Startseite Classical solution for compressible Navier-Stokes-Korteweg equations with zero sound speed
Artikel Open Access

Classical solution for compressible Navier-Stokes-Korteweg equations with zero sound speed

  • Mengqian Liu und Zhigang Wu EMAIL logo
Veröffentlicht/Copyright: 16. April 2025

Abstract

We consider the compressible Navier-Stokes-Korteweg equations describing the dynamics of a liquid-vapor mixture with diffuse interphase in R d with d 3 when the initial perturbation is suitably small. In particular, when the base sound speed P ( ρ ¯ ) = 0 , we first give the global existence and optimal L 2 -decay rate of the smooth solution, where the optimality means that the decay rate of the solution is the same as that for the corresponding linearized system, and there is no decay loss for the highest-order spatial derivatives of the solution. Then, we establish space-time behavior of the solution based on Green’s function method. It is obviously different from the case P ( ρ ¯ ) > 0 , which obeys the generalized Huygens’ principle as the compressible Navier-Stokes equations.

MSC 2010: 35A09; 35B40; 35Q35

1 Introduction

The compressible fluid models of Korteweg type were introduced by Korteweg [23] and deduced rigorously in Dunn and Serrin [9]. It governs the motions of the compressible isothermal viscous capillary fluids and is formulated as

(1.1) ρ ˜ t + div ( ρ ˜ u ) = 0 , ( ρ ˜ u ) t + div ( ρ ˜ u u ) + P ( ρ ˜ ) μ Δ u ( μ + μ ) div u = κ ρ ˜ Δ ρ ˜ .

Here, ( x , t ) R d × R + , the unknown functions ρ ˜ ( x , t ) and u ( x , t ) represent fluid density and fluid velocity, respectively. P = P ( ρ ˜ ) is the pressure given by a smooth function depending on ρ ˜ , and u u denotes the tensor product of two vectors. The viscosity constants μ and μ satisfy μ > 0 and 2 μ + μ > 0 , and κ > 0 is the capillary coefficient.

One of the most important features of system (1.1) is that the pressure is in general non-monotone. For instance, P ( ρ ¯ ) < 0 ( ρ ¯ is the background density) will lead to linear instability for some value of ρ ¯ . Thus, almost all of the mathematical results (in the whole space) only deal with the condition that P ( ρ ¯ ) > 0 , except for local-in-time results. In fact, there have been a lot of mathematical studies on system (1.1), which include the weak solutions in [2,12], the existence in Besov spaces in [4,21], the local existence of strong solutions in [24], the global existence of strong solutions in [37], the maximal L p - L q regularity in [31], the global existence of smooth solution in [13,14], and L 2 -decay rate of the smooth solution in [10,38]. With regard to the nonisothermal case, we refer to [16,17] for the global existence and the decay rate of the smooth solution with small initial energy, respectively, and [18] for the vanishing capillarity limit. When the sound speed c = P ( ρ ¯ ) > 0 , Jiang and Wu [19] verified the pointwise space-time behavior of the classical solution to the Cauchy problem of (1.1) as

(1.2) ( 1 + t ) 3 2 1 + x 2 1 + t 3 2 + ( 1 + t ) 2 1 + ( x c t ) 2 1 + t 3 2 , x R 3 ,

which means that the pointwise decription contains two diffusion waves, one is D-wave (zero sound speed) and the other is H-wave (the sound speed c > 0 ) . In other words, the classical solution obeys the generalized Huygens’ principle as that for the compressible Navier-Stokes equations in [5,6,15,27,28] due to the hyperbolic-parabolic structure of the system. In this direction, there are also similar space-time descriptions for the other compressible fluid models, for instance, the bipolar Navier-Stokes-Poisson equations [39,40]. However, we know that some compressible fluid models do not obey the generalized Huygens’ principle, such as the damped Euler equations [35,42] due to the damped mechanism and the unipolar Navier-Stokes-Poisson equations [36] due to the presence of the electronic field. Last but not least, we shall also refer to some work on the large time behavior of the other related models in [7,8,20,22,25,26,29,30,4345] and references therein.

On the contrary, for the critical case P ( ρ ¯ ) = 0 , Chikami and Kobayashi [3] recently established global existence of the strong solution in critical Besov space ( Λ ( ρ ˜ ρ ¯ ) , m ) B ˙ 2,1 d 2 3 ( R d ) B ˙ 2,1 d 2 1 ( R d ) with d 3 and also showed the decay estimates of such global solutions under additional regularity assumption B ˙ 2 , d 2 ( R d ) . This is the first global existence for this special case. To overcome the loss of lower order dissipation of the density arising from P ( ρ ¯ ) = 0 , Chikami and Kobayashi [3] gave a subtle assumption in the low-frequency and fully used the conservative structure of the system, which is the reason why they considered the unknowns: the density ρ and the momentum m . Recently, Song and Xu [33] extended the result in [3] to the L p -critical Besov space.

We are interested in the classical solution for the Cauchy problem of system (1.1) for the special case P ( ρ ¯ ) = 0 . In particular, we shall give the pointwise space-time description for this special case, which is inevitably different from the pointwise description given in [19]. To this end, we first prove that the classical solution globally exists. In fact, we need higher regularity to obtain the pointwise description of the solution when t 0 due to the quasi-linearity of the system and the singularity of Green’s function at t = 0 . Our global existence of the classical solution is partially inspired from [3], that is, we also need to impose some additional low-frequency assumption on the initial perturbation to overcome the loss of dissipation of the first derivative of the density ρ , and still consider the conservative variables ρ ˜ , m = ρ ˜ u with the initial data

(1.3) ρ ˜ ( x , t ) t = 0 = ρ ˜ 0 ( x ) , m ( x , t ) t = 0 = m 0 ( x ) .

More precisely, we are going to obtain the global solution for the small initial perturbation ( Λ ( ρ ˜ 0 ρ ¯ ) , m 0 ) B ˙ 2,1 d 2 3 ( R d ) H k ( R d ) ( k [ d 2 ] + 1 ) with d 3 . Note that when d 6 B ˙ 2,1 d 2 3 ( R d ) H k ( R d ) = H k ( R d ) , that is, one can obtain the global existence of classical solution in a pure H k -framework when d 6 .

On the contrary, we shall obtain the optimal L 2 -decay rate of the smooth solution and its derivatives. In fact, Chikami and Kobayashi [3] have established the optimal decay rate of the strong solution in B ˙ 2,1 s ( R d ) with s ( d 2 , d 2 + 1 ] and d 3 . Here, we derive the optimal L 2 -decay rate of the smooth solution and its derivatives by using two different methods. In fact, our L 2 -decay rate for the higher order derivatives of the smooth solution will be used to obtain the pointwise space-time behavior of the solution.

Finally, we provide the pointwise description of the solution by using Duhamel’s principle together with the energy estimates and L 2 -decay rate of the solution. We would like to state some differences in Green’s functions between the cases P ( ρ ¯ ) > 0 and P ( ρ ¯ ) = 0 (see the details in Section 2):

(1). The first difference is in G 22 for these two cases:

When P ( ρ ¯ ) > 0 , there exist the operators with symbols cos ( c ξ t ) ξ ξ τ ξ 2 ( e c 1 ξ 2 t e c 2 ξ 2 t ) and ( cos ( c ξ t ) 1 ) e ξ 2 t ξ ξ τ ξ 2 for positive constants c 1 and c 2 in G ˆ 22 , and they will produce so-called Riesz wave behaving as t d 2 1 + x 2 t d 2 (see (1.2)).

When P ( ρ ¯ ) = 0 , although G ˆ 22 contains the terms like ( e c 1 ξ 2 t e c 2 ξ 2 t ) ξ ξ τ ξ 2 , one can obtain the better estimate as the heat kernel t d 2 e x 2 t by using the cancellation.

(2). The second difference is in G 12 for these two cases:

When P ( ρ ¯ ) > 0 , it holds that G ˆ 12 = i e λ + t e λ t λ + λ ξ T i e ξ 2 t sin ( c ξ t ) ξ ξ T , which yields that G 12 behaves as a Huygens’ wave in the low frequency as in (1.2) due to the fact λ ± i c ξ with the sound speed c = P ( ρ ¯ ) .

When P ( ρ ¯ ) = 0 , it holds that G ˆ 12 = i e λ + t e λ t λ + λ ξ T i ξ T ξ 2 e ξ 2 t , which yields that x α G 12 behaves as ( 1 + t ) d 1 + α 2 1 + x 2 1 + t d 1 + α 2 according to Lemma A.6.

Before stating our main result, we first introduce some notations which will be used frequently throughout the rest of the article.

Notation. C denotes a generic positive constant which may vary in different estimates. We write f g instead of f C g . We note that = ( x 1 , , x d ) and for a multi-index α = ( α 1 , , α d ) , x α = x 1 α 1 x 2 α 2 x d α d , and α = i = 1 d α i . We also use x k to denote x α with α = k . The norm in the usual homogeneous and inhomogeneous Sobolev spaces H ˙ s ( R d ) and H s ( R d ) are denoted by H ˙ s and H s . In addition, let f ˆ ( ξ ) be the Fourier transform of f ( x ) with respect to x R d , that is, f ˆ ( ξ ) = ( f ) ( ξ ) = ( 2 π ) d 2 R d f ( x ) e i x ξ d x . We also define the fractional derivative operator by Λ s ( Δ ) s 2 = 1 s .

The main results in this article are stated as follows.

Theorem 1.1

( d 3 ) Assume that P ( ρ ¯ ) = 0 and ( Λ ( ρ ˜ 0 ρ ¯ ) , m 0 ) B ˙ 2,1 d 2 3 H k ( R d ) is sufficiently small with an integer k [ d 2 ] + 1 . Then, there exists a unique global smooth solution ( ρ ˜ ( x , t ) , m ( x , t ) ) to the Cauchy problem of (1.1) such that

(1.4) ( Λ ( ρ ˜ ρ ¯ ) , m ) ( t ) B ˙ 2,1 d 2 3 ( R d ) 2 + ( ( ρ ˜ ρ ¯ ) , m ) ( t ) H k ( R d ) 2 + 0 t Δ ρ ˜ ( τ ) H k ( R d ) 2 + m ( τ ) H k ( R d ) 2 d τ ( ( Λ ( ρ ˜ 0 ρ ¯ ) , m 0 ) B ˙ 2,1 d 2 3 ( R d ) 2 + ( ( ρ ˜ 0 ρ ¯ ) , m 0 ) H k ( R d ) 2 ) .

Remark 1.1

It is easy to see that one does not need the initial assumption in a negative Besov space when d 6 , that is, one can obtain the global existence of the solution in the usual H k -framework. And, it is interesting to prove the global existence of the strong (smooth) solution for d = 2 as stated in [3].

In the following, we always use f L and f H to denote the low- and high-frequency parts of a well-defined function f .

Theorem 1.2

Suppose that the assumptions in Theorem 1.1 hold.

  1. If further for d 3 that

    (1.5) Λ ( ρ ˜ 0 ρ ¯ ) L , m 0 L L 1 ( R d ) 1 ,

    then when k [ d 2 ] + 1 , the solution for the Cauchy problem of (1.1) satisfies

    (1.6) x j ( Λ ( ρ ˜ ρ ¯ ) , m ) L 2 ( R d ) C ( 1 + t ) d 4 j 2 , 0 j k .

  2. If further for d 4 and k [ d 2 ] + 2 that

    (1.7) ( Λ ( ρ ˜ 0 ρ ¯ ) , m 0 ) B ˙ 2 , d 2 ( R d ) L 1 ,

    then the solution for the Cauchy problem of (1.1) satisfies

    (1.8) x j ( Λ ( ρ ˜ 0 ρ ¯ ) , m ) L 2 ( R d ) C ( 1 + t ) d 4 j 2 , j = 0 , 1 , , k .

Remark 1.2

Our main contribution is the optimal decay rate for the higher order derivatives j 3 , where the optimality means that it is consistent in the decay rate of the linear problem, moreover, there is no decay loss for the highest-order spatial derivatives of the solution. In fact, the L 2 -decay result of the lower order derivatives ( 0 j 2 ) of the solution in (1.8) is a byproduct of that in B ˙ 2,1 s with s d 2 + 1 in [3]. Due to the quasi-linearity of the system, to derive the pointwise estimates of the lower order derivatives of the solution, we need the higher order derivatives of the solution to close the ansatz.

Remark 1.3

The second part of Theorem 1.2 is proved by a pure energy method introduced by Guo and Wang [11], where they studied the compressible Navier-Stokes equations in R 3 and obtained the optimal decay rate for j -order derivatives ( 0 j k 1 ) of the solution (except the highest order derivatives). When considering the compressible Navier-Stokes-Korteweg equations for P ( ρ ¯ ) = 0 and d 4 , we actually obtained the decay rate of the highest order derivative. In fact, the optimal decay rate of the highest order derivatives of the solution when P ( ρ ¯ ) 0 and d 3 can be derived similarly. This can also be regarded as one of the largest differences between these two fluid models.

Theorem 1.3

( d 3 ) Let the assumptions of Theorem 1.2 hold for k [ d 2 ] + 1 . If the initial data has the pointwise assumption:

(1.9) ( ρ ˜ 0 , m 0 ) C ( 1 + x 2 ) r , r > d 2 ,

then the solution has the following pointwise space-time behavior for t 1 :

(1.10) ( ρ ˜ ρ ¯ ) ( x , t ) C ( 1 + t ) d 1 2 1 + x 2 1 + t d 1 2 , w h e n 2 μ + μ ρ ¯ 2 κ ρ ¯ , ( ρ ˜ , m ) ( x , t ) C ( 1 + t ) d 2 1 + x 2 1 + t d 2 , w h e n 2 μ + μ ρ ¯ 2 κ ρ ¯ ,

(1.11) ( ρ ˜ ρ ¯ ) ( x , t ) C ( 1 + t ) d 1 2 1 + x 2 1 + t r , when 2 μ + μ ρ ¯ = 2 κ ρ ¯ , ( ρ ˜ , m ) ( x , t ) C ( 1 + t ) d 2 1 + x 2 1 + t r , when 2 μ + μ ρ ¯ = 2 κ ρ ¯ .

Remark 1.4

It is obvious that the pointwise space-time behavior of the solution for the special case P ( ρ ¯ ) = 0 is different from the case P ( ρ ¯ ) > 0 . In fact, when the bound sound speed P ( ρ ¯ ) > 0 , the solution obeys the generalized Huygens’ principle as the compressible Navier-Stokes equations [28].

Remark 1.5

We only give the pointwise estimate of the solution when t 1 . Compared with the previous pointwise results for other fluid models such as [6,19,27,28], the regularity is largely relaxed if one only focuses on the large time behavior. On the contrary, due to the singularity of the heat kernel at t = 0 and the quasi-linearity of the system, to obtain the pointwise estimate of the solution when t 0 , it needs to impose the higher regularity assumption on the initial data as our previous pointwise results for other fluid models. For instance, when d = 3 , if the assumptions in Theorem 1.2 hold with k 4 , one can also have the pointwise estimates as in (1.10) and (1.11) when t 0 .

Remark 1.6

The difference in the spacial integrability in (1.10) and (1.11) is arising from the pointwise information of Green’s function in Theorem 2.1 for two cases: 2 μ + μ ρ ¯ 2 κ ρ ¯ and 2 μ + μ ρ ¯ = 2 κ ρ ¯ .

The rest of this article is organized as follows. In Section 2, we shall obtain the representation of Green’s function in Fourier space and give the pointwise estimate for Green’s function. The global existence and L 2 -decay rate of the solution are derived in Sections 3 and 4. In Section 5, we obtain the pointwise space-time behavior of the solution for the nonlinear problem. At last, some useful lemmas are stated in Appendix.

At the end of this section, we shall state the definition of homogeneous Besov space for completeness. We first recall some basic facts on the Littlewood-Paley theory (see [1] for instance). Let χ 0 and φ 0 be two smooth radial functions so that the support of χ is contained in the ball { ξ R d : ξ 4 3 } , the support of φ is contained in the annulus { ξ R d : 3 4 ξ 8 3 } and

j Z φ ( 2 j ξ ) = 1 , ξ 0 .

The homogeneous dyadic blocks Δ ˙ j and the low-frequency cutoff operators S ˙ j are defined for all j Z by

Δ ˙ j f = 1 ( φ ( 2 j ) f ) , S ˙ j f = 1 ( χ ( 2 j ) f ) .

Remark that for any homogeneous function η of order 0 smooth outside 0, we have

p [ 1 , ] , η ( D ) Δ ˙ j f L p C Δ ˙ j f L p .

Denote by S h ( R d ) the space of tempered distributions subject to the condition

lim j S ˙ j f = 0 .

Then, we have the decomposition

f = j Z Δ ˙ j f , f S h ( R d ) .

Now, we recall the definition of homogeneous Besov spaces.

Definition 1.1

Let ( p , r ) [ 1 , + ] 2 , s R and f S h ( R d ) , which means that f S ( R d ) and lim j S ˙ j f L = 0 (see Definition 1.26 of [1]), we set f B ˙ p , r s = def ( 2 j s Δ ˙ j f L p ) r .

2 Green’s function

In what follows, we assume that the steady state of the Cauchy problem (1.1)–(1.3) is ( ρ ¯ , 0 ) . For notational simplicity, we use ( ρ , m ) to denote the perturbation ( ρ ˜ ρ ¯ , m ) . Then, system (1.1) can be rewritten in the perturbation form as

(2.1) ρ t + div m = 0 , m t μ + μ ρ ¯ div m μ ρ ¯ Δ m κ ρ ¯ Δ ρ = F ,

where the nonlinear term F 1 is defined as

(2.2) F = div m m ρ + ρ ¯ μ ρ ¯ Δ ρ m ρ + ρ ¯ μ + μ ρ ¯ div ρ m ρ + ρ ¯ P ( ρ + ρ ¯ ) κ ρ Δ ρ .

Denote Green’s function of the linearized system for (2.1) by

(2.3) G t = B G , G t = 0 = δ 0 ( x ) ,

where δ 0 is the Dirac- δ function and

(2.4) B = 0 div κ ρ ¯ Δ μ ρ ¯ Δ + μ + μ ρ ¯ div .

A direct computation gives the representation of Green’s function in Fourier space:

(2.5) G ˆ ( ξ , t ) = λ + e λ t λ e λ + t λ + λ i e λ + t e λ t λ + λ ξ T i κ ρ ¯ e λ + t e λ t λ + λ ξ 2 ξ e μ ρ ¯ ξ 2 t I + λ + e λ + t λ e λ t λ + λ e μ ρ ¯ ξ 2 t ξ ξ τ ξ 2 = λ + e λ t λ e λ + t λ + λ i e λ + t e λ t λ + λ ξ T i κ ρ ¯ e λ + t e λ t λ + λ ξ 2 ξ e μ ρ ¯ ξ 2 t I + λ + ( e λ + t e μ ρ ¯ ξ 2 t ) λ + λ λ ( e λ t e μ ρ ¯ ξ 2 t ) λ + λ ξ ξ τ ξ 2

when λ + λ . Here,

(2.6) λ ± ( ξ ) = η 2 ξ 2 ± 1 2 ξ 2 η 2 4 κ ρ ¯ , with η = 2 μ + μ ρ ¯ .

As a result, one has

(2.7) λ + = a 1 ξ 2 , λ = a 2 ξ 2 , with 0 < a 1 < a 2 , when η > 2 κ ρ ¯ , λ ± = η 2 ξ 2 i 4 κ ρ ¯ η 2 ξ 2 , when η < 2 κ ρ ¯ .

When λ + = λ (or η 2 μ + μ ρ ¯ = 2 κ ρ ¯ ), Green’s function has the following representation in the Fourier space:

(2.8) G ˆ ( ξ , t ) = 1 + η 2 ξ 2 t e η 2 ξ 2 t i t e η 2 ξ 2 t ξ T i κ ρ ¯ ( ξ 2 t η 2 ξ 4 t 2 ) e η 2 ξ 2 t ξ e μ ρ ¯ ξ 2 t I κ ρ ¯ 2 t 2 ξ 2 ξ ξ T e η 2 ξ 2 t + 0 0 0 ( e η 2 ξ 2 t e μ ρ ¯ ξ 2 t ) ξ ξ τ ξ 2 .

Now, we shall state the differences in Green’s functions between the case P ( ρ ¯ ) > 0 and the case P ( ρ ¯ ) = 0 as follows:

First, when P ( ρ ¯ ) > 0 , the lowest power of ξ in the imaginary part of λ ± for the low-frequency is one, and hence, there exists wave operators in the low frequency. As a result, one can deduce the Huygens’ wave in the low frequency and ultimately verify the generalized Huygens’ principle for the nonlinear problem. When P ( ρ ¯ ) = 0 (the sound speed is zero), the lowest power of ξ in λ ± in the low frequency is two, and there is no wave operator in the low frequency. Hence, we can verify that the pointwise space-time description of Green’s function only contain the diffusion wave with zero sound speed (like the heat kernel).

Second, there exist the operators with symbols cos ( c ξ t ) ξ ξ τ ξ 2 ( e a 1 ξ 2 t e a 2 ξ 2 t ) and ( cos ( c ξ t ) 1 ) e ξ 2 t ξ ξ τ ξ 2 for some positive constant a 1 and a 2 in G 22 when P ( ρ ¯ ) > 0 , and this Riesz operator will produce the so-called Riesz wave behaving as t d 2 1 + x 2 t d 2 in the pointwise description of G 22 and hence affects the estimate for the solution of the nonlinear problem. However, when P ( ρ ¯ ) = 0 , from G 22 in (2.5) and (2.8) and recalling (2.6)–(2.7), we find that although G 22 contains the terms like ( e a 1 ξ 2 t e a 2 ξ 2 t ) ξ ξ τ ξ 2 , one can verify that it has better estimate as the heat kernel t d 2 e x 2 C t by using the cancellation.

Third, G ˆ 12 will produce Huygens’ wave since the lowest power of ξ in λ ± in the low frequency is one when P ( ρ ¯ ) > 0 , and there is no singularity based on the reformulation. However, for P ( ρ ¯ ) = 0 , G ˆ 12 is of the form i e λ + t e λ t λ + λ ξ T or i t e η 2 ξ 2 t ξ T . These two terms have the same time-decay rate. Nevertheless, the former is singular when ξ = 0 due to the operator with the symbol ξ ξ 2 . This operator is also contained in the electronic field ϕ = Δ ρ of the compressible Navier-Stokes-Poisson equations [36]. Thus, in virtue of Lemma A.6, the former term can be bounded by t d 1 2 1 + x 2 t d 1 2 .

In particular, we have the following pointwise description when P ( ρ ¯ ) = 0 .

Theorem 2.1

There exists a constant C > 0 such that for k 0 and t 1 , it holds that

(2.9) x k G 11 + x k G 22 C ( 1 + t ) d + k 2 e x 2 C ( 1 + t ) , x k G 12 C ( 1 + t ) d 1 + k 2 1 + x 2 1 + t d 1 + k 2 , w h e n 2 μ + μ ρ ¯ 2 κ ρ ¯ , x k G 12 C ( 1 + t ) d 1 + k 2 e x 2 C ( 1 + t ) , w h e n 2 μ + μ ρ ¯ = 2 κ ρ ¯ , x k G 21 C ( 1 + t ) d + 1 + k 2 e x 2 C ( 1 + t ) .

As a result, one can immediately obtain the following L 2 -decay estimate of the low-frequency of the solution for the linear Navier-Stokes-Korteweg system (2.1) when P ( ρ ¯ ) = 0 .

Corollary 2.2

Let k 0 be an integer. Assume that ( ρ , m ) be a solution for the linear Navier-Stokes-Korteweg system (2.1).

When the initial data ( Λ ρ 0 , m 0 ) H k L q ( R d ) and q [ 1 , ] , then

(2.10) x k ( Λ ρ L , m L ) L 2 ( R d ) C ( 1 + t ) d 2 ( 1 q 1 2 ) k 2 ( Λ ρ 0 L , m 0 L ) L q ( R d ) .

When the initial data ( ρ 0 , Λ 1 m 0 ) H k L q ( R d ) and q [ 1 , ] , then

(2.11) x k ρ L L 2 ( R d ) C ( 1 + t ) d 2 ( 1 q 1 2 ) k 2 ( ρ 0 L , Λ 1 m 0 L ) L q ( R d ) , x k m L L 2 ( R d ) C ( 1 + t ) d 2 ( 1 q 1 2 ) k + 1 2 ( ρ 0 L , Λ 1 m 0 L ) L q ( R d ) .

This corollary will be used to obtain the optimal decay rate of any order derivative of the solution for the nonlinear problem.

3 Global existence

The global existence is usually derived by combining local existence with the a priori estimate as usual. In fact, the local existence can be established by a standard contraction mapping argument, and hence, we mainly focus on giving the a priori estimate. In this section, we shall consider the critical case: P ( ρ ¯ ) = 0 . The main results are stated as follows.

The local existence was given in [3] when the initial perturbation is in B ˙ 2,1 d 2 1 ( R d ) for d 2 . In the same way, one can prove the local existence in the framework of Theorem 1.1. For simplicity, we omit the details. Thus, in the following, we only need to obtain the a priori estimate as follows:

Proposition 3.1

(A priori estimate). Assume that P ( ρ ¯ ) = 0 and let the initial data satisfy ( Λ ρ 0 , m 0 ) B ˙ 2,1 d 2 3 H k ( R d ) for d 3 and an integer k [ d 2 ] + 1 . Suppose that (2.1) has a solution ( ρ , m ) satisfying

(3.1) sup 0 t T { ( Λ ρ , m ) B ˙ 2,1 d 2 3 ( R d ) + ( Λ ρ , m ) H k ( R d ) } δ 1 ,

then for any t [ 0 , T ] , it holds that

( Λ ρ , m ) ( t ) B ˙ 2,1 d 2 3 ( R d ) 2 + ( Λ ρ , m ) ( t ) H k ( R d ) 2 + 0 t ( Δ ρ ( τ ) H k ( R d ) 2 + m ( τ ) H k ( R d ) 2 ) d τ ( Λ ρ 0 , m 0 ) B ˙ 2,1 d 2 3 ( R d ) 2 + ( Λ ρ 0 , m 0 ) H k ( R d ) 2 .

Proof

We give the proof in three steps.

Step 1. Basic energy estimate.

Multiplying Δ (2.1) 1 and (2.1) 2 with κ ρ ¯ ρ and m , respectively, and then integrating the resultant equalities over R d , one has

(3.2) 1 2 d d t { κ ρ ¯ ρ ( t ) L 2 2 + m L 2 2 } + μ ρ ¯ m L 2 2 + μ + μ ρ ¯ div m L 2 2 = m , F = m , div m m ρ + ρ ¯ + μ ρ ¯ Δ ρ m ρ + ρ ¯ + μ + μ ρ ¯ div ρ m ρ + ρ ¯ + P ( ρ + ρ ¯ ) + κ ρ Δ ρ I 1 + I 2 + I 3 + I 4 .

Then, by virtue of the Hölder inequality, Cauchy inequality, Sobolev embedding theorem, and the smallness of δ , one has

(3.3) I 1 = m , div m m ρ + ρ ¯ m L 2 m L 2 d d 2 m L d δ m L 2 2 , I 2 = m , μ ρ ¯ Δ ρ m ρ + ρ ¯ + μ + μ ρ ¯ div ρ m ρ + ρ ¯ = m , μ ρ ¯ ρ m ρ + ρ ¯ + div m , μ + μ ρ ¯ div ρ m ρ + ρ ¯ m L 2 ρ L m L 2 + m L d ρ L 2 d d 2 + div m L 2 ( div m L 2 ρ L + m L d ρ L 2 d d 2 ) δ ( div m L 2 2 + m L 2 2 ) , I 4 = κ m , ρ Δ ρ = κ div m , ρ Δ ρ + κ 2 m , div ( ρ ρ ) δ ( div m L 2 2 + m L 2 2 + Δ ρ L 2 2 ) .

For I 3 , we shall give a subtle computation due to the loss of the lower order dissipation of ρ . First, denote P ( ρ + ρ ¯ ) = ( ρ P ˜ ( ρ ) ) due to the fact P ( ρ ¯ ) = 0 , where P ˜ ( ρ ) is smooth and vanishes at 0. Then,

(3.4) I 3 = m , P ( ρ + ρ ¯ ) = ρ P ˜ ( ρ ) , div m div m L 2 Λ 1 ρ L d Λ P ˜ ( ρ ) L 2 d d 2 div m L 2 Λ 4 d 2 ρ L 2 Λ ρ L 2 d d 2 div m L 2 Λ 4 d 2 ρ L 2 2 ρ L 2 div m L 2 Λ ρ B ˙ 2,1 d 2 3 2 ρ L 2 δ ( div m L 2 2 + 2 ρ L 2 2 ) δ ( div m L 2 2 + Δ ρ L 2 2 ) ,

where we have used the well-known facts in [1]:

(3.5) Λ s f L q f L p , when 0 < s < d , 1 < p < q < , and 1 q + s d = 1 p , f L 2 d d 2 s f H ˙ s , when s 0 , d 2 .

Next, we shall obtain the dissipation of Δ ρ . Multiplying (2.1) 2 by ρ , integrating the resulting equation in R d and using (A) 1 , one has

(3.6) d d t ρ , m + κ ρ ¯ Δ ρ L 2 2 = div m L 2 2 + μ ρ ¯ Δ m , ρ + μ + μ ρ ¯ div m , ρ ρ , div m m ρ + ρ ¯ μ ρ ¯ ρ , Δ ρ m ρ + ρ ¯ μ + μ ρ ¯ ρ , div ρ m ρ + ρ ¯ ρ , P ( ρ + ρ ¯ ) κ ρ , ρ Δ ρ div m L 2 2 + i = 5 12 I i .

Then, by using the smallness of δ , Hölder inequality, Sobolev inequality, Cauchy inequality, and the same process of the proof for I 1 to I 4 , one can easily conclude that

(3.7) i = 5 12 I i κ ρ ¯ 4 Δ ρ L 2 2 + C ( div m L 2 2 + m L 2 2 ) ,

which together with (3.6) implies that

(3.8) d d t ρ , m + κ ρ ¯ 2 Δ ρ L 2 2 ( div m L 2 2 + m L 2 2 ) .

Then, by taking (3.4) + η (3.8) with a suitably small constant η > 0 , one can immediately obtain

(3.9) 1 2 d d t { κ ρ ¯ ρ ( t ) L 2 2 + m L 2 2 + 2 η ρ , m } + μ 2 ρ ¯ m L 2 2 + μ + μ 2 ρ ¯ div m L 2 2 + η κ ρ ¯ 2 Δ ρ L 2 2 0 .

Step 2. The highest order energy estimate.

The process in this step is similar to the first step, and thus we only give the sketch of the proof. First, we have for j 1 that

(3.10) 1 2 d d t { κ ρ ¯ x j ρ ( t ) L 2 2 + x j m L 2 2 } + μ ρ ¯ x j m L 2 2 + μ + μ ρ ¯ x j div m L 2 2 = x j m , x j F 1 C δ ( x j m L 2 2 + x j Δ ρ L 2 2 ) ,

(3.11) d d t x j ρ , x j m + κ ρ ¯ x j Δ ρ L 2 2 = x j div m L 2 2 + μ ρ ¯ x j Δ m , x j ρ + μ + μ ρ ¯ x j div m , x j ρ + x j ρ , x j F 1 C x j m L 2 2 + C δ x j Δ ρ L 2 2 .

Then, by taking (3.10) + η (3.11) with a suitably small constant η > 0 , one can immediately obtain the following by combining the basic energy estimate and the highest order energy estimate:

(3.12) 1 2 d d t { κ ρ ¯ ρ ( t ) H ˙ k 2 + m H ˙ k 2 + 2 η x k ρ , x k m } + μ 2 ρ ¯ m H ˙ k 2 + μ + μ 2 ρ ¯ div m H ˙ k 2 + η κ ρ ¯ 2 Δ ρ H ˙ k 2 0 , k 0 .

Step 3. Energy estimate in the negative Besov (Sobolev) space.

Note that we have used the estimates Λ ρ B ˙ 2,1 d 2 3 ( R d ) with d 3 . Thus, to enclose the a priori estimate, we have to obtain the estimate for the density in the negative Besov (Sobolev) space. In fact, Chikami and Kobayashi [3] have obtained this estimate when the initial perturbation ( Λ ρ 0 , m 0 ) is small in B ˙ 2,1 d 2 3 B ˙ 2,1 d 2 1 ( R d ) . As we know, H s ( s > d 2 ) B 2,1 d 2 1 B ˙ 2,1 d 2 1 B ˙ 2,1 d 2 , which also yields that B ˙ 2,1 d 2 3 H s ( s > d 2 ) B ˙ 2,1 d 2 3 B ˙ 2,1 d 2 1 . Thus, we can directly use the following estimate in the negative Besov space given in [3]:

(3.13) ( Λ ρ , m ) B ˙ 2,1 d 2 3 B ˙ 2,1 d 2 1 ( R d ) + Λ 2 ( Λ ρ , m ) L t 1 ( B ˙ 2,1 d 2 3 B ˙ 2,1 d 2 1 ( R d ) ) C ( Λ ρ 0 , m 0 ) B ˙ 2,1 d 2 3 B ˙ 2,1 d 2 1 ( R d ) .

Finally, combining the above estimates suffices to complete the proof of Proposition 3.1.□

4 L 2 -decay rate

In this section, we mainly give the optimal L 2 ( R d ) -decay rate for the higher order derivatives of the smooth solution when d 3 , although Chikami and Kobayashi [3] have given the decay rate of the solution in B ˙ 2,1 s with s ( d 2 , d 2 + 1 ] for d 3 . In fact, these decay estimates for the higher order derivatives of the solution will also be used to derive the pointwise estimate for the solution of the nonlinear problem due to the quasi-linearity of the system in the next section. We shall use two different methods to deal with the cases d 3 and d 4 .

First, when d 3 , we shall deduce the optimal L 2 -decay result of the smooth solution in the H ˙ k -framework via a method different from that in [3] by replacing the assumption ( Λ ρ 0 , m 0 ) B ˙ 2 , d 2 L in [3] with ( Λ ρ 0 L , m 0 L ) L 1 .

Second, when d 4 , under the same assumption as in [3], we still obtain the decay result in the H ˙ k -framework by using another method different from that in [3]. In particular, our proof is inspired by the idea in Guo and Wang [11] for compressible Navier-Stokes equations, where they obtained the L 2 -decay rate of the derivatives with the order 0 j k 1 by using a family of scaled energy estimates with minimum derivative counts and interpolations among them without linear decay analysis. Unfortunately, this method is not suitable for the case d = 3 when P ( ρ ¯ ) = 0 , since the decay rate of ρ is slower than the heat kernel, which is the main difference from the compressible Navier-Stokes equations in [11]. On the contrary, due to the benefit of the capillary term, we can obtain the L 2 -decay rate of any order derivatives of the solution, which is one of the biggest differences between these two fluid models in [11] and (1.1).

4.1 Method 1

In this subsection, we mainly consider the case d = 3 , since the other case d > 3 can be treated without any new difficulty due to the fact that the solution decays faster as the dimension becomes larger.

Denote U = ( Λ ρ , m ) . If the initial data U 0 H k B ˙ 2,1 3 2 δ 1 and U 0 L L 1 is bounded, we are going to prove that

(4.1) ρ L 2 C ( 1 + t ) 1 4 , x j U L 2 C ( 1 + t ) 3 4 j 2 , j = 0 , 1 , , k , k 2 .

Proof of (4.1)

First, by using the j -level energy estimate in (3.12) and the high-low-frequency decomposition, we can rewrite (3.12) as follows:

(4.2) 1 2 d d t { κ ρ ¯ x j ρ ( t ) L 2 2 + x j m L 2 2 + 2 η x j ρ , x j m } 2 μ + μ 4 ρ ¯ x j m H 2 + η κ ρ ¯ 4 x j ρ H 2 + μ 4 ρ ¯ m H ˙ j 2 + μ + μ 4 ρ ¯ x j div m 2 + η κ ρ ¯ 4 x j Δ ρ 2 0 ,

where ρ H and m H represent the high-frequency part of ρ and m . By adding the corresponding low-frequency part ( ρ L , m L ) on both sides of (4.2) and using the smallness of η , one can obtain that there exists a positive constant C 0 such that

(4.3) d d t { κ ρ ¯ ρ ( t ) H ˙ j 2 + m H ˙ j 2 + 2 η x j ρ , x k m } + C 0 { κ ρ ¯ ρ ( t ) H ˙ j 2 + m H ˙ j 2 + 2 η x j ρ , x j m } C ( ρ L ( t ) , m L ( t ) ) H ˙ j 2 .

Denote

(4.4) M ( t ) = sup 0 τ t j = 0 k ( 1 + τ ) 3 4 + j 2 x j U ( τ ) L 2 + ( 1 + τ ) 1 4 ρ L 2 .

Note that M ( t ) is non-decreasing, and

(4.5) ρ L 2 C M ( t ) ( 1 + τ ) 1 4 , x j U ( τ ) L 2 C M ( t ) ( 1 + τ ) 3 4 j 2 , j = 0 , 1 , , k .

For the convenience of narrative, we shall first divide the nonlinear term F ( U ) in (2.2) into two parts:

(4.6) F ˜ ( U ) F ( U ) P ( ρ + ρ ¯ ) and P ( ρ + ρ ¯ )

and

(4.7) F ˜ ( U ) m x m + x 2 ( ρ m ) + ρ x 3 ρ .

By using the Hölder inequality and Sobolev inequality, we have

(4.8) F ˜ ( U ) L 1 m L 2 ( x m , x 2 ρ ) L 2 + ρ L 2 ( x 2 m , x 3 ρ ) L 2 δ M ( t ) ( 1 + τ ) 5 4 .

On the contrary, taking advantage of the Taylor expansion, the second term P ( ρ + ρ ¯ ) can be rewritten as

(4.9) P ( ρ + ρ ¯ ) = ( f ( ρ ) ρ 2 ) ,

due to the fact P ( ρ ¯ ) = 0 . Here, f ( ) is smooth and bounded.

Then, based on Corollary 2.2 and (4.6)–(4.8), we shall prove M ( T ) C by reformulating the low-frequency of the derivatives of the solution for 2 j k as follows:

(4.10) x j U L ( t ) L 2 ( 1 + t ) 3 4 j 2 U L ( 0 ) L 1 + 0 t 2 ( 1 + t τ ) 3 4 j + 1 2 ( f ( ρ ) ρ 2 ) L 1 ( τ ) J 1 + t 2 t ( 1 + t τ ) 1 2 x j 1 ( P ˜ ( ρ ) ρ ) L 2 ( τ ) J 2 + 0 t 2 ( 1 + t τ ) 3 4 j 2 F ˜ ( U ) L 1 ( τ ) J 3

+ t 2 t ( 1 + t τ ) 1 2 x j 1 ( m x m ) ( τ ) L 2 J 4 + t 2 t ( 1 + t τ ) 1 x j 2 x 2 ( ρ m ) ( τ ) L 2 J 5 + t 2 t ( 1 + t τ ) 1 x j 2 ( ρ x 3 ρ ) ( τ ) L 2 J 6 .

According to Proposition 3.1, (4.4), the Young inequality, and Sobolev inequality, we have the following estimates for J i :

(4.11) f ( ρ ) ρ 2 L 1 ( τ ) M ( t ) 2 ( 1 + τ ) 1 2 , x j 1 ( f ( ρ ) ρ 2 ) L 2 ( τ ) ρ L x j 1 ρ L 2 δ M 3 2 ( t ) ( 1 + τ ) 5 4 j 2 , F ˜ 1 ( U ) L 1 ( τ ) m L 2 ( x m , x 2 ρ ) L 2 + ρ L 2 ( x 2 m , x 3 ρ ) L 2 δ M ( t ) ( 1 + τ ) 5 4 , x j 1 ( m x m ) L 2 ( τ ) x j 1 m L 6 x m L 3 + m L x j m L 2 x j m L 2 x m L 2 1 2 x 2 m L 2 1 2 + m L 2 1 4 x 2 m L 2 3 4 x j m L 2 ( 1 + τ ) 3 4 j 2 δ 1 2 M ( t ) 3 2 ( 1 + τ ) 7 8 + ( 1 + τ ) 3 4 j 2 δ 1 2 M ( t ) 3 2 ( 1 + τ ) 3 4 , x j 2 ( x 2 ( ρ m ) ) L 2 ( τ ) x j ρ L 6 m L 3 + x j m L 2 ρ L x j + 1 ρ L 2 m L 2 1 2 x m L 2 1 2 + x j m L 2 ρ L 2 1 4 x 2 ρ L 2 3 4 ( 1 + τ ) 3 4 j 2 δ 1 2 M ( t ) 3 2 ( 1 + τ ) 5 8 , x j 2 ( ρ x 3 ρ ) L 2 ( τ ) x j 2 ρ L x 3 ρ L 2 + x j + 1 ρ L 2 ρ L x j 2 ρ L 2 1 4 x j ρ L 2 3 4 x 3 ρ L 2 + ρ L 2 1 4 x 2 ρ L 2 3 4 x j + 1 ρ L 2 ( 1 + τ ) 3 4 j 2 δ 1 2 M ( t ) 3 2 ( 1 + τ ) 1 8 .

By inserting these estimates into (4.10), one has

(4.12) x j U L ( t ) L 2 ( 1 + t ) 3 4 j 2 U L ( 0 ) L 1 + U ( 0 ) H k B ˙ 2,1 3 2 + δ M ( t ) + δ 1 2 M 3 2 ( t ) ,

which together with (4.2) gives for 2 j k that

(4.13) x j U ( t ) L 2 2 C e C 0 t x j U ( 0 ) 2 + C U L ( 0 ) L 1 2 + U ( 0 ) H k B ˙ 2,1 3 2 2 + δ 2 M 2 ( t ) + δ M 3 ( t ) ( 1 + t ) 3 4 j 2 .

In the same way, one can obtain the following decay rate for the lower order derivatives of the solution without any new difficulty:

x j U ( t ) L 2 2 C U L ( 0 ) L 1 2 + U ( 0 ) H k B ˙ 2,1 3 2 2 + δ 2 M 2 ( t ) + δ M 3 ( t ) ( 1 + t ) 3 2 j , j = 0 , 1

and

ρ ( t ) L 2 2 C U L ( 0 ) L 1 2 + U ( 0 ) H k B ˙ 2,1 3 2 2 + δ 2 M 2 ( t ) + δ M 3 ( t ) ( 1 + t ) 1 2 .

In summary, from the definition of M ( t ) and the smallness of δ , we can immediately conclude that there exists a positive constant C ¯ such that

(4.14) M 2 ( t ) C ¯ 2 { U L ( 0 ) L 1 2 + U ( 0 ) H k B ˙ 2,1 3 2 2 + δ 2 M 2 ( t ) + δ M 3 ( t ) } .

By using the Young inequality, one knows that there exist a positive constant K 0 dependent of the initial data and a positive constant C δ 1 when δ 1 such that

(4.15) M 2 ( t ) K 0 + C δ M 4 ( t ) .

Noting that M ( t ) is continuous and non-decreasing, one can conclude that M ( t ) C . This proves (4.1).□

4.2 Method 2

We basically use the pure energy method introduced by Guo and Wang [11] for the compressible Navier-Stokes equations to derive the L 2 -decay rate of the solution under the same initial assumption in the negative Besov space B ˙ 2 , d 2 ( R d ) as in [3]. Unfortunately, we can only deal with the case d 4 due to the slower decay rate of the density ρ than the heat kernel.

For simplicity, we only give the sketch of the proof.

Step 1. Prove the following estimate of the solution in the negative Besov space:

(4.16) 1 2 d d t { κ ρ ¯ ρ ( t ) B ˙ 2 , d 2 2 + m B ˙ 2 , d 2 2 } + μ 2 ρ ¯ m B ˙ 2 , d 2 2 + μ + μ 2 ρ ¯ div m B ˙ 2 , d 2 2 ( ρ , m ) L 2 ( ρ H 2 + m H 1 d 2 s + 1 ) m B ˙ 2 , s .

The proof of the above estimate is standard. In particular, by applying Δ ˙ j with j Z to equation (2.1), multiplying the resulting identity by Δ ˙ j Δ ρ , Δ ˙ j m , respectively, and integrating over R d by parts, one has

1 2 d d t { κ ρ ¯ Δ ˙ j ρ ( t ) L 2 2 + Δ ˙ j m L 2 2 } + μ ρ ¯ Δ ˙ j m L 2 2 + μ + μ ρ ¯ Δ ˙ j div m L 2 2 = Δ ˙ j m , Δ ˙ j F 1 = Δ ˙ j m , Δ ˙ j div m m ρ + ρ ¯ + μ ρ ¯ Δ ˙ j Δ ρ m ρ + ρ ¯ + μ + μ ρ ¯ Δ ˙ j div ρ m ρ + ρ ¯ + Δ ˙ j P ( ρ + ρ ¯ ) + κ Δ ˙ j ( ρ Δ ρ ) .

Further, multiplying the equality above by 2 d j and then taking the supremum over j Z , one can immediately obtain (4.16), where we have used Lemma A.1, Lemma A.5, and Hölder’s inequality.

Step 2. Derive the optimal L 2 -decay rate of the derivatives for the solution under the condition ( ρ ( t ) , m ( t ) ) B ˙ 2 , d 2 C . In fact, from the energy estimates at each j th level ( j 0 ) in (3.12), we know that there exists a positive constant C 0 such that

(4.17) d d t ( x j ρ , x j m ) L 2 2 + C 0 ( x j + 1 ρ , x j + 1 m ) L 2 2 0 .

If ( ρ ( t ) , m ( t ) ) B ˙ 2 , d 2 C , then by using Lemma A.4, one has

(4.18) ( x j + 1 ρ , x j + 1 m ) L 2 2 C 1 ( x j ρ , x j m ) L 2 1 + 1 j + d 2 ,

since the fact

(4.19) x j f L 2 f B ˙ 2 , d 2 1 j + 1 + d 2 x j + 1 f L 2 j + d 2 j + 1 + d 2 .

Combining (4.17) and (4.18), we obtain the following decay rate of the solution:

(4.20) x j ( ρ , m ) H k j C ( 1 + t ) j + d 2 2 , j = 0 , 1 , , k .

Step 3. Prove that ( ρ ( t ) , m ( t ) ) B ˙ 2 , d 2 C by using the estimate in the first step. Thus, according to (4.16), we know the key is to prove that

(4.21) 0 t ρ ( τ ) L 2 ρ ( τ ) L 2 d τ < ,

since the term ρ ( τ ) L 2 ρ ( τ ) L 2 in (4.16) has the lowest decay rate. To prove (4.21), we will use (4.19) again to have

(4.22) ρ ( τ ) L 2 ρ ( τ ) L 2 C ρ ( τ ) B ˙ 2 , ( d 2 1 ) 1 1 + d 2 1 ρ ( τ ) L 2 d 2 1 1 + d 2 1 ρ ( τ ) L 2 C Λ ρ ( τ ) B ˙ 2 , d 2 2 d ρ ( τ ) L 2 d ( d 1 ) 2 d C Λ ρ ( τ ) B ˙ 2 , d 2 2 d ( 1 + τ ) d 1 2 ,

which is integrable with respective to the time t when d 4 . Then, (4.16) and (4.21) suffice to yield ( ρ ( t ) , m ( t ) ) B ˙ 2 , d 2 C when d 4 . This proves the L 2 -decay rate of the solution when d 4 in Theorem 1.2.

5 Pointwise estimates for nonlinear system

In this section, we mainly deduce the pointwise description of the solution for the nonlinear problem when t 1 . In fact, when t 1 , one can put any order derivatives on Green’s function, which can largely relax the regularity assumption on the initial data. This is one difference from our previous pointwise results for other fluid models. Of course, if one want to obtain the pointwise estimate for t 0 , it also needs higher regularity assumption on the initial data due to the singularity at t = 0 and the quasi-linearity of the system.

First of all, by using Duhamel’s principle, we can obtain the representation of the perturbation ( ρ , m ) for the nonlinear problem (A):

(5.1) x k ρ = x k 1 G 11 Λ ρ 0 + x k G 12 m 0 + 0 t x k G 12 ( , t τ ) F ( , τ ) d τ , x k m = x k 1 G 21 Λ ρ 0 + x k G 22 m 0 + 0 t x k G 22 ( , t τ ) F ( , τ ) d τ ,

where k 0 and the nonlinear term F is defined in (2.2).

Initial propagation. Use ( ρ ˘ , m ˘ ) to denote the linear part of the solution in (5.1). By Theorem 2.1 and the initial condition (1.9) together with the representation (5.1) yield the linear estimate for t 1 as follows:

(5.2) x k ρ ˘ C ε ( 1 + t ) d 1 + k 2 1 + x 2 1 + t d 1 2 , when 2 μ + μ ρ ¯ 2 κ ρ ¯ , x k ρ ˘ C ε ( 1 + t ) d 1 + k 2 1 + x 2 1 + t r , when 2 μ + μ ρ ¯ = 2 κ ρ ¯ , x k m ˘ C ε ( 1 + t ) d + k 2 1 + x 2 1 + t r , r > d 2 .

Nonlinear Coupling. According to the above initial propagation, we should give the ansatz for the nonlinear problem when k = 0,1 :

(5.3) x k ρ 2 C ε ( 1 + t ) d 1 + k 2 1 + x 2 1 + t d 2 , when 2 μ + μ ρ ¯ 2 κ ρ ¯ , x k ρ 2 C ε ( 1 + t ) d 1 + k 2 1 + x 2 1 + t r , when 2 μ + μ ρ ¯ = 2 κ ρ ¯ , m 2 C ε ( 1 + t ) d 2 1 + x 2 1 + t r , r > d 2 .

Use ( ρ ˘ ˘ , m ˘ ˘ ) to denote the nonlinear part of the solution in (5.1). First, recall that

(5.4) F = O ( 1 ) x ( m 2 + x ( ρ m ) + ρ 2 + ρ x 2 ( ρ ) + ( x ρ ) 2 ) .

We only consider the case 2 μ + μ ρ ¯ 2 κ ρ ¯ since the other case can be treated similarly. For t 1 , one can put one derivative on the heat kernel when estimating the nonlinear convolution. Hence, we have for k = 0 , 1 that

x k ρ ˘ ˘ C 0 t ( 1 + t τ ) d + k 2 1 + x y 2 1 + t τ d 2 m 2 + x ( ρ m ) + ρ 2 + ρ x 2 ( ρ ) + ( x ρ ) 2 ( y , τ ) d y d τ C 0 t ( 1 + t τ ) d + k 2 1 + x y 2 1 + t τ d 2 ( 1 + τ ) d 2 1 + y 2 1 + τ d 2 ( x m + x 2 ρ ) ( y , τ ) d y d τ + C 0 t ( 1 + t τ ) d + k 2 1 + x y 2 1 + t τ d 2 ( 1 + τ ) ( d 1 ) 1 + y 2 1 + τ ( d 1 ) d y d τ C ( 1 + t ) d + k 2 1 + x 2 1 + t d 2 .

Here, we only need to use the L 2 -decay rate of x m L 2 + x 2 ρ L 2 C ( 1 + t ) d 4 , which has been given in Theorem 1.2. The pointwise estimate of m ˘ ˘ can be derived similarly. Hence, we have completed the proof of Theorem 1.3.

Finally, we shall remark the pointwise description of the solution when t 0 . In fact, one only needs to impose higher regularity on the initial data due to the singularity of Green’s function at t = 0 and the quasi-linearity of the system. For instance, when d = 3 , if the assumptions in Theorem 1.2 hold with k 4 , one can also obtain the pointwise estimates as in (5.3) for t 0 .

Acknowledgements

The authors express their gratitude to the anonymous referees for their helpful comments and suggestions.

  1. Funding information: The research was supported by the National Natural Science Foundation of China (No. 11971100) and Natural Science Foundation of Shanghai (Grant No. 22ZR1402300).

  2. Author contributions: Zhigang Wu designed research; Zhigang Wu and Mengqian Liu performed research; Zhigang Wu and Mengqian Liu wrote the article and Zhigang Wu revised the article.

  3. Conflict of interest: The authors declare that they have no conflict of interest.

  4. Data availability statement: All data generated or analysed during this study are included in this published article.

Appendix

In this section, we give some Sobolev inequalities and Besov inequalities, which have been used in the above sections. In the following, d denotes the dimension of spacial variable x .

Lemma A.1

(Gagliardo-Nirenberg inequality). Let 0 m , k l , then we have

x k g L p C x m g L q 1 α x l g L r α ,

where k satisfies

1 p k d = ( 1 α ) 1 q m d + α 1 r l d .

Lemma A.2

[11] Let s 0 and l 0 , then we have

x l g L 2 C x l + 1 g L 2 1 α g H ˙ s α , where α = 1 l + s + 1 .

Lemma A.3

[34] Let 0 < s < d , 1 < p < q < , 1 q + s n = 1 p , then

Λ s g L q C g L p .

Next, we give some lemmas on Besov space B ˙ 2 , s .

Lemma A.4

[32] Suppose k 0 and s > 0 , then we have

x k f L 2 C x k + 1 f L 2 1 α f B ˙ 2 , s α , w h e r e α = 1 l + 1 + s .

Lemma A.5

[32] Suppose that s > 0 and 1 p < 2 . We have the embedding L p B ˙ q , s with 1 2 + s d = 1 p . In particular, we have

f B ˙ 2 , s C f L p .

Lemma A.6

[39,41] If t R + , x R d with d 2 and f ( x , t ) C 1 + x 2 1 + t r with r > d 2 , then

( Δ ) 1 f ( x , t ) C ( 1 + t ) 1 2 1 + x 2 1 + t d 1 2 .

The following lemma is usually used to deal with the terms containing Calderon-Zygmund operator with the symbol ξ ξ T ξ 2 in the low-frequency part of Green’s function.

Lemma A.7

[39,41] Suppose that f ˆ ( ξ , t ) = ξ ξ T ξ 2 χ 1 ( ξ ) e ξ 2 t + O ( ξ 3 ) t in the Fourier space and χ 1 ( ξ ) is the cutoff function for the low-frequency part. Then, it holds that

1 ( ξ α f ˆ ( ξ , t ) ) C ( 1 + t ) d + α 2 1 + x 2 1 + t d + α 2 .

References

[1] H. Bahouri, J. Chemin, and R. Danchin, Fourier Analysis and Nonlinear Partial Differential Equations, Grundlehren Math. Wiss., vol. 343, Springer-Verlag, Berlin, Heidelberg, 2011, http://dx.doi.org/10.1007/978-3-642-16830-7.Suche in Google Scholar

[2] D. Bresch, B. Desjardins, and C. K. Lin, On some compressible fluid models: Korteweg, lubrication and shallow water systems, Comm. Partial Differential Equations 28 (2003), no. 3–4, 843–868, https://doi.org/10.1081/PDE-120020499. Suche in Google Scholar

[3] N. Chikami and T. Kobayashi, Global well-posedness and time-decay estimates of the compressible Navier-Stokes-Korteweg system in critical Besov spaces, J. Math. Fluid Mechanics 21 (2019), 31, https://doi.org/10.1007/s00021-019-0431-8. Suche in Google Scholar

[4] R. Danchin and B. Desjardins, Existence of solutions for compressible fluid models of Korteweg type, Ann. Inst. H. Poincare Anal. Non Lineaire 18 (2001), 97–133, https://doi.org/10.1016/S0294-1449(00)00056-1. Suche in Google Scholar

[5] S. J. Deng and S. H. Yu, Green’s function and pointwise convergence for compressible Navier-Stokes equations, Quarterly of Applied Mathematics 75 (2017), no. 3, 433–503, http://dx.doi.org/10.1090/qam/1461. Suche in Google Scholar

[6] L. L. Du and Z. G. Wu, Solving the non-isentropic Navier-Stokes equations in odd space dimensions: The Green function method, J. Math. Phys. 58 (2017), no. 2, 101506, https://doi.org/10.1063/1.5005915. Suche in Google Scholar

[7] R. J. Duan, H. X. Liu, S. Ukai, and T. Yang, Optimal Lp-Lq convergence rate for the compressible Navier-Stokes equations with potential force, J. Diff. Eqns. 238 (2007), 220–233, https://doi.org/10.1016/j.jde.2007.03.008. Suche in Google Scholar

[8] R. J. Duan, S. Ukai, T. Yang, and H. J. Zhao, Optimal convergence rate for the compressible Navier-Stokes equations with potential force, Math. Models Methods Appl. Sci. 17 (2007), 737–758, https://doi.org/10.1142/S021820250700208X.Suche in Google Scholar

[9] J. E. Dunn and J. Serrin, On the thermomechanics of interstitial working, Arch. Ration. Mech. Anal. 88 (1985), no. 2, 95–133, https://doi.org/10.1007/BF00250907. Suche in Google Scholar

[10] J. C. Gao, Z. Yang, and Z. A. Yao, Long-time behavior of solution for the compressible Navier-Stokes-Korteweg equations in R3, Appl. Math. Letters 48 (2015), 30–35, https://doi.org/10.1016/j.aml.2015.03.006. Suche in Google Scholar

[11] Y. Guo and Y. J. Wang, Decay of dissipative equations and negative Sobolev spaces, Comm. Partial Differential Equations 37 (2012), 2165–2208, https://doi.org/10.1080/03605302.2012.696296. Suche in Google Scholar

[12] B. Haspot, Existence of global weak solution for compressible fluid models of Korteweg type, J. Math. Fluid Mech. 13 (2009), no. 2, 223–249, https://doi.org/10.1007/s00021-009-0013-2. Suche in Google Scholar

[13] H. Hattori and D. Liii, Solutions for two dimensional system for materials of Korteweg type, SIAM J. Math. Anal. 25 (1994), 85–98, https://doi.org/10.1137/S003614109223413X. Suche in Google Scholar

[14] H. Hattori and D. Li, Global solutions of a high-dimensional system for Korteweg materials, J. Math. Anal. Appl. 198 (1996), no. 2, 84–97, https://doi.org/10.1006/jmaa.1996.0069. Suche in Google Scholar

[15] D. Hoff and K. Zumbrun, Pointwise decay estimates for multidimensional Navier-Stokes diffusion waves, Z. Angew. Math. Phys. 48 (1997), 597–614, https://doi.org/10.1007/s000330050049. Suche in Google Scholar

[16] X. F. Hou, H. Y. Peng, and C. J. Zhu, Global well-posedness of the 3D non-isothermal compressible fluid model of Korteweg type, Nonlinear Analysis: Real World Applications 43 (2018), 18–53, https://doi.org/10.1016/j.nonrwa.2018.02.002. Suche in Google Scholar

[17] X. F. Hou, H. Y. Peng, and C. J. Zhu, Global classical solutions to the 3D Navier-Stokes-Korteweg equations with small initial energy, Anal. Appl. 16 (2018), no. 1, 55–84, https://doi.org/10.1142/S0219530516500123. Suche in Google Scholar

[18] X. F. Hou, L. Yao, and C. J. Zhu, Vanishing capillarity limit of the compressible non-isentropic Navier-Stokes-Korteweg system to Navier-Stokes system, J. Math. Anal. Appl. 448 (2017), no. 1, 421–446, https://doi.org/10.1016/j.jmaa.2016.11.014. Suche in Google Scholar

[19] X. P. Jiang and Z. G. Wu, Pointwise space-time behavior of a compressible Navier-Stokes-Korteweg system in dimension three, Acta Math. Sci. 42B (2022), no. 5, 2113–2130, https://doi.org/10.1007/s10473-022-0522-0. Suche in Google Scholar

[20] S. Kawashima, System of a Hyperbolic-Parabolic Composite Type, with Applicationsto the Equations of Manetohydrodynamics, Doctoral thesis, Kyoto University, Kyoto, 1983, https://doi.org/10.14989/doctor.k3193. Suche in Google Scholar

[21] S. Kawashima, Y. Shibata, and J. Xu, The Lp energy methods and decay for the compressible Navier-Stokes equations with capillarity, J. Math. Pures Appl. 154 (2021), 146–184, https://doi.org/10.1016/j.matpur.2021.08.009. Suche in Google Scholar

[22] T. Kobayashi and Y. Shibata, Remark on the rate of decay of solutions to linearized compressible Navier-Stokes equations, Pacific J. Math. 207 (2002), no. 1, 199–234, https://doi.org/10.2140/pjm.2002.207.199. Suche in Google Scholar

[23] D. J. Korteweg, Sur la forme que prennent les équations du mouvement des fluides siĺon tient compte des forces capillaires causées par des variations de densité considérables mais continues et sur la théorie de la capillarité dansĺhypothèse daune variation continue de la densité, Archives Néerlandaises de Sciences Exactes et Naturelles 6 (1901), 1–24. Suche in Google Scholar

[24] M. Kotschote, Strong solutions for a compressible fluid model of Korteweg type, Ann. Inst. H. Poincare Anal. Non Lineaire 25 (2008), no. 4, 679–696, https://doi.org/10.1016/j.anihpc.2007.03.005. Suche in Google Scholar

[25] H. L. Li and T. Zhang, Large time behavior of isentropic compressible Navier-Stokes system in R3, Math. Meth. Appl. Sci. 34 (2011), 670–682, https://doi.org/10.1002/mma.1391. Suche in Google Scholar

[26] H. L. Li, T. Yang, and M. Y. Zhong, Green’s function and pointwise space-time behaviors of the Vlasov-Poisson-Boltzmann system, Arch. Ration. Mech. Anal. 235 (2019), 1–47, https://doi.org/10.1007/s00205-019-01438-w. Suche in Google Scholar

[27] T. P. Liu and S. E. Noh, Wave propagation for the compressible Navier-Stokes equations, J. Hyperbolic Differ. Eqns. 12 (2015), 385–445, https://doi.org/10.1142/S021989161550011. Suche in Google Scholar

[28] T. P. Liu and W. K. Wang, The pointwise estimates of diffusion wave for the Navier-Stokes systems in odd multi-dimension, Comm. Math. Phys. 196 (1998), 145–173, https://doi.org/10.1007/s002200050418. Suche in Google Scholar

[29] T. P. Liu and Y. N. Zeng, Large time behavior of solutions for general quasilinear hyperbolic-parabolic systems of conservation laws, vol. 125, Memoirs of the American Mathematical Society, Providence, RI, 1997, https://doi.org/10.1090/memo/0599. Suche in Google Scholar

[30] A. Matsumura and T. Nishida, The initial value problems for the equations of motion of viscous and heat-conductive gases, J. Math. Kyoto Univ. 20 (1980), 67–104, https://doi.org/10.1215/kjm/1250522322. Suche in Google Scholar

[31] H. Saito, On the maximal Lp-Lq regularity for a compressible fluid model of Korteweg type on general domains, J. Diff. Eqns. 268 (2020), no. 6, 2802–2851, https://doi.org/10.1016/j.jde.2019.09.040. Suche in Google Scholar

[32] V. Sohinger and R. M. Strain, The Boltzmann equation, Besov spaces, and optimal time decay rates in the whole space, Adv. Math. 261 (2014), 274–332, https://doi.org/10.1016/j.aim.2014.04.012. Suche in Google Scholar

[33] Z. H. Song and J. Xu, Global existence and analyticity of Lp solutions to the compressible fluid model of Korteweg type, J. Diff. Eqns. 370 (2023), 101–139, https://doi.org/10.1016/j.jde.2023.06.011. Suche in Google Scholar

[34] E. M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton University Press, Princeton, NJ, 1970, https://www.jstor.org/stable/j.ctt1bpmb07. Suche in Google Scholar

[35] W. K. Wang and T. Yang, The pointwise estimates of solutions for Euler equations with damping in multi-dimensions, J. Diff. Eqns. 173 (2001), 410–450, https://doi.org/10.1006/jdeq.2000.3937. Suche in Google Scholar

[36] W. K. Wang and Z. G. Wu, Pointwise estimates of solution for the Navier-Stokes-Poisson equations in multi-dimensions, J. Diff. Eqns. 248 (2010), 1617–1636, https://doi.org/10.1016/j.jde.2010.01.003. Suche in Google Scholar

[37] W. J. Wang and W. K. Wang, Decay rates of the compressible Navier-Stokes-Korteweg equations with potential forces, Discrete Contin. Dyn. Syst. 1 (2015), no. 35, 513–536, https://doi.org/10.3934/dcds.2015.35.513. Suche in Google Scholar

[38] Y. J. Wang and Z. Tan, Optimal decay rates for the compressible fluid models of Korteweg type, J. Math. Anal. Appl. 379 (2011), no. 1, 256–271, https://doi.org/10.1016/j.jmaa.2011.01.006. Suche in Google Scholar

[39] Z. G. Wu and W. K. Wang, Pointwise estimates for bipolar compressible Navier-Stokes-Poisson system in dimension three, Arch. Rational Mech. Anal. 226 (2017), no. 2, 587–638, https://doi.org/10.1007/s00205-017-1140-1. Suche in Google Scholar

[40] Z. G. Wu and W. K. Wang, Pointwise estimates of solution for non-isentropic Navier-Stokes-Poisson equations in multidimensions, Acta Math. Sci. 32B (2012), no. 5, 1681–1702, https://doi.org/10.1016/S0252-9602(12)60134-9. Suche in Google Scholar

[41] Z. G. Wu and W. K. Wang, Space-time estimates of 3D bipolar compressible Navier-Stokes-Poisson system with unequal viscosities, Sci. China Math. 67(2024), 1059–1084, https://doi.org/10.1007/s11425-022-2130-2. Suche in Google Scholar

[42] Z. G. Wu and Y. P. Li, Pointwise estimates of solutions for the multi-dimensional bipolar Euler-Poisson system, Z. angew. Math. Phys. 67 (2016), no. 3, 50, https://doi.org/10.1007/s00033-016-0651-1. Suche in Google Scholar

[43] S. H. Yu, Nonlinear wave propagation over a Boltzmann shock profile, J. Amer. Math. Soc. 23 (2010), no. 4, 1040–1118, https://www.jstor.org/stable/20753579. 10.1090/S0894-0347-2010-00671-6Suche in Google Scholar

[44] Y. N. Zeng, Thermal non-equilibrium flows in three space dimensions, Arch. Rational Mech. Anal. 219 (2016), 27–87, https://doi.org/10.1007/s00205-015-0892-8. Suche in Google Scholar

[45] Y. N. Zeng, L1 asymptotic behavior of compressible isentropic viscous 1-D flow, Comm. Pure Appl. Math. 47 (1994), 1053–1082, https://doi.org/10.1002/cpa.3160470804. Suche in Google Scholar

Received: 2023-04-28
Revised: 2023-12-10
Accepted: 2025-02-24
Published Online: 2025-04-16

© 2025 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Artikel in diesem Heft

  1. Research Articles
  2. Incompressible limit for the compressible viscoelastic fluids in critical space
  3. Concentrating solutions for double critical fractional Schrödinger-Poisson system with p-Laplacian in ℝ3
  4. Intervals of bifurcation points for semilinear elliptic problems
  5. On multiplicity of solutions to nonlinear Dirac equation with local super-quadratic growth
  6. Entire radial bounded solutions for Leray-Lions equations of (p, q)-type
  7. Combinatorial pth Calabi flows for total geodesic curvatures in spherical background geometry
  8. Sharp upper bounds for the capacity in the hyperbolic and Euclidean spaces
  9. Positive solutions for asymptotically linear Schrödinger equation on hyperbolic space
  10. Existence and non-degeneracy of the normalized spike solutions to the fractional Schrödinger equations
  11. Existence results for non-coercive problems
  12. Ground states for Schrödinger-Poisson system with zero mass and the Coulomb critical exponent
  13. Geometric characterization of generalized Hajłasz-Sobolev embedding domains
  14. Subharmonic solutions of first-order Hamiltonian systems
  15. Sharp asymptotic expansions of entire large solutions to a class of k-Hessian equations with weights
  16. Stability of pyramidal traveling fronts in time-periodic reaction–diffusion equations with degenerate monostable and ignition nonlinearities
  17. Ground-state solutions for fractional Kirchhoff-Choquard equations with critical growth
  18. Existence results of variable exponent double-phase multivalued elliptic inequalities with logarithmic perturbation and convections
  19. Homoclinic solutions in periodic partial difference equations
  20. Classical solution for compressible Navier-Stokes-Korteweg equations with zero sound speed
  21. Properties of minimizers for L2-subcritical Kirchhoff energy functionals
  22. Asymptotic behavior of global mild solutions to the Keller-Segel-Navier-Stokes system in Lorentz spaces
  23. Blow-up solutions to the Hartree-Fock type Schrödinger equation with the critical rotational speed
  24. Qualitative properties of solutions for dual fractional parabolic equations involving nonlocal Monge-Ampère operator
  25. Regularity for double-phase functionals with nearly linear growth and two modulating coefficients
  26. Uniform boundedness and compactness for the commutator of an extension of Riesz transform on stratified Lie groups
  27. Normalized solutions to nonlinear Schrödinger equations with mixed fractional Laplacians
  28. Existence of positive radial solutions of general quasilinear elliptic systems
  29. Low Mach number and non-resistive limit of magnetohydrodynamic equations with large temperature variations in general bounded domains
  30. Sharp viscous shock waves for relaxation model with degeneracy
  31. Bifurcation and multiplicity results for critical problems involving the p-Grushin operator
  32. Asymptotic behavior of solutions of a free boundary model with seasonal succession and impulsive harvesting
  33. Blowing-up solutions concentrated along minimal submanifolds for some supercritical Hamiltonian systems on Riemannian manifolds
  34. Stability of rarefaction wave for relaxed compressible Navier-Stokes equations with density-dependent viscosity
  35. Singularity for the macroscopic production model with Chaplygin gas
  36. Global strong solution of compressible flow with spherically symmetric data and density-dependent viscosities
  37. Global dynamics of population-toxicant models with nonlocal dispersals
  38. α-Mean curvature flow of non-compact complete convex hypersurfaces and the evolution of level sets
  39. High-energy solutions for coupled Schrödinger systems with critical growth and lack of compactness
  40. On the structure and lifespan of smooth solutions for the two-dimensional hyperbolic geometric flow equation
  41. Well-posedness for physical vacuum free boundary problem of compressible Euler equations with time-dependent damping
  42. On the existence of solutions of infinite systems of Volterra-Hammerstein-Stieltjes integral equations
  43. Remark on the analyticity of the fractional Fokker-Planck equation
  44. Continuous dependence on initial data for damped fourth-order wave equation with strain term
  45. Unilateral problems for quasilinear operators with fractional Riesz gradients
  46. Boundedness of solutions to quasilinear elliptic systems
  47. Existence of positive solutions for critical p-Laplacian equation with critical Neumann boundary condition
  48. Non-local diffusion and pulse intervention in a faecal-oral model with moving infected fronts
Heruntergeladen am 30.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/anona-2025-0078/html
Button zum nach oben scrollen