This paper investigates the long-time behavior for a Navier–Stokes–Allen–Cahn system, a diffuse interface model for two-phase incompressible flows with unmatched densities, non-constant viscosities, and a singular Flory–Huggins potential. First, we establish the dissipativity of strong solutions via some a priori estimates. Then, we demonstrate the regular-continuity of the semigroup, which allows us to prove the existence of the global attractor in the strong solutions space.
Citation: Chunyou Sun, Junyan Tan. Attractors for a Navier–Stokes–Allen–Cahn system with unmatched densities[J]. Communications in Analysis and Mechanics, 2025, 17(1): 237-262. doi: 10.3934/cam.2025010
[1] | Hongxia Lin, Sabana, Qing Sun, Ruiqi You, Xiaochuan Guo . The stability and decay of 2D incompressible Boussinesq equation with partial vertical dissipation. Communications in Analysis and Mechanics, 2025, 17(1): 100-127. doi: 10.3934/cam.2025005 |
[2] | Zhigang Wang . Serrin-type blowup Criterion for the degenerate compressible Navier-Stokes equations. Communications in Analysis and Mechanics, 2025, 17(1): 145-158. doi: 10.3934/cam.2025007 |
[3] | Zhen Wang, Luhan Sun . The Allen-Cahn equation with a time Caputo-Hadamard derivative: Mathematical and Numerical Analysis. Communications in Analysis and Mechanics, 2023, 15(4): 611-637. doi: 10.3934/cam.2023031 |
[4] | Yang Liu . Global attractors for a nonlinear plate equation modeling the oscillations of suspension bridges. Communications in Analysis and Mechanics, 2023, 15(3): 436-456. doi: 10.3934/cam.2023021 |
[5] | Dandan Song, Xiaokui Zhao . Large-time behavior of cylindrically symmetric Navier-Stokes equations with temperature-dependent viscosity and heat conductivity. Communications in Analysis and Mechanics, 2024, 16(3): 599-632. doi: 10.3934/cam.2024028 |
[6] | Haifa El Jarroudi, Mustapha El Jarroudi . Asymptotic behavior of a viscous incompressible fluid flow in a fractal network of branching tubes. Communications in Analysis and Mechanics, 2024, 16(3): 655-699. doi: 10.3934/cam.2024030 |
[7] | Yang Liu, Xiao Long, Li Zhang . Long-time dynamics for a coupled system modeling the oscillations of suspension bridges. Communications in Analysis and Mechanics, 2025, 17(1): 15-40. doi: 10.3934/cam.2025002 |
[8] | Andrea Brugnoli, Ghislain Haine, Denis Matignon . Stokes-Dirac structures for distributed parameter port-Hamiltonian systems: An analytical viewpoint. Communications in Analysis and Mechanics, 2023, 15(3): 362-387. doi: 10.3934/cam.2023018 |
[9] | Jiangbo Han, Runzhang Xu, Chao Yang . Continuous dependence on initial data and high energy blowup time estimate for porous elastic system. Communications in Analysis and Mechanics, 2023, 15(2): 214-244. doi: 10.3934/cam.2023012 |
[10] | Huiyang Xu . Existence and blow-up of solutions for finitely degenerate semilinear parabolic equations with singular potentials. Communications in Analysis and Mechanics, 2023, 15(2): 132-161. doi: 10.3934/cam.2023008 |
This paper investigates the long-time behavior for a Navier–Stokes–Allen–Cahn system, a diffuse interface model for two-phase incompressible flows with unmatched densities, non-constant viscosities, and a singular Flory–Huggins potential. First, we establish the dissipativity of strong solutions via some a priori estimates. Then, we demonstrate the regular-continuity of the semigroup, which allows us to prove the existence of the global attractor in the strong solutions space.
This paper considers the following Navier–Stokes–Allen–Cahn (NSAC) system modeling for two-phase flow with unmatched densities and viscosities, reading as follows:
{ρ(ϕ)∂tu−div(ν(ϕ)Du)+ρ(ϕ)u⋅∇u+∇P=−div(∇ϕ⊗∇ϕ),divu=0,∂tϕ+u⋅∇ϕ=−μ−ρ′(ϕ)|u|22+¯μ+ρ′(ϕ)|u|22,in Ω×(0,∞), | (1.1) |
where Ω⊂R2 is a bounded domain with a smooth boundary. Considering the no-slip boundary condition for u, the homogeneous Neumann boundary condition for ϕ
{u=0,∂nϕ=0,on ∂Ω×(0,∞), | (1.2) |
and the initial conditions
{u(⋅,0)=u0,ϕ(⋅,0)=ϕ0,in Ω. | (1.3) |
Here, u=u(x,t) represents the volume-averaged fluid velocity field, while P=P(x,t) denotes the pressure. The viscosity of the mixture, ν, is not constant, and ρ denotes the density of the mixture, which depends on the phase function ϕ. D is the symmetric gradient, which has the following form: D=12(∇+∇T). The chemical potential is defined by
μ=Ψ′(ϕ)−Δϕ, | (1.4) |
and ¯X=1|Ω|∫ΩXdx denotes the spatial average of the term X. As an example, though not the only possibility, we can consider the averaged density
ρ(ϕ)=ρ11+ϕ2+ρ21−ϕ2, |
and the averaged viscosity of the binary fluids
ν(ϕ)=ν11+ϕ2+ν21−ϕ2, |
where ρ1 and ρ2 denote the densities of the two fluids, and ν1 and ν2 represent their respective viscosities. The function Ψ is the double-well free energy density, also known as the Flory–Huggins potential, which is given by
Ψ(s)=θ2((1+s)ln(1+s)+(1−s)ln(1−s))−θ02s2=F(s)−θ02s2, | (1.5) |
for every s∈[−1,1], where θ and θ0 are two positive constants representing the absolute temperature of the mixture and the critical temperature, respectively, and they satisfy 0<θ<θ0.
Investigating the dynamics of two-phase flows is one of the most attractive and important problems within the hydrodynamic theory of fluids, with the Allen–Cahn equation playing a fundamental role (see [1,2]). The interface between two fluids is a (d−1)-dimensional manifold, posing great challenges both to the theoretical analysis and to the computational applications. Recently, a method called the diffuse-interface approach has emerged as a powerful technique for the study of interface theory (see [3,4,5,6]). The diffuse-interface method introduces a labeling function to replace the sharp interfaces with transition layers of width ε>0, where ε is a small parameter. Under this framework, the dynamics of interfaces between two fluids recognized as level sets of the order parameter can be naturally described (see [7]). Within the diffuse-interface framework, the phase function ϕ represents the contrast between local concentrations of the two fluids.
Two commonly used model equations in the study of the evolution of binary fluid systems with mass conservation are the following.
(1) Mass-conserving Allen–Cahn equation (see [8])
{∂tφ+u⋅∇φ+m(μ−¯μ)=0,inΩ×(0,T),∂nφ=0,on∂Ω×(0,T); | (1.6) |
(2) Cahn–Hilliard equation (see [9])
{∂tφ+u⋅∇φ−div(m∇μ)=0,inΩ×(0,T),∂nφ=∂nμ=0,on∂Ω×(0,T), | (1.7) |
where μ represents the chemical potential, and m is a physically relevant constant.
The transport equation, in contrast to the Allen–Cahn and Cahn–Hilliard equations, does not include a diffusive term, and thus fails to maintain the proper shape of the diffuse interface along the normal direction, which motivates us to study the incompressible Navier–Stokes equations coupled with either the Allen–Cahn equation (1.6) or the Cahn–Hilliard equation (1.7), without dropping the crucial conservation laws (see (4.1)).
Recent research on incompressible binary fluid mixtures has led to significant findings. For the phase-field model of binary fluids, including the case of equal densities or small density contrasts, which can be approximated by the Boussinesq equations, we refer readers to [3,10,11,12,13] and references therein. We also refer interested readers to [14] for a nice reference about the compressible Navier–Stokes equations with Onsager's regularity. Nevertheless, in most physical models, the density differences between two fluids are non-negligible. Significant contributions to the Navier–Stokes–Allen–Cahn models with constant density have been made by the authors in [15,16,17]. However, the authors chose the potential as the classical Landau double-well form as well as the lack of mass conservation (see e.g. [18] and [19] and references therein).
The model (1.1) was derived by Onsager (see [20,21]), and we also refer interested readers to [22] for the Navier–Stokes–Cahn–Hilliard system. We also mention that there are several works for coupled nonlinear parabolic systems (see [23,24,25]) by introducing the so-called potential well method for the global existence of weak solutions (see [26,27,28]). Such systems are particularly relevant in the modeling of two-phase fluid systems, where the complex interplay between the phases often results in nonlinear coupling. The authors in [21] demonstrated both the well-posedness and the existence of the global attractor associated with system (1.1) in the 2-dimensional case. Their analysis focused on a specific case when the potential is smooth. The system they considered lacks a mass-conserving law and has constant viscosity. As a result, they ensured that the phase function ϕ remained confined within the physical range [−1,1], which is essential to their analysis. The well-posedness, regularity, and existence of the global attractor for the Navier–Stokes–Cahn–Hilliard system were established by the authors in [29] and [30]. The system they considered has matched density, and consequently, the uniqueness of the weak solution was easily obtained. This, together with the higher-order regularity of the phase function ϕ in (1.7), ensured that the dynamical system they constructed was on a lower-order regularity space Hσ×H1(Ω), and they obtained the compact absorbing set by dissipativity estimates in Vσ×H2(Ω) (see definition in section 2). Nevertheless, compared to the NSCH system, the NSAC system (1.1) contains only second-order diffusion terms. As a result, the regularity of ϕ is lower and we need more delicate estimates for ϕ (for more details, we refer to Proposition 4.1 in this paper and the argument of absorbing set in Theorem 4.1 in [30]). Moreover, since there is currently no theoretical proof of the uniqueness of weak solutions for the NSAC systems with unmatched densities, we can only consider strong solutions and construct the global attractor in a higher-regularity space Vσ×H2(Ω), and therefore we need a much higher estimate of solutions in H2σ×H3(Ω) to get the existence of a compact absorbing set (see Proposition 4.4 and Proposition 4.5 for more details).
The authors in [31] established the existence of a global weak solution of (1.1) in both 2-dimensional and 3-dimensional cases, together with the uniqueness of weak solutions with matched densities in the 2-dimensional case. Additionally, they proved the existence and uniqueness of strong solutions in the 2-dimensional case and derived several entropy estimates. However, there is no successful method that gives the uniqueness of weak solutions to system (1.1) with unmatched densities in the 2-dimensional case.
Before concluding this introduction, we give some additional remarks about our work. This study investigates the long-time behavior of solutions to the NSAC system (1.1). The system we considered here is more closely related to the actual physical model since the differences between densities and viscosities are not dropped. We also consider the system added a nonlinear term (1/2)ρ′(ϕ)|u|2 representing the force which effectively models the impact of macroscopic fluid effects on the microscopic description arising from density differences (see [31]). Building on the framework established in [31], we demonstrate the dissipativity in the complete metric space Hm. Due to Theorem 3.2, the existence of strong solutions provided in [31], we focus our analysis on strong solutions and construct an absorbing set on a suitable phase space Ym (refer to Section 3 for details). Moreover, because the chosen Flory–Huggins potential has singular derivatives, the uniform bound of F′′(ϕ) as time t away from zero is obtained by a corollary of Theorem 3.3, and this result enables us to derive compactness of the trajectories by proving dissipativity in a higher-regularity function space. Finally, applying the interpolation techniques, we demonstrate the continuity of the semigroup on the phase space Ym and obtain the existence of the global attractor. For further research, one may get a higher regularity of the global attractor by the framework in [32]. This analysis lies beyond the framework of the present study and will be investigated in future work.
The plan of this paper reads as follows: In section 2, we present the function spaces, several inequalities in analysis, the theory of elliptic and the Stokes problems, as well as some Gronwall-type lemmas. In section 3, we recall the well-posedness results shown in [31], and we introduce the dynamical system in a suitable phase space generated by (1.1)-(1.3). Section 4 gives the existence of the global attractor, demonstrating the existence of a bounded and compact absorbing set in the phase space together with the continuity of the semigroup.
Throughout this paper, the notation C=C(a1,a2,...,aN) indicates that the constant C is a positive constant depending on the quantities a1,a2,...,aN. The boldface letter (e.g., L) denotes the space of vector fields. If X is a metric space, BX(R) denotes the closed ball in X with radius R, centered at the origin. In this paper, A:B is defined as the inner product of two matrices A and B, given by A:B=tr(ATB), and we denote norms ‖⋅‖Lp(Ω),‖⋅‖H1(Ω)... by ‖⋅‖Lp,‖⋅‖H1... (1⩽p⩽+∞) unless otherwise specified.
Let Ω⊂R2 be a bounded domain with a smooth boundary ∂Ω. We denote by Hσ the closure of C∞0,σ(Ω) in L2(Ω), Vσ the closure of C∞0,σ(Ω) in H1(Ω), and H2σ the closure of C∞0,σ(Ω) in H2(Ω), where
C∞0,σ(Ω)={u∈C∞0(Ω):divu=0}. |
Then they are Hilbert spaces, and for convenience, we may still use ‖⋅‖L2, ‖⋅‖H1 and ‖⋅‖H2 for the norms in those spaces.
The Korn inequality related to the symmetric gradient reads as follows:
‖∇u‖L2≤√2‖Du‖L2≤√2‖∇u‖L2,for all u∈Vσ. | (2.1) |
We also recall the following inequalities in the 2D case (see[31]):
‖f‖L4≤C‖f‖12L2‖f‖12H1,for all f∈H1(Ω), | (2.2) |
‖f‖L∞≤C‖f‖12L2‖f‖12H2,for all f∈H2(Ω), | (2.3) |
‖∇f‖L4≤C‖f‖12H2‖f‖12L∞,for all f∈H2(Ω), | (2.4) |
‖f‖L∞≤C‖f‖H1ln12(e‖f‖H2‖f‖H1),for all f∈H2(Ω), | (2.5) |
‖f‖L∞≤C(p)‖f‖H1ln12(C(p)‖f‖W1,p‖f‖H1),for all f∈W1,p(Ω),p>2. | (2.6) |
We recall the following lemma and refer interested readers to [31] for a detailed proof.
Lemma 2.1 ([31]). Let f∈H1(Ω),g∈Lp(Ω) where Ω⊂R2 is a bounded domain with a smooth boundary and p>2. Then
‖fg‖L2≤C(pp−2)12‖f‖H1‖g‖L2ln12(e|Ω|p−22p‖g‖Lp‖g‖L2), |
for some C=C(Ω).
In the following, we recall an important differential inequality in order to obtain the dissipativities later (see [29,32] for more details).
Lemma 2.2 (Uniform Gronwall lemma in logarithm). Assume f>0 is absolutely continuous on [0,∞) and g,h>0 are both locally integrable on [0,∞), satisfying
f′(t)≤g(t)f(t)ln(e+f(t))+h(t),a.e.t≥0, |
and in addition the uniform bounds: for every t≥0,
∫t+rtf(τ)dτ≤a1,∫t+rtg(τ)dτ≤a2,∫t+rth(τ)dτ≤a3, |
for some r, a1,a2,a3>0. Then for every t≥r,
f(t)≤e(a1+rr+a3)ea2. |
Now we recall two lemmas for the Stokes problem and the elliptic estimate of Neumann problems (see [3] and [31]).
Lemma 2.3. Let Ω⊂R2 be a bounded domain with a smooth boundary. ν∈W2,∞(R), and satisfies 0<ν∗≤ν(s)≤ν∗ for all s∈R. φ∈W1,r(Ω), with r>2. The force g∈Lp(Ω), with p∈(1,∞). Assume that u∈Vσ is a weak solution of
{−div(ν(φ)Du)+∇P=g,inΩ,u=0,on∂Ω, |
in the following sense
(ν(φ)Du,∇v)=(g,v),for allv∈Vσ. |
Then,
‖u‖W2,p≤C(‖g‖Lp+‖∇φ‖Lr‖Du‖L2), |
for some positive constant C=C(p,Ω) and 1p=12+1r.
Lemma 2.4. Let Ω⊂R2 be a bounded domain with a smooth boundary. Assume that φ is the solution to the Neumann problem:
{−Δφ+F′(φ)=g,inΩ,∂nφ=0,on∂Ω. |
Then we have:
(a) If g∈Lp(Ω), p∈[2,∞], then
‖F′(φ)‖Lp≤‖g‖Lp. |
(b) If g∈H1(Ω), then
‖Δφ‖L2≤‖∇φ‖12L2‖∇g‖12L2, |
and for every p≥2, there exists a positive constant C=C(p,Ω), such that
‖φ‖W2,p+‖F′(φ)‖Lp≤C(1+‖g‖H1+‖φ‖L2). |
First we assume that the density and the viscosity ρ,ν∈C2([−1,1]) satisfy
0<ρ∗≤ρ(s)≤ρ∗,0<ν∗≤ν(s)≤ν∗, | (3.1) |
for every s∈[−1,1].
Next, we recall the well-posedness and regularity theorems given in [31].
Theorem 3.1 ([31]). Let Ω⊂R2 be a bounded domain with a smooth boundary, (u0,ϕ0)∈Hσ×(H1(Ω)∩L∞(Ω)) with ‖ϕ0‖L∞≤1 and |¯ϕ0|<1. Then there exists a weak solution (u,ϕ) to problem (1.1)-(1.3) on the interval [0,∞), satisfying:
(a) For every T>0,
u∈L∞(0,T;Hσ)∩L2(0,T;Vσ),ϕ∈L∞(0,T;H1(Ω))∩L2(0,T;H2(Ω)),∂tϕ∈L2(0,T;L2(Ω)),ϕ∈L∞(Ω×(0,T)):|ϕ(x,t)|<1a.e.inΩ×(0,T),μ∈L2(0,T;L2(Ω)),F′(ϕ)∈L2(0,T;L2(Ω)). |
(b) The pair (u,ϕ) solves the problem in the following sense:
−∫T0∫Ω(ρ′(ϕ)∂tϕη(t)+ρ(ϕ)η′(t))u⋅wdxdt+∫T0∫Ω(ρ(ϕ)u⋅∇u)⋅wη(t)dxdt+∫T0∫Ων(ϕ)(Du:Dw)η(t)dxdt=∫Ωρ(ϕ0)u0⋅wη(0)dx+∫T0∫Ω((∇ϕ⊗∇ϕ):∇w)η(t)dxdt, |
for every T>0, w∈Vσ, η∈C1([0,T]) with η(T)=0, and
∂tϕ+u⋅∇ϕ=Δϕ−Ψ′(ϕ)−ρ′(ϕ)|u|22+¯Ψ′(ϕ)+ρ′(ϕ)|u|22,a.e.(x,t)∈Ω×(0,T). |
u(⋅,0)=u0, ϕ(⋅,0)=ϕ0 in Ω, and ∂nϕ=0 almost everywhere on ∂Ω×(0,T).
(c) Set the total energy of the system by
E(u,ϕ)=∫Ω12ρ(ϕ)|u|2+12|∇ϕ|2+Ψ(ϕ)dx, | (3.2) |
then the weak solutions satisfy the energy inequality as follows:
E(u(t),ϕ(t))+∫t0∫Ων(ϕ(τ))|Du(τ)|2dx+∫t0‖(∂tϕ(τ)+u(τ)⋅∇ϕ(τ))‖2L2≤E(u0,ϕ0), | (3.3) |
for all t>0.
Theorem 3.2 ([31]). Let Ω⊂R2 be a bounded domain with a smooth boundary. Assume that (u0,ϕ0)∈Vσ(Ω)×H2(Ω) such that ‖ϕ0‖L∞≤1, |¯ϕ0|<1, μ0=Ψ′(ϕ0)−Δϕ0∈H1(Ω) and ∂nϕ0=0 on ∂Ω. Then there is a strong solution (u,ϕ) to problem (1.1)-(1.3) on the interval [0,∞), satisfying:
(a) For every T>0 and for every p∈(2,∞),
u∈L∞(0,T;Vσ)∩L2(0,T;H2(Ω))∩H1(0,T;Hσ),ϕ∈L∞(0,T;H2(Ω))∩L2(0,T;W2,p(Ω)),∂tϕ∈L∞(0,T;L2(Ω))∩L2(0,T;H1(Ω)),F′(ϕ)∈L∞(0,T;L2(Ω))∩L2(0,T;Lp(Ω)). |
The solution (u,ϕ) solves the system (1.1) almost everywhere in Ω×(0,∞). Moreover, u(⋅,0)=u0, ϕ(⋅,0)=ϕ0 in Ω, ∂nϕ=0 a.e. on ∂Ω×(0,T).
(b) If in addition there exists η1=η1(E(u0,ϕ0),‖u0‖Vσ,‖ϕ0‖H2,‖F′(ϕ0)‖L2,θ,θ0), such that ‖ρ′‖L∞(−1,1)≤η1 and F′′(ϕ0)∈L1(Ω), then for every T>0,
(F′′(ϕ))2ln(1+F′′(ϕ))∈L1(Ω×(0,T)), | (3.4) |
and furthermore, the strong solution that satisfies (3.4) is unique.
Theorem 3.3 ([31]). Let the assumptions of Theorem 3.2 be satisfied. Assume in addition that ‖ρ′‖L∞(−1,1)≤η1. If (u,ϕ) is the strong solution of system (1.1), then for every ξ>0, there exists a positive constant δ(ξ), such that the absolute value of ϕ is away from one:
−1+δ(ξ)≤ϕ(x,t)≤1−δ(ξ), |
for every x∈¯Ω and t≥ξ.
Remark 3.4. In contrast to the Navier–Stokes–Cahn–Hilliard system, the phase function ϕ can approach ±1 arbitrarily closely as t goes to zero (see [31] and [30] for a detailed discussion).
For any m∈(−1,1), we define the following spaces:
Hm=Hσ×Vm, |
Ym={(u,ϕ)∈Vσ×H2(Ω):|ϕ|≤1, a.e. ,¯ϕ=m,∂nϕ=0 on ∂Ω}, |
where
Vm={ϕ∈H1(Ω)∩L∞(Ω):‖ϕ‖L∞≤1,¯ϕ=m}. |
Then Hm and Ym are two complete metric spaces.
According to Theorems 3.1-3.2, the problem (1.1)-(1.3) generates a dynamical system: for each t≥0,
S(t):Ym→Ym, |
in the following sense
S(t)(u0,ϕ0)=(u(t),ϕ(t)), |
where (u(t),ϕ(t)) is the unique solution of problem (1.1)-(1.3). The dynamical system is a semigroup S(t) on Ym satisfying:
(a)S(0)=IdYm;(b)S(t+τ)=S(t)S(τ),for every t,τ≥0;(c)t→S(t)(u0,ϕ0)∈C([0,∞),Ym),for every (u0,ϕ0)∈Ym. |
In this section, we will prove the existence of the global attractor Am of the semigroup S(t) on the phase space Ym.
Proposition 4.1. There is a bounded set B0⊂Hm, such that for any bounded subset B of Hm, there exists t0(B)>0, which depends only on the Hm-bounds of B, satisfying
(u(t),ϕ(t))∈B0,for allt≥t0(B), |
where (u,ϕ) is the weak solution of (1.1) subject to the initial value (u0,ϕ0)∈B.
Proof. Let us fix R>0. We consider (u0,ϕ0)∈BHm(R)⊂Hm. First we integrate the equation (1.1)3 over Ω to obtain the mass conservation: for every t≥0,
∫Ωϕ(t)dx=∫Ωϕ0dx, | (4.1) |
and we define
m=¯ϕ(t)=1|Ω|∫Ωϕ(t)dx. |
By Theorem 3.1, we recall the energy identity:
ddtE(t)+∫Ων(ϕ)|Du|2dx+‖∂tϕ+u⋅∇ϕ‖2L2=0, | (4.2) |
for every t>0. For (1.1)3, we take the L2-inner-product with ϕ−¯ϕ=ϕ−m to obtain
12ddt‖ϕ‖2L2+‖∇ϕ‖2L2+∫ΩF′(ϕ)(ϕ−m)dx−θ0∫Ωϕ(ϕ−m)dx+∫Ωρ′(ϕ)(ϕ−m)|u|22dx=0. |
By multiplying the above equation by ε and summing with (4.2), we arrive at
ddt(E(t)+ε2‖ϕ‖2L2)+ε‖∇ϕ‖2L2+ε∫ΩF′(ϕ)(ϕ−m)dx+∫Ων(ϕ)|Du|2dx+‖∂tϕ+u⋅∇ϕ‖2L2≤C−ε∫Ωρ′(ϕ)(ϕ−m)|u|22dx≤C+C1ε‖u‖2L2≤C+C1ε‖∇u‖2L2, |
where C=C(θ0,m,Ω,ε) and C1=C1(ρ,m,Ω). Then by the Korn inequality and (3.1), we obatin
ddt(E(t)+ε2‖ϕ‖2L2)+(12ν∗−ε)‖∇u‖2L2+ε∫ΩF′(ϕ)(ϕ−m)dx+ε‖∇ϕ‖L2+‖∂tϕ+u⋅∇ϕ‖2L2≤C. |
Next we need an inequality, which can be found in [33]:
β∫Ω|F′(ϕ)|dx≤∫ΩF′(ϕ)(ϕ−m)dx+C0, | (4.3) |
for some β, C0>0, depending only on F and m. Then we obatin
ddt(E(t)+ε2‖ϕ‖2L2)+(12ν∗−ε)‖∇u‖2L2+εβ‖F′(ϕ)‖2L1+ε‖∇ϕ‖L2+‖∂tϕ+u⋅∇ϕ‖2L2≤C. |
Taking ε=14C1ν∗, and Ψ∗:=maxs∈[−1,1]|Ψ(s)|,Ψ∗:=maxs∈[−1,1]|Ψ(s)| we obtain
ddt(E(t)+ε2‖ϕ‖2L2)+14ν∗‖∇u‖2L2+14ν∗‖∇ϕ‖2L2+∫ΩΨ(ϕ)dx+ε‖ϕ‖2L2≤~C1, |
where ~C1=~C1(F,θ0,m,ν,Ω). By the definition of E(t), we obtain
ddt(E(t)+ε2‖ϕ‖2L2)+α(E(t)+ε2‖ϕ‖2L2)≤K20, | (4.4) |
where α=α(λ1,Ψ∗), depends on parameters of system (1.1), and λ1 is the first eigenvalue of the Stokes operator A, while K20:=~C1+|Ω|Ψ∗. By the Gronwall lemma, for each t≥0,
E(t)≤(E(0)+ε2‖ϕ0‖2L2)e−αt+(K′1)2, |
where (K′1)2=K20α. Thus by the definition of the energy E(t) again, we obtain the crucial inequality:
‖u‖2L2+‖∇ϕ‖2L2≤C(ρ)(‖u0‖2L2+‖∇ϕ0‖2L2)e−αt+K2. |
As (u0,ϕ0)∈BHm(R), when t≥te(R), for some te(R)=te(ρ,R),
C(ρ)(||u0||2L2+||∇ϕ0||2L2)e−αt≤1, |
and
‖u(t)‖2L2+‖∇ϕ(t)‖2L2≤K21, | (4.5) |
where K21:=K2+1, depends only on parameters of system (1.1). Then we can finish the proof by taking B0=BHm(K1).
Remark 4.2. As a direct consequence of the above proposition, we indeed obtain the dissipativity in the weak solution space. This may allow us to construct the so-called trajectory attractor for the weak solutions of the NSAC system (1.1) without uniqueness. On the other hand, (4.5) together with (4.4) yields, for all t≥0,
‖u(t)‖2L2+‖∇ϕ(t)‖2L2+∫t+1t(‖∇u(s)‖2L2+‖∂tϕ(s)+u⋅∇ϕ(s)‖2L2+‖F′(ϕ)‖L1)ds≤M0, | (4.6) |
for some M0 that depends only on the parameters of the system and K1.
Lemma 4.3. The following estimates hold for all t≥te(R):
‖ϕ‖2H2≤C(1+‖∂tϕ+u⋅∇ϕ‖2L2+‖Du‖2L2)≤C1(1+‖∂tϕ‖2L2+‖Du‖2L2), | (4.7) |
and
‖∂tϕ‖2L2≤C2(‖∂tϕ+u⋅∇ϕ‖2L2+‖Du‖2L2), | (4.8) |
for some C1,C2 dependent on parameters of (1.1) and K1.
Proof. Multiply (1.4) by −Δϕ and integrate over Ω to obtain
‖Δϕ‖2L2+∫ΩF′′(ϕ)|∇ϕ|2dx=θ0‖∇ϕ‖2L2−∫Ω(μ−¯μ)Δϕdx. |
As t≥te(R), and by (4.5),
‖Δϕ‖2L2≤θ0K21+‖Δϕ‖L2‖μ−¯μ‖L2≤θ0K21+12‖Δϕ‖2L2+12‖μ−¯μ‖2L2. |
Thus,
‖Δϕ‖2L2≤C(1+‖μ−¯μ‖2L2). |
Then by (4.5), we obatin
‖ϕ‖2H2≤C(1+‖μ−¯μ‖2L2), |
for any t≥te(R). Also by (2.4) and (4.5),
‖u⋅∇ϕ‖L2≤‖u‖L4‖∇ϕ‖L4≤C‖u‖12L2‖∇u‖12L2‖∇ϕ‖12L2‖ϕ‖12H2≤CK121‖∇u‖12L2K121‖ϕ‖12H2≤C‖∇u‖12L2‖ϕ‖12H2. |
We can infer from (1.1)3 that
∫Ω(∂tϕ+u⋅∇ϕ)(μ−¯μ)dx+‖μ−¯μ‖2L2+∫Ωρ′(ϕ)|u|22(μ−¯μ)dx=∫Ω(¯μ+ρ′(ϕ)|u|22−¯μ)(μ−¯μ)dx=0. |
Denoting (ρ′)∗=maxs∈[−1,1]|ρ′(s)|, we obtain that
‖μ−¯μ‖2L2=|−∫Ω(∂tϕ+u⋅∇ϕ)(μ−¯μ)dx−∫Ωρ′(ϕ)|u|22(μ−¯μ)dx|≤‖∂tϕ+u⋅∇ϕ‖L2‖μ−¯μ‖L2+‖ρ′(ϕ)|u|22‖L2‖μ−¯μ‖L2. |
Thus, we obtain
‖μ−¯μ‖L2≤C‖∂tϕ+u⋅∇ϕ‖L2+(ρ′)∗‖u‖2L4≤C(‖∂tϕ+u⋅∇ϕ‖L2+‖∇u‖L2), |
where C=C(K1). Consequently, we obtain
‖ϕ‖2H2≤C(1+‖∂tϕ+u⋅∇ϕ‖2L2+‖Du‖2L2)≤C(1+‖∂tϕ‖2L2+‖u⋅∇ϕ‖2L2+‖Du‖2L2)≤C(1+‖∂tϕ‖2L2+‖Du‖L2‖ϕ‖H2+‖Du‖2L2)≤C(1+‖∂tϕ‖2L2+‖Du‖2L2)+12‖ϕ‖2H2, |
which implies
‖ϕ‖2H2≤C1(1+‖∂tϕ‖2L2+‖Du‖2L2) |
for all t≥te(R).
Since
‖∂tϕ‖2L2≤C(‖∂tϕ+u⋅∇ϕ‖2L2+‖u⋅∇ϕ‖2L2)≤C(‖∂tϕ+u⋅∇ϕ‖2L2+‖Du‖L2‖ϕ‖H2)≤C(‖∂tϕ+u⋅∇ϕ‖2L2+‖Du‖2L2)+12‖∂tϕ‖2L2, |
we obtain that for each t≥te(R),
‖∂tϕ‖2L2≤C2(‖∂tϕ+u⋅∇ϕ‖2L2+‖Du‖2L2), | (4.9) |
where the parameters C1,C2 depend only on parameters of the system (1.1) and K1.
Proposition 4.4. The dynamical system (Ym,S(t)) possesses a bounded absorbing set B1, i.e., for any bounded set B⊂Ym, there exists t1(B)>0, depending only on the Ym-bound of B, such that for any t≥t1(B), S(t)B⊂B1.
Proof. This part of the proof is similar to the higher regular estimates of the NSAC system in [31], whereas keep in mind that (4.6) is valid for any t>te(R)+1.
Multiplying (1.1) by ∂tu, integrating over Ω, we get
∫Ωρ(ϕ)|∂tu|2dx+(ρ(ϕ)u⋅∇u,∂tu)+∫Ων(ϕ)Du⋅D∂tudx=−∫ΩΔϕ∇ϕ⋅∂tudx, |
where
12ddt∫Ων(ϕ)|Du|2dx=∫Ων(ϕ)Du:D∂tudx+12∫Ων′(ϕ)∂tϕ|Du|2dx, |
we get
12ddt∫Ων(ϕ)|Du|2dx+∫Ωρ(ϕ)|∂tu|2dx=−(ρ(ϕ)u⋅∇u,∂tu)+12∫Ων′(ϕ)∂tϕ|Du|2dx−∫ΩΔϕ∇ϕ⋅∂tudx. | (4.10) |
Differentiating (1.1)3 with respect to t, multiplying by ∂tϕ and integrating over Ω, we arrive at
12ddt‖∂tϕ‖2L2+‖∇∂tϕ‖2L2+∫ΩF′′(ϕ)|∂tϕ|2dx=−∫Ω∂tu⋅∇ϕ∂tϕdx+θ0‖∂tϕ‖2L2−∫Ωρ′′(ϕ)|∂tϕ|2|u|22dx−∫Ωρ′(ϕ)u⋅∂tu∂tϕdx+∂t(¯μ+ρ′(ϕ)|u|22)∫Ω∂tϕdx. | (4.11) |
By summing up (4.10) and (4.11), we obtain that
ddt(12∫Ων(ϕ)|Du|2dx+‖∂tϕ‖2L2)+‖∇∂tϕ‖2L2+ρ∗‖∂tu‖2L2+∫ΩF′′(ϕ)|∂tϕ|2dx≤−∫ΩΔϕ∇ϕ⋅∂tudx+12∫Ων′(ϕ)∂tϕ|Du|2dx+θ0‖∂tϕ‖2L2−∫Ω∂tu⋅∇ϕ∂tϕdx−(ρ(ϕ)u⋅∇u,∂tu)−∫Ωρ′′(ϕ)|∂tϕ|2|u|22dx−∫Ωρ′(ϕ)u⋅∂tu∂tϕdx. | (4.12) |
Then by (2.2)-(2.6) and Lemma 2.1, as well as a fundamental inequality: for each x,y>0,
x2ln(yx)≤x2ln(y)+1, | (4.13) |
we obtain that
ddtG(t)+ρ∗2‖∂tu‖2L2+12‖∇∂tϕ‖2L2≤C(G2(t)ln(C‖u‖W1,p)+(1+G2(t))ln(C‖ϕ‖W2,p)+G2(t)+1), | (4.14) |
where
G(t)=12∫Ων(ϕ(t))|Du(t)|2dx+12‖∂tϕ(t)‖2L2, | (4.15) |
and
ν∗2‖Du‖2L2+12‖∂tϕ‖2L2≤G(t)≤C(‖Du‖2L2+‖∂tϕ‖2L2). | (4.16) |
From Lemma 2.3, for the Stokes problem with the force term g=−ρ(ϕ)∂tu−ρ(ϕ)u⋅∇u−Δϕ∇ϕ, we have, for any ε∈(0,1),
‖u‖W2,1+ε≤C(‖∂tu‖L1+ε+‖u⋅∇u‖L1+ε+‖Δϕ∇ϕ‖L1+ε+‖Du‖L2‖∇ϕ‖Lr)≤C(‖∂tu‖L2+G(t)+1). | (4.17) |
By the Sobolev embedding W2,1+ε(Ω)↪W1,p(Ω) for every p∈(2,∞) and 1p=11+ε−12,
‖u‖W1,p≤C(‖∂tu‖L2+G(t)+1). | (4.18) |
Consider the elliptic problem
{−Δϕ+F′(ϕ)=μ+θ0ϕ,a.e. in Ω×(0,T),∂nϕ=0,a.e. on ∂Ω×(0,T), | (4.19) |
and Lemma 2.4,
‖ϕ‖W2,p+‖F′(ϕ)‖Lp≤C(1+‖ϕ‖L2+‖μ+θ0ϕ‖Lp)≤C(1+‖ϕ‖Lp+‖μ‖Lp), | (4.20) |
for any p∈[2,∞). By (1.1)3, we have
‖μ−¯μ‖Lp≤‖∂tϕ‖Lp+‖u⋅∇ϕ‖Lp+‖ρ′(ϕ)|u|22−¯ρ′(ϕ)|u|22‖Lp≤C(‖∇∂tϕ‖L2+‖u‖H1‖ϕ‖H2+‖∇u‖2L2). |
Note that ¯μ=¯F′(ϕ)−θ0ϕ, and thus |¯μ|≤C(1+‖F′(ϕ)‖L1). Taking the L2-inner product of (1.1) with ϕ−m, we obtain
‖∇ϕ‖2L2+∫ΩF′(ϕ)(ϕ−m)dx=∫Ω(μ−¯μ)(ϕ−m)dx+∫Ωθ0ϕ(ϕ−m)dx. |
By (4.3),
‖F′(ϕ)‖L1≤C(1+‖μ−¯μ‖L2). | (4.21) |
Thus,
‖μ‖Lp≤C‖μ−¯μ‖Lp+C|¯μ|≤C(G(t)+‖∇∂tϕ‖L2+1). |
Then by (4.20)
‖ϕ‖W2,p≤C(G(t)+‖∇∂tϕ‖L2+1). | (4.22) |
Consequentially, applying the following inequality: for each x,y>0,
xy≤(xlnx−x+1)+(ey−1), |
we obtain that
G2(t)ln(C‖u‖W1,p)≤ρ∗4‖∂tu‖2L2+C(G2(t)ln(e+G(t))+G2(t)+1). |
Similarly, we can obtain that
(1+G2(t))ln(C‖ϕ‖W2,p)≤18‖∇∂tϕ‖2L2+C((G(t)+G2(t))ln(e+G(t))+G2(t)+1). |
Then by (4.14),
ddt(e+G(t))+ρ∗4‖∂tu‖2L2+14‖∇∂tϕ‖2L2≤C(1+G(t)(e+G(t))ln(e+G(t))). | (4.23) |
Due to (4.16), we have
∫t+1tG(s)ds≤C(1+∫t+1t‖Du‖2L2+‖∂tϕ‖2L2ds)≤C, |
where using (4.5) and (4.6), we obtain
∫t+1t‖Du‖2L2≤CE(t)≤K21, |
and
∫t+1t‖∂tϕ‖2L2≤C∫t+1t‖∂tϕ+u⋅∇ϕ‖2L2+‖Du‖2L2ds≤2CE(t)≤CK21. |
Then by the uniform Gronwall lemma in logarithm, we get that, for all t≥te(R)+1:=t1(R),
G(t)≤(K′2)2. |
On the other hand, from (4.7) and (4.16), for every t≥te(R)+1:=t1(R),
(4.24) |
where depends only on the parameters of the system. Therefore, we can finish the proof by taking .
Proposition 4.5. For every and for any there exists and a bounded subset of , such that
for all and therefore is the compact absorbing set for on phase space .
Recall that the set was obtained in Proposition 4.4.
Proof. Recall that is the Stokes operator, and there exists . Then a.e. in with the following estimates:
(4.25) |
Multiplying by and taking integration over , we obtain
where
and
Then we obtain
(4.26) |
By (2.2), (4.5), (4.24), and the Young inequality, we obtain
and
By (2.2), (2.4), (4.5), (4.24), (4.25), and the global bound of in , we obtain
and
Thus, by the Sobolev embeddings, (2.6) and (4.20), we arrive at
(4.27) |
Adding (4.27) and (4.23), we obtain that as ,
(4.28) |
where and depends on the parameters of system (1.1). Integrating (4.28), and by (4.24), we get there exists depending on , such that
(4.29) |
for every
Multiplying with and taking integration over , we obtain
which yields
(4.30) |
By (2.5) and (4.24), and the Sobolev embeddings, we obtain
and
By the same argument in (4.27), the Agmon inequality together with the Sobolev embeddings, we obtain
and
From the above estimates, we obtain
(4.31) |
On the other hand, differentiating with respect to , multiplying , integrating over , we obtain
(4.32) |
By Theorem 3.3, there exists , such that for every and ,
Then . By (4.5), (4.24), and (4.27), we obtain
and by the Agmon inequality (2.3), we have that
From the above estimates, we obtain
Adding (4.31) into the above inequality, we obtain
(4.33) |
Taking the gradient of , we obtain
and consequently,
(4.34) |
Then by (4.33), we arrive at
(4.35) |
where , and
(4.36) |
By (4.29), we obtain
Then by the uniform Gronwall lemma (e.g., see [32]), we can obtain that
and by (4.34),
for every , and
for each . Thus, we obtain that is a compact absorbing set by the compact Sobolev embedding theorem, where and .
The proposition above directly leads to the following corollary.
Corollary 4.6. For each bounded set , there exists , such that is relatively compact in .
In this section, we will prove that for every , the semigroup is continuous on the phase space . Let be a sequence in and , such that in the strong topology of . Let . Then by assumption, there exists , such that, for all ,
Proposition 4.7. For any , there is a constant , such that
Proof. Let be the corresponding strong solutions initiated from and let .
Following the proof of Theorem 4.1 in [31], and the uniqueness of the strong solution, we obtain
where
and
moreover, and are both in for any Let which depends on following the proof in [31], and use the Gronwall lemma to obtain
Then
Following the proof of Proposition 4.5, we obtain that for each ,
(4.37) |
Note that in the proof of Proposition 4.5, in (4.29) is independent of the initial value and (4.29) is valid for . However, here in (4.37), depends on the initial data, but we need the inequality for . Then there is a such that satisfies , and
Taking as the initial datum, there exists a global strong solution , for each , there exists a constant ,
where
Also by (4.35), there is a constant , such that
Using the uniform Gronwall lemma again, we obtain
and
for every , where and . Then we obtain that for every
by viewing as initial values of the trajectories. Now let , , . By interpolation,
Replacing , we obtain
where and .
Thanks to these propositions above, we obtain the following theorem by Theorem 1.1 in [32].
Theorem 4.8. Let the assumptions of Theorem 3.2 and Theorem 3.3 hold. The dynamical system possesses a unique global attractor , which is a connected compact set and has the following properties:
(a) is strictly invariant in , i.e., for every ;
(b) is an attracting set for on , i.e., for every bounded ball in the phase space
where the Hausdorff semi-distance between sets is defined by .
Note that from Proposition 4.5, we know that the obtained global attractor is bounded in .
Chunyou Sun: Methodology, Writing-review & editing; Junyan Tan: Writing-original draft, Writing-review & editing.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
This work was partly supported by the NSFC (Grants No. 12271227)
The authors declare there is no conflict of interest.
[1] |
S. M. Allen, J. W. Cahn, A microscopic theory for antiphase boundary motion and its application to antiphase domain coarsening, Acta Metall., 27 (1979), 1085–1095. https://doi.org/10.1016/0001-6160(79)90196-2 doi: 10.1016/0001-6160(79)90196-2
![]() |
[2] | D. T. Wasan, Interfacial transport processes and rheology: structure and dynamics of thin liquid films, Chem. Eng. Educ., 26 (1992), 104–112. |
[3] |
H. Abels, On a diffuse interface model for two-phase flows of viscous, incompressible fluids with matched densities, Arch. Ration. Mech. Anal., 194 (2009), 463–506. https://doi.org/10.1007/s00205-008-0160-2 doi: 10.1007/s00205-008-0160-2
![]() |
[4] |
D. M. Anderson, G. B. McFadden, A. A. Wheeler, Diffuse-interface methods in fluid mechanics, Annu. Rev. Fluid Mech., 30 (1998), 139–165. https://doi.org/10.1146/annurev.fluid.30.1.139 doi: 10.1146/annurev.fluid.30.1.139
![]() |
[5] | J. J. Feng, C. Liu, J. Shen, P. Yue, An energetic variational formulation with phase field methods for interfacial dynamics of complex fluids: advantages and challenges, In: Modeling of Soft Matter, New York: Springer, 2005, 1–26. https://doi.org/10.1007/0-387-32153-5_1 |
[6] |
J. Shen, X. Yang, A phase-field model and its numerical approximation for two-phase incompressible flows with different densities and viscosities, SIAM J. Sci. Comput., 32 (2010), 1159–1179. https://doi.org/10.1137/09075860X doi: 10.1137/09075860X
![]() |
[7] |
J. Lowengrub, L. Truskinovsky, Quasi-incompressible Cahn-Hilliard fluids and topological transitions, Proc. R. Soc. Lond. A, 454 (1998), 2617–2654. https://doi.org/10.1098/rspa.1998.0273 doi: 10.1098/rspa.1998.0273
![]() |
[8] |
X. Chen, D. Hilhorst, E. Logak, Mass conserving Allen-Cahn equation and volume preserving mean curvature flow, Interfaces and Free Bound., 12 (2010), 527–549. https://doi.org/10.4171/IFB/244 doi: 10.4171/IFB/244
![]() |
[9] |
J. W. Cahn, On spinodal decomposition, Acta Metall., 9 (1961), 795–801. https://doi.org/10.1016/0001-6160(61)90182-1 doi: 10.1016/0001-6160(61)90182-1
![]() |
[10] |
Z. Han, R. Xu, Improved growth estimate for one-dimensional sixth-order Boussinesq equation with logarithmic nonlinearity, Appl. Math. Lett., 160 (2025), 109290. https://doi.org/10.1016/j.aml.2024.109290 doi: 10.1016/j.aml.2024.109290
![]() |
[11] |
R. Xu, Y. Yang, B. Liu, J. Shen, S. Huang, Global existence and blowup of solutions for the multidimensional sixth-order "good" Boussinesq equation, Z. Angew. Math. Phys., 66 (2015), 955–976. https://doi.org/10.1007/s00033-014-0459-9 doi: 10.1007/s00033-014-0459-9
![]() |
[12] |
Z. Han, R. Xu, Y. Yang, The qualitative behavior for one-dimensional sixth-order Boussinesq equation with logarithmic nonlinearity, Discrete Contin. Dyn. Syst. -S, 16 (2023), 3131–3145. https://doi.org/10.3934/dcdss.2023141 doi: 10.3934/dcdss.2023141
![]() |
[13] |
R. Xu, Y. Luo, J. Shen, S. Huang, Global existence and blow up for damped generalized Boussinesq equation, Acta Math. Appl. Sin. Engl. Ser., 33 (2017), 251–262. https://doi.org/10.1007/s10255-017-0655-4 doi: 10.1007/s10255-017-0655-4
![]() |
[14] |
R. Chen, Z. Liang, D. Wang, R. Xu, On the inviscid limit of the compressible Navier-Stokes equations near Onsager's regularity in bounded domains, Sci. China Math., 67 (2024), 1–22. https://doi.org/10.1007/s11425-022-2085-3 doi: 10.1007/s11425-022-2085-3
![]() |
[15] |
R. Hosek, V. Macha, Weak-strong uniqueness for Navier-Stokes/Allen-Cahn system, Czechoslovak Math. J., 69 (2019), 837–851. https://doi.org/10.21136/CMJ.2019.0520-17 doi: 10.21136/CMJ.2019.0520-17
![]() |
[16] |
H. Wu, Well-posedness of a diffuse-interface model for two-phase incompressible flows with thermo-induced Marangoni effect, Eur. J. Appl. Math., 28 (2017), 380–434. https://doi.org/10.1017/S0956792516000322 doi: 10.1017/S0956792516000322
![]() |
[17] |
X. Xu, L. Zhao, C. Liu, Axisymmetric solutions to coupled Navier-Stokes/Allen-Cahn equations, SIAM J. Math. Anal., 41 (2010), 2246–2282. https://doi.org/10.1137/090754698 doi: 10.1137/090754698
![]() |
[18] |
Y. Chen, V. D. Radulescu, R. Xu, High energy blowup and blowup time for a class of semilinear parabolic equations with singular potential on manifolds with conical singularities, Commun. Math. Sci., 21 (2023), 25–63. https://doi.org/10.4310/CMS.2023.v21.n1.a2 doi: 10.4310/CMS.2023.v21.n1.a2
![]() |
[19] |
Q. Lin, R. Xu, Global well-posedness of the variable-order fractional wave equation with variable exponent nonlinearity, J. Lond. Math. Soc., 111 (2025), e70091. https://doi.org/10.1112/jlms.70091 doi: 10.1112/jlms.70091
![]() |
[20] | M. Giga, A. Kirshtein, C. Liu, Variational modeling and complex fluids, In: Handbook of Mathematical Analysis in Mechanics of Viscous Fluids, Cham: Springer, 2018, 73–113. https://doi.org/10.1007/978-3-319-13344-7_2 |
[21] |
J. Jiang, Y. Li, C. Liu, Two-phase incompressible flows with variable density: an energetic variational approach, Discrete Contin. Dyn. Syst., 37 (2017), 3243–3284. https://doi.org/10.3934/dcds.2017138 doi: 10.3934/dcds.2017138
![]() |
[22] |
C. Liu, J. Shen, X. Yang, Decoupled energy stable schemes for a phase-field model of two-phase incompressible flows with variable density, J. Sci. Comput., 62 (2015), 601–622. https://doi.org/10.1007/s10915-014-9867-4 doi: 10.1007/s10915-014-9867-4
![]() |
[23] |
S. Chen, Y. Salmaniw, R. Xu, Global existence for a singular Gierer-Meinhardt system, J. Differ. Equ., 262 (2017), 2940–2960. https://doi.org/10.1016/j.jde.2016.11.022 doi: 10.1016/j.jde.2016.11.022
![]() |
[24] |
S. Chen, Y. Salmaniw, R. Xu, Bounded solutions to a singular parabolic system, J. Math. Anal. Appl., 455 (2017), 936–978. https://doi.org/10.1016/j.jmaa.2017.06.012 doi: 10.1016/j.jmaa.2017.06.012
![]() |
[25] |
R. Xu, W. Lian, Y. Niu, Global well-posedness of coupled parabolic systems, Sci. China Math., 63 (2020), 321–356. https://doi.org/10.1007/s11425-017-9280-x doi: 10.1007/s11425-017-9280-x
![]() |
[26] |
R. Xu, J. Liu, Y. Niu, S. Chen, Asymptotic behaviour of solution for multidimensional viscoelasticity equation with nonlinear source term, Bound. Value Probl., 2013 (2013), 1–13. https://doi.org/10.1186/1687-2770-2013-42 doi: 10.1186/1687-2770-2013-42
![]() |
[27] |
R. Xu, J. Su, Global existence and finite time blow-up for a class of semilinear pseudo-parabolic equations, J. Funct. Anal., 264 (2013), 2732–2763. https://doi.org/10.1016/j.jfa.2013.03.010 doi: 10.1016/j.jfa.2013.03.010
![]() |
[28] |
R. Xu, Y. Niu, Addendum to "Global existence and finite time blow-up for a class of semilinear pseudo-parabolic equations'' [J. Func. Anal. 264 (2013) 2732–2763], J. Funct. Anal., 270 (2016), 4039–4041. https://doi.org/10.1016/j.jfa.2016.02.026 doi: 10.1016/j.jfa.2016.02.026
![]() |
[29] |
A. Giorgini, A. Miranville, R. Temam, Uniqueness and regularity for the Navier-Stokes-Cahn-Hilliard system, SIAM J. Math. Anal., 51 (2019), 2535–2574. https://doi.org/10.1137/18M1223459 doi: 10.1137/18M1223459
![]() |
[30] |
A. Giorgini, R. Temam, Attractors for the Navier-Stokes-Cahn-Hilliard system, Discrete Contin. Dyn. Syst. -S, 15 (2022), 2249–2274. https://doi.org/10.3934/dcdss.2022118 doi: 10.3934/dcdss.2022118
![]() |
[31] |
A. Giorgini, M. Grasselli, H. Wu, On the mass-conserving Allen-Cahn approximation for incompressible binary fluids, J. Funct. Anal., 283 (2022), 109631. https://doi.org/10.1016/j.jfa.2022.109631 doi: 10.1016/j.jfa.2022.109631
![]() |
[32] | R. Temam, Infinite-dimensional dynamical systems in mechanics and physics, New York: Springer-Verlag, 1997. https://doi.org/10.1007/978-1-4612-0645-3 |
[33] | A. Miranville, The Cahn-Hilliard Equation: Recent Advances and Applications, CBMS-NSF Regional Conference Series in Applied Mathematics, vol. 95, Society for Industrial and Applied Mathematics (SIAM), Philadelphia, PA, 2019. https://doi.org/10.1137/1.9781611975925 |