Abstract

We study the Navier-Stokes-Nernst-Planck-Poisson system modeling the flow of electrohydrodynamics. For small initial data, the global existence, uniqueness, and asymptotic stability as time goes to infinity of self-similar solutions to the Cauchy problem of this system posed in the whole three dimensional space are proved in the function spaces of pseudomeasure type.

1. Introduction

In this paper, we consider the Cauchy problem for the (normalized) Navier-Stokes-Nernst-Planck-Poisson system which governs the hydrodynamic transport of binary diffusion charge densities as follows (see [1]): 𝜕𝑡𝐮Δ𝐮+(𝐮)𝐮+𝑝=Δ𝜙𝜙in3×(0,),(1.1)𝐮=0in3×(0,),(1.2)𝜕𝑡𝑣+𝐮𝑣=(𝑣𝑣𝜙)in3×(0,),(1.3)𝜕𝑡𝑤+𝐮𝑤=(𝑤+𝑤𝜙)in3×(0,),(1.4)Δ𝜙=𝑣𝑤in3×(0,),(1.5)𝐮(𝑥,0)=𝐮0(𝑥),𝑣(𝑥,0)=𝑣0(𝑥),𝑤(𝑥,0)=𝑤0(𝑥)in3.(1.6) Equations (1.1) and (1.2) are the momentum conservation and the mass conservation equations of incompressible flow. 𝐮=𝐮(𝑥,𝑡)=(𝑢1(𝑥,𝑡),𝑢2(𝑥,𝑡),𝑢3(𝑥,𝑡))3,𝑝=𝑝(𝑥,𝑡) and 𝜙=𝜙(𝑥,𝑡) denote, respectively, the velocity field, the pressure of the fluid, and the electrostatic potential, and the right-hand side term in (1.1) is the Lorentz force caused by the charged particles. Equations (1.3) and (1.4) model the balance between diffusion and convective transport of charge densities by the flow and the electric fields. 𝑣=𝑣(𝑥,𝑡) and 𝑤=𝑤(𝑥,𝑡) denote the charge densities of the negatively and positively charged species, respectively, hence the sign difference in front of the convective term in either equation. Equation (1.5) is the Poisson equation for the electrostatic potential 𝜙, and the right-hand side is the net charge density. For simplicity, we have chosen the fluid density, viscosity, charge mobility and dielectric constant to be unit.

To start with, let us recall two special cases of (1.1)–(1.6). In the case that the flow is charge free, that is, 𝑣=𝑤=𝜙=0, the system (1.1)–(1.6) reduces into the well-known Navier-Stokes equations: 𝜕𝑡𝐮Δ𝐮+(𝐮)𝐮+𝑃=0in3×(0,),𝐮=0in3×(0,),𝐮(𝑥,0)=𝐮0(𝑥)in3.(1.7) After the pioneering work [2], the Navier-Stokes equations (1.7) has drawn great attention of researchers for many years and a huge number of works can be found from the literature, for example, [38] and the references therein. If, on the other hand, the velocity field 𝐮 is identically vanishing, then (1.1)–(1.6) reduces into the following Nernst-Planck-Poisson equations which was formulated by W. Nernst and M. Planck at the end of the nineteenth century as a basic model for the diffusion of ions in an electrolytes (cf. [9]): 𝜕𝑡𝑣=(𝑣𝑣𝜙)in3𝜕×(0,),𝑡𝑤=(𝑤+𝑤𝜙)in3×(0,),Δ𝜙=𝑣𝑤in3×(0,),𝑣(𝑥,0)=𝑣0(𝑥),𝑤(𝑥,0)=𝑤0(𝑥)in3.(1.8) In some literatures, it is also called the Debye-Hückel system (cf. [10]). It has drawn much attention of analysts during the past twenty years, and some works concerning existence of (large) weak solutions, (small) mild solutions, convergence rate estimates to stationary solutions of time-dependent solutions and other related topics can be found from the literature, cf., for example, [1016] and the references therein.

In 2002, Jerome [17] proved that the system (1.1)–(1.6) has a unique local smooth solution for smooth initial data where he verified the local existence in Kato’s semigroup framework. In [18], by using the energy inequalities and the Schauder fixed point theorem, Schmuck established global existence of weak solutions to the system (1.1)–(1.6) in a bounded domain Ω with homogeneous Neumann boundary conditions with initial data 𝐮0[𝐿2(Ω)]𝑛 and 𝑣0, 𝑤0𝐿(Ω) for 𝑛=2,3. In [19], Ryham studied existence, uniqueness, and regularity of weak solutions of (1.1)–(1.6) in a bounded domain with no-flux boundary conditions for general 𝐿2 initial data in 𝑛=2 and for small initial data in 𝑛=3. The convergence to the stationary solution with a rate is also established in [19]. In our recent work [20], by using the 𝐿𝑝-𝐿𝑞 estimates of the heat semigroup and the classical Hardy-Littlewood-Sobolev inequality, we established local well-posedness of (1.1)–(1.6) in critical and subcritical Lebesgue spaces (i.e., 𝐮0[𝐿𝑞(𝑛)]𝑛, and 𝑣0,𝑤0𝐿𝑝(𝑛), 𝑛𝑞<, 𝑛/2𝑝<𝑛) and global well-posedness for small initial data in critical Lebesgue spaces (i.e., 𝐮0[𝐿𝑛(𝑛)]𝑛,𝑣0,𝑤0𝐿𝑛/2(𝑛) and 𝐮0[𝐿𝑛]𝑛+𝑣0𝐿𝑛/2+𝑤0𝐿𝑛/2 is sufficiently small). For computational simulations of the problem (1.1)–(1.6), see [2123].

The most important results stated in this paper are theorems on the global existence, uniqueness and asymptotic stability as time goes to infinity of self-similar solutions to the system (1.1)–(1.6) in the functional spaces of pseudomeasure-type. Let us recall that the solution (𝐮,𝑝,𝑣,𝑤,𝜙) of the system (1.1)–(1.6) is called a self-similar solution if it satisfies the following scaling invariant property: 𝐮(𝑡,𝑥)=𝐮𝜆(𝑥,𝑡), 𝑝(𝑡,𝑥)=𝑝𝜆(𝑥,𝑡), 𝑣(𝑡,𝑥)=𝑣𝜆(𝑥,𝑡), 𝑤(𝑡,𝑥)=𝑤𝜆(𝑥,𝑡) and 𝜙(𝑡,𝑥)=𝜙𝜆(𝑥,𝑡) for all 𝜆>0, 𝑥3, and 𝑡0, where 𝐮𝜆(𝑥,𝑡)=𝜆𝐮𝜆𝑥,𝜆2𝑡,𝑝𝜆(𝑥,𝑡)=𝜆2𝑝𝜆𝑥,𝜆2𝑡,𝑣𝜆(𝑥,𝑡)=𝜆2𝑣𝜆𝑥,𝜆2𝑡,𝑤𝜆(𝑥,𝑡)=𝜆2𝑤𝜆𝑥,𝜆2𝑡,𝜙𝜆(𝑥,𝑡)=𝜙𝜆𝑥,𝜆2𝑡.(1.9) It is clear that if (𝐮,𝑝,𝑣,𝑤,𝜙) is a solution of (1.1)–(1.5) with initial data (1.6), then, for each 𝜆>0, (𝐮𝜆,𝑝𝜆,𝑣𝜆,𝑤𝜆,𝜙𝜆) also solves (1.1)–(1.5) with initial data 𝐮0,𝜆(𝑥)=𝜆𝐮0(𝜆𝑥),𝑣0,𝜆(𝑥)=𝜆2𝑣0(𝜆𝑥),𝑤0,𝜆(𝑥)=𝜆2𝑤0(𝜆𝑥).(1.10) Apparently such initial data do not belong to any Lebesgue and Sobolev spaces due to their strong singularity at 𝑥=0 as well as slow decay as |𝑥|. In [24], the authors found explicit formulas for a one-parameter family of stationary solutions of the three-dimensional Navier-Stokes equations with zero external force; these solutions are global but not smooth, more precisely, they are singular at the origin with a singularity of the kind 1/|𝑥| for all time. Note that 1/|𝑥|𝐿3(3) but 1/|𝑥|𝒫2 (see (1.16) below for definition of this functional space). Similar phenomenon also appeared for the Nernst-Planck-Poisson equations; see [25]. This is the reason why we consider the system (1.1)–(1.6) in the pseudomeasure-type spaces. By a standard contraction argument, we establish global existence of solution for small initial data. It is worth pointing out that this solution is unique in a ball of the functional spaces in which the existence of solutions is going to be obtained. To overcome this restrictive condition, we establish a stability result in terms of a perturbation of initial data, which allow us to give a complete answer to the uniqueness problem of solution. Moreover, we establish the asymptotic stability of self-similar solutions as time goes to infinity. Here, we refer the reader to see [2628] and the references cited there for more details related to the Navier-Stokes equations with measures as initial data.

The self-similar solution is related to an asymptotic behavior, for large time, of global solution to the system (1.1)–(1.6), and we could characterize the self-similar condition in the following way. Here, we disregard the functions 𝑝 and 𝜙 because when 𝐮, 𝑣, and 𝑤 are determined, 𝑝 and 𝜙 can be easily obtained from (1.2) and (1.5). A vector function (𝐮,𝑣,𝑤) has the self-similar property to the system (1.1)–(1.6) if and only if there exists a vector function (𝐔,𝑉,𝑊) such that 𝐮(𝑥,𝑡)=𝐔(𝑥/𝑡)/𝑡, 𝑣(𝑥,𝑡)=𝑉(𝑥/𝑡)/𝑡 and 𝑤(𝑥,𝑡)=𝑊(𝑥/𝑡)/𝑡 for all 𝑥3 and 𝑡>0. In fact, when (𝐔,𝑉,𝑊) exists, these last equalities give the definition for (𝐮,𝑣,𝑤), and it is straightforward to see that it is self-similar. Conversely, when the self-similar solution (𝐮,𝑣,𝑤) is given, we define 𝐔(𝑥)=𝐮(𝑥,1), 𝑉(𝑥)=𝑣(𝑥,1) and 𝑊(𝑥)=𝑤(𝑥,1) for all 𝑥3. Then, the self-similar condition on (𝐮,𝑣,𝑤) turns out the expected equality between (𝐮,𝑣,𝑤) and (𝐔,𝑉,𝑊) by choosing 𝜆=1/𝑡.

Now, as a standard practice, we can reformulate the problem (1.1)–(1.6) into a system of integral equations. To this end, we first solve (1.5) to get 𝜙 as a functional of 𝑤𝑣: 𝜙(𝑥)=(Δ)1(𝑤𝑣)(𝑥)=1||𝜉||2(𝑤(𝜉)𝑣(𝜉))(𝑥)(1.11) in the distributional sense, where and 1 denote the Fourier transform and the inverse Fourier transform, respectively. Next, it is convenient to eliminate the pressure 𝑝 by applying the Leray projector to both sides of (1.1), by (1.11), (1.1) and (1.2) can be transformed into the following equations: 𝜕𝐮𝜕𝑡Δ𝐮+(𝐮)𝐮=(𝑣𝑤)(Δ)1.(𝑤𝑣)(1.12) Recalling that is given formally by the formula =𝐼+(Δ)1div; that is, is the 3×3 matrix pseudo-differential operator in 3 with the symbol ()(𝜉)=(𝛿𝑗𝑘(𝜉𝑗𝜉𝑘)/|𝜉|2)3𝑗,𝑘=1, where 𝐼 denotes the unit operator and 𝛿𝑗𝑘 is the Kronecker symbol. It is obvious that all these components are bounded, that is, sup𝜉3{0}||||()(𝜉)<.(1.13) Finally, by the well-known Duhamel principle, we see that the problem (1.1)–(1.6) can be further reduced into the following system of integral equations: 𝐮=𝑒𝑡Δ𝐮0+𝑡0𝑒(𝑡𝜏)Δ𝐺1(𝐮(𝜏),𝑣(𝜏),𝑤(𝜏))𝑑𝜏,𝑣=𝑒𝑡Δ𝑣0+𝑡0𝑒(𝑡𝜏)Δ𝐺2(𝐮(𝜏),𝑣(𝜏),𝑤(𝜏))𝑑𝜏,𝑤=𝑒𝑡Δ𝑤0+𝑡0𝑒(𝑡𝜏)Δ𝐺3(𝐮(𝜏),𝑣(𝜏),𝑤(𝜏))𝑑𝜏,(1.14) where 𝑒𝑡Δ is the heat operator which can be regarded as the convolution with the heat kernel 𝐺(𝑥,𝑡)=(4𝜋𝑡)3/2exp(|𝑥|2/4𝑡), and 𝐺1(𝐮,𝑣,𝑤)=(𝐮𝐮)+(𝑣𝑤)(Δ)1,𝐺(𝑤𝑣)2(𝐮,𝑣,𝑤)=(𝐮𝑣)𝑣(Δ)1,𝐺(𝑤𝑣)3(𝐮,𝑣,𝑤)=(𝐮𝑤)+𝑤(Δ)1.(𝑤𝑣)(1.15) Later on we will work on this system of integral equations instead of (1.1)–(1.6).

Before giving the explicit meaning of solutions to the system (1.1)–(1.6), we define the functional spaces relevant to the existence of solutions of (1.1)–(1.6). Let us first define 𝒫𝑎=𝑓𝒮3𝑓𝐿1loc3,𝑓𝒫𝑎=esssup𝜉3||𝜉||𝑎||||,𝑓(𝜉)<(1.16) where 𝑎0 is a given parameter. Since 𝒫𝑎 is not separable and the heat semigroup {𝑒𝑡Δ}𝑡0 is not strongly continuous on this space but only weakly continuous, we will use the notation 𝒞𝑤 to denote by the space of functions which are weakly continuous as distributions with respect to 𝑡. Next, we construct the solution (𝐮,𝑣,𝑤) of the system (1.1)–(1.6) with the velocity 𝐮 in the space 𝒳=𝒞𝑤[0,),𝒫23(1.17) equipped with the norm 𝐮𝒳=sup𝑡0𝐮(𝑡)[𝒫2]3, and the components 𝑣 and 𝑤 in the space 𝒴𝑎=𝑣𝒞𝑤[0,),𝒫1sup𝑡>0𝑡(𝑎1)/2𝑣(𝑡)𝒫𝑎<(1.18) equipped with norm 𝑣𝒴𝑎=sup𝑡0𝑣(𝑡)𝒫1+sup𝑡>0𝑡(𝑎1)/2𝑣(𝑡)𝒫𝑎, where 𝑎 is a given parameter satisfying 1<𝑎<2 in the whole paper. For simplicity, we denote 𝒴𝑎,1 and 𝒴𝑎,2 for each part in the norm of 𝒴𝑎, and the product of Banach spaces 𝒳×𝒴×𝒵 will be equipped with the norm (𝑢,𝑣,𝑤)𝒳×𝒴×𝒵=max{𝑢𝒳,𝑣𝒴,𝑤𝒵}.

Remark 1.1. Let 𝑓𝒮(3)𝐿1loc(3); for a positive parameter 𝜆, we denote 𝑓𝜆(𝑥)=𝑓(𝜆𝑥). It is easy to verify that, for each 𝜆>0, 𝑓𝜆(𝜉)=𝜆3𝜆(𝑓)1𝜉,𝑓(𝜆)𝒫𝑎=𝜆𝑎3𝑓𝒫𝑎.(1.19) Hence, the norms in 𝒳 and 𝒴𝑎 are invariant under the scaling (1.9).

Definition 1.2. The solution of (1.1)–(1.6) is a vector function  (𝐮,𝑣,𝑤)  with components satisfying, for some 0<𝑇, 𝐮𝒞𝑤[0,𝑇),𝒫23,𝑣,𝑤𝒞𝑤[0,𝑇),𝒫1,(1.20) and the following equalities hold for all 0𝑡<𝑇: (𝐮)(𝜉,𝑡)=𝑒𝑡|𝜉|2𝐮0(𝜉)𝑡0𝑒(𝑡𝜏)|𝜉|2+()(𝜉)𝑖𝜉(𝐮𝐮)(𝜉,𝜏)𝑑𝜏𝑡0𝑒(𝑡𝜏)|𝜉|2()(𝜉)(𝑣𝑤)(Δ)1(𝑤𝑣)(𝜉,𝜏)𝑑𝜏,(𝑣)(𝜉,𝑡)=𝑒𝑡|𝜉|2𝑣0(𝜉)𝑡0𝑒(𝑡𝜏)|𝜉|2𝑖𝜉(𝐮𝑣)(𝜉,𝜏)𝑑𝜏𝑡0𝑒(𝑡𝜏)|𝜉|2(𝑖𝜉𝑣Δ)1((𝑤𝑣)𝜉,𝜏)𝑑𝜏,(𝑤)(𝜉,𝑡)=𝑒𝑡|𝜉|2𝑤0(𝜉)𝑡0𝑒(𝑡𝜏)|𝜉|2+𝑖𝜉(𝐮𝑤)(𝜉,𝜏)𝑑𝜏𝑡0𝑒(𝑡𝜏)|𝜉|2𝑖𝜉𝑤(Δ)1(𝑤𝑣)(𝜉,𝜏)𝑑𝜏.(1.21)

Now, we state the main results of this paper as follows.

Theorem 1.3. Let 𝐮0[𝒫2]3, 𝐮0=0, and 𝑣0, 𝑤0𝒫1. There exists a constant 𝜀 such that if (𝐮0,𝑣0,𝑤0)[𝒫2]3×[𝒫1]2𝜀, then the system (1.1)–(1.6) has a global solution (𝐮,𝑣,𝑤) in the space 𝒳×[𝒴𝑎]2, and this is the unique solution under the condition (𝐮,𝑣,𝑤)𝒳×[𝒴𝑎]22𝜂𝜀,(1.22) where 0<𝜀<1/(4𝜂2) and the constant 𝜂 is defined by (2.23). Moreover, the solution depends continuously on initial data in the following sense: if ̃𝐮(0,̃𝑣0,𝑤0)[𝒫2]3×[𝒫1]2𝜀, then one denotes by (̃̃𝐮,𝑣,𝑤) the unique solution of (1.1)–(1.6) with initial data (̃𝐮0,̃𝑣0,𝑤0), and ̃̃(𝐮,𝑣,𝑤)𝒳×[𝒴𝑎]22𝜂𝜀, then, one has ̃̃(𝐮𝐮,𝑣𝑣,𝑤𝑤)𝒳×[𝒴𝑎]2𝐮𝐶(𝜀,𝜂)0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0[𝒫2]3×[𝒫1]2,(1.23) where 𝐶(𝜀,𝜂)=𝜂/(14𝜀𝜂2).

Remark 1.4. In Theorem 1.3, we obtained only a partial answer to the uniqueness problem of solution; that is, under the restrictive condition (1.22), the solution of (1.1)–(1.6) is unique. For a complete answer to this problem, see Corollary 1.8 below.
Based on the uniqueness of solution in Theorem 1.3, by a standard way, we can deduce the existence of self-similar solution to the system (1.1)–(1.6).

Corollary 1.5. Assume that 𝐮0, 𝑣0, and 𝑤0 satisfy the assumptions of Theorem 1.3. Assume that, moreover, 𝐮0(𝑥)=𝜆𝐮0(𝜆𝑥),𝑣0(𝑥)=𝜆2𝑣0(𝜆𝑥),𝑤0(𝑥)=𝜆2𝑤0(𝜆𝑥).(1.24) Then, the solution (𝐮,𝑣,𝑤) constructed in Theorem 1.3 is a self-similar solution.

In order to give a complete answer to the uniqueness problem of solutions to the system (1.1)–(1.6), we will establish the following stability result.

Theorem 1.6. Let (𝐮0,𝑣0,𝑤0) and (̃𝐮0,̃𝑣0,𝑤0) belong to [𝒫2]3×[𝒫1]2 such that (𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2<1/(8𝜂2), where 𝜂 is a constant defined by (2.23), and let (𝐮,𝑣,𝑤) and (̃̃𝐮,𝑣,𝑤) be two solutions of (1.1)–(1.6) with initial conditions (𝐮0,𝑣0,𝑤0) and (̃𝐮0,̃𝑣0,𝑤0), respectively. Then, one has ̃̃(𝐮𝐮,𝑣𝑣,𝑤𝑤)𝒳×[𝒴𝑎]22𝜂(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2.(1.25)

Remark 1.7. Theorem 1.6 implies that we can measure the difference of two solutions in terms of the difference of their initial data provided that the difference between these initial data is small enough.
The direct consequence of Theorem 1.6 is the following corollary.

Corollary 1.8. Assume that 𝐮0[𝒫2]3, 𝐮0=0 in the distributional sense, and 𝑣0, 𝑤0𝒫1. Then, there exists at most one solution of (1.1)–(1.6) with initial data (𝐮0,𝑣0,𝑤0).

Finally, we study the asymptotic stability of solutions in the sense proposed in [13] and developed in [27].

Theorem 1.9. Let 𝜀 be a sufficiently small number such that 𝜀<min{1/(4𝜂2),1/̃𝜂}, where 𝜂 and ̃𝜂 are defined by (2.23) and (4.19), respectively. Assume that (𝐮0,𝑣0,𝑤0) and (̃𝐮0,̃𝑣0,𝑤0) satisfy the assumptions of Theorem 1.3, and let (𝐮,𝑣,𝑤) and (̃̃𝐮,𝑣,𝑤) be two global solutions of (1.1)–(1.6) with initial conditions (𝐮0,𝑣0,𝑤0) and (̃𝐮0,̃𝑣0,𝑤0), respectively. Then, the following two conditions are equivalent: lim𝑡𝑒𝑡Δ(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2+𝑡(𝑎1)/2×𝑒𝑡Δ(𝑣0̃𝑣0,𝑤0𝑤0)[𝒫𝑎]2=0,(1.26)lim𝑡̃̃(𝐮𝐮,𝑣𝑣,𝑤𝑤)[𝒫2]3×[𝒫1]2+𝑡(𝑎1)/2̃(𝑣𝑣,𝑤𝑤)[𝒫𝑎]2=0.(1.27)

Remark 1.10. As an interesting application of Theorem 1.9, we get the asymptotic stability of self-similar solutions to the system (1.1)–(1.6), namely, under the assumptions of Theorem 1.9, besides, assume that (𝐮0,𝑣0,𝑤0) satisfies (1.24). Then, we know that the mild solution (̃̃𝐮,𝑣,𝑤) tends to the self-similar solution (𝐮,𝑣,𝑤) as time goes to infinity as long as (̃𝐮0,̃𝑣0,𝑤0) satisfies the condition (1.26).

Notations
Let 𝐴 and 𝐵 be two real numbers; we denote 𝐴𝐵 if there is a universal constant 𝐶, which does not depend on varying parameters of the problem, such that 𝐴𝐶𝐵. We denote 𝐴𝐵 if 𝐴𝐵 and 𝐵𝐴. In the rest part of this paper, we will use “sup” instead of “esssup” for convenience.

Structure of the Paper
In Section 2, we prove Theorem 1.3 and Corollary 1.5. The purpose of Section 3 is to prove Theorem 1.6. In the last section, we present the proof of Theorem 1.9.

2. Global-in-Time Solutions

In this section, we give the proofs of Theorem 1.3 and Corollary 1.5. Thus, throughout this section, we assume that 𝐮0[𝒫2]3, 𝐮0=0, and 𝑣0,𝑤0𝒫1.

2.1. The Proof of Theorem 1.3

Let 𝑎 be a fixed number such that 1<𝑎<2, and let 𝔛=𝒳×[𝒴𝑎]2. Given (𝐮,𝑣,𝑤)𝔛, we define 𝔉(𝐮,𝑣,𝑤)=(𝐮,𝑣,𝑤), where 𝐮=𝑒𝑡Δ𝐮0+𝑡0𝑒(𝑡𝜏)Δ𝐺1(𝐮(𝜏),𝑣(𝜏),𝑤(𝜏))𝑑𝜏,(2.1)𝑣=𝑒𝑡Δ𝑣0+𝑡0𝑒(𝑡𝜏)Δ𝐺2(𝐮(𝜏),𝑣(𝜏),𝑤(𝜏))𝑑𝜏,(2.2)𝑤=𝑒𝑡Δ𝑤0+𝑡0𝑒(𝑡𝜏)Δ𝐺3(𝐮(𝜏),𝑣(𝜏),𝑤(𝜏))𝑑𝜏.(2.3)

We fulfil the proof of Theorem 1.3 through the following two lemmas.

Lemma 2.1. The map 𝔉 is well defined and maps 𝔛 into itself.

Proof. Note that although some parts of the proof were given in [27], we would rather give it for completeness. We first prove that 𝐮 is well defined and 𝐮𝒳. From (2.1), we can denote 𝐮=𝐮1+𝐮2+𝐮3, where 𝐮1(𝑡)=𝑒𝑡Δ𝐮0,𝐮2(𝑡)=𝑡0𝑒(𝑡𝜏)Δ[](𝐮𝐮)(𝜏)𝑑𝜏,𝐮3(𝑡)=𝑡0𝑒(𝑡𝜏)Δ(𝑣𝑤)(Δ)1(𝑤𝑣)(𝜏)𝑑𝜏.(2.4) For 𝐮1, since 𝐮0[𝒫2]3, it is easy to see that 𝑒𝑡Δ𝐮0𝒳=sup𝑡0sup𝜉3||𝜉||2|||𝑒𝑡|𝜉|2𝐮0|||𝐮(𝜉)0[𝒫2]3.(2.5) Thus, 𝑒𝑡Δ𝐮0𝐿([0,),[𝒫2]3). To prove the weak continuity of 𝐮1 with respect to 𝑡, due to the properties of heat semigroup 𝑒𝑡Δ, it suffices to prove this for 𝑡=0. For every 𝜑[𝒮(3)]3, by applying the Plancherel formula, we obtain ||𝑒𝑡Δ𝐮0𝐮0||=||||,𝜑3𝑒𝑡|𝜉|2𝐮10||||(𝜉)(𝜑)(𝜉)𝑑𝜉𝑡sup𝜉3||||||𝑒𝑡|𝜉|21𝑡||𝜉||2||||||𝐮0[𝒫2]3(𝜑)[𝐿1(3)]30as𝑡0.(2.6) This implies that 𝐮1𝒳,𝐮1𝒳𝐮0[𝒫2]3.(2.7) For 𝐮2, using (1.13) and the properties of the Fourier transform, we get 𝐮2[𝒫2]3=sup𝜉3||𝜉||2||||𝑡0𝑒(𝑡𝜏)|𝜉|2||||()(𝜉)𝑖𝜉(𝐮𝐮)(𝜉,𝜏)𝑑𝜏sup𝜉3𝑡0||𝜉||3𝑒(𝑡𝜏)|𝜉|2||𝜉||2||𝜉||2𝑑𝜏𝐮2𝒳sup𝜉3𝑡0||𝜉||2𝑒(𝑡𝜏)|𝜉|2𝑑𝜏𝐮2𝒳𝐮2𝒳.(2.8) Here, we have used the fact that |𝜉|2|𝜉|2|𝜉|1 (see [29, Chapter 5, Section  1, (8)]). For the weak continuity of 𝐮2 with respect to 𝑡, we can prove it by a standard argument (cf. [30]). Hence, 𝐮2𝒳,𝐮2𝒳𝐮2𝒳.(2.9) For 𝐮3, using similar calculations, 𝐮3[𝒫2]3=sup𝜉3||𝜉||2||||𝑡0𝑒(𝑡𝜏)|𝜉|2()(𝜉)(𝑣𝑤)(Δ)1||||(𝑤𝑣)(𝜉,𝜏)𝑑𝜏sup𝜉3𝑡0||𝜉||2𝑒(𝑡𝜏)|𝜉|2||𝜉||𝑎||𝜉||𝑎1𝜏(𝑎1)𝑑𝜏(𝑣,𝑤)2𝒴𝑎,2sup𝜉3𝑡0||𝜉||42𝑎𝑒(𝑡𝜏)|𝜉|2𝜏(𝑎1)𝑑𝜏(𝑣,𝑤)2𝒴𝑎,2sup𝜉3𝑡0||𝜉||(𝑡𝜏)22𝑎𝑒(𝑡𝜏)|𝜉|2(𝑡𝜏)(2𝑎)𝜏(𝑎1)𝑑𝜏(𝑣,𝑤)2𝒴𝑎,2(𝑣,𝑤)2𝒴𝑎,2.(2.10) Here, we have used the fact |𝜉|𝑎|𝜉|𝑎1|𝜉|22𝑎 and the assumption 1<𝑎<2 to ensure that the integral 𝑡0(𝑡𝜏)(2𝑎)𝜏(𝑎1)𝑑𝜏 is finite and independent of 𝑡. It remains to show the weak continuity of 𝐮3, but this is a standard argument as we mentioned before. Hence, 𝐮3𝒳,𝐮3𝒳(𝑣,𝑤)2𝒴𝑎,2.(2.11) Combining the above estimates (2.7)–(2.11), we see that 𝐮𝒳,𝐮𝒳𝐮0[𝒫2]3+𝐮2𝒳+(𝑣,𝑤)2𝒴𝑎,2.(2.12)
Next, we prove that 𝑣 is well defined and 𝑣𝒴𝑎. Note that, from (2.2), we can denote 𝑣=𝑣1+𝑣2+𝑣3, where 𝑣1(𝑡)=𝑒𝑡Δ𝑣0,𝑣2(𝑡)=𝑡0𝑒(𝑡𝜏)Δ[](𝐮𝑣)(𝜏)𝑑𝜏,𝑣3(𝑡)=𝑡0𝑒(𝑡𝜏)Δ𝑣(Δ)1(𝑤𝑣)(𝜏)𝑑𝜏.(2.13) Since 𝑣0𝒫1, as in the proof of 𝐮1, it can be easily seen that 𝑣1𝒴𝑎,𝑣1𝒴𝑎𝑣0𝒫1.(2.14) Indeed, it suffices to estimate the second term in the norm of 𝒴𝑎 as follows: sup𝑡>0𝑡(𝑎1)/2𝑣1𝒫𝑎=sup𝑡>0sup𝜉3𝑡(𝑎1)/2||𝜉||𝑎𝑒𝑡|𝜉|2||𝑣0||(𝜉)sup𝑡>0sup𝜉3𝑡||𝜉||2(𝑎1)/2𝑒𝑡|𝜉|2||𝜉||||𝑣0||𝑣(𝜉)0𝒫1.(2.15) For 𝑣2, we can do the same calculations to deal with the first term in the norm of 𝒴𝑎, and one obtains that 𝑣2𝒴𝑎,1𝐮𝒳𝑣𝒴𝑎,2.(2.16) To deal with the second term in the norm of 𝒴𝑎, we need to calculate more. Note first the following two elementary inequalities: ||𝜉||2𝑎0𝑡/2𝑒(𝑡𝜏)|𝜉|2𝜏(𝑎1)/2||𝜉||𝑑𝜏2𝑎𝑒(𝑡|𝜉|2)/20𝑡/2𝜏(𝑎1)/2||𝜉||𝑑𝜏𝑎𝑡(𝑎1)/2,||𝜉||2𝑎𝑡𝑡/2𝑒(𝑡𝜏)|𝜉|2𝜏(𝑎1)/2𝑑𝜏|𝜉|2𝑎𝑡(𝑎1)/2𝑡𝑡/2𝑒(𝑡𝜏)|𝜉|2||𝜉||𝑑𝜏𝑎𝑡(𝑎1)/2.(2.17) Hence, taking the Fourier transform to 𝑣2, we get ||𝑣2||𝑡0||𝜉||𝑒(𝑡𝜏)|𝜉|2||𝜉||2||𝜉||𝑎𝜏(𝑎1)/2𝑑𝜏𝐮𝒳𝑣𝒴𝑎,2𝐮𝒳𝑣𝒴𝑎,2||𝜉||2𝑎0𝑡/2+𝑡𝑡/2𝑒(𝑡𝜏)|𝜉|2𝜏(𝑎1)/2||𝜉||𝑑𝜏𝑎𝑡(𝑎1)/2𝐮𝒳𝑣𝒴𝑎,2.(2.18) This implies that 𝑣2𝒴𝑎,2𝐮𝒳𝑣𝒴𝑎,2. By a standard argument, we can prove that 𝑣2 is weakly continuous with respect to 𝑡. Hence, 𝑣2𝒴𝑎,𝑣2𝒴𝑎𝐮𝒳𝑣𝒴𝑎,2.(2.19) For 𝑣3, we can do the same calculations as for 𝑣2 and obtain 𝑣3𝒴𝑎,𝑣3𝒴𝑎(𝑣,𝑤)2𝒴𝑎,2.(2.20) Concluding the above estimates (2.14)–(2.20), we have already proved that 𝑣𝒴𝑎,𝑣𝒴𝑎𝑣0𝒫1+𝐮𝒳+(𝑣,𝑤)𝒴𝑎,2(𝑣,𝑤)𝒴𝑎,2.(2.21)
Similarly, for 𝑤, we can prove that 𝑤𝒴𝑎,𝑤𝒴𝑎𝑤0𝒫1+𝐮𝒳+(𝑣,𝑤)𝒴𝑎,2(𝑣,𝑤)𝒴𝑎,2.(2.22) The proof of Lemma 2.1 is complete by (2.12), (2.21), and (2.22).

From Lemma 2.1, there exists a constant 𝜂>0 such that, for any (𝐮,𝑣,𝑤)𝔛 and (𝐮,𝑣,𝑤)=𝔉(𝐮,𝑣,𝑤), one has (𝐮,𝑣,𝑤)𝔛𝐮𝜂0,𝑣0,𝑤0[𝒫2]3×[𝒫1]2+(𝐮,𝑣,𝑤)2𝔛.(2.23) Let 𝜀>0 be sufficiently small so that 4𝜂2𝜀<1. If (𝐮0,𝑣0,𝑤0)[𝒫2]3×[𝒫1]2𝜀, then from (2.23) one has (𝐮,𝑣,𝑤)𝔛𝜂𝜀+𝜂(𝐮,𝑣,𝑤)2𝔛.(2.24) Now, let 𝔅 be a closed ball in 𝔛 with radius 2𝜂𝜀, that is, 𝔅=(𝐮,𝑣,𝑤)𝔛(𝐮,𝑣,𝑤)𝔛.2𝜂𝜀(2.25) For any (𝐮,𝑣,𝑤)𝔅, from (2.24), we see that (𝐮,𝑣,𝑤)𝔛𝜂𝜀+𝜂(2𝜂𝜀)2=1+4𝜂2𝜀𝜂𝜀2𝜂𝜀.(2.26) This implies that 𝔉 maps 𝔅 into itself.

Lemma 2.2. Let 𝜀 be as before (𝜀<1/4𝜂2). If (𝐮0,𝑣0,𝑤0)[𝒫2]3×[𝒫1]2𝜀, then 𝔉 is a contraction mapping.

Proof. Let (𝐮1,𝑣1,𝑤1), (𝐮2,𝑣2,𝑤2)𝔅, and let (𝐮𝑗,𝑣𝑗,𝑤𝑗)=𝔉(𝐮𝑗,𝑣𝑗,𝑤𝑗), 𝑗=1,2. Then by a similar argument as in the proof of Lemma 2.1, we obtain the following estimate: (𝐮1𝐮2,𝑣1𝑣2,𝑤1𝑤2)𝔛𝐮𝜂1,𝑣1,𝑤1𝔛+𝐮2,𝑣2,𝑤2𝔛×𝐮1𝐮2,𝑣1𝑣2,𝑤1𝑤2𝔛4𝜂2𝜀𝐮1𝐮2,𝑣1𝑣2,𝑤1𝑤2𝔛.(2.27) Since 4𝜂2𝜀<1, we see that 𝔉 is a contraction mapping. By using Lemmas 2.1 and 2.2 and the Banach fixed point theorem, we know there exists a global solution (𝐮,𝑣,𝑤) of (1.1)–(1.6) in the space 𝒳×[𝒴𝑎]2, and this is the unique solution satisfying the condition (𝐮,𝑣,𝑤)𝒳×[𝒴𝑎]22𝜂𝜀. It remains to show that the solution depends continuously on initial data. Let (𝐮,𝑣,𝑤) and (̃̃𝐮,𝑣,𝑤) be two solutions of (1.1)–(1.6) corresponding to initial conditions (𝐮0,𝑣0,𝑤0) and (̃𝐮0,̃𝑣0,𝑤0), respectively, and (𝐮0,𝑣0,𝑤0)[𝒫2]3×[𝒫1]2(̃𝐮𝜀,0,̃𝑣0,𝑤0)[𝒫2]3×[𝒫1]2𝜀.(2.28) Then, proceeding as in (2.27), we get ̃̃(𝐮𝐮,𝑣𝑣,𝑤𝑤)𝒳×[𝒴𝑎]2𝜂(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2+4𝜂2𝜀̃̃(𝐮𝐮,𝑣𝑣,𝑤𝑤)𝔛.(2.29) Since 4𝜂2𝜀<1, (2.29) yields that ̃̃(𝐮𝐮,𝑣𝑣,𝑤𝑤)𝒳×[𝒴𝑎]2𝐮𝐶(𝜀,𝜂)0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0[𝒫2]3×[𝒫1]2,(2.30) where 𝐶(𝜀,𝜂)=𝜂/(14𝜀𝜂2). This proves Theorem 1.3.

2.2. The Proof of Corollary 1.5

Proof. On the one hand, from Theorem 1.3, we know, the system (1.1)–(1.6) admits a unique global solution (𝐮,𝑣,𝑤) with initial data (𝐮0,𝑣0,𝑤0). Moreover, (𝐮,𝑣,𝑤)𝒳×[𝒴𝑎]22𝜂𝜀. On the other hand, since 𝐮0, 𝑣0, and 𝑤0 satisfy the condition (1.24), by the scaling invariance of (1.1)–(1.6), for each 𝜆>0, the function (𝐮𝜆,𝑣𝜆,𝑤𝜆) (see (1.9)) is also a solution with the same initial data. Note that the norm of 𝔛 is invariant under the scaling (1.9), that is, 𝐮𝜆,𝑣𝜆,𝑤𝜆𝒳×[𝒴𝑎]2=(𝐮,𝑣,𝑤)𝒳×[𝒴𝑎]2.(2.31) Hence, by the uniqueness result of Theorem 1.3, the solution (𝐮,𝑣,𝑤) of (1.1)–(1.6) is self-similar.

3. Stability of Solutions

In this short section, we prove Theorem 1.6. Let us pick up any two solutions (𝐮,𝑣,𝑤) and (̃̃𝐮,𝑣,𝑤) associated with initial conditions (𝐮0,𝑣0,𝑤0) and (̃𝐮0,̃𝑣0,𝑤0), respectively, lying in [𝒫2]3×[𝒫1]2. As in the proof of Lemma 2.1, one has 𝑡0𝑒(𝑡𝜏)Δ𝐺1(𝐮(𝜏),𝑣(𝜏),𝑤(𝜏))𝐺1̃̃𝐮(𝜏),𝑣(𝜏),𝑤(𝜏)𝑑𝜏𝒳̃𝜂𝐮𝐮𝒳𝐮𝒳̃+𝐮𝒳+̃𝑤𝑣𝑣,𝑤𝒴𝑎(𝑣,𝑤)𝒴𝑎+̃𝑤𝑣,𝒴𝑎2𝜂𝑀2,(3.1) where 𝑀=max{(𝐮,𝑣,𝑤)𝔛̃̃,(𝐮,𝑣,𝑤)𝔛̃̃,(𝐮𝐮,𝑣𝑣,𝑤𝑤)𝔛}. Similarly, 𝑡0𝑒(𝑡𝜏)Δ(𝐺2(𝐮(𝜏),𝑣(𝜏),𝑤(𝜏))𝐺2(̃̃𝐮(𝜏),𝑣(𝜏),𝑤(𝜏)))𝑑𝜏𝒴𝑎2𝜂𝑀2,𝑡0𝑒(𝑡𝜏)Δ(𝐺3(𝐮(𝜏),𝑣(𝜏),𝑤(𝜏))𝐺3(̃̃𝐮(𝜏),𝑣(𝜏),𝑤(𝜏)))𝑑𝜏𝒴𝑎2𝜂𝑀2.(3.2) Now, we subtract the integral equation (1.14) for (̃̃𝐮,𝑣,𝑤) from the analogous expression for (𝐮,𝑣,𝑤), using the definition of the norm of 𝔛 and (3.1)-(3.2), we obtain ̃̃(𝐮𝐮,𝑣𝑣,𝑤𝑤)𝔛𝑒𝑡Δ(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)𝔛+2𝜂𝑀2𝜂(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2+2𝜂𝑀2.(3.3) Applying the definition of 𝑀, (3.3) yields that 𝑀𝜂(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2+2𝜂𝑀2.(3.4) Since we have assumed that (𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2<1/(8𝜂2), the continuity argument implies that 𝑀𝑀1, where 𝑀1 is the smallest root corresponding to the following quadratic equation: 2𝜂𝑀2𝑀+𝜂(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2=0.(3.5) From (3.4) we know that this root satisfies 𝑀12𝜂(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2.(3.6) Since 𝑀𝑀1, this last inequality yields that ̃̃(𝐮𝐮,𝑣𝑣,𝑤𝑤)𝔛2𝜂(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2.(3.7) The proof of Theorem 1.6 is complete.

4. Large Time Behavior of Solutions

We are now in a position to show Theorem 1.9 on the large time behavior of solutions to the system (1.1)–(1.6). Let (𝐮,𝑣,𝑤) and (̃̃𝐮,𝑣,𝑤) be two solutions of (1.1)–(1.6) constructed in Theorem 1.3 which correspond to initial conditions (𝐮0,𝑣0,𝑤0) and (̃𝐮0,̃𝑣0,𝑤0), respectively. Let us recall that, by Lemma 2.1, there exists a constant 2𝜂𝜀, the radius of 𝔅, such that (𝐮,𝑣,𝑤)𝔛(̃̃2𝜂𝜀,𝐮,𝑣,𝑤)𝔛2𝜂𝜀.(4.1) Now, to simplify the notations, we introduce the following two auxiliary functions: 𝑒(𝑡)=𝑡Δ(𝐮0̃𝐮0,𝑣0̃𝑣0,𝑤0𝑤0)[𝒫2]3×[𝒫1]2+𝑡(𝑎1)/2𝑒𝑡Δ(𝑣0̃𝑣0,𝑤0𝑤0)[𝒫𝑎]2,𝑙̃̃(𝑡)=(𝐮𝐮,𝑣𝑣,𝑤𝑤)[𝒫2]3×[𝒫1]2+𝑡(𝑎1)/2̃(𝑣𝑣,𝑤𝑤)[𝒫𝑎]2.(4.2) We first assume that (1.26) holds. It follows immediately from Lemma 2.1 that (𝑡)𝐿([0,)),lim𝑡(𝑡)=0.(4.3) By calculating the norm [𝒫2]3 of ̃𝐮𝐮, we can easily get ̃𝐮𝐮[𝒫2]3𝑒𝑡Δ(𝐮0̃𝐮𝟎)[𝒫2]3+𝑡0𝑒(𝑡𝜏)Δ̃̃(𝐮𝐮𝐮𝐮)𝑑𝜏[𝒫2]3+𝑡0𝑒(𝑡𝜏)Δ×(𝑣𝑤)(Δ)1̃𝑤(𝑤𝑣)𝑣(Δ)1̃𝑣𝑤𝑑𝜏[𝒫2]3=𝐼0+𝐼1+𝐼2.(4.4) For 𝐼1, let 𝛿(0,1) be a constant to be chosen later, we decompose the integral 𝑡0𝑑𝜏 into 0𝛿𝑡𝑑𝜏+𝑡𝛿𝑡𝑑𝜏 and estimate each term separately: 𝐼1sup𝜉30𝛿𝑡+𝑡𝛿𝑡||𝜉||2𝑒(𝑡𝜏)|𝜉|2𝐮[𝒫2]3+̃𝐮[𝒫2]3̃𝐮𝐮[𝒫2]3𝑑𝜏=𝐼11+𝐼12.(4.5) For 𝐼11, we change the variables 𝜏=𝑡𝑠 and use the following identity: sup𝜉3||𝜉||2𝑒(1𝑠)𝑡|𝜉|2=1(1𝑠)𝑡sup𝜎3|𝜎|2𝑒|𝜎|2=1.(1𝑠)𝑡𝑒(4.6) Thus, from (4.1), one has 𝐼114𝜂𝜀𝛿0(1𝑠)1̃𝐮(𝑡𝑠)𝐮(𝑡𝑠)[𝒫2]3𝑑𝑠.(4.7) For 𝐼12, we can estimate it directly that 𝐼124𝜂𝜀sup𝛿𝑡𝜏𝑡̃𝐮(𝜏)𝐮(𝜏)[𝒫2]3.(4.8) Hence, it follows immediately from (4.7) and (4.8) that 𝐼14𝜂𝜀𝛿0(1𝑠)1̃𝐮(𝑡𝑠)𝐮(𝑡𝑠)[𝒫2]3𝑑𝑠+sup𝛿𝑡𝜏𝑡̃𝐮(𝜏)𝐮(𝜏)[𝒫2]3.(4.9) Now, we compute 𝐼2, by using the same argument, we obtain 𝐼2sup𝜉30𝛿𝑡+𝑡𝛿𝑡||𝜉||42𝑎𝑒(𝑡𝜏)|𝜉|2𝜏(𝑎1)×𝜏(𝑎1)/2(𝑣,𝑤)[𝒫𝑎]2+𝜏(𝑎1)/2̃𝑤𝑣,[𝒫𝑎]2×𝜏(𝑎1)/2̃(𝑣𝑣,𝑤𝑤)[𝒫𝑎]2𝑑𝜏=𝐼21+𝐼22.(4.10) For 𝐼21, we change the variables 𝜏=𝑡𝑠, and from (4.1), 𝐼214𝜂𝜀𝛿0(1𝑠)(2𝑎)𝑠(𝑎1)(𝑡𝑠)(𝑎1)/2̃𝑣(𝑡𝑠)𝑣(𝑡𝑠),𝑤(𝑡𝑠)𝑤(𝑡𝑠)[𝒫𝑎]2𝑑𝑠.(4.11) For 𝐼22, we estimate it directly that 𝐼224𝜂𝜀sup𝛿𝑡𝜏𝑡𝑣̃𝑣𝑤(𝜏)(𝜏),𝑤(𝜏)(𝜏)[𝒫𝑎]2.(4.12) From (4.11) and (4.12), we get 𝐼24𝜂𝜀𝛿0(1𝑠)(2𝑎)𝑠(𝑎1)(𝑡𝑠)(𝑎1)/2̃(𝑣(𝑡𝑠)𝑣(𝑡𝑠),𝑤(𝑡𝑠)𝑤(𝑡𝑠))[𝒫𝑎]2𝑑𝑠+sup𝛿𝑡𝜏𝑡̃(𝑣(𝜏)𝑣(𝜏),𝑤(𝜏)𝑤(𝜏))[𝒫𝑎]2.(4.13) Combining the above estimates (4.9) and (4.13), we have already proved that ̃𝐮𝐮[𝒫2]3𝑒𝑡Δ(𝐮0̃𝐮𝟎)[𝒫2]3×+4𝜂𝜀𝛿0(1𝑠)1̃𝐮(𝑡𝑠)𝐮(𝑡𝑠)[𝒫2]3𝑑𝑠+sup𝛿𝑡𝜏𝑡̃𝐮(𝜏)𝐮(𝜏)[𝒫2]3+𝛿0(1𝑠)(2𝑎)𝑠(𝑎1)(𝑡𝑠)(𝑎1)/2̃𝑣(𝑡𝑠)𝑣(𝑡𝑠),𝑤(𝑡𝑠)𝑤(𝑡𝑠)[𝒫𝑎]2𝑑𝑠+sup𝛿𝑡𝜏𝑡̃(𝑣(𝜏)𝑣(𝜏),𝑤(𝜏)𝑤(𝜏))[𝒫𝑎]2.(4.14)By using the analogous argument above, we can estimate ̃𝑣𝑣 and obtain the following estimate: ̃𝑣𝑣𝒫1+𝑡(𝑎1)/2̃𝑣𝑣𝒫𝑎𝑒𝑡Δ(𝑣0̃𝑣0)𝒫1+𝑡(𝑎1)/2𝑒𝑡Δ(𝑣0̃𝑣0)𝒫𝑎×+4𝜂𝜀𝛿0(1𝑠)(3𝑎)/2𝑠(𝑎1)/2×(̃𝐮𝑡𝑠)𝐮(𝑡𝑠)[𝒫2]3+(𝑡𝑠)(𝑎1)/2̃𝑣𝑣(𝑡𝑠)(𝑡𝑠)𝒫𝑎+𝑑𝑠𝛿0(1𝑠)(2𝑎)𝑠(𝑎1)×(𝑡𝑠)(𝑎1)/2̃(𝑣(𝑡𝑠)𝑣(𝑡𝑠),𝑤(𝑡𝑠)𝑤(𝑡𝑠))[𝒫𝑎]2+𝑑𝑠𝛿0(1𝑠)1𝑠(𝑎1)/2×̃𝐮(𝑡𝑠)𝐮(𝑡𝑠)[𝒫2]3+(𝑡𝑠)(𝑎1)/2̃𝑣(𝑡𝑠)𝑣(𝑡𝑠)𝒫𝑎+𝑑𝑠𝛿0(1𝑠)(3𝑎)/2𝑠(𝑎1)×(𝑡𝑠)(𝑎1)/2̃𝑣(𝑡𝑠)𝑣(𝑡𝑠),𝑤(𝑡𝑠)𝑤(𝑡𝑠)[𝒫𝑎]2𝑑𝑠+sup𝛿𝑡𝜏𝑡̃𝐮(𝜏)𝐮(𝜏)[𝒫2]3+𝜏(𝑎1)/2̃(𝑣(𝜏)𝑣(𝜏),𝑤(𝜏)𝑤(𝜏)[𝒫𝑎]2.(4.15) The estimate of 𝑤𝑤 has exactly the same form as (4.15). Now, let 𝑀=limsup𝑡𝑙(𝑡)=lim𝑘,𝑘sup𝑡𝑘𝑙(𝑡).(4.16) In order to prove (1.27), it suffices to prove 𝑀=0. Note that from (4.1) we know that 𝑀 is nonnegative and finite. Hence, by applying the Lebesgue dominated convergence theorem to (4.14), (4.15), and the same estimate as (4.15) for 𝑤𝑤 and using the assumption (1.26) we obtain 𝑀4𝜂𝜀(𝐹(𝛿)+1)𝑀,(4.17) where 𝐹(𝛿) is defined by 1𝐹(𝛿)=log+(1𝛿)𝛿0(1𝑠)(2𝑎)𝑠(𝑎1)+𝑑𝑠𝛿0(1𝑠)(3𝑎)/2𝑠(𝑎1)/2+𝑑𝑠𝛿0(1𝑠)1𝑠(𝑎1)/2𝑑𝑠+𝛿0(1𝑠)(3𝑎)/2𝑠(𝑎1)𝑑𝑠.(4.18) Thus, there exists a universal constant ̃𝜂 which may depend on 𝜂 such that 𝑀̃𝜂𝜀(𝐹(𝛿)+1)𝑀.(4.19) Since we have assumed that 𝜀̃𝜂<1, we can choose 𝛿 sufficiently small such that ̃𝜂𝜀(𝐹(𝛿)+1)<1 by the fact that lim𝛿0𝐹(𝛿)=0. This implies 𝑀=0 by (4.19). We complete the proof of (1.27).

Conversely, we assume that (1.27) holds. Note that from (4.1) one has 𝑙(𝑡)𝐿([0,)),lim𝑡𝑙(𝑡)=0.(4.20) We need to prove (1.26). Repeat calculations similar to the proofs of (4.14) and (4.15), and, from the boundedness of (𝐮,𝑣,𝑤) and (̃̃𝐮,𝑣,𝑤) in (4.1), we can obtain the following estimate: (𝑡)̃𝜂𝜀(𝐹(𝛿)+1)𝑙(𝑡).(4.21) It is obvious that ̃𝜂𝜀(𝐹(𝛿)+1) is bounded and independent of 𝑡, so (1.26) follows immediately from (1.27) and (4.21). This proves Theorem 1.9.

Acknowledgments

This work is supported by the China National Natural Science Foundation under the Grant no. 11171357. The authors would like to thank the anonymous referee for invaluable suggestions.