Asymptotical Behavior of Global Solutions of the Navier–Stokes–Korteweg Equations with Respect to
Capillarity Number at Infinity
Abstract
Vanishing capillarity in the Navier–Stokes–Korteweg (NSK) equations has been widely investigated, in particular, it is well-known that the NSK equations converge to the Navier–Stokes (NS) equations by vanishing capillarity number. To our best knowledge, this paper first investigates the behavior of large capillary number, denoted by , for the global(-in-time) strong solutions with small initial perturbations of the three-dimensional (3D) NSK equations in a slab domain with Navier(-slip) boundary condition. Under the well-prepared initial data, we can construct a family of global strong solutions of the 3D incompressible NSK equations with respect to , where the solutions converge to a unique solution of 2D incompressible NS-like equations as goes to infinity.
keywords:
Navier–Stokes–Korteweg equations; asymptotical behavior; uniform estimates; global-in-time solutions.1 Introduction
A classical model to describe the dynamics of an inhomogeneous incompressible fluid endowed with internal capillarity (in the diffuse interface setting) is the following general system of incompressible Navier–Stokes–Korteweg (NSK) equations:
(1.1) |
where , and denote the density, velocity and kinetic pressure of the fluid resp. at the spacial position for time . The differential operator is defined by , where the superscript represents the transposition. The shear viscosity function , the capillarity function are known smooth functions , and satisfy , . The general capillary tensor is written as
where is the identity matrix. To conveniently investigate the asymptotical behavior of capillarity number, we considers that and are positive constants [6], thus it holds that
(1.2) |
where we have defined that . We call the parameters and the viscosity coefficient and the capillarity number, resp..
In classical hydrodynamics, the interface between two immiscible compressible/incompressible fluids is modeled as a free boundary which evolves in time. The equations describing the motion of each fluid are supplemented by boundary conditions at the free surface involving the physical properties of the interface. For instance, in the free-boundary formulation, it is assumed that the interface has an internal surface tension. However, when the interfacial thickness is comparable to the length scale of the phenomena being examined, the free-boundary description breaks down. Diffuse-interface models provide an alternative description where the surface tension is expressed in its simplest form as , i.e., the capillary tension which was introduced by Korteweg in 1901 [41]. Later, its modern form was derived by Dunn and Serrin [19].
In the physical view, it can serve as a phase transition model to describe the motion of compressible fluid with capillarity effect. Owing to its importance in mathematics and physics, there has been profuse works for the mathematical theory of the corresponding compressible NSK system, for example, see [2, 7, 24] for the global(-in-time) weak solutions with large initial data, [17, 25, 27, 40, 50] for global strong solutions with small initial data, [12, 26, 42, 43] for local(-in-time) strong/classical solutions with large initial data, [16] for stationary solutions, [20] for highly rotating limit, [15, 28] for the stability of viscous shock wave, [54] for the maximal – regularity theory, [9, 13, 39] for the decay-in-time of global solutions, [5, 14, 21, 38, 30] for the vanishing capillarity limit, [44, 45, 29] for the nonisentropic case, [48, 55, 37] for the low Mach number limit, [3, 4, 8, 57] for the inviscid case, and so on.
The incompressible NSK system (1.1) has been also widely investigated. The local existence of a unique strong solution was obtained by Tan–Wang [58]. Burtea–Charve also established the existence result of strong solutions with small initial perturbations in Lagrangian coordinates [11], and further presented that the lifespan goes to infinity as the capillarity number goes to zero. Yang–Yao–Zhu proved that as both capillarity number and viscosity coefficient vanish, local solutions of the Cauchy problem for the incompressible NSK system converge to the one of the inhomogeneous incompressible Euler equations [60]. Recently a similar result was also established by Wang–Zhang for the fluid domain being a horizontally periodic slab with Navier-slip boundary conditions [59].
In addition, it also has been investigated that the capillarity under the large capillarity number stabilizes the motion of the incompressible fluid endowed with internal capillarity. Bresch–Desjardins–Gisclon–Sart derived that the capillarity slows down the growth rate of linear Rayleigh–Taylor (RT) instability, which occurs when a denser fluid lies on top of a lighter fluid under gravity [6]. Li–Zhang proved that the capillarity can inhibit RT instability for properly large capillarity number by the linearized motion equations [46] (Interesting readers further refer to [51, 46, 62, 61] for the existence of RT instability solutions in the incompressible NSK system with small capillarity number). Later, motivated by the result of magnetic tension inhibiting the RT instability in the two-dimensional (2D) non-resistive magnetohydrodynamic (MHD) fluid in [33], Jiang–Li–Zhang mathematically proved that the capillarity inhibits RT instability in the 2D NSK equations in the Lagrangian coordinates [34]. Recently Jiang–Zhang–Zhang further verified such inhibition phenomenon for the 3D case in Eulerian coordinates [35]. Jiang–Zhang–Zhang’s result rigorously presents that the capillarity has the stabilizing effect as well as the magnetic tension in the MHD fluid.
It is well-known that the larger the magnetic tension, the stronger the field intensity in the MHD fluid. Moreover the local solutions of the system of the idea MHD fluid with well-prepared initial data tend to a solution of a 2D Euler flow coupled with a linear transport equation as the field intensity goes to infinity, see [36, 22, 10] for relevant results. This naturally motivates us to except a similar result in the incompressible NSK system. In this paper, under the well-prepared initial data, we also construct a family of the global strong solutions of the 3D incompressible NSK equations with respect to , where the solutions converge to a unique solution of a 2D incompressible Navier–Stokes-like equations as goes to infinity, see Theorem 2.2. As Lin pointed out in [49], the stratified fluids with the internal surface tension, the viscoelastic fluids, the non-resistive MHD fluids and the diffuse-interface model (or NSK model) can also be regarded as complex fluids with elasticity. Our mathematical result in Theorem 2.2 supports that the capillarity has the role of elasticity again.
2 Main results
This section is devoted to introducing our main results in details. Let us consider a rest stat of the system (1.1). We choose the equilibrium density in the rest state to satisfy
(2.1) |
Without loss of generality, we consider
(2.2) |
Denoting the perturbations around the rest state by
and then recalling the relation
we obtain the following perturbation system from (1.1) with (1.2):
(2.3) |
where represents the unit vector with the third component being and we have defined that . The initial condition reads as follows
(2.4) |
Here and in what follows, we always use the right superscript to emphasize the initial data.
We consider that the fluid domain denoted by is a slab, i.e.
The 2D domain , denoted by , is the boundary of . We focus on the following Navier(-slip) boundary conditions on the velocity on :
where denotes the unit outward normal to and the subscript “tan” means the tangential component of a vector (for example, ) [18, 47]. The Navier slip boundary conditions describe an interaction between a viscous fluid and a solid wall. Since is a slab domain, the Navier boundary conditions are equivalent to the boundary condition
(2.5) |
We mention that it is not clear to us whether our main results in Theorems 2.1 and 2.2 can be extended to the non-slip boundary condition of velocity.
However, to investigate the asymptotical behavior of solutions with respect to capillarity number at infinity and to except the vanishing of , we set
and then transform (2.3) into
(2.6) |
The corresponding initial condition reads as follows
(2.7) |
In addition, to avoid the vacuum of density, we assume that
(2.8) |
Before stating our main results, we shall introduce simplified notations, which will be used throughout this paper.
(1) Simplified basic notations: denotes an interval, in particular, . denotes the closure of a set with , in particular, and . means that for some constant , where at most depends on the domain , , , in (2.8), and may vary from line to line. Sometimes we also denote by for emphasizing that is fixed, where , . , where . Let be a vector function defined in a 3D domain, we define that and . , , , and . For the simplicity, we denote by , where represents a norm, and is scalar function or vector function for .
(2) Simplified Banach spaces, norms and semi-norms:
where is a real number, and , , , are integers.
(3) Simplified spaces of functions with values in a Banach space:
(4) Energy/dissipation functionals (generalized):
Now we state the first result, which presents that the initial-boundary value problem (2.5)–(2.7) with small initial data admits a unique global strong solution with uniform estimates with respect to .
Theorem 2.1.
Let , , be given, and be defined by (2.1). There exist constants , and , where
(2.9) |
such that, for any satisfying (2.8),
(2.10) |
the initial-boundary value problem (2.5)–(2.7) has a unique global strong solution
Moreover the solution satisfies the following estimates
(2.11) | ||||
(2.12) |
and
(2.13) |
where .
The existence of strong solutions with small initial perturbations to the initial-boundary value problem (2.3)–(2.5) in a horizontally periodic domain has been established by Jiang–Zhang–Zhang, where . Let us first roughly recall some key ideas in Jiang–Zhang–Zhang’s proof in [35].
The standing point of Jiang–Zhang–Zhang’s proof is the basic energy identity in differential version for the problem (2.3)–(2.5):
(2.14) |
We call (2.14) the zero-order energy estimate of . Then they further derived energy estimates for the high-order spacial derivatives and the temporal derivatives of . However, the integrals related to the nonlinear terms in the system (2.3) appear in the high-order energy estimates. We call such integrals the nonlinear integrals for simplicity. In particular, there exist a troublesome nonlinear integral , which is equal to due to the incompressible condition (2.3)3. Jiang–Zhang–Zhang naturally expect that enjoy a fine decay-in-time, which contributes to close the troublesome nonlinear integral. However one can not directly derive the decay-in-time of from high-order energy estimates due to the absence of the dissipation of . Fortunately, they can capture the dissipation of from the vortex equations, which can be obtained by applying to (2.3)2. The dissipation of , together with the horizontal periodicity of , results in the decay-in-time of and by energy method with extremely fine estimates.
Following Jiang–Zhang–Zhang’s ideas in [35] for the initial-boundary vaule problem (2.5)–(2.7), however it seems to be difficult to establish the desired decay-in-time of due to the unboundedness of the fluid domain . This results in that we shall develop a new alternative method to estimate for the troublesome nonlinear integral (recalling the linear relation ). More precisely, we use the transport equation (2.6)1 twice and the anisotropic Gagliardo–Nirenberg–Sobolev type estimate in A.6 to estimate , see (3.20) for the detailed derivation. Based on this new idea, we can refine the energy method in [35] to establish Theorem 2.1.
It is easy see from (2.9) that for reduces to . Consequently, we can make use of the uniform-in- estimates in (2.11)–(2.13) with to establish the asymptotical behavior with respect to capillarity number at infinity. More precisely, for any given , we choose initial data and , which satisfy (2.8) and (2.10) with in place of . In view of Theorem 2.1, the initial-boundary value problem (2.3)–(2.5) with in place of admits a global strong solution denoted by , where we have defined that . Moreover the solutions have the following asymptotical behavior with respect to at infinity
Theorem 2.2.
Let . We additionally assume that in as . There exist functions , , and such that, for any given ,
(2.15) | |||
(2.16) | |||
(2.17) | |||
(2.18) | |||
(2.19) | |||
(2.20) |
as . Moreover, , , and satisfy
(2.21) |
and
(2.22) |
Remark 2.1.
Modifying the proofs of Theorems 2.1 and 2.2 by further using the energy method with time-weight in [35], we also obtain the correspond results for the case of the horizontally periodic domain with finite height, i.e. . In addition, since for , we can not except the vanishing capillarity limit from Theorem 2.1.
3 Proof of Theorem 2.1
This section is devoted to the proof of the global(-in-time) solvability with uniform-in- estimates for the initial-boundary value problem (2.5)–(2.7). The key point is to a priori derive the uniform-in- estimate in (2.11). For this purpose, let be a fixed time and a solution to (2.5)–(2.7) on with initial data . Moreover, we assume that satisfies (2.8) and the solution is sufficiently regular so that the procedure of formal deduction makes sense.
3.1 Basic estimates
This section is devoted to deriving some basic estimates of from the initial-boundary value problem (2.5)–(2.7). Let us first recall the energy identity of .
Proof.
Next we further extend (3.1) to both the cases satisfied by the highest-order spacial derivatives and the temporal derivatives of resp..
Lemma 3.2.
It holds that
(3.2) |
and
(3.3) |
Proof.
(1) Let . Applying and to (2.6)1 and (2.6)2 resp., we obtain
(3.4) |
Taking the inner product of (3.4)1 and in , and then using the integration by parts and the boundary conditions of in (2.5), (A.16) yields
(3.5) |
Exploiting the integration by parts, (1.1)1 and the boundary condition of in (2.5), we can derive that
(3.6) |
In view of the incompressible condition, the boundary condition of in (2.5), (A.16) and the equation (2.6)2, it is to see that
(3.7) |
Taking the inner product of (3.4)2 and in , and then making use of the incompressible condition, the integration by parts, the boundary conditions of in (2.5), (3.7), (A.16), and the identity (3.6), we deduce that
(3.8) |
Putting (3.5) and (3.8) together yields
(3.9) |
where we have used the Einstein convention of summation over repeated indices. Next we shall estimate for and in sequence.
Using the definition of and the imbedding inequality (A.1), we get
(3.10) |
Making use of Hölder inequality, the product estimates in (A.12) and (3.10), we can estimate that
(3.11) |
Now we turn to estimating for . Using the incompressible condition, the integration by parts and the boundary conditions of in (2.5), (3.7), (A.16), it holds that
and
We can use the above three identities and the integration by parts to rewrite as follows:
(3.12) |
where we have defined that
Take a similar procedure as (3.11), we can easily obtain
(3.13) |
Making use of the incompressible condition, the integration by parts, the product estimate, and the boundary condition of in (A.16), we deduce that
(3.14) |
Due to the absence of the dissipation of , we can not directly estimate for . To overcome this difficulty, we shall use equation twice as follows.
(3.15) |
where in the fourth equality we have used the identity
Next we shall estimate the three terms – in sequence.
It follows from the product estimates that
(3.16) |
Exploiting the incompressible condition, the integration by parts, the product estimates and the boundary condition of in (A.16), we have
(3.17) |
Similar to , we shall use the equation twice to rewrite as follows:
(3.18) |
By the incompressible condition, the integration by parts, the product estimate, and the boundary condition of in (A.16), we have
In addition, we can use the product estimate and the anisotropic interpolation inequality (A.6) to obtain
Similarly, we also get
Thus substituting the above three estimates into (3.18) yields
(3.19) |
Now inserting (3.16), (3.17) and (3.19) into (3.15) and then using Young’s inequality, we obtain
(3.20) |
Next we shall derive the dissipative estimates for and .
Lemma 3.3.
It holds that
(3.23) |
and
(3.24) |
Proof.
(1) Taking the inner product of (2.6)2 and in , and then using the integration by parts and the boundary condition (2.5), we get
(3.25) |
Using and the boundary condition of in (A.16), we find that
Putting the above identity into (3.25) yields
(3.26) |
Exploiting Hölder inequality, the incompressible condition, the integration by parts, the product estimate, the boundary condition of in (2.5), (3.10), Poincaré-type inequality (A.3), we obtain
Similarly, we also have
(2) Applying to the mass equation (2.6)1 yields
Taking the inner product of the above identity and in , we have
Taking the inner product of (3.22)2 and in , and then using the integration by parts, the boundary conditions of in (2.5), (3.7), we arrive at
Summing up the above two identities, we obtain
(3.27) |
where we have defined that
Making use of the product estimates, (3.10) and the relation , it is easy to have
(3.28) |
Exploiting the integration by parts, the product estimate, the boundary condition (2.5), and Poincaré-type inequality (A.3), we arrive at
(3.29) |
Consequently, inserting (3.28) and (3.1) into (3.1) yields (3.3). ∎
Next we shall establish the energy estimate of and the dissipation estimate of .
Lemma 3.4.
It holds that
(3.30) |
and
(3.31) |
Proof.
(1) We can rewrite as follows
(3.32) |
where . Taking the inner product of the above identity and in , and then using the integration by parts and the boundary conditions of in (A.16), we have
We easily deduce from the above identity that
which yields (3.30).
(2) Applying to the momentum equation (2.6)2, we can obtain the vortex equation
(3.33) |
where we have defined that
Exploiting the definition of and the product estimates, we have
(3.34) |
Finally we shall derive that and can be controlled by the norms of spatial derivatives of .
Lemma 3.5.
It holds that
(3.35) |
and
(3.36) |
Proof.
It is well-known from the incompressible condition and the mass equation (1.1)1 that
(3.37) |
where . Taking the inner product of (3.32) and in , and then using the integration by parts, and the boundary condition of in (2.5), we get
Exploiting Hölder inequality, the product estimate, (3.10) with and (3.37), we can get (3.36) from the above identity. This completes the proof. ∎
3.2 A priori stability estimates
Now we are in the position to building the total energy inequality (2.11) for the initial boundary value problem (2.5)–(2.7).
Proposition 3.1 (A priori estimates).
Proof.
Let
(3.40) |
In view of (3.35) and (3.36), it is easy to see that
(3.41) |
where will be defined by (3.49). In addition, we can derive from and (3.40) that
(3.42) |
Exploiting Young’s inequality, (3.35), (3.40), (A.20) with and (A.22) with , we derive from Lemmas 3.1–3.3 and (3.31) that, for sufficiently large positive constant (independent of ),
(3.43) |
where we have defined that
In addition, making use of Young’s inequality, the definitions of , , (3.10) with , (3.30), (3.37), (3.40), (3.42), Poincaré-type inequality (A.3), (A.20) with and (A.21), we have, for sufficiently large positive constant ,
(3.44) | |||
(3.45) | |||
(3.46) |
It is easy to see from (3.43) and (3.46) that there exists such that the following estimate holds for any :
(3.47) |
Integrating the above inequality over we obtain
which, together with (3.41), (3.44) and (3.45), implies
(3.48) |
If
(3.49) |
we can derive from (3.40) and (3.48) that
(3.50) |
which yields (3.39). This completes the proof. ∎
4 Proof of Theorem 2.1
Now we introduce the local(-in-time) well-posedness of the initial-boundary value problem (2.5)–(2.7) for any fixed .
Proposition 4.1.
Proof 1.
Thanks to the a priori stability estimate (3.39) in Proposition 3.1, we can easily establish the global solvability in Theorem 2.1 based on the local solvability in Proposition 4.1. Next, we briefly describe the proof for the readers’ convenience.
Assume that satisfies (2.8) and (2.10), where and are provided by Proposition 3.1. In view of Proposition 4.1, there exists a unique local strong solution to the initial-boundary value problem (2.5)–(2.7) with the maximal existence time .
By the regularity of , we can verify that satisfies stability estimate (3.39) in Proposition 3.1, i.e.
(4.1) |
if
Let
Then, we easily see that the definition of makes sense by the fact
(4.2) |
Thus, to show the existence of a global strong solution, it suffices to verify . We shall prove this by contradiction below.
Assume , then by and the definition of and Proposition 4.1, we have
(4.3) |
Noting that
then, by (4.1) with and (4.2), we have
In particular,
(4.4) |
Making use of (4.3), (4.4) and the strong continuity , we deduce that there is a constant , such that
which contradicts with the definition of . Hence, , which implies . This completes the proof of the existence of a global solution. The uniqueness of the global solution is obvious due to the uniqueness result of local solutions in Proposition 4.1. In addition, exploiting the product estimates and (2.11), we can easily derive (2.12) from (3.32).
To complete the proof of Theorem 2.1, we shall verify (2.13). To this purpose, applying to the momentum equation (3.32)1 yields
(4.5) |
Applying to the above identity, and then using the product estimate and (2.11), we can estimate that
(4.6) |
Noting that , we can derive from (2.11), (3.30), (4.6) and Lemma A.5 that
(4.7) |
Thanks to (2.11) and (4.7), we easily deduce from (3.32) that
which, together with (4.7), yields (2.13). This completes the proof of Theorem 2.1.
5 Proof of Theorem 2.2
This section is devoted to the proof of the asymptotic behavior of solutions stated in Theorem 2.2. For the simplicity, we denote by , where .
Let and be arbitrary given. Making use of (2.11)–(2.13), Aubin–Lions theorem (see Theorem A.3), Arzelá–Ascoli theorem (see Theorem A.4) and Banach–Alaoglu theorem (see Theorem A.5), there exits a sequence (not relabeled) such that, for ,
(5.1) | |||
(5.2) | |||
(5.3) | |||
(5.4) | |||
(5.5) | |||
(5.6) |
moreover, for any , for any , for any , for any satisfying and for any satisfying ,
(5.7) | |||
(5.8) | |||
(5.9) |
We can further derive from (5.7)–(5.9) that, for a.e. ,
(5.10) | |||
(5.11) | |||
(5.12) |
In view of (5.11), (5.12) and Lemma A.6, we have
(5.13) | |||
(5.14) |
which, together with (5.10), yields
(5.15) |
In addition, it is easy to see from (5.2) and (5.5) that
(5.16) |
which, together with the regularity of , imply
(5.17) |
By virtue of the product estimate and (2.11), it is easily to see that
Thus it holds that
(5.18) | |||
(5.19) |
Making use of (5.1), (5.2), (5.3), (5.17)–(5.19) and (A.2), one has
(5.20) | |||
(5.21) | |||
(5.22) |
Let . Making use of (5.1), (5.4), (5.6), (5.13), (5.14) and (5.20) –(5.22), we easily derive from (2.6) with in place of that
(5.23) |
and
(5.24) |
which, together with (5.15), yields
(5.25) |
In addition, applying to (5.23)1, and then using (5.23)2, we arrive at
(5.26) |
Now let , , and enjoy the same regularity as well as , , and , and satisfy the following initial-boundary value problem
(5.27) |
then it is easy to check that
The uniqueness mentioned above means that any sequence of
converges to the limit function is independent of choosing the sequences of solutions. This completes the proof of Theorem 2.2.
Appendix A Analysis tools
This appendix is devoted to providing some mathematical results, which have been used in previous sections. In addition, still denotes where the positive constant depends on the parameters and the domain in the lemmas in which appears.
Lemma A.1.
Embedding inequality [1, Theorem 4.12]: Let be a domain satisfying the cone condition. It holds that
(A.1) |
for any (after possibly being redefined on a set of measure zero), and
(A.2) |
Lemma A.2.
Poincaré-type inequality [23, Lemma A.10]: For any , it holds that
(A.3) |
Lemma A.3.
-
1.
Interpolation inequality in : Let be a domain satisfying the cone condition and , then it holds that
(A.4) for any and for any .
-
2.
Interpolation inequalities in : it holds that
(A.5) (A.6)
Proof 2.
Lemma A.4.
Product estimates (see [31, Lemma A.3]): Let be a domain satisfying the cone condition, and , be functions defined in . Then
(A.12) |
if the norms on the right hand side of the above inequalities are finite.
Lemma A.5.
A Hodge-type elliptic estimate [32, Lemma A.4]: Let , then
(A.13) |
Lemma A.6.
Helmholtz decomposition in -spaces (see Lemma 2.5.1 in Chapter II in [56]): Let , , be any domain. We define that
and
Then
and each has a unique decomposition
with , , and
Lemma A.7.
An elliptic estimate for the Dirichlet boundary value condition: Let , and be given, then there exists a unique solution solving the boundary value problem:
Moreover,
(A.14) |
Proof 3.
Please refer to [32, Lemma A.7] for the proof.
Lemma A.8.
An elliptic estimate for the Neumann boundary value condition: Let be a positive constant and , then there exists a unique solution solving the boundary value problem:
where denotes the outward unit normal vector to . Moreover,
(A.15) |
Lemma A.9.
Let be the solution of (2.6) with the initial condition , then satisfies the following boundary condition:
(A.16) |
where .
Proof.
The above result (A.16) can be found in [35, Lemma 2.1]. However we provide the proofs for reader’s convenience. In view of the mass equation (2.6)1 and the boundary condition of in (2.5), it holds that
Taking the inner product of the above identity and in , and then using the integration by parts and the embedding inequality of in (A.1), we derive that
Noting that and , thus applying Gronwall’s inequality to the above inequality yields
which implies
(A.17) |
Lemma A.10.
Under the assumption of Lemma A.9, we have the following estimates:
(A.20) | |||
(A.21) | |||
(A.22) |
Proof.
Similar results can be found in [35, Lemma 2.2], however we also provide the proofs of (A.20)–(A.22) for reader’s convenience.
Theorem A.3.
Aubin–Lions theorem [53, Theorem 1.71]: Let , be Banach spaces, a sequence bounded in and bounded in , where . Then is relatively compact in for any .
Theorem A.4.
Arzelá–Ascoli theorem [53, Theorem 1.70]: Let and , be Banach spaces such that is compact. Let be a sequence of functions uniformly bounded in and uniformly continuous in . Then there exists such that strongly in at least for a chosen subsequence.
Theorem A.5.
Banach–Alaoglu theorem (see Sections 1.4.5.25 and 1.4.5.26 in [53]):
-
1.
Let be a reflexive Banach space and let be a bounded sequence. Then there exists a subsequence weakly convergent in .
-
2.
Another version of the Banach–Alaoglu theorem: Let be a separable Banach space and let be a bounded sequence. Then there exists a subsequence weakly- convergent in .
Acknowledgements. The research of Fei Jiang was supported by NSFC (Grant Nos. 12371233 and 12231016), and the Natural Science Foundation of Fujian Province of China (Grant Nos. 2024J011011 and 2022J01105) and the Central Guidance on Local Science and Technology Development Fund of Fujian Province (Grant No. 2023L3003).
References
- Adams and Fourier [2003] R.A. Adams, J.J.F. Fourier, Sobolev Spaces, Academic press, New York, 2003.
- Antonelli and Spirito [2022] P. Antonelli, S. Spirito, Global existence of weak solutions to the Navier–Stokes–Korteweg equations, Ann. Inst. H. Poincaré C Anal. Non Linéaire 39 (2022) 171–200.
- Audiard and Haspot [2017] C. Audiard, B. Haspot, Global well-posedness of the Euler–Korteweg system for small irrotational data, Comm. Math. Phys. 351 (2017) 201–247.
- Benzoni-Gavage et al. [2007] S. Benzoni-Gavage, R. Danchin, S. Descombes, On the well-posedness for the Euler–Korteweg model in several space dimensions, Indiana Univ. Math. J. 56 (2007) 1499–1579.
- Bian et al. [2014] D. Bian, L. Yao, C. Zhu, Vanishing capillarity limit of the compressible fluid models of Korteweg type to the Navier–Stokes equations, SIAM J. Math. Anal. 46 (2014) 1633–1650.
- Bresch et al. [2008] D. Bresch, B. Desjardins, M. Gisclon, R. Sart, Instability results related to compressible Korteweg system, Ann. Univ. Ferrara Sez. 54 (2008) 11–36.
- Bresch et al. [2003] D. Bresch, B. Desjardins, C.K. Lin, On some compressible fluid models: Korteweg, lubrication, and shallow water systems, Comm. Partial Differential Equations 28 (2003) 843–868.
- Bresch et al. [2019] D. Bresch, M. Gisclon, I. Lacroix-Violet, On Navier–Stokes–Korteweg and Euler–Korteweg systems: application to quantum fluids models, Arch. Ration. Mech. Anal. 233 (2019) 975–1025.
- Bresch et al. [2022] D. Bresch, M. Gisclon, I. Lacroix-Violet, A. Vasseur, On the exponential decay for compressible Navier–Stokes–Korteweg equations with a drag term, J. Math. Fluid Mech. 24 (2022) Paper No. 11, 16.
- Browning and Kreiss [1982] G. Browning, H.O. Kreiss, Problems with different time scales for nonlinear partial differential equations, SIAM J. Appl. Math. 42 (1982) 704–718.
- Burtea and Charve [2017] C. Burtea, F. Charve, Lagrangian methods for a general inhomogeneous incompressible Navier–Stokes–Korteweg system with variable capillarity and viscosity coefficients, SIAM J. Math. Anal. 49 (2017) 3476–3495.
- Charve [2014] F. Charve, Local in time results for local and non-local capillary Navier–Stokes systems with large data, J. Differential Equations 256 (2014) 2152–2193.
- Charve et al. [2021] F. Charve, R. Danchin, J. Xu, Gevrey analyticity and decay for the compressible Navier–Stokes system with capillarity, Indiana Univ. Math. J. 70 (2021) 1903–1944.
- Charve and Haspot [2013] F. Charve, B. Haspot, Existence of a global strong solution and vanishing capillarity-viscosity limit in one dimension for the Korteweg system, SIAM J. Math. Anal. 45 (2013) 469–494.
- Chen and Li [2021] Z. Chen, Y. Li, Asymptotic behavior of solutions to an impermeable wall problem of the compressible fluid models of Korteweg type with density-dependent viscosity and capillarity, SIAM J. Math. Anal. 53 (2021) 1434–1473.
- Chen and Zhao [2014] Z. Chen, H. Zhao, Existence and nonlinear stability of stationary solutions to the full compressible Navier–Stokes–Korteweg system, J. Math. Pures Appl. 101 (2014) 330–371.
- Danchin and Desjardins [2001] R. Danchin, B. Desjardins, Existence of solutions for compressible fluid models of Korteweg type, Ann. Inst. H. Poincaré C Anal. Non Linéaire 18 (2001) 97–133.
- Ding et al. [2018] S. Ding, Q. Li, Z. Xin, Stability analysis for the incompressible Navier–Stokes equations with Navier boundary conditions, J. Math. Fluid Mech. 20 (2018) 603–629.
- Dunn and Serrin [1985] J.E. Dunn, J. Serrin, On the thermomechanics of interstitial working, Arch. Ration. Mech. Anal. 88 (1985) 95–133.
- Fanelli [2016] F. Fanelli, Highly rotating viscous compressible fluids in presence of capillarity effects, Math. Ann. 366 (2016) 981–1033.
- Germain and LeFloch [2016] P. Germain, P. LeFloch, Finite energy method for compressible fluids: the Navier–Stokes–Korteweg model, Comm. Pure Appl. Math. 69 (2016) 3–61.
- Goto [1990] S. Goto, Singular limit of the incompressible ideal magneto-fluid motion with respect to the Alfvén number, Hokkaido Math. J. 19 (1990) 175–187.
- Guo and Tice [2013] Y. Guo, I. Tice, Decay of viscous surface waves without surface tension in horizontally infinitedomains, Anal. PDE 6 (2013) 1429–1533.
- Haspot [2011] B. Haspot, Existence of global weak solution for compressible fluid models of Korteweg type, J. Math. Fluid Mech. 13 (2011) 223–249.
- Haspot [2020] B. Haspot, Strong solution for Korteweg system in with initial density in , Proc. Lond. Math. Soc. (3) 121 (2020) 1766–1797.
- Hattori and Li [1994] H. Hattori, D. Li, Solutions for two-dimensional system for materials of Korteweg type, SIAM J. Math. Anal. 25 (1994) 85–98.
- Hong [2020] H. Hong, Strong solutions for the compressible barotropic fluid model of Korteweg type in the bounded domain, Z. Angew. Math. Phys. 71 (2020) Paper No. 85, 25.
- Hong [2022] H. Hong, Stability of stationary solutions and viscous shock wave in the inflow problem for isentropic Navier–Stokes–Korteweg system, J. Differential Equations 314 (2022) 518–573.
- Hou et al. [2018] X. Hou, H. Peng, C. Zhu, Global well-posedness of the 3D non-isothermal compressible fluid model of Korteweg type, Nonlinear Anal. Real World Applications 43 (2018) 18–53.
- Hou et al. [2017] X. Hou, L. Yao, C. Zhu, Vanishing capillarity limit of the compressible non-isentropic Navier–Stokes–Korteweg system to Navier–Stokes system, J. Math. Anal. Appl. 448 (2017) 421–446.
- Jiang et al. [2023a] F. Jiang, H. Jiang, S. Jiang, Rayleigh–Taylor instability in stratified compressible fluids with/without the interfacial surface tension, ariXiv:2309.13370 (2023a).
- Jiang et al. [2022] F. Jiang, S. Jiang, Y. Zhao, On inhibition of the Rayleigh–Taylor instability by a horizontal magnetic field in ideal MHD fluids with velocity damping, J. Differential Equations 314 (2022) 574–652.
- Jiang et al. [2023b] F. Jiang, S. Jiang, Y. Zhao, On inhibition of the Rayleigh–Taylor instability by a horizontal magnetic field in 2D non-resistive MHD fluids: the viscous case, CSIAM Trans. Appl. Math. 4 (2023b) 451–514.
- Jiang et al. [2023c] F. Jiang, F. Li, Z. Zhang, On stability and instability of gravity driven Navier–Stokes–Korteweg model in two dimensions, arXiv preprint arXiv:2302.01013 (2023c).
- Jiang et al. [2024] F. Jiang, Y. Zhang, Z. Zhang, On the inhibition of Rayleigh–Taylor instability by capillarity in the Navier–Stokes–Korteweg model, arXiv preprint arXiv:2402.07201 (2024).
- Jiang et al. [2019] S. Jiang, Q. Ju, X. Xu, Small Alfvén number limit for incompressible magneto-hydrodynamics in a domain with boundaries, Sci. China Math. 62 (2019) 2229–2248.
- Ju and Xu [2022] Q. Ju, J. Xu, Zero-Mach limit of the compressible Navier–Stokes–Korteweg equations, J. Math. Phys. 63 (2022) 111503.
- Jüngel et al. [2014] A. Jüngel, C.K. Lin, K.C. Wu, An asymptotic limit of a Navier–Stokes system with capillary effects, Comm. Math. Phys. 329 (2014) 725–744.
- Kawashima et al. [2021] S. Kawashima, Y. Shibata, J. Xu, The energy methods and decay for the compressible Navier–Stokes equations with capillarity, J. Math. Pures Appl. (9) 154 (2021) 146–184.
- Kobayashi et al. [2022] T. Kobayashi, M. Murata, H. Saito, Resolvent estimates for a compressible fluid model of Korteweg type and their application, J. Math. Fluid Mech. 24 (2022) Paper No. 12, 42.
- Korteweg [1901] D.J. Korteweg, Sur la forme que prennent les équations du mouvements des fluides si l’on tient compte des forces capillaires causées par des variations de densité considérables mais connues et sur la théorie de la capillarité dans l’hypothèse d’une variation continue de la densité, Archives Néerlandaises des Sciences exactes et naturelles 6 (1901) 1–24.
- Kotschote [2008] M. Kotschote, Strong solutions for a compressible fluid model of Korteweg type, Ann. Inst. H. Poincaré C Anal. Non Linéaire 25 (2008) 679–696.
- Kotschote [2010] M. Kotschote, Strong well-posedness for a Korteweg-type model for the dynamics of a compressible non-isothermal fluid, J. Math. Fluid Mech. 12 (2010) 473–484.
- Kotschote [2012] M. Kotschote, Dynamics of compressible non-isothermal fluids of non-Newtonian Korteweg type, SIAM J. Math. Anal. 44 (2012) 74–101.
- Kotschote [2014] M. Kotschote, Existence and time-asymptotics of global strong solutions to dynamic Korteweg models, Indiana Univ. Math. J. 63 (2014) 21–51.
- Li and Zhang [2023] F. Li, Z. Zhang, Stabilizing effect of capillarity in the Rayleigh–Taylor problem to the viscous incompressible capillary fluids, SIAM J. Math. Anal. 55 (2023) 3287–3315.
- Li and Ding [2021] Q. Li, S. Ding, Global well-posedness of the Navier–Stokes equations with Navier-slip boundary conditions in a strip domain, Commun. Pure Appl. Anal. 20 (2021) 3561–3581.
- Li and Yong [2016] Y. Li, W. Yong, Zero Mach number limit of the compressible Navier–Stokes–Korteweg equations, Commun. Math. Sci. 14 (2016) 233–247.
- Lin [2012] F. Lin, Some analytical issues for elastic complex fluids, Comm. Pure Appl. Math. 65 (2012) 893–919.
- Murata and Shibata [2020] M. Murata, Y. Shibata, The global well-posedness for the compressible fluid model of Korteweg type, SIAM J. Math. Anal. 52 (2020) 6313–6337.
- Nguyen [2023] T.T. Nguyen, Influence of capillary number on nonlinear Rayleigh–Taylor instability to the Navier–Stokes–Korteweg equations, arXiv preprint arXiv:2312.05536 (2023).
- Nirenberg [1959] L. Nirenberg, On elliptic partial differential equations, Estratto Ann. Sc. Norm. Super. Pisa (3) XIII(II) (1959) 1–48.
- Novotný and Straškraba [2004] A. Novotný, I. Straškraba, Introduction to the Mathematical Theory of Compressible Flow, Oxford University Press, Oxford, 2004.
- Saito [2020] H. Saito, On the maximal – regularity for a compressible fluid model of Korteweg type on general domains, J. Differential Equations 268 (2020) 2802–2851.
- Sha and Li [2019] K. Sha, Y. Li, Low Mach number limit of the three-dimensional full compressible Navier–Stokes–Korteweg equations, Z. Angew. Math. Phys. 70 (2019) 169–175.
- Sohr [2001] H. Sohr, The Navier–Stokes equations: An elementary functional analytic approach, Birkhäuser Verlag, Basel, 2001.
- Sy et al. [2006] M. Sy, D. Bresch, F. Guillén-González, J. Lemoine, M.A. Rodríguez-Bellido, Local strong solution for the incompressible Korteweg model, C. R. Math. Acad. Sci. Paris 342 (2006) 169–174.
- Tan and Wang [2010] Z. Tan, Y. Wang, Strong solutions for the incompressible fluid models of Korteweg type, Acta Math. Sci. Ser. B (Engl. Ed.) 30 (2010) 799–809.
- Wang and Zhang [2024] P. Wang, Z. Zhang, Vanishing capillarity-viscosity limit of the incompressible Navier–Stokes–Korteweg equations with slip boundary condition, Nonlinear Anal. 243 (2024) 113526.
- Yang et al. [2015] J. Yang, L. Yao, C. Zhu, Vanishing capillarity-viscosity limit for the incompressible inhomogeneous fluid models of Korteweg type, Z. Angew. Math. Phys. 66 (2015) 2285–2303.
- Zhang et al. [2023] X. Zhang, F. Tian, W. Wang, On Rayleigh–Taylor instability in Navier–Stokes–Korteweg equations, J. Inequal. Appl. (2023) Paper No. 119, 30.
- Zhang [2022] Z. Zhang, Rayleigh–Taylor instability for viscous incompressible capillary fluids, J. Math. Fluid Mech. 24 (2022) Paper No. 70, 23.