Traveling Wave Solutions to Brenner-Navier-Stokes-Fourier system
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 perturbation1991 Mathematics Subject Classification:
76N15, 35Q30, 35C07
1. Introduction
We consider the so-called Brenner-Navier-Stokes-Fourier(BNSF) system (in the Eulerian coordinates) in 1D :
(1.1) |
Here, is the density, is the mass velocity, is the volume velocity, is the pressure, and is the internal energy. In addition, and 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 which is the classical concept; the other is the volume velocity . 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, , 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 and suggested by Brenner is as follows:
(1.2) |
where is the specific heat at constant pressure. The difference between the two velocities is proportional to . 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 to denote the mass variable )
(1.3) |
where is the specific volume, is now the volume velocity, is the absolute temperature, is the total energy, and 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 and the internal energy function are given by
(1.4) |
where is the gas constant and is the adiabatic constant.
In this article, we prove that the system (LABEL:main) admits viscous shock waves connecting two end states and when these two end states are close enough and satisfy the Rankine-Hugoniot condition and Lax entropy condition as follows:
(1.5) | ||||
and either , and or , and holds. |
Here, and satisfy
In other words, for given sufficiently close constant states and satisfying (LABEL:end-con), we prove the existence of a viscous shock wave ()(), as a traveling wave solution to the following system of ODEs:
(1.6) |
Here, if , ()() is a 1-shock wave with velocity , where . If , it becomes a 3-shock wave with .
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 , there exists a constant such that for any right-end state satisfying (LABEL:end-con) and , there is a unique traveling wave solution to (LABEL:main), as a monotone profile satisfying (LABEL:shock_0), which more precisely satisfies (1.13) when , or (1.14) when .
The second result provides quantitative estimates of traveling wave solutions. Especially, we present estimates for the ratio between the components 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 , there exist positive constants , , and such that for any right-end state satisfying (LABEL:end-con) and , the following holds.
Let be the monotone solution to (LABEL:shock_0) with .
Then,
(1.7) | |||
(1.8) | |||
(1.9) |
It also holds that for all . More explicitly, we have the following:
(1.10) | ||||
(1.11) |
which satisfies
(1.12) |
where
or
In addition, if we consider the 1-shock, i.e., , then we have
(1.13) |
and for the 3-shock case, i.e., ,
(1.14) |
Furthermore, the following estimate holds:
(1.15) |
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 and . Thus, in the sequel, we only consider the 3-shock case, i.e.,
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 -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 and use the second equation of it, to find the system of ODEs of first order:
(2.1) |
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 , and (1.11) from the following system of two equations on variables only:
(2.2) |
where the above system can be derived from (2.1), and the second order term is due to the right-hand side of . 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 variable only, like (3.19) that could be derived from (2.2). We may (formally) introduce a new variable defined by , where the scale is slow while is fast. Then, by setting , , , and so , and , 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 :
(2.3) |
where , . Here, we assume and to be smooth on a set , where is open and is an open interval containing . Following [22], we say an -dimensional manifold is a critical manifold of the system (2.3) if each satisfies that . To state Fenichel’s first theorem, we need the following definitions:
Definition 2.1.
We say that the manifold is normally hyperbolic relative to (2.3) if the matrix
has eigenvalues (counting multiplicity) with nonzero real part for each .
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 is locally invariant under the flow from (2.3) if there is an open set containing such that for any and any , the trajectory starting at with also satisfies . (When , replace with ).
Note that, if is normally hyperbolic, we can express it locally as a graph of in terms of thanks to the implicit function theorem. However, in the following discussion, we only deal with the case when can be globally expressed as a graph of in terms of on a compact domain , 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 . Suppose that there is a smooth function of which graph is contained in , i.e.,
where is a compact and simply connected smooth set. Then, there exists such that for each with , there is a function which satisfies the followings: for each , the graph
is contained in an -dimensional manifold which is locally invariant under the flow of (2.3). Moreover, can be taken to be for each , jointly in and and on where denotes at .
Remark 2.1.
Since the family of functions is regular in as well, a consequence of Proposition 2.1 is that there exists a function with two variables and which satisfies
Note that the function has the same regularity with ; i.e. is a -function for each , jointly in and .
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 small enough (to be determined later), we consider the system of ODEs :
(3.1) |
where and
(3.2) | ||||
Here, is defined by
(3.3) |
It is worth mentioning that the function is smooth for any in some open interval containing , and hence so is the system (3.1). Indeed, this can be verified by the argument below. We observe that can be rewritten as follows:
From (3.3), we have
(3.4) |
Hence, it follows that
which is a smooth function near .
We then find critical points of the system (3.1) for . If is a critical point of the system, then from (3.2) together with the relation (3.3), it holds that
which proves that there are only two critical points for any .
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 and 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 and . Equivalently, we have and
so that
(3.5) |
Thus, the critical manifold of (3.1) can be parametrized by as follows:
which forms a curve in -dimensional space. Here,
is a constant depending only on the type of gas.
To use Proposition 2.1, it is required to verify that is normally hyperbolic relative to the system (3.1). Considering the definition of normally hyperbolicity, we need to compute
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 is normally hyperbolic.
We now consider a closed interval containing . From Proposition 2.1 and Remark 2.1, for each small we obtain an 1-dimensional locally invariant manifold and smooth functions and defined on , which satisfies
(3.6) |
where the functions and are sufficiently regular jointly in and , in particular, . Step 1: We will show that the system (3.1) has a solution connecting the two critical points and . Here, we will prove the existence of a solution defined on to the system (3.1) passing through a point
which stays on and connects the two points and .
Since at for sufficiently small and as , there exist with such that and on . Thus, by the succeeding lemma, we have a unique solution of the following ODE
(3.7) |
subject to the initial datum
(3.8) |
and the solution is increasing and converges to and as .
Lemma 3.1.
Let with , on , and fix . Then, there is a unique function satisfying and , which is increasing and converges to and as .
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 , we let
(3.9) |
and we will prove that the function satisfies (3.1).
For any fixed , the point is on the manifold .
From the Cauchy-Lipschitz theorem, we have a unique Lipschitz solution (locally defined near ) to the system (3.1) passing through at , i.e., .
Then, since by (3.9) and the local solution is continuous near , it holds from Proposition 2.1 that the local curve is contained in the locally invariant manifold , that is,
(3.10) |
From the definition of , we have , and so is a local solution of (3.7). Then, by the uniqueness of (3.7) with , we can deduce that near . Furthermore, since both and are parametrized by and as in (3.9) and (3.10) respectively, those are locally the same as well. This proves that satisfies (3.1) at . Since was chosen arbitrary in , the globally defined function is a solution to the system (3.1).
Lastly, we prove that connects the critical points. Since converges at , and also converge at by its constructions. Hence, the solution has a limit at . Notice that the limits of a solution of an ODE system at should be its stationary points. Thus, converges to and at because there are only two critical points of (3.1) as we observed in advance.
Step 2: In Step 1, for sufficiently small , we proved the global existence of (3.1) with
Especially, it holds from (3.1) that
(3.11) |
In this step, we will show that the function in (3.11) gives a solution to the following system:
(3.12) |
and thus, to the desired system (LABEL:shock_0).
From now on, we consider the parameter as the amplitude of the shock . First of all, we claim
Given (LABEL:end-con), we have , so that
(3.13) |
Multiplying (LABEL:end-con)2 by , and substracting it from (LABEL:end-con)3 gives
which implies
(3.14) |
From (3.13) and (3.14), we obtain the expression for :
(3.15) |
From (LABEL:end-con)1, and have the opposite sign.
Therefore, we obtain that
(3.16) |
Plugging (3.16) into (LABEL:end-con)1, we conclude that
(3.17) |
as desired.
Now, we define by
(3.18) |
Multiplying on both sides of (3.11) and using (3.18), we have
(3.19) | ||||
which can be reduced to
(3.20) | ||||
Then we define and in accordance with (3.12)1 and (3.12)2 respectively as follows:
(3.21) |
and
Since we have
(3.22) |
can be written in the following way:
(3.23) |
Hence, differentiating both sides of (3.23), we get
(3.24) | ||||
Plugging (3.21), (3.23), and (3.24) into (3.20), we obtain that
(3.25) |
Then, multiplying (3.12)2 by , and adding it to (3.25) yields (3.12)3:
Furthermore, since is increasing, the monotonicity of follows from (3.18). The monotonicity of and 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 to . Now, we will prove the uniqueness of the monotone solution to (3.12) connecting two endpoints to .
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 , where and is a critical point, i.e., . Let be the unstable manifold composed of all points converging to for .
Let (resp. ) denote the linear subspace spanned by eigenvectors of corresponding to negative (resp. positive) eigenvalues.
Suppose has no eigenvalues on the imaginary axis.
Then, there are a neighborhood of and a function such that
Here, both and its Jacobian matrix vanishes at .
This implies that the set is a -manifold of which dimension is equal to the dimension of , and this is tangent to the affine space at .
To apply Proposition 3.1, we rewrite the system (3.12) by using (3.25) as follows:
(3.26) |
It is easy to see from (LABEL:end-con) that 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 so as to determine the dimension of the unstable manifold. For (3.26), the Jacobian matrix at is given by
(3.27) |
We then compute the trace and determinant of (3.27) as follows:
Here, we utilized (3.4) for the first inequality. Upon examining the graph of the characteristic equation with , 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 .
Now, we prove the desired uniqueness. We already constructed a monotone solution which satisfies (3.12) and connects to . This solution trajectory is locally contained in the one-dimensional unstable manifold . More precisely, must coincide with the trajectory near . Note that, any other solution which connects to should be on , and especially if it is monotone, it should intersect with the trajectory . Hence, by the uniqueness for the Lipschitz autonomous system of ODE, 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, represents a positive constant that may vary from line to line, yet remains independent of .
To prove the desired estimates as in Theorem 1.2, we will proceed in the following order. First, we obtain . 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) |
Similarly, we also have
(4.2) |
On the other hand, from (3.6), we have
(4.3) |
for some compact set , and is smooth.
So, we get . Then, using (4.1)1, we obtain .
Following this, we differentiate (4.3):
Then, since is continuous and is confined to , it holds that . This together with (4.1) implies that .
To obtain the bounds for and , we differentiate (4.3) once and twice more:
Then, since all the higher derivatives of are continuous and is confined to as well, we have and .
Hence, from (4.2), we also obtain and .
Thus, by differentiating (3.22) and (3.24), we use all of these estimates to yield that
We now show (1.10) and (1.11): from (LABEL:shock_0)1, it follows that
Then, (1.12) gives the desired result (1.10). Likewise, for (1.11), we use (3.24) to get
Since is increasing, we have . Considering this and (1.12), the definition of gives the desired one (1.11). Furthermore, these two estimates with imply that 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) |
Afterwards, we analyze both sides of (LABEL:vODEre) respectively. For the left-hand side, we use (3.3) to have the following:
(4.5) | ||||
We then observe that all the terms on the right-hand side can be controlled by . Indeed,
(4.6) | ||||
Combining (LABEL:vODEre) with (4.5) and (4.6), we obtain (1.15), i.e.,
With this, the standard argument of Gronwall’s inequality (for example, see [23, Lemma2.1]) gives that
(4.7) |
Then, by plugging these estimates into (1.15), we get
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
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.