You are on page 1of 40

1.

Fundamental Equations for Newtonian Fluids

1.1 Introduction
In this chapter we recall the standard form of the classical fluid dynamics
equations in eulerian form. For the evolution of a fluid in N 1 spatial
dimensions, the description involves ( N + 2 ) fields. These are the mass
density, the velocity field and the energy. The fundamental equations of
fluid dynamics are derived following the next three universal conservation
laws:
1. Conservation of Mass.
2. Conservation of Momentum or the Newtons Second Law.
3. Conservation of Energy or the First Law of Thermodynamics.
The resulting equations are the continuity equation, the momentum
equations and the energy equation, respectively. The system of governing
equations is closed by:
1. Statements characterizing the thermodynamic behavior of the fluid.
These are called state equations. The state equations do not depend
on the fluid motion and represent relationships between the
thermodynamic variable: pressure (p), density ( ), temperature (T),
specific internal energy per unit mass (e) and total enthalpy per unit
mass (h). The microscopic transport properties like the dynamic
viscosity ( ) and the thermal conductivity (k) are also related to the
thermodynamic variables.
2. Constitutive relations, which are postulated relationships between
stress and rate of strain and heat flux and temperature gradient. The
previously mentioned microscopic transport properties appear in the
constitutive relations as proportionality factors (between stress and
rate of strain and heat flux and temperature gradient) and must be
determined experimentally.

The flow is considered as known if, at any instant in time, the velocity field
and a precise number of thermodynamic variables are known at any point.
This number equals two for a real compressible fluid in thermodynamic
equilibrium (for example, the pressure and the density).
There is some freedom in choosing a set of variables to describe the fluid
flow. A possible choice is the set of primitive or physical variables,
namely the density, pressure and velocity components along the three
directions of a reference frame, i.e. the set ( , p, u, v, w ) . Another choice is
the set of the so-called conservative variables which are the density, the
three momentum components and the total energy per unit mass, i.e. the set
( , u, v, w, E ) . The set of conservative variables results naturally from
the application of the above fundamental laws of conservation. From a
computational point of view, there are some advantages in expressing the
governing equations in terms of the conservative variables. We will call
conservative methods the class of numerical methods based on the
formulation of the governing equations in conservative variable.
The primitive and conservative variables are dependent variables. They are
depending on time and space coordinates, i.e. on the independent
variables. For example, the density is = ( t , x, z, y ) . Recall that the
number of the spatial coordinates is N.
The reader is invited here to notice an important difference between the
governing equations and the closure conditions. The governing equations
are exact mathematical conditions that must be satisfied by the flow
variables. The state equations and the constitutive relations are only
practical models and they depend on the degree of knowledge the user has
about the fluid and flow under consideration. Thus, they represent sources
of uncertainties that are introduced in the mathematical model of the flow.
The derivation of the fundamental equations mentioned above will not be
presented here. Almost all of the books devoted to fluid mechanics or to
CFD have chapters devoted to this issue. For example, more details on the
equations can be found in (Hirsch, 1990), (Ferziger and Peric, 1999) and
(Warsi, 1999).

1.2 Preliminary Concepts


Before proceeding to the derivation of the equations in integral form review
some preliminary concepts that are needed. Consider a scalar field function
( x, y, z , t ) , then the time rate of change of as registered by an observer
moving with the fluid velocity V = (u.v, w) is given by the substantial
derivative (or material derivative) defined as:
D
=
+ V grad
(1.1)
Dt t

denotes the partial derivative of with respect to time


The first term
t
and represents the local rate of change of while the second term is the
convective rate of change. If the scalar field is replaced by a vector field, the
D
operator
can also be applied in a component-wise manner. In particular,
Dt
we can obtain the substantial derivatives of V = (u.v, w) , that is:
DV V
=
+ V gradV
(1.2)
Dt
t

which in full becomes:


T
T
u v w
Du u

Dt t
x x x

Dv = v + (u , v, w) u v w
(1.3)
u y y
Dt t

u v w
Dw w

z z z
Dt t

The left hand side of (1.3) is the acceleration vector of an element of moving
fluid. Let us now consider the new scalar field
(t ) = ( x, y, z , t )dV
(1.4)

where the volume of integration V is enclosed by a piece-wise smooth


boundary surface A = V that moves with the material under consideration.
It can be shown that the material derivative of is given by:
D

= V
dV + A (n V ) dA
(1.5)
Dt
t

where n = (n1 , n2 , n3 ) is the outward pointing unit vector normal to the


surface A. Expression (1.5) can be generalized to vectors ( x, y, z , t ) as
follows:
D

= V
dV + A (n V )dA
(1.6)
Dt
t
The surface integrals in the last equalities may be transformed to a volume
integral by virtue of Gausss theorem. This states that for any differentiable
vector field = (1 , 2 , 3 ) and a volume V with smooth bounding surface A
the following identity holds:
(1.7)
(n )dA = div dV
A

The above identity also applies to differentiable scalar and tensor fields. If
we use it, for example, to transform the right hand side of equation (1.5) into
a volume integral, one obtains:
D

(1.8)
= V + div ( V ) dV
Dt
t

Please notice that in the above relations the volume V is arbitrary and
moving with the fluid particles. Let J be the jacobian of the transform that
links two successive positions of the volume V. There is a theorem due to
Euler which states that the time derivative of the jacobian is (Lions, 1996):
DJ
= J divV
(1.9)
Dt
From (1.9) we remark that mathematically, the incompressibility means:
divV = 0
(1.10)

1.3 Integral Form of the Governing Equations


1.3.1 Conservation of Mass

The law conservation of mass can now be stated in integral form by


identifying the scalar in (1.4) as the density . In this case (t ) in (1.5)
becomes the total mass in the volume V. By assuming that no mass is
generated or annihilated within the material control volume V, we have
D
= 0 , or:
Dt

dV + A n ( V )dA = 0
(1.11)
t
This is the integral form of the law of the conservation of mass. This
integral conservation law may be generalised to include sources of mass ,
which will then appear as additional integral terms. A useful reinterpretation
of the integral form result if we rewrite it as

(1.12)
V t dV = A n ( V)dA
If now V is a fixed control volume then the left hand side of (1.12) becomes:

d
(1.13)
V t dV = dt V dV
and thus is the time-rate of change of the mass enclosed by the volume V.
The right hand side of (1.12) shows that the mass enclosed by the control
volume V, in the absence of sources or skins, can only change by virtue of
mass flow through the boundary of the control volume V. One obtains from
(1.8) and (1.12):

(1.14)
V t + div( V ) dV = 0
As V is arbitrary it follows that the integrand must vanish, that is:
t + ( u ) x + ( v) y + ( w) = 0
(1.15)
which is the differential form of the continuity equation.

We notice here that the differential and integral forms (1.15) and (1.14) are
not valid in general. Under the assumption of inviscid fluids, the governing
equations admit discontinuities in the solution, such as shock waves and
contract surfaces. The integral form (1.11), however, remains valid.
1.3.2 Conservation of Momentum

As done for the mass equation, we now provide the foundations for the law
of conservation of momentum, derive its integral form in quite general
terms and show that under appropriate smoothness assumptions the
differential form is implied by the integral form. A control volume V with
bounding surface A is chosen and the total momentum in V is given by:
(t ) = VdV
(1.16)
V

The law of conservation of momentum results from the direct application of


Newtons law: the time rate of change of the momentum in V is equal to the

total force acting on the volume V. The total force is divided into surface
forces fs and volume fv given by:
fs = SdA, fv = gdV
(1.17)
A

Here g is the specific volume-force vector and may account for inertial
forces, gravitational forces, electromagnetic forces and so on. S is the stress
vector, which is given in terms of a stress tensor as S = n . The stress
tensor can be split into a spherical symmetric part due to pressure p, and
a viscous part :
= pI
(1.18)
Application of Newtons Law gives:

(1.19)
V t ( V)dV = A V (n V )dA + f s + f v
Regarding V as a fixed volume in space, independent of time, we write:
d
( V )dV = V (n V )dA + fs + f v ,
A
dt V
We interpret the above as saying that the time rate of change of momentum
within the fixed control volume V is due to the net momentum inflow over
momentum outflow, given by the first term in (1.19), plus surface and
volume forces.
Substituting the stress tensor from (1.18) into (1.19) and writing all surface
terms into a single integral we have:
d
( V )dV = [ V (n V ) + pn n ] dA + gdV . (1.20)
A
V
dt V
This is a general statement and is valid even for the case of discontinuous
solutions.
The differential conservation law can now be derived from (1.20) under
assumption that the integrand in the surface integral is sufficiently smooth
so that Gausss theorem may be invoked. The first term of the integrand of
the surface integral can be rewritten as:
V (n V ) = n V V ,
where the dyadic (tensor) product V V is a tensor. Its components are
easily determined from the next equality. The three columns of the left hand
side are:

7
T

u (n V ) = n u 2 , uv, uw ,
T

v (n V ) = n uv, v 2 , vw ,
T

w (n V ) = n uw, vw, w2 .
Now one can apply the Gausss theorem to each of the surface terms:

V t ( V) = V [ div( V V) + gradp div] dV + V gdV . (1.21)


As this is valid for any arbitrary volume V the intregrand must vanish, i.e.:

( V ) + div [ V V + pI ] = g
(1.22)
t
This is the differential conservative form of the momentum equation,
including a source term due to volume forces. In order to use it for a given
fluid flow the components of the viscous stress tensor must be provided
through a constitutive relation.
When the viscous stresses are identically zero, and the volume forces are
neglected, we obtain the so-called compressible Euler equation. If the
viscous stresses are given by a constitutive relation under the Newtonian
assumption we obtain the compressible Navier-Stokes equation in
differential conservation law form.
1.3.3 Conservation of Energy
As done for mass and momentum we now consider the total energy in a
control volume V, that is:
(t ) = EdV .
(1.23)
V

where E is the total energy per unit volume:


1
1

(1.24)
E = e + V 2 = e + ( u 2 + v 2 + w2 )
2
2

Here e is the specific internal energy and the second term represents the
specific kinetic energy.
Before we discuss the equations in more detail, let us make a physical
comment: the derivation of the energy equation relies upon the assumption
that, in a fluid in motion, the fluctuation around thermodynamic equilibrium
are sufficiently weak so that the classical thermodynamics results hold at

every point and at all times. In particular, the thermodynamical state of the
fluid is determined by the same state variables as in classical
thermodynamics, and these variables are determined by the same state
equations (see the devoted paragraph in this chapter).
The energy conservation law states that the time rate of change of total
energy (t ) is equal to the work done, per unit time, by all forces acting on
the volume plus the influx of energy per unit time into the volume. The
surface and volume forces in (1.17) respectively give rise to the following
terms:
Esurf = p(V n)dA + V (n )dA ,
A

Evol = (V g)dV .
V

The first term corresponds to the work done by the pressure while the
second term corresponds to the work done by the viscous stresses. Evol is
the work done by the volume force g. To account for the influx of energy
into the volume we denote the energy flux vector by Q = (q1 , q2 , q3 ) . The
flow of energy per unit time across a surface dA is given by the flux
(n Q)dA . This gives the total influx of energy to be included in the
equation of balance of energy:
Einf = (n Q)dA ,
A

Thus, one can write that:

D
= Esurf + Evol + Einf
(1.25)
Dt
The left hand side of (1.25) can be transformed into:
D (t )

=
EdV + (n EV )dA
(1.26)
V t
A
Dt
As done for the laws of conservation of mass momentum we now reinterpret
the volume V as fixed in space and independent of time. This leads to:
d
EdV = [n ( EV + pV + Q) V (n ] dA + (V g)dV (1.27)
A
V
dt V
The differential form of the conservation of energy law can now be derived
by assuming sufficient smoothness and applying Gausss theorem to all
surface integrals of (1.27). Direct application of Gausss theorem gives:
n ( pV )dA = div( pV )dV
A

n QdA = divQdV .
A

The second term of (1.27) can be transformed via Gausss theorem by first
observing that V (n ) = n (V ). This follows from the symmetry of the
viscous stress tensor . Hence:
V (n )dA = div(V )dV
A

Substitution of these volume integrals into the integral form of the law of
conservation of energy (1.84) gives:

{E + div ( E + p ) V V + Q (V g)} dV = 0
V

(1.28)

Since the volume V is arbitrary the integrand must vanish identically, that
is:
(1.29)
Et + div ( E + p ) V V + Q (V g ) = 0
This is the differential conservative form of the law of conservation of
energy with a source term accounting for effect of body forces; if these are
neglected we obtain the homogeneous energy equation corresponding to the
Navier-Stokes equation. The energy flux vector Q must be specified through
a constitutive relation.
Further, when viscous and heat conduction effects are neglected we obtain
the energy equation corresponding to the compressible Euler equations.
1.3.4 The Role of the Integral Form of the Equations
The previous derivation of the governing equations, such as the
compressible Euler and Navier-Stokes equations stated earlier, is based on
integral relations on control volume and their boundaries.
The differential form of the equations results from further assumptions on
the flow variables (smoothness). In the case of the Navier-Stokes equations,
the smoothness of the solution is naturally assured by the diffusion due to
the viscosity and heat conductivity. On the contrary, in the absence of
viscous diffusion and heat conduction one obtains the Euler equations.
These admit discontinuous solutions and the smoothness assumption that
leads to the differential form no longer holds true. Thus, one must return to
the more fundamental integral from involving integrals over control
volumes and their boundaries. However, for most of the theoretical
developments the differential form of the equations is preferred.

10

Form a computational point of view there is another good reason for


returning to the integral form of the equations. Discretised domains result
naturally in finite control volumes or computational cells. Local
enforcement of the fundamental equations in these volumes leads to Finite
Volume numerical methods.

1.4 Equations of State


1.4.1 Generalities
The governing equations expressed in integral or differential form are
describing the dynamics of a compressible material. However, these
equations are insufficient to completely describe the physical processes
involved during a flow. There are more unknowns than equations and thus
closure conditions are required. For example, if we take as unknowns the
five conservative variables ( , V, E ) , it is easy to see that there are terms
in the integral or differential form of the governing equations which must be
expressed in terms of the unknowns. Such terms are the pressure and the
specific internal energy. One therefore requires another relation defining e
and p in terms of quantities already given, such as the conservative
variables. For a number of applications, due to the complexity of the
physical phenomena other variables, such as temperature T or entropy s for
instance, may need to be introduced.
The formulation of the closure conditions we have referred to previously is
provided by Thermodynamics under the form of state equations. The
specific internal energy e has an important role in the First Law of
Thermodynamics, while the entropy s is involved with the Second Law of
Thermodynamics. We mention here that the entropy plays an important not
just in establishing the governing equations but also at the level of their
mathematical properties and the designing of numerical methods to solve
them. Details can be found in (Lions, 1996) and (Toro, 1997).

11

1.4.2 Equations of State for Systems in Thermodynamic Equilibrium


A system in thermodynamic equilibrium can be completely described by the
basic thermodynamic variables pressure p and the specific volume . The
specific volume is defined as
1
=
(1.30)

A family of states in thermodynamic equilibrium may be described by a


curve in the p plane each characterized by a particular value of a
variable temperature T. There are physical situations that require additional
variables. Here we are only interested in p T systems. In these, one
can relate the variables via the thermal equation of state:
T = T ( p, )
(1.31)
The p T relationship changes from substance to substance. For
thermally ideal gases one has the simple expression:
p
p
T=
=
(1.32)
R R
where R is a constant which depends on the particular gas under
consideration.
The First Law of Thermodynamics states that for a non-adiabatic system
the change e in internal energy e in a process will be given by e=W
+Q, where W is the work done on the system and Q is the heat
transmitted to the system. Taking the work done as dW = - pdv one may
write:
dQ = de + pd
(1.33)
The internal energy e can also be related to the pressure and specific volume
through a caloric equation of state:
e = e ( p , )
(1.34)
The relations (1.31) and (1.34) are invertible.
For a calorically ideal gas one has the simple expression or equation of
state (EOS):
p
e=
(1.35)
( 1)
where is a constant that depend on the particular gas under consideration.

12

The thermal and caloric equation of state for a given material are closely
related. Both are necessary for a complete description of the
thermodynamics of a system. Choosing a thermal EOS does restrict the
choice of a caloric EOS but does not determinate it. Note that for the Euler
equations one requires a caloric EOS , e.g. p = p ( , e ) unless temperature
T is needed for some other purpose, in which case a thermal EOS needs to
be given explicitly.
1.4.3 Other Thermodynamic Variables and Relations
The entropy s results as follows. We first introduce an integrating factor 1/T
so that that expression
e
e

(1.36)
de + pd = + p d + dp
p
v

becomes an exact differential. Then for the Second Law of


Thermodynamics one introduces the new variable s, called entropy, via the
relation:
Tds = de + pd
(1.37)
For any physical process the change in entropy is due to the entropy carried
into the system through the boundaries of the system, s0 , and to the
entropy generated in the system during the process, si . Examples of
entropy-generating mechanism are heat transfer, and viscosity such as may
operate within the internal structure of shock waves.
The importance of the entropy in fluid dynamics and not only resides from
the Second Law of Thermodynamics. This states that
si 0
(1.38)
in any physical process. If the process is reversible then the entropy does not
change, i.e. si = 0 . An equivalent but more useful form of the equation
(1.38) is the next one:
(1.39)
( s )t + div ( s V ) 0
As long as shocks do not form during the flow, the entropy equation is valid
and if initially s is constant, it remains constant. This is a very strong quality

13

criterion for the evaluation of different numerical schemes developed for the
Euler equations and not only.
Another variable of interest is the specific enthalpy h. This is defined in
terms of other thermodynamic variables, namely:
h = e + p
(1.40)
One can also establish other relationship amongst the basic thermodynamic
variable already defined. For instance one may choose to express the
internal energy e in terms of the variables appearing in the differentials, i.e.
e = e( s, v).
Also, taking the differential of (1.40) we have dh = de + pdv + vdp, which
becomes dh = Tds + vdp, and thus we can choose to define h in terms of s
and p, i.e. h = h( s, p ).
We call the last two relations canonical equations of state and, unlike the
thermal and caloric equations of states (1.31) and (1.34), each of these
provides a complete description of the Thermodynamics.
Two more quantities can be defined if a thermal EOS = ( p, T ) . These
are the volume expansivity (or expansion coefficient) and the isothermal
conpressibility , namely:
1 v
=
v T p
and
1 v
= .
v p T
The heat capacity at constant pressure cp and the heat capacity at
constant volume cv (specific heat capacities) are now introduced. In
general, when an addition of heat dQ changes the temperature by dT the
ratio c = dQ / dT is called the heat capacity of the system. For a
thermodynamic process at constant pressure one obtains:
dQ = de + d ( p, v) = dh,

14

where the definition of the specific enthalpy has been used. Assuming
h = h(T , p ) , the heat capacity cp at constant pressure becomes:
h
s
cp =
(1.41)
=T

T p
T p
The heat capacity cv at constant volume may be written, following a similar
argument, as:
e
s
cv =
(1.42)
=T

T v
T v
The speed of sound is another variable of fundamental interest. For flows
in which particles undergo unconstrained thermodynamic equilibrium one
defines a new state variable a, called the equilibrium speed of sound or
just speed of sound. Given a caloric equation of state such as p = p ( , s ),
one defines the speed of sound a as:

p
a=
(1.43)
s
This basic definition can be transformed in various ways using established
thermodynamic relations. For instance, given a caloric EOS in the form
h = h( p, ), we can also write:

h
h
1 p
p
dp + d = Tds + d + ds
s
s
p p
p
Settings ds=0 and using definition (1.43) we obtain successively:
h

p
2
a =
h
1

p
and for a thermally ideal gas characterized by the EOS of the form h= h(T) :
h

p p
a2 =
1
h

p

15

h
From = c p and for a thermally ideal gas (i.e., if the thermal EOS is
p
acceptable), we obtain the widely used relation for the speed of sound:
P
a = () R =
(1.44)

The reader should notice the dependence on the temperature of the


coefficient .
For a general material the caloric EOS is a functional relationship involving
the variable p--e. The derived expression for the speed of sound a depends
on the choice of dependent variables. Two possible choices and their
respective expression for a are:

p
p = p( , e), a =
p
+
p

(1.45)

e
p
e = e( , p), a =

2e p e p
where subscripts denote partial derivates.
1.4.4 Ideal Gases

We consider in what follows gases obeying the ideal thermal EOS of the
form:
pV = nRT
(1.46)
3
-1
where V is the volume, R=8.134 x 10 J kilomole , called the Universal
Gas Constant, and T is the absolute temperature measured in degrees
Kelvin (K). Recall that a mole of a substance is numerically equal to gram
and contains 6.02x1023 particles of that substance, where is the relative
atomic mass or relative molecular mass; 1 kilomole = kg. One kilomole of
substance contains NA = 6.02 x 1026 particles of that substance. The constant
NA is called the Avogadro Number. Sometimes this number is given in
terms of one mole. The Boltzmann constant k is k = RNA. The number of
kilomoles in volume V is n = N/NA and N is the number of molecules. On
division by the mass m = n have:

p = RT , R =
(1.47)

16

or
p

= RT

(1.48)

where R is called the Gas constant. Equation (1.48) is known as Mariottes


law. Further, we obtain from (1.47) the ideal gas thermal equation of state
as:
RT
v = v(T , p ) =
.
p
The volume expansivity and the isothermal compressibility defined
1
1
previously become = and = , respectively. After some
T
p
manipulations, one deduce that:
e
(1.49)
=0
v T
This means that if the ideal thermal EOS is assumed, then it follows that the
internal energy e is a function of temperature alone, that is e=e(T). A
remarkable particular case is Joules law, that is one in which:
e = cvT
(1.50)
where the specific heat capacity cv is a constant. In this case one speaks of
a calorically ideal gas, or a polytropic gas.
It is now possible to relate cp and cv via the general expression
2T
c p = cv +
. For a thermally ideal gas equations this expression

simplifies to:
c p cv = R

(1.51)

We mention here that, a necessary condition for thermal stability is cv>0 and
a condition for mechanical stability is >0, see (Toro, 1997). From (1.51) the
following inequalities result c p > cv > 0 .
The ratio of specific heats , or adiabatic exponent, is defined as:
c
= p
cv
which if used in conjuntion with (1.51) gives:

(1.52)

17

cp =

R
R
, cv =
.
1
1

(1.53)

In order to determinate the caloric EOS (1.50) we need to determine the


specific heat capacities, cv in particular. Molecular Theory and the principle
of equipartition of energy (Toro, 1997) can also provide an expression for
the specific internal energy of molecule. In general a molecule, however
complex, has M degrees of freedom, of which three are translational. Other
possible degrees of freedom are rotational and vibrational. From Molecular
Theory it can be shown that if the energy associated with any degree of
freedom is a quadratic function in the appropriate variable expressing that
1
degree of freedom, then the mean value of the energy is kT where k is the
2
Boltzmann constant. Moreover, from the principle of equipartition of energy
this is the same for each degree of freedom. Therefore, the mean total
1
energy of a molecule is e = MkT and for N molecule we have
2
1
Ne = MNkT
2
From which the specific internal energy is
1
e = MRT .
2
From the above considerations we obtain successively:
e 1
cv =
= MR,
T v 2
and
M +2
cp =
R.
2
Thus, the ratio of specific heat becomes can be expressed in terms of the
degrees of freedom M:
M +2
=
.
M
And hence:
2
M=
1
From the thermal EOS for ideal gases we have:

18

e=

1
Mpv.
2

And

pv
p
=
(1.54)
( 1) ( 1) p
Thus, we have found once again the expression for the specific internal
energy.
e=

These theoretical expressions for the specific heats and their ratio in terms
of R and M are found to be very accurate for monatomic gases, when M = 3
(three translational degrees of freedom). For polyatomic gases rotational and
vibrational degrees of freedom contribute to M but now the expression
might be rather inaccurate when compared with experimental data. A strong
5
dependence on T is observed. However, the inequality 1 < < , predicted
3
for the limiting values M = 3 and M = , holds true.
1.4.5 Covolume and van der Waal Gases

A very simple generalisation of the ideal-gas thermal EOS is the so-called


covolume equation of state:
p ( b ) = RT
(1.55)
3 -1
where b is called the covolume and in SI units has dimensions of m kg .
This EOS applies to dense gases at high pressure for which the volume
occupied by the molecules themselves is no longer negligible. There is
therefore a reduction in the volume available to molecular motion. This type
of correction to the ideal gas EOS is said to have first been suggest by
Clausius. Sometimes, this equation is also called the Noble-Abel EOS. In
the study of propulsion systems, gaseous combustion products at very high
densities are reasonably well described by the covolume EOS. In its
simplest version the covolume b is a constant and is determined
experimentally or from equilibrium thermochemical calculations. It vas
experimentally observed that for a range of solid propellants the b changes
very little, usually in the range 0.9x10 -3 b 1.1x10 3 . The best values of b
lead to errors of no more than 2% and thus there is some justification in
using (1.55) with b constant.

19

The covolume EOS can be further corrected to account for the forces of
attraction between molecules, the so-called van der Waal forces. These are
neglected in both the ideal and covolume equations of state. Accounting for
such forces results in a reduction of the pressure by an amount of c/v2,
where c is a quantity that depend on the particular gas under consideration.
Thus from (1.49) the pressure is corrected as
RT
c
p=
2.
v b v
Then we can write:
c

(1.56)
p + 2 (v b) = RT
v

This is generally known as the van der Waals equation of state for real
gases.
So far, we have presented the governing equations for the dynamics of a
compressible medium along with some EOS so that a closed system is
obtained.

1.5 Viscous Stresses and Heat Conduction

The stresses in a fluid, given by a tensor , are due to the effects of the
thermodynamic pressure p and the viscous stresses. Thus the stress tensor
can be written as
= pI +
(1.57)
where pI is the spherically tensor due to p, I is the unit tensor and is the
viscous stress tensor. It is desirable to express in terms of flow variables
already defined. For the pressure contribution this has already been achieved
by defining p in terms of other thermodynamic variables via an equation of
state. Recall that equations of state are approximate statements about
the nature of a material. In defining the viscous stress contribution one
may resort to the Newtonian approximation whereby is related to the
derivatives of the velocity field V = ( u, v, w ) via the deformation tensor:

20

1
1

ux
(v x + u y )
( wx + u z )

2
2

1
1
D = (v x + u y )
vy
( wy + v z )
(1.58)
2

2
1

( wx + u z ) 1 ( wy + vz )

wz
2
2

The Newtonian assumption is an idealisation in which the relationship


between and D is linear and homogeneous, that is will vanish only if
D vanishes, and the medium is isotropic with respect to this relation. An
isotropic medium is that in which there are no preferred directions. By
denoting the stress tensor by:
xx xy xz

(1.59)
= yx yy yz
zx
zy
zz

the Newtonian approximation becomes:


2

(1.60)
= 2 D + b (divV ) I
3

In full we have:
4
2

xx = u x ( v y + wx ) + b divV ,
3
3

4
2
yy
= v y ( wz + u x ) + b divV ,

3
3

4
2
(1.61)
zz = wz ( ux + v y ) + b divV ,
3
3

xy = yx = (u y + vx ),

yz = zy = (vz + wy ),

zx = xz = ( wx + u z ).

In the Newtonian relationship (1.60) there are two scalar quantities that are
still undetermined. These are the coefficient of shear viscosity and the
coefficient of bulk viscosity b. Approximate expressions for there are
obtained mainly from experiments.

21

In particular, for monatomic gases Molecular Theory gives for the


coefficient of bulk viscosity b = 0, which is found to agree well with
experiment. For polyatomic gases b 0 and appropriate values for b are
to be obtained experimentally.
Concerning the coefficient of shear viscosity , it is observed that, as long as
temperature are not too high, depends strongly on temperature and only
slightly on pressure. A relatively accurate relation between and T is the
Sutherland formula:
1

C2
(1.62)
T
T
where C1 and C2 are two experimentally adjustable constants. When T is
measured in Kelvin degrees, for case of air one has:
C1 = 1.46 x10 6 , C2 = 112 K .
Sutherlands formula describes the dependence of on T rather well for a
wide range of temperatures, provided no dissociation or ionisation take
place. These phenomena occur at very high temperatures where the
dependence of on pressure p, in addition to temperature T, cannot be
neglected.

= C1 1 +

In summary, the Navier-Stokes equations (momentum equation) can now be


written in differential conservation law form as:
(1.63)
( V )t + ( V V + pI ) = 0
where is given by (for monatomic gases):
2
= 2 D (divV ) I
(1.64)
3
The Euler equations written in differential conservation form are:
(1.65)
( V )t + ( V V + pI ) = 0
Influx of energy contributes to the rate of change of total energy E. We
denoted by Q = ( q1 , q2 , q3 ) the energy flux vector, which results from: i)
heat flow due to temperature gradients, ii) diffusion processes in gas
mixture and (iii) radiation.
T

In what follows we only consider the effect of the heat flux due to
temperature gradients. Thus, Q is identical to the heat flux vector caused by

22

temperature gradients. In a similar manner to that it which viscous stresses


were related to gradient of the velocity vector V, one can assume that the
fluid is isotropic and thus one can relate Q to gradients of temperature T via
Fouriers heat conduction law:
Q = k T
(1.66)
where k is a positive scalar quantity already called the coefficient of
thermal conductivity, and is yet to be determined.
Note the analogy between the two microscopic transport properties of the
fluid, and k. This analogy between and k goes further in that k, just as ,
depends on T but only slightly on pressure p.
In fact, Molecular Theory says that k is directly proportional to . Under the
assumption that the specific heat at constant pressure cp is constant, the
dimensionless quantity called the Prandtl number:
c p
Pr =
(1.67)
k
is a constant. For monatomic gases Pr is very nearly constant. For air in the
temperature range 200 K T 100 K Pr differs only slightly from its mean
value of 0.7.
There is also a formula attributed to Eucken, see (Toro, 1997), that relates Pr
to the ratio of specific heats via
4
Pr =
9 5
to account for departures from calorically ideal gas behaviour.

1.6 Differential Form of the Governing Equations

We have determined previously the integral form of the governing


equations, as well as the conservative differential form of the equations.
Here we summarize the governing equations written in conservative
differential form. We notice that the name Euler and/or Navier-Stokes
originally given to the momentum equations (or equations of motion)
transfers to the entire system of equations.
We summarize below the general laws of conservation of mass, momentum
and total energy, written both in differential and integral form:

23

a) the continuity equation:

t + div( V ) = 0 or
d
dV + n ( V )dA = 0
A
dt V
b) the momentum eqution:
( V )t + div [ V V + pI ] = g, or

( V ) + [ div( V V ) + gradp div ] dV = gdV .


V
V
t
c) the energy equation
Et + div [ ( E + p)V V + Q ] = (V g). or

d
EdV + [n ( EV + pV + Q) V (n ] dA = (V g )dV
A
V
dt V
where g = ( g1 , g 2 , g 3 ) is a body force vector.
At this point, all the terms, such as the viscous stress tensor, the heat flux,
the pressure, the temperature and the specific internal energy are expressed
as functions of the conservative variables. For practical reasons, it is useful
to write the previous equations in a more compact form, which is called
conservation law form.
1.6.1 The Euler Equations

In this section we consider the time-dependent Euler equations. These are a


system of non-linear hyperbolic conservation laws that govern the dynamics
of a compressible material, such as gases or liquids at high pressure.
When body forces are included via a source term vector but viscous and
heat conduction effects are neglected we have the Euler equations:
U t + Fx (U ) + G y (U ) + H z (U) = S(U )
(1.68)
where the vector of conservative variables is U and the convective fluxes
are F,G,H in the x, y and z directions. Their explicit forms are:

24


u
u
u 2 + p


U = v ,
F = uv ,


w
puw

E
u ( E + p)

w
v
uw
uv

G = v 2 + p , H = vw
2

w + p
vw
w( E + p).
v( E + p)

(1.69)

First, it is important to note that the fluxes are nonlinear functions of the
conserved variable vector. Any set of partial differential equations written in
the form (1.68) is called a system of conservation law (in differential
formulation). Recall that the differential form assumes smooth solutions,
that is, partial derivatives are assumed to exist. It is also clear now that the
integral form of the equations is an alternative way of expressing the
conservation laws in which the smoothness assumption is relaxed so that to
include discontinuous solutions.
Here S = S(U ) is a source or forcing term. Due to the presence of the
source term, the equations are said to be inhomogeneous. There are several
physical effects that can be included in the forcing term: body forces such as
gravity, injection of mass, momentum and/or energy. Usually, S(U) is a
prescribed algebraic function of the flow variables and does not involve
derivatives of these, but there are exceptions. When S(U ) = 0 one speaks of
homogeneous equations. We also mention here that there are situations in
which source terms arise as a consequence of approximating the
homogeneous equations to model situations with particular geometric
features (axy-symmetric flows, for instance). In this case the source term is
of geometric character, but we shall still call it a source term.
Sometimes it is convenient to express the equations in term of the primitive
or physical variables , u, v, w and p. By expanding derivatives in the
conservation law form and using the mass equation into the momentum

25

equations and in turn using these into the energy equation one can re-write
the thee-dimensional Euler equations for ideal gases with a body-force
source term as:
t + u x + v y + w z + (ux + v y + wz ) = 0

ut + uu x + vu y + wu z +
vt + uvx + vv y + wvz +

wt + uwx + vwy + wwz +

px = g1 ,
py = g2 ,

(1.70)

pz = g3 ,

pt + upx + vp y + wpz + p (u x + v y + wz ) = 0
For computational purposes it is the conservation law form (1.68) that is
most useful. The formulation in primitive variables is more convenient to
theoretical developments.
1.6.2 The Navier-Stokes Equations

When the effects of viscosity and heat conduction are added to the basic
Euler equations one has the Navier-Stokes equations with heat conduction
in conservation law form:
U t + Fxc + G cy + H cz = Fxd + G dy + H dz
(1.71)
where U is once again the vector of conserved variables, the flux vectors Fc,
Gc and Hc are the inviscid fluxes (c stand for convection) for the Euler
equations as given previously and the respective flux vectors Fd, Gd and Hd
(d stands for diffusion) due to viscosity and heat conduction are:

26

0
0

yx
xx

, Gd =
'
yy
Fd =
xy

yz

xz

u xx + v xy + w xz q
u yx + v yy + w yz q

2
(1.72)
0

zx

zy
d

H =

zz

u zx + v zy + w zz q
3

The form of the equations given by (1.71) split the effect of convection on
the left-hand side from those of viscous diffusion and heat conduction on
the right-hand side. For numerical purpose, the particular form of the
equations adopted depends largely on the numerical technique to be used to
solve the equations. One possible approach is to split the convection effects
from those of viscous diffusion and heat conduction during a small time
interval t, in which case the above from is perfectly adequate. An
alternative form is obtained by combining the fluxes due to convection,
viscous diffusion and heat conduction into new fluxes so that the governing
equations look formally like a homogeneous system (zero right-hand side)
of conservation laws:
U t + Fx + G y + H z = 0,
with

(1.73)

F = Fc Fd , G = G c G d , H = Hc Hd
This form is only justified if the numerical method employed actually
exploits the coupling of convection, viscosity and heat conduction when
defining numerical approximations to the flux vectors F, G and H in (1.73).

27

1.7 Conclusion
We have presented in this chapter the time-dependent Euler and NavierStokes equations of fluid dynamics. The equations are accompanied by
equations of state and by constitutive relations. Both models can be used for
homogeneous gases and/or liquids at high pressure. We appreciate that for
usual purposes the Euler and Navier-Stokes are the representative models.
We now conclude this chapter with a brief discussion of initial and
boundary conditions. Our goal in this book is to present numerical schemes
that can be used to solve the Cauchy problems for the models derived and
described above. In other words, we solve the systems of equations for t 0
prescribing the values of all the unknowns at t = 0 . The question of the
boundary condition is much more delicate. Simply said, one impose
physical boundary condition at the solid boundaries for some of the
variables, and these can be of Dirichlet or Neumann type (depending on the
variable type). Since the boundary conditions are replacing the governing
equations on the boundaries, the rest of the variables are determined via
numerical techniques. These are strongly related to the numerical solvers
and thus we will discuss the imposition of the boundary conditions in the
context of numerical solvers.
A short comment on the Navier-Stokes model is necessary. The equations
presented in this chapter can be used for the numerical simulation of laminar
flows only. At practical Reynolds numbers (see the introduction), the effects
of the turbulence must be taken into account. The Reynolds Averaged
Navier-Stokes (RANS) equations and the Large Eddy Simulation (LES)
equations have formally the same mathematical aspect, with minor
differences. The huge difference comes from the necessity of modeling the
turbulence effects, mainly in the vicinity of solid boundaries. As we already
mentioned, it is beyond the purposes of this book to discuss about
turbulence modeling and turbulence models.
More details about the derivation of these and other equations can be found
in (Anderson et. all, 1984), (Danaila and Berbente, 2003), (Hirsch, 1990),
(Warsi, 1999). A very rigorous discussion of the models and their solutions
can be found in (Lions, 1996).

28

Two remarks we wish to make before at the end of this first chapter. The
first is that the Navier-Stokes and Euler equations are the cornerstones for
the development of practical CFD codes. From a mathematical point of
view, these models are systems of partial differential equations. Since with a
few exceptions we cannot find exact solutions of these equations for
practical problems, the only way is to solve them is to use numerical
techniques. It is generally accepted that the discretization techniques must
be based not only on the underlying physics bust also on the mathematical
properties of the partial differential equations. It is therefore useful to begin
with the analysis of the mathematical nature of the governing equations of
the flow before trying to solve them numerically.
The second remark is that the equations of compressible fluid flow reduce to
hyperbolic conservation laws (i.e. the Euler system) when the effects of
viscosity and heat conduction are neglected. Furthermore, the hyperbolic
part represents the convection and pressure gradient effects and it can be
identified in the equations even the above physical effects are not negligible.
This is the reason why we allocate a special chapter to study the hyperbolic
partial differential equations.

29

2. Approximated and Simplified Models

2.1 Generalities
In this chapter we wish to consider successively simplified version, or
submodels, of the governing equations and their closure conditions. There
are many reasons and many ways to use and derive simplified models. On
one hand, it is clear that any simplification leads to the reduction of the
generality of the equations. On the other hand, a rational simplification may
lead to a significant reduction of the computational effort without penalties
on the quality of the solution. There are some possibilities that can be
exploited to derive simplified models, and perhaps the most common of
them are the reduction of the dimensionality of the problem and the
incompressibility assumption.
One can also augment the basic equations by source terms to account for
additional physics. This is not exactly a simplification of the basic model;
on the contrary, adding some semi-empirical source terms may enlarge the
area of practical applications that can be solved with that model.
In the previous chapter we already done a significant simplification of the
Navier-Stokes equations. We have neglected the viscosity and the heat
conduction effects and we have arrived to the Euler equations. In this
chapter we discuss further simplifications of both the Euler and NavierStokes equations.
Compressible submodels will include flows with variation; flows with axial
symmetry; flows with cylindrical and spherical symmetry; plane onedimensional flow and further simplifications of this to include linearised and
scalar submodels. Incompressible submodels will include various
formulations of the incompressible Navier-Stokes equations.

30

2.2 Compressible Submodels


2.2.1 Flow with Area Variation
Flows with area variation arise naturally in the study of fluid flow
phenomena in ducts, pipes, shock tubes and nozzles. One may start from the
two dimensional homogeneous version of Euler equations to produce, under
the assumption of smooth area variations, a quasi-two dimensional system
with a geometric source term S(U) , namely:
U t + Fx (U) = S(U)
(2.1)
where

u
uAx At
1

2
U = u , F = u + p , S = pu 2 Ax
(2.2)
A
E
u ( E + p)
u ( E + p ) Ax
Here x denotes distance along the tube, nozzle, etc.; A is the cross-sectional
area and in general is a function of both space and time, that is A = A( x, t ) .
The most common case is that in which A depends on x only, for which one
cam write governing equations in the more convenient form:
U t + Fx (U) = S (U)
(2.3)
where
A
A u
0

2
U = A u , F = A( u + p) , S = pAx
AE
Au ( E + p)
0
This is a convenient form in that by defining area weighted values for and
p, equation (2.3) may be interpreted as the usual one-dimensional equations
of Gas Dynamics plus a simple source term S.
2.2.2 Axi-Symmetric Flows
Here we consider domains that are symmetric around a coordinate direction.
We choose this coordinate to be the z-axis and is called the axial direction.
The second coordinate is r, which measures distance from the axis of
symmetry z and is called the radial direction. There are two component of
velocity, namely u (r , z ) and v(r , z ) . These are respectively the radial (r)
and axial (z) components of velocity. Then the three dimensional

31

(inhomogeneous) conservation laws are approximated by a two dimensional


problem with geometric source terms S (U ) , namely:
U t + Fr (U) + G z (U) = S(U)
(2.4)
where
u
v

u
2

uv
u

2
u + p
1 u

,F =
,G =
,S =
(2.5)
U=
uv
u 2 + p
v
r uv

E
u ( E + p )
u ( E + p )
v( E + p)
2.2.3 Cylindrical and Spherical Symmetry
Cylindrical and spherical symmetric wave motion arises naturally in the
theory of explosion waves in water, air and other media. In these situations
the multidimensional equation may be reduced to essentially onedimensional equations with a geometrical source term vector S(U) to
account for the second and third spatial dimensions. We write:
U t + Fr (U) = S(U)
(2.6)
where

u
u

2
U = u , F = u + p , S = u 2
(2.7)
r
E
u ( E + p)
u ( E + p)
Here r is the radial distance from the origin and u is the radial velocity.
When = 0 we have plane one-dimensional flow; when = 1 we have
cylindrically symmetric flow, an approximation to two-dimensional flow.
This is a special case of the axy-symmetric equations when no axial
variations are present (v = 0) . For = 2 we have spherically symmetric
flow, an approximation to three-dimensional flow. Approximations of this
kind can easily be solved numerically to a high degree of accuracy by a
good one-dimensional numerical method. These accurate one-dimensional
solutions can then be very useful in partially validating two and three
dimensional numerical solutions of the full models.

32

2.2.4 Plain One-Dimensional Flow


We first consider the one-dimensional time depend case:
U t + F(U) x = 0
where

U = u , F = u 2 + p
E
u ( E + p)

(2.8)

These equations also result from the previous equations, for example from
(2.3). They are useful for solving shock-tube type problems. Further, under
suitable physical assumptions they produce even simpler mathematical
models. In all the submodels studied so far we have assumed some
thermodynamic closure condition given by an Equation of State (EOS).
The isentropic equations result under the assumption that the entropy s is
constant everywhere, which is a simplification of the thermodynamics. Now
the EOS becomes p = p ( ) C , where C = constant. This makes the
energy equation redundant and we have 2x2 system:
U t + F(U) x = 0,
(2.9)

u
U = ,F = 2

u
u + p
with the pressure p given by the above simple EOS. Notice please that even
in the isentropic assumption, the equation are still non-linear.
The isothermal equations are even a simpler model then the isentropic
equations, still non-linear. These may be viewed as resulting from the
isentropic equations (2.9) with a simpler EOS, that is p = p ( ) a 2 ,
where a is a non-zero constant and represents the propagation speed of
sound. Thus, the isothermal equations are:
U t + Fx (U) = 0,

U = ,F = 2
2
u
u + a

(2.10)

More submodels may be further obtained by writing the isentropic equations


as:

33

t + u x + u x = 0

ut + uu x = px

(2.11)

The inviscid Burgers equation is a scalar (single equation) non linear


equation given by:
ut + uu x = 0
(2.12)
and can be obtained from the above momentum equation (2.11) by neglecting
density and thus pressure variations. In conservative form equation (1.115)
reads:
u2
ut + = 0
(2.13)
2 x
The Linearised Equations of Gas Dynamics are obtain from (2.11) by
considering small disturbances u , to a motionless gas. Set u = u and
= 0 + , where 0 is a constant density value. Recall that p = p( ) and
neglecting products of small quantities we have:
p
( 0 )
p = p ( 0 ) +

p
that is, p = p0 + a 2 with p 0 = p ( 0 ) , and a 2 =
( 0 ) = constant.

Substituting into (2.11) and neglecting squares of small quantities we obtain


the linear equations:
t + 0u x = 0

(2.14)
a2

+
u
t x = 0
0

In matrix form system (1.118)-(1.119) reads:


Wt + AWx = 0,

0
0
(2.15)

W = u , A = a2
0

0

where bars have been dropped. The coefficient matrix A is now constant
and thus the system (2.15) is a linear system with constant coefficients, the
linearised equations of gas dynamics.

34

The linear advection equation, sometimes called linear convection,


equation is:
ut + au x = 0
(2.16)
where a is a constant speed of wave propagation. This is also known as the
one-way wave equation and plays a major role in the designing, analysing
and testing of numerical methods for wave propagation problems.
2.2.5 Steady Flow
The steady, or time-independent, homogeneous version of the threedimensional Euler equations are:
Fx + G y + H z = 0
(2.17)
In the steady regime it is important to identify subsonic and supersonic flow
regions. To this end we recall the definition of Mach number M:

u 2 + v 2 + w2 2
M =

a2

where a is the speed of sound. Supersonic flow requires M>1, while for
subsonic flow we have M<1. For sonic flow M=1. The basic approach
therefore relies on the two-dimensional case. In the two-dimensional case
the differential conservation form of the steady equations are:
Fx (U) + G y (U) = 0
(2.18)
with
u
v
u 2 + p
uv

F=
,G = 2
(2.19)
uv
v + p

u ( E + p )
v( E + p)
As discontinuous solutions such as shock waves and slip surfaces are to be
admitted, one can replace the differential form by the more general integral
conservation form:
(2.20)
 ( Fdy Gdx) = 0
In (2.20) the integral is to be evaluated over the boundary of the appropriate
control volume. In numerical methods this will be a computation cell.

35

Steady linearised models can be further obtained from the steady Euler
equations. An interesting submodel is the small perturbation, twodimensional steady supersonic system of equations:
2
u x + a v y = 0
(2.21)

vx u x = 0
with
1
a2 = 2
M 1
and where M denotes the constant free-stream Mach number and
u ( x, y ), v( x, y ) are small perturbations of the x and y velocity components.
In matrix form these equations read
Wx + AWy = 0
(2.22)
with
0 a 2
u
W = ,A =
.
1
0

2.2.6 Viscous Scalar Equations


From the viscous compressible Navier-Stokes equations one can deduce,
after some manipulations, two representative scalar equations. The first is
the viscous Burgers equation, which is the viscous version of (2.13). It
reads:
u2
ut + = u xx
(2.23)
2 x
where is a coefficient of viscosity. A linearised form of this is the linear
advection-diffusion equation:
ut + au x = u xx
(2.24)
wich is the viscous version of (2.16).

36

2.3 Incompressible Submodels


2.3.1 About Incompressible Models and Low Mach Number Expansions
In the field of CFD, the incompressibility assumption is very important for
applications since many common fluids (liquids) are incompressible or only
very slightly compressible. Mathematically, the incompressibility
condition means:
divV = 0
(2.25)
Therefore, the volume occupied by a group of fluid particles at the initial
time remains constant during the flow. The continuity equation written as:
t + u x + v y + w z + (ux + v y + wz ) = 0
leads to
D
= t + V = 0
(2.26)
Dt
This means that if the density is constant initially and on the boundaries
from where the fluid comes inside the domain under consideration it
remains so. This is equivalent to say that the fluid is homogeneous.
Further, having in mind the previous derivation of the compressible models
and their EOS and constitutive models, it could be useful to make a
distinction between models obtained by:
1. Incompressibility hypothesis.
2. Low Mach number expansions.
We emphasize here that the incompressibility hypothesis does not impose
an explicit restriction on the magnitude of the velocity. Moreover, the
incompressible models aim to describe liquids where compression effects
are neglected and the density is taken as constant. For example, it is not
rational to expect from an incompressible model to describe accurately the
propagation of acoustic or pressure waves through liquids. This is due the
fact that the incompressibility condition (2.25) is normally associated with
other working hypothesis made on the EOS and on the fluid transport
properties. Further, the energy equation is firmly decoupled from the
continuity and momentum equations, by stating that the temperature field
can be calculated separately, after the velocity and pressure fields have been
determined.

37

On the other side, the compressible models presented in the previous


chapter are valid for gases. Starting from the compressible models it is
rational to discuss about the low Mach number expansions. The precise
p ' ( ) . Therefore, letting M
definition of the Mach number is: M = V
go to zero means that, keeping constant values for the density and the
temperature, the magnitude of the velocity is a small parameter. An
asymptotic analysis starting from the Navier-Stokes equation derived for a
compressible ideal gas shows that there is a lack of consistency between
compressible models and incompressible submodels, see (Lions, 1996), in
the presence of heat conduction. A possible physical explanation is the
following assertion: compressible models are valid for gases and the low
Mach number limit yields particular incompressible submodels. These
particular submodels are definitively determined by the EOS and transport
properties chosen for the gas. Further, such an asymptotic analysis reveals
that the incompressible submodel is very sensitive to the errors in the
pressure calculation.
In what follows, we assume the fluid to be incompressible, homogeneous,
non-heat conducting and viscous, with constant coefficient of viscosity .
Body forces are also neglected, only for simplicity. We study three
mathematical formulations of the governing equations in Cartesian
coordinates and restrict our attention to the two-dimensional case.

2.3.2 The Incompressible Navier-Stokes Equations in Primitive


Variable Form
The primitive variable formulation of the incompressible two dimensional
Navier-Stokes equations is given by:
ux + vy = 0

ut + uu x + vu y +
vt + uvx + vv y +

p x = v u xx + u yy

(2.27)

p y = v vxx + v yy

where the kinematic viscosity is:


v=

(2.28)

38

Recall that is the coefficient of shear viscosity. We have a set of three


equations for the three unknowns u, v, p, the primitive variables. This is a
mixed elliptic-parabolic system. Due to the mixed nature of the
mathematical model, the solution cannot be obtained directly via timemarching algorithms. In principal, given a domain along with initial and
boundary conditions for the equations one should be able to solve this
problem using the primitive variable formulation.
2.3.3 The Incompressible Navier-Stokes Equations in Stream-Function
Vorticity Form

The stream-function vorticity formulation is another way of expressing the


incompressible Navier-Stokes equations. This formulation is attractive for
the two-dimensional case but not so much in three dimensions, in which the
role of a stream function is replaced by that of a vector potential. The
magnitude of the vorticity vector can be written as:
= vx u y
(2.29)
Introducing a stream function we have for the velocity components:
u = y , v = x .
By combining the momentum equations so as to eliminate the pressure p,
and using (2.29) we obtain the vorticity transport equation:
t + u x + v y = v xx + yy
(2.30)
This is an advection-diffusion equation of parabolic type. In order to
solve it, one requires the solution for the stream function , which is in turn
related to the vorticity via:
xx + yy =
(2.31)
This is called the Poisson equation and is of elliptic type. There are
numerical schemes to solve (2.30)-(2.31) using the apparent decoupling of the
parabolic-elliptic problem to transform it into the parabolic equation vor the
vorticity and the elliptic equation for the stream function.. A relevant
observation, from the numerical point of view, is that the convection terms
of the left hand side of equation (2.30) can be written in conservative form
and hence we have:
t + ( u ) x + ( v ) y = v xx + yy
(2.32)
This follows from the fact that u x + v y = 0 , which was also used to obtain.

39

2.3.4 The Incompressible Navier-Stokes Equations in Artificial


Compressibility Form

The artificial compressibility formulation is yet another approach to


formulate the incompressible Navier-Stokes equations and was originally
put forward by Chorin, see (Chorin, 1968) for the steady case. Let us
consider the two-dimensional incompressible Navier-Stokes equations
written in non-dimensional form:
ux + vy = 0
(2.33)
ut + uu x + vu y + px = u xx + u yy
vt + uvx + vv y + p y = vxx + v yy

(2.34)

where the following non-dimensionalisation has been used:


u
v
p
u
,v
,p
,
V
V
V2

tV
x
y
, y 't '
L
L
L
V L
1
=
, ReL =
.
ReL
V
x

Multiplying (2.33) by the non-zero parameter c 2 and adding an artificial


compressibility term pt the first equations reads:
pt + ( uc 2 ) + ( vc 2 ) = 0
x

(2.35)

By using equation (2.33) the convective terms in the momentum equations


can be written in conservative form, so that the modified system becomes:
pt + ( uc 2 ) + ( vc 2 ) = 0
x

ut + (u 2 + p ) x + (uv) y = u xx + u yy

(2.36)

vt + (uv) x + (v 2 + p) x = vxx + v yy
The equations can be written in compact form as
U t + Fx (U) + G y (U) = S(U)
where the vectors of unknowns, fluxes and source terms are:

(2.37)

40

c 2u
c 2v

0
p

(2.38)
U = u , F = u + p , G = uv , S = (u xx + u yy )

(vxx + v yy )
v
uv
v + p

The above equations are called the artificial compressibility equations. Here
c2 is the artificial compressibility factor, usually taken as a constant
parameter. The source term vector in this case is a function of second
derivatives. Note that the modified equations are equivalent to the original
equations in the steady state limit only. The left-hand side of the artificial
compressibility equations form a non-linear hyperbolic system.
More recently, new formulations have been proposed for the solution of
steady and unsteady incompressible Navier-Stokes equations. Since timemarching methods cannot be applied directly, the system (2.36) must be
transformed into a more convenient one. The dual time approach requires
the addition of derivatives of a fictitious pseudo-time to each of the three
equations to give:
1
p + pt + ( u ) x + ( v ) y = 0
2

u + ut + (u 2 + p) x + (uv) y = u xx + u yy

(2.39)

v + vt + (uv) x + (v 2 + p) x = vxx + v yy
where is a parameter and the term added to the continuity equation has
the same form as the basic artificial compressibility method. A steady-state
solution in pseudo-time ( p , u , v 0 ) corresponds to an
instantaneous unsteady solution in real time. A recommended value for the
parameter in the case the governing equations are written in
dimensionless form is (1) . The convective part of the system (2.39) is
of hyperbolic type and therefore a time-marching solution procedure is
possible, (Gaitonde, 1998).

You might also like