Asymptotic behavior of a nonisothermal viscous Cahn–Hilliard equation with inertial term

Asymptotic behavior of a nonisothermal viscous Cahn–Hilliard equation with inertial term

J. Differential Equations 239 (2007) 38–60 www.elsevier.com/locate/jde Asymptotic behavior of a nonisothermal viscous Cahn–Hilliard equation with ine...

227KB Sizes 0 Downloads 17 Views

J. Differential Equations 239 (2007) 38–60 www.elsevier.com/locate/jde

Asymptotic behavior of a nonisothermal viscous Cahn–Hilliard equation with inertial term Maurizio Grasselli a,1 , Hana Petzeltová b,2 , Giulio Schimperna c,∗,1,3 a Dipartimento di Matematica “F. Brioschi”, Politecnico di Milano, Via Bonardi, 9, I-20133 Milano, Italy b Mathematical Institute AS CR, Žitná, 25, CZ-115 67 Praha, Czech Republic c Dipartimento di Matematica “F. Casorati”, Università degli Studi di Pavia, Via Ferrata, 1, I-27100 Pavia, Italy

Received 6 November 2006; revised 7 May 2007 Available online 17 May 2007

Abstract We consider a differential model describing nonisothermal fast phase separation processes taking place in a three-dimensional bounded domain. This model consists of a viscous Cahn–Hilliard equation characterized by the presence of an inertial term χtt , χ being the order parameter, which is linearly coupled with an evolution equation for the (relative) temperature ϑ. The latter can be of hyperbolic type if the Cattaneo– Maxwell heat conduction law is assumed. The state variables and the chemical potential are subject to the homogeneous Neumann boundary conditions. We first provide conditions which ensure the well-posedness of the initial and boundary value problem. Then, we prove that the corresponding dynamical system is dissipative and possesses a global attractor. Moreover, assuming that the nonlinear potential is real analytic, we establish that each trajectory converges to a single steady state by using a suitable version of the Łojasiewicz–Simon inequality. We also obtain an estimate of the decay rate to equilibrium. © 2007 Elsevier Inc. All rights reserved. MSC: 35B40; 35B41; 35R35; 80A22

* Corresponding author.

E-mail addresses: [email protected] (M. Grasselli), [email protected] (H. Petzeltová), [email protected] (G. Schimperna). 1 The author was also supported by the Italian MIUR PRIN Research Project Modellizzazione Matematica ed Analisi dei Problemi a Frontiera Libera. 2 The author was supported by the Academy of Sciences of the Czech Republic, Institutional Research Plan No. AV0Z10190503 and by Grant IAA100190606 of GA AV CR. 3 The author was also supported by the HYKE Research Training Network. 0022-0396/$ – see front matter © 2007 Elsevier Inc. All rights reserved. doi:10.1016/j.jde.2007.05.003

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

39

Keywords: Phase-field models; Bounded absorbing sets; Global attractors; Convergence to stationary solutions; Łojasiewicz–Simon inequality

1. Introduction Consider a bounded domain Ω ⊂ R3 with smooth boundary ∂Ω which contains, for any time t  0, a two-phase system subject to nonisothermal phase separation. A well-known evolution system which describes this kind of process is (see [8], cf. also [7]) 

(ϑ + χ)t − ϑ = 0,   χt −  −χ + φ(χ) − ϑ = 0,

(1.1)

in Ω ×(0, ∞). Here ϑ denotes the (relative) temperature around a given critical one, χ represents the order parameter (or phase-field) and φ is the derivative of a suitable smooth double well potential (e.g., φ(r) = r 3 − ar, a > 0). For the sake of simplicity, all the constants have been set equal to one. In the isothermal case, the following singular perturbation of Cahn–Hilliard equation has been examined in several papers (see [6,14,22,23,55,56] and references therein)   εχtt + χt −  −χ + αχt + φ(χ) = 0,

(1.2)

where ε > 0 is a small inertial parameter and α  0 is a viscosity coefficient. The inertial term εχtt accounts for fast phase separation processes (see, e.g., [20]) and, according to [21], its presence seems to give a better description of the spinodal decomposition. Regarding the viscous term αχt the reader is referred to [40] for details. The mathematical works quoted above are concerned with the analysis of the infinite-dimensional dissipative dynamical system generated by (1.2) endowed with suitable boundary conditions. We recall that the case α = 0 has been analyzed so far in one spatial dimension only, since in two and three dimensions, many issues are still open (see, however, [47,50]). In this paper we consider Eq. (1.2) in the nonisothermal case, namely, ⎧ ⎪ ⎨ (ϑ + χ)t + ∇ · q = 0, σ qt + q = −∇ϑ, ⎪ ⎩ εχ + χ − −χ + αχ + φ(χ) − ϑ  = 0, tt t t

(1.3)

where σ ∈ [0, 1]. Observe that the standard Fourier law is obtained when σ = 0. Otherwise, we have the so-called Maxwell–Cattaneo heat conduction law which entails that ϑ propagates at finite speed (see, e.g., [27–29] and their references). System (1.3) is subject to the initial conditions ϑ(0) = ϑ0 ,

σ q(0) = σ q0 ,

χ(0) = χ0 ,

χt (0) = χ1 ,

in Ω,

(1.4)

and to the no-flux boundary conditions q · n = ∇χ · n = ∇(χ) · n = 0,

on ∂Ω × (0, ∞),

(1.5)

40

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

where n stands for the outward normal derivative and · indicates the usual Euclidean scalar product. Observe that (1.3) reduces to (1.1) when ε = α = 0. Moreover, note that (1.5) are equivalent to assume the first two conditions and ∇u · n = 0, where u = −χ + αχt + φ(χ) − ϑ is the so-called chemical potential. Here we want to demonstrate first that problem (1.3)–(1.5) is well posed. Thus we can construct a strongly continuous semigroup Sσ (t) on an appropriate phase-space. This semigroup possesses a bounded absorbing set which is compact in the phase-space if σ = 0, otherwise we show the existence of a compact exponentially attracting set which entails the asymptotic compactness of Sσ (t). The latter result is based on a recent decomposition of the solution semigroup devised in [41] (see also [23]). Therefore, for any σ  0, we deduce that Sσ (t) possesses a (smooth) global attractor. Taking advantage of these results, we can also deduce that any trajectory originating from the phase-space is precompact. Then, we can proceed to analyze the asymptotic behavior of a single trajectory. More precisely, we show that if φ is real analytic, then any (weak) solution (ϑ(t), σ q(t), χ(t)) converges, as t goes to ∞, to a single equilibrium, namely, to a triplet (ϑ∞ , 0, χ∞ ), where ϑ∞ and χ∞ satisfy ⎧ −1 ⎪ ϑ = |Ω| (ϑ0 − εχ1 ), ⎪ ∞ ⎪ ⎪ ⎪ ⎪ ⎪ Ω ⎪ ⎪ ⎨ χ∞ = (εχ1 + χ0 ), ⎪ ⎪ ⎪ Ω Ω ⎪ ⎪   ⎪ ⎪ − −χ∞ + φ(χ∞ ) = 0, in Ω, ⎪ ⎪ ⎩ ∇χ∞ · n = ∇(χ∞ ) · n = 0, on ∂Ω.

(1.6)

This result is obtained by exploiting a well-known technique originated from some works of S. Łojasiewicz [37,38] and then refined by L. Simon [48]. We recall that, in more than one spatial dimension, the structure of the set of solutions to (1.6) may contain a continuum of solutions if Ω is a ball or an annulus (cf., e.g., [31] and references therein). If this is the case, it is nontrivial to decide whether or not a given trajectory converges to a single stationary state. Moreover, this might not happen even for finite-dimensional dynamical systems (cf. [5]) and there are negative results for semilinear parabolic equations with smooth nonlinearities (see [42,43]). During the last years, the Łojasiewicz–Simon technique has been modified and used by many authors (cf., e.g., [9,10,12,18,32–36,53]) to investigate a number of parabolic and hyperbolic semilinear equations with variational structure. More recently, this technique has also been used for problems with only a partial variational structure, like the phase-field systems. More precisely, nonconserved models (with or without memory effects) have been analyzed in [1,2,17, 24,54], while the case of a hyperbolic dynamics for the order parameter has been examined in [25,51]. There are also results for nonlocal models (see [16,26]). Concerning the standard Cahn–Hilliard equation, convergence to stationary states has been examined in [11,19,44,46,52], while the nonconstant temperature case, namely (1.1) with (1.5), has been first analyzed in [15] and then in [45] in the case of dynamic boundary conditions. The memory effects in the heat flux have been treated in [3,4] for the Coleman–Gurtin law and, recently, in [39] for a generalization of the Maxwell–Cattaneo law. As we shall see, here we need a particular Łojasiewicz–Simon type inequality which is a refinement of the one proved in [19] (see Lemma 4.1 and its proof in Appendix A).

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

41

2. Well-posedness and uniform bounds Let H = L2 (Ω) and H = (L2 (Ω))3 . These spaces are endowed with the natural inner product ·,· and the induced norm · . For the sake of simplicity, we will assume |Ω| = 1 and ε = 1. Then, we set V = H 1 (Ω), V = (H 1 (Ω))3 and W = H 2 (Ω), both endowed with their standard inner products, and we define the subspace of H of the null mean functions

H0 = v ∈ H : v, 1 = 0 . We also introduce the linear nonnegative operator A = − : D(A) ⊂ H → H0 with domain D(A) = {v ∈ W : ∇v · n = 0, on ∂Ω}, and denote by A0 its restriction to H0 . Note that A0 is a positive linear operator; hence, for any r/2 r ∈ R, we can define its powers Ar and, consequently, set V0r = D(A0 ) endowed with the inner product r/2 r/2 v1 , v2 V0r = A0 v1 , A0 v2 . Clearly, we have V00 ≡ H0 . In addition, we need to use the Hilbert spaces V0 = {v ∈ V: v · n = 0, on ∂Ω}, and Hσ = H × H × V × V ∗ ,

Vσ = V × V0 × D(A) × H,

endowed with the following norms, respectively,  2  2  2  2 = z1  + σ z2  + z3 V + z4 V ∗ ,          1 2 3 4 2  z , z , z , z  = z1 2 + σ z2 2 + z3 2 + z4 2 , V V W Vσ

 1 2 3 4 2  z ,z ,z ,z 



if σ > 0. Otherwise, we simply set H0 = H × V × V ∗ ,

V0 = V × D(A) × H.

Our assumptions on the function φ and on the potential Φ, defined by y Φ(y) =

φ(ξ ) dξ, 0

are the following:

∀y ∈ R,

42

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

Φ ∈ C 3 (R) such that Φ(y)  −c0 , 

   φ (y)  c1 1 + |y| , ∀y ∈ R;

∀y ∈ R;

(2.1) (2.2)

∀ > 0, there exists c > 0 such that   φ(y)  Φ(y) + c , ∀y ∈ R; ∀ς ∈ R,

there exist c2 > 0 and c3  0

(y − ς)φ(y)  c2 Φ(y) − c3 ,

φ (y)  −c4 ,

(2.3) such that

∀y ∈ R;

(2.4)

∀y ∈ R,

(2.5)

for some positive constants c0 , c1 , c4 . Here c2 and c3 continuously depend on ς . We now rewrite system (1.3) together with (1.5) in the following form ⎧

in (0, ∞), ⎪ ⎨ (ϑ + χ)t , v − q, ∇v = 0, σ qt + q, v = ϑ, ∇ · v, in (0, ∞), ⎪

⎩ χtt + χt , w + Aχ + φ(χ) + αχt − ϑ, Aw = 0, in (0, ∞),

(2.6)

for all v ∈ V , v ∈ V0 , and w ∈ D(A), endowed with initial conditions (1.4). Let us prove Theorem 2.1. Let (2.1)–(2.5) hold. Then, for any (ϑ0 , q0 , χ0 , χ1 ) such that ϑ0 ∈ H,

(2.7)

σ q0 ∈ H,

(2.8)

χ0 ∈ V ,

(2.9)



(2.10)

χ1 ∈ V ,

the Cauchy problem (2.6)–(1.4) has a (weak) solution (θ, χ) with the following properties:   ϑ ∈ C 0 [0, ∞), H ,   σ q ∈ C 0 [0, ∞); H , q ∈ L2 (0, ∞; H),   χ ∈ C 0 [0, ∞), V ,   χt ∈ C 0 [0, ∞), V ∗ ∩ L2 (0, ∞, V ∗ ), αχt ∈ L2 (0, ∞, H ),

(2.11) (2.12) (2.13) (2.14) (2.15)

and there exists a positive constant C, depending on the norms of the initial data and on φ, such that, for all t  0,    ϑ(t), q(t), χ(t), χt (t) 2 H

σ

∞       

  ϑ(τ ) − ϑ(τ ), 1 2 + q(τ )2 + χt (τ )2 ∗ + α χt (τ )2 dτ  C, + V t

(2.16)

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

43

and

(ϑ + χ)(t), 1 = ϑ0 + χ0 , 1,

χ(t), 1 = χ0 + χ1 , 1 − χ1 , e−t .



(2.17)

If α > 0, then the solution is unique and the following bound holds t+1 2  sup Aχ(τ ) dτ  C. t0

(2.18)

t

Moreover, for any fixed T > 0, if (ϑ0i , q0i , χ0i , χ1i ) ∈ Hσ , i = 1, 2, then the corresponding solutions (ϑ i , qi , χ i , χti ) satisfy  1          ϑ − ϑ 2 (t), q1 − q2 (t), χ 1 − χ 2 (t), χ 1 − χ 2 (t) 2 t Hσ 2  KT  (ϑ01 − ϑ02 , q01 − q02 , χ01 − χ02 , χ11 − χ12 )H ,  C(R)e σ

∀t ∈ [0, T ], (2.19)

for some positive constants C(R) and K, both independent of T , where   (ϑ0i , q0i , χ0i , χ1i )



 R,

i = 1, 2.

Proof. We first show inequality (2.16) arguing formally. This argument can be made rigorous within a Faedo–Galerkin scheme and it suffices to prove the existence of a solution for all α  0. From now on C will denote a generic positive constant which depends on φ and on the spatial averages of ϑ0 + χ0 and χ0 + χ1 (cf. (2.17), see also (2.37) below), at most. If a solution exists, then it is easy to show the validity of (2.17), due to the boundary conditions (1.5). Moreover, we have

χt (t), 1 = χ1 , 1e−t .

(2.20)

Let us set now ϑ˜ = ϑ − ϑ, 1,

χ˜ = χ − χ, 1,

(2.21)

and rewrite problem (2.6) in the form

⎧ ˜ in (0, ∞), ⎪ ⎨ (ϑ + χ˜ )t , v − q, ∇v = 0, ˜ ∇ · v, in (0, ∞), σ qt + q, v = ϑ, ⎪

⎩ ˜ Aw = 0, in (0, ∞), χ˜ tt + χ˜ t , w + Aχ˜ + φ(χ) + α χ˜ t − ϑ,

(2.22)

for all v ∈ V , v ∈ V0 , and w ∈ D(A). Let us take v = ϑ˜ in the first equation, v = q in the second equation, and w = A−1 0 (χ˜ t + β χ˜ ), where β > 0 will be chosen small enough. Adding together the resulting identities, we get

44

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

  −1/2 d −1/2 ˜ 2 + σ q 2 + A−1/2 χ˜ t 2 + ∇ χ ϑ ˜ 2 + 2β A0 χ˜ t , A0 χ˜ 0 dt  −1/2 2

 + β A0 χ˜  + αβ χ˜ 2 + 2 Φ(χ), 1  −1/2 2

+ 2 q 2 + 2(1 − β)A0 χ˜ t  + 2α χ˜ t 2 − 2 φ(χ), χt (t), 1

˜ χ˜  = 0. + 2β ∇ χ ˜ 2 + 2β φ(χ), χ˜ − 2βϑ,

(2.23)

Observe that, using (2.4) with ς = χ, 1, we deduce



φ(χ), χ˜  C1 Φ(χ), 1 − C2 ,

(2.24)

for some C1 > 0, while, on account of (2.3), we infer





βC1 Φ(χ), 1 − e−t . − φ(χ), χt , 1 = − φ(χ), 1 χt , 1  − 2 2

(2.25)

Hence, using (2.1), we have



−2 φ(χ), χt , 1 + 2β φ(χ), χ˜

   βC1 Φ(χ), 1 − 2βC2 − cβ e−t  −C β + cβ e−t .

(2.26)

Then, taking (2.17) and (2.20) into account, from (2.23) we deduce   −1/2 d −1/2 ˜ 2 + σ q 2 + A−1/2 χ˜ t 2 + ∇ χ ϑ ˜ 2 + 2β A0 χ˜ t , A0 χ˜ 0 dt  −1/2 2

 + β A0 χ˜  + αβ χ˜ 2 + 2 Φ(χ), 1  −1/2 2 ˜ χ˜  + 2 q 2 + 2(1 − β)A0 χ˜ t  + 2α χ˜ t 2 + 2β ∇ χ ˜ 2 − 2βϑ,    C β + cβ e−t .

(2.27)

Let us now test the third equation of (2.22) with χ˜ . We obtain  d 2χ˜ t , χ˜  + χ˜ 2 + α ∇ χ˜ 2 dt

˜ Aχ˜  = 0. − 2 χ˜ t 2 + 2 Aχ ˜ 2 + 2 φ (χ)∇ χ˜ , ∇ χ˜ − 2ϑ,

(2.28)

Moreover, in the case σ > 0, using the first two equations of (2.22), we have



d q, ∇A−1 ϑ˜ = qt , ∇A−1 ϑ˜ + q, ∇A−1 ϑ˜ t 0 0 0 dt 2



 −1/2 −1 ˜ 2  ˜ = −σ −1 q, ∇A−1 ϑ − q, ∇A−1 ∇ · q . 0 ϑ −σ 0 χ˜ t + A0

(2.29)

Let us discuss first the case σ > 0. Then, multiply (2.28) by γ1 and (2.29) by γ2 , γ1 > 0 and γ2 > 0 to be chosen later, and sum both the obtained expressions to (2.27). Note also that, by the Poincaré inequality and (2.5), for some κ1 > 0 depending only on Ω, we have

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

45



˜ χ˜  + 2γ1 φ (χ)∇ χ˜ , ∇ χ˜ − 2γ1 ϑ, ˜ Aχ˜  −2βϑ, ˜ 2.  −(β + 2γ1 c4 ) ∇ χ˜ 2 − γ1 Aχ˜ 2 − (βκ1 + γ1 ) ϑ

(2.30)

Additionally, for some κ2 > 0 depending also only on Ω, we get  

σ −1  

˜ 2 + A−1/2 χ˜ t 2 + κ2 1 + σ −1 q 2 . ˜ − q, ∇A−1 χ˜ t  ϑ −σ −1 q, ∇A−1 ϑ 0 0 0 2

(2.31)

Then, let us introduce the functional   ˜ q, χ, ˜ 2 + σ q 2 + A−1/2 χ˜ t 2 + ∇ χ ˜ χ˜ t ) = ϑ ˜ 2 Ψσ (ϑ, 0  −1/2 2 −1/2

−1/2 + 2β A0 χ˜ t , A0 χ˜ + β A0 χ˜  + αβ χ ˜ 2 + 2 Φ(χ), 1

  (2.32) ˜ 2 + α ∇ χ ˜ 2 + γ2 q, ∇A−1 ϑ˜ , + γ1 2χ˜ t , χ˜  + χ 0

and, recalling (2.30) and (2.31), let us choose, in turn, γ2 so small that  

max 2γ2 , γ2 κ2 1 + σ −1  1, and then β and γ1 so small that β  1/2, γ1 c4  β/4, and (βκ1 + γ1 )  γ2 σ −1 /4. Then, Ψσ fulfills the inequality     d ˜ q, χ˜ , χ˜ t ) + c σ −1 γ2 ϑ ˜ 2 + q 2 + A−1/2 χ˜ t 2 + α χ˜ t 2 + β ∇ χ Ψσ (ϑ, ˜ 2 + γ1 Aχ ˜ 2 0 dt   (2.33)  C β + cβ e−t . Moreover, possibly choosing a smaller γ2 (and consequently smaller β and γ1 ), we find   ˜ Ψσ ϑ(t), q(t), χ˜ (t), χ˜ t (t) 2 2  2   2    Cβ ϑ(t) + σ q(t) + χ(t)V + χt (t)V ∗ − C,

∀t  0,

(2.34)

where we stress once more that the constants C are allowed to depend on the spatial averages of the initial data specified in (2.17). On the other hand, on account of (2.1), (2.2) and (2.7)–(2.10), and recalling notation (2.21), we find R0 > 0 such that Ψσ (ϑ˜ 0 , q0 , χ˜ 0 , χ˜ 1 )  R0 . Using then [22, Lemma 2.1], we deduce that there exists t0 = t0 (R0 ) > 0 such that, for all t  t0 ,   ˜ Ψσ ϑ(t), q(t), χ˜ (t), χ˜ t (t)  R, where R is independent of R0 . Thus, recalling (2.34), we deduce that    ϑ(t), q(t), χ(t), χt (t) 2



 C(R0 ),

(2.35)

46

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

for all t ∈ [0, ∞). On account of (2.20) and (2.35), taking β = 0 in (2.23), integrating from t to T and letting T go to ∞ we also get the integral control ∞        q(τ )2 + χt (τ )2 ∗ + α χt (τ )2 dτ  C(R0 ). V

(2.36)

t

Then, using (2.35) and (2.36), from (2.29) we deduce ∞   ϑ(τ ˜ )2 dτ  C(R0 ), t

so that (2.16) is proved. In addition, integrating (2.33) from t to t + 1, we then find (2.18). The case σ = 0 is simpler. We can take the functional   −1/2 −1/2 ˜ χ˜ , χ˜ t ) = ϑ ˜ 2 + A−1/2 χ˜ t 2 + ∇ χ ˜ 2 + 2β A0 χ˜ t , A0 χ˜ Ψ0 (ϑ, 0  −1/2 2  

+ β A0 χ˜  + αβ χ˜ 2 + 2 Φ(χ), 1 + γ1 2χ˜ t , χ˜  + χ ˜ 2 + α ∇ χ ˜ 2 , and observe that     d ˜ χ˜ , χ˜ t ) + c ∇ ϑ ˜ 2 + A−1/2 χ˜ t 2 + α χ˜ t 2 + β ∇ χ Ψ0 (ϑ, ˜ 2 + γ1 Aχ˜ 2 0 dt    C β + cβ e−t . Then we can argue as above. Estimate (2.19) is standard, provided that α > 0. Indeed, it suffices to write down problem (2.6) for the difference of two solutions (ϑi , qi , χi ), i = 1, 2, and then multiply the first equation by ϑ1 − ϑ2 , the second one by q1 − q2 , and the third one by A−1 0 (χ˜ 1 − χ˜ 2 )t . Using the Gronwall Lemma and taking (2.2) into account, one easily gets the wanted estimate (see, e.g., [6] or [22] for the isothermal case). 2 From Theorem 2.1 and its proof we deduce that, letting Xσδ =



 1 2 3 4 

 

 z , z , z , z ∈ Hσ :  z1 + z3 , 1  +  z3 + z4 , 1   δ

(2.37)

for some δ  0, endowed with the metric induced by the norm of Hσ , and setting   ϑ(t), q(t), χ(t), χt (t) =: Sσ (t)(ϑ0 , q0 , χ0 , χ1 ),

∀t  0,

and for (ϑ0 , q0 , χ0 , χ1 ) ∈ Hσ , we have that Sσ (t) is a strongly continuous semigroup on Hσ . Moreover, its restriction Sσδ (t) to Xσδ admits a bounded absorbing set. Similarly, we can define a strongly continuous dissipative semigroup S0δ (t) on X0δ . Summing up, we have Corollary 2.2. Let (2.1)–(2.5) hold. For any given σ ∈ [0, 1] and δ  0, the semigroup Sσδ (t) acting on Xσδ has a bounded absorbing set.

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

47

3. Precompactness of trajectories and global attractor Here we prove Theorem 3.1. Let (2.1)–(2.5) hold and suppose α > 0. If σ ∈ (0, 1] and (ϑ0 , q0 , χ0 , χ1 ) satisfies (2.7)–(2.10), then, indicating by (ϑ, q, χ) the corresponding solution to (2.6)–(1.4) given by  Theorem 2.1, the orbit t0 (ϑ(t), q(t), χ(t), χt (t)) is precompact in Hσ . Moreover, there holds   ϑ(t) − ϑ0 − χ1 , 1 → 0,   q(t) → 0,   χt (t) ∗ → 0, V

(3.1) (3.2) (3.3)

as t goes to ∞, and the ω-limit set ω(ϑ0 , q0 , χ0 , χ1 ) consists only of equilibrium points of the form (ϑ∞ , 0, χ∞ , 0) where (ϑ∞ , χ∞ ) satisfies (1.6). Similar results hold when σ = 0. Proof. On account of [41], observe first that, thanks to (2.2), (2.5), and (2.16), we can choose  > c4 large enough, and depending on the norms of the initial data, such that   1 ∇z 2 + ( − 2c4 ) z 2 − φ χ(t) z, z  0, 2

(3.4)

for all z ∈ V and every t  0. Consequently, we set ψ(r) = φ(r) + r,

∀r ∈ R.

Then, we split the solution to (2.6) in this way     (ϑ, q, χ) = ϑ d , qd , χ d + ϑ c , qc , χ c , where ⎧  d 

ϑ + χ d t , v − qd , ∇v = 0, ⎪ ⎪ ⎪ ⎪ ⎨ σ qd + qd , v = ϑ , ∇ · v, d t

d  

d d ⎪ χtt + χt , w + Aχ + ψ(χ) − ψ χ c + αχtd − ϑ d , Aw = 0, ⎪ ⎪ ⎪ ⎩ d ϑ (0) = ϑ˜ 0 , σ qd (0) = σ q0 , χ d (0) = χ˜ 0 , χtd (0) = χ˜ 1 ,

in (0, ∞), in (0, ∞), in (0, ∞),

(3.5)

in Ω,

and 

⎧  c ϑ + χ c t , v − qc , ∇v = 0, ⎪ ⎪ ⎪ c

⎨ σ q + qc , v = ϑc , ∇ · v, ct

 

⎪ χtt + χtc , w + Aχ c + ψ χ c + αχtc − ϑ c , Aw = χ, Aw, ⎪ ⎪ ⎩ c ϑ (0) = ϑ0 , 1, σ qc (0) = 0, χ c (0) = χ0 , 1, χtc (0) = χ1 , 1, for all v ∈ V , v ∈ V0 , and w ∈ D(A).

in (0, ∞), in (0, ∞), in (0, ∞), in Ω,

(3.6)

48

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

We shall prove that (ϑ d (t), qd (t), χ d (t), χtd (t)) exponentially decays at 0 in Hσ as t goes to ∞, while (ϑ c , qc , χ c , χtc ) is bounded in a space which is compactly embedded in Hσ , uniformly in time. Let us prove first that, for any t  s  0 and every  > 0, there holds t α

 c 2 χ (τ ) dτ   (t − s) + C . t 

(3.7)

s

This estimate combined with (2.16) will allow us to use a suitable version of the Gronwall Lemma. c Let us take v = ϑ˜ c in the first equation of (3.6), v = qc in the second equation, and w = A−1 0 χ˜ t in the third one. Then we obtain       

   d  ϑ˜ c 2 + σ qc 2 + A−1/2 χ˜ c 2 + ∇ χ˜ c 2 + 2 Ψ χ c , 1 − 2 χ, χ˜ c t 0 dt  2  −1/2   2 + 2qc  + 2A0 χ˜ tc  + 2α χ˜ tc 

 

= 2 ψ χ c , χt , 1 − 2 χt , χ˜ c .

(3.8)

Here Ψ is a primitive of ψ . Observe first that it is not difficult to realize that an estimate similar to (2.16) holds for (ϑ c , qc , χ c , χtc ) as well. Therefore, on account of (2.2) and (2.20), we have, for any  > 0 and any t  0,   



 C  χt (t)2 +  e−t . 2 ψ χ c (t) , χt (t), 1 − 2 χt (t), χ˜ c (t)  2 +  Therefore, (3.7) follows from integrating (3.8) with respect to time from s to t, using the above inequality and (2.16), recalling that α > 0, and observing that (cf. (2.20)) c

χt (t), 1 = χ1 , 1e−t . In order to prove the exponential decay of (ϑ d , qd , χ d , χtd ), we first note that (cf. (2.17))



ϑ d (t), 1 = χ d (t), 1 = 0,

∀t  0,

(3.9)

so that ϑ d = ϑ˜ d and χ d = χ˜ d . We now argue as to get (2.32), namely, we take v = ϑ d in the first equation, v = qd in the d d second equation, and w = A−1 0 (χt + βχ ), with some β > 0 which will be chosen later on. d −1 Then, we add the functional γ q , ∇A ϑ d  with γ > 0. Thus, recalling (2.29), we obtain       

d  ϑ d 2 + σ qd 2 + A−1/2 χ d 2 + ∇χ d 2 + 2β A−1/2 χ d , A−1/2 χ d t t 0 0 0 dt  2  −1/2 2

 

+ β A0 χ d  + αβ χ d  + γ qd , ∇A−1 ϑ d + 2 ψ(χ) − ψ χ c , χ d  2  −1/2 2  2 γ



− ψ (χ)χ d , χ d + 2qd  + 2(1 − β)A0 χtd  + 2α χtd  + qd , ∇A−1 ϑd 0 σ  −1/2  d 2 2 2

d γ −1 d d d + ϑ  + γ q , ∇A0 χt − γ A0 ∇ · q  + 2β ∇χ  σ

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60



  d + 2β ψ(χ) − ψ χ c , χ d − 2β ϑ d , A−1 0 χ  

  2 = 2 ψ (χ) − ψ χ c χtc , χ d − ψ

(χ)χt , χ d .

49

(3.10)

Observe that, owing to (2.1), (2.16), and (2.20), we have   

 2 2 ψ (χ) − ψ χ c χtc , χ d − ψ

(χ)χt , χ d   2   C χt + χtc  ∇χ d    2  2 2   β ∇χ d  + Cβ χt 2 + χtc  ∇χ d  .

(3.11)

On the other hand, setting  2  2  −1/2 2  2

−1/2 −1/2 Λd = ϑ d  + σ qd  + A0 χtd  + ∇χ d  + 2β A0 χtd , A0 χ d  −1/2 2  2

+ β A0 χ d  + αβ χ d  + γ qd , ∇A−1 ϑ d  



+ 2 ψ(χ) − ψ χ c , χ d − ψ (χ)χ d , χ d , and observing that (cf. (2.5) and (3.4))  2 2  



1 2 ψ(χ) − ψ χ c , χ d − ψ (χ)χ d , χ d  ( − 2c4 )χ d  − φ (χ)χ d , χ d  − ∇χ d  , 2 we have that, for β and γ small enough,     1  ϑ d , q d , χ d , χ d 2  Λ d  C  ϑ d , q d , χ d , χ d 2 . t t H Hσ σ 4

(3.12)

Moreover, possibly choosing β and γ smaller than before, and using (3.11), from (3.10) we infer  2   d Λd + cβ,γ Λd  Cβ,γ χt 2 + χtc  Λd . dt Thus, on account of (2.16) and (3.7), we can apply the Gronwall Lemma reported, e.g., in [13, Lemma 2.1] and deduce the exponential decay of Λd , so that (cf. (2.20) and (3.12))  d   ϑ (t), qd (t), χ d (t), χ d (t)  t



 C(R)e−ct ,

provided that (ϑ0 , q0 , χ0 , χ1 ) Hσ  R. Moreover, taking w = χ d in the third equation of (3.5), we obtain (cf. (2.28))  2 2  d  d d  2 χt , χ + χ d  + α ∇χ d  dt  2  2



  − 2χtd  + 2Aχ d  + 2 ψ(χ) − ψ χ c , Aχ d − 2 ϑ d , Aχ d = 0,

(3.13)

50

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

which yields, using the Young inequality, (2.23), and (3.13),   2  2 2   2  d  d d  2 χt , χ + χ d  + α ∇χ d  + Aχ d   C 1 + χtd  . dt On account of (2.36) and (3.7), we have t+1 2  sup χtd (τ ) dτ  C, t0

t

so that there holds the additional bound t+1 2  sup Aχ d (τ ) dτ  C. t0

(3.14)

t

We now consider (3.6). Taking v = Aϑ c in the first equation, v = −∇∇ · qc in the second one (this procedure is just formal at this stage, but could be made rigorous through an approximation), and adding together the resulting identities, we obtain      

d  ∇ϑ c 2 + σ ∇ · qc 2 + 2 χ c , Aϑ c + 2∇ · qc 2 = 0. t dt

(3.15)

We also have (cf. (2.29))





d c q , ∇ϑ c = qct , ∇ϑ c + qc , ∇ϑtc dt  2 2

 = −σ −1 qc , ∇ϑ c − σ −1 ∇ϑ c  − qc , ∇χtc + ∇ · qc  .

(3.16)

Let us now take w = χtc + βχ c in the third equation. We find       

   d  χ c 2 + Aχ c 2 + 2β χ c , χ c + β χ c 2 + αβ ∇χ c 2 + 2 ψ χ c , Aχ c t t dt   2 2  2



  + 2(1 − β)χtc  + 2α ∇χtc  + 2β Aχ c  + 2β ψ χ c , Aχ c − 2 ϑ c , Aχtc

 

  − 2β ϑ c , Aχ c − 2 ψ χ c χtc , Aχ c = ∇χ, ∇ χtc + βχ c . (3.17) Observe that, on account of (2.16),  2    

   ψ χ c χtc , Aχ c  C 1 + χ c L6 (Ω) χtc L6 (Ω) Aχ c       C χtc V Aχ c .



Therefore, setting   2 2  2  2

Λc = ∇ϑ c  + σ ∇ · qc  + χtc  + Aχ c  + 2β χtc , χ c  2  2  



+ β χ c  + αβ ∇χ c  + 2 ψ χ c , Aχ c + γ qc , ∇ϑ c ,

(3.18)

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

51

for some γ > 0, using the Young inequality, we can choose β and γ small enough so that  2   d Λc + cβ,γ Λc  Cβ,γ 1 + Aχ c  . dt Then, on account of (2.18) and (3.14), we obtain the uniform boundedness of Λc which implies        c 2 ∇ϑ (t) + σ ∇ · qc (t)2 + χ c (t)2 + Aχ c (t)2  C, t

∀t  0.

(3.19)

The second equation of (3.6) can now be written in the strong form, namely, σ qct + qc = −∇ϑc ,

a.e. in Ω × (0, ∞),

so that   σ ∇ × qc t + ∇ × qc = 0,

a.e. in Ω × (0, ∞),

and, since (∇ × qc )(0) = 0, we have (∇ × qc )(t) = 0 for any t  0. Consequently, thanks to (3.19), qc (t) V is uniformly bounded as well. Summing up, we have shown that a given trajectory originating from Hσ is a sum of an exponentially decaying part and a term which belongs to a closed bounded subset of Vσ . Therefore the trajectory is precompact in Hσ and, due to the integral controls of (2.16) and to (2.17), we infer (3.1)–(3.3). Finally, it is not difficult to prove that

ω(ϑ0 , q0 , χ0 , χ1 ) ⊆ (ϑ∞ , 0, χ∞ , 0): (ϑ∞ , χ∞ ) satisfies (1.6) . The case σ = 0 is easier. In fact, arguing as in the isothermal case (see [6]), we can prove the bound    ϑ(t), χ(t), χt (t) 2  C, V 0

∀t  t1 = t1 (R) > 0,

provided that (ϑ0 , χ0 , χ1 ) H0  R. Hence the trajectory is precompact in H0 and we can conclude as above. 2 From the proof of Theorem 3.1, we deduce that the semigroup Sσδ (t) has a bounded attracting set in Vσ , for any σ ∈ (0, 1], while S0δ (t) has a compact absorbing set. Therefore we have (see, e.g., [30,49]) Corollary 3.2. For each σ ∈ [0, 1] and δ  0, the semigroup Sσδ (t) has a connected global attractor Aδσ which is bounded in Vσ . Remark 3.3. The above result is a first, but essential, step toward the construction of a family of exponential attractors which is stable (robust) with respect to σ and, possibly, to ε (see [22] for the isothermal case). This might be the subject of a future investigation.

52

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

4. Convergence to stationary states Let us set

1 ˜ E(v) = ∇v 2 + Φ(v), 1 2 for any v ∈ V01 , where ˜ Φ(y) =

y

˜ ) dξ, φ(ξ

∀y ∈ R,

0

and   ˜ φ(y) = φ y + χ0 + χ1 , 1 ,

∀y ∈ R.

(4.1)

The version of the Łojasiewicz–Simon inequality we need is the following (see Appendix A) Lemma 4.1. Suppose that φ is real analytic and assume (2.2) and (2.5). Let v∞ ∈ V02 be such that   ˜ ∞ ) = 0. A A0 v∞ + φ(v

(4.2)

Then there exist ρ ∈ (0, 12 ), η > 0, and a positive constant L such that   

 E(v) − E(v∞ )1−ρ  LA0 v + φ(v) ˜ ˜ − φ(v), 1 V −1 ,

(4.3)

0

for all v ∈ V01 such that v − v∞ V 1  η. 0

Then we prove Theorem 4.2. Let the assumptions of Lemma 4.1 hold and let α > 0 and σ > 0 be fixed. If (ϑ0 , q0 , χ0 , χ1 ) satisfies (2.7)–(2.10), then the trajectory (ϑ(t), q(t), χ(t), χt (t)) originated from (ϑ0 , q0 , χ0 , χ1 ) is such that

ω(ϑ0 , q0 , χ0 , χ1 ) = (ϑ∞ , 0, χ∞ , 0) ,

(4.4)

⎧ −1 ⎪ ⎪ ϑ∞ = |Ω| (ϑ0 − χ1 ), ⎪ ⎪ ⎪ ⎪ ⎪ Ω ⎨ χ∞ = (χ1 + χ0 ), ⎪ ⎪ ⎪ ⎪ ⎪ Ω Ω ⎪ ⎪  ⎩  A Aχ∞ + φ(χ∞ ) = 0.

(4.5)

where (ϑ∞ , χ∞ ) satisfies

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

53

Moreover,   lim χ(t) − χ∞ V = 0,

(4.6)

t→∞

and there exist t ∗ > 0 and a positive constant C such that   ϑ(t) − ϑ∞ 

V∗

ρ   + χ(t) − χ∞ V ∗  Ct − 1−2ρ ,

∀t  t ∗ .

(4.7)

If σ = 0 a similar result hold. Proof. Let us set σ = 1 for simplicity. On account of Theorem 3.1, we consider (ϑ∞ , 0, χ∞ , 0) ∈ ω(ϑ0 , q0 , χ0 , χ1 ), and we observe first that (3.1)–(3.3) hold and (ϑ∞ , χ∞ ) fulfills (4.5). On account of (2.17), we can rewrite (2.22) in the form

⎧ ˜ in (0, ∞), ⎪ ⎨ (ϑ + χ˜ )t , v − q, ∇v = 0, ˜ σ qt + q, v = ϑ, ∇ · v, in (0, ∞), ⎪



⎩ ˜ ˜ χ˜ tt + χ˜ t , w + A0 χ˜ + φ(χ˜ ) + α χ˜ t − ϑ, Aw = h(χ˜ ), Aw , in (0, ∞), for all v ∈ V , v ∈ V0 , and w ∈ D(A). Here we have set (cf. (4.1)) 

 ˜ χ˜ ) − φ˜ χ˜ − χ1 , e−t , ∀r ∈ R, ∀t  0. h(χ˜ ) = φ(

(4.8)

(4.9)

Arguing as in the proof of Theorem 2.1 (cf. (2.23)), we find 2   2  2   

d  ˜ L ϑ(t), q(t), χ˜ (t), χ˜ t (t) = −q(t) − χ˜ t (t)V ∗ − α χ˜ t (t) + h χ˜ (t) , χ˜ t (t) , (4.10) dt where   ˜ L ϑ(t), q(t), χ˜ (t), χ˜ t (t)           1  ˜ 2 + q(t)2 + ∇ χ˜ (t)2 + 2 Φ˜ χ˜ (t) , 1 + χ˜ t (t)2 . = ϑ(t) V∗ 2

(4.11)

Note that, due to (2.1), (2.2), (2.16) and (4.9), there holds 2  

α h χ˜ (t) , χ˜ t (t)  Cα e−2t + χ˜ t (t) , 2 using also the Young inequality. Therefore, from (4.10) we deduce  d  ˜ L ϑ(t), q(t), χ˜ (t), χ˜ t (t) dt 2   2 2 α  −q(t) − χ˜ t (t)V ∗ − χ˜ t (t) + Cα e−2t , 2 for all t  0.

(4.12)

54

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

Then, combining (2.29) with (4.10), we obtain

 d ˜ 2 L + μ q, ∇A−1 ϑ˜ + q 2 + χ˜ t 2V ∗ + α χ˜ t 2 + μ ϑ 0 dt  −1/2 2



−1  ˜ + μ q, ∇A−1 ∇ · q = h(χ˜ ), χ˜ t , 0 ϑ + μ q, ∇A0 χ˜ t − μ A0

(4.13)

for some μ > 0 to be chosen below. Following a well-known strategy (see, e.g., [12,36]) we consider the functional  −1  ˜ ˜ G(t) = A−1 0 χ˜ t , A0 A0 χ˜ + φ(χ˜ ) − φ(χ˜ ) ,

t  0,

where

˜ χ˜ ) = φ( ˜ χ˜ ), 1 , φ( and we observe that     d ˜ χ˜ ) − φ( ˜ χ˜ ) + A−1 χ˜ t , A−1 A0 χ˜ t + φ˜ (χ˜ )χ˜ t − φ˜ (χ˜ )χ˜ t G = A−1 A0 χ˜ + φ( χ˜ tt , A−1 0 0 0 0 dt   −1/2   −1   ˜ ˜ χ˜ ) 2 ˜ ˜ χ˜ ) − φ( = − A−1 A0 χ˜ + φ( 0 χ˜ t , A0 A0 χ˜ + φ(χ˜ ) − φ(χ˜ ) − A0     ˜ ˜ ˜ ˜ ˜ −1 − α χ˜ t , A−1 0 A0 χ˜ + φ(χ˜ ) − φ(χ˜ ) + ϑ, A0 A0 χ˜ + φ(χ˜ ) − φ(χ˜ )    −1/2 2  ˜ ˜ χ˜ t  + h(χ˜ ), A−1 0 A0 χ˜ + φ(χ˜ ) − φ(χ˜ ) + A0  −1  ˜ ˜ (4.14) + A−1 0 χ˜ t , A0 φ (χ˜ )χ˜ t − φ (χ)χ˜ t . Observe that (cf. (2.2))  −1/2 2  −1  ˜  ˜ A−1 χ˜ t  . 0 χ˜ t , A0 φ (χ˜ )χ˜ t − φ (χ˜ )χ˜ t  C A0



(4.15)

Then, from (4.13) and (4.14), using the Young inequality, we find (cf. also (4.9), (4.11), and (4.15)) d M + Cμ,ν N 2  0, dt

(4.16)

for μ > 0 and ν > 0 sufficiently small, where M(t) =



1 ˜ 2 + q 2 + χ˜ t 2V ∗ + E(χ) ˜ − E(χ˜ ∞ ) + μ q, ∇A−1 ϑ ϑ˜ 0 2

+ νG + Cα,ν e−2t , 2    2       ˜ 2 + ν A0 χ˜ (t) + φ˜ χ˜ (t) − φ˜ χ˜ (t) 2 −1 , N 2 (t) = q(t) + χ˜ t (t)V ∗ + μϑ(t) V 0

for all t  0.

(4.17) (4.18)

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

55

Let us introduce the unbounded set     η Σ = t  0: χ˜ (t) − χ˜ ∞ V 1  0 3 where η is given by Lemma 4.1. Then, for every t ∈ Σ , define     τ (t) = sup t  t: sup χ˜ (s) − χ˜ ∞ V 1  η , s∈[t,t ]

0

and observe that τ (t) > t, for every t ∈ Σ. Recalling (3.1)–(3.3), let t0 ∈ Σ be large enough such that       ϑ(t) ˜  + q(t) + χ˜ t (t) ∗  1, ∀t  t0 , V

(4.19)

and set   J = t0 , τ (t0 ) ,

J1 = t ∈ J : N (t) > e−2(1−ρ)t , J2 = J \ J1 . From (4.16), we have that M is decreasing, therefore it is constant on ω(ϑ0 , q0 , χ0 , χ1 ). In addition, there holds ρ ρ−1 d   d  M(t) sgn M(t) = ρ M(t) M(t), dt dt

∀t  0,

(4.20)

so that |M|ρ sgn M is decreasing as well. Observe now that, for every t ∈ J1 , thanks to (4.3) and (4.19), we have   M(t)1−ρ  CN (t), possibly choosing μ and ν even smaller than before. Then, on account of (4.16) and (4.20), we infer

τ (t0 )

N (t) dt  −C J1

ρ  d  M(t) sgn M(t) dt dt

t0

ρ   ρ    C M(t0 ) + M τ (t0 )  ,

where we mean that |M(τ (t0 ))| = 0 if τ (t0 ) = ∞. On the other hand, we easily get N (t) dt  Ce−2(1−ρ)t0 . J2

Therefore χ˜ t (·) V ∗ is integrable over J and

(4.21)

56

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60 τ (t0 )

  χ˜ t (t)

0  lim sup

t0 ∈Σ, t0 →∞

 c lim sup

V∗

dt

t0

      M(t0 )ρ + M τ (t0 ) ρ + Ce−2(1−ρ)t0 = 0.

t0 ∈Σ, t0 →∞

(4.22)

Notice that, for every t ∈ J ,   χ(t) ˜ − χ˜ ∞ 

t

V∗



  χ˜ t (s)

V∗

  ds + χ˜ (t0 ) − χ˜ ∞ V ∗ .

(4.23)

t0

Suppose now that τ (t0 ) < ∞ for any t0 ∈ Σ . Then, by definition,     χ˜ τ (t0 ) − χ˜ ∞ 

V01

= η,

∀t0 ∈ Σ.

Consider an unbounded sequence {tn }n∈N ⊂ Σ with the property   ˜ n ) − χ˜ ∞ V 1 = 0. lim χ(t

n→∞

0

By compactness, we can find a subsequence {tnk }k∈N and an element v∞ ∈ D(A) such that (ϑ∞ , 0, v∞ , 0) ∈ ω(ϑ0 , q0 , χ0 , χ1 ), v˜∞ − χ˜ ∞ V 1 = η, and 0

    lim χ˜ τ (tnk ) − v˜∞ V 1 = 0.

k→∞

0

Then, owing to (4.22) and (4.23), we deduce the contradiction  τ (tnk )          χ˜ t (s) V ∗ ds + χ˜ (tnk ) − χ˜ ∞ V ∗ = 0.  lim sup

0 < v˜∞ − χ˜ ∞ V ∗

k→∞

tnk

Hence, τ (t0 ) = ∞ for some t0 > 0 large enough and, recalling (2.20), we can deduce that χt (·) V ∗ is indeed integrable over (t0 , ∞). This yields (4.6) by precompactness. On the other hand, on account of (3.1)–(3.3), (4.4) holds as well. Finally, arguing as in [25], we can prove that ∞

ρ

N (τ ) dτ  Ct − 1−2ρ ,

∀t  t ∗ ,

t

for some t ∗ > 0. This entails (cf. (2.20) and (4.18)) ∞   χt (τ ) t

ρ

V∗

dτ  Ct − 1−2ρ ,

∀t  t ∗ .

(4.24)

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

57

Thus we have   χ(t) − χ∞ 

V∗

ρ

 Ct − 1−2ρ ,

∀t  t ∗ .

(4.25)

Recalling now (2.20), setting v = 1 in the first equation of (2.6), and integrating from t to ∞, we obtain

ϑ(t), 1 − ϑ∞ = χ1 , 1e−t .

(4.26)

Therefore, integrating the first equation of (2.22) with respect to time from t  t ∗ to ∞, we deduce

˜ + χ˜ ∞ − χ˜ (t), v = −ϑ(t)



q(τ ), ∇v dτ,

t

so that, on account of (4.18), (4.24), and (4.25), we infer   ϑ(t) ˜ 

ρ

V∗

 Ct − 1−2ρ ,

∀t  t ∗ .

(4.27)

Therefore, rate estimate (4.7) is a consequence of (4.25)–(4.27). In the case σ = 0 we can proceed in a similar (actually, simpler) way, noting that q = −∇ϑ . 2 Remark 4.3. The decay estimate (4.7) for ϑ can be slightly improved. Actually, using the decomposition       ϑ(t) − ϑ∞ 2  2ϑ d (t)2 + 2ϑ c (t) − ϑ∞ 2 ,

(4.28)

we see that, by (3.13), the first term decays exponentially. Concerning the latter, one can use (3.19), (4.27) and the interpolation inequality v 2  c v V v V ∗ , holding for all v ∈ V . Thus, (4.28) eventually gives ρ   ϑ(t) − ϑ∞   Ct − 2−4ρ ,

∀t  t ∗ .

(4.29)

Appendix A This section is devoted to demonstrate Lemma 4.1. Let us introduce the functional   1 ˜ E(v) = |∇v|2 + Φ(v) dx, 2 Ω

defined on the space V01 . As before, we assume |Ω| = 1. The differential operator associated with the gradient ∂E does not conserve null mean functions. Hence the version of the Łojasiewicz theorem given in [32] is not applicable directly. This problem was solved in [19], but [19, Assumption 5] is not exactly satisfied here. Our proof, essentially given for the reader’s convenience, follows the lines of [11] based on a general version of the Łojasiewicz–Simon theorem obtained in [9].

58

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

Proof of Lemma 4.1. We begin to observe that v∞ satisfying (4.2) is a solution to ˜ ∞ ) = 0. ˜ ∞ ) − φ(v Av∞ + φ(v

(A.1)

Moreover, v∞ is a critical point of E on V01 . Indeed, it is easy to check that, owing to our hypotheses, E is continuously differentiable on V01 , and ∂E(v∞ )h =

  ˜ ∞ )h dx ∇v∞ · ∇h + φ(v

Ω



=

  ˜ ∞ )h dx, ˜ ∞ )h − φ(v ∇v∞ · ∇h + φ(v

(A.2)

Ω

for all h ∈ V01 . We recall that the dual space (V01 )∗ is the space of classes



[f ] = f + g; g ∈ V ∗ , g, V01 = 0 ,

f ∈ V ∗,

where ·,· stands for the duality between V ∗ and V , endowed with the norm   [f ] 1 ∗ = (V ) 0

inf

g∈V ∗ , g,V01 =0

f + g V ∗ = inf f + c V ∗ = f − f V ∗ . c∈R

Consider the mapping F : V02 → H0 , F = ∂E|V 2 defined by 0

˜ ˜ F (v) = A0 v + φ(v) − φ(v). By virtue a well-known Sobolev embedding theorem, we have that v∞ ∈ V02 ⊂ L∞ (Ω), and, due to our assumptions, we can find a neighborhood U(v∞ ) in the space V02 such that F : U(v∞ ) → H0 is analytic. Further, A0 : V01 → (V01 )∗ and A0 : V02 → H0 have compact resolvents. Observe now that, when ∂ 2 E : V01 → Lin[V01 , (V01 )∗ ], then 2

∂ E(v)[w], z =



φ˜ (v)wz dx +

Ω

∇w · ∇z dx.

(A.3)

Ω

In addition, we have   ∂ 2 E = ∂F : V02 → Lin V02 , H0 ,

˜ ˜ ∂F (v)[w] = A0 w + φ(v)w − φ(v)w.

Hence, in both cases, ∂ 2 E(v∞ ) can be viewed as a bounded perturbation of A0 restricted to the respective spaces. It follows that Ker ∂ 2 E(v∞ ) ⊂ V02 and its range is closed in (V01 )∗ and H0 , respectively. Moreover, there holds      1 ∗ V0 = Ker ∂ 2 E(v∞ ) ⊕ Ran ∂ 2 E(v∞ ) ,

    H0 = Ker ∂ 2 E(v∞ ) ⊕ Ran ∂F (v∞ ) .

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

59

Now, we can apply [9, Theorem 3.10] and [9, Corollary 3.11] to obtain     E(v) − E(v∞ )1−ρ  L∂E(v) 1 ∗ , (V ) 0

and, consequently, (4.3).

2

References [1] S. Aizicovici, E. Feireisl, Long-time stabilization of solutions to a phase-field model with memory, J. Evol. Equ. 1 (2001) 69–84. [2] S. Aizicovici, E. Feireisl, F. Issard-Roch, Long time convergence of solutions to a phase-field system, Math. Methods Appl. Sci. 24 (2001) 277–287. [3] S. Aizicovici, H. Petzeltová, Asymptotic behavior of solutions of a conserved phase-field system with memory, J. Integral Equations Appl. 15 (2003) 217–240. [4] S. Aizicovici, H. Petzeltová, Convergence of solutions of phase-field systems with a nonconstant latent heat, Dynam. Systems Appl. 14 (2005) 163–173. [5] B. Aulbach, Continuous and Discrete Dynamics Near Manifolds of Equilibria, Springer, Berlin, 1984. [6] A. Bonfoh, Existence and continuity of uniform exponential attractors for a singular perturbation of a generalized Cahn–Hilliard equation, Asymptot. Anal. 43 (2005) 233–247. [7] M. Brokate, J. Sprekels, Hysteresis and Phase Transitions, Springer, New York, 1996. [8] G. Caginalp, The dynamics of a conserved phase field system: Stefan-like, Hele-Shaw, and Cahn–Hilliard models as asymptotic limits, IMA J. Appl. Math. 44 (1990) 77–94. [9] R. Chill, On the Łojasiewicz–Simon gradient inequality, J. Funct. Anal. 201 (2003) 572–601. [10] R. Chill, E. Fašangová, Convergence to steady states of solutions of semilinear evolutionary integral equations, Calc. Var. Partial Differential Equations 22 (2005) 321–342. [11] R. Chill, E. Fašangová, J. Prüss, Convergence to steady states of solutions of the Cahn–Hilliard equation with dynamic boundary conditions, Math. Nachr. 279 (2006) 1448–1462. [12] R. Chill, M.A. Jendoubi, Convergence to steady states in asymptotically autonomous semilinear evolution equations, Nonlinear Anal. 53 (2003) 1017–1039. [13] M. Conti, V. Pata, Weakly dissipative semilinear equations of viscoelasticity, Commun. Pure Appl. Anal. 4 (2005) 705–720. [14] A. Debussche, A singular perturbation of the Cahn–Hilliard equation, Asymptot. Anal. 4 (1991) 161–185. [15] E. Feireisl, F. Issard-Roch, H. Petzeltová, Long-time behaviour and convergence towards equilibria for a conserved phase field model, Discrete Contin. Dyn. Syst. 10 (2004) 239–252. [16] E. Feireisl, F. Issard-Roch, H. Petzeltová, A non-smooth version of the Łojasiewicz–Simon theorem with applications to non-local phase-field systems, J. Differential Equations 199 (2004) 1–21. [17] E. Feireisl, G. Schimperna, Large time behaviour of solutions to Penrose–Fife phase change models, Math. Methods Appl. Sci. 28 (2005) 2117–2132. [18] E. Feireisl, F. Simondon, Convergence for semilinear degenerate parabolic equations in several space dimensions, J. Dynam. Differential Equations 12 (2000) 647–673. [19] H. Gajewski, A.-J. Griepentrog, A descent method for the free energy of multicomponent systems, Discrete Contin. Dyn. Syst. 15 (2006) 505–528. [20] P. Galenko, D. Jou, Diffuse–interface model for rapid phase transformations in nonequilibrium systems, Phys. Rev. E 71 (2005), 046125 (13 pages). [21] P. Galenko, V. Lebedev, Analysis of dispersion relation in spinodal decomposition of a binary system, Philos. Mag., in press. [22] S. Gatti, M. Grasselli, A. Miranville, V. Pata, Hyperbolic relaxation of the viscous Cahn–Hilliard equation in 3-D, Math. Models Methods Appl. Sci. 15 (2005) 165–198. [23] S. Gatti, M. Grasselli, A. Miranville, V. Pata, On the hyperbolic relaxation of the one-dimensional Cahn–Hilliard equation, J. Math. Anal. Appl. 312 (2005) 230–247. [24] M. Grasselli, H. Petzeltová, G. Schimperna, Long time behavior of solutions to the Caginalp system with singular potential, Z. Anal. Anwend. 25 (2006) 51–72. [25] M. Grasselli, H. Petzeltová, G. Schimperna, Convergence to stationary solutions for a parabolic–hyperbolic phasefield system, Commun. Pure Appl. Anal. 5 (2006) 827–838.

60

M. Grasselli et al. / J. Differential Equations 239 (2007) 38–60

[26] M. Grasselli, H. Petzeltová, G. Schimperna, A nonlocal phase-field system with inertial term, Quart. Appl. Math., in press. [27] L. Herrera, D. Pávon, Hyperbolic theories of dissipation: Why and when do we need them?, Phys. A 307 (2002) 121–130. [28] D.D. Joseph, L. Preziosi, Heat waves, Rev. Modern Phys. 61 (1989) 41–73. [29] D.D. Joseph, L. Preziosi, Addendum to the paper: Heat waves [Rev. Modern Phys. 61 (1) (1989) 41–73], Rev. Modern Phys. 62 (1990) 375–391. [30] J.K. Hale, Asymptotic Behaviour of Dissipative Systems, Amer. Math. Soc., Providence, RI, 1988. [31] A. Haraux, Systèmes dynamiques dissipatifs et applications, Masson, Paris, 1991. [32] A. Haraux, M.A. Jendoubi, Convergence of bounded weak solutions of the wave equation with dissipation and analytic nonlinearity, Calc. Var. Partial Differential Equations 9 (1999) 95–124. [33] A. Haraux, M.A. Jendoubi, O. Kavian, Rate of decay to equilibrium in some semilinear parabolic equations, J. Evol. Equ. 3 (2003) 463–484. [34] S.-Z. Huang, P. Takáˇc, Convergence in gradient-like systems which are asymptotically autonomous and analytic, Nonlinear Anal. 46 (2001) 675–698. [35] M.A. Jendoubi, A simple unified approach to some convergence theorems of L. Simon, J. Funct. Anal. 153 (1998) 187–202. [36] M.A. Jendoubi, Convergence of global and bounded solutions of the wave equation with linear dissipation and analytic nonlinearity, J. Differential Equations 144 (1998) 302–312. [37] S. Łojasiewicz, Une propriété topologique des sous-ensembles analytiques réels, in: Les équations aux dérivées partielles, Paris, 1962, in: Colloques internationaux du CNRS, vol. 117, Editions du CNRS, Paris, 1963, pp. 87–89. [38] S. Łojasiewicz, Ensembles semi-analytiques, notes, I.H.E.S., Bures-sur-Yvette, 1965. [39] G. Mola, Global and exponential attractors for a conserved phase-field system with Gurtin–Pipkin heat conduction law, PhD thesis, Politecnico di Milano, Milan, 2006. [40] A. Novick-Cohen, On the viscous Cahn–Hilliard equation, in: Material Instabilities in Continuum Mechanics, Edinburgh, 1985–1986, Oxford Sci. Publ., Oxford Univ. Press, New York, 1988, pp. 329–342. [41] V. Pata, S. Zelik, A remark on the damped wave equation, Commun. Pure Appl. Anal. 5 (2006) 609–614. [42] P. Poláˇcik, K.P. Rybakowski, Nonconvergent bounded trajectories in semilinear heat equations, J. Differential Equations 124 (1996) 472–494. [43] P. Poláˇcik, F. Simondon, Nonconvergent bounded solutions of semilinear heat equations on arbitrary domains, J. Differential Equations 186 (2002) 586–610. [44] J. Prüss, R. Racke, S. Zheng, Maximal regularity and asymptotic behavior of solutions for the Cahn–Hilliard equation with dynamic boundary conditions, Ann. Mat. Pura Appl. (4) 185 (2006) 627–648. [45] J. Prüss, M. Wilke, Maximal Lp -regularity for the Cahn–Hilliard equation with nonconstant temperature and dynamic boundary conditions, in: Partial Differential Equations and Functional Analysis, in: Oper. Theory Adv. Appl., vol. 168, Birkhäuser, Basel, 2006, pp. 209–236. [46] P. Rybka, K.-H. Hoffmann, Convergence of solutions to Cahn–Hilliard equation, Comm. Partial Differential Equations 24 (1999) 1055–1077. [47] A. Segatti, On the hyperbolic relaxation of the Cahn–Hilliard equation in 3-D: Approximation and long time behaviour, Math. Models Methods Appl. Sci. 17 (2007) 411–437. [48] L. Simon, Asymptotics for a class of non-linear evolution equations with applications to geometric problems, Ann. of Math. (2) 118 (1983) 525–571. [49] R. Temam, Infinite-Dimensional Dynamical Systems in Mechanics and Physics, Springer, New York, 1997. [50] V. Vergara, Convergence to steady state for a phase field system with memory, PhD dissertation, Martin-LutherUniversität Halle-Wittenberg, Halle, Germany, June 2006. [51] H. Wu, M. Grasselli, S. Zheng, Convergence to equilibrium for a parabolic–hyperbolic phase-field system with Neumann boundary conditions, Math. Models Methods Appl. Sci. 17 (2007) 125–153. [52] H. Wu, S. Zheng, Convergence to equilibrium for the Cahn–Hilliard equation with dynamic boundary condition, J. Differential Equations 204 (2004) 511–531. [53] H. Wu, S. Zheng, Convergence to equilibrium for the damped semilinear wave equation with critical exponent and dissipative boundary condition, Quart. Appl. Math. 64 (2006) 167–188. [54] Z. Zhang, Asymptotic behavior of solutions to the phase-field equations with Neumann boundary conditions, Commun. Pure Appl. Anal. 4 (2005) 683–693. [55] S. Zheng, A. Milani, Exponential attractors and inertial manifolds for singular perturbations of the Cahn–Hilliard equations, Nonlinear Anal. 57 (2004) 843–877. [56] S. Zheng, A. Milani, Global attractors for singular perturbations of the Cahn–Hilliard equations, J. Differential Equations 209 (2005) 101–139.