Stability of steady state solutions associated with initial and boundary value problems of a coupled fluid-reaction-diffusion system in one space dimension is analyzed. It is shown that under Dirichlet-Dirichlet type boundary conditions, non-trivial steady state solutions exist and are locally stable when the system parameters satisfy certain constraints.
Citation: Hongyun Peng, Kun Zhao. On a hyperbolic-parabolic chemotaxis system[J]. Mathematical Biosciences and Engineering, 2023, 20(5): 7802-7827. doi: 10.3934/mbe.2023337
[1] | Zijuan Wen, Meng Fan, Asim M. Asiri, Ebraheem O. Alzahrani, Mohamed M. El-Dessoky, Yang Kuang . Global existence and uniqueness of classical solutions for a generalized quasilinear parabolic equation with application to a glioblastoma growth model. Mathematical Biosciences and Engineering, 2017, 14(2): 407-420. doi: 10.3934/mbe.2017025 |
[2] | Xiaoli Hu, Shengmao Fu . Global boundedness and stability for a chemotaxis model of Boló’s concentric sclerosis. Mathematical Biosciences and Engineering, 2020, 17(5): 5134-5146. doi: 10.3934/mbe.2020277 |
[3] | Lin Zhang, Yongbin Ge, Zhi Wang . Positivity-preserving high-order compact difference method for the Keller-Segel chemotaxis model. Mathematical Biosciences and Engineering, 2022, 19(7): 6764-6794. doi: 10.3934/mbe.2022319 |
[4] | Huy Tuan Nguyen, Nguyen Van Tien, Chao Yang . On an initial boundary value problem for fractional pseudo-parabolic equation with conformable derivative. Mathematical Biosciences and Engineering, 2022, 19(11): 11232-11259. doi: 10.3934/mbe.2022524 |
[5] | Xu Song, Jingyu Li . Asymptotic stability of spiky steady states for a singular chemotaxis model with signal-suppressed motility. Mathematical Biosciences and Engineering, 2022, 19(12): 13988-14028. doi: 10.3934/mbe.2022652 |
[6] | Davide Bellandi . On the initial value problem for a class of discrete velocity models. Mathematical Biosciences and Engineering, 2017, 14(1): 31-43. doi: 10.3934/mbe.2017003 |
[7] | Chichia Chiu, Jui-Ling Yu . An optimal adaptive time-stepping scheme for solving reaction-diffusion-chemotaxis systems. Mathematical Biosciences and Engineering, 2007, 4(2): 187-203. doi: 10.3934/mbe.2007.4.187 |
[8] | Qigang Deng, Fugeng Zeng, Dongxiu Wang . Global existence and blow up of solutions for a class of coupled parabolic systems with logarithmic nonlinearity. Mathematical Biosciences and Engineering, 2022, 19(8): 8580-8600. doi: 10.3934/mbe.2022398 |
[9] | Lorena Bociu, Giovanna Guidoboni, Riccardo Sacco, Maurizio Verri . On the role of compressibility in poroviscoelastic models. Mathematical Biosciences and Engineering, 2019, 16(5): 6167-6208. doi: 10.3934/mbe.2019308 |
[10] | Ana I. Muñoz, José Ignacio Tello . Mathematical analysis and numerical simulation of a model of morphogenesis. Mathematical Biosciences and Engineering, 2011, 8(4): 1035-1059. doi: 10.3934/mbe.2011.8.1035 |
Stability of steady state solutions associated with initial and boundary value problems of a coupled fluid-reaction-diffusion system in one space dimension is analyzed. It is shown that under Dirichlet-Dirichlet type boundary conditions, non-trivial steady state solutions exist and are locally stable when the system parameters satisfy certain constraints.
Chemotaxis is a phenomenon describing the influence of environmental chemical substances on the motion of various cells. Chemotaxis widely exists in various biological phenomena, such as cells aggregation [1], embryonic development [2], vascular network formation [3,4], etc. This paper is concerned with the following hyperbolic-parabolic chemotaxis system describing vasculogenesis
{∂tρ+∇⋅(ρu)=0,∂t(ρu)+∇⋅(ρu⊗u)+∇P(ρ)=−αρu+βρ∇Φ,τ∂tΦ=dΔΦ−aΦ+bρ, | (1.1) |
where (x,t)∈Ω×(0,∞). The model (1.1) was proposed in [5] to reproduce key features of experiments of in vitro formation of blood vessels showing that cells randomly spreading on a gel matrix autonomously organize to a connected vascular network (more extensive modeling details can be found in [6]), where the unknowns ρ=ρ(x,t)≥0 and u=u(x,t)∈Ω denote the density and velocity of endothelial cells, respectively, and Φ=Φ(x,t)≥0 denotes the concentration of the chemoattractant secreted by the endothelial cells. The convection term ∇⋅(ρu⊗u) models the cell movement persistence (inertial effect), P(ρ) is the cell-density dependent pressure function accounting for the fact that closely packed cells resist to compression due to the impenetrability of cellular matter, −αρu corresponds to a damping (friction) force with coefficient α>0 as a result of the interaction between cells and the underlying substratum and the quantity |β|>0 measures the intensity of cell response to the chemical concentration gradient, where β>0 (resp. β<0) means the chemotaxis is attractive (resp. repulsive) (cf. [7,8]). In this paper, we consider attractive chemotaxis. τ≥0 and d>0 are the relaxation time scale and diffusion coefficient of the chemoattractant, respectively, and the positive constants a and b denote the death and secretion rates of the chemoattractant, respectively. In the literature (cf. [9]), some parabolic-hyperbolic chemotaxis models with different structures than (1.1) have also been studied.
Chavanis and Sire obtained through asymptotic analysis in [10] that when the damping coefficient β is large, the solution of (1.1) converges to the solution of the Keller-Segel model. Subsequently, this has also been verified from the mathematical analysis in [11]. Natalini et al. in [12] numerically studied the difference and connection between the model (1.1) and the Keller-Segel chemotaxis model. By adding a viscous term Δu to the second equation of (1.1), the linear stability of the constant ground state [ˉρ,0,ˉΦ] was obtained in [13] under the condition
bP′(ˉρ)−aαˉρ>0. | (1.2) |
When the initial value [ρ0,u0,Φ0]∈[Hs(Rd)]d+2(s>d/2+1) is a small perturbation of the constant ground state [ˉρ,0,ˉΦ] with ˉρ>0 sufficiently small, it was shown in [14,15] that the system (1.1) admits global strong solutions without vacuum converging to [ˉρ,0,ˉΦ] with an algebraic rate (1+t)−34 as t→∞. In [16], when the pressure function P satisfies (1.2), the authors removed the limitation that the density is sufficiently small, obtained the global existence of classical solutions to (1.1), and improved the decay rates of the solutions. Subsequently, in [17], the authors also proved that the system (1.1) in R admits nonlinear diffusion waves which are stable against a small perturbation. Recently in [18], the well-posedness of global classical solutions to the Cauchy problem of (1.1) is established in homogeneous hybrid Besov spaces.
All the studies above are on the Cauchy problem of (1.1), and the problem becomes more complicated when boundary effects are considered. Subsequently, the stationary solutions of (1.1) with vacuum (bump solutions) in a bounded interval with zero-flux boundary condition were constructed in [19,20]. Recently, the stability of transition layer solutions of (1.1) on R+=[0,∞) was established in [21]. An interesting question is whether the stability of non-constant stationary solutions of (1.1) can be considered in bounded regions. In the following, we will consider the hyperbolic-parabolic chemotaxis system (1.1) on I=[0,1] with P=A0ρ2:
{ρt+(ρu)x=0,x∈I, t>0,ρut+ρuux+2A0ρρx=−αρu+βρΦx,x∈I, t>0,τΦt=dΦxx−aΦ+bρ,x∈I, t>0;(ρ,u,Φ)(x,0)=(ρ0,u0,Φ0)(x),x∈I;u|x=0,x=1=0,Φ|x=0,x=1=0,t>0, | (1.3) |
where A0 is a positive constant. In this paper, we shall first use delicate analysis to show that the system (1.3) has a unique non-constant stationary solution. Then we show that the stationary solution is locally asymptotically stable when the system parameters satisfy certain constraints.
To identify the stationary solution associated with the initial-boundary value problem (1.3), we first note that because of the dissipation mechanism induced by linear damping and the boundary condition for u, it is reasonable to expect that the equilibrium velocity is zero. Denote respectively the equilibrium density and concentration by ˆρ and ˆΦ. Taking into account the zero equilibrium velocity, we see that ˆρ and ˆΦ satisfy
{2A0ˆρˆρx=βˆρˆΦx,x∈I, t>0,dˆΦxx−aˆΦ+bˆρ=0,x∈I, t>0;ˆΦ|x=0,x=1=0,t>0. | (2.1) |
Lemma 2.1. The boundary value problem (2.1) admits a unique solution (ˆρ,ˆΦ). Moreover, dˆρ(x)dx and d2ˆρ(x)dx2 are small, when d is large while a, b, A0, and β are fixed.
Proof. The first equation of (2.1) implies 2A0ˆρ=βˆΦ+ˆC, for some constant ˆC which is to be determined later. Substituting ˆρ=β2A0ˆΦ+ˆC2A0 into the second equation of (2.1), we have
dˆΦxx−aˆΦ+bβ2A0ˆΦ+bˆC2A0=0. | (2.2) |
Let
Λ≡bβ−2aA02dA0,ˆD≡bˆCbβ−2aA0,Ψ≡ˆΦ−ˆD. | (2.3) |
Then we derive from (2.2) that
Ψxx−ΛΨ=0. | (2.4) |
Now we discuss the three cases regarding the sign of Λ.
Case I. If Λ=0 (i.e., bβ=2aA0), (2.2) can be written directly as
ˆΦxx+bˆC2dA0=0. |
Then we obtain ˆΦ(x)=−bˆC2dA0(x22+Ax+B) for some constants A and B. Since ˆΦ=0 when x=0 and x=1, we can deduce that ˆΦ(x)=−bˆC4dA0(x2−x) for all x∈[0,1]. This implies
ˆρ(x)=−bβˆC8dA20(x2−x)+ˆC2A0. | (2.5) |
Since the total cellular mass is conserved under the zero velocity boundary condition, we can show that
∫10ˆρ(x)dx=bβˆC48dA20+ˆC2A0=∫10ρ0(x)dx. |
This implies, if the initial total mass is positive,
0<A0∫10ρ0(x)dx<ˆC<4A0∫10ρ0(x)dx, | (2.6) |
when d>0 is sufficiently large while the other parameters are fixed. Taking into consideration that ρ represents the cell density, we should require ˆρ(x)>0 for all x∈[0,1]. Moreover, the reader will see from the asymptotic analysis presented below that ˆρ needs to be bounded from above and below away from zero, in order to obtain the stability of the non-constant stationary solution. Combining (2.5) and (2.6), and noting x−x2≤14 for all x∈[0,1], we see that
0<14∫10ρ0(x)dx<ˆρ(x)<4∫10ρ0(x)dx,∀x∈[0,1], | (2.7) |
when d is relatively large compared with the other parameters. This gives us the desired property of the equilibrium density.
Meanwhile, in the asymptotic analysis we will require dˆρ(x)dx and d2ˆρ(x)dx2 to be relatively small compared with ˆρ. To fulfill such a requirement, we observe that
dˆρ(x)dx=−bβˆCx4dA20+bβˆC8dA20→0,asd→∞,d2ˆρ(x)dx2=−bβˆC4dA20→0,asd→∞. |
Hence, in the case Λ=0, the smallness of dˆρ(x)dx and d2ˆρ(x)dx2 can be realized, when d is sufficiently large, while a, b, A0 and β are fixed.
Case II. When Λ>0, let λ=√Λ. Then we have from (2.4) that Ψ(x)=Aeλx+Be−λx for some constants A and B, which implies ˆΦ(x)=Aeλx+Be−λx+ˆD. Using the boundary conditions, we can show that
ˆΦ(x)=ˆDeλ+1(eλ+1−eλx−eλ(1−x)). | (2.8) |
Since ˆρ=β2A0ˆΦ+ˆC2A0, we obtain
ˆρ(x)=βˆD2A0(eλ+1)(eλ+1−eλx−eλ(1−x))+ˆC2A0. |
Using the conservation of total mass again, we can show that
∫10ˆρ(x)dx=βˆD(λeλ+λ−2eλ+2)2A0(eλ+1)λ+ˆC2A0=∫10ρ0(x)dx. |
Recalling the definition of ˆD (see (2.3)), we have
ˆC[bβ(λeλ+λ−2eλ+2)(bβ−2aA0)(eλ+1)λ+1]=2A0∫10ρ0(x)dx, |
which yields
ˆC=2A0(∫10ρ0(x)dx)[bβbβ−2aA0f1(λ)+1]−1, |
where
f1(λ)=λeλ+λ−2eλ+2(eλ+1)λ. |
Therefore,
ˆρ(x)=ˆC2A0[bβbβ−2aA0g1(λ;x)+1], |
where
g1(λ;x)=eλ+1−eλx−eλ(1−x)eλ+1. | (2.9) |
To guarantee ˆρ is bounded from above and below away from zero, we first note f1(λ)∈(0,1) for λ∈(0,∞) and limλ→0f1(λ)=0. Second, observe that since a>0, b>0, A0>0 and d>0, then Λ>0 if and only if bβ>2aA0>0. This implies bβbβ−2aA0>0. Hence, when a, b, A0 and β are fixed, there exists a small number λ0>0 such that
1<bβbβ−2aA0f1(λ)+1<2,∀λ∈(0,λ0). |
This implies
A0∫10ρ0(x)dx<ˆC<2A0∫10ρ0(x)dx,∀λ∈(0,λ0). | (2.10) |
Moreover, note that for all x∈[0,1], it holds that
0≤g1(λ;x)≤g1(λ;1/2)=(eλ2−1)2eλ+1→0,asλ→0. |
Therefore, there exists a small number λ1>0 such that
1<bβbβ−2aA0g1(λ;x)+1<2,∀x∈[0,1],∀λ∈(0,λ1), |
which implies
ˆC2A0<ˆρ(x)<ˆCA0,∀x∈[0,1],∀λ∈(0,λ1). |
In view of (2.10), we see that
12∫10ρ0(x)dx<ˆρ(x)<2∫10ρ0(x)dx,∀x∈[0,1],∀λ∈(0,min{λ0,λ1}). | (2.11) |
Since λ=√Λ=√bβ−2aA02dA0, the smallness of λ can be realized when d is sufficiently large, while a, b, A0 and β are fixed. We also observe that
dˆρ(x)dx=ˆC2A0⋅bβbβ−2aA0⋅−λeλx+λeλ(1−x)eλ+1→0,asλ→0,d2ˆρ(x)dx2=ˆC2A0⋅bβbβ−2aA0⋅−λ2eλx−λ2eλ(1−x)eλ+1→0,asλ→0. |
Hence, the smallness of dˆρ(x)dx and d2ˆρ(x)dx2 can be realized when λ is sufficiently small, or equivalently, when d is large while a, b, A0 and β are fixed.
Case III. When Λ<0, let λ=√−Λ. Then we have ˆΦ(x)=Acosλx+Bsinλx+ˆD. Using the boundary conditions, we get A+ˆD=0 and Acosλ+Bsinλ+ˆD=0. Note that B is uniquely determined if λ≠kπ (k∈N). In this case, we have A=−ˆD and B=cosλ−1sinλˆD. Hence, ˆΦ is given by
ˆΦ(x)=ˆD(cosλ−1sinλsinλx−cosλx+1). |
Following the same spirit as in Case II, we can show that
ˆρ(x)=ˆC2A0[bβbβ−2aA0g2(λ;x)+1], | (2.12) |
where
g2(λ;x)=sinλ−sinλcosλx+cosλsinλx−sinλxsinλ, |
and the constant ˆC is given by
ˆC=2A0(∫10ρ0(x)dx)[bβbβ−2aA0f2(λ)+1]−1, |
where
f2(λ)=λsinλ−2+2cosλλsinλ. |
Note that Λ<0 if and only if bβ<2aA0, and
bβbβ−2aA0{>0ifa>0andβ<0,<0ifa>0andβ>0. | (2.13) |
The function f2(λ) satisfies
limλ→0f2(λ)=0andf′2(λ)=2(cosλ−1)(λ−sinλ)λ2sin2λ. |
These imply when λ is sufficiently close to zero, f2(λ)<0 is sufficiently small. Regarding the two cases in (2.13), we can show that there exists a small number λ2>0, such that for all λ∈(0,λ2),
[bβbβ−2aA0f2(λ)+1]∈{(0.5,1)ifa>0andβ<0,(1,2)ifa>0andβ>0. |
In summary, the constant ˆC satisfies
A0∫10ρ0(x)dx<ˆC<4A0∫10ρ0(x)dx,∀λ∈(0,λ2). | (2.14) |
For g2(λ;x), we can show that
g2(λ;0)=0=g2(λ;1),dg2(λ;x)dx=λ(cosλ(1−x)−cosλx)sinλ,d2g2(λ;x)dx2=λ2(sinλ(1−x)+sinλx)sinλ, | (2.15) |
which imply dg2(λ;x)dx=0 when x=0.5, and g2(λ;x) is convex for x∈[0,1] when λ>0 is sufficiently small. These tell us g2(λ;x)≤0 for x∈[0,1] and
|g2(λ;x)|≤−g2(λ;0.5)=sec0.5λ−1, |
when λ>0 is sufficiently small. Hence, there exists a small number λ3>0, such that for all x∈[0,1],
[bβbβ−2aA0g2(λ;x)+1]∈{(0.5,1)ifa>0andβ<0,(1,2)ifa>0andβ>0. |
Therefore,
ˆC4A0<ˆρ(x)<ˆCA0,∀x∈[0,1],∀λ∈(0,λ3). | (2.16) |
Combining (2.14) and (2.16), we see that
14∫10ρ0(x)dx<ˆρ(x)<4∫10ρ0(x)dx,∀x∈[0,1],∀λ∈(0,min{λ2,λ3}). |
In addition, we see from (2.12) and (2.15) that
|dˆρ(x)dx|→0and|d2ˆρ(x)dx2|→0asλ→0. |
Thus, the smallness of the derivatives of ˆρ can be realized when λ>0 is sufficiently small. Combining the conclusions of the above three cases, we have completed the proof of Lemma 2.1.
With the preliminary discussions in §2.1, we now state the main results for (1.3). We first introduce some notations for convenience.
Notation 2.1. Throughout this paper, we use ‖⋅‖L2, ‖⋅‖Hs, and ‖⋅‖L∞ to denote the norms of the standard Lebesgue space L2((0,1)), Hilbert space Hs((0,1)), and Sobolev space L∞((0,1)), respectively. The total energy, of order s∈N, of the function f is denoted by
‖f(t)‖2s≡s∑k=0‖(∂ktf)(t)‖2Hs−k. | (2.17) |
In addition, we use ‖(f1,f2,...,fn)‖2⋆ to denote ‖f1‖2⋆+‖f2‖2⋆+⋯‖fn‖2⋆, where ⋆ denotes either L2, Hs, L∞, or s, whenever it is applicable. Unless otherwise specified, C will denote a generic positive constant which is independent of time. The value of the constant may vary line by line according to the context.
Theorem 2.2. Consider the initial-boundary value problem (1.3), where the parameters satisfy α>0, τ≥0, d>0, A0>0, a>0, b>0, and β∈R. Suppose the initial data satisfy ρ0(x)>0 for all x∈[0,1], (ρ0,u0)∈[H2((0,1))]2, Φ0∈H4((0,1)), and are compatible with the boundary conditions. Assume further that there is a small constant ε0>0, such that ‖(u0,ρ0−ˆρ)‖2H2+‖Φ0−ˆΦ‖2H4≤ε0, and there is a large constant d0>0, such that the diffusion coefficient d≥d0. Then there exists a unique solution to (1.3), which satisfies
‖(˜ρ,u)(t)‖22+2∑k=0‖(∂kt˜Φ)(t)‖2H4−2k+∫t0(‖(˜ρ,u)(τ)‖22+‖˜Φ(τ)‖2H4+‖˜Φt(τ)‖2H3+‖˜Φtt(τ)‖2H1)dτ≤C,∀t>0, |
where ˜ρ=ρ−ˆρ, ˜Φ=Φ−ˆΦ, and the constant C>0 is independent of t>0. Moreover, there are positive constants η1 and η2 which are independent of t>0, such that
‖(˜ρ,u)(t)‖22+2∑k=0‖(∂kt˜Φ)(t)‖2H4−2k≤η1e−η2t,∀t>0. |
We prove Theorem 2.2 by applying L2-based energy methods. First of all, it should be mentioned that the local well-posedness of classical solutions to (3.3) can be established by using some classical approaches in the literature, see e.g., [22,23], and we omit the details to simplify the presentation. The bulk of this paper is devoted to deriving the a priori estimates of the local solution, in order to extend it to a global one. We begin the proof with reformulating the first and second equations in (1.3) by using the sound speed transformation to obtain a symmetric hyperbolic system. Note that the boundary data of the spatial derivatives of the solution are unknown. Hence, the direct energy method (differentiating with respect to x, then performing L2-type estimates) is not accessible for the problem under consideration. One of the key steps in the proof is to reduce the estimate of the total (spatial and temporal) derivatives to the temporal ones only, using an iteration scheme based on the structure of the equations. Moreover, note that in the hyperbolic portion of the system, only the dissipation of u appears on the right-hand side of the second equation. We recover the dissipation mechanism of ρ by essentially working a wave-type equation of the function.
In this section, we give a proof of Theorem 2.2. The proof consists of three major steps: 1) apply the sound speed transformation to symmetrize the first two equations in (1.3); 2) reduce the estimate of the total (spatial and temporal) derivatives of the solution to the temporal ones only; 3) perform L2-based energy estimates. We first present the symmetrization process.
Since the principle part of the first two equations in (1.3) is hyperbolic, one needs to introduce an appropriate new variable to symmetrize these two equations, after which one can carry out L2-based energy estimates. For this purpose, we let σ=2√2A0ρ be the sound speed. Then the initial-boundary value problem (1.3) can be written in terms of σ, in the regime of classical solutions, as
{2σt+2uσx+σux=0,x∈I, t>0,2ut+2uux+σσx=−2αu+2βΦx,x∈I, t>0,8τA0Φt=8dA0Φxx−8aA0Φ+bσ2,x∈I, t>0;(σ,u,Φ)(x,0)=(2√2A0ρ0,u0,Φ0)(x),x∈I;u|x=0,x=1=0,Φ|x=0,x=1=0,t>0. | (3.1) |
To perform asymptotic analysis, leading to the global dynamics of the solution to (3.1), we need to write the system of equations in (3.1) in terms of the perturbed variables around the stationary solution. Since the stationary solution satisfies (2.1), letting ˆσ=2√2A0ˆρ, we can show that
{ˆσˆσx=2βˆΦx,8dA0ˆΦxx−8aA0ˆΦ+bˆσ2=0. | (3.2) |
Letting ˜σ=σ−ˆσ and ˜Φ=Φ−ˆΦ, we update (3.1) by using (3.2) as
{2˜σt+2u˜σx+˜σux+2uˆσx+ˆσux=0,x∈I, t>0,2ut+2uux+˜σ˜σx+˜σˆσx+ˆσ˜σx=−2αu+2β˜Φx,x∈I, t>0,8τA0˜Φt=8dA0˜Φxx−8aA0˜Φ+b(˜σ+2ˆσ)˜σ,x∈I, t>0;(˜σ,u,˜Φ)(x,0)=(2√2A0ρ0−2√2A0ˆρ,u0,Φ0−ˆΦ)(x),x∈I;u|x=0,x=1=0,˜Φ|x=0,x=1=0,t>0. | (3.3) |
The energy estimates derived in the rest of this section are based on the a priori assumption:
esssupt∈[0,T]X(t)≡esssupt∈[0,T]‖(˜σ,u,˜Φ)(t)‖22≤ε2, | (3.4) |
where T>0 denotes the lifespan of the local solution and ε>0 is a small number to be determined later. Note the smallness of ε can be realized by the smallness assumption of the initial perturbation in Theorem 2.2 and the local well-posedness theory. We will focus on deriving the time-independent a priori estimates of the local solution under (3.4), which, when combined with standard continuation argument, will generate the global well-posedness and long-time behavior of the solution in one stroke.
The rest of the proof consists of two major steps which are contained in two subsections. As was discussed in §2.1, the stationary solution takes on different forms, depending on the sign of bβ−2aA0. In the analysis presented below, we shall focus on the case when bβ−2aA0>0, in which the stationary solution is given by (2.8)–(2.9). The other case, i.e., bβ−2aA0≤0, can be proved in exactly the same fashion, and we omit the details for brevity.
We first deal with the case of τ>0 in §3.2 and §3.3. The proof of the case of τ=0 will be sketched in §3.4. We begin with the reduction of the total derivatives of the solution to (3.3).
Lemma 3.1. Let (˜σ,u,˜Φ) be the local solution to the IBVP (3.3) with τ>0 up to some finite time T>0. Assume (3.4) holds for some small ε>0. Then, under the conditions of Theorem 2.2, there exists a constant D0>1, which is independent of t, such that
X(t)≤D0X1(t):=D0‖(˜σt,˜σtt,u,ut,utt,˜Φx,˜Φxt,˜Φtt)(t)‖2L2. | (3.5) |
Proof. Step 1. We first derive a Poincaré-type inequality for ˜σ. From the discussions in §2.1 we infer that when the diffusion coefficient is sufficiently large, the stationary solution ˆρ satisfies (2.11). Denote the spatial integral of ρ0 by ¯ρ (which is positive by the assumptions of Theorem 2.2). Then we have
12¯ρ<ˆρ<2¯ρ. | (3.6) |
According to the definition of ˆσ, we know
2√A0¯ρ<ˆσ<4√A0¯ρ. | (3.7) |
Note that by definition,
˜σ=σ−ˆσ=2√2A0(√ρ−√ˆρ)=2√2A0ρ−ˆρ√ρ+√ˆρ. | (3.8) |
Since ρ0 is sufficiently close to ˆρ (by assumptions of Theorem 2.2) and ˆρ>12¯ρ>0, from the local well-posedness theory we know ρ(x,t) is positive within the lifespan of the local solution. Using such information, we deduce from (3.8) and (3.6) that
|˜σ|≤2√2A0√ˆρ|ρ−ˆρ|≤4√A0√¯ρ|ρ−ˆρ|, |
which implies
‖˜σ‖L2≤4√A0√¯ρ‖ρ−ˆρ‖L2. | (3.9) |
Since ρ−ˆρ is mean-free, it can be shown that
‖ρ−ˆρ‖L2≤‖(ρ−ˆρ)x‖L2. | (3.10) |
Since
ρ−ˆρ=σ2−ˆσ28A0=˜σ(˜σ+2ˆσ)8A0, |
we have
(ρ−ˆρ)x=˜σx(˜σ+2ˆσ)8A0+˜σ(˜σx+2ˆσx)8A0, |
which implies
‖(ρ−ˆρ)x‖L2≤(‖˜σ‖L∞+2‖ˆσ‖L∞)‖˜σx‖L28A0+(‖˜σx‖L∞+2‖ˆσx‖L∞)‖˜σ‖L28A0. | (3.11) |
Using (3.11), we update (3.10) as
‖ρ−ˆρ‖L2≤(‖˜σ‖L∞+2‖ˆσ‖L∞)‖˜σx‖L28A0+(‖˜σx‖L∞+2‖ˆσx‖L∞)‖˜σ‖L28A0. | (3.12) |
Substituting (3.12) into (3.9), we arrive at
‖˜σ‖L2≤12√A0¯ρ[(‖˜σ‖L∞+2‖ˆσ‖L∞)‖˜σx‖L2+(‖˜σx‖L∞+2‖ˆσx‖L∞)‖˜σ‖L2]≤12√A0¯ρ[(√2‖˜σ‖H1+8√A0¯ρ)‖˜σx‖L2+(√2‖˜σx‖H1+2‖ˆσx‖L∞)‖˜σ‖L2]≤12√A0¯ρ[(√2ε+8√A0¯ρ)‖˜σx‖L2+(√2ε+2‖ˆσx‖L∞)‖˜σ‖L2], | (3.13) |
where we used the 1D Sobolev inequality: ‖f‖L∞≤√2‖f‖H1, (3.4) and (3.7). Since
ˆσx=√2A0√ˆρˆρx, |
using (3.6), we can show that
‖ˆσx‖L∞≤√2A0√¯ρ‖ˆρx‖L∞. | (3.14) |
From the discussions in §2.1 we know when d is large enough, ‖ˆρx‖L∞ is sufficiently small. In this case, we denote (3.14) by
‖ˆσx‖L∞≤δ, | (3.15) |
where the constant δ decreases as d increases. Using (3.15), we update (3.13) as
‖˜σ‖L2≤12√A0¯ρ[(√2ε+8√A0¯ρ)‖˜σx‖L2+(√2ε+2δ)‖˜σ‖L2]. |
This implies when ε and δ are sufficiently small, such that
(√2ε+2δ)≤√A0¯ρ, | (3.16) |
it holds that
‖˜σ‖L2≤92‖˜σx‖L2+12‖˜σ‖L2. |
Hence,
‖˜σ‖L2≤9‖˜σx‖L2. | (3.17) |
Step 2. From the first equation of (3.3) we see that
ux=−2˜σ+ˆσ(˜σt+u˜σx+uˆσx). | (3.18) |
Using (3.7), Sobolev embedding, (3.4), and (3.16), we deduce that
‖˜σ+ˆσ‖L∞≥‖ˆσ‖L∞−‖˜σ‖L∞≥2√A0¯ρ−√2‖˜σ‖H1≥2√A0¯ρ−√2ε≥√A0¯ρ. | (3.19) |
Using (3.19), we deduce from (3.18) that
‖ux‖2L2≤12A0¯ρ(‖˜σt‖2L2+‖u‖2L∞‖˜σx‖2L2+‖ˆσx‖2L∞‖u‖2L2). | (3.20) |
Since u satisfies the zero boundary condition, it can be shown that
‖u‖L2≤‖ux‖L2. | (3.21) |
Using Sobolev embedding, (3.4), (3.15) and (3.21), we update (3.20) as
‖ux‖2L2≤12A0¯ρ(‖˜σt‖2L2+2ε2‖˜σx‖2L2+δ2‖ux‖2L2). | (3.22) |
Now, from the second equation of (3.3) we see that
˜σx=−1˜σ+ˆσ(2ut+2uux+˜σˆσx+2αu−2β˜Φx). | (3.23) |
Using (3.19), we can show that
‖˜σx‖2L2≤5A0¯ρ(4‖ut‖2L2+4‖u‖2L∞‖ux‖2L2+‖ˆσx‖2L∞‖˜σ‖2L2+4α2‖u‖2L2+4β2‖˜Φx‖2L2). | (3.24) |
Using (3.17), we update (3.24) as
‖˜σx‖2L2≤5A0¯ρ(4‖ut‖2L2+8ε2‖ux‖2L2+81δ2‖˜σx‖2L2+4α2‖u‖2L2+4β2‖˜Φx‖2L2). | (3.25) |
Taking the sum of (3.22) and (3.25) gives us
‖ux‖2L2+‖˜σx‖2L2≤1A0¯ρ[20‖ut‖2L2+12‖˜σt‖2L2+20α2‖u‖2L2+20β2‖˜Φx‖2L2+(40ε2+12δ2)‖ux‖2L2+(24ε2+405δ2)‖˜σx‖2L2]. |
When ε and δ are sufficiently small, we conclude that
‖ux‖2L2+‖˜σx‖2L2≤C‖(ut,˜σt,u,˜Φx)‖2L2. | (3.26) |
Step 3. Taking ∂t to (3.18), we obtain
uxt=−2˜σ+ˆσ(˜σtt+ut˜σx+u˜σxt+utˆσx)+2˜σt(˜σ+ˆσ)2(˜σt+u˜σx+uˆσx). |
Using similar arguments as in Step 1, we can derive the following estimate:
‖uxt‖2L2≤32A0¯ρ[‖˜σtt‖2L2+(2ε2+δ2)‖ut‖2L2+2ε2‖˜σxt‖2L2]+48A20¯ρ2(2ε4+ε2+ε2δ2)‖˜σt‖2L2. | (3.27) |
Taking ∂t to (3.23), we can show that
‖˜σxt‖2L2≤48A0¯ρ[4‖utt‖2L2+8ε2‖ut‖2L2+8ε2‖uxt‖2L2+δ2‖˜σt‖2L2+4α2‖ut‖2L2+4β2‖˜Φxt‖2L2]+80A20¯ρ2(8ε4+4ε2+ε2δ2+4α2ε2+4β2ε2)‖˜σt‖2L2. | (3.28) |
Taking the sum of (3.27) and (3.28), we have
‖uxt‖2L2+‖˜σxt‖2L2≤16A0¯ρ[2‖˜σtt‖2L2+12‖utt‖2L2+(28ε2+2δ2+12α2)‖ut‖2L2+3δ2‖˜σt‖2L2+12β2‖˜Φxt‖2L2+24ε2‖uxt‖2L2+4ε2‖˜σxt‖2L2]+16A20¯ρ2(46ε4+23ε2+8ε2δ2+20α2ε2+20β2ε2)‖˜σt‖2L2. |
When ε and δ are sufficiently small, there holds that
‖uxt‖2L2+‖˜σxt‖2L2≤C‖(utt,˜σtt,ut,˜σt,˜Φxt)‖2L2. | (3.29) |
Step 4. Taking ∂x to (3.18) and using Poincaré inequality for u, we can derive the following estimate:
‖uxx‖2L2≤20A0¯ρ(‖˜σxt‖2L2+(2ε2+δ2+‖ˆσxx‖2L∞)‖ux‖2L2+2ε2‖˜σxx‖2L2)+24A20¯ρ2(2ε2+δ2)(‖˜σt‖2L2+(2ε2+δ2)‖ux‖2L2). | (3.30) |
Note that
ˆσxx=√2A0√ˆρˆρxx−√2A02ˆρ√ˆρ(ˆρx)2. |
From the discussions in §2.1 we know that ˆρx and ˆρxx are small when d is large. Hence, as in (3.15), we may assume ‖ˆσxx‖L∞≤δ, as well. Then, we update (3.30) as
‖uxx‖2L2≤C(‖˜σxt‖2L2+‖˜σt‖2L2+‖ux‖2L2+ε2‖˜σxx‖2L2). | (3.31) |
Next, taking ∂x to (3.23) and using (3.17), we can show that
‖˜σxx‖2L2≤C(‖uxt‖2L2+‖ut‖2L2+‖ux‖2L2+‖˜σx‖2L2+‖˜Φxx‖2L2+‖˜Φx‖2L2+ε2‖uxx‖2L2). | (3.32) |
Taking the sum of (3.31) and (3.32) gives us
‖uxx‖2L2+‖˜σxx‖2L2≤C(‖uxt‖2L2+‖˜σxt‖2L2+‖ut‖2L2+‖˜σt‖2L2+‖ux‖2L2+‖˜σx‖2L2+‖˜Φxx‖2L2+‖˜Φx‖2L2+ε2(‖˜σxx‖2L2+‖uxx‖2L2)). | (3.33) |
It is clear that when ε is small, we can update (3.33) as
‖uxx‖2L2+‖˜σxx‖2L2≤C(‖uxt‖2L2+‖˜σxt‖2L2+(‖ux‖2L2+‖˜σx‖2L2)+‖ut‖2L2+‖˜σt‖2L2+‖˜Φxx‖2L2+‖˜Φx‖2L2). | (3.34) |
By (3.26) and (3.29), we further update (3.34) as
‖uxx‖2L2+‖˜σxx‖2L2≤C(‖utt‖2L2+‖˜σtt‖2L2+‖ut‖2L2+‖˜σt‖2L2+‖u‖2L2+‖˜Φxt‖2L2+‖˜Φx‖2L2+‖˜Φxx‖2L2). | (3.35) |
Step 5. Since ˜Φ and ˜Φt satisfy the zero boundary condition, it follows from Poincaré inequality that
‖˜Φ‖2L2≤‖˜Φx‖2L2and‖˜Φt‖2L2≤‖˜Φxt‖2L2. | (3.36) |
Moreover, using (3.36), (3.17) and (3.26), we can deduce from the third equation of (3.3) that
‖˜Φxx‖2L2≤C‖(˜Φxt,˜Φx,ut,˜σt,u)‖2L2. | (3.37) |
Substituting (3.37) into (3.35), we get
‖uxx‖2L2+‖˜σxx‖2L2≤C‖(utt,˜σtt,ut,˜σt,u,˜Φxt,˜Φx)‖2L2. | (3.38) |
Combining (3.17), (3.26), (3.29), (3.36), (3.37) and (3.38), we arrive at (3.5). This completes the proof of the lemma.
In this subsection, we examine the quantity X1(t) defined in (3.5) and derive the desired energy estimates, along with the exponential decaying of the perturbed solution.
Lemma 3.2. Let (˜σ,u,˜Φ) be the local solution to the IBVP (3.3) with τ>0 up to some finite time T>0. Then under the conditions of Theorem 2.2, the quantity
‖(˜σ,u)(t)‖22+2∑k=0‖(∂kt˜Φ)(t)‖2H4−2k+∫t0(‖(˜σ,u)(τ)‖22+‖˜Φ(τ)‖2H4+‖˜Φt(τ)‖2H3+‖˜Φtt(τ)‖2H1)dτ |
is uniformly bounded with respect to t>0, and ‖(˜σ,u)(t)‖22+∑2k=0‖(∂kt˜Φ)(t)‖2H4−2k decays exponentially rapidly to zero as t→∞.
Proof. Step 1. Taking L2 inner product of the first equation in (3.3) with ˜σ, we have
ddt‖˜σ‖2L2=−2∫10u˜σx˜σdx−∫10˜σ2uxdx−2∫10uˆσx˜σdx−∫10ˆσux˜σdx=−2∫10uˆσx˜σdx−∫10ˆσux˜σdx, | (3.39) |
where we used the zero boundary condition for u. Taking L2 inner product of the second equation in (3.3) with u gives us
ddt‖u‖2L2+2α‖u‖2L2=−2∫10u2uxdx−∫10˜σ˜σxudx−∫10˜σˆσxudx−∫10ˆσ˜σxudx+2β∫10˜Φxudx=−∫10˜σ˜σxudx−∫10˜σˆσxudx−∫10ˆσ˜σxudx+2β∫10˜Φxudx. | (3.40) |
Taking the sum of (3.39) and (3.40), we obtain
ddt(‖˜σ‖2L2+‖u‖2L2)+2α‖u‖2L2=−2∫10uˆσx˜σdx−∫10˜σ˜σxudx+2β∫10˜Φxudx≤(2‖ˆσx‖L∞+‖˜σx‖L∞)‖u‖L2‖˜σ‖L2+2β‖˜Φx‖L2‖u‖L2. | (3.41) |
Since, by Young's inequality,
2β‖˜Φx‖L2‖u‖L2≤α‖u‖2L2+α−1β2‖˜Φx‖2L2, |
we update (3.41) as
ddt(‖˜σ‖2L2+‖u‖2L2)+α‖u‖2L2≤(2‖ˆσx‖L∞+‖˜σx‖L∞)‖u‖L2‖˜σ‖L2+α−1β2‖˜Φx‖2L2. |
By Sobolev embedding and Cauchy-Schwarz inequality, we can show that
ddt(‖˜σ‖2L2+‖u‖2L2)+α‖u‖2L2≤(δ+2−12ε)(‖u‖2L2+‖˜σ‖2L2)+α−1β2‖˜Φx‖2L2, | (3.42) |
where we used (3.15) and (3.4).
Step 2. Taking ∂t to the three equations in (3.3), we have
{2˜σtt+2ut˜σx+2u˜σxt+˜σtux+˜σuxt+2utˆσx+ˆσuxt=0,2utt+2utux+2uuxt+˜σt˜σx+˜σ˜σxt+˜σtˆσx+ˆσ˜σxt=−2αut+2β˜Φxt,8τA0˜Φtt=8dA0˜Φxxt−8aA0˜Φt+2b(˜σ+ˆσ)˜σt. | (3.43) |
Taking L2 inner product of the first equation in (3.43) with ˜σt, we have
ddt‖˜σt‖2L2=−2∫10ut˜σx˜σtdx−2∫10u˜σxt˜σtdx−∫10˜σ2tuxdx−∫10˜σuxt˜σtdx−2∫10utˆσx˜σtdx−∫10ˆσuxt˜σtdx=−∫10˜σxut˜σtdx+∫10˜σut˜σxtdx−∫10utˆσx˜σtdx+∫10ˆσut˜σxtdx. | (3.44) |
Taking L2 inner product of the second equation in (3.43) with ut, we obtain
ddt‖ut‖2L2+2α‖ut‖2L2=−2∫10uxu2tdx−2∫10uuxtutdx−∫10˜σt˜σxutdx−∫10˜σ˜σxtutdx−∫10˜σtˆσxutdx−∫10ˆσ˜σxtutdx+2β∫10˜Φxtutdx=−∫10uxu2tdx−∫10˜σt˜σxutdx−∫10˜σ˜σxtutdx−∫10˜σtˆσxutdx−∫10ˆσ˜σxtutdx+2β∫10˜Φxtutdx. | (3.45) |
Taking the sum of (3.44) and (3.45), we arrive at
ddt(‖˜σt‖2L2+‖ut‖2L2)+2α‖ut‖2L2=−2∫10(˜σx+ˆσx)ut˜σtdx−∫10uxu2tdx+2β∫10˜Φxtutdx≤2(‖˜σx‖L∞+‖ˆσx‖L∞)‖ut‖L2‖˜σt‖L2+‖ux‖L∞‖ut‖2L2+2β‖˜Φxt‖L2‖ut‖L2. |
Similar to (3.42), it can be shown that
ddt(‖˜σt‖2L2+‖ut‖2L2)+α‖ut‖2L2≤(δ+2−12ε)(‖ut‖2L2+‖˜σt‖2L2)+√2ε‖ut‖2L2+α−1β2‖˜Φxt‖2L2. | (3.46) |
In completely the same fashion, we can show that
ddt(‖˜σtt‖2L2+‖utt‖2L2)+α‖utt‖2L2≤(δ+√2ε)(‖utt‖2L2+‖˜σtt‖2L2)+3√2ε(‖utt‖2L2+‖uxt‖2L2+‖˜σtt‖2L2+‖˜σxt‖2L2)+α−1β2‖˜Φxtt‖2L2. | (3.47) |
Step 3. Taking L2 inner product of the third equation in (3.3) with −˜Φxx, we have
ddt(4τA0‖˜Φx‖2L2)+8dA0‖˜Φxx‖2L2+8aA0‖˜Φx‖2L2=−b∫10(˜σ+2ˆσ)˜σ˜Φxxdx≤2b(‖˜σ‖L∞+‖ˆσ‖L∞)‖˜σ‖L2‖˜Φxx‖L2. | (3.48) |
By Cauchy-Schwarz inequality, we update (3.48) as
ddt(4τA0‖˜Φx‖2L2)+8dA0‖˜Φxx‖2L2+8aA0‖˜Φx‖2L2≤b2dA0(‖˜σ‖L∞+‖ˆσ‖L∞)2‖˜σ‖2L2+dA0‖˜Φxx‖2L2≤2b2dA0(‖˜σ‖2L∞+‖ˆσ‖2L∞)‖˜σ‖2L2+dA0‖˜Φxx‖2L2, |
which implies, by (3.7),
ddt(‖˜Φx‖2L2)+7d4τ‖˜Φxx‖2L2+2aτ‖˜Φx‖2L2≤b2τdA20(8A0¯ρ+ε2)‖˜σ‖2L2. | (3.49) |
Similarly, by taking L2 inner product of the third equation in (3.43) with −˜Φxxt, we can show that
ddt(‖˜Φxt‖2L2)+7d4τ‖˜Φxxt‖2L2+2aτ‖˜Φxt‖2L2≤b2τdA20(8A0¯ρ+ε2)‖˜σt‖2L2. | (3.50) |
Moreover, taking ∂t to the third equation in (3.43), then taking L2 inner product of the resulting equation with ˜Φtt, it can be shown that
ddt(‖˜Φtt‖2L2)+7d4τ‖˜Φxtt‖2L2+2aτ‖˜Φtt‖2L2≤b√2τA0‖˜σt‖H1‖˜σt‖L2‖˜Φtt‖L2+b2τdA20(8A0¯ρ+ε2)‖˜σtt‖2L2. | (3.51) |
For the first term on the right-hand side of (3.51), we have
b√2τA0‖˜σt‖H1‖˜σt‖L2‖˜Φtt‖L2≤b√2τA0‖˜σt‖H1‖˜σt‖L2‖˜Φxtt‖L2≤b22τdA20‖˜σt‖2H1‖˜σt‖2L2+d4τ‖˜Φxtt‖2L2, |
where we applied Poincaré inequality. Then we update (3.51) as
ddt(‖˜Φtt‖2L2)+3d2τ‖˜Φxtt‖2L2+2aτ‖˜Φtt‖2L2≤b22τdA20‖˜σt‖2H1‖˜σt‖2L2+b2τdA20(8A0¯ρ+ε2)‖˜σtt‖2L2. | (3.52) |
Step 4. Taking the sum of (3.42), (3.46), and (3.47) gives us
ddt(‖(˜σ,˜σt,˜σtt,u,ut,utt)‖2L2)+α‖(u,ut,utt)‖2L2≤(δ+4√2ε)X(t)+α−1β2‖(˜Φx,˜Φxt,˜Φxtt)‖2L2. | (3.53) |
Taking the sum of (3.49), (3.50) and (3.52), we obtain
ddt(‖(˜Φx,˜Φxt,˜Φtt)‖2L2)+3d2τ‖(˜Φxx,˜Φxxt,˜Φxtt)‖2L2≤b2τdA20(8A0¯ρ+ε2)X(t), | (3.54) |
where we threw away the non-negative terms involving a. Taking the sum of (3.53) and (3.54), and using the definition of X1(t) (c.f. (3.5)), we obtain
ddt(‖˜σ(t)‖2L2+X1(t))+α‖(u,ut,utt)‖2L2+3d2τ‖(˜Φxx,˜Φxxt,˜Φxtt)‖2L2≤(δ+4√2ε)X(t)+b2τdA20(8A0¯ρ+ε2)X(t)+α−1β2‖(˜Φx,˜Φxt,˜Φxtt)‖L2. | (3.55) |
Note that by Poincaré inequality, we have ‖˜Φx‖L2≤‖˜Φxx‖L2 and ‖˜Φxt‖L2≤‖˜Φxxt‖L2. Hence, using the assumption that d>0 is sufficiently large, we update (3.55) as
ddt(‖˜σ(t)‖2L2+X1(t))+α‖(u,ut,utt)‖2L2+dτ‖(˜Φxx,˜Φxxt,˜Φxtt)‖2L2≤(δ+4√2ε)X(t)+b2τdA20(8A0¯ρ+ε2)X(t). | (3.56) |
Again, by Poincaré inequality, we deduce from (3.56) that
ddt(‖˜σ(t)‖2L2+X1(t))+α‖(u,ut,utt)‖2L2+d2τ‖(˜Φx,˜Φxt,˜Φtt,˜Φxx,˜Φxxt,˜Φxtt)‖2L2≤(δ+4√2ε)X(t)+b2τdA20(8A0¯ρ+ε2)X(t). | (3.57) |
Step 5. Taking L2 inner product of the first equation in (3.43) with −˜σ, we obtain
ddt(∫10−˜σ˜σtdx)+‖˜σt‖2L2=12∫10(2ut˜σx+2u˜σxt+˜σtux+˜σuxt+2utˆσx+ˆσuxt)˜σdx. | (3.58) |
For the integral involving the first five integrands on the right-hand side of (3.58), we can show that
|12∫10(2ut˜σx+2u˜σxt+˜σtux+˜σuxt+2utˆσx)˜σdx|≤12(‖˜σ‖L∞+‖ˆσx‖L∞)X(t)≤12(√2ε+δ)X(t). | (3.59) |
For the integral of the last integrand, using integration by parts, we have
|12∫10ˆσuxt˜σdx|=12|∫10(ˆσxut˜σ+ˆσut˜σx)dx|≤δ4X(t)+√A0¯ρ(‖ut‖2L2+‖˜σx‖2L2). | (3.60) |
Similar to (3.25), we can derive the following estimate:
‖˜σx‖2L2≤5A0¯ρ(4‖ut‖2L2+8ε2‖u‖2L2+81δ2‖˜σx‖2L2+4α2‖u‖2L2+4β2‖˜Φx‖2L2). | (3.61) |
Since δ is small, we update (3.61) as
‖˜σx‖2L2≤C(‖ut‖2L2+‖u‖2L2+‖˜Φx‖2L2). | (3.62) |
Substituting (3.62) into (3.60), we obtain
|12∫10ˆσuxt˜σdx|≤δ4X(t)+D1‖(ut,u,˜Φx)‖2L2, | (3.63) |
where the constant D1 depends only on A0, ¯ρ. Substituting (3.59) and (3.63) into (3.58) gives us
ddt(∫10−˜σ˜σtdx)+‖˜σt‖2L2≤(√22ε+34δ)X(t)+D1‖(ut,u,˜Φx)‖2L2. |
Next, taking ∂t to the first equation in (3.43), then taking L2 inner product of the resulting equation with −˜σt, we get
ddt(∫10−˜σt˜σttdx)+‖˜σtt‖2L2=12∫10(2ut˜σx+2u˜σxt+˜σtux+˜σuxt+2utˆσx+ˆσuxt)t˜σtdx. | (3.64) |
For the integral involving the first five integrands on the right-hand side of (3.64), using integration by parts, we can show that
12∫10(2ut˜σx+2u˜σxt+˜σtux+˜σuxt+2utˆσx)t˜σtdx=12∫10(2˜σxutt+4ut˜σxt+ux˜σtt+2˜σtuxt+2ˆσxutt)˜σtdx+12∫10(2u˜σxtt+˜σuxtt)˜σtdx=12∫10(˜σxutt+4ut˜σxt−ux˜σtt+2˜σtuxt+2ˆσxutt)˜σtdx−12∫10(2u˜σtt+˜σutt)˜σxtdx. |
Similar to (3.59), we have
|12∫10(2ut˜σx+2u˜σxt+˜σtux+˜σuxt+2utˆσx)t˜σtdx|≤(‖˜σx‖L∞+‖ux‖L∞+‖˜σt‖L∞+‖ut‖L∞+‖˜σ‖L∞+‖u‖L∞+‖ˆσx‖L∞)X(t)≤(2√3ε+δ)X(t). | (3.65) |
For the integral of the last integrand on the right-hand side of (3.64), we deduce that
|12∫10ˆσuxtt˜σtdx|=12|∫10(ˆσxutt˜σt+ˆσutt˜σxt)dx|≤δ4X(t)+√A0¯ρ(‖utt‖2L2+‖˜σxt‖2L2). |
According to (3.27) and (3.28), we know
‖˜σxt‖2L2≤C‖(utt,ut,˜Φxt)‖2L2+Cε‖uxt‖2L2+C(ε+δ)‖˜σt‖2L2 | (3.66) |
and
‖uxt‖2L2≤C‖˜σtt‖2L2+C(ε+δ)‖(ut,˜σt)‖2L2+Cε‖˜σxt‖2L2. | (3.67) |
Substituting (3.67) into (3.66), we obtain
‖˜σxt‖2L2≤C‖(utt,ut,˜Φxt)‖2L2+Cε‖˜σtt‖2L2+C(ε+δ)‖˜σt‖2L2+Cε‖˜σxt‖2L2. | (3.68) |
When ε is sufficiently small, we update (3.68) as
‖˜σxt‖2L2≤C‖(utt,ut,˜Φxt)‖2L2+Cε‖˜σtt‖2L2+C(ε+δ)‖˜σt‖2L2. | (3.69) |
Substituting (3.65) and (3.69) into (3.64), we arrive at
ddt(∫10−˜σt˜σttdx)+‖˜σtt‖2L2≤(2√3ε+54δ)X(t)+C‖(utt,ut,˜Φxt)‖2L2+Cε‖˜σtt‖2L2+C(ε+δ)‖˜σt‖2L2. | (3.70) |
When ε and δ are sufficiently small, we update (3.70) as
ddt(∫10−(˜σ˜σt+˜σt˜σtt)dx)+12‖(˜σt,˜σtt)‖2L2≤(2√3ε+54δ)X(t)+D2‖(utt,ut,u,˜Φx,˜Φxt)‖2L2. | (3.71) |
We observe from (3.27), (3.28), and (3.62) that when ε and δ are sufficiently small, the constant D2 depends only on A0, ¯ρ, α, and β.
Step 6. Note that the dissipations in (3.57) and (3.71) contain a quantity that is equivalent to X1(t) defined in (3.5). Hence, we shall make a coupling of (3.57) and (3.71) to close the overall energy estimates to capture the global dynamics of the perturbed solution. However, direct summation of (3.57) and (3.71) is problematic, as some leading terms are standing on the right-hand side of (3.71) and the summation of the terms inside the time derivatives does not cover the total H2-norm of ˜σ. To overcome such a technical difficulty, we shall require d>0 to be large enough, such that
4τD2d−1≤2, | (3.72) |
and let
χ=max{2, 2D2α−1}. | (3.73) |
Dividing (3.71) by χ, we get
ddt(∫10−(˜σ˜σt+˜σt˜σtt)χdx)+12χ‖(˜σt,˜σtt)‖2L2≤1χ(2√3ε+54δ)X(t)+D2χ‖(utt,ut,u,˜Φx,˜Φxt)‖2L2. | (3.74) |
Taking the sum of (3.57) and (3.74) gives us
ddt(V(t))+W(t)≤θX(t), | (3.75) |
where
V(t)≡‖˜σ(t)‖2L2+X1(t)−∫10(˜σ˜σt+˜σt˜σtt)χdx,W(t)≡α‖(u,ut,utt)‖2L2+d2τ‖(˜Φx,˜Φxt,˜Φtt,˜Φxx,˜Φxxt,˜Φxtt)‖2L2+12χ‖(˜σt,˜σtt)‖2L2,−D2χ‖(utt,ut,u,˜Φx,˜Φxt)‖2L2,θ≡δ+4√2ε+b2τdA20(8A0¯ρ+ε2)X(t)+1χ(2√3ε+54δ). |
Note that under (3.72) and (3.73),
D2χ‖(utt,ut,u,˜Φx,˜Φxt)‖2L2=D2χ‖(utt,ut,u)‖2L2+D2χ‖(˜Φx,˜Φxt)‖2L2≤α2‖(utt,ut,u)‖2L2+d4τ‖(˜Φx,˜Φxt)‖2L2. | (3.76) |
Hence, it follows from the definition of X1(t) that
W(t)≥α2‖(u,ut,utt)‖2L2+d4τ‖(˜Φx,˜Φxt,˜Φtt)‖2L2+12χ‖(˜σt,˜σtt)‖2L2+d2τ‖(˜Φxx,˜Φxxt,˜Φxtt)‖2L2≥D3X1(t)+d2τ‖(˜Φxx,˜Φxxt,˜Φxtt)‖2L2, | (3.77) |
where
D3=min{α2, d4τ, 12χ}. |
Since χ≥2, from the definition of V(t) we see that
V(t)≥‖(˜σ,˜σt,˜σtt)‖2L2−14(‖˜σ‖2L2+2‖˜σt‖2L2+‖˜σtt‖2L2)+‖(u,ut,utt,˜Φx,˜Φxt,˜Φtt)‖2L2=14(3‖˜σ‖2L2+2‖˜σt‖2L2+3‖˜σtt‖2L2)+‖(u,ut,utt,˜Φx,˜Φxt,˜Φtt)‖2L2≥12‖(˜σ,˜σt,˜σtt,u,ut,utt,˜Φx,˜Φxt,˜Φtt)‖2L2=12‖˜σ‖2L2+12X1(t). | (3.78) |
Using Lemma 3.1 and (3.77), we update (3.75) as
ddt(V(t))+D3X1(t)+d2τ‖(˜Φxx,˜Φxxt,˜Φxtt)‖2L2≤θD0X1(t), |
Since δ→0 as d→∞, from the definition of θ we see that θ→0 as ε→0 and d→∞. Hence, when ε is sufficiently small and d is sufficiently large, it holds that
ddt(V(t))+D32X1(t)+d2τ‖(˜Φxx,˜Φxxt,˜Φxtt)‖2L2≤0. | (3.79) |
Integrating (3.79) with respect to t, we obtain
V(t)+∫t0(D32X1(t)+d2τ‖(˜Φxx,˜Φxxt,˜Φxtt)(t)‖2L2)≤V(0). | (3.80) |
Since, according to (3.78) and Lemma 3.1, 12X1(t)≤V(t)≤X(t)≤D0X1(t), the estimate (3.80) yields
‖(˜σ,u,˜Φ)(t)‖22+∫t0(‖(˜σ,u,˜Φ)(τ)‖22+‖(˜Φxxt,˜Φxtt)(τ)‖2L2)dτ≤D4,∀t>0, | (3.81) |
where the constant D4 is independent of t. Moreover, using the third equation in (3.3) and Poincaré inequality, we can show that
‖˜Φxxx‖2L2≲‖(˜Φxt,˜Φx,˜σx,˜σ)‖2L2, | (3.82) |
‖˜Φxxxx‖2L2≲‖(˜Φxxt,˜Φxx,˜σxx,˜σx,˜σ)‖2L2≲‖(˜Φtt,˜Φt,˜σt)‖2L2, | (3.83) |
‖˜Φxxxt‖2L2≲‖(˜Φxtt,˜Φxt,˜σxt,˜σt)‖2L2. |
Hence, it follows from (3.81), (3.82), and (3.83) that
‖(˜σ,u)(t)‖22+2∑k=0‖(∂kt˜Φ)(t)‖2H4−2k+∫t0(‖(˜σ,u)(τ)‖22+‖˜Φ(τ)‖2H4+‖˜Φt(τ)‖2H3+‖˜Φtt(τ)‖2H1)dτ≤D5,∀t>0, |
for some constant D5 which is independent of t.
To derive the exponential decaying of the perturbation, we note that by dropping the non-negative term d2τ‖(˜Φxx,˜Φxxt,˜Φxtt)(t)‖2L2 from the left-hand side of (3.79), and using the equivalency of V(t) and X1(t), it holds that
ddt(V(t))+D32D0V(t)≤0, |
which yields the exponential decaying of V(t), and hence of X(t). Moreover, the exponential decaying of ∑2k=0‖(∂kt˜Φ)(t)‖2H4−2k follows from the decaying of X(t) and (3.82)–(3.83). This completes the proof of Lemma 3.2.
In this subsection, we mainly consider the case of τ=0 in (3.3):
{2˜σt+2u˜σx+˜σux+2uˆσx+ˆσux=0,x∈I, t>0,2ut+2uux+˜σ˜σx+˜σˆσx+ˆσ˜σx=−2αu+2β˜Φx,x∈I, t>0,8dA0˜Φxx−8aA0˜Φ+b(˜σ+2ˆσ)˜σ=0,x∈I, t>0;(˜σ,u)(x,0)=(2√2A0ρ0−2√2A0ˆρ,u0)(x),x∈I;u|x=0,x=1=0,˜Φ|x=0,x=1=0,t>0. | (3.84) |
In this case, instead of X(t) defined by (3.4), we let
Y(t)≡‖(˜σ,u)(t)‖22, |
and derive the a priori estimates based on the assumptions that (1) Y(t) is sufficiently small within the lifespan of the local solution, and (2) the diffusion coefficient d is sufficiently large.
First, by using the third equation in (3.85), we can modify the proof of Lemma 3.1 to get the qualitative equivalency of Y(t) and ‖(˜σt,˜σtt,u,ut,utt)(t)‖2L2. Indeed, using Sobolev embedding, (3.7), (3.16) and Poincaré inequality, it can be shown that
8dA0‖˜Φx‖2L2+8aA0‖˜Φ‖2L2≤b‖˜σ+ˆσ‖L∞‖˜σ‖L2‖˜Φ‖L2≤5b√A0¯ρ‖˜σ‖L2‖˜Φx‖L2≤Cd‖˜σ‖2L2+4dA0‖˜Φx‖2L2, |
which implies
‖˜Φx‖2L2≤Cd2‖˜σ‖2L2≤Cd2‖˜σx‖2L2, | (3.85) |
where we also used (3.17). Substituting (3.85) into (3.25), we obtain
‖˜σx‖2L2≤C(‖ut‖2L2+ε2‖ux‖2L2+δ2‖˜σx‖2L2+α2‖u‖2L2+d−2‖˜σx‖2L2). | (3.86) |
Taking the sum of (3.86) and (3.22) gives us
‖˜σx‖2L2+‖ux‖2L2≤C[‖ut‖2L2+‖u‖2L2+‖˜σt‖2L2+(ε2+δ2)‖ux‖2L2+(ε2+δ2+d−2)‖˜σx‖2L2]. |
Hence, when ε and δ are sufficiently small and d is sufficiently large, such that the coefficients in front of ‖ux‖2L2 and ‖˜σx‖2L2 are smaller than 12, there holds that
‖˜σx‖2L2+‖ux‖2L2≤C(‖ut‖2L2+‖u‖2L2+‖˜σt‖2L2). | (3.87) |
Similarly, it follows from the elliptic equation that
‖˜Φxt‖2L2≤Cd2‖˜σt‖2L2and‖˜Φxx‖2L2≤Cd2‖˜σx‖2L2, | (3.88) |
by using which we can show that
‖uxt‖2L2+‖˜σxt‖2L2≤C‖(utt,˜σtt,ut,˜σt)‖2L2, |
and
‖uxx‖2L2+‖˜σxx‖2L2≤C‖(utt,˜σtt,ut,˜σt,u)‖2L2. |
Hence,
Y(t)≅Y1(t)≡‖(˜σt,˜σtt,u,ut,utt)(t)‖2L2. |
Regarding Lemma 3.2, it follows from the elliptic equation that
‖˜Φxtt‖L2≤ε2d2‖˜σt‖2L2+1d2‖˜σtt‖2L2. | (3.89) |
Similar to (3.53), by using (3.85), (3.87), (3.88), and (3.89), we can derive the following estimate:
ddt(‖(˜σ,˜σt,˜σtt,u,ut,utt)‖2L2)+α‖(u,ut,utt)‖2L2≲O(ε,δ,d−1)Y(t). | (3.90) |
Similar to (3.71) and using the modified estimates in this section, it can be shown that
ddt(∫10−(˜σ˜σt+˜σt˜σtt)dx)+12‖(˜σt,˜σtt)‖2L2≤O(ε,δ,d−1)Y(t)+O(1)‖(utt,ut,u)‖2L2. | (3.91) |
By coupling (3.90) and (3.91) together, and using the smallness of ε, δ and the largeness of d, we can derive the exponential decaying of Y1(t), and hence equivalently of Y(t). Moreover, it follows from the elliptic equation that
2∑k=0‖(∂kt˜Φ)(t)‖2H4−k≲‖˜σ(t)‖22, |
which yields the exponential decaying of ˜Φ in the corresponding topology. Thus, the proof of Theorem 2.2 is completed.
The authors are grateful to four referees for their valuable comments, which greatly improved the exposition of our paper. H.Y. Peng was partially supported by the National Natural Science Foundation of China (No. 12271112). The research of K. Zhao was partially supported by the Simons Foundation's Collaboration Grant for Mathematicians (No. 413028).
The authors declare there is no conflict of interest.
[1] | J. D. Murray, Mathematical biology I: An introduction. vol. 17 of Interdisciplinary Applied Mathematics, Springer-Verlag, New York, third ed., 2002. https://link.springer.com/book/10.1007/b98868 |
[2] |
S. G. Li, K. Muneoka, Cell migration and chick limb development: chemotactic action of FGF-4 and the AER, Dev. Biol., 211 (1999), 335–347. https://doi.org/10.1006/dbio.1999.9317 doi: 10.1006/dbio.1999.9317
![]() |
[3] |
P. Carmeliet, Mechanisms of angiogenesis and arteriogenesis, Nat. Med., 6 (2000), 389–395. https://doi.org/10.1038/74651 doi: 10.1038/74651
![]() |
[4] |
G. Helmlinger, M. Endo, N. Ferrara, L. Hlatky, R. Jain, Formation of endothelial cell networks, Nature, 405 (2000), 139–141. https://doi.org/10.1038/35012132 doi: 10.1038/35012132
![]() |
[5] |
A. Gamba, D. Ambrosi, A. Coniglio, A de Candia, S. Di Talia, E. Giraudo, et al., Percolation, morphogenesis, and Burgers dynamics in blood vessels formation, Phys. Rev. Lett., 90 (2003), 118101–118104. https://doi.org/10.1103/PhysRevLett.90.118101 doi: 10.1103/PhysRevLett.90.118101
![]() |
[6] |
D. Ambrosi, F. Bussolino, L. Preziosi, A review of vasculogenesis models, J. Theoret. Med., 6 (2005), 1–19. https://doi.org/10.1080/1027366042000327098 doi: 10.1080/1027366042000327098
![]() |
[7] | H. Y. Jin, Z. A. Wang, Global stabilization of the full attraction-repulsion Keller-Segel system, Discrete Contin. Dyn. Syst., 40 (2020), 3509–3527. https://doi.org/10.3934/dcds.2020027 |
[8] |
P. Liu, J. P. Shi, Z. A. Wang, Pattern formation of the attraction-repulsion Keller-Segel system, Discrete Contin. Dyn. Syst.-Ser. B, 18 (2013), 2597-2625. https://doi.org/10.3934/dcdsb.2013.18.2597 doi: 10.3934/dcdsb.2013.18.2597
![]() |
[9] |
Z. A. Wang, Z. Xiang, P. Yu, Asymptotic dynamics on a singular chemotaxis system modeling onset of tumor angiogenesis, J. Differ. Equ., 260 (2016), 2225–2258. https://doi.org/10.1016/j.jde.2015.09.063 doi: 10.1016/j.jde.2015.09.063
![]() |
[10] |
P. H. Chavanis, C. Sire, Kinetic and hydrodynamic models of chemotactic aggregation, Physica A, 384 (2007), 199–222. https://doi.org/10.1016/j.physa.2007.05.069 doi: 10.1016/j.physa.2007.05.069
![]() |
[11] |
M. Di Francesco, D. Donatelli, Singular convergence of nonlinear hyperbolic chemotaxis systems to Keller-Segel type models, Discrete Contin. Dyn. Syst. Ser. B, 13 (2010), 79–100. https://doi.org/10.3934/dcdsb.2010.13.79 doi: 10.3934/dcdsb.2010.13.79
![]() |
[12] |
R. Natalini, M. Ribot, M. Twarogowska, A numerical comparison between degenerate parabolic and quasilinear hyperbolic models of cell movements under chemotaxis, J. Sci. Comput., 63 (2015), 654–677. https://doi.org/10.1007/s10915-014-9909-y doi: 10.1007/s10915-014-9909-y
![]() |
[13] |
R. Kowalczyk, A. Gamba, L. Preziosi, On the stability of homogeneous solutions to some aggregation models, Discrete Contin. Dyn. Syst. Ser. B, 4 (2004), 203–220. https://doi.org/10.3934/DCDSB.2004.4.203 doi: 10.3934/DCDSB.2004.4.203
![]() |
[14] | C. Di Russo, Analysis and numerical approximations of hydrodynamical models of biological movements, Rend. Mat. Appl., 32 (2012), 117–367. http://hdl.handle.net/2307/4247 |
[15] |
C. Di Russo, A. Sepe, Existence and asymptotic behavior of solutions to a quasi-linear hyperbolic-parabolic model of vasculogenesis, SIAM J. Math. Anal., 45 (2013), 748–776. https://doi.org/10.1137/110858896 doi: 10.1137/110858896
![]() |
[16] |
Q. Q. Liu, H. Y. Peng, Z. A. Wang, Asymptotic stability of diffusion waves of a quasi-linear hyperbolic-parabolic model for vasculogenesis, SIAM J. Math. Anal., 54 (2022), 1313–1346. https://doi.org/10.1137/21M1418150 doi: 10.1137/21M1418150
![]() |
[17] |
Q. Q. Liu, H. Y. Peng, Z. A. Wang, Convergence to nonlinear diffusion waves for a hyperbolic-parabolic chemotaxis system modelling vasculogenesis, J. Differ. Equ., 314 (2022), 251–286. https://doi.org/10.1016/j.jde.2022.01.021 doi: 10.1016/j.jde.2022.01.021
![]() |
[18] | T. Crin-Barat, Q. Y. He, L. Y. Shou, The hyperbolic-parabolic chemotaxis system modelling vasculogenesis: global dynamics and relaxation limit, arXive, (2022), arXiv: 2201.06512v1. https://doi.org/10.48550/arXiv.2201.06512 |
[19] |
F. Berthelin, D. Chiron, M. Ribot, Stationary solutions with vacuum for a one-dimensional chemotaxis model with nonlinear pressure, Commun. Math. Sci., 14 (2016), 147–186. https://doi.org/10.4310/CMS.2016.v14.n1.a6 doi: 10.4310/CMS.2016.v14.n1.a6
![]() |
[20] |
J. Carrillo, X. Chen, Q. Wang, Z. Wang, L. Zhang, Phase transitions and bump solutions of the Keller-Segel model with volume exclusion, SIAM J. Appl. Math., 80 (2020), 232–261. https://doi.org/10.1137/19M125827 doi: 10.1137/19M125827
![]() |
[21] |
G. Y. Hong, H. Y. Peng, Z. A. Wang, C. J. Zhu, Nonlinear stability of phase transition steady states to a hyperbolic–parabolic system modeling vascular networks, J. London Math. Soc., 103 (2021), 1480–1514. https://doi.org/10.1112/jlms.12415 doi: 10.1112/jlms.12415
![]() |
[22] |
R. H. Pan, K. Zhao, The 3D compressible Euler equations with damping in a bounded domain, J. Differ. Equ., 246 (2009), 581–596. https://doi.org/10.1016/j.jde.2008.06.007 doi: 10.1016/j.jde.2008.06.007
![]() |
[23] | S. Schochet, The compressible Euler equations in a bounded domain: Existence of solutions and the incompressible limit, Comm. Math. Phys., 104 (1986), 49–75. https://link.springer.com/article/10.1007/BF01210792 |