There are several reasons to extend the presentation of Navier–Stokes equations to multicomponent systems. Many technological applications are based on physical phenomena that are present in neither pure elements nor in binary mixtures. Whereas Fourier's law must already be generalized in binaries, it is only with more than two components that Fick's law breaks down in its simple form. The emergence of dissipative phenomena also affects the inertial confinement fusion configurations, designed as prototypes for the future fusion nuclear plants hopefully replacing the fission ones. This important topic can be described in much simpler terms than it is in many textbooks since the publication of the formalism put forward recently by Snider [Phys. Rev. E 82, 051201 (2010)]. In a very natural way, it replaces the linearly dependent atomic fractions by the independent set of partial densities. Then, the Chapman–Enskog procedure is hardly more complicated for multicomponent mixtures than for pure elements. Moreover, the recent proposal of a convergent kinetic equation by Baalrud and Daligault [Phys. Plasmas 26, 082106 (2019)] demonstrates that the Boltzmann equation with the potential of mean force is a far better choice in situations close to equilibrium, as described by the Navier–Stokes equations, than Landau or Lenard–Balescu equations. In our comprehensive presentation, we emphasize the physical arguments behind Chapman–Enskog derivation and keep the mathematics as simple as possible. This excludes, as a technical non-essential aspect, the solution of the linearized Boltzmann equation through an expansion in Hermite polynomials. We discuss the link with the second principle of thermodynamics of entropy increase, and what can be learned from this exposition.
I. INTRODUCTION
Navier–Stokes (NS) equations1 describe the hydrodynamic evolution in time and space of fluids, with ubiquitous applications in nature and technology. Of particular interest are the experiments performed by large laser facilities, to probe extreme states of matter such as encountered in astrophysics2 or to set up configurations for inertial confinement fusion (ICF).3 ICF is an emerging technology aimed at providing the next generation of nuclear plants. It provides fusion energy from the implosion of a capsule filled with deuterium and tritium (DT). In this context, there is a renewed interest in kinetic theory since various conditions are met from hydrodynamic-like to strongly out-of-equilibrium phenomena.
It is a new circumstance that multicomponent diffusion in weakly coupled plasmas must be considered as a part of the hydrodynamic simulation of the ICF capsule implosions.4–6 Multicomponent diffusion is also an essential ingredient of many technological applications, reviewed by Krishna,7 where counter-intuitive phenomena occur as the osmotic diffusion where the presence of a third component leads to an uphill diffusion between the other two components against their concentration gradient.
The Navier–Stokes equations can be derived from the thermodynamics of irreversible processes8 under two principal assumptions: first that the system be close to thermodynamical equilibrium and second that any gradient of thermodynamic quantity be small. Then, the gradients lead to dissipative phenomena acting against them, which appear as fluxes of mass, momentum, and energy. These fluxes are proportional to the gradients, and the coefficients of proportionality are the transport coefficients. Important properties of symmetry of the transport coefficients are obtained from the second principle of thermodynamics about the increase in entropy of an isolated system. Nevertheless, this route toward the Navier–Stokes equations does not provide any criteria of thermodynamic equilibrium nor does it provide recipes to compute the transport coefficients.
The kinetic theory is another route toward the Navier–Stokes equations that draws a relationship between microscopic events and macroscopic behaviors (see Refs. 9 and 10 for instance). This leads naturally to a criterion of equilibrium and to a prescription to compute the transport coefficients. Then, the symmetry of the transport coefficients is warranted by construction. Moreover, the second principle of thermodynamics becomes a consequence of the theory. However, all this is at the cost of a restriction to only deal with systems where the interaction energy between the particles is much weaker than their kinetic energy.
The domain of validity of the kinetic theory encompasses the situations of dilute neutral gas at low density n, where there is rarely more than two particles within the range σ of the interaction potential V(r), so that , and the situations of weakly coupled plasmas at high temperature T, where the effect of the multiple collisions of a particle with the others can be summed in pairs, since the strength of the potential V0 at typical interparticle distances is far less that the mean kinetic energy, (k is the Boltzmann constant).11,12
The cornerstone of the derivation of the Navier–Stokes equations using the kinetic theory is an expansion of the equations with respect to a small parameter: the Knudsen number, which is defined later. This permits to develop analytically the foundations of the theory and to obtain exact results within a controlled validity domain. Unfortunately, the mathematical apparatus of the kinetic theory is quite intricate and it often obscures to the novices the physical principles at the heart of the theory. Here, we present a derivation of the multicomponent Navier–Stokes equations, which emphasizes the physical ingredients and keep the analytical developments as simple as possible, thanks to Snider's recent proposal for a new treatment of multicomponent diffusion.13 Indeed, the complexity inherent in the treatment of multicomponent systems stems from the appearance of a set of N linearly dependent concentrations xi, where N is the number of components i in the mixture. Fortunately, Snider13 proposed to circumvent this difficulty using the set of the independent partial densities ni instead of the concentrations xi.
Another recent breakthrough in the theory is the recent work of Baalrud and Daligault12 on a convergent kinetic equation particularly well adapted to systems close to equilibrium. It is Boltzmann equation with the potential of mean force. To find a closure of the BBGKY hierarchy, they exhibited an expansion parameter independent of the range or the strength of the interaction potential. As a result, their kinetic equation applies equally well to neutral gas and plasmas and admits a particularly large validity domain.14
We start, in Sec. II, by explaining how the fluid equations are obtained as the velocity moments of Boltzmann's equations. It is the occasion to emphasize the central role of the conservation equations of mass, momentum, and energy, in separating slow modes of variations associated with the fluid equations from fast modes associated with the collision integrals.
In Sec. III, the main steps of the Chapman–Enskog derivation are described: the fluid scaling that results in the appearance of the inverse Knudsen number ε in Boltzmann's equation; the separation in order of ε between Euler's and Navier–Stokes equations; and Snider's formulation of the driving forces associated with the different gradients of partial densities, velocity, and temperature, which leads to the corresponding general solution of the linearized Boltzmann equations. The Navier–Stokes equations are then obtained substituting this solution into the fluid equations.
We found that the analysis (made in Sec. IV) of the rate of entropy production highlights the emergence of the dissipation mechanisms associated with the transport coefficients. It is also a useful guide to introduce and define them, providing their properties of symmetry with physical insights, especially with respect to the thermal conductivity.
Section V is dedicated to a concrete illustration of the emergence of the dissipative phenomena as the number of components in a mixture increases from pure elements to binary mixtures, and, finally, ternary mixtures. We examine here how Fourier's and Fick's laws, in their simple form, must be generalized in these situations and we estimate qualitatively the impact of the generalized Fick's law on some ICF configurations.
For completeness, we provide two Appendixes on binary collisions ( Appendix A), the derivation of Boltzmann's equation, and its properties ( Appendix B). Other Appendixes are also provided that give the missing steps of some derivations.
II. FROM BOLTZMANN KINETIC EQUATION TO FLUID EQUATIONS
A. Boltzmann equation
B. Collisional invariants
It is crucial to realize that the collision operator of the Boltzmann kinetic equation represents the net effect between the collisions with species j depleting the distribution function fi of species i at a given velocity and the inverse collisions replenishing the distribution (see Appendix B 1). Since the hydrodynamic equations are obtained as the velocity moments of Boltzmann's equation, we shall see that the only way to get rid of the collision terms is to sum each moment equation over all species in the mixture. By doing so, the conservation of mass, momentum, and energy applies whatever the distribution functions and the resulting fluid equations do not involve collisional (friction) terms. The only link to the distribution functions appears through the expressions of the transport coefficients, which constitute the closures of the fluid equations.
C. Fluid equations
D. Euler's closures
III. CHAPMAN–ENSKOG FORMULATION
A. Fluid scaling at small Knudsen
The fluid scaling of kinetic equations introduces a small parameter, the Knudsen number ε representing the ratio between the small space and time scales of the microscopic processes and the large scales of the macroscopic flows. This allows one to develop a perturbation expansion of the velocity distribution function. Eventually, this perturbation development gives rise to the Navier–Stokes (NS) hydrodynamic equations.
. | Macroscopic . | Microscopic . |
---|---|---|
Length, | Smallest gradient length, L0 | Mean free path, |
Time, | Smallest time scale, T0 | Collision time, τ |
Velocity, | Sound speed, c0 | Thermal velocity, vth |
. | Macroscopic . | Microscopic . |
---|---|---|
Length, | Smallest gradient length, L0 | Mean free path, |
Time, | Smallest time scale, T0 | Collision time, τ |
Velocity, | Sound speed, c0 | Thermal velocity, vth |
The microscopic length scale to compare with L0 is the mean free path λ between two collisions. However, the only space variable of the Boltzmann collision integrals is the impact parameter b, which is of the order of the maximum impact parameter b0 for the collisions with the smallest deflection of the particles.
For weakly coupled plasmas, the maximum impact parameter b0 is of the order of the Debye length λD.11,15 This length scale characterizes the screening of a test charge by the unlike charges piled up around it and the like charges repelled from it. The Coulomb potential of the test charge Q is dressed by this shielding cloud to form a Debye–Hückel (DH) potential . This DH potential is the solution of a Poisson–Boltzmann system of equations, linearized with respect to . This linearization is a good approximation in the validity domain of the kinetic theory, often characterized by large values of the parameter , which represents the number of charges in a Debye cube. In these conditions, each charge interacts simultaneously with many other charges in a Debye sphere. Most of these collisions gives rise to razing diffusion and the mean free path λ is defined as the typical distance, where the cumulative effect of the different collisions is associated with a substantial deflection.11
In another equivalent definition, the mean free path λ is defined from a collision frequency νc itself defined from a multi-fluid approach to the hydrodynamic equations.16 When Maxwellian distribution functions of species with equal temperature but different mean velocities are introduced in the collision integrals, the velocity moments of the kinetic equations give rise to friction terms that can be put under the form to define a collision frequency for the interaction between species i and j.
B. Chapman–Enskog ansatz
The formulation proposed by Chapman and Enskog17 starts from the Boltzmann equation in the reduced units of the fluid scaling. It then introduces a close relationship between the expansion of the velocity distribution functions according to the order in ε, the Knudsen number, and the hydrodynamic equations: Euler's equations control the leading order, and the NS equations are associated with the next-to-leading order. This important point is often obscured by the complicated mathematical apparatus accompanying the calculation of the transport coefficients, i.e., the development of the solution in orthogonal Sonine polynomials. In the following, the derivation procedure is kept as simple as possible principally because we do not present the practical calculation of the transport coefficients, but only their expressions as functionals of the solution of the kinetic equations. The final expressions are naturally translated in dimensional units by equating ε to 1.
C. Snider's approach to driving forces
D. Solution of the linearized Boltzmann equation
The first argument to put forward when solving the linearized Boltzmann equation is that space and time variables, r and t, do not appear explicitly in Eqs. (18)–(20). The solution depends on them through the macroscopic variables only: , and .
The second argument concerns the velocity variable v of the distribution functions fi, which only appears in the combination in Eq. (20). In Eq. (19), changing v for c just amounts to warrant the Galilean invariance of the binary collisions. As a consequence, the distribution functions fi depend on velocity through the variable c.
Traditionally, the unknown functions , and are developed on a basis of Sonine polynomials, since at the time of these developments the digital computer was not discovered yet. We shall not dwell with the solution of this system, but we shall assume known the solution and derive its relationship with the closure of fluid equations.
IV. NAVIER–STOKES EQUATIONS
With the solution of the linearized Boltzmann equation expressed as a function of the gradients of densities ni, fluid velocity u, and temperature T, Eq. (24), the closure relations of the fluid equations, Eqs. (5e), (6b), and (7c), can be computed as functions of these gradients. The final results represent the constitutive relations known as Fick's law, Fourier's law, and Newton's law, with their associated transport coefficients of diffusion, thermal conductivity, and viscosity, respectively.
As a first step, it is useful to compute the rate of entropy production in order to identify the different dissipation mechanisms that involve the transport coefficients.
A. H-theorem and rate of entropy production
B. Thermal and mass diffusions
C. Viscosity
D. Thermal conductivity
V. EMERGENCE OF PHENOMENA IN MULTICOMPONENT MIXTURES
We shall see, in this section, how new dissipative phenomena appear as the number of components in a mixture increases from the case of pure elements, to binary mixtures, and beyond three species. For a fluid made of only one component, there are only the transport coefficients of viscosity η and thermal conductivity λ. For a mixture, there appear additional transport coefficients of thermal diffusion DTi and mutual diffusion Dij. For mixtures with more than two components, the interdiffusion coefficients Dij exhibit complex behaviors that goes beyond Fick's law in its simple form.
A. Pure elements
B. Binary mixtures
C. Ternary mixtures
1. Case A:
2. Case B: and
D. ICF applications
Very recently, the ICF program has taken an important step forward the ignition quest.18–20 Ignition is a regime where the energy produced by the fusion reactions between deuterium and tritium (DT) outperforms the leakage due to thermal conduction and Bremsstrahlung radiation so as to propagate the burn from the hot spot to the remaining fuel. To reach this regime in ICF, the experimental setup consists in heating a gold cavity (named Hohlraum) by many UV-laser beams to convert the incident energy into x-ray radiation which in turn ablates the DT-filled capsule. This leads to the implosion of the capsule by a rocket effect bringing mechanical work that compresses and heats the DT fuel.
In these recent experiments, a burning plasma regime was achieved where the fusion energy was a heat source larger than the mechanical work putting the configuration on the verge of ignition. Many obstacles had to be challenged:21–23 the Rayleigh' Taylor instabilities and a degraded symmetry of implosion can lead to mix compressed fuel with capsule material that rapidly cool it by enhanced radiation leakage; the filling of the Holhraum by the expanding bubbles of laser-irradiated gold prevent the laser beams to propagate efficiently; the backscatter of energy out of the Hohlraum by laser–plasma interaction; and so on.
In the following numerical applications, we shall qualitatively examine the role of multicomponent diffusion at the fuel–ablator interface and in a Hohlraum filled with helium (He) gas to mitigate its closure when expanding gold meets expanding capsule ablator. These numerical applications need to estimate the diffusion coefficients in thermodynamic conditions of high temperatures and varying densities. Several modelings of transport coefficients are available in the literature (see, for instance, Refs. 24–28). It is worth noticing that Kagan and Baalrud29 provided Matlab routines to compute viscosity and heat diffusion coefficients in the weakly coupled regime. We decided to use the simplest modeling, the Pseudo-Ion in Jellium (PIJ) model,25 that has been recently compared with molecular dynamics simulations of many ternary mixtures, including the plastic compound of carbon and hydrogen (CH), often chosen as ablator, mixed with silver (Ag) impurities.30,31
1. Separated reactants
An interesting experiment was dedicating to assess any mix between the DT fuel and the ablator.32 The idea was to separate the D and T reactants in the experimental setup. The DT fuel was replaced by a pure T gas, and a thin layer of CD was buried into the CH ablator. Depending on the depth where this CD layer sits, the DT reactions were observed or not. This was thought to be a good indicator of the amount of CH ablator that mixes with the DT fuel. However, the most convincing interpretation5 of the measurements requires to invoke multicomponent diffusion. This interpretation evidence that the atoms of D diffuse much more rapidly toward the T gas than the atoms of C. This can be illustrated using the oversimplified situation of case A of Sec. V C 1. Indeed, one can assume that the gradients of concentration of C and D are equal, at least as first approximation. Table II shows that the diffusion coefficient of D is around 50 times higher than the one of C, leading to a length of diffusion of D through T roughly times longer than the one of C.
2. He-filled Hohlraum
Another situation where multicomponent diffusion may be important in a ternary mixture is the plasma collision between the expansion of gold and of CH ablator within a Hohlraum filled with He gas.33 The presence of He is supposed to act as a cushion (“airbag”) against Au and CH to prevent the plasma collision within the Hohlraum and to leave room for the laser beams to propagate. Here, it is the case B of Sec. V C 2 that represents a pertinent albeit simplified description, assuming a null gradient of concentration of He. Interestingly, Table III indicates that, even without any He gradient, there is a mass flux of He toward the ablator (assumed here to be pure C to avoid examining a quaternary mixture). We think that this process deserves more investigation and a more in-depth analysis. Unfortunately, to our knowledge, this issue has not yet been addressed in any hydrodynamic simulations inclusive of multicomponent diffusion.
VI. CONCLUSION
The route, traced by Chapman and Enskog17 from the description of the binary collisions and the Boltzmann kinetic equations to the Navier–Stokes equations, is a long one. However, it is very instructive to follow it in order to highlight the most important physical arguments and assumptions.
The principal pillar of this edifice is the Knudsen number, ε. Its very definition is possible when the kinetic theory offers us the opportunity to define microscopic scales of time and space: the collision time τ and the mean free path λ.
The criterion for reaching thermodynamic equilibrium can then be formulated: one can consider in equilibrium a volume, which is homogeneous over distances much larger than λ and which is left out any solicitations for a time much greater than τ. This criterion applies equally well out of the validity domain of the kinetic theory, at least in order of magnitude.
With the Knudsen number, ε, the assumption of small gradients of the flow variables, which is invoked in the derivation from the thermodynamics of irreversible processes, becomes an operational criterion, since it can be checked that the macroscopic scales of time and length, T0 and L0, associated with these gradients, are indeed much larger than τ and λ. The Euler equations can be used to estimate these gradients, for this comparison.
If ε is small, and the conditions of applicability of the kinetic theory are met, the Boltzmann equations can be linearized with source terms, known as driving forces, arising from the decoupling of orders in ε. These driving forces dictate the form of the general solution and of the transport coefficients.
However, the edifice is very fragile. One often assumes that the Knudsen number is small without further verification. The difficulties can arise when the thermodynamic equilibrium is not complete.
In ICF, the ions and the electrons of the plasma must often be considered at different temperatures, for instance. As discussed in the seminal paper by Braginskii,34 this assumes that both distribution functions of ions and electrons relax to two Maxwellians at different temperatures, Ti and Te, quicker than the time required for the relaxation between Ti and Te. The fluid equations must, therefore, include two equations for each energy of the ions and the electrons. As a result, additional terms appear coming from the collision integrals evaluated with Maxwellians at Ti and Te. These terms warrant that the system shall relax to a common temperature. In most situations, the resulting Euler equations are then dominated by the time scale τie of temperature relaxation. Due to the large mass ratio between ions and electrons, it is, however, possible to define a small Knudsen number in order to get linearized Boltzmann equations with driving forces as the source terms.35
In other circumstances, there is a decoupling of velocity. For instance, when two fluids meet at different velocities. This has an impact on the equations of momentum and energy conservation with additional contributions coming from the collision integrals evaluated with Maxwellians centered at different velocities. The issue of defining a Knudsen number in this case is much more involved. Different approaches are developed following either the thermodynamic route36 or the kinetic one37 and confronting the hydrodynamic approximation to more detailed descriptions requiring microscopic simulations.
AUTHOR DECLARATIONS
Conflict of Interest
The authors have no conflicts to disclose.
Author Contributions
Philippe Arnault: Writing – original draft (equal). Sébastien Guisset: Writing – original draft (equal).
DATA AVAILABILITY
Data sharing is not applicable to this article as no new data were created or analyzed in this study.
APPENDIX A: BINARY COLLISIONS
It is convenient to define the origin of time t and angle θ at the turning point. Indeed, the trajectory is symmetric with respect to this point, with a value of the angle before collision, when and after collision, when . The deflection angle χ is then given by
APPENDIX B: BOLTZMANN COLLISION INTEGRAL
1. Empirical derivation
Boltzmann gave an empirical derivation of his kinetic equation by considering the net effect between the collisions depleting the distribution function at a given velocity and the inverse collisions replenishing the distribution.
In a fluid, being a dilute gas or a plasma, many binary collisions have to be considered. In Appendix A, it is shown that a binary collision can be defined by the initial velocities and of the two colliding particles of species i and j and by their impact parameter b. Equation (A9b) gives the final velocities and in this collision. We shall compute the rate of these collisions, in the fluid.
2. Moments
3. Linearized operator
4. Rotational invariance
The general form of the solution to the linearized Boltzmann equation is given in Eq. (24) using an argument of rotational invariance, which we describe in detail in this section.
Assume that we rotate the reference frame of velocities. We note the functions of vectors with coordinates in the new frame with a tilde, , whereas the same functions in the old frame are noted without tilde, .
APPENDIX C: MOMENTUM AND ENERGY EQUATIONS
APPENDIX D: DRIVING FORCES
APPENDIX E: BRACKET INTEGRALS
The rate of entropy production discussed in Sec. IV A involves bracket integrals, which are developed in the following.
1. Temperature gradient
2. Velocity gradients
3. Partial density gradients
4. Temperature and density gradients
APPENDIX F: TENSOR INTEGRALS
-
Case : (Txy; Txz; Tyz)
All the terms other than and vanishes as integrals over odd functions of the velocity components. Since is symmetricand since cx, cy, and cz are integration variables and F(c) is invariant over any interchange of cx, cy, and cz, the three components are equal to -
Case i = j: (Txx; Tyy; Tzz)
All the terms other than vanishes as integrals over odd functions of the velocity componentswith the second integral already worked out in the case and using the fact that is traceless. Since cx, cy, and cz, are integration variables and F(c) is invariant over any interchange of cx, cy, and cz, the first integral can be evaluated using spherical coordinates with ci = czWith this last result, Eq. (F2) is proven.