(t,s)=G(1-s,1-t)\) for \(t,s\in[0,1];\)
  • (2) \(t^{\alpha-2}(1-t)^2q(s)\leq G(t,s)\leq (\alpha-1)q(s)\) and \(G(t,s)\leq \frac{(\alpha-1)(\alpha-2)}{\Gamma(\alpha)}t^{\alpha-2}(1-t)^2\) for \(t,s\in[0,1]\) where \(q(s)=\frac{(\alpha-2)}{\Gamma(\alpha)}s^2(1-s)^{\alpha-2}.\)
Xu in [3] gave more properties of the function \(G(t,s)\) in (2) as follows:

Lemma 3([3]). The Green function \(G(t,s)\) defined by (2) satisfies the following conditions:

  • (1) \(G(t,s)=G(1-s,1-t)\) for \(t,s\in(0,1)\);
  • (2) \((\alpha-2)t^{\alpha-2}(1-t)^2s^2\leq\Gamma(\alpha)G(t,s)\leq M_0 s^2(1-s)^{\alpha-2}\), for \(t,s\in(0,1)\);
  • (3) \(G(t,s)>0\), for \(t,s\in(0,1);\)
  • (4) \((\alpha-2)s^2(1-s)^{\alpha-2}t^{\alpha-2}(1-t)^2\leq\Gamma(\alpha)G(t,s)\leq M_0 t^{\alpha-2}(1-t)^2\), for \(t,s\in(0,1)\), where \(M_0=\max\{\alpha-1,(\alpha-2)^2\}\).
Stanek in [2] studied the following fractional boundary value problems:

Lemma 4([2]). Suppose that \(\rho\in L^1[0,1]\). Then \(u(t)=\int_0^1 G(t,s)\rho(s)ds\) for \(t\in [0,1]\) is the unique solution of the following equation in \(C^1[0,1]\): \begin{equation*} \begin{cases} D^\alpha u(t)+\rho(t)=0,&\alpha\in(2,3)\\ u(0)=0,\,\, u'(0)=u'(1)=0,& \end{cases} \end{equation*} where

\begin{eqnarray} \label{Green2} G(t,s)=\left \{ \begin{array}{lll} \frac{t^{\alpha-1}(1-s)^{\alpha-2}-(t-s)^{\alpha-1}}{\Gamma(\alpha)},&0\leq s\leq t\leq1\\ \frac{t^{\alpha-1}(1-s)^{\alpha-2}}{\Gamma(\alpha)},&0\leq t\leq s\leq1. \end{array}\right. \end{eqnarray}
(3)

Lemma 5([2]). The function \(G(t,s)\) given in (3) has the following properties:

  • (1) \(G\in C([0,1]\times [0,1])\) and \(G^\alpha>0\) on \((0,1)\times (0,1)\);
  • (2) \(G(t,s)\leq\frac{1}{\Gamma(\alpha)}\) for \((t,s)\in [0,1]\times [0,1]\);
  • (3) \(\displaystyle\int_0^1 G(t,s)ds\geq\frac{t^{\alpha-1}}{(\alpha-1)\Gamma(\alpha+1)}\) for \(t\in[0,1]\);
  • (4) \(\frac{\partial}{\partial t}G(t,s)\in C([0,1]\times [0,1])\) and \(\frac{\partial}{\partial t}G(t,s)>0\) on \((0,1)\times (0,1)\);
  • (5) \(\frac{\partial}{\partial t}G(t,s)\leq\frac{1}{\Gamma(\alpha-1)}\) for \((t,s)\in [0,1]\times [0,1]\);
  • (6) \(\displaystyle\int_0^1 \frac{\partial}{\partial t}G(t,s)ds\geq\frac{t(1-t)}{\Gamma(\alpha)}\) for \(t\in[0,1]\).
The paper is outlined as follows. A short preliminary on second order boundary problems is given in Section 2. In Section 3, we gave the main results and their proofs while Section 4 contains a concise summary of the paper.

2. Preliminaries

Let \((a,b)\in\mathbb{R}\) be an interval and \(p,q,f:(a,b)\rightarrow\mathbb{R}\) be continuous functions. Now, consider the linear second order equation given by \(y''+p(x)y'+q(x)=f(x)\) subject to the following boundary conditions:
  • (a.) Dirichlet or first kind: \(y(a)=\eta_1,\,\,y(b)=\eta_2\),
  • (b.) Neuman or second kind: \(y'(a)=\eta_1,\,\,y'(b)=\eta_2\),
  • (c.) Robin or third or mixed kind: \(\alpha_1 y(a)+\alpha_2 y'(a)=\eta_1,\,\,\alpha_1 y(b)+\alpha_2 y'(b)=\eta_2\),
  • (d.) Periodic: \(y(a)=y(b),\,\,y'(a)=y'(b)\).

Remark 1.

  • 1. Unlike the initial value problems, the BVPs do not behave nicely because there are BVPs for which their solutions fail to exist and even when a solution does exist, there might be many of them. This of course makes existence and uniqueness for BVPs to fail generally.
  • 2. For example, \(y''+y=0\) has a unique solution \(y(x)=\sin x+\cos x\) with the boundary conditions \(y(0)=y(\frac{\pi}{2})=1\). It has no solution with the boundary conditions \(y(0)=y(\pi)=1\) and has infinitely many solutions with the boundary conditions \(y(0)=y(2\pi)=1\).
Thus, for the linear boundary value problem with the Dirichlet boundary conditions:
\begin{equation}\label{eqn-linear} \begin{cases}L[u(x)]=f(x,u),\\ u(a)=u(b)=0, \end{cases} \end{equation}
(4)
where \(L\) is a second-order differential operator given by \(L=\frac{d}{dx}\bigg[p(x)\frac{d}{dx}\bigg]+q(x).\) The Equation (4) has the following integro-differential equation as its solution: \[u(x)=\int_a^b G(x,\xi)f(\xi,u(\xi))d\xi,\] with \(G(x,\xi)\) the Green's function.

Lemma 6([10]). Let \(y(t)\in X\), then the Boundary Value Problem \begin{eqnarray*} \left \{ \begin{array}{lll} u''(t)-y(t)=0,& 0< t< 1\\ u(0)=u(1)=0,& \end{array}\right. \end{eqnarray*} has a unique solution \(\displaystyle\int_0^1G(t,s)y(s)ds,\) where

\begin{eqnarray} \label{Green3} G(t,s)=\left \{ \begin{array}{lll} t(1-s),&0\leq t\leq s\leq 1,\\ s(1-t),&0\leq s\leq t\leq 1, \end{array}\right. \end{eqnarray}
(5)
is the Green's function of BVP \begin{eqnarray*} \left \{ \begin{array}{lll} u''(t)=0,& 0< t< 1\\ u(0)=u(1)=0.& \end{array}\right. \end{eqnarray*}

Remark 2. The Green function has the following bound: \(\displaystyle\sup_{0\leq t,s\leq 1}G(t,s)\leq\frac{1}{2}.\)

3. Main results

Now, consider the following global Lipschitz continuity condition on \(\sigma\):

Condition 1. Let there exist a nonnegative function \(p\in L^2[0,1],\) such that \(|\sigma(t,x)-\sigma(t,y)|\leq p(t)|x-y|,\,\,\forall\,t\in [0,1],\,x,y\in\mathbb{R},\) with \(\sigma(t,0)=0\) for convenience, and there exists \(t_0\in [0,1]\) such that \(p(t_0)\neq 0.\)

In particular, for \(p(t)=Lip_{\sigma}\) and for all \(t\in[0,1]\), we have the following:

Condition 2. There exist a finite positive constant \(Lip_{\sigma},\) such that for all \(x,\,y\in\mathbb{R}\), we have \( |\sigma(t,x)-\sigma(t,y)|\leq Lip_{\sigma}|x-y|, \) with \(\sigma(t,0)=0\) for convenience.

Now, define \(L^2({\mathbb P})\) norm of the solution \(u\) by \[\|u\|_2:=\bigg\{\sup_{0\leq t\leq 1}{\mathbb E}|u(t)|^2\bigg\}^{1/2}.\]

3.1. Proof of results for Equation (1)

Using Condition 2, we obtain the following results for Equation (1):

Theorem 1. Suppose \(\lambda< \frac{\sqrt{\beta(\beta+1)\Gamma(\beta)\Gamma^2(\alpha)}}{Lip_{\sigma}}\), where \(\alpha\in(3,4)\) and \(\beta>0\). For a positive constant \(Lip_{\sigma}\) together with Condition 2, there exists a solution \(u\) for Equation (1) that is unique up to modification.

To proof the Theorem 2, let \(u(t)=\mathcal{A} u(t)\), where the operator \(\mathcal{A}\) is given by

\[\mathcal{A} u(t)=\lambda\int_0^1 G(t,s)\sqrt{I^\beta[\sigma^2(s,u(s))]}dw(s),\] and we will use the fixed point of \(\mathcal{A}\). The proof follows using the Lemma(s) below:

Lemma 7. Given a random solution \(u\) such that \(\|u\|_2< \infty\) and Condition 2 holds. Then \[\|\mathcal{A} u\|^2_2\leq c_{\alpha,\beta}\lambda^2Lip_{\sigma}^2\|u\|_2^2,\] where \(c_{\alpha,\beta}=\frac{1}{\beta(\beta+1)\Gamma(\beta)\Gamma^2(\alpha)}.\)

Proof. Take second moment of both sides and use Itó isometry together with Lemma 5(2) to obtain \begin{eqnarray*} \mathbb{E}|\mathcal{A} u(t)|^2&=& \lambda^2\int_0^1G^2(t,s){\mathbb E}|\sqrt{I^\beta[\sigma^2(s,u(s))]}|^2ds\\ &\leq&\lambda^2\int_0^1G^2(t,s)\bigg[\frac{1}{\Gamma(\beta)}\int_0^s(s-r)^{\beta-1}{\mathbb E}|\sigma^2(r,u(r))|dr\bigg]ds\\ &\leq&\lambda^2Lip_{\sigma}^2\int_0^1G^2(t,s)\bigg[\frac{1}{\Gamma(\beta)}\int_0^s(s-r)^{\beta-1}{\mathbb E}|u(r)|^2dr\bigg]ds\\ &\leq&\lambda^2Lip_{\sigma}^2\sup_{0\leq s\leq 1}G^2(t,s)\|u\|_2^2\int_0^1\bigg[\frac{1}{\Gamma(\beta)}\int_0^s(s-r)^{\beta-1}dr\bigg]ds\\ &\leq&\lambda^2Lip_{\sigma}^2\bigg[\sup_{0\leq s\leq 1}G(t,s)\bigg]^2\|u\|_2^2\int_0^1\bigg[\frac{1}{\Gamma(\beta)}\int_0^s(s-r)^{\beta-1}dr\bigg]ds\\ &\leq&\frac{\lambda^2Lip_{\sigma}^2}{\Gamma^2(\alpha)\Gamma(\beta)}\|u\|_2^2\int_0^1\bigg[\int_0^s(s-r)^{\beta-1}dr\bigg]ds=\frac{\lambda^2Lip_{\sigma}^2}{\beta(\beta+1)\Gamma(\beta)\Gamma^2(\alpha)}\|u\|_2^2. \end{eqnarray*} Now, by taking supremum over \(t\in [0,1]\), the result follows.

Remark 3. The operator \(\mathcal{A}\) is a contraction for \(\frac{\lambda^2Lip_{\sigma}^2}{\beta(\beta+1)\Gamma(\beta)\Gamma^2(\alpha)}< 1\).

Lemma 8. Suppose \(u\) and \(v\) are random solutions such that \(\| u\|_{2}+\|v\|_{2}< \infty\) and Condition 2 holds. Then \[\|\mathcal{A} u-\mathcal{A} v\|^2_2\leq c_{\alpha,\beta}\lambda^2Lip_{\sigma}^2\|u-v\|_2^2.\]

Proof. The proof follows similarly to the proof of Lemma 7.

Now, we present the proof of Theorem 1.

Proof of Theorem 1. From Lemma 7, we have \[\|u\|^2_2=\|\mathcal{A} u\|^2_2\leq c_{\alpha,\beta}\lambda^2Lip_{\sigma}^2\|u\|_2^2,\] which follows that \[\|u\|^2_2\bigg[1-c_{\alpha,\beta}\lambda^2Lip_{\sigma}^2\bigg]\leq 0,\] and this shows that \(\|u\|_2< \infty\) whenever \(c_{\alpha,\beta}\lambda^2Lip_{\sigma}^2< 1\) or \( \lambda< \frac{1}{\sqrt{c_{\alpha,\beta}}Lip_{\sigma}}\).

Similarly from Lemma 8, we have

\[\|u-v\|^2_2=\|\mathcal{A} u-\mathcal{A} v\|^2_2\leq c_{\alpha,\beta}\lambda^2Lip_{\sigma}^2\|u-v\|_2^2,\] and therefore \[\|u-v\|^2_2\bigg[1-c_{\alpha,\beta}\lambda^2Lip_{\sigma}^2\bigg]\leq 0.\] This implies that \(\|u-v\|_2< 0\) if and only if \(c_{\alpha,\beta}\lambda^2Lip_{\sigma}^2< 1\) and thus \(\|u-v\|_2=0\Rightarrow u=v\). This shows the existence and uniqueness result by Banach's contraction principle.

3.2. Existence and uniqueness for stochastic second order BVP

Here, we consider the following stochastic second order boundary value problem:
\begin{eqnarray} \label{eqn1} \left \{ \begin{array}{lll} \frac{d^2}{dt^2}u(t)=\lambda \sigma(t,u(t))\dot{w}(t),& 0< t< 1\\ u(0)=u(1)=0,& \end{array}\right. \end{eqnarray}
(6)
where \(\lambda>0\) is a noise level parameter, \(\sigma:[0,1]\times\mathbb{R}\rightarrow\mathbb{R}\) is continuous and \(\dot{w}(t)\) is a Gaussian white noise.

Definition 4. We say that \(\{u(t)\}_{0\leq t\leq 1}\) is a mild solution of Equation (6) if \(a. s\), \begin{equation*} u(t)=\lambda\int_0^1 G(t,s) \sigma(s,u(s))\dot{w}(s)ds=\lambda\int_0^1 G(t,s) \sigma(s,u(s))dw(s), \end{equation*} is satisfied, where \(G(t,s)\) is as given in Equation (5).

If \(\{u(t)\}_{0< t< 1}\) satisfies the additional condition \(\displaystyle\sup_{0\leq t\leq 1}{\mathbb E}|u(t)|^2< \infty,\) then we say that \(\{u(t)\}_{0\leq t\leq 1}\) is a random field solution to Equation (6).

Theorem 2. Suppose Condition 1 holds and there exists a positive constant \(\lambda^*\) such that for any \(0< \lambda\leq\lambda^*\), then Equation (6) has a unique solution.

To proof the Theorem 2, we define the operator

\[\mathcal{B} u(x,t)=\lambda\int_0^1 G(t,s) \sigma(s,u(s))dw(s),\] and use the fixed point of the operator \(\mathcal{B}\). The proof follows using the Lemma(s) below:

Lemma 9. Given a random solution \(u\) such that \(\|u\|_2< \infty\) and Condition 1 holds. Then there exists a positive constant \(\lambda^*\) such that for \(0< \lambda\leq\lambda^*\), \[\|\mathcal{B} u\|_2\leq \frac{1}{2}\|u\|_2.\]

Proof. By Itó Isometry, we obtain \begin{eqnarray*} {\mathbb E}|\mathcal{B} u(t)|^2&\leq&\lambda^2\int_0^1 G^2(t,s) {\mathbb E}|\sigma(s,u(s))|^2ds\\ &\leq&\lambda^2\sup_{0\leq t,s\leq 1}G^2(t,s)\int_0^1 p^2(s) {\mathbb E}|u(s)|^2ds\\ &\leq&\lambda^2\bigg[\sup_{0\leq t,s\leq 1}G(t,s)\bigg]^2\int_0^1 p^2(s){\mathbb E}|u(s)|^2ds\\ &\leq&\frac{\lambda^2}{4}\|u\|_2^2\int_0^1 p^2(s) ds. \end{eqnarray*} Taking suprimum of both sides over \(t\in[0,1]\) and letting \(\lambda^*:=\displaystyle\bigg(\int_0^1 p^2(s) ds\bigg)^{-2}\) such that \(0< \lambda\leq\lambda^*\), we have \(\|\mathcal{B} u\|^2_2\leq \frac{1}{4}\|u\|_2^2.\)

Following similar steps of Lemma 9, we obtain the following result.

Lemma 10. Suppose \(u\) and \(v\) are random solutions such that \(\| u\|_{2}+\|v\|_{2}< \infty\) and Condition 1 holds. Then there exists a positive constant \(\lambda^*\) such that for \(0< \lambda\leq\lambda^*\), \[\|\mathcal{B} u-\mathcal{B} v\|_2\leq \frac{1}{2}\|u-v\|_2.\]

Proof of Theorem 2. Following Lemma 9 and Lemma 12, it is clear that \(\mathcal{B}\) is a contraction. Thus by Banach fixed point theorem, the existence of a unique solution for Equation (6) follows: Let \(u=\mathcal{B}\) and \[\|u\|^2_2=\|\mathcal{B} u\|^2_2\leq \frac{1}{4}\|u\|_2^2\Rightarrow \|u\|^2_2\left[1-\frac{1}{4}\right]\leq 0\Rightarrow \|u\|_2=0\Rightarrow u=0.\] Assume a nontrivial solution \(u\) of Equation (6) and we show that it is unique. Suppose for contradiction that there exists another solution \(v\) of Equation (6) such that \[\|u-v\|^2_2=\|\mathcal{B} u-\mathcal{B} v\|^2_2\leq \frac{1}{4}\|u-v\|_2^2,\] then \(\|u-v\|^2_2[1-\frac{1}{4}]\leq 0\), which follows that \(\|u-v\|=0\). Thus \(u=v\), a unique solution.

Next, we seek to establish the existence and uniqueness of solution for Equation (6) using Condition 2 as follows:

Theorem 3. Suppose \(\lambda< \frac{2}{Lip_{\sigma}}\), for positive constant \(Lip_{\sigma}\) together with Condition 2. Then there exists solution \(u\) that is unique up to modification.

Lemma 11. Given a random solution \(u\) such that \(\|u\|_2< \infty\) and Condition 2 holds. Then \(\|\mathcal{B} u\|_2\leq\frac{\lambda Lip_{\sigma}}{2}\|u\|_2.\)

Lemma 12. Suppose \(u\) and \(v\) are random solutions such that \(\| u\|_{2}+\|v\|_{2}< \infty\) and Condition 2 holds. Then \(\|\mathcal{B} u-\mathcal{B} v\|_2\leq\frac{\lambda Lip_{\sigma}}{2}\|u-v\|_2.\)

Proof of Theorem 3. By fixed point theorem we have \(u(t)=\mathcal{A} u(t)\) and \( \|u\|^2_{2}=\|\mathcal{B} u\|^2_{2}\leq \frac{\lambda^2Lip_{\sigma}^2}{4}\|u\|^2_{2}, \) which follows that \(\|u\|^2_{2}\left[1-\frac{\lambda^2Lip_{\sigma}^2}{4}\right]\leq 0\Rightarrow \|u\|_{2}< \infty \Leftrightarrow \lambda< \frac{2}{Lip_{\sigma}}.\)

Similarly, \(\|u-v\|^2_{2}=\|\mathcal{B} u-\mathcal{B} v\|^2_{2}\leq \frac{\lambda^2Lip_{\sigma}^2}{4}\|u-v\|^2_{2},\) thus \(\|u-v\|^2_{2}\left[1-\frac{\lambda^2Lip_{\sigma}^2}{4}\right]\leq 0\) and therefore \(\|u-v\|_{2}< 0\) if and only if \(\lambda< \frac{2}{Lip_{\sigma}}.\) Hence, the existence and uniqueness result follows by Banach's contraction principle.

4. Conclusion

We studied the boundary value problems for both stochastic nonlinear fractional order differential equation and stochastic nonlinear second order equation. The existence and uniqueness result for both boundary value problems were given under different linearity conditions on \(\sigma\) using contraction fixed point theorem.

Acknowledgments

The first author wishes to acknowledge the continuous support of the University of Hafr Al Batin, Saudi Arabia.

Author Contributions

All authors contributed equally to the writing of this paper. All authors read and approved the final manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, C. F., Luo, X. N., & Zhou, Y. (2010). Existence of positive solutions of the boundary value problem for nonlinear fractional differential equations. Computers & Mathematics with Applications, 59(3), 1363-1375. [Google Scholor]
  2. Stanek, S. (2011). The existence of positive solutions of singular fractional boundary value problems. Computers & Mathematics with Applications, 62(3), 1379-1388.[Google Scholor]
  3. Xu, X., Jiang, D., & Yuan, C. (2009). Multiple positive solutions for the boundary value problem of a nonlinear fractional differential equation. Nonlinear Analysis: Theory, Methods & Applications, 71(10), 4676-4688. [Google Scholor]
  4. Yuan, C., Jiang, D., & Xu, X. (2009). Singular positone and semipositone boundary value problems of nonlinear fractional differential equations. Mathematical Problems in Engineering, 2009, Article ID 535209. [Google Scholor]
  5. Ha, S. N., & Lee, C. R. (2002). Numerical study for two-point boundary value problems using Green's functions. Computers & Mathematics with Applications, 44(12), 1599-1608.[Google Scholor]
  6. Biala, T. A., & Jator, S. N. (2017). A family of boundary value methods for systems of second-order boundary value problems. International Journal of Differential Equations, 2017, Article ID 2464759. [Google Scholor]
  7. Yongdong, S., & Liangsheng, D. (2001). On the existence of solution of a two-point boundary value problem in a cylindrical floating zone. International Journal of Mathematics and Mathematical Sciences, 25, Article ID 641523. [Google Scholor]
  8. Shi, Y., Zhou, Q., & Li, Y. (1997). A note on a two-point boundary value problem arising from a liquid metal flow. SIAM Journal on Mathematical Analysis, 28(5), 1086-1093. [Google Scholor]
  9. Horgan, C. O., Saccomandi, G., & Sgura, I. (2002). A two-point boundary-value problem for the axial shear of hardening isotropic incompressible nonlinearly elastic materials. SIAM Journal on Applied Mathematics, 62(5), 1712-1727. [Google Scholor]
  10. Xin, L., Guo, Y., & Zhao, J. (2019). Nontrivial Solutions Of Second-Order Nonlinear Boundary Value Problems. Applied Mathematics E-Notes, 19, 668-674. [Google Scholor]
  11. Bai, Z., & Lü, H. (2005). Positive solutions for boundary value problem of nonlinear fractional differential equation. Journal of Mathematical Analysis and Applications, 311(2), 495-505. [Google Scholor]
  12. Bai, C. (2008). Triple positive solutions for a boundary value problem of nonlinear fractional differential equation. Electronic Journal of Qualitative Theory of Differential Equations, 2008(24), 1-10.[Google Scholor]
  13. Kaufmann, E., & Mboumi, E. (2008). Positive solutions of a boundary value problem for a nonlinear fractional differential equation. Electronic Journal of Qualitative Theory of Differential Equations, 2008(3), 1-11. [Google Scholor]
  14. Kosmatov, N. (2009). A singular boundary value problem for nonlinear differential equations of fractional order. Journal of Applied Mathematics & Computing, 29, 125-135. [Google Scholor]
]]>
Common fixed point results of \(s\)-\(\alpha\) contraction for a pair of maps in \(b\)-dislocated metric spaces https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/common-fixed-point-results-of-s-alpha-contraction-for-a-pair-of-maps-in-b-dislocated-metric-spaces/ Wed, 16 Dec 2020 10:41:55 +0000 https://old.pisrt.org/?p=4803
OMA-Vol. 4 (2020), Issue 2, pp. 142 - 151 Open Access Full-Text PDF
Abdissa Fekadu, Kidane Koyas, Solomon Gebregiorgis
Abstract: The purpose of this article is to construct fixed point theorems and prove the existence and uniqueness of common fixed point results of \(s-\alpha\) contraction for a pair of maps in the setting of \(b\) - dislocated metric spaces. Our results extend and generalize several well-known comparable results in the literature. The study procedure we used was that of Zoto and Kumari [1]. Furthermore, we provided an example in support of our main result.
]]>

Open Journal of Mathematical Analysis

Common fixed point results of \(s\)-\(\alpha\) contraction for a pair of maps in \(b\)-dislocated metric spaces

Abdissa Fekadu, Kidane Koyas, Solomon Gebregiorgis\(^1\)
Department of Mathematics, Jimma University, Jimma, Ethiopia.; (A.F & K.K & S.G)
\(^{1}\)Corresponding Author: solomonggty@gmail.com

Abstract

The purpose of this article is to construct fixed point theorems and prove the existence and uniqueness of common fixed point results of \(s-\alpha\) contraction for a pair of maps in the setting of \(b\) – dislocated metric spaces. Our results extend and generalize several well-known comparable results in the literature. The study procedure we used was that of Zoto and Kumari [1]. Furthermore, we provided an example in support of our main result.

Keywords:

Fixed point, \(s-\alpha\) contraction, \(b\)-dislocated metric spaces.

1. Introduction

Fixed point theory is an important tool in the study of nonlinear analysis as it is considered to be the key connection between pure and applied mathematics with wide applications in all branches of Mathematics, Economics, Biology, Chemistry, Physics and almost all engineering fields.

The famous Banach contraction principle is one of the powerful tools in metric fixed point theory and It has been extended and generalized in different directions by different researchers. The notion of metric space has been extended, improved and generalized in many different ways. Bakhtin [2] introduced a \( b \) - metric space as a generalization of metric space and investigated some fixed point theorem in such spaces. Hitzler [3] introduced the notion of dislocated metric spaces. Zeyada et al., [4] generalized the results of Hitzler [3] and introduced the concept of complete dislocated quasi metric space. Aage et al., [5] proved common fixed point theorem in dislocated quasi \( b \) - metric space. Zoto and Kumari [1] constructed theorems on common fixed point results on \( b \) - dislocated metric spaces and proved the existence and uniqueness.

In this research work, we concentrate in establishing and proving common fixed point results for a pair of maps satisfying \( s-\alpha \) contraction condition in the setting \( b \) - dislocated metric spaces.

2. Preliminaries

Throughout this manuscript \(\Re^{+}\) represents the set of non-negative real numbers and \(\mathbf{N}\) represents the set of natural numbers.

Definition 1. [6] Let \( X \) be nonempty set and a mapping \( d_{l}:X\times X\rightarrow\Re^{+} \) is called a dislocated or \( d_{l} \) - metric if the following conditions hold:

  • (a) \( d_{l}(x,y)=0\Rightarrow x=y \) ;
  • (b) \( d_{l}(x,y)=d_{l}(y,x) \) ;
  • (c) \( d_{l}(x,y)\leq d_{l}(x,z)+d_{l}(z,y),\) for all \( x,y\in X \) .
Then the pair \( (X,d_{l}) \) is called a \(d_{l}\) - metric space.

Definition 2. [7] Let \( X \) be nonempty set and \( s\geq 1 \) be a real number, then a mapping \( b_{d}:X\times X\rightarrow \Re^{+}\) is called \( b \) - dislocated metric if the following conditions hold:

  • (a) \( b_{d}(x,y)=0 \Rightarrow x=y \) ;
  • (b) \( b_{d}(x,y)=b_{d}(y,x) \) ;
  • (c) \( b_{d}(x,y)\leq s[b_{d}(x,z)+b_{d}(z,y)],\) for all \( x,y,z\in X \) .
Then the pair \( (X,b_{d}) \) is called a \( b \) - dislocated metric space.

Remark 1. The class of \( b \) - dislocated metric space is larger than that of dislocated metric space.

Definition 3. Let \((X,d)\) be a metric space and \(T : X \rightarrow X\) be a self-map, then \(T\) is said to be a contraction map if there exists a constant \(k \in [0, 1)\) such that \(d(Tx,Ty) \leq kd(x,y)\) for all \(x,y \in X.\)

Definition 4. [1] Let \( (X, b_{d}) \) be a complete \( b \) - dislocated metric space with parameter \( s\geq 1 .\) If \( T:X\rightarrow X \) is self-mapping that satisfy

\begin{equation} s^{2}b_{d}(Tx,Ty)\leq \alpha \max\Bigl\{b_{d}(x,y),b_{d}(x,Tx),b_{d}(y,Ty),b_{d}(x,Ty),b_{d}(y,Tx)\Bigl\} \label{eq 4.1} \end{equation}
(1)
for all \( x,y\in X \) and \( \alpha \in [ 0,\dfrac{1}{2}) \), then \( T \) is called a \( s-\alpha \) quasi-contraction.

Lemma 1. Let \( (X,b_{d}) \) be a \( b \) - dislocated metric space with parameter \( s\geq 1 .\) Suppose that \( \{x_n\}\) and \( \{y_n\} \) are \( b \) - dislocated convergent to \( x,y\in X \) respectively. Then we have \[\frac{1}{s^{2}}b_{d}(x,y)\leq\lim\limits_{n\to\infty}infb_{d}(x_{n},y_{n})\leq\lim\limits_{n\to\infty} \text{Sup}\,b_{d}(x_{n},y_{n})\leq s^{2}b_{d}(x,y).\] In particular, if \( b_{d}(x_{n},y_{n})=0 \), then we have \( \lim\limits_{n\to\infty} b_{d}(x_{n},y_{n})=0=b_{d}(x,y) .\) Moreover, if each \( z\in X \), we have \[\frac{1}{s}b_{d}(x,z)\leq\lim\limits_{n\to\infty}infb_{d}(x_{n},z)\leq\lim\limits_{n\to\infty} \text{Sup}\,b_{d}(x_{n},z)\leq sb_{d}(x,z).\] In particular, if \( b_{d}(x,z)=0 \) , then we have \( \lim\limits_{n\to\infty} b_{d}(x_{n},z)=0=b_{d}(x,z) \).

Theorem 1. [1] Let \( (X,b_{d}) \) be complete \( b \) - dislocated metric space with parameter \( s\geq 1 .\) If \( T:X\rightarrow X\ \) is a self-map that is a \( s-\alpha \) quasi contraction, then \( T \) has a unique fixed point in \( X .\)

Example 1. [1] Let \(X=[0,\infty)\) and \( b_{d}(x,y)=(x+y)^{2} \) for all \( x,y\in X .\) Then \( b_{d} \) is a \( b \) - dislocated metric on \( X \) with parameter \( s=2 \) and is complete.

Definition 5. [8] Let \( (X,b_{d}) \) be a \( b \) - dislocated metric space and \( \{x_n\} \) be a sequence of points in \( X .\) A point \( x \in X \) is said to be the limit of the sequence \( \{x_n\} \) if \(\lim\limits_{n \to\infty} d(x_n,\ x)=0 \) and we say that the sequence \( \{x_n\} \) is \( b \) - dislocated convergent to \( x \) and denote it by \( x_{n}\rightarrow x \) as \( n\rightarrow\infty .\)

Lemma 2. [7] The limit of a convergent sequence in a \( b \) - dislocated metric space is unique.

Definition 6. [7] A sequence \( \{x_n\} \) in a \( b \) - dislocated metric space \( (X,b_{d}) \) is called a \( b \) - dislocated Cauchy sequence if and only if given \( \epsilon> 0 ,\) there exists \( n_{0}\in\mathbb{N} \) such that for all \(n,m>n_{0},\) we have \( b_{d}(x_{n},x_{m})< \epsilon \) or \(\lim\limits_{n,m\to\infty} d(x_n\,\ x_m)=0 .\)

Lemma 3. [7] Every \( b \) - dislocated convergent sequence in \( b \) - dislocated metric space is a \( b \) - dislocated Cauchy.

Definition 7. [7] A \( b \) - dislocated metric space \( (X, b_{d}) \) is called complete if every \( b \) - dislocated Cauchy sequence in \( X \) is a \( b \) - dislocated convergent.

Definition 8. [6] Let \( T:X\rightarrow X \) and \( S:X\rightarrow X \) be self-maps in \((X,b_{d}).\) An element \(x \in X \) is said to be a coincidence point of \( T \) and \( S \) if and only if \( Tx=Sx=u .\) A point \( u \in X \) is point of coincidence of \( T \) and \( S .\)

Definition 9. [9] Let \( T \) and \( S \) be two self-maps on a metric space \( (X,d) .\) Then \( T \) and \( S \) are said to be weakly compatible if they commute at their coincident point; that is \( Tu=Su \) for some \( u \in X \) implies \( STu=TSu.\)

Definition 10. [10] Two self-maps \( T:X\rightarrow X \) and \( S:X\rightarrow X \) are said to be occasionally weakly compatible (OWC) if there exists some point \( u \in X \) such that \( Tu=Su \) and \( STu=TSu.\)

Remark 2. Clearly weakly compatible maps are occasionally weakly compatible. However the converse is not true in general.

Definition 11. [11] Let \( T:X\rightarrow X \) and \( S:X\rightarrow X \) be two self-maps on a metric space \( (X,d) .\) Then \( T \) and \( S \) are said to satisfy the common limit in the range of \( S \) property, denoted by (CLRs) if there exists a sequence \(\{x_n\} \in X \), such that \[\lim\limits_{n\to\infty} Tx_{n}=\lim\limits_{n\to\infty} Sx_{n}=Sx,\] for some \( x \in X .\)

Inspired and motivated by the result of Zoto and Kumari [1], the purpose of this research was to extend and generalize their main theorem to common fixed point theorem involving pairs of \( s-\alpha \) contraction condition in the setting of \( b \) - dislocated metric space.

3. Main results

In this section, we shall state and prove our main results.

Definition 12. Let \( (X,b_{d}) \) be a \( b \) - dislocated metric space with parameter \( s\geq 1 \). If \( T,S :X\rightarrow X \) are self-mapping that satisfy

\begin{equation} \begin{aligned} s^{2}b_{d}(Tx,Ty)\leq &\alpha \max\Bigl\{b_{d}(Sx,Sy),b_{d}(Sx,Tx),b_{d}(Sy,Ty), b_{d}(Sx,Ty),b_{d}(Sy,Tx)\Bigl\} \label{eq 4.2} \end{aligned} \end{equation}
(2)
for all \( x,y\in X \) and \( \alpha \in [ 0,\dfrac{1}{2}) \), then \( T \) and \( S \) are called an \( s-\alpha \) contraction maps.

Theorem 2. Let \( (X,b_{d}) \) be a complete \( b \) - dislocated metric space with parameter \( s\geqslant 1 .\) If \( T,S \) be self-maps of \( X \) such that: The pair \( (T,S) \) satisfy common limit in the range of S property (CLRs) in \( X \) and also \( T \) and \( S \) are \( s-\alpha \) contraction maps, then

  • 1. The pair \( (T,S) \) has a coincidence point in \( X \).
  • 2. The pair \( (T,S) \) has a unique common fixed point provided that \( T \) and \( S \) are weakly compatible mapping.

Proof. Since \( T \) and \( S \) satisfy (CLRs) property, there exists a sequence \( \{x_{n}\} \in X \) such that: \[ \lim\limits_{n\to \infty}Tx_{n}=\lim\limits_{n\to \infty}Sx_{n}=Su \] for some \( u\in X .\) By Equation (2), we have

\begin{equation} s^{2}b_{d}(Tx,Ty)\leq\alpha\max\{b_{d}(Sx,Sy),b_{d}(Sx,Tx),b_{d}(Sy,Ty),b_{d}(Sx,Ty),b_{d}(Sy,Tx)\}.\label{eq 3.2} \end{equation}
(3)
By replacing \( x=u \) and \( y=x_{n} \) in the above condition, we obtain \[s^{2}b_{d}(Tu,Tx_{n})\leq\alpha\max\{b_{d}(Su,Sx_{n}),b_{d}(Su,Tu),b_{d}(Sx_{n},Tx_{n}),b_{d}(Su,Tu),b_{d}(Sx_{n},Tu)\}.\] By applying Lemma 1 and taking the upper limit as \( n\rightarrow\infty \) on Equation (3), we get \begin{eqnarray*} s^{2}b_{d}(Tu,Su)&\leq&\alpha\max\{b_{d}(Su,Su),b_{d}(Su,Tu),b_{d}(Su,Su),b_{d}(Su,Su),b_{d}(Su,Tu)\}\\ &=& \alpha\max\{b_{d}(Su,Tu),b_{d}(Su,Su)\}. \end{eqnarray*} We consider the following cases as follows;

Case 1: If \( max\{b_{d}(Su,Tu),b_{d}(Su,Su)\} =b_{d}(Su,Tu)\), then we have \(b_{d}(Tu,Su)\leq\frac{\alpha}{s^2} b_{d}(Su,Tu).\) It follows that \(b_{d}(Su,Tu)=0\) since \( 0\leq\alpha< \frac{1}{2}.\) Hence \( Su=Tu=z \) (say).

Case 2: If \( max\{b_{d}(Su,Tu),b_{d}(Su,Su)\} =b_{d}(Su,Su)\), then we have \begin{eqnarray*} s^{2}b_{d}(Tu,Su)&\leq&\alpha\ b_{d}(Su,Su)\\ &\leq&\ s \alpha [b_{d}(Su,Tu)+b_{d}(Tu,Su)]\\ &=&\ 2s\alpha b_{d}(Su,Tu). \end{eqnarray*}

It follows that \(b_{d}(Tu,Su)\leq\frac{2\alpha}{s} b_{d}(Su,Tu),\) which in turn implies that \(b_{d}(Su,Tu)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( Su=Tu=z \) (say). Therefore, \( T \) and \( S \) have a coincidence point. Now, by weakly compatibility property of the pair \( (T,S) \), we have \[ Tz=T(Su)=STu=Sz\] which implies that \( Tz=Sz .\) Now, we show existence of a common fixed point. First, we show that \( z \) is a fixed point of \( T \). By Equation (2), we have \[s^{2}b_{d}(Tz,Tx_{n})\leq\alpha\max\{b_{d}(Sz,Sx_{n}),b_{d}(Sz,Tz),b_{d}(Sx_{n},Tx_{n}),b_{d}(Sz,Tz),b_{d}(Sx_{n},Tz)\}.\] Taking the upper limit as \( n\rightarrow\infty \), we get \begin{eqnarray*} s^{2}b_{d}(Tz,z)&\leq&\alpha\max\{\lim\limits_{n\to\infty}sup[b_{d}(Sz,Sx_{n}),b_{d}(Sz,Tz),b_{d}(Sx_{n},Tx_{n}),b_{d}(Sz,Tx_{n}),b_{d}(Sx_{n},Tz)]\}\\ &=& \alpha\max\{b_{d}(Sz,Su),b_{d}(Sz,Tz),b_{d}(Su,Su),b_{d}(Sz,Su),b_{d}(Su,Tz)\}. \end{eqnarray*} Since \( Su=Tu=z \) and \( Tz=Sz \), we have \begin{eqnarray*} s^{2}b_{d}(Tz,z)&\leq&\alpha\max\{b_{d}(Tz,z),b_{d}(Sz,Sz),b_{d}(z,z),b_{d}(Tz,z),b_{d}(z,Tz)\}\\&=&\alpha\max\{b_{d}(Tz,z),b_{d}(Tz,Tz),b_{d}(z,z),b_{d}(Tz,z),b_{d}(z,Tz)\}. \end{eqnarray*} Now, we consider the following cases as follows;

Case 1: If \( max\{b_{d}(Tz,z),b_{d}(Tz,Tz),b_{d}(z,z)\} =b_{d}(Tz,z)\), then we have \(b_{d}(Tz,z)\leq\frac{\alpha}{s^2} b_{d}(Tz,z),\) which implies \(b_{d}(Tz,z)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence it follows that \( Tz=z. \)

Case 2: If \( max\{b_{d}(Tz,z),b_{d}(Tz,Tz),b_{d}(z,z)\} =b_{d}(Tz,Tz)\), then we have \begin{eqnarray*} s^{2}b_{d}(Tz,z)&\leq&\alpha\ b_{d}(Tz,Tz)\\ &\leq&\ s\alpha [b_{d}(z,Tz)+b_{d}(Tz,z)]\\ &=&\ 2s\alpha b_{d}(Tz,z). \end{eqnarray*} It follows that \(b_{d}(Tz,z)\leq\frac{2\alpha}{s}b_{d}(Tz,z),\) which implies \(b_{d}(Tz,z)=0\) Since \( 0\leq\alpha< \frac{1}{2}.\) Hence \( Tz=z \).

Case 3: If \( max\{b_{d}(Tz,z),b_{d}(Tz,Tz),b_{d}(z,z)\} =b_{d}(z,z)\), then we have \begin{eqnarray*} s^{2}b_{d}(Tz,z)&\leq&\alpha\ b_{d}(z,z)\\ &\leq&\ s\alpha [b_{d}(z,Tz)+b_{d}(Tz,z)]\\ &=&\ 2s\alpha b_{d}(Tz,z). \end{eqnarray*} It follows that \(b_{d}(Tz,z)\leq\frac{2\alpha}{s^2}b_{d}(Tz,z),\) which implies \(b_{d}(Tz,z)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( Tz=z \). But we know that \( Tz=Sz\) which gives us \( Tz=Sz=z\). Therefore, \(z\) is a common fixed point of \(T\) and \(S\).

Uniqueness

Let \(z\) and \(z\prime\) be fixed points of \(T\) and \(S\) with \(z\neq z\prime\) . Then by Equation (2), we have \begin{eqnarray*} s^{2}b_{d}(Tz,Tz\prime) &\leq&\alpha\max \{b_{d}(Sz,Sz\prime),b_{d}(Sz,Tz),b_{d}(Sz\prime,Tz\prime),b_{d}(Sz,Tz\prime),b_{d}(Sz\prime,Tz)\}\\ &=&\alpha\max\{b_{d}(z,z\prime),b_{d}(z,z),b_{d}(z\prime,z\prime),b_{d}(z,z\prime),b_{d}(z\prime,z)\}. \end{eqnarray*} We consider the following cases as follows.

Case 1: If \( max\{b_{d}(z\prime,z),b_{d}(z,z),b_{d}(z\prime,z\prime)\} =b_{d}(z,z\prime)\), then we have \(b_{d}(z\prime,z)\leq\frac{\alpha}{s^2} b_{d}(z,z\prime)\) which implies \(b_{d}(z,z\prime)=0\) since \( 0\leq\alpha< \frac{1}{2}.\) Hence it follows that \( z=z\prime \).

Case 2: If \( max\{b_{d}(z\prime,z),b_{d}(z,z),b_{d}(z\prime,z\prime)\} =b_{d}(z,z)\), then we have \begin{eqnarray*} s^{2}b_{d}(z,z\prime)&\leq&\alpha\ b_{d}(z,z)\\ &\leq&\ s\alpha [b_{d}(z,z\prime)+b_{d}(z\prime,z)]\\ &=&\ 2s\alpha b_{d}(z,z\prime). \end{eqnarray*} It follows that \(b_{d}(z,z\prime)\leq\frac{2\alpha}{s} b_{d}(z,z\prime)\) which implies \(b_{d}(z\prime,z)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( z=z\prime \).

Case 3: If \( max\{b_{d}(z,z\prime),b_{d}(z,z),b_{d}(z\prime,z\prime)\} =b_{d}(z\prime,z\prime)\), then we have \begin{eqnarray*} s^{2}b_{d}(z,z\prime)&\leq&\alpha\ b_{d}(z\prime,z\prime)\\ &\leq&\ s\alpha [b_{d}(z,z\prime) b_{d}(z\prime,z)]\\ &=&\ 2s\alpha b_{d}(z\prime,z). \end{eqnarray*} It follows that \(b_{d}(z\prime,z)\leq\frac{2\alpha}{s} b_{d}(z\prime,z\prime)\) which implies \(b_{d}(z\prime,z)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( z\prime=z \) which contradicts to our assumption \( z\neq z\prime \).

Hence \( z \) is a unique common fixed point of \( T \) and \( S .\)

Theorem 3. Let \( (X,b_{d}) \) be a complete \( b \) - dislocated metric space with parameter \( s\geqslant 1 .\) If \( T,S \) be self-maps of \( X \) such that the pair \( (T,S) \) satisfy occasionally weakly compatible property (OWC) in \( X \) and \(T\) and \(S\) are an \( s-\alpha \) contraction maps. Then the pair \( (T,S) \) has a unique common fixed point.

Proof. Since \( T \) and \( S \) satisfy (OWC) property, there exists a point \( u \in X \) such that \(Tu=Su\) and \( TSu=STu.\) This implies that \( TSu=TTu=STu=SSu.\) It follows that \( TTu=SSu. \) We claim that \( Tu \) is the unique common fixed point of \( T \) and \( S .\) First, we assert that \( Tu \) is a fixed point of \( T .\) For, if \( TTu\neq Tu ,\) then by Equation (2), we get \begin{equation*} s^{2}b_{d}(Tu,TTu)\leq\alpha\max\Bigl\{b_{d}(Su,STu),b_{d}(Su,Tu),b_{d}(STu,TTu),b_{d}(Su,TTu),b_{d}(STu,Tu)\Bigr\}. \end{equation*} Since \( Su=Tu, \) then we have \begin{eqnarray*} s^{2}b_{d}(Tu,TTu)&\leq&\alpha\max\Bigl\{b_{d}(Tu,TTu),b_{d}(Tu,Tu),b_{d}(TTu,TTu),b_{d}(Tu,TTu),b_{d}(TTu,Tu)\Bigr\}\\ &=&\alpha\max\Bigl\{b_{d}(Tu,TTu),b_{d}(Tu,Tu),b_{d}(TTu,TTu)\Bigr\}. \end{eqnarray*} We consider the following three cases as follows.

Case 1: If \( max\{b_{d}(Tu,TTu),b_{d}(Tu,Tu),b_{d}(TTu,TTu)\} =b_{d}(Tu,TTu)\), then we have \(b_{d}(Tu,TTu)\leq\frac{\alpha}{s^2} b_{d}(Tu,TTu),\) which implies \(b_{d}(Tu,TTu)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( Tu=TTu \).

Case 2: If \( max\{b_{d}(Tu,TTu),b_{d}(Tu,Tu),b_{d}(TTu,TTu)\} =b_{d}(Tu,Tu)\), then we have \begin{eqnarray*} s^{2}b_{d}(Tu,TTu)&\leq&\alpha\ b_{d}(Tu,Tu)\\ &\leq&\ s \alpha [b_{d}(TTu,Tu)+b_{d}(Tu,TTu)]\\ &=&\ 2s\alpha b_{d}(TTu,Tu). \end{eqnarray*} It follows that \(b_{d}(TTu,Tu)\leq\frac{2\alpha}{s} b_{d}(TTu,Tu)\), which implies that \(b_{d}(TTu,Tu)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( TTu=Tu \).

Case 3: If \( max\{b_{d}(Tu,TTu),b_{d}(Tu,Tu),b_{d}(TTu,TTu)\} =b_{d}(TTu,TTu)\), then we have \begin{eqnarray*} s^{2}b_{d}(Tu,TTu)&\leq&\alpha\ b_{d}(TTu,TTu)\\ &\leq&\ s \alpha [b_{d}(TTu,Tu)+b_{d}(Tu,TTu)]\\ &=&\ 2s\alpha b_{d}(TTu,Tu). \end{eqnarray*} It follows that \(b_{d}(TTu,Tu)\leq\frac{2\alpha}{s} b_{d}(TTu,Tu),\) which implies that \(b_{d}(TTu,Tu)=0\) since \( 0\leq\alpha< \frac{1}{2} .\) Hence \( Tu=TTu \) which is a contradiction with \( Tu \neq TTu .\) There fore, \( TTu=Tu \) and \( Tu \) is the fixed point of \( T. \) Since \( TTu=TSu=STu=Tu=SSu, \) it implies \(STu=Tu.\) Thus \( Tu \) is fixed point of \( S. \) Therefore, \(Tu\) is a common fixed point of \( T \) and \( S .\)

Uniqueness

Suppose that \( u, v \in X \) such that \( Tu=Su=u \) and \( Tv=Sv=v \) and \( u\neq v.\) Then by Equation (2), we get \begin{eqnarray*} s^{2}b_{d}(u,v)&=& s^{2}b_{d}(Tu,Tv)\\ &\leq&\alpha\max\Bigl\{b_{d}(Su,Sv),b_{d}(Su,Tu),b_{d}(Sv,Tv),b_{d}(Su,Tv),b_{d}(Sv,Tu)\Bigr\}\\ &=&\alpha\max\Bigl\{b_{d}(u,v),b_{d}(u,u),b_{d}(v,v),b_{d}(u,v),b_{d}(v,u)\Bigr\}\\ &=&\alpha\max\Bigl\{b_{d}(u,v),b_{d}(u,u),b_{d}(v,v)\Bigr\}. \end{eqnarray*} We consider the following cases as follows.

Case 1: If \( max\{b_{d}(u,v),b_{d}(u,u),b_{d}(v,v)\} =b_{d}(v,u)\), then we have \(b_{d}(u,v)\leq\frac{\alpha}{s^2} b_{d}(v,u),\) which implies that \(b_{d}(v,u)=0\), since \( 0\leq\alpha< \frac{1}{2}. \) Hence it follows that \( u=v .\)

Case 2: If \( max\{b_{d}(v,u),b_{d}(u,u),b_{d}(v,v)\} =b_{d}(u,u)\), then we have \begin{eqnarray*} s^{2}b_{d}(v,u)&\leq&\alpha\ b_{d}(u,u)\\ &\leq&\ s\alpha [b_{d}(u,v)+b_{d}(u,v)]\\ &=&\ 2s\alpha b_{d}(u,v). \end{eqnarray*} It follows that \(b_{d}(v,u)\leq\frac{2\alpha}{s} b_{d}(v,u),\) which implies that \(b_{d}(v,u)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( v=u \).

Case 3: If \( max\{b_{d}(u,v),b_{d}(v,v),b_{d}(u,u)\} =b_{d}(v,v)\), then we have \begin{eqnarray*} s^{2}b_{d}(u,v)&\leq&\alpha\ b_{d}(v,v)\\ &\leq&\ s\alpha [b_{d}(u,v) b_{d}(v,u)]\\ &=&\ 2s\alpha b_{d}(u,v). \end{eqnarray*} It follows that \(b_{d}(u,v)\leq\frac{2\alpha}{s} b_{d}(v,v),\) which implies that \(b_{d}(v,u)=0\), since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( u=v \). Therefore, it contradicts with our assumption \( u\neq v .\)

Hence \( u \) is a unique common fixed point of \( T \) and \( S .\)

Theorem 4. Let \( (X, b_{d}) \) be a complete \( b \) - dislocated metric space with parameter \( s\geqslant 1.\) If \( T,S \) be self-maps of \( X \) such that \[s^{2}b_{d}(Tx,Ty)\leq\alpha\max\Bigl\{b_{d}(Sx,Sy),b_{d}(Sx,Tx),b_{d}(Sy,Ty),b_{d}(Sx,Ty),b_{d}(Sy,Tx)\Bigr\}\] for all \( x,y\in X \) and \( 0\leq\alpha< \frac{1}{2}.\) Then the pair \( (T,S) \) has a unique common fixed point.

Proof. Let \( x_{0} \) be arbitrary given point in \( X \). Define the sequence \( {y_{n} }\in X \) such that; \( y_{2n}=Tx_{2n}=Sx_{2n+1} ,\) for all \( n\geq 0. \) We show that \( {y_{n} }\in X \) for all \( n\in\mathbb{N}.\) Since \(y_{2n}=Tx_{2n}=Sx_{2n+1},\) we have from (1) that \begin{eqnarray*} s^{2}b_{d}(y_{2n}, y_{2n+1})&=& s^{2}b_{d}(Tx_{2n},Tx_{2n+1})\\ &\leq&\alpha\max\Bigl\{b_{d}(Sx_{2n},Sx_{2n+1}),b_{d}(Sx_{2n},Tx_{2n}),b_{d}(Sx_{2n+1},Tx_{2n+1}),b_{d}(Sx_{2n},Tx_{2n+1}),{}b_{d}(Sx_{2n+1},Tx_{2n}x)\Bigr\}\\ &=&\alpha\max\Bigl\{b_{d}(y_{2n-1},y_{2n}),b_{d}(y_{2n-1},y_{2n}),b_{d}(y_{2n},y_{2n+1}),b_{d}(y_{2n-1},y_{2n+1}),b_{d}(y_{2n},y_{2n})\Bigr\}\\ &\leq&\alpha\max\Bigl\{b_{d}(y_{2n-1},y_{2n}),b_{d}(y_{2n-1},y_{2n}),b_{d}(y_{2n},y_{2n+1}),s[b_{d}(y_{2n-1},y_{2n})+b_{d}(y_{2n},{}y_{2n+1})],s[b_{d}(y_{2n},y_{2n+1}+b_{d}(y_{2n-1},y_{2n})]\Bigr\} \end{eqnarray*} \begin{eqnarray*} &=&\alpha\max\Bigl\{b_{d}(y_{2n-1},y_{2n}),b_{d}(y_{2n-1},y_{2n}),b_{d}(y_{2n},y_{2n+1}),s[b_{d}(y_{2n-1},y_{2n})+b_{d}(y_{2n},{}y_{2n+1})],2s[b_{d}(y_{2n-1},y_{2n})]\Bigr\}. \end{eqnarray*} If \(b_{d}(y_{2n-1},y_{2n})\leq b_{d}(y_{2n},y_{2n+1}),\) for some \( n\in\mathbb{N},\) then by (1), we have \(s^{2}b_{d}(y_{2n},y_{2n+1})\leq2\alpha b_{d}(y_{2n},y_{2n+1}),\) which implies that \(b_{d}(y_{2n},y_{2n+1})\leq\frac{2\alpha}{s^2} b_{d}(y_{2n},y_{2n+1}).\) This is not true because \(\frac{2\alpha}{s^2}< 1.\) Thus \( b_{d}(y_{2n},y_{2n+1})\leq b_{d}(y_{2n-1},y_{2n})\) for all \( n\in\mathbb{N}.\) Also, by the above inequality, we get \begin{eqnarray*} s^2b_{d}(y_{2n},y_{2n+1})&\leq& 2\alpha s b_{d}(y_{2n-1},y_{2n}),\\ b_{d}(y_{2n-1},y_{2n})&\leq&\frac{2\alpha}{s} b_{d}(y_{2n-2},y_{2n-1}),\\ b_{d}(y_{2n-2},y_{2n-1})&\leq&\frac{2\alpha}{s} b_{d}(y_{2n-3},y_{2n-4}). \end{eqnarray*} Continuing like this, we have \[b_{d}(y_{2n},y_{2n+1})\leq cb_{d}(y_{2n-1},y_{2n})\leq c^{2}b_{d}(y_{2n-2},y_{2n-1}), \cdots \leq c^{n}b_{d}(y_{0},y_{1}),\] where \( c=\frac{2\alpha}{s} \) and \( 0\leq c< 1. \) Taking the upper limit as \( n\rightarrow\infty \) in the inequality above, we have \(b_{d}(y_{2n},y_{2n+1})\rightarrow 0.\) Now, we prove that \(\{y_{m}\}\) is a \( b \) - dislocated Cauchy sequence where \( m=2n. \) To do this let \(m,n\geq 0 \) with \( m>n.\) Now, using the triangle inequality, we have

\begin{align} b_{d}(y_{n},y_{m})&\leq s\left[b_{d}(y_{n},y_{n+1})+b_{d}(y_{n+1},y_{m})\right] \notag\\ &\leq s\Bigl[b_{d}(y_{n},y_{n+1})+s\left[b_{d}(y_{n+1},y_{n+2})+b_{d}(y_{n+2},y_{m})\right]\Bigr] \notag\\ &\leq sb_{d}(y_{n},y_{n+1})+s^{2}b_{d}(y_{n+1},y_{n+2})+\cdots+s^{n}b_{d}(y_{m-1},y_{m})\notag\\ &= s\left[c^{n}b_{d}(y_{0},y_{1})+sc^{n+1}b_{d}(y_{0},y_{1})+\cdots \right]\notag\\ &= sc^{n} \left[b_{d}(y_{0},y_{1})+scb_{d}(y_{0},y_{1})+\cdots \right] \notag\\ &= sc^{n}b_{d}(y_{0},y_{1})\left[1+sc+(sc)^{2}+\cdots \right] \notag\\ &\leq\frac{sc^{n}}{(1-sc)}b_{d}(y_{0},y_{1}). \end{align}
(4)
Therefore \( b_{d}(y_{n},y_{m})\leq\frac{sc^{n}}{(1-sc)}b_{d}(y_{0},y_{1}).\) Taking the upper limit as \( m,n\rightarrow\infty, \) we have \( b_{d}(y_{n},y_{m})\rightarrow 0 \) as \( sc< 1. \) Therefore, \(\{\ y_{m}\} \) is a \( b \) -dislocated Cauchy sequence in \( b \) - dislocated metric space \( (X, b_{d}).\) So there is some \( u \in X\) such that \(\{\ y_{m}\} \) is a \( b \) -dislocated converges to \( u \). Since a subsequence of a Cauchy sequence in \( b \) -dislocated metric space is a Cauchy sequence, then \(\{Tx_{2n}\}\) and \(\{Sx_{2n+1}\}\) are also Cauchy sequences.

If \(T\) and \( S \) are continuous mappings, we get

\[ T(u)=T(\lim\limits_{n\to \infty}x_{n})=\lim\limits_{n\to \infty}Tx_{n}= S(u)= \lim\limits_{n\to \infty}Sx_{n+1}=u .\] Thus, \( u \) is a common fixed point of \( T \) and \( S \). If the self-map \( T \) is not continuous then, we consider \begin{eqnarray*} s^{2}b_{d}(y_{2n}, Tu)&=& s^{2}b_{d}(Tx_{2n},Tu)\\ &\leq&\alpha\max\Bigl\{b_{d}(Sx_{2n},Su),b_{d}(Sx_{2n},Tx_{2n}),b_{d}(Su,Tu),b_{d}(Sx_{2n},Tu),b_{d}(Su,Tx_{2n})\Bigr\}\\ &=&\alpha\max\Bigl\{b_{d}(y_{2n-1},Su),b_{d}(y_{2n-1},y_{2n}),b_{d}(Su,Tu),b_{d}(y_{2n-1},Tu),b_{d}(Su,y_{2n})\Bigr\}. \end{eqnarray*} On taking upper limit as \( n\rightarrow \infty ,\) we get \( s^{2}b_{d}(u,Tu)\leq\alpha\max \Bigl\{b_{d}(Su,Su),b_{d}(Su,Tu),b_{d}(Su,Tu)\Bigr\}\). We consider the following cases:

Case 1: If \( max\{b_{d}(u,u),b_{d}(Tu,Tu),b_{d}(u,Tu)\} =b_{d}(Tu,u)\), then we have \(b_{d}(Tu,u)\leq\frac{\alpha}{s^2} b_{d}(Tu,u),\) which implies \(b_{d}(Tu,u)=0\), since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( Tu=u \).

Case 2: If \( max\{b_{d}(u,u),b_{d}(Tu,Tu),b_{d}(Tu,u)\} =b_{d}(Tu,Tu)\), then we have \begin{eqnarray*} s^{2}b_{d}(Tu,u)&\leq&\alpha\ b_{d}(Tu,Tu)\\ &\leq&\ s\alpha [b_{d}(u,Tu)+b_{d}(Tu,u)]\\ &=&\ 2s\alpha b_{d}(Tu,u). \end{eqnarray*} It follows that \(b_{d}(Tu,u)\leq\frac{2\alpha}{s} b_{d}(Tu,u),\) which implies \(b_{d}(Tu,u)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( Tu=u \).

Case 3: If \( max\{b_{d}(u,u),b_{d}(Tu,Tu),b_{d}(Tu,u)\} =b_{d}(u,u)\), then we have \begin{eqnarray*} s^{2}b_{d}(Tu,u)&\leq&\alpha\ b_{d}(u,u)\\ &\leq&\ s\alpha [b_{d}(u,Tu) b_{d}(Tu,u)]\\ &=&\ 2s\alpha b_{d}(Tu,u). \end{eqnarray*} It follows that \(b_{d}(Tu,u)\leq\frac{2\alpha}{s} b_{d}(Tu,u),\) which implies \(b_{d}(Tu,u)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( Tu=u \).

In all cases \( b_{d}(u,Tu)=0 \) which implies that \( u=Tu. \) Thus, \( u \) is fixed point of \( T.\)

In similar cases, we have \(b_{d}(Su,u)\leq\frac{2\alpha}{s} b_{d}(Su,u),\) which implies \(b_{d}(Su,u)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence \( Su=u \). Thus, \( u \) is fixed point of \( S.\) Since, \( Su=u=Tu, \) then \( u \) is a common fixed point of \( T \) and \( S .\)

Uniqueness

Let \( u \) and \( v \) are fixed points of \( T \) and \( S \) with \( u\neq v.\) Then by using Equation (2), we have \begin{eqnarray*} s^{2}b_{d}(u,v)=s^{2}b_{d}(Tu,Tv)&\leq&\alpha\max \Bigl\{b_{d}(Su,Sv),b_{d}(Su,Tu),b_{d}(Sv,Tv),b_{d}(Su,Tv),b_{d}(Sv,Tu)\Bigr\}\\ &=&\alpha\max \Bigl\{b_{d}(u,v),b_{d}(u,u),b_{d}(v,v),b_{d}(u,v),b_{d}(v,u)\Bigr\}\\ &=&\alpha\max \Bigl\{b_{d}(u,v),b_{d}(v,v),b_{d}(v,u)\Bigr\}. \end{eqnarray*} We consider the following three cases;

Case 1: If \( max\{b_{d}(u,v),b_{d}(u,u),b_{d}(v,v)\} =b_{d}(v,u)\), then we have \(b_{d}(u,v)\leq\frac{\alpha}{s^2} b_{d}(v,u),\) which implies \(b_{d}(v,u)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence it follows that \( u=v \).

Case 2: If \( max\{b_{d}(v,u),b_{d}(u,u),b_{d}(v,v)\} =b_{d}(u,u)\), then we have \begin{eqnarray*} s^{2}b_{d}(v,u)&\leq&\alpha\ b_{d}(u,u)\\ &\leq&\ s\alpha [b_{d}(u,v)+b_{d}(u,v)]\\ &=&\ 2s\alpha b_{d}(u,v). \end{eqnarray*} It follows that \(b_{d}(v,u)\leq\frac{2\alpha}{s} b_{d}(v,u),\) which implies \(b_{d}(v,u)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence it follows that \( v=u \).

Case 3: If \( max\{b_{d}(u,v),b_{d}(v,v),b_{d}(u,u)\} =b_{d}(v,v)\), then we have \begin{eqnarray*} s^{2}b_{d}(v,u)&\leq&\alpha\ b_{d}(v,v)\\ &\leq&\ s\alpha [b_{d}(u,v)+b_{d}(v,u)]\\ &=&\ 2s\alpha b_{d}(u,v). \end{eqnarray*} It follows that \(b_{d}(u,v)\leq\frac{2\alpha}{s} b_{d}(v,u),\) which implies \(b_{d}(v,u)=0\) since \( 0\leq\alpha< \frac{1}{2}. \) Hence in all cases \( u=v \). Therefore, it contradicts with our assumption \( u\neq v \). Hence \( u \) is a unique common fixed point of \( T \) and \( S .\)

Remark 3. If we take \(S=I\) (\(I\) is the identity map on \(X\)) in Theorem 4, we get Theorem 1.

Now, we give an example in support of Theorem 4.

Example 2. Let \(X=[0,1]\) and \(b_{d}(x,y)=(x+y)^2\) for all \(x, y \in X\) when \(s=2\) is a b-dislocated metric on \(X\). Then \((X,b_{d})\) is a b-dislocated metric space.

Solution. We take the \(s-\alpha\) contraction map and define the following \[Tx=\begin{cases} \frac{1}{25}x, \text{ if } x \in [0,1)\\ \frac{1}{30}, \text{ if } x=1 \end{cases} \text{ and }\;\;\;\; Sx=\frac{1}{5}x.\] Consider the sequence \(\{x_n\}\) given by \(X_n= \frac{1}{n}\) for all \(n \in \mathbf{N}.\) \(T\) and \(S\) satisfies the common limit in the range of \(S\) property. By using (CLRs) properties, we have \(\lim\limits_{n \to \infty}Tx_n = \lim\limits_{n \to \infty}Sx_n=0,\) for \( 0 \in Tx\) or \( 0 \in Sx.\) We have the following cases with \(\alpha = \frac{4}{25}. \)

Case 1: For \(x, y \in [0,1)\), we have \begin{equation*} s^2 b_{d}(Tx,Ty) = 4 b_{d}\biggl(\frac{1}{25}x,\frac{1}{25}y\biggr) = 4 \biggl(\frac{1}{25}x+\frac{1}{25}y\biggr)^2 = \frac{4}{25} \biggl(\frac{1}{5}x+\frac{1}{5}y\biggr)^2 \leq \frac{4}{25} b_{d}(Sx,Sy). \end{equation*} Therefore, \(s^2b_{d}(Tx,Ty)\leq \alpha b_{d}(Sx,Sy).\)

Case 2: For \(y< x\) and \(x=1\), we have \begin{equation*} s^2 b_{d}(T1,Ty) = 4 b_{d}\biggl(\frac{1}{30},\frac{1}{25}y\biggr) = 4 \biggl(\frac{1}{30}+\frac{1}{25}y\biggr)^2 \leq \frac{4}{25} \biggl(\frac{1}{5}+\frac{1}{5}y\biggr)^2 = \frac{4}{25} b_{d}\biggl(\frac{1}{5},\frac{1}{5}y\biggr) \leq \frac{4}{25} b_{d}(Sx,Sy). \end{equation*} Therefore, \(s^2b_{d}(T1,Ty) \leq \alpha b_{d}(S1,Sy)= \alpha b_{d}(Sx,Sy).\)

Case 3: For \(x< y\) and \(y=1\), we have \begin{equation*} s^2b_{d}(Tx,T1) = 4 b_{d}\biggl(\frac{1}{25}x,\frac{1}{30}\biggr) = 4 \biggl(\frac{1}{25}x+\frac{1}{30}\biggr)^2 \leq \frac{4}{25} \biggl(\frac{1}{5}x+\frac{1}{5}\biggr)^2 = \frac{4}{25} b_{d}\biggl(\frac{1}{5}x,\frac{1}{5}\biggr) \leq \frac{4}{25} b_{d}(Sx,Sy). \end{equation*} Therefore, \(s^2b_{d}(Tx,T1) \leq \alpha b_{d}(Sx,S1)= \alpha b_{d}(Sx,Sy).\)

Case 4: For \(x=y=1\), we have \begin{equation*} s^2 b_{d}(T1,T1) = 4 b_{d}\biggl(\frac{1}{30},\frac{1}{30}\biggr) = \frac{4}{25} \biggl(\frac{1}{6}+\frac{1}{6}\biggr)^2 \leq \frac{4}{25} \biggl(\frac{1}{5}+\frac{1}{5}\biggr)^2 = \frac{4}{25} b_{d}\biggl(\frac{1}{5},\frac{1}{5}\biggr) = \frac{4}{25} b_{d}\biggl(S1,S1\biggr) \leq \frac{4}{25} b_{d}(Sx,Sy). \end{equation*}

Therefore, \(s^2b_{d}(T1,T1) \leq \alpha b_{d}(S1,S1)=\alpha b_{d}(Sx,Sy).\)

From cases 1 up to 4, we see that all the conditions of Theorem 4 are satisfied and \(0\) is the unique common fixed point of \(T\) and \(S\).

4. Conclusion

Zoto and Kumari [1] established the existence and uniqueness of fixed point for a mapping satisfying \( s-\alpha \) type contraction condition in a complete dislocated metric space. In this thesis, we have explored the properties of \( s-\alpha \) type contraction mapping in \( b \) - dislocated metric spaces. We established the theorem on common fixed points of two mapping satisfying \( s-\alpha \) contraction condition in the setting of \( b \) - dislocated metric spaces and proved the existence and uniqueness of common fixed point for a pair of maps T and S in the setting of \( b \) - dislocated metric space. Also we provided an example in support of our main results. Our work extended fixed point result in single map to common fixed point result in a pair of maps. The presented theorem extends and generalizes several well-known comparable results in literature.

Acknowledgments

The authors would like to thank the College of Natural Sciences, Jimma University for funding this research work.

Author Contributions

All authors contributed equally to the writing of this paper. All authors read and approved the final manuscript.

Conflicts of interest

The authors declare no conflict of interest.

References

  1. Zoto, K., & Kumari, P. S. (2019). Fixed point theorems for \(s-\alpha \) contractions in dislocated and \(b\)-dislocated metric space. Thai Journal of Mathematics, 17(1), 263-276. [Google Scholor]
  2. Bakhtin, I. (1989). The contraction mapping principle in quasimetric spaces. Functional Analysis, Gos. Ped. Inst. Unianowsk, 30, 26-37. [Google Scholor]
  3. Hitzler, P., & Seda, A. K. (2000). Dislocated topologies. Journal of Electrical Engineering, 51(12), 3-7. [Google Scholor]
  4. Zeyada, F. M., Hassan, G. H., & Ahmed, M. A. (2006). A generalization of a fixed point theorem due to Hitzler and Seda in dislocated quasi-metric spaces. Arabian journal for Science and Engineering, 31(1A), 111.[Google Scholor]
  5. Aage, C. T., & Golhare, P. G. (2016). On fixed point theorems in dislocated quasi \(b\)-metric spaces. International Journal of Advances in Matematics, 2016 (1), 55-70. [Google Scholor]
  6. Sintunavarat, W., & Kumam, P. (2011). Common fixed point theorems for a pair of weakly compatible mappings in fuzzy metric spaces. Journal of Applied mathematics, 2011, Article ID 637958. [Google Scholor]
  7. Kutbi, M. A., Arshad, M., Ahmad, J., & Azam, A. (2014). Generalized common fixed point results with applications. Abstract and Applied Analysis, 2014, Article ID 363925. [Google Scholor]
  8. Hussain, N., Roshan, J. R., Parvaneh, V., & Abbas, M. (2013). Common fixed point results for weak contractive mappings in ordered \(b\)-dislocated metric spaces with applications. Journal of Inequalities and Applications, 2013, Article No. 486. [Google Scholor]
  9. Jungck, G. (1996). Common fixed points for noncontinuous nonself maps on nonmetric spaces. Far East Journal of Mathematical Sciences, 4(2), 199-215. [Google Scholor]
  10. Jungck, G., & Rhoades, B. E. (2006). Fixed point theorems for occasionally weakly compatible mappings. Fixed Point Theory, 7(2), 287-296. [Google Scholor]
  11. Verma, R. K., & Pathak, H. K. (2012). Common fixed point theorems using property (EA) in complex-valued metric spaces. Thai Journal of Mathematics, 11(2), 347-355. [Google Scholor]
  12. Banach, S. (1922). Sur les opérations dans les ensembles abstraits et leur application aux équations intégrales. Fundamenta Mathematicae, 3(1), 133-181. [Google Scholor]
]]>
On the existence of positive solutions of a state-dependent neutral functional differential equation with two state-delay functions https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/on-the-existence-of-positive-solutions-of-a-state-dependent-neutral-functional-differential-equation-with-two-state-delay-functions/ Mon, 14 Dec 2020 14:06:32 +0000 https://old.pisrt.org/?p=4792
OMA-Vol. 4 (2020), Issue 2, pp. 132 - 141 Open Access Full-Text PDF
El-Sayed, A. M. A, Hamdallah, E. M. A, Ebead, H. R
Abstract: In this paper, we study the existence of positive solutions for an initial value problem of a state-dependent neutral functional differential equation with two state-delay functions. The continuous dependence of the unique solution will be proved. Some especial cases and examples will be given.
]]>

Open Journal of Mathematical Analysis

On the existence of positive solutions of a state-dependent neutral functional differential equation with two state-delay functions

El-Sayed, A. M. A, Hamdallah, E. M. A, Ebead, H. R\(^1\)
Faculty of Science, Alexandria~University, Alexandria, 21500, Egypt.; (E.S.A.M.A & H.E.M.A & E.H.R)
\(^{1}\)Corresponding Author: hanaaRezqalla@alexu.edu.eg

Abstract

In this paper, we study the existence of positive solutions for an initial value problem of a state-dependent neutral functional differential equation with two state-delay functions. The continuous dependence of the unique solution will be proved. Some especial cases and examples will be given.

Keywords:

Neutral differential equations, state-dependent, existence of positive solutions, continuous dependence.

1. Introduction

The differential and integral equations with deviating arguments usually involve the deviation of the argument only the time itself, however, another case, in which the deviating arguments depend on both the state variable \(x\) and the time \(t\) is important in theory and practice. This kind of equations is called self-reference or state dependent equations. Differential equations with state-dependent delays attract interests of specialists since they widely arise from application models, such as two-body problem of classical electrodynamics also have may applications in the class of problems that have past memories, for example in hereditary phenomena see [1,2,3,4]. For papers studying such kind of equations, see for example [5,6,7,8,9,10,11,12,13,14,15,16] and references therein.

One of the first papers studying this class of equations was introduced by Buicá [7], the author proved the existence and the uniqueness of the solution of the initial value problem

\begin{eqnarray*} \frac{d x(t)}{dt} &=& f(t,x(x(t))), ~~~t\in[a,b],\\ x(0) &=& x_{0}, \end{eqnarray*} where \( f\in C([a,b]\times[a,b])\) and satisfied Lipshitz condition.

EL-Sayed and Ebead [13] relaxed the assumptions of Buicá and generalized their results, they studied the functional integral equation of the more general form

\begin{eqnarray*}\label{equ8555} x(t)~=~f\bigg(t,~\int_{0}^{t}g\big(s,x(x(s))\big)ds\bigg),~~~~t\in[0,T], \end{eqnarray*} where \( g\) satisfies Carathéodory condition.

El-Sayed and Ahmed [9] studied the existence of solutions and its continuous dependence of the initial value problem

\begin{eqnarray*} \frac{d}{dt} x(t)&=& f(t,x(\int_0^{\phi(t)} g(s,x(s)) ds)),~~a.e.~~t\in(0,T],\nonumber\\ x(0)&=& x_{0}, \end{eqnarray*} where \(g: [0,T]\times R^+\rightarrow R^+\) is continuous and \(g(t,x(t))\leq1\) and \(\phi(t) \leq t.\)

El-Sayed and Ebead [11,12] studied the existence of solution and its continuous dependence of the initial value problem of the delay-refereed differential equation

\begin{eqnarray*} \frac{d}{dt} x(t)&=& f(t,x(g(t,x(t)))),~~~a.e.~~t\in(0,T],\nonumber\\ x(0)&=& x_{0},\nonumber \end{eqnarray*} with the two cases of state-delay functions
  • (i) \(g: [0,T]\times R^+\rightarrow [0,T]\) is continuous and \(g(t,x(t))\leq t\),
  • (ii) \(g: [0,T]\times [0,T] \rightarrow [0,T]\) is continuous and \(g(t,x(\phi(t)))\leq x(\phi(t)).\)
Here we shall study the initial value problem of state-dependent neutral functional differential equation with two state-delay functions
\begin{equation}\label{0} \frac{d}{dt}~\big[x(t) -f_{1}\big(t,x(g_{1}(t,(x(\phi(t)))))\big)\big]= f_{2}\big(t,x(g_{2}(t,x(\phi(t))))\big),~~a.e.~t\in[0,T], \end{equation}
(1)
with the initial data
\begin{equation}\label{00} x(0)= f_{1}(0,x(g_{1}(0,x(0)))), \end{equation}
(2)
where \(g_{i},~i=1,~2\) are continuous and \(g_{1}(t,x)\leq \phi (t)\) and \(g_{1}(t,x)\leq x(t) .\)

Our aim in this work is to study the existence of at least one and exactly one positive solution of the Problem (1)-(2). The continuous dependence of the unique solution on the two functions \(g_{1}~\) and \(~g_{2} \) will be proved. To illustrate our results some examples will be given.

In order to achieve our goal, we study the existence of positive solutions \(x \in C[0; T] \) for the state-dependent functional integral equation

\begin{eqnarray}\label{1} x(t)&=&f_{1}\big(t,x(g_{1}(t,(x(\phi(t)))))\big)+\int_{0}^{t}~f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)ds ~~~t\in[0,T], \end{eqnarray}
(3)
and we will show later that this integral equation is equivalent to the initial value Problem (1)-(2).

2. Existence of solutions

Here we study the existence of solutions \(x\in C[0,T]\) for the integral Equation (3) under the following assumptions
  • (1) \(f_{1}:[0,T]\times[0,T] \rightarrow R^+\) is continuous and there exist two positive constants \(k_{1}~and~k_{2}\) such that \[ |f_{1}(t_{2},x)-f_{1}(t_{1},y)|\leq k_{1}|t_{2}-t_{1}|+k_{2}|x-y|. \]
  • (2) \(g_{1}:[0,T]\times [0,T] \rightarrow [0,T]\) is continuous and there exist the two positive constants \(k_{3},~k_{4}~\) such that \[ |g_{1}(t_{2},x)-g_{1}(t_{1},y)|\leq k_{3}|t_{2}-t_{1}|+k_{4}|x-y| \] and \(g_{1}(t,x)\leq \phi(t)\).
  • (3) \(f_{2}:[0,T]\times [0,T] \rightarrow R^+~\) satisfies Carathéodory condition i.e. \(f(t,x)\) is measurable in \(t\) for all \(x\in C [0,T]\) and continuous in \(x\) for almost all \(t\in[0,T]\).
  • (4) There exists a bounded measurable function\(~m : [0,T] \rightarrow R^{+},~m(t)\leq A,\) and a constant \(~k_{5}\geq 0\) such that \[ f_{2}(t,x)\leq m(t) +k_{5}~x. \]
  • (5) \( g_{2}:[0,T]\times [0,T]\rightarrow [0,T]\) is continuous and \(g_{2}(t,x)\leq x\)
  • (6) \( \phi:[0,T]\rightarrow [0,T],~\phi(0)=0\) and \[ |\phi(t)-\phi(s)|\leq |t-s|, \] which implies that \(\phi(t)\leq t\).
  • (7) There exists a real positive solution \(L\in (0,1)\) of the equation \[k_{2}~ k_{4} L^{2}+(k_{2}~k_{3}-1)L+ M+k_{1}=0,\] where \(M=A+k_{5}T\).
  • (8) \(LT+x(0)\leq T.\)
Now we are in a position to prove the following existence theorem.

Theorem 1. Let the assumptions \((1)-(8)\) be satisfied, then the state-dependent integral Equation (3) has at least one positive solution \(x\in C[0,T]\).

Proof. Let \(X=C[0,T]\) be the class of real valued continuous functions defined on \([0,T]\). Define the subset \(S_{L}\) of \(X \) by \[ S_{L}=\big\{x\in X:|x(t_{2})-x(t_{1})|\leq L|t_{2}-t_{1}|\big\}. \] It is clear that \(S_{L}\) is nonempty, closed, bounded and convex subset of \(C[0,T]\).

Now define the operator \(F\) associated with Equation (3) by

\begin{eqnarray*} Fx(t)=f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big)+\int_{0}^{t}~f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)ds,~~ t\in[0,T]. \end{eqnarray*} It is clear that F makes sense and well-defined.

Now, first we prove that the class of functions \(\{Fx\}\) is uniformly bounded on the set \(S_L.\) Let \(x \in X\), then for \(t \in [0, T],\) we obtain

\begin{eqnarray*} |Fx(t)|&=&|f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big)+\int_{0}^{t}~f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)ds|\\ &\leq &|f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big)|+\int_{0}^{t}~|f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)|ds. \end{eqnarray*} Using assumptions (1),(2) and (6) we can get
\begin{eqnarray}\label{3} | f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big)| &=& | f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big)- f_{1}\big(0,x(g_{1}(0,x(0)))\big)|+ |f_{1}\big(0,x(g_{1}(0,x(0)))\big)|\nonumber \\ &\leq& k_{1}~T+ k_{2}~|x(g_{1}(t,x(\phi(t))))-x(g_{1}(0,x(0)))|+ |f_{1}\big(0,x(g_{1}(0,x(0)))\big)|\nonumber \\ &\leq& k_{1}~T+ k_{2}~L |g_{1}(t,x(\phi(t)))-g_{1}(0,x(0))|+x(0)\nonumber \\ &\leq& k_{1}~T+ k_{2}~L ~(k_{3}~T+k_{4}|x(\phi(t))-x(0)|)+x(0)\nonumber \\ &\leq& k_{1}~T+ k_{2}~L ~(k_{3}~T+k_{4}~L~\phi(t))+x(0)\nonumber \\ &\leq& (k_{1}+ k_{2}~k_{3}~L +k_{2}k_{4}~L^{2})~T+x(0). \end{eqnarray}
(4)
Using assumptions (3) and (4) we can get
\begin{eqnarray}\label{5} | f_{2}\big(t,x(g_{2}(s,x(\phi(t))))\big)|&\leq &k_{5}~x(g_{2}(s,x(\phi(t))))+m(t)\nonumber \\ &\leq &k_{5}~\{|x(g_{2}(s,x(\phi(t))))-x(0)|+x(0)\}+m(t)\nonumber \\ & \leq &k_{5}(L~g_{2}(s,x(\phi(t)))+x(0))+A\nonumber \\ & \leq &k_{5}(L~x(\phi(t))+x(0))+A . \end{eqnarray}
(5)
Now from (4) and (5) and by assumption (8), we get \begin{eqnarray*} |Fx(t)| &\leq & (k_{1}+ k_{2}~k_{3}~L +k_{2}k_{4}~L^{2})~T+x(0)+\int_{0}^{t} (k_{5}(L~x(\phi(s))+x(0))+A)ds\\ &\leq & (k_{1}+ k_{2}~k_{3}~L +k_{2}k_{4}~L^{2})~T+x(0)+\int_{0}^{t} (k_{5}(L~T+x(0))+A)ds\\ &\leq &(k_{1}+ k_{2}~k_{3}~L +k_{2}k_{4}~L^{2})~T+x(0)+(k_{5}~T+ A)\int_{0}^{t}ds \end{eqnarray*} \begin{eqnarray*} &\leq & (k_{1}+ k_{2}~k_{3}~L +k_{2}k_{4}~L^{2}+M)~T+x(0)\\ &= & L~T+x(0)\leq T. \end{eqnarray*} This proves that the class of functions \(\{~Fx~\}\) is uniformly bounded on the set \(~S_L.~\)

Next, we prove that \(~F:S_{L}~\rightarrow S_{L} \) and the class of functions \(\{Fx\}\) is equi-continuous on the set \(S_L.\) Let \(x\in S_{L}\) and \(t_{1}, ~t_{2}\in [0,T]\) with \(t_{1}< t_{2} \) such that \( |t_{2}, -t_{1}|< \delta\), then

\begin{eqnarray*}\label{2} |Fx(t_2)-Fx(t_1)|&=&\bigg| f_{1}\big(t_{2},x(g_{1}(t_{2},x(\phi(t_{2}))))\big)~+\int_{0}^{t_{2}}~f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)ds \nonumber\\ &&-f_{1}\big(t_{1},x(g_{1}(t_{1},x(\phi(t_{1}))))\big)-\int_{0}^{t_{1}}~f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)ds \bigg|\nonumber\\ &\leq& |f_{1}\big(t_{2},x(g_{1}(t_{2},x(\phi(t_{2}))))\big)- f_{1}\big(t_{1},x(g_{1}(t_{1},x(\phi(t_{1}))))\big)|+\int_{t_{1}}^{ t_{2}}|f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)|ds \nonumber\\ &\leq& k_{1}|t_{2}-t_{1}|+k_{2}|x(g_{1}(t_{2},x(\phi(t_{2}))))-x(g_{1}(t_{1},x(\phi(t_{1}))))|+ \int_{t_{1}}^{t_{2}} (k_{5}~T+ A)ds\\ &\leq& k_{1}|t_{2}-t_{1}|+k_{2}|x(g_{1}(t_{2},x(\phi(t_{2}))))-x(g_{1}(t_{1},x(\phi(t_{1}))))| +M|t_{2}-t_{1}|. \end{eqnarray*} Using assumptions (2) and \(x\in S_{L}\), we can get \begin{eqnarray}\label{20} |x( g_{1}\big(t_{2},x(\phi(t_{2})))\big)- x(g_{1}\big(t_{1},x(\phi(t_{1}))))|& \leq & L | g_{1}\big(t_{2},x(\phi(t_{2})))\big)- g_{1}\big(t_{1},x(\phi(t_{1}))))|\nonumber\\& \leq & L~k_{3}|t_{2}-t_{1}|+ L~k_{4}|x(\phi(t_{2})))-x(\phi(t_{1})))|\nonumber\\ &\leq&L~ k_{3}|t_{2}-t_{1}|+L^{2} k_{4}|\phi(t_{2})-\phi(t_{1})|\nonumber \\ &\leq&L~ k_{3}|t_{2}-t_{1}|+ L^{2}k_{4}~|t_{2}-t_{1}|.\nonumber \end{eqnarray} Then, we have \begin{eqnarray*} |Fx(t_2)-Fx(t_1)|&\leq& k_{1}|t_{2}-t_{1}|+k_{2}(L~ k_{3}|t_{2}-t_{1}|+ L^{2}k_{4}~|t_{2}-t_{1}|)+M|t_{2}-t_{1}|\\ &=& \big(k_{2}~k_{4}~L^{2}+ k_{2}~k_{3}~L+k_{1}+M\big)|t_{2}-t_{1}|\\ &=& L|t_{2}-t_{1}|. \end{eqnarray*} Hence, we proved that \(~F:S_{L}~\rightarrow S_{L}~ \) and the class of functions \(\{Fx\}\) is equi-continuous on the set \(S_L.\) Applying Arzela-Ascoli Theorem [17], we deduce that \(F\) is compact operator. Now we show that \(F\) is continuous. Let \(\{x_{n}\} \subset S_{L},~x_{n}\rightarrow x \) on \([0,T]\), i.e. \(|x_{n}(\phi(t))-x(\phi(t))|\leq \epsilon_{1}\) this implies that \( | x_{n}(g_{i}(t,x(\phi(t))))-x(g_{i}(t,x(\phi(t)))) |\leq \epsilon_{2}\) for arbitrary \(\epsilon_{1},\epsilon_{2}\geq0\), \(i=1,~2,\) then \begin{eqnarray*} &~&|x_{n}(g_{i}(t,x_n(\phi(t))))-x(g_{i}(t,x(\phi(t)))) |\\ &&\leq |x_{n}(g_{i}(t,x_n(\phi(t))))-x_{n}(g_{i}(t,x(\phi(t))))|+ | x_{n}(g_{i}(t,x(\phi(t))))-x(g_{i}(t,x(\phi(t)))) |\\ &&\leq L|g_{i}(t,x_n(\phi(t)))-g_{i}(t,x(\phi(t)))|+| x_{n}(g_{i}(t,x(\phi(t))))-x(g_{i}(t,x(\phi(t)))) |\\ &&\leq \epsilon,~~i=1,~2. \end{eqnarray*} Then \begin{eqnarray*} x_{n}(g_{i}(t,x_n(\phi(t)))) \rightarrow x(g_{i}(t,x(\phi(t))))~in~ S_{L}, ~ ~i=1,~2 \end{eqnarray*} and by using the continuity of the functions \(f_{1} ,\) we obtain \begin{eqnarray*} f_{1}\big(t,x_{n}(g_{1}(t,x_{n}(\phi(t))))\big) &\rightarrow & f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big). \end{eqnarray*} Now by using the continuity of the functions \(f_{2},\) assumption \((5)\) and Lebesgues dominated convergence theorem [17], we obtain \begin{eqnarray*} \int_{0}^{t} f_{2}\big(t,x_{n}(g_{2}(t,x_{n}(\phi(t))))\big)ds &\rightarrow & \int_{0}^{t} f_{2}\big(t,x(g_{2}(t,x(\phi(t))))\big)ds, \end{eqnarray*} and \begin{eqnarray*} \lim_{n\rightarrow \infty}\big(F x_{n}\big)(t) &=& \lim_{n\rightarrow \infty} f_{1}\big(t,x_{n}(g_{1}(t,x_{n}(\phi(t))))\big)+\lim_{n\rightarrow \infty}\int_{0}^{t} f_{2}\big(t,x_{n}(g_{2}(t,x_{n}(\phi(t))))\big)ds\\& =& f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big)+\int_{0}^{t} f_{2}\big(t,x(g_{2}(t,x(\phi(t))))\big)ds\\ &=&\big(F x \big)(t). \end{eqnarray*} This proves that the operator \(F \) is continuous.

Now all conditions of Schauder fixed point theorem [17] are satisfied, then the operator \(F\) has at least one fixed point \(x\in S_{L}\). Consequently there exists at leat one solution \(x\in C[0,T]\) of Equation (3). This completes the proof.

Now, we introduce the following equivalence theorem.

Theorem 2. Let the assumptions (1)-(8) be satisfied, then the initial value Problem (1)-(2) has at least one positive solution \(x\in C[0,T]\).

Proof. Let \(x\) be a solution of the Problem (1)-(2). Integrate (1) and substitute by (2), we obtain the integral Equation (3). Let \(x\) be a solution of (3) differentiate (3) we obtain (1) and the initial value (2). This proves the equivalence between the Problem (1)-(2) and the integral Equation (3). Then the Problem (1)-(2) has at least one positive solution \( x \in C[0,T]\).

3. Applications

As application of our results, we introduce the following corollaries;

Corollary 1. Let the assumptions \((1)-(8)\) of Theorem 2 be satisfied, if

  • (i) \(g_1(t,x(\phi(t)))=\int_0^{\phi(t)} g_{3}(s,x(s))ds,~~~g_{3}: [0,1]\times [0,1]\rightarrow R^+\) is continuous and \(g_{3}(t,x(t))\leq 1 ,\)
  • (ii) \(g_{2}(t,x(t)) = x(t).\)
Then the initial value problem \begin{eqnarray*} &~&\frac{d}{dt}~\big[x(t)-f_{1}\big(t,x\big(\int_0^{\phi(t)}~g_3(s,x(s))ds)\big)\big] =f_{2}\big(t,x(x(\phi(t)))\big), ~~a.e.~~t\in(0,1], \\&~&x(0)= f_{1}(0,x(0)) \end{eqnarray*} has at least one positive solution \(x\in C[0,T]\).

Corollary 2. Let the assumptions of Theorem 1 be satisfied with \(g_{1}(t,x(\phi(t)))~= \phi(t)\) and \(g_{2}(t,x(\phi(t))) = x(\phi(t))\). Then the integral equation \begin{eqnarray*} x(t)=f_{1}(t,x(\phi(t)))+\int_{0}^{t}f_{2}(s,x(x(\phi(s))))ds,~~~t\in[0,T] \end{eqnarray*} has at least one solution \(x\in C[0,T].\) Consequently the initial value problem \begin{eqnarray*} \frac{d}{dt}~\big[x(t) -f_{1}(t,x(\phi(t)))\big]= f_{2}(t,x(x(\phi(t)))),~~~a.e.~~t\in[0,T] \end{eqnarray*} with the initial data \( x(0) = f_{1}(0,x(0))\) has at least one positive solution \( x \in C[0,T]\).

Corollary 3. Let the assumptions of Corollary 2 be satisfied with \(f_{1}(t,x)=t,\) then the initial value problem \begin{eqnarray*} \frac{d x(t)}{dt} &=& f(t,(x(x(\phi(t)))))~~a.e,~~t\in(0,T], \label{i1} \\ x(0)&=&0\label{i2} \end{eqnarray*} has at least one positive solution \( x \in C[0,T]\) where \(f(t,x)=f_{2}(t,x)+1\).

4. Uniqueness of the solution

In this section, we prove the uniqueness of the solution for the Problem (1)-(2). Therefore, we have to assume the following assumptions
  • (\(1^{\prime}\)) \(|f_{2}(t,x)-f_{2}(t,y)|\leq k_{5}~|x-y|\),
  • (\(2^{\prime}\)) \(~\sup_{t} |f_{2}(t,0)|\leq A,~~t\in [0,T]\),
  • (\(3^{\prime}\)) \(|g_{2}(t,x)-g_{2}(t,y)|\leq k_{6}|x-y|\).

Theorem 3. Let the assumptions (1)-(3), (5)-(8) and (1\(^{\prime}\))-(3\(^{\prime}\)) be satisfied, if \(k_{2}~(L~k_{4}+1)+k_{5}~T~(L~k_{6}+1) < 1,\) then the initial value Problem (1)-(2) has a unique positive solution \(x \in C[0,T].\)

Proof. Assumption (4) of Theorem 1 can be deduced from assumptions (1\(^{\prime}\)) and (2\(^{\prime}\)) as follows \begin{eqnarray}\label{w1} |f_{2}(t,x)|&\leq & k_{5}~|x|+|f_{2}(t,0)|\nonumber\\ &\leq & k_{5}~x+A,\nonumber \end{eqnarray} then we deduce that all assumptions of Theorem 1 are satisfied and the solution of Equation (3) exists. Now let \(x,~y\) be two solutions of (3), then \begin{eqnarray*} |x(t)-y(t)| &=&\big|f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big)+\int_{0}^{t}~f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)ds \\ &&- f_{1}\big(t,y(g_{1}(t,y(\phi(t))))\big)-\int_{0}^{t}~f_{2}\big(s,y(g_{2}(s,y(\phi(s))))\big)ds\big|\\ &\leq & |f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big)- f_{1}\big(t,y(g_{1}(t,y(\phi(t))))\big)|\\ &&+\int_{0}^{t}|f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)- f_{2}\big(s,y(g_{2}(s,y(\phi(s))))\big)|ds\\ &\leq & k_{2}~|x(g_{1}(t,x(\phi(t))))-y(g_{1}(t,y(\phi(t))))|\\ &&+ k_{5}\int_{0}^{t}|x(g_{2}(s,x(\phi(s))))-y(g_{2}(s,y(\phi(s))))|ds. \end{eqnarray*} But \begin{eqnarray}\label{q6} &~&|x(g_{1}(t,x(\phi(t))))-y(g_{1}(t,y(\phi(t)))) |\nonumber\\&&= |x(g_{1}(t,x(\phi(t))))-x(g_{1}(t,y(\phi(t))))\nonumber+x(g_{1}(t,y(\phi(t))))-y(g_{i}(t,y(\phi(t)))) |\nonumber \\ &&\leq |x(g_{1}(t,x(\phi(t))))-x(g_{1}(t,y(\phi(t))))|\nonumber+| x(g_{1}(t,y(\phi(t))))-y(g_{1}(t,y(\phi(t)))) |\nonumber\\ &&\leq L|g_{1}(t,x(\phi(t)))-g_{1}(t,y(\phi(t)))|\nonumber+| x(g_{1}(t,y(\phi(t))))-y(g_{1}(t,y(\phi(t)))) |\nonumber\\ &&\leq L ~k_{4}~\|x-y\|+\|x-y\|\nonumber\\ &&=(L ~k_{4}+1)\|x-y\|.\nonumber \end{eqnarray} Similarly, we can obtain \begin{eqnarray*}\label{q06} |x(g_{2}(t,x(\phi(t))))-y(g_{2}(t,y(\phi(t)))) |\leq (L ~k_{6}+1)\|x-y\|. \end{eqnarray*} Now \begin{eqnarray*} |x(t)-y(t)|&\leq & k_{2}~(L ~k_{4}+1)\|x-y\| +k_{5}~\int_{0}^{t}(L ~k_{6}+1)\|x-y\|ds, \end{eqnarray*} and \begin{eqnarray*} \|x-y\| &\leq&\big(k_{2}~(L~k_{4}+1)+k_{5}~T~(L~k_{6}+1)\big)\|x-y\|. \end{eqnarray*} Then we deduce that \[ \bigg(1- \big(k_{2}~(L~k_{4}+1)+k_{5}~T~(L~k_{6}+1)\big)\bigg)~|| x - y || \leq 0, \] and from the assumptions \(\big(k_{2}~(L~k_{4}+1)+k_{5}~T~(L~k_{6}+1)\big)< 1~~\) we can obtain \(x=y\) and the solution of (3) is unique. Consequently, the initial value Problem (1)-(2) has a unique positive solution \(x \in C[0,T].\)

5. Continuous dependence

5.1. Continuous dependence on the function \( g_{1}\)

Here we prove that the solution of the Problem (1)-(2) depends continuously on the function \(g_{1} \).

Definition 1. The solution of the Problem (1)-(2) depends continuously on the function \(g_{1},\) if \(\forall ~\epsilon > 0, ~\exists ~\delta(\epsilon)>0~\) such that \begin{eqnarray*}\label{cont} ||g_{1}-g_{1}^{*}||\leq \delta \Rightarrow \|x-x^{*}\|\leq \epsilon, \end{eqnarray*} where \(x^{*}\) is the unique solution of the equation

\begin{equation}\label{cont01} \frac{d}{dt}~\big[x^{*}(t) -f_{1}\big(t,x^{*}(g^{*}_{1}(t,(x^{*}(\phi(t)))))\big)\big]= f_{2}\big(t,x^{*}(g_{2}(t,x^{*}(\phi(t))))\big),~~~~~a.e.~~t\in[0,T], \end{equation}
(6)
with the initial data
\begin{equation}\label{cont05} x^*(0) = f_{1}\big(0,x^{*}(g^{*}_{1}(0,(x^{*}(0))))\big). \end{equation}
(7)

Theorem 4. Let the assumptions of Theorem 3 be satisfied, then the solution of initial value Problem (1)-(2) depends continuously on the function \(g_{1}. \)

Proof. Let \(x\) and \(x^*\) be the solution of the initial value Problems (1)-(2) and (6)-(7) respectively, then we have \begin{eqnarray*} |x(t)-x^{*}(t)| & =& \big|f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big)+\int_{0}^{t}~f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)ds \\ &&- f_{1}\big(t,x^{*}(g^{*}_{1}(t,x^{*}(\phi(t))))\big)+ \int_{0}^{t}~f_{2}\big(s,x^{*}(g_{2}(s,x^{*}(\phi(s))))\big)ds\big|\\ &\leq& |f_{1}\big(t,x(g_{1}(t,x(\phi(t))))\big) -f_{1}\big(t,x^{*}(g^{*}_{1}(t,x^{*}(\phi(t))))\big) | \\ &&+ \int_{0}^{t} |f_{2}\big(s,x(g_{2}(s,x(\phi(s))))\big)- f_{2}\big(s,x^{*}(g_{2}(s,x^{*}(\phi(s))))\big)|ds\\ &\leq& k_{2}~ |x(g_{1}(t,x(\phi(t))))\big) -x^{*}(g^{*}_{1}(t,x^{*}(\phi(t)))\big) | \\ &&+ k_{5}\int_{0}^{t} |x(g_{2}(s,x(\phi(s))))\big)- x^{*}(g_{2}(s,x^{*}(\phi(s)))\big)|ds\\ &\leq & k_{2}~ |x(g_{1}(t,x(\phi(t))))\big) -x^{*}(g_{1}(t,x^{*}(\phi(t)))\big) | \\ &&+ k_{2}~ |x^{*}(g_{1}(t,x^{*}(\phi(t))))\big) -x^{*}(g^{*}_{1}(t,x^{*}(\phi(t)))\big) | + k_{5}~T (L~k_{6}+1)\|x-x^{*}\|\\ &\leq& k_{2} (L~k_{4}+1)\|x-x^{*}\|+ k_{2}~ L~|g_{1}(t,x^{*}(\phi(t)))) -g_{1}^{*}(t,x^{*}(\phi(t))) | \\ &&+ k_{5}~T (L~k_{6}+1)\|x-x^{*}\|\\ &\leq& \big( k_{2}~ (L~k_{4}+1)+k_{5}~T (L~k_{6}+1)\big)\|x-x^{*}\|+k_{2}~L ~\delta. \end{eqnarray*} Then \[ \|x-x^{*}\|\leq \frac{k_{2}~L ~\delta}{\big(1-\big(k_{2}~ (L~k_{4}+1)+k_{5}~T (L~k_{6}+1)\big)}=\epsilon, \] and by the assumption \(\big(k_{2}~ (L~k_{4}+1)+k_{5}~T (L~k_{6}+1)\big)< 1,\) then the solution of (3) depends continuously on the functions \(g_{1}.\) Consequently the solution of the Problem (1)-(2) depends continuously on the functions \(g_{1} \) which complete the proof.

5.2. Continuous dependence on the function \( g_{2}\)

Here we prove that the solution of the Problem (1)-(2) depends continuously on the function \(g_{2} \).

Definition 2. The solution of the Problem (1)-(2) depends continuously on the function \(g_{2},\) if \(\forall ~\epsilon > 0, ~\exists ~\delta(\epsilon)>0~such~that,\) \begin{eqnarray*} ||g_{2}-g_{2}^{*}||\leq \delta \Rightarrow \|x-x^{*}\|\leq \epsilon, \end{eqnarray*} where \(x^{*}\) is the unique solution of the equation

\begin{equation}\label{cont1} \frac{d}{dt}~\big[x^{*}(t) -f_{1}\big(t,x^{*}(g_{1}(t,(x^{*}(\phi(t)))))\big)\big]= ~f_{2}\big(t,x^{*}(g^{*}_{2}(t,x^{*}(\phi(t))))\big),~~~~a.e.,~~t\in[0,T], \end{equation}
(8)
with the initial data
\begin{equation}\label{cont5} x^*(0) = f_{1}\big(0,x^{*}(g_{1}(0,(x^{*}(0))))\big). \end{equation}
(9)

Theorem 5. Let the assumptions of Theorem 3 be satisfied, then the solution of initial value Problem (1)-(2) depends continuously on the function \(g_{2}. \)

Proof. The proof follow similarly as the proof of Theorem 4.

6. Examples

Example 1. Consider the following problem

\begin{eqnarray}\label{ex2} \frac{d}{dt}~\bigg[x(t) - \frac{1}{18}(1+t^{2})-\frac{1}{8}x\bigg(\frac{\beta~t}{1+x^{2}(\beta~t)}\bigg)\bigg] = \frac{1}{7-t}+\frac{e^{-t}}{16}~x\bigg(\frac{x(\beta~t)~e^{-x^{2}(\beta~t)}}{1+\sin^{2}x(\beta~t)}\bigg) \end{eqnarray}
(10)
with the initial data
\begin{equation}\label{ex22} x(0) ~=~ \frac{4}{63}. \end{equation}
(11)
where \(t\in (0,1]\) and \(\beta \in (0,1]\). Here we have

\(\phi(t)=\beta~t\),

\(g_{1}\big(t,x(\phi(t))\big)=\frac{\beta~t}{1+x^{2}(\beta~t)},\)

\(g_{1}\big(t,x)\leq \beta~t,\)

\( f_{1}\big(t,x(g_{1}\big(t,x(\phi(t)))\big)\big) =\frac{1}{18}(1+t^{2})+\frac{1}{8}x\bigg(\frac{\beta~t}{1+x^{2}(\beta~t)}\bigg),\)

\( f_{1}\big(t,x)=\frac{1}{18}(1+t^{2})+\frac{1}{8}x,\)

\(g_{2}\big(t,x(\phi(t))\big)=\frac{x(\beta~t)~e^{-x^{2}(\beta~t)}}{1+\sin^{2}x(\beta~t)},\)

\(g_{2}\big(t,x)\leq x,\)

\( f_{2}\big(t,x(g_{2}\big(t,x(\phi(t)))\big)\big) =\frac{1}{7-t}+\frac{e^{-t}}{16}~x\bigg(\frac{x(\beta~t)~e^{-x^{2}(\beta~t)}}{1+\sin^{2}x(\beta~t)}\bigg), \)

\( f_{2}\big(t,x)=\frac{1}{7-t}+\frac{e^{-t}}{16}x,.\)

Thus we have \(k_{1}=\frac{1}{9},~k_{2}=\frac{1}{8},~k_{3}=2, ~k_{4}=2, ~k_{5}=\frac{1}{16},~A=\frac{1}{6},~ M=\frac{11}{48} ,~L\simeq 0.557< 1~ and~ L~T+x(0)\approx0.62< T=1.\)

Now all the assumptions of Theorem 1 are satisfied, then the Problem (10)-(11) has at least one solution \(x\in C[0,T]\).

Example 2. Consider the following problem

\begin{equation}\label{exv2} \frac{d}{dt}~\bigg[x(t) - \frac{1}{48}~(1+t)-\frac{1}{4}t^{2}~x\bigg(\frac{\frac{1}{4}~t^{2}}{1+32~x(t^{2})}\bigg)\bigg] = \frac{1}{5+2t} ~\sin^{2}(3(t+1)) +\frac{1}{12}~x\bigg(\frac{x(t^{2})~\sin^{2}(x(t^{2}))}{1+x^{2}(t^{2})}\bigg), \end{equation}
(12)
with the initial data
\begin{equation}\label{exv22} x(0) ~=~ \frac{1}{48}, \end{equation}
(13)
where \(t\in (0,\frac{1}{2}]\). Here we have

\(\phi(t)=t^{2}\),

\(g_{1}\big(t,x(\phi(t))\big)=\frac{ \frac{1}{4}~t^{2}}{1+32~x(t^{2})} ,\)

\(g_{1}\big(t,x)\leq t^{2},\)

\( f_{1}\big(t,x(g_{1}\big(t,x(\phi(t)))\big)\big) =\frac{1}{48}(1+t)+ \frac{1}{4}~t^{2}~x\bigg(\frac{\frac{1}{4}~t^{2}}{1+ 32~x(t^{2})}\bigg), \)

\( f_{1}\big(t,x)=\frac{1}{48}(1+t)+ \frac{1}{4}~t^{2}~x,\)

\(g_{2}\big(t,x(\phi(t))\big)=\frac{x(t^{2})~\sin^{2}(x(t^{2}))}{1+x^{2}(t^{2})} ,\)

\(g_{2}\big(t,x)\leq x,\)

\( f_{2}\big(t,x(g_{2}\big(t,x(\phi(t)))\big)\big) =\frac{1}{5+2t} \sin^{2}(3(t+1)) +\frac{1}{12}~x\bigg(\frac{x(t^{2})~\sin^{2}(x(t^{2}))}{1+x^{2}(t^{2})}\bigg), \)

\( f_{2}\big(t,x)=\frac{1}{5+2t} \sin^{2}(3(t+1)) +\frac{1}{12}~x.\)

Thus, we have \(k_{1}=\frac{7}{48},~k_{2}=\frac{1}{16},~k_{3}=\frac{17}{4}, ~k_{4}=2, ~k_{5}=\frac{1}{12},~A=\frac{1}{5},~ M=\frac{29}{120} ,~L\simeq 0.586< 1~ and~ L~T+x(0)\approx 0.3139< T=\frac{1}{2}.\)

Now all the assumptions of Theorem 1 are satisfied, then the Problem (12)-(13) has at least one solution \(x\in C[0,T]\).

7. Conclusion

Here we relaxed the assumptions and generalized the results in [8,11,14,18] and [1]. We proved the existence of at lease one solution of the Problem (1)-(2). The sufficient condition for the uniqueness of the solution have been given and the continuous dependence of the unique solution have been proved. Also some examples and applications have been given.

Author Contributions

All authors contributed equally to the writing of this paper. All authors read and approved the final manuscript.

Conflicts of interest

The authors declare no conflict of interest.

References

  1. Miranda, M., & Pascali, E. (2006). On a type of evolution of self-referred and hereditary phenomena. Aequationes Mathematicae, 71(3), 253-268. [Google Scholor]
  2. Tuan, N. M., & Nguyen, L. T. (2010). On solutions of a system of hereditary and self-referred partial-differential equations. Numerical Algorithms, 55(1), 101-113. [Google Scholor]
  3. Van Le, U., & Nguyen, L. T. (2008). Existence of solutions for systems of self-referred and hereditary differential equations. Electronic Journal of Differential Equations, 51, 1-7. [Google Scholor]
  4. Yang, D., & Zhang, W. (2004). Solutions of equivariance for iterative differential equations. Applied Mathematics Letters, 17(7), 759-765. [Google Scholor]
  5. Anh, P. K., Lan, N. T. T., & Tuan, N. M. (2012). Solutions to systems of partial differential equations with weighted self-reference and heredity. Electronic Journal of Differential Equations, 2012(117), 1-14. [Google Scholor]
  6. Berinde, V. (2010). Existence and approximation of solutions of some first order iterative differential equations. Miskolc Mathematical Notes, 11(1), 13-26. [Google Scholor]
  7. Buica, A. (1995). Existence and continuous dependence of solutions of some functional–differential equations. In Babes-Bolyai Univ., Cluj–Napoca, Seminar on fixed point theory, 3, 1-13. [Google Scholor]
  8. Eder, E. (1984). The functional differential equation \(x'(t) = x(x(t))\). Journal of Differential Equations, 54(3), 390-400. [Google Scholor]
  9. El-Sayed, A. M. A., & Ahmed, R. G. (2020). Solvability of the functional integro-differential equation with self-reference and state-dependence. Journal of Nonlinear Sciences & Applications, 13(1), 1-8.[Google Scholor]
  10. El-Sayeda, A. M. A., El-Owaidyb, H., & Ahmedb, R. G.(2020). Solvability of a boundary value problem of self-reference functional differential equation with infinite point and inte-gral conditions. Journal of Mathematics and Computer Science, 21, 296-308.[Google Scholor]
  11. EL-Sayed, A. M. A., & Ebead, H. R. (2020). On an initial value problem of delay-refereed differential Equation. International Journal of Mathematics Trends and Technology, 5(66), 32-37. [Google Scholor]
  12. El-Sayed, A., Hamdallah, E. M. & Ebead, H. R. (2020). Positive solutions of an initial value problem of a delay-self-reference nonlinear differential equation. Malaya Journal of Matematik, 8(3), 1001-1006. [Google Scholor]
  13. El-Sayed, A., & Ebead, H. R. (2020). On the solvability of a self-reference functional and quadratic functional integral equations. Filomat, 34(1), 129-141. [Google Scholor]
  14. Feckan, M. (1993). On a certain type of functional differential equations. Mathematica Slovaca, 43(1), 39-43. [Google Scholor]
  15. Gal, C. G. (2007). Nonlinear abstract differential equations with deviated argument. Journal of Mathematical Analysis and Applications, 333(2), 971-983. [Google Scholor]
  16. Letelier, J. C., Kuboyama, T., Yasuda, H., Cárdenas, M. L., & Cornish-Bowden, A. (2005). A self-referential equation, \(f(f) = f\), obtained by using the theory of \((m; r)\) systems: Overview and applications. Algebraic Biology, 115-126.[Google Scholor]
  17. Goebel, K., & Kirk, W. A. (1990). Topics in metric fixed point theory (Vol. 28). Cambridge University Press. [Google Scholor]
  18. Lan, N. T. & Pascali, E. (2018). A two-point boundary value problem for a differential equation with self-refrence, Electron Journal of Mathematical Analysis and Applications, 6(1), 25-30. [Google Scholor]
]]>
Exponential decay of solutions with \(L^{p}\) -norm for a class to semilinear wave equation with damping and source terms https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/exponential-decay-of-solutions-with-lp-norm-for-a-class-to-semilinear-wave-equation-with-damping-and-source-terms/ Wed, 02 Dec 2020 11:41:35 +0000 https://old.pisrt.org/?p=4747
OMA-Vol. 4 (2020), Issue 2, pp. 123 - 131 Open Access Full-Text PDF
Amar Ouaoua, Messaoud Maouni, Aya Khaldi
Abstract: In this paper, we consider an initial value problem related to a class of hyperbolic equation in a bounded domain is studied. We prove local existence and uniqueness of the solution by using the Faedo-Galerkin method and that the local solution is global in time. We also prove that the solutions with some conditions exponentially decay. The key tool in the proof is an idea of Haraux and Zuazua with is based on the construction of a suitable Lyapunov function.
]]>

Open Journal of Mathematical Analysis

Exponential decay of solutions with \(L^{p}\) -norm for a class to semilinear wave equation with damping and source terms

Amar Ouaoua\(^1\), Messaoud Maouni, Aya Khaldi
Laboratory of Applied Mathematics and History and Didactics of Mathematics (LAMAHIS) University of 20 August 1955, Skikda, Algeria.; (A.O & M.M & A,K)
\(^{1}\)Corresponding Author: ouaouaama21@gmail.com;

Abstract

In this paper, we consider an initial value problem related to a class of hyperbolic equation in a bounded domain is studied. We prove local existence and uniqueness of the solution by using the Faedo-Galerkin method and that the local solution is global in time. We also prove that the solutions with some conditions exponentially decay. The key tool in the proof is an idea of Haraux and Zuazua with is based on the construction of a suitable Lyapunov function.

Keywords:

Wave equation, source termes, Faedo-Galerkin method, global existence, exponential decay.

1. Introduction

Consider the following problem:

\begin{align} &u_{tt}-div\left( \frac{\mid \nabla u\mid ^{2m-2}\nabla u}{\sqrt{1+\mid \nabla u\mid ^{2m}}}\right) -\omega \Delta u_{t}+\mu u_{t}=u\mid u\mid ^{p-2},\ x\in \Omega ,\ t\geq 0, \label{1} \end{align}
(1)
\begin{align} &u\left( x,0\right) =u_{0}\left( x\right) ,\ u_{t}\left( x,0\right) =u_{1}\left( x\right) , \label{2} \end{align}
(2)
\begin{align} &u\left( x,t\right) =0,\ x\in \partial \Omega ,\ t\geq 0, \label{3} \end{align}
(3)
where \(\Omega \) is a bounded regular domain in \(\mathbb{R}^{n},\) \(n\geq 1\) with a smooth boundary \(\partial \Omega \). \(\omega \), \(\mu \) and \(m\), \(p\) are real numbers.

The nonlinear wave equations

\begin{align} &u_{tt}-\Delta u-\omega \Delta u_{t}+\mu u_{t}=u\mid u\mid ^{p-2},\ x\in \Omega ,\ t\geq 0, \label{4} \end{align}
(4)
\begin{align} &u\left( x,0\right) =u_{0}\left( x\right) ,\ u_{t}\left( x,0\right) =u_{1}\left( x\right) ,\ t\geq 0, \label{5} \end{align}
(5)
\begin{align} &u\left( x,t\right) =0,x\in \partial \Omega ,t\geq 0, \label{6} \end{align}
(6)
has been investigated by many authors [1,2,3,4,5,6,7,8,9,10]. In the absence of the nonlinear source term, it is well know that the presence of one damping term ensures global existence and decay of solutions for arbitrary initial condition [5,6]. For \(\omega =\mu =0\) the nolinear term \(u\mid u\mid ^{p-2}\) causes finite time blow up of solutions with negative energy [2]. The interaction between the damping and the source terms was first considered by Levine [11]. He showed that solutions with negative initial energy blows up in finite time. When \(\omega =0\) and the linear term \(u_{t}\) is replaced by \(\mid u_{t}\mid ^{r-2}u_{t}\), Georgiev and Todorowa [12] extended Levin's result to the case where \(r>2\). In their work, the authors introduced a method different from the one know as the concavity method. The termined suitable relations between \(r\) and \(p\), for whith there is global existence or alternatively fnite time blow-up.

For the initial boundary value problem of a quasilinear equation

\begin{equation*} u_{tt}-div\left( {\mid \nabla u\mid ^{m-2}\nabla u}\right) +au_{t}\mid u_{t}\mid ^{p-2}-\Delta u_{t}=bu\mid u\mid ^{r-2}, \end{equation*}
\begin{align} &x\in \Omega ,\ t\geq 0, \label{7} \end{align}
(7)
\begin{align} &u\left( x,0\right) =u_{0}\left( x\right) \in W_{0}^{1,m}\left( \Omega \right) ,\ u_{t}\left( x,0\right) =u_{1}\left( x\right) \in L^{2}\left( \Omega \right) ,\ t\geq 0, \label{8} \end{align}
(8)
\begin{align} &u\left( x,t\right) =0,\ x\in \partial \Omega ,\ t\geq 0. \label{9} \end{align}
(9)
Yang and Chen [13,14] studied the problem (7)-(9) and obtained global existence results under the growth assumptions on the nonlinear terms and the initial value. This global existence results have been improved by Liu and Zhao [15] by using a new method. In [13], the author considered a similar problem to (7)-(9) and proved a blow-up result under the condition \(p>max(r,m)\) and the energy is sufficiently negative. Messaoudi and Said-Houari [16] improved the results in [15] and showed that blow-up takes place for negative initial data only regardless of the size of \(\Omega\) Messaoudi in [17] showed that for \(m=2\), the decay is exponential. In absence of strong damping \(-\Delta u_{t}\) equation (7) becames
\begin{equation} \ u_{tt}-div\left( {\mid \nabla u\mid ^{m-2}\nabla u}\right) +au_{t}\mid u_{t}\mid ^{p-2}=bu\mid u\mid ^{r-2},\ x\in \Omega ,\ t\geq 0. \label{10} \end{equation}
(10)
For \(b=0\), it is well known that the damping term assures global existence and decay of the solution energy for arbitrary initial value [18]. For \(a=0\), the source term causes finite time blow-up of solutions with negative initial energy if \(r>m\) (see [2]). When the quasilinear operator \(-div\left( {\mid \nabla u\mid ^{m-2}\nabla u}\right) \) is replaced by \(\Delta ^{2}u\) , Wu and Tsai [19] showed that the solution is global in time under some conditions without the relation between \(p\) and \(r\). They also proved that the local solution blows up infinite time if \(r>p\) and the initial energy is nonnegative, and gave the decay estimates of the energy function and the lifespan of solutions. In this paper, we show that the local solutions of the problem (1)-(3) can be extented in infinite time to global solutions with the some conditions on initial data in the stable set for which the solutions decay expontially with \(L_{p}\) norm. The key tool in the proof is an idea of Haraux and Zuazua [6] and [9] with is based on the construction of a suitable Lyapunov function.

2. Assumptions and preliminaries

In this section, we present some material needed in the proof in our result.

Lemma 1. \(\left( Young^{\prime }s\ inequality\right) \) \(Let\ a,\ b\geq 0\) and \(\frac{1}{p}+\frac{1}{q}\)=1 for \( 1 < p, q < + \infty \), then one has the inequality \(ab \leq \delta a^{p} +C\left( \delta \right) b^{q}, \) where \(\delta >0 \) is an arbitrary constant, and \(C\left( \delta \right) \) is a positive constant depending on \(\delta \).

Lemma 2. Let s be a number with \(2\leq s < +\infty \) if \(\ n\leq r\) and \(2\leq s\leq \frac{nr}{n-r}\) if \(n>r\). Then there is a constant \(C\) depending on \(\Omega \) and s such that \(\left\Vert u\right\Vert _{s}\) \(\leq C\left\Vert \nabla u\right\Vert _{r}\), \(u\in W_{0}^{1,r}\left( \Omega \right) .\)

We denote the total energy related to the problem (1)-(3) by

\begin{equation} E\left( t\right) =\frac{1}{2}\left\Vert u_{t}\right\Vert _{2}^{2}+\frac{1}{m} \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}dx- \frac{1}{p}\left\Vert u\right\Vert _{p}^{p}. \label{11} \end{equation}
(11)
We also introduce the following functionals:
\begin{equation} I\left( t\right) =\underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}dx-\left\Vert u\right\Vert _{p}^{p}, \label{12} \end{equation}
(12)
\begin{equation} J\left( t\right) =\frac{1}{m}\underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}dx-\frac{1}{p}\left\Vert u\right\Vert _{p}^{p}. \label{13.} \end{equation}
(13)
As in [20], we can now define the so called " Nehari manifold" as follows: \begin{equation*} \mathcal{N}\text{=}\left\{ u\in W_{0}^{1,m}\left( \Omega \right) \backslash \left\{ 0\right\} ;\text{ }I\left( t\right) =0\right\} . \end{equation*} \(\mathcal{N}\) separates the two unbounded sets: \begin{equation*} \mathcal{N}^{+}\text{=}\left\{ u\in W_{0}^{1,m}\left( \Omega \right) ;\text{ }I\left( t\right) >0\right\} \cup \left\{ 0\right\} \end{equation*} and \begin{equation*} \mathcal{N}^{-}\text{=}\left\{ u\in W_{0}^{1,m}\left( \Omega \right) ;\text{ }I\left( t\right) < 0\right\} . \end{equation*} Assumptions:
  • \(\left( A1\right) :\) Assume that \(I\left( 0\right) >0,\) and \(0< E\left( 0\right) \) such that
    \begin{equation} B=c^{p}\left( \frac{mp}{p-m}E\left( 0\right) \right) ^{\frac{p-m}{m}}< 1.% \text{ } \label{14} \end{equation}
    (14)
    where \(c\) is the Poincaré constant.
  • \(\left( A2\right) :\) \(p\) satisfies
\[(2< m< p\leq\dfrac{nm}{n-m}, n\geq m; 2< m < p\leq+\infty, n < m.\] For simplicity, we define the weak solutions of (1)-(3) over the interval \(\left[ 0,T\right) \), but it is to be understood throughout that \(T\) is either infinity or the limit of the existence interval.

Definition 1. We say that \(u\left( x,t\right) \) is a weak solution of the problem (1)-(3) on the interval \(\Omega \times \left[ 0,T\right) ,\) if \(u\in L^{\infty }\left( \left[ 0,T\right) ;W_{0}^{1,m}\left( \Omega \right) \right) ,u_{t}\ \in L^{\infty }\left( % \left[ 0,\text{ }T\right) ;\text{ }L^{2}\left( \Omega \right) \right) \cap L^{2}\left( \left[ 0,\text{ }T\right) ;\ H_{0}^{1}\left( \Omega \right) \right) \) satisfy the following conditions:

  • \(\left( i\right) \)
    \begin{equation} \left( u^{\prime \prime }\left( t\right) ,\phi \right) +\left( \frac{\mid \nabla u\mid ^{2m-2}\nabla u}{\sqrt{1+\mid \nabla u\mid ^{2m}}},\nabla \phi \right) +\omega \left( \nabla u^{\prime },\nabla \phi \right) +\mu \left( u^{\prime },\phi \right) =\left( u\lvert u\rvert ^{p-2},\phi \right) , \label{15} \end{equation}
    (15)
    for any function \(\phi \in W_{0}^{1,m}\left( \Omega \right) \) and a.e. \(t\in \left[ 0,T\right)\).
  • \(\left( ii\right) \)
    \begin{equation} u\left( x,0\right) =u_{0}\left( x\right) \in L^{2}\left( \Omega \right) ,\ u_{t}\left( x,0\right) =u_{1}\left( x\right) \in L^{1}\left( \Omega \right) . \label{16} \end{equation}
    (16)

Theorem 1. (Local existence) Suppose that \(u_{0}\in L^{2}\left( \Omega \right) ,\ u_{1}\in L^{1}\left( \Omega \right) \) and \(E\left( 0\right) >0,\) then there exists \(T>0\) such that problem (1)-(3) has a unique solution u satisfying \(u\ \in L^{\infty}\left( [0, T ];\ W_{0}^{1,m}\left( \Omega\right)\right), \) \(u_{t}\ \in L^{\infty }\left( \left[ 0,\text{ }T\right) ;\text{ }L^{2}\left( \Omega \right) \right) \cap L^{2}\left( \left[ 0,\text{ }T\right) ;\ H_{0}^{1}\left( \Omega \right) \right) .\)

3. Global existence and exponential decay of solutions

In this section we are going to obtain the existence of local solutions to the problem (1)-(3) and exponential decay of solution. We will use the Faedo- Galerkin's method approximation.

Let \(\{w_{l}\}_{l=1}^{\infty }\) be a basis of \(W_{0}^{1,m}\left( \Omega \right) \) wich constructs a complete orthonormal system in \(L^{2}\left( \Omega \right) .\) Denote by \(V_{k}=span\{w_{1},w_{2},...,w_{k}\}\) the subspace generated by the first \(k\) vectors of the basis \( \{w_{l}\}_{l=1}^{\infty }.\) By the normalization, we have \(\lVert w_{l}\rVert =1.\) for any given integer k, we consider the approximation solution

\begin{equation*} u_{k}\left( t\right) =\sum_{l=1}^{k}u_{lk}\left( t\right) v_{l}, \end{equation*} where \(u_{k}\) is the solutions to the following Cauchy problem
\begin{equation} \left( u_{k}^{\prime \prime }\left( t\right) ,v_{l}\right) +\left( \frac{\mid \nabla u_{k}\mid ^{2m-2}\nabla u_{k}}{\sqrt{1+\mid \nabla u_{k}\mid ^{2m}}},\nabla v_{l}\right) +\omega \left( \nabla u_{k}^{\prime },\nabla v_{l}\right) +\mu \left( u_{k}^{\prime },v_{l}\right) =\left( u_{k}\lvert u_{k}\rvert ^{p-2},v_{l}\right) , \label{17} \end{equation}
(17)
where \(l=1,...,k,\) with initial conditions \(u_{k}\left( 0\right) =u_{0k}\) and \(u_{k}^{\prime }\left( 0\right) =u_{1k},\) \(u_{k}\left( 0\right) \) and \( u_{k}^{\prime }\left( 0\right) \) are chosen in \(V_{k}\) such that
\begin{equation} \sum_{l=1}^{k}\left( u_{0},v_{l}\right) v_{l}=u_{0k}\longrightarrow u_{0}\ in\ L^{2}\left( \Omega \right) ;\sum_{l=1}^{k}\left( u_{1},v_{l}\right) v_{l}=u_{1k}\longrightarrow u_{1}\ in\ L^{1}\left( \Omega \right) . \label{18} \end{equation}
(18)
Well known results on the solvability of nonlinear ODE provide the existence of a solution to problem (17)-(18) on interval \(\left[ 0,\tau\right) \) for some \(\tau >0\) and we can extend this solution to the whole interval \(\left[ 0,T\right] \) for any given \(T>0\) by making use of the a priori estimates below. Multiplying equation (17) by \(u_{lk}^{\prime }\left( t\right) \) and sum for \(l=1,...,k,\) we obtain
\begin{equation} \frac{d}{dt}\left( \frac{1}{2}\left\Vert u_{k}^{\prime }\right\Vert _{2}^{2}+ \frac{1}{m}\underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx-\frac{1}{p}\left\Vert u_{k}\right\Vert _{p}^{p}\right) =-\left( \omega \underset{\Omega }{\int }\lvert \nabla u_{k}\rvert _{2}^{2}dx+\mu \underset{\Omega }{\int }\lvert u_{k}^{\prime }\rvert _{2}^{2}dx\right). \label{19} \end{equation}
(19)
Integrating (19) over \(\left( 0,t\right)\), we obtain the estimate
\begin{equation} \frac{1}{2}\left\Vert u_{k}^{\prime }\right\Vert _{2}^{2}+\frac{1}{m} \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx- \frac{1}{p}\left\Vert u_{k}\right\Vert _{p}^{p} +\omega \int_{0}^{t}\underset{\Omega }{\int }\lvert \nabla u_{k}\rvert _{2}^{2}dx+\mu \int_{0}^{t}\underset{\Omega }{\int }\lvert u_{k}^{\prime }\rvert _{2}^{2}dx\leq E\left( 0\right) . \label{20} \end{equation}
(20)
Since \(I\left( 0\right) >0,\) then there exists \(\tau < T\) by continuity such that \(I\left( t\right) \geq 0,\). We get from (12) and (13) that
\begin{equation} J\left( u_{k}\left( t\right) \right) =\frac{p-m}{mp}\underset{\Omega }{\int }% \sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx+\frac{1}{p}I\left( u_{k}\left( t\right) \right) \label{21} \end{equation}
(21)
\begin{equation} J\left( u_{k}\left( t\right) \right) \geq \frac{p-m}{mp}\underset{\Omega }{% \int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx, \forall t\in \left[ 0,\text{ }\tau \right] . \label{22} \end{equation}
(22)
Hence we have
\begin{equation} \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}} dx\leq \frac{mp}{p-m}J\left( u_{k}\left( t\right) \right) . \label{23} \end{equation}
(23)
From (11) and (13) we obvioulsy have \(\forall t\in \left[ 0,\text{ }\tau \right] ,\) \(J\left( u_{k}\left( t\right) \right) \leq E\left( u_{k}\left( t\right) \right) .\) Thus we obtain
\begin{equation} \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}} dx\leq \frac{mp}{p-m}E\left( u_{k}\left( t\right) \right) . \label{24} \end{equation}
(24)
Since \(E\) is a decreasing function of \(t,\) we have
\begin{equation} \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}} dx\leq \frac{mp}{p-m}E\left( 0\right) , \forall t\in \left[ 0,\text{ }\tau \right] \label{25} \end{equation}
(25)
By using Lemma 2, we easily have \begin{eqnarray*} \left\Vert u_{k}\right\Vert _{p}^{p} &\leq &c^{p}\left\Vert \nabla u_{k}\right\Vert _{m}^{p}=c^{p}\left( \underset{\Omega }{\int }\left\vert \nabla u_{k}\right\vert ^{m}dx\right) ^{\frac{p}{m}}\leq c^{p}\left( \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}} dx\right) ^{\frac{p}{m}} \\ &\leq &c^{p}\left( \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx\right) ^{\frac{p-m}{m}}\underset{\Omega }{\int } \sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx \end{eqnarray*} Using the inequality (25), we deduce \begin{equation*} \left\Vert u_{k}\right\Vert _{p}^{p}\leq c^{p}\left( \frac{mp}{p-m}E\left( 0\right) \right) ^{\frac{p-m}{m}}\underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx. \end{equation*} Now exploiting the inequality (14), we obtain
\begin{equation} \left\Vert u_{k}\right\Vert _{p}^{p}\leq \underset{\Omega }{\int }\sqrt{ 1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx. \label{26} \end{equation}
(26)
Hence \(\underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx-\) \(\left\Vert u_{k}\right\Vert _{p}^{p}>0,\) \(\forall t\in \left[ 0, \text{ }\tau \right] ,\) this shows that \(I\left( u_{k}\left( t\right) \right) >0,\) by repeating this procedure, \(\tau \) is extended to T.

Since \(\underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u_{k}\right\vert ^{2m}}dx>\lVert \nabla u_{k}\rVert _{m}^{m},\) it follows from (20) and (26) that

\begin{equation} \frac{1}{2}\left\Vert u_{k}^{\prime }\right\Vert _{2}^{2}+\frac{p-m}{pm}% \lVert \nabla u_{k}\rVert _{m}^{m}+ +\omega \int_{0}^{t}\underset{\Omega }{\int }\lvert \nabla u_{k}\rvert _{2}^{2}dx+\mu \int_{0}^{t}\underset{\Omega }{\int }\lvert u_{k}^{\prime }\rvert _{2}^{2}dx\leq E\left( 0\right) . \label{27} \end{equation}
(27)
From (27), we have
\begin{equation} \left\{ \begin{array}{c} \left\{ u_{k}\right\} \mathit{\ }\text{is uniformly bounded in }L^{\infty }\left( \left[ 0,T\right] ;W_{0}^{1,m}\left( \Omega \right) \right) , \\ \left\{ u_{k}\right\} \rightharpoonup u\mathit{\ }\text{is uniformly bounded in }L^{2}\left( \left[ 0,T\right] ;H_{0}^{1}\left( \Omega \right) \right) , \\ \left\{ u_{k}^{\prime }\right\} \mathit{\ }\text{is uniformly bounded in }L^{\infty }\left( \left[ 0,T\right] ;L^{2}\left( \Omega \right) \right) , \\ \left\{ u_{k}^{\prime }\right\} \mathit{\ }\text{is uniformly bounded in }L^{2}\left( \left[ 0,T\right] ;L^{2}\left( \Omega \right) \right). \end{array} \right. \label{28} \end{equation}
(28)
Furthermore, we have from Lemma 2 and (28) that
\begin{equation} \left\{ \lvert u_{k}\rvert ^{p}u_{k}\right\} \mathit{\ }\text{ is uniformly bounded in }L^{\infty }\left( \left[ 0,T\right] ;L^{2}\left( \Omega \right) \right) . \label{29} \end{equation}
(29)
By (28) and (29), we infer that there exists a subsequence of \(u_{k}\) (denote still by the same symbol) and a function \(u\) such that
\begin{equation} \left\{ \begin{array}{c} u_{k}\rightharpoonup u\mathit{\ }\text{weakly star in } L^{\infty }\left( \left[ 0,T\right] ;W_{0}^{1,m}\left( \Omega \right) \right) , \\ u_{k}\rightharpoonup u\mathit{\ }\text{weakly star in } L^{2}\left( \left[ 0,T\right] ;H_{0}^{1}\left( \Omega \right) \right) , \\ u_{k}^{\prime }\rightharpoonup u^{\prime }\mathit{\ }\text{ weakly star in }L^{\infty }\left( \left[ 0,T\right] ;L^{2}\left( \Omega \right) \right) , \\ u_{k}^{\prime }\rightharpoonup u^{\prime }\mathit{\ }\text{ weakly star in }L^{2}\left( \left[ 0,T\right] ;L^{2}\left( \Omega \right) \right) , \\ \lvert u_{k}\rvert ^{p-2}u_{k}\rightharpoonup \mathcal{X}\mathit{\ }\text{ weakly star in }L^{\infty }\left( \left[ 0,T\right] ;L^{2}\left( \Omega \right) \right) . \end{array} \right. \label{30} \end{equation}
(30)
By the Aubin-Lions compactness Lemma [7], we conclude from (30) that \begin{equation*} \left\{ \begin{array}{c} u_{k}\rightharpoonup u\ \ \text{strongly in}\ \ C\left( \left[ 0,T\right] ;L^{2}\left( \Omega \right) \right) , \\ u_{k}^{\prime }\rightharpoonup u^{\prime }\mathit{\ }\text{ strongly in }C\left( \left[ 0,T\right] ;L^{2}\left( \Omega \right) \right) , \end{array} \right. \end{equation*} and
\begin{equation} u_{k}\rightharpoonup u\mathit{\ }\text{almost everywhere in }\left[ 0,T\right] \times \Omega . \label{31} \end{equation}
(31)
It follows from Lemma 1.3 in [21] and (31)
\begin{equation} \lvert u_{k}\rvert ^{p-2}u_{k}\rightharpoonup \lvert u\rvert ^{p-2}u\mathit{ \ }\text{weakly star in }L^{\infty }\left( \left[ 0,T \right] ;L^{2}\left( \Omega \right) \right) . \label{32} \end{equation}
(32)
By the last formula (32) and (30), we obtain \(\mathcal{X}=\lvert u\rvert ^{p-2}u\) On the other hand, taking \(\phi =1\), (17) become
\begin{equation} \left( u_{k}^{\prime \prime }\left( t\right) ,1\right) +\mu \left( u_{k}^{\prime },1\right) =\left( u_{k}\lvert u_{k}\rvert ^{p-2},1\right). \label{33} \end{equation}
(33)
We have \begin{equation*} \lvert \left( u_{k}^{\prime \prime }\left( t\right) ,1\right) +\mu \left( u_{k}^{\prime },1\right) \rvert \geq \lVert u_{k}^{\prime \prime }\rVert -\mu \lVert u_{k}^{\prime }\rVert. \end{equation*} Since, the measure of \(\Omega \) is finite, by the embedding theorem, (30) and (33), we obtain \begin{equation*} \lVert u_{k}^{\prime \prime }\rVert \leq C, \end{equation*} then \begin{equation*} \left\{ u_{k}^{\prime \prime }\right\} \mathit{\ }\text{is uniformly bounded in }L^{\infty }\left( \left[ 0,T\right] ;L^{1}\left( \Omega \right) \right) . \end{equation*} Similarly, we have
\begin{equation} u_{k}^{\prime \prime }\rightharpoonup u^{\prime \prime }\mathit{\ }\text{ weakly star in }L^{\infty }\left( \left[ 0,T\right] ;L^{1}\left( \Omega \right) \right) , \label{34} \end{equation}
(34)
Setting up \(k\longrightarrow \infty \) and passing to the limit in (17), we obtain \begin{equation*} \left( u^{\prime \prime }\left( t\right) ,v_{l}\right) +\left( \frac{\mid \nabla u\mid ^{2m-2}\nabla u}{\sqrt{1+\mid \nabla u_{k}\mid ^{2m}}},\nabla v_{l}\right) +\omega \left( \nabla u^{\prime },\nabla v_{l}\right) +\mu \left( u^{\prime },v_{l}\right) =\left( u\lvert u\rvert ^{p-2},v_{l}\right) , \end{equation*} \(l=1,...,k.\) Since \(\{v_{l}\}_{l=1}^{\infty }\) is a base of \(W_{0}^{1,m}\left( \Omega \right)\), we deduce that \(u\) satisfies (1).

From (30), (34) and Lemma 3.1.7 in [22], with \(B=L^{2}\left( \Omega \right) \) and \(B=L^{1}\left( \Omega \right) ,\) respectively, we infer that

\begin{equation} \left\{ \begin{array}{c} u_{k}\left( 0\right) \rightharpoonup u\left( 0\right) \ \text{ weakly in} \text{ }\ L^{2}\left( \Omega \right) , \\ u_{k}^{^{\prime }}\left( 0\right) \rightharpoonup u^{^{\prime }}\left( 0\right) \ \text{weakly star in}\ L^{1}\left( \Omega \right) . \end{array}% \right. \label{35} \end{equation}
(35)
We get from (18) and (35) that \(u\left( 0\right) =u_{0},\) \(u^{\prime }\left( 0\right) =u_{1}.\) Thus, the proof is complete.

Lemma 3. Assume that \(p>m\) and \( u_{0}\in \mathcal{N}^{+},\) \( u_{1}\in L^{2}\left( \Omega \right) .\) If \(0< E\left( 0\right) \) and satisfy (14) then the local solution of the problem (1)-(3) is global in time.

Proof. Since the map \(t\longmapsto E\left( t\right) \) is a decreasing of the time \(t,\) we have

\begin{equation} E\left( 0\right) \geq E\left( t\right) =\frac{1}{2}\left\Vert u_{t}\right\Vert _{2}^{2}+\frac{p-m}{mp}\underset{\Omega }{\int }\sqrt{% 1+\left\vert \nabla u\right\vert ^{2m}}dx+\frac{1}{p}I\left( t\right) \label{36} \end{equation}
(36)
which give
\begin{equation} E\left( 0\right) \geq E\left( t\right) \geq \frac{1}{2}\left\Vert u_{t}\right\Vert _{2}^{2}+\frac{p-m}{mp}\underset{\Omega }{\int }\sqrt{% 1+\left\vert \nabla u\right\vert ^{2m}}dx \label{37} \end{equation}
(37)
thus, \(\forall t\in \left[ 0,\text{ }T\right) ,\) \(\left\Vert u_{t}\right\Vert _{2}^{2}+\underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}dx\) is uniformly bounded by a constant depending only on \(E\left( 0\right) ,\) \(p\) and \(m\) then the solution is global, so \(T_{\max }=\infty .\)

Theorem 2. Assume that \(p>m\). Let \(u_{0}\in \mathcal{N}^{+}\) and \(u_{1}\in L^{2}\left( \Omega \right) .\) Moreover, assume that \(0< E\left( 0\right) \) and satisfy (14). Then there exists two positive constants \( \alpha \) and \(\beta \) independent of \(t\) such that: \(0< E\left( t\right) \leq \beta e^{-\alpha t},\) \(\forall t>0.\)

Proof. Since we have proved that \(t\geq 0,\) \(u\left( t\right) \in \mathcal{N}^{+},\) we already have \begin{equation*} 0< E\left( t\right) ,\forall t\geq 0. \end{equation*} We define a Lyaponov function, for \(\epsilon >0.\)

\begin{equation} L\left( t\right) =E\left( t\right) +\epsilon \underset{\Omega }{\int } u_{t}udx+\frac{\epsilon \omega }{2}\left\Vert \nabla u\right\Vert _{2}^{2}. \label{38} \end{equation}
(38)
We prove that \(L\left( t\right) \) and \(E\left( t\right) \) are equivalent in the sens that there exist two constants \(B_{1}\) and \(B_{2}\) depending on \( \epsilon\) such that for \(t\geq 0\)
\begin{equation} B_{1}E\left( t\right) \leq L\left( t\right) \leq B_{2}E\left( t\right) . \label{39} \end{equation}
(39)
By the Lemma 1, we have \begin{equation*} L\left( t\right) =E\left( t\right) +\epsilon \underset{\Omega }{\int } u_{t}udx+\frac{\epsilon \omega }{2}\left\Vert \nabla u\right\Vert _{2}^{2} \leq E\left( t\right) +\epsilon \left( \frac{1}{4\delta }\left\Vert u_{t}\right\Vert _{2}^{2}+\delta \left\Vert u\right\Vert _{2}^{2}\right) + \frac{\epsilon \omega }{2}\left\Vert \nabla u\right\Vert _{2}^{2}. \end{equation*} Thanks of the Poincaré inequality and since \(\delta\) is an arbitrary constant, we choose \(\delta \) small suffisant for that,
\begin{equation} \delta \left\Vert u\right\Vert _{2}^{2}\leq \delta C\left\Vert \nabla u\right\Vert _{2}^{2}\leq \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}dx \label{40} \end{equation}
(40)
Then, we get \begin{equation*} L\left( t\right) \leq E\left( t\right) +\epsilon \frac{1}{4\delta } \left\Vert u_{t}\right\Vert _{2}^{2}+\epsilon \left( \delta C+\frac{\omega }{ 2}\right) \left\Vert \nabla u\right\Vert _{2}^{2} \leq E\left( t\right) +\epsilon \frac{1}{4\delta }\left\Vert u_{t}\right\Vert _{2}^{2}+\epsilon \left( 1+\frac{\omega }{2}\right) \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}dx. \end{equation*} By (37), we get
\begin{equation} L\left( t\right) \leq E\left( t\right) +\epsilon \frac{1}{2\delta }E\left( t\right) +\epsilon \left( 1+\frac{\omega }{2}\right) \frac{mp}{p-m}E\left( t\right) \leq B_{2}E\left( t\right) , \label{41} \end{equation}
(41)
where \(B_{2}=\left( 1+\epsilon \frac{1}{2\delta }+\epsilon \left( 1+\frac{\omega }{2}\right) \frac{mp}{p-m}\right)\).

On the other hand, we have

\begin{align*} L\left( t\right) &\geq E\left( t\right) -\epsilon \left( \frac{1}{4\delta } \left\Vert u_{t}\right\Vert _{2}^{2}+\delta \left\Vert u\right\Vert _{2}^{2}\right) +\frac{\epsilon \omega }{2}\left\Vert \nabla u\right\Vert _{2}^{2}\\ &\geq E\left( t\right) -\epsilon \frac{1}{4\delta }\left\Vert u_{t}\right\Vert _{2}^{2}-\epsilon \delta \left\Vert u\right\Vert _{2}^{2}\\ &\geq E\left( t\right) -\epsilon \frac{1}{2\delta }E\left( t\right) -\epsilon \delta \left\Vert u\right\Vert _{2}^{2}\\ &\geq \left( 1-\epsilon \frac{1}{2\delta }\right) E\left( t\right) -\epsilon \delta \left\Vert u\right\Vert _{2}^{2}. \end{align*} From (37) and (40), we obtain
\begin{equation} L\left( t\right) \geq \left( 1-\epsilon \frac{1}{2\delta }-\epsilon \frac{mp% }{p-m}\right) E\left( t\right) =B_{1}E\left( t\right) , \label{42} \end{equation}
(42)
where \(B_{1}=\left( 1-\epsilon \frac{1}{2\delta }-\epsilon \frac{mp}{p-m}% \right)\).

Now, we have

\begin{align*} \frac{d}{dt}L\left( t\right) &=-\omega \left\Vert \nabla u_{t}\right\Vert _{2}^{2}-\mu \left\Vert u_{t}\right\Vert _{2}^{2}+\epsilon \left\Vert u_{t}\right\Vert _{2}^{2} +\epsilon \underset{\Omega }{\int }{div}\left( \frac{\left\vert \nabla u\right\vert ^{2m-2}\nabla u}{\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}}% \right) udx+\epsilon \left\Vert u\right\Vert _{p}^{p}-\epsilon \mu \underset{% \Omega }{\int }u_{t}udx\\ &=-\omega \left\Vert \nabla u_{t}\right\Vert _{2}^{2}-\mu \left\Vert u_{t}\right\Vert _{2}^{2}+\epsilon \left\Vert u_{t}\right\Vert _{2}^{2}-\epsilon \underset{\Omega }{\int }\frac{\left\vert \nabla u\right\vert ^{2m}}{\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}}dx +\epsilon \left\Vert u\right\Vert _{p}^{p}-\epsilon \mu \underset{\Omega }{% \int }u_{t}udx. \end{align*} So that
\begin{equation} \frac{d}{dt}L\left( t\right) \leq -\omega \left\Vert \nabla u_{t}\right\Vert _{2}^{2}+\left( \epsilon \left( \frac{\mu }{4\delta }+1\right) -\mu \right) \left\Vert u_{t}\right\Vert _{2}^{2}+\epsilon \mu \delta \left\Vert u\right\Vert _{2}^{2} -\epsilon \underset{\Omega }{\int }\frac{\left\vert \nabla u\right\vert ^{2m} }{\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}}dx+\epsilon \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}dx. \label{45} \end{equation}
(43)
So
\begin{equation} \frac{d}{dt}L\left( t\right) \leq \left( \epsilon \left( \frac{\mu }{4\delta }+1\right) -\mu \right) \left\Vert u_{t}\right\Vert _{2}^{2}+\epsilon \left( 1+\mu \right) \underset{\Omega }{\int }\sqrt{1+\left\vert \nabla u\right\vert ^{2m}}dx. \label{46} \end{equation}
(44)
Using the inequality (37) and (44), we deduce \begin{eqnarray*} \frac{d}{dt}L\left( t\right) &\leq &2\left( \epsilon \left( \frac{\mu }{ 4\delta }+1\right) -\mu \right) E\left( t\right) +\epsilon \left( 1+\mu \right) \frac{mp}{p-m}E\left( t\right) \\ &\leq &-\left( 2\mu -\epsilon \left( \left( \frac{\mu }{2\delta }+2\right) +\left( 1+\mu \right) \frac{mp}{p-m}\right) \right) E\left( t\right). \end{eqnarray*} We choosing \(\epsilon \) small enough such that
\begin{equation} -\left( 2\mu -\epsilon \left( \left( \frac{\mu }{2\delta }+2\right) +\left( 1+\mu \right) \frac{mp}{p-m}\right) \right) =\zeta < 0. \label{47} \end{equation}
(45)
So
\begin{equation} \frac{d}{dt}L\left( t\right) \leq \zeta E\left( t\right). \label{48} \end{equation}
(46)
From (39), we have
\begin{equation} \frac{d}{dt}L\left( t\right) \leq \frac{\zeta }{B_{2}}L\left( t\right). \label{49} \end{equation}
(47)
Integrating the provious differential inequality (47) between \(0\) and \(t\) gives the following estimate for the function \(L:\)
\begin{equation} L\left( t\right) \leq ce^{\frac{\zeta }{B_{2}}t}, \forall t\geq 0. \label{50} \end{equation}
(48)
Consequently, by using (39) once again, we conclude
\begin{equation} E\left( t\right) \leq ke^{\frac{\zeta }{B_{2}}t}, \forall t\geq 0. \label{51} \end{equation}
(49)
By using (26) and (37) we easily have
\begin{equation} \left\Vert u\right\Vert _{p}^{p}\leq k_{1}e^{\frac{\zeta }{B_{2}}t}, \forall t\geq 0 . \label{52} \end{equation}
(50)
The proof is complete.

4. Conclusion

In this paper, we have studied a class of hyperbolic equation supplemented with Dirichlet boundary conditions as a model of wave equation with damping and source nonlinear terms. We showed that the solution with positive initial energy exponentially decay, this is mainly due to the presence of one of term of weak or strong damping.

Acknowledgments:

The authors wish to thank deeply the anonymous referee for useful remarks and careful reading of the proofs presented in this paper.

Author Contributions:

All authors contributed equally to the writing of this paper. All authors read and approved the final manuscript.

Conflict of Interests:

The authors declare no conflict of interest.

References

  1. Haraux, A., & Zuazua, E. (1988). Decay estimates for some semilinear damped hyperbolic problems. Archive for Rational Mechanics and Analysis, 100(2), 191-206. [Google Scholor]
  2. Zuazua, E. (1990). Exponential decay for the semilinear wave equation with locally distributed damping. Communications in Partial Differential Equations, 15(2), 205-235.[Google Scholor]
  3. Ball, J. M. (1977). Remarks on blow-up and nonexistence theorems for nonlinear evolution equations. The Quarterly Journal of Mathematics, 28(4), 473-486. [Google Scholor]
  4. Ono, K. (1997). On global existence, asymptotic stability and blowing up of solutions for some degenerate non-linear wave equations of Kirchhoff type with a strong dissipation. Mathematical Methods in the Applied Sciences, 20(2), 151-177. [Google Scholor]
  5. Aassila, M., Cavalcanti, M. M., & Soriano, J. A. (2000). Asymptotic stability and energy decay rates for solutions of the wave equation with memory in a star-shaped domain. SIAM Journal on Control and Optimization, 38(5), 1581-1602. [Google Scholor]
  6. Payne, L. E., & Sattinger, D. H. (1975). Saddle points and instability of nonlinear hyperbolic equations. Israel Journal of Mathematics, 22(3-4), 273-303. [Google Scholor]
  7. Berrimi, S., & Messaoudi, S. A. (2004). Exponential decay of solutions to a viscoelastic equation with nonlinear localized damping. Electronic Journal of Differential Equations, 88, 1-10. [Google Scholor]
  8. Messaoudi, S. A., & Houari, B. S. (2004). Global non-existence of solutions of a class of wave equations with non-linear damping and source terms. Mathematical methods in the applied sciences, 27(14), 1687-1696. [Google Scholor]
  9. Zuazua, E. (1990). Exponential decay for the semilinear wave equation with locally distributed damping. Communications in Partial Differential Equations, 15(2), 205-235. [Google Scholor]
  10. Georgiev, V., & Todorova, G. (1994). Existence of a solution of the wave equation with nonlinear damping and source terms. Journal of Differential Equations, 109(2), 295-308. [Google Scholor]
  11. Ikehata, R., & Suzuki, T. (1996). Stable and unstable sets for evolution equations of parabolic and hyperbolic type. Hiroshima Mathematical Journal, 26(3), 475-491. [Google Scholor]
  12. Lions, J. L. (1969). Quelques méthodes de résolution des problèmes aux limites nonlinéaires, Zbl 0189.40603. [Google Scholor]
  13. Yacheng, L., & Junsheng, Z. (2004). Multidimensional viscoelasticity equations with nonlinear damping and source terms. Nonlinear Analysis: Theory, Methods & Applications, 56(6), 851-865. [Google Scholor]
  14. Zhijian, Y. (2002). Existence and asymptotic behaviour of solutions for a class of quasi-linear evolution equations with non-linear damping and source terms. Mathematical Methods in the Applied Sciences, 25(10), 795-814.[Google Scholor]
  15. Gerbi, S., & Said-Houari, B. (2012). Exponential decay for solutions to semilinear damped wave equation. Discrete and Continuous Dynamical Systems-Series S, 5(3), 559-566. [Google Scholor]
  16. Messaoudi, S. A. (2005). On the decay of solutions for a class of quasilinear hyperbolic equations with non-linear damping and source terms. Mathematical methods in the applied sciences, 28(15), 1819-1828. [Google Scholor]
  17. Kawashima, S., & Shibata, Y. (1992). Global existence and exponential stability of small solutions to nonlinear viscoelasticity. Communications in Mathematical Physics, 148(1), 189-208.[Google Scholor]
  18. Cavalcanti, M. M., Domingos Cavalcanti, V. N., & Soriano, J. A. (2002). Exponential decay for the solution of semilinear viscoelastic wave equations with localized damping. Electronic Journal of Differential Equations, 44, 1-14.[Google Scholor]
  19. Strauss, W. (1966). On continuity of functions with values in various Banach spaces. Pacific journal of Mathematics, 19(3), 543-551. [Google Scholor]
  20. Wu, S. T., & Tsai, L. Y. (2006). On global existence and blow-up of solutions for an integro-differential equation with strong damping. Taiwanese Journal of Mathematics, 10(4), 979-1014. [Google Scholor]
  21. Wu, S. T., & Tsai, L. Y. (2009). On global solutions and blow-up of solutions for a nonlinearly damped Petrovsky system. Taiwanese Journal of Mathematics, 13(2A), 545-558. [Google Scholor]
  22. Zhijian, Y. (2003). Initial boundary value problem for a class of non-linear strongly damped wave equations. Mathematical Methods in Applied Sciences, 26(12), 1047-1066.[Google Scholor]
]]>
Global solutions and general decay for the dispersive wave equation with memory and source terms https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/global-solutions-and-general-decay-for-the-dispersive-wave-equation-with-memory-and-source-terms/ Sun, 08 Nov 2020 11:52:32 +0000 https://old.pisrt.org/?p=4653
OMA-Vol. 4 (2020), Issue 2, pp. 116 - 122 Open Access Full-Text PDF
Mohamed Mellah
Abstract: This paper concerns with the global solutions and general decay to an initial-boundary value problem of the dispersive wave equation with memory and source terms.
]]>

Open Journal of Mathematical Analysis

Global solutions and general decay for the dispersive wave equation with memory and source terms

Mohamed Mellah
Faculty of Exact Sciences and Computer Science, Hassiba Benbouali University of Chlef, Chlef Algeria.; m.mellah@univ-chlef.dz

Abstract

This paper concerns with the global solutions and general decay to an initial-boundary value problem of the dispersive wave equation with memory and source terms.

Keywords:

Dispersive wave equation, small data global solution, general decay.

1. Introduction

This paper deals with the initial boundary value problem of the dispersive wave equation with memory and source terms
\begin{equation}\label{eq1.1} u_{tt}-\Delta u+\alpha\Delta^{2} u-\int_{0}^{t}g(t-\tau)\Delta^{2} u(\tau)d\tau+u_{t}=|u|^{p-1}u,\quad x\in\Omega,\ t>0, \end{equation}
(1)
where \(\Omega\) is a bounded domain in \(\mathbb{R}^{d}\) \((d\geq1)\) with a smooth boundary \(\partial\Omega\), \(\alpha\) is a positive constant and \(g(t)\) is a positive function that represents the kernel of the memory term, which will be specified in Section 2. Here, we understand \(\Delta^{2}u\) to be the dispersive term. In the absence of the viscoelastic term and the dispersive term (that is, if \(g=\alpha= 0\)), the model (1) reduces to the weakly damped wave equation
\begin{equation} u_{tt}-\Delta u+u_{t}=|u|^{p-1}u,\quad x\in\Omega,\ t>0. \end{equation}
(2)
The interaction between the weak damping term and the source term are considered by many authors. We refer the reader to, Haraux and Zuazua [1], Ikehata [2] and Levine [3, 4]. If \(\alpha=0\) and \(g\) is not trivial on \(\mathbb{R}\), but replacing the fourth order memory term in (1) by a weaker memory of the form \(\int_{0}^{t}g(t-\tau)\Delta u(\tau)d\tau\), then (1) can be rewritten as follows
\begin{equation}\label{eq1.3} u_{tt}-\Delta u+\int_{0}^{t}g(t-\tau)\Delta u(\tau)d\tau+u_{t}=|u|^{p-1}u,\quad x\in\Omega,\ t>0, \end{equation}
(3)
The Equation (3) has been considered by Wang et al [5]. Under some appropriate assumptions on \(g\), by introducing potential wells they obtained the existence of global solution and the explicit exponential energy decay estimates. Our main goal in the present paper is to discuss the global solutions and general decay to the following weakly damped wave equation with dispersive term, the fourth order memory term and the nonlinear source term
\begin{equation}\label{eq1.4} u_{tt}-\Delta u+\Delta^{2} u-\int_{0}^{t}g(t-\tau)\Delta^{2} u(\tau)d\tau+u_{t}=|u|^{p-1}u \quad \text{in}\ \Omega\times\mathbb{R}^{+}, \end{equation}
(4)
with simply supported boundary condition
\begin{equation}\label{eq1.5} u=0,\quad \frac{\partial u}{\partial\nu}=0 \quad \text{on}\ \partial\Omega\times\mathbb{R}^{+}, \end{equation}
(5)
and initial conditions
\begin{equation}\label{eq1.6} u(\cdot,0)=u_{0} \quad \text{and} \quad u_{t}(\cdot,0)=u_{1} \quad \text{in}\ \Omega, \end{equation}
(6)
where \(\Omega\) is a bounded domain of \(\mathbb{R}^{d}\) with a smooth boundary \(\partial\Omega\) and \(p>1\). Here, \(\nu\) is the unit outward normal to \(\partial\Omega\), and \(g(t)\) is a positive function that represents the kernel of the memory term, which will be specified in Section 2. We prove that Problem (4)-(6) has a global weak solution assuming small initial data. In addition, we show the general decay of solutions. The global solutions are constructed by means of the Galerkin approximations and the general decay is obtained by employing the technique used in [6].

2. Preliminaries

Before proceeding to our analysis, we use the following abbreviations \(\|\cdot\|_{q}=\|\cdot\|_{L^{q}(\Omega)}\) \((1\leq q\leq+\infty)\) denotes usual \(L^{q}\) norm, \((\cdot,\cdot)\) denotes the \(L^{2}\)-inner product, and consider the Sobolev spaces \(H^{1}_{0}(\Omega)\) and \(H^{2}_{0}(\Omega)\) with their usual scalar products and norms. We also use the embedding \(H^{1}_{0}(\Omega)\hookrightarrow L^{q}(\Omega)\) for \(2< q< \frac{2d}{d-2}\) if \(d\geq3\) or \(2< q< \infty\) if \(d=1,2\). In this case, the embedding constant is denoted by \(C_{*}\), that is \( \|u\|_{q}\leq C_{*}\|\nabla u\|_{2}. \) We define the polynomial \(Q\) by \( Q(z)=\frac{1}{2}z^{2}-\frac{C_{*}^{p+1}}{p+1}z^{p+1}, \) which is increasing in \([0,z_{0}]\), where \( z_{0}=C_{*}^{\frac{p+1}{1-p}} \) is its unique local maximum. Next, we give the assumptions for Problem (4)-(6).
(G1) The relaxation function \(g:\mathbb{R}_{+}\rightarrow \mathbb{R}_{+}\) is a bounded \(C^{1}\) function such that \( g(0)>0,\quad 0< \eta=1-\int_{0}^{\infty}g(\tau)d\tau\leq1-\int_{0}^{t}g(\tau)d\tau=\eta(t). \)
(G2) There exist positive constants \(\xi_{1}\) and \(\xi_{2}\) such that \( -\xi_{1}g(t)\leq g'(t)\leq-\xi_{2}g(t)\quad \forall t\geq0. \)
(G3) We also assume that \( 1< p\leq \frac{d+2}{d-2}\ \ \mbox {if} \ \ d\geq3 \ \ \mbox {and} \ \ \ p>1 \ \ \ \mbox {if} \ \ d=1,2. \) where \(\lambda_{1}\) is the first eigenvalue of the following problem
\begin{equation}\label{eq2.7} \Delta^{2}u=\lambda_{1}u \quad \text{in} \ \Omega,\quad u=\frac{\partial u}{\partial\nu}=0 \quad \text{in} \ \partial\Omega. \end{equation}
(7)

Remark 1. [7] Assuming \(\lambda_{1}\) is the first eigenvalue of the problem (7), we have

\begin{equation}\label{eq2.8} \|\Delta u\|_{2}^{2}\geq\lambda_{1}\|\nabla u\|_{2}^{2}. \end{equation}
(8)
Now, we define the following energy function associated with a solution \(u\) of the Problem (4)-(6)
\begin{equation} E(t)=\frac{1}{2}\|u_{t}\|^{2}_{2}+\frac{1}{2}\left(1-\int_{0}^{t}g(\tau)d\tau\right)\|\Delta u\|^{2}_{2}+\frac{1}{2}\|\nabla u\|^{2}_{2}+\frac{1}{2}(g\circ \Delta u)(t)-\frac{1}{p+1}\|u\|_{p+1}^{p+1} \end{equation}
(9)
for \(u\in H^{2}_{0}(\Omega)\), and
\begin{equation} E(0)=\frac{1}{2}\|u_{1}\|^{2}_{2}+\frac{1}{2}\|\Delta u_{0}\|^{2}_{2}+\frac{1}{2}\|\nabla u_{0}\|^{2}_{2}-\frac{1}{p+1}\|u_{0}\|_{p+1}^{p+1} \end{equation}
(10)
is the initial total energy. To facilitate further on our analysis, we use the following notation \begin{equation*} (g\circ\Delta u)(t)=\int_{0}^{t}g(t-\tau)\|\Delta u(\tau)-\Delta u(t)\|_{2}^{2}d\tau. \end{equation*} Now, we are in a position to state our main results.

3. Main results

Theorem 1. Assume that \((G1)-(G3)\) hold, \(u_{0}\in H^{2}_{0}(\Omega)\), \(u_{1}\in L^{2}(\Omega)\). Further assume that \(\|\nabla u_{0}\|_{2}< z_{0}\) and \(E(0)< Q(z_{0})\), then the Problem (4)-(6) possesses a global weak solution satisfying; \(u\in L^{\infty}(0,\infty;H^{2}_{0}(\Omega)),\quad u_{t}\in L^{\infty}(0,\infty;L^{2}(\Omega))\) for \(0\leq t< \infty\), and the energy identity

\begin{equation}\label{eq3.11} E(t)+\int_{0}^{t}\|u_{t}(\tau)\|_{2}^{2}d\tau-\frac{1}{2}\int_{0}^{t}(g'\circ \Delta u)(\tau)d\tau+\frac{1}{2}\int_{0}^{t}g(\tau)\|\Delta u(\tau)\|_{2}^{2}d\tau=E(0), \end{equation}
(11)
holds for \(0\leq t< \infty\). Moreover, for \(\zeta:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+}\) a increasing \(C^{2}\) function satisfying
\begin{equation}\label{eq3.12} \zeta(0)=0,\quad \zeta_{t}(0)>0,\quad \lim_{t\rightarrow+\infty}\zeta(t)=+\infty,\quad \zeta_{tt}(t)< 0\quad \forall t\geq0, \end{equation}
(12)
and, if \(\|g\|_{L^{1}(0,\infty)}\) is sufficiently small, we have for \(\kappa>0\); \( E(t)\leq E(0)e^{-\kappa \zeta(t)},\quad \forall t\geq0. \)

Remark 2. From (11) and \((G2)\), we can easily obtain

\begin{equation}\label{eq3.13} \frac{d}{dt}E(t)=-\|u_{t}(t)\|_{2}^{2}+\frac{1}{2}(g'\circ\Delta u)(t)-\frac{1}{2}g(t)\|\Delta u(t)\|_{2}^{2}\leq-\|u_{t}(t)\|_{2}^{2}-\frac{1}{2}\xi_{2}(g\circ\Delta u)(t)-\frac{1}{2}g(t)\|\Delta u(t)\|_{2}^{2}\leq0. \end{equation}
(13)

Remark 3. For \(\zeta(t)=t+\frac{t}{t+1}\), we can get the exponential decay rate \( E(t)\leq E(0)e^{-\kappa t},\quad \forall t\geq0\). For \(\zeta(t)=ln(1+t)\), we can get polynomial decay rate \( E(t)\leq E(0)(1+t)^{-\kappa },\quad \forall t\geq0. \)

4. Proof of main results

In this section, we shall divide the proof into two steps. In Step 1, we prove the global existence of weak solutions by using Galerkin's approximations. In Step 2, we establish the general decay of energy employing the method used in [6].

Step 1 Global existence of weak solutions

Let \(\left\{\omega_{j}\right\}_{j=1}^{\infty}\) be an orthogonal basis of \(H^{2}_{0}(\Omega)\) with \(\omega_{j}\) being the eigenfunction of the problem \( -\Delta \omega_{j}=\lambda_{j}\omega_{j},\quad x\in\Omega,\quad \omega_{j}=0,\quad x\in\partial\Omega. \) Let \(V^{n}=\text{Span}\left\{\omega_{1},\omega_{2},\cdot\cdot\cdot,\omega_{n}\right\}\). By the standard method of ODE, we know that \( u^{n}(t)=\sum_{j=1}^{n}b^{n}_{j}(t)\omega_{j}(x) \) of the Cauchy problem as follows
\begin{eqnarray}\label{eq4.14} &&\int_{\Omega}u^{n}_{tt}\omega dx+\int_{\Omega}\nabla u^{n}\cdot\nabla \omega dx+\int_{\Omega}\Delta u^{n}\cdot\Delta \omega dx-\int_{0}^{t}g(t-\tau)\int_{\Omega}\Delta u^{n}(\tau)\cdot\Delta \omega dxd\tau\nonumber\\ &&+\int_{\Omega}u^{n}_{t}\omega dx-\int_{\Omega}|u^{n}|^{p-1}u^{n}\omega dx=0, \end{eqnarray}
(14)
\begin{equation}\label{eq4.15} u^{n}(0)=u^{n}_{0}\rightarrow u_{0},\ \ \mbox {in} \ \ H^{2}_{0}(\Omega),\quad u^{n}_{t}(0)=u^{n}_{1}\rightarrow u_{1}\ \ \mbox {in} \ \ \ L^{2}(\Omega). \end{equation}
(15)
By the standard theory of ODE system, we prove the existence of solutions of Problem (14)-(15) on some interval \([0, t_{n})\), \(0< t_{n}< T\) for arbitrary \(T>0\), then, this solution can be extended to the whole interval \([0,T]\) using the first estimate given below. Taking \(\omega=u^{n}_{t}(t)\) in (14), we obtain
\begin{eqnarray}\label{eq4.16} &&\frac{1}{2}\frac{d}{dt}\|u^{n}_{t}\|^{2}_{2}+\frac{1}{2}\frac{d}{dt}\|\nabla u^{n}\|^{2}_{2}+\frac{1}{2}\frac{d}{dt}\|\Delta u^{n}\|^{2}_{2}-\frac{1}{p+1}\frac{d}{dt}\|u^{n}\|_{p+1}^{p+1} +\|u^{n}_{t}\|^{2}_{2}\nonumber\\ &&-\int_{0}^{t}g(t-\tau)\int_{\Omega}\Delta u^{n}(\tau)\cdot\Delta u^{n}_{t}(t)dxd\tau=0. \end{eqnarray}
(16)
For the last term on the left hand side of (16) we have
\begin{eqnarray}\label{eq4.17} &&-\int_{0}^{t}g(t-\tau)\int_{\Omega}\Delta u^{n}(\tau)\cdot\Delta u^{n}_{t}(t)dxd\tau =\frac{1}{2}\frac{d}{dt}(g\circ\Delta u^{n})(t)-\frac{1}{2}\frac{d}{dt}\left(\int_{0}^{t}g(\tau)d\tau\right)\|\Delta u^{n}(t)\|^{2}_{2}\nonumber\\ &&-\frac{1}{2}(g'\circ\Delta u^{n})(t)+\frac{1}{2}g(t)\|\Delta u^{n}(t)\|^{2}_{2}. \end{eqnarray}
(17)
Inserting (17) into (16) and integrating over \([0,t]\subset[0, T]\), we obtain
\begin{eqnarray}\label{eq4.18} &&\frac{1}{2}\|u^{n}_{t}\|^{2}_{2}+\frac{\eta(t)}{2}\|\Delta u^{n}(t)\|^{2}_{2}+\frac{1}{2}\|\nabla u^{n}\|^{2}_{2}-\frac{1}{p+1}\|u^{n}\|_{p+1}^{p+1}+\int_{0}^{t}\|u^{n}_{t}(\tau)\|^{2}_{2}d\tau+\frac{1}{2}(g\circ\Delta u^{n})(t)\nonumber\\ &&-\frac{1}{2}\int_{0}^{t}(g'\circ\Delta u^{n})(\tau)d\tau+\frac{1}{2}\int_{0}^{t}g(\tau)\|\Delta u^{n}(\tau)\|^{2}_{2}d\tau=E^{n}(0). \end{eqnarray}
(18)
Now from assumption \((G3)\) and the Sobolev embedding, we have that
\begin{equation}\label{eq4.19} \|u^{n}\|^{p+1}_{p+1}\leq C_{*}^{p+1}\|\nabla u^{n}\|^{p+1}_{2}, \end{equation}
(19)
and then we have
\begin{eqnarray}\label{eq4.20} &&\frac{1}{2}\|u^{n}_{t}\|^{2}_{2}+\frac{\eta(t)}{2}\|\Delta u^{n}(t)\|^{2}_{2}+\mathcal{Q}(\|\nabla u^{n}\|^{2}_{2})+\int_{0}^{t}\|u^{n}_{t}(\tau)\|^{2}_{2}d\tau+\frac{1}{2}(g\circ\Delta u^{n})(t)\nonumber\\ &&-\frac{1}{2}\int_{0}^{t}(g'\circ\Delta u^{n})(\tau)d\tau+\frac{1}{2}\int_{0}^{t}g(\tau)\|\Delta u^{n}(\tau)\|^{2}_{2}d\tau\leq E^{n}(0). \end{eqnarray}
(20)
By using the fact that \( -\int_{0}^{t}(g'\circ\Delta u^{n})(\tau)d\tau+\int_{0}^{t}g(\tau)\|\Delta u^{n}(\tau)\|^{2}_{2}d\tau\geq0, \) estimate (20) yields
\begin{eqnarray}\label{eq4.21} \frac{1}{2}\|u^{n}_{t}\|^{2}_{2}+\frac{\eta(t)}{2}\|\Delta u^{n}(t)\|^{2}_{2}+\frac{1}{2}(g\circ\Delta u^{n})(t)+\mathcal{Q}(\|\nabla u^{n}\|^{2}_{2})+\int_{0}^{t}\|u^{n}_{t}(\tau)\|^{2}_{2}d\tau\leq E^{n}(0). \end{eqnarray}
(21)
From \(E(0)< \mathcal{Q}(z_{0})\) and (15), it follows that
\begin{equation}\label{eq4.22} E^{n}(0)< \mathcal{Q}(z_{0}), \end{equation}
(22)
for sufficiently large \(n\). We claim that there exists an integer \(N\) such that
\begin{equation}\label{eq4.23} \|\nabla u^{n}(t)\|^{2}_{2}< z_{0}\quad \forall t\in[0,t_{n})\quad n>N. \end{equation}
(23)
Suppose the claim is proved. Then \(\mathcal{Q}(\|\nabla u^{n}\|^{2}_{2})\geq0\) and from (21) and (22),
\begin{equation}\label{eq4.24} \frac{1}{2}\|u^{n}_{t}\|^{2}_{2}+\frac{\eta(t)}{2}\|\Delta u^{n}(t)\|^{2}_{2}+\frac{1}{2}(g\circ\Delta u^{n})(t)+\int_{0}^{t}\|u^{n}_{t}(\tau)\|^{2}_{2}d\tau\leq E^{n}(0)< \mathcal{Q}(z_{0}) \end{equation}
(24)
for sufficiently large \(n\) and \(0\leq t< \infty\).

Proof. [Proof of the claim] Suppose that (23) false. Then for each \(n>N\), there exists \(t\in[0,t_{n})\) such that \(\|\nabla u^{n}(t)\|_{2}\geq z_{0}\). We note that from \(\|\nabla u_{0}\|_{2}< z_{0}\) and (15) there exists \(N_{0}\) such that \( \|\nabla u^{n}(0)\|_{2}< z_{0}\quad \forall n>N_{0}. \) Then by continuity there exits a first \(t_{n}^{*}\in[0,t_{n})\) such that

\begin{equation}\label{eq4.25} \|\nabla u^{n}(t_{n}^{*})\|_{2}=z_{0}, \end{equation}
(25)
from where \( \mathcal{Q}(\|\nabla u^{n}(t)\|_{2})\geq0 \quad \forall t\in[0,t_{n}^{*}]. \) Now from \(E(0)< \mathcal{Q}(z_{0})\) and (24), there exists \(N>N_{0}\) and \(\gamma\in(0,z_{0})\) such that \( 0\leq\frac{1}{2}\|u^{n}_{t}(t)\|^{2}_{2}+\frac{\eta(t)}{2}\|\Delta u^{n}(t)\|^{2}_{2}+\frac{1}{2}(g\circ\Delta u^{n})(t)+\mathcal{Q}(\|\nabla u^{n}(t)\|^{2}_{2})\leq\mathcal{Q}(\gamma)\quad \forall t\in[0,t_{n}^{*}]\quad \forall n>N. \) Then the monotonicity of \(\mathcal{Q}\) in \([0,z_{0}]\) implies that \( 0\leq\|\nabla u^{n}(t)\|^{2}_{2}\leq\gamma< z_{0}\quad \forall t\in[0,t_{n}^{*}], \) and in particular, \(\|\nabla u^{n}(t)\|^{2}_{2}< z_{0}\), which is a contradiction to (24). From (24), we have
\begin{equation}\label{eq4.26} \|\Delta u^{n}\|^{2}_{2}< \frac{2\mathcal{Q}(z_{0})}{\eta},\quad 0\leq t< \infty, \end{equation}
(26)
\begin{equation}\label{eq4.27} \| u^{n}_{t}\|^{2}_{2}< 2\mathcal{Q}(z_{0}),\quad 0\leq t< \infty, \end{equation}
(27)
\begin{equation}\label{eq4.28} \int_{0}^{t}\|u^{n}_{t}(\tau)\|^{2}_{2}d\tau< \mathcal{Q}(z_{0}),\quad 0\leq t< \infty. \end{equation}
(28)
Using Sobolev inequality, (8) and (26), it follows that
\begin{eqnarray}\label{eq4.29} \|u^{n}\|^{2}_{p+1}\leq C_{*}^{2}\|\nabla u^{n}\|^{2}_{2}\leq C_{*}^{2}\lambda_{1}^{-1}\|\Delta u^{n}\|^{2}_{2}< \frac{2C_{*}^{2}\lambda_{1}^{-1}\mathcal{Q}(z_{0})}{\eta},\quad 0\leq t< \infty. \end{eqnarray}
(29)
Furthermore, by (29), we get
\begin{eqnarray}\label{eq4.30} |(|u^{n}|^{p-1}u^{n},u^{n})|\leq \|u^{n}\|^{p+1}_{p+1}< C_{*}^{p+1}\left(\frac{2C_{*}^{2}\lambda_{1}^{-1}\mathcal{Q}(z_{0})}{\eta}\right)^{\frac{p+1}{2}},\quad 0\leq t< \infty. \end{eqnarray}
(30)
The estimates (26)-(30) permit us to obtain a subsequences of \(\left\{u_{n}\right\}\) which from now on will be also denoted by \(\left\{u_{n}\right\}\) and functions \(u\), \(\chi\) such that
\begin{equation}\label{eq4.31} u_{n}\rightarrow u \ \ \mbox {weak star in} \ \ \ L^{\infty}(0,\infty;H_{0}^{2}(\Omega)),\quad n\rightarrow+\infty, \end{equation}
(31)
\begin{equation}\label{eq4.32} u^{n}_{t}\rightarrow u_{t} \ \ \mbox {weak star in} \ \ \ L^{\infty}(0,\infty;L^{2}(\Omega)),\quad n\rightarrow+\infty, \end{equation}
(32)
\begin{equation}\label{eq4.33} |u^{n}|^{p-1}u^{n}\rightarrow \chi \ \ \mbox {weak star in} \ \ \ L^{\infty}(0,\infty;L^{\frac{p+1}{p}}(\Omega)),\quad n\rightarrow+\infty. \end{equation}
(33)
Besides, from Lions-Aubin Lemma we also have
\begin{equation}\label{eq4.34} u^{n}\rightarrow u \ \ \mbox {strongly in} \ \ \ L^{2}(0,\infty;L^{2}(\Omega)),\quad n\rightarrow+\infty, \end{equation}
(34)
and consequently, making use of the Lemma 1.3 in [8], we deduce
\begin{equation}\label{eq4.35} |u^{n}|^{p-1}u^{n}\rightarrow \chi=|u|^{p-1}u \ \ \mbox {weak star in} \ \ \ L^{\infty}(0,\infty;L^{\frac{p+1}{p}}(\Omega)),\quad n\rightarrow+\infty. \end{equation}
(35)
Thus, we obtain that \(u\) is a global weak of problem (4)-(6). Next, we shall prove that \(u\) satisfies (11). From the discussion above, we obtain for each fixed \(t>0\) that
\begin{equation} \lim_{n\rightarrow +\infty}(g\circ\Delta u^{n})(t)=(g\circ\Delta u)(t),\quad \lim_{n\rightarrow +\infty}\|u^{n}\|_{p+1}^{p+1}=\|u\|_{p+1}^{p+1}. \end{equation}
(36)
We obtain for each fixed \(t>0\) that
\begin{eqnarray} |(g\circ\Delta u)(t)-(g\circ\Delta u^{n})(t)|&=&\left|\int_{0}^{t}g(t-\tau)\|\Delta u(\tau)-\Delta u(t)\|^{2}_{2}d\tau-\int_{0}^{t}g(t-\tau)\|\Delta u^{n}(\tau)-\Delta u^{n}(t)\|^{2}_{2}d\tau\right|\nonumber\\ &\leq& \int_{0}^{t}g(t-\tau)\|\Delta u(\tau)-\Delta u^{n}(\tau)\|_{2}\|\Delta u(\tau)+\Delta u^{n}(\tau)\|_{2}d\tau\nonumber\\ &&+\int_{0}^{t}g(t-\tau)\|\Delta u(\tau)-\Delta u^{n}(\tau)\|_{2}d\tau\|\Delta u(t)+\Delta u^{n}(t)\|_{2}\nonumber\\ &&+\int_{0}^{t}g(t-\tau)\|\Delta u(\tau)+\Delta u^{n}(\tau)\|_{2}d\tau\|\Delta u(t)-\Delta u^{n}(t)\|_{2}\nonumber\\ &&+\int_{0}^{t}g(\tau)d\tau\|\Delta u(t)+\Delta u^{n}(t)\|_{2}\|\Delta u(t)-\Delta u^{n}(t)\|_{2}\nonumber\\ &\leq&C\int_{0}^{t}g(t-\tau)\|\Delta u(\tau)-\Delta u^{n}(\tau)\|_{2}d\tau+C\int_{0}^{t}g(\tau)d\tau\|\Delta u(t)-\Delta u^{n}(t)\|_{2}\rightarrow 0,\nonumber\\&& \end{eqnarray}
(37)
as \(n\rightarrow +\infty\), and
\begin{eqnarray} \|u^{n}\|_{p+1}^{p+1}-\|u\|_{p+1}^{p+1} &\leq&(p+1)\left|\int_{\Omega}|u+\theta_{n}u^{n}|^{p-1}(u+\theta_{n}u^{n})(u^{n}-u)dx\right|\nonumber\\ &\leq&(p+1)\|u+\theta_{n}u^{n}\|_{p+1}^{p}\|u^{n}-u\|_{p+1}\leq C\|u^{n}-u\|_{p+1}\rightarrow 0, \end{eqnarray}
(38)
as \(n\rightarrow +\infty\), where \(0< \theta_{n}< 1\). Hence, we have
\begin{equation} \lim_{n\rightarrow +\infty}(g\circ\Delta u^{n})(t)=(g\circ\Delta u)(t),\quad \lim_{n\rightarrow +\infty}\|u^{n}\|_{p+1}^{p+1}=\|u\|_{p+1}^{p+1}. \end{equation}
(39)
From (15), it follows that \(E^{n}(0)\rightarrow E(0)\) as \(n\rightarrow+\infty\). Finally, taking \(n\rightarrow +\infty\) in (18), we deduce that the energy identity (11) holds for \(0\leq t< \infty\).

Step 2 General decay of the energy

Firstly, we state several Lemmas to prove the decay rate estimate of the energy.

Lemma 1. Let \(u\in L^{\infty}(0,\infty;H^{2}_{0}(\Omega))\) be the solution of (4)-(6) and \(E(0)< \mathcal{Q}(z_{0})\), \(\|\nabla u_{0}\|_{2}< z_{0}\), then we have

\begin{eqnarray}\label{eq4.40} 0\leq E(t)\leq\frac{1}{2}\|u_{t}\|_{2}^{2}+C_{1}\|\Delta u\|^{2}_{2}+\frac{1}{2}(g\circ \Delta u)(t), \end{eqnarray}
(40)
where \(C_{1}=\frac{1}{2}+(2\lambda_{1})^{-1}\).

Proof. From \(E(0)< \mathcal{Q}(z_{0})\) and \(\|\nabla u_{0}\|_{2}< z_{0}\), we can obtain \(\mathcal{Q}(\|\nabla u(t)\|_{2})\geq0\) for \(0\leq t< \infty\). Thus we have

\begin{eqnarray}\label{eq4.41} E(t)&=&\frac{1}{2}\|u_{t}\|^{2}_{2}+\frac{1}{2}\left(1-\int_{0}^{t}g(\tau)d\tau\right)\|\Delta u\|^{2}_{2}+\frac{1}{2}(g\circ \Delta u)(t)+\frac{1}{2}\|\nabla u\|^{2}_{2}-\frac{1}{p+1}\|u\|_{p+1}^{p+1}\nonumber\\ &\geq&\frac{1}{2}\|u_{t}\|^{2}_{2}+\frac{\eta}{2}\|\Delta u\|^{2}_{2}+\frac{1}{2}(g\circ \Delta u)(t)+\mathcal{Q}(\|\nabla u(t)\|_{2})\geq0, \end{eqnarray}
(41)
and
\begin{eqnarray}\label{eq4.42} E(t)\leq\frac{1}{2}\|u_{t}\|^{2}_{2}+\frac{1}{2}\|\Delta u\|^{2}_{2}+\frac{1}{2}(g\circ \Delta u)(t)+\frac{1}{2}\|\nabla u\|^{2}_{2} \leq\frac{1}{2}\|u_{t}\|_{2}^{2}+C_{1}\|\Delta u\|^{2}_{2}+\frac{1}{2}(g\circ \Delta u)(t). \end{eqnarray}
(42)

Lemma 2. The energy \(E(t)\) satisfies

\begin{eqnarray}\label{eq4.43} \frac{d E(t)}{dt}&\leq&-\|u_{t}(t)\|^{2}_{2}-\frac{1}{2}\xi_{2}(g\circ\Delta u)(t)-\frac{1}{2}\left[g(0)-\xi_{1}\|g\|_{L^{1}(0,\infty)}\right]\|\Delta u(t)\|^{2}_{2}\quad \forall t\geq0. \end{eqnarray}
(43)

Proof. From (13), we have

\begin{eqnarray}\label{eq4.44} \frac{d E(t)}{dt}&\leq&-\|u_{t}(t)\|^{2}_{2}-\frac{\xi_{2}}{2}(g\circ \Delta u)(t)-\frac{1}{2}g(t)\|\Delta u(t)\|^{2}_{2}. \end{eqnarray}
(44)
From assumptions \((G2)\) and since \(\int_{0}^{t}g'(\tau)d\tau=g(t)-g(0)\), we obtain
\begin{eqnarray}\label{eq4.45} -\frac{1}{2}g(t)\|\Delta u(t)\|^{2}_{2}&=&-\frac{1}{2}g(0)\|\Delta u(t)\|^{2}_{2}-\frac{1}{2}\left(\int_{0}^{t}g'(\tau)d\tau\right)\|\Delta u(t)\|^{2}_{2}\nonumber\\ &\leq&-\frac{1}{2}g(0)\|\Delta u(t)\|^{2}_{2}+\frac{\xi_{1}}{2}\|g\|_{L^{1}(0,\infty)}\|\Delta u(t)\|^{2}_{2}\nonumber\\ &=&-\frac{1}{2}\left[g(0)-\xi_{1}\|g\|_{L^{1}(0,\infty)}\right]\|\Delta u(t)\|^{2}_{2}. \end{eqnarray}
(45)
Then, Combining (45) and (44) our conclusion holds. Multiplying (43) by \(e^{\kappa\zeta(t)}\) \((\kappa>0)\) and using (40), we have
\begin{eqnarray}\label{eq4.46} \frac{d}{dt}\left(e^{\kappa\zeta(t)}E(t)\right)&\leq&-\|u_{t}(t)\|^{2}_{2}e^{\kappa\zeta(t)}E(t)-\frac{1}{2}\xi_{2}(g\circ\Delta u)(t)e^{\kappa\zeta(t)}E(t)\nonumber\\ &&-\frac{1}{2}\left[g(0)-\xi_{1}\|g\|_{L^{1}(0,\infty)}\right]\|\Delta u(t)\|^{2}_{2}e^{\kappa\zeta(t)}E(t)+\kappa\zeta_{t}(t)e^{\kappa\zeta(t)}E(t)\nonumber\\ &\leq&-\frac{1}{2}\left[2-\kappa\zeta_{t}(t)\right]\|u_{t}(t)\|^{2}_{2}e^{\kappa\zeta(t)}E(t)-\frac{1}{2}\left[\xi_{2}-\kappa\zeta_{t}(t)\right](g\circ\Delta u)(t)e^{\kappa\zeta(t)}E(t)\nonumber\\ &&-\frac{1}{2}\left[g(0)-\xi_{1}\|g\|_{L^{1}}-2C_{1}\kappa\zeta_{t}(t)\right]\|\Delta u(t)\|^{2}_{2}e^{\kappa\zeta(t)}E(t). \end{eqnarray}
(46)
Using the fact that \(\zeta_{t}\) is decreasing by (12), we arrive at
\begin{eqnarray}\label{eq4.47} \frac{d}{dt}\left(e^{\kappa\zeta(t)}E(t)\right)&\leq&-\frac{1}{2}\left[2-\kappa\zeta_{t}(0)\right] \|u_{t}(t)\|^{2}_{2}e^{\kappa\zeta(t)}E(t)-\frac{1}{2}\left[\xi_{2}-\kappa\zeta_{t}(0)\right](g\circ\Delta u)(t)e^{\kappa\zeta(t)}E(t)\nonumber\\ &&-\frac{1}{2}\left[g(0)-\xi_{1}\|g\|_{L^{1}(0,\infty)}-2C_{1}\kappa\zeta_{t}(0)\right]\|\Delta u(t)\|^{2}_{2}e^{\kappa\zeta(t)}E(t). \end{eqnarray}
(47)
Choosing \(\|g\|_{L^{1}(0,\infty)}\) sufficiently small so that \(g(0)-\xi_{1}\|g\|_{L^{1}(0,\infty)}=B>0\) and defining \(\kappa_{0}=\min\left\{\frac{2}{\zeta_{t}(0)},\frac{\xi_{2}}{\zeta_{t}(0)},\frac{B}{2C_{1}\zeta_{t}(0)}\right\},\) we conclude by taking \(\kappa\in(0, \kappa_{0}]\) in (47) that \( \frac{d}{dt}\left(e^{\kappa\zeta(t)}E(t)\right)\leq0,\quad t>0. \) Integrating the above inequality over \((0,t)\), it follows that \( E(t)\leq E(0)e^{-\kappa\zeta(t)},\quad t>0. \)

Competing Interests

The author declares no conflict of interest.

References

  1. Haraux, A., & Zuazua, E. (1988). Decay estimates for some semilinear damped hyperbolic problems. Archive for Rational Mechanics and Analysis, 100(2), 191-206.[Google Scholor]
  2. Ikehata, R. (1996). Some remarks on the wave equations with nonlinear damping and source terms. Nonlinear Analysis: Theory, Methods & Applications, 27(10), 1165-1175.[Google Scholor]
  3. Levine, H. A. (1974). Instability and nonexistence of global solutions to nonlinear wave equations of the form \(Pu_{tt}=Au+F(u)\). Transactions of the American Mathematical Society, 192, 1-21.[Google Scholor]
  4. Levine, H. A. (1974). Some additional remarks on the nonexistence of global solutions to nonlinear wave equations. SIAM Journal on Mathematical Analysis, 5(1), 138-146.[Google Scholor]
  5. Wang, Y., & Wang, Y. (2008). Exponential energy decay of solutions of viscoelastic wave equations. Journal of Mathematical Analysis and Applications, 347(1), 18-25.[Google Scholor]
  6. Mellah, M., & Hakem, A. (2019). Global existence, uniqueness, and asymptotic behavior of solution for the Euler-Bernoulli viscoelastic equation. Open Journal of Mathematical Analysis, 3(1), 42-51.[Google Scholor]
  7. Park, J. Y., & Kang, J. R. (2010). Global existence and uniform decay for a nonlinear viscoelastic equation with damping. Acta Applicandae Mathematicae, 110(3), 1393-1406.[Google Scholor]
  8. Lions, J. L. (1969). Quelques méthodes de résolution des problemes aux limites non linéaires. Dunod Gauthier-Villars, Paris.[Google Scholor]
]]>
Controllability for some nonlinear impulsive partial functional integrodifferential systems with infinite delay in Banach spaces https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/controllability-for-some-nonlinear-impulsive-partial-functional-integrodifferential-systems-with-infinite-delay-in-banach-spaces/ Sun, 08 Nov 2020 11:39:02 +0000 https://old.pisrt.org/?p=4651
OMA-Vol. 4 (2020), Issue 2, pp. 104 - 115 Open Access Full-Text PDF
Patrice Ndambomve, Khalil Ezzinbi
Abstract: This work concerns the study of the controllability for some impulsive partial functional integrodifferential equation with infinite delay in Banach spaces. We give sufficient conditions that ensure the controllability of the system by supposing that its undelayed part admits a resolvent operator in the sense of Grimmer, and by making use of the measure of noncompactness and the Mönch fixed-point Theorem. As a result, we obtain a generalization of the work of K. Balachandran and R. Sakthivel (Journal of Mathematical Analysis and Applications, 255, 447-457, (2001)) and a host of important results in the literature, without assuming the compactness of the resolvent operator. An example is given for illustration.
]]>

Open Journal of Mathematical Analysis

Controllability for some nonlinear impulsive partial functional integrodifferential systems with infinite delay in Banach spaces

Patrice Ndambomve\(^1\), Khalil Ezzinbi
Department of Mathematics, Faculty of Science, University of Buea.; (P.N)
Cadi Ayyad University, Faculty of Science Semlalia, Department of Mathematics, B.P. 2390, Marrakesh.; (K.E)
\(^{1}\)Corresponding Author: ndambomve.patrice@ubuea.cm;

Abstract

This work concerns the study of the controllability for some impulsive partial functional integrodifferential equation with infinite delay in Banach spaces. We give sufficient conditions that ensure the controllability of the system by supposing that its undelayed part admits a resolvent operator in the sense of Grimmer, and by making use of the measure of noncompactness and the Mönch fixed-point Theorem. As a result, we obtain a generalization of the work of K. Balachandran and R. Sakthivel (Journal of Mathematical Analysis and Applications, 255, 447-457, (2001)) and a host of important results in the literature, without assuming the compactness of the resolvent operator. An example is given for illustration.

Keywords:

Controllability, impulsive functional differential equation, infinite delay, resolvent operator, measure of noncompactness, Mönch’s fixed-point theorem.

1. Introduction

The dynamics of evolution processes is often subjected to abrupt changes such as shocks, harvesting, and natural disasters. Often these short-term perturbations are treated as having acted instantaneously or in the form of impulses [1]. The study of dynamical systems with impulsive effects is of great importance. Impulsive differential equations have become more important in recent years in some mathematical models of real processes and phenomena studied in control, physics, chemistry, population dynamics, aero- nautics and engineering. The concept of controllability plays an important role in many areas of applied mathematics. In recent years, significant progress has been made in the controllability of linear and nonlinear deterministic infinite dimensional systems, see for instance [2, 3, 4, 5, 6, 7, 8, 9, 10, 11] and the references therein. Many authors studied the controllability problem of nonlinear systems with delay in infinite dimensional Banach spaces; see for instance [2, 6, 9, 10, 11] etc and the references contained in them.

The controllability problem for nonlinear impulsive systems in infinite dimensional Banach spaces has been studied by several authors, see e.g., [6, 7, 11]. In [11], Selvi and Arjunan considered the following impulsive differential systems with finite delay
\begin{equation} \begin{cases} x'(t)=A(t)x(t)\ +\ f(t,x_t)+Cu(t),& \text{for}\ t\in J=[0,b],\ t\neq t_k,\ k=1,2,\cdots,m ,\\ \Delta x(t_k)=I_k(x_{t_k}),& k=1,2,\cdots,m,\\ x(t) = \varphi(t), & t\in[-r,0]. \end{cases} \label{eqselvi} \end{equation}
(1)
Using the Hausdorff measure of noncompactness and the Mönch fixed-point theorem and under some sufficient conditions, they obtained a controllability result for Equation (1), without assuming the compactness of the semigroup. In [6], Machado et al., considered the following impulsive mixed-type functional integro-differential system with finite delay and nonlocal conditions of the form
\begin{equation} \begin{cases} x'(t)=A(t)x(t)+\displaystyle{f\left(t, x_t,\int_0^sh(s,t,x_s)ds,\int_0^bk(t,s,x_s)ds\right)}\ +Cu(t)& \text{for}\ t\in J=[0,b], t\neq t_k, k=1,2,\cdots,m\\ \Delta x(t_k)=I_k(x_{t_k}),& k=1,2,\cdots,m\\ x_0 = \phi+g(x),& t\in[-r,0]. \end{cases} \label{eqmachado} \end{equation}
(2)
Using the Mönch fixed-point theorem via measures of noncompactness and semigroup theory, they obtained a controllability result for Equation (2) without assuming the compactness of the evolution system. However, the result obtained in [11] and [6] are only in connection with finite delay and the impulsive functions \(I_k\ (k = 1,\cdots , m)\) are assumed to be bounded. But since most often many systems arising from realistic models can be described as functional differential and integrodifferential systems with infinite delay [12], it would be natural and interesting to discuss this kind of problems. In an attempt to address this kind of problems, Chang [13] considered the following impulsive functional differential systems with infinite delay;
\begin{equation} \begin{cases} x'(t)=Ax(t)\ +\ f(t,x_t)+Cu(t)& \text{for}\ t\in J=[0,b],\ t\neq t_k,\ k=1,2,\cdots,m, \\ \Delta x(t_k)=I_k(x_{t_k}),& k=1,2,\cdots,m,\\ x(t) = \phi\in\mathcal{BM}_h, \end{cases} \label{eqyong} \end{equation}
(3)
where \(\mathcal{BM}_h\) is an abstract phase space. Assuming the compactness of the C\(_0\)-semigroup generated by \(A\) and using Schauder's fixed point theorem together with some sufficient conditions, the author obtained a controllability result for Equation (3).
Motivated by the above works, we study in this paper the controllability for some systems that take the form of the following abstract model of impulsive partial functional integrodifferential equation with infinite delay in a Banach space \((X,\ \|\cdot\|)\);
\begin{equation} \begin{cases} x'(t)=Ax(t)\ +\ \displaystyle{\int_0^t\gamma(t-s)x(s)ds}\ +\ f(t,x_t)+Cu(t)& \text{for}\ t\in J=[0,b],\ t\neq t_k,\ k=1,2,\cdots,m,\\ \Delta x(t_k)=I_k(x_{t_k}),& k=1,2,\cdots,m\\ x_0 = \varphi\in\mathcal{P}, \end{cases} \label{eq1} \end{equation}
(4)
where \(A:\mathcal{D}(A)\rightarrow X\) is the infinitesimal generator of a \(C_0\)-semigroup \(\big(T(t)\big)_{t\geq0}\) on a Banach space \(X\); for \(t\geq0\), \(\gamma(t)\) is a closed linear operator with domain \(\mathcal{D}(\gamma(t))\supset\mathcal{D}(A)\). The control \(u\) belongs to \(L^2(J,U)\) which is a Banach space of admissible controls, where \(U\) is a Banach space. The operator \( C\in\mathcal{L}(U,X)\), where \(\mathcal{L}(U,X)\) denotes the Banach space of bounded linear operators from \(U\) into \(X\), and the phase space \(\mathcal{P}\) is a linear space of functions mapping \(]-\infty,0]\) into \(X\) satisfying axioms which will be described later, for every \(t\geq0\), \(x_t\) denotes the history function of \(\mathcal{P}\) defined by \(x_t(\theta)=x(t+\theta)\ \ \text{for}\ -\infty\leq\theta\leq0.\) Here \(0< t_1< \cdots< t_m< t_{m+1}< b\) are prefixed numbers, \(f:\, J\times \mathcal{P}\rightarrow X,\ I_k:\mathcal{P}\rightarrow X\) are appropriate functions satisfying some conditions, and the symbol \(\Delta\xi(t)\) represent the jump of the function \(\xi\) at \(t\), which is defined by \(\Delta\xi(t)=\xi(t^{+})-\xi(t^{-})\). In the literature devoted to equations with finite delay, the phase space is the space of continuous functions on \([-r,0]\), for some \(r>0\), endowed with the uniform norm topology. But when the delay is unbounded, the selection of the phase space \(\mathcal{P}\) plays an important role in both qualitative and quantitative theories. A usual choice is a normed space satisfying some suitable axioms, which was introduced by Hino et al., [14]. In this work, we use resolvent operators for integral equations, the Mönch fixed-point theorem and the measure of noncompactness, without any compactness assumption on the resolvent operators. In [15], Grimmer proved the existence and uniqueness of resolvent operators for this type of functional integrodifferential equations that give the variation of parameters formula for the solution. In [16], Desch, Grimmer and Schappacher proved the equivalence of the compactness of the resolvent operator and that of the operator semigroup. In this work, we use the equivalence between the operator-norm continuity of the associated resolvent operator and that of the operator semigroup. This property allows us to drop the compactness assumption on the operator semigroup, considered by the authors in [2, 10], and prove that the operator solution satisfies the Mönch condition. The variation of parameters formula for the mild solutions of Equation (4) is given by the resolvent operator, and we prove the controllability result using the Mönch fixed-point theorem and the measure of noncompactness. This method enables us overcome the resolvent operator case considered in this work. In contrary to the evolution semigroup case considered in [6, 11], here the semigroup property can not be used because resolvent operators in general do not form semigroups. To the best of our knowledge, up to now no work has reported on controllability of impulsive partial functional integrodifferential Equation (4) with infinite delay. It has been an untreated topic in the literature, and this fact is the main aim and motivation of the present work. The work is organized as follows; Section 2 is devoted to preliminary results. In this Section, we give the definition of resolvent operator. This allows us to define the mild solution of Equation (4). In Section 3, we study the controllability of Equation (4). In Section 4, we give an example to illustrate this work.

2. Integrodifferential equations, measure of noncompactness and Mönch's theorem

In this Section, we introduce some definitions and lemmas that will be used throughout the paper. Let \(J=[0,b],\ \ b>0\) and let \(X\) be a Banach space. A measurable function \(x:J\rightarrow X\) is Bochner integrable if and only if \(\|x\|\) is Lebesgue integrable. We denote by \(L^1(J,X)\) the Banach space of Bochner integrable functions \(x:J\rightarrow X\) normed by $$ \|x\|_{L^1}=\int_0^b\|x(t)\|dt.$$ In considering the impulsive condition, it is important to introduce some additional concepts and notations. We say that a function \(x:[\mu,\eta]\rightarrow X\) is a normalized piecewise continuous function on \([\mu,\eta]\) if \(x\) is piecewise continuous, and left continuous on \((\mu,\eta]\). Let \(\mathcal{PC}([\mu,\eta],X)\) denote the space of normalized piecewise continuous functions from \([\mu,\eta]\) to \(X\). The notation \(\mathcal{PC}\) stands for the space of all functions \(x:[\mu,\eta]\rightarrow X\) such that \(x\) is continuous at \(t\neq t_k,\ x(t_k^{-})=x(t_k)\) and \(x(t_k^{+})\) exists for all \(k=1,2,\cdots,m\). In this Section, \((\mathcal{PC},\|\cdot\|_{\mathcal{PC}})\) is a Banach space endowed with the norm \(\|x\|_{\mathcal{PC}}=\sup_{s\in J}\|x(s)\|\). In this work, we will employ an axiomatic definition of the phase space \(\mathcal{P}\) introduced by Hino et al., in [14]. Thus, \((\mathcal{P},\|\cdot\|_{\mathcal{P}})\) will be a normed linear space of functions mapping \(]-\infty,0]\) into \(X\) and satisfying the following axioms;
  • (\(A_1\)) For \(\sigma>0\), if \(x:\,]-\infty,\mu+\sigma]\rightarrow X\) is such that \(x_{\mu}\in\mathcal{P}\) and \(x|_{[\mu,\mu+\sigma]}\in \mathcal{PC}([\mu,\mu+\sigma];X)\) then, for every \(t\in[\mu,\mu+\sigma]\), the following conditions hold; There exist positive constant \(H\) and functions \(K:\,\mathbb{R}^{+}\rightarrow [1,\infty)\) continuous and \(M:\,\mathbb{R}^{+}\rightarrow[1,\infty)\) locally bounded, and all independent of \(x\), such that
    • (i) \(x_t\in\mathcal{P}\),
    • (ii) \(\|x(t)\|\leq H\|x_t\|_{\mathcal{P}}\), which is equivalent to \(\|\varphi(0)\|\leq H\|\varphi\|_{\mathcal{P}}\) for every \(\varphi\in\mathcal{P}\),
    • (iii) \(\|x_t\|_{\mathcal{P}}\leq K(t-\mu)\displaystyle\sup_{\mu\leq s\leq t}\|x(s)\|+M(t-\mu)\|x_{\mu}\|_{\mathcal{P}}\).
  • \((A_2)\) For the function \(x\) in \(A_1\), \(t\rightarrow x_t\) is a \(\mathcal{P}\)-valued continuous function for \(t\in[\mu,\mu+\sigma]\).
  • (\(A_3\)) The space \(\mathcal{P}\) is complete.
Consider the following linear homogeneous equation;
\begin{equation} \left\{ \begin{array}{l} x'(t)=Ax(t)+{\displaystyle\int_0^t\gamma(t-s)x(s)ds}\ \ \text{for} \quad t\geq0, \\ x(0)=x_0\in X. \end{array} \right.%\eqno(2.1) \label{eq2} \end{equation}
(5)
where \(A\) and \(\gamma(t)\) are closed linear operators on a Banach space \(X\). In the sequel, we assume \(A\) and \(\big(\gamma(t)\big)_{t\geq0 }\) satisfy the following conditions;
  • \((H_1)\)] \(A\) is a densely defined closed linear operator in \(X\), hence \(\mathcal{D}(A)\) is a Banach space equipped with the graph norm defined by, \(|y|=\|Ay\|+\|y\|\) which will be denoted by \((X_1,|\cdot|)\).
  • \((H_2)\)] \(\big(\gamma(t)\big)_{t\geq0 }\) is a family of linear operators on \(X\) such that \(\gamma(t)\) is continuous when regarded as a linear map from \((X_1,|\cdot|)\) into \((X,\|\cdot\|)\) for almost all \(t\geq0\) and the map \(t\mapsto \gamma(t)y\) is measurable for all \(y\in X_1\) and \(t\geq0\), and belongs to \(W^{1,1}(\mathbb{R}^{+},X)\). Moreover there is a locally integrable function \(b:\mathbb{R}^{+}\rightarrow\mathbb{R}^{+}\) such that \(\|\gamma(t)y\|\ \leq\ b(t)|y|\ \ {\text and}\ \left\|\frac{d}{dt}\gamma(t)y\right\|\ \leq\ b(t)|y|\ .\)

Remark 1. Note that \((H_2)\) is satisfied in the modelling of Heat Conduction in materials with memory and viscosity. More details can be found in [17].

Let \(\mathcal{L}(X)\) be the Banach space of bounded linear operators on \(X\).

Definition 1.[18] A resolvent operator \(\big(R(t)\big)_{t\geq0 }\) for Equation \((5)\) is a bounded operator valued function $$R\, :\ [0,+\infty)\ \longrightarrow \ \mathcal{L}(X)$$ such that

  • (i) \(R(0)=Id_X\) \ and \ \(\|R(t)\|\leq Ne^{\beta t}\) for some constants \(N\) and \(\beta\).
  • (ii) For all \(x\in X\),\ the map \(t\mapsto R(t)x\) is continuous for \(t\geq0\).
  • (iii) Moreover for \ \(x\in X_1\),\ \ \(R(\cdot)x\,\in\,\mathcal{C}^1(\mathbb{R}^{+};X)\cap\mathcal{C}(\mathbb{R}^{+};X_1)\) \ and \(R'(t)x = AR(t)x+\int_0^t\gamma(t-s)R(s)xds = R(t)Ax+\int_0^tR(t-s)\gamma(s)xds.\)
\label{defresolvent}

Observe that the map defined on \(\mathbb{R}^{+}\) by \ \(t\mapsto R(t)x_0\)\ solves Equation \((5)\) for \(x_0\in\mathcal{D}(A)\).

Theorem 1.[15] Assume that \((H_{1})\) and \((H_2)\) hold. Then, the linear Equation (5) has a unique resolvent operator \(\big(R(t)\big)_{t\geq0 }\).

Remark 2. In general, the resolvent operator \(\big(R(t)\big)_{t\geq0}\) for Equation (5) does not satisfy the semigroup law, namely, \(R(t+s) \ \neq \ R(t)R(s) \ \mbox{for some} \ \ t,\, s >0\, . \)

We have the following theorem that establishes the equivalence between the operator-norm continuity of the \(C_0\)-semigroup and the resolvent operator for integral equations.

Theorem 2.[5] Let \(A\) be the infinitesimal generator of a \(C_0\)-semigroup \(\big(T(t)\big)_{t\geq0}\) and let \(\big(\gamma(t)\big)_{t\geq0}\) satisfy \((H_2)\). Then the resolvent operator \(\big(R(t)\big)_{t\geq0}\) for Equation \((5)\) is operator-norm continuous (or continuous in the uniform operator topology) for \(t>0\) if and only if \(\big(T(t)\big)_{t\geq0}\) is operator-norm continuous for \(t>0\). \label{normcontinuity}

Definition 2. Let \(u\in L^2(J,U)\) and \(\varphi\in\mathcal{P}\). A function \(x:\,]-\infty,b]\rightarrow X\) is called a mild solution of equation (4) if \(x(t)=\varphi(t)\ \ \text{for}\ \ t\in(-\infty,0],\ \Delta x(t_k)=I_k(x_{t_k}),\ \ k=1,2,\cdots,m\), the restriction of \(x\) to intervals \(J_k=(t_k,t_{k+1}]\ (k=0,\cdots,m)\) is continuous and the following integral equation is satisfied

\begin{equation} x(t)= R(t)\varphi(0)+\displaystyle{\int_0^tR(t-s)\left[f(s,x_s)+Cu(s)\right]\,ds}\ +\sum_{0< t_k< t}R(t-t_k)I_k(x_{t_k}) \ \text{for}\ t\in J\,. \label{eqn3} \end{equation}
(6)

Definition 3. Equation \((4)\) is said to be controllable on the interval \(J\) if for every \(\varphi\in\mathcal{P}\) and \( x_1\in X\), there exists a control \(u\in L^2(J,U)\) such that a mild solution \(x\) of Equation \((4)\) satisfies the condition \(x(b)=x_1\).

For proving the main result of the paper we recall some properties of the measure of noncompactness and the Mönch fixed-point theorem.

Definition 4. [19] Let \(D\) be a bounded subset of a normed space \(Y\). The Hausdorff measure of noncompactness ( shortly MNC) is defined by $$\beta(D)=\inf\Big\{\epsilon>0: D\ has\ a\ finite\ cover\ by\ balls\ of\ radius\ less\ than\ \epsilon\Big\}.$$

Theorem 3. [19] Let \(D,\ D_1,\ D_2\) be bounded subsets of a Banach space \(Y\). The Hausdorff MNC has the following properties:

  • (i) If \(D_1\subset D_2\), \ then \(\beta(D_1)\leq\beta(D_2)\), \ (monotonicity).
  • (ii) \(\beta(D)=\beta(\overline{D})\).
  • (iii) \(\beta(D)=0\) if and only if \(D\) is relatively compact.
  • (iv) \(\beta(\lambda D)=|\lambda|\beta(D)\) for any \(\lambda\in\mathbb{R}\), (Homogeneity).
  • (v) \(\beta(D_1+D_2)\leq\beta(D_1)+\beta(D_2)\), where \(D_1+D_2=\{d_1+d_2:\ d_1\in D_1,\ d_2\in D_2\}\), (subadditivity).
  • (vi) \(\beta(\{a\}\cup D)=\beta(D)\) for every \(a\in Y\).
  • (vii) \(\beta(D)=\beta(\overline{co}(D))\), where \(\overline{co}(D)\) is the closed convex hull of \(D\).
  • (viii) For any map \ \(G:\mathcal{D}(G)\subseteq X\rightarrow Y\) which is Lipschitz continuous with a Lipschitz constant \(k\), we have \(\beta(G(D))\ \leq\ k\beta(D),\) for any subset \(D\subseteq\mathcal{D}(G)\).

Let \(R_b=\displaystyle\sup_{t\in[0,b]}\|R(t)\|,\ \ K_b=\sup_{t\in[0,b]}\|K(t)\|,\ \ M_b=\sup_{t\in[0,b]}\|M(t)\|.\) We now state the following useful result for equicontinuous subsets of \(\mathcal{C}([a,b];X)\), where \(X\) is a Banach space. %The proof of the following Lemma 1 can be found in [19].

Lemma 1. [19] Let \(M\subset \mathcal{PC}([a,b];X)\) be bounded and piecewise equicontinuous on \([a,b]\). Then \(\beta(M(t))\) is piecewise continuous for \(t\in[a,b]\) and \(\beta(M)=\sup\{\beta(M(t));\,t\in [a,b]\},\ \ \ \text{where}\ M(t)=\{x(t);\,x\in M\}.\)

Lemma 2. [19] Let \(M\subset\mathcal{C}([a,b];X)\) be bounded and equicontinuous. Then the set \(\overline{co}(M)\) is also bounded and equicontinuous.

To prove the controllability for Equation (4), we need the following results.

Lemma 3.[4] If \((u_n)_{n\geq1}\) is a sequence of Bochner integrable functions from \(J\) into a Banach space \(Y\) with the estimation \(\|u_n(t)\|\leq\mu(t)\) for almost all \(t\in J\) and every \(n\geq1\), where \(\mu\in L^1(J,\mathbb{R})\), then the function \(\psi(t)\ =\ \beta(\{u_n(t):n\geq1\})\) belongs to \(L^1(J,\mathbb{R}^{+})\) and satisfies the following estimation \(\beta\left(\left\{\int_0^tu_n(s)ds\,:\ n\geq1\right\}\right)\ \leq\ 2\int_0^t\psi(s)ds.\)

We now state the following nonlinear alternative of Mönch's type for selfmaps, which we shall use in the proof of the controllability of Equation \((4)\).

Theorem 4. {[20](Mönch, 1980) Let \(\mathcal{K}\) be a closed and convex subset of a Banach space \(Z\) and \(0\in \mathcal{K}\). Assume that \(F:\mathcal{K}\rightarrow \mathcal{K}\) is a continuous map satisfying Mönch's condition, namely, \(D\subseteq \mathcal{K}\) be countable and \( D\subseteq \overline{co}(\{0\}\cup F(D))\) implies \( \overline{D}\) is compact. Then \(F\) has a fixed point.

3. Controllability result

In this Section, we give sufficient conditions ensuring the controllability of Equation (4). For that goal, we need to assume that;
  • \((H_3)\)]
    • (i) The following linear operator \(W\,:\, L^2(J,U)\rightarrow X\)\ defined by \(Wu\ =\ \int_0^bR(b-s)Cu(s)\,ds,\) is surjective so that it induces an isomorphism between \(L^2(J,U)\,/_{{\text Ker}W}\)\ and \ \(X\) again denoted by \(W\) with inverse \(W^{-1}\) taking values in \(L^2(J,U)\,/_{{\text Ker}W}\) [21].
    • (ii)There exists a function \(L_W\in L^1(J,\mathbb{R}^{+})\) such that for any bounded set \(Q\subset X\) we have \(\beta((W^{-1}Q)(t))\leq L_W(t)\beta(Q),\) where \(\beta\) is the Hausdorff MNC.
  • \((H_4)\) The function \(f\,:\, J\times \mathcal{P}\longrightarrow X\) satisfies the following two conditions;
    • (i) \(f(\cdot,\varphi)\) is measurable for \(\varphi\in \mathcal{P}\) and \(f(t,\cdot)\) is continuous for a.e \(t\in J\),
    • (ii)for every positive integer \(q\), there exists a function \(l_q\in L^1(J,\mathbb{R}^{+})\) such that \(\displaystyle\sup_{\|\varphi\|_{\mathcal{P}}\leq q}\|f(t,\varphi)\|\leq l_q(t)\) for a.e. \(t\in J\) and \(\liminf_{q\rightarrow+\infty}\int_0^b\frac{l_q(t)}{q}dt=l< +\infty,\)
    • (iii) there exists a function \(h\in L^1(J,\mathbb{R}^{+})\) such that for any bounded set \(D\subset\mathcal{P}\), \ \(\beta(f(t,D))\leq h(t)\sup_{-\infty< \theta\leq0}\beta(D(\theta))\) for a.e. \(t\in J,\) where \(D(\theta)=\{\phi(\theta):\,\phi\in D\}.\)
  • \((H_5)\) \(I_k:\mathcal{P}\rightarrow X,\ k=1,2,\cdots,m\) are continuous such that;
    • (i) There are nondecreasing functions \(L_k:\mathbb{R}^{+}\rightarrow \mathbb{R}^{+}\) such that \(\|I_k(x)\|\leq L_k(\|x\|_{\mathcal{P}}),\ \ k=1,2,\cdots,m,\ \ x\in\mathcal{P},\) and \(\liminf_{\rho\rightarrow+\infty}\frac{L_k(\rho)}{\rho}=\lambda_k< +\infty,\ \ k=1,2,\cdots,m.\)
    • (ii) There exist constants \(\alpha_k\geq0\) such that, \(\beta(I_k(D))\leq\alpha_k\sup_{-\infty< \theta\leq0}\beta(D(\theta)),\ \ k=1,2,\cdots,m,\) for every bounded subset \(D\) of \(\mathcal{P}\). \(\tau=\Big(1+2R_bM_2\|L_W\|_{L^1}\Big)\left(2R_b\|h\|_{L^1}+R_b\sum_{k=0}^m\alpha_k\right)< 1,\) where \(R_b=\displaystyle\sup_{0\leq t\leq b}\|R(t)\|\) and \(M_2\) is such that \(M_2=\|C\|\).

Theorem 5. Suppose that hypotheses \((H_3)-(H_5)\) hold and Equation (5) has a resolvent operator \(\big(R(t)\big)_{t\geq0}\) that is continuous in the operator-norm topology for \(t>0\). Then Equation \((4)\) is controllable on \(J\) provided that

\begin{equation} R_b(1+R_bM_2M_3b)K_b\left(l+\sum_{k=1}^{m}\lambda_k\right)< 1, \label{h5} \end{equation}
(6)
where \(M_3\) is such that \(M_3=\|W^{-1}\|\). \label{mainthm}

Proof Using \((H_3)\) and given an arbitrary function \(x\), we define the control as usual by the following formula; $$u_x(t)\ =\ W^{-1}\left\{x_1-R(b)\varphi(0)-\int_0^bR(b-s)f(s,x_s)\,ds-\sum_{0< t_k< t}R(b-t_k)I_k(x_{t_k})\right\}(t)\qquad \text{for}\ t\in I.$$ For each \(x\in \mathcal{PC}\) such that \(x(0)=\varphi(0)\), we define its extension \(\widetilde{x}\) from \(]-\infty,b]\) to \(X\) as follows \begin{equation*} \widetilde{x}(t)=\left\{ \begin{array}{l} x(t)\ \ \text{if}\ \ t\in[0,b],\\ \varphi(t) \ \ \text{if}\ \ t\in]-\infty,0]. \end{array} \right.%\eqno(2.2) \end{equation*} We define the space \(E_b=\Big\{x:]-\infty,b]\rightarrow X\ \text{such that}\ x|_{J}\in \mathcal{PC}\ \text{and}\ x_0\in\mathcal{P}\Big\},\) where where \(x|_{J}\) is the restriction of \(x\) to \(J\). We show, by using this control that the operator \(P:\,E_b\rightarrow E_b\) defined by \begin{equation*} (Px)(t)= R(t)\varphi(0)+\displaystyle{\int_0^tR(t-s)\big[f(s,\widetilde{x}_s)+Cu_x(s)\big]\,ds}+\sum_{0< t_k< t}R(t-t_k)I_k(x_{t_k})\, \ \ \text{for}\ \ t\in I=[0,b] \end{equation*} has a fixed-point. This fixed point is then a mild solution of Equation \((4)\). Observe that \((Px)(b)=x_1\). This means that the control \(u_x\) steers the integrodifferential equation from \(\varphi\) to \(x_1\) in time \(b\) which implies that the Equation \((4)\) is controllable on \(J\). For each \(\varphi\in\mathcal{P}\), we define the function \(y\in \mathcal{PC}\) by \(y(t)= R(t)\varphi(0)\) and its extension \(\widetilde{y}\) on \(]-\infty,0]\) by \begin{equation*} \widetilde{y}(t)=\left\{ \begin{array}{l} y(t)\ \ \text{if}\ \ t\in[0,b], \\ \varphi(t)\ \ \text{if}\ \ t\in]-\infty,0]. \end{array} \right. \end{equation*} For each \(z\in \mathcal{PC}\), set \(\widetilde{x}(t)=\widetilde{z}(t)+\widetilde{y}(t)\), where \(\widetilde{z}\) is the extension by zero of the function \(z\) on \(]-\infty,0]\). Observe that \(x\) satifies \((\ref{eqn3})\) if and only if \(z(0)=0\) and $$z(t)= \int_0^tR(t-s)\big[f(s,\widetilde{z}_s+\widetilde{y}_s)+Cu_z(s)\big]\,ds+\sum_{0< t_k< t}R(t-t_k)I_k(z_{t_k}+\widetilde{y}_{t_k})\ \ \text{for}\ t\in[0,b],$$ where \(u_z(t)\ =\ W^{-1}\left\{x_1-R(b)\varphi(0)-\int_0^bR(b-s)f(s,\widetilde{z}_s+\widetilde{y}_s)\,ds-\sum_{0< t_k< t}R(b-t_k)I_k(z_{t_k}+\widetilde{y}_{t_k})\right\}(t).\) Now let \(E_b^0=\Big\{z\in E_b\ \text{such that}\ z_0=0\Big\}.\) Thus \(E_b^0\) is a Banach space provided with the supremum norm. Define the operator \(\Gamma\, :\,E_b^0\rightarrow E_b^0\) by \begin{equation*} (\Gamma z)(t)= \displaystyle{\int_0^tR(t-s)\big[f(s,\widetilde{z}_s+\widetilde{y}_s)+Cu_z(s)\big]\,ds}+\sum_{0< t_k< t}R(t-t_k)I_k(z_{t_k}+\widetilde{y}_{t_k})\ \ \text{for}\ \ t\in[0,b]. \end{equation*} Note that the operator \(P\) has a fixed point if and only if \(\Gamma\) has one. So to prove that \(P\) has a fixed point, we only need to prove that \(\Gamma\) has one. For each positive number \(q\), let \(B_q=\{z\in E_b^0:\|z\|\leq q\}\). Then, for any \(z\in B_q\), we have by axiom \((A_1)\) that \begin{eqnarray} \|z_s+y_s\|&\leq &\|z_s\|_{\mathcal{P}}+\|y_s\|_{\mathcal{P}}\\ &\leq&K(s)\|z(s)\|+M(s)\|z_0\|_{\mathcal{P}}+K(s)\|y(s)\|+M(s)\|y_0\|_{\mathcal{P}}\\ &\leq&K_b\|z(s)\|+K_b\|R(t)\|\|\varphi(0)\|+M_b\|\varphi\|_{\mathcal{P}}\\ &\leq&K_b\|z(s)\|+K_bR_bH\|\varphi\|_{\mathcal{P}}+M_b\|\varphi\|_{\mathcal{P}}\\ &\leq&K_b\|z(s)\|+\Big(K_bR_bH+M_b\Big)\|\varphi\|_{\mathcal{P}}\\ &\leq&K_b\,q+\Big(K_bR_bH+M_b\Big)\|\varphi\|_{\mathcal{P}}. \end{eqnarray} Thus, \(\|z_s+y_s\|\leq K_b\,q+\Big(K_bR_bH+M_b\Big)\|\varphi\|_{\mathcal{P}}=:q'.\) We shall prove the theorem in the following steps;
Step 1. We claim that there exists \(q>0\) such that \(\Gamma(B_q)\subset B_q\). We proceed by contradiction. Assume that it is not true. Then for each positive number \(q\), there exists a function \(z^q\in B_q\), such that \(\Gamma(z^q)\notin B_q,\ i.e.,\ \|(\Gamma z^q)(t)\|>q\) for some \(t\in [0,b]\). Now we have that \begin{eqnarray} q &< & \Big\|(\Gamma z^q)(t)\Big\|\\ &\leq& R_b\int_0^b\Big\|f(s,\tilde{z}_s^q+\widetilde{y}_s)\Big\|\,ds+R_b\int_0^b\|Cu_{z^q}(s)\|\,ds+R_b\sum_{k=0}^mL_k(\|z_{t_k}+\widetilde{y}_{t_k}\|)\\ &\leq& R_b\int_0^b\Big\|f(s,\tilde{z}_s^q+\widetilde{y}_s)\Big\|\,ds+R_b\sum_{k=0}^mL_k(q')\\ && +R_b\int_0^b\Big\|BW^{-1}\Big[x_1-R(b)\varphi(0)-\int_0^bR(b-s)f(s,\tilde{z}_s^q)\,ds-\sum_{0< t_k< t}R(b-t_k)I_k(z_{t_k}+\widetilde{y}_{t_k})\Big]\Big\|\,ds\\ &\leq& bR_bM_2M_3\left(\|x_1\|+R_b\|\varphi(0)\|+R_b\int_0^b\|f(s,\tilde{z}_s^q)\|\,ds+R_b\sum_{k=0}^mL_k(q')\right)\\ && +R_b\int_0^b\Big\|f(s,\tilde{z}_s^q+\widetilde{y}_s)\Big\|\,ds+R_b\sum_{k=0}^mL_k(q')\\ &\leq& bR_bM_2M_3\left(\|x_1\|+R_bH\|\varphi\|_{\mathcal{B}}+R_b\int_0^bl_{q'}(s)\,ds+R_b\sum_{k=0}^mL_k(q')\right)+ R_b\int_0^bl_{q'}(s)\,ds+R_b\sum_{k=0}^mL_k(q'), \end{eqnarray} where \(q':=K_b\,q+q_0,\ \ \text{with}\ q_0:=\Big(K_bR_bH+M_b\Big)\|\varphi\|_{\mathcal{B}}.\) Hence $$q\leq \Big(1+R_bM_2M_3b\Big)\left(R_b\int_0^bl_{q'}(s)\,ds+R_b\sum_{k=0}^mL_k(q')\right)+R_bM_2M_3b\Big(\|x_1\|+R_bH\|\varphi\|_{\mathcal{B}}\Big).$$ Dividing both sides by \(q\) and noting that \(q'=K_bq+q_0\rightarrow+\infty\) as \(q\rightarrow+\infty\), we obtain that $$1\leq\Big(1+R_bM_2M_3b\Big)R_b\left(\frac{\displaystyle{\int_0^bl_{q'}(s)\,ds}+\sum_{k=0}^mL_k(q')}{q}\right)+\frac{R_bM_2M_3b\Big(\|x_1\|+R_bH\|\varphi\|_{\mathcal{B}}\Big)}{q}$$ and $$\liminf_{q\rightarrow+\infty}\left(\frac{\displaystyle{\int_0^bl_{q'}(s)\,ds}+\sum_{k=0}^mL_k(q')}{q}\right)=\liminf_{q\rightarrow+\infty}\left(\frac{\displaystyle{\int_0^bl_{q'}(s)\,ds}}{q'}+\frac{\sum_{k=0}^mL_k(q')}{q'}\right)\frac{q'}{q}=\left(l+\sum_{k=0}^m\lambda_k\right)K_b.$$ Thus we have, \(1\leq\Big(1+R_bM_2M_3b\Big)R_b\left(l+\sum_{k=0}^m\lambda_k\right)K_b\), and this contradicts \((7)\). Hence for some positive number \(q\), we must have \(\Gamma(B_q)\subset B_q\).
Step 2. \(\Gamma:\,E_b^0\rightarrow E_b^0\) is continuous. In fact let \(\Gamma:=\Gamma_1+\Gamma_2\), where $$(\Gamma_1z)(t)=\int_0^tR(t-s)f(s,\widetilde{z}_s+\widetilde{y}_s)\,ds+\sum_{k=0}^mR(t-t_k)I_k(z_{t_k}+\widetilde{y}_{t_k})\ \ \ \text{and}\ \ \ \ \ (\Gamma_2z)(t)=\int_0^tR(t-s)Cu_z(s)\,ds.$$ Let \(\{z^n\}_{n\geq1}\subset E_b^0\) with \(z^n\rightarrow z\) in \(E_b^0\). Then there exists a number \(q>1\) such that \(\|z^n(t)\|\leq q\) for all \(n\) and \(a.e.\ \,t\in J\). So \(z^n,\,z\in B_q\). By \((H_4)-(i),\,\, f(t,\tilde{z}_t^n+\widetilde{y}_t)\rightarrow f(t,\widetilde{z}_t+\widetilde{y}_t)\) for each \(t\in [0,b]\). Also, by \((H_5)-(i),\,\,I_k(z_{t_k}^n+\widetilde{y}_{t_k})\rightarrow I_k(z_{t_k}+\widetilde{y}_{t_k})\) for each \(t\in [0,b]\). And by \((H_4)-(ii)\), \(\|f(t,\tilde{z}_t^n+\widetilde{y}_t)-f(t,\widetilde{z}_t+\widetilde{y}_t)\|\leq 2l_{q'}(t).\) Then we have $$\|\Gamma_1z^n-\Gamma_1z\|_{\mathcal{P}}\leq R_b\int_0^b\|f(s,\tilde{z}_s^n+\widetilde{y}_s)-f(s,\widetilde{z}_s+\widetilde{y}_s)\|\,ds+R_b\sum_{k=0}^m\|I_k(z_{t_k}^n+\widetilde{y}_{t_k})-I_k(z_{t_k}+\widetilde{y}_{t_k})\|\longrightarrow0,\ as\ n\rightarrow+\infty$$ by dominated convergence Theorem. Also we have that $$\|\Gamma_2z^n-\Gamma_2z\|_{\mathcal{P}}\leq R_b^2M_2M_3b\left(\int_0^b\|f(s,\tilde{z}_s^n+\widetilde{y}_s)-f(s,\widetilde{z}_s+\widetilde{y}_s)\|\,ds+\sum_{k=0}^m\|I_k(z_{t_k}^n+\widetilde{y}_{t_k})-I_k(z_{t_k}+\widetilde{y}_{t_k})\right)\longrightarrow0,$$ by dominated convergence Theorem. Thus \(\|\Gamma z^n-\Gamma z\|\leq\|\Gamma_1z^n-\Gamma_1z\|+\|\Gamma_2z^n-\Gamma_2z\|\longrightarrow0,\ as\ n\rightarrow+\infty.\) Hence \(\Gamma\) is continuous on \(E_b^0\).
Step 3. \(\Gamma(B_q)\) is equicontinuous on \([0,b]\). In fact let \(t_1,\,t_2\in J_k,\ \ t_1< t_2\) and \(z\in B_q\), we have \begin{eqnarray} &&\|(\Gamma z)(t_2)-(\Gamma z)(t_1)\|\leq\int_0^{t_1}\|R(t_2-s)-R(t_1-s)\|\|f(s,\widetilde{z}_s+\widetilde{y}_s)+Cu_z(s)\|\,ds\\ && +\sum_{0< t_k< t_1}\|R(t_2-t_k)-R(t_1-t_k)\|\|I_k(z_{t_k}+\widetilde{y}_{t_k})\|+\sum_{t_1\leq t_k< t_2}\|R(t_1-t_k)\|\|I_k(z_{t_k}+\widetilde{y}_{t_k})\|\\ && +\int_{t_1}^{t_2}\|R(t_2-s)\|\|f(s,\widetilde{z}_s+\widetilde{y}_s)+Cu_z(s)\|\,ds\\ &&\leq\int_0^{t_1}\|R(t_2-s)-R(t_1-s)\|l_{q'}(s)\,ds\\ && +\int_0^{t_1}\|R(t_2-s)-R(t_1-s)\|M_2M_3\left(\|x_1\|+R_bH\|\varphi\|_{\mathcal{B}}+R_b\int_0^bl_{q'}(\tau)\,d\tau+\sum_{k=0}^mL_k(q')\right)\,ds\\ && +\sum_{0< t_k< t_1}\|R(t_2-t_k)-R(t_1-t_k)\|L_k(q')+R_b\sum_{t_1\leq t_k< t_2}L_k(q')+\int_{t_1}^{t_2}\|R(t_2-s)\|l_{q'}(s)\,ds\\ && +\int_{t_1}^{t_2}\|R(t_2-s)\|M_2M_3\left(\|x_1\|+R_bH\|\varphi\|_{\mathcal{B}}+R_b\int_0^bl_{q'}(\tau)\,d\tau+\sum_{k=0}^mL_k(q')\right)\,ds. \end{eqnarray} By the continuity of \(\big(R(t)\big)_{t\geq0}\) in the operator-norm toplogy, the dominated convergence Theorem, we conclude that the right hand side of the above inequality tends to zero and independent of \(z\) as \(t_2\rightarrow t_1\). Hence \(\Gamma(B_q)\) is equicontinuous on \(J\).
Step 4. We show that the Mönch's condition holds. Suppose that \(D\subseteq B_q\) is countable and \(D\subseteq\overline{co}(\{0\}\cup \Gamma(D))\). We shall show that \(\beta(D)=0\), where \(\beta\) is the Hausdorff MNC. Without loss of generality, we may assume that \(D=\{z^n\}_{n\geq1}\). Since \(\Gamma\) maps \(B_q\) into an equicontinuous family, \(\Gamma(D)\) is also equicontinuous on \(J\). By \((H_3)-(ii)\), \((H_4)-(iii)\) and Lemma 3, we have that \begin{align*} & \beta\Big(\{u_{z^n}(t)\}_{n\geq1}\Big) = \beta\left(W^{-1}\left\{x_1-R(b)\varphi(0)-\int_0^bR(t-b)f\Big(s,\{\tilde{z}_s^n+\widetilde{y}_s\}_{n\geq1}\Big)\,ds-\sum_{0< t_k< t}R(b-t_k)I_k(z_{t_k}^n+\widetilde{y}_{t_k})\right\}\right)\\ &\leq L_W(t)\beta\left(\left\{x_1-R(b)\varphi(0)\right\}\right)+L_W(t)\beta\left(\left\{\int_0^bR(t-b)f\Big(s,\{\tilde{z}_s^n+\widetilde{y}_s\}_{n\geq1}\Big)\,ds\right\}_{n\geq1}\right) \end{align*} \begin{align*} &\;\;\;+L_W(t)\beta\left(\left\{\sum_{0< t_k< t}R(b-t_k)I_k(z_{t_k}^n+\widetilde{y}_{t_k})\right\}_{n\geq1}\right)\\ &\leq 2R_bL_W(t)\left(\int_0^bh(s)\beta\left(\left\{\tilde{z}_s^n\right\}_{n\geq1}+\{\widetilde{y}_s\}\right)ds\right)+R_bL_W(t)\sum_{k=0}^m\beta\left(\left\{I_k(z_{t_k}^n+\widetilde{y}_{t_k})\right\}_{n\geq1}\right)\\ &\leq 2R_bL_W(t)\left(\int_0^bh(s)\Big[\beta\left(\left\{\tilde{z}_s^n\right\}_{n\geq1}\right)+\beta\Big(\{\widetilde{y}_s\}\Big)\Big]\,ds\right)+R_bL_W(t)\sum_{k=0}^m\alpha_k\sup_{-\infty< \theta\leq0}\beta\left(\left\{z_{t_k}^n+\widetilde{y}_{t_k}\right\}_{n\geq1}\right)\\ &\leq 2R_bL_W(t)\left(\int_0^bh(s)\beta\left(\left\{\tilde{z}_s^n\right\}_{n\geq1}\right)\,ds\right)+R_bL_W(t)\sum_{k=0}^m\alpha_k\sup_{-\infty< \theta\leq0}\beta\left(\left\{z_{t_k}^n\right\}_{n\geq1}\right), \end{align*} since \(\Big\{\widetilde{y}_s:\ s\in[0,b]\Big\}\)is compact, so \begin{align*} &\leq 2R_bL_W(t)\left(\int_0^bh(s)\displaystyle\sup_{-\infty< \theta\leq0}\beta\left(\left\{\tilde{z}_s^n(\theta)\right\}_{n\geq1}\right)\,ds\right)+R_bL_W(t)\sum_{k=0}^m\alpha_k\sup_{-\infty< \theta\leq0}\beta\left(\left\{z_{t_k}^n\right\}_{n\geq1}\right), \end{align*} by Lemma 1, since \(D=\{z^n\}_{n\geq1}\) is equicontinuous, we obtain \begin{align*} &\leq 2R_bL_W(t)\left(\int_0^bh(s)\,ds\right)\displaystyle\sup_{0\leq t\leq b}\beta\left(\left\{z^n(t)\right\}_{n\geq1}\right)+R_bL_W(t)\sum_{k=0}^m\alpha_k\sup_{0\leq \tau_k\leq t_k}\beta\left(\left\{z^n(\tau_k)\right\}_{n\geq1}\right). \end{align*} This implies that \begin{align*} &\beta\Big(\{(\Gamma z^n)(t)\}_{n\geq1}\Big)\leq \beta\left(\left\{\int_0^tR(t-s)f(s,\{\tilde{z}_s^n+\widetilde{y}_s\}_{n\geq1})\,ds\right\}_{n\geq1}\right)+\beta\left(\left\{\int_0^tR(t-s)u_{z^n}(s)\,ds\right\}_{n\geq1}\right)\\ &\;\;\;+\beta\left(\left\{\sum_{0< t_k< t}R(b-t_k)I_k(z_{t_k}^n+\widetilde{y}_{t_k})\right\}_{n\geq1}\right)\\ &\leq2R_b\left(\int_0^bh(s)\,ds\right)\displaystyle\sup_{0\leq t\leq b}\beta\left(\left\{z^n(t)\right\}_{n\geq1}\right)+R_b\sum_{k=0}^m\alpha_k\sup_{0\leq \tau_k\leq t_k}\beta\left(\left\{z^n(\tau_k)\right\}_{n\geq1}\right)\\ & \;\;\;+2R_bM_2\left(\int_0^bL_W(s)\,ds\right)2R_b\left(\int_0^bh(s)\,ds\right)\displaystyle\sup_{0\leq t\leq b}\beta\left(\left\{z^n(t)\right\}_{n\geq1}\right)\\&\;\;\;+2R_b^2M_2\left(\int_0^bL_W(s)\,ds\right)\sum_{k=0}^m\alpha_k\sup_{0\leq \tau_k\leq t_k}\beta\left(\left\{z^n(\tau_k)\right\}_{n\geq1}\right)\\ &\leq 2R_b\|h\|_{L^1}\displaystyle\sup_{0\leq t\leq b}\beta\left(\left\{z^n(t)\right\}_{n\geq1}\right)+R_b\sum_{k=0}^m\alpha_k\sup_{0\leq \tau_k\leq t_k}\beta\left(\left\{z^n(\tau_k)\right\}_{n\geq1}\right)\\&\;\;\; +2R_bM_2\|L_W\|_{L^1}2R_b\|h\|_{L^1}\displaystyle\sup_{0\leq t\leq b}\beta\left(\left\{z^n(t)\right\}_{n\geq1}\right)+2R_b^2M_2\|L_W\|_{L^1}\sum_{k=0}^m\alpha_k\sup_{0\leq \tau_k\leq t_k}\beta\left(\left\{z^n(\tau_k)\right\}_{n\geq1}\right). \end{align*} It follows that \begin{align*} & \beta\Big(\Gamma (D)(t)\Big) \leq 2R_b\|h\|_{L^1}\displaystyle\sup_{0\leq t\leq b}\beta\left(D(t)\right)+R_b\sum_{k=0}^m\alpha_k\sup_{0\leq t\leq b}\beta\left(D(t)\right) +2R_bM_2\|L_W\|_{L^1}2R_b\|h\|_{L^1}\displaystyle\sup_{0\leq t\leq b}\beta\left(D(t)\right)\\ & \;\;\;+2R_b^2M_2\|L_W\|_{L^1}\sum_{k=0}^m\alpha_k\sup_{0\leq t\leq b}\beta\left(D(t)\right)\\ &\leq \left(2R_b\|h\|_{L^1}+R_b\sum_{k=0}^m\alpha_k+2R_bM_2\|L_W\|_{L^1}2R_b\|h\|_{L^1}+2R_b^2M_2\|L_W\|_{L^1}\sum_{k=0}^m\alpha_k\right)\sup_{0\leq t\leq b}\beta\left(D(t)\right)\\ &\leq \Big(1+2R_bM_2\|L_W\|_{L^1}\Big)\left(2R_b\|h\|_{L^1}+R_b\sum_{k=0}^m\alpha_k\right)\sup_{0\leq t\leq b}\beta\left(D(t)\right). \end{align*} Since \(D\) and \(\Gamma(D)\) are equicontinuous on \([0,b]\) and \(D\) is bounded, it follows by Lemma 1 that \(\beta\Big(\Gamma(D)\Big)\leq \tau\beta\Big(D\Big)\), where \(\tau\) is as defined in \((H_5)\). Thus from the Mönch condition, we get that \(\beta\Big(D\Big)\leq\beta\Big(\overline{co}(\{0\}\cup \Gamma(D)\Big)=\beta\Big(\Gamma(D)\Big)\leq \tau\beta\Big(D\Big),\) and since \(\tau< 1\), this implies \(\beta\Big(D\Big)=0\), which implies that \(D\) is relatively compact as desired in \(B_q\) and the Mönch condition is satisfied. We conclude by Theorem 4, that \(\Gamma\) has a fixed point \(z\) in \(B_q\). Then \(x=z+y\) is a fixed point of \(P\) in \(E_b\) and thus equation \((4)\) is controllable on \([0,b]\).

Numerical example

Now, we illustrate our main result by the following example.

Example 1. Consider the partial functional integrodifferential system of the form

\begin{equation}\label{exampleequation} \begin{cases} \displaystyle\frac{\partial v}{\partial t}(t,\xi) = \displaystyle\frac{\partial^2 v}{\partial \xi^2}(t,\xi) +{\displaystyle\int_0^t \zeta'(t-s)\frac{\partial^2 v}{\partial \xi^2}(s,\xi)\,ds}&\\\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+\displaystyle\int_{-\infty}^0\alpha(\theta)g(t,v(t+\theta,\xi))\,d\theta+\eta u(t,\xi) & \text{for}\ t\in J=[0,b]\ \text{and}\ \xi\in(0,\pi)\\ v(t,0)= 0 = v(t,\pi)\ &\text{for}\ t\in[0,b], \\ v(t_k^{+},\xi)-v(t_k^{-},\xi)=\displaystyle\int_{-\infty}^{t_k}\mu_k(t_k-s)v(s,\xi)ds,\ & \xi\in(0,\pi),\ k=1,2,\cdots,m,\\ v(\theta,\xi) = \phi(\theta,\xi)\ &\ \text{for}\ \theta\in]-\infty,0] \text{and}\ \xi\in(0,\pi),\ \end{cases} \end{equation}
(8)
where \(\ \eta>0,\ \phi\in\mathcal{P},\ I_k>0,\ k=1,2,\cdots,m\), \(u\in L^2((0,\pi))\), \(g:\,[0,1]\times\mathbb{R}\rightarrow\mathbb{R}\) is continuous and Lipschitzian with respect to the second variable, the initial data function \(\phi:\,\mathbb{R}^{-}\times\Omega\rightarrow\mathbb{R}\) is a given function, \(\alpha:\,\mathbb{R}^{-}\rightarrow\mathbb{R}\) is continuous, \(\alpha\in L^1(\mathbb{R}^{-},\mathbb{R})\) and \(\zeta\in \mathcal{C}^{2}([0,b])\) and \(\zeta(0)>0\). Let \(X=L^2(0,\pi)\), and the phase space \(\mathcal{P}= \mathcal{PC}_0\times L^2(\tilde{h},X)\) (\(\tilde{h}:]-\infty,-r]\rightarrow\mathbb{R}\) be a positive function), as introduced in [22]. We define \(A:\mathcal{D}(A)\subset X\rightarrow X\) by $$ \left\{ \begin{array}{l} \mathcal{D}(A)=\Big\{v\in X:\,v\ \text{and}\ v'\ \text{are absolutely continuous},\ v''\in X,\ v(0)=v(\pi)=0\Big\}\\ Av\,=\, v''\ \ \text{for each}\ v\in\mathcal{D}(A). \end{array} \right.%\eqno(4.1) $$ Then, \(Av=\sum_{n=1}^{\infty}n^2\langle v,\,v_n\rangle v,\ \ v\in\mathcal{D}(A),\) where \(v_n(s)=\sqrt{2/\pi}\sin(ns),\ \ n=1,2,3,\cdots\) is the orthogonal set of eigenvectors of \(A\). It is well known that \(A\) is the infinitesimal generator of an analytic semigroup \(\big(T(t)\big)_{t\geq0}\) in \(X\) as is given by \(T(t)v=\sum_{n=1}^{\infty}exp(-n^2t)\langle v,\,v_n\rangle v,\ \ v\in X.\) Moreover, \(\big(T(t)\big)_{t\geq0}\) generated by \(A\) above, is compact for \(t>0\) and operator-norm continuous for \(t>0\). Then by Theorem 2, the corresponding resolvent operator is operator-norm continuous. Now define $$x(t)(\xi)\ =\ v(t,\xi),\ \ \ x'(t)(\xi)\ =\ \frac{\partial v(t,\xi)}{\partial t},\ \ \ u(t,\xi)=u(t)(\xi), \ \ \varphi(\theta)(\xi)=\phi(\theta,\xi)\ \ \text{for}\ \ \theta\in ]-\infty,0]\ \text{and}\ \xi\in(0,\pi).$$ $$I_k(\varphi)(\xi)= \displaystyle\int_{-\infty}^{0}\mu_k(-s)\varphi(s,\xi)ds,\ \ \xi\in(0,\pi),\ k=1,2,\cdots,m.$$ $$f(t,\psi)(\xi)\ =\ \int_{-\infty}^0 \alpha(\theta)g(t,\psi(\theta)(\xi))\,d\theta\ \ \text{for}\ \theta\in ]-\infty,0]\ \text{and}\ \xi\in(0,\pi).$$ Now \(C:X\rightarrow X\) be defined by \(\Big(Cu(t)\Big)(\xi)\,=\,Cu(t)(\xi)=\,\eta u1_{\Gamma}(t,\xi).\) $$(\gamma(t)x)(\xi)\ =\ \zeta(t)\Delta v(t,\xi)\ \ \text{for}\ \ t\in [0,b],\ \ x\in\mathcal{D}(A)\ \text{and}\ \ \xi\in(0,\pi).$$ We suppose that \(\varphi\in \mathcal{P}\). Then, Equation \((8)\) is then transformed into the following form
\begin{equation}\label{abstractequation} \begin{cases} x'(t)=Ax(t)\ +\ {\displaystyle\int_0^t\gamma(t-s)x(s)ds}\ +\ f(t,x_t)+Cu(t)&\text{for}\ t\in J=[0,b], \\ \Delta x(t_k)=I_k(x_{t_k}), & k=1,2,\cdots,m, \\ x_0=\varphi\in\mathcal{P}. & \end{cases} \end{equation}
(9)
Suppose there exists a continuous function \(p\in L^1(J;\mathbb{R}^{+})\) such that \(|g(t,y_1)-g(t,y_2)|\leq p(t)|y_1-y_2|\ \ \text{for}\ t\in J\ \text{and}\ y_1,\,y_2\in\mathbb{R}\) and \(g(t,0)=0 \ \ \text{for}\ t\in J.\) One can see that \(f\) is Lipschitz continuous with respect to the second variable and moreover for \(\varphi\in\mathcal{P}\), we have we have \( \displaystyle\sup_{\|\varphi\|_{\mathcal{B}}\leq q}\Big\|f(t,\varphi)\Big\|\leq q\,\|\alpha\|\,p(t).\) So \(f\) satisfies \((H_4)-(i)\ \text{and}\ (H_4)-(ii)\) with \(l_q(t)\,=\,q\,\|\alpha\|\,p(t)\). Also \(f\) satisfies \( (H_4)-(iii)\) by condition \( (viii)\) of Theorem 3, since \(f\) is Lipschitz. Now for \(\xi\in(0,\pi)\), the operator \(W\) is given by \((Wu)(\xi)=\eta\int_0^bR(b-s)u(s,\xi)\,ds.\) Assuming that \(W\) satisfies \((H_3)\), then all the conditions of Theorem 5 hold and Equation \((9)\) is controllable.

Conclusion

In this work, we have shown the controllability of some impulsive partial functional integrodifferential differential equation with infinite delay in Banach spaces by using the Hausdorff Measure of Noncompactness and the Mönch fixed point theorem. We achieved this without assuming the compactness of the resolvent operator for the associated undelayed part.

Author Contributions

All authors contributed equally to the writing of this paper. All authors read and approved the final manuscript.

Competing Interests

The author(s) do not have any competing interests in the manuscript.

References

  1. Benchohra, M., Henderson, J., & Ntouyas, S. (2006). Impulsive differential equations and inclusions (Vol. 2). New York: Hindawi Publishing Corporation.[Google Scholor]
  2. Balachandran, K., & Sakthivel, R. (2001). Controllability of functional semilinear integrodifferential systems in Banach spaces. Journal of Mathematical Analysis and Applications, 255(2), 447-457.[Google Scholor]
  3. Chang, Y. K., Nieto, J. J., & Li, W. S. (2009). Controllability of semilinear differential systems with nonlocal initial conditions in Banach spaces. Journal of Optimization Theory and Applications, 142(2), 267-273.[Google Scholor]
  4. Wang, J., Fan, Z., & Zhou, Y. (2012). Nonlocal controllability of semilinear dynamic systems with fractional derivative in Banach spaces. Journal of Optimization Theory and Applications, 154(1), 292-302.[Google Scholor]
  5. Ezzinbi, K., Degla, G., & Ndambomve, P. (2015). Controllability for some partial functional integrodifferential equations with nonlocal conditions in Banach spaces. Discussiones Mathematicae, Differential Inclusions, Control and Optimization, 35(1), 25-46.[Google Scholor]
  6. Machado, J. A., Ravichandran, C., Rivero, M., & Trujillo, J. J. (2013). Controllability results for impulsive mixed-type functional integro-differential evolution equations with nonlocal conditions. Fixed Point Theory and Applications, 2013(1), 66.[Google Scholor]
  7. Li, M., Wang, M., & Zhang, F. (2006). Controllability of impulsive functional differential systems in Banach spaces. Chaos, Solitons & Fractals, 29(1), 175-181.[Google Scholor]
  8. Atmania, R., & Mazouzi, S. (2005). Controllability of semilinear integrodifferential equations with nonlocal conditions. Electronic Journal of Differential Equations(EJDE), 2005.[Google Scholor]
  9. Sakthivel, R., Mahmudov, N. I., & Nieto, J. J. (2012). Controllability for a class of fractional-order neutral evolution control systems. Applied Mathematics and Computation, 218(20), 10334-10340.[Google Scholor]
  10. Baghli, S., Benchohra, M., & Ezzinbi, K. (2009). Controllability results for semilinear functional and neutral functional evolution equations with infinite delay. Surveys in Mathematics and its Applications, 4, 15-39.[Google Scholor]
  11. Selvi, S., & Arjunan, M. M. (2012). Controllability results for impulsive differential systems with finite delay.Journal of Nonlinear Sciences and Applications, 5(3), 206-219.[Google Scholor]
  12. Hernández, E., & Henriquez, H. R. (1998). Existence results for partial neutral functional differential equations with unbounded delay. Journal of Mathematical Analysis and Applications, 221(2), 452-475.[Google Scholor]
  13. Chang, Y. K. (2007). Controllability of impulsive functional differential systems with infinite delay in Banach spaces. Chaos, Solitons & Fractals, 33(5), 1601-1609.[Google Scholor]
  14. Hino, Y., Murakami, S., & Naito, T. (1991). Functional differential equations with infinite delay. Lecture Notes in Mathematics, Springer-Verlag, Berlin.[Google Scholor]
  15. Grimmer, R. C. (1982). Resolvent operators for integral equations in a Banach space. Transactions of the American Mathematical Society, 273(1), 333-349.[Google Scholor]
  16. Desch, W., Grimmer, R., & Schappacher, W. (1984). Some considerations for linear integrodifferential equations. Journal of Mathematical analysis and Applications, 104(1), 219-234.[Google Scholor]
  17. Liang, J., Liu, J. H., & Xiao, T. J. (2008). Nonlocal problems for integrodifferential equations. Dynamics of Continuous, Discrete & Impulsive Systems. Series A, 15(6), 815-824.[Google Scholor]
  18. Ezzinbi, K., Toure, H., & Zabsonre, I. (2009). Existence and regularity of solutions for some partial functional integrodifferential equations in Banach spaces. Nonlinear Analysis: Theory, Methods & Applications, 70(7), 2761-2771.[Google Scholor]
  19. Banas, J. (1980). On measures of noncompactness in Banach spaces. Commentationes Mathematicae Universitatis Carolinae, 21(1), 131-143.[Google Scholor]
  20. Mönch, H. (1980). Boundary value problems for nonlinear ordinary differential equations of second order in Banach spaces. Nonlinear Analysis: Theory, Methods & Applications, 4(5), 985-999.[Google Scholor]
  21. Quinn, M. D., & Carmichael, N. (1985). An approach to non-linear control problems using fixed-point methods, degree theory and pseudo-inverses. Numerical Functional Analysis and Optimization, 7(2-3), 197-219.[Google Scholor]
  22. Hernández, E., Rabello, M., & Henríquez, H. R. (2007). Existence of solutions for impulsive partial neutral functional differential equations. Journal of Mathematical Analysis and Applications, 331(2), 1135-1158.[Google Scholor]
]]>
On hyper-singular integrals https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/on-hyper-singular-integrals/ Thu, 22 Oct 2020 13:29:44 +0000 https://old.pisrt.org/?p=4581
OMA-Vol. 4 (2020), Issue 2, pp. 100 - 103 Open Access Full-Text PDF
Alexander G. Ramm
Abstract: The integrals \(\int_{-\infty}^\infty t_+^{\lambda-1} \phi(t)dt\) and \(\int_0^t(t-s)^{\lambda -1}b(s)ds\) are considered, \(\lambda\neq 0,-1,-2...\), where \(\phi\in C^\infty_0(\mathbb{R})\) and \(0\le b(s)\in L^2_{loc}(\mathbb{R})\). These integrals are defined in this paper for \(\lambda\le 0\), \(\lambda\neq 0,-1,-2,...\), although they diverge classically for \(\lambda\le 0\). Integral equations and inequalities are considered with the kernel \((t-s)^{\lambda -1}_+\).
]]>

Open Journal of Mathematical Analysis

On hyper-singular integrals

Alexander G. Ramm
Department of Mathematics, Kansas State University, Manhattan, KS 66506, USA.; ramm@math.ksu.edu

Abstract

The integrals \(\int_{-\infty}^\infty t_+^{\lambda-1} \phi(t)dt\) and \(\int_0^t(t-s)^{\lambda -1}b(s)ds\) are considered, \(\lambda\neq 0,-1,-2…\), where \(\phi\in C^\infty_0(\mathbb{R})\) and \(0\le b(s)\in L^2_{loc}(\mathbb{R})\). These integrals are defined in this paper for \(\lambda\le 0\), \(\lambda\neq 0,-1,-2,…\), although they diverge classically for \(\lambda\le 0\). Integral equations and inequalities are considered with the kernel \((t-s)^{\lambda -1}_+\).

Keywords:

Hyper-singular integrals.

1. Introduction

In [1] the following integral equation is of interest;

\begin{equation} \label{e1} b(t)=b_0(t)+\int_0^t (t-s)^{\lambda -1}b(s)ds, \end{equation}
(1)
where \(b_0\) is a smooth functions rapidly decaying with all its derivatives as \(t\to \infty\), \(b_0(t)=0\) if \(t< 0\). We are especially interested in the value \(\lambda=-\frac 1 4\), because of its importance for the Navier-Stokes theory, [1], Chapter 5, [2,3]. The integral in (1) diverges in the classical sense for \(\lambda\le 0\). Our aim is to define this hyper-singular integral. There is a regularization method to define singular integrals \(J:=\int_{\mathbb{R}}t_+^{\lambda}\phi(t)dt\), \(\lambda< -1\), in distribution theory, [4]. However, the integral in (1) is a convolution, which is defined in [4], p.135, as a direct product of two distributions. This definition is not suitable for our purposes because although \(t_+^{\lambda-1}\) for \(\lambda\le 0\), \(\lambda\neq 0,-1,-2,...\) is a distribution on the space \(C^\infty_0(\mathbb{R}_+)\) of the test functions, but it is not a distribution in the space \(K=C^\infty_0(\mathbb{R})\) of the test functions used in [4]. Indeed, one can find \(\phi\in K\) such that \(\lim_{n\to \infty}\phi_n=\phi\) in \(K\), but \(\lim_{n\to \infty}\int_{\mathbb{R}} t_+^{\lambda-1} \phi(t)dt=\infty\) for \(\lambda\le 0\), so that \(t_+^{\lambda-1}\) is not a linear bounded functional in \(K\), i.e., not a distribution. On the other hand, one can check that \(t_+^{\lambda-1}\) for \(\lambda\in R\) is a distribution (a bounded linear functional) in the space \(\mathcal{K}=C^\infty_0(\mathbb{R}_+)\) with the convergence \(\phi_n\to \phi\) in \(\mathcal{K}\) defined by the requirements: a) the supports of all \(\phi_n\) belong to an interval \([a,b]\), \(0< a\le b< \infty\), b) \(\phi_n^{(j)}\to \phi^{(j)}\) in \(C([a,b])\) for all \(j=0,1,2,....\). Indeed, the functional \(\int_0^\infty t_+^\lambda\phi(t)dt\) is linear and bounded in \(\mathcal{K}\): \[ \left|\int_0^\infty t_+^\lambda\phi_n(t)dt\right|\le (a^\lambda+b^\lambda) \int_a^b |\phi_n(t)|dt. \] A similar estimate holds for the derivatives of \(\phi_n\). Although \(t_+^{-\frac 5 4}\) is a distribution in \(\mathcal{K}\), the convolution
\begin{equation} \label{e2} h:=\int_0^t(t-s)^{-\frac 5 4}b(s)ds:= t_+^{-\frac 5 4}\star b \end{equation}
(2)
cannot be defined similarly to the definition in [4] because the function \(\int_0^\infty \phi(u+s) b(s)ds\) does not, in general, belong to \(\mathcal{K}\) even if \(\phi\in \mathcal{K}\).

Let us define the convolution \(h\) using the Laplace transform

\[L(b):=\int_0^\infty e^{-pt}b(t)dt, \quad Re p>0.\] Laplace transform for distributions is studied in [5]. One has \(L(t_+^{-\frac 5 4}\star b)=L(t_+^{-\frac 5 4})L(b)\). To define \(L(t^{\lambda-1})\) for \(\lambda\le 0\), note that for Re\(\lambda>0\) the classical definition
\begin{equation} \label{e3} \int_0^\infty e^{-pt}t^{\lambda-1}dt= \frac{\Gamma(\lambda)}{p^\lambda} \end{equation}
(3)
holds. The right-side of (3) admits analytic continuation to the complex plane of \(\lambda\), \(\lambda\neq 0,-1,-2,....\). This allows one to define integral (3) for any \(\lambda\neq 0,-1,-2,...\). Recall that the gamma function \(\Gamma(\lambda)\) has its only singular points, the simple poles, at \(\lambda=-n\), \(n=0,1,2,...\) with the residue at \(\lambda=-n\) equal to \(\frac{(-1)^n}{n!}\). It is known that \(\Gamma(z+1)=z\Gamma(z)\), so
\begin{equation} \label{e3a} \Gamma(-\frac 1 4)=-4\Gamma(3/4):=-c_1, \quad c_1>0. \end{equation}
(4)
Therefore, we define \(h\) by defining \(L(h)\) as follows:
\begin{equation} \label{e4} L(h)=-c_1p^{\frac 1 4}L(b), \quad \lambda=-\frac 1 4, \end{equation}
(5)
and assume that \(L(b)\) can be defined. That \(L(b)\) is well defined in the Navier-Stokes theory follows from the a priori estimates proved in [1], Chapter 5. From (5) one gets
\begin{equation} \label{e4a} L(b)=-c_1^{-1}p^{-\frac 1 4} L(h). \end{equation}
(6)

2. Convolution of special functions

Define \(\Phi_\lambda=\frac {t_+^{\lambda-1}}{\Gamma(\lambda)}\).

Lemma 1. For any \(\lambda, \mu \in \mathbb{R}\) the following formulas hold;

\begin{equation} \label{e5} \Phi_\lambda\star \Phi_\mu=\Phi_{\lambda+\mu}, \quad \Phi_{\lambda +0}\star \Phi_{-\lambda}=\delta(t). \end{equation}
(7)

Proof. For Re\(\lambda>0\), Re\(\mu>0\) one has

\begin{equation} \label{e6} \Phi_\lambda\star \Phi_\mu= \frac 1 {\Gamma(\lambda)\Gamma(\mu)} \int_0^t(t-s)^{\lambda -1}s^{\mu -1}ds =\frac{t_+^{\lambda+\mu-1}}{\Gamma(\lambda)\Gamma(\mu)}\int_0^1 (1-u)^{\lambda-1}u^{\mu -1}du=\frac{t_+^{\lambda+\mu -1}}{\Gamma(\lambda+\mu)}, \end{equation}
(8)
where we used the known formula for beta function: \[B(\lambda, \mu):=\int_0^1u^{\lambda -1}(1-u)^{\mu -1}du= \frac{\Gamma(\lambda)\Gamma(\mu)} {\Gamma(\lambda+\mu)}. \] Analytic properties of beta function follow from these of Gamma function. The function \(\frac 1 {\Gamma(z)}\) is entire function of \(z\).

Let us now prove the second formula (7). We have \(\Gamma(\epsilon)\sim \epsilon\) as \(\epsilon \to 0\). Therefore

\begin{equation} \label{e6'} \frac {t_+^{\lambda+\epsilon -\lambda -1}}{\Gamma(\epsilon)}\sim \epsilon t_+^{\epsilon-1}. \end{equation}
(9)
If \(f\) is any continuous rapidly decaying function then
\begin{equation} \label{e6a} \lim_{\epsilon\to 0}\epsilon\int_0^\infty t^{\epsilon-1}f(t)dt=f(0). \end{equation}
(10)
Indeed, fix a small \(\delta>0\), such that \(f(t)\sim f(0)\) for \(t\in [0,\delta]\) as \(\delta\to 0\). Then, as \(\epsilon \to 0\), one has
\begin{equation} \label{e6b} \lim_{\epsilon\to +0}\epsilon\int_0^\delta t^{\epsilon-1}f(t)dt=\lim_{\epsilon\to +0} \epsilon f(0)\frac {t^\epsilon}{\epsilon}|_0^\delta=f(0)\lim_{\epsilon\to +0}\delta^\epsilon=f(0). \end{equation}
(11)
Note that
\begin{equation} \label{e6c} \lim_{\epsilon\to 0} \epsilon\int_\delta^\infty t^{\epsilon-1}f(t)dt=0, \quad \delta>0, \end{equation}
(12)
because \(|\int_\delta^\infty t^{\epsilon-1}f(t)dt|\le c\) and \(\epsilon\to 0\). From (11) and (12) one obtains (10). So, the second formula (7) is proved. Lemma 1 is proved.

Remark 1. The first formula (7) of Lemma 1 is proved in [4], pp.150-151. Our proof of the second formula (7) differs from the proof in [4] considerably.

Remark 2. A different proof of Lemma 1 can be given: \(L(\Phi_{\lambda}\star\Phi_{\mu})=\frac 1 {p^{\lambda+\mu}}\) by formula (3), and \(L^{-1}(\frac 1 {p^{\lambda+\mu}})=\Phi_{\lambda+\mu}(t)\). If \(\lambda=-\mu\), then \( \frac 1 {p^{\lambda+\mu}}=1\) and \(L^{-1}(1)=\delta(t)\).

3. Integral equation and inequality

Consider equation (1) and the following inequality:
\begin{equation} \label{e7} q(t)\le b_0(t)+t_+^{\lambda-1}\star q, \quad q\ge 0. \end{equation}
(13)

Theorem 1. Equation (1) has a unique solution. This solution can be obtained by iterations by solving the Volterra equation

\begin{equation} \label{e8} b_{n+1}=-c_1^{-1}\Phi_{1/4}\star b_{n} +c_1^{-1}\Phi_{1/4}\star b_0, \quad b_{n=0}= c_1^{-1}\Phi_{1/4}\star b_0, \quad b=\lim_{n\to \infty}b_n. \end{equation}
(14)

Proof. Applying to (1) the operator \(\Phi_{1/4}\star\) and using the second equation (7) one gets a Volterra equation \[\Phi_{1/4}\star b=\Phi_{1/4}\star b_0-c_1b,\quad c_1=|\Gamma(-\frac 1 4)|,\] or

\begin{equation} \label{e9} b=-c_1^{-1}\Phi_{1/4}\star b +c_1^{-1}\Phi_{1/4} \star b_0, \quad c_1=4\Gamma(3/4). \end{equation}
(15)
The operator \(\Phi_\lambda\) with \(\lambda>0\) is a Volterra-type equation which can be solved by iterations, see [1], p.53, Lemmas 5.10, 5.11. If \(b_0\ge 0\) then the solution to (1) is non-negative, \(b\ge 0\). Theorem 1 is proved.

For convenience of the reader let us prove the results mentioned above.

Lemma 2. The operator \(Af:=\int_0^t(t-s)^pf(s)ds\) in the space \(X:=C(0,T)\) for any fixed \(T\in [0,\infty)\) and \(p>-1\) has spectral radius \(r(A)\) equal to zero, \(r(A)=0\). The equation \(f=Af+g\) is uniquely solvable in \(X\). Its solution can be obtained by iterations

\begin{equation} \label{e9a} f_{n+1}=Af_n+g, \quad f_0=g; \quad \lim_{n\to \infty}f_n=f, \end{equation}
(16)
for any \(g\in X\) and the convergence holds in \(X\).

Proof. The spectral radius of a linear operator \(A\) is defined by the formula \(r(A)=\lim_{n\to \infty}\|A^n\|^{1/n}\). By induction one proves that

\begin{equation} \label{e9b} |A^nf|\le t^{n(p+1)}\frac{\Gamma^n(p+1)}{\Gamma(n(p+1)+1)}\|f\|_X, \quad n\ge 1. \end{equation}
(17)
From this formula and the known asymptotic of the gamma function the conclusion \(r(A)=0\) follows. The convergence result (16) is analogous to the well known statement for the assumption \(\|A\|< 1\). Lemma 2 is proved.

If \(q\ge 0\) then inequality (13) implies

\begin{equation} \label{e9c} q\le -c_1^{-1}\Phi_{1/4}\star q +c_1^{-1}\Phi_{1/4}\star b_0. \end{equation}
(18)
Inequality (18) can be solved by iterations with the initial term \(c_1^{-1}\Phi_{1/4}\star b_0\). This yields
\begin{equation} \label{e11} q\le b, \end{equation}
(19)
where \(b\) solves (1). See also [6,7].

Conflict of Interests

The author declares no conflict of interest.

References

  1. Ramm, A. G. (2019). Symmetry Problems. The Navier-Stokes Problem, Morgan & Claypool Publishers, San Rafael, CA. [Google Scholor]
  2. Ramm, A. G. (2019). Solution of the Navier-Stokes problem. Applied Mathematics Letters, 87, 160-164. [Google Scholor]
  3. Ramm, A. G. (2020). Concerning the Navier-Stokes problem. Open Journal of Mathematical Analysis, 4(2), 89-92. [Google Scholor]
  4. Gel'fand, I. & Shilov, G. (1959). Generalized functions, Vol.1, GIFML, Moscow. (in Russian) [Google Scholor]
  5. Brychkov, Y. A., & Prudnikov, A. P. (1986). Integral transforms of generalized functions. Journal of Soviet Mathematics, 34(3), 1630-1655. [Google Scholor]
  6. Ramm, A. G. (2018). Existence of the solutions to convolution equations with distributional kernels. Global Journal of Mathematical Analysis, 6(1), 1-2.[Google Scholor]
  7. Ramm, A. G. (2020). On a hyper-singular equation. Open Journal of Mathematical Analysis, 4(1), 8-10. [Google Scholor]
]]>
A differential inequality and meromorphic starlike and convex functions https://old.pisrt.org/psr-press/journals/oma-vol-4-issue-2-2020/a-differential-inequality-and-meromorphic-starlike-and-convex-functions/ Thu, 08 Oct 2020 17:16:53 +0000 https://old.pisrt.org/?p=4551
OMA-Vol. 4 (2020), Issue 2, pp. 93 - 99 Open Access Full-Text PDF
Kuldeep Kaur Shergill, Sukhwinder Singh Billing
Abstract: In the present paper, we study a differential inequality involving certain differential operator. As a special case of our main result, we obtained certain differential inequalities implying sufficient conditions for meromorphic starlike and meromorphic convex functions of certain order.
]]>

Open Journal of Mathematical Analysis

A differential inequality and meromorphic starlike and convex functions

Kuldeep Kaur Shergill\(^1\), Sukhwinder Singh Billing
Department of Mathematics, Sri Guru Granth Sahib World University, Fatehgarh Sahib-140407(Punjab), India.; (K.K.S & S.S.B)
\(^1\)Corresponding Author: kkshergill16@gmail.com

Abstract

In the present paper, we study a differential inequality involving certain differential operator. As a special case of our main result, we obtained certain differential inequalities implying sufficient conditions for meromorphic starlike and meromorphic convex functions of certain order.

Keywords:

Meromorphic function, meromorphic starlike function, meromorphic convex function, multivalent function, differential operator.

1. Introduction

Let \(\Sigma_{p,n}\) denote the class of functions of the form \[f(z)=\frac{a_{-1}}{z^p}+\sum_{k=n}^\infty a_{k-p}z^{k-p}~(p,n\in \mathbb N=\{1,2,3,\ldots\}),\] which are analytic and \(p\)-valent in the punctured unit disc \( \mathbb E_0=\mathbb E\setminus\{0\},\) where \(\mathbb E = \{z\in\mathbb C:|z|< 1\}\). Define \begin{eqnarray*} D^0f(z)&=&f(z),\\ D^1 f(z)&=&\frac{a_{-1}}{z^p}+2a_0+3a_1z+4a_2z^2+\ldots=\frac{(z^2 f(z))'}{z},\\ D^2f(z)&=&D^1(D^1 f(z)), \end{eqnarray*} and for \(n=1,2,3,\ldots\) \[D^nf(z)=D^1(D^{n-1} f(z))=\frac{(z^2 D^{n-1}f(z))'}{z}.\] Let \(\mathcal {MS}^*_n(p,\alpha)\) denote the class of functions \(f \in \Sigma_{p,n} \) if \[-\Re\frac{1}{p}\left(\frac{D^{n+1}f(z)}{D^nf(z)}-2\right)>\alpha,(\alpha< 1;z \in \mathbb E).\] and let \(\mathcal {MK}_n(p,\alpha)\) denote the class of functions \(f \in \Sigma_{p,n} \) if \[-\Re\frac{1}{p}\left(\frac{(D^{n+1}f(z))'}{(D^nf(z))'}-2\right)>\alpha,(\alpha< 1;z \in \mathbb E).\] The classes of meromorphic starlike functions of order \(\alpha\) and meromorphic convex functions of order \(\alpha\) are denoted by \(\mathcal{MS}^*(\alpha)\) and \(\mathcal{MK}(\alpha)\), respectively and are defined as: \[ \mathcal{MS}^*(\alpha)=\left\{f\in\Sigma:-\Re\left(\frac{zf'(z)}{f(z)}\right)>\alpha,(0\leq\alpha< 1;z \in \mathbb E)\right\},\] and \[\mathcal{MK}(\alpha)=\left\{f\in\Sigma:-\Re\left(1+\frac{zf''(z)}{f'(z)}\right)>\alpha,(0\leq\alpha< 1;z \in \mathbb E)\right\}.\] Note that \(\mathcal{MS}^*(\alpha)=\mathcal{MS}^*_0(1,\alpha)\) and \(\mathcal{MK}(\alpha)=\mathcal{MK}_0(1,\alpha).\) In the theory of meromorphic functions, there exists a variety of results for starlikeness and convexity of meromorphic functions, we state some of them below. Wang et al. [1] proved the following results;

Theorem 1. If \(f(z)\in \Sigma_p\) satisfies the following inequality \[\left|\frac{f(z)}{z f'(z)}\left(1+\frac{z f''(z)}{f'(z)}-\frac{z f'(z)}{f(z)}\right)\right|< \mu~\left(0< \mu< \frac{1}{p}\right),\] then \(f\in\mathcal{MS}_p^*\left(\frac{p}{1+p\mu}\right)\).

Theorem 2. If \(f(z)\in \Sigma_p\) satisfies the inequality \[\left|\frac{z f'(z)}{f(z)}-\frac{z f''(z)}{f'(z)}-1\right|< \delta~(0< \delta< 1),\] then \(f\in\mathcal{MS}_p^*(p(1-\delta))\).

Theorem 3. If \(f(z) \in \Sigma_p\) satisfies the following inequality \[\Re\left(\frac{zf'(z)}{f(z)}+\beta\frac{z^2 f''(z)}{f(z)}\right)< \beta\lambda\left(\lambda+\frac{1}{2}\right)+\frac{1}{2}p\beta-\lambda \hspace{0.5 cm} (\beta\geq 0, p-\frac{1}{2}\leq \lambda \leq p),\] then \(f\in\mathcal{MS}_p^*(\lambda).\)

Goswami et al. [2] proved the following results;

Theorem 4. If \(f(z) \in \Sigma_p,n\) with \(f(z)\neq 0\) for all \(z\in\mathbb E_0\), satisfies the following inequality \[\left| [z^p f(z)]^{\frac{1}{\alpha-p}}\left(\frac{z f'(z)}{f(z)}+\alpha\right)+p-\alpha\right|< \frac{(n+1)(p-\alpha)}{\sqrt{(n+1)^2+1}}, z\in\mathbb E, \] for some real values of \(\alpha~(0\leq\alpha< p)\),then \(f\in\mathcal{MS}_{p,n}^*(\alpha).\)

Theorem 5.If \(f(z) \in \Sigma_p,n\) with \(f(z)\neq 0\) for all \(z\in\mathbb E_0\) satisfies the following inequality \[\left|\frac{ \gamma [z^p f(z)]^\gamma}{z}\left(\frac{z f'(z)}{f(z)}+p\right)\right| \leq\frac{(n+1)}{2\sqrt{(n+1)^2+1}}, z\in\mathbb E, \] for \(\gamma\leq-\frac{1}{p}\),then \(f\in\mathcal{MS}_{p,n}^*\left(p+\frac{1}{\gamma}\right).\)

Theorem 6. If \(f(z) \in \Sigma_p,n\) with \(f(z)\neq 0\) for all \(z\in\mathbb E_0\), satisfies the inequality \[\left| \left(\frac{z^{p+1} f'(z)}{-p}\right)^{\frac{1}{\alpha-p}}\left(1+\frac{z f''(z)}{f'(z)}+\alpha\right)+p-\alpha\right|< \frac{(n+1)(p-\alpha)}{\sqrt{(n+1)^2+1}}, ~z\in\mathbb E ,\] for some real values of \(\alpha~(0\leq\alpha< p)\),then \(f\in\mathcal{MK}_{p,n}(\alpha).\)

Theorem 7. If \(f(z) \in \Sigma_p,n\) with \(f(z)\neq 0\) for all \(z\in\mathbb E_0\), satisfies the inequality \[\left| \frac{1}{z} \left(\frac{z^{p+1} f'(z)}{-p}\right)^{\frac{1}{\alpha-p}}\left(1+\frac{z f''(z)}{f'(z)}+p\right)\right|\leq\frac{(n+1)(p-\alpha)}{2\sqrt{(n+1)^2+1}}, ~z\in\mathbb E ,\] for some real values of \(\alpha~(0\leq\alpha< p)\),then \(f\in\mathcal{MK}_{p,n}(\alpha).\)

From above stated results, we notice that a number of sufficient conditions for meromorphic starlike and meromorphic convex functions have been obtained in terms of differential inequalities in the literature of meromorphic functions. The study of such results is a source of motivation for us to produce the present paper. In the present paper, we study differential inequalities involving a differential operator. As particular cases of our main results, we derive certain sufficient conditions for meromorphic starlike and meromorphic convex functions.

2. Preliminaries

We shall use the following lemma of [3] to prove our result.

Lemma 1. Suppose w is a nonconstant analytic function in \(\mathbb E\) with \(w(0)=0\). If \(|w(z)|\) attains its maximum value at a point \(z_0 \in\mathbb E\) on the circle \(|z|=r< 1,\) then \(z_0 w'(z_0)=m w(z_0),\) where \(m\geq1\), is some real number.

3. Main Results

Theorem 8. Let \(f(z)\in\Sigma_p\) satisfy

\begin{equation} \left|\frac{D^{n+1}[f](z)}{D^n[f](z)}-1\right|^\gamma\left|\frac{D^{n+2}[f](z)}{D^{n+1}[f](z)}-1\right|^\beta < M(p,\alpha,\beta,\gamma),~z\in\mathbb E,\label{eqn1} \end{equation}
(1)
for some real numbers \(\alpha, \beta\) and \(\gamma\) such that \(0\leq\alpha< p\), \(\beta\geq0\), \(\gamma\geq0,\beta+\gamma>0,\) then \(f(z)\in\mathcal {MS}_n^*(p,\alpha)\), where \(n\in\mathbb N_0=\mathbb N\cup\{0\} \) and \begin{eqnarray} M(p,\alpha,\beta,\gamma)& = &\begin{cases}\left(1-\frac{\alpha}{p}\right)^\gamma\left(\frac{1}{2}-\frac{\alpha}{p}\right)^{\beta}, 0\leq\alpha< \frac{p}{2},\nonumber \\ \left(1-\frac{\alpha}{p}\right)^{\gamma+\beta}\left(\frac{2}{2-\frac{\alpha}{p}}\right)^{\beta},~\frac{p}{2}\leq\alpha< p. \end{cases} \end{eqnarray}

Proof. We consider the following two cases separately.

Case (i). When \( 0\leq\alpha< \frac{p}{2}\). Writing \(\frac{\alpha}{p}=\mu\), we see that \(0\leq\mu< \frac{1}{2}.\) Define a function \(w\) as

\begin{equation} 2-\frac{D^{n+1}[f](z)}{D^n[f](z)}=\frac{1+(1-2\mu)w(z)}{1-w(z)},\label{eqn2} \end{equation}
(2)
where \(w\) is an analytic function in \(\mathbb E, ~w(0)=0\) and \(w(z)\neq 1\) in \(\mathbb E\). Differentiating (2) logarithmatically, we get
\begin{equation} \label{eqn3} \frac{z(D^{n+1}[f](z))'}{D^{n+1}[f](z)}-\frac{z(D^{n}[f](z))'}{D^n[f](z)}=\frac{2(\mu-1)zw'(z)}{(1-w(z))(1+(2\mu-3)w(z))}. \end{equation}
(3)
From the definition, we have the identity
\begin{equation} z(D^n[f](z))'=D^{n+1}[f](z)-2D^n[f](z).\label{eqn4} \end{equation}
(4)
Using (4) in (3), we get \(\frac{D^{n+2}[f](z)}{D^{n+1}[f](z)}=\frac{1+(2\mu-3)w(z)}{1-w(z)}+\frac{2(\mu-1)zw'(z)}{(1-w(z))(1+(2\mu-3)w(z))}\). So, we have
\begin{eqnarray} \left|\frac{D^{n+1}[f](z)}{D^{n}[f](z)}-1\right|^\gamma\left|\frac{D^{n+2}[f](z)}{D^{n+1}[f](z)}-1\right|^\beta&=&\left|\frac{2(1-\mu)w(z)}{1-w(z)}\right|^\gamma\left|\frac{2(1-\mu)w(z)}{1-w(z)}+ \frac{2(1-\mu)z w'(z)}{(1-w(z))(1+(2\mu-3)w(z))}\right|^\beta\nonumber\\ &=&\left|\frac{2(1-\mu)w(z)}{1-w(z)}\right|^{\gamma+\beta}\left|1+\frac{z w'(z)}{w(z)(1+(2\mu-3)w(z))}\right|^\beta \end{eqnarray}
(5)
We claim that \(|w(z)|< 1,~z\in\mathbb E.\) Suppose, to the contrary, that there exists a point \(z_0\in\mathbb E\) such that \(\max_{|z|\leq|z_0|} |w(z)|=|w(z_0)|=1\). Then by Lemma 1, we have \(w(z_0)=e^{i\theta},0\leq\theta< 2\pi\) and \(z_0w'(z_0)=mw(z_0),m\geq1.\) Therefore \begin{eqnarray} \left|\frac{D^{n+1}[f](z)}{D^{n}[f](z)}-1\right|^\gamma\left|\frac{D^{n+2}[f](z)}{D^{n+1}[f](z)}-1\right|^\beta&=&\left|\frac{2(1-\mu)w(z_0)}{1-w(z_0)}\right|^{\gamma+\beta}\left|1+ \frac{m}{1+(2\mu-3)w(z_0)}\right|^\beta \nonumber\\ &=&\frac{2^{\gamma+\beta}(1-\mu)^{\gamma+\beta}}{|1-e^{i\theta}|^{\gamma+\beta}} \left|1+ \frac{m}{1+(2\mu-3)e^{i\theta}}\right|^\beta \nonumber\\ &\geq&(1-\mu)^{\gamma+\beta}\left(1- \frac{m}{2(1-\mu)}\right)^\beta \nonumber\\ &\geq&(1-\mu)^{\gamma+\beta}\left(1- \frac{1}{2(1-\mu)}\right)^\beta \nonumber\\ &=&(1-\mu)^\gamma \left(\frac{1}{2}-\mu\right)^\beta,\nonumber \end{eqnarray} which contradicts (1) for \( 0\leq\alpha< \frac{p}{2}.\) Therefore we must have \(|w(z)|< 1\) for all \(z\in\mathbb E,\) and hence from (2), we conclude that \(f\in\mathcal {MS}_n^*(p,\alpha)\).

Case (ii). When \(\frac{p}{2}\leq\alpha< p\), therefore we must have \(\frac{1}{2}\leq\mu< 1\), where \(\mu=\frac{\alpha}{p}.\) Let \(w\) be defined by

\begin{equation} 2-\frac{D^{n+1}[f](z)}{D^n[f](z)}=\frac{\mu}{\mu-(1-\mu)w(z)},\label{eqn5} \end{equation}
(6)
where \( w(z)\neq\frac{\mu}{1-\mu}\) in \(\mathbb E\). Then w is analytic in \(\mathbb E\) with \(w(0)=0\). Proceeding as in Case (i) above, we obtain
\begin{eqnarray} &&\left|\frac{D^{n+1}[f](z)}{D^{n}[f](z)}-1\right|^\gamma\left|\frac{D^{n+2}[f](z)}{D^{n+1}[f](z)}-1\right|^\beta\nonumber\\ &=&\left|\frac{(1-\mu)w(z)}{\mu-(1-\mu)w(z)}\right|^\gamma\left|\frac{(1-\mu)w(z)}{\mu-(1-\mu)w(z)}+ \frac{\mu(1-\mu)z w'(z)}{(\mu-(1-\mu)w(z))(\mu-2(1-\mu)w(z))}\right|^\beta \nonumber\\ &=&\left|\frac{1-\mu}{\mu-(1-\mu)w(z)}\right|^{\gamma+\beta}|w(z)|^\gamma \left|w(z)+\frac{\mu z w'(z)}{\mu-2(1-\mu)w(z)}\right|^\beta. \end{eqnarray}
(7)
We shall prove that \(|w(z)|< 1,~z \in\mathbb E.\) If not, suppose there exists a point \(z_0\in\mathbb E\) such that there exists a point \(z_0\in\mathbb E\) such that \(\max_{|z|\leq|z_0|} |w(z)|=|w(z_0)|=1\). Then by Lemma 1 we have \(w(z_0)=e^{i\theta},0\leq\theta< 2\pi\) and \( z_0w'(z_0)=mw(z_0),m\geq1.\) Therefore \begin{eqnarray*} \left|\frac{D^{n+1}[f](z)}{D^{n}[f](z)}-1\right|^\gamma\left|\frac{D^{n+2}[f](z)}{D^{n+1}[f](z)}-1\right|^\beta&=&\frac{(1-\mu)^{\gamma+\beta}\left(1+\frac{m\mu}{\mu+2-2\mu}\right)^\beta}{|\mu-(1-\mu)w(z_0)|^{\gamma+\beta}}\nonumber\\ &\geq& \left(1-\mu\right)^{\gamma+\beta}\left(\frac{2}{2-\mu}\right)^\beta,\nonumber \end{eqnarray*} which contradicts (1) for \(\frac{p}{2}\leq\alpha< p.\) Therefore, we must have \(|w(z)|< 1\) for all \(z\in\mathbb E\), and hence in view of (6), we conclude that \(f\in\mathcal {MS}_n^*(p,\alpha)\). This completes the proof of the theorem.

Theorem 9. Let \(f(z)\in\Sigma_p\) satisfy

\begin{equation} \left|\frac{(D^{n+1}[f](z))'}{(D^n[f](z))'}-1\right|^\gamma\left|\frac{(D^{n+2}[f](z))'}{(D^{n+1}[f](z))'}-1\right|^\beta < M(p,\alpha,\beta,\gamma),~z\in\mathbb E,\label{eqn7} \end{equation}
(8)
for some real numbers \(\alpha, \beta\) and \(\gamma\) such that \(0\leq\alpha< p\), \(\beta\geq0\), \(\gamma\geq0,\beta+\gamma>0,\) then \(f(z)\in\mathcal {MK}_n(p,\alpha)\), where \(n\in\mathbb N_0=\mathbb N\cup\{0\}\) and \begin{eqnarray} M(p,\alpha,\beta,\gamma)& = &\begin{cases}\left(1-\frac{\alpha}{p}\right)^\gamma\left(\frac{1}{2}-\frac{\alpha}{p}\right)^{\beta}, 0\leq\alpha< \frac{p}{2},\nonumber \\ \left(1-\frac{\alpha}{p}\right)^{\gamma+\beta}\left(\frac{2}{2-\frac{\alpha}{p}}\right)^{\beta},~\frac{p}{2}\leq\alpha< p. \end{cases} \end{eqnarray}

Proof. Again,we consider the following two cases separately.

Case (i). When \( 0\leq\alpha< \frac{p}{2}\). Writing \( \frac{\alpha}{p}=\mu\), we see that \(0\leq\mu< \frac{1}{2}.\) Define a function \(w\) as

\begin{equation} 2-\frac{(D^{n+1}[f](z))'}{(D^n[f](z))'}=\frac{1+(1-2\mu)w(z)}{1-w(z)},\label{eqn8} \end{equation}
(9)
where \(w\) is an analytic function in \(\mathbb E, ~w(0)=0\) and \(w(z)\neq 1\) in \(\mathbb E\). Differentiating (9) logarithmatically, we get
\begin{equation} \frac{z(D^{n+1}[f](z))''}{(D^{n+1}[f](z))'}-\frac{z(D^{n}[f](z))''}{(D^n[f](z))'}=\frac{2(\mu-1)zw'(z)}{(1-w(z))(1+(2\mu-3)w(z))}.\label{eqn9} \end{equation}
(10)
From the following identity
\begin{equation} z(D^n[f](z))'=D^{n+1}[f](z)-2D^n[f](z),\label{eqn10} \end{equation}
(11)
we have
\begin{equation} z(D^n[f](z))''=(D^{n+1}[f](z))'-3(D^n[f](z))'.\label{eqn11} \end{equation}
(12)
Using the identity (12), Equation (10) may be written as \[\frac{(D^{n+2}[f](z))'}{(D^{n+1}[f](z))'}=\frac{1+(2\mu-3)w(z)}{1-w(z)}+\frac{2(\mu-1)zw'(z)}{(1-w(z))(1+(2\mu-3)w(z))}.\] So, we have
\begin{eqnarray} &&\left|\frac{(D^{n+1}[f](z))'}{(D^{n}[f](z))'}-1\right|^\gamma\left|\frac{(D^{n+2}[f](z))'}{(D^{n+1}[f](z))'}-1\right|^\beta \nonumber\\ &=&\left|\frac{2(1-\mu)w(z)}{1-w(z)}\right|^\gamma\left|\frac{2(1-\mu)w(z)}{1-w(z)}+ \frac{2(1-\mu)z w'(z)}{(1-w(z))(1+(2\mu-3)w(z))}\right|^\beta\nonumber\\ &=&\left|\frac{2(1-\mu)w(z)}{1-w(z)}\right|^{\gamma+\beta}\left|1+\frac{z w'(z)}{w(z)(1+(2\mu-3)w(z))}\right|^\beta. \end{eqnarray}
(13)
The remaining part of the proof is similar to that of Theorem 8.

4. Criteria for Starlikeness and Convexity

When we assign particular values to various parameters involved in Theorem 8 and Theorem 9, we obtain following special cases. Setting \(n=0\) in Theorem 8, we obtain the following result.

Corollary 1. Let \(f\in\Sigma_p\) satisfy the condition \begin{eqnarray} \left|\frac{1}{p}\left(\frac{z f'(z)}{ f(z)}\right)-1\right|^\gamma \left|\frac{1}{p}\left(\frac{z f''(z)+3f'(z)}{z f'(z)+2f(z)}\right)-1\right|^\beta &\hspace{-0.3cm}<&\hspace{-0.3cm}\begin{cases}\left(1-\frac{\alpha}{p}\right)^\gamma \left(\frac{1}{2}-\frac{\alpha}{p}\right)^\beta, ~0\leq\alpha< \frac{p}{2},\nonumber \\ \left(1-\frac{\alpha}{p}\right)^{\gamma+\beta} \left(\frac{2}{2-\frac{\alpha}{p}}\right)^\beta,~\frac{p}{2}\leq\alpha< p, \end{cases} \end{eqnarray} for all \(z\in\mathbb E\) and for some real numbers \(\alpha(0\leq\alpha< 1), \beta\geq0\) and \(\gamma\geq0\) with \(\beta+\gamma>0\), then \(f\in\mathcal{MS}^*(p,\alpha)\).

For \(p=1\), Theorem 8 reduces to the following;

Corollary 2. For some real numbers \(\alpha(0\leq\alpha< 1), \beta\geq0\) and \(\gamma\geq0\) with \(\beta+\gamma>0\), if \(f\in\Sigma\) satisfies \begin{eqnarray} \left|\frac{D^{n+1}[f](z)}{D^n[f](z)}-1\right|^\gamma \left|\frac{D^{n+2}[f](z)}{D^{n+1}[f](z)}-1\right|^\beta &\hspace{-0.3cm}<&\hspace{-0.3cm}\begin{cases}\left(1-\alpha\right)^\gamma \left(\frac{1}{2}-\alpha\right)^\beta, ~0\leq\alpha< \frac{1}{2},\nonumber \\ \left(1-\alpha\right)^{\gamma+\beta} \left(\frac{2}{2-\alpha}\right)^\beta,~\frac{1}{2}\leq\alpha< 1, \end{cases} \end{eqnarray} in \(\mathbb E\), then \(\mathcal{MS}^*_n(1,\alpha)\), where \(n\in\mathbb {N}_0.\)

Setting \(n=0\) in above corollary, yields the following result.

Corollary 3. Let \(f(z)\in\Sigma\) satisfy the condition \begin{eqnarray} \left|\frac{z f'(z)}{f(z)}+1\right|^\gamma\left|\frac{z f''(z)+3f'(z)}{z f'(z)+2f(z)}-1\right|^\beta & < &\begin{cases}\left(1-\alpha\right)^\gamma \left(\frac{1}{2}-\alpha\right)^\beta, ~0\leq\alpha< \frac{1}{2},\nonumber \\ \left(1-\alpha\right)^{\gamma+\beta} \left(\frac{2}{2-\alpha}\right)^\beta,~\frac{1}{2}\leq\alpha< 1, \end{cases} \end{eqnarray} where \(z\in\mathbb E, \alpha~(0\leq\alpha< 1), \beta\geq0\) and \(\gamma\geq0\) with \(\beta+\gamma>0\),then \(f\in\mathcal{MS}^*(\alpha)\).

Setting \(\beta=\gamma=1\) and \(\alpha=0\) in above corollary, we obtain the following result.

Remark 1. If \(f(z)\in\Sigma\) satisfies \[\left|\frac{z f'(z)}{f(z)}+1\right|\left|\frac{z f''(z)+3f'(z)}{z f'(z)+2f(z)}-1\right|< \frac{1}{2},~z\in\mathbb E,\] then \(f\in\mathcal{MS}^*\).

By writing \(\beta=1\) and \(\gamma=0\), Theorem 8, we get

Corollary 4. If for all \(z\in\mathbb E\), a function \(f\in\Sigma_p\) satisfies the condition \begin{eqnarray} \frac{D^{n+2}[f](z)}{D^{n+1}[f](z)}&\hspace{-0.3cm}\prec&\begin{cases} 1+\left(\frac{1}{2}-\frac{\alpha}{p}\right)z ,~0\leq\alpha< \frac{p}{2},\nonumber \\ 1+\left[\frac{2\left(1- \frac{\alpha}{p}\right)}{2-\frac{\alpha}{p}}\right]z,~\frac{p}{2}\leq\alpha< p, \end{cases} \end{eqnarray} then \[2-\frac{D^{n+1}[f](z)}{D^n[f](z)}\prec\frac{1+\left(1-\frac{2\alpha}{p}\right)z}{1-z},~z\in\mathbb E, \] i.e. \[\Re\left(2-\frac{D^{n+1}[f](z)}{D^n[f](z)}\right)>\frac{\alpha}{p}.\]

Setting \(n=0\) in Theorem 9, we obtain the following result.

Corollary 5. Let \(f\in\Sigma_p\) satisfy the condition \begin{eqnarray} \left|\frac{1}{p}\left(\frac{z f''(z)+3f'(z)}{ f'(z)}\right)-1\right|^\gamma \left|\frac{1}{p}\left(\frac{z^2 f'''(z)+7zf'(z)+9f'(z)}{z f''(z)+3f(z)}\right)-1\right|^\beta &<&\hspace{-0.3cm}\begin{cases}\left(1-\frac{\alpha}{p}\right)^\gamma \left(\frac{1}{2}-\frac{\alpha}{p}\right)^\beta, ~0\leq\alpha< \frac{p}{2},\nonumber \\ \left(1-\frac{\alpha}{p}\right)^{\gamma+\beta} \left(\frac{2}{2-\frac{\alpha}{p}}\right)^\beta,~\frac{p}{2}\leq\alpha< p, \end{cases} \end{eqnarray} for all \(z\in\mathbb E\) and for some real numbers \(\alpha(0\leq\alpha< 1), \beta\geq0\) and \(\gamma\geq0\) with \(\beta+\gamma>0,\) then \(f\in\mathcal{MK}(p,\alpha)\).

For \(p=1\), Theorem 9 reduces to the following;

Corollary 6. For some real numbers \(\alpha(0\leq\alpha< 1), \beta\geq0\) and \(\gamma\geq0\) with \(\beta+\gamma>0\), if \(f\in\Sigma\) satisfies \begin{eqnarray} \left|\frac{(D^{n+1}[f](z))'}{(D^n[f](z))'}-1\right|^\gamma \left|\frac{(D^{n+2}[f](z))'}{(D^{n+1}[f](z))'}-1\right|^\beta &<&\begin{cases}\left(1-\alpha\right)^\gamma \left(\frac{1}{2}-\alpha\right)^\beta, ~0\leq\alpha< \frac{1}{2},\nonumber \\ \left(1-\alpha\right)^{\gamma+\beta} \left(\frac{2}{2-\alpha}\right)^\beta,~\frac{1}{2}\leq\alpha< 1, \end{cases} \end{eqnarray} in \(\mathbb E\),then \(\mathcal{MK}_n(1,\alpha)\), where \(n\in\mathbb {N}_0.\)

Setting \(n=0\) in above corollary, yields the following result;

Corollary 7. Let \(f(z)\in\Sigma\) satisfy the condition \begin{eqnarray} \left|\frac{z f''(z)}{f'(z)}+2\right|^\gamma\left|\frac{z^2 f'''(z)+7zf''(z)+9f'(z)}{z f''(z)+3f(z)}-1\right|^\beta & < &\begin{cases}\left(1-\alpha\right)^\gamma \left(\frac{1}{2}-\alpha\right)^\beta, ~0\leq\alpha< \frac{1}{2},\nonumber \\ \left(1-\alpha\right)^{\gamma+\beta} \left(\frac{2}{2-\alpha}\right)^\beta,~\frac{1}{2}\leq\alpha< 1, \end{cases} \end{eqnarray} where \(z\in\mathbb E, \alpha~(0\leq\alpha< 1), \beta\geq0\) and \(\gamma\geq0\) with \(\beta+\gamma>0\),then \(f\in\mathcal{MK}(\alpha)\).

Setting \(\beta=\gamma=1\) and \(\alpha=0\) in above corollary, we obtain the following result;

Remark 2. If \(f(z)\in\Sigma\) satisfies \[\left|\frac{z f''(z)}{f'(z)}+2\right|\left|\frac{z^2 f'''(z)+7zf''(z)+9f'(z)}{z f''(z)+3f(z)}-1\right|< \frac{1}{2},~z\in\mathbb E,\] then \(f\in\mathcal{MK}\).

Acknowledgments

The authors are grateful to the referee for careful reading of the paper and valuable suggestions and comments.

Author Contributions

All authors contributed equally to the writing of this paper. All authors read and approved the final manuscript.

Conflict of Interests

The authors declare no conflict of interest.

References

  1. Wang, Z. G., Liu, Z. H., & Xiang, R. G. (2011). Some criteria for meromorphic multivalent starlike functions. Applied Mathematics and Computation, 218(3), 1107-1111. [Google Scholor]
  2. Goswami, P., Bulboaca, T., & Alqahtani, R. T. (2016). Simple sufficient conditions for starlikeness and convexity for meromorphic functions. Open Mathematics, 14(1), 557-566. [Google Scholor]
  3. Jack, I. S. (1971). Functions starlike and convex of order \(\alpha\). Journal of the London Mathematical Society, 2(3), 469-474. [Google Scholor]
]]>