This paper focuses on investigating the initial-boundary value problem of incompressible heat conducting Navier-Stokes equations with variable coefficients over bounded domains in R3, where the viscosity coefficient and heat conduction coefficient are powers of temperature. We obtain the global well-posedness of a strong solution under the assumption that the initial data and the measure of the initial vacuum region are sufficiently small. It is worth mentioning that the initial density is allowed to contain vacuum, and there are no restrictions on the power index of the temperature-dependent viscosity coefficient and heat conductivity coefficient. At the same time, the exponential decay-in-time results are also obtained.
Citation: Jianxia He, Qingyan Li. On the global well-posedness and exponential stability of 3D heat conducting incompressible Navier-Stokes equations with temperature-dependent coefficients and vacuum[J]. Electronic Research Archive, 2024, 32(9): 5451-5477. doi: 10.3934/era.2024253
[1] | Jie Zhang, Gaoli Huang, Fan Wu . Energy equality in the isentropic compressible Navier-Stokes-Maxwell equations. Electronic Research Archive, 2023, 31(10): 6412-6424. doi: 10.3934/era.2023324 |
[2] | Dandan Song, Xiaokui Zhao . Large time behavior of strong solution to the magnetohydrodynamics system with temperature-dependent viscosity, heat-conductivity, and resistivity. Electronic Research Archive, 2025, 33(2): 938-972. doi: 10.3934/era.2025043 |
[3] | Linlin Tan, Bianru Cheng . Global well-posedness of 2D incompressible Navier–Stokes–Darcy flow in a type of generalized time-dependent porosity media. Electronic Research Archive, 2024, 32(10): 5649-5681. doi: 10.3934/era.2024262 |
[4] |
Guochun Wu, Han Wang, Yinghui Zhang .
Optimal time-decay rates of the compressible Navier–Stokes–Poisson system in |
[5] | Yue Cao . Blow-up criterion for the 3D viscous polytropic fluids with degenerate viscosities. Electronic Research Archive, 2020, 28(1): 27-46. doi: 10.3934/era.2020003 |
[6] | Keqin Su, Rong Yang . Pullback dynamics and robustness for the 3D Navier-Stokes-Voigt equations with memory. Electronic Research Archive, 2023, 31(2): 928-946. doi: 10.3934/era.2023046 |
[7] | Guoliang Ju, Can Chen, Rongliang Chen, Jingzhi Li, Kaitai Li, Shaohui Zhang . Numerical simulation for 3D flow in flow channel of aeroengine turbine fan based on dimension splitting method. Electronic Research Archive, 2020, 28(2): 837-851. doi: 10.3934/era.2020043 |
[8] | Shuguan Ji, Yanshuo Li . Quasi-periodic solutions for the incompressible Navier-Stokes equations with nonlocal diffusion. Electronic Research Archive, 2023, 31(12): 7182-7194. doi: 10.3934/era.2023363 |
[9] | Jie Qi, Weike Wang . Global solutions to the Cauchy problem of BNSP equations in some classes of large data. Electronic Research Archive, 2024, 32(9): 5496-5541. doi: 10.3934/era.2024255 |
[10] | Wei Shi, Xinguang Yang, Xingjie Yan . Determination of the 3D Navier-Stokes equations with damping. Electronic Research Archive, 2022, 30(10): 3872-3886. doi: 10.3934/era.2022197 |
This paper focuses on investigating the initial-boundary value problem of incompressible heat conducting Navier-Stokes equations with variable coefficients over bounded domains in R3, where the viscosity coefficient and heat conduction coefficient are powers of temperature. We obtain the global well-posedness of a strong solution under the assumption that the initial data and the measure of the initial vacuum region are sufficiently small. It is worth mentioning that the initial density is allowed to contain vacuum, and there are no restrictions on the power index of the temperature-dependent viscosity coefficient and heat conductivity coefficient. At the same time, the exponential decay-in-time results are also obtained.
In this paper, we consider the nonhomogeneous incompressible heat conducting Navier-Stokes equations that can be written as
{ρt+div(ρu)=0,(ρu)t+div(ρu⊗u)−div(2μD(u))+∇P=0,cv[(ρθ)t+div(ρuθ)]−div(κ∇θ)=2μ|D(u)|2,divu=0, | (1.1) |
supplemented with the initial data:
(ρ,u,θ)(x,0)=(ρ0,u0,θ0)(x),x∈Ω, | (1.2) |
and the boundary condition:
u=0,∇θ⋅n=0,on∂Ω. | (1.3) |
Here Ω⊂R3 is a bounded smooth domain, and n is the unit outward normal to ∂Ω. ρ,u,θ, and P stand for the density, velocity, absolute temperature, and pressure of the fluid, respectively. D(u)=12(∇u+(∇u)T) is the deformation tensor. The coefficients μ, cv, and κ denote the viscosity, specific heat at constant volume, and heat conductivity, respectively.
From a physical perspective, viscosity and heat conductivity coefficients depend on temperature. In fact, when deriving Navier-Stokes equations from the Chapman-Enskog expansion based on the Boltzmann equation, Chapman and Cowling [1], Liu, Xin and Yang [2] confirmed that viscosity and heat conductivity coefficients depend on temperature. For more details, one can refer to [3,4,5,6]. This played a guiding role in later mathematical research, and scholars began to focus on investigating the case of variable coefficients. Specifically, there is a special case
μ=c1θb, κ=c2θb, |
where c1,c2,b are positive constants. For this situation, if the intermolecular potential of the cut-off reverse power force model varies with r−a(a∈(0,+∞], r represents the intermolecular distance), then there is the following interesting discovery in [1]
b=a+42a. |
Furthermore, b=12 corresponds to rigid elastic spherical molecules; b=1 corresponds to Maxwellian molecules; b=52 corresponds to ionized gas. For these reasons, we mainly concentrate on the case that μ,κ satisfy the following physical restrictions:
μ=μ(θ)=θα, κ=κ(θ)=θβ, forα,β≥0. | (1.4) |
The incompressible heat conduction Navier-Stokes equations with constant coefficients have been widely studied, and a large amount of literature can be found on their well-posedness. The weak solvability of homogeneous systems was proved by Lions [7]. For inhomogeneous fluids, Zhang and Tan [8] obtained the global well-posedness of a strong solution for the initial density with a positive lower bound in the whole 3D space. When the initial density includes vacuum, the unique local strong solution was guaranteed by [9] with the help of compatibility conditions. With it, Zhong [10] extended this local solution to the global solution under the condition that μ−4‖√ρ0u0‖2L2‖∇u0‖2L2 is small in 3D bounded domain. Wang, Yu and Zhang [11] studied problems with general external force and also obtained a similar result. The existence and uniqueness of the global strong solution to the Cauchy problem were proved by Zhong [12,13], that is, 2D allows for large initial data, while 3D required ‖ρ0‖L∞‖√ρ0u0‖2L2‖∇u0‖2L2 to be sufficiently small. Furthermore, Zhong [14] also studied the 3D initial-boundary value problem with special boundaries, namely the Navier-slip boundary with velocity and Neumann boundary with temperature, and obtained a globally strong solution under the assumption that ‖√ρ0u0‖2L2‖u0‖2L2 is small. Without thermal conductivity, it is the classic incompressible Navier-Stokes equations. Abidi et al. [15,16], and Craig et al. [17] studied the well-posedness and decay behavior in critical space. There were many important research results on the well-posedness of solutions to nonhomogeneous incompressible Navier-Stokes equations, which can be referred to [18,19,20,21,22], and their references.
In general, the Navier-Stokes equations with variable coefficients can more accurately describe the motion of fluids in practice. When considering that viscosity and heat conduction depend on temperature or density, the strong coupling makes the study of system (1.1) more difficult. For 3D homogeneous incompressible heat conduction Navier-Stokes equations with μ=μ(θ),κ=κ(θ), weak solutions were established by Naumann [23] in bounded domains and by Feireisl and Málek [24] in periodic domains, respectively. Amann [25] considered the initial-boundary value problem with coefficients depending on D(u) and θ, and showed the unique solvability under the assumption of small data through the linearization method. Later, Frehse et al. [26] proved the existence of a global weak solution for shear-thickening heat conducting incompressible fluids with μ=μ(ρ,θ,|D(u)|2),κ=κ(ρ,θ) in bounded regions. When the initial vacuum is taken into account, one needs to put more effort into the structure of the equation. In the 3D domain (both bounded or unbounded), Cho and Kim [9] considered the cases μ=μ(ρ,ρθ),κ=κ(ρ,ρθ) and obtained the unique local strong solution under the compatibility conditions
{−div(2μ0D(u0))+∇P0=√ρ0g1,−div(κ0∇θ0)−2μ0|D(u0)|2=√ρ0g2, |
for some P0∈H1,(g1,g2)∈L2. On this basis, for the case μ=μ(ρ,θ),κ=const., Xu and Yu [27,28] extended this strong solution globally under the small initial energy. Specifically, the global strong solution to the initial-boundary value problem was established in [28], and the global existence and algebraic decay of the strong solution to the Cauchy problem were established in [27]. For the 2D case, if μ=μ(ρ)≥μ_>0,κ=κ(ρ)≥κ_>0, the global strong solution was established by Zhong [29] under the smallness on ‖∇μ(ρ0)‖Lq(2<q<∞). Although these research results considered variable coefficients, it is necessary that the viscosity or heat conductivity does not degenerate. Recently, for the degenerate compressible Naiver-Stokes equations for non-isentropic flow, Xin and his collaborators [30,31] obtained several important progresses on the local-in-time well-posedness of regular solutions with vacuum. Meanwhile, Zhang and Fang [32] also established the unique local strong solution and blow-up criterion for the 3D case. In the absence of vacuum, Guo and Li [33] showed the existence and uniqueness of a global strong solution to the problem (1.1)–(1.4) and also obtained the large-time behavior.
Based on the above research results, we are very interested in the problem of nonhomogeneous heat conduction incompressible Navier-Stokes equations with temperature-dependent viscosity and heat conductivity. Little is known about the global solvability of the initial-boundary value problem (1.1)–(1.4) with vacuum. The problem becomes quite difficult when vacuum is taken into account. To study this issue, we must deal with the following main obstacles:
(i) Strong coupling caused by temperature-dependent viscosity. Compared with the constant coefficients equation, the temperature-dependent equations have a stronger coupling between the momentum equations and the temperature equations, which makes the structure of the equations more complex. The temperature-dependent viscosity makes it difficult to estimate the velocity and temperature in the higher order, and even the lower order estimates cannot be obtained directly. Based on the regularity theory of the Stokes equations (see [21]), to get an H2-estimate of u, we need to first deal with ‖∇μ(θ)‖Lq. Similarly, to get the estimate of θ, we need to deal with ‖κ(θ)‖L∞ first. Dealing with ‖∇μ(θ)‖Lq and ‖κ(θ)‖L∞ is essentially a matter of getting a consistent bound on temperature. It is precisely because we have no restrictions on the power index of viscosity μ and thermal conductivity κ that this creates a certain hindrance to energy estimation. To overcome this difficulty, we need to simultaneously estimate the energy of both velocity and temperature, and we need to make very detailed estimates. Meanwhile, we make full use of the bootstrap argument.
(ii) Strong nonlinearity due to temperature-dependent heat conductivity. For the case where the heat conductivity is a constant, the key is to obtain a consistent estimate of temperature. If the heat conductivity depends on temperature, the estimation of temperature is a huge challenge. In the process of controlling the additional terms generated by nonlinear terms in the energy equation, we do time-weighted energy estimates. We first obtain the upper bound of t12‖∇2θ‖L2, and then combine it with the local existence theorem to obtain a consistent estimate of ‖∇2θ‖L2.
(iii) Degradation caused by vacuum. Vacuum leads to the degeneracy of time evolution in momentum equations and temperature equations. When it comes to the initial vacuum, it is usually natural to introduce the compatibility condition.
These obstacles require us to make full use of its structural characteristics and make precise estimates starting from the equations. The main objective of this article is to investigate the global well-posedness and large time behavior of strong solutions to the (1.1)–(1.4) under the assumption of small initial data.
Without loss of generality, we abbreviate
∫fdx≜∫Ωfdx. | (1.5) |
For 1≤r≤∞,k≥1, the standard Sobolev spaces are abbreviated as follows:
Lr=Lr(Ω), Wk,r=Wk,r(Ω)={f∈Lr|∇af∈Lr, ∀|a|≤k}, Hk=Hk(Ω)=Wk,2,H10,σ=H10,σ(Ω)={f∈H1|divf=0, f=0 on ∂Ω},H2n=H2n(Ω)={f∈H2|∇f⋅n=0 on ∂Ω}. |
The definition of strong solutions to be established in this article is as follows:
Definition 1.1. Given T∈(0,∞) and q∈(3,6], (ρ,u,θ,P) is called a strong solution to the initial-boundary value problem (1.1)–(1.4) on [0,T]×Ω, if
{ρ∈C([0,T];W1,q),ρt∈C([0,T];Lq),u∈C([0,T];H10,σ∩H2)∩L2([0,T];W2,q),θ∈C([0,T];H2n)∩L2([0,T];W2,q),P∈C([0,T];H1)∩L2([0,T];W1,q),(√ρut,√ρθt)∈L∞([0,T];L2),(ut,θt)∈L2([0,T];H1), | (1.6) |
and (ρ,u,θ,P) satisfies the system (1.1)–(1.4) a.e. on [0,T]×Ω.
We prove the global existence of a strong solution, which allows for an initial density containing vacuum. The specific conclusion is as follows:
Theorem 1.1. For some positive constants θ_ and q∈(3,6], suppose that the initial data (ρ0,u0,θ0) satisfies
0≤ρ0∈W1,q,u0∈H10,σ∩H2,θ_≤θ0∈H2n, | (1.7) |
and the compatibility conditions
{−div(2μ(θ0)D(u0))+∇P0=√ρ0g1,−div(κ(θ0)∇θ0)−2μ(θ0)|D(u0)|2=√ρ0g2, | (1.8) |
for some P0∈H1 and (g1,g2)∈L2. Then there exist two positive constants ϵ0 and ϵ1 depending on Ω,θ_,α,β and the initial data, such that if
|V|≤ϵ0,‖√ρ0u0‖L2+β‖√ρ0θ0‖L2≤ϵ1, | (1.9) |
where V={x∈Ω|ρ0(x)=0}, the initial-boundary value problem (1.1)–(1.4) admits a unique global strong solution (ρ,u,θ,P), and the following large-time behavior holds:
‖u‖2H2+‖P‖2H1+‖√ρut‖2L2≤Ce−C−1t, | (1.10) |
‖θ−1¯ρ0|Ω|E0‖2H2+‖√ρθt‖2L2≤Ce−C−1t, | (1.11) |
where
E0=∫ρ0(θ0+12|u0|2)dx,¯ρ0=1|Ω|∫ρ0dx. |
Remark 1.1. The authors in [34] studied the well-posedness of strong solutions to the system (1.1)–(1.4) in 2D and also proposed the condition (1.9)1. In addition, for the initial-boundary value problem of temperature-dependent non-isentropic compressible Navier-Stokes equations with vacuum, the unique local strong solution was established by Cao and his collaborators [35,36] under the assumption that the size of the initial vacuum domain is sufficiently small and compatibility conditions. From this perspective, the conditions (1.8) and (1.9)1 we proposed are also reasonable. Establishing strong solutions without compatibility conditions as in [37] is meaningful and is left for further research.
Remark 1.2. When the initial density is far away from vacuum and ‖√ρ0u0‖L2+‖√ρ0θ0‖L2 is sufficiently small, the authors in [33] established the global existence of a strong solution to the system (1.1)–(1.4) and obtained the exponential decay-in-time results. Compared to their results, our results do not require the initial density to be far away from the vacuum, which is also the highlight of this article.
Here, we would like to express some opinions on the analysis of this article. The local existence and uniqueness of strong solutions are guaranteed by Lemma 1. Our main idea is to reasonably combine the bootstrap argument with time-weighted estimates. We are committed to establishing consistent a prior estimates that are independent of time. As pointed out in [33], the core is to obtain the bound of θ. But the method they use is aimed at the initial density with a positive lower bound and is not suitable for situations with vacuum. In order to overcome the difficulties caused by vacuum, we borrow the technique from [36] and divide the domain Ω into two parts: {x∈Ω|ρ(x)≤c0} and {x∈Ω|ρ(x)>c0}. Further utilize the smallness of the initial data and measure of the initial vacuum region to obtain consistent estimates. With global a priori estimates, we can successfully extend the local solution to the global one. Finally, we also obtain exponential decay.
The remaining content of this article is arranged as follows: in the second section, we reviewed the existence of local strong solutions, basic inequalities and lemmas. In the third section, we mainly establish a priori estimates and prove Theorem 1.1.
In this section, we review the known results and basic inequalities, which are crucial for our subsequent calculations.
We first elaborate on the conclusion of the local existence of a strong solution, which is guaranteed by [9].
Lemma 2.1. Assume that the initial data (ρ0,u0,θ0) satisfy (1.7) and (1.8). Then there exists a small time T0>0 and a unique strong solution (ρ,u,θ,P) to the system (1.1)–(1.4) on Ω×(0,T0).
The following Poincaré-type inequality is extremely helpful for studying problems involving vacuum (see [24]).
Lemma 2.2. Let f∈H1,0≤g≤C1, and ∫gdx≥(C1)−1. Then for p∈[1,6], one has
‖f‖Lp≤C‖gf‖L1+C‖∇f‖L2, |
where C depends only on p,C1,Ω.
The direct application of Lemma 2.2 is to control term ‖θt‖L6. Specifically, we can take g=ρ and combine ∫ρdx=∫ρ0dx to easily derive
‖f‖Lp≤C‖ρf‖L1+C‖∇f‖L2, |
where C only depends on p,‖ρ0‖L1,Ω. This inequality is frequently used in the a priori estimates in Section 3.1.
The following estimation of high-order derivatives is based on the regularity theory of the Stokes equations with variable viscosity coefficient; please refer to [38] for details.
Lemma 2.3. Assume that μ(θ) satisfies
μ(θ)∈C1[0,∞), 0<μ_≤μ(θ)≤¯μ<∞, ∇μ(θ)∈Lq (3<q≤6). |
Let (u,P)∈H10,σ×L2 be the unique weak solution to the following system
{−div(2μ(θ)D(u))+∇P=F,inΩ,divu=0,inΩ,u=0,on∂Ω,∫Pμ(θ)dx=0. | (2.1) |
There are the following conclusions:
(1) If F∈Lr with 2≤r<q, then (u,P)∈W2,r×W1,r and
‖u‖W2,r+‖Pμ(θ)‖W1,r≤C(1+‖∇μ(θ)‖qq−3×5r−62rLq)‖F‖Lr. | (2.2) |
(2)If F∈H1,∇2μ(θ)∈L2, then (u,P)∈H3×H2 and
‖u‖H3+‖Pμ(θ)‖H2≤C(1+‖∇μ(θ)‖4+qq−3H1)‖F‖H1. | (2.3) |
Here C depends only on μ_,¯μ,q,r,Ω.
Proof. The proof of (2.2) has been given in [38], but for simplicity, we have omitted it here. We now provide a detailed proof process for (2.3). We rewrite (2.1) as
−Δu+∇(Pμ(θ))=Fμ(θ)+2∇μ(θ)⋅D(u)μ(θ)−P∇μ(θ)μ(θ)2. |
According to the regularity theory of the Stokes system, Gagliardo-Nirenberg inequality, we have
‖u‖H3+‖Pμ(θ)‖H2≤C‖Fμ(θ)‖H1+C‖∇μ(θ)⋅D(u)μ(θ)‖H1+C‖P∇μ(θ)μ(θ)2‖H1≤C‖F‖L2+C‖∇F‖L2+C‖F∇μ(θ)‖L2+C‖∇μ(θ)⋅D(u)‖L2+C‖∇μ(θ)Pμ(θ)‖L2+C‖|∇2μ(θ)||D(u)|‖L2+C‖|∇2μ(θ)||Pμ(θ)|‖L2+C‖|∇μ(θ)||∇2u|‖L2+C‖|∇μ(θ)||∇(Pμ(θ))|‖L2+C‖|∇μ(θ)|2|D(u)|‖L2+C‖|∇μ(θ)|2|Pμ(θ)|‖L2≤C‖F‖H1+C‖F‖L4‖∇μ(θ)‖L4+C‖∇μ(θ)‖Lq‖(∇u,Pμ(θ))‖L2qq−2+C‖∇2μ(θ)‖L2‖(∇u,Pμ(θ))‖L∞+C‖∇μ(θ)‖L4‖(∇2u,∇(Pμ(θ)))‖L4+C‖∇μ(θ)‖2L6‖(∇u,Pμ(θ))‖L6≤C(1+‖∇μ(θ)‖H1)‖F‖H1+C‖∇μ(θ)‖H1‖(∇u,Pμ(θ))‖H1+C‖∇μ(θ)‖H1(‖∇u‖14L2‖∇3u‖34L2+‖Pμ(θ)‖14L2‖∇2(Pμ(θ))‖34L2+‖(∇u,Pμ(θ))‖L2)+C‖∇μ(θ)‖H1(‖∇2u‖14L2‖∇3u‖34L2+‖∇(Pμ(θ))‖14L2‖∇2(Pμ(θ))‖34L2+‖(∇2u,∇(Pμ(θ)))‖L2)+C‖∇μ(θ)‖2H1‖(∇u,Pμ(θ))‖H1≤12‖(∇3u,∇2(Pμ(θ)))‖L2+C(1+‖∇μ(θ)‖H1)‖F‖H1+C(‖∇μ(θ)‖H1+‖∇μ(θ)‖2H1+‖∇μ(θ)‖4H1)‖(∇u,Pμ(θ))‖H1, |
where we have used the following inequality
‖∇2u‖L4≤C‖∇2u‖14L2‖∇3u‖34L2+C‖∇2u‖L2. |
By virtue of (2.2),
‖u‖H2+‖Pμ(θ)‖H1≤C(1+‖∇μ(θ)‖qq−3Lq)‖F‖L2. |
Hence,
‖u‖H3+‖Pμ(θ)‖H2≤C(1+‖∇μ(θ)‖4+qq−3H1)‖F‖H1, |
which implies that (2.3) is valid. We finish the proof of Lemma 2.3.
This section is divided into two subsections. The first subsection is to establish time-weighted a priori estimates. The second subsection extends the unique local strong solution to the global strong solution using the bootstrap argument based on the obtained estimates. It also proves the conclusion of exponential decay.
In this subsection, we will establish the necessary a priori estimates, which can extend the local solution to the global solution. In order to get consistent estimates that do not depend on time, we do some time-weighted estimates. Let (ρ,u,θ,P) be the strong solution of Navier-Stokes equations (1.1)–(1.4) with initial data (ρ0,u0,θ0) on Ω×(0,T], and the initial data satisfies conditions (1.7)–(1.9). For simplicity, let us take cv=1. For a general positive constant cv, the calculation is the same. The letter C represents a generic positive constant that depends on ˜ρ≜‖ρ0‖L∞,θ_, and Ω but does not depend on T.
Our main idea is to apply the bootstrap argument. To this end, we first propose the following a priori assumptions:
sup0≤t≤T(‖∇u‖2H1+‖θ‖2H2+‖√ρθt‖2L2)+∫T0‖(√ρθt,∇θt,∇2u,∇2θ)‖2L2dt≤M. | (3.1) |
Here M is a positive constant that depends only on the initial data.
We first use the standard method to find the upper and lower bounds of density and the lower bound of temperature.
Lemma 3.1. It holds that ∀(x,t)∈Ω×[0,T],
0≤ρ(x,t)≤˜ρ,θ_≤θ(x,t), | (3.2) |
sup0≤t≤T(ti‖√ρu‖2L2)+∫T0ti‖∇u‖2L2dt≤C‖√ρ0u0‖2L2,i=0,...,8. | (3.3) |
Proof. The proof of (3.2)1 is provided by [7,24,33]. Then, applying the standard maximum principle (see [36] or [24, p.43]) to (1.1)3 along with θ0≥θ_ shows (3.2)2. Finally, multiplying (1.1)2 by u and then integrating over Ω, using integration by parts, one has
12ddt∫ρ|u|2dx+2∫θα|D(u)|2dx=0. | (3.4) |
By integrating the above equation on [0,t] and with the help of θ≥θ_ and 2‖D(u)‖2L2=‖∇u‖2L2, it can be inferred that
sup0≤t≤T‖√ρu‖2L2+∫T0‖∇u‖2L2dt≤C‖√ρ0u0‖2L2. | (3.5) |
Furthermore, multiplying (3.4) by t, integrating with respect to t, and taking advantage of Poincaré inequality and (3.5) yield
sup0≤t≤T(t‖√ρu‖2L2)+∫T0t‖∇u‖2L2dt≤C∫T0‖√ρu‖2L2dt≤C∫T0‖∇u‖2L2dt≤C‖√ρ0u0‖2L2. | (3.6) |
Similarly, for i=2,...,8, the following formula can also be obtained
sup0≤t≤T(ti‖√ρu‖2L2)+∫T0ti‖∇u‖2L2dt≤C‖√ρ0u0‖2L2, | (3.7) |
which, combined with (3.5) and (3.6), can prove that (3.3) holds.
Lemma 3.2. Suppose (ρ,u,θ,P) satisfies the assumptions (3.1). There exists a small positive constant ϵ1 depending on Ω,θ_,α,β and the initial data, such that for any i=1,2,3,4,
sup0≤t≤T‖∇u‖2L2+∫T0‖√ρut‖2L2dt≤C, | (3.8) |
sup0≤t≤T(ti‖∇u‖2L2)+∫T0ti‖√ρut‖2L2dt≤ϵ131. | (3.9) |
Proof. Notice that Lemma 2.2 can directly provide
‖θt‖Lp≤C‖ρθt‖L1+‖∇θt‖L2, ∀p∈[1,6]. | (3.10) |
Multiplying (1.1)2 by ut and integrating it over Ω, according to (3.10) and the Gagliardo-Nirenberg inequality, we can easily show
ddt∫θα|D(u)|2dx+‖√ρut‖2L2=∫αθα−1θt|D(u)|2dx−∫ρ(u⋅∇u)⋅utdx≤C(M)‖θt‖L6‖∇u‖2L125+C‖√ρu‖L3‖∇u‖L6‖√ρut‖L2≤C(M)‖(√ρθt,∇θt)‖L2‖∇u‖32L2‖∇u‖12H1+12‖√ρut‖2L2+C‖√ρu‖L2‖∇u‖L2‖∇u‖2H1. | (3.11) |
Then, integrating the last identity over [0,t], by virtue of Holder's inequality, (3.1) and Lemma 3.1, we can establish
sup0≤t≤T‖∇u‖2L2+∫T0‖√ρut‖2L2dt≤C‖∇u0‖2L2+C(M)sup0≤t≤T‖∇u‖L2(∫T0‖(√ρθt,∇θt)‖2L2dt)12(∫T0‖∇u‖2L2dt)14(∫T0‖∇u‖2H1dt)14+Csup0≤t≤T(‖√ρu‖L2‖∇u‖L2)∫T0‖∇u‖2H1dt≤C+C(M)ϵ121+C(M)ϵ1≤C, |
provided that ϵ1≪1.
Similarly, multiplying (3.11) by ti(i≥1), then integrating it over [0,t], one has from (3.1) and Lemma 3.1 that
sup0≤t≤T(ti‖∇u‖2L2)+∫T0ti‖√ρut‖2L2dt≤C∫T0ti−1‖∇u‖2L2dt+C(M)(∫T0t4i‖∇u‖2L2dt)14+C(M)sup0≤t≤T(ti‖√ρu‖L2)≤Cϵ21+C(M)ϵ121+C(M)ϵ1≤ϵ131. |
The proof of Lemma 3.2 is completed.
Lemma 3.3. Suppose (ρ,u,θ,P) satisfies the assumptions (3.1). There exist two small positive constants ϵ0 and ϵ1 depending on Ω,θ_,α,β and the initial data, such that
sup0≤t≤T((1+t)‖√ρut‖2L2)+∫T0(1+t)‖∇ut‖2L2dt≤C, | (3.12) |
sup0≤t≤T‖∇ρ‖Lq≤C. | (3.13) |
Proof. Differentiating (1.1)2 with respect to t, multiplying the result identity by ut, then integrating over Ω yields
12ddt‖√ρut‖2L2+2∫θα|D(ut)|2dx=∫div(ρu)|ut|2dx+∫div(ρu)u⋅∇u⋅utdx−∫ρut⋅∇u⋅utdx−2α∫θα−1θtD(u):∇utdx≤2∫ρ|u||ut||∇ut|dx+∫ρ|u||∇(u⋅∇u⋅ut)|dx+∫ρ|ut|2|∇u|dx+C(M)∫|θt||∇u||∇ut|dx. | (3.14) |
It follows from Hölder's inequality, the Gagliardo-Nirenberg inequality, Lemmas 3.1–3.2, and (3.1) that
∫ρ|u||∇(u⋅∇u⋅ut)|dx≤C‖u‖L6(‖∇u‖L2‖∇u‖L6‖ut‖L6+‖u‖L6‖∇2u‖L2‖ut‖L6+‖u‖L6‖∇u‖L6‖∇ut‖L2)≤C‖∇u‖2L2‖∇u‖H1‖∇ut‖L2≤14∫θα|D(ut)|2dx+C‖∇u‖4L2‖∇u‖2H1, |
2∫ρ|u||ut||∇ut|dx+∫ρ|ut|2|∇u|dx≤C‖√ρut‖L3‖u‖L6‖∇ut‖L2+(‖√ρut‖14L2‖√ρut‖34L6)2‖∇u‖L2≤C‖√ρut‖12L2‖∇ut‖32L2‖∇u‖L2≤14∫θα|D(ut)|2dx+C‖√ρut‖2L2‖∇u‖4L2. |
The estimate of the last term in inequality (3.14) is more subtle. In view of ρ0∈W1,q↪C, we know that there exists a positive constant c0 such that
|Vc0|≤2|V|, |
where Vc0={x∈Ω|ρ0(x)≤c0} and V={x∈Ω|ρ0(x)=0}. Then
C(M)∫|θt||∇u||∇ut|dx=C(M)∫ρ≤c0|θt||∇u||∇ut|dx+C(M)∫ρ>c0|θt||∇u||∇ut|dx≤C(M)‖1‖L12(ρ≤c0)‖θt‖L6‖∇u‖L4‖∇ut‖L2+C(M)c−120‖√ρθt‖L2‖∇u‖L∞‖∇ut‖L2≤C(M)|Vc0|112‖(√ρθt,∇θt)‖L2‖∇u‖14L2‖∇u‖34H1‖∇ut‖L2+C(M)c−120‖∇u‖L∞‖∇ut‖L2≤14∫θα|D(ut)|2dx+C(M)|V|16‖∇u‖12L2‖(√ρθt,∇θt)‖2L2+C(M)c−10‖∇u‖2L∞ | (3.15) |
owing to |{x∈Ω|ρ(x,t)≤c0}|=|Vc0| (see [7, Theorem 2.1]), (3.1), (3.10) and the Gagliardo-Nirenberg inequality. Moreover, since
‖∇2u‖L6≤C(M)‖ρut+ρu⋅∇u‖L6≤C(M)‖ut‖L6+C(M)‖u‖L∞‖∇u‖L6≤C(M)‖∇ut‖L2+C(M)‖∇u‖2H1, |
then according to the Gagliardo-Nirenberg inequality, we find
‖∇u‖L∞≤C‖∇u‖14L2‖∇2u‖34L6+C‖∇u‖L2≤C(M)‖∇u‖14L2(‖∇ut‖34L2+‖∇u‖32H1)+C‖∇u‖L2. | (3.16) |
Now (3.14) implies
12ddt‖√ρut‖2L2+∫θα|D(ut)|2dx≤C(M)(1+c−40)‖∇u‖2L2+C(M)|V|16‖∇u‖12L2‖(√ρθt,∇θt)‖2L2+C(M)c−10‖∇u‖12L2‖∇2u‖3L2+C‖√ρut‖2L2‖∇u‖4L2. | (3.17) |
Next, multiplying (3.17) by 1+t and integrating it over [0,t], we obtain after using Grönwall's inequality, Lemmas 3.1–3.2, and (3.1) that
sup0≤t≤T((1+t)‖√ρut‖2L2)+∫T0(1+t)‖∇ut‖2L2dt≤exp{Csup0≤t≤T‖∇u‖2L2∫T0‖∇u‖2L2dt}[C+C∫T0‖√ρut‖2L2dt+C(M)(1+c−40)ϵ21+C(M)ϵ160+C(M)c−10(∫T0(1+t)4‖∇u‖2L2dt)14(∫T0‖∇2u‖2L2dt)34]≤C+C(M)c−10ϵ121≤C, |
provided that ϵ0≪1, ϵ1≪1.
Furthermore, we calculate from (1.1)1 that
∂iρt+u⋅∇∂iρ+∂iu⋅∇ρ=0. |
Then, we multiply the last identity by q|∇ρ|q−2∂iρ and integrate it over Ω to obtain
ddt‖∇ρ‖Lq≤C‖∇u‖L∞‖∇ρ‖Lq. | (3.18) |
Combining (3.16) and Lemmas 3.1 and 3.2, we ascertain
∫T0‖∇u‖L∞dt≤C(M)(∫T0‖∇u‖2L2dt)18((∫T0‖∇ut‖2L2dt)38T12+(∫T0‖∇u‖2H1dt)34T18)+C(∫T0‖∇u‖2L2dt)12T12≤C(M)ϵ141+Cϵ1≤C,if0≤T≤1, | (3.19) |
and
∫T0‖∇u‖L∞dt=∫10‖∇u‖L∞dt+∫T1‖∇u‖L∞dt≤C(M)ϵ141+C(M)(∫T1t8‖∇u‖2L2dt)18(∫T1‖∇ut‖2L2dt)38(∫T1t−2dt)12+C(M)(∫T1t2‖∇u‖2L2dt)18(∫T1‖∇u‖2H1dt)34(∫T1t−2dt)18+C(∫T1t2‖∇u‖2L2dt)12(∫T1t−2dt)12≤C(M)ϵ141+Cϵ1≤C,ifT≥1. | (3.20) |
Now according to (3.18) and Grönwall's inequality, one has
sup0≤t≤T‖∇ρ‖Lq≤‖∇ρ0‖Lqexp(C∫T0‖∇u‖L∞dt)≤C. |
The proof of this lemma is finished.
Next, we will make estimates for each order regarding θ. This is the key to all a priori estimates.
Lemma 3.4. Suppose (ρ,u,θ,P) satisfies the assumptions (3.1). It holds that
sup0≤t≤T‖√ρθ‖2L2+∫T0‖∇θ‖2L2dt≤C‖√ρ0θ0‖2L2+ϵ1, | (3.21) |
sup0≤t≤T‖ρθβ+2‖L1+∫T0‖∇θβ+1‖2L2dt≤C‖ρ0θβ+20‖L1+ϵ1, | (3.22) |
sup0≤t≤T‖∇θβ+1‖2L2+∫T0‖√ρθβ2θt‖2L2dt≤C, | (3.23) |
sup0≤t≤T(t‖∇θβ+1‖2L2)+∫T0t‖√ρθβ2θt‖2L2dt≤C‖ρ0θβ+20‖L1+ϵ121, | (3.24) |
sup0≤t≤T‖θ‖H1≤C. | (3.25) |
Proof. Firstly, multiplying (1.1)3 by θ and integrating it over Ω yields
12ddt∫ρ|θ|2dx+∫θβ|∇θ|2dx=2∫θα+1|D(u)|2dx. |
By integrating the above equation on [0,t], we obtain from Lemma 3.1 that
sup0≤t≤T‖√ρθ‖2L2+∫T0‖∇θ‖2L2dt≤C‖√ρ0θ0‖2L2+C(M)∫T0‖∇u‖2L2dt≤C‖√ρ0θ0‖2L2+C(M)ϵ21≤C‖√ρ0θ0‖2L2+ϵ1. |
Similarly, multiplying (1.1)3 by θβ+1 and integrating it over Ω×[0,t], it can be inferred that
sup0≤t≤T‖ρθβ+2‖L1+∫T0‖∇θβ+1‖2L2dt≤C‖ρ0θβ+20‖L1+ϵ1. | (3.26) |
Next, taking (1.1)1 and (1.3) into account, multiplying (1.1)3 by θβθt, and integrating by parts over Ω, we obtain
12ddt‖∇θβ+1‖2L2+‖√ρθβ2θt‖2L2=−∫(ρu⋅∇θ)θβθtdx+∫2θα|D(u)|2θβθtdx≤‖√ρu‖L3‖∇θ‖L6‖√ρθβθt‖L2+2‖θ‖α+βL∞‖D(u)‖2L125‖θt‖L6≤C(M)‖√ρu‖12L2‖∇u‖12L2‖∇2θ‖L2‖√ρθβ2θt‖L2+C(M)‖∇u‖32L2‖∇u‖12H1‖(√ρθt,∇θt)‖L2≤12‖√ρθβ2θt‖2L2+C(M)‖√ρu‖L2‖∇2θ‖2L2+C(M)‖∇u‖L2‖(√ρθt,∇θt)‖L2, | (3.27) |
by the Gagliardo-Nirenberg inequality and (3.1). On the one hand, integrating (3.27) with respect to t leads to
sup0≤t≤T‖∇θβ+1‖2L2+∫T0‖√ρθβ2θt‖2L2dt≤‖∇θβ+10‖2L2+C(M)sup0≤t≤T‖√ρu‖L2∫T0‖∇2θ‖2L2dt+C(M)(∫T0‖∇u‖2L2dt)12(∫T0‖(√ρθt,∇θt)‖2L2dt)12≤C+C(M)ϵ1≤C, | (3.28) |
where in the next to last inequality one has used Lemma 3.1 and (3.1). On the other hand, it follows from (3.27), (3.1), (3.26), and Lemma 3.1 that
sup0≤t≤T(t‖∇θβ+1‖2L2)+∫T0t‖√ρθβ2θt‖2L2dt≤∫T0‖∇θβ+1‖2L2dt+C(M)sup0≤t≤T(t‖√ρu‖L2)+C(M)(∫T0t2‖∇u‖2L2dt)12≤C‖ρ0θβ+20‖L1+ϵ1+C(M)ϵ1≤C‖ρ0θβ+20‖L1+ϵ121. |
By applying Lemma 2.2, (3.26) and (3.28), one can derive
sup0≤t≤T‖θ‖H1≤Csup0≤t≤T(‖ρθ‖L1+‖∇θ‖L2)≤C. |
Therefore, we finish the proof of Lemma 3.4.
Lemma 3.5. Suppose (ρ,u,θ,P) satisfies the assumptions (3.1). It has
sup0≤t≤T‖(√ρθt,∇2θ)‖2L2+∫T0‖(∇2θ,θβ2∇θt)‖2L2dt≤C. | (3.29) |
Proof. Differentiating (1.1)3 with respect to t yields
ρθtt+ρu⋅∇θt−div(θβ∇θt)=−ρtθt−ρtu⋅∇θ−ρut⋅∇θ+2αθα−1θt|D(u)|2+2θα∂t(|D(u)|2)+div(βθβ−1θt∇θ). | (3.30) |
Multiplying (3.30) by θt, then integrating it over Ω, we obtain
12ddt‖√ρθt‖2L2+‖θβ2∇θt‖2L2=−∫ρtθ2tdx−∫ρtu⋅∇θθtdx−∫ρut⋅∇θθtdx+2∫θα(|D(u)|2)tθtdx+2α∫θα−1θt|D(u)|2θtdx−β∫θβ−1θt∇θ⋅∇θtdx≜6∑i=1Ji. | (3.31) |
Now, we will use (1.1), the Gagliardo-Nirenberg inequality, (3.1), (3.10), Lemmas 3.1–3.4, and (3.15) to estimate each term on the right hand of (3.31) as follows:
J1≤∫ρ|u||θt||∇θt|dx≤‖ρu‖L3‖θt‖L6‖∇θt‖L2≤C‖√ρu‖12L2‖∇u‖12L2‖(√ρθt,∇θt)‖L2‖∇θt‖L2≤C(M)ϵ121‖(√ρθt,θβ2∇θt)‖2L2, |
J2=∫u⋅∇ρu⋅∇θθtdx≤‖u‖2L6qq−3‖∇ρ‖Lq‖∇θ‖L2‖θt‖L6≤C‖u‖2q−32qL2‖∇2u‖2q+32qL2‖(√ρθt,∇θt)‖L2≤18‖(√ρθt,θβ2∇θt)‖2L2+C‖u‖2q−3qL2‖∇2u‖2q+3qL2, |
J3≤‖√ρut‖L6‖∇θ‖L2‖√ρθt‖L3≤C‖∇ut‖L2‖√ρθt‖12L2‖(√ρθt,∇θt)‖12L2≤18‖θβ2∇θt‖2L2+C‖(∇ut,√ρθt)‖2L2, |
J4≤C‖∇ut‖2L2+C(M)ϵ160‖(√ρθt,θβ2∇θt)‖2L2+C(M)c−10‖∇u‖2L∞, |
J5≤C(M)∫ρ≤c0|∇u|2θ2tdx+C(M)∫ρ>c0|∇u|2θ2tdx≤C(M)‖1‖L3(ρ≤c0)‖∇u‖2L6‖θt‖2L6+C(M)c−10∫ρ>c0|∇u|2ρθ2tdx≤C(M)ϵ130‖(√ρθt,θβ2∇θt)‖2L2+C(M)c−10‖∇u‖2L∞‖√ρθt‖2L2, |
and
J6≤C(M)β∫ρ≤c0|θt||∇θ⋅∇θt|dx+C(M)c−120β∫ρ>c0√ρ|θt||∇θ⋅∇θt|dx≤C(M)‖1‖L6(ρ≤c0)‖θt‖L6‖∇θ‖L6‖∇θt‖L2+C(M)c−120β‖√ρθt‖L4‖∇θ‖L4‖∇θt‖L2≤C(M)|V|16‖(√ρθt,θβ2∇θt)‖2L2+C(M)c−120β‖√ρθt‖14L2‖(√ρθt,∇θt)‖34L2‖∇θ‖14L2‖∇2θ‖34L2‖∇θt‖L2≤(C(M)ϵ160+18)‖(√ρθt,θβ2∇θt)‖2L2+C(M)c−40β8‖√ρθt‖2L2‖∇θ‖2L2. |
Substituting all the estimates of Ji(i=1,...,6) into (3.31), then taking ϵ0 and ϵ1 satisfy
38+C(M)ϵ121+C(M)ϵ160+C(M)ϵ130≤12, |
we deduce
ddt‖√ρθt‖2L2+‖θβ2∇θt‖2L2≤C‖(∇ut,√ρθt)‖2L2+C‖u‖2q−3qL2‖∇2u‖2q+3qL2+C(M)c−10‖∇u‖2L∞+C(M)c−10(‖∇u‖2L∞+c−30β8‖∇θ‖2L2)‖√ρθt‖2L2. | (3.32) |
In view of Poincaré's inequality, Lemma 3.1 and (3.1), one has
C∫T0ti‖u‖2q−3qL2‖∇2u‖2q+3qL2dt≤Csup0≤t≤T‖∇2u‖2L2(∫T0ti2q2q−3‖∇u‖2L2dt)2q−32q(∫T0‖∇2u‖2L2dt)32q≤C(M)ϵ2q−3q1≤C,i=0,1. | (3.33) |
Next, similar to (3.19), we infer that
C(M)c−10∫T0ti‖∇u‖2L∞dt≤C(M)c−10(∫T0t4i‖∇u‖2L2dt)14((∫T0‖∇ut‖2L2dt)34+(∫T0‖∇u‖4H1dt)34)+C(M)c−10∫T0ti‖∇u‖2L2dt≤C(M)c−10ϵ121+C(M)c−10ϵ21≤C,i=0,1. | (3.34) |
Furthermore, we employ Lemma 3.4 to obtain
(3.35) |
Combining (3.32)–(3.35), Lemmas 3.3, 3.4, and Grönwall's inequality gives
(3.36) |
We then rewrite as
(3.37) |
It follows from the standard -theory of elliptic equations that
(3.38) |
where in the last inequality we have used (3.1). Note that
thus
(3.39) |
On the one hand, according to (3.38), (3.1), Lemma 3.2, (3.36), and Lemma 3.4, we compute
(3.40) |
According to the local existence theorem of the solution (see Lemma 2.1), there exists a small positive time such that
This, together with (3.40), yields
(3.41) |
Moreover, Lemma 3.4, the Gagliardo-Nirenberg inequality, and (3.41) imply
(3.42) |
(3.43) |
On the other hand, by (3.1), (3.38), and (3.39), we discover
(3.44) |
Consequently, combining (3.36), (3.41) with (3.44), we complete the proof of this lemma.
By synthesizing Lemmas 3.1–3.5, the following corollary can be obtained.
Corollary 3.1. There exist two small positive constants and as described in Theorem 1.1, such that if
then it holds for any that
Proof. Now, we consider the following Stokes equations:
(3.45) |
According to Lemma 2.3, we have the following regularity results:
(3.46) |
and
(3.47) |
which indicates that
(3.48) |
Using Lemmas 3.1–3.3, it can be inferred that
(3.49) |
(3.50) |
Hence,
(3.51) |
On the other hand, by virtue of (3.37) and the regularity theory to the elliptic equation, one has
(3.52) |
Combining (3.47), (3.52) and estimates we have obtained, it can be inferred that
(3.53) |
Considering the estimates obtained in Lemmas 3.1–3.5, combined with (3.49)–(3.51) and (3.53), we have completed the proof of the Corollary 3.1. In this way, we close the a priori assumptions (3.1).
With all the a priori estimates established in Corollary 3.1 at hand, the global well-posedness part of Theorem 1.1 then follows by standard procedures.
From Lemma 2.1, we know that there exists a such that the system (1.1)–(1.4) has a unique local strong solution on . We now extend this local solution to the global solution by contradiction. Therefore, from now on, we assume that holds, where and are the same as in Corollary 3.1. Lemma 2.1 and (1.7) indicate that there is a such that (3.1) holds at . Denote
(3.54) |
Obviously, . With the help of Corollary 3.1 and the standard embedding, one can deduce that for any ,
(3.55) |
Now we prove
(3.56) |
Otherwise, we assume that . Since , Lemmas 3.1–3.5 and the continuity argument show that the global estimates stated in Corollary 3.1 hold on . Let , it can be inferred from (3.55) that
(3.57) |
satisfies regularity condition
Thus
with
satisfying , which can be guaranteed by Corollary 3.1. Therefore, taking as the initial data, according to Lemma 2.1, the local strong solution can be extended beyond . This contradicts the definition of . So, (3.54) has been proven.
To complete the entire proof of Theorem 1.1, we only need to prove (1.10) and (1.11), i.e., the following proposition.
Proposition 3.1. It holds that
(3.58) |
(3.59) |
where the positive constants and are defined in Theorem 1.1.
Proof. The proof is divided into the following two steps.
Step 1. The proof of (3.58).
It follows from the Poincaré inequality that
(3.60) |
Substituting (3.60) into (3.4) yields
(3.61) |
Then multiplying (3.61) by and integrating it over , we arrive at
(3.62) |
Next, substitute (3.46) into (3.11), then (3.11) becomes
owing to Young's inequality, Lemma 3.1 and Corollary 3.1. Multiply the last inequality by , using Grönwall's inequality, Lemma 3.4, (3.36), (3.62) and
imply
(3.63) |
Similarly, in view of the proof of Lemma 3.3, (3.48) and
(3.64) |
we thereby obtain
Multiplying the last inequality by , applying (3.62), (3.63), Grönwall's inequality, and Corollary 3.1, we deduce
(3.65) |
Thus, collecting (3.62)–(3.65), we get (3.58) immediately.
Step 2. The proof of (3.59).
The method in [34] is used in this process. Integrating over , we arrive at
(3.66) |
Combining (3.4) and (3.66), it can be inferred that
Hence
(3.67) |
Multiplying by and integrating it over , one obtains
It follows from Poincaré's inequality that
where . Thus, we have
and this combined with (3.62) gives
(3.68) |
Now, we re-estimate (3.27) as follows:
(3.69) |
Multiplying it by , according to (3.58), and Corollary 3.1, we have
Consequently,
(3.70) |
Recalling (3.31), (3.64) and Corollary 3.1 leads to
Then, multiplying the last inequality by , combining the result with Grönwall's inequality, (3.58), (3.70) and Corollary 3.1 gives
(3.71) |
Therefore, we complete the proof of (3.59) by (3.68), (3.70), (3.71), and the following estimate
The proof of Proposition 3.1 is finished.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
The authors would like to thank the editor and reviewers for careful reading, helpful comments and suggestions. The work is supported by the Fundamental Research Funds for the Central Universities, CHD (No.300102122115).
The authors declare there is no conflicts of interest.
[1] | S. Chapman, T. Cowling, The Mathematical Theory of Non-uniform Gases. An Account of the Kinetic Theory of Viscosity, Thermal Conduction and Diffusion in Gases, 3 edition, Cambridge University Press, London, 1970. |
[2] |
T. P. Liu, Z. P. Xin, T. Yang, Vacuum states of compressible flow, Discrete Contin. Dyn. Syst., 4 (1998), 1–32. https://doi.org/10.3934/dcds.1998.4.1 doi: 10.3934/dcds.1998.4.1
![]() |
[3] | E. Feireisl, Dynamics of Viscous Compressible Fluids, Oxford University Press, Oxford, 2004. |
[4] |
H. Grad, Asymptotic theory of the Boltzmann equation, Phys. Fluids, 6 (1963), 147–181. https://doi.org/10.1063/1.1706716 doi: 10.1063/1.1706716
![]() |
[5] |
Y. Guo, S. Q. Liu, Incompressible hydrodynamic approximation with viscous heating to the Boltzmann equation, Math. Models Methods Appl. Sci., 27 (2017), 2261–2296. https://doi.org/10.1142/S0218202517500440 doi: 10.1142/S0218202517500440
![]() |
[6] |
S. Q. Liu, T. Yang, H. J. Zhao, Compressible Navier-Stokes approximation to the Boltzmann equation, J. Differ. Equations, 256 (2014), 3770–3816. https://doi.org/10.1016/j.jde.2014.02.020 doi: 10.1016/j.jde.2014.02.020
![]() |
[7] | P. Lions, Mathematical Topics in Fluid Mechanics. Vol. 1: Incompressible Models, Oxford Lecture Ser. Math. Appl., Oxford University Press, New York, 1996. |
[8] |
X. Zhang, Z. Tan, The global wellposedness of the 3D heat-conducting viscous incompressible fluids with bounded density, Nonlinear Anal. Real World Appl., 22 (2015), 129–147. https://doi.org/10.1016/j.nonrwa.2014.08.001 doi: 10.1016/j.nonrwa.2014.08.001
![]() |
[9] |
Y. Cho, H. Kim, Existence result for heat-conducting viscous incompressible fluids with vacuum, J. Korean Math. Soc., 45 (2008), 645–681. https://doi.org/10.4134/JKMS.2008.45.3.645 doi: 10.4134/JKMS.2008.45.3.645
![]() |
[10] |
X. Zhong, Global strong solutions for 3D viscous incompressible heat conducting Navier-Stokes flows with non-negative density, J. Differ. Equations, 263 (2017), 4978–4996. https://doi.org/10.1016/j.jde.2017.06.004 doi: 10.1016/j.jde.2017.06.004
![]() |
[11] |
W. Wang, H. B. Yu, P. X. Zhang, Global strong solutions for 3D viscous incompressible heat conducting Navier-Stokes flows with the general external force, Math. Methods Appl. Sci., 41 (2018), 4589 –4601. https://doi.org/10.1002/mma.4915 doi: 10.1002/mma.4915
![]() |
[12] |
X. Zhong, Global well-posedness to the Cauchy problem of two-dimensional nonhomogeneous heat conducting Navier-Stokes equations, J. Geom. Anal., 32 (2022), 200. https://doi.org/10.1007/s12220-022-00947-7 doi: 10.1007/s12220-022-00947-7
![]() |
[13] |
X. Zhong, Global well-posedness to the 3D Cauchy problem of nonhomogeneous heat conducting Navier-Stokes equations with vacuum and large oscillations, J. Math. Fluid Mech., 24 (2022), 14. https://doi.org/10.1007/s00021-021-00649-0 doi: 10.1007/s00021-021-00649-0
![]() |
[14] |
X. Zhong, Global existence and large time behavior of strong solutions for 3D nonhomogeneous heat conducting Navier-Stokes equations, J. Math. Phys., 61 (2020), 111503. https://doi.org/10.1063/5.0012871 doi: 10.1063/5.0012871
![]() |
[15] |
H. Abidi, G. L. Gui, P. Zhang, On the decay and stability to global solutions of the 3-D inhomogeneous Navier-Stokes equations, Commun. Pure Appl. Math., 64 (2011), 832–881. https://doi.org/10.1002/cpa.20351 doi: 10.1002/cpa.20351
![]() |
[16] |
H. Abidi, G. L. Gui, P. Zhang, On the wellposedness of three-dimensional inhomogeneous Navier-Stokes equations in the critical spaces, Arch. Ration. Mech. Anal., 204 (2012), 189–230. https://doi.org/10.1007/s00205-011-0473-4 doi: 10.1007/s00205-011-0473-4
![]() |
[17] |
W. Craig, X. D. Huang, Y. Wang, Global strong solutions for the 3D inhomogeneous incompressible Navier-Stokes equations, J. Math. Fluid Mech., 15 (2013), 747–758. https://doi.org/10.1007/s00021-013-0133-6 doi: 10.1007/s00021-013-0133-6
![]() |
[18] |
H. Y. Choe, H. Kim, Strong solutions of the Navier-Stokes equations for nonhomogeneous incompressible fluids, Commun. Partial. Differ. Equations, 28 (2003), 1183–1201. https://doi.org/10.1081/PDE-120021191 doi: 10.1081/PDE-120021191
![]() |
[19] |
R. Danchin, Local and global well-posedness results for flows of inhomogeneous viscous fluids, Adv. Differ. Equations, 9 (2004), 353–386. https://doi.org/10.57262/ade/1355867948 doi: 10.57262/ade/1355867948
![]() |
[20] | G. P. Galdi, An Introduction to the Mathematical Theory of the Navier-Stokes Equations, Springer, New York, 1994. |
[21] |
J. Simon, Nonhomogeneous viscous incompressible fluids: existence of velocity, density, and pressure, SIAM J. Math. Anal., 21 (1990), 1093–1117. https://doi.org/10.1016/j.jmaa.2004.11.022 doi: 10.1016/j.jmaa.2004.11.022
![]() |
[22] |
O. A. Ladyzhenskaya, V. A. Solonnikov, Unique solvability of an initial and boundary value problem for viscous incompressible nonhomogeneous fluids, J. Soviet Math., 9 (1978), 697–749. https://doi.org/10.1007/BF01085325 doi: 10.1007/BF01085325
![]() |
[23] |
J. Naumann, On the existence of weak solutions to the equations of non-stationary motion of heat-conducting incompressible viscous fluids, Math. Methods Appl. Sci., 29 (2006), 1883–1906. https://doi.org/10.1002/mma.754 doi: 10.1002/mma.754
![]() |
[24] |
E. Feireisl, J. Málek, On the Navier-Stokes equations with temperature-dependent transport coefficients, Int. J. Differ. Equations, 2006 (2006), 090616. https://doi.org/10.1155/DENM/2006/90616 doi: 10.1155/DENM/2006/90616
![]() |
[25] | H. Amann, Heat-conducting incompressible viscous fluids. Navier-Stokes equations and related nonlinear problems, Plenum Press, New York, 1995. |
[26] |
J. Frehse, J. Málek, M. Rẑiĉka, Large data existence result for unsteady flows of inhomogeneous shear-thickening heat-conducting incompressible fluids, Commun. Partial Differ. Equations, 35 (2010), 1891–1919. https://doi.org/10.1080/03605300903380746 doi: 10.1080/03605300903380746
![]() |
[27] |
H. Xu, H.B. Yu, Global regularity to the Cauchy problem of the 3D heat conducting incompressible Navier-Stokes equations, J. Math. Anal. Appl., 464 (2018), 823–837. https://doi.org/10.1016/j.jmaa.2018.04.037 doi: 10.1016/j.jmaa.2018.04.037
![]() |
[28] |
H. Xu, H. B. Yu, Global strong solutions to the 3D inhomogeneous heat-conducting incompressible fluids, Appl. Anal., 98 (2019), 622–637. https://doi.org/10.1080/00036811.2017.1399362 doi: 10.1080/00036811.2017.1399362
![]() |
[29] |
X. Zhong, Global strong solution for viscous incompressible heat conducting Navier-Stokes flows with density-dependent viscosity, Anal. Appl., 16 (2018), 623–647. https://doi.org/10.1142/S0219530518500069 doi: 10.1142/S0219530518500069
![]() |
[30] |
Q. Duan, Z. P. Xin, S. G. Zhu, On regular solutions for three-dimensional full compressible Navier-Stokes equations with degenerate viscosities and far field vacuum, Arch. Ration. Mech. Anal., 247 (2023), 71. https://doi.org/10.1007/s00205-022-01840-x doi: 10.1007/s00205-022-01840-x
![]() |
[31] | Q. Duan, Z. P. Xin, S. G. Zhu, Well-posedness of regular solutions for 3-D full compressible Navier-Stokes equations with degenerate viscosities and heat conductivity, preprint, arXiv: 2307.06609. https://doi.org/10.48550/arXiv.2307.06609 |
[32] | T. Zhang, D.Y. Fang, Existence and uniqueness results for viscous, heat-conducting 3-D fluid with vacuum, preprint, arXiv: math/0702170. https://doi.org/10.48550/arXiv.math/0702170 |
[33] |
Z. H. Guo, Q. Y. Li, Global existence and large time behaviors of the solutions to the full incompressible Navier-Stokes equations with temperature-dependent coefficients, J. Differ. Equations, 274 (2021), 876–923. https://doi.org/10.1016/j.jde.2020.10.031 doi: 10.1016/j.jde.2020.10.031
![]() |
[34] | W. C. Dong, Q. Y. Li, Global well-posedness for the 2D incompressible heat conducting Navier-Stokes equations with temperature-dependent coefficients and vacuum, preprint, arXiv: 2401.06433. https://doi.org/10.48550/arXiv.2401.06433 |
[35] |
Y. Cao, Y. C. Li, S. G. Zhu, Local classical solutions to the full compressible Navier-Stokes system with temperature-dependent heat conductivity, Methods Appl. Anal., 28 (2021), 105–152. https://doi.org/10.4310/MAA.2021.v28.n2.a2 doi: 10.4310/MAA.2021.v28.n2.a2
![]() |
[36] |
Y. Cao, Y. C. Li, Local strong solutions to the full compressible Navier-Stokes system with temperature-dependent viscosity and heat conductivity, SIAM J. Math. Anal., 54 (2022), 5588–5628. https://doi.org/10.1137/21M1419544 doi: 10.1137/21M1419544
![]() |
[37] |
J. K. Li, Y. S. Zheng, Local existence and uniqueness of heat conductive compressible Navier-Stokes equations in the presence of vacuum without initial compatibility conditions, J. Math. Fluid Mech., 25 (2023), 14. https://doi.org/10.1007/s40818-019-0064-5 doi: 10.1007/s40818-019-0064-5
![]() |
[38] |
X. D. Huang, Y. Wang, Global strong solution of 3D inhomogeneous Navier-Stokes equations with density-dependent viscosity, J. Differ. Equations, 259 (2015), 1606–1627. https://doi.org/10.1016/j.jde.2015.03.008 doi: 10.1016/j.jde.2015.03.008
![]() |