Loading [MathJax]/jax/output/SVG/jax.js
Research article Special Issues

Energy decay for a porous system with a fractional operator in the memory

  • In this work, we examine a porous-elastic system with a fractional operator incorporated in the memory term, which acts exclusively on one equation within the system. Under appropriate conditions on the polynomially decreasing kernels of the memory type, we establish the result of polynomial decay.

    Citation: Chahrazed Messikh, Soraya Labidi, Ahmed Bchatnia, Foued Mtiri. Energy decay for a porous system with a fractional operator in the memory[J]. Electronic Research Archive, 2025, 33(4): 2195-2215. doi: 10.3934/era.2025096

    Related Papers:

    [1] Wenjun Liu, Zhijing Chen, Zhiyu Tu . New general decay result for a fourth-order Moore-Gibson-Thompson equation with memory. Electronic Research Archive, 2020, 28(1): 433-457. doi: 10.3934/era.2020025
    [2] Ibtissam Issa, Zayd Hajjej . Stabilization for a degenerate wave equation with drift and potential term with boundary fractional derivative control. Electronic Research Archive, 2024, 32(8): 4926-4953. doi: 10.3934/era.2024227
    [3] Antonio Magaña, Alain Miranville, Ramón Quintanilla . On the time decay in phase–lag thermoelasticity with two temperatures. Electronic Research Archive, 2019, 27(0): 7-19. doi: 10.3934/era.2019007
    [4] Jiwei Jia, Young-Ju Lee, Yue Feng, Zichan Wang, Zhongshu Zhao . Hybridized weak Galerkin finite element methods for Brinkman equations. Electronic Research Archive, 2021, 29(3): 2489-2516. doi: 10.3934/era.2020126
    [5] Zayd Hajjej . Asymptotic stability for solutions of a coupled system of quasi-linear viscoelastic Kirchhoff plate equations. Electronic Research Archive, 2023, 31(6): 3471-3494. doi: 10.3934/era.2023176
    [6] Jie Qi, Weike Wang . Global solutions to the Cauchy problem of BNSP equations in some classes of large data. Electronic Research Archive, 2024, 32(9): 5496-5541. doi: 10.3934/era.2024255
    [7] Linyan Fan, Yinghui Zhang . Space-time decay rates of a nonconservative compressible two-phase flow model with common pressure. Electronic Research Archive, 2025, 33(2): 667-696. doi: 10.3934/era.2025031
    [8] Nisachon Kumankat, Kanognudge Wuttanachamsri . Well-posedness of generalized Stokes-Brinkman equations modeling moving solid phases. Electronic Research Archive, 2023, 31(3): 1641-1661. doi: 10.3934/era.2023085
    [9] Qi Peng, Shaobo Yang, Guangcan Shan, Nan Qiao . Analysis of the diffusion range of the sleeve valve pipe permeation grouting column-hemispheric under the variation of the permeability of porous media. Electronic Research Archive, 2023, 31(11): 6928-6946. doi: 10.3934/era.2023351
    [10] Guochun Wu, Han Wang, Yinghui Zhang . Optimal time-decay rates of the compressible Navier–Stokes–Poisson system in $ \mathbb R^3 $. Electronic Research Archive, 2021, 29(6): 3889-3908. doi: 10.3934/era.2021067
  • In this work, we examine a porous-elastic system with a fractional operator incorporated in the memory term, which acts exclusively on one equation within the system. Under appropriate conditions on the polynomially decreasing kernels of the memory type, we establish the result of polynomial decay.



    In this paper, we are concerned with the porous system where the damping mechanism is presented by a fractional term of memory type:

    {ρ1φtt(μφxx+bψx)=0,in(0,L)×(0,),ρ2ψttδψxx+bφx+ξψ+0g(s)Bθψ(ts)ds=0,in(0,L)×(0,), (1.1)

    with the boundary conditions

    φ(0,t)=φ(L,t)=ψx(0,t)=ψx(L,t)=0,t(0,), (1.2)

    and the initial conditions

    {φ(x,0)=φ0(x),φt(x,0)=ψ1(x),ψ(x,0)=ψ0(x),ψt(x,0)=ψ1(x),ψ(x,s)=ϕ0(x,s),s>0,x(0,L). (1.3)

    Here, φ represents the longitudinal displacement, while ψ denotes the volume fraction of the solid elastic material. The parameters ρ1, μ, b, ρ2, δ, and ξ are positive constitutive constants that satisfy the inequality μξ>b2. The operator B corresponds to the differential operator (xx), and the parameter θ is taken within the range θ(0,1).

    The selection of boundary conditions (BCs) in Eq (1.2) and initial conditions (ICs) in Eq (1.3) plays a fundamental role in both the mathematical analysis and the physical interpretation of the problem. Below, we briefly discuss their significance:

    Importance of boundary conditions (Eq (1.2))

    ● The conditions φ(0,t)=φ(L,t)=0 correspond to a longitudinal displacement that is fixed at both ends of the domain, modeling a clamped or fixed boundary. This is a standard assumption in structural mechanics and implies that no axial motion occurs at the boundaries.

    ● The conditions ψx(0,t)=ψx(L,t)=0 imply that the flux of porosity vanishes at the endpoints, i.e., no net transfer of the volume fraction across the boundaries. This is physically reasonable in many porous materials where the microstructure remains static near the edges.

    ● From a mathematical standpoint, these boundary conditions are critical in establishing the energy framework of the problem, particularly in demonstrating the dissipation mechanism introduced by the memory term. They also play a central role in the stability and decay analysis carried out in later sections.

    Importance of initial conditions (Eq (1.3))

    ● The initial values φ(x,0), φt(x,0), ψ(x,0), and ψt(x,0) specify the initial configuration and velocity fields for the displacement and volume fraction. These are standard requirements for second-order hyperbolic systems.

    ● In addition, due to the presence of a memory-type damping term, the model requires an initial history condition ψ(x,s)=ϕ0(x,s) for s>0. This condition is essential for capturing the hereditary effects inherent in fractional damping mechanisms and ensures the proper functioning of the convolution integral.

    ● Without appropriate initial conditions—especially for the memory component—the system would be ill-posed, leading to issues of non-uniqueness or lack of existence of solutions.

    Relevance to the present work

    Physical accuracy: The choice of BCs and ICs reflects physically meaningful constraints typical in porous elastic materials subject to damping.

    Well-posedness: These conditions are essential in ensuring that the problem admits a unique, physically interpretable solution.

    Energy and stability framework: The BCs, in particular, facilitate the derivation of energy inequalities and decay rates, which are central contributions of this paper.

    Mathematical coherence: The imposed conditions are compatible with the differential structure of the system and the properties of the fractional damping operator.

    The function g serves as the kernel of the memory term and satisfies the following conditions:

    H: The function g:R+R+ is a C1-decreasing function that satisfies

    g(0)>0,g(t)m(t)gp(t),1<p<32,
    +0g(s)ds<b(πL)2(1θ),

    where m:R+R+ is a C1 non-increasing function.

    Noting that short memory can be considered in our model. This corresponds to replacing the integral over (0,+) with an integral over (0,t) or, more generally, over (t0,t). Mathematically, this can be achieved by setting the function g to zero for s(t,+) and ensuring that hypothesis H remains satisfied. This modification would still allow us to analyze the effect of fractional damping within a finite memory framework (see [1,2]).

    The fundamental evolution equations governing the one-dimensional theory of porous materials are given by

    ρ1φtt=Tx,ρ2ψtt=Hx+G. (1.4)

    Here, T represents the stress, H denotes the equilibrated stress, and G corresponds to the equilibrated body force. The variables φ and ψ describe the displacement of the solid elastic material and the volume fraction, respectively.

    The constitutive relations governing the system are expressed as

    T=μφx+bψ,H=δψx+0g(s)Bθ1/2ψ(ts)ds,G=bφxξψ. (1.5)

    Substituting these expressions into Eq (1.4), we derive the governing system given by Eq (1.1). The development of this model follows the fundamental balance laws of continuum mechanics, combined with constitutive equations that incorporate memory effects.

    The porous-elastic system studied in this paper is based on the well-established theory of poroelasticity, which describes the behavior of fluid-saturated porous materials. The governing equations are derived from the fundamental balance laws of continuum mechanics, incorporating constitutive relations that account for the interaction between the solid matrix and the fluid within the pores. The memory term in our model represents the hereditary effects observed in real-world porous materials, where the stress–strain relationship is influenced by past deformations.

    This framework is particularly relevant in applications where energy dissipation and wave propagation are significantly affected by the internal structure of the material. For instance, in geophysics, porous-elastic models are widely used to study seismic wave propagation in sedimentary rocks, where fluid flow and viscoelastic effects play a crucial role. Similarly, in biomechanics, these models help describe the behavior of biological tissues, such as bones and cartilage, which exhibit poroelastic properties due to their fluid-filled microstructure. In engineering, porous-elastic materials are used in acoustic insulation, filtration systems, and energy absorption applications.

    By incorporating a fractional damping term in the memory effect, our model extends classical poroelasticity theories to account for more complex dissipation mechanisms. This provides a more accurate description of materials that exhibit long-term relaxation behavior, making the results applicable to a broader range of porous media, including viscoelastic foams, polymer-based composites, and other engineered materials with internal fluid interactions.

    Porous materials represent a highly significant area of materials science due to their broad range of applications. They are widely used in various fields, including soil mechanics, engineering, power technology, biology, and material science, among others. The theoretical framework for porous elastic materials was introduced by Cowin and Nunziato [3], who developed a nonlinear theory of elastic materials with voids.

    Theory is based on the observation that, in addition to the usual elastic effects, these materials possess a microstructure with the property that the mass at each point is obtained as the product of the mass density of the material matrix and the volume fraction. For more details, we refer the reader to [4,5,6,7] and the references therein.

    Recent works, such as [8,9,10,11,12,13], have focused on the asymptotic behavior of solutions under different damping conditions, revealing intricate stability properties that depend on the interaction between porosity and viscoelasticity. In [14], Quintanilla considered (1.1) with a linear damping term τut (τ is constant) in the second equation (without the memory term g=0) and initial and mixed boundary conditions. He obtained a decay result, but it is non-exponential decay.

    In [15], the authors introduced a viscoelastic damping term of the form τutxx in the first equation of (1.1), assuming g=0. They established that the decay rate of the solution is polynomial and cannot be exponential. In [8], Apalara investigated system (1.1) with Neumann–Dirichlet boundary conditions, incorporating a finite memory term t0g(s)ψxx(ts)ds instead of an infinite memory term +0g(s)Bθψ(ts)ds. Under the assumption of equal wave propagation speeds and an exponentially decaying relaxation function, a general decay result was obtained, encompassing both exponential and polynomial decay as special cases. Recently, this result in [13] was extended to the case of non-equal wave speeds, which is more realistic from a physical perspective.

    When μ=b=ξ=K, it is well known that system (1.1) reduces to the Timoshenko system with a fractional operator in the infinite memory:

    {ρ1φttK(φxx+ψx)=0,in(0,L)×(0,),ρ2ψttδψxx+K(φx+ψ)+0g(s)Bθψ(ts)ds=0,in(0,L)×(0,). (1.6)

    Astudillo and Oquendo [16] investigated system (1.6) under the assumption of an exponentially decreasing kernel and Dirichlet–Neumann boundary conditions. Using semigroup theory and the spectral approach, they established polynomial decay rates. Specifically, they demonstrated that if the wave propagation speeds are different and θ1, the solutions decay polynomially with a rate of t1/(42θ), while if the wave propagation speeds are equal, the solutions decay polynomially with a rate of t1/(22θ). In addition, they proved that these decay rates are optimal. Moreover, when θ=1 and the wave propagation speeds are equal, they obtained exponential decay of the solutions. In their study, they also established the global existence of weak solutions.

    From a numerical perspective, advanced finite element and spectral methods have been employed to approximate solutions to complex poroelastic systems, enabling more precise simulations of wave propagation and energy dissipation in porous structures [17,18]. These studies provide a solid foundation for further exploration of porous-elastic models with memory effects, as considered in this work.

    In the present work, we address the following question: By applying the multiplier method to the systems (1.1)–(1.3) under assumption (H), do we obtain the same polynomial decay result as in [16]? It is important to emphasize that the method used in our study differs from the approach used in [16].

    The remainder of this paper is organized as follows: In Section 2, we present some preliminary results that are essential for proving our main result. In Section 3, we establish the well-posedness of the systems (1.1)–(1.3). Finally, in Section 4, we state and prove the main result concerning the energy decay of the system using the multiplier technique.

    In this section, we introduce some preliminary material needed for the proof of our results. Throughout this paper, C denotes a generic positive constant.

    First, we define the functional spaces used in our analysis. The space L2=L2(0,L) denotes the usual Lebesgue space, equipped with the norm L2. For simplicity, we will use instead of L2 and , instead of ,L2.

    Let s be a non-negative number. The Sobolev space Hs=Hs(0,L) consists of functions in L2(0,L) whose weak derivatives up to order s also belong to L2(0,L), and it is endowed with the norm Hs.

    Next, we introduce the following Hilbert spaces:

    L2(0,L)={fL2(0,L):L0f(x)dx=0},
    H1(0,L)=H1(0,L)L2(0,L).

    Next, we define the operators:

    B=xx:D(B)L2(0,L)L2(0,L),
    B=xx:D(B)L2(0,L)L2(0,L),

    where

    D(B)=H2(0,L)H10(0,L),

    and

    D(B)={ψH2(0,L)L2(0,L):ψx(0)=ψx(L)}.

    The operators B and B are positive, self-adjoint, and have a compact inverse. Consequently, the operators Bσ and Bσ are positive, bounded for σ0, and self-adjoint for all σR. Furthermore, the embeddings

    D(Bσ1)D(Bσ2),D(Bσ1)D(Bσ2),

    are continuous for σ1>σ2.

    We define the norms:

    Bσφ=φD(Bσ),Bσφ=φD(Bσ),

    for σ0.

    If φD(Bσ+1/2) and ψD(Bσ+1/2), we have

    Bσ+1/2φ=Bσxφ,Bσ+1/2ψ=Bσxψ.

    In the case when σ=0, it follows that

    B1/2φ=xφ,B1/2ψ=xψ.

    If φD(Bσ0) and ψD(Bσ0) with σ0=max(σ,1/2), then we have

    Bσψ,φx=φx,Bσφ.

    If ψD(B) and +0g(s)ds<b(πL)2(1θ), then Bσψ and Eσψ are equivalent for all σR, where

    Eψ=δBψ(+0g(s)ds)Bθψ.

    For more details on this context, we refer to [16,19].

    Now, let η(t,s)=ψ(t)ψ(ts); then the system (1.1) becomes

    {ρ1φtt(μφxx+bψx)=0,in(0,L)×(0,),ρ2ψtt+Eψ+bφx+ξψ++0g(s)Bθη(s)ds=0,in(0,L)×(0,),ηtψt+sη=0. (2.1)

    We end with the following crucial lemma, which will be used in the proof of our main result.

    Lemma 2.1. Let α, c1, and c2 be three positive constants; F, m, and h be positive functions such that F is differentiable and m and h are continuous on R+, satisfying

    t>0,F(t)c1mα+1(t)Fα+1(t)+c2h(t).

    Then, for some constant C>0, we have

    F(t)C(1+t)1αmα+1α[1+t0(s+1)1αmα+1αh(s)ds]t>0. (2.2)

    Proof. In order to prove the relation (2.2), we follow the same steps as in [20] (page 598).

    In this section, we study the existence of solutions for the porous system. For this purpose, we consider the following Hilbert space:

    H=H10(0,L)×H1(0,L)×L2(0,L)×L2(0,L)×L2g(R+;D(Bθ/2)).

    The energy associated with the solution of the problem is given by

    E(t)=ρ12φt2+ρ22ψt2+μ2φx2+12E1/2ψ2+ξ2ψ2+bL0ψφxdx+12η2L2g(R+;D(Bθ/2)), (3.1)

    for all (φ,ψ,φt,ψt,η)H.

    Lemma 3.1. Let (φ,ψ,η) be a regular solution of the problem (2.1). Then, the energy functional defined by (3.1) satisfies

    E(t)=12L0+0g(s)[Bθ/2η]2dsdx0.

    Proof. Multiplying (2.1)1 by φt and (2.1)2 by ψt and integrating over (0,L), we obtain

    ddt[ρ1φt2+μφx2]+bL0ψφxtdx=0, (3.2)

    and

    12ddt[ρ2ψt2+E1/2ψ2+μφx2]+bL0φxψtdx+L0+0g(s)Bθ/2η(s)Bθ/2ψtdsdx=0. (3.3)

    We remark that

    L0+0g(s)dds(Bθ/2η)2dsdx=L0+0g(s)(Bθ/2η)2dsdx. (3.4)

    Using the relation (3.4) and (2.1)3, we obtain

    L0+0Bθ/2ψtg(s)Bθ/2η(s)dsdx=12ddt[L0+0g(s)(Bθ/2η)2(s)dsdx]12L0+0g(s)(Bθ/2η(s))2dsdx. (3.5)

    Inserting (3.5) into (3.3), we obtain

    12ddt[ρ2ψt2+E1/2ψ2+μφx2+L0+0g(s)(Bθ/2η)2(s)dsdx]+bL0φxψtdx12L0+0g(s)(Bθ/2η(s))2dsdx. (3.6)

    Summing (3.2) and (3.6), we arrive at

    E(t)=12L0+0g(s)(Bθ/2η)2dsdx0.

    Hence, the proof of this lemma is achieved.

    For completeness, we state and prove the existence result of (2.1), (1.2), (1.3) by the Galerkin method together with some energy estimates.

    Theorem 3.2. For (φ0,ψ0,φ1,ψ1,η)H and T>0. Assume that (H) is satisfied; then the problems (1.1)–(1.3) has a unique weak solution such that

    (φ,ψ)C([0,T],H10(0,L)×H1(0,L))C1([0,T],L2(0,L)×L2(0,L)).

    Proof. The proof can be done in two steps:

    Step 1: Faedo-Galerkin approximation: We construct an approximation of the solution (φ,ψ,η) using the Faedo–Galerkin method. Specifically, let Wn=span(w1,,wn) be a Hilbert basis of the space H10(0,1).

    We choose a sequence

    (φn0,ψn0,ηn0)Wn,(φn1,ψn1,ηn1)Wn,

    such that

    (φn0,φn1,ψn0,ψn1,ηn0,ηn1)(φ0,φ1,ψ0,ψ1,η0,η1)strongly in H.

    We now define the approximation:

    (φn(x,t),ψn(x,t),ηn(x,t,s))=nj=1(fnj(x,t),gnj(x,t),gnj(x,t)gnj(x,ts))wj,

    which satisfies the following problem:

    {ρ1L0φnttwjdx+L0(μφnx+bψn)wjxdx=0,ρ2L0ψnttwjdx+L0E1/2ψnE1/2wjdx+L0(bφnx+ξψn)wjdx+L0(+0g(s)Bθ/2ηn(s)ds)Bθ/2wjdx=0, (3.7)

    with the initial conditions:

    (φn(0),ψn(0))=(φn0,ψn0)and(φnt,ψnt)=(φn1,ψn1). (3.8)

    According to the standard theory of ordinary differential equations, the finite-dimensional problems (3.7) and (3.8) has a solution (fnj,gnj)j=1,...,n defined on [0,tn). The following estimate allows us to conclude that tn=T.

    Step 2: Energy estimate: Multiplying (3.7)1 by fnj, (3.7)2 by gnj, summing over j=1,,n for each obtained equation, and finally integrating over (0,t), we obtain:

    {ρ1t0L0φnttφntdxdtt0L0(μφnxx+bψx)φntdxdt=0,ρ2t0L0ψnttψntdx+t0L0Eψnψntdx+t0L0(bφnx+ξψn)ψntdx+t0L0(+0g(s)Bθηn(s)ds)ψntdx=0.

    By integrating by parts, as Bθ and Eθ are positive self-adjoint operators, we obtain using (3.5)3:

    {ρ12t0L0(φnt)2dxdt+t0L0(μφnx+bψn)φnxtdxdt=ρ12L0(φnt)2(0)dx,ρ22L0(ψnt)2dx+12L0(E1/2ψn)2dx+ξ2L0(ψn)2dx+12L0(+0g(s)Bθ/2(ηn(s))2ds)dx+bt0L0φnxψntdxdx=ρ22L0(ψnt)2(0)dx+12L0E1/2ψn2(0)dx+ξ2L0(ψn)2(0)dx+12L0(+0g(s)(Bθ/2ηn(0,s))2ds)dx+12t0L0(+0g(s)Bθ/2(ηn(s))2ds)dx. (3.9)

    Now, we denote:

    En(t)=ρ12φnt2+ρ22ψnt2+μ2φnx2+12E1/2ψ2+ξ2ψ2+bL0ψnφnxdx+12ηn2.

    Summing up, we obtain:

    En(t)En(0).

    Since the sequence converges, we can find a positive constant C independent of n such that:

    En(t)C.

    From there we can pass to the limit in (3.7) and (3.8). The rest of the proof follows.

    Remark 3.3. If the condition μξ>b2 is satisfied, then the energy E(t) defined in (3.1) is equivalent to the norm

    ρ12φt2+ρ22ψt2+μ2φx2+12E1/2ψ2+ξ2ψ2+12η2L2g(R+;D(Bθ/2)).

    This equivalence ensures that the energy E(t) properly measures the total dynamics of the system and provides a useful tool for stability analysis.

    In this section, we prove a decay result for the energy of the systems (1.1)–(1.3) using the multiplier technique. To this end, we first establish the following lemmas.

    Lemma 4.1. Assume that g satisfies (H). Then, for all tR+, we have:

    (i) For 1<p<32, there exists a constant C>0 such that

    m(t)L0t0g(s)[Bθ/2η]2dsdxC[E(t)]12p1.

    (ii) If there exists a positive constant n0 such that B1/2ϕ0(t)2n0, then for σ(0,1], we have:

    m(t)L0+tg(s)[Bσ/2η]2dsC(+tg(s)ds)m(t)=Ch(t). (4.1)

    Proof. For the proof of (ⅰ), we refer to Corollary 2.1 in [21].

    Now, we prove (ⅱ). We estimate:

    L0+tg(s)[Bσ/2η]2ds2Bσ/2ψ(t)2+tg(s)ds+2+tg(s)Bσ/2ψ(ts)2ds2E(0)+tg(s)ds+2supz0Bσ/2ψ(z)2+tg(s)ds(2E(0)+n0)+tg(s)ds=C+tg(s)ds.

    Thus, relation (4.1) is verified.

    Corollary 4.2. Assume that g satisfies (H). Then, for all tR+, we have

    m(t)(L0+0g(s)[Bθ/2η]2ds+L0+tg(s)[B1/2η]2ds)C([E(t)]12p1+h(t)).

    Proof. This result follows directly from (ⅰ) and (ⅱ) in Lemma 4.1.

    Lemma 4.3. The functional

    F1(t)=ρ2L0ψψtdx+bρ1μL0ψx0φt(y)dydx

    satisfies

    F1(t)12E1/2ψ2(ξb2μ)ψ2+ε1φt2+C(1+1ε1)ψt2+CgBθ/2η, (4.2)

    where

    gBθ/2η=L0+0g(s)[Bθ/2η]2dsdx.

    Proof. We compute the derivative of F1 with respect to t. Using (2.1)1, (2.1)2, and integration by parts, we obtain

    F1(t)=E1/2ψ2(ξb2μ)ψ2+ρ2ψt2+0g(s)L0Bθ/2ηBθ/2ψ(t)dxds+bρ1μL0ψtx0φt(y)dydx.

    Now, we estimate the last two terms on the right-hand side as follows:

    Using Young's inequality, we estimate

    bρ1μL0ψtx0φt(y)dydxε1L2L0(x0φt(y)dy)2dx+(bρ1μ)2L24ε1L0ψ2tdxε1L0φ2tdx+Cε1L0ψ2tdx. (4.3)

    For the integral involving g(s), using Young's and Cauchy–Schwarz's inequalities along with the fact that

    E1/2ψB1/2ψ,andB1/2ψBθ/2ψ,

    we obtain

    +0g(s)L0Bθ/2ηBθ/2ψ(t)dxdsδ1L0+0[Bθ/2ψ]2dx+14δ1L0[+0g(s)Bθ/2ηds]2dxδ1L0+0[Bθ/2ψ]2dx++0g(s)ds4δ1L0+0g(s)[Bθ/2η]2dsdxc1δ1E1/2ψ2+Cδ1gBθ/2η.

    Choosing δ1=12c1, we obtain

    +0g(s)L0Bθ/2ηBθ/2ψ(t)dxds12E1/2ψ2+CgBθ/2η. (4.4)

    By combining (4.3) and (4.4), we verify inequality (4.2), completing the proof.

    Lemma 4.4. Let T be a positive constant, and let the functional

    F2(t)=ρ2L0ψt+tg(s)η(s)dsdx

    satisfy, for any ε2>0 and ε3>0, the inequality for all tT

    F2(t)ρ2g02ψt2+2ε2E1/2ψ2+ε3φx2CgBθ/2η+Cε2˜gB1/2η+C(1ε2+1ε3+1)gBθ/2η,

    where

    g0=Tg(s)dsfor alltT,
    gBθ/2η=L0+0g(s)[Bθ/2η]2dsdx,

    and

    ˜gB1/2η=L0+tg(s)[B1/2η]2dsdx.

    Proof. Differentiating F2 with respect to t, we obtain

    F2(t)=ρ2L0ψtt+tg(s)η(s)dsdx+ρ2L0ψtg(t)η(t)dxρ2L0ψt+tg(s)ηt(s)dsdx.

    Using (2.1)2 and (2.1)3, and also the self-adjointness of E1/2 and Bθ/2, we obtain

    F2(t)=L0E1/2ψ(+tg(s)E1/2η(s)ds)dx+L0(bφx+ξψ)(+tg(s)η(s)ds)dx+L0(+0g(s)Bθ/2η(s)ds)(+tg(s)Bθ/2η(s))dxρ2L0ψt(+tg(s)η(s)ds)dxρ2(+tg(s)ds)ψt2.

    Now, using Young's and Cauchy–Schwarz inequalities, and noting that D(Bθ/2)L2(0,L) and D(E1/2)D(B1/2), we obtain the following estimates:

    J1=L0E1/2ψ(+tg(s)E1/2η(s)ds)dxε2E1/2ψ2+14ε2L0[+tg(s)E1/2ηds]2dxε2E1/2ψ2++tg(s)ds4ε2L0+tg(s)[E1/2η]2dsdxε2E1/2ψ2+Cε2˜gB1/2η. (4.5)

    Similarly to (4.5), we find

    J2=bL0φx(+tg(s)η(s)ds)dxε3φx2+Cε3gBθ/2η, (4.6)
    J3=ξL0ψ(+tg(s)η(s)ds)dxε2E1/2ψ2+Cε2gBθ/2η, (4.7)
    J4=L0(+tg(s)Bθ/2η(s)ds)(+0g(s)Bθ/2η(s)ds)dxL0(+0g(s)Bθ/2η(s)ds)2dxCgBθ/2η, (4.8)

    and

    J5=ρ2L0ψt(+tg(s)η(s)ds)dxρ2[δ1L0ψ2tdx+14δ1L0(+tg(s)η(s)ds)2dx]ρ2δ1L0ψ2tdxρ24δ1(+tg(s)ds)L0+0g(s)(η(s))2dsdxρ2δ1ψt2Cδ1gBθ/2η.

    Putting δ1=g02, we obtain

    J5ρ2g02ψt2CgBθ/2η. (4.9)

    From (4.5)–(4.9) and since g is a decreasing function, it follows that the relation (4.4) is verified. Hence, the proof is completed.

    Lemma 4.5. Assume (H) holds and that μρ1=δρ2. Then, the functional

    F3(t)=L0φt[ψx+ρ1μρ2+0g(s)Bθ1/2[ψ(t)η(s)]ds]dx+L0φxψtdx

    satisfies, for any ε4>0,

    F3(t)C(1+ε4)E1/2ψ2b2ρ2φx2+ε4φt2+Cε4gBθ/2ηCε4gBθ/2η. (4.10)

    Proof. We take the derivative of F3(t), which gives

    F3(t)=L0φtt[ψx+ρ1μρ2+0g(s)Bθ1/2[ψ(t)η(s)]ds]dx+L0φt[ψxt+ρ1μρ2+0g(s)Bθ1/2[ψt(t)ηt(s)]ds]dx+L0φxtψtdx+L0φxψttdx.

    Using Eqs (2.1)1 to (2.1)3, we obtain

    F3(t)=μρ1L0φxxψxdx+1ρ2(+0g(s)ds)L0Bθ1/2ψφxxdx1ρ2L0φxx+0g(s)Bθ1/2η(s)dsdx+bρ1ψx2+bμρ2(+0g(s)ds)L0Bθ1/2ψψxdxbμρ2L0ψx+0g(s)Bθ1/2η(s)dsdx1ρ2L0φxEψdxbρ2φx2ξρ2L0φxψdx1ρ2L0φx+0g(s)Bθη(s)dsdxρ1μρ2L0φt+0g(s)Bθ1/2η(s)dsdx. (4.11)

    We recall that ψx=B1/2ψ and Bθ1/2ψx=Bθψ. Then, applying integration by parts to Eq (4.11) and taking into account the boundary conditions, as well as the self-adjointness of E1/2 and Bσ/2 for all σR, we obtain

    F3(t)=bρ1ψx2bρ2φx2+bμρ2(+0g(s)ds)L0Bθ1/2ψψxdxbμρ2L0ψx+0g(s)Bθ1/2η(s)dsdxρ1μρ2L0φt+0g(s)Bθ1/2η(s)dsdxξρ2L0φxψdx. (4.12)

    Next, we estimate the terms on the right-hand side as follows:

    - Using Young's and Cauchy–Schwarz's inequalities and the fact that D(Bθ/2)D(Bθ1/2), we obtain

    I1=bμρ2(+0g(s)ds)L0Bθ1/2ψψxdxCE1/2ψ2. (4.13)

    - For the second term,

    I2=bμρ2L0ψx+0g(s)Bθ1/2η(s)dsdxε4ψx2+(bμρ2)24ε4L0[+0g(s)Bθ1/2η(s)ds]2dxε4ψx2+(bμρ2)2+0g(s)dsε4gBθ1/2ηCε4E1/2ψ2+Cε4gBθ/2η. (4.14)

    - For the third term,

    I3=ρ1μρ2L0φt+0g(s)Bθ1/2η(s)dsdxε4φt2Cε4gBθ/2η. (4.15)

    - For the fourth term,

    I4=ξρ2L0φxψdxδ1φx2+(ξρ2)214δ1ψ2δ1φx2+C1(ξρ2)214δ1E1/2ψ2. (4.16)

    Putting δ1=b2ρ2, we obtain

    I4b2ρ2φx2+CE1/2ψ2. (4.17)

    Inserting inequalities (4.13) through (4.17) into Eq (4.12), we obtain the desired result, and thus inequality (4.10) holds.

    Lemma 4.6. The functional

    F4(t)=ρ1L0φtφdx

    satisfies the following inequality:

    F43μ2φx2ρ1φt2+CE1/2ψ2. (4.18)

    Proof. We take the derivative of F4, using (2.1)1, integrating by parts, and applying the boundary conditions. This leads to the expression

    F4(t)=ρ1μL0(φx)2dxρ1L0φ2tdx+bL0ψφxdx.

    Next, we apply Young's inequality to the term L0ψφxdx, and since B1/2ψE1/2ψ we arrive at the inequality

    F43μ2φx2ρ1φt2+CE1/2ψ2,

    which completes the proof.

    We are now ready to state and prove the main result, which concerns the decay of energy in our system and its important physical implications, particularly in the context of porous-elastic materials with memory effects. The derived decay rates describe how the system dissipates energy over time, reflecting the internal damping mechanisms induced by the memory term. Physically, this corresponds to the stabilization of mechanical vibrations and the gradual attenuation of wave propagation within the material. Such behavior is crucial in applications involving viscoelastic or porous structures, where controlling long-term stability is essential for maintaining structural integrity.

    Theorem 4.7. Assume (H) holds. If

    μρ1=δρ2

    and

    n0>0such thatB1/2ϕ0n0,

    where ϕ0 is defined in (1.3), then for any T>0, there exists a positive constant C such that for all tT, the energy functional E(t) given in (3.1) satisfies the following inequality:

    E(t)C(1+t)12p2m2p12p2[1+t0(s+1)12p2m2p12p2h2p1(s)ds], (4.19)

    where h(t)=m(t)tg(s)ds.

    Proof. We define a Lyapunov functional

    L(t)=NE(t)+N1F1(t)+N2F2(t)+N3F3(t)+F4(t), (4.20)

    where N,N1,N2, and N3 are positive constants to be chosen later. By differentiating (4.20) and using Lemmas 4.3–4.6, we find

    L(t)[N122ε2N2CCN3(1+ε4)]E1/2ψ2(ξb2μ)N1ψ2[ρ1ε4N3ε1N1]φt2[b2ρ2N332με3N2]φx2[ρ22g0N2C(1+1ε1)N1]ψt2+C[N1+N2(1ε3+1ε2+1)+N3ε4]gBθ/2η+[N2CN3ε4CN2]gBθ/2η+CN2ε2˜gB1/2η.

    By setting ε1=ρ14N1,ε2=N18N2,ε3=b4ρ2N3N2,ε4=ρ14N3, we obtain

    L(t)[N14CCN3]E1/2ψ2[ξb2μ]N1ψ2ρ12φt2[b4ρ2N332μ]φx2[ρ22g0N2CN1(1+N1)]ψt2+C[N1+N2(N2N1+N2N3+1)+N23]gBθ/2η[N2CN2CN23]gBθ/2η+CN22N1˜gB1/2η.

    First, we choose N3 large enough such that α1=bN34ρ232μ>0, then we choose N1 large enough such that α2=N14CN3(1+N3)C>0, and finally, we choose N2 large enough such that α3=ρ2N2g02CN1(1+N1)>0, so we have

    L(t)α2E1/2ψ2α0ψ2ρ12φt2α1φx2α3ψt2+CgBθ/2η+[N2C]gBθ/2η+C˜gB1/2η, (4.21)

    where α0=(ξb2μ)N1 and g0=Tg(s)ds.

    On the other hand, we have

    |L(t)NE(t)|N1|F1(t)|+N2|F2(t)|+N2|F3(t)|+|F4(t)|.

    Exploiting Young's, Cauchy–Schwarz, and Poincaré inequalities, and recalling that

    E1/2ψB1/2ψandB1/2ψψx,

    we can estimate each of |F1(t)|,|F2(t)|,|F3(t)|,|F4(t)| one by one.

    Specifically, we have

    |F1(t)|ρ2L0|ψψt|dx+bρ1μL0|ψ|(x0|φt(y)|dy)dxρ22(ψ2+ψt2)+bρ1μL0L0|ψ(x)||φt(y)|dydxρ22(ψ2+ψt2)+bρ1L2μ(φt2+ψ2)C(ψ2+ψt2+φt2).

    Using D(Bδ1)D(Bδ2) for δ1>δ2 (we take δ2=0,δ1=θ2), we have the following estimate:

    |F2(t)|ρ2L0|ψt||+tg(s)η(s)ds|dxρ2ψt+tg(s)η(s)dsρ22(ψt2++0g(s)η(s)ds2)ρ22(ψt2+(+0g(s)ds)gBθ/2η)C(ψt2+gBθ/2η).

    Using again D(B1/2)(Bθ/2)D(Bθ1/2) for all θ(0,1), we obtain the following estimate:

    |F3(t)|L0|φtψx|dx+ρ1μρ20g(s)dsL0|φtBθ1/2ψ|dx+ρ1μρ20g(s)(L0|φt||Bθ1/2η(s)|dx)ds+L0|φxψt|dx12(φt2+ψx2+ψt2+φx2)+ρ1μρ20g(s)ds(φt2+Bθ1/2ψ2)+ρ12μρ20g(s)(Bθ1/2η(s)2+φt2)dsC(φt2+ψt2+φx2+B1/2ψ2+Bθ1/2ψ2+gBθ1/2η)C(φt2+ψt2+φx2+E1/2ψ2+gBθ/2η).

    Using Poincaré's inequality, we have:

    |F4(t)|ρ1L0|φtφ|dxC(φt2+φx2).

    From here, we can deduce that:

    L(t)NE(t)C(φt2+φx2+ψt2+ψ2+gBθ/2η+E1/2ψ2)CE(t).

    Thus, we obtain the estimate:

    (NC)E(t)L(t)(N+C)E(t).

    Next, we return to the estimation (4.21) and choose N large enough such that N2C>0 and NC>0. Therefore, this means that L(t)E(t). Since gBθ/2η<0, Eq (4.21) yields:

    L(t)K1E(t)+K2gBθ/2η+K3˜gB1/2η.

    Multiplying the above inequality by m(t) and using Corollary (4.2), we obtain:

    m(t)L(t)K1m(t)E(t)+C[E(t)]1/(2p1)+Ch(t).

    Next, multiplying the above inequality by (mE)α(t), where α=2p2, and applying Young's inequality, we get for any ϵ>0:

    mα+1(t)L(t)Eα(t)(K12ϵ)mα+1(t)Eα+1(t)CϵE(t)+Cϵhα+1(t).

    Now, we choose ϵ small enough such that

    K2=K12ϵ>0,

    and set

    F(t)=mα+1(t)L(t)Eα(t)+CϵE(t).

    Since both E and m are decreasing functions, we have:

    F(t)mα+1(t)L(t)Eα(t)+CϵE(t)K2mα+1(t)Eα+1(t)+Cϵhα+1(t).

    On the other hand, it is easy to remark that F(t)E(t), from where we deduce:

    F(t)K2mα+1Fα+1(t)+Cϵhα+1(t).

    Thanks to Lemma 2.1, this implies:

    F(t)C(1+t)12p2m2p12p2[1+t0(s+1)12p2m2p12p2h2p1(s)ds],

    and hence, we conclude that the estimate (4.19) is satisfied. Therefore, the proof of this theorem is complete.

    To illustrate the energy decay rates obtained by Theorem 4.7, we give the following example.

    Example 1. Let g(t)=β1(1+t+1)β2, where β1 and β2 are two positive constants, which we will choose later. Indeed,

    g(t)=m(t)gp(t),

    where m(t)=β22β1/β21t+1 and p=β2+1β2.

    We first choose β2>2 so that p[1,32) and +0g(s)ds are bounded. Then, we choose β1 such that

    +0g(s)ds<(πL)2(1θ).

    Therefore, all conditions of Theorem 4.7 are fulfilled, and we can apply the estimate of the energy decay (4.19). Indeed, we start by

    (1+t)12p2m2p12p2(1+t)32p2(2p2).

    Note that 32p2(2p2)<0 because p(1,32).

    We have

    h2p1=(Cm(t)tg(s)ds)2p1C(1+t)(1β2)(2p12).

    Thus, we deduce that

    I=t0(s+1)12p2m2p12p2h2p1(s)dsCt0(s+1)1+2p12(2p12p2β2)ds. (4.22)

    Putting β3=1+2p12(2p12p2β2), the relation in (4.22) yields

    {IC((t+1)β3+11β3+1)ifβ31,ICln(1+t)ifβ3=1.

    From where it follows that

    {IC(t+1)β3+1if2<β2<2p12p2,ICln(1+t)ifβ2=2p12p2,ICifβ2>2p12p2.

    We see that 2p12p2>2 because p(1,32). So, finally, we arrive at

    {E(t)C(t+1)1+(1β2)(2p12)if2<β2<2p12p2,E(t)C(t+1)32p2(2p2)ln(1+t)ifβ2=2p12p2,E(t)C(t+1)32p2(2p2)ifβ2>2p12p2.

    It is easy to remark that in the case when 2<β2<2p12p2, the functional E(t) exhibits polynomial decay if we choose β2(1+2p2p1,2p12p2).

    In this paper, we have studied the stability properties of a porous-elastic system with fractional damping in the memory term. By carefully analyzing the interplay between the wave propagation speeds and the memory effect, we established a polynomial decay rate for the energy, highlighting the crucial role of the memory kernel in dictating the system's asymptotic behavior. Our findings provide a deeper understanding of the dissipative mechanisms governing porous-elastic materials and extend previous results by refining the decay estimates under minimal assumptions on the kernel.

    Future research directions include investigating the impact of nonhomogeneous boundary conditions, exploring the extension of the model to fractional-order time evolution equations, and conducting numerical simulations to further validate the theoretical results. Additionally, it would be of interest to analyze the system under more general geometric configurations or in the presence of external forcing terms, which could offer new insights into real-world applications.

    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 Khaled University for funding this work through a large research project under grant number RGP2/368/45.

    The authors declare no competing interests.

    The data that support the findings of this study are available from the corresponding author upon reasonable request.



    [1] M. Malendowski, W. Sumelka, T. Gajewski, R. Studziński, P. Peksa, P. W. Sielicki, Prediction of high-speed debris motion in the framework of time-fractional model: Theory and validation, Arch. Civ. Mech. Eng., 23 (2023). https://doi.org/10.1007/s43452-022-00568-5 doi: 10.1007/s43452-022-00568-5
    [2] Y. Zhou, Y. Zhang, Noether symmetries for fractional generalized Birkhoffian systems in terms of classical and combined Caputo derivatives, Acta Mech., 231 (2020), 3017–3029. https://doi.org/10.1007/s00707-020-02690-y doi: 10.1007/s00707-020-02690-y
    [3] J. W. Nunziato, S. C. Cowin, A nonlinear theory of elastic materials with voids, Arch. Ration. Mech. Anal., 72 (1979), 175–201. https://doi.org/10.1007/BF00249363 doi: 10.1007/BF00249363
    [4] S. C. Cowin, The viscoelastic behavior of linear elastic materials with voids, J. Elast., 15 (1985), 185–191. https://doi.org/10.1007/BF00041992 doi: 10.1007/BF00041992
    [5] S. C. Cowin, J. W. Nunziato, Linear elastic materials with voids, J. Elast., 13 (1983), 125–147. https://doi.org/10.1007/BF00041230 doi: 10.1007/BF00041230
    [6] D. Iesan, R. Quintanilla, Decay estimates and energy bounds for porous elastic cylinders, Z. Angew. Math. Phys., 46 (1995), 268–281. https://doi.org/10.1007/BF00944757 doi: 10.1007/BF00944757
    [7] R. Quintanilla, Uniqueness in nonlinear theory of porous elastic materials, Arch. Mech., 49 (1997), 67–75.
    [8] T. A. Apalara, General decay of solutions in one-dimensional porous-elastic system with memory, J. Math. Anal. Appl., 469 (2019), 457–471. https://doi.org/10.1016/j.jmaa.2017.08.007 doi: 10.1016/j.jmaa.2017.08.007
    [9] T. A. Apalara, On the stability of porous-elastic system with microtemperatures, J. Therm. Stresses, 42 (2019), 265–278. https://doi.org/10.1080/01495739.2018.1486688 doi: 10.1080/01495739.2018.1486688
    [10] J. R. Fernández, R. Quintanilla, Two-temperatures thermo-porous-elasticity with microtemperatures, Appl. Math. Lett., 111 (2021), 106628. https://doi.org/10.1016/j.aml.2020.106628 doi: 10.1016/j.aml.2020.106628
    [11] W. Liu, D. Chen, S. A. Messaoudi, General decay rates for one-dimensional porous-elastic system with memory: The case of non-equal wave speeds, J. Math. Anal. Appl., 482 (2020), 123552. https://doi.org/10.1016/j.jmaa.2019.123552 doi: 10.1016/j.jmaa.2019.123552
    [12] M. L. Santos, A. D. S. Campelo, M. L. S. Oliveira, On porous-elastic systems with Fourier law, Appl. Anal., 98 (2018), 1181–1197. https://doi.org/10.1080/00036811.2017.1419197 doi: 10.1080/00036811.2017.1419197
    [13] B. Feng, M. Yin, Decay of solutions for a one-dimensional porous elasticity system with memory: The case of non-equal wave speeds, Math. Mech. Solids, 24 (2018), 2361–2373. https://doi.org/10.1177/1081286518757 doi: 10.1177/1081286518757
    [14] R. Quintanilla, Slow decay for one-dimensional porous dissipation elasticity, Appl. Math. Lett., 16 (2003), 487–491. https://doi.org/10.1016/S0893-9659(03)00025-9 doi: 10.1016/S0893-9659(03)00025-9
    [15] A. Magaña, R. Quintanilla, On the time decay of solutions in one-dimensional theories of porous materials, Int. J. Solids Struct., 43 (2006), 3414–3427. https://doi.org/10.1016/j.ijsolstr.2005.06.077 doi: 10.1016/j.ijsolstr.2005.06.077
    [16] M. Astudillo, H. P. Oquendo, Stability results for a Timoshenko system with a fractional operator in the memory, Appl. Math. Optim., 83 (2021), 1247–1275. https://doi.org/10.1007/s00245-019-09587-w doi: 10.1007/s00245-019-09587-w
    [17] L. Guo, J. C. Vardakis, D. Chou, Y. Ventikos, A multiple-network poroelastic model for biological systems and application to subject-specific modelling of cerebral fluid transport, Int. J. Eng. Sci., 147 (2020), 103204. https://doi.org/10.1016/j.ijengsci.2019.103204 doi: 10.1016/j.ijengsci.2019.103204
    [18] M. Botti, D. A. Di Pietro, O. Le Maître, P. Sochala, Numerical approximation of poroelasticity with random coefficients using polynomial Chaos and hybrid high-order methods, Comput. Methods Appl. Mech. Eng., 361 (2020), 112736. https://doi.org/10.1016/j.cma.2019.112736 doi: 10.1016/j.cma.2019.112736
    [19] H. Dridi, A. Djebabla, Timoshenko system with fractional operator in the memory and spatial fractional thermal effect, Rend. Circ. Mat. Palermo, 70 (2021), 593–621. https://doi.org/10.1007/s12215-020-00513-6 doi: 10.1007/s12215-020-00513-6
    [20] M. M. Cavalcanti, A. Guesmia, General decay rates of solutions to a nonlinear wave equation with boundary condition of memory type, Differ. Integr. Equ., 18 (2005), 583–600. https://doi.org/10.57262/die/1356060186 doi: 10.57262/die/1356060186
    [21] S. A. Messaoudi, W. Al-Khulaifi, General and optimal decay for a quasilinear viscoelastic equation, Appl. Math. Lett., 66 (2017), 16–22. https://doi.org/10.1016/j.aml.2016.11.002 doi: 10.1016/j.aml.2016.11.002
  • Reader Comments
  • © 2025 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(330) PDF downloads(34) Cited by(0)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog