This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Traveling Wave Solutions to Brenner-Navier-Stokes-Fourier system

Saehoon Eo Department of Mathematical Sciences,
Korea Advanced Institute of Science and Technology
Daejeon 34141, Korea
[email protected]
Namhyun Eun Department of Mathematical Sciences,
Korea Advanced Institute of Science and Technology
Daejeon 34141, Korea
[email protected]
Moon-Jin Kang Department of Mathematical Sciences,
Korea Advanced Institute of Science and Technology
Daejeon 34141, Korea
[email protected]
 and  HyeonSeop Oh Department of Mathematical Sciences,
Korea Advanced Institute of Science and Technology
Daejeon 34141, Korea
[email protected]
Abstract.

As a continuum model for compressible fluid flows, Howard Brenner proposed the so-called Brenner-Navier-Stokes-Fourier(BNSF) system that improves some flaws of the Navier-Stokes-Fourier(NSF) system. For BNSF system, the volume velocity concept is introduced and is far different from the mass velocity of NSF, since the density of a compressible fluid is inhomogeneous. Although BNSF was introduced more than ten years ago, the mathematical study on BNSF is still in its infancy. We consider the BNSF system in the Lagrangian mass coordinates. We prove the existence and uniqueness of monotone traveling wave solutions to the BNSF system. We also present some quantitative estimates for them.

Key words and phrases:
Traveling Wave Solutions, Brenner-Navier-Stokes-Fourier, Viscous Conservation Laws, Geometric singular perturbation
1991 Mathematics Subject Classification:
76N15, 35Q30, 35C07
Acknowledgment. The authors thank Professor Alexis Vasseur for valuable comments on BNSF system. This work was supported by Samsung Science and Technology Foundation under Project Number SSTF-BA2102-01.

1. Introduction

We consider the so-called Brenner-Navier-Stokes-Fourier(BNSF) system (in the Eulerian coordinates) in 1D :

(1.1) {tρ+x(ρum)=0,t(ρuv)+x(ρuvum+p)=x(μxuv),t(ρ(uv22+e))+x(ρ(uv22+e)um+puv)=x(μuvxuv+κxθ).\displaystyle\left\{\begin{aligned} &\partial_{t}\rho+\partial_{x}(\rho u_{m})=0,\\ &\partial_{t}(\rho u_{v})+\partial_{x}(\rho u_{v}u_{m}+p)=\partial_{x}\left(\mu\partial_{x}u_{v}\right),\\ &\partial_{t}\left(\rho\left(\frac{u_{v}^{2}}{2}+e\right)\right)+\partial_{x}\left(\rho\left(\frac{u_{v}^{2}}{2}+e\right)u_{m}+pu_{v}\right)=\partial_{x}\left(\mu u_{v}\partial_{x}u_{v}+\kappa\partial_{x}\theta\right).\end{aligned}\right.

Here, ρ\rho is the density, umu_{m} is the mass velocity, uvu_{v} is the volume velocity, pp is the pressure, and ee is the internal energy. In addition, μ\mu and κ\kappa represent the viscosity coefficient and the heat-conductivity coefficient, respectively, which are some positive constants depending only on the type of gas.

In a series of papers [2, 3, 4, 5], Howard Brenner proposed the system (1.1) to improve some defects of the Navier-Stokes-Fourier(NSF) system. His idea is based on so-called ‘bi-velocity theory’ which indicates the existence of two different velocities: one is the mass velocity umu_{m} which is the classical concept; the other is the volume velocity uvu_{v}. Note that these two velocities are implicitly assumed to be identical in the vast majority of studies on continuum fluid mechanics. This perspective has remained unchallenged since Euler’s era. However, Brenner contended that, in general, umuvu_{m}\neq u_{v}, and the inconsistency increases as the density gradient becomes larger. In fact, there have been numerous studies pointing out and attempting to improve the imperfections of the NSF system(see [1, 6, 8, 9, 10, 11, 12, 13, 14, 17, 18, 19, 21, 28, 29, 30, 31, 32, 33]), but it can be said that Brenner’s approach is the most systematic and well-established.

The necessity of the concept of a new velocity can be explained as follows: consider the work generated by the displacement of gas particles due to the pressure within the gas. To calculate the amount of the work, the displacement distance must be defined. However, it is crucial to point out that it is not the individual gas particle that forms pressure and is displaced by pressure, but rather a collection of gas particles. Since the boundary of a collection of gas particles cannot be clearly defined, it intuitively seems impossible to rigorously define the displacement distance. This is related to the fact that the volume of gas is not a mass-point property. The volume of gas cannot be understood as a property possessed by particles at a point but can be considered statistically as a property of a collection of particles at best. Hence, the volume velocity, as a new velocity, is required to describe the motion of a collection of particles, as opposed to the conventional velocity. Therefore, in the existing NSF system, if we only use the mass velocity for the mass transportation and the convection terms, but replace all other velocities by the volume velocity, it results in the BNSF system (1.1). Note that the concept of volume velocity is inherently elusive. Thus, it must be defined by a certain constitutive equation.

Brenner proposed the volume velocity concept through the following argument. First, he derived the constitutive equations from linear irreversible thermodynamics. He then demonstrated that these equations align with the Burnett’s solution to the Boltzmann equation for a dilute monatomic case. Additionally, it turns out that they better match the experimental data. One remark is that when the no-slip boundary condition at solid surfaces is described by the volume velocity, it more closely corresponds to the experimental data as well. (See [5])

The constitutive relation between uvu_{v} and umu_{m} suggested by Brenner is as follows:

(1.2) uv=um+κρcplnρ,u_{v}=u_{m}+\frac{\kappa}{\rho c_{p}}\nabla\ln\rho,

where cpc_{p} is the specific heat at constant pressure. The difference between the two velocities is proportional to ρ\nabla\rho. Thus, in incompressible fluids with uniform density, the concept of volume velocity becomes insignificant. Brenner mentioned that the NSF system is a PDE limited to incompressible fluids, and it seems it took a long time to discover the flaws of the NSF system because, during its initial validation, data from liquids or essentially incompressible gases were mainly used. (See [5])

Although this amendment to the NSF system were proposed over ten years ago, the mathematical study on BNSF is still in its infancy. We refer to Feireisl-Vassuer [16], which proved the existence of the global weak solutions to the initial boundary value problem.

In this paper, we consider the one-dimensional BNSF system (1.1) in the Lagrangian mass coordinates: (still use the same notation xx to denote the mass variable yy)

(1.3) {vtux=(τvxv)x,ut+p(v,θ)x=(μuxv)x,Et+(p(v,θ)u)x=(κθxv)x+(μuuxv)x,\left\{\begin{aligned} &v_{t}-u_{x}=\left(\tau\frac{v_{x}}{v}\right)_{x},\\ &u_{t}+p(v,\theta)_{x}=\left(\mu\frac{u_{x}}{v}\right)_{x},\\ &E_{t}+(p(v,\theta)u)_{x}=\left(\kappa\frac{\theta_{x}}{v}\right)_{x}+\left(\mu\frac{uu_{x}}{v}\right)_{x},\end{aligned}\right.

where vv is the specific volume, uu is now the volume velocity, θ\theta is the absolute temperature, E=e+u22E=e+\frac{u^{2}}{2} is the total energy, and τ=κ/cp\tau=\kappa/c_{p} is the Brenner coefficient. When converting to Lagrangian mass coordinates with the constitutive relation (1.2), the convection terms disappear, thus Brenner’s modification only remains in the mass conservation law. We also consider the ideal polytropic gas where the pressure law pp and the internal energy ee function are given by

(1.4) p(v,θ)=Rθv,e(θ)=Rγ1θ+const.p(v,\theta)=\frac{R\theta}{v},\qquad e(\theta)=\frac{R}{\gamma-1}\theta+const.

where R>0R>0 is the gas constant and γ>1\gamma>1 is the adiabatic constant.

In this article, we prove that the system (LABEL:main) admits viscous shock waves connecting two end states (v,u,E)(v_{-},u_{-},E_{-}) and (v+,u+,E+)(v_{+},u_{+},E_{+}) when these two end states are close enough and satisfy the Rankine-Hugoniot condition and Lax entropy condition as follows:

(1.5) σs.t.{σ(v+v)(u+u)=0,σ(u+u)+p(v+,θ+)p(v,θ)=0,σ(E+E)+p(v+,θ+)u+p(v,θ)u=0,\displaystyle\exists~{}\sigma\quad\text{s.t.}~{}\left\{\begin{aligned} &-\sigma(v_{+}-v_{-})-(u_{+}-u_{-})=0,\\ &-\sigma(u_{+}-u_{-})+p(v_{+},\theta_{+})-p(v_{-},\theta_{-})=0,\\ &-\sigma(E_{+}-E_{-})+p(v_{+},\theta_{+})u_{+}-p(v_{-},\theta_{-})u_{-}=0,\end{aligned}\right.
and either v>v+v_{-}>v_{+}, u>u+u_{-}>u_{+} and θ<θ+\theta_{-}<\theta_{+} or v<v+v_{-}<v_{+}, u>u+u_{-}>u_{+} and θ>θ+\theta_{-}>\theta_{+} holds.

Here, θ\theta_{-} and θ+\theta_{+} satisfy

E=Rγ1θ+u22,E+=Rγ1θ++u+22.E_{-}=\frac{R}{\gamma-1}\theta_{-}+\frac{u_{-}^{2}}{2},\qquad E_{+}=\frac{R}{\gamma-1}\theta_{+}+\frac{u_{+}^{2}}{2}.

In other words, for given sufficiently close constant states (v,u,E)(v_{-},u_{-},E_{-}) and (v+,u+,E+)(v_{+},u_{+},E_{+}) satisfying (LABEL:end-con), we prove the existence of a viscous shock wave (v~,u~,θ~)(ξ)=(v~,u~,θ~\tilde{v},\tilde{u},\tilde{\theta})(\xi)=(\tilde{v},\tilde{u},\tilde{\theta})(xσtx-\sigma t), as a traveling wave solution to the following system of ODEs:

(1.6) {σv~u~=(τv~v~),σu~+p(v~,θ~)=(μu~v~),σ(Rγ1θ~+u~22)+(p(v~,θ~)u~)=(κθ~v~)+(μu~u~v~),limξ±(v~,u~,θ~)(ξ)=(v±,u±,θ±).(=ddξ)\left\{\begin{aligned} &-\sigma\tilde{v}^{\prime}-\tilde{u}^{\prime}=\left(\tau\frac{\tilde{v}^{\prime}}{\tilde{v}}\right)^{\prime},\\ &-\sigma\tilde{u}^{\prime}+p(\tilde{v},\tilde{\theta})^{\prime}=\left(\mu\frac{\tilde{u}^{\prime}}{\tilde{v}}\right)^{\prime},\\ &-\sigma\left(\frac{R}{\gamma-1}\tilde{\theta}+\frac{\tilde{u}^{2}}{2}\right)^{\prime}+(p(\tilde{v},\tilde{\theta})\tilde{u})^{\prime}=\Bigg{(}\kappa\frac{\tilde{\theta}^{\prime}}{\tilde{v}}\Bigg{)}^{\prime}+\Bigg{(}\mu\frac{\tilde{u}\tilde{u}^{\prime}}{\tilde{v}}\Bigg{)}^{\prime},\\ &\lim_{\xi\to\pm\infty}(\tilde{v},\tilde{u},\tilde{\theta})(\xi)=(v_{\pm},u_{\pm},\theta_{\pm}).\end{aligned}\right.\qquad\left({}^{\prime}=\frac{d}{d\xi}\right)

Here, if v>v+v_{-}>v_{+}, (v~,u~,θ~\tilde{v},\tilde{u},\tilde{\theta})(xσtx-\sigma t) is a 1-shock wave with velocity σ=p+pv+v\sigma=-\sqrt{-\frac{p_{+}-p_{-}}{v_{+}-v_{-}}}, where p±:=p(v±,θ±)p_{\pm}:=p(v_{\pm},\theta_{\pm}). If v<v+v_{-}<v_{+}, it becomes a 3-shock wave with σ=p+pv+v\sigma=\sqrt{-\frac{p_{+}-p_{-}}{v_{+}-v_{-}}}.

1.1. Main results

Our first result is for the existence and uniqueness of traveling wave solutions of (LABEL:main) as monotone profiles satisfying (LABEL:shock_0).

Theorem 1.1.

(Existence and Uniqueness) For a given left-end state (v,u,θ)+××+(v_{-},u_{-},\theta_{-})\in{\mathbb{R}}^{+}\times{\mathbb{R}}\times{\mathbb{R}}^{+}, there exists a constant ε0>0\varepsilon_{0}>0 such that for any right-end state (v+,u+,θ+)+××+(v_{+},u_{+},\theta_{+})\in{\mathbb{R}}^{+}\times{\mathbb{R}}\times{\mathbb{R}}^{+} satisfying (LABEL:end-con) and ε:=|vv+|<ε0\varepsilon:=\left|v_{-}-v_{+}\right|<\varepsilon_{0}, there is a unique traveling wave solution (v~ε,u~ε,θ~ε):+××+({\tilde{v}_{\varepsilon}},{\tilde{u}_{\varepsilon}},{\tilde{\theta}_{\varepsilon}})\colon{\mathbb{R}}\to{\mathbb{R}}^{+}\times{\mathbb{R}}\times{\mathbb{R}}^{+} to (LABEL:main), as a monotone profile satisfying (LABEL:shock_0), which more precisely satisfies (1.13) when v>v+v_{-}>v_{+}, or (1.14) when v<v+v_{-}<v_{+}.

The second result provides quantitative estimates of traveling wave solutions. Especially, we present estimates for the ratio between the components v~,u~,θ~\tilde{v},\tilde{u},\tilde{\theta} of the traveling wave, and estimates for the exponential tail of the wave. This could be useful in stability estimates for traveling waves.

Theorem 1.2.

(Estimates) For a given left-end state (v,u,θ)+××+(v_{-},u_{-},\theta_{-})\in{\mathbb{R}}^{+}\times{\mathbb{R}}\times{\mathbb{R}}^{+}, there exist positive constants ε0\varepsilon_{0}, CC, and C1C_{1} such that for any right-end state (v+,u+,θ+)+××+(v_{+},u_{+},\theta_{+})\in{\mathbb{R}}^{+}\times{\mathbb{R}}\times{\mathbb{R}}^{+} satisfying (LABEL:end-con) and ε:=|vv+|<ε0\varepsilon:=\left|v_{-}-v_{+}\right|<\varepsilon_{0}, the following holds.
Let (v~ε,u~ε,θ~ε)({\tilde{v}_{\varepsilon}},{\tilde{u}_{\varepsilon}},{\tilde{\theta}_{\varepsilon}}) be the monotone solution to (LABEL:shock_0) with v~ε(0)=v+v+2{\tilde{v}_{\varepsilon}}(0)=\frac{v_{-}+v_{+}}{2}.
Then,

(1.7) |(v~ε(ξ)v,u~ε(ξ)u,θ~ε(ξ)θ)|CεeC1ε|ξ| for all ξ0,|(v~ε(ξ)v+,u~ε(ξ)u+,θ~ε(ξ)θ+)|CεeC1ε|ξ| for all ξ0,\displaystyle\begin{aligned} &\left|({\tilde{v}_{\varepsilon}}(\xi)-v_{-},{\tilde{u}_{\varepsilon}}(\xi)-u_{-},{\tilde{\theta}_{\varepsilon}}(\xi)-\theta_{-})\right|\leq C\varepsilon e^{-C_{1}\varepsilon\left|\xi\right|}\text{ for all }\xi\leq 0,\\ &\left|({\tilde{v}_{\varepsilon}}(\xi)-v_{+},{\tilde{u}_{\varepsilon}}(\xi)-u_{+},{\tilde{\theta}_{\varepsilon}}(\xi)-\theta_{+})\right|\leq C\varepsilon e^{-C_{1}\varepsilon\left|\xi\right|}\text{ for all }\xi\geq 0,\\ \end{aligned}
(1.8) |(v~ε(ξ),u~ε(ξ),θ~ε(ξ))|Cε2eC1ε|ξ| for all ξ,\displaystyle\left|(\tilde{v}^{\prime}_{\varepsilon}(\xi),\tilde{u}^{\prime}_{\varepsilon}(\xi),\tilde{\theta}^{\prime}_{\varepsilon}(\xi))\right|\leq C\varepsilon^{2}e^{-C_{1}\varepsilon\left|\xi\right|}\text{ for all }\xi\in{\mathbb{R}},
(1.9) |(v~ε′′(ξ),u~ε′′(ξ),θ~ε′′(ξ))|Cε|(v~ε(ξ),u~ε(ξ),θ~ε(ξ))| for all ξ.\displaystyle\left|(\tilde{v}^{\prime\prime}_{\varepsilon}(\xi),\tilde{u}^{\prime\prime}_{\varepsilon}(\xi),\tilde{\theta}^{\prime\prime}_{\varepsilon}(\xi))\right|\leq C\varepsilon\left|(\tilde{v}^{\prime}_{\varepsilon}(\xi),\tilde{u}^{\prime}_{\varepsilon}(\xi),\tilde{\theta}^{\prime}_{\varepsilon}(\xi))\right|\text{ for all }\xi\in{\mathbb{R}}.

It also holds that |v~ε||u~ε||θ~ε|\left|{\tilde{v}_{\varepsilon}}^{\prime}\right|\sim\left|{\tilde{u}_{\varepsilon}}^{\prime}\right|\sim\left|{\tilde{\theta}_{\varepsilon}}^{\prime}\right| for all ξ\xi\in{\mathbb{R}}. More explicitly, we have the following:

(1.10) |u~ε(ξ)+σv~ε(ξ)|\displaystyle\left|\tilde{u}^{\prime}_{\varepsilon}(\xi)+\sigma_{*}\tilde{v}^{\prime}_{\varepsilon}(\xi)\right| Cε|v~ε(ξ)|\displaystyle\leq C\varepsilon\left|\tilde{v}^{\prime}_{\varepsilon}(\xi)\right|
(1.11) |θ~ε(ξ)+(γ1)pRv~ε(ξ)|\displaystyle\left|\tilde{\theta}^{\prime}_{\varepsilon}(\xi)+\frac{(\gamma-1)p_{-}}{R}\tilde{v}^{\prime}_{\varepsilon}(\xi)\right| Cε|v~ε(ξ)|\displaystyle\leq C\varepsilon\left|\tilde{v}^{\prime}_{\varepsilon}(\xi)\right|

which satisfies

(1.12) |σεσ|Cε,\left|\sigma_{\varepsilon}-\sigma_{*}\right|\leq C\varepsilon,

where

σ=σε:=p+pv+vand σ:=γpv=γRθv2,\sigma=\sigma_{\varepsilon}:=-\sqrt{-\frac{p_{+}-p_{-}}{v_{+}-v_{-}}}\,\,\text{and }\sigma_{*}:=-\sqrt{\frac{\gamma p_{-}}{v_{-}}}=-\frac{\sqrt{\gamma R\theta_{-}}}{v_{-}^{2}},

or

σ=σε:=p+pv+vand σ:=γpv=γRθv2.\sigma=\sigma_{\varepsilon}:=\sqrt{-\frac{p_{+}-p_{-}}{v_{+}-v_{-}}}\,\,\text{and }\sigma_{*}:=\sqrt{\frac{\gamma p_{-}}{v_{-}}}=\frac{\sqrt{\gamma R\theta_{-}}}{v_{-}^{2}}.

In addition, if we consider the 1-shock, i.e., σε<0\sigma_{\varepsilon}<0, then we have

(1.13) v~ε<0,u~ε<0,θ~ε>0 for all ξ,\tilde{v}^{\prime}_{\varepsilon}<0,\quad\tilde{u}^{\prime}_{\varepsilon}<0,\quad\tilde{\theta}^{\prime}_{\varepsilon}>0\text{ for all }\xi\in{\mathbb{R}},

and for the 3-shock case, i.e., σε>0\sigma_{\varepsilon}>0,

(1.14) v~ε>0,u~ε<0,θ~ε<0 for all ξ.\tilde{v}^{\prime}_{\varepsilon}>0,\quad\tilde{u}^{\prime}_{\varepsilon}<0,\quad\tilde{\theta}^{\prime}_{\varepsilon}<0\text{ for all }\xi\in{\mathbb{R}}.

Furthermore, the following estimate holds:

(1.15) C1(v+v~ε)(v~εv)v~εC(v+v~ε)(v~εv).C^{-1}(v_{+}-{\tilde{v}_{\varepsilon}})({\tilde{v}_{\varepsilon}}-v_{-})\leq{\tilde{v}_{\varepsilon}}^{\prime}\leq C(v_{+}-{\tilde{v}_{\varepsilon}})({\tilde{v}_{\varepsilon}}-v_{-}).

Note that it suffices to prove the Theorem 1.2 for 3-shocks. Indeed, the results for the 1-shocks can be obatained by the change of variables xx,uu,σεσεx\mapsto-x,u\mapsto-u,\sigma_{\varepsilon}\mapsto-\sigma_{\varepsilon} and σσ\sigma_{*}\mapsto-\sigma_{*}. Thus, in the sequel, we only consider the 3-shock case, i.e.,

σε=p+pv+v>0and σγpv=γRθv2\sigma_{\varepsilon}=\sqrt{-\frac{p_{+}-p_{-}}{v_{+}-v_{-}}}>0\,\,\text{and }\sigma_{*}\coloneqq\sqrt{\frac{\gamma p_{-}}{v_{-}}}=\frac{\sqrt{\gamma R\theta_{-}}}{v_{-}^{2}}
Remark 1.1.

The estimates of Theorem 1.2 would be useful in the stability estimates of traveling wave solutions, based on the method of aa-contraction with shifts, as in previous results [7, 20, 23, 24, 25, 26, 27]. Recently, the contraction (up to shift) of large perturbations from the traveling waves was proved in [15], for which the estimates of Theorem 1.2 is crucially used.

2. Main ideas and methodology

2.1. Main ideas for the proofs

First of all, we may integrate the ODE system (LABEL:shock_0) over (,ξ](-\infty,\xi] and use the second equation of it, to find the system of ODEs of first order:

(2.1) {σε(v~v)(u~u)=τv~v~,σε(u~u)+(p~p)=μu~v~,σε(Rγ1(θ~θ)+12(u~2u2))+(p~u~pu)=κθ~v~.\left\{\begin{aligned} &-\sigma_{\varepsilon}(\tilde{v}-v_{-})-(\tilde{u}-u_{-})=\tau\frac{\tilde{v}^{\prime}}{\tilde{v}},\\ &-\sigma_{\varepsilon}(\tilde{u}-u_{-})+(\tilde{p}-p_{-})=\mu\frac{\tilde{u}^{\prime}}{\tilde{v}},\\ &-\sigma_{\varepsilon}\left(\frac{R}{\gamma-1}(\tilde{\theta}-\theta_{-})+\frac{1}{2}(\tilde{u}^{2}-u_{-}^{2})\right)+(\tilde{p}\tilde{u}-p_{-}u_{-})=\kappa\frac{\tilde{\theta}^{\prime}}{\tilde{v}}.\end{aligned}\right.

For the existence theory of solutions to the system (2.1), it might be enough to use some typical approach such as Cauchy-Lipschitz theory. However, in order to get the ratio estimates as in (1.10)-(1.11), we need to first show that the first derivative of wave dominates its second derivative as in (1.9), since we may verify (1.10) from the second order equation (LABEL:shock_0)1\eqref{shock_0}_{1}, and (1.11) from the following system of two equations on v~,θ~\tilde{v},\tilde{\theta} variables only:

(2.2) {σε2(v~v)+τσεv~v~+(p~p)=μ(σεv~v~τv~′′v~2+τ(v~)2v~3),σε(Rγ1(θ~θ)12(σε(v~v)+τv~v~)2)p(σε(v~v)+τv~v~)=κθ~v~,\left\{\begin{aligned} &\sigma_{\varepsilon}^{2}(\tilde{v}-v_{-})+\tau\sigma_{\varepsilon}\frac{\tilde{v}^{\prime}}{\tilde{v}}+(\tilde{p}-p_{-})=\mu\left(-\sigma_{\varepsilon}\frac{\tilde{v}^{\prime}}{\tilde{v}}-\tau\frac{\tilde{v}^{\prime\prime}}{\tilde{v}^{2}}+\tau\frac{(\tilde{v}^{\prime})^{2}}{\tilde{v}^{3}}\right),\\ &-\sigma_{\varepsilon}\left(\frac{R}{\gamma-1}(\tilde{\theta}-\theta_{-})-\frac{1}{2}\left(\sigma_{\varepsilon}(\tilde{v}-v_{-})+\tau\frac{\tilde{v}^{\prime}}{\tilde{v}}\right)^{2}\right)-p_{-}\left(\sigma_{\varepsilon}(\tilde{v}-v_{-})+\tau\frac{\tilde{v}^{\prime}}{\tilde{v}}\right)=\kappa\frac{\tilde{\theta}^{\prime}}{\tilde{v}},\end{aligned}\right.

where the above system can be derived from (2.1), and the second order term v~′′\tilde{v}^{\prime\prime} is due to the right-hand side of (2.1)1\eqref{temsys}_{1}. This is very different from the case for the Navier-Stokes-Fourier system that has a simpler structure as in [27].

Since we do not know the monotonicity of solutions prior to the existence of them, the proof for the desired estimates of Theorem 1.2 is not obvious. Thus, for the proofs of the main results, especially to verify the quantitative estimates as desired, we will employ a geometric approach to the singular perturbation system of ODEs associated to (2.1). To derive the singular perturbation system, we would consider the third order ODE for v~\tilde{v} variable only, like (3.19) that could be derived from (2.2). We may (formally) introduce a new variable v(z)v(z) defined by v~(ξ)=εv(εξ)+v\tilde{v}(\xi)=\varepsilon v(\varepsilon\xi)+v_{-}, where the scale z:=εξz:=\varepsilon\xi is slow while ξ\xi is fast. Then, by setting w0vw_{0}\coloneqq v, w1vw_{1}\coloneqq v^{\prime}, w2εv′′w_{2}\coloneqq\varepsilon v^{\prime\prime}, and so w0=w1w^{\prime}_{0}=w_{1}, εw1=w2\varepsilon w_{1}^{\prime}=w_{2} and εw2=ε2v′′′\varepsilon w_{2}^{\prime}=\varepsilon^{2}v^{\prime\prime\prime}, we may have the singular perturbation system (3.1), which is a starting point in our analysis. We may consider a locally invariant manifold near a critical manifold of (3.1), where the locally invariant manifold could be explicitly described based on Fenichel’s first theorem.

2.2. Fenichel’s First Theorem and locally invariant manifolds

We here present the main concepts for the geometric approach. Consider the following system of ODEs with a small parameter ε\varepsilon:

(2.3) {εx=f(x,y,ε)y=g(x,y,ε),\left\{\begin{aligned} \varepsilon x^{\prime}&=f(x,y,\varepsilon)\\ y^{\prime}&=g(x,y,\varepsilon),\end{aligned}\right.

where xnx\in{\mathbb{R}}^{n}, yly\in{\mathbb{R}}^{l}. Here, we assume ff and gg to be smooth on a set U×IU\times I, where Un+lU\subset{\mathbb{R}}^{n+l} is open and II is an open interval containing 0. Following [22], we say an ll-dimensional manifold M0n+lM_{0}\subset{\mathbb{R}}^{n+l} is a critical manifold of the system (2.3) if each (x,y)M0(x,y)\in M_{0} satisfies that f(x,y,0)=0f(x,y,0)=0. To state Fenichel’s first theorem, we need the following definitions:

Definition 2.1.

We say that the manifold M0M_{0} is normally hyperbolic relative to (2.3) if the n×nn\times n matrix

Dxf(x,y,ε)|ε=0\left.D_{x}f(x,y,\varepsilon)\right|_{\varepsilon=0}

has nn eigenvalues (counting multiplicity) with nonzero real part for each (x,y)M0(x,y)\in M_{0}.

Fenichel’s first theorem proves the existence of a manifold on which the flow of the solution of (2.3) is confined. To be precise, we define the concept of a locally invariant manifold:

Definition 2.2.

A set Mn+lM\subseteq{\mathbb{R}}^{n+l} is locally invariant under the flow from (2.3) if there is an open set Vn+lV\subseteq{\mathbb{R}}^{n+l} containing MM such that for any x0Mx_{0}\in M and any TT\in{\mathbb{R}}, the trajectory xx starting at x0x_{0} with x([0,T])Vx([0,T])\subset V also satisfies x([0,T])Mx([0,T])\subset M. (When T<0T<0, replace [0,T][0,T] with [T,0][T,0]).

Note that, if M0M_{0} is normally hyperbolic, we can express it locally as a graph of xx in terms of yy thanks to the implicit function theorem. However, in the following discussion, we only deal with the case when M0M_{0} can be globally expressed as a graph of xx in terms of yy on a compact domain KlK\subset{\mathbb{R}}^{l}, and in this case, Fenichel’s first theorem can be stated as follows.

Proposition 2.1.

[22] Consider the system (2.3) with normally hyperbolic critical manifold M0M_{0}. Suppose that there is a smooth function h:Knh\colon K\to{\mathbb{R}}^{n} of which graph is contained in M0M_{0}, i.e.,

{(x,y)x=h(y),yK}M0\{(x,y)\mid x=h(y),y\in K\}\subset M_{0}

where KlK\subset{\mathbb{R}}^{l} is a compact and simply connected smooth set. Then, there exists ε0>0\varepsilon_{0}>0 such that for each ε\varepsilon with |ε|<ε0\left|\varepsilon\right|<\varepsilon_{0}, there is a function hε:Knh^{\varepsilon}\colon K\to{\mathbb{R}}^{n} which satisfies the followings: for each ε0\varepsilon\neq 0, the graph

{(x,y)x=hε(y),yK}\{(x,y)\mid x=h^{\varepsilon}(y),y\in K\}

is contained in an ll-dimensional manifold MεM_{\varepsilon} which is locally invariant under the flow of (2.3). Moreover, hεh^{\varepsilon} can be taken to be CrC^{r} for each r<r<\infty, jointly in yy and ε\varepsilon and h=h0h=h^{0} on KK where h0h^{0} denotes hεh^{\varepsilon} at ε=0\varepsilon=0.

Remark 2.1.

Since the family of functions hεh^{\varepsilon} is regular in ε\varepsilon as well, a consequence of Proposition 2.1 is that there exists a function ss with two variables ε\varepsilon and yy which satisfies

hε(y)=h(y)+εs(y,ε).h^{\varepsilon}(y)=h(y)+\varepsilon s(y,\varepsilon).

Note that the function ss has the same regularity with hεh^{\varepsilon}; i.e. ss is a CrC^{r}-function for each r<r<\infty, jointly in yy and ε\varepsilon.

3. Proof for existence and uniqueness

In this section, we present the proof of Theorem 1.1. Instead of proving the existence of the solution to (LABEL:shock_0) directly, we consider an alternative ODE system (3.1) to which Proposition 2.1 can be applied. We will first show the existence and uniqueness of the solution to the ODE system (3.1), and then to the system (LABEL:shock_0) through a proper transformation.

For any fixed parameter ε\varepsilon\in\mathbb{R} small enough (to be determined later), we consider the system of ODEs :

(3.1) {w0(z)=w1(z),εw1(z)=w2(z),εw2(z)=f(w0(z),w1(z),w2(z),ε),\left\{\begin{aligned} w_{0}^{\prime}(z)&=w_{1}(z),\\ \varepsilon w_{1}^{\prime}(z)&=w_{2}(z),\\ \varepsilon w_{2}^{\prime}(z)&=f(w_{0}(z),w_{1}(z),w_{2}(z),\varepsilon),\end{aligned}\right.

where zz\in{\mathbb{R}} and

(3.2) f(w0,w1,w2,ε)\displaystyle f(w_{0},w_{1},w_{2},\varepsilon)
1ε(1μτκ)[σ¯ε2w0Rσ¯ε(εw0+v)3γ1+R(εw0+v)2(σ¯εpγ1w0+σ¯εpw0)]=:𝒥(ε)\displaystyle\quad\coloneqq\frac{1}{\varepsilon}\left(\frac{1}{\mu\tau\kappa}\right)\underbrace{\left[-\overline{\sigma}_{\varepsilon}^{2}w_{0}\frac{R\overline{\sigma}_{\varepsilon}(\varepsilon w_{0}+v_{-})^{3}}{\gamma-1}+R(\varepsilon w_{0}+v_{-})^{2}\left(\frac{\overline{\sigma}_{\varepsilon}p_{-}}{\gamma-1}w_{0}+\overline{\sigma}_{\varepsilon}p_{-}w_{0}\right)\right]}_{=:\mathcal{J}(\varepsilon)}
+Rσ¯ε(εw0+v)3(γ1)κ(σ¯ε(μ+τμτ)w1εw0+vw2(εw0+v)2+ε2w12(εw0+v)3)\displaystyle\quad+\frac{R\overline{\sigma}_{\varepsilon}(\varepsilon w_{0}+v_{-})^{3}}{(\gamma-1)\kappa}\left(-\overline{\sigma}_{\varepsilon}\left(\frac{\mu+\tau}{\mu\tau}\right)\frac{w_{1}}{\varepsilon w_{0}+v_{-}}-\frac{w_{2}}{(\varepsilon w_{0}+v_{-})^{2}}+\varepsilon^{2}\frac{w_{1}^{2}}{(\varepsilon w_{0}+v_{-})^{3}}\right)
+R(εw0+v)2μτκ(σ¯ε32w02εσ¯ε2τw0w1εw0+vε2τ2σ¯ε2w12(εw0+v)2+pτw1εw0+v)\displaystyle\quad+\frac{R(\varepsilon w_{0}+v_{-})^{2}}{\mu\tau\kappa}\left(-\frac{\overline{\sigma}_{\varepsilon}^{3}}{2}w_{0}^{2}-\varepsilon\overline{\sigma}_{\varepsilon}^{2}\tau\frac{w_{0}w_{1}}{\varepsilon w_{0}+v_{-}}-\varepsilon^{2}\tau^{2}\frac{\overline{\sigma}_{\varepsilon}}{2}\frac{w_{1}^{2}}{(\varepsilon w_{0}+v_{-})^{2}}+p_{-}\tau\frac{w_{1}}{\varepsilon w_{0}+v_{-}}\right)
+(εw0+v)(1μτ((pσ¯ε2v2εσ¯ε2w0)w1σ¯ε(μ+τ)w2)\displaystyle\quad+(\varepsilon w_{0}+v_{-})\Bigg{(}\frac{1}{\mu\tau}\left((p_{-}-\overline{\sigma}_{\varepsilon}^{2}v_{-}-2\varepsilon\overline{\sigma}_{\varepsilon}^{2}w_{0})w_{1}-\overline{\sigma}_{\varepsilon}(\mu+\tau)w_{2}\right)
+ε23w1w2(εw0+v)2ε42w13(εw0+v)3).\displaystyle\quad\hskip 79.66771pt\qquad+\varepsilon^{2}\frac{3w_{1}w_{2}}{(\varepsilon w_{0}+v_{-})^{2}}-\varepsilon^{4}\frac{2w_{1}^{3}}{(\varepsilon w_{0}+v_{-})^{3}}\Bigg{)}.

Here, σ¯ε\overline{\sigma}_{\varepsilon} is defined by

(3.3) σ¯εγpv+γ+12ε.\overline{\sigma}_{\varepsilon}\coloneqq\frac{\sqrt{\gamma p_{-}}}{\sqrt{v_{-}+\frac{\gamma+1}{2}\varepsilon}}.

It is worth mentioning that the function f(w0,w1,w2,ε)f(w_{0},w_{1},w_{2},\varepsilon) is smooth for any ε\varepsilon in some open interval containing 0, and hence so is the system (3.1). Indeed, this can be verified by the argument below. We observe that 𝒥(ε)\mathcal{J}(\varepsilon) can be rewritten as follows:

𝒥(ε)\displaystyle\mathcal{J}(\varepsilon) =Rσ¯εγ1(εw0+v)2w0(γpσ¯ε2(εw0+v))\displaystyle=\frac{R\overline{\sigma}_{\varepsilon}}{\gamma-1}(\varepsilon w_{0}+v_{-})^{2}w_{0}(\gamma p_{-}-\overline{\sigma}_{\varepsilon}^{2}(\varepsilon w_{0}+v_{-}))
=Rσ¯εγ1v2w0(γpσ¯ε2v)+εRσ¯εγ1(εw02+2w0v)w0(γpσ¯ε2v)\displaystyle=\frac{R\overline{\sigma}_{\varepsilon}}{\gamma-1}v_{-}^{2}w_{0}(\gamma p_{-}-\overline{\sigma}_{\varepsilon}^{2}v_{-})+\varepsilon\frac{R\overline{\sigma}_{\varepsilon}}{\gamma-1}(\varepsilon w_{0}^{2}+2w_{0}v_{-})w_{0}(\gamma p_{-}-\overline{\sigma}_{\varepsilon}^{2}v_{-})
εRσ¯εγ1(εw0+v)2σ¯ε2w02.\displaystyle\qquad-\varepsilon\frac{R\overline{\sigma}_{\varepsilon}}{\gamma-1}(\varepsilon w_{0}+v_{-})^{2}\overline{\sigma}_{\varepsilon}^{2}w_{0}^{2}.

From (3.3), we have

(3.4) γpσ¯ε2v=γpγ+12εv+γ+12ε.\gamma p_{-}-\overline{\sigma}_{\varepsilon}^{2}v_{-}=\gamma p_{-}\frac{\frac{\gamma+1}{2}\varepsilon}{v_{-}+\frac{\gamma+1}{2}\varepsilon}.

Hence, it follows that

1ε𝒥(ε)\displaystyle\frac{1}{\varepsilon}\mathcal{J}(\varepsilon) =Rσ¯εγ1v2w0γpγ+12v+γ+12ε+Rσ¯εγ1(εw02+2w0v)w0(γpσ¯ε2v)\displaystyle=\frac{R\overline{\sigma}_{\varepsilon}}{\gamma-1}v_{-}^{2}w_{0}\gamma p_{-}\frac{\frac{\gamma+1}{2}}{v_{-}+\frac{\gamma+1}{2}\varepsilon}+\frac{R\overline{\sigma}_{\varepsilon}}{\gamma-1}(\varepsilon w_{0}^{2}+2w_{0}v_{-})w_{0}(\gamma p_{-}-\overline{\sigma}_{\varepsilon}^{2}v_{-})
Rσ¯εγ1(εw0+v)2σ¯ε2w02,\displaystyle\qquad-\frac{R\overline{\sigma}_{\varepsilon}}{\gamma-1}(\varepsilon w_{0}+v_{-})^{2}\overline{\sigma}_{\varepsilon}^{2}w_{0}^{2},

which is a smooth function near ε=0\varepsilon=0.

We then find critical points of the system (3.1) for ε0\varepsilon\neq 0. If (w0,w1,w2)(w_{0},w_{1},w_{2}) is a critical point of the system, then from (3.2) together with the relation (3.3), it holds that

0\displaystyle 0 =σ¯ε(εw0+v)γ1(σ¯ε2w0)+σ¯εpγ1w0+σ¯εpw0εσ¯ε32w02\displaystyle=\frac{\overline{\sigma}_{\varepsilon}(\varepsilon w_{0}+v_{-})}{\gamma-1}(-\overline{\sigma}_{\varepsilon}^{2}w_{0})+\frac{\overline{\sigma}_{\varepsilon}p_{-}}{\gamma-1}w_{0}+\overline{\sigma}_{\varepsilon}p_{-}w_{0}-\varepsilon\frac{\overline{\sigma}_{\varepsilon}^{3}}{2}w_{0}^{2}
=σ¯εw0[σ¯ε2(εw0+vγ1+εw02)+γpγ1]\displaystyle=\overline{\sigma}_{\varepsilon}w_{0}\left[-\overline{\sigma}_{\varepsilon}^{2}\left(\frac{\varepsilon w_{0}+v_{-}}{\gamma-1}+\frac{\varepsilon w_{0}}{2}\right)+\frac{\gamma p_{-}}{\gamma-1}\right]
=σ¯ε3w0[(εw0+vγ1+εw02)+v+εγ1+ε2]=σ¯ε3ε(1γ1+12)w0(1w0),\displaystyle=\overline{\sigma}_{\varepsilon}^{3}w_{0}\left[-\left(\frac{\varepsilon w_{0}+v_{-}}{\gamma-1}+\frac{\varepsilon w_{0}}{2}\right)+\frac{v_{-}+\varepsilon}{\gamma-1}+\frac{\varepsilon}{2}\right]=\overline{\sigma}_{\varepsilon}^{3}\varepsilon\left(\frac{1}{\gamma-1}+\frac{1}{2}\right)w_{0}(1-w_{0}),

which proves that there are only two critical points (0,0,0),(1,0,0)(0,0,0),(1,0,0) for any ε0\varepsilon\neq 0.

3.1. Proof of Theorem 1.1: Existence

In this subsection, we prove the existence part of Theorem 1.1. The proof consists of two steps: we first prove that the system (3.1) has a solution connecting the two critical points (0,0,0)(0,0,0) and (1,0,0)(1,0,0) by using Proposition 2.1, and then we demonstrate that this gives a desired solution to the original system (LABEL:shock_0).

First of all, we calculate the critical manifold of the system (3.1); we have w2=0w_{2}=0 and f(w0,w1,w2,0)=0f(w_{0},w_{1},w_{2},0)=0. Equivalently, we have w2=0w_{2}=0 and

0\displaystyle 0 =f(w0,w1,0,0)\displaystyle=f(w_{0},w_{1},0,0)
=1μτκ(Rσγ1vγpγ+12w0Rσγ1vγpw02)\displaystyle=\frac{1}{\mu\tau\kappa}\left(\frac{R\sigma_{*}}{\gamma-1}v_{-}\gamma p_{-}\frac{\gamma+1}{2}w_{0}-\frac{R\sigma_{*}}{\gamma-1}v_{-}\gamma p_{-}w_{0}^{2}\right)
Rγpv(γ1)κμ+τμτw1Rσvγpμτκ12w02+Rvμτκpτw11μτ(γ1)vpw1\displaystyle\qquad-\frac{R\gamma p_{-}v_{-}}{(\gamma-1)\kappa}\frac{\mu+\tau}{\mu\tau}w_{1}-\frac{R\sigma_{*}v_{-}\gamma p_{-}}{\mu\tau\kappa}\frac{1}{2}w_{0}^{2}+\frac{Rv_{-}}{\mu\tau\kappa}p_{-}\tau w_{1}-\frac{1}{\mu\tau}(\gamma-1)v_{-}p_{-}w_{1}

so that

(3.5) w1=RγσRτ+Rγμ+(γ1)2κγ+12(w0w02).w_{1}=\frac{R\gamma\sigma_{*}}{R\tau+R\gamma\mu+(\gamma-1)^{2}\kappa}\frac{\gamma+1}{2}(w_{0}-w_{0}^{2}).

Thus, the critical manifold M0M_{0} of (3.1) can be parametrized by w0w_{0} as follows:

M0={(w0,w1,w2)w1=A(w0w02),w2=0}M_{0}=\{(w_{0},w_{1},w_{2})\mid w_{1}=A(w_{0}-w_{0}^{2}),\,w_{2}=0\}

which forms a curve in 33-dimensional space. Here,

ARγσRτ+Rγμ+(γ1)2κγ+12A\coloneqq\frac{R\gamma\sigma_{*}}{R\tau+R\gamma\mu+(\gamma-1)^{2}\kappa}\frac{\gamma+1}{2}

is a constant depending only on the type of gas.

To use Proposition 2.1, it is required to verify that M0M_{0} is normally hyperbolic relative to the system (3.1). Considering the definition of normally hyperbolicity, we need to compute

(w2w1w2w2w1f(w0,w1,w2,ε)w2f(w0,w1,w2,ε))|ε=0=(011μτκvpγ1(Rτ+Rγμ+(γ1)2κ)σvμτκ(Rγ1μτ+(μ+τ)κ)),\left.\begin{pmatrix}\frac{\partial w_{2}}{\partial w_{1}}&\frac{\partial w_{2}}{\partial w_{2}}\\ \frac{\partial}{\partial w_{1}}f(w_{0},w_{1},w_{2},\varepsilon)&\frac{\partial}{\partial w_{2}}f(w_{0},w_{1},w_{2},\varepsilon)\end{pmatrix}\right|_{\varepsilon=0}\\ =\begin{pmatrix}0&1\\ -\frac{1}{\mu\tau\kappa}\frac{v_{-}p_{-}}{\gamma-1}(R\tau+R\gamma\mu+(\gamma-1)^{2}\kappa)&-\frac{\sigma_{*}v_{-}}{\mu\tau\kappa}\left(\frac{R}{\gamma-1}\mu\tau+(\mu+\tau)\kappa\right)\end{pmatrix},

which is followed by (3.2). Then, from simple calculations, we deduce that the matrix above has two real eigenvalues and both of them are negative, which proves that M0M_{0} is normally hyperbolic.

We now consider a closed interval KK containing [0,1][0,1]. From Proposition 2.1 and Remark 2.1, for each small ε\varepsilon we obtain an 1-dimensional locally invariant manifold MεM_{\varepsilon} and smooth functions s1s_{1} and s2s_{2} defined on KK, which satisfies

(3.6) {(w0,w1,w2)w0K,w1=A(w0w02)+εs1(w0,ε),w2=εs2(w0,ε)}Mε\{(w_{0},w_{1},w_{2})\mid w_{0}\in K,\,w_{1}=A(w_{0}-w_{0}^{2})+\varepsilon s_{1}(w_{0},\varepsilon),\,w_{2}=\varepsilon s_{2}(w_{0},\varepsilon)\}\subset M_{\varepsilon}

where the functions s1s_{1} and s2s_{2} are sufficiently regular jointly in w0w_{0} and ε\varepsilon, in particular, C3C^{3}. Step 1: We will show that the system (3.1) has a solution connecting the two critical points (0,0,0)(0,0,0) and (1,0,0)(1,0,0). Here, we will prove the existence of a solution defined on \mathbb{R} to the system (3.1) passing through a point

(w0,w1,w2)(0)=(1/2,A/4+εs1(1/2,ε),εs2(1/2,ε))(w_{0},w_{1},w_{2})(0)=(1/2,A/4+\varepsilon s_{1}(1/2,\varepsilon),\varepsilon s_{2}(1/2,\varepsilon))

which stays on MεM_{\varepsilon} and connects the two points (0,0,0)(0,0,0) and (1,0,0)(1,0,0).

Since Fε(w0):=A(w0w02)+εs1(w0,ε)>0F_{\varepsilon}(w_{0}):=A(w_{0}-w_{0}^{2})+\varepsilon s_{1}(w_{0},\varepsilon)>0 at w0=1/2w_{0}=1/2 for sufficiently small ε\varepsilon and Fε(w0)F_{\varepsilon}(w_{0})\to-\infty as |w0|\left|w_{0}\right|\to\infty, there exist a,ba,b\in{\mathbb{R}} with a<1/2<ba<1/2<b such that Fε(a)=Fε(b)=0F_{\varepsilon}(a)=F_{\varepsilon}(b)=0 and Fε>0F_{\varepsilon}>0 on (a,b)(a,b). Thus, by the succeeding lemma, we have a unique solution w¯0:\bar{w}_{0}\colon{\mathbb{R}}\to{\mathbb{R}} of the following ODE

(3.7) w0(z)=Fε(w0(z))=A(w0(z)w0(z)2)+εs1(w0(z),ε)w^{\prime}_{0}(z)=F_{\varepsilon}(w_{0}(z))=A(w_{0}(z)-w_{0}(z)^{2})+\varepsilon s_{1}(w_{0}(z),\varepsilon)

subject to the initial datum

(3.8) w0(0)=1/2,w_{0}(0)=1/2,

and the solution is increasing and converges to aa and bb as z±z\to\pm\infty.

Lemma 3.1.

Let B𝒞1([a,b])B\in\mathcal{C}^{1}([a,b]) with B(a)=B(b)=0B(a)=B(b)=0, B>0B>0 on (a,b)(a,b), and fix f0(a,b)f_{0}\in(a,b). Then, there is a unique 𝒞1\mathcal{C}^{1} function f:[a,b]f\colon{\mathbb{R}}\to[a,b] satisfying f(z)=B(f(z))f^{\prime}(z)=B(f(z)) and f(0)=f0f(0)=f_{0}, which is increasing and f(z)f(z) converges to aa and bb as z±z\to\pm\infty.

This lemma is a classical result in the theory of ordinary differential equations, specifically in the context of the autonomous case. For the unique profile w¯0\bar{w}_{0}, we let

(3.9) w¯1A(w¯0w¯02)+εs1(w¯0,ε),w¯2εs2(w¯0,ε)on ,\bar{w}_{1}\coloneqq A(\bar{w}_{0}-\bar{w}_{0}^{2})+\varepsilon s_{1}(\bar{w}_{0},\varepsilon),\quad\bar{w}_{2}\coloneqq\varepsilon s_{2}(\bar{w}_{0},\varepsilon)\quad\mbox{on }{\mathbb{R}},

and we will prove that the function w¯:-(w¯0,w¯1,w¯2)\bar{w}\coloneq(\bar{w}_{0},\bar{w}_{1},\bar{w}_{2}) satisfies (3.1).

For any fixed z0z_{0}\in{\mathbb{R}}, the point w¯(z0)\bar{w}(z_{0}) is on the manifold MεM_{\varepsilon}. From the Cauchy-Lipschitz theorem, we have a unique Lipschitz solution w~(z)=(w~0,w~1,w~2)(z)\tilde{w}(z)=(\tilde{w}_{0},\tilde{w}_{1},\tilde{w}_{2})(z) (locally defined near z0z_{0}) to the system (3.1) passing through w¯(z0)\bar{w}(z_{0}) at z0z_{0}, i.e., w~(z0)=w¯(z0)\tilde{w}(z_{0})=\bar{w}(z_{0}).
Then, since w~(z0)=w¯(z0)Mε\tilde{w}(z_{0})=\bar{w}(z_{0})\in M_{\varepsilon} by (3.9) and the local solution w~(z)\tilde{w}(z) is continuous near z0z_{0}, it holds from Proposition 2.1 that the local curve w~(z)\tilde{w}(z) is contained in the locally invariant manifold MεM_{\varepsilon}, that is,

(3.10) w~1=A(w~0w~02)+εs1(w~0,ε),w~2=εs2(w~0,ε),near z0.\tilde{w}_{1}=A(\tilde{w}_{0}-\tilde{w}_{0}^{2})+\varepsilon s_{1}(\tilde{w}_{0},\varepsilon),\quad\tilde{w}_{2}=\varepsilon s_{2}(\tilde{w}_{0},\varepsilon),\quad\mbox{near }z_{0}.

From the definition of w~\tilde{w}, we have w~0=w~1\tilde{w}_{0}^{\prime}=\tilde{w}_{1}, and so w~0\tilde{w}_{0} is a local solution of (3.7). Then, by the uniqueness of (3.7) with w~(z0)=w¯(z0)\tilde{w}(z_{0})=\bar{w}(z_{0}), we can deduce that w~0=w¯0\tilde{w}_{0}=\bar{w}_{0} near z0z_{0}. Furthermore, since both (w¯1,w¯2)(\bar{w}_{1},\bar{w}_{2}) and (w~1,w~2)(\tilde{w}_{1},\tilde{w}_{2}) are parametrized by w¯0\bar{w}_{0} and w~0\tilde{w}_{0} as in (3.9) and (3.10) respectively, those are locally the same as well. This proves that w¯\bar{w} satisfies (3.1) at z0z_{0}. Since z0z_{0} was chosen arbitrary in {\mathbb{R}}, the globally defined function w¯\bar{w} is a solution to the system (3.1).

Lastly, we prove that w¯\bar{w} connects the critical points. Since w¯0\bar{w}_{0} converges at ±\pm\infty, w¯1\bar{w}_{1} and w¯2\bar{w}_{2} also converge at ±\pm\infty by its constructions. Hence, the solution w¯\bar{w} has a limit at ±\pm\infty. Notice that the limits of a solution of an ODE system at ±\pm\infty should be its stationary points. Thus, w¯\bar{w} converges to (0,0,0)(0,0,0) and (1,0,0)(1,0,0) at ±\pm\infty because there are only two critical points of (3.1) as we observed in advance.

Step 2: In Step 1, for sufficiently small ε\varepsilon, we proved the global existence of (3.1) with

limzw0(z)=0,limzw0(z)=1,limz±w0(z)=0, and limz±w0′′(z)=0.\lim_{z\to-\infty}w_{0}(z)=0,\quad\lim_{z\to\infty}w_{0}(z)=1,\quad\lim_{z\to\pm\infty}w_{0}^{\prime}(z)=0,\quad\text{ and }\quad\lim_{z\to\pm\infty}w_{0}^{\prime\prime}(z)=0.

Especially, it holds from (3.1) that

(3.11) ε2w0′′′(z)=f(w0(z),w0(z),εw0′′(z),ε).\varepsilon^{2}w^{\prime\prime\prime}_{0}(z)=f(w_{0}(z),w^{\prime}_{0}(z),\varepsilon w^{\prime\prime}_{0}(z),\varepsilon).

In this step, we will show that the function w0(z)w_{0}(z) in (3.11) gives a solution to the following system:

(3.12) {σε(v~v)(u~u)=τv~v~,σε(u~u)+(p~p)=μu~v~,σε(Rγ1(θ~θ)+12(u~2u2))+(p~u~pu)=κθ~v~+μu~u~v~,\left\{\begin{aligned} &-\sigma_{\varepsilon}(\tilde{v}-v_{-})-(\tilde{u}-u_{-})=\tau\frac{\tilde{v}^{\prime}}{\tilde{v}},\\ &-\sigma_{\varepsilon}(\tilde{u}-u_{-})+(\tilde{p}-p_{-})=\mu\frac{\tilde{u}^{\prime}}{\tilde{v}},\\ &-\sigma_{\varepsilon}\left(\frac{R}{\gamma-1}(\tilde{\theta}-\theta_{-})+\frac{1}{2}(\tilde{u}^{2}-u_{-}^{2})\right)+(\tilde{p}\tilde{u}-p_{-}u_{-})=\kappa\frac{\tilde{\theta}^{\prime}}{\tilde{v}}+\mu\frac{\tilde{u}\tilde{u}^{\prime}}{\tilde{v}},\end{aligned}\right.

and thus, to the desired system (LABEL:shock_0).

From now on, we consider the parameter ε\varepsilon as the amplitude of the shock ε=v+v>0\varepsilon=v_{+}-v_{-}>0. First of all, we claim

σ¯ε=σε.\overline{\sigma}_{\varepsilon}=\sigma_{\varepsilon}.

Given (LABEL:end-con), we have (p+p)(v+v)=(u+u)2(p_{+}-p_{-})(v_{+}-v_{-})=-(u_{+}-u_{-})^{2}, so that

(3.13) θ+=v+R(p(u+u)2v+v)=θ+pR(v+v)v+R(u+u)2v+v.\theta_{+}=\frac{v_{+}}{R}\left(p_{-}-\frac{(u_{+}-u_{-})^{2}}{v_{+}-v_{-}}\right)=\theta_{-}+\frac{p_{-}}{R}(v_{+}-v_{-})-\frac{v_{+}}{R}\frac{(u_{+}-u_{-})^{2}}{v_{+}-v_{-}}.

Multiplying (LABEL:end-con)2 by u+u_{+}, and substracting it from (LABEL:end-con)3 gives

σεRγ1(θ+θ)+σε2(u+u)2+p(u+u)=0,-\sigma_{\varepsilon}\frac{R}{\gamma-1}(\theta_{+}-\theta_{-})+\frac{\sigma_{\varepsilon}}{2}(u_{+}-u_{-})^{2}+p_{-}(u_{+}-u_{-})=0,

which implies

(3.14) θ+=θ+γ12R(u+u)2γ1Rp(v+v).\theta_{+}=\theta_{-}+\frac{\gamma-1}{2R}(u_{+}-u_{-})^{2}-\frac{\gamma-1}{R}p_{-}(v_{+}-v_{-}).

From (3.13) and (3.14), we obtain the expression for (u+u)2(u_{+}-u_{-})^{2}:

(3.15) (u+u)2=γp(v+v)γ12+v+v+v=γp(v+v)2v+γ+12(v+v).(u_{+}-u_{-})^{2}=\frac{\gamma p_{-}(v_{+}-v_{-})}{\frac{\gamma-1}{2}+\frac{v_{+}}{v_{+}-v_{-}}}=\frac{\gamma p_{-}(v_{+}-v_{-})^{2}}{v_{-}+\frac{\gamma+1}{2}(v_{+}-v_{-})}.

From (LABEL:end-con)1, u+uu_{+}-u_{-} and v+vv_{+}-v_{-} have the opposite sign.
Therefore, we obtain that

(3.16) u+u\displaystyle u_{+}-u_{-} =γpv+γ+12(v+v)(v+v).\displaystyle=-\frac{\sqrt{\gamma p_{-}}}{\sqrt{v_{-}+\frac{\gamma+1}{2}(v_{+}-v_{-})}}(v_{+}-v_{-}).

Plugging (3.16) into (LABEL:end-con)1, we conclude that

(3.17) σε=u+uv+v=γpv+γ+12(v+v)=γpv+γ+12ε=σ¯ε\sigma_{\varepsilon}=-\frac{u_{+}-u_{-}}{v_{+}-v_{-}}=\frac{\sqrt{\gamma p_{-}}}{\sqrt{v_{-}+\frac{\gamma+1}{2}(v_{+}-v_{-})}}=\frac{\sqrt{\gamma p_{-}}}{\sqrt{v_{-}+\frac{\gamma+1}{2}\varepsilon}}=\overline{\sigma}_{\varepsilon}

as desired.

Now, we define v~=v~(ξ)\tilde{v}=\tilde{v}(\xi) by

(3.18) v~(ξ)εw0(εξ)+v,εξ=z.\tilde{v}(\xi)\coloneqq\varepsilon w_{0}(\varepsilon\xi)+v_{-},\quad\varepsilon\xi=z.

Multiplying ε2μτκR(εw0+v)2\frac{\varepsilon^{2}\mu\tau\kappa}{R(\varepsilon w_{0}+v_{-})^{2}} on both sides of (3.11) and using (3.18), we have

(3.19) 0\displaystyle 0 =σεv~γ1(σε2(v~v)σε(μ+τ)v~v~μτv~′′v~2+μτ(v~)2v~3)+σεpγ1(v~v)\displaystyle=\frac{\sigma_{\varepsilon}\tilde{v}}{\gamma-1}\left(-\sigma_{\varepsilon}^{2}(\tilde{v}-v_{-})-\sigma_{\varepsilon}(\mu+\tau)\frac{\tilde{v}^{\prime}}{\tilde{v}}-\mu\tau\frac{\tilde{v}^{\prime\prime}}{\tilde{v}^{2}}+\mu\tau\frac{(\tilde{v}^{\prime})^{2}}{\tilde{v}^{3}}\right)+\frac{\sigma_{\varepsilon}p_{-}}{\gamma-1}(\tilde{v}-v_{-})
σε32(v~v)2σε2τ(v~v)v~v~σετ22(v~)2v~2+σεp(v~v)+pτv~v~\displaystyle\qquad-\frac{\sigma_{\varepsilon}^{3}}{2}(\tilde{v}-v_{-})^{2}-\sigma_{\varepsilon}^{2}\tau(\tilde{v}-v_{-})\frac{\tilde{v}^{\prime}}{\tilde{v}}-\frac{\sigma_{\varepsilon}\tau^{2}}{2}\frac{(\tilde{v}^{\prime})^{2}}{\tilde{v}^{2}}+\sigma_{\varepsilon}p_{-}(\tilde{v}-v_{-})+p_{-}\tau\frac{\tilde{v}^{\prime}}{\tilde{v}}
+κRv~((p2σε2v~+σε2v)v~σε(μ+τ)v~′′μτ(v~′′′v~3v~v~′′v~2+2(v~)3v~3))\displaystyle\qquad+\frac{\kappa}{R\tilde{v}}\left((p_{-}-2\sigma_{\varepsilon}^{2}\tilde{v}+\sigma_{\varepsilon}^{2}v_{-})\tilde{v}^{\prime}-\sigma_{\varepsilon}(\mu+\tau)\tilde{v}^{\prime\prime}-\mu\tau\left(\frac{\tilde{v}^{\prime\prime\prime}}{\tilde{v}}-\frac{3\tilde{v}^{\prime}\tilde{v}^{\prime\prime}}{\tilde{v}^{2}}+\frac{2(\tilde{v}^{\prime})^{3}}{\tilde{v}^{3}}\right)\right)

which can be reduced to

(3.20) σεγ1[v~(pσε2(v~v)σε(μ+τ)v~v~μτv~′′v~2+μτ(v~)2v~3)pv]\displaystyle-\frac{\sigma_{\varepsilon}}{\gamma-1}\left[\tilde{v}\left(p_{-}-\sigma_{\varepsilon}^{2}(\tilde{v}-v_{-})-\sigma_{\varepsilon}(\mu+\tau)\frac{\tilde{v}^{\prime}}{\tilde{v}}-\mu\tau\frac{\tilde{v}^{\prime\prime}}{\tilde{v}^{2}}+\mu\tau\frac{(\tilde{v}^{\prime})^{2}}{\tilde{v}^{3}}\right)-p_{-}v_{-}\right]
+12σε(σε(v~v)τv~v~)2+p(σε(v~v)τv~v~)\displaystyle+\frac{1}{2}\sigma_{\varepsilon}\left(-\sigma_{\varepsilon}(\tilde{v}-v_{-})-\tau\frac{\tilde{v}^{\prime}}{\tilde{v}}\right)^{2}+p_{-}\left(-\sigma_{\varepsilon}(\tilde{v}-v_{-})-\tau\frac{\tilde{v}^{\prime}}{\tilde{v}}\right)
=κRv~((p2σε2v~+σε2v)v~σε(μ+τ)v~′′μτ(v~′′′v~3v~v~′′v~2+2(v~)3v~3)).\displaystyle\quad=\frac{\kappa}{R\tilde{v}}\left((p_{-}-2\sigma_{\varepsilon}^{2}\tilde{v}+\sigma_{\varepsilon}^{2}v_{-})\tilde{v}^{\prime}-\sigma_{\varepsilon}(\mu+\tau)\tilde{v}^{\prime\prime}-\mu\tau\left(\frac{\tilde{v}^{\prime\prime\prime}}{\tilde{v}}-\frac{3\tilde{v}^{\prime}\tilde{v}^{\prime\prime}}{\tilde{v}^{2}}+\frac{2(\tilde{v}^{\prime})^{3}}{\tilde{v}^{3}}\right)\right).

Then we define u~\tilde{u} and θ~\tilde{\theta} in accordance with (3.12)1 and (3.12)2 respectively as follows:

(3.21) u~uσε(v~v)τv~v~,\displaystyle\tilde{u}\coloneqq u_{-}-\sigma_{\varepsilon}(\tilde{v}-v_{-})-\tau\frac{\tilde{v}^{\prime}}{\tilde{v}},

and

θ~pRv~+σεRv~(u~u)+μRu~.\tilde{\theta}\coloneqq\frac{p_{-}}{R}\tilde{v}+\frac{\sigma_{\varepsilon}}{R}\tilde{v}(\tilde{u}-u_{-})+\frac{\mu}{R}\tilde{u}^{\prime}.

Since we have

(3.22) u~=σεv~τv~′′v~+τ(v~)2v~2,\tilde{u}^{\prime}=-\sigma_{\varepsilon}\tilde{v}^{\prime}-\tau\frac{\tilde{v}^{\prime\prime}}{\tilde{v}}+\tau\frac{(\tilde{v}^{\prime})^{2}}{\tilde{v}^{2}},

θ~\tilde{\theta} can be written in the following way:

(3.23) θ~=v~R(pσε2(v~v)σε(μ+τ)v~v~μτv~′′v~2+μτ(v~)2v~3).\tilde{\theta}=\frac{\tilde{v}}{R}\left(p_{-}-\sigma_{\varepsilon}^{2}(\tilde{v}-v_{-})-\sigma_{\varepsilon}(\mu+\tau)\frac{\tilde{v}^{\prime}}{\tilde{v}}-\mu\tau\frac{\tilde{v}^{\prime\prime}}{\tilde{v}^{2}}+\mu\tau\frac{(\tilde{v}^{\prime})^{2}}{\tilde{v}^{3}}\right).

Hence, differentiating both sides of (3.23), we get

(3.24) θ~\displaystyle\tilde{\theta}^{\prime} =1R(pv~2σε2v~v~+σε2vv~σε(μ+τ)v~′′μτ(v~′′′v~v~v~′′v~22v~v~′′v~2+2(v~)3v~3))\displaystyle=\frac{1}{R}\left(p_{-}\tilde{v}^{\prime}-2\sigma_{\varepsilon}^{2}\tilde{v}\tilde{v}^{\prime}+\sigma_{\varepsilon}^{2}v_{-}\tilde{v}^{\prime}-\sigma_{\varepsilon}(\mu+\tau)\tilde{v}^{\prime\prime}-\mu\tau\left(\frac{\tilde{v}^{\prime\prime\prime}}{\tilde{v}}-\frac{\tilde{v}^{\prime}\tilde{v}^{\prime\prime}}{\tilde{v}^{2}}-\frac{2\tilde{v}^{\prime}\tilde{v}^{\prime\prime}}{\tilde{v}^{2}}+\frac{2(\tilde{v}^{\prime})^{3}}{\tilde{v}^{3}}\right)\right)
=1R((p2σε2v~+σε2v)v~σε(μ+τ)v~′′μτ(v~′′′v~3v~v~′′v~2+2(v~)3v~3)).\displaystyle=\frac{1}{R}\left((p_{-}-2\sigma_{\varepsilon}^{2}\tilde{v}+\sigma_{\varepsilon}^{2}v_{-})\tilde{v}^{\prime}-\sigma_{\varepsilon}(\mu+\tau)\tilde{v}^{\prime\prime}-\mu\tau\left(\frac{\tilde{v}^{\prime\prime\prime}}{\tilde{v}}-\frac{3\tilde{v}^{\prime}\tilde{v}^{\prime\prime}}{\tilde{v}^{2}}+\frac{2(\tilde{v}^{\prime})^{3}}{\tilde{v}^{3}}\right)\right).

Plugging (3.21), (3.23), and (3.24) into (3.20), we obtain that

(3.25) σε(Rγ1(θ~θ)12(u~u)2)+p(u~u)=κθ~v~.-\sigma_{\varepsilon}\left(\frac{R}{\gamma-1}(\tilde{\theta}-\theta_{-})-\frac{1}{2}(\tilde{u}-u_{-})^{2}\right)+p_{-}(\tilde{u}-u_{-})=\kappa\frac{\tilde{\theta}^{\prime}}{\tilde{v}}.

Then, multiplying (3.12)2 by u~\tilde{u}, and adding it to (3.25) yields (3.12)3:

σε(Rγ1(θ~θ)+12(u~2u2))+(p~u~pu)=κθ~v~+μu~u~v~.-\sigma_{\varepsilon}\left(\frac{R}{\gamma-1}(\tilde{\theta}-\theta_{-})+\frac{1}{2}(\tilde{u}^{2}-u_{-}^{2})\right)+(\tilde{p}\tilde{u}-p_{-}u_{-})=\kappa\frac{\tilde{\theta}^{\prime}}{\tilde{v}}+\mu\frac{\tilde{u}\tilde{u}^{\prime}}{\tilde{v}}.

Furthermore, since w0w_{0} is increasing, the monotonicity of v~\tilde{v} follows from (3.18). The monotonicity of u~\tilde{u} and θ~\tilde{\theta} follows from (1.10) and (1.11), which will be proved in Section 4. ∎

3.2. Proof of Theorem 1.1: Uniqueness

In Section 3.1, we proved the existence of a solution of (3.12) connecting (v,u,θ)(v_{-},u_{-},\theta_{-}) to (v+,u+,θ+)(v_{+},u_{+},\theta_{+}). Now, we will prove the uniqueness of the monotone solution to (3.12) connecting two endpoints (v,u,θ)(v_{-},u_{-},\theta_{-}) to (v+,u+,θ+)(v_{+},u_{+},\theta_{+}).

To achieve this, we will use the following proposition on the unstable manifold at a critical point.

Proposition 3.1.

[34, Theorems 9.4, 9.5] Consider a system of ODEs z=F(z)z^{\prime}=F(z), where F𝒞k(n;n)F\in\mathcal{C}^{k}({\mathbb{R}}^{n};{\mathbb{R}}^{n}) and x0x_{0} is a critical point, i.e., F(x0)=0F(x_{0})=0. Let W(x0)W^{-}(x_{0}) be the unstable manifold composed of all points converging to x0x_{0} for tt\to-\infty. Let E+E^{+} (resp. EE^{-}) denote the linear subspace spanned by eigenvectors of Dx0FD_{x_{0}}F corresponding to negative (resp. positive) eigenvalues. Suppose Dx0FD_{x_{0}}F has no eigenvalues on the imaginary axis.
Then, there are a neighborhood U(x0)=x0+UU(x_{0})=x_{0}+U of x0x_{0} and a function h𝒞k(EU;E+)h^{-}\in\mathcal{C}^{k}(E^{-}\cap U;E^{+}) such that

W(x0)U(x0)={x0+a+h(a)aEU}.W^{-}(x_{0})\cap U(x_{0})=\{x_{0}+a+h^{-}(a)\mid a\in E^{-}\cap U\}.

Here, both hh^{-} and its Jacobian matrix vanishes at 0.

This implies that the set W(x0)W^{-}(x_{0}) is a CkC^{k}-manifold of which dimension is equal to the dimension of EE^{-}, and this is tangent to the affine space x0+Ex_{0}+E^{-} at x0x_{0}.

To apply Proposition 3.1, we rewrite the system (3.12) by using (3.25) as follows:

(3.26) {v~=v~τ[σε(v~v)(u~u)],u~=v~μ[σε(u~u)+(p~p)],θ~=v~κ[σε(Rγ1(θ~θ)+12(u~u)2)+p(u~u)].\left\{\begin{aligned} &\tilde{v}^{\prime}=\frac{\tilde{v}}{\tau}\left[-\sigma_{\varepsilon}(\tilde{v}-v_{-})-(\tilde{u}-u_{-})\right],\\ &\tilde{u}^{\prime}=\frac{\tilde{v}}{\mu}\left[-\sigma_{\varepsilon}(\tilde{u}-u_{-})+(\tilde{p}-p_{-})\right],\\ &\tilde{\theta}^{\prime}=\frac{\tilde{v}}{\kappa}\left[-\sigma_{\varepsilon}\left(\frac{R}{\gamma-1}(\tilde{\theta}-\theta_{-})+\frac{1}{2}(\tilde{u}-u_{-})^{2}\right)+p_{-}(\tilde{u}-u_{-})\right].\end{aligned}\right.

It is easy to see from (LABEL:end-con) that (v±,u±,θ±)(v_{\pm},u_{\pm},\theta_{\pm}) are critical points of the system (3.26). Then we need to compute the sign of eigenvalues of the linearized system at the critical point (v,u,θ)(v_{-},u_{-},\theta_{-}) so as to determine the dimension of the unstable manifold. For (3.26), the Jacobian matrix J(v~,u~,θ~)J(\tilde{v},\tilde{u},\tilde{\theta}) at (v,u,θ)(v_{-},u_{-},\theta_{-}) is given by

(3.27) J(v~,u~,θ~)|(v~,u~,θ~)=(v,u,θ)=(σεvτvτ0pμσεvμRμ0vpκRγ1σεvκ).\left.J(\tilde{v},\tilde{u},\tilde{\theta})\right|_{(\tilde{v},\tilde{u},\tilde{\theta})=(v_{-},u_{-},\theta_{-})}=\begin{pmatrix}-\frac{\sigma_{\varepsilon}v_{-}}{\tau}&-\frac{v_{-}}{\tau}&0\\ -\frac{p_{-}}{\mu}&-\frac{\sigma_{\varepsilon}v_{-}}{\mu}&\frac{R}{\mu}\\ 0&\frac{v_{-}p_{-}}{\kappa}&-\frac{R}{\gamma-1}\frac{\sigma_{\varepsilon}v_{-}}{\kappa}\end{pmatrix}.

We then compute the trace and determinant of (3.27) as follows:

detJ((v,u,θ))\displaystyle\det J((v_{-},u_{-},\theta_{-})) =1μτκRσεv2γ1(γpσε2v)>0,\displaystyle=\frac{1}{\mu\tau\kappa}\frac{R\sigma_{\varepsilon}v_{-}^{2}}{\gamma-1}(\gamma p_{-}-\sigma_{\varepsilon}^{2}v_{-})>0,
trJ((v,u,θ))\displaystyle\operatorname{tr}J((v_{-},u_{-},\theta_{-})) =σεv(Rγ11κ+1μ+1τ)<0.\displaystyle=-\sigma_{\varepsilon}v_{-}\left(\frac{R}{\gamma-1}\frac{1}{\kappa}+\frac{1}{\mu}+\frac{1}{\tau}\right)<0.

Here, we utilized (3.4) for the first inequality. Upon examining the graph of the characteristic equation λ3+Aλ2+Bλ+C=0\lambda^{3}+A\lambda^{2}+B\lambda+C=0 with C<0C<0, we deduce that there is at least one positive real root. Since the sum of three eigenvalues are negative, the sum of the other two roots is negative and their product is positive; those can be complex roots with negative real part, or two negative real roots. Hence, the dimension of the unstable manifold, which is equal to the number of eigenvalues with positive real part, is 11.

Now, we prove the desired uniqueness. We already constructed a monotone solution (v~,u~,θ~)(\tilde{v},\tilde{u},\tilde{\theta}) which satisfies (3.12) and connects (v,u,θ)(v_{-},u_{-},\theta_{-}) to (v+,u+,θ+)(v_{+},u_{+},\theta_{+}). This solution trajectory (v~,u~,θ~)(\tilde{v},\tilde{u},\tilde{\theta}) is locally contained in the one-dimensional unstable manifold W((v,u,θ))W^{-}((v_{-},u_{-},\theta_{-})). More precisely, W((v,u,θ)){vv,uu,θθ}W^{-}((v_{-},u_{-},\theta_{-}))\cap\{v\geq v_{-},u\leq u_{-},\theta\leq\theta_{-}\} must coincide with the trajectory (v~,u~,θ~)(\tilde{v},\tilde{u},\tilde{\theta}) near (v,u,θ)(v_{-},u_{-},\theta_{-}). Note that, any other solution which connects (v,u,θ)(v_{-},u_{-},\theta_{-}) to (v+,u+,θ+)(v_{+},u_{+},\theta_{+}) should be on W((v,u,θ))W^{-}((v_{-},u_{-},\theta_{-})), and especially if it is monotone, it should intersect with the trajectory (v~,u~,θ~)(\tilde{v},\tilde{u},\tilde{\theta}). Hence, by the uniqueness for the Lipschitz autonomous system of ODE, (v~,u~,θ~)(\tilde{v},\tilde{u},\tilde{\theta}) is the unique solution in the class of monotone solutions, up to a translation.

4. Proof of Theorem 1.2

We present the proof of Theorem 1.2. In this section, CC represents a positive constant that may vary from line to line, yet remains independent of ε\varepsilon.

To prove the desired estimates as in Theorem 1.2, we will proceed in the following order. First, we obtain (|v~ε′′(ξ)|,|u~ε′′(ξ)|,|θ~ε′′(ξ)|)Cε|v~ε(ξ)|(\left|{\tilde{v}_{\varepsilon}}^{\prime\prime}(\xi)\right|,\left|{\tilde{u}_{\varepsilon}}^{\prime\prime}(\xi)\right|,\left|{\tilde{\theta}_{\varepsilon}}^{\prime\prime}(\xi)\right|)\leq C\varepsilon\left|{\tilde{v}_{\varepsilon}}^{\prime}(\xi)\right|. We then prove (1.10) and (1.11), so that (1.14) and (1.9) are automatically established. Subsequently, we show (1.15), and this with Gronwall’s inequality gives (LABEL:decay) and (1.8).

Recall (3.1) and (3.18): we have the following relations.

(4.1) v~ε(ξ)=ε2w0(z),v~ε′′(ξ)=ε2w2(z)=ε3w1(z).\displaystyle{\tilde{v}_{\varepsilon}}^{\prime}(\xi)=\varepsilon^{2}w^{\prime}_{0}(z),\quad{\tilde{v}_{\varepsilon}}^{\prime\prime}(\xi)=\varepsilon^{2}w_{2}(z)=\varepsilon^{3}w_{1}^{\prime}(z).

Similarly, we also have

(4.2) v~ε′′′(ξ)=ε4w1′′(z),v~ε′′′′(ξ)=ε5w1′′′(z).\displaystyle{\tilde{v}_{\varepsilon}}^{\prime\prime\prime}(\xi)=\varepsilon^{4}w^{\prime\prime}_{1}(z),\quad{\tilde{v}_{\varepsilon}}^{\prime\prime\prime\prime}(\xi)=\varepsilon^{5}w^{\prime\prime\prime}_{1}(z).

On the other hand, from (3.6), we have

(4.3) w1=A(w0w02)+εs1(w0,ε),w0K,\displaystyle w_{1}=A(w_{0}-w_{0}^{2})+\varepsilon s_{1}(w_{0},\varepsilon),\quad w_{0}\in K,

for some compact set KK, and s1s_{1} is smooth.
So, we get |w1|=|w0|C\left|w_{1}\right|=\left|w^{\prime}_{0}\right|\leq C. Then, using (4.1)1, we obtain |v~ε(ξ)|Cε2\left|{\tilde{v}_{\varepsilon}}^{\prime}(\xi)\right|\leq C\varepsilon^{2}. Following this, we differentiate (4.3):

w1=A(12w0)w0+εD1s1(w0,ε)w0.w^{\prime}_{1}=A(1-2w_{0})w^{\prime}_{0}+\varepsilon D_{1}s_{1}(w_{0},\varepsilon)w^{\prime}_{0}.

Then, since D1s1D_{1}s_{1} is continuous and w0w_{0} is confined to KK, it holds that |w1(z)|C|w0(z)|\left|w^{\prime}_{1}(z)\right|\leq C\left|w^{\prime}_{0}(z)\right|. This together with (4.1) implies that |v~ε′′(ξ)|Cε|v~ε(ξ)|\left|{\tilde{v}_{\varepsilon}}^{\prime\prime}(\xi)\right|\leq C\varepsilon\left|{\tilde{v}_{\varepsilon}}^{\prime}(\xi)\right|.

To obtain the bounds for |u~ε′′|\left|{\tilde{u}_{\varepsilon}}^{\prime\prime}\right| and |θ~ε′′|\left|{\tilde{\theta}_{\varepsilon}}^{\prime\prime}\right|, we differentiate (4.3) once and twice more:

w1′′\displaystyle w^{\prime\prime}_{1} =2A(w0)2+A(12w0)w0′′+εD11s1(w0,ε)(w0)2+εD1s1(w0,ε)w0′′\displaystyle=-2A(w^{\prime}_{0})^{2}+A(1-2w_{0})w^{\prime\prime}_{0}+\varepsilon D_{11}s_{1}(w_{0},\varepsilon)(w^{\prime}_{0})^{2}+\varepsilon D_{1}s_{1}(w_{0},\varepsilon)w^{\prime\prime}_{0}
=2Aw1w0+A(12w0)w1+εD11s1(w0,ε)w1w0+εD1s1(w0,ε)w1,\displaystyle=-2Aw_{1}w^{\prime}_{0}+A(1-2w_{0})w^{\prime}_{1}+\varepsilon D_{11}s_{1}(w_{0},\varepsilon)w_{1}w^{\prime}_{0}+\varepsilon D_{1}s_{1}(w_{0},\varepsilon)w^{\prime}_{1},
w1′′′\displaystyle w^{\prime\prime\prime}_{1} =6Aw0w0′′+A(12w0)w0′′′+εD111s1(w0,ε)(w0)3\displaystyle=-6Aw^{\prime}_{0}w^{\prime\prime}_{0}+A(1-2w_{0})w^{\prime\prime\prime}_{0}+\varepsilon D_{111}s_{1}(w_{0},\varepsilon)(w^{\prime}_{0})^{3}
+3εD11s1(w0,ε)w0w0′′+εD1s1(w0,ε)w0′′′\displaystyle\qquad+3\varepsilon D_{11}s_{1}(w_{0},\varepsilon)w^{\prime}_{0}w^{\prime\prime}_{0}+\varepsilon D_{1}s_{1}(w_{0},\varepsilon)w^{\prime\prime\prime}_{0}
=6Aw1w1+A(12w0)w1′′+εD111s1(w0,ε)w12w0\displaystyle=-6Aw_{1}w^{\prime}_{1}+A(1-2w_{0})w^{\prime\prime}_{1}+\varepsilon D_{111}s_{1}(w_{0},\varepsilon)w_{1}^{2}w^{\prime}_{0}
+3εD11s1(w0,ε)w1w1+εD1s1(w0,ε)w1′′.\displaystyle\qquad+3\varepsilon D_{11}s_{1}(w_{0},\varepsilon)w_{1}w^{\prime}_{1}+\varepsilon D_{1}s_{1}(w_{0},\varepsilon)w^{\prime\prime}_{1}.

Then, since all the higher derivatives of s1s_{1} are continuous and w0w_{0} is confined to KK as well, we have |w1′′(z)|C|w0(z)|\left|w^{\prime\prime}_{1}(z)\right|\leq C\left|w^{\prime}_{0}(z)\right| and |w1′′′(z)|C|w0(z)|\left|w^{\prime\prime\prime}_{1}(z)\right|\leq C\left|w^{\prime}_{0}(z)\right|.
Hence, from (4.2), we also obtain |v~ε′′′(ξ)|Cε2|v~ε(ξ)|\left|{\tilde{v}_{\varepsilon}}^{\prime\prime\prime}(\xi)\right|\leq C\varepsilon^{2}\left|{\tilde{v}_{\varepsilon}}^{\prime}(\xi)\right| and |v~ε′′′′(ξ)|Cε3|v~ε(ξ)|\left|{\tilde{v}_{\varepsilon}}^{\prime\prime\prime\prime}(\xi)\right|\leq C\varepsilon^{3}\left|{\tilde{v}_{\varepsilon}}^{\prime}(\xi)\right|.
Thus, by differentiating (3.22) and (3.24), we use all of these estimates to yield that

|u~ε′′(ξ)|,|θ~ε′′(ξ)|Cε|v~ε(ξ)|.\left|{\tilde{u}_{\varepsilon}}^{\prime\prime}(\xi)\right|,\left|{\tilde{\theta}_{\varepsilon}}^{\prime\prime}(\xi)\right|\leq C\varepsilon\left|{\tilde{v}_{\varepsilon}}^{\prime}(\xi)\right|.

We now show (1.10) and (1.11): from (LABEL:shock_0)1, it follows that

|u~ε+σεv~ε|=|τv~ε′′v~ετ(v~ε)2v~ε2|Cε|v~ε|.\left|{\tilde{u}_{\varepsilon}}^{\prime}+\sigma_{\varepsilon}{\tilde{v}_{\varepsilon}}^{\prime}\right|=\left|\tau\frac{{\tilde{v}_{\varepsilon}}^{\prime\prime}}{{\tilde{v}_{\varepsilon}}}-\tau\frac{({\tilde{v}_{\varepsilon}}^{\prime})^{2}}{{\tilde{v}_{\varepsilon}}^{2}}\right|\leq C\varepsilon\left|{\tilde{v}_{\varepsilon}}^{\prime}\right|.

Then, (1.12) gives the desired result (1.10). Likewise, for (1.11), we use (3.24) to get

|θ~ε1R((p2σε2v~ε+σε2v)v~ε)|Cε|v~ε|.\left|{\tilde{\theta}_{\varepsilon}}^{\prime}-\frac{1}{R}((p_{-}-2\sigma_{\varepsilon}^{2}{\tilde{v}_{\varepsilon}}+\sigma_{\varepsilon}^{2}v_{-}){\tilde{v}_{\varepsilon}}^{\prime})\right|\leq C\varepsilon\left|{\tilde{v}_{\varepsilon}}^{\prime}\right|.

Since v~ε{\tilde{v}_{\varepsilon}} is increasing, we have |v~εv||v+v|=ε\left|{\tilde{v}_{\varepsilon}}-v_{-}\right|\leq\left|v_{+}-v_{-}\right|=\varepsilon. Considering this and (1.12), the definition of σ\sigma_{*} gives the desired one (1.11). Furthermore, these two estimates with v~ε>0{\tilde{v}_{\varepsilon}}^{\prime}>0 imply that |v~ε||u~ε||θ~ε|\left|{\tilde{v}_{\varepsilon}}^{\prime}\right|\sim\left|{\tilde{u}_{\varepsilon}}^{\prime}\right|\sim|{\tilde{\theta}_{\varepsilon}}^{\prime}| and (1.14).

In order to prove the exponential decays (LABEL:decay) and (1.8), it is required to show (1.15) first. To this end, we rewrite (3.20) in the following form:

(4.4) σε3v~εγ1(v~εv)σεpγ1(v~εv)+σε32(v~εv)2σεp(v~εv)=σεv~εγ1(σε(μ+τ)v~εv~εμτv~ε′′v~ε2+μτ(v~ε)2v~ε3)σε2τ(v~εv)v~εv~εσετ22(v~ε)2v~ε2+pτv~εv~ε+κRv~ε((p2σε2v~ε+σε2v)v~εσε(μ+τ)v~ε′′μτ(v~ε′′′v~ε3v~εv~ε′′v~ε2+2(v~ε)3v~ε3)).\displaystyle\begin{aligned} &\frac{\sigma_{\varepsilon}^{3}{\tilde{v}_{\varepsilon}}}{\gamma-1}({\tilde{v}_{\varepsilon}}-v_{-})-\frac{\sigma_{\varepsilon}p_{-}}{\gamma-1}({\tilde{v}_{\varepsilon}}-v_{-})+\frac{\sigma_{\varepsilon}^{3}}{2}({\tilde{v}_{\varepsilon}}-v_{-})^{2}-\sigma_{\varepsilon}p_{-}({\tilde{v}_{\varepsilon}}-v_{-})\\ &=\frac{\sigma_{\varepsilon}{\tilde{v}_{\varepsilon}}}{\gamma-1}\left(-\sigma_{\varepsilon}(\mu+\tau)\frac{{\tilde{v}_{\varepsilon}}^{\prime}}{{\tilde{v}_{\varepsilon}}}-\mu\tau\frac{{\tilde{v}_{\varepsilon}}^{\prime\prime}}{{\tilde{v}_{\varepsilon}}^{2}}+\mu\tau\frac{({\tilde{v}_{\varepsilon}}^{\prime})^{2}}{{\tilde{v}_{\varepsilon}}^{3}}\right)-\sigma_{\varepsilon}^{2}\tau({\tilde{v}_{\varepsilon}}-v_{-})\frac{{\tilde{v}_{\varepsilon}}^{\prime}}{{\tilde{v}_{\varepsilon}}}-\frac{\sigma_{\varepsilon}\tau^{2}}{2}\frac{({\tilde{v}_{\varepsilon}}^{\prime})^{2}}{{\tilde{v}_{\varepsilon}}^{2}}+p_{-}\tau\frac{{\tilde{v}_{\varepsilon}}^{\prime}}{{\tilde{v}_{\varepsilon}}}\\ &\qquad+\frac{\kappa}{R{\tilde{v}_{\varepsilon}}}\left((p_{-}-2\sigma_{\varepsilon}^{2}{\tilde{v}_{\varepsilon}}+\sigma_{\varepsilon}^{2}v_{-}){\tilde{v}_{\varepsilon}}^{\prime}-\sigma_{\varepsilon}(\mu+\tau){\tilde{v}_{\varepsilon}}^{\prime\prime}-\mu\tau\left(\frac{{\tilde{v}_{\varepsilon}}^{\prime\prime\prime}}{{\tilde{v}_{\varepsilon}}}-\frac{3{\tilde{v}_{\varepsilon}}^{\prime}{\tilde{v}_{\varepsilon}}^{\prime\prime}}{{\tilde{v}_{\varepsilon}}^{2}}+\frac{2({\tilde{v}_{\varepsilon}}^{\prime})^{3}}{{\tilde{v}_{\varepsilon}}^{3}}\right)\right).\end{aligned}

Afterwards, we analyze both sides of (LABEL:vODEre) respectively. For the left-hand side, we use (3.3) to have the following:

(4.5) LHS\displaystyle LHS =σε3(v~εv)[γpσε2(γ1)+(v~εγ1+12(v~εv))]\displaystyle=\sigma_{\varepsilon}^{3}({\tilde{v}_{\varepsilon}}-v_{-})\left[\frac{-\gamma p_{-}}{\sigma_{\varepsilon}^{2}(\gamma-1)}+\left(\frac{{\tilde{v}_{\varepsilon}}}{\gamma-1}+\frac{1}{2}({\tilde{v}_{\varepsilon}}-v_{-})\right)\right]
=σε3(v~εv)(v++γ12(v+v)γ1+v~εγ1+12(v~εv))\displaystyle=\sigma_{\varepsilon}^{3}({\tilde{v}_{\varepsilon}}-v_{-})\left(-\frac{v_{+}+\frac{\gamma-1}{2}(v_{+}-v_{-})}{\gamma-1}+\frac{{\tilde{v}_{\varepsilon}}}{\gamma-1}+\frac{1}{2}({\tilde{v}_{\varepsilon}}-v_{-})\right)
=σε3(1γ1+12)(v~εv)(v~εv+).\displaystyle=\sigma_{\varepsilon}^{3}\left(\frac{1}{\gamma-1}+\frac{1}{2}\right)({\tilde{v}_{\varepsilon}}-v_{-})({\tilde{v}_{\varepsilon}}-v_{+}).

We then observe that all the terms on the right-hand side can be controlled by v~ε{\tilde{v}_{\varepsilon}}^{\prime}. Indeed,

(4.6) RHS\displaystyle RHS =(σε2(μ+τ)γ1+pτ1v~ε2κRσε2+κRv~ε(p+σε2v)+𝒪(ε))v~ε\displaystyle=\left(-\frac{\sigma_{\varepsilon}^{2}(\mu+\tau)}{\gamma-1}+p_{-}\tau\frac{1}{{\tilde{v}_{\varepsilon}}}-2\frac{\kappa}{R}\sigma_{\varepsilon}^{2}+\frac{\kappa}{R{\tilde{v}_{\varepsilon}}}(p_{-}+\sigma_{\varepsilon}^{2}v_{-})+\mathcal{O}(\varepsilon)\right){\tilde{v}_{\varepsilon}}^{\prime}
=(γ(μ+τ)p(γ1)v+pτ1v(γ1)κpRv+𝒪(ε))v~ε\displaystyle=\left(-\frac{\gamma(\mu+\tau)p_{-}}{(\gamma-1)v_{-}}+p_{-}\tau\frac{1}{v_{-}}-\frac{(\gamma-1)\kappa p_{-}}{Rv_{-}}+\mathcal{O}(\varepsilon)\right){\tilde{v}_{\varepsilon}}^{\prime}
=pv(γγ1μ+1γ1τ+γ1Rκ+𝒪(ε))v~ε.\displaystyle=-\frac{p_{-}}{v_{-}}\left(\frac{\gamma}{\gamma-1}\mu+\frac{1}{\gamma-1}\tau+\frac{\gamma-1}{R}\kappa+\mathcal{O}(\varepsilon)\right){\tilde{v}_{\varepsilon}}^{\prime}.

Combining (LABEL:vODEre) with (4.5) and (4.6), we obtain (1.15), i.e.,

C1(v+v~ε)(v~εv)v~εC(v+v~ε)(v~εv).C^{-1}(v_{+}-{\tilde{v}_{\varepsilon}})({\tilde{v}_{\varepsilon}}-v_{-})\leq{\tilde{v}_{\varepsilon}}^{\prime}\leq C(v_{+}-{\tilde{v}_{\varepsilon}})({\tilde{v}_{\varepsilon}}-v_{-}).

With this, the standard argument of Gronwall’s inequality (for example, see [23, Lemma2.1]) gives that

(4.7) |v~ε(ξ)v|CεeCε|ξ| for ξ0,|v~ε(ξ)v+|CεeCε|ξ| for ξ0.\displaystyle\left|{\tilde{v}_{\varepsilon}}(\xi)-v_{-}\right|\leq C\varepsilon e^{-C\varepsilon\left|\xi\right|}\text{ for }\xi\leq 0,\quad\quad\left|{\tilde{v}_{\varepsilon}}(\xi)-v_{+}\right|\leq C\varepsilon e^{-C\varepsilon\left|\xi\right|}\text{ for }\xi\geq 0.

Then, by plugging these estimates into (1.15), we get

|v~ε(ξ)|Cε2eCε|ξ|,\left|{\tilde{v}_{\varepsilon}}^{\prime}(\xi)\right|\leq C\varepsilon^{2}e^{-C\varepsilon\left|\xi\right|},

and this together with (1.10) and (1.11) gives the full strength of (1.8).
In addition, applying (4.7) and (1.8) to (3.12)1 and (3.25), we obtain that

|u~ε(ξ)u|CεeCε|ξ| for ξ0,\displaystyle\left|{\tilde{u}_{\varepsilon}}(\xi)-u_{-}\right|\leq C\varepsilon e^{-C\varepsilon\left|\xi\right|}\text{ for }\xi\leq 0, |u~ε(ξ)u+|CεeCε|ξ| for ξ0,\displaystyle\left|{\tilde{u}_{\varepsilon}}(\xi)-u_{+}\right|\leq C\varepsilon e^{-C\varepsilon\left|\xi\right|}\text{ for }\xi\geq 0,
|θ~ε(ξ)θ|CεeCε|ξ| for ξ0,\displaystyle\left|{\tilde{\theta}_{\varepsilon}}(\xi)-\theta_{-}\right|\leq C\varepsilon e^{-C\varepsilon\left|\xi\right|}\text{ for }\xi\leq 0, |θ~ε(ξ)θ+|CεeCε|ξ| for ξ0.\displaystyle\left|{\tilde{\theta}_{\varepsilon}}(\xi)-\theta_{+}\right|\leq C\varepsilon e^{-C\varepsilon\left|\xi\right|}\text{ for }\xi\geq 0.

This completes the proof of (LABEL:decay) and Theorem 1.2. ∎

References

  • [1] Errol B Arkilic, Kenneth S Breuer, and Martin A Schmidt. Mass flow and tangential momentum accommodation in silicon micromachined channels. Journal of fluid mechanics, 437:29–43, 2001.
  • [2] Howard Brenner. Kinematics of volume transport. Physica A: Statistical Mechanics and its Applications, 349(1-2):11–59, 2005.
  • [3] Howard Brenner. Navier–stokes revisited. Physica A: Statistical Mechanics and its Applications, 349(1-2):60–132, 2005.
  • [4] Howard Brenner. Fluid mechanics revisited. Physica A: Statistical Mechanics and its Applications, 370(2):190–224, 2006.
  • [5] Howard Brenner. Beyond navier–stokes. International Journal of Engineering Science, 54:67–98, 2012.
  • [6] Suman Chakraborty and Franz Durst. Derivations of extended navier-stokes equations from upscaled molecular transport considerations for compressible ideal gas flows: Towards extended constitutive forms. Physics of Fluids, 19(8), 2007.
  • [7] K. Choi, M.-J. Kang, Y.-S. Kwon, and A. F. Vasseur. Contraction for large perturbations of traveling waves in a hyperbolic-parabolic system arising from a chemotaxis model. Math. Models Methods Appl. Sci., 30(2):387–437, 2020.
  • [8] S Kokou Dadzie and Jason M Reese. A volume-based hydrodynamic approach to sound wave propagation in a monatomic gas. Physics of Fluids, 22(1), 2010.
  • [9] S Kokou Dadzie and Jason M Reese. Analysis of the thermomechanical inconsistency of some extended hydrodynamic models at high knudsen number. Physical Review E, 85(4):041202, 2012.
  • [10] S Kokou Dadzie and Jason M Reese. Spatial stochasticity and non-continuum effects in gas flows. Physics letters A, 376(8-9):967–972, 2012.
  • [11] S Kokou Dadzie, Jason M Reese, and Colin R McInnes. A continuum model of gas flows with localized density variations. Physica A: Statistical Mechanics and its Applications, 387(24):6079–6094, 2008.
  • [12] Nishanth Dongari, Franz Durst, and Suman Chakraborty. Predicting microscale gas flows and rarefaction effects through extended navier–stokes–fourier equations from phoretic transport considerations. Microfluidics and Nanofluidics, 9:831–846, 2010.
  • [13] Nishanth Dongari, Rajamani Sambasivam, and Franz Durst. Extended navier-stokes equations and treatments of micro-channel gas flows. Journal of Fluid Science and Technology, 4(2):454–467, 2009.
  • [14] Franz Durst, Navaneetha Krishnan, David Filimonov, Rajamani Sambasivam, Sarit K Das, et al. Treatments of flows through micro-channels based on the extended navier-stokes-equations. 2011.
  • [15] S. Eo, N. Eun, and M.-J. Kang. Stability of a Riemann shock in the physical class: from Brenner-Navier-Stokes-Fourier to Euler. in preparation.
  • [16] Eduard Feireisl and Alexis Vasseur. New perspectives in fluid dynamics: Mathematical analysis of a model proposed by howard brenner. New Directions in Mathematical Fluid Mechanics: The Alexander V. Kazhikhov Memorial Volume, pages 153–179, 2010.
  • [17] Mohamed Gad-el Hak. The fluid mechanics of microdevices—the freeman scholar lecture. 1999.
  • [18] JD Goddard. On material velocities and non-locality in the thermo-mechanics of continua. International journal of engineering science, 48(11):1279–1288, 2010.
  • [19] Christopher J Greenshields and Jason M Reese. The structure of shock waves as a test of brenner’s modifications to the navier–stokes equations. Journal of Fluid Mechanics, 580:407–429, 2007.
  • [20] S. Han, M.-J. Kang, J. Kim, and H. Lee. Long-time behavior towards viscous-dispersive shock for Navier-Stokes equations of Korteweg type. arXiv preprint arXiv:2402.09751, 2024.
  • [21] John C Harley, Yufeng Huang, Haim H Bau, and Jay N Zemel. Gas flow in micro-channels. Journal of fluid mechanics, 284:257–274, 1995.
  • [22] C. K. R. T. Jones. Geometric singular perturbation theory. In Dynamical systems (Montecatini Terme, 1994), volume 1609 of Lecture Notes in Math., pages 44–118. Springer, Berlin, 1995.
  • [23] M.-J. Kang and A. F. Vasseur. Contraction property for large perturbations of shocks of the barotropic Navier-Stokes system. J. Eur. Math. Soc., 23(2):585–638, 2021.
  • [24] M.-J. Kang and A. F. Vasseur. Uniqueness and stability of entropy shocks to the isentropic Euler system in a class of inviscid limits from a large family of Navier-Stokes systems. Invent. Math., 224(1):55–146, 2021.
  • [25] M.-J. Kang and A. F. Vasseur. Well-posedness of the Riemann problem with two shocks for the isentropic Euler system in a class of vanishing physical viscosity limits. J. Differential Equations, 338:128–226, 2022.
  • [26] M.-J. Kang, A. F. Vasseur, and Y. Wang. Time-asymptotic stability of composite waves of viscous shock and rarefaction for barotropic Navier-Stokes equations. Adv. Math., 419:Paper No. 108963, 66, 2023.
  • [27] M.-J. Kang, A. F. Vasseur, and Y. Wang. Time-asymptotic stability of generic riemann solutions for compressible navier-stokes-fourier equations. arXiv preprint arXiv:2306.05604, 2023.
  • [28] Yu L Klimontovich. On the need for and the possibility of a unified description of kinetic and hydrodynamic processes. Theoretical and Mathematical Physics, 92(2):909–921, 1992.
  • [29] Yu L Klimontovich. From the hamiltonian mechanics to a continuous media. dissipative structures. criteria of self-organization. Theoretical and Mathematical Physics, 96:1035–1056, 1993.
  • [30] Yu L Klimontovich. A unified approach to kinetic description of processes in active systems. (No Title), 1995.
  • [31] T Koide and T Kodama. Navier–stokes, gross–pitaevskii and generalized diffusion equations using the stochastic variational method. Journal of Physics A: Mathematical and Theoretical, 45(25):255204, 2012.
  • [32] Tomoi Koide and Takeshi Kodama. Navier-stokes equation by stochastic variational method. arXiv: Statistical Mechanics, 2011.
  • [33] Jason M Reese, Michael A Gallis, and Duncan A Lockerby. New directions in fluid dynamics: non-equilibrium aerodynamic and microsystem flows. Philosophical Transactions of the Royal Society of London. Series A: Mathematical, Physical and Engineering Sciences, 361(1813):2967–2988, 2003.
  • [34] Gerald Teschl. Ordinary differential equations and dynamical systems, volume 140. American Mathematical Soc., 2012.