You are on page 1of 168

1

Notes for the 2nd revised edition of


TRANSPORT PHENOMENA
by
R. B. Bird, W. E. Stewart, and E. N. Lightfoot
by R. B. Bird
9 Aug 2011
These "Notes," begun in 2009, are intended for use by students
and instructors to fill in the missing steps in some of the derivations
in the text. Comments and corrections will be greatly appreciated.
Also, if there are places in the text where you feel additional
explanation is needed, it would be appreciated if you would let us
know.
Many of these notes involve the Leibniz formula for
differentiating integrals, the hyperbolic functions, the error function,
Taylor series, and the gamma functionsall topics for which many
undergraduates have received inadequate instruction. These topics
are all reviewed in Appendix C.
Pages in the text for which "Notes" have been prepared:
p. 5 Conservation laws in binary collisions
p. 18 Evaluation of stress-tensor components
p. 26 Dimensional consistency in fluid dynamics
p. 35 Convective momentum flux
p. 50 Flow in tubes with elliptical cross section
p. 51 Flow of kinetic energy in tubes
p. 52 Conduits with circular and triangular cross sections
p. 54 Velocity distribution in annular flow
p. 55 Limiting cases of annular flow
p. 58 Flow of immiscible fluids
p. 59 Flow around a sphere
p. 78 Normal stresses at solid surfaces
p. 81 Equation of change for mechanical energy
p. 82(i) Proof that

t:Vv
( ) is positive for Newtonian fluids
p. 82(ii) Conservation equation for angular momentum
p. 86(i) Vector identity needed for the Bernoulli equation
p. 86(ii) Another way to look at the Bernoulli equation
2
p. 90(i) Tangential flow between cylinderspressure distribution
p. 90(ii) Torque balance
p. 117 Slope of the complementary error function
p. 121 Fluid motion near an oscillating plate
p. 123 Equation for the stream function
p. 125 Alternative method of getting Stokes' law
p. 139 The Falkner-Skan equation
p. 154 Average velocity in turbulent flow in tubes
p. 170 Turbulent flow in a circular jet
p. 198 Derivation of macroscopic balances
p. 199 Efflux from a spherical tank
p. 202 The lawn sprinkler
p. 218 Unsteady flow from a cylindrical tank
p. 241 The Weissenberg-Rabinowitsch equation
p. 242 Power-law flow in circular tubes
p. 248 Viscoelastic flow near an oscillating plate
p. 250 The corotational Maxwell model
p. 255 Polymer flow analyzed with a FENE-P model
p. 259 The Casson equation
p. 275 Dimensional consistency in heat transfer
p. 286 Enthalpy of an ideal monatomic gas
p. 299 Temperature profile in flow with viscous heating
p.309 Checking the cooling fin solution
p. 315 Temperature profile in tube flow
p. 337 Alternative equation of change for temperature
p. 341 The equation of change for entropy
p. 343 Tangential annular flow with viscous heating
p. 346 Temperature profile for transpiration cooling
p. 352 Stationary shock wave velocity distribution
p. 375 One-dimensional time-dependent heat conduction
p. 377 Two solutions for the slab heating problem
p. 379 Unsteady heat conduction with sinusoidal heating
p. 386 Steady-state potential flow of heat in solids
p. 388 Boundary layer flow with heat transfer
p. 413 Turbulent flow in tubes with heat transfer
p. 415 Turbulent flow in circular jets with heat transfer
p. 454 Derivation of macroscopic energy balance
p. 494 Planck's radiation law and Wien's displacement law
p. 529 Dimensional consistency in diffusion
p. 534 Binary formulas from multicomponent formulas
p. 535 Two formulations of Fick's law of diffusion
p. 547(i) Diffusion through a stagnant gas film
3
p. 547(ii) Taylor expansion of diffusion problem result
p. 555 Diffusion with chemical reaction
p. 557 Simplifying Eq. 18.4-18
p. 563(i) Diffusion with solid dissolution
p. 563(ii) Evaluation of mass flux from concentration profiles
p. 565 Verification of solution of diffusion problem
p. 584 Form of diffusion equation in molar units
p. 585 Diffusion with convection and chemical reaction
p. 589 Energy equation for multicomponent mixtures
p. 590 Simplification of the combined energy flux e
p. 591 Euler's theorem for homogeneous functions
p. 615 Time-dependent evaporation of a liquid
p. 622 Diffusion with time-dependent interfacial area
p. 626 Diffusion with chemical reaction
p. 628 Forced convection from flat plate
p. 692 Interaction of phase resistances
p. 766 Driving force in multicomponent diffusion
p. 767 Simplification of multicomponent diffusion result
p. 768 Relating Maxwell-Stefan and Fick diffusivities
p. 769 Illustrating interrelations between diffusivities
4
Note to p. 5
Section 0.3 is important for emphasizing some of the basic
concepts and definitions. Here we work through some of the missing
steps in Section 0.3, going from Eq. 0.3-3 to Eq. 0.3-4, and from Eq.
0.3-5 to Eq. 0.3-6.
(a) In Eq. 0.3-3, replace

r
A1
by

r
A
+ R
A1
and make analogous
replacements for

r
A2
,

r
B1
, and

r
B2
. We also let

m
A1
m
A2

1
2
m
A
. With
these substitutions, we then get:

1
2
m
A
r
A
+

R
A1
( )
+
1
2
m
A
r
A
+

R
A2
( )
+
1
2
m
B
r
B
+

R
B1
( )
+
1
2
m
B
r
B
+

R
B2
( )
(1)
But, according to the drawing in Fig. 0.3-2,

R
A1
R
A2
and,
analogously,

R
B1
R
B2
, so that

m
A
r
A
+ m
B
r
B
m
A
r
A
+ m
B
r
B
(2)
This is the law of conservation of momentum in terms of the
molecular masses and velocities.
(b) We start with Eq. 0.3-5, and replace

r
A1
by

r
A
+ R
A1
as above. We
also let

m
A1
m
A2

1
2
m
A
. Then Eq. 0.3-5 becomes:

1
2
1
2
m
A
r
A
r
A
( ) + 2 r
A

R
A1
( )
+

R
A1

R
A1
( )
( )

+
1
2
1
2
m
A
r
A
r
A
( ) + 2 r
A

R
A2
( )
+

R
A2

R
A2
( )

_
,

+
A

+
1
2
1
2
m
B
r
B
r
B
( ) + 2 r
B

R
B1
( )
+

R
B1

R
B1
( )
( )

+
1
2
1
2
m
B
r
B
r
B
( ) + 2 r
B

R
B2
( )
+

R
B2

R
B2
( )

_
,

+
B

1
2
1
2
m
A
r
A
r
A
( ) + 2 r
A
R
A1
( ) +

R
A1

R
A1
( )
( )

1
2
m
A
r
A
+

R
A1
( )
+
1
2
m
A
r
A
+

R
A2
( )
+
1
2
m
B
r
B
+

R
B1
( )
+
1
2
m
B
r
B
+

R
B2
( )

5

+
1
2
1
2
m
A
r
A
r
A
( ) + 2 r
A

R
A2
( )
+

R
A2

R
A2
( )

+
A

+
1
2
1
2
m
B
r
B
r
B
( ) + 2 r
B

R
B1
( )
+

R
B1

R
B1
( )
( )

+
1
2
1
2
m
B
r
B
r
B
( ) + 2 r
B

R
B2
( )
+

R
B2

R
B2
( )

+
B
(3)
The single-underlined terms just exactly cancel the doubly
underlined terms in the following line, because

R
A1
R
A2
and also

R
B1
R
B2
. Hence we get

1
2
m
A
r
A
r
A
( ) +
1
2
m
A1

R
A1

R
A1
( )
+
1
2
m
A2

R
A2

R
A2
( )
+
A

+
1
2
m
B
r
B
r
B
( ) +
1
2
m
B1

R
B1

R
B1
( )
+
1
2
m
B2

R
B2

R
B2
( )
+
B

1
2
m
A
r
A
r
A
( ) +
1
2
m
A1

R
A1

R
A1
( )
+
1
2
m
A2

R
A2

R
A2
( )
+
A

+
1
2
m
B
r
B
r
B
( ) +
1
2
m
B1

R
B1

R
B1
( )
+
1
2
m
B2

R
B2

R
B2
( )
+
B
(4)
In the first line of the equation above, the terms have the following
significance: Term 1 is the kinetic energy of molecule A in a fixed
coordinate system; Term 2 is the kinetic energy of atom A1 in a
coordinate system fixed at the center of mass of molecule A; Term 3 is
the kinetic theory of atom A2 in a coordinate system fixed at the
center of mass of molecule A; Term 4 is the potential energy of
molecule A as a function of

r
A2
r
A1
, the separation of the two atoms
in molecule A. The sum of terms 2 to 4 we call the "internal energy"

u
A
of molecules A, and Eq. 4 may be rewritten in the form of Eq. 0.3-
6.
This discussion of the collision between two diatomic
molecules is interesting, for several reasons. It shows how the idea of
"internal energy" arises in a very simple system. We encounter this
concept later in 11.1 where the terms "kinetic energy" and "internal
energy" are used in connection with a fluid regarded as a continuum.
When the fluid is regarded as a continuum, it may be difficult to
understand how one goes about splitting the energy of a fluid into
kinetic and internal energy, and how to define the latter. In
considering the collision between two diatomic molecules, however,
the splitting is quite straightforward.
6
Another point is that, having seen the need for splitting the
energy into two parts, one might be led to ask: why don't we need to
split the momentum into two parts in a similar way? Here again, for
the collision of diatomic molecules, the need for dividing the
momentum into two parts is not necessary.
The subject of transport phenomena is built up on the laws of
conservation of mass, momentum, angular momentum, and energy.
The application of these laws to the system of two colliding diatomic
molecules is relatively straightforward. However, when applying
them to a moving fluid, some notational problems arise associated
with the necessity of dealing with fluid bodies in three dimensions.
7
Note to p. 18
In Fig. 3B.2 there is shown a duct with cross-section of an
equilateral triangle. The height of the triangular cross-section is H,
and the side length is

2H 3 . We want to evaluate the viscous stress
tensor components for the incompressible flow in the z-direction, for
which the velocity in the z-direction is given as a function of x and y
in Eq. 3B.2-1:

v
z
x, y
( )

P
0
P
L
( )
4LH
y H
( )
3x
2
y
2
( )
(1)
and

v
x
0 and

v
y
0. Here L is the length of the duct (which goes
from z = 0 to z = L) and

P
0
P
L
is the difference in modified
pressure between the ends of the duct. What are the stresses at the
surface y = H according to Eq. 1.2-6?
For the velocity distribution given above, the nonzero
components are

yz

zy
and

xz

zx
. From Eq. 1.2-6, we get:

t
yz

v
z
y

P
0
P
L
( )
4LH

y
3x
2
y y
3
3Hx
2
+ Hy
2
( )

P
0
P
L
( )
4LH
3x
2
3y
2
+ 2Hy
( )
(2)

t
xz

v
z
x

P
0
P
L
( )
4LH

x
3x
2
y y
3
3Hx
2
+ Hy
2
( )

P
0
P
L
( )
4LH
6x y H
( ) ( )
(3)
At the surface y = H:

t
yz
yH

P
0
P
L
( )
4LH
3x
2
H
2
( )
;

xz
yH
0 (4,5)
8
Note to p. 26
It is very important to make a habit of checking equations for
dimensional consistency. Show that the following equations in the
text are dimensionally consistent: Eq. 1.4-14, Eq. 1.5-11, and Eq. 1.7-2.
Do this by replacing the symbols in the formulas by the dimensions
corresponding to the symbols in the table beginning on p. 872. Omit
any numerical factors that appear.
(a)

M
Lt


M
( )
ML
2
t
2
T

T
( )
L
2
( )

( )
(1)
(b)

M
Lt


1
moles

ML
2
t

L
3
moles

(2)
(c)

M
Lt
2


M
Lt
2

+
M
L
3

L
t

L
t


M
Lt
2

+
M
Lt
2

+
M
L
3

L
t

L
t

(3)
In each case, dimensional consistency is found. In (c) the unit tensor
is a dimensionless quantity.
9
Note to p. 35
In Fig. 3B.7 we show the flow into a slot of width 2B. Far away
from the slot, the velocity components are given by Eqs.3B.7-2 to 4.
The convective momentum flux tensor is given by

vv , where vv is the
dyadic product of v with v. The components of

vv in Cartesian
coordinates are shown in the last three columns of Table 1.7-1.
a. What are these components for the flow in Problem 3B.7?
b. What is the convective momentum flux through a plane
perpendicular to the x axis?
a. When we make use of the Cartesian components of the
velocity given in Eqs. 3B.7-2 to 4, we get

vv
( )
xx
v
x
v
x
+
1

2w
rW

_
,

2
x
6
x
2
+ y
2
( )
4
(1)

pvv
( )
xy
pv
x
v
y
+
1
p
2w
rW

_
,

2
x
5
y
x
2
+ y
2
( )
4
(2)

vv
( )
xz
v
x
v
z
0 (3)

pvv
( )
yx
pv
y
v
x
+
1
p
2w
rW

_
,

2
x
5
y
x
2
+ y
2
( )
4
(4)

pvv
( )
yy
pv
y
v
y
+
1
p
2w
rW

_
,

2
x
4
y
2
x
2
+ y
2
( )
4
(5)

vv ( )
yz
v
y
v
z
0 (6)

vv
( )
zx
v
z
v
x
0 (7)

vv ( )
zy
v
z
v
y
0 (8)

vv
( )
zz
v
z
v
z
0 (9)
b. For a plane perpendicular to the x axis, the vector n is

x
.
The components of the momentum flux are then
10

x
vv

1
]
x
v
x
v
x
+
1

2w
W

_
,

2
x
6
x
2
+ y
2
( )
4
(10)

x
vv

1
]
y
v
x
v
y
+
1

2w
W

_
,

2
x
5
y
x
2
+ y
2
( )
4
(11)

x
pvv

z
pv
x
v
z
0 (12)
Then

x
pvv

x
is the amount of x momentum flowing per unit time
through a unit area of surface perpendicular to the x axis,

x
pvv

y
is the amount of y momentum flowing per unit time through a unit of
surface perpendicular to the x axis, and

x
pvv

z
is the amount of z
momentum flowing per unit time through a unit of surface
perpendicular to the x axis.
Verify that the units of the expressions on the right sides of Eqs.
10 and 11 do indeed have the units of momentum per area per time.
11
Note to p. 50
Information about the flow in tubes of elliptical cross-section, and
comparison with the flow in a circular tube. Label the semi-major axis
of the ellipse as a and the semi-minor axis of the ellipse as b. The
cross-sectional area of the elliptical tube is

A rab.
The velocity distribution for laminar axial flow in a tube of
elliptical cross-section is

v
z

P
1
P
2
( )
a
2
b
2
2L a
2
+ b
2
( )
1
x
a

_
,

y
b

_
,

1
]
1
1
(1)
as given by J. Happel and H. Brenner, Low Reynolds Number
Hydrodynamics, Prentice-Hall, Englewood Cliffs, NJ (1965), p. 38.
When a = b = R, this velocity distribution reduces to Eq. 2.3-18 of BSL-
2e, since

x
2
+ y
2
r
2
.
To get the mass rate of flow, we integrate the velocity
distribution over the cross section, thus

w 4
P
1
P
2
( )
a
2
b
2

2L a
2
+ b
2
( )
ab 1
2
q
2
( )
dqd
0
1
2
[
0
1
[
Here the dimensionless variables

x a and

q y b have been
introduced.
Integration over q then gives

w 4
P
1
P
2
( )
a
3
b
3
p
2L a
2
+ b
2
( )
1
2
( )
q
1
3
q
3

1
]
0
1
2
0
1
[
d

4
2
3

P
1
P
2
( )
a
3
b
3
p
2L a
2
+ b
2
( )
1
2
( )
3 2
d
0
1

(2)
The integral may be found in H. B. Dwight, Tables of Integrals and
Other Mathematical Data, Macmillan, New York, 4th edition (1961),
Formula 855.41, and the result is
12

w 4
2
3

P
1
P
2
( ) a
3
b
3
p
2L a
2
+ b
2
( )

5
2
( )

1
2
( )
2 3
( )

4
2
3

P
1
P
2
( ) a
3
b
3
p
2L a
2
+ b
2
( )

3
2
( )
1
2
( )

1
2
( )

1
2
( )
2 2!

P
1
P
2
( )
a
3
b
3
p
4L a
2
+ b
2
( )
(3)
since

1
2
( )
. When a = b = R, this mass rate of flow expression
reduces to Eq. 2.3-21 of BSL-2e.
We can now compare the mass rates of flow for the circular
tube and the elliptical tube of the same cross sectional area, as follows
Circular tube:

w


P
1
P
2
( )
A
2

2
( )
p
8L
(4)
Elliptical tube:

w
ell

P
1
P
2
( )
A
2

2
( )
ab
4L a
2
+ b
2
( )
(5)
so that

w
ell
w

2
r
1 + r
2

_
,

(6)
where

r b a s 1. Thus for tubes of the same cross-sectional area, the
ratio

w
ell
w

decreases from unity (both tubes circular) to values less


than unity as r becomes smaller.
13
Note to p. 51
On this page several quantities are obtained from the velocity
distribution of Eq. 2.3-18. We could also ask ourselves: how much
kinetic energy is flowing per unit time in the axial laminar flow of a
fluid in a circular tube?
The volume rate of flow through an element of cross section

rdrd is

v
z
rdrd . The kinetic energy per unit volume of the fluid is

1
2
v
z
2
, since the only nonzero velocity component is

v
z
. Therefore the
total amount of kinetic energy per unit volume flowing through the
tube is:

1
2
pv
z
2
( )
v
z
rdrd
0
R
[
0
2r
[
2r
1
2
pv
z
2
( )
v
z
rdr
0
R
[

2r
1
2
p R
2
v
z
3
0
1
[
d

pR
2
v
max
3
1
2
( )
0
1
[
3
d (1)
In the second line, we have introduced the dimensionless coordinate

r R and the maximum velocity

v
max
P
0
P
L
( )
R
2
4L. Now all
that remains is to evaluate the integral:

pR
2
v
max
3
1 3
2
+ 3
4

6
( )
0
1
[
d pR
2
v
max
3
1
2

3
4
+
3
6

1
8

_
,


pR
2
v
max
3
4 6 + 4 1
8

_
,


1
8
pR
2
v
max
3

1
4
rR
2
v
max
( )
1
2
pv
max
2

_
,

(2)
The last form suggests a volume rate of flow multiplied by a kinetic
energy per unit volume, both quantities evaluated for the maximum
velocity.
14
Note to p. 52
The laminar flow in a circular tube with radius R is discussed in
2.3, and the laminar flow in tubes with equilateral triangular cross-
section of height H is described in Problem 3B.2. Both tubes have the
same length, L. We want to compare these two flow problems.
a. Compare the mass rates of flow for the two tubes when their
cross-sectional areas are the same.
b. Compare the mass rates of flow for the two tubes when the
perimeters of their cross sections are the same.
a. For flow in circular and triangular tubes we have for the
mass flow rates (see Eq. 2.3-21 and Eq. 3B.2(b)):

w


P
0
P
L
( )
R
4
p
8L

w
A

3 P
0
P
L
( ) H
4
p
320L
(1,2)
In Eq. 1, R is the tube radius; in Eq. 2, H is the height of the triangular
cross section, and

2H 3 is the length of a side of the triangle. To
make the comparison, we need to express the flow rates in terms of
the cross-sectional areas. Since for circular tubes

A

rR
2
, and for
equilateral triangular tubes,

A
A

1
3
H
2
,

w


r P
0
P
L
( )
A
2
r
2
( )
p
8L
;

w
A

3 3 P
0
P
L
( ) A
2
p
180L
(3,4)
Therefore

w
A
w

3 3 16
( )
180
8r 0.726 (5)
b. The perimeters of the two tubes are

P

2rR for circular


tubes, and

P
A
2 3H for triangular tubes. Therefore
15

w


r P
0
P
L
( )
P 2r
( )
4
p
8L

w
A

3 P
0
P
L
( )
P 2 3
( )
4
p
180L
(6,7)
Taking the ratio, as before

w
A
w

3
2 3
( )
4
180
( )

8 ( ) 2r ( )
4
r
0.265 (8)
For the square cross section (see Problem 3B.3), the ratios
corresponding to Eqs. 5 and 8 may be found to be

w

0.884 (same cross-sectional areas) (9)



w

0.545 (same perimeters) (10)


What, if any, conclusions can you draw from this problem?
[The triangular duct problem is discussed on p. 58 of Landau and
Lifshitz, Fluid Mechanics, Addison Wesley (1959); our H is their a
multiplied by

1
2
3 .]
16
Note to p. 54
Let's check a few things about Fig. 2.4-1.
a. There the graph shows the transport of z-momentum in the
positive r-direction. This quantity is negative when

r < iR, positive
when

r > iR, and 0 when

r iR, where i is defined by Eq. 2.4-12.
We need to verify that this graph is consistent with Eq. 2.4-13.
b. The figure also shows the velocity distribution for flow in an
annulus, as given in Eq. 2.4-14. What is the location of the maximum
velocity? Show that the position of the maximum is nearer the inner
cylinder of the annulus.
c. What is the velocity at the maximum in the curve?
a. Eq. 2.4-13 may be rewritten as

t
rz

P
0
P
L
( )
R
2L
r
R

_
,

i
2
R
r

_
,

1
]
1

P
0
P
L
( )
R
2L
1 i
2
R
r

_
,

1
]
1
1
R
r
(1)
If

r R < i , then the bracket in Eq. 2b.6-1 is negative, whereas if

r R > i , then the bracket in Eq. 2b.6-1 is positive. This is in
agreement with the graph of

rz
vs.

r in Fig. 2.4-1.
b. To find the maximum of the expression in brackets in Eq. 2.4-14, as
a function of

r R s , we have to differentiate
[ ]
with respect to s as
follows 9 (alternatively, one may set

rz
equal to zero):

d
ds
1 s
2

1 x
2
ln 1 x
( )
ln
1
s

1
]
1
0 2s +
1 x
2
ln 1 x
( )

_
,

1
s
(2)
Setting the right side equal to zero, and solving for

s gives

s
max

1 x
2
2ln 1 x
( )
(3)
Hence, choosing the plus sign (why?), we get for the location of the
maximum in the velocity curve
17

r
max
R
1 x
2
2ln 1 x
( )
(4)
The half-way point between the inner and outer cylinders is given by

s
1
2
(1 +x ) . Therefore we now have to prove that

1 x
2
2ln 1 x
( )
<
1
2
(1+x ) or

2 1 x ( )
1 +x
( )
ln 1 x
( )
< 1 (5a,b)
It is easy to calculate the left side of Eq. 5b as a function of x :
x

2 1 x ( )
1 +x
( )
ln 1 x
( )
x

2 1 x ( )
1 +x
( )
ln 1 x
( )
0.1 0.711 0.6 0.980
0.2 0.829 0.7 0.991
0.3 0.895 0.8 0.999+
0.4 0.935 0.9 0.999+
0.5 0.962
It is evident that for x > 0.8, the maximum is very close to being half
way between the two cylinders, and that is to be expected, inasmuch
as the annular-slit flow approaches a flat-slit flow. Problem 2B.5 gives
a discussion of the interrelation of the flow in a plane slit and the
flow in a narrow annulus.
c. The maximum velocity is then

v
z,max

P
0
P
L
( )
R
2
4L
1
r
max
R

_
,

1 x
2
ln 1 x
( )
ln
R
r
max

_
,

1
]
1
1

P
0
P
L
( )
R
2
4L
1
1 x
2
2ln 1 x
( )

1 x
2
ln 1 x
( )
ln
2ln 1 x
( )
1 x
2

1
]
1
1
(6)
This result may also be written in terms of i , as in Eq. 2.4-15.
18
Note to p. 55
It is good practice to check limiting cases whenever possible.
For example, one should show that Eq. 2.4-17 becomes the Hagen-
Poiseuille formula for tube flow (Eq. 2.3-21) in the limit that

x 0.
(See the comment in the paragraph that begins four lines after Eq. 2.4-
14.)
Solution: We have to show that the bracket quantity in Eq. 2.3-21
becomes equal to unity when

x 0. In this limit, the various terms
inside the bracket become:

lim
x0
1 x
4
( )
1 (1)

lim
x0
1 x
2
( )
2
1 (2)

l im
x0
1 x ( ) ~ (3)
Thus the bracket quantity becomes:

1 x
4
( )

1 x
2
( )
2
ln 1 x ( )

1
]
1
1
lim
x 0
1
1
~

1
]
1
1 (4)
and the Hagen-Poiseuille formula is recovered.
19
Note to p. 58
Let us verify that the average velocities in the two regions are
given by Eqs. 2.5-20 and 21.
In region I, the average velocity is given by

v
z
I

p
0
p
L
( )
b
2
2
I
L
1
b
2
I

I
+
II

_
,

+

I

II

I
+
II

_
,

x
b

_
,


x
b

_
,

1
]
1
1
dx
b
0
[

p
0
p
L
( )
b
2
2
I
L
2
I

I
+
II

_
,

+

I

II

I
+
II

_
,

1
]
1
d
1
0
[
where

x b

p
0
p
L
( )
b
2
2
I
L
2
I

I
+
II

_
,


1
2

I

II

I
+
II

_
,


1
3

1
]
1

p
0
p
L
( )
b
2
2
I
L
6 2
I
3
I

II
( )
2
I
+
II
( )

I
+
II

1
]
1
1

p
0
p
L
( ) b
2
2
I
L
7
I
+
II

I
+
II

1
]
1
(1)
Similarly for Region II we have

v
z
II

p
0
p
L
( )
b
2
2
II
L
1
b
2
II

I
+
II

_
,

+

I

II

I
+
II

_
,

x
b

_
,


x
b

_
,

1
]
1
1
dx
0
b
[

p
0
p
L
( )
b
2
2
II
L
2
II

I
+
II

_
,

+

I

II

I
+
II

_
,

1
]
1
d
0
1
[

p
0
p
L
( )
b
2
2
II
L
2
II

I
+
II

_
,

+
1
2

I

II

I
+
II

_
,


1
3

1
]
1

p
0
p
L
( )
b
2
2
II
L
6 2
II
+ 3
I

II
( )
2
I
+
II
( )

I
+
II

1
]
1
1

p
0
p
L
( ) b
2
2
II
L

I
+ 7
II

I
+
II

1
]
1
(2)
20
Note to p. 59
The verification of some equations requires a lot of algebraic
detail that falls in the category of "straightforward but tedious."
Nonetheless, such derivations should be done. Here are two
examples:
a. Verify the expressions for the stress components

t
rr
and

t
r
in Eqs.
2.6-5 and 2.6-6. From Eqs. B.1-8 and 2.6-1, we get

t
rr
2
ov
r
or
2v
~
1
R

_
,

+
3
2
R
r

_
,

3
2
R
r

_
,

1
]
1
1
cos

3v
~
R

R
r

_
,

2
+
R
r

_
,

1
]
1
1
cos (1)
And from Eq. B.1-11 and Eqs. 2.6-1 and 2, we find

t
r
r
o
or
v

_
,

+
1
r
ov
r
o

1
]
1
v
~
1
R

_
,

R
r

_
,


3
2
R
r

_
,

R
r

_
,

1
]
1
1
sin

v
~
1
R

_
,

R
r

_
,


3
2
R
r

_
,

1
2
R
r

_
,

1
]
1
1
sin
( )

+
3
2
v
~
R
R
r

_
,

4
sin

+
3
2
v
~
R
R
r

_
,

4
sin (2)
This is the "form drag" result given in Eq. 2.6-8.
b. Show how Eqs. 2.6-9 and 2.6-12 are obtained by doing the
necessary integrations.
To get the normal force acting on the sphere (from Eq. 2.6-7),
we start by noting that

t
rr
on the surface of the sphere is zero. (This is
a special case of the general result given in Example 3.1-1.) The
pressure p on the surface of the sphere is given in Eq. 2.6-8. Therefore
the z-component of the force acting normal to the surface of the
sphere is:
21

F
n ( )
p
0
+ pgRcos +
3
2
v
~
R
cos

1
]
1
0
r
[
0
2r
[
cos
( )
R
2
sinddo

2r p
0
+ pgRcos +
3
2
v
~
R
cos

1
]
1
0
r
[
cos
( )
R
2
sind

2rp
0
R
2
cos
0
r
[
sin d + 2rR
2
pgR +
3
2
v
~
R

_
,

cos
2
sin d
0
r
[
(3)
The first integral is zero, and the second is 2/3, so that

F
n ( )
2rR
2
pgR +
3
2
v
~
R

_
,

2
3

_
,


4
3
rR
3
pg + 2rRv
~
(4)
To get the z-component of the tangential force acting on the
sphere, we substitute Eq. 2.6-11 into Eq. 2.6-10 and integrate:

F
(t)
2r
3
2
v
~
R
sin

_
,

0
r

sin
( )
R
2
sin d

3rRv
~
sin
3

0
r

d 3rRv
~
4
3

_
,

4rRv
~
(5)
This is the "friction drag" result displayed in Eq. 2.6-12.
22
Note to p. 78
The general result for normal stresses at solid surfaces given in
Eq. 3.1-6 is very important. We have already seen in Eq. 2.6-5 that the
normal stresses for creeping flow around a sphere are exactly zero at
the sphere surface, r = R. Verify that Eq. 2.6-5 is correct for

t
rr
, t

,
and t

by using Eqs. B.1-15, 16, and 17 and Eqs. 2.6-1, 2, and 3.


Solution:
a. First, we get

t
rr
from Eq. B.1-15:

t
rr
2
ov
r
or
2v
~
1
R
+
3
2
R
r

_
,

2
3
1
2

_
,

R
r

_
,

1
]
1
1
cos

3v
~
R

R
r

_
,

2
+
R
r

_
,

1
]
1
1
cos (1)
b. Then we get t

from Eq. B.1-16:



t

2
1
r
v

+
v
r
r

_
,


2v
~
1
R

R
r

_
,

+
3
4
R
r

_
,

2
+
1
4
R
r

_
,

1
]
1
1
cos


2v
~
1
R
R
r

_
,


3
2
R
r

_
,

2
+
1
2
R
r

_
,

1
]
1
1
cos


2v
~
R

3
4
R
r

_
,

2
+
3
4
R
r

_
,

1
]
1
1
cos


3v
~
2R

R
r

_
,

2
+
R
r

_
,

1
]
1
1
cos (2)
c. And, finally, we get t

from Eq. B.1-17:


23

t
oo
2
1
r sin
ov
o
oo
+
v
r
+ v

cot
r

1
]
1

2v
~
1
R
R
r

_
,

1
3
2
R
r
+
1
2
R
r

_
,

_
,

cos +
R
r

_
,

1 +
3
4
R
r
+
1
4
R
r

_
,

_
,

cos

1
]
1
1


3v
~
2R

R
r

_
,

2
+
R
r

_
,

1
]
1
1
cos (3)
When r = R, all of the normal stresses are zero, in agreement with
Example 3.1-1.
24
Note to p. 81
Here we give the details of the derivation of Eq. 3.3-1 from Eq.
3.2-9. Although Eq. 3.3-1 itself is not much used, the integral of Eq.
3.3-1 over large flow systems is widely used. We call this the
"macroscopic mechanical energy balance"; the term "engineering
Bernoulli equation" is also used.
The derivation we work through here is an excellent exercise in
using some of the "del" relations given in Appendix A (see
particularly A.4).
We start by forming the dot product of the local velocity v with
Eq. 3.2-9. The last term presents no problems:

v pg
( )
p v g
( ) (1)
The term involving V t
[ ]
may be rearranged using Eq. A.4-29 in
Example A.4-1:

v V t
[ ] ( )
V t v
[ ] ( )
+ t : Vv ( )
(2)
The term containing

Vp may be similarly rearranged by using Eq.
A.4-19:

v Vp
( )
V pv
( )
+ p V v
( ) (3)
We now tackle the remaining two terms by first putting both of
them on the left side of the equation:

v
o
ot
pv

_
,

+ v V pvv
[ ] ( )

pv
o
ot
v

_
,

+ v v
( )
o
ot
p + v pv Vv
[ ] ( )
+ v v
( )
V pv
( )
(4)
------------- ----------------
In the first term, we differentiate the product with respect to t, and in
the second term, we use Eq. A.4-24. In the second line, the dashed
underlined terms then sum to zero by using the equation of
25
continuity; furthermore, the first term is split up into two terms, and
the third term is rearranged, thus:

o
ot
1
2
p v v
( )

_
,


1
2
v v
( )
o
ot
p + pv v Vv
[ ] ( )
(5)
We again use the equation of continuity to rewrite the second term in
Eq. 5:

o
ot
1
2
p v v
( )

_
,

+
1
2
v v
( )
V pv
( )
+ pv v Vv
[ ] ( )
(6)
Now the second and third term may be combined by using Eq. A.4-
19 with

s replaced by

1
2
v v
( ) and v replaced by

pv:

o
ot
1
2
p v v
( )

_
,

+ V
1
2
p v v
( )
v

_
,

o
ot
1
2
pv
2

_
,

+ V
1
2
pv
2
v

_
,

(7)
Clearly knowing that the final result is Eq. 3.3-1 is very helpful in
doing the last several steps.
26
Note to p. 82 (i)
We want to verify that

t:Vv
( ) , when written for a Newtonian
fluid, may be written as a sum of squares as shown in Eq. 3.3-3, and is
hence positive.
First define a tensor

y Vv + Vv ( )

, and then

t:Vv
( ) may be
written for Newtonian fluids (for which t is symmetric) as:

t:Vv
( )

1
2
t:

y
( )

1
2

y
2
3
V v
( )
o

1
]
:

y
( )
+
1
2
x V v
( )
o:

y
( )
(1)
where Eq. 1.2-7 has been used. Next, since

o:

y
( )
2 V v
( ) , we get

t:Vv ( )
1
2
y: y ( )
4
3
V v ( )
2

1
]
+x V v ( )
2
(2)
This is equivalent to:

t:Vv
( )

1
2

y
2
3
V v
( )
o

1
]
:

y
2
3
V v
( )
o

1
]
( )
+x V v
( )
2
(3)
which is another way of writing Eq. 3.3-3. The last step may be seen
to be true by expanding the double-dot product in Eq. 3 thus:

y
2
3
V v
( )
o

1
]
:

y
2
3
V v
( )
o

1
]
( )

y: y ( )
4
3
V v ( )
2

4
3
V v ( )
2
+
4
9
3 V v ( )
2

y: y ( )
4
3
V v ( )
2
(4)
which is the coefficient of

1
2
in Eq. 2.
27
Note to p. 82 (ii)
Eq. 3.4.1 is obtained by taking the cross product of the position
vector with the equation of motion in Eq. 3.2-9. To do this, we form
the cross product, term by term:
a. The time-derivative term is straightforward; for the ith component:

r
o
ot
pv

1
]
1
i

o
ot
r pv
[ ]
i
(1)
because the position vector r is independent of the time t.
b. For the next term in the equation, we consider only the i
component and expand the expression in terms of its components:

r V pvv
[ ]

1
]
i
r
ijk
x
j

x
l
pv
l
v
k
l

1
]
1
k


r
ijk
l
_
k
_
j
_
o
ox
l
px
j
v
l
v
k
r
ijk
pv
l
v
k
l
_
k
_
j
_
o
jl


x
l
pv
l
r
ijk
x
j
l
_
k
_
j
_
v
k
r
ijk
pv
j
v
k
k
_
j
_
(2)
In the second line, we have moved the

x
j
inside the differentiation
and subtracted off a compensating term. In the third line, we have
rearranged the first term and performed the sum on l in the second
term. It can be seen that the second term is zero (inasmuch as it
involves a double sum on a pair of indices that appear symmetrically
in one factor (

v
j
v
k
) and antisymmetrically in another (

r
ijk
); see
Exercise 5 on p. 815). Now we convert Eq. 2 back to bold-face
notation:

r V pvv
[ ]

1
]
i
V pv r v
[ ]

1
]
i
(3)
c. Next we examine the term containing the pressure:
28

r Vp
[ ]
i
r
ijk
k
_
j
_
x
j
o
ox
k
p r
ijk
k
_
j
_
o
ox
k
x
j
p r
ijk
k
_
j
_
o
jk
p (4)
The last term is zero, since it involves a double sum on a pair of
indices that appear symmetrically in one factor and antisymmetric in
another. The other term can be rearranged as follows:

o
ox
l
r
ijk
x
j
po
kl
{
l
_
k
_
j
_

o
ox
l
r
ijk
x
j
po
kl
k
_
j
_

l
_
(5)
On the right side, the quantity within the braces can be recognized as
the cross product of a vector with a tensor (see text just after Eq. A.3-
19). Therefore the above result in Eq. 5 can be written as:

o
ox
l
r po
{
il

l
_
o
ox
l
r po
{
li

l
_
V r po
{

1
]
i
(6)
In order to write Eq. 6 as the divergence of a tensor, the indices must
be as in Eq. A.4-13. This requires introducing the transpose of the
cross product as indicated above.
d. The term containing the tensor t can be treated in somewhat the
same manner as in part (c) above:

r V t
[ ]

1
]
i
r
ijk
x
j
k
_
j
_
V t
[ ]
k
r
ijk
x
j
k
_
j
_

x
l
t
lk
l
_
(7)
Next we write this intermediate result as the sum of two terms:

r
ijk
l
_
k
_
j
_

x
l
x
j
t
lk
r
ijk
l
_
k
_
j
_
t
lk

x
l
x
j

o
ox
l
r
ijk
x
j
t
kl

k
_
j
_

l
_
r
ijk
l
_
t
lk
k
_
j
_
o
jl

o
ox
l
r t

{
il

l
_
r
ijk
t
jk
k
_
j
_


x
l
r t

{
li

+ r
ikj
t
jk
k
_
j
_
l
_
29

V r t

1
]
1
i
+ r:t
[ ]
i
(8)
If the stress tensor is symmetric, the

r:t
[ ]
i
term vanishes, and

t

may
be replaced by t in the first term.
e. The ith component of the external force term,

r pg
[ ]
i
, is
straightforward.
When all the terms are collected, Eq. 3.4-1 results.
30
Note to p. 86(i)
In Example 3.5-1 it was pointed out that, to derive the Bernoulli
equation, we need the vector identity:

v v V v
[ ]

1
]
( )
0 (1)
To show that this relation is true, we use A.2 and Eq. A.4-10.
The proof requires replacing the dot and cross operations by
their expressions in terms of vector components. This is most
efficiently done by making use of summations. First we write the dot
product in terms of the components:

v v V v
[ ]

1
]
( )
v
i
i
_
v V v
[ ]

1
]
i
(2)
Next we write the cross product operations using the

r
ijk
symbol:

v
i
i
_
r
ijk
v
j
V v
[ ]
k
k
_
j
_
v
i
i
_
r
ijk
v
j
r
klm
o
ox
l
v
m
m
_
l
_
k
_
j
_
(3)
Then we rearrange the expression and make use of the cyclic
property of the permutation symbol, i.e.,

r
klm
r
mkl
r
lmk
:

r
ijk
r
klm
v
i
m
_
l
_
k
_
j
_
i
_
v
j
o
ox
l
v
m

r
ijk
r
lmk
v
i
m
_
l
_
k
_
j
_
i
_
v
j
o
ox
l
v
m
(4)
We can now use Eq. A.2-7 to replace the sum on k of products of two
permutation symbols:

o
il
o
jm
o
im
o
jl
( )
m
_
v
i
v
j
o
ox
l
v
m
l
_
j
_
i
_
(5)
After doing the sums on l and m we get two terms:
31

v
i
v
j
j
_
i
_
o
ox
i
v
j
v
i
v
j
j
_
i
_
o
ox
j
v
i
(6)
In the second summation, we replace i by j and j by i to get:

v
i
v
j
j
_
i
_
o
ox
i
v
j
v
j
v
i
j
_
i
_
o
ox
i
v
j
0 (7)
It may now be seen that both terms are the same. Consequently their
difference is zero. Therefore, we have proven the identity in Eq. 1.
32
Note to p. 86(ii)
Did you notice some similarity between Eq. 3.5-11 and Eq. 3.3-
2? In this "Note," we demonstrate the connection.
If we assume steady state, inviscid flow (as we did in Example
3.5-1), then Eq. 3.3-2 can be rewritten, with the help of Eq. 3.5-4 and
Eq. A.8-19, as

p
D
Dt
1
2
v
2
+

d
( )
v Vp
( ) (1)
where the inviscid flow assumption has been used to get rid of the
terms containing t . Next we use the steady-state flow assumption to
rewrite Eq. (1) as

p v V
1
2
v
2
( )
+ p v V

d
( )
v Vp ( )
(2)
Then we write the potential energy term using

V

d gVh, where h is
the coordinate in the direction opposite to the direction of gravity.
This gives

v V
1
2
v
2
( )
+ v gVh
( )
+
1
p
v Vp
( )
0 (3)
We can now divide each term by

v and introduce the unit vector

s v v , and then recognize that

s V
( )
d ds , where s is the
coordinate along a streamline. Equation (3) may now be written as

d
ds
1
2
v
2
( )
g
dh
ds

1
p
dp
ds
0 (4)
which is the same as Eq. 3.5-11. Equation (4) may then be integrated
from point "1" to point "2" along the streamline to get the Bernoulli
equation. Thus it is seen that the Bernoulli equation can be obtained
directly from the mechanical energy balance.
33
Note to p. 90(i)
In Example 3.6-3, we give Eq. 3.6-20, but we don't obtain the
pressure distribution. If one knows the pressure at one point in the
system (after a steady state velocity distribution has been obtained),
then it is possible to determine the constant of integration and obtain
the pressure distribution.
Here we attack a different problem. We now show how Eq. 3.6-
20 can be used to get the pressure distribution in the space between
the cylinders in terms of the pressure

p
0
that exists in the system
before the outer cylinder is rotated.
We start by temporarily relaxing the assumption of incom-
pressibility. It can be shown that the equation of motion in Eq. 3.2-9
and Newton's law of viscosity as given in Eq. 1.2-7 are still valid for a
compressible fluid, provided that one assumes that the viscosity is
independent of the pressure. The important new feature that we have
to inject into the development is the equation of statethat is, the
density as a function of the pressure. We do this via a Taylor series
for the density as a function of the Gibbs free energy,

G H TS

p p
0
+
op
oT

_
,

0
G G
0
( )
+ (1)
in which the subscript zero refers to properties of the system when
the fluid is at rest before the outer cylinder is rotated. Next we
truncate the Taylor series after two terms and write

p p
0
1 + b
0
G G
0
( )

1
]
p
0
1 + b
0
p
1
dp
p
0
p
[

1
]
1
(2)
in which

b
0
is the value of

b 1 p
( )
op oG
( )
T
op op
( )
T
at

p p
0
.
The integration of Eq. 3.6-20 may now be performed by
inserting the velocity distribution of Eq. 3.6-29 and assuming that the
density is a known function of the pressure

p
1
dp
p
0
p
[

v

2
r
[
dr
x
2
R
2
O
o
2
1 x
2
( )
2

x

x

_
,

2
[
1

d
34

x
2
R
2
O
o
2
1 x
2
( )
2

x
2

2

+
x
2

_
,

[
d

x
2
R
2
O
o
2
1 x
2
( )
2
1
2

2
x
2

x
2

_
,

2ln + C

1
]
1
1
(3)
The integration constant C is determined by the mass conservation
statement

p
0
d
x
1
[
pd
x
1
[
or

p
1
dpd 0
p
0
p
[
x
1
[
(4,5)
where Eq. (2) has been used. Next, substitution of Eq. (3) into Eq. (5)
gives a relation from which C can be determined

1
2

2
x
2

x
2

_
,

2ln + C

1
]
1
1
x
1
[
d 0 (6)
Performing the integration gives

1
2x
2

4
4

1
2
x
2
ln 2
1
2

2
ln
1
4

_
,

+
1
2
C
2

1
]
1
x
1
0 (7)
This gives for the integration constant

C 1
1 +x
2
4x
2

3x
2
lnx
1 x
2
(8)
Hence Eq. (3) becomes

p
1
dp
p
0
p
[

x
2
R
2
O
o
2
1 x
2
( )
2
1
2

2
x
2

x
2

_
,

2ln
1
1 +x
2
4x
2

3x
2
lnx
1 x
2

1
]
1
1
1
1
1
(9)
35
For an incompressible fluid, the left side of Eq. (9) becomes simply

(p p
0
) p
0
, so that we get for the pressure distribution in the two-
cylinder system with the outer cylinder rotating

p p
0
p
0

x
2
R
2
O
o
2
1 x
2
( )
2
1
2

2
x
2

x
2

_
,

2ln
1
1 +x
2
4x
2

3x
2
lnx
1 x
2

1
]
1
1
1
1
1
(10)
A similar development for the system with the inner cylinder
rotating with angular velocity

O
i
and the outer cylinder fixed, gives

p p
0
p
0

x
4
R
2
O
i
2
1 x
2
( )
2
1
2

_
,

2ln
1
1 +x
2
4

1 + 2x
2
1 x
2
lnx

1
]
1
1
1
1
1
(11)
36
Note to p. 90(ii)
An alternative method of solving the problem in Example 3.6-3
is given here. Eq.3.6-21 may be set up by making a shell torque
balance on a shell of thickness

Ar and height L. Then the torque at
radius r is equal to

2rrLt
r
r
r , whereas the torque at radius

r + Ar
is

2r r + Ar ( )Lt
r
r+Ar
r + Ar ( )
. The torque is (force per unit area) x
(area) x lever arm. When these torques are equated, we get after
dividing by

Ar and letting

Ar go to zero:

d
dr
r
2
t
r
( )
0 or

d
dr
r
2
r
d
dr
v

_
,

1
]
1

_
,

0 (1a,b)
Eq. B.1-11 has also been used. Here we show that this is equivalent to
Eq. 3.6-21, which is

d
dr
1
r
d
dr
rv

( )

_
,

0 or

d
2
v

dr
2
+
1
r
dv

dr

r
2
0 (2)
We start by performing the differentiations in the large
parentheses in Eq. 1b:

r
2
r
d
dr
v

_
,

1
]
1
r
2
r
1
r
dv

dr

r
2

_
,

1
]
1
r
2
dv

dr
rv

(3)
Next do the differentiation with respect to r (in Eq. 1b) and set the
result equal to zero:

d
dr
r
2
dv

dr
rv

_
,

2r
dv

dr
+ r
2
d
2
v

dr
2
r
dv

dr
v


r
2
d
2
v

dr
2
+
1
r
dv

dr

r
2

_
,

0 (4)
37
Therefore either

r
2
is zero, or the quantity in parentheses is zero. But

r
2
cannot be zero, and hence Eq. 2which came from Example 3.6-
3must be the same as Eq. 1b (from the torque balance).
38
Note to p. 117
When you look at Fig. 4.1-2, you might wonder whether the
slope at y = 0 is

1. You can answer that question by differentiating
Eq. 4.1-15 with respect to q:

d
dq
erfcq
q0

d
dq
1
2
r
e
q
2
dq
0
q

_
,

q0


2
r
e
q
2
q0

2
r
1.1287 (1)
Here we have used the Leibniz formula for differentiating integrals in
C.3 and the definition of the complementary error function in C.6.
As may be seen from Fig. 4.1-2, the slope is somewhat steeper than
minus 1.
Notice that in Eq. 4.1-14, as well as in Eq. 1 above, we used a
bar over the q to make a distinction between the variable of
integration and the upper limit on the integral. When applying the
Leibniz theorem, it is vital to make this distinction.
39
Note to p. 121
Example 4.1-3 should be straightforward, except possibly for
(a) the line immediately after Eq. 4.1-49, and (b) the line following Eq.
4.1-53. Here we show how to obtain the expressions given in these
two locations.
(a) Let the real and imaginary parts of the arguments be:
for

z
1
:

z
1r
and

z
1i
for

z
2
:

z
2r
and

z
2i
for w:

w
r
and

w
i
Then the expression

J z
1
w
{
J z
2
w
{ becomes

J z
1r
+ iz
1i
( ) w
r
+ iw
i
( ) {
J z
2r
+ iz
2i
( ) w
r
+ iw
i
( ) {
(1)
Then, taking the real parts of both sides, we get

z
1r
w
r
z
1i
w
i
z
2r
w
r
z
2i
w
i
(2)
Rearranging gives:

w
r
z
1r
z
2r
( )
w
i
z
1i
z
2i
( ) (3)
Since w, and hence

w
r
and

w
i
, are arbitrary, the only way that this
equation can be satisfied is if

z
1r
z
2r
and

z
1i
z
2i
(i.e.,

z
1
z
2
).
(b) The square root of i will be some number in the complex
plane, say

a + bi , so that

i a + bi (4)
where the real quantities a and b must be determined. When the left
and right sides of Eq. 4 are squared, we get

i a
2
b
2
( )
+ 2abi (5)
40
To find a and b, we equate the real and imaginary parts of the left and
right sides:

a
2
b
2
0 and

2ab 1 (6)
Eliminating b between these two equations gives

a
4

1
4
(7)
Hence

a
2

1
2
,
1
2
and

a
1
2
,
1
2
i (8)
whence

b
1
2
,
1
2
1
i
(9)
and the two square roots of i are

i

a + bi
1
2
1 + i
( )
(10)
Alternatively, one can write i in polar form and use the fact that
i has a unit length:

i re
i
1e
ir 2
(11)
Then the square root of i is

i e
ir 4
cos
r
4
+ i sin
r
4

_
,


1
2
+ i
1
2

_
,

(12)
41
Note to p. 123
Here we show how Eq. (A) of Table 4.2-1 is obtained.
The velocity is of the form

v x, y ( ) o
x
v
x
x, y ( ) + o
y
v
y
x, y ( )
(1)
The velocity components are related to the stream function according
to the relations in the 3rd column.
The first term in Eq. (A) comes from the first term in Eq. 4.2-1,
which is

t
V v
[ ]
o
z

t
V v
[ ]
z
o
z

t
v
y
x

v
x
y

_
,


o
z
o
ot
o
2

ox
2
+
o
2

oy
2

_
,

o
z
o
ot
V
2
(2)
The right side of Eq. (A) comes from the right side of Eq. 4.2-1:

vV
2
V v
[ ]
v
o
2
ox
2
+
o
2
oy
2

_
,

V v
[ ]

o
z
v
o
2
ox
2
+
o
2
oy
2

_
,

o
2

ox
2
+
o
2

oy
2

_
,

o
z
vV
4
(3)
The second term on the left side of Eq. (A) can be taken care of
similarly:

V v V v
[ ]

1
]

1
]
V v o
z
V
2

1
]

1
]

V o
x
v
x
+ o
y
v
y
( )
o
z
V
2

1
]

1
]

V o
y
v
x
V
2
+ o
x
v
y
V
2

1
]

1
]
42

o
x
o
y
o
z
o
ox
o
oy
o
oz
0 v
x
V
2
0

o
x
o
y
o
z
o
ox
o
oy
o
oz
v
y
V
2
0 0

+o
z
o
ox
v
x
V
2

( )
+ o
z
o
oy
v
y
V
2

( )

o
z
ov
x
ox
+
ov
y
oy

_
,

V
2
+ o
z
v
x
o
ox
V
2
+ v
y
o
oy
V
2

_
,

(4)
The first term is zero, because of the assumption of incompressibility
(see statement just above Eq. 4.2-1). Then we get finally

o
z

o
oy
o
ox
V
2
+
o
ox
o
oy
V
2

_
,

o
ox
o
oy
o
ox
V
2

o
oy
V
2

o , V
2

( )
o x, y
( )
(5)
When the results in Eqs. 2, 3, and 5 are combined, we get same
expression that is given in Eq. (A) of the table.
43
Note to p. 125
Here we want to fill in the details of getting Eq. 4.2-20 from Eqs.
4.2-18 and 4.2-19. This is a "straightforward but tedious exercise."
We begin by evaluating each of the four squared terms in Eq.
4.2-19 using the velocity components in Eqs. 4.2-13 and 14; it is
convenient to introduce the dimensionless variable

r R.

2
ov
r
or

_
,

9
2
v
~
2
r
2

1

3

1
]
cos
( )
2
(1)

2
1
r
ov

o
+
v
r
r

_
,

2v
~
2
r
2
1 +
3
4

1
+
1
4

3

1
]
cos + 1
3
2

1
+
1
2

3

1
]
cos
( )
2

9
8
v
~
2
r
2

1
+
3

1
]
cos
( )
2
(2)

2
v

cot
r
+
v
r
r

_
,

2v
~
2
r
2
1 +
3
4

1
+
1
4

3

1
]
cos + 1
3
2

1
+
1
2

3

1
]
cos
( )
2

9
8
v
~
2
r
2

1
+
3

1
]
cos
( )
2
(3)

r
o
or
v

_
,

+
1
r
ov
r
o

_
,

ov

or

r
+
1
r
ov
r
o

_
,

v
~
2
r
2

3
4

1
+
3
4

3

1
]
sin + 1
3
4

1

1
4

3

1
]
sin
1
3
2

1
+
1
2

3

1
]
sin

_
,

v
~
2
r
2

3
2

3
sin

1
]
2
(4)
When the above results are combined, we get

t:Vv
( )
r
2
v
~
2
27
4

2

27
2

4
+
27
4

6

1
]
cos
2
+ v
~
2
9
4

6

1
]
sin
2

(5)
44
Then the kinetic contribution to the force on the sphere is given by

F
k
v
~
2r t:Vv
( )
r
2
dr
R
~

0
r

sind

2r
2
3
v
~
2
27
4

2

27
2

4
+
27
4

6

1
]
dr + 2r
4
3
v
~
2
9
4

6

1
]
dr
R
~
[
R
~
[
(6)
since

cos
2

0
r
[
sin d
2
3
and

sin
3

0
r
[
d
4
3
. Then, finally

F
k
9rv
~
R
2
2
4
+
6

1
] 1
~
[
d + 6rv
~
R
6

1
] 1
~
[
d

9rv
~
R
1
+
2
3

3

1
5

5

1
]
1
~
+ 6rv
~
R
1
5

5

1
]
1
~

9rv
~
R 1
2
3
+
1
5

1
]
+ 6rv
~
R
1
5

1
]

9rv
~
R
2
5

+ 6rv
~
R
1
5


24
5
+
6
5
( )
rv
~
R

6rv
~
R (7)
which is Stokes' law.
45
Note to p. 139
We show here how to get the Falkner-Skan equation in Eq. 4.4-35 for
the flow near a corner. For this system, the external flow

v
e
was
found earlier (see Eqs. 4.3-42 and 43) to be

v
e
x
( )

2c
2
x
2 ( )
c' x
2 ( )
(1)
We then have to solve Eq. 4.4-11 to get the velocity distribution in the
neighborhood of the wedge shown in Eq. 4.3-4:

v
x
ov
x
ox
+ v
y
ov
x
oy
v
e
dv
e
dx
+v
o
2
v
x
oy
2
(2)
This equation can be rewritten in terms of the stream function

x, y
( ) , by using the expressions for the velocity components in the
first row of Table 4.2-1:

o
oy

_
,


o
2

oxoy

_
,

+
o
ox

_
,


o
2

oy
2

_
,

v
e
dv
e
dx
+v
o
2
oy
2

o
oy

_
,

(3)
Insertion of Eq. 1 into this equation then gives Eq. 4.4-32:

o
oy
o
2

oxoy

o
ox
o
2

oy
2

c'
2

_
,

1
x
23 ( ) 2 ( )
v
o
3

oy
3
(4)
Next we want to rewrite this equation in terms of f and q:

x, y ( ) c'v 2 ( )x
1 2 ( )
f q ( ) Ax
1 2 ( )
f q ( ) (5)

q x, y
( )

c'
v 2
( )
y
x
1 ( ) 2 ( )

B
y
x
1 ( ) 2 ( )
(6)
46
We start by converting the various derivatives from

x, y
( ) to
derivatives of

f q
( ) :

o
oy
Ax
1 2 ( )
df
dq
oq
oy
ABx
1 2 ( )
f '
1
x
1 ( ) 2 ( )
ABx
2 ( )
f ' (7)

o
2

oy
2
ABx
2 ( )
oq
oy
f " AB
2
x
2 ( )
x
1 ( ) 2 ( )
f " AB
2
1
x
12 ( ) 2 ( )
f "(8)

y
3
AB
2
1
x
12 ( ) 2 ( )
q
y
f AB
3
1
x
23 ( ) 2 ( )
f (9)

o
ox
A
d
dx
x
1 2 ( )

_
,

f Ax
1 2 ( )
df
dq
oq
ox

A
x
1 2 ( )
2
( )
1
x
f Ax
1 2 ( )

By 1
( )
2
( )
x
1 ( ) 2 ( )1

_
,

f '

A
1
2
( )
x
1 2 ( )
x
f + A
1
( )
2
( )
x
1 2 ( )
x
q f ' (10)

o
2

oyox
A
1
2
( )
x
1 2 ( )
x
oq
oy
f '+ A
1
( )
2
( )
x
1 2 ( )
x
oq
oy
d
dq
q f '
( )

AB
1
2
( )
x
1 2 ( )
x
1 ( ) 2 ( )
x
f '+ AB
1
( )
2
( )
x
1 2 ( )
x
1 ( ) 2 ( )
x
q f "+ f '
( ) (11)
The terms on the right side of Eq. 4.4-32 are then:

c'
2

_
,

1
x
23 ( ) 2 ( )
+ vAB
3
1
x
23 ( ) 2 ( )
f
c'
2
2
1
x
23 ( ) 2 ( )
+ f
( )
(12)
Next we write down the terms on the left side of Eq. 4.4-32:

o
oy
o
2

oxoy

o
ox
o
2

oy
2
A
2
B
2
1
2
( )
x
2 ( )
x
1 2 ( )
x
1 ( ) 2 ( )
x
f
2
47

A
2
B
2
1
( )
2
( )
x
2 ( )
x
1 2 ( )
x
1 ( ) 2 ( )
x
q f f "+ f
2
( )

A
2
B
2
1
2
( )
x
1 2 ( )
x
12 ( ) 2 ( )
x
f f

+ A
2
B
2
1
( )
2
( )
x
1 2 ( )
x
12 ( ) 2 ( )
x
q f f "

c'
2

2
x
23 ( ) 2 ( )
f
2
c'
2
1
2
( )
x
23 ( ) 2 ( )
f f

c'
2
1
2
( )
x
23 ( ) 2 ( )
f
2
f f
( )
(13)
Combining the results in Eqs. 12 and 13 gives

c'
2
1
2
( )
x
23 ( ) 2 ( )
f
2
f f
( )
c'
2
1
2
( )
x
23 ( ) 2 ( )
+ f
( )
(14)
or

f
2
f f + f (15)
which is the same as Eq. 4.4-35, the Falkner-Skan equation.
[Note: In earlier printings of the textbook, the prime was omitted from
c', and the quantity c' was not defined.]
48
Note to p. 154
To get the average flow velocity from Eq. 5.1-4, we integrate the
velocity distribution over the circular tube cross-section:

v
z

v
z
rdrd
0
R
[
0
2r
[
rdrd
0
R
[
0
2r
[
v
z,max
1 r R ( )

1
]
1 7
rdrd
0
R
[
0
2r
[
rdrd
0
R
[
0
2r
[

v
z,max
2r
rR
2
1 r R ( )

1
]
1 7
rdr 2v
z,max
0
R
[
1
[ ]
1 7
0
1
[
d (1)
where

r R . To evaluate the integral, we make a change of variable

1 o . Then

v
z
2v
z,max
o
1 7
1 o ( )
0
1
[
do (2)
This can then be written as the sum of two integrals, which can be
evaluated:

v
z
2v
z,max
o
1 7
0
1
[
do o
8 7
0
1
[
do
( )
2v
z,max
o
8 7
8 7

o
15 7
15 7

_
,

0
1
0.82v
z,max
(3)
Since 0.82 is approximately 4/5, relation in Eq. 5.1-5 is verified.
49
Note to p. 170
Here we fill in some of the missing steps following Eq. 5.6-18.
Setting

C
1
0 in Eq. 5.6-18 and rewriting the equation, we get:

F F F + F
( )
2 F (1)
where the primes indicate differentiation with respect to . Next we
note that Eq. 1 may be put into the form

1
2
F
2
( )

F ( )

2 F (2)
Each term may now be integrated with respect to to give

1
2
F
2
F 2F C
2
(3)
which is the same as Eq. 5.6-19. The constant

C
2
is zero according to
Eq. 5.6-16, with a, b, and d set equal to zero. [Note: The comment
about setting

ln does not seem to be helpful.]
Eq. 3 is a separable, first-order differential equation

dF
d
2F +
1
2
F
2
or

dF
2F 1 +
1
4
F
( )

d

(4)
Then, according to a table of integrals (e.g., Formula 101.1 of
Dwight's Table of Integrals), Eq. 4 gives, on integration

1
2
ln
1 +
1
4
F
F
ln + lnC
3
or

ln
1 +
1
4
F
F
ln + lnC
3
(5a,b)
Next take the antilog of the equation and then square the result to
obtain
50

F
1 +
1
4
F
C
3

( )
2
(6)
Now either a "plus" sign or a "minus" sign may be inserted inside the
absolute value bars. Since we have no reason to prefer one over the
other, we consider the two cases separately:
Case I (plus sign):

F + 1 +
1
4
F
( )
C
3

( )
2
or

F
C
3

( )
2
1
1
4
C
3

( )
2
(7)
Case II (negative sign):

F 1 +
1
4
F
( )
C
3

( )
2
or

F
C
3

( )
2
1 +
1
4
C
3

( )
2
(8)
When F in Eq. 7 is plotted against

C
3

( )
2
, it tends toward
infinity as

C
3

( )
2
4 is approached from below, and approaches
minus infinity when approached from above. Hence Case I is
physically unreasonable behavior for the stream function.
When F in Eq. 8 is plotted versus

C
3

( )
2
, is it monotone
decreasing over the entire range of

C
3

( )
2
. Since this is physically
reasonable, we choose the solution in Case II (which agrees with Eq.
5.6-20 in the textbook).
When Eq. 8 (or Eq. 5.6-20) is inserted into Eq. 5.6-12 and 13, Eqs.
5.6-21 and 22 follow immediately.
Then substitution of Eq. 5.6-21 into the expression for J in Eq.
5.6-2 gives:

J 2r pv
z
2
0
~
[
rdr 2rpv
t ( )2
2C
3
2
( )
2
1 +
1
4
C
3

( )
2

1
]
1
4 0
~
[
d (9)
We now let

1
4
C
3

( )
2
x
2
so that

d 4 C
3
2
( )
xdx ; then
51

J 32rpv
t ( )2
C
3
2
xdx
1 + x
2
( )
4 0
~
[

32rpv
t ( )2
C
3
2
1
6 1 + x
2
( )
3

1
]
1
1
1
0
~

16
3
rpv
t ( )2
C
3
2
(10)
whence Eq. 5.6-23 follows at once.
Similarly the mass rate of flow is

w 2r pv
z
0
~
[
rdr 2rp
v
t ( )
z
2C
3
2
( )
1 +
1
4
C
3

( )
2

1
]
1
2 0
~
[
d z
2

4rv
t ( )
pzC
3
2

4
C
3
2
xdx
1 + x
2
( )
2 0
~
[
16rpv
t ( )
z
1
2 1 + x
2
( )

1
]
1
1
0
~

8rpv
t ( )
z (11)
52
Note to p. 198
Here we show how to obtain the macroscopic mass and momentum
balances from the corresponding equations of change.
Macroscopic Mass Balance
The equation of change for conservation of mass is given in Eq.
3.1-4. We want to integrate this equation over the system pictured in
Fig. 7.0-1 on p. 197:

op
ot
dV V pv
( )
dV
V t ( )
[
V t ( )
[
(1)
We now apply the 3-dimensional Leibniz formula (Eq. A.5-5) to the
left side and the Gauss divergence theorem (Eq. A.5-2) to the right
side:

d
dt
pdV n pv
S
( )
dS
S t ( )
[
V t ( )
[
n pv
( )
dS
S t ( )
[
(2)
in which n is the outwardly directed unit vector on the surface

S t
( ) .
The surface is a function of time t, because there may be moving parts
in the system. Eq. 2 may be rewritten as

d
dt
pdV
V t ( )
[
n p v v
S
( ) ( )
dS
S t ( )
[
(3)
We note that the mathematical surface

S t
( ) defining the system
consists of several parts that we identify as follows:
the "inlet" surface

S
1
(on which

v
S
= 0)
the "outlet" surface

S
2
(on which

v
S
= 0)
the "fixed" surface

S
f
(on which

v v
S
0)
the "moving" surface

S
m
(on which

v v
S
= 0)
with v being the fluid velocity and

v
S
the surface velocity. The
surface integrals are then split into four parts corresponding to the
four types of surfaces.
53
The integral on the left side is the total mass

m
tot
in the system.
The surface integrals over

S
f
and

S
m
are zero, because v =

v
S
.
Therefore we are left with

d
dt
m
tot
n pv
( )
dS n pv
( )
dS
S
2
[
S
1
[
(4)
We now introduce the vectors

u
1
and

u
2
, which are unit vectors in
the direction of flow at planes "1" and "2", respectively. Thus the n in
the

S
1
integral will be

u
1
, whereas the n in the

S
2
integral will be

+ u
2
. Now we make the assumptions that (i) the density is a constant
over the cross section, and (ii) that the velocity is always parallel to
the walls of the entry and exit tubes, so that

v u at plane "1" and

v u at plane "2". Then Eq. 4 becomes

d
dt
m
tot
+p
1
vdS p
1
vdS
S
2
[
S
1
[
(5)
Here v is the velocity in the direction of flow, which varies across the
cross section. Therefore integrations over the surfaces

S
1
and

S
2
give

d
dt
m
tot
p
1
v
1
S
1
p
2
v
2
S
2
w
1
w
2
(6)
where is the average value over the cross section;

w
1
and

w
2
are
the mass rates of flow at the inlet and outlet, respectively. Eq. 6 is just
the same as Eq. 7.1-1 in the textbook, which was written down by
using elementary arguments (i.e., common sense).
Macroscopic Momentum Balance
When the equation of motion of Eq. 3.3-9 is integrated over the
volume of the flow system in Fig. 7.0-1, we get

o
ot
pv

_
,

V t ( )
[
dV V pvv
[ ]
V t ( )
[
dV VpdV V t
[ ]
V t ( )
[
V t ( )
[
dV + pgdV
V t ( )
[
(7)
54
Next apply the Leibniz formula to the left side and the Gauss
divergence theorem (or Eq. A.5-2) to the surface integrals on the right
side to get

d
dt
pvdV pv n v
S
( )
dS
S t ( )
[
V t ( )
[
n pvv
[ ]
S t ( )
[
dS npdS n t
[ ]
S t ( )
[
S t ( )
[
dV
+

pgdV
V t ( )
[
(8)
The integral on the left side is just the total momentum

P
tot
in the
flow system. If g does not change over the volume of the flow system,
then it may be removed from the integral, which gives

m
tot
. Then

d
dt
P
tot
n p v v
S
( )
v

1
]
S t ( )
[
dS npdS n t
[ ]
S t ( )
[
S t ( )
[
dS + m
tot
g (9)
We now consider the three surface integrals seriatim:
The first integral is zero on fixed and moving surfaces, and

v
S
is zero at the entry and exit planes, so that

n p v v
S
( )
v

1
]
S t ( )
[
dS + u
1
u
1
u
1

1
]
S
1
[
pv
2
dS u
2
u
2
u
2

1
]
S
2
[
pv
2
dS

+p
1
v
1
2
S
1
u
1
p
2
v
2
2
S
2
u
2
(10)
Here it has been assumed that the flow at the inlet and outlet planes is
parallel to the container wall.
The second integral will contribute both at the inlet and outlet
planes and also to the force on the various solid surfaces:

npdS
S t ( )
[
+ u
1
S
1
[
pdS u
2
S
2
[
pdS npdS
S
f
+S
m
[

p
1
S
1
u
1
p
2
S
2
u
2
F
f s
p ( )
(11)
the last contribution being the force exerted by the fluid on the solid
surfaces by the pressure.
Finally the third integral will be

n t
[ ]
dS
S t ( )
[
+ u
1
t

1
]
S
1
[
dS u
2
t

1
]
S
2
[
pdS n t
[ ]
dS
S
f
+S
m
[
55

- F
f s
t ( )
(12)
Note that we have omitted the contributions at the entry and exit
planes because they would normally be quite small compared to the
pressure terms. Therefore we are left with just the force exerted by
the fluid on the solid surfaces because of the viscous forces.
When all the forces are combined we get for the macroscopic
momentum balance (with

F
f s
p ( )
+ F
f s
t ( )
F
f s
F
sf
):

d
dt
P
tot
p
1
v
1
2
S
1
u
1
p
2
v
2
2
S
2
u
2
+ p
1
S
1
u
1
p
2
S
2
u
2
+ F
sf
+ m
tot
g (13)
This is the same as Eq. 7.2-11 (or 7.2-12) in the textbook, obtained by
elementary reasoning.
[Note: The derivation of the macroscopic mechanical energy balance is
given in 7.8, the derivation of the macroscopic energy balance is given
in the Note to p. 454.]
56
Note to p. 199
Here we work through all the details of Example 7.1-1, (a)
explaining how to get Eq. 7.1-4, (b) giving the details of how the
difference in the modified pressures is obtained, and (c) working
through the algebraic details of the remainder of the problem.
a. (This development was given by Professor L. E. Wedgewood,
University of Illinois at Chicago)
We have to find the volume of liquid in the sphere below the
liquid level. We imagine that the sphere is generated by a circle in the
xz-plane, with its center at z = R and x = 0. The tank is draining in the
negative z-direction, and the exit from the sphere into the attached
tube is located at z = 0. The sphere is created by rotating the
generating circle around the z-axis.
The generating circle has the equation:

x
2
+ z R ( )
2
R
2
or

x
2
2Rz z
2
(1a,b)
Then we visualize the liquid volume as being made up of a stack of
thin circular disks of thickness dz, each with a volume of

dV rx
2
dz r 2Rz z
2
( )
dz (2)
Then the total volume of the liquid is:

V r 2Rz z
2
( )
dz
0
h
[

r Rz
2

1
3
z
3
( )
0
h

r Rh
2

1
3
h
3
( )
(3)
Thus the liquid volume is:

V rRh
2
1
1
3
h
R

_
,

(4)
We may check this result at three points where we know the result:
Tank full:

h 2R

V
4
3
rR
3
Tank half full:

h R

V
2
3
rR
3
57
Tank empty:

h 0

V 0
b. Next we want to apply Eq. 7.1-2 to the system. No liquid is
entering at plane "1" so that

w
1
0 , and, if the diameter of the exit
tube is sufficiently small that the flow in it is laminar, then

w
2
will be
given by the Hagen-Poiseuille formula (Eq. 2.3-21). Hence

d
dt
rRh
2
1
1
3
h
R

_
,

p
r P
2
P
3
( )
D
4
p
128L
(5)
To get the modified pressures we must specify a datum plane; we
choose this to be at the tube outlet (i.e., plane "3"), so that:

P
2
p
2
+ pgh
2
pgh + p
atm
( )
+ pgL (6)
That is,

p
2
is the pressure due to the liquid in the sphere above plane
"2" plus the atmospheric pressure, and

h
2
is the height of plane "2"
above the datum plane (i.e., plane "3"). Furthermore,

P
3
p
3
+ pgh
3
p
atm
+ 0 (7)
since

p
3
is the pressure atmospheric pressure at the tube outlet, and

h
3
is the distance upward from the datum plane (i.e., plane "3"),
which is zero. Hence, Eq. 5 becomes

d
dt
rRh
2
1
1
3
h
R

_
,


r pgh + pgL ( )D
4
128L
(8)
c. Eq. 8 may be rewritten as

R
h + L
d
dt
h
2

1
3
h
3
R

_
,


pgD
4
128L
A (9)
(which defines the quantity A) or
58

R
h + L
2h
h
2
R

_
,

dh
dt
A (10)
Now we introduce a new variable

H h + L in order to facilitate the
solution of the differential equation

2R H L
( )
H L
( )
2
H
dH
dt
A (11)
or

H 2 R + L
( )
+ 2R + L
( )
L
1
H

1
]
1
dH
dt
A (12)
Now we integrate this equation and make use of the initial and final
conditions:

1
2
H
2
2 R + L ( ) H + 2R + L ( )LlnH

1
]
1
2R+L
L
At
0
t
efflux
(13)
This gives

1
2
L
2

1
2
2R + L
( )
2
2 R + L
( )
L + 2 R + L
( )
2R + L
( )
+ 2R + L
( )
Lln
L
2R + L

At
efflux
(14)
or after some cancellations

t
efflux

1
A
2R
2
+ 2RL 2R + L
( )
Lln
2R + L
L

1
]
1
(15)
When

L
2
is factored out of the bracket expression, Eq. 7.1-8 of the
textbook is obtained.
59
Note to p. 202
A lawn sprinkler has four arms of length L and cross-sectional area S.
Water enters the sprinkler at the center at a mass flow rate w and
splits into four streams. It leaves the sprinkler at right angles to the
sprinkler arms. It is desired to find the angular velocity O of the
sprinkler, when there is a frictional torque of

T
f
.
The flow velocity is each arm is

v w 4pS, so that the velocity at the
outlet stream, relative to the sprinkler arm is (at plane "2")

v
2
o
y
w
4pS
LO

_
,

Therefore, Eq. 7.3-2 gives for the angular momentum balance at


steady state (here we take the arm to be projecting in the x direction,
the fluid to be issuing in the y direction, and the torque to be in the z
direction)

0 w Lo
x
o
y
w
4pS
LO

_
,

1
]
1
T
f
o
z
the entering stream contributing nothing to the angular momentum
balance. The unit vectors may now be removed

T
f
wL
w
4pS
LO

_
,

Then, solving for O we get



O
w
4pSL

T
f
wL
2
for the angular velocity of the sprinkler.
[For more on this problem, see Surely You're Joking Mr. Feynman," by
R. P. Feynman, Bantam Books, New York (1986), pp. 51-53.]
60
Note to p. 218
Here we work through the details from Eq. 7.7-8 to the end of the
example.
The first term of Eq. 7.7-5 may be transformed as follows:

2h
d
2
h
dt
2
2h
d
dt
u
( )
2h
1
2
u
1 2
du
dt

_
,

hu
1 2
du
dh
dh
dt
h
du
dh
(1)
from which Eq. 7.7-8 follows at once:

h
du
dh
N 1
( )
u + 2gh 0 (2)
To verify that Eq. 7.7-9 is the solution to Eq. 7.7-8, we substitute the
former into the latter to get

h C N 1
( )
h
N2
+
2g
N 2

1
]
1
N 1
( )
Ch
N1
+
2gh
N 2

1
]
1
+ 2gh 0 (3)
where C is the constant of integration. We then see that the terms in
"g" and the terms in "C" separately sum to zero:
g-terms:

2gh
N 2
N 1
( )
2gh
N 2
+
N 2
N 2
2gh 0 (4)
C-terms:

C N 1
( )
h
N1
N 1
( )
Ch
N1
0 (5)
We next apply the initial conditions (Eqs. 7.7-6 and 7) to the solution
in Eq. 7.7-9

dh
dt

_
,

2
Ch
N1
+
2gh
N 2
(6)
to get an equation for the constant of integration
61

2gH
R
0
R

_
,

4
CH
N1
+
2gH
N 2
(7)
When this is solved for C, we get (noting that

N R R
0
( ) as defined
immediately after Eq. 7.7-3):

C
2g
H
N2
1
N

1
N 2

_
,


4g
H
N2
1
N N 2
( )

_
,

(8)
Even though the factor containing

N 2
( ) would cause C to become
infinite for

N 2 , this need cause no alarm, since N is going to be a
large number, i.e., when the radius outlet hole is small compared to
the radius of the tank.
When we take the square root of Eq. 7.7-9 and introduce the
dimensionless liquid height

q h H, we then get Eq. 7.7-10:

dq
dt

2g
N 2
( )
H
q
2
N
q
N1

_
,

(9)
We then choose the minus sign, because we know that the height of
the fluid will be decreasing with increasing time, and therefore

dq dt
must be negative. To get the efflux time, we integrate Eq. 9 from t = 0
when

q 1 (full tank) to

t t
efflux
when

q 0 (empty tank):

t
efflux
dt
N 2
( )
H
2g
1
q 2 N
( )
q
N1
dq
0
1
[
0
t
efflux
[

NH
2g
o N ( )
(10)
Keep in mind that

t
efflux
NH 2g is the quasi-steady-state solution
in Eq. 7.7-3the solution that we would expect to be valid when N is
extremely large, i.e., for the case that the outlet hole is so small that
the system is never far from steady state.
The function

o N
( ) is then given by
62

o N ( )
1
2
N 2
N
1
q 2 N ( )q
N1
dq
0
1
[
(11)
This integral can probably not be performed analytically. However,
in the expression under the square-root sign, for large N, the first
term will predominate, over the range of integration (from

q 0 to

q 1). Therefore we take the first term outside of the radical and
write

o N
( )

1
2
N 2
N
1
q
1
2
N
q
N2

_
,

1 2
dq
0
1
[
(12)
The quantity to the

1
2
power can then be expanded in a Taylor
series (see Eq. C.2-1) about

2 N
( )
q
N2
0 to get, after integrating
term by term

o N
( )

1
2
N 2
N
1
q
1 +
1
1!
1
2
2
N
q
N2

_
,

+
1
2!
1
2
3
2
2
N
q
N2

_
,

1
]
1
1
dq
0
1
[

1
2
N 2
N
2 +
1
N N
3
2
( )
+
3
4N
2
N
7
4
( )
+

_
,

(13)
When this expression is expanded in a Taylor series around

1 N 0,
we get

o N ( ) 1
1
N

1
2
1
N
2
+ O
1
N
3

_
,

1
]
1
+
1
2
1
N
2
+ O
1
N
3

_
,

1
]
1
+
3
8
O
1
N
3

_
,

1
]
1
+

1
1
N
+O
1
N
3

_
,

(14)
in which

O( ) means "term of the order of ( )." Thus we have arrived
at Eq. 7.7-13 at the end of the example.
63
Note to p. 241
In the text it is shown how to get the function

q

y
( ) from data
obtained with a cone-and-plate viscometer. In the absence of this
kind of device, one may have to extract the function from tube-flow
data. Show how to get the relation for the non-Newtonian viscosity

q

y
( ) from experimental data on mass rate of flow

w vs pressure
drop

P
0
P
L
for flow through circular tubes.
We know from 2..3 that, for any kind of fluid

t
rz
t
R
r R,
where

t
R
P
0
P
L
( )
R 2L is the shear stress at the tube wall

r R.
The mass rate of flow through the tube is

w 2rp v
z
rdr
0
R
[
2rp
dv
z
dr

_
,

r
2
2

_
,

dr
0
R
[
+rp

y r
2
dr
0
R
[
(1)
the second form being obtained by an integration by parts. In the
third form we have made the replacement

y dv
z
dr .
Next we change the variable of integration from r to

t
rz

w
rR
3
p

1
t
R
3

yt
rz
2
0
t
R
[
dt
rz
(2)
In this equation, the shear rate

y is to be regarded as a function of the
shear stress. Equation 2 tells us that data taken in tubes of different
lengths and radii should collapse onto a single curve when plotted as

w rR
3
p vs

t
R
. If now we multiply Eq. 2 by

t
R
3
and then differentiate
with respect to

t
R
, we get

d
dt
R
t
R
3
w
rR
3
p

_
,

y
R
t
R
2
(3)
where the Leibniz formula for differentiating an integral has been
used (see C.3). This is the Weissenberg-Rabinowitch equation. It tells
64
how the wall shear rate

y
R
can be obtained by differentiating the
flow-rate vs pressure-drop data.
We now put Eq. 3 into a different form. Differentiating the
product with respect to

t
R
gives

w
rR
3
p
3t
R
2
+ t
R
3
d
dt
R
w
rR
3
p

_
,

y
R
t
R
2
(4)
Then division by

t
R
2
w rR
3
p
( )
gives

3 +
dln w rR
3
p
( )
dlnt
R
y
R
1
w rR
3
p
(5)
Then finally

q

y
R
( )

R
y
R

R
w rR
3
p
3 +
dln w rR
3
p
( )
dlnt
R

1
]
1
1
1
(6)
Equation 6 gives the viscosity as a function of shear rate at the wall
from experimental measurements of w and

P
0
P
L
. If we assume
that

q t
R
( ) at the wall is the same as

q

y
( ) throughout the tube, then
Eq. 6 gives

q

y
( ) . This assumption seems to be valid for typical
polymeric fluids. It would not, however, be expected to hold for
suspensions of fibers, because of the change in the fluid
microstructure near the wall. The above analysis is applicable only
when there is no appreciable viscous heating.
65
Note to p. 242
By working through all the intermediate steps in Example 8.3-1
we get a better idea how to solve problems involving the power-law
model. First we have to obtain the expression for the scalar

y that
appears in Eq. 8.3-5:

y
1
2

y :

y
( )

1
2

y
ij

y
ji
j1
3
_
i1
3
_
and

y Vv + Vv ( )

(1,2)
where i and j take on the values 1 = r, 2 = , and 3 = z, since we are
dealing with cylindrical coordinates. The components of

y in
cylindrical coordinates may be obtained from Eqs. (S) to (AA) in
Table A.7-2. Since the only component of v that is nonzero is the z-
component and since that is a function of r only, the only components
of

y that we need are the rz- and zr-components,

y
rz
Vv
( )
rz
+ Vv
( )
zr
ov
z
or + ov
r
oz ov
z
or + 0 (3)

y
zr
Vv
( )
zr
+ Vv
( )
rz
ov
r
oz + ov
z
or 0 + ov
z
or (4)
Therefore

y
1
2
y : y ( ) dv
z
dr ( )
2
dv
z
dr (5)
where the minus sign must be chosen, since

dv
z
dr is negative. Then
the shear stress

t
rz
will be

t
rz
m
dv
z
dr

_
,

n1
dv
z
dr

dv
z
dr

_
,

n
(6)
By going through the above procedure, it is guaranteed that, when
we take the fractional powers of the quantities in parentheses (see
Table 8.3-2 for sample values of n), we will not get any imaginary
quantities. When Eq. 6 is substituted into Eq. 2.3-13, we then get:
66

m
dv
z
dr

_
,

P
0
P
L
2L

_
,

r or

dv
z
dr

P
0
P
L
2mL

_
,

1 n
r
1 n
(7,8)
Integration of Eq. 8 gives

v
z

P
0
P
L
2mL

_
,

1 n
r
1 n ( )+1
1 n
( )
+1
+ C (9)
Application of the no-slip condition requires that

v
z
0 at

r R:

0
P
0
P
L
2mL

_
,

1 n
R
1 n ( )+1
1 n
( )
+1
+ C (10)
Subtraction of Eq. 10 from Eq. 9 eliminates the integration constant C
and leads to

v
z

P
0
P
L
2mL

_
,

1 n
r
1 n ( )+1
R
1 n ( )+1
1 n
( )
+ 1
(11)
Rearrangement then gives

v
z

P
0
P
L
( )
R
2mL

_
,

1 n
R
1 n
( )
+ 1
1
r
R

_
,

1 n ( )+1

1
]
1
1
(12)
The mass rate of flow w is then obtained by integrating the velocity
distribution over the tube cross-section:

w p v
z
0
R
[
0
2r
[
rdrd

2rR
2
p
P
0
P
L
( )
R
2mL

_
,

1 n
R
1 n
( )
+1
1
1 n ( )+1
( )
0
1
[
d (13)
where

r R. Performing the integration gives
67

w 2rR
2
p
P
0
P
L
( )
R
2mL

_
,

1 n
R
1 n
( )
+ 1
1 n
( )
+ 1
2 1 n
( )
+ 3

1
]

_
,


rR
3
p
P
0
P
L
( )
R
2mL

_
,

1 n
1
1 n
( )
+ 3
(14)
This is the power-law analog of the Hagen-Poiseuille equation for
Newtonian fluids.
68
Note to p. 248
Here we work through the missing steps in Example 8.4-2. In
Eq. 8.4-20, we make the change of variable

s t t after the first line
in order to get a factor

e
iut
to appear explicitly on the right side of the
equation (to match a similar factor on the left side):

pJ iuv
0
e
iut
{
J
d
2
v
0
dy
2
e
iut
q
0
i
1
e
s i
1
e
ius
ds
0
~
[

(1)
Next, we perform the integration over s:

pJ iuv
0
e
iut
{
J
d
2
v
0
dy
2
e
iut
q
0
i
1
1
1 i
1
( )
iu

_
,

e
s i
1
e
ius
0
~


J
d
2
v
0
dy
2
e
iut
q
0
1 + ii
1
u

_
,

(2)
We may now remove the real-operator sign from both sides, as well
as the common multiplier

e
iut
, to get:

piuv
0

d
2
v
0
dy
2
q
0
1 + ii
1
u

_
,

or

d
2
v
0
dy
2

piu 1 + ii
1
u
( )
q
0

1
]
1
v
0
(3a,b)
Eq. 3b is a differential equation for the complex function

v
0
y
( ) . The
equation is of the form of Eq. C.1-4, where

[ ]
is

a
2
. Since this is a
complex quantity, we write it as

o + i ( )
2
, where o and are real, so
that we can write the solution as

v
0
Ae
+ o+i ( )y
+ Be
o+i ( )y
(4)
Then according to Eq. 8.4-19,
69

v
x
y, t ( ) J v
0
y ( )e
iut
{
J Ae
+ o+i ( ) y
+ Be
o+i ( )y
( )
e
iut
{
(5)
Now we know that the oscillatory disturbance will not propagate
with increasing amplitude as the distance from the wall increases, so
that A will have to be set equal to zero and o will have to be positive.
Also we know that the amplitude of the disturbance right at the wall
will be

v
0
. Therefore B will have to be set equal to

v
0
. Therefore Eq. 5
can be rewritten as

v
x
y, t ( ) J v
0
e
o+i ( )y
e
iut
{
v
0
e
oy
J e
i uty ( )
{
v
0
e
oy
cos ut y ( )
(6)
Now we must find out what o and are. To do this, we have to go
back to Eq. 3:

piu 1 + ii
1
u ( )
q
0
o + i ( )
2
(7)
We next equate the real and imaginary parts of both sides:

o
2

2

pi
1
u
2
q
0
and

2o
pu
q
0
(8, 9)
This gives us two equations for the two unknowns o and . We can
eliminate between the two equations and get an equation for o :

o
2

pu
2q
0

_
,

2
1
o
2

pi
1
u
2
q
0
(10)
or, after multiplying through by

o
2

o
4
+
pi
1
u
2
q
0
o
2

pu
2q
0

_
,

2
0 (11)
70
This can be regarded as a quadratic equation in

o
2
, which has the
solution:

o
2

pi
1
u
2
2q
0

1
2
pi
1
u
2
q
0

_
,

2
+ 4
pu
2q
0

_
,

2


pi
1
u
2
2q
0
1 1 +
1
i
1
2
u
2

1
]
1
1

pu
2q
0
i
1
u 1 + i
1
2
u
2

1
]
1
(12)
We now extract the square root of both sides to get:

o
pu
2q
0
i
1
u 1 + i
1
2
u
2

1
]
1
1 2
(13)
Because o must be real and positive, the sign must be chosen to be
+, and the

sign must also be chosen to be "plus." Thus we finally
get:

o +
pu
2q
0
i
1
u + 1 + i
1
2
u
2

1
]
1
1 2
(14)

+
pu
2q
0
i
1
u + 1 + i
1
2
u
2

1
]
1
1 2
(15)
However, we could just as well have chosen both of the signs to be
"minus." Then we would have

o
pu
2q
0
1 + i
1
2
u
2
i
1
u

1
]
1
1 2
(16)


pu
2q
0
1 + i
1
2
u
2
i
1
u

1
]
1
1 2
(17)
The quantity o given in Eq. 16 is still real and positive. So, how do
we make a choice between Eqs. 14 and 15 on the one hand, and Eqs.
16 and 17 on the other? It is easy to show that the former do not
satisfy Eq. 8, whereas the latter do. Therefore, we conclude that Eqs.
71
16 and 17 are the proper quantities to insert in Eq. 6. Thus Eqs. 8.4-24
and 25 of the textbook are correct.
72
Note to p. 250
The simplest nonlinear viscoelastic model for polymers is the
corotational Maxwell model, which is obtained by replacing the partial
time derivative in Eq. 8.4-3 by the Jaumann derivative of Eq. 8.5-2.
This gives

t + i
D
Dt
t q
0
y (1)
This equation contains just two parameters:

q
0
, the zero-shear-rate
viscosity, and i , the relaxation time.
Here we show how to get the viscosity, the first normal-stress
coefficient, and the second normal-stress coefficient for the model
given in Eq. (1). The imposed flow is

v
x

y y with

v
y
0 and

v
z
0.
In learning how to analyze and evaluate nonlinear viscoelastic
models, it is useful to display the various parts of the models in
matrix form. We start by getting

Vv,

Vv ( )

,

y , and u for the flow
field being considered.

Vv
0 0 0
1 0 0
0 0 0

_
,

y

Vv ( )

0 1 0
0 0 0
0 0 0

_
,

y (2,3)

y
0 1 0
1 0 0
0 0 0

_
,

y

u
0 1 0
1 0 0
0 0 0

_
,

y (4,5)
Since the stress tensor t does not depend on position and time, in the
Jaumann derivative of t , the substantial derivative will not
contribute and we are left with only

1
2
u t t u
{ .
73

u t {
0 1 0
1 0 0
0 0 0

_
,

t
xx
t
xy
t
xz
t
yx
t
yy
t
yz
t
zx
t
zy
t
zz

_
,

y
t
yx
t
yy
t
yz
t
xx
t
xy
t
xz
0 0 0

_
,

y (6)

t u {
t
xx
t
xy
t
xz
t
yx
t
yy
t
yz
t
zx
t
zy
t
zz

_
,

0 1 0
1 0 0
0 0 0

_
,

y
t
xy
t
xx
0
t
yy
t
xy
0
t
zy
t
xz
0

_
,

y (7)
Hence the Jaumann derivative is (keep in mind that the stress tensor
is symmetric)

D t
D t

t
yx
1
2
t
xx
t
yy
( )

1
2
t
yz
1
2
t
xx
t
yy
( )
t
yx
1
2
t
xz

1
2
t
yz
1
2
t
xz
0

_
,

y (8)
Thus the corotational Maxwell model in matrix form is

t
xx
t
xy
t
xz
t
yx
t
yy
t
yz
t
zx
t
zy
t
zz

_
,

+ i
t
yx
1
2
t
xx
t
yy
( )

1
2
t
yz
1
2
t
xx
t
yy
( )
t
yx
1
2
t
xz

1
2
t
yz
1
2
t
xz
0

_
,

y

q
0
0 1 0
1 0 0
0 0 0

_
,

y (9)
From this matrix equation, we can write down the following
algebraic equations

t
xx
t
yy
( )
2t
yx
i

y 0 (10)

t
yy
t
zz
( )
+t
yx
i

y 0 (11)
74

t
yx
+
1
2
t
xx
t
yy
( )
i

y q
0

y (12)

t
xz

1
2
t
yz
i y 0 (13)

t
yz
+
1
2
t
xz
i

y 0 (14)
From the last two equations we see that

t
xz
t
yz
0. From solving
the first three simultaneous equations for

t
yx
,

t
xx
t
yy
, and

t
yy
t
zz
we find

t
yx
q
0
y
1
1 + i y ( )
2
or

q
q
0

1
1 + i y ( )
2
(15,16)

t
xx
t
yy
q
0
y
2i y
1 + i y ( )
2
or

+
1
q
0
i

2
1 + i y
( )
2
(17,18)

t
yy
t
zz
+q
0
y
i y
1 + i y ( )
2
or

+
2
q
0
i

1
1 + i y
( )
2
(19,20)
We know that the shear stress keeps increasing as the velocity
gradient increases, but Eq. 16 indicates otherwise. In fact, the shear
stress goes through a maximum at

i y = 1. The second normal stress
is negative, and that is in agreement with measurements for
polymeric liquids. However, generally one expects

+
2
+
1
to be in
the range from

0.1 to

0.4 . Hence from examination of this one
particular flow, it is evident that caution must be used when drawing
any conclusions from this model. However, for a model with just two
parameters, it seems to be promising for predicting qualitative
behavior.
75
Note to p. 255
Since the FENE-P dumbbell model (a molecular model)
describes many of the observed rheological phenomena of polymer
solutions, it is important to understand how we go from the
constitutive equation in Eqs. 8.6-2 to 4 to the expressions for the
rheological properties. Here we show how to get the non-Newtonian
viscosity and the first normal-stress coefficient in steady shear flow.
For the steady shear flow,

v
x

y y, the rate-of-strain tensor



y
may be displayed as a matrix (see A.9) thus:

y
xx

y
xy

y
xz

y
yx

y
yy

y
yz
y
zx
y
zy
y
zz

_
,

y 0
0 0 0
0 0 0

_
,

(1)
Then Eq. 8.6-4 may be written in matrix form as follows:

Z
t
p,xx
t
p, xy
t
p, xz
t
p, yx
t
p, yy
t
p, yz
t
p, zx
t
p, zy
t
p, zz

_
,

i
H

y
2t
p,xy
t
p, yy
t
p, yz
t
p, yy
0 0
t
p, yz
0 0

_
,


nKTi
H

y
0 1 0
1 0 0
0 0 0

_
,

(2)
From this we get at once that the only nonzero components of

t
p
are

t
p, xx
and

t
p, xy
=

t
p, yx
. Then from the matrix equation, Eq. 2, we can
write down the two nonvanishing component equations as:

Zt
p, xx
2t
p, yx
i
H

y and

Zt
p, yx
nKTi
H

y (3,4)
in which

Z 1 + 3 b ( ) 1 tr t
p
3nKT
( )

1 + 3 b ( ) 1 t
p, xx
3nKT
( )

(5)
76
In order to make the algebraic manipulations somewhat easier, we
introduce the following dimensionless quantities:
Dimensionless shear stress:

S t
p, yx
3nKT (6)
Dimensionless normal stress:

N t
p, xx
3nKT (7)
Dimensionless shear rate:

i
H

y (8)
Then Eqs. 3, 4, and 5 become:

ZN 2S (9)

ZS
1
3
(10)

Z 1 + 3 b
( )
1 N
( ) (11)
Use the second of these equations to eliminate Z and rewrite Eqs. 9
and 11 thus:

N
3S
2S and


3S
1 +
3
b
1 N
( ) (12,13)
Then N may be eliminated between these two equations to get a cubic
equation for the dimensionless shear stress S (for method of solving
cubic equations, see a mathematics handbook):

S
3
+
b + 3
18
S +
b
54
0 or

S
3
+ 3pS + 2q 0 (14,15)
Eqs. 14 and 15 serve to define p and q. Eq. 15 has three solutions, two
imaginary solutions (of no interest here) and one real solution:

S 2p
1 2
sinh
1
3
arcsinh p
3 2
q
( )
(16)
or, in terms of the original variables:
77

t
p, yx
3nKT
2
b + 3
54

_
,

+1 2
sinh
1
3
arcsinh
b + 3
54

_
,

3 2
b
108
i
H

1
]
1
1
(17)
From this result one can plot the non-Newtonian viscosity as a
function of the shear rate.
If one wants only the limiting values of the non-Newtonian
viscosity at zero and infinite shear rates, this information can be
obtained from Eq. 17 or 16. Very small shear rate corresponds to very
small q, so that

S 2p
1 2
sinh
1
3
p
3 2
q

1
]
( )
2p
1 2
1
3
p
3 2
q

1
]
+
{

2
3
p
1
q
(18)
if we keep only the terms linear in q in the Taylor series expansion of
the hyperbolic sine and the arc hyperbolic sine functions:

sinh x x +
1
6
x
3
+

arcsinhx x
1
6
x
3
+ (19,20)
In terms of the original variables, Eq. 18 is

t
p, yx
3nKT
2
3
b + 3
54

_
,

1
b
108
i
H

y nkT
b
b + 3

_
,

i
H

y (21)
This corresponds to

q q
s
nkTi
H
b
b + 3

_
,

(22)
in the limit of zero velocity gradient.
In the limit of infinite velocity gradient, we make use of the fact
that for large values of the argument, sinh x -

1
2
exp x (see Appendix
C.5). Therefore

S 2p
1 2
sinh
1
3
ln 2p
3 2
q
( ) ( )
2p
1 2
1
2
exp
1
3
ln 2p
3 2
q
( ) ( )
2q
( )
1 3
(23)
78
In the original variables, this becomes

t
p, yx
3nKT 2
b
108
i
H

_
,

1 3
nKT 54
b
108
i
H

_
,

1 3
(24)
This corresponds to a "power-law function":

q q
s
nKTi
H
b
2
1
i
H
y

_
,

1 3
(25)
in the infinite velocity gradient limit.
Now the results in Eqs. 22 and 25 can be obtained directly from
Eq. 15 in another way. If the velocity gradient (and hence, q) is quite
small, the cubic term in S may be omitted, and one gets then

S
2
3
p
1
q (26)
directly (see Eq. 18). Similarly, if the velocity gradient is quite large,
then the linear term in S may be omitted, and one gets

S 2q
( )
1 3
(27)
immediately (see Eq. 23).
From Eq. 17, one can also find the molecular stretching as a
function of the shear rate, as described at the top of p. 255. In
addition, from Eq. 3 the first normal stress coefficient can be obtained

+
1

2 q q
s
( )
2
nKT
(28)
this formula being predicted for the entire shear-rate range.
[Note: For more information on the FENE-P dumbbell model, see
Teaching with FENE Dumbbells, by R. B. Bird, Rheology Bulletin,
January 2007.]
79
Note to p. 259
The Bingham fluid is not the only empirical equation with a
yield stress. Closely related to the Bingham equation is the two-
constant Casson equation, proposed for pigment-oil suspensions, but
widely used for blood
1,2,3
q ~ when

t s t
0
(1)

q
0
+
t
0
y
when

t > t
0
(2)
in which

t
0
is the yield stress, and

0
is a parameter with dimensions
of viscosity. Derive the expression for the mass rate of flow of a
Casson fluid through a circular tube of radius R, and length L.
1
N. Casson, Rheology of Disperse Systems, C. C. Mill, ed., Pergamon, London
(1959), pp. 84-104.
2
M. M. Lih, Transport Phenomena in Medicine and Biology, Wiley, New York (1975),
378-38.
3
E. N. Lightfoot, Transport in Living Systems, Wiley, New York (1974), pp. 35, 430,
438, 440.
The expression for the shear stress in a circular tube for any
kind of fluid was derived in Chapter 2 and found to be

t
rz

P
0
P
L
2L

_
,

r (3)
At some radius

r
0
, the yield stress will be exceeded. The region

0 s r s r
0
will be a plug-flow region, which moves as a solid, and

r
0
is
defined by

t
0
P
0
P
L
( ) 2L
( )
r
0
. We further note that

y dv
z
dr
for flow in circular tubes, according to the definition given just before
Eq. 8.3-2. For the region

r
0
s r s R, we then get from Eq. 2

q y
0
y + t
0
or

t
rz

0
y + t
0
(4a,b)
80
Combining Eqs. 3 and 4b and rearranging gives

P
0
P
L
( )r
2L

0
y + t
0
(5)
Solving Eq. 5 for

y we get

y
1

0
P
0
P
L
( )
r
2L
2
P
0
P
L
( )
rt
0
2L
+ t
0

1
]
1
1

t
0

0
r
r
0
2
r
r
0
+ 1

_
,

(6)
Inserting

y dv
z
>
dr , we can get a differential equation for the
dimensionless velocity profile

o
>

0
r
0
t
0
( )
v
z
>
as a function of the
dimensionless radial coordinate

r r
0

do
>
d
2 + 1 (7)
Integration of this equation, and using the boundary condition that

o
>
0 at

R r
0
, we get for

r
0
s r s R

o
>

1
2
R
r
0

_
,

2
1
r
R

_
,

1
]
1
1

4
3
R
r
0

_
,

3 2
1
r
R

_
,

3 2

1
]
1
1
+
R
r
0
1
r
R

_
,

(8)
and for

0 > r > r
0

o
<

1
2
R
r
0

_
,

4
3
R
r
0

_
,

3 2
+
R
r
0

1
6
(9)
The mass rate of flow can then be obtained by using Eq. 2.3-21 after
doing an integration by parts. The dashed underlined terms are zero.
81

w 2rp
1
2
v
z
r
2
r0
rR

1
2
r
2
dv
z
dr
dr
0
R
[

_
,


rp r
2
dv
z
dr
dr + r
2
dv
z
dr
dr
r
0
R
[
0
r
0
[

_
,

----------- --------------

rp
r
0
3
t
0

_
,


2
do
>
d
d
1
R r
0
[
+ rp
r
0
3
t
0

_
,


2
2 + 1
( )
d
1
R r
0
[

rp
r
0
3
t
0

_
,

1
4
R
r
0

_
,

4
1

1
]
1
1

4
7
R
r
0

_
,

7 2
1

1
]
1
1
+
1
3
R
r
0

_
,

3
1

1
]
1
1


rp
r
0
t
0
R
4

_
,

1
16
7
r
0
R

_
,

1 2
+
4
3
r
0
R

_
,


1
21
r
0
R

_
,

1
]
1
1

r P
0
P
L
( )
R
4
p
8
0
L
1
16
7
2Lt
0
P
0
P
L
( )
R
+
4
3
2Lt
0
P
0
P
L
( )
R

1
21
2Lt
0
P
0
P
L
( )
R

_
,

1
]
1
1
1
1
1
(10)
When

t
0
0 and

0
, the Newtonian result is recovered.
82
Note to p. 275
As an exercise in identifying the dimensions of thermal
quantities, we verify that the following equations are dimensionally
consistent: Eq. 9.3-12, Eq. 9.8-6, and Eq. 9.8-8. We do this by
replacing the symbols in these equation by the corresponding
dimensions, making use of the "Notation" table on pp. 872 et seq.
(a)

ML
t
3
T

_
,


M
( )
ML
2
t
2
T

_
,

T
( )
L
2
ML
2
t
2
T

_
,

M
(1)
(b)

M
t
3

_
,


M
L
3

_
,

L
t

_
,

2
L
t

_
,

+
M
L
3

_
,

L
2
t
2

_
,

L
t

_
,

+
M
Lt
2

_
,

L
t

_
,

+
M
t
3

_
,

(2)
(c)

L
2
t
2

_
,


L
2
t
2
T

_
,

T ( ) +
L
3
M

_
,

M
Lt
2

_
,

+ T ( )
L
3
M

_
,

T
( )
M
Lt
2

_
,

(3)
In each of these cases, each term has the same dimensions.
83
Note to p. 286
We illustrate the use of Eq. 9.8-8 for an ideal monatomic gas.
For the ideal monatomic gas, the heat capacity at constant pressure is
given by

C
p

5
2
R
M
(1)
as given two lines after Eq. 9.3-15.
The ideal gas law equation of state is

p

V
RT
M
(2)
Therefore, the bracket in Eq. 9.8-8 is

V T
o

V
oT

_
,

V T
o
oT
RT
pM

_
,

_
,

V T
R
pM

_
,

0 (3)
Hence, the only term that survives is the heat-capacity term, which is:

H
o

5
2
R
M
T T
o
( )
(4)
84
Note to p. 299
In connection with Fig. 10.4-2, we want to find the location of
the maximum in the temperature vs. distance curve. We will express
the result in terms of the Brinkman number, Br. We also need to
know what happens when the Brinkman number goes to zero (i.e.,
negligible viscous heating).
To simplify the discussion, we introduce the following
dimensionless variables:

O
T T
0
T
b
T
0
(temperature)


x
b
(distance) (1)
so that Eq. 10.4-9 becomes:

O
1
2
Br 1
( )
+ (2)
To get the location of the maximum of the temperature, we
differentiate with respect to and set the derivative

dO d equal to
zero:

dO
d

1
2
Br 1 2
( )
+ 1 0 (3)
From this we get the location of the maximum in the temperature
curve:

max

1
2
+
1
Br
(4)
When there is negligible viscous heating,

Br 0 , and, according to
Eq. 4,

max
~. This result is nonsense, since the system extends only
to

1. Therefore, Eq. 4 has to be restricted to

max
s 1, and when

max
1, Br = 2. Alternatvely, Eq. 4 has to be limited to

2 s Br<~.
85
Note to p. 309
It's always a good idea to check the solutions to problems. Here
we verify that the expression in Eq. 10.7-13 for the dimensionless
temperature in the cooling fin satisfies the differential equation in Eq.
10.7-9, and the boundary conditions in Eqs. 10.7-10 and 11.
First calculate the derivatives from Eq. 10.7-13 and Eqs. C.5-10
and 11:

dO
d

sinh N 1
( )
cosh N
N
( )
(1)

d
2
O
d
2

cosh N 1
( )
cosh N
+N
2
( )
(2)
When the expression for O (in Eq. 10.7-13) and its second derivative
(in Eq. 2) are substituted into Eq. 10.7-9, an identity results. Therefore,
Eq. 10.7-13 satisfies the differential equation.
The boundary condition at

0 is satisfied, since

O
cosh N 1
( )
cosh N
0
1 (3)
and the boundary condition at

1 is also satisfied, since

dO
d
1

sinh N 1 ( )
cosh N
N ( )
1
0 (4)
The hyperbolic sine is shown in Figure C.5-2.
86
Note to p. 315
Here we verify the determination of the constants of integra-
tion,

C
0
,

C
1
, and

C
2
in Eq. 10.8-27.

C
1
: Boundary condition 1 requires that

O 0,
( )
finite:

O 0,
( )
C
0
+ C
1
ln0 + C
2
(1)
This can be satisfied only if

C
1
0.

C
0
: Boundary condition 2 requires that

oO o 1at

1:

oO
o
1
C
0
2
4

4
3
16

_
,

1
4
C
0
(2)
from which it follows that

C
0
4 .

C
2
: The dimensionless form of condition 4 is given in Eq. 10.8-25.
Substitution of Eq. 10.8-27 into that equation gives:

4 + 4

2
4


4
16

_
,

+ C
2

1
]
1
1
0
1
[
1
2
( )
d (3)
Next we evaluate the integrals:

4 1
2
( )
d 4
3
( )
d 4
1
2

1
4

_
,

0
1
[
0
1
[
(4)

4

2
4


4
16

_
,

0
1
[
1
2
( )
d 4

3
4

5
5
16
+

7
16

1
]
1
0
1
[
d

4
1
16

5
6 16
+
1
8 16

_
,


7
24 4
(5)
87

C
2
1
2
( )
d
0
1
[
1
4
C
2
(6)
Combining the results of Eq. 3 to 6, we find that

C
2
7 24.
88
Note to p. 337
In the textbook, we transformed Eq. 11.2-2 into an equation for
the enthalpy (Eq. 11.2-3) and then used an equilibrium thermo-
dynamic formula to get the equation of energy in terms of

C
p
and T,
given in Eq. 11.2-5 (and also Eq. (J) in Table 11.4-1). Specifically, we
used the equation for

H T, p
( ) :

d

H
o

H
oT

_
,

p
dT +
o

H
op

_
,

T
dp

C
p
dT +

V T
o

V
oT

_
,

1
]
1
1
dp (1)
appropriately rewritten for a "particle" of fluid moving with the local
velocity v.
Alternatively, we can begin with Eq. 11.2-2 and use another
equilibrium thermodynamic formula to get the equation of energy in
terms of

C
V
and T (see Eq. (I) in Table 11.4-1). Specifically, we use
the equation for

U T,

V
( )
:

d

U
o

U
oT

_
,

V
dT +
o

U
o

_
,

T
d

C
V
dT + p + T
op
oT

_
,

1
]
1
d

V (2)
Then we can use this equation to obtain Eq. (I) in Table 11.4-1.
First we rewrite Eq. 2 for a fluid particle moving with the fluid,
and then we multiply by the density of the fluid. This gives:

p
D

U
Dt
p

C
V
DT
Dt
+ p p + T
op
oT

_
,

1
]
1
D

V
Dt
(3)
Since

V 1 p , we may rewrite

pD

V Dt as follows:

p
D

V
Dt
p
D
Dt
1
p

_
,

p
1
p
2

_
,

Dp
Dt
(4)
89
Then if we use Eq. (A) of Table 3.5-1, we may rewrite this last relation
as:

p
D

V
Dt
p
1
p
2

_
,

p V v
( ) ( )
V v
( ) (5)
Now substitute this expression into Eq. 3, and then make use of Eq.
11.2-2 to get:

p

C
V
DT
Dt
+ p + T
op
oT

_
,

1
]
1
V v
( )
V q
( )
p V v
( )
t:Vv
( ) (6)
When the terms involving

V v
( ) are moved to the right side, there is
some cancellation and Eq. (I) of Table 11.4-1 is obtained:

p

C
V
DT
Dt
V q ( ) T
op
oT

_
,

V
V v ( ) t:Vv ( )
(7)
[Note: See footnote 1 at the bottom of p.338 for a discussion of
"incompressible fluids."]
90
Note to p. 341
Verify that Eq. (T) of Table 11.4-1 is the equation of change for
entropy.
We can start with the thermodynamic expression

d

U Td

S pd

V (1)
which is written for a small isolated mass of fluid that is stationary.
For a small mass of fluid that is moving with the fluid, we can then
write:

D

U
Dt
T
D

S
Dt
p
D

V
Dt
(2)
where we are now assuming that this expression can be applied
locally. When this equation is multiplied by p and then combined
with Eq. (G) of the table we get:

V q
( )
p V v
( )
t:Vv
( )
pT
D

S
Dt
pp
D

V
Dt
(3)
Next, divide by T and replace

V by

1 / p to get

p
D

S
Dt

1
T
V q
( )

1
T
p V v
( )

1
T
t:Vv
( )
+
pp
T
D
Dt
1
p

_
,

(4)
The second and fourth terms on the right side may be seen to cancel
one another is one makes use of the equation of continuity in the
form of Eq. (A) of Table 3.5-1, so that we get

p
D

S
Dt

1
T
V q
( )

1
T
t:Vv
( ) (5)
When we make use of Eq. 3.5-4, Eq. 5 becomes
91

o
ot
p

S + V p

Sv
( )

1
T
V q
( )

1
T
t:Vv
( ) (6)
This is equivalent to Eq. (T) in Table 11.4-1. Eq. (T) is written in a
form that emphasizes the entropy flux

q / T
( ) and the entropy
production (the terms in brackets:

o
ot
p

S V p

Sv
( )
V
q
T

_
,

+
1
T
2
q VT
( )

1
T
t:Vv
( )

1
]
1
(7)
See Problem 11D.1 and 24.1 for more on this subject.
92
Note to p. 343
Verify that Eq. 11.4-14 gives the location of the maximum
temperature. We'd also like to know whether the maximum is nearer
the inner cylinder or the outer.
Differentiate O in Eq. 11.4-13 with respect to and set the
derivative equal to zero:

dO
d

1
lnx
+ N
2

3
1
1
x
2

_
,

1
lnx

1
]
1
0 (1)
Solving for

2
2
gives:

2

2

1
Nlnx
+ 1
1
x
2

_
,

1
lnx
(2)
whence, the location of the maximum temperature rise is

max

2ln 1 x
( )
1 x
2
( )
1 1 N
( )
(3)
Now is

max
greater than or less than

1
2
x +1
( ) ? Clearly this cannot be
answered until N is known; but N depends on the geometry, the
temperature difference, the thermal conductivity, and the viscosity. If
we take N to be infinity, then we get:
x

1
2

1
3

1
4

1
2
x +1
( ) 0.75 0.67 0.625

max
0.67 0.51 0.42
This suggests that the maximum occurs nearer the inner wall, where
the velocity gradient is larger.
93
Note to p. 346
Here we want to derive Eq. 11. 4-27 from Eq. 11.4-26 by follow-
ing the instructions in the text. First we rewrite Eq. 11.4-26 in terms of
dimensionless variables:

O
T T
1
T
x
T
1
= dimensionless temperature (1)


r
R
= dimensionless radial coordinate (2)
Then Eq. 11.4-26 becomes:

dO
d

R
R
0
d
d

2
dO
d

_
,

where

R
0

w
r

C
p
4rk
(3a,b)
Now introduce the change of variable

u
2
dO d
( ), and rewrite the
differential equation as:

u

2

R
R
0
du
d
(4)
This equation may be solved to get:

R
R
0
lnu +
R
R
0
lnC
1
or

C
1
u exp
R
0
R
1

_
,

(5a,b)
where the constant of integration has been written as

R R
0
( )
lnC
1
.
Reverting to the original variable O gives:

C
1

2
dO
d
exp
R
0
R
1

_
,

(6)
Further integration then gives:
94

C
1
O
2
exp
R
0
R
1

_
,

1
[
d + C
2

exp
R
0
R
v

_
,

1
v
[
dv + C
2

e
R
0
R ( )v
e
R
0
R ( )
R
0
R
( )
+ C
2

e
R
0
R ( )
e
R
0
R ( )
R
0
R
( )
+ C
2
(7)
The constants of integration may be determined from the boundary
conditions that: at x ,

O 1, and at

1,

O 0. From the second
of these boundary conditions it is evident that

C
2
0. From the first
boundary condition,

C
1
may be found. Then the final expression for
the dimensionless temperature is:

O
e
R
0
R ( )
e
R
0
R ( )
e
R
0
Rx ( )
e
R
0
R ( )
(8)
which agrees with Eq. 11.4-27.
95
Note to p. 352
Fill in the missing steps, showing how to arrive at Eqs. 11.4-67, 11.4-
68, and 11.4-75.
We begin by multiplying Eq. 11.4-66 by

v
x
p
1
v
1
to get

4
3

p
1
v
1
v
x
dv
x
dx
v
x
2
+
pv
x
p
1
v
1

C
III
v
x
p
1
v
1
(1)
Eq. 11.4-61 may be used to simplify the second term on the right side
to

p p, and this may be further rearranged by using ideal gas
relations, thus:

p
p

RT
M

C
p


C
V
( )
T
M

C
p

C
V
( )
T
y 1
y

_
,

C
p
T
y 1
y

_
,

C
I

1
2
v
x
2
( )
(2)
In the last step, use has been made of Eqs. 11.4-64 and 65, the second
of which should have been written as

C
I

C
p
T
1
+
1
2
v
1
2
. Then we get Eq.
11.4-67

4
3

p
1
v
1
v
x
dv
x
dx

y +1
y

_
,

v
x
2
+
y 1
y

_
,

C
I

C
III
v
x
p
1
v
1
(3)
To put this equation in dimensionless form, we insert the expressions
for the constants of integration, multiply the equation by

3ip
1
4v
1
and then introduce the quantities defined in Eqs. 11.4-69,
70, 71, 73 and 74, as follows:

o
do
d

3ip
1
4v
1
y + 1
2y

_
,

v
x
2
p
1
v
1
2
+
RT
1
M

_
,

v
x
p
1
v
1
+
y 1
y

_
,

C
p
T
1
+
1
2
v
1
2
( )

1
]
1
96

Ma
1
i
Ma
1

3p
1
4v
1

_
,

y + 1
2y

_
,

v
1
2
o
2
v
1
2
+
RT
1
M

_
,

o +
y 1
y

_
,

C
p
T
1
+
1
2
v
1
2
( )

1
]
1

Ma
1
2y
y +1

_
,

y +1
2y

_
,

o
2
1 +
RT
1
Mv
1
2

_
,

o +
y 1
y

_
,

C
p
T
1
v
1
2
+
1
2

_
,

1
]
1
1

Ma
1
o
2

2y
y +1

_
,

1 +
1
y Ma
1
2

_
,

o + 2
y 1
y + 1

_
,

1
y 1
1
Ma
1
2
+
1
2

_
,

1
]
1
1
(4)
The coefficient of o may be seen by Eq. 11.4-72 to be

2o and the
constant term is + o . Therefore, Eq. 4 finally becomes

o
do
d
Ma
1
o 1
( )
o o
( )
(5)
which is in agreement with Eq. 11.4-68. This equation is separable

odo
o 1
( )
o o
( )
Ma
1
d
[ [
Ma
1

0
( ) (6)
where

0
is the constant of integration. The integral on the left side is
to be found on p. 29 of Table of Integrals and Other Mathematical Data,
H. B. Dwight, MacMillan, New York, 4th edition (1961). There we
find

xdx
a + x
( )
c + x
( )
[

1
a c
aln a + x c ln c + x

(7)
For the problem at hand

a 1 and

c o , so that

odo
o 1
( )
o o
( )

1
o 1
[
1ln 1+o + o ln o+o

1
]
Ma
1

0
( ) (8)
In order to remove the absolute value signs, we note that

o < o < 1, so
that
97

1ln 1 o ( ) +o ln o o ( )

1
]
o 1 ( )Ma
1

0
( ) (9)
so that finally

1 o
( )
o o
( )
o
exp 1 o
( )
Ma
1

0
( )

1
]
(10)
which is the same as Eq. 11.4-75.
98
Note to p. 375
a. First we'll verify the formula for differentiation of the error
function in Eq. C.6-2.
b. Then we'll show that Eq. 12.1-8 is a solution to Eq. 12.1-3 and
the associated boundary and initial conditions.
a. The differentiation of the error function is given by Eq. C.6-2:

d
dx
erf u
2
r
e
u
2

du
dx
(1)
where it is understood that u is a function of x. Now differentiate the
error function in Eq. C.6-1 with respect to x using the Leibniz formula
of C.3:

d
dx
erf u
d
dx
2
r
e
u
2
du
0
u
[

_
,

2
r
o
ox
e
u
2
du + e
u
2
ou
ox
e
0
2
o0
ox
0
u
[

_
,

(2)
Since

u is a dummy variable of integration, it is not a function of x,
and therefore the first term is zero. The last term is also zero. Hence
the second term is the only term in the parenthesis that contributes to
the derivative of erf u, and that leads directly to Eq. 1. [Note: It is
important to designate the dummy variable of integration,

u , and the
upper limit in the integral, u, by two different symbols. This example
emphasizes the importance of this statement.]
b. Now turn to Eq. 12.1-8, where the left side is the dimension-
less temperature difference O. Form the derivatives that appear in
Eq. 12.1-3:

oO
ot

2
r
e
y
2
4ot
y
4o

1
2

_
,

t
3 2

_
,

(3)

oO
oy

2
r
e
y
2
4ot
1
4ot

_
,

(4)

o
2
O
oy
2

2
r
e
y
2
4ot
1
4ot

_
,


2y
4ot

_
,

(5)
99
When the results in Eqs. 3 and 5 are inserted into Eq. 12.1-3 it is found
that the equation yields an identity. Therefore, Eq. 12.1-8 does satisfy
the differential equation.
Eq. 12.1-8 also satisfies the boundary conditions at

y 0 and

y ~, as may be seen from Fig. 4.1-2. At

t 0, Fig. 4.1-2 tells us that

O 0.
[Note: Here, and elsewhere we have made use of the Leibniz formula
for differentiating an integral. For additional information and anec-
dotes regarding the Leibniz formula see R. P. Feynman, Surely You're
Joking Mr. Feynman, Bantam Books, New York (1986), p. 72 and p. 93.
Professor Feynman was a Nobel Prize winner in physics.]
100
Note to p. 377
As pointed out in the footnote, there are two solutions to this
problem, one that converges rapidly for long times (see Eq. 12.1-31):

O 2
1
( )
n
n +
1
2
( )
r
exp n +
1
2
( )
2
r
2
t

1
]
1
n0
~
_
cos n +
1
2
( )
rq (1)
and one that converges rapidly for small times:

O 1 1
( )
n
erfc
n +
1
2
( )

1
2
q
t
+ erfc
n +
1
2
( )
+
1
2
q
t

1
]
1
1 n0
~
_
(2)
We want to verify that both of these solutions satisfy the partial
differential equation (Eq. 12.1-14), the initial condition (Eq. 12.1-15),
and the boundary conditions (Eq. 12.1-16).
a. Solution in Eq. 1:
When Eq. 1 is substituted into Eq. 12.1-14, we get by
differentiating the exponential once with respect to t and the cosine
twice with respect to q

2
1
( )
n
n +
1
2
( )
r
exp n +
1
2
( )
2
r
2
t

1
]
1
n0
~
_
cos n +
1
2
( )
rq n +
1
2
( )
2
r
2

1
]
1


2
1
( )
n
n +
1
2
( )
r
exp n +
1
2
( )
2
r
2
t

1
]
1
n0
~
_
cos n +
1
2
( )
rq n +
1
2
( )
r

1
]
n +
1
2
( )
r

1
]
(3)
and this is clearly an identity, so that the partial differential equation
is satisfied.
When t is set equal to zero and

O 1in Eq. 1, we get

1 2
1
( )
n
n +
1
2
( )
r
n0
~
_
cos n +
1
2
( )
rq (4)
101
We have to prove that this is an identity. To do this, multiply both
sides by

cos m+
1
2
( )
rq and integrate over q from 1 to +1:

cos m+
1
2
( )
rqdq 2
1
( )
n
n +
1
2
( )
r
n0
~
_
cos m+
1
2
( )
rq
1
+1
[
1
+1
[
cos n +
1
2
( )
rqdq
(5)
Then performing the integrations we get

1
m+
1
2
( )
r
sin m+
1
2
( )
rq
1
+1
2
1
( )
n
n +
1
2
( )
r
n0
~
_
o
mn
cos
2
n +
1
2
( )
rq
1
+1
[
dq
or

2
m+
1
2
( )
r
sin m+
1
2
( )
r 2
1
( )
m
m+
1
2
( )
r
(7)
This is an identity, since

sin m+
1
2
( )
r 1
( )
m
.
When q is set equal to +1 or 1, we get

O 0, since

cos m+
1
2
( )
r 0.
b. Solution in Eq. 2
When Eq. 1 is substituted into Eq. 12.1-14, we get by
differentiating the complementary error functions once with respect
to t and then twice with respect to q

o
ot
erfc
n +
1
2
( )

1
2
q
t

2
r
exp
n +
1
2
( )

1
2
q
t

_
,

1
]
1
1
n +
1
2
( )

1
2
q

1
]

1
2
( )
t
3 2
(8)

o
oq
erfc
n +
1
2
( )

1
2
q
t

2
r
exp
n +
1
2
( )

1
2
q
t

_
,

1
]
1
1

1
2

1
]
t
(9)

o
2
oq
2
erfc
n +
1
2
( )

1
2
q
t

2
r
exp
n +
1
2
( )

1
2
q
t

_
,

1
]
1
1
102


2
n +
1
2
( )

1
2
q
t

_
,

1
]
1
1

1
2

1
]
t

1
2

1
]
t
(10)
Thus, the nth term satisfies the partial differential equation, and
hence the entire series does.
At

t 0 , all of the complementary error functions are zero, so
that

O 1, and the initial condition is satisfied.
At

q 1,

O 0 as may be seen as follows. First consider

q +1:

O 1 erfc
0
t
+ erfc
1
t

_
,

erfc
1
t
+ erfc
2
t

_
,

+ erfc
2
t
+ erfc
3
t

_
,

1
]
1

(11)
Hence there is cancellation of between the nth and (n + 1)th terms,
and, since erfc 0 = 1, the dimensionless temperature is zero. A similar
argument can be made for

q 1.
103
Note to p. 379
In Example 12.1-3 going from Eqs 12.1-38 and 39 to Eq. 12.1-40
presents a few problems. Therefore we go through the details.
Substitution of Eq. 12.1-38 into Eq. 12.1-39 gives:

k T
0
T
( )
q
0
e
u 2oy
cos ut
u
2o
y

_
,

dy
y
~
[
(1)
Bars have been added to the variable of integration to distinguish it
from the lower limit on the integral. It is easier to perform the
integration if the cosine is converted into an exponential of a complex
quantity, thus:

T T
0

q
0
k
e
u 2oy
J e
iuti u 2oy
{
y
~
[
dy

q
0
k
J e
iut
e
u 2oyi u 2oy
y
~
[
dy
{

q
0
k
J e
iut
e
1+i ( ) u 2oy
1 + i ( ) u 2o
y
~

q
0
k
J e
iut
e
1+i ( ) u 2oy
1 + i
( )
u 2o

(2)
Next we remove from the braces that portion of the exponential that
is real; we also multiply numerator and denominator by

1 i
( ) . Then
we have

T T
0

q
0
k
e
u 2oy
J
e
iut
e
i u 2oy
1 i
( )
2 u 2o

q
0
2k
2o
u
e
u 2oy
J e
iut
e
i u 2oy
1 i
( )
{
(3)
104
In order to proceed, we need to rewrite

1 i
( ) in the form

re
i
. Then
we find r and as follows:

1 i re
i
r cos + i sin
( ) (4)
so that equating real and imaginary parts gives

1 r cos and

1 r sin (5,6)
Taking the ratio of these two equations we get

r sin
r cos

1
1
or

tan 1 or


3
4
r,
1
4
r (7,8,9)
Since

1 i
( ) is in the 4th quadrant, the appropriate choice is


1
4
r .
Next we square both Eq. 5 and Eq. 6

1 r
2
cos
2
and

1 r
2
sin
2
(10)
Adding the two equations then gives:

r
2
2 or

r 2 (11)
The plus sign must be chosen, since r must be non-negative.
Therefore we have shown that

1 i 2e
ir 4
(12)
Returning now to Eq. 3, we get:

T T
0

q
0
2k
2o
u
e
u 2oy
J 2e
iuti u 2oyir 4
{

q
0
k
o
u
e
u 2oy
cos ut u 2o y
1
4
r
( )
(13)
This agrees with Eq. 12.1-40 in the textbook.
105
Note to p. 386
The derivation of Eq. 12.3-6 from Eq. 12.3-5 is given here. First
we note that, if

z x + iy re
i
, then

lnz lnr + i

ln x
2
+ y
2
+ i arctan y x ( ) (1)
Now we have to resolve Eq. 12.3-5 into its real and imaginary parts.
We introduce the abbreviated notation

Z rz b,

X rx b, and

Y r y b . Then

w
1
r
ln
sinZ 1
sinZ+ 1

_
,


1
r
ln
sin X + iY
( )
1
sin X + iY
( )
+ 1

_
,

+ + iO

1
r
ln
sinXcos iY + cosXsin iY 1
sinXcos iY + cosXsin iY + 1

_
,

1
r
ln
sinXcoshY + i cosXsinhY 1
sinXcoshY + i cosXsinhY + 1

_
,

1
r
ln
sinXcoshY 1
( )
sinXcoshY + 1
( )
+ cos
2
Xsinh
2
Y
+i cosXsinhY sinXcoshY + 1
( )
cosXsinhY sinXcoshY 1
( )
sinXcoshY + 1
( )
2
+ cosXsinhY
( )
2

_
,

1
r
ln
sin
2
Xcosh
2
Y 1 + cos
2
Xsinh
2
Y
( )
+ 2i cosXsinhY
sinXcoshY +1
( )
2
+ cosXsinhY
( )
2

_
,

(2)
In going from the third to the fourth line, we have multiplied the
numerator and denominator by the complex conjugate of the
denominator.
The imaginary part of the expression in Eq. 2 is then
106

O
1
r
arctan
2cosXsinhY
sin
2
Xcosh
2
Y 1 + cos
2
Xsinh
2
Y

_
,

1
r
arctan
2cosXsinhY
sinh
2
Y cos
2
X

_
,

1
r
arctan
2 cosX sinhY
( )
1 cosX sinhY
( )
2

_
,

1
r
arctan
2tanA
1 tan
2
A

_
,

(3)
The last expression serves to define A. But the quantity in
parentheses is just tan 2A (see, e.g., formula 406.02 of Dwight's Tables
of Inegrals and Other Mathematical Data, 4th edition), and the "angle
whose tangent is tan 2A" is just 2A (i.e., arctan(tan2A) = 2A.
However, A = arctan

cosX sinhY
( ) so that, finally

O
2
r
arctan
cosX
sinhY

_
,


2
r
arctan
cosrx b
sinhr y b

_
,

(4)
which is the result in Eq. 12.3-6.
107
Note to p. 388
Here we work through the missing steps to get the result in Eq.
12.4-16. We begin by evaluating the integral in the first term on the
right side of the equals sign in Eq. 12.-4-4 (the second term is zero,
because

v
e
v
~
a constant in this problem):

pv
x
v
~
v
x
( )dy pv
~
2
o x ( )
v
x
v
~
0
1
[
0
~
[
1
v
x
v
~

_
,

dq (1)
Here

o x
( ) is the velocity boundary-layer thickness, and

q y o x
( ) is
the dimensionless coordinate in the y-direction. In the second integral
we have changed the upper limit to "1" because

1 v
x
v
~
( ) in Eq.
12.4-6 and 7 is zero beyond

q 1. Then substituting the assumed
velocity profile into Eq. 1 gives:

pv
x
0
~
[
v
~
v
x
( )
dy

pv
~
2
o x
( )
2q 2q
3
+q
4
( )
1 2q + 2q
3
q
4
( )
dq
0
1
[

pv
~
2
o x
( )
2q 4q
2
2q
3
+ 9q
4
4q
5
4q
6
+ 4q
7
q
8
( )
dq
0
1
[

pv
~
2
o x ( ) 1
4
3

1
2
+
9
5

2
3

4
7
+
1
2

1
9
( )

pv
~
2
o x ( )
315+56718035
315
( )

37
315
pv
~
2
o x ( ) (2)
Similarly, the integral appearing in Eq. 12.4-5 may be
evaluated:

p

C
p
0
~
[
v
x
T
~
T
( )
dy

p

C
p
v
~
T
~
T
0
( )o
T
x ( )
v
x
v
~
T
~
T
T
~
T
0

_
,
0
1
[
dq
T

p

C
p
v
~
T
~
T
0
( )o
T
x ( )
v
x
v
~
T
~
T
0
T
~
T
0

T
0
T
T
0
T
~

_
,
0
1
[
dq
T
108

p

C
p
v
~
T
~
T
0
( )o
T
x ( )
v
x
v
~
1
T
0
T
T
0
T
~

_
,
0
1
[
dq
T
(3)
in which

q
T
y o
T
x
( )
y Ao x
( ) , and A is assumed to be inde-
pendent of x. Then, inserting the postulated profiles for velocity and
temperature into Eq. 3, we get:

p

C
p
0
~
[
v
x
T
~
T
( )
dy

p

C
p
v
~
T
~
T
0
( )
o
T
x
( )
2q
T
A 2q
T
3
A
3
+q
T
4
A
4
( )
0
1
[

1 2q
T
+ 2q
T
3
q
T
4
( )
dq
T

p

C
p
v
~
T
~
T
0
( )o
T
x ( )
2
15
A
3
140
A
3
+
1
180
A
4
( )
(4)
When the expressions in Eqs. 2 and 4 as well as Eqs. 12.4-6 to 9 are
substituted into Eqs. 12.4 and 5, we get differential equations for the
boundary-layer thicknesses as a function of the distance along the
plate:

2v
~
o

37
315
pv
~
2
d
dx
o (5)

2k T
~
T
0
( )
o
T

2
15
A
3
140
A
3
+
1
180
A
4
( )
p

C
p
v
~
T
~
T
0
( )
d
dx
o
T
(6)
Eq. 5 for

o x
( ) may be solved as follows:

o
d
dx
o
315
37
2v
~
pv
~
2
;

odo
0
o
[

630
37

pv
~
dx
0
x
[
;

o
1260
37
x
pv
~

_
,

(7,8,9)
and Eq. 6 for

o
T
x
( ) may also be solved:

2k T
~
T
0
( )
o
T

2
15
A
3
140
A
3
+
1
180
A
4
( )
p

C
p
v
~
T
~
T
0
( )
d
dx
o
T
(10)

o
T
d
dx
o
T

1
2
15
A
3
140
A
3
+
1
180
A
4
( )
2k
p

C
p
v
~
(11)
109

o
T
do
T

2
2
15
A
3
140
A
3
+
1
180
A
4
( )
o
v
~
dx
0
x
[
0
o
T
[
(12)

o
T

4
2
15
A
3
140
A
3
+
1
180
A
4
( )
ox
v
~

_
,

(13)
It remains to find

A o
T
x
( )
o x
( ) as a function of the physical
properties. Forming the ratio we get:

A
o
T
o

4 37 1260
( )
2
15
A
3
140
A
3
+
1
180
A
4
( )
op

_
,

(14)
Squaring both sides and collecting all the A terms on the left side, we
find:

2
15
A
3

3
140
A
5
+
1
180
A
6

35
315
1
Pr


Pr=
p

k
p

C
p

C
p
k

_
,

(15)
For large Prandtl numbers, we can drop all but the lead term on the
left side and get

A
35
315
15
2
3
Pr
1 3
0.880
3
Pr
1 3
0.958Pr
1 3
(16)
As pointed out in the textbook, 0.958 may be replaced by 1 to fit the
exact curve in Eq. 12b.2-15 within 5%. This replacement leads to Eq.
12.4-16 in the textbook.
[Note: In earlier printings of the textbook, the second integrals in Eqs.
12.4-10 and 11 had an upper limit of infinity rather than 1.]
110
Note to p. 413
The object here is to fill in the missing steps between Eq. 13.4-11
and Eq. 13.4-16. First we multiply Eq. 13.4-11 by

d
d
1 +
o
t ( )
o

_
,

d+
d

_
,

C
0
o (1)
Then first integration with respect to gives:

1 +
o
t ( )
o

_
,

d+
d
C
0
o
0

[
d + C
1
C
0
I ( ) + C
1
(2)
where o
( ) is the dimensionless velocity defined just after Eq. 13.4-6,
and the abbreviation

I
( ) is introduced. Eq. 2 may now be rewritten:

d+
d

C
0
I ( )
1 + o
t ( )
o
( )

1
]
+
C
1
1 + o
t ( )
o
( )

1
]
(3)
A second integration with respect to gives:

+ C
0
I
( )
1 + o
t ( )
o
( )

1
]
0

[
d + C
1
1
1 + o
t ( )
o
( )

1
]
d + C
2
0

[
(4)
in which we must remember that

o
t ( )
is now a function of . If next
we consider the limit of the above expression as

0, it may be seen
that the first term goes to zero and the second term goes to infinity
(and therefore violates B. C. 1); therefore

C
1
must be taken to be zero.
Hence the expression for the dimensionless temperature becomes:
111

O ,
( )
C
0
+ C
0
I
( )
1 + o
t ( )
o
( )

1
]
d
0

[
+ C
2
(5)
Next, apply the boundary condition at

1:

1 C
0
I
( )
1 + o
t ( )
o
( )

1
C
0
I 1
( ) or

C
0
I 1
( )

1
]
1
(6)
inasmuch as

o
t ( )
is zero at the wall. Next we want to get the driving
force

O
0
O
b
, which is the dimensionless wall temperature minus
the dimensionless bulk temperature (defined in Eq. 10.8-33):

O
0
O
b
C
0
I
( )
1 + o
t ( )
o
( )

1
]
d
0
1
[
C
0
o
I
( )
1 + o
t ( )
o
( )

1
]
dd
0

[
0
1
[
od
0
1
[


C
0
I
( )
1 + o
t ( )
o
( )

1
]
d
0
1
[

C
0
I 1
( )
od

1
[ ( )
I
( )
1 + o
t ( )
o
( )

1
]
d
0
1
[
(7)
In the second expression we have interchanged the order of integra-
tion. The second term in Eq. 7 may be rewritten:

C
0
I 1
( )
od od
0

[
0
1
[
( )
I
( )
1 + o
t ( )
o
( )

1
]
d
0
1
[


C
0
I 1
( )
I 1
( )
I
( )

1
] 0
1
[
I
( )
1 + o
t ( )
o
( )

1
]
d
112

C
0
I
( )
1 + o
t ( )
o
( )

1
]
0
1
[
d +
C
0
I 1
( )
I
( )

1
]
2
1 + o
t ( )
o
( )

1
]
0
1
[
d (8)
The first term in Eq. 8 above just cancels the first term in Eq. 7, and
hence we are left with:

O
0
O
b

C
0
I 1
( )
I
( )

1
]
2
1 + o
t ( )
o
( )

1
]
0
1
[
d
I
( )
I 1
( )

1
]
2
1 + o
t ( )
o
( )

1
]
0
1
[
d (9)
This is in agreement with Eq. 13.4-16 in the textbook.
113
Note to p. 415
Some additional material is given here on 13.5.
Perform the indicated substitutions into Eq. 13.5-1 to get for the
partial differential equation for O:

v
t ( )
z
F

_
,

O
r
+
v
t ( )
z
F

_
,

O
z

v
t ( )
Pr
t ( )
1
r

r
r
O
r

_
,

(1)
where the primes indicate differentiations with respect to . We next
make the change of variables

r z and

pv
t ( )
w
( )
z, and further
let

O r, z
( )

O ,
( )
f
( )
, so that

oO
or

O
o
o
or
+
o

O
o
o
or

O
o

1
z
+
o

O
o
0
f

1
z
(2)

O
z

z
+

o
oz

O
o

r
z
2

_
,

+
o

O
o
pv
t ( )
w

_
,

r
z
2

_
,

_
,

(3)
Substitution of these expressions into Eq. 1 gives

v
t ( )
z
F

_
,

1
z

1
]
1

v
t ( )
z
F

_
,


r
z
2

_
,

_
,

1
]
1

v
t ( )
Pr
t ( )
1

o
o

of
o

_
,


1
z
2
(4)
Multiplication of the entire equation by

z
2
v
t ( )
gives

F

_
,

f
[ ]
+ F
( )
f +
f

1
]
1

1
Pr
t ( )
1

o
o

of
o

_
,

(5)
114
On the left side, the terms involving

F f cancel, and the equation
may be rewritten as follows:

1

d
d
Ff
( )

1
Pr
t ( )
1

o
o

of
o

_
,

(6)
This equation can be multiplied by

Pr
t ( )
and integrated once to give

Pr
t ( )
Ff
of
o
+ C (7)
According to Eq. 5.6-20, F = 0 at

0 , which means that C = 0. A
further integration from 0 to then yields

ln
f
( )
f 0
( )
Pr
t ( )
F

[
d Pr
t ( )
C
3
2

1 +
1
4
C
3

( )
2
d
0

[

Pr
t ( )
ln 1 +
1
4
C
3

( )
2

1
]
1
2
(8)
or, taking the antilogarithm of both sides

f
( )
f 0
( )
1 +
1
4
C
3

( )
2

1
]
1
2Pr
t ( )
(9)
When this result is compared with Eq. 5.6-21 for the velocity profile
in a circular jet, and use is made of the definition in Eq. 13.5-8, we
obtain finally

O
O
max

v
z
v
z,max

_
,

Pr
t ( )
(10)
which is a rather simple, and apparently fairly satisfactory, result.
115
Note to p. 454
Verify that Eq. 15.1-1 can be obtained from Eq. 11.1-9 by the method
described on p. 454.
We start by integrating Eq. 11.1-9 over the volume of the flow
system shown in Fig. 7.0-1:

o
ot
1
2
pv
2
+ p

U + p

d
( )

_
,

dV V
1
2
pv
2
+ p

U + p

d
( )
v
( )
dV
V t ( )
[
V t ( )
[

V q
( )
V t ( )
[
dV V pv
( )
V t ( )
[
dV V t v
[ ] ( )
V t ( )
[
dV (1)
We next apply the Leibnitz formula to the left side of the equation
and the Gauss divergence theorem to the right side:

d
dt
1
2
pv
2
+ p

U + p

d
( )
dV n
1
2
pv
2
+ p

U + p

d
( )
v
S ( )
S t ( )
[
V t ( )
[
dS

n
1
2
pv
2
+ p

U + p

d
( )
v
( )
S t ( )
[
dS

n q
( )
S t ( )
[
dS

n pv
( )
S t ( )
[
dS n t v
[ ] ( )
S t ( )
[
dS (2)
The integral in the first term on the left side is the total energy
(kinetic + internal + potential energy). The second term on the left
side can be combined with the first term on the right side. Thus we
get

d
dt
K
tot
+ U
tot
+ d
tot
( )
n
1
2
pv
2
+ p

U + p

d
( )
v v
S
( )
( )
S t ( )
[
dS

n q
( )
S t ( )
[
dS n pv
( )
S t ( )
[
dS n t v
[ ] ( )
S t ( )
[
dS (3)
We now analyze the terms on the right side seriatim:
The first term can be seen to contribute nothing on the fixed
surface

S
f
and the moving surface

S
m
. At the inlet cross section

S
1
and the outlet section

S
2
, the surface velocity

v
S
is zero and collinear
with the outwardly directed unit vector n. The fluid velocity vector v
is assumed to point in the direction opposite to the n vector at the
entry plane, and in the same direction as as the n vector at the exit
116
plane; therefore, at the entry

n v
( )
v , and at the exit

n v
( )
+v .
We make the further assumption that the internal energy and the
potential energy are constant over the cross section. Then when the
integration over the cross sectional area is performed we get:

n
1
2
pv
2
+ p

U + p

d
( )
v v
S
( )
( )
S t ( )
[
dS

1
2
p
1
v
1
3
S
1
+ p
1

U
1
v
1
S
1
+ p
1

d
1
v
1
S
1

1
2
p
2
v
2
3
S
2
+ p
2

U
2
v
2
S
2
+ p
2

d
2
v
2
S
2
(4)
The second term on the right side (the q-term) is the integral
over all surfaces of the normal component of the heat flux vector and
is thus the rate of total heat addition to the system, Q:

n q
( )
S
f
+S
m
+S
1
+S
2
[
dS Q (5)
It is assumed that the heat addition at surfaces

S
1
and

S
2
is usually be
small compared to the heat added at the solid surfaces.
The third term on the right (the p-term) has to be evaluated at
all the surfaces. At the inlet and outlet planes, we will get

n pv
( )
S
1
+S
2
[
dS p
1
v
1
S
1
p
2
v
2
S
2
(6)
by the same arguments leading to the internal energy terms in Eq. 4.
These terms represent the rate of doing work on the system at the
entry and exit planes. On the solid surfaces we get

n pv
( )
S
f
+S
m
[
dS W
m
p
(7)
This term is the rate that pressure does work on the system at the
moving surfaces

S
m
; there is no work done at the fixed surfaces

S
f
,
inasmuch as the rate of doing work is a force times a velocity, and at

S
f
the surface velocity is zero.
The fourth term on the right (the t -term) is evaluated similarly
to the p-term. First the contributions at surfaces

S
1
and

S
2
are
considered, but it is assumed that these will be small compared to the
117
pressure terms in Eq. 6. On the solid surfaces there will be a
contribution similar to that for the pressure forces on the moving
surfaces:

n t v
[ ] ( )
S
f
+S
m
[
dS W
m
t
(8)
and, here again, the contribution at

S
f
will be zero. The contributions
from Eqs. 7 and 8 will be added to give

W
m
p
+ W
m
t
W
m
, the total
work done on the system through the moving surfaces.
When all the contributions in Eqs. 4 through 8 are added up we
get Eq. 15.1-1 of the textbook:

d
dt
K
tot
+ U
tot
+ d
tot
( )

1
2
p
1
v
1
3
+ p
1

U
1
v
1
+ p
1

d
1
v
1 ( )
S
1

1
2
p
2
v
2
3
+ p
2

U
2
v
2
+ p
2

d
2
v
2 ( )
S
2
+ Q + W
m

+ p
1
v
1
S
1
p
2
v
2
S
2
( )
(8)
or, introducing the enthalpy

d
dt
K
tot
+ U
tot
+ d
tot
( )

1
2
p
1
v
1
3
+ p
1

H
1
v
1
+ p
1

d
1
v
1 ( )
S
1

1
2
p
2
v
2
3
+ p
2

H
2
v
2
+ p
2

d
2
v
2 ( )
S
2
+ Q + W
m
(9)
Either Eq. 8 or Eq. 9 is referred to as the unsteady-state macroscopic
energy balance.
118
Note to p. 494
The derivations of the Stefan-Boltzmann law and Wien's law
from the Planck black-body distribution law are quite important and
therefore it is a good idea to understand all the intermediate steps in
the development.
a. The Stefan-Boltzmann law
We integrate Planck's distribution law over all wavelengths:

q
b
e ( )
q
bi
e ( )
0
~
[
di
2rc
2
h
i
5
1
e
ch iKT
1
0
~
[
di (1)
Next we make a change of variable

x ch iKT . Then

2rc
2
h
1
i
5
di 2rc
2
h
KT
ch

_
,

5
x
5

ch
KT

_
,


1
x
2

_
,

dx

2r KT
( )
4
c
2
h
3
x
3
( )
dx (2)
Hence the integral becomes

q
b
e ( )


2r KT ( )
4
c
2
h
3
x
3
e
x
1
0
~
[
dx (3)
Then expand the denominator of the integrand as a Taylor series
about x = 0, to get:

e
x
1 1 + e
x
+ e
2x
+ e
3x
+
( )
1 (4)
Then a term by term integration of Eq. 3 gives:

2r KT ( )
4
c
2
h
3
x
3
e
nx
0
~
[
dx
n1
~
_

2r KT ( )
4
c
2
h
3
6
1
n
4
n1
~
_

2r KT ( )
4
c
2
h
3
r
4
15

_
,

(5)
119
On p. 171 of Planck's book, The Theory of Radiation, Dover, New York
(1959), which is a translation of Vorlesungen ber die Theorie der
Wrmestrahlung, 5th edition, Barth, Leipzig (1923), the Stefan-
Boltzmann equation is obtained as shown above. However, Planck
did not evaluate the summation in Eq. 5 exactly. Instead, he simply
evaluated the sum numerically as:

o 1 +
1
2
4
+
1
3
4
+
1
4
4
+ 1.0823 (5)
Nowadays, even in a small integral table (such as H. B. Dwight,
Tables of Integrals and Other Mathematical Data, Macmillan, New York,
Fourth Edition (1961)), the integral over x in Eq. 3 may be found; see
Formula 860.33 on p. 231.
When Eq. 5 is compared with the Stefan-Boltzmann law for a
black body

q
e ( )
oT
4
, then we get the Stefan-Boltzmann constant:

o
2
15
r
5
K
4
c
2
h
3
(6)
which interrelates key constants from several different fields of
physics.
b. Wien's displacement law
First rewrite Eq. 16.3-7 in terms of x thus:

q
b
e ( )

2rc
2
h
i
5
1
e
ch iKT
1
2rc
2
h
KT
ch

_
,

5
x
5
e
x
1
(7)
We can then differentiate this with respect to x to get

dq
b
e ( )
dx
2rc
2
h
KT
ch

_
,

5
5x
4
e
x
1

x
5
e
x
e
x
1
( )
2

_
,

(8)
Then setting the derivative equal to zero gives the value of x at which
the maximum in the

q
b
e ( )
curve occurs:
120

5
x
max
e
x
max
e
x
max
1
0 (9)
whence

x
max
5 1 e
x
max
( )
(10)
from which one can find, by trial and error,

x
max
4.9651... , or

i
max
T 0.2884cm K. This is Wien's displacement law.
121
Note to p. 529
We want to verify that Eq. 17.4-4, Eq. (I) of Table 17.8-1, and Eq.
(E) of Table 17.8-2 are dimensionally consistent.
If we put the dimensions of the quantities into the equation
instead of the mathematical symbols (and omit the numerical factors)
we get:
(a)

L
2
t

_
,

M
Lt

_
,

ML
2
t
2
T

_
,

T ( )

1
L
(1)
(b)

moles
L
2
t

_
,


moles
L
3

_
,

L
t

_
,

(2)
(c)

M
L
3

_
,

L
t

_
,


M
L
3

_
,

L
2
t

_
,

1
L

_
,

(3)
In each case, the dimensions on the left side are the same as the
dimensions on the right side.
122
Note to p. 534
It is important to know how to simplify multicomponent
relations to their corresponding binary equations. We illustrate this
procedure by showing how to get the binary formula in Eq. (Q') from
the multicomponent formula in Eq. (Q) in Table 17.7-1.
In the sum in Eq. (Q), the indices o and y can take on only the
values of A and B in a binary system. Then in the sum, y must be B.
Therefore, Eq. (Q') becomes for a binary system:

Vu
A

M
A
M
2
M + x
A
M
B
M
A
( )

1
]
Vx
B
(1)
Next use Eqs. (M) and (J) of the table, written for a binary system:

Vu
A

M
A
x
A
M
A
+ x
B
M
B
( )
2
x
A
M
A
+ x
B
M
B
( )
+ x
A
M
B
M
A
( )

1
]
Vx
A
( )
(2)
Within the bracket, the terms

x
A
M
A
cancel each other, and the
remaining terms may be combined, since

x
A
+ x
B
1 . We are then left
with:

Vu
A
+
M
A
M
B
x
A
M
A
+ x
B
M
B
( )
2
Vx
A
(3)
which is just Eq. (Q').
123
Note to p. 535
Here we want to verify that Eq. 17.7-4 can be derived from Eq.
17.7-3 using the relations in Tables 17.7-1 and 2.
First we transform the last term in Eq. 17.7-3 into the analogous
term in Eq. 17.7-4, with a multiplying factor:

pD
AB
Vu
A
cM ( )D
AB
M
A
M
B
M
2
( )
Vx
A

cD
AB
Vx
A
M
A
M
B
M
( ) (1)
Here, Eq. (F) of Table 17.7-1 was used, as well as Eq. (

Q ).
Next we transform the mass concentration times the diffusion
velocity as follows:

p
A
v
A
v ( ) c
A
M
A
( ) v
A
u
A
v
A
+u
B
v
B
( )

1
]

c
A
M
A
( )
u
B
v
A
v
B
( )
c
A
M
A
( )
x
B
M
B
M
( )
v
A
v
B
( )

c
A
M
A
( )
M
B
M
( )
v
A
x
A
v
A
+ x
B
v
B
( ) ( )

c
A
v
A
v*
( )
M
A
M
B
M
( )
(2)
In this development, we end up with the same factor

M
A
M
B
M
( )
appearing. In the first step above, we used Eq. (B) of Table 17.7-2. In
the second step we used Eq. (K) of Table 17.7-1. In the third step Eq.
(O) of Table 17.7-1 was used. In the fourth step we used Eq. (J) of
Table 17.7-1, and in the last step Eq. (C) of Table 17.7-2.
The rest of the proof makes use of the definitions

j
A
p
A
v
A
v
( )

J
A
-
c
A
v
A
v*
( ) (3a,b)
which are given in Eqs. (E) and (I) of Table 17.8-1. It remains, then, to
verify that

j
A
J
A
-
M
A
M
B
M
( ) (4)
This may be done by rewriting Eq. 4 as
124

j
A
J
A
-
M
( )
M
A
M
( )
M
B
M
( )
J
A
-
p c
( )
u
A
x
A
( )
u
B
x
B
( ) (5)
Here Eqs. (G) and (O) of Table 17.7-1 have been used. The result in
Eq. 5 may also be written thus:

j
A
pu
A
u
B

J
A
-
cx
A
x
B
(6)
This important equation is also given in Eq. 17B.3-1.
125
Note to p. 547 (i)
Starting with the concentration profile for species A in Eq. 18.2-
11, derive the subsequent results up through Eq. 18.2-16.
First obtain the expression for

x
B,avg
:

x
B,avg
x
B1

x
B2
x
B1
( )

d
0
1
[
d
0
1
[

x
B2
x
B1
( )

ln x
B2
x
B1
( )
0
1
(1)
where we have used the integral

a
x
[
dx a
x
ln a
( )
+ C. Therefore

x
B,avg
x
B1

x
B2
x
B1
( ) 1
ln x
B2
x
B1
( )
or

x
B,avg

x
B2
x
B1
ln x
B2
x
B1
( )
x
B
( )
ln
(2a,b)
Hence the rate of evaporation of A at the gas-liquid interface is:

N
A
zz
1

cD
AB
x
B1
dx
B
dz
zz
1

cD
AB
x
B1
dx
B
d
0
d
dz

cD
AB
z
2
z
1
d x
B
x
B1
( )
d
0
(3)
Then, using the derivative

da
x
dx a
x
ln a , we get:

N
A
zz
1

cD
AB
z
2
z
1
x
B2
x
B1

_
,

ln
x
B2
x
B1

_
,

1
]
1
1
0

cD
AB
z
2
z
1
ln
x
B2
x
B1

_
,

(4)
Then, multiplying the numerator and denominator by

x
B2
x
B1
( ) :

N
A
zz
1

cD
AB
z
2
z
1
x
B2
x
B1
( )
x
B
( )
ln

cD
AB
z
2
z
1
x
A1
x
A2
( )
1 x
A
( )
ln
(5)
Next we obtain the solution for very small values of

x
A1
and

x
A2
:
126

N
A
zz
1

cD
AB
z
2
z
1
x
A1
x
A2
( )
ln 1 x
A2
( )
1 x
A1
( )

1
]
1 x
A2
( ) 1 x
A1
( )

cD
AB
z
2
z
1
ln 1 x
A2
( )
ln 1 x
A1
( )

1
]

cD
AB
z
2
z
1
x
A2

1
2
x
A2
2

1
3
x
A2
3

( )
+ x
A1
+
1
2
x
A1
2
+
1
3
x
A1
3

( )

1
]

cD
AB
x
A1
x
A2
( )
z
2
z
1
1 +
1
2
x
A1
2
x
A2
2
( )
x
A1
x
A2
( )
+
1
3
x
A1
3
x
A2
3
( )
x
A1
x
A2
( )
+

1
]
1
1

cD
AB
x
A1
x
A2
( )
z
2
z
1
1 +
1
2
x
A1
+ x
A2
( ) +
1
3
x
A1
2
+ x
A1
x
A2
+ x
A2
2
( )
+

1
]
(6)
The Taylor expansion in Eq. C.2-3 has been used for expanding the
logarithms in line 2.
127
Note to p. 547 (ii)
a. The result in Eq. 18.2-11 may be written as follows:

x
B
x
B1

x
B2
x
B1

_
,

in which


z z
1
z
2
z
1
(1,2)
We want to expand this result in a Taylor series in to get a result of
the form

x
B
x
B1
1
( )
+ (3)
b. Next we rework the problem in 18.2 by omitting the

x
A
term in the denominator of Eq. 18.2-1 (that is, assume that

x
A
is so
small that it can be neglected with respect to unity). We can then
show that this gives the first two terms of the series in Eq. 3.
a. From Eq. C2.1 we get by expanding about =0:

x
B
x
B1

x
B2
x
B1

_
,

0
+
d
d
x
B2
x
B1

_
,

0
0
( )
+

1 + 1 ln
x
B2
x
B1

_
,

1
]
1
1
+ (4)
Now from Eq. C.2-3, with (1 + x) replaced by x, and x replaced by
(1 x), we have, for

0 < x s 2

lnx x 1 ( )
1
2
x 1 ( )
2
+
1
3
x 1 ( )
3

1
4
x 1 ( )
4
+ (5)
Therefore,
128

x
B
x
B1
1 +
x
B2
x
B1
1

_
,


1
2
x
B2
x
B1
1

_
,

2
+
1
3
x
B2
x
B1
1

_
,

3
+

1
]
1
1
+ (6)
If

x
B2
is only slightly greater than

x
B1
(which would be the case if
species A is present only in a small amount), then we need retain
only the first term inside the bracket. The final result is then:

x
B
x
B1
1 +
x
B2
x
B1
1

_
,

(7)
b. When

x
A
can be neglected with respect to unity in Eq. 18.2-1,
the differential equation for

x
A
as a function of z is:

d
2
x
A
dz
2
0 (8)
Since

x
A
+ x
B
1 , the same differential equation (with A replaced by
B) is valid for species B. Furthermore, since z and are related by the
linear expression given in Eq. (2) we may write:

d
2
x
B
d
2
0 (9)
This equation may be integrated to give:

x
B
x
B1
1 +
x
B2
x
B1
1

_
,

(10)
in agreement with Eq. 7.
129
Note to p. 555
We wish to verify that Eq. 18.4-9 does satisfy Eq. 18.4-7 by
substituting into the differential equation.
Differentiate Eq. 18.4-9 with respect to , using Eq. C.5-10:

d
d

sinh o 1 ( )

1
]
o ( )
cosho
(1)
A second differentiation then gives, using Eq. C.5-11:

d
2

d
2

cosh o 1
( )

1
]
+o
2
( )
cosho
(2)
By substituting the second derivative from Eq. 2 and the expression
in Eq. 18.4-9 into Eq. 18.4-7, it is seen that an equality is obtained.
130
Note to p. 557
Here we show that Eq. 18.4-18 can be rearranged into a simpler form.
First we put the expression in the large parentheses over a common
denominator

N
o
sinho
cosh
2
o + V So ( )o cosho sinho 1
cosho + V So
( )
o sinho

_
,

(1)
Next make use of Eq. C.5-5 to get

N
o
sinho
sinh
2
o + V So ( )o cosho sinho
cosho + V So
( )
o sinho

_
,

(2)
Finally cancel the

sinho appearing in the numerator and
denominator to get

N o
sinho + V So
( )
o cosho
cosho + V So
( )
o sinho

_
,

(3)
As

V So
( )
0,

N o tanho , and as

V So
( )
~,

N o cotho .
131
Note to p. 563 (i)
Here we verify that Eq. 18.6-8 is a solution to the differential
equation given in Eq. 18.6-2.
Because of the definition of f just above Eq. 18.6-6,

f c
A
c
A0
must satisfy Eq. 18.6-2. Therefore, we calculate the derivatives using
the Leibniz formula given in C.3:

of
oz

1

4
3
( )
exp q
3
( )
oq
oz


1

4
3
( )
exp q
3
( )

1
3
( )
yz
4 3
a 9D
AB
( )
1 3
(1)

of
oy

1

4
3
( )
exp q
3
( )
oq
oz
of
oy

1

4
3
( )
exp q
3
( )
oq
oy


1

4
3
( )
exp q
3
( )
a 9D
AB
z ( )
1 3
(2)

o
2
f
oy
2

1

4
3
( )
exp q
3
( )
a 9D
AB
z
( )
1 3
3q
2
( )
a 9D
AB
z
( )
1 3
(3)
Then, substituting these expressions into Eq. 18.6-2, we get:

ay +
1
3
( )
yz
4 3
a 9D
AB
( )
1 3
D
AB
+3q
2
( )
a 9D
AB
z
( )
2 3
(4)
If both sides are multiplied by

9D
AB
z a ( )
1 3
, and replace

q
2
by

y
2
a 9D
AB
z ( )
2 3
, we get

ay
2
3z
3D
AB
y
2
a 9D
AB
z
( ) (5)
which is an identity. This concludes the proof.
132
Note to p. 563 (ii)
The concentration profiles are given by Eq. 18.6-8.

c
A

c
A0

4
3
( )
e
q
3
q
~
[
dq (1)
where

q y a 9D
AB
z ( )
1 3
. We show how to get the next two equations
from this result.
Notice in Eq. 1 the bar over q, which is used to make a
distinction between the dummy variable of integration and the
dimensionless distance. This distinction is vital when we apply the
Leibniz rule (Eq. C.3-2) for differentiation of an integral to get the
local molar flux at the wall is then:

N
Ay
y0
D
AB
oc
A
oy
y0

D
AB
c
A0

4
3
( )
d
dq
e
q
3
q
~
[

dq
dy

1
]
1
y0
(2)
Note that only the third term in Eq. C.3-2 contributes to the
derivative (i.e., the term involving the lower limit of the integral):

N
Ay
y0

D
AB
c
A0

4
3
( )
e
q
3
( )

a
9D
AB
z

_
,

1 3

1
]
1
1
q0

D
AB
c
A0

4
3
( )
a
9D
AB
z

_
,

1 3
(3)
Then the total molar flow across the surface of width W and
length L is given by the integral over that surface:

W
A
N
Ay
y0
dzdx
0
L
[
0
W
[
D
AB
c
A0

4
3
( )
a
9D
AB

_
,

1 3
W z
1 3
dz
0
L
[

D
AB
c
A0

4
3
( )
a
9D
AB

_
,

1 3
W
z
2 3
0
L
2
3

2D
AB
c
A0
WL
4
3

4
3
( )
a
9D
AB
L

_
,

1 3
(4)
133
In the last step, Eq. C.4-4 can be used to replace

4
3

4
3
( )
by

7
3
( )
to
agree with Eq. 18.6-10.
134
Note to p. 565
Let us verify that Eq. 18.7-9 satisfies the differential equation in
Eq. 18.7-6. To simplify the problem, it is a good idea to introduce
dimensionless variables:


r
R
;


c
A
c
AR
;

A
k
1
aR
2
D
A
(1,2,3)
Then the differential equation in Eq. 18.7-6 and the solution in Eq.
18.7-9 may be rewritten as:

1

2
d
d

2
d
d

_
,

A
2
I ;

I
1

sinhA
sinhA
(4,5)
First we use Eq. 5 to evaluate the derivative (see also C.5):

d
d

1

2
sinhA
sinhA
+
A

coshA
sinhA
(6)
Multiplication by

2
then gives:

2
dI
d

sinhA
sinhA
+ A
coshA
sinhA
(7)
Further differentiation gives:

d
d

2
dI
d

_
,

A
coshA
sinhA
+ A
coshA
sinhA
+ A
2

sinhA
sinhA
(8)
Division by

2
gives:

1

2
d
d

2
dI
d

_
,

A
2
1

sinhA
sinhA
(9)
135
But the right side of the equation is just I , according to Eq. 5. We
have therefore shown that the differential equation is satisfied.
136
Note to p. 584
Here we show how to get Eq. 19.1-15 from the preceding equations.
Start with Eq. 19.1-10, and let

c
o
cx
o
, so that we get

c
ox
o
ot
+ x
o
oc
ot
V N
o
( )
+ R
o
(1)
Next we use Eq. 19.1-12 to obtain

c
ox
o
ot
+ x
o
V cv *
( )
+ R

1
N
_

1
]
1
V N
o
( )
+ R
o
(2)
Then we go to Eq. 17.8-2 to find

N
o
J
o
-
+ c
o
v *, so that (2) becomes


c
ox
o
ot
x
o
V cv *
( )
+ V c
o
v *
( )
V J
o
-
( )
+ R
o
x
o
R

1
N
_
(3)
------------------------------
Next we have to show that the two dashed-underlined terms on the
left side of (3) are the same as the

c v * Vx
o
( )
in Eq. 19.1-15:

x
o
V cv *
( )
+ V c
o
v *
( )
x
o
V cv *
( )
+ V x
o
cv *
( )

x
o
V cv *
( )
+ x
o
V cv *
( )
+ cv * Vx
o
( )

c v * Vx
o
( )
(4)
Here we have made use of Eq. A.4-19 to differentiate the product of a
scalar with a vector.
The last two terms in Eq. 19.1-17 for a binary system are obtained thus:

R
A
x
A
R
A
+ R
B
( )
1 x
A
( )
R
A
x
A
R
B
x
B
R
A
x
A
R
B
(5)
because for a binary system N in the upper limit of the sum is 2, there
being just two components, A and B.
137
Note to p. 585
The equation to be solved is in Example 19.1-1 is

v
0
dc
A
dz
D
AB
d
2
c
A
dz
2
k
1
c
A
(1)
which is of the form of Eq. C.1-7a. We know that the concentration of
A will decrease with increasing distance from the porous plug, so we
assume that it will have the form

c
A
exp az
( ) , where a is a positive
constant. When this is substituted into the Eq. 1, we get (after
canceling

exp az
( ) from each term):

v
0
a D
AB
a
2
k
1

or

D
AB
a
2
+ v
0
a k
1

0 (2)
This quadratic equation can be solved for a to get:

a
v
0
v
0
2
+ 4D
AB
k
1

2D
AB
(3)
To insure that a is positive, we must choose the plus sign. Hence a is
given by:

a 1 + 1 + 4D
AB
k
1

v
0
2
( )

1
]
1
v
0
2D
AB

_
,

(4)
and the concentration profile is:

c
A
exp +1 1 + 4D
AB
k
1

v
0
2
( )

1
]
1
v
0
z
2D
AB

_
,

(5)
which is the result in Eq. 19.1-20.
138
Note to p. 589
It is desired to show how to obtain Eq. (H) of Table 19.2-4 from Eq.
(E).
We move the

V q
( ) term to the left side of the equation and
then perform mathematical operations on the new left side. First we
use Eq. 3.5-4 to rewrite the new left side as:

p
D

H
Dt
+ V q
( )

o
ot
p

H + V pv

H
( )
+ V q
( ) (1)
For a multicomponent mixture, the heat flux vector q may be written
as described in Fn. (a) of Table 19.2-4, and then neglect the
contribution

q
x ( )
. Then Eq. 1 may be rewritten as

p
D

H
Dt
+ V q
( )

o
ot
c
o
H
o
o1
N
_
+ V pv

H
( )
V kVT
( )
+ V
H
o
M
o
j
o
o1
N
_

_
,

(2)
Here we have also used the relation Eq. 19.3-9 to rewrite the first
term on the right side. Next we combine the second and fourth terms
on the right side to get

p
D

H
Dt
+ V q ( )
o
ot
c
o
H
o
o1
N
_
+ V vc

H + J
o
H
o
o1
N
_

1
]
1

_
,

V kVT ( )

o
ot
c
o
H
o
o1
N
_
+ V v c
o
H
o
o1
N
_
+ J
o
H
o
o1
N
_

1
]
1

_
,

V kVT ( )
(3)
Then, since

c
o
v + J
o
N
o
from Table 17.8-1, Eqs. (G) and (H),

p
D

H
Dt
+ V q
( )

o
ot
c
o
H
o
o1
N
_
+ V N
o
H
o
o1
N
_

_
,

V kVT
( )
(4)
Combining this with Eq. (E) then gives Eq. (H).
139
Note to p. 590
Simplification of the expression for the combined energy flux e,
given in the first line of Eq. 19.3-4. First we insert Eq. 19.3-3 for q to
get the second line of Eq. 19.3-4:

e p

U +
1
2
v
2
( )
v kVT +
H
o
M
o
j
o
+ pv + t v
[ ]
o1
N
_
(1)
(a) (b) (c) (d) (e) (f)
We next move terms (c) and (d) to the left, and terms (b) and (f) to the
right:

e kVT +
H
o
M
o
j
o
+ p

U + p

V
( )
o1
N
_
v +
1
2
pv
2
v + t v
[ ]
(2)
(c) (d) (a) (e) (b) (f)
where used has been made of the expression

V 1 p to get term (e).


We next combine terms (a) and (e) to introduce the enthalpy:

e kVT +
H
o
M
o
j
o
+ p

H
o1
N
_
v +
1
2
pv
2
v + t v
[ ]
(3)
(c) (d) (a+e) (b) (f)
Next we make use of the relation

p

H c

H. Then we can use the fact
that enthalpy is "homogeneous of degree 1" (see Example 19.3-1 on p.
591 and in particular Eq. 19.3-9),

H Z
o
n
o
H
o
. Therefore, the
enthalpy per mole is

H H n Z
o
x
o
H
o
. Hence term (a+e) can be
rewritten so that (3) becomes

e kVT +
H
o
M
o
j
o
+ c
o
H
o
o1
N
_
o1
N
_
v +
1
2
pv
2
v + t v
[ ]
(4)
(c) (d) (a+e) (b) (f)
Then we use Eq. H of Table17.7-1, Eq. S of Table 17.8-1, and Eq. P of
Table 17.8-1 to rewrite parts of the summands in term (d+a+e) thus:
140

j
o
M
o
+ c
o
v
j
o
+ p
o
v
M
o

n
o
M
o
N
o
(5)
When this is substituted into (4) we get

e kVT + H
o
N
o
o1
N
_
+
1
2
pv
2
v + t v
[ ]
(6)
(c) (d+a+e) (b) (f)
Then, for systems in which terms (b) and (f) may be neglected, we
finally arrive at Eq. 19.3-6.
141
Note to p. 591
The theorem of Euler (pronounced "Oiler") for homogeneous
functions is used in order to get Eq. 19.3-9. We here give a proof of
Euler's theorem.
First we give a definition of a homogeneous function. A function
of n variables is said to be "homogeneous of degree k" if

f ix
1
, ix
2
, ix
3
,ix
n
( )
i
k
f x
1
, x
2
, x
3
,x
n
( )
(1)
That is, if in the function

f x
1
, x
2
, x
3
,x
n
( )
, we replace

x
1
by

ix
1
,

x
2
by

ix
2
, etc., then this will give a result that is the same as multiplying
the original function by

i
k
.
If we differentiate both sides of Eq. 1 with respect to i , we get

of
o ix
1
( )
o ix
1
( )
oi
+
of
o ix
2
( )
o ix
2
( )
oi
+
of
o ix
n
( )
o ix
n
( )
oi
ki
k1
f x
1
, x
2
, x
3
,x
n
( )
(2)
where, on the left side, it is understood that by f we mean the
function

f ix
1
, ix
2
, ix
3
,ix
n
( )
. It is understood that the function f
has continuous first partial derivatives. We now perform the
differentiations on the left side to get

of
o ix
1
( )
x
1
+
of
o ix
2
( )
x
2
+
of
o ix
n
( )
x
n
ki
k1
f x
1
, x
2
, x
3
,x
n
( )
(3)
Next, we set i = 1, to get

x
1
of
ox
1
+ x
2
of
ox
2
+x
n
of
ox
n
kf x
1
, x
2
, x
3
,x
n
( )
(4)
which is Euler's theorem. Here, the functionality of f is exactly the
same on both sides of the equation. Since the enthalpy is a
homogeneous function of degree "1" Eq. 19.3-9 follows directly.
142

n
o
o
_
oH
on
o

_
,

=o ( ),T, p
H (5)
or

n
o
o
_
H
o
H (6)
This result is frequently used in discussions of mixtures. An example
is the relation immediately after Eq. 17C.1-3, where, for a binary
mixture

n
A
V
A
+ n
B
V
B
V
or, when the entire equation is divided by V,

c
A
V
A
+ c
B
V
B
1
Another example of Euler's equation is in going from Eq. 1 to Eq. 2 in
the Note to p. 589.
143
Note to p. 615
We want to verify that Eqs. 20.1-16 and 17 are a solution to 20.1-
9 and 10. We make use of C.6 on the error function.
First we find the first and second derivatives of X with respect
to Z:

X
1 erf Z
( )
1 + erf
(1)

dX
dZ

1
1 + erf
2
r
e
Z ( )
2
(2)

d
2
X
dZ
2

1
1 + erf
2
r
e
Z ( )
2
2 Z
( )

1
]
(3)
When the derivatives in Eqs. 1, 2, and 3 are substituted into Eq. 20.1-
9, it is seen that the latter equation is satisfied.
Next we substitute the first derivative from Eq. 2 into Eq. 20.1-
10 to get:


1
2
x
A0
1 x
A0

_
,

1
1 + erf
2
r
e
Z ( )
2

_
,

Z0

x
A0
1 x
A0

_
,

2
1 + erf ( ) r
(4)
which agrees with Eq. 20.1-17.
144
Note to p. 622
Here we give a more detailed discussion of the development
between Eqs. 20.1-67 and Eq. 20.1-74.
When the bracket in Eq. 20.1-67 is set equal to unity, we get:

d
2
g
d
2
+ 2
dg
d
0 (1)
where

g c
A
c
A0
and

z o t
( ) , which has the solution (by analogy
with Example 4.1-1)

c
A
c
A0
1 erf
z
o
(2)
with o given by Eq. 20.1-70. To get Eq. 20.1-72, we differentiate the
concentration profile:

N
Az0
D
AB
oc
A
oz
z0
D
AB
c
A0

o
oz
erf
z
4D
AB
S t
( )
S t
( )

1
]
2
dt
0
t
[

1
]
1
1
1
z0
(3)
Use Eq. C.6-2 to differentiate the error function, and get

N
Az0
+D
AB
c
A0
2
r
4D
AB
S t
( )
S t
( )

1
]
2
dt
0
t
[

1
]
1
1 2

c
A0
D
AB
rt
1
t
S t
( )
S t
( )

1
]
2
dt
0
t
[

1
]
1
1 2
(4)
The total number of moles of A that have crossed the mass-transfer
surface S(t) at time t is then given by:

M
A
S t
( )
c
A0
1 erf z / o
( )

1
]
0
~
[
dz S t
( )
c
A0
erfc z / o
( )
0
~
[
dz (5)
145
The complementary error function "erfc (...)" is defined in C.6. We
now insert this function into Eq. 5:

M
A
S t
( )
c
A0
2
r
e

2
z/o
~
[
0
~
[
ddz S t
( )
c
A0
o
2
r
e

2
d z o
( )
d
0

[
0
~
[
(6)
In the second form, the order of integrations over and

z o has been
interchanged. Now the integral over

z o can be performed to give

M
A
S t
( )
c
A0
o
2
r
e

2
0
~
[
d S t
( )
c
A0
o
2
r

1
2
(7)
Inserting the expression for o from Eq. 20.1-70, we then get

M
A
S t
( )
c
A0
1
r
4D
AB
S t
( )
S t
( )

2
dt
0
t
[

c
A0
4D
AB
r
S t
( )

2
dt
0
t
[
(8)
which is Eq. 20.1-73.
Another expression can be obtained from integrating Eq. 20.1-
72:

M
A
S t
( )
N
Az0
( t )dt
0
t
[

c
A0
4D
AB
r
S t
( )
S t
( )
S t
( )

2
dt
0
t
[
0
t
[
dt (9)
It is relatively easy to show that Eqs. 8 and 9 are the same. We first
note that the

S t
( )
in the denominator can be removed from the
integral, so that Eq. 9 can be rewritten as

M
A
c
A0
4D
AB
r
S t
( )

2
S t
( )

2
dt
0
t
[
0
t
[
dt (10)
146
Then we make use of the fact that the square root in the denominator
contains no

t and hence can be removed from the integral over

t to
give

M
A

c
A0
4D
AB
r
S t
( )

2
dt
0
t
[
S t
( )

2
dt
0
t
[

c
A0
4D
AB
r
S t
( )

2
dt
0
t
[

c
A0
4D
AB
r
S t
( )

2
dt
0
t
[
(11)
This is the same as Eq. 8. Proving that the two expressions for

M
A
are
the same is a stronger statement than that the two expressions give
the same results for

dM
A
dt .
This example is also a good illustration of the importance of
using a symbol for the dummy variable of integration that is different
from that used as one of the limits in the integral.
147
Note to p. 626
We want to verify that the limiting solutions for slow and fast
reactions in Eqs. 20.2-16 to 18 and 20.2-19 are correct.
(a) Slow reactions
We start by rewriting Eq. 20.2-13 in terms of dimensionless
variables. We do this by using the definition of in Eq. 20.2-17:

1
Sc

4
3

d
d
A
3
+ A
3
+ A
2
(1)
Then, using the expansion in Eq. 20.2-16, we have

A Sc
1 3
1 + a
1
+ a
2

2
+
( )
(2)

A
2
Sc
2 3
1 + 2a
1
+ a
1
2
+ 2a
2
( )

2
+
( )
(3)

A
3
Sc
1
1 + 3a
1
+ 3a
1
2
+ 3a
2
( )

2
+
( )
(4)
Substitution of these three expressions into Eq. 1 gives

1
4
3
3a
1
+ 2 3a
1
2
+ 3a
2
( )

2
+
( )
+ 1 + 3a
1
+ 3a
1
2
+ 3a
2
( )

2
+
( )

+ Sc
1 3
1 + 2a
1
+ a
1
2
+ 2a
2
( )

2
+
( )
(5)
We now equate terms in the same powers of :
Zeroth power of : 1 = 1 (6)
First power of :

0 4a
1
+ 3a
1
+ Sc
1 3

or

a
1

1
7
Sc
1 3
(7)
Second power of :

0
8
3
a
1
3
+ 4a
1
a
2
( )

2
+ a
1
3
+ 4a
1
a
2
( )

2
+ 2Sc
1 3
a
1

2
or

0 11 a
1
2
+ a
2
( )
+ 2Sc
1 3
a
1
148
or

0 11
1
49
Sc
2 3
+ a
2
( )
+ 2Sc
1 3

1
7
Sc
1 3
( )
or

a
2
+
3
539
Sc
2 3
(8)
Thus we have obtained the first two coefficients in the expansion of
Eq. 20.2-16, and hence the first few terms in the expression for A as a
function of .
(b) Fast reactions
When the trial function for A is taken to be of the form

A K
m

m < 0 (9)
When this is substituted into Eq. 1, we get

1
Sc

4
3
3m
( )
K
3

3m
+ K
3

3m
+ K
2

2m+1
(10)
The ratio of the last term to either of the first two terms is
proportional to

m+1
, which is a positive quantity. For large , the
last term is the dominant term on the right side. Therefore Eq. 1
becomes

1
Sc
A
2
(11)
and therefore the solution to Eq. 1 for this case is

A Sc ( )
1 2
(12)
in agreement with Eq. 20.2-19.
149
Note to p. 628
Example 20.2-2 is, for the most part easy to follow, except for the
steps going from Eq. 20.2-30 to Eq. 20.2-34. Here we provide those
missing steps.
First, we assume H will be a function of q alone. Then the
derivatives of H with respect to x and y must be converted to
derivatives with respect to q, using Eq. 20.2-33:

oH
ox

_
,

dH
dq
oq
ox

_
,

dH
dq
y
v
~
2v

1
2
x
3 2

_
,


dH
dq

1
2
q
x

_
,

(1)

oH
oy

_
,

dH
dq
oq
oy

_
,

dH
dq
v
~
2vx

dH
dq
q
y

_
,

(2)

o
2
H
oy
2

_
,

o
oy
dH
dq
v
~
2vx

_
,


d
2
H
dq
2
v
~
2vx

d
2
H
dq
2
q
y

_
,

2
(3)
Then Eq.20.3-30 becomes

H
v
dH
dq

1
2
q
x

_
,

+
v
0
x ( )
v
~

o
ox
H
v
dy
0
y
[

_
,

dH
dq
q
y

_
,


v
v
~
A
d
2
H
dq
2
q
y

_
,

2
(4)
Now multiplication by

v
~
v
( )
y q
( )
2
leads to

H
v
dH
dq
q +
v
0
x
( )
v
~

o
ox
H
v
dy
0
y
[

_
,

2v
~
x
v
dH
dq

1
A
d
2
H
dq
2
(5)
(a) (b) (c) (d)
Our attempt to use a combination of variables to get an equation for
H q
( ) has not been successful because of the appearance of the term
(b), which contains both x and q. Also term (c) may contain both x
and q. We shall show that term (c) contains only q, and later (look
ahead to Eq. 20.2-37) arguments are given that

v
0
x
( )
v
~
( )
2v
~
x v
( )
may be set equal to a constant, K.
150
Let us, therefore, proceed to an analysis of term (c). We begin
by replacing the integral over y by an integral over q:

o
ox
H
v
dy
0
y
[

o
ox
H
v
dq
2vx
v
~
0
q
[

_
,



v
2v
~
x

_
,

H
v
dq
0
q
[ ( )

2vx
v
~
o
ox
H
v
dq
0
q
[ ( )
(6)
Then the derivative with respect to x is replaced by a derivative with
respect to q:

o
ox
H
v
dy
0
y
[

v
2v
~
x

_
,

H
v
dq
0
q
[ ( )

2vx
v
~
d
dq
H
v
dq
0
q
[

_
,

y
v
~
2v

1
2
x
3 2

_
,



v
2v
~
x

_
,

H
v
dq
0
q
[ ( )
+
v
2v
~
x

_
,

H
v
q (7)
When Eq. 7 is multiplied by

2v
~
x v dH dq , we then get

o
ox
H
v
dy
0
y
[

_
,

2v
~
x
v
dH
dq
H
v
dq
0
q
[ ( )
dH
dq
+ H
v
q
dH
dq
(8)
(

c
1
) (

c
2
)
Thus, term

c
2
cancels term a, and the remaining term appears in the
final equation for H:

K x
( )
H
v
dq
0
q
[ ( )
dH
dq

1
A
d
2
H
dq
2
(9)
which is Eq. 20.2-34.
When K is taken to be constant, and the quantity in parentheses
is set equal to

f , then the velocity profile

H
v
may be calculated
from Eq. 20.2-39 with the associated boundary conditions in Eqs.
20.2-40 to 42. The remaining profiles may be obtained from Eq. 20.2-
151
43, which is obtained by setting

dH dq=A q
( ) in Eq. 9 and then
solving the first-order differential equation for A q
( ) . One obtains

dA
dq
+ Af A 0 (10)
the solution of which is

dH
dq
A C
1
exp A fdq
0
q
[ ( )
(11)
A further integration gives

H q, A, K
( )
C
1
exp A f q, K
( )
dq
0
q
[ ( )
dq
0
q
[
+ C
2
(12)
Then the constants of integration are obtained from the boundary
conditions in Eqs. 20.2-35 and 36, with the final result for the profiles

H q, A, K ( )
exp A f q, K
( )
dq
0
q
[ ( )
dq
0
q
[
exp A f q, K
( )
dq
0
q
[ ( )
dq
0
~
[
(13)
which is the same as Eq. 20.2-43.
152
Note to p. 692
We begin by formulating the problem as in Problem 12.1-4:
Solid Liquid

oC
s
ot

1

2
o
o

2
oC
s
o

_
,

(1)

oC
s
o
1
NC
l
(5)

C
s
finite at

0 (2)

C
s
C
l
at

1 (3)

C
s
1 at

t 0 (4)
Taking the Laplace transform of the problem we get
Solid Liquid

pC
s
1
1

2
d
d

2
dC
s
d

_
,

(6)

dC
s
d
1
NC
l
(9)

C
s
finite at

0 (7)

C
s
C
l
at

1 (8)
The solution to the homogeneous equation corresponding to Eq. 6 is
the complementary function

C
s,cf

K
1

cosh p +
K
2

sinh p (10)
The particular integral of Eq. 6 is, by inspection

C
s,pi

1
p
(11)
The complete solution to Eq. 6 is then

C
s

K
1

cosh p +
K
2

sinh p +
1
p
(12)
153
The boundary condition in Eq. 7 requires that

K
1
0. The boundary
condition in Eq. 8 requires that

C
s
1
K
2
sinh p +
1
p
C
l
(13)
and combination of Eqs. 9 and 12 gives

NC
l
K
2
sinh p + p cosh p
( )
(14)
Elimination of

C
l
between Eqs. 13 and 14 leads to an expression for
the integration constant

K
2
, and hence also to

C
s
:

C
s

1
p

N
p

_
,

sinh p
p cosh p + N 1 ( )sinh p
(15)
Next, the Laplace transform of the total amount of A within the
sphere is

M
A
4rR
3
c
0
C
s
0
1
[

2
d

1
3p

N
p
2
xsinh x
p cosh p + N 1 ( )sinh p
0
p
[
dx

1
3p

N
p
2
xsinh x
p cosh p + N 1 ( ) p
0
p
[
dx

1
3p

N
p
2
xcosh x sinhx
p cosh p + N 1
( )
sinh p

_
,

0
p

1
3p

N
p
2
p cosh p + N 1
( )
sinh p Nsinh p
p cosh p + N 1
( )
p

_
,

1
3p

N
p
2
1
Nsinh p
p cosh p + N 1
( )
p

_
,

1
3p

N
p
2
p cosh p + N 1 ( )sinh p Nsinh p
p cosh p + N 1
( )
sinh p

_
,

154

1
3p

N
p
2
1
Nsinh p
p cosh p + N 1
( )
sinh p

_
,

1
3p

N
p
2
+
N
2
p
2
p coth p + N 1
( )

1
]
(16)
Now we take the inverse Laplace transform of the above; the
transforms of the first two terms may be found in an elementary table
of transforms:

M
A
4rR
3
c
0

1
3
Nt + N
2
L
1
1
p
2
p coth p + N 1
( )
( )

(17)
To get the inverse transform of the last term, we can use the
Heaviside partial fractions expansion theorem for repeated roots,
which is:
If

f p
( )
N p
( )
D p
( ) with

D p ( ) p a
1
( )
m
1
p a
2
( )
m
2
p a
n
( )
m
n
,

N p
( ) is a polynomial of degree less than

Zm
j
( )
1, and

a
i
= a
k
for

i = k , then

f t ( )
d
kl
a
k
( )
m
k
l ( )! l 1 ( )!
l1
m
k
_
k1
n
_
t
m
k
l
e
a
k
t
with

d
kl
p
( )

d
l1
dp
l1
N p
( )
D
k
p
( )

_
,

and

D
k
p
( )

D p
( )
p a
k
( )
m
k
.
The contribution from the factor

p
2
is then (with

a
1
0 ,

m
1
2, k=1)

d
11
0
( )
2 1
( )
! 1 1
( )
!
t +
d
12
0
( )
2 2
( )
! 2 1
( )
!
d
11
0 ( )t + d
12
0 ( )
(18)
where
155

d
11
0
( )
t
N
2
p coth p + N 1
( )
p0
t
N
2
1 + N 1
( )
t Nt (19)

d
12
0 ( )
d
dp
N
2
p coth p + N 1 ( )

1
]
1
1
p0

N
2
d dp
( )
p coth p + N 1
( )
( )
p coth p + N 1
( )
( )
2

1
]
1
1
1
p0


1
3
(20)
The contributions in Eqs. 19 and 20 just exactly cancel the first two
terms in Eq. 17.
The contribution from the remaining factor in

L
1
{ is

M
A
t ( )
4rR
3
c
0


L
1
N p
( )
D p
( )

N a
k
( )
D' a
k
( )
k2
~
_
e
+a
k
t
(21)

N
2
e
+a
k
t
p
2
1
2
p
1 2
coth p
1
2
csch
2
p
( )
+ 2p p coth p + N 1
( )
( )

1
]
pa
k
k2
~
_
------------------------------
where the prime on D indicates differentiation with respect to p, and
the

a
k
are the zeros of the denominator in the braces in Eq. 17 (except
for the double zero from

p
2
, which we have already taken into
account). This means that the dashed underlined term in Eq. 21 may
be omitted.
We know on physical grounds that the quantity M(t) must be a
decreasing function of time. Therefore, the

a
k
must be negative. This
can be guaranteed by setting

a
k
equal to

i
k
2
where the

i
k
are real
numbers. Then we have

L
1
N p ( )
D p
( )

2N
2
sinh
2
p
p
2
p
1 2
sinh p cosh p 1
( )

1
]
pi
k
2
k2
~
_
e
i
k
2
t
156

2N
2
i
2
sin
2
i
k
ii
k
i
k
4
( )
i sini
k
cosi
k
ii
k
( ) k2
~
_
e
i
k
2
t

2N
2
sin
2
i
k
i
k
3
( )
i
k
sini
k
cosi
k
( ) k2
~
_
e
i
k
2
t
(22)
Therefore, the total amount of A within the sphere is at any time t is:

M
A
t ( )
4
3
rR
3
c
0
6N
2
B
n
n1
~
_
exp i
n
2
D
AB
t / R
2
( )
(23)
where the

i
n
are determined from

i
n
cot i
n
+ N 1
( )
0 (24)
and the

B
n
are

B
n

N
2
sin
2
i
n
i
n
3
i
n
sini
n
cosi
n
( )
(25 )
For infinite N these last two expressions may be simplified. If, in Eq.
24, N is infinite,

sini
k
must be zero, and therefore

i
k
must be

nr . If,
in Eq. 25,

N ~

B
n

i
n
cot i
n
( )
2
sin
2
i
n
i
n
3
i
n
sini
n
cosi
n
( )

i
n
2
i
n
4

1
nr
( )
2
(26)
Thus we obtain the results in Eqs. 22.4-34 and 35.
Two references should have been cited here:
E. N. Lightfoot, in Lectures in Transport Phenomena, AIChE, New
York (1969), pp. 59-60.
H. Grber, S. Erk, and U. Grigull, Die Grundgesetze der Wrme-
157
bertragung, Springer-Verlag, Berlin, 3rd edition (1961), pp. 55-62.
158
Note to p. 766
Here we work through the details of the development on p. 766,
leading up to the expression for the generalized diffusional driving
force at the bottom of the page.
When Eq. 24.1-2 is applied to a moving element of fluid, we
write

D

U
Dt
T
D

S
Dt
p
D

V
Dt
+
G
o
M
o
Du
o
Dt
o1
N
_
(1)
in which

p
D

U
Dt
V q
( )
r:Vv
( )
+ j
o
g
o
( )
o1
N
_
Table 19.2-4, Eq. D (2)
[Note footnote b to Eq. D!!]

D

V
Dt

D
Dt
1
p

_
,


1
p
2
Dp
Dt
+
1
p
V v
( )
Table 19.2-3, Eq. A (3)

p
Du
o
Dt
V j
o
( )
+ r
o
Table 19.2-3, Eq. B (4)
Substitution of these expressions into Eq. 1 gives

p
D

S
Dt

1
T
V q
( )
r:Vv
( )
+ j
o
g
o
( )
o1
N
_

_
,

+
p
T
V v
( )

1
T
G
o
M
o
V j
o
( )
+ r
o
( )
o1
N
_


1
T
V q
( )

G
o
M
o
V j
o
( )
o1
N
_

_
,


1
T
t:Vv
( )

1
T
G
o
M
o
r
o
o1
N
_

+
1
T
j
o
g
o
( )
o1
N
_
159

V
1
T
q
G
o
M
o
j
o
o1
N
_

_
,

_
,

q
1
T
2
VT

_
,

+

j
o
V
1
T
G
o
M
o

1
T
g
o

1
]
1

_
,

o1
N
_

1
T
t:Vv
( )

1
T
G
o
M
o
r
o
o1
N
_
(5)
Comparison of Eq. 5 with Eq. 24.1-1 gives the entropy flux vector and
the entropy production rate (Eqs. 24.1-3 and 4):

s
1
T
q
G
o
M
o
j
o
o1
N
_

_
,

(6)

g
S
q
1
T
2
VT

_
,

j
o
V
1
T
G
o
M
o

1
T
g
o

1
]
1

_
,

o1
N
_

1
T
t:Vv
( )

1
T
G
o
M
o
r
o
o1
N
_
(7)
Next we rewrite Eq. 6 by replacing q by

q
h ( )
+ H
o
M
o
( )
j
o
o1
N
_
--that is,
we subtract off the heat flux associated with the diffusion of the
chemical species. This gives us then for the entropy flux (Eq. 24.1-5):

s
1
T
q
h ( )

S
o
M
o
j
o
o1
N
_

_
,

(8)
The entropy flux is now written as the sum of two terms: the first
term is the entropy flux associated with heat flow, and the second
term is that connected with diffusion of the chemical species. (For
more on the reasons for replacing q by

q
h ( )
+ H
o
M
o
( )
j
o
o1
N
_
, see S. R.
de Groot and P. Mazur, Non-Equilibrium Thermodynamics, North-
Holland, Amsterdam (1962), pp. 24-25.)
In order to rewrite the entropy production term, we make use
of the Gibbs-Duhem relation in the entropy representation (see H. B.
Callen, Thermodynamics and an Introduction to Thermostatics, Wiley,
New York (1985), pp. 60-62):
160

Ud
1
T

_
,

+Vd
p
T

_
,

n
o
d
G
o
T

_
,

o1
N
_
0 (9)
where

n
o
is the number of moles of species o . Division by V and
doing some elementary manipulations gives us the form of the
Gibbs-Duhem equation that we need:

p
o
H
o
M
o
o1
N
_
1
T
2
VT +
1
T
Vp p
o
V
1
T
G
o
M
o

_
,

0
o1
N
_
(10)
Next, in Eq. 7, we add four terms inside the bracket in the second
term on the right side in such a way that the term is not changed:

g
S
q
1
T
2
VT

_
,


j
o
p
o

p
o
V
1
T
G
o
M
o

1
T
p
o
p
Vp + p
o
H
o
M
o
1
T
2
VT
p
o
H
o
M
o
1
T
2
VT
1
T
p
o
g
o
+
1
T
p
o
p
p

1
N
_

1
]
1
1
1
1
1

_
,

o1
N
_

1
T
t:Vv
( )

1
T
G
o
M
o
r
o
o1
N
_
(11)
The terms containing

H
o
clearly cancel one another, and the terms
containing

Vp and

Z

do not contribute to the sum, inasmuch as



Z

0 . Thus, Eq. 11 may now be rewritten as:



g
S
q
h ( )

1
T
2
VT

_
,


j
o
p
o

p
o
V
1
T
G
o
M
o

1
T
p
o
p
Vp + p
o
H
o
M
o
1
T
2
VT

1
T
p
o
g
o
+
1
T
p
o
p
p

1
N
_

1
]
1
1
1
1
1

_
,

o1
N
_

1
T
t:Vv
( )

1
T
G
o
M
o
r
o
o1
N
_
(12)
or
161

Tg
S
q
h ( )
VlnT
( )

j
o
p
o
cRTd
o

_
,

o1
N
_

1
T
t:Vv
( )

1
T
G
o
M
o
r
o
o1
N
_
(13)
The

d
o
are the generalized driving forces for diffusion . Note that the
above development guarantees that

Z
o
d
o
0, as is required, since

Z
o
j
o
0. We can see that

Z
o
d
o
0, since the first three terms in the
bracket in the first line of Eq. 12 sum to zero according to the Gibbs-
Duhem equation, and the fourth and fifth terms clearly combine to
sum to zero.
The above development has led us to the expression for the
generalized driving forces for diffusion:

cRTd
o
c
o
TV
G
o
T

_
,

+ c
o
H
o
VlnT u
o
Vp p
o
g
o
+u
o
p

1
N
_
(14)
To get the alternative form for

d
o
given in the second line of Eq. 24.1-
8, we follow the discussion of Ref. 5 on p. 766.
The quantities

G
o
,

H
o
, and p are functions of the state of a fluid
element, which may be described by the temperature T, the pressure
p, and the set of (N 1) mole fractions

x
o
where o = 1, 2,3, ... ( N 1).
Then the first term in Eq. 14 may be written by expanding

VG
o
using
the chain rule of partial differentiation to give:

c
o
TV
G
o
T

_
,

c
o
VG
o
c
o
G
o
VlnT

c
o
oG
o
ox

_
,

Vx

+ c
o
oG
o
oT

_
,

VT + c
o
1
N1
_
oG
o
op

_
,

Vp c
o
G
o
VlnT

c
o
oG
o
ox

_
,

1
N1
_
Vx

c
o
TS
o
VlnT + c
o
V
o
Vp c
o
G
o
VlnT (15)
162
When Eq. 15 is combined with Eq. 14, it is seen that the

VlnT terms
exactly cancel, and the

Vp terms may be combined. Furthermore,
when the relation

dG
o
RTdlna
o
is used, we get finally

cRTd
o
c
o
RT
olna
o
olnx

_
,

T, p, x
x

1
N1
_
+ o
o
u
o
( )
p p
o
g
o
+ u
o
p

1
N
_
(16)
in which

o
o
c
o
V
o
is the volume fraction of species o , and the
subscript x stands for all the

x

except

x
o
and

x
N
. Thus, we see that
the driving forces

d
o
include the contributions from the mole fraction
gradients, the pressure gradient, and the external forces.
Equation 16 is in agreement with Eq. 7.8 of Ref. 5 at the bottom
of p. 766, and also with Eq. 11.1-27 of Molecular Theory of Gases and
Liquids, by Hirschfelder, Curtiss, and Bird, Wiley, New York (1964).
Therefore, the first term of the second line of Eq. 24.1-8 of BSL may be
misleading. Here and elsewhere (Eqs. 24.2-8, 24.2-9, 24.2-10, 24.4-1,
and 24.5-4) the abbreviated notation

olna
o
olnx

_
,

T, p,x
Vx

1
N1
_
V
T, p
lna
o
(17)
is used (this is just the chain rule applied to

Vlna
o
with the

Vp and

VT terms omitted). This notation has also been used by E. N.
Lightfoot, Transport Phenomena and Living Systems, Wiley, New York
(1974), Eqs. 1.2.7 and 12, on p. 163, and W. M. Deen, Analysis of
Transport Phenomena, Oxford University Press (1998), Eq. 11.6-2.
Deen, however, states that his treatment is for dilute mixtures. Both
of these authors, however, appear to have their summations going
from 1 to N (instead of 1 to (N 1).
The discussion in Multicomponent Mass Transfer, by R. Taylor
and R. Krishna, Wiley, New York (1993), also uses essentially the
notation of Eq. 17 above, with the summation going from 1 to (N 1)
as may be seen on pp. 23, 24, and 29 (their Eq. 2.3.10 is, aside from
some notational differences, the same as Eq. 16 above).
163
In Advanced Transport Phenomena, by J. C. Slattery, Cambridge
University Press (1999), p. 450, Eq. 8.4.3-4, gives a result that is
consistent with Eq. 16 above, although the notation is considerably
different from ours.
There is also a discussion of multicomponent systems in
Transport Phenomena Fundamentals, by J. L. Plawsky, pp. 66-69 (heat
flux) and pp. 69-73 (mass fluxes); however, no derivations of the
expressions are given.
Thermodynamics of Irreversible Processes, by G. D. C. Kuiken,
Wiley, New York (1994), has a discussion of the driving force

d
o
on
p. 183.
164
Note to p. 767
Just as a check, we show that Fick's law for a binary system
may be obatained by simplifying Eq. 24.2-3. This equation is, for a
binary system with the two components labeled "A" and "B":

j
A
+p
A
D
AA
d
A
+ D
AB
d
B
( ) (1)
Use Eqs. (A) and (C) of Table 24.2-1 and the equation

d
A
+ d
B
0 (see
line 2 on p. 767), to eliminate the generalized Fick diffusivities in
favor of the Maxwell-Stefan diffusivities:

j
A
+p
A

u
B
2
x
A
x
B
D
AB
d
A

u
A
u
B
x
A
x
B
D
AB
d
A

_
,

(2)
This may be rewritten, using the fact that

u
A
+ u
B
1, to give

j
A

p
A
u
B
x
A
x
B
D
AB
d
A
(3)
Next use the second line of Eq. 24.1-8, omitting the terms for thermal
diffusion, pressure diffusion, and forced diffusion. This yields:

j
A

p
A
u
B
x
A
x
B
D
AB
x
A
Vln a
A
( )
p
A
u
B
x
A
x
B
D
AB
oln a
A
oln x
A

_
,

Vx
A
(4)
Then, to obtain the molar flux of species "A" we use Eq. 17B.3-1 to get
the relation

J
A
*
cx
A
x
B
pu
A
u
B
( )
j
A
. Finally, combine this equation
with Eq. 4 and introduce the binary diffusivity

D
AB
from Eqs. 24.3-2
and 4, to get:

J
A
*
cD
AB
Vx
A
(5)
This is the form of Fick's (first) law given in Eq. B of Table 17.8-2 for a
binary system.
165
Note to p. 768
Here we show how to derive the generalized Maxwell-Stefan
relations (Eq. 24.2-4) from the generalized Fick equations (Eq. 24.2-3).
This proof was first given by H. J. Merk, Appl. Sci. Res., 73-99 (1959),
5, Eqs. 89 and 90.
We start by introducing the abbreviation

j
o
#
j
o
+ D
o
T
VlnT, in
order to avoid having to carry along the thermal diffusion term
throughout the derivation. Then the generalized Fick equation, Eq.
24.2-3, for species o may be written as

j
o
#
p
o
D
oy
d
y
y 1
N
_
(1)
where

j
y
#
y
_
0,

d
y
y
_
0 ,

D
oy
D
yo
, and

u
o
o
_
D
oy
0 (see above Eq.
24.2-3). Next, write a similar equation for species and form the
difference of the two equations; then multiply both sides of the
equation by the quantity

x
o
x

D
o
to get

x
o
x

D
o
j

#
p

j
o
#
p
o

_
,


x
o
x

D
o
D
y
D
oy
( )
d
y
y 1
N
_
(2)
Now we sum both sides over the index

x
o
x

D
o
=o
_
j

#
p

j
o
#
p
o

_
,


x
o
x

D
o
D
y
D
oy
( )
=o
_

1
]
1
1 y 1
N
_
d
y
(3)
If the bracket quantity on the right side is set equal to o
oy
u
o
, we
see that we get

x
o
x

D
o
o
_
j

#
p

j
o
#
p
o

_
,

o
oy
u
o
( )
y 1
N
_
d
y
d
o
u
o
d
y
d
o
u
o
0
( )
y 1
N
_
d
o
166
or

x
o
x

D
o
=o
_
j

#
p

j
o
#
p
o

_
,

d
o
(4)
This is exactly the generalized Maxwell-Stefan equation for species o
(see Eq. 24.2-4). At the same time we get

x
o
x

D
o
D
y
D
oy
( )

o
_
o
oy
u
o
(5)
From this we will get the relations between the

D
o
and the

D
o
.
One might wonder why we replace the bracket expression in
Eq. 3 by o
oy
u
o
rather than simply o
oy
, which would, after all, also
lead to the generalized Maxwell-Stefan equation. If Eq. 5 is multiplied
by u
y
and summed over
y
, one gets the identity 0 = 0, because of

u
o
o
_
D
oy
0. But if the u
o
is not included on the right side of Eq. 5,
multiplication of the equation by u
o
and summing over y will give 0
= u
o
. Similar arguments may be made for not setting the bracket
expression equal to

o
oy
Au
o
, where A is an arbitrary constant.
Let us now return to 5 and get the relation between the Fick
and the Maxwell-Stefan multicomponent diffusivities. We start by
defining a matrix

B
o
with matrix elements

B
o
( )
y
D
y
+D
oy
; the
matrix

B
o
is an

N 1
( )
N 1
( ) matrix, with the ( o )-row and the
(y o )-column missing. We now rewrite Eq. 5, omitting the equation
for y o :

x
o
x

D
o
B
o
( )
y

=o
_
u
o
(y ) (6)
Next we perform some operations on the relation

u

0:

D
y
D
oy
( )
+

D
oy
0 or

u

( )
y
+

_
D
y
0 (7)
167
Then we multiply Eq. 6 by

B
o
1
( )
yv
and sum on y to get:

x
o
x
v
D
ov

o
B
o
1
( )
yv
y

(8)
Multiplying the second form of Eq. 7 by

B
o
1
( )
yv
and summing on y
gives

v
D
oy
B
o
1
( )
yv
y

(9)
Then multiplying the reciprocal of Eq. 8 by Eq. 9, and replacing the
index v by the index , gives Eq. 24.2-7 of the textbook:

D

D
y
adjB

( )
y
y =

adjB

( )
y
y =

(10)
In the last step we have also made use of the fact that

B
o
1
( )
adjB
o
( )
det B
o
( )
(11)
where

adjB

is the matrix adjoint to



B

and

det B

is the determinant
of the matrix

B

. The matrix

adjB

is the transpose of the matrix of


cofactors (or "signed minors") of

B

.
168
Note to p. 769
As a check on Eq. 24.2-7, we use it to get Eq. (B) of Table 24.2-2.
We first write Eq. 24.2-7 for the specific case of the 1-2 pair of a
ternary system:

D
12

x
1
x
2
u
1
u
2
D
12
adjB
1
( )
22
+D
13
adjB
1
( )
32
adjB
1
( )
22
+ adjB
1
( )
32
(1)
The 2 x 2 matrix

B
1
( ) may be displayed thus:

B
1
( )

B
1
( )
22
B
1
( )
23
B
1
( )
32
B
1
( )
33


D
22
+D
12
D
23
+ D
13
D
32
+ D
12
D
33
+D
13

(2)
The adjoint matrix is the transpose of the matrix of cofactors (a
cofactor is a "signed minor"):

adj B
1
( )

adj B
1
( )
22
adj B
1
( )
23
adj B
1
( )
32
adj B
1
( )
33


B
1
( )
33
B
1
( )
23
B
1
( )
32
B
1
( )
22

D
33
+ D
13
D
23
D
13
D
32
D
12
D
22
+ D
12

(3)
We are now ready to substitute into Eq. 24.2-7:

D
12

x
1
x
2
u
1
u
2
D
12
D
33
+D
13
( )
+ D
13
D
32
D
12
( )
D
33
+D
13
( )
+ D
32
D
12
( )

x
1
x
2
u
1
u
2
D
12
D
33
+ D
13
D
32
D
12
D
33
+ D
13
+ D
32

_
,

x
1
x
2
u
1
u
2
+D
12
D
33
D
13
D
23
D
12
+D
33
D
13
D
23

_
,

(4)
In the last step, we have made use of the symmetry of the

D

You might also like