Sunteți pe pagina 1din 22

Contemporary Physics

ISSN: 0010-7514 (Print) 1366-5812 (Online) Journal homepage: http://www.tandfonline.com/loi/tcph20

Beyond Navier–Stokes equations: capillarity of


ideal gas

A. N. Gorban & I. V. Karlin

To cite this article: A. N. Gorban & I. V. Karlin (2016): Beyond Navier–Stokes equations:
capillarity of ideal gas, Contemporary Physics, DOI: 10.1080/00107514.2016.1256123

To link to this article: http://dx.doi.org/10.1080/00107514.2016.1256123

Published online: 24 Nov 2016.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tcph20

Download by: [University of Otago] Date: 24 November 2016, At: 21:54


CONTEMPORARY PHYSICS, 2016
http://dx.doi.org/10.1080/00107514.2016.1256123

Beyond Navier–Stokes equations: capillarity of ideal gas


A. N. Gorbana and I. V. Karlinb
a Department of Mathematics, University of Leicester, Leicester, UK; b Department of Mechanical and Process Engineering, ETH Zürich, Zürich,
Switzerland

ABSTRACT ARTICLE HISTORY


The system of Navier–Stokes–Fourier equations is one of the most celebrated systems of equations Received 25 July 2016
in modern science. It describes dynamics of fluids in the limit when gradients of density, velocity Accepted 31 October 2016
and temperature are sufficiently small, and loses its applicability when the flux becomes so non- KEYWORDS
equilibrium that the changes of velocity, density or temperature on the length compatible with the Kinetics; fluid dynamics;
mean free path are non-negligible. The question is: how to model such fluxes? This problem is still Boltzmann; capillarity;
open. (Despite the fact that the first ‘final equations of motion’ modified for analysis of thermal creep invariant manifold; thermal
in rarefied gas were proposed by Maxwell in 1879.) There are, at least, three possible answers: (i) use creep
molecular dynamics with individual particles, (ii) use kinetic equations, like Boltzmann’s equation or
(iii) find a new system of equations for description of fluid dynamics with better accounting of non-
equilibrium effects. These three approaches work at different scales. We explore the third possibility
using the recent findings of capillarity of internal layers in ideal gases and of saturation effect in
dissipation (there is a limiting attenuation rate for very short waves in ideal gas and it cannot increase
infinitely). One candidate equation is discussed in more detail, the Korteweg system proposed in
1901. The main ideas and approaches are illustrated by a kinetic system for which the problem of
reduction of kinetics to fluid dynamics is analytically solvable.

1. Introduction • 1845 George Stokes published the Navier–Stokes


equations.
1.1. Timeline (the history of equations)
• 1872 Ludwig Boltzmann published ‘Kinetic Theory
Historical timelines are seldom comprehensive and are of Gases’ and introduced the Boltzmann equation
often subject to debates but they do provide a background for evolution of the particle distribution in the space
to examine the present. Older chronology seems to be of possible positions and momenta.
more reliable and less disputable (it may happen because • 1879 James Clerk Maxwell used Boltzmann’s kinet-
some details went forgotten). For understanding the sit- ics to come up with additional terms in the Navier–
uation ‘beyond Navier–Stokes’ the following milestones Stokes equations (the ‘final equations of motion’), in
from the distant past are crucial: order to explain the thermal creep in rarefied gases
which cannot be captured by the Navier–Stokes–
• 1687 Isaac Newton stated that the shear stress be- Fourier equations.
tween layers of a fluid is proportional to the velocity • 1893 Johannes Diderik van der Waals introduced
gradient in the direction perpendicular to the layers. the thermodynamic theory of capillarity under the
• 1738 Daniel Bernoulli proved that the gradient of hypothesis of a continuous variation of density.
pressure is proportional to the acceleration. • 1900 David Hilbert presented his prominent list of
• 1759 Leonhard Euler applied Newton’s second law problems. The sixth Hilbert problem includes ‘the
of motion to fluid dynamics (no viscosity) and pub- problem of developing mathematically the limiting
lished the Euler equations of fluid motion. processes, ... , which lead from the atomistic view to
• 1807 Jean Baptiste Joseph Fourier presented his heat the laws of motion of continua’.
conduction equation to the Institut de France. • 1901 Diederik Johannes Korteweg proposed the
• 1822 Claude Navier introduced viscosity in the dynamics equations for fluids with capillarity, based
Euler equation. on van der Waals’ approach.
• 1823 Augustin-Louis Cauchy published his general • 1917 David Enskog defended his thesis in which he
theory of stress. gave a systematic derivation of the Navier–Stokes–

CONTACT A. N. Gorban ag153@le.ac.uk


© 2016 Informa UK Limited, trading as Taylor & Francis Group
2 A. N. GORBAN AND I. V. KARLIN

Fourier equations from the Boltzmann equation and


predicted the thermodiffusion effect in the mixtures
of gases. The latter was used in early uranium sepa-
ration technologies.
• 1932 Steven Chapman and Steven Cowling refined
the method of Enskog and wrote the classical treatise
‘Mathematical theory of non-uniform gases’.
Thus, at the beginning of the twentieth century, the
main equations we intend to discuss were invented. The
Navier–Stokes–Fourier equations and the Boltzmann
equation were known and Hilbert considered reduction
from kinetics (‘the atomistic view’) to fluid mechanics as
a crucially important problem. Van der Waals proposed
to include in the energy density a new term, ∼ c(∇ρ)2 ,
where ρ is the density and c is the capillarity coeffi-
cient [1]. His doctorate student, Korteweg, found how
this term affects the motion of fluids, and created the
dynamics of fluids with capillarity [2]. Figure 1. Crookes radiometer. It is made from a glass bulb with low
Since then the research tree has branched enormously. air pressure inside. Within the bulb there is a rotor with several
The van der Waals capillarity term played an essential role vertical metal vanes, which are polished on one side and black on
in a number of important achievements: in the Landau the other. When exposed to light the vanes turn, the dark sides
theory of phase transitions [3] and the Ginzburg–Landau retreating from the radiation source and the light sides advancing.
(Public domain picture.)
theory of superfluids and superconductors [4], in Cahn–
Hilliard models of phase separation [5] and in the Langer
Bar-On and Miller theory of spinodal decomposition [6],
1.2. The problem
to name a few. Many new equations of motion were in-
vented and studied for various continuous media, In 1946, John von Neumann stated that computational
and a special discipline, rational thermodynamics, was fluid dynamics would make experimental fluid dynamics
developed to manage this world of continuum mechanics obsolete [11] (see also [12]).
models [7,8]. In particular, the Korteweg equations were The computational approach to the problem as dis-
slightly revised by means of rational thermodynamics [9]. cussed by von Neumann has been developing in many
The fluid dynamics equations were rigorously derived relevant works published meanwhile including e.g. very
from Boltzmann’s kinetics in some rescaling limits: if we recent publications [13,14].
consider a flow of an almost equilibrium gas with very But rephrasing the famous quotation of Lord Kelvin
small space derivatives of velocities and density and then [15] we can say that the beauty and clearness of the the-
change the time scale (go to ‘slow time’) then the Euler ory, which explains fluid dynamics equations as special
equations hold for this flow with high accuracy. For some limits of kinetic equations, is at present obscured by a
scaling procedures, the Navier–Stokes–Fourier equations cloud. Higher order corrections to the Navier–Stokes–
can be achieved as well. (Exact statements, more details, Fourier equations produced from kinetics beyond the
more results and further references can be found in the degenerated scaled flows demonstrate some non-physical
book [10].) properties and cannot be used without regularisation.
Several families of methods were developed in theo- Moreover, there are relatively simple experiments with
retical physics for deduction of fluid dynamic equations rarefied gases (known to Reynolds and Maxwell [16,17])
from kinetic equations. We introduce some of them later. which can be captured by neither the Navier–Stokes–
They usually work well for the derivation of the ‘expected’ Fourier equations nor by the higher order corrections
Euler and Navier–Stokes–Fourier equations but the next obtained from the Boltzmann equations by regular pro-
step, a correction to the Navier–Stokes equations for cedures. These are, for example, effects of thermal creep
more non-equilibrium regimes, was far less successful. (or transpiration), i.e. the steady motion of a rarefied gas
Therefore, the most challenging goal of these methods, induced by a temperature gradient on the boundary of
the correct new post–Navier–Stokes–Fourier equations, the flow domain, which are demonstrated by the Crookes
remained unachieved. radiometer (the light mill, Figure 1).
CONTEMPORARY PHYSICS 3

Another example of such a system is the Knudsen The exact solutions demonstrate that the van der
pump [18]. In Knudsen’s experiments, it was a pipe Waals capillarity emerges in the non-equilibrium ideal
with alternating narrow (diameter 1/3 mm) and wide gas. Therefore, Korteweg’s equations can be a candidate
(diameter 10 mm) segments of 5 cm length (Figure 2). for the post–Navier–Stokes–Fourier equations. The fact
Every second pipe joint was heated by a special heat- that the van der Waals capillarity energy emerges from a
ing element. Metallic wire was used for heat removal rigorous analytic solution of the reduction problem for
from other joints. The temperature difference between kinetic models of ideal gas was unexpected.
the heated and unheated pipe joints was 500 ◦ C. For On the other hand, the exact solutions demonstrate
normal pressure, there was no difference in pressure at saturation of dissipation: the attenuation rate of the
the opposite ends of the pipe, whereas for the pressure at acoustic waves has a finite limit when the wave length
one end, p1 ∼ 0.5 mmHg (65 Pa) the ratio between these tends to zero [24,27]. In that respect, the original
pressures is ∼ 10. Recently, this system was revisited, Korteweg’s equation may be not sufficient: for them,
both by the reproduction of experiments and by theoret- the attenuation rate grows as (frequency)2/3 (for high
ical analysis [19]. The effect was also demonstrated on frequency, see (26) and Figure 7(b)). Two different
two joined pipes of different inner diameters, 12 and 24 hypotheses seem to be possible at present:
mm, each 60 mm in length. The joint was heated to about • For the genuine Boltzmann equation with realistic
500 ◦ C, the ends were cooled by thick copper plates. The interaction between the particles, the asymptotic of
flux from the thinner to the thicker pipe was indicated the attenuation rate of acoustic waves is the same as
for the pressure 5–50 Pa. At pressure 100 Pa (and above) for the Korteweg equations;
there was no motion observed in this system [19]. • The proper hydrodynamics for non-equilibrium gas
Of course, these effects do not contradict the kinetic should include the mixed derivatives, which look
theory and can be explained in the framework of Boltz- like the operator (1 − α)∂t in the left-hand side of
mann’s kinetics and some of its simplifications [20,21]. the equations (in the simplest form), where  is the
The problem is in the appropriate continuum mechanics Laplace operator.
model, which is produced systematically from kinetics
and explains thermal creeps. The prophecy of von Neu- There are some arguments in favour of the Korteweg
mann is unfulfilled not only because the computational asymptotic reported very recently [31]. Nevertheless, the
tools are insufficient, but also because the proper equa- second hypothesis is strongly supported by the exactly
tions for some situations remain unknown. solved problem, as we demonstrate below.

1.3. Solutions proposed: capillarity of ideal gas 1.4. Capillarity of ideal gas in words and numbers
Van der Waals introduced capillarity energy for non- In high school textbooks and popular science literature,
ideal multiphase fluids [1,22]. However, the terms which capillarity is the ability of a liquid to flow in narrow gaps
look similar to Korteweg’s stress tensor have been recog- without the assistance of, and in opposition to, external
nised in the Chapman–Enskog expansion for non- forces such as external pressure or gravity. In scientific
equilibrium ideal gas many times. Everyone can compare literature, the term ‘capillarity’ is used in wider sense.
the Korteweg stress (Equation (11) below) with Burnett’s For example, the pressure difference across the curve
equations, the first post-Navier–Stokes correction de- interface is termed the ‘capillary pressure’. Gibbs theory
rived from the Boltzmann equation (see, for example, of capillarity [33] carries the title ‘Influence of Surfaces
the Chapman and Cowling book [23]). of Discontinuity upon the Equilibrium of Heterogeneous
Nevertheless, it was not before the exact summation Masses’. This title includes two important ideas: equilib-
of the entire Chapman–Enskog series was found [24–27] rium and discontinuity. On the contrary, van der Waals
that the following rule was discovered [28,29]: Chapman– [1,22] proposed the theory of continuous transition layers
Enskog ⇒ Viscosity + Capillarity. which was developed later in the dynamic phase field
The problem with Burnett’s equations is, among oth- theory and numerics. In the limit of the thin equilibrium
ers, in the wrong sign of the higher order correction terms film, the van der Waals theory gives results very similar
leading to a parasite instability of short waves (Bobylev’s to the Gibbs theory.
instability [30]). Maxwell, in his attempt to describe ther- Gibbs studied energy and equilibria of interface sur-
mal creep, encountered the wrong sign of the effect (ac- faces. We can call them infinitely ‘thin films’, films with-
cording to Burnett’s equations the light mill rotates in out thickness. Van der Waals and, later, his student
the direction opposite to the experimentally observed Korteweg studied continuous media with energy depen-
rotation). dent on density gradients. We can call the van der Waals
4 A. N. GORBAN AND I. V. KARLIN

Figure 2. Knudsen pump. Bold lines correspond to thick pipes. The direction of flux (for open pipe ends) is indicated. For closed pipe ends,
p1 > p2 . Original Knudsen’s experimental data for stationary pressure are in the table. The plot of stationary temperature difference δT
from non-heated pipe joints is presented schematically.

transition layers ‘thick films’, they have non-negligible


thickness and volume. Now, we aim to demonstrate that
even very simplified piecewise continuous approxima-
tion of a thick film can give useful insights.
Consider two chambers with fluid connected by a
capillary tube (Figure 3). Assume that a system of heaters Figure 3. Two chambers connected by a capillary tube.
and coolers keeps the temperatures, θ1 > θ2 , in these
chambers constant in time. In a steady state, there is no on non-equilibrium thermodynamics [36]. All of them
fluid flux through the tube but a heat flux exists and so include unknown parameters, which should be found
this is not a thermodynamic equilibrium. Two limit cases experimentally. Here we will demonstrate two simple
for gases are very well known: models based on capillarity (perhaps, the simplest mod-
• The Navier–Stokes limit, the mean free path λ  d. els). First of all, assume that the thermomechanical effect
In this limit zero flux implies zero pressure differ- is produced by a thin layer at the wall of the capillary. The
ence. In the steady state P1 = P2 ; property of this surface depends on the thermodynamic
• The Knudsen limit λ  d, the flux in the capil- conditions and on the material of the wall and the fluid.
lary tube is collisionless. The gas flux from the hot In the Gibbs approach [33], we operate with a thin
chamber to the cold chamber is const × n1 v̄1 and film (a surface). The force in the thermomechanical effect
the inverse flux is const × n2 v̄2 with the same const should be proportional to the length of the section of this
(we presume that the capillary tube is symmetric). surface, that is the circle, F = γ πd, and the pressure
Here, ni is density, the ideal gas equation gives difference should be (Figure 4)
√i = Pi , v̄i is the thermal velocity, proportional
ni kθ
F 4γ P2 P2 d
to kB θi , and kB is the Boltzmann constant. The P1 − P2 = = and = .
zero flux condition gives
1
4 πd
2 d P1 P2 d + 4γ

 This model includes one unknown coefficient γ ,


P1 P2 P2 θ2
√ =√ or = . (1) which can be easily evaluated from the empiric curves
θ1 θ2 P1 θ1 (Figure 4). Of course, this simplest model belongs to Type
5 of Peierls’ classification [32]: ‘Instructive model (No
In the Knudsen limit, we see how the steady state dif- quantitative justifications but gives insight)’ or even to
ference between temperatures produces the difference Type 6: ‘Analogy (Only some features in common)’. This
between pressures. This is called the thermomechanical model works well in comparison with experiment (Figure
effect, and the effect of the capillary flux induced by the 4) for not very high Knudsen numbers Kn < 0.1, gives
temperature gradient is called thermoosmosis. The ratio the proper Navier–Stokes limit, but has one fundamental
P2 /P1 is called the thermal transpiration ratio. defect: P2 /P1 → 0 when P2 d → 0 (or, which is the same,
The thermomechanical effect and thermoosmosis Kn→ ∞).
were discovered in the twentieth century for various flu- To improve the model, let us follow the van der Waals
ids, from rarefied gases to superfluid liquids and water capillarity idea. Consider a ‘thick film’ near the border
[34,35]. Now, they are considered as a promising driving with thickness ς, if d > 2ς. If d < 2ς then the thick sur-
mechanism for micro- or nanoscale engines [37]. There face fills the whole capillary. Assume that the thickness is
are many approaches to the theory of these effects based proportional to the mean free path, ς = aλ. The force of
CONTEMPORARY PHYSICS 5

coefficient α (a function of temperature, which can be


easily extracted from the Knudsen asymptotic) and the
ratio of the film thickness to the mean free path, a = ς/λ.
We express the thermal transpiration ratio through the
Knudsen number:

1
P2 if aKn < 12 ,
= 1+αKn(1−aKn)
1 (2)
P1 1+α/(4a) if aKn ≥ 12 .

The thick surface model has the proper high Kn and


low Kn asymptotics and can be used for the understand-
ing the data. Of course, it is not far from the simple
dimensional analysis and belongs to Type 5 or 6 of Peierls’
classification. It does not reveal the intrinsic physical
mechanism of the thermomechanic effect and the ther-
moosmosis. We just hypothesised that these effects are
connected to the energy (enthalpy) density in the layers
near the interface between gas and capillary and to the
transport in these layers (thermoosmosis slip). The mod-
els can be improved if we account carefully for variations
Figure 4. Thermal transpiration values P2 /P1 of hydrogen. The of the films along the tube. Solving the Korteweg equa-
simplest thin film capillarity model (solid line) and the thick tions instead of postulating of the thick surface can pro-
film model (bold dotted line) versus experimental data (bold
dashed line). P1 is the pressure at the ‘hot end’, T1 = 473 ± 1K, vide the further improvements, but this next step requires
P2 is the pressure at the ‘cold end’, T2 = 295 ± 1K (room more technical effort. Some calculations of boundary
temperature). Experimental data are taken from [38]. The simplest layers and transpiration were performed recently on the
capillarity model gives: P2 /P1 = P2 d/(P2 d + 1/15) (fitted to data basis of Korteweg’s equation [39]. The problem is not
in the interval 0.01 < Kn < 0.1). Kn (the ratio of the mean in an accurate calculation of these effects. The kinetic
free path λ to the diameter d) is evaluated at the cold end.
equations or even direct molecular simulation do this job
We used the collisional diameter of H2 , DH2 ≈ 2.3 × 10−10 m
estimated from viscosity data using mean free path theory. In quite well. The question is about an appropriate contin-
the analysis of molecular seeds a larger diameter is used usually, uum model that describes the bulk motion, the thermal
DH2 ≈ 2.89 × 10−10 m [41]. In the thick film model a ≈ 0.25 for transpiration and the other similar effects in a unified
the experimental data presented. way.
In any case, already very simple models demonstrate
that the transport phenomena of the rarefied gas near the
the thermomechanical effect in this model is proportional
interfaces may be described using the idea of capillarity.
to the section of the surface orthogonal to the axis of the
We might imagine that the light mill is a close relative of
capillary (to the surface area), and to the particle density
the soap boat [40] because both are moving by capillarity
(empty space does not produce force): F = κAn, where A
effects.
is the area of the section. The coefficient κ has the physical
sense of ‘energy per particle’ and describes the energetic
difference between the boundary layer of thickness ς and 1.5. The structure of the paper
the rest √of the volume of the capillary tube. Recall that
nλ = ( 2πD2 )−1 , where D is the collision diameter The goal of this paper is to highlight the origin of the
of molecules, and P = nkB θ. Evaluate λ at the cold capillarity of ideal gas. For this purpose, we give a com-
end. After simple algebra we get the formula for the prehensive introduction into modern theory of dynamic
transpiration ratio in the thick film approximation: model reduction from kinetics to fluid dynamics and
analyse approximate and recently found exact solutions
 of this problem.
P2 √ P2 d if aKn < 12 ,
= P2 d+4κa( 2πD2 )−1 (1−aKn) In the next Section, we start the unhurried introduc-
1
P1 1+κ(kB θ2 )−1
if aKn ≥ 12 . tion of the main equations: from the conservation of mass
and Cauchy equations to the Navier–Stokes–Fourier
Here, κ is still an unknown coefficient. The thick film equations. In Section 2.6, Korteweg’s stress tensor is
fills the whole capillary at Kn = 1/(2a). It is convenient described. After that, we discuss a simple ‘mean free path’
to gather all the coefficients into two unknown numbers: theory for Korteweg’s stress. To that end, the idea of
6 A. N. GORBAN AND I. V. KARLIN

Ehrenfests’ coarse-graining is utilised: we take the col- vector n = (ni ) the stress tensor σ be given. Then the
lisionless Knudsen gas with periodic equilibration to the force applied to the body through this surface fragment is
local equilibrium and analyse the post-Navier–Stokes 
equation derived by this approach. F = σ · nS, that is, Fi = S σij nj .
After that, a short but comprehensive explanation of j
the Chapman–Enskog procedure is presented. We intro-
duce the invariance equation in Section 4.2 and construct The stress tensor is symmetric (to provide conservation
the Chapman–Enskog as a formal solution to this equa- of angular momentum), σij = σji .
tion (Section 4.3). The geometric language allows us to The momentum Equation (4) is the second Newton
avoid the usual bulky calculations while describing the law written for a small brick of the material: Imagine a
main constructions. small cubic brick (a ‘parcel’) with soft but impenetrable
Analysis of exactly solvable problem of reduction boundaries and edges parallel to the coordinate axes,
allows us to prove that the capillarity effects in ideal which moves with the material. The outer normals of the
gas is not a by-product of special approximations and opposite faces differ by sign only, therefore the sum of
 ∂σ
series expansions but appears as a direct consequence of the contact forces from all faces is j ∂xijj L3 + o(L3 ),
kinetic equations. In Section 5 the energy formula for where L is the edge of the brick (use the Taylor formula
hydrodynamics of non-equilibrium ideal gas is proved. for σij ). The possibility to represent the motion of contin-
It includes the pseudodifferential capillarity term, which uum as the flight of many infinitesimal parcels with soft
coincides with the van der Waals energy of capillarity deformable but impenetrable boundaries is in the essence
in first approximation. Finally, hypotheses are discussed of the mechanics of materials.
about the plausible forms of non-equilibrium fluid
mechanics.
2.4. Energy equation
2. Fluid dynamics equations If the material body  experiences infinitisemal displace-
2.1. Hydrodynamic fields ment x  → x + δr(x), which vanishes with derivatives
on the boundary, then the work of the contact forces is
Consider the classical fluid, which is defined by the (use the Stokes formula for integration by parts):
hydrodynamic fields: density ρ (scalar), velocity u (vector)
   
and specific internal energy e, that is internal energy per ∂σij 3 ∂δri 3
δW = δri d x = − σij d x. (5)
unit of mass (scalar).  i,j ∂xj  i,j ∂xj

2.2. Conservation of mass Take into account symmetry of σij and find that δW
depends on the symmetric part of ∂δri /∂xj :
1
∂ρ  ∂(ρuj )   
∂δri ∂δrj

+ = 0. (3) δW = −
1
σij + d3 x.
∂t ∂xj 2 ∂xj ∂xi
j  i,j

2.3. Cauchy momentum equations The power is the work per time. For a moving of con-
tinuum media, the power produced by the contact forces
∂(ρui )  ∂(ρui uj )  ∂σij is
+ = + fi . (4)  
∂t ∂xj ∂xj ∂ui 3
j j
P=− σij d x
 i,j ∂xj
where f = (fi ) is the body force density, σii (i = 1, 2, 3)    
1 ∂ui ∂uj
are normal stresses, and σij (i = j) are shear stresses. The =− σij + d3 x. (6)
pressure p is 2  ∂xj ∂xi
i,j

1 1 According to the Cauchy equation and the formula for


p = − trσ = − σii .
3 3 power we can write the conservation of energy
i

The Cauchy stress tensor σ = (σij ) (i, j = 1, 2, 3) ∂(ρ)  ∂(ρuj ) 


∂t + j ∂xj = σij ∂u
∂x
ij
i
describes the so-called contact force: Let for a small ele-  j  ∂qi
(7)
ment of a body surface with area S and outer normal + i fi u i − i ∂xi ,
CONTEMPORARY PHYSICS 7

where  is the energy density per unit mass: 1 2 3R


∂(ρ[ 2M θ + 12 u2 ])  ∂(ρ[ 2M3R
θ + 12 u2 ]uj )
 2= e + 2 u , +
e is the specific internal energy, u =
2
i ui , and q = ∂t ∂xj
j
(qi ) is the vector of heat flux, which appears because of
2
thermal conductivity. For the perfect (monoatomic) gas, R  ∂ui  ∂ui
e = 2M 3R
θ, where R is the ideal gas constant, M is the + ρθ =λ
M ∂xi ∂xi
i i
molar mass and θ is the absolute temperature.  
μ  ∂ui ∂uj 2   ∂  ∂θ 
+ + + fi u i + κ .
2 ∂xj ∂xi ∂xi ∂xi
ij i i
(9)
2.5. Stress tensor and heat flux for ideal fluid: the
Navier–Stokes–Fourier equations The power produced by the viscous stress is
Cauchy equations are simple and nice, and can be applied
2  
to many materials (and even beyond the mechanics of  ∂ui μ  ∂ui ∂uj 2
P = −λ − + .
materials). The question is: how can we express the stress ∂xi 2 ∂xj ∂xi
i ij
tensor and the heat flux through the hydrodynamic fields:
density ρ, velocity u and specific internal energy e? If μ > 0 and 3λ + 2μ ≥ 0 then P ≤ 0 (i.e. viscosity is
The simplest answer is given by the Navier–Stokes friction, indeed).
stress tensor and Fourier heat flux: According to the model classification proposed by R.
Peierls [32], the Navier–Stokes viscous stress tensor is
σij = σijE + σijV ; a typical example of the linear response models together
σijE = −pδij ; with Ohm’s and Hooke’s laws, Fourier’s law and many
 ∂uk   others. ‘This refers to a situation in which one is inter-
∂ui ∂uj
σijV = λδij +μ + ; (8) ested, by definition, in the response of a system to some
∂xk ∂xj ∂xi
k parameter in the limit in which this parameter may be
∂θ treated as infinitesimal’.
qi = −κ .
∂xi

2.6. Fluids with capillarity: van der Waals energy


Here (and in what follows), δij is Kronecker’s delta, σijE ,
and Korteweg stress tensor
σijV are elastic and viscous contributions to the Cauchy
stress tensor, the coefficients λ, µ and κ depend on den- In 1877, Gibbs published the theory of capillarity based
sity and temperature. The thermodynamic pressure p in on the idea of surfaces of discontinuity [33]. He intro-
the elastic stress is defined through the equation of state. duced and studied thermodynamics of two-dimensional
For example, for ideal gas p = nRθ = ρθR/M, where n is objects – surfaces. On the contrary, van der Waals pro-
the molar density (n = ρ/M). The standard assumption posed the theory of capillarity using the hypothesis that
is that the viscous tensor has zero trace and, therefore, density of the body varies continuously at and near the
λ = − 23 μ. Strictly speaking, this is just an assumption transition layer and the energy depends on the gradient
and the mechanical pressure should not necessarily co- of density [1] (just square of gradient is added). This is
incide with the thermodynamic pressure and may have a an energetic ‘penalty’ for high gradients, which does not
viscous component. allow discontinuities to appear and defines the thickness
With the constitutive Equation (8) we can immedi- and the energy of the layer between phases. Modern
ately write the Navier–Stokes–Fourier equations for the ‘phase field’ approaches are also based on this idea [42,
density, velocity and temperature of ideal gas (p = M R
θ, 43]. After van der Waals, the standard approach to the
3R
e = 2M θ): continuous theory of capillarity is: represent the Gibbs
free energy of a body  as a sum, G = G0 + GK , where
∂ρ  ∂(ρuj ) 
+ = 0;
∂t ∂xj G0 = g0 (ρ, θ) d3 x;
j 
∂(ρui )  ∂(ρui uj ) R ∂(ρθ)    ∂ρ 2
+ + GK = K(ρ, θ) d3 x. (10)
∂t ∂xj M ∂xi  ∂xi
j i

∂  ∂uk  ∂  ∂ui ∂uj



= λ + μ + + fi ; Here, K(ρ, θ) is the capillarity coefficient. Korteweg
∂xi ∂xk ∂xj ∂xj ∂xi found the corresponding GK addition to the stress tensor:
k j
8 A. N. GORBAN AND I. V. KARLIN

The hydrodynamic variables are (for particles of unit



 ∂ ∂ρ mass):
σijK = ρ ∂xk K(ρ, θ) ∂x δij
k
k
(11) 
∂ρ ∂ρ
− K(ρ, θ) ∂x i ∂xj
. ρ= f (x, v, t) d3 v;

Detailed analysis of this stress tensor and its various ρu = vf (x, v, t) d3 v;
modifications from the rational thermodynamics point 
of view is performed by Dunn and Serrin [9]. Thermo- 1
ρ = v 2 f (x, v, t) d3 v. (16)
dynamics requires to complement the Korteweg stress 2
tensor by the additional contributions to the internal
energy and the heat flux: Integration of the advection Equation (15) in v gives
the conservation of mass Equation (3). If we multiply the
   ∂ρ 2 advection equation by vi then the integration gives the
eK = K − θ ∂K
1
2ρ ∂θ ∂xi ; momentum equation and after multiplication with 12 v 2
i
∂ρ  ∂uj (12) the integration produces the energy equation:
qiK = K(ρ, θ)ρ ∂x i ∂xj .
∂ρ  ∂(ρuj )
j
+ = 0;
∂t ∂xj
To modify the Navier–Stokes equation we should just add j
the Korteweg stress to the elastic and viscous stresses, and ∂ρui  ∂ij
+ = 0;
also add the capillarity contributions to the energy and to ∂t ∂xj
j
the heat flux
∂ρ  ∂Qj
+ = 0. (17)
σij = σijE + σijV + σijK ; ∂t ∂xj
j
1
 = e + u2 + e K ; Here, tensor ij of momentum flux and vector Qi of
2
∂θ energy flux are defined as moments of the distribution f :
qi = −κ + qiK . (13)
∂xi  
1
ij = vi vj f (x, v, t) d3 v; Qi = vi v 2 f (x, v, t) d3 v.
With these additions, the Navier–Stokes–Fourier equa- 2
tions for ideal gas (9) turn into Korteweg equations for (18)
gas with capillarity. Now we have to understand how it It is convenient to separate the flux with the mean flow
may happen that the ideal gas gains capillarity. velocity u from other forms of transport. For this pur-
pose, the central moments of f are convenient. Simple
algebra gives:
3. Kinetics 
1
3.1. Stress tensor for collisionless gas ρe = (v − u)2 f (x, v, t) d3 v,
2
Collisionless gas consists of particles which are moving 1
ρ = ρe + ρu2 ; (19)
without collisions between them. It is described by the  2
one-particle distribution function f (x, v, t), where x is σij = − (vi − ui )(vj − uj )f (x, v, t) d3 v,
particle’s position, v is particle’s velocity and t is time.
The integral of the density over a domain  in the 6- ij = ui uj − σij ; (20)

dimensional space of positions and velocities is the num- 1
qi = (vi − ui )(v − u)2 f (x, v, t) d3 v,
ber of particles in :  f (x, v, t)d 3 x d3 v = N(t). 2

Time evolution of f (x, v, t) is given by a simple explicit Qi = qi − σij uj + ui ρ (21)
expression: j

f (x, v, t) = f (x − vt, v, 0). (14) (the minus sign in the definition of σ appears just to
provide similarity to Cauchy stress). Recall that here for
It satisfies the advection equation simplicity, we take unit mass of the particles and measure
ρ in number of particles per unit volume. To return to
∂f (x, v, t)  ∂f (x, v, t) the standard absolute temperature θ we have to notice
+ vi = 0. (15)
∂t ∂xi that e = 32 kB θ, where kB is Boltzmann’s constant.
i
CONTEMPORARY PHYSICS 9

With usage of central moments, the transport chain for the collisionless gas and present the results of
equations look very similar to the qCauchy and energy the coarse-graining.
equations: First of all, for each value of hydrodynamic variables,
the equilibrium distribution should be defined:
∂ρ  ∂(ρui ) f eq (v|ρ, u, e). It is the Maxwellian distribution. In the
+ = 0;
∂t ∂xi units used in this section (we count neither moles nor
j
grams but particles) the Maxwellian is
∂(ρui )  ∂(ρui uj )  ∂σij
+ = ; (22)
∂t ∂xj ∂xj  3/2 
j j 3 3(v − u)2
f eq (v|ρ, u, e) = ρ exp − .
∂(ρ)  ∂(ρuj )  ∂qj  4πe 4e
+ + = σij uj . (24)
∂t ∂xj ∂xj
j j j
For the given hydrodynamic fields, ρ(x), u(x), e(x), we
use the local Maxwellian, f eq (v|ρ(x), u(x), e(x)), that is
The standard definition of pressure gives:
a distribution in six-dimensional space of positions and
 velocities, which is the Maxwellian for each x.
1 1 2
p=− σii = (v − u)2 f (x, v, t) d3 v = ρe. Define Ehrenfests’ chain with time step τ . Let the
3 3 3
i initial values of the hydrodynamic fields be given, ρ(x, 0),
(23) u(x, 0), e(x, 0). The initial distribution is
Particles of collisionless gas do not interact and there
are no forces. Motion of such a gas cannot be represented
f (x, v, 0) = f eq (v|ρ(x, 0), u(x, 0), e(x, 0)).
as motion of continuum. Nevertheless, the Cauchy stress
tensor can be defined through the momentum flux and
Start the advection from this distribution. After time τ it
the Cauchy transport equation holds (22).
approaches the distribution f (x − vτ , v, 0). The hydro-
The elastic part (pressure) of the stress tensor for colli-
dynamic fields ρ(x, τ ), u(x, τ ), e(x, τ ) are defined
 using
sionless gas has the very common form (23). It is not sur-
the moments of this distribution: ρ(x, τ ) = f (x −
prising that the rest of the stress tensor does not satisfy the
vτ , v, 0) d3 v, etc.
Navier–Stokes constitutive relations (8). Nevertheless,
If the hydrodynamic fields ρ(x, kτ ), u(x, kτ ), e(x, kτ )
Ehrenfests’ idea of coarse-graining allows us to deduce,
are given then
step by step, the Euler equations, the Navier–Stokes–
Fourier equations and the Korteweg equations for fluid
f (x, v, kτ ) = f eq (v|ρ(x, kτ ), u(x, 0), e(x, kτ )),
motion from the simple model of collisionless gas.

and the hydrodynamic fields ρ(x, (k + 1)τ ), u(x, (k +


1)τ ), e(x, (k + 1)τ ) are defined using the moments of
3.2. Ehrenfests’ coarse-graining: from collisionless
the distribution f (x − vτ , v, kτ ). Ehrenfests’ chain is
gas to Korteweg fluid dynamics through
the sequence ρ(x, kτ ), u(x, kτ ), e(x, kτ ) (k = 0, 1, 2, . . .)
coarse-graining
(Figure 5). The hydrodynamic fields do not change in
In 1911, Paul and Tanya Ehrenfest in their paper for the equilibration steps and their trajectories between the
the scientific Encyclopedia [44] discussed emergence of equilibration jumps are just ‘shadows’ of the advection
irreversibility and introduced a special operation, that (14). Equilibration returns the distribution to the local
of coarse-graining. This operation transforms a proba- Maxwellian form and free flight destroys this form.
bility density in phase space into a ‘coarse-grained’ den- As a result, a discrete-time dynamical system is defined
sity. That is a piece-wise constant function, a result of in the space of hydrodynamic fields. We can look for
density averaging in cells. The size of cells is assumed a continuous-time system which transforms ρ(x, kτ ),
to be small, but finite, and does not tend to zero. The u(x, kτ ), e(x, kτ ) into ρ(x, (k + 1)τ ), u(x, (k + 1)τ ),
coarse-graining (‘shaking’) models thermalisation of the e(x, (k + 1)τ ) during time τ .
system by uncontrollable and weak interaction with Let us describe formally this procedure. Denote five
surroundings. hydrodynamic fields as M(t) = {ρ(x, v, t), u(x, v, t),

We generalise Ehrenfests’ idea of coarse-graining by e(x, v, t). We aim to find a closed system ∂t M = Jτ (M),
combining the genuine motion with the periodic where Jτ is a (nonlinear) operator, which depends on
partial equilibration. The result is Ehrenfests’ chain [45] the coarse-graining time τ as a parameter, and Jτ (M) is
(Figure 5). The general theory with the proof of the en- also a set of five fields (scalar, vector, and scalar again).
tropy production formula and various examples can be Let operator m map the distribution f onto hydrody-
found elsewhere [46,47]. Here we just define Ehrenfests’ namic fields: M = m(F) (16). Select an initial state:
10 A. N. GORBAN AND I. V. KARLIN

Figure 5. Ehrenfests’ chain. Two spaces are schematically


represented: the space of distribution functions (the upper
rectangle) and the space of hydrodynamic fields (the bottom
Figure 6. Matching condition for hydrodynamic equations.
parallelogram). The vertical lines with bidirectional arrows
Two spaces are schematically represented: the space of
illustrate coupling between the hydrodynamic fields and the
distribution functions (upper rectangle) with the manifold of
corresponding local Maxwellians. The coarse-graining is the
local Maxwellians and the space of hydrodynamic fields (the
projection of a distribution function onto the corresponding local
bottom parallelogram). The vertical lines with bidirectional arrows
Maxwellian with preservation of the hydrodynamic fields.
illustrate coupling between the hydrodynamic fields and the
corresponding local Maxwellians. The matching condition is:
M(τ ) = f eq (x − vτ , v|M0 ).
M(0) = M0 and f0 is the corresponding local Maxwellian,
f0 = f eq (x, v|M0 ). Represent the result of advection dur-
ing time τ as a power series in τ : of kinetic coefficients: μ ≈ 13 ρl v̄ = 13 ρτf v̄ 2 (where v̄ is

the thermal velocity and τf is the mean free time, see,
 τk
f (τ ) = f0 + ( − 1)k (v, ∇)k f0 . for example [48]). The thermal velocity for Maxwellian

k=1
k! distribution is v̄ ∼ 4e/3 (that is the most probable
velocity, whereas the average magnitude of the velocity
Find the power series for the result of hydrodynamic of the particles is √2π v̄). Therefore, the coarse-graining
motion according to a hypothetical equation Ṁ = Jτ (M): time τ can be considered as the analogue of the mean
 free time τf with a constant coefficient of order 1.
t d Jτ (M) 
∞ k k−1
 It is worth mentioning that viscosity and thermal con-
M(t) = M(0) + tJτ +  , ductivity grow with the coarse-graining time, and dissi-
k! dt k−1 
k=2 M=M0 pative processes become more intensive with the growth
of the mean free path. Collisions in gas delay diffusive
Jτ (M)
k−1
where the sequence of derivatives d dt k−1 is calculated transport and decrease kinetic coefficients.
by iterations of the chain rule from the equation Ṁ = Thus, the simple model of non-equilibrium gas based
Jτ (M). We use this series for t = τ . on Ehrenfests’ coarse graining produces the Navier–
Recall that Jτ (M) depends also on τ as on a parameter. Stokes–Fourier equations in the first order of the coarse-
Represent this dependence as a power series as well and graining time τ . This model has only one free parameter,
substitute it into the series for M(t) (t = τ ). The match- τ , and, therefore, the viscosity and the thermal conduc-
ing condition should hold: M(τ ) = m(f (τ )) (Figure 6). tivity are proportional (and Prandtl number Pr = 1). This
Coefficients at the same powers of τ in M(τ ) and m(f (τ )) is not a miracle, and just demonstrates that the Navier–
should coincide. As a result, we recover the macroscopic Stokes–Fourier equations provide an unavoidable first
vector field Jτ order-by-order. dissipative correction to the Euler equations. The prob-
The zeroth-order terms give the Euler equations and lem is in the next approximation. In the second order of
the first-order terms give the Navier–Stokes–Fourier τ , the coarse graining produces the Korteweg terms from
equations with viscosity and heat conductivity propor- the collisionless gas.
tional to τρe with constant coefficients (ρe is the density Consider the linearised fluid dynamics near a state
of internal energy): λ = − 23 μ, μ = 13 τρe [47]. These with u = 0, ρ = ρ0 , e = e0 . Introduce the dimensionless
formulas recall the Maxwell mean free path estimations deviation from this state:
CONTEMPORARY PHYSICS 11

ρ − ρ0 (a) (b)
M0 = ,
ρ0
ui
Mi = (i = 1, 2, 3),
v̄  
3 ρ − ρ0 e − e0
M4 = + ,
2 ρ0 e0


4
where v̄ = 3 e0 . Use the new space scale in which
v̄ = 1: xnew = x/v̄. In these variables, the linearised
second approximation in τ for the coarse-grained free
flight advection is:

∂M0  ∂Mi
3
Figure 7. Dispersion curves for the second approximation (25). (a)
= ; Dependence of attenuation rates on |k|: solid λ1,2 , dashed λ3 ,
∂t ∂xi
i=1 dotted λ4,5 ; (b) Curves λ(k) on complex plane for λ4,5 .
∂Mi 1 ∂M4
=−
∂t 3 ∂xi λ1,2 1 λ3

= − ; 2 = O(1/k2 ) → 0;
τ  ∂ 2  ∂Mk
3 3
∂Mi ∂Mj k 2 4 k 
+ + − δij
4 ∂xj ∂xj ∂xi 3 ∂xk λ4,5 17 59 · 89
j=1 k=1 = − ± i|k| . (26)
  k 2 12 9 · 108
∂ 1 89
+ τ2 M0 + M4 (i = 1, 2, 3);
∂xi 8 108 The second approximation in τ for the coarse-grained
5  ∂Mk 5τ 
3 3 collisionless gas provides a sort of ‘mean free path model’
∂M4
=− + M4 of capillarity effects and Korteweg stress in ideal gas.
∂t 2 ∂xk 2
k=1
3
k=1

It seems to be surprising that such a simple approach
59  ∂M k generates the Korteweg equations (with some restrictions
+ τ2  . (25) on the relationships between parameters, like Pr = 1).
9 ∂xk
k=1 Nevertheless, two question remain: (i) are there capil-
larity effects in more realistic models of non-equilibrium
3 ∂ 2 gas, and (ii) what will happen in the coarse-grained col-
Here,  = k=1 ∂x 2 is the Laplace operator. Terms
k lisionless gas beyond the second order in τ ? Of course, if
without τ correspond to the linearised Euler equation, the we summarise the whole power series then the solutions
first-order terms give the Navier–Stokes–Fourier dissipa- of the resulting continuum mechanics equations will go
tion (viscosity and thermal conductivity). Terms of the exactly through the points of Erenfests’ chain but what
second order in τ correspond to Korteweg’s stress tensor will happen on the way?
(in equations for Mi , i = 1, 2, 3) and to the contribution The more realistic model of non-equilibrium gases is
of capillarity into heat flux (in equation for M4 ). well-known, that is the Boltzmann equation.
The parameter τ can be eliminated from Equation (25)
by rescaling. Recall that in these equations space and time
are measured by the same time units (to provide v̄ = 1). 4. From Boltzmann kinetics to fluid dynamics:
Let us select the new time and space unit τ . In this scale, model reduction
τ = 1 in (25). From some point of view, this means
4.1. The model reduction problem
that the equations are the same for all values τ > 0. For
example, it is sufficient to analyse stability just for one Let us consider kinetic equations which describe the evo-
value τ = 1. lution of a one-particle gas distribution function f (x, v, t)
Let us look for the solutions of Equation (25) in the
form Mj = Aj exp (λt + ikx). Here, k is the wave vector ∂f  ∂f 1
+ vi = Q(f ). (27)
and real parts of λ describe dissipation. It is necessary ∂t ∂xi Kn
i
for stability that Reλ ≤ 0 for all real vectors k. The
characteristic equation for λ(k) has five roots (Figure 7). The only difference from the free flight advection (15) is
For all of them Reλ < 0 (k = 0). In the short wave the collision operator Q(f ) in the right-hand part of (27).
asymptotic of λ (k2 → ∞) the roots are: For the Boltzmann equation, Q is a quadratic operator
12 A. N. GORBAN AND I. V. KARLIN

and, therefore, the notation Q(f , f ) is often used. Kn is


the Knudsen number, which is a dimensionless parameter
defined as Kn = Ll , where l is the mean free path and L is
‘representative physical length’ scale. It aims to measure
how important the microscopic effects are (associated
with l) at the macroscopic scale (L).
The collision term Q(f ) in (27) describes the change
in distribution f due to collisions. The term Q(f ) does
not contribute directly into the time derivatives  of the
hydrodynamic  variables, ρ = f d 3 v, u = vf d3 v
1
and e = 2 (v − u) f d v because, due to the mass,
2 3

momentum and energy conservation in collisions:



{1; v; v 2 }Q(f ) d3 v = 0. Figure 8. Fast–slow decomposition. Bold dashed lines outline
the vicinity of the slow manifold where the solutions stay
For the space-uniform distributions the collision opera- after initial layer. The projection of the distributions onto the
hydrodynamic fields and the parametrisation of this manifold
tor in (27) provides relaxation to Maxwellian distribution
by the hydrodynamic fields are represented. (A derivative work
(24) [49] (about the general problem of existence and based on Figure 1 from [59].)
stability for the Boltzmann equations we refer readers to
[50]).
Therefore, the following qualitative ‘nonrigorous pic-
ture of the Boltzmann dynamics’ [51] is expected for the
solutions: (i) the collision term goes quickly almost to its
equilibrium (the system almost approaches a local equi-
librium) and during this fast initial motion the changes
of hydrodynamic variables are small, (ii) after that the
distribution function is defined with high accuracy by
the hydrodynamic variables (if they have bounded space
derivatives). The relaxation of the collision term almost
to its equilibrium is supported by monotonic entropy
growth (Boltzmann’s H-theorem). This qualitative pic-
ture is illustrated in Figure 8.
Figure 9. McKean diagram. All the model reduction approaches
Perhaps, McKean gave the first clear explanation of the aim to create a lifting operation, from the hydrodynamic variables
problem as a construction of a ‘nice submanifold’ where to the relevant distributions on the invariant manifold. IM stands
‘the hydrodynamical equations define the same flow as for Invariant Manifold. The part of the diagram in the dashed
the (more complicated) Boltzmann equation does’ [52]. polygon is commutative, that is any superposition of operations
He presented the problem by a diagram and we reproduce following the arrows does not depend on the path but only on the
start and end points. (A derivative work based on Figure 2 from
his idea in slightly revised form in Figure 9. How to find [59].)
this ‘nice submanifold’? Perhaps, the first task is to write
an equation for them (for more detail we refer to the
The essence of the invariance equation and the
chapter ‘Invariance equation’ of the book [53]).
Chapman–Enskog method become more transparent in
the abstract form when the bulky details of the genuine
Boltzmann equation do not obscure the simple geometric
4.2. Invariance equation
sense. Let us consider an equation in a domain U of a
The invariance equation expresses the fact that the normed space E with analytical right-hand side
vector field is tangent to the manifold. In 1892, A.M.
Lyapunov introduced and studied this equation in his ∂t f = J(f ). (28)
doctoral thesis ‘The general problem of the stability of
motion’ (the Lyapunov ‘auxiliary theorem’ [55], which We call f a microscopic variable. A space of macroscopic
is recently used in various applications, from model variables (moment fields) is defined with a linear map
reduction in chemical kinetics to optimal control prob- m : f  → M (M are macroscopic variables). Assume
lems [56,57]). that the image of m is the whole space of M (i.e. m
CONTEMPORARY PHYSICS 13

is surjective). We are looking for an invariant mani- The following assumptions connect the macroscopic
fold (McKean’s nice manifold’) parameterised with the variables to the singular perturbation:
macroscopic fields M. For such manifolds we use the • m(Q(f )) = 0;
notation f M . The self-consistency condition m( f M ) = M • for each M ∈ m(U) the system of equations
is necessary: the manifold is parametrised by its own value Q(f ) = 0, m(f ) = M has a unique solution fM
eq
of macroscopic variables. (in Boltzmann kinetics it is the local Maxwellian);
For the reduction of Boltzmann’s kinetics hydrody- eq
• fM is asymptotically stable and globally attracting
namics, the microscopic variable is the one-particle dis- eq
for the fast system ∂t f = 1 Q(f ) in (fM +ker m)∩U.
tribution function f and the macroscopic variables are
hydrodynamic fields. In extended irreversible thermody- Let the QM be the differential of the fast vector field
eq
namics (EIT) the larger sets of macroscopic variables are Q(f ) at equilibrium fM : QM = (DM Q(f ))f =f eq . For the
M
considered [58]. Chapman–Enskog method it is important that QM is
The invariance equation for f M is invertible in ker m.
The invariance equation for the singularly perturbed
J(fM ) = (DM fM )m(J(fM )). (29) system (32) with the moment parametrisation m is:

Here, the differential DM of fM is calculated at the point


M = m(fM ). 1
Q(fM ) = A(fM ) − (DM fM )(m(A(fM ))). (33)
Equation (29) means that the time derivative of f on 
the manifold fM can be calculated by a simple chain
rule: calculate the derivative of M using the map m, The self-consistency condition m(fM ) = M gives
Ṁ = m(J(fM )), then write that on the invariant manifold m((DM fM )m(J)) = m(J) for all J, hence,
the time dependence f (t) can be expressed through the
time dependence of M(t): f (t) = fM(t) and the time m[A(fM ) − (DM fM )m(A(fM ))] = 0. (34)
derivatives also coincide. If we find the approximate so-
lution to Equation (29) then the approximate reduced If we find an approximate solution fM of (33) then the cor-
model (hydrodynamics) is responding macroscopic (hydrodynamic) Equation (30)
is
∂t M = m(J(fM )). (30)
∂t M + m(A(fM )) = 0. (35)
The invariance equation can be represented in the form Let us represent all the operators in (33) by the Taylor
series (for the Boltzmann equation A is the linear free
∂tmicro fM = ∂tmacro fM , (31) flight operator, A = v · ∇, and Q is the quadratic collision
operator). We look for the invariant manifold in the form
where the microscopic time derivative, ∂tmicro fM is just a of the power series:
value of the vector field J(fM ) and the macroscopic time
derivative is calculated by the chain rule, ∞

eq (i)
f M = fM +  i fM (36)
∂tmacro fM = (DM fM )∂t M i=1

with the self-consistency condition m(fM ) = M, which


under the assumption that dynamics of M follows the implies m(fM ) = M, m(fM(i) ) = 0 for i ≥ 1. After
eq
projected Equation (30). For more details we refer to the matching the coefficients of the series in (33), we obtain
review paper [59]. for every fM(i) a linear equation

(i) (1) (i−1)


QM fM = P (i) (fM , fM , . . . , fM
eq
4.3. Chapman–Enskog expansion in brief ), (37)

The Chapman–Enskog approach assumes the special sin- where the polynomial operator P (i) at each order i can
gularly perturbed structure of the equations and looks be obtained by straightforward calculations from (33).
for the invariant manifold in a form of the series in the Due to the self-consistency, m(P (i) ) = 0 for all i and the
powers of a small parameter . A one-parametric system Equation (37) is solvable. The first term of the Chapman–
of equations is considered: Enskog expansion has a simple form
1
fM(1) = Q−1
eq eq
∂t f + A(f ) = Q(f ). (32) M (1 − (DM fM )m)(A(fM )). (38)

14 A. N. GORBAN AND I. V. KARLIN

Many books and papers are devoted to the detailed analy- (more precisely, it is the locally Gaussian distribution
sis of this formula for the Boltzmann equation and other with space-dependent parameters defined by hydrody-
kinetic equations after the classical book [23]. Most of namic fields). Substitute this function into Botzmann’s
the physical applications of kinetic theory, from transport equation, calculate the corresponding time derivatives of
processes in gases to modern numerical methods (lattice macroscopic variables and the closed 10-moment Max-
Boltzmann models [54]) give examples of the practi- Ent Grad system is ready. Linearise and study the solu-
cal applications and deciphering of this formula. For tions that depend on one space coordinate x with the
the Boltzmann kinetics, the zero-order approximation, velocity oriented along the x axis. Use dimensionless
(0) eq
fM ≈ fM produces in projection on the hydrodynamic variables. (Here, σ is the dimensionless xx-component
fields (35) the compressible Euler equation. The first- of the stress tensor.)
(1) eq (1)
order approximate invariant manifold, fM ≈ fM +fM , Let us illustrate the invariance equation and the
gives the compressible Navier–Stokes equation and pro- Chapman–Enskog series on the simplest example (39).
vides the explicit dependence of the transport coefficients ⎛ ⎞
on the collision model. p(x)  
⎝ ⎠ 100
The calculation of higher order terms needs nothing f = u(x) , m = ,
010
but differentiation and calculation of the inverse operator σ (x)
⎧⎛ ⎞⎫
Q−1M , although it may be rather bulky. The second or-   ⎨ 0 ⎬
p(x)
der in  hydrodynamic Equation (30) are called Burnett M= , ker m = ⎝ 0 ⎠ ,
equations (with  2 terms) and super-Burnett equations u(x) ⎩ ⎭
y
for higher orders. ⎛ 5 ⎞ ⎛ ⎞
3 ∂x u 0
Alas, the Burnett equations produce non-physical ef- A(f ) = ⎝ ∂x p + ∂x σ ⎠ , Q(f ) = ⎝ 0 ⎠ ,
fects, instability of short waves (Bobylev’s instability) and 4
−σ
3 ∂x u
negative viscosity at the space scale near mean free path,
i.e. close to the scales, where they are needed. What will Q−1
M = QM = −1 on ker m,
⎛ ⎞ ⎛ ⎞
happen if we sum up the whole Chapman–Enskog series p(x) 10
fM = ⎝ u(x) ⎠ , DM fM = ⎝ 0 1 ⎠ ,
eq eq
(at least, hypothetically)? For Ehrenfests’ chain the sum
is expected to coincide with the chain in discrete time 0 00
⎛ ⎞
moments nτ . For the Boltzmann equation, it remains 0
(1)
unclear. Let us choose a simplified system, for which the fM = ⎝ 0 ⎠ .
model reduction can be performed explicitly for all time − 43 ∂x u
scales.
We hasten to remark that (39) is a simple linear sys-
4.4. Exact analytic solution of reduction problem for tem and can be integrated immediately in explicit form.
a simple kinetic equation However, that solution contains both the fast and slow
components and it does not readily reveal the slow
The simplest model and the starting point in our analysis hydrodynamic manifold of the system. Instead, we are
is interested in extracting this slow manifold by a direct
5 method. The Chapman-Enskog expansion is thus the for
∂t p = − ∂x u,
3 this which we shall address first.
∂t u = −∂x p − ∂x σ , (39) The projected equations in the zeroth (Euler) and the
4 1 first (Navier–Stokes) order of  are
∂t σ = − ∂x u − σ ,
3 
To obtain this system from Boltzmann’s equation we ∂t p = − 53 ∂x u,
(Euler)
have to select the set of macroscopic variables: ‘Hydrody- ∂t u = −∂x p;
namic fields plus stress tensor’. For these 10 variables (1 ∂t p = − 53 ∂x u,
(Navier–Stokes)
– density, plus 3 – momentum density, plus 1 – energy, ∂t u = −∂x p +  43 ∂x2 u.
and plus 5 – traceless symmetric stress tensor) find the
representative (quasiequilibrium) density function
 f by It is straightforward to calculate the two next terms (Bur-
conditional maximisation of entropy, S = − f ln f d3 v, nett and super-Burnett ones) but let us introduce conve-
for given values of these 10 macroscopic moments. The nient notations to represent the whole Chapman-Enskog
conditional maximisation of entropy subject to given series for (39). Only the third component of the invari-
first and second moments is a standard exercise in the ance Equation (33) for (39) is non-trivial because of the
MaxEnt approach. The result is the Gaussian distribution self-consistency condition (34), and we can write
CONTEMPORARY PHYSICS 15

− 1 σ(p,u) = 43 ∂x u − 53 (Dp σ(p,u) )(∂x u) For the super Burnet equations Reλ > 0 if k2 > 3. This is
(40)
− (Du σ(p,u) )(∂x p + ∂x σ(p,u) ). an example of Bobylev’s instability. The dispersion curves
are presented in Figure 10.
Here, M = (p, u) and the differentials are calculated by Let us analyse the structure of the Chapman–Enskog
the elementary rule: if a function  depends on values series given by the recurrence formula (41). The terms
of p(x) and its derivatives,  = (p, ∂x p, ∂x2 p, . . . ) then in the series alternate: For odd i = 1, 3, . . . they are
Dp  is a differential operator, proportional to ∂xi u and for even i = 2, 4, . . . they are
proportional to ∂xi p. It follows from the parity properties:
∂ ∂ ∂ 2 u and ∂x change sign after spatial reflection (vectors),
Dp  = + ∂x + ∂ + ...
∂p ∂(∂x p) ∂(∂x2 p) x whereas p (a scalar) and σ (a second-order tensor) are
invariant with respect to inversion. This global structure
The equilibrium of the fast system (the Euler approx- of the Chapman–Enskog series gives the following repre-
(0)
imation) is known, σ(p,u) = 0. We have already found sentation of the stress σ on the hydrodynamic invariant
(1)
σ(p,u) = − 43 ∂x u (the Navier–Stokes approximation). In manifold
each order of the Chapman–Enskog expansion i ≥ 1 we
get from (40): σ (x) = A( − ∂x2 )∂x u(x) + B( − ∂x2 )∂x2 p(x), (45)

(i+1) (i) (i)


σ(p,u) = 53 (Dp σ(p,u) )(∂x u) + (Du σ(p,u) )(∂x p) where A(y), B(y) are yet unknown functions and the sign
 (j) (l) (41) ‘−’ in the arguments is adopted for simplicity of formulas
+ j+l=i (Du σ(p,u) )(∂x σ(p,u) ).
in the Fourier transform. For the stress tensor (45) the
(i+1) reduced equations have the form
This chain of equations is non-linear but every σ(p,u)
is a linear function of derivatives of u and p with constant 5
coefficients because this sequence starts from − 43 ∂x u and ∂t p = − ∂x u,
3 (46)
the induction step in i is obvious. 2 2 2
∂t u = −∂x p − ∂x [A( − ∂x )∂x u + B( − ∂x )∂x p].
Simple algebra gives for the Burnett term (i + 1 = 2)
(2)
σ(p,u) = − 43 ∂x2 p and for the super Burnett term (i + 1 =
(3)
3) the σ(p,u) = − 49 ∂x3 u. The projected equations have the 4.5. Exact invariant manifold in Fourier
form representation
∂t p = − 53 ∂x u, It is convenient to work with the pseudodifferential oper-
Burnett; (42)
∂t u = − ∂x p +  43 ∂x2 u +  2 43 ∂x3 p; ators like (45) in Fourier space. Let us denote pk , uk and
∂t p = − 53 ∂x u, σk , where k is the ‘wave vector’ (space frequency).
∂t u = − ∂x p +  43 ∂x2 u +  2 43 ∂x3 p super Burnett. (43) The Fourier-transformed kinetic Equation (39) takes
+  3 49 ∂x4 u; the form ( = 1):

5
Compute the dispersion relation for these hydrodynamic ∂t pk = − ikuk ,
modes. Exclude  using a new space-time scale, 3
x  =  −1 x, and t  =  −1 t. Look for wave solutions ∂t uk = −ikpk − ikσk , (47)
u = uk ϕ(x  , t  ), and p = pk ϕ(x  , t  ), where ϕ(x  , t  ) = 4
∂t σk = − ikuk − σk .
exp (λt  + ikx  ), and k is a real-valued wave vector. The 3
following dispersion relations λ(k) are the conditions of a We know already that the result of the reduction
non-trivial solvability of the corresponding linear system should be a function σk (uk , pk , k) of the following form:
with respect to uk and pk :

2 1  σk (uk , pk , k) = ikA(k2 )uk − k2 B(k2 )pk , (48)


λ± = − k2 ± i|k| 15 − 4k2 , Navier–Stokes;
3 3
2 2 1  where A and B are unknown real-valued functions of k2 .
λ± = − k ± i|k| 15 + 16k2 , Burnett; The solution of the invariance equation amounts to
3 3
2 2 finding the two functions, A(k2 ) and B(k2 ). Let us rewrite
λ± = − k (3 − k2 ) the invariance equation for these unknown functions.
9
1  Compute the time derivative of σk (uk , pk , k) in two dif-
± i|k| 135 + 144k2 + 24k4 − 4k6 , super Burnett; ferent ways. First, use the right-hand side of the third
9
(44) equation in (47). We find the microscopic time derivative:
16 A. N. GORBAN AND I. V. KARLIN

(a) Re λ Re λ
1 |k|
10 0

1 |k|
0 -0.5

-5

-10 -1

(b) Im λ
10 1

Re λ
-5 5 10 -1 1

-5

-10 -1

Figure 10. Dispersion curves (44) for various hydrodynamic approximations obtained from the simple kinetic Equation (39): solid -
exact solution, dotted - Navier–Stokes approximation, dashed - Burnett equation, dash and dotted – super Burnett approximation. (a)
Dependence of attenuation rates on |k| (for the Navier–Stokes and Burnett curves Reλ coincide if 4k 2 < 15; they differ in Imλ); (b)
Curves λ(k) on complex plane.

  
4 4 5  
∂tmicro σk = −ik + A uk + k2 Bpk . (49) −A − − k2 B + A2 = 0, −B + A 1 − k2 B = 0.
3 3 3
(52)
Secondly, let us use the chain rule and the first two Solving the system (52) for B, and introducing a new
equations in (47). We find the macroscopic time function, X(k2 ) = k2 B(k2 ), we obtain an equivalent cubic
derivative: equation:

∂σk ∂σk  
∂tmacro σk = ∂t uk + ∂t pk 5 2 4 X
∂uk ∂pk − (X − 1) X + = 2. (53)
  3 5 k
  2 5
= ikA −ikpk − ikσk − k B − ikuk
3
  We need the real-valued functions A(k2 ) and B(k2 ) (48).
5 2  
= ik k B + k 2 A u k + k 2 A − k 2 B pk . The real-valued root X(k2 ) of (53) is unique and negative
3 for all finite values k2 . Moreover, the function X(k2 ) is a
(50) monotonic function of k2 and

The microscopic time derivative should coincide with


the macroscopic time derivative for all values of uk and lim X(k2 ) = 0, lim X(k2 ) = −0.8. (54)
|k|→0 |k|→∞
pk . This is the invariance equation:
Therefore, both B = k2 X and A = B/(1−X) are negative.
∂tmacro σk = ∂tmicro σk . (51) For the given A(k2 ) and B(k2 ) the dispersion relation
for the hydrodynamic modes are (just use the stress ten-
For the kinetic system (47), it reduces to a system of two sor σ (48) in the first two equations of system (47) and
quadratic equations for functions A(k2 ) and B(k2 ): express A and B through X):
CONTEMPORARY PHYSICS 17


X |k| 5X 2 − 16X + 20 In x-space the energy equation has the standard form
λ± = ±i , (55) (7):
2(1 − X) 2 3
  ∞  ∞   ∞
1 3 2 2
where X = X(k2 ) is the real-valued root of Equation (53). ∂t p dx + u dx = σ ∂x u dx.
2 5 −∞ −∞ −∞
Since 0 > X(k2 ) > −1 for all |k| > 0, the attenuation (58)
rate, Re(λ± ), is negative for all |k| > 0, and the exact Note that the usual factor ρ in front of u2 is absent because
acoustic spectrum of the reduced equations is stable for we work with linearised equations and dimensionless
arbitrary wave lengths (Figure 10, solid line). In the short- variables.
wave limit, from (54), (55) we obtain the saturation of Let us use in (58) the representation (45) for σ and
dissipation: notice that ∂x u = − 35 ∂t p:
2 Imλ± √  ∞  ∞
lim Reλ± = − ; lim = ± 3. (56) σ ∂x u dx = (∂x u)(A( − ∂x2 )∂x u) dx
|k|→∞ 9 |k|→∞ |k|
−∞ −∞

3 ∞
Thus, we found the invariant hydrodynamic manifold − (∂t p)[B( − ∂x2 )∂x2 p] dx.
in two steps: 5 −∞

(1) We used the invariance equation, Chapman– The operator B( − ∂x2 )∂x2 is symmetric, therefore,
Enskog procedure and the symmetry properties  ∞
to find a linear space where the hydrodynamic (∂t p)[B( − ∂x2 )∂x2 p] dx
invariant manifold is located. This space is param- −∞
 ∞ 
eterised by two functions of one variable (48); 1 2 2
= ∂t p[B( − ∂x )∂x p] dx .
(2) We used the invariance equation again and 2 −∞
defined an algebraic manifold in this space. For
the simple kinetic system (39), (47) this manifold The quadratic form,
is given by the system of two quadratic equations 
3 ∞
which depend linearly on k2 (52). Uc = p(B( − ∂x2 )∂x2 p) dx
5 −∞

3 ∞
=− (∂x p)(B( − ∂x2 )∂x p) dx (59)
5. Van der Waals capillarity energy in ideal gas 5 −∞
5.1. The energy formula and ‘capillarity’ of ideal gas may be considered as a part of the energy. Moreover,
Let us look on the stress tensor (48). Traditionally, σ in the function B(y) is negative, hence, this form is positive.
kinetics of gases is considered as a viscous stress tensor Finally, the energy formula in x-space is
but the second term, B( − ∂x2 )∂x2 p(x), is proportional to   
1 ∞ 3 2 3
second derivative of p(x) and it does not meet usual 2 2
∂t p + u − (∂x p)(B( − ∂x )∂x p) dx
expectations (σ ∼ ∇u). Slemrod [28,29] noticed that 2 −∞ 5 5
 ∞
the proper interpretation of this term is the capillarity
tension rather than the viscous stress. This is made clear = (∂x u)(A( − ∂x2 )∂x u) dx.
−∞
by inspection of the energy equation. Let us derive the
(60)
energy equation for the simple model (39). Find the time
derivative of the kinetic energy due to the first two Equa- It is crucially important that the functions A(k2 ) and
tion (39): B(k2 ) are negative, indeed, despite the fact that some of
 ∞  ∞ the Taylor coefficients may be positive and, therefore, the
1
∂t u2 dx = u∂t u dx truncation of the formula at some powers of ∂x may not
2 −∞ −∞
 ∞  ∞ work. We have to use either the whole series or special
=− u∂x p dx − u∂x σ dx approximations which preserve negativity of A and B.
−∞ −∞
  ∞ Slemrod [28] represents the structure of the obtained
1 3 ∞ 2 energy formula as
= − ∂t p dx + σ ∂x u dx.
2 5 −∞ −∞
(57) ∂t (MECHANICAL ENERGY)
+ ∂t (CAPILLARITY ENERGY) (61)
Here we used integration by parts under the standard
= VISCOUS DISSIPATION.
assumptions at infinity. Note, that 12 ∂t (p2 ) = − 53 p∂x u.
18 A. N. GORBAN AND I. V. KARLIN

5.2. Matched asymptotics: from k 2 = 0 to k 2 = ∞ These equations give us a clue about the proper asymp-
totic of the continuum mechanic equations for rarefied
For large values of k2 , an analogue of the Chapman–
non-equilibrium gas: we can expect the appearance of
Enskog expansion at an Infinitely distant point is useful.
several factors of the form (1 − α), where  is the
Let us rewrite the algebraic equation for the invariant
Laplace operator.
manifold (52) in the form
 
5 2 4
B + A = −ς + A , AB = ς(A − B), (62) 6. Other approaches: conclusion and outlook
3 3
We presented the main continuum mechanics equations
where ς = 1/k2 .
For the analytic solutions
∞ near the for compressible fluids, from the Euler to the Navier–
point
∞ ς = 0 the Taylor series is: A = l=1 α l ς l, B = Stokes–Fourier and Korteweg equations. The problem
4 4 80
l=1 βl ς , where α1 = − 9 , β1 = − 5 , α2 = 2187 ,
l of deduction of the proper equations for highly non-
4
β2 = 27 , ... . The first term gives for the frequency (55) equilibrium fluxes was formulated. The essential part of
the same limit: Hilbert’s sixth problem is model reduction from kinetics
to continuum mechanics [29,59]. We demonstrated two
2 √
λ± = − ± i|k| 3, (63) classical approaches: solution of the invariance Equations
9 (29), (31) by the Chapman–Enskog series and the Ehren-
and the higher order term give some corrections. fests coarse graining. We solved the reduction problem
Let us match the Navier–Stokes term and the first term exactly for a simple kinetic system (39). This system
in the 1/k2 expansion. Find rational functions A ≈ Ã(k2 ) provided us with a benchmark for comparison of various
and B ≈ B̃(k2 ) such that Ã(0) = B̃(0) = − 43 (the methods and for the explicit demonstration of the van
Navier–Stokes limit) and k2 Ã(k2 ) → − 49 , k2 B̃(k2 ) → der Waals capillarity of ideal gases (60), (61).
There are many attempts to solve the reduction prob-
− 45 when k2 → ∞ (the short wave limit). Solution with
lem and deduce the continuum mechanics equations for
the minimal powers of k2 is:
non-vanishing Knudsen number from the Boltzmann
4 4 equation. We can solve the invariance equation by the di-
A≈− , B≈− (64) rect Newton (or Newton–Kantorovich) method [60]. The
3 + 9k2 3 + 5k2
Newton iterations for the invariance equations provide
and much better results than the Chapman–Enskog expan-
sion. The first iteration gives the Navier–Stokes asymp-
σk = ikA(k2 )uk − k2 B(k2 )pk totic for long waves and the qualitatively correct
4ik 4k2 behaviour with saturation for short waves. The second
≈− u k + pk . (65) iteration gives the proper higher order approximation in
3 + 9k2 3 + 5k2
the long wave limit and the quantitatively proper asymp-
This simplest non-locality captures the main effects: totic for short waves.
the Navier–Stokes approximation for small Knudsen and Another idea is extension of the set of independent
Mach numbers (small k2 ) and the proper asymptotic for variables. Grad proposed to write the equations for higher
short waves (large k2 ) with the saturation of dissipation. moments [65]. His method in combination with the Max-
This saturation is a universal effect [60–64] and hydro- imum Entropy approach got the name ‘Extended irre-
dynamics that do not take it into account cannot pretend versible thermodynamics’ (EIT) [58]. We can start from
to be an universal asymptotic equation. any equation of EIT and apply the method of invariant
For the matched asymptotic (64) we obtain from (46) manifold: write the invariance Equation (29), find the
first terms of the Chapman–Enskog expansion, etc. [66].
5 There appears also a group of methods, which take ac-
∂t p = − ∂x u,
  3 count of some of the higher order terms in the lower order
5 truncation of the Chapman–Enskog expansion [67–69].
(1 − 3∂x2 ) 1 − ∂x2 ∂t u = −∂x p
3 These terms may regularise the singularities in the lower
  orders.
4 5 2
+ ∂x 1 − ∂x ∂x u A rich family of mesoscopic lattice Boltzmann meth-
3 3
ods was developed for applications in fluid dynamics
2 2
+ (1 − 3∂x )∂x p , and beyond [54]. They can be successfully applied to
microfluidics and various other problems between fluid
(66) dynamics and kinetics.
CONTEMPORARY PHYSICS 19

Most of these methods lead beyond continuum me- Notes on contributors


chanics. The Korteweg equation is the first post Navier–
Stokes–Fourier equation, which remains inside A. N. Gorban holds a personal chair in
continuum mechanics but captures some nonequilib- Applied Mathematics at the University
rium kinetic phenomena like the capillarity of ideal gas. of Leicester since 2004. He had worked
for Russian Academy of Sciences,
Exact solutions of the reduction problem (from kinetics
Siberian Branch (Krasnoyarsk, Russia)
to hydrodynamics) give us hints of how the post Navier– and ETH Zürich (Switzerland), had
Stokes equations may look. been a visiting professor and research
Finally, it should be noted that the dissipation and the scholar at Clay Mathematics Institute
capillarity terms in the energy equation resulting from (Cambridge, US), IHES (Bures–sur-
the exact summation are of same order if the gradients Yvette, Île de France), Courant Institute
of Mathematical Sciences (NY, US) and Isaac Newton Institute
are not small. Thus, Korteweg’s term is not just a small for Mathematical Sciences (Cambridge, UK). Main research
correction to dissipation but rather a contribution to the interests include Dynamics of systems of physical, chemical and
energy balance on the same scale. Korteweg’s equations biological kinetics; Biomathematics; Data mining and model
were originally introduced in relation to non-ideal gas reduction problems.
equation of state to capture the effect of surface tension
between different phases. I.V. Karlin is a faculty member at the
So, after all, why capillarity emerges in ideal gas? The Department of Mechanical and Process
answer to this question is in the nature of the inter- Engineering, ETH Zurich, Switzerland.
face of the brick of matter in Cauchy stress construc- He was Alexander von Humboldt Fellow
at the University of Ulm (Germany),
tion. Whenever the gradients of the hydrodynamic fields
CNR Fellow at the Institute of Applied
become commensurable with the mean free path, there Mathematics CNR ‘M. Picone’ (Rome,
is an energy price to be paid for their maintenance. The Italy) and senior lecturer in Multiscale
highly idealised conventional picture of continuous me- Modeling at the University of Southamp-
dia assumes an almost impenetrable elastic interface ton (England). His main research
(Euler) with only a small smearing (Navier–Stokes interests include Exact and non-perturbative results in kinetic
theory; Fluid dynamics; Entropic lattice Boltzmann method;
–Fourier) of the order of a mean free path. However, Model reduction for combustion systems.
when the gradients increase, also the dispersion effects
come into play which is precisely what the surface energy
is responsible for in Korteweg’s picture. It is clear that
the non-locality associated with this effect is essential References
since no truncation of the Chapman–Enskog series is
possible. [1] J.D. van der Waals, The thermodynamic theory of
capillarity under the hypothesis of a continuous variation
We demonstrated how the van der Waals capillarity of density, J. Stat. Phys. 20 (1979), pp. 197–200 (original
appears in dynamics of ideal gas. The above assertion is work published 1893; translated by J.S. Rowlinson).
based on the linearised model equations, the only case [2] D.J. Korteweg, Sur la forme que prennent les équations
so far which was amenable to the exact solution. The du mouvements des fluides si l’on tient compte des forces
non-linear case still requires work but the useful clues capillaires causées par des variations de densité [On the
form of equations of motion of fluids when capillary
are provided by Korteweg’s equations and kinetic models
forces due to variations of density are taken into account],
with exactly solvable reduction problem. Archives Néerl. Sci. Exactes Nat. Ser. II 6 (1901), pp. 1–24.
[3] L.D. Landau, On the theory of phase transitions, Zh. Eksp.
Notes Teor. Fiz. 7 (1937), pp. 19–32 (Reprinted in Collected
papers of L.D. Landau, Pergamon Press, London, 1965,
1. The most important equations are boxed. pp. 193–216.).
[4] V.L. Ginzburg and L.D. Landau, Zh. Eksp. Teor. Fiz. 20
Disclosure statement (1950), pp. 1064–1082 (Reprinted in Collected papers of
L.D. Landau, Pergamon Press, London, 1965, pp. 546–
No potential conflict of interest was reported by the authors. 568).
[5] J.W. Cahn and J.E. Hilliard, Free energy of a nonuniform
Funding system. I. Interfacial free energy, J. Chem. Phys 28 (1958),
pp. 258–266.
ANG was supported by the EPSRC [grant number EP/N022653/1], [6] J.S. Langer, M. Bar-On, and H.D. Miller, New
IVK acknowledges support by the ERC [grant number 291094- computational method in the theory of spinodal
ELBM]; SNSF [grant number 200021_149881]. decomposition, Phys. Rev. A. 11 (1975), pp. 1417–1429.
20 A. N. GORBAN AND I. V. KARLIN

[7] C. Truesdell and W. Noll, The Non-linear Field Theories [24] A.N. Gorban and I.V. Karlin, Structure and approxima-
of Mechanics, Springer, Berlin-Heidelberg, 2004. tions of the Chapman-Enskog expansion, Sov. Phys. JETP
[8] Y.I. Dimitrienko, Nonlinear Continuum Mechanics 73 (1991), pp. 637–641.
and Large Inelastic Deformations, Springer, Berlin- [25] A.N. Gorban and I.V. Karlin, Short-wave limit of
Heidelberg, 2011. hydrodynamics: a soluble example, Phys. Rev. Lett. 77
[9] J.E. Dunn and J. Serrin, On the thermomechanics of (1996), pp. 282–285.
interstitial working, Arch. Ration. Mech. Anal. 88 (1985), [26] I.V. Karlin, Exact summation of the Chapman-Enskog
pp. 95–133. expansion from moment equations, J. Phys. A 33 (2000),
[10] L. Saint-Raymond, Hydrodynamic Limits of the Boltz- pp. 8037–8046.
mann Equation, Vol. 1971, Lecture Notes in Mathematics. [27] I.V. Karlin and A.N. Gorban, Hydrodynamics from Grad’s
Springer, Berlin, 2009. equations: What can we learn from exact solutions?, Ann.
[11] J. Von Neumann and H.H. Goldstine, On the Principles of Phys. (Leipzig) 11 (2002), pp. 783–833.
Large Scale Computing Machines, in Collected Works, by [28] M. Slemrod, Chapman–Enskog ⇒ viscosity-capillarity,
John Von Neumann, Vol. 5, Pergamon Press, New York Quart. Appl. Math. 70 (2012), pp. 613–624.
1961, pp. 1–33 (original work published 1946). [29] M. Slemrod, From Boltzmann to Euler: Hilbert’s 6th
[12] K.C. Harper, Weather by the Numbers, MIT Press, problem revisited, Comput. Math. Appl. 65 (2013), pp.
Cambridge, MA, 2012. 1497–1501.
[13] A. Pikovsky and A. Politi, Lyapunov Exponents: A Tool to [30] A.V. Bobylev, The Chapman-Enskog and Grad methods
Explore Complex Dynamics, Cambridge University Press, for solving the Boltzmann equation, Sov. Phys. Dokl. 27
Cambridge, UK, 2016, . (1982), pp. 29–31.
[14] S.L. Brunton, J.L. Proctor, and J.N. Kutz, Discovering [31] F. Huang, Y. Wang, and T. Yang, Justification of limit for
Governing Equations from Data by Sparse Identification the Boltzmann equation related to Korteweg theory, Quart.
of Nonlinear Dynamical Systems, Proceedings of the Appl. Math. 74 (2016), pp. 719–764.
National Academy of Sciences 113 (2016), pp. 3932–3937. [32] R. Peierls, Model-making in physics, Contemp. Phys. 21
[15] L.I. Kelvin, Nineteenth century clouds over the dynamical (1980), pp. 3–17.
theory of heat and light, Philos. Mag. J. Sci. 2 (1901), pp. [33] J.W. Gibbs, On the Equilibrium of Heterogeneous
1–40. Substances, in The Scientific Papers of J. Willard Gibbs,
[16] O. Reynolds, On certain dimensional properties of matter Vol. 1, H.A. Bumstead and R.G. Van Nameeds. Ox Bow
in the gaseous state. Part I. Experimental researches on Press, Woodbridge, CT, 1993, pp. 55–354.(original work
thermal transpiration of gases through porous plates and published 1876–1878).
on the laws of transpiration and impulsion, including an [34] B. Derjaguin and G. Sidorenkov, Thermoosmosis
experimental proof that gas is not a continuous plenum. at ordinary temperatures and its analogy with the
Part II. On an extension of the dynamical theory of gas, thermomechanical effect in helium II, Doklady Acad. Sci.
which includes the stresses, tangential and normal, caused USSR 32 (1941), pp. 622–666.
by a varying condition of gas, and affords an explanation [35] H.P. Hutchison, I.S. Nixon, and K.G. Denbigh, The
of the phenomena of transpiration and impulsion, Proc. thermo-osmosis of liquids through porous materials,
Philos. Trans. Roy. Soc. 170 (1879), pp. 727–845. Discuss Faraday Soc. 3 (1948), pp. 86–94.
[17] J.C. Maxwell, On stresses in rarefied gases arising from [36] R.P. Rastogi, Introduction to Non-equilibrium Physical
inequalities of temperature, Phil. Trans. Roy. Soc. 170 Chemistry, Elsevier, Amsterdam, 2007.
(1879), pp. 231–256. [37] S. Semenov and M. Schimpf, Thermoosmosis as driving
[18] M. Knudsen, Eine Revision der Gleichgewichtsbedingung mechanism for micro- or nanoscale engine driven by
der Gase, Thermische Molekularströmung, Ann. Phys. 31 external temperature gradient, J. Phys. Chem. C 119
(1910), pp. 205–229. (2015), pp. 25628–25633.
[19] Y. Sone, Flows induced by temperature fields in a rarefied [38] T. Takaishi and Y. Sensui, Thermal transpiration effect of
gas and their ghost effect on the behavior of a gas in the hydrogen, rare gases and Methane, Trans. Faraday Soc. 59
continuum limit, Annu. Rev. Fluid Mech. 32 (2000), pp. (1963), pp. 2503–2514.
779–811. [39] Y.J. Kim, M.G. Lee, and M. Slemrod, Thermal creep of
[20] T. Ohwada, Y. Sone, and K. Aoki, Numerical-analysis of a rarefied gas on the basis of non-linear Korteweg-theory,
the shear and thermal creep flows of a rarefied gas over Arch. Ration. Mech. Anal. 215 (2015), pp. 353–379.
a plane wall on the basis of the linearized Boltzmann- [40] C. Renney, A. Brewer, and T.J. Mooibroek, Easy
equation for hard-sphere molecules, Phys. Fluids A 1 demonstration of the Marangoni effect by prolonged and
(1989), pp. 1588–1599. directional motion: ‘Soap Boat 2.0’, J. Chem. Ed. 90 (2013),
[21] S.K. Loyalka, N. Petrellis, and T.S. Storvick, Some pp. 1353–1357.
numerical results for the BGK model: thermal creep and [41] D.W. Breck, Zeolite Molecular Sieves: Structure,
viscous slip problems with arbitrary accommodation at the Chemistry, and Use, Wiley, New York, 1974.
surface, Phys. Fluids 18 (1975), pp. 1094–1099. [42] I. Steinbach, F. Pezzolla, B. Nestler, M. Seeßelberg, R.
[22] J.S. Rowlinson and B. Widom, Molecular Theory of Prieler, G.J. Schmitz, and J.L. Rezende, A phase field
Capillarity, Oxford University Press, Oxford, 1989. concept for multiphase systems, Phys. D 94 (1996), pp.
[23] S. Chapman and T. Cowling, Mathematical Theory of 135–147.
Non-uniform Gases, 3rd ed., Cambridge University Press, [43] L.K. Antanovskii, A phase field model of capillarity, Phys.
Cambridge, UK, 1970. Fluids 7 (1995), pp. 747–753.
CONTEMPORARY PHYSICS 21

[44] P. Ehrenfest and T. Ehrenfest, Begriffliche Grundlagen der [57] N. Kazantzis and T. Good, Invariant manifolds and
statistischen Auffassung in der Mechanik, in Enzyklopädie the calculation of the long-term asymptotic response of
der Mathemathischen Wissenschaften, Bd. IV 2, II, Heft nonlinear processes using singular PDEs, Comput. Chem.
6, (Reprinted in P. Ehrenfest, Collected Scientific Papers, Eng. 26 (2002), pp. 999–1012.
North-Holland, Amsterdam 1959 (1911), pp. 213–300). [58] D. Jou, J. Casas-Vázquez, and G. Lebon, Extended
[45] A.N. Gorban, I.V. Karlin, P. Ilg, and H.C. Öttinger, Irreversible Thermodynamics, Springer, Berlin, 1993.
Corrections and enhancements of quasi-equilibrium states, [59] A.N. Gorban and I. Karlin, Hilbert’s 6th Problem: Exact
J. Non-Newtonian Fluid Mech. 96 (2001), pp. 203–219. and approximate hydrodynamic manifolds for kinetic
[46] A.N. Gorban, I.V. Karlin, and H.C. Öttinger, H.C., equations, Bull. Amer. Math. Soc. 51 (2014), pp. 186–246.
and L.L. Tatarinova, Ehrenfests’ argument extended to a [60] A.N. Gorban and I.V. Karlin, Method of invariant
formalism of nonequilibrium thermodynamics, Phys. Rev. manifolds and regularization of acoustic spectra, Trans.
E 63 (2001), article number 066124. Theory Stat. Phys. 23 (1994), pp. 559–632.
[47] I.V. Karlin, L.L. Tatarinova, A.N. Gorban, and [61] P. Rosenau, Extending hydrodynamics via the regulariza-
H.C. Öttinger, Irreversibility in the short memory tion of the Chapman-Enskog expansion, Phys. Rev. A 40
approximation, Phys. A 327 (2003), pp. 399–424. (1989), pp. 7193–7196.
[48] F. Reif, Fundamentals of Statistical and Thermal Physics, [62] A.N. Gorban and I.V. Karlin, Quasi-equilibrium
Waveland Press Inc, Nong Grove, 1965. approximation and non-standard expansions in the theory
[49] T. Carleman, Sur la théorie de l’équation intégrod- of the Boltzmann kinetic equation, in Mathematical
ifférentielle de Boltzmann [On the theory of the Modelling in Biology and Chemistry. In New Approaches,
integrodifferential Boltzmann equation], Acta Math. 60 R. G. Khlebopros, ed., Nauka, Novosibirsk, 1991, pp. 69–
(1933), pp. 91–146. 117.
[50] R.J. DiPerna and P.L. Lions, On the Cauchy problem for [63] A. Kurganov and P. Rosenau, Effects of a saturating
Boltzmann equations: global existence and weak stability, dissipation in Burgers-type equations, Comm. Pure Appl.
Ann. Math. 130 (1989), pp. 321–366. Math. 50 (1997), pp. 753–771.
[51] L. Desvillettes and C. Villani, On the trend to [64] M. Slemrod, Renormalization of the Chapman–Enskog
global equilibrium for spatially inhomogeneous systems: expansion: Isothermal fluid flow and Rosenau saturation,
the Boltzmann equation, Invent. Math. 159 (2005), J. Stat. Phys. 91 (1998), pp. 285–305.
pp. 245–316. [65] H. Grad, On the kinetic theory of rarefied gases, Comm.
[52] H.P. McKean Jr, A simple model of the derivation of Pure Appl. Math. 2 (1949), pp. 331–407.
fluid mechanics from the Boltzmann equation, Bull. Amer. [66] I.V. Karlin, A.N. Gorban, G. Dukek, and T.F. Nonnen-
Math. Soc. 75 (1969), pp. 1–10. macher, Dynamic correction to moment approximations,
[53] A.N. Gorban and I.V. Karlin, Invariant Manifolds for Phys. Rev. E 57 (1998), pp. 1668–1672.
Physical and Chemical Kinetics, in Lecture Notes in [67] A.V. Bobylev, Instabilities in the Chapman-Enskog
Physics, Vol. 660, Springer, Berlin–Heidelberg, 2005, . expansion and hyperbolic Burnett equations, J. Stat. Phys.
[54] S. Succi, The Lattice Boltzmann Equation for Fluid 124 (2006), pp. 371–399.
Dynamics and Beyond, Clarendon Press, Oxford, 2001. [68] Y. Sone, Kinetic Theory and Fluid Dynamics, Springer,
[55] A.M. Lyapunov, The General Problem of the Stability of New York, 2012.
Motion, Taylor & Francis, London, 1992. [69] H. Struchtrup, Macroscopic Transport Equations for
[56] N. Kazantzis and C. Kravaris, Nonlinear observer design Rarefied Gas Flows, Springer, Berlin-Heidelberg, 2005.
using Lyapunov’s auxiliary theorem, Systems & Control
Lett. 34 (1998), pp. 241–247.

S-ar putea să vă placă și