Вы находитесь на странице: 1из 288

Course Reader for ENSC3007 Heat and Mass Transfer

By

Professor Hui Tong Chua


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Contents
Chapter 1 Introduction to Conduction ................................................................................ 3
1.1 The conduction rate equation .................................................................................... 3
1.2 The heat diffusion equation ...................................................................................... 8
1.3 Boundary and initial conditions .............................................................................. 16
Chapter 2 One-dimensional, steady-state conduction ....................................................... 21
2.1 The plane wall ......................................................................................................... 21
2.1.1 Thermal resistance ........................................................................................... 22
2.2 The composite wall ................................................................................................. 24
2.3 Cylindrical system .................................................................................................. 28
2.4 Conduction with thermal energy generation ........................................................... 35
2.4.1 The plane wall .................................................................................................. 35
2.4.2 Radial systems ................................................................................................. 41
2.5 Heat transfer from finned surfaces .......................................................................... 45
2.5.1 A general conduction analysis ......................................................................... 46
2.5.2 Fins of uniform cross-sectional area ................................................................ 48
2.6 Fin performance ...................................................................................................... 54
Chapter 3 Introduction to convection ............................................................................... 60
3.1 A basic convection heat transfer issue .................................................................... 60
3.2 The convection boundary layers ............................................................................. 63
3.2.1 The velocity boundary layer ............................................................................ 63
3.2.2 The thermal boundary layer ............................................................................. 64
3.2.3 The concentration boundary layer ................................................................... 66
3.3 Laminar and turbulent flows ................................................................................... 68
3.4 Physical significance of the dimensionless parameters .......................................... 70
3.5 The boundary layer approximations ....................................................................... 73
3.6 Boundary layer similarity ....................................................................................... 73
3.7 Evaporative cooling ................................................................................................ 79
3.8 Analogy between momentum and heat transfers .................................................... 82
Chapter 4 External flow .................................................................................................... 84
4.1 The flat plate in parallel flow .................................................................................. 84
4.1.1 Laminar flow .................................................................................................... 84
4.1.2 Turbulent flow ................................................................................................. 86
4.1.3 Mixed boundary layer conditions .................................................................... 86
4.2 The cylinder in cross flow....................................................................................... 93
Chapter 5 Internal flow ................................................................................................... 101
5.1 Flow conditions ..................................................................................................... 101
5.2 The mean velocity ................................................................................................. 102
5.3 Pressure gradient and friction factor in a fully developed flow ............................ 102
5.4 Thermal conditions ............................................................................................... 105
5.5 The mean temperature........................................................................................... 105
5.6 Energy balance ...................................................................................................... 106
5.6.1 Constant surface heat flux .............................................................................. 107
5.6.2 Constant surface temperature ......................................................................... 109

Prof. Hui Tong Chua 1


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
5.7 Laminar flow in circular tubes .............................................................................. 112
5.7.1 The entry region ............................................................................................. 114
5.8 Turbulent flow in circular tubes ............................................................................ 115
Chapter 6 Transient conduction ...................................................................................... 119
6.1 The lumped capacitance method ........................................................................... 119
6.2 Validity of the lumped capacitance method.......................................................... 120
Chapter 7 Radiation: Processes and Properties ............................................................... 126
7.1 Fundamental concepts ........................................................................................... 126
7.2 Radiation intensity ................................................................................................ 127
7.2.1 Definitions...................................................................................................... 128
7.2.2 Irradiation ....................................................................................................... 128
7.2.3 Radiosity ........................................................................................................ 129
7.3 Blackbody radiation .............................................................................................. 130
7.3.1 The Planck distribution .................................................................................. 131
7.3.2 Band emission ................................................................................................ 133
7.4 Surface emission ................................................................................................... 137
7.5 Surface absorption, reflection and transmission ................................................... 141
7.6 Kirchhoff’s law ..................................................................................................... 143
Chapter 8 Radiation exchange between surfaces ............................................................ 150
8.1 The view factor ..................................................................................................... 150
8.1.1 View factor relations ...................................................................................... 150
8.2 Blackbody radiation exchange .............................................................................. 156
8.3 Radiation exchange between diffuse, gray surfaces in an enclosure .................... 159
8.3.1 Net radiation exchange at a surface ............................................................... 159
8.3.2 Radiation exchange between surfaces ........................................................... 160
8.3.3 The two-surface enclosure ............................................................................. 164
8.3.4 Radiation shields ............................................................................................ 165
8.3.5 The reradiating surface .................................................................................. 167
Chapter 9 Diffusion Mass Transfer................................................................................. 171
9.1 Physical origins and rate equations ....................................................................... 171
9.2 Mass transfer in nonstationary media ................................................................... 177
9.2.1 Evaporation/desorption in a column .............................................................. 179
9.2.2 Diffusivities of gases and vapors ................................................................... 184
9.3 Conservation of species for a stationary medium ................................................. 185
9.4 Stationary media with specified surface concentrations ....................................... 188
9.5 Boundary conditions and discontinuous concentrations at interfaces .................. 191
9.5.1 Evaporation and sublimation ......................................................................... 191
9.5.2 Solubility of gases in liquids and solids......................................................... 192
Tutorial questions and solutions ..................................................................................... 197

Prof. Hui Tong Chua 2


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Chapter 1 Introduction to Conduction

1.1 The conduction rate equation

Consider a cylindrical rod of say, aluminum, which is insulated on its lateral surface as
shown in figure 1.1. Its two end faces are maintained at different temperatures T1 and T2
with T1>T2 for a long time. To fix ideas, we can take T1 to be 100C and T2 to be 50C.

A = 2cm2, A = 2cm2,
T1 = 100C T = T1-T2 = 50C T2 = 50C

qx qx = 23.7W

x = 0.1m
x
Fig 1.1 A steady-state heat conduction experiment.

The rod has a constant cross sectional area, A and a length, x. Let us say A = 2cm2 and
x = 0.1m. If we measure the heat transfer rate, qx, at the two end faces, we find that heat
flows from the left end face to the right end face and qx is about 23.7W.

Let us consider the situation when we fix everything to be constant and increase the cross
sectional area by 50%, so that A becomes 3cm2. In this case, we find that qx also
increases by 50% so that it is now about 35.6W. In fact, we find that qx is directly
proportional to A, qx  A.

Now, let us shorten the rod so that it is one tenth the original length or 0.01m, we will
find that qx increases ten folds and becomes 237W. If the rod is one fifth the original
length or 0.02m, then qx increases five folds and is 118.5W. In short, we find that qx is
inversely proportional to the length or x, qx  1/x.

Finally, let us increase the temperature difference, T, by 100% to 100C, we find that qx
increases by 100% as well. In contrast, if T1 is increased to 120C and T2 to 70C so that
T remains unchanged, we find that qx remains the same. After a few more experimental
trials, we can then conclude that qx  T.

Prof. Hui Tong Chua 3


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

From the above few experiments, we can state that

T
qx  A (1.1).
x

Let us now change the material of the cylindrical rod, say from aluminum to pyrex. We
will find that while the heat transfer rate is significantly reduced for the same conditions,
relation (1.1) still holds. This implies that the proportionality constant is material
dependant and is therefore a property of the material. We can therefore write

T
q x  kA (1.2)
x

where k is the thermal conductivity of the material (in W/mK). Allowing x to approach
zero, we obtain the differential relation for the heat transfer rate

dT
q x  kA (1.3).
dx

Note that the minus sign is necessary as heat is always transferred in the direction of
decreasing temperature. We can also accordingly write down the differential relation for
the heat flux

qx dT
q x   k (1.4).
A dx

Equation (1.4) is known as the Fourier’s law. Since heat flux is a measure of the amount
of heat transferred across an area, it is a vectorial quantity. In other words, it is a quantity
with both magnitude and direction. More specifically, the direction of heat flux is normal
to a surface with a constant temperature. We call such a surface an isothermal surface.

Prof. Hui Tong Chua 4


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Normal line, n
Normal line, n

q n
q n in the direction of
decreasing temperature,
T<T1

T = T1 =constant
An isothermal surface
Figure 1.2 The directions of heat flux vectors from an
isothermal surface

From figure 1.2, we can also write the Fourier’s law as

dT
q n  k (1.5)
dn

Note that the empirical Fourier’s law1 is used to define the thermal conductivity. In other
words,

q x
k (1.6)
dT dx

The Fourier’s law is applicable to the solid, liquid and gaseous phases.

1
The Fourier’s law is established simply on experimental observations, hence it is empirical.

Prof. Hui Tong Chua 5


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

1.3a Thermal conductivities of graphite and diamond

10000

9000
Pyrolytic graphite, parallel to layers

8000 Pyrolytic graphite perpendicular to layers


Thermal conductivity (W/m.K)

7000 Diamond

6000

5000

4000

3000

2000

1000

0
0 200 400 600 800 1000 1200 1400 1600 1800 2000
Temperature (K)

Fig 1.3b Thermal conductivities of selected solids

500
Aluminum
450 Aluminum alloy 2024
Copper
400 Gold
Iron
Thermal conductivity (W/m.K)

350 Stainless steel 304


Platinum
300 Silver
Tungsten
250 Aluminum oxide

200

150

100

50

0
100 500 900 1300 1700 2100 2500
Temperature (K)

Prof. Hui Tong Chua 6


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Fig 1.4a Thermal conductivities of selected liquid metals

100

90

80
Thermal conductivity (W/m.K)

70
Sodium
60

50

40
Potassium
30

20
Bismuth

10
Mercury
0
200 300 400 500 600 700 800 900 1000 1100 1200
Temperature (K)

Fig 1.4b Thermal conductivities of selected liquids

0.80

0.70
Water
0.60
Thermal conductivity (W/m.K)

0.50

0.40

0.30
Glycerin

0.20
Engine oil
(unused)
0.10

0.00
250 300 350 400 450 500 550 600
Temperature (K)

Prof. Hui Tong Chua 7


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Fig 1.5 Thermal conductivity of various gases

0.45

0.40 Helium
Hydrogen Hydrogen
0.35 Carbon dioxide
Air
Thermal conductivity (W/m.K)

0.30 Water vapor at 1 atm

0.25
Helium
0.20

0.15

0.10

0.05
Air
0.00
0 100 200 300 400 500 600 700 800 900 1000
Temperature (K)

1.2 The heat diffusion equation

Referring to the figure 1.6 below, consider a scenario where an electric current passes
through a long cylindrical conducting rod. For safety reason, the rod is insulated with an
electrically nonconducting material. The passage of electricity gives rise to uniform
resistive heating within the rod. To prevent overheating, fluid is passed transversely
across the rod to provide convective cooling. A pertinent question to ask is the
temperature distribution within the rod if we know the material properties?

A current
carrying rod

Cladding Cooling fluid

Fig 1.6 What is the temperature distribution within the rod?

Prof. Hui Tong Chua 8


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Referring to figure 1.7 below, consider a plane wall which is fixed with a heater on one
side. The plane wall is initially at the room temperature. The heater is then switched on
while at the same time air is being blown along the other side of the wall. How does the
temperature distribution within the wall change with time?

Air
Wall
Insulator

Heater

Fig 1.7 How does the temperature


distribution in the wall change with time?

To answer questions such as these, we have to develop the equation that governs the
behavior of a temperature field or distribution in a material. We call such an equation a
heat diffusion equation.

Consider a differential control volume in a material medium depicted using the Cartesian
coordinate system as shown in figure 1.8 below.

Prof. Hui Tong Chua 9


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

z
y
x

qz+dz
qy+dy

dz

E g
qx qx+dx
E st

qy
dy
dx

qz
Fig 1.8 Differential control volume for a heat diffusion analysis

The conduction heat rates at the (x, y, z) coordinate location along the x, y and z locations
are given by qx, qy and qz, respectively. The conduction heat rates at the opposite
surfaces can then be expressed as2

q x
q x dx  q x  dx (1.7a)
x
q y
q y dy  q y  dy (1.7b)
y
q
q z dz  q z  z dz (1.7c)
z

Within the material medium, there may also be an energy source term, such as that due to
the passing of an electric current over a resistive material. This term can be represented
as

2
We make use of the Taylor series expansion here and neglect higher order terms.

Prof. Hui Tong Chua 10


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

E g  q dxdydz (1.8)

where q is the rate of energy generation per unit volume (W/m3). Additionally, the
amount of internal energy stored in the control volume may also change with time. This
is due to an imbalance in the rate of influx and outflux of heat as well as energy
generation within the control volume. The rate of change of internal energy stored in the
control volume can be expressed as

T
E st  c p dxdydz (1.9)
t

T
where c p is the rate of change of internal energy per unit volume.
t

Now we invoke the conservation of energy which requires

E in  E g  E out  E st (1.10)

where E in is the total influx of heat into the control volume and E out the total outflux of
heat from the control volume, so that we have

T
q x  q y  q z  q dxdydz  q x dx  q ydy  q z dz  c p dxdydz (1.11)
t

From (1.7a-c) we have

q x q y q T
q x  q y  q z  q dxdydz  q x  dx  q y  dy  q z  z dz  c p dxdydz
x y z t

or

q x q y q T
 dx  dy  z dz  q dxdydz  c p dxdydz (1.12)
x y z t

From the Fourier’s law, we have

T
q x  kdydz (1.13a)
x
T
q y  kdxdz (1.13b)
y
T
q z  kdxdy (1.13c)
z

Prof. Hui Tong Chua 11


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Substituting (1.13a-c) into (1.12), we accordingly have

  T    T    T  T
   kdydz dx    kdxdz dy    kdxdy dz  q dxdydz  c p dxdydz
x  x  y  y  z  z  t

  T    T    T  T
k    k    k   q  c p (1.14).
x  x  y  y  z  z  t

Equation (1.14) is the general form of the heat diffusion equation (or otherwise known as
the heat equation) in Cartesian coordinates. From its solution, we can determine the
temperature within the material medium as a function of space and time.

If we are not dealing with large temperature range, the thermal conductivity can often be
treated as a constant, so that the heat equation can be simplified into

 2 T  2 T  2 T q c p T 1 T
     (1.15)
x 2 y 2 z 2 k k t  t

where  is the thermal diffusivity. If we are dealing with steady-state conditions, so that
there is no change in the energy stored within the control volume, (1.14) can be
simplified as

  T    T    T 
k   k   k   q  0 (1.16).
x  x  y  y  z  z 

In addition, if we are only dealing with one-dimensional heat transfer (say in the x
direction) with no heat generation within the medium, (1.16) reduces to

  T 
k 0 (1.17).
x  x 

(1.17) implies that under steady-state, one-dimensional conditions with no energy


generation within the medium, the heat flux ( qx  k T x ) remains a constant in the
direction of heat transfer since qx x  0 .

The heat equation can also be expressed in the cylindrical and spherical coordinate
systems. In the cylindrical coordinate system fig 1.9, it can be shown that

Prof. Hui Tong Chua 12


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

z
q z
r
q 

T(r, , z)

R q r

x
Fig 1.9 The cylindrical coordinate system.

1   T  1   T    T  T
 kr  2  k    k   q  c p (1.18),
r r  r  r     z  z  t

and

T k T T
q r  k q    q z  k (1.19).
r r  z

Similarly in the spherical coordinate system as shown in fig 1.10, it can be shown that the
heat equation takes the form

1   2 T  1   T  1   T  T
 kr  2  k   2  k sin    q  c p (1.20),
r r 
2
r  r sin      r sin   
2
  t

and

T k T k T
q r  k q    q    (1.21).
r r sin   r 

Prof. Hui Tong Chua 13


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

q r

q 

r

y
 q 

x
Fig 1.10 The spherical coordinate system.

Example 1.1 (Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer,


1990, John Wiley & Sons)

The temperature distribution across a wall 1m thick at a certain instant of time is given as

T(x) = 900 - 300x - 50x2

where T is in degree Celsius and x is in meters. The wall has a cross sectional area of 10
m2, and is subjected to a uniform heat generation, q  1000 W m 3 . The relevant
thermophysical properties of the wall are  = 1600 kg/m3, k = 40 W/mK, and cp = 4
kJ/kgK.

Prof. Hui Tong Chua 14


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

T=900-300x-50x2
2
A=10m

Internal heat
generation

qin qout
Internal energy
storage

L = 1m

Determine the rate of heat transfer entering the wall (x = 0m) and leaving the wall
(x = 1m).

The rate of heat transfer entering the wall, qin, is given by

dT
q in  q x x  0  kA  kA 300  50  2  0
dx x 0
q in  kA 300  50  2  0  40  10   300  120kW .

Similarly,

dT
q out  q x x  1m   kA  kA 300  50  2  1
dx x 1m
q out  kA 300  50  2  1  40  10   300  100  160kW

Determine the rate of change of energy storage in the wall.

The rate of change of energy storage can be determined by applying the principle of the
conservation of energy. We note that

E in  E g  E out  E st

where E g  q AL  1000  10  1  10kW . Hence,

Prof. Hui Tong Chua 15


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

E  E  E  E
  q  q AL  q
st in g out in out
E  120kW  10kW  160kW
st

 30kW

Determine the rate of change of temperature at x = 0, 0.25 and 0.5m.

The rate of change of temperature at any location within the material medium can be
determined by applying the heat equation, so that from (1.15)

c p T  2 T q
 
k t x 2 k

T k  2T q
 
t c p x 2
c p

Now, since T(x) = 900 - 300x - 50x2

 2 T   T 
  
x 2 x  x 

  300  100x 
x
 100  C m 2

Hence,

T k  2T q
 
t c p x 2 c p
40 1000
   100 
1600  4000 1600  4000
4 4
 6.25  10  1.56  10
 4.69  10  4  C s

We can therefore conclude that the rate of change of temperature is uniform everywhere
within the wall, so that everywhere within the wall is cooling down at the same rate.

End of example

1.3 Boundary and initial conditions

Prof. Hui Tong Chua 16


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
The heat equation described above dictates how the temperature distributes in space and
time within the material medium. In order to get to the particular solution for the
problem at hand, we need to further specify the conditions at the boundary of the medium
and, if the problem is time dependant, the spatial temperature distribution at a particular
point in time, which is called the initial condition.

Referring to the heat equation

  T    T    T  T
k    k    k   q  c p (1.14),
x  x  y  y  z  z  t

we see that it is second order in each space coordinate and first order in time. Hence we
require two boundary conditions for each of the three space coordinates and one initial
condition.

Prof. Hui Tong Chua 17


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Table 1.1 delineates the three types of boundary conditions commonly encountered.

1. Constant surface temperature


T(0,t) = Ts (1.22) Ts

T(x, t)

2. Constant surface heat flux


(a) Finite heat flux
T
k  q s (1.23) q s
x x 0
T(x, t)

(b) Adiabatic or insulated surface


T
0 (1.24)
x x 0

T(x, t)

3. Convection surface condition


T T, h T(0, t)
k  hT  T0, t  (1.25)
x x 0

T(x, t)

The constant surface temperature condition is closely approximated when the surface is
in contact with a boiling liquid, a melting solid or, in general, a medium undergoing
phase change. Thus heat transfer at the surface will continue while the temperature is
maintained constant.

The finite heat flux condition at the surface is realized when heat is supplied by an
electric heater. An adiabatic or a perfectly insulated surface condition is closely
approximated when the surface is well insulated.

Prof. Hui Tong Chua 18


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Finally, the convection surface condition is realized when the surface is subjected to
convective heating or cooling.

Example 1.2 (Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer,


1990, John Wiley & Sons)

A long rectangular copper bar, with its width w being much greater than its thickness t, is
maintained in a good thermal contact with a heat sink at its lower surface. The
temperature of the heat sink is To. Initially, the temperature of the entire bar is identical
with that of the heat sink so that it is also at To throughout. At a point in time, an electric
current I is allowed to pass through the bar while a stream of air at a temperature T  is
blown over the top of the bar as shown in the figure below. All the while, the bottom
surface of the bar is maintained at To.

Air: T, h
T, h z
T(L,t)
y
q T(x,y,z,t)
t x
x x w  T(x,t)

t Heat sink
T(0,t) = To To
I

Find the differential equation, boundary conditions and initial condition needed to
determine the temperature as a function of space and time within the bar.

We note that w >> t, so that heat transfer is primarily one dimensional along the x
direction. Hence from (1.14)

  T    T    T  T
k    k    k   q  c p ,
x  x  y  y  z  z  t

and considering constant thermophysical properties, we have

 2 T q c p T
  (1).
x 2 k k t

The two appropriate boundary conditions are

T
T(0, t) = To and k  hTL, t   T  (2),
x x L

Prof. Hui Tong Chua 19


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
and the initial condition is

T(x, 0) = To (3).

If the thermophysical properties are known together with the resistivity of the copper bar,
the equation together with the two boundary conditions and the initial condition can then
be solved for the spatial and temporal temperature field.

Intuitively, when the electrical power is sufficiently large, the spatial and temporal
temperature profiles can be readily sketched out as shown below. Interestingly, we can
see that the free stream temperature serves as a reference temperature so that when T(L,t)
< T heat is being transferred from the free stream to the copper bar, when T(L,t) = T no
heat transfer occurs between the free stream and the bar, and when T(L,t) > T the
direction of heat transfer reverses such that heat flows from the bar to the free stream.
T(x,t)

T(x,)

T T

T(x,0)
To

0 Distance, x L

The corresponding heat fluxes at x = 0 and x = L are sketched below.


Heat flux

+
qx L, t 

0
qx 0, t 

-
0 Time, t 

End of example

Prof. Hui Tong Chua 20


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Chapter 2 One-dimensional, steady-state conduction

2.1 The plane wall

Let us consider the situation when a plane wall separates two streams of fluids as shown
in figure 2.1 below. You can think of this wall being the wall of your house.

x
qx
T, 1
Ts, 1
Cold fluid
Ts, 2 T, 2, h2
Hot fluid
T, 2
T, 1, h1
L
T, 1 Ts, 1 Ts, 2 T, 2

qx
1 L 1
h 1A kA h 2A
Figure 2.1 Steady-state heat transfer through a wall.

The temperature distribution in the wall can be determined by solving the heat diffusion
equation by employing the appropriate boundary conditions. From the heat equation, its
relevant form for the present problem is

d  dT  d
k 0 or q x   0 (2.1).
dx  dx  dx

From (2.1), it is clear that the heat flux across the wall is a constant. Treating the thermal
conductivity as being practically a constant, we have from (2.1)

dT
 C1
dx

T = C1x + C2 (2.2).

Prof. Hui Tong Chua 21


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

The relevant boundary conditions are

T(0) = Ts,1 and T(L) = Ts,2.

By applying the first boundary conditions, it then follows that

C2 = Ts,1, and

by applying the second boundary condition, we have at x = L

Ts, 2  Ts,1
Ts,2 = C1L + Ts,1 or C1  .
L

Hence, from (2.2), the temperature distribution within the wall is given by

x
Tx   Ts, 2  Ts,1   Ts,1 (2.3).
L

We conclude that the temperature varies linearly within the wall under a steady-state
condition and when there is no internal heat generation.

From the Fourier’s law, we can now determine the heat transfer rate across the wall so
that

dT kA
q x  kA  Ts,1  Ts,2  (2.4).
dx L

The corresponding heat flux is then

qx k
q x   Ts,1  Ts, 2  (2.5).
A L

2.1.1 Thermal resistance

It is now timely to introduce the concept of thermal resistance which is analogous to the
concept of electrical resistance. According to the Ohm’s law, we have for the electrical
resistance, Re

E s,1  E s, 2
Re  (2.6)
I

where E is the electric potential and I the current. Similarly, the thermal resistance for
conduction is defined to be

Prof. Hui Tong Chua 22


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Ts,1  Ts, 2
L
R t ,cond   (2.7).
qx kA
We can also ascribe a thermal resistance to the convection heat transfer at a surface.
From the Newton’s law of cooling,

q = hA(Ts - T) (2.8),

the thermal resistance for convection is defined to be

Ts  T 1
R t ,conv   (2.9).
q hA

The equivalent thermal circuit for the plane wall with convection on both sides is also
shown in figure 2.1. Based on the thermal circuit, the heat transfer rate can be expressed
as

T,1  Ts,1 Ts,1  Ts, 2 Ts, 2  T, 2


qx    (2.10).
1 h 1A L kA 1 h 2A

The same heat transfer rate can also be expressed in terms of the total temperature
difference, T,1 - T,2, and a total thermal resistance, Rtot as

T ,1  T , 2
qx  (2.11)
R tot

where Rtot is given by

1 L 1
R tot    (2.12).
h 1A kA h 2 A

In addition to the heat transfer resistances discussed above, the heat transfer surface may
also interact with its surrounding via radiation. In which case, we define a thermal
resistance for radiation as

Ts  Tsur 1
R t ,rad   (2.13).
q rad hrA

Now the net rate of radiation heat exchange between the surface and its surroundings is
expressed as


q rad   A Ts4  Tsur
4
 (2.14)

Prof. Hui Tong Chua 23


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
where  is the emissivity of the surface,  the Stefan-Boltzmann constant ( = 5.6710-8
W/m2K4) and the unit of temperature is in Kelvin3. (2.14) can be expressed alternatively
as

q rad   ATs2  Tsur


2
Ts  Tsur Ts  Tsur 

Ts  Tsur 
q rad 
1  T s
2
 Tsur
2
Ts  Tsur A
Hence, hr is given by


h r   Ts2  Tsur
2

Ts  Tsur  (2.15).

Surface radiation and convection resistances act in parallel. Furthermore, if T = Tsur, the
two resistances can be combined into one effective resistance, Rt,conv+rad as follow

1 1 1
  (2.16).
R t ,conv rad 1 h r A 1 hA

2.2 The composite wall

Let us now consider a slightly more complicated situation involving more than one layer
of wall, which we call a composite wall. Consider a series composite wall as in figure
2.2.

3
More will be said about radiative heat transfer in a later chapter.

Prof. Hui Tong Chua 24


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

x
T, 1
Ts, 1 T2
Cold fluid
T, 4, h4
Hot fluid Ts, 4
T3
T, 1, h1
T, 4
A B C

T, 1 Ts, 1 T2 T3 Ts, 4 T, 4

qx 1 LA LB LC 1
h 1A kAA k BA kCA h 4A

Figure 2.2 Steady-state heat transfer through a composite wall.

The heat transfer rate, qx, for this system can be expressed as

T ,1  T , 4
qx  (2.17)
R t

where the total thermal resistance is given by

1 L L L 1
R t   A  B  C 
h1A k A A k B A k C A h 4 A
(2.18).

The same heat transfer rate can also be expressed in terms of the temperature drop across
individual component as

T,1  Ts,1 Ts,1  T2 T2  T3 T3  Ts, 4 Ts, 4  T, 4


qx      (2.19).
1 h 1A LA k AA LB k BA LC k CA 1 h 4A

It is also expeditious to introduce the idea of an overall heat transfer coefficient, U in the
following form

q x  UAT,1  T, 4   UAT (2.20)

Prof. Hui Tong Chua 25


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
where T is the overall temperature difference. Comparing (2.20) and (2.17), it is clear
that

1
UA 
Rt
or

1 1
U  (2.21).
R tot A 1 h1   L A k A   L B k B   L C k C   1 h 4 

In short, we conclude that

T 1
R tot   R t   (2.22).
q UA

Example 2.1 (Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer,


1990, John Wiley & Sons)

A manufacturer proposes a self-cleaning oven design that makes use of a composite


window separating the oven cavity from the room air. The composite composes of two
layers of high temperature plastics (A and B) of thicknesses LA = 2LB and thermal
conductivities kA = 0.15 W/mK and kB = 0.08 W/mK. During the self-cleaning process,
the oven wall temperature and the air temperature within the oven, Tw and Ta, are 400C.
The room air temperature, T, is 25C. The inside convection and radiation heat transfer
coefficients hi and hr, as well as the outside convection coefficient ho, are each 25
W/m2K. Determine the minimum window thickness, L = LA + LB, needed to ensure a
temperature that does not exceed 50C for safety reason at the outer surface of the
window.

Oven cavity Composite


LA LB LA = 2LB
window
hr = 25 W/m2K
Tw = 400C Ts,o  50C
Air
Air
Ta = 400C
hi = 25 W/m2K kB = 0.08 W/mK
T = 25C kA = 0.15 W/mK
ho = 25 W/m2K

Prof. Hui Tong Chua 26


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
For our design purposes, we shall assume that the condition is steady state. This is
satisfactory, even though we know that the self-cleaning process will be a transient affair
as the temperature difference will be the greatest during steady state. So we will
introduce a good safety factor here as well aside from the fact that the analysis will be
more straightforward.

Let us sketch out the thermal circuit.

1
hrA
Tw
Ts, o T
Tw = Ta
LA LB 1 q
Ta
1 kAA k BA hoA
hiA

Now, we know that Ts,o  50C. This implies that

Ts,o  T
q  h o A50  25  25h o A
1 hoA

It is given that Tw = Ta. From energy balance, we also know that

Ta  T
q
Rt
where

1
 1 1  L L 1
 R t   1 h A  1 h A   k AA  k BA  h A
 i r  A B o

1 1 L L 1 
   A  B   .
A  hi  hr kA kB ho 
1  1  2 1  1 
     L   L A  2L B
A  h i  h r  k A k B  h o 
B

Hence, we deduce that

Prof. Hui Tong Chua 27


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Ta  T
 q  25h o A
1  1  2 1  1 
    L 
A  h i  h r  k A k B  h o 
B

Ta  T 1 1
 
25h o hi  hr ho
LB 
2 1

kA kB
400  25 1 1
 
 25  25 25  25 25 .
2 1

0.15 0.08
 0.0209m

Hence, the minimum thickness of the composite window is

L  0.0209  2  0.0209  0.0627m or 62.7mm .

End of example

2.3 Cylindrical system

Conduction heat transfer analyses for cylindrical systems are very common – just think of
the insulated piping that you commonly see. So imagine a hollow cylinder, whose inner
and outer surfaces are in contact with fluids at different temperatures as in figure 2.3
below.

Prof. Hui Tong Chua 28


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

r1 r2

Ts,2 Cold fluid


L
T,2, h2

Ts,1
Ts,1

Ts,2

Hot fluid
r1 r2
T,1, h1
Fig 2.3 Conduction across a hollow cylinder with convective
surface conditions.

For steady-state and one-dimensional (i.e., radial direction only) conditions without any
internal heat generation, the heat equation in cylindrical coordinates can be expressed as

1 d  dT 
 kr 0 (2.23)
r dr  dr 

where k may be a function of radius. Now the rate at which heat is transported in the
radial direction is, according to the Fourier’s law,

dT dT
q r  kA  k 2rL (2.24).
dr dr

Comparing (2.23) with (2.24), we conclude that the heat transfer rate, qr, is a constant in
the radial direction. Note that the radial heat flux, q r , which is given by  k dT dr , is
NOT a constant in the radial direction. In the context of figure 2.3, q r decreases with
radius.

Treating k to be a constant, (2.23) can then be integrated for the temperature field where

Prof. Hui Tong Chua 29


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

1 d  dT 
 kr 0
r dr  dr 

d  dT 
r 0
dr  dr 

dT
r  C1
dr

T  C1 ln r  C2 (2.25).

We require two boundary conditions to fully specify the temperature field. Let us
consider

T(r1) = Ts,1 and T(r2) = Ts,2, so that

Ts,1  Ts, 2 r


Tr   ln    Ts, 2 (2.26).
ln r1 r2   r2 

Inspecting (2.26), we note that the radial temperature distribution in a cylindrical system
is logarithmic instead of linear, as in the case of a plane wall.

Combining with (2.24), we can now obtain the expression for qr as

dT
q r  k 2rL 
dr
(2.27).
Ts ,1  Ts , 2

ln r2 r1  2Lk

Accordingly, the thermal resistance is of the form

ln r2 r1 
R t ,cond  (2.28).
2Lk

The thermal circuit commensurate with fig 2.3 is shown below.

Prof. Hui Tong Chua 30


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

T,1 Ts,1 Ts,2 T,2


qr
1 ln r2 r1  1
h 1 2r1 L 2kL h 2 2r2 L

Fig 2.4 Thermal circuit for a hollow cylinder with


convective surface conditions

Let us now consider a composite cylindrical wall as shown in figure 2.5.

r3

r1 r2

Ts,3 T,1

T2 Cold fluid Ts,1


L
T,3, h3
T2
Ts,1

Ts,3
T,3
Hot fluid
r1 r2 r3
T,1, h1

T,1 Ts,1 T2 Ts,3 T,3


qr
1 ln r2 r1  ln r3 r2  1
h 1 2r1 L 2k A L 2k B L h 3 2r3 L

Fig 2.5 Conduction across a hollow composite cylinder with convective surface conditions.

The heat transfer rate can be expressed as

Prof. Hui Tong Chua 31


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
T ,1  T ,3
qr  (2.29).
1 ln r2 r1  ln r3 r2  1
  
h 1 2r1 L 2Lk A 2Lk B h 3 2r3 L

We can also express the heat transfer rate in terms of an overall heat transfer coefficient,
where

T,1  T,3
qr   U1 A1 T,1  T,3  (2.30)
R tot

and where A1 = 2r1L and

1
U1  (2.31).
1 r r  r r  r1 1
 1 ln  2   1 ln  3  
h 1 k A  r1  k B  r2  r3 h 3

It is good to note that the choice of using U1A1 to calculate heat transfer coefficient is
completely arbitrary. One can choose to use U2A2 or U3A3 instead such that

U1A1 = U2A2 = U3A3 = Rtot-1 (2.32).

The specific form of U2, and U3 can readily be inferred from (2.30 and 31).

Example 2.2 (Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer,


1990, John Wiley & Sons)

A thin walled copper tube with a radius ri carries a low temperature refrigerant. The
refrigerant temperature, Ti, is less than the ambient temperature, T, around the exterior
of the tube. Insulation is needed to reduce heat loss from the refrigerant, if the insulation
material cost is negligible, does it make sense to keep increasing the insulation thickness?
In other words, will the insulation capability of the insulator increase with its thickness?

The answer to this question seems obvious, as we know from our common sense that
when we insulate a plane wall, the efficacy of insulation increases with the insulator’s
thickness. But less us be patient and set up the heat transfer equation for an analysis.

Since the thickness of the copper tube is thin, we can ignore the conduction thermal
resistance of the copper tube so that we can sketch the thermal circuit as

Prof. Hui Tong Chua 32


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

T Ti
qr
1 ln r ri 
h 2rL 2kL

The heat transfer rate can be expressed as

T  Ti T  Ti
qr    .
1 ln r ri  R tot

h 2rL 2Lk

There are therefore two terms contributing to the total thermal resistance, R tot. The effect
of the insulation thickness is manifested as r, so that if r = ri, its thickness is zero and if
r > ri, its thickness is equal to r-ri.

If we increase r, we appreciate that the convection thermal resistance decreases


monotonically, whereas the conduction thermal resistance increases monotonically.
Hence, it is clear that there exists a critical insulation thickness such that thermal
resistance is minimized! Let us cast this conclusion in concrete mathematical terms.
Since

1 ln r ri 
R tot   , to minimize it, we set
h 2rL 2Lk

dR tot
 0 . Hence,
dr

1 1
  0
h 2r L 2rLk
2

or

k
r .
h

To confirm that the above result does indeed yield a minimum thermal resistance, we
d 2 R tot 2 1
examine the sign of (=  ) at r = k/h. We find that we have
dr 2
h 2r L 2r 2 Lk
3

2h 2 h2
 0
2k 3 L 2k 3 L

and thus the thermal resistance is indeed a minimum at r = k/h.

Prof. Hui Tong Chua 33


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
We call this radius the critical insulation radius

rcr  k/h.

Mathematically speaking, below the critical radius, introducing an insulation results in a


greater heat transfer rate! It is only when the insulation radius goes beyond the critical
radius will the heat transfer rate decrease as the insulation thickens.

To fix ideas, let us consider how the thermal resistance varies for a 10mm diameter tube
insulated by varying thickness of cellular glass with a thermal conductivity k = 0.055
W/mK. The outer surface convection coefficient is 5 W/m2K.

Total thermal resistance as a function of insulation radius

Insulation thickness (m)


0.000 0.005 0.010 0.015 0.020 0.025 0.030 0.035 0.040
8 7

7 6
Total resistance
6
Thermal resistance (mK/W)

5
Conduction resistance
5
4

4
3
3

2
2
Convection resistance
1
1
rcr = 0.011 m r = 0.032 m
0 0
0.005 0.010 0.015 0.020 0.025 0.030 0.035 0.040 0.045
Insulation radius (m)

We can see that the insulation has to be more than 27mm thick before the thermal
resistance is higher than the case where no insulation is applied at all! If it is a steam
carrying pipe, the insulation thickness has better be much more than 27mm.

If the insulation radius is less than 11mm, the total thermal resistance decreases as
insulation material is added. This apparently anomalous trend is actually desirable in the
case of an electricity carrying wire. This is because the addition of electrical insulation
actually promotes heat dissipation from the wire to the surrounding!

Note that the issue of a critical insulation thickness is only important for small diameter
wires or tubes and for small convection coefficients so that rcr > ri. For a typical
insulation (k ~ 0.03 W/mK) and free convection in air (h ~ 10 W/m2K), rcr = k/h ~

Prof. Hui Tong Chua 34


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
0.003m. The small value indicates to us that normally ri > rcr and we need not be overly
concerned with the effects of a critical radius.

End of example

2.4 Conduction with thermal energy generation

We often encounter a current carrying medium, such as electrical wires, filament in a


lamp, thermoelectric material, and so on. In all these cases, electrical energy is being
converted into thermal energy within the material due to resistive heating. The rate at
which resistive energy, E g , is generated is governed by the Ohm’s law so that

E g  I 2 R e (2.33)

where I is the current and Re the electrical resistance.

If the power generation (in Watts) occurs uniformly within the material medium, then the
energy generation per unit volume (in W/m3) is then

E g I2R e
q   (2.34).
V V

Apart from resistive heating, internal energy generation can also be brought about due to
chemical reactions within the medium, or the absorption of electromagnetic radiation by
a semitransparent medium. The latter mechanism is precisely why one can use laser or
high flux polychromatic light to necrose cancerous tissue.

2.4.1 The plane wall

Consider the plane wall in figure 2.6 whose surfaces are maintained at different
temperatures Ts,1 and Ts,2 and in which there is uniform internal energy generation.

Prof. Hui Tong Chua 35


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

T(x)

Ts,1
q Ts,2

T,1, h1 T,2, h2
-L +L
x
Fig 2.6 Conduction in a plane wall with
internal heat generation.

For a material with constant thermal conductivity, the heat equation can be expressed as

d 2 T q
 0 (2.35)
dx 2 k

for which the general solution can be readily written as

q 2
T x  C1 x  C 2 (2.36).
2k

For the boundary conditions as in fig 2.6, where

T(-L) = Ts,1 and T(L) = Ts,2,

the constants are evaluated as

Ts, 2  Ts,1 q 2 Ts,1  Ts, 2


C1  and C2  L  ,
2L 2k 2

so that the temperature field is

q L2  x 2  Ts, 2  Ts,1 x Ts, 2  Ts,1


Tx   1  2    (2.37).
2k  L  2 L 2

The heat flux at any point within the wall can readily be obtained by applying the
Fourier’s law onto (2.37). As a result of the internal heat generation, the heat flux is no
longer a constant within the material medium even during steady state.

Prof. Hui Tong Chua 36


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Let us now specialize (2.37) to the situation where the two surface temperatures are
identical as in fig 2.7a so that Ts,1 = Ts,2 = Ts.

T(x)

Ts Ts
To
q

T , h T, h
-L +L
(a) x

To
Ts
T(x)

T, h
q
(b)

Fig 2.7 Conduction in a plane wall with


(a) symmetrical boundary conditions,
and (b) adiabatic boundary condition at
mid-plane.

As suggested by figure 2.7a, the temperature distribution is symmetrical about the mid-
plane and is given by

q L2  x2 
Tx   1  2   Ts (2.38).
2k  L 

The maximum temperature, To, is located at the mid-plane

q L2
T0  To   Ts (2.39).
2k

The same temperature distribution can alternatively be expressed in terms of To so that

Tx   To  x 
2

  (2.40).
Ts  To L

Prof. Hui Tong Chua 37


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Note that at the mid-plane or the plane of symmetry, the temperature gradient is zero, or
(dT/dx)x=0 = 0 as depicted in figure 2.7. Hence, the mid-plane can be equivalently
represented by an adiabatic surface as in figure 2.7b. Immediately we can extend the
result to describe the temperature distribution in plane walls where one surface is very
well insulated and the other is maintained at a fixed temperature.

What if we can not confidently measure the surface temperature? In which case how can
we make use of the temperature distributions derived so far for design and analysis?

We can definitely easily measure the free stream temperature, T. Then if we can
reasonably estimate the convective heat transfer coefficient, we can relate the surface
temperature Ts to T by energy balance

dT
k  h Ts  T  (2.41).
dx x L

Substituting (2.38) into (2.41) we have

q L
Ts  T  .
h

Hence, in the case of convective heat transfer at the surface, (2.38) can be alternatively
expressed as

q L2  x 2  q L
Tx   1  2    T (2.42)
2k  L  h

where the unknown surface temperature has been eliminated.

Example 2.3 (Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer,


1990, John Wiley & Sons)

A composite wall is made of two materials, A and B. There is a uniform internal heat
generation of q  1.5  10 6 W m 3 within the wall of material A which has a thickness LA
= 50mm, kA = 75 W/mK. The wall of material B does not have internal heat generation.
It has a thickness LB = 20mm, kB = 150 W/mK. The inner surface of material A is well
insulated, while the outer surface of material B is cooled by a stream of water with T =
30C and h = 1000 W/m2K.

Prof. Hui Tong Chua 38


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

160

140
T(0mm)=140 degC

120

T(50mm)=115 degC

100 T(70mm)=105 degC

Internal heat generation = 1.5 x 106 W/m3


kB = 150 W/m.K
80 kA = 75 W/m.K

60 Material A Material B

Water
2
h = 1000W/m .K
40 LA = 50mm LB = 20mm

30 degC

20
-10 0 10 20 30 40 50 60 70 80 90

It will be interesting if we can establish the temperature distribution within the composite
wall.

The appropriate equation to describe the temperature distribution in the composite wall is

d 2 T q
  0.
dx 2 k

q is zero in material B as there is no internal heat generation. For material A, the general
equation is

q
TA   x 2  C1 x  C 2 .
2k A

For material B, it is given by the simpler form

TB  D1 x  D 2 .

In order to evaluate the four arbitrary constants, we need four boundary conditions

dTA dTA dTB


 0,  kA   kB , TA x LA
 TB x LA
,
dx x 0 dx x LA dx x LA

Prof. Hui Tong Chua 39


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

and  k B
dTB
dx

 h TB x LA  LB

 T .
x LA  LB

Now

dTA q
 x  C1 ,
dx kA

Invoking the first boundary condition, we conclude that

C1 = 0.

Making use of the second boundary condition, we have

 q 
 k A   L A   k B D1
 kA 

q L
D1   A
kB

From the third boundary condition, we can establish that

q 2 q L2A
 LA  C2    D2
2k A kB

q L2A q L2A
C2     D2
kB 2k A

Finally from the fourth boundary condition, we have

 q L   q L 
 k B   A   h  A L A  L B   D 2  T 
 kB   kB 

 1 L  LB 
D 2  q L A   A   T
h k B 

Hence,

Prof. Hui Tong Chua 40


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

q L2A  1 L  L B  q L2A
C2    q L A   A   T
kB h k B  2k A

So,

q L2  1 L  LB 
TA 
q
2k A
 
L2A  x 2  A  q L A   A
kB
  T
k B 
h
1 L 
TB  
q L A
kB
 
x  L A  q L A   B   T
 h kB 

It is worthy to note that the heat flux that passes through material A is not a constant
which increases with the distance from the insulated surface, while it is a constant in
material B and given by q L A . This is also the amount of heat flux carried away by the
water.

End of example

2.4.2 Radial systems

qr

Ts q

Cooling fluid
T, h
Fig 2.8 Temperature distribution in a cylindrical rod with
internal heat generation.

Many material medium that generate heat internally come in the shape of cylinders (e.g, a
current carrying wire). For a material with constant thermal conductivity and uniform
internal heat generation, the steady-state heat diffusion equation can be expressed as

1 d  dT  q
r  0 (2.43).
r dr  dr  k

Prof. Hui Tong Chua 41


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Solving this equation, we have
dT q
r   r 2  C1
dr 2k

q 2
T r  C1 ln r  C 2 (2.44).
4k

The appropriate boundary conditions are

dT
0 and T(ro) = Ts.
dr r 0

The first boundary condition implies that C1 = 0. The second condition yields

q 2
C 2  Ts  ro .
4k

Hence, the temperature distribution can be expressed as

q ro2  r2 
T 1  2   Ts (2.45).
4k  ro 

The surface temperature, Ts, can be replaced in favor of the free-stream fluid temperature
T by an energy balance analysis

dT
k  h Ts  T 
dr r  ro


 q 
 k  ro   h Ts  T 
 2k 

q ro
Ts   T (2.46).
2h

Hence, the temperature distribution can alternatively be expressed as

q ro2  r 2  q ro
T 1  2    T (2.47).
4k  ro  2h

Prof. Hui Tong Chua 42


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Example 2.4 (Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer,
1990, John Wiley & Sons)

We have a long thick walled tube with uniform heat generation within the material as
shown below. A coolant is being passed through the tube so as to control its temperature.
The external surface of the tube is insulated.

Coolant
T, h

ri ro

q

What is the temperature distribution within the rod?

Invoking the general solution (2.44), we have

q 2
T r  C1 ln r  C 2 .
4k

The relevant boundary conditions are

dT dT
k 0 and k  hT  Tri  .
dr r  ro dr r  ri

From the first boundary condition, we have

q C
0 ro  1
2k ro

Prof. Hui Tong Chua 43


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

q ro2
C1  .
2k
From the second boundary conditions,

 q r q ro2   q r 2 q r 2 
 k  i    h T  i  o ln ri  C 2 
 2k 2k ri   4k 2k 

q ri  ro2  q ri2  r2 
C2   1  2   1  2 o2 ln ri   T .
2h  ri  4k  ri 

Hence, the temperature distribution within the material can be expressed as

q ri2  2
r2   2  ro2 
1  2 ro ln ri    qro ln r  qri

T 1  2   T .
4k  ri2 ri2  2k 2h  r 
  i 

We are now ready to investigate the heat removal rate per unit length of the rod by the
coolant

 dT 
q  r    k 2r 
 dr 
 r  ri 

 q  ro2  ri2
This is simply a statement that all the heat generated within the material medium must be
carried away by the coolant since the exterior surface of the tube is insulated4.

The temperature distribution within the tube material for the following parameters is
sketched below.

ri = 5mm
ro = 30mm
q  1.5  10 6 W m 3
k = 25 W/mK
h = 1000 W/m2K
T = 30C

4
The negative sign that appears in front of the Fourier’s law in the radial direction is necessary as heat is
being removed in the negative r-direction on the inner surface of the tube.

Prof. Hui Tong Chua 44


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Temperature distribution within the hollow tube material

200

180

160

140
Temperature (degC)

120

100

80

60

40
ri ro
20

0
0 5 10 15 20 25 30
Radial distance (mm)

End of example

2.5 Heat transfer from finned surfaces

Finned surfaces are prevalently used in thermal engineering applications to promote heat
transfer. You are probably aware that they are used for cooling engine heads on
motorcycles and lawn mowers or for cooling electric power transformers. You may have
also seen fins being attached to the condenser coil and evaporator of an air conditioner.
Figure 2.9 illustrates two typical finned-tube heat exchangers that you may have
encountered.

Prof. Hui Tong Chua 45


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Liquid flow
Liquid flow

Gas flow

Fig. 2.9 Two typical finned-tube heat exchangers.

Figure 2.10 below illustrates some of the typical fin configurations.

Straight fin Straight fin Pin fin Annular fin


Uniform cross section Non-uniform cross section
Fig 2.10 A few typical types of fins.

2.5.1 A general conduction analysis

As engineers, knowing the qualitative features of the efficacious effects of fins is not
good enough. It is important that we are able to quantify the analyses so as to facilitate
the design and application procedures.

Referring to figure 2.11, we can perform an energy balance to a differential fin element.

Prof. Hui Tong Chua 46


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

dAs dqconv

qx qx+dx

x Ac(x)
dx

z y
x

Fig 2.11 Energy balance for a fin.

Invoking the law of conservation of energy, we obtain


qx = qx+dx + dqconv (2.48).

From the Fourier’s law, we know that

dT
q x  kA c x  (2.49)
dx

where Ac(x) is the cross-sectional area of the fin and is in general a function of x. Now

dq x
q x dx  q x  dx (2.50)
dx

From (2.49), we can express it as

dT d  dT 
q x dx  kA c x   k  A c x  dx (2.51).
dx dx  dx 

From the Newton’s law of cooling,

dqconv = h dAs (T - T∞) (2.52)

where dAs is the surface area of the differential element. Substituting these expressions
into (2.48) we obtain

d  dT  h dA s
 A c x    T  T   0
dx  dx  k dx

Prof. Hui Tong Chua 47


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
or

d 2 T 1 dA c dT 1 h dA s
  T  T   0 (2.53).
dx 2 A c dx dx A c k dx

(2.53) is the working equation for the design and analysis of fins. We need to identify
two boundary conditions to obtain the specific solution for a problem at hand.

2.5.2 Fins of uniform cross-sectional area

Figure 2.12 shows some of the common configurations for fins with uniform cross-
sectional area.

T∞, h T∞, h

qconv qconv

Tb t Tb

Ac qf D
qf w x
x Ac
P = 2w + 2t L
L Ac = wt
P = D
Ac = D2/4
(a) (b)
Fig 2.12 Geometric characteristics of straight fins of uniform cross section. (a)
Rectangular fin. (b) Pin fin.

In order to apply (2.53), we require the geometric information about the fins. For fins
with uniform cross-sectional area such as figures 2.12a-b, Ac is a constant, and

As = Px (2.54)

where P is the perimeter of the fin. Hence (2.53) simplifies to

d 2T P h
 T  T   0 (2.55).
dx 2 A c k

Using the variable  defined as

Prof. Hui Tong Chua 48


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
(x)  T(x) - T∞ (2.56)

(2.53) is conveniently transformed into

d 2
2
 m2  0 (2.57).
dx

where

hP
m2  (2.58).
kA c

(2.57) is a standard linear, homogeneous, second-order ordinary differential equation


with the following general solution

x   C1e mx  C 2 e  mx (2.59).

We require two boundary conditions to solve for C1 and C2. The first condition is given
by

x  0  Tb  T   b (2.60).

The second boundary condition at x = L can correspond to two commonly encountered


different physical conditions. The first type of condition relates to convective heat
transfer at the tip of the fin so that

dT
 kA c  hA c TL   T 
dx x L

or

d
k  hL  (2.61).
dx x L

Invoking the two boundary conditions (2.60 and 2.61), we have

 b  C1  C 2 and kmC 2 e  mL  C1e mL   hC1e mL  C 2 e  mL  .

Solving for the two arbitrary constants and rearranging, we have

 T  T cosh mL  x   h mk sinh mL  x 


  (2.62).
 b Tb  T cosh mL  h mk sinh mL

The qualitative temperature profile along the length of the fin is shown in figure 2.13.

Prof. Hui Tong Chua 49


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

T∞, h

qconv
Tb

qf  k dT dx x L h[T(L)-T∞]

T
Tb

L x
Fig 2.13 A typical temperature profile along a uniform-cross-sectional-area fin
with convective heat loss at the tip.

What is the total amount of heat, qf, removed by the fin? The most straightforward
approach to evaluate this quantity is simply to note that

dT d
q f  kA c  kA c (2.63).
dx x 0 dx x 0

Combining with (2.62), we can readily conclude that

sinh mL  h mk cosh mL
q f  hPkAc b (2.64).
cosh mL  h mk sinh mL

The second type of boundary conditions corresponds to an insulated fin tip so that heat
loss from the tip is practically negligible and hence

d
0 (2.65).
dx x L

This implies that

Prof. Hui Tong Chua 50


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

C 1 e mL  C 2 e  mL  0 and  b  C1  C 2 5, so that

 cosh mL  x 
 (2.66).
b cosh mL

Combining with (2.63), we conclude that

q f  hPkA c  b tanh mL

Example 2.5

A 130mm-long rod with a 5mm diameter is maintained at one end at 100C. The rod is
subjected to an ambient temperature of 25C with a convective heat transfer coefficient
of 100W/m2K.

25C, 100W/m2K

100C

5mm

130mm

What are the temperature distributions if we were to use copper, aluminum and AISI316
stainless steel as materials for the rod?

The temperature distribution can be determined by (2.62) so that

 cosh mL  x   h mk sinh mL  x 


Tx    Tb  T   T .
 cosh mL  h mk sinh mL 

hP
where m  , P = D and Ac = D2/4.
kA c
We need to have the thermal conductivities of the three materials. The rough temperature
of the fin is (100 + 25)/2 or 62.5°C or 335K.

5
This is from the first boundary condition.

Prof. Hui Tong Chua 51


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Evaluation of thermal conductivities (W/mK)
Temperature (K)
Materials 200 400 335
Copper 413 393 398
Aluminum 237 240 239

Temperature (K)
300 400 335
316 Stainless Steel 13.4 15.2 14

Substituting the parameters into the equation for the temperature distribution, we obtain
the graph below.
Temperature distributions of fins of different materials

100
Copper
Aluminum
90
316 Stainless steel
Ambient temperature

80

70
Temperature (°C)

60

Copper

50
Aluminum

40

316 Stainless steel

30
Ambient temperature

20
0 20 40 60 80 100 120
Distance from the base of fin (mm)

From the graph, we can immediately appreciate that stainless steel is the least desirable
material for constructing the fin. The material starting from 30mm from the base of the
fin does not serve any purpose at all. Copper is the best material as it can maintain an
appreciable temperature difference with the ambience throughout the fin.

What is the amount of heat that is being extracted?

This can be determined by

sinh mL  h mk  cosh mL
q f  hPkA c
cosh mL  h mk sinh mL

For the copper, aluminum and 316 stainless steel, qf are 7.9W, 6.3W, and 1.6W,
respectively.

Prof. Hui Tong Chua 52


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Can we use a simpler way to estimate the temperature distribution along the fin?

We can estimate the temperature distribution by assuming that the heat loss at the tip of
the fin is negligible, so that from (2.66)

 cosh mL c  x 
Tx    Tb  T   T
 cosh mL c 

We correct for the loss in surface area for heat loss at the tip by using a corrected length
Lc. For a rectangular fin,

Lc = L+ t/2, where t is the fin thickness.

For a pin fin,

Lc = L + D/4, where D is the fin diameter.

Temperature distributions of fins of different materials

100
Copper
Aluminum
316 Stainless steel
90 Ambient temperature
Adibatic tip
Adiabatic tip
Adiabatic tip
80

70
Temperature (°C)

60

Copper

50
Aluminum

40

316 Stainless steel

30
Ambient temperature

20
0 20 40 60 80 100 120
Distance from the base of fin (mm)

The amount of heat loss from the fin can also be estimated more easily

q f  hPkA c  b tanh mL c

For the copper, aluminum and 316 stainless steel, qf are 7.9W, 6.3W, and 1.6W,
respectively. There is no change in the results!!

Prof. Hui Tong Chua 53


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

2.6 Fin performance

There are several ways to evaluate the performance of a fin. The first index is the fin
effectiveness f. It is defined as the ratio of fin heat transfer rate to the heat transfer rate
that would exist without the fin or

qf
f  (2.67)
hA c,b b

where Ac,b is the cross-sectional area of the fin at the base. Practically speaking, the use
of a fin is justified when f  2.

Let us inspect the design implications of this concept of fin effectiveness. For the
insulated-tip fin condition, f can be expressed as

kP kP  hP 
f  tanh mL  tanh L  (2.68).
hA c hA c  kA c 

We can appreciate that increasing the ratio of perimeter to cross-sectional area promotes
the fin effectiveness. This explains why thin fins which are closely packed are generally
preferred provided that the close packing does not severely militate against the fluid flow
between the fins.

 hP 
For a long fin, so that tanh L   1 , we can also appreciate using a material with a
 kA c 
high thermal conductivity heightens the fin effectiveness. Similarly the fin effectiveness
will be boosted when the convection coefficient, h, is small. This immediately implies
that the effectiveness of a fin is bigger when it is used in a gas than in a liquid. Also, it is
more effective when the convection is free instead of forced 6 . This is why in the
evaporator coil of an air conditioner, you will find the fins at the external surface of the
tube where heat is being extracted from the flowing air with a low convection coefficient,
and not at the internal surface of the tube where heat is being absorbed by the boiling
refrigerant with a high convection coefficient.

 hP 
Note that since tanh L  is a monotonic increasing function of L7,
 kA c 

kP
f  when L  ∞ (2.69).
hA c

6
In general, a free convection has a smaller convection heat transfer coefficient than a forced convection.
7
tanh(0) = 0, and tanh(∞) = 1.

Prof. Hui Tong Chua 54


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

This theoretical maximum effectiveness is achieved at the price of an exorbitant material


cost. In a practical situation, we desire the actual fin effectiveness to be

kP
 f ,actual ~ 0.98 (2.70).
hA c

This will require tanh(mL) to be 0.98, which translates to mL = 2.3. Hence, in a practical
design, L  2.3/m.

Another index to gauge the performance of a fin is via the fin efficiency f. It is defined
as the ratio of the actual heat transfer from the fin to the ideal heat transfer if the
temperature along the entire fin equals the base temperature or

qf qf
f   (2.71).
q max hA f b

For a straight fin of uniform cross section and with an adiabatic tip, (2.71) translates to

f 

hPkAc b tanh hP kA c L tanh mL

 (2.72).
hPLb mL

hP
where m  . Since we are invariably dealing with effective fins, their thickness is
kA c
2h
usually small and hence P  2w. Since Ac = wt, m can be simplified to .
kt

Prof. Hui Tong Chua 55


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Fig. 2.14 Fin efficiency of a rectangular straight fin

100

90

80

70
Fin efficiency (%)

60

50 Lc = L + t/2

40
t

30
L
20

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
mLc or (2h/kt)1/2Lc

Figure 2.15 illustrates the efficiency of annular fins of rectangular profile. Note that
2h
m .
kt
Fig 2.15 Fin efficiency of annular fins of rectangular profile

100

90

80

70
Fin efficiency (%)

60

50

1 = r2c/r1
40
t
2
30
3
r2c = r2 + t/2
Lc = L + t/2
20 5
L
r1
10 r2

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
mLc or (2h/kt)1/2Lc

We have so far been focusing on the efficiency of a single fin. Very often, we are
interested in the efficiency of an array of fins together with their affiliated base area.

Prof. Hui Tong Chua 56


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Fig 2.16a A Fig 2.16b An annular-fin


rectangular-fin array. array.

The overall efficiency, o, is defined as

qt qt
o   (2.73)
q max hA t b

where qt is the total heat transfer rate from the surface area At which includes both the
total surface area of the fins and the base area which is not covered by the fins8. At can
be represented as

At = NAf + Ab (2.74)

where N is the total number of fins in the array, Af is the surface area of each fin, and Ab
is the area of the prime surface. qt can be expressed as

q t  Nf hAf b  hAbb (2.75).

Combining with (2.74), we have

 NAf 
q t  hNf Af  A t  NAf b  hA t 1  1  f b (2.76).
 At 

Substituting this result into (2.73), we obtain

NAf
o  1  1  f  (2.77).
At

8
This exposed surface is often termed the prime surface.

Prof. Hui Tong Chua 57


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Hence, knowing the individual fin performance and the number of fins in the array, we
can calculate the overall efficiency of a fin array. Then based on (2.73), one can arrive at
the total heat transfer rate of the fin array.

Example 2.6 (Adapted from FP Incropera and DP DeWitt, Introduction to Heat Transfer,
4th ed., 2002)

An engine cylinder is constructed of 2024-T6 aluminum alloy. Its height is H = 0.15m


and outer diameter is D = 50mm. Under the rated operating conditions, the outer surface
of the cylinder is at 500 K, the ambient air is at 300 K. The affiliated convection
coefficient is 50 W/m2K. The cylinder is integrally cast with annular fins to enhance
heat transfer. Consider an array with five fins, with each fin having a thickness t of 6mm,
length L of 20mm. All the fins are equally spaced apart. What is the heat-transfer
enhancement with the use of the fins?

Tb = 500K

T∞ = 300K
h = 50 W/m2K
t = 6mm
H= 0.15m

r1 = 25mm
L = 20mm

Firstly, the thermal conductivity of 2024-T6 aluminum at 400K is 186 W/mK.

The heat-transfer rate in the presence of the array of fins is given by (2.76)

 NAf 
q t  hA t 1  1  f b .
 At 

    
2
Af  2 r22c  r12  2 45 103  6 103 2  25 103    0.0105m
2 2
, and
A t  NAf  2r1 H  Nt   0.0527  20.0250.15  5  0.006  0.0716m2 . We need to
make use of figure 2.15 to ascertain the fin efficiency. Now r2c/r1 = 1.92, Lc = 0.023m,
and (2h/kt)1/2Lc = 0.22. Hence, f = 0.98.

The total heat-transfer rate is given by

Prof. Hui Tong Chua 58


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

 0.0527
q t  50  0.07161  1  0.98 200  705W .
 0.0716 

Without the use of fins, the heat-transfer rate will be

q  h 2r1Hb  50  2  0.025  0.15  200  236W.

Thus the heat-transfer enhancement is

q  q t  q  705  236  469W .

Prof. Hui Tong Chua 59


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.

Chapter 3 Introduction to convection

3.1 A basic convection heat transfer issue

Let us consider the following scenario.

u, T

q" As, Ts

x dx L
Fig. 3.1 Convection heat transfer

A fluid with a velocity u and temperature T flows over a flat plate with a surface area
As and maintained at a uniform temperature Ts. If Ts  T, convection heat transfer will
occur. The local heat flux, q”, is given as

q  hTs  T  (3.1)

where h is the local convection coefficient. We expect that the flow condition to differ
along the length of the plate, so that both q” and h should also vary accordingly.

The total heat transfer rate q for the entire surface may be expressed as

q   qdA s (3.2).
As

From (3.1), it can in turn be expressed as

q  Ts  T   hdA s (3.3).


As

It is convenient to work with an average convection coefficient h for the entire surface,
such that the total heat transfer rate q can be expressed as

q  hAs Ts  T  (3.4),

Prof. Hui Tong Chua 60


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
from which we deduce that the average and local convection coefficients are related as
follow

1
h
As  hdA
As
s (3.5).

For our scenario of a flat plate with a constant width, (3.5) reduces to

L
1
h   hdx (3.6).
L0

The issue of quantifying the local and average convection coefficients constitutes the
central problem in convective heat transfer. This is by all means not a trivial problem as
the convection coefficient is a function of a host of fluid properties, surface geometry and
flow conditions.

Example 3.1 (Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer,


2002, John Wiley & Sons)

It is found that for flow over a very rough flat plate, the local heat transfer coefficient hx
is found to follow the relation below

h x  ax 0.1

where a is a coefficient (W/m1.9K) and x the distance (m) from the leading edge of the
plate.

u, T Boundary layer


T hx = ax-0.1

Ts

Determine the qualitative trend of hx as a function of x.

Prof. Hui Tong Chua 61


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
1.5

1
hx/a (W/m2.K)

hx = ax-0.1

0.5

0
0 1 2 3 4
x (m)

Determine the relation between the average heat transfer coefficient and the local heat
transfer coefficient.

x
1
x 0
hx  h x dx

x
a
  x 0.1dx
x0
a x 0.9

x 0.9
 1.11ax 0.1 or 1.11h x

This trend is featured below.

Prof. Hui Tong Chua 62


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
1.5

hx,average = 1.11ax-0.1 or 1.11hx

1
hx/a or hx,average/a (W/m2.K)

hx = ax-0.1

0.5

0
0 1 2 3 4
x (m)

End of example

3.2 The convection boundary layers

3.2.1 The velocity boundary layer

The concept of a velocity boundary layer is best understood by considering the flow over
a flat plate.

u Free stream u
(x)

y 

Velocity boundary layer


x
Fig. 3.2 The development of velocity boundary layer on a flat
plate.

When fluid particles engage with the surface of the plate, they are stopped by the plate so
that there is no relative velocity between the plate and the particles. These particles then

Prof. Hui Tong Chua 63


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
slow down those particles in the adjoining layer, which in turn decelerate those in the
next layer and so on. This retardation effect continues until at a vertical distance y = , it
becomes negligible. This deceleration mechanism is associated with the shear stress 
acting in layer parallel to the fluid velocity as suggested in figure 3.2.

The quantity  is referred to as the velocity boundary layer thickness and typically
u
defined as the y/L value for which  0.99 , where L is the length of the plate, in the
u
direction of the fluid flow. y/L or y* is the dimensionless height, and u/u or u* is the
dimensionless horizontal component of the fluid velocity. Non-dimensionalizing the
height and horizontal component of the fluid velocity are convenient as their values are
both bounded by 0 and 1. The boundary layer velocity profile refers to the way in which
u* increases with y* until it eventually approaches 1, which is then equivalent to the free
stream fluid velocity.

Accordingly within the velocity boundary layer, there are appreciable velocity gradients
and shear stresses; beyond this layer, those two quantities are negligible. The importance
of this velocity boundary layer cannot be overemphasized as it helps engineers relate to
the surface shear stress s and therefore surface frictional effect on the plate.

For a Newtonian fluid, the surface shear stress s can be related to the velocity profile by

u
s   (3.7)
y y0

where  is the dynamic viscosity of the fluid.

For a flow over an external surface or an external flow, s is in turn related to the
dimensionless local friction coefficient Cf as

s
Cf  (3.8)
u 2 2

which is useful for the determination of surface frictional drag.

3.2.2 The thermal boundary layer

Prof. Hui Tong Chua 64


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
T u T
t(x)

y t

Ts
x
Thermal boundary layer
Fig. 3.3 Thermal boundary layer on a flat plate.

Consider a fluid flow over a flat plate maintained at a lower temperature, T, than that of
the plate, Ts. With such a temperature difference, a thermal boundary layer must develop
as shown in figure 3.3. At the leading edge, the fluid temperature profile is uniform with
T(y) = T. Those fluid particles that come into contact with the plate are at thermal
equilibrium with the plate. These in turn heat up the fluid particles in the adjoining fluid
layer, and thus giving rise to a temperature gradient in the fluid. The layer in which such
a gradient exists is termed the thermal boundary layer. Its thickness t is defined as the
T T Ts  T
value of y/L or y* at which s  0.99 . or T* is termed the non-
Ts  T Ts  T
dimensional temperature.

Let us now establish the relationship between the thermal boundary layer and the
convective heat transfer coefficient.

At any distance x from the leading edge, the local heat flux from the plate to the fluid can
be determined by applying the Fourier’s law as

T
qs  k f (3.9).
y y0

The mode of heat transfer from the plate surface to the adjoining fluid is purely
conductive as there is no fluid motion at the interface.

Combining (3.9) with the Newton’s law of cooling or q  hTs  T  , we conclude that

 k f T y y0
h (3.10).
Ts  T

Prof. Hui Tong Chua 65


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
3.2.3 The concentration boundary layer

Mixture
of A+B
CA, u CA,
c(x)

y c

CA,s
x
Concentration boundary layer
Fig. 3.4 Concentration boundary layer on a flat plate.

Consider a binary mixture of chemical species A and B flowing over a flat surface as in
figure 3.4. The molar concentration (mol/m3) of species A at the surface is CA,s, and in
the free stream, it is CA,. If CA,s is higher than CA,, then species A will be transported
by convection from the surface to the free stream. For example, if the flat plate contains
a water film, water (species A) can evaporate into the air stream (species B) on top. As
another example, the flat surface can be made of dry ice so that carbon dioxide (species A)
sublimes into the air stream (species B) on top. In such situations, a concentration
boundary layer will develop and within which a concentration gradient will exist. The
thickness of the boundary layer, c, is defined as the point where
CA,s  CA   0.99 .
CA,s  CA, 
CA,s  CA  or C* is the dimensionless species A concentration.
CA,s  CA,  A

We are keen to determine the rate of transfer of a species between a surface and the free
stream. Specifically we ask question about the molar flux of species A,
 
NA kmol / s  m 2 . When the molar flux of species transfer is driven by diffusion, it can
be determined by the Fick’s law:

CA
NA  D AB (3.11)
y

where DAB is the binary diffusion coefficient. Within the concentration boundary layer
where y > 0, species transfer is effected by both bulk fluid motion (advection) and
diffusion. Hence (3.11) or the Fick’s law does not describe the whole picture. However,
at y = 0 or the surface, there is essentially no fluid motion and species transfer is
dominated by diffusion. This allows us to apply the Fick’s law at y = 0 to fully describe
the molar flux of species transfer so that

Prof. Hui Tong Chua 66


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
C
NA ,s  D AB A (3.12)
y y 0

where the subscript s emphasizes that we are tracking the molar flux of A at the surface.

Following the Newton’s law of cooling, we can relate the molar flux of A to the
concentration difference between the surface and the free stream by

NA ,s  h m CA,s  CA,  (3.13)

where hm (m/s) denotes the convection mass transfer coefficient. Comparing (3.12) and
(3.13) elicits the following conclusion:

 D AB C A y y 0
hm  (3.14).
C A ,s  C A , 

Example 3.2 (Adapted from Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat
and Mass Transfer, 6th ed., 2007)

A long cylindrical naphthalene rod with a diameter of 20 mm is exposed to an air stream


with an average convection mass transfer coefficient h m  0.05m / s . The molar
concentration of naphthalene vapor at the cylinder surface is 510-6 kmol/m3, and its
molecular weight is 128 kg/kmol. Determine the mass sublimation rate per unit length of
the rod.

CA, = 0
h m  0.05m / s

Air

CA,s = 510-6 kmol/m3


D = 20mm

From (3.13), the molar flux of naphthalene vapor from the cylinder is

NA ,s  h m CA,s  CA, ,

Prof. Hui Tong Chua 67


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
so that the molar transport rate of naphthalene vapor per unit length from the rod is

NA
 h m DCA,s  CA ,  .
L

Hence,

NA
L
 
 0.05  20 103 5 106  0  1.57 108 kmol s  m .

The mass sublimation rate per unit length can therefore be deduced to be

n A MA NA
  128 1.57 108 kmol s  m  2.01106 kg s  m .
L L

End of example

3.3 Laminar and turbulent flows

An essential first step in tackling convection heat and mass transfer problem is to identify
whether the boundary layer is laminar or turbulent.

Streamline u
y, v
v

u Turbulent
u x, u u region

Buffer layer
y
Laminar sublayer

Laminar Transition Turbulent


x

Fig. 3.4 The development of velocity boundary layer on a flat plate.

There are sharp differences between laminar and turbulent flow conditions. In the
laminar boundary layer, the fluid motion is highly ordered and we can identify
streamlines along which particles move.

Prof. Hui Tong Chua 68


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
On the other hand, fluid motion in the turbulent layer is highly irregular and is typified by
velocity fluctuations. These fluctuations facilitate momentum and energy transfers which
in turn increase surface friction as well as convection heat transfer rate. Consequently,
the thickness of the turbulent boundary layer is larger and the temperature and velocity
boundary layer profiles are flatter than in laminar flow. In the turbulent boundary layer,
there are three distinct regions. Firstly, there is a laminar sublayer in which transport is
dominated by diffusion and the velocity profile is practically linear. Secondly, there is an
adjoining buffer layer wherein diffusion and turbulent mixing are comparable. Thirdly,
there is a turbulent zone in which transport is governed by turbulent mixing.

Figure 3.5 below illustrates the variation of velocity boundary layer thickness  and the
local heat transfer coefficient h for flow over an isothermal plate.

h, 
h(x)

u, T (x)

Ts
y

Laminar Transition Turbulent


x
xc
Fig. 3.5 The profiles of the velocity boundary layer thickness  and the
local heat transfer coefficient h for flow over an isothermal flat plate.

A dimensionless number called the Reynolds number is used to determine the flow
regimes:

Re x  u  x  (3.15)

A critical Reynolds number exists when transition from laminar to turbulent flows begins.
For a flat plate and depending on surface roughness and the turbulence level of the free
stream, this number varies from approximately 105 to 3106.

u  x c
Re x ,c   5  105 (3.16)

is often assumed for the occurrence of a laminar-to-turbulent transition.

Prof. Hui Tong Chua 69


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
3.4 Physical significance of the dimensionless parameters

Physically the Reynolds number Re signifies the ratio of inertia to viscous forces in the
velocity boundary layer. The inertia forces are associated with an increase in the
momentum flux of fluid moving through a control volume. It can be shown that the
inertia forces increases much faster with the fluid velocity than the viscous forces.

Consequently we expect that the inertia forces to dominate for large values of Re and
viscous forces to dominate for small Re.

There are two more very important dimensionless numbers specifically related to the
transport properties of the fluid. They are the Prandtl number and the Schmidt number.

The Prandtl number Pr is defined as the ratio of the momentum diffusivity (or kinematic
viscosity)  to the thermal diffusivity  so that


Pr  (3.17).

 k
Since   and   , we have
 c p

c p
Pr  (3.18)
k

as well. The Prandtl number measures the relative effectiveness of momentum and
energy transport by diffusion in the velocity and thermal boundary layers, respectively.

Prof. Hui Tong Chua 70


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
Fig. 3.6 Prandtl number of some typical vapors

1.1

1.0

0.9 Air
Prandtl number

Carbon dioxide
Helium
Oxygen
Water vapor (Steam)
0.8

0.7

0.6
0 100 200 300 400 500 600 700 800 900 1000
Temperature (K)

Figure 3.6 indicates that the Prandtl number of gases is in the order of unity. This implies
that the magnitudes of energy and momentum transfer by diffusion are comparable. As
for liquid metals, referring to figure 3.7, Pr << 1 and the energy diffusion rates greatly
exceeds the momentum diffusion rate. This implies that the thermal boundary layer is
very thick compared with that of the velocity boundary layer particularly in the laminar
flow regime. In contrast, for oil and some chemicals and referring to figure 3.8, Pr >> 1
signifying that, especially within the laminar flow regime, the thermal boundary layer is
very thin compared with that of the velocity boundary layer.

Prof. Hui Tong Chua 71


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
Fig. 3.7 Prandtl numbers of typical liquid metals

Bismuth
Lead
Potassium
Sodium
NaK (45%/55%)
0.1 NaK (22%/78%)
Mercury
Prandtl number

PbBi (44.5%/55.5%)

0.01

0.001
200 300 400 500 600 700 800 900 1000 1100
Temperature (K)

Fig. 3.8 Prandtl numbers for some oil and chemicals

100000

10000

Engine oil (unused)


Ethylene Glycol [C2H4(OH)2]
Glycerin [C3H5(OH)3]
1000
Prandtl number

100

10

1
250 270 290 310 330 350 370 390 410 430 450
Temperature (K)


Similarly the Schmidt number defined by Sc  offers a measure of the relative
D AB
effectiveness of momentum and mass transport by diffusion in the velocity and
concentration boundary layers, respectively.

Prof. Hui Tong Chua 72


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.

3.5 The boundary layer approximations

In most cases, boundary layer thicknesses are typically very small. As such, the
following boundary layer approximations are known to apply.

Velocity boundary layer:

u  v
u u v v
 , ,
y x y x

Thermal boundary layer:

T T

y x

In short, the velocity component along the surface is much larger than that normal to the
surface, and gradients normal to the surface are much bigger than those along the surface.
It follows that the single most important shear stress in the velocity boundary layer is

u
   xy   yx (3.19)
y

yx

y
xy xy

x
yx

In addition, the rate of thermal conduction in the y direction is much larger than in the x
direction in the thermal boundary layer.

3.6 Boundary layer similarity

It turns out that given the aforementioned boundary layer approximations and especially
when the pressure gradient is negligible, which is exactly the case as depicted in figures
(3.2-4), the behaviors of the velocity, thermal and concentration boundary layers are

Prof. Hui Tong Chua 73


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
startlingly similar. These similar behaviors are germane to low-speed, forced-convection
flows which are prevalent in many engineering applications.

In particular, mathematically it can be shown that if the pressure gradient is zero or dp/dx
= 0, and if Pr = Sc = 1, then the three boundary layers, in their dimensionless forms, are
exactly equivalent. In the literature the pressure is commonly nondimensionalize as

p
p*  (3.20),
u 2

so that the condition of dp/dx = 0 is often quoted as dp*/dx* = 0, where x* = x/L and x is
the measure of distance along the direction of fluid flow.

Consequently, we can make the important statement that the boundary layer velocity,
temperature and species concentration profiles must be of the same functional form.

Based on careful observation, both experimentally and mathematically, we expect that


the dimensionless fluid velocity must depend on the location along the surface, viz. x*
and y*, flow regime, i.e., laminar or turbulent, which is governed by the Reynolds
number, and the surface profile which in turn determines the pressure gradient or dp*/dx*.
We can summarize this as

 * * dp* 
u  f  x , y , Re L , * 
*
(3.21)
 dx 

Now the shear stress at the surface, where y* = 0, can be expressed as

u  u  u
*
s      * ,
y y 0  L  y y*  0

the corresponding dimensionless friction coefficient is then given by

s 2 u *
Cf   (3.22).
u 2 2 Re L y* y*  0

From (3.21), we deduce that

u *  dp* 
 f  x * , Re L , *  .
y* y*  0  dx 

Thence, for a prescribed profile where dp*/dx* is known, (3.22) can be expressed as

Prof. Hui Tong Chua 74


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
f x * , Re L 
2
Cf  (3.23).
Re L

(3.23) is a very important conclusion made from simple observations. It points out that
for a prescribed profile, the function that relates Cf to x* and ReL is universally
applicable, independent of the types of fluid and over a wide range of u and L.

Let us now apply this powerful technique to the convection heat transfer coefficient.
Based on careful observations, both experimentally and mathematically, we know that
the dimensionless temperature must depend on the location over the surface, i.e. x* and
y*, the flow regime or the Reynolds number, the type of fluid, which is governed by the
Prandtl number, and the surface profile which dictates dp*/dx*, so that

 dp* 
T*  f  x * , y* , Re L , Pr, *  (3.24).
 dx 

From (3.10),

 k f T y y0 k f T*
h 
Ts  T L y* y*  0

which leads to the definition of an important dimensionless number – the Nusselt number,
Nu

hL T*
Nu   (3.25).
k f y* y*  0

From (3.24), it is clear that for a prescribed geometry,


Nu = f(x*, ReL, Pr) (3.26).

Inspect the similarity between the functional dependence of the Nusselt number and that
of the friction coefficient. Nu is to the thermal boundary layer what the friction
coefficient is to the velocity boundary layer. We can now make the powerful statement
that, for a given geometry, and for different fluids and a wide range of u and L, a
universal function exists between Nu, x*, ReL and Pr.

This implies that when Nu is known, we can then compute for h which in turn enables us
to quantify the local surface heat flux.

From (3.26), we can make yet another powerful statement about the average heat transfer
coefficient which is obtained by integrating with respect to x *. In which case, it must be
independent of x* and hence,

Prof. Hui Tong Chua 75


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
hL
Nu   f Re L , Pr  (3.27)
kf

Turning now to the mass transfer phenomenon over a liquid surface or a sublimating
solid surface, we can similarly elicit the following conclusion, that

 dp* 
C*A  f  x * , y* , Re L , Sc , *  (3.28).
 dx 

The mass transfer coefficient expression from (3.14) can be written as

 D AB C A y y 0 D AB C*A
hm   .
C A ,s  C A ,  L y* *
y 0

This leads us to a dimensionless number known as the Sherwood number, Sh

h m L C*A
Sh   (3.29).
D AB y* *
y 0

From (3.28) and for a prescribed geometry, we deduce that

Sh  f x* , Re L , Sc  (3.30).

Following a similar argument, we also conclude that the average Sherwood number has
the following functional form:

Sh  f Re L , Sc  (3.31).

Example 3.3 (Adapted from Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat
and Mass Transfer, 6th ed., 2007)

A solid is completely exposed to atmospheric air with a free stream temperature and
velocity of 20°C and 100m/s, respectively. The solid exhibits a surface temperature of
80°C and possesses a characteristic length of 1m. Under these prevailing conditions, the
heat flux at a particular point (call the dimensionless location x*) on this solid’s surface is
104 W/m2 while the temperature in the boundary layer above this particular point (this
dimensionless location will be [x*, y*]) is measured to be 60°C. A mass transfer process
is to be performed on a larger but similarly shaped solid with a characteristic length of
2m. Specifically, water from this solid’s surface is to be evaporated into dry atmospheric
air with a free stream velocity of 50m/s. Both the solid and the air are at 50°C.

Prof. Hui Tong Chua 76


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
Determine the molar concentration and the species molar flux of the water vapor at a
location (x*, y*) over this larger solid which corresponds to same dimensionless location
of the smaller solid at which the heat flux and temperature measurements were performed.

CA(x*,y*)
T(x*,y*) = 60°C T = 50°C
u = 50m/s
T = 20°C
u = 100m/s
p = 1 atm  
NA x *
p = 1 atm  
qs x *  10 4 W m 2 CA, = 0

Ts = 80°C Water film (A)


L = 1m L = 2m Ts = 50°C

Heat transfer Mass transfer

We can advantageously invoke the heat and mass transfer analogy to determine the
surface mass flux and the species concentration at the desired locations.

We have learnt that

T  Ts  dp* 
T*   f  x * , y* , Re L , Pr, *  , and
T  Ts  dx 
C A  C A ,s  * * dp* 
CA 
*
 f  x , y , Re L , Sc , * 
C A ,   C A ,s  dx 

For the heat transfer case,

u , HT L HT 100 1
Re L, HT    5.5 10 6 , Pr = 0.70
 18.2 10 6

For the mass transfer case,

u ,MT L MT 50  2
Re L,MT    5.5  106 , and
 18.2  10 6
 18.2  10 6
Sc    0.70
D AB 0.26  10  4

Noting that the shapes of the solids are identical, and ReL,HT = ReL,MT, Pr = Sc,
x*HT  x *MT , y*HT  y*MT , we deduce that

Prof. Hui Tong Chua 77


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
T  Ts  dp*  C  C A ,s
T*   f  x * , y* , Re L , Pr, *   C*A  A .
T  Ts  dx  C A ,   C A ,s

Thence,

 
C A x * , y*  C A ,s

 
T x * , y*  Ts 60  80
  0.33
C A ,   C A ,s T  Ts 20  80

 A,sat 0.082
Now C A,s  C A,sat 50C    0.0046 kmol m3 , and
MA 18
CA, = 0. Hence,

 
C A x * , y*  0.33C A ,s  C A ,s
 0.00461  0.33
 0.0031 kmol m 3

The molar flux can be calculated from

 
NA ,s x*  h m CA,s  CA,  .

Once again, we can apply the heat and mass transfer analogy by noting that

Sh 
h mL
D AB
 
 f x * , Re L , Sc , and Nu =
hL
kf
= f(x*, ReL, Pr) as well as realizing that

q
h .
Ts  T 

Since x*HT  x *MT , ReL,HT = ReL,MT, Pr = Sc, the governing functions for both Sh and Nu
are identical and hence,

h m L MT hL HT
 .
D AB kf

Whence

Prof. Hui Tong Chua 78


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
L D
h m  h HT AB
L MT k f
q L HT D AB

Ts  T  L MT k f
10 4 1 0.26 10  4

80  20 2 0.028
 0.077 m s

Therefore,

NA x *   0.0770.0046  0  3.54 104 kmol s  m2

3.7 Evaporative cooling

Evaporative cooling is an important heat and mass transfer phenomenon. This happens
when liquid evaporates into a flowing air stream which in turn cools the liquid.

Gas flow
(species B)

q conv
 q evap

Liquid layer q add



(species A)

Figure 3.9 Latent and sensible heat exchange between a liquid layer
and the vapor blanket.

Evaporation occurs from the liquid surface; the latent heat of vaporization carried away
by the vaporized liquid molecules provides the cooling to the liquid layer. In order to
maintain a constant temperature in the liquid layer, energy must therefore be added to
replenish the energy lost through vaporization. The latter energy may be realized by
having a heated plate underneath the liquid layer or by electric heating. Neglecting
radiation, an energy balance can be written for this system as depicted in figure 3.9 as

qconv
  qadd
  qevap
 (3.32)

where the evaporative heat flux, qevap


 , can be expressed as

Prof. Hui Tong Chua 79


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.

qevap
  nA h fg (3.33).

Considering the situation where qadd


 is zero such as in the case of cooling towers, and
employing (3.1) and (3.13), we have for (3.32)

h T  Ts   h fg h m CA ,sat Ts   CA , M A


(3.34),
 h fg h m A ,sat Ts   A , 

where MA is the molecular weight of species A. Note that the vapor concentration (in
kmol/m3 or kg/m3) at the liquid surface corresponds to the saturated conditions at the
liquid surface temperature, Ts. Working further from (3.34), we have

T  Ts  h fg  m A,sat Ts   A , 
h 
(3.35).
 h 

Let us try to obtain a generic relationship for hm/h… Referring to (3.26) and (3.30) we
have

Nu = f(x*, ReL, Pr) Sh  f x* , Re L , Sc  .

From extensive experimental observations, it turns out that Nu and Sh are typically
proportional to Prn and Scn, respectively where n typically assumes a value of 1/3. Hence
we have

Nu = f(x*, ReL)Prn Sh  f x* , Re L Sc n .

Since Nu and Sh shares an equivalent function, f(x*, ReL), we conclude that

Nu Sh
 , so that we have
Pr n Sc n

hL k h m L DAB

Pr n Sc n

n
h k  Pr  k
    (3.36a)
h m DAB  Sc  DABLe n

Sc   
where Le    . (3.36a) could alternatively be expressed as
Pr DAB  DAB

Prof. Hui Tong Chua 80


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.

h k
  cp Le1 n (3.36b).
h m DABLe n

Applying this to (3.35), and applying the ideal gas law to the vapor of species A (i.e.
pMA
A  ), we have
RT

h fg M A h fg  pA ,sat Ts  pA , 
T  Ts   T     
A , sat s A,    (3.37).
cp Le 23
Rcp Le 23
 Ts T 

In (3.37) the species B properties, to wit , cp and Le are to be evaluated at the mean
temperature of the thermal boundary layer or Tm = (Ts + T)/2. In (3.37), the value of n
is 1/3.

Example 3.4 (Adapted from Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat
and Mass Transfer, 6th ed., 2007)

In hot arid regions, a simple method to keep a container of beverage cool is to wrap it
with a fabric which is continually moistened by a volatile liquid. Consider a situation
where this wrapped container is exposed to dry ambient air at 40°C and there is forced
convection around this container. The volatile liquid used has a molecular weight of 200
kg/kmol and a latent heat of vaporization of 100 kJ/kg. The saturated vapor pressure of
the liquid at the surface temperature of the container is 5000 Pa, and the diffusion
coefficient of the vapor in air is 0.210-4 m2/s. Determine the steady-state temperature of
the beverage.

Ts
Air (B)
T = 40°C
Volatile liquid (A)
pA, = 0 hfg = 100 kJ/kg
MA = 200 kg/kmol
PA,sat (Ts) = 5000 N/m2
q conv
 DAB = 0.210-4 m2/s

q evap

The evaporative cooling effect may be quantified with

M A h fg  p A ,sat Ts  p A , 
T  Ts    .
Rc p Le 23
 Ts T 

Prof. Hui Tong Chua 81


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.


We know that pA, = 0 and Le  . We assume that the mean boundary layer
D AB
22.5 106
temperature is 300 K so that Air = 22.510-6 m2/s. Hence, Le  . Arranging
0.2 104
the abovementioned expression, we have

M A h fg p A ,sat
Ts2  T Ts  0
Rc p Le 2 3
M A h fg p A ,sat


200 100  5000 10 3   9518K 2
Rc p Le 23
 22.5 10 6

23

8.315 1.16 1.007   4



 0.2 10 

Solving for Ts we have

T  T2  4  9518 313  3132  4  9518


Ts   .
2 2

The minus sign must be nonphysical and rejected (Ts must equal T if there is no
M h p
evaporation which corresponds to pA,sat = 0 and therefore A fg A2,sat  0 ), and hence
Rc p Le 3

Ts = 278.9K = 5.9°C.

3.8 Analogy between momentum and heat transfers

As mentioned earlier, the velocity, thermal and species concentration boundary layer
equations are analogous when dp*/dx* =0, Pr  1, and Sc  1. This further implies that
the general functions in (3.23), (3.26) and (3.30), for respectively the friction coefficient,
Nusselt number, and Sherwood number are interchangeable. We therefore arrive at the
conclusion that

Cf
Re L
2
 Nu  Sh  f x * , Re L   (3.38).

Replacing Nu by the Stanton number St and Sh by the mass transfer Stanton number Stm,

h Nu hm Sh
St   St m   (3.39),
u  c p Re Pr u  Re Sc

Prof. Hui Tong Chua 82


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons and Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat and Mass
Transfer, 6th ed., 2007, John Wiley & Sons.
we arrive at

Cf
 St  St m (3.40).
2

(3.40) is known as the Reynolds analogy which quantifies the similarity between the
velocity, thermal and species concentration boundary layers. Hence if the velocity
parameter is known, the analogy may be used to obtain the heat and mass transfer
parameter, and vice versa.

The Reynolds analogy is restricted by requiring dp*/dx* =0 and Pr  1. However, a


Prandtl number correction can be added so that it applies to a wide range of Pr:

Cf
 St Pr 2 3  jH 0.6  Pr  60 (3.41),
2

Cf
 St mSc 2 3  jm 0.6  Sc  3000 (3.42).
2

These are known as the Chilton-Colburn, or the modified Reynolds, analogies while jH
and jm are known as the Colburn j factors for heat and mass transfer, respectively. For
laminar flow, (3.41-2) are only appropriate when dp*/dx*  0; but in turbulent flow, the
conditions are less sensitive to pressure gradients and are approximately valid.
Obviously, if the analogy is applicable to every point on a surface, then it is also valid for
the surface average coefficients.

Prof. Hui Tong Chua 83


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

Chapter 4 External flow

4.1 The flat plate in parallel flow

u, T

Ts
y

Laminar Turbulent
x
Fig. 5.1 The flat plate in parallel flow.

4.1.1 Laminar flow

In the laminar flow regime, it can be shown that the local friction coefficient is given by

s, x
Cf , x   0.664 Rex1 2 (4.1).
u 2
 2

Now by numerically solving the energy-balance equation for the thermal boundary layer,
it is found that the local Nusselt number can be correlated as

hxx
Nu x   0.332 Re1x 2 Pr1 3 Pr  0.6 (4.2).
k

We can then employ the heat and mass transfer analogy to obtain the local Sherwood
number correlation for mass transfer by inspecting (4.2) so that

h m,x x
Sh x   0.332 Re1x 2 Sc 1 3 Sc  0.6 (4.3)
D AB

We can now derive the corresponding averaged parameters. For the average friction
coefficient, it is defined as

Prof. Hui Tong Chua 84


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
s, x
Cf , x  (4.4)
u 2 2

x
1
x 0
where s, x  s, x dx .

Now s , x can be obtained from (4.1) and the ensuing integration yields

Cf , x  1.328 Rex1 2 (4.5).

In addition, the average heat transfer coefficient for laminar flow is

x 12x
1 k u  1
h x   h x dx  0.332  Pr1 3    x dx
  
12
x0 x 0

k
 0.664  Re1x 2 Pr1 3
x
 2h x

Hence,

hxx
Nu x   0.664 Re1x 2 Pr1 3 Pr  0.6 (4.6).
k

Invoking the heat and mass transfer analogy, we also have

h m,x x
Sh x   0.664 Re1x 2 Sc 1 3 Sc  0.6 (4.7).
D AB

It is imperative to note that all the fluid properties should be evaluated at a mean
boundary layer temperature Tf, known as the film temperature

Ts  T
Tf  (4.8).
2

(4.6) can clearly be used for various vapors, engine oil and certain chemicals but is
obviously unsuitable for liquid metals. Despite the corrosive and reactive nature of liquid
metals, their low vapor pressure, low melting point, high thermal capacity and
conductivity render them attractive coolants in situations requiring high heat transfer
rates.

Prof. Hui Tong Chua 85


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
For laminar flow over an isothermal plate, and for all Prandtl numbers, the following
correlation applies

0.3387 Re1x 2 Pr1 3


Nu x  Pex  Re x Pr  100 (4.9).
1  0.0468 Pr 
23 14

Pe is known as the Peclet number and can alternatively be expressed as u  L  . Note


that Nux  2Nux .

4.1.2 Turbulent flow

It is known that, for turbulent flows with Reynolds number up to approximately 107, the
local friction coefficient is adequately correlated by

Cf ,x  0.0592 Rex1 5 Re x  107 (4.10).

The same expression can be used for values of Rex up to 108 to within 15% accuracy.

For turbulent flow, the boundary layer development is dominated by random fluctuations
in the fluid and not by molecular diffusion. Hence the relation between thermal, velocity
and species concentration boundary layer growths does not depend on the value of Pr and
implies that the three local boundary layer thicknesses are approximately the same. This
fully justifies the use of the Chilton-Colburn analogies (3.51-52) and results in

1
Nu x  St Re x Pr  Cf Re x Pr1 3  0.0296 Re 4x 5 Pr1 3 0.6  Pr  60 (4.11),
2
and

Sh x  St m Re x Sc  0.0296 Re4x 5 Sc1 3 0.6  Sc  3000 (4.12),

It is useful to note that turbulent boundary layer grows more rapidly than the laminar
boundary layer and is typified by larger friction and convection coefficients.

4.1.3 Mixed boundary layer conditions

For laminar flow over the entire plate, equations in section 4.1.1 can be used with
confidence. An additional practical point is that if the transition into turbulence is very
much delayed such that it occurs in the range of 0.95  x c L  1, the same equations can
reasonably be used to compute those average coefficients.

Prof. Hui Tong Chua 86


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
If the transition occurs any earlier, then the determination of the average coefficients
must consider the effects of both the laminar and turbulent regimes. In which case, the
average convection coefficient can be computed as

1  c 
x L
hL   h dx   h turbulentdx  ,
L  0 
la min ar
xc 

so that invoking (4.2 and 4.11), we have

k dx  1 3
1 2 xc 45 L
u  dx u 
h L   0.332   0 x1 2  0.0296    x 4 5  Pr ,
L    xc 

Integrating, we obtain


Nu L  0.037 Re 4L 5  0.037 Re4x ,5c  0.664 Re1x ,2c Pr1 3  0.6  Pr  60, Re x ,c  Re L  108 (4.13)

where Rex,c is the critical Reynolds number. Adopting a critical Reynolds number of
5105, we have

 
Nu L  0.037 Re 4L 5  871 Pr1 3 0.6  Pr  60, Re x ,c  Re L  108 , Re x ,c  5 105 (4.14)

Applying the heat and mass transfer analogy to (4.13), we have


Sh L  0.037 Re 4L 5  871 Sc1 3  0.6  Sc  60, Re x ,c  Re L  108 , Re x ,c  5 105 (4.15)

It can similarly be shown that, for the friction coefficient,

0.074 1742
Cf ,L  15
 5 105  Re L  108 , Re x ,c  5 105 (4.16)
Re L Re L

If the turbulent boundary layer exists all over the plate which can be achieved by using a
fine wire or other turbulent promoter at the leading edge, then the following correlations
apply

Nu L  0.037 Re 4L 5 Pr1 3 (4.17)


1 5
Cf ,L  0.074 Re L (4.18)

It is important to note that the fluid properties must be evaluated at the film temperature.

Example 4.1

Prof. Hui Tong Chua 87


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
A flat plate of length 0.5m is being swept by air at a pressure of 6000 Pa, a temperature
of 300°C and a velocity of 10m/s. Estimate the cooling rate per unit width of the plate in
order to maintain its surface temperature at 27°C.

10m/s, 300°C,
6 kPa

27°C

L = 0.5m
x

The film temperature is 437K ( [300+27]/2). For air at 437K and 1atm,  = 30.8410-6
m2/s, k = 36.410-3 W/mK, Pr = 0.687. k, Pr, and  can be reasonably assumed to be
independent of pressure, but not     as it is a function of density which in turn is
sensitive to pressure.

 2 p1
From the ideal gas law,   p RT , we deduce that  , and hence at 437K and 6
1 p 2
kPa, we have

101330
  30.84 106   5.21104 m 2 s
6000

For a plate of unit width, the cooling rate needed will have to be given by

q  hLT  Ts  .
This implies that we have to determine the average heat transfer coefficient. The
Reynolds number is calculated as

uL 10  0.5
Re L   4
 9597  5 105 .
 5.2110

Hence the flow is laminar over the entire plate. Accordingly, we use

Nu x  0.664 Re1x 2 Pr1 3 Pr  0.6 , and

Nu L k 0.6649597  0.6871 3 36.4 103  4.18 W


12
h  m2  K .
L 0.5

The required cooling rate per unit width is then given by

Prof. Hui Tong Chua 88


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

q  4.18  0.5300  27  570 W m

End of example

Example 4.2

A flat plate is maintained at a uniform surface temperature of 230°C by using


independently controlled electrical strip heaters. Each heater is 50mm long. If
atmospheric air at 25°C flows at 60m/s over the plate, which heater will receive the
maximum electrical input? What is this electrical input?

60m/s, 25°C

Plate 1 230°C Plate 5 Plate 6

50mm

L6 = 300mm
x

For Tf = 400 K, and p = 1 atm,  = 26.4110-6 m2/s, k = 0.0338 W/mK, Pr = 0.690.

We know that plate 1 experiences the largest local laminar convection coefficient, while
the plate located at the laminar-to-turbulent transition point receives the largest local
turbulent convection coefficient. Hence, the first task will be to identify the heater plate
at which the flow transition occurs.

At the point of transition, Rex,c = 5105, in which case,

 26.41106
xc  Re x ,c  5 105  0.22m .
u 60

Hence, we know that the transition occurs at the fifth plate. The heater which receives
the maximum electrical input must experience the maximum average convection
coefficient. Three possibilities thus become apparent:

1. Heater 1, as it experiences the maximum local laminar convection coefficient.


2. Heater 5, as it experiences the maximum local turbulent convection coefficient.
3. Heater 6, as the entire flow regime over its surface is turbulent.

Prof. Hui Tong Chua 89


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

For the first heater, qconv,1  h1L1 Ts  T  .

u  L1 60  50 103
Now, Nu1  0.664 Re11 2 Pr1 3 , while Re1   6
 1.14 105 .
 26.4110

Hence,

h1   

Nu1k 0.664 Re11 2 Pr1 3 k 0.664 1.14  105 0.691 3  0.0338 
12

 134 W m 2  K ,
L1 L1 0.05

and

qconv,1  134  0.05230  25  1370 W m .

The power requirement of heater 5 can be obtained by subtracting the total power
requirement of the first 4 heaters from the total power requirement of all the first five
heaters, so that

qconv,5  h15 L5 Ts  T   h14 L 4 Ts  T  ,

where

h1 4   
 
12
Nu 4 k 0.664 Re14 2 Pr1 3 k 0.664 4  1.14  105 0.691 3  0.0338
 67 W m 2  K .
L4 L4 4  0.05

Heater 5 sees a transition from laminar to turbulent flow, so that

 
Nu5  0.037 Re54 5  871 Pr1 3 .
Hence,

h15  
  
  
45

Nu 5 k 0.037 Re 54 5  871 Pr1 3 k 0.037 5  1.14  105  871 0.691 3  0.0338
L5 L5 5  0.05
 74 W m 2  K

The rate of heat transfer from heater 5 is therefore given by

qconv,5  74  0.25230  25  67  0.2230  25  1050 W m .

The rate of heat transfer from heater 6 is similarly expressed as

qconv,6  h16 L6 Ts  T   h15 L5 Ts  T 

Prof. Hui Tong Chua 90


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

h16  
  
  
45
Nu 6 k 0.037 Re 64 5  871 Pr1 3 k 0.037 6 1.14 10 5  871 0.691 3  0.0338 
L6 L6 6  0.05
 85 W m 2  K

Hence,

qconv,6  85  0.3230  25  74  0.25230  25  1440 W m .

We can therefore conclude that the sixth plate has the largest power consumption.

Convection coefficients along the plate

300

250

Turbulent convection
coefficient
200
h (W/m2.K)

150

100
Laminar convection
coefficient

50

xc
0
0.00 0.05 0.10 0.15 0.20 0.25 0.30
x (m)

End of example

Example 4.3 (Adapted from Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat
and Mass Transfer, 6th ed., 2007, John Wiley & Sons)

Residential swimming pools are a common sight. Since clean water is a precious
resource, we are interested in estimating the daily water loss due to pool evaporation. We
shall apply representative conditions where both the water and ambient air temperatures
are at 25°C and the relative humidity in the air is 50%. Let’s take a typical pool
dimension of 6m by 12m. Wind blows at a velocity of 2m/s in the direction of the long
side of the pool. What is the water loss from the pool in kilograms per day?

Prof. Hui Tong Chua 91


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

Air
u = 2 m/s
T = 25°C
 = 0.5
6m Water
Ts = 25°C
A,s = A,sat(Ts)

12m

The air flow over the pool of water can be treated to be fully turbulent as the air flow is
tripped by the deck around the pool. In which case,

h mL
Sh L   0.037 Re 4L 5 Sc 1 3 .
D AB

Whence,

D AB 4 5 1 3
h m  0.037 Re L Sc .
L

At 25°C, for a water vapor-air mixture, DAB = 0.2610-4 m2/s,  of air at 25°C = 15.710-
6
m2/s, so that Sc = /DAB = 0.60. For saturated water vapor at 25°C, A,sat = 0.0226
kg/m3.

uL 2  12
Re L    1.53  106 .
 15.7  10 6

Hence,

0.26 104
h m  0.037
12

1.53 106  0.601 3  5.99 103 m s .
45

The evaporation rate of the pool of water is then

n A  h m AA,s  A, .

The free-stream water vapor can be taken to be an ideal gas so that

 A ,
  .
 A ,sat T 

Prof. Hui Tong Chua 92


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

Therefore

n A  h m A A ,s   A ,sat Ts 


 h m A A ,sat Ts 1   
 5.99 10 3  6 12  0.0226  1  0.5
 4.87 10 3 kg s
 421 kg day

End of example

4.2 The cylinder in cross flow

Let us now examine the external flow across a cylinder as shown in figure 4.2.

Forward Wake
stagnation Boundary
point layer

Fig. 4.2 A cylinder in cross flow

p p
0 0
x x
Favorable pressure gradient Adverse pressure gradient

Separation point
Flow reversal
Fig 4.2 Velocity profiles at different pressure gradient regimes for flow
across a cylinder.

Prof. Hui Tong Chua 93


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
The free stream fluid stops at the forward stagnation point with an accompanying
increase in pressure. Within the flow regime in which the pressure gradient is negative or
favorable, the fluid velocity increases until the gradient turns zero. After which, the fluid
velocity slows down until one point where the velocity gradient at the cylinder surface
becomes zero. Beyond this point, the fluid near the surface lacks sufficient momentum to
surmount the adverse pressure gradient so that the boundary layer separates from the
cylinder surface resulting in a flow reversal near the surface and the formation of a wake.
This is characterized by flow irregularity and vortex formation.

The flow characteristics across a cylinder is governed by the Reynolds number defined as

u  D u  D
Re D   .
 

It governs the occurrence of a laminar to turbulent boundary layer transition across the
cylinder and dictates the position of flow separation as shown in figure 4.3.

u u
sep sep

Re D  2 105 Re D  2 105
Laminar Laminar Separation
boundary Separation
boundary
Transition
layer layer
Fig 4.3 Turbulence effect on separation
The fluid particles in a turbulent boundary layer are significantly more energetic than
those in a laminar boundary layer. Hence, referring to figure 4.3, we expect that when
the laminar boundary layer transform to a turbulent boundary layer, the flow separation
from the cylinder is delayed. This turns out to occur at a Reynolds number greater than
2105. When the Reynolds number is smaller, the boundary layer remains laminar and
separation happens at   80°. Conversely, when it is larger than the critical Reynolds
number, the separation occurs instead at   110°9. This delayed separation results in a
reduced wake region and consequently largely reduce the drag force experienced by the
cylinder.

The above boundary-layer characteristics profoundly affect the convection heat transfer
characteristics of flow across a cylinder. The increase in the Nusselt number with
increasing ReD is associated with a progressively reducing boundary layer thickness.

9
M Coutanceau and JR Defaye. Circular cylinder wake configurations: a flow visualization survey. Appl
Mech Rev 1991; 44(6), pp. 255-305. For ReD from 4-5105 – 3.5106,   140°, and surprisingly from ReD
 3.5106,   110°.

Prof. Hui Tong Chua 94


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
The correlation for the average Nusselt number can be described as

hD
Nu D   C Re mD Pr1 3 (4.15)
k

where the values for both C and m are tabulated in table 4.1. Note that all the relevant
fluid properties must be determined at the film temperature.

Table 4.1 Values of C and m in (4.15)


ReD C m
0.4 – 4 0.989 0.330
4 – 40 0.911 0.385
40 – 4000 0.683 0.466
4000 – 40000 0.193 0.618
40000 – 400000 0.027 0.805

(4.15) can also be applied to gas flow over cylinders with noncircular cross section. The
appropriate values for C and m in addition to the characteristic length D are tabulated in
table 4.2.

Prof. Hui Tong Chua 95


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

Table 4.2 Values of C and m in (4.15) for noncircular cylinders


Geometry ReD C m
3 5
Square 510 – 10 0.246 0.588
u
D

u 5103 – 105 0.102 0.675


D

Hexagon 5103 – 1.95104 0.160 0.638


u
D
1.95104 – 105 0.0385 0.782

u 5103 – 105 0.153 0.638


D

Vertical plate 4103 – 1.5104 0.228 0.731


u
D

For circular cylinder, a single comprehensive correlation exists over the entire range of
ReD and for a wide range of Pr:

45
0.62 Re1D2 Pr1 3   Re D  
58

Nu D  0.3  1     Re D Pr  0.2 (4.16)



1  0.4 Pr  
23 14
  282000  

All fluid properties must be evaluated at the film temperature.

Example 4.4

An experiment on the convection heat transfer involving air flow across a cylinder is
conducted in a wind tunnel. The metallic cylinder concerned has a diameter of 12.7mm
and a length of 94mm. It is internally heated by an embedded resistive heater and is
externally cooled by a cross flow of air. It is observed that for an air velocity of 10 m/s
and a free-stream temperature of 26.2°C, power consumed by the heater is 46W whilst
the average surface temperature of the cylinder is maintained at 128.4°C. It is estimated
that about 15% of the electrical power is lost via surface radiation and conduction from
the two ends of the cylinder.

Prof. Hui Tong Chua 96


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

Air Ts = 128.4°C
T = 26.2 °C
u = 10 m/s

D = 12.7mm L = 94mm

Determine the average convection coefficient based on this experimental observation.

This can be determined by invoking the Newton’s law of cooling where

q 0.85  46
h 
ATs  T   12.7 10  94 10 3 128.4  26.2
3

 102 W m 2  K

Compare this experimentally determined value with that computed from the relevant
correlation.

Making use of (4.15) and table 4.1, we have

hD
Nu D   C Re mD Pr1 3 .
k

Tf = 350K so that Pr = 0.70,   20.92 10 6 m 2 s , and k = 0.030 W/mK.

u  D 10 12.7 10 3
Re D    6071 .
 20.92 10 6

According to table 4.1, we have C = 0.193 and m = 0.618. Hence,

k 0.03
h C Re 0D.618 Pr1 3  3
0.193  60710.618  0.71 3  88 W m 2  K
D 12.7 10

This is within 14% of the measured value.

Alternatively, we can make use of (4.16), so that

Prof. Hui Tong Chua 97


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

 0.626071 0.701 3
12
  6071  5 8 
45

0.03  
h 0.3  1     
12.7 10 3
 1 
 0.4 0.70 23 14
   282000   

 96.0 W m  K
2

This is within 6% of the measured value.

Note that this does not in anyway imply that (4.16) is superior to (4.15). A host of
uncertainties are associated with the measurement of air velocity, estimation of the heat
loss from the two ends of the cylinder and surface radiation, and estimation of the
average cylinder surface temperature, so that the measurement error is no better than 15%.
Hence, we can only conclude that all the three values are mutually consistent with one
another.

End of example

Example 4.5 (Adapted from Incropera, DeWitt, Bergman, Lavine, Fundamentals of Heat
and Mass Transfer, 6th ed., 2007, John Wiley & Sons)

Hydrogen has been stored in 350-bar cylinders on our Transperth buses operated by the
proton-exchange membrane fuel cells. We are keen to reduce the hydrogen storage
pressure on vehicles, cars in particular. This can be realized by adsorbing it into a metal
hydride powder. When automotive power is needed, hydrogen can then be released by
desorption through the application of heat over the entire storage cylinder.

The pressure of the desorbed hydrogen found in the midst of the metal hydride powder
depends intimately on the hydride temperature as

p H 2  exp  3550 T  12.9

where p H 2 is the hydrogen pressure in atm and T is the hydride temperature in Kelvin.
The desorption process is endothermic and the heat of desorption can be described as

 
 H 2 kg s 29.5 103 kJ kg .
E g  m

Thermal energy has to be accordingly supplied so that the hydride is maintained at a


sufficiently high temperature such that the hydrogen pressure remains above 1 atm. The
latter is to ensure that hydrogen can be reticulated to the fuel cell which operates at 1 atm.

At a cruising speed of 25 m/s, a fuel cell powered vehicle consumes


m H2  1.35 104 kg s . This amount of hydrogen is supplied from a cylindrical, stainless
steel cylinder with an inner diameter Di = 0.1m, length L = 0.8m, and thickness t =
0.5mm. The hydride loaded cylinder is installed onboard such that it is subjected to
cross-flow air at u = 25m/s, T = 23°C. Determine the amount of heating, over and

Prof. Hui Tong Chua 98


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
above what is available from the warm air, that is required to the cylinder so that
p H 2  1atm .

Metal hydride
storage
cylinder
Air
u, T

Fuel-cell
stack

Air
T = 23°C
u = 25m/s

Hydride, E g

L = 0.8m

Di = 0.1m Stainless steel thickness


t = 0.05mm

We need to determine the threshold temperature at which the metal hydride could sustain
a hydrogen pressure of 1atm. From

p H 2  exp  3550 T  12.9 , we have

 3550  3550
T   275.2K .
ln p H 2  12.9 ln 1  12.9

The thermal power required to desorb the hydrogen at this threshold temperature is given
by

 
E g  1.35 104  29.5 106  3982W .

Part of this energy comes from convection heat transfer from the incoming air. The
Reynolds number of the air flow is

Prof. Hui Tong Chua 99


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

u  Di  2t  23  0.1  2  0.005


Re D    173760
 14.56  10 6

We shall try to use


45
0.62 Re1D2 Pr1 3   Re D  
58

Nu D  0.3  1     Re D Pr  0.2 .

1  0.4 Pr 
23 14

  282000  

For the prevailing air conditions, Re D Pr  173760  0.712  123717  0.2 . Hence,

45
0.62173760 0.7121 3   173760  
12 58

Nu D  0.3   
1     315.8 .

1  0.4 0.712
23 14

  282000  

Thence,

k 25.3  10 3
h  Nu D  315.8   72.6 W m 2  K
Di  2t  0.1  2  0.005
Therefore,

T  Ti
q conv 
1 ln D i  2t  D i 

LD i  2t h 2k ss L
296  275.2

1 ln 0.1  2  0.005 0.1

  0.80.1  2  0.00572.6 2  13.4  0.8
 406W

The additional heating power needed is therefore

qadd = 3982 – 406 = 3576W.

Prof. Hui Tong Chua 100


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Chapter 5 Internal flow


For internal flow applications, in addition to ascertaining whether the flow is laminar or
turbulent, we have to be cognizant of the existence of the entrance and fully developed
flow regimes.

5.1 Flow conditions

Inviscid flow region Boundary layer region

 ro

x xfd,h
Hydrodynamic entrance region Fully developed region

Fig 5.1 Laminar hydrodynamic boundary layer


development in a tube

Referring to figure 5.1, when we have a laminar flow in a circular tube, viscous effect at
the wall results in the formation of a hydrodynamic boundary layer which thickens with
increasing x until xfd,h. At which point, the boundary layer is fully developed and
remains constant along the tube. We term xfd,h the hydrodynamic entry length. For a
laminar flow, the fully developed velocity profile is parabolic. In contrast, the velocity
profile for a turbulent flow is flatter as a result of turbulent mixing in the radial direction.

Once again, the governing dimensionless number for the various flow regimes is the
Reynolds number defined as

u m D u m D
Re D   (5.1)
 

where um is the mean fluid velocity across the tube cross section and D is the inner
diameter of the tube. In a fully developed flow, the onset of turbulence is characterized
by a critical Reynolds number

Re D,c  2300 (5.2).

A fully turbulence regime is established when ReD  10000. It is also likely that the
transition to turbulence occurs within the hydrodynamic entrance region.

Prof. Hui Tong Chua 101


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
For laminar flow, where ReD  2300, the hydrodynamic entry length is approximately

 x fd ,h 
   0.05 Re D (5.3)
 D lam

provided that the fluid enters the tube from a rounded convergent nozzle as suggested in
figure 5.1. This is to ensure that the fluid has a nearly uniform velocity profile at the tube
entrance. In contrast, for turbulent flow in tube, it is fully developed when (x/D) > 10.

5.2 The mean velocity

The mean velocity is a very convenient quantity to work with in internal flows. It is
defined as

m

um  (5.4)
A c

where m  is the mass flow rate and Ac is the internal cross sectional area of the tube. For
a steady, incompressible flow in a circular tube, (5.4) implies that the Reynolds number
can be expressed as

4m

Re D  (5.5).
D

5.3 Pressure gradient and friction factor in a fully developed flow

A sustained pressure drop is essential in maintaining an internal flow. This in turn


dictates the pump or fan power requirement. Pressure drop is most conveniently
characterized by the Moody or Darcy friction factor defined as

 dp dx D
f (5.6).
u 2m 2

Accordingly, the pressure drop can be related to the friction factor as

L u 2m
p  f (5.7)
D 2

where L is the length over which the pressure drop occurs. Accordingly, the power
required to sustain this pressure drop is given as

Prof. Hui Tong Chua 102


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

P  pV (5.8)

where V is the volumetric flow rate which can be expressed as m


  for an
incompressible fluid.

You have learnt in a fluid mechanics (or thermal-fluid) course that in a fully developed
laminar flow,

64
f (5.9).
Re D

The friction factor for a wide range of Reynolds number is classically encapsulated in the
Moody diagram (figure 5.2).

For turbulent flow, the friction factor can be adequately found from the following
equation which is within two percent of what is available from the Moody diagram:

1  6.9   D 1.11 
 1.8 log     (5.10)
f  Re  3.7  

Prof. Hui Tong Chua 103


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John Wiley & Sons.

 (m)
Drawn tubing 1.5
Commercial steel 46
Cast iron 260
Concrete 300-3000

Fig 5.2 Friction factor for fully developed flow in a circular tube.
Source: LF Moody. Friction factors for pipe flow. Trans ASME 1944; 66, pp. 671-
Prof. Hui Tong Chua
684. 104
Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

5.4 Thermal conditions

qs


ro

x Ts xfd,t Ts
Hydrodynamic entrance region Fully developed region

Fig 5.3 Thermal boundary layer development in a


heated tube.

Figure 5.3 shows the progressive development of the thermal boundary layer in a circular
heated tube. The specific development depends, in addition, to the tube surface condition.
If the tube surface is constrained by a constant surface temperature or a constant surface
heat flux condition, a fully developed thermal condition will eventually emerge as
illustrated in the figure.

For laminar flows, the thermal entry length can be expressed as

 x fd , t 
   0.05 Re D Pr (5.11).
 D lam

Comparing (5.3) and (5.11), it is evident that if Pr > 1, the hydrodynamic boundary layer
develops more rapidly than the thermal boundary layer, conversely if Pr < 1 the thermal
boundary layer develops faster than the hydrodynamic boundary layer. It is therefore
easy to imagine that for oil (Pr  100), the thermal boundary layer will develop over a
long distance.

For turbulent flows, xfd,t/D  10.

5.5 The mean temperature

Just as we define a mean velocity, it is convenient for us to define a mean temperature for
flows within a tube. The mean (or bulk) temperature of the fluid at a particular cross
section is defined with respect to the thermal energy transported by the fluid, E t , as it
crosses the particular section:

Prof. Hui Tong Chua 105


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

E t   uc v TdAc (5.12).


Ac

Hence, the mean temperature is defined so that

E t  m
 c v Tm (5.13).

Accordingly, the Newton’s law of cooling can now be expressed as

qs  hTs  Tm  (5.14)

where h is the local convection coefficient. Note that Tm increases with x if heat is added
to the fluid from the tube surface.

5.6 Energy balance

dq conv  qsPdx

dm h
m
h h
m dx
dx

x dx
Fig 5.4 A control-volume analysis
of flow in a tube.

Let us analyze the energy consideration related to flow in a tube. Referring to figure 5.4
we have

dm h 
0m  h  mh dx   dq conv
 dx 

h dTm
m
 dx  dq conv  qsPdx
Tm dx

dTm
m
 cp  qsP  PhTs  Tm  (5.15)
dx

where P is the perimeter of the tube cross section and h the specific enthalpy of the fluid.

Prof. Hui Tong Chua 106


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

5.6.1 Constant surface heat flux

For a constant surface heat flux condition, (5.15) implies that

dTm qsP
  constant
dx m
 cp

qsP
Tm  Tm,inlet  x (5.16).
m
 cp

This means that the mean temperature will evolve linearly with x along the tube as
illustrated in figure 5.5.

Fully developed
region
T
Ts(x)

Tm(x)
qs = constant

x
Figure 5.5 Mean temperature
variation along a tube for
constant heat flux condition.

Note that since the convection coefficient in the fully developed region is a constant, a
constant surface heat flux condition translates to a constant temperature difference (Ts –
Tm) in the fully developed region.

Example 5.1

A water heating system is intended to process water from an inlet temperature of Tm,i =
20°C to an outlet temperature of Tm,o = 60°C. The water is heated by passing through a
tube embedded with resistive heaters within its wall. The tube has internal and external
diameters of 20 mm and 40 mm respectively and the resistive heaters generate a uniform
heating rate of q  106 W m3 .

Prof. Hui Tong Chua 107


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

q  106 W m3

m

Di = 20mm Do = 40mm

Tm,i  20C
Tm,o  60C

x Insulation

Given a water mass flow rate of m   0.1kg s , determine the length of the tube so as to
achieve the targeted outlet temperature.

Now, at a Tm  313K , cp of water = 4.179 kJ/kgK.

At steady state, all the heat generated within the tube must be convected away by the
water. Hence,

 2
q
4

Do  Di2 L  m 
 c p Tm,o  Tm,i 


 c p Tm ,o  Tm ,i 
4m
L

q  D o2  Di2 
4  0.1 417960  20


106  0.04 2  0.02 2 
 17.7m

If the inner surface temperature of the tube is 70°C at the outlet, determine the convection
heat transfer coefficient at the same location.

We know that,

qs
ho 
Ts ,o  Tm,o

Let us assume, quite reasonably, that a uniform internal heat generation also gives rise to
a uniform surface heat flux condition at the inner surface of the tube. Hence,

Prof. Hui Tong Chua 108


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

qs 
 
q  4 D o2  Di2 L q D o2  Di2

 
Di L 4D i



106 0.04 2  0.02 2 
4  0.02
 1.5 10 4 W m 2

qs 1.5 104


Hence h o    1500 W m 2  K
Ts,o  Tm,o 70  60

One ramification of this finding is that if the flow is fully developed throughout the tube,
then the convection coefficient will remain the same throughout as ho. In addition, the
temperature difference, Ts – Tm, remains constant at 10°C along the entire tube.

5.6.2 Constant surface temperature

For a constant surface temperature condition, (5.15) implies that

dTm
m
 cp  PhTs  Tm 
dx

dTm PhTs  Tm 

dx m cp

Ts  Tm ,o 
dTs  Tm 
L
Ph

Ts  Tm ,i  Ts  Tm 
 
0
m
 cp
dx


T  T    PL  1 hdx  L

T  T  m c  L  
s m,o
ln
s m ,i p 0


T  T   exp  PL h 
s m, o

T  T 
s
 m
 c 
m,i

p
(5.17)

(5.17) suggests that the temperature difference (Ts – Tm) decays exponentially with
distance along the length of the tube. This is illustrated in figure 5.6.

Prof. Hui Tong Chua 109


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

T Ts
To
Ti Tm(x)

Ts = constant

x
Figure 5.6 Mean temperature
variation along a tube for
constant surface temperature
condition.

The total heat transfer rate, qconv, is given by

 cp Ts  Tm,i   Ts  Tm,o  .


 cp Tm,o  Tm,i   m
qconv  m

From (5.17), m
 c p can be replaced so that qconv becomes

q conv  hPL
T  T   T  T   hA T
s m,o s m,i
(5.18),
ln
T  T 
s m,o
s lm

T  T 
s m,i

where As = PL and Tlm is the log mean temperature difference.

In many other practical situations, the temperature of the external fluid, T , as shown in
figure 5.7, is fixed instead of the tube surface temperature, Ts.

Tm,i
Tm,o

Interior flow
m , hi External flow
T, h o

Fig 5.7 Heat transfer between an external fluid and a fluid


in the tube.

Prof. Hui Tong Chua 110


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

In these cases, it is readily evident that Ts and h in (5.17 and 5.18) can be replaced by T
and U respectively where U is the average overall heat transfer coefficient.
Consequently, for such situations, we have

T  Tm, o   UAs 


 exp    (5.19)
T  Tm , i 
 m
  c p

and

q conv  UAs
T   Tm, o   T  Tm,i 
 UAs Tlm (5.20).
ln
T  Tm, o 
T  Tm,i 
Example 5.2

A thin walled circular tube with a diameter of 50mm and a length of 6m has its wall
maintained at a uniform surface temperature of 100C by having steam condensing on its
outer surface. Water flows through the tube at a flow rate of 0.25 kg/s. It is heated from
an inlet temperature, Tm,i = 15C, to an outlet temperature, Tm,o, of 57C. Determine the
average convection coefficient related to this heat-transfer problem.

L = 6m

Tm,i = 15°C
Tm,o = 57°C

Water
0.25 kg s Ts = 100°C
Di = 50mm

The mean water temperature is 36C, so that cp = 4178 J/kgK. Now,

 cp Tm, o  Tm,i   hDLTlm


q conv  m

 cp Tm, o  Tm,i 
m
h ,
DLTlm

and

Prof. Hui Tong Chua 111


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

Tlm 
T  T   T  T 
s m,o s m,i

ln
T  T 
s m,o

T  T 
s m,i


100  57  100  15
ln
100  57 
100  15
 61.6 C

Hence

0.25  417857  15


h  756 W m2  K
  0.05  6  61.6

5.7 Laminar flow in circular tubes

In a circular tube subjected to constant surface heat flux and laminar, fully developed
conditions, the Nusselt number can be shown to be

hD
Nu D   4.36 qs  cons tan t (5.21).
k

In contrast, for constant surface temperature conditions and for similar flow conditions,
the Nusselt number is

NuD = 3.66 Ts = constant (5.22).

Note that k should be determined at Tm.

Example 5.3

One design option for solar energy collection involves positioning a tube along the focal
line of a parabolic trough reflector and passing a fluid to be heated through the tube.

Prof. Hui Tong Chua 112


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

Insulation
Absorber
tube

Concentrator

This design consideration can be approximated as one involving a constant heat flux at
the tube surface. The tube has a diameter of 60mm and on a good day qs  2000 W m2 .

qs  2000 W m2

Water
 = 0.01kg/s
m
Di = 60mm

Tm,i  20C
Tm,o  80C

Pressurized water with a flow rate of 0.01 kg/s enters the tube at Tm,i = 20C. Determine
the length of the tube in order to achieve an outlet water temperature of 80C.

Now Tm = 323K and water cp = 4181 J/kgK.

 c p Tm , o  Tm ,i 
qsDL  m

 c p Tm , o  Tm ,i 
m
L
qsD

0.01  418180  20


Hence, L   6.65m .
2000    0.06

Prof. Hui Tong Chua 113


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
What is the surface temperature of the tube at outlet?

At Tm,o = 353K, water k = 0.670 W/mK,  = 35210-6 Ns/m2, Pr = 2.2.

qs
Now, qs  hTs,o  Tm, o  so that Ts, o   Tm, o .
h

4m
 4  0.01
Re D    603
D   0.06  352  10 6

Hence, the flow is laminar. x fd D  0.05 ReD Pr  0.05  603  2.2  66.3 . This is
smaller than L/D = 110. Hence the flow is fully developed.

hD k 0.67
Nu D   4.36 qs  cons tan t , h  4.36  4.36  48.7 W m2  K .
k D 0.06

Hence, the tube surface temperature at outlet is

2000
Ts, o   80  121 C
48.7

The process water must be significantly pressurized to prevent boiling at the tube internal
surface.

5.7.1 The entry region

The Nusselt numbers in the flow developing region are generally higher than that at the
fully developed regime. Solutions have been obtained for two different entry conditions.
The simplest being the thermal entry length problem where the thermal conditions are
developing in the midst of a fully developed velocity profile. Such a situation occurs
when the location at which heat transfer commences is preceded by an unheated starting
length. In addition, this thermal entry length problem is also germane to laminar heat
transfer for large Prandtl number fluids, such as oils. For such fluids, the velocity
boundary layer develops far more rapidly than the thermal boundary layer so that heat
transfer can be approximated as a simple thermal entry length problem. The combined
(thermal and velocity) entry length problem is the second entry condition where both the
temperature and velocity profiles develop simultaneously.

Figure 5.8 illustrates the development of the local Nusselt number in a circular tube as a
function of the reciprocal of the Graetz number, GzD = (D/x)ReDPr. The Nusselt
numbers are infinite at x = 0 and decrease to their asymptotic values with increasing x.
For the thermal entry length problem, the Nusselt number is independent of Pr. In
contrast, for the combined entry length problem, the Nusselt number decreases with
increasing Pr and approaches the thermal entry length results as Pr  . The result for

Prof. Hui Tong Chua 114


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
Pr = 0.7, which is relevant to most gases, is shown in figure 5.8. Interesting to note that
fully developed conditions are approached at GzD-1  0.05.

Figure 5.8 Local Nusselt numbers for laminar flow in a circular tube

20

18

16
Constant surface heat
14 flux
Thermal entry length
12 Combined entry length (Pr = 0.7)
NuD

10

4
Constant surface temperature
2

0
0.001 0.01 0.1 1

Gz-1 = x/(D.ReD.Pr)

5.8 Turbulent flow in circular tubes

For fully developed turbulent flow in circular tubes, the Dittus-Boelter equation is one of
the most famous correlations used for analyzing heat transfer:

n  0.4 heating Ts  Tm 


Nu D  0.023 Re 4D 5 Pr n  (5.23),
n  0.3 cooling Ts  Tm 

 
0.7  Pr  160
 Re  10000 
 D 
 L
 10 
 D 

All properties should be evaluated at Tm. The Dittus-Boelter equation is good for small
to moderate temperature differences, i.e. Ts – Tm. For flows involving large property
variations, the following correlation is preferred:

0.14

Nu D  0.027 Re 45
D Pr  
13
(5.24),
 s 

Prof. Hui Tong Chua 115


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

 
0.7  Pr  16700
 Re  10000 
 D

 L 
 10
 D 

All properties, except s, are to be evaluated at Tm. The abovementioned two correlations
are applicable to both uniform surface temperature and heat flux conditions. Note that
despite extensive experimental validations, errors as big as 25% are to be commonly
expected. For smaller Reynolds number and for greater accuracy in the order of 10%, the
following correlation used in conjunction with the Moody diagram is recommended:

Nu D 
f 8Re D  1000 Pr (5.25),
1  12.7f 8 Pr 2 3  1
12

 0.5  Pr  2000 
3000  Re  5 106 
 D 

Example 5.4

Hot air with a mass flow rate m   0.05 kg s flows through a bare sheet metal duct
located in the crawlplace of a building. The metal duct has a diameter D = 0.15m and a
length L = 5m. The hot air temperature at the duct inlet is 103°C. At the end of the duct,
the air temperature drops to 77°C. The ambient air around the duct is T = 0°C. The
heat transfer coefficient, ho, between the ambient air and the duct external surface is 6
W/m2K.

Cold ambient T = 0°C


ho = 6 W/m2K Ts(L)

Hot air
m = 0.05kg/s
D = 0.15 m

Tm,0  103C
L = 5m Tm,L  77C
x

Calculate the rate of heat loss from the duct over its entire length.

At Tm  363K , air cp = 1010 J/kgK.

Prof. Hui Tong Chua 116


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons

 c p Tm,L  Tm,0 
qm
 0.05  101077  103  1313W

Determine the heat flux and the duct surface temperature at x = L.

At Tm,L = 350K, air k = 0.030 W/mK,  = 20810-7 Ns/m2, Pr = 0.70.

From the following thermal resistance network,

Tm,L Ts,L T
q s L 
1 1
h x L  ho

Tm,L  T
we reason that qsL   .
1 1

h x L  h o

In order to ascertain hx(L), we need to determine the flow regime at x = L.

4m 4  0.05
Re D    20404
D   0.15  208  10 7

The flow is therefore fully turbulent. Now L/D = 5/0.15 = 33.3, we therefore expect the
flow conditions to be fully developed at x = L and we can confidently apply the Dittus-
Boelter equation with n = 0.3 so that

h x LD
 0.023 Re 4D 5 Pr 0.3  0.02320404 0.700.3  57.9
45
Nu D 
k
k 0.03
h x L  Nu D  57.9  11.6 W m 2  K
D 0.15

77  0
Hence qsL    304.5 W m 2 .
1 1

11.6 6

In order to determine the duct surface temperature at x =L, we also know that
T  Ts ,L
qsL   m ,L , so
1
h x L 

Prof. Hui Tong Chua 117


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons
qsL 304.5
Ts,L  Tm,L   77   50.7  C .
h x L  11.6

Prof. Hui Tong Chua 118


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Chapter 6 Transient conduction

6.1 The lumped capacitance method

A hot metal forging is suddenly quenched in a liquid bath. Can we predict how does the
temperature of the hot metal forging relax to the temperature of the liquid bath?

Hot metal
t<0
forging
Ti T = Ti

Cooling
liquid bath
E out  q conv
t0 T(t) E st
T = T(t)

Fig 6.1 Cooling of a hot metal forging.

Metal has a high thermal conductivity, so it is safe to assume that the resistance to
conduction within the metal forging is small in relation to the resistance to convection.
Let us apply an overall energy balance to the metal forging so that

E st  E out
or
dT
c p V  hAs T  T  (6.1).
dt

Introducing the variable  (= T - T∞), (6.1) reduces to

d
c p V  hAs 
dt
or
c p V  d t

hA s 
i

  dt , where i = Ti - T∞. Solving, we have
0

 hA s
ln  t
i c p V
or

Prof. Hui Tong Chua 119


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

 T  T
 e t    e t  (6.2)
i Ti  T

c p V
where   .  has a dimension of time and is a so called thermal time constant; the
hA s
larger the , the longer it takes the object to reach thermal equilibrium when  = 0 as
shown in figure 6.2.

Fig 6.2 Temporal temperature response for solids with different thermal time constants

1.0

0.9

0.8

0.7

0.6
/i

0.5 =1 =2  =3 =4

0.4

e-1 = 0.368
0.3

0.2
e-2 = 0.135

0.1
e-3 = 0.050

0.0
0 =1 2 3 4 5 6
time

6.2 Validity of the lumped capacitance method

What should be the criterion to judge if a lumped capacitance method is appropriate?

Consider steady-state conduction through a plane wall of area A as shown in figure 6.3.

Prof. Hui Tong Chua 120


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

T
Bi << 1
Ts,1 Ts,2
1
Ts,2
Bi >> 1
Ts,2
T∞, h
x L
Fig 6.3 Effect of Biot number on
steady-state temperature distribution in
a plane wall with surface convection

Applying energy balance at the surface where x = L,

Ts, 2  Ts,1
 kA  hATs, 2  T 
L

or

Ts,1  Ts, 2 L kA  R cond hL


    Bi (6.3).
Ts, 2  T 1 hA  R conv k
hL/k is a dimensionless number and is termed the Biot number. If Bi << 1, the resistance
to conduction within the material is much less than the resistance to convection at the
material’s surface. In other words, the temperature gradient within the material is much
less than the temperature gradient between the surface of the material and the free stream.

This concept is readily extended to a transient situation, so that if

Bi = hLc/k < 0.1,

the lumped capacitance method can be used with confidence. Lc is a characteristic length
defined as V/As. This generic definition enables us to handle object with a complicated
shape.

Example 6.1 (Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer,


2002, John Wiley & Sons)

A thermocouple junction is used to measure temperature in a gas stream. The junction


can be idealized as a sphere. The convection coefficient between the junction surface and
the gas is h = 400 W/m2K. The thermophysical properties of the thermocouple junction
are k = 20 W/mK, cp = 400 J/kgK, and  = 8500 kg/m3. What is the maximum diameter
of the junction so that the time constant will not exceed 1s? With this maximum diameter,

Prof. Hui Tong Chua 121


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
if the junction is at 25°C and placed in a gas stream of 200°C, how long does it take for
the junction to reach 199°C?

Leads
Thermocouple
T∞ = 200°C junction
h = 400 W/m2K Ti = 25°C
 = 8500 kg/m3
cp = 400 J/kgK
Gas stream k = 20 W/mK

c p V
Let’s assume first that the lumped capacitance method applies. Now,   . For a
hA s
V 4r 3 3 r D
sphere,    . Hence, the maximum diameter is
As 4r 2 3 6

c p D
1
h 6
8500  400
 D
400  6
or
6
D  7.06  104 m .
8500

To check if our assumption of a lumped capacitance analysis is valid, let’s check the
corresponding Biot number

hLc h V h D 400 7.06 104


Bi      2.35  104  0.1 .
k k As k 6 20 6

Hence, our conclusion that the maximum diameter of the junction should not exceed
0.7mm is valid.

To know the time taken to reach 199°C, we use

T  T
 e t  , where  = 1s. Thence,
Ti  T

199  200
 e t 1
25  200
or

Prof. Hui Tong Chua 122


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
1
t   ln  5.2s .
 175

Temporal response of a thermocouple junction

225

200

175

150
Temperature (°C)

125

100

75

50

25

0
0 1 2 3 4 5 6 7 8 9 10
Time (s)/ (s)

Example 6.2

A 3-mm-thick panel of aluminum alloy (k = 177 W/mK, cp = 875 J/kgK, and  = 2770
kg/m3) is coated on both sides with epoxy which must be cured at a minimum
temperature Tc of 150C for at least 5 min. The curing process involves two steps: (1)
convective heating with air at T,o = 175C and a convection coefficient ho = 40 W/m2K,
and (2) air cooling at T,c = 25C and a convection coefficient hc = 10 W/m2K. If the
panel has an initial temperature of 25C and the curing process stops at a safe-to-touch
temperature of 37C, what is the minimum total elapsed time tt for the curing operation?

Prof. Hui Tong Chua 123


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

2L = 3mm

ho = 40 W/m2K hc = 10 W/m2K
Epoxy T,c = 25 C
T,o = 175 C

Aluminum, T(0) = Ti,o = 25C Aluminum, T(tt) = 37C

Step 1: Heating Step 2: Cooling

Let us check out the Biot number to ascertain if a lumped capacitance analysis is valid.

The characteristic length of the panel is given by

V/As = 2LWH/(2WH) = L = 1.5mm.

h o L 40  1.5  10 3
During the heating process, Bi h    3.4  10 4 ,
k 177
h L 10  1.5  10 3
whereas during the cooling process, Bi c  c   8.5  10 5 .
k 177

Hence, a lumped capacitance approximation is excellent.

Since the heating process holds that the panel must be heated at a minimum temperature
of 150°C for at least 5 min, the minimum heating time must be equal to the time that
takes to reach 150°C and another 5min.

To know the time taken to reach 150°C, we use

T  T t 
c p L 2770  875  1.5  10 3
 e , where     91.0s .
Ti  T ho 40

The time taken can be evaluated as

 T  T   150  175 
t   ln    91 ln    162.9s .
 Ti  T   25  175 

Prof. Hui Tong Chua 124


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Hence, the total time taken for the heating process is 162.9s + 300s = 462.9s. At which
point in time, the temperature of the panel will be

T  Ti  T et   T  25  175e462.9 91  175  174.1C .

The time taken for the cooling down process can be easily calculated as

 T  T  2770  875  1.5  103  37  25 


t   ln     ln    916.1s
 Ti  T  10  174.1  25 

Hence the total time taken for the curing process is 462.9s + 916.1s = 1379s.

Temperature profile of the panel during the curing process

185

165

145

125
Temperature (°C)

105

85
Actual curing period

65

45

25
0 200 400 600 800 1000 1200 1400
Time (s)

Prof. Hui Tong Chua 125


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Chapter 7 Radiation: Processes and Properties

7.1 Fundamental concepts

Radiation involves the transport of thermal energy by electromagnetic waves.


Consider a solid which is originally at a higher temperature Ts than that of its
surrounding Tsur (figure 7.1). A vacuum exists between the two entities so that heat
transfer cannot proceed by conduction and convection. The resultant cooling of the
solid is directly due to the net emission of thermal radiation, qrad,net, from the surface
to the surrounding. Thermal equilibrium will be reached when Ts equals Tsur.

Radiation from
surroundings qrad,net

Vacuum Solid
Surface radiation
Ts emission
Tsur

Fig. 7.1 Cooling of a solid by radiation.

Generally speaking, when thermal radiation passes through air, no significant portion
of the radiation is absorbed so that air is a nonparticipating medium. On the other
hand, fluids such as water vapor, liquids, etc. do absorb thermal radiation and are
therefore participating medium.

Thermal radiation

Thermal radiation is a form of electromagnetic radiation which propagates in vacuum


at the speed of co (= 2.998×108 m/s). The complete electromagnetic spectrum is
delineated below.

Prof. Hui Tong Chua 126


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Violet
Blue
Green
Yellow
Red
Infrared
X Rays Ultraviolet

Gamma Thermal radiation Microwave


Rays
0.4 0.7

10-5 10-4 10-3 10-2 10-1 1 10 102 103 104

, m
Figure 7.2 The electromagnetic radiation spectrum.

The spectral (viz. wavelength-dependent) nature of thermal radiation is one of two


features that complicate the description of thermal radiation. The second feature
relates to its directionality.
radiation emission
Monochromatic

Wavelength

Spectral nature of thermal radiation Directional nature of thermal radiation

Figure 7.3 Radiative emission from a surface.

7.2 Radiation intensity

In this unit, we shall not concern ourselves with the directional nature of thermal
radiation, so that thermal radiation associated with a surface is uniform in all
directions over the entire hemisphere directly above the surface.

Prof. Hui Tong Chua 127


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
7.2.1 Definitions
A surface, in general, will emit spectral radiation over the entire hemisphere directly
above it. We call that spectral hemispherical emission per unit surface area from the
surface, E. Since this surface is emitting uniformly over all possible directions over
the hemisphere, we call this a diffuse surface.

E is termed the spectral, hemispherical emissive power and has a unit of W/m 2m.
The total hemispherical emissive power, E (W/m2), which is the rate at which
radiation is emitted per unit area over all possible wavelengths and in all possible
direction over the hemisphere is accordingly given by


E   E  d . (7.1)
0

E is also known as the total emissive power.

7.2.2 Irradiation
What we have just discussed about radiation emitted by a surface could be extended
to incident radiation. In other words, we consider the amount of radiation intercepted
by the surface, or the irradiation, over the entire hemisphere directly above it.

This irradiation may originate from emission and reflection from other surfaces.

The spectral irradiation G (W/m2m) is defined as the rate at which radiation of


wavelength  is incident on a surface, per unit area of the surface and per unit
wavelength interval d about .

The total irradiation G (W/m2) can accordingly be expressed as


G   G  d . (7.2)
0

Example 7.1

The spectral distribution of surface irradiation is given as

Prof. Hui Tong Chua 128


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

1200

1000

800
G (W/m2m)

600

400

200

0
0 5 10 15 20 25
 (m)

Evaluate the total irradiation.

The total irradiation can be obtained from


G   G  d .
0

Hence,

5 20 25 
G   G  d   G  d   G  d   G  d
0 5 20 25

1 1
  5 1000  20  51000   25  201000  0
2 2
 20000 W m 2

END OF EXAMPLE

7.2.3 Radiosity
Radiosity accounts for all of the radiant energy leaving a surface. This radiation
includes the reflected portion of the radiation, as well as direct emission.

Prof. Hui Tong Chua 129


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Irradiation
Radiosity

Emission
Reflected
portion of
irradiation

Figure 7.4 Surface radiosity.

The spectral radiosity J (W/m2m) represents the rate at which radiation of


wavelength  leaves a unit area of the surface, per unit wavelength interval d about .

The total radiosity J (W/m2) associated with the entire spectrum is


J   J   d (7.3)
0

7.3 Blackbody radiation

When we describe the radiation characteristics of real surfaces, it is useful to make


use of the concept of a blackbody. The blackbody is an ideal surface having the
following properties.

1. A blackbody absorbs all incident radiation, regardless of wavelength and


direction.
2. For a prescribed temperature and wavelength, no surface can emit more
energy than a blackbody.
3. The radiation emitted by a blackbody is a function of wavelength and
temperature, and is independent of direction. In other words, a blackbody is a
diffuse emitter.

A directional A diffuse emitter


emitter

Prof. Hui Tong Chua 130


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Figure 7.5 The essential difference between a directional emitter and a diffuse emitter.

As the perfect absorber and emitter, the blackbody serves as a standard against which
the radiative properties of actual surfaces can be compared.

A blackbody could be very closely approximated by a cavity whose inner surface is at


a uniform temperature.

Diffuse emission

Isothermal surface

Figure 7.6 The signatures of an isothermal blackbody cavity: complete absorption,


diffuse emission from an aperture, and diffuse irradiation of interior surfaces.

If radiation enters the cavity through a small aperture as shown above, it is most likely
to experience many reflections before reemergence. Hence, it is almost entirely
absorbed by the cavity and blackbody behavior is very closely approximated.

If the cavity surface is isothermal, the radiation emitted by the inner surface will
stream through the small aperture after undergoing many reflections. Hence, the
emission from the small aperture will be diffuse.

Moreover, due to the cumulative effect of emission and reflection from the cavity
surface, the radiation field within the cavity must be of the same form as the radiation
leaving the small aperture. Hence, a blackbody radiation field prevails within the
cavity. Accordingly, any small surface within the cavity will experience a blackbody
irradiation. This surface will be diffusely irradiated, regardless of its orientation.

Blackbody radiation exists within the cavity regardless of whether the inner surface of
the cavity is highly reflecting or absorbing.

7.3.1 The Planck distribution

According to Planck, the spectral distribution of the spectral hemispherical emissive


power, E,b, of a blackbody emission can be obtained as:

Prof. Hui Tong Chua 131


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

C1
E  ,b , T   (7.4)
 exp C2 T   1
5

where C1 = 2h c o2 = 3.742×108 W·µm4/m2, C2 = hco/k = 1.439×104 µm·K, h is the


Planck’s constant with a value of 6.6256×10-34 Js, k the Boltzmann’s constant with a
value of 1.3805×10-23 J/K, and co is the speed of light in vacuum with a value of
2.998×108 m/s. T is the absolute temperature in Kelvin while  is the wavelength in
µm.

Equation (7.4) is known as the Planck distribution and it is plotted below for some
temperatures.

Spectral blackbody emissive power

1000000000
Visible spectral range
100000000
maxT = 2898 m.K
10000000

1000000
Spectral emissive power, E,b, W/m2.m

5800 K
Solar 2000 K
100000
radiation
10000 1000 K

1000
800 K
100

10
300 K
1

0.1
100 K
0.01
50 K
0.001

0.0001
0.4 0.7
0.1 1 10 100
Wavelength, , m

Figure 7.7

The total emissive power of a blackbody, Eb, may then be expressed as

 
C1
E b   E  ,b , T d   d  T 4 (7.5)
0 0
 5
exp C 2  T   1
where  (= 5.670×10-8 W/m2·K4) is the Stefan-Boltzmann constant. Equation (7.5) is
known as the Stefan-Boltzmann law.

From figure 7.7, it is evident that the Planck’s distribution exhibits a maximum
stationary point. Differentiating (7.4) with respect to  and setting the result to zero,
we obtain

maxT = C3 = 2897.8 mK (7. 6).

Prof. Hui Tong Chua 132


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

This result is known as the Wien’s displacement law and illustrated in figure 7.7.

7.3.2 Band emission

It is often useful to know the fraction of the total emission from a blackbody from a
certain wavelength interval or band (see figure 7.8).

Figure 7.8 Radiation emission from a blackbody in the spectral band 0 to .

1.E+08

8.E+07

6.E+07
E,b()

E
4.E+07

 ,b d
0

2.E+07


0.E+00
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
 (m)

For a prescribed temperature and the interval from 0 to , this fraction, F(0) can be
expressed as:

 

E  ,b d E  ,b d 
C1
T
C1
F0   0
 0
 d   dT 

T 4 T  exp C 2 T   1
4 5
0 T  exp C 2 T   1
5

E d 0
 ,b
0

 f T 

From the above observation, it is evident that the fraction of the radiation between any
two wavelengths 1 and 2 can be expressed as

2 1

 E ,b d   E ,b d
F1  2   0 0
 F0 2   F01 
T 4

F(0) as a function of T is tabulated in the table below.

Prof. Hui Tong Chua 133


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
T (mK) F(0)
200 0.000000
400 0.000000
600 0.000000
800 0.000016
1000 0.000321
1200 0.002134
1400 0.007790
1600 0.019718
1800 0.039341
2000 0.066728
2200 0.100888
2400 0.140256
2600 0.183120
2800 0.227897
2898 0.250108
3000 0.273232
3200 0.318102
3400 0.361735
3600 0.403607
3800 0.443382
4000 0.480877
4200 0.516014
4400 0.548796
4600 0.579280
4800 0.607559
5000 0.633747
5200 0.658970
5400 0.680360
5600 0.701046
5800 0.720158
6000 0.737818
6200 0.754140
6400 0.769234
6600 0.783199
6800 0.796129
7000 0.808109
7200 0.819217
7400 0.829527
7600 0.839102
7800 0.848005
8000 0.856288
8500 0.874608
9000 0.890029
9500 0.903085
10000 0.914199
10500 0.923710
11000 0.931890

Prof. Hui Tong Chua 134


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
11500 0.939959
12000 0.945098
13000 0.955139
14000 0.962898
15000 0.969981
16000 0.973814
18000 0.980860
20000 0.985602
25000 0.992215
30000 0.995340
40000 0.997967
50000 0.998953
75000 0.999713
100000 0.999905

0.9

0.8

0.7

0.6
F(0-)

0.5

0.4

0.3

0.2

0.1

0
0 2000 4000 6000 8000 10000 12000 14000 16000 18000 20000
Tx10-3, m.K

Figure 7.9 Fraction of the total blackbody emission in the spectral band from 0 to  as
a function of T.

Example 7.2

A large isothermal enclosure is maintained at a uniform temperature of 2000K.


Evaluate the emissive power of the radiation which emerges from a small aperture on
the enclosure surface. What is the wavelength 1 below which 10% of the emission is
concentrated? Evaluate also the wavelength 2 above which 10% of the emission is
concentrated. Ascertain the maximum spectral emissive power and the wavelength at
which this emission occurs. What is the irradiation incident on a small object placed
inside the enclosure?

Prof. Hui Tong Chua 135


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
E = Eb(T)

G = Eb(T)

Small
object

Enclosure
T = 2000K

5.E+05

4.E+05

2000K

3.E+05
E,b(T) (W/m2.m)

10%

2.E+05

1.E+05

10%

0.E+00
max
0 2 4 6 8 10 12 14 16 18 20
 (m)

We assume that the areas of the aperture and object are very small in relation to the
enclosure surface.

The emissive power from the aperture is given by

E  E b T   T 4  5.670 108  20004


 9.07 105 W m 2

The wavelength 1 below which 10% of the emission is concentrated can be found by
noting that F(01) = 0.10, so that

1T  2200mK, or 1 = 1.1m.

The wavelength 2 above which 10% of the emission is concentrated can be found
from

F(2) = 1 - F(02) = 0.1

Prof. Hui Tong Chua 136


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

or

F(02) = 0.9.

Hence, 2T = 9382mK or 2 = 4.69m.

From the Wien’s law, maxT = 2897.8 mK so that

max = 1.45 m.

The corresponding spectral emissive power can be obtained from

C1
E  ,b 
 exp C 2 T   1
5

3.742  108

   
1.455 exp 1.439  10 4 1.45  2000  1
 4.12  10 W m  m
5 2

The irradiation incident on a small object placed inside the enclosure is computed as

G  E b T   T 4  5.670 108  20004  9.07 105 W m2 .

END OF EXAMPLE

7.4 Surface emission

It is imperative to realize that, in general, the spectral radiation emitted by a real


surface differs significantly from the Planck’s distribution (figure 7.10).

Prof. Hui Tong Chua 137


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Figure 7.10 The spectral distributions of a blackbody and real surface

1.00E+08

1.00E+05
Blackbody, 2000K
E,b(,T)
1.00E+02

1.00E-01

Real surface, 2000K


E(T) (W/m 2.m)

1.00E-04

1.00E-07

1.00E-10

1.00E-13

1.00E-16
E(,T) = E,b(,T)

1.00E-19
0.01 0.1 1 max 10 100 1000
 (m)

For most engineering applications, we usually deal with surface properties that
represent directional averages. A spectral, hemispherical emissivity is therefore
defined as

E  , T 
  , T   (7.7)
E  ,b , T 

The total, hemispherical emissivity, which represents an average over all possible
directions and wavelengths, is defined as

ET 
T   (7.8)
E b T 

From (7.1) and (7.7), we therefore have

  , T E , T d


 ,b

T   0
(7.9).
E b T 

The emissivity of a diffuse emitter is a constant, independent of direction. In practice


it is taken to be equal to the normal emissivity n at  = 0°, so that

 = n (7.10).

This conclusion also applies to the spectral components.

Representative total, normal emissivities are shown in figures 7.11 and 7.12.

Prof. Hui Tong Chua 138


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Figure 7.11 Temperature dependence of the total, normal emissivity n of selected materials

0.9 Silicon carbide

0.8

0.7 Stainless steel, heavily


Total, normal emissivity, n

oxidized

0.6

Aluminum oxide
0.5

Stainless steel, lightly


0.4 oxidized

0.3

0.2
Tungsten
0.1

0
500 700 900 1100 1300 1500 1700 1900 2100 2300 2500
Temperature (K)

Metals, as received and unpolished


Metals, oxidized
Carbon, graphites
Oxides, ceramics
Minerals, glasses
Vegetation, water, skin
Special paints, anodized finishes

0 0.2 0.4 0.6 0.8 1.0

Highly polished metals, foils, films


Polished metals
Metals, as received
0 0.05 0.10 0.15
Figure 7.12 Representative values of the total, normal emissivity n.

Example 7.3
A diffuse surface at 1600K has the following spectral, hemispherical emissivity.

Prof. Hui Tong Chua 139


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

0.9

0.8
T = 1600K
1 = 2m
0.7 2 = 5m

0.6

0.5
 ()

0.4

0.3

0.2

0.1

0
0 1 2 3 4 5 6 7 8 9 10
 (m)

Determine the total, hemispherical emissivity and the total emissive power. What is
the wavelength at which the spectral emissive power is a maximum?

The total, hemispherical emissivity can be found from

 2 5

 E   ,b d 1  E  , b d   2  E  , b d
 0
 0 2

Eb Eb

or

  1F02    2 F05  F02   


1T = 21600 = 3200 mK  F(02) = 0.318
2T = 51600 = 8000 mK  F(05) = 0.856

Hence,
  0.4  0.318  0.80.856  0.318  0.558

The total emissive power is

E  E b   T 4

E  0.558 5.67  10 8  1600 4  207 kW m 2 
The location at which maximum spectral emissive power occurs is not immediately
obvious as the spectral emissivity is not a constant.

From the Wien’s law,

Prof. Hui Tong Chua 140


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

2898
 max   1.81m
1600

The spectral emissive power at max is

E   max , T      max E  , b  max , T 


C1
    max 
 exp C 2  max T   1
5

3.742  108
 0.4
 
1.815 exp 1.439  10 4 2898  1  
 54 kW m  m 2

We also need to ascertain the spectral emissive power at 2m where  = 0.8.

E  2m,1600K   0.8  E  ,b 2m, T 


3.742  108
 0.8

25 exp 1.439  10 4 2  1600  1 
 105.5 kW m  m 2

Hence the peak spectral emission occurs at 2m.

1.60E+05

1.40E+05

1.20E+05

1.00E+05
E,b, EW/m2.m

8.00E+04

6.00E+04

4.00E+04

2.00E+04

0.00E+00
0 1 2 3 4 5 6 7 8 9 10
 (m)

END OF EXAMPLE

7.5 Surface absorption, reflection and transmission

Prof. Hui Tong Chua 141


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
Firstly, spectral irradiation G (W/m2m) is defined as the rate at which radiation of
wavelength  is incident on a surface per unit area of the surface and per unit
wavelength interval d about . The irradiation may be incident from all possible
directions, and it may originate from more than one source.

In the most general case, the irradiation interacts with a semitransparent medium, such
as a layer of water or a glass plate as shown below:

Reflection G,ref
Irradiation G
Semitransparent
medium
G = G,abs + G, ref + G,tr

Absorption G,abs

Transmission G,tr

From a simple energy balance, it follows that

G = G,abs + G,ref + G,tr (7.11)

as the spectral irradiation could be absorbed, reflected and/or transmitted. Equation


(7.11) could be expressed further as

G = G + G + G


Or
1 =  +  +  (7.12)

where  is the spectral absorptivity,  the spectral reflectivity, and  the spectral
transmissivity.

If we work with the total irradiation, G (W/m2) which encompasses all spectral

contribution and could be evaluated as G   G  d , then
0

Prof. Hui Tong Chua 142


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
   

G
0
 d     G  d     G  d     G  d
0 0 0
  

  G  d  G   d  G   d
1 0

 0

 0

(7.13)
G
0
 d G
0
 d G
0
 d

1     

Thus, equation (7.13) provides the relation for the total absorptivity, , total
reflectivity, , and total transmissivity, .  and  are both equal to zero if the
material is opaque.

In general, reflection of radiation can either be specular or diffuse.

Reflected
radiation of 1 = 2
uniform
Incident ray intensity Incident ray Reflected ray
1 2

Surface Surface

Diffuse reflection Specular reflection

No surface, however, is perfectly diffuse or specular. A specular condition could be


approximately achieved by polished, mirror-like surfaces whereas a diffuse condition
could be approximately achieved by rough surfaces. Most commonly encountered
engineering applications involve rough surfaces and hence diffuse reflection.

7.6 Kirchhoff’s law

We are now going to consider the conditions under which the absorptivity and
emissivity are equal.

Consider a large, isothermal enclosure of surface temperature Ts which contains


several small bodies as shown below.

Prof. Hui Tong Chua 143


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

G = Eb(Ts)
G

E3

E1

E2
Ts

The bodies confined are small relative to the enclosure, so that they exert negligible
influence on the radiation field generated by the enclosure as a consequence of the
cumulative effect of emission and reflection by the enclosure surface.

Since the enclosure constitutes a blackbody cavity, the irradiation experienced by any
of the bodies confined is diffuse and equal to the blackbody emission at the enclosure
surface Ts. Hence, G = Eb(Ts).

Considering steady-state conditions, thermal equilibrium involving the enclosure and


all the bodies confined must prevail. Hence, the temperatures of all the bodies
confined must be identical and equal to Ts, so that T1 = T2 = … = Ts and the net rate
of heat transfer must be zero.

Applying an energy balance to a control surface about body 1, we have

1GA1 - E1(Ts)A1 = 0

E1 Ts 
or  G  E b Ts  .
1

The same argument applies to all of the other confined bodies so that

E1 Ts  E 2 Ts 
    E b Ts  (7.14).
1 2

Equation (7.14) is known as the Kirchhoff’s law. A major implication is that, since 
 1, E(Ts)  Eb(Ts). Hence, no real surface can have an emissive power more than
that of a black surface at the same temperature.

Prof. Hui Tong Chua 144


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
From the definition of the total, hemispherical emissivity, , which states that
ET 
T   , we have from equation (7.14)
E b T 

E1 Ts  E 2 Ts 
    E b Ts 
1 2

1E b Ts   2 E b Ts 
    E b Ts  (7.15)
1 2

1  2
 1
1  2

Hence, the total hemispherical emissivity of a surface at a temperature T is equal to its


total hemispherical absorptivity for radiation coming from a blackbody at the same
temperature. This is an alternative form of the Kirchhoff’s law.

Note carefully that this relation is derived under the condition that the surface
temperature is equal to the temperature of the source of irradiation. Hence, one must
exercise caution against using the Kirchhoff’s law when there is a significant
temperature difference (in excess of a few hundred degrees) between the surface and
the source of irradiation.

The above derivation can also be performed for radiation at a specified wavelength so
as to obtain the spectral form of the Kirchhoff’s law:

(T) = (T) (7.16).

Equation (7.16) is valid when the irradiation or the emitted radiation is diffuse.

It is useful to note that a surface for which its emissivity and absorptivity are
independent of direction and wavelength is termed a diffuse, gray surface (diffuse
because of the directional independence and gray because of the wavelength
independence). Such a surface satisfies equations (7.15-16), namely  =  and  = .

Example 7.4
An instrumentation transmitter pod is a box containing electronic circuitry and a
power supply for sending sensor signals to a base receiver for recording. Such a pod
is placed on a conveyor system which passes through a large vacuum brazing furnace
as shown in the sketch below.

Prof. Hui Tong Chua 145


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Coating

Vacuum furnace walls,


Tw = 1200 K, w = 0.7
Interior has
electronic
circuitry, 50 W
T = 87C dissipation Layer of PCM

Pod
Insulation
0.9

0.05
Conveyor system
0 5

, m

The exposed surfaces of the pod have a special diffuse, opaque coating with spectral
emissivity as shown in the sketch above.

To stabilize the temperature of the pod and prevent overheating of the electronics, the
inner surface of the pod is surrounded by a layer of a phase-change material (PCM)
having a fusion temperature of 87C and a heat of fusion of 25 kJ/kg. The pod has an
exposed surface area of 0.040 m2 and the mass of the PCM is 1.6 kg. Further, it is
known that the power dissipated by the electronics is 50W. Consider the situation
when the pod enters the furnace at a uniform temperature of 87C and all the PCM is
initially in the solid state. How long will it take before all the PCM changes to the
liquid state?

The irradiation received by the coating on the PCM from the furnace is diffuse.
Furthermore, the emitted radiation from the coating is also diffuse. Hence, we could
apply the spectral form of the Kirchhoff’s law so that

 = .

Considering the energy balance on the PCM, we have

Prof. Hui Tong Chua 146


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
 
m PCM H
 50  A PCM     w E  , b Tw d  A PCM    E  , b T d
 0 0

5 

     w E  , b Tw d      w E  , b Tw d  
0 5 
 50  A PCM  5  
  E T d   E T d 
   ,b    ,b 
0 5 
 5
E T  
E T  
 0.05  0.7Tw4   , b w d  0.9  0.7Tw4   , b w d  
 0
Eb 5
Eb 
 50  A PCM  
 0.05T 4 E  , b T  d  0.9T 4 E  , b T  d
5 




0
E b

5
E b

 Tw

 0.05  0.7Tw4 F05   0.9  0.7Tw4 F0   F05  
Tw


 50  A PCM 

4
T

 0.05T F05   0.9T 4 F0   F05 
T
 

 0.05  0.7Tw4 F05   0.9  0.7Tw4 1  F05 
 Tw
 
Tw


 50  A PCM 
 T

 0.05T 4 F05   0.9T 4 1  F05 
T
 

 0.05  0.7Tw4  0.737818  0.9  0.7Tw4 1  0.737818  
 50  A PCM  
 0.05T 4  0.039341  0.9T 4 1  0.039341 
 
 915.1851W

Hence, the time taken to fully melt the PCT, , is given by

1.6  25000
  44s .
915.1851

END OF EXAMPLE

Example 7.5

Components of an electronic package in an orbiting satellite are mounted on a


compartment which is well insulated on all but one side of its sides. The un-insulated
side consists of an isothermal copper plate whose outer surface is exposed to the
vacuum of outer space and whose inner surface is attached to the components. The
plate dimensions are 1m by 1m on a side.

Prof. Hui Tong Chua 147


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

q s  Isothermal plate,


 Tp

The exposed surface of the plate, which is opaque and diffuse, has a spectral
hemispherical absorptivity of  = 0.2 for   2m and  = 0.8 for  > 2m.
Consider steady-state conditions for which the plate is exposed to a solar flux of qs” =
1350 W/m2 which is incident at an angle of  = 30 relative to the surface normal. If
the plate temperature is Tp = 500K, how much power is being dissipated by the
components?

Solar spectral emissive power and plate spectral absorptivity

1.E+08 0.9

0.8

8.E+07
0.7

0.6

6.E+07
0.5

Plate 
E,b()

0.4
4.E+07

0.3

0.2
2.E+07

0.1

0.E+00 0
0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0
 (m)

Since the exposed surface of the plate is diffuse,  = .

Since the extraterrestrial solar irradiation could be considered to be a blackbody


irradiation at 5800K, it would be reasonable to assume that most of the solar
absorption occurs at   2m so that solar absorptivity is 0.2.

The amount of power dissipated by the components, P, could then be estimated as


follow.

Prof. Hui Tong Chua 148


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

P     E  ,b Tp d  0.2  1350  cos 30
0
2
E  ,b Tp  
E  ,b Tp 
 0.2T  d  0.8T  d  233.83
4 4
p p
0
Eb 2
Eb

 0.2Tp4 F02   0.8Tp4 1  F02    233.83


Tp  Tp 

 0.2Tp4  0.000321  0.8Tp4 1  0.000321  233.83


 2.6kW

END OF EXAMPLE

Prof. Hui Tong Chua 149


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Chapter 8 Radiation exchange between surfaces


We have so far been discussing processes that occur at a single surface. Now, we
shall consider radiative exchange between two or more surfaces.

This exchange depends strongly on the surface geometries and orientation, as well as
on their radiative properties and temperatures.

We assume that the surfaces are separated by a nonparticipating medium. Since such
a medium neither emits, absorbs nor scatters, it has no effect on the transfer of
radiation between surfaces. A vacuum meets these requirements exactly, and most
gases meet them to an excellent extent.

8.1 The view factor

The view factor Fij is defined as the fraction of the radiation leaving surface i, with an
area Ai, which is intercepted by surface j, which has an area Aj, so that

q i j
Fij  (8.1)
AiJi

Similarly, the view factor Fji is defined as the fraction of the radiation that leaves Aj
and is intercepted by Ai.

8.1.1 View factor relations

Based on more detailed geometrical consideration, the following reciprocity relation


has been proven

AiFij = AjFji (8.2)

For surfaces of an enclosure, energy conservation requires that all of the radiation
leaving surface i must be intercepted by the enclosure surfaces, possibly also by
surface i itself. The latter condition is possible if the surface is concave so that the
surface sees itself and Fii is nonzero. For a convex or plane surface, however, Fii = 0.
Hence for an enclosure with N surfaces,

F
j1
ij  1, 1  i  N (8.3)

Prof. Hui Tong Chua 150


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

TN
T1

T2
Ti

Figure 8.2 Radiation exchange


within an enclosure.

View factors for some two-dimensional geometries

Parallel plates with midlines connected by perpendicular

wi
Fij 
W  W 
i j
2
4   W  W 
12
j i
2
4 
12

2Wi
i
wi
Wi 
L
L wj
Wj 
L
j
wj

Perpendicular plates with a


common edge
j
Fij 

1  w j w i   1  w j w i  
2 12

2
wj

wi

Inclined parallel plates of equal width and a common


j edge
w

Fij  1  sin 
 2
i
Prof. Hui Tong
w Chua 151
Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Three-sided enclosure
wi  w j  wk
wj Fij 
wk 2w i
j
k
i

wi

Parallel cylinders of different radii

j


  C 2  R  12 1 2  

i
ri
rj


 C 2  R  12 1 2  

1   R   1   
Fij   R  1cos 1      
2 
 C   C  
 
s  R  1cos 1  R    1  
  C   C  
     
R = rj/ri, S = s/ri, C = 1+R+S

r Cylinder and parallel rectangle

j r  1 s1 1 s 2 
Fij  tan L  tan L 
L s1  s 2  
i

s2
s1

Infinite plane and row of cylinders


s

j
D
i
Prof. Hui Tong Chua 152
Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
12 12
  D 2  D  s2  D2 
Fij  1  1        tan 1  2

  s   s  D 

View factors for some 3-D geometries

Aligned parallel rectangles


XX L YY L
j
ln 

  1  X 2 1  Y 2 1 2    
12
tan 1


X
  X 1 Y
2

  1  X 2
 Y 2
  12
1 Y2  
L  
Fij 
2 
 Y 1  X 
2 12
tan 1  Y 

Y
XY  1 X2
12
  
 1 1 
i
X  X tan X  Y tan Y 
 

Perpendicular rectangles with a common edge

HZ X WY X
j
 
 
 
 
 
Z  
1 1
i  W tan 1  H tan 1 
 W H 
X 1  
Fij   H  W
2
 2 1 2

tan 1
1

Y W  H2  W2
1 2
 
 
 
 1  W 2 1  H 2  W 2 1  H 2  W 2     
2
W

 

  ln 
1  1  H 2
 W 2 
 1  W 2
H 2
 W2 
    
 4  
 
2
H
   H 2
1  H 2
 W 2
 
  

 
  1  H H  W 
2 2 2 
   
rj
Coaxial parallel disks
j
1  R 2j
L R i  ri L , R j  rj L , S  1 
R i2
ri
i
Fij 
1
2
 
S  S2  4rj ri 
2 12
 

Prof. Hui Tong Chua 153


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Figure 8.3 View factor for aligned parallel rectangles

1
 10
j 4
2

L
1.0

Y
i 0.6
X
0.4

0.1
Fij

0.2

Y/L = 0.1

0.01
0.1 1 10 100
X/L

Figure 8.4 View factor for coaxial parallel disks

1.0
8 r
0.9
j j
6
0.8 2 L
5 r
0.7 1.5 i i

0.6 4 1.25

0.5 1.0
Fij

0.4
0.8

0.3
0.6

0.2

0.4
0.1
rj/L = 0.3
0.0
0.1 1 10
L/ri

Prof. Hui Tong Chua 154


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Figure 8.5 View factor for perpendicular rectangle with a common edge.

0.5
Y/X = 0.02 j

0.05
0.4 Z
0.1 i
X
Y
0.2
0.3

0.4
Fij

0.6
0.2

1.0

1.5
2.
0.1
4
10
20
0
0.1 1 10
Z/X

It is worthwhile to observe that the results presented in figures 8.3 to 8.5 can be used
to determine other view factors. For example, the view factor for an end surface of a
cylinder (or a truncated cone) relative to the lateral surface may be obtained by using
figure 8.4 in tandem with (8.3).

Example 8.1

Determine the view factors F12 and F21 for the following geometries:

D A1
A1
A1
L=D A2
L A2 L = D

A2 A3 A3
(1) (2) (3)

(1) Sphere of diameter D in a cubical box of length L = D.


(2) Diagonal partition within a long square duct.
(3) End and side of a circular tube of equal length and diameter.

For the sphere within a cube,

F12 = 1.

From the reciprocity theorem,

A1 D 2 
F21  F12  2  1  .
A2 6L 6

Prof. Hui Tong Chua 155


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

For the partition within a square duct,

F11  F12  F13  1

where F11 = 0.

From symmetry, F12 = F13.

Hence, F12 = 0.5.

From the reciprocity theorem,

A1 2L
F21  F12   0.5  0.71 .
A2 L

For the circular tube,

From figure 8.4, with r3/L = 0.5 and L/r1 = 2, F13  0.17.

Now F11  F12  F13  1

With F11 = 0, F12 = 1 – F13 = 0.83.

From the reciprocity theorem,

A1 D 2 4
F21  F12   0.83  0.21
A2 DL

END OF EXAMPLE

8.2 Blackbody radiation exchange

In general, radiation may leave a surface due to both reflection and emission, and on
reaching a second surface undergoes absorption and reflection. In the case of
blackbodies, however, there is no reflection.

Consider radiation exchange between two black surfaces of arbitrary shape.

Prof. Hui Tong Chua 156


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

ni nj
Ji = Ebi Jj = Ebj

Aj, Tj
Ai, Ti

Figure 8.6 Radiation heat transfer between two


blackbodies.

We define qij as the rate at which radiation leaves surface i and is intercepted by
surface j. Accordingly, we have

qij = (AiJi) Fij (8.4)

For a black surface i, Ji = Ebi as there is no reflection from it. Hence,

qij = Ai Fij Ebi (8.5)

Also, qji = Aj Fji Ebj. (8.6)

The net radiative exchange between the two surfaces may then be defined as
qij = qij - qji = Ai Fij Ebi - Aj Fji Ebj. Hence,

qij = Ai Fij  ( Ti4  Tj4 ) (8.7)

From equation (8.7), we could further evaluate the net radiation transfer from any
surface in an enclosure of black surfaces. With N surfaces maintained at different
temperatures, the net radiation transfer from surface i could be expressed as

 
N
q i   A i Fij Ti4  Tj4 (8.8)
j1

Example 8.2

A cylindrical furnace cavity with a diameter 75mm and a length of 150mm is open at
one end to a large surrounding which is at 27°C. Its sides and bottom can be
approximated as blackbodies, are heated electrically, are well insulated, and are
maintained at 1350°C and 1650° respectively.

Prof. Hui Tong Chua 157


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

D
Side, T1

L Heater wire
Insulation

Bottom, T2

How much power is required to maintain the furnace condition?

Tsur = 300K
q
A3, T3 = Tsur

L = 0.15m
A1, T1 = 1350°C

A2, T2 = 1650°C
D = 0.075m

Since the surrounding is large, it can be approximated as a blackbody. The cavity


opening can be treated as a hypothetical surface with area A3 with a temperature T3 =
Tsur. Since it is through the cavity that the cylinder interacts with the surrounding,
surface A3 can be treated as a blackbody.

The heat loss to the surrounding is

q = q13 + q23.

From (8.7),

  
q  A1F13 T14  T34  A2F23 T24  T34 . 
From figure (8.4), rj/L = 0.0375/0.15 = 0.25 and L/ri = 0.15/0.0375 = 4, F23 = 0.06.

F21 = 1 – F23 = 1 – 0.06 = 0.94.

From the reciprocity theorem,

A2 0.0752 4
F12  F21   0.94  0.118
A1   0.075  0.15

Prof. Hui Tong Chua 158


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
From symmetry, F13 = F12. Hence,


q    0.075  0.15  0.118  5.67  108  16234  3004  

4

 0.0752  0.06  5.67  108  19234  3004 
q  1639  205  1844W

END OF EXAMPLE

8.3 Radiation exchange between diffuse, gray surfaces in an


enclosure

When we analyze radiation exchange between diffuse, gray surfaces in an enclosure,


it is common to assume that each surface of the enclosure is isothermal and
characterized by a uniform radiosity and irradiation. The surfaces are also assumed to
be opaque, diffuse and gray. The medium within the enclosure is taken to be
nonparticipating.

8.3.1 Net radiation exchange at a surface

The net rate at which radiation leaves surface i, qi, is the essentially directly
proportional to the difference between the surface radiosity and irradiation.

qi = Ai(Ji-Gi) (8.9)

From our previous discussion on radiosity,

Ji  Ei + iGi (8.10)

Hence, qi = Ai(Ei - [1 - i] Gi) = Ai(Ei - i Gi)

Now, equation (8.10) could also be re-expressed as

Ji = i Ebi + (1 - i)Gi = i Ebi + (1 - i)Gi

J i   i E bi
Hence, G i  . Substituting this result into equation (8.9), we obtain
1  i

1
q i  E bi  J i  (8.11)
1   i 
i Ai

Equation (8.11) can be readily interpreted as a radiative transfer with a driving


potential (Ebi – Ji) and a surface radiative resistance (1 - i)/ iAi.

Prof. Hui Tong Chua 159


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

8.3.2 Radiation exchange between surfaces

To make use of equation (8.11), the surface radiosity must be known by considering
all radiation exchanges between all the surfaces of the enclosure.

The irradiation of surface i can be evaluated from the radiosities of all the surfaces in
the enclosure, so that

N
A i G i   Fji A j J j
j1

Noting the reciprocity relation, AiFij = AjFji, we have

N
A i G i   Fij A i J j
j1


N
G i   Fij J j
j1

From equation (8.9), the net rate at which radiation leaves surface i, qi can be
expressed as

q i  A i J i  G i 
 N 
 A i  J i   Fij J j 
 j1 

N
From the summation rule, F
j1
ij  1 , we further have

 N 
q i  A i  J i   Fij J j 
 j1 
 N N 
 A i   Fij J i   Fij J j 
 j1 j1  (8.12)
q i   A i Fij J i  J j 
N

j1
N
  q ij
j1

This result naturally equates the net rate of radiation transfer from surface i, qi, to the
sum of components qij related to radiative exchange with the other surfaces. We

Prof. Hui Tong Chua 160


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
could interpret from equation (8.12) that the driving force for qij is (Ji – Jj) and (Ai Fij)-
1
is a space or geometric resistance.

Combining equations (8.11-12), we obtain

N J J
E bi  J i

i j
(8.13)
1   i   i A i j1 A i Fij 1

qi1

J1

(AiFi1)-1 qi2
J2

1  i
i Ai Ji (AiFi2)-1
qi

JN-1
Node (AiFiN) -1 (AiFi[N-1])-1
qi3
corresponding
to the surface i JN

qi4

Figure 8.7 Thermal-resistance-circuit representation of radiative exchange between


surface i and the rest of the surfaces in an enclosure.

Equation (8.13) is preferred when the surface temperature Ti is known. When,


however, the net radiation transfer rate at the surface qi is known rather than Ti,
equation (8.12) is more convenient. It could alternatively be expressed as

qi  
N J  J 
i j
(8.14)
j1 A F i ij
1

Example 8.3
During the fabrication process, the special coating on a curved solar absorber surface
is cured by exposing it to an infrared heater. The area of the absorber surface is A 2 =
15m2 while the width of the heater is W = 1m. The length, L, of both the absorber
and heater is equal to 10m

Prof. Hui Tong Chua 161


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Absorber surface, A2 = 15m2, T2, 2

H = 1m

Room walls, Tsur


W = 1m

Heater, A1, T1, 1

The heater is maintained at T1 = 1000K and has an emissivity 1 = 0.9. The absorber
is at T2 = 600K and possesses an emissivity 2 = 0.5. The system is housed in a large
room whose walls are at Tsur = 300K. What is the net rate of heat transfer to the
absorber surface?

A2 = 15m2, T2 = 600K, 2 = 0.5

H = 1m A2’

A3, T3 = Tsur, 3 = 1 Tsur = 300K


W = 1m

A1 = 10m2, T1 = 1000K, 1 = 0.9

This problem can be idealized as a three-surface enclosure, and we are after the
radiation heat transfer to surface 2.

From (8.11) we have

1
q 2  E b 2  J 2  (1)
1  2 
2A 2

where J2 is an unknown.

In view of the fact that the room is large and completely encloses the device, surface 3
is a blackbody at 300K and hence J3 = Eb3. Thus the only unknowns are J1 and J2.
From (8.13), we have

Eb2  J2 J J J J
 2 11  2 31
1  2  2A2 A2F21 A2F23 

Prof. Hui Tong Chua 162


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

where J 3  E b3  T34  459 W m 2 and Eb 2  T24  7348 W m2 .

From the reciprocity theorem, we have

A2F21 = A1F12

Where F12 = F12’ and A2 is just the rectangular base of the absorber surface. From
figure 8.3, with Y/L = 10/1 = 10 and X/L = 1/1 = 1,

F12 = 0.39
A 1 10
F21  1 F12   0.39  0.26
A2 15

Now,

F13 = 1 – F12 = 1 – 0.39 = 0.61.

From the reciprocity theorem,

A1 110
F31  F13   0.61  0.305 .
A3 2 110

Invoking symmetry,

F31 = F32’ = F32. Hence,

A3 2 110
F23  F32   0.305  0.41
A2 15

Thus the radiation balance for surface 2 is

7348  J 2 J 2  J1 J 2  459
 
1  0.5 0.5A 2 A 2  0.26 A 2  0.411
1

Or

7348 – J2 = 0.26J2 – 0.26J1 + 0.41J2 – 188


0.26J1 – 1.67J2 = -7536 (2)

Setting up the radiation balance for surface 1, we have

E b1  J1 J J J J
 1 21  1 31
1  1  1A1 A1F12  A1F13 

where E b1  T14  56700 W m 2 , so that

Prof. Hui Tong Chua 163


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
56700  J1 J1  J 2 J1  459
 
1  0.9 0.9A1 A1  0.39 A1  0.611
1

or

-10J1 + 0.39J2 = -510002


J1 = 51000 + 0.039J2 (3)

Substituting into (2),

0.26(51000 + 0. 039J2) – 1.67J2 = -7536

so that

J2 = 12528 W/m2.

Substituting into (1), the net heat transfer rate to the absorber is

1
q 2  7348  12528
1  0.5  77.7kW
0.5 15

8.3.3 The two-surface enclosure

A two-surface enclosure is the simplest example of radiation heat transfer where the
two surfaces just exchange radiation energy with each other (figure 8.8).

A2, T2, 2

q12

q1 = q12 = -q2
A1, T1, 1

1  1 1 1 2
Eb1 1 A 1 J1 A1F12 J2 2A2 Eb2

q1 -q2
E b1  J1 J 2  E b2
q1  q12  q2 
1  1  1A1 1   2   2 A 2
Figure 8.8 A two surface enclosure.

From figure (8.8), it can be observed that the total resistance to radiation heat transfer
between surfaces 1 and 2 is comprised of two surface resistances and one geometrical
resistance. Hence,

Prof. Hui Tong Chua 164


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

q12  q1  q 2 
 T14  T24   (8.15).
1  1 1 1 2
 
1A1 A1F12  2 A 2

This result is applicable to any two diffuse, gray surfaces which constitute an
enclosure.

8.3.4 Radiation shields

Radiation shields made from highly reflective (and therefore low emissivity) materials
can be used to attenuate the net radiation transfer between two surfaces. Referring to
figure 8.9, consider positioning a radiation shield between two large, parallel planes.

Radiation
shield

q1 q13 q32 -q2

A1, T1, 1 A2, T2, 2


3,1 3,2
A3, T3

1  1 1 1   3,1 1   3, 2 1 1 2
q1 1 A 1 A1F13  3,1A 3  3, 2 A 3 A 3 F32 2A2

Eb1 J1 J3,1 Eb3 J3,2 J2 Eb2


Figure 8.9 Radiation exchange between two surfaces in the presence of a
radiation shield.

Without a radiation shield, the radiation heat exchange between the two large parallel
planes can be inferred from (8.15) so that

q12 

A1 T14  T24 
1  1 1 1   2
 
1 1 2
(8.16).

A  T 4  T24
 1 1

1 1
 1
1  2

Prof. Hui Tong Chua 165


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
In the presence of a radiation shield and noting that F13 = F32 = 1, however, the heat
transfer is reduced to

q12 
A1 T14  T24  
1  1 1 1  3,1 1  3, 2 1 1   2
    
1 1 3,1  3, 2 1 2
(8.17).


A1 T14  T24 
1 1 1 1
   2
1 3,1 3, 2  2

Observe that the resistances offered by the radiation shield can be very large when 3,1
and 3,2 are very small.

Example 8.4
A cryogenic fluid flows through a long 20-mm diameter tube. The outer surface of
the tube is diffuse and gray with 1 = 0.02 and T1 = 77K. This tube is concentric with
a larger 50-mm diameter tube, the inner surface of which is diffuse and gray with 2 =
0.02 and T2 = 300K. The space between the two surfaces is maintained in vacuo.
Determine the heat gain by the cryogenic fluid per unit length of tubes. If a thin
radiation shield of 35-mm diameter and 2 = 0.02 (both sides) is inserted midway
between the inner and outer surfaces, ascertain the percentage change in heat gain per
unit length of the tubes.

T2 = 300K
D2 = 50mm
2 = 0.05

T1 = 77K
Shield
D1 = 20mm
D3 = 35mm
1 = 0.02
3 = 0.02
Without shield With shield

The thermal resistance circuit for the cryogenic system without shield is shown in
figure 8.8. The heat transfer rate can be found from

q12 

 T14  T24 
1  1 1 1  2
 
1A1 A1F12  2 A 2



 D1L T14  T24 
1  1 1 1   2  D1 
   
1 1  2  D 2 



 D1L T14  T24 
1 1   2  D1 
  
1  2  D 2 

Prof. Hui Tong Chua 166


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.

Hence,

q 5.67  108   0.02 77 4  3004
q  

L 1 1  0.05  0.02 
  
0.02 0.05  0.05 
 0.50 W m

The thermal resistance circuit for the cryogenic system equipped with a radiation
shield is shown in figure 8.9. The radiative heat transfer rate is

q

E b1  E b 2  T14  T24


R tot R tot

where

1  1 1  1  3  1 1  2
R tot    2   
1 D1L  D1L F12  3 D3L  D3L F32 2 D2 L
or

1  1  0.02 1  1  0.02  1 1  0.05 


R tot     2   
L  0.02  0.02   0.021  0.02  0.035   0.0351 0.05  0.05
1
 779.9  15.9  891.3  9.1  121.0
L
1817

L

Hence,

q  

q 5.67  108 774  3004 
 0.25 W m
L 1817

The percentage change in the heat gain is therefore

qw  qwo  0.25   0.50


 100   100  50%
qwo  0.50

8.3.5 The reradiating surface

The assumption of a reradiating surface is common to many applications. This


idealized surface is characterized by zero net radiation transfer (qi = 0). It is closely
approached by real surfaces that are well insulated on one side and for which
convection effects may be neglected on the opposite (radiating) side.

Prof. Hui Tong Chua 167


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
1
With qi = 0, equations (8.9, 8.11), viz. qi = Ai(Ji-Gi) and q i  E bi  J i  ,
1   i 
i Ai
4
implies that Gi = Ji = Ebi = Ti . This means that if the radiosity of a reradiating
surface is known, the temperature of this surface is also known. This temperature is
independent of the emissivity of the reradiating surface.

A three-surface enclosure, for which the third surface, surface R, is reradiating, is


shown below.

AR, TR, R

A1, T1, 1 A2, T2, 2

qR = 0

1  R
R AR
JR = EbR
1 1
A 1 F1R A 2 F2 R
1  1 1  2
1 A 1 q1R qR2 2A2
q1 -q2
J1 J2
1
A 1 F12

From the thermal network above, it is clear that

E b1  E b 2
q 1  q 2  (8.18)
1  1 1 1 2
 
1A1 A1F12  1 A1F1R  1 A 2 F2 R 1
2A2

Prof. Hui Tong Chua 168


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
1
Knowing q1 = -q2, equation (8.11), q i  E bi  J i  , may be applied to surface 1
1   i 
i Ai
and 2 to determine their radiosities J1 and J2. Then, knowing J1 and J2 and the
geometrical resistances, the radiosity of the reradiating surface J R may be determined
by q1R = qR2 so that

J1  J R J  J2
 R 0 (8.19)
1 A1F1R  1 A 2 F2R 
The temperature of the reradiating surface may then be determined from the
requirement that TR4  J R .

Example 8.5
A paint baking oven is made from a long, equilateral triangular duct which contains a
heated surface maintained at 1200K and an insulated surface. Painted panels,
maintained at 500K, form the third surface. The sides of the triangular duct are each
1m in width. The heated and insulated surfaces have an emissivity of 0.8. The
emissivity of the painted panels is 0.4. During steady-state operation, determine the
amount of power supplied to the heated side in order to maintain its temperature.
Ascertain the temperature of the insulated surface as well.

Equilateral
triangle
T1 = 1200K
R = 0.8, TR 1 = 0.8

T2 = 500K
2 = 0.4

W = 1m
qR = 0

JR = EbR
1 1
A 1 F1R A 2 F2 R
1  1 1  2
q12
1 A 1 2A2
q1 q2
J1 J2
1
A 1 F12

Prof. Hui Tong Chua 169


Adapted from FP Incropera, DP De Witt, Introduction to Heat Transfer, 2002, John
Wiley & Sons.
The system can be modeled as a three-surface enclosure with one surface reradiating.
The rate at which energy must be supplied to the heated surface can be obtained from
(8.18), or

E b1  E b 2
q1 
1  1 1 1  2
 
1A1 A1F12  1 A1F1R  1 A 2 F2 R 1
2A 2

From symmetry, F12 = F1R = F2R = 0.5. A1 = A2 = WL where L is the duct length.
Hence,

q1 
q1


5.67  108 12004  5004   37 kW m .
L 1  0.8  1

1  0.4
0.8  1 1  0.5  2  21 0.4  1

The temperature of the insulated surface can be obtained from the relation

JR = EbR,

where JR is obtainable from (8.19) and requires the knowledge of J1 and J2. Applying
(8.11) to surfaces 1 and 2, we have

1  1 1  0.8
J1  E b1  q1  5.67  108  12004   37000  108323 W m2
1W 0.8  1
1  2 1  0.4
J2  Eb2  q2  5.67  108  5004    37000  59043 W m2
2 W 0.4  1

From (8.19), it follows that

108323  J R J  59043
 R 0
1 W  L  0.5 1 W  L  0.5
Hence

J R  83683 W m2  E bR  TR4
14
 83683 
TR   8 
 1102K
 5.67  10 

Prof. Hui Tong Chua 170


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.

Chapter 9 Diffusion Mass Transfer

9.1 Physical origins and rate equations

The driving force of mass transfer is the difference in concentration and mass diffusion is
in the direction of decreasing concentration.

Mass diffusion happens in gases, liquids and solids. Diffusion occurs most
spontaneously in gases, followed by liquids and finally solids. Examples of diffusion
abound; for example nitrous oxide from a vehicle exhaust in air, dissolved oxygen in
water, and carbon in iron.

A mixture consists of minimum two species; the amount of any species i can be
quantified either by its mass density i (kg/m3) or molar concentration Ci (kmol/m3). The
relation between mass density and molar concentration is given by

i  Mi Ci (9.1)

where Mi is the species molecular weight in kg/kmol.

Accordingly the density of the mixture is given by

   i (9.2),
i

and the total number of moles per unit volume of the mixture is

C   Ci (9.3).
i

We can also quantify the amount of species i by its mass fraction

i
mi  (9.4)

or its mole fraction

Ci
xi  (9.4).
C

From (9.2) and (9.4), it follows that

Prof. Hui Tong Chua 171


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
 mi  1
i
(9.5)

and

x
i
i 1 (9.6).

For a mixture of ideal gases, the mass density and molar concentration of a particular
species can be related to its partial pressure:

pi
i  (9.7)
R iT

and

pi
Ci  (9.8)
RT

where Ri is the gas constant for species i and R is the universal gas constant. Combining
(9.4) and (9.8) and noting that

p   pi (9.9),
i

we have

Ci p i
xi   (9.10).
C p

The rate of mass diffusion is governed by the Fick’s law. For the transport of species A
in a binary mixture of A and B, it can be written as

jA  DABmA (9.11)

or
*
jA  CDABx A (9.12)
 
jA (kg/sm2) is the diffusive mass flux of species A whereas jA* (kmol/sm2) is the
diffusive molar flux of species A.

Prof. Hui Tong Chua 172


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
Assuming ideal gas behavior, DAB follows the following relationship with respect to
pressure and temperature expressed in Kelvin:

DAB ~ p 1T 3 2 (9.13).

This relation is useful for the estimation of DAB at pressures and temperatures for which
the data are unavailable.

For binary liquid solutions, we have to rely solely on empirical observations. Generally
speaking, for a dilute solution of A (solute) in B (solvent), DAB increases with increasing
temperature.

Example 9.1

Consider the diffusion of hydrogen (species A) in air, liquid water and iron (species B) at
T = 293K. Calculate both the molar and mass species fluxes when the concentration
dCA
gradient at a particular location is  1 kmol m3  m . Compare the value of the mass
dx
diffusivity to the thermal diffusivity. You can assume that the mole fraction of hydrogen,
xA, is much smaller than unity.

The mass diffusivity of hydrogen in air at 298K is DAB = 0.4110-4 m2/s. Using (9.13) at
T = 293K,

32 32
 T  4 293 
D AB,T  D AB, 298K    0.41 10    0.40  10  4 m 2 s .
 298   298 

The molar flux of hydrogen is dictated by the Fick’s law so that

* dx
jA  CD AB A .
dx

Since xA << 1, the total molar concentration C is approximately constant as hydrogen is a


very dilute species and hence

* dC
jA  D AB A .
dx

For the hydrogen-air mixture, we therefore have


*
jA  0.40 104 1  4 105 kmol m2  s .

The corresponding mass flux is found as

Prof. Hui Tong Chua 173


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
 
 
jA  M A jA*  2  4 105  8 105 kg s  m2 .

Since xA << 1, i.e. the hydrogen is a dilute species, the thermal properties of the air-
hydrogen mixture can be taken to be those of air. Hence the thermal diffusivity is
21.610-6 m2/s. The comparison between the mass diffusivity and thermal diffusivity is
best manifested through the use of the Lewis number which is

 21.6
Le    0.54 .
D AB 40

The results for all three mixtures are tabulated in the table below.

Species B 106 (m2/s) DAB106 (m2/s) Le 
jA 106 kg s  m2 
Air 21.6 40 0.54 80
Water 0.14 6.310-3 23 1310-3
Iron 23.1 26010-9 89106 0.5210-6

End of example

For a binary mixture of A and B with a constant total concentration and undergoing one
dimensional diffusion, equation (9.12) can be written for species A as

* dC
jA  D AB A (9.14).
dx

An equation of exactly the same form may be written for species B as

* dC
jB  D BA B (9.15).
dx

The condition of constant total concentration is particularly suitable to an idea gas


mixture at constant pressure. Hence,

dCA  C B 
 0 , or
dx

dC A dC
 B (9.16).
dx dx

Equation (9.16) can alternatively be expressed as

Prof. Hui Tong Chua 174


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
dCT x A dC x dx A dx
  T B or  B (9.16a), or
dx dx dx dx

d A M A d M dA d M
 B B or  B A (9.16b), or
dx dx dx dx M B

under isothermal and isobaric condition and from equation (9.8), we have

d pA R  T d p R T dp A dp
 B or  B (9.16c).
dx dx dx dx

A prerequisite for the pressure or molar concentration of the binary gas mixture to remain
constant is that there is no net diffusion of molecules, or that there is equimolecular
counterdiffusion, to wit
* *
jA  jB  0 .

This is only possible if DAB = DBA. Hence for brevity we can write equation (9.14) as

* dC
jA  D A (9.17).
dx

Equimolecular counterdiffusion also occurs in a distillation column when the molar latent
heats of the two components are the same. At any point in the column, a falling stream of
liquid contacts a rising stream of vapor and the system is not in equilibrium. The less
volatile component is transferred from the vapor to the liquid and the more volatile
component is transferred from the liquid to the vapor. If the molar latent heats of the two
components are equal, the condensation of the less volatile component releases exactly
the amount of latent heat necessary for the volatilization of an equal molar quantity of the
more volatile component. Thus at the interface, and throughout the liquid and vapor
phases, equimolecular counterdiffusion takes place.

Under these conditions, equation (9.17) can easily be integrated to yield

* CA  CA1
jA  D 2
x 2  x1

D
x 2  x1
 
CA1  CA 2  h D CA1  CA 2   (9.18), or

*
jA 
DCT
x 2  x1
  
x A1  x A 2  k x x A1  x A 2  (9.18a).

Prof. Hui Tong Chua 175


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
hD is known as the mass transfer coefficient with the driving force being molar
concentration difference, whereas kx is a mass transfer coefficient with the driving force
being molar fraction difference.

Under constant temperature and pressure, equation (9.17) can also be expressed as

* D p A 2  p A1
jA  
RT x 2  x1

D
  
p A  p A 2  kG p A1  p A 2
RTx 2  x1  1
 (9.18b).

k G is a mass transfer coefficient with the driving force being partial pressure difference.

Example 9.2

A simple rectifying column at atmospheric pressure consists of a tube arranged vertically


and supplied at the bottom with a mixture of benzene and toluene as vapor. At the top, a
condenser returns some of the product as a reflux which flows in a thin film down the
inner wall of the tube. The tube is well insulated. At one point in the column, the vapor
contains 70 mol% benzene and the adjacent liquid reflux contains 59 mol% benzene.
The temperature at this point is 365K. Assuming the diffusional resistance to vapor
transfer to be equivalent to the diffusional resistance of a stagnant vapor layer 0.2mm
thick, calculate the rate of interchange of benzene and toluene between vapor and liquid.
The molar latent heats of the two materials can be taken as equal. The vapor pressure of
toluene at 365K is 54.0 kN/m2 and the diffusivity of the vapors is 0.051 cm2/s.

Benzene (B) –toluene (T)


mixture

CB = 70 mol%
CB = 59 mol%

Since the latent heats are equal and the system is well insulated, there is no net change of
phase across the interface. This is an example of equimolecular counterdiffusion and the
molar flux of toluene, denoted with a subscript T, can be expressed as

Prof. Hui Tong Chua 176


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
*
jT 
D

p T  p T2
RTL 1

where L = 0.2 mm = 0.0002 m. Subscripts 1 and 2 refer to the liquid surface and vapor
side of the interface respectively.

According to the Raoult’s law, the partial pressure of toluene at the liquid surface is given
by

pT1  1  0.59 54  22.14 kN m2 .

The partial pressure of toluene in the vapor is simply given by

pT2  1  0.70101.3  30.39 kN m2 .

 0.051 10 4
For toluene, jT*  22.14  30.39  6.93 105 kmol m 2  s .
8.314  365  0.0002

The toluene flux is moving toward the liquid surface. For benzene,

p B1  101.3  22.14  79.16 kN m2 , p B2  101.3  30.39  70.91kN m2 .

 0.051104
Hence, jB*  79.16  70.91  6.93 105 kmol m2  s .
8.314  365  0.0002

The benzene flux is moving toward the vapor core. Hence the rate of interchange of
benzene and toluene is equal but opposite in direction.

End of example

9.2 Mass transfer in nonstationary media

If there is bulk motion in the medium, then just like heat transfer, mass transfer can also
occur by advection. We shall define the total or absolute flux of a species, which
includes both diffusive and advective components.

The absolute mass flux of species A in a binary mixture of A and B, n A , is related to the

species absolute velocity v A by
 
nA  A v A (9.19)

Prof. Hui Tong Chua 177


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.

v A can be associated with any point in the mixture, and is the average velocity of all the
A particles in a small element about the point. We can similarly write down the absolute
mass flux of species B in the mixture as
 
nB  B v B (9.20).

A mass average velocity for the mixture can then be obtained as


     
v  n  nA  nB  A vA  B v B (9.21),

so that
  
v  mA v A  mB v B (9.22).

Note that all the velocities are absolute velocities, and all the fluxes are absolute fluxes.
Namely they are referenced to a coordinate system that is fixed in space.

The mass flux of species A relative to the mixture mass average velocity is
  
jA  A v A  v (9.23).

 
While n A is the absolute mass flux of species A, jA is the relative or diffusive flux of the
species and relates directly to the Fick’s law, equation (9.11). From equations (9.19) and
(9.23), it is evident that
  
nA  jA  A v (9.24).

This highlights the two contributions to the absolute mass flux of species A: one due to
diffusion (i.e. due to the motion of A relative to the mass-average velocity) , and the other
due to advection (i.e. due to motion of A with the mass-average velocity of the mixture).
Applying equations (9.11) and (9.21), we obtain
  
nA  DABmA  mA nA  nB  (9.25).

We can write similar expressions for species B. The diffusive mass flux of B can be
written as
  
jB  B v B  v (9.26)

where jB  DBAmB . From equations (9.21), (9.23) and (9.26), it follows that
 
jA  jB  0 (9.27).

Prof. Hui Tong Chua 178


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.

Note that DBA = DAB = D. Hence the absolute mass flux of B can be expressed as
  
nB  DABmB  mB nA  nB  (9.28).

The foregoing expressions are expressed in mass fluxes, we can apply the same principle
and express in molar fluxes. Hence the absolute molar fluxes of A and B can be written
as
   
NA  CA v A and NB  CB v B (9.29).


The molar-average velocity for the mixture, v * , can be obtained from the following
consideration
     
N  NA  NB  Cv*  CA v A  CB v B (9.30),
so that
  
v*  x A v A  x B v B (9.31).

Note that molar average velocity should not be used in the various conservation equations.

The molar diffusive flux of A, jA* , stems from (9.12) and can also be expressed as
* 
 
jA  CA v A  v*  (9.32).

We accordingly have the following expressions as well


  
NA  jA*  CA v* (9.33), and

 
  
NA  CDABx A  x A NA  NB (9.34).

Similarly we also have


* *
jA  jB  0 (9.35).

9.2.1 Evaporation/desorption in a column

Prof. Hui Tong Chua 179


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
Binary gas
mixture A + B
xA,L , xB,L
x=L x=L
N A , x
xB xA
xA,0 , xB,0
x x=0
Liquid A xB = 1 xB = 0

Evaporation of liquid A into a binary gas mixture, A + B.

Referring to the figure above, fixed species concentrations, xA,L and xB,L are maintained
at the top of a tube containing a liquid layer of species A. The system is at constant
temperature and pressure. Since equilibrium prevails at the liquid surface, the vapor
concentration can be found from saturation condition. With x A,0 > xA,L, species A
evaporates from the liquid surface and diffuse upward. For steady, one-dimensional
conditions with no chemical reactions, species A cannot accumulate in the vapor space
within the column. The absolute molar flux of A must therefore be constant throughout
the column and hence

dNA , x
0 (9.36).
dx

Now throughout the column, C = CA + CB, xA + xB = 1. With xA,0 > xA,L, we must have
xB,0 < xB,L. Hence species B must diffuse from the top of the column to the liquid surface.
But if species B cannot be absorbed into liquid A, steady-state condition can only be
maintained if NB , x  0 everywhere within the column. This is only possible if the
downward diffusion of B is exactly nullified by the upward advection of B.

Noting that NB , x  0 , equation (9.34) simplifies to

dx A
NA , x  CD AB  x A NA , x (9.37).
dx

Rearranging, we obtain

CD AB dx A
NA , x   (9.38).
1  x A  dx

Prof. Hui Tong Chua 180


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
For constant pressure and temperature, C and DAB are also constant; equation (9.38) can
alternatively be expressed as

CD AB dC A
NA , x   (9.38a).
C B dx

This is known as the Stefan’s Law, where the bulk flow enhances the mass transfer rate
by a factor of C/CB, which is known as the drift factor.

Substituting equation (9.38) into equation (9.36), we have

d  1 dx A 
 0.
dx  1  x A dx 

For which we obtain

 ln 1  x A   C1x  C2 .

Applying xA(0) = xA,0 and xA(L) = xA,L, we arrive at

x L
1  x A  1  x A ,L 
 (9.39)
1  x A ,0  1  x A ,0 

Because 1 - xA = xB, we also have

x L
x B  x B,L 
 (9.40).
x B,0  x B,0 

To determine NA , x , we use equation (9.39) to obtain dxA/dx and substitute it into
equation (9.38) so that

CD AB  1  x A ,L 
NA , x  ln   (9.41).
L  1  x A,0 

Note that the aforementioned discussion can equally be applied to desorption condition
(e.g. desorption of water from lithium chloride solution), except that at the liquid surface
the vapor pressure is determined by Raoult’s law instead of saturation condition.

For constant pressure and temperature, equation (9.41) can be equally expressed as

Prof. Hui Tong Chua 181


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
CD AB  C B,L 
NA , x  ln  
(9.41a).
L  C B, 0 

Defining the logarithmic mean of CB,L and CB,0 as CBm as

C B, L  C B, 0
C Bm  , we have
C 
ln  B,L 
 C B, 0 

CD AB CB,L  CB,0 CD AB CA,0  CA,L


NA , x   (9.41b).
L CBm L CBm

CD AB
The factor h D  is the mass transfer coefficient. Equation (9.41b) can also be
C BmL
expressed as

pDAB p A,0  p A,L


NA ,x   k G p A,0  p A ,L  (9.41c), or
RTL p Bm

CD AB x A,0  x A,L
NA ,x   k x x A,0  x A,L  (9.41d).
L x Bm

Example 9.3

A water resistant sheet used for water proofing applications is made from an
impermeable polymeric material. Careful microstructural examination of the material
reveals that there are open pores of diameter D = 10 m that extend through the entire
thickness of the 100 m thick sheet. The pore diameter is sufficiently large to allow the
passage of water vapor but not liquid water through the sheet. Determine the rate at
which water vapor is transmitted through a single pore when saturated liquid rests on top
of the sheet while moist air at a relative humidity of  = 50% exists on the bottom of the
sheet. Evaluate the transfer rate at T = 298 K and a pressure of p = 1 atm. Compare the
transfer rates to rates that are predicted by neglecting the molar-average motion of the
mixture in the pore.

Prof. Hui Tong Chua 182


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
Liquid water
L = 100 m

Water
x resistant
material

D = 10m

Moist air,  = 50%

We assume that the pore is cylindrical and straight. At 298K, the saturation pressure of
water is pA,sat = 0.03165 bar, for water vapor-air mixture, DAB = 0.2610-4 m2/s.

The total transfer rate can be found from

D 2 CD AB  1  x A ,L 
N A , x  A poreNA , x  ln  
(E.1)
4 L  1  x A,0 

p 101.33
where C    40.9  10 3 kmol m3 .
RT 8.314  298

p A,sat0.03165
x A,0    31.23  10 3 , while
p 1.0133
 p 0.5  0.03165
x A,L   A,sat   15.62 103 .
p 1.0133

The evaporation rate per pore can therefore be evaluated according to equation (E.1) as

N A,x 

 10 106 2
40.9 103  0.26 104  1  15.62 103 
ln    13.4 1015 kmol s .
6 3 
4 100 10  1  31.23 10 

Neglecting the effects of bulk motion of the mixture,

dx A
NA , x  CD AB .
dx

Prof. Hui Tong Chua 183


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
Hence,

N A , x  A poreNA , x 
D 2 x  x A,L 
CD AB A , 0
4 L


 2
  10 10 6  40.9 10 3  0.26 10  4
 31.23  15.6210 3
6
4 100 10
15
 13.0 10 kmol s

You can therefore see that advection does augment the diffusive flux, resulting in an
increased evaporation rate.

Note that the molar average velocity may be neglected when the water vapor
concentration is small; particularly this applies when NB ,x  0 and x A  1 .

End of example

Returning now to equation (9.34), we see that for equimolecular counterdiffusion where
 
NA   NB , it reduces to the Fick’s law. For a system in which species B undergoes no

net transfer, so that NB  0 , it simplifies to the Stefan’s law.

9.2.2 Diffusivities of gases and vapors

Experimental values of diffusivities and the Schmidt numbers are given in the table
below for a number of gases and vapors in air at 298K and atmospheric pressure.

Substance D Sc = /D Substance D Sc = /D


(m2/s  106) (m2/s  106)
Ammonia 28.0 0.55 Valeric acid 6.7 2.31
Carbon 16.4 0.94 i-Caproic acid 6.0 2.58
dioxide
Hydrogen 71.0 0.22 Diethyl amine 10.5 1.47
Oxygen 20.6 0.75 Butyl amine 10.1 1.53
Water 25.6 0.60 Aniline 7.2 2.14
Carbon 10.7 1.45 Chlorobenzene 7.3 2.12
disulphide
Ethyl ether 9.3 1.66 Chlorotoluene 6.5 2.38
Methanol 15.9 0.97 Propyl 10.5 1.47
bromide
Ethanol 11.9 1.30 Propyl iodide 9.6 1.61
Propanol 10.0 1.55 Benzene 8.8 1.76

Prof. Hui Tong Chua 184


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
Butanol 9.0 1.72 Toluene 8.4 1.84
Pentanol 7.0 2.21 Xylene 7.1 2.18
Hexanol 5.9 2.60 Ethyl benzene 7.7 2.01
Formic acid 15.9 0.97 Propyl 5.9 2.62
benzene
Acetic acid 13.3 1.16 Diphenyl 6.8 2.28
Propionic 9.9 1.56 n-Octane 6.0 2.58
acid
i-Butyric 8.1 1.91 Mesitylene 6.7 2.31
acid
Note: The Schmidt numbers above are evaluated for mixtures composed largely of air.

9.3 Conservation of species for a stationary medium

As alluded to in the previous section, when the diffusion of a very small amount of
species A occurs within a stagnant species B, the molecular motion associated with the
mass transfer will not induce significant bulk motion of the medium. This is very
common in the diffusion of a dilute gas or liquid within a stagnant liquid or a solid host
medium. In these cases, the medium can be assumed to be stationary, and advection can
be neglected.

For such situations, the diffusive mass and molar fluxes are identical to the absolute mass
and molar fluxes, so that
 
nA  jA  DABmA (9.42)
 
NA  jA*  CDABx A (9.43)

Note that the total density, , or the total concentration, C, is approximately that of the
host medium, species B. With this important assumption of a stationary medium, direct
analogy between conduction heat transfer and diffusion mass transfer can be applied.

The conservation of mass for a species A can be written as

dm A
 A,in  m
m  A ,g  m
 A,out  m
 A,st (9.44)
dt

Species A generation, m  A ,g , happens when there are chemical reactions within the
control volume. Take methane cracking, CH4 → C + 2H2, for example, there is a net
generation of carbon and hydrogen and simultaneously a net depletion of methane.

Prof. Hui Tong Chua 185


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
We shall now develop the mass diffusion equation. We consider a binary mixture of
species A and B in which mass transfer occurs solely by diffusion and there is no bulk
fluid motion or advection.

z
y
x

n A ,z dz n A , y dy

dz

m
 A ,g
n A , x n A , x dx
m
 A ,st

n A , y
dy
dx

n A ,z
Fig 9.1 Differential control volume for a mass diffusion analysis

The mass diffusion rates at the opposite surfaces can then be expressed as

nA , x
nA , x dx  nA , x  dx (9.45a)
x
nA , y
nA , y dy  nA , y  dy (9.45b)
y
nA ,z
nA ,z dz  nA ,z  dz (9.45c)
z

Within the material medium, there may also be a material source term, such as that due to
a chemical reaction within the control volume. This term can be represented as

Prof. Hui Tong Chua 186


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.

 A,g  n A dxdydz
m (9.46)

where n A is the rate of increase of the mass of species A per unit volume (kg/sm3).
Additionally, the mass of species A stored in the control volume may also change with
time and can be expressed as

A
 A,st 
m dxdydz (9.47).
t

Considering the overall mass balance of species A in the control volume, we have

nA , x dydz  nA , y dxdz  nA ,z dxdy  n A dxdydz


A ,
 nA , x dxdydz  nA , y dydxdz  nA ,z dzdxdy  dxdydz
t

Substituting from (9.45) we eventually have

nA , x nA , y nA ,z A


    n A 
x y z t

From (9.11) and substituting the x, y and z components of jA , we have

  mA    mA    mA  


 D AB    D AB    D AB   n A  A (9.48a).
x  x  y  y  z  z  t

In terms of the molar concentration, we similarly have

  x    x    x   CA
 CD AB A    CD AB A    CD AB A   N A  (9.48b).
x  x  y  y  z  z  t

We shall focus on cases where both DAB and  are constant so that

 2A  2A  2A n A 1 A


    (9.49a),
x 2
y 2
z 2
D AB D AB t

similarly we focus on cases where both DAB and C are constant so that

 2CA  2CA  2CA N  1 CA


   A
 (9.49b).
x 2
y 2
z 2
D AB D AB t

Prof. Hui Tong Chua 187


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.

Note the similarity with the thermal diffusion equation.

The species diffusion equation can also be expressed in cylindrical and spherical
coordinate systems. In the cylindrical coordinate system, it is given as

1 x  1   x    x   CA
 CD ABr A   2  CD AB A    CD AB A   N A  (9.50).
r r  r  r     z  z  t

In the spherical coordinate system, it is expressed as

1  2 x A  1   x  1   x   CA
 CD ABr  2 2  CD AB A   2  CD AB sin  A   N A 
r r 
2
r  r sin      r sin      t
(9.51).

9.4 Stationary media with specified surface concentrations

Let us take a look at the one-dimensional diffusion of species A through a planar medium
of A and B (figure 9.2).

x
*
jA , x

A and B
xA,s1 xA,s2

L
Figure 9.2 Mass transfer in a stationary
planar medium.

For steady-state conditions without chemical reactions, (9.48b) reduces to

d  dx 
 CD AB A   0 (9.52).
dx  dx 

Assume the total molar concentrations and the diffusion coefficient to be constant, (9.52)
together with the boundary conditions in figure 9.2 can be solved to obtain

Prof. Hui Tong Chua 188


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.

x
x A x   x A,s 2  x A,s1   x A,s1 (9.53)
L

From (9.12), the molar flux is given as

x  x A,s1 
NA , x  CD AB  A,s 2  (9.54)
 L 

To obtain the molar flow rate and expressing it in terms of species concentration, we have

D ABA
N A,x  CA,s1  CA,s 2  (9.55)
L

Thence we can define a resistance to species transfer by diffusion in a planar medium as

C A,s1  CA ,s 2 L
R m,dif   (9.56)
N A,x D ABA

Once again marvel at the glaring analogy between heat and mass transfer by diffusion.

This analogy also applies to cylindrical and spherical systems. For one-dimensional
steady diffusion in a cylindrical, non-reacting system, (9.50) simplifies to

1d dx 
 CD ABr A   0 (9.57a).
r dr  dr 

The species concentration profile and the species diffusion resistance are accordingly
expressed as

x A ,s1  x A ,s 2 r
x A r   ln    x A ,s 2 (9.57b)
ln r1 r2   r2 

and

ln r2 r1 
R m,dif  (9.57c),
2LD AB

respectively.

For a spherical medium, we similarly have

Prof. Hui Tong Chua 189


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
d 2 dx A 
 CD ABr 0 (9.58a),
dr  dr 

x A ,s1  x A ,s 2  1 1 
x A r       x A ,s 2 (9.58b),
1 r1  1 r2  r r2 

and

1 1 1
R m,dif     (9.58c).
4D AB  r1 r2 

Example 9.4
The effectiveness of some pharmaceutical products is compromised by extended
exposure to humidity. For such products, blister packaging is employed to prevent direct
exposure to water vapor.

Let us consider tablets packed in a blister package that is comprised of a flat lidding
sheeting and a formed sheet with troughs to contain the individual tablet. The thickness
of the polymeric formed sheet is L = 50m. Each trough has a diameter D = 5mm and
depth h = 3mm. The lidding sheet is made of aluminum foil. The binary diffusion
coefficient for water vapor in the polymer is DAB = 610-14 m2/s. The aluminum foil is
impermeable to water vapor. The molar concentrations of water vapor at the outer and
inner surfaces of the polymer are CA,s1 = 4.510-3 kmol/m3 and CA,s2 = 0.510-3 kmol/m3,
respectively. Determine the rate of water vapor diffusion through the trough wall to the
tablet.

Aluminum lidding
Formed sheet
Polymer
Tablet h sheet
A
Lidding A
sheet
D
THK = L
Blister package Section A-A

The diffusion rate of water vapor through the polymer sheet can be written as

CA,s1  CA,s 2   DAB  D  Dh CA,s1  CA,s 2  .


2
D ABA
N A,x 
L L  4 

Thence

Prof. Hui Tong Chua 190


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.

N A,x 
6 10 14   5 10 3  
2

   5 10 3  3 10 3 4.5  0.510 3
50 10 6  4 

 0.32 10 15 kmol s

The shelf life of the pharmaceutical product is inversely proportional to the diffusion rate
of water vapor into the blister package.
End of example

9.5 Boundary conditions and discontinuous concentrations at


interfaces

We have already seen the following type of boundary condition earlier:

x A 0, t   x A,s (9.59)

In the abovementioned case, the surface concentration at x = 0 is constant.

The second boundary condition in analogy with constant surface heat flux condition is
constant species flux at the surface whereby

x A
 CD AB  j*A ,s (9.60).
x x 0

A special case of this boundary condition pertains to an impermeable surface so that

x A
 0.
x x 0

Mass diffusion, in contrast with heat diffusion, has an added element of difficulty though.
The species concentrations at the interface between two materials are typically
discontinuous, whereas temperature is continuous. A simple example to consider is a
pool of volatile liquid oil exposed to air. Clearly the mole fraction of oil in the liquid is
unity while the mole fraction of the oil vapor is less than unity.

9.5.1 Evaporation and sublimation

For evaporation and sublimation, the concentration of species A in the gas phase at the
interface (at x = 0) can be deduced from the Raoult’s law, so that

Prof. Hui Tong Chua 191


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
p A 0  x A 0p A,sat (9.61)

where pA is the partial pressure of A in the gas phase, xA is the mole fraction of species A
in the liquid or solid phase, and pA,sat is the saturation pressure of species A at the surface
temperature.

Gas phase with


species A

x NA , x
pA(0)
xA(0)

Liquid or solid concentrated in


species A

Figure 9.3 Evaporation or sublimation of


species A from a liquid or solid into a gas.

Raoult’s law is valid when the gas phase can be considered to be close to ideal and the
liquid or solid phase is concentrated in species A.

We have earlier dealt with the special case of Raoult’s law where the liquid or solid phase
is a pure species A so that xA(0) = 1 and pA(0) = pA,sat.

9.5.2 Solubility of gases in liquids and solids

We shall now consider the transfer of species A from a gas stream into a liquid or solid,
species B.

Prof. Hui Tong Chua 192


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
Gas phase with
species A

x NA , x
pA(0)
xA(0)

Liquid or solid, species B,


dilute in species A

Figure 9.4 Transfer of weakly soluble


species A from a gas to a liquid or solid.

If species A is only weakly soluble in a liquid so that x A is small, Henry’s law can be
employed to relate the mole fraction of A in the liquid to the partial pressure of A in the
gas phase outside the liquid:

p A 0
x A 0  (9.62).
H
H is the Henry’s constant. While H depends on temperature, it is quite insensitive to
pressure up and until 5 bar.

In the case of the gas and solid interaction, we can relate the concentration of the gas in
the solid at the interface through the solubility, S:

CA 0  Sp A 0 (9.63).

pA(0) is the partial pressure of the gas adjoining the interface. C A(0) has the units of
kmol of A/m3 of solid and the units of S is kmol of A/m3 of solid/bar partial pressure of A.
Solubility values are typically presented in units of m3 of species A (at STP, Standard
Temperature and Pressure, of 0°C and 1atm) per m3 of solid per atm partial pressure of A.
~
We denote this solubility as S . Since at STP one kilomole occupies 22.414m3, the
~
relationship between S and S is
~

S  S 22.414m3 kmol . 
Note that additional conversion between bar and atm may be needed.

Example 9.5
Gaseous helium is stored at 20°C in a spherical container of fused silica (SiO2). The
container has a diameter of 0.20m and its wall is 2-mm thick. The initial storage pressure
is 4 bar. Determine the rate at which the pressure decreases with time.

Prof. Hui Tong Chua 193


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.

D = 0.2m L = 2mm

A,2
dM A A,1
 A - helium
dt 
B - fused silica
Helium 
M A ,out
pA = 4 bar
T = 20°C

From the requirement of the conservation of species, we have


M 
A , out  M A , st .

Since the reduction in pressure is due to the diffusion of helium through the container
wall,

 
M A , out  n A , x A .

Hence,

 dA V 
 nA , x A  M A , st  .
dt

pA
A  M A CA and C A  .
RT

Hence we have

 p 
d M A A V 
d A V  dM A C A V  RT 
  
dt dt dt

dp A RT
 nA , x A
dt MAV

Now

dm A d
nA , x  D AB  D AB A
dx dx

Prof. Hui Tong Chua 194


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.

For our condition,

A ,1  A, 2
nA , x  D AB
L

The species densities A,1 and A,2 relate to quantities within the fused silica at its inner
and outer surfaces, respectively. Since A  M A CA ,

A,1  M ASp A,i  M ASp A and A, 2  M ASp A,o  0 .

pa,i and pA,o are helium pressures at the inner and outer surfaces, respectively. Hence

M ASp A
nA , x  D AB .
L

Thence

dp A RT M Sp
 D AB A A A
dt MAV L
RTADABS
 pA
LV

Since A = D2 and V = D3/6

dp A 6RTDABS
 pA
dt LD

Now for fused silica at 293K, DAB = 0.410-13 m2/s and S = 0.4510-3 kmol/m3bar. R =
0.08314m3bar/kmolK. Hence

dp A

 
60.08314293 0.4 1013 0.45 10 3
4

dt 0.002  0.2
 2.63 10 8 bar s

This is the initial and maximum leak rate for the system. As the pressure within the
container decreases, the leak rate will decrease.

Example 9.6
Hydrogen gas is kept at 3 bar and 1 bar on the two sides of a polymeric membrane. The
membrane is 0.3-mm thick. The temperature of the system is 25°C and the binary
diffusion coefficient of hydrogen in the polymer is 8.710-8 m2/s. The solubility of

Prof. Hui Tong Chua 195


Adapted from (1) FP Incropera, DP De Witt, TL Bergman, AS Lavine, Fundamentals of
Heat and Mass Transfer, 2007, John Wiley & Sons and (2) JM Coulson, JF Richardson,
JR Backhurst, JH Harker, Chemical Engineering Volume 1: Fluid Flow, Heat Transfer
and Mass Transfer, 2005, Elsevier Butterworth-Heinemann.
hydrogen in the polymer is 1.510-3 kmol/m3bar. Determine the mass diffusive flux of
hydrogen through the membrane.

 A  hydrogen
 B  plastic


 D AB  8.7 10 8 m 2 s
x SAB  1.5 10 3 kmol m 3  bar

Hydrogen Hydrogen
CA,1 CA,2
pA,1 = 3 bar pA,2 = 1 bar

CA,s1 CA,s2

0.3mm
x A,s1  x A,s 2 CA,s1  CA,s 2
Now NA , x  CD AB  D AB
L L

Invoking (9.32) we have

CA,s1  1.5 103  3  4.5 103 kmol m3


CA,s 2  1.5 103 1  1.5 103 kmol m3

Thence

4.5 103  1.5 103


NA ,x  8.7 108 3
 8.7 107 kmol s  m 2 .
0.3 10

Since the molecular weight of hydrogen is 2 kg/kmol, we accordingly have

nA ,x  8.7 107  2  1.74 106 kg s  m2 .

Prof. Hui Tong Chua 196


Tutorial questions and solutions

Prof. Hui Tong Chua 197


School of Mechanical and Chemical Engineering
The University of Western Australia
Heat and Mass Transfer (ENSC3007)

Course lecturer and Unit coordinator: Prof. Hui Tong Chua


E-mail id: huitong.chua@uwa.edu.au, Tel no.: 64881828

Adapted from FP Incropera and DP DeWitt, Fundamentals of heat and mass transfer, 5th
ed., John Wiley & Sons, 2002, TL Bergman, AS Lavine, FP Incropera, DP DeWitt,
Fundamentals of heat and mass transfer, 7th ed., John Wiley & Sons, 2011.

Tutorial 1

1. In a two dimensional body shown below, the temperature gradient at surface A is


T T T
found to be  30K / m . What are and at surface B?
y y x

Insulation
1m B, TB = 100°C

k = 10W/mK

y
2m

A, TA = 0°C x

2. A composite rod consists of two different materials, A and B, each of length 0.5L.

T1 T1 < T 2 T2

A B

x 0.5L L

198
The thermal conductivity of material A is half that of material B, so that kA/kB =
0.5. Sketch the steady-state temperature and heat flux distributions, T(x) and qx ,
respectively. Assume constant properties and no internal heat generation in either
material.

3 The steady-state temperature distribution in a one dimensional wall of thermal


conductivity 50 W/mK and thickness 50mm is observed to be T(°C) = 200 -2000
x2, where x is in meters.
(a) What is the heat generation rate q in the wall?
(b) Determine the heat fluxes at the two wall faces. In what manner are these heat
fluxes related to the heat generation rate?

199
200
201
202
School of Mechanical and Chemical Engineering
The University of Western Australia
Heat and Mass Transfer (ENSC3007)

Course lecturer and Unit coordinator: Prof. Hui Tong Chua


E-mail id: huitong.chua@uwa.edu.au, Tel no.: 64881828

Adapted from FP Incropera and DP DeWitt, Fundamentals of heat and mass transfer, 5th
ed., John Wiley & Sons, 2002.

Tutorial 2

1. A measurement technique for convective heat transfer coefficient involves


securing one side of a thin metallic foil to an insulating material and subjecting
the other side of the foil to the fluid flow conditions of interest to the
experimentalists.

T, h  , Ts)
Foil ( Pelec

Tb
Foam insulation (k)

 ,
By passing an electric current through the metallic foil, a uniform heat flux, Pelec
is generated within the foil. The thickness, L, and the thermal conductivity, k, of
the insulation is known. In addition, the temperatures of the free stream fluid (T),
foil (Ts), and the insulation (Tb) are measured and therefore known.

  2000 W m 2 , L =
For the particular conditions at hand, T = Tb = 25°C, Pelec
10mm, and k = 0.040 W/mK.

(a) When water flows over the foil surface, the foil temperature reads Ts = 27°C.
Determine the convection coefficient. What error would one incur if the
thermal power generated by the foil is assumed to be dissipated purely by
convection to the flowing water?

(b) When air flows over the foil surface, the foil temperature rises to Ts = 125°C.
What is the convection coefficient? The foil has an emissivity of 0.15 and is
in a large surrounding with a temperature of 25°C. What error would one
incur if the thermal power generated by the foil is assumed to be dissipated
purely by convection to the air?

203
(c) Heat flux gages are commonly operated at a constant temperature, Ts. In
which case, the power expenditure provides a direct measurement of the
 as a function of ho for
convection coefficient. For Ts = 27°C, plot Pelec
10  h o  1000 W m 2  K . What is the effect of ho on the error related to
omitting conduction through the insulator?

2. Approximately 106 discrete electrical components can be placed on a single


integrated circuit (chip). The attendant power dissipation can be 30000 W/m 2.
The thin chip is exposed to a dielectric liquid at its outer surface, so that ho =
1000W/m2K and T,o = 20°C. The chip’s inner surface is attached to a circuit
board. The thermal contact resistance between the chip and the board is 10-4
m2K/W. The board thickness and thermal conductivity are Lb = 5 mm and kb = 1
W/mK respectively. The outer surface of the board is subjected to flowing
ambient air with hi = 40 W/m2K and T,i = 20°C.

Dielectric
fluid Chip ( q c , Tc)
T,o, ho

Lb Thermal contact resistance


R t,c

Board (kb)
Air
T,i, hi

(a) Sketch the equivalent thermal circuit.


(b) Under steady-state conditions, what is the chip temperature?
(c) The maximum allowable heat flux, q c,m , is constrained by the requirement
that the chip temperature will not exceed 85°C. If air is used instead of the
dielectric fluid, so that ho = 100 W/m2K, what is the corresponding q c,m ?

3. Superheated steam at 575°C is directed from a boiler to the turbine of an electric


power plant via steel tubes (k = 35 W/mK) with an inner diameter of 300mm and
a wall thickness of 30mm. To reduce heat loss to the surroundings and to
maintain a safe-to-touch outer surface temperature, calcium silicate insulation is
applied around the tubes which is in turn externally wrapped by a thin aluminum
sheet with an emissivity  = 0.20. The air and wall temperatures of the power
plant are 27°C.

204
(a) Assuming that the inner surface of a steel tube corresponds to that of
the steam and an outer convection coefficient of 6 W/m2K for the
aluminum sheet, determine the minimum insulation thickness so that
the aluminum temperature is not more than 50°C.
(b) Determine the corresponding heat loss per unit length of the tube.

4. An electric cable of radius r1 and thermal conductivity kc is enclosed by an


insulating sleeve whose outer surface is of radius r2 and experiences convection
heat transfer and radiation exchange with the adjoining air and large surroundings,
respectively. When electric current passes through the cable, thermal energy is
generated within the cable at a volumetric rate q .

Tsur
Electric cable
Ambient air
T, h Insulation

Ts,1

r1 Ts,2

r2
(a) Write down the steady-state forms of the heat diffusion equation for
the insulation and the cable.

Verify that these equations are satisfied by the following temperature


distributions:

ln r r2 
Insulation: Tr   Ts, 2  Ts,1  Ts, 2 
ln r1 r2 
q r12  r 2 
Cable: Tr   Ts,1  1  
4k c  r12 
Sketch the temperature distribution, T(r), in the cable and the sleeve,
labeling key features.

(b) Applying Fourier’s law, show that the rate of conduction heat transfer
per unit length through the sleeve may be expressed as

2k s Ts,1  Ts, 2 


q r 
ln r2 r1 

205
Applying an energy balance to a control surface placed around the
cable, obtain an alternative expression for q r , expressing your results
in terms of q and r1.

(c) Applying an energy balance to a control surface placed around the


outer surface of the sleeve, obtain an expression from which Ts,2 may
be determined as a function of q , r1, h, T,  and Tsur.

(d) Consider conditions for which 250A are passing through a cable
having an electric resistance per unit length of Re’ = 0.005 /m, a
radius of r1 = 15mm, and a thermal conductivity of kc = 200 W/mK.
For ks = 0.15 W/mK, r2 = 15.5mm, h = 25 W/m2K,  = 0.9, T =
25°C and Tsur = 35°C, evaluate the surface temperatures Ts,1 and Ts,2
as well as the temperature To at the centerline of the cable.

206
207
208
209
210
211
212
213
214
215
School of Mechanical and Chemical Engineering
The University of Western Australia
Heat and Mass Transfer (ENSC3007)

Course lecturer and Unit coordinator: Prof. Hui Tong Chua


E-mail id: huitong.chua@uwa.edu.au, Tel no.: 64881828

Adapted from FP Incropera and DP DeWitt, Fundamentals of heat and mass transfer, 5th
ed., John Wiley & Sons, 2002, and from TL Bergman, AS Lavine, FP Incropera and DP
DeWitt, Fundamentals of Heat and Mass Transfer, 7th ed., John Wiley & Sons, 2011

Tutorial 3

1. A quartz window with a thickness of L is used as a viewing glass in a furnace for


annealing steel. The inner surface (x = 0) of the window is irradiated with a
uniform heat flux q o . This is attributed to the radiative emission from the hot
gases in the furnace. A fraction, , of this radiation is absorbed at the inner
surface, while the rest is partially absorbed by the quartz material. The resultant
volumetric heat absorption can be quantified as

q x   1  qoe x

where  is the absorption coefficient of the quartz. Convection heat transfer


exists at the outer surface of the viewing glass (x = L) with an ambient air
temperature of T and a convection coefficient of h. Convection and radiation
emission from the inner surface as well as the radiation emission from the outer
surface can be omitted. Determine the temperature distribution in the quartz
window in terms of those aforementioned parameters.

2. A bonding operation employs a laser to bond a thin adhesive-backed plastic film


to a metal strip as shown below. The thickness of the metal strip, d, is 1.25mm
and its width, w2, is much bigger than that of the plastic film, w1. For the metal
strip,  = 7850 kg/m3, cp = 435 J/kg.K, and k = 60 W/mK. The thermal
resistance of the plastic film with a width w1 = 40mm is negligible. The upper
and lower surfaces of the strip (inclusive of the film) undergo convective heat
transfer with the surrounding air at 25°C and a convective coefficient of 10
W/m2K. The strip and film are very long in the direction normal to the page.
The edges of the metal strip are at the air temperature, T.

(a) Derive an expression for the temperature distribution in the portion of the
steel strip with the plastic film (-w1/2  x  w1/2).
(b) If the laser heat flux is 10000 W/m2, evaluate the temperature of the
plastic film at the center (x = 0) and its edges (x = w1/2).
(c) Plot the temperature distribution for the entire strip and highlight its
salient features.

216
Laser source, qo

Plastic film
T, h
Metal strip w1
d
T, h x
w2

3. Finned passages are very frequently employed in the cores of compact heat
exchangers. A notable application of such compact heat exchangers is in
electronic cooling, where one or more air-cooled stacks are located between heat
generating electronic devices. Consider a stack with 50 fins as shown below.

100mm Aluminum

400K

12mm

350K

1mm 3mm

Air 200mm
300 K, 150W/m2K

What is the maximum possible power dissipation as a result of using such a stack?

4. A pin fin of uniform, cross-sectional area is fabricated from an aluminum alloy (k


= 160 W/mK). The fin diameter is D = 4 mm, and the fin is exposed to
convective conditions characterized by h = 220 W/m2K. The fin efficiency is f
= 0.65. Determine the fin length L and the fin effectiveness f. Account for tip
convection.
5. An annular aluminum (k = 160 W/mK) fin of rectangular profile is attached to a
circular tube having an outside diameter of 25 mm and a surface temperature of
250°C. The fin is 1 mm thick and 10 mm long, and the temperature and the
convection coefficient associated with the adjoining fluid are 25°C and 25
W/m2K, respectively. What is the heat loss per fin? If 200 such fins are spaced

217
at 5 mm increments along the tube length, what is the heat loss per meter of tube
length?
Fig 2.15 Fin efficiency of annular fins of rectangular profile

100

90

80

70
Fin efficiency (%)

60

50

1 = r2c/r1
40
t
2
30
3
r2c = r2 + t/2
Lc = L + t/2
20 5
L
r1
10 r2

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5
mLc or (2h/kt)1/2Lc

218
219
220
221
222
223
224
225
School of Mechanical and Chemical Engineering
The University of Western Australia
Heat and Mass Transfer (ENSC3007)

Course lecturer and Unit coordinator: Prof. Hui Tong Chua


E-mail id: huitong.chua@uwa.edu.au, Tel no.: 64881828

Adapted from FP Incropera and DP DeWitt, Fundamentals of heat and mass transfer, 5th
ed., John Wiley & Sons, 2002.

Tutorial 4

1. Experiments have confirmed that, for airflow at T = 35°C and u,1 = 100 m/s, the
rate of heat transfer from a turbine blade with a characteristic length L1 = 0.15m
and a surface temperature Ts,1 = 300°C is q1 = 1500 W. Determine the heat
transfer rate from a second turbine blade with a characteristic length L2 = 0.3m
operating at Ts,2 = 400°C in airflow at T = 35°C and u,2 = 50 m/s. The surface
area of the turbine blade can be safely assumed to be directly proportional to its
characteristic length.

2. To prevent ice formation on the wings of a small, private aircraft, one viable
method is to embed electric heating elements within the wings. To determine the
typical power requirements, consider a representative flight conditions for which
the plane flies at 100m/s in air which is at a temperature of -23°C (k = 0.022
W/mK, Pr = 0.72, and  = 16.310-6 m2/s). If the characteristic length of the
airfoil is L = 2m and wind tunnel measurements yield an average friction
coefficient of Cf  0.0025 for the nominal conditions, determine the heat flux
required to maintain a surface temperature of Ts = 5°C.

3. An experiment conducted to evaluate the local convection heat transfer coefficient


for uniform flow normal to a heated circular disk with a constant surface
temperature has engendered the following radial Nusselt number distribution

hr D  r 
n

Nu D   Nuo 1  a  
k   ro  

where both n and a are positive. The Nusselt number at the stagnation point, Nuo,
where r = 0, can be correlated as

hr  0D
Nuo   0.814 Re1D2 Pr 0.36
k

226
L

V, ro
T D

Ts

Obtain the average Nusselt number, Nu D  h D k , which corresponds to


convection heat transfer from an isothermal disk. You have learnt in the lectures
that boundary layer development from a stagnation point generates a decreasing
convective heat transfer coefficient profile with respect to the distance from the
same stagnation point. Offer a reasonable explanation for the observed opposite
trend in this experiment.

227
228
229
230
School of Mechanical and Chemical Engineering
The University of Western Australia
Heat and Mass Transfer (ENSC3007)

Course lecturer and Unit coordinator: Prof. Hui Tong Chua


E-mail id: huitong.chua@uwa.edu.au, Tel no.: 64881828

Adapted from FP Incropera and DP DeWitt, Fundamentals of heat and mass transfer, 5th
ed., John Wiley & Sons, 2002; and Incropera, De Witt, Bergman and Lavine,
Fundamentals of heat and mass transfer, 6th ed., John Wiley & Sons, 2007.

Tutorial 5

1. Consider the effect of wind blowing past the penthouse tower on a tall building.
The tower length in the wind direction is 8m and there are 8 window panels.

Wind

(a) Calculate the average convection coefficient for the first, third, and eighth
window panels when the wind speed is 5m/s. Use a film temperature of
300K to evaluate the thermophysical properties required in the appropriate
correlation.

(b) For the first, third, and eighth windows, on one graph, plot the variation of
the average convection coefficient with wind speed for the range 5  u 
100 km/h.

2. The roof of a refrigerated truck compartment is made of a composite material


consisting of a layer of foamed urethane insulation (t2 = 50mm, ki = 0.026 W/mK)

231
sandwiched between aluminum alloy panels (t1 = 5mm, kp = 180 W/mK). The
length and width of the roof are L = 10m and W = 3.5m respectively. The
temperature of the inner surface of the roof is Ts,i = -10°C. Consider conditions
where the truck travels at V = 105 km/h, the air temperature is T = 32°C, and the
solar irradiation is Gs = 750 W/m2. Turbulent flow can be assumed to prevail
over the entire roof.

(a) For equivalent values of the solar absorptivity and the emissivity of the
outer surface (s =  = 0.5), estimate the average temperature Ts,o of the
outer surface. What is the corresponding heat load imposed on the
refrigeration system?
(b) A special finish (s = 0.15,  = 0.8) may be applied to the outer surface.
What effect would such an application have on the surface temperature
and the heat load?
(c) If with s =  = 0.5, the roof is not insulated so that t2 = 0, what are the
corresponding values of the surface temperature and the heat load?

3. Hot water at 50°C is routed from one building in which it is generated to another
building where it is used for space heating. Transfer between the buildings is
done with a steel pipe (k = 60 W/mK) of 100-mm outer diameter and 8-mm wall
thickness. During the winter, the air is at a representative temperature of -5°C and
the cross-flow velocity V = 3 m/s over the pipe.

(a) If the cost of producing the hot water is $0.05 per kWh, what is the
representative daily cost of heat loss from an uninsulated pipe to the air
per meter of pipe length? The convection resistance associated with water
flow in the pipe may be neglected.

(b) Determine the savings associated with application of a 10-mm-thick


urethane insulation (k = 0.026 W/mK) to the outer surface of the pipe.

4. A flat plate coated with a volatile substance (species A) is exposed to dry,


atmospheric air in parallel flow with T = 20°C and u = 8m/s. The plate is
maintained at a constant temperature of 134°C by a resistive heater while the
substance evaporates. The plate has a width of 0.25m and is well insulated on the
bottom.

232
Air, u, T

Coating, A,s, Ts

L = 4m
Heater is uniformly embedded
The molecular weight and the latent heat of vaporization of species A are M A =
150 kg/kmol and hfg = 5.44106 J/kg, respectively. The mass diffusivity is DAB =
7.7510-7 m2/s. The saturated vapor pressure of the substance is 0.12 bar at
134°C. Determine the electrical power required to maintain steady-state
conditions.

5. A person installed a circular outdoor spa bath in the open area. In a typical
operating condition, when the spa water is maintained at 37°C, and when the
ambient air has a temperature of 17°C with a relative humidity of 30%, he has
to top up the water at a rate of 0.002 kg/s so as to maintain a fixed water level
in the bath. The tub is well insulated on all sides and its bottom, and the
makeup water enters at 17°C.

a) Determine the value of hmA that commensurate with the makeup water flow
rate. The symbols have their usual meaning.

b) Determine the extent of convective heat loss from the spa bath.

c) Determine the electrical power needed to maintain the spa bath temperature
at 37°C.

At 300K, air = 1.161 kg/m3, cp,air = 1.007 kJ/kgK,  = 22.5  10-6 m2/s.
At 310K, water vapour,sat = 0.0436 kg/m3, hfg = 2414 kJ/kg.
At 290K, water vapour,sat = 0.0143 kg/m3.
DAB = 26  10-6 m2/s, cp,water = 4.2 kJ/kgK

233
234
235
236
237
238
239
240
241
242
243
School of Mechanical and Chemical Engineering
The University of Western Australia
Heat and Mass Transfer (ENSC3007)

Course lecturer and Unit coordinator: Prof. Hui Tong Chua


E-mail id: huitong.chua@uwa.edu.au, Tel no.: 64881828

Adapted from FP Incropera and DP DeWitt, Fundamentals of heat and mass transfer, 5th
ed., John Wiley & Sons, 2002.

Tutorial 6

1. An air heater for an industrial application consists of an insulated, concentric tube


annulus. Air flows through a thin-walled inner tube while saturated steam flows
through the outer annulus. Condensation of the steam maintains a uniform
temperature Ts on the tube surface.

L
Tm,o, po

Ts

Air
Tm,i, pi, m

Saturated Do Insulation
steam, psat Di
The air enters a 50mm diameter tube at a pressure of 5atm, a temperature of Tm,i =
17°C and a flow rate of m   0.03 kg s , while saturated steam at 2.455 bars
condenses on the outer surface of the tube. If the length of the annulus is L = 5m,
what are the outlet temperature Tm,o and pressure po of the air? What is the mass
rate at which condensate leaves the annulus?

2. Exhaust gases from a wire processing oven are discharged into a tall stack, and
the gas and stack surface temperatures at the stack outlet must be estimated.
Knowledge of the outlet gas temperature Tm,o is useful for predicting the
dispersion of effluents in the thermal plume, while knowledge of the outlet stack
surface temperature Ts,o reflects whether the gas products will condense or not.
The thin-walled, cylindrical stack is 0.5m in diameter and 6.0m in height. The
exhaust gas flows at 0.5 kg/s and its inlet temperature into the stack is 600°C.

For an ambient air temperature of 4°C and a wind velocity of 5 m/s, estimate the
gas outlet and stack surface temperatures. The gas products can be approximated
as atmospheric air.

244
Stack Height, 6m, diameter 0.5m

Stack
base
Oven Oven
exhaust
gases

3. A pharmaceutical flowing at 0.3 m/s through a straight thin-walled stainless


steel tube of 12.7 mm diameter is sterilized by heating it from 25°C to 75°C.
A uniform heat flux is maintained by an electric resistance heater wrapped
around the outer surface of the tube.

(i) If the tube is 15 m long, what is the required heat flux?

(ii) If the fluid enters the tube with a fully developed velocity profile and a
uniform temperature profile, what is the surface temperature at the tube
exit?

(iii) What is the corresponding fluid temperature at a distance of 1m from the


entrance?

(iv) What is the corresponding tube surface temperature at a distance of 1m


from the entrance?

The following fluid properties can be used.


 = 1000 kg/m3, cp = 4000 J/kgK,  = 210-3 kg/sm, k = 0.8 W/mK, and Pr = 10

245
Figure 5.8 Local Nusselt numbers for laminar flow in a circular tube

20

18

16
Constant surface heat
14 flux
Thermal entry length
12 Combined entry length (Pr = 0.7)
NuD

10

4
Constant surface temperature
2

0
0.001 0.01 0.1 1

Gz-1 = x/(D.ReD.Pr)

246
247
248
249
250
251
252
253
School of Mechanical and Chemical Engineering
The University of Western Australia
Heat and Mass Transfer (ENSC3007)

Course lecturer and Unit coordinator: Prof. Hui Tong Chua


E-mail id: huitong.chua@uwa.edu.au, Tel no.: 64881828

Adapted from FP Incropera and DP DeWitt, Fundamentals of heat and mass transfer, 5th
ed., John Wiley & Sons, 2002.

Tutorial 7

1. A packed bed of solid spheres is commonly employed by thermal energy storage


systems. If energy is charged into the system, one simply passes a stream of hot
gas through the packed bed. Conversely, if energy requires to be taken from the
system, one just passes a stream of cold gas through the same packed bed.

Thermal energy storage system

Sphere
Gas , cp, k, Ti
Tg, i, h

Packed bed

A thermal energy storage system is packed with 75 mm-diameter aluminum


spheres ( = 2700 kg/m3, cp = 950 J/kgK, k = 240 W/mK). In one charging
process, gas enters the system at a temperature Tg,i = 300°C. If the initial
temperature of the spheres is Ti = 25°C and the convection coefficient h = 75
W/m2K, how long does it take a sphere near the inlet of the storage system to
accumulate 90% of the maximum possible internal energy? What is the affiliated
temperature of the sphere? Is there anything to gain from replacing the aluminum
spheres with copper spheres?

2. A spherical reactor for synthesizing pharmaceuticals has a 5 mm-thick stainless


steel wall (k = 17 W/mK) and an inner diameter of Di = 1.0m. During production,
the vessel is charged with reactants with  = 1100 kg/m3 and cp = 2400 J/kgK.
The ensuing exothermic reactions releases energy at q = 104 W/m3. As a first
approximation, the mixture is assumed to be well stirred and the thermal
capacitance of the vessel is neglected.

254
(a) The exterior of the vessel is subject to ambient air with T = 25°C and h =
6 W/m2K. If the initial temperature of the reactants is 25°C, determine
the temperature of the mixture after an elapsed time of five hours.
(b) Investigate the effect of the convection coefficient on the transient thermal
behavior of the mixture.

3. A major factor for defects in electronic modules pertains to thermal cycling stress
due to intermittent heating and cooling. To fix ideas, in circuit cards possessing
active and passive elements made with material of different thermal expansion
coefficients, thermal stresses are the primary culprit leading to failure in
component joints, such as soldered and wired connections. While fatigue failure
resulting from numerous thermal excursions during the product lifespan is a major
general concern, it is still possible to diagnose defective joints by conducting
accelerated thermal stress tests before a product release. In which cases, the
thermal stress test has to be rapid to minimize production disruption.

A manufacturer wishes to develop a rapid thermal cycling apparatus to test those


circuit cards by subjecting them to forced convection characterized by
Nu L  C Re0L.8 Pr 0.33 . Air (k = 0.026 W/mK,  = 210-5 m2/s, Pr = 0.71) and a
dielectric liquid (k = 0.064 W/mK,  = 110-6 m2/s, Pr = 25) are possible
candidates as working fluid. Assuming equivalent air and liquid velocities and
validity of the lumped capacitance model for the components, evaluate the ratio of
the thermal time constants for the two fluids. Which fluid provides the faster
thermal response?

4a. What is the physical significance of the Biot number?

4b. Consider heat transfer between two identical hot solid bodies and their
environments. The first solid is dropped in a large container filled with water,
while the second one is allowed to cool naturally in the air, which is at more
or less the same temperature at the water. For which solid is the lumped
system analysis more likely to be applicable? Justify your answer.

4c. In a manufacturing facility, 4 cm diameter brass balls (k = 111 W/mK,  =


8522 kg/m3, and cp = 385 J/kgK) initially at 130°C are quenched in a water
bath at 40°C for a period of 2 min at a rate of 110 balls per minute. The
convection heat transfer coefficient is 238 W/m2K.

(i) Determine the temperature of the balls at the end of the quenching and

(ii) Clearly show that the amount of energy removed from a brass ball after
  hA s 
time t can be expressed as c p VTi  T 1  exp   t  , where all
  c V 
 p 
the symbols have their usual meaning.

255
(iii) Determine the rate at which heat needs to be removed from the water in
order to maintain it constant at 40°C.

256
257
258
259
260
261
262
School of Mechanical and Chemical Engineering
The University of Western Australia
Heat and Mass Transfer (ENSC3007)

Course lecturer and Unit coordinator: Prof. Hui Tong Chua


E-mail id: huitong.chua@uwa.edu.au, Tel no.: 64881828

Adapted from FP Incropera and DP DeWitt, Fundamentals of heat and mass transfer, 5th
ed., John Wiley & Sons, 2002, and TL Bergman, AS Lavine, FP Incropera and DP
DeWitt, Fundamentals of heat and mass transfer, 7th ed., John Wiley & Sons, 2011.

Tutorial 8

1. A zirconia-based ceramic has the spectral, hemispherical emissivity shown


below and it is considered to be used as the filament of a light bulb.

0.8

0.6
el

0.4

0.2

0
0 0.4 0.7 1 2 3 4
l (mm)

In comparison, the spectral, hemispherical emissivity of tungsten is presented


below.

263
1

0.8

0.6

el = 0.45
el

0.4

0.2
el = 0.10

0
0 1 2 3 4
l (mm)

(a) What is the total, hemispherical emissivity of a zirconia filament


operating at 3000K?

(b) What is the total, hemispherical emissivity of a tungsten filament


operating at 3000K?

(c) For zirconia and tungsten filament operating at 3000K in an evacuated


bulb, which filament consumes the larger power?

(d) In terms of the production of visible radiation, which filament is more


efficient? Quantitatively explain.

2. A radiant oven for drying newsprint comprises a long duct (L = 20 m) of


semicircular cross section. The newsprint moves through the oven on a conveyor
belt at a velocity of V = 0.2 m/s. The newsprint has a water content of 0.02 kg/m 2
as it enters the oven and has to be completely dry as it leaves. For quality
assurance, the newsprint must be maintained at 300 K during drying. All systems
components and the air flowing through the oven are therefore maintained at 300
K. The inner surface of the semicircular duct has an emissivity of 0.8 and a
temperature T1. It provides the radiant heat required for the drying. The wet

264
surface of the newsprint can be taken to be black. The air throughout the oven
has a temperature of 300 K and a relative humidity of 20%. It is made completely
turbulent before the entrance to promote evaporation.

T = 300 K
T1, e1 = 0.8  = 0.20 Wet
u newsprint
T2, e2 = 1.0

V = 0.2 m/s

L = 20 m
W=1m
Dry V = 0.2 m/s
newsprint
For steady state operation,

(a) determine the required evaporation rate for the newsprint,


(b) determine the required air velocity u, and
(c) determine the corresponding temperature T1 of the semicircular duct so as
to meet the specifications.

DAB for the water vapour-air system is 2610-6 m2/s. Pr of air = 0.707, hfg of water
= 2437.2 kJ/kg, k of air = 0.02624 W/mK,  of air = 1.56810-5 m2/s, A,sat of
water vapour at 300K = 0.0256kg/m3.

3. A wet towel hangs on a clothes line. One surface receives solar irradiation of
Gs = 900 W/m2, while both surfaces receive atmospheric radiation, Gsky = 200
W/m2 and ground radiation, Gground = 250 W/m2. The prevailing conditions
are such that air flow at 27°C and a relative humidity of 60% maintains a
convection heat transfer coefficient of 20 W/m2K at both surfaces. The wet
towel has an emissivity of 0.96 and a solar absorptivity (s) of 0.65. All the
properties of the atmospheric air can be evaluated at 300 K.

Note that solar radiation is concentrated in the short wavelength region of the
spectrum and surface emission is at much longer wavelengths. Hence we
typically cannot approximate most surfaces as gray in regards to solar
irradiation, and the solar absorptivity is therefore given. For the absorption of
atmospheric and ground radiations, the surface concerned can be treated to be
gray.

The steady state towel surface temperature turns out to be 298K.

265
(i) Determine the corresponding mass transfer coefficient for the towel at
298K.

(ii) Write down the energy balance equation for this problem and express it in
terms of the towel surface temperature.

(iii) Verify that a towel surface temperature of 298 K will balance the energy
to within 4 W/m2. (Note that for this problem 0°C = 273K)

Thermophysical properties of air at 300K

 = 1.16 kg/m3, cp = 1007 J/kgK,  = 0.22510-4 m2/s

Properties of water vapour

A,sat, 300K = 0.0256 kg/m3, A,sat, 298K = 0.0226 kg/m3

Binary species diffusivity between water vapour and air = 0.2610-4 m2/s

Latent heat of vaporization at 298K = 2.442106 J/kg

4. The energy flux associated with solar radiation incident on the outer surface of the
earth’s atmosphere has been accurately measured and is known to be 1368 W/m 2.
The diameters of the sun and earth are 1.39109 m and 1.27107 m, respectively.
The distance between the sun and the earth is 1.51011 m.

(a) What is the emissive power of the sun?


(b) Approximating the sun’s surface as black, what is its temperature?
(c) At what wavelength is the spectral emissive power of the sun a maximum?
(d) Assuming the earth’s surface to be black and the sun to be the only source of
energy for the earth, estimate the earth’s surface temperature.

5. The spectral emissivity of an opaque, diffuse surface is as shown.

266
0.9

0.8

0.7

0.6

0.5
el

0.4

0.3

0.2

0.1

0
0 1 2 3 4 5 6 7 8
l (mm)

(a) If the surface is maintained at 1000 K, what is the total, hemispherical


emissivity?

(b) What is the total, hemispherical absorptivity of the surface when


irradiated by large surroundings with a temperature of 1500K?

(c) What is the radiosity of the surface ewhen it is maintained at 1000K and
subjected to the irradiation prescribed in part (b)?

(d) Determine the net radiation flux into the surface for the conditions of part
(c).

6. Consider the attic of a home. The floor of the attic has a width of 10 m, while the
roof makes an angle of 30° from the horizontal direction, as shown below. The
owner intends to reduce heat load to the home by applying bright aluminium foil
(ef = 0.07) onto the bottom of the attic roof. Prior to installation of the foil, all the
surfaces are of emissivity eo = 0.85.

A2, e2, T2

30°
Attic

A1, e1, T1 L1 = 10 m

267
Determine the ratio of the radiation heat transfer after the foil installation to
before the foil installation.

268
269
270
271
272
273
274
275
276
School of Mechanical Engineering
The University of Western Australia
Heat and Mass Transfer (ENSC3007)

Course lecturer and Unit coordinator: Prof. Hui Tong Chua


E-mail id: huitong.chua@uwa.edu.au, Tel no.: 64881828

Adapted from FP Incropera, DP DeWitt, TL Bergman and AS Lavine, Fundamentals of


heat and mass transfer, 6th ed., John Wiley & Sons, 2007, and from JM Coulson, JF
Richardson, JR Backhurst and JH Harker, Chemical Engineering Volume 1: Fluid flow,
heat transfer and mass transfer, Elsevier Butterworth Heinemann, 1999.

Tutorial 9

1. An open pan of diameter 0.2 m and height 80 mm (above the water at 27°C) is
exposed to ambient air at 27°C and 25% relative humidity. Determine the
evaporation rate, assuming that only mass diffusion occurs. Determine the
evaporation rate, considering bulk motion.

2. Ammonia gas is diffusing at a constant rate through a layer of stagnant air 1 mm


thick. Conditions are such that the gas contains 50% by volume ammonia at one
boundary of the stagnant layer. The ammonia diffusing to the other boundary is
quickly absorbed and the concentration is negligible at that plane. The
temperature is 295 K and the pressure atmospheric, and under these conditions the
diffusivity of ammonia in air is 1.810-5 m2/s. Estimate the rate of diffusion of
ammonia through the layer.

3. By what percentage would the rate of absorption be increased or decreased by


increasing the total pressure from 100 to 200 kN/m2 in the following cases?

a) The absorption of ammonia from a mixture of ammonia and air containing 10%
of ammonia by volume, using pure water as solvent. Assume that all the
resistant to mass transfer lies within the gas phase.

b) The same condition as (a) but the absorbing solution exerts a partial vapor
pressure of ammonia of 5 kN/m2.

The diffusivity can be assumed to be inversely proportional to the absolute


pressure.

4. Gaseous hydrogen at 10 bars and 27 °C is stored in a 100 mm diameter spherical


tank having a steel wall 2 mm thick. The molar concentrations of hydrogen in the
steel are 1.50 kmol/m3 at the inner surface and negligible at the outer surface.
The diffusion coefficient of hydrogen in steel is approximately 0.310-12 m2/s.
What is the initial rate of mass loss of hydrogen by diffusion through the tank
wall? What is the initial rate of pressure drop within the tank?

277
5. Oxygen gas is maintained at pressures of 2 bars and 1 bar on opposite sides of a
rubber membrane that is 0.5 mm thick, and the entire system is at 25 °C. What’s
the molar diffusive flux of oxygen through the membrane? What are the molar
concentrations of oxygen on both sides of the membrane (outside the rubber)?
The solubility of oxygen in rubber at 298 K is 3.1210-3 kmol/m3bar.

6. The presence of noncondensables, such as air, can cause a very significant


reduction in the heat rejection rate to a water-cooled steam condenser surface,
such as found in a lithium bromide-water absorption chiller. When the chiller
is well maintained, so that there is a negligible amount of noncondensable
within the system, the rate of condensation on the condenser surface is 0.020
kg/m2s, and the condenser surface temperature maintains at 20°C, while the
condensate surface temperature is at 28°C. In this case, the free stream
vapour pressure is the saturated pressure of steam at 28°C. When the
operating budget is tight, and exacerbated by the lack of a highly qualified
team of technicians, stagnant air accumulates in the condenser, so that while
the condenser surface temperature remains at 20°C, the condensate surface
temperature reduces to 24°C and the rate of condensation plummets to 0.010
kg/m2s. The total free stream pressure remains the same as before, namely
being equal to the saturated pressure of steam at 28°C.

psteam,sat(28°C) = 0.03767 bar, psteam,sat(24°C) = 0.02983 bar, Dair,steam(25°C, 1


bar) = 0.26  10-4 m2/s.

a) Qualitatively, in the case with stagnant air accumulation, briefly describe


the trend of the steam partial pressure from the free stream to the condensate
surface, as well as the trend of the partial pressure of air from the free stream
to the condensate surface. Explain your trend.

b) Determine the partial pressure of air at the condensate surface.

c) Write down the governing equation for the molar flux of steam toward the
condenser surface.

d) Based on (c), show that the expression for the partial pressure of air as a
function of distance from the condensate surface can be expressed as

 RT 
p air , y  p air ,0  exp  Nsteam
 ,y y  , where the symbols have their usual
 pDair ,steam 
meaning.

e) Work out the partial pressure of air at 5 mm from the condensate surface.

7. An apparatus to measure the diffusion coefficient of vapour-gas mixtures is


composed of a vertical and narrow cylindrical column which contains the

278
liquid that evaporates into the gas flowing over the mouth of the column. The
gas flow rate is sufficient to maintain a negligible vapour concentration at the
mouth. The height of the column, from the liquid surface to the mouth, is 200
mm high. The pressure and temperature of the environment are 0.3 atm and
310 K, respectively. For a test with water and air under the aforementioned
conditions,

(i) determine the prevailing binary species diffusivity, DAB, for the problem,

(ii) determine the molar fractions of water vapour at (a) the liquid-air
interface, and (b) the mouth of the column, and

(iii) calculate the expected evaporation rate (kg/hm2)

Binary species diffusivity at 298 K and 1 atm = 0.2610-4 m2/s

Saturation pressure of water at 310 K = 0.06221 bar

Molecular weight of water = 18 g/mol

279
280
281
282
283
284
285
286
287

Вам также может понравиться