• Aucun résultat trouvé

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

N/A
N/A
Protected

Academic year: 2022

Partager "Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php"

Copied!
28
0
0

Texte intégral

(1)

GLOBAL WELL-POSEDNESS OF THE ONE-DIMENSIONAL COMPRESSIBLE NAVIER–STOKES EQUATIONS WITH CONSTANT

HEAT CONDUCTIVITY AND NONNEGATIVE DENSITY

JINKAI LI

Abstract. In this paper we consider the initial-boundary value problem to the one-dimensional compressible Navier–Stokes equations for ideal gases. Both the viscous and heat conductive coeffi- cients are assumed to be positive constants, and the initial density is allowed to have vacuum. Global existence and uniqueness of strong solutions is established for anyH2initial data, which generalizes the well-known result of Kazhikhov and Shelukhin [J. Appl. Math. Mech.,41 (1977), pp. 273–282]

to the case that with nonnegative initial density. An observation to overcome the difficulty caused by the lack of the positive lower bound of the density is that the ratio of the density to its initial value is inversely proportional to the time integral of the upper bound of the temperature along the trajectory.

Key words. compressible Navier–Stokes equations, global well-posedness, heat conductivity AMS subject classifications. 35A01, 35B45, 76N10, 76N17

DOI. 10.1137/18M1167905

1. Introduction.

1.1. The compressible Navier–Stokes equations. In this paper, we consider the following one-dimensional heat conductive compressible Navier–Stokes equations:

tρ+∂x(ρu) = 0, (1.1)

ρ(∂tu+u∂xu)−µ∂x2u+∂xp= 0, (1.2)

cvρ(∂tθ+u∂xθ) +∂xup−κ∂x2θ=µ(∂xu)2, (1.3)

where ρ, u, θ, andp, respectively, denote the density, velocity, absolute temperature, and pressure. The viscous coefficientµand heat conductive coefficientκare assumed to be positive constants. The state equation for the ideal gas reads as

p=Rρθ, whereR is a positive constant.

The compressible Navier–Stokes equations have been extensively studied. In the absence of vacuum, i.e., the case that the initial density is uniformly bounded away from zero, global well-posedness of strong solutions to the one dimensional com- pressible Navier–Stokes equations has been well-known since the pioneer works by Kazhikhov and Shelukhin [19] and Kazhikhov [18]. Inspired by these works, global ex- istence and uniqueness of weak solutions were later established by Zlotnik and Amosov [40, 41] and Chen, Hoff, and Trivisa [2] for the initial boundary value problems, and

Received by the editors January 30, 2018; accepted for publication (in revised form) July 15, 2019; published electronically September 5, 2019.

https://doi.org/10.1137/18M1167905

Funding:The work of the author was partially supported by the South China Normal University start-up grant 550-8S0315, by the Hong Kong RGC grant CUHK 14302917, and by the National Natural Science Foundation of China grants 11971009, 11771156, and 11871005.

South China Research Center for Applied Mathematics and Interdisciplinary Studies, South China Normal University, Guangzhou 510631, China (jklimath@m.scnu.edu.cn, jklimath@

gmail.com).

3666

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(2)

by Jiang and Zlotnik [17] for the Cauchy problem. Large time behavior of solutions to the one dimensional compressible Navier–Stokes equations with large initial data was recently proved by Li and Liang [22]. The corresponding global well-posedness results for the multidimensional case were established only for small perturbed initial data around some nonvacuum equilibrium or for spherically symmetric large initial data;

see, e.g., Matsumura and Nishida [28, 29, 30, 31], Ponce [33], Valli and Zajaczkowski [35], Deckelnick [8], Jiang [15], Hoff [12], Kobayashi and Shibata [20], Danchin [7], Chikami and Danchin [3], and the references therein.

In the presence of vacuum, that is, the density may vanish on some set or tends to zero at the far field, global existence of weak solutions to the isentropic compressible Navier–Stokes equations was first proved by Lions [26, 27], with adiabatic constant γ ≥ 95, and later generalized by Feireisl, Novotn´y, and Petzeltov´a [9] toγ > 32, and further by Jiang and Zhang [16] to γ > 1 for the axisymmetric solutions. For the full compressible Navier–Stokes equations, global existence of the variational weak solutions was proved by Feireisl [10, 11], which however is not applicable for the ideal gases. Local well-posedness of strong solutions to the full compressible Navier–Stokes equations, in the presence of vacuum, was proved by Cho and Kim [6]; see also Salvi and Straˇskraba [34], Cho, Choe, and Kim [4], and Cho and Kim [5] for the isentropic case. The solutions in [34, 4, 5, 6] are established in the homogeneous Sobolev spaces, and, generally, one can not expect the solutions in the inhomogeneous Sobolev spaces, if the initial density has compact support, due to the recent nonexistence result by Li, Wang, and Xin [21]. Global existence of strong solutions to the compressible Navier–Stokes equations, with small initial data, in the presence of initial vacuum, was first proved by Huang, Li, and Xin [14] for the isentropic case (see also Li and Xin [25] for further developments), and later by Huang and Li [13] and Wen and Zhu [37] for the full case. Due to the finite blow-up results in [38, 39], the global solutions obtained in [13, 37] must have infinite entropy somewhere in the vacuum region, if the initial density has an isolated mass group; however, if the initial density is positive everywhere but tends to vacuum at the far field, one can expect the global existence of solutions with uniformly bounded entropy to the full compressible Navier–Stokes equations; see the recent work by the author and Xin [23, 24].

An interesting problem is to show the positivity of the temperature to the full compressible Navier–Stokes equations. Along this direction, some results have been proved by Mellet and Vasseur [32] and Baer and Vasseur [1], where under some appro- priate constitutive assumptions on the internal energy, the pressure law, the viscous coefficients, and the heat conductive coefficient, they have shown that the temperature has a uniform positive lower bound, with the bound depending on the time, at each time after the initial time, for the initial-boundary value problem, under the nonheat flux boundary condition. However, due to the constitutive assumptions in [32, 1], the results there do not apply to the ideal gas with constant viscous and heat con- ductive coefficients. Besides, the positivity of the temperature inside the space time domain to the system considered in [32, 1] is also not clear if replacing the nonheat flux temperature boundary condition with the zero temperature boundary condition.

Note that in the global well-posedness results for system (1.1)–(1.3) in [19, 18, 40, 41, 2, 17, 22], the density was assumed to be uniformly away from vacuum. Global well-posedness of strong solutions to system (1.1)–(1.3) in the presence of vacuum was proved by Wen and Zhu [36]; however, due to the following assumption,

κ0(1 +θq)≤κ(θ)≤κ1(1 +θq) for someq >0,

made on the heat conductive coefficientκin [36], the case thatκ(θ)≡const.was not included there. The aim of this paper is to study the global well-posedness of strong

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(3)

solutions to system (1.1)–(1.3), with both constant viscosity and constant diffusivity, in the presence of vacuum. This result will be proven in the Lagrangian flow map coordinate; however, it can be equivalently translated back to the corresponding one in the Euler coordinate.

1.2. The Lagrangian coordinates and main result. Letybe the Lagrangian coordinate, and define the coordinate transform between the Lagrangian coordinate y and the Euler coordinatexas

x=η(y, t), whereη(y, t) is the flow map determined byu, that is,

tη(y, t) =u(η(y, t), t), η(y,0) =y.

Denote by%, v, ϑ, andπthe density, velocity, temperature, and pressure, respec- tively, in the Lagrangian coordinate; that is, we define

%(y, t) :=ρ(η(y, t), t), v(y, t) :=u(η(y, t), t), ϑ(y, t) :=θ(η(y, t), t), π(y, t) :=p(η(y, t), t).

Recalling the definition ofη(y, t), by straightforward calculations, one can check that (∂xu, ∂xθ, ∂xp) =

yv

yη,∂yϑ

yη,∂yπ

yη

, (∂x2u, ∂x2ϑ) = 1

yη∂y

yv

yη

, 1

yη∂y

yϑ

yη

,

tρ+u∂xρ=∂t%, ∂tu+u∂xu=∂tv, ∂tθ+u∂xθ=∂tϑ.

Define a functionJ =J(y, t) as

J(y, t) =ηy(y, t);

then it follows

(1.4) ∂tJ =∂yv,

and system (1.1)–(1.3) can be rewritten in the Lagrangian coordinate as

t%+∂yv J %= 0, (1.5)

%∂tv−µ J∂y

yv J

+∂yπ

J = 0, (1.6)

cv%∂tϑ+∂yv J π−κ

J∂y

yϑ J

=µ ∂yv

J 2

, (1.7)

whereπ=R%ϑ.

Due to (1.4) and (1.5), it holds that

t(J %) =∂tJ %+J ∂t%=∂yv%−J∂yv J %= 0, from which, by setting%|t=0=%0 and noticing thatJ|t=0= 1, we have

J %=%0.

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(4)

Therefore, one can replace (1.5) with (1.4), by setting%= %J0, and rewrite (1.6) and (1.7), respectively, as

%0tv−µ∂y

yv J

+∂yπ= 0 and

cv%0tϑ+∂yvπ−κ∂y

yϑ J

=µ(∂yv)2 J . In summary, we only need to consider the following system:

tJ =∂yv, (1.8)

%0tv−µ∂y

yv J

+∂yπ= 0, (1.9)

cv%0tϑ+∂yvπ−κ∂y

yϑ J

=µ(∂yv)2 J , (1.10)

where

π=R%0

Jϑ.

We consider the initial-boundary value problem to system (1.8)–(1.10) on the interval (0, L), with L >0; that is, system (1.8)–(1.10) is defined in the space-time domain (0, L)×(0,∞). We complement the system with the following boundary and initial conditions:

(1.11) v(0, t) =v(L, t) =∂yϑ(0, t) =∂yϑ(L, t) = 0 and

(1.12) (J, v, ϑ)|t=0= (1, v0, ϑ0).

For 1 ≤ q ≤ ∞ and positive integer m, we use Lq = Lq((0, L)) and Wm,q = Wm,q((0, L)) to denote the standard Lebesgue and Sobolev spaces, respectively, and in the case thatq = 2, we use Hm instead of Wm,2. We always use kukq to denote theLq norm of u.

The main result of this paper is the following theorem.

Theorem 1.1. Given(%0, v0, ϑ0)∈H2((0, L)), satisfying

%0(y)≥0, ϑ(y)≥0 ∀y∈[0, L], v0(0) =v0(L) =∂yϑ0(0) =∂yϑ0(L) = 0.

Assume that the following compatibility conditions hold:

µv000−R(%0ϑ0)0 =√

%0g0, κϑ000+µ(v00)2−Rv00%0ϑ0=√

%0h0

for two functions g0, h0∈L2((0, L)).

Then, there is a unique global solution (J, v, ϑ) to system (1.8)–(1.10), subject to (1.11)–(1.12), satisfyingJ >0 andϑ≥0on [0, L]×[0,∞), and

J ∈C([0, T];H2), ∂tJ ∈L2(0, T;H2),

v∈C([0, T];H2)∩L2(0, T;H3), ∂tv∈L2(0, T;H1), ϑ∈C([0, T];H2)∩L2(0, T;H3), ∂tϑ∈L2(0, T;H1) for any T∈(0,∞).

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(5)

Remark 1.1. The same result as in Theorem 1.1 still holds if replacing the bound- ary condition∂yϑ(0, t) =∂yϑ(L, t) = 0 by one of the following three,

ϑ(0, t) =ϑ(L, t) = 0, ϑ(0, t) =∂yϑ(L, t) = 0,

yϑ(0, t) =ϑ(L, t) = 0,

and the proof is exactly the same as the one presented in this paper; the only difference is that the basic energy identity in Proposition 2.2 will then be an inequality.

Remark 1.2. The argument presented in this paper also works for the free bound- ary value problem to the same system. Because, if rewritten the system in the Lagrangian coordinates, the only difference between the initial boundary value prob- lem and the free boundary value problem is the boundary conditions forv: in the free boundary problem, the boundary conditions forv in (1.11) are replaced by

µ∂yv J −π

y=0,L= 0.

Note that all the energy estimates obtained in this paper hold if replacing the bound- ary condition onvin (1.11) with the above ones, by slightly modifying the proof.

The argument used in Kazhikhov and Shelukhin [19], in which the nonvacuum case was considered, does not apply directly to the vacuum case. One main obser- vation in [19] is the lower bound of the density is inversely proportional to the time integral of the upper bound of the temperature along the trajectory. Note that this only holds for the case that the density has a positive uniform lower bound. To over- come the difficulty caused by the lack of the positive lower bound of the density (it is of this case if the lower bound of the initial density is zero), our observation is the ratio of the density to its initial value is inversely proportional to the time integral of the temperature, along the trajectory, or, equivalently, the upper bound of J is proportional to the time integral of the upper bound of the temperature. This obser- vation holds for both the vacuum and nonvacuum cases, which, in particular, reduces to the one in [19] for the nonvacuum case; this also indicates the advantage of taking J rather than%as one of the unknowns.

The key issue of proving Theorem 1.1 is to establish the appropriate a priori energy estimates, up to any finite time, of the solutions to system (1.8)–(1.10), subject to (1.11)–(1.12). There are four main stages for carrying out the desired a priori energy estimates. In the first stage, we derive from (1.8)–(1.9) an identity

1 +R µ%0(y)

Z t

0

ϑ(y, τ)H(τ)B(y, τ)dτ =J(y, t)H(t)B(y, t)

for some functions H(t) andB(y, t). The temperature equation is not used at all in deriving the above identity, and this identity is in the spirit of the one in [19], but in different Lagrangian coordinates. The basic energy estimate implies that both H and B are uniformly away from zero and uniformly bounded, up to any finite time.

As a direct corollary of the above identity, one can obtain the uniform positive lower bound ofJ, and the control of the upper bound ofJ in terms of the time integral of ϑ. By using the positive lower bound ofJ, one obtains a density-weighted embedding inequality (which can be viewed as the replacement of the Sobolev embedding inequal- ity when the vacuum is involved) forϑ; see (ii) of Proposition 2.4, which implies that

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(6)

the upper bound of√

%0ϑcan be controlled by that ofJ, up to a small dependence on kyϑ

Jk2, i.e., the termηkyϑ

Jk2, with a small positiveη. This combined with the above identity leads us to carry out theL(L2) type estimates onϑ, or, more precisely, on

√%0ϑ. In the second stage, we carry out theL(L2) energy estimate on√

%0ϑ, and, at the same time, theL(L2) energy estimate onvwill be involved naturally, due to the coupling structure betweenv and ϑin the system; as a conclusion of this stage, by making use the control relationship between the upper bounds of √

%0ϑ and J obtained in the first stage, we are able to obtain the a priori upper bound ofJ and the a prioriL(L2)∩L2(H1) type estimates on (v, ϑ). In the third stage, by investi- gating the effective viscous fluxG:=µJyv −πand working on itsL(L2)∩L2(H1) type a priori estimate, we are able to get the a priori L(H1) estimate on (J, v);

however, due to the presence of the term µJ(∂yv)2 and the degeneracy of the leading term%0tϑin the ϑequation, we are not able to obtain the corresponding L(H1) estimate on ϑ, without appealing to higher order energy estimates thanH1. In the fourth stage, we carry out the a priori L(H2) type estimates, which are achieved through performing theL(L2) type energy estimate on√

%0tϑand L(H1) type estimate on G. It should be mentioned that the desired a prioriL(H2) estimates onϑis obtained without knowing its a prioriL(H1) bound in advance.

The rest of this paper is arranged as follows: in the next section, section 2, which is the main part of this paper, we consider the global existence and the a priori estimates to system (1.8)–(1.10), subject to (1.11)–(1.12), in the absence of vacuum, while Theorem 1.1 is proven in the last section.

Throughout this paper, we useCto denote a general positive constant which may differ from line to line.

2. Global existence and a priori estimates in the absence of vacuum.

We first recall the global existence results due to Kazhikohov and Shelukhin [19]

stated in the following proposition. The original result in [19] was stated in the Lagrangian mass coordinates, rather than the Lagrangian map flow coordinates as here, and the initial data (%0, v0, ϑ0) was assumed inH1; however, due to the sufficient regularities of the solutions established in [19], in particular theL1(0, T;W1,∞) ofv, the global existence result established there can be translated to the corresponding one in the Lagrangian map flow coordinates, and one can show that the solutions have correspondingly more regularities if the initial data has more regularities as stated in the following proposition.

Proposition 2.1. For any(%0, v0, ϑ0)∈H2, satisfyingv(0) =v(L) = 0, and min

inf

y∈(0,L)%0(y), inf

y∈(0,L)ϑ0(y)

>0,

there is a unique global solution (J, v, %) to system (1.8)–(1.10), subject to (1.11)–(1.12), satisfyingJ, ϑ >0 on(0, L)×(0,∞), and

J ∈C([0, T];H2),∂tJ ∈L(0, T;H1)∩L2(0, T;H2), v∈C([0, T];H2)∩L2(0, T;H3),∂tv∈L(0, T;L2)∩L2(0, T;H1), ϑ∈C([0, T];H2)∩L2(0, T;H3),∂tϑ∈L(0, T;L2)∩L2(0, T;H1) for any T∈(0,∞).

In the rest of this section, we always assume that (J, v, ϑ) is the unique global solution obtained in Proposition 2.1, and we will establish a series of a priori estimates of (J, v, ϑ) independent of the lower bound of the density.

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(7)

We start with the basic energy identity in the following proposition.

Proposition 2.2. It holds that Z L

0

J(y, t)dy=L and

Z L

0

%0

2 v2+cv%0ϑ dy

!

(t) =E0 for any t∈(0,∞), whereE0:=RL

0

%0

2v20+cv%0ϑ0

dy.

Proof. The first conclusion follows directly from integrating (1.8) with respect to y over (0, L) and using the boundary condition (1.11). Multiplying (1.9) by v, integrating the resultant over (0, L), one gets from integrating by parts that

1 2

d dt

Z L

0

%0v2dy+µ Z L

0

(∂yv)2 J dy=

Z L

0

yvπdy.

Integrating (1.10) over (0, L) yields cv

d dt

Z L

0

%0ϑdy+ Z L

0

yvπdy=µ Z L

0

(∂yv)2 J dy, which, summed with the previous equality, leads to

d dt

Z L

0

%0

2 v2+cv%0ϑ dy= 0;

the second conclusion follows.

2.1. A priori L2 estimates. In this subsection, we will derive the uniform positive lower and upper bounds of J and the a priori L(0, T;L2)∩L2(0, T;H1) estimates ofv andϑ.

Before carrying out the desired estimates, we first derive an equality, i.e., (2.3) below, in the spirit of [19]. As will be shown later, this equality leads to the positive lower bound ofJ, and it will be combined with theL2type energy inequalities to get the upper bound ofJ and the a prioriL(0, T;L2)∩L2(0, T;H1) estimates ofvand ϑ. Due to (1.8), it follows from (1.9) that

%0tv−µ∂ytlogJ+∂yπ= 0.

Integrating the above equation with respect totover (0, t) yields

%0(v−v0)−µ∂ylogJ+∂y

Z t

0

πdτ

= 0, from which, integrating with respect toy over (z, y), one obtains

Z y

z

%0(v−v0)dξ−µ(logJ(y, t)−logJ(z, t)) + Z t

0

(π(y, τ)−π(z, τ))dτ = 0 for anyy, z∈(0, L). Using

Z y

z

%0(v−v0)dξ= Z y

0

%0(v−v0)dξ− Z z

0

%0(v−v0)dξ,

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(8)

and rearranging the terms, we obtain Z y

0

%0(v−v0)dξ−µlogJ(y, t) + Z t

0

π(y, τ)dτ

= Z z

0

%0(v−v0)dξ−µlogJ(z, t) + Z t

0

π(z, τ)dτ ∀y, z∈(0, L).

Therefore, the function Z y

0

%0(v−v0)dξ−µlogJ(y, t) + Z t

0

π(y, τ)dτ is independent ofy, and we denote it byh(t), that is,

Z y

0

%0(v−v0)dξ−µlogJ+ Z t

0

πdτ =h(t), from which, recallingπ=R%J0ϑ, one gets

(2.1) 1

J exp R

µ%0

Z t

0

ϑ Jdτ

=H(t)B(y, t), or equivalently

(2.2) exp

R µ%0

Z t

0

ϑ Jdτ

=J(y, t)H(t)B(y, t), where

H(t) = exp h(t)

µ

, B(y, t) = exp

−1 µ

Z y

0

%0(v−v0)dξ

. Multiplying (2.1) by Rµ%0ϑone obtains

t

exp R

µ%0 Z t

0

ϑ Jdτ

=R

µ%0ϑHB, which gives

exp R

µ%0 Z t

0

ϑ Jdτ

= 1 +R µ%0

Z t

0

ϑHBdτ.

Combining this with (2.2), one gets

(2.3) 1 +R

µ%0(y) Z t

0

ϑ(y, τ)H(τ)B(y, τ)dτ =J(y, t)H(t)B(y, t) for anyy∈(0, L) and anyt∈[0,∞).

A prior positive lower bound of J and the control of the upper bound of J in terms of%0ϑare stated in the following proposition.

Proposition 2.3. We have the following estimate:

(m1f1(t))−1≤J(y, t)≤m21+R

µm31f1(t) Z t

0

%0ϑdτ for any y∈(0, L)and any t∈[0,∞), where

m1= exp 2

µ

p2k%0k1E0

, f1(t) =m1exp

Rm21E0t µcvL

.

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(9)

Proof. By Proposition 2.2, it follows from the H¨older inequality that

Z y

0

%0(v−v0)dξ

≤ Z L

0

(|%0v|+|%0v0|)dξ≤2k%0k112(2E0)12, and, thus,

(2.4) m:= exp

−2 µ

p2k%0k1E0

≤B(y, t)≤exp 2

µ

p2k%0k1E0

=:m1. Integrating (2.3) over (0, L), it follows from (2.4) and Proposition 2.2 that

L≤ Z L

0

1 + R

µ%0

Z t

0

ϑHBdτ

dy=H(t) Z L

0

J Bdy

≤m1H(t) Z L

0

J dy=m1LH(t) and

mLH(t) =H(t)m Z L

0

J dy

≤H(t) Z L

0

J Bdy= Z L

0

1 +R

µ%0 Z t

0

ϑHBdτ

dy

≤L+Rm1 µ

Z t

0

Z L

0

%0ϑdyHdτ ≤L+Rm1E0 µcv

Z t

0

Hdτ.

Therefore, we have

m−11 ≤H(t)≤m−1+Rm1E0

µcvmL Z t

0

Hdτ, from which, by the Gronwall inequality, one further obtains (2.5) m−11 ≤H(t)≤m−1exp

Rm1E0t µcvmL

=:f1(t), t∈[0, T].

Due to (2.4) and (2.5), it follows from (2.3) that 1≤1 + R

µ%0 Z t

0

ϑHBdτ=J(y, t)H(t)B(y, t)≤J(y, t)m1f1(t), and

m−11 mJ(y, t)≤H(t)B(y, t)J(y, t) = 1 +R µ%0

Z t

0

ϑHBdτ

≤1 + R

µm1f1(t) Z t

0

%0ϑdτ.

Therefore, we have

(m1f1(t))−1≤J(y, t)≤m1m−1+Rm21f1(t) µm

Z t

0

%0ϑdτ for anyy∈(0, L) and anyt∈[0, T], proving the conclusion.

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(10)

As a preparation of deriving the a priori upper bound of J and the a priori L(0, T;L2)∩L2(0, T;H1) type estimates on (v, ϑ), we prove the following proposi- tion, which, in particular, gives the density-weighted estimate ofϑ.

Proposition 2.4. We have the following two items.

(i)It holds that k%20ϑk2

E0

cv

2 8 ¯%2

L2 + 32k%00k2

+ 6 ¯%103 E0

cv

23

yϑ

√ J

4 3

2

kJk23 , kϑk≤√

L

yϑ

√J 2

+ 2E0

cvω0%¯, where

¯

%=k%0k, ω0=|Ω0|, Ω0:=n

y∈(0, L)

%0(y)≥%¯ 2 o

. (ii)As a consequence of(i), we have

k√

%0ϑk2≤η

yϑ

√J

2

2

+Cη(kJk2+ 1)

for anyη ∈(0,∞), whereCη is a positive constant depending only onηandN1, where N1:= E0

cv

+ ¯%+1

¯

%+L+ 1 L+ 1

ω0

+k%00k. Proof. (i) Choosey0∈(0, L) such that

%20(y0)ϑ(y0, t)≤ 2 L

Z L

0

%20ϑdξ≤2 ¯% Lk%0ϑk1. By the H¨older and Young inequalities, we deduce

(%20ϑ)2(y, t)≤(%20ϑ)2(y0, t) + 2 Z L

0

%20ϑ|∂y(%20ϑ)|dξ

≤ 2 ¯%

L 2

k%0ϑk21+ 2 Z L

0

(2%20ϑ%0ϑ|%00|+%20ϑ%20|∂yϑ|)dξ

≤ 2 ¯%

L 2

k%0ϑk21+ 4k%20ϑkk%0ϑk1k%00k+ 2 ¯%52

yϑ

√J 2

kJk12 k%0ϑk112k%20ϑk12

≤ 1

2k%20ϑk2+ 2 ¯%

L 2

k%0ϑk21+ 16k%0ϑk21k%00k2+ 3 ¯%103k%0ϑk123

yϑ

√J

4 3

2

kJk23

for anyy∈(0, L), and, thus, k%20ϑk2≤2

2 ¯% L

2

k%0ϑk21+ 32k%0ϑk21k%00k2+ 6 ¯%103k%0ϑk123

yϑ

√ J

4 3

2

kJk23 , from which, by Proposition 2.2, we have

k%20ϑk2≤ E0

cv

28 ¯%2

L2 + 32k%00k2

+ 6 ¯%103 E0

cv

23

yϑ

√ J

4 3

2

kJk23 .

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(11)

Noticing that

ϑ(y, t) = 1 ω0

Z

0

ϑdz+ 1 ω0

Z

0

Z y

z

yϑdξdz, we deduce, by the H¨older inequality, that

kϑk≤ 1 ω0

Z

0

%0ϑ

%0

dz+ Z L

0

|∂yϑ|dz

≤ 2

ω0%¯k%0ϑk1+ Z L

0

yϑ

√ J

2

dz

!12 Z L

0

J dz

!12 , from which, by Proposition 2.2, one obtains

kϑk≤√ L

yϑ

√J 2

+ 2E0

cvω0%¯. (ii) Thanks to (i), one has

k%20ϑk2≤C 1 +

yϑ

√ J

4 3

2

kJk23

!

, kϑk≤C

yϑ

√ J

2

+ 1

, for a positive constantC depending only onN1. Therefore, we have

k√

%0ϑk2=

(%20ϑ)14ϑ34

2

≤ k%20ϑk12 kϑk32

≤C 1 +

yϑ

√ J

4 3

2

kJk23

!14

yϑ

√ J

2

+ 1 32

,

for a positive constantCdepending only onN1, from which, by the Young inequality, (ii) follows.

We can now prove the desired a prioriL(0, T;L2)∩L2(0, T;H1) estimates on (v, ϑ) as in the next proposition.

Proposition 2.5. GivenT ∈(0,∞), it holds that sup

0≤t≤T

k(√

%0v2,√

%0ϑ)k22+kJk2 +

Z T

0

kϑk2+k(∂yϑ, v∂yv)k22 dt

≤C(1 +k(√

%0ϑ0,√

%0v02)k22)

for a positive constant Cdepending only onR, cv, µ, κ, m1, N1, andT, wherem1 and N1 are the numbers in Proposition2.3 and Proposition2.4, respectively.

Proof. DenoteE:= v22+cvϑ. Then one can derive from (1.9) and (1.10) that (2.6) %0tE+∂y(vπ)−κ∂y

yϑ J

=µ∂y

1 J∂y

v2 2

.

Multiplying (2.6) byE and integrating the resultant over (0, L), one get from integra- tion by parts that

(2.7) 1

2 d dt

Z L

0

%0E2dy+ Z L

0

1

J(κ∂yϑ+µv∂yv)∂yEdy= Z L

0

vπ∂yEdy.

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(12)

By the Young inequality, we have Z L

0

1

J(κ∂yϑ+µv∂yv)∂yEdy= Z L

0

1

J(κ∂yϑ+µv∂yv)(v∂yv+cvyϑ)dy

≥ 3κcv

4 Z L

0

yϑ

√J

2

dy−C Z L

0

v∂yv

√J

2

dy and

Z L

0

vπ∂yEdy=R Z L

0

v%0

J ϑ(v∂yv+cvyϑ)dy

≤ κcv

4 Z L

0

yϑ

√J

2

dy+C Z L

0

1

J[%20v2ϑ2+ (v∂yv)2]dy

for a positive constant C depending only on R, cv, µ, andκ. Substituting the above two inequalities into (2.7) and applying Proposition 2.2 and Proposition 2.3, one obtains

d dtk√

%0Ek22+κcv

yϑ

√ J

2

2

≤C

v∂yv

√ J

2

2

+ Z L

0

%0v2 J %0ϑ2dy

!

≤C

v∂yv

√ J

2

2

+k√

%0vk22k√

%0ϑk2m1f1(t)

!

≤C

v∂yv

√J

2

2

+E0m1f1(t)k√

%0ϑk2

!

for a positive constantC depending only onR, cv, µ, andκ, and, thus,

(2.8) d

dtk√

%0Ek22+κcv

yϑ

√ J

2

2

≤A1

v∂yv

√ J

2

2

+E0m1f1(t)k√

%0ϑk2

!

for a positive constantA1 depending only onR, cv, µ, and κ.

Multiplying (1.9) by 4v3 and integrating the resultant over (0, L), one gets from integration by parts and the Young inequality that

d dt

Z L

0

%0v4dy+ 12µ Z L

0

v∂yv

√J

2

dy

= 12 Z L

0

πv2yvdy= 12R Z L

0

%0

J ϑv2yvdy

≤6µ Z L

0

v∂yv

√ J

2

dy+6R2 µ

Z L

0

%0v2

J %0ϑ2dy,

from which, by Proposition 2.3, the H¨older inequality, and Proposition 2.2, one obtains d

dt Z L

0

%0v4dy+ 6µ Z L

0

v∂yv

√ J

2

dy≤6R2 µ

Z L

0

%0v2 J %0ϑ2dy

≤6R2

µ m1f1(t)k√

%0vk22k√

%0ϑk2≤ 12R2m1E0

µ f1(t)k√

%0ϑk2;

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(13)

that is,

(2.9) d

dtk√

%0v2k22+ 6µ

v∂yv

√J

2

2

≤12R2

µ E0m1f1(t)k√

%0ϑk2.

Multiplying (2.9) by Aµ1, adding the resultant to (2.8), and noticing thatf1(t) is nondecreasing int, one gets

d dt

k√

%0Ek22+A1

µ k√

%0v2k22

+ 5A1

v∂yv

√ J

2

2

+κcv

yϑ

√ J

2

2

1 + 12R2 µ2

A1E0m1f1(t)k√

%0ϑk2≤A2E0m1f1(T)k√

%0ϑk2(t) (2.10)

for anyt∈(0, T), where A2= (1 +12Rµ22)A1.

By Proposition 2.3 and (ii) of Proposition 2.4, we have k√

%0ϑk2(t)≤η

yϑ

√ J

2

2

(t) +Cη

1 + Z t

0

k√

%0ϑk2

for anyη∈(0,∞), and anyt∈(0, T), whereCηis a positive constant depending only on R, cv, µ, κ, m1, N1, T, and η. Multiplying both sides of the above inequality by 2A2E0m1f1(T), choosingη = 4A κcv

2E0m1f1(T), and summing the resultant with (2.10), one obtains

d dt

k√

%0Ek22+A1

µ k√

%0v2k22+A2E0m1f1(T) Z t

0

k√

%0ϑk2

+ 5A1

v∂yv

√J

2

2

+κcv

2

yϑ

√J

2

2

≤CA2E0m1f1(T)

1 + Z t

0

k√

%0ϑk2

for anyt∈(0, T), whereCis a positive constant depending only onR, cv, µ, κ, m1, N1, andT. Applying the Gronwall inequality to the above inequality, one gets

sup

0≤t≤T

k(√

%0v2,√

%0ϑ)k22+ Z T

0

(k√

%0ϑk2+k(∂yϑ, v∂yv)k22)dt

≤C(1 +k(√

%0ϑ0,√

%0v02)k22) (2.11)

for a positive constant C depending only on R, cv, µ, κ, m1, N1, and T. The desired estimate

sup

0≤t≤T

kJk2+ Z T

0

kϑk2dt≤C(1 +k(√

%0ϑ0,√

%0v02)k22) follows from (2.11), by applying Proposition 2.3 and (i) of Proposition 2.4.

As a corollary of Proposition 2.3 and Proposition 2.5, we have the following corollary.

Corollary 2.1. GivenT ∈(0,∞), it holds that 0<J ≤J(y, t)≤C(1 +k(√

%0v0,√

%0v02,√

%0ϑ0)k22)

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(14)

for all(y, t)∈(0, L)×(0, T),and sup

0≤t≤T

k(√

%0v,√

%0ϑ)k22+ Z T

0

kϑk2+k(∂yϑ, ∂yv)k22 dt

≤C(1 +k(√

%0v0,√

%0v20,√

%0ϑ0)k22)

for positive constants J and C depending only on R, cv, µ, κ, m1, N1, and T, where m1 andN1 are the numbers in Proposition 2.3and Proposition 2.4, respectively.

Proof. All estimates except that forRT

0 k∂yvk22dtfollow directly from Proposition 2.3 and Proposition 2.5. Multiplying (1.9) byv, integrating the resultant over (0, L), one gets from integration by parts that

1 2

d dtk√

%0vk22

yv

√J

2

2

= Z L

0

π∂yvdy

=R Z L

0

%0

J ϑ∂yvdy≤ µ 2

yv

√ J

2

2

+ R 2µ

Z L

0

%20 J ϑ2dy, and, thus, by Proposition 2.3,

d dtk√

%0vk22

yv

√J

2

2

≤R

µm1f1(t) ¯%k√

%0ϑk2, from which, by Proposition 2.5, the conclusion follows.

2.2. A prioriH1estimates. This section is devoted to the a priorH1type esti- mates on (J, v, ϑ). Precisely, we will carry out the a prioriL(0, T;H1)∩L2(0, T;H2) estimate on v and the L(0, T;H1) estimate onJ; however, due to the presence of the term µJ(∂yv)2on the right-hand side of the equation forϑ, (1.10), one can not get the desired a prioriH1 estimate of ϑindependent of the lower bound of the density without appealing to the higher thanH1 energy estimates.

Define the effective viscous fluxGas G:=µ∂yv

J −π=µ∂yv J −R%0

J ϑ.

Then, one can derive from (1.8)–(1.10) that (2.12) ∂tG−µ

J∂y

yG

%0

=−κ(γ−1) J ∂y

yϑ J

−γ∂yv J G.

Moreover, by (1.9), one has∂yG=%0tv, from which, recalling the boundary condi- tion (1.11), we have

yG(0, t) =∂yG(1, t) = 0, t∈(0,∞).

We have the a prioriL2 estimates onGsated in the following proposition.

Proposition 2.6. GivenT ∈(0,∞), it holds that sup

0≤t≤T

kGk22+ Z T

0

yG

√%0

2

2

dt≤C,

for a positive constantC depending only on R, cv, µ, κ, m1, N1,N2, andT, where N2:=k√

%0v02k2+k√

%0ϑ0k2+kv00k2,

andm1 andN1 are the numbers in Proposition2.3 and Proposition2.4, respectively.

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

(15)

Proof. Multiplying (2.12) byJ G, integrating the resultant over (0, L), and recall- ing∂yG|y=0,L= 0, one gets from integration by parts that

Z L

0

tGJ Gdy+µ Z L

0

yG

√%0

2

dy=κ(γ−1) Z L

0

yϑ∂yG J dy−γ

Z L

0

yvG2dy.

Using (1.8), one has Z L

0

tGJ Gdy= 1 2

d dt

Z L

0

J G2dy−1 2

Z L

0

tJ G2dy

= 1 2

d dt

Z L

0

J G2dy−1 2

Z L

0

yvG2dy.

Therefore, it follows from the H¨older, Young, and Gagliardo–Nirenberg inequalities and Corollary 2.1 that

1 2

d dt

Z L

0

J G2dy+µ Z L

0

yG

√%0

2

dy

=κ(γ−1) Z L

0

yϑ∂yG J dy+

1 2 −γ

Z L

0

yvG2dy

≤κγ√

¯

%

yG

√%0

2

yϑ

√ J

2

+γk∂yvk2kGk2kGk

≤C

yG

√%0 2

k∂yϑk2+k∂yvk2kGk232(kGk2+k∂yGk2)12

≤µ 2

yG

√%0

2

2

+Ch

(1 +k∂yvk22)kGk22+k∂yϑk22i , that is,

(2.13) d

dtk√

J Gk22

yG

√%0

2

2

≤Ch

(1 +k∂yvk22)kGk22+k∂yϑk22i

for anyt∈(0, T), whereCis a positive constant depending only onR, cv, µ, κ, m1, N1, and T. Applying the Gronwall inequality to (2.13) and using Corollary 2.1, the conclusion follows.

Based on Proposition 2.6, we can obtain the desiredH1 type estimates onJ and v as stated in the next proposition.

Proposition 2.7. GivenT ∈(0,∞), it holds that sup

0≤t≤T

(k∂yJk22+k∂yvk22) + Z T

0

(k√

%0tvk22+k∂y2vk22)dt≤C

for a positive constantC depending only onR, cv, µ, κ, m1, N1, N2, and T, wherem1, N1 andN2 are the numbers in Propositions 2.3, 2.4, and2.6, respectively.

Proof. The estimate sup0≤t≤Tk∂yvk22+RT 0 k√

%0tvk22dt≤C is straightforward from Corollary 2.1 and Proposition 2.6, by the definition of G and noticing that

%0tv=∂yG. Note that, by the Sobolev embedding inequality, it follows from Propo- sition 2.6 that

(2.14)

Z T

0

kGk2dt≤C Z T

0

kGk2H1dt≤C

Downloaded 09/06/19 to 130.102.13.19. Redistribution subject to SIAM license or copyright; see http://www.siam.org/journals/ojsa.php

Références

Documents relatifs

Recently, the notion of memory-efficient represen- tations of the orthonormal basis matrices of the second-order and higher-order Krylov subspaces was generalized to the model

Shu , Stability analysis and a priori error estimates to the third order explicit Runge-Kutta discontinuous Galerkin method for scalar conservation laws, SIAM J. Redistribution

In section 3, we use the sticky particle method to study weak solutions to the mCH equation for general Radon measure initial data.. Space-time BV estimates

Then, making use of this fractional calculus, we introduce the generalized definition of Caputo derivatives. The new definition is consistent with various definitions in the

Based on iteration (3), a decentralized algorithm is derived for the basis pursuit problem with distributed data to recover a sparse signal in section 3.. The algorithm

In the first set of testing, we present several numerical tests to show the sharpness of these upper bounds, and in the second set, we integrate our upper bound estimates into

In the important case in which the underlying sets are convex sets described by convex polynomials in a finite dimensional space, we show that the H¨ older regularity properties

For the direct adaptation of the alternating direction method of multipliers, we show that if the penalty parameter is chosen sufficiently large and the se- quence generated has a