You are on page 1of 7

Department of Mechanical & Aerospace Engineering

CARLETON UNIVERSITY

AERO 4304: Computational Fluid Dynamics


Winter 2013

Lecture 2: Review of Conservation Laws


Lecture summary
In this lecture, we will review the conservation principles of fluid dynamics, including
Conservation applied to a control volume (Reynolds transport theorem)
Mass conservation
Momentum conservation
Energy conservation
Simplified models of fluid flow

Conservation applied to a control volume

Conservation laws can be derived by considering a control mass (CM) and monitoring
the variation of its extensive properties, such as mass or momentum. Because mass
cannot be created or destroyed in flows of engineering interest, mass conservation can
be written as
dm
=0
(1)
dt
where t stands for time and m is the mass of the CM. Momentum can be changed
by the action of external forces, and its conservation is governed by Newtons second
law:
d(mu) X
=
f
(2)
dt
where u is the velocity vector, mu is the momentum, and f are the forces acting on
the control mass.
While looking at a control mass may be convenient in certain fields, in fluids, it
is difficult to follow a parcel matter, and so we instead prefer to analyze the flow
in a certain region in space, called a control volume (CV). Within a CV, instead
of considering the conservation of mass, we consider the conservation of mass per
unit volume, or density , and instead of considering the conservation of momentum,
we consider the conservation of momentum per unit mass, or velocity u. These
properties, which are independent of the amount of matter under consideration, are

AERO 4304

Lecture 2

called intensive properties. For any intensive quantity , the corresponding extensive
property can be obtained by integrating over the volume occupied by the whole
control mass:
Z
=
d
(3)
CM

where CM is the volume occupied by the control mass. Using this definition, the
left-hand side of the conservation laws expressed in Eqns. 1 and 2 can be written for
the control volume as:
Z
Z
d
d
=
d +
(u ub ) n dS
(4)
dt
dt
CV

SCV

where CV is the CV volume, SCV is the surface enclosing the CV, n is the unit
vector orthogonal to SCV and directed outwards, u is the fluid velocity and ub is the
velocity of the CV surface. and is the fluid density. The integration means that
we account for all of the that occurs within the control volume. This equation
states that the rate of change of the amount of the property in the control mass,
, equals the rate of change of the property within the CV plus the net flux of it
through the CV boundary due to fluid motion relative to the CV boundary. The
last term is frequently referred to as the convective (or advective) flux of through
the CV boundary, or simply the convection term. Equation 4 is called the Reynolds
transport theorem, and it is a general conservation equation for an intensive quantity,
from which the mass, momentum, and energy conservation equations can be derived.

Mass conservation

The integral form of the mass conservation law in Eqn. 1 follows directly from Eqn.
4 by setting = 1:
Z
Z

d + (u n)dS
(5)
t

For simplicity, the CV is assumed to be stationary, and so the ub term is omitted. By


applying Gauss divergence theorem to the convection term in Eqn. 5, we can convert
the surface integral into a volume integral. Allowing the volume of the CV to tend
towards zero produces the differential form of the mass conservation (or continuity)
equation:

+ div(u) = 0.
(6)
t

Momentum conservation

The momentum conservation equation (or just momentum equation) can be derived
from Reynolds transport theorem (Eqn. 4) by setting = u. Newtons second law

AERO 4304

Lecture 2

expressed in Eqn. 2 becomes


Z
Z
X

f
u d + u(u n) dS =
t

(7)

To express the right hand side in terms of intensive properties, one has to consider
the forces that may act on the fluid in a CV. In general, these forces fall into two
categories:
1. Surface forces (pressure, normal stresses, shear stresses, surface tension, etc.)
2. Body forces (gravity, centrifugal and Coriolis forces, electromagnetic forces, etc.)
Lets start with the surface forces. We restrict ourselves to Newtonian fluids, where
the stress tensor T is defined as


2
T = p + div u I + 2D
(8)
3
where is the dynamic viscosity, I is the unit tensor, p is the static pressure, and
D is the rate-of-strain (deformation) tensor. These are easiest to express in index
notation in Cartesian coordinates:


2 uj
Tij = p +
ij + 2Dij
(9)
3 xj


1 ui uj
Dij =
+
(10)
2 xj
xi
where ij is the Kronecker symbol (ij = 1 if i = j and 0 otherwise). The viscous
forces and pressure forces of the stress tensor are frequently separated:
Tij = ij p

(11)

where the viscous part of the stress tensor ij is defined as


2
ij = 2Dij ij div u.
3

(12)

From the continuity equation (Eqn. 6), it is easy to see that the second term on the
right hand side of ij is zero when the flow is incompressible.
With the body forces per unit mass being represented by b, the integral form of
the momentum conservation equation becomes:
Z
Z
Z
Z

u d + u(u n) dS = T n dS + b d
(13)
t

AERO 4304

Lecture 2

The surface integral terms on the left and right hand sides of Eqn. 13 respectively
represent the convective and diffusive fluxes of momentum through the CV boundary.
As with the continuity equation, the differential form of the momentum equation is
obtained by applying Gauss theorem to the convective and diffusive flux terms and
allowing the volume of the CV to tend to zero. The resulting differential momentum
equation is
ui
+ div(ui u) = div Tij + bi
(14)
t
The differential form of the mass and momentum equations (Eqns. 6 and 14) are
collectively referred to as the Navier-Stokes equations.

Energy conservation

In an analogous manner to the continuity and momentum equations, the energy


conservation equation (or just energy equation) is obtained by applying the Reynolds
transport theorem with equal to the specific energy of the fluid, E:
Z
Z
Z
Z

E d +
E(u n)dS =
p(u n) dS +
S grad(u)d
t
CV
SCV
S

Z CV
ZCV
+
kgrad(T )dS +
qE d
(15)
SCV

CV

Here, the specific energy E = e + 0.5uii incorporates the internal and kinetic energy
components, while the potential energy due to gravity is included in the energy source
term qE . The T term in Eqn. 15 is the temperature, k is the thermal conductivity
(k = cp /Pr), cp is the specific heat, Pr is the Prandtl number, S is the viscous part
of the stress tensor (S = T + pI). The first two terms on the right hand side of
Eqn. 15 represent the net rate of work down on the CV by normal and shear stresses,
the third term represents the net rate of heat addition to the CV, and the fourth
term represents the energy added due to sources. As above, application of Gauss
divergence theorem and shrinking the CV volume to zero produces the differential
form of the energy equation:
E
+ div(Eu) = div(pu) + S grad(u) + div(k gradT ) + qE
t

(16)

Conservation form of the governing equations

The continuity, momentum, and energy equations derived in the above sections are all
written in strong conservation form. This means that the coefficients multiplying the
derivatives of the PDE are either constant or their derivatives do not appear in the

AERO 4304

Lecture 2

equation. As an illustration, writing the conservation form of the continuity equation


(Eqn. 6) in terms of Cartesian velocity components u, v, w yields
u v w
+
+
+
= 0.
t
x
y
z

(17)

Notice that density gradient does not appear as a coefficient in the continuity equation
in the above form. The same equation can be written in non-conservation form as

+
+u
+
+v
+
+w
=0
t
x
x
y
y
z
z

(18)

Now, in non-conservation form, the density gradient appears as a coefficient to the


velocity derivatives. Although the conservation and non-conservation form are mathematically equivalent, discretization of the non-conservation forms may produce difficulties in situations where the coefficients may be discontinuous, as in flows containing
shock waves. The conservation form have additional benefits in the context of finitevolume methods (discussed in Lectures 16-19), where the use of the conservation
form automatically ensures global momentum conservation in the whole simulation
domain. Essentially, that means that it ensures that the total mass flux in and out
of the whole simulation domain will add to exactly zero. Retaining this property
improves the stability of the calculation.

Simplified models of fluid flow

The conservation equations presented in the above sections assume that all fluid and
flow properties vary in space and time, which makes them the most general, but also
increase the complexity of the problem. In many flows, simplifications can be made
to the governing equations that significantly reduces the computing effort required to
solve them.

6.1

Incompressible flow

In applications where the fluid density is constant, the fluid is considered to be incompressible. While compressibility is actually a property of the fluid and not the flow,
assuming that the fluid behaves as an incompressible substance is often justified if
the Mach number is below 0.3. If the flow is also isothermal, the viscosity is constant,
and the continuity and momentum equations reduce to
div u = 0
1 p
ui
+ div (ui u) = div (grad ui )
+ bi
t
xi
where = / is the kinematic viscosity.

(19)
(20)

AERO 4304

6.2

Lecture 2

Inviscid flow

Inviscid flows are those where the effect of viscosity is so small that it is neglected.
Such flows can occur, for example, far away from solid surfaces. In this case, the
stress tensor reduces to T = pI, and the Navier-Stokes equations reduce to the
Euler equations. The continuity equation is identical to Eqn. 19, and the momentum
equation is
p
(ui )
+ div (ui u) =
+ bi
(21)
t
xi
Since the flows are inviscid, the fluid does not stick to solid surfaces (no skin
friction) and boundary layers do not form. While this is an idealization, the Euler
equations are often used to study compressible flows at high Mach numbers. At such
high velocities, the Reynolds number is high, the boundary layers are very thin, and
the viscous effects are important only in a small region near the walls. Of course, the
inviscid-flow assumption also means that the effect of turbulence will also not be captured. However, because there is no boundary that needs to be resolved, simulations
can use much coarser grids and the computational cost is decreased significantly. In
fact, flows over the whole aircraft have been simulated using Euler equations; capturing the viscous regions in such a large simulation would be very expensive.

6.3

Potential flow

In a potential flow, in addition to the absence of viscous forces in the flow, the velocity
field is also irrotational:
u=0
(22)
From this condition, it follows that there exists a velocity potential such that the
velocity vector field can be defined as u = grad. The continuity equation for
incompressible flow, gradu = 0, then becomes a Laplace equation for the potential :
2 = 0.

(23)

While the Laplace equation cannot be solved analytically for arbitrary geometries,
there are simple analytical solutions for elementary flows (uniform flow, sink, source,
vortex, etc.) and the linearity of the Laplace operator allows these simple solutions
to be combined to create more complicated configurations.

6.4

Other simplified flow models

There are several other flow models that are used to simplify the solution of the
conservation equations. Boundary layer flows simplify the Navier-Stokes equations by
assuming that the diffusive transport is much smaller than the convective transport
and so can be neglected, the velocity component in the principal flow direction is
much larger than in the other directions, and the pressure gradient normal to the

AERO 4304

Lecture 2

principal flow direction is small. Integration of the boundary-layer equations using


various integral methods, which converts them from partial differential equations into
ordinary differential equations, have also been used. Other simplified models relate to
flows at very small Reynolds numbers (termed creeping flows) or buoyant flows with
small density gradients. A huge subset of simplified flows use semi-empirical models
to describe complicated behaviour that cannot be easily solved otherwise, such as
turbulence, combustion, multi-phase flow. The methods for modelling turbulence
will be discussed in further detail later in this course.

You might also like