• Nie Znaleziono Wyników

ON NONSTATIONARY MOTION OF A FIXED MASS OF A GENERAL VISCOUS COMPRESSIBLE HEAT

N/A
N/A
Protected

Academic year: 2021

Share "ON NONSTATIONARY MOTION OF A FIXED MASS OF A GENERAL VISCOUS COMPRESSIBLE HEAT"

Copied!
23
0
0

Pełen tekst

(1)

25,4 (1999), pp. 489–511

E. Z A D R Z Y ´N S K A (Warszawa)

ON NONSTATIONARY MOTION OF A FIXED MASS OF A GENERAL VISCOUS COMPRESSIBLE HEAT

CONDUCTING CAPILLARY FLUID BOUNDED BY A FREE BOUNDARY

Abstract. The motion of a fixed mass of a viscous compressible heat conducting capillary fluid is examined. Assuming that the initial data are sufficiently close to a constant state and the external force vanishes we prove the existence of a global-in-time solution which is close to the constant state for any moment of time. Moreover, we present an analogous result for the case of a barotropic viscous compressible fluid.

1. Introduction. The aim of this paper is to prove the global existence theorem for a free boundary problem for equations of a viscous compressible heat conducting capillary fluid in the general case, i.e. without assuming any conditions on the form of the internal energy.

In papers [13], [18], [19] the global existence theorem was proved under the assumption of a special form of the internal energy ε = ε(̺, θ), where

̺ is the density of the fluid and θ is the temperature. More precisely, we assumed

ε(̺, θ) = a

0

̺

α

+ h(̺, θ),

where a

0

> 0, α > 0, h(̺, θ) ≥ h

≥ 0 for ̺ ∈ [̺

, ̺

], θ ∈ [θ

, θ

]; a

0

, α, h are constants, and

̺

= min

t∈[0,T ]

min

t

̺(x, t), ̺

= max

t∈[0,T ]

max

t

̺(x, t), θ

= min

t∈[0,T ]

min

t

θ(x, t), θ

= max

t∈[0,T ]

max

t

θ(x, t), T is the time of local existence of a solution.

1991 Mathematics Subject Classification: 35A05, 35R35, 76N10.

Key words and phrases: free boundary, compressible viscous heat conducting fluids, global existence.

[489]

(2)

In this paper we consider the motion of a fluid in a bounded domain Ω

t

⊂ R

3

which depends on time t ∈ R

+

. The shape of the free boundary S

t

of Ω

t

is governed by surface tension. Let v = v(x, t) be the velocity of the fluid, ̺ = ̺(x, t) the density, θ = θ(x, t) the temperature, f = f (x, t) the external force per unit mass, r = r(x, t) the heat sources per unit mass, θ = θ(x, t) the heat flow per unit surface, p = p(̺, θ) the pressure, µ and ν the viscosity coefficients, κ the coefficient of heat conductivity, c

v

= c

v

(̺, θ) the specific heat at constant volume and p

0

the external (constant) pressure.

Then the motion of the fluid is described by the following system (see [1], Chs. 2 and 5):

(1.1)

̺[v

t

+ (v · ∇)v] + ∇p − µ∆v − ν∇ div v = ̺f in e Ω

T

,

̺

t

+ div(̺v) = 0 in e Ω

T

,

̺c

v

t

+ v · ∇θ) + θp

θ

div v − κ∆θ

− µ 2

X

3 i,j=1

(v

ixj

+ v

jxi

)

2

− (ν − µ)(div v)

2

= ̺r in e Ω

T

,

T n − σHn = −p

0

n on e S

T

,

v · n = −ϕ

t

/|∇ϕ| on e S

T

,

∂θ/∂n = θ on e S

T

,

v|

t=0

= v

0

, ̺|

t=0

= ̺

0

, θ|

t=0

= θ

0

in Ω,

where ϕ(x, t) = 0 describes S

t

, n is the unit outward normal vector to the boundary, e Ω

T

= S

t∈(0,T )

t

× {t}, Ω

0

= Ω is the initial domain, and S e

T

= S

t∈(0,T )

S

t

× {t}. Moreover, T = T(v, p) = {T

ij

}

i,j=1,2,3

= {−pδ

ij

+ µ(v

ixj

+ v

jxi

) + (ν − µ)δ

ij

div v}

i,j=1,2,3

is the stress tensor and H is the double mean curvature of S

t

which is negative for convex domains and can be expressed in the form

Hn = ∆(t)x, x = (x

1

, x

2

, x

3

), where ∆(t) is the Laplace–Beltrami operator on S

t

.

Let S

t

be determined by

x = x(s

1

, s

2

, t), (s

1

, s

2

) ∈ U ⊂ R

2

, where U is an open set. Then

∆(t) = g

−1/2

∂s

α



g

−1/2

bg

αβ

∂s

β



= g

−1/2

∂s

α



g

1/2

g

αβ

∂s

β



, α, β = 1, 2,

where the summation convention over the repeated indices is assumed, g =

det{g

α,β

}

αβ=1,2

, g

αβ

= x

α

· x

β

(x

α

= ∂x/∂s

α

), {g

αβ

} is the inverse matrix

to {g

αβ

} and {bg

αβ

} is the matrix of algebraic complements of {g

αβ

}.

(3)

Assume that the domain Ω is given. Then by (1.1)

5

, Ω

t

= {x ∈ R

3

: x = x(ξ, t), ξ ∈ Ω}, where x = x(ξ, t) is the solution of the Cauchy problem

∂x

∂t = v(x, t), x|

t=0

= ξ ∈ Ω, ξ = (ξ

1

, ξ

2

, ξ

3

).

Hence, we obtain the following relation between the Eulerian x and the Lagrangian ξ coordinates of the same fluid particle:

x = ξ +

t

\

0

u(ξ, t

) dt

≡ X

u

(ξ, t),

where u(ξ, t) = v(X

u

(ξ, t), t). Moreover, by (1.1)

5

, S

t

= {x : x = x(ξ, t), ξ ∈ S = ∂Ω}.

By the continuity equation (1.1)

2

and the kinematic condition (1.1)

5

the total mass is conserved, i.e.

(1.2)

\

t

̺(x, t) dx =

\

̺

0

(ξ) dξ = M, where M is a given constant.

Moreover, in view of thermodynamic considerations assume c

v

> 0, κ > 0, ν >

13

µ > 0.

In this paper we prove the existence of a global-in-time solution of prob- lem (1.1) near a constant state.

Assume that p

̺

> 0, p

θ

> 0 for ̺, θ ∈ R

+

and consider the equation

(1.3) p

 M

4

3

πR

3e

, θ

e



= p

0

+ 2σ R

e

.

We assume that there exist R

e

> 0 and θ

e

> 0 satisfying (1.3). Then we introduce the following definition.

Definition 1.1. Let f = r = θ = 0. By a constant (equilibrium) state we mean a solution (v, θ, ̺, Ω

t

) of problem (1.1) such that v = 0, θ = θ

e

,

̺ = ̺

e

, Ω

t

= Ω

e

for t ≥ 0, where ̺

e

= M/

43

πR

3e



, Ω

e

is a ball of radius R

e

, and R

e

> 0 and θ

e

> 0 satisfy equation (1.3).

The methods used to prove the main result of the paper, Theorem 3.5, are similar to those applied in [11], [14]–[19], [21]–[23] and [2]–[6]. To prove the global existence theorem (Theorem 3.5) we use the local existence theorem of [12], the differential inequality (3.5) which is similar to the differential inequalities derived in [11], [16], [17] and [21]–[23] and the conservation laws for energy, mass and momentum which are presented together with their consequences in Section 2. Theorem 3.5 is proved without assuming any conditions on the form of the internal energy ε = ε(̺, θ).

Theorem 3.7 is the global existence theorem for the case p

0

= 0.

(4)

In Section 4 we present the global existence theorem for the case of a viscous compressible barotropic capillary fluid (Theorem 4.3).

In contrast to [22]–[25] we do not assume any conditions on the form of the pressure p = p(̺). The case of a general p = p(̺) and σ = 0 was examined in [21], where the global existence theorem was proved. On the other hand papers [7]–[8] are devoted to the global motion of a viscous compressible barotropic fluid in the case of a general p = p(̺), σ > 0 and p

0

= 0.

Papers [10]–[11] are concerned with the global motion of a viscous com- pressible barotropic self-gravitating fluid in the case when p = A̺

κ

, where A > 0 and κ > 1 are constants.

Finally, in [9], [12], [20] and [24] local existence theorems are proved, while [2]–[6] are devoted to the motion of a viscous compressible heat- conducting fluid both in the space R

3

and in a fixed domain.

Now, we present the notation used in the paper. We denote by W

2l,1/2

(Q

T

) the anisotropic Sobolev–Slobodetski˘ı spaces of functions defined in Q

T

, where Q

T

= Ω

T

= Ω × (0, T ) (Ω ⊂ R

3

is a domain, T < ∞ or T = ∞) or Q

T

= S

T

= S × (0, T ), S = ∂Ω. We define W

2l,l/2

(Ω

T

) as the space of functions u such that

kuk

Wl,l/2

2 (ΩT)

=

 X

|α|+2i≤[l]

kD

ξα

it

uk

2L2(ΩT)

+ X

|α|+2i=[l]



T\

0

\

\

|D

ξα

ti

u(ξ, t) − D

ξα

ti

u(ξ

, t)|

2

|ξ − ξ

|

3+2(l−[l])

dξ dξ

dt

+

\

T

\

0 T

\

0

|D

αξ

ti

u(ξ, t) − D

ξα

ti

u(ξ, t

)|

2

|t − t

|

1+2(l/2−[l/2])

dt dt



1/2

< ∞, where we use generalized derivatives, D

αξ

= ∂

ξα11

ξα22

ξα33

, ∂

ξαjj

= ∂

αj

/∂ξ

jαj

(j = 1, 2, 3), α = (α

1

, α

2

, α

3

) is a multi-index, |α| = α

1

2

3

, ∂

ti

= ∂

i

/∂t

i

and [l] is the integer part of l. In the case when l is an integer the second term in the above formula must be omitted, and in the case of l/2 integer also the last term is omitted.

The space W

2l,l/2

(S

T

) is defined similarly by using local charts and a partition of unity.

By W

2l

(Q), where l ∈ R

+

, Q = Ω, S, S

1

(Ω ⊂ R

3

is a bounded domain;

S = ∂Ω, S

1

is the unit sphere), we denote the usual Sobolev–Slobodetski˘ı spaces. To simplify notation we write

kuk

l,Q

= kuk

Wl,l/2

2 (Q)

if Q = Ω

T

or Q = S

T

, kuk

l,Q

= kuk

Wl

2(Q)

if Q = Ω or Q = S or Q = S

1

.

(5)

Next, we introduce the spaces Γ

kl

(Q) and Γ

kl,l/2

(Q) of functions u defined on Q × (0, T ) (T < ∞ or T = ∞, Q = Ω, S) such that

|u|

l,k,Q

≡ kuk

Γl

k(Q)

= X

i≤l−k

k∂

ti

uk

l−i,Q

< ∞ and

kuk

Γl,l/2

k (Q)

= X

2i≤l−k

k∂

it

uk

l−2i,Q

< ∞, where l ∈ R

+

, k ≥ 0.

By u

l,0,p,ΩT

we denote the norm in the space L

p

(0, T ; Γ

0l,l/2

(Ω)) and by C

B2,1

(Q) (Q ⊂ R

3

× [0, ∞)) the space of functions such that D

xα

ti

u ∈ C

B0

(Q) for |α| + 2i ≤ 2 (where C

B0

(Q) is the space of continuous bounded functions on Q).

Finally, we introduce the seminorm u

κ,ST

=



T\

0

kuk

20,S

t

dt



1/2

.

2. Conservation laws and their consequences. The following lemma is proved in [15].

Lemma 2.1. For sufficiently regular solutions (v, θ, ̺) of problem (1.1) we have

(2.1) d dt



\

t

̺

 v

2

2 + ε



dx + p

0

|Ω

t

| + σ|S

t

|



− κ

\

St

θ ds

=

\

t

̺f · v dx (conservation of energy), where |Ω

t

| = vol Ω

t

, |S

t

| is the surface area of S

t

, and ε = ε(̺, θ) is the internal energy per unit mass. Moreover ,

d dt

\

t

̺x dx =

\

t

̺v dx and

d dt

\

t

̺v · η dx =

\

t

̺f · η dx,

where η = a + b × x and a, b are arbitrary constant vectors.

Now, assume:

f = 0, θ ≥ 0, (2.2)

̺

1

< ̺(x, t) < ̺

2

, θ

1

< θ(x, t) < θ

2

for all x ∈ Ω

t

and t ∈ [0, T ],

(2.3)

(6)

where T is the time of existence of a solution of problem (1.1); 0 < ̺

1

< ̺

2

and 0 < θ

1

< θ

2

are constants; and

(2.4) ε

1

< ε(̺, θ) < ε

2

for all ̺ ∈ (̺

1

, ̺

2

) and θ ∈ (θ

1

, θ

2

).

Integrating (2.1) with respect to t in an interval (0, t) (t ≤ T ) and using (2.2)–(2.4) we get

(2.5) ε

1

̺

α2

\

t

̺

β

dx +

\

t

̺v

2

2 dx + p

0

|Ω

t

| + σ|S

t

|

\

̺

0

 v

02

2 + ε

0



dξ + p

0

|Ω| + σ|S| + κ sup

t t

\

0

dt

\

St′

θ(s, t

) ds ≡ d, where β = α + 1, α > 0 is a constant, ε

0

= ε(̺

0

, θ

0

). Hence in the same way as in Lemma 2 of [15] we obtain

Lemma 2.2. Under assumptions (2.2)–(2.4) the following estimate holds:

 M

β

ε

1

α2



1/(β−1)

≤ |Ω

t

| ≤ d p

0

.

Let R

t

be the radius of a ball of volume |Ω

t

|. Then inequality (2.5) yields (2.6) ε

1

̺

α2

\

t

̺

β

dx + p

0

|Ω

t

| + σec|Ω

t

|

2/3

− d +

\

t

̺ v

2

2 dx + σ(|S

t

| − 4πR

2t

) ≤ 0, where ec = (36π)

1/3

. Multiplying (2.6) by |Ω

t

|

β−1

and using (1.2) we have (2.7) y(|Ω

t

|) + ε

1

̺

α2

h |Ω

t

|

β−1

\

t

̺

β

dx − 

\

t

̺ dx 

β

i

+ |Ω

t

|

β−1

\

t

̺ v

2

2 dx + σ|Ω

t

|

β−1

(|S

t

| − 4πR

2t

) ≤ 0, where

y(x) = p

0

x

β

+ σecx

β−1/3

− dx

β−1

+ ε

1

̺

α2

M

β

.

Since the last three terms in (2.7) are non-negative we have y(|Ω

t

|) ≤ 0, so we have to consider y = y(x) for x > 0 only.

To do this introduce (as in [15])

D = ν

0

0

− 2µ

30

), where

µ

0

= ecσ(β − 1/3)

3p

0

β , ν

0

= d(β − 1)

2p

0

β .

(7)

We have the following possibilities:

if ν

0

∈ (2µ

30

, ∞) ≡ I

1

, then D > 0, (2.8)

if ν

0

∈ (µ

30

, 2µ

30

] ≡ I

2

, then D ≤ 0, (2.9)

if ν

0

∈ (0, µ

30

] ≡ I

3

, then D < 0.

(2.10)

For ν

0

∈ I

i

, we define ϕ

i

, i = 1, 2, 3, by cosh ϕ

1

≡ ν

0

µ

30

− 1, where ν

0

∈ I

1

; (2.11)

cos ϕ

2

≡ ν

0

µ

30

− 1, where ν

0

∈ I

2

; (2.12)

cos ϕ

3

≡ 1 − ν

0

µ

30

, where ν

0

∈ I

3

. (2.13)

Next, set

(2.14) Φ

1

0

, ϕ

1

, p

0

, β, ε

1

, ̺

2

, M ) = p

0

µ

0

β − 1



2 cosh ϕ

1

3 − 1



3(β−1)

·

 2



cosh ϕ

1

+ 1



− β − 1 β − 1/3



2 cosh ϕ

1

3 − 1



2



− ε

1

̺

α2

M

β

, (2.15) Φ

2

0

, ϕ

2

, p

0

, β, ε

1

, ̺

2

, M ) = p

0

µ

0

β − 1



2 cos ϕ

1

3 − 1



3(β−1)

·



2(cos ϕ

2

+ 1) − β − 1 β − 1/3



2 cos ϕ

2

3 − 1



2



− ε

1

̺

α2

M

β

, (2.16) Φ

3

0

, ϕ

3

, p

0

, β, ε

1

, ̺

2

, M ) = p

0

µ

0

β − 1

 2 cos

 π 3 − ϕ

3

3



− 1



3(β−1)

·

 2



1 − cos ϕ

2



− β − 1 β − 1/3

 2 cos

 π 3 − ϕ

3

3



− 1



2



− ε

1

̺

α2

M

β

. In the same way as Theorem 1 of [15] the following theorem can be proved.

Theorem 2.3. Let conditions (2.2)–(2.4) be satisfied. Let δ

0

∈ (0, 1) be given. Assume that the parameters µ

0

, ν

0

, p

0

, β, ε

1

, ̺

2

, M satisfy one of the relations

(2.17)

i

ν

0

∈ I

i

, 0 < Φ

i

0

, ϕ

i

, p

0

, β, ε

1

, ̺

2

, M ) ≤ δ

0

,

i = 1, 2, 3, where I

i

are defined in (2.8)–(2.10), and Φ

i

are given by (2.14)–

(2.16). Then there exists a constant c

1

independent of δ

0

(it can depend on the parameters ) such that

(2.18) var

0≤t≤T

|Ω

t

| ≤ c

1

δ,

where δ

2

= cδ

0

, c > 0 is a constant.

(8)

Moreover , in the case (2.17)

i

we have

(2.19) | |Ω

t

| − Q

i

| ≤ c

2

δ, t ∈ [0, T ],

where Q

1

= µ

30

(2 cosh(ϕ

1

/3) − 1)

3

, Q

2

= µ

30

(2 cos(ϕ

2

/3) − 1)

3

and Q

3

= µ

30

[2 cos(π/3 − ϕ

3

/3) − 1]

3

, and c

2

> 0 is a constant independent of δ

0

.

Remark 2.4. It can be proved in the same way as in Lemma 4 of [15]

that for any δ

0

sufficiently small and for any 1 ≤ i ≤ 3 there exist parameters p

0

, µ

0

, ν

0

, β, ε

1

, ̺

2

, M such that relation (2.17)

i

is satisfied.

Now, consider the case p

0

= 0. Instead of (2.7) we have in this case y

0

(|Ω

t

|) + ε

1

̺

α2

h |Ω

t

|

β−1

\

t

̺

β

dx − 

\

t

̺ dx 

β

i

+ |Ω

t

|

β−1

\

t

̺ v

2

2 dx + σ|Ω

t

|

β−1

(|S

t

| − 4πR

2t

) ≤ 0, where

y

0

(x) = σecx

β−1/3

− d

0

x

β−1

+ ε

1

̺

α2

M

β

, d

0

=

\

̺

0

 v

20

2 + ε

0



dξ + σ|S| + κ sup

t t

\

0

dt

\

St′

θ(s, t

) ds.

In this case the following theorem analogous to Theorem 2 of [15] holds:

Theorem 2.5. Let p

0

= 0 and let assumptions (2.2)–(2.4) be satisfied.

Moreover , assume that

\

̺

0

v

20

2 dξ +

\

̺

0

(ε(̺

0

, θ

0

) − ε(̺

e

, θ

e

)) dξ + κ sup

t t

\

0

dt

\

St′

θ(s, t

) ds ≤ δ

0

,

\

0

− ̺

e

| dξ ≤ δ

0

,

| |S| − |S

e

| | ≤ δ

0

, (2.20) 0 <

 2

3 (β − 1)

3(β−1)/2

(β − 1/3)

−(3β−1)/2

(ecσ)

−(3β−1)/2

· (ecσ|Ω

e

|

1/3

+ ̺

e

ε(̺

e

, θ

e

))

(3β−1)/2

|Ω

e

|

(β−1)/2

− ε

1

̺

α2

̺

βe



|Ω

e

|

β

≤ δ

0

, where δ

0

> 0 is a sufficiently small constant, |S

e

| = 4πR

2e

, and ̺

e

, R

e

, Ω

e

are introduced in Definition 1.1. Then

0≤t≤T

var |Ω

t

| ≤ c

2

δ,

where c

2

> 0 is a constant independent of δ

0

, δ

2

= cδ

0

and c > 0 is a

constant.

(9)

Remark 2.6. There exist β, δ, ε

1

, ̺

2

, ̺

e

, θ

e

, |Ω

e

| such that condition (2.20) is satisfied. In fact, assuming

(2.21) ecσ|Ω

e

|

−1/3

β − 1 = ̺

e

ε(̺

e

, θ

e

) we have

(2.22)

 2

3 (β − 1)

3(β−1)/2

(β − 1/3)

−(3β−1)/2

(ecσ)

−3(β−1)/2

·

 β

β − 1 ecσ|Ω

e

|

−1/3



(3β−1)/2

|Ω

e

|

(β−1)/2

− ε

1

̺

α2

̺

βe



|Ω

e

|

β

=

 2

3 ecσ β

(3β−1)/2

(β − 1)(β − 1/3)

(3β−1)/2

|Ω

e

|

1/3

− ε

1

̺

α2

̺

βe



|Ω

e

|

β

=

 2 3

 β

β − 1/3



(3β−1)/2

̺

e

ε(̺

e

, θ

e

) − ε

1

̺

α2

̺

βe



|Ω

e

|

β

.

Taking β sufficiently close to 1 and choosing σ, ̺

e

, θ

e

, |Ω

e

|, ε

1

, ̺

2

satisfying (2.21) and (2.22) we see that condition (2.20) also holds.

3. Global existence of solutions of problem (1.2). In [12] (see also [19]) we proved the existence of a sufficiently smooth local solution of problem (1.1). In order to show the global existence we assume the following condition:

(A) Ω

t

is diffeomorphic to a ball, so S

t

can be described by (3.1) |x| = r = R(ω, t), ω ∈ S

1

,

where S

1

is the unit sphere and we consider the motion near the constant state (see Definition 1.1). Define

p

σ

= p − p

0

− 2σ

R , ̺

σ

= ̺ − ̺

e

, ϑ

0

= θ − θ

e

.

Using the Taylor formula p

σ

can be written as (see [17], formula (3.2)) (3.2) p

σ

= p

1

̺

σ

+ p

2

ϑ

0

,

where p

i

(i = 1, 2) are positive functions. Formula (3.2) yields (3.3) kϑ

0

k

20,Ωt

≤ c

3

(kp

σ

k

20,Ωt

+ k̺

σ

k

20,Ωt

).

Next, by the Poincar´e inequality we have

σ

k

20,Ωt

≤ k̺ − ̺

t

k

20,Ωt

+ k̺

t

− ̺

e

k

20,Ωt

(3.4)

≤ c

4

σx

k

20,Ωt

+ k̺

t

− ̺

e

k

20,Ωt

, where ̺

t

=

|Ω1

t|

T

t

̺ dx and c

4

> 0 is a constant depending on Ω

t

.

(10)

In the same way as the differential inequality (3.46) of [17] and by using (3.3)–(3.4), the following inequality can be proved:

dt + c

0

Φ ≤ c

5

P (X)X(1 + X

3

) 

X + Φ +

t

\

0

kvk

24,Ω

t′

dt

 (3.5)

+ c

6

F + c

7

ψ + c

8

kH(·, 0) + 2/R

e

k

42,S1

+ εc

9

(kH(·, 0) + 2/R

e

k

22,S1

+ kR(·, t) − R(·, 0)k

24,S1

) + c

10

 kR(·, t) − R(·, 0)k

24+1/2,S1

t

\

0

v dt

2 3,St

+ kR(·, t) − R(·, 0)k

23,S1

t

\

0

v dt

2

4,St

 ,

where

c

11

ϕ

0

(t) ≤ ϕ(t) ≤ c

12

ϕ

0

(t) and

ϕ

0

(t) = |v|

23,0,Ωt

+ |̺

σ

|

23,0,Ωt

+ |ϑ

0

|

23,0,Ωt

+

t

\

0

v dt

2 4,St

t

\

0

v dt

2 0,St

+ |v|

23,1,St

+ kH(·, 0) + 2/R

e

k

22,S1

, Φ(t) = |v|

24,1,Ωt

+ |̺

σ

|

23,0,Ωt

+ |ϑ

0

|

24,1,Ωt

,

X(t) = |v|

23,0,Ωt

+ |̺

σ

|

23,0,Ωt

+ |ϑ

0

|

23,0,Ωt

+

t

\

0

kvk

23,Ω

t′

dt

,

ψ(t) = kvk

20,Ωt

+ kp

σ

k

20,Ωt

+ kR(·, t) − R(·, 0)k

20,S1

+ k̺

t

− ̺

e

k

20,Ωt

, F (t) = kr

ttt

k

20,R3

+ |r|

22,0,R3

+ krk

0,R3

+ kθk

24,1,R3

+ kθk

1,R3

,

t ∈ [0, T ] (T is the time of local existence); 0 < c

0

< 1 is a constant depending on ̺

1

, ̺

2

, θ

1

, θ

2

, µ, ν and κ; c

i

> 0 (i = 5, . . . , 12) are constants depending on ̺

1

, ̺

2

, θ

1

, θ

2

, T ,

TT

0

kvk

23,Ω

t′

dt

, kSk

4+1/2

and on the constants from the imbedding lemmas and the Korn inequalities; ε > 0 is a small parameter and P is a positive continuous increasing function.

In order to prove the global existence assume also

(3.6) sup

t∈[0,T ]

F (t) ≤ δ, where δ > 0 is sufficiently small.

Next, introduce the spaces

N (t) = (v, ϑ

0

, ̺

σ

) : ϕ

0

(t) < ∞},

(11)

M (t) = n

(v, ϑ

0

, ̺

σ

) : ϕ

0

(t) +

t

\

0

Φ(t

) dt

< ∞ o . In [19] the following lemma is proved:

Lemma 3.1 (see [19], Lemma 5.1). Let the assumptions of Theorem 4.2 of [12] be satisfied. Let the initial data v

0

, ̺

0

, θ

0

, S of problem (1.1) be such that (v, ϑ

0

, ̺

σ

) ∈ N(0) and S ∈ W

24+1/2

. Let

\

̺

0

v

0

(a + b × ξ) dξ = 0,

\

̺

0

ξ dξ = 0,

for all constant vectors a, b. Let condition (A) be satisfied and let the initial data v

0

, ̺

0

, θ

0

, S and the parameters p

0

, σ, d, β, κ, M , ε

1

, ̺

2

(d, β, ε

1

and

̺

2

are defined in Section 2) be such that ϕ(0) ≤ α

1

, ω(t) = sup

t≤t

kR(·, t

) − R

e

k

20,S1

≤ α

2

for t ≤ T, χ(0) = kH(·, 0) + 2/R

e

k

22+1/2,S1

≤ α

3

,

where α

1

, α

2

, α

3

> 0 are sufficiently small constants, and T is the time of local existence. Then the local solution of problem (1.1) is such that (v, ϑ

0

, ̺

σ

) ∈ M(t) for t ≤ T and

ϕ(t) +

t

\

0

Φ(t

) dt

≤ c

14

(ϕ(0) + χ(0) + ω(t) + sup

t∈[0,T ]

F (t))

≤ c

14

1

+ α

2

+ α

3

+ δ).

Now, we prove

Lemma 3.2. Assume that there exists a local solution to problem (1.1) which belongs to M(t) for t ≤ T . Let the assumptions of Lemma 3.1 and Theorem 2.3 be satisfied. Moreover , if (2.17)

i

holds then assume that

(3.7) |Q

i

− |Ω

e

|| ≤ δ

1

,

where δ

1

> 0 is sufficiently small and (3.8)

\

̺

0

v

20

2 dξ +

\

̺

0

(ε(̺

0

, θ

0

) − ε

1

) dξ

+ p

0

(|Ω| − Q

i

+ δ

2

) + σ[|S| − ec(Q

i

− δ

2

)

2/3

] + κ sup

t t

\

0

dt

\

St′

θ(s, t

) ds ≤ δ

3

,

where t ≤ T , δ

2

∈ (0, 1/2] is a constant so small that Q

i

−δ

2

> 0 and assume that δ from (2.19) is so small that c

2

δ ≤ δ

2

. Then

(3.9) ψ(t) ≤ δ

4

for t ≤ T,

(12)

where δ

4

= c

15

(δ + δ

1

+ δ

3

+ δ

α

1

), c

15

> 0 is a constant depending on ̺

1

,

̺

2

, θ

1

, θ

2

, β, d; δ is the constant from estimate (2.18), α

1

is the constant from Lemma 3.1 and δ

∈ (0, 1) is a sufficiently small constant.

P r o o f. First, assumption (3.7) and estimate (2.19) yield

| |Ω

t

| − |Ω

e

| | ≤ c

2

δ + δ

1

. Hence, by Lemma 2.2,

(3.10) k̺

t

− ̺

e

k

20,Ωt

≤ ̺

2e

|Ω

t

| (c

2

δ + δ

1

)

2

≤ ̺

2e

 d̺

α2

M

β

ε

1



1/(β−1)

(c

2

δ + δ

1

)

2

. Next, integrating (2.1) with respect to t in an interval (0, t) (t ≤ T ) we get

(3.11)

\

t

̺

 v

2

2 + ε



dx + p

0

|Ω

t

| + σ|S

t

|

\

̺

0

 v

2

2 + ε(̺

0

, θ

0

)



dξ + p

0

|Ω| + σ|S| + κ sup

t t

\

0

dt

\

St′

θ(s, t

) ds.

Now, let R

t

be the radius of a ball of volume |Ω

t

|. Then by (2.19), ec(Q

i

− δ

2

)

2/3

≤ 4πR

2t

≤ ec(Q

i

+ δ

2

)

2/3

,

where ec = (36π)

1/3

. Hence

|S

t

| − ec(Q

i

− δ

2

)

2/3

= |S

t

| − 4πR

2t

+ 4πR

2t

(3.12)

− ec(Q

i

− δ

2

)

2/3

≥ 0 for t ≤ T.

Using (2.19), (2.3)–(2.4), (3.12) and assumption (3.8) in (3.11) we obtain

\

t

̺ v

2

2 dx +

\

t

̺(ε − ε

1

) dx + p

0

(|Ω

t

| − Q

i

+ δ

2

) + σ[|S

t

| − ec(Q

i

− δ

2

)

2/3

] ≤ δ

3

. Hence

(3.13) kvk

20,Ωt

≤ 2

̺

1

δ

3

.

Next, using the same argument as in the proof of Lemma 5.2 of [23] we obtain the estimate

(3.14) kp

σ

k

20,Ωt

≤ c

16

α

1

δ

+ c(δ

3

,

where c

16

> 0 is a constant, α

1

is the constant from Lemma 3.1, δ

∈ (0, 1)

is a sufficiently small constant, and c(δ

) is a decreasing function of δ

.

(13)

Finally, the estimate

(3.15) kR(ω, t) − R

t

k

21,S1

≤ c

17

δ

3

for t ∈ [0, T ]

follows from Lemma 2.4 of [23]. Hence by (3.15) and (2.18) we have kR(ω, t) − R(0, t)k

21,S1

≤ kR(ω, t) − R

t

k

21,S1

(3.16)

+ c

18

|R

t

− R

0

|

2

+ kR(0, t) − R

0

k

21,S1

≤ c

19

δ

3

+ c

20

δ, where R

0

=

3

|Ω| 

1/3

.

By (3.10), (3.13), (3.14) and (3.16) we get (3.9).

This completes the proof.

Remark 3.3. In the case ε = c

v

θ (c

v

> 0 is a constant), p = R̺θ (R > 0 is a constant) assumption (3.7) is satisfied if

(3.17) M |Ω

e

|

α

(β − 1)c

v

θ

1

 ̺

e

̺

2



α

− Rθ

e

< c

21

δ

5

, where c

21

> 0 is a constant and δ

5

> 0 is sufficiently small.

P r o o f. Consider Q

i

(where i = 1, 2, 3) and set x

0

= Q

i

. Then the equation determining x

0

has the form (see equation (40) of [15])

(3.18) p

0

x

β0

+ ecσ(β − 1/3)β

−1

x

β−1/30

− (β − 1)β

−1

dx

β−10

= 0.

Moreover, from assumption (2.17)

i

it follows (see [15]) that (3.19) 0 ≤ −y(x

0

) = −



p

0

x

β0

+ ecσx

β−1/30

− dx

β−10

+ ε

1

̺

α2

M

β



≤ δ

0

. Applying (3.18) in (3.19) we get

(3.20) 0 ≤ p

0

x

β0

+ 2

3 ecσx

β−1/30

− (β − 1) ε

1

̺

α2

M

β

≤ (β − 1)δ

0

. In this case equation (1.3) takes the form

e

θ

e

= 2σ R

e

+ p

0

. Hence

(3.21) p

0

|Ω

e

|

β

+ 2

3 ecσ|Ω

e

|

β−1/3

− R̺

e

θ

e

|Ω

e

|

β

= 0, where we have used the fact that 2

43

π 

1/3

=

23

ec.

In view of (3.20)–(3.21) we see that assumption (3.7) is satisfied if (3.17) holds and δ

0

is sufficiently small.

Now, Lemmas 3.1, 3.2 and inequality (3.5) imply

(14)

Lemma 3.4 (see Lemma 5.4 of [19]). Let the assumptions of Lemmas 3.1–3.2 be satisfied. Moreover , assume

ϕ(0) ≤ α

1

, kH(·, 0) + 2/R

e

k

22,S1

≤ α.

Then for sufficiently small α

1

, α, δ, δ

3

(where δ is the constant from as- sumption (3.6) and δ

3

is the constant from Lemma 3.2) we have

ϕ(t) ≤ α

1

for t ≤ T.

Now, we formulate the main result of the paper.

Theorem 3.5. Let ν >

13

µ > 0, κ > 0, c

v

= ε

θ

> 0, c

v

∈ C

2

(R

2+

), ε ∈ C

1

(R

2+

), p ∈ C

3

(R

2+

), p

̺

> 0, p

θ

> 0, f = 0, θ ≥ 0. Suppose the assumptions of the local existence theorem (Theorem 4.2 of [12]) with r, θ ∈ C

B2,1

(R

3+

× [0, ∞)) are satisfied and the following compatibility conditions hold :

D

ξα

it

(Tn − σHn + p

0

n)|

t=0,S

= 0, |α| + i ≤ 2, D

ξα

ti

(n · ∇θ − θ)|

t=0,S

= 0, |α| + i ≤ 2.

Let (v, ϑ

0

, ̺

σ

) ∈ N(0) and

ϕ(0) ≤ α

1

, (3.22)

kv

0

k

24,Ω

≤ α

1

. (3.23)

Assume that

(3.24) l > 0 is a constant such that ̺

e

− l > 0, θ

e

− l > 0 and ε

1

< ε(̺, θ) < ε

2

for ̺ ∈ (̺

1

, ̺

2

), θ ∈ (θ

1

, θ

2

), where ̺

1

= ̺

e

− l, θ

1

= θ

e

− l, ̺

2

= ̺

e

+ l, θ

2

= θ

e

+ l,

(3.25) F (t) ≤ δ for t ≥ 0,

F occurs in inequality (3.5), and (3.26)

\

̺

0

dξ = M,

\

̺

0

ξ dξ = 0,

\

̺

0

v

0

(a + b × ξ) dξ = 0, for all constant vectors a, b.

Moreover , let the parameters ν

0

, µ

0

, β, ε

1

, ̺

2

, M satisfy one of the relations

(3.27)

i

ν

0

∈ I

i

and 0 < Φ

i

0

, ϕ

i

, p

0

, β, ε

1

, ̺

2

, M ) ≤ δ

0

,

i = 1, 2, 3, (where I

i

are defined by (2.8)–(2.10) and Φ

i

are defined by (2.14)–

(2.16)) and assume the following conditions:

(3.28) |Q

i

− |Ω

e

|| ≤ δ

1

(15)

and (3.29)

\

̺

0

v

02

2 dξ +

\

̺

0

(ε(̺

0

, θ

0

) − ε

1

) dξ

+ p

0

(|Ω

t

| − Q

i

+ δ

2

) + σ[|S| − ec(Q

i

− δ

2

)

2/3

] + κ sup

t t

\

0

dt

\

St′

θ(s, t

) ds ≤ δ

3

for ν

0

∈ I

i

, where δ

2

∈ (0, 1/2] is a constant so small that Q

i

− δ

2

> 0.

Moreover , assume that Ω is diffeomorphic to a ball and let S be described by |ξ| = e R(ω), ω ∈ S

1

(S

1

is the unit sphere ), where e R satisfies

(3.30) sup

S1

|∇ e R|

2

+ k e R − R

e

k

20,S1

≤ α

2

, and R

e

is the solution of equation (1.3).

Finally , assume that S ∈ W

24+1/2

and it is very close to a sphere, so (3.31) kH(·, 0) + 2/R

e

k

22,S1

≤ α

3

.

Then for sufficiently small constants α

i

(i = 1, 2, 3), δ

0

, δ

i

(i = 1, 2, 3) and δ there exists a global solution of problem (1.1) such that (v, ϑ

0

, ̺

σ

) ∈ M(t) for t ∈ R

+

, S

t

∈ W

24+1/2

for t ∈ R

+

and

(3.32) ϕ(t) ≤ α

1

, kH(·, t) + 2/R

e

k

22,S1

≤ α

3

for t ∈ R

+

.

P r o o f. Similarly to the case when σ = 0 (see [21]) the theorem is proved step by step using the local existence in a fixed time interval.

By Remark 3.2 of [19] the local solution of problem (1.1) satisfies the estimate

(3.33) kuk

24,ΩT

+ kη

σ

k

23,ΩT

+ η

σ 23,0,∞,ΩT

+ kγ

0

k

24,ΩT

≤ ϕ

1

(T, K

0

)(kv

0

k

23,Ω

+ k̺

σ0

k

23,Ω

+ kϑ

00

k

23,Ω

+ kkk

22,ΩT

+ kΓ k

23−1/2,ST

+ D

ξ,t2

Γ

21/2,ST

+ kk(0)k

21,Ω

+ kH(·, 0) + 2/R

e

k

22+1/2,S1

)

≤ ϕ

2

(T, K

0

)(α

1

+ δ),

where ̺

σ0

= ̺

0

− ̺

e

, ϑ

00

= θ

0

− θ

e

, u(ξ, t) = v(X

u

(ξ, t), t), η

σ

(ξ, t) =

̺

σ

(X

u

(ξ, t), t), γ

0

(ξ, t) = ϑ(X

u

(ξ, t), t), k(ξ, t) = r(X

u

(ξ, t), t), Γ (ξ, t) =

θ(X

u

(ξ, t), t); ϕ

1

and ϕ

2

are continuous increasing functions of their argu-

ments, K

0

is a constant such that K

0

> c(k̺

0

k

3,Ω

+ |̺

0

|

∞,Ω

+ |1/̺

0

|

∞,Ω

+

kv

0

k

3,Ω

+ kθ

0

k

3,Ω

+ ku

t

(0)k

1,Ω

+ kγ

0t

(0)k

1,Ω

), c > 0 is a constant.

Cytaty

Powiązane dokumenty

The equations of the laminar compressible boundary layer can be transformed (like in case of incompressible equations) such that similar solution are obtained i.e...

Distributions of Mach number (a) and static pressure (b) obtained for various angular velocities by segregated solver using Spalart-Allmaras turbulence model on the finest grid

In this paper we use some results of paper [6], which is concerned with the local existence of solutions of a free boundary problem for the equations of compressible barotropic

In the L p -framework global-in-time existence of solutions for equations of viscous compressible heat-conducting fluids with Dirichlet boundary conditions was proved by Str¨ ohmer

We consider the motion of a viscous compressible barotropic fluid in R 3 bounded by a free surface which is under constant exterior pressure, both with surface tension and without

The aim of this paper is to derive a global differential inequality for the following free boundary problem for a viscous compressible heat conducting fluid (see [2], Chs... Let

On the global existence theorem for a free boundary problem for equations of a viscous compressible heat conducting fluid.. by Ewa Zadrzy´ nska and

[r]