Decaimiento de Soluciones para Mezcla de Dos Solidos Termoelasticos
Decaimiento de Soluciones para Mezcla de Dos Solidos Termoelasticos
Decaimiento de Soluciones para Mezcla de Dos Solidos Termoelasticos
1. Introduction
In recent years an increasing interest has been directed to understand the so-called nonclassical thermomechanical
theories of materials as micromorphic materials, porous materials, etc. One of them is the mixture of materials. The origin of
the modern formulation of continuum thermomechanical theories of mixtures goes back to the papers by Kelly [1], Truesdell
and Toupin [2], Eringen and Ingram [3,4], Mller [5], Green and Naghdi [6,7], Bowen [8], Bowen and Wiese [9], Atkin and
Craine [10] and Bowen and Drumheller [11]. Recent books on this subject are given by Samohyl [12] and Rajagopal and
Tao [13]. For several theories of mixture of materials [8,6,14], the spatial description is used and the independent variables
are the displacement gradients and the relative velocity. The first theory, where the Lagrangian description is proposed and
where the independent variables are the displacement gradients and the relative displacement, was presented in the works
by Bedford and Stern [15,16]. These models have been applied by Tiersten and Jahanmir [17] to derive a theory of composites
where the relative displacement of the individual constituents is infinitesimal. Theory of mixtures is amply accepted in the
scientific community and it has a wide field of applications.
Mathematical studies for this theory have been developed. Questions concerning the existence, uniqueness, continuous
dependence and asymptotic behavior have been studied in [1824]. We believe that the mathematical and physical studies
are necessary issues to clarify the scope of its applicability. We here want to deep the study of the time decay of solutions
for the one-dimensional mixture of two thermoelastic solids. Our aim is to clarify when we can expect the exponential or
polynomial decay for the solutions of our problem. From a mathematical point of view the question is of interest. We have
the coupling of two conservative equations with one dissipative equation. We want to see when the coupling between these
three equations is strong enough to guarantee the exponential or at least polynomial decay.
Corresponding author. Tel.: +39 030 3715735; fax: +39 030 3715745.
E-mail addresses: [email protected] (J.E. Muoz Rivera), [email protected] (M.G. Naso), [email protected] (R. Quintanilla).
0898-1221/$ see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.camwa.2013.03.022
42 J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155
We consider a rod which is composed by a mixture of two interacting continua occupying the finite interval (0, ). The
displacement of each constituent is denoted by u and w , respectively, where u = u(x, t ) : (0, ) (0, T ) R and w =
w(y, t ) : (0, ) (0, T ) R, with T > 0. We assume that the particles under consideration are in the same position at
time t = 0, so that x = y. We suppose also that the temperature = (x, t ) : (0, ) (0, T ) R is the same at both
constituents in each point x and at time t. The mass density of each constituent at time t = 0 is denoted by i , i = 1, 2.
Partial stresses associated with each constituent are T and S respectively, P is the internal body force, is the entropy
density, Q is the heat flux vector and 0 is the absolute temperature in the reference configuration.
When the supply terms are absent the system of equations are proposed by:
the equations of motion
1 utt = Tx P , 2 wtt = Sx + P , (1.1)
the energy equation
(1 + 2 )0 t = Qx , (1.2)
the constitutive equations
T = a11 ux + a12 wx 1 , S = a12 ux + a22 wx 2 , (1.3)
P = (u w), = 1 ux + 2 wx + , Q = K x . (1.4)
If we substitute the constitutive equations (1.3)(1.4) into evolutive equations (1.1)(1.2), we obtain the system of the field
equations
where := K 01 (1 + 2 )1 .
We note that is related with the thermal conductivity, is the thermal capacity and the parameters aij are related with
the conservative part of the system as well as . The coefficients i , i = 1, 2 are the coupling terms between the mechanical
(conservative) and the thermal (dissipative) parts and they will play a relevant role in our study.
We supplement the system (1.5) with the initial conditions
In this paper we restrict our attention to the case that 1 + 2 = 0.1 Without loss of generality we may suppose that 1 +
2 > 0.2
As we pointed out before our problem (1.5)(1.7) contains three partial differential equations. The two firsts are conser-
vative (there is no decay of energy) and the third one is dissipative. We will show that generically the coupling is so strong
that the thermal dissipation brings the whole system to the exponential/polynomial decay.
To describe the rate of decay of the solutions, it is usual to use expressions like slow decay or exponential decay. When
the solutions of the system are exponentially stable we will say that the system has exponential decay. Otherwise we will
speak about slow decay. The difference has relevant physical consequences. When the system has exponential decay, then
the thermomechanical displacements are damped in a way that they can be neglected significantly earlier compared to
the slow decay. Therefore, the nature of the solutions highly determines the temporal behavior of the system. Thus, this
mathematical aspect highly clarifies the physical nature of the system. Concerning the terminology, we talk indifferently of
stability of our problem or of its associated solution.
In a recent paper [18] Alves, et al. proved that, under suitable conditions on the coefficients of the problem, the solution
decays exponentially. We here want to continue this study and to improve the results under the following points. The first
one is that we consider null thermal Dirichlet boundary conditions (see (1.7) 3 ) which is different from the study proposed
in [18]. This is relevant, because the mathematical analysis for these boundary conditions proposes a greater difficulty than
the problem corresponding to the Neumann boundary conditions. Second aspect we want to emphasize is that we give
a complete characterization of the spectrum of the infinitesimal operator of the C0 -semigroup associated to our problem.
Consequently we find the necessary and sufficient conditions to guarantee that the imaginary axis is contained in the re-
solvent, and then we can characterize the decay of the solutions. The last point to highlight here is that, when there are not
eigenvalues in the imaginary axis and the decay is not exponential, we prove the polynomial decay of the solutions.
In classical linear thermoelasticity the asymptotic behavior of solutions, as t +, has been studied by many authors.
We refer, e.g., to the book of Liu and Zheng [25] for a survey on these topics. It is also suitable to point out that the slow decay
which can happens for a one dimensional thermoelastic mixture suggests a difference from the general behavior for the clas-
sical thermoelasticity. In fact, for the classical one dimensional thermoelasticity the decay is always exponential, however
in our case thought the decay is generically exponential, there are a family of cases where the decay is not exponential.
It is also worth recalling several results concerning time decay in thermoelastic mixtures when viscous effects are also
present [19,26,27]. Some regularity properties as analyticity have been proved in these cases, but this is because we have
more dissipation mechanisms. In our paper we only have one dissipative mechanism which is the temperature. The coupling
with the mechanical part is not strong enough to guarantee the analyticity.
Furthermore, for the problem (1.5)(1.7) we cannot expect that the solution always decays. For instance, in case that
1 + 2 = 03 and 2 (a11 + a12 ) = 1 (a12 + a22 ), the system admits isothermal solutions of the form u = w and = 0. They
are undamped and do not decay to zero. Also, when 1 = 2 = 0, the mechanical and thermal parts are uncoupled and the
displacements are described by a conservative equation which cannot decay. They are very particular cases, but we will see
that there are some other cases where there are undamped solutions or where the solutions decay in a way that cannot be
controlled by an exponential function. By applying some recent results (see, e.g., [28]), the polynomial decay will be proved
in these cases.
Our main aim is to show that the C0 -semigroup (and then the solution U (t )) associated to system (1.5)(1.7) is exponen-
tially stable if and only if suitable conditions (see Lemma 3.4, condition (1) or (iii)) on the coefficients are satisfied, whereas
in case that condition (ii) of Lemma 3.4 is verified we will obtain the polynomial decay.
This paper is organized as follows. In Section 2, we establish the well-posedness of system (1.5)(1.7). In Section 3 we
identify the conditions on the constitutive coefficients to guarantee that the imaginary axis is contained in the resolvent (see
Lemma 3.4). Then, under condition (i) or (iii) of Lemma 3.4 the exponential decay of solutions is shown (see Theorem 3.11).
In Section 4 we prove the lack of exponential decay when conditions (i) and (iii) of Lemma 3.4 are not satisfied. In Section 5
we show the polynomial decay in case that the decay is not of exponential type, analyzing separately Dirichlet thermal
boundary conditions and Neumann thermal boundary conditions. Moreover in this later case we find the optimal decay
rate.
In this section we prove the existence and uniqueness of the solutions for the problem (1.5)(1.7). It is worth noting
that a similar result was obtained in [22]. But it is suitable to state it here again, because we will consider the case that the
functions take values in the complex field. Also it will be very useful to recall the energy and the dissipation of the system
as well as the matrix operator associated to it. As the arguments are very similar we only sketch them.
With the usual notation, we introduce the spaces L2 (0, ), H01 (0, ) and H 1 (0, ) acting on a bounded interval (0, ) R.
Let , and denote the L2 -inner product and L2 -norm, respectively. We recall also the Poincar inequality
0 2 x 2 H01 ,
where 0 = 2 /2 > 0. Thus H01 can be endowed with the inner product , H 1 = x , x . In case of Dirichlet thermal
0
boundary conditions, let us consider the vectorial space
H = H01 (0, ) H01 (0, ) L2 (0, ) L2 (0, ) L2 (0, ),
with the inner product
1
(u, w, v, , ), (
u, w
,
v,
,
)H = a11 ux ,
ux + a12 ux , w
x + wx ,
ux + a22 wx , w
x
2
+ (u w), (
uw
) + 1 v,
v + 2 ,
+ ,
.
(u, w, v, , )2H c ux 2 + wx 2 + v2 + 2 + 2 ,
is satisfied. For the sake of simplicity, here and in what follows we shall employ the same symbol c for different constants,
even in the same formula.
Analogously, for Neumann thermal boundary conditions, we introduce
0 0 I 0 0
0 0 0 I 0
1
a11 a12
()xx ()xx + ()x
I I 0 0
A= 1 1 1 1 1 ,
(2.1)
a12 a22 2
()xx + ()xx I ()x
I 0 0
2 2 2 2 2
1 2
0 0 ()x ()x ()xx
with the domain
D(A) = {U H : AU H } .
The domain D(A) is dense in the Hilbert space H . Our initialboundary value problem (1.5)(1.7) can be rewritten as the
following initial value problem
d
U (t ) = AU (t ), U (0) = U0 ,
dt
where U0 = (u0 , w0 , u1 , w1 , 0 ),
Lemma 2.1. The operator A is the infinitesimal generator of a C0 -semigroup of contractions, denoted by S (t ) = eAt , t 0.
Proof. We will prove that A is a dissipative operator and 0 (A) (resolvent of the operator). Then our conclusion will be
a consequence of the LumerPhillips theorem (see, e.g., [29]).
If U D(A) then
ReAU , U H = x 2 0, (2.2)
2
and therefore the operator A is dissipative.
Given F = (f , g , h, p, q) H , we will show that there exists a unique vector U = (u, w, v, , ) in D(A) such that AU
= F , that is
xx = q 1 fx 2 gx L2 (0, ). (2.4)
J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155 45
Thus, there exists a unique H 2 satisfying (2.4). It follows from (2.3) 3 and (2.3) 4 that we have to get u and w satisfying:
B((u, w), (
u, w
)) = a11 ux ,
ux + a12 (ux , w ux ) + a22 wx , w
x + wx , x + u w, (
uw
)
is continuous and coercive, it follows using the LaxMilgram theorem (see, e.g., [30]) that there exists a unique vector (u, w)
satisfying system (2.5)(2.6). Therefore, we obtain the vector U satisfying the equation AU = F . It is easy to show that
U H c F H , for a positive constant c, and we conclude that 0 (A).
We finish this section establishing the following result
Proposition 2.2. For any U0 = (u0 , w0 , u1 , w1 , 0 ) H there exists a unique solution U = (u, w, ut , wt , ) of system (1.5)
(1.7) satisfying
3. Exponential stability
The main aim of this section is to give the necessary and sufficient conditions to guarantee the imaginary axis is contained
in the resolvent and the sufficient conditions to have the exponential decay of solutions.
It is useful to recall the following known result (see [31]):
Theorem 3.1. Let S (t ) = eAt be a C0 -semigroup of contractions on Hilbert space. Then S (t ) is exponentially stable if and only if
iR (A) and
First we need to know when the imaginary axis is contained in the resolvent (A). Putting F = (f , h, g , p, q) H , we
consider the following system
(i I A)U = F ,
or, in components,
iu v = f , (3.2)
iw = g , (3.3)
i1 v (a11 uxx u) (a12 wxx + w) 1 x = h, (3.4)
i2 (a12 uxx + u) (a22 wxx w) 2 x = p, (3.5)
i 1 vx 2 x xx = q, (3.6)
D1 := det(M1 ), D2 := det(M2 ),
(3.8)
0 := 2 1 D2 + 1 2 D1 , 1 := 2 1 1 2 .
46 J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155
Lemma 3.2. If 0 = 0, then operator A has only one imaginary eigenvalue if and only if
D 1 1 (1 + 2 )1
1+ > 0 and N.
0 0
If 0 = 0, then operator A has
m imaginary eigenvalues if and only if
1 = 0 and D1 < 0,
where m is the maximal natural number such that 2 D1 m2 + 2 (1 + 2 ) > 0;
infinitely many imaginary eigenvalues if and only if
1 = 0 and D1 > 0.
Proof. Let us suppose that there exists at least one imaginary eigenvalue i, with R. This implies that there exists U
= 0 such that (iI A)U = 0. Because of dissipative property (3.7), we obtain = 0. Hence, we get
1
(a11 2 a12 1 )uxx = 2 1 u
(1 + 2 )u, (3.12)
2 2
1 1
(a12 2 a22 1 )uxx = 2 2 u + (1 + 2 )u. (3.13)
2 2 2
That is
(1 + 2 ) = 2 2 1 D1 2 , (3.16)
(1 + 2 ) = 1 2 + D2 ,
2 2
(3.17)
where = / , with N. Then, recalling notation (3.8), it follows that there exists only one solution if and only if
2 2 2 2
0 = 0.
In this case we will have
2 1 1 2 1
2 = (1 + 2 ) = (1 + 2 ) ,
2 1 D2 + 1 2 D1 0
(1 + 2 ) D 1 1
D1 + D2
2 = (1 + 2 ) = 1+ .
2 1 D2 + 1 2 D1 2 1 0
As / must be a natural number and 2 must be positive we see
1 D1 1
(1 + 2 ) N and 1 + > 0.
0 0
When 0 = 0 there exists 0 = 0, such that
(D1 , D2 ) = 0 (2 1 , 1 2 ).
So we have from (3.16) and (3.17) that
(1 + 2 ) = 2 2 1 + 0 2 1 2 ,
(1 + 2 ) = 2 1 2 + 0 1 2 2 ,
J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155 47
and there exist solutions if and only if 1 = 0, that is 1 2 = 2 1 . In this case we will have
(1 + 2 ) = 2 1 2 D1 2 .
Therefore
2 1 2 = D1 2 + (1 + 2 ),
and we have infinitely many eigenvalues when D1 > 0. Hence the solution is given by
(1 + 2 ) D1
2 = + 2 .
1 2 1 2
If D1 < 0 then there exist at most m eigenvalues where m is the maximal number satisfying
2 D1 m2 + 2 (1 + 2 ) > 0.
When 0 = 0 and 1 = 0 then there is no imaginary eigenvalue.
Remark 3.3. Note that the case D1 = D2 = 0 is not possible, because in this case there exist 0 , 1 R such that
(a11 , a12 ) = 0 (1 , 2 ) and (a12 , a22 ) = 1 (1 , 2 ),
a11 a12
which implies that matrix a12 a22
is singular.
Lemma 3.4. Inclusion iR (A) holds if and only if one of the following conditions is verified:
D ( + )
(i) 1 + 1 1 < 0 or 1
0
2 1
N, provided 0 = 0.
0
(ii) 0 = 0 and 1 = 0.
(iii) 0 = 0, 1 = 0 and 2 D1 + 2 (1 + 2 ) < 0.
Proof. The proof is an immediate consequence of Lemma 3.2.
Remark 3.5. If condition iR (A) holds, the C0 -semigroup S (t ) is strongly stable, namely, limt S (t )U0 H = 0 for all
U0 H (see [32]).
We now introduce the following coefficients
A1 := 1 2 a11 + 2 1 a12 ,
(3.18)
A2 := 1 2 a12 + 2 1 a22 .
Note that 0 = 2 A1 1 A2 .
Lemma 3.6. Under the assumptions of Lemma 3.4, if U = (u, w, v, , ) is a solution of the resolvent system (3.2)(3.6), for
any > 0 there exists c > 0 such that
1 v + 2 2 c U H F H + c F 2H + A1 ux + A2 wx 2 + 1 ux + 2 wx 2 + u w2 .
c x 2 + A1 ux + A2 wx 2 + u w2 + c F 2H
c U H F H + A1 ux + A2 wx 2 + u w2 + c F 2H .
Finally, after the use of the arithmeticgeometric mean inequality we obtain the desired estimate.
Lemma 3.8. Under assumptions (i), (ii) or (iii) of the Lemma 3.4, we have that for any there exists a positive constant c such
that
1 ux + 2 wx 2 c F 2H + U 2H + c U H F H .
Lemma 3.9. Let us assume that || is large enough. If 0 = 0, then the following estimate
A1 ux + A2 wx 2 c F 2H + U 2H ,
holds.
J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155 49
A1 ux + A2 wx 2 = 1 2 1 v + 2 , A1 v + A2 (1 2 2 1 ) u w, A1 u + A2 w
c
|| U 2H
+ 1 2 1 v + 2 , A1 f + A2 g + (12 2 + 22 1 )x , A1 u + A2 w
+ 2 1 h + 1 2 p, A1 u + A2 w.
If is large enough, it follows that
A1 ux + A2 wx 2 c 1 v + 2 2 + c 2 + c U H F H + U 2H .
Using (3.7) and Lemma 3.6 we get
A1 ux + A2 wx 2 c F 2H + c U H F H + U 2H ,
and our conclusion follows.
A1 ux + A2 wx = 0 (1 ux + 2 wx ) , (3.24)
where
Theorem 3.11. If condition (i) or (iii) of Lemma 3.4 holds, then the C0 -semigroup S (t ) = eAt is exponentially stable.
Proof. Let us assume that condition (i) of Lemma 3.4 holds. From Lemmas 3.8 and 3.9 we get
A1 ux + A2 wx 2 c F 2H + U 2H ,
1 ux + 2 wx 2 c F 2H + U 2H .
By condition (i) we find
ux 2 + wx 2 c A1 ux + A2 wx 2 + 1 ux + 2 wx 2 c F 2H + c U 2H .
In this section we will show that when conditions (i) and (iii) fail there exists a lack of exponential stability. However, we
only propose this analysis in the case that thermal boundary conditions are of the Neumann type. That is, we assume that
the heat flux vector vanishes at x = 0, . Moreover, to make the calculations easier we will assume in this section that the
length of the interval is . That is, = .
Let us consider f = g = q = 0. So the resolvent system can be written as
p1 () = 2 1 2 0 2 , p2 () = 1 2 2 0 2 + 12 , p3 = i + 2 .
The matrix of the system is given by
12
p1 1
1 2 p2 1 .
i 0 p3
It follows that
p1 p3 i12 2
B= .
p1 p2 p3 1 1 2 p3 i2 (12 + 12 p2 )
Taking p2 () = c0 , it follows that
1 2 2 = 0 2 + 12 c0 ,
J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155 51
Remark 4.1. The analysis developed in this section is strongly based on the boundary conditions. With these boundary
conditions it is possible to obtain the precise solution to the different proposed equations. One would like to obtain a similar
result for the case of the Dirichlet thermal boundary conditions. However, we cannot develop the analysis in a similar way
for the alternative conditions, because we cannot obtain the precise solutions.
5. Polynomial decay
In this section we obtain the polynomial stability of solutions in the case (ii) proposed below. Our argument is strongly
based on the following theorem (see [28, Theorem 2.4]).
Theorem 5.1. Let (S (t ))t 0 be a bounded C0 -semigroup on a Hilbert space H with generator A such that iR (A). Then, for a
fixed > 0,
(iI A)1 L(H ) = O (|| ) , S (t )A1 L(H ) = O t 1/ , t .
Theorem 5.2. Let us suppose that condition (ii) holds. If satisfies Neumann boundary condition at least in one end, then the
corresponding solution U (t ) = eAt U0 decays polynomially as
c
U (t )HN U0 D(A) .
t 1/2
Otherwise, for satisfying Dirichlet boundary condition, we have the solution decays as
c
U (t )H U0 D(A) .
t 1/4
Moreover the rate of decay to Neumann boundary condition is optimal.
Proof. We note that in that follows we will use the same symbol R for different functions such that |R| c U H F H .
Using (3.21) and the fact that 0 = 0, there exists a constant 0 = 0 such that
1 v + 2 2 c U H F H + C F 2H + u w2H 1 . (5.2)
On the other hand from (3.23) and recalling the definition of J and K , we have that
i [1 (a22 + a12 )v 2 (a11 + a12 )] D(u w)xx + 3 (u w) + 3 x = F2 , (5.3)
where
3 = (a11 + 2a12 + a22 ), 3 = (D1 + D2 ), F2 = (a22 + a12 )h (a11 + a12 )p.
52 J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155
when 1 = 0. Multiplying Eqs. (4.7) and (5.4) by 0 and 1 2 respectively and summing up the corresponding result, it
follows
ib1 (v ) b2 (u w)xx + b3 (u w) + b4 x = F4 .
In the case 1 = 0, Eq. (5.4) is already of this type. Multiplying the above equation by (v ) and taking its real part we get
1 u w2 = 1 2 (1 v + 2 ), (v ) 0 (1 ux + 2 wx ), (ux wx )
=I0
+ (12 2 + 22 1 )x , (u w) + R, (5.6)
|R| c u w F H .
Multiplying Eq. (4.7) by (1 u + 2 w)/0
1 2 12
(1 ux + 2 wx ), (ux wx ) = (1 v + 2 ), (v ) (1 u + 2 w), (u w)
0 0
1 1
(1 u + 2 w), x (1 v + 2 )x 2 + R.
0 0 2
Substitution of I0 in (3.23) given by the above identity we get
0
1
u w2 = 1 2 1 Re(1 v + 2 ), (v ) + c2 Re(1 u + 2 w), (u w)
1 0
+ c3 Rex , (u w) + c4 Re(1 u + 2 w), x + c5 (1 v + 2 )x 2 + Re(R). (5.7)
Let us suppose that satisfies Neumann boundary condition, that is we assume that x (0) = 0. Integrating Eq. (3.6) over
[0, x] and multiplying the resulting equation by (v ), we get
x
(1 v + 2 ), (v ) = ix , (v ) x , (v ) q ds , (v ) ,
0
where
xx = , x (0) = 0, () = 0.
J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155 53
where
|R| c U HN F HN .
Inserting this inequality into (5.7) we get
u w2 c 1 v + 2 2 + c x 2 + R.
From (5.2) it follows that
v = F1 , 1 v + 2 = G1 ,
we have that
2 F1 + G1 1 F1 + G1
v= , = .
1 + 2 1 + 2
So we obtain
Therefore multiplying Eq. (3.4) by u and (3.5) by w and summing up the product results we get
|R| c U HN F HN .
Since the matrix (aij ) is positive definite, there exists positive constant such that
ux 2 + wx 2 c v2 + 2 + c U HN F HN .
which implies that the solution decay polynomially as t 1/2 . Using inequality (4.8) we conclude that the rate of decay is
optimal for the Neumann conditions in both ends. Because if there exists a better decay such as t 1/(2) , then we will
have that ||2 U H must be bounded. But this is contradictory to inequality (4.8). Finally, if satisfies Dirichlet boundary
condition, then multiplying by 2 Eq. (5.6) we arrive to
v 2 c ||4 1 v + 2 2 + 1 ux + 2 wx 2 + x 2 + ux wx 2 + ||2 R.
U 2H c ||8 F 2H ,
and then our conclusion follows.
54 J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155
Remark 5.3. As we pointed out at the end of the previous section, in the case of the Dirichlet thermal boundary conditions
we do not know how to find a similar result to the one obtained for the Neumann thermal boundary conditions. Thus, we
cannot conclude that the decay estimate, obtained for the Dirichlet thermal boundary conditions, is optimal. However, we
want to recall that results where the decay is different, depending on the boundary conditions, have been proved in other
thermomechanical situations as the Bresse system (see [33]).
Remark 5.4. It is worth noting that it is possible to obtain rate of decay for the L2 -norm of the spatial derivative for the
temperature. We do it for the situation where the decay is slowest, but it can be adapted to the other faster situation. As
c
U H U0 D(A) ,
t 1/4
we see that
c
2 U0 2D(A) .
t 1/2
Using the semigroup property and the linearity we get that
c
AU H U0 D(A2 ) .
t 1/4
In view of [34, Proposition 3.1] we conclude that
c
AU H U0 D(A) .
t 1/8
Therefore we obtain that
c
xx 2 1/4 U0 2D(A) .
t
After the use of the interpolation inequality we see
c
x 2 C xx U0 2D(A) .
t 3/8
This inequality proposes decay rates for the several spatial derivatives of the temperature.
Acknowledgments
References
[1] P. Kelly, A reacting continuum, Internat. J. Engrg. Sci. 2 (2) (1964) 129153.
[2] C. Truesdell, R. Toupin, The classical field theories, in: Handbuch der Physik, vol. Bd. III/1, Springer, Berlin, 1960, With an appendix on tensor fields by
J.L. Ericksen, pp. 226793; appendix, pp. 794858.
[3] A.C. Eringen, J.D. Ingram, A continuum theory of chemically reacting media I, Internat. J. Engrg. Sci. 3 (1965) 197212.
[4] J.D. Ingram, A.C. Eringen, A continuum theory of chemically reacting media II, Internat. J. Engrg. Sci. 5 (1967) 289322.
[5] I. Mller, A thermodynamic theory of mixtures of fluids, Arch. Ration. Mech. Anal. 29 (1968) 344369.
[6] A.E. Green, P.M. Naghdi, A dynamical theory of interacting continua, Internat. J. Engrg. Sci. 3 (1965) 231241.
[7] A.E. Green, P.M. Naghdi, A note on mixtures, Internat. J. Engrg. Sci. 6 (11) (1968) 631635.
[8] R.M. Bowen, Theory of mixtures, in: A.C. Eringen (Ed.), Continuum Physics III, Academic Press, New York, 1976, pp. 689722.
[9] R.M. Bowen, J.C. Wiese, Diffusion in mixtures of elastic materials, Internat. J. Engrg. Sci. 7 (7) (1969) 689722.
[10] R.J. Atkin, R.E. Craine, Continuum theories of mixtures: basic theory and historical development, Quart. J. Mech. Appl. Math. 29 (2) (1976) 209244.
[11] A. Bedford, D.S. Drumheller, Theories of immiscible and structured mixtures, Internat. J. Engrg. Sci. 21 (8) (1983) 863960.
[12] I. Samohyl, Thermodynamics of Irreversible Processes in Fluid Mixtures, Teubner Verlag, Leipzig, NJ, 1987.
[13] K.R. Rajagopal, L. Tao, Mechanics of Mixtures, in: Series on Advances in Mathematics for Applied Sciences, vol. 35, World Scientific Publishing Co. Inc.,
River Edge, NJ, 1995.
[14] T.R. Steel, Applications of a theory of interacting continua, Quart. J. Mech. Appl. Math. 20 (1967) 5772.
[15] A. Bedford, M. Stern, A multi-continuum theory of composite elastic materials, Acta Mech. 14 (1972) 85102.
[16] A. Bedford, M. Stern, Towards a diffusing continuum theory of composite elastic materials, J. Appl. Mech. 38 (1972) 814.
[17] H.F. Tiersten, M. Jahanmir, A theory of composites modeled as inerpenetreting solid continua, Arch. Ration. Mech. Anal. 65 (1977) 153192.
[18] M.S. Alves, J.E. Muoz Rivera, R. Quintanilla, Exponential decay in a thermoelastic mixture of solids, Internat. J. Solids Struct. 46 (78) (2009)
16591666.
[19] M.S. Alves, J.E. Muoz Rivera, M. Seplveda, O.V. Villagrn, Exponential stability in thermoviscoelastic mixtures of solids, Internat. J. Solids Struct. 46
(24) (2009) 41514162.
[20] M.S. Alves, B.M.R. Calsavara, J.E. Muoz Rivera, M. Seplveda, O.V. Villagrn, Analyticity and smoothing effect for the coupled system of equations of
Kortewegde Vries type with a single point singularity, Acta Appl. Math. 113 (1) (2011) 75100.
[21] D. Iean, R. Quintanilla, Existence and continuous dependence results in the theory of interacting continua, J. Elasticity 36 (1) (1994) 8598.
J.E. Muoz Rivera et al. / Computers and Mathematics with Applications 66 (2013) 4155 55
[22] F. Martnez, R. Quintanilla, Some qualitative results for the linear theory of binary mixtures of thermoelastic solids, Collect. Math. 46 (3) (1995)
263277.
[23] R. Quintanilla, Existence and exponential decay in the linear theory of viscoelastic mixtures, Eur. J. Mech. A Solids 24 (2) (2005) 311324.
[24] R. Quintanilla, Exponential decay in mixtures with localized dissipative term, Appl. Math. Lett. 18 (12) (2005) 13811388.
[25] Z. Liu, S. Zheng, Semigroups Associated with Dissipative Systems, in: Chapman & Hall/CRC Research Notes in Mathematics, vol. 398, Chapman &
Hall/CRC, Boca Raton, FL, 1999.
[26] M.S. Alves, J.E. Muoz Rivera, M. Seplveda, O.V. Villagrn, Analyticity of semigroups associates with thermoeviscoelastic mixtures of solid, J. Therm.
Stresses 32 (2009) 9861004.
[27] M.S. Alves, J.E. Muoz Rivera, M. Seplveda, O.V. Villagrn, Stabilization in three-dimensional porous thermoviscoelastic mixtures, Q. J. Math. 64 (1)
(2013) 3757.
[28] A. Borichev, Y. Tomilov, Optimal polynomial decay of functions and operator semigroups, Math. Ann. 347 (2) (2010) 455478.
[29] A. Pazy, Semigroups of Linear Operators and Applications to Partial Differential Equations, in: Applied Mathematical Sciences, vol. 44, Springer-Verlag,
New York, 1983.
[30] H. Brezis, Analyse Fonctionnelle, Thorie et Applications. [Theory and Applications], in: Collection Mathmatiques Appliques pour la Matrise
(Collection of Applied Mathematics for the Masters Degree), Masson, Paris, 1983.
[31] J. Prss, On the spectrum of C0 -semigroups, Trans. Amer. Math. Soc. 284 (2) (1984) 847857. MR 743749 (85f:47044).
[32] F.L. Huang, Strong asymptotic stability of linear dynamical systems in Banach spaces, J. Differential Equations 104 (2) (1993) 307324.
[33] Z. Liu, B. Rao, Energy decay rate of the thermoelastic Bresse system, Z. Angew. Math. Phys. 60 (1) (2009) 5469.
[34] J. Prss, A. Batkai, K. Engel, R. Schnaubelt, Polynomial stability of operator semigroups, Math. Nachr. 279 (1) (2006) 14251440.