1 Introduction

In this paper, we consider the following viscoelastic problem:

$$ \left\{\begin{array}{llll} u_{tt}(t)- {\Delta} u(t)+{{\int}_{0}^{t}}g(t-s){\Delta} u(s)ds= 0, \qquad \text{ in } {\Omega} \times \mathbb{R}^{+}\\ \frac{\partial u}{\partial \nu}(t)-{{\int}_{0}^{t}}g(t-s)\frac{\partial u}{\partial \nu}(s)ds+h(u_{t}(t))= 0,\quad \text{on } {\Gamma}_{1} \times \mathbb{R}^{+}\\ u(t)= 0,\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad \text{ on } {\Gamma}_{0} \times \mathbb{R}^{+}\\ u(x,0)=u_{0}(x), u_{t}(x,0)=u_{1}(x),\qquad\qquad\quad \text{in } {\Omega} \times \mathbb{R}^{+} \end{array}\right. $$
(1.1)

where u denotes the transverse displacement of waves, Ω is a bounded domain of \(\mathbb {R}^{N} (N\ge 1)\) with a smooth boundary Ω = Γ0 ∪Γ1 such that Γ0 and Γ1 are closed and disjoint, with meas. (Γ0) > 0, ν is the unit outer normal to Ω, and g and h are specific functions.

During the last half century, a great attention has been devoted to the study of viscoelastic problems and many existence and long-time behavior results have been established. We start with the pioneer work of Dafermos [5, 6], where he considered a one-dimensional viscoelastic problem of the form

$$\rho u_{tt}=c u_{xx}-{\int}_{-\infty}^{t}g(t-s)u_{xx}(s)ds, $$

and established various existence results and then proved, for smooth monotone decreasing relaxation functions, that the solutions go to zero as t goes to infinity. However, no rate of decay has been specified. Hrusa [7] considered a one-dimensional nonlinear viscoelastic equation of the form

$$u_{tt}-c u_{xx}+{{\int}_{0}^{t}} m(t-s)(\psi(u_{x}(s)))_{x}ds=f(x,t), $$

and proved several global existence results for large data. He also proved an exponential decay result for strong solutions when m(s) = es and ψ satisfies certain conditions. In [8] Dassios and Zafiropoulos considered a viscoelatic problem in \(\mathbb {R}^{3}\) and proved a polynomial deacy result for exponentially decaying kernels. In their book, Fabrizio and Morro [9] established a uniform stability of some problems in linear viscoelasticity. In all the above mentioned works, the rates of decay in relaxation functions were either of exponential or polynomial type. In 2008, Messaoudi [10, 11] generalized the decay rates allowing an extended class of relaxation functions and gave general decay rates from which the exponential and the polynomial decay rates are only special cases. However, the optimality in the polynomial decay case was not obtained. Precisely, he considered relaxation functions that satisfy

$$ g^{\prime}(t)\le -\xi(t) g(t),\text{ } t\ge 0, $$
(1.2)

where \(\xi :\mathbb {R}^{+} \to \mathbb {R}^{+}\) is a nonincreasing differentiable function and showed that the rate of the decay of the energy is the same rate of decay of g, which is not necessarily of exponential or polynomial decay type. After that a series of papers using Eq. 1.2 has appeared see, for instance, [12,13,14,15,16,17,18] and [19].

Inspired by the experience with frictional damping initiated in the work of Lasiecka and Tataru [20], another step forward was done by considering relaxation functions satisfying

$$ g^{\prime}(t)\le -\chi(g(t)). $$
(1.3)

This condition, where χ is a positive function, χ(0) = χ(0) = 0, and χ is strictly increasing and strictly convex near the origin, with some additional constraints imposed on χ, was used by several authors with different approaches. We refer to previous studies [21,22,23,24,25,26,27] and [28], where general decay results in terms of χ were obtained. Here, it should be mentioned that, in [26], it was the first time where Lasiecka and Wang established not only general but also optimal results in which the decay rates are characterized by an ODE of the same type as the one generated by the inequality (1.3) satisfied by g. Mustafa and Messaoudi [29] established an explicit and general decay rate for relaxation function satisfying

$$ g^{\prime}(t)\le -H(g(t)), $$
(1.4)

where \(H\in C^{1}(\mathbb {R})\), with H(0) = 0 and H is linear or strictly increasing and strictly convex function C2 near the origin. In [30], Cavalcanti et al. considered the following problem

$$ \left\{\begin{array}{llll} {\vert u_{t}\vert}^{\rho}u_{tt}-{\Delta} u-{\Delta} u_{tt}+{{\int}_{0}^{t}}g(t-s){\Delta} u(s)ds= 0,\qquad\qquad\text{in } {\Omega}\times\mathbb{R}^{+}, \\ u(x,t)= 0, \qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad \text{on } {\Gamma}\times\mathbb{R}^{+}, \\ u(x,0)=u_{0}(x),\quad u_{t}(x,0)=u_{1}(x),\qquad\qquad\qquad\qquad\quad \text{in}\hspace{0.05in}{\Omega}\times\mathbb{R}^{+}, \end{array}\right. $$
(1.5)

with a relaxation function satisfying Eq. 1.4 and the additional requirement:

$$\lim\inf_{x\to 0^{+}}{x^{2} H^{\prime\prime} -xH^{\prime}+H(x)}\ge 0, $$

and that \(y^{1-\alpha _{0}}\in L^{1}(1,\infty )\), for some α0 ∈ [0, 1), where y(t) is the solution of the problem

$$y^{\prime}(t)+H(y(t))= 0,\text{ } y(0)=g(0)>0. $$

They characterized the decay of the energy by the solution of a corresponding ODE as in [20]. Recently, Messaoudi and Al-Khulaifi [31] treated (1.5) with a relaxation function satisfying

$$ g^{\prime}(t)\le -\xi(t) g^{p}(t), \text{ } \forall t \ge0,\text{ } 1 \le p<\frac{3}{2}. $$
(1.6)

They obtained a more general stability result for which the results of [10, 11] are only special cases. Moreover, the optimal decay rate for the polynomial case is achieved without any extra work and conditions as in [25] and [20]. For stabilization by mean of boundary feedback, Cavalcanti et al. [32] studied (1.1) and proved a global existence result for weak and strong solutions. Moreover, they gave some uniform decay rate results under some restrictive assumptions on both the kernel g and the damping function h. These restrictions had been relaxed by Cavalcanti et al. [33] and further they established a uniform stability depending on the behavior of h near the origin and on the behavior of g at infinity. In the absence of the viscoelastic term (g = 0), problem (1.1) has been investigated by many authors and several stability results were established. We refer the reader to the work of Lasiecka and Tataru [20], Alabau-Boussouira [34], Cavalcanti et al. [35], Guesmia [36, 37], Cavalcanti [38] and the references therein.

2 Preliminaries

In this section, we present some materials needed in the proof of our results. We use the standard Lebesgue space L2(Ω) and the Sobolev space \({H_{0}^{1}}({\Omega })\) with their usual scalar products and norms and denote by V the following spac

$$V=\{v\in H^{1}({\Omega}):v = 0\hspace{0.05in}\text{on}\hspace{0.05in}{\Gamma}_{0}\}.$$

Throughout this paper, c is used to denote a generic positive constant.

We consider the following hypotheses:

(A1):

\(g: \mathbb {R}^{+}\to \mathbb {R}^{+}\) is a C1 nonincreasing function satisfying

$$ g(0) > 0, \qquad 1-{\int}_{0}^{+\infty}g(s)ds={\ell} > 0, $$
(2.1)

and there exists a C1 function G : (0, ) → (0, ) which is linear or it is strictly increasing and strictly convex C2 function on (0, r1], r1g(0), with G(0) = G(0) = 0, such that

$$ g^{\prime}(t)\le -\xi(t) G(g(t)),\qquad \forall t\ge 0, $$
(2.2)

where ξ(t) is a positive nonincreasing differentiable function.

(A2):

\(h: \mathbb {R}\to \mathbb {R}\) is a nondecreasing C0 function such that there exists a strictly increasing function \(h_{0}\in C^{1}(\mathbb {R}^{+})\), with h0(0) = 0, and positive constants c1, c2, ε such that

$$\begin{array}{@{}rcl@{}} &&h_{0}(\vert s \vert)\le \vert h(s) \vert \le h_{0}^{-1}(\vert s \vert)\qquad \text{for all}\qquad \vert s \vert \le \varepsilon,\\ &&\qquad c_{1}\vert s \vert \le \vert h(s) \vert \le c_{2}\vert s \vert \qquad \text{ for all}\qquad \vert s \vert \ge \varepsilon. \end{array} $$
(2.3)

In addition, we assume that the function H, defined by \(H(s)=\sqrt {s}h_{0} (\sqrt {s})\), is a strictly convex C2 function on (0, r2], for some r2 > 0, when h0 is nonlinear.

Remark 2.1

It is worth noting that condition (2.3) was considered first in [20].

Remark 2.2

Hypothesis (A2) implies that sh(s) > 0, for all s≠ 0.

Remark 2.3

If G is a strictly increasing and strictly convex C2 function on (0, r1], with G(0) = G(0) = 0, then it has an extension \(\overline {G}\), which is strictly increasing and strictly convex C2 function on (0, ). For instance, if G(r1) = a, G(r1) = b, G(r1) = c, we can define \(\overline {G}\), for t > r1, by

$$ \overline{G}(t)=\frac{c}{2} t^{2}+ (b-cr_{1})t+ \left( a+\frac{c}{2} {r_{1}}^{2}-b r_{1} \right). $$
(2.4)

The same remark can be established for \(\overline {H}\).

For completeness we state, without proof, the existence result of [32].

Proposition 2.4

Let (u0, u1) ∈ V × L2(Ω) be given.Assume that (A1) and (A2) are satisfied, then the problem (1.1) has a unique global (weak) solution

$$u\in C(\mathbb{R}^{+};V)\cap C^{1}(\mathbb{R}^{+};L^{2}({\Omega}).$$

Moreover, if

$$(u_{0},u_{1})\in (H^{2}({\Omega})\cap V)\times V,$$

and satisfies the compatibility condition

$$\frac{\partial u_{0}}{\partial \nu}+h(u_{1})= 0\hspace{0.05in}\text{on}\hspace{0.05in}{\Gamma}_{1},$$

then the solution

$$u\in L^{\infty}(\mathbb{R}^{+};H^{2}({\Omega})\cap V)\cap W^{1,\infty}(\mathbb{R}^{+};V)\cap W^{2,\infty}(\mathbb{R}^{+};L^{2}({\Omega})).$$

We introduce the “modified” energy associated to problem (1.1):

$$ E(t)=\frac{1}{2} {\vert\vert{u_{t}(t)}\vert\vert}_{2}^{2} + \frac{1}{2}\left( 1-{{\int}_{0}^{t}}g(s)ds\right){\vert \vert\ {\nabla u(t)} \vert\vert}_{2}^{2}+\frac{1}{2}(g \circ \nabla u)(t), $$
(2.5)

where

$$(g \circ \nabla u)(t)={{\int}_{0}^{t}} g(t-s) {\vert\vert \nabla u(t)- \nabla u(s) \vert\vert}_{2}^{2}ds. $$

Direct differentiation, using Eq. 1.1, leads to

$$ E^{\prime}(t)=\frac{1}{2}(g^{\prime} \circ \nabla u)(t)-\frac{1}{2}g(t){\int}_{\Omega}\vert \nabla u\vert^{2} dx-{\int}_{{\Gamma}_{1}}u_{t}(t) h(u_{t}(t))d{\Gamma} \le0. $$
(2.6)

3 Technical Lemmas

In this section, we establish several lemmas needed for the proof of our main result.

Lemma 3.1

Under the assumptions (A1) and (A2), the functional

$$\psi_{1}(t):={\int}_{\Omega}u(t) u_{t}(t)dx$$

satisfies, along the solution of Eq. 1.1, the estimate

$$ \psi_{1}^{\prime}(t)\le -\frac{\ell}{2}{\vert \vert\ {\nabla u(t)} \vert\vert}_{2}^{2}+{\vert\vert{u_{t}(t)}\vert\vert}_{2}^{2}+ \frac{cC_{\alpha}}{2\ell}(k \circ \nabla u)(t)+c{\int}_{{\Gamma}_{1}} h^{2}(u_{t}(t))d{\Gamma}, \hspace{0.15in} \forall t \in \mathbb{R}^{+}, $$
(3.1)

where, for any 0 < α < 1,

$$ C_{\alpha}={\int}_{0}^{\infty}\frac{g^{2}(s)}{\alpha g(s)-g^{\prime}(s)}ds \hspace{0.15in}\text{ and } \hspace{0.15in}k(t)=\alpha g(t)-g^{\prime}(t). $$
(3.2)

Proof

Direct computations, using Eq. 1.1, yield

$$\begin{array}{@{}rcl@{}} &&\psi^{\prime}(t)={\int}_{\Omega}{u^{2}_{t}}dx+{\int}_{\Omega}u{\Delta} udx-{\int}_{\Omega}u{{\int}_{0}^{t}}g(t-s){\Delta} u(s)dsdx\\ &&\hspace{0.3in}={\int}_{\Omega}{u^{2}_{t}}dx-\left( 1-{{\int}_{0}^{t}}g(s)ds\right){\int}_{\Omega}{\vert \nabla u\vert}^{2}dx-{\int}_{{\Gamma}_{1}}uh(u_{t})d{\Gamma}\\ &&\hspace{0.4in}+{\int}_{\Omega}\nabla u.{{\int}_{0}^{t}}g(t-s)(\nabla u(s)-\nabla u(t))dsdx. \end{array} $$
(3.3)

Using Young’s and Cauchy Schwarz’ inequalities, we obtain

$$\begin{array}{@{}rcl@{}} &&{\int}_{\Omega}\nabla u.{{\int}_{0}^{t}}g(t-s)(\nabla u(s)-\nabla u(t))dsdx\\ &&\hspace{0.15in}\le \delta {\int}_{\Omega}{\vert \nabla u\vert}^{2}dx+\frac{1}{4\delta}{\int}_{\Omega}\left( {{\int}_{0}^{t}}g(t-s)\vert \nabla u(s)-\nabla u(t)\vert ds\right)^{2}dx\\ &&\hspace{0.15in}\le \delta {\int}_{\Omega}{\vert \nabla u\vert}^{2}dx+\\ &&\hspace{0.15in}\frac{1}{4 \delta}{\int}_{\Omega}\left( {{\int}_{0}^{t}}\frac{g(t-s)}{\sqrt{\alpha g(t-s)-g^{\prime}(t-s)}}\sqrt{\alpha g(t-s)-g^{\prime}(t-s)}\vert\nabla u(s)-\nabla u(t)\vert ds\right)^{2}dx\\ &&\hspace{0.15in}\le \delta {\int}_{\Omega}{\vert \nabla u\vert}^{2}dx+\\ &&\hspace{0.15in}\frac{1}{4 \delta}\left( {{\int}_{0}^{t}}\frac{g^{2}(s)}{\alpha g(s)-g^{\prime}(s)}ds\right){\int}_{\Omega}{{\int}_{0}^{t}}\left[\alpha g(t-s)-g^{\prime}(t-s)\right] \vert \nabla u(s)-\nabla u(t)\vert^{2} dsdx\\ &&\hspace{0.15in}\le\delta {\int}_{\Omega}{\vert \nabla u\vert}^{2}dx+\frac{1}{4\delta}C_{\alpha}(k \circ\nabla u)(t). \end{array} $$
(3.4)

Also, use of Young’s and Poincaré’s inequalities and the trace theorem gives

$$\begin{array}{@{}rcl@{}} -{\int}_{{\Gamma}_{1}}uh(u_{t})d{\Gamma} &\le& \delta {\int}_{{\Gamma}_{1}}u^{2}d{\Gamma}+\frac{1}{4\delta}{\int}_{{\Gamma}_{1}}h^{2}(u_{t})d{\Gamma}\\ &\le& c_{p}\delta {\int}_{\Omega}{\vert \nabla u\vert}^{2}d{\Gamma}+\frac{1}{4\delta}{\int}_{{\Gamma}_{1}}h^{2}(u_{t})d{\Gamma}. \end{array} $$
(3.5)

From Eqs. 3.4 and 3.5, we have

$$\begin{array}{@{}rcl@{}} &&{\int}_{\Omega}\nabla u.{{\int}_{0}^{t}}g(t-s)(\nabla u(s)-\nabla u(t))dsdx-{\int}_{{\Gamma}_{1}}uh(u_{t})d{\Gamma}\\ &&\hspace{0.5 in} \leq (1+c_{p})\delta {\int}_{\Omega}{\vert \nabla u\vert}^{2}dx+\frac{1}{4\delta}C_{\alpha}(k \circ \nabla u)(t)+\frac{1}{4\delta}{\int}_{{\Gamma}_{1}}h^{2}(u_{t})d{\Gamma}\\ &&\hspace{0.5in} \leq c\delta {\int}_{\Omega}{\vert \nabla u\vert}^{2}dx+\frac{1}{4\delta}C_{\alpha}(k \circ \nabla u)(t)+\frac{1}{4\delta}{\int}_{{\Gamma}_{1}}h^{2}(u_{t})d{\Gamma} \end{array} $$
(3.6)

Combining Eqs. 3.3 and 3.6 and choosing \(\delta =\frac {\ell }{2c}\) leads to Eq. 3.1. □

Lemma 3.2

Under the assumptions (A1) and (A2), thefunctional

$$\psi_{2}(t):=-{\int}_{\Omega} u_{t}(t){{\int}_{0}^{t}}g(t-s)(u(t)-u(s))dsdx $$

satisfies, along the solution of Eq. 1.1, the estimate

$$\begin{array}{@{}rcl@{}} &\psi_{2}^{\prime}(t) \le \delta{\vert \vert\ {\nabla u(t)} \vert\vert}_{2}^{2}-\left( {{\int}_{0}^{t}}g(s)ds-\delta \right){\vert\vert{u_{t}(t)}\vert\vert}_{2}^{2}+\left( (\frac{3c}{\delta}+ 1)C_{\alpha}+\frac{c}{\delta}\right)(ko \nabla u)(t)\\ &\hspace{0.4in}+c{\int}_{{\Gamma}_{1}} h^{2}(u_{t}(t))d{\Gamma}, \hspace{0.2in} \forall t \in \mathbb{R}^{+}\hspace{0.05in}\text{ and }\hspace{0.08in}\forall \delta>0. \end{array} $$
(3.7)

Proof

By exploiting Eq. 1.1 and performing integration by parts, we arrive at

$$\begin{array}{@{}rcl@{}} \psi_{2}^{\prime}(t)&=&{\int}_{\Omega}\nabla u.{{\int}_{0}^{t}}g(t-s)(\nabla u(t)-\nabla u(s))dsdx\\ &&\hspace{0.3in}-{\int}_{\Omega}\left( {{\int}_{0}^{t}}g(t-s)\nabla u(s)ds\right).\left( {{\int}_{0}^{t}}g(t-s)(\nabla u(t)-\nabla u(s))ds\right)dx\\ &&\hspace{0.3in}+{\int}_{{\Gamma}_{1}}\left( {{\int}_{0}^{t}}g(t-s)(u(t)-u(s))ds\right)h(u_{t})d{\Gamma}\\ &&\hspace{0.3in}-{\int}_{\Omega}u_{t}{{\int}_{0}^{t}}g^{\prime}(t-s)(u(t)-u(s))dsdx-\left( {{\int}_{0}^{t}}g(s)ds\right){\int}_{\Omega}{u_{t}}^{2}dx\\ \hspace{0.25in}&=&\left( 1-{{\int}_{0}^{t}}g(s)ds\right){\int}_{\Omega}\nabla u.{{\int}_{0}^{t}}g(t-s)(\nabla u(t)-\nabla u(s))dsdx\\ &&\hspace{0.2in}+{\int}_{\Omega}{\left| {{\int}_{0}^{t}}g(t-s)(\nabla u(t)-\nabla u(s))ds\right|}^{2}dx\\ &&\hspace{0.3in}+{\int}_{{\Gamma}_{1}}\left( {{\int}_{0}^{t}}g(t-s)(u(t)-u(s))ds\right)h(u_{t})d{\Gamma}\\ &&\hspace{0.3in}-{\int}_{\Omega}u_{t}{{\int}_{0}^{t}}g^{\prime}(t-s)(u(t)-u(s))dsdx-\left( {{\int}_{0}^{t}}g(s)ds\right){\int}_{\Omega}{{u^{2}_{t}}}dx. \end{array} $$

Using Young’s inequality and performing similar calculations as in Eq. 3.4, we obtain

$$\left( 1 - {{\int}_{0}^{t}}g(s)ds\right) {\int}_{\Omega}\nabla u.{{\int}_{0}^{t}}g(t-s)(\nabla u(s)-\nabla u(t))dsdx\!\le \delta {\int}_{\Omega}{\vert \nabla u\vert}^{2}dx+\frac{c C_{\alpha}}{\delta}(k \circ \nabla u)(t)$$

and

$${\int}_{{\Gamma}_{1}}\left( {{\int}_{0}^{t}}g(t-s)(u(t)-u(s))ds\right)h(u_{t})d{\Gamma}\le \frac{c C_{\alpha}}{\delta} (k \circ \nabla u)(t)+ \delta {\int}_{{\Gamma}_{1}}h^{2}(u_{t})d{\Gamma}.$$

Also,

$$\begin{array}{@{}rcl@{}} &&-{\int}_{\Omega}u_{t}{{\int}_{0}^{t}}g^{\prime}(t-s)(u(t)-u(s))dsdx={\int}_{\Omega}u_{t} {{\int}_{0}^{t}}k(t-s)(u(t)-u(s))dsdx\\ &&-{\int}_{\Omega}u_{t} {{\int}_{0}^{t}}\alpha g(t-s)(u(t)-u(s))dsdx\le \delta {\int}_{\Omega}{u_{t}^{2}}dx\\ &&+\frac{1}{2\delta}{\int}_{\Omega}\left( {{\int}_{0}^{t}}\sqrt{k(t-s)} \sqrt{k(t-s)} \vert u(t)-u(s)\vert ds\right)^{2} dx\\ &&+\frac{\alpha^{2}}{2\delta}{\int}_{\Omega}\left( {{\int}_{0}^{t}}g(t-s)\vert u(t)-u(s)\vert ds\right)^{2} dx\le \delta {\int}_{\Omega}{u_{t}^{2}} dx\\ &&+\frac{\left( {{\int}_{0}^{t}}k(s)ds\right)}{2\delta} (k \circ u)(t)+\frac{\alpha^{2} C_{\alpha}}{2\delta}(k \circ \nabla u)(t)\le \delta {\int}_{\Omega}{u_{t}^{2}} dx\\ &&+\frac{c}{\delta}(k \circ \nabla u)(t)+\frac{c C_{\alpha}}{\delta} (k \circ \nabla u)(t). \end{array} $$

Combining all the above estimates, Eq. 3.7 is established. □

Lemma 3.3

Under the assumptions (A1) and (A2), the functional

$$ \psi_{3}(t)={\int}_{\Omega}{{\int}_{0}^{t}}r(t-s)\vert \nabla u(s)\vert^{2}dsdx, $$
(3.8)

satisfies, along the solution of Eq. 1.1, the estimate

$$ \psi_{3}^{\prime}(t)\le -\frac{1}{2}(g \circ \nabla u)(t)+ 3(1-\ell){\int}_{\Omega}\vert \nabla u(t)\vert^{2}dx. $$
(3.9)

where\(r(t)={\int }_{t}^{+\infty }g(s)ds\).

Proof

By Young’s inequality and the fact that r(t) = −g(t), we see that

$$\begin{array}{@{}rcl@{}} &&\psi_{3}^{\prime}(t)=r(0){\int}_{\Omega}\vert \nabla u(t)\vert^{2}dx-{\int}_{\Omega}{{\int}_{0}^{t}}g(t-s)\vert \nabla u(s)\vert^{2} dx\\ &&\hspace{0.3in}=-{\int}_{\Omega}{{\int}_{0}^{t}}g(t-s)\vert \nabla u(s)-\nabla u(t)\vert^{2} ds dx\\ &&\hspace{0.3in}-2{\int}_{\Omega}\nabla u(t).{{\int}_{0}^{t}}g(t-s)(\nabla u(s)-\nabla u(t))dsdx+r(t){\int}_{\Omega}\vert \nabla u(t)\vert^{2} dx. \end{array} $$

Now,

$$\begin{array}{@{}rcl@{}} &&-2{\int}_{\Omega}\nabla u(t).{{\int}_{0}^{t}}g(t-s)(\nabla u(s)-\nabla u(t))dsdx\\ &&\hspace{0.3in}\le 2(1-\ell){\int}_{\Omega}\vert \nabla u(t)\vert^{2} dx+\frac{{{\int}_{0}^{t}}g(s)ds}{2(1-\ell)}{\int}_{\Omega}{{\int}_{0}^{t}}g(t-s)\vert \nabla u(s)-\nabla u(t)\vert^{2} dsdx. \end{array} $$

Using the facts that r(t) ≤ r(0) = 1 − and \({{\int }_{0}^{t}}g(s)ds \le 1-\ell \), Eq. 3.9 is established. □

Lemma 3.4

There exist positive constants d andt1such that

$$ g^{\prime}(t)\le -d g(t),\hspace{0.1in}\forall t\in [0,t_{1}]. $$
(3.10)

Proof

By (A1), we easily deduce that \(\lim _{t\to +\infty }g(t)= 0\). Hence, there is t1 ≥ 0 large enough such that

$$g(t_{1})=r$$

and

$$g(t) \le r,\hspace{0.2in}\forall t\ge t_{1}.$$

As g and ξ are positive nonincreasing continuous and H is a positive continuous function, then, for all t ∈ [0, t1],

$$\left\{\begin{array}{llll} 0<g(t_{1})\le g(t) \le g(0)\\ 0<\xi(t_{1})\le \xi(t) \le \xi(0), \end{array}\right. $$

which implies that there are two positive constants a and b such that

$$a \le \xi(t) H(g(t))\le b.$$

Consequently, for all t ∈ [0, t1],

$$ g^{\prime}(t)\le -\xi(t) H(g(t))\le -\frac{a}{g(0)}g(0)\le -\frac{a}{g(0)}g(t). $$
(3.11)

Remark 3.5

Using the fact that \(\frac {\alpha g^{2}(s)}{\alpha g(s)-g^{\prime }(s)} <g(s)\) and recalling the Lebesgue dominated convergence theorem, we can easily deduce that

$$ \alpha C_{\alpha}={\int}_{0}^{\infty}\frac{\alpha g^{2}(s)}{\alpha g(s)-g^{\prime}(s)}ds \to 0 \text{ as } \alpha \to 0. $$
(3.12)

Lemma 3.6

Assume that (A1) and (A2) hold. Then there exist constantsN, N1, N2, m, m0, c > 0 such thatthe functional

$$L(t)=NE(t)+N_{1}{\psi_{1}} (t)+N_{2} {\psi_{2}}(t)+m_{0}E(t)$$

satisfies, for alltt1,

$$ L^{\prime}(t) \le -mE(t)+c{\int}_{t_{1}}^{t}g(t-s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(s)\vert}^{2}dxds +c {\int}_{{\Gamma}_{1}} h^{2}(u_{t}(t))d {\Gamma} $$
(3.13)

Proof

By using Eqs. 2.63.1 and 3.7, recalling that g = (αgk) and taking \(\delta =\frac {\ell }{4N_{2}}\), we easily see that

$$\begin{array}{@{}rcl@{}} &&L^{\prime}(t) \le -\left( \frac{\ell}{2}N_{1}-\frac{\ell}{4}\right){\vert \vert\ {\nabla u} \vert\vert}_{2}^{2}-\left( N_{2}g_{1}-\frac{\ell}{4}-N_{1}\right){\vert\vert{u_{t}}\vert\vert}_{2}^{2}+\frac{\alpha}{2}N(g \circ \nabla u)(t)\\ &&\hspace{0.35in}-\left( \frac{1}{2}N-\frac{4c}{\ell}{N_{2}^{2}}-C_{\alpha}\left( \frac{c}{2 \ell}N_{1}+\frac{12c}{\ell}{N_{2}^{2}}+N_{2}\right)\right)(k \circ \nabla u)(t)]\\ &&\hspace{0.35in}+ c(N_{1}+N_{2}) {\int}_{{\Gamma}_{1}} h^{2}(u_{t})d {\Gamma} +m_{0}E^{\prime}(t). \end{array} $$
(3.14)

At this point, we choose N1 large enough so that

$$\frac{\ell}{2}N_{1}-\frac{\ell}{4}>4(1-\ell)$$

and then N2 large enough so that

$$N_{2} g_{1}-\frac{\ell}{4}-N_{1}-1 >0.$$

Now, using Remark 3.5, there is 0 < α0 < 1 such that if α < α0, then

$$ \alpha C_{\alpha}<\frac{1}{8\left( \frac{cN_{1}}{2\ell}+\frac{12c {N_{2}^{2}}}{\ell}+N_{2}\right)}. $$
(3.15)

Now, we choose N large enough and α so that

$$\frac{1}{4}N-\frac{4c}{{N^{2}_{2}}}>0\text{ and } \alpha=\frac{1}{2N}<\alpha_{0},$$

which gives

$$\frac{1}{2}N-\frac{4c}{\ell}{N_{2}^{2}}-C_{\alpha}\left( \frac{c}{2\ell}N_{1}+\frac{12c}{\ell}{N_{2}^{2}}+N_{2}\right)>0.$$

Therefore, we arrive at

$$ L^{\prime}(t) \le -4(1-\ell){\vert \vert\ {\nabla u} \vert\vert}_{2}^{2}- {\vert\vert{u_{t}}\vert\vert}_{2}^{2}+\frac{1}{4}(g \circ \nabla u)(t)+ c {\int}_{{\Gamma}_{1}} h^{2}(u_{t})d {\Gamma}+m_{0}E^{\prime}(t). $$
(3.16)

Using Eqs. 2.6 and 3.10 we conclude that, for any tt1,

$$\begin{array}{@{}rcl@{}} &&{\int}_{0}^{t_{1}}g(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\le \frac{-1}{d}{\int}_{0}^{t_{1}}g^{\prime}(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\\ &&\hspace{2.65in}\le -cE^{\prime}(t) \end{array} $$
(3.17)

Combining Eqs. 3.16 and 3.17 and selecting a suitable choice of m0, Eq. 3.13 is established. On the other hand (see [39]), we can choose N even larger (if needed) so that

$$ L \sim E. $$
(3.18)

4 Stability

In this section, we state and prove the main result of our work. For this purpose, we have the following lemmas and remarks.

Lemma 4.1

Under the assumptions (A1) and (A2), the solution of Eq. 1.1satisfies the estimates

$$ {\int}_{{\Gamma}_{1}}h^{2}(u_{t})d{\Gamma} \le c {\int}_{{\Gamma}_{1}}u_{t} h(u_{t})d{\Gamma},\hspace{0.3in}\text{ if } h_{0} \text{ is linear} $$
(4.1)
$$ {\int}_{{\Gamma}_{1}}h^{2}(u_{t})d{\Gamma} \le cH^{-1}(J(t))-cE^{\prime}(t),\hspace{0.3in}\text{ if } h_{0} \text{ is nonlinear} $$
(4.2)

where

$$ J(t):=\frac{1}{\vert{\Gamma}_{12}\vert}{\int}_{{\Gamma}_{12}}u_{t}(t)h(u_{t}(t))d{\Gamma} \le -cE^{\prime}(t) $$
(4.3)

and

$${\Gamma}_{12}=\{x\in {\Gamma}_{1}:\vert u_{t}(t) \vert \le \varepsilon_{1} \}.$$

Proof

Case 1::

h0 is linear. Then, using (A2) we have

$$c^{\prime}_{1} \vert u_{t} \vert \le \vert h(u_{t}) \vert \le c_{2}^{\prime} \vert u_{t} \vert,$$

and hence

$$ h^{2}(u_{t}) \le c_{2}^{\prime} u_{t}h(u_{t}), $$
(4.4)

So, Eq. 4.1 is established.

Case 2::

h0 is nonlinear on [0, ε].

We establish this case, borrowing some ideas from [20]. So, we first assume that max{r2, h0(r2)} < ε; otherwise we take r2 smaller. Let ε1 = min{r2, h0(r2)}. Using (A2), we have, for ε1 ≤|s|≤ ε,

$$\vert h(s) \vert \le \frac{h_{0}^{-1}(\vert s \vert)}{\vert s \vert}\vert s \vert \le \frac{h_{0}^{-1}(\vert \varepsilon \vert)}{\vert \varepsilon_{1} \vert}\vert s \vert $$

and

$$\vert h(s) \vert \ge \frac{h_{0}(\vert s \vert)}{\vert s \vert}\vert s \vert \ge \frac{h_{0}(\vert \varepsilon_{1} \vert)}{\vert \varepsilon \vert}\vert s \vert $$

So, we deduce that

$$ \left\{\begin{array}{llll} h_{0}(\vert s \vert) \le \vert h(s)\vert \le h^{-1}_{0}(\vert s \vert)\hspace{0.1in}\text{for all}\hspace{0.08in} \vert s \vert < \varepsilon_{1}\\ c^{\prime}_{1} \vert s \vert \le \vert h(s) \vert \le c^{\prime}_{2} \vert s \vert \hspace{0.1in}\text{for all}\hspace{0.08in} \vert s \vert \ge \varepsilon_{1} \end{array}\right. $$
(4.5)

Then Eq. 4.5, yields, for all |s|≤ ε1,

$$H(h^{2}(s))=\vert h(s) \vert h_{0}(\vert h(s) \vert) \le sh(s)$$

which gives

$$ h^{2}(s) \le H^{-1}(sh(s))\hspace{0.1in}\text{for all}\hspace{0.08in} \vert s \vert \le\varepsilon_{1}. $$
(4.6)

Now, we define the following partition which was first introduced by Komornik [40]:

$${\Gamma}_{11}=\{x\in {\Gamma}_{1}:\vert u_{t}(t) \vert > \varepsilon_{1} \},\hspace{0.2in}{\Gamma}_{12}=\{x\in {\Gamma}_{1}:\vert u_{t}(t) \vert \le \varepsilon_{1} \}$$

Using Eq. 4.5, we get on Γ12

$$ u_{t}h(u_{t}(t)) \le\varepsilon_{1} h_{0}^{-1}(\varepsilon_{1}) \le h_{0}(r_{2})r_{2}=H({r_{2}^{2}}). $$
(4.7)

Then, Jensen’s inequality gives (note that H− 1 is concave)

$$ H^{-1}\left( J(t)\right)\ge c {\int}_{{\Gamma}_{12}}H^{-1}(u_{t}(t)h(u_{t}(t)))d{\Gamma}. $$
(4.8)

Thus, combining Eqs. 4.6 and 4.8, we arrive at

$$\begin{array}{@{}rcl@{}} &&{\int}_{{\Gamma}_{1}}h^{2}(u_{t}(t))d{\Gamma}={\int}_{{\Gamma}_{12}}h^{2}(u_{t}(t))d{\Gamma}+{\int}_{{\Gamma}_{11}}h^{2}(u_{t}(t))d{\Gamma}\\ &&\hspace{1.1in}\le {\int}_{{\Gamma}_{12}}H^{-1}\left( u_{t}h(u_{t}(t))\right)d{\Gamma}+{\int}_{{\Gamma}_{11}}h^{2}(u_{t}(t))d{\Gamma}\\ &&\hspace{1.1in}\le cH^{-1}(J(t))-cE^{\prime}(t) \end{array} $$
(4.9)

Lemma 4.2

Assume that (A1) and(A2) hold andh0is linear. Then, the energy functional satisfies the following estimate

$$ {\int}_{0}^{+\infty}E(s)ds < \infty $$
(4.10)

Proof

Let F(t) = L(t) + ψ3(t), then using Eqs. 3.9 and 3.16, we obtain

$$ F^{\prime}(t)\le -(1-\ell){\int}_{\Omega}\vert \nabla u\vert dx-{\int}_{\Omega}{u_{t}^{2}} dx-\frac{1}{4}(go\nabla u)(t)+ c {\int}_{{\Gamma}_{1}} h^{2}(u_{t})d {\Gamma} $$
(4.11)

Using Eqs. 2.64.1 and 4.11, we obtain

$$\begin{array}{@{}rcl@{}} &&F^{\prime}(t)\le -b E(t)+c{\int}_{\Omega}u_{t} h(u_{t})dx\\ &&\hspace{0.3in}\le -bE(t)-cE^{\prime}(t), \end{array} $$

where b is some positive constant. Therefore,

$$ b{\int}_{t_{1}}^{t}E(s)ds\le F_{1}(t_{1})-F_{1}(t)\le F_{1}(t_{1})<\infty, $$
(4.12)

where F1(t) = F(t) + cE(t) ∼ E. □

Let’s define

$$ I(t):=-{\int}_{t_{1}}^{t}g^{\prime}(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\le -cE^{\prime}(t), $$
(4.13)

Lemma 4.3

Under the assumptions (A1) and (A2), we have the following estimates

$$ {\int}_{t_{1}}^{t}g(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\le \frac{1}{q} \overline{G}^{-1}\left( \frac{qI(t)}{\xi(t)}\right),\hspace{0.1in}\text{if } h_{0} \text{ is linear} $$
(4.14)
$$ {\int}_{t_{1}}^{t}g(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\le \frac{(t-t_{1})}{q} \overline{G}^{-1}\left( \frac{qI(t)}{(t-t_{1})\xi(t)}\right),\hspace{0.1in}\text{if } h_{0} \text{ is nonlinear} $$
(4.15)

whereq ∈ (0, 1) and\(\overline {G}\)is an extension ofGsuch that\(\overline {G}\)is strictly increasing and strictly convexC2function on (0, ); see Remark 2.3.

Proof

First we establish (4.14). For this, we define the following quantity

$$\lambda (t):=q{\int}_{t_{1}}^{t}{\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds,$$

where, by Eq. 4.10, q is chosen so small that, for all tt1,

$$ \lambda (t)<1. $$
(4.16)

Since G is strictly convex on (0, r1] and G(0) = 0, then

$$ G(\theta z)\le \theta G(z),\text{ } 0\le \theta \le 1\text{ and } z\in (0,r_{1}]. $$
(4.17)

The use of Eqs. 2.24.16, and 4.17 and Jensen’s inequality leads to

$$\begin{array}{@{}rcl@{}} &&I(t)=\frac{1}{q\lambda(t)}{\int}_{t_{1}}^{t}\lambda (t)(-g^{\prime}(s)){\int}_{\Omega}{q\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\\ &&\hspace{0.3in}\ge \frac{1}{q\lambda(t)}{\int}_{t_{1}}^{t}\lambda (t)\xi(s) G(g(s)){\int}_{\Omega}{q\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\\ &&\hspace{0.3in}\ge \frac{\xi(t)}{q\lambda(t)}{\int}_{t_{1}}^{t}G(\lambda (t)g(s)){\int}_{\Omega}{q\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\\ &&\hspace{0.3in}\ge \frac{\xi(t)}{q}G\left( q{\int}_{t_{1}}^{t}g(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\right)\\ &&\hspace{0.3in}=\frac{\xi(t)}{q}\overline{G}\left( q{\int}_{t_{1}}^{t}g(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(s)\vert}^{2}dxds\right) \end{array} $$
(4.18)

This gives (4.14).

For the proof of (4.15), we define the following

$$\lambda_{1} (t):=\frac{q}{(t-t_{1})}{\int}_{t_{1}}^{t}{\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds,$$

then using (2.5) and (2.6), we easily see that

$$\lambda_{1}(t)\le \frac{8q E(0)}{\ell},$$

then choosing q ∈ (0, 1) small enough so that, for all tt1,

$$ \lambda_{1} (t)<1. $$
(4.19)

The use of Eqs. 2.24.17 and 4.19 and Jensen’s inequality leads to

$$\begin{array}{@{}rcl@{}} &&I(t)=\frac{1}{q\lambda_{1}(t)}{\int}_{t_{1}}^{t}\lambda_{1} (t)(-g^{\prime}(s)){\int}_{\Omega}{q\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\\ &&\hspace{0.3in}\ge \frac{1}{q\lambda_{1}(t)}{\int}_{t_{1}}^{t}\lambda_{1} (t)\xi(s) G(g(s)){\int}_{\Omega}{q\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\\ &&\hspace{0.3in}\ge \frac{\xi(t)}{q\lambda_{1}(t)}{\int}_{t_{1}}^{t}G(\lambda_{1} (t)g(s)){\int}_{\Omega}{q\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\\ &&\hspace{0.3in}\ge \frac{(t-t_{1})\xi(t)}{q} G\left( \frac{q}{(t-t_{1})}{\int}_{t_{1}}^{t}g(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\right)\\ &&\hspace{0.3in}=\frac{(t-t_{1})\xi(t)}{q}\overline{G}\left( \frac{q}{(t-t_{1})} {\int}_{t_{1}}^{t}g(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\right). \end{array} $$
(4.20)

This implies that

$${\int}_{t_{1}}^{t}g(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(t-s)\vert}^{2}dxds\le \frac{(t-t_{1})}{q} \overline{G}^{-1}\left( \frac{qI(t)}{(t-t_{1})\xi(t)}\right)$$

Theorem 4.4

Let (u0, u1) ∈ V × L2(Ω) be given. Assume that (A1) and (A2) are satisfied andh0is linear. Then there exist strictly positive constantsc1, c2, k1andk2such that the solution of Eq. 1.1satisfies, for alltt1,

$$ E (t)\le c_{1} e^{-c_{2}{\int}_{t_{1}}^{t}\xi(s)ds},\text{ if \textit{G} is linear} $$
(4.21)
$$ E (t)\le k_{2} G_{1}^{-1}\left( k_{1} {\int}_{t_{1}}^{t}\xi(s)ds\right),\text{ if } G \text{ is nonlinear,} $$
(4.22)

where\(G_{1}(t)={\int }_{t}^{r_{1}}\frac {1}{sG^{\prime }(s)}ds\).

Proof

Case 1::

G is linear

Multiplying (3.13) by ξ(t) and using Eqs. 2.22.64.14.3 and 4.13, we get

$$\begin{array}{@{}rcl@{}} &&\xi(t) L^{\prime}(t) \le\\ &&-m \xi(t) E(t)+c \xi(t) {\int}_{t_{1}}^{t}g(t-s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(s)\vert}^{2}dxds+c \xi(t) {\int}_{{\Gamma}_{1}} h^{2}(u_{t}(t))d {\Gamma}\\ & &\le -m \xi(t) E(t)+c {\int}_{t_{1}}^{t}\xi(s) g(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(s)\vert}^{2}dxds +c \xi(t) {\int}_{{\Gamma}_{1}} h^{2}(u_{t}(t))d {\Gamma}\\ & &\le -m \xi(t) E(t)-c {\int}_{t_{1}}^{t} g^{\prime}(s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(s)\vert}^{2}dxds +c \xi(t) {\int}_{{\Gamma}_{1}} u_{t}h(u_{t}(t))d {\Gamma}\\ &&\le -m \xi(t) E(t)-2c E^{\prime}(t) \end{array} $$

which gives, as ξ(t) is non-increasing,

$$ (\xi L + 2c E)^{\prime} \leq -m \xi(t) E(t), \forall t \geq t_{1}. $$
(4.23)

Hence, using the fact that ξL + 2cEE, we easily obtain

$$ E (t)\le c^{\prime} e^{-\bar{c}{\int}_{t_{1}}^{t} \xi (s) ds}. $$
(4.24)
Case 2::

G is non-linear.

Using Eqs. 3.134.1 and 4.14, we obtain

$$ L^{\prime}(t)\le -m E(t)+c\left( \overline{G}\right)^{-1}\left( \frac{qI(t)}{\xi(t)}\right)-cE^{\prime}(t),\hspace{0.1in} $$
(4.25)

Let \(\mathcal {F}_{1}(t)=L(t)+cE(t)\sim E\), then Eq. 4.25 becomes

$$ \mathcal{F}_{1}^{\prime}(t)\le -m E(t)+c\left( \overline{G}\right)^{-1}\left( \frac{qI(t)}{\xi(t)}\right),\hspace{0.1in} $$
(4.26)

we find that the functional \(\mathcal {F}_{2}\), defined by

$$\mathcal{F}_{2}(t):=\overline{G}^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)\mathcal{F}_{1}(t)$$

satisfies, for some α1, α2 > 0.

$$ \alpha_{1}\mathcal{F}_{2}(t)\le E(t)\le \alpha_{2}\mathcal{F}_{2}(t) $$
(4.27)

and

$$\begin{array}{@{}rcl@{}} \mathcal{F}_{2}^{\prime}(t)&=&\varepsilon_{0}\frac{E^{\prime}(t)}{E(0)}\overline{G}^{\prime\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)\mathcal{F}_{1}(t)+\overline{G}^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right){\mathcal{F}_{1}}^{\prime}(t)\\ \hspace{0.3in}&\le& -m E(t)\overline{G}^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)+c \overline{G}^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)\overline{G}^{-1}\left( \frac{qI(t)}{\xi(t)}\right). \end{array} $$
(4.28)

Let \(\overline {G}^{*}\) be the convex conjugate of \(\overline {G}\) in the sense of Young (see [41]), then

$$ \overline{G}^{*}(s)=s(\overline{G}^{\prime})^{-1}(s)-\overline{G}\left[(\overline{G}^{\prime})^{-1}(s)\right],\hspace{0.1in}\text{if}\hspace{0.05in}s\in (0,\overline{G}^{\prime}(r_{1})] $$
(4.29)

and \(\overline {G}^{*}\) satisfies the following generalized Young inequality

$$ A B\le \overline{G}^{*}(A)+\overline{G}(B),\hspace{0.15in}\text{if}\hspace{0.05in}A\in (0,\overline{G}^{\prime}(r_{1})],\hspace{0.05in}B\in(0,r_{1}]. $$
(4.30)

So, with \(A=\overline {G}^{\prime }\left (\varepsilon _{0}\frac {E^{\prime }(t)}{E(0)}\right )\) and \(B=\overline {G}^{-1}\left (\frac {qI(t)}{\xi (t)}\right )\) and using Eqs. 2.6 and 4.284.30, we arrive at

$$\begin{array}{@{}rcl@{}} \mathcal{F}_{2}^{\prime}(t)&\le& -m E(t)\overline{G}^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)+c \overline{G}^{*}\left( \overline{G}^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)\right)+c \left( \frac{qI(t)}{\xi(t)}\right)\\ \hspace{0.3in}&\le& -m E(t)\overline{G}^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)+c\varepsilon_{0}\frac{E(t)}{E(0)}\overline{G}^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)+c\left( \frac{qI(t)}{\xi(t)}\right). \end{array} $$
(4.31)

So, multiplying Eq. 4.31 by ξ(t) and using the fact that \(\varepsilon _{0}\frac {E(t)}{E(0)}<r_{1}\), \(\overline {G}^{\prime }\left (\varepsilon _{0}\frac {E(t)}{E(0)}\right )=G^{\prime }\left (\varepsilon _{0}\frac {E(t)}{E(0)}\right )\), gives

$$\begin{array}{@{}rcl@{}} &&\xi(t) \mathcal{F}_{2}^{\prime}(t)\le -m \xi(t) E(t)G^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)+c \xi(t)\varepsilon_{0}\frac{E(t)}{E(0)}G^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)+c q I(t)\\ &&\hspace{0.3in}\le-m \xi(t) E(t)G^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)+c \xi(t)\varepsilon_{0}\frac{E(t)}{E(0)}G^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)-c E^{\prime}(t) \end{array} $$

Consequently, with a suitable choice of ε0, we obtain, for all tt1,

$$ \mathcal{F}_{3}^{\prime}(t)\le -k \xi(t)\left( \frac{E(t)}{E(0)}\right)G^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)=-k\xi(t) G_{2}\left( \frac{E(t)}{E(0)}\right), $$
(4.32)

where \(\mathcal {F}_{3}=\xi \mathcal {F}_{2}+c E \sim E\) and G2(t) = tG(ε0t). Since \(G^{\prime }_{2}(t)=G^{\prime }(\varepsilon _{0}t)+\varepsilon _{0}t G^{\prime \prime }(\varepsilon _{0}t)\), then, using the strict convexity of G on (0, r1], we find that \(G_{2}^{\prime }(t), G_{2}(t)>0\) on (0, 1]. Thus, with

$$R(t)=\varepsilon \frac{\alpha_{1}\mathcal{F}_{3}(t)}{E(0)},\hspace{0.1in}0<\varepsilon<1, $$

taking in account (4.27) and (4.32), we have

$$ R(t)\sim E(t) $$
(4.33)

and, for some k1 > 0.

$$R^{\prime}(t)\le -k_{1}\xi(t) G_{2}(R(t)),\hspace{0.1in}\forall t\ge t_{1}. $$

Then, the integration over (t1, t) yields

$${\int}_{t_{1}}^{t}\frac{-R^{\prime}(s)}{G_{2}(R(s))}ds \ge k_{1}{\int}_{t_{1}}^{t}\xi(s)ds. $$

Hence, by an approprite change of variable, we get

$${\int}_{\varepsilon_{0} R(t)}^{\varepsilon_{0} R(t_{1})}\frac{1}{\tau G^{\prime}(\tau)}d\tau \ge k_{1} {\int}_{t_{1}}^{t} \xi(s)ds $$

Thus, we have

$$ R(t)\le \frac{1}{\varepsilon_{0}}G_{1}^{-1}\left( k_{1}{\int}_{t_{1}}^{t}\xi(s)ds\right), $$
(4.34)

where \(G_{1}(t)={\int }_{t}^{r_{1}}\frac {1}{sG^{\prime }(s)}ds\). Here, we have used the fact that G1 is strictly decreasing on (0, r1]. Therefore Eq. 4.22 is established by virtue of Eq. 4.33. □

Remark 4.5

The decay rate of E(t) given by Eq. 2.2 is optimal because it is consistent with the decay rate of g(t) given by Eq. 4.22. In fact,

$$g(t)\le G_{0}^{-1}\left( {\int}_{g^{-1}(r_{1})}^{t}\xi(s)ds\right),\hspace{0.1in}\forall t \ge g^{-1}(r_{1}),$$

where \(G_{0}(t)={{\int }_{t}^{r}}\frac {1}{G(s)}\). Using the properties of G, G0 and G1, we can see that

$$G_{1}(t)={\int}_{t}^{r_{1}}\frac{1}{s G^{\prime}(s)}ds\le {\int}_{t}^{r_{1}}\frac{1}{G(s)}ds=G_{0}(t). $$

This implies

$$G_{1}^{-1}(t)\le G_{0}^{-1}(t).$$

This shows that Eq. 4.22 provides the best decay rates expected under the very general assumption (2.2).

Theorem 4.6

Let (u0, u1) ∈ V × L2(Ω) be given. Assume that (A1) and (A2) are satisfied andh0is nonlinear. Then there exist strictly positive constantsc3, c4, k2, k3andε2such that the solution of Eq. 1.1satisfies, for alltt1,

$$ E(t)\le H_{1}^{-1}\left( c_{3}{\int}_{t_{1}}^{t}\xi(s)ds+c_{4}\right),\text{ if \textit{G} is linear,} $$
(4.35)

where\(H_{1}(t)={{\int }_{t}^{1}}\frac {1}{H_{2}(s)}ds\).

$$ E(t) \leq k_{3} (t-t_{1}) {W_{2}}^{-1} \left( \frac{k_{2}}{(t-t_{1}){\int}_{t_{1}}^{t}\xi(s) ds} \right), \text{ if \textit{G} is non-linear,} $$
(4.36)

whereW2(t) = tW(ε2t) and\(W=\left (\overline {G}^{-1}+\overline {H}^{-1}\right )^{-1}\).

Proof

Case 1::

G is linear

Multiplying Eq. 3.13 by ξ(t) and using Eq. 4.2, we get

$$\begin{array}{@{}rcl@{}} &&\xi(t) L^{\prime}(t) \le -m \xi(t) E(t)+c \xi(t) {\int}_{t_{1}}^{t}g(t-s){\int}_{\Omega}{\vert \nabla u(t)-\nabla u(s)\vert}^{2}dxds \\ &&\hspace{0.95in}+c \xi(t) {\int}_{{\Gamma}_{1}} h^{2}(u_{t}(t))d {\Gamma}\\ &&\hspace{0.3in} \le -m \xi(t) E(t)-cE^{\prime}(t)+c \xi(t) {\int}_{{\Gamma}_{1}} h^{2}(u_{t}(t))d {\Gamma}\\ &&\hspace{0.3in} \le -m \xi(t) E(t)-c E^{\prime}(t) + c \xi(t) H^{-1}(J(t)) -c \xi(t) E^{\prime}(t)\\ &&\hspace{0.3in} \le -m \xi(t) E(t)-c E^{\prime}(t) + c \xi(t) H^{-1}(J(t)) -c \xi(0) E^{\prime}(t)\\ &&\hspace{0.3in} \le -m \xi(t) E(t)-c E^{\prime}(t) + c \xi(t) H^{-1}(J(t)) \end{array} $$

which gives, as ξ(t) is non-increasing,

$$ (\xi L +c E)^{\prime} \leq -m \xi(t) E(t)+ c \xi(t) H^{-1}(J(t)), \forall t \geq t_{1}. $$
(4.37)

Therefore, Eq. 4.37 becomes

$$ {\mathcal{L}}^{\prime}(t) \leq -m \xi(t) E(t)+ c \xi(t) H^{-1}(J(t)), \forall t \geq t_{1}, $$
(4.38)

where \(\mathcal {L}:=\xi L + 2c E\), which is clearly equivalent to E. Now, for ε1 < r2 and c0 > 0, using Eq. 4.38 and the fact that E≤ 0, H > 0, H > 0 on (0, r2], we find that the functional \(\mathcal {L}_{1}\), defined by

$$\mathcal{L}_{1}(t):=H^{\prime} \left( \varepsilon_{1}\frac{E(t)}{E(0)}\right)\mathcal{L}(t)+c_{0}E(t)$$

satisfies, for some α3, α4 > 0.

$$ \alpha_{3} \mathcal{L}_{1}(t)\le E(t)\le \alpha_{4}\mathcal{L}_{1}(t) $$
(4.39)

and

$$\begin{array}{@{}rcl@{}} \mathcal{L}_{1}^{\prime}(t)&=&\varepsilon_{0}\frac{E^{\prime}(t)}{E(0)}H^{\prime\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right) \mathcal{L}(t)+ H^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right){\mathcal{L}}^{\prime}(t)+c_{0}E^{\prime}(t)\\ \hspace{0.3in}&\le& \!-m E(t)H^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)\!+c \xi(t) H^{\prime}\left( \varepsilon_{0}\frac{E(t)}{E(0)}\right)H^{-1}(J(t))+c_{0}E^{\prime}(t) \end{array} $$
(4.40)

Let H be the convex conjugate of H in the sense of Young (see [41]), then, as in Eqs. 4.29 and 4.30, with \(A=H^{\prime }\left (\varepsilon _{1}\frac {E(t)}{E(0)}\right )\) and B = H− 1(J(t)), using Eqs. 2.6 and 4.7, we arrive at

$$\begin{array}{@{}rcl@{}} \mathcal{L}_{1}^{\prime}(t)&\le& -m E(t)H^{\prime}\left( \varepsilon_{1}\frac{E(t)}{E(0)}\right)+c \xi(t) H^{*}\left( H^{\prime}\left( \varepsilon_{1}\frac{E(t)}{E(0)}\right)\right)+c \xi(t) J(t)+c_{0}E^{\prime}(t)\\ &\le& -m E(t)H^{\prime}\left( \varepsilon_{1}\frac{E(t)}{E(0)}\right)+c\varepsilon_{1} \xi(t) \frac{E(t)}{E(0)}H^{\prime}\left( \varepsilon_{1}\frac{E(t)}{E(0)}\right)-c E^{\prime}(t)+c_{0} E^{\prime}(t) \end{array} $$

Consequently, with a suitable choice of ε1 and c0, we obtain, for all tt1,

$$ \mathcal{L}_{1}^{\prime}(t)\le -c \xi(t) \frac{E^{\prime}(t)}{E(0)}H^{\prime}\left( \varepsilon_{1}\frac{E(t)}{E(0)}\right)=-c \xi(t) H_{2}\left( \varepsilon_{1}\frac{E(t)}{E(0)}\right), $$
(4.41)

where H2(t) = tH(ε1t). Since \(H^{\prime }_{2}(t)=H^{\prime }(\varepsilon _{1}t)+\varepsilon _{1}t H^{\prime \prime }(\varepsilon _{1}t)\), then, using the strict convexity of H on (0, r2], we find that \(H_{2}^{\prime }(t), H_{2}(t)>0\) on (0, 1]. Thus, with

$$R_{1}(t)=\varepsilon \frac{\alpha_{3}{\mathcal{L}_{1}}(t)}{E(0)},\hspace{0.1in}0<\varepsilon<1, $$

taking in account (4.39) and (4.41), we have

$$ R_{1}(t)\sim E(t) $$
(4.42)

and, for some c3 > 0.

$$R_{1}^{\prime}(t)\le -c_{3} \xi(t) H_{2}(R_{1}(t)),\hspace{0.1in}\forall t\ge t_{1}.$$

Then, a simple integration gives, for some c4 > 0,

$$ R_{1}(t)\le H_{1}^{-1}(c_{3}{\int}_{t_{1}}^{t} \xi(s) ds +c_{4}),\hspace{0.1in}\forall t\ge t_{1}, $$
(4.43)

where \(H_{1}(t)={{\int }_{t}^{1}}\frac {1}{H_{2}(s)}ds\).

Case 2. :

G is non-linear.

Using Eqs. 3.134.2 and 4.15, we obtain

$$ L^{\prime}(t)\le -m E(t)+c(t-t_{1})\left( \overline{G}\right)^{-1}\left( \frac{qI_{1}(t)}{(t-t_{1})\xi(t)}\right)+c H^{-1}(J(t))-cE^{\prime}(t). $$
(4.44)

Since \(\lim _{t\to +\infty } \frac {1}{t-t_{1}}= 0\), there exists t2 > t1 such that \(\frac {1}{t-t_{1}} < 1\) whenever t > t2. Combining this with the strictly increasing and strictly convex properties of \(\overline {H}\), setting \(\theta =\frac {1}{t-t_{1}} < 1\) and using Eq. 4.17, we obtain

$$\overline{H}^{-1}(J(t)) \leq (t-t_{1})\overline{H}^{-1}\left( \frac{J(t)}{(t-t_{1})}\right), \forall t \ge t_{2}$$

and, then, Eq. 4.44 becomes

$$\begin{array}{@{}rcl@{}} &&L^{\prime}(t)\le -m E(t)+c(t-t_{1})\left( \overline{G}\right)^{-1}\left( \frac{qI_{1}(t)}{(t-t_{1})\xi(t)}\right)+c(t-t_{1})\overline{H}^{-1}\left( \frac{J(t)}{(t-t_{1})}\right)\\ &&\hspace{1.5in}-cE^{\prime}(t),\hspace{0.5in}\forall t\ge t_{2}. \end{array} $$
(4.45)

Let L1(t) = L(t) + cE(t) ∼ E, then Eq. 4.45 takes the form

$$ L_{1}^{\prime}(t)\le -m E(t)+c(t-t_{1})\left( \overline{G}\right)^{-1}\left( \frac{qI_{1}(t)}{(t-t_{1})\xi(t)}\right)+c(t-t_{1})\overline{H}^{-1}\left( \frac{J(t)}{(t-t_{1})})\right),\hspace{0.1in} $$
(4.46)
$$\text{Let } r_{0}=\min{\{r_{1},r_{2}\}},\chi(t)=\max{\{\frac{qI_{1}(t)}{(t-t_{1})\xi(t)},\frac{J(t)}{(t-t_{1})}}\}\text{ and } W=\left( \left( \overline{G}\right)^{-1}+\overline{H}^{-1}\right)^{-1}. $$

So, Eq. 4.46 reduces to

$$ L_{1}^{\prime}(t)\le -m E(t)+c(t-t_{1}) W^{-1}(\chi(t)),\forall t\ge t_{2}\hspace{0.1in} $$
(4.47)

Now, for ε2 < r0 and using Eq. 4.44 and the fact that E≤ 0, W > 0, W > 0 on (0, r0], we find that the functional L2, defined by

$$L_{2}(t):=W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right)L_{1}(t), \hspace{0.1in}\forall t\ge t_{2}, $$

satisfies, for some α5, α6 > 0.

$$ \alpha_{5} L_{2}(t)\le E(t)\le \alpha_{6}L_{2}(t) $$
(4.48)

and, for all tt2,

$$\begin{array}{@{}rcl@{}} L_{2}^{\prime}(t)&=&\left( \frac{-\varepsilon_{2}}{(t-t_{1})^{2}} \frac{E(t)}{E(0)}+\frac{\varepsilon_{2}}{(t-t_{1})} \frac{E^{\prime}(t)}{E(0)}\right)W^{\prime\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot\frac{E(t)}{E(0)}\right){L_{1}}(t)\\ &&+W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot\frac{E(t)}{E(0)}\right){L}^{\prime}_{1}(t)\le -m E(t)W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot\frac{E(t)}{E(0)}\right)\\ &&+c(t-t_{1})W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot\frac{E(t)}{E(0)}\right)W^{-1}(\chi(t)). \end{array} $$
(4.49)

Let W be the convex conjugate of W in the sense of Young (see [41]), then, as in Eqs. 4.29 and 4.30, and with \(A=W^{\prime }\left (\frac {\varepsilon _{2}}{t-t_{1}} \cdot \frac {E(t)}{E(0)}\right )\) and B = W− 1(χ(t)), using (2.6), we arrive at

$$\begin{array}{@{}rcl@{}} &&L_{2}^{\prime}(t)\le -m E(t)W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot\frac{E(t)}{E(0)}\right)+c (t-t_{1})W^{*}\left( W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot\frac{E(t)}{E(0)}\right)\right)\\ &&\hspace{0.9in}+c (t-t_{1})\chi(t)\\ &&\hspace{0.3in}\le -m E(t)W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right)+c(t-t_{1})\frac{\varepsilon_{2}}{t-t_{1}} \cdot\frac{E(t)}{E(0)}W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot\frac{E(t)}{E(0)}\right)\\ &&\hspace{0.9in}+c(t-t_{1})\chi(t). \end{array} $$
(4.50)

Using Eqs. 4.3 and 4.13, we observe that

$$\begin{array}{@{}rcl@{}} (t-t_{1})\xi(t) \chi(t)&\le& qI(t)+\xi(t)J(t)\\ &\le& qI(t)+\xi(0)J(t)\\ &\le& -cE^{\prime}(t)-cE^{\prime}(t)\\ &\le& -cE^{\prime}(t) \end{array} $$

So, multiplying Eq. 4.50 by ξ(t) and using the fact that, \(\varepsilon _{2}\frac {E(t)}{E(0)}<r_{0}\), give

$$\begin{array}{@{}rcl@{}} \xi(t)L_{2}^{\prime}(t)&\le& -m \xi(t) E(t)W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right)+c \varepsilon_{2} \xi(t) \cdot \frac{E(t)}{E(0)}W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right)\\ &&-cE^{\prime}(t), \forall t\ge t_{2}. \end{array} $$

Using the non-increasing property of ξ, we obtain, for all tt2,

$$\begin{array}{@{}rcl@{}} (\xi L_{2}+cE)^{\prime}(t)&\le& -m \xi(t) E(t)W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right)\\ &&+c \varepsilon_{2}\xi(t) \frac{E(t)}{E(0)}W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right) \end{array} $$

Therefore, by setting L3 := ξL2 + cEE, we get

$$\begin{array}{@{}rcl@{}} &&L_{3}^{\prime}(t)\le -m \xi(t) E(t)W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right)+c \varepsilon_{2} \xi(t) \cdot \frac{E(t)}{E(0)}W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right)\\ &&\hspace{0.3in} \end{array} $$

This gives, for a suitable choice of ε2,

$$L_{3}^{\prime}(t)\le -k \xi(t) \left( \frac{E(t)}{E(0)}\right)W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right), \hspace{0.3in}\forall t\ge t_{2} $$

or

$$ k\left( \frac{E(t)}{E(0)}\right)W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right)\xi(t)\leq - L_{3}^{\prime}(t), \hspace{0.3in}\forall t\ge t_{2} $$
(4.51)

An integration of Eq. 4.51 yields

$$ {\int}_{t_{2}}^{t} k \left( \frac{E(s)}{E(0)} \right)W^{\prime}\left( \frac{\varepsilon_{2}}{s-t_{1}} \cdot \frac{E(s)}{E(0)}\right)\xi(s) ds \leq - {\int}_{t_{2}}^{t} L_{3}^{\prime}(s)ds\le L_{3}(t_{2}). $$
(4.52)

Using the facts that W, W > 0 and the non-increasing property of E, we deduce that the map \(t \mapsto E(t)W^{\prime }\left (\frac {\varepsilon _{2}}{t-t_{1}} \cdot \frac {E(t)}{E(0)}\right )\) is non-increasing and consequently, we have

$$\begin{array}{@{}rcl@{}} && k \left( \frac{E(t)}{E(0)} \right)W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right){\int}_{t_{2}}^{t} \xi(s) ds\\ &&\hspace{0.3in}\leq {\int}_{t_{2}}^{t} k \left( \frac{E(s)}{E(0)} \right)W^{\prime}\left( \frac{\varepsilon_{2}}{s-t_{1}} \cdot \frac{E(s)}{E(0)}\right)\xi(s) ds\le L_{3}(t_{2}),\hspace{0.2in}\forall t\ge t_{2} \end{array} $$
(4.53)

Multiplying each side of Eq. 4.53 by \(\frac {1}{t-t_{1}}\), we have

$$ k \left( \frac{1}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right) W^{\prime}\left( \frac{\varepsilon_{2}}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right){\int}_{t_{2}}^{t} \xi(s) ds \leq \frac{k_{2}}{t-t_{1}},\hspace{0.3in}\forall t\ge t_{2} $$
(4.54)

Next, we set W2(s) = sW(ε2s) which is strictly increasing, then we obtain,

$$ k W_{2} \left( \frac{1}{t-t_{1}} \cdot \frac{E(t)}{E(0)}\right) {\int}_{t_{2}}^{t} \xi(s) ds \leq \frac{k_{2}}{t-t_{1}},\hspace{0.3in}\forall t\ge t_{2} $$
(4.55)

Finally, for two positive constants k2 and k3, we obtain

$$ E(t) \leq k_{3} (t-t_{1}) {W_{2}}^{-1} \left( \frac{k_{2}}{(t-t_{1}){\int}_{t_{2}}^{t}\xi(s) ds} \right). $$
(4.56)

This finishes the proof. □

Example 4.7

The following examples illustrate our results:

  1. 1.

    h0 and G are linear

    Let g(t) = aeb(1 + t), where b > 0 and a > 0 is small enough so that Eq. 2.1 is satisfied, then g(t) = −ξ(t)G(g(t)) where G(t) = t and ξ(t) = b. For the frictional nonlinearity, assume that h0(t) = ct and \(H(t)=\sqrt {t} h_{0}(\sqrt {t})=ct\). Therefore,, we can use Eq. 4.21 to deduce

    $$ E(t) \leq c_{1} e^{-c_{2}t} $$
    (4.57)

    which is the exponential decay.

  2. 2.

    h0 is linear and G is non-linear

    Let \(g(t)=a e^{-t^{q}}\), where 0 < q < 1 and a > 0 is small enough so that g satisfies (2.1), then g(t) = −ξ(t)G (g(t)) where ξ(t) = 1 and \(G(t)=\frac {q^{t}}{\left (ln (a/t)\right )^{\frac {1}{q}-1}}\). For, the boundary feedback, let h0(t) = ct, and \(H(t)=\sqrt {t} h_{0}(\sqrt {t})=ct\). Since

    $$\begin{array}{@{}rcl@{}} &&G^{\prime}(t)=\frac{(1-q)+q ln (a/t)}{\left( ln (a/t)\right)^{1/q}}\\ &&\hspace{0.45in}\text{and}\\ &&G^{\prime\prime}(t)=\frac{(1-q) \left( ln (a/t)+ 1/q \right)}{\left( ln (a/t)\right)^{\frac{1}{q}+ 1}}. \end{array} $$

    then the function G satisfies the condition (A1) on (0, r1] for any 0 < r1 < a.

    $$\begin{array}{@{}rcl@{}} G_{1}(t)&=&{\int}_{t}^{r_{1}}\frac{1}{sG^{\prime}(s)}ds={\int}_{t}^{r_{1}}\frac{\left[\ln{\frac{a}{s}}\right]^{\frac{1}{q}}}{s\left[1-q+q\ln{\frac{a}{s}}\right]}ds\\ &=&{\int}_{\ln{\frac{a}{r_{1}}}}^{\ln{\frac{a}{t}}}\frac{u^{\frac{1}{q}}}{1-q+qu}du\\ &=&\frac{1}{q}{\int}_{\ln{\frac{a}{r_{1}}}}^{\ln{\frac{a}{t}}} u^{\frac{1}{q}-1}\left[\frac{u}{\frac{1-q}{q}+u}\right]du\\ &\le& \frac{1}{q}{\int}_{\ln{\frac{a}{r_{1}}}}^{\ln{\frac{a}{t}}} u^{\frac{1}{q}-1}du \le \left( \ln{\frac{a}{t}}\right)^{\frac{1}{q}}. \end{array} $$

    Then, Eq. 4.22 gives

    $$ E(t)\leq k e^{-k t^{q}} $$
    (4.58)
  3. 3.

    h0 is non-linear and G is linear

    Let g(t) = aeb(1 + t), where b > 0 and a > 0 is small enough so that Eq. 2.1 is satisfied, then g(t) = −ξ(t)G(g(t)) where G(t) = t and ξ(t) = b. Also, assume that h0(t) = ctq, where q > 1 and \(H(t)=\sqrt {t} h_{0}(\sqrt {t})=ct^{\frac {q + 1}{2}}\). Then,

    $${H_{1}}^{-1}(t)= \left( ct+ 1\right)^{\frac{-2}{q-1}}.$$

    Therefore, applying Eq. 4.35, we obtain

    $$ E(t)\leq \left( c_{1} t +c_{2} \right)^{\frac{-2}{q-1}} $$
    (4.59)
  4. 4.

    h0 is non-linear and G is non-linear

    Let \(g(t)=\frac {a}{(1+t)^{2}}\), where a is chosen so that hypothesis (2.1) remains valid. Then

    $$g^{\prime}(t)=-b G(g(t)),\hspace{0.2in}\text{with}\hspace{0.2in}G(s)=s^{\frac{3}{2}},$$

    where b is a fixed constant. For the boundary feedback, let h0(t) = ct5 and H(t) = ct3. Then,

    $$W(s)=(G^{-1}+H^{-1})^{-1}=\left( \frac{-1+\sqrt{1 + 4s}}{2}\right)^{3}$$

    and

    $$\begin{array}{@{}rcl@{}} &&W_{2}(s)=\frac{3s}{\sqrt{1 + 4s}}\left( \frac{-1+\sqrt{1 + 4s}}{2}\right)^{2}\\ &&\hspace{0.3in}=\frac{3s}{2\sqrt{1 + 4s}}+\frac{3s^{2}}{\sqrt{1 + 4s}}-\frac{3s}{2}\\ &&\hspace{0.3in}\le \frac{3s}{2}+\frac{3 s^{2}}{2\sqrt{s}}-\frac{3s}{2}=c s^{\frac{3}{2}} \end{array} $$

    Therefore, applying Eq. 4.36, we obtain

    $$E(t)\le \frac{c}{(t-t_{1})^{\frac{1}{3}}}$$