Next Article in Journal
Extensions of Bicomplex Hypergeometric Functions and Riemann–Liouville Fractional Calculus in Bicomplex Numbers
Previous Article in Journal
Fuzzy Adaptive Approaches for Robust Containment Control in Nonlinear Multi-Agent Systems under False Data Injection Attacks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Polynomial Decay of the Energy of Solutions of the Timoshenko System with Two Boundary Fractional Dissipations

by
Suleman Alfalqi
1,†,
Hamid Khiar
2,†,
Ahmed Bchatnia
3,*,† and
Abderrahmane Beniani
2,†
1
Department of Mathematics, Applied College in Mahayil, King Khalid University, Abha 61421, Saudi Arabia
2
Engineering and Sustainable Development Laboratory, Faculty of Science and Technology, University of Ain Temouchent, Ain Temouchent 46000, Algeria
3
LR Analyse Non-Linéaire et Géométrie, LR21ES08, Department of Mathematics, Faculty of Sciences of Tunis, University of Tunis El Manar, Tunis 2092, Tunisia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Fractal Fract. 2024, 8(9), 507; https://doi.org/10.3390/fractalfract8090507
Submission received: 16 June 2024 / Revised: 25 August 2024 / Accepted: 26 August 2024 / Published: 28 August 2024
(This article belongs to the Topic Fractional Calculus: Theory and Applications, 2nd Edition)

Abstract

:
In this study, we examine Timoshenko systems with boundary conditions featuring two types of fractional dissipations. By applying semigroup theory, we demonstrate the existence and uniqueness of solutions. Our analysis shows that while the system exhibits strong stability, it does not achieve uniform stability. Consequently, we derive a polynomial decay rate for the system.
MSC:
35L05; 34K35; 93D20

1. Introduction

In this study, we investigate the well-posedness and stabilization of a one-dimensional Timoshenko system of the following form:
ρ 1 φ t t d 1 ( φ x + ψ ) x = 0 , ( x , t ) ( 0 , L ) × ( 0 , + ) , ρ 2 ψ t t d 2 ψ x x + d 1 ( φ x + ψ ) = 0 , ( x , t ) ( 0 , L ) × ( 0 , + ) .
The initial conditions are
φ ( x , 0 ) = φ 0 ( x ) , φ t ( x , 0 ) = φ 1 ( x ) , x ( 0 , L ) , ψ ( x , 0 ) = ψ 0 ( x ) , ψ t ( x , 0 ) = ψ 1 ( x ) , x ( 0 , L ) ,
and the following boundary conditions:
φ ( 0 , t ) = 0 , ψ ( 0 , t ) = 0 , in ( 0 , + ) , ( φ x + ψ ) ( L ) = γ 1 t α , η φ ( L ) , ( ψ x + ψ ) ( L ) = γ 2 t α , η ψ ( L ) , in ( 0 , + ) ,
where ρ 1 , ρ 2 , d 1 , d 2 , γ 1 , and γ 2 are positive constants, η is a non-negative constant, and α is in ( 0 , 1 ) .
The notation t α , η represents the generalized Caputo fractional derivative of order α (where 0 < α < 1 ) with respect to time t. It is defined as
t α , η f ( t ) = 1 Γ ( 1 α ) 0 t ( t s ) α e η ( t s ) d f d s ( s ) d s , η 0 .
The Timoshenko system, traditionally used to model the behavior of beams in mechanical structures, is extended in this study to incorporate fractional derivatives on the boundary conditions. This extension is significant because fractional derivatives are known to provide more accurate models for systems with memory effects and complex dissipation properties, which are often encountered in practical applications, such as material microstructure analysis. The introduction of fractional derivatives into the boundary conditions is novel in the context of the Timoshenko system. This allows us to model more realistic dissipative effects that occur in various materials and structures. Our results provide new insights into the stability characteristics of systems with fractional boundary dissipation, contributing to both the theoretical understanding and practical applications in engineering and materials science.
Fractional calculus has developed into a well-established theory with a solid mathematical foundation, and its applications have gained significant interest in various research fields, including electrical circuits, chemical processes, signal processing, bioengineering, viscoelasticity, and control systems (see [1]). Fractional-order control is not only theoretically important but also has practical implications. It generalizes classical integer-order control theory, enabling more accurate modeling and enhanced control performance. Experimental observations reveal that many phenomena cannot be fully described using traditional Newtonian terms. For example, in viscoelastic materials, the material’s microstructure leads to a combined response involving both elastic solid and viscous fluid characteristics.
The literature (see [2]) establishes that the fractional derivative t α enforces dissipation in the system and ensures that the solution converges to an equilibrium state. Consequently, when applied at the boundaries, fractional derivatives can act as controllers to suppress or attenuate undesirable vibrations.
In [3], B. Mbodje explored the asymptotic behavior of solutions with the following system:
t 2 u ( x , t ) u x 2 ( x , t ) = 0 , ( x , t ) ( 0 , 1 ) × ( 0 , + ) , u ( 0 , t ) = 0 , x u ( 1 , t ) = k t α , η u ( 1 , t ) , α ( 0 , 1 ) , η 0 , k > 0 , u ( x , 0 ) = u 0 ( x ) , t u ( x , 0 ) = v 0 ( x ) .
He demonstrated strong asymptotic stability of the solutions when η = 0 and a polynomial decay rate of t 1 as time approaches infinity when η 0 . The polynomial decay rate was established using the energy method.
Kim and Renardy [4] investigated (1) with two boundary controls of the following form:
K ( φ x + ψ ) ( L ) = γ 1 α t φ ( L ) , b ψ x ( L ) = γ 2 t α ψ ( L ) , for , , t ( 0 , + ) ,
and employed multiplier techniques to prove an exponential decay result for the natural energy of (1). Additionally, Yan [5] established a polynomial decay result when examining two boundary frictional damping terms with polynomial growth near the origin.
Benaissa and Benazzouz [6] investigated the stabilization of the following Timoshenko system with two dynamic boundary control conditions involving fractional derivatives:
ρ 1 ϕ t t K ( ϕ x + ψ ) x = 0 , ( x , t ) ( 0 , L ) × ( 0 , + ) , ρ 2 ψ t t b ψ x x + K ( ϕ x + ψ ) = 0 , ( x , t ) ( 0 , L ) × ( 0 , + ) .
The system is subject to the following boundary conditions:
m 1 ϕ t t ( L , t ) + K ( ϕ x + ψ ) ( L , t ) = γ 1 t α , η ϕ ( L , t ) for t ( 0 , + ) , m 2 ψ t t ( L , t ) + b ψ x ( L , t ) = γ 2 t α , η ψ ( L , t ) for t ( 0 , + ) , ϕ ( 0 , t ) = 0 , ψ ( 0 , t ) = 0 for t ( 0 , + ) ,
where m 1 and m 2 are positive constants. They demonstrated that the system (1) is not uniformly stable using the spectrum method. Polynomial stability was established through semigroup theory and by applying a result from Borichev and Tomilov.
M. Akil et al. [7] studied the Timoshenko system with a single fractional derivative described by
a u t t ( u x + y ) x = 0 , ( x , t ) ( 0 , 1 ) × ( 0 , + ) , b y t t y x x + c ( u x + y ) = 0 , ( x , t ) ( 0 , 1 ) × ( 0 , + ) ,
where a, b, and c are positive constants. The system is subject to the following boundary conditions:
u x ( 1 , t ) + y ( 1 , t ) + γ t α , η u ( 1 , t ) = 0 , t R + , u ( 0 , t ) = y x ( 0 , t ) = y x ( 1 , t ) = 0 .
They demonstrated that the energy of the system (4) and (5) decays polynomially over time. References such as [8,9,10,11,12,13,14] present a comprehensive collection of published works that support the mathematical formulation of problems related to fractional differential equations and the decay rate of the associated energy.
This paper is organized as follows: in Section 2, we demonstrate the well-posedness of system (1) with the boundary conditions (3) using semigroup theory. In Section 3, we prove that the Timoshenko system (1) with the boundary conditions (2) is not exponentially stable, whether the wave propagation speeds are equal ( ρ 1 d 1 = ρ 2 d 2 ) or not ( ρ 1 d 1 ρ 2 d 2 ). In Section 5, we show that the solution decays polynomially to zero when η > 0 , employing a frequency domain approach and a theorem by Borichev and Tomilov.

2. Augmented Model and Well-Posedness of the System

In this section, we focus on reformulating the model (1) into an augmented system. To proceed, we first require the following theorem:
Theorem 1 
([3]). Let μ be the following function:
μ ξ = ξ 2 α n / 2 , ξ R n , 0 < α < 1 .
Consider the system governed by the equation:
t φ ξ , t + | ξ | 2 + η φ ξ , t U ( t ) μ ξ = 0 , ξ R n , η 0 , t > 0 ,
with the initial condition
φ ξ , 0 = 0 ,
and the output defined as
O t = 2 sin α π Γ n 2 + 1 n π n 2 + 1 R n μ ξ φ ξ , t d ξ .
The relationship between the ‘input’ U and the ‘output’ O is then given by
O ( t ) = I 1 α , η U ( t ) = D α , η U ( t ) ,
where
I α , η f ( t ) = 1 Γ α 0 t t τ α 1 e η t τ f τ d τ .
Lemma 1 
([15]). If λ D = λ C ( λ ) + η > 0 ( λ ) 0 , then
τ ( α ) R n μ 2 ξ λ + η + ξ 2 d ξ = λ + η α 1 ,
where τ ( α ) = π 1 sin ( α π ) .
Using the previous theorem, the system (1) can be rewritten as the following augmented model:
ρ 1 φ t t d 1 ( φ x + ψ ) x = 0 , ( x , t ) ( 0 , L ) × ( 0 , + ) , ρ 2 ψ t t d 2 ψ x x + d 1 ( φ x + ψ ) = 0 , ( x , t ) ( 0 , L ) × ( 0 , + ) , t ϕ 1 ξ , t + ξ 2 + η ϕ 1 ξ , t μ ξ t φ ( L , t ) = 0 , t ( 0 , + ) , ξ R , t ϕ 2 ξ , t + ξ 2 + η ϕ 2 ξ , t μ ξ t ψ ( L , t ) = 0 , t ( 0 , + ) , ξ R , ( φ x + ψ ) ( L ) = γ 1 sin ( α π ) π R μ ξ ϕ 1 ξ , t d ξ , t ( 0 , + ) , ψ x ( L ) = γ 2 sin ( α π ) π R μ ξ ϕ 2 ξ , t d ξ , t ( 0 , + ) ,
with the following initial conditions:
φ ( x , 0 ) = φ 0 ( x ) , φ t ( x , 0 ) = φ 1 ( x ) , x ( 0 , L ) , ψ ( x , 0 ) = ψ 0 ( x ) , ψ t ( x , 0 ) = ψ 1 ( x ) , x ( 0 , L ) .
For a solution U = ( φ , φ t , ψ , ψ t , ϕ 1 , ϕ 2 ) of (6), we define the energy by
E ( t ) = 1 2 U H 2 ,
where
U H 2 = 1 2 0 L ρ 1 | φ t | 2 + ρ 2 | ψ t | 2 + d 1 | φ x + ψ | 2 + d 2 | ψ x | 2 d x + ξ 1 2 R | ϕ 1 | 2 d ξ + ξ 2 2 R | ϕ 2 | 2 d ξ ,
with constants ξ i = γ i d i sin ( α π ) π .
Lemma 2. 
Let U = ( φ , φ t , ψ , ψ t , ϕ 1 , ϕ 2 ) be a regular solution of the problem (6). Then, the energy functional defined in (7) satisfies the following relation:
d d t E ( t ) = ξ 1 R | ξ | 2 + η | ϕ 1 ( ξ , t ) | 2 d ξ ξ 2 R | ξ | 2 + η | ϕ 2 ( ξ , t ) | 2 d ξ .
Proof. 
Multiplying Equations (6)1 and (6)3 by φ t and ψ t , respectively, integrating by parts over ( 0 , L ) , and then summing the resulting equations, we obtain
1 2 d d t 0 L ρ 1 | φ t | 2 + ρ 2 | ψ t | 2 + d 1 | φ x + ψ | 2 + d 2 | ψ x | 2 d x ( φ x + ψ ) ( L ) φ t ( L ) = 0 .
Multiplying Equations (6)2 and (6)4 by ξ 1 ϕ 1 and ξ 2 ϕ 2 , respectively, integrating over R , and adding the resulting equations gives us
1 2 d d t ξ 1 R | ϕ 1 | 2 d ξ + ξ 2 R | ϕ 2 | 2 d ξ + ξ 1 R ( ξ 2 + η ) | ϕ 1 ( ξ , t ) | 2 d ξ + ξ 2 R ( ξ 2 + η ) | ϕ 2 ( ξ , t ) | 2 d ξ ξ 1 φ t ( L ) R μ ( ξ ) ϕ 1 ( ξ , t ) d ξ ξ 2 ψ t ( L ) R μ ( ξ ) ϕ 2 ( ξ , t ) d ξ = 0 .
Combining Equations (8) and (9), we obtain
d d t E ( t ) = ξ 1 R ξ 2 + η | ϕ 1 ( ξ , t ) | 2 d ξ ξ 2 R ξ 2 + η | ϕ 2 ( ξ , t ) | 2 d ξ .
This concludes the proof of the lemma. □
We now address the well-posedness of (6). To this end, we introduce the following Hilbert space, referred to as the energy space:
H = H L 1 ( 0 , L ) × L 2 ( 0 , L ) 2 × L 2 ( R ) ,
where H L 1 ( 0 , L ) is defined as
H L 1 ( 0 , L ) = { φ H 1 ( 0 , L ) φ ( 0 ) = 0 } .
For U = ( u 1 , u 2 , u 3 , u 4 , ϕ 1 , ϕ 2 ) T and U ˜ = ( u ˜ 1 , u ˜ 2 , u ˜ 3 , u ˜ 4 , ϕ ˜ 1 , ϕ ˜ 2 ) T , we define the inner product in H as follows:
U , U ˜ H = 0 L ρ 1 u 2 u ˜ 2 ¯ + ρ 2 u 4 u ˜ 4 ¯ d x + 0 L d 1 ( u 1 x + u 3 ) ( u ˜ 1 x + u ˜ 3 ) ¯ d x + 0 L d 2 u 3 x u ˜ 3 x ¯ d x + d 1 2 ξ 1 R ϕ 1 ϕ ˜ 1 ¯ d ξ + d 2 2 ξ 2 R ϕ 2 ϕ ˜ 2 ¯ d ξ .
We transform the system described by (6) into a semigroup framework. By defining the vector function U = ( u 1 , u 2 , u 3 , u 4 , ϕ 1 , ϕ 2 ) T , we express the system (6) in the equivalent form
U = A U , t > 0 , U ( 0 ) = U 0 ,
where U 0 = ( φ 0 , φ 1 , ψ 0 , ψ 1 , ϕ 1 0 , ϕ 2 0 ) T .
The operator A is linear and defined by
A u 1 u 2 u 3 u 4 ϕ 1 ϕ 2 = u 2 d 1 ρ 1 ( u 1 x + u 3 ) x u 4 d 2 ρ 2 u 3 x x d 1 ρ 2 ( u 1 x + u 3 ) ( ξ 2 + η ) ϕ 1 + u 2 ( L ) μ ( ξ ) ( ξ 2 + η ) ϕ 2 + u 4 ( L ) μ ( ξ ) .
The domain of A is then
D A = ( u 1 , u 2 , u 3 , u 4 , ϕ 1 , ϕ 2 ) T H : u 1 , u 3 H 2 H L 1 , ξ ϕ 1 , ξ ϕ 2 L 2 ( R ) , | ξ | 2 + η ϕ 1 + u 2 ( L ) μ ( ξ ) L 2 ( R n ) , ( u 1 x + u 3 ) ( L ) = γ 1 sin ( α π ) π R μ ξ ϕ 1 ξ , t d ξ , | ξ | 2 + η ϕ 2 + u 4 ( L ) μ ( ξ ) L 2 ( R ) , u 3 x ( L ) = γ 2 sin ( α π ) π R μ ξ ϕ 2 ξ , t d ξ . .
We state the following theorem on existence and uniqueness:
Theorem 2. 
1. 
If U 0 D A , then the system (6) has a unique strong solution
U C 0 R + , D A C 1 R + , H .
2. 
If U 0 H , then the system (6) has a unique weak solution
U C 0 R + , H .
Remark 1. 
Note that, while strong solutions satisfy the differential equation pointwise and require higher regularity, weak solutions are defined in an integral sense with lower regularity requirements.
Proof. 
First, we demonstrate that the operator A is dissipative.
For any U = ( u 1 , u 2 , u 3 , u 4 , ϕ 1 , ϕ 2 ) D ( A ) , we have
R e A U , U H = d 1 2 ξ 1 R ξ 2 + η | ϕ 1 ( ξ , t ) | 2 d ξ d 2 2 ξ 2 R ξ 2 + η | ϕ 2 ( ξ , t ) | 2 d ξ 0 .
Hence, A is dissipative.
We will show that the operator I A is surjective.
Given F = ( f 1 , f 2 , f 3 , f 4 , f 5 , f 6 ) H , we prove that there exists
U = ( u 1 , u 2 , u 3 , u 4 , ϕ 1 , ϕ 2 ) D ( A ) satisfying
I A U = F .
That is,
u 1 u 2 = f 1 , u 2 d 1 ρ 1 ( u 1 x + u 3 ) x = f 2 u 3 u 4 = f 3 , u 4 d 2 ρ 2 u 3 x x + d 1 ρ 2 ( u 1 x + u 3 ) = f 4 ϕ 1 ( 1 + ξ 2 + η ) μ ( ξ ) u 2 ( L , t ) = f 5 , ϕ 2 ( 1 + ξ 2 + η ) μ ( ξ ) u 4 ( L , t ) = f 6 .
Then, (11)1, (11)2, (11)5, and (11)6 yield
u 2 = u 1 f 1 , u 4 = u 3 f 3 , ϕ 1 = f 5 1 + ξ 2 + η + μ ( ξ ) u 2 ( L , t ) 1 + ξ 2 + η , ϕ 2 = f 6 1 + ξ 2 + η + μ ( ξ ) u 4 ( L , t ) 1 + ξ 2 + η .
Inserting Equations (11)1 in (11)2 and (11)3 in (11)4, we obtain
ρ 1 u 1 d 1 ( u 1 x + u 3 ) x = ρ 1 ( f 1 + f 2 ) , ρ 2 u 3 d 2 u 3 x x + d 1 ( u 1 x + u 3 ) = ρ 2 ( f 3 + f 4 ) .
Solving system (13) is equivalent to finding u 1 , u 3 H 2 0 , L H L 1 0 , L such that
0 L ρ 1 u 1 d 1 ( u 1 x + u 3 ) x χ d x = 0 L ρ 1 f 1 + f 2 χ d x , 0 L ρ 2 u 3 d 2 u 3 x x + d 1 ( u 1 x + u 3 ) ζ d x = 0 L ρ 2 f 3 + f 4 ζ d x ,
for all χ , ζ H L 1 0 , L .
Inserting Equations (12)3 in (14)1 and (12)4 in (14)2, we obtain
0 L ρ 1 u 1 χ + d 1 ( u 1 x + u 3 ) χ x d x + k 1 u 2 ( L ) χ ( L ) = 0 L ρ 1 f 1 + f 2 χ d x d 1 2 ξ 1 χ ( L ) R f 5 μ ξ 1 + ξ 2 + η d ξ , 0 L ρ 2 u 3 ζ + d 2 u 3 x ζ x + d 1 ( u 1 x + u 3 ) ζ d x + k 2 u 4 ( L ) ζ ( L ) = 0 L ρ 2 f 3 + f 4 ζ d x d 2 2 ξ 2 ζ ( L ) R f 6 μ ξ 1 + ξ 2 + η d ξ ,
where k i = d i 2 ξ i R μ 2 ξ 1 + ξ 2 + η d ξ and with the following boundary conditions:
u 2 ( L ) = u 1 ( L ) f 1 ( L ) , u 4 ( L ) = u 3 ( L ) f 3 ( L ) .
Inserting (16) into (15), we obtain
0 L ρ 1 u 1 χ + d 1 ( u 1 x + u 3 ) χ x d x + k 1 u 1 ( L ) χ ( L ) = 0 L ρ 1 f 1 + f 2 χ d x d 1 2 ξ 1 χ ( L ) R f 5 μ ξ 1 + ξ 2 + η d ξ + k 1 f 1 ( L ) χ ( L ) , 0 L ρ 2 u 3 ζ + d 2 u 3 x ζ x + d 1 ( u 1 x + u 3 ) ζ d x + k 2 u 3 ( L ) ζ ( L ) = 0 L ρ 2 f 3 + f 4 ζ d x d 2 2 ξ 2 ζ ( L ) R f 6 μ ξ 1 + ξ 2 + η d ξ + k 2 f 3 ( L ) ζ ( L ) .
Thus, the problem (17) can be reformulated as the following problem:
a u 1 , u 3 , χ , ζ = L χ , ζ ,
where
a u , v , χ , ζ = 0 L ρ 1 u χ + d 1 ( u x + v ) ( χ x + ζ ) + ρ 2 v ζ + d 2 v x ζ x d x + k 1 u ( L ) χ ( L ) + + k 2 v ( L ) ζ ( L ) ,
and
L χ , ζ = 0 L ρ 1 f 1 + f 2 χ d x d 1 2 ξ 1 χ ( L ) R f 5 μ ξ 1 + ξ 2 + η d ξ + k 1 f 1 ( L ) χ ( L ) + 0 L ρ 2 f 3 + f 4 ζ d x d 2 2 ξ 1 ζ ( L ) R f 6 μ ξ 1 + ξ 2 + η d ξ + k 2 f 3 ( L ) ζ ( L ) .
It is straightforward to verify that a is continuous and coercive and that L is continuous. By applying the Lax–Milgram Theorem A1, we conclude that, for all ( χ , ζ ) H L 1 ( 0 , L ) × H L 1 ( 0 , L ) , the problem (18) has a unique solution ( u 1 , u 3 ) H L 1 ( 0 , L ) × H L 1 ( 0 , L ) .
Using classical elliptic regularity results, it follows from (17) that ( u 1 , u 3 ) H 2 ( 0 , L ) × H 2 ( 0 , L ) . Consequently, the operator I A is surjective. Finally, Theorem 2 follows from the Lumer–Phillips Theorem A2. □

3. Asymptotic Stability

In this section, we analyze the asymptotic stability of the system described by (1)–(3), which requires
lim t + E ( t ) = 0 , U 0 H .
We will examine the spectrum and investigate the strong stability of the C 0 semigroup associated with the system (1)–(3) using the criteria of Arendt–Batty [16].
The main results of this paper are summarized as follows:
Theorem 3. 
The semigroup of contractions ( S ( t ) ) t 0 is strongly stable on the energy space H , meaning that
lim t e A t U 0 H = 0 U 0 H .
First, we need to prove the following lemmas:
Lemma 3. 
The point spectrum of the operator A does not intersect with the imaginary axis, i.e.,
σ p ( A ) i R = ,
where
σ p ( A ) = λ C ker ( λ I A ) { 0 } .
Proof. 
For clarity, we divide the proof into two steps.
Step 1. By direct computation, the equation
A U = 0
with U D ( A ) admits only the trivial solution, i.e., U = 0 . Hence, 0 σ p ( A ) .
Step 2. Suppose that there exists β R * such that
ker ( i β I A ) { 0 } .
Thus, λ = i β is an eigenvalue of A . Let U be an eigenvector in D ( A ) associated with λ , satisfying
( i β I A ) U = 0 .
Equivalently, we have
u 2 = i β u 1 , d 1 ρ 1 ( u 1 x + u 3 ) x = i β u 2 u 4 = i β u 3 , d 2 ρ 2 u 3 x x d 1 ρ 2 ( u 1 x + u 3 ) = i β u 4 ϕ 1 ( ξ 2 + η ) + μ ( ξ ) u 2 ( L , t ) = i β ϕ 1 , ϕ 2 ( ξ 2 + η ) + μ ( ξ ) u 4 ( L , t ) = i β ϕ 1 .
First, a straightforward computation shows that
0 = < i β I A U , U > H = d 1 2 ξ 1 R ξ 2 + η | ϕ 1 ( ξ , t ) | 2 d ξ + d 2 2 ξ 2 R ξ 2 + η | ϕ 2 ( ξ , t ) | 2 d ξ .
We deduce that ϕ 1 = ϕ 2 = 0 a.e. in R .
On the other hand, by (19)5 and (19)6, we obtain,
ϕ j = μ ( ξ ) u 2 j ( L , t ) ξ 2 + η + i β , j = 1 , 2 ,
which yields u 2 ( L , t ) = u 4 ( L , t ) = 0 . Hence, from (19)1 and (19)3, we obtain
u 1 ( L , t ) = u 3 ( L , t ) = 0 , and from ( 10 ) , ( u 1 x + u 3 ) ( L , t ) = u 3 x ( L , t ) = 0 .
Otherwise, replacing (19)1 into (19)2 and (19)3 into (19)4 and setting v = u 1 x + u 3 , we obtain
β 2 u 1 + d 1 ρ 1 v x = 0 , β 2 u 3 + d 2 ρ 2 u 3 x x d 1 ρ 2 v = 0 .
We can rewrite (21) and (20) as
d d x X = B X , t > 0 , X ( L ) = 0 ,
where X : = u 1 , v , u 3 , u 3 x T . The operator B is linear and defined by
B u 1 v u 3 u 3 x = v u 3 ρ 1 β 2 d 1 u 1 u 3 x d 1 d 2 v ρ 2 β 2 d 2 u 3 .
According to Picard’s theorem for ordinary differential equations, the system (3) has a unique solution, which is X = 0 . Thus, u 1 = u 3 = 0 . It follows from (19) that u 2 = u 4 = 0 . Consequently, we obtain U = 0 over the interval ( 0 , L ) , which contradicts the assumption that U 0 . □
Lemma 4. 
The operator ( i β I A ) is surjective.
Proof. 
Let F = ( f 1 , f 2 , f 3 , f 4 , f 5 , f 6 ) H be looking for
U = ( u 1 , u 2 , u 3 , u 4 , ϕ 1 , ϕ 2 ) D ( A ) such that
i β U A U = F .
That is,
i β u 1 u 2 = f 1 , i β u 2 d 1 ρ 1 ( u 1 x + u 3 ) x = f 2 i β u 3 u 4 = f 3 , i β u 4 d 2 ρ 2 u 3 x x + d 1 ρ 2 ( u 1 x + u 3 ) = f 4 ϕ 1 ( i β + ξ 2 + η ) μ ( ξ ) u 2 ( L , t ) = f 5 , ϕ 2 ( i β + ξ 2 + η ) μ ( ξ ) u 4 ( L , t ) = f 6 ,
which is equivalent to
u 2 = i β u 1 f 1 , β 2 u 1 d 1 ρ 1 ( u 1 x + u 3 ) x = f 2 + i β f 1 , u 4 = i β u 3 f 3 , β 2 u 3 d 2 ρ 2 u 3 x x + d 1 ρ 2 ( u 1 x + u 3 ) = f 4 + i β f 3 ϕ 1 = f 5 + μ ( ξ ) u 2 ( L , t ) i β + ξ 2 + η , ϕ 2 = f 6 + μ ( ξ ) u 4 ( L , t ) i β + ξ 2 + η .
To solve the last system (22), it is enough to study the following:
β 2 ρ 1 u 1 + d 1 ( u 1 x + u 3 ) x = ρ 1 f 2 + i β f 1 , β 2 ρ 2 u 3 + d 2 u 3 x x d 1 ( u 1 x + u 3 ) = ρ 2 f 4 + i β f 3 ,
with the conditions
u 1 ( 0 ) = 0 , u 3 ( 0 ) = 0 , u 1 x + u 3 ( L ) = d 1 ξ 1 ( f 5 + i β u 1 ( L ) ) I 1 ( β , η ) f 1 ( L ) I 2 ( β , η ) , u 3 x ( L ) = d 2 ξ 2 ( f 6 + i β u 3 ( L ) ) I 1 ( β , η ) f 3 ( L ) I 2 ( β , η ) ,
where I 1 ( β , η ) = R μ ( ξ ) i β + ξ 2 + η d ξ and I 2 ( β , η ) = R μ 2 ( ξ ) i β + ξ 2 + η d ξ .
We now distinguish two cases.
Step 1. β = 0 and η > 0 : System (23) is equivalent to finding u 1 , u 3 H 2 0 , L H L 1 0 , L such that
0 L d 1 ( u 1 x + u 3 ) x χ d x = 0 L ρ 1 f 2 χ d x , 0 L d 2 u 3 x x + d 1 ( u 1 x + u 3 ) ζ d x = 0 L ρ 2 f 4 ζ d x ,
for all χ , ζ H L 1 0 , L .
Using integration by parts in (24), we deduce that (22) is equivalent to
b u 1 , u 3 , χ , ζ = M χ , ζ ,
where
b u , v , χ , ζ = 0 L d 1 ( u x + v ) ( χ x + ζ ) + d 2 v x ζ x d x ,
and
M χ , ζ = 0 L ρ 1 f 2 χ + ρ 2 f 4 ζ d x d 1 2 ξ 1 f 5 f 1 ( L ) I 2 ( 0 , η ) χ ( L ) d 2 2 ξ 2 f 6 f 3 ( L ) I 2 ( 0 , η ) ζ ( L ) .
It is straightforward to verify that the bilinear form b is continuous and coercive and that the operator M is continuous. By applying the Lax–Milgram theorem, we conclude that, for all χ , ζ H L 1 ( 0 , L ) × H L 1 ( 0 , L ) , the problem (25) has a unique solution ( u 1 , u 3 ) H L 1 ( 0 , L ) × H L 1 ( 0 , L ) . Utilizing classical elliptic regularity, it follows from (24) that ( u 1 , u 3 ) H 2 ( 0 , L ) × H 2 ( 0 , L ) . Consequently, the operator A is surjective.
Step 2. β 0 and η 0 :
Now, we consider the following system:
d 1 ( u 1 x + u 3 ) x = g 1 , d 2 u 3 x x + d 1 ( u 1 x + u 3 ) = g 2 ,
with the conditions
u 1 ( 0 ) = 0 , u 3 ( 0 ) = 0 , u 1 x + u 3 ( L ) = i β d 1 ξ 1 u 1 ( L ) I 1 ( β , η ) , u 3 x ( L ) = i β d 2 ξ 2 u 3 ( L ) I 1 ( β , η ) ,
where ( g 1 , g 2 ) L 2 ( 0 , L ) 2 .
Let us note that L : ( u 1 , u 2 ) ( d 1 ( u 1 x + u 3 ) x , d 2 u 3 x x + d 1 ( u 1 x + u 3 ) ) with domain D ( L ) = { ( u 1 , u 3 ) H L 1 ( 0 , L ) 2 , u 1 ( 0 ) = 0 , u 3 ( 0 ) = 0 , ( u 1 x + u 3 ) ( L ) = i β d 1 ξ 1 u 1 ( L ) I 1 ( β , η ) , u 3 x ( L ) = i β d 2 ξ 2 u 3 ( L ) I 1 ( β , η ) } .
Multiplying (26)1 by χ and (26)2 by ζ , one obtains:
0 L d 1 ( u x + v ) ( χ x + ζ ) + d 2 v x ζ x d x + i β d 1 2 ξ 1 I 1 ( β , η ) u 1 ( L ) χ ( L ) + i β d 2 2 ξ 2 I 2 ( β , η ) u 3 ( L ) ζ ( L ) = 0 L g 1 χ + g 2 ζ d x ,
for all ( χ , ζ ) H L 1 ( 0 , L ) 2 .
By applying the Lax–Milgram theorem once more, we deduce that there exists a unique strong solution ( u 1 , u 3 ) H L 1 ( 0 , L ) 2 D ( L ) for the variational problem (27).
Consequently, it follows that L 1 is compact in L 2 ( 0 , L ) 2 and therefore (23) is equivalent to
β 2 L 1 I U = Φ ,
where U = ( u 1 , u 3 ) and Φ = ρ 1 ( f 2 + i β f 1 ) , ρ 2 ( f 4 + i β f 3 ) and, by Fredholm’s alternative, it suffices to prove that K e r β 2 L 1 I = { 0 } .
For this purpose, let ( y 1 , y 2 ) K e r β 2 L 1 I ; then, we have
β 2 ρ 1 y 1 + d 1 ( y 1 x + y 3 ) x = 0 , β 2 ρ 2 y 3 + d 2 y 3 x x d 1 ( y 1 x + y 3 ) = 0 ,
with the conditions
y 1 ( 0 ) = 0 , y 3 ( 0 ) = 0 , y 1 x + y 3 ( L ) = i β d 1 ξ 1 y 1 ( L ) I 1 ( β , η ) , y 3 x ( L ) = i β d 2 ξ 2 y 3 ( L ) I 1 ( β , η ) .
Multiplying (28)1 by y 1 ¯ and (28)2 by y 3 ¯ , integrating over ( 0 , L ) , one obtains
0 L β 2 ρ 1 | y 1 | 2 + β 2 ρ 2 | y 3 | 2 d 2 | y 3 x | 2 d 1 | y 1 x + y 3 | 2 d x = i β I 1 ( β , η ) d 1 2 ξ 1 | y 1 ( L ) | 2 + d 2 2 ξ 2 | y 3 ( L ) | 2 .
Taking the imaginary part, we deduce that
d 1 2 ξ 1 | y 1 ( L ) | 2 + d 2 2 ξ 2 | y 3 ( L ) | 2 = 0 .
Hence, we deduce that ( y 1 , y 2 ) is the solution of
β 2 ρ 1 y 1 + d 1 ( y 1 x + y 3 ) x = 0 , β 2 ρ 2 y 3 + d 2 y 3 x x d 1 ( y 1 x + y 3 ) = 0 , y 1 ( 0 ) = y 3 ( 0 ) = y 1 x + y 3 ( L ) = y 3 x ( L ) .
Using the same argument used in Lemma 3, we infer that ( y 1 , y 2 ) = ( 0 , 0 ) .
This completes the proof of Lemma 4. □
From Lemmas 3 and 4, we conclude the following result.
Proposition 1. 
σ ( A ) i R = { 0 } .
Proof of Theorem 3. 
Due to Proposition 1, the operator A lacks pure imaginary eigenvalues, and the intersection σ ( A ) i R is countable. By applying the general criterion from Arendt and Batty in [17], the C 0 semigroup ( S ( t ) ) t 0 of contractions is strongly stable. □

4. Lack of Exponential Stability

The primary result of this section is encapsulated in the following theorem.
Theorem 4. 
The semigroup generated by the operator A fails to exhibit exponential stability in the energy space H .
Proof. 
Our objective is to demonstrate that an infinite number of eigenvalues of the operator A approach the imaginary axis, thereby preventing the Timoshenko system (1)–(3) from achieving exponential stability. To begin, we derive the characteristic equation that determines the eigenvalues of A .
Given that A is dissipative, we choose a sufficiently small α 0 > 0 and examine the asymptotic behavior of the eigenvalues λ of A within the set S = λ C : α 0 Re ( λ ) 0 .
We first establish the characteristic equation that the eigenvalues of A must satisfy. Let λ C * be an eigenvalue of A and let U = ( u 1 , λ u 1 , u 3 , λ u 3 , ϕ 1 , ϕ 2 ) D ( A ) be a corresponding eigenvector such that | U | = 1 .
The resulting eigenvalue problem is then given by
λ 2 u 1 d 1 ρ 1 u 1 x x d 1 ρ 1 u 3 x = 0 , λ 2 u 3 d 2 ρ 2 u 3 x x + d 1 ρ 2 u 1 x + d 1 ρ 2 u 3 = 0 , ϕ 1 = μ ( ξ ) u 1 ( L , t ) ξ 2 + η + λ , ϕ 2 = μ ( ξ ) u 3 ( L , t ) ξ 2 + η + λ , u 1 ( 0 ) = u 3 ( 0 ) = 0 , u 1 x + u 3 ( L ) = d 1 ξ 1 u 1 ( L ) I α ( λ , η ) u 3 x ( L ) = d 2 ξ 2 u 3 ( L ) I α ( λ , η ) .
where I α ( λ , η ) = R ξ 2 α 1 ξ 2 + η + λ d ξ .
Equivalently, we have
u 1 x x x x λ 2 ρ 1 d 1 + ρ 2 d 2 u 1 x x + λ 2 ρ 1 ρ 2 d 1 d 2 λ 2 + d 1 ρ 2 u 1 = 0 , u 1 ( 0 ) = u 3 ( 0 ) = 0 , ρ 1 d 1 λ 2 γ 1 γ 2 ( λ + η ) 2 α 2 u 1 ( L ) γ 2 ( λ + η ) α 1 u 1 x ( L ) u 1 x x ( L ) = 0 , ρ 2 γ 1 d 2 λ 2 + d 1 ρ 2 u 1 ( L ) ( λ + η ) α 1 + ρ 1 d 1 + ρ 2 d 2 λ 2 u 1 x ( L ) u 1 x x x ( L ) = 0 ,
where we used I α ( λ , η ) = γ i d i ξ i ( λ + η ) α 1 ,   i = 1 , 2 (see Lemma 2.1 in [6] for the proof).
The characteristic polynomial associated with System (29) is given by
P ( r ) : = r 4 λ 2 ρ 1 d 1 + ρ 2 d 2 r 2 + λ 2 ρ 1 ρ 2 d 1 d 2 λ 2 + d 1 ρ 2 = 0 .
Our goal is to analyze the asymptotic behavior of the large eigenvalues λ of A within the set S. A detailed examination reveals that the polynomial P has four distinct roots when ρ 1 d 1 ρ 2 d 2 2 λ 2 4 ρ 1 d 2 .
Thus, the four distinct roots of P are given by r 1 ( λ ) , r 2 ( λ ) , r 3 ( λ ) = r 1 ( λ ) , and r 4 ( λ ) = r 2 ( λ ) , where
r 1 ( λ ) = λ 2 ρ 1 d 1 + ρ 2 d 2 + ρ 1 d 1 ρ 2 d 2 2 4 ρ 1 d 2 λ 2 , r 2 ( λ ) = λ 2 ρ 1 d 1 + ρ 2 d 2 ρ 1 d 1 ρ 2 d 2 2 4 ρ 1 d 2 λ 2 .
The general solution to (29) can be expressed as
u 1 ( x ) = c 1 sinh ( r 1 ( λ ) x ) + c 2 sinh ( r 2 ( λ ) x ) + c 3 cosh ( r 1 ( λ ) x ) + c 4 cosh ( r 2 ( λ ) x ) .
Applying the boundary conditions in (29) at x = 0 yields c 3 = c 4 = 0 . Additionally, the boundary conditions at x = L in (29) can be expressed as
M λ C = f 1 ( r 1 ) sinh ( r 1 L ) f 2 ( r 1 ) cosh ( r 1 L ) f 1 ( r 2 ) sinh ( r 2 L ) f 2 ( r 2 ) cosh ( r 2 L ) H 1 sinh ( r 1 L ) + g ( r 1 ) cosh ( r 1 L ) H 1 sinh ( r 2 L ) + g ( r 2 ) cosh ( r 2 L ) c 1 c 2 = 0 0
where
f 1 ( r ) = ρ 1 d 1 λ 2 γ 1 γ 2 ( λ + η ) 2 α 2 r 2 , f 2 ( r ) = γ 2 λ ( λ + η ) α 1 r , H 1 = ρ 2 γ 1 d 2 ( λ 2 + d 1 ρ 2 ) ( λ + η ) α 1 , g ( r ) = ρ 1 d 1 + ρ 2 d 2 λ 2 r r 3   and   C = ( c 1 c 2 ) T .
Let det ( M ) represent the determinant of the matrix M. Then, it follows that
d e t M λ = f 1 ( r 1 ) f 1 ( r 2 ) H 1 sinh ( r 1 L ) sinh ( r 2 L ) f 2 ( r 1 ) g ( r 2 ) f 2 ( r 2 ) g ( r 1 ) cosh ( r 1 L ) cosh ( r 2 L ) + f 1 ( r 1 ) g ( r 2 ) + f 2 ( r 2 ) H 1 sinh ( r 1 L ) cosh ( r 2 L ) f 2 ( r 1 ) H 1 + f 1 ( r 2 ) g ( r 1 ) cosh ( r 1 L ) sinh ( r 2 L ) .
The Equation (29) has a non-trivial solution if and only if det ( M ) = 0 .
Case 1. Assuming that ρ 1 d 1 = ρ 2 d 2 , and applying the asymptotic expansion, we obtain
r 1 = λ ρ 1 d 1 1 + i d 1 ρ 1 d 2 1 λ = ρ 1 d 1 λ + i 2 d 1 d 2 + 1 8 d 1 3 / 2 ρ 1 d 2 1 λ + o ( 1 λ ) , r 2 = λ ρ 1 d 1 1 i d 1 ρ 1 d 2 1 λ = ρ 1 d 1 λ i 2 d 1 d 2 + 1 8 d 1 3 / 2 ρ 1 d 2 1 λ + o ( 1 λ ) .
Next, inserting (31) into (30), we obtain
( f 1 ( r 1 ) f 1 ( r 2 ) ) H 1 = 2 i γ 1 ρ 1 d 2 λ 2 + α + O ( λ 1 + α ) , f 2 ( r 1 ) g ( r 2 ) f 2 ( r 2 ) g ( r 1 ) = 2 i γ 2 ρ 1 d 2 λ 3 + α + O ( λ 2 + α ) , f 1 ( r 1 ) g ( r 2 ) + f 2 ( r 2 ) H 1 = i ρ 1 d 2 λ 3 + O ( λ 2 ) , f 2 ( r 1 ) H 1 + f 1 ( r 2 ) g ( r 1 ) = i γ 1 γ 2 ρ 1 d 2 ρ 1 d 2 λ 2 + 2 α + O ( λ 1 + 2 α ) .
where we used
λ + η α 1 = λ α 1 + O λ α 2 .
From (31), we obtain
sinh ( r 1 L ) = sinh ρ 1 d 1 λ L + O ( 1 λ ) cosh ( r 1 L ) = cosh ρ 1 d 1 λ L + O ( 1 λ ) , sinh ( r 2 L ) = sinh ρ 1 d 1 λ L + O ( 1 λ ) cosh ( r 2 L ) = cosh ρ 1 d 1 λ L + O ( 1 λ ) .
Therefore, from (32) and (33), we obtain
det M λ 3 + α = 2 i γ 2 ρ 1 d 2 cosh 2 ρ 1 d 1 λ L i ρ 1 d 2 1 λ α 1 2 sinh 2 ρ 1 d 1 λ L i γ 1 γ 2 ρ 1 d 2 ρ 1 d 2 1 λ 1 α 1 2 sinh 2 ρ 2 d 2 λ L + O λ 1 .
Let λ be a large eigenvalue of A . Then, according to (34), λ is an approximate root of the following asymptotic equation:
φ ( λ ) = φ 0 ( λ ) + φ 1 ( λ ) λ min ( α , 1 α ) + O ( λ 1 ) ,
where φ 0 ( λ ) = 2 i γ 2 ρ 1 d 2 cosh 2 ρ 1 d 1 λ L and φ 1 ( λ ) = c sinh 2 ρ 1 d 1 λ L .
It is important to note that φ 0 and φ 1 remain bounded within the strip α 0 Re ( λ ) 0 .
The roots of φ 0 are given by i ( 2 n + 1 ) π 2 L d 2 ρ 1 , k Z , and we conclude using Rouché’s theorem.
Case 2.  ρ 1 d 1 ρ 2 d 2 is treated in a similar way.
The proof of Theorem 4 is thus concluded. □

5. The Rate of Decay of the C 0 Semigroup

This section focuses on analyzing the asymptotic behavior of the solution to the system (1)–(3). We demonstrate the polynomial stability of the system (1)–(3):
Theorem 5. 
Let ( S ( t ) ) t 0 be the bounded C 0 semigroup on the Hilbert space H ; with generator A , we have
S ( t ) A ( I A ) 1 = O ( t 1 ) , t .
The following corollary follows from Theorem 5 and Remark 8.5 in [16].
Corollary 1. 
Given ( u 0 , u 1 ) D ( A ) R ( A ) . There exist constants C, t 0 > 0 such that, for all t t 0 ,
S ( t ) ( u 0 , u 1 ) C t ( u 0 , u 1 ) D ( A ) R ( A ) .
To establish Theorem 5, we derive a specific resolvent estimate using a result from Batty, Chill, and Tomilov as presented in [16]. More precisely, we have the following lemmas:
Lemma 5. 
The operator A defined by (2) and (10) satisfies
lim sup β R , | β | + i β I A 1 L ( H ) < .
Proof. 
By contradiction, suppose that
lim sup β R , | β | i β I A 1 L ( H ) = .
There exists a sequence of real numbers β n > 0 with β n and a sequence of vectors ( U n ) n D ( A ) such that
U n H = 1 ,
and
( i β n I A ) U n = : F n = o ( 1 ) in H .
Our objective is to show that U n converges to zero, leading to a contradiction.
Note that (35) is equivalent to
i β n u 1 n u 2 n = f 1 n 0 in H L 1 , i β n u 2 n d 1 ρ 1 u 1 x n + u 3 n x = f 2 n 0 in L 2 , i β n u 3 n u 4 n = f 3 n 0 in H L 1 , i β n u 4 n d 2 ρ 2 u 3 x x n + d 1 ρ 2 u 1 x n + u 3 n = f 4 n 0 in L 2 , ϕ 1 n ( i β n + ξ 2 + η ) μ ( ξ ) u 2 n ( L , t ) = f 5 n 0 in L 2 , ϕ 2 n ( i β n + ξ 2 + η ) μ ( ξ ) u 4 n ( L , t ) = f 6 n 0 in L 2 , ( u 1 x n + u 3 n ) ( L ) = γ 1 sin ( α π ) π R μ ξ ϕ 1 n ξ , t d ξ , u 3 x n ( L ) = γ 2 sin ( α π ) π R μ ξ ϕ 2 n ξ , t d ξ .
First, taking the real part of the inner product of (35) with U n in H , we obtain
R e i β n I A ) U n , U n H = d 1 2 ξ 1 R ξ 2 + η | ϕ 1 n ( ξ , t ) | 2 d ξ + d 2 2 ξ 2 R ξ 2 + η | ϕ 2 n ( ξ , t ) | 2 d ξ .
Then, from (35) and (37), we obtain
ϕ 1 n L 2 = ϕ 2 n L 2 = o ( 1 ) ,
and we deduce that
( u 1 x n + u 3 n ) ( L ) = o ( 1 )   and   u 3 x n ( L ) = o ( 1 ) .
Note also that we deduce from (36)1 and (36)3 that u 1 n L 2 = o ( 1 ) and u 3 n L 2 = o ( 1 ) .
Now, inserting (36)1 into (36)2 and (36)3 into (36)4, we obtain
u 2 n = i β u 1 n f 1 n , β 2 u 1 n d 1 ρ 1 ( u 1 x n + u 3 n ) x = f 2 n + i β f 1 n , u 4 n = i β u 3 n f 3 n , β 2 u 3 n d 2 ρ 2 u 3 x x n + d 1 ρ 2 ( u 1 x n + u 3 n ) = f 4 n + i β f 3 n ϕ 1 n = f 5 n + μ ( ξ ) u 2 n ( L , t ) i β + ξ 2 + η , ϕ 2 n = f 6 n + μ ( ξ ) u 4 n ( L , t ) i β + ξ 2 + η .
We will break the proof into several steps, and, for simplicity, we will omit the index n.
Step 1. Multiplying (38)2 by x u 1 x ¯ and integrating over ( 0 , L ) , one obtains
β 2 0 L | u 1 | 2 2 d x β 2 L 2 | u 1 ( L ) | 2 + d 1 L ρ 1 0 L | u 1 x | 2 2 d x d 1 ρ 1 | u 1 x ( L ) | 2 d 1 ρ 1 0 L x ( u 3 x u 1 x ¯ ) d x = 0 L x f 2 u 1 x ¯ d x i β 0 L ( f 1 + x f 1 x ) u 1 ¯ d x + i β L f 1 ( L ) u 1 ( L ) ¯ .
Using the fact that 0 L x f 2 u 1 x ¯ d x = o ( 1 ) , 0 L ( f 1 + x f 1 x ) β u 1 ¯ d x = o ( 1 ) and
| f 1 ( L ) | = | 0 L f 1 x d x | L f 1 H L 1 ( 0 , L ) = o ( 1 ) .
Therefore,
β 2 0 L | u 1 | 2 2 d x β 2 L 2 | u 1 ( L ) | 2 + d 1 ρ 1 0 L | u 1 x | 2 2 d x d 1 L ρ 1 | u 1 x ( L ) | 2 d 1 ρ 1 0 L x ( u 3 x u 1 x ¯ ) d x = o ( 1 ) 1 + ( β u 1 ( L ) ¯ .
Analogously, by multiplying Equation (38)4 by x u 3 x ¯ , we obtain
( β 2 d 1 ρ 2 ) 0 L | u 3 | 2 2 d x + L ( d 1 ρ 2 β 2 ) | u 3 ( L ) | 2 2 + d 2 ρ 2 0 L | u 3 x | 2 2 d x d 2 L | u 3 x | 2 2 ρ 2 + d 1 ρ 2 0 L x ( u 1 x u 3 x ¯ ) d x = o ( 1 ) 1 + ( β u 3 ( L ) ¯ .
Consequently, estimates (39) and (40) give
β 2 ρ 1 0 L | u 1 | 2 2 d x β 2 ρ 1 L 2 | u 1 ( L ) | 2 + d 1 0 L | u 1 x | 2 2 d x d 1 L | u 1 x ( L ) | 2 + ( β 2 ρ 2 d 1 ) 0 L | u 3 | 2 2 d x + L ( d 1 β 2 ρ 2 ) | u 3 ( L ) | 2 2 + d 2 0 L | u 3 x | 2 2 d x = o ( 1 ) 1 + ( β u 1 ( L ) ¯ + ( β u 3 ( L ) ¯ .
Step 2. From (38)5 and (38)6, we have
ϕ 1 2 = f 5 2 β 2 + ξ 2 + η 2 + μ 2 ( ξ ) u 2 2 ( L , t ) β 2 + ξ 2 + η 2 + 2 R e f 5 μ ( ξ ) u 2 ( L , t ) β 2 + ξ 2 + η 2 , ϕ 2 2 = f 6 2 β 2 + ξ 2 + η 2 + μ 2 ( ξ ) u 4 2 ( L , t ) β 2 + ξ 2 + η 2 + 2 R e f 6 μ ( ξ ) u 4 ( L , t ) β 2 + ξ 2 + η 2 .
Given that ϕ 1 L 2 = ϕ 2 L 2 = o ( 1 ) and R μ ( ξ ) β 2 + ( ξ 2 + η ) 2 d ξ > 0 , we can deduce that
u 2 ( L ) 0   and   u 4 ( L ) 0 .
Consequently, using (38)1 (38)3, and the fact that | f i ( L ) | L f i H L 1 = O ( 1 ) , i = 1 , 3 , we obtain
β u 1 ( L ) 0   and   β u 3 ( L ) 0 ,
and, next, u 1 x ( L ) = o ( 1 ) .
Combining this with (41), (38)1, and (38)3, we obtain that u 2 L 2 = o ( 1 ) and u 4 L 2 = o ( 1 ) and, next, U H = o ( 1 ) , which contradicts the hypothesis that U H = 1 .
Thus, the proof of the lemma is complete. □
Lemma 6. 
The operator A defined by (2) and (10) satisfies
lim sup β R , β 0 β i β I A 1 L ( H ) < .
Proof. 
By contradiction, suppose that
lim sup β R , β 0 β i β I A 1 L ( H ) = .
Put β = 1 γ so (42) is equivalent to
lim sup γ R , | γ | + γ 1 ( i γ 1 I A ) 1 L ( H ) = .
Then, there exists a sequence of real numbers γ n > 1 with γ n and a sequence of functions ( U n ) n D ( A ) such that
U n H = 1 ,
and
γ n ( i γ n 1 I A ) U n = : F n = o ( 1 ) , in H .
We will demonstrate that U n H = o ( 1 ) , which contradicts the assumption regarding U n . To simplify, we will drop the index n in the following and divide the remainder of the proof into two steps for clarity.
Step 1. In fact, (44) can be written as
i u 1 γ u 2 = f 1 0 i n H L 1 , i u 2 γ d 1 ρ 1 ( u 1 x + u 3 ) x = f 2 0 i n L 2 , i u 3 γ u 4 = f 3 0 i n H L 1 , i u 4 γ d 2 ρ 2 u 3 x x + γ d 1 ρ 2 ( u 1 x + u 3 ) = f 4 0 i n L 2 , ϕ 1 ( i + γ ξ 2 + γ η ) γ μ ( ξ ) u 2 ( L , t ) = f 5 0 i n L 2 , ϕ 2 ( i + γ ξ 2 + γ η ) γ μ ( ξ ) u 4 ( L , t ) = f 6 0 i n L 2 .
Since U n is bounded by 1 in H and F n converges to 0 in H , (45)1 and (45)3 imply that
u 2 L 2 = u 4 L 2 = o ( 1 ) .
By taking the real part of the inner product of (44) with U in H , we obtain
R e γ ( i γ 1 I A ) U , U H = γ d 1 2 ξ 1 R ξ 2 + η | ϕ 1 n ( ξ , t ) | 2 d ξ + γ d 2 2 ξ 2 R ξ 2 + η | ϕ 2 n ( ξ , t ) | 2 d ξ .
Then, from (43) and (44), we obtain
ϕ 1 L 2 = ϕ 2 L 2 = o ( 1 ) .
Considering that U D ( A ) , we deduce that
( u 1 x + u 3 ) ( L ) = o ( 1 )   and   u 3 x ( L ) = o ( 1 ) .
Next, we use the same steps used in the previous lemma to obtain
γ u 2 ( L ) 0   and   γ u 4 ( L ) 0 ,
and we infer that
u 1 ( L ) 0   and   u 3 ( L ) 0 .
Step 2. Now, from (45)2, (45)4, and (46), we obtain
γ ( u 1 x + u 3 ) x = ρ 1 d 1 ( f 2 i u 2 ) : = g 1 0 in L 2 , γ u 3 x x γ d 1 d 2 ( u 1 x + u 3 ) = ρ 2 d 2 ( f 4 i u 4 ) : = g 2 0 in L 2 .
Consequently, V : = u 1 , u 1 x + u 3 , u 3 , u 3 x T is the solution of
d d x V = C V + G V ( L ) = u 1 ( L ) , ( u 1 x + u 3 ) ( L ) , u 3 ( L ) , u 3 x ( L ) T .
Here, C is defined as
C = 0 1 1 0 0 0 0 0 0 0 0 1 0 d 1 d 2 0 0
and G : = ( 0 , 1 γ g 1 , 0 , 1 γ g 2 ) T .
On one hand, by Duhamel’s formula, one obtains
V ( x ) = exp C ( x L ) V ( L ) + L x exp C ( x s ) G ( s ) d s .
Conversely, a straightforward calculation yields the characteristic polynomial of C:
P ( λ ) = λ 4 .
Hence, we obtain
exp ( C x ) = 1 x d 1 x 3 6 d 2 x x 2 2 0 1 0 0 0 d 1 x 2 2 d 2 1 x 0 d 1 x d 2 0 0 , x R .
Therefore,
u 1 = u 1 ( L ) + x L d 1 ( x L ) 3 6 d 2 ( u 1 x + u 3 ) ( L ) ( x L ) u 3 ( L ) ( x L ) 2 2 u 3 x ( L ) + 1 γ L x ( x s d 1 ( x s ) 3 6 d 2 ) g 1 ( x s ) 2 2 g 2 d s ,
and, from (48) and (49), it is simple to show that u 1 converges to 0 in L 2 . We also show that u 1 x , u 3 and u 3 x converge to 0 in L 2 .
Finally, combining the last result with (46) and (47), we find that U H = o ( 1 ) , which contradicts the assumption that U H = 1 . □
Proof of Theorem 5. 
It follows immediately from Lemma 5, Lemma 6, and Theorem 7.6 in [16] that
S ( t ) A ( I A ) 1 = O ( t 1 ) , t .

Author Contributions

Conceptualization, S.A. and H.K.; Methodology, A.B. (Ahmed Bchatnia) and A.B. (Abderrahmane Beniani); Validation, A.B. (Ahmed Bchatnia); Investigation, S.A.; Resources, S.A., H.K. and A.B. (Ahmed Bchatnia); Writing—original draft, S.A., H.K., A.B. (Ahmed Bchatnia) and A.B. (Abderrahmane Beniani); Visualization, A.B. (Abderrahmane Beniani); Supervision, A.B. (Ahmed Bchatnia). All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khaled University for funding this work through Large Research Project under grant number RGP2/37/45.

Conflicts of Interest

The authors declare no conflicts of interest.

Appendix A

In this appendix, we present several key theorems from the literature that have been utilized in various proofs throughout this article. We begin with the Lax–Milgram theorem, which serves as a type of representation theorem for bounded linear functionals on a Hilbert space H :
Theorem A1. 
Let B be a bounded, coercive bilinear form on a Hilbert space H . Then, for every bounded linear functional ℓ on H , there exists a unique element x H such that ( x ) = B ( x , x ) for all x in H .
We then proceed to the Lumer–Phillips theorem, which provides a necessary and sufficient condition for a linear operator in a Banach space to generate a contraction semigroup:
Theorem A2. 
Let A be a densely defined operator on X. Then, A generates a C 0 -semigroup of contractions on X if and only if
1. 
A is dissipative;
2. 
( I A ) D ( A ) = X , where D ( A ) is the domain of the operator A .

References

  1. Podlubny, I. Fractional Differential Equations. In Mathematics in Science and Engineering; Academic Press: London, UK, 1999; Volume 198. [Google Scholar]
  2. Mbodje, B.; Montseny, G. Boundary fractional derivative control of the wave equation. IEEE Trans. Automat. Contr. 1995, 40, 368–382. [Google Scholar] [CrossRef]
  3. Mbodje, B. Wave energy decay under fractional derivative controls. SIAM J. Control Optim. 2006, 23, 237–257. [Google Scholar] [CrossRef]
  4. Kim, J.U.; Renardy, Y. Boundary control of the Timoshenko beam. SIAM J. Control Optim. 1987, 25, 1417–1429. [Google Scholar] [CrossRef]
  5. Yan, Q.X. Boundary stabilization of Timoshenko beam. Syst. Sci. Math. Sci. 2000, 13, 376–384. [Google Scholar]
  6. Benaissa, A.; Benazzouz, S. Well-posedness and asymptotic behavior of Timoshenko beam system with dynamic boundary dissipative feedback of fractional derivative type. Z. Angew. Math. Phys. 2017, 68, 94. [Google Scholar] [CrossRef]
  7. Akil, M.; Chitour, Y.; Ghader, M.; Wehbe, A. Stability and exact controllability of a Timoshenko system with only one fractional damping on the boundary. Asymptot. Anal. 2020, 119, 221–280. [Google Scholar] [CrossRef]
  8. Akil, M.; Wehbe, A. Stabilization of multidimensional wave equation with locally boundary fractional dissipation law under geometric conditions. Math. Control Relat. Fields 2019, 9, 97–116. [Google Scholar] [CrossRef]
  9. Bchatnia, A.; Mufti, K.; Yahia, R. Stability of an infinite star-shaped network of strings by a Kelvin-Voigt damping. Math. Methods Appl. Sci. 2022, 45, 2024–2041. [Google Scholar] [CrossRef]
  10. Beniani, A.; Bahri, N.; Alharbi, R.; Bouhali, K.; Zennir, K. Stability for Weakly Coupled Wave Equations with a General Internal Control of Diffusive Type. Axioms 2023, 12, 48. [Google Scholar] [CrossRef]
  11. Bagley, R.L.; Torvik, P.J. A theoretical basis for the application of fractional calculus to viscoelasticity. J. Rheol. 1983, 27, 201–210. [Google Scholar] [CrossRef]
  12. Batty, C.J.K.; Duyckaerts, T. Non-uniform stability for bounded semi-groups on Banach spaces. J. Evol. Equ. 2008, 8, 765–780. [Google Scholar] [CrossRef]
  13. Lyubich, I.; Phng, V.Q. Asymptotic stability of linear differential equations in Banach spaces. Stud. Math. 1988, 88, 37–42. [Google Scholar] [CrossRef]
  14. Soufyane, A. Stabilisation de la poutre de Timoshenko. C. R. Acad. Sci. Paris Sér. I Math. 1999, 328, 731–734. [Google Scholar] [CrossRef]
  15. Achouri, Z.; Amroun, N.; Benaissa, A. The Euler Bernoulli beam equation with boundary dissipation of fractional derivative. Math. Methods Appl. Sci. 2017, 40, 3837–3854. [Google Scholar] [CrossRef]
  16. Batty, C.J.K.; Chill, R.; Tomilov, Y. Fine scales of decay of operator semigroups. J. Eur. Math. Soc. 2016, 18, 853–929. [Google Scholar] [CrossRef]
  17. Arendt, W.; Batty, C.J.K. Tauberian theorems and stability of one-parameter semigroups. Trans. Am. Math. Soc. 1988, 306, 837–852. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alfalqi, S.; Khiar, H.; Bchatnia, A.; Beniani, A. Polynomial Decay of the Energy of Solutions of the Timoshenko System with Two Boundary Fractional Dissipations. Fractal Fract. 2024, 8, 507. https://doi.org/10.3390/fractalfract8090507

AMA Style

Alfalqi S, Khiar H, Bchatnia A, Beniani A. Polynomial Decay of the Energy of Solutions of the Timoshenko System with Two Boundary Fractional Dissipations. Fractal and Fractional. 2024; 8(9):507. https://doi.org/10.3390/fractalfract8090507

Chicago/Turabian Style

Alfalqi, Suleman, Hamid Khiar, Ahmed Bchatnia, and Abderrahmane Beniani. 2024. "Polynomial Decay of the Energy of Solutions of the Timoshenko System with Two Boundary Fractional Dissipations" Fractal and Fractional 8, no. 9: 507. https://doi.org/10.3390/fractalfract8090507

Article Metrics

Back to TopTop