On the initial-boundary value problem of two-phase incompressible flows with variable density in smooth bounded domain
Abstract.
In this work, we study the so-called Allen-Cahn-Navier-Stokes equations, a diffuse-interface model for two-phase incompressible flows with different densities. We first prove the local-in-time existence and uniqueness of classical solutions with finite initial energy over the smooth bounded domain . The key point is to transform the boundary values of the higher order spatial derivatives to that of the higher order time derivatives by employing the well-known Agmon-Douglis-Nireberg theory in [6]. We then prove global existence near the equilibrium and justify the time exponetial decay of the global solution. The majority is that the derivative of the physical relevant energy density will generate an additional damping effect under the perturbation .
Keywords. Incompressible Allen-Cahn-Navier-Stokes system; two-phase flow; global stability; time exponetial decay
AMS subject classifications. 35B45, 35B65, 35Q35, 76D03, 76T99
1. Introduction
1.1. Two-phase incompressible flows
In the past two decades, the phase field approach was employed by many researchers in various fluid models [2, 3, 9, 16, 17, 18, 20], and have carried out extensive analytical and numerical studies on the two-phase flows. For phase field model of incompressible binary fluids with identical densities, or with small density ratios where Boussinesq approximation can be applied in practice, we refer the readers to [1, 8, 11, 12, 13, 20, 21] and references therein for detailed derivations and mathematical analysis. However, in most cases, the density differences of the components are not negligible, whence studies on the incompressible two-phase fluids with non-matched densities become even more interesting and challenging.
Recently, in [14], by employing the energetic variational approach, a new two-phase incompressible flows with variable density was derived. In fact, before [14], analytical and numerical results for two-phase incompressible flow models that are valid for large density ratios between different species are quite limited [4, 17, 18]. Most studies of the phase-field models for binary fluids have been restricted to the cases with the same density or with small density differences. In the latter case, Boussinesq approximation can be used, where the variable density is replaced by a background constant density value and external gravitational force is added to model the effect of density force [16, 20]. In the model derived in [14], the density ratio between two phases can be quite large and hence the Boussinesq approximation is no more physically valid.
In this paper, we study the hydrodynamics of a diffuse-interface model describing a mixture of two immiscible incompressible fluids in the smooth bounded domain with different densities and . A phase field is introduced to characterize the two fluids such that
with a thin, smooth transition region. While the two fluids are mixed, will ranged be in . More precisely, we study the following Allen-Cahn-Navier-Stokes (briefly, ACNS) system:
| (1.1) |
Here, is the phase field function that labels different species, denotes the velocity of the fluid, and stands for the pressure. is the average density which is a given function of . is the viscosity, represents the competition between the kinetic energy and the free energy, and comes from microscopic internal damping during the mixing of two immiscible incompressible fluids. is the material derivative of with respect to the velocity . The physical relevant energy density functional that represent the two phases of the mixture usually has a double-well structure. Without loss of generality, in current paper, we assume that
| (1.2) |
with some small parameter . For the average density function , it is generally assumed that (see [14])
| (1.3) |
with being two positive constants and the following exterior convexity
| (1.4) |
In current paper, we assume that the is the parabolic average of and satisfying (1.3)-(1.4) as follows (see [17]):
| (1.5) |
Let be the outnormal vector of the boudnary . We now impose the boundary conditions:
| (1.6) |
and the initial data:
| (1.7) |
with the compatibility condition .
As shown in [14], once initially in , the Maxmal Principle of the heat equation implies that for and . We therefore know that the average density admits the lower and upper bounds, namely,
| (1.8) |
for all .
The system (1.1) was derived from employing the energetic variational approach by Jiang-Li-Liu [14], in which they considered the total energy
It consists of the first part of the macroscopic kinetic energy and the second part of the Helmholtz free energy. They also took the dissipation of the energy as
where the first part accounts for the macroscopic dissipation due to viscosity and the second part comes from microscopic internal damping during the mixing. Finally, the so-called energetic variational approach implies the system (1.1). The details can be seen in [14].
1.2. Notations and main results
In the sequel, we consider the smooth bounded domain in . We first denote by by the standard Lebesgue space with norm
For simplicity, we denote by . Let be the standard Sobolev space with norm . Here with and
As usual, simply denotes by with norm . In particular, . Moreover, we introduce a weighted space by
Remark that the corresponding vector-valued Lebesgue and Sobolev spaces will still be expressed by , , and , etc.
We also employ to denote by for some harmless constant . Moreover, means that for two harmless constants .
We then introduce the following energy functional and dissipation functional for ,
| (1.9) | ||||
Theorem 1.1 (Local well-posedness).
Let integer and be a smooth bounded domain. Assume that the initial data satisfy in and
Then there exists a , depending only on , , and the all coefficients, such that the ACNS system (1.1)-(1.2) with boudnary conditions (1.6) admits a unique solution satisfying
for . Moreover, the following energy inequality
| (1.10) |
and
| (1.11) |
hold for any and some constant , depending only on , , and all the coefficients.
Remark 1.1.
The next theorem is to prove the global well-posedness of (1.1)-(1.7) near the equilibrium . More precisely, let . We then know that satisfies
| (1.13) |
with the boundary conditions
| (1.14) |
and initial data
| (1.15) |
which satisfies the compatibility condition . Here and
| (1.16) |
We now introuce the global energy functionals and dissipations as follows: for ,
| (1.17) |
Theorem 1.2 (Global stability near ).
Let integer . Assume that and . Then there is a small positive constant , depending only on , and the all coefficients, such that if the initial energy
then (1.13)-(1.14) with initial data (1.15) admits a unique global in time solution satisfying
for . Furthermore, there hold
| (1.18) |
and
| (1.19) |
for all and for some positive constants , depending only on , and the all coefficients.
1.3. Main ideas and sketch of the proofs
The first goal of this paper is to prove the local well-posedness of the initial-boundary value problem (1.1)-(1.7) over the smooth bounded domain in the functions spaces that the spatial variables with regularity for large index , i.e., Theorem 1.1. In many known literatures that considered the well-posedness of various models, only the period domain or whole space was considered in the spaces with spatial regularity for large . Oppositely, if the bounded domains were focused, the aims were to prove the existence of weak solutions or strong solutions (in space, for example), in which cases only the information of boundary values was required, rather than that of higher order spartial derivatives of the boundary values.
In the smooth bounded domain , if one investigates the well-posedness of the ACNS model in the functions spaces with -regularity of the spatial variables, the main difficulties come from dealing with the boundary values of the higher order spatial derivatives. Generally speaking, although the boundary value of a function is finite, the higher order derivatives may be infinite. For example, the function for is continuous up to the boundary but . Note that it is impossible to control the boundary values of higher order spatial derivatives only employing the usual -theory over the period domain or whole space. For instance,
the boundary integral cannot be controlled by the “good” quantity combining with the Trace Theorem.
In order to overcome the difficulty, we employ the Agmon-Douglis-Nirenberg (briefly, ADN) theory associated with the general elliptic system in [6], which was reviewed in Section 2 below. The main ideas are as follows. The ACNS system (1.1) with boundary conditions (1.6) can be rewritten as the abstract forms
By the ADN theory, can be bounded by , namely, the second order spatial derivatives of can be transformed to the first order time derivatives of . We therefore mutate the higher order spatial derivatives problem to the higher order time derivatives problem.
The key point to dominate the higher order time derivatives is that the boundary condition (1.6), i.e., and can imply the boundary conditions (2.16) of the higher order time derivatives, i.e., and for any . The sketch of the proofs as follows.
- (1)
- (2)
- (3)
- (4)
The second goal of this paper is to investigate the global stability and long time decay of the ACNS system (1.1) near the equilibrium , hence, Theorem 1.2. In this case, under the purturbation , we reduce the equivalent system (1.13). The way to deal with the information of the boundary values is totally the same as that in proving the locall well-posedness to (1.1). The key ingredients to verify the global stability is to seek a new dissipation or damping structure on the unknown . Fortunately, by the perturbation , the derivative of the physical relevant energy density in the -equation of (1.1) will generate an additional damping effective , which gaurantees us to prove the global existence of the ACNS system near the equilibrium .
1.4. Organization of this paper
In the next section, we give some preliminaries, in particular review the general ADN theory. In Section 3, we derive three types of the a priori estimates of the system (1.1): 1) -estimates; 2) -estimates of the higher order time derivatives; 3) The estimates for the higher order time-space mixed derivatives. In Section 4, based on the a priori estimates, we prove the local well-posedness by employing the iteration methods. In Section 5, we prove the global classical solution for the system (1.13)-(1.7) near the equilibrium . Moreover, the time exponetial decay is also obtained. In Appendix A, we give the proof of Lemma 3.4.
2. Preliminaries
2.1. Agmon-Douglis-Nirenberg theory
In order to deal with the high order spatial derivatives of the solutions , we shall employ the well-known Agmon-Douglis-Nirenberg (briefly, ADN) theory [6]. For convenience of readers, we sketch the theory here. More precisely, they studied the general linear elliptic system on the bounded smooth domain with the following forms:
| (2.1) |
where the , linear differential operators, are polynomials in with coefficients depending on . The orders of these operators are be assumed to depend on two groups of integer weights, and , attached to the equations and to the unknowns, respectively, corresponding to the -th equation and to the -th dependent unknown . The manner of the dependence is represented by the inequality
| (2.2) |
where “” refers of course to the degree in , and . Moreover, if . The ellipticity of (2.1) is characterized by
| (2.3) |
where consists of the terms in which are just of the order . Furthermore, the following supplementary condition on should be imposed:
-
(SC)
is of even degree (with respect to ). For every pair of linearly independent real vectors , the polynomial in the complex variable has exactly roots () with positive imaginary part, i.e., .
Uniform ellipticity will be required in the sense that there is a positive constant such that
| (2.4) |
for every real vector and for every point in the closure of the domain , where .
In the boundary value of (2.1), is exactly . The linear boundary operator are of complex coefficients depending on . The orders of , like those of the operators , depend on two groups of integer weights, in this case the group already attached to the dependent unknowns and a new group of which pertains to the -th boundary condition, . The exact dependence is expressed by the inequality
| (2.5) |
and when . Moreover, the following complementing boundary condition on the boundary operator should also be imposed:
-
(CBC)
For any and any real, non-zero vector tangent to at , let us regard and the elements of the matrix
(2.6) as polynomial in the indeterminate . The rows of the latter matrix are required to be linearly independent modulo , i.e.,
only if the constant are all zero. Here is the normal to at , consists of the terms in which are just of the order , and denotes the matrix adjoint to .
Specifically, the operators and are of the forms
| (2.7) |
where and denote multi-indices indicative of the precise differentiation involved. Then the following results hold.
Proposition 2.1 (Theorem 10.5 of ADN [6]).
Let be an open bounded domain of class , and . Assume that , , and 111 and is equipped with the image norm where is the trace operator on .. A constant exists such that, if is finite for , then also is finite, and
| (2.8) |
is dependent on , , , , , , and .
Remark that if the solution to (2.1) is unique, the term on the right can be omitted.
Lemma 2.1 (Proposition 2.2, Chapter I of [19]).
Let be an open bounded set of class , , integer . Let us suppose that
| (2.9) |
are solutions of the generalized Stokes problem (2.10):
| (2.10) | ||||
If , and , then , , and there exists a constant such that
| (2.11) | ||||
where for , for .
Next, the usual -theory of elliptic equation can be reduced by the ADN theory in Proposition 2.1. One takes the following Laplace equation into consideration:
| (2.12) | ||||
By letting with , and , one gains a first order differential system
| (2.13) |
on with the boundary condition
| (2.14) |
where denotes the -th component of the normal vector to . When the weights
are assigned to , respectively, and
to the first, , the -th, and the -th equations, respectively, the characteristic determinant of (2.13) thereby is
whose degree is , i.e., . The system (2.13) is thus elliptic. Moreover, the equations (2.13) only need one boundary condition (2.14). It is obvious that the (SC) condition on holds, and the root of with positive imaginary is
Moreover, in (2.14), the weight is assigned to the only boundary condition. It is easy to verify that
Observe that for and . We have . A direct calculation shows that the matrix is
By and , the vector with elements is then equal to
which is zero modulo if and only if . Therefore, the complementing boundary condition (CBC) holds.
Therefore, the ADN theory in Proposition 2.1 directly concludes the results:
Lemma 2.2.
Let be an open bounded domain of class , and . Assume that , , and is a solution to the boundary value problem (2.12). Then, a constant exists such that and
| (2.15) |
2.2. Boundary conditions for ACNS system (1.1)
In this paper, the boundary values for ACNS system (1.1) is imposed on (1.6), i.e.,
The goal of this paper is to investigate the well-posedness of the ACNS equations (1.1) in the Sobolev space for large integer . To deal with the boundary values of the higher order derivatives is thereby one of the key points of this paper. However, the values of the higher order spatial derivatives restricted on the boundary is impossible to be controlled by the boundary values (1.6), i.e., . The idea is to convert the boundary values of the higher order spatial derivatives into that of the higher order time derivatives by the constitutive of the equations and ADN theory. For the boundary values of higher order time derivatives, the conditions (1.6), i.e., , show that, for ,
| (2.16) |
provided that they are all well-defined.
2.3. Some basic estimates
In this subsection, we first give some calculus inequalities, which will be frequently used later.
For the functions and satisfying the boundary conditions (2.16) with integers , one has
| (2.17) |
Then the Gagliardo-Nirenberg interpolation inequality gives the following results.
Lemma 2.3.
For any , and integer , there hold
| (2.18) | ||||
provided that the right-hand side of the quantities are all finite.
3. The A Priori Estimates
In this section, we devoted to the a priori estimate for the system (1.1)-(1.2). We divided this section into three parts. We first obtain the closed -estimates of (1.1). Then the higher order time derivative estimates are derived. At the end, the higher order spatial derivative estimates are established by employing the ADN theory.
3.1. Preparation
We first introduce the following energy functional and dissipation functional for ,
| (3.1) |
where the positive constant in will be determined later.
Moreover, the following bounds will be used later.
Lemma 3.1.
Proof.
We first dominate the quantity . Note that
By the expression of in (1.5) and Lemma 2.3, one knows that
Moreover, Lemma 2.3 indicates that
Collecting the above estimates and employing the Young’s inequality, we have
| (3.3) | |||
for any small and some . By the similar arguments in the previous estimates, one can calculate that
| (3.4) | ||||
Next we focus on the norms . By the similar arguments in (3.3), one can obtain
Recalling with , one knows
where Lemma 2.3 has been utilized. As similar as in (3.3), it infers that
Consequently, we gain
| (3.5) |
Furthermore, one observes that
By Lemma 2.3 and the Hölder inequality, it is implied that
for any small and some . Then by Lemma 2.3,
Moreover, by the similar arguments in (3.3), there hold
and
Therefore, we have
| (3.6) |
As a result, the second inequality in (3.2) is concluded by (3.5) and (3.6). Then the proof of Lemma 3.1 is completed. ∎
3.2. -estimates for ACNS equations (1.1)
We first derive the -estimates of the ACNS system (1.1), which will contain the major structures of the energy functionals. More precisely, the following time differential inequality holds.
Lemma 3.2.
Proof of Lemma 3.2.
One first takes -inner product of the first equation of (1.1) by dot with . There thereby holds
By the boundary conditions (1.6) and (2.16), one knows that
Moreover, implies . Thus, integrating by parts over reduces to
| (3.8) |
where
| (3.9) |
We then multiply the third equation of (1.1) by , which means that
Integrating by parts over and the boundary value in (1.6) imply that
Thus, one gains
| (3.10) |
where
| (3.11) |
Furthermore, from taking -inner product of the third equation of (1.1) by dot with , we deduce that
The conditions (2.16) indicate that . Then the integration by parts over tells us
Consequently, one has
| (3.12) |
Then, by the equalities (3.8), (3.10) and (3.12), there holds
| (3.13) | ||||
By the Hölder inequality, Lemma 3.1 and the bound with derived by (1.8), the first three terms in the right-hand side of (3.13) can be bounded by
| (3.14) | ||||
where the small is to be determined. Moreover, the last term in the right-hand side of (3.13) can be controlled by
| (3.15) | ||||
where Lemma 2.3 has been utilized. Then, plugging (3.14) and (3.15) into (3.13) reduces to
| (3.16) | ||||
for small to be determined.
Remark that the differential inequality (3.16) is still not closed, due to the uncontrolled quantity in the right-hand side of (3.16). In order to control this quantity, our way is to employ Lemma 2.1, a corollary of ADN theory in Proposition 2.1, and the -equation of (1.1).
From the -equation of (1.1) and (1.6), one has
| (3.17) | ||||
where
| (3.18) |
Then the inequality (2.11) in Lemma 2.1 indicates that
| (3.19) | ||||
where and the last inequality is derived from (3.18), (1.8) and the Hölder inequality. Recalling the definition of in (3.1), one derives from (3.19) and Lemma 2.3 that
| (3.20) | ||||
Next one focuses on the uncontrolled quantity . In order to deal with it, one will employ the -equation in (1.1). More precisely,
which implies that
| (3.21) |
where . By Lemma 2.3, there further hold
| (3.22) | ||||
Moreover, by (1.2) and , one easily knows
| (3.23) |
Furthermore, it infers from Lemma 2.3 that
| (3.24) | ||||
From substituting the bounds (3.22), (3.23) and (3.24) into (3.21), it thereby infers that
| (3.25) | ||||
Then, (3.20) and (3.25) imply that
| (3.26) |
3.3. -estimates for -derivatives of ACNS system with integers
Note that the functionals and in the a priori estimate in Lemma 3.2 only involve the second order spatial derivatives of and the third order spatial derivatives of . In order to investigate the information of the higher order spatial derivatives of , we initially dominate the higher order time derivatives of . Then we control the higher order spatial derivative by employing the ADN theory and the structures of the ACNS system. More precisely, the following differential inequality holds.
Lemma 3.3.
Proof.
For , by applying to (1.1) and combining with the boundary conditions (2.16), one gains
| (3.28) |
We take inner prouduct of the equation (3.28)1 with . Then there holds
By the boundary conditions in (3.28), one has
Moreover, implies . It thereby infers from integrating by parts over that
where is given in (3.9). Then, in the -equation of (3.28), and then take -inner product with . Together with the boundary conditions , integrating by parts over yields that
where is given in (3.11). Consequently, there holds
| (3.29) | ||||
By the similar arguments in (3.14), the term in the right-hand side of (3.29) can be bounded by
| (3.30) | |||
for any small . Moreover, the same arguments in (3.15) imply that
| (3.31) |
for small to be determined.
It remains to control the quantity in (3.29). By the Hölder inequality and the definition of in (3.1), one has
Recalling the definition of in (1.5) and Lemma 2.3, one has
It therefore infers that
| (3.32) | ||||
for and small . From substituting (3.30), (3.31) and (3.32) into (3.29), it is derived that
| (3.33) | ||||
Next we turn to control the quantity by using the ADN theory and the constitutive of the equations (3.28). From the -equation of (3.28), one has
| (3.34) | ||||
where
| (3.35) |
By the ADN theory in Lemma 2.1, one derives that
Employing the Sobolev embedding theory in Lemma 2.3 and the bounds (1.8) of , one implies that
Consequently, one has
| (3.36) | ||||
Here the functionals and are defined in (3.1).
3.4. Estimates for higher order spatial derivatives
As shown in Lemma 3.2 and Lemma 3.3, the energy and dissipative rates involve at most the third order spatial derivatives. In this subsection, by employing the ADN theory, we aim at investigating the information of the higher order spatial derivatives of the solutions to the ACNS system (1.1). For any fixed integer , the equations (1.1) with the boundary conditions (1.6) indicate that
| (3.41) | ||||
and
| (3.42) | ||||
where and are defined in (3.35) and (3.37), respectively. Namely,
| (3.43) | ||||
Then, the ADN theory given in Lemma 2.1-2.2 follows that for any integer ,
| (3.44) |
and
| (3.45) |
provided that the quantities and are both finite. Here .
Then, we will control the above two quantities in terms of the functionals . For notation simplicity, we denote by
| (3.46) |
Remark that, for ,
| (3.47) |
and for ,
| (3.48) |
Lemma 3.4.
For the integers and , the following estimates hold:
-
(1)
For ,
(3.49) -
(2)
For ,
(3.50) -
(3)
For ,
(3.51) Here is sufficiently small to determined.
The proof will be given in Appendix A later.
Corollary 3.1.
Let and . Then
| (3.52) |
where .
Proof.
First, by the definition in (3.46), the second inequality in (3.49) and (3.47) implies that
that is, the first bound in (3.52) holds.
We next control the quantity for as in (3.52) by induction arguments.
Case 1. .
Observe that the first inequality in (3.49) reduces to
Moreover, the second inequality in (3.50) indicates that
By taking small enough, one has
| (3.53) |
where . Namely, the second bound in (3.52) holds for the case .
Case 2. .
The bounds (3.47)-(3.48) and (3.50) indicate that
due to . Moreover, by (3.51), (3.47)-(3.48), the first inequality in (3.52) and (3.53), one has
where we have used . Consequently,
namely, the second inequality in (3.52) holds for . It remains to prove the cases .
Case 3. The Induction Hypotheses for .
Assume that
| (3.54) |
Case 4. Consider the case .
The bounds (3.51) and (3.47)-(3.48) indicate that for ,
| (3.55) |
and
One thereby has
Moreover, by (3.55), it can be easily implied that
As a consequence, there hold
Denote by
Then we have
| (3.56) |
As in (3.47)-(3.48), admits the same properties of and . The quantities and have the same property of in (3.47). Note that the Induction Hypotheses (3.54) indicate
Then, together with the induction relation (3.56),
where the last second inequality is derived from the facts and . Observe that satisfies . It then follows that
Then the Induction Principle concludes the second bound (3.52). The proof of Corollary 3.1 is completed. ∎
4. Local well-posedness of (1.1)
In this section, we prove the local well-posedness of the ACNS equations (1.1) with boundary values (1.6) and initial data (1.7). We first construct the linear approximate system by iteration scheme. The key step is to prove the existence of the uniform positive time lower bound to the approximate system and the uniform energy bounds based on the a priori estimates in Lemmas 3.2-3.3-3.5. Finally, by the compactness arguments, we can justify the local existence results of (1.1).
We first construct the approximate system by the iteration as follows: for all integer ,
| (4.1) |
The iteration starts from
| (4.2) |
By the standard linear theory, there hold:
Lemma 4.1.
Suppose that and the initial data satisfies and with . Then there is a maximal number such that the system (4.1) admits a unique solution satisfying and
for .
We remark that .
The next goal is to prove existence of the uniform lower bound of the time sequence of the approximate system (4.1)-(4.2). We introduce the approximate energy functional and the approximate dissipation functional as follows:
and
Lemma 4.2.
Proof.
From the similar arguments in Lemma 3.2 and Lemma 3.3, one easily knows that for all ,
| (4.4) |
and
| (4.5) | ||||
with . Note that for and some , where the quantity is defined in Theorem 1.1. We will take , , , such that
| (4.6) |
For , define
| (4.7) |
where we have used the convection . It is easy to see that
Proof of the Theorem1.1: Local well-posedness. By Lemma 4.2, we know that for any fixed given in Lemma 4.2, there is a such that for all integer and ,
| (4.9) |
Moreover, by the similar arguments in Lemma 3.5 and the above uniform bound (4.9), one has
uniformly in and . Then by compactness arguments and Arzela-Ascoli Theorem, we obtain that the system (1.1)-(1.2) with boudnary conditions (1.6) admits a solution satisfying
for . Moreover, there hold
and
| (4.10) |
uniformly in . Since , the maximal principle of the parabolic equation shows that . Hence the existence result in Theorem 1.1 holds.
Then we will prove the uniqueness of the solution to (1.1)-(1.2) with boundary conditions (1.6). Assume that and are the two solutions to (1.1)-(1.2)-(1.6). Denote by
Then subjects to the following system
| (4.11) |
By multiplying by the -equation and multiplying by the -equation in (4.11), one easily has
where
and
and
Together with the bound (4.10), the Hölder inequality implies that
Consequently, we have
for , where . Note that . The Grönwall inequality implies
which means that and . Then the first equation in (4.11) reduces to
Thus for any constant . Namely, the pressure is unique up to a constant. The proof of Theorem 1.1 is finished.
5. Global stability near
In this section, we will prove the global classical existence and time decay rate of the ACNS system near the constant equilibrium . More precisely, we prove the global solution to (1.13)-(1.14) with small initial data (1.15). Furthermore, the exponetial decay of the global solution is also gained. The key point is that the term in the -equation of (1.1) will generate an additional damping term under the perturbation . With this damping structure, one can derive the uniform global-in-times energy estimates of the fluctuated system (1.13)-(1.14) withe intial data (1.15). The process of deriving the global estimates is similar to the a priori estimates in Section 3. Thus we will introduce the energy functional and dissipation as follows: for ,
| (5.1) |
where
| (5.2) |
and the small constant in will be determined later.
Lemma 5.1.
Proof.
We will prove the lemma by two steps: 1. -estimates for (1.13)-(1.15); 2. -estimates for higher order time derivatives.
Step 1. -estimates. As similar in (3.8), we first Ttake -inner product of the first equation of (1.13) by . It thereby follows from integrating by parts over that
As same in (3.10) and (3.12), it infers from taking -inner product of the -equation in (1.13) by dot with and integrating by parts that
where are given in (5.2). One thereby has
We consider the term . By Lemma 2.3 and the Hölder inequality, one infers that
| (5.3) |
where and are defined in (5.1). Similarly, the term can be bounded by
| (5.4) |
Finally, the term can be bounded by
| (5.5) |
Namely,
| (5.6) |
Observe that the quantities and involved in do not occur in the dissipative structures of (5). We thus need to control them by the ADN theory in Lemma 2.1 and the constitive of the equations (1.13). More precisely, one has
where
Then the ADN theory in Lemma 2.1 and the similar arguments in (3.19) show that
for some constants .
Next we dominate the quantity . The -equation in (1.13) indicates that
which, by the similar arguments in (3.21)-(3.24), implies that
for some positive constants . Consequently, one has
| (5.7) |
We now take such that . Therefore, from adding (5) to the times of (5.7), it follows that
which concludes that
| (5.8) |
Step 2. Estimates for higher order time derivatives. For , we apply to and get
| (5.9) |
Moreover, subjects to the boundary conditions
| (5.10) |
Then, by employing the similar derivation of (3.29), i.e., combining with the boundary conditions (5.10) and taking -inner products via dot with and in the first and third equation of (5.9), respectively, it follows that
where is given in (5.2), and
By the standard Sobolev theory and the similar arguments in (3.30)-(3.32), it easily follows that
and
and
and
Consequently, one has
| (5.11) |
Note that the quantities and involved in do not occur in the dissipative structures of (5). It thereby is requrired to dominate then by the ADN theory in Lemma 2.1 and the constitive of the equations (5.9). To be more precise, one has
where
Then the ADN theory in Lemma 2.1 and the similar arguments in (3.36) indicate that
for some positive constants .
Proof of Theorem 1.2: global well-posedness with small initial data.
By Lemma 5.1, we know that
| (5.14) |
Observe that by the constitive of (1.13),
for small to be determined. Then there is a sufficiently small such that if , then
| (5.15) |
Now we define
| (5.16) |
By the continuity of and (5.15), one has .
Further claim that . Indeed, if , then the energy inequality (5.14) implies that for all
| (5.17) |
which means
It therefore follows that
By the continuity of , there is a such that for all
which contradict to the definition of in (5.16). Thus . Consequently, we have
Moreover, since and , the Poincaré inequality indicates that for all . It thereby follows that
Together with (5.17), it infers that , which means that
| (5.18) |
Appendix A Proof of Lemma 3.4
The goal of this section is to justify the computations of bounds on and , namely, to prove Lemma 3.4. Later, we will frequently use the following calculus inequalities:
| (A.1) |
We now start to prove the results in Lemma 3.4.
Proof of Lemma 3.4.
We will divide the proof into three steps.
Step 1. : To control and for .
We first estimate the quantity . By the definition of in (3.43), it suffices to estimate the norms , and .
By the second inequality in (A.1) and the Hölder inequality, it follows that
where is defined in (3.46).
Next, by Lemma 2.3 and the last two inequalities in (A.1),
| (A.2) | ||||
for small to be determined, where the last second inequality is derived from (3.44). Here is defined in (3.46).
Collecting the above three bounds, we conclude the first inequality in (3.49) about the quantity .
We then estimate the quantity . Note that, by the definition of in (3.43),
By the similar arguments in (A.2), one can easily derive that
Recall that . Then the inequalities in (A.1) indicate that
Obviously, . We summarily obtain
hence, the second inequality in (3.49) holds.
Step 2. : To control and for .
We first estimate the quantity . Note that, by (3.43),
By the last two inequality in (A.1),
Moreover, by the last two inequality in (A.1) and the estimate (3.44),
Similarly, one has
where the last second inequality is implies by the bound (3.45). Consequently, we obtain
Together with the first bound in (3.49), it follows the validity of the first inequality about in (3.50).
We then estimate the quantity . By (3.38) and (3.44)-(3.45) we know that
Together with the second inequality in (3.49), we conclude the bound of in (3.50).
Step 3. : To control and for .
We first dominate the quantity . By (3.43),
Note that, by (1.5) and the first inequality in (A.1),
| (A.3) |
Then, it follows from (A.1) and (A.3) that
Similarly, one has
and
As a result, there holds
| (A.4) |
From the estimates (3.44) and (3.45), it is deduced that
| (A.5) | ||||
Then, by (A.4)-(A.5) and (3.46), we know that the bound of in (3.51) holds.
To end the proof, we estimate the quantity for . Recalling the definition of in (3.43), we know that
By (3.45),
Moreover, it follows from (A.1) and (3.44)-(3.45) that
Similarly, it infers that
Note that . Then (A.1) and (3.45) imply that
Consequently, we have
which, combining with (3.46), concludes the bound of in (3.51). Then the proof of Lemma 3.4 is finished. ∎
Acknowledgments
The first author N. J. was supported by grants from the National Natural Science Foundation of China under contract No. 11471181 and No. 11731008. The second author Y.-L. L. was supported by grants from the National Natural Science Foundation of China under contract No. 12201220, the Guang Dong Basic and Applied Basic Research Foundation under contract No. 2021A1515110210, and the Science and Technology Program of Guangzhou, China under the contract No. 202201010497.
References
- [1] H. Abels, On a diffuse interface model for two-phase flows of viscous, incompressible fluids with matched densities. Arch. Ration. Mech. Anal., 194 (2009), no. 2, 463-506.
- [2] H. Abels, Existence of weak solutions for a diffuse interface model for viscous, incompressible fluids with general densities. Comm. Math. Phys., 289 (2009), no. 1, 45-73.
- [3] H. Abels, H. Garcke and G. Grun, Thermodynamically consistent, frame indifferent diffuse interface models for incompressible two-phase flows with different densities. Math. Models Methods Appl. Sci., 22 (2012), no. 3, 1150013, 40 pp.
- [4] H.Abels, D. Depner and H. Garcke, On an incompressible Navier-Stokes/Cahn-Hilliard system with degenerate mobility. Ann. Inst. H. Poincare Anal. Non Lineaire, 30 (2013), no. 6, 1175-1190.
- [5] S. Agmon, A. Douglis, and L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions, I, Comm. Pure Appl. Math., Vol. 12, 1959, pp. 623–727.
- [6] S. Agmon, A. Douglis, and L. Nirenberg, Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions, II, Comm. Pure Appl. Math., Vol. 17, 1964, pp. 35–92.
- [7] S. N. Antontsev, A. V. Kazhikhov and V. N. Monakhov, Boundary Value Problems in Mechanics of Nonhomogeneous Fluids, Elsevier, 1989.
- [8] F. Boyer, Mathematical study of multi-phase flow under shear through order parameter formulation. Asymptot. Anal., 20 (1999), no. 2, 175-212.
- [9] S. Ding, Y. Li and W. Luo, Global solutions for a coupled compressible Navier-Stokes/Allen-Cahn system in 1D. J. Math. Fluid Mech., 15 (2013), 335-360.
- [10] D. A. Edwards, H. Brenner and D. T. Wasan, Interfacial Transport Process and Rheology. Butter-worths/Heinemann, London., (1991).
- [11] C. G. Gal and M. Grasselli, Asymptotic behavior of a Cahn-Hilliard-Navier-Stokes system in 2D. Ann.Inst. H. Poincare Anal. Non Lineaire., 27 (2010), no. 1, 401-436.
- [12] C. G. Gal and M. Grasselli, Longtime behavior for a model of homogeneous incompressible two-phase flows. Discrete Contin. Dyn. Syst., 28 (2010), no. 1, 1-39.
- [13] M. Gurtin, D. Polignone and J. Vinals, Two-phase binary fluids and immiscible fluids described by an order parameter. Math. Models Methods Appl. Sci., 6 (1996), no. 6, 815-831.
- [14] J. Jiang, Y. Li and C. Liu. Two-phase incompressible flows with variable density: an energetic variational approach. Discrete Contin. Dyn. Syst., 37 (2017), no. 6, 3243-3284.
- [15] J. U. Kim, Weak solutions of an initial boundary value problem for an incompressible viscous fluid with nonnegative density. SIAM J. Math. Anal., 18 (1987), 89-96.
- [16] C. Liu and J. Shen, A phase field model for the mixture of two incompressible fluids and its approximation by a Fourier-spectral method. Physica D., 179 (2003), 211-228.
- [17] C. Liu, J. Shen and X. F. Yang, Decoupled energy stable schemes for a phase-field model of two-phase incompressible flows with variable density. J. Sci, Comput., 62 (2015), 601-622.
- [18] J. Shen and X. F. Yang, A phase-field model and its numerical approximation for two-phase incompressible fows with different densities and viscosities. SIAM J. Sci. Comput., 32 (2010), 1159-1179.
- [19] R. Temam, Navier-Stokes equations. Theory and numerical analysis. Studies in Mathematics and its Applications, Vol. 2. North-Holland Publishing Co., Amsterdam-New York-Oxford, 1977.
- [20] H. Wu and X. Xu, Analysis of a diffuse-interface model for the binary viscous incompressible fluids with thermo-induced Marangoni effects. Commun. Math. Sci., 11 (2013), 603-633.
- [21] X. Xu, L. Zhao and C. Liu, Axisymmetric solutions to coupled Navier-Stokes/Allen-Cahn equations. SIAM J. Math. Anal., 41 (2010), 2246-2282.