You are on page 1of 620

■ ’ ■ v * ' /: - - > ' ’.

wv 4-

FOR ENGINEERS

FRANCIS J.
DOOLITTLE HALE
Complimentary Copy For Bi

V Ur p| j 1 StiURGH

John Wiley & Sons, Inc

8HWMI
Digitized by the Internet Archive
in 2018 with funding from
Kahle/Austin Foundation

https://archive.org/details/thermodynamicsfoOOOOdool
THERMODYNAMICS
FOR ENGINEERS
«-
THERMODYNAMICS
FOR ENGINEERS
Jesse S. Doolittle
Francis J. Hale

Department of Mechanical
and Aerospace Engineering
North Carolina State University

JOHN WILEY & SONS

New York • Chichester • Brisbane • Toronto • Singapore


Copyright © 1983, by John Wiley & Sons, Inc.

All rights reserved. Published simultaneously in Canada.

Reproduction or translation of any part of


this work beyond that permitted by Sections
107 and 108 of the 1976 United States Copyright
Act without the permission of the copyright
owner is unlawful. Requests for permission
or further information should be addressed to
the Permissions Department, John Wiley & Sons.

Library of Congress Cataloging in Publication Data :


Doolittle, Jesse Seymour, 1903-
Thermodynamics for engineers.
Includes index.
1. Thermodynamics. I. Hale, Francis J. II. Title.
TJ265.D683 621.402T 82-7052
ISBN 0-471-05805-X AACR2

Printed in the United States of America

10 987654321
PREFACE

The scope of engineering thermodynamics is very broad. No one textbook


can meet the needs of all undergraduate students in this field. As a general
approach, we believe that all engineering students should gain an understand¬
ing of energy and energy transformations as formulated in the first and second
laws of thermodynamics.
We present the first and second laws first from the macroscopic or classical
viewpoint. This approach enables the students to relate these concepts to
ideas with which they are familiar. In the discussion of these laws the
necessary material is introduced relative to the properties and the behavior of
matter. Particular consideration is given to gases and vapors, whenever matter is
involved in energy transfers and transformations.
We believe that this approach will enable engineering students in one
semester to gain a basic understanding of the subject and to recognize its
importance. A very meaningful one-semester course can be devised by
studying Chapters 1 through 6 and Chapters 9 and 10. Time permitting,
Chapters 13 and 14 may be added.
For those students taking a two-semester course, it may be desirable to
include Chapters 7 and 8 in the first semester work. For second semester
students, applications of basic thermodynamic principles are directed to
engineering areas such as chemical thermodynamics, thermodynamics of fluid
flow, gas and vapor cycles, refrigerating cycles, and air conditioning. Because
we believe that kinetic theory, microscopic thermodynamics, and statistical
thermodynamics are essential for students to gain a deeper understanding of
the subject, we present the fundamental aspects of these topics.
In addition to conventional power-producing devices, an ever-growing
emphasis is being placed on direct energy conversion, particularly photovol¬
taic cells, magnetohydrodynamics, and fuel cells. The thermodynamics of
direct-energy conversion is discussed on an elementary basis for the benefit
of those students who do not wish to take elective courses in the subject, or
who do not plan to go on to graduate school.
The last Chapter deals with the elementary concepts of heat transfer. It is
made available here for the benefit of those students who do not intend to
take a separate course in the subject. Furthermore, since so much of ther¬
modynamics is concerned with heat transfers to and from systems, it is
helpful to understand the basic principles of this process.
Many problems are given at the end of each chapter. In general, they are
designed to teach students to use fundamental concepts in solving engineering
problems as well as to illustrate fundamental principles. Many of the prob-

v
vi Preface

lems have been devised to show the effects of changes in various parameters
on the performance of engineering devices. For example, the answers to
various problems in the vapor cycles chapter show the precise effect of initial
conditions, exhaust pressure, reheating, and feedwater heating on the
efficiency of steam cycles. Because many instructors believe that students
should not be given answers to problems ahead of time, no answers are given
in the text. However, if some instructors wish to give answers to their
students, they can find answers in the solution manual that will be available to
them.
Recognizing the trend toward the use of the International System of Units
(SI), we use these units throughout the book. However, since the English
system of units is still being used extensively in this country, some work using
English units is given in the first part of the text. Although changes are being
made, the dimensions of much industrial equipment are still given in inches or
feet. For example, common pipes are manufactured to specified dimensions
(both the nominal and actual diameters) in inches. Hence, when dealing with
pipes, we use the sizes as given in inches.
We express our appreciation to Linda Jackson for her excellent help in the
preparation of the manuscript.

Jesse S. Doolittle
Francis J. Hale
CONTENTS

SYMBOLS AND ABBREVIATIONS xv

Chapter 1 Introduction 1

1-1 The Nature of Thermodynamics 1


1-2 Thermodynamic Systems 7
1-3 Thermodynamic Properties 9
1-4 Units 11
1-5 Mass, Force, and Weight 12
1-6 Pressure 14
1-7 Temperature 18
1-8 Other Thermodynamic Properties 26
1-9 Processes 27
1- 10 Pressure, Temperature, and Molecular Kinetic Energy 27
Problems 29

Chapter 2 The First Law 31

2- 1 Introduction 31
2-2 Basis of the First Law 31
2-3 Energies Involved in the First Law 32
2-4 Evaluation of Potential and Kinetic Energies 34
2-5 Evaluation of Transferred Energies 35
2-6 The First Law Applied to Systems 42
2-7 Power 43
2-8 Enthalpy 45
2-9 First-Law Applications 46
Problems 53

Chapter 3 Ideal and Actual Gases 57

3-1 Introduction 57
3-2 Ideal Versus Actual Gases 61

■ ■

VII
viii Contents

3-3 The Mole 65


3-4 Equations of State for Actual Gases 68
3-5 Van der Waals Equation of State 69
3-6 Accuracy of the van der Waals Equation 71
3-7 Other Equations of State 76
3-8 Virial Equations of State 77
3- 9 Corresponding States and Reduced Properties 78
3- 10 Summary 84
Problems 85

Chapter 4 Changes in State of Gases 87

4- 1 Equilibrium and Reversibility 87


4-2 Polytropic Processes 92
4-3 Evaluation of Work for Polytropic Processes 95
4-4 Internal Energy Changes 98
4-5 Enthalpy Changes 100
4-6 Relationship Between cp and cv 100
4-7 The p-v Relationship for a Reversible Adiabatic Process 101
4-8 Specific Heats for Polytropic Processes 102
4-9 Variable Specific Heat 104
4- 10 Internal Energy and Enthalpy Changes of Gases 109
Problems 112

Chapter 5 The Second Law and Entropy 116

5- 1 Introduction 116
5-2 The Carnot Cycle 118
5-3 The Reversed Carnot Cycle 122
5-4 The Clausius Inequality 125
5-5 Entropy 127
5-6 Entropy and Irreversibility 130
5-7 Uses for Entropy 132
5- 8 Various Forms of the Second Law 134
Problems 135

Chapter 6 Some Consequences of the Second Law 139

6- 1 Temperature and the Second Law 139


6-2 Available and Unavailable Energy 141
Contents ix

6-3 Availability of Energy Entering a System 141


6-4 Loss in Available Energy During Heat Transfer 143
6-5 Change in Available Energy of Systems 145
6-6 Availability of a Closed System 149
6- 7 Entropy and Unavailable Energy 151
Problems 153

Chapter 7 Probability and the Nature of Entropy 156

7- 1 Introduction 156
7-2 The Microscopic Approach 156
7-3 Probability 157
7-4 Probability and Entropy 161
7- 5 Entropy and the Third Law 163
Problems 165

Chapter 8 Genera! Equations of Thermodynamics 166

8- 1 Introduction 166
8-2 Internal Energy and Enthalpy 168
8-3 Entropy Changes 171
8-4 Specific Heat Relations 172
8-5 Clapeyron Equation 174
8-6 Joule-Thomson Coefficients 175
8-7 Other General Equations 179
8- 8 Summary 180
Problems 181

Chapter 9 Vapors 183

9-1 Introduction 183


9-2 Vaporization of Liquids 184
9-3 Vapor Tables 188
9-4 Compressed Liquids 189
9-5 Properties of a Wet Vapor 191
9-6 Determination of Vapor Properties 193
9-7 Vapor Charts 194
9-8 Throttling of Vapors and Liquids 196
9-9 Changes in State of Vapors 197
9- 10 Measurements To Determine the State of Vapors 201
x Contents

9- 11 Nonequilibrium States for Two-Phase Mixtures 203


Problems 207

Chapter 10 Nonreactive Gaseous and Vapor


Mixtures 211
10-1 Introduction 211
10-2 Properties of Ideal Gaseous Mixtures 211
10-3 Irreversible Mixing of Gases Within a System 216
10-4 Mixing of Ideal Gases During Flow 219
10-5 Volumetric and Gravimetric Analysis 220
10-6 Gas-Vapor Mixture 221
10-7 Air-Vapor Mixtures 222
10-8 Psychrometric Charts 228
10-9 Simple Heat Transfer 229
10- 10 Humidification and Dehumidification 231
10-11 Cooling by Evaporation 234
Problems 238

Chapter 11 Elements of Chemical Thermodynamics 241

11-1 Introduction 241


11-2 Combustion 241
11-3 Combustion of Actual Fuels 245
11-4 Uses of Analyses of Products 250
11-5 Combustion in General 253
11-6 Heat of Reaction 253
11-7 Heating Values 254
11-8 Heat of Formation 260
11- 9 Chemical Energy 261
11-10 Combustion Temperatures 263
11-11 Chemical Equilibrium 264
11-12 Equilibrium in a Reactive Mixture 265
11-13 Fugacity and Activity Coefficient 271
Problems 274

Chapter 12 Thermodynamics of Fluid Flow 277

12- 1 Introduction 277


12-2 Laminar and Turbulent Flows 278
Contents xi

12-3 Continuity of Flow 281


12-4 Energy Equation and Pressure Changes 281
12-5 Stagnation Properties 284
12-6 Acoustic Velocity and Mach Number 286
12-7 Adiabatic Flow with Constant-Area Passage 288
12-8 Adiabatic Flow with Varying Area 291
12- 9 Flow in Ideal Nozzles 296
12-10 Nozzle Efficiency 299
12-11 Shock Waves in Nozzles 301
12-12 Diabatic Flow 305
12- 13 Summary 309
Problems 311

Chapter 13 Gas Cycles and Applications 315

13- 1 Introduction 315


13-2 Criteria for Cycles 316
13-3 The Otto Cycle 317
13-4 The Diesel Cycle 321
13-5 The Brayton or Joule Cycle 324
13-6 Stirling and Ericsson (Regenerative) Cycles 327
13-7 Internal Combustion Engines 331
13-8 Internal Combustion Engine Performance 334
13-9 Gas Compressors 337
13- 10 Gas Turbines 343
13-11 Actual Gas Turbines 345
13- 12 Jet Propulsion 348
Problems 352

Chapter 14 Vapor Cycles and Application 356

14- 1 Introduction 356


14-2 Carnot Vapor Cycle 356
14-3 Rankine Vapor Cycle 358
14-4 Reheating Cycle 365
14-5 Regenerative Feedwater Heating 367
14-6 Additional Vapor Cycles 371
14-7 Refrigeration 372
xii Contents

14-8 Vapor Compression Refrigeration 376


14- 9 Other Refrigerating Systems 379
Problems 384

Chapter 15 Kinetic Theory of Gases 388

15- 1 Introduction 388


15-2 Pressure, Temperature, and Kinetic Theory 388
15-3 Temperature and the Root Mean Square Velocity 392
15-4 Maxwell Speed Distribution 394
15-5 Evaluation of the Function a and b 399
15- 6 Validity of the Maxwell Speed Distribution 401
Problems 403

Chapter 16 Elementary Statistics and Quantum


Mechanics 404

16-1 Introduction 404


16-2 Energy Levels 405
16-3 Maxwell-Boltzmann Statistics 406
16-4 Bose-Einstein Statistics 409
16-5 Fermi-Dirac Statistics 411
16-6 Examination of the Three Models 412
16-7 Evaluation of the Constant (3 413
16-8 Applications to Internal Energy, Entropy, and Specific Heats 414
16-9 Energy of Translation and Specific Heat 416
16- 10 Rotational Energy and Specific Heat 418
16-11 Atomic Vibratory Energy and Specific Heat 423
16-12 Hydrogen at Low Temperatures 427
16-13 Gases at Extremely High Temperatures 428
16-14 Solids and Liquids 428
16- 15 Application to Entropy Determinations 430
Problems 433

Chapter 17 Introduction to Irreversible


Thermodynamics 434

17- 1 Introduction 434


17-2 Entropy Production 434
Contents xiii

17-3 The Onsager Relations 437


17- 4 Thermodynamics of Thermoelectricity 439
Problems 447

Chapter 18 Direct Energy Conversion 448

18- 1 Introduction 448


18-2 Thermoelectric Generation 449
18-3 Thermionic Generation 454
18-4 Magnetohydrodynamics (MHD) Generation 460
18-5 Fuel Cells 466
18- 6 Photovoltaic Cells 471
Problems 475

Chapter 19 Elements of Heat Transfer 477

19-1 Introduction 477


19-2 Conduction 478
19-3 Conduction Through a Composite Wall 481
19-4 Conduction Through Thick-Wall Cylinders 482
19-5 Radiation 485
19-6 Relationship Between Absorptivity and Emissivity 489
19-7 Net Radiant Energy Exchange 490
19-8 Convection 493
19-9 Free Convection 495
19- 10 Forced Convection 497
19-11 Film Coefficients for Phase Changes 500
19-12 Overall Coefficients of Heat Transfer 502
19-13 Mean Temperature Differences for Parallel Flow and
Counterflow Heat Exchangers 506
19-14 Multipass and Crossflow Heat Exchangers 511
19-15 Applications 513
Problems 515
Appendixes 521
Charts of Properties 571
Some Selected and Fundamental References 581
Index 583
SYMBOLS AND
ABBREVIATIONS

SYMBOL MEANING

A Area
A Ampere
A Helmholtz function
a Specific Helmholtz function
a Absorptivity
a Acceleration
a Acoustic velocity
Btu British thermal unit
B Magnetic field strength
C Constant (general)
c Specific heat
cv Specific heat for a constant-volume process
Cp Specific heat for a constant-pressure process
Cx Specific heat for a polytropic process
cm Centimeter
C.O.P. Coefficient of performance
(C.O.P.)r Coefficient of performance for refrigeration
(C.O.P .)h p Coefficient of performance for a heat pump
Chem. E. Chemical energy
D Diameter
E Voltage
E Emissive power
eV Electron volt
F Force
Fa Shape factor
Fe Emissivity factor
f Fugacity
f Saturated liquid (as a subscript)
fg Change in properties during phase change (as a subscript)

xv
xvi Symbols and Abbreviations

SYMBOL MEANING

ft Foot
G Mass rate of flow per unit area
G Gibbs function (free energy) = H - TS
g Specific Gibbs function
g Gram
g Acceleration due to gravity
g Saturated vapor (as a subscript)
gi Possible number of quantum states
Gr Grashof number
g-mole gram mole
H Enthalpy
h Specific enthalpy
h Hour
h Planck constant
h Surface or film heat-transfer coefficient
hc Convertive surface or film heat-transfer coefficient
hp horsepower
hr Radiant surface or film heat-transfer coefficient
I Moment of inertia
I Electric current
I Intensity of radiation
J Current density
J Joule
J Quantum number
Js Entropy current
K Kelvin
K Combined thermal conductivity
KP Chemical equilibrium constant
kl kiloliters
K.E. Kinetic energy
k Boltzmann constant (gas constant per molecule)
k Specific heat ratio (=cp/cj
k Thermal conductivity
kg Kilogram
kW Kilowatt
lb Pound
lb/ Pound force
lb-mole Pound mole
Symbols and Abbreviations xvii

SYMBOL MEANING

lbm Pound mass


/ Distance
In Natural logarithm
log Common logarithm
LMTD Log mean temperature difference
M Mach number
m Mass
m Meter
m Mass rate of flow
m Mass per mole
MW Megawatt
N Newton
N Number of moles
N Number of cells
N Cycles per minute
n Number of molecules
n Polytropic process exponent
n0 Number of molecules per mole
n Quantum number
Nu Nusselt number
0 As a superscript refers to stagnation
P Power
P.E. Potential energy
Pa Pascal
Pe Peclet number
Pr Prandtl number
P Pressure
psi Pounds per square inch
psia Pounds per square inch absolute
psig Pounds per square inch gage
o Heat
0
Heat flow rate
Q Heating value
Qcomb Heat of combustion
q Charge per electron
R Rankine
R Specific gas constant
Ro Gas constant per mole
xviii Symbols and Abbreviations

SYMBOL MEANING

R Electric resistance
Re Reynolds number
r Distance between magnetic poles
r Compression ratio
r Radius
rt Combined electrical resistance
r Rotational (as a subscript)
S Entropy
s Specific entropy
s Second
T Absolute temperature
T Torque
t Degrees Celsius or Fahrenheit
t Time
t Translational (as a subscript)
U Internal energy
u Overall coefficient of heat transfer
uk Internal molecular kinetic energy
Up Molecular potential energy
u Specific internal energy
V Volume
V Volts
v- Specific volume
v Vibratory (as a subscript)
V Vapor (as a subscript)
W Work
Wm Mechanical work
W, Electrical work
w Rate of doing work
W' Possible number of macroscopic states
w Watt
w Weight
w Specific weight
X or x Unknown (general)
x Quality, mass fraction of vapor in a two-phase mixture
x Thickness
y Moisture content, mass fraction of liquid in a two-phase
mixture
Symbols and Abbreviations XIX

SYMBOL MEANING

Z Compressibility factor
z Figure of merit
z Quantum partition function
2 Height
p Coefficient of volume expansion
e Emissivity
e Energy level
V Efficiency
Vm Mechanical efficiency
Vn Nozzle efficiency
Vt Thermal efficiency
A Wavelength
v Kinematic viscosity
V Frequency
Viscosity (absolute)
P' Joule-Thomson coefficient
a ohm
p Density
p Reflectivity
y Velocity
y' rms Root mean square velocity
T Transmissivity
T Thomson coefficient, thermoelectricity
a Seebeck coefficient, thermoelectricity
IT Peltier coefficient, thermoelectricity
Work function, thermionic generation
& Possible number of microscopic states
<t> Relative humidity
OJ Specific humidity
cr Symmetry number (molecular)
cr Internal electric conductivity
n Product of
*'
_ 1 _

INTRODUCTION

1-1 THE NATURE OF THERMODYNAMICS


Thermodynamics is one of the basic engineering sciences. At first, before the
nature of heat was fully understood, thermodynamics was concerned pri¬
marily with heat-work transformations. In the 1840s, James Prescott Joule
clearly demonstrated that heat was energy1 and that there was a precise
relationship between heat and work. It then became recognized that energy
had several other forms. Thus, the original scope of thermodynamics was
expanded to include a study of energy in all its forms, the transfer and
transformation of energy, and the results of these transfers and trans¬
formations.
Energy is difficult to define. Frequently, energy is defined as the capacity to
do work. To make such a definition all-inclusive, work must be viewed in its
broadest sense. It should be recognized that all objects tend to undergo
changes. Some forces act on the objects so as to resist the changes. Other
forces tend to cause changes to take place. Where the driving forces exceed
the resisting forces, then changes do take place. Work is required to produce
the changes. Energy must be drawn from some source to provide the work
required.
Work is required to increase the elevation of an object or to increase its
velocity. When a liquid is vaporized, work is required to overcome the
intermolecular forces existing between the molecules of the liquid. When a
tank of air is heated, work is required to increase the velocity of the air
molecules. When a solid is heated, with some of the valence electrons being
set free, work is required to overcome the attractive forces originally acting
on the electrons. When an electric resistance coil is used for heating purposes,
work is required to generate the electrical energy that is used.
It should be noted that in all of these illustrations, energy is associated with
changes in the physical or chemical characteristics of matter. In the broad
sense, since changes in the characteristics of all matter have a tendency to
occur, it may be said that thermodynamic processes are taking place all
around us. As a practical matter, however, engineering thermodynamics is
restricted to engineering processes.

‘The work of Count Rumford in 1798 and Sir Humphry Davy in 1799 contributed to the final
understanding of the nature of heat.

1
2 Introduction

Energy is of particular interest to engineers because it gives them the


capability to take the raw materials from nature and shape them to the use
they desire. In the early days of recorded history, human beings depended
on their own muscular energy and that of beasts of burden to furnish the
energy that they required. As such, their accomplishments were distinctly
limited. Over the centuries up to the dawn of the Industrial Revolution,
changes in life-styles were exceedingly slow. The Industrial Revolution
became possible only when mankind learned to draw energy from nature and
to utilize it. At first, the Industrial Revolution produced relatively slow
changes in life-styles, but once initiated, these changes took place with
ever-growing rapidity. For example, life-styles today bear very little resem¬
blance to those of as recently as 50 years ago. These tremendous changes
would have been utterly impossible without the use of energy drawn from
nature. In reality, if this energy were to be denied to us, our life-styles would
return to those of the Dark Ages.
Until very recently in this country, we have treated our supply of energy as
being nearly inexhaustible. In 1980, we used about 1.75 times as much energy
as we did just 20 years earlier. About 4 percent of the energy used in this
country is hydroelectric. The remaining 96 percent is obtained from non¬
renewable resources: oil, natural gas, coal, and uranium. We are using up
these resources, particularly natural gas and oil, at an alarming rate. To meet
our demands, we are now importing between 40 and 50 percent of all the oil
that is used in this country. This is causing a severe economic problem. It has
been predicted that unless immediate steps are taken to change this energy
situation, life-styles of this country will be very seriously altered. This means
that an in-depth examination must be made of methods of energy procure¬
ment, energy conversions and energy utilizations. By its very nature, ther¬
modynamics is intimately involved in energy conversion and in energy utili¬
zation processes.
In a fuel cell, the energy of the fuel is converted directly to electrical
energy. Otherwise, fuels are burned, producing high temperature products.
The energy of these high temperature products is used directly or indirectly to
accomplish desired effects. A large portion of our fuels is used for heating
purposes such as heating homes and commercial and industrial buildings. This
is known as space heating. Many of the industrial processes take place only at
elevated temperatures. Other processes take place better at elevated tem¬
peratures. In both cases, heat must be supplied. Whether heat is supplied for
space heating purposes or for industrial processes, the first law of ther¬
modynamics is valid. We will discuss the first law later on in some detail. At
this point, however, it may be said that the first law states that although
energy may be transformed or transferred, it cannot be destroyed. There must
be an accounting of all of the energy involved in any given process. For
example, when oil is burned in a household furnace, part of the energy of the
fuel is carried up the chimney by the hot gases. Another part of the energy of
the fuel is lost when combustion is not complete. If the furnace is located
The Nature of Thermodynamics 3

outside of the heated space, some of the energy is lost from the furnace as
heat (see Fig. l-l).2 The remainder of the energy of the fuel may be used to
heat the house. To conserve energy in this case, two factors must be
optimized:

1 The amount of heat required to heat the house must be minimized.


2 The amount of the energy of the fuel that can be utilized to heat the house
must be maximized. Here, thermodynamics is involved.

Approximately one-quarter of the total energy used in this country is used


for transportation purposes, largely for automobiles, trucks, buses, trains, and
airplanes. Approximately another quarter of the total energy supply is used
for electric power production. The first law of thermodynamics is perfectly
valid for both of these work-producing producing processes. The second law
of thermodynamics is also intimately involved in these work-producing pro¬
cesses. The second law will be discussed in much detail later on. However,
two aspects of the second law are very pertinent to these work-production
and work-utilization processes:

1 The second law limits the amount of heat that can be transformed into
work and, for the given conditions, specifies the maximum amount of
work that can be produced from a given quantity of heat.
2 In every actual process some of the ability to do work is lost. Thus the
second law states that energy available for doing work continuously
becomes unavailable for this purpose.

2The term “lost” as used here means that energy of the fuel becomes unavailable for its intended
use; it is not destroyed.
4 Introduction

Consider the automotive engine. By virtue of the second law only a small
portion of the energy of the fuel (perhaps 20 to 25 percent) is delivered to the
pistons. The remainder of the energy of the fuel is:

1 Carried away by the exhaust gases by virtue of their high temperature and
by the products of incomplete combustion, such as the hydrocarbons and
carbon monoxide.
2 Carried away by the jacket water to the radiator where it is dispensed to
the atmosphere.
3 Lost by the engine to the surroundings, largely by radiation and con¬
vection.

A small amount of the energy delivered to the pistons is lost by side-wall


friction and by engine-bearing friction. Of the work delivered by engines,
some of it is lost in the transmission, in the drive shaft bearings and in the
differential. The remainder of the work is delivered to the wheels to propel
the automobile. This work is ultimately dissipated in overcoming the bearing
losses in the car and in the air resistance to car motion. Some of the energy
will be lost in braking the car. To conserve energy, here, too, are two factors
that should be optimized:

1 The amount of energy required to move the car should be minimized.


2 The percentage of the energy of fuel that is available to drive the car
should be maximized. Thermodynamics, particularly the second law, is
intimately involved here.

As another example of the involvement of the first and second laws in


energy transformations, consider the machining of a metal part in a factory.
Assume that the machine is driven by an electric motor. The energy to run the
motor conceivably can be obtained from many various sources, such as a
hydroelectric power plant, a nuclear power plant, a gas turbine power plant, a
magnetohydrodynamic power plant, a thermoelectric generator, a thermionic
generator, or a fuel cell. Because approximately 80 percent of our electrical
energy is produced in fossil fuel steam power plants, it will be assumed here
that the electrical energy required to drive the motor for the machining
operation is produced in a fossil fuel steam power plant.
Figure 1-2 illustrates in general what happens to the energy originally
possessed by the fuel in the process of delivering the work required for the
machining operation. The fuel is burned in the furnace of the steam generator
of the power plant. Some of the energy of the fuel, perhaps 10 percent, is lost,
mainly being carried away by the gases passing up the chimney. The remain¬
der of the energy is added to water, producing steam that is piped to the
steam turbine. There is a small heat loss from the pipe and from the turbine
itself. The second law states that only a portion of the heat added to produce

Thermodynamics plays a very important role in all of power production processes.


c/i
E
<D
C/D

C/D

<u
£
o
D.

C
C/D
c
o
03
c7>
C/D

T3
>.
DC
i—
<u
C
W
<N

W
P4
D
O
HH

c
_co
Q.
i—
CD

o
g
Q_

5
6 Introduction

steam can, even theoretically, be delivered as work by the turbine, with the
actual work being perhaps 75 percent of the theoretical. The remainder of the
heat supplied is removed in the condenser as the exhaust steam from the
turbine is condensed and the resulting condensate is pumped back into the
steam generator. The electrical generator converts about 98 percent of the
work delivered by the turbine into electrical energy. This electrical energy is
transmitted through the transmission system. Because of the electrical resis¬
tance of the transmission system, some of the electrical energy input is
dissipated as heat. The remainder of the electrical energy is delivered to the
various places where it is to be used, such as the motor of the machine in the
factory.
The first law may be applied to the overall process. The energy of the fuel
is not destroyed and an accounting can be made of all the ways in which it is
dissipated. More importantly, the second law is very much in evidence. The
second-law limitations on the amount of work that can be produced from a
given quantity of heat are very evident here. The second aspect of the second
law also is very much in evidence as work is transferred from the turbine to
the machine in the factory. In each transfer process, some of the work is
dissipated as heat. As a result, only a portion of the work delivered by the
turbine can be used for the machining operation. In the interest of energy
conservation as well as cost considerations, it is essential that the work output
of the turbine be maximized and that the work dissipation losses be mini¬
mized.
The above three examples of energy conversions and transfers are only a
few of the exceedingly large number of similar problems encountered daily by
engineers. In all energy-involved problems, the energy produces changes in
the physical and/or chemical aspects of matter. Hence, a complete study of
thermodynamics must involve a study of matter in all of its forms: solid,
liquid, and gaseous. This study must be centered on changes in the properties
of matter produced by the application of energy. Although the engineer is
sometimes concerned with thermodynamic problems dealing with liquids and
solids, he or she is more involved with those problems dealing with gases and
vapors. Hence, the major emphasis in this book is on gases and vapors.
In the early days of thermodynamics, observations were made about the
behavior of matter. Laws were devised to express the effects of variations of
some parameters on the behavior of other parameters. These experimentally
based laws are valid provided that they truly represent the experimental
results and that the experimental results are correct. A further restriction is
that these laws must be applied only to the conditions for which they were
derived. For example, Boyle observed that the pressure of a gas varied
inversely with its volume when the temperature was held constant. This law is
perfectly valid under normal conditions of temperature and pressure for gases
that are hard to liquefy such as oxygen, nitrogen, and hydrogen. But this law
breaks down completely for these gases at very low temperatures and shows
Thermodynamic Systems 7

an appreciable error under normal conditions for gases that are easy to
liquefy.
Thermodynamics uses these experimentally based laws, but it must con¬
sider the limitations of these laws as well as their adaptation to conditions
other than those for which they were devised. It is frequently helpful in the
study of these laws to formulate hypotheses to help explain them. Desirable
as this practice may be, it is important to recognize that the science of
thermodynamics is not based in any way on these hypotheses but rather on
the observable laws.
In the preceding discussion, no attempt was made to study the behavior of
the molecules of which matter is composed, but rather attention was directed
to matter as a composite whole. Such an approach is known as a macroscopic
or classical approach. Since most engineering problems involve finite quan¬
tities of matter containing an extremely large number of molecules, the
macroscopic approach is satisfactory for these types of problems. However,
when one is dealing with matter at extremely low pressures, such as those in
the very upper reaches of the atmosphere, molecules of gases are so far apart
that serious errors may be caused by assuming that matter is continuous and
by applying the laws of classical thermodynamics. Also when gases are
heated to very high temperatures, there will be changes in the molecular
behavior that classical thermodynamics cannot explain. In both of these
examples, as well as several others, it is necessary to study the behavior of
the molecules and, from such a study, to formulate laws that govern the
behavior of the matter. Such an approach is known as the microscopic
approach.
However, because even an extremely small volume of matter under normal
conditions contains billions upon billions of molecules, it is not feasible to
consider the behavior of each and every molecule. This problem may be
solved by formulating a molecular model of matter and then predicting from
this model the probability of the behavior of a large number of molecules.
This study involves the use of statistical thermodynamics.

1-2 THERMODYNAMIC SYSTEMS


In the engineering world, objects normally are not isolated from one another.
In most engineering problems many objects enter into a given problem. Some
of these objects, all of these objects, or even additional ones may enter into a
second problem. The nature of a problem and its solution are dependent on
which objects are under consideration. Thus, it is necessary to specify which
objects are under consideration in a particular situation. In thermodynamics
this is done either by placing an imaginary envelope around the objects under
consideration or by using an actual envelope if such exists. The term system
refers to everything lying inside the envelope. The envelope, real or imag¬
inary, is referred to as the boundaries of the system. It is essential that the
8 Introduction

boundaries of the system be specified very carefully. For example, when one
is dealing with a gas in a cylinder where the boundaries are located on the
outside of the cylinder, the system includes both the cylinder and its con¬
tained gas. On the other hand, when the boundaries are placed at the inner
face of the cylinder, the system consists solely of the gas itself.
When the boundaries of a system are such that it cannot exchange matter
with the surroundings, the system is said to be a closed system (see Fig. l-3a).
The system, however, may exchange energy in the form of heat or work with
the surroundings. The boundaries of a closed system may be rigid or may
expand or contract, but the mass of a closed system cannot change. Flence,
the term control mass sometimes is used for this type of system. When the
energy crossing the boundaries of a closed system is zero or substantially so,
the system may be treated as an isolated system (Fig. l-3b).
In most engineering problems, matter, generally a fluid, crosses the bound¬
aries of a system in one or more places. Such a system is known as an open
system (see Fig. l-3c). The boundaries of an open system are so placed that
their location does not change with time. Thus, the boundaries enclose a fixed
volume, commonly known as the control volume.
Sometimes a system may be a closed system at one moment and an open
one the next. For example, consider the cylinder of an internal combustion
engine with the boundaries at the inner walls. With the valves closed (Fig.
l-4a), the system is a closed one. However, with either or both of the valves
open (Fig. l-4b), the system becomes an open system. Thus care must be
taken not only to specify the location of the boundaries of the system but also
to specify what is happening at the boundaries.
Frequently the total system to be considered may be large and complicated.
The system may be broken down into component parts and an analysis of the
component parts made. Then the performance of the entire system can be
determined by the summation of the performance of the individual com¬
ponent systems. For example, consider the liquid-vapor part of a steam power
plant as an entire system. This system, which is closed, contains the steam
generator (which includes the superheater, the reheater, and the economizer),
the steam turbine, the steam condenser, the feed-water pumps, and the

(a) (b) (c)

FIGURE 1-3 Types of systems, (a) Closed system, (b) Isolated system, (c) Open system.
Thermodynamic Properties 9

Intake and exhaust valves


X

From intake
manifold

Spark plug—5—

Water jacket

(b)
FIGURE 1-4 Internal combustion engine, (a) Valves closed. (b) Valves open.

feed-water heaters. Any or all of these units may be considered separately by


throwing a boundary around them. Since a fluid enters and leaves each of
these smaller systems, each one is an open system and must be analyzed as
such.

1-3 THERMODYNAMIC PROPERTIES


The distinguishing characteristics of a system are called its properties. Those
characteristics that are associated with thermodynamics are known as ther¬
modynamic properties. Some elementary thermodynamic properties are
pressure, temperature, specific volume, and density. It will be shown later on
that the energy possessed by a substance is a thermodynamic property. Other
thermodynamic properties, generally more complex, will be discussed at some
length in later chapters. These properties include enthalpy, entropy, and free
energy.
Properties may be classified as independent properties and dependent
properties. An independent property may be varied at will. A dependent
property is controlled by one or more independent properties. Many proper¬
ties are independent in some cases and dependent in others. Consider a gas
contained in a vertical cylinder equipped with a movable piston upon which
weights are placed. The pressure of the gas in the cylinder will vary as the
10 Introduction

weights on the piston are changed. However, the pressure is controlled


entirely by an external agency and is independent of all other properties of
the gas. Hence, in this example, pressure is an independent property of the
gas. On the other hand, consider a tank containing a gas under a certain
pressure. As heat is added to the gas, its temperature increases. In this case,
temperature is an independent property. This change in temperature produces
a corresponding increase in the pressure. Since the pressure in this instance is
wholly dependent on the temperature, pressure is a dependent property of the
gas.
Some properties depend wholly on the extent or size of the system. These
properties are known as extensive properties. Such properties are the mass of
the system, its total volume, and its total energy. Other properties are
independent of the size of the system but are dependent on the condition at
the point of the property measurement. These properties are said to be
intensive properties. Some intensive properties are pressure and temperature.
Other properties, such as specific volume and density, may be treated as
intensive properties, since they too are independent of the size of the system.
These properties will be discussed in detail later on.
A substance may exist in any one of three phases: namely, the gaseous, the
liquid, or the solid. In general, each of these phases has individual charac¬
teristics that distinguish it from the others. There are instances, which will be
discussed later, in which it is difficult or impossible to recognize the line of
demarcation between two phases. The state of a substance at any instant is its
condition of existence at that instant. The state is described by specifying
certain properties of the substance.
It will be shown in Chapter 3 that there is a very definite relationship
between the pressure, volume, and temperature of any simple substance.
When these properties are independent, the specification of two of these
properties automatically fixes the third property. In Chapter 8, we will discuss
the mathematical relationships between pressure, volume, temperature, and
other properties, such as enthalpy, entropy, specific heat, and the Helmholtz
function. A study of these equations will show that the specification of two
independent properties fixes the state of a simple substance and, hence, all
other properties.
A system may be composed of a single substance existing in a single phase.
Such a system is known as a homogeneous one. The properties of a homo¬
geneous system may be uniform throughout or may vary gradually from one
part to another with no sharp variation of properties at any place in the
system. For a multiphase system, there will be discontinuities in its properties
as we move throughout the system. Consider an ice-water mixture. In moving
from the water to the ice, there is a sharp change in such properties as density
and energy content. The multiphase system is designated as a heterogeneous
system.
Another type of heterogeneous system consists of two or more substances
Units 11

not completely mixed. There may be large or small variations in pressures,


temperatures, and velocities throughout the system. Specifying some of the
intensive properties for a heterogeneous system may be difficult in some
instances, almost impossible in others, and practically meaningless in still
others. However, it is possible to specify some mean intensive properties
for a heterogeneous system. For instance, the mean density of the system
equals the total mass of the system divided by the total volume of the
system.
Because of the different types of systems that exist and because of the
different kinds of properties that may be involved in a given problem, it is
essential not only to fix the boundaries of the system but also to describe the
system in detail before any analysis of the problem can be made.

1-4 UNITS
Before investigating the various properties of substances, it is desirable to
consider the units used in evaluating these properties. Substantially all coun¬
tries, with the exception of the United States, now use the SI system
(Systeme International). Here in the United States we are now moving from
English units to the SI system. Hence, the SI system will be used primarily in
this book. However, some examples and some problems will employ English
units to prepare students to work in the industrial world where complete
conversion to the SI system will require several years.
The SI system consists of six base units and two supplementary units. The
base units are:

Quantity Unit SI Symbol


length meter m
mass kilogram kg
time second s
electric current ampere A
thermodynamic temperature Kelvin K
luminous intensity candela cd

The supplementary units are:

plane angle radian rad


solid angle steradian sr

As various properties are discussed, various derived units, some named and
some unnamed, will be introduced.
To avoid the use of very large or very small numbers, prefixes have been
12 Introduction

established to indicate orders of magnitude. They are:

Prefix Symbol Multiplication


terra T 1012
giga G 109
mega M 106
kilo k 103
hecto h 102
deka da 101
deci d 10 1
centi c 10“2
milli m 1(T3
micro F 10"6
nano n 1(T9
pico P 1012
femto f 1(T15
atto a 10 18

1-5 MASS, FORCE, AND WEIGHT


Many properties are precisely interrelated. The specification of the units of
certain properties determines the units of other properties. For example;
mass, force, and acceleration are interrelated by Newton’s second law of
motion,
F = ma (1-1)
where F = force acting on the mass in direction of motion
m = mass
a = acceleration of the mass

The mass of an object is the quantity of matter that it possesses. The


standard mass, the kilogram, is a block of a platinum-iridium alloy maintained
in Sevres, France. The mass of other objects may be compared most readily
with that of the standard by use of a balance (see Fig. 1-5). If the balance
beam is horizontal, the mass of the secondary object equals that of the
standard.

Standard
mass
Unknown
mass

>■

Knife edge

FIGURE 1-5 Mass determination.


Mass, Force, and Weight 13

Acceleration is defined as the change in velocity per unit time. Using SI


units, the dimensions of velocity are meters per second (m/s). Hence, the
dimensions of acceleration are meters per second per second (m/s2).
Force is defined as the quantity required to produce acceleration of a given
mass. A unit force will accelerate a unit mass of one kilogram one meter per
second per second. Hence, the units of force are kg-m/s2. This unit force has
been designated as the newton (N). Another unit of force that is still used
extensively is the dyne. One dyne is that force which will give a mass of one
gram an acceleration of one centimeter per second per second. The dimen¬
sions of the dyne are g-cm/s2. There are 105 dynes per newton.
Weight is defined as the gravitational force existing between an object and
the earth. Weight may be expressed in either newtons or dynes. Since the
force of gravity varies with the distance from the center of the earth, the
weight of a given object depends on the location of the object. However, the
variation in the force of gravity over the surface of the earth is small, about
0.5 percent as a maximum. Hence, this variation is frequently neglected in
most engineering problems. However, when objects are located above the
surface of the earth, it is essential to recognize that their weights may differ
significantly from those they have when on the earth’s surface. Because
weight is a force, Eq. 1-1 can be written as

W = mg 0-2)

where W is weight
g is the acceleration of gravity

The standard acceleration of gravity (at sea level, 45 degrees N latitude) is


9.80665 m/s2.

Example 1-1. Determine the force in newtons required to produce an ac¬


celeration of 20 cm/s2 of a 40-g mass.

Solution. Using Eq. 1-1,

40 w 20
F = ma
1000 X 100
= 0.008 N Answer

Example 1-2. Determine the mass of an object that has a weight of 180 N
where the gravity acceleration is 9.802 m/s2.

Solution. Using Eq. 1-2,

W= 180
18.36 kg Answer
g 9.802

In dealing with mass, force, and weight in English units, confusion arises
because the unit of force is taken as the pound force (lb/) and at the same
14 Introduction

time the unit of mass is taken as the pound mass (lbm).4 The pound mass is
defined in terms of the kilogram mass. The pound mass equals 0.45359237 kg.
The pound force is defined as the force of gravity acting on a pound mass at
standard sea level conditions. But, at this location, the acceleration of gravity
is 32.174 feet per second per second (ft/s2). Thus the pound force produces an
acceleration of the pound mass of 32.174 ft/s2. Under these conditions it is
necessary to introduce a dimensional constant gc. This constant, not to be
confused with g, the acceleration of gravity, has a value of 32.174 lbm ft/lb/ s2.
Equation 1-1 may now be written as

ma
F = (1-la)
gc

Inserting the unit into the equation,


lbm x ft/s'
lb,=
f 32.174 lbm ft/lbf s2

In similar manner gc may be introduced into Eq. 1-2.


lbm x ft/s2
W = ^- or
Or lb/ =
I Of = ; d-3)
gc 32.174 lbm ft/lb, s2

Example 1-3. An object has a mass of 1201b. Determine its weight at a


location where the acceleration of gravity is 32.168 ft/s2.

Solution. Since the pound mass and the pound force are both to be used, Eq.
1-3 is to be used. Then

ii/ _ mg _ 120 x 32.168


119.98 lb/ Answer
" gc “ 32.174
The variation of the gravity acceleration, g, with altitude is very small. At an
elevation of 5000 m (16,400 ft), the value of g is only 0.157 percent less than
that at sea level. This means that on the earth’s surface, the values of g and gc
in Eq. 1-3 are substantially equal, numerically. Thus the weight of an object,
expressed in lb/, and the mass of the object, expressed in lbm, may be taken to
be equal numerically for most engineering problems on the surface of the earth.

1-6 PRESSURE
Pressure is defined as the force acting on a unit area. When a force is exerted
on a fluid, this force is transmitted throughout the fluid. If the fluid is
stationary, the pressure within the fluid is uniform throughout the fluid, if we
neglect the force of gravity acting on the fluid. The fluid exerts a pressure on
its containing walls which, in turn, exert the same pressure on the fluid.

^his confusion can be avoided by either taking the unit of force as the poundal or the unit of
mass as the slug. However, since these units are used so infrequently, they will not be discussed
here.
Pressure 15

In the SI system, pressure is expressed in newtons per square meter (N/m2).


This unit of pressure is sometimes called the pascal (Pa). Expressed in
fundamental SI units, the dimensions of the pascal are kg/m-s2. In the English
system, pressures are expressed generally in pounds per square inch (psi).
Thermodynamically speaking, there is only one kind of pressure and that is
absolute pressure. Although there are devices available to measure absolute
pressures, most pressure measuring devices measure pressure differences.
A very common pressure measuring device is the Bourdon-tube type of
pressure gage. Fundamentally, this gage consists of a coiled elliptical tube that
is fixed at one end and free to move at the other end. The pressure to be
measured is transmitted to the inside of the tube. There will be movement of
the free end of the tube when the pressure within the tube differs from that
outside the tube. The movement of the tube is transmitted to the needle of the
gage. Because the pressure on the outside of the tube is atmospheric, the
reading on the gage is the pressure above or below that of the atmosphere. In
general, then, gage pressure is defined as the pressure above or below that of
the atmosphere.
Another type of pressure device is the manometer. The U-tube type of
manometer is illustrated in Figure 1-6. Here the fluid pressure to be measured
is balanced against the weight of a column of a liquid. Liquids commonly used
for this purpose are water, mercury, and special manometer oil. A knowledge
of the specific weight, or the weight per unit volume, of the fluid and the
height of the column that balances the pressure permits the determination of

FIGURE 1-6 U-tube type of manometer.


16 Introduction

the pressure. Thus, the product of the specific weight in pounds per cubic inch
and the column height in inches equals the pressure in pounds per square inch
(see Fig. 1-7). In general terms, since the force exerted by liquid, F = W =
w V = wAz, then
force
p =-= wz (1-4)
area
where p = pressure
w = specific weight
z = column height

In Eq. 1-4, the units of pressure will be fixed by the selection of the units of
specific weight and the column height.
The atmospheric pressure must be added to the gage pressure to obtain the
true or absolute pressure (see Fig. 1-8). Barometers generally are used to
determine the atmospheric pressure. Hence, atmospheric pressure frequently
is referred to as the barometric pressure. Standard atmospheric pressure (i.e.,
pressure of the standard atmosphere) is defined as being equivalent to a
mercury column 760 mm high, where the mercury has a temperature of 0°C.
This is equivalent to a pressure of 14.6960 psi.
When the pressure is less than atmospheric, a vacuum is said to exist. The
magnitude of the vacuum denotes how much the pressure is below that of the
atmosphere. Vacuum generally is expressed as the height of a mercury
column.

FIGURE 1-7 Liquid pressure.


Pressure 17

Gage pressure
(negative) or vacuum
J Atmospheric
r
Absolute
pressure

pressure
t

FIGURE 1-8 Relations of various pressures.

Example 1-4. The vacuum in the evaporator of a refrigerating system is


284 mm of mercury. The barometric pressure is 742 mm of mercury. Deter¬
mine the absolute pressure in pascals.

Solution. The temperature of the mercury is not stated. Frequently, the


observed reading is corrected to 0°C before reporting. It will be assumed that
this has been done here. The density of mercury at 0°C is 13.596 g/cm3. Or,

1 3,'aaa gln— X 1,000,000 cm3/m3 = 13,596 kg/m3


1000 g/kg

From Eq. 1-2,

w = Y = y g = 13,596 kg/m3 x 9.80665 m/s2 = 133,300 kg/m2-s2

Since 1N = 1 kg-m/s2, w = 133,300 N/m3. Remembering that a vacuum is a


pressure below atmospheric, we find that the absolute pressure is (742-
284)/1000 = 0.458 m of mercury. Using Eq. 1-4, we find that the absolute
pressure
p = wz
= 133,300 N/m3 x 0.458 m - 61,060 N/m2
= 61,060 Pa Answer
Example 1-5. Determine the absolute pressure in Example 1-4 in psia.

Solution. From the tables of conversion factors in the Appendix, 1 lb/ =


4.44822 N and lm = 3.2808 ft. Converting the value of 61,060 N/m2 from
Example 1-4,
61,060 N/m2 _ o'too • a

(4.44822 N/lb,) x (3.2808 x 12)2 in2/nr “ 8'789 pSla Answer

The results of Example 1-4 point out that, even for low pressures, when
pressures are expressed in pascals, the pressures have very large values. To
avoid this difficulty, except for very low pressures, the term kilopascals (kPa)
is used frequently (1 kPa = 1000 Pa).
Two additional units of pressure are used extensively. One of these is
18 Introduction

atmospheres (atm). As stated earlier, the pressure of the standard atmosphere


is 760 mm of mercury, measured at 0°C. This is equivalent to 1.01325 x
105 N/m2 or 1.01325 x 105 Pa.
The second unit of pressure that is used rather extensively is the bar. The
bar is defined as 105 N/m2 or 10s Pa. The bar is equivalent to 0.986923 atm or
14.50377 psi. Thus a pressure of 1 bar is roughly equivalent to a pressure of
1 atm. Although the bar is not a basic SI unit of pressure, it is used in several
well-known tables, such as the Steam Tables by J. H. Keenan, F. G. Keyes,
P. G. Hill, and J. G. Moore (John Wiley & Sons, New York, 1969). Hence, it
will be used in this text as one unit of measure.
In summary, pressures may be expressed as follows:

1 N/m2 = 1 Pa
1000 Pa = 1 kPa
100,000 Pa - 100 kPa = 1 bar
1,000,000 Pa = 1 MPa = 10 bars
101,325 Pa = 1 atm

Example 1-6. Express the pressure in Example 1-5 in atmospheres and in


bars.

Solution. Using the pressure of 61,060 N/m2,

61,060
pressure ^ = 0.603 atm Answer
1.0135 x 10
or
61,060 n
—j-Q?— — 0.6106 bars Answer

1-7 TEMPERATURE
Temperature is connected with the “hotness” or “coldness” of a body.
However, these terms are purely relative and are not suited for use in a
thermodynamic sense. Experience has shown that when a “hot body” is
brought in contact with a “cold body,” there is a change in the physical
properties of each body. The changes in properties can be correlated with the
“hotness” of the body or with its temperature. Consider two blocks of steel,
A and B. Assume that our senses tell us that block A is “hot” and block B
is “cold.” Assume that the dimensions of the two blocks can be measured with
a high degree of accuracy. Now bring the two blocks together. Assume that a
running check can be made of the dimensions of the two blocks. It will be
found that there is a very small decrease in the dimensions of block A and a
very small increase in those of block B. These dimensions will change with
time, the changes becoming progressively smaller as time passes. Ultimately,
it will be found that there will be no further change in dimensions with time.
As the dimensions approach their final values, it becomes very difficult and
Temperature 19

then impossible for our senses to distinguish between the hot and cold blocks,
even before the blocks reach their final dimensions. When it becomes im¬
possible for our senses to distinguish between the hot and the cold blocks, the
blocks are said to have the same temperature. But this is only an ap¬
proximation since measurements show that the dimensions are not yet stabi¬
lized. Substituting for our senses, we may say that the temperatures of blocks
A and B are equal when there is no further change in measurements. In this
case, temperatures are associated with volumes and temperature changes with
volume changes.
Temperatures may also be associated with other properties. Suppose, as in
Figure 1-9, gases are confined in adjacent compartments separated from one
another by both a heat-conducting partition and a nonconducting partition.
Again, it will be supposed that the temperature of one gas exceeds that of the
other gas. The pressures of the two gases are carefully measured. Now
remove the nonconducting partition (but not the conducting partition), and
continue to observe the pressures. The pressure of the “hot” gas decreases,
while that of the “cold” gas increases. After a lapse of time, the pressures of the
two gases remain stable. It may be concluded that the temperatures of the two
gases are then equal.
From the experiments just described, the following conclusions may be
drawn: When objects or systems that are not at the same temperature are
brought into intimate contact, a change in one or more properties occurs.
When there is equality of temperature, no change in properties takes place.
When equality of temperature exists, it may be said that thermal equilibrium
has been attained.
In the discussion above, it was shown how to ascertain whether or not two
bodies are at the same temperature by bringing the two bodies together. In
most cases, it is not feasible to move the two bodies. Under these conditions,
a third, small, portable body may be selected (see Fig. 1-10). If by measure¬
ments of the desired properties, it is found that bodies A and C are at the
same temperature and also bodies B and C are at the same temperature, it
may be concluded that bodies A and B are at the same temperature. This
conclusion is expressed in the Zeroth law of thermodynamics that states

Pressure
gage

FIGURE 1-9 Pressure changes with temperature changes.


20 Introduction

FIGURE 1-10 Illustration of Zeroth law.

“when two bodies are each in thermal equilibrium with a third body, they are
in thermal equilibrium with each other.”
It is now possible to determine whether or not the temperature of a body
equals a readily reproducible temperature. Two common reproducible tem¬
peratures are those at the “ice point” and the “steam point.”
The ice point is the freezing temperature of air-saturated water at standard
atmospheric pressure, which is 1.01325 x 105 Pa or 14.6960 psia. Or, stated
differently, the ice point is the temperature at which ice and air-saturated
water are in equilibrium at standard atmospheric pressure. The steam point is
the temperature at which pure water boils, or the temperature at which water
and its vapor are in equilibrium at standard atmospheric pressure. By use of a
suitable thermometric measuring device, it can be ascertained when some
object has a temperature equal to that of the ice point or that of the steam
point. The temperature at each of these two points can be arbitrarily desig¬
nated. In the Celsius (formerly called centigrade) system, the temperature of
the ice point is designated as 0 degrees and that of the steam point as 100
degrees. In the Fahrenheit system, the temperature of the ice point is
specified as 32 degrees and that of the steam point as 212 degrees.
To specify a temperature that is between that of the ice point and the steam
point, it is necessary to establish a scale on our thermometric device. Such a
procedure may be illustrated by use of a mercury-in-glass thermometer (see
Fig. 1-11). If the Celsius scale is to be used, a mark may be etched on the
glass to designate the top of the mercury column when the temperature is
0°C. Likewise a second mark may be etched to designate 100°C. The distance
between the two marks may be divided into 100 equal parts, each division
being a Celsius degree.
The procedure outlined for a mercury-in-glass thermometer may be applied
to other thermometric devices. In each case, the assumption is made that the
temperature change is directly proportional to the change in property that is
measured by the device. Unfortunately, the change in various properties as
measured by various thermometric devices is not precisely linear with tem¬
perature change. Hence, the various thermometric devices, with scales
established in the manner previously described, will not indicate exactly the
same temperatures for temperatures in between the steam and ice points. The
differences in the temperature readings indicated by various thermometric
devices are not large, about a maximum of one-half degree Celsius. However,
for precise work, these differences cannot be ignored.
Temperatures vary almost exactly linearly with such properties as pressure
FIGURE 1-11 Liquid-in-glass
thermometer.
22 Introduction

and volume for those gases that are difficult to liquefy. (Such gases are
sometimes called “permanent” gases, because formerly it was thought that
they could not be liquefied.) Some of these gases are oxygen, nitrogen, carbon
monoxide, hydrogen, and helium, particularly the last two. Hence, these gases
are used for accurate temperature measurements in either a constant-pressure
or constant-volume thermometer. In the constant-pressure thermometer,
measured volume changes are correlated with temperature changes. Hence, it
may be said that temperature is a function of the volume. Thus,

T'=CV (1-5)

where T' is the indicated temperature and V is the volume observed from the
constant-pressure thermometer.
The constant-volume thermometer is shown schematically in Figure 1-12. A
capillary tube connects the thermometer bulb with a manometer type of
pressure-measuring device. The liquid in the left-hand liquid-column tube may
be brought to the reference line by adjusting the height of the leveling vessel.
A measurement of the height of the liquid in the right-hand liquid-column tube
permits the calculation of the gas pressure in the thermometer bulb. For this
thermometer,
T' = C'p (1-6)

If Eq. 1-6 is applied to the ice and steam points, the result is

where the subscript s refers to the steam point and i to the ice point.

FIGURE 1-12 Constant-volume gas thermometer.


Temperature 23

It might be expected that the ratio ps/p, should be a constant and be


independent of both the type of gas used and the magnitude of the gas
pressure. Although this expectation is very nearly correct, it is not precisely
so. Hence, the ratio T'JT\ is not a constant but depends both on the gas and
on the pressures used in the gas thermometer. In general, the higher the
pressure, the larger is this ratio. For those gases (hydrogen and helium) having
very low condensing temperatures, the ratio is lower than for other gases and,
furthermore, the ratio changes very little with a change in pressure.
Extensive investigations have shown that, as the ice point pressure is
reduced, the values of the ratio ps/p, for various gases approach each other,
and in the limiting case, become equal at zero pressure (see Fig. 1-13). Thus,

^ = limit (1-8)
Ii Pj~>0 Pi

where Ts and T, are the true temperatures at the steam and ice points. Very
careful experimental determination shows that the right-hand side of Eq. 1-8
has a value of 1.36609 ± 0.00004.
From the foregoing discussion it may be deduced that the attempt to
establish temperature as a function of a property of substances is not
satisfactory because of lack of uniformity in the variations in the various
properties of substances with a variation in temperature. However, when the
pressure in a gas thermometer at the ice-point temperature is reduced and
approaches zero as a limit, then a temperature may be measured which is
independent of the gas. It is for this reason that the temperatures in Eq. 1-8
are said to be true temperatures.
Since for the Celsius system of temperature, Ts - T, = 100, and T5/T, =
1.36609,

Tj = 273.15 degrees (1-9)

This temperature of 273.15 degrees is a temperature above the lowest con¬


ceivable temperature. Hence, it is known as an absolute temperature. Absolute

FIGURE 1-13 Effect of ice point pressure on ratio of ps/p,


(not to scale).
24 Introduction

temperatures are expressed in degrees Kelvin (K) in the Celsius system. Thus
Eq. 1-9 becomes
T, = 273.15 K

The ratio of one degree Celsius to one degree Fahrenheit equals 180/100 or
1.8. Absolute temperatures are expressed in degrees Rankine (R) in the
Fahrenheit system. Thus, Eq. 1-9 becomes

T, = 273.15 x 1.8 = 491.67 R (1-10)

In the preceding discussion, one of the two fixed points used was the ice
point. This point was chosen since, by definition, the temperature at this point
is exactly zero degrees on the Celsius scale. However, it is somewhat difficult
to reproduce this temperature precisely. It is somewhat easier to reproduce
the triple-point temperature, the temperature where ice, water, and water
vapor all are in equilibrium. The temperature at the triple point is 0.01°C or
32.018° F.
To measure a temperature other than that at the steam point, Eq. 1-8 may
be modified as follows:
T
T,
= limit (—\ (Ml)
Pi-o \PiJ

where p is the observed pressure in the gas thermometer at temperature T.


From the discussion of the Fahrenheit and Celsius systems it follows that

°F= 1.8°C + 32° (1-12)


Likewise
T-32c
>C d-13)
1.8

The relations between the various temperature scales are shown in Figure
1-14.
When Pi has a finite value, the temperature determined by use of pressure
measurements is a function of the properties of the gas that is employed
(refer to Eq. 1-11). However, when p, is reduced to zero, the temperature
determined becomes independent of the properties of the gas. Hence, this
temperature sometimes is known as a thermodynamic temperature. An
entirely different concept of a thermodynamic temperature will be discussed
in Chapter 6, as a consequence of the second law.
Although the concept of a thermodynamic temperature is a basic one, it is
very difficult to use directly in the determination of temperature. Because of
this, many efforts have been made to establish a temperature scale that is
easily reproducible, and yet coincides very closely with the thermodynamic
scale of temperature. The latest of these efforts were those of the International
Committees on Weight and Measures. They adopted the International Practical
Temperature Scale of 1968 (IPTS-68), which is reproduced in Table 1-1 with the
addition of calculated values in degrees Rankine and Fahrenheit.
Rankine Fahrenheit Kelvin Celsius
s~\ A

671.67 212 373.15 100

FIGURE 1-14 Relations between various temperature scales.

Table 1-1
Defined Fixed Point, International
Practical Temperature Scale of 1968 (IPTS-68)

Ka °Ca R °F

t.p. hydrogen 13.81 -259.34 24.86 -434.81


b.p. hydrogen at 25/76 atm 17.042 -256.108 30.68 -428.99
b.p. hydrogen 20.28 -252.87 36.50 -423.17
b.p. neon 27.102 -246.048 48.784 -410.89
t.p. oxygen 54.361 -218.789 97.850 -361.82
b.p. oxygen 90.188 -182.962 162.338 -297.33
t.p. water 273.16 +0.01 491.688 32.018
b.p. water 373.15 100 671.67 212
f.p. zinc 692.73 419.58 1246.91 787.24
f.p. silver 1235.08 961.93 2223.14 1763.47
f.p. gold 1337.58 1064.43 2407.62 1947.95

t.p. = triple point.


b.p. = boiling point at one standard atmosphere,
f.p. = freezing point.
“Defined temperatures.

25
26 Introduction

Example 1-6. The pressure in a constant-volume gas thermometer is


measured as 32 mm of mercury above atmospheric at the triple point of water.
Determine the temperature (°C) when the pressure is 76 mm of mercury above
atmospheric pressure. The barometric pressure is 752 mm of mercury.
Solution. Rewriting Eq. 1-11,
T _ . . / p \ where subscript t.p. refers to the
Tt p. ~~ ™ VPt.p./ triple point of water
T _ 752 + 76 _ .
273.15 + 0.01 752 + 32
T = 288.49 K or 15.34°C Answer
Note: Since the gas pressures are low, the gas so closely approximates an
ideal gas that the calculated temperature may be taken as being correct.

1-8 OTHER THERMODYNAMIC PROPERTIES


In addition to temperature and pressure, another common intensive ther¬
modynamic property of a system is specific volume. Specific volume is
defined as the volume per unit mass, or,
V
(1-14)

where u = specific volume


V = total volume
m = mass
Density is defined as the mass per unit volume, or,
m 1
(1-15)

where p = density. Specific weight is the weight per unit volume or


W
(116)

where w = specific weight


W = total weight
When the density is expressed in terms of pound mass and the specific
weight in terms of pounds force, the two quantities may be assumed to be
numerically equal unless very precise results are desired or unless the system
is located at a very high elevation.
In addition to the thermodynamic properties introduced in this chapter,
additional ones such as internal energy, enthalpy, and entropy will be intro¬
duced in subsequent chapters.
Pressure, Temperature, and Molecular Kinetic Energy 27

1-9 PROCESSES
Engineers are concerned with systems in which changes of some kind take
place. The term process is used to describe how changes in state occur. A
system may change from its initial to final state in one specified manner. Such
a change is said to be a single process. A change in state may start in one
manner and then proceed in one or more different manners before the
terminal state is reached. For these cases, two or more processes are said to
take place between the initial and final states. When a system undergoes a
series of processes and ultimately returns to its initial state, the series of
processes involved constitutes a cycle.
Many open systems operate under steady-state conditions (i.e., the states of
various parts of the system do not change with time). However, there is a
change in state of the fluid as it moves through the system. Thus, for an open
system, the term process is used to describe what happens to the fluid as it
moves through the system.
In some processes, certain properties remain constant throughout the
process. When the pressure remains constant throughout a process, the
process is said to be a constant-pressure or isobaric process. As the name
implies, the temperature remains constant in a constant-temperature process.
This process frequently is called an isothermal process. Likewise, a process
may be a constant-volume process.5 Another common process that will be
encountered frequently throughout this text is the adiabatic process. An
adiabatic process is defined as a process that takes place with no heat
exchange between the system and its surroundings, that is, the system is
thermally isolated from its surroundings. (Although no heat is added in an
adiabatic process, such a process definitely is not an isothermal process for
gases and vapors since work exchanges with the surroundings will cause a
temperature change in the system.)
During the change in state of a system, there may be variations in proper¬
ties, such as pressure and temperature, throughout the system. Under these
conditions the state of the system cannot be specified. In most of these cases
it is not possible to describe fully the nature of the process taking place. On
the other hand, when the properties are uniform throughout the system and
remain so, the system is said to be in equilibrium and the type of process can
be fully specified for a change in state of the system. The term equilibrium
will be discussed more extensively later on.

1-10 PRESSURE, TEMPERATURE, AND MOLECULAR


KINETIC ENERGY
The relationships between temperature, pressure, and molecular kinetic
energy are discussed in detail in Chapter 15. However, it is desirable at this
time to obtain an elementary concept of these relationships. For our purposes

Sometimes referred to as an isometric or isochoric process.


28 Introduction

here, consider an ideal gas. The volume of the molecules themselves as well
as their intermolecular attractions can be neglected. The temperature of the
gas can be determined by inserting a thermometer into the gas. If the gas
temperature is high, energy will be transferred from the gas molecules to the
thermometer. Thus, the thermometer measures the energy of the molecules of
an ideal gas. This energy is kinetic energy. Thus, it can be said that tem¬
perature is a measure of the mean kinetic energy of the molecules of an ideal
gas.
In Chapter 15, it is pointed out that the pressure exerted by a molecule as it
strikes its containing wall is a function of its mass and its velocity. The total
pressure is a function of the product of the pressure exerted per impact and
the number of impacts per unit time. The number of impacts per unit time is a
function of the mean velocity of the molecules and the number of molecules.
Thus, the total pressure is a function of the product of mass per molecule and
its velocity, and the number of molecules and their mean molecular velocities,
or, the total pressure is a function of the number of molecules, the mass of
the molecules, and the mean square of their velocity. In other words, the
pressure exerted by the molecules of an ideal gas is proportional to the product
of the mean kinetic energy per molecule and the number of molecules.
The concepts presented here for an ideal gas are of value because they
show the various factors that influence temperature and pressure. The devia¬
tions in these relationships for actual gases are discussed in subsequent
chapters.
Problems 29

PROBLEMS _____
1-1 State whether the following systems are open or closed: (a) electric
storage battery, (b) incinerator, (c) household refrigerator, (d) fuel cell,
(e) steam turbine.
1-2 Specify, if possible, where the system boundaries must be placed if
the following systems can be taken as closed systems: (a) air con¬
ditioning unit, (b) nuclear reactor, (c) household heating system, (d)
internal combustion engine.
1-3 The same as Problem 1-2 except that the systems are to be open.
1-4 The refrigerant entering the evaporator of a refrigerating system con¬
sists of a vapor that entrains drops of liquid. Is the refrigerant homo¬
geneous?
1-5 (a) Is atmospheric air normally homogeneous?
(b) Same as (a) but during a rainstorm.
1-6 In a gas-turbine locomotive, the energy output is taken as the energy
delivered by the wheels. Specify the boundaries of the system.
1-7 Determine the mass of 122 liters of water having a temperature of 50°C.
1-8 Determine the weight in newtons of the water in Problem 1-7 where the
acceleration of gravity is 960 cm/s2.
1-9 Determine the pressure, in pascals, exerted by a column of mercury
1.87 m high that has a temperature of 90° C.
1-10 A salt is used in a heat storage system for solar energy. When can the
system be taken to be homogeneous?
1-11 (a) During expansion in a steam turbine, some of the steam condenses.
Is the steam-water mixture at turbine exit homogeneous?
(b) Is the steam at turbine entrance normally homogeneous?
1-12 A mercury manometer indicates a vacuum of 62 cm of mercury (20°C)
in a condenser. Determine the absolute pressure in the condenser in
pascals if the barometer pressure is 750 mm of mercury.
1-13 A water manometer (15°C) indicates a pressure of 32 cm of water
above atmospheric. The barometric pressure is 752 mm of mercury.
Determine the absolute pressure.
1-14 A tank having a volume of 1.22 m3 contains 6.48 kg of nitrogen and
7.78 kg of oxygen. Determine the density and the specific volume of the
mixture.
1-15 The pressure in a constant-volume thermometer at the ice point is
12,300 Pa. Calculate the pressure at a temperature of 85°C. Is this
calculated pressure precise or a close approximation?
1-16 Determine the force required in newtons to give a 4200-g mass an
acceleration of 27 cm/s2.
30 Introduction

1-17 A mercury column is to be used in a U-tube manometer to measure a


pressure difference which is estimated to be about 1800 Pa. Determine
the height of the mercury column required.
1-18 The pressure in a constant volume thermometer is 103.2 kPa at the ice
point. Compute the temperature when the pressure is 112.5 kPa. Is the
computed temperature precise? Why?
1-19 A tank having a volume of 1.62 m3 contains 11.2 kg of nitrogen and
9.8 kg of carbon dioxide. Determine the density and specific volume of
the mixture.
1-20 A barometric pressure is measured as 758 mm of mercury at a tem¬
perature of 20°C. Determine the pressure in kilopascals.
1-21 A pressure difference is measured by use of a U-tube water
manometer. The difference in height of the two columns is 27.7 in. of
water at 82°F. Determine the pressure in pascals.
1-22 An open tank having a diameter of 3.82 m is filled with water to a depth
of 8.25 m. The water temperature is 28°C. The barometric pressure is
72.43 cm of mercury at 28°C. Determine the total force on the bottom
of the tank in newtons.
1-23 The same as Problem 1-22 except find the force in pounds.
1-24 A 6-in. pipe (i.d. = 6.065 in.) 3.2 m in length stands vertically. A 4-in.
pipe (i.d. = 4.026 in.) having a length of 4.8 m is welded on top of the
6-in. pipe. Water at 22°C fills the 4-in. pipe to a depth of 2.24 m. Above
the water is air under an absolute pressure of 242 kPa. Determine the
total force on the bottom of the 6-in. pipe.
1-25 At an altitude of 1800 m, the standard barometric pressure is 61.12 cm
of mercury. Determine the pressure in pascals at the bottom of an open
tank in this location that contains water to a depth of 4.82 m. The
water and air temperatures are both 30°C.
_ 2 _

THE FIRST LAW

2-1 INTRODUCTION
As stated in Chapter 1, the first Saw deals with energy transfers and energy
transformations. Energy was defined in the broad sense as the capacity to do
work. It is difficult to understand energy fully since energy cannot be seen and
does not have substance. It is manifested only by the results it produces.
When added to a system, it produces changes in the physical and/or chemical
aspects of this system. Furthermore, physical or chemical aspects of the
system cannot change unless energy is involved.

2-2 BASIS OF THE FIRST LAW


The law of conservation of energy states that energy can neither be created
nor destroyed. Although this law was established before the structure of the
atom was fully understood, it is still valid when viewed in a wide perspective.
When either fission or fusion of atoms takes place, there is transformation of
matter into energy. There is a precise relationship between the matter trans¬
formed and the energy produced.
In its more general form, the first law may be stated as follows: when
energy is either transferred or transformed, the final total energy present in all
forms must precisely equal the original total energy. There are no implications
in the first law that any particular transformation will occur. The first law is
simply a statement of the relationships among the forms of energy involved
when transformations do take place.
The first law of thermodynamics cannot be proved mathematically. Like
most of our laws of thermodynamics, it is based on experimental obser¬
vations. All such observations that have been made confirm the correctness of
the law.
Since energy transformations take place for all processes, natural or man¬
made, the first law is very far-reaching in its scope. As will be discussed
subsequently, in most cases it is very difficult to measure one or more energy
quantities. In such situations, the application of the first law may enable the
determination of an energy quantity that cannot be measured or may verify an
energy quantity whose accuracy of determination is open to question. Such an
application of the first law is termed an energy balance. Energy balances are used
very extensively in problems involving energy transformations.

31
32 The First Law

2-3 ENERGIES INVOLVED IN THE FIRST LAW


When applied to a system, the first law states that the net energy added to a
system is equal to the difference between the original and final energies of the
system. Thus, an understanding of the first law requires an understanding of
the forms of energy that may be added to a system as well as the forms of
energy possessed by the system.

a. Energy Transferred to a System


In order for energy to be added to a system, there must be a driving force, or
a potential, that will cause energy to cross the boundaries of the system.
There are three such potentials; namely, mechanical forces, electrical forces,
and temperature. The energies associated with these potentials are, respec¬
tively, work, electrical energy (frequently called electrical work), and heat.
When there is a difference in the magnitude of any of these three potentials
on the two sides of the boundaries of a system, there is a possibility that
energy may cross the boundaries of the system. However, energy cannot
cross the boundaries unless a path for transferring the energy exists. For
example, the electric potential outside a system may exceed that within a
system. However, electrical energy will not be transferred unless an electric
conductor is available to conduct the electrical energy. Likewise, the tem¬
perature outside a system may greatly exceed that within the system. But if
the system is heavily insulated, no significant amount of heat can be trans¬
ferred.
Work is defined as the product of a force and the distance through which it
acts. This definition implies that the force causes a displacement and that only
the component of the force in the direction of the displacement is involved in
the evaluation of the work (see Fig. 2-1). Thus,

8W1 = FLdl (2-1)

where W = work
Fl = the component, in the direction of motion, of the force acting on the
object
/ = distance moved or the displacement of the object

Electrical energy (electrical work) is defined as the product of a voltage


difference and the current that flows because of the voltage difference.
Since the term “heat” will be used very extensively, it must be defined very
carefully. Heat is that energy transferred across the boundaries of a system
by virtue of the temperature difference that exists across the boundaries.
Unlike mechanical work and electrical energy, the determination of the heat

'The term 8W indicates that Eq. 2-1 is not an exact differential equation. Since work is done only
during a displacement, the terms W, and W2 are wholly nonexistent. The term f2 SW may be
evaluated as WU2, designating the work done between states 1 and 2, for a particular path or
process.
Energies Involved in the First Law 33

\
V
s
s
V

s
s
s -
V

FIGURE 2-1 Work determination.

that crosses the boundary of a system is difficult. It is true that when the
thermal conductivity of the heat flow path is known, this factor together with
the temperature difference enables the calculation of the heat flow. However,
the thermal conductivity frequently is unknown and the heat flow must be
determined by indirect means. These means are discussed later on.

b. Energy Possessed by the System


The energy possessed by the system may be classified either as the energy the
system possesses as a composite whole, or the energy possessed internally by
the system.
When energy is added to a system, there will be a change in energy of the
system unless a corresponding amount of energy is removed simultaneously
from the system.
The energy of the system may change in several ways. There may be a
change in the elevation of the system, that is, a change in the potential energy
of the system. The addition of energy to a system may change its velocity,
that is, produce a change in the kinetic energy of the system. Both potential
and kinetic energies are concerned with the system as a whole and with its
relationship with its surroundings. These two energies frequently are desig¬
nated as extrinsic energies.
The addition of energy to a system may produce internal changes within the
system: the temperature of the system may increase; the system may expand;
or there may be a phase change in the system. A chemical reaction may occur
within the system. Furthermore, if the system is gaseous and is at a high
temperature, the addition of heat may produce ionization within the system.
In certain systems, nuclear fission or fusion may take place. The energy
associated with any and all of these various internal changes is termed
internal energy and is designated as U.
When there is a change in temperature of the system, there is a change in
the velocity of the molecules, and hence, a change in the molecular kinetic
energy. Molecular kinetic energy is designated as Uk. The system may expand
or contract. Hence, there is a change in the distance between the molecules.
34 The First Law

When there are intermolecular attractive forces present, there will be a


change in the molecular potential energy. Molecular potential energy is
designated as IV
When a chemical reaction takes place, there is a change in the molecular
structure of a system. The energy associated with the change in molecular
structure is known as chemical energy. Under some conditions, there may be
changes in the atomic structure within the system. Such changes may be
ionization, nuclear fission, or nuclear fusion. The energy associated with
changes in the atomic structure is known as nuclear energy. The various forms of
energy are given in Table 2-1.

Table 2-1
Forms of Energy

Transferred Energies

1. Heat Driving potential: temperature difference


2. Work (mechanical) Driving potential: unbalance in mechanical forces
3. Work (electrical) Driving potential: voltage difference

Energies of the System

of the system as a composite wholea


1. Potential energy Associated with elevation
2. Kinetic energy Associated with velocity
Energies of the internal structure of the system (intrinsic or internal)
1. Molecular
a. Kinetic Associated with absolute temperature
b. Potential Associated with intermolecular or interatomic forces
2. Atomic
Chemical Associated with changes in molecular structure
3. Subatomic
Nuclear Associated with changes in atomic structure

aSometimes designated as extrinsic energy.

2-4 EVALUATION OF POTENTIAL AND


KINETIC ENERGIES
Energy itself, having neither form nor substance, cannot be measured directly
but must be evaluated by a measurement of the results it produces. Generally
engineers are not concerned with the absolute amount of energy possessed by
a system, but rather they are concerned with changes in energy associated
with changes in the state of the system. Thus, a datum is selected for
expressing the various forms of energy of a system.
In evaluating the potential energy of a system, it is common practice to
select the surface of the earth at the location under consideration as the
Evaluation of Transferred Energies 35

FIGURE 2-2 Work against gravity.

datum. The potential energy of an object not on the surface equals the work
required to elevate it from the surface (see Fig. 2-2). From Eq. 2-1, the potential
energy,
PE. = Fgz
= Wz (2-2)

where W is the force of gravity and hence the weight of the object at the
given location, and z is the elevation above the earth’s surface. Unless there
is an extremely large change in elevation, the weight of the object may be
taken as a constant at any given location.
The kinetic energy possessed by a moving object equals the work required
to bring it from rest to its existing velocity. Thus,

dK.E. = 8W = F dl
But

dV _ dl dV dY
F = ma = m mV
dt m dt dl dl
Then

dK.E. = mV dl = mV dV
dl

Integrating, assuming the initial velocity to be zero,

mV2
K.E. = (2-3)

2-5 EVALUATION OF TRANSFERRED ENERGIES


a. Work
When there is a change in the elevation of a system, work is transferred to or
from the system. This work equals the change in the potential energy of the
36 The First Law

system. From Eq. 2-2,


W,_2= W(z2 - zx) (2-2a)

When there is a change in the velocity of a system, its kinetic energy


changes. The work entering or leaving the system due to its change in velocity
can be evaluated by use of Eq. 2-3.

W,.2 = y(V|-V?) (2-3a)

In addition to the force of gravity, other forces may act on the system,
causing work to be transferred. Some of these forces are body forces. These
forces act on the system as a whole. A common example of a body force is
the force of gravity.
In the fields of electricity and magnetism there are many examples of body
force. One example is the case of two magnets (see Fig. 2-3). Flere the north
pole of magnet A is separated from the south pole of magnet B. The force of
attraction between the two poles is

mxm2
(2-4)
r2

where Fm is the magnetic force


m, is the magnetic strength of pole N
m2 is the magnetic strength of pole S
r is the distance between the poles

By introducing a force equal to the magnetic force, but opposite in direction,


the poles may be separated. The work required is

W,„l 2 = f Fm dr (2-4a)

The value of Fm is given in terms of r in Eq. 2-4.

-s- r->■

FIGURE 2-3 Work against magnetic forces.

As the name implies, surface tension is a surface force. When a liquid


partially fills a vessel, a force, known as surface tension, exists at the surface
of the liquid. Within the body of the liquid, molecules of the liquid are acted
on by adjacent molecules in all directions and hence, the intermolecular
Evaluation of Transferred Energies 37

FIGURE 2-4 Surface tension work.

forces tend to be balanced. At the surface of the liquid, molecules are acted
on only by those molecules that are in the liquid. (The molecules in the gaseous
phase above the surface of the liquid normally are so far apart that they do
not offer a significant force on those at the liquid surface.) The net result is
that there is a downward force exerted on the molecules at the surface. This
downward force resists efforts to extend the surface by the application of a
force. Surface tension is defined as the force required per unit area to
increase the surface area of the liquid. The work to overcome this force is
illustrated in Figure 2-4. A liquid, such as a soap solution, is formed into a
very thin sheet on a wire frame. As a movable wire is pulled downward, the
films are stretched. (Note that there is a surface film on each side of the thin
sheet of liquid.) The work required to stretch the two films,

Wf = 2 a lx (2-5)

where cr = the surface tension per unit area


/ = length of the movable wire
x = distance the wire is moved

A very common example of the need to determine work is gas compression


(see Fig. 2-5a). If the force is assumed to be constant, this work is given by

/?///>//>T?m
(a)

FIGURE 2-5 Compression of a gas.


38 The First Law

the expression

W = FI (2-6)

where F = force in the direction of motion and


l = distance the piston moves

In reality, the gas pressure varies during the compression. Then Eq. 2-6
becomes, in more general form,

W,.2 = J Fdl (2-7)

Since force is a product of pressure and area, Eq. 2-7 may be written as
2 2

W,.2 = f pA dl = j p dV (2-7a)

or

8W = PdV

Note that the work delivered by the face of the piston, Eq. 2-7a, is in¬
dependent of whether or not the system is closed as in Figure 2-5 a or open as
in Figure 2-5 b.
In all cases discussed so far, care must be taken in finding the work
required for the desired effect to use the resisting force. For example, a force
greater than the force of gravity may be applied to the raising of a weight. The
extra force may accelerate the weight or, if the weight was being dragged up
the side of a building, be required to overcome friction. In the use of Eq. 2-7a,
the pressure that must be used is the gas pressure in direct contact with the
piston and not the mean gas pressure in the cylinder. (In most cases these
pressures will be equal.) Furthermore, the pressure at the piston face at all
times must be specifiable in terms of piston position.
In the SI system, since force is expressed in newtons and distance in
meters, the units of work are newton-meters (N-m). One joule (J) equals one
newton-meter. Force may also be expressed in dynes and distance in cen¬
timeters. When this is done, the unit of work is the dyne-centimeter. The
dyne-centimeter is designated as the erg. There are 107 ergs per joule. In the
English system, the common unit of force is the pound force and distance is
measured in feet. Thus, work has the dimensions of pound-feet. Since torque
in the English system also has the same dimensions but is not energy, work is
expressed in foot-pounds to distinguish between the two.

Example 2-1. An elevator having a mass of 585 kg is raised a distance of


32 m. Determine the work done, neglecting frictional effects and the work
required for acceleration.

Solution. Since it is not stated otherwise, it will be assumed that the


acceleration of gravity is substantially equal to the standard value. The weight
Evaluation of Transferred Energies 39

of the elevator,

W = mg = 585 x 9.807 = 5737 N


Work = F x 1 = 5737 x 32 = 183,600 N-m or 183,600 J Answer

Example 2-2. Same as Example 2-1 but determine the work in foot-pounds.
Solution.
585
mass of the elevator = . - - = 1289.7 Ibm
0.45359
where 0.45359 = kilograms per pound
weight of the elevator = 1289.7 lb/
change in elevation = 32 x 3.28 = 104.99 ft
where 3.28 = feet per meter
work = 1289.7 x 104.99 = 135,400 ft-lb Answer

b. Heat
Generally, it is more difficult to measure heat transferred to a system than it is
to measure work. It may be rather easy to measure the temperature difference
causing a flow of heat, but unless the conductivity of the heat flow path is
known, the heat flow cannot be determined directly. Rather, it is necessary to
determine the effects of the heat flow on the system or on the source that is
supplying the heat.
Normally, addition of heat produces a temperature rise of the system. The
amount of heat required to produce a given temperature rise of the system
depends on the nature of the system, its mas^s, and also what takes place
within the system during heat addition. It is therefore necessary to determine
the amount of heat that must be added to produce a temperature rise of one
degree of a unit mass of the system for each type of process. This amount of
heat is known as the specific heat of the system. It is also termed the specific
heat capacity and the heat capacity.
From the definition of specific heat, the amount of heat added to a system is
given as
8Q = me (IT (2-8)

where c is the specific heat of the system.2


In general, when heat is added to a system, it may deliver work at the same
time or work may be added to the system. Hence, the temperature rise of the
system and the specific heat of the system cannot be determined until the
nature of the process taking place during heat addition is specified. The
specific heat for various types of processes is discussed in Chapter 4 under
the heading of polytropic specific heats.

2As in the case of 8W, the term 8Q indicates that this is an inexact differential. Since heat is
transferred only during a change in state, Qi and Q2 are wholly nonexistent. Then /2 8Q becomes
Qi_2, designating the heat added between states 1 and 2 for a particular process.
40 The First Law

It is desirable at this point to discuss two common specific heats, the


constant-volume specific heat, cv, and the constant-pressure specific heat, cp.
When the volume of the system remains constant during addition of heat, the
heat added,
8Q-mcvdT (2-8a)

With the volume remaining constant, no work is performed. Hence, all of the
heat added must go to increasing the internal energy of the system. Or
8Qv=dU (2-9)

Combining Eq. 2-8a and Eq. 2-9,


dU = mc„ dT (2-10)

dU
c =- (2-11)
m dI

The constant-volume specific heat, c„, is a property of the system. This


statement can be justified by recognizing that internal energy is a true
property. The statement can also be justified by recognizing that the value of
cv depends solely on the nature of the system.
A second very common specific heat is the constant-pressure specific heat,
cp. This may be defined as the amount of heat that must be added per unit
mass of the system to produce a unit change in temperature when the
pressure remains constant. Thus,
SQP = mcp dT (2-8b)

This specific heat is also a true property since its value depends solely on the
properties of a system.
Originally, the unit of heat was the calorie (cal), which was defined as the
amount of heat required to increase the temperature of 1 g of water 1°C.
Unfortunately, this amount of heat required varies with the temperature and
the pressure of the water. These effects are shown in Figure 2-6. To avoid the
dependence of a fundamental unit on such variables, the unit of heat in the SI
system is taken as the joule, which as stated earlier, is one newton-meter.
Over the period of years, various international conferences have established
relationships between the joule and other forms of energy. There is a very
small difference in these conversion factors. The International Table fixed the
I.T. calorie as equal to 4.1868 International Table joules. This value will be
used here.
The unit of heat in the English system, the British thermal unit (Btu), was
also originally defined in terms of the properties of water, namely, the amount
of heat required to increase the temperature of 1 lb of water 1°F. Since there
are 453.59237 grams per pound and since 1°C equals 1.8°F, 1 Btu equals
251.9958 cal. Using these figures, the I.T. Btu equals 1055.056 J. Since 1 ft-lb
equals 1.355818 J, 1 Btu = 778.169 ft-lb.
Evaluation of Transferred Energies 41

-20 0 20 40 60 80 100 120 140 160 180 200 220 240 260 280 300 320 340
Temperature (° C)

FIGURE 2-6 Effects of temperature and pressure on the specific heat of water.

Example 2-3. A tank contains 5.75 kg of air. The addition of 64,100 cal
increases the temperature of the air from 20°C to 85°C. Determine the specific
heat of the air.

Solution. From Eq. 2-8,

c = Ql-2
m(T2 —T\)

= 5.75 x 1000(85 - 20) = 01715 Cal^g'°C Answer

Note: It is assumed that because the temperature range is sufficiently small


the specific heat may be taken to be constant.

Example 2-4. Express the amount of air in Example 2-3 in pounds, the heat
added in Btu and the temperature in degrees Fahrenheit. Determine the specific
heat, Btu/lb-°F.

Solution
5.75
mass of air = - 12.68 lb
0.45359
64,100
heat added = = 254.4 Btu
252
temperature change = (85 - 20) 1.8= 117° F
Q,- 2 _ 254.4
c = = 0.1715 Btu/lb-° F Answer
m(T2— T,) 12.68 x 117
42 The First Law

2-6 THE FIRST LAW APPLIED TO SYSTEMS


The general statement of the first law applied to systems may be stated as
follows:
For any change taking place within a system, the final energy of a
system equals the original energy of the system plus the net energy
added to the system during the period in which the changes take place.

As discussed earlier, the system possesses internal energy of various forms.


When the system is in motion, it possesses kinetic energy. If the elevation of
the system is changed, there will be a change in its potential energy. Energy
may be added to the system in the form of heat or work, both mechanical or
electrical. Arbitrarily, the heat added to the system is considered to be a
positive quantity and the work delivered by the system is also taken to be a
positive quantity.
If the system is an open one, a substance, generally a fluid, crosses the
boundary of this system in one or more places. In so doing, it increases or
decreases the energy of the system by the amount of its own energy. As with
any substance, the fluid possesses potential, kinetic, and internal energies. In
addition, since the fluid is forced across the boundary of the system, it carries
with it the energy required to force it across the boundary of the system. This
is illustrated in Figure 2-7. Consider a definite mass of material about to cross
the boundary of the system. Resisting its motion is the pressure, p, at the
boundary. Directly or indirectly, work is required to overcome this pressure.
This work equals the product of a force and a distance, namely, pA and /,
where A is the cross-sectional area of the flow channel and l is the length of
the flow channel occupied by the definite mass. Since Al = V, this work is
given by

Wf = pV (2-12)

It is difficult to describe this work term in one or two words. It is called flow
work for lack of better terminology and is designated as Wf. It should be
thought of as the equivalent amount of work required to move the fluid
against the resisting pressure. It must not be confused with the work entering

(fWill S \ YYYWS \ \ \ \ SS
System

Flow-— p — ■*—P-
JL
-e-/

FIGURE 2-7 Flow work.


The First Law Applied to Systems 43

or leaving the system by means of shafts or other such mechanical devices, or


with electrical work. As the fluid crosses the boundary of the system, it brings
its flow work into the system.
The first law as applied to any system may be expressed in equation form
as follows:
US] + P.E.S| + K.E.S] + (U + p V + K.E. + P.E.)in + Q
= US2 + P.E.S2 + K.E.S2 + (U + p V + K.E. + P.E.)out + W (2-13)

Where s refers to the system, 1 refers to the initial state of the system and 2
to the final state. V is the total fluid volume entering or leaving during the
process under consideration. Q is the heat added to the system and W is the
net work delivered by the system. In general, flow is involved in Eq. 2-13.
Hence, care must be taken to examine the system for a specified time interval
and to make certain that all of the energies are evaluated for that time
interval.
Equation 2-13 can be arranged as follows:

(U + pV + K.E. + P.E.)in+ Q = Us2~ USI + P.E.S2 — P.E.S] + K.E.S2


- K.E.S] + (L7 + pV + K.E. + P.E.)out + W (2-14)
It is to be noted in Eq. 2-14 that only changes in the energies of the system are
required rather than absolute value of these energies.
«

Equation 2-14 appears to be very involved. It may be in some cases but in


others it can be greatly simplified by recognizing that certain energy terms
may be insignificant in comparison with other energy terms. For example, the
system may be at rest or be moving very slowly. For these conditions, the
changes in the kinetic and potential energies of the system may be neglected.
Some systems may be so well insulated that no significant amount of heat
may be transferred. Hence, Q may be assumed to be zero without causing a
significant error. When there is no means of transmitting either mechanical or
electrical energy, W becomes zero.
In applying Eq. 2-14 to work-producing or work-absorbing systems, it is
frequently necessary to become involved with the concept of power. Hence,
this concept will now be discussed. In addition, it is convenient to use a term
that is designated as enthalpy. This term will also be discussed before
applying Eq. 2-14 to various types of systems.

2-7 POWER
The size of work-producing devices is dependent not on the total work to be
delivered, but on the rate at which it is delivered, that is, on the power
required. Power is defined as the time rate of doing work. In the SI system,
the fundamental unit of power is the watt (W). The watt is defined as the work
done at the rate of a joule per second. Since the watt is a small unit of power,
the kilowatt (kW), which is a thousand watts, is used extensively. For still
44 The First Law

larger amounts of power, the megawatt (MW), which is a million watts, may
be used.
Although rates of heat addition to a system may be expressed in terms of
watts, care must be taken not to think of this amount of heat as being equal to
the electrical energy deliverable. For example, a nuclear reactor, rated at
1000 MW, may produce steam for a steam turbogenerator unit. The electrical
output of the generator may be only 320 MW and not 1000 MW. To avoid this
difficulty when dealing with rates of heat transfers expressed in watts, it is
customary to express them as watts thermal. Thus, the rating of the reactor
could be expressed as 1000 MW(.

Example 2-5. A mass of 210 kg is elevated 24.5 m in 1 min 34 s. Determine the


power required.

Solution. Assume that the acceleration of gravity equals the standard value.
Then the weight of the object,

W = mg = 210 x 9.80665 = 2060 N

work = FI = 2060 x 24.5 = 50,500 N-m or 50,500 J


W 50,500
power = — = = 537 W Answer
94

In the English system, the unit of power is the horsepower (hp), which is
defined as work being done at the rate of 33,000 ft-lb/min.

Example 2-6. For Example 2-5, determine the mass in pounds and the
distance in feet. Determine the horsepower required.

Solution. The mass equals 210/0.45359 = 463 lb. Assuming that the ac¬
celeration of gravity substantially equals the standard value,

weight = 463 lb/

j. * 24.5x 100 OAnor,


distance = --7-—— = 80.38 ft
2.54 x 12

work = 463 x 80.38 = 37,200 ft-lb

37 200
p0Wer = (33,000/60) x 94 = 0 72 hp Answer

Since there are 778.169 foot-pounds per Btu, there are 33,000/778.169 =
42.407 Btu per horsepower-minute or 2544.4 Btu per horsepower-hour. Using
the figure of 1055.056 joules per Btu, there are 0.7457 kilowatts per horse¬
power. For these conditions one kilowatt hour (1 kWh) equals 3412 Btu.

Example 2-7. A solar cell system is to supply 1050 kWh/month. The average
output is 15 percent of the rated power. Determine the rated power required.
Solution. Assume 30 days per month.
Enthalpy 45

1050
average power required = - 1.458 kW
30x24
1.458
rated power = = 9.72 kW Answer
0.15

Example 2-8. A steam turbogenerator delivers 750,000 kW when the steam


flow is 2,250,000 kg/h. Calculate the work done per kilogram of steam.

Solution

750,000 kW = 750,000,000 W
= 750,000,000 J/s
2,250,000
steam flow = = 625 kg/s
3600
750,000,000
work = 1,200,000 J/kg Answer
625

Example 2-9. The power required by a plane flying at 545 km/h is 1800 kW.
Determine the average resisting force in newtons.

Solution

1 W = 1 J/s
power required = 1800 x 1000 = 1,800,000 J/s
545 x 1000
distance traveled per second = 151.4 m/s
3600

Since a joule equals one newton-meter,

r 1,800,000 11CAAXT
force = — = 11,890 N Answer

2-8 ENTHALPY
In Eq. 2-13, the terms U and pV represent the energy of a given mass, m, of
fluid entering the system. But U = mu and V-mv. Then U + pV =
m(u + pv), where u is the internal energy per unit mass and v is the specific
volume. As will be discussed in some detail later one, when the state of the
fluid is fixed, the summation of u + pv is also fixed, and hence, has a specific
value. This eliminates the necessity of evaluating each term individually.
Credit for using this combination is given to Professor Williard Gibbs of Yale
University who, in 1879, designated it as his \ (chi) function.3 Unfortunately,
at the turn of the century, other terms such as total heat and heat content,
were applied to this function. These terms are entirely erroneous in their
implications and hence led to considerable confusion. Some years later, the

'This is one of three thermodynamic functions that are credited, in part at least, to Gibbs. The
other two functions will be discussed later on.
46 The First Law

term enthalpy was selected for this function and was given the symbol H. The
defining equation for enthalpy is
H = U + pV (2-15)
and
h = u + pv (2-15a)

Since 17, p, and V all are properties, H must also be a property. Because it
is composed of several properties, it is known as a compound property. As a
property, it can be tabulated in tables of properties of various substances.
Methods will be developed later on to evaluate enthalpy changes for both
ideal and actual gases.
Although the concept of enthalpy is a very valuable one when dealing with
flow, care must be taken in its use in nonflow conditions. The term pV was
shown to be the energy necessary to force a given mass of a fluid across a
given boundary of a system; hence, pV represents energy transmitted by the
fluid as it crosses the boundary of the system. Thus, for a fluid in motion, its
enthalpy is truly energy. On the other hand, for a substance at rest, the pV
term cannot represent energy being transmitted since no energy is being
transmitted. Thus, for a substance at rest, the pV term is not energy. Hence,
enthalpy cannot be energy and must not be used as such.
Although enthalpy can be used as energy only in the case of flow, it may be
useful for other purposes. In certain cases, such as for closed systems, it may
be necessary to determine changes in internal energies. However, some tables
of properties list values of enthalpies but not internal energies. Internal
energies may be calculated readily by using the defining equation of enthalpy,
H = U+pV, or U = H-pV.

2-9 FIRST-LAW APPLICATIONS


Introducing the concept of enthalpy, Eq. 2-14 becomes

(H + K.E. + P.E.)in + Q = US2 - US] + P.E.S2 - P.E.S, + K.E.,2 - K.E.Sl


+ (H + K.E. + P.E.)out + W (2-16)

This general equation can be adapted readily to both open and closed
systems.

a. Open Systems
The application of Eq. 2-16 to open systems is illustrated in Figure 2-8.
However, before applying Eq. 2-16 to open (control-volume) systems, the
problem should be analyzed carefully to eliminate energy quantities that are
either zero or are very small in comparison with other energy terms. For
many open-systems problems, the system may so closely approach steady-
state conditions for the period of time under consideration that variations in
state conditions can be neglected. For steady-state conditions, the properties
First-Law Applications 47

v u
pV pV

FIGURE 2-8 Energy exchanges for an open system at rest.

of the system do not vary with time and the total mass rate of flows into and
out from the system are equal. Then, for a steady-state system, Eq. 2-14
becomes

(H + K.E. + P.E.)flujdsin + Q — (H + K.E. + P.E.)fluidsout-I- W (2-17)

For a steady-state system where the fluid enters at only one place and leaves
at a second place and does not change in composition within the system, Eq.
2-17 may be written for a specified time interval as

H2-H1 + K.E.2-K.E., + P.E.2-P.E.1= Q-W (2-17a)

where subscript 2 designates the fluid leaving the system and 1 the fluid
entering the system. It should be noted that the total energy of the fluid
involved for this specified time interval equals the product of the energy per
unit mass and the mass flow for the time interval. Equation 2-17a calls for the
determination of the change in enthalpy of the fluid rather than the
specification of the absolute value of the enthalpies.

b. Closed Systems
The application of Eq. 2-16 to a closed system is illustrated in Figure 2-9. For
closed systems, Eq. 2-16 reduces to

Q=U52 - USl + K.E.S2 - K.E.S| + P.E.S2 - P.E.S| + W (2-18)

For closed systems at rest, Eq. 2-18 reduces to

Q=u52-Us+W (2-18a)

Although enthalpy represents energy of a fluid only in the case of flow, it


48 The First Law

Direction of

r Mechanicalii
J Electrical f

FIGURE 2-9 Energy exchanges for a closed system.

has a special use for a closed system at rest undergoing a constant-pressure


process. Equation 2-18a may be written in the differential form as

8Q = dU + 8W

Substituting for 8W its value from Eq. 2-7a,

8Q — dU T- p dV

From the definition of enthalpy,

dH = dU + p dV+V dp

For a constant-pressure process,

(dH)p = dU + p dV

Thus, it may be seen that for a constant-pressure process (and only for a
constant-pressure process) occurring in a closed system at rest, the heat
added is equal, numerically, to the change in enthalpy of the system or

(SQ)P = dH (2-19)

c. Flow to Nonflow Systems


There are certain problems in which a fluid flows into a system for a given
period of time, with no fluid leaving the system. As an example, during the
intake stroke of an automotive engine, the air-fuel charge flows into the
cylinder and no fluid leaves the cylinder. In other problems, the fluid flows out
from the system with no fluid entering. This occurs during the exhaust stroke
of an automotive engine. For both of these conditions, Eq. 2-16 can be used
by recognizing that either there is no fluid leaving or no fluid entering the
system. The special case of a flow to nonflow process in which there is no
work done and no substantial heat transfer is illustrated in Figure 2-10. For
this case Eq. 2-16 reduces to Hfluidin = US2 - USl and thus shows that the
internal energy of the system increases by the enthalpy of the incoming fluid,
provided that the kinetic energy of the fluid can be neglected.
The method of solving nonflow to flow problems is quite similar. Consider
First-Law Applications 49

t J

air flowing from a tank, where it is stored under pressure, into a pipe line.
Assume that the velocity in the pipe line is sufficiently low so that the kinetic
energy of the air can be neglected. Unless the temperature of the air differs
significantly from that of the surroundings, heat exchanges can be neglected.
Applying Eq. 2-16, the result is 17S] - US2 = Hairout. As the air flows out from
the tank, the temperature of the air remaining in the tank drops. Hence, the
temperature of the air leaving the tank also drops. However, the total
enthalpy of the air leaving the tank in any given time interval equals the total
change in the internal energy of the air in the tank for the time interval.

Example 2-10. A closed system having a mass of 40 kg has an initial velocity


of 12 m/s. Its velocity is increased to 31 m/s and its elevation is increased by
43 m. During this process, the system receives 25,000 J of heat and 4800 J of
work. The system delivers 0.002 kWh (kilowatt hours) of electrical energy.
Determine the change in internal energy of the system.

Solution. From Eq. 2-18,

U2- 17, = Q-(P.E.2-P.E.,)-(K.E.2-K.E.,)- W


weight of the system = 40 x 9.80665 = 392.2 N
Then

P.E.2-P.E., = 392.2(43)= 16,860 N-m or 16,8601


40(312- 122)
K.E.2 - K.E., = v 2 - J = 16,340 J

Since a watt equals one joule per second,


0.002 kWh = 0.002 x 1000 x 3600 = 7200 J
Then

U2 - 17, = 25,000 - 16,860 - 16,340 + 4800 - 7200 = -10,600 J


Answer
Example 2-11. A closed system having a mass of 85 lb has an initial velocity
of 30 ft/s. Its velocity is increased to 80 ft/s and its elevation is increased by
50 The First Law

122 ft. During the process, the system receives 75 Btu of heat and 2800 ft-lb of
work and delivers 0.02 kWh of electrical energy. Determine the change in
internal energy of the system.

Solution. From Eq. 2-18,

U2 — U] = Q - (P.E.2- P.E.t) - (K.E.2- K.E.,) - W

Assuming the standard acceleration of gravity, the weight of the system is


85 lb. The factor gc = (32.174 lbm ft/lb/ s) must be introduced to determine the
kinetic energy in ft-lb/. Then

85(802 - 302)
2x 32.174x778.16
= -12.33 Btu Answer
Example 2-12. Air enters a nozzle with a velocity of 30 m/s. The decrease in
enthalpy in the nozzle is 170,000 J/kg. Determine the velocity at exit.

Solution. See Figure 2-11. Since the air passes through the nozzle in a very
small part of a second, there cannot be a significant amount of heat lost.
Because no work is delivered from the nozzle and there is no change in
elevation, Eq. 2-17a becomes

K.E.2-k.e., = h,-h2
or
K.E,= K.E, + Hi-H2

For unit kilogram basis,

Then

K.E.2 = 450 4- 170,000 - 170,450 J

rr'j1 = 170,450, r = 584 m/s Answer

K.E.
Hn FIGURE 2-11 Air nozzle.
First-Law Applications 51

Example 2-13. The power delivered by a gasoline engine is 33 kW when


supplied with 12 liters of gasoline per hour. The energy supplied by the
gasoline is 4x 107 J/liter. Water circulating through the jackets of the engine
flows at a rate of 16.2 kg/min and increase 34°C in temperature. Determine the
summation of the heat lost (joules per minute) from the engine (other than in
the jacket water) and the energy carried away by the exhaust gases.
Solution. This example has been simplified because the means of determin¬
ing the energy carried away by the exhaust gases have not yet been discussed
in detail. Assuming that the engine is operating at steady-state conditions, the
energy equation for this problem reduces to
energy supplied to the engine = energy removed from the engine

energy supplied to the engine =


12
77.
7 .
x 4 x 10 = 8x10 J/min
60
work delivered = 33 x 1000x60= 1.98 x 106 J/min
(1 W = 1 J/s)
heat to the water = 16.2 x 1000 x 1 x 34 = 5.508 x 10' cal/min
5.508 x 105 x 60
= 0.3843 x 10' W
860
0.3843 x 10' x 60 = 2.306 x 106 J/min
Then the energy carried away by the exhaust gases plus the heat lost from the
engine = 8 x 106 - (1.98 + 2.306) x 106 = 3.714 x 106 J/min. Answer
Example 2-14. Air flows from a large tank through a line where its state is
maintained constant at a pressure of 70 N/cnr and a temperature of 40°C. The
volume of a gram of air under these conditions is 1.250 x 102cm3. A valve
leading from the line to a completely evacuated tank is suddenly opened. The
pressure in this tank reaches 70 N/cnr very shortly. Neglect heat transfer
between the air and the walls of the tank. Calculate
(a) The increase in the internal energy per unit mass of the air which flows
into the tank.
(b) The final temperature of the air.
Solution. From Eq. 2-16, with both Q and W and also US] equal to zero, and
with no net change in kinetic energy and no change in potential energy,
(a) US2 = H|ine = (U +p YOline
17,2 - Ujine = p V|inc = 70 x 1.250 x 102 = 8.75 x 103 N-cm/g
= 87.5 J/g Answer
(b) From Eq. 2-10, assuming that the value of cv is a constant.
52 The First Law

The specific heat, c*, of air = 0.1715 cal/g-°C (Example 2-3). Since there
are 4.1868 joules per calorie,

87.5
T2 121.9°C
4.1868x0.1715
t2 = 161.9°C Answer
Problems 53

PROBLEMS ____
2-1 An elevator, having a mass of 972 kg, is to be raised a distance of
14.5 km. Determine the minimum work required.
2-2 A piston having a diameter of 24 cm moves a distance of 60 cm when
acted on by a gas pressure of 430 kPa. Determine the work done in
joules.
2-3 Change all data in Problem 2-2 to English units and then compute the
work done in foot-pounds.
2-4 1528 kl of water (temperature = 20°C) are elevated 26 m. Determine the
work required in joules.
2-5 Determine the kinetic energy possessed by a 1050-kg automobile when
it is traveling at a speed of 82 km an hour.
2-6 The automobile in Problem 2-5 crashes into an object which it displaces
1.8 m as it comes to rest. Determine the average force exerted on the
object.
2-7 Water exists in a reservoir at an elevation of 85.3 m above that of the
hydraulic turbine, (a) Determine the potential energy per kilogram of
water, (b) Assuming 100 percent conversion of potential energy into
work, determine the water flow per second to produce 75,000 kW.
2-8 A force of 16.5 kg, acting on an object at an angle of 20° to the motion
of the object, moves it a distance of 28.5 m. Determine the work done
in joules.
2-9 (a) In a domestic hot water heater determine the amount of heat to be
added to 72 liters of water to raise the temperature from 9°C to 55°C.
(b) Determine the kilowatt-hours required.
2-10 A diesel engine delivers 2860 kW of power when burning 615 kg of fuel
per hour. The heating value of the fuel is 42,000 kJ/kg. Determine the
total amount of the heat loss from the engine and the energy carried
away by the exhaust gases per hour.
2-11 The enthalpy of air is increased by 139.5 J/g in a compressor. The rate
of air flow is 16.7 kg/min. The power input is 48.2 kW. Determine the
heat loss from the compressor per minute.
2-12 A reciprocating pump has a displacement of 0.24 m3. During the deliv¬
ery stroke, the mean pressure on the face of the piston is 1024 kPa.
The delivery stroke takes place in 0.006 min. Determine the power
required at this time.
2-13 A spring is compressed 22.7 cm by the application of an average force
of 452 N. Neglecting heat losses, determine the increase in the internal
energy of the spring. Physically, how can the internal energy of the
spring increase?
2-14 The spring in Problem 2-13 is made of steel. Its reported temperature
54 The First Law

rise is 2.5°C. The specific heat of the steel is 0.45 J/g-°C and its mass is
76 g. Check this data and make comments.
2-15 A ball having a weight of 2.32 kg is dropped 22.2 m. It then bounces
8.2 m. Neglecting heat losses, determine the net change in internal
energy of the ball in joules.
2-16 A storage battery delivers current at the rate of 44.2 A for a 24-min
period. The heat lost from the battery during this period is 95 kJ.
Determine the change in internal energy of the battery if the voltage is
12.
2-17 Air enters a diffuser with a velocity of 280 m/s and leaves with a
velocity of 14 m/s. Determine the change in enthalpy in J/kg of air.
State what assumption must be made and justify this assumption.
2-18 (a) Water enters a steam generator at a pressure of 20,000 kPa. It has
an internal energy of 837 J/g and a specific volume of 1.138 cm3/g.
Steam leaves the steam generator at a pressure of 18,000 kPa. It has an
internal energy of 3080 J/g and a specific volume of 18.68 cm3/g.
Determine the heat added per minute if the flow of water is
2,130,000 kg/h.
(b) Comment on the necessity of obtaining accurate values of the pV
term for the water and for the steam.
2-19 (a) Determine the minimum power required to bring a 100-car train from
rest up to a speed of 84 km/h in a period of 3 min. The average mass of
each car is 85,000 kg.
(b) State the assumptions made in part (a).
2-20 An oil pipeline experiences an increase in its elevation of 112 m. Is
there a change in the enthalpy of the oil if heat transfer can be
neglected? Why?
2-21 A tank of air is heated by the addition of 720 kJ. The increase in
enthalpy of the air is 1008 kJ. Explain why the increase in enthalpy
exceeds the heat added.
2-22 The heat-transfer rate in a pressurized water nuclear reactor is
1120 MW. The water temperature increases from 60° C to 320° C. The
mean specific heat of the water is 4.37 J/g-°C. Determine the mass rate
of water flow through the reactor.
2-23 Air enters the combustion chamber of a household warm-air furnace
with an enthalpy of 305 J/g. Gases leave the furnace with an enthalpy of
622 J/g. There are 18 g of air per gram of fuel. The enthalpy of the
entering fuel is 44,500 J/g. The furnace is to deliver 54,800 kJ of heat
per hour. Determine the rate of fuel flow.
2-24 At a speed of 90 km/h, an automobile requires 20 kW of power to
overcome the wind resistance. Determine the wind resistance in new¬
tons.
Problems 55

2-25 The heat output for a nuclear reactor is 1240 MW. Determine the heat
output in kJ/s.
2-26 An elevator and its contents weigh 12,400 kg. Determine the minimum
power required to raise the elevator at a rate of 1.4 m/s.
2-27 The surface level of water in a reservoir is 92.5 m above that of a
hydraulic turbine. The efficiency (i.e., the portion of the available
potential energy the turbine uses) is 90 percent. Determine the water
flow, in kg/s, needed to produce 100,000 kW.
2-28 (a) Determine the kinetic energy in joules possessed by a 1020-kg car
traveling at 30 km/h.
(b) Same as part (a) but at 120 km/h.
2-29 (a) How much heat is required to heat 100 kg of water from 40° C to
80°C? The pressure on the water is 100 bars. Refer to Figure 2-6.
(b) Same as part (a) except from 200° C to 240° C.
2-30 A jet plane uses four jets, each developing a thrust of 15,000 N.
Determine the total power produced at a speed of 1020 km/h.
2-31 It requires 9329.4 kJ to heat 120 kg of ice from -40°C to 0°C. Deter¬
mine the mean specific heat of the ice.
2-32 The enthalpy of a refrigerant is 359 J/g at the compressor entrance and
381 J/g at the compressor exit. The rate of refrigerant flow is 62 kg/min.
The heat lost from the compressor is 12 percent of the work input.
Determine the power input.
2-33 The motor and the compressor of a heat pump are located within the
building to be heated. The heat pump delivers 2.4 times as much heat
directly to the building as the power it receives. The efficiency of the
electric motor is 82 percent. Determine the power input to the motor
when the total heat to be supplied to the building is 120,000 kJ/h.
2-34 Hot gases enter the blades of a gas turbine with a velocity of 550 m/s
and leave with a velocity of 120 m/s. There is an increase in the enthalpy of
the gases in the blade passages of 5.1 J/g. The rate of gas flow is 98 kg/min.
Determine the power produced.
2-35 Humid air enters a dehumidifier with an enthalpy of 50.2 J/g of dry air
and 2550 J/g of water vapor. There are 0.02 g of vapor per gram of dry
air at entrance of 0.009 g at exit. The dry air at exit has an enthalpy of
30.7 J/g. The vapor at exit has an enthalpy of 2522 J/g. The condensate
at exit has an enthalpy of 51.1 J/g. The rate of flow of the dry air is
132.5 kg/min. Determine the rate of heat removal in the dehumidifier.
2-36 (a) Steam enters a frictionless turbine nozzle with a velocity of 30 m/s
and leaves with a velocity of 550 m/s. Determine the change in enthalpy
in the nozzle. Justify the assumption that must be made.
(b) Same as part (a) except for an actual nozzle. Does the amount of
friction in the nozzle change the answer? Why?
56 The First Law

2-37 (a) Air flows from a very large tank where conditions are maintained
constant into a small tank that, initially, was completely evacuated.
After a short period of time, the pressure in the small tank is ap¬
proximately equal to that in the large tank. Compare the final tem¬
perature of the air in the small tank with that in the line at the entrance
to the small tank. Explain.
(b) If the temperature of the air in the large tank is 40° C, find the final
maximum temperature in the small tank.
_ 3 _

IDEAL AND ACTUAL


GASES

3-1 INTRODUCTION
Although the majority of thermodynamic problems encountered by engineers
involve gases and vapors, it is desirable at this time to examine briefly the
existence of substances in each of their three phases; namely, solid, liquid,
and gaseous. Because it is the one substance that is encountered most
frequently in its three phases, water will be used here to illustrate the
existence of substances in their three phases.
Ice at atmospheric pressure and at temperatures below 0°C will sublimate,
that is, pass directly from the solid to gaseous phase. Ice, when surrounded by
air at standard atmospheric pressure, will melt at 0°C when heat is supplied to
it. When ice, water, and water vapor are contained in an isolated system at a
vapor pressure of 611.3 Pa (0.006113 bars)1 and a temperature of 0.01°C, they
will be in a state of equilibrium, that is, there will be no change in the amount
existing in each phase. This condition is known as the triple point, with the
temperature being the triple-point temperature and the pressure the triple¬
point pressure.
Water existing at a temperature above 0°C has a tendency to vaporize.
When sufficient heat is added to it, it will boil. The temperature at which it
boils is dependent on the pressure imposed on the vapor produced (see Fig.
3-1). When the vapor pressure is low, say 1227.6 Pa (0.012276 bars)1, water
boils at a temperature of 10°C. At standard atmospheric pressure, water boils
at 100°C. At the very high pressure of 20 MPa (200 bars), water will not boil
until it is heated to 365.81°C.
Because the boiling temperature increases with an increase in pressure, the
specific volume of boiling water also increases with pressure. This is shown in
Figure 3-2. However, as the pressure is increased, the volume of the vapor
produced by boiling decreases. Thus, with an increase in pressure, the specific
volume of water and its vapor approach each other. At a pressure of
22.09 MPa and a temperature of 374.136°C, the two specific volumes are
equal. An examination of other properties of boiling water and its vapor show

'A bar equals 105 Pa or 0.9869 atm.

57
Q
FIGURE 3-1 Liquid boiling under pressure.

FIGURE 3-2 Boiling liquid and its vapor.

Temperature

FIGURE 3-3 p-T diagram for water.

58
Introduction 59

that they too approach each other as the boiling pressure is increased, and
become equal at the same conditions as do the specific volumes. This state at
which the properties of a boiling liquid and its vapor become equal is known
as the critical state. The pressure at this point is the critical pressure and the
temperature is the critical temperature. At pressures and temperatures greater
than the critical values, there is no distinction between a liquid and its vapor.
Turning to the solid phase, as the pressure on ice is increased, the melting
temperature decreases. At a pressure of approximately 210 MPa (2100 bars),
ice melts at —21°C. At higher pressures ice can have six other phases. Since
the engineer seldom encounters ice under such high pressures, these other
phases will not be discussed here.

FIGURE 3-4 p -v diagram for water.

FIGURE 3-5 p-T diagram for carbon dioxide.


60 Ideal and Actual Gases

The three phases of water are shown on the p-T plane, Figure 3-3, and the
P-ia plane. Figure 3-4. It is to be noted that water expands upon freezing.
Most other substances contract upon freezing. Figures 3-5 and 3-6 show the
p-T and p-v- planes for such a substance, in this case, carbon dioxide.
The p-v and p-T planes may be combined to form a p-v-T three-
dimensional surface for various substances. A p-^-T surface for water is
shown in Figure 3-7, and a p-v-T surface for substances that contract upon
freezing, in Figure 3-8.
Ideal Versus Actual Gases 61

FIGURE 3-8 p-v-T surface for carbon dioxide.

3-2 IDEAL VERSUS ACTUAL GASES


In a general sense, thermodynamics is concerned with substances in all of the
three phases that have been discussed. Although some of the thermodynamic
problems encountered by engineers involve liquids and, in a few instances,
solids, most thermodynamic problems ordinarily involve gases and vapors.
Such problems generally involve the intensive properties of systems. Hence,
the intensive properties of gases and vapors now will be investigated. Nor¬
mally, it is very difficult to specify the intensive properties of a heterogeneous
system because of the nonuniformity of such a system. On the other hand, the
intensive properties of a homogeneous system, which is isolated, tend to
become uniform throughout the system. In the discussion that follows it will
be assumed that the intensive properties are uniform throughout the system
under consideration.
Previously, it has been pointed out that there is an interrelationship be¬
tween the various properties of a substance. For the most common proper¬
ties—pressure, specific volume, and temperature—the relationship may be
expressed as

P=f(v,T) (3-1)

Two early investigators, Boyle and Charles, made experimental studies of the
relationships among these properties. Both worked with gases that are not
easy to liquefy, such as oxygen, nitrogen, carbon monoxide, and hydrogen.
62 Ideal and Actual Gases

For the most part, their work was done at low or moderate pressures and at
temperatures much in excess of the condensation point of the gases.
Boyle investigated the pressure-volume variations when the temperature
was held constant. His results may be expressed as

(pv)T = C (3-2)

The subscript T specifies that the temperature is constant.


In some of Charles’ experiments, the volume was held constant. In other
experiments, he held the pressure constant. Charles’ results may be expressed
as
(J) = C (3-3)
/V-

and

(3-4)

The subscripts denote the property that is held constant. Equations 3-2 and
3-4 may be combined to establish a relationship between pressure, volume,
and temperature. This is done in Appendix A-14 and shows that

p v- = RT (3-5)

pV = mRT (3-5a)

where v- = specific volume


V = total volume
m = mas's
R = the proportionality constant, known as the gas constant. Its magnitude
depends on the gas under consideration.

An ideal gas is defined as one that obeys the laws of Boyle and Charles, and
hence Eq. 3-5. Thus, Eq. 3-5 is said to be the characteristic equation of an
ideal gas.
Later investigators, such as Amagat, found that many gases deviated greatly
under many conditions from the laws of Boyle and Charles, and that even the
gases they studied showed large deviations from these laws at low temperatures
and high pressures. To investigate these variations, Amagat plotted the pv
product against pressure for various temperatures. Such curves are known as
Amagat’s isothermals. A set of Amagat’s isothermals are shown for carbon
monoxide in Figure 3-9 and for steam in Figure 3-10. For an ideal gas, these
isothermals must be horizontal lines. From these two figures it may be noted that,
for steam, the isothermals deviate greatly from the horizontal, the deviations
being greater at low temperatures. For carbon monoxide, the deviations are large
at very low temperatures, but at the relatively low temperature of 300 K, the
isotherm is almost horizontal except for pressures above 150 atm.
Ideal Versus Actual Gases 63

FIGURE 3-9 Amagat’s isothermals for carbon monoxide.

An examination of Amagat’s isothermals shows that the relationship between


pressure, volume, and temperature can be formulated as

p » = ZRT (3-6)

where R = a constant, depending on the nature of the gas


Z = a proportionality factor, called the compressibility factor.
Z is a function of both temperature and pressure.
product (p in bars, v in cm3/g)
pv
64 Ideal and Actual Gases

Multiplying each side of Eq. 3-6 by the mass, m,

pV = ZmRT (3-6a)

Equation 3-6a is known as the characteristic equation of an actual or real gas.


A comparison of Eqs. 3-5 and 3-6 shows that when the compressibility factor
is unity, the gas is an ideal one.
The compressibility factor for carbon monoxide is shown in Figure 3-11,
and for steam in Figure 3-12. Several facts may be noted from these figures:

1 Under certain conditions, very serious errors will be caused by assuming that
the gas is an ideal one. For example, for steam at a pressure of 300 bars and a
temperature of 380°C, the compressibility factor is 0.175. This means that the
actual specific volume is only 17.5 percent of that of an ideal gas.
2 At high temperatures and at low and moderate pressures relative to the
critical values, the compressibility factor is close to unity, and hence, the
gas behaves substantially as an ideal gas.
3 For a given temperature and pressure, there may be a large difference
between the compressibility factors of steam and carbon monoxide.

For the conditions used by Boyle and Charles, namely for temperatures
that are not much below normal ambient temperature and for pressures not in
excess of approximately 20 atmospheres, it is evident from Figure 3-11 that
carbon monoxide may be treated as an ideal gas. A study of the compressibility

Pressure (atm)

FIGURE 3-11 Compressibility factors for carbon monoxide.


The Mole 65

Pressure (bars)

FIGURE 3-12 Compressibility factors for steam.

factor of other hard-to-liquefy gases, such as oxygen, nitrogen, hydrogen, and


helium, shows that these gases also may be treated as ideal gases under similar
conditions of temperature and pressure.
When sufficient data are available for a given gas, the compressibility factor
may be plotted, and hence, used to determine the pressure-volume-tem¬
perature relationship for the gas in a given state. Although information to
determine the compressibility factor is available for the most common gases,
such information is not available for many other gases. Later in this chapter,
the concept of reduced properties and corresponding states will be intro¬
duced. This will enable an approximate determination of properties, including
the compressibility factor, when only a very limited knowledge of the gas is
available.

3-3 THE MOLE


As stated in Chapter 1, gas pressure is proportional to the mean molecular kinetic
energy for an ideal gas and temperature is a measure of the mean molecular
kinetic energy. Hence

pV = c'nT (3-6b)

Equation 3-6b cannot be used easily since it calls for a knowledge of the
number of molecules, n, of the gas. The number of molecules present in even
a very small volume of a gas under normal conditions is enormous. For
example, a cubic centimeter of air at standard atmospheric pressure and 25°C
contains approximately 2.46 x 1019 molecules. What is required, then, is an
unit of quantity, similar to the dozen or the gross, but for a very much larger
number of molecules. Such an unit is the mole.
66 Ideal and Actual Gases

In the SI system, the mole is the gram mole (g mole). The gram mole is
defined as the mass of the substance whose mass in grams is equivalent to its
molecular “weight.” There are approximately 6.02486 x 1023 molecules in a
gram mole. This number, designated as n0, is known as Avogradro’s number.
Although the mole has mass and volume, it is not a unit of mass or of volume
but simply is a very definite, specific number of molecules. By definition, the
number of moles present,
tn
N = ™ (3-7)
m

where m is the mass per mole, which equals, numerically, the molecular
“weight.”
Since the total number of molecules present, n, is the product of the
number of molecules per mole, n0, and the number of moles, N, Eq. 3-6b may
be written as

pV = c'n0NT (3-8)

The two constants in Eq. 3-8, c' and n0, may be combined into a single
constant, which is designated as R0. Since nothing was assumed as to the kind
of gas involved in deriving Eq. 3-6b, other than that it obeyed certain pos¬
tulates, R0 is independent of the kind of gas under consideration and, hence, it
has a specific value for all gases obeying the postulates set up in Section 1-10.
For this reason, R0 is known as the universal gas constant. Equation 3-8 now
becomes

pV = NRqT (3-9)

When the pressures are expressed in newtons per square meter, volumes in
cubic meters, and temperatures in degrees Kelvin, R0 has the value of
8.314 N-m/g mole-K or 8.314 J/g mole-K.
Comparing Eqs. 3-5a and 3-9, it is evident that the products of mR and NR0
must be equal. Substituting the value of N from Eq. 3-7, the particular gas
constant,

R=^ (3-10)
m

The value of R may be calculated readily for any gas by use of Eq. 3-10,
when the molecular “weight” of the gas is known. Under most conditions, Eq.
3-5a is more readily usable since the mass of a gas present may be directly
available. For convenience, the values of R are presented in Table 3-1.
As was done for actual gases in Eq. 6-a, the compressibility factor, Z, can
be introduced into Eq. 3-9. Thus, for actual gases,

PV = ZNR0T (3-11)
In the English system, the pound mole (lb mole) is defined as the amount of a
substance whose mass in pounds is equivalent to its molecular “weight.”
The Mole 67

Table 3-1
Properties of Gases

Molecular Gas
Gas Formula Weight Constant, R Cp C ka

Air 28.97 0.2870 1.004 0.718 1.40


Carbon
monoxide CO 28.011 0.2968 1.029 0.734 1.40
Carbon
dioxide C02 44.011 0.1889 0.845 0.647 1.31
Helium He 4.003 2.0769 5.230 3.128 1.67
Hydrogen h2 2.016 4.1240 14.186 10.08 1.41
Nitrogen n2 28.016 0.2968 1.029 0.734 1.40
Oxygen o2 32.000 0.2598 0.908 0.651 1.40
Water vapor h2o 18.016 0.4615 1.846 1.377 1.34

Note: The specific heats are for the gases at low pressures and at near room temperatures. The
units are joules per gram-degree K. The units for R are newton-meters per gram-degree K, or
N-m/g-K.
ak is defined as the ratio of cp to cv. See Sections 4-6 and 4-7.

When pressures are expressed in pounds per square foot, volumes in cubic
feet, masses in pounds, and temperatures in degrees Rankine, the value of R0
is 1545 lb/ ft/lb mole-R.

Example 3-1. Carbon monoxide exists at a pressure of 120N/cm2 and a


temperature of 92° C. There are 4.2 kg of carbon monoxide present. Determine
the volume of the carbon monoxide.

Solution. Atmospheric pressure equals 1.01325 bars. One bar equals 100,000
newtons per square meter or 10 newtons per square centimeter. Thus, the
pressure of the carbon monoxide,

P
120
10x 1.01325
t1
n-84a,m
,
The absolute temperature equals 92 + 273 = 365 K. It is evident for these
conditions, as shown in Figure 3-9, that the compressibility factor is sub¬
stantially unity.
The number of gram moles of carbon monoxide from Eq. 3-7 is
4.2 x 1000
N = 150 g moles
28

Using Eq. 3-9,


150x8.314x365
V= = 0.3793 nr Answer
120 x 10,000
68 Ideal and Actual Gases

Example 3-2. Express the pressure in Example 3-1 in psia, the temperature
in degrees Rankine, and the mass in pounds. Determine the volume in cubic
feet.

Solution. One newton per square centimeter equals 1.4504 pounds per
square inch.

pressure p = 120 x 1.4504= 174.0 psia


temperature = 92 x 1.8 -I- 32 = 197.6°F = 657 R
4.2
mass = = 9.259 lb
0.45359

9.259(1545/28)x 657
volume = 13.39 ft3 Answer
174x 144

Example 3-3. Determine the density of steam at a pressure of 15.5 MPa and
a temperature of 410°C.

Solution. From Figure 3-12 the compressibility factor for steam is 0.76.
Using Eq. 3-10, the volume of 1 mole of steam,

ZNR0T 0.76 x 1 x 8.314 x (410 + 273) 278.4 3/


V = —T- =-15.5X 1,000.000-= -W cm'/g m°le
or
278.4 1 06
1C._ 3/
"T0rXl6 = I5*47 cm’/g

1
= 0.0646 g/cm Answer
15.47

3-4 EQUATIONS OF STATE FOR ACTUAL GASES


A relationship between pressure, temperature, and volume for actual gases
can be> established when the compressibility factor is known. But the com¬
pressibility factor is dependent on both temperature and pressure. This means
that the compressibility factor must be known over the whole range of
temperatures and pressures. An alternative to this approach is to formulate an
equation of state for actual gases.
In formulating such an equation, it should be recognized that under certain
conditions all gases deviate greatly from an ideal gas. It is helpful to question
why this is true. The behavior of ideal gases can be explained by making two
assumptions: (1) The molecules are so far apart that there are no molecular
attractions. (2) The size of the molecules is so small relative to the volume
they occupy that it can be neglected. When molecules are forced closer
together by increasing the pressure, molecular attractions and molecular size
both become progressively important and deviations of the gas from the ideal
become progressively larger. At a given pressure, an increase in temperature
produces two effects. First it causes the volume, and hence, the distance
Equations of State for Actual Gases 69

between the molecules to increase. In addition, there is an increase in the


kinetic energy of the molecules, thus helping to overcome molecular attrac¬
tions. Thus, as the temperature of a gas is increased, it tends to behave more
nearly as an ideal gas.

3-5 VAN DER WAALS EQUATION OF STATE


Over the years, some 50 to 60 equations of state have been developed to give
more accurate results than the ideal gas equation. Some of these equations are
based on analytical considerations; others are empirical or semiempirical.
Some equations are relatively simple; others are very complex. All of these
equations of state have some merit in that they produce acceptable results for
some conditions. No equation, regardless of its complexity, is wholly satis¬
factory for all conditions of temperature and pressure.
One of the earliest equations of state was that proposed by J. D. van der
Waals, a Dutch physicist in 1873. The van der Waals equation is a simple one.
It is therefore valuable for illustrating the use of an equation of state to
obtain various properties of gases without becoming involved in the excessive
mathematical calculations required by use of the more complicated equations
of state. Moreover, the van der Waals equation gives fairly satisfactory
results for low to moderate pressures and moderate to high temperatures. The
van der Waals equation will be examined in some detail and then other
equations of state will be considered.
In Section 3-4, it was stated that intermolecular attractions cause deviations
from the ideal equation of state. As an illustration, consider a particular gas
molecule that is ready to strike its containing wall. According to the ideal-gas
equation of state, it will exert a certain pressure on the containing wall.
However, other molecules that are close enough to exert an attractive force
will prevent the particular molecule from exerting its full pressure on the
containing wall (see Fig. 3-13). This reduction in pressure caused by inter¬
molecular attractive forces is known as the internal pressure. Because of
intermolecular forces, the reduction in pressure exerted on a certain area of

Hemisphere
of action

FIGURE 3-13 Pressure reduction by adjacent molecules.


70 Ideal and Actual Gases

the wall by each molecule is proportional to the number of molecules in the


sphere of action around that area. Furthermore, the combined reduction in
pressure exerted by all the molecules in that partial sphere of action is
proportional to the number of molecules in the partial sphere. Hence, the
internal pressure exerted by all molecules in a sphere of action may be
considered proportional to the square of the number of molecules in a given
volume or proportional to the square of the density of the gas.2 The sum of
the external or observable pressure and the internal pressure represents the
total pressure that would be exerted by an ideal gas.
The ideal-gas equation of state is based on the assumption that the volume
of the molecules is negligible. Hence, the free volume (volume free of
molecules) equals the entire volume, which, per unit mass, is v-. In addition, it
assumes that the pressure exerted by the molecules with no intermolecular
attractions equals the observable pressure, p. To deal with actual gases, van
der Waals introduced a term b which is subtracted from the total volume to
obtain the free volume. He also introduced the term alv-2 which, when added
to the observed pressure, indicates the total molecular pressure. Thus, the van
der Waals equation of state becomes

(p + ^2)^ ~ k) = RqT (3-12)

Values of the constants a and b depend on the kind of gas and the units
used in Eq. 3-12. When the pressures are expressed in bars, the volume in
cubic meters per kilogram mole, and the temperature in degrees Kelvin, R0
has a value of 0.08314 bar-m3/kg mole-K. Using these units, the values of a
and b are given for various gases in Table 3-2.

Example 3-4. Determine the specific volume of steam at a pressure of


20.0 MPa and a temperature of 600° C:

(a) Using the ideal-gas equation.


(b) Using van der Waals equation.

Solution

(a) From Eq. 3-9,

v _ NR0T _ 1 x 8.314 x (600 + 273)


= 36.29 x 10 5 m3/g mole
p 20 x 106
= 362.9 cm3/g mole

362.9
v- = = 20.16 cm lg Answer
18

This conclusion is based on the assumption that the molecules behave as spheres and is
equivalent to stating that the internal pressure is inversely proportional to the square of the
distance between molecules. In reality, molecules may be very complex in their structures. There
is some evidence to indicate that the internal pressure may be inversely proportional to the
seventh to ninth power of the distance between molecules. This evidence, of course, explains the
serious errors in the predictions by the use of the van der Waals equation under certain
conditions.
Accuracy of the van der Waals Equation 71

Table 3-2
Constants for van der Waals Equation

Gas aa 6a

Air 1.361 0.0367


Argon 1.369 0.0386
Carbon dioxide 3.659 0.0429
Carbon monoxide 1.484 0.0396
Helium 0.0339 0.0232
Hydrogen 0.248 0.0267
Methane 2.283 0.0427
Nitrogen 1.369 0.0386
Oxygen 1.385 0.0319
Sulfur dioxide 6.844 0.0568
Water vapor 5.538 0.0305

aThe units of a are bars (m3/kg mole)2 and those of b


are nr/kg mole.

(b) van der Waals equation, Eq. 3-12, is more easily solved by trial and error.
When v = 0.314 m3/g mole, van der Waals equation becomes

(200 + 3^)(0.314- 0.0305) = 0.08314x 873

or

72.6 = 72.6

Then

0.314 x 106
specific volume = 17.44 cm3/g Answer
18 x 103

The true specific volume of steam as obtained from the Steam Tables3 for
these conditions is 18.18 cm3/g. It should be noted that the volume calculated
by the use of the van der Waals equation, while not very accurate, is much
closer to the true volume than the volume calculated by use of the ideal-gas
equation.

3-6 ACCURACY OF THE VAN DER WAALS EQUATSON


A line of constant temperature, for an ideal gas, obeys the law pv = C. A
series of such lines is shown on a p-v plane for steam as an ideal gas in Figure
3-14. In contrast with this, Figure 3-15 shows isothermal lines for steam, using
the actual values from the Steam Tables. It should be noted that, at low

3Steam Tables will be discussed in Chapter 9.


72 Ideal and Actual Gases

pressures and high temperatures, there is good agreement between the two
sets of curves; whereas, at high pressures and low temperatures, the ideal-gas
prediction is almost valueless. The van der Waals isothermal curves (shown in
Figure 3-16), although far from accurate, are a distinct improvement over
those for an ideal gas.
In addition to the isothermal lines shown in Figure 3-15, a liquid line and a
vapor line also are shown. The liquid line is the locus of the points represent¬
ing a boiling liquid. The vapor line is the locus of the points representing the
vapor formed from the boiling liquid. Since the phase change from a liquid to
a vapor occurs at constant temperature when the pressure is held constant,
isothermal curves are horizontal in this region. Note the 280° C line in Figure
3-15. As the pressure is increased, the volume change during vaporization
becomes smaller and becomes zero at a pressure of 220.9 bars, the critical
pressure. At this point the temperature is 374.14°C. (This is the critical
temperature.)
The van der Waals isothermal curve drawn at the critical temperature of
374.14°C (Fig. 3-16) shows a point of inversion, and hence, recognizes the
Accuracy of the van der Waals Equation 73

FIGURE 3-15 Actual isothermal lines for steam.

existence of a critical point, which the ideal-gas law cannot. However, at


temperatures below the critical temperature, the van der Waals equation does
show an irregularity during the phase change but does not show that a line of
constant temperature is also one of constant pressure in this region. See the
280°C line in Figure 3-16.
There is a definite relationship between the constants a and b in the van der
Waals equation and the critical pressure, temperature, and specific volume.
Since the critical point is a point of inversion of an isothermal curve on the
p-v- plane at this point (dpldv)T = 0 and (d2p/d^2)T =0. The van der Waals
equation may be arranged as follows:

R0T a
(3-12a)
v- — b v-2

Then if the subscript c denotes the critical.

RqTc
=0
T Wc-b)2
74 Ideal and Actual Gases

or

RoTr = (vc ~ b)‘ (3-13)

Also,
d2p\ _ 2RqTc 6a
0 (3-14)
sd v-2) T (v-c — b)2 v-
or

3a(v-c ~ by
RqTc — (3-14a)

When Eqs. 3-13 and 3-14 are combined, the result is

vc = 3b (3-15)

Substituting this value of v-c in Eq. 3-14 gives

T _ 8a
(3-16)
c~ 21 R0b
Accuracy of the van der Waals Equation 75

Also, substitution of these values of vc and Tc in Eq. 3-12 gives

a
Pc = (3-17)
21b

Critical properties are available for most substances, and hence, it is


common practice to use these properties, together with Eqs. 3-16 and 3-17, to
determine constants a and b. Values of these constants for common gases are
given in Table 3-2. Critical properties of various gases are given in Table 3-3.
The van der Waals equation, although satisfactory for many conditions of
temperature and pressure, is quite inaccurate in the region around the critical
state. Substituting for the critical properties their values in terms of a and b
yields
PcV-c _=a (3b) 3r
(3-18)
Tr 27i>2(8a/27R0f>) 8 0
or

Pc V-c 3
rTt 8

Table 3-3
Critical Properties3

Critical Critical Critical


Press Temp. Vol.
Substance Formula (bars) (K) (cm3/g)

Air 37.72 133.0 2.86


Alcohol, ethyl (grain) c2h6o 63.85 516.3 3.63
Alcohol, methyl (wood) CH40 79.71 513.2 3.69
Ammonia nh3 112.80 405.5 4.26
Argon A 48.75 151.2 1.88
Benzene c6h6 49.23 562.6 3.33
Carbon dioxide co2 73.91 304.2 2.14
Carbon monoxide CO 34.96 133.0 3.32
Helium He 2.29 5.26 14.60
Hydrogen h2 12.98 33.3 33.03
Methane ch4 46.40 190.7 6.19
Nitrogen n2 33.92 126.2 3.22
Octane c8h18 24.96 569.5 4.26
Oxygen o2 50.40 154.4 2.32
Sulfur dioxide so2 78.88 430.7 1.91
Water vapor h2o 220.89 647.3 3.16

aThe values for water vapor are based on values in Steam Tables, by J. H. Keenan, F. G. Keyes,
P. G. Hill, and J. G. Moore, Wiley New York, 1978. The properties of all other substances, except
air, were derived from Generalized Thermodynamic Properties of Pure Fluids, by Lydersen,
Greenkorn, and Hougen, Engineering Experiment Station, University of Wisconsin. We believe
that these values for the critical pressure and temperature are reliable. However, those for the
critical volume are subject to some question.
76 Ideal and Actual Gases

Table 3-4
PvT Relationships at the Critical

PcvJTcRo

Air 0.284
Ammonia nh3 0.243
Argon A 0.292
Carbon dioxide co2 0.275
Carbon monoxide CO 0.294
Helium He 0.306
Hydrogen h2 0.309
Nitrogen n2 0.291
Oxygen 02 0.291
Water vapor h2o 0.232

Actual values of pcvJTcR0, as shown in Table 3-4, differ significantly from the
value of | or 0.375.

3-7 OTHER EQUATIONS OF STATE


In addition to the van der Waals equation of state, some of the more common
ones are shown in Table 3-5.
In general, the more complicated equations are more accurate over a larger
range of temperature and pressure than are the simple equations. However, in
certain problems the more complex equations are difficult to use. For in¬
stance, it will be shown in Chapter 8 that in dealing with the general equations

Table 3-5
Equations of State
Name Equation Constants

van der Waals (p+fy(v-b) = R0T a, b


Clausius pO - b) = R0T b
R0T acp 1
Callender p fi-1 T" a, n
_ R°T -alBT„
Dieterici V L C a, b
v- — b
R0T a
Berthelot a, b
p v-b Tv2
Beattie-Bridgeman

-$(-!) a, A0, b, B0, c


Virial Equations of State 77

Table 3-6
Constants for the Beattie-Bridgeman Equation

A0 a Bo b c x 10"4
/atm-liters6'\ / liters3 \ / liters3 \ / liters3 \ /liters3-K\
V g mole2 ,/ Vgmole/ Vg mole/ \g mole/ V g mole )

Air 1.302 0.0193 0.0461 -0.0110 4.34


o2 1.491 0.0256 0.0463 0.0042 4.79
h2 1.344 0.0262 0.0504 -0.0069 4.19
CO 1.344 0.0262 0.0504 -0.0069 4.19
C02 5.009 0.0714 0.1048 0.0724 65.96
He 0.0218 0.0598 0.0140 0 0.004
h2 0.197 -0.0051 0.0210 -0.0436 0.050

Note: When these units are used, the value of R0 is 0.08206 atm-liter/g moie-K.

of thermodynamics, it is necessary to perform such partial differentiations as


(3p/dT)v and (dWdT)p. It is suggested that in such cases, a determination be
made as to whether or not a simple equation of state is sufficiently accurate
for the given conditions, before attempting to use a more complicated
equation.
One of the most widely used equations is the Beattie-Bridgeman equation.
The constants for this equation are listed in Table 3-6 for some common
substances.

3-8 VIRIAL EQUATIONS OF STATE


A different type of equation of state is the virial equation. An examination of
Figures 3-1 and 3-2 shows that the relationship between pressure, volume, and
temperature may be expressed as

pn - RT 1 H-I—^ ; + • • ,N] (3-19)


\ v- v-~ v- )

The constants B, D, D, and so forth, are known as the virial coefficients. These
coefficients are a function of temperature. B is called the first virial
coefficient, C the second, and D the third, and so on. Equation 3-19 may also
be expressed as

pv = RT(\ + B'p + C'p2 + D'p3+ • • •) (3-19a)

Equations 3-19 and 3-19a are known as the virial equations of state.
Values of the virial coefficients as functions of temperature can be deter¬
mined from experimental data. The virial coefficients also can be deduced
from a study of the intermolecular forces. It might appear that the analysis of
the intermolecular forces would be quite easy and that these forces should be
proportional to the mass of the molecule and inversely proportional to the
78 Ideal and Actual Gases

square of the distance between the molecules. In reality there is no simple


relationship by which the intermolecular forces may be evaluated.
Under some conditions, it appears that the intermolecular forces vary in¬
versely as the seventh and ninth power of distance between the molecules.
When the molecules are extremely close together, the attractive force is
replaced by a repelling force. Nevertheless, extensive investigations of these
forces have been made and from these investigations the virial coefficients
have been deduced for several gases. This analysis is beyond the scope of this
text.

3-9 CORRESPONDING STATES AND REDUCED


PROPERTIES
At any given condition of temperature and pressure there may be a wide
deviation in the behavior of actual gases from the ideal-gas laws. Under some
conditions, the compressibility factor of one gas may be several times that of
another gas. However, when the states of various gases are expressed in terms
of their respective critical temperatures and pressures rather than in absolute
values, then there is a fairly good correlation between their deviations from
the ideal-gas law. The van der Waals equation of state is helpful in demon¬
strating this fact.
In this connection, it is desirable to introduce the concept of reduced
properties. A reduced property is defined as the ratio of the property in a
given state to the value of the property at the critical state. Using the
subscript r to signify a reduced property, pr = p\pc, Tr = T/Tc, and v-r = W^c,
where the subscript c refers to the critical state. Substituting these values into
Eqs. 3-17, 3-16, and 3-15, respectively,

op
27 b

8a
T = Tr
27bR0

v- — 3b v

Substituting these values into Eq. 3-12,

(a , a .v 8 aR0T
(r^Pr + 9F^)(3b^-b) = 2TbRir (3-20)

or

(Pr + Wr ~ I) = f Tr
According to Eq. 3-20, all gases existing at the same reduced temperature
and pressure must have the same reduced volume. However, as noted earlier,
no gas obeys the van der Waals equation exactly, the deviations being the
Corresponding States and Reduced Properties 79

largest near the critical state. If, however, all gases deviate from the van der
Waals equation in the same amount, then gases existing at the same reduced
temperature and reduced pressure should have the same reduced properties in
addition to those of pressure and temperature. A careful examination of
various properties of various gases shows that although this statement is not
precisely true, it is approximately so.
When two gases exist at the same reduced temperature and reduced
pressure, they are said to be in corresponding states. Although various
properties have been determined for some substances, such as steam, over a
wide range of temperature and pressure, properties of most substances are
known only rather sketchily and then generally over a rather restricted range
of temperature and pressure. It should be recognized that it is extremely
difficult to determine certain properties of a substance with a high degree of
accuracy. And yet it is desirable to have available the thermodynamic
properties of the very large number of substances that are being used today.
It will be shown in Chapter 8 that when an accurate p-^-T formulation has
been obtained for any substance, it is possible to determine the changes in
other properties, such as internal energy, enthalpy, and entropy with a high
degree of accuracy. However, when an accurate p-^-T formulation is not
available, the concept of corresponding states may be used to obtain a good
approximation of the desired properties.

Example 3-5. What is the boiling temperature of ammonia at a pressure of


320 kPa? Assume that tables of ammonia are not available but those of steam
are available.

Solution. At a pressure of 320 kPa,

"• -f;- na' 002837

(The critical pressure of ammonia is obtained from Table 3-3.) Steam in a


corresponding state has a pressure equal to

PrPc — 0.02837 x 220.89 = 6.267 bars or 0.6267 MPa

From the tables of properties of steam, the boiling temperature at this


pressure is 161.1°C. Under these conditions, the reduced temperature of the
steam,

161.1 + 273.15
0.671
647.3

Ammonia, at this reduced temperature of 0.671, should have a boiling tem¬


perature,

T = TrxTc = 0.671 x 405.5 = 271.7 K or - 1.45°C

(The actual boiling temperature of ammonia at a pressure of 320 kPa is 8.9°C.)


80 Ideal and Actual Gases

Example 3-6. What is the specific volume of carbon dioxide at a pressure of


70 bars and a temperature of 47°C?

Solution

p = 70
0.947
pc 73.91

For steam, when pr = 0.947,

P - PrPc = 0.947 x 220.89 = 209.2 bars or 20.92 MPa


, _ 47 + 273
1.052
r 304.2

For steam, when Tr = 1.052,

T = TrTc = 1.052 x 647.3 = 680.9 K

For steam, when p = 20.92 MPa and T = 0.80.9 K, v- — 9.68 cm3/g. Then

9.68
3.083
3.14

For carbon dioxide,

v- = v-rv~c = 3.083 x 2.14 = 6.598 cm3/g Answer

Note: The actual volume of carbon monoxide at these conditions is


5.59cnr/g. Although the calculated volume is not very close to the
actual volume, it is much closer than the volume calculated by assum¬
ing an ideal gas, which is 8.637 cm3/g. The conditions chosen here are
very close to the critical state where the deviations should be expected
to be much larger than for other conditions. It should be noted from
Table 3-4 that the values of pcvclR0Tc for steam and carbon dioxide
are not in good agreement.

The concept of corresponding states may be applied to the determination of


compressibility factors. Some of the most extensive work on the com¬
pressibility factor was done by Leonard C. Nelson and Edward F. Obert. The
results of this investigational work were published in the Transactions of the
ASME.4 Two of their diagrams, Figures 3-17 and 3-18, reproduced by per¬
mission, are in the pocket at the back of this book. Nelson and Obert
investigated 30 substances having widely differing properties, in preparing
their compressibility-factor charts. They have shown in their charts the
regions in which the compressibility factor for many substances should be
correct within 1 percent at low pressures and within about 5 percent for very
high pressures.
From Eq. 3-20, it may be seen that the van der Waals equation is of the

4L. C. Nelson and E. F. Obert, “Generalized p-v-T properties of Gases,” Transactions of the
ASME, October 1954, pp. 1057-1066.
r-H
o
o o
o
CO
o
00
o

ad
,/J/
% -=
d
r'-
o
CD
o
d

‘jopej A^iiqissajdiuoo
ID
o
d
o
■sT
o
o
oo
o
o
CO

o
o
o
0-1
oo
d
O'

o
CD
r^.

o
o
LO
o
d

Reduced Pressure, p
'r

FIGURE 3-17 Compressibility factors for reduced pressures 0 to 1.0. (Courtesy of Leonard C. Nelson and Edward F.
Obert.)

81
82
Reduced Pressure, p

FIGURE 3-18 Compressibility factors for reduced pressures 0 to 10.0. (Courtesy of Leonard C. Nelson and Edward F. Obert.)
Corresponding States and Reduced Properties 83

following form:

f(Pn Vn Tr) = 0
It has been found experimentally that for a wide range of temperatures and
pressures, there is a serious error in this type of equation. The curves of the
compressibility chart, on the other hand, are based on the following relation¬
ship:
Z = f(Pn Tr)
Since Nelson and Obert found that over an extensive region of temperature
and pressure, the compressibility factors for various substances did not
deviate by more than a small percentage, it follows that this equation is valid
for a wide range. It should not be inferred, however, that the equation is valid
over the entire range of temperatures and pressures, since there is evidence to
show that it is in considerable error in regions near the critical conditions.
In these charts, Nelson and Obert used a pseudo-reduced volume which is
defined as follows:
= v = ZRTIp = 7Tl
RTJpc RTJpe Z pr
Since Z = f(pr, Tr),

= f'(pn Tr)
The establishment of this relationship enabled them to plot lines of constant
v-r' on their charts.
When the compressibility factor is substantially unity, a substance may be
treated as an ideal gas when its p-v-T relations are considered. This is no
way implies that the specific heat is a constant. It is possible for the specific
heat to vary greatly in regions of high pressure and moderate temperature
where the compressibility factor is close to unity. The use of Figures 3-17 and
3-18 can be shown by the following examples.

Example 3-7. Using the compressibility-factor charts, determine the specific


volume of the carbon dioxide in Example 3-6.
Solution. From Figure 3-17, when pr = 0.947 and Tr = 1.052, the com¬
pressibility factor is 0.64. Then the specific volume.

ZJR0T 0.64x8.314x320
v- =- = 24.32 x 10 3m3/g mole
P 70 x 10s
or
24.32 x 10“5x 106
V- = 5.528 cm'/g Answer
44

Note: 1 bar = 105 N/m2.

Example 3-8. A tank having a volume of 3.8 m3 contains the refrigerant F-12,
at a pressure of 8 bars and a temperature of 60°C. Determine the mass of
84 Ideal and Actual Gases

refrigerant present. The critical pressure is 41.155 bars and the critical
temperature is 112°C.

Solution.

8
0.1944
41.15
= 333
0.864
- 385

From Figure 3-17, the compressibility factor, Z, =0.88


pV = ZNRqT
8 x 105 x 3.8
1248 g moles
0.88x8.314x333

F-12 is CC12F. Its molecular weight = 120.9.

m = 1248 x 120.9 = 150,880 g = 150.88 kg Answer

From tables of properties of this refrigerant, for the given conditions, the
specific volume is 25.25 cnr/g. Then the true mass of the F-12 =
(3.8 x 106)/25.25 = 150,500 g or 150.5 kg.

A fair degree of accuracy can be obtained by assuming that the com¬


pressibility factor is a function only of the reduced pressure and the reduced
temperature. To obtain still more accuracy, investigations have been made as
to the possible effects of other parameters on the value of the compressibility
factor. In particular, the critical compressibility factor shows promise as an
additional parameter that will give a higher degree of accuracy. However,
additional investigations must be made before final conclusions can be drawn.
Certain thermodynamic properties of gases, such as thermal conductivity
and viscosity, are very difficult to measure accurately. The concept of
corresponding states has been used to a limited extent to approximate these
properties when accurate experimental data are not available. In general,
although the concept of corresponding states has been utilized to an apprecia¬
ble extent in predicting the various thermodynamic properties of substances
when accurate data are lacking, the full usefulness of this concept has not
been completely realized.

3-10 SUMMARY
In this chapter, the p-v-T relationship has been developed for ideal gases.
Methods have been presented to adjust this relationship for actual gases. In
addition, the concept of corresponding states has been utilized to determine
the properties of actual gases. However, when the properties of well-known
vapors are desired, it is more convenient to refer to tables of their properties.
A discussion of the properties of vapors has been delayed until Chapter 9.
One very important property of vapors is entropy, which is normally included
in tables of vapor properties. Hence, the discussion of vapor properties has
been delayed until the concept of entropy is developed.
Problems 85

PROBLEMS

3-1 A tank having a volume of 1.32 m3 contains nitrogen under a pressure


of 6482 kPa at a temperature of 35°C. Determine the mass of nitrogen
present. Can the nitrogen be treated as an ideal gas under these
conditions?
3-2 Air is removed from a power plant condenser at a rate of 1.2 nr/min
(measured at 102.2 kPa and 31°C). Determine the mass flow rate of the
air.
3-3 The air within the condenser of Problem 3-2 is at a pressure of 0.62 kPa
and a temperature of 28°C. Determine the volume rate of air flow
through the condenser.
3-4 A power plant furnace burns coal at the rate of 108,200 kg/h. Air at
100.8 kPa, 28°C is supplied at the rate of 13.8 g/g coal. Determine the
volume rate of air flow in nr/min.
3-5 An unknown gas having a mass of 1.36 kg occupies a volume of
0.628 m3 at 198 kPa, 35°C. Determine the gas constant of the gas.
3-6 Determine the molecular weight of the gas in Problem 3-5.
3-7 A jet engine uses 42 g of air per gram of fuel. The rate of fuel flow is
2420 kg of fuel per hour. The air is at 8.26 kPa, -35°C. Determine the
rate of air flow in m3/min.
3-8 Air exists at 30°C, 4200 kPa. Determine the specific volume by (a)
ideal-gas law, (b) by the van der Waals equation Suggestion: Use trial
and error.
3-9 The same as Problem 3-8 but at -85°C, 4200 kPa.
3-10 Solve Problems 3-8 and 3-9 using the compressibility factors based on
reduced properties. Compare the results.
3-11 The steam in the condenser of Problem 3-3 has a pressure of 3.78 kPa.
Determine its specific volume.
3-12 Determine the specific volume of steam at 40 bars, 275°C by (a)
ideal-gas law, (b) van der Waals equation, (c) use of compressibility
factor. Note: The true value of the specific volume is 54.47 cm3/g.
3-13 Using the critical values of pressure and temperature for steam as given
in Table 3-3, calculate the constants a and b in van der Waals equation.
Compare with values given in Table 3-2.
3-14 The same as Problem 3-13 except use the critical volume and tem¬
perature. Explain why the constants thus obtained differ from those in
Problem 3-13.
3-15 Determine the specific volume of air at 40bars, -20°C using (a)
ideal-gas law, (b) van der Waals equation, (c) Beattie-Bridgeman equa¬
tion, (d) compressibility factor.
3-16 An ideal gas occupies a volume of 2.209 m3 under a vacuum of 16.61 cm
of mercury. The barometric pressure is 74.47 cm of mercury. The gas
86 Ideal and Actual Gases

temperature is 32.22°C. Determine the number of moles of gas present.


Can the mass of the gas be determined?
3-17 An automotive engine uses 10.2 liters of gasoline per hour. The density
of gasoline is 0.67 g/cm3. The engine uses 13.2 g of air per gram of fuel.
Air is supplied at 101 kPa, 30°C. Determine the volume rate of air flow
in rrr/min.
3-18 Using the concept of corresponding states and the properties of steam
in the Appendix, determine the saturation (boiling temperature) of
carbon dioxide for a pressure of 62.4 bars.
3-19 Using the concept of corresponding states, determine the specific
volume of carbon dioxide vapor for Problem 3-18.
3-20 Using the compressibility factor, determine the specific volume of
carbon dioxide vapor for Problem 3-18.
3-21 Using Table 3-3, determine the critical compressibility factors for: (a)
air, (b) carbon dioxide, (c) hydrogen, (d) water vapor. Compare with
Table 3-4.
3-22 (a) Determine the reduced pressure and temperature for ammonia
boiling at a pressure of 2.5 bars. (See the Appendix for properties of
ammonia.)
(b) Determine the actual pressure of steam having the same reduced
pressure as that of the ammonia.
(c) Determine the reduced temperature of steam at its actual pressure in
part (b), using the properties of steam in the Appendix.
(d) Compare this reduced temperature with that of part (a).
■91;

_ 4 _

CHANGES IN STATE
OF GASES

4-1 EQUILIBRIUM AND REVERSIBILITY


In Chapter 3 we dealt with gases in a given state of existence. Although it is
important to determine the interrelationships among the various properties of
a substance in such a given state, it is much more important to understand the
changes in properties that accompany changes in state. In Chapter 1, we
discussed processes from a general viewpoint, paying little attention to the
internal aspects of the substance undergoing a change in state. It is essential
that we consider such aspects so that the change in state can be described
fully and all energies involved can be evaluated correctly.
A process taking place in a system at rest can be described completely if
the state of the system can be described at all times during the process. The
state of a system may be described if it is in a state of equilibrium. A system
is said to be in a state of equilibrium when no unbalanced forces exist that
can change its state of existence. Four possible internal forces that may
change the state of existence are: (I) mechanical, (2) thermal, (3) electrical,
and (4) chemical. Each and all of these forces within a system must be in
complete balance in order for the system to be in thermodynamic equilibrium.
For instance, consider an isolated system consisting of a mixture of hydrogen
and oxygen. If allowed to stand for a sufficient length of time, there will be no
pressure or temperature variations throughout the system. There may be no
electrical forces acting on the system. However, since a very small spark will
cause a radical change in state, the oxygen and hydrogen cannot be in
chemical equilibrium. Hence, the system is not in a state of complete
thermodynamic equilibrium. The application of a very small force may cause
large displacement of the system. For an unstable system, the removal of the
force that caused the displacement of the system will not permit the system to
return to its original state. Consider the block in Figure 4-1. In its given state,
the block is in an unstable equilibrium condition and the application of a small
force on the left face of the block will cause a permanent displacement of the
block. The block in Figure 4-2 is in a stable equilibrium condition. Although
the application of a horizontal force will cause a displacement of the block,
when the force is removed, the block returns to its original position. In most

87
88 Changes in State of Gases

problems that will be encountered here, a system of equilibrium is in stable


equilibrium, and hence, only stable equilibrium will be considered in this
discussion.
The sudden application of a large force to a system causes instantaneous
changes in the system and may cause a considerable departure from equili¬
brium conditions. Likewise, the removal of a restraining force that has kept a
system at a condition differing greatly from that of the surroundings may also
cause a considerable departure from equilibrium conditions within the system.
Consider, for example, that a gas is contained in the vertical, insulated
cylinder shown in Figure 4-3. Assume that the system is in equilibrium.

FIGURE 4-3 Stable equilibrium.


Equilibrium and Reversibility 89

FIGURE 4-4 Uncontrolled expansion (sud¬


den removal of weight).

Because of the large weight on the piston, the pressure of the gas in the
cylinder is high. Assume that the piston is weightless and without friction. In
the ideal case, the weight may be removed instantaneously and placed in the
position shown in Figure 4-4. Since the piston is weightless, it will move
upward instantaneously. This will cause the pressure just beneath the piston
to drop, also instantaneously. Although a pressure wave will travel throughout
the gas at the acoustic velocity, tending to equalize the pressure in the cylinder,
this is a time-consuming affair. Likewise, there will be nonuniformity of
temperature throughout the gas during the expansion. Since equilibrium does
not exist during the expansion process, the state, and hence, the properties of
the gas cannot be specified during piston motion. After the piston reaches the
end of its travel, equilibrium will be reached rather shortly, and hence, the
final state may be specified. The initial and final states, 1 and 2, are shown on
a pressure-volume diagram, Figure 4-5a. The dashed line connecting states 1
and 2 indicates that the intermediate states and, hence, the path cannot be
specified.
The system can be restored to its original state only by bringing in work
from outside the system to compress the gas. Then heat must be removed
from the gas to bring the temperature down to its original value.
In Figure 4-6 the large weight has been replaced by a large number of small
weights. Assume that each weight is removed instantaneously and slid
horizontally to a niche in the wall of the cylinder. Allow time for the system
to reach equilibrium before another weight is removed. Thus, a series of
equilibrium states may be specified for the expansion process. This is shown
in Figure 4-5c. As the weights become smaller and smaller, more and more
intermediate states may be specified. In the limiting condition with an infinite
number of weights, all intermediate states may be specified, as in Figure 4-5d.
At any time, the expansion process can be stopped and the gas can be
compressed to its original state by first exerting a sufficient force on the piston
90 Changes in State of Gases

FIGURE 4-5 Various expansion processes (see Figures 4-4 and


4-6). (a) One weight, (b) Two weights, (c) Four
weights, (d) Limiting conditions (infinite number
of weights).

so that it will move down to permit the top weight to be slid horizontally onto
the piston. Then, progressively, each weight is moved back onto the piston.
The only work required to be brought in from the outside is the work required
to depress the piston sufficiently to permit sliding the top weight onto it. As
the number of weights is increased, the work required to be brought in from
the surroundings to restore the system to its original state becomes progres¬
sively smaller and approaches zero as the weights approach zero in size.
As the weights become smaller and smaller in size, the deviations from
equilibrium conditions also become smaller. In the limit, when the weights

Many small
weights

FIGURE 4-6 Controlled expansion


(weights removed one
at a time).
Equilibrium and Reversibility 91

become zero in size, equilibrium conditions are maintained throughout the


expansion process. If this happens, the process is said to be a reversible
process. An internally reversible process is one in which equilibrium is
maintained throughout the process, and hence, no energy is required to be
brought in to restore the system to its original state.
If the system is in complete thermodynamic equilibrium, there are no
unbalanced forces to cause a change of state of the system. However, an
infinitesimal unbalanced force will cause a change of state to take place.
Under these conditions, the amount of energy required to be supplied to restore
the system to its original state is infinitesimal, and deviations from equilibrium
are also infinitesimal. Such a process has been designated as a quasistatic
process.
A very important fact should be emphasized about irreversible processes.
Let us go back to the piston with the single weight. Because the gas existed at
a pressure higher than that of the surroundings, it had the capacity of doing
work. When the weight was removed, the pressure of the gas reached that of
the surroundings immediately without doing any useful work.1 On the other
hand, when the weights became infinitesimal in size, the maximum amount of
work was performed by the expansion of the gas. Hence, it may be said that
in a reversible process there is no loss of the capacity for doing work.
Although an adiabatic expansion process was considered here to illustrate
the concept of reversibility, this concept can be applied equally well to all
processes.
When the second law of thermodynamics is fully developed (Chapter 5), it
will become possible to explore the concepts of equilibrium and reversibility
in still greater depth.
Consider a system existing at a temperature much less than that of its
surroundings. In the absence of appreciable thermal insulation, there will be a
heat flow into the system, tending to bring its temperature up to that of the
surroundings. Heat will not, of its own accord, flow back from the system to
the surroundings. However, if the system is gaseous, it may be compressed,
thus increasing its temperature to above that of the surroundings. Heat then
may flow from the system back into the surroundings. Work had to be brought
into the system to cause this reversal of heat flow. This amount of work is
minimized when the original temperature difference between the system and
its surroundings is minimized. It is evident then, that heat-flow processes
occurring with finite temperature differences are irreversible. As the tem¬
perature differences approach zero, these processes approach reversibility,
with the amount of work required to return the system and its surroundings to
their original states also approaching zero.
No process, either natural or man-made, is truly reversible since it is always
necessary to bring in some energy to restore the system to its initial state.

'The atmosphere is pushed out of the way by the piston as it moves upward. This, in itself, is not
a normal desired effect and, hence, is not classified as useful work.
92 Changes in State of Gases

Although all actual processes are irreversible, there are various degrees of
irreversibility. Consider the gas in the cylinder in Figure 4-4. When the weight
on the piston is removed instantaneously, the gas expands without doing any
work. Hence, the maximum amount of work is required to restore the gas to
its original state. We may say that this process is completely irreversible.
Now consider the piston with a large number of small weights on it in Figure
4-6. As the weights are removed, one at a time, the gas does some work in
raising the remaining weights. Hence, less work is required to be supplied
from an external source to restore the gas to its initial state. We may say that
such a process is only partially irreversible.
Many irreversible processes take place at a sufficiently low rate that it is
possible to specify at least some of the properties during the process. For
example, assuming that air expands slowly in a cylinder and does work on a
piston as heat is added to it. The process is irreversible since there will be
friction between the piston and the walls and also because there is a
temperature difference throughout the air. Nevertheless, because of the
slowness of the expansion, it is possible to specify the pressure in the cylinder
at all times. In fact, an engine indicator may be used to plot the pressure as a
function of the piston position. A knowledge of the pressure variation permits
the determination of the work delivered to the face of the piston during the
process or any part of the process. Assuming that the mass of air present is
known, the average temperature during any part of the expansion may be
calculated. Knowing the temperature change facilitates the calculation of the
internal energy change. This internal energy change plus the work delivered
from the system will equal the heat added.
Most of our devices that involve some type of a fluid are, in the overall
sense, open systems. However, many of these devices operate as closed
systems for very short periods of time. This is particularly true for recipro¬
cating devices. For example, consider a reciprocating air compressor. During
the compression process, both valves are closed, and hence, the compression
process is a closed system process. Even for devices that are not reciprocat¬
ing, it is possible and, frequently, desirable to analyze the change of state of
the fluid as it flows through the device. Consider a centrifugal air compressor.
Air enters the compressor more or less at constant pressure, is compressed,
and then flows out of the compressor, again at more or less constant pressure.
Thermodynamically, these three processes for the centrifugal compressor are
similar to those taking place in a reciprocating compressor. Thus, the analysis of
various types of processes is in no way limited to those processes taking place
within a system that remains closed at all times.

4-2 POLYTROPIC PROCESSES


Various processes have been discussed in which one property, such as
temperature, pressure, or volume, remained constant during the process. In
many processes, however, no one property remains a constant. It has been
Polytropic Processes 93

found that when no phase change is involved, the relationship between


pressure and volume of actual gases during a change in state may be
expressed quite accurately as

pVn = C (4-1)

It is assumed that the process is a quasistatic one, with the properties being
specifiable throughout the process. Equation 4-1 also approximates the pres¬
sure-volume relationship for many phase-change expansions and com¬
pressions. The value of the exponent n in Eq. 4-1 depends on the type of the
process. For instance, when n is zero, Eq. 4-1 states that the pressure is
constant. Rewriting Eq. 4-1 as pUnV = C', it may be seen that when n = oo, the
volume is constant. Various processes are shown in Figure 4-7 with the value
of n given for each process. Note that there are an infinite number of
processes that follow the law pVn = C, with n ranging from plus to minus
infinity. Processes that obey the law pVn = C are known as polytropic
processes.

V
FIGURE 4-7 Polytropic processes.
94 Changes in State of Gases

Consider now an ideal gas. From the ideal-gas relationship pV = mRT, it is


evident that
Pi V, _ p2V2
(4-2)

From Eq. 4-1,

(4-la)

Comparing Eqs. 4-la and 4-2, it may be seen that when n = 1, T, = T2, or that
when n = 1 for an ideal gas, the temperature must be constant. Since Eq. 4-2
is not valid for an actual gas, it follows that there may be a significant change
in temperature for an actual gas undergoing a process in which the value of
the exponent n is unity.
For a constant temperature process, Eq. 4-2 becomes

(piVi)r — (P2^)t (4-2a)


The subscript T denotes that the temperature is constant. Equation 4-2a is a
mathematical statement of Boyle’s law. When the pressure is constant,

a-© (4-2b)

When the volume is constant,

K
r) -9)
1 1/ V \ i 2/ V
(4-2c)

Equations 4-2b and 4-2c are expressions of Charles’ laws.


When the value for p of mJRT/V is substituted in Eq. 4-la,

mRTi xrn mRT2 xzn


V? = V\ (4-2d)
V, ' 1 V, 2
Then
T / y \(n_1)
(4-3)
T2 \Vi
Likewise, substituting the value of V = mRT/p in Eq. 4-la,

/mRT]\n (mRT2\n
\-JT) =P2VJT)
Then

I, =/piy"-1’'"
(4-4)
T2 \Pi)
Equation 4-la may be written as
Evaluation of Work for Polytropic Processes 95

Example 4-1. In a diesel engine, a temperature of 600°C is desired at the end


of the compression process. The compression process obeys the law pV135 =
C. The temperature at the start of the compression process is 60°C. Deter¬
mine the compression ratio required (i.e., the ratio of the volume before
compression to the volume after compression).

Solution. Rewriting Eq. 4-3,

/ V,\ _ /T2\ 1/("~1} _ /600 + 273\ 1/035


15.70 Answer
\vj Viy V 60 + 273 /
4-3 EVALUATION OF WORK FOR POLYTROPIC
PROCESSES
When a simple system expands or is compressed, the work delivered by the
system may be evaluated by the use of Eq. 2-6b. The work is delivered by a
movement of part of the boundary of the system. It is assumed that the
pressure in contact with the boundary is known at all times and varies
according to Eq. 4-1.
Substituting for p from Eq. 4-la in Eq. 2~7a,
dV
SW = C — (4-5)

When the system changes from state 1 to state 2, the integration of Eq. 4-5
yields
CV'~nl2 pVnVx~n pV l2
(1 ~ n)Ji (1 - n) (1 - n)Ji
or

P2V2-P1V1 (4-6)
1-n

Equation 4-6 is valid for all polytropic processes when the p-V variations are
known at the face of the work-delivering portion of the boundary of the
system. However, when n is unity, Eq. 4-6 becomes indeterminate. For this
special case, when pV = C, Eq. 4-5 becomes

dV
8W = C(4-5a)

When the system moves from state 1 to state 2, integration of Eq. 4-5a yields

W,.2=Cln(^) = pVln(^) (4-6a)

Equation 4-6a is valid for all processes in which the pV product is a constant,
that is, when the exponent n equals unity. For the special case of an ideal gas,
Eq. 4-6a gives means of evaluating the work done in an isothermal process,
since n is unity in this case.
96 Changes in State of Gases

■)

Example 4-2. The original volume in Example 4-1 is 32,000 cm and the
original pressure is 0.96 bars. Determine the work of compression.

Solution. The work can be determined by use of Eq. 4-6.

P2 = p, (y-) = 0.96(15.7)'35 = 39.51 bars = 395.1 N/cm:

V, 32,000
V2 = = 2038 cm
V,/V2 15.7
p2V2-p.Vi _ 395.1 x2038-9.6x 32,000
Wx.2 = -1,423,000 N-cm
1 — n 1-1.35
-14,230 N-m
—14,230 J Answer

The negative sign indicates that there is a work input during the com¬
pression process.

Example 4-3. Repeat Example 4-2, using English units and the pressures in
Example 4-2.

Solution

original volume = 32,000x3.531 x 10~5 = 1.124 ft3

final volume = yjy = 0.0716 ft3

original pressure = 0.96 x 14.504 = 13.92 psia


final pressure = 13.92(15.7)'15 = 572.9 psia
, 144(572.9x0.0716- 13.92 x 1.124)
work =-—r?- -10,440 ft-lbf

Answer
Example 4-4. A refrigerant expands in a cylinder from a volume of 1950 cm3
to a volume of 9800 cm3, the pressure decreasing from 9.8 bars to 1.2 bars.
Determine the work done.

Solution. Since P\Vn\ = p2V2,

Pi _ (Yi\
P2 WJ

or

9.8
(^5)" 8.17 = 5.026", n = 1.301
1.2 \1950/ ’
work = PiYpPjV, 1.2 x 10 x 9800 - 9.8 x 10 x 1950
= 240,000 N-cm
1—n 1-1.301
= 2400 J Answer
Note: 1 bar = 10 N/cm2.
Evaluation of Work for Polytropic Processes 97

FIGURE 4-8 p-V diagram indicating work.

A pressure-volume diagram has been used to show changes in these


properties during both compression and expansion processses. This type of a
diagram has an additional use. Refer to Figure 4-8.
In this figure, the crosshatched area equals p dV, but from Eq. 2-7a,
p dV = 8W. Thus, the crosshatched area on the p-V plane equals the
increment of work done during the incremental change in volume dV. The
total area beneath the curve 1-2 represents the work done during the expan¬
sion process.
It is assumed that the expansion in Figure 4-8 approaches a quasistatic one.
This process is repeated in Figure 4-9. Now assume that there is irreversibility
occurring in the system as it expands from state 1. State 2' is shown for this
process, which continues until volume V2 is reached. By the very concept of
irreversibility, the work done during the irreversible expansion is less than
work done during the reversible one; or, less of the energy of the system is
transformed into work and removed from the system. Thus, the final energy
of the system for the irreversible process must exceed that of the reversible
process. Hence, the final pressure and temperature also must be higher.
Because of irreversibility, the state of the system cannot be specified during
the irreversible expansion and, hence, cannot truly be shown on the p-v or
any other diagram. To indicate that some sort of a process takes place
between states 1 and 2', a dotted line has been drawn between the two state
points in Figure 4-9. It must be emphasized that this line has no other
meaning. It does not indicate p-v variations during the process. The area
beneath this curve does not represent work. (The area beneath curve 1-2'
exceeds the area beneath curve 1-2, yet the work done during the reversible
process exceeds that for the irreversible process.) In general, state points
98 Changes in State of Gases

FIGURE 4-9 Quasistatic vs irreversible ex¬


pansion.

cannot be specified for irreversible processes. Dotted lines placed on any sort
of diagram connecting the terminal states of any irreversible process do not
have significance. In any case, the areas beneath these curves do not have
any significance whatsoever.

4-4 INTERNAL ENERGY CHANGES


Changes in the internal energy of a system cannot be measured directly but
must be calculated. In general, since u = /(^, T),

•“•-(w ),."*(fi)/* (*7>


For an ideal gas, (du/dv)T = 0. This may be demonstrated in two ways. The
internal energy of a gas is partly molecular kinetic energy and partly molecu¬
lar potential energy. A change in molecular potential energy occurs when the
distance between the molecules is changed, with energy being required to
overcome the intermolecular attractions. For an ideal gas, there is no inter-
molecular attraction, and hence, there will be no change in the molecular
potential energy. If the temperature is held constant, the molecular velocities,
and hence, the molecular kinetic energy can not change. Hence, (du/dv-)T = 0.
A second approach to the evaluation of (du/dv)T is to consider the classical
experiment of Joule. A schematic diagram of the apparatus used in this
experiment is shown in Figure 4-10. Originally vessel A contained air at a
pressure of 22 atm while vessel B was highly evacuated. After measuring the
temperature of the water very carefully with a thermometer that could be
read to 1/200°C, Joule opened the connecting valve between the two vessels.
After allowing time for equilibrium to be reached, Joule again measured the
Internal Energy Changes 99

fi

FIGURE 4-10 Joule’s experiment on in¬


ternal energy.

temperature of the water and found no measurable change in its temperature.


This indicated that there was no energy exchange between the air and its
surroundings. Thus, there could not have been any change in the internal
energy of the air, or, (duldv)T = 0. Therefore, for an ideal gas,

* - © iT
(4-8)

But, from Eq. 2-10, cu = (duldT)». Then for an ideal gas,

du = cv dT (4-9)

Equations 4-8 and 4-9 were derived for ideal gases. As such they are valid for
all possible changes in state of ideal gases, whether with a constant volume or
not.
A more general equation for the change in internal energy will be developed
in Chapter 8, which will be applicable not only to gases and vapor but also to
solids and liquids.

Example 4-5. Determine the change in internal energy and the heat added for
Ex. 4-2. Neglect the effect of temperature on the specific heats. Assume that
air alone is being compressed.

Solution. The mass of air present,

pV (0.96 x 10) x 32,000


RT (0.287x 100)333 g

From Eq. 4-9,

U2 U\ = mcn(T2- T,) or U2- 17, = 32.14x0.718(873- 333) = 12,460 J


Answer
100 Changes in State of Gases

Since Q,_2 = U2 ~ U\+ Wx.2 for a closed system,


Qi.2 = 12,460 - 14,230 - - 1770 J Answer

Note: 1 bar = 10 N/cm\ R = J/g-K = N-m/g-K.

The negative sign indicates that heat was removed during the process.

4-5 ENTHALPY CHANGES


In general, h = f(T, p). Hence,

- (It), dT - (If)t dp <4-'0)


Since h = u T p v-, dh = du + d(p v). Then

(dh\ = /du\ ra(p^)~i


\dp ) j \dp ) j _ dp Jj

For an ideal gas (p^)T = C. Then [d(p v)/dp]T = 0. Following the same line of
reasoning used in the previous section, (duldp)T = 0 for an ideal gas.

M (w),iT
■ «-■■>

In Section 2-9, it was shown that for a closed system at rest, dh = 8q for a
constant pressure process. Since for a constant pressure process 8q = cp dT,

dhp = cp dT (4-12)

or

(4-13)
* - (w\
Substituting into Eq. 4-11,

dh = cP dT (4-14)

Equation 4-14 was derived for an ideal gas. As such it is valid for all possible
changes in state of ideal gases, whether constant pressure or not. A more
general equation for the change in enthalpy will be developed in Chapter 8
which will be applicable not only to gases and vapors but also to solids and
liquids.

4-6 RELATIONSHIP BETWEEN cp AND cv


The specific heats, cp and cv, are properties of a system. Hence, there must be
a relationship between these properties. Since h = u + p v, dh = du T d(pv).
For an ideal gas, dh = cp dT, du = c„ dT, and d(p v) = R dT. Then,

Cp dT = c„ dT + R dt
The p-v Relationship for a Reversible Adiabatic Process 101

or
cp = c„ + R (4-15)

Equation 4-15 is valid for all ideal gases.


Previously, the gas constant R was shown to be dependent on the p-v-T
relationship. In Eq. 4-15, R is shown to be also equal to the difference
between the specific heats, cp and cv, for an ideal gas, or,

R = cp - c„ (4-15a)

Since R is always positive, the value of cp for an ideal gas must always
exceed cv.
As will be seen in Section 4-7, it may be desirable to use the ratio of the two
specific heats, cp and cv. The symbol k is used for this ratio. Thus,

k= — (4-16)

Introducing the term k into Equation 4-15, and rearranging,

k =—+ 1 (4-17)
c„

4-7 THE p-v RELATIONSHIP FOR A REVERSIBLE


ADIABATIC PROCESS
Since, by definition, SQ for an adiabatic process is zero, the first-law energy
equation for a reversible adiabatic process taking place in a closed system at
rest becomes
0 = du + p dv-
or
0= dT + p dv
For an ideal gas,
p dv- + v- dp
pv = RT and dT =
~ R
p dv + v dp
0= c + p dv
v R~
Multiplying by Rlpv,
dv , dt\ d v-
+—)+R = 0
v- P / v-

dp
(c. + R) — + c = 0
P
dp
£jlA±+c = 0

Cp d v- dp
=0
v-
102 Changes in State of Gases

Integrating this equation and letting k = cp/cv,

k In v- + In p = C'
or

pvk - C (4-18)

Equation 4-18 is written on the specific volume basis. For the total volume,
the equation becomes

pVk = C (4-18a)

Comparison between Eqs. 4-1 and 4-18a shows that the reversible adiabatic
process of an ideal gas is a special form of a polytropic process with the
exponent n assuming the special value of k.

Example 4-6. Air undergoes a quasistatic adiabatic expansion from an initial


condition of 15.2 bars, 112°C until the pressure becomes 1.15 bars. The
original volume is 34,000 cm3. Determine the change in internal energy,
neglecting the effect of temperature on the specific heat.

Solution. The mass of air,

pV _ (15.2 x 10)(34,000)
= 467.7 g
RT (0.287 x 100X112 4-273)
T} = 385
= 184.2 K
(pnJp2)(k~m (15.2/1.15)04/1-4'

Using Eq. 4-9,

U2-Ui = mcJT2- T,) = 467.7 x 0.718(184.2- 385) = -67,430 J Answer

Alternate Solution. Since there is no heat transfer,

U2- Ui = - WU2
1/1.4
V, = V, (
- Sf = 34’000 ( iM) = 214,900 cm3

11/ _P2V2-p1V1 1.15x214,900- 15.2x34,000


= 6,743,000 N-cm
Wl'2 1 -k 1-1.4
or 67,430 N-m or 67,430 J Answer

Then U2-Ux = -67,4301.

4-8 SPECIFIC HEATS FOR POLYTROPIC PROCESSES


As defined in Section 2-5b, specific heat is the amount of heat added to a
system per unit mass of the system per unit change in the temperature of the
system, or,
8Q
C m dT
Specific Heats for Polytropic Processes 103

In general, the amount of heat that must be added per unit mass of the
system to produce an unit change in temperature of the system is dependent
on the nature of the process during which the heat is added. This is true
because work normally is done by the system as heat is added to it. The
temperature rise of the system thus is dependent on how much of the heat
added to the system leaves in the form of work and how much remains in the
system to produce a temperature rise.
Although the heat added to a closed system may be determined by evaluat¬
ing the change in internal energy of the system and the work delivered during
a change in state, in certain problems it is desirable to determine the heat
added in terms of the specific heat of the system for a given change in state.
Let cx represent the specific heat of the system, with the subscript x denoting
that the specific heat is unknown until the nature of the process is specified.
For an ideal gas, a relationship between cx and the polytropic exponent n may
be determined as follows. Assume a reversible process taking place in a
closed cylinder, containing a piston.

8q = du + p dv

cx dT = c„dT + pdv
or
p dv
c = c v- -f
'-'X ^ 1
dT

Since p v- = RT,

p d v- + v- dp
dT
R

Then

Rp dv
cx cv +
p dv + v- dp

From pvn = c,

np vn 'dv + vndp= 0

or
if- dp
p dv = -
n
Then

R(—v dpln)
(—if- dpln)-I- v dp
or

R
cx — c„ + (4-19)
1 —n
104 Changes in State of Gases

Since n can have any value ranging from positive to negative infinity, cx
also will have values ranging from positive to negative infinity. When n = 0, as
it does for a constant-pressure process, cx becomes cp. Similarly, for the
constant-volume processes with n = oo, cx becomes cn. When n = 1, cx is infinity.

Example 4-7. Find the heat added in Example 4-5 by use of the polytropic
specific heat.

Solution. From Eq. 4-19,

R
Cx~c,,+(\-n)

or

0 787
cx = 0.718 + - ' = -0.102
1-1.35
Q = mcx(T2- T,) - 32.14(-0.102)(873 - 333) = -1770 J Answer

Note: When only the heat added is required, the use of the polytropic specific
heat eliminates the necessity of first calculating the internal energy
changes and the work done.

4-9 VARIABLE SPECIFIC HEAT


With the exception of monatomic gases, the specific heats of gases, as well as
those of solids and liquids, are influenced by temperature. In addition, the
number of atoms per molecule and their arrangements within the molecule
affect the specific heats. The reason for these effects and the prediction of the
magnitude of these effects require a study of microscopic thermodynamics as
well as quantum mechanics. Such a study is presented in Chapter 16.
Although such a study is essential to a full understanding of specific heats, it
is only necessary in this chapter to understand how the specific heats, once
determined, can be used in the solution of engineering problems.
The specific heat-temperature relationship is a complicated one when the
temperature range is large (i.e., 500°C or more) and, hence, the precise
relationship generally is not given in equation form. In 1938, Sweigert and
Beardsley studied available information on the specific heats of several gases.
They found that it was possible to formulate empirical equations, containing
at the most four terms, which have a maximum error of 1.5 percent and an
average error much lower than this. Their equations were converted to the
metric system and are presented in Table 4-1. It should be noted that these
equations are valid only over those temperature ranges listed in the table.
Particularly at low temperatures, as the point of liquefaction is approached,
the deviations from these equations are quite large. Since equations generally
are not available for these conditions, tables of properties are desirable.
Another restriction on equations, such as those of Sweigert and Beardsley,
is that they are valid only at low pressures, generally lower than a few
Variable Specific Heat 105

Table 4-1
Equations for Specific Heats (cp is in J/g mole-K)

Temperature Maximum
Range (K) Error (%)

10,n 536.7,3558
o2 cD = 48.211 + 300-2800 1.1
Vt T
536.7 3558 , 0.376^
Cp=48.211 + + (T -2222) 2800-5000 0.3
Vt T 1000
8.07 x 103 1.499 xio6
n2 cp = 39.65 T + J2 300-5000 1.7

7.652 x 10’, 1.363 x 106


CO cp = 39.61 + T2 300-5000 1.1

h2 cCp = 24 12 1344T
1000 + 62-41 300-2222 0.8

c„ = 24.12 - ~ - — (T - 2222) 2222-5000 1.4


1000 VT 1000

h2o cp = 83.15 17,450 300-2800 1.8


Vt T
15.19x 103 , 1.822x 106
co2 Cp = 67.83 + T2 300-3500 0.8

CH4 cp = 18.92 +0.0555T 300-830 1.2


c2h4 cp = 17.71 + 0.0887T 195-610 1.5
c2h6 Cp = 16.79 + 0.1233T 220-610 1.5
c8h18 cp = 33.16 + 0.453T 220-610 est. 4
c i2H26 cp = 36.34+ 0.670T 220-610 est. 4

hundred pounds per square inch. The exact effect of pressure on the specific
heat of any substance at a given temperature is discussed in Chapter 8 and is
given by Eqs. 8-33 and 8-34. When pv = RT, these equations show that
pressure has no effect on the specific heats.
When a more exact knowledge of specific heats is desired, tables of specific
heats are available for many gases. One widely used table is that of Keenan
and Kaye, entitled Thermodynamic Properties of Gases (John Wiley & Sons,
1948). It should be noted that this table is valid only for gases at low
pressures.
In 1955, the National Bureau of Standards, U.S. Department of Commerce,
issued a very extensive study entitled Tables of Thermal Properties of Gases,
Circular 564. Figures 4-11, 4-12, and 4-13 were prepared based on data from
these tables. Figure 4-11 shows the effects of pressure and temperature on the
values of the constant-volume specific heat, cv, and the specific heat ratio, k,
for air. Figure 4-12 is similar to Figure 4-11 except that it shows the effects on
Specific heat ratio cp/c
Temperature (K)
FIGURE 4-11 Effects of pressure and temperature on cv and k for air.

Temperature (K)
FIGURE 4-12 Effects of pressure and temperature on cp for air.

106
Variable Specific Heat 107

I
CD

03

CCD
o

o
0)
a
GO

FIGURE 4-13 Effects of pressure and temperature on cp for various gases.

the constant-pressure specific heat, cp. Figure 4-13 shows the effects of
pressure and temperature on the constant-pressure specific heat, cp, of
other gases.
These figures show that at low temperatures, both pressure and temperature
have a marked effect on the specific heat. At very high temperatures, pressure
does not have a significant effect on specific heat. Furthermore, at high
temperatures, for low and moderate temperature changes the average specific
heat for a given temperature range may be used for ordinary engineering
problems.
The specific heats previously discussed are known as instantaneous specific
heats; that is, specific heats at a specific temperature. These are in contrast to
mean specific heats, which are the true mean specific heats between a fixed
reference temperature and the temperature in question. An instantaneous
specific heat may be read directly from figures such as Figure 4-11, 4-12, or
4-13, or may be approximated by substituting the desired temperature in
equations such as those given in Table 4-1. Instantaneous specific heats also
may be obtained from many published tables. Tables of mean specific heats
are also available or these specific heats may be calculated from the in¬
stantaneous specific heats. An examination of Figures 4-11 to 4-13 shows that
the instantaneous specific heat curves may be considered as straight lines
over a small temperature range, and hence, the mean specific heat may be
taken as that specific heat at the mean temperature. However, this method
108 Changes in State of Gases

cannot be used over a larger temperature range since there is a considerable


deviation from a straight line. When a specific heat is plotted against tem¬
perature, the true mean specific heat between two temperatures equals the
area beneath the curve divided by the temperature difference. Thus,

I cdT
mean c = -=-
T2- T,

For the most part, specific heats are of value in that they permit the
evaluation of heat added, internal energy changes, and enthalpy changes. In
Chapter 8 it will be shown that specific heats are useful in determining
entropy changes.
The methods of determining energy changes when there is an appreciable
variation in the specific heats are illustrated in the following examples.

Example 4-8. Nitrogen expands in a cylinder in a reversible adiabatic manner


from a temperature of 850°C to a temperature of 78°C. Determine the work
done in joules per gram of air, using Table 4-1.

Solution. It will be assumed that the pressures are sufficiently low so that the
nitrogen can be treated as an ideal gas for the given temperatures. For an
adiabatic process, W = -(U2- U\) = Uj - U2. Per mole,

8.07 x 103 | 1.499 x 1Q(


du = c„ dT = (cp - R0) dT = (^39.65 - -8.314 dT
T

31.336 — ^7x^ + 1A99x^)dT

Integrating and putting in the limiting temperatures.

1123
ui-u2= 31.336(1123 - 351) - 8.07 x 103 In
351
1 1
1.499 x 10 ^1123 351

17,740 J/g mole or 633 J/g


W = 633 J/g Answer

Example 4-9. Determine the mean specific heat of the nitrogen in Example
4-8 and compare it with that given in Table 3-1.

Solution

mean cv = j = = 0.82 J/g-°C Answer

This compares with a value of 0.734 J/g-°C given in Table 3-1.


Internal Energy and Enthalpy Changes of Gases 109

4-10 INTERNAL ENERGY AND ENTHALPY CHANGES


OF GASES
Changes in internal energy and enthalpy of gases may be approximated by use
of the specific-heat equations of Sweigert and Beardsley and other similar
equations. There are two drawbacks to this method. The integrations and the
substitutions of temperatures in the integrated equations are time consuming.
Furthermore, since the equations themselves are only approximations, the
calculated changes in internal energy and enthalpy are not as accurate as may
be desired.
To eliminate these difficulties, considerable effort has been devoted to the
determination and tabulation of internal energy and enthalpy of various gases.
For the most part such tabulations are based on the original spectroscopic
data rather than on approximate specific-heat equations, for instance, those of
Sweigert and Beardsley and, hence, give more accurate results.
In most engineering problems, changes in enthalpies and internal energies are
required rather than the absolute values of these quantities. Various datum
temperatures ranging from 0 K to 25°C have been selected for tabulating these
properties. A skeleton table of the enthalpies of common gases at atmospheric
pressure is given in Table 4-2. Internal energies can be calculated from the
relationship H = U + pV. Except for water vapor and carbon dioxide at low
temperatures, the compressibility factor of these gases is substantially equal to
unity. Hence, the relationship between internal energy and enthalpy may be
expressed as H = U + RT. For convenience, values of R0T, on the molar basis,
are given in Table 4-2.

Example 4-10. Solve Example 4-8 using Table 4-2.

Solution

850°C = 1123 K and 78°C = 351K

From Table 4-2,

At 1123 K, h = 1221.6 J/g and R^T = 9337 J/g mole or 333.5 J/g
At 351 K, h = 382.7 J/g and R0T = 2918 J/g mole or 104.2 J/g

Then

u2 U\ — h2 h\ — (RT2 - RT{) = (1221.6 - 382.7) - (333.5 - 104.2) = 709.6 J/g


Answer

Example 4-11. Air enters a system at 30° C and leaves at 340° C. The rate of air
flow is 8.52 kg/s. The power input to the system is 8.4 kW. Determine the rate of
heat addition to the system.

Solution. Since the information is not given otherwise, it will be assumed that
the kinetic energy of the air can be neglected. The first-law equation for an open
Based on Tables of Thermal Properties of Gases, U.S. Department of Commerce, National Bureau of Standards Circular 564, 1955.
o
re ON Tf NO d OO © ,—i re Tt c~ © Cl >n oo 1—1 'rf NO o Cl
E-h b re
00
NO
NO
r-
o
ON
Tf
Cl
re
NO oo
ON
Cl
oo
in
\D
oo
eT
i—l
re
c~
ON ©
©
re
©
ON
c-
©
ON
Cl
to
ON
1—1
©
r-
Cl
'Tj-
o
rS 00 d C1 re eT Tj- m © c- oo ON i—i re © 00 ON y—i re
n_i i— i—1 l-1 Cl Cl Cl
<u
a.

U. I-H d ON c- OO
<u o
o in
Enthalpy of Gases at Standard Atmospheric Pressure, Joules per Gram

On
S' re Cl re re
d ON re in
& >

«
c
NO o l-H i—1 d NO NO ©
00
re NO d d oo OO ^4 00 ci oo ON re © m c-
} , r-~ Cl 1—H Cl NO 1—1 d Cl ON © m ON oo © ©
to OO tO Cl NO r—H <o © ON © i—* T-H
1—I Cl re Tf tO d oo © Cl c- © re c- ©
Cl Cl d Cl

c ci rf re Cl ON Cl © c- r-; in © m,
o 3 l-H y—< l—H i—i
to re re to NO On re re
"t- £ NO l-H o © o ^—1 re c- Cl oo oo
a .2 1—1 Cl re ef to © r- 00 ON Cl ^1- ©

Note: The datum for this table is the internal of the gas at 0 degrees absolute.
U Q
Table 4-2

<D
C ^3 to l-H Cl d Cl d c- re in ON ■^t Cl NO NO NO © © © oo ©
2 x r4 ON _4 to ,-4 oo oo NO re ,-H ci d m] >c NO On re On
2 o »o l-H y—i Cl Cl re in © oo Cl c- Cl r- re ON te, i—i oo
& 5 Cl o re 3" »o NO r- 00 On © re m 00 © re in 00 i—i re ©
Cl Cl Cl Cl re re re
tj ^2
S

c
o i
ON d d re
, ON Cl NO oo *n in © © re oo re »o o
a>
00 ON ,-4 NO ci to ,-4 © NO ci © T—i NO in NO i-4 oo ON ci od NO
oo ON Cl d NO NO NO © © c- oo © Cl in OO Cl in ON oo re
X r—1 Cl C1 re TT *n © r- oo On Cl © oo i re to, oo © re
Cl Cl Cl Cl re re
O

c
a> 1— o Cl ef- er ON © NO © ON c- ON m, re C; to 'Tf
00 -4 ^4
c-n ci r- ON ,-4 to © NO >c NO © in ci NO d re ci to ON
o o »o l-H i—i Cl Cl re ef © c~ 1—H in © in i—i C re ON to y—i
i—1 Cl Cl re ef to NO c~ oo ON © re in oo © re in oo © re NO
Cl Cl Cl Cl re re re
2

Tf c~ o re Cl r- eT © ON re in re c- © to d to NO
l-H
• P—» 00 ON © 1—4 re d re ci re NO OO oo oo oo i—i i-4 re re
©
Os ON tO o o o o t—i Cl re r- NO i—i \D Cl r- re o to
i—i Cl re Tf >o NO r- 00 On © Cl r- © Cl to c- © Cl to
Cl Cl Cl Cl re re re

<D
u.
3
■«—>
ce o o o O o o o © © © © © © © © © © © © © ©
o o to o o o o © © © © © © © © © © © © © ©
nL/ —1
1 cire Cl to NO r~~ 00 ON © Cl rr NO oo © Cl NO oo ©
Cl Cl Cl Cl Cl re
C
£-<
<d
H

110
Internal Energy and Enthalpy Changes of Gases 111

system thus reduces to

Q = m(h2 - hi) -l- W

Using values of h from Table 4-2,

Q = 8520(621.2 - 202.5) - 8.4 x 1000 = 269,800 J/s Answer

Example 4-12. Repeat Example 4-11 using the specific heat from Table 3-1.

Solution

Q = H2 - Hi + W = mcp(T2 - Tt) + W = 8520(1.004)(340 - 30)


— 8.4 x 1000 = 264,300 J/s Answer
112 Changes in State of Gases

PROBLEMS ____
4-1 Air exists in a closed tank. State how the air can be heated in a process
approaching a reversible one.
4-2 Can the air in the tank in Problem 4-1 be expanded in close to a
reversible process? If so, how?
4-3 In a compression ignition engine (termed also as a diesel engine), air,
originally at 49°C, is to be compressed. The compression ratio is 15 to 1
and the compression obeys the law pVlM=C. Determine the tem¬
perature at the end of compression.
4-4 Gases expand in a gas turbine from a pressure of 5.9 bars to a pressure
of 1.02 bars. The expansion obeys the law pV127= C. The initial gas
temperature is 892°C. Determine the temperature of the gases at the
exit of the turbine.
4-5 Determine, per kilogram of air, the work of compression and the heat
lost per gram of air in Problem 4-3. Neglect the effect of temperature on
specific heat.
4-6 (a) Treating the gas as air, determine the change in internal energy per
kilogram of gas in Problem 4-4. Calculate a value of cv at the mean
temperature, using Table A-9.
(b) Is it reasonable to treat the gases as air?
(c) Is it reasonable to neglect heat losses?
4-7 Determine the work done per kilogram of gas passing through the
turbine of Problem 4-4, treating the gas as air and neglecting heat
losses.
4-8 Calculate the work done per kilogram of gas during its expansion in
Problem 4-4 using Eq. 4-6. Compare this answer to the answer of
Problem 4-6(a) and comment.
4-9 Air at 1.01 bars, 22°C, is compressed in a reversible adiabatic manner
until its pressure becomes 5.05 bars. It is then heated at constant
pressure until its volume becomes four times that before heating. Next,
it expands in a reversible adiabatic manner until its pressure becomes
1.01 bars. Finally it is cooled at constant pressure until its volume
becomes one-fourth that before cooling. Specify the final temperature.
4-10 Assuming constant specific heat, determine per kilogram of air the heat
added, change in internal energy, and work done for each of the four
processes in Problem 4-9.
4-11 (a) During an isothermal heating of nitrogen in a closed cylinder, its
pressure decreases from 5.16 bars to 2.12 bars. The original volume of
the nitrogen was 0.052 m3. Determine the work done during the expan¬
sion.
(b) What assumption was made for work in part (a)?
4-12 Determine the heat added in Problem 4-11.
Problems 113

4-13 Air expands adiabatically in a cylinder, its temperature dropping from


150°C to 50°C. Determine the work done per kilogram of air for both a
reversible and an irreversible expansion. Comment on your results.
4-14 Methane (CH4) enters a pipe line compressor at 28°C, 34.5 bars pres¬
sure. The flow rate is 340 m3/min. Determine the mass rate of flow.
4-15 The power input in Problem 4-14 is 10,200 kW. Neglecting heat losses,
determine the gas temperature at compressor exit.
4-16 Compare the volume rates of flow of air and hydrogen when they are
used to cool an electric generator having a power output of 500,000 kW.
Assume that the gas picks up energy equivalent to 1.2 percent of the
generator output. Each gas enters the generator at 3.2 bars, 32°C. The
temperature rise of the gas in the generator is 6.8°C.
4-17 Air enters an electric heating duct at 1.02 bars, 27°C and leaves at 42°C.
The electrical input is 108 kW. Determine the volume rate of air flow.
4-18 (a) Air expands in a reversible adiabatic manner from 0.08 nr at 2.8
bars to a volume of 0.16 m . Determine the work done.
(b) Same conditions as part (a), except the expansion is an irreversible
adiabatic expansion to the same final volume with the work done equal
to 88 percent of that in part (a). Determine the final pressure and
compare with that of part (a).
4-19 A tank contains air at 10.5 bars, 95°C. A large valve on the tank is
opened suddenly and the pressure drops almost immediately to 2.07
bars. The valve is then closed. After the tank has stood for some time,
the pressure in the tank increases to 2.65 bars and the temperature is
found to be 29.8°C. Make calculations to show whether or not the
expansion in the tank was close to a reversible adiabatic.
4-20 Air flows through a large pipe line where the conditions are maintained
constant at 32 bars, 35°C. A valve is suddenly opened allowing air to
flow into a small, completely evacuated tank having a volume of
0.015 m3. As soon as the pressure in the tank reaches 32 bars, the valve
is closed. Neglecting heat losses, determine the temperature of the air
in the tank and the amount of air present.
4-21 After standing for a short time, the temperature of the air in the tank in
Problem 4-20 is found to be 124°C. Determine the heat lost from the air.
4-22 Air flowing in a pipeline has a temperature of 40° C. A slight amount of
oil from the air compressor is carried over with the air. The oil has a
flash point of 150°C. (The flash point is the self-ignition temperature.)
Make calculations to show whether or not there may be an explosion if
a valve is closed suddenly.
4-23 Although there are very small amounts of other gases present, dry air
may be assumed to be 21 percent oxygen and 79 percent nitrogen by
volume. Its molecular mass may be taken as 28.97 g/g mole. Using
114 Changes in State of Gases

Table 4-1, derive an equation for the specific heat of air for a tem¬
perature range of 300 K to 2800 K.
4-24 In a theoretical Otto cycle using air, heat is added at constant volume at
the rate of 1860 J/g. The initial temperature is 540°C. Determine the
temperature after addition of heat by (a) using constant specific heat
from Table 3-1, and (b) using the specific-heat equation derived in
Problem 4-23. (See Eq. 4-15 to obtain c„.)
4-25 Determine the mean value of c in Problem 4-24(b). Compare it with the
value used in Problem 4-24(a).
4-26 10.4 kg of carbon dioxide are heated at constant pressure from 280°C to
840° C. Determine the amount of heat added by (a) using the constant
value of specific heat from Table 3-1, and (b) using Table 4-1.
4-27 Determine the mean specific heat for Problem 4-26(b).
4-28 Solve Problem 4-26 using Table 4-2.
4-29 Solve Problem 4-24 using Table 4-2.
4-30 Air enters an axial flow compressor at 35°C, 1.04 bars and leaves at
285°C. The rate of air flow is 102 m3/min. The power delivered to the air
is 98 percent of the power input to the compressor, that is, the
mechanical efficiency is 98 percent. Determine the power input to the
compressor. State any assumptions that must be made.
4-31 Air enters an axial flow compressor at 32°C, 1.02 bars and leaves at
292°C. The power delivered to the air is 98 percent of the power input
to the compressor. The power input is 892 kW. Determine the volume
rate of air flow. State any assumptions that must be made.
4-32 (a) Air at 6.2 bars, 65°C is heated in a reversible isothermal manner, its
volume increasing from 0.05 nr to 0.32 m3. Determine the work done,
(b) Does it make a difference in part (a) if the system is open or closed?
Why?
4-33 A system is cooled at constant volume from state 1 to state 2, the
internal energy decreasing 88 kJ. It is then heated from state 2 to state
3. The total work done between states 1 and 3 is 44 kJ. The system then
progresses to state 4, the internal energy increasing 35 kJ. The system
finally returns to its initial state, the internal energy increasing 53 kJ.
Determine the heat added to the system as it changes from state 2 to
state 3.
4-34 An ideal gas is heated at constant pressure by the addition of 113.2 J. It
has its temperature increased 110°C. During the heating, the gas does
32.45 J of work. Determine the gas constant R if the mass present is
0.998 g. Determine also the molecular mass of the gas.
4-35 A hot-air furnace heats air from 22°C to 38°C. Assume that the
pressure of the air is 101.5 kPa. Determine the volume rate of air flow
from the furnace when the furnace must deliver 60,200 kJ of heat per
hour. Neglect the water vapor present in the air.
Problems 115

4-36 Actually there are 0.015 g of water vapor per gram of dry air in Problem
4-35. Compare the amount of heat picked up in the furnace per gram of
dry air and per gram of wet air.
4-37 Air enters a centrifugal compressor at 1.018 bars, 31°C at the rate of
11.5 m3/s. The temperature at compressor exit is 281°C. Determine the
power input to the compressor if 98 percent of the power input is
delivered to the air. State any assumptions that are made.
4-38 Air flows through a heater for a magnetohydrodynamics generator,
where it is heated from 40°C to 1600°C. Determine the heat required
per kilogram of air by (a) using the specific heat in Table 3-1, and (b)
taking into account variable specific heat.
4-39 A compression ignition (diesel) locomotive engine uses 635 kg of fuel
per hour. The efficiency of the engine is 35 percent (i.e., the output is 35
percent of the energy supplied). The heating value of the fuel is
42,000 J/g. Determine the output of the engine.
4-40 The engine in Problem 4-39 uses 21 g of air per gram of fuel. Air
is supplied at 1.015 bars, 29°C. Determine the volume rate of air flow.
4-41 Carbon dioxide is heated at constant pressure in a cylinder and per¬
forms 24.2 J of work. Determine the heat added.
_ 5 _

THE SECOND LAW


AND ENTROPY

5-1 INTRODUCTION
In the latter part of the eighteenth century and the earlier part of the
nineteenth century, man learned to produce work from heat by utilizing steam in
engines. This development was to have profound and far-reaching effects. Such
an accomplishment was most remarkable, particularly since heat was not
recognized at that time as a form of energy. At first, there was no recognition that
the first law governs the heat-work transformation. There certainly was no
awareness of any limitation in the amount of heat that can be transformed into
work.
The first person credited with recognizing such a limitation was Sadi
Carnot, a remarkable French engineer. In 1824, he published his treatise
entitled Reflections on the Motive Power of Heat} In it, Carnot concluded that
.. heat cannot perform work except when it passes from a higher to a lower
temperature level.” Furthermore he stated that the amount of work that can
be produced by a steam engine from a given quantity of heat is a direct
function of the difference between the temperature at which the steam is
produced and the temperature of the steam at the engine exhaust.
The insight of Carnot was especially remarkable, since his theory was
developed at least 20 years before the brilliant demonstration of James
Prescott Joule convinced the scientific world that heat is a form of energy.
Indeed, it seems probable that Carnot was not aware of the true nature of
heat. In Carnot’s time, the caloric theory of heat was the accepted one. This
theory pictured heat to be an imponderable fluid, having neither mass nor
substance, but occupying all space. To be sure, scientists such as Count
Rumford and Mayer were questioning this concept, but it remained for Joule
to demonstrate so convincingly that heat is a form of energy.
After it was recognized that heat is a form of energy and that some of it can
be transformed into work, many other observations were made relative to
energy transformations. Some of these observations showed that certain

'Carnot was 28 years old at this time. He developed the concepts involved in this treatise several
years earlier.

116
Introduction 117

transformations may take place with the first law holding precisely, but that
these transformations cannot be reversed without energy being supplied
externally. Some examples of these transformations are as follows:

1 Whenever a heat-flow path exists, heat flows readily from a region of high
temperature to one at a lower temperature, but the reverse is not true.
Heat does not spontaneously flow “up hill.”
2 Work, of the mechanical form, can be converted completely into heat.
Consider the pushing of a block over a rough surface. The work used in
pushing the block to overcome friction produces a heating effect on the
block and the surface. To restore the temperature of the system to its
original value, heat, equivalent in amount to the work input, must be
removed from the system. But experience has shown that the converse of
this operation is impossible. The heat that was removed will not of its own
accord flow back into the system, restore the block to its original position,
and deliver an amount of work equivalent to the original work input.
Furthermore, heating the block will obviously not cause it to move, either.
3 An electric current flowing through a resistor produces a heating effect.
Heat, in an amount equivalent to the electrical energy input, may be
removed from the resistor. However, the converse of this operation is not
true. The heat that was removed will not flow back into the resistor and
produce electrical energy.
4 An uncontrolled exothermic chemical reaction produces products whose
temperature is in excess of the original temperature. Heat must be
removed from the products to restore their temperature to that of the
original mixture. Thus, there is transformation of chemical energy into
heat. A common example is a mixture of hydrogen and oxygen. The
introduction of a spark into the mixture produces a violent reaction,
forming water vapor that has a temperature much in excess of the original
temperature. The temperature of the water vapor may be restored to that
of the original hydrogen-oxygen mixture by removing heat. It is in¬
conceivable, however, that this heat that was removed will flow back into
the water vapor and dissociate it into hydrogen and oxygen.
5 Consider a chamber divided into two parts by means of a partition. Let
one part be occupied by oxygen and the second part by nitrogen. When
the partition is ruptured, mixing of the nitrogen and oxygen takes place
until there is a uniformity of the mixture throughout the chamber.
However, our experience tells us that the mixture will not of its own
accord become “unmixed,” with the oxygen and nitrogen returning to their
original positions.

In the five examples cited above, the first law governs the energies involved
in the transformations. But the first law does not answer many questions,
such as why there cannot be complete transformation of heat into work but
work can be completely transformed into heat. Neither does it answer the
118 The Second Law and Entropy

question of why some processes can proceed in one direction but not in the
other. Further, the first law in no way explains why it is possible for certain
processes to take place but impossible for other processes to occur. The
second law does provide answers to these questions.
The scope of the second law is broad and quite inclusive. At the heart of
the second law is a property called entropy. To understand entropy fully, as
well as make complete use of it, it is necessary to investigate entropy from
the microscopic viewpoint with the aid of statistical thermodynamics. As a
first approach, however, the nature of entropy will be investigated from the
classical viewpoint. This may be done by utilizing the concept of the Carnot
cycle.

5-2 THE CARNOT CYCLE


Although he stated that the amount of work that can be obtained from a given
quantity of heat is a function of the temperatures of the heat source and the
heat sink, Carnot did not specify the precise amount of the work that can be
obtained. Neither did he specify the cycle of operation for obtaining the
maximum amount of work. Lord Kelvin is credited with detailing such a cycle
of operation. Since it related directly to Carnot’s concepts of heat-work
transformation, it is known as the Carnot cycle.
The Carnot cycle engine is one of many types of heat engines. A heat
engine is a device for converting heat into work. A heat engine requires a heat
source and a heat sink. Its operation may be cyclic (see Fig. 5-la). Assume,
initially, that the working substance within the engine is at a relatively high
pressure and at a temperature somewhat below that of the heat source. When
a heat conductor is inserted between the heat source and the engine, heat will
flow into the engine. As the heat flows into the engine, the working substance
in the engine expands, delivering work. After a sufficient amount of heat has
been added, the heat conductor is removed. By virtue of its pressure the

(a)
FIGURE 5-1 Elements of a heat engine, (a) Cyclic operation intermittent heat flow, (b)
Continuous heat flow.
The Carnot Cycle 119

working substance continues expansion, delivering more work and decreasing


in temperature. After the desired amount of expansion has taken place, the
heat conductor is inserted between the engine and the heat sink. Now, as the
working substance in the engine is compressed, heat is transferred to the heat
sink. When a sufficient amount of heat has been transferred to the heat sink,
the heat conductor is removed. The working substance is now further com¬
pressed until it reaches its original state. The cycle may now be repeated.
These processes describe a common cycle of operation of a heat engine.
Other combinations of processes may be used. In most heat engines the flow
of heat to and from the engine is continuous (see Fig. 5-1 b). This is especially
true in steam power plants and for such direct energy conversion devices as
the thermoelectric, the thermionic, and the magnetohydrodynamic generators.
These operations will be considered later on.
It is common practice to assume that, thermodynamically speaking, the heat
source and the heat sink are infinite. This does not necessarily refer to their
size but means that they, because of their nature, do not experience a
measurable change in temperature as heat is being added to or removed from
them. Following this practice, it will be assumed here that the temperatures of
the heat source and heat sink remain constant during both heat addition and
heat removal.
One purpose of the Carnot cycle is to predict the maximum amount of work
that may be obtained from a given quantity of heat supplied by the heat
source. Thus all processes composing the cycle are idealized ones, and hence,
must be reversible. There can be no friction, no heat transfers with finite
temperature differences, and no extraneous heat transfers. See Section 4-1 for
a discussion of reversibility.
A pressure-volume diagram for a Carnot cycle engine is shown in Figure
5-3. Reference should be made to Figure 5-2. When the temperature of the

Heat
conductor

\
El
Heat !|| (jemperature
sink constant)

FIGURE 5-2 Carnot cycle engine.


120 The Second Law and Enthalpy

Carnot cycle.

working substance in the engine is equal to that of the heat source, the heat
conductor is inserted between the two. By virtue of high gas pressure in the
cylinder, the piston moves outward, tending to decrease the gas temperature.
As soon as this takes place, heat will flow from the heat source into the engine
and will continue to flow until the heat conductor is removed (process 1-2).
During this process, the temperature within the cylinder is substantially equal
to that of the heat source. After thermal contact with the heat source is broken,
the working substance expands adiabatically until its temperature reaches that
of the heat sink (process 2-3). Now thermal contact is established with the
heat sink. The piston is pushed to the left, causing heat to flow into the heat
sink in a constant temperature manner (process 3-4). At the correct time,
thermal contact with the heat sink is broken and an adiabatic compression
takes place (process 4-1) until the temperature reaches that of the heat source.
The cycle is now complete.
The thermal efficiency of a heat engine is defined as that portion of the heat
supplied to it which is delivered as work. Or, thermal efficiency,

W
v, = Qh (5-1}

where QH is the heat received from the source and W is the work output of
the engine.
It is desirable to obtain an expression for the thermal efficiency of the
Carnot cycle in order that the effects of the temperature of the heat source
and the heat sink can be evaluated. Carnot stated that thermal efficiency is
independent of the nature of the working substance. This will be discussed in
some detail later on in this chapter. Assume, initially that the working
substance is an ideal gas.
For a heat engine with no extraneous heat transfers (i.e., only receiving
The Carnot Cycle 121

heat from the source and rejecting heat to the sink),

W = Qh-Ql (5-2)
where QL is the heat rejected to the sink. Substituting into Eq. 5-1,

v -Qh-Ql
(5-3)
Vt Qh
Qh is transferred at a constant temperature. Hence, from Eq. 4-6a with
P\V\ — PiV2 and applying the first law,

Qh = Wi.2 = Pi V, In = mRTH In
V 1 V ]

where TH is the source temperature.


In a manner similar to that used to find QH,

Ql = = p4V4 In Tr = mRTL In ^
V4 V4

where TL is the sink temperature. Substituting into Eq. 5-3,

_ mRTH In V2IV, - mRTL In V,/V4 /r „ ,


Vl ~ raRTfi In V,/V,- (5'3a)

Since both processes 2-3 and 4-1 are reversible and adiabatic, using Eq. 4-3,
k-l
III (V2)
and
H

Tl Vv3y1
Thus

VT V, V2
and
V2 v4 v, V4
Substituting into Eq. 5-3a,

th-tl II
Vt 1 - (5-4)
Th Th

Equation 5-4 shows that the thermal efficiency of the Carnot cycle increases
as the temperature difference increases and approaches 100 percent as the
source temperature approaches infinity. Equation 5-4 also shows that the
thermal efficiency increases as the sink temperature decreases. Hence, it is
desirable to use that available sink having the lowest temperature. Although
Equation 5-4 indicates that there will be a further gain in thermal efficiency by
artificially lowering the sink temperature, there will be an overall loss by doing
this since the work required to lower the sink temperature will actually
exceed the gain in engine output.
Example 5-1. 1450 kJ of heat are supplied to a Carnot cycle engine at a
temperature of 820°C. The engine rejects heat to a sink at a temperature of
122 The Second Law and Entropy

40°C. Determine: (a) thermal efficiency, (b) work delivered, (c) heat rejected
to the sink.

Solution

(a) The thermal efficiency

Tl 40 -f 273
0.714 or 71.4 percent Answer
Th 820 + 273

(b) W = Qh x Vt = 1450 x 0.714 = 1035 kJ Answer


(C) ql=qh-W= 1450 - 1035 = 415 kJ Answer

Example 5-2. The heat supplied in Example 5-1 is supplied in a period of


43 s. Determine the power delivered in kilowatts and in horsepower.

Solution. The rate at which work is produced is

1035
= 24.1 kJ/s
43

Since 1 J/s = 1 W,
power produced = 24.1 kW Answer

Since 1 hp = 0.7457 kW,


24.1
horsepower = 32.3 hp Answer
0.7457

5-3 THE REVERSED CARNOT CYCLE


Since all processes in the Carnot cycle are reversible, the cycle itself is
reversible (see Fig. 5-2). When work in an amount equal to that delivered by
the engine, W, is put back into the engine, heat in the amount of QL is
removed from the sink and an amount of heat QH is delivered to the source.
Under these conditions, the heat source, the heat sink, and the surroundings
all will be at the same conditions as those prior to the transfer of heat from
the heat source to the engine.
A pressure-volume diagram for a reversed Carnot cycle engine is the same
as that for a Carnot cycle heat engine, except that the direction of the
processes are reversed. (See Figures 5-4 and 5-5.) One purpose of a reversed
heat engine is to maintain the temperature of the low-temperature reservoir
below that of its surroundings. For this purpose of operation, the reversed
heat engine is called a refrigerator. A second purpose of a reversed heat
engine may be to maintain the temperature of the high-temperature reservoir
above that of the surroundings, that is, to heat a house or other buildings.
Under these conditions of operation, the reversed heat engine is known as a
heat pump.
The Reversed Carnot Cycle 123

/
Heat
\
\
conductors I

L /

FIGURE 5-4 Reversed Carnot cycle engine.

For a heat engine with no extraneous heat transfers,

W = Qh- Ql (5-5)

For a reversed heat engine, operating as a refrigerator,


Ql=Qh~W (5-6)

For a reversed heat engine, operating as a heat pump,

Qh = Ql+W (5-7)

Although Eqs. 5-5, 5-6, and 5-7 are equivalent to each other, they are
presented in their different forms so as to place the desired effect on the left
side of the equations.

V FIGURE 5-5 Reversed Carnot cycle.


124 The Second Law and Entropy

Example 5-3. A reversed Carnot cycle engine receives heat at — 10°C and
delivers it at 30°C. The power input is 4.8 kW. Determine (a) the heating effect
and (b) the refrigerating effect.

Solution

(a) From Eqs. 5-1 and 5-4,


W _ Th - Tl
Qh Th

The heating effect is

Qh = 43o+0(t lof = 364 kW = 36'4 kJ/s Answer

(b) The refrigerating effect

Ql = Qh — W = 36.4-4.8 = 31.6 kW =31.6 kJ/s Answer

Example 5-4. Determine the heating effect in Example 5-3 in calories and
Btu/min.

Solution. Since 1 cal = 4.1868 J,


36.4 x 1000x60
heating effect = = 522 x 103 cal/min Answer
4.1868

Since 1 Btu = 1055.056 J,


36.4 x 1000x60
heating effect = = 2070 Btu/min Answer
1055

The thermal efficiency of the Carnot cycle engine was derived in Section
5-2 by assuming that the working substance is an ideal gas. The question must
be raised as to the effect of the nature of the working substance on the thermal
efficiency. The answer is obtained by indirect logic. Consider the possibility
that the thermal efficiency of a Carnot cycle engine using an actual working
substance is higher than that of a Carnot cycle engine using an ideal gas. In
Figure 5-6, engine A uses the actual fluid and engine B an ideal gas. Let an
amount of heat, QHa, be delivered to engine A. Let its work output be used to

Engine A
Engine B
with actual
Carnot with ideal gas
gas

FIGURE 5-6 Coupled heat engine and reversed engine.


The Clausius Inequality 125

drive engine B, now operating in reverse. When the efficiencies of the two
engines are equal, engine B will pump an amount of heat, QHb, which exactly
equals QHa, into the heat reservoir. But now assume that, because it is using
an actual gas, the efficiency of engine A exceeds that of engine B. Then
engine A will deliver more work to engine B than formerly. Because of a
larger work input, the amount of heat that engine B now delivers to the heat
reservoir will exceed QHa. This means more heat is delivered to the high-
temperature reservoir than is removed from it. An application of the first law
shows that this additional amount of heat comes from the low-temperature
reservoir. If this be true, then the arrangement in Figure 5-6 is an isolated
system in which heat flows from a low-temperature reservoir to one at higher
temperature. But all of our experience has shown that in an isolated system,
heat will not flow from a region of low temperature to one of higher
temperature. It must be concluded, then, that the conditions of operation as
assumed for Figure 5-6 are impossible, and hence, no heat engine can have an
efficiency higher than that of the Carnot cycle.
In a similar manner, it may be concluded that the thermal efficiency of any
reversible heat engine receiving the rejecting heat isothermally cannot be less
than that of the Carnot cycle. It is evident, therefore, that the thermal
efficiency of any and all reversible heat engines, which both receive and reject
heat isothermally, must equal that of the Carnot cycle heat engine, which uses
an ideal gas. This statement is in agreement with Carnot when he stated that
“The motive power of heat is independent of the agents employed to realize
it . . . ”

5-4 THE CLAUSIUS INEQUALITY


The first law simply states that when changes in state of a system take place
there must be an accounting for all energies involved in the change. It gives
no clue as to whether or not specific changes can take place. It is recognized
that when energy is added to a system, changes must take place within the
system. But for an isolated system, the first law is of no aid whatsoever in
predicting whether or not changes of state may take place. Of course, when it
can be established that an isolated system is in a state of thermodynamic
equilibrium, then changes in state cannot take place. In many cases, however,
it is difficult to establish directly whether or not a system is in thermodynamic
equilibrium. The second law establishes a new property, which can show
whether or not a system is in complete equilibrium, and hence, indicates
whether or not a possible change in state of the system can take place. This
property was originated by Clausius2 who called it entropy.
In developing the concept of entropy, Clausius first established what is
known as the Clausius inequality. Figure 5-7 illustrates an arrangement that
may be used to develop this inequality. Details of the “system” are not

2Rudolf Clausius (1822-1888), a German physicist, presented this theorem in 1850.


126 The Second Law and Entropy

FIGURE 5-7 Diagram to illustrate the Clausius inequality.

specified. It receives heat from heat reservoirs I and II. It may reject all of the
heat to the heat sink or it may deliver work up to a maximum of that of a
reversible heat engine. The work inputs to the reversed Carnot engines A and
B are so regulated that they deliver to their respective heat reservoirs exactly
as much heat as is removed from these reservoirs to supply the system. Thus,
numerically, QA| = QS] and Qb2 = Q^. The system is so operated that there is
no change in its contained energy. When every process occurring in the
system is wholly reversible, the work delivered by the system equals the
summation of the work inputs to engines A and B, and the heat rejected by
the system to the heat sink equals the total delivered by the sink to engines A
and B. An irreversibility in the system causes it to deliver less work and to
reject more heat to the heat sink. Thus, numerically

Qsj = Qa3 + Qb3 (5-8)


For Carnot engine A,
Qa3 = Qa, - WA
and from Eqs. 5-1 and 5-4,

Substituting,
Qa3 Qa,
Qa3 — Qa, Qa, or (5-9)

In a similar manner,

Qb3 Qb2
(5-10)

Consider the heat flows into the system as positive and those from it as
negative. Thus QS) and Q^ are positive and Q^ is negative. Then QAl and QBj
are positive. Since, numerically, Q^= Qa3 + Qb„
Qa, + Qb, + Qs, S 0 (5-11)
Entropy 127

Since QA] = QS], from Eq. 5-9,

Qa3 — Qs, j

Likewise, from Eq. 5-10,

Qb3 — Qs, jT

Substituting into Eq. 5-11,

or

(5-12)

Equation 5-12 was established by using three heat reservoirs to exchange


heat with the system. In general, this equation may be written as

(5-13)

This equation states that the summation of each quantity of heat transferred
to the system divided by the temperature at which it is transferred is equal to
or less than zero. Equation 5-13 may be written as

(5-14)

Equation 5-14 is known as the inequality of Clausius.


Note that the T in Eq. 5-14 is the temperature at which heat is added to a
system, and not the temperature of the system itself. Also note that the
integration is carried out over the complete cycle as indicated by the symbol f>.

5-5 ENTROPY
Assume that a completely reversible cycle takes place within a system, such
as the one shown in Figure 5-8. Consider an infinitesimal amount of heat
added at point P as the cycle is traversed in the clockwise direction. Call this
amount of heat SQd. Since the cycle is a reversible one, when the cycle is
traversed in the opposite direction at point P, an infinitesimal amount of heat,
8Qco, will be removed. Because the process is reversible,

8QC\ - -8Qco (5-15)

Let the symbol <f stand for integration around a complete cycle. When the
Clausius inequality is written for the cycle traversed in a clockwise direction,

(5-16)
128 The Second Law and Entropy

FIGURE 5-8 Reversible cycle.

Likewise, when traversed in the counterclockwise direction,

(5-17)

Since 5Qd = SQco, it is evident that neither $(8QJT) or $(8QcolT) can be less
a reversible cycle,

(5-18)
J T
But
N)

8Q\
0
;stV
II

=
T /,

or

J. U )a J,r ^
V T J B
= a constant (5-19)

The value of the constant in Eq. 5-19 is independent of the path chosen to
evaluate it and depends solely on states 1 and 2. As such, it must be a
property. This property was termed entropy by Clausius and is designated by
the symbol S. Then

(5-20)

Equation 5-20 was developed by assuming that the process or processes


followed in progressing from state 1 to state 2 are reversible. But, for a closed
system at rest,

8Q = dU + 8W
From Eq. 2-7a, 8W = p dV or

8Q = dU + p dV for reversible processes


Entropy 129

Substituting into Eq. 5-20,

dU + p dV
dS (5-21)
T

Since S is a property and is a function of U, p, V, and T, Eq. 5-21 is an


exact differential equation. It expresses a relationship between properties, and
hence, must be valid regardless of how those changes take place. Although a
reversible process was selected, because of convenience, to establish the
relationship between the properties, Eq. 5-21 must be valid for all changes of
state, whether reversible or not.
When a p-V-T relationship is available for any substance together with a
means of evaluating dU, the expression for dS can be put into a more useful
form. For example, consider an ideal gas: dU = mc„ dT and p/T = mR/V.
Then, for any process involving an ideal gas in a closed system,

mcv dT dV
dS + mR (5-22)
T V

Example 5-5. Four kilograms of air at a pressure of 1.72 bars and a


temperature of 34° C undergo a series of unknown processes until it reaches a
temperature of 282°C and a pressure of 1.72 bars. Determine the change in
entropy.

Solution. Since the final pressure equals the initial pressure, the actual
change in entropy equals the change in entropy for a reversible constant
pressure process between the initial and final states. Using Eq. 5-20,

mcp dT T
S2 mcp In jr
T 11

/282 + 273\
= 4 x 1000 x 1.004 In = 2430 J/K Answer
V 34 + 273 )

Example 5-6. Air is compressed in an unknown manner from 1.2 bars, 18°C
to 48 bars, 135°C. Determine the change in entropy per gram of air.

Solution

T2= (p2\<"'1,,n 135 + 273 _ /4S\tn~')ln


T, W ’ 18 + 273 V1 -2/

^ = 0.2428, n = 1.321
n

From Eq. 4-19,

R qio 0.2870 .. ,
cx = c„ + -= 0.718 + -j—TTTT ~ “0.J/g- C
1 —n 1 — 1.321
130 The Second Law and Entropy

From Eq. 5-20,

S2—S, = / = lx (-0.176) = -0.0573 J/g-K

Answer

Example 5-7. Determine the change in entropy per gram of air during its
compression in a diesel engine. The compression ratio is 15 to 1. The original
temperature is 35° C and the final temperature is 620° C.

Solution. Using Eq. 5-22,

T2 V-,
S2~ Si = mev In — + mR In —
I 1 V!

If precise results are required, the value of c„ as a function of temperature


can be calculated using data in Table 4-1. This value of cv can be substituted
into Eq. 5-22. Normally, however, since the temperature range in this example
is relatively low, an average value of may be approximated from Figure
4- 11. This value of c„ = 0.77 J/g-K. Then, for 1 g of air,

S2 - S, = 1 x 0.77 In 6325°^227733 + 1 X In ~ = 0.820 - 0.777

= 0.043 J/g-K Answer

Note: 8.314 is the gas constant per mole and 28.97 is the number of grams of
air per mole.

5- 6 ENTROPY AND IRREVERSIBILITY


Let a completely irreversible expansion take place from states 1 to state 2, as
shown in Figure 5-9. Assume that the expansion is that of an ideal gas in a
closed system. Since the expansion is irreversible, the exact path is not
known, and hence, is indicated by a dashed line. By definition there can be no

FIGURE 5-9 Reversible and irreversible iso¬


v thermal expansion.
Entropy and Irreversibility 131

useful work delivered in a completely irreversible expansion. Thus, the


internal energy at state 2 and also the temperature at that state must equal
those properties at state 1.
State 2 may also be reached by following a reversible isothermal process
from state 1. Heat must be added during this reversible expansion, and hence,
by Eq. 5-20, the entropy must increase, or, S2> S,. Since entropy is a function
of state, the change in entropy between states 1 and 2 is a fixed quantity.
Thus, there must also be an increase in entropy during the completely
irreversible expansion.
When a gas or vapor is compressed adiabatically in an actual process, the
process is irreversible. As such there must be an increase in entropy. This
may be demonstrated by use of Figure 5-10. Process 1-2 represents a
reversible adiabatic compression and process 1-3 the actual compression.
Because of irreversibility, the work input is higher, and hence, the internal
energy at state 3 exceeds that of state 2. State 3 may be reached from state 2
by the addition of heat. Then, by Eq. 5-20, S3 > S2. Since the entropy remains
constant in a reversible adiabatic compression, S2= STherefore, S3>S,.
Hence, there must be an increase in entropy during the actual compression.
Another type of an irreversible process is that of an electric current flowing
through a resistor. As a result of the current flow, there is an increase in the
temperature of the resistor. There will be an increase in the entropy of the
resistor equal to that which would occur if heat were to be added to the
resistor, which would produce the same temperature rise as was produced by
the electric current. Similarly, any irreversible process that produces a heating
effect will cause an increase in entropy.
In general, then, any irreversibility occurring within a system tends to
increase the entropy of the system. However, heat may be removed simul¬
taneously from the system, thus tending to decrease the entropy of the
system. Under these conditions, the entropy of the system may either
decrease or increase, depending on the magnitude of the two effects.

FIGURE 5-10 Reversible and irreversible adi¬


v abatic compression.
132 The Second Law and Entropy

5-7 USES FOR ENTROPY


Since it is a property, entropy can be used to determine the state of a
substance. When the entropy of a simple substance existing in a known phase
can be determined, the knowledge of one additional independent property will
fix the state of the substance.
A very common ideal process in connection with the compression or
expansion of vapors is a reversible adiabatic process. For such a process, Eq.
5-20 shows that there can be no change in entropy. For this reason, a
reversible adiabatic process frequently is called an isentropic process (i.e.,
one of constant entropy).3 Normally, the original state of the vapor is
specified but only the final pressure is available. For the isentropic process the
final entropy equals the original entropy, which is known if the original state
is known. Now two independent properties are available at the second state,
entropy and pressure. The final state thus is fixed and the other properties of
interest can be determined.
Because entropy is a property, it may be used as one coordinate of a graph
representing changes in state. The most common of these graphs is a tem¬
perature-entropy diagram. The process shown in Figure 5-9 is shown on a T-S
diagram in Figure 5-11, and the process of Figure 5-10 is shown in Figure 5-12.
A temperature-entropy diagram is of much value, since areas beneath a
reversible process on this plane represent the heat transferred during the
process. This is illustrated in Figure 5-13. The crosshatched area in this figure
equals T dS and the total area beneath the curve 1-2 equals /,2 T dS. From Eq.
5-20, the heat added can be found to be Q12 - Jj2 T dS.
The Carnot cycle is shown on a T-S diagram in Figure 5-14, with the state
points being numbered to agree with those of Figure 5-3. Heat is received
from the source during process 1-2. Area 1-2-6-5 represents this heat. Heat is

3It is conceivable that the removal of heat from a system may exactly neutralize the increase in
entropy produced by irreversibility within the system. Then the entropy of the system will remain
constant. This is very unusual. Hence, normally, the word isentropic denotes a reversible
adiabatic process.

FIGURE 5-11 Reversible and irreversible iso¬


thermal expansion.
Uses for Entropy 133

FIGURE 5-12 Reversible and irreversible adi¬


abatic compression.

FIGURE 5-13 Areas on T-S plane represent


heat transferred for reversible
processes.

S FIGURE 5-14 Carnot cycle on T-S plane.

rejected to the sink during process 3-4. Area 3-4-5-6 represents this heat. The
difference between these two areas, area 1-2-3-4, represents the cycle work.
An expression for the thermal efficiency of the Carnot cycle may be derived
by use of the T-S plane.

= W = area 1-2-3-4 = (TH - TL)(S2- S,) =TH-TL


Vt Qh areal-2-6-5 T„(S2-S,) T„

This expression for efficiency is identical to that derived in Section 5-2 (Eq.
5-4).
134 The Second Law and Entropy

5-8 VARIOUS FORMS OF THE SECOND LAW


As can be seen in the preceding sections, the second law is broad in its scope,
and hence, cannot be expressed in a simple statement. In essence, the second
law shows which processes can possibly take place and which processes
cannot. It also places a limitation on the amount of some forms of energy,
such as heat, which may be transformed into work. At the heart of the second
law is the property known as entropy, which was introduced in this chapter
Two interrelated statements relative to entropy and the second law are:

1 The entropy of an isolated system as well as the universe always tends to


increase.
2 No actual process, natural or man-made, can take place unless there is an
overall increase in entropy.
Although these statements do encompass the various aspects of the second
law, more explicit statements are desirable. Some of these statements are:
The Kelvin-Planck statement: “No cyclic process is possible whose sole
result is the flow of heat from a single heat reservoir and the performance of
an equivalent amount of work on a work reservoir.” This statement, of
course, is a restatement of the Carnot principle previously presented.
The Reeves statement: “Energy tends ever, so long as it undergoes no
transformation, to gravitate to a lower degree of intensity.” This statement is
in accord with our observations that heat flows readily from a region of high
temperature to one of lower temperature.
The Clausius statement: “It is impossible for a self-acting machine, un¬
aided by an external agency, to convey heat from one body to another at a
higher temperature.” This statement expresses our observations that heat will
not, of its own accord, flow “uphill,” that is, from a region of low temperature
to one at a higher temperature. However, it does not in any way preclude heat
being pumped from a region of low temperature to one at a higher tem¬
perature as is done in the case of mechanical refrigeration and of heat pumps.
The Kelvin statement: “It is impossible by means of an inanimate material
agency to derive mechanical effect from any portion of matter by cooling it
below the temperature of surrounding objects.” In accord with the Carnot
principle, a larger portion of heat supplied from a heat source to a heat engine
may be turned into work if the temperature of the heat sink is lowered. The
temperature of the heat sink may be lowered by use of a refrigerating device.
Theoretically, the gain in output of the heat engine obtained by the lowering
of the sink temperature exactly equals the work required for refrigeration to
produce the lower temperature. Any irreversibilities in either the heat engine
or in the refrigerating device mean a net loss in power output from the
system.
Problems 135

PROBLEMS _____

5-1 A Carnot cycle engine receives 1826kJ/min of heat at 420°C. It rejects


heat at 39°C. Determine the power output of the engine.
5-2 A Carnot cycle engine rejects heat at 42°C. The heat rejected is 2.42
times the work output. Determine: (a) thermal efficiency and (b) source
temperature.
5-3 A Carnot cycle engine has a sink temperature of 45°C. Its thermal
efficiency is 48.2 percent. The heat rejected is 742kJ/min. Determine:
(a) the source temperature, (b) the heat supplied per minute, and (c)
power output.
5-4 A reversed Carnot cycle engine is to produce 750 kJ of refrigeration per
minute at 12°C. It rejects heat at 42°C. Determine: (a) power required
and (b) heat rejected per minute.
5-5 A reversed Carnot cycle engine, used as a heat pump, delivers
980kJ/min of heat at 48°C. It receives heat at 18°C. Determine the
power input.
5-6 A Carnot cycle engine using air as the working substance receives heat
at 400° C and rejects it at 60° C. The minimum cylinder volume is
3.94 cm3. The maximum cycle pressure is 24 bars. During addition of
heat, the volume is doubled. Determine the pressure and volume at
each part of the cycle.
5-7 Determine the work done during each part of the cycle in Problem 5-6
and the thermal efficiency.
5-8 (a) Determine the work done per cubic centimeter of piston displace¬
ment in Problem 5-7.
(b) An engine operating under the same conditions of pressure and
temperature of Problem 5-6 delivers 30 kW at 2400 cycles per minute.
Determine the piston displacement per stroke.
5-9 Based on the results of Problem 5-8, make comments relative to the
desirability of trying to operate an engine on a cycle approaching the
Carnot.
5-10 A reportedly new design gas turbine has a mean combustion tem¬
perature of 840° C. The exhaust gases reject heat to the atmosphere at a
mean temperature of 395°C. The gas turbine is reported to deliver
3570 kW when using 798 liters of fuel per hour. The heating value of the
fuel is 40,300 kJ/liter. Is the reported performance reasonable?
5-11 In a Carnot cycle engine using air, the volume at the start of addition of
heat is 0.006 m3, the temperature is 675°C and the pressure is 82.7 bars.
The heat added is 17 kJ. Heat is rejected at 30°C. Determine the
maximum cylinder volume and the work done per unit of piston
displacement.
136 The Second Law and Entropy

5-12 A reversed Carnot cycle engine removes 140,000 kJ/h of heat from a
heat sink at 2°C and delivers heat at 55°C. Determine the power
required.
5-13 Same as Problem 5-12 except the sink temperature is -20°C.
5-14 In a Carnot cycle engine, the volume after the adiabatic compression is
8.5 percent of that before the compression. The sink temperature is
30°C. Determine the thermal efficiency.
5-15 The temperature of the hot junction of a thermoelectric generator is
750°C and the cold junction temperature is 150°C. Determine the
maximum possible power output when heat is supplied at the rate of
lOOOkJ/min (see Section 18-2 for a discussion of thermoelectric
generators).
5-16 The emitter (heat source) of a thermionic generator has a temperature
of 150Q°C and a collector (heat sink) temperature of 750°C. Determine
the minimum rate of heat input when the generator delivers 2.5 kW (see
Section 18-3 for a discussion of thermionic generators).
5-17 The radiation heat losses from the emitter of Problem 5-16 at 1500°C
are very high and the emitter life is relatively low. Repeat Problem 5-16
with an emitter temperature of 1200°C.
5-18 A thermionic generator is cascaded with a thermoelectric generator.
(The collector of the thermionic generator is the heat source for the
thermoelectric generator.) The temperature of the emitter of the ther¬
mionic generator is 1500°C and its collector temperature is 750° C. The
heat sink temperature for the thermoelectric generator is 150°C. The
thermionic generator produces 2.5 kW. For the ideal case determine: (a)
overall efficiency and (b) power output of the thermoelectric generator.
5-19 Gases enter a magnetohydrodynamic generator at 6.2 bars, 2700°C and
leave at 1.12 bars. Assume that the value of k of the gases for reversible
adiabatic expansion is 1.38. Determine: (a) theoretical gas temperature
at exit and (b) maximum thermal efficiency. (See Section 18-4 for a
discussion of magnetohydrodynamic generators.)
5-20 In the early development of the steam power plant approximately
1.12 kg of coal was required per kilowatt hour. Assume that the mean
temperature at which heat was supplied was 175° C and heat was
rejected at 100°C. Today assume that heat is supplied at a mean
temperature of 380° C and rejected at 32° C. Assume that the ratio of the
actual thermal efficiency to that of the Carnot cycle today is 1.15 times
that of the earlier years. Assume that the heating value of the coal is the
same for both cases. Calculate the amount of coal now required per
kilowatt hour.
5-21 (a) The sink temperature for a Carnot cycle engine is 30°C and the
source temperature is 200°C. Determine the thermal efficiency.
Problems 137

(b) The same as part (a) but for source temperatures of 600°C, 1000°C,
and 1500°C.
(c) Plot the thermal efficiency against source temperature.
5-22 An actual refrigerating system requires 3.2 times the amount of power
to produce a given amount of refrigeration at specified temperatures as
does a reversed Carnot cycle engine. Refrigeration is desired at — 10°C
and heat is rejected at 35°C. Determine the actual power required to
produce refrigeration at a rate of 500kJ/min.
5-23 A Carnot cycle engine using air has a clearance volume (the volume at
the end of adiabatic compression) of 0.005 m3. The maximum cylinder
volume is 0.05 m3. The source temperature is 300° C and the sink
temperature is 40°C. Determine the work per cycle if the pressure at
the start of the isothermal compression is 1.02 bars.
5-24 Repeat Problem 5-23 for source temperatures of 400° C and 500° C.
5-25 The engine in Problem 5-23 is used to produce refrigeration at 10°C.
The engine rejects heat at 40°C and the minimum pressure in the cycle
is 1.02 bars. Determine the amount of refrigeration produced per cycle
and the work required per unit of refrigeration.
5-26 The same as Problem 5-25 except for refrigeration produced at —10°C
and -30°C.
5-27 Is Eq. 5-3 precise for actual heat engines? Why?
5-28 (a) A kilogram of air is heated at constant pressure from 30°C to 80°C.
Calculate the change in entropy.
(b) If the heating in part (a) is done in a cylinder having a piston, will
piston friction change the answer?
5-29 The same as problem 5-28 except that the air is heated from 30°C to
1000°C.
5-30 2.25 kg of air are throttled from an initial temperature of 124°C and a
pressure of 21.5 bars to a pressure of 2.15 bars. Determine the change
in entropy. Hint: Assuming no net change in kinetic energy, what
happens overall to the enthalpy and temperature during the throttling?
5-31 A tank having a volume of 2.82 m3 contains air at 24.5 bars, 32.5°C.
Determine the change in entropy of the air if it receives 720 kJ of heat.
5-32 2.75 kg of air are heated in a reversible constant pressure process by
the addition of 140.5 kJ of heat. The initial temperature is 42°C.
Determine the change in entropy.
5-33 4.25 kg of air are heated in a reversible isothermal process. The air
temperature is 38.5°C. The increase in entropy is 1.5kJ/K. Determine
the ratio of the final to initial pressure.
5-34 Air is compressed from a pressure of 2.13 bars to a pressure of 8.05
bars, its temperature increasing from 36°C to 179°C. Determine the
change in entropy if there are 0.892 kg of air.
138 The Second Law and Entropy

5-35 3.2 kg of air are compressed in a reversible adiabatic manner in a closed


system, requiring 41.5 kJ of work. The air then cools at constant
volume until its initial temperature of 34°C is reached. Determine the
change in entropy of the air.
5-36 The air is cooled in Problem 5-35 by rejecting heat to the surroundings.
If the temperature of the surroundings is 22°C, determine the change in
entropy of the surroundings.
5-37 In a Carnot cycle engine, heat is added at 480°C, producing a change in
entropy of 5.65 kJ/K. The work delivered per cycle is 2360 kJ. Deter¬
mine the temperature at which heat is rejected.
5-38 4 kg of air are compressed from 40°C, 1.25 bars to 250°C, 8.75 bars. It is
then throttled until its pressure reaches 2.57 bars. Finally it is cooled
until the pressure becomes 1.25 bars and the temperature is 180°C.
Calculate the overall change in entropy.
5-39 A heat pump is to deliver 48,000 kJ/h of heat at a temperature of 46°C.
It receives heat at 8°C. It requires 2.5 times as much power input for
the heat pump as a Carnot cycle engine would for the same amount of
heat delivered. Determine the power input.
5-40 In a Carnot cycle heat is received at 480°C and rejected at 40°C. The
entropy of the sink increases 0.0785 kJ/K per cycle. Determine the
work done per cycle.
5-41 In a constant-temperature process occurring in a closed system com¬
posed of an ideal gas, the work delivered is equal to the heat added.
Does this statement violate the Kelvin-Planck statement of the second
law? Why?
5-42 A heat sink may be cooled below the temperature of the surroundings
by spraying it with water and allowing the water to vaporize. When this
is done, the efficiency of a heat engine rejecting heat to the heat sink
will increase. Do these statements conflict with the Kelvin statement of
the second law? Why?
5-43 Demonstrate that if the Clausius statement of the second law is true,
the Kelvin statement also must be true.
_ 6 _

SOME CONSEQUENCES
OF THE SECOND LAW

6-1 TEMPERATURE AND THE SECOND LAW


One of Carnot’s concepts was that the amount of work produced by a heat
engine receiving a specified amount of heat from a heat source is dependent
on the temperature difference between the heat source and the heat sink.
From this hypothesis, Kelvin reasoned that since the work is a function of
temperature differences, then temperature, as a concept, should be capable of
correlation with the work output of a Carnot engine. In this manner, it is
possible to conceive of temperature and its measurement as being in¬
dependent of the properties of any actual substance.
Kelvin’s concept may be visualized with the help of Figure 6-1. In Figure
6-1, the heat reservoir and the heat sinks are thermodynamically infinite; that
is, their temperatures remain constant. A quantity of heat Q, is transferred
from the heat reservoir at temperature T, to engine I, where work W, is
produced. A quantity of heat Q2 is rejected to heat sink I at a temperature T2.
Sink I acts as a heat reservoir for engine II. A quantity of heat equal to Q2 is
transferred from sink I to engine II. The temperature of heat sink II is
adjusted so that the work produced by engine II equals that of engine I. In a
similar manner, sink II becomes a heat reservoir for engine III, and the
temperature of heat sink III is adjusted so that the work of engine III equals
that of engine I. But the work of engine I, from Eqs. 5-1 and 5-4, is

Likewise, for engine II,

But

139
140 Some Consequences of the Second Law

{Q 3

Engine 111 -1
w3
{04

Heat sink 111


FIGURE 6-1 Concept of Kelvin scale of temperature.

Then

When the works of the two engines are equated, the result is

Thus,

T,-T2= T2-T3

This last equation states that when Carnot engines are operated in series
and the engines produce equal amounts of work, then the temperature
differences across the engines must be equal. The magnitude of the tem¬
perature difference (T,- T2) can be varied at will by varying the amount of
work produced by engine I. The quantity (Tj— T2) can be established as any
desired unit of temperature by establishing the amount of work to be
produced by engine I.
When enough Carnot engines are added in Figure 6-1, the total work
produced by all the engines will equal the heat transferred from the heat
reservoir. The temperature of the last heat sink, when all heat transferred
from the heat reservoir is converted into work, is the lowest conceivable
temperature.
No temperature lower than this temperature can be conceived since a lower
temperature will permit the addition of an additional heat engine, which could
produce an additional amount of work. The sum total of the work produced
would then exceed the heat supplied from the heat reservoir, thus violating
the first law.
Availability of Energy Entering a System 141

Although the Kelvin analysis is valuable since it gives a thermodynamic


concept of temperature and of the absolute zero of temperature independent
of the properties of actual substances, there is no direct way of utilizing this
concept in the measurement of temperature.

6-2 AVAILABLE AND UNAVAILABLE ENERGY


In a thermodynamic sense, available energy is defined as that energy which is
available for doing work. Any system existing at a temperature above ab¬
solute zero possesses energy. However, the system cannot do work by virtue
of its temperature unless its temperature exceeds that of its surroundings.
Neither can it do work by virtue of its pressure unless its pressure exceeds
that of its surroundings. The maximum amount of work that can be done by a
system existing at a given state is known as its available energy. The
remainder of its energy is said to be unavailable energy.
Based on the second-law concept is can be concluded that only a portion of
the heat added to a system can be used to do work, that is, only a portion of
the heat added to a system is available energy. The remainder of the heat is
unavailable energy.
There are three general problems in which the availability of energy will be
considered. They are:

1 The availability of energy that crosses the boundary of a system.


2 The energy that is made available when a change of state of a system
takes place.
3 The total available energy of a system.

6-3 AVAILABILITY OF ENERGY ENTERING A SYSTEM


Energy may enter a system either as heat or as work. In addition, when a fluid
or a solid enters a system, it carries its possessed energy with it. The
availability of this energy will be considered later on in this chapter. By the
very concept of available energy, the work that enters a system is available
energy. This statement does not infer that the work entering a system remains
as available energy within the system. Any irreversibility occurring within the
system will cause some of the available energy to become unavailable.
The actual process of transferring heat is an irreversible process. Heat
flows from a region of high temperature to one at a lower temperature and in
doing so loses some of its capacity to do work. Thus, it is necessary when
evaluating the amount of the available portion of a quantity of heat being
transferred to specify whether the availability of the energy is desired as it
leaves the high-temperature reservoir or as it enters the lower temperature
system.
The available portion of the heat that crosses the boundary of a system may
be determined by assuming that the given amount of heat is added to an
ideal heat engine in the same manner as it is added to the system. The work
142 Some Consequences of the Second Law

output of the ideal engine, by definition, is the available energy. In making


such a determination, it is assumed that the ideal heat engine rejects heat to a
sink at the temperature of the surroundings of the system. This temperature
commonly is known as the lowest available temperature. Normally, the
ideal heat engine selected to evaluate the available portion of heat crossing
the boundary of a system is the Carnot cycle engine.
In most cases, the temperature of the system changes as heat is added to it,
and hence, there is a change in temperature at the boundary of the system.
Thus, each small quantity of heat is added to the system at its own particular
temperature (see Figure 6-2). Consider the small quantity of heat, 8Q, added
at temperature T, which produces an entropy change dS. If this heat were to
be added to a Carnot engine at temperature T and the lowest available
temperature is TL, then the work of the Carnot engine and, hence, the
available portion of 5Q, is given as

available portion = 8Q(^ (6-1)

The remaining portion of SQ is unavailable, or

unavailable portion = 8Q — 8Q(^ ~ (6-2)

The shaded portion in Figure 6-2 represents the unavailable energy added to
the system. By integrating,

8Q
unavailable added = TL (6-3)
T
But 8Q/T = dS. Then

unavailable added = TL(S2- SO (6-4)

The available portion of the heat added to the system equals the difference
between the total heat added and the unavailable energy. Thus,

available added = Q — TL(S2- SO (6-5)

FIGURE 6-2 Addition of heat to a system.


Loss in Available Energy During Heat Transfer 143

Example 6-1. An amount of 90 kJ of heat is added to a tank of air that is


originally at a temperature of 26° C. The tank contains 2.8 kg of air. The
temperature of the surroundings is 15°C. Determine the available energy that
is added.

Solution. The process is a constant-volume one.

T - t Q 90
44.8° C
2 1 mc„ 2.8 x 0.718
T2 = 26 + 44.8 = 70.8° C or 344 K
Since

jo _
Lt kJ
&Q
r-j-<
_ mc*<
rj-,
dT ^

S2-S, = mc„ In 5 = 2.8 x 0.718 In ^ = 0.282 kJ/K


i i 2yy

Note: The units of cv are J/g-K = kJ/kg-K.

From Eq. 6-4,

unavailable energy added = T,(S2 - St) = 288(0.282) = 81.2 kJ


available energy added = 90 - 81.2 = 8.8 kJ Answer

In some cases, particularly for phase change conditions, the temperature of


the system remains constant during heat addition.

Example 6-2. An amount of 850 kJ of heat is added to a system during a


vaporization process occurring at a temperature of 180°C. The pressure is
held constant also. Determine the available energy added to the system if the
temperature of the surroundings is 22° C.

Solution. Since the pressure is held constant during vaporization, the tem¬
perature also remains constant. The change in entropy of the system,

c _ c = [ SQ = Qh = 850 = ] 876 kl/K


S2 Si J j T i8o + 273 L876kJ/K

unavailable energy added = TL(S2 - Si) = (22 + 273)1.876 = 534 kJ


available energy added = 850 - 534 = 316 kJ Answer

6-4 LOSS IN AVAILABLE ENERGY DURING


HEAT TRANSFER
Since a finite temperature difference must exist to cause a heat flow in an
actual case, there is a loss of available energy in all actual heat transfer
processes. It should be noted that there is no loss in the total energy involved.
Rather, as the result of heat flow from a region of high temperature to one at
lower temperature, some of the energy originally available for the production
144 Some Consequences of the Second Law

of work becomes unavailable during the heat transfer process see Figure 6-3.
The loss in available energy occurring when heat is transferred from a
high-temperature source to a lower temperature system may be determined by
finding the difference between the available energy leaving the source and the
available energy entering the system. This method is illustrated in the follow¬
ing example.

Example 6-3. Exhaust gases from a gas turbine are used to heat water. The
gases leave the turbine at 650°C and may be cooled to 145°C. The rate of gas
flow is 1510 kg/min and the rate of water flow is 1890 kg/min. The water enters
at 35°C. Assume that the mean specific heats of the gases and water are
1.088 J/g-K, and 4.27 J/g-K, respectively. The lowest available sink tem¬
perature is 32°C. Determine the loss of available energy resulting from the
heat transfer.

Solution. See Figure 6-3. The final water temperature can be obtained by
equating the heat given up by the gases to that received by the water. Thus,

mcp(T] — T2)gas = mcp(T2 — T\) water

or

1510 x 1.088(650- 145)


(T2-T,) water = 102.8° C
1890x4.27
water T2 = 137.8°C

The unavailable energy removed from the gas,

650 + 273
= (32 + 273)1510 x 1.088 ln( ) = 397,000 kJ/min
.145 + 273

FIGURE 6-3 Irreversible heat transfer.


Change in Available Energy of Systems 145

The unavailable energy added to the water,

Ounavail TLmc„ In ^ = (32 + 273)1890 x 4.27 ln(1^'^^3)

= 709,000 kJ/min
loss in available energy = gain in unavailable energy
= 709,000 — 397,000 = 312,000 kJ/min Answer

Example 6-4. Determine the loss in power as a result of the heat transfer in
Example 6-3.

Solution. Since 1 W = 1 J/s

loss in power = - = 5200 kW Answer

6-5 CHANGE IN AVAILABLE ENERGY OF SYSTEMS


When available energy crosses the boundaries of a system, there is a cor¬
responding change in the available energy of the system. The change in
available energy of the system, when energy is added to it, may be evaluated
by determining the amount of available energy in the energy that is added. If
the energy added is in the form of work, then the increase in available energy
of the system must be equal to the work added. If the energy added is in the
form of heat, then the available portion of the heat added may be determined
as was done in Section 6-4.

Example 6-5. Air in a cylinder containing a piston is heated at constant


pressure by the addition of 98 kJ/kg of air. The original air temperature is
28° C. The temperature of the surroundings is 26° C. Determine the change in
available energy per kilogram of air.

Solution. The final air temperature,

98
T2 = T, + -2
mcp
- = (28 + 273) +
1 x 1.005
= 398.5 K

unavailable energy added = TL(S2~ S0 = TLmcp ln(-^r


Ji
/T-.N
T\

= (26 + 273)1 x 1.005 = 86.3 kj

available energy added = 98 — 86.3 - 11.7 kJ

Because the process is a constant-pressure one, available energy, in the


form of work, will be delivered to the piston.

work = p(V2 — V,) = mR(T2- T,)= 1 x (398.5 - 301) = 28.0 kJ


Lo.y i

net change in available energy of the system = 11.7 - 28.0 = -16.3 kJ Answer
146 Some Consequences of the Second Law

Note: In an actual case, because of sidewall friction, a small amount of the


piston work will go back into the air as heat.

The unavailable energy of a system increases when heat is added to it. This
increase in unavailable energy equals the unavailable energy of the heat that
is added.

Example 6-6. Determine the increase in unavailable energy of the system in


Example 6-5 per kilogram of air.

Solution. In Example 6-5, it was found that the unavailable energy portion of
the heat added was 86.3 kJ/lb of air. Work, in the amount of 28.0 kJ, is
delivered to a piston. Lacking other information, it was assumed that this
amount of work was delivered from the system and that none of it became
unavailable because of sidewall piston friction. Then the change in unavail¬
able energy of the system equals 86.3 kJ. Answer

It is to be noted that even though available energy was added to the system
in Example 6-5, there was an overall decrease in the available energy of the
system because available energy, in an amount larger than that which was
added, was removed from the system in the form of work.
The unavailable energy of a system will also increase when an irreversible
process occurs within the system. For a definite change in the state of a
system, whether by addition of heat or by any irreversibility occurring within
the system, there is a very definite change in the unavailable energy of the
system. It should be noted that the change in unavailable energy of the
system, because of heat addition, was evaluated in terms of the entropy
change of the system. When any irreversibility that produces the same change
in entropy occurs within the system, there must be the same change in
unavailable energy of the system.

Example 6-7. Determine the change in available and unavailable energy


when 1.5 kg of air, at 28°C, undergo a free expansion. The temperature of the
surroundings is 25°C. The original pressure is 10.2 bars and the final pressure
is 1.01 bars.

Solution. Treat the air as an ideal gas and assume no heat transfer. Then
there will be no net change in temperature. From Eq. 5-22, the change in
entropy,

S? — S. = me
mffi) + mR ln(© = 0-1- mR In
©
= 0.995 kJ/K
change in unavailable energy = TL(S2~ Si) = (25 + 273)(0.995) = 296 kJ
Answer

Since no energy crossed the boundaries of the system, this increase in the
unavailable energy could have come only from the available energy of the
Change in Available Energy of Systems 147

system. Hence, the change in available energy of the system is negative and
equal to — 296 kJ. Answer

Frequently it is desired to evaluate the maximum possible amount of work


that a system can deliver as it changes from state 1 to state 2. In order that the
maximum amount of work may be obtained, the system must move from state
1 to state 2 in a reversible manner. Under this condition, the work obtainable
is a function of the change in some property of the system.
A special case is when the system is in temperature equilibrium with its
surroundings. Assume the system to be a closed one and that changes in
kinetic and potential energies may be neglected. For these assumptions, the
first law is
8Q = dU + 8W (6-6)

where 8W is the total of the mechanical and electrical work.


It is desirable here to introduce a compound property that frequently is
credited to Willard Gibbs, but generally is known as the Helmholtz function.
This function is designated as A and is defined as

A = U-TS (6-7)

From Eq. 6-7, dA = dU - T dS - S dT. Since the change in state must be a


reversible one to obtain the maximum work, T dS = 8Q, or,

dA = dU — 8Q - S dT

so that

dU = dA + 8Q + SdT

Substituting this value of dU into Eq. 6-6 and recognizing that the tem¬
perature is constant,
SQ = dA + §Q T 8W

Then

8W = -dA

and

max W = (A, - A2)t (6-8)

Equation 6-8 states that the maximum work performed by a system that
changes from state 1 to a final state without a net change in temperature
equals the decrease in the Helmholtz function.
The expansion of a system existing in the atmosphere causes work to be
done against the atmosphere. This work is part of the work given by Eq. 6-8.
This work p0dV, is not considered to be useful for most purposes (p0 is the
atmospheric pressure). Hence, it must be subtracted from the maximum work
of Eq. 6-8 to obtain the useful work, or,

8 Wusefu| = 8 Wtotal — Po dV (6-9)


148 Some Consequences of the Second Law

From Eqs. 6-6 and 6-9,

max SWusefui - 8Q ~ dU - p0 dV (6-10)

and
max WUSefui = (A, - A2)T - p0(V2- VO (6-10a)
The maximum useful work may be obtained directly for the special case
when initial and final temperatures are equal by the use of another function.
This function, known as the free-energy, or Gibbs function, is defined as

G = H - TS (6-11)

Then
dG = dH - T dS - S dT = dU + p dV+V dp-TdS-SdT

or
dU = dG - p dV - V dp+ TdS + SdT

Substituting this value of dU into Eq. 6-10 and assuming reversibility yields

SWuseful = TdS-dG + pdV+Vdp-TdS-SdT-p0dV

For the special case where the initial and final temperatures are equal and
where the pressure of the system is constant and equal to that of the
surroundings,

max SWusefui ' (dG)t,pq

or

max Wuseful = (Gj - G2)t,Po (6-12)


Equation 6-12 states that the maximum useful work deliverable by a system
that changes from an initial state to a final state with no net change in
temperature and with the pressure of the system equal to that of the
surroundings, is equal to the decrease in the Gibbs function of the system.
Equation 6-12 is of particular value in evaluating the maximum electrical
energy that can be delivered by a storage battery, such as a Daniels cell, or by
a fuel cell.1 In the Daniels cell, for low rates of discharge, the chemical action
approaches reversibility. Under these conditions, the temperature is sub¬
stantially constant. Furthermore, the pressure within the cell is substantially
equal to that of the surroundings. Under these conditions, the electrical
energy delivered by the cell approaches the decrease in the Gibbs function.
This concept can be adapted to the fuel cell. It should be noted that the output

‘in both the Daniels cell and the fuel cell a chemical reaction takes place, producing electric
energy. The supply of the reactants is self-contained in the Daniels cell. In the fuel cell, the
reactants are supplied to the cell and the products are removed from the cell in a continuous
manner.
Availability of a Closed System 149

of the Daniels cell and the fuel cell is not limited by the Carnot cycle concept,
since the Carnot cycle concept applies only to heat-work transformations.
The action in the Daniels cell and the fuel cell is a chemical energy-work
transformation.
Both the Helmholtz and Gibbs functions are potentials; they are driving
forces that can cause reactions to take place. It will be shown in Chapter 11
that a reaction cannot take place in an isolated system unless there can also
be a decrease in the Gibbs function as a result of the reaction.
It should be noted that both the Helmholtz and Gibbs functions contain a
TS term. Normally, during a chemical reaction there will be a change in the
entropy. The entropy of the products generally will not be equal to that of its
reactants, even at the same temperature. Thus, it is essential to use the
absolute values of entropy in determining the Helmholtz and Gibbs functions.
It is possible to determine the absolute values of entropy of a gas by a
step-by-step process. However, statistical thermodynamics, together with
quantum mechanics, provides a more satisfactory method of obtaining ab¬
solute values of entropy. This will be discussed in Chapter 16.

6-6 AVAILABILITY OF A CLOSED SYSTEM


The availability of a closed system is defined as the maximum amount of
work the system can deliver as it changes from its initial state to a state that is
in equilibrium, both pressure-wise and temperature-wise, with its surround¬
ings (the atmosphere). Since the amount of work obtainable from the system
is a function of the state of the system, any reversible process or series of
reversible processes may be selected to change the state of the system to that
of the surroundings. An accounting, however, must be made of all the
available energies transferred to or from the system. The problem may be
simplified by selecting processes such that heat is transferred only at the
temperature of the surroundings. In this way, no available energy will be
transferred as heat.
Consider a system existing at state 1 (Fig. 6-4). Now let the system expand

S FIGURE 6-4 Availability of a closed system.


150 Some Consequences of the Second Law

isentropically until its temperature reaches that of the surroundings, T0


(process 1-2). Since, in this figure, the pressure of the system at state 2 is less
than that of the surroundings, p0, let the system be compressed in a reversible
isothermal manner until its pressure is equal to p0 (process 2-0). Heat is
transferred during this process, but since the temperature is T0, none of this
heat is available energy. Available energy, then, is transferred to or from the
system in the form of work. The algebraic summation of these works is the
maximum work the system can perform and, hence, is the original available
energy of the system. Applying the first law to the combination of the two
processes,

Q= U0-U\ + Wmax

or

Wmax = Q + 17, - U0
But

Q = T0(S0- S2) = T0(S0- Si)


Then

^max — T0S0 ~ T0S i + U\~ Uq


or

VVmax = (17, - T0S,) - (Uo- T0S0) (6-13)

Since 17, T, and S are properties of the system and are dependent on state
only, the maximum work as determined by the use of Eq. 6-13 is the available
energy of the system in its original state.

Example 6-8. Air exists in a tank at a pressure of 3.8 bars and at a


temperature of 190°C. The mass of air present is 8.2 kg. The surroundings are
at a pressure of 1.005 bars and a temperature of 22°C. Determine the available
energy of the air.

Solution. See Figure 6-4. The available energy equals the maximum work
that can be delivered as the state of the system changes to that of the
surroundings. From Eq. 6-13,

= T0(So - St) + 17, - U0 = T0(S0 ~ S2) +17,-170

Since process 2-0 is at constant temperature,

But at constant temperature,

Pi
Q = mRT0 In
Po
Entropy and Unavailable Energy 151

Then

T„(S0-S2)=mRT„ln|2
*0

Substituting,

Wm = mRTo In ^ + mcu(Tl - T0)


iQ

Then

available energy = the maximum work

= 8.2 x x 293 + 8.2 x 0.718(190 - 22) = 818.7 kj

Answer

6-7 ENTROPY AND UNAVAILABLE ENERGY


It has been shown in this chapter that when heat is added to a system or when
any irreversibility occurs within the system, these actions cause an increase in
the unavailable energy within the system. For a given temperature of the
surroundings, the increase in unavailable energy is directly proportional to the
increase in entropy. Thus, it can be said that entropy is a direct measure of
unavailable energy.
When reversible processes occur within an isolated system, there is no
change in entropy. Hence, there can be no change in the unavailable energy of
the system. If heat is added in a reversible manner to a system, its entropy,
and hence, its unavailable energy, increases. But there is a corresponding
decrease in the entropy and thus in the unavailable energy of the source that
supplied the heat. Consequently, there is no net change in the unavailable
energy of the universe itself. All actual processes, man-made and natural, are
irreversible. As such they cause a net increase in the entropy, and hence, in
the unavailable energy of the universe. Available energy, energy that could be
used to produce work, is continuously being transformed into unavailable
energy.
Our civilization rests on the energy we draw from nature, mostly in the
form of fuels, fossil and nuclear. When these fuels are burned, a sizable
portion of their energy becomes available energy—energy that can be used to
do work. But in the utilization of this energy, irreversible processes take
place, with a sizable portion of the available energy, as such, being lost. The
magnitude of this loss can be readily determined by calculating the entropy
change resulting from the process. This concept may be applied to Example
6-3.

Example 6-9. Rework Example 6-3 by first finding the change in entropy of
the gases and the water.
152 Some Consequences of the Second Law

Solution. The change in entropy of the gases

S2~ Si = mcp In jr = -mcp In^r^ = -1510 x 1.088

= -1301 kJ/min-K

The change in entropy of the water

S2- S, = mCp In ^ = 1890 x 4.27 ln(|+) = 2324 kJ/min-K

loss in available energy = gain in unavailable energy


= Tl (net change in entropy)
= 305(2324 — 1301) = 312,000 kJ/min Answer

Example 6-10. Follow the procedures of Examples 6-3 and 6-9, with the
exception that the hot gases are used to boil water at a temperature of 120°C.

Solution

heat given up by the gases = mCp(T} - T2) = 1510 x 1.088(650 - 145)


= 830,000 kJ/min
Q _ 830,000
change in entropy of the water (for constant temperature)
T 120 + 273
= 2111 kJ/min-K
loss in available energy = gain in unavailable energy = TL(change in entropy)
= 305(2111— 1301) = 247,000 kJ/min Answer
Problems 153

PROBLEMS ___
6-1 A Carnot cycle engine receiving heat at 200°C and rejecting heat at
20°C delivers 10.8422 kJ of work when its heat input is 28.5 kJ. Cal¬
culate the conversion factor for changing degrees Celsius to degrees
Kelvin.
6-2 4.85 kg of air are heated in a closed system from 30°C to 90°C at
constant pressure. The temperature of the surroundings is 28°C.
Determine: (a) change in entropy of the air, and (b) available energy
added to air.
6-3 Same as Problem 6-2 but the air is heated as it flows through a heater.
6-4 Boiling water enters a boiler at 240°C at the rate of 4270 kg/min. During
the boiling process, 7,543,000 kJ of heat are added. The temperature of
the surroundings is 25°C. Calculate the available energy added. (The
temperature will not change during boiling.)
6-5 In an Otto cycle engine, heat is rejected to the atmosphere at constant
volume, the temperature decreasing from 190°C to 40°C. The tem¬
perature of the surroundings is 24° C. Air is the working substance in
the engine. Determine the available energy removed per kilogram of air.
6-6 The condensing temperature in the condenser of a steam power plant is
30°C. The temperature of the surroundings is 22°C. Steam flows to the
condenser at the rate of 1,287,000 kg/h. Each kilogram gives up 2190 kJ
of heat as it condenses. (Condensation takes place at constant tem¬
perature.) Express the available energy given up in the condenser in
terms of kilowatts.
6-7 In an air heater, air is heated from 45°C to 245°C by gases that cool
from 345°C to 150°C. Assume that the mean specific heat of the air is
1.017 J/g and of the gases, 1.035 J/g. The temperature of the surround¬
ings is 30° C. Determine the loss of available energy per kilogram of air
as a result of the heat transfer. Note: First find the kilograms of gas per
kilogram of air.
6-8 A tank of water, containing 893 kg, is heated from 40°C to 80°C by
steam condensing in a steam coil at 120°C. The temperature of the
surroundings is 28°C. Determine the loss of available energy resulting
from the heat transfer.
6-9 A 2.3-kg mass of iron having a temperature of 870°C is quenched by
immersing it in a tank containing 74.5 kg of water with a temperature of
28°C. The specific heat of iron is 0.465 J/gK. Determine: (a) net change
in entropy and (b) loss of available energy if the temperature of the
surroundings is 28° C.
6-10 A constant-temperature source supplies 24,000 kJ/min to a heat engine.
The source temperature is 665°C, the lowest available temperature is
24°C and the mean temperature at which the engine receives heat is
154 Some Consequences of the Second Law

480°C. Determine the loss of available energy resulting from the heat
transfer.
6-11 (a) Determine the maximum power that can be produced by the engine
in Problem 6-10.
(b) Determine the potential percentage loss of power because of heat
transfer.
6-12 During a reversible expansion of a system 65 kJ of heat are added to
the system and the system delivers 65 kJ of work. Recognizing that only
part of the heat added is available energy and all of the work is
available energy, explain how this is possible.
6-13 Liquid sodium is used to transfer heat from a nuclear reactor to
produce steam. Steam is produced at 326°C at a rate of 482,000 kg/h.
The steam picks up 1181.3 J/g. The sodium is cooled from 650°C to
345°C. If the temperature of the surroundings is 30°C, determine the
loss of available energy per hour because of heat transfer. Determine
also the maximum amount of power that can be produced by this
available energy.
6-14 Air is cooled in the intercooler of a two-stage air compressor from
105°C to 27°C. The temperature of the surroundings is 25°C. The rate
of air flow is 75.8kg/min. Determine the loss in available energy
resulting from the heat transfer.
6-15 Air at 14.5 bars 60°C is throttled to 1.15 bars. The temperature of the
surroundings is 24°C. Assume no heat lost and no net change in kinetic
energy. The air flow is 28.5 kg/min. Determine the change in entropy
and the loss in available energy.
6-16 For control purposes, steam is throttled as it enters a steam turbine,
producing an increase in entropy of 0.29kJ/kg-K. The temperature of
the surroundings is 26°C. The rate of steam flow is 52,500 kg/h. Deter¬
mine the loss in available energy resulting from the throttling and the
maximum loss in power.
6-17 A heat engine receives 1700 kJ of heat per minute from a heat source
whose temperature remains constant at 430°C. The engine receives heat
at a mean temperature of 402°C and rejects heat at a mean temperature
of 40°C. The sink temperature is 33°C. The efficiency of the engine is 55
percent of that of the Carnot engine for the given engine temperatures.
Determine: (a) power lost by heat transfer to the engine, (b) power lost
because the engine is less efficient than the Carnot, and (c) power lost
by heat transfer to the sink. Make an accounting of the available energy
taken from the source.
6-18 52 kJ of work are added to a system as its entropy increases 0.38 kJ/K.
The temperature of the surroundings is 24°C. Determine the change in
the available energy of the system.
6-19 (a) Air expands from 8.1 bars, 212°C to 1.75 bars according to the law
Problems 155

1 36 ’
pV = C. The temperature of the surroundings is 28°C. Determine, per
kilogram of air, the increase in the unavailable energy of the air.
(b) Is it necessary to know if the process in part (a) is reversible?
6-20 Water is to be heated from 27°C to 90°C at the rate of 1050 kg/h by
gases whose temperature drops from 425°C to 95°C. The temperature
of the surroundings is 23°C. Determine the loss of available energy as a
result of the heat transfer. Note: The specific heat and the mass rate of
flow of the gases are unknown.
6-21 Air enters an air turbine at 6.95 bars, 180°C and leaves at 1.28 bars.
Determine the maximum change in available energy per kilogram of air,
assuming no heat is removed.
6-22 In Problem 6-21, the air actually leaves at a temperature of 45°C.
Determine: (a) work delivered per kilogram of air and (b) change in
available energy per kilogram of air.
6-23 It is proposed to cool a compression-ignition (diesel) engine by produc¬
ing steam in the “water” jackets. The steam is to be used in a separate
cylinder to produce work. When the engine is developing 1850 kW, the
heat flow to the jackets is approximately equivalent to the work. Steam
is produced at 110°C. The temperature of the surroundings is 30°C.
Determine the maximum power that can be produced by the steam. Do
you recommend this arrangement?
6-24 In the regenerator of a gas turbine (see section 13-10), the air is heated
before its entrance to the combustion chamber by the exhaust gases
from the turbine. The air flow is 2850kg/min. There are 54 g of air per
gram of fuel. The air is heated from 132°C to 366°C. The gases enter at
534°C. The lowest available temperature is 24°C. Determine the maxi¬
mum amount of power that could be produced from the available
energy lost by heat transfer. Take the value of the specific heat of the
air at its mean temperature. The mean specific heat of the gases is
1.075 J/g-K. Hint: find the temperature of the exit gases.
6-25 Air exists in a closed system at 13.8 bars, 94°C and occupies a volume
of 1.71 m3. If the surroundings are 1.02 bars, 20°C, determine the
available energy of the system.
6-26 A system undergoes a series of processes and returns to its original
state. The net change in the Gibbs function must be zero. Does this
mean that there is no net work done? Why?
6-27 A compressor compresses air according to the law pV135 = C. Air
enters the compressor at 1.02 bars, 30°C and leaves at a pressure of
6.78 bars. The rate of air flow into the compressor is 6.75 nr/min. The
temperature of the surroundings is 30°C. Determine: (a) heat lost from
the air per minute, (b) power input, (c) unavailable energy change of the
air, and (d) available energy change of the air.
_ 7 _

PROBABILITY AND THE


NATURE OF ENTROPY

7-1 INTRODUCTION
In Chapter 5 the concept of the property known as entropy was developed
from a macroscopic viewpoint. In Chapters 5 and 6, engineering uses of this
property were presented. In addition, methods of determining entropy
changes were developed and the entropy changes were correlated to changes
in unavailable energy. Thus the engineer has been given means of utilizing
this very important property.
However, we could not investigate the fundamental nature of entropy,
since macroscopic thermodynamics does not lend itself to such an in¬
vestigation. In this chapter, we will use the microscopic approach to gain an
understanding of the nature of entropy.

7-2 THE MICROSCOPIC APPROACH


Consider a simple gas. We may use instruments to measure properties such as
temperature and pressure. Other properties, such as density, internal energy,
enthalpy, and entropy may be determined from the observed properties. If it
is suspected that there are variations in properties throughout the system,
measurements may be made of the properties at several points throughout the
system and then these properties may be averaged in an appropriate manner
to obtain the mean properties. This approach, which assumes matter to be
continuous, is quite satisfactory for most engineering problems.
Under ordinary conditions of temperature and pressure, even a small
volume of the gas contains an exceedingly large number of molecules. These
molecules have random velocities in random directions. In reality, our
measuring devices indicate average properties of the molecules in direct
contact with the measuring devices. This assumption is subject to question
only when the gas pressure is extremely low, as it is in interstellar space. If
only a relatively very small number of molecules are involved, the properties
of those molecules in direct contact with the measuring devices may differ
from those in the vicinity of the devices. Under these conditions it becomes
necessary to consider the probability of the measuring devices indicating the

156
Probability 157

true properties. This probability can be determined only by a study of


microscopic thermodynamics.
Even more importantly, microscopic thermodynamics helps to explain why
matter behaves the way it does. In addition, microscopic thermodynamics is
very useful in predicting how matter will behave when the conditions under
which it exists are changed.
In microscopic thermodynamics, it is recognized that matter is composed of
molecules which are formed from atoms. The atoms are composed of elec¬
trons, protons, and, except for hydrogen, neutrons. Unless heated to elevated
temperatures, there is no change in either the atomic or molecular structure of
a simple gas. For the sake of simplicity, it will be assumed in the microscopic
analysis that follows, that primary consideration will be given to the behavior
of molecules as composite wholes.
Since the behavior of a system depends on the behavior of its individual
components, it may seem reasonable to determine the behavior of each
molecule and then to make a summation. There are two difficulties with such
a procedure. First, the velocity and the position of each and every molecule
change rapidly with time. Furthermore, the number of molecules present in
even an extremely small volume is so very large that it is wholly unrealistic to
consider the behavior of each molecule individually. (A cube, one thousandth
of a centimeter on a side, contains approximately 2.464 x 1010 molecules of an
ideal gas at standard atmospheric pressure and 25°C.)
The alternative to attempting to deal with the behavior of each individual
molecule is to establish a method of predicting the probability of the behavior
of a group of molecules as a composite whole. In the discussion that follows,
it will be shown that when dealing with a very large number of molecules, the
laws of probability will predict their behavior with a high degree of accuracy.

7-3 PROBABILITY
The probability of predicting the outcome of an occurrence is dependent on
the number of objects involved in the occurrence. Consider two small equally
sized chambers, A and B, each containing a molecule, a and b (see Fig. 7-1).
Following common practice, these chambers are designated as cells. Let the
two cells be moved together and let the common wall be pierced to permit
the molecules access to both chambers. Since the molecules have high
velocities, they will have a tendency to move between the two cells. At any
given instant, there is a fifty-fifty chance that molecule a will be in the cell B.
The same statement may be made about molecule b. Thus, there are four
possible microscopic states for the two molecules, as shown in Figure 7-2.

a
FIGURE 7-1 Two molecules in two cells, A and B.
158 Probability and the Nature of Entropy

FIGURE 7-2 Possible microscopic states for two molecules in two cells, A and B.

Each of the four possibilities is equally probable. If we were to predict that


the two molecules would divide equally between the two cells, there are two
chances in four that we would be right and also two chances in four that we
would be completely wrong.
Now consider four molecules in the two cells. There are 16 possible
microscopic states for the four molecules, with each state being equally
probable. The 16 possible microscopic states are shown in Figure 7-3.
Normally, it might be expected that there will be two molecules in each cell.
As shown in Figure 7-3, there are 6 chances in 16 that this will be the actual
distribution and 8 more chances in 16 that there will be at least one molecule
in each cell. There are 2 chances in 16 that there will be no molecules in one
cell, which is as far as possible from the normal prediction.
Although we may predict the probable distribution of a larger number of
molecules by preparing figures similar to Figures 7-2 and 7-3, such a pro¬
cedure becomes exceedingly laborious as the number of molecules is in¬
creased. Hence, an investigation will be made of the laws of probability
governing the distribution of molecules.
First it is necessary to distinguish between a macroscopic state and a
microscopic state. For the microscopic state, each molecule is distinguishable
from all other molecules. For the macroscopic state, the molecules are not
distinguishable. For example, in the second, third, fourth and fifth figures of
Figure 7-3, there is a distinct difference on the microscopic basis. But on a
macroscopic basis, these four figures are identical, since there are three
molecules in cell A and one in cell B. Consider the case of four molecules.
There are five possible macroscopic states, namely: (1) four molecules in cell
A, (2) three molecules in cell A and one in cell B, (3) two molecules in each

A B A B A B A B A B A B
ab 1 ab 1 ac ab be 1 ab 1 cd
cd c d _d_|_b_ d c d
1 1 a
A B A B A B A B A B A B
ac 1 bd ad 1 be be I ad bd ac cd 1 ab a 1 be
1 1 1 d

A B A B A B A B
b 1 ac c i ab d I ab ab
1 d 1 d cd

FIGURE 7-3 Possible microscopic states of four molecules in two cells.


Probability 159

cell, (4) one molecule in cell A and three in cell B, and (5) four molecules in
cell B. The possible number of macroscopic states is given by the following
equation:
(N + n - 1)!
W' = (7-1)
(N — 1)! n !

where W' = the possible number of macroscopic states


N = the number of cells
n — the number of molecules

On the microscopic basis, each molecule behaves independently of the


other molecules. There is a fifty-fifty chance that molecule a will be in cell A
and an equal chance that it will be in cell B. The same statement may be made
about each of the other three molecules. Thus, as shown in Figure 7-3, there
are six different microscopic states possible for the one macroscopic state of
equal division of four molecules between the two cells. The possible number
of microscopic states for a given macroscopic state is given as

op = n n
(7-2)
nA \ nBl... ft*! product of n,!
where & = the possible number of microscopic states for a given macroscopic state
nA the number of molecules in cell A
Hb the number of molecules in cell B
rii the number of molecules in cell i
n the total number of molecules

It should be noted that when there are no molecules in either cell A or B,


there will be a 0! in the denominator. 0! must be evaluated as unity.
In Figure 7-3, it is shown that there is a total of 16 possible microscopic
states. This fact is expressed as
= Nn (7-3)

where & — the total number of possible microscopic states


N = the number of cells

It should be emphasized that each state is equally probable.


It is evident from Eq. 7-3 that as the number of molecules is increased,
there is a tremendous increase in the total possible number of microscopic
states. For example, when there are 20 molecules, there are 1,048,576 possible
microscopic states compared with 16 for four molecules. With an increase in
the number of molecules, the probability of finding one cell or the other
completely empty approaches zero. Thus, with 20 molecules, there are only 2
chances out of 1,048,576 that one cell or the other will be empty.
On the other hand, an increase in the number of molecules greatly increases
the probability of somewhere near equal division of molecules between the
160 Probability and the Nature of Entropy

two cells. For example, with a total of 10 molecules, the probability of finding
at least 40 percent of the molecules in each cell is approximately 656 chances
out of a thousand. With 20 molecules, the probability increases to ap¬
proximately 737 out of 1000 that there will be at least 40 percent of the
molecules in each cell. For two very small ceils, each having a dimension of
one thousandth of a centimeter on a side and existing at standard atmospheric
pressure and 25°C, the total possible number of microscopic states equals
24.928 x 1010. This astronomical figure means that there is almost absolutely
no chance that all the molecules will collect in one cell. Furthermore, there is
an extremely high probability that the division of the molecules between the
two cells will be so nearly equal that it will not be possible to detect a
difference in distribution by any common measuring device.
It is true, by the laws of probability, that there is a possibility of a
significant unbalance in the molecular distribution between the two cells. But,
even for the very small cells under consideration here, such a possibility is
extremely remote. Furthermore, because of the very rapid movement of the
molecules from cell to cell, such an unbalance will last for such a small
fraction of a second that the unbalance cannot be detected by normal
measuring devices.
On the macroscopic basis, when a system is caused to depart from
equilibrium conditions, there is an unbalance of forces acting on the system,
tending to cause equilibrium to be restored. On a microscopic basis, when a
system is not in a state of equilibrium, the molecules are not in their most
probable state. Consider the two interconnected cells in Figure 7-1. Assume
that there are a large number of molecules present. Let all of the molecules be
moved into cell A and then let the system be isolated. This is a most
improbable state. The molecules will move until the most probable state is
reached, which is substantially equal distribution of the molecules. Thus, an
isolated system always tends to move from a least probable state to the most
probable state.

Example 7-1. Determine the possible number of macroscopic states for 15


molecules in two interconnecting cells of equal volumes.
Solution. Using Eq. 7-1,

= (N + n — 1)! = (2+15-1)!
Answer
(N-l)lnl 1! 15!

Example 7-2. Determine the possible number of microscopic states in


Example 7-1.
Solution. Using Eq. 7-3,

& — Nn = 215 = 32,768 Answer

Example 7-3. Determine the possible number of microscopic states for Exs.
7-1 and 7-2 for 8 molecules in cell A and 7 molecules in cell B.
Probability and Entropy 161

Solution. Using Eq. 7-2,

n’ 15’
& = —.-- = = 6435 Answer
nA!nB! 8!7!

7-4 PROBABILITY AND ENTROPY


In Chapter 5, it was stated that all irreversible processes produce increases in
entropy. A common irreversible process is a free-expansion one (see Fig. 7-4).
Cylinder A contains a gas under high pressure. Cylinder B is completely
evacuated. When the valve between the two cylinders is open, gas rushes into
cylinder B until the pressure between the two cylinders becomes equalized. If
the system is isolated, its internal energy cannot change. Assuming that the
gas is an ideal one, then there can be no net change in its temperature. The
change in entropy resulting from the free expansion may be found by using
Eq. 5-22, which states that, for an ideal gas,

S2~ Si = mcv In -=? + mR In jr


1 1 V 1

Since there is no net change in temperature, the change in entropy for the free
expansion of an ideal gas,

V?
S2~ Si = mR —
Vi

During the free-expansion process, the molecules moved from a highly


improbable state to the most probable state. Equation 7-3 may be used to
evaluate the probabilities of each state. This equation involves the quantity,
N, which is the number of cells under consideration. Since these cells must be
of equal volumes, Eq. 7-3 may be rewritten as

= V" (7-4)

where V is the total number of units of volumes under consideration.


After the valve in Figure 7-4 has been opened, the possible number of
microscopic states equals (VT/VA)" where VT is the total volume and VA is the
volume of cylinder A. The probability of finding all the molecules in cylinder
A is one chance out of (VT/VA)n or the probability of this occurrence equals

FIGURE 7-4 Free-expansion system.


162 Probability and the Nature of Entropy

(VA/VT)". The probability of finding all the molecules in the total volume is
unity. The ratio of the probability ^2 of the final state to that of the initial
state, is
0>2_ 1
(7-5)
(?^ (v,/v2y
or

In tyj- In = n lnf^r^ (7-6)


\ V i/

For one mole of gas, n = n0, the number of molecules per mole. Then

\n(^A = (7-7)
\VJ no
On the mole basis the change in entropy,

s2-s, = R0ln(^) (7-8)

Substituting Eq. 7-7 into 7-8,

s2— s, = — (In &-i-In 3*,) (7-9)


no

Since Roln0= k, the gas constant per molecule,

s2 5, — k(\n ^2-In 0>,) (7-10)


Equation 7-10 states that when an isolated system moves from a less
probable state to a more probable state, the increase in entropy is directly
proportional to the increase in the natural logarithm of the probability of the
state of the system.
Return to the free-expansion process. Before the valve was opened, each
and every molecule was in cylinder A. Their positions could be specified.
After the free-expansion process, it is not possible to specify whether any
given molecule is in cylinder A or in cylinder B at any specified instant. Thus,
we say that the final state of the system is a random one or a disorderly one.
The system passed from a more orderly state to a less orderly state. Fur¬
thermore, the system passed from a less probable state to a more probable
state. In so doing, there was an increase in entropy of the system.
The concepts developed here for a free-expansion process may be adapted
to all irreversible processes. In all such processes, the system proceeds from
an orderly state to a less orderly state and from a state of low probability to
its most probable state. This change of state is accompanied by an increase in
entropy, the change in entropy being a function of the probabilities of exis¬
tence of the initial and final states.
When an isolated system is not in its most probable state, it tends to move
to this state. In so doing, its entropy must increase. When the system attains
Entropy and the Third Law 163

its most probable state, its entropy cannot increase. Thus, if an isolated
system cannot experience a change in entropy, it must be in its most probable
state. But if the entropy of the system can increase, the system cannot be in
its most probable state, and hence, the system is not in equilibrium. These
facts may be used in determining whether or not a system is in equilibrium. If
calculations show that the entropy of an isolated system can increase, then
the system cannot be in equilibrium.

7-5 ENTROPY AND THE THIRD LAW


Some authors hesitate to call that which is known as the Third Law of
Thermodynamics a law. The first law deals with energy relationships during
energy exchanges and transformations. The second law places certain limita¬
tions on energy transformations and establishes a new property, entropy,
which is needed to deal with energy transformations. The third law presents
no concepts about new areas of thermodynamics. However, since it for¬
mulates a conclusion relative to an important area of thermodynamics that is
in agreement with our experiences, it will be treated as a law in this text.
The concepts of the third law were put forth first by Nernst in 1906 as his
heat theorem. This work was augmented by Planck in 1911. Perhaps the
all-inclusive statement of what is now recognized as the third law was made
by Lewis and Randall in 1923.1 Their statement is: “If the entropy of each
element in some crystalline state be taken as zero at the absolute zero of
temperature, every substance has a finite positive entropy; but at the absolute
zero of temperature the entropy may become zero and does so become in the
case of perfect crystalline substances.”
It is difficult to prove the third law. However, there are two ways of
arriving at the reasonableness of this law. Two or more elements that exist as
pure crystalline substances at absolute zero of temperature may be heated,
changed into liquids, further heated, vaporized, and then combined chemic¬
ally. The resultant products may be cooled down, liquefied, solidified, and
finally cooled down to absolute zero of temperature. If the product becomes a
pure crystalline substance, the result will be zero when the entropy changes
for all these processes are evaluated and summed up. This means that the
entropies of all pure crystalline substances are equal at zero degrees absolute.
Either this value of entropy is zero, or it may be treated as such in all
problems.
The second approach to determining the value of the entropy of a pure
crystalline substance at an absolute zero of temperature is to make use of a
postulate formulated by L. Boltzmann in 1890 and M. Planck in 1912. This
postulate is

s = k In0> (7-11)

'G. N. Lewis and M. Randall, Thermodynamics and the Free Energy of Chemical Substances, 1st
ed., McGraw-Hill, New York, 1923, p. 448.
164 Probability and the Nature of Entropy

There is no rigid proof of this postulate but experience has shown its validity.
A pure crystalline substance is perfectly ordered at a temperature of zero
degrees absolute, that is, there is one probable state. Since & is then unity and
In & = 0, the entropy must be zero at zero degrees absolute temperature.
When changes in entropy are desired for systems in which chemical
reactions are not involved, these changes can be determined by the methods
presented in Chapter 5. However, when chemical reactions are involved,
absolute values of entropy of both the reacting substances and the products
of reactions are required. Absolute values of entropy of a gas may be
obtained by recognizing that the absolute value of entropy of a pure crystal¬
line substance is zero at zero degrees absolute and by finding the summation
of (1) the change in entropy as the substance is heated to its fusion tem¬
perature, (2) the entropy change during melting, (3) the change in entropy as
the liquid is heated to the vaporization temperature, (4) the entropy change
during vaporization, and (5) the entropy change as the gas is heated to the
temperature at which its entropy is desired. This method of determining the
absolute entropy of a gas is a laborious one and, furthermore, requires an
accurate knowledge of the specific heats of the solid, the liquid, and the gas,
the fusion and vaporization temperatures, and also the latent heats of both
fusion and vaporization.
Statistical thermodynamics provides a method for the direct determination
of the absolute value of entropy. This method will be discussed in a later
chapter.
Problems 165

PROBLEMS ____
7-1 (a) Consider the case of molecules existing in two interconnecting cells.
Determine the possible number of macroscopic states for 2, 4, 6, 8, and
10 molecules.
(b) Determine the possible number of microscopic states for part (a).
7-2 Determine for Problem 7-1 the probability that one or the other of the
two cells will be vacant.
7-3 Determine for the case of molecules existing in two interconnecting cells
the probability of there being at least one-third of the molecules in each
cell for 6, 9, and 12 molecules.
7-4 Nitrogen occupies a cube at 1.02 bars, 30°C. The cube is 0.1cm on a
side. Determine the number of molecules present. See Section 3-3.
7-5 (a) For 6 molecules in two equal-size cells, plot the possible division in
either cell against its microscopic probability (as a fraction). For exam¬
ple, for 4 molecules in one cell, the probability is 30 out of 64 or 0.469.
(b) The same as (a) except for 10 molecules.
Comment on these plots.
_ 8 _

GENERAL EQUATIONS
OF THERMODYNAMICS

8-1 INTRODUCTION
Certain properties of substances, such as pressure, temperature, volume, and
mass, are relatively easy to measure. Other properties, such as internal
energy, enthalpy, and entropy generally cannot be measured directly but must
be calculated. Methods of determining the changes in these latter properties,
rather than their absolute values, are both acceptable and desirable, since
most engineering problems involve only changes in state. Hence, absolute
values of these properties are not required. In Chapter 4, we developed the
means for the determination of the changes in the internal energy and
enthalpy of an ideal gas. After a discussion of the second law, methods were
developed for the changes in entropy of ideal gases.
Although there are many problems dealing with those gases that show very
little deviation from ideal-gas behavior, there are many more problems in
which serious errors will be made by the use of ideal-gas laws. In this chapter,
we develop the general methods of determining changes in the properties of
substances in all phases: solid, liquid, and gaseous. These changes will be
evaluated in terms of changes in pressures, temperatures, volumes and, in
some cases, specific heats. The relationships between these simple properties
were presented in Chapter 4. It will become evident that, for the conditions
under consideration, when relatively simple relations exist between these
simple properties, the equations developed in this chapter will provide a
feasible means of determining changes in the complex properties. It will also
be shown that, when p v = RT, the general relationships developed in this
chapter will reduce to the relationships developed previously for an ideal gas.
For many engineering problems, the accurate pressure-temperature-volume
relationships required are so involved that it is difficult to use them to
determine changes in other properties. Urtder these conditions, engineers
must resort to tables of the properties of various substances.
Tables of properties of substances are based on extensive laboratory tests;
however, it is very difficult to obtain all the experimental information required
to prepare extensive tables of properties. Furthermore, the accuracy of
experimental data may be subject to question. The equations developed in

166
Introduction 167

this chapter will show that for various processes there are thermodynamic
relationships between the various properties. For example, during the
vaporization of a liquid there is a very precise relationship between the
volume change and the latent heat. Thus, these equations are necessary to
validate experimental data as well as to determine complex properties that
cannot be measured.
The equations developed in this chapter will mathematically verify equa¬
tions that have been used in this text for ideal gases. In addition the equations
developed here can be used to evaluate changes in properties of those actual
gases having simple p-v-T relationships, such as those of van der Waals. This
chapter will enable the student to understand how thermodynamics is in¬
volved in the preparation of tables of properties of various substances.
Since the methods to be discussed involve the use of partial differential
equations, a brief summary of the required principles is in order.
Assume that Z is a true function of X and Y. Then

dz - (§)/x+(jf)/y (8-»
where (dZldX)Y is the rate of change of Z with X when Y is held constant
and (dZldY)x has a similar meaning.
Equation 8-1 is said to be an exact differential equation, since it involves
quantities X, Y, and Z, which are functions of each other as was originally
assumed. If Z has continuous derivatives, the equality of the second cross
partials becomes a necessary condition for exactness. This principle can be
validated by taking the second cross partials in Eq. 8-1. Thus,
d(dZldX) _ d(dZldY)
dY dX
or

a2z _ d2z
(8-la)
dXdY dXdY
Since there is a mathematical interrelationship between the various properties
of substances and also between the changes in the properties, any differential
equation expressing the relationship must be an exact one.
Some energy quantities, such as heat and work, may be expressed in terms
of the property changes of the system. For instance, for a reversible process
in a closed system, the work transferred is expressed as

8W = pdV

It may appear that W = f(p, V) and that the following differential equation
may be written as follows:

dV
168 General Equations of Thermodynamics

However, the term (dWldp)v has no significance since there is no work done
in a constant-volume process. Neither does the term (dWldV)p have
significance since the work done at a given constant pressure is a constant for
a given volume change. Obviously then, a meaningful differential equation
involving work done by a system and its properties cannot be written. A
similar statement may be made relative to the heat added to the system.
When Z = f(X, Y), a relationship can be established between the partials,
dZldX, dXldY, and dY/dZ by writing an equation similar to Eq. 8-1 based on
X = f(Y, Z). Thus,

dX = (8-2)

Substituting the value of dZ from Eq. 8-1 gives

dX\
dX = dY + dX + dY
( dY )■

Hence,

dY

Since X and Y are independent variables, any values may be assigned to


dX and dY. When a value of zero is assigned to dY, either dX may be zero or

Therefore, for a nontrivial solution,

(dZ\ _ 1 /dZ\ _(dZ\


(8-3)
\dX)y (dX/dZ)y ^ UX/y \aXjy
This relationship proves that the partial derivatives can be inverted.
In a similar manner, if a value of zero is assigned to dX, either dY may be
zero or

(8-4)

When Eq. 8-4 is divided by (dX/dY)z and the relationship in Eq. 8-3 is used,
the result is

(8-5)

8-2 INTERNAL ENERGY AND ENTHALPY


In differential form, for a closed system at rest, the energy equation for a
reversible process (per unit mass) is
8q = du + p dv (8-6)
Internal Energy and Enthalpy 169

or
du = 8q - p d v (8-7)
Since 8q = T ds,

du = T ds - p dv- (8-8)

Also, since s = f(T, v).

(8-9)

Substitution in Eq. 8-8 gives

du T — p dv (8-10)

Equation 8-10 is a perfectly valid equation for the change in internal energy
for any substance in any phase. However, the partials involving entropy
cannot be measured directly; hence, it is necessary to evaluate these partials
in terms of more readily measurable quantities. In Chapter 4, it was pointed
out that cu is a property and may be expressed as = (duldT)v. This may be
rewritten as

cv = (8-11)

From Eq. 8-8, when v- is constant,

Therefore,

(8-12)

The term (dsldv)T in Eq. 8-10 may be replaced by making use of the
Helmholtz function. This is defined as A = U - TS or, per unit mass, a =
u - Ts. Then

da = du - T ds — s dT

Substituting the value of du from Eq. 8-8,


da = —p dv — s dT (8-13)

Since a = f(v, T), and since a, v-, and T are all true properties, Eq. 8-13
must be an exact differential equation. From Eq. 8-la, it can be shown that

(8-14)

This is one of the four expressions known as the Maxwell relations. The other
three will be derived later.
170 General Equations of Thermodynamics

When the value of (ds/dT)v from Eq. 8-12 and the value of (ds/dv-)T from
Eq. 8-14 are substituted in Eq. 8-10 the result is

du = dT + \T d v- (8-15)
m-

Equation 8-15 is a general equation and is applicable to any pure substance


existing in any of its phases. The evaluation of Eq. 8-15 requires a relationship
between cv and T and also an equation of state relating p, v-, and T. When
these two relationships are known exactly, then Eq. 8-15 must yield the
accurate value of du. This statement may be illustrated by assuming that a
substances obeys the law pv = RT. Then

dp) =R
dT) V-

Substitution in Eq. 8-15 gives

RT
du = dT + p^j dv- = cM dT
v-

Since RT/v = p.
In a similar manner, an equation may be derived to evaluate the change in
enthalpy of a pure substance. Since h = u + p v-,

dh = du + p dv + v dp

Replacing du by its value from Eq. 8-8 gives

dh = T ds + v- dp (8-16)
Also, since s = f(T, p),

ds
-(£). MS)* (8-17)

When this value of ds is substituted in Eq. 8-16, the result is

ds'
dh = T
.dT.
dT +
(S)*l + v- dp (8-18)

As discussed in Chapter 3, cp is a property and may be expressed as


cp = (dhldT)p. This may be rewritten as

dh
cp (8-19)
.ds ). m.

From Eq. 8-16, when p is constant,


Entropy Changes 171

Then,

ds'
cp = T (8-20)
JT.

In order to replace (ds/dT)p in Eq. 8-18, consider the Gibbs function G.


Since G = H - TS, or g = h - Ts,

dg = dh - T ds - s dT

Replacing dh by its value from Eq. 8-16 gives

dg = v dp - s dT (8-21)

Since g = /(p, T), Eq. 8-21 must be an exact differential equation. From Eq.
8-3,

(8-22)

This is the second of Maxwell’s relations.


When the value of (dsldp)T from Eq. 8-22 and the value of (dsldT)p from
Eq. 8-20 are substituted in Eq. 8-18, the result is

dh = CpdT + (8-23)
[ v--T
A knowledge of the relation between cp and T and the equation of state
relating p, V, and T make possible the use of Eq. 8-23 for evaluating the
change in enthalpy of any pure substance for any process. This statement
may be illustrated for a substance that obeys the law p v = RT. Thus,

/dv\ _ R
\dT)p~ p

Substitution in Eq. 8-23 gives

dh = cpdT + (v - dp = Cp dT

8-3 ENTROPY CHANGES


Equation 8-9 is a general expression for entropy change. However, it involves
partials of s. From Eq. 8-12, (dsldT)v = cJT. From Eq. 8-14, (dsldv)T =
(dpidT)v. When these values are substituted in Eq. 8-9, the result is

ds = cv d v- (8-24)

when p v- = RT,
172 General Equations of Thermodynamics

Then
d v-
ds (8-24a)
v-

A second expression for entropy change may be obtained by the use of Eq.
8-17. From Eq. 8-20, (ds/dT)p = cplT; from Eq. 8-22 (ds/dp)T = —(d WdT)p.
Substituting these values in Eq. 8-17 gives

ds = cpf- (ypj dp (8-25)

When pv = RT,

(w)
\STJP -£
P

and
dT dp
ds cp (8-25a)
~T~rT

8-4 SPECIFIC HEAT RELATIONS


A general relationship for the difference between the specific heats cp and cu
can be obtained as follows: By equating the expressions for entropy change
given by Eq. 8-24 and 8-25, an equation may be obtained which contains both
cp and c„. Thus,

dv- (8-26)

Then,
^ j = T (dpi dT) T(dWdT)p
d v- + dp (8-27)

Since T = f(p, v),

dT = (8-28)

By equating either the coefficients of d^ or those of dp in Eqs. 8-27 and


8-28, the value of cp — cv may be found. Thus,

- c T (8-29)

As with other properties, a knowledge of the true p-v-T relationship


permits the determination of the difference between the specific heats, or
(cp - c„). When p = f(v, T) and Eq. 8-5 is applied, the result is
Specific Heat Relations 173

Substituting this value into Eq. 8-29 gives

Equation 8-30 was termed by Zemansky1 as one of the important equations


of thermodynamics. The following statements are based on it:

1 Since (dp/d^)T is negative for all known substances and since (dWdT)2
must be positive, cp can never be less than cv.
2 When T approaches zero, cp approaches cu.
3 When (dvldT)2p is zero (as it is for water at 4°C), cp must equal cu.

It is recognized that pressure changes at a given temperature may change


the specific heat cp. The effect of a pressure change may be determined as
follows: Equation 8-25 is

dl - T “T - (If), *
Since s = /(T, p), this is an exact differential equation. Then

(8-31)

Hence,

or
(8-32)

It is common practice to report values of cp at zero pressure.2 Large


changes in pressure, particularly at low temperature, produce measurable
changes in the specific heat cp. The magnitude of the effect of a pressure
change may be obtained by integration of Eq. 8-32 between zero pressure p0
and the pressure p at which the specific heat is desired. Thus,

(cP cp)t T (8-33)

In a similar manner, it may be shown that a change in volume produces a


change in the specific heat c„, as given by the following equation:

(c,-cJt = T f(ffV) d» (8-34)


J'V'co ' V-

where the term refers to infinity.

'Mark W. Zemansky, Heat and Thermodynamics, McGraw-Hill, New York, 1957.


2In reality, this pressure is a finite pressure sufficiently low that the value of the specific heat is
independent of pressure.
174 General Equations of Thermodynamics

8-5 CLAPEYRON EQUATION3


Many thermodynamic properties are interdependent. For instance, there is a
definite relation between the volume change and the enthalpy change when a
liquid changes into a vapor at constant pressure. The relationship desired is
one giving a value of (dh!dv-)p. It may be obtained as follows:

(8-35)

From Eq. 8-16,

dh\ _ T
ds)p

As will be discussed in Chapter 9, when the pressure of a vaporizing liquid is


constant, its temperature is also constant. Thus, (dsl d v-)p = (dsl d v)T- Then,
from Eq. 8-14,

Substitution in Eq. 8-35 gives

f!),*T(lr)„ ,8-36)
When a substance exists as a two-phase mixture, as it does when it is
changing from the liquid phase to the vapor phase, (dpi dT)v = (dpi dT)tvap,
where the subscript “evap” refers to the evaporation or phase-change con¬
dition. For the total phase change from liquid to vapor with the pressure held
constant, the ratio of the change in enthalpy to the change in volume is a
constant, or

hlK. — (Ah.\
V-fg \dv-Jp

where the subscript fg refers to the change in property during vaporization.


Substitution in Eq. 8-36 gives

(8-37)

Equation 8-37, which is the Clapeyron equation, gives the exact relationship
between the change in volume and the latent heat (enthalpy change) when a
liquid changes to a vapor. When experimental data do not agree with results
calculated from this equation, then the experimental data must be in error.

3Sometimes known as the Clausius-Clapeyron equation.


Joule-Thomson Coefficients 175

8-6 JOULE-THOMSON COEFFICIENTS4


The Joule-Thomson coefficient of a gas is a measure of its deviation from
being an ideal gas. The Joule-Thomson coefficient is defined as the overall
change in temperature with a change in pressure when a gas is throttled, there
being no net change in enthalpy.
Whenever a restriction of any sort' is placed in a flow passage, a throttling
process takes place. A throttling process through an orifice is illustrated in
Figure 8-1. Because of the much smaller area of the orifice, the velocity of the
gas at the exit of the orifice must be much greater than the upstream velocity.
To attain this larger velocity there is a transformation of some of the enthalpy

4Because Thomson later became Lord Kelvin, these coefficients are sometimes known as
Joule-Kelvin coefficients.
5This includes orifices as well as pressure reducing and pressure regulating valves which produce
irregularities in the flow pattern.

1 2 3

FIGURE 8-1 Throttling process.


176 Genera! Equations of Thermodynamics

of the gas into kinetic energy. This results in a reduction in the temperature of
the gas.
Downstream from the restriction, because of a greater flow area, the gas
stream slows down, with kinetic energy being turned back into enthalpy. This
results in an increase in the temperature of the gas. Assuming that the area of
the restriction is relatively very small, the flow rate is low. This means that
both the initial and final velocities are low and the net change in kinetic
energy may be neglected.
When the gas temperature is close to that of the surroundings or when the
throttling device is well insulated, the heat exchanged between the gas and its
surroundings can be neglected. In the absence of heat transfer and a net
change in kinetic energy, the first law shows that there cannot be an overall
change in enthalpy. Normally these conditions are assumed to be present, and
hence, it may be assumed that there is no net change in enthalpy in a
throttling process. For an ideal gas, dh = cp dT. Thus for an ideal gas, overall,
a throttling process produces no net change in temperature.
Joule and Thomson recognized that even air, oxygen, nitrogen, and similar
gases are not ideal gases and that other gases, such as carbon dioxide, show
marked deviations from the ideal-gas laws. Hence, it should be expected that
these gases might have measurable net changes in temperature when throt¬
tled. The change in temperature when a gas is throttled can be determined
precisely by use of Eq. 8-23, which is repeated here for convenience.

dh = cpdT + v-T

or (8-23)

cp dT = dh - v —
[ T

For the overall throttling process, with no net change in kinetic energy and
with no heat transfer, there cannot be a net change in enthalpy. Consequently,
the overall throttling process can be treated as a constant-enthalpy process.

(8-38)

The quantity (dT/dp)h is called the Joule-Thomson coefficient and is


frequently denoted by the symbol /x. By use of Eq. 8-38, p can be calculated
when reliable information is available for finding the mean value of cp for the
given conditions and when an accurate p-v-T relationship is known.
Since accurate values of cp and accurate p-v-T relationships are not always
available, the Joule-Thomson coefficient has been determined experimentally
for many substances. When these experimental determinations are being
made, it is essential that the gas be moisture free, that there be no measurable
overall change in the kinetic energy, and that there be no heat exchange with
the surroundings. Joule and Thomson investigated many types of throttling
Joule-Thomson Coefficients 177

Porous plug

f7/7/77/7
p2, h pi -
(variable) (fixed)
-
2 1

TTT-n'n r/-j-2ZZZZ2 s jV s ? ? s ? 7 s ? ? s s s s a.

(a)

(b)

FIGURE 8-2 Throttling from fixed initial conditions.

devices and found that the porous-plug type was the most satisfactory.
Because the porous plug is used frequently today in determining the Joule-
Thomson coefficients, this type of experimental work is frequently referred to
as the porous-plug experiment.
The variations in temperature with the pressure after a gas is throttled from
fixed initial conditions are illustrated in Figure 8-2. For some gases, at high
temperature and pressure a small amount of throttling produces an increase in
temperature.
The point on the curve where further throttling produces an inversion in the
temperature change is known as the inversion temperature. The nature of the
throttling curve is independent on the initial conditions as is shown in Fig. 8-3.
When the initial state of the gas is within the inversion curve, throttling
produces a drop in temperature. At room temperatures and at low or
moderate pressures, the states of most gases fall within the inversion curve.
When hydrogen is at room temperature, however, its state falls outside the
inversion curve. Hence, hydrogen undergoes an increase in temperature upon
being throttled.
To see why throttling produces an increase in temperature in some in¬
stances and a drop in temperature in others, the expression for dh should be
examined. Since h = u + pv,

dh = du + d(p v)

But u - uk + Up, where uk is molecular kinetic energy and up is molecular


potential energy. Then, since overall dh = 0 in a throttling process,

0 = duk + dup + d(p v)

The energy uk is a function of temperature. Hence, the nature of the


temperature change can be predicted by examining the effect of a pressure
change on uk. The relationship to be considered is

duk = -dup - d(p v)


178 General Equations of Thermodynamics

FIGURE 8-3 Throttling curves.

The molecular potential energy increases as the pressure is decreased. In a


throttling process the percentage change in the absolute temperature is
normally so small that it can be neglected. Under these conditions, the
product p v is directly proportional to the compressibility factor. From Figure
3-9, it may be seen that d(pv) can be positive, negative, or zero. When d(pv)
is positive and numerically exceeds dup, then duk is negative. But the decrease
in uk is caused by a decrease in pressure. Thus, for this case (duk/dp)h is
positive. When (dujdp)h is positive, the Joule-Thomson coefficient (dT/dp)h is
also positive. Unless d(p v) is positive and numerically exceeds dup, the
Joule-Thomson coefficient is either negative or zero.
Measured Joule-Thomson coefficients are valuable in determining ther¬
modynamic properties of gases and vapors. Joule-Thomson coefficients were
used by Marks and Davis in determining the enthalpies of superheated steam
for their steam tables.6
Linde recognized the possibility of liquefying a gas when the Joule-
Thomson coefficient is large. To obtain a commercial yield of a liquefied gas,
Linde used regenerative cooling. This process is illustrated schematically in
Figure 8-4. High-pressure gas, cooled as much as possible by available means
(such as cooling water), enters the inner tube. It is throttled as it leaves the
tube through an orifice and, hence, experiences a drop in temperature. Initially,

6L. S. Marks and H. N. Davis, The Thermal Properties of Saturated and Superheated Steam,
Longmans, Green and Company, New York, 1922.
Other General Equations 179

i High-pressure
gas

Low-pressure
gas Throttling
orifice

Valve

FIGURE 8-4 Liquefaction of gases.

the temperature of the gas after throttling is only a little lower than that
entering the liquefier. However, the lower temperature gas is circulated back
around the tube containing the higher pressure gas, thus reducing its tem¬
perature. Hence, the temperature of the gas after throttling is also reduced.
Continued operation of this apparatus produces successively lower tem¬
peratures of the gas emerging from the orifice. Ultimately the temperature
drops to the condensation temperature for the low-pressure gas, and partial
condensation occurs. The remaining low-pressure gas may be compressed to
be recirculated.

8-7 OTHER GENERAL EQUATIONS


The equations presented in this chapter are only a few of the general
equations of thermodynamics which are so necessary in formulating the
thermodynamic properties of various substances. A discussion of all of these
necessary general equations is beyond the scope of this text.
Earlier in this chapter, mention was made of the existence of the four
Maxwell relationships. Two of them were derived. One of these, Eq. 8-14, is

(8-14)

The second one, Eq. 8-22, is

(8-22)

Another equation may be derived as follows: Equation 8-8 is

du = T ds — p dv (8-8)

Since u = /(s, ^), Eq. 8-8 is an exact differential equation. Hence,

(8-39)

Equation 8-39 is a third Maxwell relationship.


180 General Equations of Thermodynamics

The fourth Maxwell equation may be derived as follows:

dh = du + p d v- + v- dp = T ds + v- dp (8-40)

Since h = f(s, p), Eq. 8-40 is an exact differential equation. Hence,

(8-41)

The first two Maxwell relationships were used here to determine certain
properties from p-v~-T measurements. The other two relationships can be
used in a similar manner.

8-8 SUMMARY
In this chapter the general equations of thermodynamics have been presented,
showing the interrelationships between the various thermodynamic proper¬
ties. When p-1< -T relationships and, in some cases, the behavior of the
specific heats of any substance are known, the other thermodynamic proper¬
ties can be determined without the necessity of trying to measure them. This
fact is very important since it is extremely difficult to measure many of these
properties.
It has been shown also that when pv = RT, these general equations reduce
to those for ideal gases, some of which were developed earlier in this text. In
other cases these equations point out relationships that had not been
developed previously. For example, Eq. 8-33 shows that when pv = RT, a
change in pressure at constant temperature produces no effect on the value of
the specific heat cp.
When the experimentally determined properties of substances do not have
the same relationships to each other that the general equations of ther¬
modynamics show they should have, then there must be errors in the
experimental data. Thus tables of properties, to be valid, must be based on
experimental data, verified, and tied together using the general equations of
thermodynamics.
Problems 181

PROBLEMS__
8-1 Derive an expression for the change in internal energy of a substance
that obeys van der Waals equation of state.
8-2 Air is compressed from 13.8 bars, 93°C to 138 bars, 205°C. Determine
accurately the change in internal energy and enthalpy per kilogram of
air. Suggestion: Use the mean compressibility factor as a constant.
8-3 Determine the change in entropy in Problem 8-2.
8-4 The lowest available temperature in Problem 8-3 is 22°C. Determine the
change in available energy of the air.
8-5 The following boiling temperatures of water have been determined:

1.3 MPa 191.64°C 1.2 MPa 187.99°C


1.6 MPa 201.41°C
1.9 MPa 209.84°C 2.0 MPa 212.42°C

At 1.6 MPa, the volume change during vaporization is 122.64 cm3/g.


Estimate the enthalpy change during vaporization.
8-6 Derive an expression for the difference between cp and c„ for a gas that
obeys the law p v- = RT.
8-7 Derive an equation to show the effect of a change in pressure at a
constant temperature on the value of cp for a gas that obeys the law
p v- = RT.
8-8 Derive an equation to show the effect of a change in pressure at a constant
temperature on the value of for a gas that obeys the van der Waals law.
8-9 Table 4 in the Steam Tables, by J. H. Keenan, F. G. Keyes, P. G. Hill, and
J. G. Moore (Wiley, New York, 1978) shows that at 0°C, an increase in the
pressure on water up to 22 MPa produces an increase in entropy. At all
other temperatures an increase in pressure produces a decrease in
entropy. Justify these statements by using one of the Maxwell relation¬
ships.
8-10 At pressures above 22 MPa, an increase in pressure at a constant
temperature of 0°C produces a decrease in entropy of water. What
must happen to the specific volume under these conditions?
8-11 Derive a general relationship, in terms of p, v, and T, for the ratio of cp
to cv.
8-12 Show that the Joule-Thomson coefficient for a gas obeying the law
p v = RT must be zero.
8-13 In the derivation of Eq. 8-15 and 8-23 it was assumed that the process
was a reversible one taking place in a closed system. Does this mean
that these equations are valid only for reversible processes within a
closed system?
182 General Equations of Thermodynamics

8-14 Derive a general expression for the change in entropy of a gas obeying
the van der Waals equation. Hint: See Eq. 8-24.
8-15 For liquids and solids it is relatively easy to measure the volume
expansion (the fractional change in volume per unit increase in
temperature at constant pressure) and the compressibility k (the frac¬
tional decrease in volume per unit increase in pressure at constant
temperature). Derive an equation for the difference between cp and
in terms of £$ and k.
8-16 Derive an expression for the change in temperature per unit volume
change for the unrestricted expansion of an actual gas. (As used here,
the unrestricted expansion is a free expansion in a closed system.) The
expression is to be in terms of p, v, T, and cv.
8-17 Using the expression developed in Problem 8-16, derive an expression
for the temperature change for a free expansion of a gas obeying the
law pv = RT.
8-18 The same as Problem 8-17 except for a gas obeying the van der Waals
law.
8-19 Derive an expression for the change in enthalpy in terms of p, v, T, and
£$ (the fractional increase in volume per unit increase in temperature at
constant pressure).
8-20 The following boiling temperatures of refrigerant F-12 have been
determined:

4.2330 bars 10° C


4.9137 bars 15°C
5.6729 bars 20° C

The change in enthalpy during vaporization at 15°C is 143.683 J/g.


Compute the volume change during vaporization.
. 9 ...

VAPORS

9-1 INTRODUCTION
There is no sharp line of distinction between a vapor and a gas. In general, a
gas is considered to be far removed from its liquid state whereas the vapor is
thought to be rather readily liquefiable. However, at very low pressures a
so-called vapor may be treated as an ideal gas for most engineering purposes.
For example, the water vapor in atmospheric air is at such a low partial
pressure that it follows the ideal-gas laws very closely. On the other hand,
when a vapor exists at very high pressures it deviates significantly from an
ideal gas, even when its temperature is much higher than the boiling tem¬
perature for the given pressure. In general, when the compressibility factor of
a vapor is substantially equal to unity for the existing temperature and
pressure, the vapor may be treated as an ideal gas (see Section 3-9).
It was pointed out in Chapter 8 that when a p-v-T relationship is available
for a substance, other properties may be calculated. However, when the
compressibility factor differs greatly from unity, an accurate p-v-T relation¬
ship is a complicated one. This makes the determination of other properties
by use of the p-v-T relationship very difficult. Under these circumstances, it
is most desirable to have tables available that give the various properties of
the substance under consideration.
Very extensive and reliable tables are available for the properties of
common vapors, such as steam. Although tables are available for the less
common vapors, generally they are much less extensive and, in some cases,
are of doubtful accuracy. It should be recognized that the compilation of
accurate and extensive tables of properties is extremely difficult. In general,
certain properties, such as pressure, temperature, specific volume, and
specific heat must be determined over a wide range of conditions. The
determination of many of these properties is far from being easy. For
example, a precise determination of the specific volume of a boiling liquid
and of its vapor at the boiling temperature for the given pressure requires
much ingenuity and patience. Once the simple properties have been deter¬
mined experimentally, these properties may be correlated by use of the
equations developed in Chapter 8. Then additional properties can be cal¬
culated, generally by use of these equations and p-v-T relationships. Since
these equations are precise ones, the accuracy of the calculated properties
depends solely on the experimentally determined properties. However, it

183
184 Vapors

should be recognized that this method of determining complex properties is


quite difficult.
The compilers of various vapor tables normally give the source of their
experimental data and also the thermodynamic methods used in calculating
other properties. Since there is no universal approach to the calculation of
these properties, a discussion of the methods used will not be considered
here.

9-2 VAPORIZATION OF LIQUIDS


Consider a simple liquid existing in a vertical cylinder with a loaded piston
resting on the liquid. If the load on the piston remains constant, as indicated
in Figure 9-1, any process taking place will be a constant-pressure one.
Addition of heat to the liquid produces an increase in both the temperature
and the entropy, as shown in Figure 9-2. This increase in temperature and
entropy continues with the addition of heat, as indicated in process 1-2, until
the boiling temperature is reached at state 2. Further addition of heat
produces a phase change of the liquid, the phase change taking place at
constant temperature, as represented by process 2-3. The liquid is completely
vaporized at state 3. Further addition of heat will cause an increase in the
temperature, as shown for process 3-4.
When this overall heating process is repeated with different loads on the
piston and, hence, different pressure on the liquid, a series of curves will be
obtained. Three such curves are shown in Figure 9-3. The temperature at
which the phase change (evaporation) takes place is a function of the pressure
exerted on the liquid. The lines representing the heating of the liquid are
distorted in Figure 9-3 to distinguish them from one another. Unless distortion
is used, it is necessary to use a very large scale to separate these lines.
Figure 9-4 shows a typical plot of the effect of pressure on the evaporating
or boiling temperature. This temperature is known as the saturation tem¬
perature for the given pressure. The pressure associated with the saturation
temperature is the saturation pressure. Specification of the saturation tem-

|W \
F’istor

Liquid "

II M a

Q FIGURE 9-1 Heating a liquid at constant pressure.


Vaporization of Liquids 185

FIGURE 9-2 T-S diagram for heating a liquid and vapor at constant pressure.

perature of a pure liquid automatically specifies the saturation pressure and


vice versa.
A vapor existing at the saturation temperature and pressure is said to be a
saturated vapor. A liquid existing at the saturation temperature and pressure
is known as a saturated liquid.
The locus of the points representing a saturated liquid is known as the
saturated liquid line (sometimes shortened to liquid line). The saturated liquid
line is almost coincidental with the lines representing the heating of the liquid
at various pressures (see Fig. 9-3). The saturated vapor line is the locus of
points representing a saturated vapor. (A saturated vapor is frequently known
as a dry vapor.) The two lines together constitute what is known as the vapor
dome. The region beneath the vapor dome is a region of a mixture of a
saturated liquid and a saturated vapor, frequently known as a wet-saturated
vapor. For normal pressures, unless the mass of a wet-saturated vapor is
almost all liquid, most of the volume occupied by a wet vapor is that occupied
Temperature

FIGURE 9-3 Constant-pressure curves for heating a liquid and vapor.

186
Vaporization of Liquids 187

FIGURE 9-5 Saturated-liquid and saturated-vapor lines on a p-v- plane.

by the vapor. Hence, it is common practice to speak of a liquid-vapor mixture


as a vapor. For example, a mixture of water vapor and liquid water droplets is
referred to as steam. (The steam vapor itself is not seen but the water droplets
that are present are seen.) The saturated-liquid and saturated-vapor lines are
shown on a p-t^ plane in Figure 9-5. The saturated-liquid and saturated-vapor
lines shown in the T-S diagram of Figure 9-3 are typical for compounds

FIGURE 9-6 T-S diagram for a substance having nonpolar molecules.


188 Vapors

composed of polar compounds, such as water, ammonia, and carbon dioxide.


Figure 9-6 shows a T-S diagram that is typical for substances having
nonpolar1 molecules, such as the hydrocarbons.
A vapor existing at a temperature in excess of the saturation temperature at
the given pressure is called a superheated vapor. Although it is common
practice to specify the state of a superheated vapor by giving the temperature
and pressure, sometimes the state is specified by giving the pressure and
degree of superheat. The degree of superheat is the difference between the
actual temperature and the saturation temperature for the given pressure.

9-3 VAPOR TABLES


As we discussed earlier in this chapter, complete vapor tables can be
compiled only after long and patient experimental work and extensive cor¬
relation of experimental results by use of the general equations of ther¬
modynamics. Although there are many tables in existence that give the
thermodynamic properties of various substances, most of these tables are
limited in their range of coverage. Furthermore, because of the lack of
accurate experimental data, most of these tables cannot be considered ac¬
curate to a fraction of one percent.
Over the years, more thought, time, and energy have been devoted to the
determination of the properties of water and its vapor than for any other
substance. Many “steam tables” have been prepared both in the United States
and abroad. Periodically, International Steam Table Conferences have been
held, which have fixed the allowable tolerance in various properties at
specified conditions. This has made it possible for certain interchangeability in
the use of steam tables published in various countries. Since their publication
in 1936, the tables in Thermodynamic Properties of Steam, by Keenan and
Keyes (Wiley, New York), have been used almost exclusively in this country.
In 1967, the American Society of Mechanical Engineers published the ASME
Steam Tables. In 1969, Keenan, Keyes, Hill, and Moore published later
tables, entitled Steam Tables.2 Since these tables are used very extensively,
they will be used here and will be referred to as the Steam Tables. These
tables fall within the limits established by the Skeleton Table of the Sixth
International Steam Table Conference of 1963. Except at very high pressures,
there is good agreement between the three sets of tables listed above. Even at
high pressures the deviations generally are less than 1 percent, although
deviations may be higher for the values of the viscosities. Since steam is used
so extensively and because such detailed tables of its properties are available,
much of the discussion in this chapter will be centered on the properties of

'In a nonpolar compound the atoms are arranged symmetrically in the molecule, and there are no
unbalanced electrical charges that tend to rotate the molecule when it is placed in an electrostatic
field.
:J. H. Keenan, F. G. Keyes, P. G. Hill, and J. G. Moore, Steam Tables, Wiley, New York, 1969.
Second edition, 1978.
Compressed Liquids 189

steam. The general considerations, however, will be equally applicable to the


properties of other vapors.
Tables 1 and 2 in Keenan, Keyes, Hill, and Moore deal with saturated
steam and Table 3 gives the properties of superheated steam. Table 1 is
compiled for various temperatures ranging from 0°C to the critical tem¬
perature of 374.136°C. Table 2, a pressure table, ranges in pressures from the
triple-point pressure of 0.6113kPa to the critical pressure of 22.09 MPa.3
Table 3, the superheated table, covers the pressure range up to 100 MPa, and
the temperature range up to 1300°C. Abstracts of these tables are given in the
Appendix of this text.
The arrangement of the properties in Table 2 is as follows:

Temp¬ Specific
Pressure erature Volume Internal Energy Enthalpy Entropy
(MPa) (°C) (cm3/g) (J/g) (J/g) (J/g-K)

Sat Sat Sat Evap Sat Sat Evap Sat Sat Evap Sat
Liquid Vapor Liquid Vapor Liquid Vapor Liquid Vapor
p t v-f v-g Uj Ufg Ug hf hfg hg Sf Sfg sg

In the headings in the tables, the subscript / denotes a saturated liquid, g a


dry saturated vapor, and the subscripts fg the change in property during
vaporization.
Since only changes in enthalpies, internal energies, and entropies are
required in engineering problems, it is not necessary to determine absolute
values. Thus, a datum is selected and values are calculated with reference to
the datum. Keenan, Keyes, Hill, and Moore selected for their tables a datum
of the triple point (0.01°C, 0.6113 kPa) for the internal energy and the entropy
of the water. Other datums are used for other substances. In this country it
has been a common practice to use a datum of -40°F (-40°C) for refri¬
gerants.

9-4 COMPRESSED LIQUIDS4


Under most conditions, liquids do not exist at their boiling or saturated state.
Hence, their properties differ from those of a saturated liquid as given in
tables of liquid-vapor properties. Liquids are only slightly compressible,
particularly at low temperatures. Thus, a relatively moderate increase in the
pressure exerted on the liquid produces only a very small change in the
properties of the liquid, provided that its temperature is not changed. Unless a
high degree of accuracy is desired when the difference between the properties
of a liquid and its vapor is wanted, the properties of a compressed liquid (one
existing at a pressure in excess of the saturation pressure for the given

The unit of pressure used in these tables is the megapascals. In dealing with steam, some authors
express pressures in either kilopascals or bars. A bar equals 100 kilopascals and is approximately
equal to 1 atmosphere.
4Sometimes called subcooled liquids.
190 Vapors

temperature) may be taken as being equal to the properties of a saturated


liquid at the same temperature, particularly for low and moderate pressures.
The properties of compressed water are given in Table 4 of Keenan, Keyes,
Hill, and Moore. The two examples that follow illustrate the use of these
tables in determining the magnitude of the effect of pressure on the properties
of compressed water.

Example 9-1

(a) Determine the difference between (1) the specific volume and (2) the
enthalpy of saturated water at 40°C and water at a pressure of 5 MPa,
40° C.
(b) Solve the same problem stated in part (a), except the second pressure is
50 MPa.

Solution

(a) From Table A-3a in the Appendix at 40°C, = 1.008 cnr/g and hf =
167.57 J/g. From Table A-3d, v = 1.0056 cm3/g and h = 171.97 J/g. Then the
increase in pressure from the saturated value of 0.007354 MPa to 5 MPa
has decreased the specific volume 0.0024 cnr/g or 0.24 percent. The
enthalpy has increased 2.6 percent.
(b) From Table A-3d in the Appendix, at 40°C, 50 MPa, v- = 0.9872 cm3/g and
h = 211.21 J/g. Then the increase in pressure to 500 bars has decreased the
specific volume by 0.0208 cm3/g or 2.06 percent. The increase in pressure
has increased the enthalpy by 43.65 J/g, or 26.1 percent.

It should be noted that increasing the pressure from a saturated value of


0.007384 MPa to 5 MPa had practically no effect on the specific volume and
only a slight effect on the enthalpy. However, when the pressure is increased
to 50 MPa, there is a significant change in these properties. It should be noted
also that since h = u + p v, a large change in pressure can produce a significant
increase in the enthalpy.
Since hot water is more compressible than cold water, the effects of
pressure changes on these properties of water are larger than those found in
this example.

Example 9-2. Determine the difference between the specific volume and the
enthalpy of saturated water at 300°C and water at 300°C under a pressure of
50 MPa.

Solution. From Table A-3a in the Appendix, the specific volume of saturated
water at 300°C is 1.4040 cm3/g. From Table A-3d, the specific volume of water
at 300°C, 500 bars is 1.2860. This represents a decrease in specific volume of
0.1150 cm3/g or 8.40 percent.
From Table A-3a, the enthalpy of saturated water at 300°C is 1344.0 J/g.
From Table A-3d, the enthalpy of water 300°C, 50 MPa is 1323.0 J/g. This
represents a decrease of 21 J/g or 1.56 percent.
Properties of a Wet Vapor 191

9-5 PROPERTIES OF A WET VAPOR


In the discussion of the vaporization of liquids, it was assumed that there was
complete separation of the vapor and the liquid from which it was formed. In
most instances, separation is not complete and the vapor leaves the vapor-
producing device with some entrained liquid. When the vapor is conveyed
through a pipe line and heat losses occur, some of the vapor condenses and
produces more liquid. Vapors, other than the hydrocarbons and similar
substances, tend to partially condense when expanded in a work-producing
device. Thus, mixtures of liquid and vapor are frequently encountered in
engineering devices. Such mixtures are known as wet vapors.
In many cases, the vapor-liquid mixture is not a uniform one. When the
vapor is moving with a relatively high velocity, the liquid droplets are
entrained by the vapor and are carried along with it, but with a lower velocity
than that of the vapor. Unless the vapor is moving with a relatively high
velocity, the liquid droplets, partially at least, will separate. However, when
no provisions are made for removing the liquid from the flow passage, and
when steady-state conditions exist, the rate of flow of vapor and liquid past a
given point is the same whether the mixture is uniform or whether the two
phases are completely separated. For this reason, wet vapor is generally
assumed to be a uniform mixture in determining the total properties. Caution
must be used, however, in the actual analysis of the flow of a wet vapor, to
make certain that separation and possible diversion of the liquid do not occur.
Furthermore, in the flow of a vapor at high velocities, consideration must be
given to the fact that the liquid and vapor velocities are not equal.
Before the properties of a wet vapor can be specified, it is necessary to
know the relative amounts of liquid and vapor present. The relative amounts
are expressed as either the quality or the moisture content. The quality x of
wet vapor is the mass of the vapor divided by the total mass of liquid and
vapor. The moisture content y equals the mass of the liquid divided by the
total mass. The sum of the quality and the moisture content obviously must
be unity, or x + y = 1. Then x = 1 — y.
The entropy of a wet vapor, sx exceeds the entropy of the liquid from
which it was formed by that portion of the entropy of vaporization which
must be added to produce a wet vapor with quality sx. Thus,
sx = sf + xsfg (9-1)
See Figure 9-7.
In many problems dealing with wet steam, such as in steam turbines, the
amount of moisture present is of prime importance. Hence, it is desirable to
find the properties of steam in terms of its moisture content. The entropy of a
wet vapor is less than that of a dry vapor by that portion of the entropy of
vaporization change produced when dry vapor is sufficiently condensed to
produce a wet vapor with moisture content, y. In equation form,
sx = sg- ysfg (9-2)
See Figure 9-7.
192 Vapors

FIGURE 9-7 Entropy of a wet vapor.

In a similar manner

hx hg yfyg (9-3)

or
hx = hf 4- xhfg (9-4)

and
vx = v-g - yvfg (9-5)

and

vx - v-f + X Vfg (9-6)

Except at high temperatures, the specific volume of the saturated liquid is


insignificant in comparison to that of a saturated vapor. Under these con¬
ditions, vg is substantially equal to vfg (which is not listed in the tables).
Equation 9-5 becomes

= v-f, - yv< (9-5 a)5

Equation 9-6 becomes


= X V-, (9-6a)

5It is preferable to leave this equation in this form when the moisture content of the steam is
known.
Determination of Vapor Properties 193

9-6 DETERMINATION OF VAPOR PROPERTIES


Thus far it has been assumed that the phase of the substance is known, that is,
either it is a liquid, a saturated vapor (liquid-vapor mixture), or a superheated
vapor. In actual cases, the phase of the substance must be determined first
before attempting to determine its properties.
The two properties easiest to measure are pressure and temperature. When
the observed temperature equals the saturation temperature for the observed
pressure, the substance is in the saturated region (i.e., it lies somewhere inside
the vapor dome). If the observed temperature exceeds the saturation tem¬
perature for the observed pressure, the substance is superheated. Other
properties can then be determined directly from the superheated tables. When
the observed temperature is less than the saturation temperature for the
observed pressure, the substance is a compressed liquid. Compressed-liquid
tables can be used to determine other properties.
It must be emphasized that, for the saturation region (under the vapor
dome), temperature and pressure are not independent properties. The ob¬
served pressure and temperature do specify that the substance is in the satura¬
tion region. But this information does not tell whether the quality is zero or
100 percent or anything in between. One additional independent property
must be determined before other properties can be specified.
In certain problems, either temperature or pressure and one other in¬
dependent property are available. When the second property exceeds that of a
dry saturated vapor for the given temperature or pressure, the substance must
be in the superheated region. If the second property is less than that of a
saturated liquid, the substance is a compressed liquid.6 In either case, other
properties can be determined by use of the superheated- or compressed-liquid
tables, as the case may be.
The methods of determining the properties of a vapor in a given state will
be illustrated by the use of examples involving steam. The methods developed
will be applicable to other substances.
Example 9-3. Specify the state of steam having a pressure of 0.6 MPa and an
enthalpy of 2570 J/g.
Solution. In Table A-3b of the Steam Tables (see Appendix), the value of hg
is given as 2756.8 J/g. Since the given value of enthalpy is less than that of dry
saturated steam, the steam must be wet. Its moisture content can be deter¬
mined by use of Eq. 9-3. Thus,
hx = hg- yhfg
2570 = 2756.8 - y(2086.3)
y = 0.0895

or the state is steam at a pressure of 0.6 MPa and a moisture content of 8.95
percent. Answer

6It is assumed here that the substance is far removed from its solid phase.
194 Vapors

Example 9-4. Steam at a pressure of 1 MPa has a specific volume of


232.7 cm3 per gram. Specify the state of the steam.
Solution. From Table A-3b in the Appendix, v-g at a pressure of 1 MPa is
194.44 cm3/g. Hence, the steam in this problem is superheated. From the
superheated tables, the specific volume of steam at 1 MPa, 250°C is
232.7 cm3/g. Hence, the state of the steam is 1 MPa, 250°C. Answer

Example 9-5. Steam at a pressure of 0.4758 MPa has an internal energy of


2559.5 J/g. Specify the state of the steam.

Solution. In Table A-3a in the Appendix, at a pressure of 0.4758 MPa, the


saturation temperature is given as 150°C and the value of ug is 2559.5 J/g.
Hence, the steam must be dry saturated steam at 0.4758 MPa. Answer

Example 9-6. Steam at a temperature of 140°C has an entropy of 6.355 J/g-


K. Specify the state of the steam.
Solution. At a temperature of 140° C the value of sg from Table A-3a in the
Appendix is 6.9299 J/g-K and the saturation pressure is 0.3613 MPa. Hence,
the steam is wet steam. Using Eq. 9-2,

6.355 = 6.9299 - y(5.1908)


y = 0.1108

Hence, the steam is at a pressure of 0.3613 MPa with a moisture content of


11.08 percent. Answer

9-7 VAPOR CHARTS


The method of determining the state of a vapor has been illustrated in Section
9-6. Various types of vapor charts are available that may be used for a quick,
approximate determination of state and changes in state of vapors. One of
these is the T-s chart. A small T-s chart is presented in Keenan, Keyes, Hill,
and Moore as Figure 9-8. A skeleton of a diagram of this type is illustrated
here in Figure 9-8. This type of diagram is valuable in determining the state of
the substance. Although specific volumes and enthalpies are plotted on this
type of chart, it is difficult to read values accurately.
A second type of chart is the enthalpy-entropy, or h-s chart, which is
generally known as a Mollier diagram.7 A skeleton h-s chart is shown in
Figure 9-9. In normal engineering practice, the properties of very wet vapors
are seldom required. Hence, that part of the h-s diagram dealing with very
wet vapors may be omitted to permit the construction of a diagram that can
be read with a fair degree of accuracy and still be of reasonable size. Such a
chart can be found in the back of Steam Tables, by Keenan, Keyes, Hill, and
Moore. This diagram is presented in skeleton form in Figure 9-10. On this

The enthalpy-entropy chart is the most common of several diagrams designed by Mollier.
Vapor Charts 195

FIGURE 9-8 Skeleton T-s diagram for steam.

chart, pressures, temperatures, and specific entropies and enthalpies can be


read with sufficient accuracy for many engineering problems. Since constant-
volume lines and constant-pressure lines are nearly parallel, particularly in the
superheated region, constant-volume lines are normally omitted from an h-s
chart.

FIGURE 9-9 Skeleton h-s diagram for steam.


196 Vapors

Entropy (J/g-K)

FIGURE 9-10 Enlarged h-s diagram for steam.

9-8 THROTTLING OF VAPORS AND LIQUIDS


Flowing fluids suffer an abrupt loss of pressure when passing through restric¬
ted openings, such as valves, orifice plates, and similar constrictions in the
flow passages. A more gradual drop in pressure is experienced by a fluid
flowing in a constant-area passage that is straight or has a gradual change in
direction. The flow through a restricted opening is, by definition, a throttling
process. However, since pipe-wall friction produces a pressure loss, the flow
in a pipe may be treated as throttling action.
The energy equation applicable to a flowing fluid is independent of the
nature of the fluid. It may be assumed that no external work is done in the
flow passage under consideration and that no chemical reaction takes place.
Under these conditions the energy equation becomes

Hin + K.E.jn — JTout + K.E.out + Q (9-7)


Changes in State of Vapors 197

The subscripts “in” and “out” refer to the fluid conditions at entrance and
exit, respectively, of the section of the flow passage under consideration, and
Q is the heat lost by the fluid in its passage through the section.
In most engineering applications, fluid velocities in pipe lines are relatively
low. Although the throttling action produces an immediate increase in velo¬
city, normally because the flow area after the throttling device increases
greatly, the fluid velocity reduces to close to its original value. Overall, then,
there will be no appreciable change in kinetic energy.
Whether the heat transferred is appreciable is dependent on the tem¬
perature difference between the fluid in the flow section and its surroundings,
the amount of insulation covering the flow section, the length of the flow
section, and the mass rate of flow.
When the throttling action occurs abruptly, as in a valve or orifice, heat
losses are generally negligible. When there are large restrictions in the flow
passages, the rates of flow are low. Hence, the kinetic energies in Eq. 9-7 may
then be neglected, and Eq. 9-7 becomes Hm = Hout. This conclusion agrees
with that developed in connection with the discussion of the Joule-Thomson
effect. The Joule-Thomson effect for a vapor is large. Therefore, when a
vapor is throttled, an appreciable temperature change takes place, and it
becomes necessary to use vapor tables to predict the state of a vapor after it
is throttled.

9-9 CHANGES IN STATE OF VAPORS


In Chapter 4, methods were discussed for determining the final conditions of
an ideal gas undergoing a change in state, as well as for evaluating the
energies involved during the change. When gases deviate only slightly from
ideal gases, these methods may be adapted to apply to changes in state of the
gases. However, unless the pressure is very low, vapors deviate so much from
ideal gases that it is exceedingly difficult to apply the methods established for
ideal gases, except possibly for constant-pressure or constant-volume pro¬
cesses. Hence, tables of properties of vapors must be utilized to determine
final states as well as the energies involved in the changes in state.
Generally, the initial state of a vapor is specified along with either one
property at the final state or the energy involved during the change in state. If
the nature of the process is specified, it now becomes possible to obtain two
independent properties at the final state, thus fixing the final state.

Example 9-7. Steam exists in a vertical cylinder at a temperature of 200° C.


The piston in the cylinder maintains the pressure constant at 1 MPa as the
steam is heated by the addition of 223.3 J/g. Specify the final state of the
steam.

Solution. For a constant-pressure process in a closed system,

dh = Sq and h2=h\ + q
198 Vapors

From Table A-3c in the Appendix, hi = 2827.9 J/g. Then h2 = 2827.9 + 223.3 =
3051.2 J/g. From Table A-3c, p = 1 MPa and h = 3051.2 J/g. Then t = 300°C.
The final state of the steam is a temperature of 300°C and a pressure of
1 MPa. Answer

Example 9-8. Steam at 3 MPa, 400°C expands isentropically to a pressure of


0.6 MPa. Specify the final condition.

Solution. From Table A-3c in the Appendix, the initial entropy is 6.9212 J/g-
K. This is also the final entropy. Since this entropy exceeds the entropy of dry
saturated steam at 0.6-MPa pressure, the steam is still superheated. From
Table A-3c, by interpolation, the final temperature is 190.4°C. Answer

In engineering problems, it is desirable to evaluate the energy changes of a


system or the energies exchanged between a system and its surroundings. The
first law may be used to solve these problems.

Example 9-9. Determine the work done per pound of steam in Example 9-8.

Solution. It was not specified whether the system is closed or open. Hence,
solutions will be given for each type.
Closed system. Q = U2~ U\ + W\.2. Since Q = 0, W\.2 = Uj — U2. Using Table
A-3c in the Appendix,

W = 2932.8 - 2622.6 = 310.2 J/g Answer

Open system. It will be assumed that the steam does not undergo a
significant change in kinetic energy as it passes through the system. Then,

W = Hi-H2= 3230.9 - 2828.9 = 402 J/g Answer

Example 9-10. Water enters a steam generator at 25 MPa and 240°C. Steam
leaves at 22 MPa and 550°C. Determine the heat added per gram.

Solution. It may be assumed that there are no significant changes in kinetic


and potential energies resulting from passing through the steam generator.
Then, since it is an open system,

Q = H2 - H\ = 3370.3 - 1041.3 = 2329.0 J/g Answer

Note: The value of Hi was obtained from Table A-3d in the Appendix.

Example 9-11. Steam enters a turbine at 3 MPa and 470° C with a flow rate of
202,000 kg/h. 115,000 kg/h are extracted at 0.24 MPa and 170°C. The turbine
exhausts at 3.8 kPa with a velocity of 140 m/s, and delivers 24,000 kW.
Determine the enthalpy of the exhaust steam.

Note: Since it is extremely difficult to measure the moisture content of very


low-pressure steam, its enthalpy cannot be determined directly, but
can be calculated by use of the first law.
Changes in State of Vapors 199

FIGURE 9-11 Energy balance on simple


steam turbine.

Solution. Refer to Figure 9-11. The turbine should be so well insulated that
the heat lost per kilogram of steam is negligible. The energy balance then is

Hm = H extracted + (H + K.E.) exhausted +W


steam steam

/ 1402 \
202,000 x 1000 x 3055.4 = 115,000 x 1000 x 2807.2 + 87,000 x 1000(h2 + 21~kV00>)

+ 24,000 x 3.6 x 106


ft2 = 2079.9 - 9.8 = 2070.1 1/g Answer

Example 9-12. Steam at 6 MPa and 480°C is throttled to 4 MPa for part-load
operation of a small turbine. Determine:

(a) Final state of the steam after throttling.


(b) Loss of available energy per kilogram of steam if the condenser pressure
is 6.5 kPa.

Solution. There will be no significant overall change in the kinetic energy of


the steam as a result of the throttling and no measurable heat is lost. Hence,
the final enthalpy equals the initial enthalpy = 3374.3 J/g.

(a) The final state = 4 MPa and 469.1°C by interpolation from the superheated
tables. Answer
(b) The change in entropy during throttling equals
6.9964 - 6.8176 = 0.1788 J/g-K

The increase in unavailable energy during throttling equals

Tl(s2 — Si) = (273.15 + 37.63)0.1788 = 55.57 J/g Answer

where 37.63 is the saturation temperature in degrees Celsius for a pressure


of 6.5 kPa.
200 Vapors

FIGURE 9-12 Throttling of saturated liquid.

Example 9-13. Liquid refrigerant F-12 enters an expansion valve at a tem¬


perature of 20°C. It is throttled in the expansion valve to a pressure of 2.191
bars. Specify the state of the refrigerant after throttling.

Solution. It may be assumed that there is no net change in kinetic energy in


the expansion valve and no heat transfer. Hence, the enthalpy after throttling
equals that before throttling. From Table A-5 in the Appendix, the value of hf
at 20°C is 54.828 J/g (see Fig. 9-12). Using Eq. 9-4,

54.828 = 26.851 + *(156.207)


x =0.1791

Then the final state is wet vapor at a pressure of 2.191 bars and a quality of
17.91 percent. Answer
i

The method of determining the state after a liquid is throttled was illus¬
trated in Example 9-7. If, instead of being throttled, the refrigerant were
permitted to expand isentropically, it could deliver work. This work equals
the loss of available energy because of throttling. The method of determining
this loss is illustrated in the following example.

Example 9-14. Determine the loss in available energy per gram of refrigerant
as a result of the throttling process in Example 9-7.

Solution. The final entropy,

s2= sf + ysfg = 0.1079 + 0.1791(0.5936)


= 0.2142 J/g-K

The change in entropy during throttling,

s2-si = 0.2142 - 0.2076 = 0.0066 J/g-K

Assume that the lowest available temperature is the saturation temperature at


Measurements to Determine the State of Vapors 201

the final pressure of 2.191 bars. This temperature is -10°C. Then

loss of available energy = gain in unavailable energy


= Tl(s2~ Sj) = (-10 + 273.15)(0.0066)
= 1.737 J/g Answer

This loss in available energy of 1.737 J/g means that if the liquid had been
expanded in a reversible adiabatic manner, 1.737 J/g of work would have been
obtained.

9-10 MEASUREMENTS TO DETERMINE THE


STATE OF VAPORS
Although not always easy to measure, the pressure and temperature of vapors
generally can be determined with a reasonable degree of accuracy. When the
vapor is superheated, a knowledge of the pressure and temperature permits
the determination of the other properties. However, when the vapor is a
saturated one, the determination of the temperature in addition to the pres¬
sure simply confirms the fact that the vapor is saturated and in no way aids in
the determination of other properties, since the vapor may have a quality
ranging from zero to 100 percent. Hence, additional information must be
obtained. This additional information generally is either the moisture content
or the quality of the vapor.
It is not feasible to analyze large quantities of a vapor and, hence, a
representative sample of the vapor must be obtained. It should be emphasized
that a wet vapor is not an uniform mixture. Droplets are held in suspension by
virtue of the vapor velocity. At low vapor velocities many of the droplets fall
out of suspension. As in many other engineering problems, it is difficult to
obtain a truly representative sample. The ASME Power Test Codes, Supple¬
ments on Instruments and Apparatus: Part II, Determination of Quality
of Steam, 1940, give details for obtaining representative samples of wet
steam.
There are several types of calorimeters available for determining the
moisture content of vapors. Some of these are the separating calorimeter, the
electrical calorimeter, and the barrel (vapor-condensing) calorimeter. The
most widely used calorimeter is the throttling calorimeter, illustrated in Figure
9-13. In many engineering problems the moisture content of steam is rela¬
tively low. When such steam is throttled sufficiently, it becomes superheated.
Under these conditions, its temperature and pressure may be measured and
its enthalpy determined. The throttling calorimeter must be designed so that
there cannot be significant heat losses. Furthermore, it must also be designed
so that the steam velocity is low enough at the point of temperature deter¬
mination that the kinetic energy may be neglected. Under these conditions the
enthalpy determined after throttling will equal the enthalpy of the steam
sample.
202 Vapors

Example 9-15. The moisture content of steam is desired in a line where the
pressure is 1.61 MPa. The barometric pressure is 740 mm of mercury. The
data from the throttling calorimeter shows that the steam pressure is 8.6 mm
of mercury and the temperature is 120°C. Determine the condition in the line.

Solution

740
atmospheric pressure = ^ x 0.101325 = 0.0987 MPa
760
line pressure = 1.61 +0.0987 = 1.7087 MPa

calorimeter pressure = X 0.101325 + 0.0987 = 0.0998 MPa

From the superheated tables, the enthalpy in the calorimeter = 2716.6 J/g.
Then, at the line pressure, since hx = hg — yhfg,

2716.6 = 2795.8 — y (1918.0)


y = 4.12 percent Answer

When the steam pressure is very low, such as it is when it enters a power
plant condenser, it is very difficult to use most calorimeters to determine the
moisture content. Generally, sufficient information can be obtained to ap¬
proximate the moisture content by application of the first law. This concept
was used in Example 9-12 to determine the enthalpy of the steam at turbine
exit.
Nonequilibrium States of Two-Phase Mixtures 203

Example 9-16. Determine the moisture content of the steam leaving the
turbine in Ex. 9-11.

Solution. Since hx = hg — yhfg and hx = 2070.1 J/g, from Example 9-12,

2070.1 = 2552.8- y(2.435.0)


y = 19.82 percent Answer
Many vapors are formed from liquids that are relatively expensive. Hence,
it is not very practical to draw off a sample to determine the moisture content
unless means can be found to condense the vapor and save its condensate. As
an alternative, the moisture content may be estimated by the application of
the first law as was done in Example 9-16.

9-11 NON EQUILIBRIUM STATES OF TWO-PHASE


MIXTURES
Thus far in this chapter it has been assumed that there is equilibrium between
the two phases of a liquid-vapor mixture. Equilibrium does exist for all
practical purposes when a given state has been reached in a relatively slow
manner. However, when there is a very rapid change of state, nonequilibrium
conditions may exist for a very short period of time. An example of a rapid
change is the expansion of steam occurring in a nozzle when the steam
initially is only slightly superheated or has a very low moisture content. Some
treatment of nonequilibrium states will be undertaken in this section to call
attention to the fact that substances may not exist in equilibrium states in
many engineering problems.
The expansion of steam in an ideal nozzle is isentropic (see Fig. 9-14). If the
steam in its initial state is only slightly superheated and it is in equilibrium
throughout its expansion, condensation should occur shortly after the start of
the expansion process. The condensation process should continue as the
expansion process proceeds. However, the condensation process takes some
time. First, two or more molecules come together and act as a nucleus. This
nucleus attracts other molecules until sufficient molecules are present to form

Constant-
pressure
line

s FIGURE 9-14 Expansion in an ideal nozzle.


204 Vapors

a recognizable droplet of water. Although this process takes place in a very


small fraction of a second, the time required for the steam to pass through a
steam-turbine nozzle is also a very small fraction of a second. Hence, there is
not sufficient time for condensation to occur at the expected point in the
nozzle. In certain cases, steam may pass through a nozzle without any
condensation occurring, even though some of it should have condensed
according to theory.
As long as the steam does not condense, it may be treated as a superheated
vapor. In this case, saturated-liquid and superheated-vapor lines on a graph
have no significance, and there is no sharp break in a constant-pressure line
when it is extended into what is normally the saturated region. It may be seen
from Figure 9-15 that the expansion to a given pressure without condensation
results in a considerably lower final temperature than would otherwise be
obtained. This result should be expected, since the kinetic energy acquired by
the expanding steam when no condensation occurs must come wholly from
the molecular kinetic energy.
Steam existing in a nonequilibrium condition is frequently called super¬
saturated steam, or it is said to be in a metastable condition. Because the
temperature of supersaturated steam is lower than the steam-table value for
the given pressure, the term uncooled is sometimes used.
When condensation of supersaturated steam does occur and very small
droplets are formed, it is possible for the droplets to be in temporary
equilibrium with the vapor at a temperature materially lower than the steam-
table value of the given pressure. Since this statement seems contrary to the
concepts discussed earlier in this chapter, it is desirable to examine the
equilibrium between small droplets and vapor.
Consider a container having a liquid in the bottom and a vapor above the
liquid. A molecule of the liquid approaching the surface has a tendency to
pass into the vapor phase by virtue of its momentum. All the molecules of the
liquid within the sphere of action of the molecule under consideration offer a
restraining influence, as shown in Figure 9-16. Only those molecules of the
liquid that have high momentum are able to leave the surface (vaporization).

s FIGURE 9-15 Supersaturated expansion.


Nonequilibrium States of Two-Phase Mixtures 205

FIGURE 9-16 Forces acting on a molecule at a flat


liquid surface.

The slower moving molecules of the vapor will be attracted into the liquid
phase (condensation) when they approach the surface of the liquid. Equili¬
brium exists between the liquid and vapor phases when the rate of vaporiza¬
tion equals the rate of condensation. For any given fluid the equilibrium
temperature is a function of the vapor pressure. This equilibrium temperature
(i.e., for a liquid with a flat surface) is listed in tables of properties of vapors
for various pressures as the saturation temperature.
Now consider a molecule of a liquid approaching the surface of a drop.
Comparing Figures 9-16 and 9-17, it may be seen that for a given molecule
approaching the surface of the drop, Figure 9-17, there are fewer molecules in
the sphere of action restraining the given molecule from leaving the surface
and passing into the vapor phase. For any given temperature and pressure,
more molecules will be leaving the liquid phase where the liquid exists in the
form of drops than where it has a flat surface. For any given pressure, the
temperature of the drops (and also the vapor) must be lowered to decrease the
rate of vaporization and thus to achieve equilibrium. Therefore, the equili¬
brium temperature between drops of a liquid and its vapor, for any given
pressure, is lower than the tabulated saturation temperature.
It is possible to establish a relationship between equilibrium and saturation
temperatures in terms of drop radii. Such a relationship shows that unless the
drop radius is extremely small, the difference between the saturation tem¬
perature and the equilibrium temperature of the drops and their vapor is
insignificant. However, it has been found that in cases of supersaturation flow
in steam-turbine nozzles, condensation at first results in the formation of

FIGURE 9-17 Forces acting on a molecule at the surface


of a drop.
206 Vapors

droplets that have very small radii. Although these droplets can be in
equilibrium with the vapor as just described, such equilibrium is only tem¬
porary. When a very small droplet loses or gains a very few molecules, its
radius changes and it is no longer in equilibrium with its vapor.
One engineering example of nonequilibrium conditions has been discussed.
Many more nonequilibrium conditions may be encountered. Although con¬
ditions in many engineering problems are so close to equilibrium that equili¬
brium may be assumed to exist, very serious errors may be introduced by
such an assumption in other cases.
Problems 207

PROBLEMS __
9-1 Specify whether steam is wet, dry, or superheated for the following
conditions:
(a) t = 200°C, p = 1.44 MPa
(b) t = 220°C, p = 2.318 MPa
(c) p = 1.0 MPa, 5 - 6.672 J/g-K
(d) p = 3.0 MPa, t = 234°C (not 234.00°C)
(e) t = 250°C, v- = 54.2 cm3/g
(f) p = 11.0 MPa, h = 2805 J/g
(g) p = 4.0 MPa, 5 = 5.897 J/g-K
(h) p = 15.0 MPa, t = 310°C
9-2 Specify when refrigerant F-12 is wet, dry, or superheated for the
following conditions:
(a) t = 10°C, p =4.12 bars
(b) t = 80° C, v = 6.974 x 10”3 m3/kg
(c) p = 5 bars, s = 0.982 J/g-K
(d) p = 6.516 bars, t = 21°C
(e) p = 1.004 bars, h = 224.14 Jig
(f) p = 1.000 bars, t = -30°C (not 30.00°C)
9-3 (a) Does an increase in pressure at constant temperature increase the
enthalpy of water?
(b) Does an increase in pressure at constant temperature decrease the
specific volume of water?
9-4 Determine the percentage of change in the properties of saturated
water (^, h, and s) at 40°C when the pressure is increased to 5 MPa.
9-5 Determine the percentage of change in the properties of saturated
water (v, h, and s) at 350°C when the pressure is increased to 50 MPa.
9-6 Draw conclusions from Problems 9-4 and 9-5 of the effect of pressure
on the properties of water.
9-7 (a) Steam at 0.125 MPa has an enthalpy of 2495 J/g. Determine the
moisture content.
(b) Steam at 250°C has an internal energy of 2527 J/g. Determine the
moisture content.
(c) Saturated water at 240°C is throttled to 0.125 MPa. Neglect heat
losses and changes in kinetic energy and determine the final quality.
(d) After expansion to 0.15 MPa steam has an entropy of 6.932 J/g-K.
Determine the moisture content.
9-8 (a) Determine the amount of pressure required to decrease the volume
of saturated water one percent at 40°C.
(b) Same as part (a) except at 300° C.
9-9 Water enters a pump at 0.12 MPa 80°C and leaves at 20 MPa, 90°C.
Determine the percentage error in determining the pump work by
neglecting the effect of pressure on enthalpy.
208 Vapors

9-10 Same as Problem 9-9 except the final conditions are 5 MPa, 82°C.
9-11 (a) Refrigerant F-12 leaves a condenser at 7.8 bars, 30°C. Within
engineering accuracy, specify the specific volume, enthalpy, and
entropy.
(b) Mercury leaves a condenser at 0.025 bar, 200°C. Within engineering
accuracy specify the specific volume, enthalpy and entropy.
9-12 Steam enters a superheater at 20 MPa with a moisture content of 0.2
percent. It leaves at 19.6 MPa, 550°C. The flow rate is 638,000 kg/h.
Determine the heat added. State any assumptions that are made.
9-13 Steam is heated at a constant pressure of 2.0 MPa to a temperature of
500°C. The original moisture content is 0.5 percent. Determine the heat
added per kilogram. Is it necessary to know whether this is an open or
closed system?
9-14 A tank having a volume of 0.685 nr is 48.2 percent full of water at
atmospheric pressure and 40°C. Neglect the water vapor in the air
above the water. How much heat must be added to increase the
pressure to 8 MPa?
9-15 Steam expands isentropically from 20 MPa, 550°C to 0.01 MPa. Cal¬
culate the work per kilogram of steam for (a) closed system and (b)
open system.
9-16 Same as Problem 9-15(b) except the final pressure is 0.005 MPa. Cal¬
culate the percentage of increase in the work.
9-17 A turbine expands steam from 20 MPa, 550°C to 0.005 MPa. The actual
work is 78 percent of the isentropic work. Specify the enthalpy at the
end of the expansion and the moisture content. What assumptions must
be made?
9-18 Water enters a steam generator at 25 MPa, 160°C. Steam leaves at
22 MPa, 550°C. Determine the percentage of error caused in finding the
heat added by neglecting the effect of pressure on the enthalpy of the
water.
9-19 (a) Dry saturated steam at 20 MPa is throttled to 3 MPa. Specify the
steam conditions after throttling.
(b) Dry saturated steam at 2.5 MPa is throttled to 0.10 MPa. Specify the
steam conditions after throttling.
9-20 Steam enters a condenser at 2.5 kPa with a moisture content of 10.2
percent. Condensate leaves at 20°C. Circulating water (to condense the
steam) enters at 8°C and leaves at 16° C. The rate of steam flow is
892,000 kg/h. Determine the rate of circulating water flow.
9-21 Refrigerant F-12 (liquid) at 35°C is throttled to 2.1912 bars and then
enters an evaporator. It leaves the evaporator at 2.1912 bars, 0°C.
Determine the rate of refrigerant flow if the evaporator is to produce
850 kJ of refrigeration per minute.
Problems 209

9-22 Refrigerant F-12 enters a condenser at 7.449 bars, 60°C, at the rate of
48.5 kg/min. Condensate leaves saturated. Calculate the amount of heat
that must be removed from the condenser per minute.
9-23 A mercury vapor turbine receives dry saturated mercury vapor at
550°C and expands it to 0.0232 bar. The turbine has an isentropic
efficiency of 78 percent. Determine the rate of mercury vapor flow to
produce 50,000 kW.
9-24 An ammonia turbine is to produce power by absorbing heat from warm
ocean surface water. Saturated ammonia vapor enters the turbine at
25°C. The vapor expands to a pressure of 728.11 kPa. The turbine has
an isentropic efficiency of 80 percent. Determine the rate of ammonia
flow for a turbine output of 100,000 kW.
9-25 Assume that the ammonia in Problem 9-24 leaves the condenser as a
saturated liquid. It is condensed by cool water brought up from the
ocean depths. This water increase in termperature from 5°C to 10°C in
the condenser. Determine the volume rate of flow of this water,
assuming that its density is 1.025 g/cnr and its specific heat is 2.279 J/g-
K.
9-26 Specify the pipe line conditions of steam at a pressure of 2.25 kPa for
the following throttling-calorimeter conditions:
(a) 0.1 MPa, 100° C
(b) 0.1 MPa, 99.6°C (not 99.63°C)
(c) 0.1 MPa, 200° C
9-27 Refrigerant F-12 is compressed isentropically from 3.0861 bars, 20°C to
a pressure of 9.6065 bars. Calculate the value of the isentropic exponent
k.
9-28 Steam expands isentropically in a closed system from 15 MPa, 550°C to
4.5 MPa. Calculate the value of k and use this to calculate the work
done per kilogram of steam. Compare this work with that calculated
using the change in internal energy.
9-29 Steam at 10.0 MPa, 350°C is throttled to 1.0 MPa. The lowest available
temperature is 40°C. Determine the loss of available energy as a result
of the throttling.
9-30 Determine the compressibility factor for steam for the following con¬
ditions:
(a) 0.01 MPa, 50°C
(b) 0.1 MPa, 100° C
(c) 1.0 MPa, 300° C
(d) 1.0 MPa, 800° C
(e) 10.0 MPa, 325°C
(f) 10.0 MPa, 800° C
9-31 From the results of Problem 9-30, is it possible to consider steam to be
an ideal gas under any conditions? If so, specify these conditions.
210 Vapors

9-32 A 0.29-m3 tank contains 165.5 kg of refrigerant F-12 at a pressure of


8.477 kPa. Determine the percentage of the refrigerant in the liquid
state.
9-33 Using the value of cp from Figure 4 in Keenan, Keyes, Hill, and Moore
at 20 MPa, 450°C and the general equations from Chapter 8, determine
the value of c„.
9-34 Steam expands in a supersaturated manner from 3.0 MPa, 300°C to
0.3 MPa. Using a mean value of the isentropic exponent k from Figure
6 in J. H. Keenan, F. G. Keyes, P. G. Hill, and J. C. Moore, Steam
Tables (Wiley, New York, 1969, 2nd ed. 1978), calculate the temperature
after expansion and compare with the saturation temperature.
9-35 The mercury vapor leaving the turbine in Problem 9-23 is condensed to
produce steam for a steam turbine. Dry saturated steam is produced at
a pressure of 1.10 MPa. Water is supplied at 150°C. Determine the
amount of steam produced per hour.
9-36 Steam enters a condenser at 3.0 kPa with a moisture content of 9.8
percent. Condensate leaves at 24°C. Circulating water (to condense the
steam) enters at 15°C at a rate of 42,000 kg/s. It leaves at 22°C.
Determine the rate of steam flow to the condenser.
9-37 (a) Refrigerant F-12 enters a compressor at 1.966 bars, 0°C. It is
compressed to 7.449 bars. The compressor efficiency (the ratio of the
theoretical to actual work) is 80 percent. Determine the actual enthalpy
at compressor exit.
(b) Determine the compressor power input for part (a), for a refrigerant
flow of 62.5 kg/min.
9-38 The vapor leaving the compressor in Problem 9-37 is condensed and
leaves the condenser with no subcooling. Water is supplied at 24° C to
condense the refrigerant. It leaves the condenser at 28°C. Determine
the rate of flow of the water.
9-39 (a) Steam expands from a pressure 12.5 MPa to 0.003 MPa. The
theoretical moisture content after expansion is not to exceed 22
percent. Determine the minimum initial steam temperature.
(b) If the turbine efficiency (ratio of actual to theoretical work) in part
(a) is 80 percent, determine the actual final moisture content.
9-40 Determine the volume rate of flow of steam to the condenser in
Problem 9-20 in m3/s.
_ 10 _

NONREACTIVE GASEOUS
AND VAPOR MIXTURES

10-1 INTRODUCTION
Thus far in this book, it has been assumed that the systems under con¬
sideration contained only a single pure substance. This assumption is
sufficiently accurate for many engineering systems, such as a steam turbine, a
single-fluid gas compressor, a pump, or a refrigerating system. Often,
however, the system may contain two or more substances in sufficient
amounts to make it necessary to examine the effects of each substance on the
properties and the behavior of the system. These substances may or may not be
in chemical equilibrium.
Unless the substances are in gaseous form and at such temperatures and
pressures that they behave as ideal gases, intermolecular forces will exist.
These intermolecular forces do not obey any simple law and, therefore, it is
difficult to determine the properties and behavior of the mixture. However, in
engineering practice many gaseous mixtures are encountered where the gases do
closely resemble ideal gases. Examples of such systems are the gas turbine, the
gas passageways in a steam generator (including the economizer and air
preheater), the internal combustion engine during compression, expansion, and
exhaust, and the gas passageway in a magnetohydrodynamic generator.
In many of these systems, the gaseous mixture is either in chemical
equilibrium or does not undergo a reaction during the time period under
consideration. Hence, it will be assumed that no chemical reaction is involved
in the systems under discussion in this chapter.

10-2 PROPERTIES OF IDEAL GASEOUS MIXTURES


As is discussed in Section 1-10, the pressure exerted by an ideal gas occupy¬
ing a given volume is directly proportional to the product of the number of
molecules of the gas and the mean molecular kinetic energy. It is independent
of the chemical composition of the gas. Consider a system of fixed volume
containing the ideal gas A. For a fixed temperature, gas A exerts a pressure
that is proportional to the number of its molecules present. Now let a second
ideas gas B having the same temperature as gas A be introduced into the

211
212 Nonreactive Gaseous and Vapor Mixtures

system. Since both gases are ideal, no intermodular attractions can exist.
Furthermore, since the two gases are at the same temperature, there can be
no net change in molecular kinetic energies. Hence, each gas exists in¬
dependently of the other gas. Gas A will exert the same pressure as it did
prior to the introduction of gas B. Likewise, gas B will exert a pressure equal
to the pressure it would have had if it existed alone in the system. The
pressure exerted by each gas in a gaseous mixture is known as its partial
pressure. Thus, the partial pressure of each gas in a mixture is the pressure
that the gas would exert if it alone occupied the entire volume at a tem¬
perature equal to that of the mixture.
The preceding observations can be formulated as Dalton's law of partial
pressures. This law states that for mixtures of ideal gases, each gas occupies
the entire volume at its own partial pressure, and the total pressure of the
gaseous mixture equals the summation of the individual partial pressures. In
equation form this law is

P = Pa + Pb+ Pc- - - (10-1)


where p represents the total pressure and each term on the right-hand side
represents the partial pressure of one gas.
Since each gas is an ideal one, the characteristic equation applies to each
gas. Thus,

_ NaRqT _ NbRqT _ NCR()T


Pa- y Pb~ y Pc- y

where N is the number of moles and R0 is the universal constant.


Substitution of these values of partial pressures in Eq. 10-1 gives

= NaR0T NbR,T NcRqT _ (Na + Nb + Nc)RqT


p y y y y WU-zj

But, since the total number of moles, N, equals NA + NB + Nc,

pV = NR0T (10-3)

As shown by Eq. 10-3, the characteristic equation of a mixture of ideal gases


is of the same form as that of a single ideal gas. Thus, mixtures of ideal gases
may be treated in the same manner as single ideal gases as far as the p-^-T
relationships are concerned.
Avogadro stated that equal volumes of ideal gases under the same con¬
ditions of temperature and pressure contain the same number of molecules.
Since, by definition, the number of molecules per mole is a constant, equal
volumes of ideal gases under the same conditions of temperature and pressure
must contain the same number of moles. Conversely, the volume of a mole of
any ideal gas depends solely on its temperature and pressure and is in¬
dependent of the nature of the gas. This statement should be evident from Eq.
10-2.
Properties of Idea! Gaseous Mixtures 213

The term partial volume is used in connection with gaseous mixtures. It


designates the volume that would be occupied by a gas in a gaseous mixture if
that gas were to be separated from the mixture and exist at the temperature
and total pressure of the mixture.

Example 10-1. A tank having a volume of 0.6 m3 contains oxygen at 25°C and
480 kPa. Nitrogen is introduced into the tank without producing a change in
temperature until the pressure becomes 920 kPa. Determine the mass of each
gas and its partial volume.

Solution. For the oxygen,

_ pV _ 480 x 1000x0.6
116.18 g moles
R0T 8.314(25 + 273.15)

Since there are 32 g/g mole oxygen,

mass of oxygen = 32 x 116.18 = 3717.8 g

For the nitrogen,


p = 920 - 480 = 440 kPa

Then
_ 440 x 1000 x 0.6
106.5 g moles
N 8.314(25 + 273.15)'

mass of nitrogen = 28 x 106.5 = 2982 g Answer

Each gas occupies the total volume at its own partial pressures. The volume it
would occupy at the total pressure = partial pressure/total pressure x total
volume. For the oxygen,
480
partial volume = x 0.6 = 0.313 Answer
920

For the nitrogen,


440 ,
partial volume = x 0.6 = 0.287 m Answer

Since the energy of a system is composed of the energies of its component


parts, the molecular energy of a gaseous mixture in a closed system must
equal the sum of the molecular energies of its constituent gases. Thus,

U = UA + UB+ (10-4)

When the gases are ideal ones, the product pV for the mixture equals the
sum of the values of pV of the individual gases. This statement may be
demonstrated as follows:
n = na + nb + nc...
214 Nonreactive Gaseous and Vapor Mixtures

Then

py _ PaVa , PbVb , PcVc


R0T R0T R0T R0T

or

pV = pAVA + PbVb + PcVc

Adding Eq. 10-4 to this,

U -I- pV = UA + pAVA + Ub + PbVb + Uc + PcVc

Since
H = U+pV
(10-5)
H = HA + HB +Hc

When the gases are not ideal, the total pressure is not necessarily equal to
the sum of the individual partial pressures. Hence, the product pV does not
necessarily equal the sum of the individual values of pV. Under these
conditions it follows that the enthalpy for a closed system of a mixture of
gases that are not ideal does not necessarily equal the sum of the enthalpies of
the individual components of the mixture.
For ideal gases, with a datum at zero degrees absolute temperature

U = Nc„T

Hence,

NcvT = (NcvT)a 4- (NcvT)b + (NcvT)c...

Thus, the specific heat of a mixture of ideal gases is

(„ —
_(Nc*)a
- + (NcJb + (Nc*)c • • •
(10-6)
N

In a similar manner,

(Ncp)a + (Ncp)fl + (Ncp)c


Cp (10-7)
N

A useful term, which may be used in connection with Eqs. 10-6 and 10-7, is
the mole fraction. The mole fraction is defined as the ratio of the number of
moles of any one gas to the total number of moles present. Thus,

where xf = the mole fraction of substance i


Nj = the number moles of substance i
N = the total number moles present
Properties of Ideal Gaseous Mixtures 215

Using the concept of the mole fraction, Eq. 10-6 becomes

c,, = (xc„)A 4- (xcv)B + (xc„)c ■ • • (10-6a)


Similarly, Eq. 10-7 becomes

cp = (xcp)A 4- (xcp)B 4- (xcp)c... (10-7a)

The partial entropy of a gas in a gaseous mixture is the entropy that the
individual gas will possess when it occupies a volume equal to the volume of
the gaseous mixture and has the same temperature. Gibbs’ law of partial
entropies states that the total entropy of a mixture of ideal gases equals the
sum of the partial entropies of the individual constituents. This law may be
demonstrated as follows.
Consider the cylinder in Figure 10-1. In the cylinder on the right-hand side
is a movable gas-tight piston, which is shown as solid. To the right of this
piston is a complete vacuum. To the left of this piston is a semipermeable
membrane, rigidly attached to the cylinder walls. It will permit the passage of
one gas, gas B, through itself but will not permit gas A to pass through. To the
left of the semipermeable membrane is a mixture of two ideal gases, gas A
and gas B. In the extreme left of the cylinder is a second semipermeable
membrane, which will permit the passage of gas A but not gas B. This
membrane, which is movable, is connected to the solid piston by means of a
piston rod.

(b)

FIGURE 10-1 Reversible separation of gases, (a) Start. (b) Intermediate, (c) Finish.
216 Nonreactive Gaseous and Vapor Mixtures

Assume that the coupled pistons have been moved an infinitesimal distance
to the right (see Fig. 10-la). Gas B in the very small space to the left of the
solid piston exerts a pressure (to the right) of pB on the solid piston. Gas A in
the very small space to the left of the membrane piston exerts a pressure (to
the right) of pA on the membrane piston. These two pressures exactly equal
the pressure of pA -I- pB acting to the left on the coupled pistons. Assuming an
isolated system and no friction, the application of an infinitesimal force to the
right on the coupled pistons will cause them to move. When the infinitesimal
force is applied intermittently, gas diffusion through the membranes will
maintain equality of the pressures of the gas involved (see Fig. 10-lb). Hence
the forces acting on the coupled piston are balanced at all times and remain
so, provided that no finite force is applied to the pistons.
Continued application of the infinitesimal force will cause the pistons to
move completely to the right (see Fig. 10- 1c). There will be complete
separation of the gases. Since only an infinitesimal force was required, the
work required for separation also is infinitesimal, and the separation process
may be assumed to be reversible. Hence, no entropy change occurred during
the separation. Thus, the entropy of a gaseous mixture equals the summation
of the entropies of the individual gases existing at their partial pressures in the
mixture and at mixture temperature, or

S = Sa + SB + Sc • • • (10-8)
where the terms SA, SB, and Sc signify partial entropies.

10-3 IRREVERSIBLE MIXING OF IDEAL GASES


WITHIN A SYSTEM
In this and in following sections it will be assumed that we are dealing with
finite quantities of gases, and hence, the qualifications placed on statements
relative to entropy changes for microscopic states are not applicable. In this
section, consideration will be given to changes in properties of gases as a
result of mixing. As a first step, assume that gas A occupies the left side of
the isolated cylinder in Figure 10-2 and that the gas expands irreversibly and
occupies the entire cylinder. The increase in entropy that occurs may be
calculated since the terminal states are fixed.

Removable partition

* ■ .m ■ • ;

Gas A Gas B
1 2 !v!v!

FIGURE 10-2 Mixing of gases in a closed system.


Irreversible Mixing of Ideal Gases within a System 217

Example 10-2. Compartment 1 in Figure 10-2 has a volume of 0.3 m3. It


contains oxygen at a pressure of 440 kPa and a temperature of 80°C. Com¬
partment 2, having a volume of 0.28 m3, is completely evacuated. Calculate
the entropy change when the partition is removed.

Solution. Assuming that the system is completely isolated, there will be no


change in its energy, and hence, no change in temperature. Then the entropy
change

S2~ S, = mR In
pv
In
©
440 x 0.3 ] 0.58
= 0.2465 kJ/K Answer
80 + 273 0.3

Now assume that with the partition in place, compartment 2 in Figure 10-2
is filled with an ideal gas B, which is not reactive with gas A and which has
the same temperature as gas A. When the partition is removed, gas A expands
and occupies the total volume. The presence of gas B does not change the
final condition of gas A. Since the initial and final conditions of gas A are
independent of the presence of gas B, the change in entropy of gas A must be
the same as it was when that gas expanded into a complete vacuum. Upon
removal of the partition, gas B also expands irreversibly. The change in
entropy for gas B can be found by adopting a procedure similar to that
described for gas A.
Upon removal of the partition, there is a driving force (i.e., the individual
gas pressure) that causes movement of each gas into the other compartment,
tending to equalize the partial pressure of each gas through the cylinder.
Putting it differently, the state of either gas is an improbable one at the time
of removal of the partition and, hence, will move to a more probable state.
As discussed in Chapter 7, when a system moves from a less probable to a
more probable state, there is an increase in entropy.
Now consider the two compartments filled with the same kind of gas under
the same conditions of temperature and pressure. Will there be an increase in
entropy when the partition is removed? The answer is no. It is true that some
molecules in compartment A will move into compartment B and vice versa
upon removal of the partition. But there is no pressure difference or driving
force to cause a change in state. Since there is no change in state of the gas,
there cannot be a change in entropy. Looking at the problem in a different
way, the most probable configuration of the molecules is an equal division
between the two compartments. Since they are in their most probable state,
removal of the partition cannot increase their probability. Hence, the entropy
cannot change.
When two gases are not under the same conditions of temperature and
pressure, there will be an exchange of energy between the two upon mixing.
Hence, the entropy change cannot be found so readily as in the previous case.
If the two gases originally occupied adjacent compartments and the mixing
218 Nonreactive Gaseous and Vapor Mixtures

took place in a closed system, the total internal energy could not change
during the mixing, provided that the system was isolated from its surround¬
ings. Since the two gases will come to equilibrium temperature, this tem¬
perature may be found. The final temperature fixes the final state of each gas.
Hence, the entropy change of each gas can be calculated.

Example 10-3. Same conditions as in Example 10-2 except that compartment


2 contains nitrogen at a pressure of 650 kPa and a temperature of 130°C.
Determine the change in entropy of each gas which occurs when the partition
is removed and the gases reach equilibrium conditions.

Solution. The moles of oxygen present are

pV 440x 1000x0.3 .. no i
or 1439 g
N°2 ~ R0T ~ 8.314(80 + 273) ~ 44 98 m0 eS

The moles of nitrogen present is

650 x 1000 x 0.28


Nn,= = 54,32 moles or 1521 g
8.314(130 + 273)

Equating the internal energy changes,

AUq2 = AUn2
1439.0 x 0.651(tf - 80) = 1521 x 0.734(130 - tf)
The final temperature
tf = 106° C

The final pressure of the oxygen is

P o. = N^°T
V
= 44,98 X
0.58
= 244,400 Pa = 244 kPa

The final pressure of the nitrogen is

54.32x8.314(106 + 273)
P N, = 295,100 Pa = 295.1 kPa
0.58

For the oxygen,

P_2
S2 — Si = ntCp In — mR In
11 Pi
/106 + 273N 8.314, /244,400\
= 1439x0.908 In 1439 x
V 80 + 273 ) \440,000/
= 92.86 + 219.82 = 0.3127 kJ/K

For the nitrogen

/106 + 273X 8.314 /295,100\


S2 -s 1 ~
1521 x 1.029 In 1521 x
\130 + 273/ 28 n \650,000/
= 0.2274 kJ/K Answer
Mixing of Ideal Gases During Flow 219

Note that the change in entropy of the oxygen exceeds that of Example
10-2. The cause of the increase is the heat the oxygen has received from the
nitrogen.
When gases are mixed in a given space, the temperature and, hence, the
pressure of the mixture are fixed by the total internal energy of the gases
forming the mixture. Therefore, all of the other properties of the mixture are
also fixed.
It has been inferred that the gases under consideration must be nonreactive
(i.e., can never react). Ideal reactive gases (can react under certain conditions)
may be treated in a similar manner, provided that the conditions during
mixing do not permit reaction. Many common reactive gaseous mixtures must
be heated to temperatures much in excess of room temperatures to cause a
measurable reaction to take place.

10-4 MIXING OF IDEAL GASES DURING FLOW


The fundamental principles developed for mixtures of gases without flow
apply equally well for flowing gases. However, instead of dealing with known
fixed volumes, mass rates of flow may be specified. For a flowing fluid, the
enthalpy of the fluid is part of its total energy. If the velocity is high, kinetic
energy must also be considered in evaluating the total energy. By means of an
energy balance, the final energy, and hence, the final temperature can be
determined. This temperature, together with the partial pressure of any gas,
permits the determination of its other final properties.

Example 10-4. A stream of oxygen at 300 kPa and 25° C, flowing at the rate
of 8 kg/min, mixes with a stream of nitrogen having a pressure of 520 kPa and
a temperature of 80°C. The nitrogen rate of flow is 14 kg/min (see Fig. 10-3).
Determine the change in entropy of the oxygen per minute.

Solution. Since the velocities of the gases are not given, it will be assumed
that they are sufficiently small to permit changes in kinetic energy to be
neglected. It will be assumed also that the mixing is adiabatic. A third

FIGURE 10-3 Mixing of gases in Example


10-4.
220 Nonreactive Gaseous and Vapor Mixtures

assumption to be made is that the pressure after mixing equals the lowest gas
stream pressure, namely 300 kPa. For the assumptions made, there can be no
change in the total enthalpy of the gases. Then the change in the enthalpy of
oxygen must equal the change in enthalpy of the nitrogen. Recognizing that
dH = mcp dt, for a time period of a minute,

8 x 1000x0.908(t/-25)= 14 x 1000 x 1.029(80 - tf)


tf = 61.53°C

The final partial pressure of each gas is proportional to the relative number
of moles of that gas.
The number of kilogram moles of oxygen equals 8/32. For nitrogen this
number equals 14/28. Then the final partial pressure of the oxygen,

8/32
Po2 x 300 = 100 kPa
8/32 4- 14/28

s2-s, = men In T - mR In Pi
T, Pi
61.53 + 273.15
= 8 x 1000 x 0.908 In 8 x 1000 x In P00
( 25 + 273.15 ) 32 V300,
= 3123 J/K Answer

10-5 VOLUMETRIC AND GRAVIMETRIC ANALYSIS


The analysis of a gaseous mixture is commonly made by chemically absorbing
each gas, one at a time, and noting the volume change that occurs in the
mixture during absorption, holding both the temperature and the total pres¬
sure constant. Each volume change equals the partial volume of the gas that
was absorbed. The ratio of the partial volume of any gas to the total volume is
the volumetric fraction of the gas that is present. Unless otherwise stated, the
analysis of gaseous mixtures is reported on the volumetric basis.
In certain cases it is desirable to convert the analysis to a gravimetric, or
weight, analysis. (Strictly speaking, this is a mass analysis.) The method of
conversion will be illustrated by using air. Air is composed mainly of oxygen
and nitrogen, but it does contain a slight amount of carbon dioxide plus traces
of rare gases, such as argon, krypton, and neon. Atmospheric air also contains
water vapor in varying amounts. Because of the varying amounts of water
vapor present in air, it is customary to speak of dry air, which comprises all
the constituents of air except water vapor. It is, furthermore, customary to
lump the rare gases, the carbon dioxide, and the nitrogen together, and to call
the combination atmospheric nitrogen. Thus, dry air may be said to be
composed of oxygen and atmospheric nitrogen.
Atmospheric nitrogen has an equivalent molecular weight of 28.16. On the
volumetric basis, dry air is 21.0 percent oxygen and 79.0 percent atmospheric
nitrogen. Consider 1 mole of dry air. Since the volume per mole of ideal gases
under the given conditions of temperature and pressure is a fixed quantity, a
Gas-Vapor Mixtures 221

volumetric analysis is also a molal analysis. The gravimetric analysis of dry


air may therefore be worked out in the following manner.

Molal Moles Grams Grams Grams


Analysis per X per = per -r- Total = per
Substance Percent Mole Air Mole Mole Air Grams Gram Air
Oxygen 21.0 0.210 X 32 6.72 - 28.97 = 0.232
Atmospheric 79.0 0.790 X 28.16 = 22.25 h- 28.97 = 0.768
Nitrogen
28.97

The gravimetric analysis of air is, then, 23.2 percent oxygen and 76.8 percent
atmospheric nitrogen. These calculations also show that the mass of a gram
mole of air is 28.97 g.
A gravimetric analysis may be converted to the volumetric basis in a
reversed manner, as illustrated in the following example.

Example 10-5. The gravimetric analysis of a gaseous mixture follows: C02,


15 percent; 02, 8 percent; N2, 77 percent. Determine (a) the volumetric
analysis, (b) the equivalent weight per mole, and (c) the gas constant for the
mixture.

Solution
(a) The calculations may be arranged as follows:
Grams per Grams Moles per Total Moles per Percent
Substance Gram Gas -r per Mole = Gram Gas -T- Moles - Mole Gas by Volume
co2 0.15 - 44 = 0.00341 -h 0.03341 = 0.102 10.2
o2 0.08 - 32 = 0.00250 -s- 0.03341 = 0.075 7.5
n2 0.77 - 28 = 0.02750 h- 0.03341 = 0.823 82.3
0.03341

(b) The calculations show that there is 0.03341 mole/g gas. Hence, the weight
per mole is
1
29.9 g
0.03341

(c) The gas constant

r = *11^ = 0.2781 N-m/g-K or 0.2781 J/g-K Answer

10-6 GAS-VAPOR MIXTURES


In Chapter 9 it was pointed out that vapors normally do not obey the ideal-gas
laws. Two of the primary reasons are that the size of a molecule of vapor is
appreciable in comparison with the total volume and that the molecular
attractions are sufficient to exert an influence on the observable vapor
pressure. The deviations from the ideal-gas laws are particularly large when
the vapor exists at a high reduced pressure and a low reduced temperature.
On the other hand, at a very low reduced pressure the behavior of a vapor is
approximated very closely by the ideal-gas laws.
222 Nonreactive Gaseous and Vapor Mixtures

The properties of vapors were discussed in Chapter 9, where it was


assumed that only the vapor of a pure substance was present. When the
vapors of two or more substances form a mixture, the determination of the
properties of the mixture is difficult if either or both vapors show marked
deviations from the ideal-gas laws. However, such mixtures are not encoun¬
tered very frequently in ordinary engineering work.
A very common mixture is a gas-vapor mixture existing under such a low
pressure that the vapor, as well as the gas, closely obeys the ideal-gas laws.
Under these conditions, the mixture may be treated as a mixture of ideal
gases, which was discussed earlier in this chapter. There is one important
difference between the gas-vapor mixture under consideration here and the
mixture of ideal gases. Unless there is a radical change in temperature, the
latter gases will remain in the gaseous phase. However, a decrease in
temperature of a few degrees may cause partial condensation of the vapor in
a gas-vapor mixture. When some of the substance that is in vapor form is
present in the liquid phase, partial vaporization will take place if any heating
occurs.
The most common gas-vapor mixture encountered in engineering work is a
mixture of air and water vapor. This mixture will be discussed in this chapter.
The principles involved will apply equally well to any other gas-vapor mixture
existing at such a low pressure that the vapor may be treated as an ideal gas.

10-7 AIR-VAPOR MIXTURES


Atmospheric air, as stated previously in this chapter, is composed of oxygen,
nitrogen, carbon dioxide, water vapor, argon, and other rare gases. Dry air
includes all constituents of atmospheric air except water vapor. In many
problems dealing with a mixture of air and water vapor, the total pressure is
substantially atmospheric. Unless otherwise stated in this discussion, it will be
assumed that the total pressure is atmospheric. Furthermore, the additional
restriction will be imposed that there will be no substantial change in the total
pressure of the mixture, even though there may be a change in the tem¬
perature of the mixture or a change in the amount of vapor present. Problems
under consideration may involve a system existing in pressure equilibrium
with its surroundings—in this case, the atmosphere—or a system in which
steady-flow conditions exist. For many common engineering problems in¬
volving steady flow of air-vapor mixtures, the rate of flow is so low that there
is no appreciable change in the absolute pressure.
The term atmospheric air does not infer that the air exists at atmospheric
pressure but rather that the dry air has a specified composition. Unless
otherwise stated, when the term “air” is used, it is implied that the mixture
consists of dry air having the same composition as dry air in the atmosphere
and a varying amount of water vapor.
Unless the temperature is above 65°C, the partial pressure of water vapor
in a mixture of air and water vapor is sufficiently low to permit it to be treated
Air-Vapor Mixtures 223

as an ideal gas in most engineering applications. As such, it occupies the


entire volume under consideration at its own partial pressure, which is the
same whether or not dry air is also present. Generally the water vapor in air is
superheated, since it exists at a temperature in excess of the saturation
temperature for its given partial pressure. When water is present in the space
occupied by the vapor, there is a tendency for vaporization to take place,
provided that the vapor is not already saturated. This tendency toward
vaporization does not depend on the presence or absence of air. It is common
practice to speak of “saturated air.” This is a misnomer, since this implies
that the space occupied by the air contains saturated water vapor.
The term humidity refers to the amount of vapor present in an air-vapor
mixture. Absolute humidity is the mass of water present in a given volume. It
may be expressed as the number of grams of water vapor in a cubic meter of
space. Relative humidity is the ratio of the actual absolute humidity to the
maximum absolute humidity at any given temperature. In equation form this
definition becomes

= /-jctugl mv_ X
\saturated mv/T V

where <f> is relative humidity. Treating the vapor as an ideal gas, Eq. 10-9 may
be rewritten as
(pV/JRT)actual vapor P actual
(10-10)
(P V/-RT) saturated vapor Psaturated

Some authors give Eq. 10-10 as the defining equation for relative humidity.
For steady-state conditions, the mass rate of flow of dry air is constant for
the system under consideration. This statement may not be true for the water
vapor, since there may be vaporization of water or condensation of water
vapor in the system. It is desirable to know the mass rate of flow of water
vapor in various parts of the system. Since the mass rate of flow of air is
constant, the mass rate of flow of vapor at any point can be determined by
establishing a relationship between the mass of vapor and the mass of dry air.
The ratio of the mass of vapor to the mass of dry air is known as the humidity
ratio, (o, also called the specific humidity. It is defined by the following
equation:

mass of water vapor


(10-11)
mass of dry air

When the vapor obeys the ideal-gas laws, Eq. 10-11 may be rewritten as
follows: Use the relationship m = pV/RT for both the vapor and dry air in
Eq. 10-11 and also R = R0/m (m is the mass per mole). Then

(pm)y _ 18.016pv _ 0.622pv


(pm) da 28.97pda Pda
224 Nonreactive Gaseous and Vapor Mixtures

or

a) = 0.622 ——— (10-12)


Pt Pv

where pt is the total pressure.


There is no convenient way1 to measure either the relative humidity or the
humidity ratio directly. They must therefore be calculated. It is to be noted
that in Eqs. 10-10 and 10-12 the pressure of the vapor is required. There are
two indirect methods that are commonly used for determining this pressure.
One of these methods involves the determination of the dew point. The dew
point is defined as the temperature at which condensation first starts when air
is cooled at a constant total pressure. During the cooling process, before
condensation starts, there is no change in the relationship between the
amounts of dry air and water vapor. Hence, there can be no change in the
relationship between the partial pressures of the two. Since the total pressure
remains constant, the vapor pressure must remain constant until the dew
point is reached (see Fig. 10-4). The vapor pressure at the dew point, and
hence, at the initial condition before cooling, is the saturation pressure for the
dew-point temperature. Unless precision equipment is available, it is difficult
to determine the dew-point temperature accurately.
The second method of determining the partial pressure of the vapor
involves the wet-bulb temperature. This temperature is the equilibrium tem¬
perature observed on a thermometer whose bulb is covered with a layer of
water. Frequently, a wet-bulb thermometer and a dry-bulb (ordinary) ther¬
mometer are mounted together. This combination is known as a psychrometer.
When the air in contact with the wet bulb is not saturated, vaporization of the
water takes place on the wet bulb. Since heat is required for vaporization,

'it is possible by chemical or physical-chemical means to remove substantially all water vapor
from air. A measurement of the amount of vapor removed permits a determination of the original
humidity of the air. Such a method, however, calls for considerable equipment.

FIGURE 10-4 Dew-point determination.


Air-Vapor Mixtures 225

Dry bulb

both the temperature of the air and that of the water on the wet bulb decrease
until equilibrium is reached. The difference between the wet-bulb and dry-
bulb temperatures is known as the wet-bulb depression. To maintain continued
vaporization, and hence, a constant wet-bulb temperature, there must be relative
motion between the wet bulb and the air surrounding it. A common form of the
wet-bulb psychrometer is the sling psychrometer, shown in Figure 10-5. This
psychrometer is twirled around, thus providing the necessary relative motion
between the air and the wet bulb.
The wet-bulb temperature is a saturation temperature. So is the dew-point
temperature. However, the wet-bulb temperature is the higher of the two.
Consider air flowing over the wet bulb. Since the air picks up moisture as it
passes over the wet bulb, the saturation temperature of the air leaving the wet
bulb (the wet-bulb temperature) must exceed that of the air approaching the
wet bulb (the dew-point temperature). The relationship between these tem¬
peratures is shown on the T-s plane of Figure 10-6.
In the ideal case, the process taking place on the surface of the wet-bulb
thermometer is an adiabatic saturation process. This process may be con¬
sidered by referring to Figure 10-7. The chamber in this figure is insulated so
that the process is adiabatic. Here air passes over the surface of water. There

FIGURE 10-6 T-s diagram for wet- and dry-bulb temperature.


226 Nonreactive Gaseous and Vapor Mixtures

FIGURE 10-7 Adiabatic saturation.

is sufficient turbulence to ensure that the air is completely saturated upon


leaving the chamber. It is assumed that the water reaches the temperature of
the leaving air, that is, the adiabatic saturation temperature. Making an energy
balance, the energy of the entering air plus the energy of the water that is
vaporized equals the energy possessed by the air leaving the chamber.
Although the energy of water at rest is internal energy, water must be
supplied continuously to maintain the water level in the chamber. Under these
conditions, the energy of the water that is vaporized is its enthalpy. Making
an energy balance per mass of dry air,

(hda "h Ojhy')] T (tt)wb OJ\)hfwb (hda "b ^^v)wb (10-13)


where subscript 1 refers to the entering air and the subscript wb refers to the
saturated property at the wet bulb temperature. The term {a>wb - <x>,)
represents the amount of water vaporized per unit mass of dry air.
Equation 10-13 may be solved for the humidity ratio of the entering air,

_ ^dawb ~ ^da, + Wwh(^vwb ~ ^/wb) nn . ^


(UM4)

Steam tables may be used to evaluate the enthalpy of water and water
vapor. At very low pressures, that is, below 0.01 MPa, the enthalpy of water
vapor is substantially independent of the pressure. Hence, the enthalpy of
superheated vapor may be taken to be equal to that of dry saturated vapor at
the given temperature. When steam tables are not available, the enthalpy of
water vapor may be determined by use of empirical equations. The following
equation is suggested by the authors:

hv = 2500 + 1.881 (J/g) (10-15)

where t is °C. The maximum error caused by the use of Eq. 10-15 is 1.9 J/g
over the temperature range of — 20°C to 100°C and for vapor pressures up to
0.01 MPa. Using a datum of 0°C for evaluating the enthalpy of air and treating
the air as an ideal gas.

hda= 1.0041 (J/g) (10-16)


Air-Vapor Mixtures 227

Example 10-6. Air at standard atmospheric pressure has a wet-bulb temper¬


ature of 20°C and a dry-bulb temperature of 28°C. Determine the humidity ratio.

Solution. From Eq. 10-12,

2.339
a>wb = 0.622 Pv = 0.622
P, “Pv 401.325 — 2.330,
= 0.0147 gig dry air

Assuming that the specific heat of water is 1 cal/g-°C or 4.1868 J/g-°C,

hfwb = 4.1868x20 = 83.7 J/g


hV] = 2500 4- 1.88 x 28 = 2552.6 J/g
hVwb = 2500 + 1.88 x 20 = 2537.6 J/g
From Eq. 10-14,

_ 1.004x20- 1.004x28 + 0.0147(2552.6-83.7)


" 2552.6-83.7
= 0.0114 gig dry air

For normal atmospheric wet-bulb temperatures, the enthalpy of the water


at the wet-bulb temperature is relatively small. If this enthalpy in Example
10-6 is neglected, there will be an error of a little over 1 percent in the
calculated humidity ratio. Unless the pertinent data is measured very ac¬
curately, the enthalpy of the water at the wet-bulb temperature, hfwb, in Eq.
10-14 can be neglected.
Investigation has disclosed that the process taking place on the wet bulb is
not a true adiabatic saturation process for the following reasons:

1 There is conduction of heat along the stem of the wet-bulb thermometer.


2 The wet bulb receives heat by radiation from its surroundings.
3 Impact between the air and the wet bulb causes a slight increase in the
wet-bulb temperature.
4 The air flowing over the wet bulb is separated from it by an air film. There
must be a temperature drop across the air film to cause heat to flow into
the wet bulb. There also must be a pressure differential across the film to
cause the vapor to diffuse outward.

Because of these four items, the observable wet-bulb temperature can,


conceivably, differ materially from the temperature of the adiabatic satura¬
tion. In fact, for certain liquids vaporizing into certain gases, the difference
between these two temperatures may amount to several degrees. Fortunately,
the observable wet-bulb temperature for water vaporizing into air is sub¬
stantially equal to the adiabatic saturation temperature for temperatures up to
65°C, provided that the relative velocity between the wet bulb and the air is
approximately 300 m/min. Hence, these two temperatures will be used inter¬
changeably from now on.
228 Nonreactive Gaseous and Vapor Mixtures

10-8 PSYCHROMETRIC CHARTS


Various types of psychrometric charts have been devised to permit easy
determination of the properties of moist air. Most of these charts have the
dry-bulb temperature as the abscissa and the specific humidity as the ordinate.
As may be seen from Eq. 10-12, the vapor pressure is almost a straight-line
function of the specific humidity and, hence, vapor pressure is frequently
shown as an auxiliary ordinate scale.2 Most of these charts are prepared for
the standard barometric pressure of 101.325 kPa or 1.01325 bars, although
some charts provide for determining properties at other barometric pressures.
Psychrometric charts are based on the principles discussed in Section 10-7.
In addition to the dry-bulb temperatures, specific humidities, and vapor
pressures, most charts show wet-bulb temperatures, relative humidities, and
specific volumes of dry air. A skeleton psychrometric chart is shown in Figure
10-8. A reproduction of a commercial chart is in the Appendix at the back of this
book.
An approximation may be made so that the heat transferred in an air-
conditioning problem may be determined rather readily with the use of a
psychrometric chart. Refer to Eq. 10-13. The right-hand side of this equation
has a fixed value for each wet-bulb temperature. The left-hand side of the
equation consists of two quantities: (1) the enthalpy of the air, or hda + (ohV];
(2) the enthalpy of the water that is vaporized, or (cowb- coi)hfwb. Except for
very high temperatures and humidities, the enthalpy of the water is extremely
small. Neglecting this enthalpy, the enthalpy of air at any dry-bulb tem-

This statement is valid for the low vapor pressures normally encountered in air-conditioning
work.

FIGURE 10-8 Skeleton psychrometric chart.


Simple Heat Transfer 229

perature must equal the enthalpy of saturated air at the given wet-bulb
temperature. The enthalpy of saturated air may be computed at each wet-bulb
temperature, and the value may be entered at the wet-bulb temperature on the
psychrometric chart. When there is a change in conditions and, hence, a
change in the wet-bulb temperature of air, the heat added or removed during
the change may be determined by taking the difference between the enthalpies
as found on the psychrometric chart for the two values of the wet-bulb
temperature. This statement assumes that the enthalpy of water that may be
added or removed during a change in state can be neglected.

10-9 SIMPLE HEAT TRANSFER


In the simple case of heat transfer, it is assumed that neither condensation nor
vaporization takes place. It will be further assumed that the process will be
one of flow with no substantial change in the total pressure. With these
assumptions, there can be no change in either the vapor pressure or the
specific humidity. However, there will be a change in the relative humidity.
During heating there will be a decrease in the relative humidity, and there will
be an increase in relative humidity during a cooling process.

Example 10-7, Air at a temperature of -10°C, 90 percent relative humidity is


heated to 20° C. Determine the relative humidity after heating.

Solution. It may be assumed that this is a flow problem with no significant


change in the total pressure. From Table 6 of Steam Tables by Keenan,
Keyes, Hill, and Moore, the saturated vapor pressure at — 10°C is 0.2602 kPa.
By Eq. 10-10, the actual vapor pressure, actual pv = 0.9 x 0.2602 = 0.2342 kPa.
Since there is no change in the amount of vapor present and since there is no
change in the total pressure, the final vapor pressure = 0.2342 kPa. At 20°C,
the saturated vapor pressure is 2.339 kPa.
Then the final relative humidity,

. 0.2342 1A
4> = 2239 x 00 ~ 10-02 percent Answer

Example 10-8. Air at 28°C has a humidity ratio of 0.012g/g dry air. The air
pressure is 98 kPa. Determine the temperature to which the air must be cooled
until its relative humidity becomes 95 percent. Determine also the initial
relative humidity.
Solution. From Eq. 10-12,

0.012 = 0.622
98 -pv
pv = 1.855 kPa
Since
P actual
Psaturated
P saturated
230 Nonreactive Gaseous and Vapor Mixtures

From the steam tables,

t = 17.12°C Answer

<t> = Paclual , initial 6 =


Psaturated 3. /oZ

= 49.05 percent Answer

In a heating problem or a cooling problem, it is generally desired to


ascertain the amount of heat that is transferred. It is desirable to consider a
pound of dry air and to evaluate the enthalpy of the dry air and its water
vapor at various parts of the system. Thus, at any point in the system, the
enthalpy per pound of dry air is

hm = 1.0041 -f a)(2500 + 1.88t) (J/g) (10-17)

where hm is the enthalpy of a pound of dry air plus the enthalpy of the vapor
accompanying it.
In a problem involving simple heating or cooling, the difference between the
enthalpies at the two points under consideration is the heat added or removed
per pound of dry air.

Example 10-9. How much heat must be added per minute for Example 10-7
if the initial air flow is 31 m '/min?

Solution. The humidity ratio of the air

" = 0 622 (t0L3252-4022342) = 0 001441 S/S air


Final enthalpy per gram of the dry air is

hrrh = 1.004 x 20 + 0.001441(2500+ 1.88 x 20)


= 23.74 J/g

Initial enthalpy per gram of the dry air is

hm. = 1.004(-10) + 0.001441[2500 + 1.88(-10)]

= 6.465 J/g

Heat added per gram of dry air equals

23.74-6.465 = 17.275 J/g

Mass flow of air per minute is

. _pV _ (101.325 -0.2342) x 103 x 31


m RT 0.2870(273.15- 10)
= 0.415 x 105 g/min

heat added = 0.415 x 105 x 17.275 = 7.17 x 105 J/min Answer


Humidification and Dehumidification 231

10-10 HUMIDIFICATION AND DEHUMIDIFICATION


The term air conditioning refers to the process of preparing air to bring it to
the conditions demanded by its use. This process may involve a change in the
temperature and the humidity of the air. It may also be necessary to remove
objectionable impurities from the air. In certain instances it is necessary to
maintain the velocity of the air at a predetermined value. Although air
cleanliness and air velocities may be of much importance in certain in¬
stallations, they are not thermodynamic problems. Hence, their consideration
is outside the scope of this text. Section 10-9 dealt with adjustment of the
temperature of the air. In this section, the thermodynamic aspects of adjus¬
ting the humidity of the air will be discussed.
The humidity of air may be decreased by physical-chemical means, as by
the use of silica gel, calcium chloride, or activated alumina. Although such
means are used in special applications, in most air-conditioning work dehu¬
midification is accomplished by cooling the air sufficiently below its dew point
to cause the desired amount of water vapor to condense. Since cold humid air
is not desired for most applications, the air must be heated after its dehu¬
midification.
Two methods are commonly employed to cool the air. Chilled water at a
sufficiently low temperature may be sprayed into the air. When a fine spray is
used, a large area is available for heat transfer between the air and the water.
The air in contact with the water droplets will be saturated as it leaves.
Whether or not all the air comes in contact with the water depends on the
extent and on the nature of the spray pattern. Unless otherwise specified, it
will be assumed that the spray is sufficiently extensive to saturate the air
completely by the time it leaves the spray chamber.
The second method of chilling air is the direct-cooling method. The air
passes directly over a series of coils which constitute the evaporator of a
refrigerating unit. Since there is considerable space between coils when only
one row of coils is used, some air will not come in contact with the coils.
Hence, when finally mixed, the air will not have a relative humidity of 100
percent. Generally, several rows of coils are used. It will be assumed, in the
absence of other information, that enough rows of coils are used to saturate
the air completely before it leaves the dehumidifying zone.
Thermodynamically, the method used for the cooling and dehumidifying
process is unimportant, and only the terminal states need to be considered.
There are two factors of particular concern in a dehumidification process.
First, to what temperature must the air be cooled to achieve the desired
dehumidification? Second, how much heat must be removed from the air?
If it is assumed that the air is saturated when leaving the dehumidifier, the
partial pressure of the vapor must be such that the air, when heated to the
desired temperature, will have the desired relative humidity. Since the total
pressure remains constant during heating, the vapor pressure must also
remain constant. Hence, the air must be cooled in the dehumidifier to the
232 Nonreactive Gaseous and Vapor Mixtures

temperature that corresponds to the desired vapor pressure of the air after
it is heated.

Example 10-10. Conditioned air is to be supplied at 26° C, 45 percent relative


humidity. Determine the temperature to which the air must be cooled in the
dehumidifier if it leaves saturated.

Solution. Using Eq. 10-10,

actual vapor pressure at 26°C = 0.45 x 3.363


= 1.513 kPa

From the Steam Tables in the Appendix, the saturation temperature cor¬
responding to 1.513 kPa is 13.2°C. Answer
Example 10-11. An air conditioner is to supply 125 m3 of air per minute at
26°C, 45 percent relative humidity. Air is supplied to the dehumidifier part of
the unit at 34°C, 65 percent relative humidity. Determine the amount of
refrigeration required for the dehumidifier.

Solution. Refer to Figure 10-9. Using Eq. 10-10,

actual vapor pressure at entrance = 0.65 x 5.329


= 3.46 kPa
From Eq. 10-12,

= 0 622 101.325-3.46 = 0 022 8/8 dry a'r


From Eq. 10-14 and Eq. 10-17a, at the entrance

hni] = 1.004 x 34 4- 0.022(2500 + 1.88 x 34)


= 90.54 J/g dry air

It may be assumed that the air leaving the dehumidifier is saturated. Then
from Example 10-10, its temperature is 13.2°C.
The humidity ratio,

W2 = 0622 ioob5-i'.5i3 = 00094 8/8 dry air


hm2= 1.004 x 13.2 + 0.0094(2500 + 1.88 x 13.2)
= 36.98 J/g dry air

Assume that the condensate is removed from the dehumidifier at the rate it
is formed and at the temperature of the air leaving the dehumidifier. The
amount of condensate equals the difference between the humidity ratios at
entrance and exit. Thus,

condensate = 0.022 - 0.0094 = 0.0126 g/g dry air

For the range of temperatures involved here, the specific heat of the water
Humidification and Dehumidification 233

may be taken as 1 cal/°C or 4.1868 J/°C. Then the enthalpy of the condensate
equals

0.0126 x 4.1868 x 13.2 = 0.70 J/g dry air

Writing an energy balance on the dehumidifier, enthalpy of air in = enthalpy


of air out + enthalpy of condensate + heat removed.

heat removed = 90.54 - 36.98 - 0.70 = 52.86 J/g dry air


The mass rate of flow of dry air is
pV _ (101.325 - 1.513)103x 125
1.453 x 10s g/min
RT 0.287 x (273.15+ 26)

The amount of refrigeration required per minute equals

52.86 x 1.453 x 105 = 76.8 x 105 J/min or 7680 kJ/min Answer

It should be noted in Example 10-11 that if the enthalpy of the condensate


is neglected there will be an error of a little over 1 percent in the amount of
refrigeration required. The error is of this magnitude in most dehumidifier
problems. Since this error is within the range of measurements made in most
air-conditioning problems, the enthalpy of the condensate is generally
neglected.
As is shown in Figure 10-9, air is cooled by means of refrigeration to
13.2°C, and then is heated by separate means to 26°C. Thermodynamically,
the removal of heat by one means from one part of the system and the
addition of heat to the system from an external source is undesirable. Such a
procedure may be avoided if the chilled air leaving the dehumidifier is passed
back through a heat-transfer coil located inside the dehumidifier. If sufficient
heat-transfer surface is provided, the chilled air may be heated to 26°C by the
incoming warm air. Desirable as this may be thermodynamically, such a
procedure is economically unsound for this particular application of
regenerative heating. The overall coefficient of heat transfer from air, through
a metal partition, to air is very low. Furthermore, the temperature difference
between the warm and chilled airs is very small. The result is that a very large

13.2° C
34° C 100 percent relative humidity 26° C
65 percent relative humidity 45 percent relative humidity

FIGURE 10-9 Air-conditioning unit for Example 10-11.


234 Nonreactive Gaseous and Vapor Mixtures

heat-transfer surface is required. The initial cost of a large heat-transfer


surface does not justify the potential gain to be made by regenerative heating
of the chilled air.
In Example 10-11, the total air supply was warm and humid. Hence, the
amount of refrigeration required is very high. To minimize the amount of
refrigeration required, it is common practice to draw the major portion of air
supplied to the dehumidifier from the conditioned space. This air normally has
a much lower absolute humidity and a lower temperature than the outside air.
Thus, there is a sizable reduction in the amount of refrigeration required for
the dehumidifier.

10-11 COOLING BY EVAPORATION


The atmosphere generally is the ultimate heat sink. When the working fluid of
a heat engine is rejected to the atmosphere, there is no real problem in
transferring heat from the working fluid to the atmosphere. Examples of this
type of heat engine are the internal combustion engine and the gas turbine. On
the other hand, a very large percentage of the electric power generated in this
country is generated in steam turbine plants. Since it is neither desirable or
feasible to discharge the exhaust steam into the atmosphere, it must be
condensed. Because of its superior heat-transfer characteristics, water rather
than air is used to remove heat in the condenser. The amount of water
required for the condensers of a large power plant is tremendous—perhaps 30
to 40m3/s. The problem of disposal of the heated water is a very real one
now, since governmental regulations have been established to prevent thermal
pollution. As a result, cooling towers are used extensively to cool the water
from the condensers and will be used much"more extensively in the future.
Cooling towers are also used widely to cool water for stationary diesel
engines, refrigeration, and many industrial processes.
In the most common type of cooling tower, the wet type, the warm water
descends through a rising mass of air (see Fig. 10-10).3 The water is broken up
into a spray to provide a very large area for heat transfer to the air. The
cooling of the water is accomplished largely by evaporation although when
the water temperature is much above that of the air, the cooler air will absorb
some heat from the water. Theoretically the water can be cooled to the
wet-bulb temperature of the entering air. Actually this temperature is not
attained, since a finite temperature difference is required for heat transfer.
The term cooling efficiency is used to designate the performance of a cooling
tower. Cooling efficiency is the actual drop in temperature of the water

3The dry-type cooling tower is used in large installations only when there is a shortage of the
makeup water required for a wet-type cooling tower. The dry type is in reality a heat exchanger
in which the water is cooled by the air. The cooled water temperature must exceed that of the
dry-bulb temperature of the air. Hence, for power plant operation, the higher temperature water
from a dry tower produces a higher condensing steam temperature and a higher steam pressure in
the condenser, thus decreasing the efficiency of the power plant.
Cooling by Evaporation 235

Warm, humid air

divided by the theoretical drop in temperature. Thus,

cooling efficiency = f1—£7 (10-18)


M *2

where = temperature of the entering water


t2 = temperature of the leaving water
t'2 = wet-bulb temperature of the entering air

As the air passes upward in the tower, it comes in contact with warmer
water. Hence, it is heated as it picks up vapor. This heating causes an increase
in the wet-bulb temperature of the air. Where there is intimate contact
between the water and the air, it may be assumed that the air is saturated as it
leaves the cooling tower. The cooling efficiency, as well as the temperature of
the air leaving the tower, is determined by the design of the tower and by its
operating characteristics. For fixed conditions, the amount of air that is to be
forced through the tower must be determined. The rate of evaporation should
also be known, in order that the proper amount of makeup water can be
furnished. The method of determining these quantities will be illustrated in
the following example.

Example 10-12. 375,000 kg of water per minute are supplied to a cooling


tower at 46°C, standard atmospheric pressure (see Fig. 10-11). Air enters at
35°C dry bulb and 25°C wet bulb, and leaves saturated at 38°C. Assume a
cooling efficiency of 80 percent. Determine the volume rate of flow of dry air
to be supplied per minute and the amount of makeup water required to
replace the amount evaporated.
236 Nonreactive Gaseous and Vapor Mixtures

Warm, humid air

35° C dry bulb


r2' — 25° C wet bulb

FIGURE 10-11 Cooling tower for Example 10-12.

Solution. From Eq. 10-18,

t2 = 46- 0.8(46 - 25) = 29.2°C

For the leaving air,

saturation pressure = 6.632 kPa (at 38° C)

and at exit

6.632
co = 0.622 (■ 0.04356 ^2E2r
101.325-6.632 ) §dry air

For the entering air,

pwb = 3.169 kPa (at 25°C)

At the wet-bulb temperature of 25° C,

a, = 0.622 = 0.0201
V101.325 - 3.169/ gdry
hdawb= 1.004x25 = 25.1 J/g
hda = 1.004x35 = 35.1 J/g
hVwb = 2500 + 1.88 x 25 = 2547 J/g
hu = 4.186x25 = 105 J/g
!iv = 2500 + 1.88 x 35 = 2566 J/g
Cooling by Evaporation 237

From Eq. 10-14, the humidity ratio of the incoming air,

25.1 -35.1 +0.0201(2547- 105) A nicn gvapor


O) =--777c-= 0.0159-iL-
2566 105 Sdryair
exit air H = ma[1.004 x 38 + 0.04356(2500 + 1.88 x 35)] = 150.26ma (J)
water at exit, H = [375,000 x 1000 - (0.04356 - 0.0159)ma][1.88 x 29.2]
= 2.059 x 1010 — 1.518ma (J)
air at entrance, H = ma[ 1.004 x 35 + 0.0159(2500 + 1.88 x 35)] = 75.94ma (J)
water at entrance, H = 375,000 x 1000(1.88 x 46) = 3.243 x 1010 J

Equating the energy out to the energy in,

150.26ma - 1.518mfl + 2.059 x 1010 = 75.94ma + 3.243 x 1010


1.184
ma 162.6 x 107 g dry air/min
72.8
mRT _ 162.6 x 107 x 0.287(35 + 273.15)
p 101.325 x 103
= 1.419 x 106 m3/min Answer
makeup water = ma(o)out - o>in)
= 162.6 x 107(0.04356 - 0.0159) = 4.498 x 107 g/min
= 4.498 x 104 kg/min Answer
238 Nonreactive Gaseous and Vapor Mixtures

PROBLEMS _

10-1 A 0.725-m1 tank contains an unknown amount of nitrogen. The amount


of 82.5 kg of carbon dioxide is introduced into the tank. The pressure
becomes 92.6 bars and the temperature is 42°C. Determine the amount
of nitrogen present.
10-2 A 0.825-m3 tank contains 0.95 kg of oxygen, 0.82 kg of nitrogen, 1.32 kg
of carbon dioxide and 0.091 kg of carbon monoxide at a temperature
of 20°C. Determine the total pressure.
10-3 (a) The carbon dioxide in Problem 10-1 is supplied at 82.8 bars, 42°C.
Determine the change in entropy as a result of the mixing in the tank,
(b) If the temperature of the surroundings for part (a) is 30°C,
determine the loss in available energy.
10-4 Flue gases are 81 percent nitrogen, 14 percent carbon dioxide, 4.2
percent oxygen, and 0.8 percent carbon monoxide, on the volumetric
basis. Determine the density of this flue gas at 1.012 bars, 47°C.
10-5 Air at 4.2 bars, 80°C is introduced at the rate of 7.8 kg/min into a
flowing air stream having a pressure of 1.022 bars at 34°C. The pressure
downstream from the point of mixing is 1.018 bars and the tem¬
perature is 42°C. Determine the mass rate of flow of the 34°C air.
10-6 Determine the net change in entropy for Problem 10-5.
10-7 Nitrogen (8.62 kg) is mixed with 7.85 kg of carbon dioxide. Determine
the volumetric analysis of the mixture and its gas constant.
10-8 Air at 35° C has a dew-point temperature of 24° C. Determine its
relative humidity and the humidity ratio. The barometric pressure is
101.3 kPa.
10-9 (a) Same as Problem 10-8 except at an altitude of 1500 m where the
pressure is 84 kPa.
(b) Draw conclusions as to the effect of neglecting the influence of
altitude on the relative humidity and the humidity ratio.
10-10 (a) Determine the absolute humidity of air at 40°C having a relative
humidity of 81.3 percent, assuming the vapor behaves as an ideal gas.
(b) Same as part (a), but take the specific volume from superheated
Steam Tables in the Appendix.
10-11 (a) Same as Problem 10-10 except at 80°C 84.4 percent relative
humidity.
(b) Draw conclusions from Problems 10-9 and 10-10(a) relative to the
desirability of treating water vapor as an ideal gas in humidity prob¬
lems.
10-12 Determine the relative humidity for Example 10-6 in the text.
10-13 (a) Check the result of Problem 10-12 with that obtained using a
psychrometric chart.
Problems 239

(b) Check the humidity ratio found in Example 10-6 in the text by
using a psychrometric chart.
10-14 Air leaves a dehumidifier saturated at 1.012 bars, 14°C. It is heated to
20°C before entering a room. Determine the relative humidity entering
the room.
10-15 (a) The air in Problem 10-14 leaves the room at 26°C. The heat input
to the room is 20,000 kJ/h. Determine the volume rate of air flow into
the room.
(b) Determine the error that will be caused if the heat picked up by
the water vapor is neglected.
10-16 (a) Air leaves a warm air furnace at 38° C, 40 percent relative humi¬
dity. The air is returned to the furnace at 18°C with no gain in
moisture. The furnace is to supply 42,500 kJ of heat per hour. Deter¬
mine the volume rate of air flow. The air pressure is 101.8 kPa.
(b) Determine the error that will be caused if the heat given up by the
water vapor is neglected.
10-17 Air enters a dehumidifier at 34°C, 101 kPa, and 60 percent relative
humidity. After being dehumidified, the air is heated to 20°C, with the
relative humidity being 48 percent. Determine the temperature leaving
the dehumidifier if the air is saturated.
10-18 Determine the heat removed per minute from the dehumidifier in
Problem 10-17 if the rate of air flow is 128 m3/min at 20°C, 48 percent
relative humidity. Neglect the enthalpy of the condensate.
10-19 For Problems 10-17 and 10-18, determine the amount of heat required
to heat the air leaving the dehumidifier.
10-20 Repeat Problem 10-18 taking the enthalpy of the condensate into
account. Assume that the condensate leaves at the temperature of the
air.
10-21 Outside air at 0°C, 85 percent relative humidity is heated to 25° C.
Determine the final relative humidity.

10-22 The outside air of Problem 10-21 is at a pressure of 101.2 kPa. It is


supplied at the rate of 72.5 m3/min. Determine the rate at which heat is
supplied.
10-23 To minimize the heat that must be removed from the dehumidifier in
Problem 10-18, 60 percent of the mass rate of air flow to the dehu¬
midifier is taken from the air conditioned room at 26° C, 52 percent
relative humidity. Determine the savings in the amount of heat that
must be removed from the dehumidifier.
10-24 (a) Using Equation 10-15, calculate the enthalpy of water vapor at
50°C. Compare this to the steam table value at 50°C, lOkPa pressure,
(b) Same as part (a), but for water vapor at 100°C, 50kPa pressure.
240 Nonreactive Gaseous and Vapor Mixtures

10-25 (a) The vapor pressure in air at a pressure of 100.5 kPa is increased
from 1.2 kPa to 2.4 kPa. Determine the percentage change in the
humidity ratio.
(b) Same as for part (a), except the vapor pressure is increased from
13.5 kPa to 27 kPa.
(c) Same as for part (a), except the air pressure is 61 kPa.
(d) Comment on the relationship between vapor pressure changes and
changes in the humidity ratio.
10-26 Air enters a heater at 101.5 kPa, 85 percent relative humidity, 0°C, at
the rate of 175 nr/m. Heat is added in the heater at a rate of
11,800 kJ/min. Determine the temperature of the air at exit.
10-27 Air enters a cooling tower at 29°C dry-bulb temperature, 24°C wet-
bulb temperature at the rate of 5100 nr/m. It leaves saturated at 38°C.
Water enters the tower at 46°C at the rate of 5950 kg/min. The
barometric pressure is 101.2 kPa. Determine the relative humidity
from the psychrometric chart in Figure A-7 of the Appendix. Deter¬
mine the cooling efficiency.
10-28 Air enters a cooling tower at 32°C dry bulb, 24°C wet bulb at the rate
of 3400m3/min, and leaves saturated at 39°C. Barometric pressure is
101.2 kPa. Water enters the tower at 45°C. Assume a cooling efficiency
of 78 percent. Use the psychrometric charts in Figure A-7 in the
Appendix for the relative humidity. Determine: (a) the temperature of
the water leaving the tower, (b) amount of cool water delivered, and
(c) minimum amount of makeup water required.
10-29 A cooling tower receives 33,900 kg of water per second at 43°C. Air
enters the tower at 34°C dry bulb, 26°C wet bulb. Barometric pressure
is 101.4 kPa. The cooling efficiency is 80 percent. Air leaves the tower
saturated at 41°C. Determine: (a) the temperature of the water leaving
the tower, (b) volume rate of air flow to the tower, and (c) minimum
amount of makeup water required.
10-30 In Problem 10-18, the condensate leaves the dehumidifier at the same
temperature as that of the air. Determine the percentage of error
caused by neglecting the enthalpy of the condensate.
10-31 Air is supplied at 101.2 kPa, 12°C and 70 percent relative humidity. It
is to be heated to 38° C and a relative humidity of 60 percent.
Determine per kilogram of dry air: (a) amount of water to be added,
(b) heat added to the water if it is supplied at 18°C, and (c) total heat
added.
_ 11 _

ELEMENTARY CHEMICAL
THERMODYNAMICS

11-1 INTRODUCTION
In Chapter 10 a study was made of mixtures of ideal gases with no chemical
reactions taking place. In many engineering situations, reactions between
gases do take place. Such reactions produce significant changes in the
properties of the system and also may involve large quantities of energy.
Some examples of common systems in which such reactions occur are the
combustor of a gas turbine, the furnace of a steam generator, the combustion
chamber of an internal combustion engine, and a fuel cell.
Chemical reactions may take place between a very large number of sub¬
stances existing in various phases. The complete analysis of all possible types
of chemical reactions is very involved and completely beyond the scope of
this book. Since the most common type of a chemical reaction encountered
by many engineers is that of combustion, the discussion in this chapter will be
centered on that particular type of chemical reaction. Many of the principles
that will be developed for combustion problems are applicable to chemical
reactions in general.

11-2 COMBUSTION
Combustion is defined as a chemical reaction in which a fuel combines with
oxygen liberating large quantities of heat and, incidentally, light. A fuel is thus
defined as a substance that combines with oxygen in a combustion process.
(The term “fuel” is also used to designate those fissionable substances that
release energy in a nuclear reactor.) Some fuels, particularly some rocket
fuels, are quite complex. The discussion of the combustion of these fuels is
beyond the scope of this text. Rather, the discussion here will be confined to
those fuels whose burnable constituents are carbon and hydrogen and, to a
minor extent, sulfur. These fuels supply the major portion of our energy
requirements.
These fuels may exist in the solid, the liquid, or the gaseous phase.
Normally a liquid fuel vaporizes before it can be burned and thus is burned in
the gaseous state. Many solid fuels, such as coal, are partly vaporized when
heated. However, there is some carbon in the coal that is not vaporized during

241
242 Elementary Chemical Thermodynamics

heating, and, hence, must be burned as a solid. The detailed mechanism of


the combustion process of a complex fuel, such as petroleum products, is not
fully understood. However, the end products of a combustion process can be
predicted without a full knowledge of the intermediate steps. It will be the
overall process that will be discussed here.
Consider a mixture of a gaseous fuel and air. For most fuels, there is no
measurable amount of reaction between the fuel and the oxygen in the air at
room temperature. The temperature of the mixture may be increased slowly
for several hundred degrees without a perceptible reaction occurring. Finally,
a temperature is reached where the fuel ignites. Without further heating from
the outside, there is a very rapid rise in the temperature of the mixture. The
ignition temperature of a fuel-air mixture may be defined as the temperature
at which there is sufficient reaction between the fuel and oxygen to produce a
temperature rise. This ignition temperature is dependent on the fuel-air ratio,
the amount of nitrogen present, and the pressure of the mixture. Hence, the
ignition temperature of a mixture of air and a given fuel depends on the
amount of air present, the pressure of the mixture, and the period of time that
the mixture is held at the given temperature.
The term complete combustion is used to describe a combustion process
that proceeds until none of the material remaining is combustible. Complete
combustion will take place only when the three following conditions are
met:

1 The temperature of the fuel-air mixture must be at least equal to the


ignition temperature.
2 There must be sufficient air present to provide the oxygen necessary for
the combustion reaction to proceed to conclusion. To ensure sufficient
oxygen for actual combustion processes, excess air must be used (see
Section 11-3 for the definition of excess air).
3 The air must be in intimate contact with the fuel.

Unless otherwise stated, it will be assumed that the preceding conditions


exist and that combustion is complete. Since, in the majority of combustion
processes, the required oxygen is supplied from air, one of the factors of
interest to an engineer is the amount of air that must be supplied to burn a
given amount of fuel. Another factor of interest is the composition of the
products of combustion. Means of determining these factors will now be
developed.
Regardless of the mechanism of combustion and the intermediate steps
before the final products are formed, it is possible to write combustion
equations that involve only the initial and final substances. Combustion takes
place between the atoms of molecules of a fuel and of oxygen, and results in
the formation of molecules of other substances called products. Hence, the
fundamental combustion equation is written on the molecular basis. This basis
may be illustrated for the combustion of carbon and also of hydrogen.
Combustion 243

For carbon, the equation is

C + 02-+C02 (11-1)
As indicated by Eq. 11-1, one molecule of carbon combines with one
molecule of oxygen to form one molecule of carbon dioxide. Since an actual
combustion process involves an extremely large number of molecules, it is
preferable to write a combustion equation on the molal basis. Because the
number of molecules per mole is the same for all substances, Eq. 11-1 may be
read as follows:

1 mole C + 1 mole 02-~► 1 mole C02 (11-la)


When it is desirable to consider the mass of the various substances
involved, the mass per mole may be introduced into Eq. 11-la. The result is

12 g C + 32 g 02 44 g C02
or

1 g C + 2i g 02 -» 3i g C02 (11-2)
An examination of Eqs. 11-la and 11-2 shows that there may be a change in
the number of moles during a combustion process but that there can be no
change in the total mass present. (This statement disregards the infinitesimal
amount of mass transformed into energy during combustion.)
When the oxygen for combustion is supplied from air, nitrogen will be
present during the combustion process. The amount of nitrogen can be
determined by the oxygen-nitrogen ratio of air. Since air is 21 percent oxygen
on the volume or molal basis, the amount of air required per mole of oxygen
equals 1/0.21 = 4.76 moles. The number of moles of atmospheric nitrogen per
mole of oxygen is 4.76- 1.00 or 1 x (79/21) = 3.76 moles. Nitrogen is some¬
times introduced in the combustion equation to indicate its presence, even
though it does not react. Thus, the equation for carbon may be written as
follows:
C + 02 + 3.76 N2^ C02 + 3.76 N2 (11-3)

The nitrogen in a combustion equation refers to atmospheric nitrogen.


As previously stated, 4.76 moles of air are required per mole of carbon.
Since there are 28.97 g of air per mole of air and there are 12 g of carbon per
mole of carbon, the number of grams of air required per gram of carbon is

4.76 x 28.97
11.49 g
12

Or, since the mass of air equals the mass of the oxygen plus that of the
nitrogen, the number of grams of air required per gram of carbon is

1 x 32 4-3.76x28.16
11.49 g
12
244 Elementary Chemical Thermodynamics

This amount of air is known as the theoretical air. It is the amount of air re¬
quired for complete combustion, with no oxygen remaining after the combustion
process. A mixture of the theoretical air and fuel is sometimes known as a
stoichiometric mixture. The equations involving the theoretical amounts of
oxygen or air, such as Eqs. 11-1, 11-la, 11-2, and 11-3, are said to be
stoichiometric equations.
The term products of combustion refers to all substances present after
combustion, whether or not they entered into the combustion process. Thus,
the products of the complete combustion of carbon in air are carbon dioxide
and nitrogen. For each mole of carbon there are 3.76 moles of nitrogen and
1 mole of carbon dioxide. The volumetric analysis of the products may be
found as follows:

Moles Molal or Volumetric Percentage


C02 1.00 -r- 4.76 = 0.21 21
N2 3.76-4.76 = 0.79 79
4.76 1.00

As a check on the accuracy of the preceding analysis, it should be recalled


that air is 21 percent oxygen and 79 percent atmospheric nitrogen on the
volume basis. (On the weight basis, air is 23.2 percent oxygen and 76.8
percent atmospheric nitrogen.) When carbon is burned completely, the
oxygen in the air is replaced by an equal amount of carbon dioxide. The
nitrogen percentage in the products must remain at 79 percent.
The fundamental combustion equation for that other important fuel ele¬
ment, hydrogen, is

2 H2 + 02-+ 2 H20

This equation may be stated as follows:

2 moles H2T 1 mole 02-*2 moles H20 (11-4)

Since quantities are desired per mole of fuel, Eq. 11-4 can be divided by 2.
The result is

1 mole H2 + \ mole 02^ 1 mole H20 (ll-4a)

On the gram basis,

lgH2 + 8g02-»9g H20 (1 l-4b)

The air required for the combustion of hydrogen is 0.5/0.21 = 2.38 moles of
air per mole of hydrogen. The number of grams of air per gram of hydrogen is

2.38x28.97
= 34.5 g/g
2
Combustion of Actual Fuels 245

11-3 COMBUSTION OF ACTUAL FUELS


Three cases will be considered: with theoretical air, with excess air, and with a
deficiency of air.

a. Theoretical Air
Theoretical air has been defined as that amount of air which contains the
amount of oxygen required for complete combustion of the fuel, with no
oxygen remaining after the combustion process. A fuel in the gaseous form
may be a mixture of combustible compounds with some inert gases being
present. Generally, it is desirable to write a combustion equation for each
combustible compound to find both the theoretical amount of air required and
the theoretical products of combustion. A summation of these quantities will
give results for the active gaseous mixture.
In writing the combustion equation for compounds, particularly complex
ones, it is desirable to start with 1 mole of the fuel, since results are normally
expressed initially per mole of fuel. Assuming that the fuel is a hydrocarbon,
the amount of carbon and hydrogen in the fuel determines the amount of
carbon dioxide and water vapor formed as well as the amount of oxygen
required. Consider the combustion of methane, CH4.

1 CH4 + a02^ bC02 + cH2G

Making a carbon balance, there must be 1 mole of C02 or b = 1. A hydrogen


balance shows that there must be 2 moles of water vapor or c = 2. Now
making an oxygen balance, since there are 4 of O or 2 of 02 on the right side,
a must equal 2. Then
CH4 + 2 02 C02 + 2 H20 (11-5)

Consider the gas ethane (C2H6). The combustion equation is

1 C2H6 4- u02—> bC02 + cH20

Balancing the carbon, b = 2. Balancing the hydrogen, c = 3. From an


oxygen balance a = 3|. Then
C2H6 + 32 02 -<■ 2 C02 + 3 H20 (11-6)

Example 11-1. The analysis of a natural gas follows: CH4 (methane), 88.1
percent; C2H6 (ethane), 11.5 percent; N2, 0.4 percent. Determine: (a) the
theoretical air and (b) the analysis of the dry products of combustion. Dry
products of combustion include all products of combustion except water
vapor. This term is widely used, since the amount of dry gaseous products is
independent of the amount of condensation of water vapor. The amount of
wet gaseous products cannot be specified until the amount of condensation is
known.

Solution. (a) A mole of the fuel contains 0.881 mole of methane and
0.115 mole of ethane. From Eq. 11-5, the oxygen required to burn the methane
246 Elementary Chemical Thermodynamics

31x0.115 = 0.403 mole


The total oxygen required equals 1.762 + 0.403 = 2.165 moles. The amount of
air required equals 2.165/0.21 = 10.31 moles of air per mole of fuel, or 10.31 m3
of air per cubic meter of fuel. Answer
(b) The dry products consist of carbon dioxide and nitrogen. The number
of moles of carbon dioxide per mole of fuel is found as follows:

from the CH4, 1 x 0.881 = 0.881 mole


from the C2H6, 2x0.115 = 0.230 mole
Total = 1.111 moles

The amount of nitrogen per mole of fuel equals 2.165 x (0.79/0.21) =


8.145 moles. Answer

The major constituent of any solid fuel, such as coal, is carbon. Other
constituents are volatile combustibles and inert matter. (Volatile com¬
bustibles are combustible substances, mostly hydrocarbons, which become
gaseous upon application of heat.) The analysis of a coal is made on the mass
basis, and shows the amounts of carbon, hydrogen, sulfur, nitrogen, and ash.
Since the analysis is on the mass basis and since the theoretical air is desired
on the mass basis, combustion calculations generally are made on the mass
basis.

Example 11-2. An analysis of a coal follows: C, 74 percent; H2, 6 percent;


02, 8 percent; S, 1 percent; N2, 1.6 percent; ash, 9.4 percent. Determine the
theoretical air-fuel ratio.

Solution. The combustion for sulfur has not been given. It is


s + o2 so2
Since both oxygen and sulfur have a molecular weight of 32,

lgS + lg02-»2g S02

The total oxygen required for 1 g of coal (see the combustion equations of
carbon, hydrogen, and sulfur) is found as follows:

Grams 02 per
Substance
Gram Coal
C 0.74 x 2\ 1.973
H 0.06 x 8 0.480
S 0.01 x 1 0.010
Total 02 required 2.463
Oxygen in coal 0.080
Oxygen from air 2.383

theoretical air = 2.383/0.232 = 10.27 g air/g coal Answer


Combustion of Actual Fuels 247

A liquid fuel may have its analysis given on the same basis as for a solid
fuel. In this case, combustion calculations may be made in a manner similar to
that shown in Example 11-2. Frequently a particular chemical compound is
chosen whose air requirements and whose products are substantially equal to
those of a liquid fuel. For instance, octane (C8H18) is used to make combustion
calculations for gasoline, and dodecane (Ci2H26) is used in place of fuel oil.
For such cases, combustion equations are written in a manner similar to that
shown for gaseous fuels.

Example 11-3. Treating gasoline as C8Hi8, determine (a) the theoretical air
required and (b) the volumetric analysis of the dry products (theoretical).

Solution, (a) The combustion equation is

C8H,8+ 125 02-+8C02 + 9H20

The number of moles of air per mole of fuel is 12.5/0.21 = 59.5 moles, and the
number of grams of air per mole of fuel is 59.5 x 28.97 = 1725 g.
The theoretical air-fuel ratio is

8x\7r+T8 = 15-12gair/e fuel


or

12.5x28.97
15.12 g air/g fuel Answer
0.21 x (8 x 12+18)

(b) The moles of dry products per mole of fuel are found as follows:

C02, 8 h- 55 = 0.1455
N2, 12.5 x (79/21) = 47- 55 = 0.8545
55

The molal or volumetric analysis is 14.55 percent C02 and 85.45 percent N2.
Answer
Note that the amount of nitrogen in the products of combustion in
Example 11-3 exceeds the amount of nitrogen in the air, because some of the
oxygen in the air combines with hydrogen to form water vapor. Since the
water vapor does not appear in the dry products, the nitrogen becomes a
larger percentage of the total.

b. Excess Air
Because it is not possible to obtain perfect mixing of air and fuel, combustion
will be far from complete unless considerable excess air is supplied. Excess
air is defined as that air supplied which is in excess of the theoretical amount.
Percentage of excess air is the excess air expressed as a percentage of the
theoretical. The percentage of excess air that is normally supplied depends on
the type of fuel (which determines the ease of mixing it with air), the type of
combustion equipment, and the type of controls. For example, with good
248 Elementary Chemical Thermodynamics

controls and a gaseous fuel, the required percentage of excess air may be only
10 to 15 percent. With hand-firing of coal, the percentage of excess air may
range from 20 percent to 100 or more percent. The excess air in a com¬
pression ignition (diesel) engine may vary from 200 or 300 percent at light load
to roughly 30 percent at full load. To obtain a high-power output in an auto¬
motive type of engine, a deficiency of air is frequently used, even though this
amount results in incomplete combustion and inefficient burning of the fuel.
Theoretically, when excess air is supplied for a fuel, combustion should be
complete. However, since mixing of the air and fuel is never perfect, there
will be at least traces of incomplete combustion in the actual products of
combustion, even with much excess air. Because the degree of mixing cannot
be predicted, the only way of predicting the amount of incomplete combustion
for an actual case is by comparison with similar actual combustion processes.
When considerable excess air is supplied, it is desirable to predict the
analysis of the products of combustion by neglecting the very small amounts
of products of incomplete combustion that may exist.

Example 11-4. A fuel oil is burned with 50 percent excess air. Determine:
(a) the volumetric analysis of the dry products and (b) the volume rate of
flow of the wet products at a pressure of 1.02 x 102kPa and a temperature of
550° C when the fuel is burned at the rate of 45 kg/min. Assume that the
combustion requirements of the fuel oil are similar to those of Ci2H26.

Solution, (a) The theoretical combustion equation is

C12H26+ 18.5 02-» 12 C02 + 13 H20

The excess oxygen per mole of fuel is

0.5(18.5) = 9.25 moles

The total nitrogen per mole of fuel is

(18.5 + 9.25) (“) = 104.4 moles

The analysis of the dry products may be found as follows:

C02, 12 - 125.65 = 0.0955

02, 9.25- 125.65 = 0.0737

XT 104.4 -T-125.65 = 0.8308


N2’ 125.65

The volumetric analysis is: C02, 9.55 percent; 02, 7.37 percent; N2, 83.08
percent. Answer
(b) The total moles of wet products per mole of fuel equals the moles of
dry products plus the moles of water vapor, or 125.65 + 13 = 138.65. The
Combustion of Actual Fuels 249

number of moles of fuel burned per minute equals

45 x 1000
= 264.7 g moles/min
(12 x 12) + 26

The number of moles of wet products equals

264.7 x 138.65 = 36,700 g moles/min

pressure - 1.02 x 102 x 1000 = 1.02 x 105 Pa = 102 kPa


Then,

NR0T _ 36,700 x 8.314(550 + 273)


V= = 2462 m3/min Answer
D 1.02 x 105

c. Deficiency of Air
Regardless of how perfect the mixing may be, there will be incomplete
combustion when there is a deficiency of air. However, because of imperfect
mixing, at least traces of oxygen may be present in the products when there is
a deficiency in the supply of air. By neglecting the oxygen in the products, it
is possible to predict the composition of the products when a single fuel
element is burned with a deficiency of air

Example 11-5. Carbon is burned with 90 percent theoretical air. Assuming


perfect mixing, predict the volumetric analysis of the products.

Solution. For the theoretical case, the equation is

C -F O2—* CO2
For the actual case, the number of moles of 02 per mole of C is 0.9(1) =
0.9 mole, and the number of moles of N2 is 0.9 x (79/21) = 3.38 moles per mole
of C.
Because of a deficiency of air, there will be both CO and C02 in the
products. The CO and the C02 together must contain the mole of carbon in
the fuel that was burned. The CO and the C02 together must also contain the
0.9 mole of available 02. A mole of CO is composed of 1 mole of C and \ mole
of 02. A mole of C02 is composed of 1 mole of C and 1 mole of 02. These
statements can be combined in equation form, as follows:

moles CO 4- moles C02 = moles available C = 1

moles CO
+ moles C02 = moles available 02 = 0.9
2
With these two equations are solved simultaneously, the results are

moles CO _ q ^ or moles CO = 0.2

moles C02 = 1 — 0.2 = 0.8


250 Elementary Chemical Thermodynamics

Products of combustion per mole of C are found as follows:

CO, 0.2 -T- 4.38 = 0.0457

C02, 0.8 -5- 4.38 = 0.1826

N2, 3.38 -j- 4.38 = 0.7717


4.38 1.000
The volumetric analysis follows: CO, 4.57 percent; C02,18.26 percent; N2,77.17
percent. Answer
When both carbon and hydrogen are present in the fuel and there is a
deficiency in the air supplied, the composition of the products of combustion
is difficult to predict. The reason for this difficulty is that the division of the
oxygen between the carbon and the hydrogen is a variable that is dependent
on the composition of the fuel, the deficiency of air, and the pressure and
temperature of the products. The basis for determining the composition of the
products of combustion will be discussed when the subject of chemical
equilibrium is investigated.

11-4 USES OF ANALYSES OF PRODUCTS


When fuel is burned at a high rate, a very large quantity of air must be
supplied. Consider, for instance, a steam-generating unit burning 100 metric
tons1 of coal per hour. Air for combustion must be supplied at the rate of
roughly 15,580 m3/min. It is difficult to determine this air flow accurately. Yet
it is essential that the excess air be kept to a minimum consistent with a
minimum amount of incomplete combustion. Any excess air that is not
required represents a loss in efficiency, since it is heated to the temperature of
the exit gases. A decrease of 1 percent in the efficiency of a unit of this size
may mean an increase of approximately $600,000 a year in the fuel bill.
An analysis of the products of combustion, together with an analysis of the
fuel, provides sufficient information for determining the amount of excess air
supplied, as well as the completeness of combustion. Various means are
available for determining the analysis of the products of combustion. Perhaps
the most common of these is the Orsat. This is a device for determining the
volumetric analysis of the dry products when combustion is reasonably
complete; that is, when carbon monoxide is the only measurable product of
incomplete combustion. The reader is referred to laboratory texts for a
discussion of the Orsat and other devices used to determine the analysis of
the products of combustion. It must be emphasized here, however, that great
care should be taken to obtain a truly representative sample of the products
before any attempt is made to analyze it. The reader is also referred to
laboratory texts for a discussion of precautions necessary to obtain represen¬
tative samples.

'A metric ton = 1000 kg = 2204.62 lb.


Uses of Analyses of Products 251

FIGURE 11-1 Burning of coal.

In the Orsat, the percentages by volume of carbon dioxide, oxygen, and


carbon monoxide in the dry products are determined by chemical absorption.
The remainder of the dry products is assumed to be nitrogen. (This assump¬
tion is in error when there is a deficiency of air, as in an automotive type of
engine. In such a case there may be a measurable amount of methane and also
some free hydrogen.) A relationship between the amount of dry products and
the amount of fuel can be established by means of the amounts of carbon
present in each. For a liquid or gaseous fuel, substantially all of the carbon in
the fuel appears in the dry products. When a solid fuel, such as coal, is
burned, a small amount of the carbon appears in the refuse and the remainder
is found in the dry products. The amount of carbon appearing in the refuse
can be determined by test. Thus for any fuel, the amount of carbon burned
per unit amount of fuel may be determined. Hence, a relationship between the
amount of fuel and the amount of dry products may be established (see Fig.
11-1). The procedure will be illustrated by several examples.
Example 11-6. The coal in Example 11-2 produces the following Orsat
analysis: C02, 12 percent; 02, 6.5 percent; CO, 0.1 percent. Determine the
moles of dry products produced per kilogram of coal burned. The refuse
contains 0.008 kg of carbon per kilogram of coal burned.
Solution. The amount of carbon burned per kilogram of coal = 0.74 - 0.008 =
0.732 kg. Since a mole of CO and a mole of C02 each contain 1 mole of
carbon, the number of moles of carbon per mole of dry products equals
0.12 + 0.001 =0.121 mole. The mass of this carbon equals
0.121 x 12 = 1.452 kg/kg mole
Hence, the amount of dry products is
0.732 + 1.452 = 0.504 kg mole/kg coal Answer
252 Elementary Chemical Thermodynamics

Example 11-7. The fuel oil in Example 11-4 produces the following Orsat
analysis: C02, 12.8 percent; 02, 3.5 percent; CO, 0.2 percent. Determine the
moles of dry products per kilogram of fuel oil.

Solution. If the fuel is treated as Cj2H26, a mole of the fuel contains 12 moles
of C. A mole of the dry products contains 0.128+0.002 or 0.13 mole of C.
Then the number of moles of dry products per mole of fuel is

12 -T- 0.13 = 92.4 moles

Since the mass of mole of fuel is (12 x 12) + 26 = 170 kg, the number of moles
of dry products per kilogram of fuel is

92.4 -T-170 = 0.543 kg mole Answer

In each of the preceding examples, a relationship has been established


between the amount of fuel and the amount of dry products. If a relationship
can be established between the amount of air supplied and the amount of dry
products, then an air-fuel ratio can be established. All of the nitrogen in the
air appears in the dry products. When the fuel contains hydrogen, part of the
oxygen combines with the hydrogen to form water vapor, and this part does
not appear in the dry products. Although a knowledge of the amount of
hydrogen in the fuel will permit the determination of the portion of the
oxygen of the air that appears in the dry products, it is more direct to relate
the amounts of air and dry products by means of the nitrogen contents of
each. The method involved will be illustrated by use of the following exam¬
ples.

Example 11-8. Determine the air-fuel ratio of Example 11-6. Determine also
the percentage of excess air used.

Solution. The number of moles of nitrogen present per mole of dry products
equals 1.000 - (0.12 + 0.065 + 0.001) = 0.814 mole. Since air is 79 percent
nitrogen, on the volumetric basis, the number of moles of air per mole of dry
products is
0.814 -T- 0.79 = 1.03 moles
Hence, the mass of air per kilogram mole of dry products is
1.03 x 28.97 = 29.85 kg
It has been assumed here that all of the nitrogen in the dry products comes
from the air. Since there is nitrogen in the coal, it will contribute to the
nitrogen in the dry products. However, it may be shown that approximately
only 0.1 percent of the nitrogen in the dry products comes from the coal.
Certain gaseous fuels may contain sufficient nitrogen to make it necessary to
consider this nitrogen.
In Example 11-6, it was determined that there was 0.504 mole of dry
products per kilogram of fuel. Hence, the mass of air per kilogram of fuel is

0.504x29.85 = 15.04 kg Answer


Heat of Reaction 253

For Example 11-2, the theoretical air fuel-ratio is 10.27 g of air per gram of
coal. Hence, the percent excess air equals

15.04-10.27
= 46.4 percent Answer
10.27“

In addition to the uses just shown, an analysis of the products of com¬


bustion is necessary when an energy balance is to be made on any fuel¬
burning device. The composition of the products is required to determine the
specific heat of the mixture, in order that the energy carried away by the
gases may be determined. A knowledge of the relative amount of water vapor
in the wet products permits the determination of the temperature where
condensation first starts as the products are cooled. This temperature is
known as the dew-point temperature. (Refer to Chapter 10.)

11-5 COMBUSTION IN GENERAL


Simple combustion calculations have been made in this chapter. It is possible
to combine many of the steps and express them in equation form. It is felt,
however, that if such equations were introduced here, the significance of the
principles involved would be lost.
No attempt has been made to examine the combustion process itself. The
mechanism of combustion, particularly that of complex hydrocarbons, is
complicated and is not fully understood. Combustion does not proceed
directly from the fuel-air mixture to the final products. Rather, partial oxida¬
tion of the fuel occurs at first, and intermediate products are formed. These
products, in turn, are oxidized and new products are formed. Ultimately, if
conditions are right, combustion becomes complete. Involved also in com¬
bustion studies are reaction rates and flame propagation. All of these proces¬
ses must be understood as fully as possible, in order to be able to design and
operate fuel-burning equipment. However, since the study of the very com¬
plicated phases of combustion is beyond the scope of this text, the reader is
referred to the references listed in the Appendix for a discussion of this
important part of thermodynamics.

11-6 HEAT OF REACTION


When a chemical reaction occurs between the components of a system, there
is a change in the forms of energy of the system. Also, unless the system is
isolated from its surroundings, there will be a change in the total energy of the
system. In some chemical processes the temperature of the system increases
during the reaction. There is then, a tendency for heat to be transferred to the
surroundings. This type of process is said to be exothermic. Some other
chemical processes cannot take place unless heat is supplied to the system.
Such a reaction is termed endothermic.
Consider a combustible gaseous mixture existing in a vertical cylinder with
the gas contained by a piston, so loaded that it maintains a constant pressure
254 Elementary Chemical Thermodynamics

on the gas. Allow combustion to take place and then cool the products down
to the original air-fuel temperature. The overall energy equation is

Q + Ur = W + Up (11-7)
where Ur = the internal energy of the reactants
Up = the internal energy of the products

The net work equals p(Vp- Vr). Then Eq. 11-7 becomes

[Q = Up - Ur + p(Vp - Vr)]T,p
or

(Qr = Hp — Hr)Tp (11-8)

The term Qr in Eq. 11-8 is termed the heat of reaction. It is defined as the
difference between the enthalpies of the products of a reaction and of the
reactants at a given temperature and pressure. Normally, heats of reactions
are specified at the standard temperature of 25° C and standard atmospheric
pressure. For these conditions the heat of reaction is termed the standard
heat of reaction.
For most engineering devices involving combustible mixtures, the system is
an open one. Assuming steady-state conditions with no significant changes in
kinetic and potential energies and no work crossing the boundaries of the
system, the energy equation becomes

Q = Hp-Hr

When the products and reactants are at the same temperature and pressure,
the energy equation for steady-state flow becomes

(Qr = Hp — Hr)T p (11-9)

Equations 11-9 and 11-8 are identical. Thus, for a given chemical reaction with
the temperature and pressure of the products equal to those of the reactants,
the heat of reaction is a fixed quantity and is independent of whether the
reaction occurs in an open or a closed system.
When the reaction is an exothermic one, heat is removed from the system.
The enthalpy of the products is less than that of the reactants for a given
temperature. Hence, the heat of reaction is negative. For an endothermic
reaction, the enthalpy of the products exceeds that of the reactants for a
given temperature. Thus, for this condition, the heat of reaction is positive.

11-7 HEATING VALUES


The heating value of a fuel is defined as the heat that must be removed from
the products of a complete combustion in order to cool the products down to
the temperature of the original air-fuel mixture (see Fig. 11-2). Thus, the
heating value of a fuel for a constant-pressure combustion process equals the
Heating Values 255

1^ Q (= heating value
when combustion
is complete)

Air, T,
Products, T2 = Ty
Fuel, T! Combustion device ----5*-
--

FIGURE 11-2 Heating value determination.

negative of the heat of reaction for that process. This means that heating
values are positive quantities.
There are various terms used to describe the heat associated with com¬
bustion processes. In addition to the term heating value, the terms calorific
value and heat of combustion are used also. The term calorific value is
synonymous with heating value. The heat of combustion generally is defined
as equal to the heat of reaction of the complete combustion of the fuel and
oxygen. Since the heat of reaction for a combustion process is negative, the
heat of combustion also is negative.
Thus far it has been assumed that all of the products exist in the gaseous
form. Most fuels contain hydrogen, which upon burning forms water vapor.
As the products are cooled down, some of this water vapor may condense.
The amount of condensation is affected by the amount of excess air used, the
humidity of the air, the original temperature of the air-fuel mixture, and the
total pressure. At one extreme, all of the water vapor of combustion will
condense, giving up its latent heat (i.e., the heat of vaporization). The heating
value thus obtained is designated as the higher heating value. It is also termed
the gross heating value. This heating value for common fuels is given in Table
11-1.
At the other extreme, there is no condensation of the water vapor of
combustion as the products are cooled. This heating value is termed the lower
heating value or net heating value.
In the previous section, it was shown that the heat of reaction for a
steady-state flow process is equal to the constant-pressure heat of reaction. In
a similar manner, it can be shown that the heating value for a steady-state
flow process equals the constant-pressure heating value.
Let the higher heating value at constant pressure be designated as QHP, and
the lower constant pressure heating value as QLP. Now write Eq. 11-8 for a
constant-pressure combustion with all of the water vapor condensing and then
again with no water vapor condensing, in each case, per gram mole of fuel.
Now subtract these two equations. Recognizing that QP is the negative of the
heat of reaction, the result will become

Qhp -Qlp ={H - Hwater)(per mole fuel)


water
vapor
256 Elementary Chemical Thermodynamics

Table 11-1
Heating Values at Constant Pressure and 25°Ca
[All fuels except carbon are in the gaseous state.
The products are C02 (gaseous) and H20 (liquid).]

Heating Value Heat of Vaporization


Fuel Formula (J/g mole) of Fuel, (h/8)(J/g mole)

Carbon (graphite) C 393,778


Carbon monoxide CO 283,123
Hydrogen h2 286,011
Methane ch4 890,909
Benzene c6h6 3,296,036 33,715
Ethyl alcohol c2h6o 1,410,030 42,370
Methyl alcohol ch4o 764,426 37,430
Butane c4h10 2,872,598 21,688
Pentane c5h12 3,516,258 26,475
Octane c8h18 5,508,175 41,240
Dodecane c i2H26 8,137,880 53,498

aBased largely on National Bureau of Standards Publications.

or

Ohp ~ Qlp = (moles H20/mole fuel)(h/g) (11-10)

where hfg is the heat of vaporization of water vapor. At the standard


temperature of 25° C, from the Steam Tables in the Appendix, the value of hfg
of water is 2442.3 J/g or 43,995 J/g mole.

Example 11-9, Determine the lower heating value of gaseous C8H18 at con¬
stant pressure at 25°C.

Solution. The combustion equation is

C8H18 4-12.5 02-* 8 C02 + 9 H20


From Eq. 11-10,

Qlp — 5,508,175 — 9(43,995)


= 5,112,220 J/g mole Answer

Note that the lower heating value is approximately 7 percent lower than the
higher heating value.

Example 11-10. Determine the higher heating value of liquid C8H18 at con¬
stant pressure at 25°C.

Solution. This heating value of the liquid is less than that of the gaseous fuel
Heating Values 257

by the amount of heat required to vaporize the liquid. Then,

Qhp of liquid = QHP of gas - hfg


= 5,508,175 — 41,240 = 5,466,935 J/g mole Answer

In some cases, fuels may be burned at constant volume. The heating values
of solid and the heavier liquid fuels are determined in a bomb calorimeter,
which is a constant-volume device. For a constant-volume combustion, the
energy equation becomes

Qv= Ur- Up
But

Qp = Hr- Hp

Then

Qv — QP +(Ur - Up) — (Hr — Hp)


- QP + (ur— Up) — (u + PV)r - (u + pv)p
= Qp + (Ur - Up) - [(U + NR0T)r ~(U + NR0T)]p
or

Qv = QP + (NR0T)p - (NR0T)r
Then

Qv — QP + (Np — Nr)R()T (11-11)

Under normal conditions the volume occupied by liquids is insignificant in


comparison with the gaseous volume. Hence, both Np and Nr must be the
number of moles in the gaseous state.

Example 11-11. Determine the higher heating value at constant volume of


gaseous C8H18 at 25°C.

Solution. In Eq. 11-11, Np = 8 and Nr = 12.5 for the gaseous state. Then

Qhp = 5,508,175 - (13.5 - 8)(8.314)(298.15)


= 5,497,541 J/g mole Answer

Example 11-12. Determine the lower heating value at constant volume of


C8H18 at 25° C.

Solution. For the lower heating value, the water vapor is gaseous. Then
Np = 17. Using the value of QLP from Example 11-9 together with Eq. 11-11,

Qlv = 5,112,220 + (17- 12.15)(8.314)(298.15)


= 5,120,895 J/g mole Answer

An alternate solution is to write an equation similar to Eq. 11-10 but for a


constant-volume process.
258 Elementary Chemical Thermodynamics

Then

Qhv — Qlv = (moles H20/mole fuel)C7/g (11-12)


Qlv = 5,497,021 - 9(41,525) = 5,120,816 J/g mole Answer

Thus far it has been assumed that the air-fuel mixture exists at the standard
temperature of 25°C. Of course, this is rarely the case. The heating value at
other temperatures can be determined by use of Eq. 11-9, recognizing that the
heating value is the negative of the heat of reaction. Thus, at any temperature

(Qhp = Hr - Hp)T2

and

(Qhp — Hr — Hp)2 5°c


From this,

(Qhp)t2 — (Qhp)25°c + (Hj2 ~ H25°c)r ~ (Ht2 — H25°c)P (11-13)


Equation 11-13 calls for the enthalpies of gases at various temperatures.
These values are given in Table 11-2.

Example 11-13. Determine the higher heating value at constant pressure of


gaseous C8H18 at a temperature of 100°C.

Solution. The reactants consist of 1 mole of C8H8, 12.5 moles of 02 and


12.5(79/21) or 47 moles of N2. The products consist of 8 moles of C02, 9 moles
of liquid water, and 47 moles of N2.
Since the amount of nitrogen in the products equals that in the reactants
and since the temperature of the products equals that of the reactants, the
enthalpies of the nitrogen cancel out.

for the reactants at 100°C, from Table 11-2, Hr = 14,527 -I- 12.5 x 2238 = 42,502

for the reactants at 25° C, Hr = 0

then (Ht, - H25°c)r = 42,502


for the products at 100°C, H2 = 8 x 2974 + 9(2564-40,662) = -319,090

for the products at 25°C, H2 = 0 - 9 x 44,000 = - 396,000

then (Ht2~ H25°c)p — 76,910


Substituting into Eq. 11-13 and taking the value of (QHp)25°c from Table 11-1,

(Qhp) ioo°c = 5,508,175 + 42,502 — 76,910


= 5,473,767 J/g mole Answer

Note that a change in the initial temperature of the air-fuel mixture from 25° C
to 100° C produces a change in the heating value of only about 0.6 of a
percent. Hence, for most engineering applications, the effect of small changes
in the initial temperature on the heating value can be neglected.
Based largely on Table 4-5 in J. S. Doolittle, Thermodynamics for Engineers, 2nd ed., International Textbook Company, Scranton, Pa., 1964.
<D
On ' in 04
£ in On © vo
r.rv
iri \0
r\ e*
co 04 i— ©
^ rt
ao

<N o r- rt oo co
O ON CO OO
S3<N
voo voe\ voo in*v
co oo o
U t-< oi

o o- on co r-
in ^ 04
S300 »n co «n
of vo ©"
u

O in oo on
in On O 04
S3 m, 04 — On
#\ o o
u — 04 04

©rot^-^»or'^©t-^inioin»n©coo4©ONO~©ooo~
(N r)OONho\^hNa\0’HvOfnin^'OM»m^
rf ©»n^©©ONONOOO\^HTfONVO'^’t^'^tVO©»n
S3 -Hrt^mooom^MiMn-HOoinfSON^TfrH
(J/g mole—datum temperature, 25°C)

oorve>.«se\#N»\r\eies#Nr\»\#ie\ne>»\

HHH(S(SffiTf<tinvo\OhooON
Molecular Enthalpy of Gases

©0\0\h'nooinooHhiD«nMn©^»nM'Cinrj
M'rH©ONTj-invNvo^voxvoa©vo'OOfn'th
O ^©vO'-’-"M(^iri-i^ONO\N'th'—1 m i—■ m
u #v»'Or\«\rsr\#N#«.eiC\#\r-rseiev#\eiri
^HtSinoOrHTfhTtOhTtMO\'£it^(XnC
^H-HrH^ffiffiTtiniovOr^xxON
Table 11-2

1—H ON ■<t 1—H m VO Tf 00 m m cn m ON r- On oo


o © 00 VO 04 VO g
1—H r~ g r- ON oo ON On oo On <N in
<N »n iq oo •n ©^ in ra ©^ oo ON vq vq On oc^ ON r| r- cq 04
S3 1—1 —H of VO ON co o-"
^“H —h
©" oo"
04 04 m
vo" in m" in vo"
m vo r- 00
vo" r-" oo" ON
ON © t-H 04

, co
(N 04 04 00 rj- >n © in On in m r4 © co in
VO Tf in o~ VO i—i «n ON 04 04 OO VO vo vo 1—! •n ©
o •n cq On q. q vO «n in rn ©^ ©^ <N Tf OO vq <N cq
u tS of of o-" T—T vo" i—H vo" r-" oo ©" nl
T—1 i—h 04 04 co
vo" oo" ©" m vo" oo"
vo r- oo ON © <N co »n
H
T"H T—1

ON r- VO © OO co ON o- 04 vo © oo m Tf CO
co i-—i © ON «n «n VO vo rf 04 T—< "'t r- oo <N © 1—1 04
S-h »n «n
©^ VO * q q <N cq in q ON OO^ 00 © rn vq ON cq oo^ cq
< T—( T of in oo" 1—H Tf" o~" ©" r-" ri On" vo" m" ^-T oo" vo"
04 m m 3
m «n VO r- oo oo ON

i—< ON NO co VO co © oo tj- VO oo m m m © oo co (N ON
<N co 1—H © ON co 1— in oo 04 ON m 1-H *-H ON CO 04 in
©^ vO q q q cq 00 vq «n m vq oo cn rq in
£ ,-T ,—T of in oo" r Tf" r-" co ©" r-" oo" vo" m" ©" oo m"
’—i 1—1 t—i <N m 3 m, •n VO oo oo On

<noomoO'-^vorNoor^ooooinr-H'sO»n»nON(n»n’—i
Conversion made to metric units.

coo4coco©coo4©vc©o4vo*-hOn'*}-voocononon
o »veseiOn^#\e>.*>.#\ryrv#sei*\^.or'*'
i --' N >n oo --iinooioo40NO~'<1'040NO-'nco^-<
*-ii-*~-H04coco''i-»nvovor^ooON©

o
o

<D
(-1
P
OJ
u,
<u
a. »n © © o
B<u 04 Tf NO 00 - - £SS„ -
VO OO © 04 Tf v£> 00
H H -H N N (N N M

259
260 Elementary Chemical Thermodynamics

It is common practice in this country to charge a combustion device with


the higher heating value of the fuel in evaluating its performance. This applies
to such devices as furnaces and internal combustion engines. The one
exception is that of gas turbines which commonly are charged with the lower
heating value. In European countries it is common practice to use the lower
heating value when evaluating the performance of combustion devices.

11-8 HEAT OF FORMATION


It can be assumed that a chemical compound may be formed by a chemical
action taking place between the elements composing the compound. The heat
of reaction for such a chemical process is known as the heat of formation. In
Section 11-6, it was shown that the heat of reaction was a function of the
states of a system. Thus, when a compound is formed from its elements with
given state points, the heat of formation is a fixed quantity and is removable
as heat during a formation. According to the first law, since the heat of
formation is dependent on state points, an amount of energy equal to the heat
of formation must be added to the system to dissociate a compound into its
elements. Since this principle was formulated by Lavoisier and Laplace
before the acceptance of the first law, it is often credited to them. A corollary
of this law, which was formulated by Hess, states that the total heat of
reaction of a chemical process is the same whether the reaction occurs in a
single process or by a series of processes.
Consider, for instance, a reaction between carbon and oxygen. Carbon
dioxide may be formed directly, or else carbon monoxide may be formed first
and then carbon dioxide may be formed by further combination with oxygen.
The energy of formation of the carbon dioxide is the same for the two cases.
Hence, the heat of formation of carbon dioxide must equal the heat of
formation of carbon monoxide plus the heat of reaction for the formation of
carbon dioxide from carbon monoxide.
When only one reaction between various reactants takes place at a given
time, the heat of formation may be measured directly. However, when two or
more reactions take place simultaneously, the heat of reaction for any
particular reaction must be calculated. Consider again the reaction of carbon
and oxygen. When sufficient oxygen is present and mixing is good, the
reaction will be complete and only carbon dioxide is formed. The heat of
formation of carbon dioxide can then be measured. However, carbon and
oxygen will not completely combine to form carbon monoxide without also
forming some carbon dioxide. Hence, the heat of formation of carbon
monoxide cannot be measured directly. Since carbon monoxide can be made
to combine completely with oxygen to form carbon dioxide, the heat of
reaction for this reaction can be measured. According to the principle just
enunciated, the heat of formation of carbon monoxide plus the heat of reaction
for converting carbon monoxide to carbon dioxide must equal the heat of
formation of carbon dioxide. Thus, for constant pressure, the heat of for-
Chemical Energy 261

mation of carbon dioxide is 393,761 J/g mole, and the heat of reaction for
converting carbon monoxide to carbon dioxide is 283,110 J/g mole. From
these values, the heat of formation of carbon monoxide is 110,651 J/g mole.
This method is of value in determining the heat of formation of a compound
which cannot be formed directly from its element and also for a compound
which cannot be formed singularly from its elements. This method may also
be used to determine the heat of reaction for a reaction process that cannot be
physically separated from other processes.
In addition to heat of combustion and heat of formation, other specialized
forms of heat of reaction are heat of solution, heat of dilution, heat of
neutralization, and heat of dissociation. The reader is referred to textbooks on
chemical thermodynamics for discussions of these particular types of heat of
reaction.

11-9 CHEMICAL ENERGY


During a combustion process, there is a change in the molecular structure of
the reactants. Since the process is a chemical one, the energy released by the
change in the molecular structure may be termed chemical energy. For a given
reaction, the chemical energy is a fixed quantity. Because of the change in the
molecular structure, there is also a change in the molecular kinetic and potential
energies. The change in this molecular energy shows some variation with
temperature.
Recognizing that the heating value is the negative of the heat of reaction,
Eq. 11-8 may be written as

Qp = (Hr - )
-Up molecular + chemical energy (11-14)
It is not feasible to determine the chemical energy directly, but it may be
calculated by use of Eq. 11-14 by using known heating values.
Various tables of enthalpies of various gases have been prepared. Using
these tables, a calculated value of the chemical energy may be obtained.
Unfortunately, various datum temperatures have been used in preparing
enthalpy tables. Hence, care must be taken to use the calculated value of
chemical energy in conjunction with the enthalpy tables from which it was
calculated. Note that the datum temperature for Table 11-2 is 25°C.
Since it is common practice to report the higher heating value of a fuel at a
constant pressure, Eq. 11-14 may be written as

chemical energy = QHP - (Hr - Hp)molecular (11-15)

In the use of Eq. 11-15, care must be taken to use the molecular enthalpy of
water in the liquid states. This enthalpy can be determined by subtracting the
heat of vaporization of water from the enthalpy of water vapor in Table 11-2.
For the normal small changes in the initial temperature of the air-fuel
temperature, the change in the heat of vaporization of water can be neglected
except where precision results are required. (The heat of vaporization of
262 Elementary Chemical Thermodynamics

water vapor is 2442.3 J/g at 25°C, 2358.5 J/g at 60°C, and 2257.0 J/g at 100°C.
This change in the heat of vaporization is very small in comparison with the
heating value of the fuels.)
Energy equations can be written for combustion processes. However, in
using these equations with either the heat of reaction or the heating value, it is
necessary first to calculate these energies at the original temperature of the
reactants. On the other hand, for a flow process, the energy equation may be
written as

Hreactants + chemical energy + Q = Hproducts 4- W (11-16)

In Eq. 11-16, the enthalpies are the molecular energies, as given in tables such
as Table 11-2. The chemical energy to be used in Eq. 11-16 is calculated by
using these enthalpies. Because chemical energy is a constant, there is an
advantage to using equations involving chemical energy rather than equations
using either heats of reaction or heating values.

Example 11-14. Determine the chemical energy of C8H18.

Solution. Refer to Example 11-3 for the combustion equation. Use Eq. 11-15
and Table 11-1 at 25°C.

chemical energy = 5,508,175 - [0 - (0 - 9 x 44,000)]

= 5,112,175 J/g mole Answer

Example 11-15. C8H18 is burned in a furnace with 20 percent excess air. The
initial air-fuel temperature is 40° C and the final temperature is 1600°C.
Determine the heat removed per gram mole of fuel. No work is performed.

Solution. The air-fuel mixture consists of 1 mole of C8H18, 12.5x1.2 =


15 moles 02 and 15(79/21) = 56.43 moles N2. The products consist of
2.5 moles 02, 8 moles C02, 9 moles N2, and 56.43 moles N2. Then

Hr = 1 x 2,547 = 2,547

15x433 = 6,495

56.43x431 = 24,321

Total = 33,813

Hp = 2.5x54,496 = 136,240

8 x 84,065 = 672,520

9x65,883 = 593,037

56.43x51,677 = 2,916,133

Total =4,317,930
Combustion Temperatures 263

Substituting into Eq. 11-16, using the chemical energy from Example 11-14,

33,813 + 5,112,175 + Q = 4,317,930

Q = - 828,058

or

heat removed = 828,058 J/g mole of fuel Answer

11-10 COMBUSTION TEMPERATURES


As we will learn in the next section, under certain conditions some of the
products of a combustion may dissociate (i.e., break down). When dis¬
sociation does occur, the final temperature is lower than that which would
have been reached without dissociation. In many furnaces, there is not
sufficient dissociation to alter significantly the maximum combustion tem¬
perature. It is the purpose of this section to determine the maximum com¬
bustion temperature, that is, the temperature with no dissociation. This is also
known as the adiabatic flame temperature.
Since ideal furnaces are open systems with no heat lost and with no work,
Eq. 11-16 reduces to

Tfreactants + chemical energy = Hproducts

Example 11-16. CO at 40°C is burned in a furnace with the theoretical


amount of air. Assuming complete combustion and no dissociation, determine
the maximum temperature.

Solution. The combustion equation is

co + 5 O’+5T (i)N’**co=+ n ©Nl


or

CO +1 02 + 1.88 N2-> C02 + 1.88 N2


Hp = H, + chemical energy

The chemical energy, by using Eq. 11-15 at 25°C, is

chemical energy = 283,123 - (0 - 0) = 283,123 J/g mole


Hr + chemical energy = 429 T 0.5 x 433 T 1.88 x 431 4- 283,123
= 284,578 J/g mole of fuel

This is the total enthalpy of the products. To determine the final temperature,
a trial-and-error solution will be used. At 2200°C,

Hp = 120,817 + 1.88(73,398) = 258,805 J


264 Elementary Chemical Thermodynamics

At 2400° C,
Hp = 133,650+ 1.88(80,733) = 285,428 J

By interpolation,
final temperature = 2394° C Answer

11-11 CHEMICAL EQUILIBRIUM


The general subject of equilibrium was discussed in Section 4-1. It was
pointed out that for a system to be in stable equilibrium, there can be no
active forces present that tend to cause a change in the state of the system. A
change in state of a system that is in complete equilibrium can be produced
only by bringing energy into the system. When a system is thermodynamically
in equilibrium, there can be no unbalanced mechanical, thermal, or chemical
forces. A system may be in equilibrium as far as one or two of these forces
are concerned and be in either a nonequilibrium or an unsteady equilibrium
state as far as the third force is concerned. For example, consider a mixture
of oxygen and carbon monoxide. When it is an isolated system, it will tend to
reach a uniform condition of temperature and pressure. The measurements of
these and other physical properties suggest that the system is in a state of
equilibrium. However, if a very small flame, a spark, or a proper catalyst is
introduced into the system, there will be a radical change in the nature of the
system. Oxygen and carbon monoxide will disappear, and carbon dioxide will
be formed. This change will be accompanied by a very high rate of tem¬
perature rise.
It is evident that although the system was in mechanical and thermal
equilibrium (as used here, the term means stable equilibrium), it was not in
chemical equilibrium. By measurement of physical properties, such as pres¬
sure and temperature, at various parts of the system, it is relatively easy to
determine whether or not mechanical and thermal equilibrium exist. It is
much more difficult to make measurements that will indicate directly whether
or not the system is in chemical equilibrium. In this connection it may be
shown that the Gibbs function may be used to determine whether or not the
system is in chemical equilibrium.
In Section 6-6, it was shown that the maximum useful work that can be
obtained from a change of state taking place with no net change in pressure
and temperature equals the decrease in the Gibbs function, stated in equation
form,

maximum useful work = (Gi - G2)t,p (11-17)

Consider a system that is in mechanical and thermal equilibrium with its


surroundings, but is not in chemical equilibrium. If a chemical reaction takes
place in a reversible manner, with no change in the temperature and pressure
of the system, then the useful work that can be delivered equals the decrease
in the Gibbs function during the reaction. When reaction rates are low in
Equilibrium in a Reactive Mixture 265

devices such as storage batteries and fuel cells, the reactions may approach
reversibility. Then the electrical energy deliverable approaches the decrease
in the Gibbs function.
In addition to predicting the maximum electrical energy deliverable during a
chemical reaction, as was stated above, the Gibbs function is helpful in
determining whether or not a system is in chemical equilibrium. Consider a
system composed of substances that possibly may react and form specific
products. Let the state of the products be so adjusted that their temperature
and pressure equal those of the reactants. Since G = H-TS=U+pV- TS,

dG = dU +pdV +V dp-TdS-SdT

(dG)T,p = dU + p dV - T dS (11-18)
The entropy of the system may be changed by the addition of heat.
Designate this entropy change as / dSe. The entropy of the system may also
be changed by irreversibilities within the system. Designate this entropy
change as / dSt. dU = 8Q - p dV = T dSe-p dV and T dS = T dSe + T dSh
Making these substitutions into Eq. 11-18 and reducing yields

(dG)T p = - T dSi (11-19)

As is discussed in Chapter 5, when the entropy of an isolated system is at


its maximum value, the system is in an equilibrium state. If the entropy can
increase, the system cannot be in equilibrium. Applying this concept to Eq.
11-19 for an isolated system, unless the Gibbs function of the products of a
possible reaction is less than those of the reactants, the system is in chemical
equilibrium and no reaction can take place. Thus, the Gibbs function is very
useful in predicting whether or not certain chemical reactions may take place
with the formation of specified products. Even though the Gibbs functions
show that it is possible for a reaction to take place, they do not give any
indication as to whether or not the reaction will actually occur. Neither do the
Gibbs functions predict what the reaction rate will be if the reaction does
occur.

11-12 EQUILIBRIUM IN A REACTIVE MIXTURE


When combustion takes place, the reaction may not be complete, that is, there
may be some combustible substances in the products. (It is assumed here that
all of the products are gaseous.) If the products are in an isolated system, they
will exist in a state of dynamic equilibrium. Some of the products will
combine with oxygen to complete the expected reaction. But an equivalent
amount of the ultimate products will dissociate into the original or other
reactants. This dissociation prevents the attainment of temperatures during a
combustion as high as might otherwise be expected. This phenomenon is
important in predicting performances, whether it be of an internal combustion
engine or the combustion chamber of a rocket engine.
266 Elementary Chemical Thermodynamics

The relative amounts of substances present after a combustion process


takes place can be calculated by the use of a so-called equilibrium constant.
This is not a true constant but is a function of the substances involved, the
temperature, and for actual gases, the pressure.
Consider that when two ideal gases A and B react, ideal gases C and D are
formed. Let the complete reaction equation be given as

aA+bB±*cC+dD (11-20)

where a, b, c, and d are the number of moles of gases A, B, C, and D,


respectively. The double arrow indicates the possibility of the reaction going
in either direction.
The factors involved in determining the equilibrium state (in which sub¬
stances A, B, C, and D are in a state of equilibrium with each other) may be
determined by the use of the van’t Hoff equilibrium box (see Fig. 11-3). Gases
A, B, C, and D are added to the box in the proper proportions until it is
established that they are in equilibrium with one another at a specified
temperature T and pressure p. Now let a moles of gas A and b moles of gas B
enter the control volume. The pressure of each gas is to be atmospheric as
supplied and the temperature is to be equal to T. Reaction will take place
within the box. Remove c moles of gas C and d moles of gas D. Maintain
conditions within the box constant. Since there are no changes in the
conditions within the box, the gases left in the box remain in an equilibrium
state. Adjust the pressure of gases C and D so that they will each leave the
control volume at atmospheric pressure and with a temperature T.
By definition, the Gibbs function, G = H - TS.
dG = dH -TdS-SdT
But
dH = dU T p dV + V dp = 8Q+Vdp = T dS + V dp
Then

dG = T dS + V dp - T dS — S dT

or
dG = V dp - S dT

Control volume

a moles of A c moles of C
-»>

b moles of B d moles of D

FIGURE 11-3 van’t Hoff equilibrium box.


Equilibrium in a Reactive Mixture 267

When there is no change in temperature,

dG = Vdp = NR"T dp
P
or

(G2-G,)T = NR0T In ^ (11-21)


Pi
Since there is no change in temperature of any gas as it enters or leaves the
box, Eq. 11-21 may be applied to the assumed system. For substance A,

(Ga2 - GOT = aR0T In(-p-)


2 ' \P atm/

where pA is the partial pressure of substance A in the box and patm is the
atmospheric pressure.
For substance B,
(GBl-GB,) = bR0T In(f5-)
2 ' \P atm/
For substance C,

(Gc2-Gc)t = cR0T ln(^f)

For substance D,

(GDl-GDl) = dR0T ln(E~)

Adding these four equations gives

(GAl + Gb2) - (Ga, + Gb) + (GC; + GO - (Gc, + Gd,)

= aRoT ln(-^-) + bR0T Inf-25-)


\P atm/ \P atm/

+ cR0Tln(^j + dRoTln(^j (11-22)

Since equilibrium conditions exist in the box, the sum of the Gibbs
functions of the reactants existing in the box must equal the sum of those of
the products, or
(GA2+Gb2)-(GCi + GDi) = 0 (11-23)

Combining Eq. 11-22 and Eq. 11-23 gives

(Gc2 + GO - (Ga, + Gb,) = RoT In (£±)‘


\P atm/
b c
+ RoT lnf^5-) +R0T lnf2^) + R0T \n(^)
VPatm/ V PC / V PD /

= R0T ln(^|4) + RoT ln((jrj*r) (11-24)


'PcPd/ 'PatmPatm'
268 Elementary Chemical Thermodynamics

When pressures are expressed in atmospheres, the last term of Eq. 11-24 is
zero (since the In 1 = 0). Under these conditions, Eq. 11-24 may be written as
follows:

PcPd

or

(11-25)

The left-hand side of Eq. 11-25 is known as the equilibrium constant and is
designated as Kp. Note that in Eq. 11-25, the Gibbs functions of the reactants
and the products are to be evaluated at atmospheric pressure and at the
temperature T. Thus, the left-hand side of Eq. 11-25 is a function only of
temperature. Although called the equilibrium constant, Kp is not a constant.
For a mixture of ideal gases, however, Kp is dependent only on temperature.
It should be noted that in Eq. 11-25 the reactants and products referred to are
those indicated by the reaction equation and not those that are in equilibrium.
The partial pressures in Eq. 11-25 are the partial pressures of the mixture. The
exponents a, b, c, and d are the numbers of moles of the various substances
as given by the reaction equation and are not the numbers of moles of the
substances present in an equilibrium mixture.
Since G is defined as H - TS, the Gibbs function may be calculated for any
substance when the absolute values of enthalpy and entropy are known.
Values of the Gibbs function at standard states have been calculated and are
tabulated in various texts on chemical engineering thermodynamics. The
equilibrium constant for a reaction of ideal gases can be calculated by use of
Eq. 11-25 when the Gibbs functions are available. Equilibrium constants for a
few reactions are given in Figure A-6 in the Appendix. It should be noted in
this figure that the common logarithm of the constants is given and not the
natural logarithm.
When a compound gas is heated, there is a tendency for the gas to
dissociate. At any temperature there is an equilibrium mixture of the gas and
its dissociated products. The relative amounts of each substance may be
determined by use of the equilibrium constant. The method of making this
determination is illustrated in the following example.

Example 11-17. Determine the amount of dissociation of carbon dioxide at


2500 K and standard atmospheric pressure.

Solution. Start with 1 mole of C02 and heat to 2500 K. Some dissociation will
occur, with CO and 02 being formed. Let X equal the number of moles of
C02 dissociated. Then X moles of CO and X/2 moles of 02 will be formed.
Equilibrium in a Reactive Mixture 269

The total number of moles present is

(1 - X)C02 + X(CO) + 02 or moles total

From Figure A-6 in the Appendix, at 2500 K the common log, log Kp = 1.45
for the reaction C0 + |02^C02.
For ideal gases, the partial pressure of each gas is proportional to the
number of moles of that gas. In using Eq. 11-25, gas A is CO, gas B is 02, and
gas C is C02. There is no gas D. Setting pt equal to the total pressure,
moles C02 _ 1-X 2(1-X)
Pc°2 “ total moles P' " (2 + X)/2 P' _ 2 + X P‘
_ moles 02 _ X/2 X
P°2 “ total moles p‘ ~ (2 + X)/2 p‘ ~ 2 + X p‘
_ moles CO _ X 2X
Pco “ total moles Pl ~ (2 + X)/2 Pl ~ 2 + X P‘

Since p, is given 1 atm, substitution into Eq. 11-25, with a = 1, b = and


c = 1, gives
2(1 — X)/(2 + X) (1 — X)(2 + X)1'2
[2X/(2 + X)][X/(2 + X)]1,5 Kp X^

Solve this equation for the value of X by trial and error so that log Kp =
1.45. If X =0.1,
(1 -0.1)(2 + 0.1)°5
41.25 and log Kp = 1.6154
Od13
If X =0.12,
(1 - 0.12)(2 + 0.12)0'5
30.8 and log Kp = 1.4855
O?5
If X =0.14,

„ (1 - 0.14)(2 + 0.14)°5
22.85 and log Kp = 1.3589
K"~ OI17
By interpolation, X- 0.126. In other words, 12.6 percent of the carbon
dioxide has dissociated at 2500 K. Of each original mole of carbon dioxide
there will be 0.126 mole of carbon monoxide, 0.063 mole of oxygen, and
0.874 mole of carbon dioxide. Answer

When a fuel burns, dissociation of the products of combustion prevents the


maximum flame temperature from reaching as high a value as it would if no
dissociation occurred. The effect of dissociation on flame temperatures is
illustrated in the following example.
Example 11-18. Carbon monoxide is supplied to a furnace at 40° C and
standard atmospheric pressure with the theoretical amount of air. Determine
270 Elementary Chemical Thermodynamics

the actual maximum furnace temperature, taking dissociation into account,


and compare it with the maximum temperature that would occur if there were
no dissociation.
Solution. With dissociation the combustion equation is CO + \ 02 + 1.88 N2-»
(1 - X)C02 + CO + X/2 02 + 1.88 N2. For this condition the energy equation is

(-H) mole CO + Hx/2 mole 02 ^ 1.88 mole N2)^0 C Chem. E. i mole CO ^


(JT(1_x)moleC02 + -HxmoleCO-^- Hx/2 mole 02 T W 1.88 moles N2) Tx + Chem. E.XmoleCO (^)

In the energy equation there are two unknowns, Tx and X. A second


equation involving these two unknowns is the equation for the equilibrium
constant. Since these two equations cannot be readily combined, they will be
solved by trial and error. If p2 is the final total pressure and N2 is the final
total number of moles, the partial pressures in the equilibrium mixture are as
follows:
Pco2 = (l-X)^

D =-X ^2-
P°2 2X N2

Pc o — X Pi
N2

From Eq. 11-25 the definition of the equilibrium constant, with a = 1, b =5,
and c — 1,
(1 -X)p2/N2 (1 — X)2°'5
(b)
p (Xp2IN2)(\l2Xp2IN2yu X'\p2IN2r
For a final assumed temperature, the value of X will be found by trial and
error, using the equilibrium constant together with Eq. (b). With the assumed
temperature and the calculated value of X, the total energy of the products
[the right side of Eq. (a)] will be calculated and compared with the energy of
the reactants [the left side of Eq. (a)]. If the two energies do not balance a
new temperature will be assumed. Assume T2 = 2200°C. From the Appendix,
at T = 2473 K, the common log, log Kp = 1.52, Kp = 33.2.
Assuming X = 0.1 and using Eq. (b) with p2= 1 atm,

N2 = 0.9 + 0.1+^y+ 1.88 = 2.93

Kp = (^I1*2 5 x 2.93°5 = 68.9

Repeating with X= 0.156, Kp = 33.3, which is close enough. Then at


2200°C, the products consist of 0.854 mole C02, 0.146 mole CO, 0.078 mole 02,
and 1.88 mole N2. The energy of these products equals

0.854(120,817)+ 0.146(74,108)+ 0.078(77,389)+ 1.88(73,398)


+ 0.146(284,578)= 299,570 J
Fugacity and Activity Coefficient 271

(The chemical energy of the CO is taken from Example 11-16.) Assume


T2 = 2000°C. From the Appendix at 2273°C, In Kp = 2.02, Kp = 105. Assume
X - 0.08, N2 = 0.92 + 0.08 + 0.08/2 4-1.88 = 2.92.

(1 - 0.08)2°'5 x 2.9205
98.2
0.0815

Repeating with X = 0.077, Kp = 104.4, which is sufficiently close. The


products consist of 0.923 mole C02, 0.077 mole CO, 0.0385 mole 02, and
1.88 moles N2. Their energy is

0.923(108,464) + 0.077(66,765) + 0.0385(69,665) + 1.88(66,110)


+ 0.077(284,578) = 278,730 J
The energy of the reactants from Example 11-16 is 284,578 J/g mole. Inter¬
polating between 2000°C and 2200°C for this value of energy, the final
temperature is 2056° C. Answer

A comparison of this answer with a temperature of 2394° C obtained in


Example 11-16 shows that dissociation has lowered the maximum attainable
temperature by 338°C.

11-13 FUGACITY AND ACTIVITY COEFFICIENT


The equation derived in Section 11-12 for the equilibrium constant of a reactive
gaseous mixture is valid for ideal gases. Since reactions between other gases are
quite common, it is desirable to set up an equation for the equilibrium constant
for a mixture of gases that are not ideal. In order that such a gaseous mixture may
be handled in the same way as a mixture of ideal gases, a new property known as
fugacity was created. Because the Gibbs function is the thermodynamic
potential for a chemical reaction, it is desirable to express the Gibbs function first
for the ideal gas and then to make modifications for the actual gas.
In general (dG)T = V dp. For an ideal gas, this becomes

(dG)T = NR0T d{In p) (11-26)

This equation is not valid for gases that are not ideal. However, there must be
some property of a substance which may be substituted for pressure in Eq. 11-26
to make the equation applicable for all gases. This property is called fugacity,
and the symbol f is used to denote it. Then,
(dG)r = NR0Td(\nf)= V dp (11-27)
The change in the Gibbs function can be obtained by integration of Eq. 11-27.
Thus,

(G2-G,)T = NRllTln{^j = jvdp (11-28)

Equation 11-28 shows that the change in the Gibbs function can be obtained if
an equation of state is available for the condition of the actual gas. As discussed
272 Elementary Chemical Thermodynamics

in Chapter 3, it is difficult to obtain an accurate equation of state for an actual gas.


Equation 11-28 also shows that the change in the Gibbs function may be
determined when the values of the fugacity of the gases are known at the two
states. Hence, it becomes desirable to investigate the nature of fugacity and
methods of determining this property.
Divide Eq. 11-27 by dp. Then

dG\ = NR0Td(\nf) = V dp =
dp) T dp dp
or
d(lnf)= V
(11-29)
dp NR0T

For an actual gas, pV = ZNR^T, where Z is the compressibility factor (see


Section 3-2). Then
V Z
NRqT p

Substituting in Eq. 11-29 gives

d(ln /)~1 =Z
dp iT P
Hence,

d{\nf)T = z(^-) =Zd(In p)T (11-30)

For an ideal gas, Z is unity. Then

d(ln f)T = d(ln p)T (11-31)

It is evident from Eq. 11-31 that fugacity must have the units of pressure.
Gases obey ideal-gas laws more closely when the pressure is reduced. If the
pressure could be reduced to zero, they would obey the ideal-gas laws and the
compressibility factor would become unity. For this condition, the fugacity
would be equal to the pressure.
To evaluate the fugacity of an actual gas, subtract d(ln p)T from each side of
Eq. 11-30. The result is

d(ln f)T — d(In p)T = Z d(In p)T — d(In p)T

or

d(lnA =(Z-l)d(lnp):

By integration,

(11-32)
Fugacity and Activity Coefficient 273

The ratio (//p)T may be evaluated by use of Eq. 11-32 when the compressibility
factor can be expressed in terms of pressure or when a mean compressibility
factor is known for the given pressure range. The ratio (Jlp)r is known as the
activity coefficient.2 It may be calculated and tabulated as a function of
temperature and pressure for various actual gases.
The equilibrium constant for mixtures of reactive actual gases may be
determined by use of the fugacity. For actual gases, Eq. 11-28 becomes

This equation may be used to evaluate the change in the Gibbs functions of
actual gases entering and leaving a van’t Hoff equilibrium box (see Section
11-12). Thus for actual gases,

fcfo _ TV ,
r aa rr hb ' ^ pr> (11-33)
Aj B

where X' is the equilibrium constant for a mixture of actual reactive gases. This
equilibrium constant is a function of both pressure and temperature.

2Sometimes referred to as the fugacity coefficient.


274 Elementary Chemical Thermodynamics

PROBLEMS __

11-1 (a) Determine the volumetric analysis of the products when 28 g of


carbon monoxide are burned completely with the theoretical amount
of air.
(b) Determine the mass of air required for part (a).
11-2 Same as Problem 11-1 except 25 percent excess air is used.
11-3 Assume that gasoline can be treated as C8H]8. Determine the theoreti¬
cal amount of air, in cubic meters, required to burn a liter of gasoline
if the gasoline has a specific gravity of 0.68. Air is supplied at 1.013
bars, 28°C.
11-4 Gasoline (C8H18) is burned completely with 30 percent excess air. The
products are at a pressure of 1.01 bars.
(a) Determine the partial pressure of water vapor in the products.
(b) Determine the dew-point temperature of the products.
11-5 A fuel oil is 83 percent carbon, 14.2 percent hydrogen on the weight
basis. The rest of the oil is inert. The oil is burned completely with 30
percent excess air at the rate of 9500 kg/h. Determine the volume rate
of flow of the products at 1.014 bars, 400°C.
11-6 A coal is 83 percent carbon, 5.8 percent free hydrogen (hydrogen free
to burn) and 1.2 percent sulfur. 28 percent excess air is supplied. Coal
is burned at the rate of 75,800 kg/h. Air is supplied at 1.013 bars 32°C.
Determine the volume rate of air flow.
11-7 The analysis of a natural gas is: CH4, 89.2 percent; C2H6, 10.4 percent;
N2, 0.4 percent. It is supplied with 12.4 m3 of air per cubic meter of
fuel. Determine: (a) the amount of excess air used and (b) the volu¬
metric analysis of the products.
11-8 The coal in Problem 11-6 produces the following Orsat analysis: C02,
12.9 percent; 02, 4.1 percent; CO, 0.4 percent. The refuse contains
0.015 g of carbon per gram of coal. Calculate (a) amount of carbon
burned per gram of coal, (b) moles of carbon per mole dry gas,
(c) moles of dry gas per gram of coal, (d) moles of nitrogen in a mole
of dry gas, (e) moles and grams of air supplied per mole of dry gas,
(f) moles of air supplied per gram of coal, and (g) percentage of
excess air supplied.
11-9 A fuel oil of Problem 11-5 is burned at the rate of 42.5 kg/min with 30
percent excess air. The humidity ratio in the incoming air is 0.015 g/g.
The products are at a pressure of 1.018 bars. Determine: (a) kilograms
of water vapor per minute in the products and (b) dew-point tem¬
perature of the products.
11-10 CH4 at 1.022 bars, 34°C, is burned at the rate of 0.98 m3/min with 20
percent excess air. Determine the heat removed per minute from the
Problems 275

products in cooling them from 600° C to 200° C. Assume complete


combustion.
11-11 A gasoline engine burns fuel at the rate of 28.5 kg/min. The Orsat
analysis shows 11.45 percent C02, 4.9 percent 02, and 0.3 percent CO
with no measurable amount of hydrocarbons. Assume that the fuel is
similar to C8H18. If air is supplied at 1.012 bars, 35°C, determine its
volume rate of flow.
11-12 The rate of air flow to the combustor of a gas turbine is 1450 kg/min
and the fuel flow is 31.8 kg/min. Treating the fuel as Ci2H26, determine
(a) the percentage of excess air and (b) the volumetric analysis of the
products. Assume complete combustion.
11-13 Determine the lower heating at constant pressure at 25°C of liquid
Ci2H26.
11-14 For gaseous Ci2H26 at 25°C, determine (a) the higher heating value at
constant volume and (b) the lower heating value at constant volume.
11-15 (a) Determine the higher heating value at constant pressure of
methane at 40° C.
(b) Same as part (a), except for gaseous C]2H26.
(c) Draw conclusions as to the necessity of knowing their tem¬
peratures in the use of heating values of fuels.
11-16 (a) Determine the lower heating value at constant pressure of
hydrogen.
(b) A coal is reported to have a higher heating value at constant
pressure of 32,000 J/g. The free-hydrogen (hydrogen free to burn)
content of the coal is 5.39 percent. Determine the lower heating value
of the coal. Compare the lower and higher heating values.
11-17 Determine the chemical energy of Ci2H26.
11-18 Determine the chemical energy of CH4.
11-19 C]2H26 is burned in a furnace with 22 percent excess air. The initial
temperature is 30°C. The products are cooled down to 400°C. Assume
complete combustion. Calculate the heat removed from the products
per gram mole of fuel.
11-20 Determine the theoretical maximum temperature (no dissociation) of
the products for Problem 11-19.
11-21 Determine the theoretical maximum temperature (no dissociation) of
the products for Problem 11-19 if no excess air is used.
11-22 C)2H26 is burned at constant volume with no excess air. The initial
temperature is 30°C. Assume complete combustion. Determine the
theoretical maximum temperature (no dissociation).
11-23 Determine the chemical energy of hydrogen.
11-24 Hydrogen is burned in a furnace with no excess air. The initial
276 Elementary Chemical Thermodynamics

temperature is 30°C. Assume complete combustion. Determine the


theoretical maximum temperature (no dissociation).
11-25 Repeat Problem 11-24 taking dissociation into account.
11-26 Same as Problem 11-24 but with 20 percent excess air.
11-27 Same as Problem 11-25 but with 20 percent excess air.
11-28 Same as Problem 11-26 but for a constant-volume combustion.
11-29 Same as Problem 11-27 but for a constant-volume combustion.
11-30 The gas turbine for Problem 11-12 is said to have a thermal efficiency
of 24.8 percent. It is not stated whether this efficiency is based on the
higher or lower heating value. Determine the power output using both
the higher and lower heating values.
_ 12 _

THERMODYNAMICS
OF FLUID FLOW

12-1 INTRODUCTION
Thermodynamics of fluid flow is concerned with those flowing fluids that
experience a change in their thermodynamic properties during the flow
process. The fluids involved may be gases, vapors, or liquids, Hence, while
some of the fluids may be treated as incompressible, the compressibility of
other fluids must be considered. In some cases, the frictional resistance of the
flow channel may be sufficient to produce significant changes in the fluid
properties. In other cases, factors such as heat addition and work production
may cause such major changes in fluid properties that the effects of side-wall
friction may be neglected.
Because of the generality of the thermodynamics of fluid flow, the prin¬
ciples of such flow are applicable to a wide range of engineering problems.
These include friction losses in pipelines, flow in both steam-turbine and
gas-turbine nozzles, air flow over airplane wings, the passage of guided
missiles through the atmosphere, flow in jet engines, the heating and cooling
of fluids in high-velocity heat exchangers, and turbo-machinery in general.
Fluid flow, in its broadest aspect, is complicated. Under some conditions,
parts of a fluid stream may be flowing in any of three planes. Furthermore,
conditions at a given point may vary with time. However, a large number of
types of flow may be approximated by making the following simplifying
assumptions:

1 Unidirectional flow. This means that all parts of the stream are considered
to flow parallel to the axis of the flow passage. This assumption does not
preclude a change in the flow area nor a change in direction of the flow
passage. Neither does it exclude turbulent flow, where cross flow of
individual particles takes place (see Section 12-2).
2 Steady flow. This means that the mass rates of flow through all cross-
sectional areas of the flow passage are equal and are constant.
3 Steady-state condition. This condition is attained when there is no varia¬
tion in properties with time at any point of the flow passage.

277
278 Thermodynamics of Fiuid Flow

Although, in general, actual flows do not follow these three simplifying


assumptions precisely, in many cases the deviations are small enough to be
insignificant. Consideration here will be restricted to internal flows.

12-2 LAMINAR AND TURBULENT FLOWS


Careful investigations of very low velocity flows show that the flow may be
described as being in layers, with each layer having a velocity differing from
its adjacent layer. This is illustrated in Figures 12-1 and 12-2. Such a flow is
known as laminar or streamline flow. It may be shown that for laminar flow,
the mean velocity equals one-half of the maximum velocity.
If the velocity of a fluid, whose flow is laminar, is increased sufficiently
there is an abrupt change in the nature of the flow. The nature of the new flow
is well characterized by its name turbulent flow (see Fig. 12-3). For turbulent
flow, the mean velocity is approximately 82 percent of the maximum velocity.
The determining factor as to whether the flow is laminar or turbulent is a
combination of relevant parameters which form a dimensional group known
as the Reynolds number, after Osborne Reynolds. This number is defined as

Re = ^ (12-1)

The Reynolds number is dimensionless when consistent units are used. In the
metric system, the diameter D of the flow system is given in meters, the
velocity T, in m/s, the density p, in kg/m3, and the viscosity p,, in kg/m-s.

FIGURE 12-1 Apparatus for studying the


nature of flow.

TZTZZZZZZ^^7777777777777777777?.

V//////7///777777777777777777777777
fmean 0-5 Imax FIGURE 12-2 Laminar flow.
Laminar and Turbulent Flows 279

FIGURE 12-3 Turbulent flow.

The product pT is the mass rate of flow per unit area. The symbol G is
used to designate this product. Substituting G in Eq. 12-1,

Re = — (12-la)

For steady flow in a constant-area passage, G is a constant.


The viscosity of fluid as used in Eq. 12-1 is known as the absolute viscosity
or dynamic viscosity. Absolute viscosity, p, may be defined as
F dx
(12-2)

where FI A = the force per unit area causing shear between adjacent layers of the fluid
dx = the distance between the layers under consideration
dT = the velocity difference between the two layers (see Fig. 12-4).
Equation 12-2 applies to true fluids but may not apply to semifluids such as
plastics. A fluid obeying this equation exactly is known as a Newtonian fluid.
In certain cases the kinematic viscosity of the fluid is useful. The kinematic
viscosity, v, is given as

(12-3)
P

FIGURE 12-4 Velocity gradient in flow.


280 Thermodynamics of Fluid Flow

The kinematic viscosity may be expressed in stokes or centistokes. A stoke


equals 10 4m2/s.
For noncircular passages, the concept of the hydraulic radius is used to
obtain an approximate equivalent flow system. The hydraulic radius is defined
as the cross-sectional area of the flow system divided by the wetted peri¬
meter. The diameter to be used in Eq. 12-1 is four times the hydraulic radius.
In general, when the dimensionless Reynolds number is less than 2000, the
flow normally is laminar. For Reynolds numbers between 2000 and 2300, the
flow is in a transition region, being partly laminar and partly turbulent. When
the Reynolds number is above 2300 the flow normally is turbulent.

Example 12-1. Water having a temperature of 30° C flows through a standard


2-in. pipe.1 Determine the maximum velocity for laminar flow.

Solution. From the table in Appendix 15, the internal diameter of the
pipe = 2.067 in. or

2.067 x 2.54
0.0525 m
100

From the table in Appendix 10,


p = 0.9956 g/cm3 or 995.6 kg/m3
p, = 801 x 10 6kg/m-s
DTp' _ Re u 2000x0.0008 A , A
Re
r= = 0.0525 x 995.6 = 0 0306 m/s A"SWer

Example 12-2. Air having a pressure of 1.01 bars and a temperature of 35° C
flows through a duct that is 30 cm by 40 cm. Determine the maximum velocity
for laminar flow.
Solution. The hydraulic radius equals
30x40
= 8.571 cm = 0.08571 m
2 x (30 + 40)
m _ p 1.01 x 105
= 1142 g/m3 = 1.142 kg/m3
P~V~RT~ 0.287(35 + 273.15)

From the table in Appendix 9,

pi = 18.2 x 10 hkg/m-s
= DTp y Re /x
Re

V = 20(?0-X-l8,2X.10.! = 0.0930 m/s Answer


(0.08571 x 4)1.142

The maximum velocities for laminar flow of water and air, as determined in

'At present, pipe sizes are still given in English units.


Energy Equation and Pressure Changes 281

Examples 12-1 and 12-2 are very much lower than those normally encoun¬
tered in pipes and ducts. Thus, generally, the flow of these fluids definitely is
turbulent. For this reason, further discussion in this chapter will be limited to
turbulent flows.

12-3 CONTINUITY OF FLOW


When it is specified that a flow system is a steady-state system, it is also
specified that the mass rate of flow is a constant throughout the system. In
other words, the total mass flowing per unit of time is the same past any plane
in the flow system. This statement holds true whether or not there is a change
in the direction of flow, in the elevation of the fluid, or in the flow area.
The product of area and velocity at any point in the flow system equals the
volume rate of flow, denoted by V, or

V = AY

But the mass rate of flow equals the product of p and V. Hence,

m = pAY (12-4)

where m = mass flow per unit of time


p = density, or mass per unit volume
A = flow area
V = fluid velocity

Since the mass rate of flow has been taken as constant,

m = pxAxYx = p2A2Y2 = • • • = pxAxYx (12-4a)

12-4 ENERGY EQUATION AND PRESSURE CHANGES


In a large number of flow systems, work is neither absorbed nor produced. The
energies involved in such a system are shown in Figure 12-5. In equation form
these energies per unit mass of fluid are
y2 y2
hx + zx + ~y + q = h2 + z2 + (12-5)

In the differential form, Eq. 12-5 can be written as

8q = dh + dz + YdY (12-6)

But dh = du + p dv- + v- dp and du + p dv = Tds. Then dh = T ds + v dp. The


entropy of the system can be changed by addition of heat to it. Designate this
entropy change as dse. The entropy of the system will be changed when
irreversibilities occur within the system. Designate this entropy change as ds*.
Then

dh = T dse T T dsi + v- dp
282 Thermodynamics of Fluid Flow

r2 y22
~2 ”2 “

FIGURE 12-5 Energies for a steady-flow system without work.

But 8q = T dse and v = 1/p, where p is the density. Substituting into Eq. 12-6,

T dse = T dse + Tdst + -— + dz + Y dY (12-7)


P
or
dp + pT dsi + Pdz + pYdY = 0 (12-8)

Each of the terms in Eq. 12-8 has the dimensions of pressure. The term pT ds,
represents the pressure loss due to irreversibility, that is, friction. The term
p dz is the change in pressure due to a change in elevation and may be termed
the gravity pressure change. pY dY is the change in pressure due to a change
in velocity, known as the velocity pressure change.
Changes in pressure in a flow system frequently are expressed in terms of
feet of the flowing fluid. This may be done by multiplying each term in Eq.
12-8 by 1 Ip:

^-+Tds, +dz + YdY = 0 (12-9)


P
In certain situations the change in pressure due to friction is so small in
comparison with other pressure changes that it may be neglected. When this
is true, Eq. 12-8 reduces to

dp + pdz +pYdY = 0 (12-8a)

Equation 12-8a frequently is known as the Bernoulli equation.


In the use of Eqs. 12-8 and 12-9, it is necessary to know how the density of
the fluid varies as it moves downstream. For some flow systems, this in¬
formation is not readily available. However, in many flow systems, the
variation that occurs in the density is so small that it can be neglected. This
Energy Equation and Pressure Changes 283

may be true for one of the following reasons:

1 The fluid is a constant-density fluid; that is, its density is not materially
influenced by a change in either temperature or pressure.
2 The fluid is an incompressible fluid; that is, its density is not materially
influenced by a change in pressure, although its density is influenced by a
change in temperature. For the given conditions there may not be a
sufficient change in temperature to affect the density materially.
3 The fluid has a density that is materially affected by a change in either
temperature or pressure. Nevertheless, the density may be considered to
be a constant when the changes in the pressure and temperature are such
small percentages of the absolute values that they can be neglected.
Under these conditions, Eq. 12-8 may be integrated and expressed as

(12-10)

where Apf is the pressure loss due to friction and equals / pT ds,.

Example 12.3. Water enters a piping system at a pressure of 250 kPa, 30° C with
a velocity of 2.1 m/s. The centerline of the piping system at exit is 14.2 m above
that at entrance. The pressure at exit is 100 kPa. Determine the frictional loss in
the piping system.

Solution. Since no information is given to the contrary, it will be assumed that


there will be no change in the diameter of the system, and hence, no change in the
velocity. From Eq. 12-10,

Ap/ = Pi~ p2 + p(zi~ z2)

From the Steam Tables in the Appendix, at 30°C,

p =j ^43 = 0.9957 g/cnr = 995.7 kg/m

Then
995.7(—14.2)9.807
A pf = 250 - 100 + = 11.34kPa Answer
1000

where 1 kg/ = 9.807 N.

Example 12-4. Determine the change in entropy per kilogram of water in


Example 12-3.
Solution

Ap, _ 11.34x1000
S' pT 995.7(273.15 + 30)
= 0.0376 N-m/kg-K = 0.0376 J/kg-K Answer
284 Thermodynamics of Fluid Flow

Example 12-5. Air enters a round horizontal duct with a diameter of 0.6 m at a
pressure of 1.01 bars and a temperature of 35°C and with a velocity of 41 m/s.
The diameter increases to 0.85 m in a distance of 2.2 m. Determine the final
pressure.

Solution. Since the duct is very short and the velocity is relatively low there
should not be a significant friction loss. Using Eq. 12-8a and assuming no
significant change in density,

p2-p, + p(z2-zO + ^U^ = 0

From Example 12-2, p = 1.142 kg/m3.


2

A, V, = A2y2, r2 = 41 (~) = 20.4 m/s

1 142(20 42 — 412)
p2 = 1.01 + ^ = 1.0028 bars Answer
F 2 x 10

Note: This very small change in pressure justifies the assumption of no


significant change in density.

12-5 STAGNATION PROPERTIES


Unless otherwise specified, the properties reported for a fluid are the static
properties. Static properties are those properties determined when there is no
relative motion between the fluid and the measuring device. For high-velocity
flows, a change in fluid velocity may produce a significant change in the static
properties of the fluid. Thus, it is desirable to devise a method of expressing
the properties of high-velocity fluids that includes the velocity effects. Such
properties are known as stagnation properties. Stagnation properties are those
properties that a fluid will possess if it should be brought essentially to rest in
an isentropic manner.2
From the definition of stagnation properties, applying the first law,

h°=h+Y (12-11)

where the superscript 0 stands for stagnation. For an ideal gas, dh = cp dT


and h°- h = cp(T° - T). With a datum temperature of 0 K, h = cpT.

h°= cpT°= cpT + ^

or
<Y2
T°=T + y— (12-12)
ZCp

2As discussed in Section 2-8c, when a fluid is brought completely to rest, there is a transformation
of its pv energy into internal energy, which produces a large temperature rise. Hence, the word
“essentially” is used here.
Stagnation Properties 285

The stagnation pressure is defined as the sum of the static pressure and the
velocity pressure, or

o pY2
P =p+ 2

The stagnation pressure can be determined for ideal gases when the static and
stagnation temperatures are known. Thus,

pi - (nyk~i (12-13)

where k is the ratio of specific heats (cp/cv). The stagnation density can be
determined by the use of the p-v-T relationship when the stagnation pressure
and stagnation temperature are known. By definition, the static and stagnation
entropies are equal.

Example 12-6. Air at a pressure of 8.2 bars and a temperature of 60°C has a
velocity of 180 m/s. Determine the stagnation pressure, temperature, and
density.

Solution. From Eq. 12-12,

1802
TO = 60 + 2x 1.004x 1000 = 76-14°C AnSWCr

(The 1000 converts J to g/s2.)


From Eq. 12-13,

o 0 „ /76.14 + 273.15314,04 . .
p = 8.21 ^q + 273 ]5 ) = 9.677 bars Answer

p RT° 287x349.29 9 653 kS/m Answer

Example 12-7. Steam in a pipeline has a velocity of 130 m/s. It has a tempera¬
ture of 215°C and a pressure of 15 bars. Specify its stagnation properties.

Solution. From Eq. 12-10, the stagnation enthalpy

i in2
h° = 2836.8 + 2^joO = 28453 J/8

s° = s =6.5379 J/g-K

From the superheated Steam Tables in the Appendix, by interpolation, with


h = 2845.3 and s = 6.5379, p° = 15.63 bars, T° = 222.2°C, and

v° = 134.3 cra’/g Answer


286 Thermodynamics of Fluid Flow

12-6 ACOUSTIC VELOCITY AND MACH NUMBER


Acoustic velocity is the velocity with which sound travels through a medium.
Thus, acoustic velocity may be defined as sonic velocity. The source of the
sound originates a pressure disturbance. This disturbance causes a wave to
move out in all directions in an attempt to overcome the pressure disturbance
and restore the system to equilibrium. Any pressure disturbance has a
tendency to propagate a pressure wave. The velocity of the pressure wave is a
function, among other factors, of the magnitude of the pressure disturbance.
When the pressure disturbance is a very weak one, the pressure wave travels
with the velocity of sound. For very large disturbances, such as detonation in
an internal combustion engine, the velocity of the pressure wave may be
several times the acoustic velocity.
To analyze the factors affecting the velocity of the pressure wave, consider
a fluid within a tube. A pressure disturbance at the right-hand end of the tube
will cause a pressure wave to travel through the tube from right to left.
However, if the fluid is caused to move through the tube toward the right with
a velocity equal to that of the pressure wave, the pressure wave will be held
in a stationary position.
When the pressure wave is a very weak one, there are small but abrupt
changes in the pressure, velocity, and density of the fluid as the pressure
wave is crossed, as indicated in Figure 12-6. For a very weak pressure wave,
or when 8plp is very small, 8p approaches dp, 8T approaches dT, and 8p
approaches dp. Since the pressure wave occupies a very narrow band, the
change in state across the pressure wave may be considered to be adiabatic.
Furthermore, because of the very small change in state, the process may be
considered to approach reversibility and, therefore, may be considered to be
an isentropic process. When Eq. 12-8a is applied,

dp + ydy = 0 (12-14)
P
From Eq. 12-4, pAT is constant. Hence,

y + X + ~° (12-15)

Source of

FIGURE 12-6 Stationary pressure wave.


Acoustic Velocity and Mach Number 287

Since A is constant,

dp dY
p r
= o (12-15 a)

From Eq. 12-6,

dh + YdY = 0 (12-16)

For an ideal gas, dh = cp dT. When this value of dh and the value of (Y dY)
from Eq. 12-14 are substituted,

cp dT ~~7~ = Q (12-17)

For an ideal gas, cp = c„ + R and c* = cplk. Hence,

Cp~ic=R or Cp=rHR
Then,

dp
(12-18)
rh8" P
When T = p/pR,

dT = ^ d —
R p

so that,

R dT = d (£-) = pdp iPdp


\pl p-

From Eq. 12-18,

k [p dp — p dp\ _ dp
k- 1 \ p1 / p

This reduces to
dp _ kp
(12-19)
dp p

From Eq. 12-14,

Substituting this value of dY into Eq. 12-15a,


288 Thermodynamics of Fluid Flow

or

y= (12-20)

When the value of dpldp from Eq. 12-19 is substituted,

r= = VkRT (12-20a)
v P

The velocity in Eq. 12-20a is the velocity of the fluid required to maintain
the pressure wave in a stationary position. It must equal the velocity the
pressure wave would have if it were to travel through a still gas. But this
velocity equals the acoustic velocity a.
Therefore, for an ideal gas, the acoustic velocity

a = VkRT (12-20b)
As will be seen later on in this chapter, the behavior of a flowing fluid is a
function of the ratio of the fluid velocity to the acoustic velocity. For
instance, when a gas enters a diverging passage with a velocity lower than the
acoustic velocity, an increase occurs in the pressure as the fluid moves
through the passage. On the other hand, when a fluid enters a diverging
passage with a velocity greater than the acoustic velocity, a decrease in
pressure takes place in the passage. It should be noted that the important
criterion is neither the magnitude of the velocity of the fluid nor that of the
acoustic velocity, but is rather the ratio of the fluid velocity to the acoustic
velocity. This ratio is called the Mach number and is symbolized by M. Thus,
y
M=— (12-21)
a

When the velocity of the fluid equals the acoustic velocity, the fluid is said
to possess sonic velocity. Any velocity less than sonic velocity is said to be
subsonic, and a velocity greater than sonic velocity is called supersonic.

12-7 ADIABATIC FLOW WITH CONSTANT-AREA


PASSAGE
In Section 12.4, consideration was given to the adiabatic flow of a constant-
density fluid. Consideration will now be given to the adiabatic flow of a fluid
that undergoes a measurable change in density. Because of the assumption of
continuous flow in a passage of constant area, work will not be involved here.
Consider a flow passage with constant area as in Figure 12-7. Since the flow
is steady, the rate of flow is constant and
G = Plyl = P2y2 (12-22)

Because there is no heat transfer, the stagnation enthalpy remains constant


Adiabatic Flow with Constant-Area Passage 289

Insulation
* •' *■- ■. .
1 1 . l 7"T|! .;•••• ••
*777777777, G/7>/yy///////777/77//7// 777777777.

7777WA

1 2
hi h2
% ^2
2 2

FIGURE 12-7 Adiabatic flow with constant area.

and equal to the original stagnation enthalpy h?. Then


Y2 y2
hx + ^=h1 + ^=h\ (12-23)

Substitution of the value of G from Eq. 12-22 gives


G2
h, + ^-i= h°, (12-24)

A line on any diagram representing the properties of a fluid during its flow
(adiabatic) through a constant-area passage is called a Fanno line. Equation
12-24 provides information for plotting a Fanno line on an enthalpy-density
plane or on an enthalpy-volume plane. A more common diagram used for
showing a Fanno line is the enthalpy-entropy diagram. Equation 12-24 may be
used as a basis for calculations for a Fanno line on the h-s plane. This
equation gives a means of determining the density for each value of the
enthalpy. Since enthalpy and density are independent properties, they fix the
values of entropy. The determination of the entropy may be relatively easy
for an ideal gas, but it may necessitate a trial-and-error solution for vapors.

Example 12-8. Air enters a constant-area duct at 30°C with a pressure of


520 kPa and leaves at a pressure of 410 kPa. The duct area is 0.02 nr. The
entering velocity is 82 m/s. Plot the Fanno line on the h-s plane.

Solution. The original density of the air is

520,000 , 3
Pl 287(30 + 273.15) 5-977kS/m
G = pY = 5.977 x 82 = 490.1 kg/m2-s

Using datum temperature of 0°C for enthalpy, Eq. 12-11 gives


8?2
h'!= 1000(1.004) x 30 + = 33,482 J/kg
290 Thermodynamics of Fluid Flow

Because there is a pressure drop, there is an increase in the specific volume,


and thus a decrease in the density. Since the area and the mass flow are
constant, the velocity, and hence, the kinetic energy increase. The resulting
decrease in enthalpy is accompanied by a decrease in air temperature.
Assume t2 = 28°C.
h2= 1000(1.004)(28) = 28,112 J/kg

From Eq. 12-24,

7 = 33,482 - 28,112 = 5370 J/kg


2P
2 490. V
P2 22.36, p2 = 4.729 kg/nr
2 x 5370

4.729 x 287(28+J-73-15) = 40g_74 kPa


P2 =
1000

S2 - S\ = C
■© - R '» ©
= 1.004In^4^-0.287 In408'74
303.15 520

= -0.006646 + 0.069095
= 0.06245 J/g-K or 62.45 J/kg-K

Similarly, when t2 = 29°C,

h2= 29,116 J/kg = 4366 J/kg


2 P:
p2 = 5.2448 kg/nr

p2 = 454.81 kPa and s2—s i = 35.126 J/kg-K

From this information just obtained, the Fanno line is plotted in Figure
12-8. The pressure line is plotted also in this figure. It shows that the specific
entropy has increased by 61.5 J/kg-K in the duct and that the static enthalpy
at the exit has decreased to 28,160 J/kg. The temperature at exit is

28,160
28.05°C
(1000 x 1.004)

Example 12-9. Determine the velocity of the air at exit in Example 12-8.

Solution. From Example 12-8, h° = 33,482 J/kg and h = 28,160 J/kg. Then

—2 = 33,482-28,160 = 5322 J/kg or Y2= 103.2 m/s Answer


Adiabatic Flow with Varying Area 291

FIGURE 12-8 Fanno lines for Example 12-8.

12-8 ADIABATIC FLOW WITH VARYING AREA


Adiabatic flow through a passage with varying area is encountered in many
different types of flow systems, such as diffusers, injectors and ejectors,
flow-measuring devices, and nozzles for hydraulic, steam, and gas turbines.
Flow-measuring devices include nozzles, orifices, and venturi meters. Many
of these flow systems operate at temperatures close to that of their surround¬
ings. In other cases, the time of passage of a particle of fluid is so short that
generally no appreciable amount of heat is transferred from the fluid particle.
In some cases, because of very high rates of flow, the total heat lost from a
device may be appreciable. Nevertheless, the heat loss per unit mass of fluid
generally is insignificant, and the flow may be considered to be adiabatic.
In addition to assuming that the flow is adiabatic, it will also be assumed
that the flow passage is continuous and that no energy in any form is
transferred between the fluid and its surroundings. Under these conditions,
the stagnation enthalpy remains constant. Thus,

h°\ = hj = • • • hx = a constant (12-25)

Since steady flow is also assumed, the mass flow is constant, or

piA 1T1 = P2A2U2 = • • * pxAxYx = a constant (12-26)


In the ideal case, friction may be neglected and the flow may be taken to be
isentropic. The effect of the change in elevation will also be neglected. From
292 Thermodynamics of Fluid Flow

Eq. 12-26,
(12-26a)
p a r
When Eq. 12-23 is expressed in the differential form, the result, with h°
being constant, is
dh + YdY = 0

But dh = T ds + dplp, or (dh)s = dplp for isentropic flows. Then

— + y dv = o

Substitution of the value of dY from Eq. 12-26a gives

dp = _y l-Y dp _ YdA\ = r21dp + dA (12-27)


P \ P A / \ p A ,

For very low Mach numbers, the percentage change in density is in¬
significant; that is, dplp is substantially equal to zero. Under these conditions,
Eq. 12-27 reduces to
2 dA
dp = Ylp (12-27a)
A

Examination of Eq. 12-27a shows that for the assumed conditions of the
isentropic flow of an ideal gas at very low Mach numbers, the pressure varies
directly with the area. Hence, a diverging passage produces an increase in
pressure, and a converging passage causes a decrease in pressure.
Equation 12-27 indicates that the pressure is proportional to the square of
the velocity. At high Mach numbers, a change in velocity produces a sizable
change in pressure and, hence, in density. Therefore, the term dplp cannot be
neglected. It is difficult to predict from Eq. 12-27 what will happen to the
pressure, and hence the velocity, when the flow area is changed. To permit
such a prediction, introduce the Mach number into the equation.
Since a2 = dpi dp,
dp
dp = a2 dp or dp (12-28)

Substitution of this value of dp in Eq. 12-27 gives

pY2 dp 2 dA Y2 dp dA
A
or
dA
A

hence,

2 dA
dp(\- M2) = pY (12-29)
A
Adiabatic Flow with Varying Area 293

When the value of dp from Eq. 12-28 is substituted in Eq. 12-29, the result
is

a2 dp(l - M2) = pV2 ^

or

dp _ a2M2 dA _ M2 dA
p <r(l-M2) A 1 - M2 A

Equating this value of dplp to its value in Eq. 12-26a gives


M2 dA dA dT
1 - M2 A A V

Hence,

dT = dA M2 dA dA / M2 \
r A 1 - JVf2 A A V 1 -M:j
or

dT dA/ 1 \
(12-30)
r a \i -m2)

An examination of Eq. 12-30 indicates that when the Mach number is less
than unity, the velocity decreases as the area increases, and vice versa. For
supersonic flow, the velocity increases as the area increases. These state¬
ments may be illustrated by considering a venturi tube and a convergent-
divergent nozzle. As shown in Figure 12-9, the two flow passages are similar

FIGURE 12-9 Venturi tube and convergent-divergent nozzle.


294 Thermodynamics of Fluid Flow

in appearance. Normally, the pressure drop in a venturi tube is relatively low,


and the velocity approaching the divergent section of the tube is subsonic.
Hence, there is a decrease in velocity in the divergent portion of the venturi
tube. On the other hand, because of the large pressure drop in a convergent-
divergent nozzle, the fluid generally attains supersonic velocity near the
entrance to the divergent section. Hence, the velocity continues to increase as
the fluid passes through the divergent section of the nozzle.
The behavior of a fluid in its expansion in a nozzle from a region of high
pressure to one of low pressure may be illustrated by considering the
isentropic flow of an ideal gas. For steady unidirectional flow,

PiYiAi = P2Y2A2 = • • • pAY = m = a constant

The stagnation enthalpy remains constant. Hence, if changes in elevation are


neglected, the velocity at any point in the nozzle is

r = V2(h°- h) = V2cp(T°- 7’) (12-31)

In Section 12-6 it was shown that cp = [kl(k — \)]R. So

r= V2(k^T)*(T#-T)
Since T/T° = (plpY~m,
/nvH« r (fc-D/k j
T°-T = T°-T° (A) = T°
1 (v)
~

Then
j \ (k-\)lk-
•a 1

1_J
n_j*

O
T—
c*

(12-32)
II

1
H
v-H

V .i
Also,
1
1

rh = pA yjl ^ j) RT° [l - |
L-

Since
(k-l)lk
p _ p /py-"'1 pp(i“t),k(p°) p° (i
P RT xftyp) RTs- RT° \t

then

Ap°
m = (12-33)
VrT*
For isentropic flow in the absence of external work, the stagnation enthalpy
remains constant. Hence, the stagnation temperature remains constant. Fur¬
thermore, the stagnation pressure is also constant. (Since dh° = T ds° +
dp°lp° = 0, it follows that for an isentropic process dp0 = -p°T° ds° = 0.) If
Eq. 12-33 is written in terms of the initial stagnation pressure and tempera-
Adiabatic Flow with Varying Area 295

ture, p? and T?, the result is

Ap?
(12-34)
VIT?
or

(12-34a)
A

The maximum rate of flow per unit area occurs at the minimum cross-
sectional area of the nozzle, which is known as the throat. The pressure at
this point can be obtained in terms of the initial stagnation pressure by
differentiating the term [(p/p?)2/k - (p/p?)tk 1)/k] from Eq. 12-34a with respect to
p/p? and setting the differential equal to 0. The result of this operation is that
kl(k-1)
(12-35)
Jc + 1

Equation 12-35 expresses the pressure at the throat of a convergent-


divergent nozzle for the isentropic expansion of an ideal gas. It is assumed
that the back pressure, or the pressure in the region into which the nozzle is
discharging, is lower than the value of the pressure given by Eq. 12-35. For
these conditions, the velocity at the throat may be obtained by substituting
the pressure ratio given by Eq. 12-35 in Eq. 12-32. First, Eq. 12-32 is rewritten
as follows:

using

or

y=

Substitution for p?/p from Eq. 12-35 gives

or
y = VkRT
This expression for the velocity at the throat is exactly the same as that given
in Eq. 12-20a for the acoustic velocity. Thus, the fluid at the throat has a sonic
velocity when the mass flow rate is at a maximum.
296 Thermodynamics of Fluid Flow

When the back pressure equals the initial pressure, and it is assumed that
there is no velocity approaching the nozzle, there is no flow and the velocity
at the throat of the nozzle is zero. As the back pressure is reduced, the
velocity at the throat of the nozzle increases, approaching the sonic velocity
as a limiting value.
When the velocity at the throat is subsonic, the mass rate of flow through
the nozzle is less than the maximum possible. The pressure at the throat is
greater than that existing for sonic velocity. As the back pressure is de¬
creased, the mass rate of flow increases and reaches a maximum value, name¬
ly, that value obtained by substituting the pressure ratio given Eq. 12-35 in
Eq. 12-34a. A further decrease in back pressure cannot increase the mass
rate of flow at the minimum cross-sectional area nor can it increase the
velocity at that point. Hence, the fluid velocity existing at the minimum
cross-sectional area when maximum rate of flow is obtained is known as the
critical velocity. This is sonic velocity. The pressure corresponding to the
critical velocity is called the critical pressure. The critical pressure may be
calculated by use of Eq. 12-35.
The conditions relative to the throat of a convergent-divergent nozzle for
the isentropic flow of an ideal gas are restated as follows:

1 The maximum velocity at the throat equals sonic velocity.


2 The minimum pressure at the throat is the critical pressure.
3 Maximum mass rate of flow occurs when the throat pressure equals the
critical pressure.
4 Decreasing the back pressure below the critical pressure does not change
the velocity or pressure at the throat or the mass rate of flow.
5 When the back pressure is lowered below the critical value, the gas
expands in the divergent part of the nozzle and the velocity becomes
supersonic.

12-9 FLOW IN IDEAL NOZZLES


In Section 12-8, general consideration was given to isentropic flow in nozzles
from a region of high pressure to one of lower pressure. In this section,
specific consideration will be given to the mass rates of flow as well as to the
velocity and density changes taking place in a convergent-divergent nozzle.
Such consideration can best be illustrated by means of an example.

Example 12-10. Air enters a nozzle at 120°C with a pressure of 1600 kPa and
expands to a pressure of 200 kPa (see Figure 12-10). The flow area approach¬
ing the nozzle is large and, hence, the initial velocity can be neglected. The
mass rate of flow is 2 kg/s. Treat the air as an ideal gas and assume an
isentropic expansion. Determine the variations in velocity, specific volume,
and area through the nozzle.
Flow in Ideal Nozzles 297

FIGURE 12-10 Convergent-divergent turbine nozzle.

Solution. The pressures will be taken at 200-kPa intervals and will be


presented in tabular form.

P2 t2 T,-T2 hi - fl; r2 V-2 A Diameter


(kPa) Pi/Pi (Pl/P2)<‘_,)'‘ (K) (K) (J/g) (m/s) (m3/kg) (cm2) (cm)

1600 393.15 0 0.07052


1400 1.1429 1.0389 378.43 14.72 14.78 171.9 0.07758 9.03 3.39
1200 1.3333 1.0857 362.12 31.03 31.15 249.6 0.08661 6.94 2.97
1000 1.600 1.1437 343.75 49.40 49.60 315.0 0.09886 6.26 2.82
v>
OO
L/\

1.893 1.2000 327.63 65.52 65.78 362.7 0.11128 6.13 2.79


800 2.000 1.2190 322.52 70.63 70.91 376.6 0.11570 6.14 2.80
600 2.667 1.3234 297.08 96.07 96.45 439.2 0.14210 6.47 2.87
400 4.000 1.4860 264.57 128.58 129.09 508.1 0.18983 7.47 3.08
200 8.000 1.8114 217.04 176.11 176.81 594.7 0.31145 10.47 3.65

aThis is the critical pressure.

The results are plotted in Figure 12-11. An examination of the curve shows
the following points.

1 Except for the very first part of the nozzle, the velocity increases at
almost a constant rate as the pressure decreases.
2 For a considerable part of the expansion, the specific volume increases
gradually. During the latter part of the expansion, the specific volume
increases very rapidly.
3 Because of the large increase in specific volume, the area of the last part
of the nozzle increases rapidly. This relationship will be particularly
noticeable if the expansion should be continued to a lower pressure.

The curves in Figure 12-11 are based on the assumption that an equal
pressure drop occurs in each unit length of the nozzle. It will be difficult to
manufacture such a nozzle. A common form of a convergent-divergent nozzle
was illustrated in Figure 12-10. In this type of nozzle there will not be the
same pressure drop in each unit length of the nozzle. Curves showing the
variations in the velocity and specific volume for this type of nozzle will have
the same general shape as those in Figure 12-11, but they will be distorted
because of unequal pressure drops.
Pressure (kPa)

FIGURE 12-11 Nozzle for Example 12-10.

298
Nozzle Efficiency 299

If the nozzle in Example 12-10 is to be designed to exhaust at any pressure


higher than 200 kPa, the plot will be exactly the same as that shown in Figure
12-11, but it will be terminated as soon as the given exhaust pressure is
reached. By use of Eq. 12-35, the critical pressure for Example 12-10 is found
to be 845 kPa. An expansion to a pressure lower than this value requires the
nozzle to have a divergent portion. If the expansion is to take place only to the
critical pressure or a higher value, a convergent nozzle is satisfactory.

12-10 NOZZLE EFFICIENCY


The fundamental purpose of a nozzle is to deliver the fluid with the maximum
amount of kinetic energy for the given conditions.
The term nozzle efficiency is used to express the efficiency of the nozzle in
fulfilling its purpose. If rjn denotes this efficiency,

actual kinetic energy


(12-36)
theoretical kinetic energy

The kinetic energy appearing at exit may come partly from kinetic energy
entering the nozzle and partly from enthalpy which is transformed into kinetic
energy because of the expansion occurring in the nozzle. For given conditions,
the maximum enthalpy that can be transformed into kinetic energy is the
enthalpy change which occurs in an isentropic expansion. Then, on the gram
basis,

rli 2 (12-37)
T|" (rill) + (h, - h2)s

where subscripts 1 and 2 refer to conditions at nozzle entrance and exit,


respectively.
When the kinetic energy entering the nozzle can be neglected,

_ (h] h2)actual
(12-37a)
V" ~ (h,-/i2)s
Extensive experimental work has shown that the nozzle efficiency is
affected by many factors, such as the material of which the nozzle is
constructed, the workmanship in manufacture, the size of the nozzle, the
shape of the nozzle (that is, round, elliptical, or rectangular), the angle of
nozzle-wall divergence, whether the nozzle passage is straight or curved, the
nature of the fluid and its state, and the fluid velocity. The investigation of the
magnitude of the effects of each of these factors is beyond the scope of this
text. However, the effect of nozzle efficiency on flow in nozzles will be
investigated.
The primary effect of friction and turbulence within a nozzle is to reduce
the nozzle efficiency and, hence, to reduce the kinetic energy of nozzle exit. A
second effect is to reduce the mass rate of flow through a nozzle of given
area. Expansion in a nozzle may be illustrated on an H-S plane, as in Figure
300 Thermodynamics of Fluid Flow

s
FIGURE 12-12 Expansion with friction.

12-12. Process 1-2 is an isentropic one, and process 1-3 is an actual expansion
to the same final pressure. If the kinetic energy at entrance is neglected, the
change in enthalpy must equal the kinetic energy at exit. The difference
between the actual and theoretical changes in enthalpy is called reheat.
Example 12-11 illustrates the effect of friction on the performance of a
nozzle.

Example 12-11. Steam enters a nozzle at 250°C with a pressure of 1.5 MPa.
The exhaust pressure is 0.9 MPa. The nozzle efficiency is 95 percent. Cal¬
culate the exit area for a mass flow of 1.2 kg/s.

Solution. It may be assumed that the approach area is sufficiently large so


that the approach velocity may be neglected. Although Eq. 12-35 was derived
for an ideal gas, it is sufficiently close for superheated steam. From page 124
of Steam Tables by Keenan, Keyes, Hill, and Moore (Wiley, New York,
1978), the mean value of k for this expansion is approximately 1.298. Then the
critical pressure equals
1.298/0.298

1.5 = 0.819 MPa


1.298+1

Since the exhaust pressure exceeds the critical pressure, the nozzle should be
a convergent one.
From the Steam Tables in the Appendix, ht = 2923.2 and s, = 6.7090. At the
end of isentropic expansion, p = 0.9 MPa and h = 2813.2. The actual change in
enthalpy in the nozzle from Eq. 12-37a equals

0.95(2923.2 - 2813.2) = 104.5 J/g


Shock Waves in Nozzles 301

The actual exit velocity equals

V2 x 104.5 x 1000 = 457.2 m/s


The actual enthalpy at exit equals

2923.2- 104.5 = 2818.7 J/g

For this enthalpy and a pressure of 0.9 MPa, the specific volume is
226.6 cm3/g. The nozzle exit area,

rinv- 1.2 x 1000x226.6


A = = 5.95 cm2 Answer
y 457.2 x 100

It is to be noted that if friction were to be neglected for the conditions of


Example 12-11, the exit velocity would be 469 m/s, the specific volume would
be 225.1 cm3/g, and the required exit area would be 5.908 cm2.

12-11 SHOCK WAVES IN NOZZLES


Because of variations in operating conditions, a nozzle frequently operates
under conditions that differ widely from those used in its design. Although the
pressures at both the nozzle entrance and exit may change, it will be assumed
here that the initial conditions are held constant and that only the back
pressure will change. This assumption will permit a ready determination of
the principles involved.
Consider the effect of an increase in the back pressure above the design
value on the performance of a convergent-divergent nozzle (see Fig. 12-13).
When the back pressure equals the initial pressure, there is no flow and hence
no change in pressure in the nozzle. If the back pressure is reduced below
that of the initial pressure, there will be a drop in pressure up to the throat of
the nozzle. Unless sonic velocity is reached at the throat, there must be an
increase in the pressure in the divergent section as shown by Eq. 12-29. This
is illustrated by curves B, C, and D in Figure 12-13, with the acoustic velocity
being reached at the throat for conditions of curve D. A reduction of the back
pressure from that of curve D will result in a supersonic velocity just beyond
the throat. The pressure in the nozzle at this point is much lower than the
back pressure. Also, from Eq. 12-29, with supersonic flow there will be a
further reduction in pressure as the area increases, rather than the required
increase in pressure necessary to bring the pressure up to that of the back
pressure. The required increase in pressure can be obtained only if there is a
discontinuity in the flow pattern. Such a discontinuity is known as a standing
shock wave or a normal compression shock, since it takes place in a plane
perpendicular to the direction of motion.
Although the actual shock wave is not accompanied by the instantaneous
pressure rise shown by curve E in Figure 12-13, extensive measurements
have shown that the pressure rise does take place in a very small distance.
Because of this fact, both the change in flow area across the shock wave and
302 Thermodynamics of Fluid Flow

FIGURE 12-13 Convergent-divergent nozzle with varying back pressure.

the side-wall friction may be neglected (see Fig. 12-14). Assuming one¬
dimensional flow,

G = piTi = p2y2 - a constant


Here state 1 is on the upstream face of the shock wave and state 2 is on the
downstream face.
From Eq. 12-8a, dp = — pY dY or

P2-P1 = Prcr2- r,) = p,v]-Plrl (12-38)

P1 P2

% r2

FIGURE 12-14 Analysis of normal compression


shock.
Shock Waves in Nozzles 303

But, for an ideal gas,

Yz = M2a2 = MlkRT
Substituting this value of Y2 into Eq. 12-38 and rearranging,

P i_ 1 + M^c
(12-39)
p2 1 + Mxk
Substituting the value of V = VM2kRT into the expression pxYx = p2Y2 and
rearranging,
pxMx p2M2
(12-40)
Vt, Vt2
Since h? = h2.
Y2
h 1 + -r- = h2 +
r2
°"'2

or, for an ideal gas,


y? y?
CpTi + — = cpT2 + —

Substituting for Y2 its value of M2kRT and rearranging,

T\ [1 + (k — 1)/2]M2
(12-41)
T2 [l + (k-l)/2]Mf
When the values of pxlp2 from Eq. 12-39 and T,/T2 from Eq. 12-41 are
substituted into Eq. 12-40, the result is

(1 + M22k)Mx (1 + M2k)M2
Vl + [(k — 1)/2]M2 V1 + [(k - l)/2]Mf
From this,

_2 Mi + 2/(k — 1)
M2— r^Ti TTil-TvITTT (12-42)
[2k/(k — l)]Mf — 1

The relationship between the Mach numbers at the two faces of the shock
wave, as given by Eq. 12-42, is plotted in Figure 12-15 for air. Note that the
larger the Mach number before the shock, the smaller will be the Mach
number after the shock.
Since a compression shock is an irreversible process, it is desirable to analyze
the entropy change that takes place. For an ideal gas, the entropy change, using
Eq. 8-25a, can be written as

Rk
52~ Si = In
k—1
or
k /l + [(k-l)/2]M?\
s2 - sx = R In - R In (12-43)
k^ Vl + [(k-l)/2]Af;/
0.6 0.8 1.0 1.2 1.4 1.6

M:

FIGURE 12-15 Relationship of Mach numbers for normal shock (for air).

FIGURE 12-16 Entropy changes for normal shocks (for air).

304
Diabatic Flow 305

The entropy changes across normal shocks in air, as determined by Eq. 12-43,
are plotted in Figure 12-16.
An examination of Figures 12-15 and 12-16 shows that when a normal shock
occurs in supersonic flow, there is a change from supersonic to subsonic flow
and the entropy increases. This increase in entropy is to be expected since the
shock is an irreversible process.
Assume that a shock can occur in a nozzle when the flow is subsonic. As
shown in Eq. 12-42, the flow changes from subsonic across the shock wave to
supersonic. By Eq. 12-39 there is a decrease in pressure. Thus, this type of
shock would be expansion shock. However, as shown by Eq. 12-43, there
must be a decrease in entropy. But the entropy cannot possibly decrease in an
adiabatic process. Hence, the second law tells us that an expansion shock
cannot occur here. It should be noted, however, that an expansion shock may
occur in other types of flow.

12-12 DIABATIC FLOW


When heat is added to a flowing fluid, the flow is said to be diabatic
(nonadiabatic). To simplify the problem it will be assumed that the flow
passage is of constant area. Hence, mechanical work will not be exchanged
with the surroundings. When a significant amount of heat is added in a short
distance, normally the effect of this heat on the properties of the fluid is very
much greater than is the effect of side-wall friction. Under these conditions,
the effect of side-wall friction can be neglected. These simplified assumptions
approximate many actual heat exchange problems.
With these assumptions

piV\ = p2Y2 = G = a constant = pY


From Eq. 12-8a, with no change in elevation,

dp + pYdY = 0=dp + GdY, since pY = G

or

Pi + GYi = p2T GY2 = a constant (12-44)


q2 q2
Pi H-= p2 H-= a constant (12-44a)
Pi P2

The magnitude of the changes in the properties of a fluid is determined by


the amount of heat added. The energy equation for steady flow, for the
conditions assumed here, reduces to
Y]
h 1 -I- ~y + q — h2 + (12-45)

or
G2
h1 + 2 T q = h2 + (12-45a)
2pi
306 Thermodynamics of Fluid Flow

where q is the heat added per unit mass. Equations 12-44 and 12-45 may be
used together to determine the conditions after a heating process.

Example 12-12. Air enters a tube having an internal diameter of 0.12 m with
a velocity of 90 m/s at a pressure of 620 kPa and a temperature of 80°C. It is
heated by the addition of 800kJ/s. Determine the pressure, temperature, and
velocity after heating.

Solution. The density of the entering air is

n =_
620 x 1000
= 6.117 kg/nr
RT, 0.287x 1000x353.15

G = piVi = 6.117 x 90 = 550.5 kg/m2-s

The total mass rate of flow equals

a ccf\ c 0.122X7T , , . .
GA = 550.5 x---= 6.226 kg/s

The heat added per kilogram equals

800
128.5 kJ/kg
6.226

From Eq. 12-44a,

irvAA , 550.52 , 550.52 ,303,050 ^AC/1A


620 x 1000 + - = p2 +- p2 H-2-= 669,540
6.117 P2 Pi

Since p = pRT, p2 ~ 287p2T2,

287p2T2 + 303’°50 = 669,540 (a)


Pi

Assuming a datum temperature of 0 K for h and using a value of cp =


1.004 kJ/kg-K, from Eq. 12-45a,

550.5: 550.5'
1.004 x (80 + 273.15) + + 128.5 = 1.004T? +
2x6.117zx 1000 2p2 x 1000
or

1.004T2+ 151325 = 487.15 (b)


P2

Multiply Eq. (b) by 285,86p2:

287p2T2 + ^^-= 138,91 lp2 (C)


P2
Diabatic Flow 307

Subtract Eq. (c) from Eq. (a):

303,050 - 43,315
= 669,540- 138,91 lp2
P2
259,735 = 669,540p2- 138,911 pi
138,911 pi - 669,540p2 + 259,735 = 0
P2~4.8199p2+ 1.8698 = 0
p2 = 4.394 kg/m1
From Eq. (b),

487.15 — 151.525/4.3942
T? =
2 1.004
= 477.39 K Answer
Then

p2 = p2RT2 = 287 x 4.394 x 477.39


= 602.3 kPa Answer
«/. _ G _ 550.5 _ loc ^ , A
v2— — a — 125.3 m/s Answer
p2 4.394

Note that there is a pressure drop of 17.7 kPa. This pressure drop occurs in
the absence of friction and is due solely to the heating. The answers found in
this example are not precise, since the true mean specific heat is slightly
higher than the value assumed. A higher degree of accuracy can be obtained
by taking a new value of cp for the temperatures just calculated and then
repeating that portion of the problem involving cp.
When a vapor, such as steam, is heated in a constant-area passage, it is
necessary to resort to a trial-and-error solution, since an accurate equation of
the state of the vapor (relating temperature, pressure, and density) is too complex
to be solved directly.
Although the velocities used in Example 12-12 are much higher than those
normally encountered in industrial heating problems, it is desirable to
investigate the effects of heating on fluid properties in order that a more
comprehensive picture can be obtained of the effects of heating fluids in
constant-area passages.
A plot showing the variations in the properties of a fluid as it is heated in a
constant-area passage is known as a Rayleigh line. In Figure 12-17, a Rayleigh
line is shown for a gas together with lines of constant entropy. In Figure
12-18, a Rayleigh line is shown together with lines of constant enthalpy. An
examination of Figure 12-17 shows that as the pressure decreases (due to
addition of heat) there is an increase in the entropy until a maximum value of
entropy is reached. Any further addition of heat results in a decrease in the
entropy. Obviously this condition is impossible since addition of heat should
produce an increase in entropy.
308 Thermodynamics of Fluid Flow

0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Specific volume (m3/kg)

FIGURE 12-17 Rayleigh lines and lines of constant entropy.

FIGURE 12-18 Rayleigh lines and lines of constant enthalpy.

From Eq. 12-44a,

p=c = C — G~u

or

dp
= -G2 = -p2r2 (12-46)
d v-
For a line of constant entropy, the slope is

(ia - m . ( ^
~P “ dp/ \dp):
(12-47)

In Eq. 12-20, the velocity V is the acoustic velocity, a, or a = V dpi dp.


Summary 309

Substituting this into Eq. 12-47,

(12-48)

At the point of maximum entropy for the Rayleigh line, its slope must equal
that of the entropy line. From Eq. 12-46a, the slope of a Rayleigh line,

Y= -p2r2 (12-49)
dv

Equating Eq. 12-48 and Eq. 12-49, it may be seen that when the entropy for
the Rayleigh line is at its maximum velocity, the fluid velocity equals the sonic
velocity.
By definition,

dh°=dh + Ydr
Sq = Tds = dfi - —
P

From Eq. 12-8a, with dz = 0,

dP-=-ydy
P

or

T ds = dh + V dT = dh° (12-50)
For the conditions discussed here, the increase in the stagnation enthalpy
equals the heat added. By Eq. 12-50, the entropy must increase as heat is
added. But, as just stated, the entropy must decrease after the fluid reaches its
sonic velocity. This means that when heat is added to fluid in a constant-area
passage that has reached sonic velocity, an unstable condition is produced
and the mass rate of flow through the nozzle will change.
When a fluid having a supersonic velocity is heated in a constant-area
passage, there will be an increase in its entropy. This means a moving to the
left in Figure 12-17, toward the point of maximum entropy. Under these
conditions, an addition of heat produces an increase in the pressure, an
increase in the density, and a decrease in the velocity until sufficient heat is
added to cause the velocity to be reduced to the sonic value.

12-13 SUMMARY
The thermodynamics of fluid flow, in the broader sense, is quite complicated
but, in this chapter it has been greatly simplified by dealing only with
one-dimensional flow and, for the most part, considering only ideal gases.
However, some fundamental concepts have been introduced. These include
the concept of sonic velocity and how it affects the performance of nozzles
310 Thermodynamics of Fluid Flow

as well as producing a limitation on heating in constant-area frictionless


passages. The effects of nozzle shape on operating conditions were in¬
vestigated. It has been shown that, for an ideal gas flowing at subsonic
velocity in a constant-area passage, friction increases the specific volume,
thus causing an increase in the velocity, a somewhat surprising effect.
Because of the complexities of this very important subject, textbooks
devoted to the subject have been prepared for various levels. The reader is
referred to some of these textbooks for a more detailed discussion of the
thermodynamics of fluid flow.
Problems 311

PROBLEMS ____
12-1 (a) Water having a temperature of 27°C flows through a standard 1-in.
pipe. Determine the maximum velocity that the water may have with
the flow being laminar.
(b) Same as part (a), except the pipe is 12 in.
(c) Draw general conclusions relative to the nature of water flow in
pipes for most engineering problems.
12-2 (a) Air at 1.012 bars, 28°C flows through a duct having an internal
diameter of 25.6 cm. Determine the maximum velocity that the air can
have with the flow being laminar.
(b) Same as part (a), except for air in a 42.5-cm-diameter duct at 20.2
bars, 200°C.
(c) Draw general conclusions relative to the nature of air flow in ducts
for most engineering problems.
12-3 Air at 1.012 bars, 28°C flows through a rectangular duct, 26.7 cm by
48.5 cm with a velocity of 40 m/s. Determine the Reynolds number.

12-4 Air enters a constant area duct, 24 cm by 38 cm, at 1.018 bars, 24°C
with a velocity of 42 m/s. It leaves at 1.012 bars, 242°C. Determine the
velocity at exit.
12-5 Same as Problem 12-4 except at exit the duct is 18 cm by 26 cm.
12-6 Water enters a 6-in. pipe at 22° C with a velocity of 2.1 m/s and leaves
at 91°C. Determine the Reynolds number at entrance and exit.
12-7 Water flows through a 4-in. pipe at the rate of 14.9 kg/s. The tem¬
perature at exit is 42°C. Determine the Reynolds number.
12-8 Carbon dioxide at 250°C, 1.82bars flows through a 4-in. pipe with a
velocity of 44 m/s. Determine the Reynolds number.
12-9 Water at 10° C enters a 2-in. horizontal pipe with a velocity of 4 m/s at
8.95 bars. Neglect friction. The water temperature at exit is 140°C.
Determine the pressure, (static) at exit in bars.
12-10 Same as Problem 12-9 except the water leaves in a 1-in. pipe.
12-11 Same as Problem 12-10 except the water leaves at an elevation of 44 m
higher than that at entrance.
12-12 Water enters a l|-in. pipe at 13.8 bars, 28°C, with a velocity of 1.8 m/s
and leaves at 120°C at an elevation of 25 m above that at entrance. The
frictional pressure loss in the pipe is 0.12 bar. Determine the pressure
at exit.
12-13 (a) Determine the change in entropy for Problem 12-12 per second.
(b) Determine the change in entropy in part (a) due to friction if no
heat had been added.
12-14 Air at a pressure of 8.2 bars, 70°C has a stagnation temperature of
86°C. Determine: (a) air velocity and (b) stagnation pressure.
312 Thermodynamics of Fluid Flow

12-15 Determine the velocity of sound in air (a) at 1.010 bars, 20°C and (b) at
1.015 bars, 300°C.
12-16 Treating steam as an ideal gas, determine the velocity of sound in
steam at 0.1015 MPa, 200°C.
12-17 Determine and compare the acoustic velocity of air and hydrogen at
0.1015 MPa, 25°C.
12-18 (a) Air at 1.012 bars, 40°C flows in a duct with a velocity of 275 m/s.
Using a datum of 0°C for enthalpy, determine (a) stagnation enthalpy,
(b) stagnation temperature, and (c) stagnation pressure.
12-19 Steam at 1.35 MPa, 200°C has a velocity of 200 m/s. Determine: (a)
stagnation enthalpy, (b) stagnation temperature, and (c) stagnation
pressure.
12-20 Air expands isentropically from 13.5 bars, 105°C to 8.1 bars. Deter¬
mine the ratio of the final to initial acoustic velocity.
12-21 (a) Air enters a constant-area duct at 1.52 bars, 43°C with a velocity
of 300 m/s. The pressure at duct exit is 1.44 bars. Neglect heat
transfer. Determine the velocity and temperature at exit.
(b) For this problem, what is the effect of friction in the duct on
velocity and temperature?
12-22 Same as Problem 12-20 except that the velocity at entrance is 580 m/s.
The pressure at duct exit is 1.60 bars.
12-23 Determine the critical pressure for a fluid entering a nozzle under the
following conditions.
(a) Air at 0.1014 MPa, 25°C.
(b) Air at 13.5 MPa, 1000°C.
(c) Steam at 0.1014 MPa, 150°C.
(d) Steam at 13.5 MPa, 550°C.
(e) Hydrogen at 0.1014 MPa, 25°C.
12-24 Air enters a nozzle at 2.21 bars, 88°C with a velocity of 80 m/s.
Neglect friction. Determine the pressure in the nozzle at the point the
velocity of the air reaches acoustic velocity.
12-25 Same as Problem 12-22 except neglect the initial velocity.
12-26 Air enters a nozzle at 21.8 bars, 92°C with negligible velocity. The exit
pressure is 3.18 bars. Neglect friction. Determine the throat and mouth
areas flow at a flow of 1.42 kg/s.
12-27 Same as Problem 12-26 except the velocity at entrance is 42 m/s.
12-28 (a) The overall nozzle efficiency in Problem 12-26 is 94 percent.
Determine the exit area.
(b) The efficiency of the nozzle up to the throat for Problem 12-26 is
97 percent. Determine the throat area.
(c) Comment on the necessity of taking the nozzle efficiency into
account in calculating the nozzle and throat areas.
Problems 313

12-29 Steam at 6.5 MPa, 350° C expands to 4.0 MPa. The nozzle efficiency is
95 percent. The nozzle exit area is 6.54 cm2. Determine the mass rate
of flow.
12-30 Steam at 6.0 MPa, 350°C expands to 0.80 MPa. Neglect friction up to
the throat. The overall nozzle efficiency is 95 percent. The throat area
is 8.6 cm2. Determine: (a) mass rate of flow and (b) exit area.
12-31 Steam at 7.0 kPa, 6.5 percent moisture content enters the last stage of
a steam turbine at the rate of 1,487,000 kg/h. The nozzle efficiency is
95 percent. The pressure at nozzle exit is 4.0 kPa. Determine the total
exit area required.
12-32 Air at 15.2 bars, 90°C expands in an ideal nozzle. An exit velocity of
560 m/s is desired. Neglect the initial velocity. Determine the exit
pressure.
12-33 Same as Problem 12-32 except that the efficiency of the nozzle is 95
percent.
12-34 A compression shock occurs in a divergent flow passage. On the
upstream side of the shock, the fluid (air) has a velocity of 420 m/s and
is at 2.06 bars, 43°C. Determine: (a) the Mach number on the down¬
stream side of the shock and (b) the velocity and pressure on the
downstream side of the shock.
12-35 Determine the change in entropy per kilogram of air as the result of
the shock in Problem 12-34.
12-36 Same as Problem 12-34 except that the initial velocity is 370 m/s.
Comment on the effect of the initial velocity on the percentage of
change in the Mach number.
12-37 Determine the change in entropy per kilogram of air as the result of
the shock in Problem 12-36.
12-38 If it is possible for a shock to occur in air having a velocity of 300 m/s,
determine the Mach number, the velocity, and the pressure on the
downstream side of the shock. The initial conditions are 2.06 bars,
43°C.
12-39 Determine the change in entropy per kilogram of air for Problem
12-38. Comment on the answer.
12-40 A compression shock occurs in steam having a Mach number of 1.25.
Assuming a value of k = 1.3, determine the ratio of the pressure after
the shock to that before the shock.
12-41 Steam enters a standard 4-in. pipe at 1.45 MPa, 205° C with a velocity
of 25 m/s. It leaves at a pressure of 1.44 MPa. Neglecting friction,
determine the rate of heat addition.
12-42 Air enters a constant-area tube at 1.42 bars, 90°C with a velocity of
460 m/s. 20J/g of heat are removed per gram of air. Determine the
velocity, pressure, and temperature at exit. Neglect friction.
314 Thermodynamics of Fluid Flow

12-43 Same as Problem 12-42 except add lOJ/g.


12-44 Air enters a diffuser with a velocity of 250 m/s and a temperature of
30°C. It leaves with a velocity of 90 m/s. Neglect friction and heat
transfer.
(a) Determine the temperature at exit.
(b) If the initial pressure is 1.25 bars, determine the pressure at exit.
(c) Determine the ratio of the exit area to entrance area.
(d) Will your answers in the first three parts change if there is some
friction? If so, how?
GAS CYCLES AND
APPLICATIONS

13-1 INTRODUCTION
When a system undergoes a series of processes and then returns to its original
state, the series of processes is known as a cycle. Cycles are further defined
as closed or thermodynamic cycles and as open cycles. In the closed cycle,
there is no exchange of mass between the system and its surroundings. The
working substance of the system may expand or contract with work and heat
being added to or removed from it. An example of such a system is the steam
in a steam power plant. The system includes all devices through which the
steam or its condensate flows, namely: (1) the steam generator including the
superheater, the economizer, and the reheater, (2) the steam turbine, (3) the
condenser, (4) the boiler feed pump, and (5) feed water heaters of various
types. Another example of the closed cycle is the vapor compression refri¬
geration system.
In the open cycle, the system undergoes a series of processes during which
some of the fluid within the system is delivered to the surroundings with
additional fluid being drawn from the outside to replace it. Consider an
internal combustion engine of the automotive type. A mixture of air and fuel
is drawn into the cylinder, compressed, and burned. The resulting gases
expand and then are exhausted from the engine. This series of processes can
now be repeated. The conditions of the system at the start of intake are the
same as originally, even though the gases themselves are not returned to their
original state. Thus the series of processes may be termed an open cycle.
Nature may attempt to reduce the carbon dioxide produced to carbon and
oxygen, and hence, in a way, nature may close the cycle. However, the series
of processes that occur do not constitute a closed thermodynamic cycle.
In general, gas turbines1 are of the open type although a very few operate
with a closed cycle. The so-called hot-air engines operate, overall, in a closed
cycle.

'Sometimes called combustion turbines.


2Air is contained in the hot-air engine system. It is heated by external combustion with the heat
being transferred to the air in a heat exchanger.

315
316 Gas Cycles and Applications

It is a common practice to use an ideal closed-cycle analysis to establish the


theoretical performance of both the open and closed cycle engines. The
various processes of the ideal cycle are selected so as to be as similar to those
of the actual engine as possible. It is assumed that the heat added through the
cylinder walls of an ideal engine produces results equivalent to the com¬
bustion taking place in the actual engine. Likewise, it is assumed that heat
may be removed through the walls of the ideal engine to produce a result
similar to the initial pressure drop that occurs when the exhaust valves of an
actual engine open.
The assumptions normally made in establishing the ideal-gas cycles may be
summarized as follows:

1 The cycle is a complete thermodynamic cycle.


2 The working substance is air.
3 The air is an ideal gas having a constant specific heat.
4 Heat may be transferred to and from the air so as to produce any desired
process, such as a constant-volume or a constant-pressure process.
5 All processes are reversible ones.

This method of analyzing cycles is known as the air-standard method.


Since the specific heat of the air is assumed to be constant, it is called the
cold-air standard. It is thus distinguished from the so-called hot-air standard
which may be used to account for the effects of the variations in the specific
heat which occur in an actual engine.

13-2 CRITERIA FOR CYCLES


In designing a heat engine to utilize a theoretical cycle, the actual engine
should have the following characteristics:

1 The thermal efficiency should be high.


2 The power output per unit of engine volume and weight should be high,
particularly for portable engines.
3 The maximum temperature and pressure should be sufficiently low to
prevent rapid deterioration of the engine.

In analyzing an air-standard cycle to discover its inherent advantages and


disadvantages, consideration must be given to ascertaining whether or not the
series of events in the actual engine will approach the theoretical cycle closely
enough to allow it to have the desired characteristics. For example, consider
the Carnot cycle. Although not stated as such, it was in reality analyzed in
Chapter 5 on the air-standard basis as described above.
Although the thermal efficiency of a Carnot cycle engine is of the highest

When an actual engine operates with a vapor, its ideal cycle is analyzed by using the particular
vapor.
The Otto Cycle 317

degree, the inherently low mean effective pressure4 means that any actual
engine approximating the Carnot cycle and using a gas as a working substance
will be a feeble engine. One type of internal combustion engine, the Lenoir,
had no compression before the combustion process. It may be discovered by
an air-standard analysis that when the cycle operates without compression, its
thermal efficiency is inherently low. Although Lenoir engines were actually
built, they became obsolete when internal combustion engines, which have
compression before combustion, came into use.
In the following sections, several cycles will be analyzed on the air-
standard basis to explore their inherent characteristics. The thermodynamic
superiorities of some cycles will be indicated, as well as their thermodynamic
and practical limitations.

13-3 THE OTTO CYCLE


In 1868 Beau de Roches first set forth a series of events required for the
efficient operation of an internal combustion engine. They are:
1 Suction (intake) of the charge during the inward stroke of the piston.
2 Compression during the outward stroke of the piston.
3 Ignition of the charge near the outward dead center followed by expansion
of the gases during the inward stroke of the piston.
4 Exhaust of the gases during the outward stroke of the piston.

The first successful engine to operate on this series of events was built by
the German Nicolaus Otto. Hence, the theoretical cycle embodying these
events is known as the Otto cycle. The air-standard Otto cycle is shown in
Figure 13-1, in the p-v and T-s planes.
Air in the Otto cycle is compressed isentropically from state 1 to state 2. The
original engines were low-speed ones. Hence, combustion was completed with
very little piston motion, thus approaching a constant-volume process. In the
air-standard cycle, this process is simulated by the addition of heat at
constant volume, process 2-3 in Figure 13-1. Then the air was expanded in an
isentropic manner until the piston reached the end of its stroke. In the original
engines, the exhaust valve was opened close to the end of the expansion
stroke, with the pressure dropping with little change in the position of the
piston. This process is simulated in the air-standard cycle by a constant-
volume cooling of the air (process 4-1) until the state of the air reaches the
original state, state 1. Although modern engines operate at such high speeds
that neither the combustion process nor the initial pressure drops through the
exhaust valves approximate constant-volume processes, the air-standard Otto
cycle is still used as a basis of comparison in judging the performance of
spark-ignition types of engines.
An objective in analyzing air-standard cycles is to obtain values of thermal

4See Eq. 13-9 for definition.


318 Gas Cycles and Applications

(a)

FIGURE 13-1 Air-standard Otto cycle.

efficiency, mean effective pressure, maximum cycle temperature, and maxi¬


mum cycle pressure. In addition, it is desirable to investigate the factors
affecting these four parameters.
Let r equal the compression ratio (the ratio of the volume before com¬
pression to the volume after compression). In Figure 13-1, r= V{/V2. The
temperature at the end of compression is

T2=T,r‘“‘ (13-1)
and the pressure at the end of compression is

Pi = Pirk (13-2)
Since heat is added at constant volume, the heat supplied is
The Otto Cycle 319

Also
T3
P3 = Pi
t2

where T3 and p3 are the maximum cycle temperature and pressure. The
thermal efficiency, by definition, is

_ - W_Q,-Q, (13-5)
v‘ ~ Qs Qs
where Qr is the heat rejected per cycle.
Since the heat is rejected at constant volume.

Qr = mCv(T3 - T4) (13-6)


Then, in terms of temperatures.

_ w _ Qs ~ Qr = mc„(T3 — T2)-mcJJ4- Tt)


v' Qs Q mc^(T3— T2)
or

(13-7)

But

so that

T4-T, . T3-T2
T, T2
Substituting into Eq. 13-7,

T,, = l-|l=l-X (13-8)

Equation 13-8 shows that the efficiency of the air-standard Otto cycle
depends solely on the compression ratio and is independent of factors such as
the heat added per cycle and the conditions at the start of compression.
A factor that is very useful in determining the work delivered per unit of
cylinder volume is the mean effective pressure (mep). Mean effective pressure
is defined as the mean pressure which, acting on the piston for one stroke,
produces work equivalent to the net cycle work. By definition

mep x piston area x stroke = net cycle work

Then

net cycle work


mep = (13-9)
piston displacement
320 Gas Cycles and Applications

Example 13-1. The conditions at the start of compression in an air-standard


Otto cycle engine are 103 kPa pressure, 26°C. The heat added per kilogram of
air is 1850 kJ. The compression ratio is 8 to 1. Determine the maximum
temperature and pressure, the thermal efficiency, and the mep.

Solution. Refer to Figure 13-1. At the end of compression,

p2 = p,rk = 103 x 8'4 = 1893 kPa


T2 = T,rk~' = (26 + 273.15)804 = 687.27 K

Q = mc„(T3-T2), T3=T, + -£-


I ills A L

1850
= 687.27 +
1 x 0.718
= 3263.9 K Answer

P3 = P2(|l) = 1893 = 8900 kPa Answer

The thermal efficiency

rjt = (l — J100 = 56.48 percent Answer

The thermal efficiency can also be found as follows:

T _ T3 3263.9
J 4 ~ “T^T = 1420.7 K
r - 8l
Qr = mcXT4-Tx) = 1 x 0.718(1420.7 - 299.15)
= 805.3 kJ/kg
_Qs~Qr_ 1850-805.3
Vt = 0.5647 or 56.47 percent
Qs 1850
The initial volume of the air is

mRTi 1 x 287x299.15
= 0.833 m3/kg
Pi 103 x 103

The piston displacement (P.D.) equals \ x 0.833 = 0.7289 m3/kg or

P.D. = 0.833 - \ x 0.833 = 0.7289 m3/kg


w = QS-Qr = 1850 - 805.3 = 1044.7 kJ/kg
From Eq. 13-9,

1044.7
1433.2 kPa Answer
mep 0.7289

An examination of the results of Example 13-1 shows that inherently the


Otto cycle has a high thermal efficiency and a high mep. On the other hand,
the mep of the Carnot cycle is very low when cycle pressures are reasonable.
The Diesel Cycle 321

Table 13-1
Thermal Efficiency
of Air-Standard Otto Cycle

Thermal Efficiency
Compression Ratio (percent)

4.0 42.6
5.0 47.5
6.0 51.2
6.0 54.1
8.0 56.5
9.0 58.5
10.0 60.2
11.0 61.7
12.0 63.0

The reason for this is that even when small amounts of heat are added in the
Carnot cycle, the cylinder volumes required are exceedingly large. For ex¬
ample, use the same conditions as in Example 13-1 except only 370 kJ of
heat are added per kilogram of air.5 The maximum cylinder volume is
5.444 m3, which is 6.535 times that of the Otto cycle, while the cycle work
produced is only one-fifth of that in Example 13-1. The mep for this Carnot
cycle is 39.13 kPa, which is only 2.7 percent of that for the Otto cycle. (This
low mep of the Carnot cycle makes it impractical to try to operate an engine
having a gaseous working substance on the Carnot cycle). The calculated
maximum temperature and pressure as determined in Example 13-1 are much
higher than those encountered in practice. One reason is that the variations in
specific heat are neglected in the air standard. Another reason is that the
actual combustion process is far from being at constant volume. In addition,
there are significant heat losses in an actual engine. However, the air-standard
analysis indicates inherently high pressures and temperatures in the Otto
cycle.
The air-standard method of analysis shows that it is essential to use a
high-compression ratio to obtain a high thermal efficiency. This relationship is
illustrated in Table 13-1, which was prepared by use of Eq. 13-8.

13-4 THE DIESEL CYCLE


A series of theoretical events based on those in an engine designed by
Rudolph Diesel is known as the Diesel cycle. Diesel injected fuel into his
engine, just before combustion, in a controlled manner so that combustion

5If 1850 kJ of heat were to be added in this case for the Carnot cycle, the total cylinder volume
would be found to be tremendous and the mep only slightly above zero.
322 Gas Cycles and Applications

(a)

(b) FIGURE 13-2 Air-standard Diesel cycle.

took place approximately at constant pressure.6 Since Diesel depended on


self-ignition of his fuel, it was essential to use a high-compression ratio to give
the necessary high air temperatures for self-ignition of the fuel. The air-
standard Diesel cycle is shown in Figure 13-2.
In the air-standard Diesel cycle, compression takes place in an isentropic
manner between states 1 and 2. Heat is added at constant pressure between
states 2 and 3. This process is followed by an isentropic expansion to state 4.
Heat is then removed at constant volume until state 1 is reached.
Since the compression is isentropic, Eq. 13-1 and Eq. 13-2 may be used to
determine the pressure and temperature at the end of compression. The
amount of heat supplied fixes the temperature rise between states 2 and 3.
Since the addition of heat is at constant pressure,

Qs = mcp(T3— T2)

6In the modern “diesel” engines, particularly those operated at high speeds, the combustion
process does not come close to a constant-pressure one. Hence, these engines are also, and more
correctly, called compression-ignition engines.
The Diesel Cycle 323

or

Also,
T
4 = (v4/v\y-'

It should be noted that the expansion ratio VJV3 in the Diesel cycle is less
than the compression ratio Vx/V2.
The basic expression for thermal efficiency applied to the Diesel cycle is

w = Qs~Qr = mcp(T3- T2)- mc„(T4- T})


Vt = 7T
Qs Qs mcp(T3— T2)
or

(13-10)

It is not possible to simplify the expression for the thermal efficiency of the
air-standard Diesel cycle in the manner shown for the air-standard Otto cycle,
since the expansion ratio in the Diesel cycle is not equal to the compression
ratio.

Example 13-2. The conditions at the start of compression in an air-standard


Diesel cycle are 103 kPa pressure and 26°C. The heat added per kilogram of
air is 1850 kJ. The compression ratio is 16 to 1. Determine the maximum
temperature and pressure, the thermal efficiency, and the mep.

Solution. Refer to Figure 13-2.


T2 = T,rk~' = (26 + 273.15) x 1604 = 906.9 K
p2 == pirk = 103 x 1614 = 4995.8 kPa = p3 Answer

Answer

v-3 v-ilv- 2 v-ilv-2 3.032

Qr = 0.718(1413.6 - 299.15) = 800.2 kJ/kg


W Qs Qr_ 1850-800.2
— = 56.75 percent Answer
v' ~ Qs Qs 1850
v-\ = 0.833 m3/kg (Example 13-1)

P.D. = ^ (0.833) = 0.7809 m3/kg


lo
w = Qs - Q, = 1850 - 800.2 = 1049.8 kJ/kg
324 Gas Cycles and Applications

From Eq. 13-9,


1049.8 x 1000
mep = 1344.3 kPa Answer
0.7809
Note that the maximum pressures and temperatures calculated here for the
Diesel cycle are lower than those for the Otto cycles of Example 13-1, in spite
of the fact that a higher compression ratio is used for the Diesel cycle. In the
Otto cycle there is a large pressure rise during the addition of heat. In the
Diesel cycle work is done during the addition of heat, thus preventing as high
a temperature rise as in the Otto cycle. Although a higher compression ratio is
used for the Diesel cycle, the efficiency of the two cycles, as may be seen by
comparing Examples 13-1 and 13-2, are substantially equal. The reason for
this is that the expansion ratio for the Diesel cycle is 5.277 to 1, whereas the
expansion ratio for the Otto cycle is 8 to 1 (the same as the compression
ratio).
Care must be taken in applying these conclusions to actual engines. In
spark-ignition engines, the combustion is far from being constant volume.
This tends to reduce the efficiency. In the so-called diesel engine, particularly
a high-speed engine, the combustion process is closer to a constant-volume
process than a constant-pressure one. This tends to increase the efficiency and
also increase the maximum cycle pressure and temperature. Thus, actual
diesel engines have higher efficiencies than do spark-ignition engines. Because
pressures are much higher in actual diesel engines, these engines must be
made heavier.

13-5 THE BRAYTON OR JOULE CYCLE


At approximately the same time, Brayton and Joule developed similar
engines. In these engines, air was compressed in a compressor and delivered
to a combustion chamber. Fuel was supplied to the combustion chamber and
caused to burn. The products of combustion passed to the expansion cylinder,
where they produced work. Part of this work was used to operate the
compressor, and the remainder was delivered as useful work. Exhaust gases
from the expander were delivered to the atmosphere. In accordance with
common practice, the ideal cycle of operation of this type of engine will be
designated as the Brayton cycle.
The air-standard Brayton cycle is shown in Figure 13-3. In this cycle, air is
first compressed in an isentropic manner (1 to 2). Heat is then added at
constant pressure (2-3). This process is followed by an isentropic expansion (3
to 4) to the initial pressure. Finally, heat is removed at constant pressure (4 to
1) until the initial state is reached. Equations 13-1 and 13-2 apply equally well
to the compression process in the Brayton cycle. The temperature at the end
of expansion is

Ti = (P3/P4)(k-|)rt
The Brayton or Joule Cycle 325

(a)

(b)

FIGURE 13-3 Air-standard Brayton cycle.

By the basic expression for thermal efficiency,

= W = Qs - Qr = mcp(T3- T2)-mcp(T4- T,)


Vt Qs Qs mcp(T3 — T2)

or

r), = 1 -14—^ (13-11)

But

Tj = (2i\(t~m = (PiYk~m = h
T4 \pj W T,
or

T,
T,
326 Gas Cycles and Applications

Hence,

or

and
T4- Ti Tx
t3 - T2 t2
Substituting in Eq. 13-11 gives
1
T2/T,
But

where r is the compression ratio. Then

(13-1 la)

Also,

where p2/pi is the compression ratio. Then,

(13-1 lb)
(pTpJ1™
Comparison of Eqs. 13-8 and 13-1 la shows that the thermal efficiencies of the
air-standard Otto and Brayton cycles are the same for the same isentropic
compression ratio. (This statement may also be applied to the Carnot cycle.)
Example 13-3. The conditions at the start of compression in an air-standard
Brayton cycle engine are 103 kPa pressure, 26°C. The heat added per kilo¬
gram of air is 1850 kJ. The compression ratio is 8 to 1. Determine the maxi¬
mum temperature and pressure, the thermal efficiency, and the mep.
Solution. Refer to Figure 13-3. The maximum pressure equals p2. From
Example 13-1,
p2 = 1893 kPa
Q = mcp(T3 — T2) Answer

1850
T3= t2 + = 687.27 + 1842.63 = 2529.90 K Answer
1 x 1.004
Stirling and Ericsson (Regenerative) Cycles 327

(T2 is obtained from Example 13-1.) From Eq. 13-1 la, the thermal efficiency

A 26 + 273 A cr An cr a^
r/t = 1 —£on = 0.5647 or 56.4/ percent Answer
687.27

In order to find the mep it is necessary to determine the cylinder specific


volume, v-4-

/0.833V2529.90\ a 3„
*3 - (t) - (s%T$ 687 ^) = °'3833 m /kg

(ui is obtained from Example 13-1.)


Ilk /t OQl\ 1/14
V-4 — /U-t, = 0.3833= 3.066 m3/kg
vP4,
0 87^
P.D. = 3.066 - = 2.962 m3/kg
8

From Eq. 13-9,

,0.5647(1850)
mep 2.962 352. / kra Answer

By comparing the mep of Example 13-3 with that of Example 13-1 it is


evident that the mep of the Brayton cycle is only a fraction of that of the Otto
cycle. This inherent low mep in the Brayton cycle is a serious drawback when
a reciprocating engine is used. Although Brayton cycle engines of the
reciprocating type were built, they were so bulky that they were impractical
and were discarded. However, the Brayton cycle became the prototype for
the simple gas turbine. Because the use of rotating machinery permits the
handling of large volumes of gases in a small volume of machinery, the low
mep of the Brayton cycle is not a serious detriment. The simple gas turbine
will be discussed in Section 13-10.
Today, consideration is being given to the use of a closed cycle Brayton
reciprocating engine, radio-isotope powered, to supply auxiliary power for
space platform needs. The inherent low mep is to be overcome by using
pressures throughout the cycle which are several times higher than those
normally encountered. Since the maximum specific volumes are only a small
fraction of normal ones, the mep will be very much higher. Hence, the power
output per unit of cylinder volume should be high.

13-6 STIRLING AND ERICSSON (REGENERATIVE) CYCLES


The thermal efficiencies discussed thus far in this chapter are much lower
than that of the Carnot cycle for a pair of given maximum and minimum cycle
temperatures. The reason for this is that for cycles other than the Carnot, heat
is added at a mean temperature much lower than the maximum cycle
temperature and is rejected at a mean temperature much higher than the
minimum cycle temperature.
328 Gas Cycles and Applications

(a)

(b)

FIGURE 13-4 Air-standard Stirling cycle.

Many individuals have investigated possible cycles to overcome this limita¬


tion. Among the cycles developed were the Stirling and Ericsson cycles. As
will be seen in the discussion of these cycles, a regeneration process is used in
both cycles. Hence, they are termed regenerative cycles.
The air-standard Stirling cycle is shown in Figure 13-4. The regenerator
needed in this cycle is shown schematically in Figure 13-5. When a hot gas is
passed to the right through the regenerator it heats up the regenerator as the
gas is cooled. At the proper part of the cycle, the gas is passed back to the left
in the regenerator, and theoretically, is heated back up to its original tem¬
perature. The operation of the cycle is so controlled that both processes
occurring in the regenerator theoretically are at constant volumes. In Figure
13-4, the air has been heated up in the regenerator to the maximum cycle
temperature at point 1. Heat is supplied from the heat source during the
constant-temperature process 1-2. The air is then forced through the
regenerator again in process 2-3, until its temperature drops to the minimum
cycle temperature. Heat is now rejected to the heat sink, process 3-4, at the
Stirling and Ericsson (Regenerative) Cycles 329

Regenerator

Step 1 —gas flow to right


Step 2—gas flow to left

FIGURE 13-5 Regeneration for Stirling cycles.

minimum cycle temperature. After heat rejection, the air is heated up to the
maximum cycle temperature, process 4-1, by being passed through the
regenerator. Because of the equality of temperatures, the heat given up by the
air as it is cooled in the regenerator is exactly equal to the heat it picks up in
passing back through the regenerator. Thus, the only exchanges of heat
between the air in the engine and the surroundings are those involving the
heat source and the heat sink. These are constant-temperature processes.
Since for a constant-temperature process of an ideal gas,

Q = W = pVln(U)

the thermal efficiency of the Stirling cycle is

Qs ~ Q, plV,ln(V2/V,)-p3V3ln(V3/V4)
7,1 Qs PiV, ln(Vj/V,)
_ mRT, ln(V2/V,) - mRT} In(V3/V4)
mRT, ln(V2/Vi)

But

V2 = V3 and V\ = V4

Then V2/ V, = V3/ V4 and

(13-12)

It should be noted that the efficiency of the Stirling cycle is exactly equal to
that of the Carnot cycle for the same maximum and minimum cycle tem¬
peratures.
The Ericsson cycle is similar to the Stirling cycle except the heat exchanges
with the regenerator take place at constant pressure. The air-standard Erics¬
son cycle is shown in Figure 13-6. Since heat is exchanged with the heat
Pressure 4 1

2
3

Volume

1 2
Temperature

Entropy

FIGURE 13-6 Air-standard Ericsson cycle.

330
Interna! Combustion Engines 331

source and the heat sink at constant temperature as it is in the Stirling cycle,
the ideal thermal efficiencies of the two cycles are equal, namely, 1- TJTH.
Many years ago, the so-called hot-air engines were designed, built, and used
extensively. Their cycles of operation were designed to approximate either
the Stirling or Ericsson cycles. In spite of the inherently high thermal
efficiency of these cycles, the actual thermal efficiencies of hot-air engines
were poor. The main reason for these low efficiencies was the fact that heat
must be transferred in all four processes through gas films on the heat-
transfer surfaces. Since gas films offer much resistance to heat transfer, the
mean cycle temperatures were very much lower than the theoretical tem¬
peratures. The resulting low thermal efficiency together with high main¬
tenance problems, high engine bulk, and the development of the spark-ignition
and the compression-ignition engines led to the disuse of hot-air engines.
However, recent engineering developments have caused a renewed interest in
Stirling cycle engines for specialized applications.

13-7 INTERNAL COMBUSTION ENGINES


In this section, the term internal combustion engine refers to an engine in
which the combustion occurs within the engine itself. This is in contrast to the
gas turbine where combustion takes place in a separate combustion chamber.
Although steam engines were used for automotive power many years ago
and interest in them has been revived recently, and although consideration is
now being given to the use of electric storage batteries and diesel engines for
automotive propulsion, the spark-ignition engine is still depended on to
furnish much of the automotive power requirements. This is true in spite of
the large amount of pollution caused by these engines in the past. The amount
of pollution from spark-ignition engine exhaust has been greatly reduced due
to government regulations.
For the larger trucks and buses, diesel engines are used almost exclusively.
Diesel engines are used extensively for marine power. Because of the growing
shortage of automotive fuel and because of their significantly higher efficien¬
cies, diesel engines now are being used extensively for automotive power.
Stationary diesel engines are used for relatively small amounts of power
generation.
At present some effort is being devoted to developing a rotary internal
combustion engine. One foreign maker has been producing this type of engine
for automotive service. American car manufacturers also are giving some
consideration to its possible use. Because of the very wide use of internal
combustion engines, consideration will be given here to the thermodynamic
problems encountered in internal combustion engines.
Internal combustion engines are classified either as two-cycle (two strokes
per cycle) or four-cycle (four strokes per cycle). They also are classified as
spark-ignition engines or compression-ignition engines. In the spark-ignition
engine, the air-fuel mixture is brought into the engine, compressed, and
332 Gas Cycles and Applications

Intake and exhaust valves

To exhaust
manifold

(c) (d)

FIGURE 13-7 Four-cycle engine, (a) Suction. (b) Compression, (c) Expansion. (d) Exhaust.

ignited by means of a spark. In the compression-ignition engine, air is brought


into the cylinder and is then so highly compressed that its temperature
exceeds the ignition temperature of the fuel. Injection of the fuel starts just
before top dead center. The high-temperature air heats the fuel up to its
self-ignition temperature and combustion starts spontaneously.
The four strokes of the four-cycle spark-ignition engine are shown in Figure
13-7. During the downward stroke, or the suction stroke, the charge of air and
fuel is drawn into the cylinder from the intake manifold. The charge is
compressed during the compression stroke. The spark then ignites a flame
which spreads throughout the combustion chamber. This is followed by the
expansion or power stroke. During the exhaust stroke, the major portion of
the combustion products is pushed out of the cylinder.
In the two-cycle engine, the suction and exhaust strokes are eliminated. To
replace the suction stroke, the charge is supplied to the cylinder under
pressure, the pressure being obtained either by crankcase compression or by a
separate blower. A two-cycle engine having crankcase compression is shown
in Figure 13-8. In Figure 13-8a, as the piston moves upward, the charge is
compressed by the upward face of the piston. At the same time, the upward
motion of the piston creates a suction in the crankcase, thus drawing in air if
it is a diesel engine or a charge for a gasoline engine. Combustion follows
(Fig. 13-8b), and the piston moves downward. The piston uncovers the
exhaust ports (Fig. 13-8c) and most of the exhaust gases rush out of the
cylinder by virtue of their higher pressure. During the downward motion of
the piston, the air (or charge) is compressed in the crankcase. Figure 13-8d
shows that the piston has uncovered the intake ports, permitting the crank¬
case contents to flow into the cylinder. This air (or charge) has a tendency to
push out much of the exhaust gases still left in the cylinder.
The actual cycle of operation of a spark-ignition engine differs greatly from
that of the air-standard Otto cycle for the following reasons:
Internal Combustion Engines 333

(a) Compression (b) Combustion (c) Exhaust (d) Scavenging


(ports closed) (ports closed) (intake port closed) and intake

Air taken into Air compressed in crankcase (Ports open)


crankcase

FIGURE 13-8 Operations of two-stroke cycle internal-combustion engine.

1 Actually an air-fuel charge is drawn into the cylinder and exhaust gases
are expelled from it.
2 There is friction between the piston and the cylinder.
3 Heat exchange takes place between the cylinder contents and the walls in
all four processes.
4 Particularly in a high-speed engine, there is much piston motion, and
hence, changes in volume during such processes as combustion, intake,
and the opening and closing of the exhaust valves.
5 In the actual engine, the cylinder contents are a real mixture, which changes
in composition throughout the cycle. In particular, the state of this real
mixture is affected by variable specific heats. The combustion process is a
complex one in which the phenomenon of dissociation is a part.

It should be evident that there are many thermodynamic problems, some of


them quite complicated, involved in the analysis of the cycle of operation of
the actual spark-ignition engine. Such an analysis is far beyond the scope of
this book.
Although the air-standard method of cycle analysis is of value in showing
inherent characteristics of various cycles, it is of little value in establishing
standards of performance, since it assumes an ideal engine having an ideal
working substance. A standard of performance of spark-ignition engines may
be established by assuming an ideal engine using an actual working substance.
In this ideal engine, there are no pressure losses through the valves, no heat
334 Gas Cycles and Applications

Table 13-2
Thermal Efficiencies, Percent

Air Percentage of Theoretical Aii• in Air-Fuel Cycle


Compression Standard
Ratio Cycle 80 100 120 200 500

4 42.5 22.0 28.5 33.3 33.3 36.3


8 56.5 30.8 39.8 41.7 45.1 48.5
12 63.1 35.3 45.7 47.5 51.1 54.2
16 67.0 38.4 49.6 51.3 54.7 57.7

exchange between the cylinder and its contents, and such events as com¬
bustion and valve openings and closings take place instantaneously. On the
other hand, an air-fuel mixture is drawn into the cylinder where it mixes with
hot clearance gases (i.e., gases left in the clearance space at the end of
exhaust). The specific heat of this mixture varies during mixing and during
compression. Dissociation, as discussed in Section 10-10, controls the tempera¬
ture reached during combustion.
By using charts entitled Thermodynamic Charts for Combustion Processes1
the effects of compression ratio and percentage of excess air on the thermal
efficiency of an ideal air-fuel cycle (i.e., an ideal engine using a real air-fuel
mixture) can be determined. The results of these determinations are shown in
Table 13-2.
The cycle of operation of a compression-ignition engine, particularly of the
high-speed type, differs greatly from the air-standard diesel cycle. One of the
main differences occurs during the combustion process. There is an ignition
lag period after injection first starts, during which time the fuel is heated up to
its ignition temperature. Then a very rapid combustion of fuel takes place,
which causes a large rise in pressure, and thus deviates greatly from the
constant-pressure addition of heat that occurs in the Diesel cycle.

13-8 INTERNAL COMBUSTION ENGINE PERFORMANCE


Previously thermal efficiency has been defined as the work delivered divided
by the heat supplied. For internal combustion engines, thermal efficiency is
defined as the work delivered divided by the heating value of the fuel. The
power delivered by the gases to the pistons is known as indicated power since
it was originally determined by means of indicators. Part of this power is used
to overcome the engine friction (the frictional horsepower). The remainder is
delivered by the engine as useful work. This is known as the brake horse-

H. C. Hottel, G. C. Williams, and G. N. Satterfield, Thermodynamic Charts for Combustion


Processes, Parts 1 and 2, Wiley, New York, 1949.
Interna! Combustion Engine Performance 335

power. Since there are two values of Work, the indicated and the brake,
there are two thermal efficiencies. The indicated thermal efficiency,

indicated work
indicated rft = (13-13)
Qhp

where QHP is the heating value of the fuel required to produce the indicated
work.
The brake thermal efficiency is

brake work
brake pt = (13-14)
Qhp

Example 13-4. The output of a compression-ignition engine is 2400 kW when


it uses 545 kg of fuel per hour. The higher heating value of the fuel at
constant pressure is 43,920 kl/kg. The frictional power of the engine is
260 kW. Determine the brake and indicated thermal efficiencies.

Solution. From Eq. 13-14,

, , 2400 x 3600 ,, ,
Answer
brake Vl = 545x437920 = 3bl perCer
(1 kW = IkJ/s)

From Eq. 13-13,

. „ , (2400 4-260)3600 _
indicated r,( = “^45^43^20~ = 40 0 percent Answer

The term mechanical efficiency is used to denote the efficiency of a mechani¬


cal device to transmit mechanical energy. In general, mechanical efficiency

mechanical output
(13-15)
mechanical input

Substituting Eq. 13-13 and 13-14 into Eq. 13-15 for an internal combustion
engine,

_ brake Tjf
(13-15a)
indicated r]t

Example 13-5. Determine the mechanical efficiency of the engine in Example


13-4.

Solution. From Eq. 13-15,

2400
Vm = 90.2 percent Answer
2400 + 260

or

36.1
Vm = 90.2 percent
40.1
336 Gas Cycles and Applications

As has been discussed previously, an important parameter in evaluating the


performance of an engine is its mep. From Eq. 13-9,

cycle work
mep = (13-16)
Lx A

where L = the length of the stroke


A = the piston area

Since the mep is related to the cycle work, it may be used to determine the
power of the engine. Thus

power = PALN (13-17)

where power = J/s = W


P = the mean effective pressure (Pa)
A = the piston area (m2)
L = the length of the stroke (m)
N = the total number of cycles per second. This equals the product of the
number of cycles per second per cylinder and the number of cylinders.

Example 13-6. An automotive engine (four-cycle) has six cylinders,


0.089 m2 x 0.114 m. The conditions at the start of compression are the same as
those in Example 13-1. The compression ratio is 8 to 1. For the air-standard
cycle, determine the power output.

Solution. Using Eq. 13-17,

..... 0.0892 x it .... 4000 .


power = 1433.2 x---x 0.114 x - - x 6

= 203.2 kW Answer

Note: The number of cycles per minute = \ revolutions per minute (rpm).

Example 13-7. The actual engine in Example 13-6 has an indicated thermal
efficiency of 75 percent of that of the air-fuel cycle, when supplied with the
theoretical amount of air. The heating value of the charge is 3722 kJ/m3. The
volumetric efficiency8 is 78 percent. Determine the power output of the
engine.

Solution. The displacement of the pistons during the suction stroke9 is

6 x 0.0892 x 77 ni1. 4000


- -x 0.114 x ——— = 8.51 m3/min
-A

4 2

8 Volumetric efficiency is the ratio of the volume actually drawn into the theoretical volume that
can be drawn in.
9In a four-cycle engine, a suction stroke occurs once every two revolutions.
Gas Compressors 337

The charge drawn in equals

8.51 x 0.78
0.1106 m3/s
60

The air-fuel cycle efficiency, Table 13-2, is 39.8 percent. Then the actual work
equals

3722 x 0.1106 x 0.398 x 0.75 = 122.9 kJ/s - 122.9 kW Answer

Another parameter that is useful in evaluating the performance of an


internal combustion engine is the torque. Torque is a turning effort. It is
defined as a force that must act on the shaft at a radius of 1 m to produce the
power delivered by the engine. From this definition,

brake power = IitNT (13-18)

where N = the rpm


T = the torque (N-m)

Example 13-8. The mechanical efficiency of the engine in Example 13-7 is 79


percent. Determine the torque.

Solution. The brake power from Example 13-7 equals

122.9 x 0.79 = 97.09 kW or 97,090 W

From Eq. 13-21,

97,090
231.8 N-m Answer
2tt x 4000/60

13-9 GAS COMPRESSORS


In the handling of gases it often becomes necessary to compress the gases.
Thermodynamics is directly involved in the compression processes. The
subject of gas compression is introduced here since compressors are an
integral part of gas turbines, which will be discussed in the next section.
However, the discussion here will cover compressors in general.
Machinery used for moving gases may be classified as fans, blowers, and
compressors. When the pressure of the delivered gas is, at the most, only a
few centimeters of water higher than that at entrance, the gas handling device is
known as a fan.
A blower builds up the gas pressure at the exit to a value of three or four
times the absolute value at the entrance. When the device builds up higher
pressures, it is known as a compressor. It should be evident that there is no
sharp line of demarcation between these three types of gas handling devices.
Compressors may be classified as reciprocating or as rotating. A rotating
compressor may be of the rotary type, the centrifugal type, or the axial-flow
type. The reciprocating compressor is well suited to compressing relatively
small volumes of gases to moderate and high pressures. A reciprocating
338 Gas Cycles and Applications

compressor is shown schematically in Figure 13-9. The valves are operated by


pressure differences. When the pressure in the cylinder drops to a value lower
than that in the suction line, the suction valve opens. Likewise, when the
pressure in the cylinder exceeds that in the discharge line, the discharge valve
opens. The ideal cycle of operation of a gas compressor without clearance is
shown in a p-V plane in Figure 13-10. As the piston of Figure 13-9 moves to
the right, gas is drawn into the left-hand side of the cylinder, 0-1. When the
piston moves on its return stroke, the gas is compressed, process 1-2, and
then discharged in process 2-3. When the piston starts again to the right, the
pressure drops immediately to that in the suction line, process 3-0, and the
cycle is repeated.
The cycle work equals the summation of the work done during each
process of the cycle, or

cycle work = W(M + WU2 + W2.3 + W3_0

= plvl + P2V12~P'V| —p2V2+0

or

cycle work = (^-j-)(p, V, - p2V2) (13-19)

FIGURE 13-10 p-V diagram for ideal-gas com¬


pressor.
Gas Compressors 339

FIGURE 13-11 p-V diagram for ideal-gas


compressor having clearance.

Since

cycle work = (^~)p,V|[l - (^) j (13-20)

In Eq. 13-20, V\ is the cylinder volume and also is the volume of gas drawn
into the cylinder. A p-V diagram for an ideal-gas compressor with clearance
is shown in Figure 13-11. At the end of the delivery of the compressed gas,
the clearance10 is filled with high-pressure gas, V3. This gas expands until its
pressure becomes equal to the suction pressure, state 0. Gas is drawn into the
cylinder between state 0 and 1. The theoretical cycle work equals VVV, 4- Wx_2
+ W2-3 + W3.4. It may be shown that this cycle work equals

i(V, - Vo)

But V] — V0 is the volume of the gas drawn into the cylinder. Thus, for an
ideal compressor, with or without clearance,
(n —l)/n
n
cycle work = (13-21)
n — 1

where V is the volume of the gas drawn into the cylinder.


Although it is sometimes assumed that the ideal compression is an iso¬
thermal one, generally the ideal compressor is assumed to have isentropic
compression. In this case, the value of n in Eq. 13-21 becomes equal to k.
Then
isentropic compressor work = (13-22)

Since no heat is transferred in an isentropic process, then

isentropic compressor work = H2- Hx = mcp(T2 — T,) (13-23)

l0The volume between the piston and the cylinder heat at the end of the stroke.
340 Gas Cycles and Applications

where T2 is the temperature at end of isentropic compression. Equations 13-22


and 13-23 are mathematically equivalent.
The work required to drive an actual compressor exceeds that given in Eq.
13-21 because of pressure drops through the valves, turbulence, unintended
heat transfer, and friction. The term internal or compression efficiency is used
to measure the degree of perfection in compressing the gas. Thus, for the
compressor,
theoretical work
internal efficiency = (13-24)
internal work

where internal work is the power delivered by the piston to the air.
The work input to the compressor, the shaft power, exceeds the work input to
the air because of mechanical frictional losses. Applying Eq. 13-15 to an air
compressor, the mechanical efficiency

indicated work
(13-25)
shaft work

Centrifugal and axial flow compressors inherently handle large volumes of


gases and develop low to moderate pressures. A four-stage centrifugal com¬
pressor is illustrated in Figure 13-12. Gas is drawn in at the center of the
shaft. Pressure is built up by the centrifugal action as the gas flows outward.
The gas also acquires a high velocity before leaving the impeller. As the gas is
slowed down in the diffuser at the exit of the impeller it acquires additional
pressure.
A sketch of an axial-flow compressor is shown in Figure 13-13. Each
moving blade imparts velocity to the gas. In the stationary-blade passages the
velocity head is converted to pressure head.
The work required to compress a given amount of a gas from fixed initial
conditions isentropically to a final pressure is independent of the mechanism
used for the compression. Hence, Eqs. 13-22 and 13-23 give the isentropic
compressor work for reciprocating compressors, centrifugal compressors,
axial-flow compressors, and rotary compressors.

Example 13-9. A centrifugal compressor receives 1990 m3/min of air at a


pressure of 101 kPa and 30°C. It discharges at a pressure of 482 kPa. The
internal efficiency is 78 percent and the mechanical efficiency is 98 percent.
Determine:

(a) Shaft power input.


(b) Temperature at the compressor exit.

Solution. The mass rate of air flow,

. pV 101x 1990/60 „ori ,


m = ... = 38.5 kg/s
RT 0.287(30 + 273.15)
Gas Compressors

•rZ7Z?//////Z22jy/////////////////??//////////////.

77ZZZZZZZZZZZZZZZZZZZZZZZZZZZZZZZZZZZZ1

£22

FIGURE 13-12 Four-stage centrifugal compressor.

The temperature at the end of isentropic compression,


0.4/1.4
/4&?\®-4^-4
T2 = (30 + 273.= 473.8 K
101

From Eq. 13-23, the isentropic power equals

38.5 x 1.004(473.8 - 303.15) = 6596 kJ/s = 6596 kW


From Eq. 13-24, the internal power equals

6596/0.78 = 8456 kW
342 Gas Cycles and Applications

FIGURE 13-13 Sketch of an axial-flow compressor.

(a) From Eq. 13-25, the shaft power equals

8456
= 8629 kW Answer
0.98

The power input to the air is the internal power and the mechanical losses
occur outside of the compressor.

internal work H2 ~
(b) internal power = Hx
rhcp(T2— Tj)
time time
8456
T2 = 30 + = 246.8° C Answer
38.5 x 1.004
Gas Turbines 343

13-10 GAS TURBINES


The gas turbine is being used increasingly in the power generating field for
peak loads, for standby service, and as topping units in conjunction with
steam turbines. Although its efficiency is relatively low, its low initial cost and
its small bulk makes it attractive for these uses. The recent development of
means of using much higher gas temperatures produces a significant im¬
provement in the efficiency of gas turbines, thus showing a greatly increasing
use of gas turbines for power generation. There are indications also of
increasing use of gas turbines for automotive service. Gas turbines are a vital
part of the propulsion systems for high-speed airplanes and for helicopters.
Figure 13-14 shows a schematic diagram of the equipment required for a
simple gas turbine. A theoretical p-V diagram is shown in Figure 13-15. The
power output of the turbine is shown shaded in Figure 13-15 b. Part of this
power output is required to operate the compressor, Figure 13-15a. The
remainder of the power developed is available for delivery by the unit, Figure
13-15 c. It should be noted that the combined diagram is similar to the air-
standard Brayton cycle in Figure 13-3. As such the ideal thermal efficiency is
given by Eq. 13-11.
Example 13-10. A theoretical gas turbine receives air at a pressure of
101 kPa, and 27°C. The maximum allowable cycle temperature is 1050°C. The
pressure ratio is 5 to 1. Determine, per kilogram of air:

(a) Turbine work.


(b) Compressor work.
(c) Net work.
(d) Thermal efficiency.
344
Pressure
Gas Cycles and Applications

(a) (b) (c)

FIGURE 13-15 Ideal gas turbine cycle, (a) Compressor. (fc>) Turbine, (c) Combined.

Solution. See Figure 13-15. The temperature at the compressor exit,


/n \(k-l)/k
T2 = Ti( —) = (27+ 273.15)504/1'4 = 475.37 K
\p,/

The temperature at the turbine exit,

r, 1050 + 273.15
-JTWO- = 835.43 K
(P3/P4)'*‘,W*

(a) Since the expansion in the ideal turbine is isentropic, the turbine work is

h3 — h4 = 1.004(1323.15 — 835.43) = 489.67 kJ/kg Answer

(b) Likewise, the compressor work is

h2- hi= 1.004(475.37-300.15)= 175.92kJ/kg Answer

(c) The net work = 489.67 - 175.92 = 313.75 kJ/kg. Answer


(d) The heat supplied equals

cp(T3-T2) = 1.004(1323.15-475.37) = 851.17 kJ/kg


W 313.75
36.86 percent Answer
7,1 Qs 851.17

As a check, from Eq. 13-1 lb.

t), = 1 — j-nTpra = 36.86 percent


Actual Gas Turbines 345

13-11 ACTUAL GAS TURBINES


In the actual gas turbine there are many losses, such as in the turbine itself
and in the compressor, pressure drops, particularly in the combustor, and heat
losses. The major losses occur in the turbine and the compressor. Losses in
the turbine mean that less power is delivered by the turbine than is indicated
in Figure 13-15fc>. Losses in the compressor require that more power be
supplied to the compressor than is indicated in Figure 13-15a. The result is
that the power output of the unit is much less than that indicated in Figure
13-15 c.

Example 13-11. Use the same conditions as for Example 13-10 but assume
identical turbine and compressor internal efficiencies1' of 80 percent and
mechanical efficiencies12 of 97 percent. Neglect the variation in specific heats
and also the difference between the mass rate of flow of the air and the mass
rate of flow of the gases.

Solution. From Eq. 13-24,

175 92
internal compressor work = — = 219.9 kJ/kg
0.8

From Eq. 13-28,

219.9
compressor shaft work = - 226.7 kJ/kg
0.97

Adapting the concept of internal efficiency to the turbine,

internal turbine work = 489.67 x 0.8 = 391.74 kJ/kg

From Eq. 13-15,


turbine shaft work = 0.97 x 391.74 = 379.98 kJ/kg

Then
net work = 379.98 - 226.7 - 153.28 kJ/kg

Because of losses within the compressor, the temperature of the air leaving
the compressor and entering the combustor is higher than the theoretical
value of 475.37 K. See Figure 13-16. Since there are no appreciable heat losses
from the compressor, the actual internal compressor work equals h2 — hj =
cp(T2~ Ti), where the subscript 2' indicates the actual property at compressor
exit. Then
rr, _ rr , internal work _ ,,AA 1. , 219.9
T2-T|+ Tp " 300 15 +L004
= 519.17 K

"compressor internal efficiency = theoretical work/compressor internal work.


12mechanical efficiency = mechanical output/mechanical input = turbine shaft work/internal
turbine work.
346 Gas Cycles and Applications

FIGURE 13-16 Gas turbine cycle.

The heat supplied equals


cp(T3 - T£) = 1.004(1323.15 - 519.17) = 807.2 kJ/kg

The thermal efficiency,

W 153.28 1QOQ . a
Vt=-Q = 8q7~2' = 18 99 percent Answer

It is to be noted that in spite of the high internal efficiencies of 80 percent for


the turbine and compressor, because of the losses in these units, the actual
thermal efficiency of the cycle is only approximately one-half that of the
theoretical gas turbine cycle. In the actual unit, the thermal efficiency will be
even less than that calculated in Example 13-11 because pressure losses and
heat losses were neglected and because variations in the specific heat were
neglected.
Because the thermal efficiency of the gas turbine unit is very sensitive to
the internal efficiencies of the turbine and the compressor, it is imperative
that these efficiencies be made as high as practical. In addition, the thermal
efficiency of the unit is quite sensitive to the maximum temperature. For
instance, if Example 13-11 were to be repeated with a maximum temperature of
650°C, it would be found that the thermal efficiency would be approximately 16
percent.
Another way to improve the thermal efficiency of the gas turbine is to use
the hot gases leaving the turbine to preheat the air before it enters the
combustor. See Figure 13-17 and also Figure 13-16. The gases leave the
Actual Gas Turbines 347

turbine at temperature T4 and enter the regenerator. Air, at temperature T '2,


leaves the compressor and also enters the regenerator. Since there is coun¬
terflow in the regenerator, in the ideal case the air can be heated up to the
temperature of the entering gases. However, this requires a regenerator of
infinite size. For the actual case, the term regenerator effectiveness is used to
describe the effectiveness of the regenerator in heating the compressed air.
When the mass rates of flow of the air and the gases do not differ significantly
and when the variations in specific heat are neglected,

~ actual temperature rise of the air


regenerator effectiveness = ~r-—r—2 ---:-7—r-—
theoretical temperature rise of the air
(13-26)

regenerator effectiveness = (13-26a)

Example 13-12. Repeat Example 13-11 using a regenerator that has a


regenerator effectiveness of 80 percent.

Solution. Neglecting the heat losses in the turbine, the decrease in enthalpy
in the turbine equals the internal turbine work. Then,

internal work 391.74


1323.15
cp 1.004
= 932.97 K

Using Eq. 13-26a, the temperature of the air leaving the regenerator,

T5 = 519.17 + 0.8(932.97 - 519.17)


= 850.21 K
348 Gas Cycles and Applications

The heat supplied in the combustor,


Qs = cp(T3-T5)= 1.004(1323.15-850.21)
= 474.83 kJ/kg
Since the conditions in the turbine and the compressor have not been
changed, the work output of the unit remains at 153.28 kJ/kg. Then the
thermal efficiency is
W 153.28 __
Answer
711 = Qs = 47483 = 32'28 perCent
This efficiency of 32.28 percent is to be compared with the efficiency of 18.99
percent when no regenerator is used.

For generation of relatively large amounts of power, compound units


frequently are used, with both the compression and expansion taking place in
two or more steps or stages. The air is cooled in between the stages of the
compressor. Fuel may be burned in between the stages of the turbine. With
these conditions, the efficiency of the gas turbine unit is significantly in¬
creased. Considerable investigational work is being done with closed-cycle
gas turbines, with the heat being supplied by either fossil fuels or nuclear
reactors. The discussion of the many interesting thermodynamic problems
involved in these various gas turbine developments, however, is beyond the
scope of this text.

13-12 JET PROPULSION


Modern high-speed aircraft depend on turbojet engines for propulsive power.
Since the propulsive force is derived from the jet itself, the aircraft is said to
be jet propelled. The elements of a turbojet engine are shown schematically in
Figure 13-18. Air enters the diffuser of the engine at a high velocity. Because

Fuel

FIGURE 13-18 Turbojet engine.


Jet Propulsion 349

of increasing area in the diffuser, there is a transformation of velocity head


into pressure head. Thus, the diffuser acts as a compressor. In the compressor
itself further compression occurs. Fuel is burned in the combustor, greatly
increasing the volume of the gases. The expansion that occurs in the turbine is
just sufficient to provide the power required to drive the compressor. Gases
leave the turbine at a moderate pressure and pass through the tail pipe to the
nozzle or jet. Expansion in the nozzle gives to the gases a high velocity and
thus provides a propulsive force which is proportional to the exit velocity.
The turbojet is quite similar to the gas turbine discussed in previous
sections but differs from it in three details. Part of the compression is
accomplished in the diffuser by diffuser action. The turbine develops only
sufficient power to drive the compressor. The useful work is made possible by
expansion in the nozzle. In the ideal case, the compression in the diffuser is
isentropic, as it is in the compressor. The combined compression process for
the ideal case is isentropic, just as it is in a gas turbine. In the ideal case, the
expansion in the nozzle is isentropic, as it is in the turbine. Thus, the overall
expansion process in the ideal case is isentropic, as it is in the gas turbine.
In the analysis that follows, it will be assumed that only ideal gases will be
involved, that the variations in specific heat may be neglected, and that the
change in the mass of gas during combustion also may be neglected. Fur¬
thermore, the analysis will be made only for an ideal turbojet engine. In this
analysis all velocities will be expressed relative to that of the engine. It makes
no difference in the analysis whether a plane is moving through still air with a
velocity say of 200 m/s or whether the engine is stationary and air enters the
engine with a velocity of 200 m/s.
If the diffuser has sufficient area at exit to reduce the air velocity to a
negligible value, the change in the enthalpy of the air in the diffuser equals the
kinetic energy of the entering air. Refer to Figures 13-18 and 13-19. For an
ideal gas, the temperature rise in the diffuser,

r? (13-27)
2cp

The pressure at the diffuser exit may be computed by using the isentropic
pressure-temperature relationship.
The action in the nozzle is the reverse of that in the diffuser. For sim¬
plification it may be assumed that the velocity at the nozzle entrance is
negligible. Assuming that the pressure at the nozzle exit is known, the
temperature, and hence, the velocity at the nozzle exit can be determined.
The propulsive force developed by the turbojet engine is

FP = m(Yn-Yd) (13-28)

where m is the mass flow rate of the air (assumed to equal the gas flow), Yn is
the gas velocity at nozzle exit, and Yd is the air velocity at diffuser entrance.
350 Gas Cycles and Applications

Example 13-13. An airplane is flying at 800 km/h at an altitude where the


pressure is 46.57 kPa and the temperature is -24.6°C. The airplane is powered
by a jet engine having a maximum cycle pressure of 280 kPa and a maximum
cycle temperature of 1090°C. The air flow is 95 kg/s. Determine the maximum
propulsive force that can be produced by the engine.

Solution. Refer to Figures 13-18 and 13-19.

air velocity at the diffuser entrance = = 222.2 m/s


3600

From Eq. 13-27,

(222.2)'
T7 = - 24.6 + = 0°C
2 x 1.004 x 1000

Then
/273 15\14/04
p2 = 46.57(|^j = 64.8 kPa
^248.55
0.4/1.4

(; = 414.94 K
t>-273 15®

compressor work = cp(T3 — T2)


= 1.004(414.94 — 273.15) = 142.36 kJ/kg

There must be sufficient expansion in the turbine to produce 142.36 kJ/kg of


Jet Propulsion 351

work. For the turbine

142.36 = cp(T4- T5) = 1.004(1090 + 273.15 - T5)


T5 = 1221.36 K
Then
j._5 \ w(k-o _/1221.36'
/1221.36\ 14,04
p5 = p4(^r I = 280 = 190.64 kPa
T4/4/ VI363.15/
'1363.15
The temperature at the nozzle exit,
(Jc-1)/Jc C7 \ 0.4/1.4
rr I Patm 1
t6 = = 1221
\190.
190^64/
= 816.52 K

The kinetic energy at the nozzle exit equals

h5-h6= cp(T5 - T6) = 1.004(1221.36 - 816.52)


= 404.46 kJ/kg

100 x 406.46 = -T r = 901.6 m/s

From Eq. 13-28, the propulsive force,


Fp = m (Tm-Td) = 95(901.6-222.2) = 64,540 N Answer
352 Gas Cycles and Applications

PROBLEMS ____
13-1 (a) Same as Example 13-1 in the text except that the compression
ratio is 12 to 1.
(b) Compare these results with those of Example 13-1.
13-2 (a) Same as Example 13-1 in the text except that the initial tem¬
perature is 70°C.
(b) Compare these results with those of Example 13-1.
13-3 Compression starts in an air-standard Otto cycle at 103 kPa and 26°C.
The compression ratio is 8 to 1. The maximum allowable pressure is
7500 kPa. Determine the mean effective pressure.
13-4 Compression starts in an air-standard Otto cycle at 103 kPa, 26°C. The
compression ratio is 8 to 1. The maximum allowable temperature is
2800 K. Determine: (a) maximum pressure and (b) mean effective
pressure.
13-5 Same as Example 13-2 in the text except the initial temperature is
70° C.
13-6 Compression starts in an air-standard Diesel cycle at 103 kPa, 26°C.
The compression ratio is 16 to 1. The maximum cycle temperature is
2200 K. Determine the thermal efficiency and the mean effective
pressure.
13-7 Same as Example 13-2 in the text except the compression ratio is 12 to
1.
13-8 Same as Example 13-3 in the text except the initial temperature is
70° C.
13-9 Compression starts in an air-standard Brayton cycle at 103 kPa and
26°C. The compression ratio is 8 to 1. The maximum temperature is
2200 K. Determine the thermal efficiency and the mean effective
pressure.
13-10 Same as Example 13-3 in the text except the compression ratio is 12 to 1.
13-11 In an air-standard Stirling cycle, heat is received at 840°C and is
rejected at 220°C. The minimum cycle pressure is 120 kPa. The ratio
of the volume at the end of addition of heat to that before addition of
heat is 3.5 to 1. Determine: (a) thermal efficiency, (b) maximum
pressure, and (c) mean effective pressure.
13-12 Determine the heat transferred per cycle in the regenerator of Prob¬
lem 13-11 per kilogram of air.
13-13 In an ideal Stirling cycle, the working fluid is helium. Heat is received
at 840°C and rejected at 220°C. The minimum cycle pressure is
120 kPa. The ratio of the volume after the end of addition of heat to
that before heat addition is 3.5 to 1. Determine: (a) thermal efficiency,
(b) maximum pressure, and (c) mean effective pressure. Compare the
results with those of Problem 13-11.
Problems 353

13-14 (a) Same as Problem 13-7 except that the maximum cycle temperature
is 1500 K.
(b) Using results from Problems 13-7 and 13-14(a), plot efficiency
against heat added. Extrapolate the curve to zero heat added. Com¬
ment on the results.
13-15 Same as Example 13-3 in the text except that the initial pressure is
400 kPa. Compare the results with those of Example 13-3.
13-16 In an air-standard Diesel cycle, the cutoff ratio (the ratio of the
volume after heat addition to that before heat addition) is 3 to 1. The
initial conditions are 101 kPa, 20°C. The compression ratio is 18 to 1.
Determine the thermal efficiency and mean effective pressure.
13-17 The thermal efficiency of an air-standard Otto cycle is 54 percent. The
mean effective pressure is 1400 kPa. The initial conditions are
100.5 kPa, 22°C. Determine: (a) maximum temperature and (b)
maximum pressure.
13-18 In an air-standard Otto cycle engine, the temperatures at the start and
end of the adiabatic expansion are 3260 K and 1420 K. The heat added
per cycle is 1850kJ/kg of air. Determine the compression ratio and the
cycle work per kilogram of air.
13-19 The compression ratio of a spark-ignition engine is 8 to 1, the in¬
dicated engine efficiency is 80 percent, and the mechanical efficiency is
76 percent when 20 percent excess air is used. Determine the brake
thermal efficiency.
13-20 Assume that the gasoline used in Problem 13-19 is equivalent in
heating value to C8Hi8. Determine the amount of fuel used per hour
when the brake power is 76 kW.
13-21 A spark-ignition engine has a compression ratio of 8 to 1. Assume that
the fuel is equivalent to C8H18. The engine uses 268 g of fuel per
kilowatt-hour. The mechanical efficiency is 78 percent. Determine the
indicated thermal efficiency.
13-22 A compression-ignition engine has an output of 3100 kW when it uses
682 kg of fuel per hour. The higher heating value of the fuel is
44,100 kJ/kg. The indicated efficiency is 40.2 percent. Determine:
(a) brake thermal efficiency and (b) frictional loss in the engine.
13-23 The charge in the cylinders of a spark-igniton engine is 0.980 bar and
40°C at the start of compression. The compression ratio is 8 to 1.
Assume the fuel is to be C8H18. The indicated engine efficiency is 80
percent when the theoretical amount of air is used. Determine the
mean effective pressure.
13-24 To make certain that the fuel in Problem 13-23 is fully vaporized at the
start of compression, the charge is heated so that its temperature at
the start of compression is 70°C. Determine the new mean effective
pressure.
354 Gas Cycles and Applications

13-25 The engine in Problem 13-23 has six cylinders and operates at
3800 rpm. It is a four-cycle engine. The bore is 9.8 cm, and the stroke
is 11.8 cm. The mechanical efficiency is 82 percent. Determine the
power output of the engine.
13-26 By use of a supercharger, the pressure in the cylinders at the start of
compression of the engine in Problem 13-23 is increased to 1.42 bars.
Determine the new mean effective pressure.
13-27 Determine the torque of the engine in Problem 13-25.
13-28 A six-cylinder, two-cycle engine with a 20.2-cm bore and a 29.8-cm
stroke has an indicated mean effective pressure of 5.94 bars and a
mechanical efficiency of 81 percent. Determine the torque at 2200 rpm.
13-29 A four-cycle compression-ignition engine uses 231 g of fuel per kilo¬
watt-hour and 22 g of air per gram of fuel when it delivers 2450 kW.
This power requires the use of a blower that receives air at 0.95 bar, 21° C
and delivers it at 1.35 bars. The theoretical power required for the blower
is 60 percent of the actual power required. (To obtain the theoretical
power of the blower, assume a rectangular p-V diagram, i.e., neglect the
change in volume of the air in the blower.) Determine the power required
for the blower.
13-30 Since a blower uses some of the power output of a supercharged
engine, will the thermal efficiency of a supercharged engine be less
than that of an engine without a supercharger? Why?
13-31 A theoretical gas turbine receives air at a pressure of 100.5 kPa, 26°C.
The pressure ratio is 6 to 1. The heat supplied per kilogram of air is
780 kJ. Determine per kilogram of air, (a) turbine work, (b)
compressor work, (c) work delivered, and (d) cycle efficiency.
13-32 Repeat Problem 13-31 for an actual turbine. Assume the internal
efficiencies of both turbine and compressor are 81 percent and the
mechanical efficiencies are both 97.5 percent. Neglect variations in
specific heats and also the difference between the mass flow of gases
and air. Assume the same maximum temperature.
13-33 Same as Problems 13-31 and 13-32 except the heat supplied per
kilogram of air is 990 kJ. Comment on the effect of the maximum
temperature on the efficiency of a gas turbine.
13-34 The fuel used in Problem 13-32 is similar to Ci2H26. (a) Determine the
mass of fuel per unit mass of air needed to provide the given amount
of heating, (b) Determine the percentage of excess air used.
13-35 Repeat Problems 13-31 and 13-32 using a pressure ratio of 9 to 1.
13-36 Determine the mass rate of flow of air for Problem 13-34 if the output
of the gas turbine unit is 12,000 kW.
13-37 A regenerator having an effectiveness of 81 percent is added to the gas
turbine of Problem 12-32. Determine the new thermal efficiency,
neglecting pressure and heat losses in the regenerator.
Problems 355

13-38 Same as Problem 13-37 but the regenerator is added to the gas turbine
in Problem 13-33.
13-39 Same as Problem 13-37 but the regenerator is added to the gas turbine
in Problem 13-35.
13-40 An airplane is flying at 900 km/h, at an altitude where the pressure is
45.2 kPa and the temperature is -25.8°C. The airplane is powered by a
jet engine having a maximum cycle pressure of 292 kPa and a maxi¬
mum cycle temperature of 1120°C. Determine for the ideal case the air
flow required to produce a propulsive force of 68,700 N.
_ 14 _

VAPOR CYCLES AND


APPLICATIONS

14-1 INTRODUCTION
The maximum thermal efficiency of a heat engine is obtained when heat is
transferred to the working substance in the engine at the source temperature
and is rejected from the engine at a temperature equal to that of the heat sink.
When the working substance is gaseous throughout the cycle, it is impractical
to hold the temperature of the working substance constant during heat
transfer. Thus, the thermal efficiencies of the gas cycles discussed in Chapter
13, such as the Otto, Diesel, and Brayton, are much lower than they should be
for the given source and sink temperatures.
On the other hand, when the pressure is held constant during a phase
change of a simple substance, the temperature remains constant. Hence, if
heat is added to a vaporizing liquid and is removed from a condensing vapor,
the desired constant-temperature heat transfer can be obtained, provided that
the pressure is held constant. A vapor-liquid cycle, commonly called a vapor
cycle, embodies these processes and, hence, has the potential for high thermal
efficiencies. Likewise, when a reversed vapor cycle is used for refrigeration,
the amount of refrigeration obtained from a given work input exceeds that
obtained by the use of most reversed gas cycles.

14-2 CARNOT VAPOR CYCLE


Because heat is added isothermally at the source temperature and is rejected
isothermally at the sink temperature, the Carnot cycle has the maximum
possible thermal efficiency for a given temperature range. Hence, the first
consideration should be given to the possibilities of this cycle when the
working substance is a vapor. A Carnot vapor cycle is shown in Figure 14-1
on the T-s plane with a line diagram being shown in Figure 14-2. The addition
of heat, process 1-2, and the rejection of heat, process 3-4, occur during phase
changes. Therefore, if the pressure is held constant during these processes,
heat will be transferred at constant temperature, as is required in the Carnot
cycle. The isentropic expansion process 2-3, may be approached in the actual
case when friction and heat transfer losses are minimized. The isentropic
compression, process 4-1, also should be approachable in the actual case.

356
Temperature

FIGURE 14-1 Carnot vapor cycle on T-s plane.

FIGURE 14-2 Carnot vapor cycle.

357
358 Vapor Cycles and Applications

However, there are two difficulties. First, the condensation must be stopped
at the right place (state 4). Second, a very wet vapor must be compressed
until, for this case, it becomes a liquid. Thus, in spite of its high thermal
efficiency, the Carnot vapor cycle is not very practical.

14-3 RANKINE VAPOR CYCLE


To eliminate the difficulties of compressing a very wet vapor, as is required in
the Carnot cycle, the vapor may be fully condensed, with the resulting liquid
being pumped into the steam generator. With this modification to the Carnot
cycle, the result is known as the Rankine cycle. A line diagram is shown for
the Rankine cycle in Figure 14-3, and Figure 14-4 is a T-s diagram. (The T-s
diagram is not to scale. In particular, the actual temperature rise in the pump
is very small.)
As with all ideal processes, the processes in the Rankine cycle are rever¬
sible ones. However, the addition of heat from the heat source to the liquid
leaving the pump to bring its temperature up to the boiling temperature is
irreversible.1 (The mean temperature of water is much lower than the source
temperature.) Hence, the thermal efficiency of the Rankine cycle is less than
that of the Carnot.
In the past, the Rankine cycle and its modifications have been used almost
exclusively in steam power plants. Therefore, steam will be used in the

‘The internal processes are reversible. The actual process of transferring heat from the sources to
the working substance is irreversible.

Heat

FIGURE 14-3 Rankine vapor cycle.


Rankine Vapor Cycle 359

FIGURE 14-4 Rankine cycle on T-s plane.

discussions of the Rankine cycle that follow. However, other fluids are now
being considered for use in the Rankine cycle in newer energy conversion
systems such as geothermal energy, ocean thermal gradients energy, and
some aspects of solar energy. The principles developed for the use of steam
apply equally well to other vapors.
The work output of the Rankine cycle is the difference between the work
output of the expander (normally a turbine) and the work input to the pump.
Since the expansion is isentropic, the turbine work,

Wt=(h2-h3)s (14-1)

In a similar manner, the work input to the pump,

Wp=(h4-h1)s (14-2)

Although tables of compressed water (Table 4 in the Steam Tables) can be


used to determine the pump work, this determination may require double
interpolation, thus causing more difficulty in determining precise results.
360 Vapor Cycles and Applications

FIGURE 14-5 Rankine cycle on p-V plane.

Particularly at low temperatures, pressures up to several hundred bars do not


produce a significant change in the specific volume of water. For example,
when saturated water at 20° C has its pressure increased to 20 MPa with its
temperature being held constant, its specific volume decreases less than 1
percent.
As shown in Figure 14-7, water is drawn into the pump, its pressure is
increased, and then it is pushed out of the pump. The shaded area equals the
work input to the pump. Since the change in volume is negligible, the work
input can be determined from

^(Phigh Plow) (14”3)

where v- is the specific volume of the water entering the pump. Applying Eq.
14-3 to the pump of Figure 14-5,

Wp = ^(p, - p4) = h{-h4 (14-3a)

Heat is supplied in the steam generator. This heat equals the difference
between the enthalpies of the steam at exit and the water at entrance. Or,
using Figures 14-3 and 14-4,

qs - h2~ h} - hi- (h4 + Wp)


= h2~ [h4 + ^4(pi -p4)] (14-4)

The thermal efficiency of the Rankine cycle,

_W _Wt-Wp_(hj-h2)s - *4(p, - p4)


Vt ds Qs fci-[h4+^4(Pi-P4)] ' '

Example 14-1. Dry saturated steam is produced at a pressure of 3.0 MPa.


The condenser pressure is 2.0 kPa. Determine and compare the cycle work
and the thermal efficiencies of the Carnot and Rankine cycles for these
conditions.
Rankine Vapor Cycle 361

Solution. Carnot (see Figures 14-1 and 14-2). The heat supplied equals

Th(s2-s,) = THSfg = (233.9+273.15X3.5412) at 3.0 MPa= 1795.6 J/g


The heat rejected equals

Tl(s3 - s4) = Tl(s2 -si) = (17.5 + 273.15)3.5412


= 1029.2 J/g

The cycle work equals

qs~qr= 1795.6 - 1029.2 - 766.4 J/g Answer

Thermal efficiency equals

w 766.4
= 42.68 percent Answer
Vt =
qs 1795.6
or
233.9- 17.5
= 42.68 percent
233.9 + 273.15

Rankine (see Figures 14-5 and 14-6).

h2 = 2802.4 s2 = 6.1869 = s3

at 2.0 kPa, 6.1869 = 8.7237 - y(8.4629)

y = 29.28 percent
h3 = 2533.5 - 0.2928(2460.0) = 1813.2 J/g

at 0.02 bars, v-f = 1.0013.


From Eq. 14-3a,

pump work = —^ (30.0 - 0.02)105 = 3.002 J/g

For this problem, the very small pump work may be neglected. Then hi = h4.
The turbine work is

Wt = hi - h2 = 2802.4 - 1813.2 = 989.2 J/g Answer

The heat supplied is

qs = h2-hl = 2802.4-73.48 - 2728.9 J/g

and the thermal efficiency is

W 989 2
rjt = — = = 36.25 percent Answer
qs LlLo.y

It should be noted that the thermal efficiency of the Rankine cycle is


approximately 85 percent of that of the Carnot. However, since much more
heat is supplied in the Rankine cycle (2728.9 J/g versus 1795.6 J/g), the cycle
work for the Rankine cycle is approximately 29 percent higher.
362 Vapor Cycles and Applications

Because the pump work does become significant when steam is generated
at very high pressures, it is desirable to illustrate how the pump work is to be
used in determining the thermal efficiency.

Example 14-2. Repeat Example 14-1 for the Rankine cycle taking into
account the pump work.

Solution. The delivered (or net) work equals the turbine work minus the
pump work. Or

Wd=Wt-Wp = 989.2 - 3.0 = 986.0 Jig

The enthalpy of the water entering the steam generator equals that leaving
the condenser plus the pump work. Or

h] = 73.48 + 3.0 = 76.48 J/g

The heat supplied is

qs = h2~h] = 2802.4 - 76.48 = 2725.9 J/g


W _ 986.0
36.17 percent Answer
qs 2725.9

The moisture content of the steam after expansion in the Rankine cycle in
Example 14-1 is almost 30 percent. Since wet steam is not a uniform mixture,
a distinct loss in efficiency will occur. There will be continuous impact
between the high-velocity steam molecules and the slower moving water
droplets, resulting in an appreciable loss in kinetic energy. In addition, there
will be mechanical difficulties including erosion in handling this very large
amount of water. These problems can be minimized by superheating the
steam before it leaves the steam generator. In addition to minimizing these
problems, in superheating the steam heat is added at a mean temperature
higher than the boiling temperature; inherently, this increases the thermal
efficiency.

Example 14-3. The conditions are the same as for the Rankine cycle in
Example 14-1 except that the steam leaving the steam generator has a
temperature of 560° C.

Solution. Refer to Figures 14-6, 14-7, and 14-8.

#12= 3591.7 s2 = 7.4022 = s3

at 2.0 kPa, 7.4022 = 8.7237 - y(8.4629)

y = 15.15 percent

h3 = 2533.5 - 0.1515 x 2460.0 = 2160.8 J/g

The turbine work is

Wt = h2-h3 = 3591.7 - 2160.8 = 1430.9 J/g


Rankine Vapor Cycle 363

FIGURE 14-6 Rankine cycle on h-s plane.

The heat supplied is

qs = h2 -hx = 3591.7 - 73.48 = 3518.2 J/g

and thermal efficiency is

W 1430.9
Vt = = 40.67 percent Answer
Qs 3518.2

The increase in the thermal efficiency from 36.25 to 40.67 percent is due solely
to the addition of some of the heat at a higher mean temperature. For the
actual case, there would be an added gain due to the reduction in the moisture
content of the steam at turbine exit. Also, because of losses within the
turbine, the moisture content of the steam at turbine exit would be much less
than the 15 percent calculated in this problem.
In the illustrative examples used thus far, the steam was generated at a
moderate pressure. In large modern plants, much higher pressures are used,
Volume

FIGURE 14-7 Feed-pump work.

364
Reheating Cycle 365

approaching the critical pressure of 22.09 MPa in many instances and exceed¬
ing it in some. For these high pressure plants, the theoretical pump work may
be approximately 1 percent of the turbine and, hence, must be taken into
account when accurate results are desired.

14-4 REHEATING CYCLE


Even with the high superheat used in Example 14-3, there is still an excessive
amount of moisture in the turbine exhaust. The amount of moisture can be
further reduced by reheating the steam after it expands part way and then
allowing it to expand down to the exhaust pressure. (The term resuperheating
is also used in this connection.)
Example 14-4. Assume the same conditions as in Example 14-3, except that
after expansion to 0.26 MPa, the steam is reheated to 560°C.
Solution. Refer to Figures 14-8, 14-9, and 14-10. As in Example 14-3,

h2 = 3591.7 s2 = 7.4022= s3

FIGURE 14-9 Reheating cycle on T-s plane.


366 Vapor Cycles and Applications

Q.
CO

c
LU

Entropy

FIGURE 14-10 Reheating cycle on h-s plane.

At 0.26 MPa, with s = 7.4022, the steam is superheated and its enthalpy is
2877.7 J/g. After resuperheating, h4 = 3615.9 and s4 = 8.5528 = s5.
At 2.0 kPa, 8.5528 = 8.7237 - y(8.4629)

y = 2.02 percent

h5 = 2533.5 - 0.0202(2460.0) = 2483.8 J/g

The total turbine work,

W, = (h2 - h3) + (h4 - h5) = (3591.7 - 2877.7) + (3615.9 - 2483.8)

= 1846.1 J/g

The total amount of heat supplied is

qs = (h2 - h0 + (h4 - h3) = (3591.7 - 73.48) + (3615.9 - 2877.7)

= 4256.4 J/g

and the thermal efficiency is

w 1846.1
= 43.37 percent
Vt =
Qs 4256.4

Not only has the moisture content at the turbine exhaust been greatly
reduced, but there is a material gain in the theoretical thermal efficiency. The
desirability of reheating the steam increases with increased initial pressures.
For very high initial pressures, it may be desirable to use two stages in
resuperheating the steam.
Regenerative Feedwater Heating 367

14-5 REGENERATIVE FEEDWATER HEATING


It was pointed out in Section 14-3 that the thermal efficiency of the Rankine
cycle is lower than that of the Carnot because the heat is added to the
feedwater at a mean temperature lower than the source temperature. The
efficiency of the Rankine cycle can be improved by heating the feedwater
regeneratively. Figure 14-11 illustrates how this may be done theoretically. In
the steam turbine the steam temperature drops progressively until it is at the
saturation temperature for the exhaust pressure. Theoretically, the conden¬
sate from the condenser is at the same temperature. In Figure 14-11, the
condensate is pumped through a heat exchanger attached to the turbine.
Theoretically, this water is heated up to the temperature of the steam entering
the turbine. This heated water is then delivered to the steam generator. Since
the heat added to the water in the steam generator is added at the maximum
temperature, the cycle efficiency equals that of the Carnot. This ideal
regenerative feedwater heating cycle is shown in Figure 14-12.

FIGURE 14-11 Ideal regenerative feedwater heating.


368 Vapor Cycles and Applications

FIGURE 14-12 T-s diagram for ideal regenerative feedwater heating.

As should be expected and as shown in Figure 14-12, the heating of the


feedwater in this manner results in an excessively high moisture content in
the steam at the turbine exit. As we discussed previously, it is essential that a
high amount of moisture in the turbine be avoided. Thus, although ther¬
modynamically desirable, this method of heating the feedwater is not feasible.
An alternate method of heating the feedwater is to use feedwater heaters.
In this method, some steam is extracted from the turbine after it has done
some work. This steam, in condensing, theoretically heats the feedwater up to
the condensing temperature. Since the temperature of the feedwater entering
the steam generator is now much higher, the mean temperature at which
heat is added in the steam generator is also higher. This results in a higher
thermal efficiency. An ideal cycle with one feedwater heater is shown in
Figure 14-13. The ideal cycle is shown on a T-s plane in Figure 14-14.

Example 14-5. The steam conditions are the same as those in Example 14-3.
Steam is extracted from the turbine at a pressure of 2.6 bars for an ideal
feedwater heater. Determine the thermal efficiency of the cycle.

Solution. Refer to Figures 14-13 and 14-14. The condition of the extracted
steam for the heater is the same as that at the exit from the higher pressure
turbine in Example 14-3; namely, p = 0.26 MPa, h = 2877.7 J/g. It will be
assumed here that there is no desuperheating zone2 in the heater and, hence,

2 A part of the heater where the superheated steam can heat the feedwater to a temperature above
the condensing steam temperature.
Regenerative Feedwater Heating 369

Heat

FIGURE 14-13 Regenerative feedwater-heating cycle.

the feedwater will be heated only to the condensing temperature correspond¬


ing to a pressure of 0.26 MPa, that is, 128.73°C. The pressures involved in this
problem are relatively low, and the effect of pressure on the enthalpy of the
liquid can be neglected unless a high degree of precision is required.
Assume 1 g of steam entering the turbine and let x equal the grams of steam
extracted. Making an energy balance around the heater of

x(h3 — h7) = (1 — x)(h8 — h6)

Neglecting the effect of pressure,

h8 = h7 = 540.90 J/g

and

h6=h5 = 73.48 J/g

Then

x(2877.7 - 540.90) = (1 - x)(540.90 - 73.48)

x =0.1667 g

1 — jc = 0.8333 g

The turbine work can be determined by making an energy balance around


the turbine. The enthalpy of the steam at the turbine exhaust is the same as in
370 Vapor Cycles and Applications

FIGURE 14-14 Regenerative feedwater-heating cycle on T-s plane. (X = lb steam to heater per
pound entering turbine.)

Example 14-3, namely, 1813.2 J/g. Per gram of steam entering the turbine,

3591.7 - 0.1667 x 2877.7 + 0.8333 x 2160.8 + Wt

Wt = 1311.4 J/g

The heated added in the steam generator is

qs = h2 - h, = 3591.7 - 540.9 = 3050.8 J/g

The thermal efficiency of the cycle is

1311.4
= 42.98 percent Answer
3050.8

A comparison of the answers of Examples 14-5 and 14-3 shows that the
efficiency of the cycle has been increased from 40.67 to 42.98 percent. This
means there is a reduction of almost 5 percent in the fuel required for a given
Additional Vapor Cycles 371

power output. For a large fossil-fuel power plant, this could result in a saving
of several millions of dollars a year in fuel costs.
For the conditions of these problems, the addition of one feedwater heater
results in a reduction of approximately 13 percent in the amount of circulating
water that must be supplied to condense the steam in the condenser. This is
particularly important because of the vast quantities of circulating water
which must be supplied (the condenser for a large turbine may require
50,000 kg or more of circulating water per second).
The gain in the thermal efficiency and the reduction in the amount of
circulating water has been shown for the addition of one feedwater heater.
Further gains can be made by the addition of more heaters. As more and more
heaters are added, the efficiency of the cycle approaches that of the Carnot.
However, as each heater is added, the gains become smaller and smaller. The
costs of the heaters, together with the complexities they introduce into the
cycle, do not justify the use of perhaps more than six feedwater heaters in a
large power plant.

14-6 ADDITIONAL VAPOR CYCLES


The resuperheating of the steam, together with the use of feedwater heating,
has resulted in a relatively high thermal efficiency for the modern power plant.
However, the use of steam as the motive vapor has limited the attainable
efficiency. Even at high pressures, water boils at a relatively low temperature.
Hence, when steam is used, much of the heat cannot be added at as high a
temperature as is desirable. Over the years several other fluids have been
considered for power plant use. Although these fluids boil at relatively high
temperatures at moderate pressures, their condensing temperatures also are
high. Thus the difference between the temperature at which heat is received
and rejected is not significantly increased, and the thermal efficiency when
using these fluids does not differ appreciably from that of steam.
To overcome this difficulty, the so-called binary vapor cycle shows prom¬
ise. Several years ago, a very few small binary vapor power plants were built.
A schematic diagram of such a plant is shown in Figure 14-15. As the
illustration shows, the mercury is vaporized, with the mercury vapor expand¬
ing in a mercury turbine and being discharged to its condenser. In this special
condenser, the mercury vapor in condensing produces steam to be used in a
regular steam turbine. Thus heat is supplied to the mercury at a high
temperature and is rejected from the steam condenser at a low temperature.
Because of the high overall temperature range, the thermal efficiency of the
binary cycle is inherently high. However, since the time when mercury
vapor-steam power plants were first introduced, increases in the temperatures
and pressures in the conventional steam power plants, together with extensive
use of feedwater heating and steam resuperheating, has brought the thermal
efficiencies of the steam plants close to that of those using mercury vapor. In
addition, the complications caused by the use of mercury, together with its
372 Vapor Cycles and Applications

limited supply, have prevented further use of mercury vapor for this purpose.
Other fluids, particularly the liquid metals such as potassium and rubidium,
have been considered as the high-temperature vapor in a binary vapor cycle.
Although they show promise, they have not, as yet, proved to be economic¬
ally feasible.
Liquid metals also show some promise for meeting auxiliary space power
demands in vapor cycles. Since heat is rejected from the condenser only by
radiation, the condensing temperature must be high to reduce the radiator
weight. To attain high power output, heat must be supplied at high tem¬
peratures.3 Some of the liquid metals fit these temperatures very well.
In certain recent developments, heat is supplied at relatively low tem¬
peratures. The resulting low thermal efficiencies necessitate large mass rates
of flow for a given power output. Since the specific volume of steam is very
large at the low pressures corresponding to the low temperatures, other fluids
are being considered for this type of operation. Some of these fluids are
ammonia, isopropane, and some of the freons. The uses considered here are
ocean thermal energy conversion (OTEC), geothermal energy from low tem¬
perature sources, and nonfocusing solar energy power generation.4 The fun¬
damental cycle for all these applications is the Rankine cycle including its
modifications. The general method of determining the power output and the
thermal efficiency is the same as that discussed for steam.

14-7 REFRIGERATION
Refrigeration is a term used to denote the maintenance of a system at a
temperature lower than that of its surroundings. Since heat has a tendency to
flow into the system, a corresponding amount of heat must be removed from
'The heat supplied here may be by radioisotopes or by solar energy.
In OTEC, use is made of the difference between the temperature of the ocean at its surface and at its
depths to generate power.
Refrigeration 373

the system to maintain its temperature. Although this removal of heat can be
accomplished in several different ways, most of the refrigeration today is
accomplished by use of the vapor-compression system.
In concept, the vapor-compression system is a reversed heat engine. Work
is put into the refrigerating system, thereby pumping heat from the refri¬
gerated space and delivering it to a higher temperature heat reservoir.
(Directly or indirectly, the atmosphere acts as the high temperature reservoir.)
Since the Carnot cycle has the highest possible thermal efficiency, the
reversed Carnot cycle should show optimum performance for producing
refrigeration. A T-s diagram of the Carnot vapor-compression system is given
in Figure 14-16. A suitable liquid is allowed to evaporate in the evaporator,
process 1-2. Heat is required for vaporization and, hence, a refrigerating
effect is produced during the vaporization. This refrigerating effect is shown
as area 1265 in Figure 14-16. The vapor leaving the evaporator enters the
compressor and, in the ideal case, is compressed isentropically to state 3. The
pressure at state 3 is sufficiently high that the vapor is condensed when heat is
removed from it by either water or air. The amount of the removed heat from
the vapor is represented by area 3456. Leaving the condenser, the refrigerant
enters an expander. In the ideal case the expansion is an isentropic process

FIGURE 14-16 Carnot vapor-compression refrigerating system.


374 Vapor Cycles and Applications

4-1. The difference between the heat given up in the condenser and the heat
picked up in the evaporator is the work input, area 1234.
The term thermal efficiency has no significance in determining the per¬
formance of a vapor-compression system. To judge this performance, the
term coefficient of performance (C.O.P.) was developed. By definition,

desired effect
C.O.P. (14-6)
work input

In the case of a vapor-compression refrigerating system, the desired effect is


the refrigeration produced. Then,

(C.O.P.), = refrlgerfng efect (14-7)


work input

Since the vapor-compression refrigerating system delivers heat to a heat


reservoir, it is also a heat pump as well as a refrigerating system. If the
desired effect of the vapor-compression system is to deliver heat, it is
designated as a heat-pump. Then, using Eq. 14-6, for a heat pump,

heating effect
(C.O.P.kp, (14-8)
work input

For the Carnot cycle, from Figure 14-17, the refrigerating effect equals
T\ow(s2 ~ Si), the heat delivered = Thigh(s2 - s}), and the work input equals
(^high~ T]ow)(s2- 5,). Then, for a Carnot cycle refrigerating system,

T|Gw(S2 S]) T,low


(C.O.P.)r (14-9)
(Thigh Tiow)(52 Si) Thigh T|low

FIGURE 14-17 Vapor-compression system.


Refrigeration 375

Similarly, for a Carnot cycle heat pump,

(C.O.P.kp = T- ThyT- (14-10)


* high * low

A comparison of Eqs. 14-9 and 14-10 shows that the coefficient of per¬
formance for the Carnot cycle is (C.O.P.)hp = (C.O. P*)r + 1.
Example 14-6. Refrigeration is desired at a temperature of — 5°C. Heat is
rejected at a temperature of 30°C. Using the Carnot cycle, determine the
coefficient of performance for both refrigeration and as a heat pump.
Solution. For refrigeration, using Eq. 14-9,

-5 + 273.15
(C.O.P.)r Answer
30-(-5)

As a heat pump, using Eq. 14-10,

30 + 273.15
(C.O.P .)h p 8.66 Answer
30-(-5)

Example 14-7. The conditions are the same as in Example 14-6 except that
refrigeration is desired at -80°C.

Solution. From Eq. 14-9,

-80 + 273.15
(C.O.P.)r 1.756 Answer
30-(-80)
From Eq. 14-10,

30 + 273.15
(C.O.P.)hp. 2.756 Answer
30-(-80)

It should be evident from the results of Examples 14-6 and 14-7, that the
coefficient of performance is very sensitive to the low temperature and to the
temperature difference. Example 14-7 shows that when refrigeration is desired
at very low temperatures or when a heat pump receives its heat at very low
temperatures, the coefficients of performance are very low. This means that
much work must be supplied to produce a desired effect. As we will discuss
shortly, the coefficients of performance for actual systems are much lower
than those for the Carnot cycle.
The coefficient of performance for the Carnot cycle refrigerating system is
the highest possible for the given temperatures. However, this system
requires an expansion cylinder which, in our example, receives a saturated
liquid and expands it into a very wet vapor. Not only does the expansion
cylinder add to the cost of the system but it also presents possible operating
difficulties. For these reasons, the Carnot vapor refrigerating cycle is not
practical.
376 Vapor Cycles and Applications

14-8 VAPOR-COMPRESSION REFRIGERATION


When a saturated liquid is throttled into a region of low pressure, a very wet
vapor is formed having a temperature much lower than the initial tem¬
perature. Thus an expansion valve provides a simple, inexpensive substitute
for the expansion cylinder used in the Carnot cycle. The line diagram of the
vapor-compression refrigerating system, having an expansion valve, is shown
in Figure 14-17. The ideal cycle is shown on the T-s plane in Figure 14-18. As
with the Carnot cycle, area 1287 represents the refrigerating effect and area
3468 represents the heat rejected to the heat reservoir. Then, area 123467
represents the heat equivalent of the work input. Note that the substitution of
the expansion valve for the expansion cylinder of the Carnot cycle decreases
the amount of refrigeration by the area 1567 and increases the work required
by the same amount.
Conditions in a given evaporator, such as the temperature, the mass rate of
refrigerant flow, and the amount of heat transfer, control the condition of the
vapor leaving the evaporator. It may leave as a wet vapor, dry saturated
vapor, or as a superheated vapor (states 2, 2', or 2" in Fig. 14-18).

FIGURE 14-18 T-s diagram for vapor-compression system (theoretical).


Vapor-Compression Refrigeration 377

FIGURE 14-19 T-s diagrams for two refrigerants, (a) Refrigerant 12 (Freon F-12). (b) Carbon
dioxide.

To compare the performance of the vapor-compression refrigerating cycle


with that of the Carnot cycle, it is necessary to select a refrigerant. Originally
carbon dioxide was used extensively. Although a “safe” refrigerant, its
performance was poor, since it operated normally near its critical state (see
Fig. 14-19). Ammonia and sulfur dioxide were also used extensively because
their performance characteristics were excellent. However, both are highly
toxic and have been largely displaced with more recently developed nontoxic
refrigerants. Although some other refrigerants were promising, none has
proved as satisfactory as those of the Freon group. These refrigerants are
compounds of fluorine. One of the first was F-12, dichlorodifluoromethane,
CC12F2, the thermodynamic properties of which are shown in the Appendix
A-5. There are at least six to eight more freon refrigerants in use today. The
saturation pressure as well as the specific volume for a given temperature
show considerable variation for the various freons. Hence, the operating
temperature range for a given type of service governs the particular type of
freon to be used.
Example 14-8. Assume the same conditions as in Example 14-6, except an
ideal vapor-compression system is used, having Freon 12 as the refrigerant.
Assume that the vapor is dry saturated at the end of compression.
Solution. Refer to Figures 14-17 and 14-18. From Table A-5 in the Appendix,

h3= 199.475 S3 — 0.6848 — s2


at -5°C,
0.6848 = 0.6986 — y(0.5736) y = 2.406 percent
h2= 185.243-0.02406(153.823) h2= 181.542 kJ/kg
378 Vapor Cycles and Applications

Then the work input equals

h3-h2 = 199.475- 181.542

= 17.933 kJ/kg

Since the process in the expansion valve is one of throttling, there is no net
change in enthalpy. Or, h{ = h4 = 64.539 kJ/kg. The refrigerating effect equals

h2~ hi= 181.542-64.539= 117.003 kJ/kg

From Eq. 14-7,

117 003
(C.O.P.)r = = 6.524 Answer

The substitution of an expansion valve for the expansion cylinder of


Example 14-6 has reduced the coefficient of performance from 7.66 to 6.524.
In practice, this loss is taken to avoid the difficulties involved in the use of an
expansion cylinder.

A unit of the amount of refrigeration produced that is still being used


widely today is the ton of refrigeration. The ton of refrigeration originally was
conceived of as being equivalent to the melting of a ton of ice in a 24-hour
period. This is equivalent to the refrigeration produced at the rate of
200 Btu/min. (12,000 Btu/h). In the SI system of units, this equals 211 kJ/min.

Example 14-9. Determine the power input theoretically required to produce


20 tons of refrigeration for the conditions of Example 14-8.

Solution

20 tons = 20 x 211 = 4220 kJ/min

From Example 14-8, the refrigeration per kilogram of refrigeration equals


117.003 kJ/kg. The mass rate of flow of refrigeration equals 4220/117.003 =
36.07 kg/min. From Example 14-8, the work input equals 17.933 kJ/kg. Then
the total input equals

36.07 x 17.933 = 646.8 kJ/min

The power input equals

= 10.78 kW (1 kW = 1 kJ/s) Answer

Alternate Solution. From Example 14-8, the (C.O.P.)r equals 6.524. For
4220 kJ of refrigeration per minute, the power input equals

4220
= 646.8 kJ/min
6.524

Example 14-10. Assume the same conditions as in Example 14-6. Determine the
power input for 20 tons of refrigeration.
Other Refrigerating Systems 379

Solution. From Example 14-6, the (C.O.P.)r equals 7.66. For 20 tons or
4220 kJ of refrigeration per minute, the power input equals 4220/7.66 =
550.9 kJ/min.

“7r“ = 9.18 kW Answer


ou

Example 14-11. The compressor of Example 14-8 has an isentropic com¬


pression efficiency (the ratio of the isentropic work to the actual shaft work)
of 74 percent. Determine the shaft power input for the given refrigerating effect
and the (C.O.P.)r.

Solution. The actual power input, from the definition of the isentropic
compression efficiency, t

, - theoretical power 10.78


shaft power =---= - - = 14.57 kw Answer
Vt 0.74

(C.O.P.)r = 14^2^60 = 4.827 Answer

Example 14-12. Determine the heating effect and the coefficient of per¬
formance as a heat pump for the conditions of Example 14-11.

Solution. Neglecting the heat losses from the compressor, the heat delivered
equals the refrigerating effect plus the work equals

4220 + 14.57 x 60 = 5094.2 kJ/min Answer


5094 2
(C.O.P.)hp, = 145^6o = 5'827 Answer

Note here that (C.O.P.)hp. = (C.O.P.)r + 1, since the heating effect equals
refrigerating effect plus work. This is not true if the power input is used as the
power input to the motor driving the compressor.

14-9 OTHER REFRIGERATING SYSTEMS


As is described in the previous section, a considerable amount of power is
required to operate the vapor-compression refrigerating system. The reason
for this is that power is necessary to compress the vapor to a sufficient
pressure so that it can be condensed by readily available cooling agents. In
the absorption system, the compressor is eliminated by using a liquid to
absorb the vapor as it leaves the evaporator (see Fig. 14-20). The resulting strong
liquor is pumped into a generator. The application of heat to the generator
vaporizes much of the refrigerant as well as some of the liquid that absorbed
the refrigerant. The vapors thus produced pass over to the condenser. It
should be observed that the absorption system is similar to the vapor-
compression system except that the compressor has been replaced by the
absorber-pump-generator combination. In the actual system, additional pieces
of equipment are added to improve the performance.
380 Vapor Cycles and Applications

Flow control >. Weak liquor


valve

FIGURE 14-20 Elements of vapor-absorption system of refrigeration.

Since the volume of the liquid to be pumped is only a very small fraction of
the vapor volume, the pump work is extremely small in comparison to the
work required to compress the vapor. However, a large amount of heat must
be supplied to the generator of the absorption system. In many instances,
waste heat from industrial processes may be used or steam may be extracted
from a steam turbine for this purpose. Under these conditions, the absorption
system may be economically sound.
For many years, most of the absorption refrigerating systems used am¬
monia as the refrigerant and water as the absorbent. At one time perhaps as
many as one-fifth of all household refrigerators were of this type, with the
heat being supplied by a gas flame. At that time, the gas required was
relatively cheap in comparison with the electricity needed. Today this is no
longer true. In addition, the danger of ammonia leakages together with the
complications of operation has eliminated the use of ammonia in the ab¬
sorption system in all except very large units. In the smaller absorption
systems, a lithium bromide-water combination is used extensively.
When large quantities of chilled water are desired, the water vapor refri¬
gerating system is used extensively. A line diagram for this system is shown
in Figure 14-21. The processes for the ideal case are shown in a T-s plane in
Figure 14-22. The water to be chilled, state 1, is sprayed into the evaporator.
Because of very low pressures in the evaporator, some of the warm water will
flash into vapor, with both the vapor and the remainder of the water reaching
the saturation temperature for the given evaporator pressure. The vapor, state
2, may be drawn into a compressor. Here it is compressed, process 2-3, to a
Moderate vacuum

Warm Compressor Condenser

t water

1
Vapor

Heat Condensate
,h A *
U pump
1I I III. .''' ll;'1 Evaporator
:.'!!!•! (high vacuum) 1
I . I

• Chilled water
■^r 5
| Pump

FIGURE 14-21 Water-vapor refrigerating system.


Temperature

Entropy

FIGURE 14-22 T-s diagram for water-vapor refrigeration (see Example 14-13).
382 Vapor Cycles and Applications

sufficiently high pressure so that it will condense in the condenser when heat
is removed from it (process 3-4). The chilled water, state 5, is removed from
the evaporator.

Example 14-13. A water vapor refrigerating system is to produce 2000 kg of


chilled water per minute at a temperature of 5°C. Warm water is supplied to
the evaporator at a temperature of 30°C. The vapor will condense in the
condenser at a temperature of 36.16°C. For the ideal case, determine the
power required to operate the compressor.

Solution. Refer to Figures 14-21 and 14-22. Making an energy balance around
the evaporator,
(hit; F mcw)t\jww tilvhv “I- mcwhfcw

where v stands for vapor


cw stands for chilled water
ww stands for warm water

Assume that the vapor leaving the evaporator is dry saturated. Then

(m v + 2000)125.9 = rfi„(2510.6) + 2000(20.98)


mv = 87.99 kg/min

The pressure in the condenser is the saturation pressure for the condensing
temperature of 36.16°C or 6.0 kPa. At the compressor entrance, h2 = 2510.6,
s2 = 9.0257 = s3. From the superheated steam tables, with p = 6.0 kPa and
s2 = 9.0257, h3 = 2827.3. Since this is the ideal case,

Compressor W = h3-h2 = 2827.3 - 2510.6 = 316.7 kJ/kg

87.99x 316.7
Compressor power = = 464.4 kW Answer
60

For normal evaporator pressures, the vapor volume is extremely large. For
Example 14-13, the specific volume of the vapor leaving the evaporator is

FIGURE 14-23 Jet compressor.


Other Refrigerating Systems 383

147,120 cm3/g. Then the volume rate of flow per minute is

V = -^T-0 X 1000 x 87.99 = 12,945 m’/min

To handle such a large volume rate of flow, it becomes necessary to use either
an axial flow or a centrifugal compressor. As an alternative, when steam is
available at moderate pressures and at a low cost, a steam jet compression
system may be desirable. As the name implies, a steam jet provides the means
of compressing the vapor. The elements of the system are shown in Figure
14-23. Steam expands in the nozzle, leaving at high velocity. In the combining
chamber it entrains the vapor coming from the evaporator and imparts
velocity to it. The mixture enters the diffuser with a high but subsonic
velocity. Since the velocity is subsonic, there will be an increase in the
pressure in the diffuser, because of the increase in area, as shown by Eq.
12-29. With proper design, the pressure will be built up in the diffuser to the
value necessary for condensation in the condenser. Because of the high
velocities used in the jet compression, it can readily handle the very large
vapor volumes leaving the evaporator.
384 Vapor Cycles and Applications

PROBLEMS _
14-1 Dry saturated steam is produced at a pressure of 4.0 MPa. The
condenser pressure is 2.5 kPa. The power to be delivered is
500,000 kW. Determine for both the Carnot and Rankine cycles (a) rate
of steam flow, (b) rate at which heat must be supplied, and (c) cycle
efficiency. Neglect the work of the boiler feed pump in the Rankine
cycle.
14-2 Repeat Problem 14-1 for the Rankine cycle with the steam being
produced at 4.0 MPa, 540°C.
14-3 (a) Determine the theoretical pump work for the Rankine cycle of
Problem 14-1.
(b) Determine the thermal efficiency taking the pump work into ac¬
count.
14-4 The same as Example 14-3 in the chapter discussion except that the
initial steam pressure is 20 MPa.
14-5 The same as Problem 14-4 except the condenser pressure is (a)
7.5 kPa, and (b) 20 kPa.
14-6 The same as Example 14-4 in the chapter discussion except after
expansion to 1.8 MPa, the steam is reheated to 560°C.
14-7 Dry saturated steam is produced at a pressure of 4.0 MPa. The
condenser pressure is 2.5 kPa. After expansion to a pressure of
0.3 MPa, sufficient steam is extracted for an ideal feedwater heater.
Assume ideal conditions in the turbine. Neglect pump work. Calculate
the thermal efficiency and compare with the efficiencies of the Carnot
and Rankine cycles of Problem 14-1.
14-8 Steam is produced at 19.5 MPa, 560°C. The condenser pressure is
2.5 kPa. Assume ideal conditions in the turbine. Neglect pump work.
Determine the thermal efficiency.
14-9 Steam is produced at 19.5 MPa, 560°C. After expansion to a pressure of
1.0 MPa, the steam is reheated to 560°C and then expands down to the
condenser pressure of 2.5 kPa. Assume ideal conditions in the turbine.
Neglect pump work. Determine the thermal efficiency.
14-10 Steam is produced at 19.5 MPa, 560°C. After expansion to a pressure
of 1.0 MPa, sufficient steam is extracted for an ideal feedwater heater.
The remainder of the steam expands down to the condenser pressure
of 2.5 kPa. Assume ideal conditions in the turbine. Neglect pump
work. Determine the thermal efficiency.
14-11 Steam is produced at 19.5 MPa, 560°C. After expansion to a pressure
of 1.0 MPa, sufficient steam is extracted for an ideal feedwater heater.
The remainder of the steam is heated to 560° C before expanding to the
condenser pressure of 2.5 kPa. Assume ideal conditions in the turbine.
Neglect pump work. Determine the thermal efficiency.
Problems 385

14-12 The same as Problem 14-9 except that the reheat pressure is 0.5 MPa.
14-13 (a) The same as Problem 14-9 except that the reheat pressure is
1.5 MPa.
(b) Compare the results of Problems 14-9, 14-12, and 14-13(a).
14-14 The same as Problem 14-10 except that the extraction pressure is
0.5 MPa.
14-15 (a) The same as Problem 14-10 except that the extraction pressure is
1.5 MPa.
(b) Compare the results of Problems 14-10, 14-14, and 14- 15(a).
14-16 The same as Problem 14-11 except that the extraction pressure is
0.5 MPa.
14-17 (a) The same as Problem 14-11 except that the extraction pressure is
1.5 MPa.
(b) Compare the results of Problems 14-11, 14-16, and 14-17(a).
14-18 (a) Steam is produced at a pressure of 4.0 MPa. The condenser
pressure is 2.5 kPa. Determine, for the ideal turbine, the initial steam
temperature. The final moisture content of the steam is not to exceed 10
percent, (b) Will the initial temperature for an actual turbine be as high as
that in part (a)? Why?
14-19 Steam is produced for an ideal turbine at a pressure of 4.0 MPa. The
condenser pressure is 2.5 kPa. The heat added per kilogram is 3500 kJ.
Neglect the pump work. Determine the thermal efficiency.
14-20 For Problem 14-8, the power output is 750,000 kW. The circulating
water (water to condense the steam in the condenser) increases 9°C in
the condenser. Determine the rate of flow of circulating water.
14-21 The same as Problem 14-20 except for the conditions of Problem 14-9.
14-22 The same as Problem 14-20 except for the conditions of Problem
14-10.
14-23 The same as Problem 14-20 except for the conditions of Problem
14-11.
14-24 (a) Determine the volume rate of steam flow to the condenser of
Problem 14-20. (b) Determine the flow area required for part (a) if the
steam has a velocity of 150 m/s.
14-25 Determine the flow area required for the circulating water of Problem
14-20 if its velocity is 1.9 m/s.
14-26 An ideal steam turbine receives steam at 4.0 MPa pressure, 400°C. The
turbine exhausts at a pressure of 0.25 MPa, the exhaust steam to be
used for heating purposes. The turbine produces 3500 kW of power.
Calculate the rate of steam flow.
14-27 Determine the initial steam temperature for Problem 14-26 if the
moisture content of the steam is not to exceed 2 percent.
386 Vapor Cycles and Applications

14-28 (a) Saturated mercury vapor is produced at a temperature of 600° C.


The mercury condenser temperature is 200°C. The rate of flow of
mercury is 500 k/s. For the ideal case determine the power output of
the turbine.
(b) If the condensing mercury produces dry saturated steam at 195°C
when receiving water at 30° C, determine the rate of steam production.
14-29 (a) The steam in Problem 14-28 condenses at 30°C. For the ideal case,
determine the power output of the steam turbine.
(b) Determine the overall thermal efficiency for the mercury vapor steam
turbine cycle of Problems 14-28 and 14-29(a).
14-30 Refrigeration is desired at a temperature of 5°C. Heat is rejected at
30°C. Using the Carnot cycle, determine the coefficient of per¬
formance for both refrigeration and as a heat pump.
14-31 The same as Problem 14-30 except heat is rejected at a temperature of
40° C.
14-32 The same as Example 14-6 in the chapter discussion. Determine the
work per cycle during expansion and also during compression. There
are 100 g of air in the cylinder. The maximum cycle volume is six
times the minimum.
14-33 Refrigeration is desired at —5°C. The refrigerant is Freon 12. The
refrigerant is dry saturated leaving the evaporator. The isentropic
compressor efficiency is 75 percent. The condensing temperature is
30°C. Assume no subcooling in the condenser. Determine the power
required per 100 kJ/s of refrigeration.
14-34 The motor driving the compressor in Problem 14-33 has an efficiency
of 91 percent when the input to the motor is 10 kW. Determine the rate
of production of refrigeration.
14-35 The same conditions as in Problem 14-34 except that the unit is to be
run for its heating effect. Determine the total heat delivered per
minute.
Note: the unit is located within the building that it is to heat.
14-36 The pressure in the evaporator of an ammonia compression system is
290.83 kPa. The condensing pressure is 1200 kPa. The vapor leaving
the evaporator is dry saturated. There is no subcooling in the conden¬
ser. The isentropic compressor efficiency is 76 percent. Determine the
power input to the compressor if the refrigerating effect is 4800 kJ/s.
14-37 Repeat Problem 14-33 assuming that the refrigerant can be cooled to
18°C before entering the expansion valve.
14-38 Since the results of Problem 14-37 show a reduction in the power
requirements when the liquid is cooled before entering the expansion
valve, it is proposed to use a counterflow heat exchanger (see Chapter
19) in which the cool vapor leaving the evaporator is used to cool the
Problems 387

liquid refrigerant down to 0°C before it enters the expansion valve, (a)
Make calculations to show whether or not this is feasible, (b) If
feasible, is this procedure to be recommended?
14-39 A water vapor refrigerating system is to produce chilled water at 5°C
when receiving warm water at 35°C. The pressure in the condenser is
8.0 kPa. The compressor will handle 12,000 m3 of vapor per minute.
Determine the rate at which chilled water is delivered from the
evaporator.
14-40 The compressor in Problem 14-39 has an isentropic compression
efficiency of 74 percent. Determine the power input to the compressor.
14-41 The power input to the compressor of a water vapor refrigerating
system is 600 kW. The isentropic compression efficiency is 73 percent.
Water is to be delivered at 5°C. The condenser pressure is 8.0 kPa.
Warm water is received by the evaporator at 35° C. Determine the rate
at which chilled water is delivered from the evaporator.
14-42 Water used to condense the water vapor in Problem 14-41 increases
8°C in temperature as it passes through the condenser. Assume no

subcooling in the condenser. Determine the rate of water flow through


the condenser.
14-43 Repeat Problem 14-39 for a chilled water temperature of 10°C.
14-44 The compressor for Problem 14-43 has an isentropic compression
efficiency of 74 percent. Determine the power input to the compressor.
14-45 (a) Determine the amount of chilled water produced per kilowatt input
for Problem 14-40. (b) The same as for part (a) except for the
conditions of Problem 14-44.
_ 15 _

KINETIC THEORY OF
GASES

15-1 INTRODUCTION
In general, up to this point, we have treated gases as if they were continuous
media. In other words, our approach has been that of macroscopic ther¬
modynamics. There have been two exceptions to this approach. In Chapter 7,
“Probability and the Nature of Entropy,” it was recognized that matter is
composed of molecules and that for fluids, these molecules are in continuous
motion. Furthermore, we established that the molecules tend to move from a
least probable state to the most probable state and, in doing so, produce an
increase in the entropy of the system. However, we made no attempt to
evaluate various molecular velocities and, hence, the energies throughout the
system. Neither did we attempt to evaluate the energy of the system as a
whole, based on microscopic thermodynamics.
In Chapter 11, “Elements of Chemical Thermodynamics,” consideration
was given to chemical reactions taking place between molecules of the
reactants. However, the original reactants were taken as continuous media, as
were the final products after the reactions were completed. No analysis was
made of the behavior of individual molecules.
In this chapter, we study the relationship between molecular velocities and
pressure and temperature. In addition, we investigate molecular velocity
distribution and its effect on temperature. This chapter provides the back¬
ground for the discussion of quantum mechanics in Chapter 16, which aids in
the determination of the energy of a system and also in the determination of
specific heats.

15-2 PRESSURE, TEMPERATURE, AND THE


KINETIC THEORY
The relationship between pressure and temperature at constant volume for¬
mulated by Charles and the relationship between pressure and volume at
constant temperature formulated by Boyle were presented in Chapter 1.
These relationships were developed from experimental observations.
However, these observations give no indication why all gases deviate from

388
Pressure, Temperature, and the Kinetic Theory 389

these relationships under certain conditions. Neither do they predict the


factors causing these deviations nor the magnitude of these deviations.
The answers to some of these questions may be obtained by using the
kinetic theory. The kinetic theory deals with the behavior of the molecules of
which matter is composed. The behavior of molecules in the solid and liquid
phase of matter is quite complex. On the other hand, the behavior of gas
molecules is much less complex. Since much of thermodynamics deals with
the gaseous phase, the discussions here will be restricted to this phase.
On the basis of experimental evidence, it may be stated that gas molecules
are in continuous motion. As a result, they possess kinetic energy. There is a
wide variation between the velocities of the individual molecules. Because
there is a continuing series of collisions between the molecules, energy is
transferred from molecule to molecule. The energy transferred produces a
change in the velocities of the individual molecules. However, if the system is
assumed to be an isolated one and at rest, there cannot be a change in the
total energy possessed by the molecules.
To further simplify the behavior of gas molecules, the following assump¬
tions may be made:

1 The molecules behave as if they were perfectly elastic spheres.


2 Since they behave as elastic spheres, there cannot be any frictional effects
between colliding molecules.
3 The mean distance between molecules is so great that intermolecular
forces can be neglected, and it may therefore be assumed that molecules
travel in straight lines.
4 The mean distance between molecules is so great that the time of impact
can be neglected in comparison with the time in free flight.
5 The size of the molecules is so small in comparison with the total volume
that it can be neglected.

A gas whose molecules obey the preceding postulates may be termed an


ideal gas. In the work that follows, attention will not be paid to each
individual molecule. Instead, the performance of a molecule will be examined,
and then the performance of the system will be formulated from the per¬
formance of an average molecule.
The relationship between the number of molecules per unit volume, the
mass of the molecules per unit volume, the mass of the molecules, and their
velocities may be established by assuming that an ideal gas is contained in a
rectangular vessel, as represented in Figure 15-1. Assume that the walls of the
vessel are perfectly smooth and perfectly elastic. Select any molecule at
random. Assume that the molecule is moving perpendicular to the YZ faces.
Since the size of the molecules is considered negligible and the intermolecular
attractions are neglected, the behavior of any given molecule is not influenced
by the other molecules present. The molecule will tend to bounce between the
390 Kinetic Theory of Gases

(a)

FIGURE 15-1 Molecular velocity components.

two YZ faces. To indicate that the molecule is moving in the X direction,


designate its velocity as Yx.
From Newton’s second law of motion,

Force = change in momentum per unit time (15-1)

The momentum equals m'Yx where m' is the molecular mass. Since the
molecule and the walls are perfectly elastic, the molecule, after striking the
wall, rebounds with a velocity numerically equal to Yx but opposite in
direction. Hence, the change in momentum per impact is 2m'Yx. After
rebounding from the YZ face, the molecule will travel a distance X before
striking the second YZ face. Again rebounding, it travels to the first YZ face.
The distance traveled per impact on a given YZ face is 2X. The number of
impacts per unit time on a given YZ face is YJ2X Substituting into Eq. 15-1,

Force = (change in momentum per impact)(impacts per unit time)


= 2m,Yx(YJ2X)
Or
m'Y2
Force = (15-la)

Since pressure equals force per unit area, the pressure exerted on the YZ
face by one molecule,
,_F_ m'Yl m'Yl
(15-2)
P A X(YZ) V "

where V = volume of vessel.


Equation 15-2 was derived on the assumption that the molecule under
consideration had a velocity only in the X direction. On the whole, the
molecules have random directions. Consider, in Figure 15-1 b, a molecule
having a velocity Y, with components Yx, Yy, and Yz. The pressure exerted by
this molecule on one YZ face is a function of the component Yx and may be
evaluated by use of Eq. 15-2.
Pressure, Temperature, and the Kinetic Theory 391

Equation 15-2 states that the pressure exerted by a molecule is proportional


to the square of its velocity perpendicular to the given plane. The total
pressure exerted on the YZ plane is

nm'Y2
Pyz — (15-3)
V
where n = the number of molecules
Y\ = average of the square of all values of Yx

Note that Yx is sometimes called the root mean square velocity.


In a similar manner,

nrn'Yi
Pxz = —y (15-4)

nm'Y]
Pxy- y (15-5)

Neglecting the effects of gravity, the pressure in the three directions, pyz, pxz,
and Pxy must be equal. Or p = pxy ~ Pxz — Pyz-
Therefore,
c^*2 _ c ^/»2 _ cy-2
(15-6)
But

Y2=Y] + Y2y + Y] or Y] = \Y2 (15-7)


Then

1 nm'Y2
(15-8)
P-3 V

The mean molecular kinetic energy equals m'T2/2. Hence,


2
p = ^ y (mean molecular kinetic energy) (15-9)

Equation 1-6 states that for a given volume, the absolute temperature is
directly proportional to the pressure. Equation 15-9 states that, for a given
amount of gas in a given volume, the pressure is directly proportional to the
mean kinetic energy of the molecules. Combining these two concepts, we see
that temperature is a measure of the mean kinetic energy of the molecules of
an ideal gas. This statement has been restricted to an ideal gas, since Eq. 15-9
was derived by making the assumptions listed earlier, which define an ideal
gas for the kinetic theory approach.
Using the relationship between temperature and mean molecular kinetic
energy, Eq. 15-9 becomes
392 Kinetic Theory of Gases

The concepts of temperature and pressure, as presented here for an ideal


gas, are of value because they show the various factors that influence these
two properties.

Example 15-1. A tank having a volume of 0.4 nr contains 581.7 g of nitrogen


at a pressure of 1.4 x 10 Pa. There are 1.25 x 10 molecules present. Deter¬
mine:

(a) The mean molecular kinetic energy.


(b) The root mean square velocity (i.e., the velocity based on the mean
molecular kinetic energy).

Solution

3pV _ 3x 1.4 x 105 x 0.4


(a) Mean K.E./molecule =
In 2 x 1.25 x 102'
= 6.72 x 1(T21 Nm Answer
Note: 1 Pa = 1 N/m\

mV2
(b) Kinetic energy =
2
or
25
6.72 x 10“21 x 1.25 x 1023 x2
2.888 x 104
581.7 x 1(T3
r = 537.4 m/s Answer

Note: 1 N = 1 kg m/s2.

15-3 TEMPERATURE AND THE ROOT MEAN


SQUARE VELOCITY
In Section 15-2, recognition was made of the fact that gases are composed of
molecules and that these molecules are in continuous motion. The pressure
exerted by a gas results from the molecular motions. A model for an ideal gas
was established by formulating five postulates (see Section 15-2). For this
model, it was shown in Eq. 15-8, that,

p V = itim'Vrms
or

Y
v rms =

where Trms is the root mean square velocity.


Since pV = NR0T,

yy rms I3NR0T
V nm'
Temperature and the Root Mean Square Velocity 393

Because the number of molecules, n, equals the product of the number of


molecules per mole, n0, and the number of moles, N,

Tms=V^
Mom
(15-12)

But R0/n0 = /c,1 the gas constant per molecule, which is also known as the
Boltzmann constant. Then

=a5_i2a)

Example 15-2. Determine the root mean square velocity of nitrogen at a


temperature of 300 K.

Solution. Using Eq. 15-12a,

_ /3 kT

where k equals the gas constant per mole divided by the number of molecules
per mole

m' = mass per mole divided by the number of molecules per mole
or
13 x 8.314 x 1000x300
516.8 m/s Answer
> 28.016

The total molecular kinetic energy = nm'°E?ms. From Eq. 1-24, Y2 = \Y2ms,
where Y2 is the root mean square velocity in the x plane. Similarly, Y2 =
1Yrms and Y2z~lY2ms. But Y2 = Y2 = Y2Z. Thus the mean molecular kinetic
energies in the three planes, nm'Y\l2, nm'Y]!2, and nm'Y\l2 are equal. Since
the total mean molecular kinetic energy equals the summation of the energies
in the three planes.
nm'Y rms
Mean molecular kinetic energy per plane = (15-13)

The molecular enetgy in Eq. 15-13 is designated as the energy of translation in


one plane. Because of the equality of energy of translation in the three planes,
the concept of equipartition of energy may be formulated; namely, the
energies of translation in the three planes are equal. The mean energy of
translation per molecule for one plane is obtained by substituting the value of
Y2ms from Eq. 15-12a into Eq. 15-13. Thus for one plane,

Mean kinetic energy per molecule = \kT (15-14)


and
• • 3
Total mean kinetic energy per molecule = (15-15)

'This k is not to be confused with the specific heat ratio, cplcv.


394 Kinetic Theory of Gases

Example 15-3. Determine the mean molecular kinetic energy per molecule
for nitrogen at 300 K.
Solution. Since there are 6.02486 x 1023 molecules per g mole,
8.314
k 6.02486 x 1023

and the total mean kinetic energy per molecule equals

(1) X (6 02486^ T(p) X 300 = 6-209 X 10 2' J pCr moiecule Answer

15-4 MAXWELL SPEED DISTRIBUTION


In Sections 15-2 and 15-3, consideration was given to the root mean square
velocity of gas molecules and to the mean kinetic energy of the molecules. In
reality, the speed of some molecules at any given instant approaches a zero
value and, for that instant, the speed of a very small number of other
molecules is exceedingly high. In this section the speed distribution of
molecules will be investigated.
The speed distribution function is credited to James Clark Maxwell,^ who
established it in 1859. Later Ludwig Boltzmann strengthened this theory with
the use of statistical mechanics.
Consider the molecules of an ideal gas that extends indefinitely in all
directions into space. At any given instant, each molecule will have its own
specific speed in a specific direction (i.e., its own specific velocity). Imagine
that it is possible to translate each molecule to the origin of axes established
at any convenient point in space. Since the gas is an ideal one, there will be no
intermolecular attractions. Nor can the molecules have a significant size.
Assume also that in the translation of the molecules to the axes that the
velocity (including both the magnitude and direction) remains unchanged. The
coordinate axes establish eight regions in space, one of which is shown in
Figure 15-2. In the following discussion it will be designated as the region.
Based on the laws of probability, discussed in Chapter 6, because of the very
large number of molecules involved, there cannot be a measurable unequal
distribution of the molecules among the eight regions. Hence, the results
obtained from the consideration of one region will apply equally well to the
other seven regions and to space in general.
Now set up a vector representing the direction of magnitude of the velocity
of each molecule in the given region. The velocity of one such molecule is
designated as Y in Figure 15-2 and has components Yx, Yy, and Yz. For this
velocity vector, Y2 = Y2 + Y2 + Y2. The specification of the end point of each
vector establishes both the magnitude and direction of each velocity vector.
Hence, it is necessary to be concerned only with these end points. Consider a

See The Scientific Papers of James Clark Maxwell, Dover Publishing Company, New York, Vol.
1, p. 377.
Maxwell Speed Distribution 395

FIGURE 15-2 Molecular velocity vectors in space.

cell at the end of vector Y in Figure 15-2, having a volume 8YX 8Yy 8YZ. As
discussed previously, under normal conditions of temperature and pressure,
even an extremely small cell contains so large a number of molecules that the
laws of probability will indicate equal distribution of molecules throughout
the cell. Thus, the cell 8YX 8Yy 8YZ may be reduced in size until it approaches
in volume the cell dYx dYy dYz without materially disturbing the laws of
probability regarding the velocity distribution.
Although only those vectors, having a velocity between Y and Y -I- dY and
the direction of Y, terminate in elementary volume dx dy dz, the x com¬
ponents of a large number of vectors terminate in the slab whose volume is
YyYz dYx. In fact, all vectors whose x component ranges from Yx to Y + dYx
must terminate in this slab, regardless of their other components. Let n equal
the number of molecules and, hence, the number of velocity vectors and end
points lying in the region under consideration. Then the number of velocity
vectors terminating in the slab YyYzYx will be dnYx and the proportion thus
terminating will be dnrJn. This fraction is a function of the location of the
slab and its thickness. Thus,

= f(yx) dYx (15-16)


n
396 Kinetic Theory of Gases

In a similar manner, slabs YxYy dYz and YXYZ dYy may be established and the
following equations derived:

nr ’lm "• (,s-|7)


and

= f(Vy) dVs (15-18)

Because of the large number of molecules involved, it may be assumed that


for each value of Yx, of Yy, and of Yz, there will be a uniform distribution of
end points throughout the slabs.
Now consider the rectangular parallelepiped formed by the intersection of
slabs YyYz dYx and YXYZ dYy, having as a base the area dYx dYy. The fraction
of the Yx component molecules with y components between Yy and Yy 4- dYy is
given as d2nrr x y
ldnYx. Because of the very large number of molecules
involved, the fractional distribution of the molecules is the same as in the slab
YXYZ dYy. Thus
d2Uyxyy _ dnyy

dtiyx it

substituting the value of dttyx from Eq. 15-16 and dnYy from Eq. 15-18,

d2nVxyy = nf(Yx) dYxf(Yy) dYy (15-19)

By the same line of reasoning, the fractional number of the molecules


having their velocity vectors terminating in the volume element dYx dYy dYz
is

d3nyxyyyz = nf(Yx)f(Yy)f(Yz) dYx dYy dYz (15-20)

The point density for this volume (i.e., the number of terminating vector
points per unit of volume) is
d tty y y
P = xy;dfydy7 = nf(rx)f(Vy)f(Yz) (15-21)

Because of the number of molecules involved, the velocity distribution is


isotropic, that is, the point density in all volume elements, dYx dYy dYz
located at a distance Y from the axis is the same and independent of the
values of Vx, Yy, and Yz.
Now select a volume element adjacent to dYxdYydYz, which is the ter¬
minus for velocity vectors that are outside the range of Y to Y + dY. The
difference in density of the elements is,

(15-22)
Maxwell Speed Distribution 397

Since, in general, f(Yy) and f(Yz) are independent of Yx, then from Eq. 15-21

dp df(Vx)
(15-23)
wx L drx
Equation 15-23 may be rewritten as

jJ~=nf'(rx)f<ry)f(Yz) (15-24)

Similar expressions may be obtained for dpldYy and dpldVz.


If the second volume element lies at the same distance as the first from the
axis, namely, V (i.e., in a segment of a spherical shell), then dp = 0. Now
substitute values of dpldYx, dpldYyy and dpjdYz into Eq. 15-22 and divide by
nf(Yx)f(Yy)f(Yz).
f\Yx) f(rz) dr2
dVx + dry + =o (15-25)
f(Vx) f(Yz)
But Yx + Yy + YZ = Y2. With Y being constant,

Yx dYx + Yy dYy + Yz dYz = 0 (15-26)

Equation 15-26 places a restraint on Eq. 15-25. In Eq. 15-26 the values of dYx,
dYy, and dYz cannot be completely independent. For example, if we arbi¬
trarily allow dYx and dYy both to be zero, dYz must also equal zero, since
Yz^ 0. However, by the use of the Lagrange method of undetermined
multipliers, Eq. 15-25 may be solved. Select an undetermined multiplier and
designate it as A. Multiply Eq. 15-26 by A and add the result to Eq. 15-25:

\t (Tx) + | drx + + \vy | dr, + rcr-j + AYr \dY7 = 0


] ]
L nrx)
ft } f(%)
(15-27)

Select a value of A to make the term [/'(^x)//(^)] + AYx - 0. This value is

v _
(15-28)
f(Yx)Yx
Then Eq. 15-27 becomes

f\rz)
dYy + + A Y7 dY7 = 0 (15-29)
L/(ry) y L f(Yz)

Any two of the quantities dYx, dYyy and dYz may be taken to be independent.
Since the coefficient of dYx in Eq. 15-27 is zero, dYx may have any desired
value. Now assume that dYz = 0 and dYy^ 0. Then from Eq. 15-29,

®+AT =0 (15-30)
f(Yy) ATy U
398 Kinetic Theory of Gases

In like manner,

fj^ + \r2 = 0 (15-31)

Substituting for f'(Yx) in the coefficient of dYx of Eq. 15-27 its value of
d[f(Yx)ldYx] and remembering that the coefficient equals zero,

1 d[f(Yx)]
+ \YX = 0
f(Yx) dYx

or
d[f(Yx)]
-A YxdYx (15-32)
f(Yx)

Integrating,

A Y2
In f(Yx) = - + In a
2
where In a is the constant of integration, or

f(Yx) = ae~b2r* (15-33)

where b2 = A/2.
Similarly,

f(ry) = ae~b2r2y (15-34)

and

/(rz) = ae~b2n (15-35)

Substituting into Eq. 15-20,

d3nTxyyrz = n dYx dYy dYz a3 exp[-b\Y2x + Y) T Y2Z)]

or

d3nyxyyvz = ndYxYyYza3e-b2y2 (15-36)


The point density is
d3ny
P = ^— = na 3 e -b2Y2 (15-37)
dYx dYy dYz

Equation 15-37 gives the number of end points per unit of volume. A very
desirable quantity is the number of molecules that have a given speed; that is,
the number of molecules that have a speed numerically between Y and
Y + dY. The terminal points of a molecular speed ranging in values numeric¬
ally equal to Y and a thickness equal to dY. The volume of this shell is given
as 4ttY2 dY. Multiplying this volume by the density as given in Eq. 15-37,

dny = nd34irY2 exp(— h2T2) dV


Evaluation of the Functions a and b 399

or

^ = 4ir a3r3e-b2r2 <15-38)

The right-hand side of Eq. 15-38 is known as the Maxwell (or the Maxwell-
Boltzmann) speed distribution function.

15-5 EVALUATION OF THE FUNCTIONS a AND b


The terms a and b must be evaluated in order that Eq. 15-38 may become
usable. First, let a be evaluated in terms of b.

= | driy = 4irnai j
n-f exp(-b2Y2)Y2 dY
Jc0 *'o

or from tables of definite integrals

3 VjT
n = 4trna (15-39)
~w

Then
3 i 3/ 3/2
a = b ITT (15-39a)

Substituting this value into Eq. 15-38,


■i
dnv 4irnb ,
V2 exp(-b2y2)=
, - 4nb3Y2
- exp(- b2V2)
. - (15-40)
dr tt212 ' Vir

The average speed of the molecules, V, is

\ Ydnr c
f = lo_= “4h3r3 exp(-h2r2) dr
n a 7T

From tables of definite integrals,

y _ Y\y_ 1
(15-41)
V77 2b4 b \/77
or

b = v=f (15‘41a)

As shown in Section 15-2, the pressure exerted by the molecules of an ideal


gas is proportional to the average of the square of the velocity of the
molecules. The root-mean-square velocity is

1/2
4 nbV4
y
v rms exp(- b2r2 dY)
V TT
400 Kinetic Theory of Gases

- = / 4bi 3Vjy'2 = 1 13
(15-42)
rms Wv Sbs ) b V2

Equating the Trms from Eqs. 15-42 and 15-12a,

1 /3 _ /3kT
b \2 ^ m'
or

(15-43)

When this value of b is substituted into Eq. 15-39a,

m
a (15-44)
2nkT

Substitution of the value of b from Eq. 15-43 into Eq. 15-40 gives

driy _ 4n f m' \ 1 y2e-(m'/2kT)r^


dr Vn^lkT'
or

(15-45)

The speed distribution function (the quantity inside the brackets on the
right-hand side of Eq. 15-45) is plotted in Figure 15-3 for a specified tem-

FIGURE 15-3 Maxwell speed distribution function.


Validity of the Maxwell Speed Distribution 401

perature as a function of Y. The crosshatched area, driyldY, in this figure


represents the fraction of the total number of molecules that have a speed
between Y and Y + dY. The fraction of the total number of molecules that
have speeds less than Y is represented by the area to the left of the
crosshatched area.
The value of the mean velocity is obtained by substituting the value of b
from Eq. 15-44 into Eq. 15-41. Then

f = 2yj—,= (15-46)
v 7im y m

The most probable velocity, that is, the velocity of the largest number of
molecules, may be obtained by differentiating the quantity within the brackets
of Eq. 15-45 with respect to Y and setting the result equal to zero. When this
is done,

l2kT
Most probable velocity = (15-47)

Example 15-4. Determine the mean velocity and the most probable velocity
of nitrogen at 300 K.

Solution. From Eq. 15-46, the mean velocity,

18.314 x 1000x300
Y = 1.596 1.596 1.596
y 28.016
= 476.2 m/s Answer

From Eq, 15-47, the most probable velocity,

12 x 8.314 x 1000x300
y 28.016
= 422 m/s Answer

It should be noted that the root mean square velocity as determined in


Example 15-2 is 516.8 m/s.

15-6 VALIDITY OF THE MAXWELL SPEED


DISTRIBUTION
Some questionable assumptions were made by Maxwell in establishing his
speed distribution function. For example, he assumed that the molecular
velocities varied from zero to infinity. However, more elaborate and rigorous
treatments of an ideal gas in a state of equilibrium lead to the verification of
Maxwell’s speed distribution function. In addition, several very carefully
made experiments have demonstrated the validity of Maxwell’s results. Some
of these experiments are described by King.3

3 Allen L. King, Thermophysics, W. H. Freeman, San Francisco, 1962, p. 167.


402 Kinetic Theory of Gases

There is a need for a knowledge of the speed distribution in an electron


gas.4 Extensive measurements have shown that electron gases at low pres¬
sures and at quasi-static equilibrium conditions behave as ideal gases and do
have an electron speed distribution that is the same as the Maxwell molecular
speed distribution.

4An electron gas is composed of free electrons rather than molecules.


Problems 403

PROBLEMS ___
15-1 Determine the temperature for Example 15-1: (a) using the ideal gas p,
V, T relationship, (b) using the root mean square velocity.
15-2 Nitrogen has a root mean square velocity of 600 m/s. Determine its
temperature.
15-3 The same as Problem 15-2 except the root mean square velocity is
1200 m/s.
15-4 Determine the kinetic energy of the molecules of nitrogen occupying a
1.5 nr container at 300 K, standard atmospheric pressure.
15-5 Determine the root mean square velocity of hydrogen at a temperature
of 300 K.
15-6 Determine the mean molecular kinetic energy per molecule of
hydrogen at a temperature of 300 K.
15-7 Plot the Maxwell velocity distribution function for nitrogen at 300 K
for velocities up to 1000 m/s.
15-8 Plot the Maxwell speed distribution function for nitrogen at 300 K of
velocities up to 1000 m/s.
15-9 Assuming a Maxwellian distribution, determine the mean velocity and
the most probable velocity of nitrogen at 400 K.
15-10 The same as Problem 15-9 except for hydrogen.
15-11 The same as Problem 15-9 except at 1000 K.
_ 16 _

ELEMENTARY
STATISTICS AND
QUANTUM MECHANICS

16-1 INTRODUCTION
Systems have been considered on a microscopic basis in both Chapter 7 and
Chapter 15. In Chapter 7, a relationship was established between entropy and
probability. In Chapter 15, the Maxwell speed distribution of molecules was
formulated. In both it was assumed that the system was an isolated one and,
at any given instant, was in a state of dynamic equilibrium. In order that the
entire behavior of the system may be determined, another constraint must be
placed on the system. The reason for this constraint is that energy is
quantized.
The concept of quantum mechanics was introduced by Max Planck. In 1900
he set out to devise a theoretical formula to describe blackbody radiation. He
found that such a formula could be derived by making a statistical evaluation
of the energy exchanged between light waves and their surroundings. He
conceived of the light that emerges from a hot body as being emitted by many
“harmonic oscillators” on the body. This hypothesis may be described as
follows.
An oscillator having a given frequency, v, emits or absorbs energy in
integral multiples of an energy unit of the size kv, h being a universal
constant. This constant, now known as the Planck constant, has a value of
6.625 x 1027 erg-seconds.1 The energy unit, hv, is known as a quantum.
Planck’s hypothesis states that the quantum is the smallest amount of energy
that may be transferred. Furthermore, the quantum is not subdivisable.
Hence, the total energy that may be transferred is the energy represented by a
whole number of quanta. Overwhelming evidence, both direct and indirect,
has established Planck’s hypothesis as a law—one of the few fundamental
laws of the physical universe.

'Or 6.625 x 1(T20 J-s.

404
Energy Levels 405

16-2 ENERGY LEVELS


When a particle receives or gives off energy, this energy will be in quantum
amounts, as discussed above. Because of a net energy exchange, the energy
of the particle will increase or decrease by one or more quanta. At any given
instant, a certain number of particles of the system will possess the same
amount of energy. These particles are said to be at the same energy level.
Similarly, the other particles of the system are grouped at other energy levels.
The total energy of the system is given as

U=2niel (16-1)

where n, is the number of particles having energies of e,.


The total energy of a particle is made up of various forms or modes. For
example, a molecule may have energy of translation in three planes, energy of
rotation about two or three axes, and atomic vibratory energy as well as
electronic energy. The magnitude of each of these energy modes is restricted
by the quantum principle. The term quantum state is used to describe the
state of a particle, including not only its energy level but also the magnitude
of each of its energy modes. When energy is exchanged by other than direct
collision, this exchange occurs through quanta that behave as elastic waves.
The Schrodinger wave equation is used to describe this exchange. From this
equation it is possible to determine the possible number of allowable quantum
states for particles.
When particles of a system, existing at a given energy level, have more than
one quantum state, the system is said to be degenerate. For those particles
existing at an energy level e„ the possible number of quantum states is
designated as g,. When the number of particles, nh having an energy level e, is
less than g„ some quantum states are not occupied at the given instant.
The degeneracy, g„ for those particles of a system existing at the energy
level e„ is dependent on the modes of energy that the particles may possess
and also on the value of €j. The effect of the value of on g, for three modes
of energy, e,^, e,2 and e,3 is shown in Table 16-1. In this table values of €,• are
assumed to be 0, 2, and 4 quanta.
The relationship between e, and g, for the conditions of Table 16-1 may be
expressed as

_ (e, + l)(e, + 2)
&« 2 (16-2)

It is evident from Table 16-1 and Eq. 16-2 that there is a very rapid increase in
the possible number of quantum states as the energy level is increased. Since
the quantum is such a small energy unit, at room temperatures the energy
level of an average particle is normally a very large number of quanta. Thus
the value of g, at room temperature may greatly exceed the number of
particles, n„ which are at the energy level e,.
406 Elementary Statistics and Quantum Mechanics

Table 16-1
Effects of Energy Level on Degeneracy (g;)a

€i = 0 6, =2 €i = 4

^il ^i2 ^i3 €i, ei2 ei2


0 0 0 1 1 0 1 1 2
gi = 1 1 0 1 1 2 1
2 0 0 1 0 3
0 1 1 1 3 0
0 2 0 2 1 1
0 0 2 2 2 0
gi =6 2 0 2
3 1 0
3 0 1
4 0 0
0 4 0
0 0 4
0 1 3
0 3 1
0 2 2
gi = 15

aThe basic energy unit is one quantum. The figures denote the number of
quanta at each quantum state.

16-3 MAXWELL-BOLTZMANN STATISTICS


The energy of a system can be evaluated by the use of Eq. 16-1 when the
number of particles, nh at each energy level, eh is known. When tt; is known
as a function of eh other properties of the system, such as entropy and
specific heat, may be determined. We pointed out in Chapter 7 that it is
possible to determine the absolute value of the entropy of substances by a
laborious step-by-step process. It will be shown subsequently that the ab¬
solute value of entropy may be determined directly by the use of statistical
thermodynamics and quantum mechanics. Specific heats were introduced and
used in Chapter 4. Empirical equations were presented to permit deter¬
mination of specific heats as a function of temperature. These equations were
designed to agree with the specific heats as determined by the use of
spectroscopic analysis combined with statistical thermodynamics and quan¬
tum mechanics.
To evaluate nh it is necessary to design a model of the system being
considered. Three different models have been devised. These models are the
Maxwell-Boltzmann (MB), the Bose-Einstein (BE), and the Fermi-Dirac (FD).
The MB model assumes that the particles of the system are distinguishable
and are distributed at the various quantum energy levels. There are no
Maxwell-Boltzmann Statistics 407

restrictions set on the number of particles that may occupy each energy level
nor on the number of quantum states at each energy level.
Consider a system of n particles and with many energy levels. For any
given macroscopic state, the possible number of microscopic states, as given
by Eq. 7-2 is

(7-2)
product of n,! Fin, !

This equation assumes equality of the states of all particles at energy level e,.
However, as we discussed previously, there may be several quantum states at
a given energy level. For a given energy level and gf quantum states, the
possible number of microscopic states is, from Eq. 7-3,

= product of g?‘ = fig"1 (7-3a)

The total possible number of microscopic states at all energy levels,

(16-3)

Equation 16-3 calls for the evaluation of n!


When a large number of molecules is involved, the evaluation of n!
becomes very involved. The factorial of a very large number may be ap¬
proximated very closely by the use of Stirling’s approximation. Since n! =
2 x 3 x 4 x 5 x (... n), then In n! = In 2 + In 3 + In 4 + In 5 + ... In n. From Figure
16-1, it may be seen that the shaded area is equal to In n\, since the width of
each rectangle is one unit. For small values of n, this area is much larger than
the area beneath the curve y = In n. With increasing values of n, the difference
between the shaded area and the area beneath the curve decreases. For very
large values of n, the difference in the two areas is very small. Thus, for large
value of n, a close approximation of n! is given as

(16-4)

FIGURE 16-1 Stirling’s approximation of In n!


408 Elementary Statistics and Quantum Mechanics

or

In n \ = n In n - n + 1 (16-4a)

When n is very large, the 1 in Eq. 16-4a may be neglected. Then

In n ! = n In n — n (16-5)

Equation 16-5 is known as Stirling’s approximation.


Applying Stirling’s approximation to Eq. 16-3,

In = In n! + 2 /i; In gi - 2 In n, ! = n In n - n + 2 nf In gf - 2 n, In n* + 2 n,

But 2 tii = n. Then

In = n In n + 2 nx In — (16-6)
Hi

There is one state of the system where ^ is a maximum. At this state


d& = 0 and also dg, = 0. In Eq. 16-6, n In n is a constant. Then d(n In n) = 0.
Then

d In = 0 + 2 rii(d In gt - d In n,-) + 2 in — dnx = 0


Yli

when 0* is at a maximum.
Since dg, = 0, d In g, = 0, and

dnx
__
d In rii •
nx d In rii = dn,

2 nx = n = C, 2 drii = 0
Then

d In ^ = 2 In (16-7)

Or

In — drii + In — dn2 + • • • In — dtt; = 0 (16-8)


nx n2 Hi

or

2 In ^ dn, = 0 (16-8a)
Hi

But

dti\ T dn2 + • • • drii = 0 (16-9)

Equation 16-9 states that dnu dn2, and so forth, cannot be independent.
This is one condition imposed on 2 dn,. A second condition is imposed on the
particles shifting from one energy level to a second one. Assuming that the
Bose-Einstein Statistics 409

system is isolated. Then

U = 'Zniei = C (16-10)

When there is a shifting of the particles,

dU = 2 6, drii = €i dri\ -I- e2 dn2 + • • • e; dn, = 0 (16-11)

Using the Lagrange method of undetermined multipliers, multiply Eq. 16-9 by


one multiplier, -In A. Multiply Eq. 16-11 by a second multiplier, -0. Add the
two new equations to Eq. 16-8a.

2 (in — - 06/ - In drii = 0 (16-12)

In Eq. 16-12, drii is not necessarily zero.


Therefore,

In — - j3e; - In A = 0 (16-13)
n,
or

Hi . — ePei
ln ix= n,A

nt = gi (16-14)

From Eq. 16-14,

_1_
n = 2n, = 2g,e -ftei (16-15)
A

The term 2 g,e 0€| in Eq. 16-15 is known as the partition function and is
designated as Z. From Eq. 16-15,

(16-16)

For a state of maximum thermodynamic probability, the number of particles


at energy level eh from Eq. 16-14, is

n, = (16-17)

16-4 BOSE-EINSTEIN STATISTICS


The model for the Bose-Einstein statistics is similar to that for the Maxwell-
Boltzmann statistics in that there are no restrictions on the number of
particles that can occupy a given energy level nor on the number of quantum
states for each energy level. The Bose-Einstein model differs from the
Maxwell-Boltzmann model in that it assumes that the particles are not
distinguishable.
410 Elementary Statistics and Quantum Mechanics

Consider two particles existing at energy level e„ which has three quantum
states (i.e., g, = 3). The possible arrangement of the particles are:

State 1 State 2 State 3


1 1 0
1 0 1
0 1 1
2 0*0
0 2 0
0 0 2

The figures denote the number of particles at a given quantum state. In this
case the possible number of arrangements is six.
Now consider three particles at energy level e„ which has four quantum
states (g, = 4). The possible arrangements of the particles are:

State 1 State 2 State 3 State 4


1 1 1 0
1 1 0 1
1 0 1 1
0 1 1 1
2 1 0 0
2 0 1 0
2 0 0 1
1 2 0 0
0 2 1 0
0 2 0 1
1 0 2 0
0 1 2 0
0 0 2 1
1 0 0 2
0 1 0 2
0 0 1 2
3 0 0 0
0 3 0 0
0 0 3 0
0 0 0 3

There are 20 possible arrangements. The possible number of arrangements of


energy level e, can be expressed as

op — M) 1)!
(16-18)
' *!,!(*,--1)!

The total possible number of arrangements, at all energy levels equals the
Fermi-Dirac Statistics 411

product of the arrangements at each energy level. Or

(g, + n, - 1)!
=n n> !(g, - 1)!
(16-19)

In ^ = 2 [ln(n, + g, - 1)! - In m! - ln(g, - 1)!] (16-20)

g, and n, may be assumed to be much larger than unity and, hence, the 1 s in
Eq. 16-20 may be neglected. Using the Stirling approximation,

In 0* = 2 [(ni + g,-) ln(n, 4- g,) - n, - g, - n, In n{ + n:- - g, In g, + g, j


(16-21)

There will be a continuous shifting of particles from one energy level to


another. Hence, nt is not a constant. However, g„ the number quantum states
at energy level e„ is a constant. For the most probable state, ^ is a maximum
and d In & = 0. Differentiating Eq. 16-21 with respect to dn, and setting the
result equal to zero,

0 = 2 In,('- +-gi> dn, (16-22)


Mi

Equations 16-9 and 16-11 apply equally well to the Bose-Einstein model. In
a manner similar to that used for the Maxwell-Boltzmann model, two
Lagrange undetermined multipliers, -In A and -(3 are used to establish the
relation:

(rij + gj)
In
tii ] In A + (3€i (16-23)

Then

gi (16-24)
Hi =
Ae^- 1

16-5 FERMI-DIRAC STATISTICS


The model for the Fermi-Dirac statistics differs from the Bose-Einstein model
in that no two particles can simultaneously occupy the same quantum state.
This statement is known as the Pauli exclusion principle. Like the Bose-
Einstein model and unlike the Maxwell-Boltzmann model, the particles of the
Fermi-Dirac model are indistinguishable from each other.
To arrive at a relationship governing the possible number of states for the
Fermi-Dirac statistics, consider two particles existing at energy level with
three quantum states at this level. (The number of quantum states is assumed
to exceed the number of particles). The total possible arrangements are:

g1 g2 g3
1 1 0
1 0 1
0 1 1
412 Elementary Statistics and Quantum Mechanics

where the 1 stands for one particle. Thus there are three possible arrange¬
ments.
For three particles at energy level €j with four quantum states, the possible
arrangements are:

gi g2 g3 g4
-4

1 i 1 0
1 0 1 1
0 1 1 1
1 1 0 1

Here there are four possible arrangements. The possible number of arrange¬
ments at energy level 6, can be expressed as:

op = gVgj} g? (16-25)
nj! n2! n,

In & — Hi In g, + n2 In g2 + • • • ni In gi — On nd + In n2! + • • • In nf!) (16-26)

or
In 0* = 2 n, In g, - 2 In Hj!

using Stirling’s approximation for In nj


In ^ = 2 [n, In g{ - n, In + n,] (16-27)

There is a shifting of particles from one energy level to another but there is no
change in g„ the number of allowable quantum states at energy level e,-. There is
one most probable state. At this state d In & = 0. Then

2 ln(^——dn, = 0 (16-28)
\ Hi /

Using Equations 16-9, 16-11, and 16-28, together with the two Lagrange
undetermined multipliers, -In A and — ]8,

"i = A^r (16'29)


16-6 EXAMINATION OF THE THREE MODELS
The Maxwell-Boltzmann model, having a large number of identical particles
that are distinguishable, has been found to describe reasonably well the
behavior of gas molecules, except for those at temperatures close to absolute
zero and also for those electrons at extremely high temperatures.
The Bose-Einstein model, having a large number of identical but indis¬
tinguishable particles, applies well to photons of radiation as well as to gas
molecules at temperatures close to absolute zero.
The Fermi-Dirac model, similar to the Bose-Einstein model except that no
two particles can occupy the same energy state, describes the behavior of
electrons in metals except at very high temperatures.
Evaluation of the Constant /3 413

The statistics for the three models can be compared by repeating the
equations for the number of molecules, n„ at the energy level e(.
p.
Maxwell-Boltzmann n, =
AeK ■

Bose-Einstein Hi =
gi
Ae*1- 1

Fermi-Dirac n, = gi
Ae^1 + 1
It is evident that when the term is very large in comparison with unity,
the Bose-Einstein and the Fermi-Dirac equations approach the Maxwell-
Boltzmann equation.

16-7 EVALUATION OF THE CONSTANT 0


Since gaseous molecules are not distinguishable, the Bose-Einstein model will
be used here rather than the Maxwell-Boltzmann model, which assumes that
the molecules are distinguishable.
From Eq. 8-8,

(16-30)

Thus, if an expression can be obtained for s in terms of /3, then it will be


possible to evaluate /3.
Equation 7-10 states that s2 — Si = k(ln <fi2 ~ In </>i). In Section 7-5, we pointed
out that the entropy of a pure crystalline solid is zero at a temperature of
absolute zero. Hence, the entropy of any substance is given by

s = k In (/> (16-31)

In Eq. 7-5, the ratio of the probability of the final state to the initial state is
given as

1
(Viiv2y
Or

<t>2l<fo = {V2lVl)n

But (V2/Vi)n is the total number of possible microscopic states for state 2,
which is (3>. Then

s = k ln0> (16-32)

where s is the absolute value of entropy per mole, k is the Boltzmann con¬
stant, and SP is the probability; that is, the possible number of microscopic
states.
414 Elementary Statistics and Quantum Mechanics

The Bose-Einstein molecular distribution, as given by Eq. 16-24, is

Hi - 1
(16-33)
gi Ae$€i 1

The logarithm of the thermodynamic probability for this distribution, as


determined from Eq. 16-21, is

ln0> = S L i- "i + g« , _ ni + gil (16-34)


Lfl‘ " rii~+ " gi J

Recognizing that gi at energy level e, is a constant but that n, is not constant,

d m^ + gln^l
L n, ® gi J

= In drii-^— dtii -4--— dn, = In + dn, (16-35)


rii rii + gi rii + gi nt

Substituting Eq. 16-35 into Eq. 16-32 and into Eq. 16-30,

^ lnr' + gA/anA (16-36)


dll Hi J\du

Using the value of ln[(n, + gi)/n,)] from Eq. 16-23,

ds
du
= k 2 (In A + fie
■>(£)
ds_' dn A ,VQ /3n,
Sin A + 2 /3e( ' (16-37)
.du du du

dllj
n = 2 nh2 =0
,du

du\ dftj
u = Eitii, du =2 ex dnh = S e, = 1
.du)

Substituting in Eq. 16-37, (ds/du)v = k(3. Equating Eq. 16-37 and Eq. 16-30,
1/T = k/3. Or

1
13 = kT (16-38)

16-8 APPLICATIONS TO INTERNAL ENERGY, ENTROPY,


AND SPECIFIC HEATS
Substituting the value of In from Eq. 16-27 into Eq. 16-32,

s = k 2 [n, In g, - n, In n, + n,] (16-39)


or

s = k 2 fn, ln^+ n,l (16-40)


Applications to Internal Energy, Entropy, and Specific Heats 415

Using the value of gj/rc,- from Eq. 16-17, Eq. 16-40 becomes

s = nk + k 2 n, (in — 4- €i ^ (16-41)
n kT /

. . , Z U
s = nk + nk In —h ~ (16-42)
n T

Equation 16-42 calls for an evaluation of the internal energy, U.

U = 2 e,n,

From Eq. 16-172 and Eq. 16-38,

rii =~gie~e'lkT

Then

—Cj/kT
U = 2 2 8ie'e (16-43)

By definition, Z = 1 g,e e‘/kT

1 — €j//cT
12 g-^e (16-44)
kT

Substituting the value of 2 g,e,e €,/kT from Eq. 16-44 into Eq. 16-43,

nkT2
U =
Z 'U

or

U = nkT2(~^) (16-45)

This value of U may now be substituted into Eq. 16-42.


Then

..•4+i%+T(i|r)] (16-46)

Expressing the amount of the substance in moles N,

s NR0 1 + In - + T (16-47)
n

In Chapter 3, values were given for the specific heats of various gases.
Included were the effects of temperature on the specific heats. Some of the data
presented were experimental. Other data were empirical formulations of the

2Here it is assumed that the term Ae^ in Eq. 16-24 is much larger than unity and, hence, this
equation reduces to Eq. 16-17.
416 Elementary Statistics and Quantum Mechanics

results obtained through spectroscopic analysis. The spectroscopic data are


combined with quantum mechanics to obtain the true specific heats. This
method is now illustrated.
The constant volume specific heat, as given by Eq. 2-10, is

(16-48)

From Eq. 16-45, on the mole basis,3

(16-49)

Although the value of cv can be determined directly from Eq. 16-49, generally
it is preferable to determine u as a function of temperature and then use Eq.
16-48.
In the determination of internal energy, entropy, and specific heat, the use
of the partition function Z, is involved. By definition, the value of Z is
affected by the particle energy, eh the total energy of the particle. If the
particle under consideration is a molecule, it may have energy of translation,
energy of rotation, and vibratory energy of its atoms as well as electronic
energy of its component atoms. Thus

e = €r + €r + €„ + ee (16-50)

where the subscript t, r, r, and e refer respectively to translation, rotation,


atomic vibration, and electronic energy.
Since there are various forms of energy, there are various forms of the
partition function. Thus Zt = 2 gte~e,lkT. The degeneracy g = gtgrgvge• Then,
assuming all forms of the energy exist,

Z = ZtZrZ,Z, (16-51)

16-9 ENERGY OF TRANSLATION AND SPECIFIC HEAT


Using the kinetic theory of gases, we gave the mean kinetic energy of
translation in Eq. 15-5 as equal to 3/2/cT. On the mole basis then,

U,=|nJ?0T (16-52)

where Ut is the energy of translation of the molecules of the system. But

C =
* N\dT)

Then

C., — 2 (16-53)

^Capital C denotes molar specific heat.


Energy of Translation and Specific Heat 417

where CVf is the portion of the mole specific heat associated with the energy
of translation of the molecules of the system.
For an ideal gas, cp = cv + R0
Then

CPt — Cv.t + Ro-2R, + Rq — 2 Ro (16-54)

The ratio of the specific heats, k = Cp/Cn


or

1.667

From Eq. 16-53,

C„, = | (8.314) = 12.471 J/gmole-K


From Eq. 16-54,

Cp, = | (8.314) = 20.785 J/g mole-K

At normal room temperatures, the experimental values of Cv of argon,


helium, and mercury vapor are 12.60, 12.48, and 12.48 J/g mole-K, respec¬
tively. Furthermore, there is no material change in their specific heats with a
change in temperature. It should be noted that these three gases are mona¬
tomic. On the basis of this experimental observation, it is apparent that
monatomic gases possess only energy of translation.
In the above derivations of the equations for the energy of translation and
the specific heats, no consideration is given to the quantum nature of energy.
Although the kinetic theory approach has given results that are close to
experimental values, the quantum approach should be investigated, since it is
known that energy is quantized.
The energy of translation in a single plane for a molecule has been estab¬
lished by the use of wave mechanics as

n2h2
(16-55)
8m'/2

where n = the quantum number (0, 1, 2, etc)


h = the Planck constant
m' = the molecular mass
l = the length of the container in the direction of the velocity component
It has been shown in quantum mechanics that the degeneracy, g„ is unity
for translating particles. The partition function for translation in one plane is

rj _ •y oo —eJkT _ ■yoc -n2h2/8m'l2kT


Zjt] — Zj0 e — Zj0 e (16-56)
418 Elementary Statistics and Quantum Mechanics

In reality, the energy difference between various groups of molecules


equals one or more quanta and, therefore, is a stepwise function. However,
because the value of the quantum is so very small in comparison with the
total energy of translation, the variation in energy between groups of mole¬
cules may be considered as a continuous function. Hence, Eq. 16-56 may be
rewritten as

Z = f e-WIUmVkT dn
(16-57)
Jo
From tables of definite integrals,

„ _(2ttm'kT)'12,
1 I (16-58)

Equations similar to Eq. 16-58 may be written for the partition function for
the other two planes of translation. But Zt| = Z,2 = Zt3 and Zt = Zt| x Zr, x Zt.
Then
'(2irm'kT)l,2f|3 _ (2wm'kT)3,2V
Z,= (16-59)
h h
where l3 = V. But V = NR0Tlp. Hence,

^ (2irm’kT)312
Zf - NR0T (16-60)

From Eq. 16-45,

(16-61)

By comparing Eq. 16-52 and Eq. 16-61, we can see that the translation
energy of molecules as formulated by means of the kinetic theory is exactly
equal to that derived by means of quantum mechanics. Since the energies
derived by the two methods are equal, the corresponding specific heats also
must be equal.

16-10 ROTATIONAL ENERGY AND SPECIFIC HEAT


Although experimental values of the specific heats of monatomic gases agree
well with the theoretical values derived in the previous section, the experi¬
mental values for other gases greatly exceed those that are calculated by
Rotational Energy and Specific Heat 419

assuming the gases possess only energy of translation. Evidently these other
gases must possess additional forms of energy. We will evaluate this ad¬
ditional amount of energy first by use of the kinetic theory and we will
compare the results thus obtained with those obtained by the use of quantum
mechanics.
Consider first a diatomic molecule. Assume that it consists of two atoms
connected by a rigid bond. Then consider that the mass of the atoms is
concentrated at their centers. (This assumption is in accord with the evidence
that the mass of the atom is concentrated in its nucleus.) From these
assumptions it may be stated that, in addition to its energy of translation, the
diatomic molecule may possess energy of rotation about two of its three
principal axes. It cannot possess any appreciable energy of rotation about
the third principal axis, since the masses are concentrated at this
axis.
Assuming that the theory of equipartition of energy is applicable to rotation
as well as translation, then the energy of rotation about one axis equals ikT.
Then the total energy of a diatomic molecule consists of the energy associated
with three modes of translation and two of rotation, or 5/2/cT. From this, for
the diatomic molecule,

(c.)w = fk (16-62)
or

(C „),, = !« o (16-62a)

and
(Cp),,r = \ Ri, (16-63)

and
fc,r = 1=1.4 (16-64)

From these equations,


(C„),.r = | (8.314) = 20.785 J/g mole-K

(Cp),.r = | (8.314) = 29.099 J/g mole-K


At room temperatures, measured values of CT and k for some diatomic
gases are:

Gas C, (J/g mole-k) k


Hydrogen 20.422 1.410
Nitrogen 20.791 1.400
Oxygen 21.054 1.397
Carbon monoxide 20.866 1.404
Nitric oxide 20.950 1.398
Chlorine 25.036 1.34
420 Elementary Statistics and Quantum Mechanics

Note that, with the exception of chlorine, the specific heats as calculated by
assuming equipartition of energy between the five degrees of freedom agree
very well with experimental values. Experimentally, it has been found that
when the diatomic gases are heated, there is an increase in the specific heat.
Later we will examine the reason for this increase in specific heat and its
magnitude, as well as the discrepancy in the specific heat of chlorine at room
temperature.
When more than two atoms per molecule are involved, the gas molecule
will have the additional energy of rotation about the third principal axis. Thus,
it will possess energy associated with six degrees of freedom. Then,

uUr = 6(ikT)

(cjt,r = 3k (16-65)

(Cv)t,r = 3Rq (16-66)

(Cp)tJ = 4R0 (16-67)


and k = 4/3
From Eq. 16-66,

(CJt.r = 24.942 J/gmole-k

From Eq. 16-67,

(Cp)u = 33.256 J/g mole-K

At room temperatures, measured values of Cv and k for some polyatomic


gases are:

Gas Cv(J/g mole-K) k


Carbon dioxide 28.163 1.298
Methane 27.112 1.307
Nitrous oxide 30.390 1.275
Sulfur dioxide 32.350 1.257
Water vapor 25.224 1.307

It is evident that, with the exception of water vapor, the measured values of
specific heats for these polyatomic gases are much higher than those predic¬
ted by assuming energies associated with six degrees of freedom. Further¬
more, as with the diatomic gases, the specific heats of these gases increase
with an increase in temperature. These variations will be examined shortly.
We can also evaluate molecular rotational energy by using the quantum
approach. Thus, the molecular rotational energy, ur, at any given rotational
energy level may be expressed in terms of the quantum number, J, where J
may be zero or an integer. By means of quantum mechanics and by simplify¬
ing assumptions, the following expression gives an excellent approximation of
Rotational Energy and Specific Heat 421

the molecular energy of rotation for linear molecules.

ur = hcBJ(J +1) (16-68)


where c = the velocity of light
B = a characteristic constant of the molecule
J = the quantum number

When the molecules are symmetrical (such as N2, CO, etc.), J may have
alternate values (0, 2, 4, etc., or 1, 3, 5, etc.). For unsymmetrical molecules, J
may have all integer values. The value of B is given as

®-8Trf <'«9>
where I = the moment of inertia of the molecule about its center of gravity.
The value of I may be determined spectroscopically or from the dimensions
of the molecule.
Combining Eqs. 16-68 and 16-69,

h2J(J + 1)
ur = (16-70)
817-I

Hence, the rotation partition function becomes

Z, = -20(2J + l)exp[ 8n,IkT J

where 2J + 1 = the statistical weight factor; a = a symmetrical number. For


symmetrical diatomic molecules, a = 2. For unsymmetrical molecules, cr =
unity, or

Zr = -1-2J(2J + l)exp[-J(J + l)y] (16-71)


(J

h2
y 8tr’lkT

At normal temperatures, say 25°C, the value of y is very small for most
diatomic gases, as is shown in the following table.

Gas f, g cm2 y

Chlorine 114.6 x lO"40 0.001174


Oxygen 19.5 x 10“" 0.00686
Carbon monoxide 14.5 x lO-40 0.00923
Nitrogen 13.9 x KT40 0.00963
Hydrogen chloride 2.66 x 10“" 0.0503
Hydrogen 0.459 x lO 40 0.292
422 Elementary Statistics and Quantum Mechanics

For small values of y, (i.e., less than 0.01), Eq. 16-71 may be written as

Zr=—[ (2J + 1) exp[-J(J + l)y] dJ (16-72)


cr Jo
Integration of Eq. 16-72 gives

Zr = — (16-72a)
cry

When the value of y exceeds 0.01 (such as for hydrogen chloride and
hydrogen), the rotational partition function may be determined by making the
summation called for in Eq. 16-71. It may appear that this will be a very
laborious process. However, as J increases, the exponential term decreases
much more rapidly than the (2J + 1) term increases. Hence, only relatively
few values of J need to be considered.
For values of y between 0.01 and 0.3, it has been shown that the rotational
partition function may be expressed as4

Zr = ^(l + 3 + lJ+'") (16-73>

Note that when y is less than 0.01, Eq. 16-73 reduces to Eq. 16-72a.
When y is small
y2
ln(l + y) = y -y+ • • •

Then

In Zr = -In y-ln<r+^ + ^Q+ - • • (16-74)

From Eq. 16-45

Id In Zr\
Ur = NRqT2 (16-75)
V dT /

When y is less than 0,01, substitute the value of Zr from Eq. 16-74 in Eq.
16-75.

U, = NR0T2(y) = NR»T (16-76)

Then

(G,)r = ( jjf ) =R„ (16-77)

Note that Eqs. 16-72 through 16-75 are not restricted to diatomic molecules
but are applicable to all linear molecules. Also note from these equations that

4J. E. Mayer and M. G. Mayer, Statistical Mechanics, Wiley, New York, 1940, p. 153.
Atomic Vibratory Energy and Specific Heat 423

when the value of y is less than 0.01, the values of rotational internal energies
and specific heats of these gases agree with those obtained by use of the
kinetic theory.
Assuming that nonlinear molecules may be treated as rigid rotators, it has
been shown that the rotational partition function is given as

8it2(8tr3IAIBIc)m(kT)m
(16-78)
crh7

where IA, IB, and Ic equal the moments of inertia with respect to the three
principal axes.
In a manner similar to that used for the linear molecules and placing a
restriction of small values on y, it has been shown that the substitution of the
value of Zr from Eq. 16-78 into Eq. 16-75 for nonlinear molecules gives

ur = ^NR0T (16-79)

and

(C„), = |j?o (16-80)

These values are substantially correct for all nonlinear polyatomic molecules
at about 25°C and are usually reliable for most permanent gases at tem¬
peratures above 100 K.

16-11 ATOMIC VIBRATORY ENERGY AND


SPECIFIC HEAT
It has been pointed out that the experimental specific heats of some gases at
room temperature exceed that predicted from a summation of the energies of
translation and rotation. Furthermore, all gases, except the monatomic ones,
have specific heats that increase with an increase in temperature. These
factors may be explained by assuming that the bonds connecting the atoms of
the molecules become elastic under certain conditions, thus permitting the
atoms to move in relation with each other.
The atoms undergoing such motion will possess both vibratory kinetic and
vibratory potential energies. The relative magnitude of these energies of a
given molecule changes with time. For a large number of molecules, statistic¬
ally speaking, the total vibratory energy is equally divided between the kinetic
and potential forms. Assuming that the theory of equipartition of energy holds
for the vibratory as well as the translational and rotational energies, the total
energy of a diatomic molecule will consist of energy of translation in three
planes, energy of rotation about two axes, and vibratory kinetic and vibratory
potential energy, or u = 1(NR0TI2) = 1I2NR0T. Then, Cw = 7/2R0. Numerically
then, Cv — 29.099 J/g mole-K. This value of C„ greatly exceeds known values of
any diatomic gas, not only at room temperatures but also at elevated tern-
424 Elementary Statistics and Quantum Mechanics

peratures. Furthermore, the concept of equipartition of energy demands that a


given molecule possess either no vibratory energy or the full amount.
Experience has shown that this is not the case at all. Hence, classical
mechanics does not describe the nature and magnitude of the vibratory
energy of molecules. On the other hand, quantum mechanics enables the
formation of an accurate description of the vibration of the atoms and of the
energy associated with the vibration.
When a molecule consists of more than two atoms, it will have more than
one mode of vibration. The total vibratory energy may be found by making a
summation of the energies of vibration of the various modes. In the dis¬
cussion that follows, the method will be established for evaluating the energy
of vibration of one mode.
The energy of a harmonic oscillator at any given quantum energy level is
given as

uv = (v+\)hcco (16-81)
where v = the quantum number with values of 0, 1, 2, 3,...
(a = the vibration frequency of the oscillator. It is the number of vibrations per
unit wave length.

Since by Eq. 16-81, the zero level vibratory energy equals {heco,
uv = u0 = vheco (16-82)

Then the vibratory partition function,

Zv = So exp[(-€„ - e0)/kT] = So exp[-(vheco/kT)] (16-83)


Let x = hcco/kT; then

Zv = So (16-84)
or
Zv = l + FUe"2x + F3xL- • (16-84a)
or
Zv = [1 — e~x]~x (16-84b)

Using values of h = 6.624 x 10 27 erg-sec, c = 2.9978 x 10'° cm/s and k =


1.3805 x 10 6 erg/deg K, x = 1.4387co/T, where co is the number of vibrations
per centimeter and T is in degrees Kelvin. Then

Zv = [1 — exp(—1.4387 to/T)]-1 (16-85)


Substituting from Eq. 16-84b into Eq. 16-45,

U - NnuT2r3{1~exp(~1 4387a>/T)r'
Atomic Vibratory Energy and Specific Heat 425

or

Uy = NR0T (16-86)

Substituting the value of x,

1.4387 (olT
NRqT T43B7WT j (16-86a)

1 / duv\ x2exR0
(CJy (16-87)
T\dT /

or

(1.4387WT)2e (1.4387o»/T)d
Kq
(C.)v 1.4387«/T (16-87a)

Example 16-1. Using a value of of 565 cm \ determine the vibratory


specific heat of chlorine at 273 K and 2000 K.

Solution. Using Eq. 16-83a, at 273 K,

1.4387W __ 1.4387(565) = „
T 273 Z^77
2.977V'967 2.9772(19.65)
(Cv)y jRo — 0.50Rq Answer
"(eT977)r (19.65-l)’2
At 2000 K,

1.4387w 1.4387(565)
0.4064
T 2000
0.4064V 4064 „ 0.40642(1.50) „
(CJV 0.988R0 Answer
(e0 4064- 1)J R° (1.501 - 1)' K°

Example 16-2. Using the value of (CU)V obtained in Example 16-1, together
with the translational and rotational specific heats as determined by the
kinetic theory, determine the value of Cn for chlorine at 273 K. Compare the
results with the value of Cu given in Section 16-8.

Solution. Using the kinetic theory, (C„)t = 3/2R0 and (Cn)r = R0. Then

— 2^0 + R() + 0.50i^0 — 3i?o

Cv = 8.314 — 24.942 J/g mole-K Answer

In Section 16-8, Cu is given as 25.036 J/g mole-K.

An examination of Eq. 16-87 shows that as 1 lx approaches zero, (C„)v also


approaches zero. When 1 lx approaches infinity, (C„)v approaches R0. From the
definition of x, it may be stated that as T approaches zero, the vibratory
specific heat also approaches zero and as T is increased, the vibratory speci-
426 Elementary Statistics and Quantum Mechanics

tie heat increases, reaching as a maximum a value of R0 when T becomes


infinity. Recalling that the vibratory energy on the macroscopic basis is half
potential and half kinetic, then the contribution of each mode to the specific
heat equals R0I2 in the limiting condition. This value is exactly equal to the
contribution made by each mode of translation and also by each mode of
rotation. Thus, in the limiting condition, the principle of equipartition of
energy holds true for vibratory atoms of molecules.
We should comment here on the significance of the equations that have
been developed. Particularly at low temperatures, the atoms of some mole¬
cules behave as if they were connected by a rigid bond and, hence, cannot
possess any vibratory energy. At room temperature this applies to sub¬
stantially all of the molecules of diatomic gases like oxygen, carbon monox¬
ide, and nitrogen. As temperatures are increased, some of the molecules
receive one or more quanta of energy and their atoms start to vibrate. At any
given temperature, the total vibratory energy of the atoms of a large group of
molecules is dependent on the statistical probability of the total number of
quanta of energy that will be received by the atoms under the given con¬
ditions. This statistical probability is recognized in Eqs. 16-86 and 16-87. As
temperatures are further increased, the atoms of more and more molecules
start vibrating, and the vibration energy levels of the various atoms are
increased. As the temperatures become very high, the energy level of the
atoms of all of the molecules approach the maximum value of kT.
To obtain some idea of the magnitude of the atomic vibratory energy,
consider oxygen. The vibration frequency of the atoms of a molecule of
oxygen is 1580 per centimeter. This means that at a temperature of 273 K, the
vibratory specific heat of oxygen equals 0.017jR0- Since the specific heat of
oxygen due to translation and rotation equals 2.5R0, it is evident that the
vibratory specific heat of oxygen at room temperature is less than 0.7 percent
of the total specific heat. The vibration frequencies of other diatomic gases
such as hydrogen, nitrogen, and carbon dioxide exceed that of oxygen and,
therefore, their vibratory specific heats are negligible at room temperatures.
On the other hand, the vibration frequencies of heavier diatomics, such as
chloride and iodine, are much lower than that of oxygen and, hence, their
vibratory specific heats are appreciable even at room temperature.
As stated, when the molecule consists of more than two atoms, there will
be more than one mode of vibration. For linear molecules, the number of
modes of vibration equals 3n — 5, where n is the number of atoms per
molecule. For nonlinear molecules, the number of modes is 3n -6. Thus, it is
necessary to know the molecular structure of a gas before its vibratory energy
can be determined. By analyzing the spectra of a heated gas, the frequency of
vibration can be obtained for each mode of vibration.
When gases are heated, it is conceivable that some electrons within the
atoms may move from one energy level to a higher one, thus increasing the
internal energy and hence the specific heat. However, particularly for dia¬
tomic gases, the separation between two electronic energy levels is so great
Hydrogen at Low Temperatures 427

that it is highly improbable that the transition will occur except at very high
temperatures. For the various diatomic gases, oxygen has one of the smallest
differences between the two electronic energy levels. On the other hand,
when dealing with gases at very high temperatures, like those encountered in
MHD generators, rocket nozzles, and in the boundary layers surrounding
reentry vehicles, the electronic energy must be considered.
There is evidence that in complex molecules, such as the hydrocarbons,
internal rotation of some of the atoms may occur. This rotation contributes to
increasing the specific heat. The determination of the magnitude of this effect
is beyond the scope of this book.

16-12 HYDROGEN AT LOW TEMPERATURES


The experimental value of C„ of hydrogen, as given in Section 16-10, is
slightly lower than those for the other diatomic gases and is slightly lower
than the predicted value of 5I2R0. When hydrogen is cooled below room
temperatures, there is a decrease in the value of Cu until a temperature of
about 60 K is reached. At this temperature, hydrogen has a value of Cv of
3/2R0. This is the value of Cu for a monatomic gas. But hydrogen exhibits
none of the other characteristics of a monatomic gas, even at this low
temperature.
From Eq. 16-70 we can see that the minimum energy of rotation is inversely
proportional to the moment of inertia of the molecule. Since the moment of
inertia of the hydrogen molecule is on the order of 2 to 3 percent of that of
diatomic gases such as oxygen, nitrogen, and carbon monoxide, the minimum
energy of rotation of the hydrogen molecules is relatively very high. Fur¬
thermore, the difference between the various rotational energy levels of
hydrogen is also very high. In addition, at very low temperatures, the
minimum energy of rotation of the hydrogen molecule is large compared with
the energy of translation. Under these conditions it is highly improbable that
sufficient energy can be communicated to the molecule to cause it to rotate
and, hence, it behaves as a monatomic molecule insofar as its energy is
concerned. As the temperature of hydrogen is increased, accompanied by an
increase in the translational energy, it becomes statistically probable that
more and more molecules will receive at least one quantum of energy and will
rotate. With a further increase in temperature, various molecules will move to
higher rotational levels of energy. At temperatures around 300 K, almost all of
the hydrogen molecules possess the normal amount of rotational energy.
The analysis of the effect of temperature of the energy of hydrogen
molecules between 60 and 300 K is complicated by the fact that hydrogen may
exist in two forms—the para and the ortho forms. At room temperatures
hydrogen is composed, roughly, of three parts of the ortho form and one part
of the para form. If hydrogen is cooled below 300 K and equilibrium is
maintained, the para form percentage increases until, at about 20 K, the
hydrogen is virtually 100 percent of the para form.
428 Elementary Statistics and Quantum Mechanics

16-13 GASES AT EXTREMELY HIGH TEMPERATURES


The methods described in this chapter provide the means of determining
specific heats of various gases within the range of accuracy that is required
for most engineering problems. These methods are valid provided that there is
no change in the atomic or molecular structure of the gas. In some situations,
however, such an assumption relative to the nature of the gas would be
invalid. For instance, for its successful operation the gas of a magneto¬
hydrodynamics generator must be ionized. In a cesium-charged thermionic
converter, the cesium vapor is ionized. In ion propulsion devices, ionized
gases are involved. Another instance where the methods developed here to
determine the specific heats would be totally inaccurate is that of air in
immediate contact with space vehicles and missiles as they reenter the earth’s
atmosphere. As air is heated to high temperatures (say, around 2800 K), first
dissociation of oxygen occurs and then, as the temperatures are increased,
dissociation of nitrogen becomes significantly large. At the same time there is
a tendency for the dissociated oxygen and nitrogen to become ionized. The
relative magnitude of these effects is dependent on the temperature and
pressure of the air.
Under certain conditions, values of the constant pressure specific heat of
air in its dissociated and ionized forms have been reported that are on the
order of ten to fifty times its normal value of 1.004 J/g-K. The methods of
determining the effects of temperature and pressure on the composition of air
and on the specific heats are entirely beyond the scope of this text.

16-14 SOLIDS AND LIQUIDS


An ideal elementary solid may be considered to consist of a space lattice
formed by atoms vibrating about their equilibrium positions. The atoms do
not interact with one another in any way. Using the equipartition of energy
theory, each vibratory mode would contribute RoT5 energy per gram or pound
atom of solid. Since, in theory, each atom is free to vibrate in all directions in
space, it will possess three independent modes of vibrations. Hence, its total
energy will be 3R0T and the value of Cu will be 24.96 J/g atom-K.
In 1819, P. L. Dulong and A. T. Petit examined the constant pressure
specific heats of many solids at room temperatures and found that the values
of these specific heats were approximately 26.79 J/g moie-K. This is known as
the Dulong-Petit value. The constant volume specific heat of a solid is very
difficult to measure but may be calculated by using one of the general
equations of thermodynamics, Eq. 8-30. The value of Cv thus calculated, using
Cp = 26.79 results in a value of most solids somewhat larger than the value of
24.96 predicted in the first paragraph of this section. A careful examination of
the constant pressure specific heats of a large number of solids at room
temperatures shows considerable deviation from the value of 26.79 for some

This includes both kinetic and potential energies.


Solids and Liquids 429

solids. Certain solid values CL, based on experimental determinations, are much
lower than 24.96.
For carbon in the diamond form, boron, beryllium, and silicon, the experi¬
mentally based values are only a fraction of 24.96. Furthermore, the specific
heats of all solids show a material decrease with a decrease in temperature
and approach zero at zero temperature. Thus, the model described in the first
paragraph of this section to account for the behavior of the atoms must be
modified to account for these deviations.
Perhaps the first theory presented to account for these variations in the
specific heats of solids was that of Albert Einstein. Einstein in 1907 applied
the quantum theory of Max Planck and deduced that the constant volume
specific heat of solids was
3RoQxp(hc(olkT) (hcoA2
(16-88)
[exp(hc(olkT)-\]2 \ kT /
Equation 16-88 should be compared with Eq. 16-87a. Einstein’s equation,
although predicting specific heats very well at moderately low temperatures,
shows a discrepancy on the low side at low temperatures with the discrepancy
increasing as the temperature is decreased. Apparently, the actual behavior of
the atoms is more complicated than that of atoms vibrating with a single
common frequency.
In 1912, Peter Debye published his theory of specific heats. He assumed
that internal energies of solids were associated with stationary elastic sound
waves set up by the vibrating atoms. He made use of an expression for
examining standing elastic wave trains that had been developed by Lord
Rayleigh in connection with blackbody radiation. His analysis resulted in a
complicated integral. This complicated integral may be evaluated by con¬
sidering two separate cases. In the region of moderate and high temperatures,
the value of the atomic specific heat becomes
1
C„ = 3R o (16-89)
500

The characteristic temperature or the Debye temperature is represented by


6. It is defined as 0 = hccom/k, with o)m being the maximum frequency of
vibration.
When the temperature is less than approximately 0/10, a solution of the
complete Debye equation becomes

CL = 3R o

or
= 464.4 cal/deg per gram atom (16-90)

Since the specific heat in this region is proportional to the cube of the
temperature, this equation is sometimes referred to as the T law.
430 Elementary Statistics and Quantum Mechanics

The analysis of the specific heats of liquids is much more involved than that
of solids, primarily because of the motion of the molecules. Since the
temperatures encountered with most liquids are not in the extreme ranges, it
is relatively easy to determine the constant pressure specific heat and then to
calculate the constant volume value by adapting Eq. 8-30.

16-15 APPLICATION TO ENTROPY DETERMINATIONS


We pointed out in Section 5-5 that when changes in the state of systems
occur without chemical reactions being involved, entropies based on any
convenient datum may be used to determine entropy changes. However,
when chemical reactions take place, entropy changes can be determined only
by using absolute values of entropies for the original and final states. As we
stated in Section 5-5, absolute values of entropies may be determined by a
step-by-step process. This process is laborious and requires a considerable
amount of experimental information. The accuracy of the final value of
entropy is dependent on the accuracy of this experimental information.
In Section 16-8, an expression was developed by use of quantum mechanics
for the determining of the absolute entropy of substances. This expression is
given in Eq. 16-58. Since in general, Z = ZxZrZvZe, then Eq. 16-46 may be
written per mole of substance as

s = 1 + In Zt + In Zr + In Z„ + Ze - In n0

(16-91)

(a) Monatomic Gases


The energy of rotation and the atomic vibratory energies are nonexistent for
monatomic gases and, therefore, Zr and Zv drop out of Eq. 16-91. As we
discuss in Section 16-11, the difference between electronic energy levels for
gaseous molecules is so great that changes in electronic energy are in¬
significant except at temperatures of at least 2000 K. Hence, for gases at all
except very high temperatures, In Ze may be taken as zero. The value of the
In Zt is given in Eq. 16-60. From Eq. 16-45, the value of (d In Z/dT)v is given
as U/nkT2. Then, per mole,

/alnZA _ Ut __ 3/2R0T
(16-92)
l 3T )v RoP

/ d In Zt \ _ 3
(16-92a)
V dT ) v- ~ 2T

Substituting into Eq. 16-91,

s + In Ze — In n0+ 2
Application to Entropy Determinations 431

or

'2tt\312 , 5 3, . , R0 . 3
= Ro[ j +2ln T _lnp + 2ln m +lnn+2ln ^ + ln
bln'

so that

n \5 , , (27T)mkm , 5, T 3 ^
= i?0[2 + “—rr-!- ^ In T = In p + - In m + In Ze (16-93)
h3 2 2

The evaluation of the electronic partition function, Ze, requires a knowledge


of the energy levels in the various electron states. It is beyond the scope of
this book, but is discussed in detail in texts on statistical thermodynamics. For
most gases, except at high temperatures, the electron energy, ee, is so small in
comparison with the other energies involved that In Ze may be neglected.

(b) Polyatomic Gases


It was shown in Section 16-11 that most diatomic gases do not possess a
significant amount of vibration energy at room temperatures. Then the terms
involving In Zv in Eq. 16-91 may be neglected. Under these conditions the
entropy of a diatomic gas at room temperatures exceeds that of a monatomic
gas only by those terms in Eq. 16-91 that involve In Zr. Since this entropy
difference is due to rotation, it will be designated as entropy of rotation.

s, = R0[lnZr + T(-|^) (16-94)

For those diatomic gases such as oxygen, nitrogen, and carbon monoxide, the
value of y is less than 0.01, where y = h2/%ir2IkT (see Section 16-10). Under
this condition, Zr = 1/cry. Then

(d InZA _ 1
\ aT / t
and
In Z, = In -
cry

Substituting in Eq. 7-39,

(16-95)

When y is small, but exceeds a value of 0.01, the value of Zr from 16-73
should be used in place of the value of 1/y.
Referring to Eq. 16-91, we can see that when atomic vibration occurs, the
entropy is increased by two terms that involve the vibration partition function.
Designate this as the vibratory entropy. Thus
432 Elementary Statistics and Quantum Mechanics

In Section 16-11, it was shown that Zv = [1 — e x] \ Then

/ d In Zv \ _ 1 x
\ 3T ) ~ Te7^
V-

where
h c co
x ~JT
and In Zv = — ln( 1 — e~x). Then

s» = R0[^i-ln(l-e^)] (16-97)

It should be observed that Eq. 16-97 is restricted to one mode of vibration.


If there are more than two atoms per molecule, then the entropy must be
determined for each mode of vibration and a summation made to determine
the total vibratory entropy.
The total entropy of a gaseous system may be obtained by adding the
results from Eqs. 16-93, 16-97 (or the equivalent for small values of y > 0.01),
and 16-95. This is equivalent to the evaluation of Eq. 16-91. Note that the
internal rotation of the atoms has been neglected in the determinations above.
Normally, the atoms do not have measurable values of internal energy of
rotation. However, in complex molecules, such as the petroleum hydro¬
carbons, there may be measurable amounts of energy involved and, con¬
sequently, this must be taken into account in the determination of the total
entropy. This determination is beyond the scope of this book.
Problems 433

PROBLEMS ____
16-1 (a) Determine the error caused in evaluating n! by using Stirling’s
approximation when there are (i) 5 molecules, (ii) 10 molecules, (iii) 30
molecules, (b) Repeat part (a), using Eq. 16-4a.
16-2 Determine the rotation value of Cv for hydrogen chloride at 25°C.
16-3 Plot the quantity exx2l(ex - l)2 versus 1 lx. Carry this plot up to a value
of 1/x of 2.5. Be careful in plotting this curve in the region of 1/x
between 0.1 and 0.2.
16-4 (a) Using the plot in Problem 16-3, determine the vibratory specific heat
of oxygen at 2000 K. (b) Determine the total specific heat and the value
of k(CpICv) of oxygen at 2000 K.
16-5 Determine the limiting value of Cv for water vapor as its temperature is
increased. What factors must be neglected to arrive at this value.
16-6 The vibratory frequency of oxygen is 1580 per centimeter. Determine
the vibratory specific heat, C*, of oxygen at 500 K.
16-7 The vibration frequencies of water vapor in the three modes are 1654,
3825, and 3935 per centimeter. Determine the total vibratory specific
heat of water vapor (C at 500 K.
16-8 Using the results of Problem 16-7 determine the total value of Cn of
water vapor at 500 K.
16-9 The same as Problems 16-7 and 16-8 except at 1000 K.
_ 17 _

INTRODUCTION TO
IRREVERSIBLE
THERMODYNAMICS

17-1 INTRODUCTION
For any actual process to take place within a system, there must be a driving
potential or an unbalance of forces acting on or within the system. Thus no
process, can take place within a system unless the system originally is in a
nonequilibrium state, or unless unbalanced forces are applied to the system.
Furthermore, the process that does take place must be an irreversible one.
Yet, throughout this book, most of the processes that are discussed are
assumed to be reversible ones. As we have pointed out, although no process
can reach reversibility, it can approach it. Thus, reversible processes are
valuable in predicting maximum possible performance. In particular, in
reversible processes there is conservation of available energy.
Especially in connection with the study of entropy, consideration has been
given to the results produced by the irreversibility of actual processes. But
the nature of these irreversible processes has not been analyzed. In this
chapter we study briefly some irreversible processes.
Some processes, particularly those in which there are rapid changes in state
with time, are most difficult to analyze. Such processes include the free
expansion of a gas through a rupture in its containing walls and the explosion
caused by introducing a spark into an inflammable mixture.
On the other hand, when irreversible processes occur in a steady-state
system, the analysis of the process is relatively much easier. This is parti¬
cularly true when the imbalance of forces is not large. For example, consider
a rod that receives heat at one end and has heat removed from it at the same
rate at the other end.

17-2 ENTROPY PRODUCTION


Care must be taken in specifying the state of a nonequilibrium steady-state
system. The properties of such a system vary throughout. However, the

434
Entropy Production 435

1
Heat -^ U—AT Heat
source sink

\
///////////////AV {

I
\
- \ Heat-—
Heat /////////////////;

\
XT'—Thermal conducting rod
T^ t2
'—Thermal insulator

FIGURE 17-1 Entropy production by heat flow.

properties of each part of the system do not change during the time period
under consideration. Consider the rod shown in Figure 17-1. The temperature
of each infinitesimally thin vertical slice of the rod differs from that of its two
adjacent vertical slices. But the temperature of each vertical slice is held
constant. It is not possible to measure the temperature of such a thin slice
because temperature sensing devices have finite size and, hence, measure the
average temperature of many very thin slices. However, except for very high
temperature differences, the thermal conductivity of the rod varies in a
straight line with temperature. Hence, a plot of the temperature of the rod
along its length will be that of a straight line. Under these conditions, it is
necessary only to determine the temperature at the two ends of the rod. The
temperature at any point in the rod can be determined from the straight-line
plot of the rod temperatures.
For the rod in Figure 17-1, the rate of change in the entropy of the source is

Q (17-1)
source
T,

where t is time, Q is the rate of heat flow, and T, is the temperature of the
source. The term Q/T sometimes is referred to as the rate of entropy flow.
Since heat flows from the rod into the sink, there is also an entropy flow
from the rod into the sink. The rate of entropy change of the sink is

(17-2)

The rod itself, being in a steady-state condition, cannot experience any


change with time of its entropy. But since T2 in Eq. 17-2 is lower than Tt of
Eq. 17-1, the entropy flow into the sink exceeds that removed from the
source. Thus the irreversible flow of heat has produced entropy. The amount
of entropy produced

(17-3)
produced
T, T2

where Sq is the entropy produced by heat flow.


The temperature change across an element of the rod in Figure 17-1 is AT.
436 Introduction to Irreversible Thermodynamics

The entropy produced in this element, by use of Eq. 17-3, is

=_0 + _Q_
v dt > produced
, , T T + AT
Q AT
(17-4)
T(T + AT)

When AT is taken to be very small in comparison with T, it can be neglected


in the denominator. Then Eq. 17-4 becomes

(m =-0AT = _0(AT) (17-4a)


\ d( / J2 T\j J
produced

Since the flow of heat is from a high temperature to a lower temperature, AT


must be negative.
The system in Figure 17-2 is similar to that in Figure 17-1 except that the
rod is thermally isolated from the heat source and an electric current flows
through the rod. Because of the resistance to current flow, a heating effect is
produced. Consider an element across which there is a voltage difference AE.
The I2R loss in this element equals I AE. The heating thus produced is known
as the Joule heating. To maintain a steady state, the element must transfer this
amount of heat to the right. The rate of entropy flow to the right,

dSE\ ==Qe = IAE


(17-5)
dt > produced
. . T T

where dSE is the entropy produced by electric current flow and QE is the rate
at which the Joule heating is produced. (This entropy flow leaving the element
is the rate of entropy production in the element, since no entropy has been
assumed to enter the element).
Designate the entropy flow, Q/T, of Eq. 17-4a as Js. Then Eq. 17-4a
becomes

(17-4b)

A comparison between Eqs. 17-4b and 17-5 shows the similarity between
the two methods of entropy production. In the case of heat flow, there is an

Electrical source

Heat Heat
- sink
source
GT^77777777777777.A/////////////A//

////////\///////7777//////////////

G \ —Electrical conducting rod t2


\ \—Thermal insulator

Thermal barrier

FIGURE 17-2 Entropy production by flow of electrical current.


The Onsager Relations 437

entropy current, Js. For the electrical flow, there is an electrical current, I. In
each case there is a potential or driving force. For the heat flow, the driving
force is the temperature difference AT. For the electrical flow, the driving
force is the voltage drop, AF.
Thus far, an analysis has been made for two cases where only one driving
force existed at a given time. In some systems, more than one driving force
exists at once. A common example of this occurs when differences in both
temperature and voltage exist along a conducting rod. As we will discuss
later, this situation occurs in a thermocouple and in a thermoelectric
generator.

17-3 THE ONSAGER RELATIONS


It has been established experimentally that when both a temperature differen¬
tial and a voltage differential exist along a conducting rod, the heat flow
depends not only on the temperature differential but also on the voltage
differential. The electric current flow likewise depends on both the voltage
differential and on the temperature differential. The two flows, then, depend
on the two driving forces.
Onsager first showed that when a proper pair of related flows are selected, a
simple equation exists that relates these flows and their driving forces. These
two related flows are known as coupled or conjugate flows. The method of
selecting flows that constitute conjugate flows is involved and, hence, will be
summarized here. It has been found that when AF and AT are selected as the
driving forces, the entropy current Js and the electric current I constitute
conjugate flows, but this is not true for the heat current Q and the electric
current I.
The expressions for entropy flow and electric current flow are1

I = Lu AF + L\2 AT (17-6)

Js = L2\ AF + L22 AT (17-7)

It is possible to establish three equations relating the coefficients of AT and


AF to the electric resistance, thermal conductivity, and thermoelectric pro¬
perties of the material. Using statistical mechanics, Onsager showed that for a
pair of conjugate flows the coefficients Ln and L2\ are equal. Thus
Lu=L2\ (17-8)

Equation 17-8 is known as the Onsager reciprocal relation. The equations


involving four unknowns can be solved for the four coefficients.
To make certain that it is differentiated from heat flows in general, the heat
current or rate of heat flow associated, in this instance, with entropy produc¬
tion has been designated as Jq. Since Jq = TJS, substitution into Eq. 17-7

'Some authors choose to divide both AT and AE by T in presenting the equations for entropy
flow and electric current flow.
438 Introduction to Irreversible Thermodynamics

yields,
Jq = tl2, ae + tl22at (17-9)

With no temperature gradient along the rod, from Eq. 17-6,

(I)at=o = En Ail

But I = AE/ARe, where ARe is the electrical resistance. Then

1
En — (17-10)
ARf
Now consider a rod having a heat current but no electrical current. From
Eq. 17-6,
~ Ei2
(AE)I=0 = -f^AT (17-11)
r-ii

Substituting the value of AE from Eq. 17-11 into Eq. 17-9,

(E\\L22 ~ Ei2)
(J0)i=0=TAT (17-12)
En
In deriving Eq. 17-12, use was made of Eq. 17-8.
The rate of heat flow per unit area Jq is inversely proportional to the
thermal resistance, ARth, of the element. Or,

AT
Jn — (17-13)
A R th
Inserting this value into Eq. 17-12,

E\\L22 ~ E212 1
(17-14)
Ln TARth
From Eq. 17-6, with AT = 0,

(Oat=o — En AE (17-15)
From Eq. 17-7, with AT = 0,

(Js)at=o — E21 AE (17-16)


From Eq. 17-9, with AT = 0

(Jq)at=o — TL2i AE (17-17)


It is desirable to devise two new parameters, Q* and S*. The parameter Q*
is defined as the ratio of the heat current to the electric current. Or,

Q* = (E) (17-18)
v i 7 AT=0
Using Eqs. 17-15, and 17-17,
(l2I\
Thermodynamics of Thermoelectricity 439

The parameter S* is defined as

S* = (^) (17-20)
Xi 7 AT =0

Using Eqs. 17-15 and 17-16,


t n*
5* = -^ = ^ (17-21)
-L11 I

For a given material, both Jq and Js, when AT = 0, are functions of the
properties of the material at the specified conditions. Hence, the quantities Q*
and S* are dependent solely on the material properties and the current flow.
A knowledge of Q* and S* permits an evaluation of both Js and JQ for a
given electric current flow with AT = 0. When AT is not zero, Js and Jq will
have larger values than those given by Eqs. 17-18 and 17-20.
When the value of AE from Eq. 17-6 is substituted into Eq. 17-7, and the
result simplified by use of Eqs. 17-8, 17-14, and 17-21, the general expression
for the rate of entropy flow is obtained as,

'’-'s'-rar, (l,-22>
Likewise, when the value of AE from Eq. 17-6 is substituted into Eq. 17-9,
and the result is simplified by use of Eqs. 17-8, 17-14, and 17-19, the general
expression for the rate of heat flow is obtained,

jq = Q*i + TIr^ (17'23)

It is evident from Eq. 17-11 that when there is a temperature gradient


across a rod, there is a voltage gradient even when there is no current flow.
This voltage gradient is obtained by substituting values from Eqs. 17-21 and
17-8 into Eq. 17-11. The result is

(AE),=0= -S* AT (17-24)


In this section, we have considered heat flows, electric current flows, and
entropy flows, which were caused by their driving forces. Similarly, analyses
may be made for other types of flows, such as fluid flows.

17-4 THERMODYNAMICS OF THERMOELECTRICITY


An excellent application of irreversible thermodynamics is the study of the
thermoelectric phenomena. A complete study of the thermoelectric
phenomena involves an analysis of the formation of free electrons in conduc¬
tors and semiconductors and the behavior of the free electrons, particularly as
affected by temperature and by “doping” the basic material. However, our
consideration here will be restricted to the application of irreversible ther¬
modynamics.
440 Introduction to Irreversible Thermodynamics

External load FIGURE 17-3 Elementary thermoelectric device.

Certain thermoelectric phenomena can occur in a single element rod; others


require a junction of dissimilar materials. Following common practice, our
attention here will be focused on the two element device, as shown in Figure
17-3. In this illustration, an external load is connected to the cold ends of
materials A and B. When the circuit is closed, a current will flow to the
external ioad. This device may be used as a thermocouple to determine the
temperature difference Ti~T2 by determining the current flow set up by the
temperature difference or by determining the voltage produced with no
current flow. This device also may be used as a thermoelectric generator or as
a thermoelectric cooler when electric current is delivered to it from the
external circuit. The thermoelectric generator will be discussed in the next
chapter.
The three thermoelectric phenomena of prime interest are the Seebeck
effect, the Peltier effect, and the Thomson effect. Although in actual operation
all three effects may occur simultaneously, it is desirable to analyze each one
separately.

a. The Seebeck Effect


Refer to Figure 17-3 and assume that the external circuit is open. Then Eq.
17-24 may be applied to each of the materials A and B. Consider the element
used to develop Eq. 17-24 to be infinitesimally thin, with the temperature drop
across it being dT. Equation 17-24 now becomes

(17-25)

Applying Eq. 17-25 to element A,


Thermodynamics of Thermoelectricity 441

For element B,

(Ec Ea)I=o — S%dT

adding these two equations,

(Ec — Ei,)i=o = f (S%-S*B) dT

or

(E„ - Ec)/=0 = J('T' (S5 - SS) dT (17-26)


1c

This voltage produced in the thermoelectric device, by virtue of a temperature


differential, is known as the Seebeck effect. The thermoelectric device under
consideration here is the common device having two dissimilar materials
joined at their ends. By virtue of the temperature difference, the Seebeck
effect causes a voltage difference to be set up in each material, with the total
voltage being the summation of the voltages produced by the two materials.
The voltage produced in one material per unit temperature difference is
known as the Seebeck coefficient, a. This is defined as

dE
(a)j=o — (17-27)
dT

Although Eq. 17-26 is valuable to the understanding of the Seebeck effect, it


is not readily usable. Integrating Eq. 17-27 to obtain the voltage produced by
the Seebeck effect:
Eh- Ec = a(Th - Tc) (17-27a)

or
A E = a(Th-Tc) (17-27b)

This equation assumes that a is a constant or, for relatively small temperature
ranges, the average value of a may be used.
When two dissimilar materials are joined together to form a common
thermoelectric device, the total voltage set up equals the summation of the
voltage produced in each material. Thus

AEab = A Ea + A Eb — «a(T^ — Tc) + ag(Th — Tc)


or (17-27c)
AEab = (aA + aB)(Th - Tc)

In the use of Eq. 17-27c, absolute rather than algebraic values of a are to be
used.
442 Introduction to Irreversible Thermodynamics

b. The Peltier Effect


It has been found experimentally that when an electric current is passed
through the junction of two dissimilar materials, the heat that must be
removed from the junction to maintain a constant junction temperature may
be more than or less than that required to account for the IR loss. This
difference between the actual heat removed and the I R heating is known as
the Peltier effect. The magnitude of the Peltier heating depends on the
materials involved at the junction and the magnitude of the electric current.
The direction of current flow determines whether the Peltier heating must be
added to or removed from the junction.
The magnitude of the heat removed to maintain the junction temperature
constant may be obtained by making an energy balance at the junction (see
Fig. 17-4). Approaching the junction from the left, the rate of heat flow is Jqa.
Flowing to the right away from the junction, the rate of heat flow is Jqb. The
I2R heating at the junction is I2Rh where Rj is the junction resistance. Then
the heat removed from the junction

JQr = Jqa+I%~Jqb
But

Jq = IQ* = ITS*
Then

Jqr = I2Rj + ITS* ~ ITSb


or

jQr = I% + IT(S*A-S*B) (17-28)


The term IT(S*-S%) is the Peltier heating. For a given junction tern-
perature and given materials at the junction, it depends solely on the electric
current. The Peltier heating is given the symbol Qp.

QP = IT(S%— Sb) = 7tabI, (17-29)


where ttab is the Peltier coefficient of the two junction materials A and B.
Then, from Eq. 17-29,

(17-29a)

FIGURE 17-4 Peltier effect.


Thermodynamics of Thermoelectricity 443

Each material contributes to the Peltier heating and, hence, each has its
own Peltier coefficient. Assume that material A is of the p-type and that
material B is of the n-type.2 It has been shown that ttab = ttp - irn where ttp is
the Peltier coefficient for the p-type material and irn is the Peltier coefficient
for the n-type.

c. The Thomson Effect


Assume that a temperature gradient exists in the rod in Figure 17-5. Then
there will be a conducted heat flow through the rod. Assume that none of this
heat flow is lost from the sides of the rod. Now pass an electric current
through the rod and remove sufficient heat from the rod to maintain all
temperatures in the rod at the same values as those before the flow of electric
current. It is found experimentally that the heat to be removed to maintain
temperatures may be more than or less than the I2R heating. This differential
in the heat to be removed is known as the Thomson effect. This effect is
caused by the flow of an electric current in a region where there is a
temperature differential. The magnitude as well as the sign of the Thomson
effect depends on the material, the magnitude, and direction of the electric
current flow as well as on the temperature gradients.
Consider an element of the rod in Figure 17-5. The heat removed from the
element is

I AE + Jn.

where Jq^ is the rate of heat flow along the rod into the element and Jq is the
rate of heat flow out. Then

J0r = I AE + I(T + AT)S*-ITS*ut

:When an electron is knocked out of an atom, a hole is left. This atom, because of its hole, is
positively charged. So-called dopes may be added to semiconductors to create an excess of holes.
Then this material is known as a p-type material. Other dopes may be added to create an excess
of free electrons. Since electrons are negatively charged, this material is designated as an n-type.

E + AE E

FIGURE 17-5 Thomson effect.


444 Introduction to Irreversible Thermodynamics

But
(A C*\
Sfn=Sout+(^r]AT

r ds*
J0r = I [AE + (T + AT)SJut + (T + AT) AT - TS* t

When AT is made very small in comparison to T, T + AT is substantially


equal to T. Then

jQr = I (aE + S*,AT + T^At) (17-30)

From Eq. 17-6,

AE = -p— “ AT
Tn Tn

Using Eqs. 17-8, 17-10, and 17-21,

Jq=1 (I ARe + T AT)

or

Jo, = I AEe + IT AT (17-31)

Since I2 ARe is the heat lost due to electric resistance, then the quantity
IT(dS*/dT) AT is the Thomson heating. Throughout this book, we have
assumed that the heat flow into the system is positive. Equation 17-30
expresses the Thomson heat removed from the system. Then the Thomson
heating may be written as -It AT, where the Thomson coefficient,

dS*
T = (17-32)
dT

Since the Thomson heating equals It AT, the Thompson coefficient can be
expressed as

__ Qt
(17-32a)
"l AT

There is an interrelationship between the three coefficients that have been


discussed. For the thermoelectric device shown in Figure 17-3, from Eq.
17-26,

Eab= (T\s*A-S*B)dT (a)


JT
1 r
Thermodynamics of Thermoelectricity 445

or
dEAB
(b)
dT
From Eq. 17-29,

ttab = T(S*a-S*b) (c)


From Eq. 17-32,
dS% dS%\
tA ~ rB = - T (d)
dT dT )
From Eqs. (b) and (d),

7TAB — 1 (17-33)

From Eq. 17-27, dE/dT = a. Then

7Tab = olabE (17-33a)


Since
aAB = aA + aB, ttAb = («a + aB)T (17-33b)
From Eqs. (b) and (d),

ta ~ tb = — T (17-34)

Since a = dEldT and d2EldT2 = daldT,

= (17-34a)

Equations 17-33 and 17-34, giving the interrelationships between the three
thermoelectric coefficients, are known as the Kelvin relationships. Kelvin
derived these relationships before the development of irreversible ther¬
modynamics. He assumed an ideal case in which there was no conducted heat
flow between the junctions and no I2R loss. He further assumed that the three
thermoelectric effects discussed here produced no net change in entropy of
the universe. These assumptions cannot be justified and yet experimental
evidence has substantiated the validity of the relationships.
Equations 17-33 and 17-34 are written for the complete thermoelectric
device. Similar equations may be written for each material in the device. It is
evident from these equations that when the Seebeck coefficient is obtained,
then both the Peltier and Thomson coefficients may be calculated. Because of
the Seebeck effect, when two dissimilar materials are joined at their ends and
one junction is heated, a voltage differential is produced. When an external
circuit is connected, electrical energy may be delivered. Such a device is
known as a thermoelectric generator. We will discuss this subject in detail in
the next chapter.
When a much higher voltage than the Seebeck voltage is impressed on a
446 Introduction to Irreversible Thermodynamics

thermoelectric device, heat can be pumped from the cold junction to the hot
junction. The thermoelectric device is then known as a thermoelectric cooler
or refrigerator.
Since, as is shown by Eq. 17-27b, the voltage produced in a specific
thermoelectric device is directly proportional to the temperature difference, a
measurement of the voltage with no current flow will indicate the temperature
difference. The thermoelectric device is now known as a thermocouple.
Problems 447

PROBLEMS ___
17-1 Heat flows through a thermal conductor at the rate of 11,200 J/min. It
enters the conductor at 70°C and leaves at 25°C. The lateral sides of the
conductor are well insulated. Determine the rate of entropy production.
17-2 The conductor in Problem 17-1 is insulated from its heat source but
remains connected to its heat sink. An electrical current of 40 A is
supplied to the conductor. There is a voltage drop in the conductor of
6.5 V. After the rod reaches equilibrium conditions, determine the rate
of entropy production.
17-3 The same as Problem 17-1 except that the heat source temperature is
200° C.
17-4 A thermally insulated conductor, resistance equal to 15 Cl, carries an
electric current of 20 A for 2 s. The initial temperature is 0°C. The
conductor has a mass of 22 g and a specific heat of 0.95 J/g-K. Deter¬
mine the entropy produced.
17-5 A strip of Pb-Te, p-type (a = +300pV/°C) is joined to a strip of
Pb-Te, n-type (a = -300 pV/°C). The temperature of the junction is
600° C and the outerface temperatures are 100° C. Determine the voltage
produced.
17-6 A current of 18 A is caused to flow from the n-type to the p-type
material of the thermoelectric device of Problem 17-5 when the junc¬
tion is maintained at 500°C. Is heat, chargeable to the current flow,
added or removed to maintain temperature constant? How much?
17-7 A semiconductor has a Seebeck coefficient of 275 /xV/°C at its cold end
of 120°C and 292 p,V/°C at its hot end of 425°C. The electrical resis¬
tance is 0.0135 Cl. The current flow is 38 A. Determine (a) joule heating,
joules per hour, and (b) Thomson heating, joules per hour.
17-8 (a) The Seebeck coefficient for a given semiconductor may be taken to
be constant over a temperature range of 200° C. Can the Peltier
coefficient be taken to be constant? Why? (b) What can you say about
the Thomson coefficient in part (a)?
_ 18 _

DIRECT ENERGY
CONVERSION

18-1 INTRODUCTION
Electrical energy is required for a large percentage of our energy applications,
both industrial and residential. With the exception of hydroelectric power,
substantially all of this energy today is developed first as mechanical energy
through the use of heat engines. The added step of first producing mechanical
energy and then transforming it into electrical energy means added com¬
plications and also the opportunity for a loss of thermal efficiency.
Many methods have been proposed over a period of years for the direct
conversion of either heat or chemical energy into electrical energy. The
fundamental principles are well understood for most of these conversion
methods. Laboratory tests have demonstrated the technical feasibility of most
of them. Some have been used to produce small amounts of electrical energy
for special applications, such as power in very remote areas and auxiliary
power for spacecraft operation. In the past, mainly because of high initial costs
or the lack of suitable materials, it has not been economically desirable to use
any of these methods to produce large amounts of power. Today, with the
development of newer materials and with methods to reduce the initial costs,
several direct energy conversion processes show promise of becoming feasi¬
ble for the production of larger quantities of electrical energy. In the sections
that follow, five of these direct energy conversion methods will be discussed.
They are:
1 Thermoelectric generation.
2 Thermionic generation.
3 Magnetohydrodynamic generation.
4 Fuel cells.
5 Photovoltaic cells.
The first three of these devices produce electrical energy directly from heat.
As such they are heat engines. Their thermal efficiencies are limited by the
Carnot cycle principle as are all heat engines. Neither the fuel cell nor the
photovoltaic cell is a heat engine and, hence, the performance of these two
devices is not governed by the Carnot cycle principle.

448
Thermoelectric Generation 449

18-2 THERMOELECTRIC GENERATION


Later in this section we show that materials must have a high value of the
Seebeck coefficient to be satisfactory for thermoelectric generation. The
Seebeck coefficient for metals is very low and, hence, they are not satis¬
factory for this purpose. However, because of other properties such as
durability and flexibility, metals are used almost exclusively for thermocouple
applications. In the discussion that follows, it will be assumed that the
thermoelectric materials are semiconductors.
As the name implies, semiconductors are a class of materials having an
electrical resistance part way between that of insulators and of metals. When
heated to temperatures above room temperatures there is a tendency to drive
off some of the electrons in the outer shell of the atom (the valence electrons).
The freed electrons (called free electrons) now move throughout the material
in much the same way as molecules move throughout a body of gas. When a
valence electron is freed from an atom, a hole is created in the atom. The
remaining portion of the atom then becomes positively charged. The process
of driving off some of the valence electrons and producing holes is known as
electron-hole generation.
A pure semiconductor is said to be intrinsic. When a small amount of
another substance having a different number of valence electrons is added to
a semiconductor, it is said to become extrinsic. This process is termed doping.
When the dope that is added contains more valence electrons than does the
basic material, an excess of free electrons is created by heating. Such a
semiconductor becomes negatively charged and is said to be of the rc-type.
When dope is added containing less valence electrons than does the basic
materials, the heating of the semiconductor produces more holes than there
are free electrons. This semiconductor being positively charged is said to be
of the p-type.
When a n-type and a p-type semiconductor are joined together and heat is
added at the junction, a voltage differential is set up. However, it is not very
feasible to add much heat directly to the junction. This difficulty may be
eliminated by bridging the two materials, as is shown in Figure 18-1, with a
highly heat conducting material, such as copper. When the junction of leg A
and the copper is held at the same temperature as that of the junction of leg B
and the copper, the thermoelectric performance is exactly the same as that
when the two legs are joined directly. When the temperatures of the two
junctions differ only slightly, the thermoelectric performance is substantially
the same as that for equal temperatures, since the Seebeck and Peltier
coefficients for copper are very low.
The analysis of the performance of the generator may be greatly simplified
by assuming that the Seebeck coefficient is independent of temperature.
Actually, the Seebeck coefficient does vary significantly over large tem¬
perature ranges. But a mean value of the Seebeck coefficient may be used
without a serious error. Since, in Eq. 17-33, the term dEldT is the Seebeck
450 Direct Energy Conversion

coefficient, the Peltier coefficient, 7r, varies directly with temperature when
the Seebeck coefficient is held constant.
In Eq. 17-34, the term d2E/dT2 represents the change in the Seebeck
coefficient with temperature. If the Seebeck coefficient is held constant, then,
according to this equation, the Thomson coefficient and, hence, the Thomson
effect is zero.
As with other power-producing devices, the two performance quantities of
interest for the thermoelectric generator are the thermal efficiency and the
power output. In analyzing the performance of the thermoelectric generator,
in addition to the three thermoelectric effects that have been discussed,
consideration must be given also to the conducted heat, Qk, and the joule
heating (I2R).
The Seebeck effect is dependent on the material of the generator and on the
two temperatures; it is independent of the geometry of the generator.
Similarly, both the Peltier and Thomson coefficients are independent of the
geometry. On the other hand, the geometry of the generator has a direct effect
on both the conducted heat and on the joule heat.
The conducted heat for the generator is the sum of the conducted heat for
each leg. Thus

Qk - QkA + QkB

The variation in thermal conductivity of the materials of the generator with


temperature is sufficiently low for the temperature range normally encoun¬
tered in thermoelectric generators so that the mean thermal conductivity may
be used without a serious error. (See Section 19-2 for a detailed discussion of
conducted heat flow.) The conducted heat flow is, then.

kA kA
(Th - Tc)a and (Th - Tc)b
l l
Thermoelectric Generation 451

where A is the cross-sectional area, l is the length of an element of the


generator, and k is the thermal conductivity. For the complete generator.
Qk = K(Th-Tc) (18-1)
where

Note that the cross-sectional area A must be made as small as possible and
that the length l of the legs must be made as long as practical to minimize the
conducted heat flow.
As with the thermal conductivity, the electrical resistivity of the materials
of the thermoelectric generator may be assumed to be constant for the
following analysis. The joule heat loss for the generator is the summation of
the I~R loss for each element; namely,

I2R = 12 08-2)

where p is the electrical resistivity for the material with the units of ohm-
centimeters. Thus, the electrical resistance of the generator is

rt = (18-3)

As is shown in Eq. 18-3, it is desirable to minimize the length of the legs of


the generator and to maximize the cross-sectional area in order that the
resistance and, thus, the joule heat may be minimized. However, this opti¬
mum geometry for minimizing the joule heat will result in a maximum value
of conducted heat.

a. Power Output
The power output of the generator is
P = I AV (18-4)
where A V is the voltage across the generator terminals. Using the definition of
the Seebeck coefficient, and assuming it to be a constant, the theoretical voltage
produced, that is the voltage with no electrical current flow, is
E = a(Th - Tc) (18-5)
The internal voltage loss equals Irt. Then
A V = a(Th — tc) — Irt (18-6)
Substituting into Eq. 18-4,
P = Ia(Th - Tc) - I2r, (18-7)
452 Direct Energy Conversion

The power output, as is shown in Eq. 18-7, is dependent on the current I. The
current may be varied by varying the external resistance. The current for
maximum power can be obtained by differentiating Eq. 18-7 with respect to I
and setting the result equal to zero. When this is done,

<*(Th-Tc)
Ip = (18-8)
2 rt
The substitution of this value of I into Eq. 18-7 gives

_a\Th-Tc)2
(18-9)
4 rt
b. Thermal Efficiency
The thermal efficiency of any heat engine device is that portion of the heat
supplied to it that is delivered as work. Or

TIt = WIQS

The work output of a thermoelectric generator is given by Eqs. 18-7 and 18-4.
The heat supplied, Qs (Qh in Figure 18-1), may be found by making an energy
balance at the hot junction. This balance is

Qh +212^ - Tthl + Qk (18-10)

where i\hI is the Peltier heat removed from the hot junction. It has been
shown that, assuming no surface losses from the legs of the generator,
one-half of the joule heat passes to the hot junction and one-half to the cold
junction. Recognizing that tt = aT, and using Eqs. 18-1, 18-7, and 18-10,

Ia(Th-Tc)-I2rt
(18-11)
V' IaTh + K(Th-Tc)-UJr,

Equation 18-11 may be rewritten as

al - I2r,/(Th - Tc)
Vt = (18-12)
K(Th-Tc) FT,
al +
Ti, 2ThJ

Note that when the thermal conductivity and the electrical resistivity are
reduced to zero, both K and rt become zero and Eq. 18-12 becomes

Vt = (18-12a)

This, of course, is the thermal efficiency of the Carnot cycle. This efficiency
can be approached only when K and rt are minimized.
As is seen by Eq. 18-12, the thermal efficiency is a function of the current I.
The maximum value of the thermal efficiency can be obtained by differentiat-
Thermoelectric Generation 453

ing Eq. 18-12 with respect to I and by setting the result equal to zero. Then

K(Th - Tc) Th + Tc
CO max rj, 1+ (18-13)
2

It is to be noted that the thermal efficiency is dependent on the quantity


a2IKrt. This quantity is called the figure of merit and is designated as Z. It is
dependent on the properties of the materials, that is, a, k, and p, and on the
geometry of the generator. Substituting into Eq, 18-13,

Th + TAU2
a(Th-Tc)
i+z - 1
Lmax n, (18-14)
(Th + Tc)

When the value of I as given in Eq. 18-14 is substituted into Eq. 18-12,

(18-15)

Example 18-1. Determine the maximum power output of a thermoelectric


generator having the following data:

Length of each leg = 4 cm.


Area of leg A = 2 cm2,
Area of leg B = 1 cm .
k for each leg = 0.02 W/cm-°C.
p for leg A = 0 001 Tt-cm, for leg B = 0.008 O-cm.
a for leg A = 180 guV/°C, for leg B = -180° gi V/°C.
Th = 675 K, Tc = 325 K.

Solution. From Eq. 18-3,

0.001 x 4 | 0.0008 x 4
0.0052 n

For the generators, a = aA - aB = 180 - (—180) = 360 /a V/K. From Eq. 18-9,

(3.6 x 10“4)2(675 - 325)'


max = 0.763 W Answer
4(0.0052)

Example 18-2. Determine the maximum thermal efficiency for Example 18-1.
454 Direct Energy Conversion

Solution. From Eq. 18-1,

K = - - + 0 02 X 1 = 0.015 W/°C
4 4
a2 (3.6 x 1(T4)2
z= Kr, 0.015x0.0052
0.00166/K

From Eq, 18-15, with

Th + Tc = 675 + 325 = 50Q K

675 - 325 [1 + 0.00166(500)]1/2 — 1


675 375
1 +0.00166(500)l,2 + |^
O/J

9.98 percent Answer

The thermal efficiency as found in Example 18-2 is higher than that


achieved with actual thermoelectric generators. The actual generators nor¬
mally have efficiencies below 8 percent. There are several reasons why actual
efficiencies are lower than that found in Example 18-2. The thermoelectric
properties of otherwise satisfactory thermoelectric materials change
significantly with temperature and, hence, the figure of merit decreases
rapidly with temperature at the higher temperatures and may be lower than
that determined in Example 18-2. Thus, with a smaller value of Z being used
in Eq. 18-15, the thermal efficiency will be lower. Furthermore, Eq. 18-15 was
derived by using a mean value of the thermoelectric properties rather than by
expressing these properties as a function of temperature. Finally, extraneous
heat losses, like those from the hot junction, have been neglected.

18-3 THERMIONIC GENERATION


Edison is credited with discovering, while working with vacuum tubes, that
metals heated to elevated temperatures emit electrons. The fundamental
concept of a thermionic generator involves the use of a higher heated metal
electrode, the emitter, to emit electrons. A second (low temperature) elec¬
trode, the collector, is used to collect the electrons. From the collector the
electrons flow to an external circuit, thus delivering electrical energy (see
Fig. 18-2).
At elevated temperatures metals contain a large number of free electrons.
Within the metal these negatively charged particles have random motions,
much as do the molecules of a gas. This free motion within the metal is
possible, since the free electron is acted on in all directions by the positively
charged particles created when the electrons are freed from the atoms. At the
surface of the metal, the free electrons are acted on only by positively
charged particles within the metal. Thus, at the surface there is a restraining
force tending to prevent the electrons from leaving the surface. Only those
electrons reaching the surface with sufficient kinetic energy to overcome this
Thermionic Generation 455

Insulation

FIGURE 18-2 Elements of a thermionic generator.

restraining force are able to escape from the surface. The energy required to
escape (i.e., to overcome the restraining forces of the surface) is known as the
work function, which is designated as cfi. Normally the work function is
expressed in electron-volts per electron.
The current density leaving the emitter is given by the Richardson-Dush-
man equation. This is
Js = AT2e{~*lkT) (18-16)
where Js = the saturated current density, amperes per square centimeter
A = Richardson constant (theoretically this equals 120 A/cm2 - K2)
T = surface temperature, K
(f> = work function, electron-volts (eV) per electron
k = Boltzmann constant, 8.61 x 10 5 eV/K
The saturation current density is that current which exists with no restriction
in the interelectrode space on the current flow. The value of the Richardson
constant can be determined theoretically by assuming that the work function
is independent of temperature and that the emitting surface is both absolutely
flat and smooth. Since these assumptions are not true for actual emitters, the
measured values of the Richardson constant differ significantly from its
theoretical value, as is shown in Table 18-1.
The interelectrode space of a thermionic generator is filled with electrons.
Since like charges repel, the electrons in the interelectrode space act as a
retarding force on the electrons leaving the emitter. Only those electrons
having a high velocity are able to traverse the interelectrode space and to
reach the collector; the remainder are turned back to the emitter. In the
simple thermionic generator (i.e., the vacuum type that operates at high
vacuum) the so-called space charge is minimized by making the width of the
interelectrode space very small, perhaps less than 0.0004 cm in width.
456 Direct Energy Conversion

Table 18-1
Measured Thermionic Emission Constants

</> (electron-volts Melting


Surface Material A (A/cm2 - K2) per electron) Temperature (K)

Cesium (Cs) 162 1.81 302


Molybdenum (Mo) 55 4.2 2893
Tantalum (Ta) 55 4.19 3302
Thorium (Th) 60 3.35 2118
Tungsten(W) 60 4.52 3643
Th on W 3 2.63 —

The work functions are shown in Figure 18-3. In this figure <j)m is the
maximum value of the space charge. The total energy required for an electron
to escape from the emitter surface and to overcome the space charge is <\>e + </>m.
Then the actual current leaving the emitter is given by a modified Richardson-
Dushman equation:

J = AT V(^m)/kT (18-17)
The space charge decreases from its maximum value as the collector is
approached, as is shown in Figure 18-3. If the value of the work function of
the collector is subtracted from the potential energy at the collector surface,
the quantity qE is obtained. The term E is the voltage across the terminals of
the generator, and q is the charge per electron. Hence, the term qE is the
work per electron delivered to the external circuit. In order that qE can be
reasonably large, the work function of the collector must be kept small.
Since the surface of the collector is a free one, it has a tendency to emit
electrons that flow back to the emitter. When the collector temperature is

FIGURE 18-3 Thermionic generator work functions.


Thermionic Generation 457

maintained at a low value, an application of Eq. 18-16 shows that the back
flow of electrons from the emitter can be neglected.

Example 18-3. A vacuum-type thermionic generator has an emitter of tung¬


sten and operates at 2800 K. The collector is thorium on tungsten and
operates at 1200 K. The space charge has a maximum value of 0.2 eV and
occurs at the collector surface. Determine:

(a) Current density.


(b) External voltage.
(c) Power density.

Solution
(a) From Eq. 18-17,

60 x 28Q01 2 3 4
, [4.52+0.2/(8.61 x 10-5x2800)]
1.478 A/cm" Answer

(b) See Figure 18-3. (f>e = 4.52, <j>m = 0.2, cfrc = 2.63.
Then qE = 4.52 + 0.2 — 2.63 = 2.09 eV per electron.

external voltage = 2.09 V Answer


(c) Power density = EJ

= 2.09 x 1.478 = 3.089 W/cm: Answer

The thermal efficiency of the thermionic generator is the portion of the heat
supplied to the emitter that is delivered as electrical energy to the external
load. The heat supplied may be obtained by making an energy balance at the
emitter. The energy leaves the emitter as:

1 The potential energy possessed by the electrons that is required to leave


the emitter.
2 The kinetic energy of the electrons.
3 The heat radiation to the collector.
4 The extraneous heat losses.

The extraneous heat losses include the heat conducted along the lead wires to
the external load and the heat loss to parts of the generator other than the
collector. These losses depend on the geometry of the generator. They must
be minimized to obtain maximum performance of the generator. As a limiting
case, they will be neglected in this analysis.
The potential energy possessed by the electrons per unit area equals the
product of the energy per electron and the number of electrons per unit area.
This equals (Jlq)((f)e + </>m). It has been shown that the kinetic energy of an
electron is 2kTE. The kinetic energy per unit area equals (T/q)2kTe.
The heat exchanged by radiation between the emitter and the collector is
dependent on the two temperatures, the heat radiation emissivities of the two
458 Direct Energy Conversion

electrodes, and on their geometry. As is discussed in Section 19-7, the heat


radiation exchange per unit area is given as

(18-18)

where

or = the Stefan-Boltzmann constant = 5.725 x 10 !2 W/cm2-K4


ee = emissivity of the emitter
ec = emissivity of the collector

It may be shown by using Eq. 18-18 that when Tc is low, say 40 percent of Te,
the back heat radiation from the collector may be omitted without causing an
error of more than 2.5 percent. Since values of the emissivities, ee and ec,
cannot be obtained within several percent, the back heat radiation may be
neglected in this case.
Neglecting extraneous heat losses, the thermal efficiency of the thermionic
generator is

EJ
Vt =
— ((f)e + </>m) + ^ (2 kTe) + a (18-19)

Example 18-4. Determine the thermal efficiency for Example 18-3 if ee = 0.19
and ec = 0.17.

Solution. Using Eq. 18-19, rjt equals

2.09 x 1.478
1.478(4.52 + 0.2) + 2x 1.478x8.61 x 10~5(2800) +
1 12
5.725 x 10~12 [2.8J - 1.24]10
1 + 1 - 1
0.19 0.17
3.089
7.49 percent Answer
6.976 + 0.713 + 33.526

It should be noted that in spite of an emitter temperature of 2800 K the


thermal efficiency of the thermionic emitter, even neglecting extraneous heat
losses, is very low. The thermal efficiency may be improved by eliminating the
space charge and by using an emitter material having a lower work function in
order that the current density and the power density may be made higher.
Care must be used, however, to avoid operating the emitter at temperatures
close to the melting temperature since there will be excess mass transport of
the emitter to the collector. It should be noted also from Example 18-4 that
Thermionic Generation 459

the major loss in the thermionic converter is that of heat radiation. Thus, a
prime requirement of both the emitter and the collector is that they have low
heat emissivities.
The space charge in the interelectrode space may be minimized by neu¬
tralizing it. This may be done by introducing a rather readily ionized vapor,
particularly cesium, into the interelectrode space. When the cesium atoms
come in contact with the positively charged emitter, some of them will be
ionized, provided that the work function of the emitter is greater than the
ionizing potential of the cesium (3.875 eV). When sufficient positively charged
ions are formed, they will neutralize the space charge. Not only will this
reduce the energy needed by the electrons to reach the collector, but it will
also permit the widening of the interelectrode space to perhaps as much as ten
times that required for the vacuum-type of generator. In addition, at normal
collector temperatures, the cesium will coat the collector surface, thus reduc¬
ing its work function to that of cesium, 1.81 eV. Emitter temperatures nor¬
mally are too high to permit coating with cesium.

Example 18-5. A vapor-type thermionic generator has conditions similar to


that in Example 18-3 except that cesium is used to completely neutralize the
space charge. Assume no change in the heat emissivities. The external
resistance is adjusted to obtain maximum voltage without restricting the
current flow from the emitter. Determine

(a) Power density.


(b) Thermal efficiency.

Solution. The external voltage equals the difference between the work func¬
tions of the two electrodes. It may be assumed that the collector is completely
covered with cesium. Then

E = 4.52- 1.81 = 2.73 V


The current density is
60(2800)2
3.39 A/cm2
[4.52/(8.61 xl(r5(2800)]

power density = (3.39)(2.73) = 9.26 W/cm2 Answer

The potential energy required for the electrons to escape from the emitter =
3.39(4.52) = 15.32 W/cm2. The kinetic energy of the electrons =
2(3.39)(8.61 x 10_5)2800 = 1.635 W/cm2. The heat radiation loss remains at
39.1 W/cm2. The thermal efficiency equals

9.26
16.5 percent Answer
Vt 15.32+ 1.635 + 39.1

Under the best conditions, thermal efficiencies of close to 17 percent have


been obtained for thermionic generators. The current density and, hence, the
460 Direct Energy Conversion

power output at any given temperature may be increased by using emitters


that have lower work functions. However, such emitters normally have lower
melting temperatures and, hence, must be operated at lower temperatures. In¬
vestigations have been made of ways to attain high current densities at
relatively low emitter temperatures by using devices such as magnetic fields
and areas in the interelectrode space. Although these devices show promise
they are not, as yet, commercially practical.

18-4 MAGNETOHYDRODYNAMICS (MHD) GENERATION


When a solid conductor is pushed through a magnetic field, a charge is
induced in it. If the conductor is connected to an external circuit, electrical
energy is delivered. When an ionized gas is passed through a magnetic field,
some of the electrons in the ionized gas are diverted to one of the two
electrodes (see Fig. 18-4). If the electrodes are connected to an external
circuit, electrical energy may be delivered. As will be seen later, this electrical
energy comes from the enthalpy of the gas rather than from the work input
required to move the conductor through the magnetic field. Thus the need of
first producing mechanical work is eliminated.
Most gases must be heated to prohibitively high temperatures to produce
the desired amount of ionization. Hence, a relatively easy-to-ionize substance,
for instance, cesium, potassium, or some of their compounds, is introduced
into the gaseous stream to seed it. Even then, temperatures of approximately
2500°C are required to produce the desired degree of ionization. Normally
such temperatures are produced by burning a fuel in highly heated air (see
Fig. 18-5). The air is preheated by the hot gases leaving the MHD genera¬
tor. In order that the MHD generator can attain the desired performance, the
gas velocity through it must be high and may even be supersonic. Such
velocities are attained by expanding the gases in a nozzle. A compressor is
required to build up the desired air pressure.
The elements of the MHD generator are shown in Figure 18-6. The gas flow
is in the x direction, the magnetic field is in the y direction, and the electron
flow is in the z direction. It should be noted that, although acted on by the

Negative electrode
\yyyyyyyyyyyyy777yyyyyyyyyyy77777yyyfyyA

Ionized
gas

^\\\\\\\\\\\\^
Positive electrode

-—VvVA—^
External load

FIGURE 18-4 Charge separation in an ionized gas.


Magnetohydrodynamics (MHD) Generation 461

MHD

FIGURE 18-5 Elements of a MHD plant.

Negative
electrode

FIGURE 18-6 Elements of a MHD generator.

same magnetic forces as those acting on the electrons, the positive ions are
not diverted significantly from a straight path. This is true because the mass
of the positive ion is several thousand times that of the electron.
Normally, the velocity of the gas does not vary significantly as it moves
through the generator. It will be assumed to be constant in these analyses. It
will be further assumed that the magnetic field strength is constant throughout
462 Direct Energy Conversion

the generator. The voltage induced by the magnetic field is given as


Ei = hYx B (18-20)

where h is the average height of the flow channel. The gas velocity, Y, is in
the jc direction and the magnetic field strength B, is in the y direction. Y x B
becomes YB when expressed in scalar quantities. Then

Ei = hYB (18-20a)

Since there is no change in the gas velocity in the system, the summation of
the forces, acting on the gas in the direction of motion, must be equal to zero.
It may be shown that the magnetic force in the direction of gas motion equals
~JB per unit volume, where J is the current density. The other forces are the
gas pressures and the sidewall force in the direction of motion. Since the
generator is relatively short, the sidewall frictional forces may be neglected.
When the summation of these forces is set equal to zero and the equation is
simplified,

Adp = -JABdl
where l is the length of the generator. When the above equation is divided
through by A and the remainder is integrated across the length of the
generator, /,

P\~ p2~ JBl (18-21)


The induced voltage, E„ as given by Eq. 18-20a is also called the open circuit
voltage. For closed circuit conditions, E,- equals the summation of the voltage
lost within the generator and the external voltage, Ee. The internal resistance,
rint = hlcrbl, where cr is the internal conductivity and b is the channel width. Since
J = I/bl, the internal voltage loss = Jh/cr. Then

Ei = — + Ee = hYB

or

Ee = h (VB - ^-) (18-22)

Even though operating temperatures are very high, the mass rate of flow is
so high that the heat lost by the gas to the walls may be neglected, particularly
for a short generator. Under these conditions,

mCpT, = mcpT2+ W = mcpT2 + IEe

where m is the mass rate of flow.


Using a mean value of cp, with I = Jbl,

JblEe
(18-23)
mcp
Magnetohydrodynamics (MHD) Generation 463

Since the velocity is a constant,

Ai/A2 — p\lp2 (18-24)


Then the density ratio,

(18-25)

The pressure ratio may be determined by use of Eq. 18-21 and the tem¬
perature ratio by use of Eq. 18-23.

Example 18-6. A MHD generator has a gas velocity of 1000 m/s. The height
at the entrance is 0.15 m and at the exit is 0.25 m. The width is 0.15 m. The
magnetic field strength is 15,000 G. The electrical conductivity is 50/H-m.
Determine the external voltage if the current density is 20,000 A/m2.

Solution. From Eq. 18-22,

0.15 + 0.25\/1000x 15,000 20,000


= 220 V Answer
2 A K? 50 /

Note: there are 10 4 V-s/m2 per gauss.


Example 18-7. The mean specific heat of the gas in Example 18-6 is 1.17 J/g-K.
At the entrance of the generator the gas is at a pressure of 275 kPa and a
temperature of 2490°C. Determine

(a) Length of the generator.


(b) Pressure at exit.
(c) Temperature at exit.

Solution. The true gas constant, R, is close to that of air. Using this constant,
the density at the entrance is

P 275 x 1000
p RT 0.287 x (2490 + 273)

The mass rate of flow is


m = pAV = 346.8x0.15 x 0.15 x 1000
= 7803 g/s

From Eq. 18-23,


20,000x0.15/ x 220
T2 = 2763 - = (2763-72.3 l)K
7803 x 1.17

From Eq. 18-21,


20,000 x 15,000/
= (275 - 301) kPa
P2 275 lO4 x 103
464 Direct Energy Conversion

Note:
ampere ampere volt-second
—S— x gauss -’f X -’f-
ITT nr nr
watt-second joule _ N Pa
m m m m

The gas density at exit, from Eq. 18-25, is

275-30/ 2763 —^ = 3484 (-HI—


p2 = 346.8 (
275 .2763 - 72 31/ V2763-72.3/

But,
/ 275-30) \
p,Ai — P2A2 — 3484
V2763-72.31)

364.8(0.15 x 0.15) = 3484 (2753 Sffjj) (0'15 x °'25)

l = 4.29 m Answer
T2 = 2763 - 72.3 x 4.29 = 2453 K Answer
p2 = 275 - 30 x 4.29 = 146.3 kPa Answer

As Example 18-7 shows, the temperature at the exit of the MHD generator
is also high. (A high exit temperature is required to maintain the desired
degree of ionization to provide electrons for charge separation and to main¬
tain a high electrical conductivity of the gas.) Thus, although the initial
temperature is high, the temperature range is relatively small because the
thermal efficiency of a MHD generator is relatively low. Hence, normally it is
not very practical to use the MHD generator by itself. Instead, the hot gases
leaving the generator, after being used to preheat the air for the combustor,
are used to produce steam at conditions of temperature and pressure com¬
parable with those in a modern fossil-fuel plant. The steam thus produced is
used in a steam turbine. The overall temperature range of the MHD generator
and the steam plant is large and, therefore, the thermal efficiency of the two
units combined is much larger than that of a steam power plant alone.
Because the fuel burned in the combustor is burned to supply energy for the
two units combined, it is not feasible to determine the thermal efficiency of either
unit individually.
In the absence of heat losses and internal losses, the ideal expansion in a
MHD generator is a reversible adiabatic process. This expansion is shown in
the T-s diagram in Figure 18-7. Process 1-2' is the theoretical expansion and
the actual one is 1-2. This diagram resembles the cycles used for both steam
and gas turbines. The performance of turbines is determined by finding the
ratio of the work actually delivered to the theoretical work. The term “turbine
efficiency” is used to designate this performance. Since the expansion in the
Magnetohydrodynamics (MHD) Generation 465

S FIGURE 18-7 MHD generator performance.

MHD generator resembles that in turbines, the term turbine efficiency will be
used for judging the performance of a MHD generator.
Since there is no change in the gas velocity in the MHD generator,

actual work (fll H 2)act


turbine efficiency = (18-26)
theoretical work (h]-h2)s

where the subscript s designates a constant entropy expansion.1 Assuming


that the specific heats for the actual and theoretical expansions are equal, Eq.
18-26 becomes

(T i T2) act
turbine efficiency = (18-27)
(T\ - T2)s

where T2 is the temperature after isentropic expansion.

Example 18-8. Determine the turbine efficiency for the MHD generator of
Example 18-7.

Solution. The temperature at the end of isentropic expansion is

k== cp - R = 1.17 - 0.287 = 0.883 J/g-K

1.17
k = 1.325
0.883
/ 1 46 0.325/1.325
2763 = 2367 K

Substituting in Eq. 18-27,

2763 - 2453
turbine efficiency = = 78.39 percent Answer
2763 - 2367

'The theoretical work is obtained with an isentropic expansion.


466 Direct Energy Conversion

18-5 FUEL CELLS


In all of the work-producing devices we have discussed thus far, a fuel
reacts with oxygen. The chemical energy released by the chemical action
produces high temperature gases. Directly or indirectly, these hot gases are
used to produce mechanical work in engines or electrical energy in the direct
energy conversion devices studied in this chapter. In the fuel cell, the
intermediate step of producing hot gases is eliminated, and electrical energy is
produced directly from chemical energy. In storage batteries, electrical energy
is also produced directly from chemical energy, but the storage battery is a
closed system with all of the reactants being contained within the battery. As
such, the battery can deliver power for only a limited length of time before
the reactants must be restored to their original state by charging. The charging
requires more power than is delivered by the battery. On the other hand, the
reactants are supplied continuously to the fuel cell, thus permitting power
generation for an indefinite length of time.
Several different fuels have been considered for use in fuel cells. These
include hydrogen, carbon, carbon monoxide, and the lighter hydrocarbons. The
heavier hydrocarbons may be reformed outside the fuel cell itself, and the
lighter fuels thus formed are then used in the fuel cell. Either pure oxygen or
air is always supplied to the cell to provide the oxidizing agent. At present,
the most common cell is the hydrogen-oxygen cell. The discussion that
follows will be centered on this type of cell. The basic principles we develop
are applicable, however, to other types of cells.
The elements of a hydrogen-oxygen fuel cell are shown in Figure 18-8. At

FIGURE 18-8 Hydrogen-oxygen fuel cell.


Fuel Cells 467

the fuel electrode the fuel is ionized. Thus

2H2->4e 4- 4H+ (18-28)

The four electrons pass from the fuel electrode to the external load and return
to the oxidizer electrode. The four positive hydrogen ions pass through the
electrolyte to the oxidizer electrode where they combine with the four
electrons and with the oxygen. Thus

4e + 4H+ 4- 02-* 2H20 (18-29)


When Eq. 18-28 and Eq. 18-29 are added,

2H2 + 02-»2H20 (18-30)

Although the overall results are the same as would be obtained when
hydrogen is burned in air, it must be emphasized that the hydrogen and
oxygen do not combine directly in the fuel cell. The actions at the two
electrodes of the hydrogen-oxygen cell as given here are the most common
ones. It is possible, with different electrodes, to have different actions. The
net result, however, is the same.
Equation 11-16 may be adapted to the fuel cell. In Eq. 11-16, the enthalpy
of the reactants is the molecular enthalpy, that is, the enthalpy of the
molecules as composite wholes. This enthalpy may be combined with the
chemical energy to determine the combined enthalpy of the reactants. Thus

H, 4- chemical E = Hrc

When this is done, Eq. 11-16 becomes

Hrc + Q = HP + W (18-31)

The maximum work is obtained when the process is reversible and there is no
net change in temperature. Under these conditions, Eq. 18-31 becomes

Wmax = Hrc - Hp + Qiso (18-3la)

where Qiso is the amount of heat that must be added to maintain the cell
temperature constant.
For a reversible process, the overall change in entropy must be zero. Or,
2 AS = 0. But
2 AS - ASsys 4- 2 ASsurr — 0

where “sys” refers to the system and “surr” to the surroundings. Since the
system must be in a steady-state condition, ASsys = 0. Then ASsurr = 0. The
reactants are drawn from the reactants supply, where the entropy change
equals — Sr. (Sr is the entropy of the reactants that are supplied to the cell.)
The products of the reaction are delivered to the surroundings causing their
entropy to increase by an amount equal to Sp. The addition of Qiso to the
system causes a change in entropy of surroundings by the amount of (QjSO/To),
468 Direct Energy Conversion

where T0 is the temperature of the surroundings. Then

2ASsurr = -Sr + S„-% = 0


1 0

Or
Q iso — Tq(Sp Sr)
Substituting into Eq. 18-3la,

Wmax = (Hrc - T0Sr) - (Hp - T0SP) (18-32)

Since G = H - TS,
Wmax=Grc-Gp (18-32a)
Observe that the absolute values of entropy for the various substances must
be used in Eq. 18-32. The absolute values are tabulated in Table 18-2.
Equation 18-32a states that the maximum amount of electrical energy de¬
livered by a fuel cell equals the difference between the combined Gibbs
function of the reactants and the Gibbs function of the products.
The term thermal efficiency has no relevance in connection with a fuel cell,
since the concept of thermal efficiency was developed for heat engines. To
make valid comparisons between the performance of a fuel cell and a heat
engine for a specific fuel it is customary to define the efficiency as

electrical energy output


fuel cell efficiency =
heating value of fuel
We
Vf.c. (18-33)
Qcomb

In Chapter 11, it was shown that -Qcomb = Hr — Hp. Then

Vf.c.
w. (18-33a)
Hrc - Hp
For the ideal fuel cell the electrical output, as given by Eq. 18-32a, is

GRc ~ Gp
Then

ideal rjf c = gRc ~ 9/ (18-34)


** rc

Using the definition of G, Eq. 18-34 may be written as

(Hrc — Hp) - T(Sr - Sp)


ideal rjfc (18-34a)
Hrc - Hf
Normally, Sr exceeds Sp and thus the ideal efficiency is less than 100 percent.
For a very limited number of reactions, Sp exceeds Sr. In this case the ideal
efficiency will exceed 100 percent. The student is invited in Problem 18-15 to
discuss whether or not this statement is a violation of the first law.
Fuel Cells 469

Table 18-2
Entropy of Gases at Atmospheric Pressure

Temperature J/g mole-K

K CO o2 n2 h2 co2 h2o

250 192.39 199.88 186.31 125.53 207.17 182.82


300 197.71 205.23 191.63 130.75 213.80 188.92
350 202.21 209.79 196.13 135.21 219.75 194.12
400 206.12 213.79 200.03 139.11 225.15 198.83
450 209.59 217.36 203.49 142.55 230.13 202.56
500 212.72 220.61 206.60 145.63 234.76 206.31
550 215.57 223.60 209.43 148.43 239.08 209.72
600 218.20 226.37 212.03 150.97 243.15 212.88
650 220.66 228.96 214.48 153.33 246.99 215.83
700 222.96 231.39 216.72 155.51 250.62 218.59
750 225.12 233.68 218.90 157.53 254.08 221.20
800 227.16 235.88 220.88 159.45 257.36 223.68
850 229.11 237.90 222.79 161.24 260.52 226.05
900 230.97 239.86 224.62 162.95 263.52 228.31
950 232.74 241.72 226.36 164.57 266.41 230.49
1000 234.43 243.51 228.03 166.11 269.18 232.59
1100 237.62 246.85 231.17 274.41
1200 240.57 249.94 234.09 279.27
1300 243.32 252.81 236.80 283.81
1400 246.90 255.49 239.35 288.07
1500 248.32 258.00 241.74 292.08
1600 250.60 260.37 244.00
1700 252.76 262.61 246.14
1800 254.80 264.74 248.17
1900 256.75 266.76 250.11
2000 256.80 286.69 251.94

Source. This table is based largely on Tables of Thermal Properties of Gases, National
Bureau of Standards Circular 569, November 1, 1955.

The electrical energy output of a reversible cell per gram mole of fuel is

W = Erneq (18-35)

The term Er is the voltage of a reversible cell, ne is the number of electrons


produced per gram mole of fuel, and q is the charge per electron.
The number of electrons produced per gram mole of fuel, ney equals the
product of the number of electrons produced per molecule of fuel, n'e and the
number of molecules per mole. There are 6.02486 x 1023 molecules per gram
mole. The charge per electron is 1.60186 x 10 26C. The product of these two
numbers, known as the Faraday constant F, equals 96,510C/g mole or
470 Direct Energy Conversion

96,510 J/g mole. Using these figures, Eq. 18-35 becomes


=96,510Ern; (18-36)

where Wm = the work of a reversible cell, J/g mole


n' = the number of free electrons per molecule of fuel

Example 18-9. A hydrogen-oxygen fuel cell operates at 300° C and at atmo¬


spheric pressure. Determine the following, per gram mole of fuel.

(a) Ideal work.


(b) Minimum amount of heat exchanged with the surroundings.
(c) Ideal efficiency.
(d) Ideal voltage.

Solution. Rewriting Eq. 18-32,

Wmax = (Hrc - Hp) - T0(Sr - Sp)

The overall reaction equation is iU + ICb-^IDO.


It is necessary to determine the chemical energy to evaluate Hrc. Refer to
Eq. 11-11. At 25°C, using Table 11-2 and Table 11-1,

chemical energy = 286,011 — [0 — (0 — 1 x 44,000)]


= 242,011 J/g mole

Hrc = 242,011 + (8030 + = 254,259 J/g mole

Hp = 9565 J/g mole


224.88'
(a) Wmax - (254,259 - 9565) - (300 + 273.15) 149.61 + 210
2 / ■77]
= 244,649-29,383 = 215,311 J/g mole Answer

(b) The heat exchanged with the surroundings

= T0(Sr - Sp)

= (300 + 273.15)
( 149.61++4^1-210.77 ]
= 29,383 J/g mole Answer

(c) From Eq. 18-33a,

W,max 215,311
ideal efficiency = = 88 percent Answer
Hrc - Hp 254,259 - 9565

(d) From Eq. 18-36,

W,max 215,311
ideal voltage = = 1.120 V Answer
2x96,150 2x96,150
Photovoltaic Cells 471

18-6 PHOTOVOLTAIC CELLS


In 1887, Hertz reported that an interaction of a light photon with an electron
could give the electron sufficient energy to escape from its parent metal. This
principle is the basis of the photovoltaic cell. Today, the source of light for
practically all photovoltaic cells is sunlight. Hence, these cells are known as
solar cells.
The solar energy that the earth receives is composed of discrete particles of
energy known as photons. The energy of each photon is determined by its
frequency of vibration. Thus,
6 = he (18-37)
where e = energy per photon, joules
h = 6.625 x 10“34 J-s = 6.625 x 10 27 g-cm2/s2
v = frequency of vibration, 1/s
The frequency of vibration,
v = c/A (18-38)
where c = velocity of light, 3 x 10locm/s
and A = wavelength, cm

Combining Eqs. 18-37 and 18-38,

(18-39)

Since it is moving at the speed of light, the energy of the photon can also be
determined by the use of the famous Einstein relation:
e = me2 (18-40)
It should be noted that the photon is a quantum of energy, the magnitude of
which is given by Eqs. 18-37 and 18-39. If the photon loses any of its energy,
it ceases to exist.
The energy required to drive off an electron from its parent materials is
known as its gap energy. Values of gap energies are given for some common
solar-cell materials in Table 18-4.

Table 18-4
Typical Gap Energies

Material Gap Energy, eg (eV)

Silicon (Si) 1.12


Gallium arsenide (GaAs) 1.35
Germanium (Ge) 0.68
Aluminum antimonide (AlSb) 1.53
Zinc telluride (ZnTe) 2.10
Cadmium sulphide (CdS) 2.40
472 Direct Energy Conversion

When a photon strikes an electron, it can free the electron only when its
energy exceeds the gap energy of the electron. With less energy than this, the
entire energy of the photon is dissipated. When a photon possesses energy in
excess of the gap energy of the electron, this excess energy is also dissipated.
Since there is a wide variation in the wavelengths of the photons in solar light
and therefore in their energies, many photons do not possess sufficient energy
to free any electrons in a given material. Other photons possess much more
energy than the gap energy of the electrons, with this excess energy produc¬
ing only a heating effect.
The portion of the solar energy, having a specified wavelength, that is
received at a given location is dependent on the location, the time of the day
and the season, and also on weather conditions. An approximation of the
distribution of the solar energy is given in Table 18-5 for a clear day with the
sun directly overhead.

Example 18-10. Determine the energy in electron volts of the photons in light
having a wavelength of 0.5 /a. Note: 1 eV = 1.602 x 10 ergs.

1 erg = 1 g-cm2/s: = 10 7 J

Solution. From Eq. 18-39,

6.625 x 10-27 x 3 x 1010


6 = = 3.975 x 1012 g-cm2/s2
5 x UT5
3.975 x 10~12
= 2.48 eV Answer
1.602 x W™

Example 18-11. Determine the maximum wavelength, in microns, of the


photons in sunlight that can release the electrons in silicon.

Solution. From Table 18-4, for silicon is 1.12eV.

1.12 x 1.602 x 10“l2 = 1.794 x 10-12 g-cm2/s2 = 1.794 x 10 19 J

Table 18-5
Distribution of Solar Energy

Portion of Solar Energy in Specific


Wavelength Interval (p) Interval (%)

0.3 to 0.5 17
0.5 to 0.7 28
0.7 to 0.9 20
0.9 to 1.1 13
Above 1.1 22

Note: 1,000,000 p = 1 m.
Photovoltaic Cells 473

From Eq. 18-39,

. 6.625 x 1(T27x 3 x 1010 , 1AO 1A_„ , ...


A =-1~794 >00^-= x cm = 1- 108jUL Answer

(Since 1 cm = 104 /x.)


An examination of Eq. 18-39, together with Table 18-4, shows that only
those photons having a rather short wavelength can possibly free electrons in
materials such as cadmium sulphide. From Table 18-5, even silicon, with its
low gap energy, cannot use any energy of 22 percent of the photons in
sunlight.
When a p-type semiconductor is brought into intimate contact with an
n-type of semiconductor, at first there is an unstable state right at the
junction. There is an excess of free electrons in the n-type and an excess of
holes in the p-type. Then, there will be movement of electrons into the p-type
material where they will be combined with holes and vanish. This leaves an
excess of electrons in the p-type material at the junction and it becomes
negatively charged. In a similar manner, because of hole movement into the
n-type material, at the junction it tends to become positively charged. When
the sunlight strikes the junction, it tends to create an excess of free electrons
in the n-type material. Because the p-type material is negatively charged at
the junction, there is a barrier that tends to resist the flow of the free
electrons created in the n-type material. Hence, a voltage is created in the
n-type material. When a connection is provided to an external circuit, elec¬
trons will flow through the circuit and then back into the p-type material,
combining with the holes created there by the sunlight. Note that although a
barrier exists to restrict the flow of the free electrons created in the n-type
material into the p-type of material, nevertheless there will be some flow of
these electrons into the p-type material. The higher the voltage set up in the
n-type material, the greater will be the flow into the p-type material. This
flow, known as drift current, reduces the efficiency of the solar cell.
It is essential that the solar rays strike both the n-type and the p-type
material at the junction. To accomplish this, one material, normally the p-type
material, is made so extremely thin that the sun’s rays can penetrate it and
reach the n-type material. It is also essential that the basic material of the
solar cell be made extremely pure. Because of these requirements, the cost of
solar cells at present is so great that power generated by them is extremely
expensive. Initial costs are being reduced, but it is difficult to predict if these
costs can be reduced enough to make it practical to use solar cells for large
amounts of power generation.
In general, the efficiency of some types of solar cells has approached 20
percent in laboratory tests. It is also difficult to predict how high the efficiency
will be in the future. In summary, it should be emphasized that several losses
must be minimized to improve the efficiency. These losses include:

1 Reflection of solar energy from the surface.


474 Direct Energy Conversion

2 Those photons that have less energy than the gap energy of the electrons
cannot free any electrons and, hence, their energy is wasted.
3 Any energy of photons in excess of the gap energy is wasted.
4 Electron-hole generation that does not take place close to the junction
may be only temporary as recombination may take place.
5 Drift current across the junction is wasted energy.
Problems 475

PROBLEMS ______
18-1 A thermoelectric generator has the following properties:
Ap — 5 cm2 A„ = 4 cm'
fcp =0.0215 W/°C =0.0215 W/°C
ap = 300 jLtV/°C = —300 /a V/°C
pp = 0.00105 n-°c pn = 0.00092 n-°c
Th = 527°C Tp = 127°C
lp = ln = 4.5 cm
Determine: (a) figure of merit, (b) maximum power output, and (c)
voltage at maximum power output.
18-2 Determine the thermal efficiency at maximum power output for Prob¬
lem 18-1.
18-3 The same as Problem 18-1 except that Th = 1000°C.
18-4 Determine the thermal efficiency at maximum power output for Prob¬
lem 18-3.
18-5 The same as Example 18-3 in the Chapter discussion but the emitter
has a temperature of 2200 K.
18-6 (a) The same as Example 18-4 in the Chapter discussion but with an
emitter temperature of 2200 K. (b) The heat radiation loss varies as the
fourth power of the absolute temperature. Explain why the reduction
of the emitter temperature from 2800 K to 2200 K should not increase
the efficiency.
18-7 A vapor-type thermionic converter has conditions similar to those in
Example 18-3 in the Chapter discussion except that the emitter tem¬
perature is 2200 K. Cesium is used to completely neutralize the space
charge. Assume no change in the heat emissivities. The external
voltage is adjusted to obtain maximum voltage without restricting the
current flow from the emitter. Determine (a) power density and (b)
thermal efficiency.
18-8 Same as Example 18-6 in the Chapter discussion except the gas
velocity of 1500 m/s.
18-9 The same as Example 18-7 in the Chapter discussion but for a gas
velocity of 1500 m/s.
18-10 A constant velocity MHD generator operates with a gas velocity of
1600 m/s and a magnetic field strength of 18,000 G. The voltage at the
terminals is 150 V. Gases enter the generator at 207 kPa, 2480°C. The
channel width is 15 cm. The height at entrance is 14 cm and at exit is
21 cm. The electrical conductivity is 48/D-m. Assume that the gas
constant equals that for air. Determine (a) current density, (b) length
of the generator, and (c) pressure and temperature at generator exit.
476 Direct Energy Conversion

18-11 Determine for Problem 18-10, (a) total current flow, (b) power de¬
livered, and (c) turbine efficiency.
18-12 The same as Example 18-9 in the Chapter discussion except the
operating temperature is 600°C.
18-13 The same as Example 18-9 in the Chapter discussion except that the cell
operates at a pressure of 5 atm.
18-14 The cell in Example 18-9 in the Chapter discussion is to operate at
average load of 75 percent of rate load of 3.5 kW for 24 hours a day
for 20 days. Determine the minimum amount of hydrogen and oxygen
required.
18-15 As discussed in the text, it is possible for the ideal efficiency of certain
fuel cells to exceed 100 percent. Is this a violation of the first law?
18-16 The same as Example 18-11 in the Chapter discussion but for a
wavelength of 1.0 ja.
18-17 The same as Example 18-11 in the Chapter discussion but for cad¬
mium sulfide.
18-18 Using Table 18-5 in the text, determine the maximum percentage of
solar energy that would theoretically release electrons from silicon.
Suggestion: using the average wavelength in each interval, determine
the percentage of the solar energy for that interval which can release
electrons of silicon. Then make a summary for all intervals.
18-19 The same as Problem 18-18 except use cadmium sulfide.
_ 19 _

ELEMENTS OF HEAT
TRANSFER

19-1 INTRODUCTION
Throughout this book a major consideration has been the heat transferred to
or from a system or within a system. However, no consideration has been
given to the mechanism of the heat transfer process, other than to recognize
that temperature is the driving force causing heat to be transferred. Nor has
consideration been given to the direct determination of rates of heat transfer
and to the factors controlling these rates. In this chapter we analyze these
various heat transfer factors.
Engineers are concerned about heat transfer with two different objectives
in mind. In some situations they deal with systems that must be maintained at
temperatures differing from that of the surroundings. Since energy must be
expended to maintain the system temperature, it is essential that heat transfer
between the system and its surroundings be minimized. In other situations,
the processes occurring within the system tend to cause the system tem¬
perature to exceed a safe operating value. In these instances, it is essential
that rates of heat transfer from the system be made high.
The rate of heat transfer is a function of a temperature difference and of
the resistance in the heat transfer path. There is a similarity in the flow of
heat, the flow of fluids, and the flow of electrical energy. In each case, the rate
of flow is a function of the driving force (potential) and of the resistance to
flow.
There are three modes of heat transfer: conduction, convection, and
radiation. Generally, a heat-transfer problem is concerned with two, or
perhaps all three, of these modes simultaneously. However, it is desirable to
study each mode separately in order to understand the effects of each type of
heat transfer on the overall rate. A study of each of these modes is necessary
to ascertain how the overall rate of heat transfer may be increased or
decreased as may be desired. In the following sections each of these three
modes of heat transfer will be investigated. Then two and three modes will be
combined to give the means to determine the overall rate of heat transfer as
encountered in engineering work.
In a large number of heat-transfer problems, results may be approximated

477
478 Elements of Heat Transfer

by assuming steady-state and unidirectional heat flow. Steady-state heat flow


is said to exist when the temperature at each point under consideration
remains unvarying with time. Unidirectional heat flow infers that the heat
flow must take place in one direction. Frequently, there is flow in two or more
directions, but the flow in any direction other than the main direction is so
small that unidirectional heat flow is approximated. Unless otherwise stated
steady-state, unidirectional heat flow will be assumed in this chapter.

19-2 CONDUCTION
Heat transfer by conduction may be thought of as the heat transferred
through a substance (or combination of substances) from a region of high
temperature to a region of low temperature by the progressive exchange of
energy between the molecules of the substance. In the process of transferring
heat by conduction, no bodily displacement of the molecules occurs. In the
case of metals, however, electron movement greatly assists in heat transfer by
conduction.
The fundamental law of conduction is credited to Fourier. This law may be
illustrated as follows. Consider steady-state, unidirectional heat flow through
a solid, as is indicated in Figure 19-1. Take a slab of the solid having a
cross-sectional area A normal to the path of heat flow. Let the thickness of
the slab be dx, and let the temperature difference across the slab be dt. From
his experimental work Fourier developed the following relationship:1

Q = -kAg (19-1)

where Q = heat flow per unit of time


k = proportionality factor, called the thermal conductivity
dT/dx = rate of change in temperature with distance in the direction of heat flow

'This is a restricted form of the more general form of Fourier’s law of heat conduction for heat
flow in three planes, which is applicable to both steady-state heat flow and other conditions of
flow.

FIGURE 19-1 Fourier’s law of heat conduction.


Conduction 479

In the SI system of units, thermal conductivity may be expressed as

W/m2 -r K/m = W/mK

Extensive experimental investigations have established the values of ther¬


mal conductivities of many substances and the effect of temperature on these
conductivities. Values of thermal conductivities of several classes of sub¬
stances are given in the Appendix. Also Figures A-3 to A-5 in the Appendix
show the effects of temperature on thermal conductivity. Note that the
thermal conductivity of any metal is very high in comparison with that of any
gas. The reported values of thermal conductivities of metals are valid only for
metals of a given degree of purity. Particularly for those metals with the
highest values of thermal conductivity, the introduction of a slight amount of
another metal will cause a significant change in the thermal conductivity.
The best heat-insulating solids owe their insulating properties to the air or
to other gases contained in cells within the material. These cells cause the
heat to flow through the solid material through a long tortuous passage. In
addition, the available cross-sectional area of the solid material is much less
than the projected area. Experimental evidence shows that many small
unicellular pockets of gas are much more effective than a series of connected
cells having the same total volume in giving insulating value to a substance.
There may be considerable variation in the thermal conductivity of any given
insulating material because the conductivity depends on its density, the size
and number of its air cells, and its absorbed moisture.
There are several accepted methods of experimentally determining the
thermal conductivity of solids. When proper care is used, fairly accurate
values can be obtained for the thermal conductivity of a given solid of
specified composition. It is much more difficult, however, to determine the
thermal conductivity of a gas, a vapor, or a liquid, since it is almost
impossible to eliminate the heat transferred by convection, which occurs
simultaneously with that transferred by conduction, without introducing
difficulties in the accurate measurement of other factors. For these reasons
there are differences of perhaps 10 to 25 percent in reported values of the
thermal conductivities of fluids.
Figure 19-2 shows the heat conduction in a simple wall. It is assumed that
the width and height of the wall are so large in comparison with the thickness
of the wall that the heat flow may be considered to be unidirectional. One face
of the wall is maintained at a uniform temperature tu and the other face is
kept at temperature t2. The heat flow through the wall may be obtained by
integration of Eq. 19-1.
An examination of the thermal conductivities of the various materials given
in the Appendix shows that, for many materials, the thermal conductivity may
be taken as constant over an appreciable range in temperature. Furthermore,
for most materials, the thermal conductivity is a straight-line function of
temperature within the range of temperature for which information is avail¬
able. Thus, the arithmetic mean thermal conductivity km may be used as the
480 Elements of Heat Transfer

true thermal conductivity. For the simple wall, Eq. 19-1 may be integrated as
follows:

(T, - T2) (19-2)

According to Eq. 19-2, the rate of heat flow is proportional to the heat-flow
area, the temperature difference causing heat flow, and the term kJX. This
term is known as the thermal conductance.
When the thermal conductivity does not vary linearly with temperature
(approximately, at least), the mean thermal conductivity krn cannot be deter¬
mined readily. In such a case it becomes desirable to express the thermal
conductivity as a function of temperature in Eq. 19-1 and then to perform the
integration.
Equation 19-2 may be rewritten as follows:

Q_(T,-T2)
(19-3)
A X/km

According to Eq. 19-3 the rate of heat flow per unit area is directly propor¬
tional to the temperature difference and is inversely proportional to the term
XIkm. Equation 19-3 is similar to the equation for the flow of electric current,
which is

E
I 09-4)
R

where I = flow of current


E = difference in voltage that causes flow
R = resistance to flow

A comparison of Eqs. 19-3 and 19-4 shows that the term Xlkm is similar to the
term R. Thus, XIkm is known as the thermal resistance of a substance.
Conduction through a Composite Wall 481

19-3 CONDUCTION THROUGH A COMPOSITE WALL


Consider a composite wall, as is shown in Figure 19-3, composed of sub¬
stances A, B, and C. When the heat flow is at a steady rate and when it may
be assumed to be unidirectional, then the heat flow through the wall may be
determined by calculating the heat flow through any of the three materials
composing the wall. This calculation requires a knowledge of an interface
temperature. Since the determination of an interface temperature is
frequently difficult, it is desirable to establish an expression for heat flow
through the wall as a function of the temperature of the two free surfaces.
When Eq. 19-3 is applied to each of the three substances composing the
wall and the terms are rearranged, the following equations are obtained for
the temperature drop across each substance:

(T,-T2)- Q(X)
A \krn J a

(T2-T3) = Q( x\
A\km)
m/ B

(t3~t4) = Q(*\
A\km/c

But
T, - T4 = (T, - T2) + (T2 - T3) + (T3 - T4)
.0
A (&+©, ■* © j
Hence,

A(T| — T4)
(19-5)
Q = (X/fcm)A + (X/km)B + (X/kJc

Area

FIGURE 19-3 Conduction through a composite wall.


482 Elements of Heat Transfer

It should be apparent from Eq. 19-5 that the resistance of a series heat-flow
path equals the summation of the resistances of the individual parts of the
path. In this respect, heat resistance is similar to electrical resistance.
When the expression for the temperature drop for any material composing
the wall is divided by the expression for the overall temperature drop, it
becomes apparent that the temperature drop is directly proportional to the
resistance.

Example 19-1. A flat wall is composed of 0.2 m of fireclay brick (k =


1.039 W/m-K), 0.1m of insulation (k = 0.278 W/m-K), and 0.1m of common
brick (k = 0.694 W/m-K). The innerface temperature of the fireclay brick is
870° C and the outer face temperature of the common brick is 38° C. Deter¬
mine the rate of heat flow per unit area through the wall.

Solution. The resistance of the wall (the denominator of Eq. 19-5)

0.2 0.1 0.1


0.696 m2K/W
1.039 0.278 0.694
Then
1 x (870 - 38)
= 1195 W/m2 Answer
0.696

Example 19-2. Determine the inner face temperature between the common
brick and the insulation for Example 19-1.

Solution. The resistance of the common brick

f = 0^4 = 0144m2K/W
The temperature drop is proportional to the resistance. The temperature drop
across the common brick,
0 144
AT = ~!hz(870 - 38) - 172°C.
u.oyo

The inner face temperature

T = 38 + 172 = 210°C Answer

As an alternate solution, when the rate of heat flow is known, from Eq. 19-3,
O X
Ti — T2 = = 1195x0.144= 172°C

19-4 CONDUCTION THROUGH THICK-WALL CYLINDERS


The heat flow through the wall of a cylinder differs primarily from that
through a straight wall only because the heat-flow area is not constant for a
cylinder. Consider the hollow cylinder illustrated in Figure 19-4. Assume that
Conduction through Thick-Wali Cylinders 483

FIGURE 19-4 Conduction through a single cylinder.

the inner-surface temperature is maintained at tx and the outer-surface tem¬


perature is kept at t2. Assume also that the length of the cylinder is such that
end effects can be neglected and, hence, the flow is unidirectional.
Consider the heat flow through a thin hollow cylinder having an inner radius
r and a thickness dr. If Eq. 19-1 is applied to a cylinder of length L, the result
is

dT
0 -k2irrL
dr

or

Q — = -2k7rL dT
r

Integrating between the limits r, and r2, gives

Q lng) = 2kmirL(T, - T2)

or

a = 2km7rL(Tf - T2)
(19-6)
V ln(r2/r,)
Equation 19-6 is applicable to a cylinder of any thickness. However, when
the cylinder wall is very thin, that is, when r2/r, is only slightly above unity, it
becomes difficult to obtain an accurate value of the denominator of Eq. 19-6.
In such a case, it is sufficiently accurate to assume that the hollow cylinder is
unwrapped to form a flat sheet having a width equal to the mean circum-
484 Elements of Heat Transfer

FIGURE 19-5 Conduction through concentric cylinders.

ference and a length and thickness equal to the length and thickness of the
cylinder. When the ratio of outside to inside radii, or r2lrx, is as large as 1.25,
the error caused by treating the hollow cylinder as a flat plate is ap¬
proximately 0.4 percent.
When heat transfer occurs through a series of concentric cylinders, it is
desirable to calculate the rate of heat flow without the necessity of determin¬
ing interface temperatures. (It is very difficult to measure these temperatures.)
This problem is similar to that of conduction through a composite wall
(Section 19-3). For the three concentric cylinders in Figure 19-5,

T, - T4 = (T, - T2) + (T2 - T3) + (T3 - T4)

The temperature drop across each cylinder may be evaluated by use of Eq.
19-6 and the three expressions may be added and equated to the overall
temperature drop, Tx — T4. From this expression,
2ttL(T, - T4)
(19-7)
lnO^/nt InOi/rTi InOJrd

The term [\n(r2l rx)]l km may be thought of as an equivalent resistance for a


hollow cylinder, and the sum of these similar terms for concentric cylinders
may be considered as the total equivalent resistance. The temperature drop
across any cylinder is proportional to the ratio of its equivalent resistance to
the total equivalent resistance.

Example 19-3. A 4-in. steam pipe (of mild steel) is covered with a 2.5-cm
layer of 85 percent magnesia. The steam pressure is 1.14 MPa. The outer
Radiation 485

surface temperature is 65° C. Determine (a) the heat lost per hour per meter of
the pipe, and (b) the inner face temperature between the pipe and the
insulation.

Solution. Since the steam pressure is relatively low, a standard weight pipe
will be used. From the Appendix, the outer pipe diameter is 4.50 in. (0.1143 m)
and the inner diameter is 4.026 in. (0.1023 m). The value of k for the insulation
is 0.0727 W/m-K, and k for the pipe is 55.4W/m-K. Assuming that the steam
is saturated, its temperature, from the steam tables, is 185.68°C. There is a
condensate film on the inner pipe wall, but its resistance is so very small in
comparison with the total resistance that it can be neglected. From Eq. 19-7,

2tr x 1(185.68-65)
0 . /0.1143\ / 0.1143 + 2x0.025'
n\0.1023/ | H 0.1143
55.4 0.0727
758.3 758.3
= 151.9 W/m Answer
(0.00200 + 4.991) 4.993

The resistance of the pipe, as just determined, is 0.00200 m-K/W and the
total resistance is 4.993 m-K/W. Then the temperature drop across the pipe is

0.00200
(185.68-65) = 0.048°C
4.993

The temperature between the pipe and the insulation is

185.68 - 0.05 = 185.63°C Answer

Since in practice temperatures cannot be measured closer than a fraction of


a degree for engineering purposes, Example 19-3 shows that the temperature
drop across a pipe can be neglected when it is covered by a reasonable
thickness of insulation.

19-5 RADIATION2
Every free surface emits energy in the form of electromagnetic waves; the
amount of energy is a function of the surface temperature. This emitted
energy is known as radiant thermal energy. The nature of this radiant energy
is not completely understood, but laws have been formulated that describe its
behavior. It is recognized that, as with other forms of radiant energy, radiant
heat energy is transmitted in the form of electromagnetic waves. The com¬
plete formulation of the laws governing radiant heat energy must consider
that this energy is quantized, that is, the energy is transferred in quanta. In

2The radiation discussed in this chapter is restricted to heat radiation, that is, the radiation
emitted from a surface by virtue of its temperature.
486 Elements of Heat Transfer

contrast with other modes of heat transfer, no medium is required to transmit


radiant energy. In fact some gases, for instance, carbon dioxide and water
vapor, absorb some of the radiant energy passing through them.
For a fixed set of conditions, any free surface emits radiant energy of
varying wavelengths. The frequency of vibration (v) of radiant waves is
dependent solely on the source of radiation and is independent of the medium
through which they pass. The velocity of radiant waves (V) is a function
solely of the medium through which they pass. Thus, the wavelength (A =
V/v) is a function of both the source and the medium.
All free surfaces receive radiant energy from all other surfaces that they
can “see,” that is, surfaces in direct line of sight. Most problems in radiation
deal with the net radiant energy exchanged between a given surface and those
that surround it. In common parlance, the term “heat exchanged by radiation”
is used. It must be emphasized, however, that radiation is not heat. Heat is
conducted to a surface. By virtue of the temperature of a surface, elec¬
tromagnetic waves transmit energy from the surface. When these strike
another surface, part of the energy will be absorbed, tending to increase the
temperature of the surface struck by them, and part will be reflected. When
the object is transparent, or partially so, to radiant waves, some or all of the
radiant energy received by the surface will pass into the object. The trans¬
parency of an object to radiant energy is a function of the wavelength of the
radiant waves. These statements relating to the radiant energy received by a
surface may be put in equation form as follows:

a +p+t = 1 (19-8)
where a = absorptivity, or the portion of the radiant energy that is absorbed;
p = reflectivity, or the portion of the radiant energy that is reflected;
r = transmissivity, or the portion of the radiant energy that is transmitted.

A black surface has an absorptivity close to unity. For this reason the term
blackbody has been used to designate an imaginary object whose surface has
an absorptivity of unity. Since no known surface completely absorbs radiant
energy, the term blackbody refers to an ideal surface. Kirchhoff conceived a
method of completely absorbing radiant energy. Assume that a hollow sphere
contains a very small opening, as is indicated in Figure 19-6. Radiation
entering this opening will be received by the back wall of the sphere. Here it
will be partially absorbed and partially reflected to other parts of the walls of
the sphere. The reflected waves are, in turn, partially reflected, so that each
reflected portion is a progressively smaller portion of the energy entering the
sphere until ultimately all of it is absorbed. Strictly speaking, some of the
reflected radiant energy will pass out through the hole. However, the surface
area of the sphere is ttD2. Hence, when the diameter of the sphere is chosen
to be 50 times that of the opening, the inside surface area is 10,000 times that
of the opening, and it may be assumed that the hollow sphere absorbs all of
the radiant energy:
Radiation 487

The amount of radiant energy emitted by a surface is a function of the


nature of the surface and its temperature. The term blackbody is also used to
denote a surface that emits the maximum conceivable amount of radiant
energy at any given temperature. There is no actual surface that is a perfect
emitter, but the hollow-sphere concept may be used to establish a standard.
The process of emission from the inner surface of the sphere is the reverse of
that of absorption.
The total radiant energy emitted in a unit time by a unit area is known as
the total emissive power and is designated by E. Since radiant energy is
emitted over a range of wavelengths,
p\=<*

E = Ex d\ (19-9)
J \ =o
where Ex is the monochromatic emissive power.3 It is assumed that Ex in Eq.
19-9 is a continuous function of A.
Several investigators attempted to establish a relationship between Ex and
A. None of these relationships agreed with observed values over the entire
range of wavelengths until the introduction of the quantum theory by Max
Planck around 1900. One application of the quantum theory is a relationship
between Ex and A for blackbody radiation that fits experimental values over
the entire range of wavelengths. This relationship is

3.658 x 10V5
14,300/XT (19-10)

where Ex = monochromatic emissive power of a blackbody, W/nr-/a


A = wavelength, microns
T = temperature of the radiating surface, K

3Monochromatic refers to a specific wavelength.


488 Elements of Heat Transfer

A micron is equal to 10 6 m. When the value of £A from Eq. 19-10 is substituted


into Eq. 19-9 and an integration is performed, the result is

E" = 5-725©4 (19-11)

where Eb is in W/m2.
The effect of temperature on the intensity of radiation, and the distribution
of the radiation at various wavelengths, are shown for blackbody radiation in
Figure 19-7. It is to be noted that the wavelengths of visible light range from
0.4 to 0.8 i±. The area below the emissivity curve for any given temperature is
the total emissive power of a blackbody at that temperature.
Some 20 years before the establishment of the quantum theory, Eq. 19-11
was independently obtained by Stefan by a study of Tyndall’s experiments
and by Boltzmann through analytical means. Hence, it is known as the
Stefan-Boltzmann law.

FIGURE 19-7 Blackbody radiation.


Relationship between Absorptivity and Emissivity 489

19-6 RELATIONSHIP BETWEEN ABSORPTIVITY


AND EMISSIVITY
Absorptivity has been defined in the preceding section. Emissivity is defined as
the ratio of the total emissive power of a surface at a given temperature to that of
a blackbody at the same temperature. Kirchhoff is credited with the establish¬
ment of the relationship between absorptivity and emissivity.
Consider a body suspended, by means of a very fine thread, within a hollow
sphere that has been completely evacuated. The thread being neglected,
energy can be exchanged between the body and its enclosing sphere only by
radiation. Ultimately, the body will reach a temperature equal to that of the
inner surface of the sphere. At this temperature the radiant energy emitted by
the body must be equal to that received from the sphere. Thus

EXA = la i A (19-12)
where A = surface area of the body
Ei = total emissive power of the body
I = radiant energy received by the body per unit time per unit area
a i = absorptivity of the body

Then

E = la or 1 ——
a

When the suspended body is truly a blackbody,

Eb = Iah or I =—
ab
Equating the values of I for the two cases,
EEb
— = — = Eb or E - aEb
a ab
(ab = unity) (19-13)

Equation 19-13 is a statement of Kirchhoff’s law of radiation; namely, that the


emissive power of a body equals the product of its absorptivity and the
emissive power of a blackbody for the specified conditions.
By definition, the emissivity of a body is

Substituting this definition in Eq. 19-13 gives

e = a (19-14)

A restriction must be placed on Eq. 19-14 because of the method by which


it was derived. When equilibrium was reached between the body and its
490 Elements of Heat Transfer

enclosing sphere, the body received radiation from a surface at its own
temperature. Hence, the full statement of Eq. 19-14 is that the emissivity of a
surface at any temperature equals its absorptivity of radiant energy from a
surface at the same temperature.
By definition, the absorptivity of a blackbody is independent of temperature.
The absorptivity of an actual surface usually depends on the temperature of
the surface emitting the radiation. For example, the absorptivity of an oil
paint is almost independent of the color of the paint when it is receiving
radiation from any surface whose temperature does not exceed 100 to 150°C,
the absorptivity ranging from 0.92 to 0.96. On the other hand, the absorptivity
of solar radiation is 0.97 to 0.99 for black paint and 0.12 to 0.26 for white
paint.
By Eqs. 19-11, 19-13, and 19-14, the total emissive power of any surface is

E = 5.725«(^)4 (19-15)

For any surface the radiant heat emitted is

Q = EA (19-16)

Equation 19-15 provides a means of determining the radiant energy given


off per unit of surface area. If the surface temperature is assumed to be
uniform, the total energy radiated by a blackbody is

5.125 A (19-17)

where Q = radiant energy in watts


A = surface area, in square meters
T = surface temperature, in degrees Kelvin

The value of the emissivity of a surface is a function not only of its


temperature but also of its smoothness and, in many instances, of its color.
Any highly polished metal has a very low value of emissivity, but as the metal
becomes oxidized, there is a marked increase in the emissivity. For example,
at room temperatures the emissivity of highly polished copper is 0.03. Heavily
oxidized copper, however, has an emissivity of approximately 0.75. Values of
the emissivities of representative surfaces are given in the Appendix.

19-7 NET RADIANT ENERGY EXCHANGE


The net heat added to a system is heat added to the system minus the heat
removed from the system. Similarly, the net radiant energy received by a
system equals the radiant energy which it receives minus the radiant energy
which it gives up to its surroundings.
One factor governing the net radiant energy exchanged by a surface
depends on the geometry of the surface in relation to its surroundings.
Net Radiant Energy Exchange 491

Consider two infinite parallel planes A and B, whose surfaces are truly black
bodies. Each surface absorbs all the radiant energy emitted by the other
surface. Thus the net radiant energy received on a given area of surface A
equals the difference between the radiant energy it receives and the radiant
energy it emits. Or

Qa = 5.725 Aa IYJjlV- ( _ZaV1 (19-18)


Luooj liooj J
where Aa is the area of surface A being considered. Equation 19-18 is equally
valid for the radiant energy exchanged for a blackbody that is totally enclosed
by a second body.
When radiation occurs between two infinite parallel planes that are not true
blackbodies, part of the radiant energy will be reflected instead of absorbed.
Hence, the net radiant energy gained by surface A will differ from that given
by Eq. 19-18. The magnitude of the net radiant-energy exchange is a function
not only of the two temperatures but also of the emissivities of the two
surfaces. In addition to the fact that there are no actual blackbodies, the two
bodies under consideration are usually so located that the energy radiated by
one surface is received only in part by the second surface. (One exception is a
surface totally enclosed by a second surface.)
Suppose, for instance, that two parallel planes of finite extent are exchang¬
ing radiant energy. Since a considerable portion of the energy radiated by
either surface is not received by the other one, there is a further reduction in
the net radiant energy exchanged between the two surfaces. Hence, the net
radiant energy exchanged between two surfaces, per unit area, depends on the
surface temperatures, the emissivities of the surfaces, and the configuration
of the surfaces. To account for the last two effects, the terms emissivity factor
and shape factor have been originated. These are introduced in Eq. 19-18 to
make it valid for actual radiant-energy exchange. Thus

Qa = 5.725AAFeFA IYTjlV- /'JkVl (19-19)


Luoo/ uooj J
where Fe equals the emissivity factor, which takes into account the emis¬
sivities and reflections of both surfaces; and FA equals the shape factor,
which takes into account the relative positions of the two surfaces arid the
effect of size and shape on the percentage of radiant energy received by the
two surfaces.
For any given problem, it is possible to establish mathematically values of
both the emissivity factor and the shape factor as a function of the emis¬
sivities of the two surfaces and the configuration of the two surfaces. The
method of establishing these factors is discussed in various textbooks on heat
transfer. In 1930, H. C. Hottel analyzed several common arrangements of
surfaces and determined values of Fe and FA.4 His values for the simplest
cases are given in Table 19-1 and Figure 19-8.

4H. C. Hottel, “Radiant Heat Transmission,” Mechanical Engineering, 52 (1930), p. 699.


492 Elements of Heat Transfer

Table 19-1
Emissivity Factors and Shape Factors

Surfaces Fe fa

1
1. Infinite parallel planes 1
±+±-i
€i e2
2. Completely enclosed body (1)
small compared with enclosing
body (2) €l 1
3. Completely enclosed body (1) 1
1
large compared with enclosing
body (2) e2
4. Parallel disks or squares <ue2 See Figure 19-8

FIGURE 19-8 Shape factors for parallel squares and disks.

Example 19-4. Two parallel squares, A and B, are each 1.2 mx 1.2 m. They
are 3.3 m apart. Square A has a temperature of 280°C and square B has a
temperature of 840°C. Treating the two surfaces as blackbodies, determine
the net radiant energy exchanged.

Solution. Since both bodies are blackbodies, the value of Fe in Eq. 19-19 is
unity. The ratio of a side to distance = 1.2/3.3 = 0.364. From Figure 19-8,
Convection 493

Fa = 0.05. Substituting into Eq. 19-19,

Q = 5.725(1.2 x 1.2X0.05)[(^) - (jg)

= 5940 W Answer
Example 19-5. Assume the same conditions as Example 19-4 except surface
A has an emissivity of 0.8 and surface B has an emissivity of 0.7.

Solution. From Table 19-1, the emissivity factor


F = e2€\ = 0.8 x 0.7 = 0.56

Then

0 = 5.725 x 1.2 x 1.2x0.05 x0.56


= 3326W Answer

Note that if the surfaces in Example 19-3 were part of parallel planes, the
radiant energy exchange would be 118,800 W, since Fa would be unity.
In the foregoing discussion, it is assumed that the space separating the two
surfaces under consideration either is a complete vacuum or is filled with
gases that do not absorb radiant energy. Under conditions normally encoun¬
tered in engineering practice, many common gases, such as oxygen and
nitrogen, do not absorb or emit sufficient radiant energy to necessitate
consideration. However, it should be recognized that other gases, such as
water vapor, carbon monoxide, carbon dioxide, sulfur dioxide, and hydro¬
carbons, do absorb or emit sufficient radiant energy at specific wavelengths to
materially affect radiant-energy transmission.
Radiation through gases that absorb radiant energy, emission of radiant
energy from these gases, and radiation from flames are important in the
transfer of radiant heat. Although much work has been done in regard to
these topics, results have not been well formulated and methods of cal¬
culation have not been sufficiently simplified to permit their presentation in
this text.

19-8 CONVECTION
The term convection heat transfer refers to the heat exchanged between a
surface and a fluid moving over the surface. The amount of heat thus
transferred is dependent on the nature of the surface and its geometry and
with the nature of the fluid and with its velocity over the surface as well as
the temperature differences. There are two types of convection heat transfers;
namely, free (or natural) convection and forced convection.
Assume that the temperature of a fluid in contact with a surface is lower
than that of the surface. The fluid will have a tendency to be heated by the
surface. This heating causes the density of the fluid in direct contact with the
surface to become less than that of the main body of the fluid. Because of its
494 Elements of Heat Transfer

lower density, this fluid will move upward carrying with it the energy that it
absorbed. This fluid is replaced at the surface by other fluid which, in turn, is
also heated and rises. This process continues, with a continuous fluid motion
over the heated surface as long as the main body fluid temperature is less than
that of the surface. A reverse of this process takes place when the surface
temperature is lower than the main body fluid temperature. Since this process
is a natural one, it is termed natural or free convection.
As the name implies, forced convection occurs when the fluid is forced to
flow over the heat transfer surface by external means. A fan, a blower, a
compressor, or a pump may be the forcing agency.
The exact nature of convection heat transfer is very difficult to describe
because of the large number of variables involved and because of the
difficulty of measuring them accurately. For instance, assume that the tem¬
perature of the fluid in the immediate vicinity of the surface is about 10 deg
lower than that of the surface. The surface temperature may not be uniform
throughout and the fluid temperature also may not be uniform. Thus there
could easily be a 2 deg or 20 percent error in determining the true temperature
difference. For such conditions, the precise but very elaborate analysis of
convection heat transfer cannot be justified.
The problem of convection heat transfer can be simplified by assuming that
a stagnant film of the fluid is in immediate contact with the surface and that
heat is transferred through this film by conduction. Even though this assump¬
tion does not describe the true situation, particularly for high fluid velocities
over the surface, it gives approximate results that are satisfactory for many
engineering purposes.
It may appear that Eq. 19-2 can be used to determine the heat conducted
through the assumed stagnant film. However, this equation calls for the
thickness of the stagnant film. Since this film is an assumed one, its thickness
cannot be measured. In the discussion of Eq. 19-2, it was pointed out that the
heat flow is proportional to the thermal conductance of the heat flow path,
k/X. Designating this term as hc and calling it the film coefficient,5 Eq. 19-2
becomes

Q = Ahc(Tl-T2) (19-20)

It is possible that a significant amount of heat may be transferred by radi¬


ation from a surface bounded by a fluid. In some situations, the fluid may be
completely enclosed by the heating surface. The fluid may be a gas, such as
air, which does not absorb a significant amount of radiant energy. In such a
case the only heat transferred from the heated surface is by convection.
When the fluid can absorb significant amounts of radiant heat energy and
where the heated surface is exposed to other surfaces, the reader should refer
to heat transfer texts, where this complicated method of heat transfer from a
heated surface is explained.

^Sometimes called the surface coefficient or the film conductance.


Free Convection 495

19-9 FREE CONVECTION


The factors that may influence the film coefficient hc for free-convection heat
exchange between a fluid and a vertical wall are the following:

The height of the wall L.


The temperature difference across the film AT.
The gravity acceleration g.
The mean fluid properties of the film: density p, viscosity p, thermal con¬
ductivity k, specific heat cp, and the coefficient of thermal expansion )3.

It is obvious that with this very large number of parameters that it is not
feasible to determine experimentally the interrelated effects of all these
parameters on hc. Here, dimensional analysis can be of considerable help. All
variables are expressed in terms of the fundamental units of length, tem¬
perature, time, mass or force, and heat. The variables then are placed in
dimensionless groups. The dimensionless group containing the film coefficient
hc is then expressed as the product of a constant, and the remaining dimen¬
sionless groups are raised to unknown powers. The constant and the
exponents of the dimensionless groups are determined experimentally. Since
there are relatively few dimensionless groups, the experimental work is far
less extensive than that required by considering the individual parameters
separately.
By use of dimensional analysis the following relationship has been
established for free convection:

tkL = c(LL£LM4 AT) (^j^j (19-21)

KL
where = Nusselt number, Nu;
k
Lip2gB ATlgi2 = Grashof number, Gr;
pcp/k = Prandtl number, Pr;
C = a constant, to be discussed shortly.

Extensive experimental work has shown that the exponents m and n are
substantially equal. Under these conditions, Eq. 19-21 will reduce to an
expression for hc of
h,,cL{mi^r (19-22)

or

hc = Cj-(aL3AT)m (19-23)

where a = g/3p2cp/jak. For an ideal gas a is dependent solely on the mean film
temperature. It has been found that for low values of (aL3 AT), the free flow
over the surface is laminar in nature. When the value of (aL AT) becomes as
496 Elements of Heat Transfer

high as approximately 109, the flow becomes turbulent. For laminar flow [i.e.,
(aL3 AT) < 109] experimental results show that the value of m is approxi¬
mately For turbulent flow, m has a value of approximately For extremely low
values of (aL3 AT) (i.e., <103), the value of m is uncertain.
Note from Eq. 19-23 that for turbulent flow with m = |, hc is independent
of L and thus the Eq. 19-23 becomes
hc = Ck(a AT)1'3 (19-23a)

For laminar flow, with m =\, Eq. 19-23 reduces to


(A T* \
a -j— J (19-23b)

When the value of L is several meters or larger, there is question about the
accuracy of Eq. 19-23b. However, normally with large values of L, the flow is
turbulent and Eq. 19-23a applies. In the use of Eq. 19-23, care must be taken
to use the mean film properties, that is, at the mean film temperature.
It has been found that Eq. 19-23 is applicable to both the upper and lower
faces of horizontal plates. There is some uncertainty as to the value of
(aL3 AT) where transition from laminar to turbulent flow takes place for
horizontal plates, but the transition may be at values as low as perhaps 107.
Equation 19-23 may be used for long horizontal cylinders if the diameter of
the cylinder, D, is substituted for L.
Values of the coefficient C of Eq. 19-23 are given in Table 19-3. In this table
the coefficient for laminar flow is designated as C'.
Example 19-6. A horizontal tube having an outside diameter of 5 cm, has its
outer surface temperature maintained at a temperature of 50° C. The ambient
air temperature is 26°C. Determine the convective film coefficient.

Table 19-3
Values of C and C To Be Used in Eqs. 19-23a and 19-23b

Laminar Flow Turbulent Flow


Type of Surface (C') (C)

Vertical plate 0.55 0.13


Upper face of horizontal plate
receiving heat 0.35 0.08
Upper face of horizontal plate
giving off heat 0.70 0.17
Lower face of horizontal plate
receiving heat 0.70 0.17
Lower face of horizontal plate
giving off heat 0.35 0.08
Long horizontal cylinder 0.45 0.13
Forced Convection 497

Table 19-4
Values of a = gf3p2cp/pk

Temperature (°C) Air at Atmospheric Pressure

-10 168
0 143
10 120
20 102
30 86
40 74
50 65
60 57
70 50
80 44
90 38
100 34

Note: (aL~ AT) is a dimensionless group. L is in meters and AT is


in °C.

Solution. The mean film temperature is 38°C. From Table 19-4, a = 76 x 106
(by interpolation)

{aV AT) = 76 x 106 x (0.05)3 x 24 = 2.28 x 105(< 107)

thus the flow is laminar. From Table 19-3, C = 0.45. Using Eq. 19-23, with
the value of k from the Appendix,

hc = C j-(aL3 AT)m = 0.45(°<.'/0|^^(228.0()0)'M = 5.34 W/m2K

Answer

19-10 FORCED CONVECTION


When heat transfer rates from natural convection are not sufficiently high,
forced convection is used. As the name implies, in forced convection, the
fluid is moved mechanically, generally by pumps, fans, blowers, or com¬
pressors. In most forced convection heat transfers, the flow is turbulent and
will be so assumed here. Flowever, when the fluid is quite viscous (for
example, heavy oils, particularly at low temperatures), the flow may be
laminar. In case of doubt, the Reynolds number should be calculated to make
certain as to the type of flow taking place.
One of the most common types of forced convection heat transfer is that
occurring inside pipes or ducts. In this section we are concerned largely with
this type of heat transfer. However, the general principles developed here are
applicable to other types of forced convection heat transfer problems.
498 Elements of Heat Transfer

The value of the film coefficient, hc, for forced convection depends on the
diameter of the passage D, the fluid velocity Y, the mean stream density p,
and the following properties of the mean film temperature: viscosity /x,
specific heat cp, and thermal conductivity k. Dimensional analysis may be
used, in a manner similar to that used in considering free convection, to
combine the variables into dimensionless groups, thus greatly reducing the
experimental work required to determine the interrelationship of the various
variables with hc.
As a result of dimensional analysis, the following equation has been
developed for turbulent flow,

(19-24)

Often, there is a variation in both the fluid density p and the velocity Y as the
fluid flows through its channel. However, for steady stable flow, the product
pY( = G) remains a constant, provided that the flow area is constant. The fluid
density to be used in determining G must be the mean stream density at the
point where Y is determined.
The values of the constant C and of the exponents e and / in Eq. 19-24 are
difficult to determine exactly but the most widely accepted values are:
C = 0.023, e = 0.8 and / = 0.4. Using these values, Eq. 19-24 becomes

(19-25)

where the subscript / denotes that the properties to be used are those at the
mean film conditions. It is to be noted that hcD/k is the Nusselt number Nu,
DG/p, is the Reynolds number Re, and cp/x/k is the Prandtl number Pr.
At this point it should be emphasized that experimental results obtained by
careful investigators may differ from the results computed by applying Eq.
19-25 by at least 25 percent. Some reasons for this difference are:

1 The assumption of a stagnant film is incorrect.


2 In many instances it is very difficult to evaluate the true mean film proper¬
ties. This is particularly true when there are wide variations in temperature
throughout the film.
3 It is very difficult to measure either the true temperature drop across the
film or the rate of heat flow.
4 An oxide or other type of scale may form on the heat-transfer surface.
Often the scale resistance cannot be readily determined separately and,
hence, it is generally added to the resistance of the fluid film.

Inside film coefficients may vary from less than unity to several thousand.
Although Eq. 19-25 does not always give the desired accuracy, in some cases
it does narrow the expected range of values for the film coefficient, which
may be satisfactory for engineering use if a small safety factor is included.
Forced Convection 499

Example 19-7. Water flows through a standard 2-in. pipe with a mean
velocity of 1.83 m/s. The mean water temperature is 38°C. The mean tem¬
perature of the inner surface of the pipe is 71°C. Determine the inside film
coefficient.

Solution. Use Eq. 19-25. From the Appendix, the internal diameter of the
2-in. pipe is 2.067 in. = 5.25 cm. Also from the Appendix, the density of water
at 38°C is 0.993 g/cm3. Then

G = pY =- jqqq-— = 0.1817 kg/cnrs

= 1817 kg/m2s

From the Appendix, the properties of water at the mean film temperature of
54.5°C are:

k = 0.652 W/mK; /x = 0.0005116 kg/ms; cp = 4158 J/kg-K.

Solving Eq. 19-25,


0.8
/ 5.25
x 1817
0.652 100 4.158x0.0005116V
hr = 0.023
5.25 \ 0.0005116 ( 0.652 )
100
= 7542 W/itrK Answer

Equation 19-25 breaks down for fluids with Prandtl numbers less than 0.1.
This is particularly true for the alkali liquid metals such as sodium, potassium,
and lithium. These fluids have very high thermal conductivities and, hence,
much of the heat is transferred by conduction across the film rather than by
momentum transfer. Lyon applied an analysis made by Martinelli and
presented the so-called Lyon-Martinelli equation

Nu = 7 + 0.25 Pe°8 (19-26)

where Pe is the Peclet number, which equals Re x Pr = DYpcplk.


This equation was derived on the basis of uniform heat flux. It has been
found to hold reasonably well for sodium, potassium, or sodium-potassium
alloy when the substance is in the pure state, but gives values much too high
when the substance contains more than traces of impurities. Experimental
evidence indicates that it predicts values that are too high for other liquid
metals. Lubarsky and Kaufman have presented the following empirical equa¬
tion, which is said to hold reasonably well for liquid metals containing slight
amounts of contaminants:

Nu = 0.625 Pe04 (19-27)

Example 19-8. Liquid sodium, having a mean temperature of 400°C, flows


through a tube having an internal diameter of 1.5 cm with a velocity of
160 cm/s. Determine the film coefficient.
500 Elements of Heat Transfer

Solution. It will be assumed that the sodium is very clean and, hence, Eq.
19-26 will apply.
^DTpCpj
Nu = 7 + 0.025

/0.015 x 1.6 x 854,000 x 1.278


= 7 + 0.025
\ 71.2
= 7 + 0.025(368)°8 = 9.82
Nu k 9.82x71.2 ^ //\a mi 2tt-
hc — — = 46,600 W/m K Answer
D

This example illustrates the fact that very high film coefficients are obtain¬
able with the liquid alkaline metals.
Extensive investigational work has not established an accurate means of
evaluation of film coefficients outside the tubes. For flow across a single tube
or wire, the following equation seems to be the best available one for gases:

Nu = 0.3 Re0'57 (19-28)

In this equation, the outside diameter is to be used in both the Nusselt and
Reynolds numbers.
For flow across a bank of 10 rows of tubes, the following equation is
suggested:

Nu = 0.33 Re06Pr°'33 (19-29)

In this equation, the velocity to be used in finding the Reynolds number is the
velocity at the minimum cross-sectional area.

19-11 FILM COEFFICIENTS FOR PHASE CHANGES


Many industrial heat-transfer problems involve either condensation or boiling.
Although these problems may be considered as special problems involving
convective transfer of heat, the mechanism of heat transfer for either con¬
densation or boiling introduces variables that have not been considered
previously.
When a cool surface is coated with a suitable promoter, so-called dropwise
condensation occurs. The condensate does not wet the surface but collects in
drops that fall off the surface as soon as they become large enough. In the
absence of a promoter, so-called film condensation occurs, in which case the
condensate forms a continuous film over the heat-transfer surface. Since the
film of condensate offers considerable resistance to heat transfer, the film
coefficient for dropwise condensation may be several times larger than that
for film condensation.
Although various promoters, which are added to vapors to produce drop-
wise condensation, have been tried over a period of years, none has yet
proved to be of commercial importance. Many of these promoters are added
to the vapor continuously and, hence, do not permanently coat the surface.
Film Coefficients for Phase Changes 501

Some investigations have been made using surfaces coated with tetrafluoro-
ethylene(teflon). Satisfactory dropwise condensation has been obtained, but
the teflon coating must be made exceedingly thin, since it has poor thermal
conductivity. More recently, dropwise condensation has been obtained by
plating surfaces with very thin layers of gold or silver. At the present time,
however, it must be assumed, unless otherwise stated, that the condensation
is of the film type.
Nusselt studied film-type condensation with laminar flow of the condensate
over the heat transfer surface and arrived at a relationship by dimensional
analysis. His equation is

(19-30)

where hc the film coefficient W/m2K


C 1.666 for a vertical surface and 1.282 for a horizontal tube
L height, in meters, for a vertical surface or outside diameter for a single
horizontal tube or the sum of the diameters of horizontal tubes in a vertical
row
k thermal conductivity at the mean film temperature W/rnK
P density at mean film temperature g/m3
P viscosity at mean film temperature, g/m-s
hfg latent heat at condensing temperature, J/g
AT condensing temperature minus surface temperature, °C

Note: Equation 19-30 is based on the standard gravity acceleration. The value
of hc varies as the \ power of the gravity acceleration. Equation 19-30
applies to natural convection of a vapor over a surface which is not
influenced by the presence of other surfaces. The vapor must be
substantially free of noncondensible gases.

Example 19-9. Steam at a temperature of 104° C condenses on the outer


surface of a horizontal tube having a diameter of 0.05 m. The outer surface of
the tube is 93°C. Determine the film coefficient.

Solution. The mean film temperature is 98.5°C. From Eq. 19-30.


1/4
0.6823 x 950,3003 x 2245\
hr = 1.282
( 0.05 x 0.2893 x 11 )
10,220 J/s-m2K

10,220 W/m2K Answer

Boiling is perhaps the most complicated and least understood of all modes
of heat transfer. Consider a heating surface immersed in a liquid. Keep the
temperature of the heating surface a very few degrees above the saturation
temperature for the pressure at the surface of the liquid. By convective ac¬
tion, the heated liquid will rise, and bubbles of vapor will be formed in the
liquid but not on the heating surface. This type of boiling resembles heat
502 Elements of Heat Transfer

transfer by natural convection. The film coefficient is approximately propor¬


tional to (AT)1/3. An increase in the temperature difference between the
heating surface and the liquid results in bubble formation on the heating
surface. This action is termed nucleate boiling. Under these conditions there
is a very large increase in the film coefficient. For nucleate boiling it appears
that the film coefficient is proportional to (AT)", where the value of n may
range, roughly, from 2 to 4. As the temperature difference is further in¬
creased, bubbles occupy a larger part of the heating surface and ultimately
blanket it completely. Under these conditions there is a decrease in the film
coefficient for boiling.
Most of the heat transfer during boiling occurs in the nucleate region. In
this region, the film coefficient is greatly influenced by the rate of bubble
release. Many factors affect the rate of bubble release. Perhaps one of the
most important factors is the cleanliness of the surface. Observations show
that a very thin coating of oxide on a surface may greatly reduce the film
coefficient. Other factors to be considered are the shape of the heating
surface, the arrangement of the surface, the pressure on the liquid, the
surface tension, viscosity, and thermal conductivity of the liquid, and the
amount of condensable gases present. Although attempts have been made to
do so, no one equation can be devised which will account for the effects of all
the variables on the film coefficient for boiling. The evaluation of surface
conditions is especially difficult.
Because of the impossibility of formulating one equation which is valid for
all types of boiling heat transfer, very many equations have been established,
with each equation applying to only one specific boiling heat transfer. The
reader should refer to heat transfer texts, like those listed in the Appendix,
for some of these equations.

19-12 OVERALL COEFFICIENT OF HEAT TRANSFER


Thus far the various modes of heat transfer have been considered separately.
Although the engineer must be concerned with each of these modes, he or she
must also be able to combine whichever modes exist together in a given
problem to predict the amount of heat that will be transferred.
For any problem of steady-state heat transfer, the rate of heat flow in any
part of a series path must equal the total heat flow rate. Consider two gases
separated by a thin metal wall, as is shown in Figure 19-9. The steady-state
heat flow from one gas to the second gas can be determined by finding the
heat flow through the metal wall. However, since the thermal conductivity of
the metal is high, there is very little temperature drop across the wall.
Sometimes, this drop may be as low as one or two degrees. Furthermore, it is
difficult to measure surface temperatures accurately. Hence, there may be an
error of 100 percent, or even more, in the calculated heat flow. The tem¬
peratures that can be measured with some degree of accuracy are those of the
gases. This section deals with the development of equations for calculating
rates of heat transfer when overall differences in temperature are known.
Overall Coefficient of Heat Transfer 503

■u

FIGURE 19-9 Series heat flow through a wall.

The rate of heat flow through the wall in Figure 19-9 is proportional to the
cross-sectional area A in the path of heat flow and to the temperature
difference between the two fluids. This statement may be expressed by the
equation

Q = UA(T\ - T4) (19-31)

where the proportionality constant U is known as the overall coefficient of


heat transfer.
A method of determining the overall coefficient of heat transfer may be
developed by considering the heat flow through each film and through the
metal wall. The heat flow through the outer film is
Q = h0A(Tl-T2)

The heat flow through the metal wall is

0 = |m(T2-T,)

The heat flow through the inner film is

Q = hiA(T3-T4)

When each of the foregoing equations is solved for the temperature


difference, and use is made of the fact that the sum of the temperature
differences equals the overall temperature difference, the result is

T\-T4 = (Tx- T2) + (T2- T3) + (T3- T4)

= Q(1 + l
+
A\h0 kJX i)

or

Q = A |(T, - T4)
1
U + ±
h0 kJX hi
504 Elements of Heat Transfer

When this equation is compared with Eq. 19-31, it is seen that

u = 1-\-r (19-32)

Fo + kjx + hi
Although Eq. 19-32 gives a mathematically correct method of obtaining the
overall coefficient of heat transfer, its use necessitates the evaluation of the
two film coefficients and the conductance of the metal wall. When small
temperature drops are involved, these factors may be determined with a
reasonable degree of accuracy. On the other hand, when the temperature
drops are very large, the mean surface temperatures must be determined by
trial and error in order that the mean film temperatures and, hence, the mean
film coefficients can be obtained.
The overall coefficient of heat transfer between two fluids separated by a
composite wall may be derived in a manner similar to that just described. For
this reason,

1
U = (19-33)
1 1 1 +1
+
ho kmAlkm !Xg kmJXn h,

A method of determining the overall coefficient of heat transfer between


two fluids separated by a cylinder, as in Figure 19-10 may be developed as
follows. For the inner film,

Q= hiir D,L(Ti - T2)


for the cylinder,

\ = 2irLkm(T2- T3)
1 ln(D0/D,)
For the outer film,

Q = h0ir D0L(T3 — T4)

Q FIGURE 19-10 Series heat flow through a cylinder.


Overall Coefficient of Heat Transfer 505

Then

T, - T4 = (T, - T2) + (T2 - T3) + (T3- T4)


_ Qr 1 , lnCDo/Dj) | 1 -
7rLLhiDi 2km h0D0_
or

^ = Wl 1
\ hjDi
ln(D0/D|)
2km
1 VT| ~
h0 D0 /
Ti)

When this equation is equated to Eq. 19-31, it becomes necessary to decide


whether to use the inside area or the outside area. If the outside area 7tD0L is
used,

(19-34)
Dp _|_ Dp ln(D,/Dp) + 1
h(D, 2/cm h0
where l/0 is the overall coefficient based on the outside area.
In a similar manner, the overall coefficient based on the inside area of ttD.L
is

(19-35)
1 | D, ln(D,/D0) [ D,
h/ D0h0
Example 19-10. Water flows through a 2 in. standard weight pipe with a mean
velocity of 1.83 m/s. The mean water temperature is 60°C. The water is heated
by condensing steam having a temperature of 104°C. The inside film
coefficient is 6805 W/m2K and the outside film coefficient is 8517W/m2K.
Determine the heat transferred per meter length of pipe.

Solution. From the Appendix, D0 = 2.375 in. = 0.06033 m and D, = 2.067 in. =
0.0525 m. From Eq. 19-34,
1
0.06033
0.06033 In
0.06033 0.0525 1
+ T
0.0525 x 6805 2x50.55 8517
1 _ 1
~ 0.0001689 + 0.000083 + 0.0001174 0.0003693
= 2708 W/m2K Answer
Area per meter = 0.06033 x 7r = 0.1895 m2/m
Q = 2708 x 0.1895(104 - 60) = 22,580 W/m-length Answer

Example 19-11. Determine the inner and outer pipe surface temperatures of
Example 19-10.
506 Elements of Heat Transfer

Solution. From Example 19-10, the total resistance is 0.0003693, the resis¬
tance of the outer film is 0.0001174, and the pipe wall resistance is 0.000083.
Since the temperature drop is proportional to the resistance, the temperature
drop across the outer film,

AT = (O003693)(104-60)= 14 C
The temperature drop across the pipe,

/ 0.000083
AT V104-6O) 9.9° C
V0.0003693

The outer pipe wall temperature = 104- 14 = 90° C Answer

The inner pipe wall temperature = 90 - 9.9 = 80.1°C Answer

19-13 MEAN TEMPERATURE DIFFERENCES FOR


PARALLEL FLOW AND COUNTERFLOW
HEAT EXCHANGERS
In Section 19-12 we assumed that either the temperatures of the two fluids
were constant throughout or that their true mean temperatures were known.
Under these conditions, the temperature difference to be used in calculating
the heat flow can be obtained by simple subtraction. In engineering situations,
the determination of the temperature difference is generally much more
difficult.
Consider two fluids flowing in the same planes either in the same direction
(parallel flow) or in opposite directions (counterflow). When a fluid changes
phase, there is no change in temperature provided the pressure is held
constant. When this condition exists, it makes no difference as to which
direction the fluid is flowing. On the other hand, when neither fluid changes
phase, the direction of flow is very important. This condition will now be
considered.
In the simple case, the fluids will be considered to flow in the same planes.
If they also flow in the same direction, the flow is said to be parallel flow.
When the two fluids flow in opposite directions, the flow is termed coun¬
terflow. Counterflow is illustrated in Figure 19-11, and parallel flow is illus¬
trated in Figure 19-12.
It should be noted for counterflow (Figure 19-11) that as the flow passage is
lengthened, the temperature difference between fluids A and B at any point in
the flow passage becomes smaller and smaller. Thus, with a very long flow
passage, the temperature of fluid B at its exit approaches that of fluid A at its
entrance. Likewise the temperature of fluid A at its exit approaches that of
fluid B as it enters. As the length of the flow passage increases without limit,
the temperature difference between the two fluids in any point in the flow
passage approaches zero and the heat transfer process approaches a rever¬
sible one.
Mean Temperature Differences 507

FIGURE 19-11 Counterflow.

On the other hand, for parallel flow as shown in Figure 19-12, the tem¬
perature of the two fluids approaches each other at the exit. The longer the
flow passage, the closer together the temperature of the two fluids will be at
exit. However, the parallel flow heat transfer process cannot approach rever¬
sibility unless the temperature difference between the two fluids is small at
entrance.
It is apparent from examination of Figures 19-11 and 19-12 that the true
mean temperature difference of two fluids in a heat exchanger, where there is
no change in phase, cannot be determined by simple subtraction of tem-

FIGURE 19-12 Parallel flow.


508 Elements of Heat Transfer

AWWWWVWWWWWW^WWWWWWVXWWWWWKWVWWW^X^^W'^W^'^

—>- ->

l r

FIGURE 19-13 Log mean temperature difference.

peratures. For a method of determining the true mean temperature difference,


refer to Figure 19-13. The following simplifying assumptions will be made:

1 There is no heat exchange with the surroundings.


2 The specific heats of the two fluids are constant.
3 The overall coefficient of heat transfer is uniform throughout the heat
exchanger.

Consider a small section of the heat exchanger, having a heat transfer area
dA. Use the symbols h and c to designate the hot and cold fluids, respec¬
tively. Then the heat transfer is
dQ = U dA(Th - Tc) = mhcPh dTh = rhccPc dTc (19-36)

It follows that

1 _ dTh
mhcPh U dA(Th - Tc)
and
1 = —dTc
mccPc U dA(Th - Tc)
Hence,
1 , 1 = d(Th-Tc)
(19-37)
mhcPh mccPc U dA(Th - Tc)
Mean Temperature Differences 509

Also

0= UA ATm = rinhcPh(Tl - Tr)h = mccPc(T, - Tr)c (19-38)

where subscript / refers to the left side of the heat exchanger, and r refers to
the right side.
From Eq. 19-38,

1 = (T| - Tr)h
mhcPh UA A Tm

and

1 _ (T, - Tr)c
mccPc UA ATm

Hence,

1 , 1 _(T,-Tr)h-(T,-Tf)c
(19-39)
mhcPh mccPc UAATm

Combining Eq. 19-37 and Eq. 19-39 gives

d(Th - Tc) = (Tt — Tr)h - (T[ — Tr)c


UdA(Th-Tc) L7A ATm
Then

-1 d(T„ - Tc) (T, - T,)h - (T, - Tr)c [‘


r Th — Tc A A Tm Jr dA
or

r(Th-rc),1 = [~(Tf - Tr)h — (T| — Tr)c


L (h - c)r j L A ATm
hence,

(Th - Tc)i - (Th - Tc)r


ATm = LMTD (19-40)
ln[(Th - Tc),/(Th - Tc)r]

vSince Eq. 19-40 contains the logarithm of a temperature ratio, the temperature
difference ATm computed by this equation is known as the log mean tem¬
perature difference (LMTD).
The mean temperature difference may also be approximated by assuming
that the temperatures vary in straight lines throughout the heat exchanger.
The temperature difference based on this assumption is known as the arith¬
metic mean temperature difference. Since this arithmetic mean is only an
approximation, its use is not recommended, unless the temperature changes
of both fluids are very small.
The expression for the log mean temperature was derived for counterflow.
As may be shown, it holds true equally well for parallel flow and for a heat
exchanger in which one of the fluids changes phase. Temperature variations
510 Elements of Heat Transfer

FIGURE 19-14 Heating with condensing vapor.

FIGURE 19-15 Cooling with vaporing liquid.

throughout the heat exchanger are illustrated in Figure 19-14 for a fluid being
heated by a condensing vapor. In Figure 19-15 are illustrated temperature
variations for the vaporization of a liquid caused by the cooling of a second
fluid.

Example 19-12. Water is heated in a double pipe heat exchanger from 38° C
to 205° C by gases that cool from 425° C to 150°C. Determine the log mean
temperature difference.

Solution. Since the exit water temperature cannot be greater than the exit
gas temperature for parallel flow, this flow must be counterflow. (If the exit
water temperature was less than that of the gas at exit, the flow could be
either parallel or counterflow.) Use Eq. 19-40 and refer to Figure 19-11

(425-205)-(150-38)
LMTD = 160°C Answer
425 - 205
In
150-38
Multipass and Crossflow Heat Exchangers 511

19-14 MULTIPASS AND CROSSFLOW HEAT


EXCHANGERS
In many types of heat exchangers, the two fluids do not flow in parallel planes
throughout the heat exchanger. Hence, Eq. 19-40 is not applicable for finding
the log mean temperature difference. A simple form of such a heat exchanger
is illustrated in Figure 19-16. This heat exchanger provides for a single tube
pass, as the fluid inside the tube flows in one direction. It also provides for a
single shell pass, as the fluid in the shell flows in one general direction. Heat
exchangers may also provide for more than one tube pass and for several
shell passes. It is difficult to predict the true mean temperature difference for
a crossflow heat exchanger, since the number and arrangements of the baffles
have an influence on the flow pattern. In addition, since it is necessary to
allow for expansion, tight contact between the tubes and the baffles is
precluded and an unpredictable amount of leakage will occur around the
tubes. A sizable amount of tube leakage will affect the true mean temperature
difference. However, this whole problem has been investigated by many
individuals, and much information is available in literature.
Bowman, Mueller, and Nagle investigated much of the published in¬
formation on crossflow heat exchangers and presented graphs for selecting a
correction factor F to be applied to the log mean temperature difference
computed by Eq. 19-40 for counterflow, to obtain the true mean temperature
difference.6 Thus,

true mean temperature difference = Fx LMTD (19-41)

Three of the several graphs prepared by Bowman, Mueller, and Nagle are
presented here as Figures 19-17 to 19-19. These graphs cover some of the

6R. A. Bowman, A. C. Mueller, and W. M. Nagle, “Mean Temperature Differences in Design,”


Trans. ASME, Vol. 62 (1940), pp 283-294.

FIGURE 19-16 Cross-flow heat exchanger.


T,
_1
u. — ; u 12
h

t2

FIGURE 19-17 Correction factor for heat exchangers with 1 shell pass and 2,4, or any multiple of 2
tube passes.

FIGURE 19-18 Correction factor for heat exchangers with 2 shell passes and 4, 8, or any multiple
of 4 tube passes.

t2

FIGURE 19-19 Correction factor for heat exchangers with 3 shell passes and 6,12, or any multiple
of 6 tube passes.

512
Applications 513

more simple types of heat exchangers and are to be used as approximations


when more accurate information is not available.

Example 19-13. Oil is cooled from 93° C to 38° C in the tubes of a heat
exchanger having a single shell pass and two tube passes. Water in the shell
side of the heat exchanger is heated from 27° C to 43° C. Determine the true
mean temperature difference.

Solution. Refer to Figure 19-17.


(27 - 43)
R = : 0.29
38 - 93
38-93
V = 27 - 93' 0.83

From Figure 19-17, F = 0.62.


true mean temperature difference F x LMTD
(38 - 27) - (93 - 43)
= 0.62 x
38-27
In
93-43
= 16.0°C Answer

19-15 APPLICATIONS
A very large number of heat-transfer problems involve the transfer of heat
from one region to another region. Two different classes of problems are
involved. In a problem of the first class, there is a desire for operation at a
temperature either above or below that of the surroundings. Many chemical
processes take place only at high temperatures and the substances involved
must, therefore, be heated. On the other hand, heat must be removed from
food so that it may be frozen and thus preserved. Because of the limitations
of the second law of thermodynamics, the working substance of a heat engine
must be heated to as high a temperature as possible, consistent with the
reliability of operation of the heat engine.
In a problem of the second class, heat must be removed from a region in
order that the temperatures of various pieces of equipment will not exceed
safe operating values. Heat must be removed from an internal-combustion
engine to prevent burning of the parts, excessive thermal stresses that may
cause failure of the parts, the deterioration of the lubricant. Present-day
practice is to miniaturize electronic equipment. This practice reduces the
surface area available for natural dissipation of heat and, hence, limits the
power of the equipment as well as its life. During the early setting of concrete
in a large dam, there is a relatively great rise in temperature. Hence, heat must
be removed rapidly to reduce stresses during the later cooling process that
might result in failure. Heat must be removed from the fuel elements of
nuclear reactors to keep their temperatures down to safe values. These are
514 Elements of Heat Transfer

but a very few of the many examples of how the capacity of engineering
equipment is limited by the rate at which heat can be removed to prevent
temperatures from exceeding safe values.
In every problem involving heat transfer, the engineer is confronted with
the selection of a suitable method of transferring heat from its source to
another region. As in all other engineering problems, the designing engineer
must consider initial costs, cost of operation, ease of maintenance, and
perhaps weight and size of equipment. Vital as these considerations are, they
are outside the scope of thermodynamics and will not be discussed further
here.
When the temperature in a region cannot be kept down to a safe value by
natural removal of heat, several possibilities are open to the engineer.

1 He may provide a better heat-flow path between the source of heat and the
region to which it is to be transferred.
2 He may resort to forced convection of a fluid over the source of heat.
3 He may increase the heat transferred by radiation by increasing the
emissivity of the radiation surface.

It is not the purpose of this section to describe in detail the methods for
reducing temperatures of various pieces of equipment. The subject of heat
removal is discussed here merely to emphasize the applications of the
principles covered in this chapter.
Consider, for example, a piece of electronic equipment enclosed in a
container of some sort. Heat is transferred from the equipment to its con¬
tainer mainly by radiation and natural convection, and this heat is then
dissipated to the surroundings. When the equipment is miniaturized to save
space, the surface area available for transferring heat is reduced and, hence,
the temperature must increase in order that the heat may be transferred.
When a maximum allowable temperature is reached, means must be found to
increase the rate of heat transfer. One such method is to insert some highly
conductive metal, such as aluminum, between the equipment and its con¬
tainer. When this is done, care must be taken to ensure good contact with the
metal, since air spaces are good insulators.
In general, liquids are much better heat-transfer agents than is air. As we
discuss in Section 19-11, film coefficients for boiling liquids are high. This fact
suggests the possibility of using a liquid that will boil at a desired temperature
at a reasonable pressure. The boiling of the liquid will cool the electronic
equipment, and the vapor will condense on the containing wall, which will
thus give up heat more rapidly. Hydrogen, particularly under pressure, is a
good heat-transfer agent. It is used to cool electric generators by forced
convection. Hydrogen pressures as high as 400-kPa gage have permitted very
large increases in the capacities of electric generators without increases in the
physical sizes of the units.
Problems 515

PROBLEMS _____
19-1 The rate of heat flow through a 25-cm wall composed of ordinary brick
is 237,000 J/m;-hr. The outerface temperature of the wall is 51°C.
Determine the inner face temperature of the wall.
19-2 The reported rate of heat flow through a 0.48 cm thick steel plate is
7.5 x 107 J/m2-hr at room temperature. The temperature drop across the
steel plate is reported to be 1.7°C. Make necessary calculations and
comments on the accuracy of the reported data.
19-3 An older furnace wall is composed of 8 in. of fireclay brick, 4 in. of
mineral wool, and 4 in. of common brick. The surface temperature of
the common brick is 60° C and the interface temperature between the
common brick and the mineral wool is 103° C. Determine the surface
temperature of the fireclay brick and the rate of heat flow.
19-4 Liquid metal flows through a 6 in. extra heavy steel pipe. It has a mean
temperature of 595° C. Assume that this is also the innerface tem¬
perature of the pipe. The pipe is insulated with 4 in. of 85 percent
magnesia. The temperature of the outer surface of the insulation is
51°C. Determine the inner face temperature between the magnesia and
the pipe.
19-5 Steam, having a temperature of 215°C flows through a 4-in. pipe. The
pipe is covered with 2 in. of insulation. The temperature of the outer
surface of the insulation is 49°C. Steam condenses in the pipe at the
rate of 165 g/hr per meter length. Determine the thermal conductivity
of the insulation assuming that the inner pipe wall temperature equals
that of the steam.
19-6 A glass electronic tube has a mean surface temperature of 61°C. Its
surface area is 198 cm2. The mean temperature of the surroundings is
30° C. Determine the rate of heat lost from the tube by radiation.
19-7 If the tube in Problem 19-6 can safely operate at 75°C, determine the
minimum surface area if the tube is to lose the same amount of heat as
in Problem 19-6, all by radiation.
19-8 The temperature of the outer surface of a steel vessel (rough steel) is
465° C when the temperature of the surroundings is 26° C. The surface
area is 41,500 cm2. Determine the reduction in the heat lost by radia¬
tion if the tank is painted with aluminum paint (a) after the paint is first
applied, and (b) after the paint has oxidized.
19-9 Two parallel squares, A and B, 212 cm on a side, are 365 cm apart.
Square B, painted with a medium shade of oil paint, has a surface
temperature of 28° C. The net radiation received by square B is
173 kJ/h. Determine the emissivity of square A if its temperature is
49° C.
516 Elements of Heat Transfer

19-10 A radiator supplies 54 percent of its heat by radiation. The radiator is


painted with an oil paint. Its surface is 98° C when the room tem¬
perature is 20° C. Determine the surface area required if the radiator is
to supply 29,500 kJ/h.
19-11 Determine the reduction in heat flow from the radiator in Problem
19-10 when it is painted with aluminum paint.
19-12 An object has a surface area of 2620 cm2, a surface temperature of
215°C, and an emissivity of 0.87. The emissivity of the surrounding
objects is 0.93, with a mean temperature of 28°C. It is questionable
whether the object is large or small in comparison with its surround¬
ings. Calculate the heat lost for both cases and compare the results.
19-13 A pipeline is covered with insulation protected by aluminum sheath¬
ing. The outer diameter is 27 cm and the length is 85.3 m. The outer
surface temperature is 65° C. The temperature of the surroundings is
26° C. Determine the rate of heat lost by radiation.
19-14 Water fills a steel vat. The mean water temperature of the water is
68° C and the mean sidewall temperature is 88° C. Each side wall is
120 cm x 190 cm. Determine the heat exchange by convection between
the walls and the water.
19-15 A tube is placed in a vacuum chamber. The initial pressure in the
chamber is 101 kPa. For this condition the convection film coefficient
is 1.64 J/cm2-hr-°C. Determine the convection film coefficient when the
pressure is reduced to 2.06 kPa.
19-16 When the product (Gr x Pr) equals 109, should the values of hc deter¬
mined for laminar and turbulent flow be equal? Are they? If not, why?
19-17 Assume that the air surrounding the pipe in Problem 19-13 has a
temperature of 26°C. The pipe is a horizontal one. Determine the heat
lost by convection and compare it with that lost by radiation.
19-18 Air at a mean pressure of 102 kPa flows through a duct having a
diameter of 25 cm with a velocity of 40 m/s. The mean air temperature
is 45°C, and the mean duct wall temperature is 60°C. Determine the
rate of heat transfer between the duct and the air per meter of length.
19-19 The same as Problem 19-18 except that the mean duct wall tem¬
perature is 75°C. Compare the answers.
19-20 Oil flows through a pipe with a velocity of 240 cm/s. Assuming all
temperatures remain constant, determine the velocity required to
increase the heat exchange between the pipe and oil by 50 percent,
19-21 Water flows through a 4-in. pipe with a velocity of 150 cm/s. The mean
water temperature is 35°C and the mean pipe wall temperature is 50°C.
Determine the heat transferred per meter length of pipe.
19-22 The same as Problem 19-21 except that the mean pipe wall tem¬
perature is 80°C. Compare the results.
Problems 517

19-23 Liquid sodium flows through a 6 in. extra-heavy (schedule 80) steel
pipe with a velocity of 300 cm/s. The mean sodium temperature is
500°C. Assuming that the sodium is of average purity, determine the
film coefficient.
19-24 Steam condenses on the outside surface of a vertical tube that is
250 cm in length. The outside diameter of the tube is 25 cm. The
temperature of the condensing steam is 110°C, and the mean outer-
surface tube surface is 90° C. Determine the rate of heat exchange
between the steam and the tube.
19-25 The same as Problem 19-24 except that the mean tube wall surface
temperature is 70°C. Compare the results.
19-26 Steam condenses on the outside of a horizontal tube having a diameter
of 2.8 cm. The condensing temperature is 50°C and the value of the
film coefficient is 2870 J/cm2-hr-°C. Determine the temperature of the
tube surface.
19-27 Steam enters a condenser at a pressure of 3 kPa, at a rate of
1,720,000 kg/h. Its enthalpy is 2100 J/g. The tubes (horizontal) have an
outside diameter of 2.5 cm and a mean surface temperature of 15°C.
(a) Determine the condensing film coefficient, neglecting the presence
of other tubes.
(b) Determine the surface area required. Assume that the condensate
leaves at condensing temperature.
19-28 Heat is supplied during a boiling process at the rate of 73,200 kJ/h. The
surface area is 630 cm2 and the temperature difference is 28°C.
Determine the film coefficient and compare it with the coefficients for
other types of heat transfer.
19-29 The rate of water flow in Example 19-10 in the Chapter discussion is
increased, changing the rate of heat transfer to 27,500 W/m length.
Determine (a) the new overall coefficient of heat transfer, and (b) the
new inside film coefficient.
19-30 Water flows through a 2 in. standard weight pipe. Its mean tem¬
perature is 60° C. It is heated by condensing steam having a tem¬
perature of 104°C. The outer film coefficient is 8580 W/m-K. The heat
transferred is 26,000 W/m length. Determine (a) the overall coefficient
of heat transfer, and (b) the inside film coefficient.
19-31 Water enters a double pipe heat exchanger at 21°C and leaves at 82°C.
The water is heated by gases that enter at 149°C and leave at 93°C.
Determine the log mean temperature difference for (a) parallel flow,
and (b) counterflow.
19-32 Steam condenses in a heat exchanger at 82°C. Water enters at 43°C
and leaves at 67°C. Determine the log mean temperature difference. Is
it necessary to know the direction of steam flow in the heat
exchanger?
518 Elements of Heat Transfer

19-33 Water flows through a 2 in. standard weight pipe with a velocity of
198 cm/s. The mean water temperature is 60°C. The water is heated by
steam condensing at 120°C. The overall coefficient of heat transfer is
10,500 kJ/m2-h-°C. Determine the actual condensing film coefficient.
Suggestion: assume a mean water side film temperature and then
check it.
19-34 The steam pressure in a condenser is 10 kPa. It enters the condenser
at 50°C. Condensate leaves at 40°C. Circulating water enters the
condenser at 35°C and leaves at 43°C. Determine the approximate
mean temperature difference. Note: How much of the heat given up by
the steam is given up as it condenses?
19-35 The regenerator of a gas turbine unit has an overall coefficient of heat
transfer of 100.1 kJ/m2-h-°C. Gases enter at 550°C and leave at 285°C.
Air enters at 160°C and leaves at 435°C. Determine the heat transfer
area required if the rate of heat transfer is 16,800,000 kJ/h.
19-36 A totally enclosed air-cooled motor has a rate output of 75 kW and an
efficiency of 94 percent. The overall coefficient of heat transfer from
the motor to the air is 139 kJ/m2-h-°C. Air enters at 27°C and leaves at
32.5°C. The mean surface temperature of the motor is 75°C. Deter¬
mine (a) the surface area required, and (b) the mass rate of air flow
required.
19-37 Water flows through a 1-in. pipe with an average velocity of 240 cm/s.
Its average temperature is 32° C. The pipe is immersed in condensing
steam whose temperature is 104°C. The steam side film coefficient is
30,600 kJ/m2-h-°C. Determine the overall coefficient of heat transfer.
19-38 The pipe in Problem 19-37 is coated with 0.0025 cm of teflon to
promote dropwise condensation. The thermal conductivity of the
teflon is 0.87 kJ/m2-h-°C/m. The new steam-side film coefficient is
140,200 kJ/m2-h-°C. Determine the overall coefficient of heat transfer.
19-39 The same as Problem 19-38 except that the thickness of the teflon is
0.0005 cm.
19-40 Air enters a single shell-two pass intercooler at 82°C and leaves at
26°C. Water enters at 15°C and leaves at 32°C. Determine the true
mean temperature difference.
19-41 The same as Problem 19-37 except that the water velocity is 350 cm/s.
Assume no change in the steam-side film coefficient.
19-42 Water enters a single shell, two-pass heat exchanger at 16° C and
leaves at 40°C. It is heated by gases that enter at 120°C and leave at
38°C. Determine the true mean temperature difference.
_APPENDIXES_

A-l CONVERSION TABLES 521


A-2 ATOMIC MASSES OF COMMON ELEMENTS 525
A-3 STEAM TABLES 526
A-4 SATURATED VAPOR PRESSURES OF REFRIGERANTS 541
A-5 REFRIGERANT FREON-12 TABLES 542
A-6 AMMONIA TABLES 551
A-7 SATURATED MERCURY VAPOR TABLE 555
A-8 PROPERTIES OF GASES 557
A-9 PROPERTIES OF AIR 558
A-10 PROPERTIES OF WATER 559
A-ll PROPERTIES OF LIQUID METALS 560
A-12 THERMAL CONDUCTIVITY OF NONMETALLIC SOLIDS 561
A-13 THERMODYNAMIC PROPERTIES OF AIR AT LOW
PRESSURES 562
A-14 EMISSIVITIES OF SURFACES AT NORMAL
ATMOSPHERIC TEMPERATURE 566
A-15 PIPE SIZES 567
A-16 CHARACTERISTIC EQUATION OF IDEAL GASES
FROM BOYLE’S AND CHARLES’ LAW 568

519
_APPENDIX 1_
CONVERSION TABLES

LENGTH
1 m= 100 cm
1 m = 1000 mm
1 m = 106 microns (/x)
1 cm = 2.54 in.
12 in. = 1 ft
1 ft = 0.3048 m
3 ft = 1 yd.
16.5 ft = 1 rod
1 m= 3.2808 ft
1 m= 1.0936 yd
1 mile = 1.6094 km, 5280 ft
1 nautical mile = 6076.1 ft
1 nautical mile = 1852 m

VOLUME
1 rrr 1,000,000 cm3
1 nr 1000 dm3 (cubic decimeter)
1 liter 1 dm3 a = 1000 cm3
1 in3 16.3871 cm3
1 liter 61.02374 in.3
1 rrr 35.3147 ft3
1 ft3 28316.85 cm3
1 ft3 1728 in.3
1 gas (US) 231 in.3
1 gas (US) 3.7854 liters
1 ft3 7.4805 gal (US)
1 barrel petroleum 42 gal (US)

AREA
1 m2 = 10,000 cm2
1 hectare = 10,000 m2
1 m2 = 6.4516 cm2
1 m2 = 10.7639 ft2
1 m2 = 1.1960 yd2
“Exact by definition of the 1964 General Conference on Weights and Measures.

521
522 Appendix 1

1 acre = 43,560 ft2


1 sq mile = 640 acres
1 acre = 4046.86 m2
1 acre = 0.404686 hectares
1 ft2 = 0.0929 m2

MASS
1 kg = 1000 g
1 metric ton = 1000 kg
1 kg = 2.2046226 lb (avoirdupois)
1 lb = 453.59237 g
1 lb = 16 oz (avoirdupois)
1 ton = 2,000 lb
1 ton = 907.18474 kg
1 metric ton = 2204.6226 lb
1 lb = 7000 grains

TEMPERATURE
°F = 1.8 x °C + 32
°F- 32
or^ _ __ ^^
1.8
r = °F +459.67
K = °C +273.15
R = Kx 1.8

PRESSURE AND FORCE


1 Newton (N) = 1 kg-m/s2
1 dyne = 1 g-cm/s2
1 N = 1 x 105 dynes
1 lb, = 4.44822 N
Standard gravity
acceleration = 9.80665 m/s2
1 N/m2 = 1 Pascal (Pa)
1 bar = 105 Pa = 102 kPa
1 atmosphere = 760 mm mercury at 0°C
1 atmosphere = 101,325 Pa — 1.01325 bars
1 bar = 0.986923 atmospheres
1 atmosphere = 14.6960 lb/in.2
1 bar = 14.50377 lb/in.2
1 lb/in.2 = 6894.757 Pa
1 Pa = 1.450377 x 10 4 lb/in.2
1 N/cm2 = 1.450377 lb/in.2
1 kg/ = 9.80665 N

ENERGY AND POWER


1 Nm = 1 Joule (J)
1 J = 107 ergs
Conversion Tables 523

1 erg = 1 dyne-cm
1 watt (W) = 1 J/s = 1 Nm/s
1 kilowatt (kW) = 1000 w
1 megawatt (mW) = 1,000,000 w
1 International
(Table (I.T.)
calorie = 4.1868 I.T. J
1 I.T. Btu = 251.0958 I.T. calories
1 I.T. Btu = 778.169 ft lb
1 I.T. Btu - 1055.056 J
1 ft-lb = 1.355818 J
1 horsepower (hp) = 33,000 ft-lb/min
1 hp = 42.407 Btu/min
1 hp = 2544.4 Btu/hr
1 kilowatt hour (kWh) = 3412.14 Btu
1 hp = 0.7457 kW
1 Btu/ft2 hr-R 7 5.6780 W/m2K
1 Btu/hr ft-R = 62.303 J/hr-cm-K
1 Btu/lb-R = 4.1866 J/gK

VISCOSITY
1 poise = 1 g/cm-s 7 0.1 kg/ms
1 poise = 100 centipoise
1 poise = 0.1 N s/m2
1 centipoise = 2.4191 lbm/ft-hr *
IK hr
1 centipoise = 5.8015 x 10 3

1 stoke = 100 centistokes


1 stoke 7 1 cm2/s
1 stoke = 0.001076 ft2/s

GAS CONSTANT
N-m
R0 = 8.31434 J/g mole-K = 8.31434 g moie_K
Ro = 1.9858 cal/g mole-K
R0 = 1.9858 Btu/lb mole-R
R0 = 1545.3 ft Ib/lb mole-R
'
_ APPENDIX 2 _
ATOMIC MASSES OF
COMMON ELEMENTS

Atomic Atomic
Substance Symbol Mass Substance Symbol Mass

Aluminum A1 26.98 Molybdenum Mo 95.95


Argon A 39.944 Nickel Ni 58.69
Bismuth Bi 209.00 Nitrogen N 14.008
Bromine Br 79.916 Oxygen O 16.000
Cadmium Cd 112.41 Phosphorus P 30.975
Carbon C 12.011 Platinum Pt 195.23
Cesium Cs 132.91 Potassium K 39.10
Chlorine Cl 35.457 Silicone Si 28.09
Chromium Cr 52.01 Silver Ag 107.88
Cobalt Co 58.94 Sodium Na 22.991
Copper Cu 63.54 Sulfur S 32.066a
Gold Au 197.0 Tantalum Ta 180.95
Helium He 4.003 Thorium Th 232.05
Hydrogen H 1.008 Tin Sn 118.70
Iron Fe 55.85 Titanium Ti 47.90
Lead Pb 207.21 Tungsten W 183.92
Lithium Li 6.940 Uranium U 238.07
Magnesium Mg 24.32 Vanadium V 50.95
Manganese Mn 54.94 Zinc Zn 65.38
Mercury Hg 200.61

‘There may be a variation of 0.003 in this value, since the amount of the various isotopes of sulfur
present may vary.

525
CN r~ oo d- CN o co o oo m
• o NO m oo cn co r- ’t NO 1—H
P Q. 00 m cn oo NO >0 lO m m NO r- ON
cd “ co NT, CN
o ON r-~ NO d- co o ON
m > ON on 00 oo 00 OO od od oo OO oo r-

00
M
,
CN NO oo ON NO m d- oo in m ■'d-
a _ NO ON ON NO o © NO NO r^i r^i
cd «0 d- m r~~ CK 1—* d" 00 r) r-
> “5 ON r- >n co *—1 o OO NO v-\ m CX
>.
a W 0\ oo 00 oo oo OO od td r- r-~
o
■4->
C
W -o o «0 NO d- ON cn r- oo ON
NO d- NO r- NO in <N 00 r-
p '3 m cn ON NO cn © r~ m o NO
CN cn m m NO
co .2* w o p cn d- r-
h4

■^r NO oo On >— CJ ro rn m <N »—i ON


O
a
oc _4 o On OO od r4 NO vd ■O rc CN o
cd O H CN m V-) NO r- oo ON o
VN vn V4 wn VN NO
« > o rJ (N rj r) (N <N CN CJ CnI CN CN
oo
STEAM TABLES
APPENDIX 3

rc NO l"; ON m N/4 NO r- 00 r~; r-


_*
C3
00
v*-i
i-4 ON r4 vd d CN d OO NO d CN d
o OO r- NO V', 3: o ON oo r-
> -c tj- ■O
Tf N" N" ■'d- -d- C^i m m
W n CN CN CN CN CN CN CN CN CN CN CN
Td
-c
■<—>
E _4> c
03 W ▼-H cn
s OO ON NO ON ON 00 r- «n m
<D cd O On O ON ON NO in N- m CN
*■» 3 S-, °®
CO H d CN CN cd N’ id NO oo On d
4) CN 'd- NO 00 O CN 'd- NO oo O m
5 .Sf* T—1 1 CN CN
i-4

co cd
i—
0) <u
a. cn co CN — On OO NO p — in —
CO OO
CO a) E „j O c>c id CN On NO CN On NO cd d NO cd d
4) rt £< r- OO 00 ON CN m,
s o O CO CO Tt
Je CL H 00 co m co co CO Tt N" Tf O’ ■d- ■d-
< o CN CN CN CN CN CN CN CN CN CN
4
CN
N
CN
o) a!

>> p ro CN O O) oo r- p ■dj CN ON
H E 00
J-l
ex
cd
00 id i-4 cd On d d NO CN od 4 ON
CO 4) r- NO -d- co O On r- NO ■d- CO 1-H
c c
> sT
m CO CO CO co CO CN CN CN CN CN 04
>n
W CN CN CN CN CN CN CN CN CN CN CN CN
W
■a
o Id
C
E I—
4) o r- o ON m oo oo r- NO •d; CN _i
*-> o on o ON p oo r~; NO 4 co CN
<D c p 3 p d CN cd 4 >d NO rd od On d
cd CT" 3 cn NO oo o CN ■d- NO oo co
O
I—*
3 1 1 i—1 CN CN

00
M
oo
\Z Cl ^ 00 co ON 'sO ON CN CN NO m NO
**
CO cd co ON P m OO CN m CN p »n
4) no r- NO rd td cn CN id On »d fN ON
© d- O o- m co CN
E CN
3
o
> CNcO'd’NOOO©CN»n
4) JO O O O O ^ —<
o o o o o o o
’3 H-.
’5 cd
4> co -2“ ^
O- h4
co oooooooooooo

E C/D _* co NO
cd c/5 O CN r- in ON ON NO oo d" CO ON OO
4>
■>-> 4) Oh CX C" CN O CO NO ■d- CN 00 ON d- m
*-> M NO OO CN CO *“H CN NO CO m CO
CO
Oh d d -d -d CN rn 4 >d (d ON CN >d
T3
4>
-t—>
cd
i—
P
■*->
cd CX ©
CO EUh © m © »n o m O © m o m
1> o
cn cn co co ■d" d- iri ir,
H

526
© o re d 0 m — ON ON oo d re NO<n ON d ON re ON <n Cl OO re NO d m ON OO re m in _ re NO _ d
On —• IT) Cl Cl d ON IT) d m oo re ONd NO d ON re d re o d NO 1C, m NO d ON Cl m oo Cl NO o d ON re
© re oo *— d d m ON re oo Cld Cl d Cl oo re ON m o NO Cl oo d o NO re On »n ci oo m d d
On oo t ; NO NO m Tf d re ci Cl p p © © ON ON oo oo d d d NO NO m *n in d d re re re Cl Cl Cl
d d d d d d d d d d d d d d d NO NO NO NO NO NO NO NO NO NO NO NO NO NO NO NO NO NO NO NO NO NO

d m d On On Cl © ON o oo Cl O o Cl >n d oo NO o o »n co d — NO o re d Cl r- d re re d co Cl
OO d o NO NO o NO m oo Cl O o Cl NO Cl O Cl NO —1 d to 5 d SO OO Cl NO d co o oo NO d m Cl
d re o NO re *— oo NO d re ci — o ON On ON £ ON ON o o i—i Cl re 1C) d OO © — m m NO OO o 04
© ON oo NO in d Cl © ON oo d NO d re Cl © ON ON oo NO m, 3 re Cl © Os oo d NO NO to
d NO NO NO NO NO NO NO NO in mi in m in m in to m d d d d d ■d d d ■d d- p co re re re re co co

04 in ON re co 1C, O ON o m tj- NO m e|- o t— c- oo m c~ in ON On NO On ON m On o oo d oo ON ON oo m


T-H re d »o m d o NO m oo m c- H ct- d ON o Cl Cl Cl w—« O ON C~- mi m O oo d * d m ON mi
co On «n d m ON m o NO *— c- C1 OO m oo (O ON d- ON d On d ON m oo m oo m c~ Cl r- NO O mi o
oc OO ON p © p Cl m m m to NO NO r- oo OO ON ON o o <—< Cl Cl m cr d d mi in NO NO d
*—< p >—< -- —< —^ —- —^ T~* —■ »—I —I —: — ci ci ci ci ci ci ci ci ci ci ci ci ci ci ci

NO m oo re d ON 1 — *■— oo >n o re >n m re ON re m d —■ *n d NO Cl d ■d o Cl O in mi — re p d re


ON oo NO m’ re .—^ o oo NO re ON NO re o d re O NO cioo re oo re oo ci NO o re NO oo O ci re ■d d- re
o T—« d re d m NO NO d oo ON ON o d d re d d m m NO NO d d oo OO ON On ON On O o O o o O
NO NO NO NO NO NO NO NO NO NO NO NO d d d d d d d d d d d d d d d d d d d OO oo 00 oo 00 00
Cl Cl d d d d Cl d Cl d d d d Cl d d d d d d d d d d d d d d d d Cl Cl d d d d Cl

in d 0© d oo o Cl d p d d >n NO m d NO d NO NO p d m, d p i— OO o d p d ON m, t^i oo m in
od NO re — oo NO re o d re o NO ci od d- ON d ON 00 ci NO On ci mi d od o o — o oi OO \6 rn o NO
in d re d o ON oo d m, d re i— o oo d m, d d ON oo NO d re i— ON d NO d d o d m, m • On NO
re re re re re d d d d d d d d T— —i T— ■f— H o o o o o O ON ON ON ON ON ON oo oo oc oo d d
d d d d d d d d d d d d d d d d d d d d d Cl d d — — —< 1—1 — — — —

m NO oo m ,—- o d NO d «n o oo — ON — ON re re o d m, d >— d d d d oo m d NO d d OO d Cl d
O ON Os o\ ON ON p p re d d P re NO p NO Cl oo «n re d p Cl re NO ON d p d NO \D d NO re
ci ci ro mi NO d ON O P ci re d- NO d ON o ci re mi d ON P re mi d ON ci mi d o re NO o re d
to d ON 1 ro m d ON, i— d NO oo O d d NO OO re m d ON i— d NO 00 o d m, d On d d NO ON re
04 d d m m re re re d d d d m, m m, m, m NO NO NO NO NO d d d d oo oo oo oo oo ON ON ON On o o

\D p NO p Cl d in NO mi d p d re NO ON © © p m p d mi »n Cl d © © oo re m m p d re p — ©
NO re On mi ci od d © NO ci od re ON d ON mi © d ON d OO ci NO © re d © ci mi d ON P ci re re d d
m, NO NO d oo 00 o © © T— —H d d re re d >n m <n NO NO d d oo oo oo o ON ON ON ON © © © © © ©
d d d d d d d «n m, m, m m >n m, m «n m m <n m m, in m, m m m, m, m m m, m, NO NO NO NO NO NO
d d d d d d d d d Cl Cl d d d d d d d d d d d d d Cl d d d d d d d Cl Cl d Cl Cl

>n T—4 NO © d NO d d NO re © d oo ON p d m d P OO <n © — © NO ON oo d d m, ON p m NO Cl Cl oo


mi NO ci d ci d ci d ci d P mi ON re d d d © re NO od © P ci re d d d re ci P ©i d d ©
© ON d NO d re © OO d m d d © ON d so d d ON d mi d d © oo NO d d © 00 NO re i— On d
Cl — — —H —H t— © © © © © © ON ON os ON ON ON oo OO oo oo 00 oo d d d d d NO NO NO NO m. m,
d d d d d d d d Cl d d d d Cl '—i —« '—' — 1— — 1—1 1— —1 1— — — 1 — —< —i 1— i— »— —

--- d m, o NO d m oo d d d © © d d m. d oo OO d d NO m d ON © ON d >n d re d d re d ON 1—1


© p Os oo oo oo oo p © p re in d © re d i-1 p Cl OO m m •—1 © p i—1 re NO © in p OO d d oo <N
ci ci rn ■d mi NO d od © P ci re d NO d od © p re d NO oc © ci d NO od © re mi od c> ci NO On rn
ty^i d ON re »n d ON d NO oo © d d NO oo re mi d ON *— d NO oo © Cl mi d ON — d NO OO © m
r4 d d m re re re re d d d d m, >n m, m m NO NO NO NO NO r- d d d OO oo oo oo OO ON On ON ON © O

m, ON d «n NO ON ON ON oo d NO
© © in © re Cl d d d m, re d
ON d Cl NO ON NO «n Cl ON re oo oo d oo oo
d Cl d OC Cl d On © NO © oo Cl OO NO Cl NO d Cl Cl NO d d NO i " d m d d NO oo m ON
d ON d m © (N © oo d —* i— re ON d NO oo © d ON d © d d i— ON d »n d Cl — © ON OO d d NO m
p © d OC re ON p d Cl © OO d NO m, m d re re re Cl Cl Cl ’—1 -—1 <— '—1 •—1 1 © © © © © ©
d NO mi re r4 ci P P P p P ci © © © © © © © © © <© © <© ci © © © © © © © © © © ©

re NO ON re NO © oo Cl NO © m, © in © in r— NO Cl oo 1—1 d d ■—i ON d re —i ON ON On ON
d ©
T— Cl Cl Cl Cl re re d 5 d m. m, NO NO d d oo oo ON ON © © 1—1 Cl Cl re d d <n 3 d OO ON © Cl
© © © © © © © © © © © © © © © © © © © © l—1 l-1 ■—1 ’—1 1—1 ■—1 l-1 ’— '—1 Cl Cl d
__ ___ __ ___ ,__ ,_ y—| .__ __ y— __ ,— .—, ___ .— ,—i 1— i—i -- T— 1— ,— » '1 .— .— .— i— *— ,— r— «— r— i— —
c © © © © © © c © O ©■ © © © © © © © © © © © © © C © c © © c © © © © © c
© © © © © © © p © © © © © © © © © © © © p p © © © © © © © © © © © © © © ©
© © ci © © © © ci © © © © © © © ci O © © © © ci © ci © © © ci © © © © © © © © ci

to Cl d NO re
oo Cl © m — — o m d oo — oo in — oo oo Cl
d © d d ©
© 1—1 © re ON oo Cl © m i—i m m re d © T— Cl Cl Cl d d re re © d OO oo <n ©
d re Os oo OS d mi o Cl d © On re d i—i SO d d i— © On On © Cl >n ON <n Cl © © 1-d d ON ©
On © m, oo — in —1 — ■— Cl Cl d d m © d d oo © p Cl re m d On p >n d © ro
ON mi oo © d o © © © © © © o o © © © © © © © p p P P P P P ci 04 ci ci re m
’—i Cl re to d oo

© to) o tO) m o tOi »n © tOi © in © to © mi m, © m © m © «n © to


©
©
m
©
©
d r- oo
m>
oo § os
©
© O
1—'
1—1
1—
i-^
i—1
©
Cl
1—
Cl ro
i—•
m
i—
©
d
—1
d mi
—■
to
I—1
©
—1
© d
—•
r-
i—1
oo
—*
oo
—1 1
On
■—
©
Cl
©
Cl Cl Cl
Cl
Cl
Cl
Cl
re m
Cl 04
©
d
Cl

527
Adapted from Joseph H. Keenan, Frederick G. Keyes, Philip G. Hill, and Joan G. Moore, Steam Tables (New York: John Wiley & Sons Inc., 1969).
U rr © er, Os (3 oo © t" ir, rr © 3 ri r- t" rj © oo
G oc rr r- sO O rr r- © r4 rr 3 3 rr © © IT, (3 r~ Os
•*—> oc © r- m O sO rr Os ir, oo 3 © © (3 oo rr rr * IT, Os (3
cd Q- CO © IT IT, rr <N © r- 3
cd o o o p p oo oo OO p p p ©
GO ir, ir, IT,
,_v > sd sd sd sd >d iri id <d ir, >d ir, IT IT, IT id 3 3

oo
P
— ri n <N oo O rr © r- IT) rr r~ y—y <3 © 3, Os IT ©
— d. o Os 00 sO ir, rr O (3 r- rr rr (3 oo © © t" ©
cd Oc © oo Os m ir. r- Os © (3 rr ir, © r-~ oo oo Os r- m rr 00
> CO rr p — p p p p p p >r, 3 rr P p © OC © ’—; sO
kT ri ri ri ri ri ri
W rd rd m <d rd ri ri ri ri ri ’ 1 1

Reprinted from Fundamentals of Classical Thermodynamics, Van Wylen and Sonntag, John Wiley & Sons, Inc.. 1976. By permission.
a.
o

C
W <3 r- m oo 3^ oo oo © 3 <3 3 © rr (3 © r- 3 r- r- © OO
. T3 r- D oo rr O ir, O so rr © © rr ■ © oo oo © © r- 3 © Os
-♦—» '3 ■r Os ro oo <3 r- rj SO ir, © ir, © 3 Os 3 IT, IT, r- (3
cd O' CO p p oo oo P p p »—1 — (3 ri rr rr rr 3 IT © r- p p 3
GO P
ri ri ri ri <3 ri rd rd rd rd rd rd rd rd rd rd rd rd rd rd 3 3
P

L. © it >r, Os p p O p rr p
, © p rr «T _ p © p ©
___ rr
•4—> o 00 rd P Os sd rd Os ■d Os rd sd oo Os 00 d 3 <d it] ri rd P ri OS
cd Cl O O Os os O OO oo r- r- © >ri 3 rr <3 © © <3 © oo rr Os
C/3 cd oo OO r- r- r- r~ r~~ r- r- r- r- r- r-~ © © IT, 3 rr ©
> <3 n n ri (3 (3 (3 (3 (3 rJ ra rJ (3 r4 n (3 CJ rj r4 (3 rj r)
'oc
p
---
'i p p oo ir, 3; ri p p © — oo p 3 © IT © © p 3 >T, so ©
d. P sd Os (N d >d d rd .—: d P 3 sd sd rd OO © d rd cd P
cd 3 oo so rr o r- 3 ^—i r- 3 © © rJ oo rr 3 (3 Os rj 3
i-~ r- sD SO so sO irs ir, ir. 3 3 3 rr rr', rj <3 — © OO r-~ 3
U-L U — — ”■
1
cd
JO
■*->
c
JO W rr CD m r~ OO — r- Os — r-
cd T3 p rr r- rr r j p o p rr © rr © 3 rr © IT rr P © IT IT, rr
E— w '5 P >d cS d On d cd ir ri Os sd ri — P P IT, 3 cd cd cd Os
cd cr © oo o rr IT) 00 rr © OO y-^ r~ © rr © <3 Os r-~ © Os OS
OJ C/3 O o — r—‘ <3 <3 <3 <3 rr rr rr 3 3 3 IT IT © r- OO ©
P (3
3
cd
l*
0/
Cl. S_ 3 3 OS p p p (3 3 © p © (3 3 © IT p © 3 IT, IT, ©
E -*—>o
Oc rd ri <d On sd rd cd sd ,=: sd as rd id sd sd id OO 3 OO P 0© OS
<u cd d. 3
cd
o o o O Os O Os oo oo r- © © ir, 3 rr (3 os © IT (3 (3
H OX) GO © SO sO ur ir, ir, ir. iri ir >ri ir, ir, IT ITi IT 3 3 3 rr (3 ©
> <3 <N rj <3 n ri r4 <3 n rJ <3 rJ ra C3 <3 <3 <3 r4 (3 (3 C3 <3

X*
s—^
>>
OX)
. r- © t"- so p rr ON P 3 P © p 3 — Os p rr © rr IT, ©
Cl Oc sd ri sd cd rr sd d OO OO d 3 P >d Os P cd rd 3 sd sd 3
cd 3 Tl OS r- OO ir. <3 Os © rr Os IT <3 oo Os Os r- <3 OO
D >
c IT T 3" 3 rr rr rr <3 (3 ri —< T—1 © Os OO r~ © rr
tti ’-- ,—l ' *“ T—'
w
Id
>
Uh — Os oo On 3" SO «r> so © (3
•4—* TD p m ri rr p rr p 3 © p (3 © rr p T, © rr rr P <3 © ©
c -4—' '3
cd cr s
sd o d oo (3 d ri d rd OO >d ri On d IT 3 id cd P IT 3- Os
it oo o (3 ir. r- o (3 ir, r- © rr lO OO 3 © r- 3 (3 3 <3
C/3 © o » 1 <3 (3 r4 <3 rr rr rr rr 3 3 «T T © r- OO ©
i—i (3

oo © r- oo
r- rr IT IT IT
©
'oc m oo i—i t" 3 On r- r- 3 r- 3 IT © Os oo Os
3 <3 IT
p J— f- OS <3 r~- SO r- r- ir © OS rr oo 3 Os
r~ oo Os Os *—
m • o 3 o ITi n oo ir, <3 o r- ir, rr — OS oo © Ti (3
© oo © 3 rr
'cd cx oc «T sr, 3- 3" rr rr rr rr <3 rJ (3 <3 y-* y—y y-* y-~ © © © ©
^E cd © O o p O O O O © © © © © © © © © © © © © ©
C/3
<u > O c> cd cd cd cd cd cd cd © cd cd © cd cd cd cd cd cd cd cd ©
E
jO
O
> o m so On (3 r- (3 oo © 3 3 I/'S r- (3 os _ OO © rr rr IT
o p-i 3 so r- OO O rr 3 © OO © rL 3 r- OS © rr 3 Os IT,
eg . w
3
<3 n <3 <3 (3 rr rr rr rr rr rr 3 3 3 3 3 IT © r- 00 <3 —
"a cd <3 rr
<D C/3 .2"
$7 © o o O o O o © © © © © © © © © © © © © © ©
cl p
f 4
o o o o o O O © © © © © © © © © © © © © © ©
CO cd cd cd cd cd cd cd cd cd cd cd cd cd © cd cd cd cd cd cd cd cd

E
Press.

cd cd 00 m os OO Os <3 <3 Os © rr <3 © r- 3 T © rr


a> Oh ►P-l
3" r- oo oo Os 3 © rr Os OC © IT 3 r- 3 oo IT rr On
■>—' SO On C3 so p 3 p 3 p 3 p IT) ri oo IT <3 OO IT T sO © ©
CO j*:
rd rd d d >d »d id sd sd d d OO Os os cd ri 3 sd oc ri
T3 <3 <3
<u
<—>
“ ~
cd
i_
3
> 3
cd CX
CO E u h< >r o er, o IT) o »T) © ITi © IT © «T © © © © © © 3
a> o 3 er, SO so r- r- oo OO os © © y—* r— <3 rr 3 IT © r- r-
H <3 n D <3 (3 rj <3 <3 (3 <3 <3 rr rr rr rr rr rr rr rr rr rr rr

528
s- ci © Os c Cl >n of —

7.2844
Cl m

7.3594
in OS of

7.2233
7.1717
© © ©

7.1271
o © «n c CC m c of LC, © oo 00 oo CC ©
cd c. r—.00 »n c- Cl Cl of c r- OS LC, LC, cc © fC © os >n
oo cd — Os 00 c © «n of cc Cl — © © oo r~- VC >n of
> Os od od oo od od oo od OO OO od c c c c c

>>
a Cl Cl os

5.9104

5.5970
r~~ c- © os © © cc c Os of

6.0568

5.6868
5.7897
©
c. © Os Cl Cl m Cl oo <C cc © oo of Cl m
o cd >n © cc © cc Cl in ^—1 c ®1 >n c cc Cl © of
S-H
■>-> > oo © of re Cl © On © in Cl © © 00 no <n Cl
c UJ Os oo od od od OC od c c c c © © no © ©
W

Os r- c © >n © cc Os ©

1.3740
OS © ©

1.3026

1.4336
1.4849
O ©

1.5301
o >n <n © Cl Cl © C o\ ■cf Cl cc in cc
o © On © >n Cl c c- It >n cc os Cl Q\ 1—1
$ .sr o 1—. Cl m m Tf ’rT «n © c oo oo © © © Cl

Uh —

2685.4
2693.6
2700.6
Cl cc in ©

2706.7
of in >n oo c Cl rc oo Os

2675.5
cc c
w ^d
Cl >n cd © in t}- ^r OS OS od mi © >n
cd © Cl cc Tf rr >n © c oo © © T— Cl rc "d" ©
C/3 >n in >n »n >n «n «n «n >n <n «n © © © © ©
> Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl

>> 1—H

2241.0
2258.0

2213.6
2201.9
2226.5
o. rn On © © © >n os c © OO CO co — Cl or ©
D. _
cd cd •£$ Tt © © 'd- ci cd so ci cd od NO NO os in od
> -C © OO c © »n 3: co Cl © Os c m Ot CO © c
-C >n ■ct 'd- 'd- or ot ot CO co CO CO co CO co Cl
c W Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl
W
Saturated Steam: Pressure Table

417.46
444.32

486.99
504.70
oo OS >n © Cl Os CO ot © CO co oo 0\ ©

467.11
c or ot © Ot oo c oo OS Ot OS Cl in of CO
Os od- cd OO — c od L in T-^ Os c © Ot
3.1-*' in c- oo ©
T—1
Cl
L—<
co © Os Cl »n
T—l >—1 Cl Cl
c oo
Cl Cl co
Of
CO
oo
CO
hJ

© rn >n or «n Cl in <n OS C c L—H of © OS c

2524.9
2519.7
2513.5

2529.5
CO
oc in >n On Of od «n © © c OO © cd od c cd © 2506.1
13 g- s c oo ON © © © L—L Cl co CO 3: <n © © c oo o\
co co co of of Of -d- of of Of of of Ot of of
00 > Cl Cl <N Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl
>>
oc
t-H
<u
2025.0
c
2069.3
2052.7
2088.7

m r~- © os in c c ’-o oo Of Cl Cl «n
2038.1
© c -d; -d;
W a. © >n c cd ci -d © ci »n yL Os os cd ci
cd ^
> s C m c<~, ci
Cl L-H © Os 00 © of Cl © Os c »n -d-
co co CO Cl Cl Cl Cl Cl Cl L—
"aj
c W Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl
1-
a>
■*->
c
of
466.94
486.80
504.49
417.36
444.19

oo oo >n oo Cl Cl oo © © CO -d-
o of of
o 3 F © of oo c oo Os co On Cl <n -d- co
cd OO t-H c od *-d >n T—H os C © of
3 £2 c- oo ©
L—1
Cl co © Os
L—1
Cl
Cl
in c oo
Cl Cl Cl co
i or
CO
OO
co
hJ
1.6940
1.3749

1.0036
0.8857
1.1593

<u OS of os © c
i— of oo © «n c © Os of c Cl of © Cl ON Of
E ° Cl OS © Cl © oo Cl © © SO Cl Cl ON Cl
3 ed g- £
Oc
© OS c c of >n of od On Of © c sd m r^ cd
O © Cl oo © m of co Cl -“01 ’ ’ 1
^ > Cl
>
o
tc
'o41
0.001 048

© © — ^- ci rc -d- i/~, go © of ^ o ci c © c
0.001 043

0.001 053
0.001 057
0.001 061

©©©©©©©©©o-i^«T-HcicicicCfC
Cc ooooooooooooooooo
CO ^ a- $> o o o o © © © o
w5
o o o o o o o ©
©©©ooooo©©©©©©©©©

r—l C m
116.06
105.99
111.37

120.23

oo © oo oo © oo os c © c © OO
99.63

Cl, © On O >n © © OS oo Cl OO Os © On L—1 OO m c


£Uh
QJ
o
© sd r of od ci in cd © of On in t-H «d
>—1 ci Cl Cl cn $ of m © © © c OO Os
H

cn
0.100

0.150

0.200
0.125

0.175

C/2
C/2 cd cd
a> 0* © © »n © in © © <© m Cl
U. © —; - ri ri m t h m © »/-) O © © *n
Oh - n M r, ia) r- s

529
1— 00 d ON On d 10 in ON in cn m 0 0 r- 00 — xO — «n, xO CO cn cn oc xO ON
• o d d 0 —4 in 0 d >n xO —4 ON 0 00 ■*t Cl Cl Cl ^t xO CO CO >n ON P ON O
00 Cl 0 OO m Cl ON NO OO Ot

Sal
cx 00 inCl ON xo Ot —4 ON m d OO xO ro 0 oc xo p
03 p p O ON ON ON On 00 00 00 r- p p p xO xO p xO m m. m p P 3 cn cn
> d d d xd xd xd xd xd xd xd xd xd ^o xd xd xd xd xd NO xd xd xd xd NO xd xd xd

cn >n ^H xOO r- ON xO O oc co oc r~~ xO — O oc On r- 00 O 00 >n


d d in O O p m On «n O <N 00 O m -t XO —* 00 r~~ 3xO xo co >n ON 3 cn
00 On
03 OO C4 xO xO 1—1 m xO ON Cl —» NO —i r-~ Cl e'¬
OC O "P 00 d
> p cn rn CJ<N p p O OC p p p p p »n »n 3 en p CO Cl p p p OC
W m >n in in >n in «n «n d d d d d d d d d d d d d d d d d cn

•a xO d 00 00 NO «n 00 xo r- m <N Cl 0 Cl 0 NO Cl r- Cl xo m d O Ot
• •— O d O r~ <N xO 0 0 r- y—< Cl Cl 0 xo 1— ■^t c- oc ON xO —» 'P >n >n d
3 1—< CO r- —- in oc —4 oc
Sal

CO d 0 r~ O n >n r- Cl xO ON m xO ON Cl "Ct c~ ON
O'
• yN in p 1 p p p p 00 OO OO O) p Os p O p 0 p Cl Cl Cl cn cn 3
J d ^—1 d d d d d d 1—* d d ci ci ci ci ci ci ci ci ci ci ci ci ci
1

, p oc p Cl >n
kH On cn enOotxOxOONdO oc co >n p — p On — 0 or
O P d d
00 ci xd —; m ON Cl m 00 C"> 00 m xd d cd xd ON d cd xd od d ci xd On
a u Cl <N Cl m m m m, in xo xo xo xo r- r- c- c- 00 00 oc On ON ON ON
00 03 d d d r- r- c c~ c- r- c~ r- c- r- r- c- t- c~ r- c- r- d d d d d
> d d d Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl d d d d d

cn m p oc p oc 00 c~ p p co p co p p ^t p p p ci p d cn ON d
d
03
oc ci rn in od d rn d 00 K xd xd xd d oc ON y-H cd >n 0 xd ci On d d d
O 00 d xo *n Tf Tt- m Cl O ON oc r~~ NO m p CO Cl 1— tp 00 d m or —* ON
> y—H r-H
H H H-H y-M O 0 0 0 O 0 O O O O 0 ON ON ON On On oc
W d d d Cl 04 Cl 04 04 Cl Cl Cl Cl Cl Cl Cl Cl Cl C4 Cl d Cl

-7—H d d ON C' m m y—l in m m xo OC Cl r-' y-H Cl co Cl ot in CO O ON O ON


d cn 00 p Cl co P p Cl Cl p m Cl Cl ■O' Cl oc p p CO p Os cn OO p d
-*-H ’3 1-H ci
cd O’ -ST 0 >n 06 d rn d d d cd d in d d id ON ci cd ci p od d d od od
Cl cn ot xo c- 00 ON O Cl >n r~- 00 ON O (N CO Tt m xO 00 On —- cn 3 d
Table A-3b (continued)

00 m >n m »n
IT) «n »n xo xo xo xo xo xo XO c- r- c- r- c- c- c- r- oc 00 oc oc 8
J

k_ p m p p ON
Cl m p p Cl p p N—1 >n p oc p >n p p p oc p oc p oc cn
•*-n o d m d od d ci
Cx 00 cn d d cd xd OO d td d ci d xd 00 d ci cd xd d d d
cd 3 cn cn -t Tf it in >n >n xo xo xO r- c- r- c- r— 00 00 oc oc 00 ON ON ON ON 0
c/5 03 »n in m >n >n m >n m >n «n <n >n »n in >n «n m in «n in >n >n m >n m, m xO
> d d d Cl Cl Cl Cl Cl Cl Cl Cl Cl Cl c-i Cl Cl Cl Cl Cl Cl Cl d d d d d d

, On p p O ON m On p Cl p >n p T—' xO p p Cl p CO p, p , cn p 00
Cx —H P —H cn
03
00 rn ci d Cl cd in xd On d d ON rd xd xd xd xd d od d ci xd d d
1— O On 00 xO m Tt m Cl O ON 00 r-~ xO »n '■t CO CO Cl 0 ON d xO m d ON
O O ON ON ON ON ON ON On ON Ch OO 00 00 oc 00 oc oc 00 00 00 r- d d d d xO
w d Cl H 1—H H y-^ y—H 1—H y—H

•7—1 d O On >n P >n O r—> c~ OO Cl O xO p "t Cl r- CO in O', ON


00 O XO XO
4 ot p p On Os p cn p xo p On p p p Cl Cl oc 0 0 d 3 d p
p Ot 3
Sat

’3 d in OO d ci cd d d ci On in On cd xd oc d d d d — d d cn OC cn xd xd
cr 3 d cn ct xo 00 ON Cl m
c- O >n xo 00 0 0 Cl CO 'It >n xo 00 ON —4 d of d
m in m in «n m m xO xo XO xo xO xo XO r-~ c- c~ c~ r- c- e'¬ r- oc 00 00 00 8
d

en cn <n it d On cn
Xh cn d cn 00 O cn ^t m O On r- c~ c- 0 xo it 0 Cl0
m cn d 00 d ot NO
0 cn 00 d »n Cl it 1—1 Cl ^t '7 Cl in Cl Cl m 0 c~ >n or ot r- cn —H 0 — cn ON
Oc —1
cd Cx ON m O xo Cl ON xO C' 'it 0 t" »n it Cl 1—1 0 0 r- xo in P cn —< ON
GO 03 d d xO xo >n «n rr ^t 'it m cn co Cl Cl Cl Cl Cl Cl Cl N-1 4-1 4— O
> d d d dddddddd d d d d d d d d d d d d d d d d

ord 0 NO ON 1—H ’’t 00 CO 'O oc Cl m Ot d cn 00 ON ON or NO d


NOxO d r- r- C" OC 00 00 ON ON O 0 0 Cl Cl Cl cn
—1 1—1 7— cn or in xo d
. "O O O 0 0 0 0 0 O O O O y-H >—1 1 >—1 >—1
»—1 4-1 4—4 4—
4—>
1
3 N-.—,
(73 0 y—1 yH h-H y—1 n-H —H —H i—i y-H —H —. _ —H —4 „ _
GO 3 O 0 O O O O O O O g g g g g g 3
0 O 0 O O O O O O O O 8 8 8 8 8 8 8 8 8
d
0 O d O O d O d d d d d d d d d d d d d d d d d d d d

8 0 r4 O 00 m co xo 00 m r- 00 CO xo 00 0 ON On d d xo d
d- 5 NOp rn p p p p 00 0 Os c- It 0 co XO 0 p On 3 p cn p p
E U E-i or d d NO 00 y-H cd cd d in 00 Cl r- 0 Cl m r- d d d y—. in od «n ci
a> 0 d d cn m m m Tf it 'it >n m in xO 3 xO r- r- r- r~ r- 00 OO ON ON ON O —4
H y—H H H « i—i 1— *— 1— d d

m 0 >n >n 0 >n


C/3 cd d n d g Cl m C" in 0 >n m 0 «n 0 m m O O O O «n
C/D
a/
t-H
CU Cx d d d co co co co 5
it m m 3 xo r- oc 00 £ O 8 p d cn 8 p d 8
3 d d d d d d d d d d d d d d d d d d d y-H d d y-H y-H —i 4—i ci
Oh

530
04 «n ON CO HH P 04 rn 04 04 ^h O' P rn r- 00 on O' p oo ON m o~ 00
O' r- NO «n © rn ON m m O' P 04 04 04 i—< ON m O' P 04 NO i— 04 ON
Os >n 00 04 o~ o~ oo i—* P o~ I*"! on ON m r- o P o 04 04 © rn 04
04 04 p p p On oo oo p p NO on P P m m 04 p O ON 00 on P
n© no NO NO NO »n »n on on on on on on on on on on on on on P P P P

oo 04 O O' 04 O', 04 p on p m 04 OO on ON P oo ON ON o oo NO
m 04 pi m m 04 04 NO 1-H P m NO --1 00 P ON On 04 rn cn m
ON o P O' on NO ON m ON on 04 ON O' p 04 ON NO m oo 1—< ON 04
p p on P 04 p OO SO on rn 04 *—1On oo p p P CO 04 p ON NO 04
rn rn rn rn rn rn oi oi oi oi oi oi 1—i r-i ^H —i 1—i H-

on O' O' m P 04 O' 1-H oo oo NO on 04 NO 04 00 ^H ON on oo On on o oo


rn p on on NO O NO 1—* NO on ON ON NO o m p NO O' -—h oo m O' On
o on p 04 On 04 04 <N o oo on 04 ON, NO 04 oo P o O' m i-h
o 1—H 04
on »n p p p ON O 04 04 m P p p p p P oo oo p p rn P
04 04 04 04 04 04 on rn on on rn rn rn rn rn rn rn rn rn rn P p P

p p 04 P P m rn p O p P p o\ 04 p >04 NO 04 «—< >04 O' p p m


i—I rn P rn — P P 04 OO 04 P
04
NO p 04 O^ O
T—
o 0^ ON P ON P >04 ON
O O O O O ON oo O' UON P o 00 NO m oo p o NO o m NO ON
OO oo 00 00 OO o- O' O' O' O' O' O' NO NO NO NO >on >04 >04 P p m o
04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04

04 p P p — 1—1 p — m p — p 0^ >04 o NO p — p p 04 P o
>04 >/n rn p o i—i >04 1-H OO 0^ >04 rn o NO o o NO o^ oo rn NO rn
NO p ON >04 —• p O' O O' 1-H <04 On m NO m >04 O' oo oo P P
OO oo O' O' O' NO >04 >04 m m 04 i—i o o ON OO O' NO >04 P
1 V—H ( ^-4 y—* T—^ i— < —H

On i—< 04 >04— m O P NO NO
P p P O' m <N rn O NO 04 p — m tn >04 i—" m O p, m p 04 m
NO 04 OO On rn O' NO rn O' o 1—H -- O o o 04 NO NO oo 04 ON
m NO o P oo 1-H NO —i NO o >04 ON m r- >04 ON m O' 04 oo 04 ON
ON ON O O o fN 04 m m p p NO NO NO O' O' OO oo O o
— —' 1-H i—< — —1 i—^ i-H 1-H —^ 1^ i—i —1 1— 04 04

O p p p m p P >/n OO °® P oo p i—i oo p p O m p NO p NO
04 rn P rn 04 O' ON o ON O' P ON rn NO NO t/n 1—H un P 00 rn o ON
O O o O o ON OO oo NO >/n P 04 i—i ON O' >o m o O' m On m oo 04
NO NO NO NO NO </n >n, >n >0, lO) «m >m >0) p p P P p m m 04 04 o O
04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 O! 04 04 04

04 o m m o rn rn o 04 m P p O' p 04 OO p oo p p p p 04
00 ON oo o cK
NO s’ ON >m 04 oo
rn P
NO o
1-H NO o vn
ON p oo
OO
04
ON 0\
NO o s O'
loi oo
ON
0^
o
oo
oo
wn
04
NO NO >Ol >o, i/n s m m 04 04 o o ON ON oo oo O' NO >/0 un m 1—1

m —H OO m 1-H T— >-n O' t—H P


oo —H P P rn OO s p p O p O NC| p P 04 Os
ON NO p ON p
rn On P >rn
m >Ol o p
04 t' >o>
OO p o NO,
>oi o rn rn rn
O
OO
ON m O' 1-H P
>n
oo
04
04
O OO On in 04 —4
NO ON m OO P NO 04
Os
ON ON o o o 04 04 m m m P p un un >n NO NO NO O' r-' OO ON O
▼“H —H — 1-H — HH hh 1-H —H 1-H —H —H 1-H 04

NO O' m o m O' NO p ON p 04 oo
>n
O'
oo
On
oo
NO
O'
o
oo
O' s s O'
rn
04
«n
oo
p
04
o
oo
ON
NO
04
oo
O'
oo
p
m
m
o
m
NO
m
oo
p
>n
NO
m
oo
>n
ON <n
SO
oo ON ■<o O' ON ON 04 O' m o oo m p 04 1—1 o On oo O' NO >n P m m
oo
o
O'
o
NO
o
«n
o
p
o
m
o
m
o
04
o
04
o
04
o
1—,

o
T-H

o o
1-H —H

o
1-H

o o
—H
o
o
rn
o
o
o
©
©
©
©
© ©
© © o
o o o o o o o o o o o o o o o o o o © © © © © o

r- r- »n
oo ON i—i m
04 so
ON 1-H
>n oo 1-H
OO
00 1-H
04 ON O' r- ^H oo i-H
>n oo 04 NO 1-H >n 1-H
© © p
O' P 04 m
SO O'
©
04 lo
P
fN 04 04 04 m m m P p >n NO © r- O' oo ON © 04 O' 1
1-H 1-H 1-H i—' 1—1 — 1-H 1-H 1-H —H HH ^H 1-H i-H i-H 1-H HH 1-H 1-H 04 04 04 m
o o o o o o o o © © © o
o o © o o o o o o o
p o o ©
© © © o © © o © o © © © © © © © © © © © ©> © © ©

m ON © © © ON P OO © © © <n m io. p P O' © P —H ON © P


P On On p P On NO oo © P © p On p 04 P m © p, OO oo oo p
oo rn rn 04 © rn >n in in rn T-i OO P © NO 04 0^ 04 o^ >n On rn P
1-H 04 m P m © O' oo On © i-H i—i 04 m m P p m m © © © O' O'
04 04 04 04 04 04 04 04 04 m m m m m m m m m m m m m m m

>n ON
04 m o m ©
04 04 rn rn p >n © O' 04
9

14

16

19
20
10

12

17

22
8

13

18
15
11

21

04

531
T}- Tf T}- m m IC Cl NO
00 00 Cl c On m ON ON NO 0 Cl »o> CO
ON m «C m
m c ON ic, NO NO >C NO oc 1C, On 0 oc NO oc
»c NO T“H r<) c*~) OP ON m NO c c NO •rt 1—« O'. Cl c r- NO ON ON
m NO 00 0 Cl >C 00 O r<) »c c ON m »c 00 ON CO to, oc
c c c c od OO OO 06 ON 0\ ON cK ON O 0 0 NO NO c c c c od

re ON
<c r i er m re re Cl c Cl NO — m O NO IC NO 00 IC Cl 00
so
«c NO NO to Tj- OO od 00 od NO 0 1—4 c ON cr, od ci 0 NO CO Tt
ON ON et- O re IC, NO NO NO oc
c c c c c C r- 00 0 Cl NT, On -rf c
ON
NO NO c 00 ON O d 'i- c ON 1—H m NO OO ■rt c- r- OO ON 0 Cl Tt-
ci d d d Cl m re m m Tf ''t ■'t tO >C w Cl Cl Cl Cl CO CO c,
CC
03
CL
(X
£
O — c OO c Tf ON NO ON Cl >C OO OO m c <C NO IC, 00 — 00 Cl

NO NO ci 00 rd O c *—1 *—< ON cn er ci c c re
O
CO NO
NO ON
o’ O O 00 «c m NO O c NO >c m >C NO 00 IC) NO 'Tf Cl 0 NO n
«c >0 <C NO c OO ON 1—H m NO 00 O Cl NO 11 1C, IC, NO c~ 00 ON
II d d d d <vi Cl Cl re m m m -'fr ■'t Cl Cl Cl Cl Cl Cl CO
CX cx

0 00 xr ic, 00 Cl 00 NO C^J
Cl" 'C NO d NO On m IC 00 0 Cl <C c ON 0 Cl O •<t IC, •*t Cl On
On ON m c O m 0 NO Cl On IC, c- c*p ON
■r—( NO NO c CO ON ic c 00
NO NO ON P NO IC O ■*T ON 00 re c~ Cl tT IC, *c, NO c oc

1-1 T— ci Cl ci re •*fr er >C >c NO NO c

< ,
ON c O NO re Cl NO 00 ON Cl c ON Cl Cl ON OO IC NO Cl 0
re er O oc IC c c ON ic NO NO ic NO 00 « r- NO Cl re IC,
ON ON «c »c re NO IC ic oc ON ON oc NO re ON c y—* 0 re Cl
05 ON ON
IC NO re «c oc NO 00 O Cl NO oc 0 re ic c O re
c c c od od od od ON ON ON ON O 0 0 O 0 NO c c c c od od
T—1 ■ 1 ■ 1 ' ' ■ 1

CO On IC c 0 IC ON C ic ON re IC ' * c NO re 0 NO NO re p p
to
Superheated Vapor

CO
ici ci 0 c NO «c od od ic od od NO O c ON ic] IC c ON IC NO
oc c c ON ON •r}-
rn c- OC
'5f 00 c 00 0 Cl IC '’t O Cl NO NO NO
00 NO NO c 00 ON 0 Cl c ON i re NO OO r—< 'Tt
m c c OO On O Cl
v—^ Cl Cl Cl Cl Cl re re re Cl re e|- ic ic Cl Cl Cl Cl re re re
Table A-3c

03
03
Cu
Cu
£
s
IC, ON NO NO ON O re IC O Cl NO ON ON oc NO NO 00 c c c p O
re
0 re »c On <c od ci ci 0\ re ef- ci C c re re 0 0 od NO IC, O
0 oc 00 <c re NO re 0 c- NO »c ic ic NO oc O "Tt c ic Cl O nO re
er »c »c NO c- OC ON ^—1 re ef NO 00 0 Cl NO ic *c NO c oc ON
II Cl Cl Cl Cl Cl Cl Cl re re re re re e- ¥ Tt II Cl Cl Cl Cl Cl Cl re

CX Cx

oc ON re re ic c
$> O oc ON NO 0 ON ef c 'Tt 00 1 Tf c »— ic re NO NO «c y—* nD
er 1 OC ic Ci oc O re <c OO Cl IC c ON Cl 0 re ON c re OO
Cl OO re oc Cl Cl < 0 ON S? 00 c- NO >c ic NO NO C c oc p
re re re 'Tf Tf- ic NO c od od ON 0 — ci re "Tt —1

1 1

Cl ON ON Cl OC Cl re c oc oc 00 , NO re c Cl IC NO NO NO oc m
0 c- oc m 0 H r- r-~ 0 Cl oc ON On 00 ON y—M c ON NO OO Cl
05
ic c 00 0 OO 0 ON NO 0 Cl re re Cl O OC Cl c O O ON Ci
p p 5 p S ’—1 Cl p OC »-* ’O NO 00 O Cl p «c Cl p c p Cl t/p
od od od od od ON ON ON On 0 0 0 O P —• p c c cl c c od od
1—H
Saturation temperature, °C, in parenthesis.

03
r~ p p p IC, re «c NO p c 0 ’Tt NO Cl OC c re c 00 ic p 00 p
OO r»* ci c re ON c NO ON ON «c od ON NO O p c ON Cl NO od 0 _L p NO
tO 00 ON 00 00 C c c c OO 0 Cl «C ON ON O O 0 NO c c c c 00
tT <c IC, NO c OO ON 0 Cl c On re OO Cl c c 00 ON 0 Cl
Cl Cl Cl Cl Cl Cl re re re re re rr ’'t ic «c Cl Cl Cl Cl re re
ct3
Ph Cd
CL,
£
o\ ON ic. o\ re p ,
p re p p OO p O P p p £ ic p p Cl p c oc
0 c re •c c p NO ci od ci ci ON re ic re c c re Cl
3 <oi NO od NO O
0 re T—1 oc NO re r—H NO re O C NO ic >c »c NO 00 0 Cl c IC, re O NO m
"rt IC ICi NO c 00 ON re NO 00 O Cl NO ic ic NO c 00 ON T—
II Cl Cl Cl Cl Cl Cl Cl Cl re re re re re II Cl Cl Cl Cl Cl Cl m
cx cx

r~ NO re 00 Cl re
•<fr ON NO Cl ic NO «c re On >c 1 NO y—< r- Cl r- Cl »c ON O oc NO ON
C NO ON Cl re NO C On 1—1 Cl IC c- 00 0 oc •C 00 ON -rf 00
p OO p «c 00 p 5 O P Cl On IC p c re p p 00 p O re >C r-
Tf c On T—< NO P IC O ON •'I' od re c ci _x _L P
Cl Cl Cl re re 3 •c IC NO NO c

.000
E-
y
C3
>0 O
—,
VI
r—< n r) m t »c no c re
t5 ^< ci
8^ 8 8
d re Tt «c
00 C/)

532
\

o© d Cl o NO o o oo oo Ct oo ON NO CO re o CO CO CO o wc wc CO wc d TJ* o NO r-
wc oo 5 NO NO wc NO 00 Cl wc 00 Cl oo 1—H d CO d re wc wc wc wc d On d NO CO NO d fN
wc ON Cl CO CO Cl o d NO «re m o d NO re d O T-H T-H o oo wc NO ON Tf tn CO o o
Tt NO ON cc wc d oo NO oo O Cl Tt wc oo re NO OO O d re wc Tt Tt d On CO NO
00 OO OO ON ON ON ON ON NO NO c~ d d d d oo oo OO OO ON ON ON ON NO NO NO NO d d d

co r-
"'t wo re ot Cl OO
oo H-< CO o wc r- d NO cr d NO d d T-HON ON o o CO d
Tt wc wc T-H
ci NO d WC ON NO OO o ON ON o NO t-H d o ON "t- we re oo ON wc d l/^i o CO d o ON d Tf
o Cl wc ON CC s ot O NO co wc wc NO NO oo ON d WC ON CO OO Tt o Q\ On o Cl Tf wc d
d On V— CC NO oo * ot d 00 ON O 1—< d Tt NO ON T-H re NO OO d oo ON o ^H d Tf*
co ro ot ot ot ot wc wc CI Cl Cl m re CO re re ■*t ^t ■*t Tt wc wc d d d CO re CO CO
cd
Cu CU
S -H s
d On ON © ot WC © oo oo NO wc Cl Cl d o On d — oo o NO OO
o
oo — CO d d to —
© d ci CO ci NO d ci oo NO O r- 00 ON NO d NO T-H ci t-h WC NO T-H Tt ci re oo wc ON ci T—^
o d NO wc WC wc NO oo d m r—« ON c~ WCd ON d NO wc wc wc NO oo ON o On oo NO WC d
CO or NO oo o Cl NO o WC NO r- r- oo ONT™^ Cl ^t NO oo o d NO WC NO NO d oo ON
re CO CO cc Ot ot Ot Cl Cl Cl Cl Cl d CO CO CO CO CO cr ^t Tt Tt d Cl d d n d CO
II II
Cl, cx

Tt Cl o oo
WC wo Cl On wc O NO 1—1 ^t OO CO oo T—1 1—H ^H
o On d NO
oo o wc d CO oo
WC T-H d Cl OO ot ON wc o o m ^t 5 CO T-h o OO TT
NO t-h ON d o CO CO d o d d
o d re wc NO OO ON HH ^t NO On Cl wc oo CO NO t-h d
re ON Tt Tt ■t NO oo o wc
o o CO or wc NO oo Cl Cl Cl r<) re re wc wc NO NO d d OO T-H I—* d d d
--
o-"

d ON NO Cl o wc ON o o wc NO Tt On T-H
■'t d d NO wc WC NO re oo Tt d cf ON
ON h-H d ON ON oo OO 1-H o NO T—H Cl NO d Cl d O NO OO oo OO oo o CO ON ON d d wc
OO CO wc NO NO wc co T-H
NO NO OO C' cr O o NO X-H CO TT CO T-H ON d 00 d CO T—H d d
WC oo o Cl or NO oo o d ON re 1C, d o d wc d On CO wc NO wc wc OO o Cl CO NO
OO oo ON ON ON ON On © NO NO d d d oo oo oo oo oo ON ON ON ON NO NO NO d d d d

, , »0 ,
d OO Ot d ot o OO OO Cl NO C- m oo ON CO wc •^t OO NO CO re
ON
OO ON o oo NO d CO
CO d d WC ON © d On oo
NO o d we o ci o wc NO Tt oo ON NO oo
r-
Tf WC wc wc CO o NO
o Cl wc ON CO ON Ot O
wc WC wc wc NO NO d oo o d wc ON re OO Tt o
oc
OO T-h CO Tt we NO d
d ON t—h CO NO OO T—1 ot d OO ON O T—X d ct d ON t—H CO NO 00 T-H d oo ON o Cl
re CO ot or ot WC wc V_✓ Cl Cl Cl m m m CO CO CO ^t eT Tf wc to d d d CO CO re CO
a a
Ph (X
s s
oo On Cl
or cc OO Cl O ef On ON o Cl NO o OO
er wc T-^ wc CO oo oo Cl d d ON oo
o
© oo ci ot ci NO d co so d OO O t-H t-H ci d ON d CO
^H ^H
NO NO ci 04
oo d Tt O'- ci er ci
o d NO wc wc wc NO oo NO m Cl o oo NO d ON d NO wc wc wc NO oo oo T-H
o oo d IT) d
CO or NO oo o ci ot NO o wc r- oo oo ON t-h d ’t NO oo O d NO T-H wc NO d d oo ON t-h
CO CO CO CO Ot Ot ot Ot Cl d Cl Cl Cl d CO CO r<) CO CO ■xt Tt 5 ■Tt Cl d Cl d d Cl re
II II
Cl. cx

re o Tt
d On T—<^H d o oo Tt Cl d o d d wc d OO ON o ^H CO CO CO oo wc oo NO
or T— t-h

wo On Ot OO Cl NO O wc Cl m Tt er m d On d er »—H oo WC re O CO On d CO 'Tt Ct ■'t


or On Ot o wc t-H NO Cl wc ON re c- H-H On NO CT d o d wc re T-H NO NO ON T—H CO wc ON
CO or NO oo ON '-- Cl ot co m ^r ■'t to wc NO d oo On On o d d Cl d d
7—1 ci Cl ci T-H

o Ot ON NO CC oo Cl Cl re Cl ON On On oo re d d On oo Tt ON ON wc o d ON d
d On ot NO NO wc NO oo T-H ON o On d m d d wc t-h d Cl d d "'t NO Tt ■^t d t-h wc d
d T-H ot wc WC or Cl ON Cl wc c~ wc m ON oo wc On d CO re d o d oo ON Cl d o NO NO
d O Cl or NO oo o r-H oo O Cl ^t NO d o CO wc oo o d cr NO d wc NO ON CO ';t d
oo ON oi On ON ON o © NO d d C~ d d oo d oo oo CT\ ON On On NO NO NO d d d d
T-H

vO
o NO Cl OO O r~- co co
OO
d cr c~ Cl d On Os d On On d ON NO NO
ON
ON NO Cl d ON WC
or d oo wc © © d On OO wc o ■rt d T—1 re f-H w-. NO ON ON NO oo oo d ci t-H d re oo
ON
o n WC On ot ON Ot O cr wc NO NO NO d oo O d wc ON CO OO 't O d Cl O’ wc wc NO d
d ON T-H CO NO OO *-H Ot d oo On O 1—1 d cr d ON 1—H CO NO oo t-h er d oo ON O T-H d er
CO CO or ot wc wc Cl Cl Cl re CO CO CO CO Tt wc o Cl d d CO CO CO CO
03 cd

ex Ph

S S
or oo wc wc o Cl
wc o Cl On wc ON NO d CT NO wc 1 NO oo CO oo wc o NO ON On d d CO ''t

T-H oo CO ot ci d d co to ci cc ci ci re oo ON d d CO* t-h NO NO ci o CO ^H ON CO wc d ef-


o d NO wc wc wc NO oo
©
NO cr Cl o oo NO d ON d NO wc wc wc NO oo 1_\ oo d O On d to Cl
CO ot NO oo o Cl ot NO WC NO c- oo oo ON T-H Cl ■^t NO oo o d eT NO wc NO d d oo ON T-H
CO CO CO CO ot Ot II Cl Cl Cl Cl Cl d CO CO CO CO CO "Tt er II d d d d d co
ex cx

On ON NO re On On NO Cl d Cl NO o d ON wc NO
re »0 NO d OO ON o cr cT 5 Cl o d o "'t NO ON Cl Tt d d
ON Tt NO Cl d Cl sO cr
t—I d o cc NO ON re d Cl c- Cl d T—H t-h o ON OO oo d NO wc wc On o CO wc oo o to
o Cl ot d On co NO re cT Ct wc wc NO d oo OO ON o Cl CO er d Cl Cl d CO CO
Cl ci Cl ci ci co co co T H t-h

o o o
0
o o o o o o © © O O o o o O o o o o o o O o o
o o o o o o o o wc o wc O o o O o <o o o o o wc o to o
NO d oo £ o T-H Cl co a Cl Cl m re eT wc NO d OO o T—H d CO cd Cl d CO co Tt wc
▼—< ■»—< *—< GO CO

533
o o so os r~ Os ■o
1—H SO sO oo Os r- rr it m — r~ ■"t r~ <r <r —« o r- OS m os
co IT «T «r
r Tf IT *r N- tN os o «T SO sO r- —- (N oo so OS © ©> o tN it tj-
Tt "f SO IT <N rr O Tf r- OO o. oo sO m n r~
oo co »T 3 Os 3 tN rr »r 3 Os *“
rr N- >T l"; On — 3 3 On 3
ri oo oo OO oo OO OS Os oo od OO OO O^ sd sd
sd sd sd sd sd t-i t~i ri ri oo

oo oo © IT Tt >T 3 n
r- Tf >r OO IT it © sO SO — 3 m Tt so OS OS sO 3 tN
© Tt" o rd -X sd r~i 3 sd oi Os id (N rd ri r~i © ri o 0\ N" >o (N id ri rd Os
it Os tN >r Os rr oo 3 o Os tN co Os —« m oo m OO ^t o Tt o tN
-c —H cr l""1 Tt‘
N rr © 3 so
On sO OS SO oo oo O 1—H <N 3 so OS m SO 00 Tt n 00 00
rr rr rr •'t It IT >r 04 t"
<N (N tN rr co rr rr rr m N- ^t N" Tt ITS >^s '—' tN tN
03 03
CX
03
Q-
CX
s 3 so O —> «T — 3 3 s rr m tN <N OS os rr r; m o
s r r~;
so SO oo o O o
o
O 3 rd Os —C OS 3 3 © lO
3 3
Os r~ IT) it, 3 it sO oo
CN 3 so oo o <N SO
©
ri
O
o
od
tN
OS
r-
ri OS
r- ITi
«d
N-
o
o>
o
r-
Os od
itn "^t
ri
ir,
rd OS
so r-
rd
O
rd
n
SO so sO r- OO OS tN ^t SO 00 o tN it sO sO sO
rr rr m rT 3 3 Tt
II II <N tN tN <N tN tN m m m rr m N- rt ■^t ■^t ii n tN

Cx
ex ex

tN t- Tt r- r o OO o r- tN
O oo On
ir SO sor~ Tt iri tN so so tN r- o rr sOoo T-i o r-
tN >r OO
SO Q\
OO 1—H
n NO
ir, OO IT) oo Osrr 1—H oo 1—H IT Os m so o rr sOOs rr r- oo
tN rr rr 3 Tt ■^t IT
OSO i-M n m IT r- Os n it r- Os sO IT IT
o 1 1 1—^ n rj n tN m m m o o

IT oo tN N" tN r oo oo ©
OS '^t oo so NO © "O >T rr m © oo On tN
rT oo r r- r Os tN Os tN tN oo IT 00 0\ os sO r-
if oo tN tN os SO o SO ©
rr r- oo o n r OO IT o\ tN rr rr tN © oo OO oo
OS 3 SO oo o rr 3 SO oo © 3 r Os tN Tt so OO © H — tN
ri od oo OO od On OS Os
sd sd sd sd ri ri ri ri ri OO OO oo od OS os sd sd
Table A-3c (continued)

rr © oo tN © © OS © IT Os ITi (N N" © tN OO
r- © tN tN OO r; tN 3 so OS
sd ri rd ri ri OO ^t ri
id rd «d id
ri sd On © On od sd
OS n IT os
© os CT
rr
©
00
oo
© ri
©
Os N"
OO
tN
Os ©
'it IT ©
tN N-
OS
©
>T £
rr
rr
©
oo
oo
N" ©
■^t
rd
m ©
OO
IT
OO
rr m ■*t N- ^t IT IT tN r-i
n tN tN m rr rr rr m rr N" ■^t N1 IT «T tN <N
03 03
CX CL

SO Tt r; so © so On s S
3 © © os © r; Os oo sO Os IT <N r IT ©
id Tt os © 'd id © © ©
Os r- IT IT IT IT © OO oo oo sd sd sd rd ri ri ri ri Os OO rd rd Os o 'Tt ^t
os CT OO © N" OS r- T 'it rt IT © r © Tt
tN '»N4 © OO © tN N © r-
rr rr 1N♦ TMf IT © © r oo OS n N" © oo © n 3 © rd © ©
CT C<) Tt ■^t rr rr tN
tN rr <N tN tN tN rr rr m ^t Tt N" tN
II II
ex ex

tN r- r~ r- © oo oo
On Os OO «T tN oo IT T-—i N- r- OS tN IT N- IT © IT,
n 1—• r- © IT © n <N r-H © oo © it
© OS © © OO IT n oo © 1 r- rr © ©
m r- T—t IT OO tN © © © ©
m rr N" '<t <T IT ©
1—( T—^ r^ "it >T © Os rj ■t- r © ri IT r~ © © r~
I—1 1—* 1—* tN r4 tN m rr m rr ^t © ©

© © oo OS r- tN m m "Tf
m os H 1—1 T— n rt oo oo tN 3 "O "O © © *T OO IT oo r- IT Os IT
—H 1—1 rr OS r- Os oo rr © m rr m -o © r rr OO
tN r- OS t-H 1—1 © oo *T n
n 3 *T r~ oo © rr rr © IT oo OS Os oo © m >T © ©
© tN 3 r OS 3 IT SO oo tN IT oo © tN so oo n
© 3 © n tN 3
OO OO od oo OO OS OS OS
sd sd sd sd ri ri ri ri OO od oo od od Os OS sd On sd

OS r^ On NO SO 3 3 rn n OC — Os Os — m —<
On t4
XI © 3 tN © tN r~ oo oo © © Os
ri rn Tt ri ri OO <d ri •
i—i QN >d N" ri rd os ri id ri rd sd rd sd ©
OS -c OS <N IT os rr 00 N- © rd
t" © ON
rr
rr
it N-
©
Tt
oo i-^ ^t
■o IT, IT
©
tN
ON
oo
C^)
o\ O i—
>T r~
tN ^t
Os i—i
© Os
IT £
CO
rr
©
oo
oo
^t
i—
©
^t rJ
rj
©
oo
©
oo
OO
OO
<N rr rT rr rr rr it N- ^t »T >T tN tN tN
03 03 «3
CX CL a.
s oo rn 3 tN >T 3 so rr
cn
S
O o O r- 1 *—1 3 m «T © r tN OS ©
3 NO NO
O sd © ri © <d id i—i •
3 SO NO <N sd © ON ri od cd Os rd N- Os IT, rd >d ri
os r- © IT «T IT © oo
tN © oo © n © o\ Nl ON oo © IT 1—H ON IT >T it >T © r ri © © ©
'sC VO r- oo Os 1—1 N- © OO © n it © © © ©
II rr rr rr O’ N" "0 it
II <N (N tN tN rr m m it 'it II tN tN n
cx CX ex

© r- 't n © »T OO r- ©
oo m r- 1—1 IT oo r—H oo oo oo © IT © OS n ©
i—i r~ Tf © r rr os © CO © N" © r- oo oo oo OO
Os 't oo m r- tN rr (N OO 't © © os oo r- © IT Tf CO os © r-
©
N- 5 ■'t IT >T NO © r~
tN m •Nt to r tN
tN
r © CO ©
n n CO CO CO
On
CO
tN >T
^t ■Tf
r- oo 00
© © ©

o o o o o o . »r o o o So o o o o o o §© o
© © © © o o o o o o © o o
it
hn so r~ oo os tN m 3 <N IT © IT o <N >T
<5 <N <N rr m Tj- IT so r OO On O <n m 03 tN tN
GO OO

534
Os «n 04 04 Os 0- ot © 00 On) rp ot m Os © Os Os 04 © m On) © m
6.4461

04 ©
O' o m O' rp cp 0- oo OS On) m 5 oo Os lO) 00 «n oo 04 Os On) m in m
>n ot m rp OO 1—H On) 04 r—i O’ O' O' m © of 1—1 m m © >n cp © g
sO oo o SO Os co «n O' oo OS © On) © 00 OS On) in O' Os m >n O' O' O'
sd SO od 0^ fd od o' OO 00 oo oo oo in sd so © sd sd sd od od O' O' od od od od mi mi

©T
O m 04 ot oo O' r-H m m m of O' 04 >n
2977.5

Os OS O' OS rp oo oo O' © O' O') o © ©


ot o) © 00 oo m ot © 1—1 OS 1—I rp ot oo of od >n © cd sd © O' od mi od mi OO od mi
o o) rp m O' of oo rp 00 on © Os rp On) © Os 1—1cp © cp O' On| cp cs mi
r-H © O' Os m OO
04 rp ot SO £ 1-H rp © oo i—< of o) O' 00 Os © 1-H cp ’t © £ 1—1 cp © 00 1-H cp On) O' O'
cp m rp m rp m ot ot of ot m m 04 On) O') cp cp cp on cp cp It ot of 'It »n <n On) On)
Cd
Ph O)
s s
2738.0

rp ot co O eo oo © On) oo m rp © O' © O' © © © © O' of © m © 00 os


o
in SO in rp 04 ot r-Hm OS Cfi mi © O' r-H OO 00 sd Os --- od sd © © >n sd 0 4 os
m 04 1“H o 00 SO m of of •n O' OS m os © Os Os O' >n of ot of of in O' oo © £
oo Os o 1-H 04 ot so oo © On) of © in >n © © oo £ Os © 04 it © oo © 04 © m m
04 r i cp rp cp cp rp rp ot Ot or of On) On) 04 04 On) 04 m cp m cp cp of of 5 'it II On) On)
II
cx
a.

SO oo ot Os SO O') m © m OS On) Of OS OS
.068 42

oo rP © O' On) O' OO O' © On)©


O'm Os 04 Os m © ot OO of 5 't rp Os OO cn >n © of 1 © © ot oo O) «n 04
soot 1—H OS rp so © Of O' © of O' Os 1—1 l—H O' m oo oo oo oo O' O' © m <n m Tf
O'00 Os OS 1-H O') of m © 00 Os © rp or Of m in © © O' oo os © On) cp it On) On)
o o o o 04 © © © © © © © © © © r-H ,~”1 © ©

© 1=H © i—< i—)


6.5390

OO O') 't oo >n On) Os OS 04 © On) oo 00 0" O') oo m m OS m m,


On) 1—' cn m oo © OS © O') 1—1 OS© 04 ^t Tt- © i—i m rr Os © r—H 04 ~t ro ©
It On) oo m © m oo OS © Os O' !—< 1—H
It 00 l-H © O' m r-H © Os © r-H © 00 >n r—( ro
O' OS © On) in O' Os ''t >n O' OS © l-H O') m, O' OO © ro in O' © 04 Tt m O' OO OS
© sd 0^ 0^ r-i 0^ 0^ OO od od od OO sd sd sd sd so sd O' O' 0^ O' oo oo OO oo od mi to

o\
m OS © m m O' OS OS © ro m oo cn On) i—( © O' m © m © m © ■'t
2993.5

't' 04 ro OS oo
mi © it sd 04 in mi 1—H rn oc
© 04 od rn cn © It rn OS © rn Os © O' Os sd Os 04 od
l—l it >n OO i—l 't' 00 m 00 "'t © IT) os © it oo © On) O' ro oo 04 O' m Os mi O' ro
i—1 On) m © Os m © oo i—i ■'t fNj O' oo OS © O') m © i—i m © oo r—H m oo O- oo
m m m it It m m 04 C4 04 m m m m CO ro 't 't it It m m 04 04 04
03
Plh c^3
CX
5
O' oo ’t © © m >n <n "f on Os © m © oo m © m © Os ro n) © oo m s m 04
2750.1

rn 04 c> od mi sd rn sd mi © © sd in © © 04 0^ m mi mi sd Os oo 04 t-H sd O' ro © © 04


^t m On) © 00 © >n it it «n © O' © m 1—^ 1—H i—l © Os O' m ^t 'f ■^t It m O' 00 ro
O'”
oo Os © l—l On) © oo © On) "t © © © O' 00 Os © © O') 'O © oo © 04 ''t © mi ©
On) On) m m m CO CO •or ^t it 't On) 04 On) O') 0 4 m CO ro ro ro m it it it ^t II 04 04
II
cx Cx

m © O' OS cn oo Tf ©
.081 14

OO O') © © >n 't «n O' r-H m, ro © oo ON O'


‘O m oo it m 00 Tf OS m © © © O'
© m cn ^t O' O' m © ■’t © >n os ro m ''t
OS O' © 0 4 oo ^t os m, © © 04
m It O' oo 'it © © O' oo OS OS O © © O' OS
Os © © O' Os n) ^t ^t in m © O' O' oo Os © ro ■'t m © O') 04
© © O') On) 04 © © © © © © © l-H
© © © ©

, O) Os , , 1—H oo O' 04 O' © ©


6.6438

m oo © 'f © mi © ro m >n r-H


© ro oo 04 04
© 'it "t ro © ro 04 m, © © © OS © oo r-H 04 Os © © oo Os © ot © © O' ©
Tt O' 04 Os it O' OO oo O' <n 04 O' 04 © 00 © m Os © 1-H in © © •n m r-H OS O'
OO © ro m, oo © 04 © oo © © 04 ro «n O' Os © ro © oo © 04 Ot © oo OO ©
Tt oo o
sd o 0^ 0^ O^ OO oo od od od Os sd sd sd sd © © 0^ O' O' o oo OO od oo od mi sd

m m oo ^H ro m, 04 © © O' © o Ot 04 O' »n © m ro OS m ro O' © r-H m m 04


3008.8

Tj- ot
sd Os © 04 sd it oo O' ro Ot r-H of 1“H sd © 04 CO © mi ot mi r-H 04 od © od © © Ot ot
04 ro «n © oo it oo ro oo "t © >n © oo © os l-H ro O' ot OO 04 oo O'. © mi 00 OO
1—< 04 ro •^t © Os ro © oo ot OO oo Os © 04 ro 5 © £ i-H m © oo r-H Of O' O' oo
04
TO ro ro ro ro ro or "t Tt of m m 04 04 04 ro m ro ro ro ro Ot ot m m 04 04 04
cd c3
Ph Oh
s S
m, mi oo ro O' Os 04 m © © os © © rp O' 04
2761.6

Os © I"; ro Os O' 1—H


© ro Os
©
Os mi 04 oo oo mi O' sd 04 o^ © (N O^ t/o sd Os © Os Os 04 © rd 04 od od OS O'
>n ro 04 oo © in ■«t ^t <n © O' © © 04 04 r-H i-H Os O' © m ot ot Ot m O' sd oo ©
oo Os © r-H 04 'it © oo © O') of © © © O' OO Os © © 04 of © OO © 04 of © in ©
04 04 ro ro ro ro ro ro Of- Of It ot 04 04 04 04 04 ro ro ro ro m ro ot ot ot 1! 04 04
II
a
a

■t
oo © 04 m 04 ro m <n O' OS »n O' O' © ot ©
09890

© © © OO O' ot
O' OS ro ro l—H
£ VP (N oc IT) O' in OO Of ot © Ot oo Os oo © ot oo m ot 1—H
OS © © Os OS oo O' m It ro r-H Os ot oo © CO © © oo © 04 ot © oo os r-H 04 ©
© 04 ro ro m O' OS l-H ro in O' 8 ot m «n © O' OO OO OS r-H 04 ro Ot m © oo m CP
04 O) 04 04 04 © © © © © © © © © ©

© © © © © © © «n © © © © © © © © ©
>n © m o © © © © •H O' O’ m O’ m © O’ © O’ © o © ©
m 'it m, © O' oo CT\ ©
r-H 1“H
04 ro
i-H i—H CO
04 ro ro Ot ot m © t OO o r-H
i—(
04 r*N
1
£ cd
CO
CP

535
p OO NO 04 m It NO) 04 It 00 t'
i—
©
O—«
OO
^H
© © OO © NO) 04 O' >Ol O' O'
o m NO, 4 o 1—1 O' non 00 1-M 04 H— l-H ON © OO ON © oo -
S Tt
m NO non 04 O' 04 OO ON ON 1—4 00 NO it O' © 04 OO 04 NO) ON ^H 04 ON O'
r^i no, O' OO o 04 non O' ON ON It p © 04 ^t © ON © 04 non P ON © 04
NO NO NO NO NO t'' O- O' o^ O' OO OO oo NO) «o] NO NO NO NO NO NO ol O^ O^ O^ OO OO

ON
© co o-i o ON O NO, non OO oo 04 cn oo OC 04 © ON © NO) OO OO © -'t
o p OO ON ON
© -e
>o] O' OO 04 OO H 04 04 00 NO ® 00 r-] ON NO cK <> 1—^ •ON © NO ON 04 ON oo oo
oo ON O' ON 04 P oo 04 NO O' 04 ON 04 O' 04 m On O' © ON «A) © NO) © NO, HH oo
ON 1—^ no-, co OO m "O’ OO ON ^H
ON 04 ON NO 00 ON NO oo ON © o © O' © oo ON
04
04 ON m CO it It It »ON NO) 04 04 ON ON ON ON ON ■^t it Tt It NO) NO)
03
03
CL
Oh

S3 ,
s
o- oo o- ON OO P ON O OO NO >Ol © NO, p © ON NO) P © p p ON © — © 04 ON 0©
© 3 -L
NO) ON <o] -L
NO P On P ro NO 04 04 oo ON NO) oi 04 no] it 04 © ON 00 On It
od p NO NO NO NO, NO-, ON m m ON © © 04 00 04 04 04 04 04 04 04 04 ON NO)
O' OO ON o 1-^ 04 NO oo o 04 © NO) © O' ON © 04 O, © OO © 04 it ©
II 04 04 04 ON ON m r^i Tt ^r ■'t 't II 04 04 04 04 ON ON ON ON ON ON ON it Tt Tt ■rt
cx cx

NO) ©
<o, 04 O' NO) NO non O' NO o\ © ON 04 0 Q\ © ON oo © ON O' OO NO) © m
CN
ON m 1 P oo ON o O ON oo OO it o ON © © 04 ■^t © © © NO) Tt
ON p OO —H non oo ’-r O t" ON OO m © © (N NO) OO © 04 it OO 04 © © it oo
04 m P or p NO, NO NO O- I-" oo § 1-^ 04 <N 04 04 m ON m ON -<t NO, NO) >Ol
o o o O © o o o o o ® © © © © © © © © © © © © © © © © ©

m oo o- NO NO p NO 04 o H— OO ON © © <—i o\ OO O' 04 NO, NON


r- © O' O' ON
Vi oo r- 04 4 oo ON O' 04 ON 04 ON Tt © P 04 On © © 04 ON oo O' O' 1-H ON NO) OO
04 ON ON p OO •*fr OO ON o ON H HH
r; NO tp ON ro non r- NO) P C\ NO © ON © © 04 ON 04 © O'
5 O O; p m © O' ON 04 P NO) O' On © p p © OO 04 ON
©
NO NO NO NO NO o^ O^ o! O' oo OO
oo oo «o] •ON >o] © © © © © O^ O'] O'] O'] O'] 00 OO od
Table A-3c (continued)

oc o
,
ON m m
NO
p — P ON © ON 04 OO 04 —
oc -c
04 oo oo OO On O' p NO) P NO) © oo ©
NO 00 O' o o o oo OO On 04 cn p ON ON © © ON © ir] oo © p P P «o] ON O'
©
m oo 1—< ON •O') oo 04 r" i—H O- m © 04 © 04 On P O' © 04 P O' © © 04 oo
oo o i—' 04 P NO) NO oo 1—1 ON NO OO v—m m O' OO On © 04 ON NO) © O' oo HH ON © oo ON
04 co
m ON m m m 'NT Tt NO) NO) 04 04 04 ON ON ON ON ON ON ON p P P p NO) NO)

o3 CS
P CL
S3 "O' NO © of 04 O' ION non o ON O'
s P P 04 P P OO © P 04 P ON ON OO © o\ NO)
o 3 o
ON OO oo ON O'’ © oo On NON ION ® fH O^ P © On 04 ON »o] P P OO P OO © K P- w ©
O' NO r-" O' r" NO NO "t m m ON NO) o' '-H ON ON P p P P ON ON 04 04 04 ON w
^t © ©
o~ oo ON o 1-^ 04 NO oo o 04 NO, © © OO ON © l-H 04 ON P © OO © 04 p
It
II 04 04 m 04 04
04 ON m ON It ■'t Tt it II
04 04 04 ON ON ON ON ON ON ON P P P
cx
Cx

©
04 © 04 _ IT) ON p
, oo ON ON 04 04 ON NO)
£ P ON NO or NON NON m ON o O' m O' © oo P P t' O' © © NO) NO) P m 1—1 OO
04 ON '-- ON NO oo OO •o, 04 ©
NO © O'
NO) ON P oo 1—t non 04 ON NO ON rn OO ON 04 © ON 04 NO) oo ON OO m oo ON O' 04
t'- 1—1 1-^ 04 04 04 m ON P P p NO) NO)
ON P NON non NO r" © © O-
P NO OO £ ON © © © © © © © © © © © © © © © © ©
o o o O o o o o o o ® ©

04 04 H-H P P © 04 ON ON — © ON — © © p
NO) oo m ON oo O' TT NO rt ON O H— NO)
m o ON o oo O' rn 04 non ON
© P O' P OO P 04 ON OO 04 p NO) oo
NO NO O- O ON NO) NO)
co P 1-H oo 04 NO 04 NO-, O' oo 00 ON 04 NO) o OO O' NO) 04
O' O' NO © oo © 04 OO ON 04
ION O' oo o NO OO 04 NO
p © © P © oo © 04 P
® 5
non >o] © © © © © © O O o o] o od OO OO
NO NO NO NO O' O^ O' oo OO OO oo

p o 04 OO 04 NO p 04 O' m O' It m © o p © © OO © 1 © ON NON c^< oo © O'; 04 04


© -s p
cd r- 04 © oo 04 NON 04 non ON NO 04 © © O'] © © HH cd non © p p o] © ON
>o, ^H
P r- o 04 p non On ON O' 04 O' m On cd P NO, NO) NO, oo HH CO NO) O' 1—H © © 04 OO
o~ o i—-• m of NON NO OO !—H ON NO oo l—H m o OO ON H—1 04 CO >ON © O' oo 1—H CO OC
© CO
04 co ON ON m m m m TT NO) co 04 04 04 m co CO CO co co CO NO) NO)
It It p p p
03 03
Oh CL
S NO On ON 04 NO ON
, l
OO OO m © ©
s oo © P P 04 04 04 © co NO, 04 CO CO 04 P
3
O p o] © P OO «o] to] 04 od rc] ON 04 On © © t' 04
ON 04 OO 04 P NO ON o^ t'' ON o^ NO)
© On P 04 p NO) NO) NO) P P co CO 04 CO co p ©
oo ON oo OO O' NO NON -r it NO NO) h-H
1—H © O' oo ON © 04 co P © OO © 04 P ©
II r- OO On O 04 NO OO o 04 ©
II 04 04 04 04 04 co CO co co co co co p P P P
04 04 04 rn m (**) m It it It It
Cx Cx

co ON of NON non 04 o oo On NO NO OO r' © co © O' NO) P O' ON © NO, © 04


04 ON NO o 04 NO) NO NON Tt m 04 © p 04 OO On NO) r' oo OO O' NO) © NO) OO p O'
04 O' 04 NO non ON 1 ON O' NO, ON fH © NO) ON m © ON 04 NO) NO)
OO P ON p © ©
P or NO non NO NO r- OO oo On © 04 04 04 04 04 m CO P P p NO) NO, © O' O' OC
O o O O o o o o o O © © © © © © © © © © © © © © © ©

o
no.
o
o
o
«o)
o
NON
o
o
o
o <^l
o ©
©
©
© ©
. NO)
04
O
«o>
8 0 0 0 o o o o
«o, o «o> p NO) o o
p P non NON NO t'' OO £ o 04 ON 03
00
co co P Tt- NO) NO) NO NO o- oo £ 04 co

536
On r~ OO OO CN cn 0 m 3 r- CN CN NO r- CN CN NO On O O NO r- O CO
O 8 1—m O 3 OO On 5 CO CN r- 0 Tf CN CN 3 NO CO CN r- cn OO 3
CN »n "it cn 8 «n ON in 00 ON oc t"- r- 1—1 t"- ON CN O NO Tf OO 0 ON NO
On in 8 *“H cn <n NO r- O CN 3 NO 00 O 00 CN NO 8 cn 3 ON CN 8 «n r-
3 in >n NO NO NO NO NO r~- r-~ r- 00 CO 3 »n «n mi in NO NO NO NO NO r-

00 t"~ CN in NO cn O r-~ 0 m CN 0 , 3 NO it 3 it O <n ON >n m ON NO CO NO


mi ON 00 O QO cn r~i m ON ON NO CN O 1-H >n CN cn CN 3 CO in On CO i—I Tf Tf* 06 CO
vC> O 1—4NO cn On cn O NO CN 00 3 O NO 03 NO 00 r- r- On 1-H ON m r- 0 NO CO
ro 3 00 O CN cn m NO OO O CO m 00 1—< to Pu r- ON cn NO ON CN CO in r- CN m QO 0 CO
cn cn cn cn cn cn cn cn 3 3 3 3 «n m i—1 CN CN CN cn CO CO CO it N" Tf »n >n
S
Oh 0
in
O cn CN On 3 0 3 3 r~- in cn 00 O CO ON — 3 r- ON
O 1—1 O 00 00
0 O r- CO r- f-H
O cn On NO cn cn 3" i-H NO CN r~ cn t—H CN 06 II CN It CO 00 1—4 1—4CN On ON NO NO 00 <n
cn ON H»H O 3 NO r- 00 00 On ON O 1—4 CN cn O in ON m CN NO OO 0 CO >n NO r- 8 O
CN NO OO ON O 1—4 CN CO m r~ O CN NO r On CN 3 r~ ON O cn >n r- ON cn NO
II CN cn CN CN cn cn cn CO CO CO 3 3 3 CN CN CN CN cn CO cn CO CO cn it 'it Tt
Cx

cn
'it cn m OO >n 00 cn O O 00 r-~ m l"~ »nCO 00 CO 0 m O m
cn 3 ON NO m r- ON CO m in r- in 1—< NO O O CN NO CN it CN r-CO I-' 00 OnNO
oc ON NO r- in NO 00 3 On 3 ON cn r- *“< it ON ON cn m <nm CN 00 "t 00CO 00
m ON CN 3 NO OO ON 1-H CO NO 00 i-H cn NO 1-H CN cn 3 NO 00 0 O 1—< cn Tt NO t"ON O
O O 1—4 1—< 1-H 1—< 1-H CN M CN CN CO <n cn O O O O O Cn O 1—1 1—( 1—« CN
O O O O O O O O 0 O O 0 0 O O O 0 O O O O O O 0 O O 0 O

cn it cn O NO NO ON —1 0 cn m 00 N- it >n CN OO NO CN OO r- m, CN CN
OO 00 cn NO m cn 0 OO cn NO ON 0 O CN O 't cn m O cn
CN NO N- ON cn
CN cn CN OO cn CN m m m cn 0 cn m 3- On cn cn O NO cn OO OO NO it
t- O CN it m r- 00 cn m ON ; ON it i—1 O CN Tt >n 00 0 CN "it NO 00
mi NO NO NO NO NO NO t' 00 cn it mi mi mi NO NO NO NO NO

m
p QO Os — it —< ON NO — a—H
<n It NO m, in CN Tt — Tt ON ON NO CN O) cn O p
3 OO
CN
CN ON
CN
Tt t-H O
ON CN 00 cn
cn It mi ON NO NO O —
m,
it" a—H mi
CN ON >n
cn 00 mi t r-a 4 NO ON 3-
in & O r- NO It OO O r- 03 ON CN
r~ 3- ONOO -t r- 3
cn m Os CN it <n NO oc O cn >n 00 cn 1 Q.
HH
r- NO CN 4 >n
OO O O CN m 00 e'¬ 0 cn
CN CN cn cn cn cn cn cn it it N- It <n m CN CN CN cn cn cn cn en it t it 't m m
cs s
Cu w*
s 0
cn ,
m O CN cn ON >n O 00 r- cn 0, cn m. 00 N; cn r- cn »n p 00 m. in 00 CN cn O
in
0 mi 3: O cn a—i NO 06 a—i n; Os NO 06 cn II K mi oi 0 O O mi mi 06 00 oi P NO
ON OO r- 00 ON ON ON O O 0 a-H CN 't cn NO a-H CN
m, cn r- O CN
NO OO O m
cn NO 00 ON O S--I CN
cn 3- NO NO 00 0 CN cx r- O it NO OO ON a-H CN cn r- Os ^H 3- NO m
II CN CN CN CN cn cn cn cn cn cn It it 't N- s-1 CN CN CN CN CN cn cn cn cn cn cn Tt 3 3
Cx
CN
O r- 3 OO 00 O 0 m OO OO NO NO QO NO <^1 ON NO
CN r- «n 00 NO t OO 1—4 NO r-~ NO OO ON m r- NO ON NO CN ON
ON m CN 0
3
r- m m m 1—( ON r-
3
m, r- r-
0
cn r- NO >n NO NO
OO NO
m CN 8
r- re m t- ON CN •0
1—I 0 >n 00 1—1 1—1 CN m NO OO O 1—4 CN m m c— ON 0 CN it
0 1^ 1—4 1—4 CN CN CN «N
1—4 m m m 3 0 O O O O 1-H i—l 1—1 1—l 1-H 1-H CN CN CN
0 O O 0 O O O O O 0 0 0 0 O 0 O O O O O O O 0 O O O O O O

00 3 ON NO 3 CN O ON 00 cn 00 O 0 OO cn CN m CN ON r- m, O m m CN
CN
On CN 1-H O ON CN r- 3 3 00 0 3 CN 1-H CN 5 ON NO O CN 0 3 O NO O 3
QO
O OO 3 5 i-H CN m, O CN CN 1—4 OO cn 3 t" r- m, r-~ NO CN cn Q0 t- NO cn
NO
cn OO ^H m m NO OO ON CN 3 0 OO 0 1-H 0 3 0 ON cn m NO ON cn in t"- Os
mi >n mi NO NO NO NO NO NO r- r- r- 00 06 3 mi mi m <n NO NO NO NO NO r-" 4

4^ m 3 m, CN NO NO m cn l 3 00 NO 0 m O 0 CN cn r- 3 NO m
3 ^H m
3
[ CN ON 3
CN
fsj O CN <n NO OO 00 CN CN 0 CN m NO CN CN NO 00 O NO Os CN mi 1—4
r- K oi od 00 oi 3
1 ON r- «n O OO 1—4 3 ON 3 ON m 1-H r- cd 3 OO O 3 NO cn ON cn r- 3' O NO CN QOm
NO NO ON i-H cn 5 m, I'' 00 O m m 00 i-H cn p_ 00 m OO On 1—4 cn 3 NO r- O cn mi 00 O cn
rJ CN CN cn cn cn cn m 3 3 3 3 m m CN CN CN cn cn cn cn cn 3 3 3 3 m m
03 S
PH 0
mi
s CN
m 3 m» Nc> p ON ON
°) 3 NO oc p p CN C^ cn >n Os cn cn O ON CN p
NO ON
0
mi O 0 On NO 3 06 0 O O 1—4 mi «n On
CN II od O oi O 3’ 1—4
3’ cn
i-H (O CN NO
mi in CN r- 1—4 1—4 1—4 1—H 1-H cn 3 Os cn O CN 1-H cn •n r- 00 8 O1 i-H
3 Os O O CN 00 NO CN
'f <n 00 ON 1—4CN m O 3 00 O CN O ex r- 3 NO r-~ QO O i-H CN cn m f" On CN NO
II
CN CN CN CN CN cn cn cn m m m 3 3 CN CN CN CN cn r^i cn cn cn cn cn 3 5 3

Cx

r- O ON m cn CN cn 3 r- cn NO CN >n
cn r- O cn i-H
O i-H 0 NO m o> cn <n r- OO O CN CN cn cn 3 1—4 3 0 CN 1—H
O
cn 3 5 OO On ON 00 NO i—1 3 r- O CN 3 ON O OO 1 1—H 3 NO ON O 1—H

0 1-H >n 00 O CN 3 NO 00 CN »n 00 CN m OO 1—4 VO ON 1-H CN 3 in NO OO 1—H cn in Q\


1-H i-H 1—4 1—4 CN CN CN CN CN cn m m 3 3 O’ 1-H 1—4 1—4 1—H 1—H 1—4 CN CN CN CN <N
O O O O O O O O O O 0 0 8 O O O O O O O O O O O O 0

.000
ts »n © m
o © o
«n © «n
g m © m OOO
t-~ p CN m © in 8
5 cn 3 3 in m no no r-~- cn 3 3 3 m m ^ vS P CN cn
00

537
i

oo © „ ,_ _ 04 ON -d- Os m, m on r-
04 04 O 1 mi 04 04 O O 04 ON oo on
C*~) © on 4 p OO oo OO i-^ i—t o oo
ON p ON on © OO © p NO ON p on p
rn ''d' ■"d" Tf mi mi mi NO NO NO NO r~i r-i t-i

mi p o ON 04 <N m mi — m. 04 © 04 ON
cd ON ON T-4 ON r-i NO r-H p ON ON i—i mi OO O' 'O'
ON ■d- o~ NO ON NO mi OO ON r- p 00
P NO oo mi OO 1—1 ON mi oo i—« of O' o 04
tJ- p mi mi
S <N 04 04 04 c*~i ON ON ON

o
o
©
P p ip ON NO OO p OO 04 mi O O' Ip 04 p
oi mi 04 ON © OO 1-pH OO t"i ,—i — NO O' OO 1—H
© O' ON mi ©i mi NO 04 O' P OO o 04 ON mi
NO O' OO O ON NO 00 o O' NO Ch 1—1 ON mi
1—1 — 04 04 04 04 ON ON ON ON ON O' O' O

OO mi mi
04 ON NO mi NO © O' mi 04 ON 00 © © O' mi
O m r—1 oo mi mi m ON O' mi © oo © 4—1 4—^
mi NO oo © ON ON oo mi 04 O' mi O' O' 04
r-H H 04 04 m O' mi © O' oo ON © i—i 04
O O O © © © © *o © © 1—» 4-H 4—1
o o © © © © © p © p © © © ©

© 1—1 O' 1—1 © mi oo 04 On © 04 © O' OO oo


ON C*~) ON oo 04 oo O' O' OO © OO O' oo mi ©
© O' 00 O' O' 1—1 ON i—i 04 oo 1-^ —- © oo
© 04 m. p m, OO © 04 m> rp © 04 p m.
ON O' O' mi mi mi NO NO NO NO O-i O^ o^ oi
A-3c (continued

© © © © »—i mi © OO OO vc> p ip mi 04 ©
NO O' © O' © ON O-i 4—^ NO O' P © 0~i ON
cd T—1 O' © 00 04 1-H O' P i—i r^\ 04 © 0" ON ©
P oo © 04 O' © 04 P © Q\ 04 mi O' © ON
i 1—! 04 04 04 ON ON ON ON r^i O' O' O' mi m

O
m
NO i—; O' © mi © © mi mi OO mi oo
OO oo oi ON mi ON 04 ON © oi o © mi oi 04
ON oo mi mi 04 © O' ON ON O' 4—< ON O' mi O'
© O' © —H mi O' ON © 04 O' r- ON ON mi
i—1 I—1 --1 04 04 04 04 ON ON ON ON O' O' O'

0)
n O' ON
CO ON © O' © 04 oo 04 © O' © T-H
© 4—H ©
mi © OO ON 4—1 4—1 © 04 O' oo 4—1 ON ©
m O' © O' QO 4—< ON O' © 04 O' O' m ©
i r-H 04 04 m m, © © O' ON © 04 m O'
© © © © © © © <^> 1—H 4-H
© o © © © © © o © o © © © © ©

p On ON © m O' O' © 04 © © ■o O' ON


<© m 04 m © oo i—i m m © m m NO 04 ©
04 4—1 © O' O' T' i-H © O' © i—i ON OI 04 ON
OO m, Os O; P © 04 ON © On ON m. ©
ON O' O' mi mi NO NO NO NO NO o^ O^ O^ O^

oo ON OO OI O' © 04 P OS © 1—1 O'; m


04 © od 04 oi oi NO © —i OO r' ON 0^ oi
cd O' ON ON p O' O' 04 OO O' m 04 Os m 04
P O' ON m Os ON m © ON 04 m O' © ON
1 04 04 04 ON ON ON ON ON O' m m
S O' O'

o
o
p © Os 1-4 O; O; © © © OO O; © p — ON
O' O' N6 mi od ON 04 oo oi o^ oi Tf oi © Tf
O' m ON © O' © 04 mi oo 1—1 ON m © OO ON
© oo © ON © OO © i—i 04 m O' ON ON m
,—i '~l 04 04 04 04 ON ON ON ON ON ON ■4f

04 ON 04 -'fr ON T—1 ON 04 id- 04 © Os


ON ON 04 oo Os © "d- 04 © 04 04
m © © ON © © Os m ON ON © Os 04
04 ON m OO ON ON 04 m © OO
o o o o o o

mi O Q O
04 m © m
in it)

538
m

3.4247
3.6546
oo t- SO d SO n >n Cl © © 't © m ©
■t <^ (O SO re in n it © 1—H © >n d d Os ©
00 *o OS SO d so Os 1—H Cl Cl Cl f—< Os d >n re d
© o Cl m, OO o Cl «n d Os re it © 00 © d
re -P-H 1—H ci ci Cl ci ci co re

>n Os 00 d oo OS d OS ©
o P—1 p—i

1591.9
1453.2
"t Os d SO OO d d © in Cl «n © m
fN
-C o «n
1—H
d o re so © n- oo It © oo d Os rn d d
ci 1—t Os oo sO "t re P—H OS oo d >n it re cn <0 co
SO i—h d re f «n >n © d oo Os © i—< d re
m i—< pn p—H p.h

cs
(X
S <n so © i-H oo ■t © © P—i ©
■"t

1431.1
1567.5
so O d «n d "t © d d © Os oo © <n ©
o
to
a «n <o >n oo i—< it oo ci d re P“^ Os © it ci
p—H
SO
p—(
OO oo SO ■"t <0 i—i OS oo © in ef d d P—H
in i—* d re Tt It in © d oo Os © 1—1 d m
P—H
II

— _

001 472 4
os

001 631 1
oo O re >n Cl d d 00 re oo "t © 't ©
OO d <n P-H o
1—H
d © Cl © ,—4 <n re et --1 >n oo d
in Os Os O d <o n d Os P-H f d p—I >n © d
$> sO Os Os O O o © © © © *—■ d d co ro
p—H o P— — P—H p— p—H p—H p-H p-H p—1 P—H p—P p—- p—11 p—-
o © o o o © © © © o © © © © © © ©
o o o o o © © © © p © © © © P p ©

© d >n © oo oo d os d r- «n oo Os d os oo Os
Os © ■t oo in oo OS oo Os p— d d re d Os It ©
S«5 in © OS © d © Os »— d m d 1—1 © oo © in ''t
re © d >n oo © d >n d p re >n © oo © Cl
re X p—H ci ci ci ci ci re re

't
m reoo Os re © ef d oo Tt
© © © p oo >n © © oo © OS T- -- re
© r<]
Compressed Liquid

-C d © rn s6 Os ci sd © >n d sd >n 00 rn it ci
© ON d »n ■'t d Os oo © in re m re t
it d re it >n in © d oo Os © d re
re 1—H
Table A-3d

03
X
S Os © »n © OS d oo oo rem
©
© © re m © © © m © p~* OS It
a rn rn © os ci sd © f © sd -f © p—l © oo
© OS oo © 't re i-^ © oo d in 3 r^) d d d d
r^) i—* d 't in >n © d oo Os © i—i d re
II
X

't d d t d <n <n OS d rn Os © in d in © d


d «n d re d 't oo t re in os oo © OO P-H d
in Os Os © 1-H d »n d Os 1—H 't oo P-H 3 d OS
* it Os Os © © © © © © © *—■ d d m re
p—t O P-H P-H —i p—i P-H p—11 P— P—1 P-H P-H p—- p-H P-H p—H
© o © © o © © o © © © © o ©
© ©
© © o © © © © © © p © © © © p ©

d © »n in © © re re «n „ in oo Os ©
© >n © 00 d re re ef d it >n d d re
>2 d 8 Os d d d © d re re re d OS oo
os © Cl in oo © re «n d os re in so oo
ci *"H ci ci ci ci ci

•^t «n d © >n d Os >n d in


Cl © SO os re oo d © p Cl OS t «n re
Os'
Os -C •'t in OO «n 00 ci d d oo in re d 't
m oo d m re d © OS d © m 3 re re
re T—H d re 't >n in © d oo Os ©
©
d

cd
X
Ct <n >n re d d © © d re
S 00 © so os d d «n 00 d so so it "t OS
© 3 d re sd © re d X sd ci Os oo OO X d
>n t oo © in re © oo d m It re re d
*—1 i—i d re t m >n © d oo OS © ’~i

cx
Os d m NO Os OO © © oo oo © © © 't OS
in d Os uo Tt © d © oo Tt re © © it
00 OS Os i—l d 't <n d OS d >n oo d d
d OS Os o © © © © © © »—1 >—1 *—1 d d

tiOOOOOOOOOOOOQOOQOQ
ci -t- © oo © d t- © oo © o et © oo © ci "t-
GO >—i •—i i '—i (N d d d d M r'l M

539
P OO O' 04 O co r- «o, —c P P ON oo O O' o oo O' i—i i—i
1 p 04 <o, P o •o, —H ON ON co i—i «0) NO ro (O NO >o) On o
© oo <0) O 4 O' oo ON OO NO OO »o> fN oo >o, 04 T—<
© 04 NO) OO © 04 P NO OO p <N 4 ©> p On <N p p OO
4 4 4 4 4 04 oi 04 04 c4 oi cn ro ro co CO

CO 04 l—1 ON O On ON NO) 04 NO)


© o <N p P OO CO CO On 04 p p <N p CO © 04 p OI NO
-s ON © 9—3 oi 4
NO ON oi no) © NO) i—< ON OO Os cn © oi © NO
cd P CO 1—1 ON o «o> CO 04 © ON r- NO P CO 04 <N 04 04 CO p
rN 04 co P NO) NO o~ 00 ON o 04 m p ST: NO O'
CU zn, H i—i
s
o
© o O NO oo P oo NO) O' i—l ON
NO, 04 O OO ON co oo p O' P p O' O'; p 04 P cn p o 04
II 3 —- 4 04 4 iri 4 Os 04 no] On 4 ® oo o' OO cn 04 NO 4
OO NO P 04 o 00 NO NO) CO o &\ O' NO NO, NO, NO) NO
a. 1—1 04 CO P P NO, NO O' OO ON Os o i—l 04 m or NO, NO
1—H 1—H i—^

NO p 04 04 CO oo NO) CO 04 so OO 04 ■or 40) © OO 04 OO


NO o O' NO O' o 't 9— < © •ct © © 9-H so OO CO OO
O' 00 OO ON o 04 NO) O' Os 1—( 't O' © 00 m © oo 00
On ON ON ON o o o o © © 04 04 04 co 'Nt NO,
O <® 9—1 9—1 9—1 9—1 9—1 9—1 9— 9—1 9—1 9—1 9-^ f—l 9-^ 9—) 9— 9—1

o o o o o o o o © © © © © © © © © © © ©
o o o o o o o © © © © © © © © © © © © ©

—0 ON O' oo oo © ro ro sO 9—1 On SO P 04
© ON © NO) so 5 1-^ On ON 04 ON 8 P OO m 04 On
© OO SO NO,
i— 00 © © © OO O' p 04 ON r- NO) P P ©
© 04 NO, OO © 04 p, p 8 p 04 p p 00 © ro p p ©
4 4 4 4 r-H 04 04 04 04 04 04 rn ro co CO P-
Table A-3d (continued)

04 ON On O' so co NO, oo ro
OO OO OO p p ON p 04 p CO © rn © OO p NO) P p
ON 9~H CO NO OO 4 P oo CO oo NO) CO 04 ON 4 04 NO NO) 4
04 r-H On O' >0) 04 © On O' © NO) p m 04 04 CO P O' CO
’~l 04 CO 5 NO, © © O' oo ON © 04 CO p NO) © oc

NO) O' p © © oo ON OO CN ON
04 p © © CO O' p oo OO p ON P O' ON P © P
4 p NO oo © CO NO © NO) 9-J CXD NO 4 4 4 © 4
oo © P 04 9-^ ON O' © P m r-H © ON 8 oo 8 © 04 oo
04 CO P P NO) © O' 00 ON © © i—i 04
1—H
m NO)
i—h — —H V-H © O'
—( 1—H9 9 9

© © 1-H 04 © © NO) 04 © ©
1-H O' NO) P O' © NO)
•O) OO NO) P NO) On
1—1 04
04 04 © ON 04 ©
p NO) © ON 04 so On
OO oo ON © O' co ON ON 04 ©
© 00 © co NO, ON co
ON ON ON © © © © © © i—' *—■ 04
04 m m P sO OO
© o o 9—< 9—- 9—( 9—1 9—1 9— 9—H 9-H 9—1 9—1 9—1 9—1 9—1 9-H
® ® ® © © © © © ® © © © ® © © © © © ® ©
p © © © © © © © © © © © © © ©

ON P co SO SO p O' 04 m P O' i—i © P ON OO ON NO) 04


co © 04 P © 04 i—i © ON © p CO O' NO)p O' O' O^ O'
<o © ON NO 04 SO ON i— i—t 04 i—i © 00 SO P 04 © ON © O'
© © 04 NO) OO © 04 p On p CO p SO OO © p co NO oo
P 4 4 4 1—H 4 4 04 04 04 04 CO CO CO CO co

rn
i—i 04 SO NO) © © © P 04 ©
© SO p oc OO © p © p 04 NO) CO © p © rn NO © co
NO co co
-c (N
© 04 NO) 4 © P 4 04 4 © On © © rn p’ 4 On
c3 OO 04 © OO so NO) CO 1“H © oo O' © P P CO co m P O' co
1 1
04 co p NO) © © O' oo ON © 04 m P NO) O'
cu
s
O
© On O' oo ON >0)
04 NO
O'
P NO
©
P ro
©
p
ON NO)
© CO ON p ON © NO P P O' OO
II 3 NO) 04 NO) 4 © co NO © NO) © 4 NO) NO OO P NO NO) ON 04
oo OO © p CO ON OO © NO) CO 04 —H © © © i-^ co ©
a O' T—1 04 ro P P NO) © O' OO ON © 04 CO P NO) O'

P OO 04 p on O' © oo «o)
NO) o OO CO © 04 NO) © O' P ©
© © 04 ON oo on ro on O' OO <N © ©
1 3
oo On ON co OO 04
ro ON ON © 1 ro p no 00 !—< © P ON NO) ©
©
04
On On On © O O © © ©
1—1 1—1
CO

1—1
N-
1—I
fN
i—i
04
i—i
04
i—i
CO
1—1
3
9—1
NO)
i—l
04
OO
1—1

© © © © © OOQOOOOOOOOO
CO
M 't vO X 8 8 § 'OXO(N^\COOO(NTt\OM
,o4o4 04 04 C4rorororoco

540
APPENDIX 4
SATURATED VAPOR
PRESSURES OF
REFRIGERANTS

0°C 40° C

Temperatures (°C) Pressure (kPa)

Water vapor 0.611 7.38


Ammonia 424.5 1547.2
Sulfur dioxide 156.6 624.8
a
Carbon dioxide 3484.0
Propane 472.9 1361.1
Trichloromonofluoromethane (R ll)b 40.2 174.7
Dichlorodifluoromethane (R 12)b 308.6 1219.4
Dichloromonofluoromethane (R21)b 70.7 286.2
Monochlorodifluoromethane (R 22)b 500.1 1548.6
Trichlorotrifluoromethane (R 113)b 14.8 79.0
Dichlorotetrafluoroethane (R 114)b 88.7 341.2

aThe critical temperature of carbon dioxide is 31.1°C


bThe term “Freon,” a registered trademark of DuPont, is often used for these
refrigerants.
C'
<D
t-H

P
c3
Lh U O O O O O >o o •o o
1> oo o- no *o © co co 04 04
a
B
<u
H

o~ co co o ON NO o- 00 04
u o On O' r-H co ON o NO O'
1 rv 04 co OO *o ON *o © ©
oc *o co i-^ ON 00 r- o- o- ©
03 *n *n
NO vO vO *o *o •O *o
> ^-H r-^
h““j r™^ r^ *~H

'—'

CO Cl
)h
*13
'5
cr
co
ON
©
ON
00
04
00
co
o-
O'
ON
o-
©
vO
ON
©
i—i
1—1
i—i
co
*n
o
00
»o
NO
o-
NO
r-
NO
i—i
*o
On
r-
04
co
1—1

Lii c vO o- OO© oo oo OO ©
APPENDIX 5

til 3 O o © © © O © © ©
REFRIGERANT

Uh VN co NO ON co 04 04
o © oo co Q\ © 00 © 04 OO
Cl, ON ir, 04 OO VOi © 1 © NO
CO a © On © 00 co «o oo © 04
<D > i—i 1—H 04 04 CO CO CO © ©
co co co co co CO co co co
■f-> oo
<D
Q. l““>
O o- 04 04 00 CO © CO O' oo
L a co © o 04 © oo CO © T—l
CL a> o^ o- t - o *o © co o\
>1 w
© c3 »n i—i o co On o *o co ©
c oo oo O' O' © © NO © NO
(0 o 03 hJ r-H H ^H fH rH i fH
U5
■ -t—' -4—1
cc G
< i_
<D U w 00 © 00 ON © *o of
*-> T3
04 04 NO NO © © r—1 O' ©
5 CO "3 Cl OO © —H C\ CO oo 04 O;
(0 CO ©
■ ON o^ NO «o CO oo ri ol ^H
1— 04 co © »o NO © r- O' OO
3 r-H
CNJ 1 1

c
O
1^1

o © L- — VO © OO VO, © O'
<D 1 o «o oo 1—1 © ON O' © Tf
X © CO 04 ON 1 ON co CO OO
(73 oo CO *o —’ oo
00 o’ o^ ON

LU > CO
*r-H
04 co
—H NO
04 1—1
oo © ©
CO 04 1—1
»o
T—
co ©
1—1

DC a>
£
T“1
w ©
©
04
NO
©
00
04
OO
©
©
co
NO
©
NO
O'
©

LL P
'o
>
"3

3
©
NO
i—i
NO
©
NO
04
NO
NO
CO
NO
© ©
O'
©
NO
©
•o
NO
© ©
*o
NO
©
©
04
O'
NO
©
OO
O'
NO
©
*n
00
NO
©

a>
Li O' O' 04 *o O' r-H 1-H oo co
D C/D 1—1 04 NO i—i T—H O' NO ©
C/3
J—( NO 04 04 © © © co ©
C/3 a3 04 CO NO oo 04 *o
© ©
u X)
Lh © © © © © © r—H
1 T—1
CL

af
>-(
P
■*->
c3
t-H
0)
a
E 0 0 0 0 0*00
oo o~ o> *o rj- co m 04 04
*o o

(S^

542
to ©
^H
«o
|
O »0 O «0 © »o o tO © »o o to o to o »o o to o to O to o
1-H
04
4-H 04 04 co co rf to to VO VO o- o- oo 00 o\ on © ©
1 1 ^H 1-H 1—H

H 1—4 4—4
1-H
CO
r-
ON
ON
O
^h
04
to
o
04
04
o
o-
ON
On
i—i
OO
to
p
p
00
p
co
CO
Os
i-H
to
CO
1—1
On
©
O'
o
04
VO
O'
to
to
o
04
VO
VO 04
oo
VO
04
«o
3
1-H
oo
o
O
m
i-h
1—H
VO
On
04
00
to
04
VO
«o
CO
On

VO
to
to
to
to
•o
'tfr
to
to
to
to »o
p
to
VO
p
to
p
«o
o
p
to
co
to
o-
co
to
CO
to
p
co
to
CO
to
o
co
to
04
to
04
»o
(N 1—H 1-H O
to
8 VO
p
VO
04
p
1—4 1-H T—1 1—4 1—4 i—i r—> 1—4 1-H ,—4 1-H 1—H i—4 r—1 1-^ 1—4 4—4 4—4 1—H 1—4 i-H —H i-H 1-H rmH 1—4 4—4

00 CO p to o 04 P CO to o ON to O- 04 ON 00 CO
C' o
4—4 1-HOO O O 4—1 »o 4—1 4—4

VO o to to 00
4—1 On ON oo r- O'
On o- oo VO 4—4 p VO p ON CO CO OO CO ON
00 VO CO VO ON to O' CO ON •o iH O- CO o VO co o 00 VO O' O
1 co VO ON VO
1—1
O VO VO 4—4

or VO oo CO p VO 1-H 00 ON 04 p VO
4—4
H O', 04 1-H p VO O' ON 4— 04
ON ON On © p p p p p p p p 04 04
1— ^H
04 04 04 04 cq co CO CO P
o o © 4—^ ^H
1—4 i-M 1—H 1—H 4-4 1—H p 1-H
4—4 iH 1-H 4—4 4—^ 4-4 4—4
i-H ^H iH —-
4—4 4—4

© It ^H O' © VO © On vo VO 04 VO oo VO to © vo O' p O' to oo O' © to ^H ON


04 CO 04 O' © oo co 04 O' VO ON P ^H ON VO 4—4 © 04 CO o^ oo It 00 O' 4—4
p vo
on i-H CO p VO VO o O' vq «o co 1-H oo co oo 04 p P 04 O' ON O' oo © to ON co
P o ON 1=-H CO to O'* ON 1-H CO to O oo © 1-H CO p »o vd vd vd vd to p © 4—4 o^
p p P to to to »o »o VO VO VO VO VO O' O' O' O' O' O' o o- O' O' O' vo p
CO co CO CO CO CO co CO CO CO co CO CO co CO CO CO CO co CO CO co CO CO CO CO CO

1-H 04 vo ON CO CO O' vo vo It 1-H 04 ON to VO vo © VO ON oo »o o


O' 00 O- ^t

Copyright © by E. I. DuPont DeNemours & Company, Inc. Abstracted by permission.


p 4—H 04 to vo oo 04 04 On 1-H to ^H 5 i-H ON 04 CO CO © vo © ©
vo CO ON p ON CO vo 8 © © 0© VO 1-H p © p 4—4 "f p ON o- OO © 5 o
oo vd CO 1-H 00 vd CO © 00 to 1—4 oc to 1-H o^ ON tt ON co O^ ON 1—4 4—4 ON o6 o
to to
1-H
to
1-H
to
1-H
p It
1-H
p
4—4
'It
1-H
CO
1-H
CO
1-H
CO
—4
<N
1-H
04
T—1
(N
1—H
4—1
r—H
1-H
1-H
©
1-H
©
1-H
On On oo r- O' vo it 04
4—4

ON 04 to © 04 CO OO 4—4 © © OO to ON "it 04 © ON 04 © 4—H ON OO i-H t-H O' vo ON


o- 04 On © or 04 Tt 04 to ON CO to OO 04 ON © © © It 00 oo vo VO ON VO
04 OO CO © p CO © OO vo to 'f »o vo oo 04 VO co p co oo r- 04 p ^t p
vd © to © Tt ON •tt oo CO oo CO 00 CO 00 ^t ON to 1-H O^ CO ON vd It 04 4—4 CO o
00 ON ON © © © 1-H 4—4 04 04 CO CO ■^t It >o to •sO O' O' oo 00 ON © 4—H 04 CO Tt
^H H 4-4 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 04 CO co CO CO co

© 04 00 CO 'it ON ^H © vo vo to
04 ^t ON 04 CO r- co 04 04 04 © vo vo 1-H '^t © to, 'tt ON to 'it oo OO OO
OO VO 04 ON to CO co © '^t OO oo © ^H © 04 i-H O- to 04 4—H O' CO 04 04 i-H 4-H
1-H p VO OO oo i-H 4^ OO to © It O' O' it 1-H O 04 04 04 © to vo ''t VO ON
© P O) co p ON ^t O' 00 to vo 4^ © p to OO O P OO © 04 to 8 04 p O
1—4 vd p to o © to © vd CO © OO vd It 04 1-H ON OO O vd NO to 'tt CO co 04 4—4
ON o- VO »o 'it CO CO 04 04 04 4—H 1-H 4-4 1—*

—« '^t
1 oo o 04 vo 04 o CO to NO ON 04 4—4 4—4 1—H
© to
vo
04 O'
©
00 8
to
CO
''t
04
CO
VO
04
04
VO
''t
CO
VO
OO
04
VO
OO to
4—4
to
vo 1-H © 1—H 04
© cn O'
oo 4—1 to,
oo vo O' © vo VO
04 4—1
O'
8-H
1
oo
to,
O
i-H
vo
co oo
-'t ^H
ON © © 1—H 04 CO TT to vo O' 00 ON H^H 04 It to, O' ON 04 Tt o i-H to 1-H ON vO ON
vo p o O o O- O O' O' O' 00 oo oo oo oo 00 On On ON © © ^H p o
© © © © © © © © © © © © © © © © 1-H 4-H 1-H 4-J 4—H 1-H
© © © © ©

© 04 VO 4—H to © O- On 04 © 04 to
NO ON VO to co co 04 VO O' vo co co CO ON oo OO to VO © »o ON 4—1 ON to »o
04 ON © OO 04 CO 4—1 O- Tt O © 'tt ON vo O oo SO ''t 00 OO vo P © oo »o
00 1—H p © VO 04 Os VO to 4 Tt vo oo ’—1 vo 04 On OO oo © CO oo to P to O' i-H
H 04 04* co co Tt 'cf to vd o 00 ON © 04 CO to vd oo © CO to © co vd ON ^H
i-H 4—4 4-H 1—H 4”1 4“1 04 04 04 04 co co co CO P

to © »o © «o © «o © to © to © •o © to © •o © »o © to © »o © »o © 04
l—H 4-H 04 04 co co p p to »o vo vo O' O' oo oo On ON © ©
i-H Pj
T T 1

543
so OS oo T—1 o tn © SO Os 04 '*t tr, >o co 1-H
Tf OO
co r- o co SO oo Os O O 1—H »-H ^H 1—1 ^H '—iO ohH
02 co SO oo o 04 no O' 00 OS o H 04 sO r-
so so SO so p p p p p p oo oc OO oo oo oo 00 oo
i—i 1—i H ^H ,—i t-h ,—^ T—H i—i 1-H 1—i 1—i 1-H T-H T-H 1-H V-H 1-H 1—«

o
o
O
SO
so so ^H «o O' NO 04 ^fr 04 CO os OS 04 oo O' 00 1-H
to
CO SO o tj- Os >o O' hH oo O' oo 04 ON O' oo ^H O'
04 TT p 04 co 1 o o OS OS os o O p 04 p to p
C/2 ,—i 1—1
U, Tf on © »o © co to 00 to OC 1-H rf o^ © CO
04 (N co Tf •o «o SO so SO so r- o- O' oo oo 00 eFi OS
co CO co CO CO co co co CO CO CO co co CO co co co co
PQ
04
SO
04
04
,-H o o to to 04 CO 04 Tt T—1 to O' O' O' OS CO OS
1-H 00 so oo oo oo OS 04 CO CO 1-H 00 Os co SO os © 04
On •tt oo © 1 T-H © «o p CO O' © p O' I—1 p p Of
o Os © 04 CO Tf to © to —H sd 04 04 oo CO oo Of On
CO SO © CO SO OS 04 Tf to O' oo © CO Tf so OS <©
so SO O' o- O' oo oo 00 oo oo os os Os Os os Os OS ©
“Freon” 12—Superheated Vapor Properties

OS ^H oo Os so oo so © so © ^f O' 00 os
oo O' Tf OS
04 © co so 00 © 04 04 CO CO CO COco CO co
co CO
co 00 © 04 'Cf o- On © 04 '^f
CO SO O' oo Os
to
so p p o: p p 00 oc 00 oc oc oc oc oc 00 p
—i p T—t r—H ▼-H 1-H t-H H 1—1 p i p p
1— < T—H 1—

co © 1—1 © O' CO oo SO os so so OS •O)


CO
oo 04 oo so SO oo © 'tfr to 00 04 co SO 04 © © CO
to p © Tf OS p CO 04 1—' 1—* 04 CO p p
OS 'Cf OS to © to © CO sd On 04 to oo T—i Tf o^
04 04 CO '^f «o to so so so so O' O' O' 00 oo oo
Table A-5b

co co co co CO CO CO co CO CO CO co co co CO CO co CO

oo 04 os CO O' co oo 1—1 CO co 04 Os »05 SO


04 t"; p OS r- to 1 O' p 04 ^f 3 oo © I—1 CO p p
'Cf oo to 04 os »o CO 04 © oo sd tn CO 1—i os 0^
04 oo ^f OS to 1—I so 04 to oo 1—1 CO so os 04 to. O' ©
1—1 1—1 04 04 co "^f ''t to, «o uo SO so so so O' O' O' oo
1—' f-H 1—H 1—1 1—1 1—H 1—1 1-H T—1 V—H H H 1—I t-h ▼-H

to 1—1 so O' co to 04 to to oo co m ro 04
so to oo 1—1 Tf so oo os © © © T-H 1—1 ▼—< H 1—1
02 to, © 04 to. O' OS 1-H co to SO oo os T—< 04
p p p p p p oc 00 oc oc oc p 8 ON p
1—i 1—H 1—i 1—i i—i p 1—i p T—H p 1—i p —i V—H 1—i
u
o
o
oo
to oo oo OS oo '*f so oo 00 co 04 to
1—1 1-H

so co CO tOi to oo co
1—1 —1 04 to © oo 1—1
C/2 p oc OS 04 so I—1 p p CO 04 04 04 04 04 CO co
u, -c
C<3 On to © sd I © co sd On 04 to oo
1— i ,—

PQ
1—1 04 04 co Tf to to, so SO so so O' O' O' oo
CO co co CO co CO CO co CO co co co CO CO co co

so
o 1—H
to © to, CO © SO 'tt to •o co © so © so
co p so p p 04 © os oo p so ■^f co Os
oo sd oo 1—i co to o^ 04 oo © sd 04 oo co
CO sd O' oo © t-H 04 CO On ^f © so 1-H O' 04 oo
t-H co Tf to. oo OS © © r-H 04 04 co co 'Cf Tf
04 04 04 04 04 04 04 co co m CO co CO CO co CO

o o o o o o o o o to © «o © to © to O »o o »o o
OO O' SO t/~> tj- co 04 <-< -h rj (N m co Tf tf tn to so

544
© i—l t ON •o oo C4 co i—1 NO
to oc m •O NO NO NO NO to, 04
o~ 04
to
ON
«o NO 3 NO
NO
oo
NO
© 04 Tf NO
O';
oo
O';
©
OO
04
OO
—H
1.0041 Bars (-30° C)

1 T—1 1 1 ,"H

CO NO Tt ON oc 04 co On oo
<N O' © © O', to © 04 04 On O'
’■"J ON r- r- o^ r- ON 04 •O On co ON •O
od co ON to ^4 O' co © NO 04 On to 04
co Tf »o NO NO o~ OO OC On ON © 1—1
co co co co co co co co co co CO 'Tf

to, NO 00 to ON © O' NO to rr co
0- OO O' O' ON to © 04 O' *—i CO
co ON 00 V—H 4 NO OO ON © © 1■
ON NO 4 1-H ON NO CO © o^ to 04 ON NO
•O'. NO O' oo 00 ON © © 1—1 04 co co
▼—H ’—i '—1 *■“< ’—i 04 04 04 04 04 04 04 04

oo NO ON O' © to, NO •'t OO OO NO


to, 00 1—H co to O' OO oo oo 00 O' NO to,
OO © to O' On co to, O' ON 1-H co
to. NO NO NO NO O' O' O; O' O; OO oo
t-H r-H ▼-H T-H 1—1 ,—i —* T-H ,-4 t-H 1—i
0.6417 Bars (-40° C)

On to, 1-H CO oo ▼■H


© co 00 O'
'tt oo oo 04 NO NO © to O'
to O; to co CO co NO ON CO O' 04
co ON 4 © NO 04 od 4 © NO co ON NO
CO co 4 to, to. NO NO O' 00 00 ON ON ©
CO CO CO co co CO CO CO co co co co Tt“

© oo OO OO OO CO © O' NO oo © to OO
—H ON ''t oo CO 1—1 04 oo © oo NO NO
ON ''t ON 04 to O' OO oo Oh oo oo O; NO
1—1 co 4 NO 4 od ON © 1-4 04 co 4 •o
''St to NO O' 00 ON © 04 CO ■'t to NO O'
04 04 04 04 04 04 CO co CO co CO CO CO

1-H O' OO CO rr © i—i On CO co © CO CO


oo i—i ''t O' ON 1—«1 04 04 co CO CO 04 1-^
ON 04 4 no oo T—H CO to O' ON 1-H co to
to NO NO NO NO O; O; O'; O; 00 oo oo
—H ,-4 1—H 1-4 i—i i-4 i—i i—i 1—H
0.3915 Bars (-50°C)

O' to to, O' ON i-h 1—1 NO ''t co © 04 O'


ON oo O' NO to to 04 © O' CO O' On
OO 04 O' CO © oo O' O' 00 ON 04 to On
OO 4 ON to i-4 NO 04 oo 4 © o^ CO ON
04 co co ''t to to NO NO O' oo 00 On ON
co CO co CO co CO co CO CO CO CO CO co

O' oo © NO ''t © NO 04 to •o CO 1-H co


r" oo O' ''t co »o © —* O' O' © <o O'
NO © co to NO NO NO to co co 04 o O;
-4 © od NO 4 04 © od NO NO 4 (N On
© 04 CO •o O' ON i—i 04 ''t NO OC ON
''t ''t to »o to to •O to

o o o o ©©©©to©»o©to©tO©©©©
T)- M
I

545
vo QC co t-H co ON t—1 00 0 ON ■rr NO VO
Tf 04 co co Tf co co 1—1 O OO NO
y/~, O' 04 NO 00 O 04 Tf NO r- ON
»0 vo NO NO 3 NO NO r- 0 0 I"; O; r-
T-P T-P t-H t-h T-H t-P T'H t— < »—< t-h T-P *—

v
o
o T—H 1—1 T—H .— 04 OO
o~ NO NO VO 04 oT
>0 co co 3 CO ^t O T—H 0 on NO OO
r-~
^r ON ON tt 0 NO co O o~ VO Tf co
(/5
Ih -C T-P oP ^f O rP of rP of 0 oP of r—«
«o NO O' o~ QC ON O T-H T—1 04 CO
cd vo 8
co co CO co co co co CO Tf Tf oT
PQ

NO
QC
O
04 00 0 T—1 0 VO vo 0 0 or co NO vo
CO ON 04 04 oT 04 vo 04 ch ON O «o r-
OO NO VO 0 or NO »o 0 or OT or T—1 O'
CO 00 'Ct 0 VO 0 •0 ON CO O' ' or
£
VO 00 0 co NO QC ,-P co «o 06 0 CO of
VO VO NO NO NO NO O' O' 0 0 00 00 00

0 04 r- NO OO NO ON OO 04 co O
0 CO VO r-~ QC ON ON ON ON 00 O' VO co
NO 00 O 04 NO QC O 04 ■*fr NO OO O
VO vo NO NO 3 NO NO O; r- O'; O'; 00
,-P T-P ,-P ^P ^H T-P T-P T-P T-P »-P ^-P T-P

co 0 ON vo «o 04 00 •0 or •0 O' 0
■O CO 04 T—1 NO NO 04 co T-—1 vo VO r—l or
NO
<L>
’—1 co NO ON 04 NO NO 04 00 •O co ▼“H
C/5
13 -c
t-H
0^ CO ON vo 04 QC •O ^P OO "Vt ^P oc vo
C o3 vo QC ON 1—1 t-H 04
»o NO t^ ON O
+-> PQ co CO CO CO co CO co co CO or or or or
c C4
o
o ON
or vo 00 O' t-H
0 ON
A c4 NO O' vo ON m O' ON ON Os or O r- O
LO or 00 CO O (N 00 O ON VO On t-H
0 ON
■ £ CO NO vo O CO t"; O t ; ON
NO 0
< NO 0 of rP T-H
of QC t-H of QC T—I of rP
QC On ON ^H H—1
00 00 O O O O
a> t-H T-H t-H t-H t-H T—H
A
(0

t" O vo VO) ON OO CO co 0 CO 04 OO 0
NO 04 or vo NO r- r" r- NO VO co 04
05 NO t-H co VO r- ON T-H co vo O' On T—
VO NO NO NO NO NO r^; r-; t"; r- QC
,-P T-l ^P ^P tP T-H ,-P tP T-P -P ^P
U
o
O
<N
04 00 O' NO 00 vr> CO T—1 ON NO t-H
co
OO 04 CO 0 or co 00 ON 00 NO
C/5 NO NO O'; 00 0 04 •ON ON CO OO or OO
u,
cd 04 00 of 0 Op co On VON 04 00 VON (N OO
or or >0 NO NO O' O' OO ON ON 0 t-H tH
PQ co co co CO CO CO co CO co or or
co
ON
o
>o
r' t-H On t-H co
or
00
00

p
or
0
04
or
04
CN
04
CO
p
t-H
0
or
QC
3
NO
r-'
T-H
or
NO
NO
H—1
NO
On
OO
00
O
NO
od of ON of ON of o\ CO 00 co QC c4 rP
0 t-H h-H 04 04 CO m or or vo VO NO NO
t-H t-H t-H T—H t-H t-H T—H T—H t-H T—H t-H t-H T-H

<u
u-
3
■*->
ci
u.
<U
cx
IG OOOOOvoO»0
04 ' 04 04 CO CO
o
r}-
»o
tJ-
O
vo
«o
vo
o
NO
o
O'
o
00
o o
— 04
H ^

546
CO CO i—h O 1—H VO o© VO i—i 1—1 oo
CO CO CO CO CO
uo
O'
vO
VO
8
*0
04

VO
CO vo r- on
vq vq vq vq
04
1—H
O
CO
O;
oo

O;
VO
VO
O;
OO
O;
1—i 4 f—H i—h 1—1 H ▼-H 1—1 i—i T—H 1—i
7.4490 Bars (30°C)

VO 04 VO VO *40 ON ON vo co O' 04 o
VO o oo 04 VO O' i—i O' vo r- 04
vo ON 04 CO VO vq r- oo o *“* CO vo OO
CO o OO vo 04 ON vd cd 4 oo vd 04 On
VO o~ O' oo ON ON o i—i 04 04 co 3:
CO co co co co CO Tf 4 tJ-

04 vo VO Xf o © OO co vo co co vo
OO ON oo 04 CO 5\ vo 1—1 vo vo O CO vO
o vo vo ON oo CO vo O' vo 04 OO 04 vo
vo 00 CO vo O' oo oON 1 04 04
cd 4 vd 4 od ON o —i cd 4 vd vd 4
04 04 04 04 04 04 co co co CO CO CO CO

i—i O' vo vo 1-H o vo vo o 04 O vo


VO o 04 vo vo VO vo CO On vo
Ot- O' ON co vo ON i—i co vo VO oo
vo vo vo
vq vq vq vq vq «- O; O'; O; o~
i—i =H i—i i 4 4 4 4 i—i 4 i—i 1—I —1
5.6729 Bars (20°C)

On vo 1-1 O ■*fr ON oo oo O' 00 vo co


04 O' vo On T—1 CO O' co 04 '3- O O
O' O'; O' VO vq vo p Tf p 4 vo VO O'
ON vd cd O 4 4 1—i od vd 04 ON vd cd
VO VO O' OO oo ON o O i—< 04 04 co
CO co CO co CO co 'O'

04 O' o O' vo ON OO vo co iH ON
O vo ON CO 04 4 vo 04 O' vo oo OO vo
00 vo vo o o vO ON O oo vo o ■*fr O'
O' ■or O vo 1—1 vo ON p O' vo oo
o 04 4 vd 4 od ON 04 4 vd vd od
co co CO CO co CO co rr 'Cf

i—i vo ^-1 oo vo vo 1—! 04 ON co CO On


o co vo O' £ ON ON ON oo VO vo co O
vo O' On i—i co VO O' On 1—H co vo O' ON
vo vo vo vq vq vq vq vq O'; O'; O'; O;
4 1—i i i_i i i H 4 i—i i—i
4.2330 Bars (10°C)

vO vo Tf vo vo oo ON vo CO On 04 CO
OO CO vo vO oo 04 ON P vo 4 O' VO ON
vO ■'t 1—1 00 vo CO o CO O' VO vo vo vo
vd 04 ON vd 04 ON vd 04 ON vd cd O 4
vo vo VO O' OO OO ON o o i—i 04 CO co
co CO co co co co co "'t

O' ON © CO CO i-i oo 1-1 "3- 04 vo vo


co O' Tf oo co OO ON rr vo O' 04 co 04
y—t co oo vo ON VO ON © ON O' CO oo
ON © © © © OO O'; VO 'Cf p ON O';
© cd vd 4 ON © 04 4 vd od On 1—H cd
tJ- vo vo vo vo vo vo vo vo

©©vo©vo©vo©vo©©© o o o o o
*—1 04 CO Tf VO

547
NO tO m T—H »o o ON ON i—i ON 04 04
Tf O 0" 00 On ON OO NO »o 04 O 0"
CO NO oo o 04 NO oo o ON ''t NO r-
to io IO NO NO 3 NO NO
T—H 1—i ,-H ,—i ,—i i—i ^H ^■H 1-H ▼-H —I —i

On ''t co OO oo o i—i OO 04 'SO 04 OO 1—H


O O' 04 o 1—1 1—H 04 O' O o © ON ©
-s: 04 ON co ■'t co OO 'f co OO ION
tn —i —i ON Tf 04 ON
1— cd © oo NO d
r- OO ON On o 1-H 04 04 co 'O’ io> »o> NO
a
co co co co ''t 't "t Tf •*t
CQ
04
ON
m,
04
(N rr On 04 m © ''t •o 04 — to © ©
td NO O' 04 Tf ■o © «o "*t On 04 NO
£
V-H

*—<
r-H*
O'
ON
—i
to
O'
04
OO

cd
NO
i—1
't
V—<

oo
It
i
to
Tf
©
NO
co
NO
NO
©
04
O^
O-
O;
O^
04
CO
OO
O'
00
00
i—H 1-H ^H T—H 1-H 1—1 1—H 1-H 1—H 1-H ▼-H 1—H

O' O' 04 ^t O' 04 1-H co © CO 1-H tn NO


O' 04 © OO ON © © ON oo NO 1-H OO
co
to
NO
to
OO
tO
© 04
NO
LO O'
NO
OO © 04
O;
5 NO O'
NO NO NO O; O; O'; O';
—- ▼—H t—i i—i 1-H 1—H 1—i 1—H 1-H ^H ▼—H —H

NO CO 1—H ON NO in', 04 ON © OO © CO co
On © co O' © to to 1-H O' V—H O' CO 1-H
CO NO to 04 On ^t ON "t OO m O' 04 O';
© OO NO "t 1 ON NO ^t 1-H On NO ’’t —i
O' O' OO ON © © 1-H 04 co co Tf to NO
co CO co CO "'t '’t ''t "*t tj- Tf '’t

NO On NO ON 04 oo NO 04 ''t NO co 04
NO NO 04 NO ON to 1-H © NO oo 1-H
Tf ON ON NO T-H ^t IO 4 04 ON NO
i”1 © OO NO tF 04 ON NO CO © NO CO
to NO NO 0^ oo On ON © —I 04 04 cd
1-H *—H —i 1-H 1-H 1-H —i 04 04 04 04 04

NO ON © © 04 NO IO O' «o On OO co IO,
o OO © rH 1-H 1-H © ON r- IO co ©
to NO oo co to r' On 04 NO oo
©
»o to to NO NO NO NO O; O; r-
— 1— 1-H ^H —i 1—H —i 1-H —H —i —H —H

U
o
O NO to NO On to 04 04 —
VC co IO
'■'t © O' 04 ON 04 04 1-H
© © to ^H

C/D
On 'Cf ON 04 NO On 04 •O OO It OO
Uh o^ TT 04 ON 'Nf ON NO cd ^H oo to
03 NO O' 00 OO ON © ^■H 1-H 04 co 'f IO
PQ co co CO CO co 'Cf 4 Tf ■'t
to
NO
o
NO NO 04 1—1 ON ON On ^H
04 O' co NO OO NO
o^ o O' 1-H © 00 co 04 OO O' © ON ON
0" © O' OO ''t oo ON O' ^t NO OO
co co co co 04 1-H © ON oo NO 04
oo ON © 1—i 04 cd 't td id NO O' OO On
1-H 04 04 04 04 04 04 04 04 04 04 04

<u
3
*-*
o3
i-
<D
O- © »o> © »/■-> © © ©> o © o o o o o o o
t ^ in 'C h oo ON © ON CO Tfr IO> NO r- oo

548
r-' ir, o ^ ’t 't ^ ^ On nO c*l
Tft^©rNTrNOOO©<Nm»/~>r'-'
»n»r)'>C'>o^C5D5or^t^t^t^r'
28.4071 Bars (91°C)

ONONoNooNOoomONocNOON’—
rtoo'sooo^oooomo’—1 r~- ^
Q5Dri,t(NOO^',t5D^tv;00
ododoo^vd'trn^asr^'on
CnnTfTtTj-'t^TfTtTt'ltTf

m oo 1 O »/r~> O oo <N
m oo tj- r- m no r- NO
N- r- On nO to oo r-~ o m no cn
m VO w 50 50 Tl e'¬
oo ^ 05 TT oo rJ 50 o en r- o m
tO 56 t-^ od oo on on on o o

nx5oniriV55oa\50oo no </n
l/NTtO^’50fvf''OTlffl OO m
04^x0(^^50000^ Nt VN
m'mn'O'O'OiO'Ohhhhh
23.0460 Bars (80°C)

r- ci ri h x 't h <N C4 <n r-- t-~


hTfvO'-^XXhXmtfNa
i'-mooooTfr^o\0000\ooNq
5dh5dv3TfoidO\hV3(SOX
r-oooNO’-^r^mmTti^NNOt^-r~
ri l^i r; Tf Tj Tt Tf' Tf Nf Tf Nf Nf

m n r- r-~ © 540 ^ _
Tj-in^Hf^Tj-mooor~'NC|ooooo
>—‘TfroO’—1000’—
nX(NO^OOON^t^(N
X'nnhfnMNh^'oafOh
vdhodxo^ddd^^^MN

r^-or^-r-^imoNv-icNmom
©ooeN*or--ooo©r--NOTl-<NONNO
m vn oo ©
- <N
- - tT
• NO oo O <N N" r-
«rNio»oNDNONONDNor^t^r^r^'r^
18.8578 Bars (70°C)

NOoo©©^rnr-~©ooo\t-'cnNO
C4 M 05 h O NO O 1^ ■’T 05 NO
1 ON 50 ^ 1-1
*OTtrn<NOodND'^:<NONr--*‘om
hX0\O’H’H(N(f)^'1’^N0h
r^iror^.N-N-N-N-N-N-N-N-N-N-

00 ON
ON m On O tr-, iy^i »o O B' I' O C l rfr
Tf On loOOoo^'^t’—'trroO’—1 no
0 4 ’“_l ^HVN't0\(riTt,^r40\'0'H
I"- to ONQO'^-ONU^O^O^tON''^'
OO ON © o <n m m Tt Tf ^ »/~>

o o o o o OOOOOOOOOOO
t— X On © 1 (Nm-rf^'or^xONO'—'<n
,—i.—i ,—< i-H i—i »—i i-H i— M (N M

549
-c

£
Table A-5b (continued)

Copyright © by E. I. DuPont DeNemours & Company, Inc. Abstracted by permission.

O OO OO VO © X Tt o ON ^ OO Os t'~-
»o vo ^o ^ m o r- n-
o r- o tN ‘ VO 00 O <N tT IT, t"
\r, >Ti iti vO vC vC vD vO h h h t'- h

U
o
O
o
o^'^-'or-'^O’-^oor-r^r-ONoo
</) r^voor^<Nmm^^—
Uh Ooo*or-m^tr4oomvooN^m
cd
TfONrHHFHOONh'OtrJHON
PQ h x o h tNmm^t»ovor^QOoo
vo rnriTtTtTfTtTt^tTtTtTtrtTj-
o
''t
o\ © o *oTfoor--mr~«omTt
m h X (N 5 oo <N «o X O h
r^oovrNr^^tm'^ O C4 cn vO <N
_ -- -
oovr^'^ooor^mr- - . o ^ ^ o a
o\ t— m
~ oo <n
- - vo — m
- o — r" O m vo oo
ro Tfr u^u^vovot'-t'^r-'-oooooooo

ooooooooooooo

550
oc oooc^ooh— x oo ri
ci <n ivs <n -d- oc vo d
^hMOIN^hf^
0\oot^h0'cjd,d;
O
O- ITi IT) ICl IC) IC) 1C) VI
cd
>
Ofl oc
«1 OO rf O h O in, (N
(N r-~ in ■— r i c i
vor--oo<Ninovc'dOs
a. <0t"'»n'd;<N®0\t~';
cd
> iv", in in n \C !n, d' d
D.
O W
u,
■>->
C
W

X m oo Tt I—* rn
d- d x M 't ®
APPENDIX 6 _
AMMONIA TABLES*

rs O On O t vo in m
'3 O <N CN <n "d IV") X
cr o o o o o o o

oc
-c
O ON X O O *n O ON
Of'icifndvdf'ix
o O o - m <n m m
a fnmd-’t'd’dd'd’
cd
03 >
"c
o oc
E oo

E -ST O VO ON 00 <N <N vO


< >. a,
O mi C mi C d
ONr^'Od’cn^ONW
C

_a cd
mrvnmrvnmr^cNCN
TD
(0 a> cd >
to -t—< x: W
i CO
< i_
u
c
W
■«—'
a>
CO oc
n CO -c
m h M oo m d; r)
to
rs (N Tt r- cK ni in oo
’3 Ci rj- vo oc *— m »n
</) a-
0)

r
<D
Q.
O oc
00 ?> ro s® oo oo in On
<N vO ro m oo oo vo
iv", VO t" <N o ^t
o iv") (N ON r- \0 m d; m
CL.
cd © o © © ©
<u >
E
"o
> Os <N VO o t oi d- o
o O © m <n
3 ^
d ^
d d m iv" , iv" , iv" ,
s
o
4)
3
0, cr
00
oooooooo

4)
OOCNVOCNcnOsCOCJN
<Z) hMmivCMDOCOC
X) £ Oh h< m o\ © vd © d-
< <D -*! h 0\ H IV", Os cn 0\ m>
— a rj rr-,

4)

cd
<u U

a
E
a>
O in © »n O m O in
H Tt co rn CJ *—' i—1

551
re <3 OO O r^) oo o O (3 Os O
O X re 3 t^i 00 3 o r- 3 re
3 r-~ ID ON re OC re r- <3 r-
o re (3 <3 — O p Os p oc OO p
a, *D <D id] ID ID 3 3 3 3" 3-
Cd
>
00
X
(3 3 ID X — X re o m X <3
«D, r-~ <3 Q h-h 3 O oo r- OC
Cl (3 r- re 0\ ID, i— OC 3 i—* oo «D,
Cd X 3 re — O p p X to re DC
> re re re re re .2
a 3" 3 3 3- 3-
'c
o w o
u.
E
C
W E
<
T-H oc re 3 <3 <3 r- O Os m oc
-a to oc re 3 3 re <3 Os r-
1—3 os oc X 3 <3 o OC >D, m —-1 C/D
5 r- oc Os © <3 <3 re p to <D
U*
© © © © •—1 —1
j (U
a.
o

o
3 X <3 X X oc p 3 cn ID) p
E
o 3- 00 <3 ID, OC X 3 3 3
a.
ID ID, X X X r- r- r- c
3 3 3 3 3 3 3 3
cd
"O
CO > o
'c E
o 00
i—
a>
x
"D
E X
H
CD E -c
<3 -H >D (3 3 Os OC ON p oc
3 < >>
. re idd X X
<3 o
X
oo
X
X
ID rn
<N O
r- ri
>D, <3
C o. X 3 3 r-~
"O O. 03 (3 (3 <3 (3 y—i h—i t-h i—i i—' o o 3
» > r— i—< h-« h— y—> *—< —*
Q) os
c -t—> x w
o C3
c
_cd
o 3
W O
L.
CO CO oc
CD CO -c u
(N <D P 3 C3 p ID ON oc p T3
L-
< o 3
r^i 3- X ID oo X i—3 X X
OS (3
3
Q> OO o <3 ID r- os <N 3 r- T3
CD a" *—- <3 (3 (3 C'J <3 re m re 3 C
A CD 3
CO
■HH
GO
CD
Q.
O oc 3
00 ?> cS
X «D re X oo re re X oo re X ID <v
OS re ID 3 OS oc o ID re DJ re L-
o oo 3 o t- 3 (3 —+ OS oo r- X 3
a. (3 (3 <3 p o o p O CQ
os o O o o o O o o O o o
<u >
E c
3 _o
X
O 3
> Z
?> IDre OS os Os o (3 X o C"
o ^-' X oo o re ID oo o <3 id r-
eg IDID X X X X X r- D- r- E
rs 1—- »—< 1—H 1—H — 1—3 1—H T—1 o
o
<u '3 O o o o o o o O O o o
X O" o o o o o o p O O p o
GO o O o o o o o O O o o "3
X <U
>
-n
<u
"3
<u
1-c _ 3
OS ID «D i—> ID 1—
C/3
3 P p '—[ p ID X o p p X 3
X>
. £ Oh
OS ID 3" OO X (3 X © X X ri £
< <D X i—i
(3 (3 ID O X <D 3 OO re
3 ID X r- OO O re ID r- X
© X
•—1 1—1 *—< 1 <3

"3
<U C
J-H oS
D cd
-*—>
ctf d—s X
I
»- CJ
<D ^
D- 2-
<
C/D
E _3
<u
O »D O >D O >D o >D O «D o X
H 3
M N m n t(- xf in
H

552
in m3/kg, h in kJ/kg, s in kJ/kg-K)

04 C4 or It co CC| 00 NO r- r~ 04 00 04 o ON wn ©
00 r- 04 NO, NO ON O V—< t/p ON NO 00 r- co WN ■t ©
i—• co ION OO ON r- NO 04 ON NO 04 co oo © ON wo
o to nO r-~ to o ■'t 04 00 cn ON r- 04 00 — co *-« ON © © o 04 P
•*t —i — NO © NO
tO
NO © f-H
tO
to © tO tO ©
co
»/P © co
04
to © © tO © oo
ON
w-~i © Tf t-
NO •^r 00
to •O tO to to t/P ^r ,'T
< i— « 1-^

© © r~ ON Tt o NO r- ON to © i—i 04 ON
<D r- © © r4 ON O1 TT 1— ro 00 04 1—
Lm i—« TT ro i—i CO CO ON
© r4 t" ©
O
■*—*
a
o
cc>
p
— ON
CO r~ fN © m
00
00 m
(N
nO
to
04 »o
ON
Tf
to
— ON
to
<N
to
NO ©
'J-,
Urn © © NO © O O © o
m m (N <N O ON 00
<D to »o
Cl
Table A-6b

E<u
H r- OO © to o r- — ON O' to
CCl ON 00 1— 04 i-^ CO »o 04
ON ON CCl 04 NO n 04
$
Properties of Superheated Ammonia

C4 © 00 ON o CO 04 to r4 04 CO CO
i-H OO © © wi © to to o 00 to © CO »o to
1—^ 1—1 o ON 00 NO
to to to Tf
' 1 1 1 ■ 1 ' 1

04 t" r4 NO 04 co ON
r- oo Tf ON CO o> to
NO * r~ »o 04 r-
CO CO <N © CO 00 co 04 or
.
NO © © o< to © to © CO to
ON 00 00 O'
Tf

r)
l~-
-rt
©
It
r-~
C4
O' T ON
t—
CO
04 04 oo
O
CO On — 3 r~ 04 to co CO
- Nf \C © r- to loi
r~- to
rf 2

NO — On NO
no On On On
NO W~l — oo
C4 w~> © NO © NO
—• r4 no © co wo
m
it 3

co co
04 co
o — r-
04 04 NO ON
o© »o
04

U
o
a
Oh <D
u.
D
a> -*—*
cd
U>
Z3
<u -c £ -c _C ^ $> .c w £ -r £ -c £ -C 05 ^ -C «o
C/D
C/D
<u
a-
E I s r~- g P g P ON NO g © g P
C4
oo ® ^ ^ 00
04 00 00 On r4 P

^ H
a>
00 ON *- ON Tf
04
— ©CCl
—« ©
CO

I *
< 00

553
5.8650
0.1108
1897.0
©4

r-~

0.1054
On ON r--

5.7521
©4 00 r~ o
O NO) NO
oo

1845.2
oo —« ON On o — O oc
d NO) NO) © 0 On NO)
>Ol NO, ■Nf
OO 00 oo

—H r-

0.0999

5.6368
sO (O o m On
00 c©

to
N- m
3
fN
r-
to
•'t
On
o
8

1793.5
nC to ©^ ON p 00 so p p NO NO
d NO O) >©) d to to d T—^ to d >©’
o O o ON
00 oo 00 00 r-

^—1 m 00

5.5140
0.0943
O' oo r- C" o TT IT) to >©) 1—4 to
r-
NO oo 00 co 3 00 ON »o ©4 8 r- r- N—- On Os to m
o NO O Os NO to 3 os ON »o os m <o o r-
—1 8 1

1741.6
■Nt NO (Nj IO © CO so <N N so oo o »—• N; p ON p N" to
© © NO d oo NO O fO sd d ON sd o d d d to d to to d o to d sd »o
oo r- r- NO sO NO to to
r- r- r- r- r- ©- r- r-
■ 1 ,™-

00 m

5.3826
0.0885
NO TJ- m r- On Os r- o r- o NO ON
m ©4 ©4
1-5
c© c© r- <o o Os N" — fN fN OS r-
o c© on Tf i c© 00 00 r- to, r- fN 00 —
©4 3 ©4 00 00 1

1688.6
NO 00 ro (N (O o SO t l c© SO p to —" NO >Oi
d d no d NO O sd sd o d to d to d d d to d »o to d On no) d d ^
c© m fN ©4 »—H —H O On On
r- r- ©- r- r- NO NO
»—<

c© ©4 00 m

0.0825

5.2403
ON ©4 On NO Os NO o os On to r- oo c© NO NO O
c© r~ ON c© r- Ol >o so ©4 ©4 m i—i i m l©) NO) ©4 NO,
On r- NO Os —* sO r- ■'Tf fN to O r- ON O
8 3 no, ©-- ©4 3 ©4 3 ©4 1 O ON CO)

1634.2
r- p (N OS NO IT) to OS p i—' c© ro
d g NO d d no o N- NO o oo to d ©4 »o d NO to d d to d (O to O
d d no' d d no
00 00 r- »o
NO NO SO
r-
SO NO
NO
NO
NO
NO
' 1
so
' 1
3

m 00

5.1646
On

0.0793
m ©4 ON fN r- ON N- to r- to to NO ON ©4
r- 2 8 >©) c© »0 CO r- fN ©4 oo 00 to 00 fN fN — ©4 ON o
nO NO r- oo m 00 to OS SO r- m r- i— 00 o O oo CO)
o r- ©f r-~ 8 00 ©4 00 o

1605.9
00 — Tf NO) o ©4 N- SO fN ■^ so SO to r- p no c© 00 ©4
ON
^ ; no d OO NO NO) NO d sd d to to o 00 to d to d d to d to o o to d NO)
r- NO to c©
NO NO 3 SO
sO so 3 3 NO
fN
SO
fN
SO 3

tj- c© c© ri

0.0760

5.0816
NO <o to N- r- so Os fN so r~ co NO ©4
©- On c© c© NO r- fN OS oo SO O fN to (O m 00 r- sO NO)
— On NO1 NO, NO Tt fN Os 00 8 m SO o (O o ON fN 00 NO)
o r-~ — no 1 o

1576.5
OO NO ff) NO, so Os fN fN ON SO NO to On p o fN fN
oo
— ON NO © NO NO O M NO © 00 to d to d d to d NO to d 00 to d T-H to d (O »0 NO) NO)
m ©4 »—«
3 3 3 so SO NO SO
O
NO
o
so
ON OO
NO)
' 1

o ■nJ- On (N 00 4.9934
0.0725

© NO) o o OS fN oo ON O NO)
oo NO OO 00 Ch ©4 Os (O OS 00 Os os r- fN 8 8 00 c© ©I NO
o 8 c©
NO
©4
00 ©4
00
©4
Tf OO
rn
m SO y-* Os »o •o m fN
00
fN o m On 3 OO
o
1545.7

NO) NO) O § On fN »©) to fO cn p fN O ON


r-
r-~ NO d d no d On NO d to d »o d to d to fN »o rn to
©4
NO
©4
NO 3 so
o
sO »o
Os
to
d
00
to
d

»o
d
3 NO)
NO)
NO)
NO)

c© c© 00
4.8962

00 o oo
0.0689

o O fN fN OS OS <o NO r- — NO)
oo ©4 c© NO) to fN so T-H to m to ON r- 00 rt
r~ O ©4 c© * so S OS 00 to to fN 3 00 tO r~
o NO 00 On ©4 00 00 3 m 1 1 -<r
1513.0

>©) •O) O CC) <N fO SO fN o m o o ON


NO
NO, NO © © NO NO d (N to d (O to d »o d to to d to to d to to d tO tO odd
o OS 00 r- SO to m ©4
NO »o to to to to to NO)
r 1

00 m ©4
0.0649

4.7863

r- fO o fN so so (O »o fN On —
©4 N- m N- 00 to Tf oo Os SO i©, c© ©4
0© Os to r- OO Tf r- y-* to 8 to OO ©~ r-~
o CO 3 1 5
1477.8

m fN to »o so fN so *—■ o o p o O oo oo
1©I
o Os to d On to d OS to d 00 to d to d sd to d c©
NO NO) ■NI¬ c© ©4 o ® 8 *
NO) NO) NO, NO) NO) NO) NO)

$ -c -c no $> .e $> no
v

£ -C ~C NO
2000 h
(49.38) s

©> -c -C CO

OJ c© oo I* g $
oo ^
^o ' tO

554
CX ^
to © to © to o to ©
E U M iri h O N >ri r-~ ©
H> o ^ rH p- M (N (N 04 re
H w

o <u £ ©
02a o\ O' —* ON OO
o § on on 04 © no Tf
<D " ^ tJ- O' O'; to re ; 04 NO
a o ^ o^ no »o oo od Tt r4 —: ©
^ > E o r i d- -
Tf *—<

—i o o r-~ ^ oo o *— O'
o ^ On oo *—' oo O Tt O 04 re
a % J? re on no 04 o r- to re ’—i
oo no no NO NO
w ~ © © © © © © © © ©

*
i
Ofl
a OO in 00d1 KN M f^l to
cx ^ to to © NOO Tf 04 © 04
cd « ON tJ- O reNO ON ON NO —*
> 03 h- r' r- »Cno no io ir> SO
O- W w o © © © © © © © ©
o

X +-»

W
c

^ 3 ^
c4
re
Tf
04
on
to
to
o
no
re
to
ON
re
oo
no
no
to
r—
no
o4
oo
O'
on
oo
04

©
© —

o
.2" w o © p p © © o
l-J © © o o © © © © ©

oc
42 r- re oo re to »o to 04
Uh 1 © © ON oo © NO tO)
cd O X ef o^ © 04 to ON — ef-'
on © v—H T—1 04 04 04 04 re re
cd re re re re re re re re
>

a. oc

cx ^
© re NO
re oo re
04 to rf —
©
ef
to

cl
Tf ON Tfr
ja cd £ no no >n >o nj- rt Tf re
> -c On ON On ON ON ON 0s, ON ON
pC W w c i 04 04 r j 04 04 04 04 04
c

< W

."E
ti '3 ^
r5 cr -s
C/J • r*
hJ
O' © 04 re re 04

rncONtcoit O^ ©
' Tj" OO
oo m t— *— to on no NO ON

-h ^ M m M n n re tj-

c/d
C/D re
<D c x c d-
l-H C/D re M O' oo 04 NO O
re On re On On Tf 04
cd 04 -rf On OO NO
C/D
X> O o * o '—I re
X) o © © ©
<

CX to O »0 © NO) O to ©
E U
<L) o
M in o O 04 >n O' ©
—I —I t-h 04 04 04 04 re
H

555
d ^
m © m O © © © © © © ©
E U ci m c~ o © in © «n © in
o re re re Tf Tt m *n © © r- c~
H w

C* © ©
o Tf r- oo r- m a—i ©
lC m 1 nO Cl © Tt r- r- r-
NO Cl «n ON £ Tf >Ci © r- in
'5 Is i OO ON re © re Cl ©
<u Cl i—i 1—1 © © © © © ©
a. © ©
GO > E d d d d © © © © ©
Li

Compiled from “Thermodynamics Properties of Mercury Vapor,’’ Lucian A. Sheldon, General Electric Company.
L- o Cl NO © © © «n Cl
o r- Cl oo m re in ON in re Cl re
Ci. &C On oo NO m re i—i ON oo r- © in
GO cS ^ in m m m in «n Tf Tf Tt Tl- Tf
> d d d d © © © © © © ©

X
.
<N r- oo re © Cl ON ON © oo
Oh ^ O ON o re Cl m <—< ON © Tt oo
£ £ ON NO m re © c~ m Cl a—a ON r~
> !>5 ti- Tf tt Tf Tt re re re re Cl Cl
Cl w w d d d d © © © © © © ©
o
Li
*->
C
W
T3 ONre ti- re >n © ON Cl in
NOCl r- Cl © c~ m Cl oo 3
o 1—1 —-I Cl re Tf Tf m © © c-
cr d1 T”1 T-^ J
GO
J W d d d d © © © © © © ©

'Zq
s: oo tt r- a—a oo OO Tf © Cl re
+-> Li re Cl o S? in © in © Ti¬ oo 1-*
cl o d d re in d d ci OO re oo d
go a. re Ti¬ Ti¬ Ti¬ in in \o © c- c- oo
c3 re re re re re re re re re re re
>
N—^
_oc
>—>

c- o re NO Cl oo re ON in c-
ja d o NO i—i NO r- r- oo OO ON © ©
CO £ re ci ci d © ON oo d NO NO in
> -C ON ON ON ON ON OO oo OO OO oo oo
„c W — Cl Cl Cl Cl Cl Cl Cl Cl Cl d Cl
c
W

rn Tt Tl- Tf ON © in m ©
'3 ^ re NO ON Cl re r- i—i »n oo ©
cd cr -c
GO d d d d © d re © NO ci ON
tJ- tt in m no NO c~ oo OO ON ON
J w

c/3
C/3
<D ^
Li C/3
ci Cl
^ 55 m ON ON r- © m
»n OO re o re T—H re in Cl re
c/3
X) d d ci d oo d re in ci d
< Cl re »n c

d ^ in o in © © © © ©
E O Cl m C~ m in © in in
<U o re re re d Tt «n in NO NO r- r-
H w

556
2 Tf uo t—' r-" os Os} T—1 ON to co co 04 04
I tn co »/~> tj- © X t3- 04 to to on o
o
23 E
M M - - (N 04 Tf 0- CO 04 04 O 04
o> O © O O O © T—1 © © © o o
X TD
C £ O © O © O © © O © © © o ©
H o
U

VS vs vs vs »r> vs vs vs vs vs vs vs vs
*-> 1 1 1 1 1 1 1 1 1 1 1 1
pH C/51
_APPENDIX 8_

O o o O O O o O o
PROPERTIES OF GASES'


t—H o o © ©
C/5 r—l i—< *■“( ^—1 V—1 <
o _E X X X X X X X X X X X X X
o
C/S r- ON 00 1 CO 04 CO C4 CO ON to CO
* ” 44 oo X 1-H to ON to, X t" ON co ON
> ON 04 Tfr t—h OO 04 oo © 04 ©
r—H *—1

a o tON X 04 00 X oo On 00 00 00 O' co
O o ON X ON © © to © © © ON X 04
Tt 04 X 04 Tf 04 X Tt co Tfr CO 04 co
H—1 X X X X X X X X X

00
00 ’—I Tj- IT) h o h ^ -< t" co 04 O
X

Properties are given at 15°C and at so low pressures that they are independent of pressure.
to
° 3
r''-xcoxr-'co^©xo'- X to 'cr
o o o co o ©

a
<u
PC to co co oo X co X oo CO X to
o a O 04 co co i—< co oo ' co CO
<o o to OO O X 04 ^ 04 O ON X

o 04 © © X »o Tf 04 © o ©
o>
a
oo

c O4ON0O»/OON©O40C oo oo »o
$ 3 —
60 r-"00ooooxxXTj-oox ON ON i-h
w tfl v XXOOOO\hhN-^ON to 04 X
04 Tj- 04 04 04 O ’—1 to 04 04 —< Tf
O
"" G
O o o © © © © 04 ^ © © o ©
U

u,
iS 04 co X co X 04 © O' X X
3 .60
42
O' co r—H r- © Tf CO © ON X
o On © On © © © © © © © 04 © © © ©
QJ ’S X T}- oo
oo X On OO <© X OO X
o £ 04 T—H co Tt 04 CO ▼—H 04 r—H CO 5 X H—H

o vO
X
VO
X>
E X o
<N
Ct PC o
C4
PC PC O
PC(*S
(N
o(N
OS O
>> £ <UDUXXUXU O U CO PC
oo

<D
TO
• 1—<
<D
X a>
3 o
'* a 3
o o 'x o
a> • P>H
c o 03
a
o a3 T3 E 4> a> c
c c 00 S2 D <u •3 >
o a C (2
(33 00 a>
c c
aJ
*-> o O c 1 o s— jc o c a>
oo
03
23 -t—1
a>
go E o X X « .2 © *-v l-l 03 O. ^■■ ■
X
05 ■«-< /" . O' tail
00 l-l O ~ 03
23 •is E b*
l_< ro
03 03
a> >s ,<u •- O X t-H OJ
00 <<<UUWXDCS £ O O (X CO 3:

557
_ APPENDIX 9_
PROPERTIES OF AIR

At Low Pressures Where Properties Are independent of Pressure

Density
(g/cm3) Specific Viscosity Thermal
Temperature (Std. Atm. Heat (cp) Conductivity
106)
(°C) Press.) (J/g-K) \ms / (W/m-K)

0 0.001295 1.003 17.33 0.0245


50 0.001095 1.005 19.60 0.0279
100 0.000947 1.010 21.78 0.0312
150 0.000836 1.017 23.94 0.0343
200 0.000747 1.025 25.80 0.0374
250 0.000675 1.033 27.59 0.0412
300 0.000612 1.043 29.31 0.0443
350 0.000567 1.056 31.02 0.0473
400 0.000524 1.068 32.65 0.0503
450 0.000495 1.078 33.93 0.0528
500 0.000456 1.091 35.73 0.0565
550 0.000429 1.105 37.24 0.0596
600 0.000404 1.146 38.72 0.0627

This table was calculated from various sources including the International Critical Tables,
McGraw-Hill Book Company, Inc., New York.

558
_ APPENDIX 10_
PROPERTIES OF WATER
(STANDARD
ATMOSPHERIC
PRESSURE)

Thermal
Temperature Density Specific Heat Viscosity Conductivity
(°C) (g/cm3) (J/g-K) (kg/ms x 106) (W/m-K)

0 0.9999 4.216 1793 0.562


4 1.0000 4.203 1572 0.571
10 0.9997 4.187 1310 0.582
20 0.9982 4.182 1009 0.600
30 0.9956 4.178 801 0.617
40 0.9920 4.178 648 0.634
50 0.9884 4.181 549 0.648
60 0.9832 4.184 469 0.659
70 0.9796 4.189 407 0.669
80 0.9718 4.197 357 0.675
90 0.9653 4.207 317 0.679
100 0.9584 4.214 284 0.682

This table was calculated from various sources including the International Critical Tables,
McGraw-Hill Book Company, Inc., New York.

559
APPENDIX
PROPERTIES OF
LIQUID METALS

This table is based largely on the Liquid Metals Handbook, 2nd ed. US Government Printing Office, Washington, D.C., 1952. There is considerable
Conductivity
(W/cm-K)
Thermal

© f On
On mi co ON — ON m sC ON © OO —
m, in m, in ■t oo oo
co co
i— ©
If
©
co
1—
OO r- NO IT,
— —; O
© © © © © © © o o O © © © © © o © o ©
Viscosity

o — — OO Tt © ''t © m,
m, 00
(kg/ms)

© © O' ON O' 1—1 On oo co on tj- 04 ©


© ON © O', OO LO 04 O oo tJ- 04 04
© © 04 1—1 o o ©
©
© © ©
© ©
©
© o
© o
o>
o
o
o
© o
© p © ©
© © © © © O o o © o © © © © © © © © ©
Specific

(J/g-K)

04 Of 04 co o© if
Heat

OO O' NO co On m, NO NO © O' m, in, © Tfr if OO o ON — m,


Nt IO NO o in in OO OO OO co co ro ON © © oo Nf o n m,
O' O' O' O' CO 04 04 04
© © © © © © it Tf Tt © © © © © © ©
Density
(g/cm3)

O' SO co rd — m, O' «—1 04 it © oo ©


© © 1—' ON O' © r- ON r- i—« oo ON If © m, © m. © ©
ON © it co 04 in, un co i—* oo O' O' O' © ON oo oo O'
ON ON ON o © © © © o co co co 04 © © © © © © © ©

uncertainty about much of these data but it is generally believed to be the best available.
Temp. (°C)

© © © © © © © © © © © © © © © © © © © © © ©
© © © © © © © © © IO) © © © © © © © © © © ©
it © oo Nt in © 04 It © - 04 m 04 rf © OO 04 ''f © OO
Temp. (°C)
Normal
Boiling

O' O' r- r- © co
O' co 1—1 © oo
It O' co co O' oo
Temp. (°C)
Melting

O'
© oo o- OO
1 ON oo ro oo
O' <N O' ro © ON
04 co 1

© __
■—i
Atomic
Weight

© it r—H © ON
© 04 ON © On
ON O' NO © On 04
© © © ro 04
04 04 04

JC Z3
E l-H
3 C/5 p
G
3 3 C/5
T? CJ 3
E cd IE u. CT3
■*—»
cn <u ■*—> <D ©
O O
PQ J £ Cu CO

560
_ APPENDIX 12_
THERMAL CONDUCTIVITY
OF NONMETALLIC
SOLIDS (W/mK)

Asbestos 0.158
Bricks
Common 0.69
Fireclay (Missouri) 1.16 at 200° C
1.47 at 600° C
1.64 at 1000°C
1.77 at 1400° C
Cardboard, corrugated 0.064
Concrete
Cinder 0.35
Stone 0.93
Cork 0.043
Glass 1.09
Granite 1.64
Ice (at 0°C) 2.25
Kapok 0.066
Magnesia (85%) 0.058 at 20°C
0.059 at 50° C
0.063 at 100°C
0.066 at 150°C
0.069 at 200° C
Mineral wool 0.039
Paper 0.13
Portland cement 0.29
Rubber (hard) 0.15
Sandstone 1.83
Sawdust 0.05
Wood
Oak, maple, pine (across grain) 0.15 to 0.21
Pine (parallel to grain) 0.35

Note: The values in this table are average ones. There may be
considerable variation, particularly with composition and, in
some instances, with the moisture content of a solid.

561
_ APPENDIX 13 _
THERMODYNAMIC
PROPERTIES OF AIR
AT LOW PRESSURE

The ratio of the pressures pa and pb corresponding to the temperatures Ta and


Tb, respectively, along a given isentropic is equal to the ratio of the relative
pressures pra and prb as tabulated for Ta and Tb, respectively. Thus,

'ra

s =constant Prb

Similarly,
V-a ra

K'V'b / s= constant ^ rb

T, h, u, s°,
K kJ/kg Pr kJ/kg V-r kJ/kg K

100 99.76 0.029 90 71.06 2230 1.4143


110 109.77 0.04171 78.20 1758.4 1.5098
120 119.79 0.056 52 85.34 1415.7 1.5971
130 129.81 0.074 74 92.51 1159.8 1.6773
140 139.84 0.096 81 99.67 964.2 1.7515
150 149.86 0.123 18 106.81 812.0 1.8206
160 159.87 0.154 31 113.95 691.4 1.8853
170 169.89 0.190 68 121.11 594.5 1.9461
180 179.92 0.232 79 128.28 515.6 2.0033
190 189.94 0.281 14 135.40 450.6 2.0575
200 199.96 0.3363 142.56 396.6 2.1088
210 209.97 0.3987 149.70 351.2 2.1577
220 219.99 0.4690 156.84 312.8 2.2043
230 230.01 0.5477 163.98 280.0 2.2489
240 240.03 0.6355 171.15 251.8 2.2915
250 250.05 0.7329 178.29 227.45 2.3325
260 260.09 0.8405 185.45 206.26 2.3717
270 270.12 0.9590 192.59 187.74 2.4096

562
T, h, u, s°,
K kJ/kg Pr kJ/kg V-r kJ/kg K

280 280.14 1.0889 199.78 171.45 2.4461


290 290.17 1.2311 206.92 157.07 2.4813
300 300.19 1.3860 214.09 144.32 2.5153
310 310.24 1.5546 221.27 132.96 2.5483
320 320.29 1.7375 228.45 122.81 2.5802
330 330.34 1.9352 235.65 113.70 2.6111
340 340.43 2.149 242.86 105.51 2.6412
350 350.48 2.379 250.05 98.11 2.6704
360 360.58 2.626 257.23 91.40 2.6987
370 370.67 2.892 264.47 85.31 2.7264
380 380.77 3.176 271.72 79.77 2.7534
390 390.88 3.481 278.96 74.71 2.7796
400 400.98 3.806 286.19 70.07 2.8052
410 411.12 4.153 293.45 65.83 2.8302
420 421.26 4.522 300.73 61.93 2.8547
430 431.43 4.915 308.03 58.34 2.8786
440 441.61 5.332 315.34 55.02 2.9020
450 451.83 5.775 322.66 51.96 2.9249
460 462.01 6.245 329.99 49.11 2.9473
470 472.25 6.742 337.34 46.48 2.9693
480 482.48 7.268 344.74 44.04 2.9909
490 492.74 7.824 352.11 41.76 3.0120
500 503.02 8.411 359.53 39.64 3.0328
510 513.32 9.031 366.97 37.65 3.0532
520 523.63 9.684 374.39 35.80 3.0733
530 533.98 10.372 381.88 34.07 3.0930
540 544.35 11.097 389.40 32.45 3.1124
550 554.75 11.858 396.89 30.92 3.1314
560 565.17 12.659 404.44 29.50 3.1502
570 575.57 13.500 411.98 28.15 3.1686
580 586.04 14.382 419.56 26.89 3.1868
590 596.53 15.309 427.17 25.70 3.2047
600 607.02 16.278 434.80 24.58 3.2223
610 617.53 17.297 442.43 23.51 3.2397
620 628.07 18.360 450.13 22.52 3.2569
630 638.65 19.475 457.83 21.57 3.2738
640 649.21 20.64 465.55 20.674 3.2905
650 659.84 21.86 473.32 19.828 3.3069
660 670.47 23.13 481.06 19.026 3.3232
670 681.15 24.46 488.88 18.266 3.3392
680 691.82 25.85 496.65 17.543 3.3551
690 702.52 27.29 504.51 16.857 3.3707
700 713.27 28.80 512.37 16.205 3.3861
710 724.01 30.38 520.26 15.585 3.4014
720 734.20 31.92 527.72 15.027 3.4156

563
T, h, u, 5°.

K kJ/kg Pr kJ/kg V-r kJ/kg K

730 745.62 33.72 536.12 14.434 3.4314


740 756.44 35.50 544.05 13.900 3.4461
750 767.30 37.35 552.05 13.391 3.4607
760 778.21 39.27 560.08 12.905 3.4751
770 789.10 41.27 568.10 12.440 3.4894
780 800.03 43.35 576.15 11.998 3.5035
790 810.98 45.51 584.22 11.575 3.5174
800 821.94 47.75 592.34 11.172 3.5312
810 832.96 50.08 600.46 10.785 3.5449
820 843.97 52.49 608.62 10.416 3.5584
830 855.01 55.00 616.79 10.062 3.5718
840 866.09 57.60 624.97 9.724 3.5850
850 877.16 60.29 633.21 9.400 3.5981
860 888.28 63.09 641.44 9.090 3.6111
870 899.42 65.98 649.70 8.792 3.6240
880 910.56 68.98 658.00 8.507 3.6367
890 921.75 72.08 666.31 8.233 3.6493
900 932.94 75.29 674.63 7.971 3.6619
910 944.15 78.61 682.98 7.718 3.6743
920 955.38 82.05 691.33 7.476 3.6865
930 966.64 85.60 699.73 7.244 3.6987
940 977.92 89.28 708.13 7.020 3.7108
950 989.22 93.08 716.57 6.805 3.7227
960 1000.53 97.00 725.01 6.599 3.7346
970 1011.88 101.06 733.48 6.400 3.7463
980 1023.25 105.24 741.99 6.209 3.7580
990 1034.63 109.57 750.48 6.025 3.7695
1000 1046.03 114.03 759.02 5.847 3.7810
1020 1068.89 123.12 775.67 5.521 3.8030
1040 1091.85 133.34 793.35 5.201 3.8259
1060 1114.84 143.91 810.61 4.911 3.8478
1080 1137.93 155.15 827.94 4.641 3.8694
1100 1161.07 167.07 845.34 4.390 3.8906
1120 1184.28 179.71 862.85 4.156 3.9116
1140 1207.54 193.07 880.37 3.937 3.9322
1160 1230.90 207.24 897.98 3.732 3.9525
1180 1254.34 222.2 915.68 3.541 3.9725
1200 1277.79 238.0 933.40 3.362 3.9922
1220 1301.33 254.7 951.19 3.194 4.0117
1240 1324.89 272.3 969.01 3.037 4.0308
1260 1348.55 290.8 986.92 2.889 4.0497
1280 1372.25 310.4 1004.88 2.750 4.0684
1300 1395.97 330.9 1022.88 2.619 4.0868
1320 1419.77 352.5 1040.93 2.497 4.1049
1340 1443.61 375.3 1059.03 2.381 4.1229

564
T, h, u, *°,
K kJ/kg Pr kJ/kg V-r kJ/kg K

1360 1467.50 399.1 1077.17 2.272 4.1406


1380 1491.43 424.2 1095.36 2.169 4.1580
1400 1515.41 450.5 1113.62 2.072 4.1753
1420 1539.44 478.0 1131.90 1.9808 4.1923
1440 1563.49 506.9 1150.23 1.8942 4.2092
1460 1587.61 537.1 1168.61 1.8124 4.2258
1480 1611.80 568.8 1187.03 1.7350 4.2422
1500 1635.99 601.9 1205.47 1.6617 4.2585

Adapted from Table 1 in Gas Tables, by Joseph H. Keenan and Joseph Kaye.
Copyright 1948, by Joseph H. Keenan and Joseph Kaye. Published by John
Wiley & Sons, Inc., New York. Reprinted from Fundamentals of Classical
Thermodynamics, Van Wyler and Sonntag, John Wiley & Sons, Inc., copyright
© 1976. By permission.
_APPENDIX 14 _
EMISSIVITIES OF
SURFACES AT NORMAL
ATMOSPHERIC
TEMPERATURE
Metals Nonmetallic Surfaces

Surface Emissivity Surface Emissivity

Aluminum Asbestos 0.93-0.96


Polished 0.04-0.18
Brick
Commercial sheet 0.09
Rough red 0.93
Heavily oxidized 0.20-0.30
Fireclay (at 1000°C) 0.75
Brass
Carbon
Polished 0.03-0.10
Lampblack 0.96
Dull plate 0.22
Candle soot 0.95
Heavily oxidized 0.60
Enamel, white,
Copper
fused on iron 0.90
Polished 0.02-0.06
Heavily oxidized 0.75 Glass, smooth 0.94

Iron and Steel Oak, planed 0.90


Polished steel 0.05-0.07
Paints
Polished iron 0.14-0.38
Flat black lacquer 0.97
Rolled steel plate 0.66
Oil paints, all colors 0.92-0.96
Oxidized steel 0.79
Aluminum paints 0.27-0.67
Heavily oxidized
cast iron 0.64-0.95 Paper 0.92-0.94
Rough steel plate 0.95
Plaster, rough lime 0.91
Nickel
Roofing paper 0.91
Polished 0.05-0.07
Heavily oxidized 0.37-0.48 Water 0.95-0.96
Tin
Bright 0.06
Commercial tin-plated
sheet iron 0.07-0.08
Zinc
Galvanized sheet iron 0.28-0.08

566
APPENDIX 15
DIMENSIONS OF
WELDED AND
SEAMLESS STEEL PIPE
c
>, <u
> 3
-3
«
<D ^ VO 3 »o © OO ON ir, vo X 3 On co o
i_
173 vo co oo 04 04 co q© ^ © co on © <u
o <u '2 s .s 3 VO oo ^ co vo ^ VO 3 CO —< oo *0 — IT) a.
vo Xi c (73 w
© © © 04 04 CO Tt »r, VO 0© © ’— 04 •o
3 oi
JJ u,
■*->
=s X C/5

TS 3
0) w
C/D
4=
-C « C/D
o
o
00 X5

75
^
c o r- oo © © »— co «o 00 ro to x to M VO O 3
3
oo ' IC) «o oot h CO — 04 — 04 ^- © »r, E
—< 04 04 04 04 co co 3" vr, VO r- ^ fO t h
O C/5
£ .Si w © © © © © © © © © © 3
Q 43
H
3
>
>.
3
« £ O' oo © on co vo,
E
tO) M 3 VO (N ©3vOO-
"3 <D O »on r- © CO 04 C/5
1 © 04 3 3 © VO 04 — vO 04v0 00© —
0)
•33 E = (N fO t O h On 04 vr, on co on cq oo oo O' VO to CO to CO C/5
o >» C 3 w © O © ^ C/5
oo > © © © 04 04 CO cd 3 vr (W ON 04 Tt
a>
c
o> S
(J
3 X
L
(V) CZ
i-H U. C/D
C/D
O X „ <L> 3
c o to On vo O' 3 on © oo no © OO O' to, 04 Sco O' © co
^ w g '—I 04 3 vo, o- On © '—< CO O' CO On OO vo 3 £
^ 44 C 04 04 04 co co co co 3 *o vr, vo O' oo
> .Si w ©©©©© ©©©©© ©©©©© © © © © ©
C/5
<u
43 _3
H 3
>
0)
Ofi
3
<u u_
<u ■4-> ON
a/
-C
C/5 c
<u
^ c
On
vO
04
3
CO
co
ON
3-
04
04
©
3-
04
OO
3
©
©
00
co
©

vq
O'
VO
©
ON
VO
3
OO
vo
©
oo
3
vr,
vo
04
©
O'
3
©
vo
vo
©
oo 04
ON ©
3
ON
CO
>
3
o c 3 <U
3 ^ M © © © © © 04 04 CO CO 3 •ri vd w © CO I—
s 3

3 _ <u
-3
<U a>
3
”3
C/D
E
o C/D
_3
oo
c _ <U
3 00 OO ^ ON CO CO © >o 3 co vo VO O' OO © 04 VO O © "5
75 C O x On © rn co Nt t VN © 1 04 CO *r, 00 04 ^O © CO ©
CO ^ M c O © ^ ^ ' 1 '—'04 04 04 04 04 04 co ro 3 3 »n 04
£ .a w 3
©©©©© ©©©©© ©©©©© © © © © ©
43 C/5
c
H
<u
CJ
c
<u
a> <u J3
*3 •r, © »o © © •ON © © vo vo 8 co *o •o
<u O 3 O' 3 »r) ^ vo © O' O' VO 04 04 VO, vr, 3
£ G c to vo oo © co vo On CO oo •o © •O VO vo VO O' O' © © c
3
3 3 w 04 04 CO 3
© © © © 3 vo VO OO © 04 3 VO
°5 c/5
<U
C/5
C/5
a>
c
3 o
.2 w ix *-« i3f ix — m
— I^t ^ l(N _ , — l(N QQ © 04 3 VO
g -Si d — co, i(N |Tt — — 04 04 CO To ^ «o vo
o w -c
3
Z
£

567
APPENDIX 16
CHARACTERISTIC
EQUATION OF IDEAL
GASES FROM BOYLE’S
AND CHARLES’ LAWS

In Chapter 3 it was stated that the characteristic equation for ideal gases may
be derived by the use of Boyle’s and Charles’ laws. The basis of this deri¬
vation is to establish a p- v-T relationship between two state points for any
gas obeying Boyle’s and Charles’ laws. Select at random any two states, as
states 1 and 2, shown in the figure below. Since p, v, and T are properties,
and hence point functions, the relationship between them can be evaluated by
choosing any path or series of paths between the two states.

Characteristic equation from Boyle’s and Charles’ laws.

568
Appendix 16 569

First, let the gas be compressed in a reversible isothermal manner (process


1-3) until its pressure equals that at the final state. Now let the gas be heated
at constant pressure (process 3-2) until the final state (state 2) is reached.
Since process 1-3 is isothermal,

or Pi_ Pi
PiV-\ — p3^3
P3 Pi
For process 3-2,

V-3 'V-i
or
t3_ t,
V-3 — V-2 rj, — V-2 Tf-
T~3“T2 12 12
Equating the two values of v3, gives

„ Pi_ T,
-lAl - #2 T •
P2 12
Hence,

Pi— = = a constant (since this ratio is fixed when the


i i i2
state is specified).

Then

p v = CT
When both sides of the last equation are multiplied by the mass m which is
present, the result is

pm v- = mCT

To distinguish the constant C from other constants, use the symbol R for it.
Then,

pV = mRT

This equation is based on the experimental observations of Boyle and


Charles. Hence it must be a valid relationship for those gases which, for the
given conditions of temperature and pressure, obey Boyle’s and Charles’
laws.
-
CHARTS OF
_PROPERTIES_ ft

Fig. A-l VISCOSITY OF GASES AT LOW PRESSURES 573


Fig. A-2 VISCOSITY OF LIQUIDS 574
Fig. A-3 THERMAL CONDUCTIVITY OF GASES 575
Fig. A-4 THERMAL CONDUCTIVITY OF LIQUIDS 576
Fig. A-5 THERMAL CONDUCTIVITY OF METALS 576
Fig. A-6 EQUILIBRIUM CONSTANTS 577
Fig. A-7 MOLLIER DIAGRAM FOR STEAM 578
Fig. A-8 PSYCHROMETRIC CHART 580

571
44

42

40

38

36

34
yyy.
32 yy
w
30
ID
O4
o
28

O 26
O) y
> 24
o
o
C/5 22
>
20
rk'j
..cv/
/
Jt/
4?y
N 6y
/
ayy
18

16
w
14

12

10

8
0 100 200 300 400 500
Temperature, °C

FIGURE A-l Viscosity of gases at low pressures (i.e., where viscosity is independent of pressure.
There is some uncertainty as to viscosity of some gases.

573
Viscosity, g/sec-cm x 10

FIGURE A-.2 Viscosity of liquids. There is some uncertainty as to the viscosity of some liquids.

574
Thermal conductivity, W/m— 0.08

0 100 200 300 400 500


Temperature, °C

FIGURE A-3 Thermal conductivity of gases at atmospheric pressure. There is some uncertainty
as to the thermal conductivity of gases.

575
0.7

\
0.6

0.5

I
E
§
>
4-»

’>
*+-*
o
D
"O
c
o
o
Id
E
k_
<D Met*VI al
coho/
St
'°ho l

-B enzen e-

"Refr igeran t F-1 2

0 20 40 60 80 100
Temperature, C

FIGURE A-4. Thermal conductivity of liquids.

v:
Thermal conductivity, W/m—

0 100 200 300 400 500 600


Temperature, °C
FIGURE A-5 Thermal conductivity of metals.

576
Common log K

Common log K

Temperature, K

FIGURE A-6 Equilibrium constants Kp.

577
Please refer to back of book for enlargement of this diagram.

FIGURE A-7 Steam tables (International System of Units—SI) Keenan, Keyes, Hill, and Moore,
copyright (c) 1969, John Wiley & Sons, Inc. Reproduced by permission.

578
knlhalpy kj/h|

579
jojoe-j leaH aiqjsuas

<u
l-H
3 .
cn
C/2 o
c
(U '-3
1—
cx cds-
CJ O

Uh CX
h-> k.
<U o
e u
o Urn
Urn
a .Si
X *G
Uh
* c3
(/)
u
c 0)
Z3
•4—1
#o
«4—I
*G
o
<L>
6
c
o
HH
1/2
OQ C/D

<D
u. i—i
E
3 <u
-*->
cd CX
u,
<D
Dh -D
E T3
<U
•4—» o
_ 3
"O
E
U-
O
s-
O CX
c <D

+->
_.
1
I-H
cd J
x:
o W

o >
*C W
-4—1
0>

E C
o
W
X oo
o cd
>>
C/2 CXh
cx
k. >/->
<u <N
'C C<2
u,
cd
O
U

oo

<

W
&
P
o
HH
Uh

580
SOME SELECTED
AND FUNDAMENTAL
REFERENCES

Angrist, S. W., Direct Energy Conversion, Allyn and Bacon, Boston, 1976.
Binder, R. C., Fluid Mechanics, 3rd. Ed., Prentice-Hall, Englewood Cliffs, N. J., 1973.
Brown, A. I., and S. M. Marco, Introduction to Heat Transfer, McGraw-Hill, New York,
1958.
Callen, H. B., Thermodynamics, John Wiley, New York, 1960.
Chang, S. S. L., Energy Conversion, Prentice-Hall, Englewood Cliffs, N. J., 1963.
Glasstone, S., Thermodynamics for Chemists, Van Nostrand, New York, 1947.
Jacob, M., and G. A. Hawkins, Elements of Heat Transfer, John Wiley, New York,
1957.
Keenan, J. H., Thermodynamics, John Wiley, New York, 1941.
Lee, J. F., and F. W. Sears, Thermodynamics, Addison-Wesley, Reading, Mass., 1955.
Lewis, G. N., and M. Randall, Thermodynamics, McGraw-Hill, New York, 1961.
McAdams, W. H., Heat Transmission, McGraw-Hill, New York, 1954.
Reynolds, W. C., Thermodynamics, McGraw-Hill, New York, 1968.
Shapiro, A. H., The Dynamics and Thermodynamics of Compressible Flow, Ronald
Press, New York, 1953.
Smith, J. M., and H. C. Van Ness, Introduction to Chemical Engineering Thermo¬
dynamics, 3rd. Ed., McGraw-Hill, New York, 1975.
Van Wylen, G. J., and R. E. Sonntag, Fundamentals of Classical Thermodynamics,
2nd. Ed., John Wiley, New York, 1978.
White, F. M., Fluid Mechanics, McGraw-Hill, New York, 1979.
Zemansky, M. W., Heat and Thermodynamics, McGraw-Hill, New York, 1951.

581
INDEX
INDEX

Absorptivity, 486, 489 Classical, see Macroscopic, approach


Acoustic velocity, 286 Clearance, 338
Activity coefficient, 273 Coefficient of performance (COP), 374
Air, atmospheric, 220 Combustion, 241, 253
dry, 220, 222 complete, 242
excess, 247 incomplete, 249
theoretical, 244 Combustion temperature, 263
Air conditioning, 225 Compressibility factor Z, 63, 80
Amagat’s isothermals, 62 Condenser, 357
Analysis, gravimetric, 220 Conduction, 478
volumetric, 220, 248 Control mass, 8
Atmosphere, 18 Control volume, 8
Atmospheric nitrogen, 220 Convection, 493
Available energy, 141 forced, 497
Available temperature, 142 free, 495
Availability, 149 Cooling tower, 234
cooling efficiency, 235
Bar, 18 Corresponding states, 79
Barometric pressure, 17 Critical point, 60, 73, 186
Battery, 148 pressure, 59
Bernoulli’s equation, 282 properties, 75, 78
Blackbody, 486 state, 59, 75
Boiling, 501 temperature, 59
nucleate, 502 Cycle, 27
Bose-Einstein statistics, 409, 412 air-standard, 316
Boundaries, 7 Carnot, 118
Boyle, 6, 61 gas, 316
law, 94 open, 315
Brayton cycle, 324 vapor, 356
British thermal unit (Btu), 40
Dalton’s Law, 212
Calorimeter steam, 201 Daniels cell, 148
Carnot cycle, 118 Debye temperature, 429
engine, 139 Degeneracy, 405
reversed, 122 Dehumidification, 231
thermal efficiency, 121 Density, 26
Celsius, 20 Dew point, 224
Charles, 61 Diesel cycle, 321
law, 94 Differential equations, 167
Chemical energy, 261 Differentials, exact, 32, 39
Chemical reaction, 241, 253 Dissociation, 265
endothermic, 253
exothermic, 117, 253 Efficiency, compressor, 340
Clapeyron equation, 174 internal, 340
Clasius inequality, 125, 127 mechanical, 335
Clasius statement, 134 thermal, 120
586 Index

Emissivity, 487 adiabatic, 291


Energy, 1 diabatic, 305
available, 141 laminar, 278, 496
chemical, 34, 261 Newtonian, 279
electrical, 33, 448 transition, 496
equation, 168 turbulent, 278, 496
equipartition, 393 Flow work, 42
extrinsic, 33 Force, 13
internal, 33 Four-cycle engine, 331
kinetic, 35 Fuel cell, 148, 466
nuclear, 34 Fugacity, 271
potential, 33, 35
rotational, 418
Gap energy, 471
solar, 472
Gas compressors, 337
translational, 416
Gas constant, 63
unavailable, 142, 146
universal, 66
vibratory, 423
Gases, 183
Enthalpy, 45
actual (real), 64
changes in, 100, 109
ideal (perfect), 62
Enthalpy-entropy diagram, 194
liquefaction, 178
Entropy, 128, 131, 134, 151, 435
permanent, 22
absolute, 149, 164, 430
Gas turbine, 343
changes in, 171
Gibbs function, 148, 171, 264
current, 437
Grashof number, Gr, 495
increase of, 131, 151
Gravity acceleration, 14
and irreversibility, 130
partial, 215
and probability, 161 Heat, 27
production, 435 of formation, 260
Equation of state, 62 of reaction, 254
actual gas, 64, 76 reservoir, 139
Beattie-Bridgeman, 77 sink, 118
ideal gas, 62 source, 118
van der Waal’s, 60 specific, 39, 416, 418, 423
virial, 77 transfer, 477
Equilibrium, 27, 87 Heat engine, 118, 448
chemical, 265 Heat exchangers, 506
thermal, 19 Heating valve, 254
thermodynamic, 91 Heat pump, 122, 374
Equilibrium constant, 268 Heat transfer, 141, 477, 512
Equipartition of energy, 393 overall coefficient, 503
Ericsson cycle,, 329 Helmholtz function, 47, 169
Humidity, 223
Fahrenheit, 20 absolute, 223
Fanno line, 289 ratio, 223
Feedwater heating, 367 relative, 223
Fermi-Dirac statistics, 411 Hydraulic radius, 282
Film coefficient, 494
Film condensation, 500 Ice point, 20
First law, 2, 31 Ideal gas, 28, 62, 389
applied to systems: Internal combustion engine, 331
for closed systems, 47 Internal energy, 33
for open systems, 46 changes in, 98, 109
Flow, 277 Irreversible process, 97, 131, 141
Jet propulsion, 348 Partial volume, 213
Joule heating, 436 Partition function, 409, 416
Joule-Thomson coefficient, 175 Pascal, 15
Pauli exclusion principle, 411
Kelvin-Planck statement, 134
Peclet number, Pe, 499
Kelvin relationships, 445
Peltier coefficient, 442, 450
Kelvin statement, 134
Phase, 10, 53
Kinetic energy, 30
change heat transfer, 500
molecular, 35
Photovoltaic cells, 471
Kinetic theory, 388
Planck constant, 404
ideal gas, 389
Polytropic process, 93
Kirchhoff’s law of radiation, 489
Potential energy, 35
Lagrange multipliers, 397 Power, 43
Liquefaction, 178 indicated, 334
Liquid, 185 Prandtl number, Pr, 495, 498
compressed, 189 Pressure, 15, 128
line, 72 absolute, 15
saturated, 185 atmospheric, 16
Log mean temperature difference (LM' 509 back, 295
Lyon-Martinelli equation, 499 critical, 59
gauge, 15
partial, 212
Mach number, 288
Probability, 157, 161
Macroscopic, approach, 7
Process, 27
state, 158
adiabatic, 27
Magnetohydrodynamic generator, 460
adiabatic saturation, 225
Mass, 12
constant-volume, 27
Maxwell-Boltzmann statistics, 406, 412
irreversible, 91, 97, 141, 436
Maxwell relations, 169, 171, 179
isentropic, 132
Maxwell speed distribution, 394
isothermal, 127
Mean effective pressure, 319
polytropic, 93
Microscopic, approach, 7, 156
quasistatic, 91
state, 158
reversible, 91
thermodynamics, 388
reversible adiabatic, 101
Mixtures, 211
throttling, 175
gas-vapor, 221
Properties, 9
ideal gaseous, 211
critical, 75
liquid-vapor, 185
dependent, 9
stoichiometric, 244
extensive, 10
Moisture content, 191
independent, 9
Mole, 65
intensive, 10
fraction, 214
reduced, 78
Mollier diagram, 194
stagnation, 284

Newton, 13 Psychrometer, 224


Non-equilibrium state, 203, 434 Psychrometric charts, 228

Nozzle, 293, 296


efficiency, 299 Quality, 191
Nusselt number, Nu, 495, 498 Quantum, 404, 471
state, 405
Onsager relations, 437 theory, 487
Orsat analysis, 250
Otto cycle, 317 Radiant energy exchange, 490
Radiation, 485
Partial pressure, 212 Rankine cycle, 358
588 Index

Rayleigh line, 307 homogeneous, 10


Reduced property, 78 isolated, 8
Reeves statement, 134 nonflow, 48
Refrigeration, 372 open, 8, 42
absorption, 379
vapor-compression, 376 Temperature, 18, 28, 391
water vapor, 380 absolute, 23
Refrigerator, 122 critical, 59
Regenerator, 347 ignition, 242
Regenerator effectiveness, 347 lowest available, 142
Reheating cycle, 365 saturated, 184
Reversible process, 91, 168 triple-point, 24
internally, 91 Temperature-entropy (T-s) diagram, 132, 194
isentropic, 132 Thermal efficiency, 120
Reynolds number, Re, 278, 498 Thermionic generator, 454
Richardson-Dushman equation, 455 Thermodynamics, 1
Root-mean-square (RMS) velocity, 399 of thermoelectricity, 440
Thermoelectric generator, 449
Saturated state, 184 Thermometer, 22
Schrodinger wave equation, 405 Third law, 163
Second law, 3, 134 Thomson coefficient, 444, 450
Seebeck coefficient, 441, 450 Throttling process, 175, 196
Semiconductor, 449, 473 Ton of refrigeration, 378
Shock wave, 301 Triple point, 24, 57
Solar energy, 472 Two-cycle engine, 331
Space charge, 455
Specific heat, 39, 100, 172 Unavailable energy, 141, 146, 151
constant-pressure, 40 Units, 11
constant-volume, 40
gases at extremely high temperatures, 428 Vacuum, 16
hydrogen at low temperatures, 427 Vapor, 183, 185
polytropic process, 102 charts, 194
solids and liquids, 428 dome, 185, 193
variable, 104 dry, 185
Specific volume, 26 line, 185
Specific weight, 15 saturated, 185
Stagnation properties, 284 superheated, 188
State, 10 tables, 188
critical, 59 wet, 185, 191
Statistical thermodynamics, 7, 404 Vapor cycle, 356
Steady-state conditions, 27, 46 binary, 371
Steam, 183 Viscosity, 279
superheated, 188 Volume, 10
supersaturated, 204 partial, 203
tables, 188 specific, 26
Steam point, 20, 22
Stefan-Boltzmann law, 488 Weight, 13
Stirling cycle, 327 Work, 1, 32, 35, 38
Stirling’s approximation, 407 flow, 40
Superheat, 188 maximum, 147, 149
Surface tension, 36 polytropic process, 95
Surroundings, 8, 142 useful, 149
System, 7 Work function, 459
closed, 8
heterogeneous, 10 Zeroth law, 19
Entropy, kJ/(kg'K)
.

You might also like