• Keine Ergebnisse gefunden

Weak solutions for a diffuse interface model for two-phase flows of incompressible fluids with different densities and nonlocal free energies

N/A
N/A
Protected

Academic year: 2021

Aktie "Weak solutions for a diffuse interface model for two-phase flows of incompressible fluids with different densities and nonlocal free energies"

Copied!
20
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

DOI: 10.1002/mma.6111

R E S E A R C H A R T I C L E

Weak solutions for a diffuse interface model for two-phase flows of incompressible fluids with different densities and nonlocal free energies

Helmut Abels

1

Yutaka Terasawa

2

1

Fakultät für Mathematik, Universität Regensburg, Regensburg, Germany

2

Graduate School of Mathematics, Nagoya University, Nagoya, Japan

Correspondence

Helmut Abels, Fakultät für Mathematik, Universität Regensburg, 93040 Regensburg, Germany.

Email: helmut.abels@mathematik.

uni-regensburg.de

Communicated by: M. Groves

Funding information

Japan Society for the Promotion of Science, Grant/Award Number: 17K17804 and 26287022; Universitätsstiftung Hans Vielberth

We consider a diffuse interface model for the flow of two viscous incompressible Newtonian fluids with different densities in a bounded domain in two and three space dimensions and prove existence of weak solutions for it. In contrast to earlier contributions, we study a model with a singular nonlocal free energy, which controls the H

𝛼/2

-norm of the volume fraction. We show existence of weak solutions for large times with the aid of an implicit time discretization.

K E Y WO R D S

Cahn-Hilliard equation, diffuse interface model, mixtures of viscous fluids, Navier-Stokes equation, nonlocal operators, two-phase flow

M S C C L A S S I F I C AT I O N

35Q30; 35Q35; 76D03; 76D05; 76D27; 76D45

1 I N T RO D U CT I O N

In this contribution, we consider a two-phase flow for incompressible fluids of different densities and different viscosities.

The two fluids are assumed to be macroscopically immiscible and to be miscible in a thin interface region; ie, we consider a diffuse interface model (also called phase field model) for the two-phase flow. In contrast to sharp interface models, where the interface between the two fluids is a sufficiently smooth hypersurface, diffuse interface model can describe topological changes due to pinch off and droplet collision.

There are several diffuse interface models for such two-phase flows. Firstly, in the case of matched densities, ie, the densities of both fluids are assumed to be identical, there is a well-known model H, cf Hohenberg and Halperin or Gurtin et al.

1,2

In the case that the fluid densities do not coincide, there are different models. On one hand, Lowengrub and Truskinovsky

3

derived a quasi-incompressible model, where the mean velocity field of the mixture is in general not diver- gence free. On the other hand, Ding et al

4

proposed a model with a divergence free mean fluid velocities. But this model is not known to be thermodynamically consistent. In Abels et al,

5

a thermodynamically consistent diffuse interface model for two-phase flow with different densities and a divergence free mean velocity field was derived, which we call AGG model for short. The existence of weak solutions of the AGG model was shown in Abels et al.

6

For analytic result in the case of matched densities, ie, the model H, we refer to Abels

7

and Giorgini et al

8

and the reference given there. Existence of weak and strong solutions for a slight modification of the model by Lowengrub and Truskinovsky was proven in Abels.

9,10

Concerning the Cahn-Hilliard equation, Giacomin and Lebowitz

11,12

observed that a physically more rigorous deriva- tion leads to a nonlocal equation, which we call a nonlocal Cahn-Hilliard equation. There are two types of nonlocal

This is an open access article under the terms of the Creative Commons Attribution License, which permits use, distribution and reproduction in any medium, provided the original work is properly cited.

© 2020 The Authors. Mathematical Methods in the Applied Sciences published by John Wiley & Sons Ltd.

Math Meth Appl Sci. 2020;43:3200–3219.

wileyonlinelibrary.com/journal/mma

© 2020 John Wiley & Sons, Ltd.

3200

(2)

Cahn-Hilliard equations. One is the equation where the second order differential operator in the equation for the chem- ical potential is replaced by a convolution operator with a sufficiently smooth even function. We call it a nonlocal Cahn-Hilliard equation with a regular kernel in the following. The other is one where the second order differential operator is replaced by a regional fractional Laplacian. We call it a nonlocal Cahn-Hilliard equation with a singular ker- nel, since the regional fractional Laplacian is defined by using singular kernel. The nonlocal Cahn-Hilliard equation with a regular kernel was analyzed in previous works.

12-16

On the other hand, the nonlocal Cahn-Hilliard equation with a sin- gular kernel was first analyzed in Abels et al,

17

where they proved the existence and uniqueness of a weak solution of the nonlocal Cahn-Hilliard equation, its regularity properties, and the existence of a (connected) global attractor.

Concerning the nonlocal model H with a regular kernel, where the convective Cahn-Hilliard equation is replaced by the convective nonlocal Cahn-Hilliard equation with a regular kernel, first studies were done in references

18-20

; see also Frigeri

21

and the references there for more recent results. More recently, the nonlocal AGG model with a regular kernel, where the convective Cahn-Hilliard equation is replaced by the convective nonlocal Cahn-Hilliard equation with a regular kernel, was studied by Frigeri,

22

and he showed the existence of a weak solution for that model. The method of the proof in Frigeri

22

is based on the Faedo-Galerkin method of a suitably mollified system and the method of passing to the limit with two parameters tending to zero. The method is different from Abels,

6

which is based on implicit time discretization and a Leray-Schauder fixed point argument.

In this contribution, we consider a nonlocal AGG model with a singular kernel, where a convective Cahn-Hilliard equation in the AGG model is replaced by a convective nonlocal Cahn-Hilliard equation with a singular kernel. Our aim is to prove the existence of a weak solution of such a system.

In this contribution, we consider existence of weak solutions of the following system, which couples a nonhomogeneous Navier-Stokes equation system with a nonlocal Cahn-Hilliard equation:

𝜕

t

(𝜌v) + div(v (𝜌v + ̃ J)) − div(2𝜂(𝜑)Dv) + ∇p = 𝜇∇𝜑 in Q, (1)

div v = 0 in Q , (2)

𝜕

t

𝜑 + v · ∇𝜑 = div (m(𝜑)∇𝜇) in Q, (3)

𝜇 = Ψ

( 𝜑 ) + 𝜑 in Q , (4)

where 𝜌 = 𝜌(𝜑) ∶=

̃𝜌1+2̃𝜌2

+

̃𝜌2̃𝜌1

2

𝜑, ̃ J = −

̃𝜌2̃𝜌1

2

m(𝜑)∇𝜇, Q = Ω × (0, ∞). We assume that Ω R

d

, d = 2, 3, is a bounded domain with C

2

-boundary. Here and in the following v, p, and 𝜌 are the (mean) velocity, the pressure, and the density of the mixture of the two fluids, respectively. Furthermore, ̃𝜌

𝑗

, j = 1, 2, are the specific densities of the unmixed fluids, 𝜑 is the difference of the volume fractions of the two fluids, and 𝜇 is the chemical potential related to 𝜑 . Moreover, Dv =

1

2

(∇v + ∇v

T

), 𝜂(𝜑) > 0 is the viscosity of the fluid mixture, and m(𝜑) > 0 is a mobility coefficient. The term ̃ J describes the mass flux; ie, we have

𝜕

t

𝜌 = −div ̃ J.

It is important to have the term with ̃ J in (1) in order to obtain a thermodynamically consistent model, cf Abels et al

5

for the case with a local free energy.

Finally,  is defined as

u(x) = p.v.

Ω

(u(x) − u(𝑦))k(x, 𝑦, x𝑦)d𝑦

= lim

𝜀→0

Ω⧵B𝜀(x)

(u(x) − u(𝑦))k(x , 𝑦, x𝑦)d 𝑦 for x ∈ Ω (5) for suitable u ∶ Ω → R . Here, the kernel k ∶ R

d

× R

d

× ( R

d

⧵ {0}) → R is assumed to be (d + 2)-times continuously differentiable and to satisfy the conditions

k(x, 𝑦, z) = k(𝑦, x, −z), (6)

|𝜕

𝛽x

𝜕

𝑦𝛾

𝜕

𝛿z

k(x , 𝑦, z) | ⩽ C

𝛽,𝛾,𝛿

| z |

−d−𝛼−|𝛿|

, (7)

c

0

|z|

−d−𝛼

k(x, 𝑦, z)C

0

|z|

−d−𝛼

. (8)

for all x, 𝑦, z ∈ R

d

, z ≠ 0 and 𝛽, 𝛾, 𝛿 ∈ N

d0

with |𝛽 | + |𝛾| + |𝛿| ⩽ d + 2 and some constants C

𝛽,𝛾,𝛿

, c

0

, C

0

> 0. Here, 𝛼 is

the order of the operator, cf Abels and Kassmann.

23

We restrict ourselves to the case 𝛼 ∈ (1 , 2). If 𝜔C

d+2b

( R

d

), then

k(x, y, z) = 𝜔(x, y)|z|

−d−𝛼

is an example of a kernel satisfying the previous assumptions.

(3)

We add to our system the boundary and initial conditions

v |

𝜕Ω

= 0 on 𝜕Ω × (0 , ∞), (9)

𝜕

n

𝜇|

𝜕Ω

= 0 on 𝜕Ω × (0 , ∞), (10)

(v, 𝜑) |

t=0

= (v

0

, 𝜑

0

) in Ω. (11)

Here, 𝜕

n

= n · ∇ and n denotes the exterior normal at 𝜕Ω. We note that (9) is the usual no-slip boundary condition for the velocity field and 𝜕

n

𝜇|

𝜕Ω

= 0 describes that there is no mass flux of the fluid components through the boundary.

Furthermore, we complete the system above by an additional boundary condition for 𝜑, which will be part of the weak formulation, cf Definition 1. If 𝜑 is smooth enough (eg, 𝜑 (t) ∈ C

1,𝛽

(Ω) for every t ≥ 0) and k fulfills suitable assumptions, then

n

x0

· ∇𝜑(x

0

) = 0 for all x

0

𝜕Ω, (12)

where n

x0

depends on the interaction kernel k, cf Abels et al,

17, theorem 6.1

and x

0

𝜕Ω.

The total energy of the system at time t ≥ 0 is given by

E

tot

(𝜑, v) = E

kin

(𝜑, v) + E

free

(𝜑), (13) where

E

kin

( 𝜑, v) =

Ω

𝜌 | v |

2

2 dx , E

free

( 𝜑 ) =

Ω

Ψ( 𝜑 ) dx + 1 2  ( 𝜑, 𝜑 ) are the kinetic energy and the free energy of the mixture, respectively, and

(u, v) =

Ω

Ω

(u(x) − u(𝑦))(v(x) − v(𝑦))k(x, 𝑦, x𝑦) dx d𝑦 (14) for all u, vH

𝛼2

(Ω) is the natural bilinear form associated to , which will also be used to formulate the natural boundary condition for 𝜑 weakly. Every sufficiently smooth solution of the system above satisfies the energy identity

d

dt E

tot

(𝜑, v) = −

Ω

2𝜂(𝜑)|Dv|

2

dx

Ω

m(𝜑)|∇𝜇|

2

dx

for all t ≥ 0. This can be shown by testing (1) with v, (3) with 𝜇, and (4) with 𝜕

t

𝜑, where the product of 𝜑 and 𝜕

t

𝜑 coincides with

(𝜑(t), 𝜕

t

𝜑(t)) under the same natural boundary condition for 𝜑(t) as before, cf (12).

We consider a class of singular free energies, which will be specified below and which includes the homogeneous free energy of the so-called regular solution models used by Cahn and Hilliard

24

:

Ψ(𝜑) = 𝜗

2 ((1 + 𝜑) ln(1 + 𝜑) + (1𝜑) ln(1 − 𝜑)) − 𝜗

c

2 𝜑

2

, 𝜑 ∈ [−1 , 1], (15) where 0 < 𝜗 < 𝜗

c

. This choice of the free energies ensures that 𝜑(x, t) ∈ [−1, 1] almost everywhere. In order to deal with these terms, we apply techniques, which were developed in Abels and Wilke

25

and extended to the present nonlocal Cahn-Hilliard equation in Abels et al.

17

Our proof of existence of a weak solution of (1) to (4) together with a suitable initial and boundary condition follows

closely the proof of the main result of Abels et al.

6

The following are the main differences and difficulties of our paper

compared with Abels et al.

6

Since we do not expect H

1

-regularity in space for the volume fraction 𝜑 of a weak solution of

our system, we should eliminate ∇𝜑 from our weak formulation taking into account the incompressibility of v. Implicit

time discretization has to be constructed carefully, using a suitable mollification of 𝜑 and an addition of a small Laplacian

term to the chemical potential equation taking into account of the lack of H

1

-regularity in space of 𝜑. While the arguments

for the weak convergence of temporal interpolants of weak solutions of the time-discrete problem are similar to Abels

et al,

6

the function space used for the order parameter has less regularity in space since the nonlocal operator of order less

(4)

than 2 is involved in the equation for the chemical potential. For the convergence of the singular term Ψ

(𝜑), we employ the argument in Abels et al.

17

The only difference is that we work in space-time domains directly. For the validity of the energy inequality, additional arguments using the equation of chemical potential and the fact that weak convergence together with norm convergence in uniformly convex Banach spaces imply strong convergence are needed.

The structure of the contribution is as follows: In Section 2, we present some preliminaries, fix notations, and collect the needed results on nonlocal operator. In Section 3, we define weak solutions of our system and state our main result concerning the existence of weak solutions. In Section 4, we define an implicit time discretization of our system and show the existence of weak solutions of an associated time-discrete problem using the Leray-Schauder theorem. In Section 5, we obtain compactness in time of temporal interpolants of the weak solutions of time-discrete problem and obtain weak solutions of our system as weak limits of a suitable subsequence.

2 P R E L I M I NA R I E S

As usual, a b = (a

i

b

𝑗

)

di,𝑗=1

for a , b ∈ R

d

and A

s𝑦m

=

1

2

(A + A

T

) for A ∈ R

d×d

. Moreover,

⟨𝑓, g⟩ ≡ ⟨𝑓, g⟩

X,X

= 𝑓 (g), 𝑓X

, gX

denotes the duality product, where X is a Banach space and X

is its dual. We write X →→ Y if X is compactly embedded into Y. For a Hilbert space H, its inner product is denoted by, ·)

H

.

Let M R

d

be measurable. As usual L

q

(M), 1 ≤ q ≤ ∞, denotes the Lebesgue space, ||.||

q

its norm and (. , .)

M

= ( . , . )

L2(M)

its inner product if q = 2. Furthermore, L

q

(M; X) denotes the set of all fMX that are strongly measurable and q-integrable functions/essentially bounded functions. Here, X is a Banach space. If M = (a, b), we denote these spaces for simplicity by L

q

(a , b; X) and L

q

(a , b). Recall that f ∶ [0 , ∞) → X belongs L

qloc

([0 , ∞); X) if and only if fL

q

(0 , T; X) for every T > 0. Furthermore, L

quloc

([0, ∞); X) is the uniformly local variant of L

q

(0, ∞; X) consisting of all strongly measurable f ∶ [0 , ∞) → X such that

||𝑓 ||

Lquloc([0,∞);X)

= sup

t≥0

||𝑓 ||

Lq(t,t+1;X)

< ∞.

If T < ∞, we define L

quloc

([0, T); X) ∶= L

q

(0, T; X).

For a domain Ω R

d

, m ∈ N

0

, 1 ≤ q ≤ ∞, the standard Sobolev space is denoted by W

qm

(Ω). W

q,0m

(Ω) is the closure of C

0

(Ω) in W

qm

(Ω), W

q−m

(Ω) = (W

qm,0

(Ω))

, and W

q−m,0

(Ω) = (W

qm

(Ω))

. H

s

(Ω) denotes the L

2

-Bessel potential of order s ≥ 0.

Let 𝑓

Ω

=

1

|Ω|

Ω

𝑓 (x) dx denote the mean value of fL

1

(Ω). For m ∈ R , we define L

q(m)

(Ω) ∶= { 𝑓L

q

(Ω) ∶ 𝑓

Ω

= m} , 1 ≤ q ≤ ∞ . Then the orthogonal projection onto L

2(0)

(Ω) is given by

P

0

𝑓 ∶= 𝑓𝑓

Ω

= 𝑓 − 1

|Ω|∫

Ω

𝑓 (x) dx for all 𝑓L

2

(Ω) . For the following, we denote

H

1(0)

H

(0)1

(Ω) = H

1

(Ω) ∩ L

2(0)

(Ω) , (c , d)

H1

(0)(Ω)

∶= (∇c , ∇d)

L2(Ω)

. Because of Poincaré's inequality, H

(0)1

(Ω) is a Hilbert space. More generally, we define for s ≥ 0

H

(0)s

H

(0)s

(Ω) = H

s

(Ω) ∩ L

2(0)

(Ω), H

(0)s

(Ω) = (H

(0)s

(Ω))

, H

0s

(Ω) = (H

s

(Ω))

, H

s

(Ω) = (H

0s

(Ω))

.

Finally, 𝑓H

locs

(Ω) if and only if f|

Ω

H

s

) for every open and bounded subset Ω

with Ω

Ω.

(5)

We denote by L

2𝜎

(Ω) is the closure of C

0,𝜎

(Ω) in L

2

(Ω)

d

, where C

0,𝜎

(Ω) is the set of all divergence free vector fields in C

0

(Ω)

d

. The corresponding Helmholtz projection, ie, the L

2

-orthogonal projection onto L

2𝜎

(Ω), is denoted by P

𝜎

, cf, eg, Sohr.

26

Let I = [0, T] with 0 < T < ∞ or I = [0, ∞) if T = ∞ and let X is a Banach space. The Banach space of all bounded and continuous fIX is denoted by BC(I; X). It is equipped with the supremum norm. Moreover, BUC(I; X) is defined as the subspace of all bounded and uniformly continuous functions. Furthermore, BC

w

(I; X) is the set of all bounded and weakly continuous fIX. C

0

(0, T; X) denotes the vector space of all smooth functions f ∶ (0, T)X with suppf ⊂⊂ (0, T).

By definition 𝑓W

p1

(0 , T; X), 1p < ∞, if and only if 𝑓,

ddt𝑓

L

p

(0 , T; X). Furthermore, W

p1,uloc

([0 , ∞); X) is defined by replacing L

p

(0, T; X) by L

puloc

([0, ∞); X), and we set H

1

(0, T; X) = W

21

(0, T; X) and H

uloc1

([0, ∞); X) ∶= W

2,uloc1

([0, ∞); X).

Finally, we note the following:

Lemma 1. Let X, Y be two Banach spaces such that Y → X and X

→ Y

densely. Then L

(I; Y)∩BUC(I; X) → BC

w

(I; Y).

For a proof, see, eg, Abels.

9

2.1 Properties of the nonlocal elliptic operator

In the following, let  be defined as in (14). Assumptions (6) to (8) yield that there are positive constants c and C such that cu

2

H𝛼2(Ω)

⩽ | u

Ω

|

2

+  (u , u)Cu

2

H𝛼2(Ω)

for all uH

[𝛼2]

(Ω).

This implies that the following norm equivalences hold:

 (u , u) ∼u

2

H𝛼2(Ω)

for all uH

𝛼 2

(0)

(Ω) , (16)

 (u, u) + |u

Ω

|

2

∼ ‖u‖

2

H𝛼2(Ω)

for all uH

𝛼2

(Ω), (17)

cf Abels et al.

17, lemma 2.4 and corollary 2.5

In the following, we will use a variational extension of the nonlocal linear operator  (see (1)) by defining  ∶ H

𝛼2

(Ω) → H

𝛼 2

0

(Ω) as

⟨ u , 𝜑⟩

H0𝛼2,H𝛼2

=  (u , 𝜑 ) for all 𝜑H

𝛼2

(Ω) . This implies

⟨ u , 1 ⟩ = (u , 1) = 0 .

We note that  agrees with (1) as soon as uH

loc𝛼

(Ω) ∩ H

𝛼2

(Ω) and 𝜑C

0

(Ω), cf Abels and Kassmann.

23, lemma 4.2

But this weak formulation also includes a natural boundary condition for u, cf Abels et al,

17, theorem 6.1

for a discussion.

We will also need the following regularity result, which essentially states that the operator  is of lower order with respect to the usual Laplace operator. This result is from Abels et al.

17, lemma 2.6

Lemma 2. Let gL

2(0)

(Ω) and 𝜃 > 0. Then the unique solution u ∈ H

1(0)

(Ω) for the problem

−𝜃 ∫

Ω

∇u · ∇𝜑 + (u, 𝜑) = (g)

L2

𝜑 for all 𝜑H

(0)1

(Ω) (18)

belongs to H

loc2

(Ω) and satisfies the estimate

𝜃||∇u||

2L2(Ω)

+ ||u||

2H𝛼∕2(Ω)

C||g||

2L2(Ω)

, where C is independent of 𝜃 > 0 and g.

For the following, let 𝜙 ∶ [a , b] → R be continuous and define 𝜙(x) = +∞ for x ∉ [a , b]. As in Abels set al,

17, section 3

we fix 𝜃 ⩾ 0 and consider the functional

F

𝜃

(c) = 𝜃

2 ∫

Ω

|∇c|

2

dx + 1

2 (c, c) +

Ω

𝜙(c(x)) dx (19)

(6)

where

domF

0

= {

cH

𝛼∕2

(Ω) ∩ L

2(m)

(Ω) ∶ 𝜙(c) ∈ L

1

(Ω) } ,

domF

𝜃

= H

1

(Ω) ∩ domF

0

if 𝜃 > 0 for a given m ∈ (a , b). Moreover, we define

𝜃

(u , v) = 𝜃

Ω

∇u · ∇v dx +  (u , v)

for all u, vH

1

(Ω) if 𝜃 > 0 and u, vH

𝛼/2

(Ω) if 𝜃 = 0.

In the following, 𝜕 F

𝜃

(c) ∶ L

2(m)

(Ω) → (L

2(0)

(Ω)) denotes the subgradient of F

𝜃

at c ∈ domF, ie, w𝜕 F

𝜃

(c) if and only if (w, c

c)

L2

F

𝜃

(c

) − F

𝜃

(c) for all c

L

2(m)

(Ω).

The following characterization of 𝜕 F

𝜃

(c) is an important tool for the existence proof.

Theorem 1. Let 𝜙 ∶ [a, b] → R be a convex function that is twice continuously differentiable in (a, b) and satisfies lim

x→a

𝜙

(x) = −∞, lim

x→b

𝜙

(x) = +∞. Moreover, we set 𝜙(x) = +∞ for x ∉ (a , b) and let F

𝜃

be defined as in (19). Then

𝜕F

𝜃

∶ (𝜕F

𝜃

) L

2(m)

(Ω) → L

2(0)

(Ω) is a single valued, maximal monotone operator with

(𝜕 F

0

) = {c ∈ H

loc𝛼

(Ω) ∩ H

𝛼∕2

(Ω) ∩ L

2(m)

(Ω) ∶ 𝜙

(c) ∈ L

2

(Ω), ∃𝑓 ∈ L

2

(Ω) ∶

(c , 𝜑) +

Ω

𝜙

(c)𝜑 dx = ∫

Ω

𝑓𝜑 dx𝜑H

𝛼∕2

(Ω)}

if 𝜃 = 0 and

(𝜕 F

𝜃

) = {c ∈ H

loc2

(Ω) ∩ H

1

(Ω) ∩ L

2(m)

(Ω) ∶ 𝜙

(c) ∈ L

2

(Ω), ∃𝑓 ∈ L

2

(Ω) ∶

𝜃

(c , 𝜑 ) +

Ω

𝜙

(c) 𝜑 dx =

Ω

𝑓𝜑 dx𝜑H

1

(Ω)}

if 𝜃 > 0 as well as

𝜕F

𝜃

(c) = −𝜃Δc + c + P

0

𝜙

(c) in

(Ω) for 𝜃 ⩾ 0.

Moreover, the following estimates hold

𝜃||c||

2H1

+ ||c||

2H𝛼∕2

+ ||𝜙

(c)||

22

C (

||𝜕F

𝜃

(c)||

22

+ ||c||

22

+ 1 )

Ω

Ω

(𝜙

(c(x)) − 𝜙

(c(𝑦)))(c(x) − c(𝑦))k(x, 𝑦, x𝑦) dx d𝑦

C (

||𝜕F

𝜃

(c)||

22

+ ||c||

22

+ 1 ) 𝜃

Ω

𝜙

′′

(c)|∇c|

2

dxC (

||𝜕F

𝜃

(c)||

22

+ ||c||

22

+ 1 )

(20)

for some constant C > 0 independent of c ∈ (𝜕F

𝜃

) and 𝜃 ⩾ 0.

The result follows from Abels et al.

17, corollary 3.2 and theorem 3.3

3 W E A K S O LU T I O N S A N D M A I N R E S U LT

In this section, we define weak solutions for the system (1)-(4) and (9)-(11) together with a natural boundary condition

for 𝜑 given by the bilinear form  , summarize the assumptions, and state the main result.

(7)

Assumption 1. Let Ω R

d

, d = 2 , 3, be a bounded domain with C

2

-boundary. The following conditions hold true:

1. 𝜌(𝜑) =

12

( ̃𝜌

1

+ ̃𝜌

2

) +

1

2

( ̃𝜌

2

̃𝜌

1

)𝜑 for all 𝜑 ∈ [−1, 1].

2. mC

1

( R ), 𝜂C

0

( R ) and there are constants m

0

, K > 0 such that 0 < m

0

m(s), 𝜂(s)K for all s ∈ R . 3. Ψ ∈ C([−1, 1]) ∩ C

2

((−1, 1)) and

s

lim

→±1

Ψ

(s) = ±∞ , Ψ

′′

(s) ≥ − 𝜅 for some 𝜅 ∈ R . (21) A standard example for a homogeneous free energy density Ψ satisfying the previous conditions is given by (15). Since for solutions we will have 𝜑 (x , t) ∈ [−1 , 1] almost everywhere, we only need the functions m , 𝜂 on this interval. But for simplicity, we assume m , 𝜂 to be defined on R .

Definition 1. Let v

0

L

2𝜎

(Ω) and 𝜑

0

H

𝛼/2

(Ω) with |𝜑

0

| ≤ 1 almost everywhere in Ω and let Assumption 1 be satisfied. Then (v , 𝜑, 𝜇) such that

vBC

w

([0, ∞); L

2𝜎

(Ω)) ∩ L

2

(0, ∞; H

01

(Ω)

d

) ,

𝜑BC

w

([0, ∞); H

𝛼∕2

(Ω)) ∩ L

2uloc

([0, ∞); H

𝛼loc

(Ω)) , Ψ

(𝜑) ∈ L

2uloc

([0, ∞); L

2

(Ω)) , 𝜇L

2uloc

([0 , ∞); H

1

(Ω)) with ∇𝜇 ∈ L

2

(0 , ∞; L

2

(Ω))

is called a weak solution of (1)-(4) and (4)-(9) if the following conditions hold true:

−( 𝜌 v , 𝜕

t

𝜓 )

Q

+ (div( 𝜌 v v) , 𝜓 )

Q

+ (2 𝜂 ( 𝜑 )Dv , D 𝜓 )

Q

− (

(v ⊗ ̃ J) ,𝜓 )

Q

= −( 𝜑𝜇, 𝜓 )

Q

(22)

for all 𝝍C

0

(Ω × (0 , ∞))

d

with div 𝝍 = 0,

−(𝜑, 𝜕

t

𝜓 )

Q

+ (v · ∇𝜑, 𝜓 )

Q

= −(m(𝜑)∇𝜇, ∇𝜓)

Q

(23)

0

Ω

𝜇𝜓 dx dt =

0

Ω

Ψ

( 𝜑 ) 𝜓 dx dt +

∞ 0

 ( 𝜑 (t) , 𝜓 (t)) dt (24) for all 𝜓C

0

((0, ∞); C

1

(Ω)) and

(v, 𝜑)|

t=0

= (v

0

, 𝜑

0

) . (25)

Recall ̃ J = −

̃𝜌2̃𝜌1

2

m( 𝜑 )∇ 𝜇. Finally, the energy inequality E

tot

(𝜑(t), v(t)) +

t

s

Ω

2𝜂(𝜑) |Dv|

2

dx d𝜏 + ∫

t

s

Ω

m(𝜑)|∇𝜇|

2

dx d𝜏

E

tot

( 𝜑 (s) , v(s)) (26)

holds true for all t ∈ [s , ∞) and almost all s ∈ [0 , ∞) (including s = 0). Here E

tot

is as in (13).

The main result of this contribution is as follows:

Theorem 2 (Existence of weak solutions). Let Assumption 1 hold and 𝛼 ∈ (1 , 2). Then for every v

0

L

2𝜎

(Ω) and 𝜑

0

H

𝛼/2

(Ω) such that |𝜑

0

| ≤ 1 almost everywhere and (𝜑

0

)

Ω

∈ (−1, 1), there exists a weak solution (v, 𝜑, 𝜇) of (1)-(4) and (9)-(11).

Remark 1. Using, eg, 𝜑∇𝜇L

2

(0, ∞; L

2

(Ω)), one can consider this term in (1) as a given right-hand side and

obtain the existence of a pressure such that (1) holds in the sense of distributions in the same way as for the single

Navier-Stokes equations, cf, eg, Sohr.

26

(8)

4 A P P ROX I M AT I O N BY A N I M P L I C I T T I M E D I S C R ET I Z AT I O N

Let Ψ be as in Assumption 1. We define Ψ

0

∶ [−1 , 1] → R by Ψ

0

(s) = Ψ(s) + 𝜅

s22

for all s ∈ [a , b]. Then Ψ

0

∶ [−1 , 1] → R is convex and lim

s→±1

Ψ

0

(s) = ±∞. A basic idea for the following is to use this decomposition to split the free energy E

free

into a singular convex part E and a quadratic perturbation. In the equations, this yields a decomposition into a singular monotone operator and a linear remainder. To this end, we define an energy EL

2

(Ω) → R ∪ {+∞} with domain

dom E = {𝜑 ∈ H

𝛼∕2

(Ω) | − 1 ≤ 𝜑 ≤ 1 a.e.}

given by

E(𝜑) = {

1

2

 (𝜑, 𝜑) + ∫

Ω

Ψ

0

(𝜑) dx for 𝜑 ∈ dom E ,

+∞ else . (27)

This yields the decomposition

E

free

(𝜑) = E(𝜑) − 𝜅

2 ||𝜑||

2L2

for all 𝜑 ∈ dom E .

Moreover, E is convex and E = F

0

if one chooses 𝜙 = Ψ

0

and F

0

is as in Subsection 2.1. This is a key relation for the following analysis in order to make use of Theorem 1, which in particular implies that 𝜕 E = 𝜕 F

0

is a maximal monotone operator.

To prove our main result, we discretize our system semi-implicitly in time in a suitable manner. To this end, let h =

1

N

for N ∈ N and v

k

L

2𝜎

(Ω), 𝜑

k

H

1

(Ω) with 𝜑

k

(x) ∈ [−1, 1] almost everywhere and 𝜌

k

=

1

2

( ̃𝜌

1

+ ̃𝜌

2

) +

1

2

( ̃𝜌

2

̃𝜌

1

)𝜑

k

be given. Then Ψ(𝜑

k

) ∈ L

1

(Ω). We also define a smoothing operator P

h

on L

2

(Ω) as follows. We choose u as the solution of the following heat equation: { 𝜕

t

u − Δu = 0 in Ω × (0 , T) ,

u |

t=0

= 𝜑

on Ω ,

𝜕

𝜈

u |

𝜕Ω

= 0 on 𝜕 Ω × (0 , T) ,

where 𝜑

L

2

(Ω), and set P

h

𝜑

∶= u |

t=h

. Then P

h

𝜑

H

2

(Ω) and P

h

𝜑

𝜑

in L

2

(Ω) as h → 0 for all 𝜑

L

2

(Ω).

Moreover, we have |P

h

𝜑

| ≤ 1 in Ω if |𝜑

(x)| ≤ 1 almost everywhere and P

h

𝜑

h→0

𝜑

in H

𝛼2

(Ω) as h → 0 for all 𝜑

H

𝛼2

(Ω).

Now, we determine (v, 𝜑, 𝜇) = (v

k+1

, 𝜑

k+1

, 𝜇

k+1

), k ∈ N , successively as solution of the following problem: Find vH

10

(Ω)

d

L

2𝜎

(Ω), 𝜑 ∈ (𝜕 E) and

𝜇H

n2

(Ω) = {u ∈ H

2

(Ω) | 𝜕

n

u |

𝜕Ω

= 0 on 𝜕 Ω} , such that

( 𝜌v𝜌

k

v

k

h , 𝜓 )

Ω

+ (div(𝜌(P

h

𝜑

k

)v v), 𝜓)

Ω

+ (2𝜂(𝜑

k

)Dv, D𝜓)

Ω

+ (

div(v ⊗ ̃ J), 𝜓 )

Ω

= −((P

h

𝜑

k

)∇𝜇, 𝜓 )

Ω

(28)

for all 𝜓C

0,𝜎

(Ω),

𝜑𝜑

k

h + v · ∇P

h

𝜑

k

= div (m(P

h

𝜑

k

)∇𝜇) almost everywhere in Ω, (29) and

Ω

( 𝜇 + 𝜅 𝜑 + 𝜑

k

2

) 𝜓 dx = (𝜑, 𝜓 ) + ∫

Ω

Ψ

0

(𝜑)𝜓 dx + h

Ω

∇𝜑 · ∇𝜓 dx (30)

for all 𝜓H

𝛼/2

(Ω), where

̃ J̃ J

k+1

∶= − ̃𝜌

2

̃𝜌

1

2 m(P

h

𝜑

k

)∇𝜇

k+1

= − ̃𝜌

2

̃𝜌

1

2 m(P

h

𝜑

k

)∇𝜇 . For the following, let

E

tot,h

(𝜑, v) =

Ω

𝜌 |v|

2

2 dx +

Ω

Ψ(𝜑) dx + 1

2 (𝜑, 𝜑) + h

2 ∫

Ω

|∇𝜑|

2

dx (31)

denote the total energy of the system (28)-(30).

(9)

Remark 2.

1. As in 6. Abels et al,

6

we obtain the important relation

𝜌𝜌

k

hv · ∇𝜌(P

h

𝜑

k

) = div ̃ J , by multiplication of (29) with −

̃𝜌2̃𝜌1

2

=

𝜕𝜌(𝜑)

𝜕𝜑

. Because of div(v ⊗ ̃ J) = (div ̃ J)v + ( ̃ J · ∇

)

v, this yields that ( 𝜌 v𝜌

k

v

k

h , 𝜓 )

Ω

+ (div( 𝜌 (P

h

𝜑

k

)v v) , 𝜓 )

Ω

+ (2 𝜂 ( 𝜑

k

)Dv , D 𝜓)

Ω

+ ((

div ̃ J𝜌𝜌

k

hv · ∇𝜌(P

h

𝜑

k

) ) v

2 , 𝜓 )

Ω

+ (( ̃ J · ∇

) v, 𝜓 )

Ω

= −((P

h

𝜑

k

)∇𝜇, 𝜓 )

Ω

(32) for all 𝜓C

0,𝜎

(Ω) to (28), which will be used to derive suitable a priori estimates.

2. Integrating (29) in space, one obtains ∫

Ω

𝜑 dx = ∫

Ω

𝜑

k

dx because of div v = 0 and the boundary conditions.

The following lemma is important to control the derivative of the singular free energy density Ψ

(𝜑).

Lemma 3. Let 𝜑 ∈ (𝜕F

h

) and 𝜇H

1

(Ω) be a solution of (30) for given 𝜑

k

H

1

(Ω) with |𝜑

k

(x)| ≤ 1 almost everywhere in Ω such that

𝜑

Ω

= 1

|Ω| ∫

Ω

𝜑 dx = 1

|Ω| ∫

Ω

𝜑

k

dx ∈ (−1 , 1) . Then there is a constant C = C(∫

Ω

𝜑

k

, Ω) > 0, independent of 𝜑, 𝜇, 𝜑

k

, such that

||Ψ

0

(𝜑)||

L2(Ω)

+ ||

|| ∫

Ω

𝜇 dx ||

|| ≤ C(||∇𝜇||

L2

+ ||∇𝜑||

2L2

+ 1) and

||𝜕 F

h

(𝜑)||

L2(Ω)

C (||𝜇||

L2

+ 1) .

Proof. The proof is an adaptation of the corresponding result in Abels et al.

6

For the convenience of the reader, we give the details. First, we choose 𝜓 = 𝜑𝜑

Ω

in (30) and get

Ω

𝜇(𝜑𝜑

Ω

) dx + ∫

Ω

𝜅 𝜑 + 𝜑

k

2 (𝜑 − 𝜑

Ω

) dx

= (𝜑, 𝜑) + ∫

Ω

Ψ

0

(𝜑)(𝜑 − 𝜑

Ω

) dx + h

Ω

∇𝜑 · ∇𝜑 dx . (33)

Let 𝜇

0

= 𝜇𝜇

Ω

. Then ∫

Ω

𝜇(𝜑𝜑

Ω

) dx = ∫

Ω

𝜇

0

𝜑 dx.

In order to estimate the second term in (32), we use that 𝜑 ∈ (−1 + 𝜀, 1 − 𝜀) for sufficiently small 𝜀 > 0 and that lim

𝜑→±1

Ψ

0

(𝜑) = ±∞. Hence, for sufficiently small 𝜀 , one obtains the inequality Ψ

0

(𝜑)(𝜑 − 𝜑

Ω

) ≥ C

𝜀

0

(𝜑)| − C ̃

𝜀

, which implies

Ω

Ψ

0

(𝜑)(𝜑 − 𝜑

Ω

) dxC

Ω

0

(𝜑)| dxC

1

. Together with (32), we obtain

Ω

0

(𝜑)| dxC ||𝜇

0

||

L2(Ω)

||𝜑||

L2(Ω)

+ C

Ω

𝜅

2 |𝜑 + 𝜑

k

||𝜑 − 𝜑

Ω

| dx + C

1

C(||𝜇

0

||

L2(Ω)

+ ||𝜑||

2L2(Ω)

+ 1)

C(||∇𝜇||

L2(Ω)

+ 1) ,

because of |𝜑| , |𝜑

k

| ≤ 1. Next, we choose 𝜓 ≡ 1 in (30). This yields

Ω

𝜇 dx = ∫

Ω

Ψ

0

(𝜑) dx − ∫

Ω

𝜅

2 (𝜑 + 𝜑

k

) dx .

(10)

Altogether, this leads to

|| || ∫

Ω

𝜇 dx ||

|| ≤ C(||∇𝜇||

L2(Ω)

+ 1) .

Finally, the estimates of 𝜕F

h

(𝜑) and Ψ

0

(𝜑) in L

2

(Ω) follow directly from (30) and (20).

Now, we will prove existence of solution to the time-discrete system. We basically follow the line of the corresponding arguments in Abels et al

6

here. As before, we denote

H

2n

(Ω) ∶= {u ∈ H

2

(Ω) ∶ n · ∇u |

𝜕Ω

= 0}.

Lemma 4. For every v

k

L

2𝜎

(Ω), 𝜑

k

H

1

(Ω) with |𝜑

k

(x)| ≤ 1 almost everywhere, and 𝜌

k

=

1

2

( ̃𝜌

1

+ ̃𝜌

2

) +

1

2

( ̃𝜌

2

̃𝜌

1

)𝜑

k

, there is some solution (v , 𝜑, 𝜇 ) ∈ (

H

01

(Ω)

d

L

2𝜎

(Ω) )

×  ( 𝜕 F

h

) × H

2n

(Ω) of the system (29)-(30) and (32). Moreover, the solution satisfies the discrete energy estimate

E

tot,h

(𝜑, v) +

Ω

𝜌

k

| vv

k

|

2

2 dx + ∫

Ω

|∇𝜑 − ∇𝜑

k

|

2

2 dx + 1

2 (𝜑 − 𝜑

k

, 𝜑𝜑

k

) + h

Ω

2 𝜂 ( 𝜑

k

) | Dv |

2

dx + h

Ω

m( 𝜑

k

) | ∇ 𝜇|

2

dxE

tot,h

( 𝜑

k

, v

k

) . (34)

Proof. As first step, we prove the energy estimate (34) for any solution (v , 𝜑, 𝜇 ) ∈ (

H

01

(Ω)

d

L

2𝜎

(Ω) )

×  ( 𝜕 F

h

) × H

n2

(Ω) of (29)-(30) and (22).

We choose 𝝍 = v in (32) and use that

Ω

(

(div ̃ J) v 2 +

( ̃ J · ∇ )

v )

· v dx = ∫

Ω

div ( ̃ J | v |

2

2 )

dx = 0.

Then we derive as in Abels et al

6

, proof of lemma 4.3

Ω

( div(𝜌(P

h

𝜑

k

)v v) − (∇𝜌(P

h

𝜑

k

) · v) v 2 )

· v dx = ∫

Ω

div (

𝜌(P

h

𝜑

k

)v |v|

2

2

)

dx = 0 ,

due to divv = 0. Next, one easily gets 1

h (𝜌v − 𝜌

k

v

k

) · v = 1 h

( 𝜌 | v |

2

2 − 𝜌

k

| v

k

|

2

2

) + 1

h (𝜌 − 𝜌

k

) | v |

2

2 + 1

h 𝜌

k

| vv

k

|

2

2 .

Therefore, (32) with 𝝍 = v yields 0 = ∫

Ω

𝜌| v |

2

𝜌

k

| v

k

|

2

2h dx + ∫

Ω

𝜌

k

| vv

k

|

2

2h dx + ∫

Ω

2 𝜂(𝜑

k

)| Dv |

2

dx + ∫

Ω

P

h

𝜑

k

∇𝜇 · v dx . (35) Moreover, multiplying (29) with 𝜇 and using the boundary condition for 𝜇 , one concludes

0 =

Ω

𝜑𝜑

k

h 𝜇 dx +

Ω

(v · ∇P

h

𝜑

k

) 𝜇 dx +

Ω

m(P

h

𝜑

k

)|∇𝜇|

2

dx . (36) Furthermore, choosing 𝜓 =

1

h

(𝜑 − 𝜑

k

) in (30), we obtain

0 = ∫

Ω

∇𝜑 · ∇(𝜑 − 𝜑

k

) dx + ∫

Ω

Ψ

0

(𝜑) 𝜑𝜑

k

h dx + 1

h  (𝜑, 𝜑 − 𝜑

k

)

− ∫

Ω

𝜇 𝜑𝜑

k

h dx − ∫

Ω

𝜅 𝜑

2

𝜑

2k

2h dx . (37)

(11)

Summation of (35) to (4) yields

0 = ∫

Ω

𝜌|v|

2

𝜌

k

|v

k

|

2

2h dx + ∫

Ω

𝜌

k

|v − v

k

|

2

2h dx + ∫

Ω

2𝜂(𝜑

k

)|Dv|

2

dx + ∫

Ω

m(P

h

𝜑

k

)|∇𝜇|

2

dx + ∫

Ω

Ψ

0

(𝜑) 𝜑𝜑

k

h dx − ∫

Ω

𝜅 𝜑

2

𝜑

2k

2h dx

+ ∫

Ω

∇𝜑 · ∇(𝜑 − 𝜑

k

) dx + 1

h  (𝜑, 𝜑 − 𝜑

k

)

≥∫

Ω

𝜌|v|

2

𝜌

k

|v

k

|

2

2h dx + ∫

Ω

𝜌

k

|v − v

k

|

2

2h dx + ∫

Ω

2𝜂(𝜑

k

)|Dv|

2

dx + ∫

Ω

m(P

h

𝜑

k

)|∇𝜇|

2

dx + 1

h

Ω

0

(𝜑) − Ψ

0

(𝜑

k

)) dx − ∫

Ω

𝜅 2

𝜑

2

𝜑

2k

h dx + ∫

Ω

|∇𝜑 − ∇𝜑

k

|

2

2 dx + ∫

Ω

( |∇𝜑|

2

2 − |∇𝜑

k

|

2

2

) dx + 1

h

(𝜑, 𝜑)

2 − 1

h

(𝜑

k

, 𝜑

k

)

2 + 1

h

 (𝜑 − 𝜑

k

, 𝜑𝜑

k

)

2 ,

because of ∫

Ω

P

h

𝜑

k

∇𝜇 · v dx = −∫

Ω

(v · ∇P

h

𝜑

k

)𝜇 dx,

Ψ

0

(𝜑) (𝜑 − 𝜑

k

) ≥ Ψ

0

(𝜑) − Ψ

0

(𝜑

k

) ,

∇𝜑 · ∇(𝜑 − 𝜑

k

) = |∇𝜑|

2

2 − |∇𝜑

k

|

2

2 + |∇𝜑 − ∇𝜑

k

|

2

2 , and

 ( 𝜑, 𝜑𝜑

k

) =  ( 𝜑, 𝜑 )

2 −  ( 𝜑

k

, 𝜑

k

)

2 +  ( 𝜑𝜑

k

, 𝜑𝜑

k

)

2 .

This shows (34).

We will prove existence of weak solutions with the aid of the Leray-Schauder principle. In order to obtain a suitable reformulation of our time-discrete system, we define suitable 

k

,

k

XY , where

X = (

H

01

(Ω)

d

L

2𝜎

(Ω) )

× (𝜕 F

h

) × H

2n

(Ω) , Y = (

H

01

(Ω)

d

L

2𝜎

(Ω) )

× L

2

(Ω) × L

2

(Ω) and

k

(w) =

( L

k

(v)

−div(m(P

h

𝜑

k

)∇𝜇) + ∫

Ω

𝜇 dx 𝜑 + 𝜕F

h

(𝜑)

)

for every w = (v , 𝜑, 𝜇) ∈ X and

L

k

(v), 𝜓 ⟩ = ∫

Ω

2 𝜂(𝜑

k

)Dv ∶ D 𝜓 dx for all 𝜓H

01

(Ω)

d

L

2𝜎

(Ω).

Moreover, we define

k

(w) =

⎛ ⎜

⎜ ⎜

𝜌v−𝜌kvk

h

− div(𝜌(P

h

𝜑

k

)v v) − ∇𝜇 P

h

𝜑

k

( diṽ J

𝜌−𝜌k

h

v · ∇𝜌(P

h

𝜑

k

) )

v

2

− ( ̃ J · ∇

) v

𝜑−𝜑k

h

v · ∇P

h

𝜑

k

+ ∫

Ω

𝜇 dx 𝜑 + 𝜇 + ̃𝜅

𝜑+𝜑2 k

⎞ ⎟

⎟ ⎟

⎠ for w = (v, 𝜑, 𝜇) ∈ X. By construction, w = (v, 𝜑, 𝜇) ∈ X is a solution of (28) to (30) if and only if

k

(w) − 

k

(w) = 0 .

(12)

In Abels et al, [section 4.2]

6

it is shown that

L

k

H

01

(Ω)

d

L

2𝜎

(Ω) → (

H

01

(Ω)

d

L

2𝜎

(Ω) )

is invertible and that for every fL

2

(Ω),

−div(m(P

h

𝜑

k

)∇ 𝜇 ) +

Ω

𝜇 dx = 𝑓 in Ω , 𝜕

n

𝜇|

𝜕Ω

= 0 (38) has a unique solution 𝜇H

n2

(Ω). This follows from the Lax-Milgram Theorem and elliptic regularity theory.

Moreover, in Abels et al,

6, section 4.2

the estimate

||𝜇||

H2(Ω)

C

k

(

||𝜇||

H1(Ω)

+ ||𝑓 ||

L2(Ω)

) (39)

is shown.

Because of Theorem 1, 𝜕 F

h

is maximal monotone, and therefore, I + 𝜕F

h

∶ (𝜕F

h

) → L

2

(Ω)

is invertible. Moreover, (I + 𝜕F

h

)

−1

L

2

(Ω) → H

1

(Ω) is continuous, which can be shown as in the proof of proposition 7.5.5 in Abels.

27

Since now, a nonlocal operator is involved, we provide the details for the convenience of the reader.

Let f

l

l→∞

f in L

2

(Ω) such that f

l

= u

l

+ 𝜕F(u

l

) and f = u + 𝜕F(u) be given. Then u

l

u in H

1

(Ω) since

|| u

l

u ||

2L2

+ h ||∇u

l

− ∇u ||

2L2

+ (u

l

u , u

l

u) ≤ || u

l

u ||

2L2

+ (𝜕 F

h

(u

l

) − 𝜕 F

h

(u), u

l

u)

L2

≤ || u

l

+ 𝜕 F

h

(u

l

) − (u + 𝜕 F

h

(u))||

L2

|| u

l

u ||

L2

≤ 1

2 ||𝑓

l

𝑓 ||

2L2

+ 1

2 || u

l

u ||

2L2

. Altogether, 

k

XY is invertible with continuous inverse 

−1k

YX.

We introduce the following auxiliary Banach spaces X ̃ ∶= (

H

01

(Ω)

d

L

2𝜎

(Ω) )

× H

1

(Ω) × H

n2

(Ω), Y ̃ ∶= L

32

(Ω)

d

× W

13

2

(Ω) × H

1

(Ω)

in order to obtain a completely continuous mapping in the following. Because of the considerations above, 

−1k

YX ̃ is continuous. Because of the compact embedding Y ̃ →→ Y,

−1k

Y ̃X ̃ is compact.

Next, we show that 

k

X ̃Y ̃ is continuous and bounded. To this end, one uses the estimates:

||𝜌 v ||

L32(Ω)

C || v ||

H1(Ω)

(||𝜑||

L2(Ω)

+ 1) , || div(𝜌(P

h

𝜑

k

)v v)||

L32(Ω)

C

k

|| v ||

2H1(Ω)

,

||∇𝜇 P

h

𝜑

k

||

L32(Ω)

C

k

||∇𝜇||

L2(Ω)

, ||(diṽ J)v ||

L32(Ω)

C

k

|| v ||

H1(Ω)

||𝜇||

H2(Ω)

,

||( ̃ J · ∇)v||

L32(Ω)

C||v||

H1(Ω)

||𝜇||

H2(Ω)

, ||v · ∇𝜑

k

||

W13

2

(Ω)

C

k

||v||

H1(Ω)

. Note that P

h

𝜑

k

and therefore 𝜌(P

h

𝜑

k

) belong to H

2

(Ω)). More precisely,

1. For the estimate of div(𝜌(P

h

𝜑

k

)v v) in L

32

(Ω), one has to estimate a term of the form 𝜌(P

h

𝜑

k

)𝜕

l

v

i

v

j

in L

32

(Ω), which are a product of functions in L

(Ω), L

2

(Ω) and L

6

(Ω). Therefore, the term is bounded in L

32

(Ω). Moreover, there are terms of the form 𝜕

l

𝜌(P

h

𝜑

k

)v

i

v

j

in L

32

(Ω), where each factor belongs to L

6

(Ω).

2. To estimate (div ̃ J)v in L

32

(Ω), one has terms of the form m

(P

h

𝜑

k

)𝜕

i

P

h

p

k

𝜕

j

𝜇v

l

and of the form m(P

h

𝜑

k

)𝜕

i

𝜕

j

𝜇v

l

. For

the first type of terms, the first factor is in L

(Ω), and the other three are in L

6

(Ω), which yields the bound in

L

32

(Ω). The second type are products of functions in L

(Ω), L

2

(Ω), and L

6

(Ω).

Referenzen

ÄHNLICHE DOKUMENTE

The free boundary problem for which existence and uniqueness of a classical solution is shown in the preceding chapter is considered in its New- tonian version, and a numerical

Studies of other groups reveal that the main source of spurious velocities in two-phase flow with surface tension is the fact that discontinuous functions allowing for jumps at

A new diffuse interface model for a two-phase flow of two incompressible fluids with different densities is introduced using methods from rational con- tinuum mechanics. The

R¨ oger, Existence of weak solutions for a non-classical sharp interface model for a two-phase flow of viscous, incompressible fluids, Ann.. Amann, Quasilinear parabolic problems

A new diffuse interface model for a two-phase flow of two incompressible fluids with different densities is introduced using methods from rational con- tinuum mechanics. The

Gasi ´nski, L., Winkert, P.: Sign changing solution for a double phase problem with nonlinear boundary condition via the Nehari manifold.. Gu, L.: Second order parabolic

• As an example for high Reynolds number flow, the plane channel flow at Re τ = 4800 is surveyed with two models of hybrid approach, Smagorinsky model with wall functions and SADES

[38] Pierre-Louis Lions and Panagiotis E. Fully nonlinear stochastic partial differential equations. Paris S´ er. Fully nonlinear stochastic partial differential equations: