Você está na página 1de 70

Momentum Heat and Mass Transport

School of Chemical Engineering and Analytical Science


The University of Manchester
Seve Pandiella and Robin Curtis
Fall Term 2015

Contents
1

Introduction: The Basics


1.1 Balance equations . . . . . . . . . . . . . . . . . . . . . .
1.1.1 One dimensional balances . . . . . . . . . . . . .
1.1.2 Differential approach: Three dimensional balances
1.2 Index notation . . . . . . . . . . . . . . . . . . . . . . . .
1.3 Mass balance . . . . . . . . . . . . . . . . . . . . . . . .
1.4 Material versus spatial coordinates . . . . . . . . . . . . .
1.5 Useful theorems . . . . . . . . . . . . . . . . . . . . . . .
1.6 Dimensional analysis . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

3
3
4
5
5
6
7
8
10

Momentum transport
2.1 Balance for frictionless flow . . . . . . . . . . . . .
2.2 Newtons law of viscosity and momentum diffusivity
2.2.1 The stress tensor . . . . . . . . . . . . . . .
2.3 One dimensional momentum transport problems . . .
2.3.1 Introduction . . . . . . . . . . . . . . . . . .
2.3.2 Flow of a falling film . . . . . . . . . . . . .
2.3.3 Laminar flow in a pipe . . . . . . . . . . . .
2.3.4 Horizontal pipe . . . . . . . . . . . . . . . .
2.3.5 Vertical pipe . . . . . . . . . . . . . . . . .
2.4 Creeping flow around a sphere . . . . . . . . . . . .
2.5 Dimensional analysis for the equations of change . .

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

12
12
14
16
19
19
19
21
21
24
25
27

.
.
.
.
.
.

29
29
30
31
32
34
36

.
.
.
.
.
.
.
.
.

39
39
40
40
41
41
42
43
45
48

Multicomponent mass transfer


3.1 Introduction . . . . . . . . . . . . . . . . . .
3.2 Diffusive flux . . . . . . . . . . . . . . . . .
3.3 Examples of steady state diffusion . . . . . .
3.3.1 Arnold diffusion . . . . . . . . . . .
3.3.2 Heterogeneous combustion . . . . . .
3.3.3 Diffusion with homogeneous reaction

.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.

Energy transport
4.1 Energy balances . . . . . . . . . . . . . . . . . . .
4.1.1 Convection and diffusion . . . . . . . . . .
4.1.2 Pressure work . . . . . . . . . . . . . . . .
4.1.3 Viscous work . . . . . . . . . . . . . . . .
4.1.4 Final energy balance . . . . . . . . . . . .
4.2 Problems in heat transfer . . . . . . . . . . . . . .
4.2.1 Heat transfer with an electrical heat source
4.2.2 Heat transfer through composite walls . . .
4.2.3 Heat conduction with a nuclear heat source
1

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

4.2.4
4.2.5
4.2.6
5

Heat conduction with a viscous heating source . . . . . . . . . . . . . . . .


Heat conduction in a cooling fin . . . . . . . . . . . . . . . . . . . . . . . .
Free convection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

Turbulence
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.2 Brief overview of statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.3 Reynold stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.4 Approximate models for turbulence . . . . . . . . . . . . . . . . . . . . . . .
5.4.1 Boussinesq hypothesis or the mean-velocity field closure . . . . . . . .
5.4.2 Prandtl mixing-length theory and von Karman similarity hypothesis . .
5.4.3 Three-region model for momentum transport (universal velocity profile)
5.5 Turbulent flow in a pipe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

50
52
55

.
.
.
.
.
.
.
.

58
58
58
59
61
61
62
64
65

Analogies in momentum, heat, and mass transfer


6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2 Comparison of diffusive transport . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.2.1 Reynolds analogy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

67
67
67
68

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

Lecture 1
Introduction: The Basics
The aim of this module is to describe the transport of momentum, heat, and mass transport and apply
the knowledge to solving fundamental and practical problems in chemical engineering. The first
question is what causes the flow of properties through a system. Thermodynamics tells us that at
equilibrium the temperature, pressure, and chemical potential of all components must be indepedent
of the locotion in the system. Mathematically, this condition is written as
T (r) = T (r0 )
p(r) = p(r0 )
(r) = (r0 )
for all positions r and r0 (note any variable in boldface denotes a vector). If there is a gradient in any of
these quantities, the system is not thermodynamically stable. As a consequence, there will be a flow
of momentum, energy and mass through the system to restore the equilibrium. Thermodynamics tells
us the direction of the flows, but does not provide any information about the rate of change towards
equilibrium. The rates of change are instead described by the general field of transport phenomena,
fluid flow, and heat and mass transport.
The purpose of this chapter is to introduce some tools and concepts which will be used throughout
the remainder of the course.

1.1

Balance equations

The starting point for any problem in transport is to provide a balance equation for a property B going
into and coming out of either a macroscopic or a differential volume element. The property B can be
any quantity, for instance, in the context of heat tranport, B could be energy, for fluid flow, it could
be the x-component of the momentum or the fluid density, and for mass transport, B would be the
concentration of a component . The balance equation can be written as
(accumulation B) = (flow in B flow out B) + (generation B)

(1.1)

Here the accumulation describes the time rate of change for the concentration of B in the control
volume. In this course, the balance equation will be applied to differential volume elements to derive
differential equations to describe the property flows in terms of the location in the system r and time
t.
In order to solve the problem, we first need a convenient way of describing the flows into and out
of the system in terms of the flux of a property B, denoted by JB , which is defined as the amount of
3

a B flowing through a plane normal to the direction of flow for a given amount of time. Note that the
flux is a vector, whose direction is defined by the direction of flow.
To demonstate the meaning of flux, consider a small plane at a position r in the system. The area
of the plane is small enough such that the flow of B across the plane does not depend on the position
and can thus be treated as a constant. The orientation of the plane is given by the unit vector n
, which
is perpendicular to the surface of the plane. The rate of B that crosses the plane with differential area
dA is given by JB n
dA, which corresponds to the projection of the flux on the unit normal. If the
plane is perpendicular to the x-direction, the rate of B that crosses the plane, per unit area is given by
JB
ex = (JB,x
ex + JB,y
ey + JB,z
ez )
ex = JB,x

(1.2)

which is the x-component of the flux vector.


ex ,
ey , and
ez correspond to the unit vectors pointing in
the x-, y-, and z-directions.
In general the flow of the property B is due to two factors, convection and diffusion
JB = Jconv
+ Jdiff
B
B .

(1.3)

The convective flux arises from the macroscopic flow of the fluid that carries B into and out of the
control volume
Jconv
= cB v
B

(1.4)

where cB is the volume concentration of B and v is the velocity field. The diffusive flux is due
to molecular motion and arises from gradients in thermodynamic properties. Examples of diffusive
transport include Newtons law of viscosity (in momentum transport), Fouriers law (in heat tranport),
and Ficks law (in mass transport).

1.1.1

One dimensional balances

In order to demonstrate the flux balances, we first consider a simple system where the property B
only changes in the x-direction. In this case, the control volume can be considered as a rectangular
slab where the planes of the rectangle are located in the yz plane and the rectangular area is equal to
A. The rectangular slab is located between the position x and x + x. The width of the slab x is
chosen small enough such that the properties of the fluid within the control volume can be treated as
uniform. In this case, the toal amount of B in the control volume is given by cB xA. The rate of
production of B is B xA, where B is the generation rate of B per unit volume. Lastly, because
there is only a gradient of B along the x-direction, the rate of flow into the control volume is given by
JB,x (x, t)A and the rate out of the control voume is JB,x (x + x, t)A

[cB (x, t)Ax] = AJB,x (x, t) AJB,x (x + x, t) + B (x, t)Ax


t

(1.5)

Dividing both sides of the equation by x gives


cB (x, t)
JB,x (x + x, t) JB,x (x, t)
=
+ B (x, t).
t
x

(1.6)

Taking the limit as x goes to zero, the first term on the right side of the equation then corresponds
to the definition of derivative, so that
cB (x, t)
JB,x (x, t)
=
+ B (x, t).
t
x

(1.7)

The balance equation for a one dimensional flow problem in cartesion coordinates can always be
written in the form given by Eq.1.7.
4

In the above discussion, you should consider how to choose the control volume. For the onedimnsional, there is no need to do a differential balance along the y and z axis as the properties are
uniform along these directions. As a consequence, a balance is only needed across a differential
element in the x-direction.
Exercise Derive a balance equation that describes the change in a property B along the radial direction
in a sphere. In this case, the properties are uniform in the angular directions usually denoted as and
. For the steady state case, compare the flux of B in the radial direction in a sphere to the flux of B
in the x-direction along a rectangular channel.

1.1.2

Differential approach: Three dimensional balances

Now we consider the system where the property B varies in all directions, x, y, and z. In this case, the
control volume must have a differential thickness along all directions. Thus we choose a rectangular
prism with dimensions x, y, and z. The dimensions are chosen small enough such that the
properties within the control volume can be considered uniform. In this case, there is a net flow of B
in all directions and the balance needs to consider the flux of B across all surfaces.

(cB xyz) = yzJB,x (x + x, y, z, t) + yzJB,x (x, y, z, t)


t
xzJB,y (x, y + y, z, t) + xzJB,y (z, y, z, t)
xyJB,z (z, y, z + z, t) + xyJB,z (x, y, z, t)
+ B xyz
cB
JB,x (x + x, y, z, t) JB,x (x, y, z, t)
=
t
x
JB,y (z, y + y, z, t) JB,y (x, y, z, t)

y
JB,z (z, y, z + z, t) JB,z (x, y, z, t)

+ B
z

(1.8)

In the preceding equation, concentrations and fluxes are functions of the spatial and time variables.
The quantities in parenthesis corresponds to the position where the property is being evaluated. Taking
the limit as x, y, and z go to zero, we find
cB
JB,x JB,y JB,z
=

+ B ,
t
x
y
z

(1.9)

which can be written in vector form as


cB
= JB + B .
t

(1.10)

is the gradient operator, defined by


e x

1.2

+ e y
+ e z .
x
y
z

(1.11)

Index notation

Here we introduce the use of index notation which allows us to write the equations in more compact
form, while maintaining the meaning and facilitating their algebraic manipulation. In index notation,
5

vectors and matrices are referred to by their various components. The expresssion vi refers to the ith
component of the vector v, which could be either the x, y, or z component (i.e. v1 = vx ,v2 = vy ,v3 =
vz ). The expression Aij denotes the ijth element of the matrix A.
If an index is repeated, this implies that the index is summed over all components. Thus,
ui v i ux v x + uy v y + uz v z = u v

(1.12)

When applied to two vectors u and v, this operation corresponds to the dot product, which in turn is
the product of a row vector multiplied by a column vector. Similarly, the multiplication of a matrix
by a vector can be written as
v ij vj ,

(1.13)

where the product is a vector. The multiplication of two matrices A and B is given by in index
notation as
A B Aij Bjk .

(1.14)

Using index notation, the general balance equation (Eq. 1.9) can be written as
cB
JB,i
=
+ B
t
xi

(1.15)

where xi is the ith component of the position vector (i.e. x1 = x, x2 = y, x3 = z) and JB,i is the ith
component of the flux vector JB . The summation convention has been used on the first term on the
right side of Eq. 1.15.

1.3

Mass balance

The first balance to consider when solving transport problems in fluids is the mass balance. The mass
concentration is usually denoted as the density . Because mass cannot be created or destroyed, there
is no generation term in the balance equation. In addition, there is no diffusive flux of mass, so only
the convective flux of mass needs to be considered (Jconv
= v). Thus, the balance Eq. 1.9 gives
B

= v
t

(1.16)

This is typically referred to as the continuity equation and is written in the form

+ v = 0.
t

(1.17)

Using the index notations, Eq 1.17 can be rewritten as

+
(vi ) = 0
t xi

(1.18)

For an incompressible fluid the density remains constant, thus all derivatives of density are equal to
0. With this assumption the equation of continuity reduces to
v = 0.

(1.19)

1.4

Material versus spatial coordinates

The aim of this section is to introduce the concept of material coordinates. The discussion in the
previous sections has made use of the spatial coordinates x, y, z, which correspond to a stationary
coordinate system in which all points of the system can be located. The fluid can be broken up into
many small differential volume elements dVm . The volume element is so small that it contains the
same material at all times. For this reason, we can consider the volume element as a particle in the
fluid. These particles can be identified unambiguously by their location R at some reference time
chosen at t = 0 as no two particles can occupy the same position in space at the same time. The
location R is called the material coordinates of the fluid element or particle.
In order to calculate some property B we could either consider B as a function of the spatial
coordinates and time, or alternatively, B could be considered as a function of the particle label (as
defined by the material coordinates) and time. In the former case, we would have some equation that
tells us how B changes at a fixed position in the coordinate system with time. In the latter case, we are
treating our indepenent variables as the material properties (or the particle label) and time. Material
coordinates are often used because balance equations applied to particles lead to simpler expressions
than when applied to a fixed location in space. This will become more clear when considering the
useful theorems given in the next section.
When material coordinates are used as independent variables, the location in the fixed coordinate
system can be determined from the time evolution of the particle positions. This can be expressed
mathematically as
r = r(R, t);

(1.20)

the position of any particle is a function of the particle label (i.e material coordinates) and time. In
other words, if we want to know some property at a fixed position in space, we first need to identify
what particle is located at that position.
The time rate of change of the spatial position vector for a fluid element is given by
 
r
Dr
=
=v
(1.21)
t R
Dt
By holding the material property constant in the derivative, we are determining the change in position
for the particle located at r, which is the velocity of the fluid at r. Partial derivatives where the material
properties are being held constant are called material derivatives and denoted by using
D
.
Dt

(1.22)

In order to get a better understanding for the meaning of material derivatives, let us break up the
fluid into many small particles. Consider the x-coordinate of the pth particle as denoted by xp . The
material derivative of xp is defined as




dxp
xp (t + t) xp (t)
Dxp
=
= limt0
= vx,p
(1.23)
Dt
dt
t
Note that the material derivative is no longer given by a partial derivative as there is no reason for
holding R constant. This is because holding R constant has the same meaning as labelling the particle
using the subscript p. The derivative corresponds to the rate of change of the x-coordinate for the pth
particle. Similarly we can calculate the rate of change of the temperature for the pth particle as




DT
dTp
Tp (t + t) Tp (t)
=
= limt0
(1.24)
Dt
dt
t
7

The above derivative could only be calculated if we had a thermometer that moved with the particle
(or with the fluid). However, we would generally measure the temperature at a fixed position in space,
which would be defined by
 


T
T (t + t) T (t)
= limt0
(1.25)
t r
t
r
where we have used a partial derivative becuse the temperature is measured as time evolves while
holding the coordinates constant (i.e. at the same position).
The material derivative can be calculated from the spatial derivative. We first express the temperature as a function of the variables x, y, z, and t.
 
 
 
 
T
T
T
T
dT =
dx +
dy +
dz +
dt
(1.26)
x y,z,t
y x,z,t
z x,y,t
t r
The partial derivative is obtained by
 
 
 
 
 
 
 
 
T
T
T
T
T
x
y
z
=
+
+
+
t R
x y,z,t t R
y x,z,t t R
z x,y,t t R
t r

(1.27)

and noting that the velocity components of the particle or fluid are equal to the corresponding derivaties
of the spatial coordinates while holding the particle label constant, we obtain
 
 
 
 
 
DT
T
T
T
T
T

=
+ vx
+ vy
+ vz
.
(1.28)
Dt
t R
t
x
y
z
which can be expressed in more compact notation as
 
 
DT
T
T

=
+ v T.
Dt
t R
t

(1.29)

The meaning of the material derivative can also be clarified by thinking about the temperature that
would be experienced by a diver in the ocean. If the diver is anchored to the ocean floor, they would
experience the temperature reflected by the spatial derivative, that is, the change in temperature at a
fixed position in the ocean. However, if the diver floated with the current of the ocean, the change in
temperature would be given by the material derivative. Following on from this argument, if the diver
ended up getting on a boat moving with a velocity w, the change in temperature of the water next to
the boat would be given by
 
 
dT
T
T

=
+ w T.
(1.30)
dt
t R
t
In the last three equations, we have used the shorthand notation that any partial derivative implies that
the independnt variables are x, y, z, nd t. The material derivative is always given by the symbols
D/Dt.

1.5

Useful theorems

The divergence theorem provides a very usefool mathematical tool that can be used for setting up
balance equations and solving equations. The theorem can be applied to a particle or fluid that takes
up the region space with volume V bounded by the surface defined by A
Z
Z
G dV =
G n dA
(1.31)
V

where G is a vector, and n is the unit normal to the surface. Thus the theorem relates an integral over
the volume of the fluid region to an integral over the surface of the region. A similar equation can be
written for a scalar property S
Z
Z
S dV =
Sn dA.
(1.32)
V

To demonstrate the applicability of the divergence theorem, a force balance can be applied to a
region of fluid, when it is static. In this case, the surface force balances the body force. The surface
force is given by the pressure force, which is directed toward the inside of the fluid region (i.e. pn).
The net surface force is obtained by integrating the pressure force over the entire surface area of the
fluid region. The body force is given by the gravitational force, which is the mass of the fluid inside
the region multiplied by the gravitational constant g. The mass is obtained by integrating the density
of the fluid over the entire region, thus we have
Z
Z
0=
g dV
pn dA
(1.33)
V

and applying the divergence theorem in scalar form Eq 1.32


Z
(g p) dV.
0=

(1.34)

In Eq. 1.34 we could choose to carry out the integration over any region of fluid. This implies that the
integrand of the integral must be equal to 0, thus
0 = g p.

(1.35)

Another very useful equation is the Reynolds transport theorem. A special form of the equation
can be written as
Z
Z
Z
D
S
S dV =
Sv n dA
(1.36)
dV +
Dt Vm (t)
Vm (t) t
Am (t)
where the integrals are carried out over a material volume Vm and a material area A, and the time
derivative is taken with the material properties being held constant. The theorem can be applied to
any scalar property S. The Reynolds transport theorem provides an alternative approach to balance
equations for deriving the equations of change for mass, momentum, and energy balances. Here, we
show how the equation of continuity can be derived. The mass M contained in a material volume
Z
M=
dV.
(1.37)
Vm (t)

The conservation of mass requires that M be a constant, thus,


Z
DM
D
=0=
dV = 0.
Dt
Dt Vm (t)
Application of the Reynolds transport theorem gives
Z
Z
Z
D

dV +
v n dA
dV =
Dt Vm (t)
Vm (t) t
Am (t)

(1.38)

(1.39)

and the area integral can be trasnformed using the divergence theorem to give the final result


Z
Z
D

dV =
+ (v) dV = 0.
(1.40)
Dt Vm (t)
Vm (t) t
9

Because the integration is over an arbitrary region, it follows that the integrant must be equal to 0,
hence the equation of continuity follows

+ (v) = 0.
t

(1.41)

The equation of continuity can be rewritten in terms of the material derviative to give
D
+ v = 0.
Dt

1.6

(1.42)

Dimensional analysis

Consider the problem of a frictionless pendulum consisting of a mass M attached to a string of length
L that is swinging back and forth under the influence of a gravtational field with strength given by
g. In order to use the pendulum as a method for keeping track of time, we need to determine the
pendulum period T or time it takes the pendulum to make one oscillation. Mathematically we seek a
function that relates the system properties to the period T
T = (M, L, g)

(1.43)

In the current form, the function depends on the choice of units. If we choose to measure time in
hours instead of seconds, then the function increases by a factor of 3600. In addition, the mathematical form of the function will depend on the unit choices for M , L and g. However, we know
that there will be a universal function if the equation is written in terms of dimensionless variables.
For instance, many years ago the unit of time corresponded to how long it took the earth to make
one full rotation, approximately 24 hours, which we will call t. We can make the pendulum period
dimensionless using the ratio T /t, which is independent of the choice for units. Thus for T equal to 1
minute, t becomes 1440 min (the number of minutes in a day), and the ratio T /t is equal to 0.000694.
Alternatively if T is reported as 60 sec, then we use a value of t equal to 86400 sec, but the ratio T /t
remains the same. Thus, we can write down the following expression


T
L M gt2
=
, ,
(1.44)
t
l m l
where the value of is now a universal function that is does not depend on the units.
We are free to choose what we want for the values of m and l, so we could choose m to corresponds
to the mass of 1 kg (or 1000g) and t to be one hour (or 60 min). Alternatively, we can simply choose
m to be equal p
to the mass of the pendulum M , and the length l to be equal to L, and the unit of time
to be equal to L/g. In this case the function is now given by
T
p
= (1, 1, 1) = A constant.
L/g

(1.45)

Thus, without doing any physics, we


phave found by dimensional analysis that the pendulum oscisllations have a period proportional to L/g.
We can apply the same approach based on dimensional analysis to characterize the pressure drop
for flowing some fluid through a pipe. In general, we know that the pressure drop per unit length of
pipe, p/L will be a function of the properties for the system, which include the pipe diameter D,
the fluid viscosity , the fluid density and the average velocity in the pipe, which we call v z
p
= (D, , , v z )
L
10

(1.46)

where denotes an undetermined function. Thus at first glance, it appears that the pressure drop is a
function of four different variables. However, as before, the function depends on the choice of units
and here we seek a universal function in terms of the dimensionles variables. The variables can be
made dimensionless by measuring them relative to standard values of mass m, time t, and length l.


p l2 t2
D l3 lt v z t
=
,
,
,
(1.47)
L m
l m m l
The problem can be simplified further by choosing the standard values to reflect the properties of the
system. Thus the unit of mass is given by m = D3 , the unit of time t = D/v z , and length l = D.


p D2 (D/v z )2
D D3 D(D/v z ) v z (D/v z )
=
,
,
,
L
D3
D D3
D3
D


p D

,1
= 1, 1,
2
Dv z
v z L
f = f (Re)
(1.48)
The problem has now been greatly simplified where we now have one function f which depends only
on one variable, the Reynolds number.

11

Lecture 2
Momentum transport
In this section we will first apply the balance equation to the quantity momentum. Because momentum
is a vector, the balance equation will yield three equations corresponding to each component,x, y, and
z. In the first part we only consider transport of momentum by convection and neglect the diffusive
flux. This leads to a differential balance that can be integrated to give Bernoullis equation. Bernoullis
equation does not include any terms ariring from friction losses to the surroundings, which in part,
are due to the effect of the wall on the diffusive flux into the fluid. In the next section is covered the
basics of calculating the diffusive flux, which leads to Newtons law of viscosty. In the last section
we derive the Navier-Stokes equations from the momentum balance.
There are two general approaches that can be used for solving the balance equations. In the control
volme approach, the starting point is to set-up a differential balance about an appropriate differential
volume as has been done in setting up Eq. 1.8. As mentioned before the differential volume needs to
be chosen such that the properties are uniform throughout. If we choose a control volume, which has
differential elements in all coordinates, we can derive a generalized differential equation that can be
applied to all problems. As an example, when the mass is used as the property B, the corresponding
balance equation is the continuity equation Eq. 1.17. The generalized differential equation can also
be derived following a procedure based on the Reynolds Transport Theorem. The generalized balance
equations for momentum are termed the Navier-Stokes equations.

2.1

Balance for frictionless flow

The first step in deriving the balance equation is to provide an expression for the momentum density or
momentum per unit volume, which corresponds to B of Eq. 1.8 when applied to momentum balance.
The momentum of a fluid particle is equal to mv, where m and v are the mass and the velocity of
the particle, respectively. Thus momentum per unit volume is the number of fluid particles per unit
volume multiplied by their average velocity, i.e. momentum density is equal to v. As momentum is
a vector there are three compenents to the momentum vx , vy , and vz .
The next term to consider is the generation of momentum, which can be addressed using Newtons
second law:

(mv) = F
t

(2.1)

where m is the mass and F is the force acting on the mass. Thus, forces generate momentum. Note
that the force has units of momentum per unit time, which are the same units for each term in the
momentum balance, such as the rate of accumulation of momentum or the flowrate of momentum
into and out of the control volume.
Consider the possible forces that act on the control volume, which, in this case, is a cubic element
of fluid, which has dimensions x, y, and z. The pressure force acts on each side of the cube in
12

a direction normal to the corresponding side and pointing in towards the center of the cube. Thus the
total pressure force acting on the cube is due to the pressure force applied to all six sides of the cube
pressure force = [(p(x, y, z) p(x + x, y, z)] yz
ex
+ [(p(x, y, z) p(x, y + y, z)] xz
ey
+ [(p(x, y, z) p(x, y, z + z)] xy
ez .

(2.2)

In addition, we need to consider the graviational force on the cube, which is equal to xyz g.
The generation term momentum has units of momentum per unit time per unit volume or force per
volume. Thus, to calculate the generation term, we need to divide the forces by the volume of the
cube (equal to xyz) and take the limit as the volume goes to zero. Applying this procedure
leads to
momentum = p + g.

(2.3)

The convective flux of a property B is given by Jconv


= cB v, where cB is the concentration of
B
property B. The concentration of momentum is equal to v, so that the flux of momentum is given
by
Jconv
momentum = (v)v.

(2.4)

Note that the flux of a scalar property is a vector. However, the flux of a vector property becomes a
three by three tensor. The components of the tensor Jconv
momentum are given by

vx vx vx vy vx vz
vy vx vy vy vy vz
(2.5)
Jconv
momentum =
vz vx vz vy vz vz
In order to better understand the origin of the tensor Jconv
momentum consider the cube with sides given by
xyz, which contains three orthogonal planes. There are three possible components of momentum that can be transported across any of the orthogonal planes. Thus the individual components of
the tensor corresponds to the transport of the jth momentum component across the ith plane
conv
Jmomentum,ij
= vi vj .

(2.6)

The general balance equation derived in chapter 1 applied to momentum


cmomentum
= Jmomentum + momentum
t
gives (upon substitution of the momentum flux and generation terms)

(2.7)

v
= (vv) p + g
(2.8)
t
As noted earlier, Eq. 2.8 corresponds to three balances, one for each component of the momentum
flux. This is apparent when considering the left hand side of the equation, which is a vector.
With the help of the continuity equation, the momentum balance can be rearranged to give

+v
t
t
v

t
v

t
v
t

= v (v) v v p + g



= v
+ (v) v v p + g
t
= v v p + g
= v v

p
+ g.

13

(2.9)

If we assume that the fluid is incompressible (that is is a constant) and the the flow is irrotational
(i.e. v = 0), then we find


v
1 2 p
=
v + gr .
(2.10)
t
2

For a steady-state system, all derivatives with respect to time vanish, so the right side of the above
equation can be integrated to yield
1 2 p
v + g r = constant.
2

(2.11)

This is just Bernoullis equation. Fo for an incompressible fluid at steady state, the momentum equation reduces to Bernoullis equation when the diffusive momentum flux can be neglected.
Exercise Derive the balance equation for the x-component of the momentum by doing a differential
momentum balance around a control volume. Once complete, write out Eq. 2.8 using the index
notation introduced in Chapter 1.

2.2

Newtons law of viscosity and momentum diffusivity

Consider a fluid located between large parallel plates. The plates are located in the z-planes at positions z = 0 and z = H. Initialy the system is at rest. At time t = 0, the lower plate is moved at
a constant velocity V in the x-direction, while the top plate is held motionless. At any given instant
in time, the velocity of the fluid at the top plate is equal to 0 and at the bottom plate the velocity is
equal to 0. This is called the no-slip boundary condition. As time progresses, the fluid at the bottom
of the channel drags along the fluid at higher z. Thus initially only the fluid near to the bottom plate
moves. As time progresses, the fluid velocity in the top section increases as well until a steady state
is reached corresponding to a linear velocity profile with respect to z. When a steady state is reached,
if the flow is laminar, the force exerted on the bottom plate is given by
F
V
=
A
H

(2.12)

thus the constant of proportionality between the shear rate and the shear force is the viscosity. These
types of forces are termed viscous forces. Note that the force is proportional to the velocity difference
of the two plates divided by their separation.
If we break up the fluid into small differential planes of thickness z, then the velocity everywhere
in the plane is a constant. Newton found that the force per unit area exerted by the plane at z by the
fluid below is given by
zx =

dvx
dz

(2.13)

Here zx is the force per unit area in the x direction exerted on the z-plane. The first subscript defines
the plane that is being acted upon by the force, and the second subscript defines the direction of the
force. Eq. 2.13 states that the force exerted on the plane is proportional to the velocity gradient in
the z-direction. The direction of the force is in the direction opposite to the velocity gradient. In this
case, the x-component of the velocity is decreasing with increasing z. Thus the direction of the force
is in the positive x-direction, or we can say that the plane with the greater velocity pulls along the
plane of smaller velocity. Eq. 2.13 is an empirical observation, rather than a universal law. However,
it accurately captures the behaviour of gases and liquids as long as the molecules in the liquid have
14

molecular weights less than 5000 Da. Gases and liquids that obey this relation are referred to as
Newtonian fluids.
There is another way of interpreting the meaning of Eq. 2.13. By moving the bottom plate, xmomentum is being transferred to the neighbouring fluid. This fluid in turn transfers some of its
x-momentum to the plane above and so forth. zx can be defined as the rate of x-momentum transfer
to the z-plane per unit area. The viscous forces can also be interpreted as the contribution to the
diffusive flux of momentum. There is a gradual diffusion of x-momentum in the z-direction from
a region of high momentum (or high velocity) to one of low momentum (or low velocity), which is
consistent with the definition of as a diffusive flux. The velocity gradient can thus be interpreted
as a driving force for momentum transport in the same way that a temperature gradient drives energy
transport, and a mole fraction (or molar concentration) gradient drives mass transport.
Eq 2.13 corresponds to the special case where there is a velocity gradient only in one direction.
The generalized equation that applies to a three dimensional case where velocity gradients can exist
in all directions is given by

 


vi
vj
2
vx vy vz
ij =
+
+

+
+
ij
(2.14)
xj xi
3
x
y
z
where ij is the kronecker delta, which is equal to 1 when i = j and equal to 0 when i 6= 0. is
called the dilatational viscosity, which is equal to 0 for gases. Generally, liquids are incompressible,
in which case the last term on the right hand side of Eq 2.14 is equal to 0 since v = 0.
Eq. 2.14 contains terms of the form ii that are referred to as normal stresses. Accordingly, they
correspond to situations where there is a non-zero gradient of the velocity in the direction of flow (i.e.
terms where (vi /xi ) 6= 0). This can occur when the cross sectional area for flow is changing along
the pipe or channel such as in a converging geometry or in a sudden expansion. For incompressible
liquids or gases the normal stresses are then given by ii = 2 (vi /xi ).
As with the convective flux, the diffusive flux can be expressed as a 3 3 tensor, that is
Jdiff
momentum =

(2.15)

xx xy xz
= yx yy yz
zx zy zz

(2.16)

where the stress tensor is given by

There are three components of momentum which can be transferred to each of the three orthogonal
planes.
Adding the contribution of diffusive flux to the momentum equation, we find
v
= (vv) p + g.
t

(2.17)

It is important to note that the left hand side of Eq 2.17 is a vector, so that there are three independent
balance equations, one for each direction. The equation can be written in simpler notation
v
= + g
t

(2.18)

where is called the combined momentum-flux tensor. The components of the tensor are given by
ij = ij + ij p + vi vj .

15

(2.19)

Using index notation, the momentum balance equation (2.17) can be rewritten as
ji
p
vi

(vj vi )

+ gi .
=
t
xj
xj
xi

(2.20)

Using the continuity equation, the momentum equation can also be written as

vi
ji
p

vi
vj
+ vi
= vj
vi

+ gi
t
t
xj
xj
xj
xi
ji
p
vi
vi

+ gi .
= vj

t
xj
xj
xi

(2.21)

The balance equation can also be written in terms of the velocity gradients by substituting Newtons
law of viscosity to give
 
 

vi

vi
vj
2
vj
p
=
(vj vi ) +

+
+

+ gi .
(2.22)
t
xj
xj
xj xi
3
xi xj
xi
For an incompressible fluid (where v = 0) with a constant viscosity, the equation reduces to
v

1
+ v v = 2 v p + g.
t

(2.23)

The ratio / is called the kinematic viscosity usually denoted by the symbol .
The derived momentum balances are known as the Navier-Stokes equations. Here, we have derived them using cartesian coordinates. The analogous equations can be derived in either cylindrical
or spherical geometries, although the mathematical expressions become more complicated. The expressions are not given here, but can be found in almost any Fluid Mechanics textbook.
exercise Derive the balance equation for the x-component of the momentum by doing a differential
momentum balance around a control volume. Consider the case of a fluid flowing between two
parallel plates in the x-direction. Once complete, write out the Eq. 2.17 using the index notation
introduced in Chapter 1.

2.2.1

The stress tensor

In Fig. 2.1 is shown the projection of a differential volume element on the xy-plane. The coordinate
axis are given by the variables, x, y, and z, which should be distinguished from the symbols denoting
the corners of the volume element given by x0 , y 0 , and z 0 . The quantities denoted by ij |i0 corresponds
to the stress exerted on the plane defined by constant i0 . The shear stress on each plane i0 contains one
contribution from the shear tensor ij |i0 , whereas the normal force is the sum of the shearing force
and the normal force due to pressure. When doing a momentum balance we need to define the forces
acting on the cube, which have a direction given by the arrows. Below we describe how we define the
direction of the arrows and the corresponding forces.
We first consider the direction of the pressure force. The pressure force acts in a direction normal
to the surface. However, there are actually two possibilities for the direction of the force, as the vector
could be pointing away from either side of the surface. In order to overcome this ambiguity, we always
define the normal to point outwards from the surface into the surroundings. With this definition, it is
clear that the pressure force can be defined by
pressure force = pni0 Ai0
16

(2.24)

Figure 2.1: Schematic of the various forces acting on a differential volume element in cartesian
coordinates. Only the projection of the volume on the z plane is shown. The quantities denoted by
correspond to the corners of the volume element. Each surface on the volume element is defined by
the coordinate that remains constant throughout the surface.
where the subscript i0 defines the surface, Ai0 corresponds to the area of the surface, and the ni0 is
the outwarding pointing unit normal vector to the surface i0 . So for instance, the surface defined by
x so that the pressure force points in the direction of
constant x0 + x has a normal equal to e
ex ,
which is consistent with the Fig. 2.1.
Similarly when defining the direction of the shear stress components ij , we also need to use a
convention for defining the orientation of the surface that is being acted upon by the shear force.
Here, we use a convention, which is used in some textbooks, but not in all. When the surface i0 has
an outward directed normal in the negative i direction, the shear stress ij is represented as acting
in the positive j direction. This direction corresponds to the arrows shown on the schematic in Fig.
2.1. Note that this direction does not actually correspond to the direction of the force as we have
not yet specified the velocity profile, a key point that we clarify below. When we follow the above
convention, we must use the following form of Newtons law of viscosity for an incompressible fluid


vi
vj
ij =
+
.
(2.25)
xj xi
If we had used the alternative definition for the orientation of the surface; i.e. when the surface has
an outward directed normal in the positive i direction, the shear stress ij is represented as acting in
the positive j direction. In this case, we would need to use the following equation for Newtons law
17

of viscosity:

ij =

vj
vi
+
xj xi


.

(2.26)

Thus the definition of Newtons law of viscosity depends on our definition for the direction that is
being acted upon by the shear. You will find in some textbooks Newtons law of viscosity given by
2.26, in this case, they are using the latter convention. If a textbook is using the latter convention, the
arrows shown in Fig. 2.1 would be pointing in the opposite directions.
For the convention used in the notes, there is an alternative way for remembering the direction
of ij . The force is acting in the positive j direction, when the surroundings are located at a lesser
value of the i coordinate than the surface being acted upon. Conversely, the force acts in the negative
j direction when the surroundings are located at a greater value than the corresponding surface.
Now, if we specify a velocity profile, we can then determine the direction of the force. For
instance, consider the one dimensional flow problem along the x-axis between two infinite plates,
where the y-direction is perpendicular to the plates. When the top plate is moved at a constant velocity
in the positive x-direction, at steady state, there will be a positive gradient for the x-component of the
velocity for all values of y between the plates, that is
vx
>0
y

(2.27)

In this case, the shear stress on the surface defined by constant y 0 , according to Eq. 2.25 is given by
yx |y0 =

vx
|y0 < 0
y

(2.28)

so that the force is in the opposite direction of the arrow for the corresponding shear direction shown in
Fig. 2.1, that is the force is in the negative x-direction. Following a similar argument. yx |y0 +y < 0
as the velocity gradient is positive for all y. Thus the shear force on the plane defined by y 0 + y is in
the positive x-direction. These forces are consistent with the inuition that fluid with greater velocity
is dragging along the fluid with lower velocity located.
For the one-dimensional flow problem described in the previous paragraph, there will also be
non-zero components of xy = yx following the definition given by Eq. 2.25
xy |x0 =

vx
|x0
y

(2.29)

According to our convention the shear stress on the x0 suface is acting in the positive x direction, but
the force is in the negative x direction. Conversely, the force on the x0 + x surface is in the positive
x direction. Because the velocity gradient is only a function of y, we find that
xy |x0 = xy |x0 +x .

(2.30)

Thus the forces on the planes of constant x will be exactly opposite to each other and cancel out
in the momentum balance for the one-dimensional flow problem. This applies to all surfaces that
are perpendicular to the flow in one-dimensional flow fields when the cross-sectional area for flow
remains constant. For this reason, these terms are often not included in shell balances covered in the
next section.
According to Newtons law of viscosity, we will always find that ij = ji . Physically, this equality
is required such that the volume element does not gain any angular momentum when being sheared
by the surroundings. If the only non-zero shear forces corresponded to yx then the applied forces
would cause the element to rotate. The shear forces xy exacly balance the forces yx so that there is
no change in the angular momentum of the element.
18

2.3
2.3.1

One dimensional momentum transport problems


Introduction

In the previous section, we derived the general equation for the momentum balance. In the remaining
sections of this chapter, we will apply the balance equations to solve specific examples. The general
approach to solving these problems is to first apply the equation of continuity to determine the dependence of the velocity on the spatial coordinates. In order to do this, it is very important to consider
the boundary conditions to the problem. Once done, the momentum balance is then used to determine
properties such as the velocoity and shear stress profile, the pressure drop, the maximum and average
velocity, and/or the volumeteric flowrate. The problems can be solved by either doing a force balance
about a differential control volume, or by applying the Navier-Stokes equations.
In this chapter, we only consider steady-state problems where the fluid velocity varies with respect to one dimension. In later chapters, we will consider multi-variate problems, such as occurs in
boundary layer problems.

2.3.2

Flow of a falling film

Consider the problem of a film of water flowing down an inclined plane. We will define the coordinate
axis such that the x-coordinate is parallel with the plane and the y-coordinate is perpendicular to the
plane. The plane is treated as infinite in the x and z coordinates. The angle of the plane with respect
to the vertical direction is taken to be and the thickess of the film is given by , which is independent
of the x-coordinate.
The approach we take to solving this problem is to carry out a shell balance. First we consider the
boundary conditions to this problem. We apply the no-slip boundary condition to the fluid next to the
surface, where the velocity of the fluid is equal to the velocity of the solid. Furthermore, the velocity
of the fluid in the y-direction is equal to 0 everywhere. We choose the origin of the y coordinate to be
located at the top of the fluid layer, and the positive y-direction to be in the downwards direction.
The equation of continuity only applies to flow in the x-direction. Because the flow is at steadystate,
vx
=0
x

(2.31)

or the velocity in the x-direction only depends on y, which we represent mathematically using the
notation vx (y). The velocity will be a maximum at the air exposed surface of the fluid and decrease
with increasing y according to our definition of the coordinate system.
A momentum balance can be applied in the y-direction. In order to have transport of y- momentum by either diffusion or convection into a surface would require having a non-zero velocity
in the y-direction vy . Thus, the momentum balance only needs to consider balancing the body (or
gravitational) force and the pressure force.
0 = gravitational force + pressure force

(2.32)

The gravitional force needs to consider the projection of the vertical direction on the y-axis, which is
given by gy = g sin . Thus the momentum (or force) balance is given by
0 = g sin xyz + (xz p)|y (xz p)|y+y ,

(2.33)

Dividing through by the differential volume element and taking the limit y 0
0 = g sin

19

dp
.
dy

(2.34)

The above equation corresponds to the static fluid. The pressure increases with an increase in y due to
the head of the fluid. The other key conclusion to draw is that the pressure does not vary with x. The
pressure at y = 0 is the atmospheric pressure and is the same everywhere on the air-exposed surface
of the fluid.
The force balance in the x direction needs to consider the momentum transport by diffusion (pressure force is not considered since pressure is independent of x)
0 = gravitational force + viscous force

(2.35)

It is important to note the direction of the viscous force. Here we follow the common notation that
the viscous force acts on the y-surface by fluid at lower y in the positive x-direction (or equivalently
yx corresponds to the transfer of x-momentum from the region of liquid at lower y)
0 = g cos xyz + (xz yx )|y (xz yx )|y+y .

(2.36)

Dividing through by the differential volume element and taking the limit y 0
0 = g cos

yx
y

(2.37)

Another more rapid way to arrive at Eq. 2.37 is to start from the general momentum balance
equation. Considering only the x-component gives

vx
vx jx p
= vj

+ gx .
t
xj
xj
x

(2.38)

Because the system is at steady state, the left hand side of the equation is equal to zero. Applying this
equation requires knowing which components of the velocity are equal to zero, and the dependence on
spatial coordinates of the non-zero velocity components. Here, vy and vz are equal to zero and vx (y).
Thus we can arrive at Eq. 2.37 by noting that pressure is independent of x and using gx = g cos .
Eq. 2.37 can be integrated with respect to y to give
yx = gy cos + A

(2.39)

where A is an integration constant. A can be determined from the boundary conditions. At the airexposed surface of the fluid, there is no mechanism for transferring momentum from the air to the
fluid, thus yx (y = 0) = 0. This condition requires that A equal 0. Next, Newtons law of viscosity
yx =

vx
y

(2.40)

is substituted into Eq. 2.39. The velocity profile is then obatined upon integrating with respect to y
vx =

g cos 2
y + B.
2

(2.41)

B is an integration constant that needs to be determined from the no-slip boundary condition, vx (y =
) = 0. Substituing this condition into Eq 2.41 and solving for B gives
B=

g cos 2
,
2

which needs to be substituted back into the equation for the velocity profile Eq 2.41

 y 2 
g 2 cos
.
vx =
1
2

20

(2.42)

(2.43)

Once the velocity profile is fully determined, we can calculate properties such as the maximum and
average velocity, the volumetric flowrate, and the pressure drop. The average velocity is determined
from integrating across the cross-sectional area for flow
RW R
vx (y) dzdy
hvx i = 0 R W0 R
dzdy
0
0
Z
1
=
vx (y) dy
0
Z 
 y 2   y 
g 2 cos 1
1
d
=
2

0
2
g cos
(2.44)
=
3
where W corrsponds to the width of the plate. The volumetric flowrate is given by the average velocity
multiplied by the corss sectional area for flow
gW 3 cos
Q = W hvx i =
.
3

(2.45)

Lastly, we should consider what the overall momentum balance on a macroscopic volume tells
us. Generally, the overall balance includes a friction-loss term which is the force exerted by the wall
on the fluid, or the loss of momentum in the fluid due to the presence of the wall. This term can be
evaluated from integrating the shear stress profile along the area of the wall contacting the fluid
Z LZ W
Fx =
yx |y= dxdz
0
0
Z LZ W
dvx
=

|y= dxdz
dy
0
0
= gLW cos .
(2.46)
We find that that the shear or frictional force exerted by the wall on the fluid balances the gravitational
force exerted in the x-direction. The same equation can be obtained by carrying out a momentum or
force balance on a macroscopic control volume including the entire cross sectional area for flow.

2.3.3

Laminar flow in a pipe

2.3.4

Horizontal pipe

Lets consider steady-state flow through a cylindrical pipe of radius R, where we designate the z-axis
as parallel to the direction of flow. As before, we should first consider the various velocity components
and the boundary conditions using cylindrical coordinates. The velocities in the radial and angular
directions are qual to zero, that is vr = v = 0. The no-slip boundary condition is applied to the wall,
such that vz (r = R) = 0. In addition we note that the velocity must obtain its maximum value in the
centre of the pipe. The equation of continuity indicates vz is independent of z.
We choose to do a momentum balance in the z direction about a differential ring with thickness
given by z and r. Because vz is independent of z, there is no net flow of momentum due to
convection. Thus, at steady-state the momentum balance reduces to
0 = pressure force + viscous force.

21

(2.47)

The pressure force acts on the surface defined by constant z, whereas the shear force acts on the radial
surface (i.e. defined by constant r). Thus,
0 = (2rr p)|z (2rrp )|z+z + (2rz rz )|r (2rz rz )|r+r

(2.48)

Dividing though by 2rrz and taking the limit as z 0 and r 0 gives


p
1 (rrz )
=
z
r r

(2.49)

Because vz is only a function of r, the right hand side of Eq. 2.49 is independent of z. In addition,
we can assume the pressure in the pipe does not depend on r (the change of pressure due to the fluid
head is much smaller than the pressure drop along the length of the pipe). Thus, the left hand side of
Eq 2.49 is independent of r. Thus, we can set either side of Eq. 2.49 to a constant A. The pressure
drop can be obtained by integrating
Z Pout
Z L
dp =
A dz
(2.50)
Pin

to give
A=

P
p
Pin Pout
=
=
L
L
z

(2.51)

where the subscripts in and out correspond to the positions at z = 0 and at z = L respectively.
Inserting Newtons law of viscosity
rz =

vz
r

into
P
1 (rrz )
=
L
r r

(2.52)

gives the differential equation

vz
r
r


=

P
r,
L

(2.53)

which can be integrated twice to give


vz (r) =

P 2
r + A ln r + B.
4L

(2.54)

A and B are unknown constants from the integration. A = 0 from the requirement that the velocity
at the center of the pipe must be finite (i.e. the logarithm of 0 is infinite). B is determined from
substituting the no-slip boundary condition vz (R) = 0 into Eq 2.54
B=

P 2
R
4L

(2.55)

which gives

 r 2 
P R2
vz (r) =
1
.
4L
R
Eq 2.56 correspnds to the well-kniwn velocity profile in a pipe under laminar flow conditions.
22

(2.56)

The volumetric flowrate in the pipe is calculated from integrating the velocity profile over the
cross-sectional area for flow A
Z R
Q =
2rvz (r) dr
0

Z R
 r 2 
P R2
1
dr
=
2r
4L
R
0
Z

P R4 1
= 2
x 1 x2 dx
4L 0
P R4
=
.
(2.57)
8L
The average velocity can then be calculated by dividing the volumetric flowrate by the cross sectional
area A
Q
Q
=
A
R2
P R2
=
.
8L

vz =

(2.58)

The Reynolds number is defined by


Re =

Dv z

(2.59)

where the pipe diameter is D = 2R. The friction factor f for flowing fluid in a pipe is defined by the
equation
f=

P D
.
2v 2z L

(2.60)

After subtituting the definition of the Reynolds number and the expression for the average velocity v z
into Eq. 2.60, we find
f=

16
.
Re

(2.61)

It is also helpful to calculate the force exerted by the wall on the fluid, which is given by integrating
the shear stress at r = R over the entire length of the pipe
Z L
Z 2
Fw =
dz
d(rz r)|R .
(2.62)
0

The shear stress at the wall is calculated using using Newtons law of viscosity
rz |R =

vz
P R
|R =
r
2L

(2.63)

where vz /r has been evaluated from the derivative with respect to r of Eq. 2.56. Substituting the
result into Eq. 2.62 and carrying out the integration gives
Fw = R2 P.

(2.64)

Thus the force exerted by the wall exactly balances the pressure forces acting on the fluid going into
and out of the tube. Alternatively, we can say the pressure drop in the pipe arises only from the friction
due to the wall. The same deduction could be obtained by carrying out a balance on the entire volume
of the fluid contained in the pipe of length L.
23

2.3.5

Vertical pipe

For flow in a vertically oriented pipe, we need to include the gravitational force in our balance equation. Here, we use the same definition of coordinates as for the horizontal pipe, except that now the z
axis is pointing downwards. We label a length of pipe L using the subscripts in and out, to define the
start and endpoints of the length.
As before, we choose to do a momentum balance about a differential ring with thickness given by
z and r. A momentum balance in the r direction will only include the pressure forces acting on
the ring. In this case, we would find that the pressure does not depend on the radial direction, i.e.
p
=0
(2.65)
r
Thus, we carry out a momentum balance in the z-direction:
0 = (2rr p)|z (2rrp )|z+z + (2rz rz )|r (2rz rz )|r+r + g2rzr. (2.66)
Dividing the equation by 2rrz and taking the limit as z,r 0 gives the partial differential
equation
p
1 (rrz )
g =
(2.67)
z
r r
As the left side of the equation only depends on z and the right side only depends on r, we can set
either side of the equation equal to a constant A.
p
1 (rrz )
g =
z
r r
We can now integrate the z-component of the equation to give
p
= g + A
z
Z Pout
Z L
dp =
[g + A] dz.
A=

Pin

(2.68)

(2.69)

Carrying out the integration yields


Pout Pin = [g + A] L

(2.70)

and upon rearrangement


Pout Pin
Pout g Pin
g =
.
(2.71)
L
L
The pressure drop along the horizontal tube occurs only due to the frictional losses from the force
exerted by the pipe wall on the fluid. In a vertical pipe, there is the additional contribution from
the gravitational force on the pressure change across the pipe. Here, we define another pressure
P = P gz which subtracts out the effect of the hydrostatic head of the fluid on the pressure. Thus,
we have Pin = Pin and Pout = Pout gL. With this definition, the constant A now becomes
Pout Pin
P
A=
=
.
(2.72)
L
L
The result can be substitued back into the differential Eq. 2.67 to give
A=

P
1 (rrz )
=
,
(2.73)
L
r r
which is analogous to Eq. 2.52, except that the actual pressure drop has been replaced by P. The
remainder of the problem is solved in the exact same way as for the horizontal pipe.
Thus, the force of the wall acting on the fluid will be given by
Fw = R2 P,
that is, the pressure drop P is only due to the frictional forces.
24

(2.74)

2.4

Creeping flow around a sphere

In this chapter, so far we have been concerned with problems dealing with fluids flowing through pipes
or on surfaces. However, another important class of problems involves flow of fluids about submerged
objects, such as would occur in the sedimantation of solid particles in a tank, in the motion of a hot air
balloon, or in the flight of an airplane, In each of these cases, we would be interested in calculating the
drag force on the particle. For instance, the fuel consumption of an airplane will in part be determined
by the energy required to overcome the drag force. In addition, the velocity of a particle settling in
a tank is determined by the competing effects of gravitational forces against a buoyancy force and
a drag force. The drag force corresponds to the force exerted by the surrounding fluid on the solid
object or particle and is determined by the sum of the normal and shear forces integrated over the
particle surface as discussed further below.

Figure 2.2: Sphere of radius R around which a fluid is flowing. The text uses the spherical coordinate
system described in the figure
Problems involving flow about objects involve curved streamlines, which can not be dealt with by
the tools we have introduced in this chapter. However, due to the important nature of these problems,
it is still worthwhile to review some key equations describing the motion of fluids about a sphere
as depicted in Fig. 2.2. The sphere has a radius R, the fluid has a viscosity and density and
approaches the sphere from below along the z-axis at a velocity given by v (note the denotes the
velocity of the fluid at any location that is far away from the sphere). At low values of the Reynolds
number (less than 0.1), the equations of momentum have been solved to determine analytical expressions for the velocity components, the pressure distribution, and the non-zero components of the
stress tensor. The shear force exerted on the r-surface in the direction is given by
r

3 v
=
2 R

 4
R
sin .
r

(2.75)

The above equation corresponds to any location in the fluid outside of the sphere. The pressure
distribution in the fluid is given by
3 v
p = p0 gz
2 R
25

 2
R
cos
r

(2.76)

where p0 corresponds to the pressure at z = 0 but at an x,y coordinate located far from the sphere.
The pressure distribution reduces to the hydrostatic pressure far from the sphere p = p0 gz as
expected. The components of the velocity distribution are given by
"
 
 3 #
3 R
1 R
vr = v 1
+
cos
(2.77)
2 r
2 r
and
"
v = v

3
1
4

 
 3 #
R
1 R

sin .
r
4 r

(2.78)

The velocity distributions reduce to the expected values based on the boundary conditions to the
problem, that is at the surface of the sphere, the no-slip boundary condition requires that v (R) =
vz (R) = 0 and far from the sphere v = 0 and vz = v .
What we need to determine are the forces exerted by the fluid on the particle surface. The force
can be broken down into a normal and a tangential component. Furthermore, intuitively, we know
that the x and y components of these forces will sum to zero as there is no motion of the sphere in
these directions. Thus we first consider calculating the normal force (given by the pressure) exerted
on the sphere in the z-direction. The normal force is given by the projection of the pressure force on
the z-axis p cos multiplied by the differential surface area R2 sin d d. The net normal force is
obtained from integrating the local force over the entire surface of the sphere
Z 2 Z
Fn =
(p cos ) R2 sin d d.
(2.79)
0

The pressure at the surface of the sphere given by


p|r=R = p0 gR cos

3 v
cos
2 R

(2.80)

is inserted into Eq. 2.79 and the integration is carried out to give
4
Fn = R3 g + 2Rv .
3

(2.81)

The first term on the right side of Eq. 2.81 is the buoyancy force and the second term of Eq. 2.81 is
called the form drag. The buoyancy force corresponds to the gravitational force acting on the mass
of fluid that is displaced by the sphere, but acting in the opposite direction of gravity. If the sphere
density is the same as the surrounding fluid, the sphere will be motionless as the buoyancy force
cancels the gravitational forces. Furthermore. balloons filled with helium move upwards because the
mass of the helium in the balloon is less than that of the surrounding air being displaced by the ballon.
There is also a tangential force at each point along the surface of the sphere, given by R , which
is the shear stress acting on the constant R surface in the -direction. The z-component of the force
is given by the shear stress (R )( sin ) acting upon the differential area equal to R2 sin d d.
The net tangential force is obtained by integrating the local force over the surface of the sphere
Z 2 Z
Ft =
(r |r=R sin ) R2 sin d d
(2.82)
0

using the relation for the shear stress evaluated at r = R


r |r=R =

3 v
sin .
2 R
26

(2.83)

Carrying out the integration gives


Ft = 4Rv .

(2.84)

which is called the friction drag force because it arises from the tangential or shear force exerted by
the fluid on the surface of the sphere.
The total force exerted by the fluid on the particle is given by
4
F = R3 g + 6Rv .
3

(2.85)

The second term on the right side of the equation is called the drag force on the sphere and arises only
from the fluid motion. As a consequence, the second term is also known as the kinetic force and is
oftern referred to as Stokes law
Fk = +6Rv

(2.86)

In this section, the motion of the fluid relative to the solid particle arises due to the gravitational force
acting upon the particle. However, Stokes law is also valid for describing the motion of particles in
an electric field, or under a chemical potential gradient field as occurs in the diffuion of colloidal
particles.
The terminal velocity of a spherical particle falling through a fluid corresponds to the case where
the forces balance and there is no net acceleration. In general the time to reach the terminal velocity
after the particle is initially at rest is very small. Thus, the particle trajectory can be accurately
calculated by just considering the terminal velocity, which can be obtained by setting the forces acting
on the particle equal to zero to give
4 3
4
R s g = R3 g + 6vt R.
3
3

(2.87)

vt = 2R2 (s )g/9.

(2.88)

After rearrangement, this gives

The result is only valid for values of Re less than 0.1.


Exercise Determine the velocity of particle as a function of time for the case of when a particle is
initially at rest in the fluid.

2.5

Dimensional analysis for the equations of change

Eq. 2.23 can be rearranged in terms of the substantial derivative to give

Dv
= p + 2 v + g
Dt

(2.89)

We seek to write the above equation in terms of dimensionless variables. The first step is to determine
what properties of the system will be used to define the basic units of mass, length, and time. In all
systems we will encounter here, there will be a characteristic length (D), velocity (V ) and density
(). So for instance, if we want to calculate the trajectory of a hot air balloon, D would be the balloon
diameter, V would be the velocity of the wind, and would be the air density. Conversely for flow in a
pipe, D is the pipe diameter, is the density of the flowing fluid, and V could be either the maximum
velocity or an average velocity of the fluid relative. Using these characteristic properties, our basic
27

units of mass, length, and time, can be made dimensionless using the relations l = D, m = D3 , and
t = D/V . Using these units we can then define the following dimensionless variables denoted by a
superscript *
v =

v p p0 tV
; p =
; t =
V
V 2
D

2 D

2
=
;

=
;

=
;
=
xi
D xi
D
D2 Dt

V
D

D
Dt

(2.90)

These relations can be substituted into Eq. 2.89 to obtain the dimensionless form for the equation of
momentum

 

1 2
V
D
1 2

(v

p
V
+

(v V ) + g

V
)
=

D Dt
D
D2





gD g
Dv

2
= p +
v +
.
(2.91)
Dt
DV
V2 g
In Eq. 2.91, the terms in the square brackets are dimensionless numbers that reflect the physical
properties of the system and the system dimensions. The Reynolds number Re given by V D/ and
the Froude number Fr is given by gD/V 2 , thus
g
1 2
Dv
= p +
v + Fr

Dt
Re
g

(2.92)

Two systems that are characterized by the same set of dimensionless numbers will be governed by the
same differential equation. Furthermore, if the boundary conditions are the same for the two systems
when written in dimensionless numbers, then the equations of change become mathematically idential. This means that the dimensionless velocity distribution v (x , y , z , t) and the dimensionless
pressure distribution p (x , y , z , t) are the same. In this case, if the behaviour of one of the systems
is known, either experimentally or from a calculation, then the properties of the second system can be
calculated using the appropriate scale factors.
The dimensionless numbers have a further use when examining the equations of change, they
provide insight into which terms are significant. The factors (i.e. Dv /Dt;2 v ; p ,g/g) are on
the order of 1 (or equal to 0). Thus the dimensionless numbers determine the contributions of each of
the terms to the equations of change. A low Reynolds number indicates the viscous contribution to the
momentum equation (2 v ) is larger than any of the other terms. Under these conditions, the main
mechanism for transport is by viscous forces or by diffusion of momentum. This case corresponds to
creeping flow which was used in the analysis of determining the drag force exerted on a sphere in the
previous section. Conversely, for large values of the Re, diffusive transport is no longer significant,
and convective or inertial forces control the behaviour of the flowing system. Similarly, a low value of
the Froude number indicates gravitational forces are insignificant and can be neglected in the analysis.

28

Lecture 3
Multicomponent mass transfer
3.1

Introduction

In many chemical engineering unit operations, the key element is the transport of a particular chemical
species from one location to another. For example, in the case of a chemical reactor, we need the
reactants to mix and diffuse together in order for them to react to form the product we desire. In a
liquid-liquid extractor, we want a chemical species to transfer from one liquid phase to another. In a
gas absorption unit, we want a particular species in the gas phase to diffuse into the contacting liquid
phase. The key to designing these processes is understanding mass transfer.
In mass transfer problems, we generally deal with multicomponent systems. Based on the general
balance equations that we developed in Chapter 1, deriving the balance equation for a species in a
multicomponent system is fairly straightforward. For a particular molecular species , we have
c
= N +
t

(3.1)

where c is the molar concentration of , N is the absolute molar flux of species , and is the
rate of generation of per unit volume (e.g., by chemical reaction).
In a multicomponent system, different species generally travel at different velocities. Therefore,
for each component that is present in the system, there is a velocity v . The absolute molar flux of
is given by
N = c v .

(3.2)

Now lets consider what happens when a drop of ink is placed in a moving stream of water. The
molecules of ink move as a result of two processes. First, the drop of ink moves along with the
general flow of the stream. This contribution to the motion of the ink is referred to as convective
flux. In addition to moving along with the flow of the stream, the ink also spreads outwards. This
contribution to the motion is referred to as diffusive flux.
Before we can define a diffusive flux, we must first define a reference velocity with respect to
which diffusion occurs. This choice of the reference velocity v . is quite arbitrary, but, in general,
takes the form
X
v =
a v
(3.3)

where a are arbitrary weighting functions that must satisfy the normalization condition
X
a = 1.

29

(3.4)

If we let a = w , the mass fraction of component in the system, then v = v, the center-of-mass
velocity of the system:
X
v=
w v
(3.5)

This is the velocity that appeared in the continuity equation, momentum balance equation, and the
energy balance equation that were developed in the previous sections. If we let a = x , the mole
fraction of component in the system, then v = v , the molar-average velocity:
X
v =
x v
(3.6)

In what follows, we will use v as our reference velocity.


With the definition of the reference velocity, the convective flux is given by c v , and the diffusive
flux J is defined as
J = c (v v ) .

(3.7)

To finish off this subsection, we demonstrate how the absolute flux can be written in terms of the
diffusive flux and the convective flux.
J = c (v v )
X
= N c
x0 v0
0

= N x

N0

(3.8)

and so we find
N = J + x

(3.9)

The first term on the right side of the equation represents the diffusive flux, while the second term
represents the convective flux.

3.2

Diffusive flux

In order to complete the description of the transport of species , we need an expression for the
diffusive flux J . We need a way to relate the diffusive flux to the various driving forces in the
system. In this subsection, we will develop this expression for binary mixtures, composed of species
A and B.
The main driving forces for diffusion are gradients in the chemical potentials of the various species
in the system. Let us consider a binary mixture consisting of molecules of type A and type B. A
given species A will tend to move from areas where its chemical potential A is high to areas where
its chemical potential is low. The effective force felt by a molecule A due to nonuniformities in its
chemical potentials is A .
The A molecules, however, are impeded from freely moving (to even out their chemical potential)
by the presence of the B molecules. The B molecules get in the way. The force exerted on the A
molecules by the B molecules is given by
RT
xB (vA vB )
DAB
30

(3.10)

where R is the gas constant, T is the absolute temperature, and DAB is the diffusion coefficient.
This drag force exerted by the B molecules on the A molecules is proportional to the amount of B
molecules (therefore the xB factor) and to the relative velocities between A and B molecules.
The diffusion coefficient DAB is an empirical coefficient that describes diffusion, like the thermal conductivity describes heat conduction and the viscosity describes momentum diffusion. The
diffusion coefficient has the units of length squared over time (e.g., m2 /s). The larger the diffusion
coefficient, the weaker the drag of B on A, and as a result, the A molecules can diffuse more quickly.
The smaller the diffusion coefficient, the slower the A molecules diffuse.
If we balance the driving force for motion of species A against the drag exerted by species B,
which opposes the motion, we find
A =

RT
xB (vA vB )
DAB

(3.11)

This equation can be mainpulated to yield


RT
(vA xA vA xB vB )
DAB
RT
=
(vA v)
DAB
RT
=
cA (vA v)
cA DAB
cA DAB
A
JA =
RT

A =

(3.12)

where cA is the molar concentration of A. Eq. 3.12 is known as the Maxwell-Stefan equation. It
shows the diffusive flux is proprtional to chemical potential gradients.
From thermodynamics, we know that the chemical potential of a species can be written as
= o (T, p) + RT ln x

(3.13)

If we substitute this into the Maxwell-Stefan equation, we find


J =

cA DAB
[oA (T, p) + RT ln xA A ]
RT

(3.14)

If we assume that the temperature and pressure of the system are uniform (i.e., constant with respect
to position), the Maxwell-Stefan equation reduces to


ln A
xA
(3.15)
J = cDAB 1 + xA
xA
where c is the overall molar concentration of molecules. If we further assume that the system behaves
as an ideal mixture (i.e., A = 1), we have
JA = cDAB xA

(3.16)

This is known as Ficks law.

3.3

Examples of steady state diffusion

In this section, we will go over some simple examples of how to apply Ficks law to various mass
transport problems.
31

3.3.1

Arnold diffusion

In this example, we consider a cylinder containing pure liquid A (see Fig. 3.1). Immediately above
the liquid surface, the mole fraction of A is yA0 . The distance between the surface of the liquid and
the top of the cylinder is H, and this region is filled with a stagnant gas B. Outside the cylinder, there
is a fast moving stream of gas, which maintains the mole fraction of A at yAH . We assume that the
system is at steady state.

Figure 3.1: Arnold diffusion cell.


We begin this problem, as with all diffusion problems, by performing a balance on species A on
a thin slab of gas located within the cylinder. The location of the bottom surface of the slab is z,
the thickness of the slab is dz, and its cross-subsectional area is S. Because we are dealing with
a steady-state problem, there should be no accumulation of A. In addition, because there are no
chemical reactions occuring in the system, there should be no generation/consumption of A. From
the symmetry of the system, we assume that motion only occurs in the z-direction. Therefore, the
influx of A through the bottom surface of the slab should be equal to the outflux of A through the
upper surface
0 = (NA,z S)|z (NA,z S)|z+dz

(3.17)

Dividing by the volume of the slab Sdz,




1 (NA,z S)|z (NA,z S)|z+dz
0=
S
dz

(3.18)

[NA,z (z)S(z)] = 0
z

(3.19)

Taking the limit dz 0,

where we have multiplied both sides of the equation by S. This equation can be integrated to yield
NA,z (z)S = wA

(3.20)

where wA is an unknown integration constant, which represents the total molar flowrate of A up the
cylinder.
32

The absolute flux of A in the system is given by


NA,z = JA,z + yA (NA,z + NB,z )

(3.21)

The first term on the right side of the equation is the flux of A due to diffusion; the second term
is the flux of A due to the overall motion of the system (i.e., convection). In this example, the A
molecules will diffuse through a stagnant layer of B. The B molecules do not move. Therefore, we
have NB = 0. The expression for the absolute flux of A then becomes
NA,z = JA,z + yA (NA,z + 0)
1
=
JA,z .
1 yA

(3.22)

If we assume that the gas mixture of A and B behaves as an ideal gas and that the temperature and
pressure are uniform along the height of the cylinder, the diffusive flux follows Ficks law (see Eq.
3.16)
JA,z = cDAB

yA
z

(3.23)

and, consequently, the absolute flux of A is given by


NA,z =

cDAB yA
.
1 yA z

(3.24)

Substituting this expression into Eq. 3.20, we find

cDAB yA
S = wA
1 yA z
wA
ln(1 yA ) =
z + C.
ScDAB

(3.25)

where C is a constant. Substituting the boundary conditions (ie, yA = yA0 at z = 0 and yA = yAH at
z = H) into Eq. 3.25, we have
ln(1 yA0 ) = C
ln(1 yAH ) =

wA
H +C
ScDAB

(3.26)

which can be solved for the integration constants wA and C.


C = ln(1 yA0 )
ScDAB 1 yAH
wA =
ln
H
1 yA0

(3.27)

Substituting these expressions for the integration constants into Eq. 3.25, we find
ln

1 yA
z
1 yAH
=
ln
1 yA0
H
1 yA0

z/H
1 yA
1 yAH
=
1 yA0
1 yA0

Exercise Determine the loss rate for A from the container.

33

(3.28)

3.3.2

Heterogeneous combustion

In this example, we will examine the combustion of a spherical coal particle of radius R. Very
far from the coal particle, the mole fraction of oxygen (which we label as species A) is given by
yA = yA = 0.21. On the surface of the coal particle, the oxygen reacts with carbon to form carbon
monoxide
2C + O2 2CO

(3.29)

We assume that this reaction proceeds so quickly that the concentration of oxygen on the surface of
the particle is zero (i.e., yA = 0 at r = R). In order for the coal particle to burn, oxygen needs to
diffuse from far from the particle to the surface of the particle. In this problem, the rate of the reaction
is controlled by the transport of oxygen to the coal particle.
For every oxygen molecule that reacts, two molecules of CO are created. These CO molecules
diffuse in the opposite direction as the oxygen molecules. Therefore, we have NCO = 2NA . In
addition, we assume that the nitrogen in the atmosphere is stationary (i.e., NN2 = 0).
We start the problem by performing a balance on oxygen molecules A within a spherical shell of
inner radius r and thickness dr located outide of the coal particle. Because the system is at steady
state, there is no accumulation in the shell. In addition, because the only chemical reaction that occurs
is on the surface of the coal particle, there is no creation/consumption of oxygen in the shell. We make
the assumption that transport of species only occurs in the r-direction. So in this case, the influx of A
at the inner surface, which has an area S(r) = 4r2 (the surface area of a sphere of radius r) must be
balanced by the outflux of A from the outer surface:
0 = (NA,r S)|r (NA,r S)|r+dr
If we divide by the volume of the spherical shell S(r)dr = 4r2 dr, we find


1 (NA,r S)|r+dr (NA,r S)|r
0=
S
dr

(3.30)

(3.31)

If we take the limit dr 0, we arrive at the following differential equation

[NA,r (r)S(r)] = 0
r

(3.32)

where we have multiplied both sides of the equation by S. This equation can be integrated to obtain
NA,r (r)S(r) = wA

(3.33)

where wA is an integration constant. The physical meaning of wA is the total moles of oxygen that
diffuse toward the coal particle per unit time.
The absolute flux of oxygen can be written as
NA,r = JA,r + yA (NA,r + NN2 ,r + NCO,r )
= JA,r + yA (NA,r 2NA,r )
1
=
JA,r
1 + yA

(3.34)

According to Ficks law


JA,z = cDAB

34

yA
r

(3.35)

If we substitute Ficks law into Eq. 3.34 and then insert the resulting expression into Eq. 3.33, we find
cDAB yA
4r2 = wA
1 + yA r
cDAB yA
wa

=
1 + yA r
4r2
wA
cDAB ln (1 + yA ) =
+C
4r

(3.36)
where we have used the fact that S(r) = 4r2 .
We can determine the unknown integration constants wA and C by substituting the boundary
conditions into the expression for yA . By doing this, we find
cDAB ln(1 + yA ) = C
0=

wA
+C
4R

(3.37)

which can be solved to give


wA = 4RcDAB ln(1 + yA )

(3.38)

Note that wA represents the rate at which oxygen is diffusing towards the coal particle. This is the
same rate at which oxygen is being consumed by the combustion reaction.
By substituting the expressions for the integration constants back into the original expression for
yA , we get
1 + yA
4RcDAB
=
ln(1 + yA )
1 + yA
4r
R
1 + yA
= ln(1 + yA )
ln
1 + yA
r

cDAB ln

(3.39)

From the solution of this problem, we can also estimate how long it will take for the coal particle to
completely react. Eq. 3.38 tells us how quickly oxygen reacts to form CO. This is precisely twice
the rate at which carbon atoms (coal) are consumed at the surface of the particle. Therefore, we have
wC = 2wA .
To determine the rate at which the coal particle is shrinking, we perform a mole balance for
carbon. For a particle of radius R, the amount of carbon present is given by 4R3 cC /3, the volume of
the particle times the molar density of carbon. The rate at which the particle shrinks is then given by:


4R3
cC = wC
t
3


4R3
cC = 8RcDAB ln(1 + yA )
t
3
cC R
R
= cDAB ln(1 + yA )
2 t
4cDAB t
ln(1 + yA ) + R02
(3.40)
R2 =
cC
where R0 is the initial radius of the particle. The time t required for the particle to disappear is
t=

cC R02
4cDAB ln(1 + yA )
35

(3.41)

3.3.3

Diffusion with homogeneous reaction

In this example, we consider the absorption of species A from a gas phase into a liquid phase composed mostly of B (see Fig. 3.2). The mole fraction of A just inside the liquid phase (i.e., at z = 0)
is equal to yA0 . At a distance beneath the surface of the liquid layer, there is an impenetrable solid
wall. In addition to the diffusion of A in the liquid phase, there is also a first order reaction that
converts species A to species C. The kinetics of this reaction are given by
A = k1 cA
= k1 cyA

(3.42)

where A is the rate of generation of A per unit volume, k1 is a kinetic constant, and c is the overall
molar concentration of the liquid.

Figure 3.2: Absorption of A from a gas phase to a liquid phase.


To begin the problem, we perform a species balance for A on a thin slab in the liquid. Because
we are dealing with a steady-state problem, there is no accumulation in the slab. From the geometry
of the problem, we assume there is only motion in the z-direction. So there is an influx of A from the
left side of the slab and an outflux of A from the right side of the slab. Unlike the previous examples
we considered in this chapter, there is a chemical reaction that consumes A. Therefore, the balance
equation is given by
0 = (NA,z S)|z (NA,z S)|z+dz (k1 cyA )Sdz


1 (NA,z S)|z+dz (NA,z S)|z
(k1 cyA )
0=
S
dz

(3.43)

Taking the limit dz 0,


1
(NA,z S) = (k1 cyA )
S z
NA,z
= k1 cyA
(3.44)
z
We assume that B is stationary (i.e., NB = 0). In addition, we assume that species C flows in the
opposite direction of species A (i.e., NC = NA ). The absolute flux is then given by
NA,z = JA,z + yA (NA,z + NB,z + NC,z )
= JA,z + yA (NA,z + 0 NA,z )
= JA,z
yA
= cDAB
z
36

(3.45)

Putting this expression into 3.44

yA
cDAB
= k1 cyA
z
z
2 yA
= ya /l2
z 2

(3.46)

where l = (DAB /k1 )1/2 . This is a second-order, linear differential equation with constant coefficients.
The general solution to this equation is given by
yA (z) = C1 ez/l + C2 ez/l

(3.47)

where C1 and C2 are constants, which for the moment are unknown. The flux can be determined by
substituting Eq. 3.47 into Eq. 3.45:
NA,z = cDAB

yA
z


cDAB
C1 ez/l C2 ez/l
(3.48)
l
We can determine the constants C1 and C2 by using the boundary conditions. At z = 0, we have
yA = yA0 , which leads to
=

yA0 = C1 + C2 .
Since the wall is impermeable to species A, the flux of A at z = must be zero:

cDAB
NA,z () =
C1 e/l C2 e/l
l
0 = C1 e/l C2 e/l

(3.49)

(3.50)

The equations can be solved to give the coefficients C1 and C2


e/l
e/l + e/l
e/l
C2 = yA0 /l
e + e/l
Substituting these coefficients into Eq. 3.47, we find
 (z)/l   (z)/l 
e
e
+
yA (z) = yA0
/l
/l
/l
e +e
e + e/l
 (z)/l

e
+ e(z)/l
= yA0
e/l + e/l
cosh( z)/l
= yA0
cosh /l
C1 = yA0

(3.51)
(3.52)

(3.53)

To determine the rate at which A is absorbed from the gas phase, we need to determine the flux of A
at the surface of the liquid. The flux is given by
yA
z 

yA0 e(z)/l e(z)/l
= cDAB
l
e/l + e/l
 (z)/l

cA0 DAB e
e(z)/l
=
l
e/l + e/l
cA0 DAB sinh(( z)/l)
=
l
cosh(/l)

NA,z (z) = cDAB

37

(3.54)

where cA0 is the concentration of A at the left surface of the liquid at z = 0. The flux of A at the
surface is given by
cA0 DAB sinh(/l)
l
cosh(/l)
cA0 DAB sinh /l
=
(DAB /k1 )1/2 cosh /l

NA,z (z = 0) =

= cA0 (k1 DAB )1/2 tanh /l


This is the rate (per unit area of liquid) at which A is absorbed from the gas phase.

38

(3.55)

Lecture 4
Energy transport
4.1

Energy balances

In this section, we derive the balance equation for energy. The total energy of the system is the sum
of three terms: (i) the internal energy of the system, (ii) the kinetic energy of the system, and (iii) the
potential energy of the system. The total energy per unit mass e can then be written as:
1
e = U + v 2 + (r)
2

(4.1)

where U is the internal energy per unit mass of the system, and is the potential energy per unit
mass of the system, which is, in general, a function of position. Examples of potential energy are
gravitational energy (where (r) = gr) or electrostatic energy (e.g., a charged system in the presence
of an electric potential).
Now lets perform an energy balance on a small subsection of a system. Our control volume will
be a small rectangular box of dimensions x, y, and z. The concentration of energy inside the
box is equal to e. Thus, the rate of accumulation of energy in the control volume is
accumulation = xyz

(e)
t

(4.2)

The rate of accumulation per unit volume is then given by


(e)
accumulation
=
xyz
t

(4.3)

From the first law of thermodynamics, we know that energy is conserved. Thus, any accumulation
of energy in our control volume must be due to flow of energy from other parts of the system. This
can occur through several mechanisms. In the following, we will consider convection, conduction,
viscous work, and pressure work. Thus, the total energy balance is given by
(e)
= J + W
t

(4.4)

where W is the work done on the system and J is the total energy flux, which is composed of a
diffusive and convective contribution
J = Jconv + Jdiff .
The individual contributions to the energy balance are described in more detail below.

39

(4.5)

4.1.1

Convection and diffusion

Diffusive energy (heat) flux q is caused by gradients in temperature and chemical potential. We will
only consider heat flux which is driven by temperature differences, which is known as conduction.
One simple relationship between heat flux and the temperature gradient is given by Fouriers law of
conduction
Jdiff = q = kT

(4.6)

where k is the thermal conductivity.


There is yet another type of flux for energy, which has no analogy in the transport of other quantities (such as mass and momentum). This is transport of energy due to radiation. We will not consider
this mode of heat transfer in this module.
The convection of energy into the system occurs when material enters (or leaves) the control
volume due to the flow of the fluid. The convective flux of energy is given by the energy density
(energy per unit volume) multiplied by the velocity of the fluid
Jconv = (e)v

(4.7)

Transport can occur across the six different surfaces of the control volume
energy flux in energy flux out = (Jx |x Jx |x+x ) yz
+ (Jy |y Jy |y+y ) xz
+ (Jz |z Jz |z+z ) xy

(4.8)

Dividing the above equation by the total volume of the system and taking the limit that x, y, and
z 0 yields
in out
Jx Jy Jz
=

xyz
x
y
z
= J
= (e)v q
(e)vi
qi
=

.
xi
xi

4.1.2

(4.9)

Pressure work

Now lets determine the work due to pressure forces. As a simple example consider a piston-cylinder
assembly aligned along the z direction. In this case, the work to push the cylinder a diistance z is
just given by the force multiplied by the displacement, Fz z. If the displacement is carried out over
a time perioud of t, the rate of work is given by Fz z/t or Fz vz which is the force multiplied
by the velocity of the piston surface. Thus, the rate of pressure work on a surface is given by the
force (which has a direction normal to the surface) multiplied by the velocity of the surface along the
direction of the displacement (which is also normal to the surface). The total work due to pressure is
obtained by summing over all surfaces
pressure work = [(pvx )|x (pvx )|x+x ] yz+
[(pvy )|y (pvy )|y+y ] xz+
[(pvz )|z (pvz )|z+z ] xy

(4.10)

and the rate of the pressure force on the system per unit volume is then:
pressure work

=
(pvi ).
dxdydz
xi
40

(4.11)

4.1.3

Viscous work

Because the material outside the control volume moves at a different velocity than the material inside, viscous stresses arise on the surfaces of the control volume. These viscous stresses lead to the
conversion of kinetic energy into internal energy in the system. The rate of conversion on a particular
surface is given by the force acting on the surface multiplied by the velocity of that surface along the
direction of the force (this is analogous to the pressure force). The total rate at which viscous forces
perform work on the system is then given by the sum of the work performed on each of the surfaces:
[(vx xx )|x (vx xx )|x+x + (vy xy )|y (vy xy )|y+y + (vz xz )|z (vz xz )|z+z ] yz+
[(vx yx )|x (vx yx )|x+x + (vy yy )|y (vy yy )|y+y + (vz yz )|z (vz yz )|z+z ] xz+
[(vx zx )|x (vx zx )|x+x + (vy zy )|y (vy zy )|y+y + (vz zz )|z (vz zz )|z+z ] xy
= viscous work (4.12)
The viscous work per unit volume is then given by

(xx vx ) +
(xy vy ) +
(xz vz )+
x
x
x

(yx vx ) +
(yy vy ) +
(yz vz +
y
y
y

(zx vx ) + (zy vy ) + (zz vz )


z
z
z

=
(ij vj ).
xi

viscous work =

(4.13)

The viscous work per unit volume per unit time is denoted by Sv in the text below.

4.1.4

Final energy balance

Applying the general balance equation to energy, we find


(e)

qi

=
(evi )

(ij vj )
(pvi )
t
xi
xi xi
xi

(4.14)

Using the continuity equation, this can also be written as

e
e
qi

= vi

(ij vj )
(pvi )
t
xi xi xi
xi

(4.15)

Now let us look at some simpified versions of this equation that can be applied to commonly encountered problems in heat transfer. If the system is entirely motionless and there is no potential energy
(e.g. gravitational forces are not important), then e = U , v = 0 and the energy balance equation is
given by

U
= q.
t

(4.16)

If we make the further assumption that the thermal conductivity of the system is constant, and we
make use of the thermodynamic relation dU = Cv dT at constant volume (density), where Cv is the
isochoric heat capacity, then
Cv

T
= k2 T
t

which is known as the heat or diffusion equation.


41

(4.17)

We will now rearrange the energy equation (Eq. 4.15) in a more convenient form. The first thing
we need to do is develop a balance equation for the kinetic energy. By multiplying Eq. 2.21 by vi , we
find
vi
vi
ji
p
vi
= vj vi
vi
vi
+ gi vi
t
xj
xj
xi
1 v 2
1
ji
p
v 2
vi
vi
+ gi vi

= vj
2 t
2
xj
xj
xi




ji
1 2

1 2
p
v =
v vj vi
vi
+ gi vi
(4.18)
t 2
xj 2
xj
xi
Subtracting Eq. 4.18 from Eq. 4.15, we find





qi
1 2
1 2

U + v gi xi = vj
U + v gi xi

(ij vj )
(pvi )

t
2
xj
2
xi xi
xi

qi
vi

vi
(U gi xi ) = vj
(U gi xi )
ji
p
gi vi
t
xj
xi
xj
xi
qi
vi
U
vi
U

ji
p
(4.19)

= vj
t
xj
xi
xj
xi
If we assume that the material in the control volume is in local thermodynamic equilibrium, we can
relate the internal energy to the local temperature and pressure of the system




U
U
dU =
dT +
dp
T p
p T
"  

#
  

 
S
V
V
S
= T
p
dT + T
p
dp
T p
T p
p T
p T
"  
"

#



 #
S
V
V
V
= T
p
dT + T
p
dp
T p
T p
T p
p T


p p
dp
= Cp
dT (p T T p)
(4.20)

where p is the thermal expansivity of the material and T is the isothermal compressibility of the
material.
For a liquid or solid, the pressure dependence of the internal energy is quite weak (i.e. the second
term in Eq. is small). In this case, we find that the energy balance can be written as
Cv

4.2

T
T
qi
vi
vi
= Cv vj

+ ji
p
t
xj
xi
xj
xi

(4.21)

Problems in heat transfer

In this following sections, we will study commonly encountered heat transfer problems. As with the
momentum balance equtions, there are two general approaches to solving these problems. One could
either start from the generalized energy balance as given by Eq. 4.15 or one can solve a shell balance
for energy transport. We use the latter approach throught the remainder of this chapter. In some cases,
it might be necessary to couple the energy balance with a momentum balance, for instance, in order
to calculate a velocity profile, which could relate to the transport of energy by convection or be the
cause of a viscous heating source.
The problems we consider in the following sections are meant to teach the principles of heat
transfer, as such they are chosen both for their physical importance, but also for their mathematical
solvability.
42

4.2.1

Heat transfer with an electrical heat source

In Figure 4.1 is shown a schematic of an electrical wire with a cross-section radius given by R. The
electrical current in the wire leads to an irreversible conversion (or dissipation) of electrical energy
into heat, which will cause a temperature rise in the wire. The rate of heat production per unit volume
is given by
Se =

I2
ke

(4.22)

where I is the cuurent desnity in amp s1 , and ke is the electrical conductivity in ohm1 cm1 , and
Se has units of J cm3 s1 . The thermal and electrical conductivity can be assumed constant for the
change in temperature expected to occur in the wire. The main purpose of the energy balance is to
calculate the temperature rise in the wire and the heat lost to the surroundings.

Figure 4.1: An electrically heated wire showing the cylindrical shell over which the energy balance
is made.
In order to do the calculation, we assume the wire is at steady-state. The surface temperature of the
wire is maintained at a constant temperature of T0 and we assume there is no variation in properties
along the length of the wire. In this case, there will be a maximum temperature in the centre of the
wire, temperature will decrease with increasing r as energy diffuses radially outwards.
In order to solve for the temperature profile, we carry out a shell balance about a cylindrical shell
of thickness r and length L. The energy balance reduces to
accumulation = 0 = flux in flux out + generation

(4.23)

The only mechanism for heat transport is by conduction as there is no convective term for a solid wire
(i.e. the veloicity is zero everywhere).. In this case the shell balance reduces to
0 = (2rLqr )|r (2rLqr )|r+r + 2rLrSe
43

(4.24)

where qr is the energy flux in the radial direction. Dividing through by 2Lr gives
(rqr )|r+r (rqr )|r
r0
r

Se r = lim

(4.25)

and taking the limit as r 0 gives the differential equation


d
(rqr ) = Se r.
dr

(4.26)

The above equation is a first order differential equation. Because Se is independent of r, the equation
can be integrated to give
qr =

Se r C1
+
2
r

(4.27)

where C1 is an integration constant to be evaluated from the boundary conditions. Because the heat
flux can not be inifinite in the centre of the wire at r = 0, the integration constant C1 must be equal to
zero. Thus, we have
qr =

Se r
2

(4.28)

which states that the heat flux increases linearly with r. Fouriers law of conduction qr = k(dT /dr)
can be substituted into Eq. 4.28 to give another first order differential equation
k

Se r
dT
=
.
dr
2

(4.29)

Integrating the equation with respect to r gives


Se r 2
T =
+ C2 .
4k

(4.30)

C2 is the integration constant to be evaluated from the boundary condition that at r = R , the temperature is equal to T0 . Using this condition gives C2 = T0 + (Se R2 /4k) and we obtain the final result
that

 r 2 
Se r 2
T T0 =
.
(4.31)
1
4k
R
According to the above equation, there is a parabolic temperature rise going away from the centre of
the wire.
Once the temperature rise has been calculated, all the other properties of the system can be calculated. The maximum temperature will occur at the centre of the wire
Tmax T0 =

Se R 2
.
4k

(4.32)

The average temperature rise can be calculated from integrating across the cross sectional area of the
wire to give
R 2 R R
(T (r) T0 ) r drd
< T > T0 = 0 0 R 2 R R
r drd
0
0
Se R 2
=
.
(4.33)
8k
44

Thus the temperature rise in the wire is one half the temperature change from the centre to the edge
of the wire.
In addition the heat flow from the surface of the wire to the surroundings (i.e. the heat lost) can
be calculated from integrating the heat flux qr over the surface of the wire (with area given by 2RL
Q|r=R = 2RLq|r=R
Se R
= 2RL
2
2
= R LSe .

(4.34)

This just indicates that all the energy dissipated from the current is lost to the surroundings. This
should not be surprising since the system is at steady state, all the energy generated by the heating
must be lost to the surroundings, otherwise there would be an accumulation of energy in the wire. A
similar result would be obtained by carrying out an energy balance over the entire volume of the wire.

4.2.2

Heat transfer through composite walls

A very commonly encountered problem in the application of transport phenomena to chemical engineering design is to calculate the heat transfer through walls made up of several materials, with
different resistances (or conductivities). Examples of this include deriving expressions for the overall
heat transfer coefficient to describe a heat exchanger, or calculating the heat lost through processing
equipment with and without insulation. The equipment, for instance, could include a reactor or a
distillation column. In most of the cases outlined above, we are interested in determining the heat
transport through cylindrical shells (i.e. a tube, or a column). As such, we consider the problem of
heat transfer through composite walls in cylindrical geometry.
Here we develop an expression for the rate of heat transfer though a cylindrical shell composed
of three different materials, each with a different thickness as shown in Fig. 4.2 below. The purpose
of this exercise is to develop an expression for the rate of heat transport (denoted by Q with units of
energy per time) as a function of the temperature inside the tube (Ta ) the temperature outside the tube
Tb , the diameters of each tube (as denoted in Fig. 4.2), and the thermal conductivity of each tube (as
denoted by the values for k). For this problem the boundary conditions will be given by an equation
for the heat flux in terms of a heat transfer coefficient at the inside surface of the smallest tube h0 and
the outside surface of the largest tube h3 . We refer to Region01 as the region inside the smallest tube,
Region12 as the region in the middle-sized tube, and Region23 as that of the largest tube.
As always with these problems, the first step is to set-up a balance through a cylindrical shell of
a tube. We choose to carry out the shell balance first in the region of the smallest tube (Region01)
of length L. Because the tube is at steady state and there are no generation terms, the heat balance
reduces to
2rLqr |r 2rLqr |r+r = 0.

(4.35)

Dividing by 2rL and taking the limit as r goes to zero give the differential equation for the heat
flux in the radial direction qr
d
(rqr ) = 0.
dr

(4.36)

rqr = C1 = r0 q0

(4.37)

Intergating the above equation gives

45

Figure 4.2: Heat transfer through a composite wall tube, which could be a model for a laminated heat
exchanger tube, or an insulated cylinrcal body of a distillation column.
The integration constant can be evaluated at either the outside or inside surface of the tube. Here we
have chosen to evaluate the integration constant at the insidersurface of the tube. The meaning of the
integration constant becomes more clear if we multiply both sides of the equation by 2L to give
2Lrqr = C1 = 2Lr0 q0 = Q

(4.38)

where we note that the total rate of heat transfer Q, which is a constant as a function of r, is just given
by the heat flux at any value of r multiplied by the surface area at r which is perpendicular to the
direction of the heat flux.
Inserting Fouriers law into Eq. 4.38 gives a differential equation for the temperature inside Region01
dT
Q
=
.
(4.39)
dr
2L
A similar shell balance can be applied to Region12 and Region23 to give equations for the temperature
profiles in the corresponding tubes
k01 r

k12 r

dT
Q
=
dr
2L

(4.40)

and
Q
dT
=
.
(4.41)
dr
2L
Eqs. 4.39 to 4.41 can be integrated with respect to r. A definite integration can be carried out to
avoid the use of integration constants. The integral is carried out from the inside surface to the outside
surface of each tube. The resulting equations are given by
k23 r

T0 T1 =

Q ln(r1 /r0 )
2L k01
46

(4.42)

and
T1 T2 =

Q ln(r2 /r1 )
2L k12

(4.43)

T2 T3 =

Q ln(r3 /r2 )
.
2L k23

(4.44)

and

Eqs. 4.42 to 4.44 provide relationships for the temperature difference across each tube. However,
generally, we will not know the surface tube temperatures. We seek a relationship for the rate of
heat transfer in terms of the temperature far from the tube surfaces Ta and Tb . This requires using
the boundary conditions, where we introduce the heat transfer coefficients at the inside surface of the
smallest tube h0 and the outside surface of the largest tube h3 . The heat flux at these surfaces is given
by
qr |r=r0 = h0 (Ta T0 )

(4.45)

qr |r=r3 = h3 (T3 Tb ).

(4.46)

and

Rearranging the equations for the temperature difference gives


Ta T0 =

q0
Q 1
=
h0
2r0 L h0

(4.47)

T3 Tb =

q3
Q 1
=
h3
2r3 L h3

(4.48)

and

Eqs. 4.42 to 4.44 and 4.47 and 4.48 can be summed to give an expression for the total heat transfer Q
in terms of the temperature different Ta Tb . The resulting equation is given by
Q0 = 2Lr0 q0
=

2L(Ta Tb )
1
r 0 h0

ln(r1 /r0 )
k01

ln(r2 /r1 )
k12

ln(r3 /r2 )
k23

1
r 3 h3

(4.49)

We can define an overall heat transfer coefficient based on the inside surface area of the smallest tube
U0
Q0 = 2Lr0 q0 = U0 (2Lr0 )(Ta Tb )

(4.50)

where the overall heat transfer coefficient is then given by


1
=
r 0 U0

X ln(rj /rj1 )
1
1
+
+
r0 h0 j=1
kj1,j
rn hn

!
.

(4.51)

Eq. 4.51 provides a relationship between the overall heat transfer resisatance (U01 ) and a sum over
the indivdual resistances due to each of the walls and from the boundary layers on the inside and
outside of the composite walls.

47

4.2.3

Heat conduction with a nuclear heat source

In this section we will consider the use of a nuclear heat source to heat up the contents of a reactor. The
heat source comes from spherical particles containing fissionable material, which can be modelled as
shown in Fig. 4.3. In the sperical particle, the fissionable material is contained within a radius Rf ,
which is surrounded by aluminum cladding out to a radius of RC . Here the superscript f is used to
refer to the region containing the fissionable material and the superscript C refers to the aluminum
cladding. Inside the fuel element, fission fragments are produced that have very high kinetic energies.
Collisions between these fragments and the atoms of the fissionable material provide the major source
of thermal energy. The rate of energy production from the fissionable material Sn (with units of energy
per unit time per unit volume) depends on the location within the material. An approximate form to
be used here is given by

 r 2 
Sn = Sn0 1 + b
(4.52)
Rf
Here Sn0 is the volume rate of heat production at the center of the sphere and b is a dimensionless
positive constant. In this case, the energy production is the smallest at the center of the sphere and
grows as a parabolic function with r.

Figure 4.3: A spherical nuclear fuel assembly, showing the temperature distribution within the system.
The purpose of this problem is to determine the temperature profile within the particle and the
rate of energy transfer from the outside surface of the particle to the surroundings. The first step is to
carry out a differential balance around a spherical shell in the fissionable material. In this case, there
is a generation of energy term that will be equal to the difference between the heat flux into and out
of the differential shell. Thus, the balance yields
0 = (4r2 qrf )|r (4r2 qrf )|r+r + 4r2 rSn

(4.53)

where the volume of the differential element is given by 4r2 r. Dividing the above equation by
4r gives
(r2 qrf )|r (r2 qrf )|r+r
= Sn r 2
r0
r
lim

48

(4.54)

and taking the limit as r goes to zero gives a differential equation for the heat flux within the
fissionable material

 r 2 
d 2 f
r2 .
(4.55)
(r qr ) = Sn0 1 + b
f
dr
R
Integrating the above equation with respect to r gives


r
b r3
C1f
f
qr = Sn0
+
+ 2
3 Rf 2 5
r

(4.56)

where C1f corresponds to an integration constant from the spherical shell balance within the fissionable material. In the aluminum cladding, there is no generation of energy term, so a spherical shell
balance gives
d 2 C
(r qr ) = 0
dr

(4.57)

and upon integration with respect to r


C1C
r2

qrC =

(4.58)

where C1C is the integration constant for the heat flux in the cladding.
The integration constants need to be evaluated using the boundary conditions. The first boundary
condition to implement is that the heat flux is finite at r equal to 0. This requires that the integration
constant C1f is equal to zero, such that


r
b r3
f
+
(4.59)
qr = Sn0
3 Rf 2 5
To evaulate C1C we require that the heat fluxes are equal (q C = q f ) at r = Rf . This gives
!
f3
b
R
+
C1C = Sn0
3
5
and upon substitution into Eq. 4.58 we get the heat flux in the cladding to be

 3
1 b Rf
C
qr = Sn0
.
+
3 5 r2
The next step is to apply Fouriers law to the fissionable material to obtain


f
r
b r3
f dT
= Sn0
+
k
dr
3 Rf 2 5

(4.60)

(4.61)

(4.62)

and for the aluminum cladding to obtain


k

C dT

dr


= Sn0

1 b
+
3 5

Rf
r2

(4.63)

where k f and k C correspond to the conductivities in the fissionable material and the cladding, respectively. Integrating to get the temperature profile in the fissionable material, we have


Sn0 r2
b r4
f
(4.64)
T = f
+ f2
+ C2f
k
6
R 20
49

Similarly the temperature profile in the cladding is obtained from indefinite integration

 3
Sn0 1 b Rf
C
T = C
+
+ C2C .
k
3 5
r

(4.65)

In order to determine the integration constants, we use the boundary conditions that at r = Rf ,
T f = T C and that at r = RC , T C = T0 where T0 is the temperature at the outside surface of
the cladding. Solving for the integration constants and inserting the resulting expressions into the
temperature profile in the fissionable material and in the cladding gives the final expressions of

2 
 r 2 
 r 4 
Sn0 Rf
3
f
+ b 1
T =
1
6k f
Rf
10
Rf




2
Sn0 Rf
3
Rf
+
1
+
b
1

+ T0
(4.66)
3k C
5
RC
and
T

Sn0 Rf
=
3k C

3
1+ b
5



Rf
Rf
C
r
R


+ T0 .

(4.67)

The maximum temperature occurs in the center of the particle at r = R. Another quantity of interest
is the energy transferred from the particle to the surroundings.
Exercise Determine the energy transferred from the particle to the surroundings.

4.2.4

Heat conduction with a viscous heating source

In this section we consider the problem of heating fluids through viscous dissipation. This is especially important for fluids undergoing high shear rates (or fluids with large velocity gradients).
Problems of this nature occur in engines where lubricants need to be used between rapidly moving
parts, flow of polymers in extrusion processes, and the flow of air in boundary layers near to rockets
or satellites when at high velocities when they re-enter the earths atmosphere. Here we consider the
heating of a viscous fluid between two coaxial cylinders where the outer cylinder is moving with an
angular velocity equal to and the inner cylinder is stationary as shown in Figs. 4.4 and 4.5 below.
The surfaces of the cylinders are maintained at constant temperatures of T0 for the inner cylinder and
Tb for the outer rotating cylinder.
As the outer cylinder rotates, each cylindrical shell of fluid rubs against an adjacent shell of fluid.
This friction between adjacent layers produces heat, a process referred to as viscous dissipation,
which corresponds to the conversion of mechanical energy (from the rotating cylinder) to thermal
energy (reflected by the rise in temperature of the fluid.) The energy generated per unit volume per
unit time by viscous dissipation is denoted by Sv . Here we consider the simplified case when the
thickness of the fluid in the annular region is much smaller than the radii of the cylinders. This case is
especially relevant because it is the high velocity gradients that give rise to large amounts of viscous
dissipation. The high velocity gradients will occur for very thin fluid films between moving objects.
In this limit, we can neglect the effect of curvature and thus analyze the system shown in Fig. 4.5. For
the cartesian coordinate system we have already derived the expression for the viscous dissipation per
unit volume per unit time Sv (see Eq. 4.13).
For this system, the only non-zero component of the velocity corresponds to the z direction. The
z component of the velocity is also only a function of x. In this case, the only non-zero components
of the shear tensor correspond to zx and xz . We carry out a shell balance on a differential volume of
50

Figure 4.4: Flow between cylinders with viscous heat generation. That part of the system enclosed
within the dotted lines is shown in Fig. 4.5 below

Figure 4.5: A zoomed in view of the fluid located between the walls of the coaxial cyliners. When
the thickness of the fluid region is much smaller than the radius of the cylinder, the curvature of the
system can be neglected.
thickness x with the other dimensions given by the width W and length L of the channel. The shell
balance gives
W Lqx |x W Lqx |x+x + W Lvz xz |x W Lvz xz |x+x = 0

(4.68)

where we have neglected the term zx vx because vx = 0. Dividing through by the differential volume
W Lx and taking the limit as x goes to 0 gives the differential equation

dqx
dxz
vz
= 0.
dx
dx

(4.69)

The equation can be integrated with respect to x to give


qx + vz xz = C1

(4.70)

where C1 is an integration constant to be determined from the boundary conditions. Next we substitute
in Fouriers law of heat conduction
qx = k
51

dT
dx

(4.71)

and Newtons law of viscosity


xz =

dvz
dx

(4.72)

into Eq. 4.69 to give


k

dvz
dT
vz
= C1 .
dx
dx

(4.73)

If we neglect the temperature dependence of viscosity, the velocity gradient in the channel will be
linear (this is the planar couette flow problem discussed in the beginning of Section 2.2, the linear
velocity profile can be obtained by performing a momentum balance). In this case, the velocity
profile is given by vz = xvb /b and the derivative dvz /dx = vb /b, where vb = R is the velocity of
the top plate (or the outside cylinder). Substitution of these results gives
k

 v 2
dT
b
x
= C1 .
dx
b

(4.74)

The equation can be integrated with respect to x to obtain the temperature profile in terms of the
integration constants C1 and C2
T =

 vb 2 x2 C1

x + C2 .
k b
2
k

(4.75)

The integration constants are determined from the boundary conditions that at x = 0, T = T0 and at
x = b, T = Tb . Solving for the contants and substitution into the temperature profiles gives the final
result


T T0
1 x
x x
= Br
1
+ ,
(4.76)
Tb T0
2 b
b
b
where Br = vb2 /k(Tb T0 ) is the dimensionless Brinkman number, which is a measure of the
importance of the viscous dissipation term. For the case where Tb = T0 , the resulting temperature
profile is given by
T T0
1 vb2 x 
x
=
1
.
T0
2 kT0 b
b

(4.77)

When the Brinkmann number is very small, there will be no temperature gradient because the viscous
dissipation term is negligible.

4.2.5

Heat conduction in a cooling fin

Another practical application of heat conduction is the calculation of the efficiency of a cooling fin,
which is used to increase the overall rate of heat transfer during cooling processes. In heat exchangers,
fins are used to increase the area for heat transfer between metal walls and poorly conducting fluids
such as gases. Here. we consider the soluble problem corresponding to a simple rectangular fin as
shown in Fig. 4.6, where the wall temperature is Tw and the ambient gas temperature is given by Ta .
In order to solve the problem, we need to make various approximations. In the model, we assume
(1) temperature is only a function of the z coordinate, (2) the heat lost from the end or edges of the
fin is small compared to the heat lost from the surface of the fin (this is accurate because the area of
the fin surface is much greater than the edges or end areas), and (3) the heat flux at the fin surface is
given by the heat lost to the surroundings q = h(T Ta ), where the local heat transfer coefficient h
is treated as a constant.
52

Figure 4.6: A simple cooling fin with the thickness 2B much less than the length L and the width W .
The shell balance is carried out over a differential volume with thickness z and dimensions W
and L. We only need to consider the transfer of heat by conduction along the fin and the heat lost to
the surroundings. The shell balance gives
2BW qz |z 2BW qz |z+z h(2W z)(T Ta ) = 0.

(4.78)

Note that in the above equation, the rate of heat transfer to the surroundings (the term containing
q = h(T Ta )) is included in the shell balance. In previous problems, this term is incorporated into the
boundary condition. For the cooling fin problem, the heat transfer coefficient is included in the shell
balance because the differential volume element includes the surface exposed to the surroundings.
Dividing the shell balance equation by 2BW z and taking the limit as z goes to zero gives the
differential equation

h
dqz
= (T Ta ).
dz
B

(4.79)

Next we substitute in Fouriers law qz = k(dT /dz) to give


d2 T
h
=
(T Ta ).
2
dz
kB

(4.80)

The above equation is to be solved with the boundary conditions that at z = 0, T = Tw , and at z = L,
dT /dz = 0. The latter boundary condition follows from the assumption that no heat is lost through
the end of the fin.
In order to solve the differential equation, we can simplify the algeabra by introducing the following dimensionless quantities:
=

N2 =

T Ta
= dimensionless temperature
Tw Ta
z
= = dimensionless distance
L

hL2
= dimensionless hreat transfer coefficient
kB

(4.81)
(4.82)
(4.83)
(4.84)

53

Rewriting Eq. 4.80 in terms of the dimensionless quantities gives


d2
= N 2
d 2

(4.85)

where the boundary conditions are that = 1 when = 0 and d/d = 0 at = 1. The differential
equation has the general solution given by
= C1 exp(N ) + C2 exp(N )

(4.86)

where C1 and C2 are the integration constants that need to be determined from the boundary conditions. Applying the boundary conditions gives the following equations
1 = C1 + C2

(4.87)

d
|=1 = 0 = C1 N exp(N ) C2 N exp(N ).
d

(4.88)

and

Solving the above equations for the constants C1 and C2 and substituing the results into the temperature profile gives the final result that
exp [N (1 )] + exp [N (1 )]
exp(N ) + exp(N )
cosh [N (1 )]
=
.
cosh N

(4.89)

We can define an effectiveness factor for the fin which corresponds to the ratio of the actual heat
lost to the surrounding divided by a theoretical loss if the fin was at the same temperature as the wall
R1
RW RL
d
h(T

T
)
dzdy
a
0
= R0 1
(4.90)
= RW
R L0
d
h(Tw Ta ) dzdy
0
0
0
Carrying out the integration gives the effectiveness factor as

1
1
1
tanh N
=
sinh N (1 ) =
.
cosh N
N
N
0

54

(4.91)

4.2.6

Free convection

The last problem we will deal with in this section is concerned with free convection. Free convection
corresponds to the situation where temperature gradients cause gradients in density, which in turn,
causes convection. A good example is the heat transfer through double-glazed windows. During
sunny weather, free convection is caused by sunlight heating up the outside surface of the window
giving rise to a temperature difference between the inside and outside glazing. This causes the gas to
rise on the outside surface and to descend on the inside surface of the window. The free convection
will determine the heat transfer coefficient that is used to describe the rate of heat loss through the
window.
Free convection has a much larger effect on the flow profile of a gas versus a liquid because gas
density is much more sensitive to changes in temperature. Thus we consider a gas flowing between
two flat plates at different temperatures T1 and T2 as depicted in Fig. 4.7. The separation between the
plates is equal to 2B. The gas has a density equal to and a viscosity . The plate at y = B is
heated to a temperature T2 , which is greater than the temperature T1 of the plate located at y = B. At
the hotter plate, the gas rises due to being at a lower density, whereas the colder gas descends. The
system is closed at the top and the bottom, so that the fluid is continuously circulating between the
plates. At steady state, there is no accumulation such that the total gas flowrate across the z-plane is
equal to 0. We assume that the plates are very tall such that end-effects can be neglected. In this case,
we can assume there is only a temperature gradient in the y-direction.

Figure 4.7: Laminar flow between flat plates driven by temperature gradients. The velocity is a cubic
function of the coordinate y.
The first step is to carry out a shell balance across a thin slab of fluid with thickness y in the
y-direction. Since there is no flow in the y-direction, transport of energy only occurs through thermal
diffusion or conduction. Applying Fouriers law to the shell balance will give the following result
k

d2 T
=0
dy 2

(4.92)

where k is the thermal conductivity of the gas. The temperature equation is to be solved with the
boundary condtions that at y = B, T = T2 and at y = B, T = T1 . The solution to the differential
equation is given by
y
1
(4.93)
T = T T
2
B
55

where the temperature difference is defined by T = T2 T1 and we define an average temperature


T = (T1 + T2 )/2.
Carrying out the momentum balance in the z-direction gives the differential equation
d2 vz
dp
2 =
+ g
dy
dz

(4.94)

where we assume the viscosity is constant and does not depend on temperature. The main reason for
the temperature-driven flow is due to differences in density of the gas as a function of temperature.
Thus we need to take into account how density changes with temperature. In order to do that, we
consider the problem where the temperature different T is small in which case we can carry out a
taylor-series expansion for the density in terms of the average temperature T . Thus, we obtain the
following equation when considering only the first order term in the taylor-series expansion

| (T T )
T T
= (T T )

= |T +

(4.95)

where and are the density and the coefficient of volume expansion evaluated at temperature T .
The coefficient of volume expansion is defined as


 
1 V
1
=
=
(4.96)
V T p
T p
Introducing the taylor series expansion for the gas density into the equation of motion gives

d2 vz
dp
=
+ g g(T T ).
2
dy
dz

(4.97)

The last term on the right side of the equation accounts for the effect of temperature on the gravitational force, which gives rise to the thermally driven motion. Thus this term is referred to as a
buoyancy force. Substituting the temperature distribution determined from the energy balance (Eq.
4.93) gives


dp
1
y
d2 vz
+ g + gT
(4.98)
2 =
dy
dz
2
B
The equation can be rearranged to give
y
d2 vz 1
2 gT =
dy
2
B

dp
+ g
dz


(4.99)

The left side of the equation only depends on the y-coordinate, whereas the right side only depends
on the z-coordinate. Thus both sides of the equation must be equal to a constant. The differential
equation can be solved by applying the no-slip boundary conditions at the walls (i.e. vz = 0 at
x = B and at x = B). The solution is given by



  

gT B 2  y 3  y 
B 2 dp
y 2
vz =

+
+ g
1 .
(4.100)
12
B
B
2 dz
B
The equation for the velocity distribution can be further reduced by applying the constraint that
the total mass flow rate in the z-direction must be equal to 0 (since the system is closed at the top and
the bottom of the plates). Thus, we require that
Z +B
vz dy = 0.
(4.101)
B

56

Substituting the expression for vz into the above integral and neglecting all terms which contain the
square of T gives
dp
= g.
dz

(4.102)

The result indicates the pressure distribution in the system is only due to the average weight of the
fluid. The velocity distribution thus reduces to


gT B 2  y 3  y 
.
(4.103)
vz =

12
B
B
The average velocity in the upward moving direction can be calculated to be
hvz i =

gT B 2
.
48

(4.104)

The velocity distribution can also be rewritten using a dimensionless velocity given by vz = Bvz /
and a dimensionless coordinate y = y/B.
vz =

1
Gr(
y 3 y)
12

(4.105)

Gr corresponds to the dimenionless Grashof number, which is defined as


Gr =

2 gT B 3
2

(4.106)

and arises in the dimensional analysis for the equations of change for non-isothermal systems. The
Grashof number represents the ratio of the buoyancy force to the viscous forces. Analogous to the
Reynolds number, large values of the Grashof number lead to turbulent flow conditions in fluids
undergoing free convection.

57

Lecture 5
Turbulence
5.1

Introduction

In turbulent flows, the various properties of the fluid, such as the velocity, pressure, etc., vary in
a very complicated manner with position and time. The instantaneous values of these properties
appear to behave almost randomly with time. Luckily, however, we are generally not interested in the
instantaneous properties of a fluid flow; rather, we want information on how the average properties
vary. These average properties vary in a regular manner and can be modeled (at least approximately)
even for a turbulent flow.

5.2

Brief overview of statistics

Before we can proceed, we need to define precisely what is meant by an average property. Consider
the situation where we are trying to perform measurements on a turbulent flow. In order to obtain
statistics for the measurement, we repeat the experiment N times. The time variation of the property
f that we are measuring (e.g., pressure, velocity, etc.) will be different each time we measure it, due
do slight differences in the initial conditions of the experiments. The average variation of the property
f with time, which we denote as hf (t)i, is defined as
N
1 X
hf (t)i =
fi (t)
N i=1

(5.1)

where fi (t) is the time dependence of f measured during the ith experiment.
For a given experiment, the quantity f can always be written as a sum of two terms:
f (t) = hf (t)i + f 0 (t)

(5.2)

where the first term is the average value of f , and the second term is the deviation of the quantity
from its average value (note that in this case, f 0 is not the derivative of f ). If we take the average of
both sides of the equation, we find
hf (t)i = hf (t)i + hf 0 (t)i
hf 0 (t)i = 0

(5.3)

which is just a statement of the fact that the average deviation of a property from its mean is zero.
We will make frequent use of this fact shortly. Although the average of f is zero, the average of f 2 is
generally not (i.e., hf 2 i =
6 0). The larger the value of hf 2 i, the larger the fluctuations in the property.

58

5.3

Reynold stresses

The various properties of a fluid can be written as the sum of an average value and an instantaneous
fluctuation from the average
vi = hvi i + vi0

(5.4)

ij = hij i + ij0

(5.5)

p = hpi + p0

(5.6)

Note that for a laminar flow, the fluctuations of the various properties are identically equal to zero (i.e.
vi0 = ij0 = p0 = 0).
We are not directly interested in the instantaneous values vi , ij , and p of the flow, which are governed by the Navier-Stokes equation. We are only concerned in their average values (i.e., hvi i,hij i,
and hpi). So we need to develop the equations that govern the average properties. To do this, we just
insert Eqs. 5.4 to 5.6 into the Navier-Stokes equations.
For the continuity equation, this becomes:
vi
=0
xi

(hvi i + vi0 ) = 0
xi

(5.7)

Taking the average of both sides of the equation, we find


hvi i
=0
xi

(5.8)

where we have used the fact that hvi0 i = 0. This is the same as the original equation, but now involves
the average velocity rather than the instantaneous velocity.
The relationship between the stress and the velocity gradients becomes:


vi
vj
ij =
+
xj xi



0
0
0
hij i + ij =
(hvi i + vi ) +
(hvj i + vj )
(5.9)
xj
xi
Taking the average of both sides of this equation, we find


hvi i hvj i
hij i =
+
xj
xi

(5.10)

where we have used the fact that hf 0 i = 0. Again, we find that this is the same as the original equation,
but with the instantaneous values replaced by their averages.
For the momentum equation, we find
vi

ji
p
=
(vj vi )

+ gi
t
xj
xj
xi

(hvi i + vi0 ) =
(hvj i hvi i + vj0 hvi i + vi0 hvj i + vj0 vi0 )
t
xj

(hij i + ij0 )
(hpi + p0 ) + gi

xj
xi
59

(5.11)

Taking the time average of both sides of the equation, we find


hvi i
hji i hpi

(hvi i hvj i + hvi0 i vj0 )

+ gi
=
t
xj
xj
xi

hpi

hvi i hvj i
(hji i + vi0 vj0 )
+ gi
=
xj
xj
xi
hpi

(t)
hvi i hvj i
(hji i + ji )
+ gi
=
xj
xj
xi

(5.12)

(t)

where ij is known as the Reynolds stress and is given by


(t)
ij = vi0 vj0

(5.13)

For the momentum equation, unlike the previous two equations, we find that instantaneous properties
are not simply replaced by their average values. We actually have an extra stress term. This additional
(t)
stress ij is due to the turbulent motion of the fluid.
The average total stress ij within a fluid can, therefore, be decomposed into two contributions:
(t)
(i) viscous forces hij i and (ii) turbulent eddies ij
(t)

ijtotal = hij i + ij

(5.14)

(t)

When the flow is laminar ij = 0.

Figure 5.1: Typical velocity profile for a system in turbulent flow conditions.
In Fig. 5.1, we plot a typical velocity profile of a fluid in turbulent flow near a wall. The flow
can be divided into three general regions: (i) the laminar sublayer (or viscous sublayer), (ii) the buffer
layer, and (iii) the turbulent core. In the laminar sublayer, the flow is relatively slow, due to the no-slip
boundary conditions imposed by the nearby wall. In this region of the flow, the level of turbulence is
extremely low, and viscous stress is much greater than turbulent stress.
60

Away from the wall, however, the Reynolds stress dominates over viscous stress. This region
is known as the turbulent core. In the region between the laminar sublayer and the turbulent core,
viscous transport and turbulent transport of momentum play roughly equal roles. This region of the
flow is known as the buffer layer. In general for a system in turbulent flow, most of the fluid is in the
turbulent core. The laminar sublayer is typically only a thin slab near the wall; the thickness of this
layer varies inversely with the Reynolds number.

(t)

Figure 5.2: Relative contributions of viscous hrz i and Reynolds rz stresses as a function of radial
position in a pipe.
In order to get a feel for the relative contributions of the viscous and Reynolds stresses, we show
a typical distribution of stress for the turbulent flow of a fluid through a pipe of radius R in Fig. 5.2.
Near the wall of the pipe (located at r = R), the viscous stresses dominate the flow. It is only when
we move away to the pipe wall, do the turbulent stresses begin to make a significant contribution.
The size of the region where the Reynolds stresses are significant depends on Reynolds number of
the flow. The higher the Reynolds number, the larger the region. For low Reynolds number where the
flow is laminar, the Reynolds stresses make no contribution.

5.4
5.4.1

Approximate models for turbulence


Boussinesq hypothesis or the mean-velocity field closure

According to Newtons law of the viscosity, the shear stress is related to the velocity gradients


hvi i hvj i
+
(5.15)
hij i =
xj
xi
Boussinesq suggested that the Reynolds stress can be described in an analogous manner to the viscous
stresses:


hvi i hvj i
(t)
(t)
ij
+
(5.16)
xj
xi
where (t) is the eddy viscosity, which plays an analogous role to the viscosity for viscous stresses.
The relation given in Eq. 5.16 is not exact. In many cases it is known to be a poor approximation;
however, in general, it provides a reasonable description of turbulence and is the starting point of
many theories.
Unlike the viscosity, the eddy viscosity is not simply a property of the fluid. That is, given the
temperature and pressure of a fluid, one cannot specify the value of the eddy viscosity, as is the case
for viscosity. The eddy viscosity is, in fact, dependent on the local flow conditions, and, consequently,
its value varies with the position in the flow.
61

If we consider flow in a circular pipe, then the total stress in the fluid can be written as:
(t)
total
hrz i + rz
= rz
hvz i
hvz i
total
(t)
= rz

r
r
hvz i
total
( + (t) )
= rz
r

(5.17)

where we have used the Boussinesq approximation for the Reynold stress. From this equation, we
can see that the ratio of the eddy viscosity to the molecular viscosity is a measure of the relative
contribution of viscous transport and turbulent transport of momentum. In Fig. 5.3, we plot (t) /
as a function of radial position for turbulent flow in a circular pipe of radius R. From this plot we
again see that viscous stresses dominate near the wall, while turbulent stresses dominate in most of
the region away from the wall.

Figure 5.3: Ratio of the eddy viscosity to the molecular viscosity as a function of radial position for a
fluid in turbulent flow in a pipe of radius R.
The Boussinesq hypothesis has essentially shifted the problem of determining the Reynolds stress
(t)
ij to determining the eddy viscosity (t) , which is a somewhat simpler problem since the eddy viscosity is a scalar quantity while the Reynolds stress is a tensor (matrix). However, the eddy viscosity
is still an undetermined function of position, and so our description of turbulence is still incomplete.

5.4.2

Prandtl mixing-length theory and von Karman similarity hypothesis

In order to develop a simple model for the Reynolds stresses, lets examine a turbulent flow with an
average flow in the x-direction (see Fig. 5.4). In turbulent flow, the main structures are swirling eddies
that spontaneously appear and then disappear due to dissipation. Prandtl assumed that these eddies
would transport a subsection of fluid in the flow a distance l, known as the mixing length, before they
disappear. For example, a piece of fluid that was located at position y + l, with a velocity given by
the average velocity of the flow hvx (y + l)i would be transported by the eddy to a position y, where
the average velocity of the fluid is hvx (y)i. Immediately after being transported, the velocity of this
62

Figure 5.4: Illustration of Prandtls mixing length theory.


piece of fluid would differ from the average velocity of the surrounding fluid at position y. From this
physical argument, we would expect the order of magnitude of the fluctuations in the x-component of
the velocity vx0 to be
vx0 hvx (y + l)i hvx (y)i
hvx (y)i
l
y

(5.18)

where we have used the Taylor series expansion. The fluctuations in the y-component of the velocity
are expected to be of the same order of magnitude (i.e., vy0 vx0 ).
Therefore we expect

2

0 0
vx
2
(5.19)
vy vx l
y
and so Prandtls expression for the Reynolds stress is


(t)
yx
= vy0 vx0




2 hvx (y)i hvx (y)i
.
= l
y
y

(5.20)

Comparing this to the Boussinesq expression for the Reynolds stress, we can make the identification:




(t)
2 hvx (y)i
= l
(5.21)
y
We still have the difficulty of determining how the mixing length l depends on position. This mixing
length physically represent, more or less, the size of the turbulent eddies. Away from the walls, we
expect these can be fairly large; near a wall, however, these eddies are limited in size. von Karman
suggested that the mixing length is proportional to the distance y from a surface.
l = y

(5.22)

where is an empirical constant, which has been found to be approximately equal to 0.4. Note that
the constant is independent of the geometry of the flow. With this relation, we now have a complete
(although fairly crude) description of turbulent flow.
63

5.4.3

Three-region model for momentum transport (universal velocity profile)

In this subsection, we will discuss the universality of the velocity profile of a turbulent flow. What we
mean by this is that the velocity profile is to a large extent independent of the geometry of the flow
(e.g., pipe diameter) and only depends on a few properties of the fluid.
To begin, we examine the velocity profile of a fluid confined between two planes separated by a
distance H. The bottom plane is stationary, while the upper plane is moving at a speed U in the x
direction. In order to maintain the motion of the upper plane, a stress w must be applied. For this
system, the flow is driven entirely by the motion of the upper plane, and there is no pressure gradient.
We perform a force (momentum) balance on a small slab of fluid of thickness dy, located at y
above the bottom plane. The system is at steady state, so the forces acting on the slab must sum to
zero.
total
total
0 = yx
(y + dy)A yx
(y)A

(5.23)

Dividing through by A dy and taking the limit as dy 0, we find


total
yx
=0
y

(5.24)

where we have made use of the fact that the cross-subsectional area A is constant. This equation can
be integrated to give
total
yx
= w

(5.25)

where the integration constant w is the shear stress at the wall. This equation states that the total
shear stress in the problem is independent of position. For a system in turbulent flow condtions, the
total shear stress is made up of two contributions: (i) a laminar contribution hyx i, and (ii) a turbulent
(t)
(t) contribution yx , the Reynold stresses. Therefore, we have
(t)
hyx i + yx
= w

(5.26)

Before we proceed any further in analyzing this problem, we need to identify the characteristic scales
in the system. These characteristic scales determine how large or small a quantity is, with respect
to the system. For example, we say that we are at a position far from the wall, this means that the
distance from the wall is much greater than the characteristic length of the system.
The characteristic scales are based upon the physical properties of the system; in this case, these
are the wall shear stress w , the fluid density , and the fluid viscosity . From these properties, we
find that the characteristic velocity V is
 1/2
w
V =
(5.27)

and the characteristic length l is


l=

.
w

(5.28)

Very close to a wall in the laminar sublayer (when the distance from the wall is much less than the
characteristic length), viscous transport of momentum dominates over turbulent transport. In this
case, the force balance simplifies to:
hyx i w
hvx i

w .
y
64

(5.29)

Integrating this equation, we find


hvx i

w
y

(5.30)

where the integration constant is equal to zero, due to the fact that the velocity must vanish at the wall.
Introducing the dimensionless velocity hvx+ i = hvx i /V and the dimensionless distance y + = y/l, we
find

+
vx y +
(5.31)
In this form, we see that the velocity profile in the laminar sublayer is actually independent on the
nature of the fluid (i.e., the properties of the fluid) and the nature of the flow.
In the turbulent core, far from any walls, the Reynolds stress is much larger than the viscous stress
(t)
(i.e., yx >> hyx i) Therefore, the viscous stress can be neglected.
(t)
yx
w

(5.32)

If we use the Boussinesq hypothesis along with the Prandtl mixing-length expression for the eddy
viscosity, we find
hvx i
(t)
w
y

2
hvx i
2
l
w
y
2

hvx i
2
w
(y)
y
 1/2
1 w
hvx i

y
y


1
+
+
hvx i V
ln y + C

+ 1
vx ln y + + C +

(5.33)

where C + is an integration constant. Experimentally, we find = 0.4 and C + 5.5. Again, we find
that the velocity profile is independent of the nature of the fluid and the nature of the flow.
In summary, there are three different regions in turbulent flow past a surface. These are the
laminar sublayer, the turbulent core, and the buffer layer (which interpolates between the previous
two regions). The velocity profiles in each of these layers can be written in a universal form, which is
independent of the nature of the fluid and the details of the flow:
laminar sublayer v + = y +
buffer layer v + = 5 ln y + 3.05
turbulent core v + = 2.5 ln y + + 5.5

5.5

for 0 < y + < 5


for 5 < y + < 30
for 30 < y +

Turbulent flow in a pipe

In this subsection, we will use the universal velocity profile developed in the previous subsection to
get a relation between the friction factor and the Reynolds number for turbulent flow in a circular
pipe with radius R. The thickness of the laminar sublayer is inversely proportional to the Reynolds
65

number. Therefore, at very high flowrates, nearly all of the flow in a pipe should be occupied by the
turbulent core. This means we can use the expression for the velocity profile obtained for the turbulent
core as the velocity profile for the entire flow in the pipe.
Z R
1
vx =
2rdr hvx (r)i
R2 0


Z R
1 y
+
2
(5.34)
= 2R
ln + C
rdrV
l
0
The coordinate y represents the distance from the wall. This is related to the coordinate r in the pipe
by the relation y = R r. Substituting the variable y, we find


Z
1 y
2V 0
+
(R y)(dy)
ln + C
vx = 2
R R
l


Z 1
1 R
+
= 2V
(1 ) d
ln
+C

l
0


3/2
1 R
+
ln + C
(5.35)
=V

l
This expression can be written in a more familiar form by introducing the friction factor f and the
Reynolds number Re. The Reynolds number is defined as
Re =

2Rv x
Dv x
=

(5.36)

p D
2v 2x L

(5.37)

where D = 2R is the diameter of the pipe.


The friction factor is defined as
f=

This can be expressed in terms of the shear stress at the wall w . At steady state, the pressure force
acting on the fluid should be precisely equal to the shear stress on the fluid due to the wall of the pipe.
p

D2
= DLw
4
 
L
p = 4
w
D

(5.38)

Substituting this into the definition of the friction factor, we get


f=

2w
v 2x

(5.39)

If we insert the definition of the friction factor and the Reynolds number into Eq. 5.35, we find
r
1
3/2
1 Re f
p
= ln
+ C+

2
2

f /2
p
1
= 4.06 log Re f 0.60
(5.40)
f
This relation was first derived by von Karman. Nikuradse empirically fit data from a series of flow
experiments and found the following relation
p
1
= 4.0 log Re f 0.40
(5.41)
f
So we see that the mixing length theory of Prandtl provides a reasonable description of turbulent pipe
flow.
66

Lecture 6
Analogies in momentum, heat, and mass
transfer
6.1

Introduction

In the previous sections, we have seen the equations that govern the transport of momentum, energy,
and mass. If we look more closely at these equations, we can see strong analogies between the three
types of transport.

6.2

Comparison of diffusive transport

To begin with, we will compare diffusive transport. Lets first look at the change in the velocity profile
around a plate immersed in an initially stationary liquid. At time t = 0, the plate is instantaneously
accelerated to a velocity V . We are interested in how the velocity profile changes as a function of
time. For this problem, the momentum balance equation reduces to:
vx
2 vx
= 2
t
y

(6.1)

where = / is the kinematic viscosity, which has units of length squared over time (e.g., m2 s1 ).
The boundary conditions are that vx = V at y = 0 and vx = 0 as y . The initial condition is that
vx = 0 when t = 0. The solution of this equation is given by



y

vx (y, t) = V 1 erf
2 t



vx (y, t)
y

(6.2)
= 1 erf
V
2 t
where erf is the error function, which is defined as
Z x
2
erf(x) =
dt exp(t2 ).
0

(6.3)

Now, lets look at the change in the temperature of a plate that is suddenly immersed in a pool of
liquid initially at a uniform temperature T0 . The temperature of the plate is kept constant at T1 . If
there is no convection in the fluid (i.e., all heat transport is due to conduction), then the governing
equation is
Cv

T
2T
=k 2
t
y
67

(6.4)

where = k/(Cv ) is the thermal diffusivity, which also has unites of length squared over time. The
boundary conditions are T = T1 at y = 0 and T = T0 as y . The initial conditions is T = T0
when t = 0. The solution of this equation is given by


T (y, t) T0
y

= 1 erf
.
(6.5)
T1 T0
2 t
Finally, lets look at the change in the concentration profile of a dye A about a plate (which is
saturated with the dye) that is suddenly immersed in a pool of pure liquid B. The concentration of A
on the plate remains constant at cA0 . The governing equation for this problem is
cA
2 cA
= DAB 2
t
y

(6.6)

where DAB is the diffusion coefficient of species A in B, which again has units of length squared over
time. The boundary conditions are cA = cA0 at y = 0, and cA = 0 as y . The initial condition is
cA = 0 when t = 0. The solution of this equation is given by


y
cA (y, t)

= 1 erf
(6.7)
cA0
2 DAB t
If we compare Eqs. 6.2, 6.5, and 6.7, we can see that they are essentially identical, if we make
the identification DAB . In the first case, momentum diffuses from the plate into the bulk
fluid; in the second case, thermal energy diffuses into the bulk fluid, while in the third case, species A
diffuses into the bulk fluid. As time progresses, the property (i.e., momentum, heat, or mass) slowly
diffuses into the bulk of the system. The distance from the wall at which this property is significantly
different from the bulk value varies with time as:

t
for momentum

t
for energy
p
DAB t for mass
The relative rate at which heat diffuses with respect to momentum is given by the Prandtl number
Pr =

Cp
=
.

(6.8)

The relative rate at which mass diffuses with respect to momentum is given by the Schmidt number
Sc =

6.2.1

.
DAB

(6.9)

Reynolds analogy

In the previous subsection, we compared the diffusive transport of momentum, heat, and mass. In this
subsection, we will compare the transport of these properties under turbulent flow conditions. In the
fully turbulent region, the turbulent stresses dominate the momentum transport and the eddy thermal
conductivity dominates heat transport. In this situation, the momentum balance is approximately
given by
(t)
yx
w

2
hvx i
l2
w
y

68

(6.10)

The energy balance is approximately given by


q (t) qw




hT i
2 hvx i
Cv l
qw
y
y

(6.11)

Dividing the energy balance by the momentum balance yields


Cv

qw
hvx i /y
=
hT i /y
w
qw
hT i
=
hvx i
Cv w

(6.12)

If we assume that the entire fluid can be considered turbulent, even near the wall, then this equation
can be integrated from the bulk (where the temperature is T , and the velocity is V ) to the wall (where
the temperature is Tw , and the velocity is zero)
qw
T Tw =
V.
(6.13)
Cv w
The heat transfer coefficient h on the wall is defined by the relation
qw = h(Tw T ).

(6.14)

If we then introduce the Nusselt number


hL
(6.15)
k
where L is some characteristic length of the system, such as the pipe diameter, the Reynolds number
Nu =

Re =

LV

(6.16)

f=

2w
V 2

(6.17)

and the friction factor:

then we can arrange Eq. 6.13 to the form


Nu
f
= .
(6.18)
Pr Re
2
The relation is known as the Reynolds analogy and has been found to agree well with experimental
dat for systems where Pr 1. It allows us to predict heat transfer coefficients from knowledge of the
system dimensions and the properties of the fluid.
The Reynolds analogy provides a similar expression for mass transfer. If we consider a component
A in a system that is diffusing from the bulk fluid, which is at concentration cA, , to a surface surface,
which is at concentration cA,w , then the flux of A is given by
NA = kc (cA,w cA, )

(6.19)

where kc is the mass transfer coefficient. Then the Reynolds analogy for mass transfer is
Sh
f
=
Sc Re
2
where Sh is the Sherwood number, which is defined as
Sh =

kc L
DAB

Similarly to the case of heat transfer, this relationship is only accurate when Sc 1.
69

(6.20)

(6.21)

Você também pode gostar