• Aucun résultat trouvé

Existence of global smooth solutions for Euler equations with symmetry (II)

N/A
N/A
Protected

Academic year: 2022

Partager "Existence of global smooth solutions for Euler equations with symmetry (II)"

Copied!
17
0
0

Texte intégral

(1)

Existence of global smooth solutions for Euler equations with symmetry (II)

Tong Yang

a;∗

, Changjiang Zhu

a;b

, Yongshu Zheng

c

aDepartment of Mathematics, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong

bWuhan Institute of Physics and Mathematics, Chinese Academy of Sciences, People’s Republic of China

cDepartment of Mathematics, Huaqiao University, Fujian, People’s Republic of China Received 12 January 1998; accepted 11 April 1998

Keywords:A priori estimates; Euler equations; Characteristic method; Spherical symmetry;

Global smooth solutions

1. Introduction

Consider the compressible Euler equations governing the gas ow surrounding a solid ball with mass M and frictional damping in n dimensions,

˜

t+∇ ·( ˜u) = 0;

( ˜u)t+∇ ·(u˜ u) +∇P( ˜) =−M˜ x

|x|n 2u;˜ (1.1)

where ˜;u; PandM are the density, velocity, pressure and mass of the gas, respectively, n3 is the dimension of x, and ¿0 is the frictional constant. In the following discussion, we assume the pressure satises the law and 1¡¡3, i.e. P=K2˜; K is a positive constant. We will study the existence and non-existence of global smooth solutions for the initial boundary problem of Eq. (1.1).

Firstly, we will show that regular solutions cannot be global if the initial density has compact support. This is a generalization of Theorem 2.1 in [7] to the case when there is an extra term due to the external force caused by the mass of the ball. In fact, the authors in [7] have noted that the Theorem 2.1 there is valid for any value of by studying an ordinary dierential equation ofH(t) as shown in Section 2 of this paper.

Corresponding author. Tel.: +852-2788-8646; fax: +852-2788-8561.

E-mail address:matyang@cityu.edu.hk (T. Yang)

0362-546X/00/$ - see front matter?2000 Elsevier Science Ltd. All rights reserved.

PII: S0362-546X(98)00273-9

(2)

For Euler equations without damping in multi-dimensional space, Sideris [12] gave sucient conditions for non-global existence of C1 solutions when inf0(x)¿0, where 0(x) is the initial density. The non-existence ofC1 solutions in [12] is related to the shock formation. In Section 2, we will study the singularity of the solutions at the vacuum states when the solutions contain no shocks. Thus, the time when the regular solutions blow up in our discussion is before the time when shock forms.

When inf0(x) = 0, Makino et al. [10] proved the non-global existence of regular solutions by assuming the initial data (0(x);u0(x)) to be of compact support, where u0(x) is the initial velocity. For the Euler–Poisson equations governing gaseous stars, Makino and Perthame [9] proved the non-global existence of tame solutions under the condition of spherical symmetry. Local existence of tame or regular solutions for these two systems was proved by Makino et al. [8, 10] by using Kato’s [3] theory for quasilinear symmetric hyperbolic system.

Secondly, we study the global existence of smooth solutions of the initial boundary value problem of Eq. (1.1) with spherical symmetry. Even though the global existence of regular solutions for Cauchy and initial boundary value problem of one-dimensional quasilinear hyperbolic systems has been extensively investigated (see [2,4–6,14]), much less is known for systems in high dimensions. To study the problem with spherical symmetry, where the system reduces to a one-dimensional system with singular source terms, is an initial stage to achieve this goal in high dimensions. The singularity is at r= 0 or ∞, where r is the radius. In Section 3, we study Eq. (1.1) with spherical symmetry and damping outside a core region of ball with mass M. Some sucient conditions are given for the global existence of smooth solutions when the frictional damping is suciently large. The study is a generalization of the one in [13] to the case when M6= 0, which is based on technical estimation of the C1-norm of the solutions.

2. No global existence

In this section, we consider the initial-boundary value problem of Eq. (1.1) with initial and boundary conditions given as follows:

˜

(x;0) = ˜0(x)0; u(x;0) =u0(x); when |x| ≥1;

u·x= 0; when |x|= 1: (2.1)

We consider the regular solution of Eqs. (1.1) and (2.1) dened as follows.

Deÿnition 2.1. A solution of Eqs. (1.1) and (2.1) is called a regular solution in [0; T)×, if

(i) ( ˜;u)C1([0; T)×); ˜0, (ii) ˜−1C1([0; T)×),

and

ut+u· ∇u=−u−Mx

|x|n

holds on the exterior of the support of ˜. Here =Rn\B1; B1={x: |x|¡1}.

(3)

Theorem 2.2. Let ( ˜(x; t);u(x; t)) be a regular solution of Eqs. (1.1) and (2.1) on 0t¡T. If the support of the initial data˜0(x) has compact support and ˜0(x)6≡0 and the mass M is suciently small; then T is ÿnite.

Remark 2.3.The following proof is similar to that of Makino et al. for the case when = 0; M= 0 [10] and Liu and Yang for the case ¿0; M= 0 [7]. We generalize their proof to the case when ¿0; M¿0.

Proof.Let (t) = supp ˜(x; t); S(t) =@(t). By Eq. (1.1) and the denition of regular solution, for any xS(t0), there exists x0S(0) and a curve x(t) connecting x0 and x such that

dx(t)

dt =u(x(t); t); x(t)S(t); 0tt0:

Let L=hx;xi; I=hu;ui, where h·;·i denotes inner product in Rn. Then we have dL

dt = 2hu;xi;

d2L dt2 = 2

du dt;x

+ 2hu;ui=dL dt 2M

|x|n−2 + 2I (2.2)

and dI dt = 2

du dt;u

=−2I−2M

u; x

|x|n

: (2.3)

It follows from Eqs. (2.2) and (2.3) that d2L

dt2 =dL dt 1

dI

dt 2M

|x|n−2 M

L−n=2dL

dt: (2.4)

Integrating Eq. (2.4), we have dL

dt ≤ −L−1

I+ 2M

(n2)L1−n=2+C0; (2.5)

where C0 is a constant. Thus dL

dt ≤ −L+ 2M

(n2)L1−n=2+C0: (2.6)

By letting L1=Ln=2, we have dL1

dt ≤ −n

2 L1+C0n

2 L1−2=n1 + Mn (n2);

which implies that there exists a constant L1 depending only on the initial data, such that

1L1L1: (2.7)

(4)

Now as in [10], let H(t) =1

2 Z

(t)(x; t)|x|˜ 2dx: (2.8)

From Eq. (1.1), we have H0(t) =1

2 Z

(t)·xdx and

H00(t) = Z

(t)( ˜|u|2+nP) dxH0(t) + Z

D0

PdSM Z

(t)

˜

|x|n−2dx; (2.9) where D0={x: |x|= 1}.

Let m=R

(t)(x; t) dx˜ be the total mass. Then we have from the Holder inequality that

m= Z

(t)(x; t) dx˜ Z

(t)˜dx 1= Z

(t)dx 1=0

;

where 1=+ 1=0= 1. Thus there exists a constant V¿0 such that Z

(t)˜dx 1=

m(V)−1=0: Thus,

H00(t) +H0(t)(nK2m−1(V)−=0M)m:

When M is suciently small, there exists a constant ¿0, such that H00(t) +H0(t):

Integrating the above inequality yields H(t)H(0) +

t+1

e−t1

H0(0)

(e−t1):

Thus,

H(t)→ ∞ as t→ ∞;

which is a contradiction to Eqs. (2.7) and (2.8). Hence, T must be nite and this completes the proof of Theorem 2.2.

(5)

3. Global existence with spherical symmetry

Now, we are going to study the global existence of smooth solutions for Eq. (1.1) with spherical symmetry outside a core region. That is, we look for solutions of the form

˜

= ˜(|x|; t); u= x

|x|·u(|x|; t):

By denoting r=|x|, Eq. (1.1) becomes

˜ t+ 1

rn−1(rn−1u)˜ r= 0;

˜

(ut+uur) +Pr=

rn−1 2u;˜ (3.1)

where P( ˜) =K2˜, 1¡¡3. We consider system (3:1) with the following initial and boundary conditions:

˜

(r;0) = ˜0(r); u(r;0) =u0(r); u(1; t) = 0: (3.2) Let =rn−1, then we can rewrite Eq. (3.1) as˜

t+ (u)r= 0;

ut+uur+ K2−2

r(n−1)(−1)r= M

rn−1 2u+K2(n1)−1

rnr(n−1)(−2) : (3.3)

In the following discussion we use the Lagrangian coordinates as follows:

=t; = Z r

1 (r; t) dr: (3.4)

Then ¿0 as long as ¿0 forr¿1, and Eq. (3.3) becomes +2u= 0;

u+ K2−1

r(n−1)(−1)= M

rn−1 2u+K2(n1)−1

rnr(n−1)(−2) : (3.5)

Now we let =−1, and replace (; ) by (x; t). Eq. (3.5) can be rewritten as tux= 0;

utK2−−1

r(n−1)(−1)x= M

rn−1 2u+K2(n1)1−

rnr(n−1)(−2)

(3.6) with the following initial and boundary conditions:

(x;0) =0(x); u(x;0) =u0(x); u(0; t) = 0:

By transformation (3:4), we have r= 1 +

Z x

0 (s; t) ds; x0: (3.7)

(6)

To symmetrize system (3:6), we use the following Riemann invariants:

w=u+2K

1r−(1=2)(n−1)(−1)−(−1)=2; z=u2K

1r−(1=2)(n−1)(−1)−(−1)=2:

(3.8)

By using rt=u and rx=, Eq. (3.6) becomes wt+wx= M

rn−1 (w+z)(n1)(1)

8r (wz)(w+z);

zt+zx= M

rn−1 (w+z) +(n1)(1)

8r (wz)(w+z)

(3.9)

with the corresponding initial and boundary conditions given by w(x;0) =w0(x); z(x;0) =z0(x); w(0; t) +z(0; t) = 0:

Here ==KArn−1(wz)(+1)=(−1) are the two characteristic speeds and A=

1 4K

(+1)=(−1) :

For later use, we list some equations as follows:

dr dt=1

2(w+z) +1

4 (wz);

dr dt=1

2(w+z)−−1

4 (wz);

d(wz)

dt =2K

Arn−1(wz)(+1)=(−1)zx

(n1)(1)

4r (wz)(w+z);

d(wz)

dt =2KArn−1(wz)(+1)=(−1)wx

(n1)(1)

4r (wz)(w+z);

d(w+z)

dt =2M

rn−1 2(w+z) + 2KArn−1(wz)(+1)=(−1)zx; d(w+z)

dt = 2M

rn−1 2(w+z)2KArn−1(wz)(+1)=(−1)wx; (3.10) where we have used the notation d=dt=@=@t+(@=@x) and d=dt=@=@t+(@=@x):

(7)

From now on, we assume the initial and boundary conditions satisfy w0(x); z0(x)Cb1[0;∞);

(r0(x))n=2−1(w0(x)z0(x)) (3.11)

and

w0(0) +z0(0) = 0;

0(0)w0x(0) +0(0)z0x(0) =−2M; (3.12)

where is a positive constant, and 0(x) and 0(x) are the given characteristic speeds at t= 0.

Remark 3.1. There exists a special solution of Eq. (3.9) satisfying Eqs. (3.11) and (3.12), i.e.

w(x; t) = z(x; t) =M1(r(x; t))1−n=2; where M1= 2

M=p

(n2)(1).

Now, we are ready to prove the following lemma on the C0-norm of the solutions inside the smoothness region of the solutions.

Lemma 3.2. Under the conditions of Eqs. (3.11) and (3.12),if

|(r0(x))n=2−1w0(x)|;|(r0(x))n=2−1z0(x)| ≤M1; (3.13) then there exists a positive constant 0 depending only on M1 and such that when

¿0, we have

|(r(x; t))n=2−1w(x; t)|;|(r(x; t))n=2−1z(x; t)| ≤M1; (3.14)

w(x; t)z(x; t)¿0; (3.15)

where (x; t) is inside the smoothness region of the solutions and M1 is the constant deÿned in the above Remark 3:1:

Proof. We now prove Eq. (3.15) rst. By Eq. (3.10), we have d(wz)

dt =−D(x; t)(w−z); (3.16) where

D(x; t) = 2KArn−1(wz)2=(−1)wx+(n1)(1)

4r (w+z):

Integrating Eq. (3.16) along the -characteristic curve, we have (wz)(x; t)(r0(x))1−n=2exp

Z t

0 D(x(); ) d

¿0;

as long as the solution is smooth.

(8)

To prove Eq. (3.14), we introduce the following transformations:

w(x; t) = w(x; t) +M1(r(x; t))1−n=2;

z(x; t) = z(x; t)M1(r(x; t))1−n=2: (3.17) Then Eq. (3.9) becomes

wt+wx=−A(x; t)( w+ z) +br−n=2w;

zt+zx=−B(x; t)( w+ z)br−n=2z; (3.18) where

A(x; t) =+(n1)(1) 8r ( wz)

+(3)(n2)2(n1)(1)

8 M1r−n=2;

B(x; t) =(n1)(1) 8r ( wz)

(3)(n2)2(n1)(1)

8 M1r−n=2;

b=12p

(n2)(1)M: (3.19)

Let w=

W +N

L(x+Cet)

ebt;

z=

Z+N

L(x+Cet)

ebt;

(3.20)

where C and L are two positive constants to be determined later. A smooth solution exists at least locally under conditions (3:11) and (3:12), cf. [1]. We use N to denote the bound for |w(x; t)| and|z(x; t)| of the local solution. Using Eq. (3.20), the system (3:18) for ( w(x; t);z(x; t)) can be written as follows:

Wt+Wx+N

L(Cet+) +N

L(x+Cet)b 1r−n=2

=−A(x; t)(W Z)bW +br−n=2W;

Zt+Zx+N

L(Cet+) +N

L(x+Cet)b(1 +r−n=2)

=B(x; t)(W Z)bZbr−n=2Z: (3.21)

(9)

If we consider system (3:21) in the region [0; L]×R+, then the initial and boundary conditions have the following properties:

W(x;0) = w(x;0)N

L(x+C)¡0;

Z(x;0) =z(x;0)N

L(x+C)¡0;

W(L; t) = w(L; t)e−btNN LCet¡0;

Z(L; t) =z(L; t)e−btN N

LCet¡0:

(3.22)

From Eqs. (3.21) and (3.22), we claim that

W(x; t)¡0; Z(x; t)¡0; (x; t)[0; L]×[0; T]: (3.23) If not, we let t= supt{t: W(x; )¡0; Z(x; )¡0; ∀x∈[0; L]; (0; t)}, then 0¡tT

¡+∞. By the continuity ofW(x; t) andZ(x; t), there exists (x; t) with 0 x¡L, such that one of the following cases holds:

Case 1: When x(0; L),

W(x;t) = 0; Z(x;t)0; @W

@t

(x;t)0; @W

@x

(x;t) = 0;

or

Z(x;t) = 0; W(x;t)0; @Z

@t

(x;t)0; @Z

@x

(x;t) = 0:

Case 2: When x= 0, W(x;t) =Z(x;t) = 0; @Z

@t

(x;t)0; @Z

@x

(x;t) 0:

For the above two cases, if we assume the following a priori estimate:

|(r(x; t))n=2−1w(x; t)|; |(r(x; t))n=2−1z(x; t)| ≤M1; (3.24) then for suciently large , we have

A(x; t)¿0; B(x; t)¿0:

By applying the “maximum principle”, cf. [13, 14], when C¿sup for all w; z under consideration, we have a contradiction. Therefore, Eq. (3.23) holds.

Equality (3:20) and inequality (3:23) imply that

w(x; t)¡

N

L(x+Cet)

ebt;

z(x; t)¿−

N

L(x+Cet)

ebt:

(10)

Since L can be arbitrary, letting L→ ∞ yields

w(x; t)0; z(x; t) 0:

Hence by Eq. (3.17), we have

w(x; t)M1(r(x; t))1−n=2; z(x; t)≥ −M1(r(x; t))1−n=2: (3.25) It is easy to get Eq. (3.14) from Eqs. (3.15) and (3.25). This completes the Proof of Lemma 3.2.

Now, we estimate the derivatives of w(x; t); z(x; t) with respect tox.

Let P(x; t) =wx(x; t); Q(x; t) =zx(x; t). By dierentiating Eq. (3.9) with respect tox, we have the following system for P; Q:

Pt+Px=−(P+Q)KA+ 1

1rn−1(wz)2=(−1)(PQ)P

(n1)(1)

4r (wz)P(n1)(1)

8r (wz)(P+Q)

(n1)(1)

8r (w+z)(PQ) +(n1)M

r2n−1 B(wz)−2=(−1) +(n1)(1)

8rn+1 B(wz)−(3−)=(−1)(w+z);

Qt+Qx=−(P+Q)K A+ 1

1rn−1(wz)2=(−1)(QP)Q +(n1)(1)

4r (wz)Q+(n1)(1)

8r (wz)(P+Q) +(n1)(1)

8r (w+z)(PQ) +(n1)M

r2n−1 B(wz)−2=(−1)

(n1)(1)

8rn+1 B(wz)−(3−)=(−1)(w+z); (3.26) where

B= 1

4K

−2=(−1) : Let

P= (wz)lrP; Q= (wz)lrQ; (3.27)

(11)

wherel=−(+ 1)=2(1)¡0 andis a constant to be determined later. Then system (3:26) becomes

Pt+Px=−(P+Q)KA+ 1

1rn+−1(wz)(3−)=2(−1)P2

(3n+ 23)(1)

8r (wz)P(n1)(1)

8r (wz)Q

(n1)+ 2

4r (w+z)P+(n1)(1)

8r (w+z)Q +(n1)M

r2n+−1 B(wz)−(3−)=2(−1)

+(n1)(1)

8rn++1 B(wz)(3−5)=2(−1)(w+z);

Qt+Qx=−(P+Q)KA+ 1

1rn+−1(wz)(3−)=2(−1)Q2 +(n1)(1)

8r (wz)P+(3n+ 23)(1)

8r (wz)Q

+(n1)(1)

8r (w+z)P(n1)+ 2

4r (w+z)Q +(n1)M

r2n+−1 B(wz)−(3−)=2(−1)

(n1)(1)

8rn++1 B(wz)(3−5)=2(−1)(w+z): (3.28)

Let

P=F+a1r−−n+1(wz)m+a2r−−n(wz)m+1 +a3r−−n(wz)m(w+z)

Q=G+a1r−−n+1(wz)ma2r−−n(wz)m+1

+a3r−−n(wz)m(w+z); (3.29)

where m=−(3−)=2(1)¡0. Then we have Ft+Fx

=−K√A+ 1

1rn+−1(wz)(3−)=2(−1)P2P

(3n+ 23)(1)

8r (wz)P

(n1)+ 2

4r (w+z)P+ (2KAma1)Q

(12)

+

2KA(m+ 1)a22KAa3(n1)(1) 8

1

r(wz)Q +

2KAma3+(n1)(1) 8

1

r(w+z)Q +(n1)(1)

4 ma1r−−n(wz)−(3−)=2(−1)(w+z) +

1

2(+n1)a1+ 2a3

r−−n(wz)−(3−)=2(−1)(w+z) +1

4 (+n1)a1r−−n(wz)(3−5)=2(−1)

+1

4 (+n)a2r−−n−1(wz)(5−7)=2(−1)

+(n1)(1)

4 (m+ 1)a2r−−n−1(wz)(3−5)=2(−1)(w+z) +

1

2(+n)a2+1

4 (+n)a3

r−−n−1(wz)(3−5)=2(−1)(w+z) + ((n1)B+ 2a3)Mr−−2n+1(wz)−(3−)=2(−1)

+(n1)(1)

8 Br−−n−1(wz)(3−5)=2(−1)(w+z) +

(n1)(1)

4 m+1

2(+n)

a3r−−n−1(wz)−(3−)=2(−1)(w+z)2;

Gt+Gx

=−K√A+ 1

1rn+−1(wz)(3−)=2(−1)Q2Q +(3n+ 23)(1)

8r (wz)Q

(n1)+ 2

4r (w+z)Q+ (2KAma1)P

2KA(m+ 1)a22KAa3(n1)(1) 8

1

r(wz)P +

2KAma3+(n1)(1) 8

1

r(w+z)P +(n1)(1)

4 ma1r−−n(wz)−(3−)=2(−1)(w+z)

(13)

+ 1

2(+n1)a1+ 2a3

r−−n(wz)−(3−)=2(−1)(w+z)

−−1

4 (+n1)a1r−−n(wz)(3−5)=2(−1)

+1

4 (+n)a2r−−n−1(wz)(5−7)=2(−1)

(n1)(1)

4 (m+ 1)a2r−−n−1(wz)(3−5)=2(−1)(w+z)

1

2(+n)a2+1

4 (+n)a3

r−−n−1(wz)(3−5)=2(−1)(w+z) + ((n1)B+ 2a3)Mr−−2n+1(wz)−(3−)=2(−1)

(n1)(1)

8 Br−−n−1(wz)(3−5)2(−1)(w+z) +

(n1)(1)

4 m+1

2(+n)

a3r−−n−1(wz)−(3−)=2(−1)(w+z)2: (3.30) Now choose ai; i= 1;2;3; as follows:

2KAma1= 0;

2KA(m+ 1)a22KAa3(n1)(1)

8 = 0;

2KAma3+(n1)(1)

8 = 0;

i.e.

a1= (1) KA(3)¡0;

a2= (n1)(1)2(+ 1) 8KA(35)(3); a3=(n1)(1)2

8KA(3)¿0;

(3.31)

where 1¡¡53 or 53¡¡3. Then the system for F andG becomes Ft+Fx=−K√+ 1

1Arn+−1(wz)(3−)=2(−1)F2 +

31 3+1

F

22(1)2 KA(3)2 +1

r−−n+1(wz)−(3−)=2(−1);

(14)

Gt+Gx=−K√+ 1

1Arn+−1(wz)(3−)=2(−1)G2 +

31 3+2

G

22(1)2 KA(3)2 +2

r−−n+1(wz)−(3−)=2(−1); (3.32) where i; i are continuous and bounded functions of the corresponding variables, i= 1;2.

Based on Lemma 3.2 and system (3:32), we are now ready to prove the following theorem.

Theorem 3.3. Under the conditions of Lemma 3.2, if 1¡¡53 or 53¡¡3, and is suciently large, then there exists a constant M2 depending only on M1 and ,such that if

|M1|1;

F(x;0); G(x;0)M2;

then there exists a global smooth solution for system (3:1)and (3:2).

Proof. Firstly, we assume a priori estimate

(r(x; t))n=2−1(w(x; t)z(x; t))1¿0; (3.33) where 1 is a positive constant depending only on M1; M2 and . By Lemma 3.2, we have

|i|;|i|¡M3; i= 1;2;

where M3 is a positive constant depending only on M1; M2 and . When is large enough and 1¡¡3, we know that the equation

0 =−K√+ 1

1Arn+−1(wz)(3−)=2(−1)y2+

31 3+i

y

22(1)2 KA(3)2 +i

r−−n+1(wz)−(3−)=2(−1); has two positive roots given by

y1i =

2(1)2

KA(+ 1)(3)+1i1=2

r−n−+1(wz)−(3−)=2(−1);

y2i =

1

KA(3)+2i1=2

r−n−+1(wz)−(3−)=2(−1); i= 1;2;

(15)

with y1i¡y2i; i= 1;2, where |1i| and |2i| have an upper bound depending only on M1; M2 and . Therefore, when 1¡¡53 or 53¡¡3, if we choose

=(n2)(3)

4(1) n+ 1;

then supy1i¡infy2i, provided1−M1 is suciently small. Thus, there exists a constant M2= max{supy11;supy12} depending only on M1 and , such that if

F(x;0); G(x;0)M2; we have

M2F(x; t); G(x; t)sup{F(x;0); G(x;0)}: (3.34) Combining Lemma 3.2 and Eq. (3.34) yields the global existence of a smooth solution for Eqs. (3.1) and (3.2).

Now, it remains to verify the a priori estimate (3:33). To do this, let

H(x; t) = (r(x; t))n=2−1(w(x; t)z(x; t)): (3.35) Then, from Eqs. (3.10), (3.29) and (3.34), we have

1

2KAH−(3−)=2(−1)−1 d dtH+

(1) KA(3) +3

H−(3−)=2(−1)M4; (3.36) where 3Cb1 and M4 is a positive constant depending only on M1; M2 and .

From Eq. (3.36), we have d

dt

H−(3−)=2(−1)exp Z t

0(+4()) d

M5exp Z t

0(+4()) d

; (3.37)

where

4() =KA(3) 1 3();

and M5 is a positive constant depending only on M1; M2 and . When is suciently large, we have

+4()¿

2:

(16)

Hence, Z t

0 exp Z s

0 (+4()) d

ds

2

Z t

0(+4(s)) exp Z s

0 (+4()) d

ds

=2

exp

Z t

0(+4()) d

1

: (3.38)

Integrating Eq. (3.37) and using Eq. (3.38), we have H−(3−)=2(−1)M6;

where M6 is a positive constant depending only on M1; M2 and . Therefore, when 1¡¡53 or 53¡¡3, we have

(r(x; t))n=2−1(w(x; t)z(x; t))1¿0:

This completes the proof of Theorem 3.3.

Remark 3.4. When =53, the conclusion of Theorem 3.3 still holds.

Acknowledgements

The research of the rst author was supported in part by the RGC Competitive Earmarked Research Grant # 9040190. The research of the second author was supported in part by the grant of the National Youth Science Foundation of China #19301038.

References

[1] A. Douglis, Existence theorems for hyperbolic systems, Comm. Pure Appl. Math. 5 (1952) 119–154.

[2] F. John, Formation of singularities in one-dimensional nonlinear wave propagation, Comm. Pure Appl.

Math. 27 (1974) 377–405.

[3] T. Kato, The Cauchy problem for quasi-linear symmetric hyperbolic systems, Arch. Rat. Mech. Anal.

58 (1945) 181–205.

[4] P.D. Lax, Development of singularities of solutions of nonlinear hyperbolic partial dierential equations, J. Math. Phys. 5 (1964) 611–613.

[5] Lin Longwei, Yang Tong, Existence and nonexistence of global smooth solutions for dampedp-system with “really large” initial data, J. Part. Di. Eqs. 4 (1991) 45–51.

[6] T.-P. Liu, Development of singularities in the nonlinear waves for quasilinear hyperbolic partial dierential equations, J. Di. Eqs. 33 (1979) 92–111.

[7] T.-P. Liu, T. Yang, Compressible Euler equations with vacuum, J. Di. Eqs. 137 (1997).

[8] T. Makino, Blowing up solutions of the Euler–Poisson equation for the evolution of gaseous stars, Transport Theory Statist. Phys. 21 (1992) 615–624.

[9] T. Makino, B. Perthame, Sur les solutions a symmetric spherique de l’equation d’Euler–Poisson pour l’evolution d’etoiles gazeuses, Japan J. Appl. Math. 7 (1990) 165–170.

(17)

[10] T. Makino, S. Ukai, S. Kawashima, Sur la solution a support compact de l’equation d’Euler compressible, Japan J. Appl. Math. 3 (1986) 249–257.

[11] T. Nishida, Nonlinear hyperbolic equations and related topics in uid dynamics, Publications Mathematiques D’ocsay 78.02, Department de Mathematique, Paris-Sud, 1978.

[12] T. Sideris, Formation of singularities in three-dimensional compressible uids, Comm. Math. Phys. 101 (1985) 475–485.

[13] L.A. Ying, T. Yang, C.J. Zhu, Existence of global smooth solutions for Euler equations with symmetry, Comm. Partial Di. Eqs. 22 (1997) 1361–1387.

[14] C.J. Zhu, Global resolvability for a viscoelastic model with relaxation, Proc. Roy. Soc. Edinburgh 125A (1995) 1277–1285.

Références

Documents relatifs

Keywords: Two-fluid flows; Euler–Maxwell equations; Partially dissipative hyperbolic systems; Global smooth solutions; Long-time behavior;..

V´eron, Nonlinear second order elliptic equations involving measures, Series in Nonlinear Analysis and Applications 21 , De Gruyter, Berlin/Boston (2013).

However the strategy used in [4] applies: this is the strategy one uses to construct Leray solutions in the context of the Navier-Stokes equations, which consists in first

Since the water waves equations are a quasi-linear system, usual normal forms approaches would face the well known problem of losses of derivatives in the unbounded transformations.

Abstract This short note studies the problem of a global expansion of local results on existence of strong and classical solutions of Navier-Stokes equations in R 3..

This chemical reaction is typical of general reversible reactions and contains the major difficulties encountered in a large class of similar problems as regards global existence

Afrouzi and Hadjian [1], established results of existence and uniqueness of classic infinite solutions for Caputo fractional differential equations with limit value movement, using

Since the non-linearity in the right-hand side is a singular ab- sorption term near the boundary, a non-trivial weak solution may not be positive everywhere in Ω and compact