You are on page 1of 240

EVERYDAY HEAT

TRANSFER
PROBLEMS
Sensitivities To
Governing Variables

by M. Kemal Atesmen

2009 by ASME, Three Park Avenue, New York, NY 10016, USA (www.asme.org)
All rights reserved. Printed in the United States of America. Except as permitted
under the United States Copyright Act of 1976, no part of this publication may be
reproduced or distributed in any form or by any means, or stored in a database or
retrieval system, without the prior written permission of the publisher.
INFORMATION CONTAINED IN THIS WORK HAS BEEN OBTAINED BY THE
AMERICAN SOCIETY OF MECHANICAL ENGINEERS FROM SOURCES BELIEVED
TO BE RELIABLE. HOWEVER, NEITHER ASME NOR ITS AUTHORS OR EDITORS
GUARANTEE THE ACCURACY OR COMPLETENESS OF ANY INFORMATION
PUBLISHED IN THIS WORK. NEITHER ASME NOR ITS AUTHORS AND EDITORS
SHALL BE RESPONSIBLE FOR ANY ERRORS, OMISSIONS, OR DAMAGES ARISING
OUT OF THE USE OF THIS INFORMATION. THE WORK IS PUBLISHED WITH
THE UNDERSTANDING THAT ASME AND ITS AUTHORS AND EDITORS ARE
SUPPLYING INFORMATION BUT ARE NOT ATTEMPTING TO RENDER
ENGINEERING OR OTHER PROFESSIONAL SERVICES. IF SUCH ENGINEERING
OR PROFESSIONAL SERVICES ARE REQUIRED, THE ASSISTANCE OF AN
APPROPRIATE PROFESSIONAL SHOULD BE SOUGHT.
ASME shall not be responsible for statements or opinions advanced in papers
or ... printed in its publications (B7.1.3). Statement from the Bylaws.
For authorization to photocopy material for internal or personal use under those
circumstances not falling within the fair use provisions of the Copyright Act, contact
the Copyright Clearance Center (CCC), 222 Rosewood Drive, Danvers, MA 01923,
tel: 978-750-8400, www.copyright.com.

Library of Congress Cataloging-in-Publication Data


Atesmen, M. Kemal.
Everyday heat transfer problems : sensitivities to governing variables / by M. Kemal
Atesmen.
p. cm.
Includes bibliographical references and index.
ISBN 978-0-7918-0283-0
1. HeatTransmissionProblems, exercises, etc. 2. MaterialsThermal properties
Problems, exercises, etc. 3. Thermal conductivityProblems, exercises, etc.
4. Engineering mathematicsProblems, exercises, etc. I. Title.
TA418.54.A47 2009
621.4022dc22

2008047423

TABLE OF CONTENTS
Introduction ................................................................................................... 1
Chapter 1 Heat Loss from Walls in a Typical House ............................. 5
Chapter 2 Conduction Heat Transfer in a Printed Circuit Board .... 13
Chapter 3 Heat Transfer from Combustion Chamber Walls.............. 25
Chapter 4 Heat Transfer from a Human Body During
Solar Tanning ............................................................................ 33
Chapter 5 Efficiency of Rectangular Fins .............................................. 41
Chapter 6 Heat Transfer from a Hot Drawn Bar .................................. 51
Chapter 7 Maximum Current in an Open-Air Electrical Wire .......... 65
Chapter 8 Evaporation of Liquid Nitrogen in a
Cryogenic Bottle ...................................................................... 77
Chapter 9 Thermal Stress in a Pipe ........................................................ 85
Chapter 10 Heat Transfer in a Pipe with Uniform Heat
Generation in its Walls ......................................................... 93
Chapter 11 Heat Transfer in an Active Infrared Sensor .................. 103
Chapter 12 Cooling of a Chip ................................................................. 113

iii

Everyday Heat Transfer Problems


Chapter 13 Cooling of a Chip Utilizing a Heat Sink
with Rectangular Fins ......................................................... 121
Chapter 14 Heat Transfer Analysis for Cooking in a Pot ................ 131
Chapter 15 Insulating a Water Pipe from Freezing ........................... 139
Chapter 16 Quenching of Steel Balls in Air Flow .............................. 147
Chapter 17 Quenching of Steel Balls in Oil......................................... 155
Chapter 18 Cooking Time for Turkey in an Oven .............................. 161
Chapter 19 Heat Generated in Pipe Flows due to Friction ............. 169
Chapter 20 Sizing an Active Solar Collector for a Pool ................... 179
Chapter 21 Heat Transfer in a Heat Exchanger ................................. 195
Chapter 22 Ice Formation on a Lake .................................................... 203
Chapter 23 Solidification in a Casting Mold ....................................... 213
Chapter 24 Average Temperature Rise in Sliding Surfaces
in Contact .............................................................................. 221
References .................................................................................................... 233
Index .............................................................................................................. 235

iv

I NTRODUCTION

Everyday engineering problems in heat transfer can be very


complicated and may require solutions using nite element or nite
difference techniques in transient mode and in multiple dimensions.
These engineering problems might cover conduction, convection and
radiation energy transfer mechanisms. The thermophysical properties
that govern a particular heat transfer problem can be challenging to
discover, to say the least.
Some of the standard thermophysical properties needed to
solve a heat transfer problem are density, specic heat at constant
pressure, thermal conductivity, viscosity, volumetric thermal
expansion coefcient, heat of vaporization, surface tension, emissivity,
absorptivity, and transmissivity. These thermophysical properties
can be strong functions of temperature, pressure, surface roughness,
wavelength and other properties. in the region of interest.
Once a heat transfer problem's assumptions are made, equations
set up and boundary conditions determined, one should investigate
the sensitivities of desired outputs to all the governing independent
variables. Since these sensitivities are mostly non-linear, one should

Everyday Heat Transfer Problems

analyze them in the region of interest. The results of such sensitivity


analyses will provide important information as to which independent
variables should be researched thoroughly, determined accurately,
and focused on. The sensitivity analysis will also provide insight into
uncertainty analysis for the dependent variable, (Reference S. J. Kline
and F. A. McClintock [9]). If the dependent variable y is dened as a
function of independent variables x1, x2, x3, xn as follows:
y = f(x1, x2, x3, xn)
then the uncertainty U for the dependent variable can be
written as:
U = [(y/x1 u1)2 + (y/x2 u2)2 + (y/x3 u3)2 + + (y/xn un)2]0.5
where y/x1, y/x2, y/x3, , y/xn are the sensitivities of the
dependent variable to each independent variable and u1, u2, u3, ,
un are the uncertainties in each independent variable for a desired
condence limit.
In this book, I will provide sensitivity analyses to well-known
everyday heat transfer problems, determining y/x1, y/x2,
y/x3, , y/xn for each case. The analysis for each problem will
narrow the eld of independent variables that should be focused
on during the design process. Since most heat transfer problems
are non-linear, the results presented here would be applicable only
in the region of values assumed for independent variables. For the
uncertainties of independent variablesfor example, experimental
measurements of thermophysical propertiesthe reader can nd the
appropriate uncertainty value for a desired condence limit within
existing literature on the topic.
Each chapter will analyze a different one-dimensional heat transfer
problem. These problems will vary from determining the maximum
allowable current in an open-air electrical wire to cooking a turkey
in a convection oven. The equations and boundary conditions for
each problem will be provided, but the focus will be on the sensitivity
of the governing dependant variable on the changing independent

Introduction

variables. For the derivation of the fundamental heat transfer


equations and for insight into the appropriate boundary conditions,
the reader should refer to the heat transfer fundamentals books listed
in the references.
Problems in Chapters 1 through 6 deal with steady-state and
one-dimensional heat transfer mechanisms in rectangular coordinates.
Chapters 7 through 10 deal with steady-state and one-dimensional
heat transfer mechanisms in cylindrical coordinates. Unsteady-state
problems in one-dimensional rectangular coordinates will be tackled
in Chapters 11 through 14, cylindrical coordinates in Chapter 15, and
spherical coordinates in Chapters 16 through 18.
The following six chapters are allocated to special heat transfer
problems. Chapters 19 and 20 deal with momentum, mass and heat
transfer analogies used to solve the problems. Chapter 21 analyzes
a counterow heat exchanger using the log mean temperature
difference method. Chapters 22 and 23 solve heat transfer problems
of ice formation and solidication with moving boundary conditions.
Chapter 24 analyzes the problem of frictional heating of materials in
contact with moving sources of heat.
I would like to thank my engineering colleagues G. W. Hodge,
A. Z. Basbuyuk, E. O. Atesmen, and S. S. Tukel for reviewing some of
the chapters. I would also like to dedicate this book to my excellent
teachers and mentors in heat transfer at several universities and
organizations. Some of the names at the top of a long list are
Prof. W. M. Kays, Prof. A. L. London, Prof. R. D. Haberstroh,
Prof. L. V. Baldwin, and Prof. T. N. Veziroglu.
M. Kemal Atesmen
Ph. D. Mechanical Engineering
Santa Barbara, California

CHAPTER

HEAT LOSS

FROM WALLS
IN A TYPICAL HOUSE

eat loss from the vertical walls of a house is analyzed under


steady-state conditions. Walls are assumed to be large and
built in a planar fashion, so that one-dimensional heat transfer
rate equations in rectangular coordinates may be used, and only
conduction and convection heat transfer mechanisms are considered.
In this analysis, radiation heat transfer effects are neglected. No air
leakage through the wall was assumed. Also, the wall material thermal
conductivities are assumed to be independent of temperature in the
region of operation.
Assuming winter conditionsthe temperature inside the house is
higher than the temperature outside the housethe convection heat
transferred from the inside of the house to the inner surface of the
inner wall is:
Q/A = hin (Tin Tinner wall inside surface)

(1-1)

Most walls are constructed from three types of materials: inner wall
board, insulation and outer wall board. The heat transfer from these
wall layers will occur by conduction, and is presented by the following
rate Eqs., (1-2) through (1-4):

Everyday Heat Transfer Problems

Q/A = (kinner wall/tinner wall) (Tinner wall inside surface Tinner wall outside surface) (1-2)
Q/A = (kinsulation/tinsulation) (Tinner wall outside surface Touter wall inside surface) (1-3)
Q/A = (kouter wall/touter wall) (Touter wall inside surface Touter wall outside surface) (1-4)
The heat transfer from the outer surface of the outer wall to the
atmosphere is by convection and can be expressed by the following
rate Eq. (1-5):
Q/A = hout (Tout Touter wall outer surface)

(1-5)

Eliminating all the wall temperatures from Eqs. (1-1) through (1-5),
the heat loss from a house wall can be rewritten as:
Q/A = (Tin Tout)/[(1/hin) + (tinner wall/kinner wall) + (tinsulation/kinsulation)
(1-6)
+ (touter wall/kouter wall) + (1/hout)]
The denominator in Eq. (1-6) represents all the thermal resistances
between the inside of the house and the atmosphere, and they are
in series.
In the construction industry, wall materials are rated with their
R-value, namely the thermal conduction resistance of one-inch
material. R-value dimensions are given as (hr-ft2-F/BTU)(1/in). The
sensitivity analysis will be done in the English system of units rather
than the International System (SI units). The governing Eq. (1-6)
for heat loss from a house wall can be rewritten in terms of R-values
as follows:
Q/A = (Tin Tout)/[(1/hin) + Rinner wall tinner wall + Rinsulation tinsulation
+ Router wall touter wall + (1/hout)]

(1-7)

where the denitions of the variables with their assumed nominal


values for the present sensitivity analysis are given as:
Q/A = heat loss through the wall due to convection and conduction
in Btu/hr-ft2

Heat Loss From Walls In A Typical House

Tin = 68F (inside temperature)


Tout = 32F (outside temperature)
hin = 5 BTU/hr-ft2-F (inside convection heat transfer coefcient)
Rinner wall = 0.85 hr-ft2-F/BTU-in (wall inside board R-value)
tinner wall = 1 in (wall inside board thickness)
Rinsulation = 3.5 hr-ft2-F/BTU-in (insulation layer R-value)
tinsulation = 4 in (insulation layer thickness)
Router wall = 5 hr-ft2-F/BTU-in (wall outside board R-value)
touter wall = 1 in (wall outside board thickness)
hout = 10 BTU/hr-ft2-F (outside convection heat transfer coefcient).
The heat loss through a wall due to changes in convection heat
transfer is presented in Figures 1-1 and 1-2. Changes in the convection
heat transfer coefcient affect the heat loss mainly in the natural
convection regime. As the convection heat transfer coefcient increases
into the forced convection regime, heat loss value asymptotes.
Resistances from both inside and outside convection heat transfer are
too small to cause any change in heat loss through the wall.
The heat loss through a wall due to changes in insulation material
R-value is presented in Figures 1-3 and 1-4. Higher R-value insulation

1.8

Q/A, Btu/hr-ft2

1.79
1.78
1.77
1.76
1.75

10

15

20

25

Outside Convection Heat Transfer Coefficient, Btu/hr-ft2-F

Figure 1-1 Wall heat loss versus outside convection heat transfer coefcient

Everyday Heat Transfer Problems

Q/A, Btu/hr-ft2

1.8
1.78
1.76
1.74
1.72

Inside Convection Heat Transfer Coefficient, Btu/hr-ft2-F

Figure 1-2 Wall heat loss versus inside convection heat transfer coefcient

material is denitely the way to go, depending upon the cost and
benet analysis results. The thickness of the insulation material is also
very crucial. Thicker insulation material is denitely the best choice,
depending upon the cost and benet analysis results.

Q/A, Btu/hr-ft2

2
1.8
1.6
1.4
1.2

3.5

Insulation "R" Value,

4.5
hr-ft2-F/Btu-in

Figure 1-3 Wall heat loss versus insulation R-value

Heat Loss From Walls In A Typical House

Q/A, Btu/hr-ft2

4
6
8
Insulation Thickness, in

10

12

Figure 1-4 Wall heat loss versus insulation thickness

The effects on heat loss of inner and outer wall board R-values
and thicknesses are similar to the effects of insulation R-value and
thickness, but to a lesser extent. Sensitivities of heat loss to all the
governing variables around the nominal values given above will be
analyzed later.
Sensitivity of heat loss to the outside convection heat transfer
coefcient can be determined in a closed form by differentiating the
heat loss Eq. (1-7) with respect to hout:
(Q/A)/hout = (Tin Tout)/{h2out [(1/hin) + Rinner wall tinner wall
+ Rinsulation tinsulation + Router wall touter wall + (1/hout)]2}

(1-8)

Sensitivity of heat loss to the outside convection heat transfer


coefcient is given in Figure 1-5. Similar sensitivity is experienced
for the inside convection heat transfer coefcient. The sensitivity
of heat loss to the convection heat transfer coefcient is high in
the natural convection regime, and it diminishes in the forced
convection regime.
Sensitivities of heat loss to insulation material R-value and
insulation thickness are given in Figures 1-6 and 1-7 respectively.

Everyday Heat Transfer Problems

(Q/A) / hout, F

0.09

0.06

0.03

10
15
5
Outside Convection Heat Transfer Coefficient,
Btu/hr-ft2-F

20

Figure 1-5 Sensitivity of house wall heat loss per unit area to the outside

convection heat transfer coefcient

These two sensitivities are similar, as can be expected, since the


linear product of insulation material R-value and insulation thickness
affects the heat loss, as shown in the governing heat loss Eq. (1-7).
Absolute sensitivity values are high at the low values of insulation

(Q/A) / Rinsulation,
(BTU/hr-ft2)2(in/F)

0
0.5
1
1.5
2
2.5

Insulation Material R-Value, hr-ft2-F/BTU-in

Figure 1-6 Sensitivity of house wall heat loss per unit area to insulation

material R-value

10

(Q/A) / t insulation, BTU/hr-ft2-in

Heat Loss From Walls In A Typical House

0.5

1.5

4
8
6
Insulation Thickness, in

10

12

Figure 1-7 Sensitivity of house wall heat loss per unit area to insulation

thickness

material R-value and insulation thickness. Sensitivities approach


zero as insulation material R-value and insulation thickness values
increase.
A ten-percent variation in independent variables around the
nominal values given above produces the sensitivity results given in
Table 1-1 The sensitivity results are given in a descending order and
they are applicable only in the region of assigned nominal values,
due to their non-linear effect to heat loss. The one exception is
temperature potential, (Tin Tout), which will always be 10% due
to its linear effect on heat loss. Material R-value and its thickness
change have the same sensitivity, since their linear product affects the
governing heat loss equation.
Heat loss through the wall is most sensitive to the temperature
potential between the inside and outside of the house. Changes in
wall insulation R-value and thickness affect heat loss as much as
the temperature potential. Continuing in order of sensitivity, wall
outer board R-value and thickness changes affect heat loss the most,
followed by wall inside board R-value and thickness. Wall heat loss is

11

Everyday Heat Transfer Problems

Table 1-1 House wall heat loss change per unit area due to a
10% change in variables nominal values

Variable

Nominal
Value

House Wall
Heat Loss
Change Due To
A 10%
Decrease In
Nominal Value

Tin Tout

36F

10%

+10%

Rinsulation

3.5 hr-ft -F/BTU-in

+7.467%

6.497%

House Wall
Heat Loss
Change Due To
A 10%
Increase In
Nominal Value

tinsulation

4 in

+7.467%

6.497%

Router wall

5 hr-ft2-F/BTU-in

+2.545%

2.545%

touter wall

1 in

+2.545%

2.545%

Rinner wall

0.85 hr-ft -F/BTU-in

+0.424%

0.424%

tinner wall

1 in

+0.424%

0.424%

hin

5 BTU/hr-ft -F

0.110%

+0.090%

hout

10 BTU/hr-ft -F

0.055%

+0.045%

least sensitive to both the inside and outside heat transfer coefcient
changes. Wall heat loss sensitivity to both the inside and outside
heat transfer coefcient changes is an order of magnitude less than
sensitivity to temperature potential changes.

12

CHAPTER

CONDUCTION

HEAT TRANSFER
IN A PRINTED
CIRCUIT BOARD

onduction heat transfer in printed circuit boards (PCBs) has been


studied extensively in literature i.e., B. Guenin [4]. The layered
structure of a printed circuit board is treated using two different
thermal conductivities; one is in-plane thermal conductivity and the
other is through-thickness thermal conductivity. One-dimensional
conduction heat transfers in in-plane direction and through-thickness
direction are treated independently. Since the signicant portion of
the conduction heat transfer in a PCB occurs in the in-plane direction
in the conductor layers, this is a valid assumption. Under steady-state
conditions and with constant thermophysical properties, the in-plane
(i-p) conduction heat transfer equation for a PCB can be written as:
Qin-plane = Q1i-p + Q2i-p + + Qni-p

(2-1)

where the subscript refers to the layers of the PCB. Using the
conduction rate equation in rectangular coordinates for a PCB with
a width of W, a length of L, layer thicknesses ti and layer thermal
conductivities ki, Eq. (2-1) can be rewritten as:

13

Everyday Heat Transfer Problems

W ti kin-plane (TL=0 TL=L)/L = W t1 k1 (TL=0 TL=L)/L


+ W t2 k2 (TL=0 TL=L)/L
+ W tn kn (TL=0 TL=L)/L

(2-2)

In-plane conduction heat transfer in a PCB represents a parallel


thermal resistance circuit which can be written as:
(1/Rin-plane) = (1/R1) + (1/R2) + + (1/Rn) where Ri = L/(kitiW) (2-3)
where
kin-plane = (kiti)/ti

(2-4)

Through-thickness (t-t) conduction heat transfer in a PCB represents


a series thermal resistance circuit, and the through-thickness conduction
heat transfer equation for a PCB can be written as:
Qthrough-thickness = Q1t-t = Q2t-t = = Qnt-t

(2-5)

which can be expanded into following equations:


W L kthrough-thickness (Tt=0 Tt=ti)/ti = W L k1 (Tt=0 Tt=t1)/t1
= W L k2 (Tt=t1 Tt=t2)/t2 = = W L kn (Tt=tn-1 Tt=tn)/tn

(2-6)

Inter-layer temperatures can be eliminated from Eqs. (2-6), and a


series thermal resistance equation extracted as follows:
Rthrough-thickness = R1 + R2 + + Rn where Ri = ti/ki

(2-7)

kthrough-thickness = ti/(ti/ki).

(2-8)

where

A printed circuit board is commonly built as layers of conductors


separated by layers of insulators. The conductors are mostly alloys
of copper, silver or gold, while the insulators are mostly a variety of

14

Conduction Heat Transfer In A Printed Circuit Board

epoxy resins. Therefore the in-plane thermal conductivity Eq. (2-4)


for a PCB can be rewritten as:
kin-plane = [kconductor tconductor + kinsulator (ttotal tconductor)]/ttotal (2-9)
Similarly, the through-thickness thermal conductivity Eq. (2-8) for
a PCB can be rewritten as:
kthrough-thickness
= ttotal/[(tconductor/kconductor) + (ttotal tconductor)/kinsulator] (2-10)
The sensitivities of these two PCB thermal conductivities are
analyzed for a 500 m-thick printed circuit board, with the assumed
nominal values for thermal conductivity of the conductor and
insulator layers given below:
kconductor = 377 W/m-C for copper conductor layers and
kinsulator = 0.3 W/m-C for glass reinforced polymer layers.
In-plane thermal conductivity versus percent of conductor layers
to total printed circuit board thickness is given in Figure 2-1. In-plane
thermal conductivity starts at the all-insulator thermal conductivity
value of 0.3 W/m-C, and increases linearly to conductor thermal
conductivity at no insulator layers.
Sensitivities of in-plane thermal conductivity to changes in kconductor
and kinsulator are represented in Figure 2-2. As you can see, the two
sensitivities are opposite.
The sensitivity of in-plane thermal conductivity to conductor
thickness is a constant, 0.75 W/m-C-m. The sensitivity of in-plane
thermal conductivity to insulator thickness is also a constant,
and it is the opposite of sensitivity to conductor thickness, namely
0.75 W/m-C-m.
A ten percent variation in variables around the nominal values
given above produce the sensitivity results in Table 2-1 for in-plane
thermal conductivity. For these nominal values, in-plane thermal

15

Everyday Heat Transfer Problems

kin-plane, W/m-C

400
300
200
100
0

20

40
60
80
% Conductor Layers Thickness

100

Figure 2-1 In-plane thermal conductivity versus percent of conductor layers

kin-plane / kconductor &


kin-plane / kinsulator

thickness to total printed circuit board thickness

kin-plane / kinsulator

0.8
0.6
0.4
0.2
0

kin-plane / kinsulator
0

20
40
60
80
% Conductor Layers Thickness

100

Figure 2-2 Sensitivity of in-plane thermal conductivity to kconductor and

to kinsulator

16

Conduction Heat Transfer In A Printed Circuit Board

Table 2-1 In-plane thermal conductivity change due to a 10%


change in variables around nominal values for a
500 micron thick PCB

Nominal
Value

In-Plane Thermal
Conductivity Change
Due To A 10%
Decrease In
Nominal Value

In-Plane Thermal
Conductivity Change
Due To A 10%
Increase In
Nominal Value

kconductor

377 W/m-C

9.99%

+9.99%

tconductor

250 m

9.98%

+9.98%

tinsulator

250 m

+9.98%

9.98%

kinsulator

0.3 W/m-C

0.01%

+0.01%

Variable

conductivity is 188.65 W/m-C. Insulator thermal conductivity is


the least effective independent variable in this case due to its low
value. Variations in the sum of conductor thicknesses and the sum of
insulator thicknesses affect in-plane thermal conductivity in opposing
directions, but with the same magnitude.
Through-thickness thermal conductivity versus percent of
conductor thickness has a non-linear behavior, and it is given
for a 500-micron PCB in Figure 2-3. Through-thickness thermal
conductivity is similar to insulator layer thermal conductivity for up
to 80% conductor layer thickness of the total printed circuit board,
and therefore is not a good conduction heat transfer path for printed
circuit boards.
Sensitivities of through-thickness thermal conductivity to
kconductor and to kinsulator are given in Figure 2-4. The sensitivity of
through-thickness thermal conductivity to changes in conductor
thermal conductivity is negligible throughout the percent conductor
layer thickness. The sensitivity of through-thickness thermal
conductivity to changes in insulator thermal conductivity increases
and becomes signicant as the thickness percentage of the insulator
layers decreases.

17

Everyday Heat Transfer Problems

kthrough-thickness, W/m-C

20
15
10
5
0

20

40
60
80
% Conductor Layers Thickness

100

Figure 2-3 Through-thickness thermal conductivity versus percent

kthrough-thickness / kconductor &


kthrough-thickness / kinsulator

conductor layer thickness

20

kthrough-thickness /
kinsulator

16
12
8
4
0

50
% Conductor Layers Thickness

100

kthrough-thickness /
kconductor

Figure 2-4 Sensitivity of through-thickness thermal conductivity to

conductor and insulator thermal conductivities versus percent


conductor layer thickness

18

kthrough-thickness / kconductor &


kthrough-thickness / kinsulator,
W/m-C-um

Conduction Heat Transfer In A Printed Circuit Board

0.05
kthrough-thickness /
kconductor

0
0.05
0.1
0.15
0.2

20
40
60
80
% Conductor Layers Thickness

100

kthrough-thickness /
kinsulator

Figure 2-5 Sensitivity of through-thickness thermal conductivity to

conductor and insulator thickness

Sensitivities of through-thickness thermal conductivity to


conductor and insulator thickness are given in Figure 2-5. The
sensitivities become signicant as conductor thickness
approaches 100%.
A ten-percent variation in independent variables around the
nominal values given above produce the sensitivity results given in
Table 2-2 for through-thickness thermal conductivity, which has a
nominal value of 0.6 W/m-C. In this case, through-thickness thermal
conductivity is very resistant to conductor thermal conductivity
variations in the region of interest.
A second and a similar analysis can be performed for a plated or
sputtered thinner circuit. A 10-m thick circuit is considered with the
assumed nominal thermal conductivities below:
kconductor = 377 W/m-C for copper conductor layers and
kinsulator = 36 W/m-C for aluminum oxide insulating layers.
In-plane thermal conductivity versus percent of conductor layers
to total thickness is given in Figure 2-6. In-plane thermal conductivity

19

Everyday Heat Transfer Problems

Table 2-2 Through-thickness thermal conductivity change


due to a 10% change in variables around nominal
values for a 500 micron thick PCB
Through-Thickness
Through-Thickness
Thermal Conductivity Thermal Conductivity
Change Due To
Change Due To
Nominal
A 10% Decrease In
A 10% Increase In
Value
Nominal Value
Nominal Value

Variable
tinsulator

250 m

+11.09%

9.08%

kinsulator

0.3 W/m-C

9.99%

+9.99%

tconductor

250 m

9.08%

+11.09%

kconductor

377 W/m-C

0.01%

+0.01%

kin-plane, W/m-C

400
300
200
100
0
0

20

40
60
80
% Conductor Layers Thickness

100

Figure 2-6 In-plane thermal conductivity versus percent of conductor layers

to total thickness

starts at the all insulator thermal conductivity value of 36 W/m-C and


increases linearly to a conductor thermal conductivity of 377 W/m-C.
Sensitivites of in-plane thermal conductivity to kconductor and kinsulator are
given in Figure 2-7. As you can see, the two sensitivities are opposite.

20

Conduction Heat Transfer In A Printed Circuit Board

kin-plane / kconductor &


kin-plane / kinsulator

kin-plane / kconductor

0.8
0.6
0.4
0.2
0

kin-plane / kinsulator
0

20
40
60
80
% Conductor Layers Thickness

100

Figure 2-7 Sensitivity of in-plane thermal conductivity to kconductor and

to kinsulator

The sensitivity of in-plane thermal conductivity to conductor thickness


is a constant at 34.1 W/m-C-m. The sensitivity of in-plane thermal
conductivity to insulator thickness is a constant, and it is the opposite of
sensitivity to in-plane thermal conductivity, namely 34.1 W/m-C-m.
A ten percent variation in variables around the nominal values given
above produce the sensitivity results in Table 2-3 for in-plane thermal
conductivity, which has a nominal value of 87.15 W/m-C. Insulator
thickness is the dominant independent variable in this region of interest.
Through-thickness thermal conductivity versus percent of
conductor thickness has a non-linear behavior, and it is given in
Figure 2-8. The percentage of the conductor layers thickness to total
circuit thickness affects the through-thickness thermal conductivity
at all conductor layer thicknesses. Through-thickness conduction heat
transfer is much more prominent in thin-plated or sputtered circuits.
Sensitivities of through-thickness thermal conductivity to kconductor
and to kinsulator are given in Figure 2-9. The sensitivity of throughthickness thermal conductivity to changes in conductor thermal
conductivity is negligible at low percentages of conductor layer
thickness. On the other hand, the sensitivity of through-thickness

21

Everyday Heat Transfer Problems

Table 2-3 In-plane thermal conductivity change due to a 10%


change in variables around nominal values for a
10-micron thick circuit

Nominal
Value

In-Plane Thermal
Conductivity
Change Due To
A 10% Decrease
In Nominal Value

In-Plane Thermal
Conductivity
Change Due To
A 10% Increase
In Nominal Value

tinsulator

8.5 m

+33.3%

33.3%

kconductor

377 W/m-C

6.5%

+6.5%

tconductor

1.5 m

5.9%

+5.9%

36 W/m-C

3.5%

+3.5

Variable

kinsulator

kthrough-thickness, W/m-C

thermal conductivity to changes in insulator thermal conductivity


is signicant; it increases to a maximum at around 90% conductor
layer thickness, and nally decreases sharply as conductor thermal
conductivity starts to dominate.

400
300
200
100
0

20

40

60

80

100

% Conductor Layers Thickness

Figure 2-8 Through-thickness conduction heat transfer coefcient versus

percentage of conductor layer thickness

22

kthrough-thickness / kconductor &


kthrough-thickness / kinsulator

Conduction Heat Transfer In A Printed Circuit Board

3.5
3

kthrough-thickness /
kinsulator

2.5
2
1.5
1
0.5
0

20
40
60
80
% Conductor Layers Thickness

100

kthrough-thickness /
kconductor

Figure 2-9 Sensitivity of through-thickness thermal conductivity to

kthrough-thickness / kinsulator &


kthrough-thickness / kconductor,
W/m-C-m

conductor and insulator thermal conductivities versus


percentage of conductor layer thickness

10

kthrough-thickness /
kconductor

0
10
20
30
40
50

20
60
40
80
% Conductor Layers Thickness

kthrough-thickness /
kinsulator
100

Figure 2-10 Sensitivity of through-thickness thermal conductivity to

conductor thickness and to insulator thickness

23

Everyday Heat Transfer Problems

Table 2-4 Through-thickness thermal conductivity change


due to a 10% change in variables around nominal
values for a 10-micron thick circuit
Through-Thickness
Through-Thickness
Thermal Conductivity Thermal Conductivity
Change Due To A 10% Change Due To A 10%
Decrease In
Increase In
Nominal Value
Nominal Value

Variable

Nominal
Value

kinsulator

36 W/m-C

9.85%

+9.18%

tinsulator

8.5 m

+9.76%

8.17%

tconductor

1.5 m

1.55%

+1.59%

kconductor

377 W/m-C

0.18%

+0.15%

Sensitivities of the through-thickness heat transfer coefcient


to conductor and insulator thickness are given in Figure 2-10. The
sensitivity to insulator thickness becomes signicant as the percent of
conductor thickness approaches 100%.
A ten percent variation in independent variables around the
nominal values given above produce the sensitivity results in
Table 2-4 for through-thickness thermal conductivity, which has a
nominal value of 41.65 W/m-C. In this case, through-thickness thermal
conductivity is most sensitive to insulator thermal conductivity and
insulator thickness variations.

24

CHAPTER

3
FROM COMBUSTION
HEAT TRANSFER

CHAMBER WALLS

ooling the walls of a combustion chamber containing gases at


high temperatures, i.e., 1000C, results in parallel modes of heat
transfer, convection and radiation. This problem can be approached
by assuming one-dimensional steady-state heat transfer in rectangular
coordinates and with constant thermophysical properties.
Convection heat transfer per unit area, from hot gases to the hot
side of a wall that separates the cold medium and the hot gases, can
be written as:
(Q/A)convection = hcg (Tg Twh)

(3-1)

Radiation heat transfer per unit area from hot gases which are
assumed to behave as gray bodies to the hot side of a wall can be
written as:
(Q/A)radiation = hrg (Tg Twh) = = g (Tg4 Twh4)

(3-2)

And so the total heat transfer from the hot gases to a combustion
chamber wall is:

25

Everyday Heat Transfer Problems

(Q/A)total = (Q/A)convection + (Q/A)radiation = (hcg + hrg) (Tg Twh) (3-3)


These two heat transfer mechanisms act as parallel thermal
resistances, namely:
(1/Rtotal) = (1/Rconvection) + (1/Rradiation)

(3-4)

where hcg is the convection heat transfer coefcient between gas


and the hot side of a wall, hrg is the radiation heat transfer coefcient
between gas and the hot side of a wall, Tg is average gas temperature
and Twh is average hot side wall temperature. The radiation heat
transfer coefcient hrg is dened as:
hrg = g (Tg4 Twh4)/(Tg Twh)

(3-5)

where g is emissivity of gas and is the Stefan Boltzmann constant.


Heat transfer occurs through the wall by conduction and is
dened as:
(Q/A)total = (kwall/L)(Twh Twc)

(3-6)

where kwall is wall material thermal conductivity, L is thickness of the


wall and Twc is the wall temperature at the cold medium side of the wall.
Heat transfer between the cold medium side of the wall and the
cold medium occurs by convection and is dened as:
(Q/A)total = hc (Twc Tc)

(3-7)

where hc is the convection heat transfer coefcient between the cold


medium side of the wall and the cold medium, and Tc is the average
temperature of cold medium.
In this example, Twh is going to be the dependent variable, and it
will be solved by iterating a function using a combination of above
Eqs. (3-3), (3-5), (3-6) and (3-7), as follows:
(Tg Tc)/{[1/(hcg + hrg)] + (L/kwall) + (1/hc)}
(hcg + hrg)(Tg Twh) = 0

26

(3-8)

Heat Transfer From Combustion Chamber Walls

The above governing equation can be rewritten as an iteration


function K as follows:
K = (Tg Tc) + {[1/(hcg + hrg)]
+ (L/kwall) + (1/hc)}(hcg + hrg)(Tg Twh)

(3-9)

During iteration to determine Twh, all temperatures should be in


degrees Kelvin because of the fourth power behavior of radiation
heat transfer. Also, all thermophysical properties are assumed to
be constants. Nominal values of these variables for the sensitivity
analysis are assumed to be as follows:
Tg = 1000C (1273 K)
Tc = 100C (373 K)
kwall = 20 W/m-K
L = 0.01m
hcg = 100 W/m2-K
hc = 50 W/m2-K
g = 0.2
= 5.67108 W/m2-K4
For these nominal variables, the iteration function crosses zero at
1075.93 K as shown in Figure 3-1, and 42.5% of the total heat transfer
from hot gases to the hot side of the wall comes from radiation mode;
the rest, 57.5%, comes from convection mode.
The effects of hot gas temperature and cold medium temperature
to hot side wall temperature are shown respectively in Figures 3-2
and 3-3. Hot gas temperature affects hot side wall temperature almost
one-to-one, namely a slope of 0.925. However, cold side medium
temperature affects hot side wall temperature almost ve-to-one,
namely a slope of 0.221.
The effects of wall parametersthermal conductivity and wall
thicknesson hot side wall temperature are shown in Figures 3-4
and 3-5. Changes in wall thermal conductivities below 10 W/m-K
are more effective on hot side wall temperature. Hot side wall
temperature sensitivity to wall thickness is pretty much a constant,
3.5 C/cm.

27

Everyday Heat Transfer Problems

Iteration Function, K

2000

1000

1000
400

600

800

1000

1200

1400

Twh, K

Figure 3-1 Iteration function versus Twh

Hot Side Wall Temperature, C

The convection heat transfer coefcients on both sides


of the wall have opposite effects on hot side wall temperature,
as shown in Figures 3-6 and 3-7. As the hot gas side convection
heat transfer coefcient increases, hot side wall temperature

2000
1600
1200
800
400
0

500

1000
1500
Gas Temperature, C

Figure 3-2 Hot side wall temperature versus gas temperature

28

2000

Hot Side Wall Temperature, C

Heat Transfer From Combustion Chamber Walls

1000

900

800

700

200

400
600
800
Cold Medium Temperature, C

1000

Figure 3-3 Hot side wall temperature versus cold medium temperature

Hot Side Wall Temperature, C

increases as well. Sensitivity of hot side wall temperature to


variations in the hot gas side convection heat transfer coefcient
is more prominent at lower convection heat transfer coefcient
values.

860

840

820

800

10

20
30
40
Wall Thermal Conductivity, W/m-K

50

Figure 3-4 Hot side wall temperature versus wall thermal conductivity

29

Hot Side Wall Temperature, C

Everyday Heat Transfer Problems

840
830
820
810
800

0.02

0.04
0.06
Wall Thickness, m

0.08

0.1

Figure 3-5 Hot side wall temperature versus wall thickness

Hot Side Wall Temperature, C

The variation of hot side wall temperature at different hot


gas emissivities is given in Figure 3-8. Hot side wall temperature
is more sensitive to changes in lower values of hot gas
emissivity.

1000
900
800
700
600

100

200

300

400

Hot Gas Side Convection Heat Transfer Coefficient,

500
W/m2-K

Figure 3-6 Hot side wall temperature versus hot gas side convection heat

transfer coefcient

30

Hot Side Wall Temperature, C

Heat Transfer From Combustion Chamber Walls

1000
800
600
400
200

100

200

300

400

500

Cold Medium Side Convection Heat Transfer Coefficient,


W/m2-K

Figure 3-7 Hot side wall temperature versus cold medium side convection

heat transfer coefcient

Hot Side Wall Temperature, C

When the nominal values of the variables given above are


varied 10%, the results shown in Table 3-1 are obtained. Hot
side wall temperature sensitivities to a 10% change in the
governing variables are given in descending order of importance,

950
900
850
800
750
700

0.2

0.4

0.6

0.8

Hot Gas Emissivity

Figure 3-8 Hot side wall temperature versus hot gas emissivity

31

Everyday Heat Transfer Problems

Table 3-1 Effects of 10% change in nominal values of


variables to hot side wall temperature

Nominal
Value

Change In Hot
Side Wall
Temperature For
A 10% Decrease
In Nominal Value

Change In Hot
Side Wall
Temperature For
A 10% Increase
In Nominal Value

Tg

1273 K

11.493%

+11.981%

hc

50 W/m2-K

+2.070%

1.991%

hcg

100 W/m2-K

1.262%

+1.136%

0.2

0.912%

+0.860%

Tc

373 K

0.295%

+0.295%

20 W/m-K

+0.056%

0.046%

0.01 m

0.051%

+0.051%

Variable

kwall
L

and they are applicable around the nominal values assumed for
this study.
Hot side wall temperature is most sensitive to variations in hot
gas temperature. Next in order of sensitivity are the convection heat
transfer coefcients on both sides of the wall. Changes to emissivity
of hot gases affect the dependent variable at the same level as the
convection heat transfer coefcients. Next in order of sensitivity is
the cold medium temperature. Hot side wall temperature is least
sensitive to variations in wall thermal conductivity and wall thickness.
This variable order of sensitivity is applicable around the nominal
values assumed for this case, due to the nonlinear relationship
between the dependent variable and the independent variables.

32

CHAPTER

HEAT TRANSFER

FROM A HUMAN
BODY DURING
SOLAR TANNING

he solar tanning of a human body was analyzed under


steady-state conditions with one-dimensional rate equations
in rectangular coordinates, and using temperature-independent
thermophysical properties. Human skin that is exposed to direct solar
radiation is considered to be in an energy balance. Energy goes into
the skin from both direct solar radiation and solar radiation scattered
throughout the atmosphere. Energy leaves the skin through a variety
of means and routes: by convection heat transfer and radiation heat
transfer (into the atmosphere), by conduction (to the inner portions
of the body), by perspiration, and by body basal metabolism. Other
energy gains and losses, such as those due to terrestrial radiation,
breathing and urination, are negligible.
Energy balance at the human skin gives the following heat transfer
equation:
Qsolar radiation absorbed + Qatmospheric radiation absorbed Qconvection
Qconduction to body Qradiation emitted Qperspiration Qbasal metabolism = 0 (4-1)

33

Everyday Heat Transfer Problems

For the present sensitivity analysis, the heat transfer rate equations
that will be used, and the nominal values that will be assumed for
energies outlined in Eq. (4-1), are given below.
Qsolar radiation absorbed = 851 W/m2

(4-2)

which assumes a gray body skin with an absorptivity, = , of 0.8, on


a clear summer day at noon, with full sun exposure
Qatmospheric radiation absorbed = 85 W/m2

(4-3)

which is assumed to be about 10% of Qsolar radiation absorbed.


Qconvection = h(Tskin Tenvironment)

(4-4)

where h is the heat transfer coefcient between the skin surface that
is being tanned and the environment. In the present analysis, h is
assumed to be 28.4 W/m2-K and Tenvironment is 30C.
Qconduction to body = (kbody/tbody)(Tskin Tbody)

(4-5)

where kbody = 0.2 W/m-K, tbody = 0.1 m, and Tbody = 37C.


Qradiation emitted = (T4skin T4environment)

(4-6)

where emissivity of skin surface = 0.8 and = 5.67 108 W/m2-K4.


Qperspiration = 337.5 W/m2

(4-7)

which corresponds to a 1 liter/hr perspiration rate for a human body


with a perspiration area of 2 m2.
Qbasal metabolism = 45 W/m2

(4-8)

which represents a 30-year-old male at rest.


There are ten independent variables that govern the dependent
variable Tskin in this heat transfer problem. Sensitivities to these ten

34

Heat Transfer From A Human Body During Solar Tanning

variables are analyzed in the region of the nominal values given above.
The governing Eq. (4-1) takes the following form, and can be solved
for Tskin by trial and error.
C1T4skin + C2Tskin = C3

(4-9)

where C1 = , C2 = h + kbody/tbody, and


C3 = Qsolar radiation absorbed + Qatmospheric radiation absorbed Qperspiration
Qbasal metabolism + T4environment + hTenvironment + kbody/tbodyTbody
All the calculations are performed in degrees Kelvin for
temperature, since the governing equation is non-linear in
temperature. These sensitivities are presented in Table 4-1 below
in the order of their signicance.
The most effective variable on the skin temperature is
Tenvironment and the least effective is the thermal conductivity of
human tissue. The skin temperature is an order of magnitude less
sensitive to changes in the thermal conductivity of human tissue
skin-to-body conduction heat transfer length, heat transfer due to
basal metabolism, body temperature, atmospheric radiation absorbed,
and emissivity of skin surfacethan changes in the temperature of
the environment, solar radiation absorbed, convection heat transfer
coefcient, and heat lost due to perspiration. Changes in some
variables, such as the environmental temperature and heat transfer
due to perspiration, behave linearly in the region of interest, and give
equal percentage changes to the dependent variable on both sides of
the variable's nominal value.
It is important to remind the reader that the order shown in
Table 4-1 is only useful in this region of the application due to
non-linear behavior of the sensitivities. The non-linear affects of
variables such as the convection heat transfer coefcient are given
in Figure 4-1 for Qsolar radiation absorbed = 851 W/m2 and for three different
perspiration rates.
The sensitivity of skin temperature to the convection heat transfer
coefcient is signicant up to 50 W/m2-K. The heat transfer coefcient

35

Everyday Heat Transfer Problems

Table 4-1 Effects of 10% change in nominal values of


variables to skin temperature

Variable

Nominal
Value

Skin
Temperature,
Tskin, Change
Due To A 10%
Decrease In
Nominal Value

Tenvironment

30C

6.03%

+6.03%

5.11%

+5.13%

Qsolar radiation absorbed

851 W/m

Skin
Temperature,
Tskin, Change
Due To A 10%
Increase In
Nominal Value

28.4 W/m -K

+2.93%

2.51%

Qperspiration

337.5 W/m

+2.03%

2.03%

0.8

+0.53

0.51

85 W/m2

0.51%

+0.51%

37C

0.45%

+0.45%

+0.27%

0.27%

Emissivity of
skin surface,
Qatmospheric radiation absorbed
Tbody

Qbasal metabolism

45 W/m

Skin-to-body
conduction
length, tbody

0.1 m

0.12%

+0.10%

0.2 W/m-K

+0.11%

0.11%

ktissue

from the skin surface to the environment can be determined from


appropriate empirical relationships found in References [6] and [10].
The heat transfer coefcient in the natural convection regime is
around 5 W/m2-K. If the wind picks up to, say, 8.9 m/s (20 mph),
then the heat transfer coefcient is in the turbulent ow regime, and
it increases to 20 W/m2-K. As the perspiration rate goes down, this
sensitivity increases. The sensitivity curves are given in Figure 4-2.
Similar results are obtained for an afternoon solar radiation
by assuming half the noon solar radiation, i.e., Qsolar radiation absorbed =
425.5 W/m2, and they are given in Figures 3-3 and 3-4.
As the solar radiation goes down, the skin temperature and its
sensitivity to the convection heat transfer coefcient decreases.

36

Heat Transfer From A Human Body During Solar Tanning

Tenvironment=30C, Tbody=37C, Qsolar radiation absorbed=851 W/m2


Skin Temperature, C

70
65
Perspiration=0.5
liters/hr
Perspiration=1.0
liter/hr
Perspiration=1.5
liter/hr

60
55
50
45
40
35
30

100

50

150

200

Convection Heat Transfer Coefficient, W/m2-K

Figure 4-1 Skin temperature versus convection heat transfer coefcient

for Qsolar radiation absorbed = 851 W/m2 and for three different
perspiration rates

Tskin / h, m2-K2/W

0
0.2

Perspiration=0.5
liters/hr
Perspiration=1.0
liters/hr
Perspiration=1.5
liters/hr

0.4
0.6
0.8
1

50

100

150

200

Convection Heat Transfer Coefficient, W/m2-K

Figure 4-2 Skin temperature sensitivity to convection heat transfer

coefcient versus convection heat transfer coefcient for


Qsolar radiation absorbed = 851 W/m2 and for three different perspiration
rates

37

Everyday Heat Transfer Problems

Tenvironment=30C, Tbody=37C, Qsolar radiation absorbed=425.5 W/m2


Skin Temperature, C

45
Perspiration=0.5
liters/hr
Perspiration=1.0
liter/hr
Perspiration=1.5
liter/hr

40
35
30
25

50

100

150

200

Convective Heat Transfer Coefficient, W/m2-K

Figure 4-3 Skin temperature versus convection heat transfer coefcient

for Qsolar radiation absorbed = 425.5 W/m2 and for three different
perspiration rates

Tskin / h, m2-K2/W

0.2
0.1
0

Perspiration=0.5
liters/hr
Perspiration=1.0
liters/hr
Perspiration=1.5
liters/hr

0.1
0.2
0.3
0.4
0.5

50

100

150

200

Convection Heat Transfer Coefficient, W/m2-K

Figure 4-4 Skin temperature sensitivity to convection heat transfer

coefcient versus convection heat transfer coefcient for


Qsolar radiation absorbed = 425.5 W/m2 and for three different
perspiration rates

38

Heat Transfer From A Human Body During Solar Tanning

60

Tskin, C

55
50
45
40
35
30

20

24

28
32
Tenvironment, C

36

40

Figure 4-5 Skin temperature versus environment temperature

In high perspiration rates, the skin temperature is below


the environmental temperature, and it approaches the
environmental temperature as the heat transfer coefcient
increases.

Tskin, C

50

45

40

35
0.5

0.6
0.7
0.8
0.9
Emissivity Of Human Skin Surface,

Figure 4-6 Skin temperature versus emissivity of human skin surface

39

Everyday Heat Transfer Problems

Tskin / , C

30

28

26

24
0.5

0.6

0.7

0.8

0.9
Emissivity Of Human Skin Surface,

Figure 4-7 Skin temperature sensitivity to emissivity of human skin surface

versus emissivity of human skin surface

Temperature of the skin behaves as shown in Figure 4-5 with


a constant sensitivity to the changes in the temperature of the
environment, in the domain of interest. The sensitivity behavior is
almost one-to-one; namely, Tskin/Tenvironment equals 0.92.
Another variable that is analyzed in detail in the region of interest
is the emissivity of human skin surface. Skin temperature increases
with an increase in emissivity as shown in Figure 4-6.
Figure 4-7 provides the skin temperature sensitivity to emissivity
of human skin surface. The sensitivity decreases in a linear fashion as
the emissivity increases.

40

CHAPTER

EFFICIENCY OF

RECTANGULAR
FINS

eat transfer from a surface can be enhanced by using ns.


Heat transfer from surfaces with different types of ns has been
studied extensively, as seen in References by Incropera, F. P. and
D. P. DeWitt [6] and by F. Kreith [10].
The present sensitivity analysis represents rectangular ns under
steady-state, one-dimensional, constant thermophysical property
conditions without radiation heat transfer. Energy balance to a
cross-sectional element of a rectangular n gives the following second
order and linear differential equation for the temperature distribution
along the length of the n.
d2T/dx2 (hP/kA) (T Tenvironment) = 0

(5-1)

where h is the convection heat transfer coefcient between the


surface of the n and the environment in W/m2-C, k is the thermal
conductivity of the n material in W/m-C, P is the n cross-sectional
perimeter in meters, and A is the n cross-sectional area in m2.
There can be different solutions to Eq. (5-1) depending upon the
boundary condition that is used at the tip of the n. If the heat loss

41

Everyday Heat Transfer Problems

to environment from the tip of the n is neglected, the following


boundary conditions can be used:
T = Tbase at x = 0 and (dT/dx) = 0 at x = L

(5-2)

The solution to Eqs. (5-1) and (5-2) can be written as


(T Tenvironment) = (Tbase Tenvironment) [cosh m(L-x)/cosh (mL)] (5-3)
where L is the length of the rectangular n in meters and m =
(hP/kA)0.5 in 1/m.
The heat transfer from the rectangular n can be determined from
Eq. (5-3) by nding the temperature slope at the base of the n,
namely
Qn = kA(dT/dx) at x = 0 or

(5-4)

Qn = (Tbase Tenvironment) sqrt(hkPA) tanh(mL)

(5-5)

Here the sensitivities of variables that affect the efciency of a


rectangular n will be analyzed. Fin efciency is generally dened by
comparing the n heat transfer to the environment with a maximum
heat transfer case to the environment, where the whole n is at the
n base temperature, namely = Qn/Qmax where Qmax = hAn(Tbase
Tenvironment). For a rectangular n, the n heat transfer efciency
is approximated by using Eq. (5-5), and by adding a corrected n
length, Lc, for the heat lost from the tip of the n.
= tanh(mLc)/(mLc)

(5-6)

where m = [h 2(w + t)/k wt ]0.5 and Lc = L + 0.5t.


For cases where the n width, w, is much greater than its
thickness, t, m becomes
m = (2h/kt)1/2

42

(5-7)

Efciency Of Rectangular Fins

There are four independent variables that affect the rectangular


n heat transfer efciency. These are the convection heat transfer
coefcient, h; the thermal conductivity of the n material, k; length of
the n, l; and thickness of the n, t.
The sensitivity of efciency to these four independent variables can
be obtained in closed forms by differentiating the efciency equation
with respect to the desired independent variable. For example:
/h = (0.5/h)[(1/cosh2(mLc)) (tanh(mLc)/(mLc))]

(5-8)

Fin efciency as a function of the convection heat transfer


coefcient for two different thermal conductivitiesaluminum and
copperis given in Figure 5-1. The sensitivity of rectangular n
efciency with respect to the convection heat transfer coefcient is
given in Figure 5-2.
Fin efciency is good in the natural convection regime and
degrades as high forced convection regimes are used. Sensitivity of
n efciency to the convection heat transfer coefcient is high in the

1
Fin Efficiency

0.9
kcu=377.2
W/m-C

0.8
0.7

kal=206
W/m-C

0.6
0.5
0.4

100

200

300

Convection Heat Transfer Coefficient,

400

L=0.0508 m
t=0.002 m

W/m2-C

Figure 5-1 Rectangular n efciency versus convection heat transfer

coefcient for two different n materials with L = 0.0508 m and


t = 0.002 m

43

Everyday Heat Transfer Problems

h/ h, m2-C/W

0.001

kcu=377.2
W/m-C
kal=206
W/m-C

0.002
0.003

L=0.0508 m
t=0.002 m

0.004
0.005

0
100
200
300
400
Convection Heat Transfer Coefficient, W/m2-C

Figure 5-2 Sensitivity of rectangular n efciency to convection heat

transfer coefcient versus convection heat transfer coefcient


for two different n materials with L = 0.0508 m and t = 0.002 m

natural convection regime and decreases as the forced convection


heat transfer coefcient increases.
Fin efciency as a function of n material thermal conductivity
for two different convection heat transfer coefcientsnatural
convection regime and forced convection regimeis given in
Figure 5-3. The sensitivity of rectangular n efciency with respect
to n material thermal conductivity is given in Figure 5-4.
Fin material thermal conductivity does not affect n efciency in
the natural convection regime except in the region of low thermal
conductivity materials. However, in the forced convection regime
the behavior is quite different. Fin material thermal conductivity
affects n efciency, and high thermal conductivity materials have to
be used in order to achieve high n efciency. The sensitivity of n
efciency to n material thermal conductivity is high for low thermal
conductivities. The sensitivity diminishes as high n material thermal
conductivities are utilized.

44

Efciency Of Rectangular Fins

Fin Efficiency,

1
h=5
W/m2-C

0.8

h=100
W/m2-C

0.6

0.4

100
200
300
400
Fin Thermal Conductivity, W/m-C

500

L=0.0508 m
t=0.002 m

Figure 5-3 Rectangular n efciency versus n material thermal

conductivity for two different convection heat transfer


coefcients with L = 0.0508 m and t = 0.002 m

/ k, m-C/W

0.0084
0.0063

h=5
W/m2-C

0.0042

h=100
W/m2-C

0.0021
0

L=0.0508 m
t=0.002 m
0

100

200

300

400

500

Fin Thermal Conductivity, W/m-C

Figure 5-4 Sensitivity of rectangular n efciency to n material thermal

conductivity versus n material thermal conductivity for two


different convection heat transfer coefcients with L = 0.0508 m
and t = 0.002 m

45

Everyday Heat Transfer Problems

Fin Efficiency,

1
0.9

h=5
W/m2-C

0.8

h=100
W/m2-C

0.7
0.6

t=0.002 m
k=377.2 W/m-C
0

0.02

0.04

0.06

0.08

0.1

Fin Length, m

Figure 5-5 Rectangular n efciency versus n length for two different

convection heat transfer coefcients with t = 0.002 m and


k = 377.2 W/m-C

Fin efciency as a function of n length for two different convection


heat transfer coefcientsnatural convection regime and forced
convection regimeis given in Figure 5-5. The sensitivity of rectangular
n efciency with respect to n length is given in Figure 5-6. Figure 5-6
shows sensitivities for combinations of two different convection heat
transfer coefcients and two different thermal conductivities.
Figure 5-5 shows that n efciency is a weak function n length in
the natural convection regime, but this weakness becomes a strong
function of n length in the forced convection regime. These results
can also be seen in Figure 5-6. In the natural convection regime,
sensitivity of n efciency to n length is low, but increases as the n
length increases. In the forced convection regime, sensitivity of n
efciency to n length starts low, goes through a maximum as the n
length increases, and decreases as the n length increases further.
Fin efciency as a function of n thickness for two different
convection heat transfer coefcientsnatural convection regime
and forced convection regimeis given in Figure 5-7. The sensitivity
of rectangular n efciency with respect to n thickness is given in
Figure 5-8. Figure 5-8 shows sensitivities for combinations of two

46

Efciency Of Rectangular Fins

8
h=5 W/m2-C & @
k=377.2 W/m-C
h=100 W/m2-C & @
k=377.2 W/m-C
h=5 W/m2-C & @
k=206 W/m-C
h=100 W/m2-C & @
k=206 W/m-C

7
/ L, 1/m

6
5
4
3
2
1
0

t=0.002 m
0

0.05
Fin Length, m

0.1

Figure 5-6 Sensitivity of rectangular n efciency to n length versus n

length for combinations of two different convection heat transfer


coefcients and two different thermal conductivities with t = 0.002 m

Fin Efficiency,

0.9
h=5
W/m2-C
h=100
W/m2-C

0.8
0.7
0.6

k=377.2 W/m-C
L=0.0508 m

0.5
0.4

0.001

0.002
0.003
Fin Thickness, m

0.004

0.005

Figure 5-7 Rectangular n efciency versus n thickness for two different

convection heat transfer coefcients with k = 377.2 W/m-C and


L = 0.0508 m

47

Everyday Heat Transfer Problems

9
8

/ t, 1/m

h=5 W/m2-C & @


k=377.2 W/m-C
h=100 W/m2-C & @
k=377.2 W/m-C
h=5 W/m2-C &
k=206 W/m-C
h=100 W/m2-C & @
k=206 W/m-C

6
5
4
3
2
1
0

L=0.0508 m
0

0.001

0.002

0.003

0.004

0.005

Fin Thickness, m

Figure 5-8 Sensitivity of rectangular n efciency to n thickness versus

n thickness for combinations of two different convection heat


transfer coefcients and two different thermal conductivities
with L = 0.0508 m

different convection heat transfer coefcients and two different


thermal conductivities.
Figure 5-7 shows that n efciency is a weak function of n
thickness in the natural convection regime, but this weakness
becomes a strong function of n thickness in the forced convection
regime. In Figure 5-8, in the natural convection regime, sensitivity of
n efciency to n thickness starts high at low n thickness values,
but decreases as the n thickness increases. In the forced convection
regime, sensitivity of n efciency to n thickness starts low, goes
through a maximum as the n thickness increases, and decreases as
the n thickness increases further.
A ten percent variation in independent variables around the
nominal values produces the following sensitivity results (Table 5-1)
for n efciency. The results are given for the natural convection
regime in descending order, from the most sensitive variable to the
least. Table 5-2 gives similar results for the forced convection regime.

48

Efciency Of Rectangular Fins

Table 5-1 Rectangular n efciency change due to a 10%


change in variables nominal values for the natural
convection regime

Variable

Nominal
Value

Rectangular
Fin Efciency
Change Due To
A 10% Decrease
In Nominal Value

0.0508 m

+0.218%

0.239%

377.2 W/m-C

0.129%

+0.106%

0.002 m

0.124%

+0.102%

5 W/m -C

+0.117%

0.116%

Rectangular
Fin Efciency
Change Due To
A 10% Increase
In Nominal Value

The order of signicance in n efciency change follows the same


pattern with respect to variables in both the natural convection and
forced convection regimes. However, in the forced convection regime,
n efciency changes are an order of magnitude higher than the
natural convection regime.

Table 5-2 Rectangular n efciency change due to a


10% change in variables nominal values for the
forced convection regime

Variable

Nominal
Value

Rectangular
Fin Efciency
Change Due To
A 10% Decrease
In Nominal Value

Rectangular
Fin Efciency
Change Due To
A 10% Increase
In Nominal Value

0.0508 m

+3.436%

3.478%

377.2 W/m-C

1.916%

+1.639%

0.002 m

1.844%

+1.574%

100 W/m -C

+1.806%

1.729%

49

CHAPTER

HEAT

TRANSFER FROM
A HOT DRAWN BAR

hot drawn bar, assumed to be moving at a constant velocity out of


a die at constant temperature, will be treated as a one-dimensional heat transfer problem. The Biot numberfor the bar, htD/2k, will be
assumed to be less than 0.1, to assure no radial variation of temperature in the bar. Here ht is the total heat transfer coefcient from the
bar surface in W/m2-K, the sum of the convection heat transfer coefcient and radiation heat transfer coefcient, D is the bar diameter,
and k is the bar thermal conductivity in W/m-K. Conduction, convection, and radiation heat transfer mechanisms affect the temperature
of the drawn bar.
Energy balance can be applied to a small element of the bar with a
width of dx:
Conduction heat transfer into the element Conduction heat
transfer out of the element Convection heat transfer out of the
element to the environment Radiation heat transfer out of the
element to the environment = Rate of change of internal energy
of the element

51

Everyday Heat Transfer Problems

The energy balance on the element can be written as follows:


Qconduction at x Qconduction at x+dx Qconvection from dx Qradiation from dx
= cpAdx(dT/d)
(6-1)
where is the density of the bar in kg/m3, cp is the specic heat of the
bar at constant pressure in J-kg/K, A is the bar cross-sectional area
in m2, T is the temperature of the bar element in K, and dx/d is the
drawn bar velocity in m/s.
Assuming that all the bar thermophysical and geometrical
properties are constants, the following one-dimensional, second-order
and non-linear differential equation is obtained:
d2T/dx2 (hP/kA)(T-Tenvironment) (P/kA)(T4 T4environment)
= (cpU/k)dT/dx

(6-2)

where P is the bar perimeter in m, is Stefan Boltzmann constant,


5.6710-8 W/m2-K4, is the bar surface emissivity, and U=dx/d the
speed of the hot drawn bar. The differential equation (6-2) reduces
to steady-state heat transfer from ns with a uniform cross-sectional
area; if the radiation heat transfer and the rate of change of internal
energy are neglected, see References by F. Kreith [10] and by
Incropera, F. P. and D. P. DeWitt [6].
The boundary condition for this heat transfer problem can be
specied as follows:
T = Tx=0 (temperature of the drawn bar at the die location)
at x = 0
(6-3)
and
T = Tenvironment as x goes to 4

(6-4)

The governing differential equation (6-2), along with boundary


conditions (6-3) and (6-4), can be solved by nite difference methods
and iteration, in order to determine the temperature at the i'th
location along the bar.

52

Heat Transfer From A Hot Drawn Bar

Another method to solve this non-linear heat transfer problem is to


dene a radiation heat transfer coefcient utilizing the temperature of
the previous bar element, i-1, as follows:
hradiation = (T4i-1 T4environment)/(Ti-1 Tenvironment)

(6-5)

and write the differential equation for location i along the bar as
d2Ti/dx2 [(hconvection + hradiation) P/kA](Ti -Tenvironment)
= (cpU/k)dTi/dx

(6-6)

The solution reached by linear differential equation (6-6) for


location i along the bar is valid for small x increments along the bar,
i.e. < 0.01 m, since the radiation heat transfer coefcient is calculated
using the temperature of the previous element i-1.
The solution to the above second order linear differential equation
(6-6) which satises both boundary conditions, (6-3) and (6-4), is:
(Ti Tenvironment)/(Tx=0 Tenvironment)
= exp{[(U/2) sqrt((U/2)2 + m2)]x}

(6-7)

where = k/cp is the thermal diffusivity of the bar in m2/s and


m2 = (hconvection + hradiation)P/kA in 1/m2.

(6-8)

This temperature distribution solution reduces to steady-state heat


transfer from rectangular ns with a uniform cross-sectional area
and the above applied boundary conditions, (6-3) and (6-4), if the
radiation heat transfer and the hot drawn bar velocity are neglected,
i.e., hradiation=0 and U=0 (see References by F. Kreith [10] and by
Incropera, F. P. and D. P. DeWitt [6]).
Sensitivity to governing variables is analyzed by xing the drawn bar
temperature at the die, i.e., Tx=0=1273 K. There are eight independent
variables that affect the temperature distribution of the hot drawn bar.
The sensitivities of bar temperature to these variables are analyzed by
assuming the following nominal values for a special steel bar:

53

Everyday Heat Transfer Problems

D = 0.01 m
= 8000 kg/m3
cp = 450 J/kg-K
k = 40 W/m-K
= 0.5
U = 0.02 m/s
Tenvironment = 298 K
The last independent variable is the convention heat transfer
coefcient between the surface of the bar and the environment.
The convection heat transfer coefcient can be determined from a
drawn bar temperature requirement at a distance from the die. In
the present analysis, Tx=10 is specied to be 373 K. At approximately
x=10 m the radiation heat transfer contribution almost diminishes.
The convection heat transfer coefcient that meets the Tx=10 = 373 K
requirement is determined from the above solution to be:
hconvection = 46.17 W/m2-K
which is in the turbulent region of forced cooling air over the
cylindrical bar, i.e., ReD = 1083 where ReD= VairD/air. The empirical
relationship for the convection heat transfer coefcient for air owing
over cylinders is given in Reference [10] by F. Kreith as:
hconvectionD/kair = 0.615 (VairD/air)0.466 for 40 < ReD < 4000

(6-9)

where Vair is mean air speed over the cylinder (2.18 m/s in this case),
kair is air thermal conductivity, and air is air kinematic viscosity.
Air thermophysical properties are calculated at lm temperature,
namely the average of bar surface temperature and environmental
temperature.
A comparison of convection and radiation heat transfer coefcients
as a function of distance from the die is given in Figure 6-1.
Heat transfer due to radiation is at the same order of magnitude
around the die. As the bar travels away from the die, radiation heat
transfer diminishes rapidly.

54

Heat Transfer Coeffieicents


W/m2-K

Heat Transfer From A Hot Drawn Bar

50
hconvection

40
30
20
10
0

hradiation
0

4
6
8
Distance From Die, m

10

Figure 6-1 Radiation and convection heat transfer coefcients as a function

of distance from die

Hot drawn bar temperature distributions, both with and without


radiation heat transfer, are shown in Figure 6-2. Radiation heat
transfer effects on bar temperature cannot be neglected below x=4
meters from the die.
In the initial sensitivity analysis, radiation heat transfer effects
will be neglected, namely hradiation=0. Hot drawn bar temperatures
as a function of distance from the die for different convection heat
transfer coefcients are given in Figure 6-3. Temperatures are very
sensitive to low convection heat transfer coefcients.
The sensitivity of bar temperature at x=10 m to the convection heat
transfer coefcient is given in Figure 6-4. Bar temperature sensitivity
is high at natural convection and at low forced-convection heat
transfer regions. As the forced-convection heat transfer coefcient
increases, bar temperature sensitivity to the convection heat transfer
coefcient decreases.
Bar temperature at x=10 m as a function of the convection heat
transfer coefcient is shown in Figure 6-5. Increasing the convection
heat transfer coefcient reduces its effects on bar temperature.

55

Everyday Heat Transfer Problems

Temperature, K

1400
1200
1000

Without Radiation
With Radiation

800
600
400
200
0

2
6
4
8
Distance From Die, m

10

Figure 6-2 Hot drawn bar temperature with and without radiation heat

transfer effects

1400
hconvection=5
W/m2-K
hconvection=20
W/m2-K
hconvection=40
W/m2-K
hconvection=60
W/m2-K

Temperature, K

1200
1000
800
600
400
200

8
6
Distance From Die, m

10

Figure 6-3 Hot drawn bar temperature for different convection heat trans-

fer coefcients with hradiation=0

56

T/ hconvection, m2-K2/W

Heat Transfer From A Hot Drawn Bar

0
10
20
30
40
50

10

20

30

40

50

hconvection, W/m2-K

Figure 6-4 Bar temperature sensitivity @ x=10 m to convection heat trans-

Bar Temperature @ x=10 m, K

fer coefcient

1000
800
600
400
200

10

20
30
40
hconvection, W/m2-K

50

60

Figure 6-5 Bar temperature @ x=10 m as a function of convection heat

transfer coefcient

57

Bar Temperature @ x=10 m, K

Everyday Heat Transfer Problems

400

380

360

340
7000

7500

8000
Bar Density,

8500

9000

kg/m3

Figure 6-6 Bar temperature @ x=10 m versus bar density

Bar temperature at x=10 m varies close to a linear behavior with


bar density, as shown in Figure 6-6. ?T/? is 0.0239 K-m3/kg in the
7000 to 8000 kg/m3 bar density range.
Bar temperature at x=10 m also varies close to a linear behavior
with bar-specic heat at constant pressure, as shown in Figure 6-7.
?T/?cp is 0.4254 K2-kg/J in the 400 to 500 J/kg-K bar specic heat
range.
Bar temperature at x=10 m varies linearly with bar thermal
conductivity, as shown in Figure 6-8. Bar temperature is a weak
function of bar thermal conductivity in this problem. ?T/?k is 0.0007
K2-m/W in the 20 to 60 W/m-K bar thermal conductivity range.
Bar temperature at x=10 meters versus bar velocity is given in
Figure 6-9.
Hot drawn bar velocity does not affect bar temperature at low
velocities, i.e., U<0.01 m/s. As bar velocity increases above 0.01 m/s,
bar temperature increases rapidly. Bar temperature at x=10 meters
has a sensitivity of about 10,000 K-s/m to changes in bar velocity
around 0.03 m/s. As bar velocity increases further, bar temperature
sensitivity at x=10 meters starts to decrease.

58

Bar Temperature @ x=10 m, K

Heat Transfer From A Hot Drawn Bar

400

380

360

340
400

420

440
460
Bar Specific Heat, J/kg-K

480

500

Figure 6-7 Bar temperature @ x=10 m versus bar specic heat at constant

Bar Temperature @ x=10 m, K

pressure

373.04
373.03
373.02
373.01
373

20

30

40

50

60

Bar Thermal Conductivity, W/m-K

Figure 6-8 Bar temperature @ x=10 m versus bar thermal conductivity

Bar temperature at x=10 meters versus bar diameter is given


in Figure 6-10. Bar temperature at x=10 meters is not sensitive to
bar diameter changes in small diameter values, i.e., D<0.005 m. As
bar diameter increases, bar temperature sensitivity goes through a

59

Bar Temperature @ x=10 m, K

Everyday Heat Transfer Problems

700
600
500
400
300
200

0.01

0.02
0.03
Bar Velocity, m/s

0.04

0.05

Bar Temperature @ x=10 m, K

Figure 6-9 Bar temperature @ x=10 m versus bar velocity

1000
800
600
400
200

0.01

0.02
0.03
Bar Diameter, m

0.04

0.05

Figure 6-10 Bar temperature @ x=10 m versus bar diameter

maximum of about 19,500 K/m at around D=0.02 meters, and starts to


decrease as the bar diameter increases.
Bar temperature at x=10 meters versus environmental temperature
is given in Figure 6-11. The relationship is linear as expected.

60

Bar Temperature @ x=10 m, K

Heat Transfer From A Hot Drawn Bar

380
375
370
365
360

15

20
25
Environment Temperature, C

30

Figure 6-11 Bar temperature @ x=10 m versus environment temperature

The slope of the curve is 0.9231 K/C under the assumed nominal
conditions.
When the nominal values of the independent variables given above
are varied +-10%, the results shown in Table 6-1 are obtained. The
sensitivity analysis is conducted by neglecting radiation heat transfer
at x=10 meters.
The convection heat transfer coefcient, bar diameter, bar velocity,
bar density and bar-specic heat at constant pressure have the same
order of magnitude sensitivity on bar temperature at x=10 m. Changes
in the temperature of the environment affect bar temperature at x=10
m, at an order of magnitude less. Changes to the thermal conductivity
of the bar have the least effect on bar temperature at x=10 m. The
sensitivity magnitudes and order that are shown in Table 6-1 are only
valid around the nominal values that are assumed for the independent
variables for this analysis. Bar velocity, bar density and bar-specic
heat at constant pressure have the same sensitivity effects on the
temperature of the bar as can be seen in Eq. (6-7).
Another interesting sensitivity analysis can be performed around
x=0.5 m, where both the convection and the radiation heat transfers
are in effect.

61

Everyday Heat Transfer Problems

Table 6-1 Effects of 10% change in nominal values of variables to bar temperature @ x=10 m

Nominal
Value

Bar Temperature
@ x=10 m Change
Due To A 10%
Decrease In
Nominal Value

Bar Temperature
@ x=10 m Change
Due To A 10%
Increase In
Nominal Value

46.17 W/m2-K

+5.879%

4.549%

0.01 m

4.986%

+5.280%

0.02 m/s

4.986%

+5.279%

8000 kg/m

4.986%

+5.279%

cp

450 J/kg-K

4.986%

+5.279%

25C (298 K)

0.619%

+0.619%

40 W/m-K

0.0007%

+0.0007%

Variable
hconvection

Tenvironment
k

Table 6-2 Effects of 10% change in nominal values of


variables to bar temperature @ x=0.5 m, including
radiation heat transfer

Nominal
Value

Bar Temperature
@ x=0.5 m Change
Due To A 10%
Decrease In
Nominal Value

Bar Temperature
@ x=0.5 m Change
Due To A 10%
Increase In
Nominal Value

0.01 m

1.386%

+1.183%

0.02 m/s

Variable

1.386%

+1.183%

8000 kg/m

1.386%

+1.183%

cp

450 J/kg-K

1.386%

+1.183%

46.17 W/m -K

+0.788%

0.783%

hconvection

Tenvironment
k

62

0.5

+0.500%

0.482%

25C (298 K)

0.046%

+0.046%

40 W/m-K

0.0003%

+0.0003%

Heat Transfer From A Hot Drawn Bar

Sensitivities to a +-10% change in independent variables are given


in descending order of importance in Table 6-2.
All the sensitivities to governing independent variables at x=0.5
meters are at the same order of magnitude, except Tenvironment and k.
Variations in radiation and convection heat transfer losses from the
bar have similar effects on bar temperature close to the die.

63

CHAPTER

MAXIMUM

CURRENT IN
AN OPEN-AIR
ELECTRICAL WIRE

aximum current in an open-air single electrical wire (not in a


bundle), will be analyzed under steady-state and one-dimensional
cylindrical coordinates, with constant property conditions. The wire is
assumed to be a cylindrical conductor with a certain diameter and with
certain material characteristics; i.e., resistivity. The wire conductor is
insulated with concentric layers of insulation material that can stand
up to a certain wire conductor temperature, Tc, which will be the
temperature rating of the wire. Heat generated in the wire conductor
is I2R and the temperature within the wire conductor is assumed to be
uniformthe conductor temperature does not vary radially from the
center to the outer radius of the wire conductor.
Heat is transferred by conduction through the wire insulator and by
convection from the surface of the wire insulator to the environment.
Radiation heat transfer from the surface of the wire insulator is
neglected. The conduction heat transfer from the conductor to the
outer radius of the wire insulator can be written by the rate equation
in cylindrical coordinates:
Q = 2Lkins (Tc Tinsulation outer radius)/ln(rw/rc)

(7-1)

65

Everyday Heat Transfer Problems

The convection heat transfer from the outer surface of the wire
insulator to the environment can be written by the rate equation in
cylindrical coordinates:
Q = 2rwLh (Tinsulation outer radius Tenv)

(7-2)

The heat transfer mechanisms in Eqs. (7-1) and (7-2) are in a series
thermal resistance path. The energy balance for this heat transfer
problem, heat generated by the conductor equals heat lost to the
environment, can be written as follows, by eliminating Tinsulation outer radius
from Eqs. (7-1) and (7-2):
I2R = (Tc Tenv)/{[ln(rw/rc)/(2Lkins)] + (1/2rwLh)}

(7-3)

where the rst term in the denominator is the conduction heat


transfer resistance in the wire insulation, and the second term is
the convection heat transfer resistance at the insulated wire's outer
surface.
Denitions of variables:
I = Current through the conductor in amps
R = Resistance of the wire can also be written as L/rc2 where
= Resistivity of the conductor material in -m
L = Length of the conductor in meters
rc = Conductor radius in meters
Tc = Temperature rating of the wire in C
Tenv = Temperature of the environment in C
rw = Radius of insulated wire in meters
kins = Thermal conductivity of wire insulation in W/m-C
h = Convection heat transfer coefcient in W/m2-C
The maximum current that a wire can stand can be written from
Eq. (7-3) by replacing resistance of the wire with resistivity of the
conductor material:
Imax = rc {(2/)(Tc Tenv)/[ln(rw/rc)/kins + (1/rwh)]}1/2

66

(7-4)

Maximum Current In An Open-Air Electrical Wire

There are six independent variables that govern the maximum


current allowed in the wire. These are the conductor radius, rc, the
conductor resistivity, , the temperature potential (the temperature
rating of the wire minus the temperature of the environment), Tc
Tenv, the radius of the insulated wire, rw, the thermal conductivity of
the insulating material, kins, and the open air convention heat transfer
coefcient, h.
The sensitivities of maximum current allowed with respect to
these six variables will be analyzed as a function of wire conductor
diameter. Since wire gauges are used in the industry, the sensitivity
results are presented for different wire gauges. Table 7-1 provides the
standard wire gauges and their conductor diameters.
In this problem it is fortunate that the sensitivities can be obtained
in a closed form by differentiating the above maximum current

Table 7-1 Wire conductor diameter versus AWG


Wire Conductor Diameter, mm

American Wire gauge (AWG)

0.8128

20

1.0236

18

1.2900

16

1.6280

14

2.0520

12

2.5900

10

3.2640

4.1150

5.1890

6.5430

7.3406

8.2550

1/0

9.2710

2/0

10.4140

3/0

11.6840

4/0

67

Everyday Heat Transfer Problems

Imax/ (2rc), A/mm

40

30

20
AWG20
10

AWG6
4

AWG2
6

AWG1/0 AWG3/0
8

10

12

Wire Conductor Diameter, mm

Figure 7-1 Maximum wire current to wire conductor diameter sensitivity as

a function of wire diameter

equation with respect to a desired variable while holding the other


variables constant, namely Imax/(2rc), Imax/Tenv, Imax/, Imax/rw,
Imax/kins, and Imax/h.
For example,
Imax/Tenv = rc {(2/)(Tc Tenv)/[ln(rw/rc)/kins + (1/rwh)]}1/2
{(1/)/[ln(rw/rc)/kins + (1/rwh)]}

(7-5)

The sensitivity of maximum wire current to conductor radius is


given in Figure 7-1.
Change in maximum wire current with respect to wire conductor
diameter is high at small diameters, and decreases as the wire
diameter increases.
Figure 7-1 was obtained by using the following values for the other
variables:
= 1.60E-08 -m
Tc = 60C
Tenv = 30C

68

Maximum Current In An Open-Air Electrical Wire

Maximum Current
Allowed, A

400
NEC
Current
Allowed, A

300
200

Maximum
Current
Allowed, A

100
0
0

10

12

Wire Conductor Diameter, mm

Figure 7-2 Maximum wire current allowed versus wire conductor diameter

@ a wire temperature rating of 60C

rw = 3rc
kins = 0.3 W/m-C
h = 9 W/m2-C
Under the above conditions the maximum wire current calculated
is compared to National Electric Code (NEC) recommendations in
Figure 7-2.
The calculated maximum wire currents are very much in line with
the NEC-recommended values up to AWG6. The variable assumptions
given above start to predict higher maximum current values for large
diameter wires, and therefore have to be adjusted for large diameter
wire applications.
The sensitivity of maximum wire current to environmental
temperature, Imax/Tenv, is given in Figure 7-3.
Sensitivities in Figure 7-3 are for a wire temperature rating of
60C. As the environmental temperature approaches the wire rating
temperature, the maximum wire current allowed becomes more
sensitive to environmental temperature changes.

69

Everyday Heat Transfer Problems

Imax/ Tenv, A/C

0
2
4
Tenv=0C
Tenv=20C
Tenv=40C
Tenv=50C

6
8
10
12

10

12

Wire Conductor Diameter, mm

Figure 7-3 Maximum wire current to environment temperature sensitivity as

a function of wire conductor diameter at different environmental


temperatures for a wire temperature rating of 60C

The sensitivities of Maximum wire current allowed to wire


insulation thickness, rw rc, are given in Figures 7-4 and 7-5.
For a given insulation thickness, the change in maximum allowable
wire current reaches its peak at a certain wire conductor diameter
and then starts to decrease, as seen in Figure 7-4. As the insulation
thickness increases, the maximum sensitivity point moves towards
smaller diameter conductors. Another way to look at the maximum
allowable wire current sensitivity to insulation thickness is shown
in Figure 7-5, where the sensitivities decrease as the wire insulation
thickness increases for different gauge wires.
The sensitivities of maximum wire current allowed to
wire insulation thermal conductivity, Imax/kins, are given in
Figure 7-6.
Maximum wire current is more sensitive to wire insulation thermal
conductivity variations at lower thermal conductivity values and in
large gauge wires. Maximum wire current is almost insensitive to
wire insulation thermal conductivity variations below 3 mm diameter
(AWG8) conductors.

70

Maximum Current In An Open-Air Electrical Wire

Imax/ (rw-rc), A/m

10000
8000
6000
4000
2000
0
0.000 0.002 0.004 0.006 0.008 0.010 0.012

Ins Thickness=
0.5*Wire Dia
Ins Thickness=
0.75*Wire Dia
Ins Thickness=
1.0*Wire Dia
Ins Thickness=
1.25*Wire Dia
Ins Thickness=
1.5*Wire Dia

Wire Conductor Diameter, m

Figure 7-4 Maximum wire current to insulation thickness sensitivity as

a function of wire conductor diameter for different insulation


thicknesses

Imax/ (rw-rc), Amp/m

8000
AWG20
AWG16
AWG12
AWG8
AWG4
AWG1
AWG2/0
AWG4/0

7000
6000
5000
4000
3000
2000
1000
0.5*d

0.75*d

1.25*d

1.5*d

Insulation Thickness

Figure 7-5 Maximum wire current to insulation thickness sensitivity as a

function of insulation thickness for different wire diameters

71

Everyday Heat Transfer Problems

Imax/ kins, A-m-K/W

300
AWG20
AWG16
AWG12
AWG8
AWG4
AWG1
AWG2/0
AWG4/0

250
200
150
100
50
0

0.2

0.25

0.3

0.35

0.4

Thermal Conductivity Of Wire Insulation, W/m-K

Figure 7-6 Maximum wire current to thermal conductivity of wire insulation

sensitivity for different wire gauges

Imax/ h, A-m2-K/W

500
AWG20
AWG16
AWG12
AWG8
AWG4
AWG1
AWG2/0
AWG4/0

400
300
200
100
0

12

16

20
2

Convection Heat Transfer Coefficient, W/m -K

Figure 7-7 Maximum wire current to convection heat transfer coefcient

sensitivity for different wire gauges

72

Maximum Current In An Open-Air Electrical Wire

Imax/ h, A-m2-K/W

600
500
h=7 W/m2-K

400

h=8 W/m2-K

300

h=9 W/m2-K
h=12 W/m2-K

200

h=15 W/m2-K
h=20 W/m2-K

100
0
0

0.002 0.004 0.006 0.008

0.01

0.012

Wire Diameter, m

Figure 7-8 Maximum wire current to convection heat transfer coefcient

sensitivity as a function of wire diameter for different convection


heat transfer coefcients

The sensitivities of maximum wire current allowed to convection


heat transfer coefcient, Imax/h, are given in Figures 7-7 and 7-8.
Maximum wire current is more sensitive to convection heat transfer
coefcient variations at high convection heat transfer coefcient
values and in large gauge wires. As the wire surface area increases
in large gauge wires, changes in the heat transfer coefcient gain
importance.
The sensitivities of maximum wire current allowed to wire
conductor resistivity, Imax/, are given in Figure 7-9.
The value of Imax/ is always negative; when wire resistivity
increases maximum wire current decreases, and vice versa. Maximum
wire current is more sensitive to wire conductor resistivity variations
at low wire conductor resistivity values and in large gauge wires.
The above sensitivity graphs show that maximum wire current
sensitivity to all the governing variables behaves non-linearly.
Sometimes it is more appropriate to analyze these sensitivities in
the region of interest, and rank them according to their effects on
maximum wire current.

73

Imax/ x 10-8, A/Ohm-m

Everyday Heat Transfer Problems

0
AWG 20
AWG 16
AWG 12
AWG 8
AWG 4
AWG 1
AWG 2/0
AWG 4/0

50
100
150
200
1.00E-08

1.50E-08

2.00E-08

2.50E-08

3.00E-08

Wire Resistivity, , Ohm-m

Figure 7-9 Maximum wire current to wire conductor resistivity sensitivity as

a function of wire conductor resistivity for different wire gauges

For example, maximum wire current sensitivities are analyzed for


a 4.115 mm diameter, AWG6, conductor with a temperature rating of
60C for the following nominal values of the variables:
Conductor diameter: 2rc = 4.115 0.4115 mm
Conductor resistivity: = 1.60E-08 0.16E-08 -m
Temperature Rating Of Wire Temperature of the environment:
Tc Tenv = 30 3C
Wire insulation thickness: rw rc = 2rc 0.2rc = 4.115 0.4115 mm
Wire insulation thermal conductivity: kins = 0.3 0.03 W/m-C
Convection heat transfer coefcient: h = 9 0.9 W/m2-C
The nominal values are varied 10%, and the effects of these
variations are presented in Table 7-2.
The sensitivity effects to maximum wire current shown in
Table 7-2 are given in descending order. The most effective variable
is the wire conductor diameter, and the least effective is the thermal
conductivity of the wire insulation material. There is an order of
magnitude difference in the effects of these two variables. In this
region of interest, it would be more advisable to focus on the accuracy

74

Maximum Current In An Open-Air Electrical Wire

Table 7-2 Effects of 10% change in nominal values of


variables to maximum wire current for a 4.115 mm
diameter, AWG6, conductor with a temperature
rating of 60C

Variable

Maximum Wire Maximum Wire


Current Change Current Change
Due To A 10%
Due To A 10%
Nominal
Decrease In
Increase In
Value
Nominal Value Nominal Value

Conductor
Diameter, 2rc

4.115 mm

13.9%

+14.5%

Conductor
Resistivity,

1.60E-08
-m

+5.4%

4.7%

Temperature Rating Of
Wire Tc Temperature
Of Environment, Tenv

30C

5.1%

+4.9%

Convection Heat
Transfer Coefcient, h

9 W/m2-C

4.32%

+4.01%

Insulation Thickness,
rw rc

2rc

2.35%

+2.17%

Thermal Conductivity
Of Wire insulation, kins

0.3 W/m-C

0.93%

+0.78%

of the wire conductor diameter rather than the accuracy of the wire
insulation thermal conductivity. It is important to remember that the
maximum wire current change order shown in Table 7-2 is good only
in this region of the application, due to non-linear behavior of the
sensitivities.

75

CHAPTER

EVAPORATION

OF LIQUID
NITROGEN IN A
CRYOGENIC BOTTLE
H

eat transfer in cryogenic bottles involves conduction, convection


and radiation modes. Dewar invented the vacuum ask at the
beginning of the twentieth century to minimize heat transfer and
contain low or high-temperature uids in it.
In this example, a cryogenic bottle with a stainless steel inner tube,
a vacuum gap and an outer insulation layer will be utilized to store
liquid nitrogen. The bottle will have a venting system to release the
evaporating nitrogen. Heat transfer from the sides of the cryogenic
bottle will be considered. The top and the bottom surfaces of the
bottle are assumed to be well-insulated. The temperature of the inner
wall of the inner tube is assumed to be that of liquid nitrogen, namely
a negligible convection heat transfer resistance between the liquid
nitrogen and the inner wall of the inner tube.
Heat transfer occurs from the environment to the nitrogen
under steady-state conditions and in one-dimensional cylindrical
coordinates. The heat transfer from the sides of the tube can be
calculated by using the following series circuit:
Q = (Tenvironment Tliquid nitrogen)/ Rij

(8-1)

77

Everyday Heat Transfer Problems

where Q is the steady-state heat transferred from the environment


to the liquid nitrogen in Watts, T is the temperature in K, and Rij
represents the thermal resistances on the thermal path in series
between Tenvironment and Tliquid nitrogen.
R21 is the conduction heat transfer thermal resistance between
the inner surface of the inner tube, radial location (1), to the outer
surface of the inner tube, radial location (2); in other words, thermal
resistance through the thickness of the inner tube.
R21 = ln(r2/r1)/(2Lkss)

(8-2)

where
ln is natural logarithm of argument (r2/r1),
r1 = inner radius of inner tube in m,
r2 = outer radius of inner tube in m,
L = height of the tube in m, and
kss = thermal conductivity of stainless steel inner tube in W/m-K.
R32 is the radiation heat transfer thermal resistance between the
inner surface of the outer insulation tube, radial location (3), and the
outer surface of the inner tube, radial location (2); in other words,
through the vacuum gap.
R32 = 1/(2r3Lhr)

(8-3)

where
r3 = inner radius of outer insulation tube in m,
hr = (T34 T24)/(T3 T2) in W/m2-K,
= emissivity of inner surface of outer insulation tube,
= Stefan-Boltzmann constant, namely 5.67 108 W/m2-K4
T3 = temperature of inner surface of outer insulation tube in K,
T2 = temperature of outer surface of inner tube in K
R43 is the conduction heat transfer thermal resistance between the
outer surface of the outer insulation tube, radial location (4) and the
inner surface of the outer insulation tube, radial location (3).

78

Evaporation Of Liquid Nitrogen In A Cryogenic Bottle

R43 = ln(r4/r3)/(2Lkinsulation)

(8-4)

where
r4 = outer radius of outer insulation tube in m and
kinsulation = thermal conductivity of insulation material of the outer
tube in W/m-K.
Renvironment-4 is the convection heat transfer thermal resistance between
the environment and the outer surface of outer insulation tube, radial
location (4). Radiation heat transfer between the environment and
the outer surface of outer insulation tube is assumed to be negligible.
Renvironment-4 = 1/(2r4ho)

(8-5)

where
ho = convection heat transfer coefcient between the environment
and the outer surface of the outer tube in W/m2-K.
Since hr depends on unknown temperatures T3 and T2, an iterative
method is used to calculate these temperatures, and thereafter hr
(i.e., the radiation heat transfer coefcient between the inner surface
of the outer insulation tube and the outer surface of the inner tube, or
through the vacuum gap). All calculations are done using the Kelvin
temperature scale due to fourth power calculations in the radiation
heat transfer coefcient. The thermophysical properties used in these
calculations are assumed to be constants.
Nominal values of the variables used in this problem are as follows:
Tenvironment = 293 K
Tliquid nitrogen = 77 K
r1 = 0.1 m
r2 = 0.106 m, namely inner tube stainless steel wall thickness is 6 mm
r3 = 0.116 m, namely vacuum gap is 0.01 m
r4 = 0.166 m, namely outer tube insulation thickness is 0.05 m
L = 0.8 m
kss = 10 W/m-K

79

Everyday Heat Transfer Problems

kinsulation = 0.002 W/m-K


ho = 10 W/m2-K, namely in natural convection regime
= 0.5, for inner surface of outer insulation tube
Hfg = 2 105 J/kg, latent heat of evaporation for liquid nitrogen
liquid nitrogen = 800 kg/m3
It is assumed that the initial volume of liquid nitrogen in the bottle
is 5 liters, or that it contains 4 kg, Mliquid nitrogen, of liquid nitrogen. The
dependent variable in the calculations is the time, t in seconds, that
it takes to completely evaporate the liquid nitrogen. Heat losses,
Q, from the cryogenic bottle, times time, t, should equal the total
latent heat of vaporization for the liquid nitrogen, Mliquid nitrogen Hfg, for
complete evaporation.
t = Mliquid nitrogen Hfg/Q

(8-6)

It takes 198 hours, or 8.25 days, for 5 liters of liquid nitrogen to


evaporate under the above nominal conditions.
Unknown temperature T3, temperature of the inner surface of the
outer insulation tube, is obtained by iteration, as shown in Figure 8-1.

Temperature, K

180
160
140
120
100
21

19

17

15

13

11

80
Number Of Iterations

Figure 8-1 Convergence of temperature, T3, at the inner surface of an outer

insulation tube

80

Evaporation Of Liquid Nitrogen In A Cryogenic Bottle

Temperature calculations converge to a nal value after ten iterations,


with less than 1% error.
Liquid nitrogen evaporation time is very sensitive to the cryogenic
bottle's insulation layer thermal conductivity, kinsulation. Five liters
of liquid nitrogen evaporation time versus insulation layer thermal
conductivity is given in Figure 8-2. The sensitivity of evaporation
time increases rapidly as thermal conductivity decreases below
0.001 W/m-K, as shown in Figure 8-3.
The second dominant heat transfer mechanism is the radiation
heat transfer R32. The sensitivity of evaporation time to the emissivity,
, of the inner surface of the outer insulation tube, is shown in
Figure 8-4. Evaporation time increases as emissivity of the insulation
inner surface, and therefore the radiation heat transfer, decreases.
When the nominal values of the variables given above are varied
10%, the results shown in Table 8-1 are obtained. Sensitivities of
the time that it takes for ve liters of liquid nitrogen to evaporate to a
10% change in independent variables are given in descending order

Evaporation Time, hr

1500

1000

500

0
0

0.002

0.004

0.006

0.008

0.01

Insulation Layer Thermal Conductivity, kinsulation, W/m-K

Figure 8-2 All-liquid nitrogen evaporation time versus bottle insulation

layer thermal conductivity

81

Everyday Heat Transfer Problems

t/ kinsulation, hr-m-K/W

0.0E+00

1.0E+06

2.0E+06

3.0E+06
0

0.001

0.002

0.003

0.004

0.005

Insulation Layer Thermal Conductivity, kinsulation, W/mK

Figure 8-3 Sensitivity of all-liquid nitrogen evaporation time to bottle

insulation layer thermal conductivity

t/

100

200

300

400

0.2

0.4

0.6

0.8

Emissivity, , Of Inner Surface Of Outer Insulation Tube

Figure 8-4 Sensitivity of all liquid nitrogen evaporation time to emissivity of

the inner surface of an outer insulation tube

82

Evaporation Of Liquid Nitrogen In A Cryogenic Bottle

Table 8-1 Effects of 10% change in the nominal values of


variables to time for 5 liters of liquid nitrogen to
evaporate

Variable

Nominal
Value

Change in Time
Change in Time
For 5 Liters Of
For 5 Liters Of
Liquid Nitrogen
Liquid Nitrogen
To Evaporate For To Evaporate For
A 10% Decrease
A 10% Increase
In Nominal Value In Nominal Value

0.8 m

+11.111%

9.091%

0.002 W/m-K

+9.323%

7.694%

r1

0.1 m

+8.062%

6.931%

r4 r3 (outer
tube insulation
thickness)

0.05 m

7.212%

+6.947%

0.5

+1.671%

1.434%

Tenvironment

293 K

+1.064%

1.043%

r3 r2
(vacuum gap)

0.01 m

+0.751%

0.740%

r2 r1 (stainless
steel inner tube
wall thickness)

0.006 m

+0.449%

0.445%

ho

10 W/m2-K

+0.0315%

0.0258%

kss

10 W/m-K

+0.00006%

0.000049%

L
kinsulation

of importance, and they are applicable only around the nominal values
assumed for this study.
The time required for ve liters of liquid nitrogen to evaporate
is most sensitive to cryogenic tube height, insulation layer thermal
conductivity, inside radius of the inner tube, and outer tube insulation
thickness. The second tier of sensitivities are an order of magnitude
less; emissivity of inner surface of the outer insulation tube,
temperature of environment, vacuum gap thickness, and stainless
steel inner tube wall thickness. Time for ve liters of liquid nitrogen to

83

Everyday Heat Transfer Problems

evaporate is least sensitive to the convection heat transfer coefcient


between the environment and the outer surface of the outer tube, and
the thermal conductivity of the stainless steel inner tube.
R43, conduction heat transfer thermal resistance between the
outer surface of the outer insulation tube and the inner surface of the
outer insulation tube, dominates the results in this case. The second
dominant heat transfer mechanism is R32, radiation heat transfer
thermal resistance between the inner surface of the outer insulation
tube and the outer surface of the inner tube; in other words, the
vacuum gap.
It is important to remember that these calculated sensitivities are
good only around the nominal values assumed for this case. Changing
these nominal values will change the magnitudes and orders of these
sensitivities, a result of the non-linear form of governing heat transfer
equations.

84

CHAPTER

9
STRESS IN A PIPE
THERMAL

hermal stresses generated by temperature variations in the


wall of a pipe have been studied extensively in Reference by
Timoshenko, S. and J. N. Goodier [17]. The stress, strain, radial
displacement relationships in cylindrical coordinates are treated
in detail in Reference [17]. To calculate the thermal stresses in
a pipe wall, the temperature distribution in the pipe wall has to
be known. The temperature distribution in the radial direction,
R, can be obtained from a steady-state, one-dimensional heat
conduction equation in cylindrical coordinates. By assuming constant
thermophysical properties and no heat sources in the pipe wall, the
heat conduction equation for the temperature distribution, T, is:
d2T/dR2 + (1/R)dT/dR = 0

(9-1)

If the temperatures at the inner surface, Ti, and the outer surface,
To, of the pipe wall are known, Eq. (9-1) can be solved by using the
following boundary conditions:
T = Ti at R = Ri

(9-2)

85

Everyday Heat Transfer Problems

and
T = To at R = Ro

(9-3)

Equation (9-1) can be integrated and the following temperature


distribution in the pipe wall can be obtained by applying the boundary
conditions in Eqs. (9-2) and (9-3):
(T To) = (Ti To) [ln(Ro /R) / ln(Ro /Ri)]

(9-4)

Derivation of thermal stress integral equations for R, , Z, thermal


stresses in three cylindrical coordinates, are detailed in Reference [17].
The thermal stress integral equations for R, and Z were developed
by applying three conditions to the stress-strain relationships. One of
the conditions results from the fact that the strain along the length of
a long pipe is zero. The integration constants are determined from two
boundary conditions. These boundary conditions come from the radial
stress, R, being zero at the inner and outer surfaces of the pipe wall.
The thermal stress integral equations are:
R = [E/(1 )](1/R2){[(R2 Ri2)/(Ro2 Ri2)]
}

(9-5)

= [E/(1 )](1/R2){[(R2 + Ri2)/(Ro2 Ri2)]


TR2}
Z = [E/(1 )]{[(2/(Ro2 Ri2)]

(9-6)
-T}

(9-7)

where
R, , and Z are the thermal stress distributions at the pipe wall in
MPa (Mega Pascals),
is the coefcient of thermal expansion of the pipe wall material in
m/m-C,
E is modulus of elasticity for the pipe wall material in MPa, and
is Poisson's ratio for pipe wall material.

86

Thermal Stress In A Pipe

The temperature distribution in Eq. (9-4) is substituted into


Eqs. (9-5) through (9-7) to determine thermal stress distributions in
three directions in cylindrical coordinates. If the heat is owing from
inside the pipe to the environment, maximum compressive stresses
occur at the inside radius of the pipe wall, and the maximum tensile
stresses occur at the outside radius of the pipe wall.
In the present sensitivity analysis, the maximum tensile stress
condition that occurs at the outside radius of the pipe wall will be
investigated.
In the heat transfer analysis, only convection and conduction heat
transfer mechanisms are considered. Radiation heat transfer from
the outer surface of the pipe to the environment is neglected. The
convection heat transfer per unit length from the uid in the pipe to
the inner surface of the pipe wall is:
Q = 2Ri hi (Tuid Ti)

(9-8)

The conduction heat transfer per unit length through the pipe wall
can be written as:
Q = 2k (Ti To)/ln(Ro/Ri)

(9-9)

The convection heat transfer per unit length from the outer surface
of the pipe wall to the environment is:
Q = 2Ro ho (To Tenvironment)

(9-10)

The heat is transferred in a series thermal circuit. Combining


Eqs. (9-8) through (9-10) demonstrates that heat transferred from
the uid in the pipe to the environment is:
Q = (Tuid Tenvironment)/[(1/2Rihi) + ln(Ro/Ri)/2k + (1/2Roho)] (9-11)
where
Q = heat transfer through pipe wall in W/m
Tuid = Mean temperature of uid in the pipe in C

87

Everyday Heat Transfer Problems

Tenvironment = Environment temperature outside the pipe in C


Ri = Pipe inside radius in meters
hi = Heat transfer coefcient between inner surface of pipe wall
and uid in pipe in W/m2-C
Ro = Pipe outside radius in meters
ho = Heat transfer coefcient between outer surface of pipe wall
and environment in W/m2-C
k = Thermal conductivity of pipe wall material in W/m-C
Maximum tensile stress due to the temperature distribution given
by Eq. (9-4) occurs at the outer radius of the pipe. At the outer radius
Ro, the maximum axial stress is the same as in the circumferential
direction, Z = . Eqs. (9-6) and (9-7) produce the same results
at R = Ri and at R = Ro. Also, it is important to remember from the
boundary conditions that R is zero at R = Ri and at R = Ro. Eq. (9-11)
can be combined with Eq. (9-6) to provide the thermal stress at R = Ro.
at R=Ro = { E Q/[4(1 )k]}{1 [2Ri2/(Ro2 Ri2)]ln(Ro/Ri)} (9-12)
Assumed nominal values for the independent variables for the
sensitivity analysis in Eqs. (9-11) and (9-12) are as follows:
Tuid = 50C
Tenvironment = 50C
Ri = 0.2 m
Ro = 0.3 m
hi = 500 W/m2-C
ho = 100 W/m2-C
= 2 105 m/m-C
E = 210000 MPa
= 0.3
k = 20 W/m-C
Thermal stress at the outside radius of the pipe wall is a linear
function of heat transferred from the uid to the environment, or
(Tuid Tenvironment), as shown in Figure 9-1. The slope of this curve is
0.828 MPa/C.

88

Thermal Stress In A Pipe

Thermal Stress At Ro,


MPa

120
100
80
60

Tenvironment= 50C

40
20
0

20

40

60

80

Temperature Of Fluid Flowing In Pipe, C

Figure 9-1 Thermal stress at outside radius of pipe versus temperature of

uid owing in pipe

As the pipe outside radius, Ro, increases for a constant wall


thickness of 0.1 m, the heat transfer increases, and therefore the
thermal stress increases, as shown in Figure 9-2. Thermal stress
sensitivity to pipe outside radius decreases as the pipe gets larger.
The effects of pipe wall thickness on thermal stress are studied
for a constant outside radius pipe, where Ro = 0.3 m, in Figure 9-3.
Thermal stress increases as pipe wall thickness increases, reaching
a maximum at around 0.15 to 0.2 meters of wall thickness. Thermal
stress then starts to decrease as pipe wall thickness further increases,
due to decreasing heat transfer.
The inside and outside heat transfer coefcients, hi and ho, affect
thermal stress similarly, as shown in Figures 9-4 and 9-5. Thermal
stress sensitivity to changes in the heat transfer coefcient are
signicant at smaller values of hi and ho.
When the nominal values of the variables given above are varied
10%, the results shown in Table 9-1 are obtained. Maximum thermal
stress at Ro sensitivities to a 10% change in the governing independent
variables are given in descending order of importance, and they are
applicable only around the nominal values assumed for this study.

89

Thermal Stress At Ro, MPa

Everyday Heat Transfer Problems

86
84
82
80
78
76
0.2

0.25

0.3

0.35

0.4

0.45

0.5

Pipe Outside Radius, Ro, m

Figure 9-2 Thermal stress at Ro versus pipe outside radius for a constant

wall thickness of 0.1 meters

Thermal Stress At Ro,


MPa

120

80

40

0
0

0.05

0.1

0.15

0.2

0.25

0.3

Pipe Wall Thickness, m

Figure 9-3 Thermal stress at Ro versus pipe wall thickness for a pipe outside

radius of 0.3 meters

90

Thermal Stress At Ro,


MPa

Thermal Stress In A Pipe

100
80
60
40
20
0

200

400

600

800

1000

Heat Transfer Coefficient, hi, W/m2-C

Figure 9-4 Thermal stress at Ro versus the heat transfer coefcient at inside

surface of pipe wall

Thermal Stress At Ro,


MPa

160
120
80
40
0

200

400

600

800

1000

Heat Transfer Coefficient, ho, W/m2-C

Figure 9-5 Thermal stress at Ro versus the heat transfer coefcient at

outside surface of pipe wall

91

Everyday Heat Transfer Problems

Table 9-1 Effects of 10% change in nominal values of


variables to maximum thermal stress at Ro

Variable
Pipe Wall
Thickness, Ro-Ri

Nominal
Value

Change In
Change In
Thermal Stress Thermal Stress
At Ro For A 10% At Ro For A 10%
Decrease In
Increase In
Nominal Value
Nominal Value

0.1 m

+10.049%

13.011%

2 105 m/m-C

10%

+10%

210000 MPa

10%

+10%

100C

10%

+10%

20 W/m-C

+7.311%

6.378%

ho

100 W/m2-C

5.502%

+5.002%

0.3

4.110%

+4.478%

hi

500 W/m -C

1.717%

+1.450%

Ro

0.3 m

+0.920%

1.248%

Tuid-Tenvironment

Thermal stress at the pipe outside radius is most sensitive to


changes in pipe wall thickness, the coefcient of thermal expansion,
modulus of elasticity, temperature potential between the uid and
the environment, and thermal conductivity of pipe material. The
second tier of independent variables affecting thermal stress at
the pipe outside radius are the heat transfer coefcient between
outside surface of pipe wall and environment, and Poisson's ratio
of pipe material. Thermal stress at the pipe outside radius is least
affected by changes to the heat transfer coefcient between inside
surface of pipe wall and uid in pipe, and outer radius of pipe. This
variable sensitivity order is applicable only around the nominal values
assumed for this case.

92

CHAPTER

10

HEAT

TRANSFER IN A
PIPE WITH UNIFORM
HEAT GENERATION IN
ITS WALLS
U

nder steady-state conditions, constant thermophysical properties


and uniform heat generation in the walls, the one-dimensional
conduction heat transfer equation in radial direction of a pipe can be
written as (see Reference by Carslaw, H. S. and J. C. Jaeger [17]):
d2T/dR2 + (1/R)dT/dR + Q/k = 0

(10-1)

where the radial heat ux is positive in the negative radial direction


(towards the center of the pipe), Q is uniform heat generation in the
pipe walls in W/m3, and k is pipe wall thermal conductivity in W/m-C.
This differential Eq. (10-1) can be solved for the radial temperature
distribution in the pipe wall by specifying the pipe wall temperatures
with the inner and outer wall radii as boundary conditions.
T = Ti at R = Ri

(10-2)

T = To at R = Ro

(10-3)

and

93

Everyday Heat Transfer Problems

The radial temperature distribution in the pipe wall which satises


boundary condition Eqs (10-2) and (10-3) is:
T(R) = To + (QRo2/4k)[1 (R2/Ro2)] {(To Ti) + (QRo2/4k)
(10-4)
[1 (Ri2/Ro2)]} [ln(Ro/R)/ln(Ro/Ri)]
In this sensitivity study, convection heat transfer and conduction
heat transfer equivalence at the pipe wall surfaces at R = Ri and
at R = Ro will be used to determine the thermal energy generation
requirements in the pipe wall. Energy balances at the inside and
outside surfaces of the pipe wall provide:
(2RiL) k dT/dR = (2RiL) hi (Tuid Ti) at R = Ri

(10-5)

and
(2RoL) k dT/dR = (2RoL) ho (To Tenvironment) at R = Ro (10-6)
where hi is the convection heat transfer coefcient at inside surface
of the pipe in W/m2-C, Tuid is mean uid temperature in C in the pipe,
ho is the convection heat transfer coefcient at outside surface of the
pipe in W/m2-C, and Tenvironment is environmental temperature in C.
L is the length of the pipe in meters and it cancels out from the
energy balance equations.
Obtaining dT/dR at R = Ri from Eq. (10-4), the governing energy
balance Eq. (10-5) at R = Ri becomes:
hi (Tuid Ti) = 0.5 Q Ri
[k/Ro ln(Ro/Ri)][(To Ti) + (Q/4k)(Ro2 Ri2)] (10-7)
Obtaining dT/dR at R = Ro from Eq. (10-4), the governing energy
balance Eq. (10-6) at R = Ro becomes:
ho (To Tenvironment)
= 0.5 Q Ro [k/Ro ln(Ro /Ri)][(To Ti) + (Q/4k)(Ro2 Ri2)] (10-8)

94

Heat Transfer In A Pipe With Uniform Heat Generation

The uniform thermal energy generation required per unit length


in the pipe wall, Q, is treated as the dependent variable, and its
sensitivity to other independent variables is analyzed for a specied
inner surface temperature, Ti. Eqs. (10-7) and (10-8) can be
combined to eliminate To and obtain a relationship for thermal energy
requirement as a function of eight independent variables.
Nominal values of independent variables for the present sensitivity
analysis are as follows:
Ri = 0.10 m
Ro = 0.11 m
k = 15 W/m-C
hi = 30 W/m2-C
Tuid = 50C
ho = 10 W/m2-C
Tenvironment = 10C
Ti = 40C
With these nominal values, the uniform thermal energy generation
required is 23,530 W/m3. Uniform thermal energy generation
requirements behave non-linearly with pipe inner radius variations,
as shown in Figure 10-1. The uniform thermal energy generation
requirement increases with decreasing pipe wall thickness. As the
wall thickness decreases, the conduction heat transfer resistance
decreases, and it gets difcult to keep the inner radius pipe wall
temperature at a constant 40C.
Uniform thermal energy generation requirements increase as pipe
outer radius increases, as shown in Figure 10-2. Uniform thermal
energy generation requirements decrease with increasing pipe wall
thickness. As wall thickness increases, conduction heat transfer
resistance increases, and it becomes easier to keep the inner radius
pipe wall temperature at a constant 40C.
Thermal energy requirement versus pipe wall thermal conductivity
is shown in Figure 10-3. Thermal energy generation requirement is
mostly a constant for thermal conductivities above 20 W/m-C, and it
begins rapidly to decrease as the wall material becomes an insulator.

95

Energy Generation, W/m3

Everyday Heat Transfer Problems

250000
200000
150000
100000
50000
0
0.09

0.095

0.1

0.105

0.11

Pipe Inner Radius, m

Figure 10-1 Thermal energy generation requirement versus pipe inner

Energy Generated, W/m3

radius

200000
150000
100000
50000
0
0.1

0.105

0.11

0.115

0.12

Pipe Outer Radius, m

Figure 10-2 Thermal energy generation requirement versus pipe outer

radius

96

Energy Generated, W/m3

Heat Transfer In A Pipe With Uniform Heat Generation

24000
23000
22000
21000
20000

20

40

60

80

100

Pipe Wall Thermal Conductivity, W/m-C

Figure 10-3 Thermal energy generation requirement versus pipe wall

thermal conductivity

Thermal energy generation requirement decreases as the


convection heat transfer coefcient at pipe inner surface increases,
as shown in Figure 10-4. The slope of the curve is 955.8 C/m, and
no thermal energy generation is required at hi = 54.6 W/m2-C. At
higher convection heat transfer coefcients at the pipe inner surface,
thermal energy has to be taken out of the pipe walls in order to keep
the pipe inner walls at 40C.
Thermal energy generation requirement increases as the
convection heat transfer coefcient at pipe wall outside surface
increases, as shown in Figure 10-5. The slope of the curve is 5062.1 C/m,
and no thermal energy generation is required at ho = 5.27 W/m2-C,
namely in the natural convection region. At lower convection heat
transfer coefcients at the pipe wall outside surface, thermal energy
has to be taken out of the pipe walls in order to be able to keep the
pipe inner walls at 40C.
The present case assumes a fully developed ow inside the
pipe. Thermal energy generation requirement versus mean uid
temperature inside the pipe is shown in Figure 10-6. Thermal energy

97

Energy Generated, W/m3

Everyday Heat Transfer Problems

60000
40000
20000
0
20000
40000

20

40

60

80

100

Convection Heat Transfer Coefficient At Pipe Wall


Inside Surface, W/m2-C

Figure 10-4 Thermal energy generation requirement versus convection heat

Energy Generated, W/m3

transfer coefcient at pipe wall inside surface

500000
400000
300000
200000
100000
0
100000

100
20
40
60
80
Convection Heat Transfer Coefficient At Pipe Wall
Outside Surface, W/m2-C

Figure 10-5 Thermal energy generation requirement versus convection heat

transfer coefcient at pipe wall outside surface

98

Heat Transfer In A Pipe With Uniform Heat Generation

Energy Generated, W/m3

150000
100000
50000
0
50000
100000
150000

20
40
60
80
Mean Fluid Temperature Inside The Pipe, C

100

Figure 10-6 Thermal energy generation requirement versus mean uid

temperature inside the pipe

generation requirement decreases as the mean uid temperature


increases. The slope of the curve is 2867.4 W/m3-C, and no thermal
energy generation is required at Tuid = 58.2C. For higher mean uid
temperatures inside the pipe, thermal energy has to be taken out from
the pipe walls in order to be able to keep the pipe inner walls at 40C.
Thermal energy generation requirement versus environmental
temperature is shown in Figure 10-7. Thermal energy generation
requirement increases as the environment temperature decreases. The
slope of the curve is 1044.1 W/m3-C, and no thermal energy generation
is required at Tenvironment = 12.5C. Pipe walls have to be cooled above
the 12.5C environmental temperature in order to keep Ti at 40C.
Thermal energy generation requirement versus pipe inside surface
temperature, Ti, is shown in Figure 10-8. Thermal energy generation
requirement increases as the pipe inner surface temperature
requirement increases. The slope of the curve is 3911.5 W/m3-C,
and no thermal energy generation is required at Ti = 34C.
When the nominal values of the independent variables given
above are varied 10%, the results shown in Table 10-1 are obtained.

99

Energy Generated, W/m3

Everyday Heat Transfer Problems

60000
40000
20000
0
20000
40000
40

20

20

40

60

Environment Temperature, C

Figure 10-7 Thermal energy generation requirement versus environmental

temperature

Energy Generated, W/m3

250000
200000
150000
100000
50000
0
50000
100000
150000

30
60
Pipe Inner Surface Temperature, C

90

Figure 10-8 Thermal energy generation requirement versus pipe inner

surface temperature

100

Heat Transfer In A Pipe With Uniform Heat Generation

Table 10-1 Effects of a 10% change in nominal values of


variables to thermal energy generation requirement

Nominal
Value

Change In
Required
Thermal Energy
Generation For
A 10% Decrease
In Nominal Value

Change In
Required
Thermal Energy
Generation For
A 10% Increase
In Nominal Value

Ti

40C

66.49%

+66.49%

Tuid

50C

+60.93%

60.93%

Ro

0.11 m

+104.35%
@ 5% decrease
in nominal value

45.38%

Ri

0.10 m

41.75%

+84.65%
@ 5% increase
in nominal value

ho

10 W/m2-C

22.08%

+22.06%

30 W/m -C

+12.19%

12.19%

10C

+4.44%

4.44%

15 W/m-C

0.13%

+0.11%

Variable

hi
Tenvironment
k

Thermal energy requirement sensitivities to a 10% change in


the governing variables are given in descending order of importance,
and they are applicable only around the nominal values assumed for
this study.
Thermal energy generation required is most sensitive to pipe
inner surface temperature, mean uid temperature inside the pipe,
pipe outer radius and pipe inner radius. Convection heat transfer
coefcients at the outside and inside surfaces of the pipe wall, and
environmental temperature, are the next independent variables in
order of sensitivity. Thermal energy generation requirement has
the lowest sensitivity to pipe wall thermal conductivity around the
assigned nominal values for this case.

101

CHAPTER

HEAT

11

TRANSFER
IN AN ACTIVE
INFRARED SENSOR
I

n an active infrared sensor, the surface temperature is held


constant during the measurement process by providing controlled
energy to the sensor's surface. The sensor surface receives radiation
heat transfer energy from the surface of the object being measured.
The sensor also loses thermal energy to its environment. The
temperature at the sensor's surface can be analyzed by using
unsteady-state and one-dimensional heat transfer rate equations in
rectangular coordinates.
Unsteady-state heat transfer in an active infrared sensor is detailed
in the Reference by J. Fraden [3]. The energy balance of a sensor
element can be written as follows:
Change in internal energy of the sensor with respect to time =
Control energy supplied to regulate the surface temperature of
the sensor Energy lost from the sensor to the environment by
conduction and by convection heat transfer + Net radiation heat
transferred from the object being measured to the sensor
Net radiation heat transfer between the object and the sensor
is assumed to occur between two gray bodies that are opaque to

103

Everyday Heat Transfer Problems

radiation. If that is true, then the emissivity, , and reflectivity, ,


characteristics of a gray surface have the following relationship:
+=1

(11-1)

In this analysis, it is assumed that the emissivity and reflectivity of


the object and the sensor surfaces are constants in the infrared region
of the electromagnetic radiation spectrum.
The radiation emitted by the object to the sensor can be written as:
Qradiation emitted by object = Aobject T4object

(11-2)

Some of the irradiation reaching the surface of the sensor is


reflected due to the reflectivity of the sensor surface. The reflected
portion of the radiation emitted by the object can be written as:
Qradiation reflected by sensor = sensor (Aobject T4object)

(11-3)

Eqs. (11-1) through (11-3) can be combined to get the net


radiation emitted by the object to the sensor:
Qnet radiation emitted by object = Asensor object T4object

(11-4)

In a similar fashion, the net radiation emitted by the sensor to the


object can be obtained:
Qnet radiation emitted by sensor = Aobject sensor T4sensor

(11-5)

The net radiation heat transfer between the object and the sensor
is determined by combining Eqs. (11-4) and (11-5):
Qnet radiation = Asensor object (T4object T4sensor)

(11-6)

Control energy supplied to regulate the surface temperature of


the sensor is assumed to be in the form of I2R. Energy lost from
the sensor to the environment by conduction and convection heat

104

Heat Transfer In An Active Infrared Sensor

transfer can be written as a rate equation, where the two heat transfer
mechanisms act in series:
Qconduction + convection = (1/RT) (Tsensor Tenvironment)

(11-7)

Energy balance for the sensor can be written as a first order and
non-linear differential equation:
cpV(dTsensor/d) = I2R (Tsensor Tenvironment)/Rtotal
+ Aobjectsensor (T4object T4sensor) (11-8)
where
cpV is the sensor thermal capacitance in W-s/K
dTsensor/d is the time rate of change of sensor temperature in K/s
I2R is control power supplied to regulate the surface temperature of
the sensor in W
Tenvironment is the environment temperature in K
Rtotal is the total heat transfer resistance between the sensor and
the environment due to conduction and convection heat transfer
in K/W
A is the sensor area in m2
= 5.6710-8 W/m2-K4 is the Stefan-Boltzmann constant
object is the surface emissivity of the object being measured
sensor is the surface emissivity of the sensor
Tobject is the temperature of the object being measured in K
Tsensor is the temperature of the sensor surface in K
For the steady-state sensor temperature, Eq. (11-8) can be
rewritten by eliminating the left-hand side of the equation, which
means that the change in internal energy of the sensor with respect to
time becomes negligible. Tsensor is calculated from the following quartic
equation, by trial and error:
T4sensor + C1 Tsensor = C2

(11-9)

C1 = 1/(RT Aobjectsensor)

(11-10)

where

105

Everyday Heat Transfer Problems

and
C2 = (I2R/Aobjectsensor) + (Tenvironment/Rtotal Aobjectsensor) + T4object
(11-11)
The transient solutions to Eq. (11-8) are obtained from the following
explicit finite difference equation, using small time intervals :
cpV(Tsensor @ (i+1) Tsensor @ i /) = I2R (Tsensor @ i Tenvironment)/Rtotal
(11-12)
+ Aobjectsensor (T4object T4sensor @ i)
For the present sensitivity analysis, nominal values of the
independent variables are assumed to be as follows:
cpV = 0.014 W-s/K
I2R = 0.1 W
Tenvironment = 20C (293 K)
Rtotal = 100 K/W
A = 0.0001 m2 (a 1 cm 1 cm sensor surface area)
object = 0.9
sensor = 0.8
The sensitivities are analyzed for three different object temperatures,
namely 100C (373 K), 500C (773 K) and 1000C (1273 K).
Sensor temperature as a function of time for three different
object temperatures is given in Figures 11-1, 11-2, and 11-3. As the
temperature of the object increases, the sensor time constant, or
the time the sensor reaches 63.2% of its steady-state temperature,
decreases. Also as the sensor emissivity increases, the sensor time
constant decreases.
As the object temperature approaches the environmental
temperature, the sensor temperature deviates from the object
temperature because fixed nominal values are used for the analysis
of this heat transfer problem. The control circuit of the sensor
has to respond and change control energy, I2R, in order to achieve
accurate results. The details of the control circuit are explained in the
Reference by J. Fraden [3].

106

Heat Transfer In An Active Infrared Sensor

200

Tsensor, C

150

Emissivity
Sensor=0.9
Emissivity
Sensor=0.8
Emissivity
Sensor=0.7

100
50
0
0

2
Time, s

Figure 11-1 Sensor temperature versus time for different sensor emissivities

and for an object at 1000C

200

Tsensor, C

150

Emissivity
Sensor=0.9
Emissivity
Sensor=0.8
Emissivity
Sensor=0.7

100
50
0
0

Time, s

Figure 11-2 Sensor temperature versus time for different sensor emissivities

and for an object at 500C

107

Everyday Heat Transfer Problems

Tsensor, C

40
35
30
25
20
15
10
5
0

Emissivity
Sensor=0.9
Emissivity
Sensor=0.8
Emissivity
Sensor=0.7
0

2
Time, s

Figure 11-3 Sensor temperature versus time for different sensor emissivities

 

and for an object at 100C












 

 

 




 

Figure 11-4 Sensor temperature versus time for different total heat transfer

resistances between the sensor and the environment, due to


conduction and convection heat transfer for an object at 1000C
temperature

108

Heat Transfer In An Active Infrared Sensor

Table 11-1 Thermal time constant change due to a 10% change


in variables around the nominal values for an
object temperature of 1000C

Variable
Sensor Thermal
Capacitance, cpV

Nominal
Value

Thermal Time
Thermal Time
Constant Change Constant Change
Due To A 10%
Due To A 10%
Decrease In
Increase In
Nominal Value
Nominal Value

0.014 W-s/K

20.39%

+10.38%

Tobject

1000C
(1273 K)

+16.73%

15.38%

0.0001 m2

+8.12%

7.44%

object

0.9

+8.12%

7.44%

sensor

0.8

+8.12%

7.44%

Rtotal

100 K/W

3.04%

+2.31%

Tenvironment

20C
(293 K)

0.15%

+0.15%

0.1 W

+0.049%

0.049%

I 2R

Another variable affecting the performance of the sensor is the


heat loss from the sensor to the environment due to conduction
and convection heat transfer. The sensitivity of the sensor response
to total heat transfer resistance between the sensor and the
environment, due to conduction and convection heat transfer, is given
in Figure 11-4 for an object at 1000C temperature.
As the total thermal resistance between the sensor and the
environment increases, as seen in Figure 11-4, the sensor is better
insulated for losses due to conduction and convection heat transfer. In
such cases, the sensor temperature approaches the object temperature.
A ten percent variation in variables around the nominal values
given above produces the sensitivity results given in Tables 11-1 and
11-2, for thermal time constant and for steady-state temperature,
respectively, for an object that is at 1000C.

109

Everyday Heat Transfer Problems

Table 11-2 Steady-state sensor temperature change due to


a 10% change in variables around the nominal
values for an object temperature of 1000C

Variable

Nominal
Value

Steady-State
Steady-State
Sensor
Sensor
Temperature
Temperature
Change Due To
Change Due To
A 10% Decrease A 10% Increase
In Nominal Value In Nominal Value

Tobject

1000C
(1273 K)

18.33%

+18.22%

Rtotal

100 K/W

4.13%

+3.48%

0.0001 m

4.07%

+3.44%

object

0.9

4.07%

+3.44%

sensor

0.8

4.07%

+3.44%

20C
(293 K)

0.11%

+0.11%

0.1 W

0.054%

+0.054%

0.014 W-s/K

0%

0%

Tenvironment
I 2R
Sensor Thermal
Capacitance, cpV

A change in sensor thermal capacitance has the most significant


effect on the thermal time constant in this region of operation.
The thermal time constant experiencees non-linear sensitivity
behaviors from all the variables in the present region of operation,
with the exception of the environmental temperature and the I2R
power input to the sensor. The thermal time constant experiencees
similar magnitude sensitivity behaviors from the sensor area, sensor
emissivity, and object emissivity variables, as expected.
A change in the object temperature has the most significant effect
on the steady-state sensor temperature in this region of operation.
The next set of independent variables that most affect the sensor
temperature are the total heat transfer resistance between the sensor

110

Heat Transfer In An Active Infrared Sensor

Table 11-3 Thermal time constant change due to a 10%


change in variables around the nominal values for
an object temperature of 100C

Variable
Tenvironment
Sensor Thermal
Capacitance, cpV
RT

Thermal Time
Constant
Change Due
To A 10%
Decrease In
Nominal Value

Thermal Time
Constant
Change Due
To A 10%
Increase In
Nominal Value

20C
(293 K)

14.97%

+13.16%

0.014 W-s/K

10.33%

+10.39%

100 K/W

9.90%

+9.83

Nominal
Value

0.0001 m

0.45%

+0.45%

object

0.9

0.45%

+0.45%

sensor

0.8

0.45%

+0.45%

0.1 W

0.028%

+0.028%

100C
(373 K)

+0.023%

0.023%

IR
Tobject

and the environment, the sensor area, and surface emissivities of the
sensor and the object. Sensor thermal capacitance does not affect
the steady-state sensor temperature, as expected. The steady-state
sensor temperature experiences non-linear sensitivity behaviors
from all the variables except the environmental temperature and
the I2R power input to the sensor, in the present region of operation.
The steady-state sensor temperature experiences similar magnitude
sensitivity behaviors from the sensor area, sensor emissivity, and
object emissivity variables, as expected.
A similar sensitivity analysis is performed for a low object
temperature case, namely Tobject = 100C, and the results are given in
Tables 11-3 and 11-4.

111

Everyday Heat Transfer Problems

Table 11-4 Steady-state sensor temperature change due to a


10% change in variables around the nominal values
for an object temperature of 100C

Variable

Nominal
Value

Steady-State
Sensor
Temperature
Change Due To
A 10% Decrease
In Nominal
Value

Steady-State
Sensor
Temperature
Change Due To
A 10% Increase
In Nominal
Value

Tenvironment

20C
(293 K)

5.58%

+5.57%

RT

100 K/W

3.99%

+3.95%

0.1 W

2.79%

+2.79%

100C
(373 K)

2.27%

+2.46%

0.0001 m2

+1.18%

1.19%

IR
Tobject
A
object

0.9

+1.18%

1.19%

sensor

0.8

+1.18%

1.19%

0.014 W-s/K

0%

0%

Sensor Thermal
Capacitance, cpV

Temperature of the environment and thermal resistance to heat


loss from the sensor by conduction and convection to the environment
become the prominent variables in this low object-temperature
application. Both the thermal time constant and the steadystate temperature of the sensor are very sensitive to variations in
the temperature of the environment and the total thermal resistance.
It is important to remember that these calculated sensitivities are
good only around the nominal values assumed for this case. Changing
these nominal values will change magnitudes and orders of these
sensitivities, resulting from the non-linear form of governing heat
transfer equations.

112

CHAPTER

COOLING

OF A CHIP

12

ower dissipation in electronic chips is a challenging heat transfer


phenomenon, as the chips get smaller and smaller. Most chips
or chip sets use copper or aluminum heat sinks to enhance the heat
transferred out. These heat sinks are attached in a variety of ways to
the chips in order to minimize the thermal resistance between the chip
and the heat sink. In this analysis, a chip in the shape of a rectangular
box, 10 mm 10 mm 0.1 mm, is considered. The chip is attached to
its copper heat sink, also in the shape of a rectangular box, 10 mm
10 mm 10 mm, by a thermally conductive epoxy of 10 m thickness.
Transient heat transfer which occurs during the cooling of a chip
can generally be solved by using the same energy balance equation
as in Chapter 11, without the radiation heat transfer effects. The
temperature of a chip can be investigated by using unsteady-state
and one-dimensional heat transfer rate equations in rectangular
coordinates. Energy balance for a chip can be written as follows:
Change in internal energy of the chip with respect to time =
Power generated by the chip Energy lost from the chip to
its heat sink and to the environment by conduction and by
convection heat transfer.
113

Everyday Heat Transfer Problems

This energy balance can be written as a first order and linear


differential equation:
cpV(dTchip/d) = P (Tchip Tenvironment)/RTotal

(12-1)

where
cpV is the chip thermal capacitance in W-s/C
dTchip/d is the time rate of change of chip temperature in C/s
P is the energy generated in the chip in W
Tenvironment is the environment temperature in C
RTotal is the total heat transfer resistance between the chip, its heat
sink and the environment due to conduction and convection heat
transfer in C/W
Tchip is the temperature of the chip in C.
Since the governing equation is linear in temperature, centigrade
dimension is used instead of Kelvin. Total heat transfer resistance
between the chip, its heat sink and the environment has two
parallel components. One component is convection heat transfer
resistance between the chip and the environment. The other is the
heat loss in series circuit from the chip through the adhesion layer
by conduction, from the adhesion layer through the heat sink by
conduction, and from the heat sink to the environment by convection.
The heat transfer resistance between the surface of the chip and the
environment can be written as:
Rchip to environment = 1/(hAchip for convection)

(12-2)

The series circuit heat transfer resistance between the chip and
the environment going through the adhesive and the heat sink can be
written as:
Rchip through heat sink = [Ladhesive/(kadhesiveAadhesive for conduction)]
+ [Lheat sink/(kheat sinkAheat sink for conduction)]
(12-3)
+ [1/(hAheat sink for convection)]

114

Cooling Of A Chip

The total heat transfer resistance from the chip is a parallel


combination of Eqs. (12-2) and (12-3):
(1/RTotal) = (1/Rchip to environment) + (1/Rchip through heat sink)

(12-4)

The solution to the governing first order differential Eq. (12-1) can
be written as follows by using the initial condition of Tchip = Tenvironment:
Tchip = Tenvironment + PRTotal [1 exp(/cpVRTotal)]

(12-5)

where cpVRTotal is the thermal time constant for the chip, which is the
product of its thermal capacitance and its total thermal resistance.
Under steady-state conditions, the solution in Eq. (12-5) provides
the allowable chip power dissipation as follows:
P = (Tchip Tenvironment)/RTotal

(12-6)

For the present sensitivity analysis, the nominal values of the above
independent variables are assumed to be as follows:
cpV = 0.0197 W-s/C
(assuming a silicon dioxide chip with 10 mm 10 mm 0.1 mm
dimensions)
h = 200 W/m2-C (convection heat transfer coefficient at the chip
and heat sink surfaces)
Tenvironment = 30C
Tchip = 90C
Achip for convection = 0.000104 m2
Ladhesive = 0.00001 m
Aadhesive for conduction = Aheat sink for conduction = 0.0001 m2
Lheat sink = 0.01 m
Aheat sink for convection = 0.0005 m2 (assuming a copper alloy heat sink
with 10 mm 10 mm 10 mm dimensions)
kadhesive = 10 W/m-C (assuming silver epoxy adhesive)
kheat sink = 400 W/m-C

115

Everyday Heat Transfer Problems

200

RTotal, C/W

150
100
50
0
0

100

200

300

400

Convection Heat Transfer Coefficient, W/m2-C

Figure 12-1 Total chip heat transfer resistance versus the convection heat

transfer coefficient

Chip Power Disipation, W

Total heat transfer resistance between the chip, its sink and the
environment is given as a function of the convection heat transfer
coefficient in Figure 12-1. As the convection mechanism goes into
forced convection, and especially forced convection in liquids,

35
30
25

with heat sink

20
15

without heat sink

10
5
0
0

200

400

600

Convection Heat Transfer Coefficient,

800

1000

W/m2-C

Figure 12-2 Chip power dissipation versus the convection heat transfer

coefficient with and without a heat sink

116

Cooling Of A Chip

total heat transfer resistance decreases dramatically. This decrease


in total heat transfer resistance allows the chip to be operated at
higher power dissipation, as shown in Figure 12-2. Figure 12-2 also
shows the effects of using a heat sink. The slope of the chip's power
dissipation with a heat sink versus the heat transfer coefficient,
(P/h), is 0.033 m2-C. The slope of the chip's power dissipation
without a heat sink versus the heat transfer coefficient is much lower,
P/h = 0.012 m2-C.
A ten-percent difference in variables around the nominal values
given above produce the sensitivity results given in Table 12-1 for
the power dissipation capability of the chip. The results are given in
descending order from the most sensitive variable to the least.

Table 12-1 Steady-state chip power dissipation capability


change due to a 10% change in variables around
the nominal values

Variable
Tchip Tenvironment
h
Aheat sink for convection
Achip for convection

Nominal
Value

Steady-State
Chip Power
Dissipation
Capability
Change Due To
A 10% Decrease
In Nominal
Value

Steady-State
Chip Power
Dissipation
Capability
Change Due To
A 10% Increase
In Nominal
Value

60C

10%

+10%

9.81%

+9.77%

8.05%

+8.01%

1.76%

+1.76%

200 W/m-C
2

0.0005 m

0.000104 m

Aadhesive for conduction =


Aheat sink for conduction

0.0001 m2

0.23%

+0.19%

kheat sink

400 W/m-C

0.22%

+0.18%

Lheat sink

0.01 m

+0.20%

0.20%

kadhesive

10 W/m-C

0.009%

+0.007%

Ladhesive

0.00001 m

+0.008%

0.008%

117

Chip Time Constant, s

Everyday Heat Transfer Problems

3.5
3
2.5
2
1.5
1
0.5
0
0

100

200

300

400

Convection Heat Transfer Coefficient, W/m2-C

Figure 12-3 Chip thermal time constant versus the convection heat transfer

coefficient

Changes in chip temperature potential, convection heat transfer


coefficient and heat sink surface area for convection heat transfer
affect the chip power dissipation capability the most. Chip surface
area for convection heat transfer comes next in the order of
sensitivity. Changes in independent variables affecting the conduction
heat transfers, Aadhesive for conduction, Aheat sink for conduction, kheat sink Lheat sink,
kadhesive, and Ladhesive, contribute the least to sensitivities in chip power
dissipation. These sensitivity results shown in Table 12-1 are valid
only in the region of analysis for the governing independent variables.
The thermal time constant, cpVRTotal , is an important parameter in
chip design and testing. Thermal time constant, which is the product
of thermal capacitance and total thermal resistance, is analyzed as
a function of the convection heat transfer coefficient, and is given
in Figure 12-3. As the convection mechanism goes into forced
convection, and especially forced convection in liquids, the thermal
time constant decreases dramatically.
A ten-percent difference in variables around the nominal values
given above produces the sensitivity results given in Table 12-2 for the

118

Cooling Of A Chip

Table 12-2 Chip thermal time constant change due to a 10%


change in variables around the nominal values

Variable
h
cpV

Nominal
Value

Chip Thermal
Time Constant
Change Due To
A 10% Decrease
In Nominal
Value

Chip Thermal
Time Constant
Change Due To
A 10% Increase
In Nominal
Value

200 W/m-C

+10.88%

8.90%

0.0197 W-s/C

10%

+10%

0.0005 m

+8.76%

7.42%

0.000104 m2

+1.79%

1.79%

Aadhesive for conduction =


Aheat sink for conduction

0.0001 m2

+0.233%

0.190%

kheat sink

400 W/m-C

+0.223%

0.183%

Lheat sink

0.01 m

0.201%

+0.201%

kadhesive

10 W/m-C

+0.009%

0.007%

Ladhesive

0.00001 m

0.008%

+0.008%

Aheat sink for convection


Achip for convection

thermal time constant of the chip. The results are given in descending
order, from the most sensitive variable to the least.
Changes in the convection heat transfer coefficient, chip thermal
capacitance and heat sink surface area for convection heat transfer
affect the chip thermal time constant the most. Sensitivities of the
chip thermal time constant to the rest of the independent variables
follow the same order as in Table 12-1. Changes in independent
variables affecting the conduction heat transfer contribute the least to
chip thermal time constant sensitivities.

119

CHAPTER

13
A CHIP UTILIZING
COOLING OF

A HEAT SINK WITH


RECTANGULAR FINS
H

eat transfer from a surface can be enhanced by using fins. This


chapter combines ideas from Chapters 5 and 12, and analyzes the
cooling of a chip utilizing a heat sink with rectangular fins.
The temperature of a chip can be analyzed using unsteady-state
and one-dimensional heat transfer rate equations in rectangular
coordinates. Energy balance for transient heat transfer during the
cooling of a chip can be written as follows:
Change in internal energy of the chip with respect to time =
Power generated by the chip Energy lost from the chip to its
heat sink and to the environment by conduction and convection
heat transfer

This energy balance can be written as a first order and linear


differential equation:
cpV(dTchip /d) = P (Tchip Tenvironment)/RTotal

(13-1)

121

Everyday Heat Transfer Problems

where
cpV is the chip thermal capacitance in W-s/C
dTchip /d is the time rate of change of chip temperature in C/s
P is the heat generated in the chip in W
Tenvironment is the environment temperature in C
RTotal is the total heat transfer resistance between the chip, its
heat sink with fins and the environment, due to conduction and
convection heat transfer in C/W
Tchip is the temperature of the chip in C
Since the governing equation is linear in temperature, centigrade
dimension is used instead of Kelvin. Total heat transfer resistance
between the chip, its sink and the environment has two parallel paths.
One path is the direct convection heat transfer between the chip and
the environment. The other is the heat loss in series from the chip to
the adhesion layer by conduction, from the adhesion layer to the heat
sink body by conduction, and from the heat sink and its fins to the
environment by convection. The convection heat transfer from the
heat sink and its fins to the environment has two parallel components;
one is the convection heat transfer from the sink base surface to the
environment, and the other is the convection heat transfer from the
fins surfaces to the environment.
The convection heat transfer resistance between the heat sinks
un-finned surfaces and the environment can be written as follows:
Rheat sink convection from un-finned surfaces = 1/(hAheat sink convection un-finned surfaces) (13-2)
The convection heat transfer between heat sink fin surfaces and
the environment can be written by combining Eqs. (5-5) and (5-6):
Qfin total = NhAfin (Tfin base Tenvironment)

(13-3)

The convection heat transfer resistance between the heat sinks fin
surfaces and the environment can be written as follows:
Rheat sink convection from fin surfaces = 1/(NhAfin)

122

(13-4)

Cooling Of A Chip Utilizing A Heat Sink With Rectangular Fins

N is the number of rectangular fins and is the rectangular fin


heat transfer efficiency as defined in Chapter 5. h represents the
convection heat transfer coefficient at the chip surfaces, the heat
sink fin surfaces and the un-finned surfaces. It is assumed that the
convection heat transfer coefficient does not vary with a change in the
number of fins and fin lengths.
The series circuit heat transfer between the chip and the
environment that goes through the adhesive layer and the heat sink as
conduction mechanisms, and from the heat sink to the environment
as parallel convection mechanisms between the un-finned and finned
heat sink surfaces, can be written as follows:
Rchip through heat sink = [Ladhesive /(kadhesiveAadhesive for conduction)]
+ [Lheat sink /(kheat sinkAheat sink for conduction)]
+ [1/(NhAfin + hAheat sink convection un-finned surfaces)] (13-5)
The heat transfer resistance between the surface of the chip and
the environment can be written as:
Rchip to environment = 1/(hAchip for convection)

(13-6)

The total heat transfer resistance between the chip, its heat
sink with fins and the environment can be written by combining
Eqs. (13-5) and (13-6) in a parallel thermal circuit:
RTotal = {1/[(1/Rchip to environment) + (1/Rchip through heat sink)]}

(13-7)

The solution to the governing first order differential Eq. (13-1) can
be written as follows, using the initial condition of Tchip = Tenvironment
and assuming temperature-independent thermophysical properties:
Tchip = Tenvironment + PRT [1 exp(/cpVRTotal)]

(13-8)

where cpVRTotal is the thermal time constant for the chip, which is the
product of its thermal capacitance and its total thermal resistance.

123

Everyday Heat Transfer Problems

For the present sensitivity analysis, the nominal values of the above
independent variables are assumed to be as follows:
cpV = 2.52 W-s/C
(assuming a silicon dioxide chip in the shape of a rectangular box
with 0.03 m 0.03 m 0.002 m dimensions)
h = 50 W/m2-C
Tenvironment = 30C
Tchip = 90C
Achip for convection = 0.00114 m2 which can be detailed as [0.03 0.03
+ 2 (0.03 + 0.03) 0.002]
Ladhesive = 0.002 m
Aadhesive for conduction = Aheat sink for conduction = 0.0009 m2 which can be
detailed as (0.03 0.03)
Lheat sink = 0.005 m
N = 7 (seven rectangular fins which are nominally 0.002 m thick
and 0.03 m in length)
Aheat sink convection un-finned surfaces = 0.00108 m2 which can be detailed as
[with 0.002 m thick seven fins, namely 0.03 (0.03 7 0.002)
+ 2 (0.03 + 0.03) 0.005]
Afin = 0.00186 m2 which can be detailed as [for a 0.002 m thick, 0.03 m
long, and 0.03 m wide fin, namely 2 0.03 (0.03 + 0.5 0.002)]
kadhesive = 100 W/m-C (assuming silver epoxy adhesive)
kheat sink = 300 W/m-C
It is assumed that the heat sink and the fins are machined from the
same block material, and the minimum machinable spacing between
the fins is 0.002 m. For seven fins, the resulting fin thickness is 0.002 m.
The heat transfer efficiency for a rectangular fin can be calculated
from Eq. (5-6) given in Chapter 5, = tanh(mLc)/(mLc) where
m = [2h(w + t)/kwt ]1/2, Lc = L + 0.5t. For cases where the fin width,
w, is much greater than its thickness t, m becomes m = ( 2h /kt )1/2.
The rectangular fin efficiency for this nominal case is 0.95.
Using the above nominal values, and changing the fin length
and the number of fins on the heat sink, the chip heat dissipation
characteristics given in Figure 13-1 are obtained. As fin length
increases, so does chip power dissipation. The chip power dissipation

124

Chip Power Dissipation, W

Cooling Of A Chip Utilizing A Heat Sink With Rectangular Fins

100
80

Lfin=0.01 m
Lfin=0.03 m
Lfin=0.05 m
Lfin=0.07 m
Lfin=0.09 m
Lfin=0.15 m

60
40
20
0
0

2
4
6
Number Of Rectangular Fins

Figure 13-1 Chip power dissipation versus number of fins for different fin

lengths

asymptotes around Lfin = 0.15 meters for an eight-fin heat sink. With
increasing fin length, the efficiency of the fin decreases, even if the
convection heat transfer area increases. Adding the maximum amount
of fins (eight in this case due to machinability constraints), with
lengths of up to 0.15 m per fin, can enhance the chip heat dissipation
by as much as 12-fold, as compared to a non-finned heat sink.
The sensitivity of chip heat dissipation to fin length is given
in Figure 13-2. The sensitivity approaches zero as the fin length
increases. By increasing the fin length, the chip power dissipation
improves less and less, and choosing the right fin length becomes a
cost-benefit issue. For example, using Figure 13-1, if the desired chip
power dissipation is 40 W, choosing a four-fin heat sink design, with a
fin length of 0.05 m, will suffice.
Chip power dissipation versus the convection heat transfer
coefficient is shown in Figure 13-3. Cases with seven fins and with
no fins are compared for a fixed fin length of 0.03 meters. Chip
power dissipation varies linearly with the convection heat transfer
coefficient, since conduction heat transfer resistances through the

125

Everyday Heat Transfer Problems

1400
P/ Lfin, W/m

1200
1000

N=4
N=5
N=6
N=7
N=8

800
600
400
200
0
0.01

0.04

0.07

0.1

0.13

Rectangular Fin Length, m

Chip Power Dissipation, W

Figure 13-2 Chip heat dissipation sensitivity to fin length versus fin length

160
140
120
100

N=7
N=0

80
60
40
20
0
0

50

100

150

200

h, W/m2-C

Figure 13-3 Chip power dissipation versus convection heat transfer coefficient

with no fins and with seven fins for fin length = 0.03 m

126

Thermal Time Constant, s

Cooling Of A Chip Utilizing A Heat Sink With Rectangular Fins

40
30
20
10
0
0

50
100
150
Convection Heat transfer Coefficient, W/m2-C

200

Figure 13-4 Chip thermal time constant versus convection heat transfer

coefficient

adhesive and through the heat sink are negligible for this case
study. The slope of the chip power dissipation versus the convection
heat transfer coefficient line for a seven-fin design, (P/h), is
0.71 m2-C. The slope of the chip power dissipation versus the
convection heat transfer coefficient line for the finless design is
0.13 m2-C. It is apparent that the present nominal heat sink design,
with seven fins, enhances the chip power dissipation by over five-fold.
The thermal time constant for the chip is the product of thermal
capacitance and total thermal resistance, cpVRT, and it is given as a
function of the convection heat transfer coefficient in Figure 13-4.
The thermal time constant for the chip is a strong function of the
convection heat transfer coefficient at low values of forced convection
heat transfer regime; i.e., h < 50 W/m2-C.
In this finned heat sink design, the heat sink base and the fins
are assumed to be the same material. The thermal conductivity of
the heat sink and fin material also starts to affect the chip power
dissipation at lower values of kheat sink , i.e., kheat sink < 100 W/m-C. The
chip power dissipation versus heat sink and fin material thermal
conductivity is given in Figure 13-5 for a seven-fin design with a fin

127

Everyday Heat Transfer Problems

Power Dissipation, W

45
40
35
30
25
20
0

100
200
300
Heat Sink & Fin Thermal Conductivity, W/m-C

400

Figure 13-5 Chip power dissipation versus heat sink and fin thermal

conductivity for Lfin = 0.03 m and N = 7

0.7
P/ ksink, m-C

0.6
0.5
0.4
0.3
0.2
0.1
0
0

100
200
300
Heat Sink & Fin Thermal Conductivity, W/m-C

400

Figure 13-6 Chip power dissipation sensitivity to heat sink and fin thermal

conductivity versus heat sink and fin thermal conductivity for


Lfin = 0.03 m and N = 7

128

Cooling Of A Chip Utilizing A Heat Sink With Rectangular Fins

Table 13-1 Steady-state chip power dissipation capability


change due to a 10% change in variables around
the nominal values for Lfin = 0.03 m and N = 7

Variable
Tchip Tenvironment

Nominal
Value

Steady-State
Chip Power
Dissipation
Capability
Change Due To
A 10% Decrease
In Nominal Value

Steady-State
Chip Power
Dissipation
Capability
Change Due To
A 10% Increase
In Nominal Value

60C

10%

+10%

50 W/m-C

9.41%

+9.29%

kheat sink

300 W/m-C

0.574%

+0.476%

Lheat sink

0.005 m

0.289%

+0.287%

kadhesive

100 W/m-C

0.148%

+0.122%

Ladhesive

0.002 m

+0.134%

0.134%

length of 0.03 meters. The sensitivity of chip power dissipation to


heat sink and fin thermal conductivity is shown in Figure 13-6, again
for a seven-fin design with a fin length of 0.03 meters. The sensitivity
increases fast for low values of heat sink and fin material thermal
conductivity.
A ten-percent difference in independent variables around the
nominal values given above produces the sensitivity results given
in Table 13-1, shown in descending order for the power dissipation
capability of the chip. The chips physical dimensions are assumed
to be constants. This sensitivity analysis is performed for a seven-fin
heat sink design with a fin length of 0.03 meters.
Changes in chip-to-environment temperature potential and the
convection heat transfer coefficient affect the chip power dissipation
capability the most. Changes in conduction heat transfer variables
about the nominal values chosen for this analysis have the least affect
on chip power dissipation capability.

129

CHAPTER

HEAT

14

TRANSFER
ANALYSIS FOR
COOKING IN A POT
F

ood cooking in a pot placed on a gas burner can be a very


complicated heat transfer problem, if multi-dimensional transient
heat transfer for the pot and for the food in the pot are considered.
The heat transfer mechanisms can get challenging, if boiling heat
transfer regime is treated. The best way to approach such a heat
transfer problem is to make simplifying assumptions and create
a simple model, then verify the results of the model by reliable
experiments. In the present heat transfer model, to simulate cooking
in a pot, an unsteady-state and a coupled one-dimensional heat
transfer analysis will be utilized with the following assumptions.
A cylindrical pot gets its cooking energy from a gas burner at a
constant rate. The pot is assumed to be made out of copper and
have a uniform temperature throughout, namely negligible internal
conduction resistance. The pot conduction heat transfer resistance,
Lc/k, is small as compared to the pot surface's convection heat
transfer resistance, 1/h, or Biot Number = hLc/k < 0.1, where Lc is
a characteristic length for the pot, which is the volume of copper
divided by the pot's outer surface area.

131

Everyday Heat Transfer Problems

The pot loses energy to food from its bottom and its side by natural
convection heat transfer. The natural convection heat transfer
mechanism between the pot and the food is analyzed up to a food
temperature of 105C at sea level atmospheric conditions; boiling heat
transfer regime is not considered. The pot also loses energy to the
environment from its bottom and its side by natural convection and
radiation.
The food receives energy from the bottom and the side of the
pot by natural convection. Temperature gradients in the food are
neglected, and the natural convection heat transfer coefficient
between the pot and the food is calculated using an average food
temperature. Conduction and radiation heat transfer mechanisms
between the pot and the food are neglected. Also, the natural
convection and radiation heat transfers from the top of the food to
the environment are neglected. Neglecting these secondary heat
transfer mechanisms introduces an initial error of about 7% in the net
heat transferred to the food, but this error diminishes fast to zero as
cooking time increases.
Food is assumed to have the same thermophysical properties as
water. Temperature-dependent variations for all of the thermophysical
properties are considered. The thermophysical properties are
calculated at a film temperature, which is the average of the surface
and the medium temperatures.
Energy balance for the food can be written as follows:
Change in internal energy of the food with respect to time =
Energy input by convection from the bottom of the pot to the
food + Energy input by convection from the sides of the pot to
the food
Energy balance for the pot can be written as follows:
Change in internal energy of the pot with respect to time =
Energy input from the heater to the bottom of the pot Energy
output by convection from the bottom of the pot to the food
Energy output by convection from the sides of the pot to

132

Heat Transfer Analysis For Cooking In A Pot

the food Energy output by convection from the bottom of


the pot to the environment Energy output by convection from
the sides of the pot to the environment Energy output by
radiation from the bottom and sides of the pot to the environment
The governing coupled and transient first-order differential
equations for the food and the pot are detailed as follows:
(cpV)food(dTfood/d)
= hpot bottom to food natural convection Apot bottom (Tpot Tfood)
+ hpot sides to food natural convection Apot sides (Tpot Tfood)

(14-1)

(cpV)pot(dTpot/d)
= Qin hpot bottom to food natural convection Apot bottom (Tpot Tfood)
hpot sides to food natural convection Apot sides (Tpot Tfood)
hpot bottom to environment natural convection Apot bottom (Tpot Tenvironment)
hpot sides to environment natural convection Apot sides (Tpot Tenvironment)
hpot to environment radiation (Apot bottom + Apot sides) (Tpot Tenvironment) (14-2)
All the natural convection heat transfer coefficients can be
obtained from empirical relationships in literature for the appropriate
geometries; for example, for natural convection heat transfer from
heated vertical plates and horizontal plates, see References [6] and
[10]. The natural convection heat transfer coefficient from the bottom
of the pot to the food is:
(hL/k) = 0.5 RaL0.25

for 104 < RaL < 107

(14-3)

(hL/k) = 0.15 RaL0.33

for 107 < RaL < 1011

(14-4)

The natural convection heat transfer coefficients from the sides


of the pot to the food and the environment are obtained from the
empirical relationship:
(hL/k) = {0.825 + 0.387RaL /[1 + (0.492/Pr)9/16]8/27}2

(14-5)

133

Everyday Heat Transfer Problems

The natural convection heat transfer coefficient from the bottom


of the pot to the environment is obtained from the empirical
relationship:
(hL/k) = 0.27 RaL0.25

for 105 < RaL < 1010

(14-6)

where the Rayleigh number, RaL, is the product of the Grashof


number and the Prandtl number, RaL = g(Tpot Tfood)L3/(). The
Grashof number represents the ratio of buoyancy forces to viscous
forces in a natural convection heat transfer system. The Prandtl
number is the ratio of momentum diffusivity to thermal diffusivity,
namely Pr = /. Tfood in the RaL equation becomes Tenvironment when
the natural convection heat transfer coefficient considered is between
the pot and the environment.
Thermophysical properties, variables and constants in the above
Eqs. (14-1) to (14-6) are defined as:
(cpV) = thermal capacitance, namely product of density, specific
heat at constant pressure and volume, in W-hr/C
T = Temperature in C or in K when radiation calculations are used
= Time in hr
h = heat transfer coefficient for natural convection or radiation heat
transfer mechanisms in W/m2-C
A = Heat transfer area in m2
Qin = Heat input from the gas burner to the pot in W
k = Thermal conductivity in W/m-C
L = Characteristic length for the surface area, where natural
convection heat transfer occurs and is defined as the ratio of heat
transfer surface area to its perimeter in m
g = Gravitational acceleration in m/s2
= Volumetric thermal expansion coefficient in 1/K
= Kinematic viscosity in m2/s
= (k/cp) thermal diffusivity in m2/s
Governing Eqs. (14-1) and (14-2) are solved using the explicit
finite difference method with a ten-second time interval.
Temperature-dependent thermophysical properties of water and

134

Heat Transfer Analysis For Cooking In A Pot

air are obtained from References [6] and [10]. The radiation heat
transfer coefficient is obtained from the following equation:
hradiation = (T4pot T4environment)/(Tpot Tenvironment)

(14-7)

where is the Stefan-Boltzmann constant, 5.67 108 W/m2-K4, and


is the emissivity of the copper pot outer surface. The heat input
from the gas burner to the pot is 750 W. Dimensions of the pot and
the food height are as follows:
Pot inside diameter = 0.127 m
Pot height = 0.076 m
Pot side thickness = 0.0018 m
Pot bottom thickness = 0.0063 m
Food height in the pot = 0.051 m
Food and pot temperatures are given as functions of time in
Figure 14-1. Food temperature is calculated up to 105C, where
the applicability of the natural convection heat transfer coefficient
equations end and the boiling heat transfer regime starts. In addition,
water heating experiments are performed in a copper pot with the
above geometrical dimensions to verify the present model. The
results of these experiments are also given in Figure 14-1. The results
of the present simple heat transfer model match the experimental
results closely. Deviations between the model and the experiments
are seen at initial cooking times; i.e., less than one minute, and as the
water temperature approaches the boiling temperature. The present
heat transfer model can be improved by including the temperature
gradients in the water, by using a more appropriate empirical
relationship for the natural convection heat transfer coefficient as the
water temperature approaches its boiling temperature, and by taking
smaller time increments during calculations.
The natural convection heat transfer coefficients from the
bottom of the pot and the sides of the pot to the food are given in
Figure 14-2. The heat transfer coefficients increase as the pot and
the food temperatures increase, mainly due to the temperature
dependent properties of water. As the food temperature increases,

135

Food and Pot Temperature, C

Everyday Heat Transfer Problems

120
100
80

Tfood, C
Tpot, C
Experiment, C

60
40
20
0

Time, min

Figure 14-1 Food and pot modeling temperatures, and food experiment

Natural Convection Heat


Transfer Coefficient, W/m2-C

temperatures, as functions of time

1600
1200
800

hpot bottom to
food, W/m2-C

400

hpot sides to food,


W/m2-C

0
0

Time, min

Figure 14-2 Pot bottom to food and pot sides to food natural convection

heat transfer coefficients

136

Heat Transfer Analysis For Cooking In A Pot

Table 14-1 Changes in time for the average food temperature


in the pot to reach 105C due to a 10% change in
variables around the nominal values

Nominal
Value

Change in Time
for the Food in
the Pot to Reach
105C Due To
A 10% Decrease
In Nominal Value

Change in Time
for the Food in
the Pot to Reach
105C Due To
A 10% Increase
In Nominal Value

750 W

+11.16%

9.09%

Thermal
Capacity
of Food

Varies
with time,
W-hr/C

8.36%

+8.36%

Apot bottom

0.01267 m2

2.71%

+2.84%

Thermal
Capacity
of Pot

0.1286 W-hr/C

1.65%

+1.65%

0.02027 m2

0.84%

+0.84%

Heat Loss
from Pot to
Environment

Varies
with
time, W

0.30%

+0.30%

hpot bottom to food

Varies with
time, W/m2-C

+0.17%

0.15%

Varies with
time, W/m2-C

+0.13%

0.12%

Variable
Qin

Apot sides

natural convection

hpot sides to food


natural convection

its volumetric thermal expansion coefficient increases and its


kinematic viscosity decreases. The natural convection heat transfer
between the bottom of the pot and the food is more dominant than
the one between the side of the pot and the food, mainly due to the
difference in characteristic lengths that affect the Rayleigh number.
A ten-percent variation around the nominal values of independent
variables given above produces the sensitivity results given in
Table 14-1, for the time that the average food temperature in the pot

137

Everyday Heat Transfer Problems

reaches 105C. The sensitivity results are given in descending order,


and they are applicable only in the region of assigned nominal values
due to their non-linear effects.
The time it takes for the food to reach 105C is most sensitive
to heat input from the gas burner into the pot, and to the thermal
capacity of the food. The sensitivity order continues with the pot
geometry and the pot thermal capacity. The time it takes for the food
to reach 105C is an order of magnitude less sensitive to changes
in energy loss to the environment, and to changes in the natural
convection heat transfer coefficients between the pot surfaces and
the food.
If all of the 750 W heat input from the gas burner goes into the
food, the time for the food to reach 105C is 5 minutes. All the
heat losses from the pot to the environment, as well as the thermal
capacity of the pot, lower the efficiency of the food cooking system.
It takes 6.23 minutes for the food to reach 105C under nominal
conditions.

138

CHAPTER

HEAT

TRANSFER AND
INSULATING
A WATER PIPE

15

rotecting water pipes from freezing requires transient heat


transfer calculations and extensive knowledge of the thermo
physical properties of insulating materials. Under severe environmental
conditions, insulating the water pipe alone might not be sufficient to
prevent it from freezing. Other protections, such as running water
in the pipe or heating tape around the pipe, might be necessary. In
this heat transfer study, a long pipe with an insulation layer wrapped
around it is considered. The interface between the pipes outer
surface and the insulation inner surface is assumed to be in good
contact, with no air gaps or other imperfections to cause any
contact resistance. The pipe is filled with stationary water, and is
only exposed to the environment where there is convection heat
transfer between the outer surface of the insulation and the
environment. The time that it takes the water in the pipe to start
freezing is analyzed, and its sensitivities to governing independent
variables are investigated.
Solutions to the one-dimensional form of the transient heat
conduction equation in cylindrical coordinates, for a long cylinder
and with appropriate boundary conditions, govern this heat transfer

139

Everyday Heat Transfer Problems

problem and are given in the Reference by P. J. Schneider [16]


as follows:
2T/r2 + (1/r)T/r = (1/)T/

(15-1)

where T is the temperature in centigrade along the radial position, r,


of the insulated pipe in meters, is time in seconds, and is thermal
diffusivity of pipe or insulation material, k/cp, in meter squared
per second.
In literature, Eq. (15-1) has closed form solutions as a series of
Bessel functions for a variety of boundary conditions (see
Reference [1]). In the present analysis, Eq. [15-1] is solved using the
explicit finite difference method, (see References [6] and [16]), with
the appropriate boundary conditions. Eq. (15-1) can be written in a
finite difference form, and a radial node temperature at a new time
step is determined by known temperatures at the surrounding nodes
from previous time step calculations, as follows:
Tr+1 =[1 (2+r/r)Fo] Tr + (1+r/r)Fo Tr+1 + Fo Tr1

(15-2)

where Fo is the Fourier number defined as /(r)2, is the time


interval in seconds used in calculations, and r is the radial node
interval in meters used in calculations. In order to achieve stable
step-by-step solutions in time to Eq. (15-2), the coefficient of Tr
has to be positive (see References [6] and [16]). In the present
calculations, appropriate time and spatial intervals are used so that
the Fourier number is always less than 1/(2 + r/r).
The initial condition and the boundary conditions to solve
Eq. (15-2) are as followsinitial condition, = 0, for the finite
difference calculations is:
Tr = Twater

(15-3)

Boundary conditions for the outer pipe-insulation interface and the


pipe inner radial nodes are obtained from applying an energy balance
to a control volume around that particular node. At the inner-most
node of the pipe, it is assumed that the water and half the pipe wall

140

Heat Transfer And Insulating A Water Pipe

thickness have the same temperature. This assumption is good for


small-diameter pipes and for pipe wall materials with high thermal
conductivities. The pipe inner node energy balance per meter of pipe
is given below. The inner node starts at the center of the pipe and
goes out half way in wall thickness of the copper pipe:
Energy lost due to conduction out from the radial element =
Change in stored thermal energy in nodal volume. [This nodal
volume starts at r = 0 and goes to 0.5(r2 r1), which includes the
water in the pipe and half the pipe wall thickness.]
or
kcopper 2 (Tr1 Tr2)/ln(r2/r1)
= {watercp waterr12 + coppercp copper2[r1 + 0.25(r2 r1)] 0.5(r2 r1)}
(15-4)
(Tr1+1 Tr1)/
where
r1 = inner radius of copper pipe in meters
r2 = outer radius of copper pipe in meters
kcopper = thermal conductivity of copper pipe in W/m-C
copper = density of copper pipe in kg/m3
cp copper = specific heat of copper pipe in J/kg-C
water = density of water in kg/m3
cp water = specific heat of water in J/kg-C
It should be noted that thermophysical properties are assumed to
be constant during the present calculations.
The interface node energy balance per meter of pipe for the node
at the interface between pipe outer surface and insulation layer inner
surface is:
Energy gained due to conduction into radial element Energy
lost due to conduction out from the radial element = Change in
stored thermal energy in nodal volume. [This nodal volume starts
at the middle of the pipe wall thickness, 0.5(r2 r1), and goes
out to the middle of the first insulation layer increment r,
(r2 + 0.5r).]

141

Everyday Heat Transfer Problems

or
kcopper 2 (Tr1 Tr2)/ln(r2/r1) kinsulation 2 (Tr2 Tr2+r)/ln(r2 + r/r2)
= {coppercp copper2[r2 0.25(r2 r1)] 0.5(r2 r1)
+ insulationcp insulation2[r2 + 0.25(r2 + r)] 0.5r} (Tr2+1 Tr2)/
(15-5)
where
kinsulation = thermal conductivity of insulation material in W/m-C
insulation = density of insulation material in kg/m3
cp insulation = specific heat of insulation material in J/kg-C
The outer node energy balance per meter of pipethe node at
the interface between the insulation layer outer surface and the
environment is:
Energy gained due to conduction into the radial element
Energy lost to environment due to convection out from the radial
element = Change in stored thermal energy in nodal volume
[This nodal volume starts at 0.5r from the outer radius of the
insulated pipe and goes out to the outer radius, r3.]
or
kinsulation 2 (Tr3-r Tr3)/ln(r3/r3 r) ho 2r3(Tr3 Tenvironment)
(15-6)
= insulationcp insulation2r3 0.5r (Tr3+1 Tr3)/
where
ho = Convection heat transfer coefficient between the outer surface
of insulation and the environment in W/m2-C
r3 = Outer radius of insulated pipe in meters
Nominal values of independent variables used in the present
sensitivity analysis are as follows:
r1 = 0.023 m
r2 = 0.025 m
r3 = 0.035 m

142

Heat Transfer And Insulating A Water Pipe

Twater = 12C
Tenvironment = 20C
water = 1000 kg/m3
cp water = 4200 J/kg-C
kcopper = 50 W/m-C (Note: water pipe is assumed to be copper alloy)
copper = 8800 kg/m3
cp copper = 400 J/kg-C
kinsulation = 0.04 W/m-C
insulation = 26 kg/m3
cp insulation = 835 J/kg-C
ho = 10 W/m2C
rcopper pipe = 0.001 m
rinsulation 0.001 m
= 0.2 s
The dependent variable in the calculations is the time it takes
the water in the pipe to reach 0C. After water reaches 0C at the
inner radius of the copper pipe, it will take extra time for it to freeze
completely. This freezing time can be calculated from the energy
required to compensate for the total latent heat of fusion, Hfg,
required for the water in the pipe:
Hfg water = 334800 W-s/kg
After the temperature of the inner node reaches 0C, the time
required for complete freezing is as follows:
for complete freezing
= waterr12 Hfg /[kcopper2(0 Tr2=time @ r1=0C)/ln(r2/r1)]

(15-7)

From Eq. (15-7), the flow rate required to prevent complete


freezing also can be calculated:
Flow rate to prevent complete freezing in kg/s
= [kcopper2(0 Tr2=T @ r1=0C)/ln(r2/r1)] L /Hfg

(15-8)

143

Time To Start Freezing, hr

Everyday Heat Transfer Problems

10
8
k insulation=0.02
W/m-C
k insulation=0.04
W/m-C
k insulation=0.06
W/m-C

6
4
2
0
0

10
20
30
Insulation Thickness, mm

40

Figure 15-1 Time for water to start freezing versus insulation thickness for

three different insulation thermal conductivities

Time For Water To Start


Freezing, hr

12
10
8
6
4
2
0
50

40

30

20

10

Environment Temperature, C

Figure 15-2 Time for water to start freezing versus environmental

temperature

144

Heat Transfer And Insulating A Water Pipe

Table 15-1 Effects of 10% change in nominal values of


variables to time for water to start freezing

Nominal
Value

Change In Time
For Water To
Start Freezing
For A 10%
Decrease In
Nominal Value

Change In Time
For Water To
Start Freezing
For A 10%
Increase In
Nominal Value

water

1000 kg/m3

8.645%

+8.645%

cp water

4200 J/kg-C

8.645%

+8.645%

kinsulation

0.04 W/m-C

+8.282%

6.774%

Twater

12C

8.115%

+7.816%

Tenvironment

20C

7.363%

+8.669%

Insulation
thickness, r3 r2

0.01 m

5.766%

+5.640%

ho

10 W/m2-C

+2.829%

2.314%

copper

8800 kg/m3

1.315%

+1.315%

cp copper

400 J/kg-C

1.315%

+1.315%

Copper pipe
thickness, r2 r1

0.002 m

+0.243%

0.240%

insulation

26 kg/m3

0.041%

+0.041%

cp insulation

835 J/kg-C

0.041%

+0.041%

kcopper

50 W/m-C

+0.003%

0.003%

Variable

where Tr2 is the temperature at r2 at the time when r1 reaches 0C and


L is the length of the insulated pipe in meters.
Under the above given conditions, water reaches 0C in 3.31 hours
and complete freezing will occur in an additional 14.94 hours.
Time for the water to reach 0C is calculated for different insulation
thicknesses and for three different insulation thermal conductivities.
The results are given in Figure 15-1. At lower insulation thermal
conductivities, time for the water to reach 0C becomes more
sensitive to insulation thickness.

145

Everyday Heat Transfer Problems

Another significant variable affecting the time for water to start


freezing is the environmental temperature. Figure 15-2 shows
the effects of environmental temperature on time for water to
start freezing under the nominal conditions. Time sensitivity to
environmental temperature increases rapidly above 10C.
When the nominal values of the independent variables given above
are varied 10%, the results shown in Table 15-1 are obtained.
Sensitivities to a 10% change in the governing independent variables
of time for water to start freezing are given in descending order of
importance, and these results are applicable only around the nominal
values assumed for this study.
The changes in variables most affecting the time for water to start
freezing are water thermal capacitance, insulation material thermal
conductivity, and water and environmental temperatures. Insulation
thickness, the convection heat transfer coefficient at the outer
surface of insulation layer, and the copper pipe thermal capacitance
properties compose the middle of the pack in order of sensitivity. The
time for water to start freezing is least sensitive to changes in copper
pipe thickness, insulation material thermal capacitance, and copper
pipe thermal conductivity.

146

CHAPTER

QUENCHING

OF STEEL
BALLS IN AIR
FLOW

16

uenching, or rapid and controlled cooling, has been used for


centuries to harden steel and increase the toughness of metal
alloys. Quenching poses a difcult heat transfer problem, as it deals
with high-temperature materials being cooled in a controlled way, in
mediums such as air, water, oil, liquid nitrogen, or some other
special quenching uid. The material properties and the uid
properties change fast, and during the quenching process, heat
transfer between the material body and the medium can go through
several regimes. The medium surrounding the hot body can have
lm boiling, transition from lm boiling to nucleate boiling, nucleate
boiling, and natural convection as quenching time progresses.
In this chapter, the quenching of a small spherical ball made out of
steel is analyzed in a forced convection air ow. The conduction heat
transfer resistance within the ball is assumed to be much less than
the heat transfer resistance between the surface of the ball and the
quenching mediumthe Biot number that is dened in Eq. (16-2) is
less than 0.1. This assumption allows the use of the following energy
balance in the present unsteady-state heat transfer application (see
Reference by J. P. Holman [5]).

147

Everyday Heat Transfer Problems

Rate of decrease of internal energy in spherical steel ball =


Heat lost to surrounding air from the surface of the spherical ball
Since the air properties, convection heat transfer coefcient, and
radiation heat transfer coefcient change by time, the nite difference
form of the energy balance can be written as follows:
(Vcp) (Tsphere (i+1) Tsphere (i))/dt = hi A (Tsphere (i) Tair)

(16-1)

Tsphere (i+1) is temperature of the ball at time (i + 1) in C,


Tsphere (i) is temperature of the ball at time (i) in C,
Tair is temperature of air away from the ball in C,
h is the total heat transfer coefcient in W/m2-C, between the
surface of the steel ball and air, which is the sum of the convection
and radiation heat transfer mechanisms, and varies with time. It is
always calculated at the ith time in order to be able to determine
the temperature at i + 1th time.
A is surface area of the steel ball in m2,
dt is time interval used in calculations in seconds,
is density of the steel ball in kg/m3,
cp is specic heat of the steel ball in J/kg-C, and
V is volume of the steel ball in m3.
cpV/hA is dened as the thermal response time for the quenching
process in seconds.
Tsphere initial is initial temperature of the ball in C when the quenching
process starts.
The Biot number is dened as
Biot number = hmax R/ksteel

(16-2)

where hmax is the maximum heat transfer coefcient, the sum


of convection and radiation heat transfer mechanisms between
the surface of the steel ball and the quenching medium encountered
during the quenching process, R is the radius of the sphere ball
in meters, and ksteel is thermal conductivity of the steel ball in
W/m-C.

148

Quenching of Steel Balls in Air Flow

The convection heat transfer coefcient is calculated from an


empirical relationship, given in literature, for forced convection heat
transfer from the surface of a sphere to air (see Reference [5]).
hforced convection = 0.37 (kair /D) (UmD/air)0.6 for 17 < ReD < 70000 (16-3)
where
kair is the temperature-dependent thermal conductivity of air in
W/m-C,
D is diameter of sphere ball in m,
Um is mean velocity of air away from the ball in m/s,
air is temperature dependent kinematic viscosity of air in
m2/s, and
ReD is the Reynolds number dened as (UmD/air).
The radiation heat transfer coefcient is calculated from
hradiation = (T4sphere T4air)/(Tsphere Tair)

(16-4)

where
is emissivity of steel ball surface and
is Stefan-Boltzmann constant, namely 5.67 108 W/m2-K4.
During the calculations, instantaneous thermophysical properties
of air, kair and air, are calculated at mean temperatures of Tsphere and
Tair, namely at (Tsphere + Tair)/2.
Steel ball properties , cp, , and D are assumed to be constants.
The nominal values of governing independent variables for the
present sensitivity calculations are as follows:
D = 0.01 m,
= 7800 kg/m3
cp = 430 J/kg-C
ksteel = 40 W/m-C
= 0.2
Um = 5 m/s

149

Everyday Heat Transfer Problems

Tair = 27C
Tsphere initial = 800C
dt = 1 s
The present case is analyzed below using different time intervals.
As the time interval is reduced to around one second, the resulting
temperatures-time proles are overlaid, and therefore dt = 1 second is
used for the results presented in the present sensitivity calculations. For
the nominal case, the Biot number is 0.022 for a hmax of 174 W/m2-C. Steel
ball temperature versus time is given in Figure 16-1. For the nominal
case, the steel ball reaches 36.8% of (Tsphere initial Tair), the thermal
response time for the quenching process in 57.0 s, and it reaches 50C
in 153.4 s.
Initially radiation heat transfer is 18% of the total heat transfer.
Radiation heat transfer effect diminishes fast and convection heat
transfer dominates the cooling process, as shown in Figure 16-2.
The effect of mean air velocity is analyzed without violating
the convection heat transfer coefcients range of application in
Eq. (16-3), 17 < ReD < 70000, and the Biot number being less than
0.1. The time for the steel ball to reach 50C and the time for it to
reach 36.8% of (Tsphere initial Tair) are shown as a function of mean

Temperature, C

800
600
400
200
0

50

100
Time, s

Figure 16-1 Steel ball temperature versus time

150

150

200

Quenching of Steel Balls in Air Flow

h, W/m2-C

200
150
h convection
h radiation

100
50
0

50

100
150
Time, s

200

250

Figure 16-2 Convection and radiation heat transfer coefcients versus time

air velocity in Figure 16-3. Changes in mean air velocity below 5 m/s
affect the quenching process time signicantly.
Figure 16-4 shows that for the present analyses, the internal
conduction resistance for the steel ball is much less than the

400
Time To
Reach
50C, s

Time, s

300

Time To
Reach
36.8% of
(Tsphere
initial-Tair), s

200
100
0

5
10
15
Mean Air Velocity, m/s

20

Figure 16-3 Time for the steel ball to reach 50C and to reach 36.8% of

(Tsphere initial Tair) versus mean air velocity

151

Everyday Heat Transfer Problems

Biot Number

0.06

0.04

0.02

10

15

20

Mean Air Velocity, m/s

Figure 16-4 Biot number versus mean air velocity

Table 16-1 Effects of 10% change in nominal values of


independent variables to time for the steel ball
temperature to reach 50C, t50C

Nominal
Value

t50C
Change For
A 10% Decrease
In Nominal
Value

t50C
Change For
A 10% Increase
In Nominal
Value

0.01 m

14.25%

+14.79%

kair

varies with
temperature

+11.07%

9.18%

7800 kg/m3

10.51%

+10.44%

cp

430 J/kg-C

10.51%

+10.44%

Um

5 m/s

+6.53%

5.66%

air

varies with
temperature

6.17%

+5.90%

773C

3.70%

+3.02%

0.2

+0.47%

0.47%

Variable

Tsphere initial Tair

152

Quenching of Steel Balls in Air Flow

convection heat transfer resistance between the surface of the


steel ball and the environment, for the range of mean air velocities
considered in the present calculations, i.e., Biot number < 0.1.
When the nominal values of the independent variables given
above are varied 10%, the results shown in Table 16-1 are obtained.
Time for the steel ball to reach 50C, t50C and sensitivities to a
10% change in the governing independent variables are given in
descending order of importance, and they are applicable only around
the nominal values assumed for this study. Cooling time to 50C
is most sensitive to independent variables such as ball diameter,
density, and specic heat which make up the steel balls thermal
capacitance and the mediums thermal conductivity. Next in the order
of sensitivity is the mean air velocity and the air kinematic viscosity.
Changes in the initial temperature potential, (Tsphere initial Tair), affect
t50C about 3% in this case study. Steel ball surface emissivity is the
least effective of independent variables to t50C.

153

CHAPTER

QUENCHING 17

OF STEEL
BALLS IN OIL

n Chapter 16, the quenching medium analyzed was air. When water,
oil and other similar mediums are used for quenching, heat transfer
gets more complicated because the quenching medium goes into
different heat transfer regimes, such as lm boiling, transition
between lm boiling and nucleate boiling, nucleate boiling, and free
convection, as the surface temperature of the material being
quenched decreases.
Heat transfer in the boiling regime depends on material surface,
quenching medium and material surface combination, and on the
quenching medium's saturated liquid and vapor thermophysical
characteristics. The best way to achieve heat transfer coefcients
for boiling heat transfer in a quenching process is by experimental
measurements (see Reference by Lee, W. J., Kim, Y. and,
E. D. Case [12]).
In order to assume a uniform temperature in the steel ball, namely
Biot number < 0.1, (see Reference by J. P. Holman [5]), oil quenching
instead of water quenching is considered. Oil quenching heat transfer
coefcients are in the order of 1000 W/m2-C

155

Everyday Heat Transfer Problems

If water is used as a quenching medium, the heat transfer


coefcients can easily exceed 10,000 W/m2-C. To satisfy the Biot
number criteria, a steel ball with a diameter of 0.0006 m or less
has to be used. In small surface areas, boiling heat transfer gets
more complicated as the ratio of buoyant and capillary forces play
signicant roles in boiling characteristics. For Biot number > 0.1
and for time varying heat transfer coefcients, nite difference or
nite element methods in multi-dimensional and unsteady-state heat
transfer have to be utilized.
The energy balance for the steel ball in oil quenching medium can
be written as:
Rate of decrease of internal energy in spherical steel ball =
Heat lost to surrounding oil from the surface of the spherical ball
Since the heat transfer coefcient will change by time, the nite
difference form of the energy equation is the same as Eq. (16-1):
(Vcp) (Tsphere (i+1) Tsphere (i) )/dt = hi A (Tsphere (i) Toil) (17-1)
Tsphere (i+1) is temperature of the ball at time (i+1) in C,
Tsphere (i) is temperature of the ball at time (i) in C,
Toil is temperature of oil in C,
hi is the experimentally obtained total heat transfer coefcient in
W/m2-C, between the surface of the steel ball and oil, which is the
sum of the convection and radiation heat transfer mechanisms, and
varies with time.
A is surface area of the steel ball in m2,
dt is time interval used in calculations in seconds,
is density of the steel ball in kg/m3,
cp is specic heat of the steel ball in J/kg-C, and
V is volume of the steel ball in m3.
In the present calculations, experimentally obtained heat transfer
coefcients for oil quenching are used (see Reference [12]). These are
assumed to be total heat transfer coefcients; namely, they account

156

Quenching of Steel Balls in Oil

htotal, W/m2-C

3000
nucleate
boiling

2000

free
convection

1000

C
A

transition

200

400
600
Steel Ball Temperature, C

film boiling
(vapor
blanket)
800

Figure 17-1 Total heat transfer coefcient versus steel ball temperature for

oil quenching

for both boiling and radiation heat transfer mechanisms. For oil
quenching, the total heat transfer coefcient as a function of steel ball
surface temperature is shown in Figure 17-1.
In temperatures above 750C, point C in Figure 17-1, the steel
ball encounters lm boiling heat transfer and is covered with a vapor
blanket. The temperatures between points B and C, 600C to 750C,
constitute the transition region from boiling heat transfer to lm
boiling heat transfer. The temperatures between points A and B, 200C
to 600C, are in the boiling heat transfer regime. Temperatures below
200C are considered to be in the free convection heat transfer regime.
Steel ball properties , cp, ksteel, and D are assumed to be constants
during the oil quenching process. The nominal values of governing
independent variables for the present sensitivity calculations are as
follows:
D = 0.012 m,
= 7800 kg/m3
cp = 430 J/kg-C

157

Everyday Heat Transfer Problems

ksteel = 60 W/m-C
Toil = 70C
Tsphere initial = 900C
dt = 0.1 s
For the present calculations, a time interval of 0.2 seconds is used
because the thermal time constant for quenching in oil has dropped
an order of magnitude as compared to quenching in air (see Chapter
16). Calculations are also repeated with a 0.1 second time interval,
which improves the results by only one percent. The Biot number
for this case, hR/ksteel, is 0.1 for an average heat transfer coefcient
of 1000 W/m2-C, where R = D/2. Steel ball temperatures versus time
are shown in Figure 17-2 for nominal heat transfer coefcients, and
for a 10% variation about the nominal heat transfer coefcient
values given in Figure 17-1. Figure 17.2 shows the initial seconds
of the quenching process to emphasize the effects of heat transfer
coefcient variations. Boiling heat transfer phases occur during the
initial seconds of the oil quenching process. For the nominal case,
after 8.6 seconds, the natural convection heat transfer regime starts.

Temperature, C

800
Nominal h
(Figure 17-1)

600

10% higher h
than nominal

400
200

10% lower h
than nominal

10

Time, s

Figure 17-2 Steel ball temperature versus time for oil quenching for three

different heat transfer coefcient distributions

158

Quenching of Steel Balls in Oil

Table 17-1 Effects of a 10% change in nominal values, of


heat transfer coefcients given in gure 17-1, to
time for the steel ball to reach 100C, and to time
for it to reach 36.8% of (Tsphere initial Toil)

Variable
h

Variable
h

Nominal
Value

Change in
Time to Reach
100C For A
10% Decrease
In h Nominal
Value

Change in
Time to Reach
100C For A
10% Increase
In h Nominal
Value

Figure 17-1

+11.3%

9.6%

Nominal
Value

Change in
Time to Reach
36.8% of
(Tsphere initial Toil)
For A 10%
Decrease In
h Nominal
Value

Change in
Time to Reach
36.8% of
(Tsphere initial Toil)
For A 10%
Increase In
h Nominal
Value

Figure 17-1

+12.3%

9.6%

Sensitivities of time for the steel ball to reach a temperature of


100C and time for the steel ball to reach 36.8% of (Tsphere initial Toil),
thermal response time for the quenching process, are analyzed as
the dependent variables to the changing heat transfer coefcients.
The results are provided in Table 17-1. Oil quenching time
sensitivities to changes in heat transfer coefcients are signicant,
and they are one-to-one. These results emphasize the importance of
process controls during quenching.

159

CHAPTER

COOKING

TIME FOR
TURKEY IN
AN OVEN

118
8

very year when Thanksgiving comes around, the question of a


turkeys cooking time in an oven looms. This is an unsteady-state
conduction heat transfer problem in spherical coordinates, whose
solution can be found in References [1], [6], [11], and [16]. The turkey
is assumed to be stuffed and in a spherical shape in which the spatial
variations of temperature are only in the radial direction. Turkey also
has constant thermophysical and physical properties in space and in
time. Energy balance in the radial direction, r, for a small increment
of time, t, gives the following partial differential equation for the
temperature-time distribution for a turkey:
2Tturkey /r2 + (2/r)Tturkey /r = (1/)Tturkey /t

(18-1)

A closed form solution to Eq. (18-1) can be obtained by applying


the following initial condition:
Tturkey = Tturkey initial at time t = 0

(18-2)

and the following two boundary conditions, one at the center

161

Everyday Heat Transfer Problems

at radius r = 0, r(Tturkey /r) = 0

(18-3)

the other at the outer surface


at radius r = R, k(Tturkey /r) = h(Tturkey Toven)

(18-4)

where
= (k/cp) is the thermal diffusivity of turkey in m2/s,
k is the thermal conductivity of turkey in W/m-C,
is the density of turkey in kg/m3,
cp is the specific heat of turkey at constant pressure in
J/kg-C, and
h is the convection heat transfer coefficient between the surface
of the turkey and the oven environment in W/m2-C.
The solution to Eqs. (18-1) through (18-4) can be found in
References [1], [6], [11], and [16], and it is as follows:
(Toven Tturkey @ r)/(Toven Tturkey initial) =
(18-5)

where n is the nth positive root of the following transcendental


equation:
1 n cot(n) = Biot number = hR/k

(18-6)

The center temperature of the turkey is of particular interest for


cooking. The USDA recommends that for a cooked turkey, the center
of the stuffing should reach a temperature of 74C. In the present
sensitivity analysis, the temperature time history at the center of the
bird is analyzed. Eq. (18-5) reduces to the following relationship at
the center, r = 0, of the turkey:

162

Cooking Time For Turkey In An Oven

(Toven Tturkey @ r=0)/(Toven Tturkey initial) =


(18-7)

There are eight independent variables that govern the present


sensitivity analysis. They are:
W = Weight of turkey in kg
= Density of turkey in kg/m3
k = Thermal conductivity of turkey in W/m-C
cp = Specific heat of turkey at constant pressure in kJ/kg-C
Tturkey initial = Initial temperature of turkey in C
Tturkey final = Final temperature desired at the center, r = 0, of
turkey in C
Toven = Oven temperature in C
h = Convection heat transfer coefficient in W/m2-C
The radius, R, for the turkey is obtained from its weight and its
spherical assumption, using R = (3W/4)1/3. There are three
non-dimensional variables that can capture all the temperature time
distributions resulting from Eq. (18-7):
The dependant variable is a non-dimensional temperature, , namely
= (Toven Tturkey @ r=0)/(Toven Tturkey initial).

(18-8)

The independent variable is the dimensionless time, namely the


Fourier number,
Fo = t/R2.

(18-9)

The dimensionless parameter is the Biot number,


Bi = hR/k,

(18-10)

which is the ratio of internal to external thermal resistance for


the turkey. Most solutions to unsteady-state heat transfer problems

163

Everyday Heat Transfer Problems

shown in literature, i.e., References [10] and [16], are shown in


graphical form by using the non-dimensional variables given in
Eqs. (18-8), (18-9) and (18-10).
To find these solutions, five terms of Eq. (18-5) are utilized:
= C1 exp(12Fo) + C2 exp(22Fo) + C3 exp(32Fo)
(18-11)
+ C4 exp(42Fo) + C5 exp(52Fo) +
The five-term expansion solution in Eq (18-11) gives very accurate
results. The fifth term only contributes at very small Fo numbers,
i.e., contribution to for Fo = 0.05 is less than 105. First the 1, 2,
3, 4, and 5 have to be determined from the positive roots of the
transcendental Eq. (18-6) for different Biot numbers.
The ovens that will be considered in the present sensitivity
analysis will have a convection heat transfer coefficient range of 1 to
60 W/m2-K. The turkeys will have a weight range of 2 to 20 kg. The
turkeys thermophysical properties are assumed to be = 1000 kg/m3,
cp = 3.33 kJ/kg-K, and k = 0.5 W/m-C. With these assumptions, the
diameter of turkey ranges from 16 cm to 34 cm. The Biot number
range will be 0.13 to 16.
The temperature time history of interest for cooking the turkey
falls in a time range between one and ten hours. For this cooking time
range, the Fourier number range is 0.1 to 0.2.
The temperature time history for the turkey can be presented by
using the three non-dimensional variables in Eqs. (18-8), (18-9) and
(18-10), as shown in Figure 18-1.
A typical turkey is cooked at an oven temperature of 190C, with
an initial turkey temperature of 5C and a final desired temperature at
the center of the cooked turkey to be 80C. These temperature values
provide a region of interest for the dimensionless temperature to be
at 0.43. The present sensitivity analysis is for a 7.3 kg stuffed turkey
with a 24 cm diameter.
There are two variables that characterize the oven that is being
used, oven temperature and the convection heat transfer coefficient
on the surface of the turkey inside the oven. For a fixed oven
temperature, Toven = 190.5C , the cooking times can be obtained

164

Dimensionless Temperature
At Center Of Turkey

Cooking Time For Turkey In An Oven

1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0.1

Bi=0.13
Bi=0.5
Bi=1
Bi=2
Bi=3
Bi=5
Bi=7
Bi=10
0.2
0.3
0.4
Fourier Number, Fo

0.5

Bi=16

Figure 18-1 Dimensionless temperature at the center of a stuffed turkey

versus Fourier number for different Biot numbers in the region


of interest for cooking a turkey

from Figure 18-1 for different convection heat transfer coefficients


or Biot numbers. The cooking times versus the convection heat
transfer coefficient for three different size turkeys are given in Figure
18-2. The results in Figure 18-2 are for a final desired temperature of
73.9C at the center of the turkey. The USDA recommended cooking
times for different size turkeys are also shown in Figure 18-2. USDA
recommended cooking times require a good convection oven with a
convection heat transfer coefficient ranging from 15 to 20 W/m2-C.
As the convection heat transfer coefficient increases cooking
time decreases, and this decrease in cooking time is very sensitive
to changes in the convection heat transfer coefficients, especially in
low values. The behaviors of these curves for different stuffed turkey
weights are similar. The slopes of the curves in Figure 18-2 are given
in Figure 18-3.
Figure 18-3 shows that (Cooking Time)/h decreases sharply
as the forced convection heat transfer coefficient values approach
natural convection heat transfer coefficient values. The other variable
that is governed by the oven is oven temperature. The sensitivities

165

Cooking Time, hr

Everyday Heat Transfer Problems

4.5 kg (10 lb)


stuffed turkey
7.3 kg (16 lb)
stuffed turkey
10 kg (22 lb)
stuffed turkey

15
14
13
12
11
10
9
8
7
6
5
4
3
2

USDA recommended
cooking times:
for a 4.5 kg (10 lb)
stuffed turkey 3-3.5 hrs,
for a 7.3 kg (16 lb)
0
5
10
15
20 stuffed turkey 4-4.5 hrs
for a 10 kg (22 lb)
Convection Heat Transfer Coefficient, W/m2-K
stuffed turkey 4.75-5.25
hrs

Figure 18-2 Stuffed turkey cooking time as a function of the convection

(Cooking Time) / h, hr-m2-C/W

heat transfer coefficient for stuffed turkeys of different weights


for Toven = 190.5C and for Tturkey final = 73.9C

0
0.1

W=4.5 kg
(10 lb)
W=7.3 kg
(16 lb)
W=10 kg
(22 lb)

0.2
0.3
0.4
0.5
0

10

Convection Heat Transfer

20

30

Coefficient,W/m2-C

Figure 18-3 Stuffed turkey cooking time to convection heat transfer

coefficient sensitivity as a function of the convection heat


transfer coefficient for stuffed turkeys of different weights, for
Toven = 190.5C, and for Tturkey final = 73.9C

166

(Cooking Time) / (Toven), min/C

Cooking Time For Turkey In An Oven

0.8
1

Turkey Center Final


Temperature=73.9C

1.2

Turkey Center Final


Temperature=79.4C

1.4

Turkey Center Final


Temperature=85.0C

1.6
1.8
180

190
200
210
Oven Temperature, C

220

Figure 18-4 A 7.3 kg stuffed turkey cooking time to oven temperature

sensitivity versus oven temperature for different turkey center


final temperatures

of the cooking time to oven temperature for different stuffed turkey


centerline temperatures are given in Figure 18-4.
The absolute value of the cooking time sensitivity to oven
temperature decreases as the oven temperature increases. Sensitivity
curves are similar in this application region for different final turkey
center temperatures.
The above sensitivity graphs show that the cooking time sensitivity
to governing independent variables behave non-linearly. Therefore,
it is more appropriate to analyze these sensitivities in the region
of interest and rank them according to their effects on the cooking
time. The results in Table 18-1 are obtained from Figure 18-1 by
interpolating between appropriate non-dimensional temperatures,
Biot numbers and Fourier numbers.
The effects shown in Table 18-1 to cooking time are given in
descending order. The most effective variable is the specific heat of
the stuffed turkey at constant pressure. The cooking time sensitivities
to all the independent variables are at the same order of magnitude.

167

Everyday Heat Transfer Problems

Table 18-1 Effects of a 10% change in nominal values of


variables to cooking time

Nominal
Value

Cooking
Time Change
Due To A 10%
Decrease In
Nominal
Value

Cooking
Time Change
Due To A 10%
Decrease In
Nominal
Value

3.33 kJ/kg-C

10%

+10%

Tturkey final @ r=0

73.9C

8.12%

+8.12%

Toven Tturkey initial

186.1C

+8.10%

6.41%

Thermal
Conductivity, k

0.5 W/m-C

+7.10%

6.14%

7.3 kg

6.02%

+6.02%

Convection
Heat Transfer
Coefficient, h

10 W/m2-C

+5.13%

4.42%

Density,

1000 kg/m3

4.32%

+4.63%

Variable
Specific Heat @
Constant
Pressure, cp

Weight, W

Some variables, such as the turkeys specific heat at constant


pressure, final temperature desired at the center of turkey, and
weight of the turkey behave linearly in the region of interest, and give
the same magnitude cooking time change percentages on both sides
of their nominal values. It is important to remember that the ranking
order shown in Table 18-1 is good only in the region of nominal
values for the present calculations, due to non-linear behavior of the
sensitivities.

168

CHAPTER

HEAT

GENERATED IN
PIPE FLOWS DUE
TO FRICTION

19

eat generated in pipes or in orices due to uid friction in


high-viscosity uids can be substantial. In the present analysis,
the heat generated in steady-state and fully developed pipe ows is
investigated for uids of different viscosities. The pressure drop in a
pipe due to uid friction, P, is generally dened as the product of the
uid friction factor, f, non-dimensional length of the pipe, L/D, and
kinetic energy of the uid, Vm2/2, owing in it.
P = f (L/D) (Vm2/2)

(19-1)

Eq. (19-1) has been determined by dimensional analysis (see


Reference [15]). The friction factor is a function of Reynolds
number, ReD = VmD/, and pipe inner surface roughness, e/D. The
friction factors, f, for fully developed pipe ows have been obtained
experimentally for different Reynolds numbers and pipe inner surface
roughness conditions, and they are given in References [6] and [10] as
graphs, called Moody Diagrams, or as empirical equations.

169

Everyday Heat Transfer Problems

The pressure loss in the pipe is related to the heat generated in the
pipe by the rst law of thermodynamics:
Q = (VmA) P

(19-2)

where
P is the pressure drop in pipe in N/m2,
L is the length of pipe in m,
D is the internal diameter of pipe in m,
is the density of uid in kg/m3,
Vm is the average uid velocity in pipe in m/s,
Q is heat generated due to friction in W, and
A is cross sectional area of pipe in m2.
The rst example analyzes a high-viscosity engine oil, and ow
rates are in the laminar ow region; i.e., ReD < 2000. The friction
factor for laminar ow is given as follows, (see Reference [10]), and it
is independent of pipe surface roughness:
f = 64/ReD

(19-3)

Then the heat generated due to friction, after combining Eqs. (19-1),
(19-2) and (19-3), becomes:
Q = (128/) L (VmA)2/D4

(19-4)

Nominal values of the independent variables for this example


are assumed to be as follows; the density of oil cancels out of the
governing Eq. (19-4).
L = 400 m,
D = 0.1 m,
VmA = 1000 liters/min (0.0167 m3/s), and
= 0.486 N-s/m2 @ 27C.
Temperature effects on the kinematic viscosity of engine oil, /,
from 5C to 55C, are given in Figure 19-1. The kinematic viscosity of

170

Kinematic Viscosity, m2/s

Heat Generated In Pipe Flows Due To Friction

0.003

0.002

0.001

0.000

15

25
35
45
Mean Oil Temperature, C

55

Figure 19-1 Kinematic viscosity of oil as a function of mean oil

temperature
oil decreases as its temperature increases. Similar behavior is seen in
Figure 19-2 for the heat generated due to friction.
The mean temperature increase in oil behaves the same way as shown
in Figure 19-3. Tmean is calculated from Q/(VmAcp), assuming that all

Heat Generated, kW

150

100

50

15

25

35

45

55

Mean Oil Temperature, C

Figure 19-2 Heat generated by friction for oil owing in a pipe

171

Temperature Increase, C

Everyday Heat Transfer Problems

5
4
3
2
1
0

15

25
35
Mean Oil Temperature, C

45

55

Figure 19-3 Temperature increase in oil due to friction

the heat generated due to uid friction goes to increase the temperature
of oil. cp is the specic heat of oil at constant pressure, in J/kg-C.
Sensitivities of heat, generated by oil owing in a pipe in the
laminar ow region, to the governing independent variables can be
obtained by differentiating Eq. (19-4):
Q/L = (128/) (VmA)2/D4

(19-5)

Q/ = (128/) L (VmA)2/D4

(19-6)

Q/D = (512/) L (VmA)2/D5

(19-7)

Q/(VmA) = (256/) L (VmA)/D4

(19-8)

The sensitivity of heat generated to pipe internal diameter, Eq.


(19-7), is shown in Figure 19-4. The sensitivity is prominent at
low pipe internal diameters, and approaches zero fast as the pipe
diameter increases for this case. The sensitivity of heat generated
to pipe ow rate, Eq. (19-8), is given in Figure 19.5. The sensitivity,

172

Heat Generated In Pipe Flows Due To Friction

Q/ D, W/m

0.E+ 00

1.E+ 07

2.E+ 07

3.E+ 07
0.05

0.07

0.09

0.11

0.13

0.15

Pipe Inside Diameter, m

Figure 19-4 Sensitivity of heat generated due to friction to pipe internal

diameter for fully developed laminar ows

Q/(VmA) increases linearly with the increasing ow rate. These


sensitivities are calculated at a mean oil temperature of 27C.
When the nominal values of the independent variables given above
are varied 10%, the results in Table 19-1 are obtained. These heat
generation sensitivities are given in descending order of importance,
and they are applicable only around the nominal values assumed for this
case. Heat generated due to friction is most sensitive to pipe internal
diameter and to ow rate. Heat generated due to friction has a one-to-one
sensitivity relation to the length of the pipe and to uid viscosity.
Similar sensitivity calculations are done for water, which has a
three order of magnitude lower viscosity than the engine oil analyzed
previously. Nominal values of the independent variables for this
example are assumed to be as follows:
L = 400 m,
D = 0.1 m,
VmA = 1000 liters/min (0.0167 m3/s), and
= 0.855 103 N-s/m2 @ 27C.
= 997 kg/m3 @ 27C

173

Everyday Heat Transfer Problems

Q/ (VmA), W-s/m3

1.5E+07

1.0E+07

5.0E+06

0.0E+00
0.01

0.03

0.05
Flow Rate, m3/s

0.07

0.09

Figure 19-5 Sensitivity of heat generated due to friction to pipe ow rate for

fully developed laminar ows

In this case friction factors are lower, ReD is higher in the


order of 105, and this case falls into a steady-state and fully
developed turbulent ow region in a pipe. For the friction
coefcient in smooth pipes in the turbulent ow region, the

Table 19-1 Effects of 10% change in nominal values of


variables to heat generated due to friction for
steady-state and fully developed laminar ow of
engine oil in a pipe

Variable
D
VmA
L
@ 27C

174

Nominal
Value

Heat Generated
Due To Friction
Change For A 10%
Decrease In
Nominal Value

Heat Generated
Due To Friction
Change For A 10%
Increase In
Nominal Value

0.1 m

+52.42%

31.70%

1000 L/min

19%

+21%

400 m

10%

+10%

10%

+10%

0.486 N-s/m

Heat Generated In Pipe Flows Due To Friction

following empirical equation is used (see Reference [6]):


f = (0.790 ln(ReD) 1.64)2 for 3000 < ReD < 5 106

(19-9)

Kinematic Viscosity, m2/s

The heat generated due to friction for fully developed turbulent


ow in a smooth pipe can be obtained by combining Eqs. (19-1),
(19-2) and (19-9). Heat generated by water ow in a pipe is
analyzed between the mean water temperatures of 5C and 55C.
The kinematic viscosity of water is three orders of magnitude less
than the engine oil kinematic viscosity, as shown in Figure 19-6.
Heat generated due to friction, and therefore the mean temperature
increase for water, is two orders of magnitude less than the previous
oil ow case, as shown in Figures 19-7 and 19-8.
When the nominal values of the variables given above are varied
10%, the results in Table 19-2 are obtained. These heat generation
sensitivities are given in descending order of importance, and they are
applicable only around the nominal values assumed for this case. Heat
generated due to friction is most sensitive to pipe internal diameter
and then to the ow rate, as in the previous case. A change in the

1.50E 06
1.30E 06
1.10E 06
9.00E 07
7.00E 07
5.00E 07

15

25

35

45

55

Mean Water Temperature, C

Figure 19-6 Kinematic viscosity of water as a function of mean water

temperature

175

Everyday Heat Transfer Problems

Heat Generated, kW

2.6

2.4

2.2

15

25

35

45

55

Mean Water Temperature, C

Figure 19-7 Heat generated by friction for fully developed and turbulent

water ow in a pipe

Temperature Increase, C

viscosity of water is the least effective independent variable to heat


generated due to friction.
Changes in pipe inner surface roughness also effect the heat
generated due to friction in a pipe. The effects of pipe inner surface

0.038
0.036
0.034
0.032
0.030
0.028

15

25

35

45

55

Mean Water Temperature, C

Figure 19-8 Temperature increase in mean water temperature due to friction

176

Heat Generated In Pipe Flows Due To Friction

Table 19-2 Effects of 10% change in nominal values of


variables to heat generated due to friction for a
steady-state and fully developed turbulent ow of
water in a smooth pipe
Heat Generated
Heat Generated
Due To Friction
Due To Friction
Change For A 10% Change For A 10%
Decrease In
Increase In
Nominal Value
Nominal Value

Nominal
Value

Variable
D

0.1 m

+65.95%

36.75%

1000 L/min

25.60%

+30.67%

400 m

10%

+10%

997 kg/m

8.15%

+7.99%

2.02%

+1.86%

VmA

@ 27C 0.855 10 N-s/m

% Increase in Heat Generated

-3

160
120
80
40
0

200
400
600
800
Pipe Inner Surface Roughness, m

1000

Figure 19-9 Percent increase in heat generated due to friction as a

function of pipe inner surface roughness for water owing in a


pipe, ReD = 2.47 10+5

177

Everyday Heat Transfer Problems

roughness are analyzed for ReD = 2.47 10+5 using the Moody
Diagram (see Reference [10]). The results are given in Figure 19-9
for steady-state, fully developed, and turbulent water owing in a
pipe. The effect of pipe inner surface roughness to heat generated
due to friction increases linearly above e = 100 m. A pipe with an
inner surface roughness of 100 m generates 38% more heat than
a smooth one.

178

CHAPTER

20
ACTIVE SOLAR
SIZING AN

COLLECTOR
FOR A POOL
S

olar collectors used to heat water have been improving steadily in


solar energy conversion efciency as the demand to go to greener
energy sources increases. To be able to design a pool solar collector
system, a good knowledge of year-around environmental conditions
has to be gathered. Average daily incident solar radiation, average
daily daytime and nighttime environment temperatures, and average
daily daytime and nighttime relative humidity must be known.
Sizing an active (directly circulating the pool water) solar
collector to heat a pool's water requires knowledge of the average
solar insolation at the location of the collector, and the heat transfer
efciency of the solar collector.
A pool's physical and thermophysical properties must be known
to deal with radiation heat transfer, convection heat transfer, and
evaporative cooling from the water's surface during the day and night.
Conduction heat transfer from the pool's structure to the earth also
has to be treated. A plastic pool surface cover will be considered in
order to minimize heat losses to the environment when the pool is
not in use. Therefore, a plastic cover's physical and thermophysical
properties must also be known.

179

Everyday Heat Transfer Problems

The present heat transfer model for the pool solar collector
system neglects temperature gradients in pool water. The model
is for unsteady-state heat transfer in one-dimensional rectangular
coordinates. There are 25 independent variables that govern the
dependent variable, size of the solar collector, and are considered
for sensitivity analysis.
The present heat transfer model has the following assumptions:
The pool is covered at night for 14 hours and open for swimming
during the day for ten hours. Energy lost from the water in the solar
collector at night and energy lost in piping between the pool and the
solar collector are considered to be lumped into the solar collector
efciency. Average heat transfer properties are used during both
day and night hours. When a pool cover is used at night, there is no
evaporation from the pool's surface.
The following energy balances can be constructed for the water in
the pool for the day-time and night-time hours.
Energy balance during the day for the pool:
Qin solar collector + Qin convection from environment during day Qout evaporation during day
+ Qin net radiation on pool surface Qout conduction to earth
(20-1)
= mw cpw (dTpool/d)
Energy balance during the night for the pool with a cover:
Qout convection to environment at night Qout conduction to earth
Qout radiation from pool cover surface = mw cpw (dTpool/d)

(20-2)

Energy balance during the night for the pool without a cover:
Qout convection to environment at night Qout conduction to earth
Qout radiation from pool surface Qout evaporation at night
= mw cpw (dTpool/d)

(20-3)

where mw is weight of water in the pool in kg, cpw is specic heat


of water at constant pressure in W-hr/kg-C, Tpool is average pool
temperature in C, and is time in hours.

180

Sizing An Active Solar Collector For A Pool

Each term in energy balance Eqs. (20-1), (20-2) and (20-3)


is dened as follows. Net solar energy the water receives while
circulating in the solar collector is:
Qin solar collector = q Ac

(20-4)

where is the solar collector heat transfer efciency, or the


percentage of solar insolation that can be converted to heating the
pool water by the collector and its connecting pipes, which can vary
between 0.4 and 0.7.
q is the average solar insolation on the collector in W/m2. Ac is the
area of the solar collector in m2.
Convection heat transferred from the environment to the water
surface during the day is:
Qin convection from environment during day
= hday Apool water surface (Tenvironment during day Tpool)

(20-5)

where hday is the average convection heat transfer coefcient between


pool water surface and environment during the day, Apool water surface
is pool water surface area (pool length times pool width in m2) and
Tenvironment during day is average environmental temperature during ten
hours of day in C.
Heat lost from the water due to evaporation to the environment
during the day is:
Qout evaporation during day = M Hfg

(20-6)

where M is evaporation rate in kg/hr-m2 and Hfg is latent heat of


evaporation for water in air in W-hr/kg.
The mass transfer rate equation can be dened similar to the
convection heat transfer rate equation. By applying the perfect gas
law, the mass transfer rate equation can be written as follows:
M = (hdiffusion day/R Tpool )(pw pa)

(20-7)

181

Everyday Heat Transfer Problems

By assuming thermal diffusivity of air, in m2/hr, being equal to the


diffusion coefcient of water in air, D in m2/hr, and by assuming the
water in air concentration gradient at the water-air interface to
be the same as the temperature gradientNusselt number (hday L/ka) =
Sherwood number (hdiffusion day L/D), the mass transfer coefcient can
be obtained from the heat transfer coefcient (see Reference [2]):
hdiffusion day = hday/a cpa

(20-8)

hdiffusion day is the mass transfer coefcient between pool water surface
and air in m/hr, R is the universal gas constant which is equal to
0.08205 m3-atm/kmol-K, pw is the saturation pressure of water vapor on
pool water surface at pool temperature, pa is partial pressure of water
vapor in air at average environmental temperature (calculated from
pa = psaturation, where is relative humidity in the air and psaturation is
the saturation pressure of water vapor in air at average environmental
temperature), L is a characteristic length of pool surface, namely pool
length, ka is thermal conductivity of air in W/m-K, a is density of air in
kg/m3, and cpa is specic heat of air at constant pressure in W-hr/kg-C.
The net solar radiation that is absorbed by water is:
Qin net radiation on pool surface = w G Apool water surface

(20-9)

where w is average solar radiation absorptivity at water surface


and G is average solar insolation on pool water surface in W/m2, that
includes both direct and diffused solar energy.
Conduction heat transfer from the walls of the pool to earth is:
Qout conduction to earth
= (kpool wall Apool-earth surfaces/tpool wall)(Tpool Tearth)

(20-10)

where kpool wall is thermal conductivity of pool walls in W/m-K,


Apool-earth surfaces are surface areas between the pool and earth in m2,
dened as [(2 pool length + 2 pool width) pool water height +
pool length pool width], tpool wall is average pool wall thickness in m,
and Tearth is average earth temperature.

182

Sizing An Active Solar Collector For A Pool

Heat lost during the night from the pool with a cover to the
environment can be written as a series thermal circuit of convection
and conduction heat transfer:
Qout convection to environment at night
= Apool water surface (Tpool Tenvironment during night)/[(1/hnight )
+ (tcover /kcover)] (20-11)
where Tenvironment during night is average environmental temperature at
night in C, hnight is the average convection heat transfer coefcient at
night integrated over the length of the pool, given by the following
empirical relationship (see Reference [6]), in W/m2-K:
(hnight L/ka) = (0.037 ReL0.8 871) Pr0.333
for 0.6 < Pr < 60, 5 105 < ReL < 1 108, and Recritical = 5 105 (20-12)
where ReL = VL/ is Reynolds number, with V being average air
speed in m/hr over the length of the pool L in m and being
kinematic viscosity of air in m2/hr at lm temperature, the average
of water surface and environmental temperature. Pr = /t is the
Prandtl number, where t is thermal diffusivity of air in m2/hr
at lm temperature, the average of surface and environmental
temperature. Recritical is the transitional Reynolds number from
laminar to turbulent ow on a at plate. tcover is thickness of pool
cover in meters and kcover is thermal conductivity of pool cover at
pool temperature.
Heat radiated from the pool cover surface to the night sky is:
Qout radiation from pool cover surface = pool cover surface T4pool cover surface

(20-13)

where pool cover surface is emissivity of pool cover surface and the is
Stefan-Boltzmann constant, namely 5.67 108 W/m2-K4.
Energy balance during the night for the pool without
a cover has convection and radiation heat transfers dened
as follows:

183

Everyday Heat Transfer Problems

Qout convection to environment at night = hnight Apool water surface


(Tenvironment during night Tpool) (20-14)
Qout radiation from pool surface = pool water surface T4pool

(20-15)

where pool water surface is emissivity of pool water surface.


Qout evaporation at night is calculated the same way as Qout evaporation during day
except that hdiffusion night is calculated from the heat transfer coefcient
integrated over the length of the pool; Eq. (20-12), using an analogy
between heat and mass transfer.
(hdiffusion night L/D) = (0.037 ReL0.8 871) Sc0.333
for 0.6 < Sc < 3000, 5 105 < ReL < 1 108, and Recritical = 5 105 (20-16)
where D is the diffusion coefcient for water in air in m2/hr and Sc is
the Schmidt number dened as /D.
Governing Eqs. (20-1) and (20-2) are solved by explicit nite
difference method and by iteration, to determine the required solar
collector area that will heat the pool water to a minimum of 25C
in two days, specically one ten-hour day followed by a 14-hour
night, followed by another ten-hour day. Nominal conditions for the
independent variables are assumed to be as follows:
Solar collector variables:
= 0.7
q = 500 W/m2
Pool variables:
L = 50 m (pool length)
W = 25 m (pool width)
H = 2 m (pool water height)
G = 350 W/m2
w = 0.96
pool water surface = 0.96
w = 1000 kg/m3

184

Sizing An Active Solar Collector For A Pool

cpw = 1.162 W-hr/kg-K


kpool wall = 0.7 W/m-K
tpool wall = 0.1 m (average wall thickness)
kcover = 0.04 W/m-K
tcover = 0.02 m
pool cover surface = 0.1
Tpool initial = 20C (pool water temperature at = 0)
Environmental variables:
Tenvironment during day = 30C (average environmental temperature
during the ten-hour day)
Tenvironment during night = 15C (average environmental temperature
during the fourteen-hour night)
Tearth = 20C (average earth temperature around pool walls)
Other properties during the day:
hday = 2 W/m2-K
a = 1.1614 kg/m3
cpa = 0.28 W-hr/kg-K
= 40% (relative humidity of air during the day)
pw = saturation pressure of water vapor on pool water surface at
pool temperature is obtained from dry saturated steam temperature
tables (see Reference [8]).
psaturation = saturation pressure of water vapor in air at average
environmental temperature is obtained from dry saturated steam
temperature tables (see Reference [8]).
Other properties at night:
D = 0.0936 m2/hr
= 0.0572 m2/hr
Pr = 0.707
ka = 0.0263 W/m-K
V = 3600 m/hr
= 40% (relative humidity of air during the night)
pw = saturation pressure of water vapor on pool water surface at
pool temperature is obtained from dry saturated steam temperature
tables (see Reference [8]).

185

Everyday Heat Transfer Problems

psaturation = saturation pressure of water vapor in air at average


environmental temperature is obtained from dry saturated steam
temperature tables (see Reference [8]).
Under these nominal conditions, and with a pool cover at night, a
solar collector of 1192 m2 (95% of the pool surface area) is required
to heat the pool water from the initial temperature of 20C to 25C
in two days. Without a pool cover at night, the solar collector area
requirement almost doubles to 2054.5 m2. Radiation heat loss to night
sky and evaporation heat loss to environment cause the required
collector size to double without a pool cover at night.
Solar collector efciency and average solar insolation are the two
dominant variables that affect similarly the required solar collector area.
Required solar collector area is very sensitive to the product of solar
collector efciency and the average solar insolation. Solar collector area
versus its efciency is shown in Figure 20-1. The sensitivity of solar
collector area to its efciency is shown in Figure 20-2. Sensitivity of
required solar collector area to these variables increases fast as collector
efciency and solar insolation values decrease.

Collector Area, m2

8000
6000
4000
2000
0

0.2

0.4
0.6
Collector Efficiency

0.8

Figure 20-1 Solar collector area versus solar collector efciency at average

solar insolation of 500 W/m2

186

(Collector Area) / , m2

Sizing An Active Solar Collector For A Pool

10000

20000

30000
0.2

0.4

0.6

0.8

Collector Efficiency,

Figure 20-2 Sensitivity of solar collector area to collector efciency at

average solar insolation of 500 W/m2

Initial pool water temperature, average environmental temperature


during the day, and the earth's temperature under the pool have
signicant effects on sizing the solar collector area. These effects are
shown in Figures 20-3, 20-4 and 20-5, respectively, and their behaviors

Collector Area, m2

3000

2000

1000

15

17

19

21

23

Pool Water Initial Temperature, C

Figure 20-3 Solar collector area versus pool water initial temperature

187

Everyday Heat Transfer Problems

Collector Area, m2

1250

1220

1190

1160

25
27
29
31
33
35
Average Environment Temperature During The Day, C

Figure 20-4 Solar collector area versus average environmental temperature

during the day

Collector Area, m2

2000

1500

1000

500

10

15

20

25

Average Earth Temperature Under The Pool, C

Figure 20-5 Solar collector area versus average earth temperature under

the pool

188

Sizing An Active Solar Collector For A Pool

Collector Area, m2

1500
1300
1100
900
700

30

35

40

45
50
Pool Length, m

55

60

Figure 20-6 Solar collector area versus pool length

are almost linear. Required solar collector area decreases 384 m2


per degree centigrade increase in the initial pool water temperature.
Similarly, required solar collector area decreases at a much smaller
rate, 7.2 m2, per degree centigrade increase in average environmental

Collector Area, m2

2000
1600
1200
800
400

10

15

20

25

30

35

40

Pool Width, m

Figure 20-7 Solar collector area versus pool width

189

Everyday Heat Transfer Problems

Collector Area, m2

2000
1600
1200
800
400
0

1.2

1.4

1.6

1.8

2.2

2.4

2.6

Pool Water Height, m

Figure 20-8 Solar collector area versus pool water height

temperature during the day. Required solar collector area sensitivity to


average earth temperature under the pool is 52.7 m2/C.
Another set of variables signicantly affecting required solar collector
area are the pool's physical dimensions. Sensitivities for pool length,
width and water height are shown in Figures 20-6, 20-7 and 20-8,
respectively. These sensitivities are constants. A 23.7 m2 increase in
solar collector area is required for every meter of increase in pool length,
a 47 m2 increase in solar collector area is required for every meter of
increase in pool width, and a 1049 m2 increase in solar collector area is
required for every meter of increase in pool water height.
Solar insolation on the pool water surface during the day also
affects required solar collector area signicantly. The effects on
required solar collector area of average solar insolation on pool water
surface, and of water absorptivity, are depicted in Figures 20-9 and
20-10, respectively. Required solar collector area decreases with
increasing average solar insolation on pool water surface by 3.4 m4/W.
Required solar collector area decreases 12.5 m2 per 1% increase in
water surface absorptivity of incoming solar radiation.
Relative humidity also has an effect on sizing the solar collector.
This effect is compared in Figure 20-11 for a pool with and without

190

Sizing An Active Solar Collector For A Pool

Collector Area, m2

2000

1500

1000

500
200

250

300

350

400

450

500

Average Solar Insolation On Pool Surface, W/m2

Figure 20-9 Solar collector area versus average solar insolation on pool

surface

a cover at night. Required solar collector area decreases 4.8 m2 for


every percent increase in relative humidity for a pool that is covered
at night. The sensitivity for a pool without a cover at night is a 6.5 m2
collector area decrease per 1% increase in relative humidity. This

Collector Area, m2

1800
1600
1400
1200
1000
0.5

0.6
0.7
0.8
0.9
Absorptivity Of Pool Water Surface

Figure 20-10 Solar collector area versus absorptivity of pool water surface

191

Everyday Heat Transfer Problems

Collector Area, m2

2800
Without
Cover At
Night

2300
1800

With Cover
At Night

1300
800

0.2

0.4

0.6

0.8

Relative Humidity

Figure 20-11 Solar collector area versus relative humidity

higher sensitivity comes from higher evaporation heat loss to the cold
night atmosphere.
If a region with a relative humidity of 10% is considered to build a
solar heated pool, as compared to a region with a relative humidity
of 60%, with everything else being the same and considering a pool
cover at night, the dry region requires 240 m2 more solar collector
area because more heat is lost due to evaporation to the dry
atmosphere.
There are over 25 independent variables that govern this heat
transfer problem. When the nominal values of the variables given
above are varied 10%, the results shown in Table 20-1 are obtained
for a pool that is covered at night. Required solar collector area
sensitivities to a 10% change in the governing variables are given in
descending order of importance, and they are applicable only around
the nominal values assumed for this study.
Required solar collector area is most sensitive to changes in initial
pool water temperature, followed by pool water height and pool water
specic heat at constant pressure. The next set of variables in order
of sensitivity are solar collector efciency, average solar insolation on
collector, average solar insolation on pool water surface, average solar

192

Sizing An Active Solar Collector For A Pool

Table 20-1 Effects of a 10% change in nominal values of


variables to required solar collector area for a
pool covered at night (Note: Effects less than 1%
are not included in Table 20.1)

Variable

Nominal
Value

Change In
Required Solar
Collector Area
For A 10%
Decrease In
Nominal Value

Tpool initial

20C

+64.35%

64.35%

H, pool water
height

2m

17.53%

+17.62%

1.162
W-hr/kg-K

17.37%

+17.45%

, solar collector
efciency

0.7

+11.16%

9.06%

q, average solar
insolation on
collector

500 W/m2

+11.16%

9.06%

G, average solar
insolation on
pool water surface

350 W/m2

+10.07%

10.07%

cpw

Change In
Required Solar
Collector Area
For A 10%
Increase In
Nominal Value

w, average
solar radiation
absorptivity at
water surface

0.96

+10.07%

4.19% with
only a possible
4.2% increase
to 1

L, pool length

50 m

9.90%

+9.98%

W, pool width

25 m

9.82%

+9.90%

Tearth

20C

+8.89%

8.81%

Tenvironment during day

30C

+1.85%

1.76%

, relative humidity

0.40

+1.60%

1.59%

tpool wall

0.1 m

+1.26%

1.01%

kpool wall

0.7 W/m-K

1.09%

+1.17%

193

Everyday Heat Transfer Problems

radiation absorptivity at water surface, pool length, pool width, and


average earth temperature around pool walls.
The sensitivities then drop an order of magnitude to less than 2%.
The average environmental temperature during the day continues in
the order of sensitivity, followed by relative humidity, average pool
wall thickness, and pool wall thermal conductivity.
The rest of the independent variables have less than 1%
sensitivities on the required solar collector area. These variables,
in order of descending sensitivity, are the emissivity of the pool
cover, temperature of the environment at night, the convection heat
transfer coefcient during the day, thermal conductivity of pool cover,
thickness of pool cover, mean air speed over the pool at night, and
thermophysical properties of air.

194

CHAPTER

HEAT

TRANSFER
IN A HEAT
EXCHANGER

21

he design of a heat exchanger can be a good challenge for


engineers. The design methods cover a vast variety of engineering
disciplines, such as heat transfer, uid mechanics, stress analysis,
corrosion, materials, economics, etc. There are two popular heat
transfer design methods covered in literature; see References [5], [6],
[10] and [15].
One heat transfer design method is called the log mean
temperature difference method. If the hot and cold uid inlet and/or
outlet temperatures are specied, the heat exchanger design can be
performed by using this method. On the other hand, if the hot and
cold uid inlet and/or outlet temperatures are not specied, and a
comparison between various types of heat exchangers is required,
then the effectiveness method is preferred.
In this chapter, the sensitivity of a heat exchanger design to
governing independent variables is analyzed. Heat transfer in a
counterow concentric pipe liquid-to-liquid heat exchanger is
considered using the log mean temperature difference design method.
Hot engine oil ows in the inner pipe and the cold water ows in
the outer pipe. Under steady-state conditions, by neglecting heat

195

Everyday Heat Transfer Problems

transfer to the environment, and assuming constant thermophysical


properties, the heat transferred through a small element of the heat
exchanger from the hot oil equals the heat received by the cold water.
dQ = hVmhAh cph dTh = cVmcAc cpc dTc

(21-1)

Heat transferred through a small element of the heat exchanger in


Eq. (21-1) can also be expressed in terms of an overall heat transfer
coefcient between the hot uid and the cold uid, through the
separating wall:
dQ = U (Th Tc) dAwall

(21-2)

Eqs. (21-1) and (21-2) can be combined to eliminate dQ, and then
integrated between the inlet and outlet temperatures of the heat
exchanger to give the following log mean temperature difference
design method equations for a counterow heat exchanger:

Q=

Q = hVmhAh cph (Th in Th out)

(21-3)

Q = cVmcAc cpc (Tc in Tc out)

(21-4)

UAwall [(Th out Tc in) (Th in Tc out)]


ln[(Th out Tc in)/(Th in Tc out)]

(21-5)

where
UAwall = UhAwall hot uid side = UcAwall cold uid side

(21-6)

and Awall hot uid side represents inside surface area of the inner tube and
Awall cold uid side represents outside surface area of the inner tube. The
overall heat transfer coefcient from the hot uid side is:
Uh = 1/[(1/hh) + Rh foul + (Dh ln(Dc/Dh)/2kss)
+ (Awall hot uid side Rc foul/Awall cold uid side)
+ (Awall hot uid side/Awall cold uid sidehc)]

196

(21-7)

Heat Transfer In A Heat Exchanger

All the variables in these governing Eqs., (21-1) through (21-7), are
dened as follows:
h = Hot oil density in kg/m3
VmhAh = Hot oil ow rate in liters/minute
cph = Hot oil specic heat at constant pressure in J-kg/C
Th in = Hot oil inlet temperature to the heat exchanger in C
Th out = Hot oil outlet temperature from the heat exchanger in C
c = Cold water density in kg/m3
VmcAc = Cold water ow rate in liters/minute
cpc = Cold water specic heat at constant pressure in J-kg/C
Tc in = Cold water inlet temperature to the heat exchanger in C
Tc out = Cold water outlet temperature from the heat exchanger in C
Uh = Overall heat transfer coefcient based on the inside surface of
the inner tube in W/m2-C
Awall hot uid side = Inside surface area of the inner tube, DhL, in m2
Uc = Overall heat transfer coefcient based on the outside surface
of the inner tube in W/m2-C
Awall cold uid side = Outside surface area of the inner tube, (Dh + 2t)L =
DcL, in m2
Dh = Inside diameter of the inner tube in m
Dc = Outside diameter of the inner tube, Dh + 2t, in m, and where t
is the thickness of the inner tube
kss = Wall tube material thermal conductivity in W/m-C
hh = Convection heat transfer coefcient between the hot oil and
the inner tube inside surface in W/m2-C
hc = Convection heat transfer coefcient between the cold water
and the inner tube outside surface in W/m2-C
Rh foul = Fouling resistance for the inside surface of the inner tube in
m2-C/W
Rc foul = Fouling resistance for the outside surface of the inner tube
in m2-C/W
In a heat exchanger, fouling resistances affecting heat transfer
on surfaces of walls are caused by corrosion or by foreign material
deposited over time, and they are determined by experimental
methods during heat exchanger life tests.

197

Everyday Heat Transfer Problems

In order to be able to determine the convection heat transfer


coefcients, rst the Reynolds numbers for inner and outer tubes
have to be determined.
For the inner tube the Reynolds number denition is given by
ReDinner tube = hVmhDh/h , where h is the viscosity of hot oil in N-s/m2.
For the outer tube the Reynolds number denition is given by
ReDouter tube = cVmcDc-hydraulic/c , where c is the viscosity of cold
water in N-s/m2. Dc-hydraulic is the hydraulic diameter of the outer tube
given as Dc-hydraulic = 4 ow cross-sectional area/ow wetted perimeter,
or Dc-hydraulic = Dinside diameter of outer tube Dc in meters.
The nominal values for the independent variables of the present
heat exchanger sensitivity analysis are assumed to be as follows:
For hot oil variables:
VmhAh = 5 liters/minute
Th in = 70C
Th out = 40C
h = 870 kg/m3
cph = 2000 J-kg/C
h = 0.1 N-s/m2
kh = 0.14 W/m-C
For cold water variables:
VmcAc = 25 liters/minute
Tc in = 20C
c = 1000 kg/m3
cpc = 41.8 J-kg/C
c = 0.00096 N-s/m2
kc = 0.6 W/m-C
For geometry and other variables:
Dh = 0.02 m
t = 0.001 m
Douter tube = 0.025 m
kss = 15 W/m-C
Rh foul = 0.0008 m2-C/W
Rc foul = 0.0002 m2-C/W

198

Heat Transfer In A Heat Exchanger

ReDinner tube is calculated to be 46, which is less than 2000, and


therefore the hot oil ows in the laminar region. It is assumed that the
laminar ow is fully developed and has constant heat ux at the wall.
Therefore the Nusselt number, NuDh, is a constant (see Reference [6]):
NuDh = (hhDh)/kh = 4.36

(21-8)

ReDouter tube is calculated to be 11,758, which is greater than 4000, and


therefore the cold water ows in the turbulent ow region. It is assumed
that the turbulent ow is fully developed and the Nusselt number, NuDc,
is determined from an empirical relationship (see Reference [6]):
NuDc = (hcDc-hydraulic)/kc = 0.023 (ReDouter tube)0.8 Prc0.4

(21-9)

Heat Exchanger Length, m

where Prc is the Prandtl number, the ratio of momentum to thermal


energy diffusion, dened as (c/c)/(kc/ccpc) or (ccpc/kc). c is the
viscosity of cold uid and kc is the thermal conductivity of cold uid.
The counterow heat exchanger length is considered as the main
dependent variable. Heat exchanger length versus hot oil ow rate is
given in Figure 21-1. The behavior is linear since the hot oil is owing

300

200

100

10

15

Hot Oil Flow Rate, L/min

Figure 21-1 Heat exchanger length versus hot oil ow rate

199

Length of Counterflow Heat


Exchanger, m

Everyday Heat Transfer Problems

160
120
80
40
0

30

35

40

45

50

Hot Oil Outlet Temperature, C

Figure 21-2 Heat exchanger length versus hot oil outlet temperature for

Length of Counterflow Heat


Exchanger, m

Th in = 70C and Tc in = 20C

300

200

100

20
30
10
Cold Water Inlet Temperature, C

40

Figure 21-3 Heat exchanger length versus cold water inlet temperature for

Th in = 70C and Th out = 40C

200

Heat Transfer In A Heat Exchanger

Table 21-1 Effects of a 10% change in nominal values of


independent variables to heat exchanger length
for a concentric tube counterow heat exchanger

Variable
Th out

Nominal
Value

Heat
Exchanger
Length
Change For A
10% Decrease
In Nominal
Value

Heat
Exchanger
Length
Change For A
10% Increase
In Nominal
Value

40C

+24.98%

20.27%

70C

16.78%

+14.65%

0.14 W/m-C

+10.74%

8.79%

5 liters/minute

10.27%

+10.33%

870 kg/m

10.27%

+10.33%

cph

2000 J-kg/C

10.27%

+10.33%

Tc in

20C

6.25%

+7.21%

Dh

0.02 m

+0.448%

0.377%

25 liters/minute

+0.347%

0.282%

1000 kg/m

+0.347%

0.282%

cpc

41.8 J-kg/C

+0.341%

0.277%

0.0008 m -C/W

0.235%

+0.236%

0.025 m

0.116%

+0.126%

0.0002 m -C/W

0.049%

+0.049%

15 W/m-C

+0.040%

0.033%

0.001 m

+0.013%

0.014%

kc

0.6 W/m-C

+0.0090%

0.0077%

0.00096 N-s/m

0.0057%

+0.0054%

0.1 N-s/m2

0%

0%

Th in
kh
VmhAh

VmcAc

Rh foul
Douter tube
Rc foul
kss

201

Everyday Heat Transfer Problems

in the laminar ow region and the convection heat transfer coefcient


is independent of ReDinner tube.
Hot oil inlet and outlet temperatures, and cold water inlet
temperature, have dominant effects on the heat exchanger length.
Heat exchanger length as a function of hot oil outlet temperature
is given in Figure 21-2. The heat exchanger length decreases
logarithmically as the requirement for the hot oil outlet temperature
increases.
Heat exchanger length is also a strong function of cold water inlet
temperature, and is shown in Figure 21-3. The heat exchanger length
increases as the cold water inlet temperature increases, and the
length increase behaves exponentially as the required hot oil outlet
temperature is approached.
When the nominal values of the independent variables given above
are varied 10%, the results shown in Table 21-1 are obtained. These
heat exchanger length sensitivities are given in descending order of
importance, and they are applicable only around the nominal values
assumed for this case study.
Heat exchanger length is most sensitive to hot oil outlet and
inlet temperature requirements. Hot oil ow rate and hot oil
thermophysical properties also affect the heat exchanger length at the
same order of magnitude as the hot oil outlet and inlet temperatures.
From the cold water variables, the heat exchanger length is most
sensitive to the cold water inlet temperature. The sensitivities of
the heat exchanger length to the rest of the independent variables
diminish fast in the present region of application. The heat exchanger
length is least sensitive to changes in fouling resistances, inside
diameter of the outer tube, thickness of the inner tube, thermal
conductivity of the tube wall material, viscosity of water, and thermal
conductivity of water.

202

CHAPTER

I CE

FORMATION
ON A LAKE

22

eat transfer during ice formation has been studied in detail, in


papers by Lin, S. and Z. Jiang (Reference [13]), and by London,
A. L. and R. A. Seban (Reference [14]). In the present sensitivity
study, planar ice formation is considered. References [13] and [14]
also treat ice formation in cylindrical and spherical shapes. In the
present heat transfer analysis, thermal capacitance of ice is neglected;
in other words, a linear temperature prole through the ice thickness
is assumed. Ice thermophysical properties are assumed to be
constants and are considered at sea level conditions. Lake water is
assumed to be fresh. For proper ice formation heat transfer modeling,
lake water temperature is assumed to be greater than zero degrees
centigrade, and the atmospheres temperature is assumed to be less
than zero degrees centigrade:
Tlake water > Tfreezing > Tatmosphere .
Heat transfer from the lake water, through the ice layer, to
the atmosphere is assumed to be in an unsteady state and in

203

Everyday Heat Transfer Problems

one-dimension rectangular coordinates. Governing heat transfer rate


equations are as follows:
Convection heat transfer from the lake water to the growing ice
surface is:
Qlake water to growing ice layer = hlake water-growing ice surface (Tlake water Tfreezing) (22-1)
Heat transfer from the growing ice layer to the atmosphere is:
Qgrowing ice layer to atmosphere
= (Tfreezing Tatmosphere)/[(1/hice upper surface-atmosphere) + (x/kice)] (22-2)
Energy balance at the growing ice layer for the latent heat of fusion
required per unit area of ice layer formation is:
iceHice (dx/dt) = Qgrowing ice layer to atmosphere
Qlake water to growing ice layer

(22-3)

T represents temperature in C. h represents the convection heat


transfer coefcient in W/m2-C. kice is the thermal conductivity of ice
in W/m-C. ice is the density of ice in kg/m3. Hice is the latent heat of
fusion for water in W-hr/kg. x is the ice thickness in meters at time, t,
in hours.
The governing differential Eq. (22-3) can be rewritten after
separating the time and the space variables in the following
non-dimensional form:
dt* = (1 + x*)dx*/[1 Z(1 + x*)]

(22-4)

where the non-dimensional time is


t* = [h2ice upper surface-atmosphere (Tfreezing Tatmosphere) t/(ice Hice kice)] (22-5)
where the non-dimensional ice thickness is
x* = hice upper surface-atmosphere x/kice

204

(22-6)

Ice Formation On A Lake

and where Z is a non-dimensional parameter which is the ratio of two


convection heat transfer mechanisms that govern this heat transfer
problem:
Z=

[hlake water-growing ice surface (Tlake water Tfreezing)]


[hice upper surface-atmosphere (Tfreezing Tatmosphere)]

(22-7)

Eq. (22-4) is integrated using the initial condition


x* = 0 at t* = 0

(22-8)

to obtain the following ice formation rate equation:


t* = (1/Z2) ln[(Z 1)/(Z(x* + 1) 1)] (x*/Z)

(22-9)

Nominal values of the independent variables used in the present


sensitivity analysis are as follows:
hice upper surface-atmosphere = 30 W/m2-C
hlake water-growing ice surface = 10 W/m2-C
Tlake water = 6C
Tatmosphere = 20C
Tfreezing = 0C
ice = 920 kg/m3
kice = 1.88 W/m-C
Hice = 93 W-hr/kg
The time to form a certain thickness of ice under the nominal
conditions given above is shown in Figure 22-1. The ice thickness
limit under the above given nominal conditions that will support the
latent heat of fusion required for ice formation is 0.564 m.
Maximum ice thickness versus the convection heat transfer
coefcient between ice upper surface and atmosphere is given in
Figure 22-2. This condition occurs when the convection heat transfer
from the upper surface of the ice to the atmosphere starts to reach
the same level as the convection heat transfer from the lake water to

205

Everyday Heat Transfer Problems

Ice Thickness Limit = 0.564 m


Under Given Nominal Conditions
Time To Form Ice, hr

2000
1500
1000
500
0

0.1

0.2
0.3
0.4
Thickness Of Ice, m

0.5

0.6

Figure 22-1 Time to form ice versus thickness

Maximum Ice Thickness, m

the growing ice surface. Then there is no more energy left to support
the latent heat of fusion of water to keep forming the ice layer.
Under this condition, the natural logarithm term in the denominator
of the ice formation rate Eq. (22-9) approaches zero, or:
0.6

0.4

0.2

10

20

30

40

50

Heat Transfer Coefficient, hice upper surface-atmosphere, W/m2-C

Figure 22-2 Maximum ice thickness versus heat transfer coefcient

between ice upper surface and atmosphere

206

Ice Formation On A Lake

x* = (1/Z) 1

(22-10)

t0.2 m / Tatmosphere, hr/C

Maximum ice thickness is very sensitive to variations in the


convection heat transfer coefcient between ice upper surface and
atmosphere at low values.
The time to achieve a certain thickness of ice depends on the values
of seven independent variables. In this study, sensitivities of time to
form 0.2 meters of ice thickness are investigated. Sensitivities of time
to form 0.2 meters of ice thickness to temperatures of atmosphere and
of lake water are shown in Figures 22-3 and 22-4, respectively.
In Figure 22-3, as the temperature of the atmosphere increases, the
sensitivity of time to form 0.2 meters thick of ice increases. Under the
present conditions, the temperature of the atmosphere has a limit of
8.4C for the ice to be able to reach a thickness of 0.2 meters
and support the energy requirement for freezing. Similarly, in
Figure 22-4, as the temperature of the lake water increases, the
sensitivity of time to form 0.2 meters thick of ice increases. Under the
present conditions, the temperature of the lake water has a limit of
14.3C for the ice to be able to reach a thickness of 0.2 meters.

300
250
200
150
100
50
0
50

40

30

20

10

Temperature Of Atmosphere, C

Figure 22-3 Sensitivity of time to form 0.2 m thick of ice to temperature of

atmosphere

207

Everyday Heat Transfer Problems

t0.2 m / Tlakewater, hr/C

80
60
40
20
0

6
8
10
Temperature Of Lake Water, C

12

14

Figure 22-4 Sensitivity of time to form 0.2 m thick of ice to temperature of

lake water

Sensitivities of time to form 0.2 meters thick of ice to


hice upper surface-atmosphere and hlake water-growing ice surface are shown in
Figures 22-5 and 22-6, respectively.
The convection heat transfer coefcients on both sides of the ice
have opposite effects to the growth of ice. In Figure 22-5, as the
convection heat transfer coefcient between the ice upper surface
and the atmosphere decreases, the absolute value of ice formation
time sensitivity increases, or the time for formation of 0.2 meters of
ice increases. On the other hand, in Figure 22-6, as the convection
heat transfer coefcient between the lake water and the growing ice
surface increases, the ice formation time and its sensitivity increase.
The ice formation time sensitivities to ice density and to the
latent heat of fusion of water are constants. The ice formation time
sensitivity to ice density is positive, 0.115 hr-m3/kg. The ice formation
time sensitivity to latent heat of fusion of water is also positive,
1.14 kg/W. The ice formation time sensitivity to thermal conductivity
of ice is non-linear, and is given in Figure 22-7.
As the thermal conductivity of ice increases, the time to form 0.2
meters thick of ice decreases, as does the absolute value of its sensitivity.

208

t0.2 m / hice upper surfaceatmosphere,


hr-m2-C/W

Ice Formation On A Lake

0
5
10
15
20
25

100

20
40
60
80
hice upper surfaceatmosphere, W/m2-C

Figure 22-5 Sensitivity of time to form 0.2 m thick of ice to the convection

t0.2 m / hlake water-growing ice surface,


hr-m2-C/W

heat transfer coefcient between ice upper surface and


atmosphere

30

20

10

10

15

20
hlake water-growing ice surface, W/m2-C

25

Figure 22-6 Sensitivity of time to form 0.2 m thick of ice to the convection

heat transfer coefcient between lake water and growing ice


surface

209

Everyday Heat Transfer Problems

t0.2 m / kice, hr-m-C/W

0
50
100
150
200
250
1

1.5

2.5

Thermal Conductivity Of Ice, W/m-C

Figure 22-7 Sensitivity of time to form 0.2 m thick of ice to thermal

conductivity of ice

Table 22-1 Effects of a 10% change in nominal values of


variables to time for ice thickness to reach 0.2 m

Variable

Nominal
Value

Tatmosphere

20C

Change in
Change in
Time For Ice
Time For Ice
Thickness to
Thickness to
Reach 0.2 m For Reach 0.2 m For
A 10% Decrease A 10% Increase
In Nominal
In Nominal
Value
Value
12.67%

+17.05%

kice

1.88 W/m-C

+10.87%

8.28%

ice

920 kg/m

10%

+10%

Hice

93 W-hr/kg

10%

+10%

hice upper surface-atmosphere

30 W/m -C

+5.80%

4.61%

hlake water-growing ice surface

10 W/m -C

4.31%

+4.78%

6C

4.31%

+4.78%

Tlake water

210

2
2

Ice Formation On A Lake

When the nominal values of the independent variables given


above are varied 10%, the results shown in Table 22-1 are obtained.
Sensitivities of time for the ice thickness to reach 0.2 meters to a
10% change in the governing independent variables are given in
descending order of importance, and they are applicable only around
the nominal values assumed for this study.
Ice thickness formation time is most sensitive to the temperature of
the atmosphere. Ice thermophysical properties are next in the order
of sensitivity. Convection heat transfer coefcients on both sides of
the ice surfaces, and the lake water temperature, round up the ice
formation time sensitivity. Ice formation time sensitivities to all the
independent variables are non-linear, except for ice density and latent
heat of fusion for water. All the ice formation time sensitivities for
the nominal independent variables values assumed for the present
analysis are at the same order of magnitude.

211

CHAPTER

SOLIDIFICATION 23

IN A CASTING
MOLD

xact temperature distribution solutions to unsteady-state onedimensional heat conduction equations in rectangular, cylindrical,
and spherical coordinates, for initial condition and different boundary
conditions, have been provided in literature (see References [1], [6],
[10] and [16]). The reference by Carslaw, H. S. and J. C. Jaeger [1]
extended these exact solutions to moving boundary conditions with
phase change.
In this chapter, sensitivity analysis for the solidication front
of a semi-innite liquid in a semi-innite mold is considered. The
unsteady-state one-dimensional heat conduction equation in
rectangular coordinates for the mold region, solidied cast material
region and liquid cast material region, with constant property
assumptions, is given by:
2T/x2 = (1/) T/

(23-1)

where temperature T varies with space x and with time . is


thermal diffusivity, k/cp, of the region in m2/s, where k is the thermal

213

Everyday Heat Transfer Problems

conductivity of the region in W/m-C, is the density of the region in


kg/m3, and cp is the specic heat of the region at constant pressure in
J/kg-C. Initial and boundary conditions for each region are as follows:
Minus innity to x = 0 is assumed to be the mold region. In this
case, it is assumed that the mold is made out of sand. x > 0 to plus
innity is the cast materials region, which is initially lled with liquid
silver. The solidication front for the liquid silver starts initially at
x = 0 and grows in the +x direction.
Eq. (23-1) is applied to three different regions, in order to obtain
the temperature distributions in each region with the following initial
and boundary conditions:
At zero time, all the mold region from x = 0 to x = is at the
initial temperature of the sand, Tsand:
T(x, 0) = Tsand

(23-2)

The temperature of the mold region far away from x = 0 always


stays at the initial temperature of the sand, Tsand:
T(, ) = Tsand

(23-3)

The interface between the mold region and the solidied material
at x = 0 has the following energy balance boundary condition:
ksand Tsand/x = ksolid Tsolid/x

(23-4)

The solidied region initial condition is:


L = 0 at = 0

(23-5)

where L is the solidication front in meters.


The solidication region boundary at x = L is always at the
solidication temperature of the liquid cast material:
T(L, ) = Tsolidication

214

(23-6)

Solidication In A Casting Mold

At zero time, all the liquid cast material region from x = 0 to x = +


is at the initial temperature of the liquid cast material, Tliquid:
T(+x, 0) = Tliquid

(23-7)

The temperature of the liquid cast material region far away from
x = 0 always stays at the initial temperature of the liquid cast
material, Tliquid:
T(+, ) = Tliquid

(23-8)

The interface between the liquid cast material region and the
solidied cast material region at x=L has the following boundary
condition, which is the result of an energy balance at the solidication
front that supports the latent heat of fusion for the liquid cast
material:
kliquid Tliquid/x ksolid Tsolid/(x = solid HdL/d)

(23-9)

where H is the latent heat of fusion in W-s/kg. Exact solutions


to Eq. (23-1) have been obtained using the conditions in
Eq. (23-2) through Eq. (23-9) for three regions, and they can be
found in Reference [1].
In the present sensitivity analysis, propagation of the solidication
front is analyzed instead of temperature distributions for each region.
Reference [1] shows that in order to satisfy the temperature
continuity at the solidication front all the time, the solidication
front L has to move proportionately to the square root of time,
namely:
L = a0.5

(23-10)

where a is the proportionality constant, has the dimensions of


m/s0.5, and can be obtained by satisfying the energy balance,
Eq. (23-9), at the interface of the solidied cast material and liquid
cast material (see Reference [1]):

215

Everyday Heat Transfer Problems

exp(a2/4solid)/{[(solidcp solidksolid)/(sandcp sandksand)]0.5


+ erf(a/20.5solid)} [(liquidcp liquidkliquid)/(solidcp solidksolid)]0.5
[(Tliquid Tsolidication)/(Tsolidication Tsand)]
exp[a2(solid/liquid)2/4liquid]/erfc[a(solid/liquid)/20.5liquid]
aH(solid/ksolidcp solid)0.5/2(Tsolidication Tsand) = 0

(23-11)

where the error function erf is dened as


z

erf(z) = (2/0.5) 0 exp(u2) du

(23-12)

and the complementary error function erfc is dened as


erfc(z) = 1 erf(z)

(23-13)

Eq. (23-11) is solved for a by trial and error. For the present
analysis, liquid silver is considered as the cast material, and sand is
considered as the mold material. For solidication to occur there are
two conditions that have to be met. One condition is that the liquid
cast materials temperature cannot exc eed the melting temperature
of the mold material:
Tliquid < Tsand melting point

(23-14)

The second condition is that the liquid cast material has to have
enough heat loss initially through the mold by conduction heat
transfer to be able to start solidifying. Eq. (23-11) provides this
condition by setting a equal to zero:
(Tliquid Tsolidication)/(Tsolidication Tsand)
< [(sandcp sandksand)/(liquidcp liquidkliquid)]0.5 (23-15)
For the present sensitivity analysis, the following nominal values
are assumed for the independent variables that also satisfy Eqs.
(23-14) and (23-15):
Sand values for the mold region:

216

Solidication In A Casting Mold

Tsand = 100C
Tsand melting point = 1430C
sand = 2330 kg/m3
cp sand = 712 J/kg-C
ksand = 42 W/m-C
Solidied silver values for the solidied cast region:
Tsolidication = 960C
solid = 10500 kg/m3
cp solid = 235 J/kg-C
ksolid = 400 W/m-C

Solidification Front L, m

Liquid silver values for the liquid cast region:


Tliquid = 1000C which is less than Tsand melting point of 1430C to satisfy
Eq. (23-14), and is less than 1180C, which is the a equals zero
condition in Eq. (23-15) to start solidication.
liquid = 9300 kg/m3
cp liquid = 318 J/kg-C
kliquid = 360 W/m-C
H = 104400 W-s/kg

0.4
0.3
0.2
0.1
0

10

15
Time, min

20

25

30

Figure 23-1 Cast solidication front L versus time

217

Everyday Heat Transfer Problems

a is obtained by iteration for these nominal values, and is 0.008137


meters/second0.5. The solidication front versus time for the nominal
case is given in Figure 23-1.
When the nominal values of the variables given above are varied
10%, the results shown in Table 23-1 are obtained. a sensitivities to
a 10% change in the governing variables are given in descending order
of importance, and they are applicable only around the nominal values
assumed for this study. The variable Tliquid is varied on the negative
side 4% in order not to fall below Tsolidication. The variable Tsolidication is
varied on the positive side +4.2% in order not to exceed Tliquid.

Table 23-1 Effects of a 10% change in nominal values of


variables to a in solidication front Eq. (23-10),
L = a0.5

Nominal
Value

Change in a
For A 10%
Decrease In
Nominal Value

Change in a
For A 10%
Increase In
Nominal Value

Tliquid

1000 C

20.738% for
4% change

+69.208%

Tsolidication

960 C

+58.119%

17.677% for
4.2% change

104400W-s/kg

+9.745%

8.206%

solid

10500 kg/m

+9.641%

8.062%

sand

2330 kg/m

3.192%

+2.933%

cp sand

712 J/kg-C

3.192%

+2.933%

ksand

42 W/m-C

3.192%

+2.933%

cp liquid

318 J/kg-C

1.520%

+1.517%

ksolid

400 W/m-C

1.324%

+1.143%

Variable

Tsand

100 C

+0.773%

0.777%

liquid

9300 kg/m3

0.717%

+0.686%

kliquid

360 W/m-C

0.717%

+0.686%

cp solid

235 J/kg-C

+0.673%

0.652%

218

Solidication In A Casting Mold

Solidication front propagation parameter a is most sensitive


to the temperature of the liquid cast material and the solidication
temperature of the cast material. The second tier of a sensitivity
belongs to the latent heat of fusion and to density of solid region. In
order of sensitivity, the next set of independent variables are mold
region thermophysical properties, namely sand, cp sand, and ksand.
a is least sensitive to variations in independent variables cp liquid ,
ksolid , Tsand , liquid , kliquid and cp solid. Changes to these least sensitive
independent variables affect a two orders of magnitude less than
changes to temperature of the liquid cast material and changes of
solidication temperature of the cast material.

219

CHAPTER

AVERAGE

TEMPERATURE
RISE IN SLIDING
SURFACES IN
CONTACT

24

rictional heating of materials in contact has been studied


thoroughly in heat transfer literature. In this case, an approximate
method developed by J. C. Jaeger (see Reference [7]) for frictional
temperature rise on a sliding square contact area will be analyzed for
sensitivity.
In order to determine the steady-state average temperature
at contacting surfaces, Jaeger [7] used the temperature solution
resulting from an instantaneous point source in an innite solid for
the following unsteady-state and three-dimensional conduction
heat transfer differential equation, in rectangular coordinates
(see Reference [1]):
2T/x2 + 2T/y2 + 2T/z2 = (1/)(T/)

(24-1)

The temperature distribution and history resulting from an


instantaneous point source at (x1, y1, z1) of strength Q that satises
Eq. (24-1) can be expressed as:
T = [Q/8()0.5] exp{[(x x1)2 + (y y1)2 + (z z1)2]/4}

(24-2)

221

Everyday Heat Transfer Problems

T in Eq. (24-2) represents the temperature distribution and


history in an innite medium at location (x, y, z) and at time . is
the constant thermal diffusivity of the medium dened as (k/cp),
where k is the mediums thermal conductivity, is mediums density,
and cp is mediums specic heat at constant pressure. Eq. (24-2) can
be integrated to obtain steady-state temperatures for moving and
stationary line and square contact areas, where heat transfer was
restricted only to contact area (see References [1] and [7]).
In the present sensitivity analysis, a square contact area, 2L 2L,
that is sliding with a constant velocity, V, over a semi-innite body is
considered. For LV/2 > 5, the average temperature at the surface of
the semi-innite medium that is under the sliding square contact area,
designated as area 1, is approximated as follows (see Reference [7]):
Taverage at surface of contacted semi-innite medium = (1.064 Q/k1) (1L/V)0.5 (24-3)
The sliding square contact area also has a conducting semi-innite
body behind it, and its temperature rises due to the heat generated
during contact. The average temperature of the square sliding contact
area designated as 2 is (see Reference [7]):
Taverage for sliding square contact area = 0.946 L Q/k2

(24-4)

Eq. (24-3) assumes that all the heat generated during the contact
goes to the semi-innite medium under the sliding contact area, but
in reality this is not the case. A portion, mQ, of the heat generated
during the sliding contact goes into the semi-innite body and the
rest, (1-m)Q, goes into the square sliding contact area. Equating the
average temperatures in Eqs. (24-3) and (24-4), the proportionality
constant m can be determined:
m = k1 (LV)0.5/[1.125 k2 10.5 + k1 (LV)0.5]

(24-5)

The heat ow proportionality constant, Eq. (24-5), is inserted


into Eq. (24-3) to determine the sliding contact area average
temperature:

222

Average Temperature Rise In Sliding Surfaces In Contact

Tsliding contact area average temperature


= (1.064 Q L 10.5)/[1.125 k2 10.5 + k1 (LV)0.5]

(24-6)

The heat generated during the contact comes from the mechanical
energy that is dissipated during the contact:
Q = fd W g V/(4 L2 J)

(24-7)

where fd is the dynamic coefcient of friction during the contact,


W is the load at the contact area in kg, g is the gravitational constant,
9.8 m/s2, V is the relative velocity of the sliding contact area over
the semi-innite medium in m/s, L is the half width of square sliding
contact area in m, and J = 1 kg-m2/s3-W, which is a proportionality
constant that equates mechanical work to heat from the rst law
of thermodynamics. Combining Eqs. (24-6) and (24-7) gives the
nal form of the sliding contact area average temperature for
(LV)/(2 1) > 5:
Taverage = (0.266 fd W g V 10.5)/{L J [1.125 k2 10.5 + k1 (LV)0.5]}

(24-8)

where
1 = thermal diffusivity of semi-innite medium in m2/s,
k1 = thermal conductivity of semi-innite medium in W/m-C,
k2 = thermal conductivity of the square sliding contact area in W/m-C.
The following nominal values for the independent variables are
assumed for the present sensitivity analysis. Also both the
semi-innite medium and the square sliding contact area are assumed
to be stainless steel.
fd = 0.5,
W = 1000 kg,
V = 10 m/s,
1 = 4 106 m2/s,
k1 = 15 W/m-C,
k2 = 15 W/m-C,
L = 0.01 m

223

Everyday Heat Transfer Problems

Sliding Contact Area


Average Temperature, C

With these nominal values, sliding contact area average


temperature rise is 546C. Average temperature rises linearly with
load at the sliding contact area and with the dynamic coefcient of
friction. Average temperature sensitivity to load at the sliding contact
area is 0.546 C/kg, and to the dynamic coefcient of friction is 109.1C
per 0.1 change in fd.
Sliding contact area average temperature rise versus relative
velocity of two bodies is shown in Figure 24-1. Average temperature
increases with increasing relative velocity as a V0.5 function. Sliding
contact area average temperature rise sensitivity to relative velocity
of two bodies is shown in Figure 24-2. Average temperature sensitivity
decreases as V-0.5 as relative velocity increases.
Sliding contact area average temperature versus thermal diffusivity
of semi-innite body under sliding contact area is shown in Figure
24-3; average temperature increases as 10.5. Average temperature
sensitivity to 1 is given in Figure 24-4, and it behaves as 10.5.
Sliding contact area average temperature versus thermal
conductivity of semi-innite body under the sliding contact area is
shown in Figure 24-5; average temperature decreases as k11.

1000
800
600
400
200
0

10

15

20

25

30

Relative Velocity Of Bodies, m/s

Figure 24-1 Sliding contact area average temperature versus relative

velocity of bodies

224

Average Temperature Rise In Sliding Surfaces In Contact

Taverage / V, C-s/m

400
300
200
100
0

10

15

20

25

30

Relative Velocity Of Bodies, m/s

Figure 24-2 Sliding contact area average temperature sensitivity to relative

Sliding Contact Area


Average Temperature, C

velocity of bodies

2000
1600
1200
800
400
0
0.E+00

1.E05

2.E05

3.E05

4.E05

5.E05

Thermal Diffusivity, 1, Of Semi-Infinite Body Under Sliding


Contact Area, m2/s

Figure 24-3 Sliding contact area average temperature versus thermal

diffusivity, 1, of semi-innite body under sliding contact area

225

Everyday Heat Transfer Problems

Taverage / 1, C-s/m2

1.2E+08
1.0E+08
8.0E+07
6.0E+07
4.0E+07
2.0E+07
0.E+00

1.E05

2.E05

3.E05

4.E05

5.E05

Thermal Diffusivity, 1, Of Semi-Infinite Body Under Sliding


Contact Area, m2/s

Figure 24-4 Sliding contact area average temperature sensitivity to thermal

Sliding Contact Area


Average Temperature, C

diffusivity, 1, of semi-innite body under sliding contact area

1600
1200
800
400
0

50

100

150

200

250

300

Thermal Conductivity, k1, W/m-C

Figure 24-5 Slider contact area average temperature versus thermal

conductivity of semi-innite body under sliding contact area

226

Average Temperature Rise In Sliding Surfaces In Contact

Average temperature sensitivity to thermal conductivity of semi-innite


body under the sliding contact area is given in Figure 24-6, and it
behaves as k12.
Sliding contact area average temperature versus thermal
conductivity, k2, of sliding square contacting body is shown in
Figure 24-7; average temperature decreases slightly, and almost
linearly, with increasing thermal conductivity of the sliding square
contacting body. Average temperature sensitivity to k2 is given in
Figure 24-8, and it can be considered a constant at 0.25 C2-m/W.
Under the given nominal conditions, slider contact area average
temperature is most sensitive to the size of the square contact area.
Sliding contact area average temperature versus half width, L, of
square contact area is depicted in Figure 24-9. Average temperature
decreases as L1.5 as side of contact area increases. Average
temperature sensitivity to half width of square contact area is given in
Figure 24-10, and it behaves as L2.5.
When the nominal values of the independent variables given above
are varied 10%, the results shown in Table 24-1 are obtained. Sliding
contact area average temperature sensitivities to a 10% change in

Taverage / k1, C2-m/W

10
30
70
110
150

50

100
150
200
250
Thermal Conductivity, k1, W/m-C

300

Figure 24-6 Sliding contact area average temperature sensitivity to thermal

conductivity of semi-innite body under sliding contact area

227

Sliding Contact Area


Average Temperature, C

Everyday Heat Transfer Problems

550

540

530

520

100
0
20
40
60
80
Thermal Conductivity Of Sliding Square Contacting body, k2,
W/m-C

Figure 24-7 Sliding contact area average temperature versus thermal

Taverage / k2, C2-m/W

conductivity of sliding square contacting body

0.2

0.25

0.3

20

40

60

80

100

Thermal Conductivity Of Sliding Square Contacting Body,


k2, W/m-C

Figure 24-8 Sliding contact area average temperature sensitivity to thermal

conductivity of sliding square contacting body

228

Average Temperature Rise In Sliding Surfaces In Contact

Sliding Contact Area


Average Temperature, C

3000
2500
2000
1500
1000
500
0

0.02

0.04

0.06

0.08

0.1

Half Width, L, Of Square Contact Area, m

Figure 24-9 Sliding contact area average temperature versus half width, L,

of square contact area

Taverage / L, C/m

0
50000
100000
150000
200000
250000
300000

0.02

0.04

0.06

0.08

0.1

Half Width, L, Of Square Contact Area, m

Figure 24-10 Sliding contact area average temperature sensitivity to half

width, L, of square contact area

229

Everyday Heat Transfer Problems

Table 24-1 Effects of a 10% change in nominal values of


independent variables to sliding contact area
average temperature

Nominal
Value

Change In
Sliding Contact
Area Average
Temperature
For A 10%
Decrease
In Nominal
Value

Change In
Sliding Contact
Area Average
Temperature
For A 10%
Increase
In Nominal
Value

0.01 m

+17.08%

13.30%

15 W/m-C

+11.02%

9.03%

10%

+10%

W, load at the
sliding contact
area

1000 kg

10%

+10%

V, relative velocity
of sliding square
contact area and
the semi-innite
body

10 m/s

5.17%

+4.92%

1, thermal
diffusivity of
semi-innite body

4 106 m2/s

5.10%

+4.84%

15 W/m-C

+0.07%

0.07%

Variable
L, half width of
sliding square
contact area
k1, thermal
conductivity of
semi-innite body
fd, dynamic
coefcient of
friction0.5

k2, thermal
conductivity of
sliding square
contact area

230

Average Temperature Rise In Sliding Surfaces In Contact

the governing independent variables are given in descending order of


importance, and they are applicable only around the nominal values
assumed for this analysis.
Sliding contact area average temperature is most sensitive to the
size of the contact area. Thermal conductivity of semi-innite body
under the sliding contact area, dynamic coefcient of friction and
contact area load are next in the order of sensitivity. However, these
sensitivities have the same order of magnitude as sensitivity to the
size of the contact area. Relative velocity of two bodies and thermal
diffusivity of semi-innite body under the sliding contact area come
next in the average temperature order of sensitivity. Sliding contact
area average temperature sensitivities to these two independent
variables are signicant, and they have half the magnitude as the
sensitivities to previous tiers independent variables.
Sliding contact area average temperature has its lowest sensitivity
to thermal conductivity of the sliding square contacting area. This
sensitivity is two orders of magnitude lower than the sensitivity to the
size of the contact area.

231

REFERENCES

[1]
[2]
[3]
[4]
[5]
[6]
[7]

[8]
[9]

Carslaw, H. S. and Jaeger, J. C., Conduction Of Heat In Solids, Second


Edition, Clarendon Press, Oxford, 1986.
Eckert, E. R. G. and Drake, Jr., R. M., Heat And Mass Transfer,
McGrawHill Book Company, Inc., New York, 1959.
Fraden J., Handbook Of Modern Sensors, Second Edition, SpringerVerlag New York, Inc., New York, 1996.
Guenin, B. Conduction Heat Transfer In A Printed Circuit Board,
Electronics Cooling Magazine, Volume 4, Number 2, 1998.
Holman, J. P., Heat Transfer, Second Edition, McGraw-Hill Book
Company, New York, 1968.
Incropera, F. P. and DeWitt, D. P., Fundamentals Of Heat And Mass
Transfer, Fourth Edition, John Wiley & Sons, New York, 1996.
Jaeger, J. C., Moving Sources Of Heat And The Temperature At Sliding
Contacts, Journal And Proceedings Of The Royal Society Of New South
Wales, Volume 79, pp. 203224, 1942.
Jones, J. B. and Hawkins, G. A., Engineering Thermodynamics, John
Wiley & Sons, Inc., New York, 1960.
Kline, S. J. and McClintock, F. A., Describing Uncertainties in
Single-Sample Experiments, Mechanical Engineering, Volume 75,
pp. 38, 1953.

233

Everyday Heat Transfer Problems


[10]
[11]
[12]

[13]

[14]
[15]
[16]
[17]

234

Kreith, F., Principles Of Heat Transfer, Second Edition, International


Textbook Company, Scranton, Pennsylvania, 1965.
Kutz, M. Heat Transfer Calculations, Mc-Graw Hill, New York, 2005.
Lee, W. J., Kim, Y., and Case, E. D., The Effect Of Quenching Media
On The Heat Transfer Coefcient Of Polycrystalline Alumina, Journal
Of Materials Science, Volume 28, pp. 20792083, 1993.
Lin, S. and Jiang, Z., An Improved Quasi-Steady Analysis For Solving
Freezing Problems In A Plate, A Cylinder And A Sphere, ASME
Journal Of Heat Transfer, Volume 125, pp. 11231128, 2003.
London, A. L. and Seban, R. A., Rate Of Ice Formation, Transactions
Of The ASME, Volume 65, pp. 771778, 1943.
Roshenow, W. M. and Choi, H. Y., Heat, Mass, And Momentum
Transfer, Prentice-Hall, Inc., Englewood Cliffs, New Jersey, 1961.
Schneider, P. J., Conduction Heat Transfer, Addison-Wesley
Publishing Company, Inc., Reading, Massachusetts, 1957.
Timoshenko, S and Goodier, J. N., Theory Of Elasticity, Second
Edition, McGraw-Hill Book Company, Inc., New York, 1951.

You might also like