Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article Special Issues

The influence of damping on the asymptotic behavior of solution for laminated beam

  • This paper dealt with a laminated beam system along with structural damping, past history, distributed delay, and in the presence of both temperatures and micro-temperatures effects. The damping terms left the system dissipative. Employing the semigroup approach, we established the existence and uniqueness of the solution. Additionally, with the help of convenient assumptions on the kernel, we demonstrated a general decay result for the solution of the considered system, with no constraints regarding the speeds of wave propagation. The main aim was to address how specific behaviors of the system were related to memory and delays. We aimed to investigate the joint impact of an infinite memory, distributed delay and micro-temperature effects on the system. We found a new relationship between the decay rate of solution and the growth of g at infinity. The objective was to find studies that use no- trivial results and their applications to relevant problems from mathematical physics.

    Citation: Abdelkader Moumen, Fares Yazid, Fatima Siham Djeradi, Moheddine Imsatfia, Tayeb Mahrouz, Keltoum Bouhali. The influence of damping on the asymptotic behavior of solution for laminated beam[J]. AIMS Mathematics, 2024, 9(8): 22602-22626. doi: 10.3934/math.20241101

    Related Papers:

    [1] Tijani A. Apalara, Aminat O. Ige, Cyril D. Enyi, Mcsylvester E. Omaba . Uniform stability result of laminated beams with thermoelasticity of type Ⅲ. AIMS Mathematics, 2023, 8(1): 1090-1101. doi: 10.3934/math.2023054
    [2] Hicham Saber, Fares Yazid, Fatima Siham Djeradi, Mohamed Bouye, Khaled Zennir . Decay for thermoelastic laminated beam with nonlinear delay and nonlinear structural damping. AIMS Mathematics, 2024, 9(3): 6916-6932. doi: 10.3934/math.2024337
    [3] Fatima Siham Djeradi, Fares Yazid, Svetlin G. Georgiev, Zayd Hajjej, Khaled Zennir . On the time decay for a thermoelastic laminated beam with microtemperature effects, nonlinear weight, and nonlinear time-varying delay. AIMS Mathematics, 2023, 8(11): 26096-26114. doi: 10.3934/math.20231330
    [4] Cyril Dennis Enyi, Soh Edwin Mukiawa . Dynamics of a thermoelastic-laminated beam problem. AIMS Mathematics, 2020, 5(5): 5261-5286. doi: 10.3934/math.2020338
    [5] Adel M. Al-Mahdi, Maher Noor, Mohammed M. Al-Gharabli, Baowei Feng, Abdelaziz Soufyane . Stability analysis for a Rao-Nakra sandwich beam equation with time-varying weights and frictional dampings. AIMS Mathematics, 2024, 9(5): 12570-12587. doi: 10.3934/math.2024615
    [6] Soh Edwin Mukiawa . Well-posedness and stabilization of a type three layer beam system with Gurtin-Pipkin's thermal law. AIMS Mathematics, 2023, 8(12): 28188-28209. doi: 10.3934/math.20231443
    [7] Houssem Eddine Khochemane, Ali Rezaiguia, Hasan Nihal Zaidi . Exponential stability and numerical simulation of a Bresse-Timoshenko system subject to a neutral delay. AIMS Mathematics, 2023, 8(9): 20361-20379. doi: 10.3934/math.20231038
    [8] Hasan Almutairi, Soh Edwin Mukiawa . On the uniform stability of a thermoelastic Timoshenko system with infinite memory. AIMS Mathematics, 2024, 9(6): 16260-16279. doi: 10.3934/math.2024787
    [9] Soh E. Mukiawa, Tijani A. Apalara, Salim A. Messaoudi . Stability rate of a thermoelastic laminated beam: Case of equal-wave speed and nonequal-wave speed of propagation. AIMS Mathematics, 2021, 6(1): 333-361. doi: 10.3934/math.2021021
    [10] Said Mesloub, Hassan Altayeb Gadain, Lotfi Kasmi . On the well posedness of a mathematical model for a singular nonlinear fractional pseudo-hyperbolic system with nonlocal boundary conditions and frictional damping terms. AIMS Mathematics, 2024, 9(2): 2964-2992. doi: 10.3934/math.2024146
  • This paper dealt with a laminated beam system along with structural damping, past history, distributed delay, and in the presence of both temperatures and micro-temperatures effects. The damping terms left the system dissipative. Employing the semigroup approach, we established the existence and uniqueness of the solution. Additionally, with the help of convenient assumptions on the kernel, we demonstrated a general decay result for the solution of the considered system, with no constraints regarding the speeds of wave propagation. The main aim was to address how specific behaviors of the system were related to memory and delays. We aimed to investigate the joint impact of an infinite memory, distributed delay and micro-temperature effects on the system. We found a new relationship between the decay rate of solution and the growth of g at infinity. The objective was to find studies that use no- trivial results and their applications to relevant problems from mathematical physics.



    Mathematical modeling is indispensable in engineering, natural science, and applied mathematics to capture the effects of both memory and delay ingrained in the studied actualities. To this end, the inclusion of both of them is often simplified for presentation purposes, as a specific description of basic operations can be intricate for mathematical manipulation. A key question to address is how certain behaviors are related to memory and delays. In this study, we investigate the joint impact of an infinite memory, distributed delay, and micro-temperature effects on the system (1.1).

    In the current work, we study the following thermoelastic laminated beam, together with structural damping, infinite memory, distributed delay, and micro-temperatures effects:

    {ϱϖtt+G(ϕϖx)x+γθx=0,Iϱ(3ψϕ)ttD(3ψϕ)xxG(ϕϖx)mθ+drx+0g(s)(3ψϕ)xx(x,ts)ds=0,3Iϱψtt3Dψxx+3G(ϕϖx)+4δψ+4βψt+4ς2ς1|μ2(ς)|ψt(x,tς)dς=0,cθtk0θxx+m(3ψϕ)t+γϖtx+k1rx=0,αrtk2rxx+k3r+k1θx+d(3ψϕ)tx=0, (1.1)

    where

    (x,ς,t)(0,1)×(ς1,ς2)×R+,

    and the initial and boundary conditions are given by

    {ϖ(x,0)=ϖ0,ψ(x,0)=ψ0,ϕ(x,0)=ϕ0,θ(x,0)=θ0,r(x,0)=r0,x(0,1),ϖt(x,0)=ϖ1,ψt(x,0)=ψ1,ϕt(x,0)=ϕ1,x(0,1),ϖx(0,t)=ϕ(0,t)=ψ(0,t)=θ(0,t)=r(0,t)=0,t>0,ϖ(1,t)=ϕx(1,t)=ψx(1,t)=θx(1,t)=r(1,t)=0,ψt(x,t)=f0(x,t)t>0. (1.2)

    Here, ϖ denotes the transverse displacement, ϕ represents the rotation angle, ψ is relative to the amount of slip occurring along the interface, θ is the temperature difference and r is the micro-temperature vector. The coefficients δ,β,ϱ,Iϱ,G, and D, are positive constants representing the adhesive stiffness, the adhesive damping parameter, the density, the shear stiffness, the flexible rigidity and the mass moment of inertia, respectively. We denote by the positive constants c,k0,k1,k2,k3,d,γ,α,m, the physical parameters describing the coupling between the various constituents of the materials.

    Herein, ς1,ς2 are positive numbers such that 0<ς1<ς2, and μ2 is an L function satisfying the following assumption:

    ● The function μ2:[ς1,ς2]R is bounded and it fulfills

    βς2ς1|μ2(ς)|dς>0.

    To motivate our work, let us recall some earlier related results. For the problems with the Timoshenko system with/without thermal law, one can see the works [4,6,7,16,23,24] and for problems related to thermoelasticity, we mention for instance [8,10,11,13,17,18].

    We start with the laminated beam model, which has become quite popular, and both scientists and engineers are interested in it. This model is a pertinent study topic, because of the wide industry applicability of such materials. Hansen and Spies in [12] were the first to introduce the following beam with two layers by developing this mathematical model:

    {ρ1ϖtt+G(ϕϖx)x=0,ρ2(3ψϕ)ttG(ϕϖx)D(3ψϕ)xx=0,ρ3ψtt+G(ϕϖx)+43γψ+43βψtDψxx=0. (1.3)

    The laminated beam equations have produced some results so far, most of which are focused on the system's stability and existence. Provided that the assumption of equal wave speeds holds, it was demonstrated that system (1.3) is exponentially stable, when linear damping terms are incorporated in two of the three equations. However, if they are included in the three equations, then the system decays exponentially with no restriction on the speeds of wave propagations, see, for instance [1,22].

    Lately, a renewed focus on investigating the asymptotic behavior of the solutions of several thermoelastic laminated beams has grown. For more details about this topic the reader may consult [2,9,20].

    The thermoelastic laminated beam problem together with nonlinear weights and time-varying delay was the study topic of Nonato et al. in [20], where the authors considered two cases (with and without the structural damping) and proved an exponential decay result for both of them. Distributed delay is one of the main damping factors in our model. It is used to model systems in which there is a delay of uncertain duration. The physical interpretation of this term differs from the delayed differential equation, as it can take several values. For example, in incoming signals, distributed delay shortens the setup and lengthens the hold time. Even moderate distributed delay likely makes setup time negative on those inputs that are directly connected to the register.

    The infinite memory is a critical aspect in addressing problems, and it has been explored in various contexts such as the work of Liu and Zhao [14], in which they considered a thermoelastic laminated beam model with past history. The authors managed to establish both exponential and polynomial stabilities, depending on the kernel function for the system involving structural damping and with no constraint on the wave speeds. Moreover, concerning the system in the absence of structural damping, they were able to establish both exponential and polynomial stabilities, in case of equal wave speeds and lack of exponential stability in the opposite case.

    The time delays problems are one of the most significant and active research areas recently. Numerous studies have demonstrated that delay can lead to instability unless certain conditions are incorporated, and it also can lead to distinct solutions that differ from those found in prior studies. Therefore, the issue of stability for systems that involve delay is highly crucial. To learn more about this term, we refer the reader to the following papers [3,5,21].

    In [19], Nicaise and Pignotti made a study on the following wave equation, together with linear frictional damping and internal distributed delay:

    uttΔu+μ1ut+a(x)τ2τ1μ2(s)ut(ts)ds, in Ω×(0,),

    and assuming that

    aτ2τ1μ2(s)ds<μ1,

    the authors managed to prove that the solution is exponentially stable.

    Problem (1.1) is considered as a delayed system, and it is also called hereditary systems, posteffect systems, and deviating argument. Distributed delay is a physical phenomenon which is found in a multitude of applications: Many real systems whose temporal evolution is not defined from a simple vector of state (expressed in the present tense) but depends irreducibly on the history of the system. This situation is encountered in the cases-numerous-where a transport of matter, energy or information generates a "dead time" in the reaction: in information and communication technologies (high-speed communication networks, control of networked systems, quality of service in Moving Picture Experts Group (MPEG) video transmissions, tele-operated systems, parallel computing, realtime computing in robotics), in population dynamics and epidemiology (gestation or incubation time), and in mechanics (viscoelasticity). Even if the process does not intrinsically contain a post-effect, its control chain can introduce distributed delays (for example, if the sensors require a significant acquisition/transmission time). For these reasons, it seems reasonable to consider distributed delay as a universal characteristic of the interaction between man and nature (hence, of sciences for engineers). The aim of our study then concerns the interaction between the different damping terms which intervene in the qualitative properties of the energy associated to the system. Before this analysis, we must ensure the existence of unique solution and then we can pass to see the asymptotic behavior of the solution with respect to damping terms. We used classical semigroup theory to find nontrivial results regarding the well-posedness of solutions. Then, under minimal restrictions on the kernel, we found qualitative properties of the solution by contracting an appropriate Lyapnov functional. The main goal is to present fundamental and new techniques for modern models applying science and technology that can stimulate research interest for exploration of mathematical applications in real life sciences.

    The rest of the current paper is structured this way: In Section 2, we provide some resources required for our research, then highlight our major results. In Section 3, we establish the well-posedness of the system. In Section 4, we introduce some fundamental lemmas required in the proof later. In Section 5, we demonstrate our general decay result.

    In this section, we provide some materials required in the proof later, then state our major results.

    ● (A1) Let g:R+R+ be a C1 function which satisfies

    g(0)>0,Dg0=ˉl>0, where g0:=0g(s)ds. (2.1)

    ● (A2) There exists a strictly increasing convex function G:R+R+ of class C1(R+)C2(]0,+[) which satisfies

    {G(0)=G(0)=0,lim

    such that

    \sup\limits_{s\in \mathbb{R}_+}\int_0^\infty \frac{ g(s)}{ {G}^{-1}(-g^\prime(s))} \; ds +\int_0^\infty \frac{ g(s)}{ {G}^{-1}(-g^\prime(s))} \;ds < +\infty.

    Now, we present the following useful inequalities.

    Lemma 2.1. The following inequalities are valid,

    \begin{equation} \int_0^1\left[ \int_0^\infty g(s)\left((3\psi-\psi)(t)-(3\psi-\phi)(t-s)\right)ds \right]^2 dx\leq c_1 (g\diamond (3\psi-\phi)_x)(t), \end{equation} (2.2)
    \begin{equation} \int_0^1\left[ \int_0^\infty g^\prime(s)\left((3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s)\right)ds \right]^2 dx\leq-g(0)(g^\prime\diamond (3\psi-\phi)_x)(t), \end{equation} (2.3)
    \begin{equation} \int_0^1\left[ \int_0^\infty g(s)\left((3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s)\right)ds \right]^2 dx\leq g_0 (g\diamond (3\psi-\phi)_x)(t), \end{equation} (2.4)
    \begin{equation} \int_0^1\left[ \int_0^\infty g^\prime(s)\left((3\psi-\phi)(t)-(3\psi-\phi)(t-s)\right)ds \right]^2 dx\leq -c_2 (g^\prime\diamond (3\psi-\phi)_x)(t), \end{equation} (2.5)

    where c_1, \; c_2 > 0, and

    (g\diamond v)(t) = \int_0^1 \int_0^\infty g(s) (v(x, t)-v(x, t-s))^2dsdx.

    Let us start by introducing (see [19])

    \begin{equation} \begin{cases} \eta^t(x, s) = (3\psi-\phi)(x, t)-(3\psi-\phi)(x, t-s), \\ \mathcal{S}(x, p, \varsigma, t) = \psi_t(x, t-\varsigma p), \end{cases} \end{equation} (2.6)

    where

    (x, p, \varsigma, s, t)\in ((0, 1))^2\times (\varsigma_1, \varsigma_2)\times\mathbb{R}_+\times \mathbb{R}_+.

    Then, the variables \eta^t and \mathcal{S} surely satisfy

    \begin{equation} \begin{cases}\eta^t_t+\eta^t_s = (3\psi-\phi)_t, \\ \varsigma \mathcal{S}_t(x, p, \varsigma, t)+ \mathcal{S}_p(x, p, \varsigma, t) = 0, \\ \mathcal{S}(x, 0, \varsigma, t) = \psi_t(x, t). \end{cases} \end{equation} (2.7)

    Hence, system (1.1) can be rewritten as

    \begin{equation} \left\{ \begin{array}{ll} \varrho \varpi_{tt}+G(\phi-\varpi_x)_x+\gamma \theta_x = 0, \\ \\ I_\varrho (3\psi-\phi)_{tt}-D(3\psi-\phi)_{xx}-G(\phi-\varpi_x)-m \theta+dr_x+\int_0^\infty g(s) (3\psi-\phi)_{xx}(t-s)ds = 0, \\ \\ 3I_\varrho \psi_{tt}-3Ds_{xx}+3G(\phi-\varpi_x)+4\delta \psi+4\beta \psi_t +4\int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}(x, 1, \varsigma, t) d\varsigma = 0, \\ \\ c\theta_t-k_0\theta_{xx}+m (3\psi-\phi)_t+\gamma \varpi_{tx}+k_1r_x = 0, \\ \\ \alpha r_t-k_2r_{xx}+k_3 r+k_1 \theta_x+d(3\psi-\phi)_{tx} = 0, \\ \\ \eta^t_t+\eta^t_s = (3\psi-\phi)_t, \\ \\ \varsigma \mathcal{S}_t+ \mathcal{S}_p = 0. \end{array} \right. \end{equation} (2.8)

    Certainly, system (2.8) is depending on the initial and boundary conditions below

    \begin{equation} \left\{ \begin{array}{ll} \varpi(x, 0) = \varpi_0, \; \psi(x, 0) = \psi_0, \; \phi (x, 0) = \phi_0, \; \theta (x, 0) = \theta_0, \; r(x, 0) = r_0, \; x\in (0, 1), \\ \varpi_t(x, 0) = \varpi_1, \; \psi_t(x, 0) = \psi_1, \; \phi_t(x, 0) = \phi_1, \quad x\in (0, 1), \\ \varpi_x(0, t) = \phi(0, t) = \psi(0, t) = \theta(0, t) = r(0, t) = 0, \quad t > 0, \\ \varpi(1, t) = \phi_x(1, t) = \psi_x(1, t) = \theta_x(1, t) = r(1, t) = 0, \; \psi_t(x, -t) = f_0(x, t)\quad t > 0, \\ \eta^t(0, s) = \eta^t_x(1, s) = 0, \; \eta^t(x, 0) = 0, \; \eta^0(x, s) = \eta_0(x, s), \quad t, s > 0, \\ \mathcal{S}(x, p, \varsigma, 0) = f_0(x, p\varsigma ), \quad x, p\in (0, 1), \; \varsigma \in (\varsigma_1, \varsigma_2) , \;t, s > 0. \end{array} \right. \end{equation} (2.9)

    Now, let

    \begin{equation*} \begin{cases} \zeta = 3\psi-\phi, \\ \zeta(0, t) = \zeta_x(1, t) = 0, \; \zeta(x, 0) = \zeta_0, \; \zeta_t(x, 0) = \zeta_1, \; (x, t)\in (0, 1)\times \mathbb{R}_+. \end{cases} \end{equation*}

    Then, system (2.8) is equivalent to

    \begin{equation} \left\{ \begin{array}{ll} \varrho \varpi_{tt}+G(3\psi-\zeta-\varpi_x)_x+\gamma \theta_x = 0, \\ \\ I_\varrho \zeta_{tt}-D\zeta_{xx}-G(3\psi-\zeta-\varpi_x)-m \theta+dr_x+\int_0^\infty g(s) \zeta_{xx}(t-s)ds = 0, \\ \\ 3I_\varrho \psi_{tt}-3D\psi_{xx}+3G(3\psi-\zeta-\varpi_x)+4\delta \psi+4\beta \psi_t +4\int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}(x, 1, \varsigma, t) d\varsigma = 0, \\ \\ c\theta_t-k_0\theta_{xx}+m \zeta_t+\gamma \varpi_{tx}+k_1r_x = 0, \\ \\ \alpha r_t-k_2r_{xx}+k_3 r+k_1 \theta_x+d\zeta_{tx} = 0, \\ \\ \eta^t_t+\eta^t_s = \zeta_t, \\ \\ \varsigma \mathcal{S}_t+ \mathcal{S}_p = 0. \end{array} \right. \end{equation} (2.10)

    Taking advantage of (2.6), we can rewrite the second equation of (2.10) as

    I_\varrho\zeta_{tt}-\bar{l}\zeta_{xx}-G(3\psi-\zeta-\varpi_x)-m\theta+dr_x-\int_0^\infty g(s)\eta^t_{xx}(x, s)ds = 0.

    At this step, let us introduce the vector function U = (\varpi, u, \zeta, \nu, \psi, y, \theta, r, \eta^t, \mathcal{S})^T, with

    \begin{aligned} u = \varpi_t, \\ \nu = \zeta_t, \\ y = \psi_t, \end{aligned}

    then, system (2.10) becomes

    \begin{equation} \begin{cases} \frac{d}{dt} U(t) = \mathcal{A} U(t), \quad t > 0, \\ U(0) = U_0 = (\varpi_0, \varpi_1, \zeta_0, \zeta_1, \psi_0, \psi_1, \theta_0, r_0, \eta_0, f_0)^T, \end{cases} \end{equation} (2.11)

    here, \mathcal{A}: D(\mathcal{A})\subset \mathcal{H}: \rightarrow \mathcal{H} stands for a linear operator indicated by

    \mathcal{A} U = \begin{pmatrix} u\\ -\frac{1}{\varrho}\left( G(3\psi-\zeta-\varpi_x)_x+\gamma \theta_x\right)\\ \nu\\ \frac{1}{I_\varrho}\left( \bar{l}\zeta_{xx}+G(3\psi-\zeta-\varpi_x)+m \theta-dr_x+\int_0^\infty g(s) \eta^t_{xx}(x, s) ds\right)\\ y\\\\ \frac{1}{I_\varrho}\left( D\psi_{xx}-G(3\psi-\zeta-\varpi_x)-\frac{4}{3}\delta \psi-\frac{4}{3}\beta y -\frac{4}{3}\int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}(x, 1, \varsigma, t) d\varsigma\right)\\\\ \frac{1}{c}\left( k_0\theta_{xx}-m\nu-\gamma u_x-k_1r_x\right)\\\\ \frac{1}{\alpha} (k_2r_{xx}-k_3r-k_1\theta_x-d\nu_x)\\\\ \nu-\eta^t_s \\ \\ -\frac{1}{\varsigma} \mathcal{S}_p \end{pmatrix}.

    Now, we shall consider the ensuing energy space

    \begin{array}{c} \mathcal{H} = \tilde{\mathbb{J}}^1_*(0, 1)\times L^2(0, 1)\times \mathbb{J}^1_*(0, 1)\times L^2(0, 1)\times \mathbb{J}^1_*(0, 1)\times L^2(0, 1)\times L^2(0, 1)\times L^2(0, 1)\\ \times \mathbb{L}_g \times L^2\left( (0, 1)\times (0, 1)\times(\varsigma_1, \varsigma_2) \right), \end{array}

    where

    \begin{array}{ll} \mathbb{J}^1_*(0, 1)& = \left\{ \varphi\in H^1(0, 1) :\; \varphi(0) = 0\right\}, \\\\ \tilde{\mathbb{J}}^1_*(0, 1)& = \left\{ \varphi\in H^1(0, 1) :\; \varphi(1) = 0\right\}, \\\\ \mathbb{J}^2_*(0, 1)& = H^2(0, 1)\cap \mathbb{J}^1_*(0, 1), \\\\ \tilde{\mathbb{J}}^2_*(0, 1)& = H^2(0, 1)\cap \tilde{\mathbb{J}}^1_*(0, 1), \end{array}

    and

    \mathbb{L}_g = \left\{ \varphi : \mathbb{R}_+\rightarrow \mathbb{J}^1_*(0, 1) , \; \int_0^1 \int_0^\infty g(s) \varphi_x^2 \; dsdx < \infty\right\}.

    For the space \mathbb{L}_g, we take the following inner product

    \langle \varphi_1, \varphi_2 \rangle_{ \mathbb{L}_g} = \int_0^1 \int_0^\infty g(s) \varphi_{1x}\varphi_{2x} \; dsdx.

    Furthermore, we consider the following domain

    \mathcal{L}_g \left( \mathbb{R}_+, \mathbb{J}^1_*(0, 1)\right) = \left\{ \eta^t \in \mathbb{L}_g, \; \eta^t_s \in \mathbb{L}_g, \; \eta^t(x, 0) = 0 \right\}.

    Then, we introduce

    \begin{eqnarray} \langle U, \bar{U}\rangle_\mathcal{H}& = & \varrho\int_0^1 u\bar{u} \;dx+I_\varrho \int_0^1 \nu \bar{\nu}\; dx+3I_\varrho \int_0^1 y \bar{y}\; dx +c \int_0^1 \theta \bar{\theta} \;dx+\alpha \int_0^1 r \bar{r}\; dx\\ &+&\bar{l} \int_0^1\zeta_x \bar{\zeta}_x \;dx +G \int_0^1 (3\psi-\zeta-\varpi_x)(3\bar{\psi}-\bar{\zeta}-\bar{\varpi}_x) dx+4\delta \int_0^1 \psi \bar{\psi}\; dx \\ &+&3D\int_0^1 \psi_x \bar{\psi}_x \;dx+\int_0^1 \int_0^1 g(s) \eta^t_x(x, t)\bar{\eta}^t_x(x, s) \; ds dx\\ &+&4\int_0^1 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \varsigma \vert \mu_2(\varsigma)\vert \mathcal{S} \bar{ \mathcal{S} } \; d\varsigma dpdx. \end{eqnarray} (2.12)

    We deduce that \mathcal{H} together with (2.12) is a Hilbert space, once we do that, we define D(\mathcal{A}) by

    D(\mathcal{A}) = \begin{Bmatrix} U\in \mathcal{H}:\; \varpi \in \tilde{\mathbb{J}}^2_*(0, 1); \; \zeta, \; \psi \in \mathbb{J}^2_*(0, 1);\\ u\in \tilde{\mathbb{J}}^1_*(0, 1); \; \nu, \; y\in \mathbb{J}^1_*(0, 1), \; \theta \in \mathbb{J}^1_*(0, 1), \; \theta_t \in L^2(0, 1);\\ r\in H^2(0, 1)\cap H^1_0(0, 1), \; \eta^t \in \mathcal{L}_g \left( \mathbb{R}_+, \mathbb{J}^1_*(0, 1)\right);\\ \mathcal{S}, \; \mathcal{S}_p \in L^2\left( (0, 1)\times (0, 1)\times(\varsigma_1, \varsigma_2) \right), \; \mathcal{S}(x, 0, \varsigma, t) = y\\ \varpi_x(0, t) = \zeta_x(1, t) = \psi_x(1, t) = \theta_x(1, t) = \eta^t_x(1, s) = 0. \end{Bmatrix}.

    Obviously, D(\mathcal{A}) is dense in \mathcal{H}.

    Now, we are ready to state our results.

    Theorem 2.1. Let U_0 \in D(\mathcal{A}), then problem (2.9)-(2.10) admits a unique solution

    U\in C(\mathbb{R}_+, D(\mathcal{A}))\cap C^1(\mathbb{R}_+, \mathcal{H}).

    In addition, if U_0\in \mathcal{H}, then

    U\in C(\mathbb{R}_+, \mathcal{H}).

    We give the energy of the solution of problem (2.8)-(2.9) by

    \begin{equation} {E}(t) = \frac{1}{2} \int_0^1 \left\{ \varrho \varpi^2_t+G(\phi-\varpi_x)^2+I_\varrho (3\psi_t-\phi_t)^2+\bar{l}(3\psi_x-\phi_x)^2+3I_\varrho \psi_t^2\right. \\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\left.+3D\psi_x^2+4\delta \psi^2+c \theta^2+\alpha r^2\right\} dx +\frac{1}{2} \left(g\diamond (3\psi-\phi)_x\right)(t)\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+2 \int_0^1\int_0^1\int_{\varsigma_1}^{\varsigma_2} \varsigma \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx. \end{equation} (2.13)

    Then, we have the following stability result.

    Theorem 2.2. Let (\varpi, \phi, \psi, \theta, r, \eta^t, \mathcal{S}) be the solution of (2.8)-(2.9), suppose that ( \boldsymbol{T} ), ( \boldsymbol{A}_1 ) and ( \boldsymbol{A}_2 ) hold. Then, for any initial data U_0\in D(\mathcal{A}) satisfying, for some p_0\geq0,

    \begin{equation} \int_0^1 \eta^2_{0x}(x, s) dx\leq p_0, \quad \mathit{\text{for all}}\; s > 0, \end{equation} (2.14)

    there exist positive constants \alpha_1, \alpha_2, and \alpha_3, such that

    \begin{equation} {E}(t)\leq \alpha_1 {G}^{-1}_*(\alpha_2 t+\alpha_3), \end{equation} (2.15)

    where

    {G}^{-1}_*(t) = \int_t^\infty \frac{ds}{ {G}_0(s)} , \quad {G}_0(t) = t {G}^\prime(\epsilon_0t), \quad \mathit{\text{for all}} \; \epsilon_0\geq 0.

    In this part, we utilize the semigroup approach to prove our well-posedness result.

    Proof of Theorem 2.1. Let's us establish the dissipativity of \mathcal{A}. By (2.12) and for any U\in D(\mathcal{A}), we have

    \;\;\;\langle \mathcal{A}U, U\rangle_\mathcal{H} = -4\beta \int_0^1 y^2 \; dx -k_3\int_0^1 r^2 \; dx-k_2 \int_0^1 r_x^2 \;dx -k_0 \int_0^1 \theta^2_x \; dx\\ -4 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert y \mathcal{S}(x, 1, \varsigma, t) \; d\varsigma dx -4 \int_0^1 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}_p \mathcal{S} \; d\varsigma dp dx\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+\frac{1}{2} (g^\prime \diamond \zeta_x)(t)\leq 0.

    One can notice that

    \begin{equation} \begin{aligned}- 4\int_0^1 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}_p \mathcal{S} \; d\varsigma dp dx& = -2 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \int_0^1 \vert \mu_2(\varsigma)\vert \partial_p \mathcal{S}^2 \; dp d\varsigma dx\\ & = -2 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, 1, \varsigma, t) \; d\varsigma dx\\ &+2 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, 0, \varsigma, t) \; d\varsigma dx. \end{aligned} \end{equation} (3.1)

    Applying Youg's inequality, we obtain

    \begin{equation*} \begin{aligned}- 4\int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert y \mathcal{S}(x, 1, \varsigma, t) \; d\varsigma dx&\leq 2 \left( \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \; d\varsigma \right)\int_0^1 y^2 \; dx\\ &+ 2 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, 1, \varsigma, t) \; d\varsigma dx, \end{aligned} \end{equation*}

    therefore, by (T) and given \mathcal{S}(x, 0) = y, we end up with

    \begin{array}{c} \langle \mathcal{A}U, U\rangle_\mathcal{H} = -4\left(\beta- \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \; d\varsigma \right) \int_0^1 y^2 \; dx -k_3\int_0^1 r^2 \; dx-k_2 \int_0^1 r_x^2 \;dx \\ -k_0 \int_0^1 \theta^2_x \; dx+\frac{1}{2} (g^\prime \diamond \zeta_x)(t)\leq 0. \end{array}

    Thereby, \mathcal{A} is dissipative.

    Thereafter, we establish the surjectivity of ( I-\mathcal{A} ), that is, we show that

    \begin{array}{c} \forall H = (h_1, h_2, h_3, h_4, h_5, h_6, h_7, h_8, h_9, h_{10})^T \in \mathcal{H}, \; \exists U\in D( \mathcal{A}):\;\\ (I-\mathcal{A})U = H. \end{array} (3.2)

    Now, we have

    \begin{equation} \begin{cases} \varpi-u = h_1, \\ \varrho u+G(3\psi-\zeta-\varpi_x)_x+\gamma \theta_x = \varrho h_2, \\ \zeta-\nu = h_3, \\ I_\varrho\nu-\bar{l}\zeta_{xx}-G(3\psi-\zeta-\varpi_x)-m \theta+dr_x-\int_0^\infty g(s) \eta^t_{xx}(x, s) ds = I_\varrho h_4, \\ \psi-y = h_5, \\ 3 I_\varrho y-3D\psi_{xx}+3G(3\psi-\zeta-\varpi_x)+4\delta \psi+4\beta y+4\int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}(x, 1, \varsigma, t) d\varsigma = 3 I_\varrho h_6, \\ c\theta-k_0\theta_{xx}+m \nu+\gamma u_x+k_1r_x = c h_7, \\ (\alpha +k_3)r-k_2 r_{xx}+k_1\theta_x+d\nu_x = \alpha h_8, \\ \eta^t-\nu+\eta^t_s = h_9, \\ \varsigma \mathcal{S}+\mathcal{S}_p = \varsigma h_{10}. \end{cases} \end{equation} (3.3)

    Solving (3.3) _{10} and using \mathcal{S}(x, 0, \varsigma, t) = y(x, t), we find

    \mathcal{S}(x, p, \varsigma, t) = y(x, t)e^{-\varsigma p}+\varsigma e^{-\varsigma p} \int_0^p e^{\varsigma\sigma } h_{10}(x, \sigma, \varsigma, t) \; d\sigma.

    Hence,

    \begin{equation} \mathcal{S}(x, 1, \varsigma, t) = y(x, t)e^{-\varsigma } +\varsigma e^{-\varsigma } \int_0^1 e^{\varsigma\sigma } h_{10}(x, \sigma, \varsigma, t) \; d\sigma. \end{equation} (3.4)

    Now, we solve Eq (3.3) _9 , and we find

    \begin{equation} \eta^t = e^{-s}\int_0^s e^\sigma (\nu+h_9(\sigma))\; d\sigma. \end{equation} (3.5)

    Inserting (3.5), (3.4), and

    \begin{cases} u = \varpi-h_1, \\ \nu = \zeta-h_3, \\ y = \psi-h_5, \\ \end{cases}

    into (3.3) _2 , (3.3) _4 , (3.3) _6 , (3.3) _7 and (3.3) _8 , we get

    \begin{equation} \begin{cases} \varrho \varpi +G(3\psi-\zeta-\varpi_x)_x+\gamma \theta_x = \varrho(h_1+h_2), \\ I_\varrho \zeta-\left(\bar{l}+\int_0^\infty (1-e^{-s})g(s) ds\right)\zeta_{xx}-G(3\psi-\zeta-\varpi_x)-m \theta+dr_x = I_\varrho (h_3+h_4)+ \tilde{h}, \\ \mu_1 \psi-3D\psi_{xx}+3G(3\psi-\zeta-\varpi_x) = \tilde{\mu}_1h_5+3I_\varrho h_6-4\int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \varsigma e^{-\varsigma}\int_0^1 e^{\varsigma\sigma} h_{10} \; d \sigma d\varsigma, \\ c \theta -k_0\theta_{xx}+m \zeta +\gamma \varpi_x+k_1r_x = \gamma h_{1x}+c h_7+mh_3, \\ (\alpha+k_3) r -k_2r_{xx}+k_1 \theta_x+d \zeta_x = \alpha h_8+dh_{3x}, \end{cases} \end{equation} (3.6)

    where

    \tilde{h} = \int_0^\infty g(s)\int_0^s e^{\sigma-s}(h_9-h_3)_{xx}\; d\sigma ds,
    \mu_1 = 3I_\varrho+4\delta+4\beta +4\int_{\varsigma_1}^{\varsigma_2} e^{-\varsigma}\vert \mu_2(\varsigma)\vert \; d\varsigma,

    and

    \tilde{\mu}_1 = 3I_\varrho+4\beta+4\int_{\varsigma_1}^{\varsigma_2} e^{-\varsigma}\vert \mu_2(\varsigma)\vert \; d\varsigma.

    We take the following variational formulation to solve (3.6):

    \begin{equation} {Q}((\varpi, \zeta, \psi, \theta, r), (\bar{\varpi}, \bar{\zeta}, \bar{\psi}, \bar{\theta}, \bar{r})) = L(\bar{\varpi}, \bar{\zeta}, \bar{\psi}, \bar{\theta})), \quad \forall (\bar{\varpi}, \bar{\zeta}, \bar{\psi}, \bar{\theta}, \bar{r}))\in X, \end{equation} (3.7)

    where,

    X = \tilde{\mathbb{J}}^1_*(0, 1)\times \mathbb{J}^1_*(0, 1)\times \mathbb{J}^1_*(0, 1)\times L^2(0, 1)\times H^1_0(0, 1),

    is a Hilbert space endowed with

    \Vert (\varpi, \zeta, \psi, \theta, r) \Vert^2_X = \Vert 3\psi-\zeta-\varpi_x \Vert^2_2+ \Vert \varpi \Vert^2_2+ \Vert \zeta_x \Vert^2_2+ \Vert \psi_x \Vert_2^2+\Vert \theta_x \Vert^2_2 +\Vert r \Vert^2_2+\Vert r_x \Vert^2_2.

    As a part of this step, we provide definitions for both the bilinear form {Q}: X\times X\rightarrow \mathbb{R} and the linear form L: X\rightarrow \mathbb{R} as follows:

    \begin{equation} \begin{aligned} & {Q}((\varpi, \zeta, \psi, \theta, r), (\bar{\varpi}, \bar{\zeta}, \bar{\psi}, \bar{\theta}, \bar{r}))\\ = & \varrho \int_0^1 \varpi \bar{\varpi} \; dx+I_\varrho \int_0^1 \zeta \bar{\zeta} \; dx+\mu_1 \int_0^1 \psi \bar{\psi} \; dx +c \int_0^1 \theta \bar{\theta} \; dx\\+&(\alpha+k_3)\int_0^1 r\bar{r}\; dx+k_2\int_0^1 r_x\bar{r}_x\; dx+\gamma \int_0^1 (\theta_x \bar{\varpi}+\varpi_x \bar{\theta}) \; dx\\+&k_0 \int_0^1 \theta_x \bar{\theta}_x \; dx +G \int_0^1 (3\psi-\zeta-\varpi_x)( 3\bar{\psi}-\bar{\zeta}- \bar{\varpi}_x)\; dx\\ +&\left(\bar{l}+\int_0^\infty (1-e^{-s}) g(s) \; ds\right)\int_0^1 \zeta_x \bar{\zeta}_x\; dx+3D \int_0^1 \psi_x \bar{\psi}_x \; dx\\ +& d \int_0^1 (r_x \bar{\zeta}+\zeta_x \bar{r}) \; dx+k_1 \int_0^1 (r_x \bar{\theta}+\bar{r} \theta_x) \; dx\\ +&m \int_0^1 (\zeta \bar{\theta}-\bar{\zeta} \theta) \; dx, \end{aligned} \end{equation} (3.8)

    and

    \begin{aligned} L(\bar{\varpi}, \bar{\zeta}, \bar{\psi}, \bar{\theta}, \bar{r})& = \varrho \int_0^1 \bar{\varpi}(h_1+h_2)\; dx+I_\varrho\int_0^1 \bar{\zeta} (h_3+h_4)\; dx+\int_0^1 \bar{\zeta} \tilde{h}\; dx\\ &+\int_0^1 \bar{\theta}(\gamma h_{1x}+m h_3+c h_7) \; dx +\int_0^1 \bar{r} (\alpha h_8+dh_{3x})\; dx \\& + \int_0^1 \bar{\psi}\left[ \tilde{\mu}_1h_5+3I_\varrho h_6 -4\int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \varsigma e^{-\varsigma}\int_0^1 e^{\varsigma\sigma} h_{10} \; d \sigma d\varsigma\right]\; dx. \end{aligned}

    We can easily prove the continuity of {Q} and L . Moreover, from (3.8) together with integration by parts, we arrive at

    \begin{aligned} & {Q}((\varpi, \zeta, \psi, \theta, r), (\varpi, \zeta, \psi, \theta, r))\\ = & \varrho \int_0^1 \varpi^2 \; dx+I_\varrho \int_0^1 \zeta^2 \; dx+\mu_1 \int_0^1 \psi^2 \; dx +c \int_0^1 \theta^2 \; dx\\+&(\alpha+k_3)\int_0^1 r^2\; dx+k_2\int_0^1 r_x^2\; dx+k_0 \int_0^1 \theta_x^2 \; dx \\+&G \int_0^1 (3\psi-\zeta-\varpi_x)^2\; dx+3D \int_0^1 \psi_x^2 \; dx\\ +&\left(\bar{l}+\int_0^\infty (1-e^{-s}) g(s) \; ds\right)\int_0^1 \zeta_x^2\; dx\\ &\geq M \Vert (\varpi, \zeta, \psi, \theta, r) \Vert^2_X, \; M > 0.\\ \end{aligned}

    From this, we conclude the coercivity of {Q} . It follows from the Lax-Milgram lemma that (3.6) admits a unique solution satisfying

    \varpi \in \tilde{\mathbb{J}}^1_*(0, 1),
    \zeta, \; \psi \in \mathbb{J}^1_*(0, 1),
    \theta \in L^2(0, 1),

    and

    r\in H^1_0(0, 1).

    If we substitute \varpi, \zeta, and \psi into (3.3) _1 , (3.3) _3 and (3.3) _5 , we find

    u\in\tilde{\mathbb{J}}^1_*(0, 1),

    and

    \nu, \; y \in \mathbb{J}^1_*(0, 1).

    In addition, taking (\bar{\zeta}, \bar{\psi}, \bar{\theta}, \bar{r})\equiv(0, 0, 0, 0) \in \left(\mathbb{J}^1_*(0, 1) \right)^2\times L^2(0, 1)\times H^1_0(0, 1), (3.7) becomes

    \begin{equation} G \int_0^1 \bar{\varpi} \varpi_{xx} \; dx = \int_0^1 \bar{\varpi}\left( \varrho \varpi +3G\psi_x-G\zeta _x+\gamma \theta_x- \varrho (h_1+h_2) \right)dx, \end{equation} (3.9)

    for all \bar{\varpi} \in \tilde{\mathbb{J}}^1_*(0, 1), which indicates that

    \begin{equation} G\varpi_{xx} = \varrho \varpi+3G\psi_x-G\zeta_x+\gamma \theta_x- \varrho (h_1+h_2)\in L^2(0, 1). \end{equation} (3.10)

    The standard elliptic regularity implies that

    \varpi \in \tilde{\mathbb{J}}^2_*(0, 1).

    We note that (3.9) remains valid for \bar{\varphi}\in C^1([0, 1])\subset \tilde{\mathbb{J}}^1_*(0, 1), that is \bar{\varphi}(1) = 0 . Then, we obtain

    G \int_0^1 \bar{\varphi}_x \varpi_{x} \; dx = \int_0^1 \bar{\varphi}\left( -\varrho \varpi -3G\psi_x+G\zeta _x-\gamma \theta_x+ \varrho (h_1+h_2) \right)dx.

    Integrating by parts, it follows that

    \varpi_x(0)\bar{\varphi}(0) = 0, \quad \text{ for all } \; \bar{\varphi}\in C^1([0, 1]).

    Hence

    \varpi_x(0) = 0.

    Likewise, we show that

    \begin{array}{c} (\zeta, \psi)\in \left( \mathbb{J}^2_*(0, 1)\right)^2, \quad \theta \in \mathbb{J}^1_*(0, 1), \quad r\in H^2(0, 1)\cap H^1_0(0, 1), \\ \text{ and } \quad\zeta_x(1) = \psi_x(1) = \theta_x(1) = 0.\end{array}

    The standard elliptic regularity guarantees the existence of a unique U \in D(\mathcal{A}) which fulfills (3.2). Thereby, \mathcal{A} is surjective.

    As a consequence, we infer that \mathcal{A} is a maximal dissipative operator. Then, the well-posedness result follows using Lumer-Philips theorem [15].

    The main purpose of this section is to establish the essential practical lemmas required to prove our stability results. To attain this goal, we apply a specific approach known as the multiplier technique, which enables us to prove the stability results of problem (2.8). Nevertheless, this method necessitates creating an appropriate Lyapunov functional equivalent to the energy and we will clarify on this in the next section. To simplify matters, we will employ \chi_* > 0 to represent a generic constant.

    Lemma 4.1. Let (\varpi, \phi, \psi, \theta, r, \eta^t, \mathcal{S}) be the solution of (2.8) and (2.9), then, the energy functional satisfies

    \begin{array}{c} \frac{d}{dt} {E}(t) \leq -m_0 \int_0^1 \psi_t^2 dx-k_0\int_0^1 \theta_{x}^2 dx-k_2\int_0^1 r_x^2 dx-k_3\int_0^1 r^2 dx\\ \;\;\;\;\;+\frac{1}{2}(g^\prime\diamond(3\psi_x-\phi_x))(t)\leq 0, \quad \mathit{\text{where}} \; m_0 > 0.\end{array} (4.1)

    Proof. As a start, we multiply (2.8) _1 , (2.8) _2 , (2.8) _3 , (2.8) _4 and (2.8) _5 by \varpi_t, \; (3\psi_t-\phi_t), \; \psi_t, \; \theta and r respectively, then, we integrate over (0, 1) and use integration by parts together with boundary conditions (2.9) and (2.6) to find

    \begin{array}{l} \frac{1}{2} \frac{d}{dt} \int_0^1 \left\{ \varrho \varpi^2_t+G(\phi-\varpi_x)^2+I_\varrho (3\psi_t-\phi_t)^2+\bar{l}(3\psi_x-\phi_x)^2+3I_\varrho \psi_t^2+3D\psi_x^2\right. \\ \;\;\;\;\;\;\left.+4\delta \psi^2+c\theta^2+\alpha r^2\right\} dx +4\beta \int_0^1 \psi_t^2 dx +k_0 \int_0^1 \theta^2_x dx+k_2 \int_0^1 r_x^2 dx\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+k_3 \int_0^1 r^2 dx-\int_0^1(3\psi-\phi)_t\int_0^\infty g(s) \eta^t_{xx} (x, s) dsdx \\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+4\int_0^1 \int_{\varsigma_1}^{\varsigma_2} \psi_t \vert \mu_2(\varsigma) \vert \mathcal{S} (x, 1, \varsigma, t) \; d\varsigma dx = 0. \end{array} (4.2)

    It follows from the sixth equation in (2.8) and the integration by parts that

    \begin{equation} \begin{aligned} &\int_0^1(3\psi-\phi)_t\int_0^\infty g(s) \eta^t_{xx} (x, s) dsdx\\ = & \int_0^\infty g(s)\left( \int_0^1 \eta^t_t \eta^t_{xx} (x, s) dx \right)ds\\&+ \int_0^\infty g(s)\left( \int_0^1 \eta^t_s \eta^t_{xx} (x, s) dx \right)ds\\ & = -\frac{1}{2} \frac{d}{dt} (g\diamond (3\psi_x-\phi_x))(t)\\&+\frac{1}{2} (g^\prime\diamond (3\psi_x-\phi_x))(t). \end{aligned} \end{equation} (4.3)

    Applying Young's inequality, we find

    \begin{equation} \begin{aligned} \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \psi_t \vert \mu_2(\varsigma) \vert \mathcal{S} (x, 1, \varsigma, t) \; d\varsigma dx &\leq \frac{1}{2} \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma) \vert \mathcal{S}^2 (x, 1, \varsigma, t) \; d\varsigma dx\\ &+ \frac{1}{2} \left( \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma) \vert d\varsigma\right)\int_0^1 \psi_t^2 \; dx. \end{aligned} \end{equation} (4.4)

    Next, we multiply (2.8) _7 by \mathcal{S} \vert \mu_2(\varsigma) \vert and integrate the result over (0, 1)\times(0, 1)\times(\varsigma_1, \varsigma_2). We get

    \begin{equation} \begin{aligned} &\frac{1}{2}\frac{d}{dt} \int_0^1 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \varsigma \vert \mu_2(\varsigma)\vert \mathcal{S}^2 (x, p, \varsigma, t)\; d\varsigma dp dx\\& = - \int_0^1 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}_p \mathcal{S} (x, p, \varsigma, t)\; d\varsigma dp dx\\ & = - \frac{1}{2} \int_0^1 \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \partial_p \mathcal{S}^2(x, p, \varsigma, t) \; d\varsigma dp dx\\ & = \frac{1}{2} \left( \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma) \vert d\varsigma\right)\int_0^1 \psi_t^2 \; dx\\&- \frac{1}{2} \int_0^1 \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, 1, \varsigma, t) \; d\varsigma dx, \\ \end{aligned} \end{equation} (4.5)

    which, together with (4.2)–(4.4) and (T) gives us

    \begin{array}{c} \frac{d}{dt} {E}(t) \leq -4\left(\beta- \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma) \vert d\varsigma\right)\int_0^1 \psi_t^2 dx-k_0\int_0^1 \theta_{x}^2 dx-k_2\int_0^1 r_x^2 dx-k_3\int_0^1 r^2 dx \\ \;\;\;\;+\frac{1}{2}(g^\prime\diamond(3\psi_x-\phi_x))(t)\leq 0.\end{array}

    We have then reached the desired result.

    Lemma 4.2. Consider the functional

    \begin{equation} {I}_1(t): = -I_\varrho \int_{0}^1 (3\psi_t-\phi_t)\int_0^\infty g(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))dsdx, \end{equation} (4.6)

    then, it satisfies

    \begin{equation} \begin{aligned} {I}_1^\prime(t)&\leq \frac{-I_\varrho g_0}{2}\int_{0}^1 (3\psi_t-\phi_t)^2 dx+\epsilon_1 \int_{0}^1 (3\psi_x-\phi_x)^2dx\\&+\epsilon_1\int_{0}^1 (\phi-\varpi_x)^2 dx+\epsilon_1 \int_{0}^1 \theta_x^2 dx+\chi_*\int_{0}^1 r^2 dx\\ &+\chi_*\left( 1+\frac{1}{\epsilon_1}\right)(g\diamond(3\psi_x-\phi_x))(t)-\chi_*(g^\prime\diamond(3\psi_x-\phi_x))(t), \; \forall\epsilon_1 > 0. \end{aligned} \end{equation} (4.7)

    Proof. First, we notice that

    \begin{equation} \begin{aligned} &\partial_t \left( \int_0^\infty g(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))ds \right)\\& = \partial_t\left( \int_{-\infty}^t g(t-s)((3\psi-\phi)(t)-(3\psi-\phi)(s) )ds \right)\\& = \int_{-\infty}^t g^\prime(t-s)((3\psi-\phi)(t)-(3\psi-\phi)(s) )ds\\&+\int_{-\infty}^t g(t-s)(3\psi-\phi)_t(t) ds\\ & = \int_0^\infty g^\prime(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))ds\\ &+g_0(3\psi-\phi)_t(t). \end{aligned} \end{equation} (4.8)

    Next, we proceed by differentiating {I}_1(t) and using both (2.8) _2 and relation (4.8), then, integrating by parts, we get

    \begin{equation} \begin{aligned} {F}_1^\prime (t) = &-I_\varrho\int_{0}^1 (3\psi_{tt}-\phi_{tt})\int_0^\infty g(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))dsdx\\&-I_\varrho \int_{0}^1 (3\psi_t-\phi_t) \frac{\partial}{\partial t} \left(\int_0^\infty g(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))dsdx\right)\\ & = D \int_{0}^1 (3\psi_x-\phi_x) \int_0^\infty g(s)((3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s))dsdx\\&-G \int_{0}^1 (\phi-\varpi_x) \int_0^\infty g(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))dsdx\\&-m \int_{0}^1 \theta \int_0^\infty g(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))dsdx-I_\varrho g_0\int_{0}^1 (3\psi_t-\phi_t)^2 dx\\&-I_\varrho\int_{0}^1 (3\psi_t-\phi_t) \int_0^\infty g^\prime(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))dsdx\\&-d \int_{0}^1 r \int_0^\infty g(s)((3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s))dsdx\\ &-\int_{0}^1 \left( \int_0^\infty g(s)(3\psi-\phi)_x(x, t-s)ds \right)\\ &\times \left( \int_0^\infty g(s) ( (3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s) )ds\right)\; dx. \end{aligned} \end{equation} (4.9)

    The last term in (4.9) can be rewritten as

    \begin{equation} -\int_{0}^1 \left( \int_0^\infty g(s)(3\psi-\phi)_x(x, t-s)ds\right) \left( \int_0^\infty g(s) ( (3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s) )ds\right) dx\\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\; = \int_{0}^1 \left( \int_0^\infty g(s) ( (3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s) )ds \right)^2 dx\\ \;\;\;\;\;\;\;\;\;\;-g_0 \int_{0}^1 (3\psi-\phi)_x \left( \int_0^\infty g(s) ( (3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s) )ds\right) dx. \end{equation} (4.10)

    Now, replacing (4.10) into (4.9), leads to

    \begin{equation*} \begin{aligned} {F}_1^\prime (t) & = \bar{l} \int_{0}^1 (3\psi_x-\phi_x) \int_0^\infty g(s)((3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s))dsdx\\&-G \int_{0}^1 (\phi-\varpi_x) \int_0^\infty g(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))dsdx\\&-m \int_{0}^1 \theta \int_0^\infty g(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))dsdx-I_\varrho g_0\int_{0}^1 (3\psi_t-\phi_t)^2 dx\\&-I_\varrho\int_{0}^1 (3\psi_t-\phi_t) \int_0^\infty g^\prime(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))dsdx\\&-d \int_{0}^1 r \int_0^\infty g(s)((3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s))dsdx\\ &+\int_{0}^1 \left( \int_0^\infty g(s) ( (3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s) )ds \right)^2\; dx. \end{aligned} \end{equation*}

    Finally, applying Young's inequality and making use of Lemma 2.1, we obtain (4.7).

    Lemma 4.3. Consider the functional

    \begin{equation*} {I}_2(t): = -c\varrho \int_0^1 \varpi_t \left( \int_x^1 \theta (y) dy\right)dx, \end{equation*}

    then, it satisfies

    \begin{array}{c} {I}_2^\prime(t)\leq \frac{-\gamma \varrho}{2} \int_0^1 \varpi_t^2\; dx+\chi_* \int_0^1 (3\psi_t-\phi_t)^2 \; dx+\epsilon_2 \int_0^1 (\phi-\varpi_x)^2\; dx\\ \;\;+\chi_* \int_0^1 r^2 \; dx+\chi_* \left( 1+\frac{1}{\epsilon_2}\right)\int_0^1 \theta_x^2 \; dx, \; \forall \epsilon_2 > 0. \end{array} (4.11)

    Proof. Simple calculations, using (2.8) _1 , (2.8) _4 and integration by parts, we get

    \begin{aligned} {I}^\prime_2(t)& = -c\varrho \int_0^1 \varpi_{tt} \left( \int_x^1 \theta (y) dy\right)dx-c\varrho \int_0^1 \varpi_t \left( \int_x^1 \theta_t (y) dy\right)dx\\ & = cG \int_0^1 (\phi-\varpi_x)\theta \; dx+k_0\varrho \int_0^1 \theta_x \varpi_t \; dx+\gamma c\int_0^1 \theta^2\; dx\\ &-\gamma \varrho \int_0^1 \varpi_t^2 \; dx-k_1\varrho \int_0^1 r \varpi_t \; dx+m\varrho \int_0^1 \varpi_t \int_x^1 (3\psi_t-\phi_t) (y) dy dx. \end{aligned}

    Now, thanks to Young, Poincaré's and Cauchy–Schwarz inequalities, we get, for any \epsilon_2 > 0,

    \begin{array}{c} {I}_2^\prime(t)\leq \frac{-\gamma \varrho}{2} \int_0^1 \varpi_t^2\; dx+\chi_* \int_0^1 (3\psi_t-\phi_t)^2 \; dx+\epsilon_2 \int_0^1 (\phi-\varpi_x)^2\; dx\\ +\chi_* \int_0^1 r^2 \; dx+\chi_* \left( 1+\frac{1}{\epsilon_2}\right)\int_0^1 \theta_x^2 \; dx. \end{array}

    The proof is then completed.

    Lemma 4.4. Consider the functional

    \begin{equation} {I}_3(t): = \varrho \int_0^1 \varpi _t \varpi\;dx+\varrho \int_0^1 \phi \left( \int_0^x \varpi_t(y) dy\right)\; dx , \end{equation} (4.12)

    then, it satisfies

    \begin{array}{c} {I}_3^\prime(t)\leq -\frac{G}{2} \int_0^1 (\phi-\varpi_x)^2 \; dx+\varrho\int_0^1 (3\psi_t-\phi_t)^2\;dx +\frac{3\varrho}{2} \int_0^1\varpi_t^2 \;dx \\ +\chi_* \int_0^1 \theta_x^2 \; dx + 9\varrho\int_0^1 \psi_t^2 dx. \end{array} (4.13)

    Proof. We differentiate {I}_3 , using (2.8) _1 together with integration by parts, to get

    \begin{equation*} \begin{aligned} {I}_3^\prime(t)& = \varrho \int_0^1 \varpi_t^2 \; dx+\varrho \int_0^1 \varpi_{tt} \varpi \; dx +\varrho \int_0^1 \phi_t \left( \int_0^x \varpi_t(y) dy\right)\; dx\\&+\varrho \int_0^1 \phi \left( \int_0^x \varpi_{tt}(y) dy\right)\; dx\\ & = \varrho \int_0^1 \varpi_t^2 \; dx-G \int_0^1 (\phi-\varpi_x)_x \varpi\; dx-\gamma \int_0^1 \varpi \theta_x \; dx\\ &+\varrho \int_0^1 \phi_t \left( \int_0^x \varpi_t(y) dy\right)\; dx-G\int_0^1 (\phi-\varpi_x) \phi \; dx -\gamma \int_0^1\theta \phi \; dx\\ & = \varrho \int_0^1 \varpi_t^2 \; dx-G \int_0^1 (\phi-\varpi_x)^2 \; dx-\gamma \int_0^1(\phi- \varpi_x) \theta \; dx\\ &+\varrho \int_0^1 \phi_t \left( \int_0^x \varpi_t(y) dy\right)\; dx. \end{aligned} \end{equation*}

    Notice that

    \int_0^1 \phi_t^2 \; dx \leq 2\int_0^1 (3\psi_t-\phi_t)^2\;dx+18\int_0^1 \psi_t^2 \; dx.

    By Young, Poincaré's, and Cauchy-Schwarz inequalities, we easily prove (4.13).

    Lemma 4.5. Consider the functional

    \begin{equation} {I}_4(t): = I_\varrho\int_{0}^1 (3\psi-\phi)_t (3\psi-\phi)\; dx, \end{equation} (4.14)

    then, it satisfies

    \begin{equation} \begin{aligned} {I}_4^\prime(t)&\leq -\frac{\bar{l}}{2} \int_0^1 (3\psi_x-\phi_x)^2\; dx+I_\varrho \int_0^1 (3\psi_t-\phi_t)^2\; dx+\chi_* \int_0^1 (r^2+\theta_x^2)\; dx\\ &+ \chi_* \int_0^1 (\phi-\varpi_x)^2\; dx+\chi_* (g\diamond (3\psi_x-\phi_x))(t). \end{aligned} \end{equation} (4.15)

    Proof. We proceed by differentiating the functional {I}_4 and using Eq (2.8) _2 together with integration by parts, which leads to

    \begin{equation} \begin{aligned} {I}_4^\prime(t) = & I_\varrho\int_{0}^1 (3\psi-\phi)_{tt} (3\psi-\phi)\; dx+I_\varrho\int_{0}^1 (3\psi_t-\phi_t)^2\; dx\\ = &I_\varrho\int_{0}^1 (3\psi_t-\phi_t)^2\; dx-\bar{l} \int_{0}^1 (3\psi_x-\phi_x)^2\; dx\\ +&G\int_{0}^1 (3\psi-\phi)(\phi-\varpi_x)\; dx+m \int_{0}^1 (3\psi-\phi)\theta \; dx\\+&d \int_{0}^1 (3\psi-\phi)_x r\; dx- \int_{0}^1 (3\psi-\phi)_x \int_0^\infty g(s)( (3\psi-\phi)_x (t)-(3\psi-\phi)_x (t-s))\; dsdx. \end{aligned} \end{equation} (4.16)

    By virtue of Young's inequality and (2.4), we have

    \begin{aligned} {I}_4^\prime(t) &\leq -\frac{\bar{l}}{2} \int_0^1 (3\psi_x-\phi_x)^2\; dx+I_\varrho \int_0^1 (3\psi_t-\phi_t)^2\; dx+\chi_* \int_0^1 (r^2+\theta_x^2)\; dx\\ &+ \chi_* \int_0^1 (\phi-\varpi_x)^2\; dx+C^1 \int_0^1\left[ \int_0^\infty g(s)\left((3\psi-\phi)_x(t)-(3\psi-\phi)_x(t-s)\right)ds \right]^2 dx\\ &\leq -\frac{\bar{l}}{2} \int_0^1 (3\psi_x-\phi_x)^2\; dx+I_\varrho \int_0^1 (3\psi_t-\phi_t)^2\; dx+\chi_* \int_0^1 (r^2+\theta_x^2)\; dx\\ &+ \chi_* \int_0^1 (\phi-\varpi_x)^2\; dx+\chi_* (g\diamond (3\psi_x-\phi_x))(t). \end{aligned}

    This completes the proof of (4.15).

    Lemma 4.6. Consider the functional

    \begin{equation} {I}_5(t): = 3I_\varrho\int_{0}^1 \psi_t \psi\; dx+2\beta \int_0^1 \psi^2 \;dx , \end{equation} (4.17)

    then, it satisfies the estimate

    \begin{equation} \begin{aligned} {I}_5^\prime(t)&\leq-2\delta\int_0^1 \psi^2 \; dx-3D\int_0^1 \psi_x^2\; dx+3I_\varrho\int_0^1 \psi_t^2 \; dx+\chi_* \int_0^1 (\phi-\varpi_x)^2\; dx \\ & +\chi_* \int_0^1 \int_{\varsigma_1}^{\varsigma_2}\vert \mu_2(\varsigma)\vert \mathcal{S}^2 (x, 1, p, t) \; d\varsigma dx. \end{aligned} \end{equation} (4.18)

    Proof. Simple computations using Eq (2.8) _3 and integration by parts, yield

    \begin{array}{c} {I}^\prime_5(t) = 3I_\varrho\int_{0}^1 \psi_t^2 \; dx-3D \int_{0}^1 \psi_x^2 \; dx-4\delta \int_{0}^1 \psi^2 \; dx-3G \int_{0}^1 (\phi-\varpi_x) \psi \; dx \\ \;-4 \int_0^1 \int_{\varsigma_1}^{\varsigma_2}\psi\vert \mu_2(\varsigma)\vert \mathcal{S} (x, 1, p, t) \; d\varsigma dx. \end{array} (4.19)

    Employing Young's inequality, we conclude (4.18).

    Lemma 4.7. Consider the functional

    \begin{equation} {I}_6(t): = \int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2}\varsigma e^{-\varsigma p} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx, \end{equation} (4.20)

    then, it satisfies

    \begin{equation} \begin{aligned} {I}^\prime_6(t)& \leq\beta \int_{0}^1 \psi_t^2 \; dx -m_1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, 1, \varsigma, t)\; d\varsigma dx\\ &-m_1 \int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2}\varsigma \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx, \end{aligned} \end{equation} (4.21)

    where m_1 is a positive constant.

    Proof. Taking the derivative of {I}_6 and using (2.8) _7 and \mathcal{S}(x, 0, t) = \psi_t, we have

    \begin{aligned} {I}^\prime_6(t)& = -2\int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2} e^{-\varsigma p} \vert \mu_2(\varsigma)\vert \mathcal{S}_p\mathcal{S}(x, p, \varsigma, t)\; d\varsigma dp dx\\ & = -\int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2}\varsigma e^{-\varsigma p} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx\\ &-\int_{0}^1\int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert\left\{ e^{-\varsigma } \mathcal{S}^2(x, 1, \varsigma, t)-\psi_t^2(x, t)\right\} d\varsigma dx. \end{aligned}

    From e^{-\varsigma }\leq e^{-\varsigma p}\leq 1, where 0 < p < 1, we arrive at

    \begin{aligned} {I}^\prime_6(t) &\leq-\int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2}\varsigma e^{-\varsigma } \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx\\ &+\left( \int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert\; d\varsigma \right) \int_{0}^1 \psi_t^2(x, t)\; dx\\ & -\int_{0}^1\int_{\varsigma_1}^{\varsigma_2} e^{-\varsigma } \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, 1, \varsigma, t) d\varsigma dx. \end{aligned}

    Since -e^{-\varsigma} is an increasing function, then

    -e^{-\varsigma}\leq -e^{-\varsigma_2}, \quad \text{ forall } \; \varsigma \in [\varsigma_1, \varsigma_2].

    Hence, if we denote m_1 = e^{-\varsigma_2} and use (T), we easily prove (4.21).

    Let us now prove our stability result by using the lemmas in Section 4.

    Proof of Theorem 2.2. We proceed by introducing a Lyapunov functional

    \begin{equation} {L}(t) = N {E}(t)+\sum\limits_{j = 1}^6 N_j {I}_j (t), \end{equation} (5.1)

    where constants N, N_j > 0, \; j = 1, \cdots, 6, will be chosen later.

    From (5.1), we write

    \begin{aligned} | {L}(t)-N {E}(t)|&\leq I_\varrho N_1 \int_0^1 \left| (3\psi-\phi)_t\int_0^\infty g(s)((3\psi-\phi)(t)-(3\psi-\phi)(t-s))ds \right| dx\\& +c \varrho N_2\int_0^1 \left| \varpi_t\int_x^1 \theta(y)dy\right|dx+\varrho N_3\int_0^1 \left| \varpi_t \varpi\right| dx+\varrho N_3 \int_0^1 \left| \phi \int_0^x \varpi_t (y) dy\right|dx \\ &+ I_\varrho N_4 \int_0^1 \left|(3\psi-\phi)_t(3\psi-\phi)\right| dx+3I_\varrho N_5 \int_0^1 \vert \psi_t \psi\vert dx+2\beta N_5 \int_0^1 \psi^2 dx\\&+N_6 \int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2}\varsigma e^{-\varsigma p} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx. \end{aligned}

    Thanks to Young, Cauchy-Schwarz and Poincaré's inequalities, we get

    | {L}(t)-N {E}(t)| \leq \vartheta_1 {E} (t), \quad \text{ where } \; \vartheta_1 > 0,

    i.e.,

    \begin{equation} (N-\vartheta_1) {E} (t) \leq {L}(t) \leq (N+\vartheta_1) {E} (t). \end{equation} (5.2)

    Now, differentiating the Lyapunov functional {L}(t), using (4.1), (4.7), (4.11), (4.13), (4.15), (4.18), and (4.21), and fixing

    N_4 = N_5 = 1, \; \epsilon_1 = \frac{\bar{l}}{4N_1}, \; \epsilon_2 = \frac{GN_3}{4N_2}.

    We find

    \begin{equation} \begin{aligned} \frac{d}{dt} {L}(t)&\leq -\left( \frac{\gamma\varrho}{2}N_2-\frac{3\varrho}{2}N_3\right) \int_0^1 \varpi_t^2 dx\\ &-\left( \frac{I_\varrho g_0}{2}N_1-\chi_*N_2-\varrho N_3-I_\varrho \right)\int_0^1 (3\psi_t-\phi_t)^2 dx\\ &-\left(m_0 N-9\varrho N_3 -\beta N_6 -3I_\varrho\right)\int_0^1 \psi_t^2 dx-\frac{\bar{l}}{4}\int_0^1(3\psi_x-\phi_x)^2 dx\\ &-\left(\frac{G}{4}N_3-\left(\frac{\bar{l}}{4}+2\chi_*\right) \right) \int_0^1 (\phi-\varpi_x)^2 dx\\ &-2\delta \int_0^1 \psi^2 dx-\left(k_0 N-\chi_* \left(1+\frac{N_2}{N_3}\right)N_2-\chi_* N_3-\chi_*-\frac{\bar{l}}{4} \right)\int_0^1 \theta_x^2 dx\\ &-3D\int_0^1 \psi_x^2 dx-k_2N\int_0^1 r_x^2 dx-\left( k_3N-\chi_*N_1-\chi_* N_2-\chi_*\right)\int_0^1 r^2 dx\\ &-m_1 N_6 \int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2}\varsigma \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx\\ &-( m_1 N_6 -\chi_*) \int_{0}^1\int_{\varsigma_1}^{\varsigma_2} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, 1, \varsigma, t)\; d\varsigma dx\\ &+\left( \frac{N}{2}-\chi_* N_1\right) (g^\prime \diamond (3\psi_x-\phi_x))(t)+\left( \chi_* \left(1+\frac{4N_1}{\bar{l}}\right)N_1+\chi_*\right) (g \diamond (3\psi_x-\phi_x))(t). \end{aligned} \end{equation} (5.3)

    Next, we choose our coefficients in (5.3), in a way that, they all except the last two become negative. We start by selecting N_6 big enough so that

    m_1 N_6 -\chi_* > 0,

    then, we take N_3 fairly wide, such that

    \frac{G}{4}N_3-\left(\frac{\bar{l}}{4}+2\chi_*\right) > 0,

    after that, we choose N_2 large enough, so that

    \frac{\gamma\varrho}{2}N_2-\frac{3\varrho}{2}N_3 > 0,

    now, we select N_1 sufficiently large such that

    \frac{I_\varrho g_0}{2}N_1-\chi_*N_2-\varrho N_3-I_\varrho > 0.

    We can now select N large enough so that we have (5.2) and

    \begin{cases} \frac{1}{2}N-\chi_* N_1 > 0, \\ m_0 N-9\varrho N_3-\beta N_6-3I_\varrho > 0, \\ k_3N-\chi_*N_1-\chi_* N_2-\chi_* > 0, \\ k_0 N-\chi_* \left(1+\frac{N_2}{N_3}\right)N_2-\chi_*N_3-\chi_*-\frac{\bar{l}}{4} > 0.\\ \end{cases}

    Hence, relation (5.3) becomes

    \begin{equation} \frac{d}{dt} {L}(t)\leq -\vartheta_2 \int_0^1 \left\{ \varpi_t^2+(\phi-\varpi_x)^2+(3\psi_t-\phi_t)^2 +\psi^2+(3\psi_x-\phi_x)^2+\psi_x^2\right. \\ \;\;\left.+\psi_t^2+\theta^2+r^2 \right\} dx- \vartheta_2 \int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2}\varsigma e^{-\varsigma p} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx \\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;+\vartheta_3(g \diamond (3\psi_x-\phi_x))(t), \quad \vartheta_2, \vartheta_3 > 0. \end{equation} (5.4)

    Now, exploiting (2.13) and Poincaré's inequality, we obtain

    \begin{aligned} {E} (t)&\leq \vartheta_4 \int_0^1 \left\{ \varpi_t^2+(\phi-\varpi_x)^2 +(3\psi_t-\phi_t)^2+\psi^2 +(3\psi_x-\phi_x)^2+\psi_x^2\right. \\ & \left. +\psi_t^2+\theta^2+r^2 \right\} dx+\vartheta_4 \int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2}\varsigma e^{-\varsigma p} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx\\ &+\vartheta_4 (g \diamond (3\psi_x-\phi_x))(t), \quad \text{ where } \; \vartheta_4 > 0, \end{aligned}

    from which

    \begin{array}{l} \;\;\;\;\;-\int_0^1 \left\{ \varpi_t^2+(\phi-\varpi_x)^2+(3\psi_t-\phi_t)^2+\psi^2 +(3\psi_x-\phi_x)^2 +\psi_x^2\right. \\ \left. +\psi_t^2+\theta^2+r^2 \right\} dx - \int_{0}^1\int_{0}^1\int_{\varsigma_1}^{\varsigma_2}\varsigma e^{-\varsigma p} \vert \mu_2(\varsigma)\vert \mathcal{S}^2(x, p, \varsigma, t)\; d\varsigma dp dx \\ \;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;\;- (g \diamond (3\psi_x-\phi_x))(t)\leq -\vartheta_5 {E} (t), \end{array} (5.5)

    where \vartheta_5 > 0. Thereby, if we combine (5.5) and (5.4), we have

    \begin{equation} \frac{d}{dt} {L}(t)\leq -\vartheta_6 {E}(t)+\vartheta_7 (g \diamond (3\psi_x-\phi_x))(t), \quad \text{ where } \; \vartheta_6, \vartheta_7 > 0. \end{equation} (5.6)

    Next, we multiply (5.6), by

    {G}^\prime\left( \frac{\epsilon_0 {E}(t)}{ {E}(0)} \right),

    we find

    \begin{equation} {G}^\prime\left( \frac{\epsilon_0 {E}(t)}{ {E}(0)} \right)\frac{d}{dt} {L}(t)\leq -\vartheta_6 {G}^\prime\left( \frac{\epsilon_0 {E}(t)}{ {E}(0)} \right) {E}(t)+\vartheta_7 {G}^\prime\left( \frac{\epsilon_0 {E}(t)}{ {E}(0)} \right)(g \diamond (3\psi_x-\phi_x))(t). \end{equation} (5.7)

    Now, we estimate the last term in (5.7) and use both (A _2 ) and (2.14), we find

    \begin{equation} \vartheta_7 {G}^\prime\left( \frac{\epsilon_0 {E}(t)}{ {E}(0)} \right)(g \diamond (3\psi_x-\phi_x))(t)\leq - \vartheta_7' {E}^\prime(t)+\vartheta_7' \epsilon_0 {G}_0\left( \frac{ {E}(t)}{ {E}(0)} \right), \; \vartheta_7' > 0. \end{equation} (5.8)

    We insert (5.8) in (5.7) and set \epsilon_0 = \frac{\vartheta_6 {E}(0)}{2\vartheta_7'}, we get

    \begin{equation} {G}^\prime\left( \frac{\epsilon_0 {E}(t)}{ {E}(0)} \right)\frac{d}{dt} {L}(t)+\vartheta_7' {E}^\prime(t)\leq -\Gamma {G}_0\left( \frac{ {E}(t)}{ {E}(0)} \right), \quad \Gamma > 0. \end{equation} (5.9)

    We consider now the functional

    {L}_1(t): = {G}^\prime\left( \frac{\epsilon_0 {E}(t)}{ {E}(0)} \right) {L}(t)+\vartheta_7' {E}(t).

    It is clear that

    {L}_1(t) \sim {E}(t),

    moreover, noticing that {E}^\prime(t)\leq0, \; {G}^{\prime \prime}(t) > 0, we obtain

    \begin{equation} \frac{d}{dt} {L}_1(t)\leq -\Gamma {G}_0\left( \frac{ {E}(t)}{ {E}(0)} \right). \end{equation} (5.10)

    Next, we present the functional

    {L}_2(t): = b_1\frac{ {L}_1(t)}{ {E}(0)}\sim {E}(t), \quad \text{ such that }
    \begin{cases} {L}_2(t)\leq 1, \\ \frac{d}{dt} {L}_2(t) \leq -\alpha_2 {G}_0( {L}_2(t)), \end{cases}

    where, \alpha_2 is a positive constant, therefore,

    {G}^\prime_*( {L}_2(t))\geq \alpha_2.

    We integrate over (0, t) to find

    {L}_2(t)\leq {G}^{-1}_*(\alpha_2 t+\alpha_3),

    from which, we deduce that

    {E}(t)\leq \alpha_1 {G}^{-1}_*(\alpha_2 t+\alpha_3),

    where, \alpha_1 and \alpha_3 are positive constants. The proof is then completed.

    The article is about the laminated beam system along with structural damping, past history, distributed delay, and in the presence of both temperatures and micro-temperature effects introduced in (1.1). By the semigroup approach, we established the existence and uniqueness of the solution which can be considered as the first main result. In addition, as a second novelty, a general decay result for the solution unusually with no constraints regarding the speeds of wave propagation is found. This last new result is considered, as far as we know, the first similar result in the literature for such a system, where we succeed to improve the earlier works known for the case of finite history, to the case of infinite history. The relaxation function becomes intended to satisfy a broader class of relaxation functions.

    We mention here that the distributed delay in our system makes a good interaction between the past history and the other damping terms of system (1.1). This type of damping gives more information and qualitative properties on the solution and also its impact on stability is very important as it is shown in the requirement of Theorem 2.2. Of course, the other terms (both temperatures and micro-temperature effects) act as balances in the stability of the system.

    Fares Yazid and Fatima Siham Djerad: Writing—original draft preparation; Abdelkader Moumen and Moheddine Imsatfia, Tayeb Mahrouz: Writing—review and editing; Keltoum Bouhali: Supervision. All authors have read and approved the final version of the manuscript for publication.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University for funding this work through small group research project under grant number RGP1/21/45.

    The authors declare that there is no conflict of interest.



    [1] M. S. Alves, R. N. Monteiro, Exponential stability of laminated Timoshenko beams with boundary/internal controls, J. Math. Anal. Appl., 482 (2020), 123516. https://doi.org/10.1016/j.jmaa.2019.123516 doi: 10.1016/j.jmaa.2019.123516
    [2] T. A. Apalara, On the stability of a thermoelastic laminated beam, Acta Math. Sci., 39 (2019), 1517–1524. https://doi.org/10.1007/s10473-019-0604-9 doi: 10.1007/s10473-019-0604-9
    [3] L. Bouzettouta, S. Zitouni, K. Zennir, A. Guesmia, Stability of Bresse system with internal distributed delay, J. Math. Comput. Sci., 17 (2017), 92–118.
    [4] A. Choucha, D. Ouchenane, K. Zennir, B. Feng, Global well-posedness and exponential stability results of a class of Bresse-Timoshenko-type systems with distributed delay term, Math. Meth. Appl. Sci., 2020 (2020). https://doi.org/10.1002/mma.6437.
    [5] N. Doudi, S. Boulaaras, Global existence combined with general decay of solutions for coupled Kirchhoff system with a distributed delay term, RACSAM Rev. R. Acad. A, 114 (2020), 1–31. https://doi.org/10.1007/s13398-020-00938-9 doi: 10.1007/s13398-020-00938-9
    [6] H. Dridi, B. Feng, K. Zennir, Stability of Timoshenko system coupled with thermal law of Gurtin-Pipkin affecting on shear force, Appl. Anal., 101 (2022), 1–15. https://doi.org/10.1080/00036811.2021.1883591 doi: 10.1080/00036811.2021.1883591
    [7] H. Dridi, K. Zennir, Well-posedness and energy decay for some thermoelastic systems of Timoshenko type with Kelvin-Voigt damping, SeMA J., 78 (2021), 385–400. https://doi.org/10.1007/s40324-021-00239-0 doi: 10.1007/s40324-021-00239-0
    [8] A. S. El-Karamany, M. A. Ezzat, On the phase-lag Green-Naghdi thermoelasticity theories, Appl. Math. Model., 40 (2016), 5643–5659. https://doi.org/10.1016/j.apm.2016.01.010 doi: 10.1016/j.apm.2016.01.010
    [9] D. Fayssal, Well posedness and stability result for a thermoelastic laminated beam with structural damping, Ric. Mat., 2022 (2022), 1–25. https://doi.org/10.1007/s11587-022-00708-2 doi: 10.1007/s11587-022-00708-2
    [10] E. I. Grigolyuk, Nonlinear behavior of shallow rods, Dokl. Akad. Nauk, 348 (1996), 759–763.
    [11] E. I. Grigolyuk, E. A. Lopanitsyn, Large axisymmetric flexures of thin short shells of revolution under small deformations, Dokl. Akad. Nauk, 346 (1996), 753–756.
    [12] S. W. Hansen, R. D. Spies, Structural damping in laminated beams due to interfacial slip, J. Sound Vib., 204 (1997), 183–202. https://doi.org/10.1006/jsvi.1996.0913 doi: 10.1006/jsvi.1996.0913
    [13] H. E. Khochemane, General stability result for a porous thermoelastic system with infinite history and microtemperatures effects, Math. Meth. Appl. Sci., 45 (2022), 1538–1557. https://doi.org/10.1002/mma.7872 doi: 10.1002/mma.7872
    [14] W. Liu, W. Zhao, Stabilization of a thermoelastic laminated beam with past history, Appl. Math. Optim., 80 (2019), 103–133. https://doi.org/10.1007/s00245-017-9460-y doi: 10.1007/s00245-017-9460-y
    [15] Z. Liu, S. Zheng, Semi-groups associated with dissipative systems, CRC Press, 1999.
    [16] A. Moumen, D. Ouchenane, A. Choucha, K. Zennir, S. A. Zubair, Exponential stability of Timoshenko system in thermoelasticity of second sound with a memory and distributed delay term, Open Math., 19 (2022), 1636–1647. https://doi.org/10.1515/math-2021-0117 doi: 10.1515/math-2021-0117
    [17] V. F. Nesterenko, Dynamics of heterogeneous materials, Springer, New York, 2001.
    [18] V. F. Nesterenko, Propagation of nonlinear compression pulses ingranular media, J. Appl. Mech. Tech. Phys., 24 (1984), 733–743. https://doi.org/10.1007/BF00905892 doi: 10.1007/BF00905892
    [19] S. Nicaise, C. Pignotti, Stabilization of the wave equation with boundary or internal distributed delay, Differ. Integral. Equ., 21 (2008), 935–958. https://doi.org/10.57262/die/1356038593 doi: 10.57262/die/1356038593
    [20] C. Nonato, C. Raposo, B. Feng, Exponential stability for a thermoelastic laminated beam with nonlinear weights and time-varying delay, Asymptotic Anal., 126 (2022), 157–185.
    [21] E. Pişkin, J. Ferreira, H. Yuksekkaya, M. Shahrouzi, Existence and asymptotic behavior for a logarithmic viscoelastic plate equation with distributed delay, Int. J. Nonlinear. Anal., 13 (2022), 763–788. https://doi.org/10.22075/IJNAA.2022.24639.2797 doi: 10.22075/IJNAA.2022.24639.2797
    [22] C. A. Raposo, O. V. Villagran, J. E. Muñoz Rivera, M. S. Alves, Hybrid laminated Timoshenko beam, J. Math. Phys., 58 (2017), 101512. https://doi.org/10.1063/1.4998945 doi: 10.1063/1.4998945
    [23] S. P. Timoshenko, On the correction for shear of the differential equation for transverse vibrations of prismatic bars, Lond. Edinb. Dublin. Philos. Mag. J. Sci., 41 (1921), 744–746. https://doi.org/10.1080/14786442108636264 doi: 10.1080/14786442108636264
    [24] X. Tian, O. Zhang, Stability of a Timoshenko system with local Kelvin-Voigt damping, Z. Angew. Math. Phys., 68 (2017), 20. https://doi.org/10.1007/s00033-016-0765-5 doi: 10.1007/s00033-016-0765-5
  • Reader Comments
  • © 2024 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(858) PDF downloads(35) Cited by(0)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog