Full Multiscale and Multiphysics Flow Simulations of Using The Boltzmann Equation Applications To Porous Media and MEMS Jun Li Ebook All Chapters
Full Multiscale and Multiphysics Flow Simulations of Using The Boltzmann Equation Applications To Porous Media and MEMS Jun Li Ebook All Chapters
com
https://textbookfull.com/product/multiscale-and-
multiphysics-flow-simulations-of-using-the-boltzmann-
equation-applications-to-porous-media-and-mems-jun-li/
OR CLICK BUTTON
DOWNLOAD NOW
https://textbookfull.com/product/modeling-phenomena-of-flow-and-
transport-in-porous-media-1st-edition-jacob-bear-auth/
textboxfull.com
https://textbookfull.com/product/the-lattice-boltzmann-equation-for-
complex-states-of-flowing-matter-sauro-succi/
textboxfull.com
https://textbookfull.com/product/multiphase-fluid-flow-in-porous-and-
fractured-reservoirs-1st-edition-wu/
textboxfull.com
https://textbookfull.com/product/routes-to-absolute-instability-in-
porous-media-antonio-barletta/
textboxfull.com
Physicochemical Fluid Dynamics in Porous Media
Applications in Petroleum Geosciences and Petroleum
Engineering Mikhail Panfilov
https://textbookfull.com/product/physicochemical-fluid-dynamics-in-
porous-media-applications-in-petroleum-geosciences-and-petroleum-
engineering-mikhail-panfilov/
textboxfull.com
https://textbookfull.com/product/the-physics-of-composite-and-porous-
media-1st-edition-t-j-t-tim-spanos/
textboxfull.com
https://textbookfull.com/product/wireless-mems-networks-and-
applications-1st-edition-uttamchandani/
textboxfull.com
https://textbookfull.com/product/album-of-porous-media-structure-and-
dynamics-ezequiel-f-medici-alejandro-d-otero/
textboxfull.com
https://textbookfull.com/product/heat-and-mass-transfer-in-drying-of-
porous-media-1st-edition-peng-xu-editor/
textboxfull.com
Jun Li
Multiscale and
Multiphysics
Flow Simulations
of Using the
Boltzmann Equation
Applications to Porous Media and MEMS
Multiscale and Multiphysics Flow Simulations
of Using the Boltzmann Equation
Jun Li
123
Jun Li
Center for Integrative Petroleum Research
College of Petroleum Engineering and Geosciences
King Fahd University of Petroleum and Minerals
Dhahran, Saudi Arabia
This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Dedicated to open-minded you.
Preface
This book presents a concise introduction to the kinetic theory, based on which
different simulation methods that were independently developed for solving prob-
lems of different fields can be naturally related to each other. Then, the advantages
and disadvantages of different methods will be discussed with reference to each
other. It mainly covers four advanced simulation methods based on the Boltzmann
equation (i.e., direct simulation Monte Carlo method, direct simulation BGK
method, discrete velocity method, and lattice Boltzmann method) and their appli-
cations with detailed results.
I tried to present an introduction to different theories of fluid dynamics starting
from the traditional macroscopic description (i.e., Navier-Stokes-Fourier equations),
which ends with discussion on its limitation in solving gas flows of high Knudsen
number (Kn), including low-speed gas flows in vacuum systems, micro/nano-
electro-mechanical systems and unconventional rock, as well as supersonic gas
flows. This limitation is usually due to the invalidity of the ordinary constitutive
equations and boundary conditions at high Kn. Subsequently, the microscopic sta-
tistical description based on the Boltzmann equation at the molecular scale as well as
the relevant simulation methods with applications is introduced.
In discussing the existing simulation methods, it was necessary to omit some
topics, which are detailed elsewhere as indicated in the references, to make the book
concise. This omission does not compromise the aim of the book, which is to
provide a basic understanding of the potentials of different simulation methods
based on the Boltzmann equation in solving different problems at multiscale. The
choice of reference coverage certainly reflects a personal preference and I apologize
for any omission of other important and relevant work. This book can serve readers
who are interested in the mathematical derivations of different theories in Chaps. 1
and 2, or the relevant simulation methods and applications in the subsequent
chapters.
vii
Contents
ix
x Contents
BGK Bhatnagar-Gross-Krook
Ca Capillary number, the ratio of viscous force to interfacial tension
CFD Computational fluid dynamics
CFL Courant-Friedrichs-Lewy
D2Q9 Two-dimensional nine-velocity-grids model
DSBGK Direct simulation BGK
DSMC Direct simulation Monte Carlo
DUGKS Discrete unified gas-kinetic scheme
DVM Discrete velocity method
EOS Equation of state
FEM Finite element method
GDVM Godunov-type discrete velocity method
GHS Generalized hard sphere
GSS Generalized soft sphere
HPC High-performance-computing
HS Hard sphere
IFT Interfacial tension coefficient
Kn Knudsen number, the ratio of molecular mean free path to flow
characteristic length
LBM Lattice Boltzmann method
LES Large eddy simulation
LVDSMC Low-variance deviational simulation Monte Carlo
Ma Mach number, the ratio of flow speed to sound speed
MD Molecular dynamics
MEMS Micro-electro-mechanical systems
NEMS Nano-electro-mechanical systems
N–S–C–H Navier–Stokes and Cahn–Hilliard
N–S–F Navier–Stokes–Fourier
N–S–K Navier–Stokes–Korteweg
NTC No-time-counter
xi
xii Acronyms
The study of flow problem is very important for many applications and the approach
varies across disciplines. The widely used theory is based on the continuum assump-
tion by using macroscopic quantities (e.g., flow velocity, density, pressure and
temperature) that can be conveniently measured and are close to our ordinary
concepts about the flow problem. Its derivation could be classified into the Eulerian
and Lagrangian descriptions, and the later can be presented in a neat mathematical
form with clear implementations of the basic physical laws (e.g., mass, momentum
and energy conservations) to a closed fluid parcel composed of fixed material par-
ticles, as introduced in this chapter. The obtained system of governing equations
will be closed by introducing the constitutive equations, which are valid for ordinary
flow problems but become inaccurate as the characteristic scale of flow problem or
the density of gas media decreases, where the kinetic theory based on a statistical
description at the molecular scale can be used as discussed at the end of this chapter.
Additionally, the governing equations will be changed in non-inertial reference frame
and the component form in orthogonal curvilinear coordinate will be different from
that in the Cartesian coordinate, which are also discussed.
Fluid mechanics is a branch of physics and can be divided into fluid statics, i.e.,
the study of fluids at rest, and fluid dynamics, i.e., the study of the effects of force
and heat on fluid motion. The traditional macroscopic theory of fluid mechanics is a
branch of continuum mechanics, which models matter without using the information
at microscopic scale (e.g., molecular position and velocity) [1]. The discussion of
this chapter will focus on the macroscopic description of fluid dynamics problems of
single-component system, where chemical reaction and ionization are neglected for
simplicity. The solution to a fluid dynamics problem typically involves calculating
© Springer Nature Switzerland AG 2020 1
J. Li, Multiscale and Multiphysics Flow Simulations of Using
the Boltzmann Equation, https://doi.org/10.1007/978-3-030-26466-6_1
2 1 Fluid Mechanics Based on Continuum Assumption
various distributions of the fluid properties, including flow velocity, pressure, density
and temperature, as functions of space and time.
All matters are made of atoms, have voids and thus are discrete at microscopic
scale. Consequently, the spatial distributions of physical properties are discontin-
uous. Nevertheless, many flow problems can be modeled as continuous using the
spatial and temporal distributions of macroscopic properties in partial differential
equations (PDEs), which is based on the continuum assumption. Under the contin-
uum assumption, macroscopic properties (e.g., density, pressure, temperature, and
bulk velocity or flow velocity) are taken to be well-defined at infinitesimal volume
elements, which are smaller than the characteristic length of the system but much
larger than the molecular average distance. Fluid macroscopic properties can vary
continuously from one volume element to another and are measured as the average
values of the relevant molecular properties within the volume element concerned [1].
Therefore, when continuum mechanics refers to a mathematical point in a continuous
body, it means an infinitesimal element of the body enclosing that point, instead of a
physical point in the interatomic space. This continuum assumption holds valid down
to a critical scale, above which the body concerned can be continually sub-divided
into smaller infinitesimal volume elements with measured macroscopic properties
of interest being almost unchanged. Further dividing the volume element to make it
much smaller than that critical scale will bring noticeable stochastic noise to the mea-
sured macroscopic properties due to insufficient samples (i.e., insufficient molecules)
within the volume element concerned. Temperature can be measured using the kinetic
energy of molecular random velocity and much more molecule samples are required
to obtain smooth measurement for temperature than for density that is subject only to
the randomness of molecular position when counting the molecular number within
the concerned volume element. Once the required molecular number for smooth
measurement and molecular average distance are known, the critical scale of each
macroscopic property can be determined, e.g., the critical scale will be 400 nm if the
molecular average distance is 4 nm and a million molecules are required to smooth
the particular measurement. The critical scale can be reduced if the time-averaging or
ensemble-averaging process is suitable to increase the samples. On the other hand, if
the size of volume element is very large and comparable to the characteristic length
of the flow problem, the measured macroscopic properties are usually not constant
due to their spatial variations at the macroscopic scale when the volume element is
continually sub-divided.
As we can see from the above discussion, the validity of macroscopic description
of flow problems based on the continuum assumption requires the existence of a
definition scale L d , around which the measurements of macroscopic properties are
definite and independent of the size of volume element, namely without the influence
due to spatial variation at the macroscopic scale or randomness at microscopic scale.
In two-dimensional problems, we can always choose an arbitrary scale much smaller
than the characteristic length as L d and assume the dimension L 3 in the direction
perpendicular to the two-dimensional plane is large enough such that the samples
within the volume element L 2d × L 3 are sufficient to smooth the measurements of
macroscopic properties. In some micro-scale three-dimensional problems, the tran-
1.1 Continuum Assumption 3
When analyzing the motion of fluid, it is necessary to describe the evolution of con-
figuration (e.g., spatial distribution of material particles) with time. The Lagrangian
description uses time t together with the material coordinate x0 , which usually is
the initial spatial coordinate of material particle, to describe the motion and focuses
on the evolution of each individual material particle with fixed x0 . By contrast, the
Eulerian description takes the motion as dynamic field distributions, which change
with the spatial coordinate x and time t, and gives attention to what is occurring at a
fixed spatial point x as time goes on.
In the Eulerian description, the object of study is an open control volume that usu-
ally has fixed spatial position and shape, through which the fluxes of mass, momentum
and energy are studied according to the balance laws to derive the governing equa-
tions of density, flow velocity and temperature, etc. In the Lagrangian description,
the object of study is fluid parcel, which is a closed and moving system composed
of fixed material particles, and thus the conservation laws can be directly applied to
obtain the governing equations. As a closed system, the mass of fluid parcel remains
constant but its volume may change in compressible flows. Its shape usually changes
due to the distortion by motion. The fluid parcel needs to be small enough so that
the macroscopic properties measured within each fluid parcel can accurately reflect
the spatial variation at the macroscopic scale. It also needs to be large enough so
that its macroscopic properties are not subject to stochastic noise associated with the
random molecular movements at the microscopic scale. The Lagrangian description
will be used in the following derivations.
According to the classical Newton’s laws, the motion of fluid parcel depends on
its external forces: surface force and body forces. Surface force or contact force,
expressed as force per unit area, can act either via the bounding surface of environ-
ment (i.e., external surface force) or an imaginary surface that separates the fluid
parcel from the surrounding ones (i.e., internal surface force). Body forces originate
from sources outside of the fluid parcel and act via force fields (e.g., gravitational
field, electromagnetic field) instead of direct contact. It is usually specified in terms
of force per unit mass or per unit volume. The macroscopic properties carried by
4 1 Fluid Mechanics Based on Continuum Assumption
each fluid parcel will change with motion due to force and heat interactions between
adjacent fluid parcels and via external fields.
In Lagrangian description, the transient position x of the concerned material par-
ticle with given x0 can be expressed using a mapping function, i.e., x = f (x0 , t).
Since x0 is the original position, we have f (x0 , 0) = x0 . The mapping function needs
to have various properties so that it is physically admissible (also see [6]):
• one-to-one bijection, so that the inverse mapping function x0 = f −1 (x, t) exists
and thus can trace backwards where the material particle currently located at x
was located at the initial state,
• at lest twice continuously differentiable with respect to t, so that differential equa-
tions describing the motion can be formulated,
• globally invertible at all times, so that the body cannot intersect itself, e.g., the
material particles initially located within an imaginary/material surface cannot
move out of the material surface that has its shape changed constantly during the
motion but will keep closed for material particles at any subsequent instant,
• orientation-preserving, because a body cannot be continuously deformed into its
mirror image.
The Lagrangian description of arbitrary property Q can be transformed into a
Eulerian description using the inverse mapping function, and vice versa:
which will be simply denoted by Q(x0 , t) = Q(x, t) in the following discussion and
Q as a function will take different form when using different independent variables.
The time derivative of arbitrary property for a given material particle with fixed
x0 is called the material derivative or substantive derivative, D/Dt, which can be
thought as being measured by an observer traveling with the concerned material
particle. Note that the transient position x is also a property of material particle and
its material derivative is the instantaneous flow velocity u of the material particle
concerned, namely:
D
x(x0 , t) = u(x0 , t). (1.2)
Dt
Since x is the transient position of material particle x0 at t, Eq. (1.2) indicates that
the new position at t + Δt will be x + Δtu. Then, Eq. (1.1) takes the following form
at t + Δt:
which, according to the chain rule, implies the following transformation of derivatives
between the Lagrangian and Eulerian descriptions:
D ∂ Q(x, t) ∂ Q(x, t)
Q(x0 , t) = +u· , (1.4)
Dt ∂t ∂x
1.2 Macroscopic Descriptions of Flow Phenomenon 5
where the first term on the right-hand side is the local derivative and the second term
is the convective derivative due to material particle changing spatial position.
Following the rigorous and concise derivations given in [10], the macroscopic gov-
erning equations of flow problems will be derived in Sect. 1.3 using the Lagrangian
description. The obtained governing equations can be transformed into Eulerian
description using Eq. (1.4) if needed.
As mentioned before, the movement of fluid parcel depends on its surface force and
body forces. The surface force can be defined according to the momentum exchange
during the motion through an imaginary surface that separates the concerned fluid
parcel from its surrounding ones. Similar to the fundamental interaction forces (e.g.,
gravitation and electromagnetic forces), the surface force happens between two bod-
ies that need to be closed systems. Thus, it makes sense to discuss the surface force
between fluid parcels. By contrast, it is not appropriate to define the force between
two adjacent open control volumes in the Eulerian description.
To calculate the net surface force of fluid parcel, we divide its closed surface
into many oriented differential surfaces d A. Each d A has an outward unit vector
normal to d A, n+ , and an inward unit normal vector, n− = −n+ . Correspondingly,
the surface force exerted by the surrounding fluid on the concerned fluid parcel via
d A and its counterforce are denoted by dFn+ and dFn− , respectively. Note that the
actual direction of dFn+ has nothing to do with the direction of n+ and will be outward
when the fluid is stretched (if physically possible to fluid), or inward when being
compressed. In non-static state, dFn+ is usually not normal to d A.
We use ei∈[1,2,3] as the unit vectors of Cartesian coordinate system of the reference
frame, where the flow problem is observed or modeled. Surface forces per unit area
or stresses are denoted by Pn+ = dFn+ /d A and Pn− = dFn− /d A=−Pn+ . Usually,
Pn+ can be expressed as a function of velocity gradient using constitutive equation.
But, it will be complicated to calculate the net effect of Pn+ on the whole closed
surface of fluid parcel concerned if the function form depends on n+ . Actually, there
is an inherent correlation between arbitrary Pn+ and the three special ones, i.e., Pe1 ,
Pe2 and Pe3 , which can be specifically expressed using the velocity gradient in a
constitutive equation.
As shown in Fig. 1.1, we study the correlation between stresses of different n+ at
the same spatial point d using a tetrahedron that has one vertex located at the point
d. The tetrahedron has three differential surfaces, d A1 , d A2 and d A3 , located within
the coordinate planes having outward unit normal vectors as −e1 , −e2 and −e3 . The
6 1 Fluid Mechanics Based on Continuum Assumption
outward
‚ unit normal vector of the fourth differential surface, d An , is denoted by n+ .
Since n+ d A = 0 for arbitrary closed surface, we have:
which implies:
dAj
n +, j = n+ · e j = , ( j = 1, 2, 3). (1.6)
d An
Note that the values of four stresses in Eq. (1.7) might not be taken exactly at the
point d (instead, should be at a point determined by the mean value theorem for
integral) because, for example, we used P−e1 d A1 as the net surface force exerted
via d A1 , within which P−e1 still has possible spatial variation. But, Eq. (1.8) shows
the correlation between surface forces associated with different orientations, whose
values are taken at the same point d because the tetrahedron will shrink to the point
d as d An → 0 with fixed n+ .
1.3 Derivations of Macroscopic Governing Equations 7
According to Eq. (1.8), the stress Pn+ of arbitrary n+ can be generally expressed
using the three special stresses Pe j ( j = 1, 2, 3) measured on the coordinate planes.
Since Pe j ( j = 1, 2, 3) denoted with one subscript are vectors, they can be assembled
into a matrix P denoted without subscript:
⎛ ⎞
P11 e1 e1 P12 e1 e2 P13 e1 e3
P = P ji e j ei = ⎝ P21 e2 e1 P22 e2 e2 P23 e2 e3 ⎠ , (1.9)
P31 e3 e1 P32 e3 e2 P33 e3 e3
Pn+ = n +, j Pe j = n +, j P ji ei = n+ · P. (1.10)
αi j = ei · ej , (1.11)
which implies:
The stress matrix P is an objective quantity and invariant when being observed in
different R. Additionally, to be a tensor, its components measured in different R
(i.e., measurements of using different coordinate systems) must satisfy a certain
correlation. In general, the components of an arbitrary nth-order tensor Q (n ≥ 1)
measured in different reference frames satisfy the following inherent correlation:
Q i1 i2 ···in ei1 ei2 · · · ein = Q = Q j1 j2 ··· jn ej1 ej2 · · · ejn . (1.14)
According to the convention used in Eq. (1.9), the measurements of P in the refer-
ence frame R is P j1 j2 = Pej · ej2 , which can be expressed using the measurements
1
of P in R as follows [7]:
8 1 Fluid Mechanics Based on Continuum Assumption
where Eq. (1.8) has been substituted with n+ being replaced by ej1 . Note that
Eq. (1.12) implies αi j αk j = δik and αi j αik = δ jk , where the Kronecker delta δik equals
1 for i = k or 0 for i = k. Now, Eq. (1.15) can be rewritten as follows:
which proves that the stress P is a second-order tensor according to Eq. (1.13).1 This
conclusion holds at both static and dynamic states and also for the stress of solid
body, where the essential condition of Eq. (1.8) is satisfied as well. Additionally, the
stress tensor can be proved symmetric using a derivation of angular momentum.
The stress as a symmetric second-order tensor can be modeled using other sym-
metric second-order tensors. Newtonian fluid is defined to be a fluid whose shear
stress is linearly proportional to the velocity gradient. This definition means regard-
less of the magnitude of external shear force acting on a fluid, it continues to flow
since the internal shear stress needs to be nonzero according to force balance. In con-
trast, solid can withstand shear forces at static state. The usual constitutive equation
of stress for Newtonian fluids is as follows:
2
Pi j = − p + μv − μ Skk δi j + 2μSi j , (1.17)
3
1 Similar proof of existence of stress tensor in Sect. 3.3.3 of [6] ends at Eq. (1.10) here.
1.3 Derivations of Macroscopic Governing Equations 9
fluid (e.g., honey) usually causes the viscosity to decrease and so the fluid appears
shear-thinning. Some fluids are shear-thickening, which is usually not observed in
pure materials but can occur in suspensions. Most fluids with long molecular chains
can react in a non-Newtonian manner. Additionally, fluids can be roughly divided
into ideal and non-ideal fluids in some applications. An ideal fluid is inviscid and
thus has no resistance to shear forces, which does not exist in reality but is an
acceptable model for the fluid far away from the solid boundaries since the viscous
effect is usually concentrated near the solid surfaces (i.e., within the boundary layers).
Correspondingly, the reduced momentum equation without the viscous term is called
the Euler equation [1].
In addition to the stress interaction between fluid parcels, there is also heat flux q,
which is specified in terms of thermal energy exchange per unit area per unit time.
According to the Fourier’s law, the usual constitutive equation of q is as follows:
∂T
qi = −ζ , (1.18)
∂ xi
where ζ is the thermal conductivity coefficient, T is the temperature, and the negative
sign means that heat flux moves from higher temperature regions to lower temperature
regions.
The mass density ρ and volume ΔV = Δx1 Δx2 Δx3 of a fluid parcel might change
with motion due to non-uniform distribution of the flow velocity u. But, as a closed
system, the total mass of fluid parcel is constant:
D Dρ DΔV
(ρΔV ) = ΔV +ρ = 0, (1.19)
Dt Dt Dt
where ρ, by a slight abuse of notation, should be the average density within ΔV .
The material derivative of ΔV can be calculated as follows:
1 DΔV 1 DΔx1 1 DΔx2 1 DΔx3
= + +
ΔV Dt Δx1 Dt Δx2 Dt Δx3 Dt
1 Dx1 1 Dx2 1 Dx3
= Δ + Δ + Δ (1.20)
Δx1 Dt Δx2 Dt Δx3 Dt
Δu k
= .
Δxk
Substituting Eq. (1.20) into Eq. (1.19), we obtain the following continuity equation
at Δx1 , Δx2 , Δx3 → 0:
10 1 Fluid Mechanics Based on Continuum Assumption
Dρ
+ ρ∇ · u = 0, (1.21)
Dt
where ρ is the local density of a particular spatial point, to which the fluid parcel
shrinks as Δx1 , Δx2 , Δx3 → 0. Equation (1.21) holds inside the whole flow domain
because the fluid parcel can be selected arbitrarily. It shows that the density of fluid
parcel doesn’t change as long as ∇ · u = 0 that is the criterion for incompressible
flows. In this sense, gas at the standard temperature and pressure (STP) condition is
very easy to compress but its flow could be deemed incompressible if ∇ · u ≈ 0 due
to very low Mach number (Ma), which is defined as the ratio of the flow speed to the
sound speed. The Eulerian description of continuity equation can be obtained using
Eqs. (1.4) and (1.21):
∂ρ
+ ∇ · (ρu) = 0. (1.22)
∂t
The Newton’s second law applied to a fluid parcel has the following form:
˚ ˚ ‹
D
uρdV = gρdV + n+ · Pd A, (1.23)
Dt
where Pn+ = n+ · P of Eq. (1.10) has been substituted, the volume integral and
surface integral are over the volume and closed surface of the fluid parcel concerned,
respectively.
˝ Since the essence of integral is summation, the order of D/Dt and
can be changed and consequently we have D(uρdV )/Dt = uD(ρdV )/Dt +
ρdV Du/Dt = ρdV Du/Dt according to Eq. (1.19). The surface integral of Eq. (1.23)
can be transformed into a volume integral using the Gauss’s theorem (i.e., divergence
theorem) and thus Eq. (1.23) can be rewritten into:
˚
Du
ρ − gρ − ∇ · P dV = 0. (1.24)
Dt
Since Eq. (1.24) holds for arbitrary integral domain occupied by the corresponding
fluid parcel, the integrand must be zero inside the whole flow domain according to
the localization theorem:
Du
ρ = ρg + ∇ · P. (1.25)
Dt
Substituting the constitutive equation of Eq. (1.17) into Eq. (1.25), we obtain the
following Navier–Stokes (N–S) momentum equation:
1.3 Derivations of Macroscopic Governing Equations 11
Du 2
ρ = ρg − ∇ p + μ − μv ∇ · u + ∇ · (2μS). (1.26)
Dt 3
The Eulerian description of momentum equation can be obtained using Eqs. (1.4)
and (1.26):
∂u ∂u 2
ρ +u· = ρg − ∇ p + μ − μv ∇ · u + ∇ · (2μS), (1.27)
∂t ∂x 3
where the left-hand side can be further transformed to ∂(ρu)/∂t + ∇ · (ρuu) accord-
ing to the continuity equation of Eq. (1.22).
The contributions to the energy change of fluid parcel include the work done by exter-
nal body force, the work done by surface force, and the heat flux from the surrounding
fluid to the fluid parcel concerned. The internal energy (associated with molecular
random translational motion, rotation, vibration and intermolecular potential energy,
but not chemical energy since chemical reaction and ionization are not considered as
stated at the beginning) and kinetic energy (associated with the macroscopic velocity
u) of fluid parcel per unit mass are denoted by e and u 2 /2 = u · u/2, respectively.
Note that similar to the cases of other conservative body forces, the gravitational
potential energy of fluid parcel should be neglected in the total energy when the
work done by gravity is considered in calculating the energy increment. For prob-
lems without volumetric heat sources/sinks, the conservation law of energy applied
to a fluid parcel has the following form:
˚ ˚ ‹
D u2
+ e ρdV = u · gρdV + (n+ · P) · ud A+
Dt 2
‹ (1.28)
−n+ · qd A,
where −n+ · q = n− · q is the heat flux toward the fluid parcel and (n+ · P) · u =
n+ · (P · u). Similar to Eq. (1.23), Eq. (1.28) can be rewritten into:
˚
D u2
ρ + e − u · gρ − ∇ · (P · u) + ∇ · q dV = 0, (1.29)
Dt 2
which holds for arbitrary integral domain occupied by the corresponding fluid parcel
and thus its integrand must be zero inside the whole flow domain:
D u2
ρ ( + e) = u · gρ + ∇ · (P · u) − ∇ · q. (1.30)
Dt 2
12 1 Fluid Mechanics Based on Continuum Assumption
D u2
ρ = u · gρ + (∇ · P) · u, (1.31)
Dt 2
which can be subtracted from Eq. (1.30) to obtain a simpler form as follows:
D
ρ e = P : ∇u − ∇ · q, (1.32)
Dt
Entropy s of fluid parcel per unit mass is also a state variable (or state function)
of thermodynamics and, as a measurement of disorder, is related to the number of
microscopic configurations that a thermodynamic system can have at an equilibrium
state specified by several macroscopic state variables. For a closed system, entropy
1.3 Derivations of Macroscopic Governing Equations 13
will increase with temperature if the volume is fixed because the set of molecular
velocities has more admissible configurations at higher temperature. At a constant
temperature, the number of admissible configurations of molecular position increases
with the volume and thus entropy increases as well. This interpretation of entropy
will become clearer by using a probability distribution function defined inside a six-
dimensional phase space (a superposition of the unbounded molecular velocity space
and the physical space), which will be introduced in the next chapter.
For usual fluid systems of single component, there are only two independent state
variables and thus the material derivative of s can be calculated by applying the chain
rule to e(s, ρ):
De ∂e Ds ∂e Dρ
ρ =ρ |ρ + |s
Dt ∂s Dt ∂ρ Dt
(1.33)
Ds
= ρT − p∇ · u,
Dt
Ds
ρT = p∇ · u + P : ∇u − ∇ · q, (1.34)
Dt
where the right-hand side degenerates to φ = 2μSi j Si j ≥ 0 when ∇ · u = 0, q = 0
and P is modeled by Eq. (1.17), which means that the constitutive equation of
Eq. (1.17) complies with the second law of thermodynamics in the cases of incom-
pressible flows. For compressible flows with q = 0, the right-hand side of Eq. (1.34)
is equal to (μv − 2μ/3)(∇ · u)2 + φ when using Eq. (1.17) and then attentions are
needed to make it always positive.
J = −D∇C, (1.35)
14 1 Fluid Mechanics Based on Continuum Assumption
where D (m2 s−1 ) is the diffusion coefficient. We assume that C is sufficiently low
and thus its influences on ρ and u are negligible. For problems without volumetric
sources/sinks of the species concerned, the conservation law of substance amount
applied to the domain occupied by a fluid parcel has the following form:
˚ ‹
D
CdV = −n+ · Jd A. (1.36)
Dt
Substituting D(dV )/Dt = (∇ · u)dV of Eq. (1.20) into Eq. (1.36), we obtain:
DC
+ C∇ · u = −∇ · J. (1.37)
Dt
Note that the material particles of fluid will not move across the closed imaginary
surface of fluid parcel during the motion when selecting fluid parcel as the object of
study but the species concerned here is allowed to freely penetrate the surface.
The N–S–F equation system is independent of any particular inertial reference frame
but its form will change in non-inertial reference frame. Most of physical quantities
contained in this equation system are objective, including P, q, J, ρ, p, T , e, s and
C. In contrast, the spatial position x, flow velocity u and acceleration a of material
particle depend on the reference frame R, where the flow problem is observed.
Additionally, the constitutive equations of Eqs. (1.17), (1.18) and (1.35) contain other
quantities S, ∇ · u, ∇T , and ∇C whose objectivities determine the objectivities of
these constitutive equations.
In the inertial reference frame R having the origin of coordinates located at the
−
→
point o, we use x for the vector od, where the point d is an arbitrary material particle
(or spatial point) concerned, and have the component form x = xi ei , where ei∈[1,2,3]
are the unit vectors of Cartesian coordinate system of the reference frame R. The
origin of coordinates of the arbitrary non-inertial reference frame R is located at o
−→ −
→
and vector o d is denoted by X. Additionally, let y denote the vector oo and thus:
y = x − X. (1.38)
The velocities of the same material particle observed in two different reference
frames R and R are two different quantities, denoted by u and U, respectively.
They are not equal when the non-inertial R has translational or rotational motion
relative to the inertial R and thus the velocity is variant. But, the divergences of
the two different velocities are always equal, namely ∂Uk /∂ X k ≡ ∂u k /∂ xk , which
implies that the velocity divergence is objective. Additionally, the strain rate tensor
1.4 Transformation to Arbitrary Non-inertial Reference Frame 15
and the gradient of temperature or any other objective scalar quantities are objective
as well, but the vorticity and gradient of velocity are variant. Note that when talking
about whether a quantity is invariant/objective, we compare two counterparts defined
equally but in different reference frames, e.g., compare U with u in discussing the flow
velocity, or compare ∂Uk /∂ X k with ∂u k /∂ xk in discussing the velocity divergence.
For a given vector, which is objective and thus independent of the reference frame,
it has different component forms when being measured in different reference frames:
x = xi ei = x j ej for example, where ej∈[1,2,3] are the unit vectors of Cartesian coor-
dinate system of the reference frame R . Since the Cartesian coordinate system is
used, ei and ej do not change with coordinates when observing the movement of
material point in R and R , respectively. Then, the flow velocity u of material particle
d observed in R can be simply calculated using the corresponding component form
of x, namely u = (Dxi /Dt)ei . Similarly, the flow velocity U of the same material par-
ticle d observed in R can be calculated as U = (DX j /Dt)ej . Note that (Dx j /Dt)ej
is the velocity of d relative to o observed in R , which is different from U that is the
velocity of d relative to o observed in R and also different from u that is the velocity
of d relative to o observed in R. Similarly, (DX i /Dt)ei = u and (DX i /Dt)ei = U.
Note that the temporal variation of a given vector might be different when being
observed in different reference frames. Thus, in Sect. 1.4, we apply the material
derivative only to scalar quantities (i.e., components of tensors) to avoid confusions
incurred when using it directly for tensors. For example, if Dx/Dt and DX/Dt are
used to denote u and U, respectively, we will literally obtain u = U if x ≡ X, which
is wrong because R can still have rotational motion relative to R making u = U
although their origins of coordinates remain overlapped, i.e., x ≡ X.
In general, we have the following correlation between u and U:
Dxi D(yi + X i )
u= ei = ei
Dt Dt
(1.39)
Dyi D(αi j X j ) Dyi Dαi j
= ei + ei = ei + U + X ei ,
Dt Dt Dt Dt j
where (Dyi /Dt)ei and (Dαi j /Dt)X j ei correspond to the translational and rotational
motions of the reference frame R relative to R, respectively. We introduce ω = ωi ei
to make the physical meaning of (Dαi j /Dt)X j ei clear and define ωi as follows:
1 Dαl j
ωi = lik αk j , (1.40)
2 Dt
Dαi j Dαi j
X ei = αk j X k ei = i jk ω j X k ei = ω × X, (1.41)
Dt j Dt
16 1 Fluid Mechanics Based on Continuum Assumption
where ω × X is the cross product of ω and X. Equation (1.41) shows that ω defined
by Eq. (1.40) is the rotational/angular velocity of R relative to R.
Note that yi , y j , αi j , ei and ej are independent of x and X in the Cartesian
coordinate system, which can simplify the calculation of each quantity and thus is
used in R and R for the convenience of discussion. Additionally, Eq. (1.12) implies
x j = αi j xi and xi = αi j x j . For the velocity divergence, we have:
∂ Dyi Dαi j
∇ · u = ek · ei + U + X j ei
∂ xk Dt Dt
∂U j ∂ X j Dαi j
= 0 + ek · e j + ek · ei
∂ xk ∂ xk Dt
∂ X i ∂U j ∂(x j − y j ) Dαi j
= αk j + e k · ei
∂ xk ∂ X i ∂ xk Dt
∂(xi − yi ) ∂U j Dαi j
= αk j + ek αk j · ei − 0
∂ xk ∂ Xi Dt (1.42)
∂U j Dαi j
= αk j αki − 0 + αi j
∂ X i Dt
∂U j 1 D(αi j αi j )
= δ ji +
∂ Xi 2 Dt
∂U 1 Dδ
= ei · ej +
j jj
∂ X i 2 Dt
= ∇ · U,
∂T ∂ X j ∂ T
∇T = ei = ei
∂ xi ∂ xi ∂ X j
(1.43)
∂(x j − y j ) ∂ T ∂T ∂T
= ei = αi j ei = e = ∇ T.
∂ xi ∂Xj ∂Xj ∂ X j j
reference frames although their forms are usually validated in inertial reference frame
by experiments. Detailed derivations and discussion are given in [7].
The N–S momentum equation of Eq. (1.26) was derived in Sect. 1.3.3 using the
Newton’s second law in an inertial reference frame R. According to the above objec-
tivity analyses of different quantities, the objective quantities of Eq. (1.26) defined in
R can be replaced by the corresponding quantities defined in R and thus Eq. (1.26)
can be rewritten into:
Du i 2
ρ ei = ρg − ∇ [ p + ( μ − μv )∇ · U] + ∇ · [μ(∇ U + ∇ UT )], (1.44)
Dt 3
Du i D2 yi DUi D( i jk ω j X k )
ei = ei + ei + ei
Dt Dt 2 Dt Dt
D2 yi D(αi j U j ) Dω j DX k
= ei + ei + i jk X k ei + i jk ω j ei
Dt 2 Dt Dt Dt
D2 yi DU j Dαi j Dω j DX k
= ei + ej + αk j Uk ei + e j × X + i jk ω j ei
Dt 2 Dt Dt Dt Dt
D2 yi DU j Dω j DX k
= ei + e + i jk ω j Uk ei + e j × X + i jk ω j ei
Dt 2 Dt j Dt Dt
D2 yi DU j Dω j DX k
= ei + e +ω×U+ e j × X + i jk ω j ei , (1.45)
Dt 2 Dt j Dt Dt
DX k D(αkl X l )
i jk ω j ei = i jk ω j ei
Dt Dt
Dαkl
= i jk ω j αml X m + αkl Ul ei (1.46)
Dt
= i jk ω j ( klm ωl X m + Uk )ei
= ω × (ω × X) + ω × U.
18 1 Fluid Mechanics Based on Continuum Assumption
Substituting Eqs. (1.45) and (1.46) into Eq. (1.44), we obtain the momentum equation,
which is valid in an arbitrary non-inertial reference frame R . Similar transformations
can be applied to other governing equations.
In the above Sect. 1.4, we discussed the replacement of a quantity defined in an iner-
tial reference frame R by its counterpart defined in an arbitrary non-inertial reference
frame R . This two counterparts will have the same form if the quantity concerned
is objective (e.g., divergence of velocity and strain rate tensor) but additional terms
are needed if not (e.g., acceleration in Eq. (1.45) and vorticity). By contrast, in this
section, the discussion will focus on the transformation between different compo-
nent forms of each given quantity contained in the governing equations of interest
such that the calculations of those quantities can be simplified in axisymmetric or
spherically symmetric flows by using non-Cartesian coordinate system. Note that
this transformation between component forms is also applicable to the governing
equations of R , e.g., Eqs. (1.44)–(1.46), although the quantities defined in R will
be used as examples in the following discussion that is based on the derivations of
[3].
For a given reference frame where the flow problem is observed, it is still possible
to use different coordinate systems for measurement to obtain different component
forms of each quantity of interest. As shown in Fig. 1.2, in the arbitrary orthogonal
curvilinear coordinate system S̃ of using (l1 , l2 , l3 ) as coordinates and (ẽ1 , ẽ2 , ẽ3 )
as coordinate vectors, each spatial point has three coordinates as in the Cartesian
coordinate system S of using (x1 , x2 , x3 ) as coordinates and (e1 , e2 , e3 ) as coordinate
vectors. There is a one-to-one bijection between (x1 , x2 , x3 ) and (l1 , l2 , l3 ) and this
correlation is given once S̃ is selected according to S. ẽi∈[1,2,3] are orthogonal at
arbitrary point but change with (l1 , l2 , l3 ), which is different from the case of S.
−→ −−→ −−→ −→
The increments of l1 associated with bb1 , b2 d3 , b3 d2 and d1 d are the same but
the corresponding spatial distances are usually different. Correspondingly, the Lame
coefficient Hi=1 (similar for i = 2, 3) at arbitrary point b is defined as follows:
∂ x1 2 ∂ x2 2 ∂ x3 2
= ( ) +( ) +( ) .
∂l1 ∂l1 ∂l1
Most people start at our website which has the main PG search
facility: www.gutenberg.org.
Our website is not just a platform for buying books, but a bridge
connecting readers to the timeless values of culture and wisdom. With
an elegant, user-friendly interface and an intelligent search system,
we are committed to providing a quick and convenient shopping
experience. Additionally, our special promotions and home delivery
services ensure that you save time and fully enjoy the joy of reading.
textbookfull.com