2015.462984.mechanical Engineering Text

Download as pdf or txt
Download as pdf or txt
You are on page 1of 545

MECHANICAL

ENGINEERING
THERMODYNAMICS
By
DAVID A. MOONEY
ORIENT I^ONGMANS PRIVATE ETD
CALCUTTA UOMBAY IVfADRAS
DRLHI HYOERABAD OACCA

F'zrst eiittzan ptiblished iQ33


PREFACE

This is an introductoiy book on the principles of thermodynamics and their


applications in mechanical engineering. Sufficient descriptive matter on
heat engines is included for those who have not studied the subject. The
reader is presumed to have a knowledge of college physics and mathematics,
including elementary calculus.
There is more than enough material in the book for a two term course.

By selection of chapters or sections for detailed study, leaving the others

for reference use, the book may be adapted to courses of different lengths.

As far as practicable the chapters on applications have been made mdepend-


ent of each other to permit change of order according to the instructor’s
preference.

Thermodynamics is treated in this book as a logical formulation of facts


known from experience. Emphasis is placed on the generality of the subject

and its independence of particular pictures of the structures of substances


and the mechanisms of processes. At the same time, it is recognized that a
book for engineers, rather than for mathematicians, cannot be effective if it

is limited to an economical logical development. Engineers need explana-


tions of the reasons for the steps in a logical development, and examples of
the use of new concepts as soon as possible after their introduction. I have
tried to satisfy these needs, and, as a further aid to understanding, have
identified the following types of concepts as they appear; facts from experi-
ence; principles, accepted because they are confirmed by experience; arbi-
traiy definitions, chosen solely on the basis of utility for their intended
purposes; approximations such as the gas bw, used because actual physical
data do not conform to simple mathematical formubtions; conventions,
devices for facilitating thought and communication.
The general order of presentation is; basic concepts, principles, data on
substances, engmeering applications. The basic concepts of work and heat
are dbcussed in detml, with a number of examples which anticipate the
subject matter of bter chapters. This not only gives more concrete pictures,
but provides for repetition, a valuable learning device.
The First Law b presented in terms of the equality of work and heat in
a cycle, and internal energy b then shown to be a property. The steady
flow energy equation b developed as a special restricted application of the
vi PREFACE

Firet Law; the misleading implication that it is in any sense a more general
equation than the First Law is avoided.
The Second Law is presented as a formal statement of the empirical fact
that irreversibility exists. The thermodynamic temperature scale is defined,
and entropy shown
is to be a property, as corollaries of the Second Law.
The properties of substances are presented first from a general viewpoint,

using plots to show the relationships among gaseous, liquid, and solid states.
Then the special methods of handling data are explained; steam table data,
perfect gas data, and modified gas law data, including gas tables, are discussed.
In the chapters on applications the approach is always from the thermo-

dynamic rather than the descriptive viewpoint, but care is taken to point
out the considerations other than efficiency which always influence the use
ofa particular cycle or apparatus. The distinction is drawn between internal
combustion power plants and cyclic heat engines. In the treatment of com-
bustion an outline of the material balance for a combustion reaction is fol-
lowed by the direct application of the First Law to a system in which a
reaction occurs. It is shown that the same basic methods of analysis can
be used for processes involving chemical reactions as for purely physical
processes,and without any new ad hoc forms of energy.
At many points in the chapters on principles and data there are detailed
explanations and examples which may seem too elementary to some readers.
It has been my experience, however, that students waste much valuable time,
and accumulate antipathy to thermodynamics, because of difficulty with
essentially trivial points of procedure which, I hope, this book will make
clear. In the discussion of applications I have gone into considerable detail,
with the object of giving the reader some idea of the relationship of the
thermodynamic analysis to the whole engineering problem, so that he will
not be working in an intellectual vacuum. However, no attempt is made
to provide a book on plant engineering for any application.
No one can trace all the influences which affect a work of this kind.
I am well aware of my debt to former teachers and colleagues at M. I. T.
As student and staff member under the guidance of Joseph H. Keenan I

saw the power and the scope of the thermodynamic method. I owe to
first

Warren M. Rohsenow the benefits of many hours of illuminating discussion,


on principles and presentation. Many of the problems used in this book
are based on problems compiled by Carl L. Svenson. Most of the illustra-
tions were drawn by Donald E. Eeyt. I acknowledge the kindness of manu-
facturers who supplied illustrations, and of authors and publishers who
granted permission to reprint material; specific credit is given for each item.
Without the help and cooperation of my wife, Eleanor F. Mooney, I could
hardly have completed the work.
David A. Mooney

CONTENTS
1-

2-
1. INTRODUCTION
1-1 Energy. 1-2 Thermodynamics. 1-3 Thermodynamic
systems. 1-4 Properties. 1-5 State changes, and cycles.
3-
6 Units and definitions.

2. WORK
3-
8
1 Work —
definition. 2-2 Application of the definition of
work.
4- 2-3 Work in a stationary system. 2-4 Work in var^

ions processes graphical representation. 2-5 Indicator
diagrams.
4- 2-6 Summary.
3. TEMPERATURE
5-
AND HEAT 24
1 Temperature. 3-2 Temperature equilibrium. 3-3
Thermometers. 3-4 Temperature scales. 3-5 Gas ther-
mometers. 3-6 Absolute temperature. 3-7 Heat. 3-8
Units of heat. 3-9 Specific heats. 3-10 Latent heats.
11 Summary.
6-
4. PROPERTIES OF SYSTEMS 41
1 Properties —
physical concept and general concept. 4-2
Properties mathematically derived. 4-3 Properties derived
from general laws. 4-4 Point functions and path functions.
5 Heat and temperature.

5. THE FIRST LAW OF THERMODYNAMICS ... 46


1 Background of the first law. 5-2 Equality of heat and
work in a cycle. 5-3 Units of energy. 5-4 The first law
and internal energy. 5-5 Internal energy general and
special concepts. 5-6 Specific heat at constant volume.

^7 Enthalpy. 5-8 Specific heat at constant pressure. 5-9
Latent heats. 5-10 Applications of the first law to sta-
tionary systems. 5-11 Summary.

6. FLOW PROCESSES—FIRST LAW ANALYSIS . . 67


1 Flow processes —control surface. 6-2 Steady flow
vii
VIU CONTENTS
definition. 6-3 Material balance and energy balance in
simple steady flow processes. 6-4 Applications of the
steady flow equations. 6-5 Use of the control surface in
7-
simple variable flow processes. 6-6 Examples of variable

flow problems system and control surface analysis. 6-7
Summary.
7. THE SECOND LAW OF THERMODYNAMICS . . 92
8- 1 Introduction to the second law. 7-2 Heat engine cy-
cles. 7-3 The second law. 7-4 Reversibility. 7-5 Irre-
versible processes.
8- 7-6 Mechanical and thermal reversi-
bility-reversible systems. 7-7 The Carnot cycle. 7-8
The reversed heat engine. 7-9 Clausius’ statement of the
second law. 7-10 Summary.
9-
8. BASIC APPLICATIONS OF THE SECOND LAW . Ill
1 The efficiency of reversible cycles. 8-2 The
absolute
thermodynamic temperature scale. 8-3 Entropy. 8-4 The
inequality of Clausius. 8-5 The temperature-entropy plot.
6 The availability of energy. 8-7 Entropy and reversi-
bility. 8-8 Examples of second law problems. 8-9 Sum-
mary.
10.
9. PROPERTIES OF PURE SUBSTANCES
10- .... 137
Methods of presenting property relations. 9-2 Solid,
1
liquid and gas phases. 9-3 Saturation states. 94 Equilib-
rium between phases. 9-5 Phase diagram on the pressure
volume
11-
plane. 9-6 Enthalpy and internal energy of the
pure substance. 9-7 Entropy of the pure substance:
temperature-entropy plot. 9-8 General property relations
forpure substances.

TABULATED PROPERTIES-STEAM TABLES . . 154


12- 1
Saturation states. 10-2 Liquid-vapor mixtures. 10-3
Superheated vapor. 104 Compressed liquid. 10-5 Charts
of thermodynamic properties. 10-6 Measurement of steam
quality.

11. PROPERTIES OF GASES


1 p, V, T relations for the perfect gas. 11-2 Specific
heats,mtemal energy and enthalpy. 11-3 Entropy
of the
^rfect gM. .114 The reversible anabatic
process. 11-5
The ^lytropic process. 11-6 Gas tables.
11-7 Modified
p, t), T relations—compressibility charts.
12. PROPERTIES OF GASEOUS MIXTURES
1 Basic rules for mixture properties.
12-2 Perfect gas

CONTENTS IX

mixtures: p, V, T relationships. 12-3 Parts by mass



parts by volume mol fractions. 12-4 Internal energy,
enthalpy, and specific heats of gas mixtures. 12-5 Entropy
of gas mixtures. 12-6 Mixtures of gases and water vapor.

PROCESS CALCULATIONS FOR STATIONARY


13. 13-
SYSTEMS 203
13-1 Constant-volume process. 13-2
Constant-pressure
process. 13-3 Constant-temperature process (isothermal).
14-4 Reversible adiabatic process. 13-5 Reversible poly-
tropic process. 13-6 Irreversible process in which pi)" =
constant.
14-
14. VAPOR CYCLES—THE RANKINE CYCLE ... 214
1 Thermal power plants and ideal cycles. The 14-2
Rankine
15-
15-
cycle. —
14-3 Rankine cycle influence of pressure
and temperature. 14-4 Actual vapor cycle processes. 14-5
The actual engine or turbine. 14-6 The actual pump.
7 Actual pipe lines. 14-8 Comparison of Rankine and
16-
Carnot cycles.

16-
15. VAPOR CYCLES—MORE EFFICIENT CYCLES . 239
1 The reheat cycle. 15-2 The
regenerative vapor cycle.
3 By-product power and heat. 15-4 Two-fluid or binary
vapor cycles.
17-
16. GAS CYCLES 265
1 The
Stirling and Ericsson cycles. 16-2 Air-standard
cycles. 16-3 The Otto cycle (constant-volume cycle).
4 The Diesel cycle (constant-pressure cycle). 16-5 The
mixed cycle (limited pressure cycle). 16-6 The Brayton
18-
cycle. 16-7 EflBciency and capacity characteristics of the
Brayton cycle. 16-8 More complex gas turbine cycles.

17. FLUID FLOW—NOZZLES AND ORIFICES ... 293


1 Nozzles —mass and energy
relations. 17-2 The revers-
ible nozzle. 17-3 Real nozzle coefficients. 17-4 Converg-
ing nozzles. 17-5 Converging-diverging nozzles. 17-6 Real
nozzle computations. 17-7 Diffusers. 17-8 Supersatura-
tion in nozzle flow. 17-9 Nozzles and orifices as flow
meters.

18. TURBINES 322


1 Turbines. 18-2 Force relations for a fluid stream in
X CONTENTS
steady flow. 18-3 Force and work on a turbine bucket.
18-5 Turbine stages—pas-
18- Nozzle-bucket
18-4 efflcisncy.
sage area. 18-6 The simple impulse stage. 18-7 The
velocity-compounded stage. 18-8 The reaction stage.
9 Multistage turbines. 18-10 Comparison of turbine
tjrpes.
19- 18-11 Turbine characteristics.

19. RECIPROCATING EXPANDERS AND COMPRES-


SORS 360
1Cylinder and piston arrangements. 19-2 Ports and
valves. 19-3 Work in a reciprocating expansion engine or
20-
compressor: indicator card analysis. 19-4 Steady flow
analysis. 19-5 Machine efficiency. 19-6 Flow capacity
of reciprocating machines. 19-7 Liquid pumps.

THE STEAM ENGINE


20. 21- 373
1 The
slide-valve steam engine. 20-2 Comparison of
real and
idealized indicator diagrams. 20-3 Cylinder con-
densation. 20-4 Methods of reducing cylinder condensa-
tion. 20-5 Conclusion.

21. GAS COMPRESSION 383


1Compression processes. 21-2 Work of compression
in steady flow. 21-3 Efficiency of a compressor. 21-4
22- —
Reciprocating compressors work of compression. 21-5
Volumetric efficiency of reciprocating compressors, 21-6
Volumetric efficiency and clearance. 21-7 Volumetric effi-

ciency and pressure ratio multistage compression. 21-8
Intercooling. 21-9 Minimum work in two-stage compres-
sion with intercooling. 21-10 Comparison of compressor
types.

22. COMBUSTION PROCESSES—FIRST LAW


ANALYSIS 405
1 Fuels. —
22-2 Combustion reactions material balance.
22-3 Products analysis. 22-4 Material balance for a com-
bustion process. 22-5 Energy equation for a chemical re-
action. 22-6 Heating values of fuels. 22-7 Temperature
of products.

23. INTERNAL COMBUSTION POWER PLANTS . . 435


M-1 ^
Spark-ig;nition engine process. 23-2 Compression-
ignition engine process. 23-3 Supercharged enginpn 23-4
The ^ simple gas turbine plant. 23-5 More complex gas

^

turbine plants jet propulsion.


CONTENTS xi

24. REFRIGERATION 457


24-1 Reversed heat engine cycles. 24-2 Vapor compression
refrigeration cycles. 24-3 Capacity of a vapor compres-
sion plant. 24-4 Power consumption of a vapor compres-
24-
sion plant. 24-5 Effect of irreversible heat transfer on
plant performance. 24-6 Expansion engines—gas cooling.
25- Liquefaction of gases. 24-8 Absorption r^rigeration.
24-7
9 Heat pumps.

25. AIR-WATER VAPOR MIXTURES 480


1 Definitions. 25-2 Humidity relations. 25-3 Wet-
bulb temperature. 25-4 Adiabatic saturation process.
25-5 The psychrometric chart. 25-6 Humidif3ring proc-
esses. 25-7 Mixing processes. 25-8 Dehumidifying proc-
esses. 25-9 Cooling towers.

APPENDIX 499

INDEX 529

BACK-COVER CHARTS
El: Temperature-Entropy Diagram for Carbon Dioxide
E2: Temperature-Entropy Diagram for Air

E3: Mollier Chart for Steam


E4: Pressure-Enthalpy Chart for Freon-12
E5: Properties of Chemically-Correct Fuel-Air Mixtures
E6: Psychrometric Chart, Normal Temperatures
MATTER AND ENERGY
All that we know about matter relates to the series of phenomena
in which energy transferred from one portion of matter to another,
is

till in some part of the series our bodies are affected, and we become

conscious of a sensation. . . .

Hence we are acquainted with matter only as that which may


have energy communicated to it from other matter, and which may,
in its turn, communicate energy to other matter.
Energy, on the other hand, we know only as that which in all
natural phenomena is continually passing from one portion of matter
to another.
James Clerk Maxwell
Matter and Motion^ 1877

A PRECEPT FOR ENGINEERS


Economy of fuel is only one of the conditions a heat engine must
many cases it is only secondary, and must often give way
satisfy ; in
to considerations of safety, strength and wearing qualities of the
machine, of smallness of space occupied, or of expense in erecting.
To know how to appreciate justly in each case the considerations of
convenience and economy, to be able to distinguish the essential
from the accessory, to balance all fairly, and finally to arrive at the
beat result by the simplest means, such must be the principal talent
of the man called on to direct and co-ordinate the work of his fellows
for the attainment of a useful object of any kind.
S. Carnot
Reflexiona sur la Puissance Motrice du Feu^ 1824
Translation by H. L. Callendar
Chapter i

INTRODUCTION

Of all the technical achievements of man the most important is the


capacity to control large quantities of energy. Only through the
wide availability of power, or controlled energy, can the results of
science be put to effective use for the general population. Industry,
transportation, agriculture, even scientific research itself, if denied
the use of power-driven machinery, would still be struggling along
the same roads they followed centuries ago.
The broad purpose of this book is to explain a method of analysis,

based on the science of thermodynamics, for processes involving the


control or use of energy.
1-1 Energy. What is energy? Since it is known only through
its effects, it is impossible to visualize what energy is. The existence
of energy is accepted, however, because logic demands it. Experi-
ence shows that a physical effect such as the hoisting of a weight can
be accomplished by maintaining a flow of electric current through a
motor, or by maintaining a flow of steam through a turbine, or by
maintaining a fire under a hot air engine; furthermore there are
quantitative relations between the amounts of electricity, steam, or
fuel used, and the amount of weight raised through a given elevation.
Something must exist, common to the different physical systems, by
which they can all accomplish the same effect. It is therefore as-

sumed that energy is transmitted from the electric conductors to the


hoisting mechanism, or from the steam to the hoisting mechanism,
or from the fire to the hoisting mechanism. The justification for this

assumption is found in the agreement between experimental facts


and the results of reasoning based upon the assumption. The exist-

ence of energy is inferred from the physical effects of energy transfer.


Experience shows that all cases of energy transfer may be classified

as either heat transfer or work transfer. A fundamental difference


2 INTRODUCTION [
1-2

exists between these two forms of energy as will appear in later


chapters.
Although energy can be directly identified only through its trans-

fer from one body or system of bodies to another, yet the possibility
of storage of energy within bodies must be considered, because some-
thing which can be quantitatively accounted for during its transfer
from one place to another must either be stored or created and de-
stroyed before and after its transfer. The fact that energy can be
stored in bodies is the substance of the First Law of Thermodynamics.
Work, heat, and the First Law will be the subjects of the first part of
this book.
The engineering processes which form the subject of the greater
part of this book, involve the conversion of energy from one form to
another, the transfer of energy from place to place, and the storage
of energy in various forms. These operations are accomplished by
utilizinga working svbsiance which can absorb, hold, or release energy
depending upon the conditions imposed upon it. The working sub-
stance is almost invariably a fluid (liquid or gas) because fluids are
easy to transport from place to place in pipe systems and because
fluids are conveniently made to absorb and
release energy. Fxamples
of working fluids are: fuel, air,
and combustion products in internal
combustion power plants; liquid water and steam in steam
power
plants; ammonia in refrigeration plants; air and water
in air-condi-
tioning plants.
1-2 Thermodynamics. Thermodynamics is the science of en-
ergy transfer in ite relation
to the physical properties of substances.
It is concerned with the general laws of
energy applying to all types
of systems, mechanical, electrical, and chemical.
It is also concerned
with the overall effects of energy transfer rather
than with the physi-
cal mechanisms by which these effects
are accomplished.* Engineer-
ing thermodynamics is the special development
of thermodynamics
as applied to mechanical engineering
operations such as power gener-
ation from fuels, refrigeration, and
air-conditioning.
1-3 Thermodynamic systems.
A thermodynamic system is a
body or group of bodies upon which attention
is to be concentrated

applied to the characteristics of


moleculM but only to f
the statistical characteristics of
individual
large of mole-

pj^mena rather than microseopic phenomena.


ond Themodytumtes. New
See Zemansky
York: McGraw-Hill, 1»43, phfip
M W
i.
1-4] INTRODUCTION 3

in the analysis of a problem. The whole structure of thermodynamic


anal}rsis is based upon an accounting scheme in which energy quan-
tities are checked into or out of the i^stem. Hence it is necessary
to choose a definite system and to keep it under close scrutiny
throughout the course of the operation being analyzed in order that
all energy quantities involved shall be properly entered in the ac-
count. As the term is used in this book, a system consists of the
same bodies throughout an operation; no material may be added or
taken away. It is usually convenient to visualize a system as en-
closed %vitbin an imaginary envelope which may change its size, shape,
or location but which must always contain the system and nothing
but the system. The selection of a suitable system is the most
important step in a thermodynamic analysis; in the examples appear-
ing in the later chapters of this book some of the factors which
determine the choice of a system will become evident.
1-4 Properties. Every system has certain characteristics by
which its physical condition may be described; examples are mass,
volume, pressure, temperature, and electrical resistivity. Such char-
acteristics are called properties of the system.
1-5 State changes, processes, and cycles. When all the prop-
erties of a system have definite values, the system is said to be in a
definite state. When a different value exists for any one or more of
the properties, the system is in a different state. Any operation in
which one or more of the properties of a system changes is called a
change of stale. The path of a change of state is the succession of
states passed through during the change of state. When the path
is completely specified, the change of state is called a process. A
process is by stating that some particular property
often specified
remains constant; an example is a constant-pressure process. An-
other way of specifying a process is by some algebraic relation
between properties; an example is a process in which the product of
pressure and volume remains constant. A thermodynamic cyde is a
series of state changes such that the final state is identical with the
initial state. Observation of a system before and after the execution
of a cycle would reveal no difference in any property of the system,
although all the properties might have had different values at some

time during the execution of the cycle.


1-6 Units and The system of units used in this
definitions.
book is the pound-second-foot system. The unit of mass is the pound
4 INTRODUCTION [1-8

mass, which determined by reference to a carefully preserved stand-


is

ard. The unit of force is the pound force, which is the force required
to accelerate a pound mass at the standard rate of 32.17 ft/sec*.
These units are related by Newton's Second Law of Motion, which
is expressed by

F = kma (a)

where F = force in pounds force (Ibf).

m= mass in pounds mass (Ibm).

a acceleration in ft/sec*,
/ Ibf sec^ \
and k a proportionality constant ^
Vlbm ft /
In the system used here the units of force and mass are arbitrarily
chosen so that the weight of a body at the earth's surface will have
the same numerical value as its mass. At the earth's surface the
acceleration of gravity is 32.17 ft/sec*. So, if F is numerically equal
to m, then fc = 1/32.17. The statement of Newton’s law then
becomes
im
(b)
32.17

which contrasts with the usage of physicists who arbitrarily assign


the value unity to the constant k. Note that the 32.17 in Eq. (b)
is not an acceleration, nor doesit change for different locations.*

The weight w a body of mass m, at a location where the accelera-


of
tion of gravity has any value g will be

mg
^ (c)
32.17

However, at any elevation normally encountered, even in high-flying


airplanes, the value of g differs by only
a fraction of a percent from
32. ] 7 hence for practical purposes
; w = m (numerically). (The sym-
bol go will be used for the dimensional constant 32.17 ft Ibm/lbf sec*.)
The unit of time for all dimensionally correct equations is the
second.
The unit of length is the foot.

• See Hawkins, L. A. and S.

1946; pp, 143-144, 660-662.


A. Moss, Mechanical Engineering, Vol. 68,
•>!,,, No. 2,
1 -6] INTRODUCTION
The unit of volume is the cubic foot. However it may be con-
venient to work in terms of specific volume, which is the volume of a
unit mass (dimensions, cu ft/lb).
Density is the reciprocal of specific volume (dimensions, Ib/cu ft).

Pressure is the normal force exerted by a system against a unit


area of its bounchng surface. Pressures acting outward against the
confining walls are taken as positive. Thus, the pressure of a body
in compression is positive, but the pressure of a body in tension is

negative. In this book only positive pressures are considered, since


fluids do not normally sustain appreciable tension.
The basic units of pressure in this book are Ib/sq ft. Unfortu-
nately, standard practice in engineering calls for the use of a variety
of units. Pressures are variously stated in Ibf/sq in. (psi), inches of
mercury, and feet, or inches, of water. Since the density of liquids
varies with temperature, pressures in terms of liquid head imply a
liquid of standard density. Taking the density of water as 62.4
Ib/cu ft (corresponding to ordinary room temperatures), we find that

1 ft water = 62.4 Ib/sq ft = 0.433 psi

Since the specific gravity of mercury at room temperature may be


taken as 13.6, a pressure of

1 in. mercury = 62.4 X 13.6 X ^


= 70.7 Ib/sq ft = 0.491 psi.

Fig. 1-1. Pressure gages, (a) Bourdon gage measures difference between
system pressure inside tube and atmospheric pressure, (b) Open U-tube indi-
cating gage pressure, (c) Open U-tube indicating vacuum, (d) Closed U-tube
indicating absolute pressure. If P is atmospheric pressure this is a barometer.
6 INTRODUCTION [1-6

When presfjures are given in atmospheres, a standard atmosphere of


760 mm of mercury (29.92 in. of mercury, or 14.696 psi) is implied.
Most instruments indicate pressure relative to the atmospheric pres-
sure, whereas the pressure of a system is its pressure above zero, or
relative to a perfect vacuum. The pressure relative to the atmos-
phere is called gage pressure; the pressure relative to a perfect vacuum
is called absolute pressure,

absolute pressure = gage pressure -f atmospheric pressure

When the pressure in a system is less than atmospheric pressure, the


gage pressure becomes negative but is frequently designated by a
positive number and called vacuum. Figures 1-1 and 1-2 show rela-
tionships among the various pressure scales.

GAGE PRESSURE ABSOLUTE PRESSURE


ISptIg —— 29.7 polo

0 psIg 0 In Hg Vac. — — 29.92 InHg Abo « l4.7pola • lotm

15 In Hg Voc. —— 14.92 InHg Abo

ZaSE In Hg Voc. —~ 0 * Perfect Vacuum

psig « pounds per squore inch


goge
ptio « pound! per tquore Inch
obiolute
Dlogrom booed on otmoophere pretture
l4.7ptlo

Fig. 1-2. Absolute pressure and gage pressure.


INTRODUCTION 7

PROBLEMS*
1-1. Find the weight Id Ibf of a 5 kilogram mass in a location where the
acceleration of gravity is 32.00 ft/sec^ How much force, in Ibf, would be neceik
sary to accelerate the mass at the rate of 32.17 ft/sec*7
1-2. The acceleration of gravity is given as a function of elevation above sea
level by p = 980.6 - 3.086 X 10“*H, where g is in cm/sec®, and H is in centi-

meters. If an airplane weighs 10,000 Ib at sea level, what is the gravity force
upon it at 20,000 ft elevation? What is the percentage difference from the sea-
level weight?

1-3. Convert the following readings of pressure to psia, assuming the barom-
eter reads 29.92 in. mercury (Hg); 76 in. Hg gage; 32 psig; 16 in. Hg vacuum;
3.3 ft HsO gage.
1-4. A certain pressure is measured by a water-column gage in which the
water is at SST. If the density of water at 39‘’F is taken as an arbitrary stand-
ard, what will be the percentage correction factor to reduce the reading to a
standard density reading? At what water temperature would the correction be

1 percent? Obtain the specific volume of water from the column headed ‘^Sat.

Liquid** in the steam table in the Appendix.


1-S. The volume coefficient of expansion of mercury at ordinary temperatures
is 101 X lO"^ cu ft/cu ft deg F. If pressures in standard inches of mercury are
understood to be in terms of mercury density at 32*’F, at what temperature would
a mercury column be 1 percent higher than a standard mercury column exerting
the same pressure?
1-6. Assume that the pressure p and the specific volume v of the atmosphere
are related according to the equation pv^’* = 75,500, where p is in Ib/sq ft, and
V is in cu ft/lb. If the acceleration of gravity is constant at 32.17 ft/sec®, what
is the depth of the atmosphere necessary to produce a pressure of 14.7 psia at
the earth’s surface? (Consider the atmosphere as a fluid column.)
1-7. A common fallacy regarding the engineering system of units is that the
engineering form of Newton’s Second Law is

ff

where w is weight and g is the local acceleration of gravity. Explain the dis-
tinction between this equation and equation (b) on page 4. Failure to under-
stand the engineering system of units is often the result of ignorance of this
distinction.

*
Conversion factors and physical data for these and problems of succeeding
chapters may be found in the Appendix.
.

Chapter 2

WORK

2-1 Work-definition. Work, one of the basic forms in which


energy is transferred, is encountered in the science of mechanics as
the result of the action of a force on a moving body. A force is a
means of transmitting an effect from one body to another, but a
force is not a form of energy, because it never accomplishes a physical
effect except when combined with motion. An effect such as the
raising of a certain weight through a certain distance can be accom-
plished by using a small force through a large distance, or a large
force through a small distance. Similarly the velocity of a body
may be increased a certain amount by
using a small force through a large dis-

1 tance, or a large force through a small dis-


tance.
of force
In every case, however, the prodxict
and distance is the same if the
same effect is produced. The following
I

H
1

I
definition of work is therefore given in
I J mechanics;
b
Work is done by a force when the force
Fig. 2-1.
acts upon a body moving in the direction of
the force.
The magnitude of the work is given by the product of the force
and the distance moved parallel to the force.
In Fig. 2-1 the body shown moves from o to 6
while acted on by
the horizontal force F
The horizontal component of the distance
moved is I, The work, W, is given by

W = FI
The positive directions of force and motion are taken to be the same.
Therefore, the work wiU be positive if the force and displacement
have the same direction.

8
2- 1 ] WORK 9

The action of a force through a distance (and the equivalent ac-


tion of a torque through an angle)
medumieal work, because
is called

other forms of work can be recognized in thermodynamics. Energy


transfer by an electric current is one example. Most work in the
scope of this book is mechanical work. Nevertheless, a definition of
work in thermodynamics should allow for forms other than mechani-
cal work, and this definition should be stated in terms of a system.
It is also desirable to state the definition in terms of a simple ph3rsical
picture such as the raising of a weight. The following definition of
work IS therefore given:
Work is transferred from a system by a given action of the system if
the total external effect of the given action can be redticed to the raising
of a weight*
The work is measuied by the amount of weight that would be
raised through a specified height if the total external effect were the
raising of a weight. Work transferred from a system is taken as
positive; conversely work transferred to a system is negative with
respect to that system.
The definition includes mechanical work because mechanical work
can always be converted by a suitable mechanism to a vertical force
raising a weight. Similarly an electric current flowing from a system,
through an external circuit in which the potential drops, and back
into the system, constitutes a flow of work from the system. For,
no matter what is actually done with the current in the external cir-
cuit, it could have been used in an electric motor to raise a weight.

Hence, the system transfers work when the current flows, even though
the actual use of the current might be in a resistance heating unit.
It should be observed that, by the definition, the total external
effect of an actionmust be reducible to the raising of a weight if the
action is to constitute a work transfer. There may be some ques-
tion as to whether or not external mechanisms and motors could
convert the action of the system into the raising of a weight without
causing extraneous effects due to mechanical friction and electrical
resistance. The answer is that, in using the definition of work, it is
assumed that and resistance can be eliminated from a hypo-
friction
thetical apparatus upon which the system acts. If, using hypotheti-
cal frictionless mechanisms, the total effect of an action can be con-
verted to the raising of a weight, then that action constitutes a work
* An external effect ia an effect on things outside the system.
10 WORK [2-2

transfer. The reason for requiring that the totoi efifect of the action

must be reducible to the raising of a weight is that another form of


energy transfer from a system, namely heat, may cause the raising
ofa wfli gb t as 'part of its external effect; but even when frictionless
mechanisms are permitted, the U>tol effect of heat transfer can never
be reduced to the raising of a weight.
The work does not tell how to identify work trans-
definition of
ferred irdo a system. To identify work transferred U> a given system
it is sufficient to show that the action of a second system upon the
given system constitutes a transfer of work from the second system.
Tliis indirect approach is necessary because work flow to a system
can not always be identified by its effects in that system. For exam-
ple, the temperature of a block of wood may be raised by rubbing

the block, thus doing work upon it, or by exposing the block to a
fire, in which case no work is done. Observation of the block alone,
without consideration of the systems acting upon it, would disclose

that the temperature had risen, but would give no evidence as to


vdiether this effect was the result of a work transfer or of some other
action.
2-2 Application of the definition of work. The action of a
force over a distance represents work in thermodynamics as well as
in mechanics if the system is so chosen that the force acts across its
boundary. Considering a man as a thermodynamic system, he does
work when he lifts a weight; the work is positive with respect to the
man as a system, and its magnitude is wz ft lb where w is the weight
in pounds and z is the distance in feet through which the weight is

raised. While raising the weight the man also raises his arm,which
has some weight, but he does no work upon his arm because it is a
part of the system (the man). Work exists only as an effect trans-
ferred across the boundary of a system. Taking a locomotive as a
system, it does work upon the train it draws because the force trans-
mitted through the draw-bar as the train moves could have been used
for the sole effect of raising a weight. The work is positive because
it is done by the system. If the train were taken as the system, the
work would be unchanged in magnitude but would be negative be-
cause it would be done on the system. If the locomotive and train
together were taken as the system, the work would be zero because
ti>e force involved would not act across the boundary of the system.
2-2] WORK 11

These examples illustrate the importance of working with a definite


system.
Consider a cylinder and piston machine, Fig. 2-2. Let the gas
in the cylinder be a system having initially the pressure pi, volume
Fi, and length U. Now allow the piston to move out to a new posi-
tion and let the new pressure, volume, and length of the system be
respectively pj, Fj, and k- At any point in the travel of the piston
let the pressure be p, the volume V, and the length 1. It is desired

GAS SYSTEM

to find the work done by the system as the piston moves from posi-
tion 1 to position 2. Examining the boundaries of the system, it is
seen that the piston is the only boundary that moves. Therefore no
work can be done by the pressure force against the other walls, but
work will be done upon the piston when it moves. The force against
the piston will probably vary as the piston moves, but it may be
written, for any position of the piston, as

F — pa

where F is the pressure force on the piston and a is the area of the
piston. For an element of piston travel the element of work may
be written
dW * Fdl

because the motion of the piston dl is parallel to the force upon it.

Then
dW = pa dl
12 WORK 2^]

But a dl may be written as dV, an element of volume; then

dW = pdV

If p is known as a function of V, W can be evaluated from this


equation.
The work done by a system as a result of a change in volume is

given in general by dV for any system in which the pressure p


J^ p
is uniform over the whole surface of the system at all times (though
it may vary with time). The pressure will not be uniform over the
surface of a system if gravity or acceleration forces are present.*
However, with gaseous systems in particular, there are many cases
in which the effects of gravity and acceleration forces are small; in
such cases the work due to a change in volume may be calculated
with good accuracy by the integral of p dV.
2-3 Work in a stationary system. A system in which motion
is is called a stationary system.
n^ligible This deffnition permits
relative motion of parts of the system, incidental to state changes,
but excludes any motion, in translation or rotation, of the system as
a whole, t Work may cross the boundary of a stationary system as
the result of a volume change, and such work is measured by the
integral of p dF if pressure equilibrium prevails in the system. The
integral of p dV, however, does not necessarily represent all the work
transferred across the boundary of a stationary system in a given
change of state. Consider the stationary fluid system showTi in Fig.
2-3. If the system is a real fluid, rotation of drum B in the fluid can
be maintained only by the continuous action of a torque
(from the
weight w) upon the rotating shaft. The work thus done
upon the
drum is transferred into the fluid system through the friction forces
• Surf^ tension, magnetism, and
electricity may sometimes affect the pres-
sure distnbution m fluid systems,
but such effects are outside the scope of this
development. In sohd systems uniform pressure is
the exceptional case.
T The pressure of a gas on a moving wall
depends on the velocity of the gas
molecules relative to the wall. The average
velocity of a gas molecule at a tem-
I»rature of is about 1600 fps, and at higher
temperatures is greater. Since
the speeds of pistons seldom exceed 50 fps,
pistons may usually be considered
station^ relative to the gas molecules, and the
gas confined by a piston may
be considered stationary.
2-4] WORK 13

at the surface of the drum. If a paddle wheel is substituted for the


drum, a torque is still necessary to maintain rotation of the paddle
wheel, and the work involved is transferred to the fluid. Work intro-
duced into a ^stem by such means as the friction drum or the paddle
wheel may be called 'paddle-wheel work or birring loork. Such work
cannot, in general, be measured by a function of the properties of
the system but must be evaluated from some other knowledge of the
circumstances.

Fig. 2-3.

It is of interest to consider the effect of using an ideal frictionless


fluid for the system of Fig. 2-3. The work due to a volume change
would still be given by the integral of p dV but there would be no
paddle-wheel work. Experience shows that as the frictional effects
in a approach zero the torque required to maintain steady rota-
fluid
tion of a paddle wheel approaches zero. Hence with a frictionless
fluid in a stationary system the paddle-wheel work is zero. The only
work transfer that will occur, to or from such a system, is the work
transfer due to normal on moving boimdaries. If pressure
forces
equilibrium prevails, this work will be measured by the integral of
pdV.*
2-4 Work in various processes —graphical representation.
Since many problems involve work given by /"p dV, it is convenient
to make a graphical presentation on the coordinates of pressure and
volume. The integral will appear as an area upon such a plot. Fig-
ure 2-4 represents a general process in which a stationaiy eystm,
* Work transfer by means of surface tension, magnetism, and electricity is

possible but is excluded here.


14 WORK 2-4]

having initially the pressure and volume indicated 1, by point changes,

by the path shown, to a state in which the pressure


and volume
point 2. If the shape of the path is known as a
function
^
indicated by
can he computed by calculus; and if the
fpdV
of p and V, the
mathematical equation is
path is an experimental plot for which no
known, the fpdV can still be computed by
graphical integration.

Fig. 2*^4. Graphical representation of a process.

Several examples are given below of the functional and graphical


representation of J'p dV for various processes in a stationary system,
(a) Ccnstant-Pressure Process.

V
Fig. 2-5. Constant-pressure process.

For this case J^pdV^p(V,-Vj)


M] WORK 15

(b) Conttaid-Volume Process.

Fig. 2-6. Constant-volume process.

For this case f^pdr~o


(c) Process in Which pV = Constant. (This and the following
relation between p and V are functions which approximate the rela-
tions observed by experiment in certain types of processes with certain
fluid systems.)

Fig. 2-7.

For this case J^pdV=>ptV^(biV,-]nVi)

Tr 1 Vs .
16 WORK [2-4

(d) Proem in Which pV* = Constant, Where n Is Any Constant.

Fig. 2-8.

Figure 2-8 shows the shapes of the pV plots of the function pF* *= C
for positive values of n between zero and infinity. (For negative n
the curves occupy the other two quadrants.)
For this case

P.F. IP2V2 _
1 - nLpiFi J 1 - nLVPi/
These equations give an indetenninate result for the case of n = 1,

but that case was considered in (c) above.

Example 1. A stationary fluid system is subjected to a process in which


the pressure and volume change according to the relation pV^'* = C, The
initial and volume are respectively 100 psia and 3 cu ft, the final
pressure
pressure is 20 psia. (a) Find the magnitude and direction of the p dV work
for this process, (b) Is this the total work for the process? Why?
Solution: (a)

Ji 1 - n
AU factoiB are knonm except Vt.

(IOO)(H4)(3)
fpir.
JX -I 1.4

Note the oonvemon of pressures to Ib/sq ft.


:

WORK 17

(b) It is impossible to say whether this is the total work because the
problem does not say whether the process was frictionless or not. If the

fluid were frictionless,no shearing forces and no paddle-wheel work


or if

were involved, the answer to (a) would be the total work.

Example 2. In a stationary fluid system. Fig. 2-3, a paddle wheel sup-


plies work at the rate of 1 horsepower. During a certain period of 1 minute
the system expands in volume from 1 cubic foot to 3 cubic feet while the
pressure remains constant at 69.4 psia. Find the total system work during
the 1-minute period.

Solution:

Paddle-wheel work, Wa-


1 hp for 1 min = 33,000 ft lb

Wa = —33,000 ft lb; negative because flowing into system

Piston work, Wa:


Wa - - Pi) = (69.4)(144)(3 - 1)

= -1-20,000 ft lb

Total or net work, W


W =‘Wa + Wb
= 20,000 - 33,000
= —13,000 ft lb; net work to the system because negative

2-5 Indicator diagrams. The plots discussed in the preceding


section were all state diagrams; the pressureand volume coordinates
of any point represented the corresponding properties of a fluid syB-
tern at a particular state. Another kind of plot which may represent
work is the indicator diagram. This is the trace made by a recording
pressure gage, called an engine indicator, attached to a chamber in
which the pressure varies. The coordinates of the record in the con-
ventional case are pressure vs. piston travel. An arrangement for
making such a record shown in Fig. 2-9. Beferring to this figure,
is

the indicator is connected by a short pipe to the engine cylinder so


that the pressure upon the indicator piston A will be always the
same as the pressure upon the engine piston F. The indicator piston
is loaded by a spring so that it moves in direct proportion to a change

in pressure. The motion of the indicator piston causes a pencil on


the end of the pencil linkage B to move in a vertical line in direct
proportion to the pressure change. The pencil at the end of the
18 VORK [24

linkage B writes upon a strip of paper wrapped about the drum C.


If the drum C remained stationary, the record of pressure changes
would be a vertical line. However, the drum is rotated about its
axis by the cord D, which isconnected through a reducing motion E

to the piston F of the engine. The result is that the surface of the
drum C moves horizontally under the pencil while the pencil moves
vertically over the surface, and a plot of pressure upon the piston
vs. piston travel is obtained.

Since an engine normally operates in a succession of essentially

Dtrtdton of tracing of diogrom


when cylinder volumt increom
0$ Indicated
PRESSURE

VOLUME
PISTON TRAVEL
Fig. 2-10. Indicator diagram.

identicalmachine cycles, the indicator diagram wiU be a


closed curve
such as shown in Fig. 2-10. It will
be observed that the diagram
OW8 only relative pressures and volumes,
there being no reference
or zero axes. In practice it is
customary to provide a pressure
reference line at 1 atmosphere by
connecting the indicator tb the
M] WORK 19

atmosphere and drawing a line at constant pressure of 1 atmosphere


before taking the diagram. Figure 2-11 shows the diagram of Fig.
2-10 superimposed upon the atmospheric line.
The ordinary use of the indicator diagram is to obtain the net
work done upon the piston during a complete cycle of the machine.
The work done upon the piston F is given by TT = y*pa dl where p

Fig. 2-11.

is the absolute pressure of the fluid, a is the area of the piston, and
I is the distance travelled by the piston. The indicator diagram is

a plot of p vs. 1] a is constant. Therefore the area between the zero


pressure lineand any line on the diagram represents to a certain
work done upon the piston while that line was being drawn.
scale the
To obtain the work of a complete cycle, however, it is necessary to
know only relative pressures, not absolute pressures. For every ele-
ment of piston travel dl\ in the positive direction there is a corre-

Fig. 2-12.

/mz
20 WORK [24 »

sponding element of travel dk in the negative direction as shown in


Fig. 2-12, The net work for the two elements of travel is the element
of work dW.
dW * PjO dli + P20 dfe
but dZj = — dll and dli = dl

hence dW = (pi — p2)o dl

The work for a complete cycle of the machine is then

W = fj(pi - Pi)a dl

which may be written

W=
where Pm (called mean effective pressure) is the average value of
(pi — Pi) over the length of the diagram, and I is the length of stroke
of the engine piston. The value of Pm is directly proportional to
the average height of the diagram.The average height of the dia-
gram is by measuring the area Oj of the diagram
usually obtained
with a planimeter and dividing the area by the length k of the
diagram. Then

where C, the indicator spring constant, is a constant of the engine


indicator, giving the scale relation between diagram height and pres-
sure. The dimensions of C are psi per in.

Exaicplb 3. An engine has a bore and stroke of 11 by 15 inches. An


indicator diagram taken from this engine has an area of 1.60 sq in., and
length of 2.40 in. The indicator spring constant is 80 psi per in. How
much net work was done by the f uid in the cylinder upon the engine piston
during the engine cycle represented by the diagram?

Solution; Bore means cylinder (and piston) diameter; stroke means


piston travel (distance between extreme positions).

Piston area: a = — = 95 sq in.


4

Piston travel; L = ~= 1.25 ft


2-6] WORK 21

1 60
Mean effective pressure: P* = 80 = 53^ pm
Net work: W =« P»aL = 6,360 ft lb

An indicator diagram is similar to a pressure-volume plot but it

does not necessarily represent all the work done by a fluid system,
even if the fluid is frictionless. The diagram represents only the
work done by the fluid on the engine piston. This work will not, in
general, be the same as the J’pdV for the fluid because in most
engines the fluid flows in and out at certain times in the engine
cycle, through the valves G in Fig. 2-9. Hence the fluid in the
engine does not constitute a stationary system and its work is not
given by J'p dV, nor does the indicator diagram represent a series
of states of a fluid system. The indicator diagram, in fact, represents
the series of states of the engine piston by giving its surface pressure
and its position. Indicator diagram characteristics for various types
of apparatus will be discussed in detail in connection with the analysis
of the particular apparatus.
2-6 Summary. Definition of Work: Work is transferred from a
system by a given action of the system if the total external effect of the
given action can be reduced to the raising of a weight. From this defini-
tion it is clear that a name for energy in transit from one sys-
work is

tem to another; energy which does not cross the boundary of a system
does not qualify as work with respect to that system. In order to
compute a particular work quantity it is necessary to select a system
having a boundary across which the particular quantity flows. Work
in thermodynamics may include not only mechanical work but also
electrical work.
Work transferred out of a system is positive with respect to that
system. Work transferred m is negative. Since the definition tells

only how to identify positive work, it is necessary to identify the


negative work of a system by showing it to be positive work of some
other system. Any work transfer may appear positive or negative
depending upon the system chosen.
In a stationary fluid system (changes of volume being permitted)
work may be transferred in or out by pressure force against a moving
boundary. If pressure is uniform in the system, this work is given
by W= XpdV and is positive or negative as dF is positive or
negative. (Only positive pressures are considered.) If the station-
ary fluid syston consists of a real fluid (having friction), it will be
22 WORK
possible to transferwork to the system (negative work) by boundary
friction forces or by paddle-wheel action. Such work cannot be
measured in terms of properties of the fluid system; some additional
information is necessary to evaluate it. If the stationary fluid sys-

tem consists of an ideal frictionle.ss fluid, the work of pressure forces

is the only work that will be transferred (assuming


gravity, electricity,

magnetism, and surface tension have negligible influence upon the


system).
A plot of a proce&s in a stationary fluid system may be used to
represent graphically the work given by J'p dV .

An indicator diagram is similar to a pressure-volume plot but the


indicator diagram is not a state diagram showing the pressure and
volume of a fluid system; it is a plot of pressure vs. piston travel and
as such may represent work done upon the piston by the fluid. This
work, however, will not equal fp dV for the fluid except in the spe-
cial case that no fluid flows into or out of the engine cylinder during
the engine cycle.

I'ROBLF.MS
A mass of 2000 lb is sus])(‘nd(*d from a pulley block. Find the ma^ni-
2-1.
tud(‘and direction of the work transfer wlion th(‘ masi? is hoisted against gravity
through 0.5 ft. Show in a sketch tin* system boundary across which the work is
transf*‘rn'd.
2-2. A pump forces 550 gallons per minute of water horizontally from an
0 ])on well to a closed tank wIktc tiu? pressure is 120 psig. Th(^ waiter tempera-
ture is 05^1'\ How much work must llu* pump do upon the water in an liour,
solely to force the water into tin* lank against the pressure?
Sketch the system
upon which the work is done, showing it both before and after the process.
2-3. If the work clone in Problem 2-2 upon the water had boon used solely
to raise the same amount of water vertically against gravity without change of
pressure, how many
feet would the water have been elevated?
2-4. (a) If thework done in Problem 2-2 upon tlie water had been used
solely to accelerate the water from zero velocity witViout change of pressure or
elevation, w'hat velocity would the water have reached? (b) If the work had
been used to accelerate the w’atcr from an initial velocity of 30 fps what would
have been the final velocity?
2-5. Gas from a bottle of compressed helium is used to inflate a balloon,
originally folded completely flat, to a volume of 8.3 cu ft. If the barometer is
30.13 in. Hg how much work is done upon the atmosphere by the balloon?
Sketch the system before and after the process.
2-6. The balloon of Problem 2-5 requires no stretching to reach the specified
volume: the pressure of the helium in the balloon is negligibly higher than the
atmospheric pressure. The helium in the steel bottle is initially at 2000 psia
WORK 23

and finally at 1925 psia. Taking the total masB of hdium as a system sketch
the system before and after the process, find the magnitude and sign of the system
work. Why can the work be calculated by fpdv when the pressure is not
uniform over the whole boundary of the system? Is the p used in calculating
the work the pressure of the system during the process? Does the system ha.ve a
definite pressure?
2-7. In a frictionless cylinder and piston machine, Fig. 2-2, the piston is
forced against the gas by a spring which exerts a force directly proportional to
the volume of the gas. In addition to the spring force the atmospheric pressure
of 1 5 psia acts upon the outer side of the piston, (a) Considering the gas as a
system find the work when, from an initial state of 1 cu ft, 30 psia, the gas vol-
ume increases to 3 cu ft. Plot the state change on the pV plane, (b) Consider-

ing the spring as a system find the work for the same process, (c) Account for

the difference in magnitude of the work in (a) and (b).


2-8. A propeller shaft in a ship turns steadily at 200 rpm. The torque ap-
plied to the shaft by the driving turbines is 790,000 ft lb. The torque transmitted
to the propeller is 780,000 ft lb, the remainder being used to overcome friction
at the shaft bearings, At what rate in horsepower does the shaft receive
(a)

work from the turbines? (b) At what rate does the shaft deliver work to the
propeller? (c) Considering the shaft as a system, what is the net rate at which

work crosses the boundaries of the system? Sketch the system.


2-9. A stationary mass of gas is compressed in a frictionless way from 12 psia,
3 cu ft, to 60 psia, 0.8 cu ft. Assuming that the pressure and volume are related
by py" = constant, find the work done by the gas system.
2-10. A mass of 1 lb of air is to be compressed from 15 psia to 110 psia in a
process for which pv = constant; the initial density of the air is 0.075 Ib/cu ft.

Find the work done by a piston, to compress the air.

;^2-ll. The indicator diagram for the process in a water pump is a rectangle
3.00 in. long (on the axis of piston travel) and 1.45 in. high; the indicator spring
scale is 100 psi/in. The pump process is repeated 60 times per minute. The
pump cylinder diameter and the piston stroke is 12 in. Find the rate in
is 8 in.

horsepower at which the piston does work upon the water.


2-12. A series of indicator diagrams for a diesel engine cylinder yields the fol-
lowing average results: area of diagram 0.232 sq in., length of diagram 2.10 in.
The indicator spring constant is speed is 295 rpm, the
600 psl/in., the engine
engine cylinder bore and stroke are respectively 10 and 18 in., and the engine in.

takes two shaft revolutions to complete each machine cycle. Find the horse-
power transferred from the gas to the piston (indicated horsepower).
2-13. A piston and cylinder machine containing a fluid system has a stirring
device in the cylinder, Fig. 2-3; the piston is frictionless, and the force F holding
it against the fluid system is due only to the atmospheric pressure, 14.7 psia.
The stirring device is turned 10,000 revolutions with an average torque against
the fluid of 10 in. lb. Meanwhile the piston 2.0 ft in diameter moves out 2.0 ft.

Find the net work for the system.


Chapter 3

TEMPERATURE AND HEAT

3-1 Temperature. The concept of temperature comes originally

from the reactions of the senses to “hot” and “cold” objects; a body
which feels hotter than another is considered to have higher tempera-
ture. Such physiological reactions are too crude and inconsistent for
quantitative measurement. Experience shows, however, that certmn
measurable properties of substances are related to the temperature.
For example, a bar of metal is usually found to increase in length
when it changes from cold to hot to the touch a quantity of mercury
;

enclosed in a glass tube may expand in volume taster than the glass

and fill a greater length of the tube when it changes from cold to hot;
the electrical resistivity of a metallic conductor usually increases as
the conductor becomes hotter to the senses. A measurement of
length, volume, or 'electrical resistivity might therefore serve as a
measurement of the temperature of a body, but such arbitrary tem-
peratures would apply only to the particular body whose length,
volume, or resistivity had been measured.
3-2Temperature equilibrium. It is a matter of common ex-
perience that when a hot body is brought into close contact with a
cold body the difference in temperature between the two bodies dimin-
ishes with time. If the two bodies are separated from all else so that
each is affected only by the other, they will finally reach a state in
which no further change occurs. The two bodies are then smd to be
in temperature equilibrium. The manner of approaching temperature
equilibrium depends upon the natures of the bodies involved. Usu-
ally the hot body becomes
cooler and the cold body warmer, as when
a piece of hot irondropped into a bucket of water to cool off. If,
is

however, the iron were dropped into the middle of a lake, the change
would appear to be confined to the iron; it could be said that for all
practical purposes the iron came to the temperature of the lake and
24
3-3] TEMPERATURE AND HEAT 25

the lake wae unaffected. This situation suggests the possibility of


comparing the temperatures of different bodies by using a comparison
body (or thermometer) small enough to have a negligible effect upon
the bodies with which it is brought into contact.
3-3 Thermometers. Suppose a mass of mercury is placed in a
glass bulb with a long narrow extension so that the mercury fills the
bulb and part of the extension or stem, Fig. 3-1. The level of the
mercury in the stem then becomes a sensitive indication of the mer-
cury volume relative to the glass volume. If an arbitrary scale is
marked on the stem, the result is a device which can indicate its own
temperature in terms of volume.
This is the mercury-in-glass ther- "ta
mometer, famiUar to everyone. 3.1 Mercury-in-glass
Now if this thermometer is placed thermometer,
in close contact with a body and
isolated from everything else, the thermometer will eventually reach
temperature equilibrium with the body and the reading of the ther-
mometer can be taken as an indication of the temperature of the
body. Three practical points should be noted here: (1) The prob-
lem of isolating the thermometer from everything but one body can-
not always be satisfactorily solved, but the general method is to
immerse the thermometer in a hole in a solid body, or directly in
a fluid body. (2) The thermometer will give the temperature of
both itself and the body after temperature equilibrium is reached,
but this may not be the same temperature as the body had before
the immersion of the thermometer. The thermometer should be
small relative to the body so that the thermometer may have only a
small effect upon the body. (3) The thermometer must not be sub-
ject to effects such as pressure changes, which might change the vol-
ume independently of temperature.
Assume that a thermometer is brought to equilibrium with each
of two bodies in succession and that the thermometer gives the same
reading with each body. Now if the two bodies are placed together,
Experience shows that no change due to temperature difference will
occur in them. That is to say: Two bodies each in temperature equi-
librium with a third body will be in temperature equilibrium with each
other* This means that an arbitrarily marked thermometer could
be used to compare the temperatures of bodies not themselves in
* This statement is sometimes called the Zeroth Law of Thermodynamics.
26 TEMPERATURE AND HEAT [3-4

contact. Also, any thennometer could be calibrated by comparison


with an arbitrary standard and would then show the same tempera-
ture for a given body as would the standard thermometer.
3-4 Temperature scales. Among the physical characteristics
which are found to be related to the temperature of a substance are
the changes from solid to liquid, and from liquid to gas. For exam-
ple, if an arbitrarily calibrated thermometer
is used as a measuring

device it is found that pure-water


always melts into liquid water
ice
at a definite temperature, and that pure water always boils into steam
at a definite temperature, provided the pressure is kept at a given
value. Therefore two points on a thermometer scale can be defined
in terms of the melting point and boiling point of pure water at one
atmosphere pressure, and these points will be reproducible in any
laboratory. On the Centigrade scale the melting point, or ice point,
is marked 0 degrees and the boiling point, or steam point, is marked
100 degrees. On the Fahrenheit scale the ice point is marked 32
degrees and the steam point marked 212 degrees. On both scales
is

the intermediate points are obtained by dividing the distance be-


tween the ice point and the steam point into equal subdivisions of
scale length. With a mercury thermometer the temperature that
would cause the mercury to reach a point midway between the
marks
at 32“F and 212*^ would be called 122®F. With a metal bar ther-
mometer the temperature that would bring the bar to a length
mid-
way between its length at 32°F and its length at 212‘’F would
be
called 122 F. With an electrical resistance thermometer the tem-
perature that would bring the resistance of the
conductor to a value
midway between the resistance corresponding to 32‘’F
and the resist-
ance corresponding to 212°F would be called
122®F. Now all these
types of thermometers, of whatever
materials they may be con-
structed,must agree at 32“F and at 212‘>F by the definition
of their
scales, but if two thermometers of different types or different mate-
rials, calibrated as described above, are
brought to equilibrium in a
place where one of them reads 122»F, the other will, in general, read
something different from 122‘’F It is an experimental fact that dif-
.

ferent materials have different


relations between expansion and tem-
perature change, such that a
temperature scale based upon the
^pa^on of one material will not agree at all points
with a scale
bas^ u^n the expansion of another
material. SimUar discrepancies
exist with respect to electrical thermometere. Furthermore, at cer-
3-6 ] TEMPERATURE AND HEAT 27

tain temperature levels one scale may even show a reversal of direc-
tion as compared with some other scale. Such characteristics limit
the choice of a scale for a basic standard of temperature.
3-5 Gas thermometers. There exists one group of substances
which exhibit excellent agreement among themselves, over wide
ranges of temperature, when used as thermometric substances.
These are the gases such as hydrogen, nitrogen, oxygen, helium,
which are difficult to condense to liquids. (Such gases were once
called “permanent gases” in the belief that they had no liquid or
solid states.)

A gas thermometera mass of gas in a bulb fitted with means


is

for measuring pressure and volume as shown schematically in Fig.


3-2. The device may be used in either of two ways. In the constant-
pressure gas thermometer the mercuiy is adjusted to maintain the
pressure constant in the gas, and the level of the mercury at « is a
measure of the gas volume which is taken as the indication of tem-
perature. In the constant-volume gas thermometer the mercury is
adjusted to keep the level z always at a fixed reference point, and the
pressure required on the gas to do this is taken as the indication of
temperature.
3-6 Absolute temperature. The gas thermometer has the fol-
lowing important characteristics:
(a) Both constant-volume and constant-pressure thermometers
using permanent gases, when individually calibrated at the ice point
and the steam point, show good agreement among themselves at all
28 TEMPERATURE AND HEAT [3-6

other temperatures not close to the condensation or dissociation tem-


peratures of the gases used. The agreement among different gases
becomes better as the pressure of the gases is reduced. By operating
a constant-pressure thermometer at several pressures and extrapolat-
ing to find the result that would be obtained at zero pressure, the
same results are obtained for all gases. Temperatures so found may
be called ideal gas temperatures, being based upon an ideal pressure-
less gas scale.

(b) If the scale of


a constant-pressure gas thermometer is extra-
polated beyond the ice point to lower temperatures,
there must even-
tually be a point reached at which the
extrapolated gas volume
become zero. This point found to correspond closely to -460*F
is
(-273‘’C) for all “permanent" gases. The extrapolation is indicated
^aplucally in Mg. 3-3. It may be expressed algebraically as follows
(for the Fahrenheit scale):

32-<o = (212-32):;^-i (a)


3-6] TEMPERATURE AND HEAT 29

The volume been found experimentally to be 1.366 for the


ratio has

ideal gas thermometer. This makes the temperature corresponding


to zero volume, <o, equal —459.7*^.
(c) If the scale of any constant-volume gas thermometer is simi-
larly extrapolated to zero pressure, the same temperature ( — 459.7°F)
is closely reached.
These facts in themselves would be sufficient to justify the accept-

ance of the ideal gas thermometer scale as an arbitrary standard.


But it is also a fact that this ideal scale, approached as a limit by
the real gas scales, is the same as a thermodynamic temperature scale
which will be introduced as a consequence of the Second Law' of
Thermodynamics. Until the latter scale is introduced in Chap. 8,

the ideal gas scale will be accepted as an arbitrary standard of tem-


perature, called absolute temperature, and defined as follows: (1) The
ratio of steam-point temperature to ice-point temperature is the same
as the ratio of the corresponding volumes of gas in a constant-pressure
ideal gas thermometer. (2) The difference between the steam-point
temperature and the ice-point temperature is 180 degrees for the
Fahrenheit absolute scale or 100 degrees for the Centigrade absolute
scale.*
The relations between the absolute temperature scales and the
conventional scales are given w'ith sufficient accuracy for general
purposes by
yF.b.
= f + 400 (b)

and = f + 27Z (0

where T®'*'* = Fahrenheit absolute temperature;

t^ = Fahrenheit temperature;

= Centigrade absolute temperature;

= Centigrade temperature.

It is customary to use capital T as a symbol for absolute temperatures

and small t as a symbol for other temperatures. The Fahrenheit


absolute scale is frequently called the Rankine scale (symbol R), and

* The International Scale of Temperature is a working scale which gives abso-


lute temperatures in terms of more practicable devices than the gas thermometer.
See the references at the end of this chapter.
30 TEMPERATURE AND HEAT [3-7

the Centigrade absolute scale is called the


Kelvin scale (symbol K).
Conversion equations are as follows:

= 1.8^^ + 32 (d)

(e)

3-7 Heat. It was pointed out in Sec. 3-2 that whenever two
bodies at different temperatures are brought into close contact with
each other and are isolated from other bodies, the temperature differ-
ence between the two bodies will diminish with time. If this happens
it follows that at least one of the bodies must change its temperature.
not necessarily true that both bodies must change temperature.
It is

For example, a piece of hot iron dropped into a bucket of ice and
water at 32®F may be finally cooled to 32°F while the mixture of
ice and water is still found to be at 32°F. This does not mean, how-
ever, that no change occurred in the water-ice mixture; in fact it will
be found that some of the ice has melted to liquid. A corresponding
situation exists in connection with the vaporization of a liquid. Hot
iron dropped into a bucket of water at 212°F will not raise the tem-
perature of the water finall}^, but will cause some of the water to
evaporate at 212°!^' while the iron cools to 2]2°1<\
It is known from experiment that the temperature changes or
other effects of bringing two bodies to tcmperatUi*e equilibrium are
quantitatively reproducible. For example, 1 lb of iron at 70°F
dropped into 1 lb of water at GO^'F will come to equilibrium at 60.9°F;
the water will rise in temperature 0.9 degrees while the iron will fall

9.1 degrees. One lb of iron at 70°!'


dropped into a mixture of ice
and water at 32°F will come to equilibrium at 32°F, but 0.0268 lb
of ice will melt, assuming at least that much ice was present initially.
Identical effects in the water and ice can be produced by materials
other than iron; identical effects in the iron can be produced by
materials other than water or ice. These faefs may be summarized
as follows:When t^vo bodies at different temperatures are placed in
contact with each other and isolated from other bodies, there will
always be a change in temperature or physical structure in both
bodies. These changes are always quantitatively related and, in
the case of two temperature changes, are in opposite directions. The
magnitudes of the effects produced depend upon the material and
the mass of each body and upon the magnitudes of their initial
temperatures.
8-7] TEMPERATURE AND HEAT 31

The conclusion drawn from these facts is that energy transfer


occurs when two bodies at different temperatures are placed in con-
tact. This energy transfer is not a work transfer because experience
shows that it cannot, even with the aid of frictionless mechanisms,
have the sole effect of raising a weight. Hence a new term, heat, is
defined as follows: HeoA is energy transferred from one system to another
solely by reason of a temperature difference between the systems*
The transfer of heat between two bodies in direct contact is called
conduction. Heat may also be transferred between bodies separated
by empty space through the mechanism of radiation. Radiation
may travel not only through empty space but also through gases and
some liquids and solids. A third method of heat transfer between
two bodies is fluid convection, which is a combination of three opera-
tions: (1) conduction from one body to the fluid; (2) motion of the
fluid from one body to the other; (3) conduction from the fluid to

the second body. Regardless of the mechanism of heat transfer in-


volved, the effect produced is always the result of a temperature
difference between two systems.
The direction of heat transfer is taken to be from the high-
temperature ^stem to the low-temperature system. The algebraic
sign relative to a given system is positive for heat flow to that ^rstem.
The quantity of heat transferred in a given operation is measured by
the mass of water that could be raised through a specified tempera-
ture interval, if the water were substituted for the low-temperature
system and the identical operation were executed in the high-
temperature system. (The water might also be substituted for the
high-temperature system while the given operation is executed in
the low-temperature system. The water would then fall through
the specified temperature interval.)
To show that heat is transferred in a given process, it is necessary
to show that a physical effect is the consequence of a temperature
difference between two systems. This is necessary because a par-
ticular effect could be obtained perhaps by other means than heat
transfer; thus a rise in temperature of a body of water may be
* Some people feel that heat, like work, should be defined independently of
the concept of energy, because energy has not been defined in a rigorous way.
A definition of heat which does not use the word energy follows: When an effect
in a system occurs solely as a result of a temperature difference between the system
and some other system, the process in which the effect occurs shall he caUed a transfer
of heat from the system at the higher temperature to the system at the lower tem-
perature.
32 TEMPERATURE AND HEAT [3-8

obtuned by txaasferring paddle-wheel work to the water. Conse-


quently the fact that the temperature rises is no guarantee that heat

transfer has occurred. On the other hand it is true that when bodies
at different temperatures are placed in contact, or are placed so thaf .

radiation can travel between them, heat will always be transferred.


The essential relation between heat and temperature is that a tem-
perature difference between two ^sterns is necessaiy for heat to be
transferred. The fact that the quantity of heat is measured in terms
of the temperature rise of water is absolutely arbitrary. It could
just as well have been said that the quantity of heat is measured by
the mass of ice that could be melted. In that case there would have
been no temperature change involved in the definition of the quantity
of heat.
A process in which no heat crosses the system boundary in either
direction is called an adiabatic process.
3-8 Units of heat. The unit of heat was formerly defined as
the
amount of heat required to cause a unit rise in temperature of a unit
mass of water at atmospheric pressure. One British thermal unit
(Btu) would raise the temperature of lb of water 1
1 deg F; one
calorie fcal) would raise the temperature of 1 gm of water 1 deg C.
Experience showed, however, that these units, being functions
of the
properties of water, varied in size depending
upon the initial tem-
perature of the water. A
new definition of the unit of heat, made
possible through the First Law of Thermodynamics,
is given in terms
of basic electrical energy units (see Sec.
5-3). The magnitude of the
unit is not changed by the new definition.
3-9 Specific heats. For small temperature
changes, experiment
show that the temperature rise of liquid
water due to heat transfer
to the water is given by

Q = mc{Ti - Ti) (1)


where Q is the heat transferred to the water,
Btu;
ffi is the mass of water, lb;

?i ~ Ti is the temperature rise of the water,


deg F;
c is an experimental factor, Btu/lb F, called specific heat
TJe spe^c heat is found to be a function of temperature
so
c-q. (1) must be rewritten for generality as
3-9] TEMPERATURE AND HEAT 33

dQ = me dT (2)

or Q=fmcdT (3)

Equation (2) is not limited to use with water, but may be written
forany homogeneous system in which a temperature change results
from heat transfer in the absence of work other than p dv work* In
general the value of c will depend upon the substance in the qrstem,
the type of state change involved, and the particular state of the
system at the time of transferring the element dQ of heat. For
liquids and solids the value of c does not differ much for different
processes but does change appreciably with temperature. For gases
the value of c differs greatly with different processes and also varies
appreciably with temperature; variation with pressure is generally
small. Because of the above facts the specific heats for solids and
liquids are often quoted without reference to the type of process in
which the heat transfer occurs, whereas for gaseous substances it is

customary to quote specific heats for two particular types of process,


the constant-pressure process and the constant-volume process. The
symbols for these two quantities are respectively c, and c«. In all

cases the specific heat may be given either as a single numerical value,
appropriate to a certain range of temperature, or as an equation in
terms of temperature.!
From the old definitions of the Btu and the calorie (Sec. 3-8) it

is evident that the specific heat of water would, by definition, be unity


in both British and metric systems of units. Consequently the spe-
cific any substance would be identical, numerically, in both
heats of
systems of units. In making the present definitions of heat units,
the relative magnitudes of the units were preserved so that specific
heats have the same numerical values whether in Btu/Ib F or
cal/gm C. Moreover, the specific heat of water remains unity (with
a small variation) at temperatures between 32®r and 212®F.
The use of specific heat data in heat transfer problems is illus-
trated in the following examples.

Example 1. It is desired to cool iron parts from SOO^F to 100°F by


dropping them into .water initially at TS'F. The specific heat of the iron
* This restrictionnecessary because in some cases the effects of heat and
is
work are indistinguishable, and the temperature change is therefore not neces-
sarily determined by the heat transferred.
t For values of specific heats of various substances see the Appendix.
34 TEMPERATURE AND HEAT 3-9]

IB 0.120 Btu/lb F and the specific heat of water may be assumed to be


1.00 Btu/lb F. Assuming that
the heat from the iron goes to the water
all

and that none of the water evaporates, how many pounds of water are
needed per pound of iron?

Solution: The iron and water come to temperature equilibrium at 100®F.


Therefore the heat flow to the water is

Quf “ TfluCw (Tsio

= m«, 1.00[100 + 460 - (75 + 460)]


= m«, 1.00(100 - 75)
The heat flow from 1 lb of iron is

-Qt * - m,Ct{T - Tu) 2i

* -(1.0)(0.120)[100 + 460 - (500 + 460)]


= -(1.0) (0.120) (100 - 500)

Note the negative sign for heat flow from the iron. Note
also that when
temperature differences are involved it is possible to use
either Fahrenheit
temperatures or absolute temperatures because the 460
subtracts out. The
heat flow to the water is equal to the heat flow from the iron:

Qw * -0, = -(1.0)(0.120)(100 - 500)


Qw “ +48.0 Btu/lb iron cooled
The pounds of water required are

“ 1.0(1(W - “ water/lb iron.


75) If
Example 2. The specific heat of a certain gas at constant
pressure is
given in Btu/lb F by
Cp = 0.248 + 0.448 X lO"* P
where T is in Fahrenheit absolute degrees.
How much heat must be trans-
ferred from a system consisting of 5 lb of
this gas to cool it at constant
pressure from 1540“F to 540‘’F7

Solution: The heat transferred to the system is given by

r») dT
3-10 ] TEMPERATURE AND HEAT 35

-ir,
m» 0
0.2483’
.:
0.448 X 10~*

Jri
r, 1540 -1- 460 = 2000«
Ti 540 -I- 460 = 1000®

- 0.448 X 10-* ( 1000^ - 2000®^


Q 5^0.248(1000 2000) -|-

^
Q 5[-248 -I- (0.448/3)(-70)] = -1290 Btu

Q is the heat transferred to the system. The heat transferred from the
system is then 1290 Btu. The numerical work is shown in detail so that
the signs can be kept right. This is very important in more complicated
problems since the sign of the result is not always self-evident.

3-10 Latent heats. When the result of heat transfer is a change


in physical structure of a substance, instead of a change in tempera-
ture, the quantity of heat in the absence of work other than p dv work*
is a function of the quantity of substance changed from one form to
the other. For example, when, by heat transfer, ice is melted to
water at 1 atm pressure, approximately 143 Btu will flow in for each
pound of ice melted; the same quantity of heat will flow out for each
pound frozen at 1 atm pressure. When, by heat transfer, liquid
water boils into vapor at 1 atm pressure, approximately 970 Btu
will flow in for each pound of water vaporized; the same quantity

of heat will flow out for each pound condensed at 1 atm pressure.
Such heat quantities are called latent heats. Latent heat related to
the melting of a solid is called heai of fusion; latent heat related to
the vaporization of a liquid is called heat of vaporization. Other latent
heats exist related to sublimation (vaporization of a solid) and to
changes in crystalline structure of solids.

The latent heat of fusion differs for various substances but is little
affected by pressure. The latent heat of vaporization is different for
various substances and is greatly affected by pressure. Latent heats
are always given for constant-pressure state changes (temperature is

also constant).

Latent heats and have other significance than that


specific heats
of a measure of heat transferred in a certain type of process; more
general definitions will be given in subsequent chapters, but for the

*'See first footnote, p. 33.


36 TEMPERATURE AND HEAT [3-10

present these quantities may be used in the computation of heat


transferred in the absence of work other than p dv work.
Example 3. Ten pounds of solid sulfur at ItfV are to be heated at con-
stant pressure \mtil it is a vapor at 1 atm pressure. How much heat will
be required?
Data for sulfur at 1 atmosphere

Melting point 235°F


Boiling point 832*F
Specific heat of solid 0.180 Btu/lb F
Specific heat of liquid 0.235 Btu/lb F
Latent heat of fusion 15.8 Btu/lb
Latent heat of vaporization 120 Btu/lb

Solution: Four separate heat quantities must be calculated: (a) to raise


solid temperature to 235°F; (b) to melt solid at 235‘'F; (c) to rmse liquid
temperature to 832°F; (d) to vaporize liquid at 832‘’F. (Change of solid
crystal forms is ignored.)

Q. - mc,{Ti - Ti) = (10)(0.180)(235 - 70) » 297 Btu

Qh = mL = (10)(15.8) = 158 Btu

Qc = mc,(T, - Ti) = (10)(0.235)(832 - 235) = 1403 Btu

(10)(120) = 1200 Btu

Q = + + + - 3058Btu
Example 4. A piece of ice having an initial temperature of 22“F is
dropped into an insulated tank, which contains 40 lb of water at 70'T.
If
the temperature of the water, after equilibrium is reached,
is 40T, how
many pounds did the ice weigh? Assume no heat transfer with other bodies
has occurred.
Data
Specific heat of ice 0.501 Btu/lb F
Specific heat of water 1.00 •Btu/lb F
Latent heat of fusion 143.3 Btu/lb
Melting point of ice 32'’F

^lution: Four separate heat quantities must be calculated: (a) to raise


^ *** ** 32®F; (c) to raise melted ice from
32 * to 40 F; (d) to lower water from 70‘’F to 40*F. Let the mass of ice
bettijlb.

0. - «<v(r, - Ti) - »m(0.501)(32 - 22) - 5.01m,


Btu
- mL - m,143.3 - 143.3m, Btu
Q, - »ui(r, - r,) - m,(1.00)(40 - 32) - 8.00m,
Btu
3-11] TEMPERATURE AND HEAT 37

Qi - mCfiTz - Ti) = (40)(1.00)(40 - 70) = -1200 Btu


Oa + 0* + Qe = —Qi
(6.01 + 143.3 + 8.00)m< = 1200

irn = = 7.68 lb of ice


loo.o

Suppose the water had been contained in a tank of copper, weighing


10 lb,which was in equilibrium with the water at the beginning and end of
the process. How much additional ice would have been required? Specific
heat of copper is 0.092 Btu/lb F.

Solution: An additional heat quantity must be calculated for the copper:


Q, = mc^arz - Ta) = (10)(0.092)(40 - 70) = -27.6 Btu
(5.01 + 143.3 + 8.00)m; = 27.6

^ 156.3
— 0.176 lb ice additional

3-11 Summsury. Temperature can tesensed directly but an ob-


jective measurement of temperature can be made only indirectly
through measurements of other properties such as volume, pressure,
or electrical resistivity.
Any group of bodies at different temperatures, if they could be
brought together and isolated from all other bodies, would eventually
reach a common temperature at which they would be in temperature
equilibrium.
A thermometer is a device calibrated to read its own temperature
in terms of some property such as volume and arranged to be con-
veniently brought to temperature equilibrium with other bodies,
while exerting a Ttiininnnn effect upon the temperature of the other
bodies.
Ordinary thermometer scales are arbitrary. The melting and
boiling temperatures of water at 1 atm pressure are assigned definite
values, and the scale is filled in by assuming a linear variation of
temperature with the reading of some arbitrarily selected standard
thermometer. The most suitable standard has been found to be the
gas thermometer, which is used in physical laboratories to determine
the reference points at which other thermometers are calibrated.
An absolute temperature scale may be constructed by taking the
temperature as directly proportional to the volume of the gas in a
constant pressure ideal gas thermometer. This scale, based upon a
gas at zero pressure, is the physical realization of a logicaUy formu-
38 TEMPERATURE AND HEAT [3-11

lated absolutethermodynamic temperature scale, which will be intro-


duced in connection with the Second Law of Thermodynamics. For
ordinary purposes absolute temperatures may be found from Fahren-
heit and Centigrade temperatures by the equations

rr**- = <*"
-f 460

+ 273
Heat energy transferred from a system at a given temperature
is

to a system at a lower temperature solely as a consequence of the


existence of a temperature difference between the two systems. The
quantity of heat is measured by the number of pounds of water that
could be raised through a given temperature interval by the given
heat flow. Heat flowing to a system is positive with respect to that
system.
The unit of heat is the British thermal unit (Btu), which was
formerly defined as the heat required to raise the temperature of a
pound of water 1 degree F at a specified temperature level. The
Btu is now defined in terms of electrical energy units but its magni-
tude is essentially the same as before.
Specific heat is the rate of heat flow per unit temperature rise

per unit mass, when the temperature of a mass is changed by heat


flow in the absence of work other than pdv work. Specific heats
have a broader significance which will be brought out in Sec. 5-6.
Specific heats for solids and liquids are not appreciably affected by
the type of process but do vary with temperature. Specific heats
for gases and vapors are affected by the type of process and are
therefore usually given for two particular types of process, the
constant-pressure process and the constant-volume process. Both
of these specific heats vary with temperature and, to a much smaller
extent, with pressure.
Latent heat is the heat flow, in the absence of work other than
p dv work, associated with changes such as melting or vaporization
in the physical structure of substances. It is measured in Btu per
unit mass of substance changed from one structure to the other, while
at constant pressure and temperature. Latent heat of fusion is the
heat required to melt a unit mass of solid substance at constant
pressure and temperature. Latent heat of vaporization is the heat
required to vaporize a unit mass of liquid at constant pressure and
TEMPERATURE AND HEAT 39

temperature. Heats of fusion are little affected by pressure change.


Heats of vaporization are greatly affected by pressure change.
w

PROBLEMS
3-1* A thermometer is by arranging an aluminum rod so that its
constructed
length can be measured accurately by a scale made of a nickel alloy, invar. The
thermometer is calibrated at the melting point and boiling point of water, and
the distance between these two points is divided uniformly into 100 degrees of
equal length on the scale. Find the difference between the reading of this ther-
mometer and the gas-thermometer temperature at 0, 25, 50, and 100®C on the
gas thermometer scale. The linear expansion of aluminum may be represented
by the equation
It = Zo(l + 0.2221 X 10“* f + 0.114 X 10“^ f)

where It is the length at t on the Centigrade gas thermometer scale, and k is


the length at 0°C. The coefficient of expansion of invar between 0®C and 100®C
is 0.9 X lO"* cm/cm deg C.
3-2. Two mercury-in-glass thermometers are made of identical materials and
are accurately calibrated at 0®C and 100®C. One has a tube of constant diam-
eter, while the other has a tube of conical bore, ten percent greater in diameter
at 100°C than at 0°C. Both thermometers have the length between 0 and 100
subdivided uniformly. What
will the straight bore thermometer read in a place

where the conical bore thermometer reads 50®C?


3-3. In a constant-volume gas thermometer the following pairs of pressure
readings were taken at the boiling point of water and at the boiling point of
sulfur, respectively:

Water bp: 50.0 100 200 300


Sulfur bp: 96.4 193 387 582

The numbers are the gas pressures, mm Hg, each pair being taken with the same
amount of gas in the thermometer, but the successive pairs being taken with
different amounts of gas in the thermometer. Plot the ratio of SbpiHaObp against
the reading at the water boiling point, and extrapolate the plot to zero pressure
at the water boiling point. This gives the ratio of Sbp:H20bp on a gas ther-
mometer operating at zero gas pressure, i.e. an ideal gas thermometer. What
is the boiling point of sulfur on the gas scale, from your plot?

3-4. In testing electrical apparatus the average temperature of a coil of wire


is by measuring the electrical resistance of the wire. The resist-
often obtained
ance of a certain magnet coil at room temperature (86'^F) was 1239 ohms; after

operation under test the resistance was 1433 ohms. The resistance of a copper
wire at temperature is /Z = Rm[1 + 0.00393(^ — 20)] where Rn is the resist-

ance at 20®C. Find the average temperature of the coil after test, in degrees F.
3-5. Experience shows that the time rate of heat transfer by conduction be-
tween two systems is closely proportional to the temperature difference between
the systems; the rate of heat transfer by radiation is proportional to the difference
between the fourth powers of the absolute temperatures of the systems. For
40 TEMPERATURE AND HEAT
both conduction and radiation, what value does the rate of heat transfer approach
as the system temperatures approach equality? (In later sections of this book
hypothetical heat transfer processes will be discussed in which the temperature
difference between the two systems is infinitesimal.) How long wduld it take
to transfer a finite quantity of heat with infinitesimal temperature difference?
3-6. (a) Explain why the temperature change of a system in a process is not
necessarily a measure of the heat transferred in the process, (b) Devise a process
in which there is heat transfer without temperature change in either of the sys-

tems involved, (c) Devise a process in which, for one system, the heat transfer
is of opposite sign from the temperature change.

3-7. The average specific heat of green vegetables is about 0.9 Btu/lb deg F.
A ton of vegetables at is placed in a cold room in which the air is main-

tained constantly at 40‘’F. The refrigeration plant can remove heat at a maxi-
mum rate of 400 Btu/min, but 150 Btu/min of this capacity is required to
compensate for heat leakage from the surroundings to the cold room. What
will be the minimum time in which the vegetables can be cooled to 40*’F? (The
and will depend upon the heat transfer rates that can
actual time will be longer,
be obtained; thermodynamics can determine only the limiting time, with perfect
heat transfer.)
3-8. Fifteen hundred pounds ofat 40^F are to be frozen and stored at
fish

10®F. The specific above freezing is 0.76 Btu/lb F, and below


heat of the fish

freezing is 0.41 Btu/lb F, The freezing point is 28°F and the latent heat of fusion
is 101 Btu/lb. How much heat must be removed to cool the fish, and what
percent of this is latent heat?
3-9. The heat of fusion of aluminum at 657®C is 94 cal/gm. If the average
specific heat of solid aluminum between 0®C and 667®C! is 0.247 cal/gm deg C,
and of liquid aluminum is 0.256 cal/gm deg C, how much heat will be required
to raise the temperature of 25 pounds of aluminum from 80°F to 1350®F?
Chapter 4

PROPERTIES OF SYSTEMS

4-1 Properties—physical concept and general concept. A


property of a system was defined in Sec. 14 as a characteristic by
which the state of a system may be described. The simple physical
concept of a property involves observable characteristics such as
pressure, temperature, mass, velocity, and location relative to the
earth. When all the observable characteristics have definite values,
the state of the system is definite; hence it is possible to define a
property in a more general way as any quarUity which has a definite

valw for each definite stale of the system. From


more general
this

viewpoint observable properties are not the only properties of a


system.
4-2 Properties mathematically derived. Given the observable
properties of a system, it is possible to form any number of mathe-
matical functions of these properties such that the value of the func-
tion will be fixed if the values of the observable properties are fixed.
Such functions then become properties of the system under the gen-
eral definition of a property. Consider, for example, the ratio of the
mass to the volume of a system. When the mass and the volume
are fixed, the ratio of mass to volume will have a definite value; it is

therefore a property of the system (it will be recognized as the den-


sity). As another example consider the product of pressure and
temperature. This quantity, chosen at random, has no phyacal
meaning, yet it is a property of a system. A number of property
functions which appear frequently in thermodynamic analysis, and
which are themselves properties, have been given names and symbols
as a matter of convenience. Some of these properties will be encoun-
tered in subsequent chapters.
4-3 Properties derived from general laws. In addition to ob-
41
42 PROPERTIES OF SYSTEMS [44

servable propertiesand mathematically derived properties, there exist


two properties which are derived from the two general laws of thermo>
dynamics. These properties, internal energy derived from the First
Law, and entropy derived from the Second Law, are not observable
properties nor can they be arrived at by mathematical manipulations
of other properties. They can be shown to be properties only through
the use of the general laws of thermod}mamics. It is not intended
to discuss these two properties at this point but only to indicate the
existence of this third class of properties.
4>4 Point functions and path functions. The general require-
ment for a property, that it have a definite value for any state of the

system, may
be expressed by saying that a property is a point func-
tion. This term refers to the fact that in a graphical representation
of properties a point on the plot will
correspond to a definite value
for each property. If a plot on the coordinates of pressure
and vol-
ume constructed to show the relation between pressure,
is
volume,
and temperature for a unit mass of a gas, the result
will look like
Fig. 4-1. At any point on the
such as A, there can be read off
plot,
definite values of pressure, volume, and temperature. It is only
necessary to mark the point A and the
properties at A are fixed. If
another point, such as B, is marked, the
properties at B are fixed.
Hence the name, point function, for a property.
It is often more important to know how much a property
changes
4^] PROPERTIES OF SYSTEMS 43

when a change of state occurs than to know its value at either end
state. The change in any property during a change of state from
AtoB would be given by the difference between the value at B and
the value at A. For example

= Tb — Ta
where ATab signifies the change in temperature during a change of
state from A to B, and Ta and Tb are, respectively, the temperatures
at states A and B. Similarly

AVab =Vb- Va

and APab = Pb — Pa
These relations would hold whether the change of state followed the
path X or the path y in Fig. 4-1. The change in the value of a property
during a change of stale is independent of the path of the change of slate
and depends only upon the end points.
Work is not a property. It will be recalled that in Chap. 2 com-
putations were made of the work done during certain processes. The
work done by a system through pressure against a moving boundary
was given by

B'. .£pdv
Now, referring to Fig. 4-1, be clear that the p dv work for the
it will

state change along the path x from A io B would be represented by


the area between the line AxB and the axis of volumes, whereas the
p dv work for the state change along the path y from A to B would
be represented by the area between Ihe line AyB and the axis of
volumes. It is obvious that these areas are not equal. It is then
obvious that the p dv work is entirely dependent upon the path fol-
lowed between A and B rather than upon the location of points
A and B. Hence the work is called a path function to distinguish it

from a property.
It was emphasized in Chap. 2 that work is not always related to
the properties of a system by any function. Now it is shown that
even when the work is related to the properties of the system there
is a fundamental difference between work and a property. Work is
determined by the path between two states and not by the end states;
44 PROPERTIES OF SYSTEMS [4^

property changes are determined by the end states and not by the
path.
Although it cannot be readily demonstrated at this point, an
it is

experimental fact that heat, like work, is a path function. Heat is


therefore not a property of a system.
4>5 Heat and temperature. It is well to point out here some
possible sources of confusion in connection with heat, which is not a
property, and temperature, which is a property.
The word heat has often been used by physicists to denote two
entirely different things. (1) The energy transferred between two
systems due to temperature difference; this is not a property. (2)
The and potential energy of the molecules and atoms of a
kinetic
substance; this is a property. This dual usage of the word is bad
enough in itself but the difficulty has been compounded in some cases
by defining heat as (2) and using it as (1). It has become generally
accepted in the field of thermodynamics that the word heat shall be
used solely to denote energy transfer due to a temperature difference.
This usage will be followed strictly in this book.
Temperature is frequently defined in physics as a measure of the
kinetic energy of molecules and atoms.
This definition, though per-
fectly proper in its place, cannot
be used in theimodjniamics because
the broad approach of thermodynamics does not take account
of the
detailed mechanisms of energy storage in molecules
and atoms. That
subject belongs to other branches of science which
are beyond the
scope of tlm book. We may recognize that energy is stored as poten-
tial and kinetic energy in molecules
and atoms but we are not in a
position to study these factors in detail. Instead, we utilize empirical
information in regard to the energy stored in
substances, and avoid
the problems of molecular and atomic
structure. The science of
molecular and atomic structure is, in fact,
of great aid to thermo-
dynamics in supplying useful data on the
properties of substances.
These data are determined on the basis of
a certain physical picture
or m^hanism for the internal structure
of a substance, but they are
used mfhermody^ics simply as empirical facts without reference
to the physi^ picture by which they
were computed. Similarly
temperature is accepted in thermodynamics
as a fact, without at-
tempting to explain the fact.
The use of a physical picture or mechanism may
often be a valu-
able aid m the formulation of a problem, but it may also lead to false
4^J PROPERTIES OF SYSTEMS 45

results if the picture is not well chosen. The great value of the
thermodynamic method (and its great limitation) lies in its inde-

pendence of any particular picture of the mechanism of energy


transfer and energy storage in systems.
Chapter 5

THE FIRST LAW OF THERMODYNAMICS

5-1 Background of the first law. Work and heat, which have
already been discussed separately, are related by the First Law of
Thermodynamics. This law is a statement of the general principle
of the conservation of energy. The hypothesis of energy conservation
was developed by a number of scientists during the first half of the
nineteenth century but direct experimental tests of the hypothesis
were firstmade by J. P. Joule in the eighteen-forties.
The early theories of heatwere based on the assumption that heat
was a fluid which could be stored in substances and transferred from
one substance to another. Some phenomena can be explained on this
basis. For example, a certain amount of heat transferred into a sys-
tem a mass of water at constant pressure will raise the
consisting of
temperature of the water from h to fe; if then the same amount of
heat is transferred out of the system, the temperature will fall from
k to ^j. This might lead to the theory that the heat supplied is stored
in the system at k and released when the temperature
returns to ti.
However, when this idea is put to the test of general experience it is
found that all the heat supplied to a system during a change
of state
need not always reappear when the i^stem is returned to
its initial
state.

It is also a matter of experience that the


work done on a system
during a change of state need not always reappear
when the system
is returned to its initial state; in
fact it often happens that nme of
the work done in a state change reappears
when the system is re-
turned to its initial state.
Inall cases, however, when a system
is caused to execute a cycle,
it isfound that a failure to recover all the heal
supplied is accom-
panied by the transfer of more work from the
system than was put
in during the cycle. Also, a failure to recover all the work supplied

46
6-2 ] THE FIRST LAW 47

is accompanied by the transfer of more heat from the ^stem than was
put in during the cycle. This indicates that neither heat nor work
is conserved as such but that some means exists whereby a system

working in a cycle can take in one of these entities and give out the
other.
The facts stated above lead to the following hypothesis: Heat and
work are different forms of a single entity which is conserved. The
name energy is applied to this entity. On the basis of this hypothesis,
energy which enters a system as heat may leave the system as work,
or energy which enters the system as work may leave as heat; con-
sequently the energy in storage in the system is neither work nor heat

but requires another name. The name internal energy is used for
energy in storage in a system.
5-2 Equality of heat and work in a cycle. In order to test
the hypothesis of energy it is necessary first to determine whether a
quantitative relation exists between work and heat in cases where
they may produce identical effects upon a system; and second, using
this quantitative relation, to determine whether the net energy trans-
ferred across the boundary of a system during a cycle of operations
is zero. A simple experiment for this purpose might be carried out
as follows:
Referring to Fig. 5-1, a mass of water is placed in an insulated
tank containing a thermometer and a paddle wheel. Taking the
48 THE FIRST LAW [M

water as a system, work can be transferred to the system 1^ the


wheel, and the quantity of work can be measured by the fall
of a weight which drives the paddle wheel.
Assume that an experiment starts with the system at state 1, in
which state the temperature is and the pressure is 1 atmosphere.
Previous experiments have shown that if the pressure and tempera-
ture of liquid water are fixed, all other observable properties will be
fixed. Therefore, if the pressure is kept constant, the temperature
will identify the state of the system completely. The experiment
consists of transferring a measured quantity of work TTi-a into the

system, and measuring the rise in temperature U — U that results.


Now it is known from experiments on heat transfer that the same
temperature rise could be produced in the given system by transfer^
r ing a definite amount of heat to the system in the absence of work.
Thus a connection is established between heat and work in terms of
a common effect.

Joule’s experiments of this general nature were made under vari-


ous conditions, sometimes using water and sometimes mercury as a
fluid system, and also using a solid system of metal blocks which
absorbed work by when rubbed against each other. Other
friction,

experiments involved the supplying of work in an electric current.


In every case the same ratio was found between the number of foot
pounds of work and the number of British thermal units of heat
that would produce identical effects in the system. which
This ratio,
was called the “mechanical equivalent of heat,” has the value 778
ft Ib/Btu. (Joule’s result was slightly smaller.)
By completing a cycle with the system a further result is obtained.
Assume that the system has been brought from state 1 to state 2 by
the suppl 3ring of work lFi_s. Now, with the paddle wheel at rest, a
second process is carried out by bringing the system into contact
with a cold body so that heat transfer will occur from the system.
The cold body has been calibrated so that by its temperature rise the
heat transferred from the system may be measured. When the tem-
perature of the system reaches U, the original value before the first
part of the experiment, the observable properties of the system are
allfound to have their original values. Hence the system has re-
turned to state 1, or has executed a cycle. This cycle consisted of
the tnmsfer of a definite amount of work Wi_s to the system followed
hy the transfer of an amount of heat Q*_j from the system. The result
64] THE FIRST LAW 49

of such an experiment is always the following: The work ITi-t trans-


ferred to the system is proportional to the heat Qt-i transferredfrom
the system, and the constant (tf proportionality is the so-called me-
chanical equivalent of heat. In the simple example given here there
are only two energy transfer quantities, but experience shows that
with any number of work quantities and heat quantities the same
result is found. Expressed algebraically

(2Tr)oyole ® •f(20)o3rcl* (1)

where (2Tr)«you is the algebraic sum of all the work transfers in' a
cycle, (ZQ),yeie is the algebraic sum of all the heat transfers in the
cycle, and J is the experimental constant of proportionality. When
work is in foot pounds and heat is in Btu, J has the value 778.16.
These experimental facts confirm the hypothesis of energy (Sec. 5-1).
Experience has shown that Eq. (1) holds for any cycle and any
system whatever. The cycle may involve chemical reactions, or
physical state changes, the only restriction being that no material
crosses the system boundary.
Equation (1) is the First Law of Thermodynamics. It is accepted
as a general law of nature because no violation of it has ever been
demonstrated. No independent general law of nature can ever be
proved except in this n^ative way because there are always new
experiments which might disprove the law; but when years of experi-
ence produce many experiments which confirm the law, and yet fail

to produce the single contrary experiment necessary to disprove it,

the law becomes as firmly established as any human knowledge can


be.*
5-3 Units of energy. Up to this point work and heat have been
measured in different units, each in terms of the physical picture by
which it was defined. But now the quantitative relation between
work and heat and the acceptance of the hypothesis that they are
both forms of energy, permit the use of a single unit for both. Heat
transfer quantities can be expressed in foot pounds by substituting
778 foot pounds for each Btu; work can be expressed in Btu by sub-
stituting 1/778 Btu for each foot pound. If either of these were

* No concern
need be felt about tiie effect upon this law of the recent discov-
the conversion of mass to onergy; mass simply becomes another form in
eries in
which energy may appear. In any event, within the scope of Engineering Ther-
modynamics, transitions of mass to energy are not encountered.
50 THE FIRST LAW [54

done, the constant of proportionality J would not be needed in


Eq. (1). From this point on
be assumed that equations in-
it will

volving work and heat are written with the same units for both work
and heat. Equation (1) will now be written as

(STF)eyel. = (SQ)cycle (2)

where the same units are used for both and Q. W


Since work to a system is negative and heat to a system is posi-
tive, Eq. (2) states that, for a cycle, the algebraic sum of all energy

transfers to a system is zero.


The fact that work and heat may be expressed in the same units
does not mean that work and heat are interchangeable in all opera-
tions, for they are not. Nevertheless, in all operations in which
work and heat are interchangeable, 778 ft lb of work will accomplish
the same effect as 1 Btu of heat. Therefore 1 Btu can represent a
perfectly definite amount of work, or 1 ft lb can represent a definite
amount of heat. It must always be remembered, however, that
changing the units of heat to foot pounds does not give heat the
characteristics of work. The basic definitions of work and heat are
unchanged, whatever the units of measurement.
It may be noted at this point that the definition of the Btu is no
longer based upon the properties of water but is arbitrarily set at a
certain ratio to the foot pound. Actually the calorie (Sec. 3-8) is
defined as 1/860 watt hour, and when these units are converted to
Btu and ft lb, the relation obtained is 1 Btu = 778.16 ft lb.
5-4 The first law and internal energy. Equation (1) is a state-
ment of the First Law of Thermodynamics but it is in an
inconvenient
form for general use, since it applies only to
cycles. If a system
executes a change of state which is not a cycle,
the summation of all
energy transfers across the boundary of the
system is not necessarily
zero. In accordance with the hypothesis of energy
conservation, a
net energy trailer to a system results
in an equal increase of internal
energy stored in the system. This may be written as follows:

Q = AE + W (3)
where Q is the heat transferred to the system during
the process;
W isthe work transferred from the system
during the process;
is the change in the internal
energy of the system during
the process, and all these terms are expressed in the
same units.
W] THE FIRST LAW 51

For convenience the equation is written in terms of a single heat


quantity and a single work quantity; if several heat quantities are
involved in the process, Q symbolizes the algebraic sum of these
quantities, and similarly W is the algebraic sum of the work quan-
tities.

Equation (3 ) is the usual statement of the First Law. It says


that in any change of state the heat supplied to a system is equal to
the increase of internal energy in the system plus the work done by
the system. This means that energy is conserved in the operation.
From this viewpoint the First Law is a particular formulation oi the
principle of the conservation of energy. Equation ( 3) may also be
considered the definition of internal energy. The equation does not
contain the internal energy E explicitly, but contains the change of
internal energy AE for the process. Hence the definition does not
give an absolute value for in-
ternal energy. Although no
absolute value can be ob-
tained for the internal energy of
a system, it is possible to show
that the internal energy has
a definite value at every state
of a system and is, therefore,
a property of the system.
Consider a system to be
initially in a state 1 which is

changed to state 2, as repre- Fig 5.2.


sented on the pressure-volume
plane by the points and 2 in Fig. 5-2
1 Let the change of state .

from I to 2 be represented by the path A, for which the following


relation will hold;

Qa = AEa + 11/1 (a)

Now complete a cycle by returning from 2 to 1 via a different path B.


Then for path 2?,

Qb = AEb + ll’a (b)

The combined processes A and B constitute a cycle for which the


sum of the work transfers must equal the sum of the heat transfers
according to the First Law, Eq. (2). Hence
52 THE FIRST LAW [5-5

Qa-\- Qb = + Wb
or Qa ~ Wa — Wb ~ Qb

Substituting from (a) and (b)

^Ea = - AEb (c)

Now suppose the system had returned from 2 to 1 by any other


path C. By the same reasoning,

^Ea — — AEc

then (d)

Hence the change of internal energy Affj-i is the same for any path
between 2 and 1. Assuming an arbitrary value assigned to the in-
ternal energy at 1, the value at any other point, 2, will depend solely
upon the location of the point 2 and not upon the path traversed in
going from 1 to 2. This shows that the internal energy E ib a point
function and therefore is a property of the system. (See Chap. 4
for the characteristics of a property.)
Since the internal energy is a property, if an arbitrary value is

assigned to it at any one state, it is possible to determine the values


at all other states by substitution in Eq. (3) of measured quantities
of heat and work. In subsequent discussion of the internal energy
it is to be understood that no absolute value is known but some
arbitrary base value may have been assigned. In many cases it is

unnecessary to establish a scale of internal energy since the change


of internal energj’^ AE satisfies the needs of the problem.
5-5 Internal energy —
general and special concepts. In the
discussion of internal energy no mention was made of the form in
which the energy might be stored in a system. In some cases this
might be known; for example, suppose a system to consist of a fly-
wheel as shown in Fig. 5-3. Work may be introduced into the system
by a torque in the rotating shaft. Assuming that friction is negli-
gible, the work input will go entirely into increasing the kinetic energy
of the system, as is known from mechanics. From the viewpoint of
thermodynamics, however, all that is said is that the internal energy
of the system has increased.
Another means of energy storage known from mechanics is the
potential energy of gravity. This may also be considered in thermo-
6-5] THE FIRST LAW 53

dynamics as simply a form of internal energy.* (If electromagnetic


or electrostatic potential energy were to be considered, they could
also be classified as internal energy.)

But there are other forms of energy in storage which cannot be


* When work is utilized to raise a weight, the work does not actually go into
internal energy of the weight. For, consider a system consisting of a weight
acted upon by gravity as shown in Fig. 5-a,
The system is acted upon by the forces F (**ex*

ternal force'') and Fg (“gravity force*'), which


are equal and opposite, while the system moves
from state 1 to state 2. State 2 is higher than
state 1
the system
by a distance

— TFi-a
is

“ Fz +
z. The work done upon

but Fg = ”F,
i7^
so Wi _2 = 0.

For the system comprising the weight,

Qi-2 = A Fi„2 -f* Qi_2 = 0.

Therefore AFi _2 * 0, and it is strictly incorrect


to say that the weight gains internal energy
when its elevation increases. Actually a system
comprising the weight and the earth gains inter-
nal energy. However, since all elevations are
measured a numerically
relative to the earth,
correct energy balance for the weight is ob-
tained by leaving out the work of the gravity force
Fg and including the gravity potential energy in
the equation. Since this is customary procedure,
it will be followed in this book.
54 THE FIRST LAW 5^]

accounted for in terms of the velocity of the system or its position


in the gravity field. A stationary system is able to store energy in
the form of kinetic and potential energy of the atoms and molecules
of the system. There is evidence of this energy in the temperature
and pressure of the is no simple way of measuring
system but there
it in detail. Here the concept of internal energy is of great value,
for without knowing anything about the details of the energy storage
it can still be completely accounted for in terms of the work anH heat

transfers experienced by the system. (Review Sec. 4-5 in this con-


nection.)
The internal energy represented by the symbol in Eq. (3) is a E
general internal energy which includes all the forms of energy in
storage in the system whether they are readily susceptible to detailed
measurement or not. The quantity E is a. property of a ^stem and
as such it is a definite function of the other properties of the system.
Since part of E may be kinetic energy associated with observable
velocities, E must depend upon theobservable motion of the system.
Since part of E may
be gravity potential energy, E must depend
upon the elevation of the system. These portions of E may, if
desired, and if sufficient data are known, be written in
detail and sub*
tracted from E leaving a quantity Eu. Then,

E= Ek -i- Ep Ev (4)

where Ek the kinetic energy associated with observable velocities,


is

Epthe gravity potential energy of the system,


is

Eu
is the remaining internal energy
of the system as defined
by (4); tlmt is, the internal energy which is not a
function of motion
and gravity.*
The quantity Eu is the portion of the internal energy which is
stored in the molecular and atomic structure of the system. For a
particular system this portion of the
internal energy can be evalu-
ated, without going into details
of atoms and molecules, by measure-
ments of work and heat transfer. But a more general approach is
available through the properties of
pure substances.
A pure substance is a substance of uniform constant cliAmirg,!
composition throughout its mass. This includes mixtures having

E
will also depend upon surface
tension electrical, and magnetic effects all
,
of which are assumed absent in
this book. If they were present, additional terms
Would appear for them in Eq. (4).
M] THE FIRST LAW 55

constant composition. In a system composed of a pure substance,


experience shows that in general the quantity Eu is fixed if the mass
and any two other independent properties (excluding velocity and
elevation) are fixed. Furthermore at any given state of such a sys-
tem, Eu is directly proportional to the mass of the system. Hence
/or a unit mass of a pure substance there exists at each state a value of
Eu which can be determined once and for all by mmsuremerds of heat
and work. These values can then be formulated as a function of two
observable properties such as temperature and pressure. The value
of Eu for a unit mass of a pure substance is called the specific internal
energy of the substance* The specific internal energy of a pure sub-
stance willbe designated by a small M.t The corresponding property
for a system consisting of m lb of a pure substance will be designated
by capital U. Then U = mu. The First Law for a system com-
posed of a pure substance in the absence of internal energy changes
due to motion and gravity may then be written

Q = AU +W (5)

The property u is a function of two independent properties of a


substance and may be formulated either as an equation, u = f{p,t)
for example, or as a tabulation of the same relationship, t This spe-
cial concept of internal energy is of great importance, first because
many systems encountered thermod3Tiamics may for
in engineering
purposes of analysis be considered as pure substances, and second
because it provides a basis for the computation of the internal energy
of more complicated systems.
To summarize: The general s3Tnbol E refers to the internal energy
of any system, and includes all forms in which energy may be stored
within the system. The special symbol U refers to the internal
energy of a system composed of a pure substance and does not in-
clude internal energy associated with motion, gravity, and the effects

* Specific internal energy may be evaluated not only by mcasuremenUs of


work and heat, but in part by other methods based upon theories of atomic and
molecular structure. The manner in which the data are obtained has no bearing
upon their use in thermodynamics.
t The distinction between Eu and u is that E„ applies to any system, where
u applies only to a pure substance.
t See, for example, Keenan, J. H. and F. G. Keyes, Thermodynamic Properties
of Steam. New York: Wiley, 1936; Keenan, J. H. and J. Kaye, Gas Tables;
Thermodynamic Properties of Air. New York: Wiley, 1948. Sec also the tables
in the Appendix.
56 THE FIRST LAW [5-6

pagf oAtAil in Eq. i. E is & function of motion, position in the

gravity field, mass, composition, and the states of the eystem’s com-
ponents. a function of the mass and state
is of a syston composed

of a pure substance and is different for different substances. U is


not a function of motion or position in the gravity field.
5-6 Specific heat at constant volume. In Chap. 3 specific
heats were presented as measures of heat transferred in processes
involving no work other than p do work. More general definitions
can now be given. The specific heat of a substance at constant vol-
ume is defined as the rate of change of specific internal energy with
respect to temperature when the volume is held constant. In the
notation of the calculus this is written

From (6) it follows that for any constant-volume process,

(Au)t — Cx dT (7)

In order to show that the specific heat at constant volume as defined


in Chap. 3 is a special case of the specific heat defined in Eq. (6), it
is necessary to show that in that special case (Au), is equal to the
heat transferred. This is done as follows. The First Law may be
written for a stationary system composed of a unit ttiass Qf g,

substance,

0 = Am + IF (8)

or in differential form, dQ = du + dW (9)

Now for a process in the absence of work other than p do work

dW — p do
and dQ = dM + p dv (10)

If the volume is held constant, dr =


0 and (dQ), = (du),. Hence,
in the absence of work other than p dv work, Eq. (7) may be written
for a constant-volume process as

which is Eq. (3) of Chap. 3 written for a unit mass in a constant-


volume process.
5-7] THE FIRST LAW 57

The specific heat at constant volume as defined in tins section is

a property of a pure substance and may be expressed as a function


of any two observable properties such as temperature and pressure.
The specific heat at constant volume may be used to calculate the
internal enei^ change in any constant-volume process, but it may
be used to calculate heat transferred only when all work is p do work
(in which case work will be zero).
5*7 Enthalpy. The specific heat at constant pressure may also
be defined in terms of properties of a substance. For this purpose a
property is needed which bears the same relation to heat in a constant-
pressure process as internal energy does in a constant-volume process.
Internal energy change is equal to the heat transferred in a constant-
volume process involving no work other than p do work. From Eq.
(10) it is possible to derive an expression for the heat in a constant-
pressure process involving no work other than p do work. In such a
process in a stationary system composed of unit mass of a pure
substance

dQ = du pdv (10)

At constant pressure pdv = dipv ) ;


then

idQ)p = du + d(pv)
or {dQ)p = d{u -f pv). (11)

The quantity (m -t- pv) will obviously be a property of a substance


because for each set of definite values assigned to u, p, and v, the
quantity (u -f pv) will have a definite value. A new property called
enthalpy (en-thal'-py) is therefore arbitrarily defined as the sum of
the internal energy and the pressure-volume product. Using the
symbol k for specific enthalpy,

h = u + pv (12)

Observe that Eq. (12) is an arbitrary definition. The reasons for


choosing this definition involved heat, and a particular type of process
at constant pressure, but the definition itself contains no restrictions.
The enthalpy of mass m is given hy H = mh. No special
a system of
symbols are used to distinguish between the enthalpy of a pure sub-
stance and the enthalpy of a general system. For a general system

H ^Eu + pV (18)
58 THE FIRST LAW

and for a pure substance


H— U pV (14)

In any given problem only one of these concepts will ordinarily be


involved so no confusion need arise from lack of another symbol.
Since no absolute value for internal energy is known there can be

no absolute value for enthalpy. However, since pressure and volume


have absolute values, if any arbitrary base is set for internal energy,
the values of enthalpy at all states will be fixed. In the steam tables
of Keenan and Keyes an arbitrary base was set for enthalpy, and the
values of internal energy thereby became fixed. Care must be exer-
cised in substituting numerical values in the equation h = u + pv
since u and h are frequently in Btu/lb while the pv product will be
in ft Ibf/lbm when the usual units are used. The equation as writ-
ten demands consistent units.
5-8 Specific heat at constant pressure. The specific heat of
a substance at constant pressure is defined as the rate of change of
specific enthalpy with respect to temperature when the pressure is
held constant. This is written

= (i). (15)

from this it follows that for any consUint-pressure process

Cp dT (16)

The heat at constant pressure defined in Chap. 3 is a special


specific
case of the specific heat defined in Eq. (15). It is evident from
Eqs. (11) and (12) that for a constant-pressure process in
the absence
of work other than
p dv work, the heat transferred could be substi-
tuted for (AA)p in Eq. (16), giving

which is Eq. (3) of Chap. 3 written for a unit mass in a constant-


pressure process.
The heat at constant pressure is a property of a
specific
pure
substance. It may
be used to calculate the enthalpy change in any
constant-pressure proc^ess, but it may be used to
calculate the heat
transferred only when no work other than
p dv work is transferred.
5-9] THE FIRST LAW 59

The reason for defining specific heats in terms of properties rather


than in terms of heat quantities is that the definition automatically
makes the specific heats properties. This permits the development
of many useful relations among the and other proper-
specific heats

ties. The specific heats still serve a useful purpose in heat transfer
computations as before but they are now useful in many more ways.
5-9 Latent heats, which were discussed in Sec. 3-10, may be
defined as enthalpy changes. Hereafter in this book, when the term
latent heat is used it will mean change of enlhalpij at constant pressure
for a given phase change.
5-10 Applications of the first law to stationary systems. In
this section some examples will be given of the use of the First Law
in problems involving stationary systems.

Example 1. A mass of gas is compressed without friction


stationary
from an initial state of and 15 psia to a final state of 5 cu ft and
10 cu ft

15 psia, the pressure remaining constant during the process. There is a


transfer of 34.7 Btu of heat from the gas during the process. How much
does the internal energy of the gas change?

Solution: Let the mass of gas be the system. By the First Law
Q = AE + W. Q is given, AE is to be found; hence seek a means of de-
termining W. In Sec. 2-3 it was pointed out that when a stationary fluid

system of uniform pressure is frictionless the only w^ork associated with it is

that given by J*p dV. Hence W = J*p dV. For constant pressure

W= p{V2 - Fi)

W= (15)(144)(5 - 10)

W= -10,800 ft lb

the negative sign indicating work done on the system, Now^ substituting
in the First Law equation and using units of Btu,

10,800
-34.7 = AE - 778'

Note that the heat is negative since it is transferred /row the system. Then
AE = -34.7 + 13.9 = -20.8 Btu

The internal energy of the gas decreases 20.8 Btu in the process. Observe
the necessity for adherence to consistent units and sign conventions.

Example 2. A stationary system of constant volume experiences a tem-


perature rise of 35°F when a certain process occurs. The heat transferred
in the process is 34 Btu. The specific heat at constant volume for the pure
W

60 THE FIRST LAW [W6

substance in the system is 1.2 Btu/lb F, and the system contains 2 lb of


substance. Detemine the internal energy change and the work done.
Solution: The First Law applies in the form

Q = At;+
Q is given; AU can be found in a constant-volume process from

AU = mj
k
Then AU = 2j Ti 1.2 dT = 2.4(rs -TO
AU = 2.4 (35) = 84.0 Btu
Then W = g- AC/ = 34-84 = -50 Btu
Observe that even though the volume was constant the work was not zero.

Example 3. In the steam tables of Keenan and Keyes it is set forth


that during the evaporation of 1 lb of water at 500 psia and 467.0rF the

specific volume increases from 0.0197 cu ft/lb to 0.9278 cu ft/lb,


while the
enthalpy increases from 449.4 to 1204.4 Btu/lb. How much work is
done
by a stationary system consisting of 1 lb of water when, because of an
inflow of heat, the system changes from liquid to vapor at
500 psia? How
much does the internal energy change?
Solution: The only work that can be done by a stationary system
of
uniform pressureis given by dV (in the absence of surface tension etc.).
fp
Therefore the work of the system in this constant-pressure
process is
W = p(Vt - Yi)
W = (500)(144)(0.9278 - 0.0197)
W - 65,380 lb = 84.0 Btu
ft

Since enthalpy, pressure, and volume are given for


both end states of the
process, the internal energy change may be
found from the definitinn of
enthalpy,

A “«+ pv
Then Ah = A(u -f- pv) = Au + A(jw)
or Au » AA - A(jw)

Au = 1204.4 - 449.4 - (0.9278 - 0.0197)

Au - 755.0 - 84.0 - 671.0 Btu


Obs^ the conversion of psia to psfa (Ib/sq ft abs) and the conversion of
the ft lb units of the pv product to Btu.
Another method may be used to
6- 10 ] THE FIRST LAW 61

find the internal energy change. In a constant-pressure process in which


the only work is given by Sp t^he heat transferred is given by

Q ^ Ah
asshown in Sec. 5-7, Eqs. (11) and (12). But for any process in a stationary
system consisting of a pure substance the First Law may be written
Q--AU + W
Therefore in this problem
Ah ^ Au +W
or Au ^ Ah — W
Then Au « 755.0 - 84.0 -= 671.0 Btu
which is the result obtained by the first method.

Example 4. The internal energy of a certain substance is given by the


following equation:
u = 0.480pf; + 35
where u given in Btu/lb, p is in psia, and v is in cu ft/lb. A system com-
is

posed of 3 lb of this substance expands from an initial pressure of 75 psia


and volume of 6 cu ft to a final pressure of 15 psia in a process in which
and volume are related by
pressure = constant.
(a) If the expansion is frictionless, determine Q, AU and W for the
process.
(b) In another process the same system again expands according to the
same pressure-volume relationship as in part (a), and from the same initial
state to the same final state as in part (a), but the heat in this case is 30 Btu.
Find the work for this process.
(c) Explain the difference in work in parts (a) and (b).

Solution: (a) AU can be determined from the equation given.


AU ^ Uz — Uv ^ m{u — ui) 2

AU ^ 3iOASOp^2 + 35 - 0.480piVi - 35)


AU = 3(0.480) (P V — PiVi)
2 2

but V == V/m = 7/3


so AU = (0.480)(p2F2 - piVj)

Pi, Vi and p2 are given, and piFl-* = p272*

^2 = 6(3.83) = 22.98 cu ft
AU = (0.480)[(16)(22.98) - (75)(6)]

AU = (0.480)(275.8 - 450) = -83.6 Btu

C*
62 THE FIRST LAW [6-11

Since the process is frictionless, W ’“fp dV. For the relation » =•

constant,
w
^ - Pi7j
1 - 1.2

- 450)(144)
^
^ _

(275.8
- 0.2
W = 125,400 ft lb = 161 Btu
The heat can now be found from the First Law.

Q = AU + W
Q = -83.6 + 161 = 77.4 Btu

(b) The system changes along a path identical to that for part (a).

Therefore f'pdV the same as for part (a). But it is also


’i^ true that the
system changes between the same end states as in part (a). Therefore AU
is the same as for part (a).

AU = -83.6 Btu
Q is given = 30 Btu
By the First Law,

IF = 0 - Af/ = 30 - (-83.6) = 113.6 Btu.

(c) The work in (b) is not equal to fp dV. This simply means that
process (b) is not a frictionless process. The point in this exaTnple is that
W is not always equal to fp dV but depends upon the type of process as
well as the path. The change on the other hand,
of internal energy, At/,
depends only upon the end states and not at upon the process. The all

First Law also applies regardless of the nature of the process. Hence the
work in (b) can be determined from the known AU and the given Q, by
use of the First Law.

5-11 Summary. The First Law of Thennodynamics is a special


statement of the law of conservation of energy. It may be written
for a cycle as

(S0)oyc..= (STDeyC.
or for a process as
AE +W
The latter equation may be taken as a definition of the quantity
internal energy, E. Internal energy has no absolute value but its
change between any two states of a system does have a definite value.
Hence internal energy is a property of a system.
The general concept of the internal energy of a system includes
energy in storage as kinetic energy associated with observable veloci-
Ml] THE FIRST LAW 63

ties, as gravity potential energy of the system as a whole, and as the


potential and atoms and molecules of the sys-
kinetic energy of the
tem.* The first two may often be computed separately by the laws
of mechanics but the third, from the viewpoint of thermodynamics, is
obtained from measurements of heat and work relative to the i^stem,
or to the substances of which the system is composed.
A special concept of internal energy is the specific internal energy
of a pure substance. This is the internal energy of a unit mass of a
pure substance in the absence of gravity and motion (and surface
tension, electricity and magnetism). Although the specific internal

energy change Au can be determined by experimental


originally only
measurements, it can be formulated as a function of any two inde-
pendent properties of a substance for subsequent use. The symbol
U is used to represent mu, the internal energy of m pounds of a
substance.
The specific heat at constant volume of a pure substance is defined
in terms of the internal energy of the substance by

This definition includes as a special case the definition given in


Chap. 3. By the use of the more general definition the specific heat
at constant volume provides a means of calculating internal energy
changes in any constant-volume process while still providing the
means of calculating heat transfer in constant-volume processes in-
volving no work.
The specific enthalpy A of a substance is a property arbitrarily
defined by h = u + pv. All terms in this equation must have the
same units, for example ft lb. The s}rmbol H is used to represent
mh, the enthalpy of m pounds of a substance.
The specific heat at constant pressure of a pure substance is de-
fined in terms of the specific enthalpy of the substance by

This definition includes as a special case the definition given in


Chap. 3. By the use of the more general definition the specific heat
at constant pressure provides a means of calculating enthalpy changes

* See footnote on p. 54.


64 THE FIRST LAW

in any constant-pressure process while still providing the means of


calculating heat transfer in constant-pressure processes involving no
work other than p dv work.
In a frictionless constant-volume process in the absence of motion
the heat transferred is equal to the change of internal energy. In a
frictionless constant-pressure process, in the absence of motion, the
heat transferred is equal to the change of enthalpy. These two state-
ments presume no work from electrical, magnetic or surface-tension
effects.
PROBLEMS*
5-1. (a) A system consists of the air in a tire pump and the connected tire.

The tire-pump plunger is pushed in and the temperature of the pump cylinder
is observed to rise. Give the signs of Q and W for the process, (b) In the
same apparatus the fluid system is a frictionless, incompressible liquid; are W
and AE positive or negative for the fluid system when the pump plunger is

pushed in?
5-2. A is made by burning a mixture of fuel
certain combustion experiment
and oxygen a constant- volume ^^bomb’’ surrounded by a water bath. During
in
the experiment the temperature of the water bath is observed to rise. For the
system comprising the fuel and oxygen give the signs of Q, AE and for the W
process.
5-3. (a) A truck powered by an electric storage battery climbs a hill at
constant speed; the storage battery and the motor are at higher temperatures
than the surroundings. Considering the entire truck as a system, indicate on a
sketch the streams of heat and work crossing the boundary of the system; would
AE be positive or negative? (b) A trolley car climbs a hill at constant speed;
the motor is at a higher temperature than the suiToundings. Considering the
a system, indicate on a sketch the streams of heat and work crossing
entire car as
the boundary of the system; would AE be positive or negative? (c) A truck
powered by a gasoline engine climbs a hill at constant speed; the engine takes
in air from the atmosphere and discharges hot exhaust gases. Why cannot a
simple j^stem be chosen to analyze this process, similarly to cases (a) and (b),
above?
5-4. A slow chemical
reaction takes place in a fluid at the constant pressure
of 15 psia. The surrounded by a perfect heat insulator during the reac-
fluid is
tion which begins at state 1 and ends at state 2. The insulation is then removed
and 100 Btu of heat flows to the surroundings as the fluid goes to state 3. The
following data are observed for the fluid at states 1, 2, and 3;

State V (cu ft) t(F)


1 0.1 70
2 10 700
3 2 70
• In these problems, when the sign of a quantity is asked for, if the quantity
is zero write zero as the answer.
THE FIRST LAW 65

For the fluid system determine the internal energy at states 2 and 3 if the internal
energy at state 1 is zero.
5-5. A piston and cylinder machine contains a fluid system which passes
through a cycle composed of four processes a-b, b-c, c-d, and d-a. During the
cycle, the total negative heat transfer is 150 Btu. The system completes 100
cycles per minute* (a) Complete the following table; show your method for
each item.

Process Q (Btu/min) W (ft Ib/min) AF (Btu/min)


a-b 0 -1.6 X 10*

b-c 20,000 0

c-d -2,000 -34,900

d-a

(b) Find the net rate of work output in horsepower.


A household refrigerator is loaded with fresh food and closed. Con-
5-6.
sider the whole refrigerator and contents as a system. The machine uses 1 kwh
of electric energy (work) in cooling the food, and the internal energy of the
system decreases 5000 Btu as the temperature drops. Find the magnitude and
sign of the heat transfer for the process.
5-7. of liquid having a constant specific heat of 0.6 Btu/lb F is
One pound
stirred in a well insulated chamber, Fig. 5-1, causing the temperature to rise
25°F. Find the change of internal energy of the liquid; find the wbrk for the
process; explain your solution on the basis of the First Law.
5-8. The same liquid as in Problem 5-7 is stirred in a conducting chamber.
During the process 15 Btu of heat arc transferred from the liquid to the sur-
roundings, while the temperature of the liquid is rising 25*’F. Find the change
of internal energy of the liquid, and the work for the process; compare with the
previous problem, and explain the solution on the basis of the First Law.
5-9. The properties of a certain substance are related as follows:

u « 673+ 0.320f
pv = 0.600(f + 460)
where u is specific internal energy, Btu/lb; t is temperature, F; p is pressure,
psia; V is specific volume, cu ft/lb.
(a) Find the specific heat at constant volume, (b) Find the specific heat at
constant pressure.
5-10. A system composed of 2 lb. of the substance of Problem 5-0 expands
in a cylinder and piston machine from an initial state of 150 psia, 200*T to a
final temperature of 60°F. If there is no heat transfer find the net work for the
procesB.
5-11. If all the work in the expansion of Problem 5-10 is done on the moving

piston, show that the equation representing the path of the expansion in the
pv plane is pt/-** « constant.
5-12. (a) A mixture of gases expands at constant pressure from 100 psia,
1 cu ft to 100 psia, 2 cu ft, with 80 Btu positive heat transfer; there is no work
66 THE FIRST LAW
other than that done on a piston. Find the change of internal energy of the
gaseous mixture, The same mixture expands through the same state path
(b)
while a stirring device does 10 Btu of work on the gas. Find the change of
internal energy, the work, and the heat transfer for the process.
5-13. A fluid is confined in a cylinder by a spring-loaded frictionless piston
BO that the pressure in the fluid is a linear function of the volume (p » a -f- hv).
The internal energy of the fluid is given by the following equation:

C/ = 32 + 0.004pF
where is in Btu,
p and V in cu ft. If the fluid changes from an
in Ib/sq ft,
initial state of 25 psia, 1 cu ft, to a final state of 60 psia, 2 cu ft, with no work

other than that done on the piston, find the direction and magnitude of the
work and the heat transfer.
5-14, In a certain process in a stationary system, in which all work is
p dV
work, the pressure is given by
p = (300F -h 1500) and the internal energy is
given by 1/ = 125 +
0.012pF where C/ is in Btu, p in Ib/sq ft, and F in cu ft.
Find the work and the heat transfer if F changes from 6 to 12 cu ft.
5-15. A well-insulated constant-pressure chamber (1 atm pressure) contains
10 lb of liquid water and 1 lb of chopped ice ata certain instant. A moto>driven
stirrerhas been operating in the mixture for some time, and continues to operate,
^th a constant motor output to the stirrer of 25 watts. The specific heat of
liquid water is 1 Btu/lb F, the specific heat of ice is 0.5 Btu/lb
F, and the latent
heat of fusion of water is 143 Btu/lb. The volume decrease of the water
during
melting is 0.M145 cu ft/lb; so small that A(pV) is negligible, and = AU AH.
How much will the temperature change, (a) in one hour? (b) in three hours?
Chapter 6

FLOW PROCESSES-FIRST LAW ANALYSIS

6-1 Flow processes—control surface. The development of the


First Law in Chap. 5 resulted in an equation which applies to any
system in any process,
Q = A£? +F 5-(3)

Since is a function of many variables it was expanded in Eq. 5-(4)


as follows:
E= Eic Ep ) Eu 5"(4)

the terms on the right side representing, in order, the kinetic energy
of the system, the gravity potential energy of the S3r8 tem, and the
residual internal energy not otherwise accounted for. From this point
Chap. 5 was devoted to the special case of a pure substance in the
absence of motion and gravity. Since many practical engineering
operations involve flow of fluids it becomes desirable to develop the
First Law for convenient application to flow processes.

From Eqs. 5-(3) and 5-(4) a general equation which accounts spe-
cifically for the effects of motion and gravity may be written as
follows:

Q AEk + AEp + AEu + W (1)

As in Chap. 5 the system may now be considered to consist of a pure


substance. On this basis, Eq. (1) may be written

Q = AEk "I" AEp -}• *1* W (2)

Equation (2) is the First Law for a system composed ofa pure sub'
stance in the absence of surface tension, electricity, and magnetism.
The system is a particular mam of substance, but it is free to move
from place to place. In Fig. 6-1 a portion of a power plant is shown
diagrammatically; steam flows to the turbine at high pressure, ex-
pands in the turbine while doing work on the rotor, and then leaves
67
68 FLOW PROCESSES [W

at low pressure through the exhaust pipe. If a thermodynamic sys-


tem consisting of a mass of steam is used to analyze this operation
by means of Eq. (1) or (2), the system must be followed in its travels
t^ugh the turbine, account being taken of work and heat transfers
all the way through. Although the system approach is quite valid
there is another approach which wdl be presented here. Instead of
concentrating attention upon a certain mass of working substance,
which constitutes a moving system, attention is concentrated upon
a certain faed region in space called a control volume, through wUch
the working substance flows. In Fig. 6-1 the broken line represents
the surface of a control volume; this surface is known as a control
surface. The method of analysis is to inspect the control surface and
account for all quantities of energy transferred throu^ this surface.
By the conservation of energy the net inflow of energy through the
control surface in a certain time must be equal to the increase during
the same time of the energy stored wiihin the control volume. Since
the control volume experiences not only energy transfer, but also mass
transfer, and since the energy transfer is a function of the trans-
fer, it becomes necessary to make an accounlhig of as well as
of energy. By the principle of the conservation of the net
inflow of mass through the control surface must equal the increase of
6- 1 ] FLOW PROCESSES
mass stored within the control volume. In the mechanics of fluids
the principle of the conservation of momentum is similarly applied
by means of the control surface.

APPLICATION OF THE FIRST LAW


TO A CONTROL VOLUME
The First Law was
developed in Chap. 5 for a system of fixed mass; its appli-
cation to a control volume may be arrived at as follows. Given a control volume,
Fig. 6-a, containing initially a sys-
tem c. During a certain process 1-2 CONTROL VOLUME
let a system d cross the control sur-
face and enter the control volume.
In this process let c transfer heat
and work arbitrarily with the sur-
roundings but restrict d to a fixed
state. Then for the system c in the
given process

Qe — AEc -1- We = Eel “ Eel + We


and for the system d, Ed is constant.
Now before process 1-2 the in-
ternal energy of the system within
the control volume is
Fig. 6-a.
Ev Eel

after the process the internal energy of the system within the control volume is

Evt ™ Eti “H Ed
since the system within the control volume is then c + d. Then the change of
internal energy within the control volume during the process is

AJEJv = Ev% “ Evi Etz — Eel + Qc “ We + Ed


The system c occupies the whole control volume except for the space taken
by system d, and system d has no state change; therefore the heat Qe and the
work We must cross the control surface and must be the net heat and work
crossing that surface. Hence they may be written respectively as Qy and Wy,
the heat and work transfers for the control volume. Then

AEy = Qy — Wy + Ed (a)

This equation depends upon the following premises: (1) the First Law; (2) no
change of state in system d while it is entering the control volume; (3) Qy and Wy
are the net heat and work quantities crossing the entire control surface during
the process of introducing system d into the control volume. In particular, note
that Wy includes the work involved in introducing the system d against the
pressure forces and shear forces from system c*
* It is usual to select control surfaces so that the work involved in introducing
a fluid element is done against pressure forces only.
70 FLOW PROCESSES [6-2

The application of the First Law toa control volume is carried out in detail
in Sec. 6-3. Eq. (8) in tiiat section is an expanded form of £q. (a) above for
the particular case that Er is constant and that material is flowing both into
and out of the control volume.

6-2 Steady flow —definition. In the general case, an arrange-


ment would be subject to changing
such as illustrated in Fig. 6-1
conditiopa from time to time so that the rates of flow of mass and
energy through the control surface would change with time and the
quantities of mass and energy within the control volume would also
ffhnngft with time. For many purposes, however, it is satisfactory
to make analyses upon the assumption of steady flow conditions.
A steady flow process is one in which there is no change in conditions
within the control volume during the time under consideration. This
means that the velocity, internal energy, specific volume, and all

other properties of the working fluid at each point within the control
volume do not vary with time. The operation of a machine like a
reciprocating engine in which conditions do change from time to time
may be treated from the external viewpoint as a steady flow process
if the operation consists of the repetition, at a constant rate, of a

particular process; for, during a period of time involvinga large num-


ber of identical operations, the conditions withiq the control volume
have average values which remain constant as time goes on. Hence
no overall change is observed from an external viewpoint. The con-
ditions of a steady flow process, from the external viewpoint, are
two: (1) the streams of mass crossing the control surface are constant
and add to a net value of zero; (2) the streams of energy crossing the
control surface are constant and add to a net value of zero.
Considering the process illustrated in Fig. 6-1 as a steady flow
process, the material balance (or mass balance) would have only two
terms since there only one inflow (at 1) and one outflow (at 2).
is

For steady flow, the mass entering at 1 per unit time must equal the
mass leaving at 2 per unit time. The energy balance would have at
least four terms to account for heat flow through the control surface,
work flow through the control surface, and energy carried through
the control surface as the internal energy of each of the two »»«««
streams. The algebraic sum of all the energy transfer quantities per
unit timemust be zero for steady flow.
6-3 Material balance and energy balance in simple steady
flow processes. In many processes involving flow only two at.wMuna
M] FLOW PROCESSES 71

of fluid appear, one entering and one leaving the apparatus. Figure
6-2 is a diagram of such a case in which the apparatus under consid-
eration is enclosed by a control surface. The control surface is lo-

cated so that it cuts the fluid streams entering and leaving the
apparatus at sections where flow patterns and properties are steady

with time- The inflowing stream is cut normal to its flow in sec-

tion 1, and the outflowing stream similarly in section 2. The remain-


der of the control surface is so located as to facilitate accounting for
any work or heat that may be flowing through the control surface.
Since the control surface is stationary in space, work can cross it only
in connection with the fluid flow or by means of shafts or other
mechanisms extending through the control surface. (Work could
also cross the control surface by electrical or magnetic effects, and
would be classified as “shaft work” in the energy balance.)

The following quantities are defined with reference to Fig. 6-2.

Ai, cross-section of stream, sq ft

Vh, Wi mass flow rate, Ib/sec

pi, pj pressure, absolute, Ib/sq ft

V\f Vi specific volume, cu ft/lb


«i. Hi specific internal energy, ft Ibf/lbm
Vi, Vi velocity, ft/sec

Zi, Zi elevation above an arbitrary datum, ft


dQ/dt net rate of heat transfer through the control surface,
ft Ib/sec
• Subscripta 1 and 2 refer to the corresponding sections.
72 FLOW PROCESSES [6-3

dWJdt “shaft work”—net rate of work transfer througji the

control surface, exclusive of vwrk done at sections 1 and %


in transferring the fluid through the control surface,
ft Ib/sec

It is assumed that at sections 1 and 2 all properties (pressure,

volume, velocity, elevation etc.) are constant throughout the area of


the section. In practice average values are used for properties which
vary over the section. Though not strictly correct, this is normaUy
a satisfactory approximation when the flow through the control sur-
face is parallel and steady over the area of the section. It is further

assumed that no state change is experienced by the fluid while cross-

ing the control surface.


By the conservation of mass, if no mass change occurs in the

control volume, the mass flow rate entering must equal the mass flow
rate leaving or
wi = to* (3)

Figure 6-3 an enlarged view of the region near section 1. The


is

cross-hatched region represents the fluid which will cross section 1

in the time interval dt; this fluid has a volume AiVi dl. The mass
of this volume is
_
^
Therefore the mass crossing section
(ft = 101 (ft

1 in unit time is
(4)

= AyVi
tOi ( 6)
Vi

By similar reasoning at section 2


:

M] FLOW PROCESSES 73

= 4*1*
Wi (6 )

and from Eqs. (3), (5) and (6)

_ A,Vt
(7)
Vi V2

This equation, known as the Equation of Continuity, expresses the


conservation of mass for the process.
By the conservation of energy the total rate of flow of all energy
streams entering the control volume must equal the total rate of flow
of all energy streams leaving the control volume. This may be ex-
pressed in the following equation

wjjci + wiWi + ^ = (8)

where ei is the internal energy carried into the control volume


with unit mass of fluid at section 1 ,
Wi is the work transferred into the control volume while
introducing unit mass of fluid at section 1,

62 is the internal energy carried out of the control volume


with unit mass of fluid at section 2 ,

11^2 is the work transferred out of the control volume while


extracting unit mass of fluid at section 2, and
wi, "usi, Q, and are as defined in relation to Fig. 6-2.
The internal energy ci can be expressed by writing Eq. 5-(4) for
a S3 8tem of unit mass,
r

e = ei, + e, + eu (9)

where e* is kinetic energy per unit mass go is the dimen-


sional constant 32.17 ft Ibm/lbf sec*.
Cp is gravity potential energy per unit mass = z(g/go)- The
fraction g/go may be taken as numerically equal to unity.
6u is internal energy of the fluid substance per unit mass
— u, the specific internal energy for a pure sul)stance. Then

= ^+
^go go
+ ( 10 )

The work Wi, often called flow work, may be derived as follows.
Beferring to Fig. 6-3, the fluid crossing section 1 exerts a normal
pressure pi against the area Ai giving a total force piAi. In time dt
74 FLOW PROCESSES
this force moves in its own direction a distance Vi dt, thereby doing
work. The work in time dt is piA^i dt, or the work per unit time is
piAjV^. But by the definitions of Wi and TTi the work per unit time
is also W\Wi. Then _
WiWi = piAiVv ( 11 )

Substituting from Eq. (5)


Wi = piVi. ( 12 )

A similar equation can be written for Wt.


Equation (8) may now be rewritten, using Eqs. (10) and (12)
and their counterparts for section 2, giving

arr » ,,_v
+ WtpiVt
, ,
-|
^ (13)

Since Wi = Wt, let w— wi ~ m = dm/di, and divide through by


dmidt, giving

H
2go
,
,dQ H.lii I
(14)

This is the steady flow energy equaiion for the special case of a single
stream of fluid entering and a single stream of fluid leaving the con-
trol volume. All terms representing energy flow iido the control
volume are on the left side of the equation and all terms representing
energy flow out are on the right side. All terms in Eq. (14) represent
energy flow per unit mass of fluid flow. Equation (13) is the same
equation written in terms of energy flow per unit time. Either basis
may be used as may be convenient in any particular problem. It is
common practice to write simply Q and Wx in place of the derivatives
with respect to time or mass, since in any one problem all terms will
be reduced to one basis or the other, and the time rate or mass rate
is understood.
The basis of energy flow per unit mass is usually more convenient
when only a single stream of fluid entering and leaving is involved.
The basis of energy flow per unit time is more convenient when more
than one fluid stream is involved. The equations as presented would
have units of ft lb for all energy terms, but obviously the units tnay
be changed to Btu by dividing through by 778. In solving prob-
&4] FLOW PROCESSES 75

lems a definite choice of basis should be made before attempting to


write the energy equation. It should be decided at the beginning
whether the analysis will be on a time basis or on a mass basis and,

before substitutingany numerical values, a definite choice of energy


units should be made.
Equations (13) and (14) contain the group of properties (m + jm)
which has already been given the name enthalpy and the symbol h
(Sec. 6-7). Thus Eq. (14) may be, and frequently is, written

n
2^0
h 2i
fifo
+ Ai + dm = n—
2^0
h ^ "h ht -i- ^
-t~
dm
k (15)

It is largely this fact which accounts for the great practical utility

of the enthalpy property, since many engineering processes are steady


flow processes, and the use of Eq. (15) reduces computation when
the enthalpy known.
is

When more than one stream of substance enters or leaves the


control volume, the mass equation and the energy equation for steady
flow are easily written by adding the terms for the individual streams

of fluid. Consider, for example, the control volume shown in Fig. 6-4.
Fluid is 1 and 2, and leaving at sections 3 and
entering at sections 4.

For steady flow the mass balance will give


wi + Vh = Wi + Wt (16)
+ + ®

76 FLOW PROCESSES [64

The energy balance will give, on a time basis,

(*‘
S + ‘'p- + S + "p”’ + (*"

-(‘+^ + aJ;)“' + + ^ + ''^)w + (*- '^* («)

The symbols Q and Vfz are understood to denote time rates of energy
flow.

VALIDITY OF THE STEADY FLOW ENERGY EQUATION


The assumptions upon which the above development of the steady flow
energy equation depends are generally satisfied for pipe line flow, but should be
scrutinized with some care when applied to the internal analysis of machinery.
With a stream, the square of the average
large variations in velocity across
velocity may not give the average kinetic energy.
With large temperature de-
ferences, and in combustion processes, the assumption of no state change in the
substance crossing the control surface may not be valid. The existence of shear
stresses in the fluid crossing the control surface may result in a work transfer not
accounted for in the equation.
In a pipe line with axis normal to the control surface, even though appreciable
shear forces exist, they do no work because there is no velocity parallel to the
control surface. In the internal analysis of a machine may be impossible to
it
locate a control surface to isolate the desired energy quantities and simultane-
ously satisfy all assumptions of the theory. The results are then open to
question.

6-4 Applications of the steady flow equations. The steady


flow energy equation applies to a wide variety of processes including
pipe line flow, heat transfer processes, combustion processes, and
mechanical power generation in engines and turbines. The relative
importance of the various terms in the equation differs in the different
applications and some terms are usually zero or negligible in
each
problem. In solving a problem, however, it is best to write the com-
plete equation first and then eliminate the unnecessary terms.
There
can be no general rule as to when a given term may be considered
negligible; thismust be decided in the light of the circumstances and
the objectives of the particular problem. Several examples
will now
be given to show possible orders of magnitude for the different terms
and to illustrate the use of the steady flow equations in various types
of processes.

Example 1. A turbine operates under steady flow conditions, receiving


steam at the following state: pressure 170 psia, temperature 368.4°F,
specific
6-4] FLOW PROCESSES 77

volume 2.675 cu ft/lb, internal energy 1111.9 Btu/lb, velocity 6000 ft/min,
and elevation 10 ft. The steam leaves the turbine at the following state:
pressure 3.0 psia, temperature 141. S^F, specific volume 100.9 cu ft/lb, in-
ternal energy 914.6 Btu/lb, velocity 300 ft/sec, and elevation 0 ft. Heat is
lost to the surroundings at the rate of 1000 Btu/hr. If the rate of steam
flow through the turbine is 2500 Ib/hr what is the power output of the
turbine in horsepower?

SoLtmoN: Sketch the process and locate a control surface. Choose a


basis and units; take energy per pound as basis, Btu as energy unit, and
Ib/hr as flow-rate unit. Write the equations for mass and energy conserva-
tion.
Wi = Wt w
Ul
n+
+ PlVi + ;H *l'^ +Q Ws + PjKj + ^ + *1 gt + Wm
2gt g» 209
Substitution of numerical values with conversion to units of Btu/lb gives:

(170)(144)(2.675) . (6000/60) ,
10 1000
1111.9
778 2(32.2)(778) '
778 2500

914.6 4
(3)(144)(100.9)
778
mi W.
+ 2(32.2)(778) +0+.2600
78 FLOW PROCESSES [6-4

1111.9 + 84.2 + 0.2 + 0.013 -


IF, -
0.4 =
658,500 Btu/hr
914.6

=
+ 56.1 +
219 hp
1.79 +0+ ^
The values appearing in the above problem are typical for a steam power
plant problem. The steam and exhaust pipes
kinetic energy terms for the
are not very large, and the potential energy terms are exceedingly small.
However, the importance of the various terms depends upon the conditions
of the particular problem. Observe that it is not necessarily permissible to
ignore some terms in the equation simply because they are small compared
to some other terms. It is necessary to consider the magnitude of the final
answer rather than the magnitudes of individual terms to determine the
relative importance of the terms. In the above example, in order to obtain
1 percent precision in the final result it is necessary to have 0.2 percent pre-
cision in the internal
energy values; however, the jw product need be known
within only 3 percent, and the kinetic energy within only 100 percent to
obtain the same overall precision.

Example 2. A certain water heater operates under steady flow condi-


tions receiving 500 Ib/min of water at temperature 165°F and enthalpy
132.9 Btu/lb. The water is heated by mixing with steam which is supplied
to the heater at temperature 215®F and enthalpy 1150 Btu/lb. The mixture

w, twiXTURE

Example 2.
M] FLOW PROCESSES 79

leaves the heater as liquid water at temperature 212®F, enthalpy 180 Btu/lb.
How many pounds per hour of steam must be supplied to the heater? State
the assumptions necessary in the solution.

Solution: Sketch the process and locate a control surface, By conser-


vation of mass
Wi +W = 2 Wz (a)

Choose a basis for the energy equation. Since more than one stream is

entering the control volume a basis of energy per unit time will be best.
Use energy units of Btu/min and flow units of Ib/min. The energy
equation may be written

VI
Wi(hi
\
^+
2^0
Zi
goj
h,
+Q

= «>// S+^>-!!-] + w. (b)


\ 2(/o 9^/
By the nature of the process there is no shaft work. Potential and kinetic
energy terms are usually small, and since no information is available they
must be assumed to balance to zero. Since no information is given on heat
transfer it is necessary either to estimate the heat transfer or to assume it is

negligible. It will be assumed that the heater is insulated so that the heat
transfer to the surroundings is negligible. Then Eq. (b) reduces to

Wihi + Wihz = Wzhz (c)

Substituting values, and using Eq. (a) for wz,

(500)(132.9) + u;2(1150) = (500 + w^KlSO)


W2 = 24.3 Ib/min = 1458 Ib/hr

Example 3. Steam is flowing steadily through a pipe line 2.00 inches


in diameter in which there is The pipe is
a pressure drop due to friction.
thoroughly insulated so that heat loss At a certain section in
is negligible.

the pipe the steam pressure is 100 psia, the specific volume is 4.937 cu ft/lb
and the enthalpy is 1227.6 Btu/lb. At another section, downstream from
the first, the pressure is 90 psia, the specific volume is 5.434 cu ft/lb and the

Example 3.
80 FLOW PROCESSES
enthalpy is 1223.9 Btu/lb. Find the average velocity at each of the sections
mentioned, and the rate of flow in Ib/sec.

Solution: Sketch the process and locate a control surface. Choose a


basis for the computation; use ft Ib/lb of steam and Ib/sec for units of
energy and flow rate respectively. By the conservation of mass

wi = Wi —w (a)

By the conservation of energy

hi + ^ + Zi ^ + Q = hi + ~ + zt — + Wx (b)

From the nature of the problem IT, = 0; Q is given = 0; assume the eleva-
tion change negligible. Then

*+? 2go
j. J.Y1

The enthalpies are given but there are two unknown velocities in the equa*
tion. However, Eq. (a) may be written in terms of velocities:
AiVi _ AiVi _ ...

Also, since the pipe is of constant diameter Ai At, and


5.434
^ pj ^
Pi 4.937

or Vt = 1.100 Vi (f)

Substituting (f) in (c)

Vi = 940 fps
Vt = (1.10)(940) = 1034 fps

From (d) w = = 4.15 Ib/sec


Vi

Example 3 illustrates the simultaneous use of the equations of moss and of


energy in solving a problem. This is the basis of many flow-metering
schemes.

6*5 Use of the control surface in simple variable flow proc-


esses. Not all moving fluid may be treated as
processes involving
steady flow processes. For example, consider the discharging of a
gas bottle containing oxygen under pressure; fluid flows during the
6-5] FLOW PROCESSES 81

process but the ptessure in the bottle changes as the mass within
the bottle decreases, so the flow is not steady. Another example
is the series of processes in the operation of a cylinder and piston
machine' such as an engine. From the external viewpoint, over a
period of time, a reciprocating engine may be considered a steady
flow device; however, no detailed investigation of the internal oi)era-v

tions of the engine can be made on the steady flow basis because the
internal operations of the engine are not steady flow operations.
Some attention to what may be called variable flow processes is
therefore desirable.
Consider an apparatus through which fluid is flowing under non-
steady state conditions. In Fig. 6-5 such an apparatus is indicated.

surrounded by a control surface. A stream of fluid is entering the


control volume at section 1 and a stream is leaving at section 2 but
these streams are not equal. The conservation of mass requires that
the rate of increase of mass within the control volume be equal to
the net rate of mass inflow. Then
dm\ dm\ dmt
Wl — tt>2 (19)
dt di dt

where mv is the mass of fluid within the control volume at any


instant,
Wi => dmi/dt is the inward mass flow rate at the same instant,
tvj dmt/dt is outward mass flow rate at the same instant.
This is the equation of continuity for the case of one stream entering
82 FLOW PROCESSES [6-6

and one stream leaving the control volume. Over any finite period
of time Eq. (19) becomes

Amr = Ami — Amj (20)

The conservation of energy requires that the rate of increase of


internal energy within the control volume be equal to the net rate
of energy inflow. Thus
dEv _ dQ dW^ dEi dEt dWi dWt
( 21 )
dt dt dt dt dt dt dt

where, at any instant,

Er is the internal energy within the control volume,


dQ/dt is the net rate of heat flow into the control volume,
dWzIdt is the net rate of shaft work leaving the control volume,
dEildt is the rate of internal energy inflow with mass flow at 1,

dEtIdi is the rate of internal energy outflow with mass flow at 2,


dWi/dt is the rate of work inflow with mass flow at 1,

dW /dt 2 is the rate of work outflow with mass flow at 2.

Referring to the development of the steady flow energy equation,


(13), it will be clear that

^1 ^1 ^ / Ff
+ Ml + plVl
dm\
( 22 )
dt dt \2go 'go ) dt

and ^
dt dt
/li
\2j;o
, 1
^ go + Ml + PiVi
dmt
) dt
(23)

Substituting (22) and (23) in (21), and using the enthalpy in place
of (m -f- pv),

dEv dWz n+ Zl
gVmi
dt dt dt 2^0 go) dt
(24)
For any finite time interval, Eq. (24) becomes

^v^Q-Wz +f(k, + ^; + 2. ^Jdmi - /(ai +


(25)

TIm equation may not always be easy to solve since, in general,


h, V, and z may not be constant; in any case dm is some function of
time if the flow is not steady. The extension of Eqs. (20) and (25)
6^] FLOW PROCESSES 83

to processes involving other numbers of streams of fluid is obvious.


It is apparent that Eq. (25) reduces to the steady flow equation
when AEr = 0, and the terms in the integrands are independent of
time. •

Two examples of the application of the control surface method


to a simple variable flow process are given in the next section.
6>6 Examples of variable flow problems —system and con-
trol surface analysis. Variable flow processes may be analyzed by
either oftwo techniques, the system technique or the control surface
technique. The general development of these two methods, has al-
ready been given so that it remains only to show how they may be

applied. Consider a process in which a gas bottle is being charged


from a pipe line while conditions in the pipe line are maintained con-
stant. This is illustrated in Fig. 6-6. Let the conditions in the pipe
line be indicated by subscript p, while conditions in the bottle are

indicated by subscripts 1 and 2 for the initial and final states, respec-
tively. In the beginning the bottle contains gas of mass mi at state
pi, ti, vi, hi, Ui. The valve is opened and gas flows into the bottle until
the mass of gas in the bottle is mi at state ps, k, vi, hi, ui. The supply
to the pipe line is large enough so that conditions in the pipe line are

unaffected by the flow to the bottle. The state in the pipe line is
given by pp, tp, Vp, hp, Up, and Vp. Elevation is neglected.
Analysis hy System Technique. In order to set up an energy bal-
ance for this process on the basis of a system it is necessary to select
a suitable system. It will be found that a suitable system consists
84 FLOW PROCESSES [6-6

of all the gas that is in the bottle at state 2. In accordance with the
discussion of systems in Sec. 1-3 it is necessary to be able to keep the
system distinct from its surroundings during the whole process. This
isdone by imagining an extensible envelope which contains at aU
times the gas that will eventually be in the bottle at state 2. This

envelope is indicated, at its location for state 1, The


in Fig. 6-7.
system is then all the gas within the envelope. As the process is
carried out, the envelope, always containing^ the same mass of gas,
collapses until it is entirely contained within the bottle.
It is now possible to set up the First Law for this process. This
will be in the form
Q = HE + W 5-(3)

Assuming the gas is a pure substance, the internal energy may be


written as
E= E, + Ep+V (26)

This can be evaluated at state1 by taking the sum of the two internal

energy quantities associated with the portions of the gas within the
bottle and within the pipe respectively. (The volume of the small
connecting tube is assumed negligible.) Then

El TIliUi -j- (27)

The first term is for the gas in the bottle and the second term for the
gas in the envelope but outside the bottle. At state 2 all the system
is in the bottle so
6- 6 ] FLOW PROCESSES 85

Et = miiit (28)
Then
^ Ei — El = mtUa — -H (mi — mi)( Y1
2{7o

The potential energy terms have been omitted, being assumed negli-
gible. The kinetic energy terms have been omitted for the gas in the
bottle because it is not in motion at the initial and final states.

Kinetic energy is included for the gas in the pipe although it might
often be negligible in specific cases.
Taking next the work term of 5-(3) it is observed that if shear
forces at the surface of the envelope are neglected, the conditions of
the problem do not permit of any work except that due to volume
change.* Now referring to Fig. 6-7, the volume change of the process
consists solely in the collapse of the portion of the envelope outside
of the bottle to zero volume. This portion of the envelope, being
always within the pipe, is always subjected to the pressure Pp. The
work is therefore given by Pp AF. The initial volume of the portion
of the system external to the bottle is (nit — m\)Vp, and its final vol-
ume is zero. Then
W = Pp(7i - Fi) = pp[0 - (mj - mi)vp] = -(m* - m\)p^p (30)

Substituting (29) and (30) in 5- (3)

Q = m2U2 — niiUi — (m2 - OTi)^


Y1 + - (m* - mi){.ppVp) (31)
2j/o

By combining the terms containing Up and pjVf, this may be written

F|
Q = W2 W2 wiiWi — (m2 - mi)^
2^0

which is the energy balance for the process.


Analysis by Control Surface Technique. The same problem may be
analyzed by the control surface technique as follows. Select a con-
trol surface as shown in Fig. 6-8. The fluid in the tube between the
pipe line and the bottle will be assumed of negligible mass. The only
place that fluid crosses the control surface is at the entrance to the
tube. At this point all properties are those in the pipe line and are
* Shear forces be negligible, (1) if the fluid is frictionless, (2) if the veloci-
will
ties arc small (of the order of ordinary pipe line velocities), or (3) if the flow to
the bottle is regulated to make the velocity entering the tube the same as the
velocity in the pipe.
86 FLOW PROCESSES [6-6

PIPE LINE

-^TUBE
CONTROL
SURFACE

VALVE-^

L. J
Fig. 6-8. Control surface analysis.

constant; it will be seen that this simplifies the solution. The work
of shear forces by the flowing fluid in the pipe line, upon the fluid
entering the tube, is neglected. This is permissible for frictionless
fluid or small velocities and in this connection ordinary pipe line
velocities may be considered small.
Applying the energy equation, (24), to this case the following
energy balance may be written on a time rate basis:

^ ^ j.
' (33)
dt dt dt ‘^V^'^2po ‘^go/dl
Since shaft work is absent and elevation changes are considered
negligible
dEv dQ ,

(34)
dt d/ V 2j7o/ df

Since hp and F„ are constant, this equation is easily integrated to


give, for the total process

AEv (35)

where (mi — mi) is the mass which flows into the control volume.
Equation (35) is an expression for AEr obtained by accounting for
all energy quantities crossing the control surface during the process.
But it is also possible to write an expression for AEv in terms of the
kno^vn properties of the fluid within the control volume before and
after the process. Thus

AEy = Ui — Ui nitUi — Wi«i (36)


6-6] FLOW PROCESSES 87

No potential or kinetic energy terms appear because the fluid is

stationary and at the same elevation before and after the process.
Combining (35) and (36),

Q = vhih - mm - -I-
~ mi) (37)

Observe that Eq. (37) is identical with Eq. (32), as it must be if

both solutions are correct. Observe also that both the system and
the control surface were so selected as to take advantage of the con-
stant conditions in the pipe line to simplify the analysis. If the pipe
had been a finite chamber instead
line of an infinite source the analysis
would have been more complex.
Anexample involving the intake process of a cylinder and piston
machine will now be given. In a reciprocating engine the piston
never moves far enough to reduce the cylinder volume to zero. The

minimum cylinder volume reached during the operation of the ma-


chine is called the clearance volume, and the fluid occupying that
volume is called the clearance fluid. In a certain engine cylinder.
Fig. 6-9, the clearance gas is known by certain measurements to have
a mass of me The amount of gas taken in through the inlet valve
lb.

during the intake process has also been measured as m/ lb. The state
of the clearance gas just before the inlet valve opens is known to be
given by pi, ui, Vi, hi. The state of the gas in the cylinder just after
the inlet valve closes, at the conclusion of the intake process, is known
to be given by pi, ut, vt, Aa. The state of the gas in the supply pipe
is constant and is given by Pp, Up, Vp, hp, Vp. How much heat is
transferred between the gas and the cylinder walls during the intake
process?
88 FLOW PROCESSES [6-7

Analysis by CarUrol Surface Technique. Draw a control surface

around the apparatus as shown in Fig. 6-9. The only place at which
material crosses the control surface is at the section p, normal to the
intake pipe axis. Writing the energy balance for this case on a time
rate basis,

dEy dQ dW, g\dOT


dt dt dt
+ ( /ip + 2g(,
-f" Zp
go) dt
(38)

It has been given that hp and Vp are constant and Zp is assumed


negligible. Equation (38) may then be integrated to give

AEv = Q-Wp + (^hp + (39)

for the total process. It is also possible to express LEy in terms of


the known masses and properties of the fluid in the cylinder before
and after the intake process:

AEy = U2 — Ui — (me -f- m/jUi — mcUi (40)

Kinetic energy is assumed negligible in the cylinder before and after


the process.
The work Wx is done upon the engine piston during the
shaft
intake process. This work could be computed if the relation between
pressure and piston travel during the process were known. Normally
an indicator diagram would give this information. The solution of
the problem is then:

Q = (me + m/)u.i - mcUi - m^h, + + Wp (41)

It is easy to verify this result by the system analysis as used in the


preceding example.
6-7 Summary. The First Law may be applied to flow processes
either by the system technique or by the control surface technique.
A control surface is the boundary of a control volume, which is a
fixed region in space upon which attention is to be concentrated in
analyzing a problem.
A steady flow process is a process in which all conditions within
the control volume remain constant with time. The steady flow
analysis may
be applied to a process which is not a true steady flow
process, provided conditions at the control surface satisfy the require-
FLOW PROCESSES 89

meats for steady flow. That is, the mass streams and the energy
streams must be constant with time and must respectively add to a
net value of zero for the entire surface. If the operation within the
control volume is a continuous repetition of a process, at a constant
can satisfy this requirement.
rate, it
For a simple steady flow process involving one mass stream enter-
ing and one mass stream leaving the control volume the following
equations will hold: mass balance or continuity equation, written on
the basis of unit time.

Vi V2

Energy balance or steady flow energy equation^ written on the basis


of unit mass of fluid,

2ffo go
+ Wi + piVi +Q= —+
^Qo
22
go
+ W2 + P2V2 + Wz
Mass balances and energy balances can easily be made for a
steady flow process involving several streams crossing the control
surface by summation of the quantities belonging to the individual
streams of fluid.

For variable flow problems energy equations may be set up by


means of system technique or control surface technique. In these
cases account must be taken not only of mass and energy quantities
crossing the control surface but also of accumulations within the
control volume.

PROBLEMS
6-1. Water at yO^F flows steadily from the bottom of a large open tank
through a nozzle to the atmosphere. The tank is filled to a
kept constantly
depth of 12 above the nozzle, (a) If the velocity of the stream leaving the
ft
nozzle is uniform, and if no heat transfer or internal energy change occurs for
the water passing through the nozzle what will be the water velocity? (b) If
the nozzle diameter is 1.5 in. what is the flow rate in Ib/hr?
6-2. In a real nozzle there will be some friction, so the velocity will not be
as great as found in Problem 6-1. If the measur&d velocity is 0.985 times the
computed velocity how much does the internal energy of the water change while
passing through the nozzle? Assume no heat transfer.
6-3. A
steam condenser consists of a metal shell or tank through which
metal tubes pass; the metal tubes carry a flow of cooling water. Steam fills the
space around the tubes and condenses on the tubes, transferring the latent heat
of vaporization to the water. The resulting liquid (condensate) drips from the
90 FLOW PROCESSES
tubes to the bottom of the she]] where there is a drain hole to remove the con-
densate. In a certain condenser steam enters at a steady rate and condensate
leaves at the same rate. Cooling water flows steadily through the tubes.
Assuming there is no heat transfer with the surroundings, write the steady
flow energy equation for (a) the steam-condensate stream; (b) the cooling water
stream; (c) the condenser process as a whole. Define each term used, referring
to a schematic flow diagram.
6-4. Air flows steadily at the rate of 50 Ib/min through an air compressor,

entering at 20 fps velocity, 14.5 psia pressure, and 13.5 cu ft/lb volume, and
leaving at 15 fps, 100 psia, and 2.6 cu ft/lb; the internal energy of the air leaving
is 38 Btu/lb greater than that of the air entering. Cooling water in the com-
pressor jackets absorbs heat from the air at the rate of 3350 Btu/min. (a) Com-
pute the rate of shaft work input to the air in horsepower, Find the ratio(b)

of inlet pipe diameter to outlet pipe diameter. State assumptions made.


6-5. A gasoline engine operates steadily with a work output rate of 50 horse-
power. Heat transfer from the engine to a stream of cooling water is at the
rate of 120,000 Btu/hr. The sum of all other heat transfer to the surrounding
atmosphere and the engine foundations is 20,000 Btu/hr. The only stream of
fluid other than the cooling water is the stream which enters as fuel-air mixture
at SO^'F and leaves as combustion products (exhaust gas) at 15(X)°F. Determine
the enthalpy increase of this stream (Btu/hr) in passing through the engine.
6-6. A certain heat exchanger consists of a bundle of metal tubes submerged
in a stream of water; a stream of oil passes through the tubes and is heated by
the water. Under steady flow conditions the water enters at 200®F and leaves
at 110°F while the oil enters at 60°F and leaves at 150°F. There are no net
changes in elevation or velocity, and the heat exchanger is perfectly insulated
from the surroundings. The enthalpy of water is given as a function of Fahren-
heit temperature by /i = t~32; the enthalpy of the oil is given by h=i0At
+ 0.00025^. What is the water flow rate required to heat 10, (XX) Ib/hr of oil?
6-7. An
incompressible fluid (specific volume constant) flows steadily in a
horizontal insulated converging tube. The pressure, velocity, and tube diameter
at the inlet section are fixed. If the flow is frictionless the specific internal
energy of the be constant; for this case derive an expression for the
fluid will
pressure in the fluid as a function of the tube diameter,
6-8. Pressure drop due to friction in a pipe is usually calculated by the
equation

where I is the length of pipe, ft;


D is the pipe diameter, ft;
V is the average velocity, ft/sec;
V is the average specific volume, cu ft/lb;
and / is the friction factor, dimensionless.

(The friction factor depends upon the fluid properties, the pipe geometry and
the velocity; valued are given in treatises on fluid mechanics.) When water at
70®F flows through a pipe line of 12 in. diameter at a velocity of 10 fps the fric-
FLOW PROCESSES 91

tion factor has the value 0.014. In a certain case these conditions exist for a
pipe 500 feet long.
(a) Calculate the pressure drop due to friction, (b) If the pipe is perfectly
insulated against heat transfer how much energy of the
will the specific internal

fiuid change during flow through the pipe? (c) If the same pipe, instead of
being insulated, is kept at constant temperature, the internal energy of the
water will remain constant. How much heat will be transferred (Blu/lb)?
6-9. ^eam flowing in a pipe line is at a steady state represented by pp, fp,
Upj Vpy hpy Vp. A small amount of the total flow is led through a small tube to
an evacuated chamber which is allowed to fill slowly until the pressure is equal
to the pipe line pressure. If there is no heat transfer, derive an expression for
the final specific internal energy in the chamber, in terms of properties in the
pipe line.

6-10. The internal energy of air is given, at ordinary temperatures, by


u = izo + 0.172/
where u is in Btu/lb;
ifo is any arbitrary value of u at 0°F, Btu/lb;
t is temperature, degrees F, for which u is given.

Also, for air, pv — 53.34(/ + 460)


where p is in Ib/sq ft,

V is in cu ft/lb.

Anevacuated bottle (Fig. 6-7 or 6-8) is fitted with a valve through which
air from the atmosphere, at 30 in. Hg and 72‘’F, is allowed to flow slowly to fill
the bottle. If no heat is transferred to or from the air in the bottle, what will its
temperature be when the pressure in the bottle reaches 30 in. Hg?
6-11. Solve Problem 6-10 if the bottle initially contained 1 cu ft of air at
15 in. Hg and 72®F.
6-12. A mass of air (having properties as given in the preceding problems)
is stirred by a paddle wheel in an insulated constant-volume tank; the velocities

due to stirring make a negligible contribution to the internal energy of the air.
Air flows out through a small valve on the tank at a rate controlled to keep the
temperature in the tank constant. At a certain instant the conditions are as
follows: tank volume 3 cu ft; pressure 150 psia; temperature 300'’F; power to
the paddle wheel 0.1 kw. Find the rate of flow of air out of the tank at this
instant.
Chapter 7

THE SECOND LAW OF THERMODYNAMICS

7-1 Introduction to the second law. The First Law of thermo-


dynamics states that, when a process is carried out, a certain energy

balance will hold. The fact that a given hypothetical process would
satisfy the First Law is, however, no guarantee that the process could
be carried out. It is a necessary but not a sufficient condition.

There are certain directional laws in regard to events in the physical


world: bodies at different temperatures, if placed in contact, always
approach a common temperature; rivers always run down-hill; people
always grow old; the reverse of these processes is never observed.
In thermod3rnamics the fact that a directional law exists results in a
limitation on energy transformation other than that imposed by the
First Law. For example, when an automobile is stopped by a fric-

tion brake the brake gets hot; the internal energy of the brake in-
creases by the absorption of the kinetic energy of the automobile.
This satisfies the First Law, but the First Law would be equally well
satisfied if the hot brake suddenly cooled off and gave its internal
energy back as kinetic energy, causing the automobile to resume its

motion. The latter process, however, never happens. The action of


the brake in stopping the car is an irreversible process. Because such
a general law, called the Second Law of
situations constantly arise,
thermodynamics, has been formulated to express the directional char-
acteristic of processes. The Law has been
existence of the Second
tacitly assumed from the beginning of this book; it is the existence
of this law which necessitates the distinction between work and heat
as different forms of energy. For engineering purposes the Second
Law is best expressed in terms of the conditions which govern the
production of work by a thermodynamic system operating in a cycle.
Hence a brief discussion of heat engine cycles is in order.
7-2 Heat engine cycles. A heat engine cycle is a thermodynamic
92
7-2] THE SECOND LAW 93

cycle in which there is a net heat flow to the system and a net work
flow from the system. The system which executes a heat engine cycle
is a heat engine. A heat engine may be of various forms such as a
mass of gas in a cylinder and piston machine, Fig. 7-la, or it may be

0 .

Wp HEAT
ENGINE

O2

(b)
Fig. 7-1.

a mass of water passing in steady flow through the apparatus of a


steam power plant; in any case the heat engine may be represented
by the conventional block diagram of Fig. 7-lb.* The heat and
work quantities indicated in Fig. 7-1 are defined as follows:

Qi, heat supplied; the heat transferred to (he system during a


cycle,

Q2 ,
heat rejected; the heat transferred from the system during a
cycle,
Wbj engine work; the work done by the system during a cycle,
Wpf pump work; the work done on the system during a cycle.
All of these quantities are customarily treated as positive numbers
even though two of them represent negative transfers with respect
to the system. Hence for a heat engine cycle

2Q =Qnet =Qi-02 (1)

and ZW = ir„et = Wb - Wp (2)

Diagrams are drawn consistent with this usage so that Q2 is asso-


ciated with an arrow pointing out from the system whereas the gen-
eral symbol Q is associated with an arrow pointing into the system.

* Note that a heat engine ia not at all an engine in the sense of a piece of ma-
chinery, although it may utilize such apparatu*^.

D*
94 THE SECOND LAW [7-3

The purpose of a heat engine cycle is to produce work; hence the


net work is of {Himary interest. When the terms cyde work or work
of the cyde are used they refer to the net work. For many purposes
it is unnecessary to break down
the work into positive a^d negative
parts. The heat quantities, on the other hand, are always treated
separately because the positive heat Qi represents a cost or input
and is therefore of primary interest.
The efficiency of a heat engine (or of its cycle) is defined by
_ net work output _ TTpet
^
gross heat input ~ Qi
where n (eta) is the symbol for efficiency. The efficiency here defined
is often called thermal efficiency. By the First Law,

ZQ = ZW (4)
Substituting from (1) and (2),

Qi - Q» = We ~Wp = (5)

~
then
’_
_ Qi
0 .
Qi 1

'"Q,
Qi
(e)

J^ote that Eq. (6) is simply a deduced result; Eq. (3) is the definition
of the efficiency of a cycle.

When heat is transferred to or


from a heat engine there must exist
other systems with which the heat
engine exchanges heat.A more com-
pletediagram than that of Fig. 7-lb
would show (Fig. 7-2) not only the
heat engine, but a body called a
source, from which heat Qi is trans-
ferred to the engine, and a body called
a sink, to which heat Qt may be trans-
ferred from the engine. A typical
source would be the mass of bu rning
fuel in a furnace; a typical sink
would
be a body of water such as a lake
Fig. 7-2.
or a river. Usually, but not nec-
fissarily, sources and sinks are as-
sumed to maintain constant temperature.
7-3 The second law. The First
Law requires of a heat engine
7-3] THE SECOND LAW 95

only that the net work be equal to the net heat. Experience shows,
however, that other restrictions exist; the net work is always lesd
than the heat supplied Qi, that is to say there always exists some
heat rejected Qt. Moreover there is no net work in a cycle unless a
temperature difference exists between the source and the sink. The
formal statement of these empirical facts constitutes the Second Law
of Thermodynamics. The earliest recognition of this fundamental
principlewas by Carnot who said, with reference to a cyclic engine,
“Wherever there exists a difference of temperature, motive power
*
can be produced.” This positive assertion, however, is not as pre-
cise as the negative assertions of later writers since it may imply,
but does not state categorically, that the temperature difference is

essential to the production of network in a cycle. Kelvin, Clausius,


and Planck each produced a commonly quoted statement of the Sec-
ond Law. The following statement, after Kelvin and Planck, is
suitable for the purposes of this book: The Second Law: It is impos-
sible for a heat engine to produce net work in a complete cycle if it

exchanges heat only with bodies at a single fixed temperature.^


This statement appears rather limited in scope but its conse-
quences are extremely broad in their applications. It will be noticed
that the form of the law is negative; it states that a certain class of
heat engine cycles is impossible of realization. Now the possibility,
under the Second Law, of any proposed process may be tested as
follows: construct a cycle composed of the proposed process and some
other processes known by experience to be possible. If such a cycle
can be constructed so that it would produce net work while the sys-
tem exchanges heat solely with bodies at a single fixed temperature,
then the proposed process is impossible. The hypothetical heat en-
gine whose cycle would thus violate the Second Law is frequently
called a perpetual motion machine of the second kind. (A perpetual

* Reflections on the Motive Power of Heal by N. L. Sadi Carnot, Paris, 1324.

Trans, by H.. H. Thurston, Amer. Soc. Mech. Engrs., New York, 1943. Carnot’s
work preceded the establishment of the First Law, hence his reasoning is incon-
sistent with the present concept of heat. Nevertheless his conclusions are sound
because the Second Law is a general law of nature entirely independent of the
First Law. Carnot not only presented the first analysis of the effects of what
is now known as the Second Law but he also originated the use of cycles in

thermodynamic analysis.
t See Section 7-9 for Clausius’ statement of the Second Law. For a discus-
aioD of the statements of the Second Law see Zemansky, M. W., Heat and Ther-
modynamics, New York: McGraw-Hill, 1943, chap. 7.
96 THE SECOND LAW [ ^
7

motion machine of the first kind is a device which violates the First
Law by creating energy.) An example is now given of the method
of proving a process impossible under the Second Law.

Example 1. Given a system, Fig. 7-3, composed of a flywheel mounted


on frictionless bearings and a friction brake which can be pressed against
the flywheel to stop its rotation. Suppose the flywheel has been brought

r n

SYSTEM
BOUNDARV

from rest up to 1800 rpm by a motor which is then disconnected from the
flywheel. By the First Law it is known that the motor must have supplied
work to the flywheel equal to its final kinetic energy (neglecting air friction).
It is well known that the wheel can be brought back to rest by applying force
to the brake block. In this process no work crosses the boundary of the
system since the operating force on the brake moves through a negligible
distance. If the process takes place quickly it may be assumed that heat
transfer is also essentially zero. Then, by the First Law, the internal energy
of the system remains constant; the kinetic energy which disappears becomes
molecular internal energy of the brake block and the wheel as evidenced by
the temperature rise always observed under such conditions.
Now it is also well known that the reverse of the braking process never
occurs, that is to say the wheel and the brake will not be found to give up
their molecular internal energy with the sole effect of producing kinetic energy
of the flywheel as a whole. It is desired to prove that this process, which is
never encountered in experience, is impossible under the Second Law. The
proof requires the construction of a cycle composed of one or more processes
7-4] THE SECOND LAW 97

known to be possible, plus the process X to be proved impossible. Construct


the cycle as follows:
(1) Process X: Starting with the fly wheel at rest, and the wheel and brake
at a high temperature, let the flywheel go from rest to 1800 rpm at the
expense of the internal energy of the wheel and brake block, During this
process the temperature of the wheel and brake block will decrease, the

reverse of what happens in a braking process. (Note that there is no need


to visualizehow the above process could be made to happen; after all, if it is
impossible, it could not be made to happen. This, however, is no bar to
the tentative assumption that the process did occur, since it has not yet been
proved impossible.)
(2) Process A: Bring the flyw’heel to rest by connecting it to a hoisting
mechanism through which weights are raised. In this way the kinetic energy
of the flywheel can he used to do work on the weights with no temperature
change in the system as the flywheel comes to rest.
(3) Process B: Supply heat from some source at a fixed high temperature
to the flywheel and brake to raise their temperatures to the original values.
This completes a cycle, since the system is brought back to the state existing
at the beginning of fhe first process (at rest, and at high temperature).
Summarizing, the following energy transfers occurred during the cycle:

Process X: No energy transfer across the system boundary,


Process A: Work done by the system, no heat transfer,
Process B: Heat transferred to the system fiom a body at fixed tempera-
ture, no work.

This cyple is precisely that which is asserted by the Second Law to be


impossible (or in academic jargon, it is a j^rpetual motion machine of the
second kind). Now since p recesses A and B are operations that are known
by experience to be possible it follows that the impossibility of the cycle
must reside in process X. Thus process X
is proved impossible under the

Second Law.

The acceptance of the Second Law as a fundamental law of nature


rests upon the fact that no one has ever been able to carry out phys-
icallya process which, by the above type of reasoning, could be proved
impossible under the Second Law. There is no way to prove that
experience will always thus confirm the Second Law but since no
single violation has ever been demonstrated the validity of the law
is unquestioned.
7-4 Reversil^ility. In Sec. 7-1 it was pointed Ait that the Second
Law is essentially a guide to the possible directions in which processes
may occur. Considering the flywheel and brake system in Exam-
98 THE SECOND LAW [ 7^5

pie above, the process of stopping the wheel by the brake may be
1,

considered a process of exchange of gross kinetic energy for molecular


internal energy within the system. The reverse process, X in the

example, would be the exchange of molecular internal energy for gross


kinetic energy within the system. Thus the two processes may be
thought of as two opposed directions of the same general process.
According to the Second Law only one of these directions is possible.
A formal way of stating this fact is to say that the braking process
is irreversible.

Although all real processes are probably irreversible, some proc-


esses are capable of being carried out almost precisely in reverse.
For example a free pendulum starts with potential energy of eleva-
tion at one extreme of its swing, exchanges the potential energy for
kinetic energy during its and reverses the process during its rise.
fall,

Experience shows that the reversalis not perfect, for the pendulum

gradually loses amplitude and eventually comes to rest. But by


placing the pendulum in a vacuum the coming to rest can be delayed,
since air friction is eliminated. By care in construction pivot friction
may be minimized; thus the process of the downward swing of the
pendulum may be made to approach the condition of a reversible proc-
ess. It is thus possible to imagine the approach to a reversible
process as a limit, and to a hypothetical
treat the limiting case as
process. Such limiting cases are of the utmost importance in the
development of the applications of the Second Law; consequently
the concept of reversibility must be formalized.

Demnition: A process is reversible if, after it has been carried oni,


it ispossible by any means whatsoever to restore both the system and its
entire surroundings to exactly the same states they were in before the

process. A process that is not reversible is irreversible.

This definition means that, if a process is reversible, it is possible


to undo it in such a way that there will be no trace anywhere of the
fact that the process occurred. A process that is irreversible may be
undone in the sense that a specific system
restored to its initin.1
is
state, but it is not possible to do this without leaving some permanent
change in things external to the system.
7-5 Irrevenil Ue processes. The definition of a reversible process
is general; it specifies no particular physical conditions necessmy for

reversibility. However, the Second Law requires, and experience


;

7-6] THE SECOND LAW 99

confirms, that the presence of certain specific conditions during a


process will render the process irreversible; among these conditions
are:
(1) heat transfer through a finite temperature difference;
(2) lack of pressure equilibrium between the system and its con-
fining walls;

(3) free expansion; the expansion of a system to a larger volume


in the absence of work done by the system;
(4) friction, solid or fluid (and electrical resistance)
(5) transfer of paddle-wheel work to a system.
The first three items involve a lack of equilibrium during the process;
the others involve so-called dissipative effects in which work is done
without producing finally an equivalent increase in the kinetic or
potential energy of any system.
An example of the effect of friction in causing irreversibility was
given as Example 1 above. Any process involving friction can be
shown in a similar manner to be irreversible.

Example 2. To demonstrate that heat transfer through a finite tem-


perature difference is irreversible, consider a cyclic heat engine E (Fig. 7-4)

producing work W while receiving heat Q\ from a source A at temperature


(.1 and rejecting heat Q: to a sink B at ta. Experience shows that this can
happen if > ta. It is assumed that the source and sink are capable of

(a) (b)
Big. 7-4.
100 THE SECOND LAW [7-6

remaining at any given temperatures while the engine operates. Now it is

known that heat can flow directly from A to 5, as for example Qa-^b in Fig.

7-4a. To prove this process irreversible assume that it can be reversed as


shown by Qb^a in Fig. 7-4b. Let Qb^a equal then there is no net heat
transfer to B, and B passes through a cycle when the engine E passes through
a cycle. Then the combined system E and B passes through a cycle and
produces net work while exchanging heat solely with the constant tempera-
ture source A, This violates the Second Law. Since the cyclic operation
of E and B is known to be possible, the impossible process must be the heat
transfer Qb-a] therefore the heat transfer Qa-b is irreversible. Since the
cycle described could be constructed in conjunction with any arbitrary heat
transfer Qb-a from a lower to a higher temperature it follows that any such
process is impossible, and every heat transfer from a higher to a lower tem-
perature is irreversible.

.
Although every real heat tarnsfer process requires a temperature differ-

ence and is therefore irreversible, it is useful to reason in terms of hypothetical


reversible heat transfer processes in which there is no temperature difference
between the two systems involved. Experience shows that the time rate of
heat transfer between two systems alw'ays approaches zero as the tempera-
ture difference approaches zero; therefore a finite quantity of heat can be
transferred reversibly only in an infinite time.

Example 3. Demonstration of the irreversibility of a process of free


expansion: An insulated container. Fig. 7-5, is divided into two regions,

Fig. 7-5.

A and B, by a thin diaphragm. Region A contains a mass of gas and


region B is completely evacuated. If the diaphragm is punctured, the gas
willexpand into region B until the pressures in A and B become equal, but
there will be no effects upon the surroundings. This process is judged by
intuition to be irreversible, but a formal proof is desired, basedupon the
Second Law. Let the mass of gas be the system.
Following the procedure used before, assume that the free expansion is
reversible. Let the reverse process occur, by which the gas in B returns
into A with an increase in pressure, without causing any net effects in the
7-5] THE SECOND LAW 101

surroundings. (Since the free expansion caused no effects in the surround-


ings its reversal must leave no net change in the surroundings.) To construct
the cycle which shall violate the Second Law, install an engine (a machine,
not a heat engine) between A and fi, Fig. 7-6, and permit the gas to expand

Assumed REVERSED free


exponsion for compressing

through the engine from A to B instead of using a free expansion, The


engine can obtain work from the sysUjm at the expense of the internal energy
of the gas as it expands. But after the expansion through the engine, the
internal energy of the system (the gas) can be restored to its initial value by
heat transfer from a source at a fixed temperature. Thus the system can
be brought to the same state as if the free expansion had occurred. Now,
by use of the reversed free expansion, the system can be restored to the
initial state of high pressure in A
and vacuum in B. The result is a cycle.
Fig. 7-6, which violates the S cond Law; hence the reversed free expansion
is impossible and the free expansion is irreversible.

Example 4. Demonstrate that a process involving paddle-wheel work


is irreversible.

The total effect of a paddle-wheel process is a change of state in the


system and a transfer of work to the system. (See Sec. 2-3.) Hence in a
reversed paddle-wheel process the total effect would be the reversed change
of state and a transfer of work /rom the system. Now experience shows that
the identical state change produced by paddle-wheel work can be produced
by a transfer of heat to the system. Therefore the reversed paddle-wheel
process in combination with a process of heat transfer to the system can
form a cycle in which the total external effects are a work transfer oid and a
heat transfer in. From this it follows that the paddle-wheel process is
irreversible.
102 THE SECOND LAW [7-6

7-6 Mechanical and thermal reversibility reversible sys- —


tems. Experience shows that no real process can be strictly revers-
ible. Nevertheless, in many cases, certain effects of real processes

differ little from the computed by assuming a reversible proc-


effects

ess. For example the expansion of a real gas in a cylinder and piston
Tn«/^hinp. generally involves so little friction and so little disturbance

of pressure equilibrium that the work done is given accurately by


computations based on a reversible expansion. It is often said that

such a process is considered mechanically reversible; a process involving


paddle-wheel work would be mechanically irreversible. Similarly, if

the process involves heat transfer it may be designated thermally


reversible or thermally irreversible depending upon the absence or pres-
ence of a temperature difference.
A
thermodynamic system is frequently chosen in such a way as
to exclude from the system most causes of irreversibility; the system
is then assumed to experience no irrever-

sible effects within its boundaries. Such


a system is an internally rever-
called
sible system or simply a reversible system.*

7-7 The Carnot cycle. In the ap-


plication of the Second Law much use
is made of reversible cycles, which are
hypothetical cycles composed entirely
of reversible processes. The classical
reversible cycle for the development of
theory is the cycle first proposed by
Carnot and named for him . The Carnot
cycle is a heat engine cycle in which all

heat supplied to the engine is transferred


Fig. 7-7. reversibly from a constanbttemperature
source, all heat rejected is transferred
reversibly to a constant-temperature sink, and all processes
in the
cycle are reversible. Referring to Fig. 7-7, if the
engine is to operate E
in a Carnot cycle, the heat Qi must be transferred while the system
is at temperature and the heat
t\,
Q2 must be transferred while the
system is at temperature tt. To complete the cycle, the system must
* For a discussion of various specific cases of reversible and irreversible proc-
esses sec Zemansky, M. W., Heat and Tkermodynamies. New York: McGraw-
Hill, 1943; chap. 8.
7-7] THE SECOND LAW 103

go from ti tott, and back from k to ^i, by adiabatic (zero heat transfer)

processes. Hence the cycle consists of four processes: (1) a revers-


ible isothermal (constant-temperature) process in which heat Qi enters
the system; (2) a reversible adiabatic process in which the tempera-
ture changes from htoh', (3) a reversible isothermal process in which
heat Qi leaves the system; (4) a reversible adiabatic process in which
the temperature changes from h to <i. The details of the mechanisms
by which these processes could be carried out are not important, but
ifa physical picture is desired the system may be considered to be a
mass of gas in a cylinder and piston machine, Fig. 7-8. There are

Fig. 7-8. Carnot heat engine: stationary system.

available a source of heat at temperature h and a sink at tempera-


ture U (which is less than U). It is assumed that means are available
to bring either the source or the sink into contact with the cylinder
head or to cover the cylinder with heat insulation. Then by bring-
ing the source into contact with the cylinder head and allowing the
piston to move out slowly a reversible isothermal heat transfer at
k
would be carried out (hypothetically, of course). By covering the
cylinder with insulation and allowing the piston to move out farther
the system temperature would be reduced to ^ in a reversible adiap
batic process. By bringing the sink into contact with the cylinder
head and forcing the piston slowly toward the cylinder head, a revers-
ible isothermal heat transfer at ti would be carried out. By insulating
104 THE SECOND LAW [7-8

the cylinder and forcing the piston farther toward the cylinder head,
the system temperature would be increased to h in a reversible
adiabatic process. This is the classical picture of the Carnot cycle,
but other pictures are possible, for example, a steady flow system
such as shown in Fig. 7-9. In this case the isothermal heat transfer

processes take place in heat exchangers while the reversible adiabatic


processes take place in the turbine and the pump. To satisfy the
conditions for the Carnot cycle there must be neither friction nor
heat transfer in the pipe lines through which the working fluid flows.
Carnot, by ingenious reasoning, showed that the cycle described
above is capable ot producing the maximum work output for cycles
working between a source and a sink at given temperatures. In the
next chapter Carnot's conclusion will be shown to be a consequence
of the Second Law.
7-8 The reversed heat engine. A reversible heat engine cycle
such as the Carnot cycle need not be used only for a heat engine.
Since all the processes of the Carnot cycle are reversible it is possible
to imagine that the processes are individually reversed and carried
out in reverse order. When a reversible process is reversed all the
7-8 ] THE SECOND LAW 105

energy transfers associated with the process are reversed in direction


but remain the same in magnitude. Hence, in a reversed Carnot
cycle, heat Qiwould flow from the system at temperature U, heat Qi
would flow to the system at temperature
fe, and the net work would

be done on the system instead of by the system. Compare Fig. 7-10,


showing the steady flow arrangement for the reversed cycle, with
Fig. 7-9, showing the steady flow heat engine cycle.

Fig. 7-10. Reversed Carnot heat engine: steady flow process.

The reversal of a cycle may also be visualized by using the block


diagrams of Fig. 7-11 in which (a) shows the normal heat engine
and (b) the reversed heat engine. If is a reversible engine, the

quantities Qi, Q 2, and W will have the same magnitudes, respectively,


whether the engine operates normally or reversed. Only the direc-
tions of these quantities will change upon reversal of the engine.
The reversed heat engine takes heat from a low-temperature body,
discharges heat to a high-temperature body, and receives an inward
flow of net work. The names heat pump and refrigerating machine
are applied to the reversed heat engine. For the heat pump the same
energy relation holds as for the heat engine.
106 THE SECOND LAW [7-9

Kg. 7-11.

F = Q, - Q*

It is possible to operate actual heat pump cycles; they are used

in refrigerators and in heating systems of various kinds. However,


since an actual cycle cannot be reversible, an actual heat pump cycle
is only approximately the reverse of an actual heat engine cycle.
7-9 Qausius’ statement of the second law. The Second Law
may be stated in terms of the heat pump: Clausius stated that it is

impossible for a system working in a complete cycle to accomplish


as the transfer of heat from a body at a given tempera-
its sole effect

ture to a body at a higher temperature. This means that a heat


pump must have a net work input. The complete equivalence of
the above statement and the statement of the Second Law given in
Sec. 7-3 is now shown in two steps.

(1) Assume that a cyclic heat pump could transfer heat from a
low-temperature region to a high-temperature region with no other
effect. Let the heat pump, P in Fig. 7-12, transfer heat Qi from a
sink at <2 to the heat engine E at temperature h. E receives heat Qi
at ti, and puts out net work W. Obviously the
rejects heat Qt at U,
condoned S3rstem P and E constitutes a heat engine working in cycles
and producing net work while exchanging heat only with a body
at a single fixed temperature. Hence any system which violates
Clausius’ statement will also violate the Second Law as originally
stated in Sec. 7-3.
7- 10 ] THE SECOND LAW 107

(2) Assume that a heat engine E could produce net work while
exchanging heat only with bodies at a temperature k- This net work
could be used to drive a heat pump P receiving heat at k and dis-
charging heat at some higher temperature as in Fig. 7-13. It is
obvious that the combined system E and P constitutes a heat pump
working in cycles and producing the sole effect of transferring heat
from k to <j. Hence any system which violates the Second Law
statement of Sec. 7-3 will also violate Clausius’ statement.
Thus the complete equivalence of the two statements is proved
since a violation of either statement necessarily involves a violation
of the other.

7-10 Summary. A heat engine cycle is a thermodynamic cycle


in which there is a net flow of heat to the system and a net flow of
work from the system. The system which executes a heat engine
cycle is a heat engine. The efficiency of a heat engine, or of its
cycle, is
:

108 THE SECOND LAW


l^net

where Q\ the heat transferred to the system in a cycle, and ITnet


is

is the net work of the cycle. This efficiency is called thermal efficiency.
A heat pump or refrigerating machine is a system to which there
is a net flow of work and from which there is a net flow of heat in a
cycle.
The Second Law is stated as follows (after Kelvin and Planck)
It is impossible for a heat engine to produce net work in a complete
cycle if it exchanges heat only with bodies at a single fixed tem-
perature.
An equivalent statement (after Clausius) follows: It is impossible
for a system working in a complete cycle to accomplish as its sole

effect the transfer of heat from a body at a given temperature to a


body at a higher temperature.
A process is reversible if, after the process has been carried out,
it is possible by any means whatsoever to restore both the system
and the entire surroundings to exactly the same states they were in
before the process. A process that is not reversible is irreversible.

A a process that can be undone in such a way


reversible process is
that no trace remains anywhere of the fact that the process occurred.
Reversible processes exist only as ideal limiting cases which real
processes may approximate to greater or lesser degree. The presence
of any of the following conditions will make a process irreversible:
(1) heat transfer through a finite temperature difference; (2) lack of
pressure equilibrium; (3) free expansion; (4) friction of any kind;
(5) paddle-wheel work. A truly reversible process would require
infinite time since any process at a finite rate involves some dis-
equilibrium.
A reversible cycle is a cycle composed entirely of reversible proc-
esses. The classical example is the Carnot cycle which consists of
two reversible isothermal processes and two reversible adiabatic
processes.'

PROBLEMS
7-1. Heat is supplied to a cj'<!lic hcjit engine at the rate of 1000 Btu/min,
and the engine docs work at the rate of 10 hp. Find the efficiency of the heat
engine and the rate of heat rejection.
THE SECOND LAW 109

7-2. A dry-cell battery can produce work in the form of an electric current

while exchanging heat solely with a constant-temperature atmosphere. Does


this violate the Second Law? Explain.
7-3. An automobile engine produces net work while exchanging heat only
with the atmosphere; why does this not violate the Second Law?
7-4. Would an atomic energy power plant violate either the First or the
Second Law of thermodynamics? Explain.
7-5. A household refrigerator is in thermal communication solely with the
atmosphere of a constant-temperature room. The refrigerator com-
works in a
plete cycle, receiving work from the electrical lines and exchanging heat only with
the atmosphere. Does this violate the Second Law? Explain.
7-6. An electric storage battery which can exchange heat only with a constant-
temperature atmosphere goes through a complete cycle of two processes. In
process 1 ;
2800 watt-hours of electrical work flow into the battery, while 700 Btu
of heat flow out to the atmosphere; in process 2; 2400 watt-hours of work flow
out from the battery, (a) Find the magnitude and direction of the heat trans-
ferred in process 2. (b) If process 1 has occurred as described above, does the
Law or
First the Second Law limit the maximum possible work of process 2?
How much is the maximum possible work? (c) If the maximum possible work
were obtained in process 2 how much heat would be transferred in process 2,
and in which direction?
7-7. If a hard steel ball is dropped on a hard steel plate it may rebound
almost to the original elevation. If the same ball is dropped on a lead plate it

will rebound to a much smaller height, (a) Explain how the First Law is satis-

fied in each case, (b) Is the process in each case almost reversible or irreversible?
7-8. A fluid flows steadily at constant velocity, without heat transfer, through
a pipe in which there is a pressure drop due to friction. Prove, by the Second
Law, that this process is irreversible.
7-9. Considering a reversible process to be one which approaches reversibility
as a limit (Sec. 7-6), each process described below falls into one of three cate-
gories: (1) it is reversible; (2) it might be reversible; (3) it cannot possibly be
reversible. Place each process in its proper category, and identify the sources
of irreversibility.
(a) A
pound of water is evaporated at constant temperature by an inflow of
heat, A pound of water is evaporated at constant temperature by an inflow
(b)
of work, (c) A mass of air in a constant-volume container is heated slowly from

100°F to 200®F by an inflow of heat, (d) A stationary mass of air is compressed


slowly in a frictionlcss, non-conducting cylinder and piston machine, (e) A gas

goes through the isothermal expansion process of a Carnot cycle, (f) A stream
of steam at 212T mixes with liquid water at 50T in a non-conducting heater,
(g) Hot stagnant gases in the cylinder of a non-conducting, frictionless cylinder
and piston machine expand as the piston moves out slowly, (h) Hot turbulent
gases in the cylinder of a water-cooled automobile engine expand as the piston
moves out rapidly.
7-10. A gaseous system in a cylinder and piston machine executes a cycle in
no THE SECOND LAW
which there are three heat transfers as follows :Qi « — 180Btu/cycle;Q2 " 60,000
ft Ib/cycle; Qi »
75,000 IT calories/cycle. Is the system acting as a heat engine
or a heat pump? Show reasoning.
7-11. A refrigerator utilizing a heat pump cycle has a power input of 0.75 kw,
and a heat flow rate from the cold region of 150 Btu/min. How much heat must
be transferred per hour to the hot region?
Chapter 8

BASIC APPLICATIONS OF THE


SECOND LAW

The Second Law is seldom used directly but several general corol-
laries or propositions derived from it are of great usefulness in the
development of theory, and in the solution of problems. In this
chapter these immediate consequences of the Second Law are de-
veloped.
8-1 The efficiency of reversible cycles. The first application
is to demonstrate the following proposition:

(1) Of all heat engines operating between a given constant-tem-


perature source and a given constant-temperature sink, none has a
higher efiSciency than a reversible engine.
Proof: Let any heat engine Ea and my revartMe heat engine Eb
be allowed to operate between a source at temperature h and a sink
at temperature U, as shown in Fig. 8-1. Assume that the theorem
to be proved is not true; thus assume Ea more efficient than Eb.
Now let the rates of working of the engines be such that

Qu “ Qia =* Qi (a)

111
11^ BASIC APPLICATIONS [8-1

By assumption VA > VB

or
^
Qia
Wb
Qib
(b)

From Eq. (a) and inequality (b) it follows that

Wa > Wb (c)

Now let Eb be reversed. Since Eb is reversible the magnitudes of


the heat and work quantities will remain unchanged but their direc-
tions will be reversed, giving the result shown in Fig. 8-2. Now
t,

ta

Fig. 8-2.

take part of Wa to supply Wb; also, use Qw to supply Qu- This


shown
will permit the elimination of the source at U, giving the result

in Fig. 8-3. The combined systems Ea and Eb now form a heat


engine which is producing net work — Wb) while exchanging
heat only with a single constant-temperature body at U- This is a
direct violation of the Second Law.* Consequently the assumption

Fig. 8-3.

*
Observe that the operation in Fig. 8-3 does not violate the First Lav
8-2 ] BASIC APPLICATIONS 113

that the efficiency of Ea is greater than the efficiency of Eb must be


false, and proposition (1) is proved.
From proposition (1) a second proposition follows: (2) The effi-

ciency of all reversible heat engines operating between the same


temperature levels is the same. Proposition (2) is proved by letting

both Ea and Eb be reversible heat engines. Then by proposition (1)


the efficiency of Ea cannot be greater than the efficiency of Eb, and
the efficiency of Eb cannot be greater than the efficiency of Ea]
hence their efficiencies must be equal and the proposition is proved.

Observe that in the above reasoning no restrictions were placed


upon the type of engine or system to be used or upon any details of
operation; thus the results are completely general. No matter wdiat
new types of engine or new working substances may be invented or
discovered, so long as the Second Law holds, these two propositions
will hold. Consequently, if the efficiency of one reversible cycle (for
example the Carnot cycle) can be determined as a function of tem-
perature, the limit of efficiency for all cycles will be known. This
information is of great practical value in guiding the development of
heat power cycles.
8-2 The absolute thermodynamic temperature scale. The
efficiency of any heat engine cycle receiving heat Qi and rejecting
heat Qj is given by

_ E _ Qi-ft _~ 1 _
. Qi Qi (1)

By the Second Law it is necessary to have a temperature difference


Ut — k) to obtain work from any cycle. For a reversible cycle oper-
ating between fixed temperatures the efficiency depends solely upon
the temperatures k and U at which heat is transferred, or

V = fik, k) (2)

where / signifies some function of the temperatures. Combining (1)


and (2), for a reversible cycle operating between k and k

1-| =/«!,<*) (3)

In terms of a new function F

= E(k,k) (4)
g
114 BASIC APPLICATIONS [8-2

If h and are temperatures on an existing scale, Eq. (4) defines the


<*

function F, but if a certain functional relationship is assigned between


h, U, and Qi/Qi, Eq. (4) becomes the definition of a temperature
scale. Such a scale is independent of the properties of particular
substances, and is known as an absolute thermodynamic temperature
scale. The choice of the function relating temperature to Qi/Qt is
to some extent arbitrary but it must be consistent with the Second
Law; the necessary condition is now derived.

Fig. 8-4. Fig. 8-5.

Consider two reversible engines in series (Fig. 8-4) Ea receiving


heat from the source at and rejecting heat at k to Eb which in turn
rejects heat to the sink at
t». Equation (4) may be written for each
engine separately and for the combination of both engines.

= ^(<1. <»)
§
Qi _ Q1/Q3
Now
Q. Q2/Q,

Fdi, ij)
or Fill, U) ( 5)
F{t2j U)

Since ti, it, and k are independent, Eq. (5) can hold only if tt divides
out of the quotient; then

( 6)
8-2 } BASIC APPLICATIONS 115

where ^ a new function of h (or ii) remaining after is divided out.


is

Since is any arbitrary function, the simplest possible way to


<^(t)

define the absolute thermodynamic temperature T is to let ^(t) = T,


as proposed by Kelvin. Then definition

(7 )
T% Qt

where Ti and T2 are temperatures on the absolute thermodynamic


temperature scale, and Qi and Qt are respectively the heat quantities
exchanged by a reversible heat engine with a source at Ti and a
sink at Tt.
The definition (7) does not fix the scale completely since it gives
only the ratio of any two temperatures; the size of the degree of
temperature is arbitrary and is taken, for convenience, the same as
the conventional degree. Thus there is a Centigrade absolute scale
called the Kelvin scale and a Fahrenheit absolute scale called the
Rankine scale.
That the thermodynamic absolute temperature scale has a definite
zero point can be shown by imagining a series of reversible engines
such as shown in Fig. 8-5, extending from a source at Ti to lower
temperatures. Assuming equal temperature intervals, the work of
all engines will be equal; for if T\ — Ti = Ti — Tj it follows from

(7) that
Q\ — Qt = Qi — Qi

and Wi =
If enough engines are placed in series to make the total work output
equal to Qi, then by the First Law the heat rejected from the last
engine will be zero. By the Second Law, however, operation of a
cyclic heat engine with zero heat rejection cannot be achieved, al-

though it may be approached as a limit. When the heat rejected


approaches zero the temperature of heat rejection also approaches
zero as a limit. Thus it appears that a definite zero point exists on
the absolute temperature scale but that this point cannot be reached
without a violation of the Second Law.
It is an experimental fact that the ideal gas temperature scale
described in Sec. 3-6 is identical with the absolute thermodynamic
scale. Therefore the following relations hold:

T® = -I- 459.7
116 BASIC APPLICATIONS [8-3

+ 273.2
r® = 1.8 T®

where T signifies absolute thermodynamic temperature, t signifies

conventional temperature, K signifies Kelvin scale, and R signifies

Rankine scale.
Whenever temperature is referred to in thermodynamic analysis
the absolute thermodynamic temperature is implied, although data
are frequently given in terms of the conventional temperature scales.
By substitution from Eq. (7) in Eq. (1) the efficiency of a reversible
heat engine cycle in which heat is received solely at Ti and heat is

rejected solely at Tt is found to be

Equation (8) sets a limit to the possible efficiency of a heat engine


operating between Ti and Tt because no such engine can exceed in
efficiency the reversible engine.
It is not necessary to operate a reversible heat engine in order to
establish the relationship of the thermodynamic temperature scales
to the conventional scales. It is possible to derive certain relations
between the absolute temperature and other properties of substances,
by which the absolute temperature may be computed from measure-
ments of other properties.*
8-3 Entropy. Because temperature is defined as a function of
the heat quantities transferred in the operation of a reversible cycle
it is possible to set up, for a reversible process, a relation between
heat transferred and the properties of a system. It was shown in
Chap. 2 that under certain special conditions it is possible to express
work as a function of pressure and volume; now it will be shown
that, for a reversible process, it is possible to express heat as a function
of temperature and a new property called entropy (en'tropy). The
existence of this new property must now be proved as a consequence
of the Second Law.
Consider a reversible heat engine E which has available a large
number of sources, a, b, c, etc., at different temperatures, Tia, Tn,,
• Planck, Max, Treatise on Thermodynamics; tr., Ogg. New York: Dover,
1945; pp. 134-138. Keenan, J. H., Thermodynamics. New York; Wiley, 1941;
chap. 21. Zemansky, M. W., Heat and Thermodynamics. New York: McGraw-
Hill, 1943; chap. 9.
BASIC APPLICATIONS 117

Tu, etc., and a single sink at a fixed temperature Tj, as shown in

Fig. 8-6. Let the engine operate through a cycle la, 2a, 3a, 4a as
shown in Fig. 8-7, receiving heat Qm from source a and rejecting
heat Qta to the sink. (It will be observed that the lines representing

reversible adiabatic paths are plotted to show increasing temperature


with decreasing volume. This is a characteristic of most substances.
A familiar result of this fact is the temperature rise noted when gas

is compressed rapidly, as in a tire pump.) Next let the engine oper-


ate through cycle ib, 2b, 3b, 4b, in which the state 3b coincides with
state 2a, and the same reversible adiabatic path is followed (in the
reverse direction) for the compres.sion in cycle b as for the expansion

Fig. 8-7.
118 BASIC APPLICATIONS

in cycle a. Next, let the engine operate through cycle Ic, 2c, 3c, 4c,

in which state 3c coincides with state 2b, and cycles b and c use the

same adiabatic path through 2b in opposite directions. The cycles


must share the common paths la-2a and lb-2b because only one
reversible adiabatic path can pass through a given point such as
2o or 26. If two different reversible adiabatic paths passed through
a single point they would form, with a reversible isothermal path,
the plot of a cycle which would violate the Second Law.
From Eq. (7) the following relations hold:

Qu^Tu Qifc _ n Qic Tu


Qla Ti Qjfc T, Qu Tt
Rearranging,
Qla _ Ota Qi6 Qib Qic Qu
" (9)
Tla T, Tu>" Tt Tu Ti

Adding Eqs. (9),

Qla
Tu
^ I
^
r.6
, Qu
Tic
Qia + Q» + Qjc (10)

Now if the heat rejection processes were carried out in succession as


a single process, the heat Q2 transferred in this combined process at
temperature Ti would be

Q2 = Q20 + Qa> + Qu
Equation (10) may now be rewritten,

Qi
( 11 )

Now if the system E changes reversibly from any given state A


to any other state B (Fig. 8-8), the path of the change may be
approximated as closely as desired by a succession of alternate revers-
ible isothermal and reversible adiabatic processes. The reversible
adiabatic paths may be extended to cross a constant-temperature
path at Tj, as shown in Fig. 8-8. Since no two reversible adiabatic
paths may intersect, a series of cycles will be formed for which Eq.
(11) will hold. If number of subdivisions of A-B is increased
the
without limit, the summation of Eq. (11) may be rewritten as an
int^al, thus
8-3] BASIC APPLICATIONS 119

Sicce Ti any arbitrary temperature, the same value of Qz/Tz must


is

exist for any temperature Tz and the given process A-B. If any

other path had been followed reversibly from A to as indicated


by the dashed line in Fig. 8-8, the path C-D at Tt would be unchanged
and Eq. (12) would still hold. Hence, for a reversible process on
any path between A and B, the integral of dQ/T has the same fixed
value. This is the characteristic of a property. The property en-
tropy is therefore defined as follows: The entropy s of a system is a
quantity which satisfies the equation

ds
^ (13)

where Qnr is the heat transferred in a reversible process. Integrating,

(13a)
120 BASIC APPLICATIONS [ 8^

In words, the change of entropy in a system is the integral, for a


reversible process, of dQ/T.
Since the change in entropy between any two states and A B is
independent of the path followed in evaluating it, the entropy is a
property, but like the internal energy it has no absolute base value
at any one state. If a base value is arbitrarily assigned to the
entropy of a system at any one state the entropy at all other states
becomes fixed. The entropy of a system is usually given in units of
Btu/F^. The specific entropy of a substance is given in units of
Btu/lb In a pure substance the entropy, like other properties,
F.b..
is a function of two independent variables such as pressure and vol-

ume, pressure and temperature, or internal energy and volume.


It is desirable to dwell upon the significance of Eq. (13) by which
entropy is defined. The equation says that in a reversible process
the entropy change is measured in terms of heat; it does not say that
entropy changes only in a reversible process. The entropy change
from A to B is identical for all processes between A and B. fdQfT
is equal to the entropy change for all reversible processes between
A and B, but has different values for irreversible processes. These
different values of y* dQ/T in irreversible processes are immaterial
to the determination of entropy change, since by definition the entropy
change from A to B is y* dQ/T for a reversible process. Hence if it
is desired tocompute the entropy change for any process, reversible
or irreversible, between A and B, all that need be done is to assume
a convenient reversible path between A and B and compute the value
of y* dQ/T for this assumed reversible path.
Since the definition of adiabatic is zero heat transfer, the integral
of dQ/T will be zero in any adiabatic process. Therefore in a revers-
ible adiabatic process the change of entropy is zero, or the entropy
of the system is Such a process is therefore called isen-
constant.
tropic, meaning constant entropy. Observe that an isentropic proc-
ess is not necessarily either adiabatic or reversible; isentropic means
nothing but constant entropy. If, however, a process is both isen-
tropic and it must be adiabatic; if it is both adiabatic and
reversible
reversible it must be isentropic; and if it is both adiabatic and isen-
tropic it must be reversible.
In applications of thermodynamics to heat power engineering, the
most common use for the entropy is to locate the path of a reversible
adiabatic process. The reversible adiabatic is a convenient ideal or
84] BASIC APPLICATIONS 121

hypothetical process to use as a standard of comparison for many


actual processes such as the expansion of a working fluid in an engine
or turbine.
8-4 The inequality of Qausius. It has been pointed out that,
in an f dQ/T is not equal to the entropy change.
irreversible process,

A more specific statement is possible; / dQ/T is equal to or less than


the entropy change in any process, reversible or irreversible. To
prove this statement consider the cycle A-B-C-D in Fig. 8-8. Let
A-B be a general process, either reversible or irreversible, while the

other processes are reversible. Then for any of the elementar}' cycles
into which the large cycle is divided

dQt
,= 1 - (14)
dQ
where dQ and dQj are respectively the heat supplied at T and the
heat rejected at Tt in the elementary cycle. By proposition (1),
Sec. 8-1, the efficiency of a general cycle will be equal to or less than
the efficiency of a reversible cycle, thus

1 IV 1 (15)

Then (16)
dQo xdQi/rw

or, since
/m ^ r
\dQ2 Aev T 2

Al
(17)

Then, since all factors are positive.

dQ _ dQi
(18)
r ^ Tt

for any process A-B, reversible or irreversible. But for a reversible


process

» rfQrev dQi
- rr ^ rp (19)
1 I 2

Hence for any process A-B


122 BASIC APPLICATIONS

Then for any cycle, integrating around the cycle,

/f </*
or
/^<0 (21)

since entropy is a property and has zero change in a cycle.

Equation (21), called Ihe inequality of Clausius, is useful as a cri-

terion of reversibility, since in a reversible cycle the equality will hold,


while in an irreversible cycle the inequality will hold. Moreover a
cycle in which the integral is greater than zero violates the Second
Law and is impossible.
8-5 The temperature-entropy plot. Since entropy is a prop-
erty, it may be used as a coordinate in a graphical representation of
the property changes in a process or cycle. The temperature-entropy
plot is particularly useful because, for reversible processes, areas upon
this plot are directly proportional to heat quantities. Thus the
temperature-entropy plot permits visualizing heat quantities, just as
the pressure-volume plot permits visualizing work quantities.

Fig. 8-9. Temperature-entropy plot of a process.

Consider a process A-B for which the path is plotted in Fig. 8-9
upon the temperature-entropy plane. The area under the curve A-B
is proportional to ST ds. By the definition of entropy, Eq. (13),

Tda^dQr„ (22)

or fTd8 = Q,^ (23)


8-6 ] BASIC APPLICATIONS 123

Hence if A-B is a reversible process, the area under the curve is

proportional to the heat transferred in the process. If A-B is an


irreversible process, the area under the curve does not represent the
heat transferred.
Some examples of representation upon the T-s plane follow.

Reversible ConstmtrTemperature Process: In this case, Fig. 8-10,

Q = fTds = TAs

Fig. 8-10- Constant-temperature process.

Reversible Adiabatic Process: In this case. Fig. 8-11,

Fig. 8-11. Reversible adiabatic process.

Q=0= Jt ds or ds = 0

Carnot Cycle. The Carnot cycle is conveniently shown in the


T-s plane because it always forms a rectangle in that plane. In
124 BASIC APPLICATIONS [8-6

Fig. 8-12 a-b-c-d is a Carnot cycle. The area under the


reversible constant-temperature path a-b, represents the heat sup-
plied. The area c-m-n-d represents the heat rejected. Hence the
area a-b-c-d enclosed by the paths of the cycle represents the ncf

heat transferred, which is equal to the net work of the cycle. The
T-s plot of any reversible cycle will enclose an area proportional to
the net work of the cycle. Such plots are useful in comparing the
work outputs and efficiencies of various ideal reversible cycles.
For the Carnot cycle, Fig. 8-12,

( T s T, >

0*0 Jo«o
S* CONST. I^S- CONST.

d T.T* c

-JL in

Fig. 8-12. Carnot cycle.

0i = Ti As

Qi ^ As

W^net ” 0l “ Q =
2 (^1 “ (24)

8-6 The availability of energy. One of the basic general prob-


lems of heat engineering is to obtain the maximum work
output from
a cycle with the minimum heat input. The theoretical maximum
work output of a heat engine supplied with a given quantity of heat
under given conditions is called the available energy supplied or the
available part of the energy supplied. The remainder of the heat
supplied is called unavailable energy. These terms refer solely to the
possibility of work production by a cyclic heat engine.
The maximum efficiency of any heat engine working between the
two fixed temperatures Ti and is given by
8- 6 ] BASIC APPLICATIONS 125

Ti-T,
( 8)
T1
Therefore if Ti is given, the maximum efficiency will be obtained by
making 7*2 as small as possible. But, for practical cases, Tt must be
the temperature of some naturally
available body which can absorb
a large amount of heat without appreciable temperature rise. Such
bodies are the atmosphere, ocean, lakes, and rivers. The tempera-
tures of these bodies vary from place to place and from time to time
so it is customary to reason in terms of a “temperature of the sur-
rniindinir.s” which is a hvnothetical averaire temnerature.

z *1

HW
“"iKOi
/

>-dQ,
_J '

T*, Tcnipcrotufc of
the surroundings

s
k-ds

Fig. 8-13. Availability of energy.

Consider a finite process a-b, Fig. 8-13, in which heat is supplied


reversibly to a heat engine. Taking an elementary cycle, the engine
receives heat dQi at temperature T\, rejects heat to the surroundings
at temperature To, and goes between these temperatures by reversible
adiabatic processes. For the elementary reversible cycle the work
will be a maximum under the given conditions; it is given by

The quantity of energy {dQ\


^ —
dQ,

dTFmax) which
= dOi - dQ,

must necessarily be
(25)

rejected as heat is said to be unavailable for work. The unavailable


energy for the elementary cycle is then
E*
126 BASIC APPLICATIONS [ 8-6

dQi-dTr^ = ^®dQi (26)

If the heat engine receives heat through the whole process a-b, and
rejects heat at To, the maximum work that can be obtained, or the
avaiUAle energy, for the finite cycle will be the sum of the elementary
quantities of Eq. (25), or

(27)

The unavailable part of the energy supplied is

(28)

Then, since To is constant,

Qa, - W„ T, # r,(St - S«) (29)

The unavailable part of the heat supplied to a heat engine is equal


to the product of the lowest temperature at which heat may be
rejected and the entropy change of the system during the process of
supplying heat. Figure 8-14 is a graphical presentation of this fact.
The fact that the possible efficiency of heat engines increases with
increasing temperature of heat supply may be expressed in terms of
the availability of energy. The following proposition will be demon-
strated in order to show the connection.
M] BASIC APPLICATIONS 127

Whenever heal is transferred through a finite temperature difference


Con-
there is a decrease in the availability of the energy so traneferred.

sider two energy and To respectively. Let a revers-


reservoirs at Ti
ible heat engine operate between Ti and To receiving heat Qi, = Ti As,

Fig. 8-16.

and rejecting heat Qt, =ToAs, as illustrated on the temperature-


entropy plane in Fig. 8-15. The work of the reversible engine will
also be the available part of Qi.

F = Qi - Q* = (T, - To) As (a)

Now assume that the heat Qi is transferred through a finite tem-


perature difference from the reservoir at Ti to the engine at Tj,
which must be lower than Ti. The availability of Qi as received by
the engine at T[ may be found by allowing the engine to operate
reversibly in a cycle between TJ and To, receiving Qi at Ti and
rejecting Q2 at To, as shown in Fig. 8-16. The same quantity of
heat is supplied at TJ as at Tj; then since TJ < Ti and

Qi = Ti As = T[ As' (b)

it follows that As' > As (c)

But Qo s To As and Qj = To As'; hence

Qfs > Qt, (d)

Therefore the work of the reversible engine in the second case is less

than in the first case.


128 BASIC APPLICATIONS [M

Fig- 8-16. Loss of availability.

If ' = 0i - Q2 < 0i - Q* = 1^’ (e)

Whenever heat is allowed to flow irreversibly (through a finite tem-


perature difference) a certain quantity of energy is lost from the
category of possible sources of work. The magnitude of the ‘dost
work” is given by

W-W' = {Qi- Q2) - (0i - Qi) = Q2 -Q 2 = To(As' - As) (f)

where To is the lowest usable temperature for heat rejection, and


(As' — As) is the additional entropy change in the system while re-
ceiving heat irreversibly, compared to the case of reversible transfer
from the same source.
8-7 Entropy and reversibility. There is a direct relation be~
tween entropy change an isolated system and the reversibility of
in
processes in that system. An isolated system is a system which
cannot ex;change energy in any way with the surroundings. Obvi-
ously any process may be made to occur in an isolated system by
including in the system everything affected by the process.
By definition every process in an isolated system must be adia-
batic; hence for such processes

( 30)

But by Ekj. (20) (Sec. 8-4), for any process in any system
S-7] BASIC APPLICATIONS 129

then for an isolated system


As > 0 (31)

Equation (31) is a statement of the Principle of Increase of Entropy.


By this principle an isolated system in a given initial state can change
only to states of the same or greater entropy; but by the First Law
an isolated system must have constant internal energy, therefore an
isolated system may change only to states having the same internal
energy and the same or greater entropy than the initial state. The
internal energy here considered is the most general internal energy
and includes effects of motion and gravity (and surface tension, elec-
tricity and magnetism if they were to be permitted in any particular

case).
The principle of increase of entropy provides a criterion of revers-
ibility for, in an isolated reversible system, the entropy must be
constant, all processes being both adiabatic and reversible. There-
fore if the entropy of an isolated system increases, the process taking
place must be irreversible.

Example: It is easily shown that a transfer of heat through a finite tem-


perature difference results in an increase of the entropy of the total system
involved. Let heat Q, be transferred from an energy reservoir 1 at T*! to
an energy reservoir 2 at Tt, when Ti > Ti. The entropy changes in the
individual reservoirs are respectively

Asi = (a)

and A*, = ^ (b)

The entropy change Asi is negative because heat is transferred from the
reservoir at Ti. In each case the entropy change is given by Q/T because
the entropy change computed by assuming a reversible process between
is

the same end states as the actual process. Another way of expressing this
is to say that the processes in the separate reservoirs are assumed internally

reversible. The process in the combined system, however, is irreversible


because

< T, (c)

Combining (a), (b) and (c) the result is obtained

m+ aa, = ^;-|>o (d)

or for the system comprising both reservoirs


130 BASIC APPLICATIONS [8-8

As = Asi ^ 0 (0 )

Thus the irreversible process involves an entropy increase for the isolated

system.

8-8 Examples of second law problems. This section contains


examples of the use of the ideas and principles presented in this and
the preceding chapter.

Example 1. It is proposed to obtain power from the hot surface water


of tropical seas using the cold water from the depths as a sink for heat
rejection. The surface water is at 85®F, the deep water at 50®F, In the
light of the Second Law is such a scheme possible? If so, what is the maxi-
mum thermal efficiency possible under the Second Law?
Solution: The scheme is possible since the Second Law permits net work
to be produced by a heat engine if a temperature difference is available.
The maximum possible efficiency is given by

TjlDMX —

r, = 85 + 460 = 545*^

r* = 50 -f- 460 = 510*

W _ J5
- 545 0.064 or 6.4 percent

An actual plant, being subject to losses and irreversibility, would have a


much lower efficiency.

Example 2. A proposed steam power plant will provide for supplying


heat to steam at temperatures up to 1050°F. The temperature of heat
rejection is about 90®F. It is stat^ that the thermal efficiency of the plant
will approach 34 percent. Does this seem reasonable in the light of the
Second Law?

Solution: The maximum possible efficiency for the given temperature


range is

W = ^ = 0.63 = 63 percent.

The predicted efficiency does not seem unreasonably high since it is only a
little more than half the maximum possible efficiency. Actual plants seldom
approach the maximum efficiency, partly because most of the heat is actually
supplied at temperatures well below the maximum temperature of the cycle.

Example 3. How much does the entropy of a pound of air change when
the air is heated reversibly from 1 atm pressure, 50®F to 1 atm, 150®F, if

the specific heat of air at constant pressure is 0.240 Btu/lb F?


8-8 ] BASIC APPLICATIONS 131

SoLxmoN; Since the given states are at the same pressure, assume a
reversible constant^pressure process between them and compute the entropy
change for the assumed process. (Note that a constant-pressure process
was not given in the example.)

As
T
For a reversible constant-pressure process, from Eq. 3-2,

dQ = mcp (U
then

A« = f^ As
= nu:,J^^mc,OnT,-]nTy)

= 1(0.240) (In 610 - In 510)

= 0.0431 Btu/F„b.

Example 4. Suppose the air inExample 3 is changed from the initial


state to the final state of that problem by an adiabatic process involving
work against friction. How much does the entropy of the air change in
this irreversible process?

SoLxmoN: The same as in Example 3. If this is not clear, review the


whole discussion of entropy.

Example 5. In a steam boiler, hot gases from a fire transfer heat to


water which vaporizes at constant temperature. In a certain case the gases
are cooled from 2000 ®r to 1000®F while the water evaporates at 400®F. The
specific heat of the gases is 0.24 Btu/lb F, and the latent heat of the water
is 826.0 Btu/lb. All the heat transferred from the gases goes to the water.
How much does the total entropy of the combined system of gas and water
increase as a result of the irreversible heat transfer? Obtain the result on
the basis of one pound of water evaporated.

Solution: The entropy change of the combined system is the sum of the
entropy changes of its parts. The individual entropy changes of the gas
and water systems can be found by assuming in each case a reversible process
between the actual end states of the individual system. For the water, since
temperature is constant,

A«h.o = f = I5 = 0-965 Btu/R


For the gas, per pound of water evaporated, Q = —826 Btu. Also

Q = m CpdT - mCp AT,*,


J*
since Cp is constant. Then, for the gas,
132 BASIC APPLICATIONS [8-8

^
= /«v f°

= 0.826 (In 1460 - In 2460) = -0.431 Btu/R

The total change is

As = As\w + ASgaa = 0.534 Btu/R


Observe that the total change of entropy is positive, as it must be by the
principle of increase of entropy.

Example 6. For the process of Example 5 find the increase in unavail-


able energy due to the irreversible heat transfer. Assume the temperature
of the surroundings is 80°F.

Solution: Sketch on the T-s plane the two assumed reversible processes,
heat transfer from the gas and heat transfer to the water. Use a basis of
one pound of water evaporated. The gas is cooled from 1 to 2 by trans-
ferring heat to the evaporating water which changes from a to 6. If the
heat from the gas had been transferred reversibly to a reversible heat engine
the heat engine would have received the heat along the path 2-1 because
reversible heat transfer requires zero temperature difference as each incre-
ment of heat is transferred; the heat engine would have rejected at To a
heat quantity To Asn or — To Asg^. If a reversible heat engine received the
8-9] BASIC APPLICATIONS 133

heat reversibly at 400®F the heat engine would have rejected at To a heat
quantity To Aso6 or To AshsO- The increase in unavailable energy due to the
transfer of heat from the gas to the water is then

ToAsh^ — (— To ASgaa) = ToCAshiO + ASgaa) = To As

where As applies to the combined system as in Example 5. The numerical


result is then
To As = 540(0.534) = 289 Btu

It is of interest to note that the available energy in the heat transferred


from the gas was, at the temperature level of the gas,

Q - ToAsn = 826 - 540(0.431) = 594 Btu.-

Therefore the irreversible transfer of heat has resulted in a loss of 289/594,


or almost 49 percent of the available energy. The situation here depicted
is typical of what is happening constantly in even the best power plants.
It shows that, despite the great advances so far made, there is still room for
improvement in basic heat power operations.

8-9 Summary, No heat engine operating between fixed temper-


ature levels can be more efficient than a reversible engine operating
between the same temperatures.
The absolute thermodynamic temperature scale is defined by the
relation

Qi
Q
where Qi is the heat received from a source at Ti, and Q2 is the heat
rejected to a sink at T2 by a reversible heat engine
,
which exchanges
heat only at Ti and ^2 Equation (7) establishes
. the ratio of any
two temperatures and the zero point of the scale, but does not estab-
lish the size of the degree of temperature, which is taken the same
as the conventionarFahrenheit degree (or the Centigrade degree) for
the Rankine scale (or the Kelvin scale). The absolute thermody-
namic scale is identical with the ideal gas scale discussed in Chap. 3.
In thermodynamic analysis temperature implies the absolute themo-
dynamic scale, although data are frequently presented with reference
to the conventional temperature scales.
By the definition of the temperature scale, the efficiency of a
reversible heat engine recei^ung heat solely at Ti and rejecting heat
solely at Tz is given by
134 BASIC APPLICATIONS [8-9

(8)

This equation sets an upper limit on the possible eflSciency of any


heat engine working between Ti and Tj.
The entropy s of a system is a quantity which satisfies the relation

ds = ^ (13)

The change in entropy between any two states of a system is meas-


ured by the integral of dQ/T for a reversible process between the two
states. Entropy a property of a system but has no definite value
is

until a value is assigned for one state of the system; then the values
at all other states become fixed. In an irreversible process the change
of entropy is the same as for a reversible process between the same
end states, but the integral of dQ/T is less for an irreversible process
than for a reversible process between the same end states.
A process in which the entropy of a system remains constant is
called isenlropic. A process in which heat transfer is zero is called
adiabatic. A reversible adiabatic process is isentropic but an adia-
batic process is not necessarily isentropic, nor is an isentropic process
necessarily reversible or adiabatic. Of the three characteristics, re-
versible, adiabatic, and isetitropic, no one necessarily implies the
others, but any two together necessarily require the third.
Becau^ entropy a property, the integral of ds around a cycle
is

must always be zero. Then, for a reversible cycle, the integral of


dQ/T around the cycle is zero. For any cycle the inequality of
Clausius holds:

/^ < 0 (21 )

The plot of a reversible process upon the cocxtlinates of tempera-


ture and entropy has the useful property that the area between the
path of the process and the line of zero temperature is proportional
to the heat transferred in the process. The area enclosed within the
plot of a reversible cycle on the T-s plane is proportional to the
work oi the cycle.
Given a certain process in which heat is supplied to a cyclic engine
and given that heat is to be rejected at the temperature of the sur-
roundings, the Second Law determines the maxiTnnm possible work
BASIC APPLICATIONS 135

that can be obtained from the engine. - This quantity of work is

called the available part of the heat supplied, or the available energy
supplied' to the engine. The remainder of the heat supplied is called
unavailable. If the lowest usable sink temperature (temperature of
the surroundings) is To the unavailable part of the heat supplied is

given by To As where As is the increase in the entropy of the system


during the process of supplying heat. The available portion of the
heat supplied to a heat engine decreases as the temperatiu^ at which
the heat is supplied decreases. 'Whenever heat flows through a finite

temperature difference some energy becomes unavailable.


In any isolated system the entropy may increase or remain con-
stant but it cannot decrease. This is the principle of increase of
entropy. By this principle, if a system is so chosen as to include all

bodies affected by a process, the entropy of the system will increase


if the process is irreversible, and will remain constant if the process
is reversible.

PROBLEMS
8-1. A cyclic heat engine receives heat solely from a source at lOOOT and
rejects heat solely to cooling water at lOO^F. What is the maximum possible
efficiency of the heat engine?
8-2. On the basis of the First Law fill the blank spaces in the following table
of hypothetical heat engine cycles; on the basis of the Second Law classify each
cycle as either irreversible, reversible, or impossible.

Rate of
Work
Cyde Temperatures Rates of Heat Flow Output Efficiency

Source Sink Supply Rejection

(a) 640T 40°F 100 Btu/sec 55 Btu/sec hp

(b) lOOOT 100“F Btu/min 1000 Btu/min kw 65%

(c) IQOO^R 600®Il Btii/hr Btu/hr 35 hp 60%

(d) 1200‘‘R 630®R 10000 Btu/hr Btu/hr 1 kw


8-3. The latent heat of fusion of water at 32°F is 143 Btu/lb. How much
does the entropy of a pound of ice change as it melts to liquid in each of the
following ways? (a) Heat is supplied reversibly to a mixture of ice and water
at 32‘’F. (b) A mixture of ice and water at 32”F is stirred by a paddle wheel.
8-4. Liquid water flows through a hydraulic turbine in which friction causes
the water temperature to rise from 65* to 67*F. If there is no heat transfer
136 BASIC APPLICATIONS
in the turbine the entropy of the water change in passing through
how much does
the turbine? (Note that liquid water is essentially incompressible, then compute
the entropy change for a reversible constant-volume process between 66®F and

8-5. pounds of water at 140“F are mixed with 3 pounds of water at


Two
40'’F in a constant-pressure process at 1 atmosphere. Find the increase of
entropy of the total mass of water due to the mixing process.
8-6. In a Carnot cycle heat is supplied at 380°F and rejected at 80**F; the
working fluid is water, which, whije receiving heat, evaporates from liquid at
380®F to steam at 380®F. According to the steam tables the entropy change
for this process is 1.0059 Btu/lb R. (a) If the cycle operates on a stationary

mass of 1 Ib of water, Fig. 7-8, how much work is done per cycle, and how much
heat is supplied? (b) If the cycle operates in steady flow, Fig. 7-9, with a power
output of 25 hp, what is the steam flow rate?
8-7. A heat engine receives reversibly 100 Btu of heat per cycle from a source
at 540?F, and rejects heat reversibly to a sink at 40°F; there are no other heat
transfers. For each of three hypothetical amounts of heat rejected, in (a), (b),
and (c), below, compute the cyclic integral of dQ/T. From these results
show which case is irreversible, which reversible, and which impossible, (a) 50
Btu/cycle rejected; (b) 75 Btu/cycle rejected; (c) 25 Btu /cycle rejected.
8-8. As shown in Fig. 7-10 a heat engine cycle may be operated in reverse,
as a heat pump. Show, by the method used in Sec. 8-1, that no heat pump
cycle working between two reservoirs at fixed temperatures can have a larger
ratio of Qi to Wnet than a reversible cycle.
8-9. Show by pump cycle working
the inequality of Clausius that no heat
between two reservoirs at fixed temperatures can require less net work per unit
of heat received than a Carnot cycle.
8-10. In a certain process a vapor, while condensing at 800®F, transfers heat
to water evaporating at 500®F. The resulting steam is used in a power cycle
which rejects heat at lOO^’F. What fraction of the available energy in the heat
transferred from the process vapor at 800®F is lost due to the irreversible heat
transfer to the water at 500°F?
8-11. Exhaust gases leave an internal combustion engine at 1500®F and
1 atmosphere, after having done 400 Btu of work per pound of gas in the engine.

The specific heat of the gases at 1 atmosphere is 0.26 Btu/lb F, and the tempera-
ture of the surroundings is SO^F, (a) How much available energy per pound of
gas is lost by throwing away the exhaust gases? (The available energy is the
work of a reversible heat engine which receives heat reversibly from the exhaust
gases, while cooling them to 80°F, and rejects heat at 80®F.) (b) What is the
ratio of the lost available energy to the engine work?
Chapter 9

PROPERTIES OF PURE SUBSTANCES

In the foregoing chapters the basic laws of thermodynamics and


some of their general applications have been presented. In subse-
quent chapters it is desired to apply this information to the analysis
of specific engineering processes. However, to obtain quantitative
results, knowledge of the properties of the substances involved is

necessary. This chapter, therefore, contains a general description


of the relations among properties, and the following chapteils give
the common methods of presenting and using data.
9-1 Methods of presenting property relations. The proper-
ties usually needed in applying thermodynamics to mechanical engi-
neering problems are pressure, temperature, volume, internal energy,
enthalpy, and entropy. If the portions of the internal energy due
to gravity, observable motion, surface tension, electricity, or mag-
netism are separately accounted for, a pure substance has in ^neral
only two independent properties. If pressure and specific vblume,
for example, are fixed then all the other properties become fixed.

This makes it possible to express any property of a substance as a


function of two other properties or to plot lines of constant value for
any one property on coordinates of two other properties.* Of the
properties listed above, the most evident to the senses are pressure,
volume, and temperature; therefore these properties are often used
to describe the states that a particular substance may assume. The
equation relating pressure, volume, and temperature is called an
equation of state. Some other information in addition to the equa-
tion of state is necessary to determine the remaining properties.
An example of a simple equation of stete which is satisfactory
for many gaseous substances is the perfect gas equaiion
*
See the appendix for tables of properties, and envelope charts at back of
book for plots.

137
138 PURE SUBSTANCES [9-2

pv^RT (1 )

inwhich p is the pressure, v the specific volume, T the temperature,


and R a constant for the particular gas. The same information may
be shown graphically, as in
Fig. 9-1. In the general case
of substances which may pass
from solid to liquid and to
gaseous states, no such simple
equation is even approxi-
mately correct. The best way
to describe the general nature
of the property relations is by
a graphical exposition. In
order to make the discussion
specific the important sub-
stance water will often be used

Fig. 9-1. Perfect gas equation of state. as an example. The char-


acteristics of other substances
are similar to those of water but s le peculiarities of water will be
noted.
and gas phases. Assume that a unit mass of
9*2 Solid, liquid
ice (solid water) at0°F and 14.7 psia fills the clearance space of a
cylinder and piston machine. Fig. 9-2. It in HprippH ijn fmcA the
changes which occur in the mass of water
as the temperature is increased while the 14.7 ptio

pressure is held constant. Use


be madewill
of two plots. Fig. 9-3,on coordinates of
pressure vs. volume and temperature vs.
volume respectively. Let heat flow slowly
to the ice so that its temperature is always
uniform; then changes in volume and tem-
perature will be observed as plotted in Fig.
9-3 (data from Keenan and Keyes):
Fig. 9-2.
1-2 the temperature of the solid in-
creases from O^F to 32°F with a small increase in volume,
2-3 the solid melts to liquid at constant temperature with a iwimlH

decrease in volume,
9-2] PURE SUBSTANCES 139

3-

4-

Fig. 9-3. Heating water at constant pressuro.

4 the temperature of the liquid increases from 32®F to 212‘’F;


the volume decreases to a minimum at 39°F and then increases,
^5 the liquid boils to a vapor or gas at constant temperature
with a large increase in volume.
;

140 PURE SUBSTANCES [ 9 ^

Beyond 5 both the temperature and the volume of the vapor


increase * The decrease in volume during melting is a peculiarity
of water. Most substances show a continuous increase in volume
as indicated in Fig. 9-4. In a slow cooling process the paths of Figs.
9-3 and 9-4 would be followed in the reverse direction.

Fig. 9-4. Constant-pressure heating; general substance.

9-3 Saturation states. In the process described above the water


homogeneous states of aggregation, or phases
existed in three different
solid between 1 and between 3 and 4; and gas beyond 5.
2; liquid
Between 2 and 3 the water was not homogeneous but consisted of a
mixture of t^vo phases, solid and liquid; between 4 and 5 it consisted
of a mixture of liquid and gas. The states indicated by points 2, 3,
4, and 5 are known as scUuration states. A saturation state is a state
from which a change of phase may occur without change of pressure
or temperature. Thus state 2 is a saturated solid state because the
solid can change to liquid at constant pressure and temperature from
state 2. State 1 is not a saturation state because in a constant
pressure process from state 1 the temperature will change before
melting occurs. States 3 and 4 are both saturated liquid states; in
state 3 the liquid is saturated with respect to solidification, whereas
in state 4 the liquid is saturated with respect to vaporization. State 5
* The terms vapor and gas are often not rigorously distinguished. Vapor is
generally used for sUtes not far from saturation; gas is sometimes restricted to
temperatures above the critical temperature. See Sec. 9-3 for the meaning of
saturadon and critical temperature.
9^] PURE SUBSTANCES 141

is a saturated vapor state because, from state 5, the vapor can con-
dense to liquid without change of pressure or temperature.
Experience shows that for any given substance, in saturation
states at various pressures, there is a definite temperature correspond-
ing to each pressure. Thus if the process 4-5 in Fig. 9-3, correspond-
ing to vaporization, is plotted for a number of different pressures,
a, b, c, etc., a series of saturation states 4 and 5 appears as in Fig. 9-5.
A curve drawn through the points 4 is called the saturated liquid
line, and a curve drawn through the points 5 is called the saturated

Fig. 9-5. Saturation curve on p-V and T-V coordinates.


142 PURE SUBSTANCES [&4

vapor line. It is found that when these two curves are extended to
higher pressures and temperatures they do not continue indefinitely
but turn toward each other and join in a continuous curve at a state
called the critical state. Thus the two saturation lines are actuaUy
a ainglft line forming what is often called the saluraiion dame or
vapor dome.
In the plots of Fig. 9-5 each point in the plane represents a par-
ticular state of the substance considered. States to the right of the
saturated vapor line are called superheated vapor states or gas states.
States to the left of the saturated liquid line are called compressed
liquid states. States between the saturation lines are called mixture
states or wet vapor states. The pressure and temperature at the
critical state are called the critical pressure and the critical temperature
respectively.
It should be understood that at pressures higher than the critical

pressure the phenomenon of boiling does not exist. At such pressures


there is never any time in a constant-pressure heating process when
the temperature remains constant, and liquid and gas become distinct
phases in a mixture. Instead, the fluid, remaining always homo-
geneous, simply becomes gradually less dense, with no distinction
between liquid and gas. Sometimes the arbitrary rule is followed
that the liquid becomes a gas when it passes the critical temperature.
9-4 Equilibrium between phases. The saturation states are
states in which two different phases of a substance can exist together
in equilibrium. If liquid water at 1 atm and 212®F is brought into

contact with water vapor at 1 atm and 212®F there is no disturbance


of either phase by the other. Since pressures are equal no work is
.transferred; since temperatures are equal no heat is transferred. If
liquid water at 1 atm and 200°F is brought into contact with water
vapor at 1 atm and 212‘’F there is equilibrium of pressure but the
difference of temperature will cause heat transfer until the two phases
have equal temperatures. Assuming the pressure held constant by
some external restraint, the only temperature at which the liquid
and vapor can reach equilibrium is 212°F; therefore sufficient vapor
must condense to raise the temperature of the liquid to 212°F. Simi-
larly, if liquid and vapor at different pressures are placed in contact,

they must assume the same pressure, exchanging work or heat as


necessary to accomplish this.

It should be apparent that all this discussion of property relations


94] PURE SUBSTANCES 143

involves an assumption that the substance is in equilibrium through-


out its mass. A diagram or plot showing the conditions imder which
the various phases of water can be in equilibrium with each other is

given in Fig. 9-6. This shows the relations between saturation tem-
peratures and saturation pressures for water, plotted on coordinates
of temperature vs. pressure. It is seen that the solid-liquid equi-
librium, or freezing, temperature is little affected by pressure; the

P (pslo)

Fig. 9-6. Equilibrium among phases of water.

variation is only about — 1®F per 1000 psi pressure increase up to


pressures of several thousand psia.* Usually such a variation is

negligible and the freezing temperature is considered independent of


pressure. The liquid-vapor equilibrium, or boiling, temperature is

seen to be very much a function of pressure; this is of great import-


ance in the design of steam power apparatus.
The liquid-vapor saturation line does not extend indefinitely in
either direction. It is limited at the upper end by the critical state,

*For most Bubstances the variation is of opposite sign, but still of small
magnitude. For water above 2000 atm the freezing temperature increases with
pressure.
144 PURE SUBSTANCES [(MS

above which no boiling process occurs. At the lower end the cir-
cumstances are as shown in Fig. 9-7 which is an enlarged plot of part
of Fig. 9-6. The enlarged plot shows that, whereas gaseous and
liquid water can exist together in equilibriiun above 32“F, gaseous

and water can exist together in equilibriiun below 32“F. The


solid
direct change from solid to gas is called sublimation. The point at
the intersection of the boiling line and the freezing line is called the
triple point because at that state the three phases of gas, liquid, and
solid may exist together in equilibrium. Since the triple point for

p { p*lo)

Tig. 9-7. Triple point for water.

water is at a pressure of about 0.09 psia while the thermometric


calibration point of 32'’F is the melting point at 14.7 psia pressure,
it follows that the triple point for water is not exactly at 32°F, but
the difference is insignificant for engineering purposes.
The triple point for carbon dioxide is at about — 70®F and about
5 atm pressure. Hence when solid COj is exposed to 1 atm pressure
it begins to change directly to gas, and cools toward an equilibrium
temperature below — 70®F. This phenomenon is applied commer-
“dry ice” for refrigeration.
cially in the use of
9-5 Phase diagram on the pressure-volume plane. A t3rpical
phase diagram on pressure-volume coordinates is shown in Fig. 9-8.
9-5] PURE SUBSTANCES 145

Fig. 9-8. Presaure-volume phase diagram.

The points and 5 correspond to like points in Fig. 9-4. The


1, 2, 3, 4,

scales of the diagram have been distorted because in true scale the
liquid and solid regions would be microscopic compared to the vapor
regions.
In Fig. 9-9 is shown the liquid-vapor portion of Fig. 9-8 with

Fig. 9-9. Isothermal lines.


146 PURE SUBSTANCES [04

some typical lines of constant temperature superimposed. In the


gas region, particularly with cur/es such as a, for temperatures well
above the critical, the constant-temperature curve approximates a
rectangular hyperbola according to Eq. (1). The curves like 6, for
temperatures below the critical, consist two approximately hyper-
of
bolic segments in the liquid and vapor regions joined by a straight
segment parallel to the v axis in the mixture region. As the tempera-
ture increases, the length of the straight segment decreases until at
the critical temperature, curve c, there is only a point of tangency
with the horizontal at the critical state.
9-6 Enthalpy and internal energy of the pure substance.
In the constant-pressure heating process described in Sec. 9-2 the

heat supplied to the water is equal to the change of enthalpy of the


water. Hence by making a continuous measurement of the heat
supplied, data could be obtained for a plot of enthalpy vs.
tempera-
ture as shown in Fig. 9-10. The similarity to the T-t; plot.
Fig. 9-4,
is obvious. The enthalpy change of the solid between 1 and 2 is
given (Sec. 5-7) by

The enthalpy change during the melting process from 2 to 3 is the


loient heat of fusion. The enthalpy change of the liquid between
3 and 4 is given by
M] PURE SUBSTANCES 147

The enthalpy change during the boiling process from 4 to 5 is the


latent heat of vaporization. The enthalpy change of the vapor beyond
5 is given by

The variation of enthalpy at constant pressure can be obtained,


as indicated here, from a knowledge of specific heats and latent
heats; but to have a complete knowledge of the enthalpy it is neces-
sary to know how it changes when the pressure changes. This in-
formation may be obtained in several ways, of which two will be
described briefly.
The Joule-Thomson experiment. Fig. 9-11, consists in passing a
fluid in steady flow through a porous plug in a duct under such con-

Pi< Pi

Vi* 7,

Zf Zi

0 » 0
W,* 0

Fig. 9-11. Joule-Thomson experiment.

ditions that heat transfer, kinetic energy change, and potential energy
change are negligible. Such a process is called a throttling process.
In the steady flow energy eqtuUion for this process all terms except
the enthalpies will disappear leaving

h\ — hi

By the throttling experiment, states having the same enthalpy may


be identified at different pressures.
Another method of obtaining the enthalpy difference between two
states at different pressures is to carry out a frictionless constant-
volume beating process in which the heat supplied and the tempera-
ture change are measured. Such an experiment will yield the change
in internal energy at constant volume since, for such a process
148 PURE SUBSTANCES [M

(Au), = Qv

If the p, V, T relation is known, the pressure change can be found,


and the enthalpy change follows from the definition h = u + pv.
Considering the p-v plot. Fig. 9-12, any two points such as A and
B can be connected by a combination of two constant-pressure lines,
A-C and B-D, with a particular constant-volume line C-D. Then,
if the specific heats are known, the enthalpy change from A to B

can be obtained by computing for the three steps A-C, C-D, D-B

Fig. 9-12.

in succession. On the other hand, A and B may be connected by a


pair of constant-pressure lines and any constant-enthalpy line found
by a throttling experiment. If E-F is such a line, the enthalpy change

from A to B can be obtained by computing for the path A-E-F-B.


By the means described, the enthalpy can be determined for all
states of a substance. From the definition of enthalpy the internal
energy can also be obtained for Since there is no absolute
all states.

scale for internal energy or enthalpy, an arbitrary zero point may be


selected for one of them. In the steam tables the enthalpy is taken
as zero at 32°F saturated liquid; then the value of internal energy at
that state is fixed at (— pv) by the definition of enthalpy.
9-7 Entropy of the pure substance: temperature-entropy
plot. As shown in Sec. 8-3 the entropy change between any two
states of a system can be computed by assuming a reversible process
9-7] PURE SUBSTANCES 149

connecting those states and computing fdQ/T for this reversible


process. If the constant-pressure heating process of Sec. 9-2 be as-
sumed reversible, the entropy change can be computed from knowl-
edge of the specific heats, the latent heats, and the temperatures.
The entropy change from 1 to 2 in Fig. 9-10 is given by

/.r. c, AT

and the entropy change from 2 to 3, since T is constant, is given by

latent heat ^3 —— ^2

From such computations a plot, Fig. 9-13, of a constant-pressure


process may be constructed on the coordinates of temperature and
entropy.

Fig. 9-13. Temperature-entropy plot.

A complete knowledge of the entropy requires not only a knowledge


of its variation in constant-pressure processes but also a knowledge of
its variation from one pressure to another. This information may be
obtained by the use of a constant-volume process to connect the
constant-pressure processes, exactly as was shown in the case of the
enthalpy. For the constant-volume process

Cv dT
150 PURE SUBSTANCES 9^
[

By such means it is possible to determine a value of entroj^ corr©*


spending to each point of the jm> phase diagram, Fig. 9-8. Then all
the boundaries between phases may be replotted on the T-s plane,
giving a result like Fig. 9-14. The path 1, 2, 3, 4, 5 in Fig. 9-14
corresponds to the similarlymarked path in Figs. 9-4 and 9-8. Tem-
perature entropy plots, to scale, for air and for carbon dioxide are
folded in the back cover envelope, and a similar chart for water
appears in the steam tables of Keenan and Keyes. Figure 9-14 is a
t5rpical diagram not to scale for any substance.
,

SOLID

Fig. 9-14. Temperature-entropy phase diagram.

9-8 General property relations for pure substances. Since


a pure substance has only two independent properties, it must be
possible to express any property as a function of two others. But
this general fact provides no clue as to the nature of the relations for
any particular substance. Hence the p, v, T relation for a substance
is determined by making measurements of pressure, volume, and tem-

perature. In the same way other properties such as specific heats


may be obtained as functions of pressure and temperature. How-
ever, not all properties need be measured directly as functions of
other properties because there exist certain general relations which
apply to the properties of all pure substances. These general rela-
94] PURE SUBSTANCES 151

tions, when combined with empirical data, permit the computation


of properties not directly measured.
Of the many relations which must hold among the properties of
all pure substances some are simpiy definitions; for example

h = u+ pv

and c. =

Other relations are derived from the application of the First and
Second Laws to a generalized process in a pure substance. Consider
any reversible process in a system of unit mass of a pure substance
in the absence of effects of motion and gravity. From the First Law

Q = Au + W
or for a small change
dQ = du + dW (2 )

Now from the Second Law development, Chap. 8, since the process
is reversible
dQ = Td8 (3)

and, since motion and gravity effects are absent,

dW = pdv (4)

Then for a reversible process in the given system

T ds = du + pdv (5)

Equation (5) is a relation solely among the properties of a system.


If the system passes from one given state to another given state, the
property changes are the same for all processes by which the S3^tem
could pass from the one state to the other. Hence Eq. (5) is true
for aU between states which could hypothetically be con-
processes
nected by a reversible process. A pure substance can pass by a
hypothetical reversible process from any equilibrium state to any
other equilibrium state. Therefore Eq. (5) is a true relation among
the properties of a pure substance for all processes during which
the System is in equilibrium throughout its mass. This restriction
is obvious from the viewpoint of physical measurement of the prop-
erties for, if equilibrium did not exist, there would not exist any
particular properties which could be called the pressure and (he tern-
152 PURE SUBSTANCES

perature of the system. The equation would then be a relation

among non-existent entities.


Note that Eqs. (2) and (5) are applicable to both reversible and
irreversible processes;Eq. (2) because it is the First Law, and Eq. (5)
because a relation solely among properties, and is therefore inde-
it is

pendent of the type of process. Equation (5) is not the First Law;
it is a property relation that is dependent upon both the First and

Second Laws.
From Eq. (5) and the definitions of properties, numerous useful
property relations can be derived. As an example, a simple deriva-
tion of a very useful relation follows:

By definition A = w + pv
or dh = du + p dt; + r dp

Substituting in (5) T ds = dh - v dp (6)

Other relations will be derived in subsequent' chapters. For gen-


eral discussion of such relations and examples of their use see the
references below.

PKOBLEMS
9-1. From the temperature-entropy diagram for carbon dioxide (in the back
cover envelope) find the phase (liquid, solid, vapor, or mixture) in which carbon
dioxide exists in equilibrium at each of the following states; (a) t = ISO^’F,

p « 1200 psia; (b) t *= — 100®F, t; = 10 cu ft/lb; (c) p = 100 psia, h = 180


Btu/lb; (d) p « 1200 psia, t = 0°F; (e) t = -80°F, p = 75 psia; (f) v - 0.2
cu ft/lb, p « 400 psia.
9-2. In the temperature-entropy diagram for carbon dioxide find the data
needed to answer the following questions; (a) What are the critical pressure
and temperature of carbon dioxide? (b) How much does the volume of carbon
dioxide change in going from saturated liquid at O^F to saturated vapor at O'^F?
(c) Find the saturation temperature of carbon dioxide at 100 psia and at KXK)
psia. (d) At what temperature will solid carbon dioxide be in equilibrium with
its vapor under a pressure of 1 atm? (This is the temperature of *'dry ice.”)
(e) At the triple-point temperature, how much does the enthalpy of carbon
dioxide change in going from saturated solid to saturated liquid from saturated ;

liquid to saturated vapor; from saturated solid to saturated vapor?


9-3. In the temperature-entropy diagram for air (in the back cover) find the
data to answer the following; (a) What are the critical pressure and temperature
of air? (b) Find the enthalpy of superheated air at — IfiO^'F and 1(X) psia.
(c) Find the constant-pressure specific heat of air at — 200°F, 14.7 psia, and

at — 2()0T, 1000 psia. (Find how much h changes for a small temperature
change along the constant-pressure line, and compute 6h/Ai.) (d) Find Cp for
PURE SUBSTANCES 153

air at240®F, 14.7 psia, and at 240®F, 1000 psia. (e) At 100 peia find the satura-
tiontemperatures for vapor and for liquid. (The saturation temperature is not
constant at constant pressure because air is a mixture of oxygen and nitrogen;
oxygen-rioh liquid condenses and a progressively lower temperature is
first,

required to condense the vapor as becomes richer in nitrogen.)*


it

9-4. From the pressure-enthalpy diagram for Freon-12 (in the back cover
envelojjfe) find the data to answer the following: (a) What are the critical pres-
sure and temperature of Freon-12? what phase is Freon-12 at 200 psia,
(b) In
400°F? (c) At = 55 Btu/lb, t = 40®F, what is the state of aggregation (liquid,
vapor, mixture)? (d) What is the normal boiling point (at 1 atm pressure) for
rreon-12? (e) What is the density of Freon-12 saturated vapor at 0°F?
9-5. A Joule-Thomson experiment (Fig. 9-11) is carried out with Freon-12;
the initial state is at 100 psia, 100®I\ and the final pressure is 20 psia. Using
the chart find (a) the final temperature of the fluid; (b) the change of internal
energy during the process.
9-6. Using either the enthalpy-entropy diagram, or the temperature-entropy
diagram in Keenan and Keyes’ steam tables, find: (a) the saturation pressure
for water at 400°F (b) the enthalpy of saturated water vapor at 100 psia; (c) the
;

approximate specific heat at constant pressure for water vapor at 1 atm, 300®F.
(Find the ratio of Ah to At for a small temperature change at constant pressure.)
9-7. From the temperature-entropy diagram for carbon dioxide find the
change in the entropy and the temperature of carbon dioxide when it goes through
a throttling process beginning at 800 psia, 100®F, and ending at 100 psia. Sketch
the area under the path of the process in the T-s plane, and explain why the
area is not zero even though the process is adiabatic.
Chapter 10

TABULATED PROPERTIES-
STEAM TABLES

In this chapter the use of tabulated property data will be explained


with particular reference to the steam tables.* The properties of
water are arranged in the steam tables as functions of pressure and
temperature. Separate tables are provided to give the properties of
the substance in the liquid and vapor phases and in the saturation
states. The several tables will be considered separately. In all the
tables the zero points for both enthalpy and entropy are taken at
32°F saturated liquid.

10*1 Saturation states. Tables 1 and 2 give the properties of


saturated liquid and saturated vapor. Either the pressure or the
temperature is sufficient to identify a saturation state so Tables
1 and 2 need only one independent variable. In Table 1 the inde-
pendent variable is temperature; on each line the following data are

tabulated:

t, °F, temperature

p, psia, saturation pressure corresponding to t

Vf, cu ft/lb, specific volume of saturated liquid


Vff, cu ft/lb, change in v during evaporation
Vf, cu ft/lb, specific volume of saturated vapor
hf, Btu/lb, specific enthalpy of saturated liquid
hft, Btu/lb, change in h during evaporation (latent heat)
hg, Btu/lb, specific enthalpy of saturated vapor
Sf, Btu/lb®R, specific entropy of saturated liquid
8/g, Btu/lb*R, change in s during evaporation
8g, Btu/lb*R, specific entropy of saturated vapor

• Keenan, J. H. and F. G. Eeyee, Themodynamie


Propertiet of Steam. New
York: Wilejr, 1936. Extracts from these tables appear in the Appendix.
154
10- 1 ] STEAM TABLES 155

The following relations will obviously hold;

^fg “ ^^0 * ^0 **

Table 2 is similar to Table 1 but has pressure as the independent


variable. Table 2 also contains columns for internal energy:

u/, Btu/lb, specific internal energy of saturated liquid


Ufg, Btu/lb, change in u during evaporation
Ug, Btu/lb, specific internal energy of saturated vapor

Ufg — Ug — Uf

Tables 1 and 2 may be used interchangeably. The reason for


two tables is to reduce the amount of interpolation required. The
use of the tables is best illustrated by examples.

Example 1. What is the heat of vaporization of water at 100 peda?


The heat of vaporization is the change of enthalpy from saturated liquid
to saturated vapor,
AA — hg A/ = hfg

At 100 psia, from Table 2, A/, = 888.8 Btu/lb.


How much work is done by 1 lb of water as it evaporates at 100 psia?
For a stationary system with no friction

W = Jpdv
At constant pressure W = p A» = p{vg — v/)

From Table 2,

Vg = 4.432 cu ft/lb and V/ = 0.01774 cu ft/lb.

Then W= (100) (144) (4.432 - 0.01774)

= 63,700 ft Ib/lb = 81.8 Btu/lb

The work may also be obtained by use of the First Law:

W = Q — Au
Am = Ufg = 807.1 Btu/lb (Table 2)

W = 888.8 - 807.1 = 81.7 Btu/lb


Since the process of evaporation at constant pressure also takes place at
constant temperature, the heat of vaporization could have been obtained
by asBnming a reversible process for which

Q= Tdg = TAs * Tsfg


J*
156 STEAM TABLES [ 10-2

From Table 2, at 100 psia, t = 327.8®F {T = 788°R), and S/g = 1.1286

Btu/lb°R. Then
Q = (788)(L1286) = 889 Btu/lb

Etcamfle 2. Saturated steam has entropy of 1.6315 Btu/lb°R; what are


its pressure, temperature, and enthalpy?
In Table 1 are found entropy values of 1.6326 at 302°F and 1.6302 at
304®F. By linear interpolation t is found to be very close to 303®F, h =
1180.6 Btu/lb, and p is close to 70 psia. Checking in Table 2 the exact
entropy value of 1.6315 Btu/lb®R happens to appear in the table at 70 psia.
Then from Table 2,

p = 70 psia t = 302.92°F and h = 1180.6 Btu/lb.

Example 3, What is the internal energy of saturated water vapor at


250°r?
Table 1 gives no values of u; however, u = h — pv. At 250°F hg =
1164.0 Btu/lb, p «= 29.825 psia, and Vg = 13.821 cu ft/lb. Then

- (^9-825) (1^) (13.82^


u, = 1164.0 ^ g Btu/lb.

Another method is to interpolate in Table 2. At 29 psia t = 248.40 and


Ug = 1087.3. At 30 psia t = 250.33 and Ug = 1087.8. Using linear inter-
polation, when t is 250.00®F Ug will be 1087.7 Btu/lb.

10-2 Liquid-vapor mixtures. When two identical masses of


substance are placed together to form a single system there are some
properties (called intensive properties) which have the same value for
the combined system as for the individual masses. For example, the
pressure and temperature are of this class. Other properties (called
extensive properties) are twice as great for the single system of two
identical masses as for either mass alone; for example, the volume,
internal energy, enthalpy and entropy are of this class. If any two
masses of a given substance having equal pressures and temperatures
are placed in contact to form a single system, the pressure and the
temperature of the combined system will be the same respectively as
the pressure and temperature of the separate systems. The volume
of thecombined system will however be equal to the sum of the
volumes of the separate systems and a similar summation will apply
to the internal energy, enthalpy and entropy.
Consider a mixture of saturated liquid water and water vapor in
equilibrium at pressure p and temperature t. The pressure and tem-
perature of both components of the mixture must also be p and t
10*2 ] STEAM TABLES 157

Let the composition of the mixture be given as the fraction, by moss,


of vapor in the mixture; the symbol x and the name quality are used
for this fraction. Then if the mixture is 10 percent liquid by mass
the quality, x, is 0.90; if the mixture is equal parts by mass of liquid
and vapor the quality is 0.50. In a mixture of total mass m and
quality x the mass of liquid, rrif, is (1 — x)fn and the mass of vapor
mg is xm. Now as stated above, each extensive property of the mix-
ture will be equal to the sum of the values of the same property for
the two components. Thus, letting subscript x denote the property
of the mixture at quality x, for total volumes

F, = F/ -h Vg

or, since total volume is mass times specific volume,

nwz mjV/ -|- MgVg

Then m«, (1 — x)mvf -|- xmvg

or = (1 — x)v/ -f xvg

similarly w, = (1 — x)w/ -|- xug

h, = (1 — x)hf -t- xhg (a)

8, = (1 — x)S/ -|- XSg

The relations (a) can be changed in form as follows:

t>, = (1 - X)Vf -H XVg

= V/ + x{vg - Vf)

- Vf + XVjg (b)

or alternatively t'* = (1 — x)»/ -f- xVg

= (1 - x)v] - (1 - x)i;, + Vg

= Vg - - X)Vjg (c)

The physical interpretation of (b) is that the volume of one pound


of mixture at quality x is equal to the volume of one pound of liquid
plus the increase of volume due to the evaporation of x pounds. The
physical interpretation of (c) is that the volume of one pound of mix-
ture of quality x is equal to the volume of one pound of vapor minus
the decrease of volume due to condensation of (1 — x) pounds. Pre-
.

158 STEAM TABLES [10-2

cisely analogous equations and interpretations can be given for inter-

nal energy, enthalpy and entropy. The quantity (1 - x) is called


the moisture fraction.

Example 4. Find the properties of a mixture of steam and liquid water


at 100 psia, containing 40 percent liquid.
The quality, x, is 0.60.

Using (a) Vx = 0.40»/ + 0.6Q»»,


From Table 2, V/ = 0.01774, V, = 4.432

Vx (0.4)(0.0m4) + (0.6)(4.432)

2.666 cu ft/lb

Using (b) hx A/ + (0.6)A/j

hf 298.4, A/, = 888.8 (Table 2)

hx = 298.4 + 533.3 = 831.7 Btu/lb

Using (c) 8x — (0-4)s/e

8g = 1.6026, Sfg = 1.1286 (Table 2)

= 1.6026 - 0.4514 - 1.1512 Btu/lb°R

In thisjR^amplq it makes little difference whether the equation form (a), (b)

or (c) is used; however, when the quality is close to unity the precision of
computation is improved if the form (c) is used, because the larger part of
the result then comes directly from the tables and only a small subtractive
quantity is computed.

Example 5. Find the enthalpy of wet steam, 0.97 quality, at 100 psia.

Using (b) K = + xh/g


hf

h, = 298.4 + (.97)(888.8)

By slide rule K = 298.4 + 862 = 1160 Btu/lb


Precision achieved is not better than to the nearest whole Btu/lb.

Using (c) K^hg- {I - x)hfg

K= 1187.2 - (0.03)(888.8)

By slide rule A, = 1187.2 - 26.6 = 1160.6 Btu/lb

Precision here is good to the first decimal place, assuming that the tables
are precise.

It may seem unnecessary to obtain such precision as in 1100


but it must be remembered that most calculations involve differences
1 ^3 ]
STEAM TABLES 159

between tabulated values. If a difference of 10 Btu/lb were involved,


precision to the first decimal place would be highly desirable, even
though the actual tabulated values are of the order of 1000.
In many problems a mixture is described not by its quality but
by a pair of properties such as entropy and pressure. The procedure
in obtaining the other properties is to find the quality from the given
properties, and then find the other properties from the quality.

Example 6. Water at 15 psia has entropy of 1.7050 Btu/lb^R. Find


its enthalpy and volume. To find the composition of the mixture,

«* = «»-(!- x)st,
From Table 2 1.7050 = 1.7549 - (1 - z) (1.4415)
1 — z = 0.0346 moisture fraction

Then hz ” ” (1 ^ x)hfg
From Table 2 A, = 1150.8 - (0.0346) (969.7)
= 1117.2 Btu/lb

Similarly »* = — (1 — x)vfg

= 26.29 - (0.0346) (26.27)

= 25.38 cu ft/Ib.

10-3 Superheated vapor. When the temperature of a vajmr is

greater than the saturation temperature corresponding to the exist-


ing pressure, the vapor is said to be superheated. The difference be-
tween the existing temperature and the saturation temperature cor-
is called the superheat; thus steam
responding to the existing pressure
at 14.7 psia and 220°F has 8 degrees superheat. In a superheated
vapor at a given pressure the temperature may have any value greater
than the saturation temperature; therefore the superheated vapor
table (Table 3 in Keenan and Keyes) is arranged with two independ-
ent variables, pressure and temperature. Three properties, volume,
enthalpy and entropy are given for each tabulated pair of values of
pressure and temperature.
Example 7. Find the enthalpy of steam at 150 psia, 400°F. Inspec-
tion of the saturation data shows that the given state is superheated. Then
in Table 3 at the intersection of the 150 psia line with the 400°F column
find A = 1219.4 Btu/lb.

Example 8. Find the temperature of steam having enthalpy 1303.7


Btu/lb and pressure 300 psia. First determine whether the steam is a mix-
160 STEAM TABLES [KM

ture or is superheated. Since hg at 300 psia is found to be less than the


given h, the steam is superheated. In Table 3 at 300 psia the given value
of h appears in the column for 580°F. If the exact value of h had not
appeared in the table it would have been necessary to interpolate to obtain
the temperature.

In some cases it may be necessary to interpolate in two directions


in the tables.For example, the properties at 296 psia and 565®F
can be found only by interpolating between 29o and 300 psia and
between 560 and 580®F. It is desirable to construct an auxiliary
table such as shown below to facilitate the work. The values tabu-
lated in Table 3 are filled in first,

V 560"F SSO^’F

from from
295 Table 3 Table 3

296

from from
300 Table 3 Table 3

and the 565® column (or the 296 row) may be filled in two end spaces
by interpolation. A second interpolation will then fill the center
space. As an example, the entropy for the state 296 psia and 565®F
may be found; the value is 1.0099 Btu/lb®Il.
10-4 Gimpressed When the temperature of a liquid is
liquid.
less than saturation temperature for the existing pressure, the liquid
is called compressed liquid. The
pressure and temperature of com-
pressed liquid may
be varied independently; consequently a table of
properties could be arranged like the superheated vapor table to give
the properties at any pressure and temjjerature. However, the prop-
erties of liquids change little with pressure; hence for pressures of a
few hundred psi, little error is made if the properties are assumed
independent of pressure. In tliis case properties are taken from the
saturation tables at the temperature of the compressed liquid.

Example 9. Find the enthalpy of liquid water at 2(K) psia and 70®F.
From Table 1 at 70°F, h/ = 38.04 Btu/lb; therefore A at 200 psia, 70°r is
approximately 38.0 Btu/lb.
10-6 ] STEAM TABLES 161

When the pressures are high, or great precision is desired, the


properties of compressed liquid water may be obtained from Table 4
of Keenan and Keyes. Table 4 does not give property values di-

rectly; instead it gives the properties in terms of differences from the


properties of saturated liquid at the same temperature, Such a table
of correction factors is less convenient but much more compact than
a complete table.

Example 10. Find tlie volume and enthalpy of liquid water at 100®F
and 1000 psia.

From Table 4 at 100°F and 1000 psia

V — V/ = —0.1 X 10“‘

h - h, = -1-2.70

From Table 1 at KKFF v, = 0.01 013

h, = (>7.!»7

Then at 100®]' and 1000 jjsia

r = 0.01G08 cu ft/lb

It = 70.07 Htu/lb

In using Table 4, care inu.st be taken with signs and decimal points. An-
other method of obtaining the enthalp}' of compressed liquid water will be
found in the discussion of the Rankine cycle, Sec. 14-2, Example 1.

10-5 Charts of thermodynamic properties. There are certain


advantages in the presentation of the properties of substances in
graphical form. The manner of variation of the properties is made
more evident and the problems of interpolation are simplified. How-
ever, it is not usually convenient to use a chart large enough to give
the same precision as tables. Charts are commonly used to deter-
mine the general nature of a problem before proceeding with a
detailed computation and to provide data for computation when the
required precision can be obtained with the charts.
Keenan and Keyes present two charts for steam, a temperature-
entropy plot and an enthalpy-entropy plot. The temperature-
entropy plot shows the vapor dome and lines of constant pressure,
constant volume, constant enthalpy, constant quality and constant
superheat. The scale of this plot is so small that its use is limited
to the general orientation of problems. For example, given the vol-
162 STEAM TABLES [IM,

ume and enthalpy of a mass of steam, to find the exact state in the
tables is a tedious task since neither property is an independent vari-
able in the tables. With the chart, however, the desired state can
easily be located approximately, thus reducing the search in the
tables to a narrow range. Processes and cycles may be plotted on
the chart for convenient visualization of property relations.
The enthalpy-entropy chart, commonly called a MoUier chart,
contains the same data as the T-s chart with the exception of the
constant-volume lines. The scale of the h~8 chart is large enough to
provide data suitable for many computations.
Some authors use coordinates other than enthalpy-entropy for
charts to be used in computation. Enthalpy and volume have been
used for a steam chart and the most common coordinates for charts
relating to refrigerants are logarithm of pressure vs. enthalpy.*
10-6 Measurement of steam quality. In general all the prop-
erties of two independent properties are
a pure substance are fixed if

given. The properties which are most convenient to measure are


pressure and temperature; therefore whenever the pressure and tem-
perature are independent properties it is customary to measure them
to determine the state of the substance. As shown in Fig. 10- 1 ,
measured values of pressure and temperature which fall in the super-
heated vapor region or in the compressed liquid region will fix definite

points on the plot of properties. In these regions pressure and tem-


perature are independent. Measured values of pressure and temperar
ture which correspond to saturation conditions, however, could apply
equally well to saturated liquid point /, saturated vapor point g, or
to mixtures of any quality, points *i, or I3. In order to fix the
some other property such as specific volume, enthalpy or com-
state,
position of the mixture (quality)
must be measured. Specific volume
measurements are usually impractical in engineering work; devices
known as calorimeters are used for determining the enthalpy or com-
position of mixtures.
In the separating calorimeter a stream of fluid is separated me-
chanically into liquid and vapor by a sudden change in direction of
flow. The liquid and vapor are measured separately by a calibrated

* Elllenwood, F. O. and C. O. Mackey, Thermodynamic Charts. New York:


Wiley, 1944. USBureau of Standards, Thermodynamic Properties of Ammonia,
Circuhr 142.^ .^er. Soc. Refrigerating Engineers, Refrigeraling Data Book,
;

10^] STEAM TABLES 163

Fig. 10-1.

receiver and a calibrated orifice. This device is used when the mois-
ture fraction is more than a few percent.
In the electric calorimeter, Fig. 10-2, a stream of wet steam is

passed in steady flow through an insulated duct containing an electric


heater. Sufficient energy is supplied by the heater to give super-
heated steam at the outlet; then the pressure and temperature at the
outlet will give the exact state at the outlet. The heat supplied is

determined by electrical meters and the sample flow rate is measured


then the steady flow energy equation will give the enthalpy of the

Fig. 10-2. Electric calorimeter.


164 STEAM TABLES [KMt*

steam at the inlet. Since elevation, kinetic energy change, and shaft
work may be eliminated the equation is

hi hi ~Q
The throttling calorimeter,
Fig. 10-3, is a steady flow device
in which a sample of steam is put
through a throttling process.
In this process the fluid flows
in a well-insulated duct through
a localized restriction (such as a
partially closed valve or a small
orifice)and then resumes its origi-
nal velocity downstream from
the restriction. For this process
the steady flow energy equation
reduces to

Fig. 10-3. Throttling calorimeter. hi = hi


In use the inlet pipe and valve are cov-
ered with insulation. The pressure in In the use of the throttling cal-
the calorimeter body, where the tem- orimeter the sample of wet steam
perature is measured, is atmospheric
entering the device becomes su-
because of the largo discharge passage.
(In calorimeters that arc not open to
perheated in the throttling process
the atmosphere the pressure in the so that the final state can be de-
calorimeter must be measured.) The termined by measurements of
steam leaving through the annular pressure and temperature. Since
jacket space reduces heat loss from the
the enthalpy of the initial state
inner chamber. (Courtesy Ellison Draft
Gage Company.) is the same as the enthalpy of the
the initial state can
final state

then be determined. a sample of steam at 100


For example, if

psia containing 3 percent moisture is expanded to 15 psia in a


throttling calorimeter, the location of the initial and final states, 1
and 2, would appear on the h-s and t~s diagrams as shown in Fig.
10-4. This may be checked on the diagrams in the steam tables.
The process in the calorimeter is not a constant-enthalpy process;
nevertheless A* = Ai and the determination of state 2 fixes state 1.

Example 11 . A sample of steam at 200 psia flows to a throttling calorim-


eter in which the pressure is 15 psia and the temperature 280°F. Find the
quality of the sample. At 15 psia and 280°F the enthalpy is found in
* 10-6 ] STEAM TABLES 165

Fig. 10-4. Throttling calorimeter process.

Table 3 to be 1183.2 Btu/lb. Therefore hi = 1183.2 Btu/lb. To determine


the quality, x,
hi = hg (1 ” ^'jhfg

1183.2 = 1198.4 - (1 - x)843.0 (Table 2)

1 - X = 0.018

X = 0,982

A solution may be made on the h-s chart: find point 2 at 15 psia, 280^F, and
proceed on a constant h line to 200 psia; there read moisture 1.7 percent.
166 STEAM TABLES [m
The range of usefulness of the throttling calorimeter is limited
by the fact that for given values of pi and p2 there is a maximum
moisture content at 1 for which the steam at 2 will be superheated;
with greater moisture the method fails. This situation is illustrated

in Fig. 10-5. For the pressures pi and p2 the method will work with
a sample at state 1, but will fail with a sample at state 1' or any
state of greater moisture. If the pressure in the calorimeter can be

reduced to ps, the moisture range is increased as is evident from the


diagram. The throttling calorimeter also has a limitation on initial

s
Fig. 10-5.

pressure; if p2 is fixed at 1 atm, the moisture range increases for in-


creased Pi up to about 700 psia and then decreases to zero at about
1800 psia.
In the use of a throttling calorimeter the measured superheat in
the calorimeter should be sufficient to give reasonable assurance that
the apparent superheat is not the result of errors in measurement.
If the steam were not really superheated in the calorimeter the results
would be worthless. An apparent superheat of 10°F is a desirable
m ini mum. The same would apply to the electric calorimeter. All
calorimeters and their connections should be well insulated and have
sufficient steam flow through so that external heat transfer will be
negligible. They must also be run long enough to warm up thor-
STEAM TABLES 167

ou^y before use. In any case a calorimeter can give only the
quality of the sample it receives; the most difficult part of the prob-
lem of determining steam quality is to get a sample which truly repre-
sents the fluid whose quality is desired.
Evdry problem in which tables are used is to some degree a cut-
and-try problem. The search for data, at first, appears laborious but
with tables, as with any other tool, experience leads to greater facility.
In the use of tables the units should be checked, as no standards
are generally followed.
In many cases for which complete tabulations of data are not
available, it is possible to find specific heats, coefficients of expansion,
and the other data from which the needed properties can be computed.

PROBLEMS*
10-1. Complete the following table for water (liquid, vapor, or mixtures).
(Insert a dash for irrelevant items.)

p t V X Superheat h s
(psia) CF) (cu ft/lb) % (F) (Btu/lb) (Btu/lb R)

(a) 100 350.4

(b) 0.01700 218.48

(c) 400 0.9

(d) 14.696 1.4600

(e) 140 600

(f) 60 9,403

(g) 60 10.000

(h) 800 1416.4

(i) 4.002 1476.2

* In solving the problems of this chapter use the abridged tables in the

appendix or the complete tables if available; use charts for preliminary orienta-
tion and for check solutions.
168 STEAM TABLES
280 100
(j)

(k) 226 — 1300.0

A pound of water in a closed container is one-third liquid and two-thirds


10-2.
vapor, by volume; the temperature is 300°F. Find the pressure, the quality,
the volume, and the enthalpy of the mixture.
10-3. Three pounds of liquid water at 1 atm is heated at constant pressure
until it becomes saturated vapor. Find the changes in volume, temperature,
and internal energy during the process.
10-4. Steam initially at 400''F, 40 psia, is cooled at constant volume, (a) At

what temperature does the steam become saturated vapor? (b) What is the
quality when the temperature reaches 200°F? (c) How much heat is trans-

ferred, per pound of steam, in the process between 400®F and 200°F?
10-5. Steam from a pipe where the pressure is 110 psia flows steadily through
an electric calorimeter and comes out at 100 psia, 430°F; the electric power input
to the calorimeter is 1 kw, and the amount of steam that flows through the
calorimeter in 5 min is 4.1 lb.Find the moisture fraction or the superheat of
the steam taken from the pipe.
10-6. A sample of steam from a boiler drum at 390 psia
is put through a

throttling calorimeter inwhich the pressure and temperature are found to be


14.7 psia and 318T. Find the quality of the sample taken from the boiler.
10-7. A sample of steam taken from the supply pipe for a turbine, where the
pressure is 235 psia, goes to a throttling calorimeter where the pressure is 20 psia.

(a) If the calorimeter temperature is 325°F what is the state of the sample taken
from the pipe? (b) If the calorimeter must show at least lO^F superheat for
acceptable results, what is the maximum moisture content that could be meas-
ured under the given pressures?
Chapter H

PROPERTIES OF GASES

The properties of gases are frequently correlated by means of cer-


tain simple rules called the Per/ect Gas Laws. These rules were
originally formulations of the best experimental data, but more ac-
curate experiments showed them to be only approximations. It is

often of great convenience, however, to assume that a gas is a so-called


“perfect gas,” a hypothetical substance which obeys the perfect gas
laws. The two characteristics that define a perfect gas are the form
of the equation of state and the fact that the specific heats are
constant.
11-1 p, V, T relations for the perfect gas. The equation of
state of a perfect gas is

pv = RT (1)

where p is absolute pressure, v is specific volume, T is absolute tem-

perature, and R is an experimental constant called the gas constant.


In the usual case, R is in units of ft Ibf/lbm ‘’R. The gas constant
has a different value for each gas (see Table A-1 on p. 502). I

Equation (1) is satisfactory for real gases at high temperatures


(more than twice the critical temperatiu'e) or at low pressures (of the
order of 1 atmosphere for temperatures below the critical). For ex-
ample, the critical temperature of nitrogen is 227®R; at room tem-
perature, (SSCR), nitrogen agrees with Eq. (1) to within 1 percent
at pressures as high as 10 atmospheres. Water (critical temperature
1166®R) deviates from Eq. (1) by about 3 percent at 1 atm, 800°R,
and by about 1 percent at 1 psia, 800®R.
Equation (1) may be used with gas mixtures such as air so Icmg
as the composition of the mixture is not changed by condensation,
evaporation, or chemical reactions (see Chap. 12).
If both sides of (1) are multiplied by the mass m of a gaseous
(system, the result is

170
11 - 1 ] PROPERTIES OF GASES 171

pV = mRT (2)

where V, which equals mo, is the volume of the mass m.


Another form of the equation of state, much used by physicists
and chemists, is based on the mol as a unit of quantity of gas. A mol
of substance has a mass numerically equal to the molecular weight
of the substance. Thus a pound mol of oxygen is 32 lb, and a gram
mol of oxygen is 32 gm. Since the molecular weight is a number
proportional to the mass of a molecule it follows that 1 mol contains
the same number of molecules for any substance. Avogadro’s law
states that a given volume will contain the same number of molecules
of any gas at given pressure and temperature. (This law is based on
the volumes entering into chemical reactions.) Then for all gases,
given the same p, V and T, the number of mols is the same. This
is valid for real gases under the same conditions as is the equation of

state, (1). Now for n mols of gas the mass is


m = nM
where M is the molecular weight. Substituting in (2)

pV = nMRT
If p, V and T are fixed at po, Vt, and To, n will have a fixed value no
the same for all gases. Then
PoVo
MR noTo

or MR is the same for all gases at po, Vo and To. But for any one
gas both M and R are constant for all states, therefore MR is a con-
stant for all gases in all states. The name universal gas constant
and the symbol R are used for the product MR. Then

pV = nRT (3)

The value of the universal gas constant is 1545 ft Ib/lb mol ®R; in
heat units this is 11)45/778 = 1.986 Btu/lb mol “R.
The gas constant for any perfect gas may then be obtained from

R= ~= ^ ft Ibf/lbm “R (4)

or R« Btu/lb “R
172 PROPERTIES OF GASES [ 11-1

Another characteristic number for all gases is the volume of a mol


of gas at standard pressure and temperature, 1 atm and 32“F. This
is called the molecular volume. The pound molecular volume is

Vo RTo _ (1545)(492)
“ n 359 cu ft
po (14.7)(144)

The gram molecular volume, used by scientists, is 22.4 liters.


The volume of n mols of perfect gas at any temperature and
pressure, T and p, is given by

V = Vmu Lii.3sgJsJL.nr
i (5)
ioP M492
where p is in psia and T is in °R.
The actual values of Rj R/M, and Vm for certain gases are listed
in Table A-l, p. 502. It will
be observed that gases of
simpler molecular structure
show better agreement with
Eq. (4) at 1 atm, 32®F, because

of their low critical tempera-


tures.
The graphical representa-
tion of Eqs. (1), (2) and (3)
upon the pressure-volume
plane is a series of rectangular
hyperbolas for lines of constant
temperature, Fig. 11-1.

Fig. 11-1. P-v-T relation for a perfect Example 1. A certain piece


gas.
of apparatus of constant volume
is filled with nitrogen at 15 psia,
80®F. From a nitrogen bottle on a weighing scale exactly 3 lb of nitrogen
isadded to the ai)paratus. The final pressure and temperature are 25 psia,
75°F. Find the volume of the apparatus.

In the initial state PiFi = tniRTi


In the final state p^W =
Since volume is constant Vi = l'^

It is given that = ?wi + 3

V\V\ _ vi\RTi
Then
m^RTi
11 - 2 ] PROPERTIES OF GASES 173

Hh = = i§ — 0 5Q4
m% P2 T1 25 540

m =
2 0.594m2 + 3

m =
2 7.39 lb

agr. . . 60.9 ft
Pi (25) (144)

Note how the problem of units is simplified by dealing as far as possible


with ratios in which units cancel.

11-2 Specific heats, internal energy and enthalpy. A perfect


gas not only satisfies the equation of state (1), but its specific heats
are constant. Real gases deviate appreciably from this rule, as shown
in Tables A-4-A-6. The variation of specific heats is primarily with
temperature change, but there is a small variation with pressure.
For many purposes the use of constant specific heats is satisfactory,

particularly if suitable average values are chosen. When this is not


satisfactory, the Gas Tables may be used (see Sec. 11-6).

A corollary of the perfect gas equation of state is that the internal


energy of a gas is solely a function of temperature. This is approxi-
mately true for real gases (Joule’s law, based on experiment) and it

can be proved analytically for any pure substance having the equation
of state (1). The proof follows. The properties of any pure sub-
stance are related by Eq. (9-5)

T ds = du + p dv

or ds = p + ^dv
Taking T and v as independent variables this may be expanded into

Also, since s — f{T,v)j by the rules of differentiation

Since T and v are independent, the coefficients of dT and dv in (a)


and (b) must be respectively equal; then
174 PROPERTIES OF GASES [114

II
and (d)

Differentiating (c) and (d)

d^s 1 ± /du\ '^


/d2\ _±^ (e)
dvdT ~ T dvdT T* \3t; ) t T \dT)v T*

1 / a*w \
' (f)
afU' T\dTdv)
But the order of differentiation makes no difference, hence the left

side of (e) is equal to the left side of (f) and their right sides may
be equated;

1 a*« 1 3*m 1 r/dn\


1
1 /^\
(g)
TdTdv T dvdT r* l\dv)T T \dT),
Multiplying through by f* and eliminating the second derivatives,
which are equal,

(h)

The results so far apply to any pure substance. Now for any sub-
stance for which pv = RT,

II II

or (i)

Then from (h) and (i), for a perfect gas

®
Since T and v are the only independent properties and u does not
change when v changes at constant temperature, then u does not
change at all unless the temperature changes. Thus u is solely a
function of temperature for the perfect gas.
In general, if w is a function of T and v,

•But by (6) the last term is zero and by definition


11.2] PROPERTIES OF GASES 175

“ {df\
Therefore for the perfect gas

du = CtdT (7)

not only for a constant-volume process but for any process. Since
Cv is constant for the perfect gas, it follows that

Au = Cv AT (8)

The enthalpy of any substance is given hy h = u+ pv. Then


for the perfect gas
h — u JtT

or A is a function of temperature only. By definition, for any sub-


stance,

= (il
Then for the perfect gas, since A is a function of T only,

dh = CpdT (9)

for any process. Since Cp is constant for the perfect gas

Ah = Cp AT (10)

If lines of constant internal energy or of constant enthalpy are


plotted on the p-v plane, they will be identical with the lines of con-
stant temperature in Fig. 11-1.
Observe that Eqs. and (9) depend solely upon the assumption
(7)
that pv = RT;
(8) and depend upon the further assumption that
(10)
the specific heats are constant. Starting with (7) and (9) (i.e. as-
suming only that pv = RT) an important relation between the spe-
cific heats can be derived. From the definition of enthalpy it follows
that
dh = du + d(pv)

Then for a gas dh + RdT


= du

or CpdT = CvdT + R dT

Then Cp — c, — R (11)

This relation will hold even if specific heats are permitted to vary
with temperature. Hence for a substance satisfjdng the perfect gas
176 PROPERTIES OF CASES [ 11-2

equation of state a value of one of the specific heats leads inunedi-


ately to the value of the other at the same temperature. It is w^ll

to recall here that the usual units for specific heats are Btu/lb degree
whereas the usual units for R are ft Ib/lb degree. It should be obvi-
ous that consistent units are necessary in Eq. (11); it is customary
to use the heat units.
The ratio of specific heats c,/c, is of importance in perfect gas
computations; it will be designated by the symbol k. According to
the classical kinetic theory of gases the ratio k should have the values
5/3, 7/5, and 4/3 respectively for monatomic, diatomic and poly-
atomic gases. This theory is not accurate, as is indicated by the
variation in measured values of k shown in Tables A-4-A-6. Never-
theless, at ordinary temperatures, for monatomic and diatomic gases
the actual values of k are not far from the values 1.67 and 1.40
respectively. For polyatomic gases there is considerable deviation
from the value 1.33. It may be observed that, if k can be predicted
in any way or measured experimentally (several common phenomena
are functions of k), the specific heats of a perfect gas can be deter-
mined without making heat measurements. From (11)

—= _ 1
Cy Cy

SO c, = (12)

Also from (4) R —


1.986
“ M{k - 1)

Only the molecular weight and k are needed to determine the specific
heats.
Modem theories of the structure of gases can predict specific heats
accurately, but such computations are outside the realm of thermo-
dynamics. Moreover the results are far too complicated for xise in
gas law computations; they are used in the preparation of tables or
charts of properties.

Example 2. Given that oxygen has a gas constant of 48.3 ft Ib/lb ®Il,
and a diatomic gas, compare its actual specific heats at lOO*’?, SOOT, and
is

1500“F with the values computed from the simple kinetic theory value of k.
c

ll-8> PROPERTIES OF GASES 177

f^ - c. = 48.3/778 = 0.0622 Btu/lb “R

Cp = ~ “ oli

- 0.218 Btu/lb °R
Cp * 0.218/1.4 = 0.156 Btu/lb ®R
From the plot, Table A-4,

= 100 500 1500

» 0.219 0.234 0.263

c. = 0.157 0.172 0.201

k = 1.40 1.36 1.31

Example 3. Show that on a molal basis the specific heats of all gases
having the same value of k are identical.

k(Cp — Cp) kR
Prom (12),
A: — 1 A; - 1

On a molal basis Cp M times as large as on a pound basis because each mol


is

of gas contains M pounds.


molal ^ Cp
^

But MR » R which is the same for all gases; therefore the molal Cp is a
function of k only.

11-3 Entropy of the perfect gas. The entropy change between


any two states of a substance may be determined by assuming a

V »

1 p*c

Fig. 11-2.
178 PROPERTIES OF GASES [11-3

convenient reversible process connecting the states, and computing


f dQ/T for the reversible process. For a perfect gas a convenient
process is a combination of a constant-pressure process and a constant-
volume process. In Fig. 11-2 any two states 1 and 2 are connected
by the constant-pressure and constant-volume paths intersecting at 3.
Assuming the processes 1-3 and 3-2 reversible, for a unit mass of gas

Since Cp and c, are constant

= Cp In
^+ c. In (b)

It is desirable to eliminate the properties at state 3; since the same


gas is dealt with at each state,

Pi»>i

Tx
_ p _
"
W_W
Tt T»
(c)

Then since p» - pi and »8 = i;2 ,

Tl
Ti t»i Tt Pi

Finally As^ = Cp In — -f- c. In ^ (13)


I'l pi

By a simple exercise in algebra, using the relations (11) and (c)


above, this can be converted to the forms

As,* = Cp In ^
1
B In ^ (14)
1 Pi

Asi* = c»ln'5^-f-i21n-^ (15)


1 1 Vi

The same results may be arrived at in a more elegant way start-


ing with the general property relation

T ds = du + pdv (9-5)

or ds = — + ^dv
Then for the perfect gas
dT ,
11-4] PROPERTIES OF GASES 179

or A«B = c.ln^ + J?In-


i 1

11-4 The reversible adiabatic process. The reversible adia-


batic process may be described in terms of properties by = 0.

From the general property relation (9-5) for a pure substance it fol-

lows that in a reversible adiabatic process

Tds=‘0 = du-\-pdv
or du = — p dv (a)

for any pure substance. For the perfect gas, then,

c, dT

dT ,
,dv

Integrating, In r = (1 — k) In v + constant

In = constant

Tv^-^ = constant 06)


Substituting from pv = RT,
pv’‘ = constant (17)

and TpO-*)/* = constant (18)

The internal energy change and enthalpy change for a reversible


adiabatic process in a perfect gas can be derived from (a) above.

(b)

Substituting from (17)

, PiVi
- pit>i
" fc - 1

Am„ = (T, - Ti) (c)


180 PROPERTIES OF GASES [IM

From Eq. (12) it is seen that this result agrees with the general rule
for the perfect gas, Au = c„ AT. Rearranging (c)

^
RTr /Ti \
(d)

and substituting from (18)

(19)

From the definition of h,

dA = du + d(p«) = du + pdx + vdy


Substituting from (a)
dh = V dp (e)

for a reversible adiabatic process in any pure substance. Then

From (f), by a development similar to that for the internal energy,


the result is obtained

Equations (19) and (20) are consistent witli (8) and (10) from wliich,
for any perfect gas process,

—=^=
' Z'
( 21 )
All Cv

Example 4. Nitrogen, in a frictionlcss adiabatic process, expands from


an initial state of 100 psia, 140®F to a final pressure of 10 psia. How much
does the enthalpy change?

One method is to use Eq. (20).

*
(1.4)(55.16)(600)
.
. ^ - 1
1.4-1 L\100/ ]
= 116,000(0.518 - 1)

= -56,000 ft Ib/lb = -72 Btu/lb

Another method is to use the relation

AA = c,AT
11-5] PROPERTIES OF GASES 181

which applies to any perfect gas process. From (18), for a reversible adia-
batic process,

T2 = 600(0.518) - 31 rR
A/i = 0.248(311 - 600) = -71.8 Btu/lb

11-5 The polytropic process. Many processes which occur in


practice can be described approximately by an equation of the form
= constant, where n is a constant. Such a process is called a
polijtropic process. The constant-temperature process and the re-
versible adiabatic process in a perfect gas are obviously special cases
of the polytropic. By substitution from pv = RT the following rela-
tions will hold for the perfect gas poly tropic process:

fpn-i = constant
'Pp(i-n)in _ constant

The paths of various perfect gas polytropic processes on the


pressure-volume plane and on the temperature-entropy plane are
shown in Fig. 11-3.
It was shown in See. 2-4 that for a process in which pv" = constant

The similarity to Eq. (19) is apparent. Note, however, that


Aw = — fp dv only for a reversible adiabatic ; (22) does not give
— A?^ unless n = k.

Also for the polytropic

This equation is similar to (20) but J'v dp is not AA unless n ^ k.

It is possible to derive what is called a '^polytropic specific heat*'

G
11 -6] PROPERTIES OF GASES 183

as follows: for a reversible polytropic process with a unit mass of


perfect gas the First Law gives

Q Aw + y*p dv
Now
A(pr) _ BAT _ Cf(k — 1) AT
tiu = CfLT and y*p dx
1 — n~l — n~ 1 —n

Then Q

The quantity in brackets in (24) is the heat transferred per unit


temperature change in a reoersible polyfpropic process and is therefore
called the polytropic specific heat.
11-6 Gas tables. In many problems encountered by engineers
the equation of state pv = RT is satisfactory but the assumption of
constant specific heats does not give the desired accuracy. The vari-
ation of specific heats has been accounted for in the past by assuming
that the specific heats are linear functions of the temperature:

c, = o -|- 6T
Cj, = (c -|- B) -f" hT
where a and b are constants. Substitution of these values in the
various equations involving specific heats leads to clumsy equations
which are still only approximate. The modem trend is to use tables
for w, A and « whenever the variation of specific heats is of importance.
The Gas Tables of Keenan and Kaye* provide data for variable
specific heat computations. These tables will be discussed briefly.
Since the perfect gas equation of state is assumed valid, the enthalpy
and internal energy are functions of temperature only and a table for
these data need have only one independent variable, temperature.
Entropy change, however, is a function of two independent proper-
ties, for example pressure and temperature; this is accounted for as

follows. For any pure substance

Tds = dh — V dp (9-6)

* Keenan, J. H. and J. Kaye, Gat Tables; TkemodyTumie Properties of Air


New York: Wiley, 1948. Also see Appendix Table A-8, p. 512.
184 PROPERTIES OF GASES [ 11-6

Then for any substance for which pt> = RT

ds = Cpy - (a)

Take the arbitrary zero of entropy at absolute zero temperature and


any given pressure po. Then the entropy Si at any given pressure
Pi and temperature Ti is

«!
0, Jio
ds
T
li
Jv,
r ^V
f-/?ln^
i po
(b)

The change of entropy Asu between any two states 1 and 2 is then
.r.
dT
ai f
A«ii ""
0 T Jo 1 Pi

In the tables the quantity

(c)
T
is tabulated as a function of temperature. Also the quantity /{ In r
is given in a second table, for a range of values of r. Then the
entropy change is given by

Asij = 4^ — ^1 — /2 In r (d)

where ^ is defined by (c) above, r = pi/pi, and all three terms on


the right side can be taken directly from the tables.
The path of a reversible adiabatic process in a perfect gas is
given by po* = constant. With variable specific heats no such sim-
ple relation exists. However, from (a), for the gas with variable
specific heats in a reversible adiabatic process

(e)

Then ^“ 1 dT
p
or for a reversible adiabatic process between pg and p
11-6] PROPERTIES OF GASES 185

Then for any given baae pressure po and base temperature To the
pressure pis a unique function of the temperature IT. Also from

pv
_ Wo
T To

the volume at any temperature must be a unique function of the


temperature.
In Fig. 11-4, for any of the constant-entropy paths between
Ti and Tt the ratio of the pressures is the same even though the

Fig. 11-4.

pressures are different, since the right side of (f) is the same for
each path; then

=
«
Pi
= f
p.i Pc
/(T’l, To) (g)

For the particular isentropic 012, passing through the base point
Po, To

2? = Tlila = Pro . .

Pi Pi/Po Pn

where p, = p/po is a function defined by (f) above, and called the


relative pressure.

The values of p, are tabulated, so for any isentropic path between


given temperatures Ti and To the ratio of the pressures can be found.
Similarly, values of relative volume v, are tabulated so that the ratio
of volumes may be found from the tables if desired.
186 PROPERTIES OF GASES [ 11-6

Exampue 5. Air initially at 100 psia, 2000°?, expands reverdbly and


adiabatically to 15 pcda. Fmd the change of enthalpy and of specific volume
by perfect gas laws, and by variable specific heats using the gas tables.
(1) Specific heats constant at room temperatufe values.

(53.34) (2460) /lOOV"*


(100)(144) \15)
= 9.12^ = 35.4 cu ft/lb

PiVi _ Jhfit

Ti r,

= r,22!!*
piVi
= 2460^^
100 9.12

= 1425'’R = 965“?

Ah = Cp AT = (0.240)(965 - 2000) = -249 Btu/lb

Av = 35.4 - 9.12 - 26.3 cu ft/lb


• (2) Specific heats constant at average values suitable to the process.
From the plot, Table A-4, take values at 1500°F as reasonable average values.
Then Cp = 0.276, c, = 0.207 Btu/lb *R, k = 1.33. Solving as before,

Ti
V2 =

=
(9.12)

1520‘'R
^ =
=

1060*F
37.8 cu ft/lb

Ah = 0.276(1060 - 2000) = -259 Btu/lb


Av = 37.8 - 9.12 = 28.7 cu ft/lb

(3) Variable specific heats (gas tables). At Ti, 2460‘’R, from Table 1,

hi = 634.34

Prt = 407.3

For a revertibie adiabatic process

p,* = Pri^* = 407.3


^= 61.1

In Table 1, p, = 61.1 at 1535*R. Then


T» = 1535®R =
hi » 378.44
11-7] PROPERTIES OF GASES 187

M- At - Ai - 378.44 - 634.34 = -256.90 Btu/lb

. ^ _ ml.. = 37.9 cu ft/lb


Pi (15) (144)

Av = 37.9 - 9.12 = 28.8 cu ft/lb

11-7 Modified p, v, T relations compressibility charts. —


When it is necessary to deal with substances not adequately described
by the perfect gas law and for which tables or charts are not available,
recourse must be had to more laborious methods. There are two
general methods of handling p, v, T data in such cases: (1) use of an
equation of state more complex than the perfect gas equation; (2) use
of plotted correction factors to apply to the perfect gas equation.
Many equations of state have been developed, some in attempts
to relate the properties to a theory of the structure of the substance,
and some solely to give a good empirical representation of experi-
mental data. Such equations are generally of more interest to physi-
cists and constructors of tables than to engineers with problems to
solve.
One of the best known of such equations, and one which is some-
times used in the absence of detailed data, is Van de Waals equation,

+ I) (^ - *>)
=
(p

In this equation a and b are constants which differ for different sub-
stances. Values of the Van de Waals constants, a and b, for a num-
ber of substances are tabulated on p. 503 in the Appendix. The
Van de Waals equation represents the behavior of real gases better
than the simple gas law equation does at high pressures, particularly
above the critical pressure.

Correction factors for the perfect gas equation are much used by
chemical engineers. A brief outline of one approach to this method
follows. An equation may be written for any gaseous substance,

pv = ZRT (a)

where the Z is an empirical factor having such values as necessary to


make the equation For any one substance Z will be a func-
balance.
tion of p and T so
a plot can be made of lines of constant temp>era-
ture on coordinates of p and Z. From this plot Z can be obtained
for any value of p and T and the volume can then be obtained from
188 PROPERTIES OF GASES

(a). The advantage of using Z instead of a direct plot of t; is a smaller


**
range of values in plotting.
would be possible to make a separate Z chart for each substance.
It
However, the general shapes of the vapor dome and of the constant-
temperature lines on the p-v plane are similar for all substances,
although the scales may be different. This similarity may be ex-
ploited by using dimensionless properties called reduced properties.

The reduced pressure is the ratio of the existing pressure to the


critical pressure of the substance, and similarly for reduced tempera-
ture and reduced volume. Then

Pit = ^ Tr =
^ “ (b)
Pc i C VjC

where subscript It denotes the reduced property, and subscript C


denotes the property at the critical state. When Tr is plotted as a
function of reduced pressure and Z, a single plot is found to be satis-

factory for a great variety of substances encountered in engineering.


Such generalized plots prepared by Weber are shown in Appendix
Table A-7, p. 510. Although necessarily approximate, the plots are
extremely useful for situations where detailed data on a particular
gas are lacking but its critical properties are available. The com-
pressibility charts may be used to estimate the accuracy of the perfect
gas law for any particular problem.
Methods of determining approximate values of other properties
have been developed by chemical engineers. *

PROBLEMS
11-1. Nitrogen is to be stored at 2000 psia, 80°F, in a steel flask of 1.5 cu ft
volume. The flask is to bo protected against excessive pressures by a fusible
plug which will melt and allow the gas to escape if the temperature rises too
high, (a) How many pounds of nitrogen will the flask hold at the designated
conditions? (b) At what temperature must the
fusible plug melt in order to
a maximum of 2400 psia?
limit the pressure of a full flask to
11-2. Tanks of 200 cu ft volume are to be used for storing gases at 250 psia,
100*F. The gases are oxygen, nitrogen, helium, hydrogen, and carbon dioxide.
Assuming these are all perfect gases, find: (a) the number of pounds of each gas
that can be stored in a tank; (b) the number of pound-mols of each gas that
can be stored in a tank.
11-3. For each of the following cases find the volume by the perfect gas

* See, for example, Weber, Harold C., Therriwdynamica for Chemical Engineers.
New York: Wiley, 1039, chaps. 5, 6, and 17.
PROPERTIES OF GASES 189

rules, and compare with the true volume shown in the tables or charts for the

substance: (a) water vapor at 300®F; (b) water vapor at 60 psia, 300®F;
1 psia,

(c) water vapor at 60 psia, 1000®F; (d) carbon dioxide at 10 psia, 0°F; (e) carbon

dioxide at 76 psia, 160®F; (f) carbon dioxide at 1100 psia, 160®F.


11-4. A tank contains dry air at 15 psia, 70°F. A flask of dry air contain-

ing 2 lb of air at 150 psia and 70'^F is connected to the tank by a tube of negligible
volume, and the pressures are allowed to equalize. Eventually the pressure is
30 psia and the temperature is 70°F in both containers. What is the volume of
the tank?
11-5. A certain gas has specific heats as follows: Cp = 0.47 Btu/lb F, c«
= 0.36 Btu/lb F. Find the molecular weight and the gas constant for the gas.
11-6. The molecular weight of acetylene, C2H 2 is 26; an experimental de-
,

termination of the specific heat ratio gives k = 1.26. Find the two specific heats.
11-7. Nitrous oxide, N2O, is a triatomic gas of molecular weight 44; esti-
mate its specific heat at constant pressure from this information. The actual
value at ordinary temperatures is about 0.21 Btu/lb P\
11-8. It was shown in Example 3, Sec. 11-2, that the molal specific heats of
a perfect gas depend only upon the specific heat ratio. Find, to two significant
figures, the molal specific heat at constant pressure and at constant volume for
perfect gases os follows: (a) monatomic, k = 6/3; (b) diatomic, k = 7/5; (c) poly-
atomic, k = 4/3.
11-9. A perfect gas has a molecular weight of 20 and Cv = 0.30 Btu/lb F.
The gas enters a pipe line at 100 psia and 1000®F#bs, and flows steadily to the
end of the pipe where the pressure is 75 psia and the temperature is 500°Fai».
No shaft work is done in the pipe, and velocities are small. How much heat is
transferred per pound of gas, and in which direction?
11-10. A constant-volume chamber of 10 cu ft capacity contains 2 lb of air
at 40°F. Heat is transferred to the air until its temperature is 300°F. Find
the work done by the air, the heat transferred, the change in internal energy,
the change in enthalpy, and the change in entropy.
11-11. A constant-pressure process begins with 2 lb of air at 40^F, occupying
a volume of 10 cu ft. Heat is transferred to the air until its temperature is
300®F; there is no work other than pdv work. Find the work, the heat trans-
ferred, the change in internal energy, the change in enthalpy, and the change
in entropy.
11-12* A quantity of nitrogen is cooled in a reversible constant-pressure
process in which 300 Btu of heat are transferred. Find: (a) the work; (b) the
change of internal energy.
11-13. A mass of hydrogen originally having a volume of 1 cu ft and a pressure
of 180 psia expands in a reversible process at constant temperature of 400^F;
the final pressure 15 psia. Find: (a) tfio work of the process; (b) the heat
is

transferred; (c) the entropy change of the gas.


11-14. Solve Problem 11-13 if the gas is nitrogen having: (a) the same
original gas volume as in Problem 11-13; (b) tJie same gas mass as in Problem
11-13.
11-15. Air flows steadily through a porous plug in a Joule-Thomson experi-
ment, Fig. 9-11. The initi|il pressure and temperature are 30 psia and SO^’F.
190 PROPERTIES OF GASES

and the final pressure is 15 psia. Find the change of temperature, of internal
energy, and of entropy, per pound of air.
11-16. Oxygen expands in a reversible adiabatic process from 45 psia,
to 30 psia. If the process occurs in a stationary system find the final tempera-
ture, the change of specific enthalpy, the heat transferred per pound of gas, and
the work done per pound of gas.
11-17. If the process of Problem 11-16 occurs in steady flow find the final
temperature, the change of specific internal energy, the heat transferred per
pound of gas, and the shaft work done per pound of gas; changes in elevation
and in kinetic energy are negligible.
11-18. Helium is compressed reversibly in a non-conducting cylinder from
70®F, 10 cu ft, 15 psia, to 1 cu ft. Find the change of enthalpy, the work done,
and the final temperature.
11-19. In the compression of air in a certain cylinder cooled by a water jacket
the indicator diagram shows that pressure and volume are related by s
constant. The process begins at 80°F, 14 psia, and the pressure rises to 80 psia.
If the process is reversible how much heat is transferred per pound of air?
11-20. A perfect gas cycle of three processes uses argon (cp » 0.125 Btu/lb F,
k = 1.67) as a working substance. Process 1-2 is a reversible adiabatic expan-
sion from 0.5 cu ft, 100 psia, 540®F, to 2 cu ft. Process 2-3 is a reversible iso-
thermal process. Process 3-1 is a constant-pressure process in which the heat
transfer is sero. Sketch the cycle in the p-v and T-$ planes, and find: (a) the
work for process 1-2; (b) the work for process 2-3; (c) the net work of the cycle.
11-21. Is the cycle of Problem 11-20 reversible? Give proof.
11-22. A Carnot cycle uses 1 lb of nitrogen as a working fluid. The maximum
and minimum temperatures of the cycle are lOOO^’R and 500‘'R, the maximum
pressure of the cycle is 100 psia^ and the volume of the gas doubles during the
isothermal expansion. Taking nitrogen as a perfect gas (specific heats constant)
show by detailed computation of the heat supplied and the work done that the
efficiency of the Carnot cycle agrees with that obtained in terms of temperature
in Chapter 8.
11-23. A mass of air is initially at 500^F and 100 psia, and occupies 1 cu ft.
The air is expanded at constant pressure to 3 cu ft; a polytropic process with
n a 1.50 is then carried out, followed by a constant-temperature process which
completes a cycle. All of the processes are reversible, (a) Sketch the cycle in
the p-^ and T-s. planes, (b) Find the heat received and the heat rejected in
the cycle, (c) Find the efficiency of the cycle, (d) Describe qualitatively any
differences that would have been found in (a), (b), and (c), if the gas had been
helium instead of air.

11-24. Air is to be heated in steady flow from 60®F to 1300®F. Find the
''heat required p^ pound of air by each of three methods: (a) use room-temperature
specific heats; (b) use a constant specific heat chosen from Table A-4, p. 507, as
a good average value for the given temperature range; (c) use enthalpy values
from the air table.
11-25. Solve Problem 11-10 by air tables.
11-26. Solve Problem 11-11 by air tables.
PROPERTIES OF GASES 191

11-27. For nitrogen, oxygen, carbon dioxide, water vapor, and hydrogen find
the enthalpy change between fiOO^Fabt and by: (a) the relation Ah
^ Cp aT, using room-temperature Cp; (b) the same relation, using Cp as an aver-

age value estimated from the plots of ^ vs. T in the Appendix, Tables A-4 and A-6;
(c) the plots of h/T vs. T in the Appendix, Tables A-5 and A-6.
11-28. One-tenth pound of air is compressed reversibly and adiabatically
from 12 psia, 140T, to 60 psia, and is then expanded at constant pressure to
the original volume, (a) Sketch these processes on the jm) and T~s planes,
(b) Compute the beat transfer and the work for the total path by simple gas-law
methods, (c) Compute the heat transfer and the work for the total path by

the air tables.


11-29. Solve Problem 11-19 by air tables.

11-30. One pound of air, initially at 80T, is heated reversibly at constant


pressure until the volume is doubled, and is then heated reversibly at constant
volume until the pressure is doubled, (a) For the total path find the work, the
heat transfer, and the change of entropy by simple gas-law methods, (b) For
the total path find the work, the heat transfer, and the change of entropy by
the air tables.
11-31. A Carnot cycle with 0.2 lb of air as working fluid operates between
2000''R and 1000^’R. The volume of the air at the beginning of isothermal
expansion is 1 cu and at the end of isothermal expansion is 2 cu ft. (a) Find
ft,

the work per cycle and the maximum volume of the gas by simple gas-law meth-
ods. (b) Find the work per cycle and the maximum volume of the gas by the
air tables.

11-32. Solve Problem ll-23(a), (b), and (c) by air tables.

11-33. Determine the specific volume of air at -100®F, 1000 psia, by: (a) the
perfect gas law; (b) the compressibility chart, Table A-7, p. 510.
11-34. Find the specific volumes for Problem ll-3(c), (e), and (f), by the
compressibility chart, and compare with the values found in Problem 1 1-3.

REFERENCES
For properties of gases, see the list of references at the end of Chap. 10, p. 168.
Chapter i2

PROPERTIES OF GASEOUS MIXTURES

Since a pure substance a substance of constant and uniform chem<


is

ical composition, a homogeneous mixture of gases would appear from


the thermodynamic or macroscopic viewpoint to be a pure substance,
and its properties could be formulated in the same way as for a single
gas.* One important example of this type of pure substance is dry
air. Mixtures' cease to be pure substances when, by reason of phase
changes or chemical reactions, the chemical composition of the mix-
ture ceases to be constant and uniform. For example, moist air is
not a pure substance in any process in which condensation or vapor-
ization of moisture occurs.
For certain mixtures of great importance (dry air, moist air, cer-
tain combustion products) thermodynamic data are available,, but for
many mixtures with which the engineer has to deal it is necessary to
compute the properties of the mixture from the composition of the
mixture and the properties of the components. In this chapter rela-
tions will be given for mixtures of components which can be treated
as perfect gases. Gas mixture properties are used in combustion
calculations (Chap. 22) and in humidity calculations (Chap. 25).
12-1 Basic rules for mixture properties. From experiments
on the mixing of gases, Dalton’s Law of Partial Pressures was devel-
oped. Datum’s Law: The 'pressure of a mixture of gases is equal to the
sum of the pressures of the individual components taken each at the
temperature and volume of the mixture.
When gases at equal pressures and temperatures are mixed adia-
batically without work, as by a constant-volume
inter-diffusion in
container, the First Law
requires that the internal energy of the
gaseous system remain constant, and experiments show that the tem-
perature remains constant. From this fact this rule may be deduced
• Even chemically pure gases arc often mixtures of i8ot(^)es.

192
12-2 ] GASEOUS MIXTURES 193

for internal energy: The internal energy of a mixtitre of gases is equal


to the sum of the internal energies of the individval components taken
each at the temperature and volume of the mixture.

mixing processes
It is possible to devise hypothetical reversible

fbr which the entropy change during mixing may be computed.*


Gibbs’s Theorem may thereby be deduced: The entropy of a mixture of
gases is equal to the sum of the entropies of the individval components
taken each at the temperature and volume of the mixture.
The three rules above are sometimes presented as deductions from
a more general theorem known as the Gibbs-Dalton Law; hence
properties obtained by these rules are. said to be obtained by the
Gibbs-Dalton Law.
12-2 Perfect gas mixtures p, Fy T relationships. The pressure
:

of an individual component of a mixture, taken at the temperature


and volume of the mixture, is called the partial pressure of the com-
ponent. For a mixture of three perfect gas components A, B, and C
let Pa, Pb and pc be the respective partial pressures; then

Pa\' = mARAT = uaRT


PbV = mBRBT = hbRT
pcT^ — mcRcT = uqRT
where subscripts indicate properties of the individual components,
and the absence of a subscript indicates a property of the mixture.
By Dalton’s Law
p = Pa + Pb + Pc (2)

In the absence of chemical reactions the mols of mixture will be the


sum of the mols of components;

n = nA + ub + ric (3)
From (1), (2) and (3)
pV = nRT
Thus if Dalton’s Law is satisfied, a mixture of perfect gases follows
the perfect gas equation of state. Writing this equation in terms of
the mass and gas constant of the mixture, pV = mRT, Also, from
(1) and (2),
pV = mARAT -j- mBRBT H- mcRcT
*
See Zemansky, M. W., Heot and Thermodynamia. New York; McGraw-
Hill, 1943, chap. 16.
194 GASEOUS MIXTURES

Then the gas constant for the mixture is given by

mR = iHiRji + maRB + wicRc (4)

The gas constant of the mixture is thus the weighted mean, on a


masH *
basis, of the gas constants of the components.

Strictly a mixture has no molecular weight but an equivalent


molecular weight is defined by
tn R

Then from (3) and (5)

niA I
ma . me
Ma^ Ma^ Me
A quantity called the partial volume of a component of a mixture
is the volume that the component alone would occupy at the pressure
and temperature of the mixture. Designating the partial volumes
by Va, Vb and Vc,
PVa = tiiaRaT ^

pVa “ maRaT
pVc = mcRcT
From (1), (2) and (7)

V=Va + Vb + Vc
Observe that in the mixture each component occupies the total
volume;* therefore the specihe volume of each component is the
volume of the mixture divided by the mass of the component. In the
mixture •

V
mA

V
Vc
me
In a gOB which behaves like a perfect gas the molecules are so thinly dis*
*

tributed through the volume occupied that no component of a mixture interferes


appreciably with the freedom of the molecules of another component to roam
throughout the volume.

12-3] GASEOUS MIXTURES 196

V _ V
therefore ^
iw tnji *4* ma 4"

and ( 10)
V Va Vs Vc

COMPONENTS

Fig. 12-1. Partial preasures.

Mixture relations may be shown in diagrams; Fig. 12-1 shows


the mixture and the components each at the temperature and volume
of the mixture.
If each of the components
is compressed at constant tem-
perature until its pressure be-
comes p the diagram will be
as in Fig. 12-2.
12-3 Parts by mass
parts by volume—mol frac-
tions. Three common ways
of specifying the composition
of a mixture are: (1) in parts 12. 2 .

by mass; (2) in parts by vol-


ume; (3) in mol fractions. The part by mass of a component is

the ratio of the mass of the component to the mass of the mixture.
The sum of all the parts by mass is unity.

m m m m ^ ^

The part by volume a component of a mixture is the ratio of the


of
partial volume component to the volume of the mixture. The
of the
sum of the parts by volume is unity.
196 GASEOUS MIXTURES [m
Ll Is I
^ V = V=
. .

( 12 )
V V
The mol fraction of a component is the ratio of the number of mols
of the component to the number of mols of mixture. The mol frac-
tion is designated by x with a subscript; thus

il
(13)

The sum of the mol fractions is unity.

Xa
LI
+ Xb + Xc ”
n
= 1
1 (14)

Parts by volume are related to parts by mass as follows:

pVa = niARAT = ^

and pV = mRT = m^T


then
Va _mA M
(15)
V m Ma
Both the part by volume and the mol fraction of a component
are equal to the ratio of its partial pressure to the pressure of the
mixture.

pVa = PaV = HaRT


and pV = nRT

then xa = —=^ (16)^


^
• n p V
Parts by volume are encountered in analyses of flue gases or ex-
haust gas. Mol fractions are frequently of convenience in dealing
with mixtures because the use of mols in gas computations reduces
the number of different constants and properties that have to be
dealt with.
The above results depend upon Dalton’s Law which is not exact.
However, in cases for which the perfect gas law is a satisfactory rep-
resentation of the properties of the components, Dalton’s Law will
be satisfactory for mixtures of uniform constant composition.
124] GASEOUS MIXTURES 197

12-4 Internal energy, enthalpy, and specific heats of gas


mixtures. The rule for the internal energy of a gas mixture, for a
mixture of three components, leads to the equation

mu = MaUa + MbUb + rticUc (17)

From (2), (9), (17), and the definition of enthalpy

mh = + mbHb + mchc (18)

From the definitions of the specific heats it follows that

mcv = mACvA + mBCvB + mcc,c (19)

and mcp = mAC,A + mbCpb + mcCpc (20)

All of the equationsabove apply to components at the temperature


and volume of the mixture.
12-5 Entropy of gas mixtures. From the Gibbs Theorem,

ms = mASA + niaSs + (21)

where the components are at the temperature and volume of the


mixture. It is instructive to consider the entropy change of a system
during adiabatic mixing at constant vol-
ume. Let two gases A
and B be con-
tained in the two halves of an insulated
rigid chamber divided by a thin wall.
Fig. 12-3. With the pressures and
temperatures respectively equal in A
and B the wall is broken and the gases
are allowed to mix. Experience shows
that with perfect gases there will be no
temperature or pressure change. Taking
both gases as the system, the entropy before mixing is

msi = -h mBSBi

and the entropy after mixing is

mSi = mASAs + mBSBi


At state 1 each component has the pressure and temperature of the
mixture; at state 2 each component has the volume and temperature
of the mixture. For each component
198 GASEOUS MIXTURES [m
m Asu *= m^Cp In
^— fl In

Then the nha-ngR of entropy for the system comprising both compo-
nents, in the constant-temperature process, is

m As — MaRa In ^ — maRa In ^
p p
Since the partial pressures are less than the total pressure, the change
of entropy is positive. Recall from Sec. 8-7 that an increase of
entropy in an isolated system is a criterion of an irreversible process.
It is well known from experience that the mixing of gases by diffusion
is irreversible. Hence the result conforms to the Second Law.
12*6 Mixtures of gases and water vapor. Superheated water
vapor at pressures less than one atmosphere is approximately a per-
fect gas. Therefore the properties of gas mixtures (for example,
atmospheric air) containing superheated water vapor at low pressure
may be obtained by the Gibbs-Dalton Law. However it is also
necessary to consider mixtures of water and gas in wliich the water
vapor not superheated. If a mixture of superheated water vapor
is

and air comes into contact with a body of liquid water at the same
temperature, the partial pressure of the superheated vapor will be
leas than the saturation pressure of the liquid, and liquid will evapo-

rate until the vapor pressure (partial pressure in this case) equals the
saturation pressure for the existing temperature. The mixture is

then said to be a saturated mixture.* So long as a gas mixture is in


contact with a body of liquid water, the partial pressure of the water
vapor will be determined by the temperature of the system and the
pressure-temperature relation for saturation as given in the steam
tables. It is assumed here that time is permitted for equilibrium to
be established; it may take much longer for water vapor to distribute
itself uniformly throughout a space full of air than to fill to satura-
tion pressure the same space empty of air.
In a saturated mixture in contact with liquid water the Gibbs-
Dalton relations will hold for any given state, but the composition
of the mixture will be a function of the temperature and pressure,
and must be determined at each state by reference to the properties

* There u s email difference between the saturation preeeure of the pure eub-
etanoe and the partial pressure of the vapor in the saturated mixture, but this
will be neglected here.
12-6] GASEOUS MIXTUflES 199

of saturated vapor. be covered in more detail iu


This subject will

Chap. 25 but some elementary relations will be givei) below.


H a gas mixture containing superheated water vapor is cooled at
constant pressure, the mixture will eventually reach the saturation
temperature corresponding to the partial pressure of the water vapor.
This is called the dew-point temperature because it is often associated
with the condensation of liquid drops or dew from a gas mixture. A
common example of this t3rpe of process is the clouding of a window
on a cold day, when the temperature of the surface in contact with

the air becomes less than the dew-point temperature. Such a process
is indicated in Fig. 12-4 which is a temperature-entropy plot for the

water vapor alone. If the initial state of the vapor is at 1 (partial

pressure Pw, temperature Ti), constant-pressure cooling will change


the state eventually to d. The partial pressure of the vapor at d is
still pv because the mixture remains at constant composition during
cooling from 1 to d. But the partial pressure at d is also the satu-
ration pressure corresponding to the dew-point temperature Td’, there-
fore Td is the saturation temperature corresponding to p».
An apparatus in which a solid surface is cooled until condensation
appears may be used to determine the water-vapor content of moist
air as shown in the following example.

Example 1. In a dew-point apparatus a metal beaker is cooled by grad-


ually adding ice water to the water initially in the beaker at room tempera-
ture. The moisture from the room air circulating around the beaker begins
f

200 GASEOUS MIXTURES [12-6

to condense on the beaker when its temperature is 55^F. If the room tern*

perature 70^ and the barometric pressure is 15.0 psia, find: (1) the partial
is

pressure of the water vapor in the room air and (2) the parts by mass of
water vapor in the room air.

Solution: The partial pressure of the water vapor in the room air is

equal to the saturation pressure corresponding to the dew-point temperature


55®r; from the steam tables this is 0.2141 psia. By Dalton^s Law

p = PA + Pw
where Pa is the partial pressure of the air and p^, that of the water vapor.
Then
Pa = 15.00 - 0.2141 = 14.79 psia

By (16) the mol fractions are

^
Xa = — Va
^A = *—
Xv> — ^U>
— Vw
^
n V

But the mass of each component is given by

m = nM
therefore the parts by mass are

VIA - PaMa
m nM pM

Since the molecular weight of the mixture is unknown, eliminate the mixture
properties by taking the mass ratio of the two components; combining the
two equations,

wiA
^ VaMa
my, py,My, (0.2141) (18)

= 0.991
111 + 1

^m = -7T
112
= 0.00893

The composition of a mixture of and water vapor is often


air
given in terms of specific humidity, which is the ratio of mass of
water vapor to mass of air in the mixture. The composition is also
often stated in terms of relatwe humidity, which is the ratio of the
vapor in the mixture to the saturation
partial pressure of the water
pressure for water vapor at the temperature of the mixture. In
GASEOUS MIXTURES 201

Example 1 the partial pressure of the water v^or was 0.2141 psia;
at 70‘*F the saturation pressure p, is 0.3631 psia. Therefore in Ex-
ample 1 the relative humidity ^ was
^ p, 0.2141
^ “ 0.59
p, 0.3631

the specific humidity y was

_ _ VaM„ 0.0090 lb vapor/lb air


~ mA~ PaMa
Example 2. Air supplied to a furnace has relative humidity 75 percent,
temperature SOT, and gage pressure 10 inches of water. The barometer
reads 29.50 inches of mercury. How many pounds of water vapor enter the
furnace per pound of dry air?

SoLtmoN: As in Example 1,

wta = = VyMnt
.y
wiA PaMa
= 4^, = (0.75) (0.5069) = 0.380 psi

PA = P-Pv,

p = ^29.50 + 3^y0.491) = 14.85 psia

Pa - 14.47 psi

—=
niy,

niA
0.380 18
TT
14.47 29
= nnicoiu
0.0163 lb vapor/lb
/lu •

air

The brief discussion above of mixtures of water vapor and air is


sufficient for certain purposes, such as combustion calculations, in
which humidity is a secondary factor. A more detailed discussion
of humidity is given in Chap. 25.

PROBLEMS
12-1. One cu ft of1 atm, 70®F, is mixed with one cu ft of nitrogen
helium at
at 1 atm, For the mixture at 1 atm, 70°F, find: (a) the volume of the
70'’F.

mixture; (b) the partial volumes of the components; (c) the partial pressures of
the components; (d) the parts by mass of the components; (e) the mol fractions
of the components; (f) the specific heats of the mixture; (g) the gas constant
of the mixture; (h) the adiabatic exponent, k, of the mixture.
12-2. Show that the molal specific heat (Bj)u/lb mol F) of a gas mixture
is the weighted mean, on a mol basis, of the molal specific heats of the compo-
nents. Check the numerical results of Problem 12-1 against this rule.
12-3.One pound of helium and one pound of nitrogen are mixed at 70*’F
202 GAS£OU$ MIXTURES

and at a total pressure of 1 atm. Find all the quantitieB called for in Plroblem
12- 1 .
12-4. The mixture of Problem 12-1 is compressed reversibly and adiabati*
cally from 20 psia, 140*’F, to 40 psia^ Find: (a) tUe final specific volume of the
mixture; (b) the entropy change of the mixture; (c) the entropy change of the
helium; (d) the entropy change of the nitrogen. What should be the relation
among the answers to (b), (c)^ and (d)7

12-S. A stream of oxygen is mixed adiabatically in steady flow with a stream


of hydrogen, the mass ratio of oxygen to hydrogen being 8 to 1. The veloci-

ties of the mixed stream and the two component streams are equal. For one
pound of the total stream of oxygen and hydrogen find the change, during the
mixing process, in: (a) enthalpy; (b) internal energy; (c) entropy.
12-6. The products of combustion of a hydrocarbon fuel have the followmg
composition by mass: Ns, 0.72; CO2 ,
0.18; H O, 0.07; O
2 2, 0.03. Using the charts
in the appendix for properties of the components, find Cp and k for the mixture

at 1600T.
12-7. The products of combustion from a furnace have the following com-
position in fractions by volume: Ns, 0.77; CO2 , 0.13; O2 ,
0.05; H O,
2 0.05. How
much heat must be transferred, per pound of mixture, to cool the gas from 2000T
to 500T in steady flow at constant velocity? Take the properties of the com-
ponents from the chart of h/T in the appendix.

12-8. (a) Find the composition by volume of the gas m Problem 12-6.
(b) Find the dew point of the gas at a total pressure of 14.5 psia.

12-9. A mixture of air and water vapor containing 0.015 lb of water vapor
and 0.986 lb of diy air occupies a constant-volume tank at 14.7 psia and 80®F.
(a) What is the dew point of the mixture? (b) What is the relative humidity?
(c) To what temperature must the mixture be cooled at constant volume to reach
saturation?
12-10. An air compressor takes in air from the atmosphere at 75*’F, 14.0 psia,

and 75 percent relative humidity. The air is discharged from the compressor at
70 psia, 280^F. At the compressor discharge find: (a) the specific humidity;
(b) the relative humidity.

12-11. The air discharged from the compressor in Problem 12-10 is passed
through an aftercooler in which heat is transferred to metal tubes cooled by
circulating water. Some of the moisture in the air condenses on the tubes, and
the liquid is drained ofl separately from the air. The air then leaves the cooler
as saturated air at 68 psia, lOO^’F. How much water is removed from each pound
of mixture flowing into the cooler?

REFERENCES
For properties of mixtures and components see the references at the end of
Chap. 10, p. 168.
Chapter 13

PROCESS CALCULATIONS FOR


STATIONARY SYSTEMS

In this chapter examples Tidll be given of the solution of problems


involving several common types of processes in stationary fluid sys-
tAiTiH Ffl/th problem is solved for air and for steam to show how
the form of the available property data may affect the details of the

solution, even though the basic principles are identical for all cases.

13-1 Constant-vbliime process.

If frictionleas = 0

Q = au

^~^
For perfect gas

Example: A fluid at 100 psia and SOO'F has a volume of 10 cu ft. In a

frictionless process at constant volume the pressure changes to 50 paa. Find

the final temperature and the heat transferred.


(a) The fluid is air.

V S

Fig. 13-l(a).

Ti jh
203
204 PROCESS CALCULATIONS [13-1

960 ^
luO
480°R = 20“F

Q= me, AT =
^ c, Ar

= (2.81)(0.l72)(-480) = -231 Btu

Using gas tables, Q = rn Au

From Table 1 at 960°R, wi = 165.26

at480°R, Wo = 81.77

Aw = -83.49 Btu/lb
Q= (2.81)(-83.5) = -235 Btu

(b) The fluid is steam.

Fig, 13-1 (b).

From Table 3 at 100 psia, 500"F,

Vi = 5.589 cu ft/lb

m= —
Vi
= 10 1 70 lb
1.79 Ik
Vi 5.589

V2 = Vi
— 5.589

From the T-s chart at this volume and 50 psia the steam is at about 0.6
quality. Then using data from Table 2

Vx==V/ + XVjg
5.589 = 0.017 + a;8.498
X2 = 0.654

From Table 2,

Tt ^ 281.0rF

m^u
13-2] PROCESS CALCULATIONS 205

Ml = Ai - ^ PiBi = 1279.1 - 103 = 1176 Btu/lb

M* = M/ + iM/, = 249.93 + (0.654) (845.4) = 802 Btu/lb


Q= (1.79)(-374) = -670 Btu
13-2 Constant-pressure process.

If frictionless, Q = AH
W = p AV
For perfect gas Yl^L
Example: A fluid having a temperature of 350^^ and a specific volume
of 6 cu ft/lb at its initial state expands at constant pressure, without friction,
until the volume is doubled. Find, for 1 lb of fluid, the work, the heat
transferred, and the final temperature.
(a) The fluid is air.

V
Fig. 13-2(a).

pv = RT
(53.34)(810) ^
^
(6)(144)

TT = p = (50.1)(144)(12 - 6) = 43,400 ft Ib/lb

- 55.9 Btu/lb

Tt =Ti^=
Ki
(810)(2) = 1620"R = llfiOT

Q = c, AT = (0.240) (810) = 194 Btu/lb

Using gas tables, Q = Ah


From Table 1 at 810°R, hi » 194.25

at 1620“R, ht = 401.09
206 PROCESS CALCULATIONS [IS^

M= 207.84 Btu/lb

Note the difference when variation of the specific heat is accounted for; the
first answer was about 7 percent low.
(b) The fluid is steam.
z

»i = 6 cu ft/lb

At 350‘F Vf = 3.342 cu ft/lb; therefore at 1 the steam is superheated.


From Table 3, Pi = 77.5 psia

hi = 1204.8 Btu/lb

At state 2, p, = p, = 77,5

t;, = 2»i = 12 cu ft/lb

From Table 3 these properties correspond to

Ti = iioe^F

fh = 1586.7 Btu/lb

Q= M= 381.9 Btu/lb

W = pAv= (77.5)(144)(12 - 6) = 67,000 ft Ib/lb

= 86.0 Btu/lb

13-3 Constant-temperature process (isothermal).

If reversible, Q = T AS
For perfect gas, pV = constant

AU = AH = 0
Example; A 400°F and 50
fluid at p^
is compressed revermbly and

isothermally to 1/10 of volume, find the final pressure, the


its original

work done and the change of internal energy per pound of fluid.
(a) The fluid is air.
13-3] PROCESS CALCULATIONS 207

W= fpdv = piVila- = RTila-


J Vi »i

= (53,34)(860) In 1/10

= (46,000)(-2.3026) = -106,000 ft Ib/lb

= -136.5 Btu/lb

A« = 0
(b) The fluid is steam. From the tables at 400*^ and 50 psia,

Fig. 13.3(b).

i>i = 10.065 cu ft/lb

vt = (1/10)», = 1.0065 cu ft/lb

At 400°F this is in the mbcture region.

», = »,- (1 - i)iy,
1.0065 = 1.8633 - (1 - *)(1.8447)
1 — X » 0.463 moisture fraction

Pi is the saturation pressure,


208 PROCESS CALCULATIONS [13-4

P2 « 247.3 psia

The work could be obtained by fp do but this is inconvenient because


of the irregularity of the p-v curve for the process. A simpler way is to use
the First Law:
IF = Q — Am
Now for this process Q = T As and W can be found.
Ml = Ai — piVi = 1141.6 Btu/lb from Table 3

M2= ^2 — P 2V2 = 773 Btu/lb from Table 1

Am = —369 Btu/lb
81 = 1.7349 Btu/lb^R

82 = 1.082 Btu/lb^^R

W= (860)(-0.653) - (-369) = -193 Btu/lb


= -150,000 ft Ib/lb

13-4 Reversible adiabatic process.

In general, 0 = 0

AS = 0
W = -AU
For the perfect gas, pv^ = constant.

Example: A reversible adiabatic process begins with a fluid at pi = 150


psia, ti = 400®F, and ends with p2 = 15 psia. Find the final specific volume
and the work per pound of fluid.
(a) The fluid is air.

>

Fig. 13.4(a).
13-4 ] PROCESS CALCULATIONS 209
210 PROCESS CALCULATIONS [IM

«1 = 81

From Table 3,
=» hi — piVi

- 1219.4 - 89.6

= 1129.8 Btu/lb

ti = 1.5995 Btu/lb'’R

At Pj = 15 peia and st 8i = 1.5995 Btu/lb^R, the steam is a mixture.

8l = 8, - (1 - x)8/,

1.5995 = 1.7549 - (1 - 1)1.4415

1 — I = 0.107 moisture fraction

From Table 2, Mj = m, — (1 — a;)w/, = 981.3 Btu/lb .

W= ui-Ui= 148.5 Btu/lb

= 115,000 ft Ib/lb

wj = p, - (1 — x)v/, = 23.48 cu ft/lb

13-5 Reveraible poly tropic process.

pv” = constant

ExAHPiiE: A reversible polytropic process begins with a fluid at pi = 160


psia, Ti — 400®F, and ends with pj = 15 psia. The exponent n has the
value 1.15. Find the final specific volume, the final temperature and the
heat transferred per pound of fluid.

(a) The fluid is air.

- “ = (2.12)(7.40) - 16.7 cu ft/lb


-

13-6] PROCESS CALCULATIONS 211

V\Vl _
Tx
~ Tt

Tt - (880)^^ = - 180T

Q = Am h TT - c. AT -H
-
1 »

= (0.172 - 0.458)(-220)

= 63.0 Btu/lb

(b) The fluid is steam.

From Table 3, vi - 3.223 cu ft/lb

vt = (3.223)(7.40) = 23.9 cu ft/lb

At 15 psia, 23.9 cu ft/Ib, steam is a mixture.

From Table 2, 1* * 213®F

w^fpdv —n
J 1

^ (144)(15)(23.9) - (144)a50)(3.223)
-0.15

= 120,000 ft Ib/lb

« 155 Btu/lb

From Table 3, ui = Ai — piVi = 1129.8 Btu/lb


From Table 2, »i = — (1 — x)»/.
212 PROCESS CALCULATIONS [13-6

23.9 = 26.29 - (1 - 1)26.27

1 — I = 0.091 moisture content

ttj = ttj — (1 — x)u/f = 996.1 Btu/lb

Q = Am + IF
= -133.7 + 155

= 21 Btu/lb

13-6 Irreversible process in which pc'* = constant. It is pos-

sible to use the relation = constant to describe the path of a


process in which pressure and volume are suitably related, regardless

of whether the process is reversible or not. When the process is

irreversible, however, the work and heat cannot be determined from


the path alone. For example, when there is a lack of pressure equi-
librium or when any frictional effects exist, the work cannot be ob-
tained from J'pdv. If either the work or the heat is known from
other considerations, then the First Law can be used to evaluate the
unknown energy quantity. Actual processes may occur in which
the path is satisfactorily described by an equation of the polytropic
form but the system is well insulated so that the heat transfer may
properly be assumed zero. Such a process is adiabatic, but the expo-
nent n is not equal to the ratio of specific heats, k, because the
process is not reversible.

Example : A process in which pt>” <= constant has the same initial state

and final pressure and the same e.xponent n as in the example of Sec. 13-5
above; however, the process is not reversible but adiabatic. Find the final

specific volume, the final temperature and the work per pound of fluid.

(a) The fluid is air. Since the given relation, pv" = constant, and the
given conditions fix the initial and final states regardless of the nature of
the process, it follows that the properties at the final state must be the
same as in Sec. 13-5; then

»5 = 15.7 cu ft/lb

Ti = 640’’R = ISOT
Am = e,iiT = (0.172)(-220) = -37.8 Btu/lb

The work is not Xv ^ because the process is not given as frictionless.


*
Work must be found by the First Law;
TF = Q-Am = 0- (-37.8) = 37.8 Btu/lb
13-6] PROCESS CALCULATIONS 213

Compare this with the 101 Btu/lb work lor the process of Sec. 13-5 having

an identical path.

(b) The fluid is steam. By the same reasoning as above the property

changes are identical with those of Sec. 13-5:

V2 = 23.9 cu ft/lb

U = 213°F

Also W = -M= 133.7 Btu/lb

This compares with 155 Btu/lb for the process of Sec. 13-5 having an identi-
cal path.
Chapter iU

VAPOR CYCLES-THE RANKINE CYCLE

Beginiiing with this chapter the principles and methods developed


in the preceding chapters will be applied to engineering processes.
One of the first and most important applications of thermodynamics
is the study of the production of power from the combustion of fuel.

Thermal power plants may utilize the energy released by combustion


in either of two ways; (1) by transfer of heat from the combustion
products to a cyclic heat engine which produces work, as in steam
power plants, or (2) by the extraction of work directly from the com-
bustion products, as in the internal combustion engine. Heat power
cycles will be discussed first.

Up to the present time the only important cyclic power plants


have been those using condensable vapor, almost invariably steam,
as the working fluid. Internal combustion engines are not cyclic
heat engines; hot-air engines or closed-cycle gas turbine plants have
yet to achieve an important role in practice. First consideration will
therefore be given to vapor cycles.
14-1 Thermal power plants and ideal cycles. A cyclic ther-
mal power plant utilizes a series of processes each of which involves
some unnecessaiy or undesirable effects in addition to the intended
effect of the process. For example, in a steam power plant the pur-
pose of the boiler or steam generator is to convert liquid water to
superheated vapor at constant pressure. Actually, because of fric-
tion, the pressure of the superheated vapor may be appreciably less

than the pressure of the liquid. Similarly the purpose of the turbine
is to obtain shaft work from the expansion of steam to a lower
pressure; frictional effects and heat transfer are not desired but they
always occur. For each process in the power plant it is possible to
specify a hypothetical or ideal process which represents the basic
intended operation and involves no extraneous effects. For the
214
14-2] VAPOR CYCLES 215

steam boiler this would be a reversible constant-pressure heating


process; for the turbine the ideal process would be a reversible adia-
batic expansion. A cycle composed of such processes is called an
ideal cycle.
The efi&ciency and capacity (work output) of an ideal cycle, when
compared with the efficiency and capacity of the actual plant, give
a measure of the perfection of the plant apparatus. On the other
hand, comparison of different ideal cycles gives a measure of the
perfection of the thermodynamic design. Thus the ideal cycle is an
invaluable aid in understanding existing plants and in developing
improved plants.
14-2 The Rankine cycle. The basic operations involved in the
simplest type of steam power plant are (1) pumping liquid water

SAFETY STEAM
PRESSURE . VALVE
GAGE

EXHAUST TO
ATMOSPHERE
WATER
LEVEL
GAGE

FEED
WATER
VALVE

TURBINE

FROM WATER SOURCE


BOILER FEED PUMP

Fig. 14-1. Simple non-condensing steam power plant. The valves wd


shown are those which are essential for safe operation and control.

into a steam boiler or steam generator, (2) evaporation of the water


at a hi^ pressure by heat transfer in the boiler, (3) expansion of
the vapor through an engine or turbine to a low pressure, producing
shaft work. In the simplest case the vapor leaving the engine at
low pressure (exhaust steam) may be discharged into the atmosphere
and a fresh supply, or makeup, of boiler feed water must be used con-
216 VAPOR CYCLES [14-2

stantly to keep the plant in operation (see Fig. 14-1). It is possible,

however, to condense the exhaust steam by heat transfer, and to


return the resultant liquid {condensate) to the boiler feed pump for
return to the boiler (see Fig. 14-2). When this is done the plant
operates in a cycle according to the flow diagram. Fig. 14-3. The

Fig. 14-2. Simple condensing steam power plant. Details shown in Fig. 14-1
are omitted. The condenser
a shell-and-tubc heat exchanger. The exhaust
is

pressure may be below atmospheric pressure if an air pump is used to remove


non-condensable gas which leaks in from the atmosphere. When the non-
condensable gas content of the steam in the condenser is kept very small the
pressure in the condenser isthe saturation pressure for the temperature to which
the steam is cooled. This may be 20 or 30° above the cooling water temperature.
When operating with exhaust pressure below atmosphere, a condensate pump is

used to extract the condensate from the hot well and supply it to the suction of
the boiler feed pump.

two types of plant are called respectively non-condensing and con-


densing plants. Although the non-condensing plant does not operate
in a complete cycle it may be analyzed as a cyclic plant because the
make-up feed water is the same, thermodynamically, as the con-
densate that might have been obtained by completing the cycle.
The natural surroundings may be considered as the condenser, receiv-
ing the vapor and returning an equal amount of condensate.
14-2] VAPOR CYCLES 217

The simple steam plant is analyzed by the aid of the following


ideal cycle: (a) reversible adiabatic expansion, starting at the state
at which steam enters the turbine and ending at the pressure of the
exhaust steam leaving the turbine; (b) reversible constant-pressure
heat transfer from the steam, ending at the saturated liquid state;
(c) reversible adiabatic compression of the liquid, ending at the initial
pressure; (d) reversible constant-pressure heat transfer to the fluid,
ending at the initial state. Such a cycle, called a Rankine cycle, is
plotted on the p-r and T-a planes in Fig. 14-4; the numbers on the

Fig. 14-3. Elementary steam plant flow diagram.

plots correspond to the numbers on the flow diagram. Fig. 14-3.


For any given pressures the steam approaching the turbine may have
any of a variety of states as indicated by points 1, 1', and 1", but the
fluid approaching the boiler feed pump may, in the ideal cycle, have

only one state, that of saturated liquid.


For purposes of analysis the Rankine cycle is assumed to be car-
ried out in a steady flow operation. Applying the steady flow energy
equation to the processes of the cycle the equations for work and
heat quantities may be set up on a basis of unit mass of fluid, assum-
ing negligible changes of potential and kinetic energy.
Process 1-2, the turbine or engine process, is adiabatic; then
218 VAPOR CYCLES IlA-2

Fig. 14-4. Rankine cycle.

hi = ht Wxi-t

and the turbine work is given by

Wt = TT*!-* = Ai - ft* (1)

Process 2-3, the condenser process, involves no shaft work; then

ft* + Qi-t ~ ft*

The heat rejected Q*, is taken of opposite sign from Qt-» (see Sec.
7-2) ;
then
Qi = —Qi-z — ht — h» ( 2)
14-2 ] VAPOR CYCLES 219

Process 3-4, the pump process, is adiabatic; then

hi = hi

and the pump work is

Wp=-W^^ = hi-h (3)

Process 4-1, the boiler process, involves no shaft work; then

hi -j- Qi-i = hi

and the heat supplied is given by

Qi = 0+-1 — hi — hi (4)

The efficiency of the Rankine cycle is given by

_ l^net _ Wt — Wp _ h\ — hi — {hi — hi)


“ 5)
Oi Qi ~ hi -hi (

Cycle efficiency may be expressed alternatively as hecU rate; this

is the heat supplied per unit work output, Btu/hphr or Btu/kwhr.


Since the compression process is carried out with liquid water,
the specific volume is small and consequently the work of compres-
sion is small and is sometimes neglected. Assuming the pump work
negligible is equivalent to setting hi equal to hi. If this is done

Eq. (5) reduces to

approximate v = (5a)

In practice it is important to know not only the efficiency but


also the capacity ofa power plant. One measure of capacity is the
steam rate, the pounds of fluid that flow to the turbine or engine
per unit time per unit power output. Steam rate is usually given
in pounds per horsepower-hour or pounds per kilowatt-hour. For
the ideal Rankine cycle the steam rate will be

Btu of work in one hphr 2545 lb


(6 )
Btu of work done per lb of steam Wr — Wp hphr
or

Bs
Btu of work in one kwhr — 3413 lb
(6a)
Btu of work done per lb of steam Wt - IFj.kwhr

Example A
Rankine cycle operates with steam conditions 200 psia,
1.

750°F, and exhaust pressure 1 psia. Find the heat supplied, the turbine
220 VAPOR CYCLES [14*2

work, and the pump work per pound of steam. Find the cycle efficiency

and steam rate.

Solution: Obtain the enthalpy at states 1, 2 , 3, and 4, Fig. 14-4. From


the tables at 200 psia, 750°F, h= 1399.2 Btu/lb, 5i = 1.7448 Btu/lb^R.
By the definition of the cycle S2 = si. At 1 psia, s = 1.7448, find from the
tables - 0.874, = 976 Btu/lb. At 1 psia saturated liquid, hz = 69.70
Btu/lb By the definition of the cycle Sa = S3 in the liquid region, however,
;

it is inconvenient to use the tables to trace a constant-entropy path. A


convenient method of obtaining the enthalpy change in a reversible adiabatic
process in a liquid is by use of the general property relation from Chap. 9 ,

T ds = dh — V dp (9-6)
since ds is zero.
dh = V dp

In the liquid the volume change due to compression to 200 psia is negligible,
as can be verified in Table 4 of Keenan and Keyes. Assuming v is constant

Ah — V Ap
144
or ViipA - pz) - 0.01014^ (200 - 1) = 0.595 Btu/lb
778

hi = 69.70 + 0.60 = 70.3 Btu/lb


Then
Qi^ hi- hi = 1329 Btu/lb

Wt = hi — hi — 423 Btu/lb

Wp = hi- hi = 0.595 Btu/lb

=
Wr- Wp = 0.318
V
Q.

2545
w= = 6.01 lb steam per hphr

The efficiency of the llankine cycle may be presented graphically


by the use of the temperature-entropy plot, Fig. 14-5. Because the
Rankine cycle is a reversible cycle, the areas representing the integral
of T ds for the various processes are proportional to the heat quan-
tities transferred. Thus Qi is proportional to area 4156, Qt is pro-
portional to area 3256, and the net work, being equal to the net
heat, is proportional to the area 1234 enclosed by the cycle diagram.
The efficiency can readily be estimated if the plot is made to scale.
Such scale plots are useful in comparing the efficiencies of cycles
working at different pressures and temperatures, and in comparing
different types of cycles.
14-3] VAPOR CYCLES 221

Fig. 14-5.

14-3 —
Rankine cycle influence of pressure and tempera-
ture, The efficiency and steam rate of the Rankine cycle, as deduced
above, are dependent upon the initial steam state and the exhaust
pressure. It has been shown by the Second Law, however, that the
maximum efficiency of a heat engine cycle is limited by the tempera-
ture range employed. Therefore it is of interest to investigate the
effect of an increase in the initial temperature upon the efficiency of

Fig. 14-6.
222 VAPOR CYCLES [14-3

Fig. 14-7. Efficiency of Rankine cycle vs. initial pressure for cycles using satu-
rated steam.

the Rankine cycle. Assuming a reference cycle using saturated


steam, two possible ways of increasing the initial temperature would

be (1) by going to higher pressure with saturated steam, or (2 ) by


going to superheated steam at the same pressure.
The steam at higher pressure is shown
effect of using saturated
in Fig. 14-6. from the plot that the change of initial
It is evident
state from 1 to 1' has resulted in a decrease in the heat rejected,
since the area under 2-3 is less than the area under 2-3. The

Fig. 14-8.
14^] VAPOR CYCLES 223

effect upon the heat supplied is not so evident since the repre-
senting Qi is decreased by the portion shaded with horizontal hatching
and increased by the portion shaded with vertical hatching. The
net effect is an increase in Qi for increased initial pressure up to
about 450 psia. Above this pressure the trend reverses. Since the
efficiency of the cycle depends upon both Qi and Qj, and since Qt
constantly decreases with increasing pressure, the efficiency does not
reach a maximum when Qi does. The efficiency continues to in-

Fig. 14>9. Efficiency of Rankinc cycle vs. initial temperature for cycles using a
constant initial prrssun*.

crease until the rate of decrease of Qi exceeds the rate of decrease


of Q2 sufficiently to reverse the trend. The maximum efficiency is
reached (for usual exhaust pressures) at about 2500 psia as indicated
in the plot, Fig. 14-7.
The effect of increasing the initial temperature at constant pres-
sure is shown in Fig. 14-8. In this plot it is evident that when the
initial state changes from 1 to 1', Qi increases faster than Qt and
consequently the efficiency of the cycle will increase. A plot of effi-

ciency vs. initial temperature for certain initial and exhaust pressures
isshown in Fig. 14-9. The efficiency in this case does not pass
through a maximum.
224 VAPOR CYCLES [14-3

The efficiency a Rankine cycle may also be expected to increase


of
with decreasing temperature of heat rejection, that is, with decreas-
ing exhaust pressure. This effect is shown in Fig. 14-10 for a fixed
initial state. When the exhaust pressure is reduced from pi to
the heat supplied is increasedby the area under 4'-4. The heat
rejected shows an increase represented by the double hatching, and
a decrease represented by the single hatching. In the usual range

Fig. 14-11. Efficiency of Rankine cycle vs. exhaust pressure for cycles using a
constant initial state.
14-4 ] VAPOR CYCLES 225

of operation these quantities are about equal, so the heat rejected


(.hanges little with exhaust pressure. Hence the efficiency of the

cycle increases with decreasing exhaust pressure as shown in Fig.


14-11.
The characteristics of the ideal cycle as described above do not
alone determine the best pressure and temperature for an actual plant
because the characteristics of the actual apparatus, and especially
cost considerations, influence the final result. For example, the use
of superheated steam benefits the actual plant more than it does the
ideal cycle, giving increased turbine efficiency and reduced mainte-
nance. On the other hand the use of very low exhaust pressure is

not as beneficial in the actual plant as might appear in the ideal


it

cycle, because of the cost of apparatus and the work required to


maintain the low pressure.
14-4 Actual vapor cycle processes. The processes of an actual
cycle differ from those of the ideal cycle; in analysis the difference is
accounted for by certain conventional methods as described in the
following sections, so that by the use of experimental factors the ac-
tual cycle characteristics may be predicted from the ideal cycle
computations.
By definition the thermal efficiency of an actual cycle is

e.

where the work and heat (|uantities are the measured values for the
actual cycle; these will differ from the corresponding quantities for
the ideal cycle. In the actual cycle conditions might be as indicated
in Fig. 14-12. The machinery and pipe lines are not frictionless and
not adiabatic. The heat supplied and heat rejected are not the sole
heat transfers, consequently the net work is not equal to their differ-

ence. The work produced by the turbine is less than the work that
would have been produced by a reversible turbine and the work input
to the pump is greater than that required by a reversible pump.
(There may be several pumps; Wp is the total work for all of them.)
The various processes of the actual cycle are considered separately
in developing relations between the actual and the ideal cycle.
The actual turbine or engine, the actual pump, the actual boiler,
the actual condenser and the actual pipe lines are taken, each under
its actual imposed conditions, and compared with ideal apparatus
.

226 VAPOR CYCLES [14-5

Fig. 14-12. Steam plant flow diagram showing losses.

working under the same imposed conditions in each case. In this


chapter particular consideration will be given to actual engines,
pumps and pipe lines. It should also be noted that in boilers and
condensers there is some pressure drop between inlet and outlet;
moreover in the condenser the fluid is often cooled below the satu-
rated liquid state.Although these effects are not studied here they
cannot be ignored in the analysis of an actual plant.
14-5 The actual engine or turbine. The work output of the
actual turbine is compared to the output of a reversible adiabatic
turbine by means of the engine efficiency defined as follows:

r^T = ( 7)

where Wt is the actual turbine work, hi is the actual enthalpy of


the steam entering the turbine, and hzs is the enthalpy at a state
having the actual exhaust pressure and lying on an isentropic path
through the state 1
The actual steam rate, Wtj is given by
2545 lb
Wt (8 )
Wt hphr
14^] VAPOR CYCLES 227

The ideal steam rate, wi, is given by


2545 Ib
(9)
Ai — hta bphr

Then from (7), (8) and (9)

ijr
m ( 10)
Wt

_ 2545
Finally ( 11 )
Wrihi - hts)

Thus the engine efficiency can be determined experimentally by meas-


uring the turbine steam rate^ the initial state of the steam, and the
exhaust pressure.
The actual engine or turbine is certainly not reversible and it

may Applying the steady flow energy equation


not be adiabatic.
to the turbine in Fig. 14-12, assuming negligible elevation change
and velocity change,
+ 0 = ^2 + Wx
ftl

or Wx = Wt = — +Q= Qiom (12)

If a plot is made on the Ts plane for the process of expansion in


the turbine it may a number of paths as shown in Fig.
follow any of
14-13. For the reversible adiabatic turbine the path will be l-2s. For
an ordinary real turbine the heat loss is small and the turbine work

Fig. 14-13. Actual exhaust states.


228 VAPOR CYCLES [14-5

may be obtained from (12) as— A2 by setting Q equal to zero.


But work is equal to hi — h2s] then since
for the ideal turbine the
the actual turbine produces less work than the reversible adiabatic
turbine, it follows that for an adiabatic turbine A2 must be greater
than his^ The actual expansion might end at 2 in Fig. 14-13 for a
turbine of high efficiency, or at 2^ for a turbine of low efficiency. If

the heat loss is not negligible, as in a small turbine or engine, the


value of h 2 need not always be greater than h 2 s- As Eq. (12) shows,
the value of /t2 may be decreased at will by increasing the heat loss.
Thus it is possible for the end state of the expansion to be at point
2b in Fig. 14-13, if the heat loss is exceptionally large. It is also
possible to imagine a case in which the heat loss and the engine effi-

ciency would be so related that the entropy would remain constant


during the expansion process; such an expansion would be isentropic
but it would not be either reversible or adiabatic. The above reason-
ing is consistent with the Second Law; in an actual adiabatic process
the entropy must increase, but in a process involving heat transfer
out there is no restriction on the direction of entropy change.

Example 2. A turbine operating under the same steam conditions as


given for the cycle of Example 1 has a measured steam rate of 8 Ib/hphr.
Find the engine efficiency of the turbine and the state of the exhaust steam.

Solution: From Example 1, hi = 1399.2 Btu/lb, h 2 a = 976 Btu/lb.


The actual turbine work per pound of steam is, from the definition of the
steam rate,

Wt = —Wt
= 318.1 Btu/lb

then = ^ 318J 0.751


A, - his 423

The same result would be obtained from Ecp (10) taking the Rankine cycle
steam rate from Example 1 and assuming pump work negligible, so that

iVT = wa.

fi.Ol
n nr.
8.00

The state of the exhaust steam can he obtained only by applying the steady
flow equation, (12), Wt and
In this equation
hi are known, h 2 and Qioas
are unknown. Lacking information we may assume negligible and
obtain
A2 - - TTr = 1399.2 - 318.1 = 1081 Btu/lb
14- 6 ] VAPOR CYCLES 229

If an estimate or a measurement of heat loss could be made, a more accurate


result would be possible.

Except for very small turbines, heat loss from turbines is gener-
ally negligible. Heat losses from engines are relatively larger than
from turbines, but may still be small in the energy balance. The
kinetic energy of the steam leaving a turbine may in some oases be
appreciably larger than the kinetic energy approaching the turbine;
in such cases Eq. (12) needs a kinetic energy term.
14-6 The actual pump. The actual pump is compared to the
ideal pump by means of the pump efficiency, r)p, defined as follows:

(13)

where As is the actual enthalpy entering the pump, fus is the enthalpy
at a state having the actual discharge pressure and lying on the
isentropic path through state 3, and Wp is the actual work supplied
to the pump. By the steady flow equation, neglecting velocity and
elevation,

Wp ~ hi — hz Qiom (14)

The actual hi is found in the same manner as the actual hz in the


case of the turbine.
In discussing vapor cycles the pump work is often neglected but
it can be of great importance as will be shown later when the llankine
and Carnot cycles are compared. In any event the designer of a
plant must know what actual piunp work to expect if he is to provide
a suitable power supply for the pump.
14-7 Actual pipe lines. Although friction and heat transfer oc-
cur in all pipe lines in a power plant the major effects upon plant
efficiency come from losses in the turbine exhaust connection and in
the steam line connecting the boiler to the turbine. Losses in the
steam line are of special importance in modern high-pressure high-
temperature plants because the steam pipe is an extremely costly
item and its design may be a problem of economic balance between
plant efficiency, which demands a large pipe, and plant cost, which
demands a small pipe.
The evaluation of the effect of steam pipe friction and heat loss
upon the plant is made in terms of the effect of the pipe line losses
upon the performance of an ideal engine supplied by the pipe. Con-
230 VAPOR CYCLES [14-7

sider the pipe line of Fig. 14-14. The


states 1' and 1 at the entrance
and on the T-s plane in Fig. 14-15,
exit of the pipe line are plotted
with the ideal expansions that would occur if the steam were passed
through an ideal engine from each of these states. The difference
in the work of the two ideal engines is taken to be the work lost as
a result of the pipe line losses. The pressure drop in the pipe is
Pi “ Pi» the heat loss is h[ — hi. The work lost because of these
effects is
TV' - F = fi'i
- - (^i - h)
This can be analyzed to give the loss due to pipe friction and the loss

due to heat transfer.

Fig. 14-15. Steam pipe losses.


14-7] VAPOR CYCLES 231

Fig. 14-lC. Throttling loss.

Friction caudes the pressure to change at constant enthalpy as in


the throttling process (Sec. 9-6). Considering the friction loss alone
(Fig. 14-16), the lost work is given by

hi ~ h2 — (hia — hta)

since h'l = hu, the lost work is hia —


The work lost due to fric-
tion, shown graphically as the cross-hatched area in Fig. 14-16, is a
loss of availability (Sec. 8-6).

Fig. 14-17. Heat transfer loss.


232 VAPOR CYCLES [14-8

Heat loss alone causes the enthalpy to decrease at constant pres-


sure as indicated in Fig. 14-17. The lost work is given by

h'l
— h2 — {ha — ha)

The lost work, shown graphically as the cross-hatched area in Fig.


14-17, is not a loss in availability of energy supplied but is in effect

equivalent to a reduction of Qi by the amount of Qio«,. The lost


work is less than Q\on because in any event only a part of this energy
could have appeared as work, hence only this part represents a loss
of work.
14-8 Comparison of Rankine and Carnot cycles. Since the
Carnot cycle has the maximum possible efficiency for given limits of

Fig. 14-18.

temperature, it seems reasonable to compare the Rankine cycle with


the Carnot cycle and to inquire why the Carnot cycle is not preferred
in practice.
A Rankine cycle working with saturated steam differs thermo-
dynamically from a Carnot cycle only in the fact that heat is supplied
to the water at temperatures below the maximum temperature of
the cycle. In Fig. 14-18 the triangular cross-hatched area represents
additional heat supplied and additional net work for the Carnot cycle
as compared mth the Rankine cycle. Since the heat rejected is the
same for both Rankine cycle must be the less efficient.
cycles, the

Why then is the Carnot cycle not used? This can best be answered
by means of two illustrative cases.
14-8] VAPOR CYCLES 233

Take pi as 200 psia and p2 as 1 psia; it can be found in Table 4


of Keen an and Keyes that the isentropic compression of liquid water
from saturation at 1 psia to 6000 psia (the limit of the' table) would
raise thetemperature from about 102“F to about 104*^. To reach
point 4c of Fig. 14-18, at 382“F, by isentropic compression from
saturated liquid at 1 psia would be utterly beyond reason in a 200
psi power plant. A rough estimate indicates that a pressure of
the order of 100,000 psi would be required.
A different Carnot cycle having the same efficiency as that of
Fig. 14-18 can be arranged so that the isentropic compression process

Fig. 14-19. Comparison of Rankine and Carnot cycles, using saturated steam.

ends on the saturated liquid line, process 3c-4c in Fig. 14-19. It is

clear from the diagram that this Carnot cycle has a smaller capacity
(net work per pound of steam) than the Rankine cycle, but it will
still have higher efficiency, and the range of states through which

the fluid must pass is now even smaller than for the Rankine cycle.
However, this cycle also suffers from practical difficulties; first that
of carrying out the specified process, and second the large pump work
or back work of the cycle.
The process of condensing steam from state 2 to some particular
mixture state, 3c, is not easy in practice because of the difficulty of
measuring and controlling the quality continuously. The problem
of back work has two aspects: (1) it reduces the capacity of the
plant even in the ideal case, and (2) when actual machines are used
234 VAPOR CYCLES [IAS

it reduces both the capacity and the efficiency intolerably. This can
best be shown by an example.

Example Compare the Ranhine and Carnot cycles shown in Fig.


3.
14-19 with respect to efficiency, steam rate, and back work. Take pi as
200 psia and p: as 1 psia. Consider first the ideal cycles and second, cycles
in which the engine efficiency and pump efficiency are both 70 percent.

Solution: The enthalpies at the various states shown in Fig. 14-19 are
tabulated below.
hi = 1198.4 Btu/lb

hts = 863.5

A,je = 69.7

h,R = 70.3

htc
~ 309.7

hic = 355.4

Then, for the Rankine cycle,

Wt = 334.9 Btu/lb

Wi> = 0.6

TV„t= 334.3

Qi = 1128.1

n = 0.295

wr ” 7.6 Ib/hphr
P’or the Carnot cycle
Wr = 334.9 Btu/lb

Wp = 45.7

289.2

Qi = 843.0

=ij 0.341

wc = 8.8 Ib/hphr

The actual cycles are sketched in Fig. 14-20. For cycles using turbines
and pumps of 70 percent efficiency the turbine work will be 70 percent of
the ideal turbine work, and the pump work will be 1/0.7 times the ideal
pump work. For the Rankine cycle with actual machines

Wt = 234.4 Btu/lb

Wp - 0.9
14-8] VAPOR CYCLES 235

Fig. 14-20.

TV„t= 233.5

actual hiB — hiB -f" Wp — 70.6

Qi = 1127.8

ri = 0.207

wb = 10.9 Ib/hphr

For the Comot cycle with actual machines

Wt = 234.4 Btu/lb

Wp = 65.3

Tr„t= 169.1

actual Kc = h^c + Wp = 375.0

Q, = 823.4

f) = 0.205

Wc = 15.0 Ib/hphr

From the results above the serious effects of turning back a large
amount of work into the cycle to drive the pump become obvious.
The back work in the Bankine cycle is only about 0.2 percent of the
turbine work; hence the turbine need be only of a size corresponding
to the net work output. In the Carnot cycle the back work is almost
14 percent of the turbine work; hence the turbine must be about
16 percent larger than the Rankine turbine for the same output
This is the only objection to the back work in the ideal cyde, but in
the actual cycle the situation is far worse.
236 VAPOR CYCLES [14-8

In the actual Rankine cycle the influence of the pump upon the
cycle is negligible because the pump work is so small; hence a machine

efficiency of 70 percent, reducing the turbine work by 30 percent,


results in essentially the same reduction of cycle efficiency and ca-
pacity. In the actual Carnot cycle the machine efficiency of 70
percent also reduces the turbine output by 30 percent, but since the
ideal cycle net work is only 86 percent of the turbine output the net
work is reduced by 40 percent because of turbine inefficiency alone.
When the pump inefficiency is considered, there is a further reduction
of cycle efficiency. The final result is that under the given conditions
the Carnot cycle has no better efficiency than the Rankine cycle,
and has a far greater steam rate.

Fig. 14-21. Comparison of Ilankine and Carnot cycles, using su{}erheated steam.

The regenerative cycle, described in the next chapter, is a modifi-


cation of the Rankine cycle by which, if saturated steam is used, the
Carnot cycle may be approximated in a practical manner.
The Rankine cycle using superheated steam
is not a good approxi-

mation to the Carnot cycle. Fig. 14-21 shows a Rankine cycle


l-2-3-4jj with a Carnot cycle l-2-3-4c working between the same
extreme temperatures. It is apparent that the Rankine cycle has
much smaller efficiency than the Carnot cycle because the net work
area is much smaller while the heat rejected is identical for the two
cycles. However, even though with increasing superheat the Rankine
cycle becomes relativeli/ poorer compared to the Carnot cycle, it is

still true that the thermal efficiency and the capacity of the Rankine
VAPOR CYCLES 237

cycle are increased. Since superheated steam also contributes to


greater engine efficiency of turbines, it is invariably used in steam
power plants where efficiency is important.
238 VAPOR CYCLES
phere at 16 psia. The feed water is taken from a lake at 65^F, and before entering
the boiler it is heated to ISO^F in a feed water heater; the source of heat is the
exhaust steam on its way to the atmosphere. The steam rate of the plant is
82 Ib/hphr. The boiler efficiency is 70 percent (Le. 70 percent of the fuel heating
value is transferred to the steam).
(a) Sketch a flow diagram for the process, (b) Find the thermal efficiency
of the plant (work outputA^ating value of the fuel consumed), (c) Find the
efficiency of an ideal Rankine cycle based upon the steam and exhaust conditions
at the actual engine, (d) Find the difference between the heat rates of the ideal
cycle and the actual plant, (e) Show approximately how much the heat rate
could be reduced by substituting an idealized component for each of the follow-
ing: (1) the steam pipe; (2) the engine; (3) the feed water heater; (4) the boiler.
In each case only one component is to be idealized.
Chapter 15

VAPOR CYCLES-MORE EFFICIENT


CYCLES

The simple Rankine cycle has limited efficiency, depending upon


the properties of the working substance, as shown in Sec. 14-3.
Moreover the characteristics of available apparatus set limits upon
the pressures and temperatures that can be used, thereby preventing
full exploitation of the cycle possibilities. This chapter shows how
the limits of efficiency may be extended by modifications of the cycle
or by using different working fluids. The cycles described in this
chapter are: the reheai cycle, used in large-capacity central station
plants; the regenerative cycle, now used in every steam power plant
of any significance; the hy-jrroduct power cycle, long used in industrial
plants and now being adopted, where possible, by many public utili-

ties; and the mercury-steam binary cycle, used in a few cases where
very high efficiency is required in plants of medium capacity.
15>1 The reheat cycle. It is a general deduction from the Sec-
ond Law that the efficiency of a heat engine cycle may be increased
by supplying heat at higher temperature. For a cycle like the
Rankine cycle, in which heat is supplied through a range of tempera-
tures, the efficiency will be increased if a greater proportion of the
heat is supplied at higher temperatures. One way of doing this is
by using the reheat cycle. Fig. 15-1.

In the reheat cycle the expansion of steam from the initial state

is carried out in two (or possibly more) steps. The first step is from
the initial state to approximately the saturation line, process 1-2 in
Fig. 15-1. Following this partial expansion the steam is resuper-
heated at constant pressure, process 2-^, and the remaining expan-
sion, process 3-4, is carried out. The resuperheating is usually done
in heating tubes incorporated in the boiler structure to absorb heat
15- 1 ] MORE EFFICIENT CYCLES 241

^ Wt-Wp ^ - ih-K)
^ Qi. hi — ht fh — hi
If the pump work is neglected,

= hi — hi hj — hj
approximate
hi — hi hi — hi

It may be observed that in practice the reheat cycle is used with


high pressures where the pump work begins to be appreciable.
The steam rate of the reheat cycle of Fig. 15-1 is

2545 2545
w = — —
Ib/hphr
hi — hi hi hi — {hi hi)

The reheat cycle may be evaluated relative to either of two basic


Rankine cycles, depending upon the point of view. An obvious as-
sumption is that the area 2-3-4-2' in Fig. 15-1 has been added to
the basic cycle l-2'-5-6. On this basis a decrease in steam rate
(increase in work per pound of steam) is apparent on the T-s plot;
whether the efficiency improves depends upon whether the average
temperature of heat supply in process 2-3 is higher than the average
temperature of heat supply in process 6-1. Usually in the ideal
cycle there is a small improvement in efficiency. From a practical

point of view, however, the reheat cycle is used only when the
Rankine cycle 1-2 -5-6 is unattainable for reasons to be explained
below. Hence the proper comparison between the reheat cycle
is

and the best Rankine cycle which could have been used. This
Rankine cycle would have the same initial temperature but a lower
initial pressure than the reheat cycle. To understand this situation
some information on the characteristics of turbines is needed.
The engine efficiency of a turbine is reduced by the presence of
moisture in the steam; but what is worse, the presence of more than
about 10 or 12 percent moisture results in excessive erosion of turbine
blading. This reduces the efficiency still further and necessitates
costly repairs. Consequently good practice requires that the maxi-
mum moisture content of the steam be limited to about 10 or 12
percent. The T-s diagram in Fig. 15-2 shows that with a given
exhaust pressure, if the moisture is limited to 10 percent, there is a
limit on the steam pressure that can be employed with any given
initial temperature. If T# is the maximum steam temperature per-

missible with the available materials of construction then pa is the


242 MORE EFFICIENT CYCLES [15-1

maximum permissible pressure for an isentropic turbine. (In the


real turbine the expansion will not be isentropic, so the limiting
pressure will be somewhat higher, but a limit will still exist.) If

better high-temperature materials become available, the limits may

r.

Fig. 15-2.

increase to T\, and p6. Fig. 15-2. A table of reasonable steam pres-
sures at given temperatures, for condensing cycles in modem practice,
is as follows:
<1 Pi
760'’F 400 psi
825 600
900 850
950 1250
1050 1500

For comparison, reheat cycles with temperature TSO^F have been


used with initial pressure 1450 psi; and with temperature 950°F
reheat cycles have been used with 2400 psi initial pressure.
The actual utilization of the reheat cycle has been related rather
closely to the development of high-temperature metals. When car-
bon steel limited steam temperature to 750°F, reheat was used to
obtain higher efficiencies than the Rankine cycle could give at per-
missible pressures. Then alloy steels came into use and temperatures
could be raised to OOO^'F, permitting a corresponding increase of
pressure and giving improved efficiency without the complication of
15-1] MORE EFFICIENT CYCLES 243

reheating; thus the reheat cycle was for a time less attractive. The
economic incentive for still higher efficiencies has again forced designs
to the upper limits of temperature, which have gradually been raised
to about llOO^F} but improvements in metals have not kept pace
with the need for more efficient cycles, so it has again become neces-
sary to accept the complication of reheating in order to satisfy this
need*
The following example is given to illustrate the above discussion.

Example 1. A reheat cycle is to operate with turbines of 85 percent


engine efficiency, but otherwise with idealized processes. The imtial pressure
is 2000 400 psia, and exhaust pressure 0.6 psia. The
psia, reheat pressure
maximum temperature of lOOO^F is to be used for both initial and reheat
temperature. Find the efficiency and steam rate of this cycle; also of a
Rankine cycle working between 2000 psia, 1000®F, and 0.5 psia; also of a
Rankine cycle working between 1400 psia, 1000®F, and 0.5 psia, this being
taken as the cycle of maximum permissible pressure without reheat. Use
turbines of 85 percent engine efficiency in the Rankine cycles.

Solution: The cycles to be computed are illustrated in Fig. 15-3. Tie


properties at the various states are given below. For the reheat cycle:

ki = 1474.5 Btu/lb Table 3

Si = 1.5603 Btu/lb *R
hts = 1277.5 Btu/lb Mollier chart

hi = hi — 0.85(Ai “ h*s)

= 1307 Btu/lb

hi — 1522.4 Btu/lb Table 3

Si = 1.7623 Btu/lb “R
his = 948 Btu/lb Mollier chart

hi — hi ~ 0.85(Ai ~ fcis)

= 1033 Btu/lb

hi — 47.6 Btu/lb Table 2

hi = 53.5 Btu/lb Fig. 3, K and K.


For the first Rankine cycle:

hts — 840 Btu/lb Mollier chart

At = Ai - 0.85(Ai - hia)

= 935 Btu/lb
15-2] MORE EFFICIENT CYCLES 245

For the second Rankine cycle:

h= 1493.2 Btu/lb Table 3

58 = 1.6093 Btu/lb

hqs = 866 Btu/lb Mollier chart

hq = hi — 0.85(/l8 — hqs)

= 960 Btu/lb

An = 51.5 Btu/lb Fig. 3, K and K.

The efficiencies and steam rates computed from the above data by equations
given in this and the preceding chapter are tabulated below.

Moisture content
V w of exhaust steam

Reheat cycle 0.401 3.9 Ib/hphr 6 percent


First Rankine cycle 0.375 4.75 15.3
Second Rankine cycle . . . 0.366 4.81 12.9

The improvement in efficiency by using the reheat cycle, compared to the


Rankine cycles is for the first case

0.401 - 0.375
7 percent
0.375
and for the second case

0.401 -
0.366
— =9.6
0.366 ^ ,
percent

The first Rankine cycle, however, would not be used; therefore the reheat
cycle is approximately 10 percent more efficient than the Rankine cycle it

replaces and its steam rate is 18 percent less.

The actual reheat cycle will lose a percentage point or two from
these figures because of pressure drop in the reheating piping system;
the actual Rankine cycle will have slightly lower turbine efficiency
because of higher moisture. To determine
any given case whether in
the reheat cycle should be used, the net gain in efficiency must final

be balanced against the additional installation and operating costs


of the reheat plant.
It should be observed that actual reheat cycles are always built
upon the regenerative cycle described in the following section, rather
than upon the simple Rankine cycle. This does not alter the conclu-
sions from the above example.
15-2 The regenerative vapor cycle. It is possible to increase
the average temperature of heat supply to a vapor cycle not only by
246 MORE EFFICIENT CYCLES [15-2

Fig. 15-4.

increasing theamount of heat supplied at high temperatures, but also


by decreasing the amount of heat supplied at low temperatures. In
the saturated steam Rankine cycle of Fig. 15-4 a considerable amount
of heat is supplied in the process 4-4' to water at a temperature
lower than Ti, the maximum temperature of the cycle. For maxi-
mum efficiency all heat should be supplied at but this requires
Tj,

that means be found of raising the temperature of the liquid without


16-2 ] MORE EFFICIENT CYCLES 247

taking heat from a source external to the system. This is done by a


cycle in which heat transfer between two parts of the thermodynamic
system takes the place of externally supplied heat at low temperatures.
Such a cycle is called a regenerative cycle.
The purpose of the ideal regenerative cycle can be accomplished
only if means can be found, by reversible processes, to make available
at all temperatures between T* and Ti energy which was received
by the system as heat at Ti 74 ). In the ideal Rankine
(or in this case
cycle the fluid expanding through the turbine passes reversibly

through all the temperatures from Ti to T2 ;


therefore it forms a
suitable source of heat for the process 4-4'. Imagine a turbine, Fig.
15-5, containing a heat exchanger consisting of tubes so arranged that
the boiler feed water can pass through them on its way from the
pump to the boiler. The water passes in the direction opposite to
the steam flow, absorbing heat from the steam and thereby rising in
temperature to the initial steam temperature by reversible heat

transfer. In such a case the ideal expansion process in the turbine


is not a reversible adiabatic process but a reversible process in which
heatis transferred. In Fig. 15-6 the resulting cycle is plotted on the
T~s plane; the quantity of heat Q, is transferred from the steam to
the liquid water.
248 MORE EFFICIENT CYCLES [15-2

Assuming reversible heat transfer, for any small step in the process
of heating the water,

Ar (water) = —AT (steam)

and As (water) = — As (steam)

Then the and 4'-3 on the T-s plot will be identical


slopes of lines 1-2'
at every temperature and the lines must be identical in contour.
Now for the complete cycle all external heat is supplied at Ti and all
heat is rejected at Ti. Then
Qi = hi — = Ti(si — si)

Qi — hi — hs = Ti(ii — Sj)

But by the similarity of the paths 1-2' and 4'-3


si — s» = Si — si

or Si " 54 ““ S2 — S3

Then Qi^Ii
Qt Ti

and the efficiency of the regenerative cycle is equal to the efficiency


of the Carnot cycle. The two cycles are compared graphically in
Fig. 15-7.
15-21 MORE EFFICIENT CYCLES 249

The ideal regenerative cycle using saturated steam will be more


efficient than the Eankine cycle but will have a larger steam rate;
the transfer of heat in the expansion process results in a loss of work,
represented in Fig. 15-6 by the area
1-2-2'. The pump work of the
ideal regenerative cycle however, the same as that of the Rankine
is,

cycle and is much smaller than that of the Carnot cycle (see Sec.
14-8).
The ideal cycle described above cannot be used for the following
reasons: (a) reversible heat transfer is unobtainable in finite time,

(b) the heat exchanger in the turbine is mechanically impractical,


(c) the process described would increase the moisture content of the
steam in the turbine, an undesirable result for reasons explained
under the reheat cycle. However, it is possible to arrange an extrac-
tion cycle which approximates the regenerative process in a practical
way. The staged regenerative process, which is universally used in
modern vapor cycles, differs from the ideal process as follows: (a) the
boiler feed water is heated in a number of steps or stages, using tem-
perature differences that are finite but much smaller than in the
water-heating process of the Rankine cycle, (b) instead of condensing
a portion of the steam within the turbine and permitting the moisture
to flow on through the machine, the required amount of steam is
extracted from the turbine and completely condensed in a separate
feed water heater.
A flow diagram of a regenerative cycle using two stages of feed
water heating is shown in Fig. 15-8, with the corresponding tempera-
ture-entropy diagram in Fig. 15-9. The total steam flow to the
turbine, tci, expands to state 2 where steam is extracted at the rate
w%-, this leaves wi-wi as the flow rale from 2 to 3, where wz is extracted.
The remaining steam flows at rate Wi to the condenser. From the
condenser the water is pumped to the pressure of state 3 and then
mixes in a heater with the steam extracted at 3. From the heater
at state 6 a stream consisting of Wz plus Wz is pumped to the next
heater where the extracted steam Wz rejoins the boiler feed water;
the total flow is then pumped back to the boiler.
In the staged heating process described above the heating of the
feed water is not reversible but in each stage the temperature differ-

ence, forexample Tj — Tc, is only one-third of the difference Ti — Tt


between the boiler feed water in the Rankine cycle and the saturated
water in the boiler, which is the source of heat for the feed water in
250 MORE EFFICIENT CYCLES [15-2

the Rankine cycle. If more stages of heating were used the tempera-
ture difference could be reduced further and the heating could be
made more nearly reversible. The relation to the reversible cycle
may be seen in Fig. 15-9; the path 1-2-3-4: represents the states of

Fig. 15-9. Practical regenerative cycle.


15-2 ] MORE EFFICIENT CY*CLES 251

a decreasing mass of fiviid. If the extracted steam had been con-


densed within the turbine, and flowed through to the condenser, the
states of the total mass of fluid would be represented by the dotted
path It is seen that the stepped cycle l-2'-3'-4'
c-c'-h-b'-a-a' approximates the ideal regenerative cycle of Fig. 15-6,
and that a greater number of stages would give a closer approxima-
tion.
The work and heat quantities and the efficiency of a staged regen-
erative cycle are easily written in terms of the flow rates and en-
thalpies. For the cycle of Figs. 15-8 and 15-9 the following are
found:

Turbine work

Wt = wi{h - hi) + (wi - Wi)(hi - hi) + (wi- Wi- Wt)(hi - h)


Heat supplied
Qi = WiQii — ha)
Cycle efficiency

where "LWp is the sum of all


Wt

pump work
~~
^ STFp

quantities.
To obtain numerical values, the enthalpies and flow rates must
be known. In an idealized cycle the temperature levels in the heat-
ing of the feed water are usually chosen arbitrmily to divide the total
temperature range between condenser and boiler into equal steps.
Then, if the turbine process is reversible and adiabatic, the diagram
of Fig. 15-9 is completely defined and all the necessary enthalpies
can be found.* The flow rates may then be found by applying the
steady flow equations as follows. Set up a control surface around
the highest pressure heater as shown in Fig. 15-8, and write the equa-
tions for energy and mass, assuming heat transfer through the control
surface, kinetic energy change, and potential energy change are
negligible.

Then wjii -I- Wih, = wjha + wiWn


and Wi + Wt, — Wa
Now Wa — Wi

* must be
With a real turbine the detailed path of the expansion process
known to locate the extraction points such as 2 and 3, Fig. 15-9.
252 MORE EFFICIENT CYCLES [16-2

so Wi = Wi — Wt

Then combining and rearranging

Wihi + (wi — Wi){hb + Wn) = WiK


Wiiha — hb — W Pi) = Wt(hi — hb — Wpb)

yh _ h„ — hb — Wpb
Wi hi — hb — W pb
For the next heater, by similar reasoning,
vh _ hb — K — Wpc
Wi — Wi hi — he — Wpc
The procedure can be repeated for any number of heaters, the equa*

tion giving for each extraction point the fraction of the steam ap-
proaching that point which is there extracted. Starting with any
given throttle flow, Wi, the flow in all branches of the system can thus
be determined.
When superheated steam is used in a regenerative cycle the cycle
cannot, even in the ideal case, approach the Carnot cycle as a limit
l)ccause only an infinitesimal amount of heat is supplied at Ti; never-
theless the elimination of the external heat supply to the feed water
is still advantageous.
As illustrated in Fig. 15-10, an extraction point occurring in the
superheat region is located not by the temperature of the feed water
heater, To, but by the saturation pressure pt corresponding to this

Fig. 15-10, Kegonorative cyclo with supc'rh(*aU‘(l stt^am.


16-2] MORE EFFICIENT CYCLES 253

temperature. Since T2 is greater than Ta, the heating process suffers


some added irreversibility compared to a process receiving saturated
steam at Ta- This effect can be minimized by suitable heater design.
In actual regenerative cycles the steam and water may be mixed
in open or conUut heaters as shown in Fig. 15-8, but almost always
some of the heaters are dosed heaters in which the two fluids do not
mix. In this case the feed water passes through tubes which are
immersed in a shell containing the heating steam; heat must be trans-
ferred through the walls of the tubes. Eventually the condensed
heating steam (called the heater drip) must be returned to the boiler

EXTRACTED EXTRACTED

Fig. 15-11. Feed heater drip flow diagram.

in order to complete the cycle. One method of doing this is shown


in Fig. 15-11. In the arrangement shown here the drip from the
higher pressure heater is removed by a trap which will pass liquid
but not vapor. This liquid at the temperature of the higher pressure
heater flows to the shell of the lower pressure heater and thereby
reduces to some extent the steam required by that heater. The drip
from the second heater could be similarly trapped to the condenser
but this would be throwing away energy to the condenser cooling
water. To avoid this waste the drip pump returns the drip directly
to the feed water stream leaving the heater at about the same tem-
perature as the drip. If the lower pressure heater had been an open
heater the drip pump would have been unnecessary. Thermody-
namically the use of the drip pump makes a closed heater equivalent
254 MORE EFFICIENT CYCLES [16,2

to an open heater in its general effect on the cycle, althou^ differ-

ences exist in details.


The advantages of the open heater are simplicity, lower cost and
high heat-transfer capacity. The disadvantage is the necessity of a
pump at each heater to handle the large feed water stream. A closed-
heater system requires only a sinfdc pump for the main feed water

stream regardless of the number of heaters. The drip pumps, if

Fig. 15-12. Reduction in heat consumption obtained by regenerative feed water


heating. (Courte^ Westinghouse Electric Corporation.)

used, are relatively small. Closed heaters are costly and may not
give as high a feed water temperature as open heaters. In most feed
water systems closed heaters are favored but at least one open heater
is used, primarily for the purpose of feed water de-aeration.

The choice of the number of stages of feed water heating and of

the temperatures at the various stages in an actual cycle is controlled

by economic and physical considerations. An increase in feed water


temperature may, in some cases, cause a reduction in the efficiency
of the steam boiler while improving the efficiency of the cycle. The
hnlftiwing of such effects often makes the most efficient cycle condi-
15-2] MORE EFFICIENT CYCLES 255

tion differ from the most economical plant condition. The physical
structure of the turbine limits the choice of extraction pressures; also
the desirability of utilizing “waste heat” from various sources for
heating feed water may influence the choice of heater temperatures.
Numerous heaters, especially for hi^ pressures, add greatly to the
cost of a plant.
It is of interest to observe the general trends of cycle efficiency
with changes in the number of stages of heating and in the final

Fig. 15-13. Typical modem regenerative cycle. (Courtesy Westinghouse Elec-


tric Coijxmtion.)

temperature of the feed water. Fig. 15-12 shows the improvement


in efficien<^ over a Rankine cycle for regenerative cycles using various
numbers of heaters. It will be observed that if n heaters are used,
the greatest gain in efficiency occurs when the overall temperature
rise is about n/(n + 1} times the difference between condenser and
boiler saturation temperatures. It will also be observed that the
efficiency gain follows the law of diminishing returns as the number
of heaters is increased.
Figure 15-13 shows the flow diagram for a typical modem regen-
erative eycle.
256 MORE EFFICIENT CYCLES [IM

15-3 By-peoduct power and heat. The concept of the effi-

ciency of a heat engine cycle attaches ultimate value only to energy


in the form of work. However, in many places there is use for heat
as well as work, and then the value of a heat supply is not measured
solely by the work that can be obtained from it. Steam is used for
heating in many industrial processes as well as for comfort heating of
buildings.

The steam necessary for a heating system can be generated espe-


cially for that purpose; in this case the steam system is considered
to have performed perfectly if the heat transferred for a useful pur-
pose is equal to the heat received by the steam at the same tempera-
ture. The flow diagram and T-s plot for such an idealized ssrstem
are shown in Fig. 15-14 in which Qu is the heat transferred to some
process.
It is also possible to generate steam at a higher pressure than pi,
expand it through a turbine to obtain work, and exhaust from the
15-3] MORE EFFICIENT CYCLES 257

Fig. 15-15. By-product power Rcheme.

turbine at pa to the process heating plant. Such an arrangement is

pictured in Fig. 15-15. It is apparent that the process heater takes


the place of a condenser for the power cycle but the rejected heat
has become useful energy rather than waste energy. Assuming that
there is use for both the net work and the heat rejected, all the heat
supplied can serve a useful -purpose. In a real plant there would be
some losses to the surroundings, but there is no inherently wasted
energy as in the case of a plant existing to produce work only. The
arrangement is called a by-product power cycle on the assumption
that the process heat is the basic need and the power is produced
incidentally. In this case the cost of power is only the aMitional
cost over the cost of the process heat alone. Since part of the installa-
tion and labor cost is borne by the process heating, and since the
additional heat supplied for power is utilized at a very high efficiency,
the by-product power scheme may offer great economies when prop-
erly applied.

Example 2. A steam plant operates with initial pressure 250 psia and
temperature 700^F, and exhausts to a heating system at 25 psia. The
condensate from the heating system is returned to the boiler plant at 150®F,
and the heating system utilizes for its intended purpose 90 percent of the
energy transferred from the steam it receives. The turbine efficiency is 70
258 MORE EFFICIENT CYCLES [ 15 ^

percent, (a) What fraction of the enei^ supplied to the steam plant serves

a useful purpose? (b) If two separate steam plants had been set up to pro-
duce the same useful energy, one to generate heating steam at 25 psia and
the other to generate power through a cycle working between 250 psia,
700®F, and 1 psia, what fraction of the energy supplied would have served
a useful purpose?

Solution: Refer to Fig. 15-15.


(a) From the tables and mollier chart,

hi = 1371 Btu/lb

his = 1149

h= 118

Then, neglecting pump work,

Qi = 1253 Btu/lb

W = 156 Btu/lb

The thermal efficiency as a power plant m


W = 0.124
17 = ^
VI

The net energy supplied to the heating system is

Qii ^ h2 hi

Assuming the turbine adiabatic

hi = hi — W
and Qh = hi - W -hi = 1096 Btu/lb

Of this, 90 percent is usefully applied; then the useful energy is

W + 0.9Q^ = 1142 Btu/lb

The fraction of the energy supplied that serves a useful purpose is then

1142
0.91
1253

It is seen that this is greater than the individual efficiency of cither the
power plant or the heating plant.
(b) If two separate plants were used, the plant for heating steam would
be supplied with Qh or 1096 Btu/lb of heating steam. The power plant,
under the given conditions, would have a thermal efficiency of about 23
percent. Thus, to produce the work of 156 Btu as in part (a) would require
a heat input of 156/0.23 or 680 Btu. The total input per pound of heating
steam for both heating and power, would be
15-4] MORE EFFICIENT CYCLES 259

1096 680 - 1776 Btu

The useful energy, as in part (a), is 1142 Btu; then the fraction of the energy
supplied which appears as useful energy is

1142
0.65
1776

To obtain the same total output under method (b) it would be necessary to

bum 0.91/0.65 = 1.4 times as much fuel as under method (a); moreover
method (b) requires an additional boiler and a condensing plant which are
not needed under method (a).

This example shows clearly the basic thermodynamic advantage


of the by-product power scheme. Whether it is economical in any
particular application depends upon the cost of fuel, the cost of pur-
chased power, the existence of simultaneous demands for power and
process heat, and other factors which cannot be considered here.
15-4 Two-fluid or binary vapor cycles. It will have become
evident by now that the efficiency of a vapor cycle is limited by the
characteristics of the working substance. The complications of re-
generative cycles and reheat cycles could be avoided if it were possible
to specify at will the property relations of the working fluid and to
obtain the specified fluid. Some of the characteristics that would be
specified in a made-to-order fluid are listed below, with the reasons
therefor;
(a) A high critical temperature so that the heat of vaporization
could be supplied at the maximum temperature of the cycle, which
would be the maximum temperature for which apparatus could be
built.

(b) A large enthalpy of vaporization and a small specific heat of


the liquid so that relatively little of the heat supplied would be used
to raise the liquid temperature to the boiling point.
(c) Saturation pressures, at the maximum and minimum tem-
peratures of the cycle, within an order of magnitude of the atmos-
pheric pressure (in the range from 0.1 atm to 10 atm) so that stress
problems and leakage problems would be minimized.
(d) Freezing point below room temperature to facilitate draining

and filling the apparatus.


(e) Saturated vapor line close to the path of a turbine expansion
process (slightly increasing entropy with decreasing pressure) so that
excessive moisture would qot appear during the expansion.
260 MORE EFFICIENT CYCLES [16-4

(f) Chemical stability and inertness toward materials of construc-


tion at all attainable temperatures so that neither the fluid nor the
apparatus would deteriorate in use.
(g) The fluid should be non-toxic and low in cost.
The only fluid at present that is able to satisfy reasonably well
all the above conditions is it is somewhat deficient in
water, though
categories (a), (c), and The critical temperature of water (TOS^F)
(e).

is not as high as the maximum temperature of modern steam cycles


(1]00®F), and the saturation pressures, even at moderate tempera-
tures, are high (2200 psia at 650°F). A more suitable fluid at high
temperatures is mercury, which has a critical temperature above

2700'’F and a saturation pressure of only 180 psia at 1000®F. How-


ever, mercury is unsuitable at low temperatures because its saturation
pressure becomes exceedingly low and it would be impractical to
maintain in a condenser the low pressure ne^ed to condense mercury
at atmospheric temperatures.
These circumstances have led to the development of a binary
(two-fluid) cycle in which mercury is used in a Rankine type cycle
for the high-temperature range. The heat rejected at relatively high

Pij?. 15-10. Morcury-stpam plant flow diagram.


15- 4 ] MORE EFFICIENT CYCLES 261

temperature from the mercury cycle is used to evaporate water in a


regenerative steam cycle which operates to the usual low-temperature
limit for a steam cycle. The diagrams, Figures 15-16, 15-17 show
how this method is used, the cycle conditions being based upon the
operating conditions of the Schiller Station of the Public Service
Company of New Hampshire, which went into operation in 1950.

The mercury cycle, a-h-c-d, is a simple Rankiae type of cycle


using saturated vapor. Heat is supplied to the mercury in process
d-a, heating the liquid from 523°F to 94 1®F and evaporating it at
944®F, 128 psia. Most of this heat is received by the mercury at
the maximum temperature of 944°F. The mercury expands in a
turbine, process a-h, and is then condensed at about 2.58 psia and
523‘’F, process h-c. The feed pump process, c-d, completes the cycle.
The heat rejected by the mercury is utilized to boil water at 615
psia, 489®F. The water is supplied as approximately saturated liquid
to the condenser-boiler, and leaves as saturated vapor. From seven
to eight pounds of mercury must condense to vaporize one pound of
262 MORE EFFICIENT CYCLES [IM

water and the plant is so designed and operated that the proper ratio
of flow rates exists. The mercury condenser thus takes the place of
a boijer for the evaporation process of the steam cycle. However it
is necessary to heat the boiler feed water and to superheat the vapor,
by heat from some other source. The flue gases from the mercury
boiler furnace are used for this purpose. The steam cycle then con-
sists of the following processes: expansion in the turbine, 1-2; con-
densation at 0.75 psia, 2-3; pumping to 615 psia, 3-4; heating by
flue gas in the economizer, 4-5; evaporation by heat from condensing

mercury, 5-6; superheating by flue gas, 6-1. In an actual plant the


steam cycle is always a regenerative cycle, but this complication has
been omitted here in the interest of simplification.
It is evident from the T-s plot, Fig. 15-17, that the addition of
the mercury cycle to the steam cycle results in a marked increase in
the average temperature of heat supply to the plant as a whole and
consequently the efficiency is increased. For example, an actual
steam plant operating on a regenerative cycle with limiting conditions
as shown in Fig. 15-17 would have a heat rate of about 12,500
Btu/kwhr. The Schiller Station, using the binary cycle shown, and
having 40,000 kw rated capacity, has a heat consumption of less
than 9,500 Btu/kwhr. This economy has been equaled with a steam
plant only by using units of 150,000 kw rating, and operating on a
reheat cycle with 2000 psi initial pressure, 1050°F initial temperature,
and lOOO^F reheat temperature.
The concept of a binary fluid power plant is more than a century

old; early experimental work was done with steam as the high-
temperature fluid and sulfur dioxide as the low-temperature fluid.
The mercury-steam cycle has been in actual commercial use for a
quarter of a century, but it has never attained wide acceptance be-
cause there has always been the possibility of improving steam cycles
by and temperature, and by reheat. Since mer-
increasing pressure
cury is expensive, limited in supply, and highly toxic, unusual incen-
tives are necessary to overcome the natural reluctance to adopt it.
It now appears, however, that the possibilities for improvement of
the steam cycle are diminishing, whereas economic incentives for fuel
economy are increasing. For these reasons the mercury-steam cycle,
or gas-steam plants, may receive more favorable consideration in the
future than such plants have in the past.
MORE EFFICIENT CYCLES 263

PROBLEMS
15-1. An ideal steam power cycle has a turbine inlet steam state of 600 psia,
725^F; the steam expands reversibly and adiabatically in a turbine to 120 psia
after which it is reheated to 725^F, 120 psia, and expands reversibly and adiabat-
ically to the exhaust pressure of 1.5 in. Hg.bi. Find: (a) the steam rate, Ib/kwhr;
(b) the cycle efficiency.
15-2. For comparison with the reheat cycle of Problem 15-1, find the turbine
inlet pressure ofa Rankine cycle which has a turbine inlet temperature of 725^F,
an exhaust pressure of 1.5 in. Hgaba, and an exhaust moisture content of
12 percent (use the h-a chart). For this cycle find: (a) the steam rate, Ib/kwhr;
(b) the cycle efficiency.
15-3. In a plant to which the basic conditions of Problem 15-1 apply, the
turbine efficiency for each part of the expansion is 75 percent; there is a 10 psi
pressure drop in the reheater (steam comes out at 110 psia, 726®F); the feed
pump efficiency is 60 percent. Heat losses from the machines are negligible.

Find: (a) the steam rate, Ib/kwhr; (b) the plant cycle efficiency.
15-4. In a certain steam power plant the original installation operated be-
tween 400 psia, 750°F, and 1.5 in. Hgabi- In order to increase the power output
a new high-pressure boiler and turbine were installed; the new turbine receives
steam at 2000 psia, 950°F, and exhausts at 420 psia; the exhaust steam is reheated
and supplied to the old turbines at 400 psia, 750°F. (a) With a turbine efficiency
of 80 percent,how much power (kw) is delivered, per Ib/hr of steam, by the new
turbine? (b) If the old turbine has 75 percent efficiency how much power does
it deliver per Ib/hr of steam? (c) WTiat Is the total turbine output, new and
old, when the new turbine exhausts enough steam for the old turbine to produce
its rated output of 70,000 kw? (d) Under conditions in (c) what fraction of
the total heat supplied to the steam is transferred in the reheater? (e) Neglect-
ing pump work, compare the thermal efficiencies of the old and new cycles.
15-5. In a single-heater ideal regenerative steam cycle the steam enters the
turbine at 400 psia, 760‘*F, and the exhaust pressure is 2 psia. The feed water
heater is a direct-contact type which operates at 50 psia. Find: (a) the effi-

ciency and steam rate of the cycle; (b) the percentage improvement in efficiency,
compared to the Rankine cycle; (c) the percentage increase in steam rate, com-
pared to the Rankine cycle.
15-6. In a cycle operating under the conditions of Problem 15-5 the efficiency
of the portion of the turbine before the extraction opening is 65 percent, and of
the remaining portion is 70 percent. Both feed water pumps operate at 60 per-
cent efficiency. There are no external heat transfers except in the boiler and in
the condenser; there are no pressure losses due to friction in piping or apparatus.
Find the efficiency and the steam rate of cycle.
15-7. In an ideal cycle operating under the conditions of Problem 15-5 the
feed water heater a closed heater from which the drip is returned to the con-
is

denser. The is heated to the saturation temperature of the extracted


feed water
steam, and the drip leaves the heater as saturated liquid. Find the efficiency
and the steam rate of the cycle and compare with the results of Problem 15-5.
264 MORE EFFICIENT CYCLES
15-8. In a two-heater ideal regenerative steam cycle the steam and exhaust
conditions are the same as in Problem 15-5. The high-pressure heater is a closed
heater using oteam extracted at 120 psia; both the feed water and the drip leave
this heater at the saturation temperature for 120 psia. The low-pressure heater
is an open heater operating at 20 psia. The drip from the high-pressure heater
is passed through a trap to the low-pressure heater in the manner shown in

Fig. 15-11. (Note that in this problem the low-pressure heater differs from that
of Fig. 15-11.) Find the efficiency and the steam rate for this cycle, and compare
with the results of Problem 15-5.
15-9. An industrial plant requires 20,000 Ib/hr of heating steam at 25 psia,
and 1000 kw of power from a turbine. Consider an ideal cycle in which steam
goes tx) the turbine at 200 psia, 600°F, and the exhaust steam goes to the heating
system at 25 psia. (a) Find the state of the exhaust steam, (b) Will the ex-
haust steam flow be sufficient for the heating requirement?
15-10. Solve Problem 15-9 with a turbine efficiency of 70 percent.
15-11. For the. conditions of Problem 15-10 all boiler feed water is taken from
a lake at 60°F. Exhaust steam which is not used in the heating system is dis-

charged to the atmosphere; heating system condensate is thrown away. Com-


pare the heat required per hour to make steam in two cases: (a) operating as in
Problem 15-10; (b) using the same steam conditions and turbine efficiency as
in Problem 15-10, with 15 psia exhaust pressure, to generate power; and using
a separate boiler operating at 25 psia to supply saturated heating steam.
15-12. A power plant design is to be based upon the following conditions:
power output 2500 kw; heating steam, 60,000 Ib/hr at 30 psia, approximately
saturated vapor, is to be extracted from the turbine; the turbine inlet pressure
and temperature are respectively 175 psia and 500®F; the exhaust pressure is
2 psia; the turbine efficiency is 75 percent for each part (before and after the
extraction opening), (a) How much steam must be supplied to the turbine?

(b) The extracted steam is returned to the boiler as liquid at 200®F; the exhaust
steam as liquid at 1 20'’F. How much heat must be supplied per hour?
15-13. Solve Problem 15-12 if sufficient additional steam is extracted at
30 psia to heat all the feed water to 215°F in a direct-contact heater. What
percentage saving in heat supplied results from this change?
15-14. Taking the conditions shown in Fig. 15-17 for a simple mercury-steam
cycle, and assuming the efficiency of each turbine is 80 percent, find: (a) the
pounds of mercury circulated per pound of steam (the heat rejected by the con-
densing mercury must equal the latent heat of the evaporating water); (b) the
thermal efficiency of the mercury-steam cycle.
15-15. An inventor has proposed using Freon-12 as a working fluid for a vapor
power cycle. Discuss the probable advantages and disadvantages of this fluid,
considering efficiency, size and strength of equipment, and any other points on
which you can comment.

REFERENCES
See the references at the end of Chap. 14, p. 238.
Chapter 16

GAS CYCLES

As pointed out in the introduction to Chap. 14, no cyclic heat engines

using gas for working fluid have yet attained great importance in
practice. The study of gas cycles is important, however, for two
reasons; (1) considerable insight into the characteristics of non-cyclic
gas power plants of the greatest practical importance (internal com-
bustion engines and turbines) can be obtained from the study of gas
cycles; (2) as a result of improved materials and designs the prospects
for the practical use of cyclic gas power plants seem better now than
at any time in the past.
16-1 The Stirling and Ericsson cycles. As in the case of vapor
cycles it is reasonable to inquire into the possibilities of the Carnot
cycle for use with gases. It is found that the large back work of the
Carnot cycle is a disadvantage in this case as in the case of vapor
(Sec. 14-8); consequently gas cycles have been devised to reduce the
back work while retaining the efficiency of the Carnot cycle. An
example of such a cycle is the Stirling cycle shown in Fig. 16-1.
The Stirling cycle consists of the following processes, with a sys-
tem composed of a perfect gas (air): (a-fe) heat supplied and work
transferred out at constant temperature T\, (b-c) temperature re-
duced at constant volume by heat transfer out; (c-d) heat rejected

and work transferred in at constant temperature Tj; (d-a) tempera-


ture raised at constant volume by heat transfer in.

In its basic form the Stirling cycle is less efficient than the Carnot
cycle because of the heat transfers in the constant-volume processes.
However, if a regenerative arrangement is used so that the heat trans-
ferred out during the process b-c is returned reversibly during the
process d-a, there no external heat transfer except at Ti and Tj.
is

Then the Stirling cycle becomes equivalent to the Carnot cycle shown
dotted in Fig. 16-1. The reasoning here is entirely analogous to that

265
266 GAS CYCLES [16-1

Fig. 16-1. Stirling cycle.

in the case of the regenerative vapor cycle using saturated vapor.


(Sec. 15-2). The mechanism by which the regeneration is accom-
plished is usually an assembly of metal plates or wires throu^ which
the gas passes during processes b-c and d-a. In process b-c some
portion of the metal becomes heated to each temperature between
7*1 and Tt; during process c-d the energy is stored in the metal; dur-
ing process d-a the metal transfers heat back to the gas. In the ideal
case these operations are reversible.
The Ericsson cycle differs from the Stirling cycle only in using
constant-pressure processes in place of the constant-volume processes.
16-2] GAS CYCLES 267

The regenerative Stirling and Ericsson cycles have the same effi-

ciency as the Carnot cycle, but much less back work; it has
therefore
been possible to operate successfully hotr-air engines of these types.
It was found, however, that the transfer of heat to the gas raised
two serious obstacles: (1) it was difficult to transfer heat to a gas at
high rates, so the engine speeds were limited, and consequently the
power output was small for a given size of engine; (2) the metal
through which the heat was transferred deteriorated rapidly since its
temperature had to be much higher than the maximum gas tempera-
ture. If the metal temperature were kept low, not only would
the efficiency be reduced but the rate of beat transfer would also
be reduced, thereby reducing the capacity (power, output) as well.
Improvements in heat-resisting metals and in the design of heat ex-
changers have in recent years made possible greatly improved hot-air
engines. Since hot-air engines can use cheap fuels, they may even-
tually find a field of use where the cost of internal combustion engine
fuels is excessive.
16-2 Air-standard cycles. Historically, the obstacles to the ef-
fective use of the hot-air engine were overcome by producing high
temperature through a combustion process directly in the working
fluid (internal combustion). In this way the gas can reach a high
temperature while the metal of the cylinder can be cooled by external
air or water circulation in order to avoid rapid deterioration. More-
over, it is easy to raise the gas temperature very rapidly so that the
engine can operate at high speed and have a large power output
capacity for its size.

When the internal combustion process is examined, it is seen to


differfrom the process of heat supply in a heat engine cycle because
there is a permanent change in the working fluid during combustion.
Therefore the fluid does not pass through a cycle and the internal
combustion engine is not a thermodynamically cyclic heat engine.
There is, in fact, no heat supplied to the engine, so it cannot be a
heat engine in the sense of Sec. 7-2. Nevertheless a heat engine
cycle can be constructed to correspond approximately to the opera-
tion of an internal combustion engine by substitution of heat transfer
processes for some of the actual engine processes. Such cycles, called
air~standard cycles, are useful in the elementary study of internal
combustion engines. More detailed study requires analysis of com-
bustion processes.
268 GAS CYCLES [IM

(b)

Fig. 16-2. (a) Spark ignition engine, (b) Indicator diagram.

16-3 The Otto cycle (constant- volume cycle). One of the


most commonly encountered types of internal combustion engine,
used in automobiles and aircraft, is the spark-ignition gasoline engine,
for which the air-standard cycle is the Otto cycle. The sequence of
processes in the elementary operation of such an engine is given be-
low, with reference to the sketch and indicator diagram. Fig. 16-2.
Process 1-2, intake. With the inlet valve open the piston moves
to increase the cylinder volume, taking in fuel-air mixture ft constant
pressure.
1!V-3J GAS CYCLES 269

Process 2-3, compression. With the valves closed the piston re-
turns, compressing the combustible mixture to the minimum cylinder
volume.
Process 3-4, cornbvstion. At the minimum cylinder volume a
spark ignites the mixture, which burns with resulting increase in
temperature and pressure.
Process 4-5, expansion. The moves out to maximum cyl-
piston
inder volume, the average pressure being greater than in the com-
pression process.

S
l''ig. 10-3. Otto cycle.
'

270 GAS CYCLES [lft4

Process 5-6, blow-doum. At the maxiiniun cylinder volume the


exhaust valve opens and fluid leaves the cylinder until the pressure
drops to the initial pressure, pi.

Process 6-1, exhaust. With the exhaust valve open, the piston
moves to expel the combustion products from the cylinder at constant
pressure.
While the series of processes described above constitutes a me-
chanical cycle, not a thermodynamic cycle. The designation
it is

four-stroke cyde, commonly applied to such a sequence, refers to the


mechanical action.
Figure 16-3 shows the air-standard cycle corresponding to the
engine of Fig. 16-2 (the Otto cycle). It consists of the following

processes, carried out with a perfect gas system composed of air:

Process a-b, heat supplied at constant volume, reversibly.


Process b-c, reversible adiabatic expansion.
Process c-d, heat rejected at constant volume, reversibly.
Process d-a, reversible adiabatic compression.
Heat transfer processes have been substituted for the combustion
and blow-down processes of the engine; the intake and exhaust proc-
esses of the engine are not needed in the cycle, for they would merely
cancel each other. Note the distinction between the indicator dia-
gram of Fig. 16-2 and the state diagram of Fig. 16-3; the former refers
to pressure and volume in a cylinder containing varying masses of
gas, while the latter refers to the properties of a fixed mass of air.
The efficiency analysis of the air-standard Otto cycle for a system
of mass m follows.
Heat supplied Qi = Qa-h = mcviTb — Ta) (1)

Heat rejected Q? = —Qc-d = rnc^{Tc — Tf) ( 2)

Wnet _ 1 _ ^_ 1 _ inCviT, - Td)


Efficiency
Qi Qi mc,{Tb-Ta)

= 1 - n - Ta
(3)
Tb- Ta

This can be simplified still further by the relations among the tem-
peratures. Since d-a and b-c are reversible adiabatic processes it

follows that
^Sa—b —
164] GAS CYCLES 271

Tj ^0
or men m^ =
,
Tnc, In
1

^
or li L
Ta Ti

Then

or, from (3), (4)

Since d-ii is a reversible adiabatic process, (4) can be rewritten

i; = 1 - (5)

For the cycle of Fig. 16-3 or the engine of Fig. 16-2 the compres-
sion ratio n is defined as the ratio of the volume at the beginning of
compression to the volume at the end of compression;

( 6)

Then (7)

Equation (7) shows that efficiency increases with increasing com-


pression ratio. The equation applies strictly only to an ideal cycle
using a perfect gas, but it serves to indicate within limits the trend
of efficiency variation in actual engines. As an example, Fig. 16-4
shows the relation between air-standard efficiency and the brake
thermal efficiency of a certain engine when the compression ratio was
changed, other conditions being constant. Brake thermal efficiency
is the ratio of the work output of the engine to the heating value of
it bums.
the fuel
A
somewhat closer approach to actual engine conditions may be
made by taking account of the variation of the specific heat of air at
high temperature, using the variable specific-heat Otto cycle. In this
case the heat transferred in a reversible constant-volume process is

given by the change of internal energy, which is most conveniently


obtained from the Gas Tables; then referring to Fig. 16-3,

- Qi - Uc — Uj
n- 1 = 1 (8 )
Qi ut — u,
272 GAS CYCLES [154

Fig. 164. Otto cycle efficiency.

In the case of the air-standard cycle the efficiency is independent of


the temperature level at which the cycle operates, and of the rate of
working, but this is not true for the variable specific-heat cycle. In
the latter case efficiency is a function of compression ratio, tempera-
ture at beginning of compression, and heat supplied per unit mass of
working fluid.

In general, as pointed out above, the efficiency of an actual engine


working on the Otto cycle increases with increased compression ratio;
however, there is a limit to the compression ratio that can be used

in any particular case in practice. In a given engine under given


operating conditions, the compression ratio may not be increased
beyond a certain limit without encountering detonation, a noisy and
destructive combustion phenomenon (“knocking”). The avoidance
of detonation requires different values of compression ratio depend-
ing upon the fuel, the engine design, and the operating conditions,
but it fixes some limit on the efficiency of all ordinary gasoline engines.
16-4 The Diesel cycle (constant-pressure cycle). One method
of avoiding the limitation on compression ratio in the spark-ignition
engine is by compressing air alone, and injecting the fuel into the
cylinder only when combustion is desired. By this method the air
can be raised to a temperature higher than the ignition temperature
of the fuel so that no spark is necessary to start combustion. The
16-4 ] GAS CYCLES 273

Fig. 16-5. Diesel cycle.

fuel can begin to bum spontaneously, soon after it is sprayed into


the hot air, and the rate of burning can be controlled to some extent

by the rate of injection of the fuel. An engine operating in this way


is called a compression-ignition engine.

The classical air-standard cycle for the compression-ignition engine


is the Diesel cycle. Fig. 16-5. In this cycle the processes are:
Process a~b, heat supplied at constant pressure, reversibly.
Process b-c, reversible adiabatic expansion.
Process e-d, heat rejected at constant volume, reversibly.
274 GAS CYCLES [IW

Process d-a, reversible adiabatic compression.


The efficiency analysis of the cycle for a system of mass m follows.
Heat supplied Q\ = = mc,(Tk — Tb)

Heat rejected Qt =» —Qe-t = mcv(Te — Ti)

1^ net _ Tj)
__ < S?— T
Efficiency
Qi
" Qi~ mc,(n-r.)
Tc-Ti
(9 )
Kn - Ta)
The efficiency of the air-standard Diesel cycle may be expressed
in terms of volume ratios but not so simply as in the case of the Otto
cycle. The efficiency of the Diesel cycle depends upon the amount
of heat supplied as well as upon the compression ratio, for as the
T-s diagram shows, the ratio of heat supplied to heat rejected de-

creases as the line b-c moves toward greater entropy. Thus the
efficiency may be expressed in terms of any two of the three ratios
below.

Compression ratio, r*, = ^


Expansion ratio, r..

Cut-off ratio, Tc,

Probably the most enlightening form of efficiency expression is in


terms of n and rc, as follows:

_-L -
1
( 10 )
Hr. 1)

This equation gives a family of curves on the plane of efficiency


vs. compression The curve for unity cut-off ratio
ratio, Fig. 16-6.

(zero heat supply) is the same as the Otto cycle curve, Fig. 16-4;
with increasing cut-off ratio the curves become lower. Thus, for the
same compresnon ratio at zero load, the Diesel cycle is as efficient
as the Otto cycle, but it is less efficient at normal loads. However,
engines using the Diesel cycle usually operate at higher compression
ratios than engines using the Otto cycle, so the resulting efficiencies
16-6 ] GAS CYCLES 275

Fig. 16- 6 . Dieael cycle eflSciency.

under operating conditions are not necessarily much different for the
two types.
The variable specific-heat Diesel cycle efficiency may be seen by
inspection of the cycle diagram to be

( 11 )

16-5The mixed cycle (Umited pressure cycle). The operation


of modem compression-ignition engines is better represented by a
mixed cycle than by the Diesel cycle. In the mixed cycle the heat
is supplied partly in a constant-volume process and partly in a
constant-pressure process. In the actual engine, fuel injection is

started before the end of the compression stroke so that a portion of


the combustion is completed while the piston is reversing its motion
and not moving rapidly. This causes a pressure rise and combustion
is completed at approximately constant pressure by the continuing
injection of fuel during the first part of the outward piston travel.
The air-standard mixed cycle is shown in Fig. 16-7.
The efficiency of the air-standard mixed cycle is given by

_ ^ ,
mCv(Ti - Te)
’ " mc.(7’» - n) + mcATc - Ti)

_ ( 12 )

“ 1
- n)
^
Tfc - n Wc -I-
276 GAS CYCLES [ 1&4

Fig. 16-7. Mixed cycle.

By substitutions as in the preceding sections

rA - 1

1 + krAe - 1)
(13)

where r* and r* are V,/Va and Ve/Vi, respectively, and r, is the


constant-volume pressure ratio pt/po. For any given compression
ratio the efficiency of the mixed cycle is intermediate between that
of the Otto cycle and that of the Diesel cycle, depending upon the
ratio of the heat supplied at constant volume to that supplied at
constant pressure.
IM] GAS CYCLES 277

Fig. l&S. Brayton cycle.

16-6 The Brayton cycle. The Brayton cycle is an air-standard


cycle composed of two reversible constant-pressure processes and
two reversible adiabatic processes as shown in Fig. 16-8. A recipro-
cating engine operating on this cycle and patented in the 1870’s by
George B. Brayton, was the first successful gas engine built in the
United States.
The efficiency of the Brayton cycle is given by

, _ , mc.(r, - Td)
e, mcp(n-r.)
278 GAS CYCLES [16-6

By the same reasoning as for the Otto cycle, Sec. 16-3, this reduces to

1
T, , ,
fVaV-^ - 1 ,,,,

where rt is the adiabatic compression ratio. Thus the efficiency of


the Brayton cycle is the same as that of the Otto cycle for the same
adiabatic compression ratio.
The Brayton same adia-
cycle differs from the Otto cycle of the
batic compression ratioand work output, in having a larger range of
volume and smaller ranges of pressure and temperature. The larger
volume of low-pressure gas cannot be handled efficiently in a recipro-
cating engine because of increased friction losses in the mechanism
and in the exhaust process; also the Brayton engine is more bulky
than the Otto engine for the same power capacity. Therefore the
Brayton cycle cannot compete with the Otto cycle in the reciprocat-
ing engine field. For turbine machinery, however, the Brayton cycle
is much better adapted than the Otto cycle because turbine machinery

can handle large volumes of gas more efficiently than small volumes.
Moreover, in steady flow machinery it is much easier to carry out
heat transfer or combustion processes at constant pressure than at
constant volume; also the turbine is not adapted to as high maximum
gas temperatures and pressures as is the reciprocating engine. For
these reasons the Brayton cycle is the basic air-standard cycle for all

modem gas turbine plants, whether of the truly cyclic type or of the
internal combustion type.

Fig. 16-9. Gas-turbine plant flow diagram.


16-6] GAS CYCLES 279

A flow diagram for a cyclic gas turbine plant using the Brayton
cycle is shown in Fig. 16-9; this arrangement is popularly called a
“closed cycle” gas turbine plant. In Fig. 16-9 the points a, b, c,

and d correspond to the identically marked points in the state dia-


grams, Fig. 16-8. The compressor work Wc is taken from the turbine
directly to the compressor, leaving the net work available for use. W
By the steady flow analysis

— Kh ba

Wt = hb — he

Qz ““ he hd

Wc = ha- ha

Waet = Wt-Wc
__ net hb hr hg -j- ha
~ Ql hb- ha

(15)
^
hb - ha

The enthalpies may be obtained from the gas tables or by use of the
perfect gas rules.

Example 1. Find the efficiency, and the work per pound of fluid circu-
lated, for a Brayton cycle working between pressures of 15 psia and 75 psia,
if the minimum temperature in the cycle is 550°R and the maximum tem-
perature is 1700°R.

(a) Gas law solution.

From Fig. 16-8 the minimum temperature occurs at d and the maximum
at b. The temperature ratios for the reversible adiabatic processes are
given by

Ts = li =
Ta \paj Te \pe)

Pa ^ Pt — 75 _ g
Pa p, 15

k 1.4

1.58
Ta Te

Ta = (550)C1.58) - 870'’R
280 GAS CYCLES [164

r. - 7^ = 1076’R
l.Oo

Q, - CfiTi - T.) - (0.24)(1700 - 870) - 199 Btu/lb

Qi - CpCr. - Ti) = (0.24)(1075 - 550) - 126 Btu/lb

- Qi - <3i - 73 Btu/lb

, . 0.367
VI

Also 1 - f!
“ ^
^ ^

(b) Gaa Table solution.


In the tables, for reversible adiabatic processes,

"Bs — ‘Eh Ei ^E2*


Pd Prd Pe Pn
At 560®R, Prd = 1.4779; hi = 131.46 Btu/lb.

At nOO^R, prh =‘ 90.95; h, = 422.59 Btu/lb, Table 1, Keenan and Kaye.


Then p„ = 5(1.478) - 7.390; T. = 868“R; K = 208.41.

p„ = 90.95/5 = 18.19; T* •= lllS^R; K= 269.27.

Qi=^h-ha = 214.18 Btu/lb

Qi = A. - Ad = 137.81 Btu/lb

W„t = Qj - Qi = 76.37 Btu/lb

, - = 0.357
Qi

The efficiency of the Brayton cycle may sometimes be increased


by the use of a regenerator or heai exchanger to transfer heat from the
turbine exhaust gas to the gas leaving the compressor, thereby reduc-
ing the heat supplied from an external source. Such a cycle is illus-

trated in Fig. 16-10.In this case the gas leaving the turbine at c
has higher temperature than the gas leaving the compressor at a.
Consequently if the two streams are passed through a heat exchanger,
the temperature of the gas leaving the compressor can be raised by
heat transfer from the turbine exhaust gas. The maximum tempera-
ture to which the cold gas can be heated is the temperature of the
hot gas'entering the exchanger; this could be done only with an infi-

nitely large heat exchanger operating reversibly. In a real case the


temperature at o'. Fig. 16-10, cannot reach the temperature at c]
16-6 ] GAS CYCLES 281

EXHAUST HEAT EXCHANGER OR REGENERATOR

Fig. 16-10. Gas-turbine plant with regenerator,

the fraction of the maximum possible temperature rise actually ob-


tained is called the effectiveness (sometimes the efficiency) of the regen-
erator. For the case illustrated,

Effectiveness =7 ^ (16)
Ic lo

As indicated in Fig. 16-10, when the regenerator is used in the


idealized cycle, the heat supplied and the heat rejected are each
282 GAS CYCLES [16^

reduced by the same amount Q* hence the work of the cycle is


unchanged but the efficiency is increased. The heat and work quan-
tities are as follows:

Qi ~ hb hn

Q2 ~ K — hd

Wt — hb — he

Wc = ha- hd

Waet = Wt -Wc = Ql-Qi


The effect of the regenerator upon the cycle is best shown by an
example.

Example 2. Repeat Example 1 for a cycle using a regenerator of 75


percent effectiveness.

Solution: The states at a, b, c, and d, Fig. 16-10, will be the same as in


Example 1. The state at a is determined by Eq. (16).

= Ti - 870
0.75
1075 - 870

T' = 1024^
Te is determined by the energy balance for the regenerator.

Cp(T' - Ta) = - T'e)


1024 - 870 = 1075 - T'e
T'e
= 92rR
Qi = CriTb - T'a) = 162 Btu/lb

= c,{T'e - Ti) = 89 Btu/lb

= 73 Btu/lb

„ = = 0.45

Compare this with 0.367 for the same cycle without the regenerator.

In practice the regenerator is costly, heavy; and bulky, and causes


pressure losses which reduce the efficiency of the cycle. These factors
have to be balanced against the ideal efficiency gain to decide whether
it is worthwhile to use the regenerator.
If the turbine and compressor of the Brayton cycle are real
.

16-6] GAS CYCLES 283

(irreversible)machines the states a and c will have to be located by


use of the machine efficiencies together with the isentropic end states

a, and c„ Fig. 16-11. In the simple Brayton cycle the use of real
machines will reduce the heat supplied by the amount ha — hat, and

Fig. 16-11. Brayton cycle with real machines.

will increase the heat rejected by K— Thus the net work of


ha.
the cycle will be reduced by the sum, ha + he — he,. The ef-
— hat
fects of using machines of 80 percent efficiency in Examples 1 and 2,
above, are tabulated below, using air table solutions.

Examine 1 Example 2
Machine Efficiency 1.00 0.80 1.00 0.80
Wt 153 123 153 123
Wc 77 96 77 96
W„t 76 27 76 27
Qi 214 195 169 142
V 0.36 0.13 0.45 0.19

Example 1 had no regenerator, Example 2, a regenerator of 75 percent effec-


tiveness.

The evident sensitivity of the cycle efficiency to machine efficiency


occurs because of the large back work or compressor work, as ex-
plained in Sec. 14-8. Because of its large back work, the gas turbine
plant must use machines having a total power rating of five to ten
times the net cycle power rating, and the efficiency of these machines
must be very high to obtain a satisfactory plant efficiency. In fact
the cycle efficiency may go to zero with machine efficiencies of 60 to
284 GAS CYCLES [16-7

70 percent for cycle conditions that give good reversible efficiencies.


With a vapor cycle a machine efficiency of 60 to 70 percent would
give a cycle efficiency of almost 60 to 70 percent of the reversible
cycle efficiency. These facts, and the problem of temperature, dis-
cussed in the following section, show why the gas turbine plant has
been so far behind the steam plant in its development.
16>7 Efficiency and capacity characteristics of the Brayton
cycle. Although the efficiency of the Bra3d;on cycle may be expressed
in terms of pressure ratio, Eq. (14), there are other influences that
require consideration if the characteristics of the cycle in its practical

application are to be understood. The following discussion refers


only to design characteristics.
Referring to Fig. 16-11 if Td, the lowest temperature of the cycle,
is the temperature of the surroundings, the maximum efficiency at-
tainable will be limited by temperature of the cycle.
Tt, the highest
In a turbine plant Tb is limited by the characteristics of metals avail-
able for burner and turbine construction. In aircraft turbines of
short life the maximum temperature may
be of the order of 18(X)®F,
while for long-life power plants the limit is currently from 1400® to
1500°F. These limits may be extended in the future by improved
metals or methods of cooling, but there will always be a definite limit
for any particular combination of design, materials, and life expect-
ancy. The reason turbine plants are much more sharply limited in
this respect than reciprocating engines is that the hot parts of the
steady flow machine are constantly exposed to hot gases, with no
intervals of contact with a cool gas.
Assuming that the temperature limits for an ideal cycle are fixed
as shown in Fig. 16-12, the cycle shape will change as the pressure
ratio changes. Cycles of low pressure ratio {a-b-c-d) have low effi-
ciency, since the average temperature of heat supply is little greater
than the average temperature of heat rejection; moreover the work
capacity is small, as shown by the area enclosed in the diagram. At
the lower limit of pressure ratio (unity) both work and efficiency
become zero.
As the pressure ratio is increased, the efficiency steadily increases
because the average temperature of heat supply steadily approaches
Toms whilc the average temperature of heat rejection steadily ap-
proaches Tain. In the limit, when the compression process ends at
7'max) the Carnot efficiency is reached. The work capacity, however,
16- 7] GAS CYCLES 285

Fig. 16-12. BraytoD cycles of various pressure ratios.

passes through a maximum as the pressure ratio increases, and when


the Carnot efficiency is reached the work output is zero. This can
happen because the heat supplied approaches zero simultaneously
with the work. Considering cycles and Fig.
16-12, it is obvious that to obtain reasonable work capacity some
reduction of efficiency must be accepted.
If the cycle uses real (irreversible) machines, the efficiency will
not increase continuously to the maximum pressure ratio, because
some energy is used to supply machine losses. The net work output

Fig. 16-13. Effect of pressure ratio on efficiency; simple Bray ton cycle.

K*
286 GAS CYCLES

S
Fig. 16-14.

willthen go to zero before the heat supplied reaches zero. Figure


16-13 shows how the efficiency will vary with pressure ratio at fixed
maximum temperature for the cases of reversible machines and irre-

versible machines. In the latter case, operation at maximum effi-

ciency will not necessarily involve much loss of capacity.

When an exhaust-gas heat exchanger, or regenerator, is used, as

explained in Sec. 16-6, the efficiency of the ideal cycle may be im-
proved, while the work capacity is unaffected. As shown in Fig.
16-14, at low pressure ratios,
if a reversible heat exchanger
is used, most of the heat re-
jected in the simple cycle can
be recovered. Heat can be
transferred until the gas ap-
proaching the turbine reaches
the turbine exhaust tempera-
ture Te. Thus the heat sup-
plied from an external source
is only that for the process
a'-b\ the average temperature
Fig. 16-15. of heat supply therefore ap-
proaches Too. as the pressure
ratio approaches unity. Similarly, the average temperature of heat
rejection approaches Tain- Thus the efficiency of the cycle ap-

proaches the Carnot efficiency as the pressure ratio approaches unity.


16-7] GAS CYCLES 287

Fig. 16-16. Effect of pressure ratio on efficiency; regenerative cycle with


reversible machines.

When the pressure ratio is increased, the possible amount of heat


recoveiy decreases because Ta and 7** approach each other; when
Ta - Te, cycle a-b-c-d, Fig. 16-15, no recovery is possible. If the
regenerator is used with Ta greater than Tc, cycle a'-b'-c'-d, Fig.
16-15, heat transfer will be reversed, wasting energy from the air
supply directly to the exhaust; therefore the regenerator would not
be used with such high pressure ratios. The characteristic of the

Fig. 16-17. Effect of pressure ratio on efiicicDcy; irreversible cycles.


288 GAS CYCLES [IM

cycle with reversible regenerator is shown in Fig. 16-16, together with

the simple cycle curve.


If a real (irreversible) exchanger is used with an otherwise revers-
ible cycle, the temperature will always be less than Te] therefore
when the pressure ratio approaches unity, 7^ will approach some value
less than T —
and some heat will be supplied even when the work
,

goes to zero. Then the efficiency must go to zero at unity pressure


ratio as shown in Fig. 16-16. If a real exchanger is used with real

machines the efficiency curve will be as indicated in Fig. 16-17.


From the plots shown it is apparent that for best efficiency the
choices of pressure ratio and of heat-exchanger effectiveness are

closely interrelated. For a plant which cannot accept a bulky, heavy


exchanger (an aircraft plant, for example) it may be necessary to go
to high pressure ratios to obtain high efficiency with a simple cycle.
For a stationary plant it may be better to use low pressure ratios and
a large heat exchanger; here, however, economics will usually deter-

mine the final choice.

More complex gas turbine cycles. The efficiency or the


16-8
work capacity of a gas turbine plant may often be increased by using
stored compression with intercooling, or by using staged heat supply,
called reheat. The first scheme is illustrated in Fig. 16-18. It is
clear that the staged compression with intercooling increases the work
of the cycle by the area 2-3-4-2', but the additional work involves
a corresponding additional heat supply from 4 to 2'. Since this heat
is supplied at a lower temperature than any heat supplied in the basic
cycle, the intercooling results in a reduced efficiency for the cycle.
16-8] GAS CYCLES 289

However, when a regenerator is used, intercooling can give a net gain


in efficiency of the reversible cycle because the additional low-
temperature heat supply comes from further cooling of the exhaust
gas rather than from an external source. With a real compressor
intercooling may so reduce the back work as to result in improved
cycle efficiency even without a regenerator.*
A cycle using reheat is shown in Fig. 16-19. Here it can be seen
that there is a gain in work,
area 4-5-6-4', which is ob-
T
tained from high-temperature
heat; however, the small Bray-
ton cycle 4-5-6-4' added to
the basic cycle has a smaller
pressure ratio than the basic
cycle, and therefore a smaller
efficiency. Thus as in the
case of intercooling, the simple
cycle efficiency is reduced by
reheating. With a regenerator
the cycle efficiency may be
Fig. 16-20. Approximation to the
improved by reheating, because
Ericsson cycle by multistage compression
the higher temperature ex- with intercooling, and reheating.
haust permits greater heat
recovery and smaller external heat supply for the basic part of the
cycle. With machines the increased work may give increased
real
efficiency through reduction in the relative back work.
* 8ee Chap. 21 for details on gas compression with intercooling.
290 GAS CYCLES

It is possible to combine in one cycle regeneration, several stages


of compression, and several stages of reheat. If this is done the cycle.
Fig. 16-20, will approach as a limit the Ericsson cycle, with heat
supply and rejection at constant temperature.
Although the simple flow diagrams presented here show only a
single shaft with all compressors and turbines on that shaft, it is often
desirable for operating convenienceand part-load efficiency to divide
the machinery between two For example, there may be a
shafts.
turbine driving the compressor on one shaft, and a separate turbine
driving the load on another shaft. The detailed study of such ar-
rangements belongs in the field of plant design rather than in that of
thermodynamics.
The gas cycles discussed in this chapter are only approximations
of the real internal combustion engines which most of them represent.
Internal combustion engines and turbine power plants are discussed
in some detail in Chap. 23.

PROBLEMS
16-1. In a Stirling cycle thevolume varies between 1 and 2 cu ft, the maxi-
mum pressure is 30 and the temperature varies between 1000°F and 500‘^F.
psia,
The working fluid is air (a perfect gas), (a) Find the efficiency and the work
per cycle for the simple cycle, (b) Find the efficiency and the work per cycle
for the cycle with an ideal regenerator, and compare with the Carnot cycle having
the same isothermal heat supply process and the same temperature range.
16-2. Plot the efficiency of the air-standard Otto cycle as a function of com-
pression ratio for compression ratios from 4 to 16.
16-3. Find the efficiency of the variable specific-heat Otto cycle for compres-
sion ratio of 8, using air tables, and taking the maximum and minimum tempera-
tures as 3500®R and 600°R, respectively. Compare with the results of Problem
16-2.
16-4. Tabulate the efficiencies at a compression ratio of 6 for Otto cycles
using perfect gases having specific heat ratios of 1.3, 1.4, 1.67. What would be
the advantages and disadvantages of helium as a working fluid for an Otto cycle?
16-5. In an air-standard Otto cycle the compression ratio is 7, and compres-
sion begins at 100°F, 14 psia; the maximum temperature of the cycle is 2000°F.
Find: (a) the heat supplied per pound of air; (b) the work done per pound of air;
(c) the cycle efficiency; (d) the temperature at the end of the isen tropic expansion;
(e) the maximum pressure of the cycle.
16-6. Solve Problem 16-5 if the maximum temperature is 3000®F.
16-7. Solve Problem 16-5, using air tables.
16-8. For an air-standard Diesel cycle with compression ratio of 12 plot
the efficiency as a function of cut-off ratio for cut-off ratios from 1 to 4. Com-
pare with the results of Problem 16-2.
GAS CYCLES 291

16-9. In an air-standard Diesel cycle the compression ratio is 13, and com-
pression begins at 15 psia, 140®F; the maximum temperature of the cycle is

2540®F. Find: (a) the heat supplied per pound of air; (b) the work done per
pound of air; (c) the cycle efficiency; (d) the temperature at the end of the
isentropic expansion; (e) the cut-off ratio; (f) the maximum pressure of the cycle.
16-10. Solve Problem 16-9, using air tables.
16-11. Derive Eq. (16-13) from Eq. (16-12).
16-12. An air-standard mixed, or limited-pressure, cycle has a compression
ratio of 10, and compression begins at 15 psia, 140°F; the maximum pressure of
the cycle is 600 psia, and the maximum temperature is 2540°F. Find: (a) the
heat supplied at constant volume, per pound of air; (b) the heat supplied at
constant pressure, per pound of air; (c) the work done per pound of air; (d) the
cycle efficiency; (e) the temperature at the end of the constant- volume heating
process; (f) the cutroff ratio.
16-13. Solve Problem 16-5 for an air-standard Brayton cycle having the same
adiabatic compression process as the Otto cycle of Problem 16-5. Compare the
maximum specific volumes and the maximum pressures for the Brayton and Otto
cycles having the same compression process and the same maximum temperature.
16-14. In the Brayton cycle of Problem 16-13 it is proposed to use a regener-
ator to increase cycle efficiency. Investigate this proposal and give an opinion
of it^ value; show your reasoning.
16-15. A gas turbine plant operates on the Brayton air cycle between a mini-
mum temperature of 40°F and a maximum temperature of 1540®F. (a) Find the
pressure ratio at which the cycle efficiency equals the Carnot efficiency, (b) Find
the pressure ratio at which the work per pound of air is maximum. (Take your
choice of an algebraic solution or a plot of computed values of work at various
pressure ratios.) (c) Compare the efficiency at this pressure ratio with the
Carnot efficiency for the given temperatures.
16-16. Theplant of Problem 16-15 is changed to have compressor efficiency
of 80 percent and turbine efficiency of 85 percent. Make a plot of efficiency,
and work per pound of air, vs. pressure ratio, for pressure ratios from 5 to 10.
Find: (a) the pressure ratio at which the work per pound of air is maximum;
(b) the efficiency and the work per pound of air at this pressure ratio; (c) the
pressure ratio at which the efficiency is maximum; (d) the efficiency and the
work per pound of air at this pressure ratio. Compare results with Problem
16-15.
16-17. In a gas turbine plant working on the Brayton air cycle the air pressure
and temperature before compression are respectively 15 psia and 80®F. The
ratio of maximum pressure to minimum pressure is 6.25, and the temperature
before expansion in the turbine is 1440®F. The turbine and compressor effi-
ciencies are each 80 percent. Find: (a) the compressor shaft work per pound of
air; (b) the turbine shaft work per pound of air; (c) the heat supplied per pound
of air; (d) the cycle efficiency; (e) the turbine exhaust temperature.
16-18. Solve Problem 16-17 if a regenerator of 75 percent effectiveness is
added to the plant.
16-19. Solve Problem 16-17 if the compression is divided into two steps,
each of pressure ratio 2.5 and efficiency 80 percent, with intercooling to 80®F.
292 GAS CYCLES

16-20. Soke Problem 16-19 if a regenerator of 75 percent effectiveness is

added to the plant.


16-21. Solve Problem 16-17 if a reheat cycle is used; the turbine expansion
is divided into two steps, each of pressure ratio 2.5 and efficiency 80 percent,
with reheat to 1440°F.
16-22. Solve Problem 16-21 if a regenerator of 75 percent effectiveness is

added to the plant.


16-23. Solve Problem 16-22 if the staged compression of Problem 16-19 is

used in the plant.


Chapter 17

FLUID FLOW-NOZZLES AND ORIFICES

In Chap. 6 equations were developed for applying the First Law to


processes involving fluids in motion. In this chapter these equations
will be applied to certain important processes in which flow of the
fluid is the principal factor. Such processes include the production
of fluid jets to drive turbine wheels and the use of flow nozzles and
orifices in fluid metering. The subject matter of this chapter is also
treated in books on Fluid Mechanics, being at least as much a part
of that field as of thermodynamics.

17-1 Nozzles —^mass and energy relations. A nozzle is a duct


or passage in which the kinetic energy of a flowing fluid increases as
a result of a drop in pressure along the stream. The following analysis
of nozzles is based upon the assumption of a steady flow process with
uniform properties across the stream at every cross-section (see Sec.
6-3). It is further assumed that potential energy change and heat
transfer are negligible, and it is noted that there is no shaft work
involved. Then forany such nozzle, Fig. 17-1, having inlet section 1
and outlet section 2 the steady flow equations for mass and energy
will be
293
294 FLUID FLOW [17-1

OiVi _
(1)
1>, Vi

-fc ( 2)

It is convenient to introduce here the concept of stagnation enthalpy

which is the enthalpy that would be reached if the stream were


brought to rest under the conditions assumed above. Using the sub-
script zero for stagnation conditions

Fo = 0

hence hQ = hi —=h
U?
2
Vi
-\-'^ (3)

If the nozzle is supplied with fluid from a very large tank the enthalpy
in the tank will approach ho as the cross-section of the tank approaches
infinite area.*

It should be observed that while (3) defines the stagnation en-


thalpy it does not fix a definite stagnation state. Stagnation proper-

s
Fig. 17-2, Nozzle process state paths.

* Stagnation enthalpy
is analogous to total head in hydraulics. When appro-
priate,a gravity potential energy term, relative to some fixed datum, may be
included in the st^nation enthalpy as it is in the total head.
17-1] FLUID FLOW 295

ties other than enthalpy are taken for an isentropic stagnation state.
The isentropic stagnation state corresponding to any section of the
nozzle is the state fixed by Ao and the entropy at the section in ques-
tion. Thus at the inlet stagnation state and at the exit So = si

stagnation state Sq = S2 , but in both cases


(3). Hence ho is given by
there may be many stagnation pressures or temperatures for a given
adiabatic nozzle process, but only one stagnation enthalpy. This is

illustrated in Fig. 17-2.

Stagnation pressure is often called total or impact pressure, while pressure


independent of velocity is called static pressure. Total pressure is measured by
a tube with open end turned directly against the flow so the stream is brought

to rest at the opening. Static pressure would be measured by a gage moving


along with the stream at the stream velocity. When the flow is parallel to a
long straight wall the static pressure may be measured through a small hole
in the wall.
Stagnation temperature (impact or total temperature) is measured by a sta-
tionary thermometer in communication solely with fluid which has been brought
to rest reversibly and adiabatically from the stream velocity; this condition is
approximated by a thermometer at the nose of a rod pointed directly against the
flow. Static temperature can be measured only by a thermometer moving along
with the fluid stream. A thermometer in the side wall of a duct (analogous to a
static pressure tap) gives a temperature nearer to the impact ten^perature than
to the static temperature.
Stagnation properties should always be identified as such; an unqualified
reference to a property always signifies the static property.

It is possible to solve (3) for the velocity at any section along the
length of a nozzle; thus

V = ^2goihi - A) + Ff = ^^9o(ho - h) (4)

where symbols lacking subscripts refer to any specified section of the


nozzle. The velocity at any point is seen to be determined by the
enthalpy of the fluid at that point and the stagnation enthalpy. The
flow rate is given from (1) by

_aV _ aiVi
(5)
V t>i

Equations (4) and (5) apply to any nozzle process, reversible or


irreversible, subject to the assumptions made above. The given, or
imposed, conditions for a nozzle process are usually the inlet state,
the inlet velocity, and the exit pressure; these are not sufficient to
obtain the exit velocity and the flow rate because they do not fix
296 FLUID FLOW [ 17-2

definite states within the nozzle. If, however, the analysis is made
for a reversible nozzle process the states are fixed by their entropy,
and solutions of (4) and (5) can be made. In Fig. 17-2 the paths of
a reversible nozzle process and an irreversible nozzle process are shown
on the enthalpy-entropy plot. State 2. is a definite state fixed by
Pj and 8i, whereas state 2 depends upon the friction or irreversibility
in the nozzle; for the irreversible adiabatic nozzle 2 must lie to the
right of 2,. The figure also shows the inlet stagnation state, 0, fixed
by ho and Sj. The outlet stagnation state for the reversible nozzle
is also 0, but for the irreversible nozzle it is a different state fixed by
the intersection of the lines for ho and $2 . In Fig. 17-2, pi and pt are
static pressures, while po is a stagnation pressure.
17-2 The reversible nozzle. If the process in an adiabatic nozzle
is frictionless or reversible, the path of the process is at constant
entropy. This fact, with Eqs. (4) and (5), provides relations for
enthalpy, velocity and cross-sectional area of the stream as functions
of the pressure. From the tables of properties (or the equations for
gases) the enthalpy may be found at any pressure, and the corre-
sponding velocity is (4); similarly, when the volume
then given by
is obtained from the tables the area can be found by (5). For exam-
ple, Fig. 17-3 shows the relations for a particular case of reversible
flow of air. The velocity, specific volume and cross-sectional area
for unit flow rate are plotted against the dimensionless ratio p/po-
The general shape of the curves shown is the same for all gases and
vapors when plotted in this way. It is seen that the velocity, start-
ing from zero at the stagnation pressure, increases continuously as
the pressure falls. The specific volume, however, is initially finite;

it more rapidly, as the pressure falls. The


increases, first slowly, then
result is that the ratio of voliune to velocity v/V decreases rapidly
from infinity, passes a minimum, and increases again as the pressure
drops from stagnation pressure toward zero. The ratio a/w, area
per unit flow rate, is ^ual to v/V at all points, from Eq. (5). The
minimum value of v/V (and of a/w) for gases and vapors always
occurs in the neighborhood of p/po = 0.5. Hence if a reversible flow
is desired when the nozzle exit pressure is greater than about 0.5po

the nozzle should be of converging form; if the nozzle exit pressure is

than about O.fipo the nozzle should converge to a uniniTnnTn section


less

and then diverge to the required exit area. This is illustrated in


Fig. 17-4.
17-2] FLUID FLOW 207

P/P„

? h V V afw
psia Btu/lb ftVlb ft/sec ft*Bec/lb
100 240.98 3.70 0 00

80 226.11 4.35 840 0.00518


70 217.00 4.78 1100 0.00434
60 208.41 5.35 1280 0.00418
55 203.28 5.70 1375 0.00415
50 197.66 6.10 1470 0.00415
40 185.48 7.16 1665 0.00430
20 152.13 11.75 2110 0.00567
10 124.75 19.85 2410 0.00824
2 78.54 164.5 2850 0.0677

Fig. 17-3. Reversible adiabatic expansion of air in a nozzle; stagnation state


100 psia, 1000” F|*,.

Fig. 17-4. Types of nozzles.


208 FLUID FLOW [17-2

The situation is different for incompressible fluids (liquids). lu


such cases the speciflc volume is constant and by (5) the cross-
sectional area of the stream is inversely proportional to the velocity.
Therefore a fire hose nozzle, for example, must always be a converg-
ing nozzle to obtain maximum velocity at the exit.
A more detailed analytical approach to the characteristics of
nozzle flow can be made as follows. For any pure substance

T ds — dh — V dp (9-6)

Then for a reversible adiabatic process

dh = V dp

Using this relation in (4), the velocity is given by

V = V2go{ho - h) i> dp (6)

To make the integration, the relation between v and p must be known


for the reversible adiabatic process. For a perfect gas the relation
was found
pa* = constwt (1H7)

For real gases and vapors the same form a satisfactory


of equation is
approximation if the value of A; is properly chosen. In any event the
general trend is correctly shown by (11-17). Substituting from
(11-17) in (6)

V- - P)
fe)"
From (5)-and (11-17)

a v_ (po/p)* ^* Vo
w V
and substituting from (7),

The minimum value of a/w, corresponding to the throat of the


nozzle, can be located by taking the derivative of the right side of (8)
17-2] FLUID FLOW 299

with respect to p/po and setting it equal to zero. This gives the fol-

lowing result for the condition of minimum area:

i-{vhr
The value of the pressure ratio for minimum area is called the critical
'pressure ratio for a nozzle. The values of k applying to gases and
vapors lie between 1.1 and 1.67; the following table gives the values
of critical pressure ratio corresponding to some values of k in this
range.
k 7>/Po (critical)

1.10 0.585
1.20 0.565
1.30 0.546
1.40 0.528
1.67 0.487

It is a matter of interest that the velocity corresponding to the


critical pressure ratio in reversible nozzle flow is the local velocity of
sound.* The velocity at higher pressures is less than the local sound
velocityand the velocity at lower pressures is greater than the local
sound velocity, hence the terms sub-sonic and supersonic are applied
to the flows in the respective regions.
The manner of using the npzzle equations is best shown by an
example.

Example 1. (a) Find the exit area of a reversible nozzle to pass 1 Ib/sec
of steam if the inlet state is 100 psia, 500°F, 100 fps velocity, and the exit

pressure is 10 psia. (b) Will the nozzle be converging or converging-


diverging and, if the latter, what will be the throat area?

Solution: (a) The properties at the inlet state are hy = 1279.1 Btu/lb,
Si = 1.7085 Btu/lb ®R, At the exit state S2 = P =
2 10 psia; then from
the tables h2 = 1091.7 Btu/lb and Vi = 36.41 cu ft/lb.

02 —
V2
zf* = - -- V2 •' —
^2go{k - hi) + 7l
Substituting values in consistent units

36.4
= 1 = 0.01185 sq ft = 1.705 sq in.
o,0o0

*
The sound velocity is given by Vkg^pv. Therefore, it has different values
at different states of the fluid.
300 FLUID FLOW [ ^
17

Note the order of magnitude of the nozzle exit velocity and the relative
insignificance of the inlet velocity in this particular case. Inlet velocity of
224 fps is equivalent to 1 Btu/lb enthalpy.
(b) The exit pressure for this nozzle is well below 0.5po so the nozzle
should be converging-diverging. To find the area at the throat, the critical

pressure ratio must be known. Take from Keenrn and Keyes, Fig. 8, at

100 psia, 500®F, the value of fc = 1.30. From Eq. (9) the critical pressure
ratio is 0.55. Then the properties at the nozzle throat are as follows:

Pt = 0.55(100) = 55 psia

(The difference between pi and po, about 0.2 psi, is negligible.)

Si = 8i

From the tables hi = 1221.5 Btu/lb

vt = 8.841 cu ft/lb
Solving as in (a) above,

a, = 1 = 0.0052 sq ft = 0.75 sq in.

It is not necessary to know the


critical pressure ratio very pre-

cisely to obtaina precise value of throat area, because the area of


the nozzle changes very slowly with pressure near the throat; as an
example see Fig. 17-3. Thus if the critical pressure ratio had been
assumed to be anything between 0.53 and 0.58 the computed area
would have been the same within slide-rule accuracy. Although both
the volume and the velocity would change appreciably their ratio
would be almost constant.
17>3 Real nozzle coefficients. A real nozzle process is never
reversible; therefore the measured characteristics of real nozzle proc-
esses are referred to the characteristics of reversible processes by
means of certain coefficients defined below.
The velocity coefficient Cv of a nozzle is defined as the oatio of the
velocity of the stream at the nozzle exit section to the exit velocity
for a reversible nozzle, both nozzles having the same inlet conditions

and discharging into a space at the same pressure.

( 10)
v'2go(A) - hu)

Note that the velocity coefficient has absolutely no connection


with the velocity anywhere but at the exit section of a nozzle.
The efficiency of a nozzle is defined as the ratio of the kinetic
17-4 ] FLUID FLOW 301

energy per unit mass of the stream at the exit section of the hozzle
to the exit kinetic energy per unit mass for the reversible nozzle, both
nozzles having the same inlet conditions and discharging into a space
at the same pressure.

J1/2go _ Vl/2a,
- (U)
Vl,/2gt ho hi.

From (10) and (11)


Vn = {Cvy ( 12 )

As in the case of the velocity coefficient, the nozzle efficiency has


no connection with the kinetic energy anywhere but at the exit sec-

tion of the nozzle.


The coefficient of discharge Cy, of a nozzle is defined as the ratio
of the actual flow rate to the flow rate of a reversible nozzle of the
same minimum cross-section, both nozzles having the same inlet con-
ditions and both discharging into a space at the same pressure.

w
C„ = H. (13)
w.
^am

where a„ is the minimum area; V,m and are taken in the reversible
nozzle at minimum area section.
its

Note that the minimum cross-section may be either an exit sec-


tion or a throat section, depending upon whether the nozzle is a
converging nozzle or a converging-diverging nozzle. Furthermore
the reversible reference nozzle is not necessarily the same as the ac-
tual nozzle in this respect. For example, if an actual converging
nozzle is discharging to a region at a pressure ratio lower than the
critical pressure ratio, the reversible reference nozzle must be a
converging-diverging nozzle and its minimum area, which is its throat

area, must be the same as the minimum area of the actual nozzle,
which is its exit area.
17-4 Converging nozzles. A real nozzle may or may not be
subjected to pressure conditions for which it is properly shaped. For
example, a converging nozzle would properly be used with pressure
ratios greater than the critical pressure ratio, but it is perfectly pos-
sible to impose upon it a pressure ratio less than the critical. A brief
outline of the characteristics of converging nozzles under both condi-
tions is given below.
302 FLUID FLOW [17^

The shape of a converging nozzle is usually a smoothly curved


contour leading in a short length from the inlet to the minimum cross-
section, and often including a short length of constant area at the
exit end, to induce parallel flow in the emerging jet. Figure 17-5
shows a typical shape. The manner of pressure variation along the
stream, in and after such a nozzle, depends upon the magnitude of

Fig. 17-G. Pressure variation in flow through a converging nozzle.


17-4] FLUID FLOW 303

the pressure ratio pi/po for the exhaust space, relative to the critical
pressure ratio (p/po)«- In Fig- 17-6 several possible cases are illus-
trated in a plot of pressure vs. distance along the nozzle axis. When
the pressure ratio for the exhaust space is equal to or greater than the
critical pressure ratio (cases a, b, and c) the pressure in the nozzle at
the exit is equal to the exhaust pressure. When the pressure ratio

Fig. 17-7. Rate of flow through a converging nozzle.

for the exhaust space is less than the critical ratio, the pressure in the
nozzle will not drop below the value for the critical ratio.* Immedi-
ately upon leaving the nozzle, however, the stream finds itself in a
region of lower pressure and it expands rapidly. The inertia of the
fluid carries the expansion beyond the exhaust pressure and a series
of waves of compression and expansion is set up as indicated at d.
These waves are damped out and eventually the pressure becomes p2
by an irreversible process.
The flow rate per unit of minimum area for a converging nozzle
varies as a function of the exhaust pressure ratio in the manner shown
in Fig. 17-7. For pressure ratios greater than the critical the flow

*A physical explanation for this is that the fluid velocity at the critical
pressure ratio is the velocity of sound, which is the velocity of propagation of
small pressure disturbances in the fluid. If the fluid is moving at sonic velocity
no small pressure change can move upstream through the nozzle against the flow,
so the conditions downstream of the region of sonic velocity cannot affect the
flow upstream of this region.
304 FLUID FLOW [17-6

rate varies from zero at unity pressure ratio to a muTinnim at the


critical pressure ratio; for lower pressure ratios the flow rate is con-
stant at the TnaYinnnm value.* The latter case is sometimes called
choking flow.
The a well-made nozzle such as described
velocity coefficient of
above maybe very close to unity (0.99 or more) when the exhaust
pressure ratio is equal to or greater than the critical pressure ratio.
For lower exhaust pressures the velocity coefficient becomes smaller
as the exhaust pressure decreases, because the irreversibility after the
exit prevents the full utilization of the pressure drop in accelerating
the fluid.

The discharge coefficient is close to unity for all exhaust pressures


because the flow is fixed by conditions at the section of minimum
area, and up to this point the flow is approximately reversible in a
well-made nozzle. The irreversibility after the nozzle exit does not
affect the quantity of flow. The discharge coefficient may be slightly
greater than unity in some cases because, with exhaust pressure ratios
greater than critical, the pressure within the nozzle may drop slightly
below the exhaust pressure, thereby increasing the flow.t
17-5 Converging-diverging nozzles. The shape of a converging-
diverging nozzleis often that of a converging nozzle followed by walls

which diverge at a small constant angle, as shown in Fig. 17-8. A


slow divergence is used in order to avoid separation of the fluid from
the wall, with consequent dissipation of kinetic energy in eddies.
However, when a smaller angle is used the nozzle must be longer to
obtain a specified expansion ratio (the ratio of exit area to throat
area). This is undesirable because, in general, wall friction losses in
ducts are directly proportional to the duct length. Therefore the
angle of divergence is made as great as possible while still avoiding
separation of the jet from the wall. The angle of divergence between
the walls of nozzles is frequently between 10 and 20 degrees.
The manner of pressure variation along the stream depends upon
the magnitude of the pressure ratio P2 /P0 for the exhaust space. In
Fig. 17-9 several possible cases are illustrated in a plot of pressure
vs. distance along the nozzle axis. At least three distinct ranges of

* This constant flow rate applies to nozzles with parallel flow at the exit;
itdoes not apoly to sharp-edged orifices with converging flow at the exit.
t H. Kraft, “Reaction Tests of Turbine Nozzles for Subsonic Velocities.”
Trans. ASME, Vol. 71, October 1949, p. 781,
17-6] FLUID FLOW 305

Fig. 17-8. Converging-diverging nozzle.

exhaust pressure ratio may be identified by particular types of pres-


sure variation. The lines of demarcation between these ranges are
at the pressures shown and p<i.
in Fig. 17-9 as pt
For very low exhaust pressures such as pa the pressure drops con-
tinuously to a minimum value p6 at the exit, passing through the
critical pressure ratio at the throat. The velocity increases continu-

Fig. 17-9. Pressure variation in flow through a converging-diverging nozzle.


306 FLUID FLOW [17-6

ously, being subsonic before the throat and supersonic after the
throat. In the exhaust space the pressure drops abruptly from pf,

and there is wave formation as described for the similar case with the
converging nozzle. This situation, called under-expansion, exists for
all exhaust pressures less than the flow pattern within the nozzle
is the same for all such exhaust pressures.
When the exhaust pressure equals pt the stream flows smoothly
from the nozzle into the exhaust space and the nozzle is said to be at
its correct operating condition.
For exhaust pressures higher than pb the pressure in the nozzle
reaches a minimum lower than the exhaust pressure and then rises
to the exhaust pressure at the exit; this is called over-expansion.
There is a certain exhaust pressure, pd in Fig. 17-9, for which the
minimum pressure in the nozzle corresponds to the critical pressure
ratio and occurs at the throat of the nozzle. For exhaust pressures
such as Pc somewhat lower than pd, the pressure in the nozzle follows
the same path as in the correct operating condition to a pressure
below the exhaust pressure and then rises sharply by an irreversible
process called a normal shock. The velocity attains supersonic values
before the shock but decreases to subsonic values in passing through
the shock. After the shock there is a gradual pressure rise to the.

exhaust pressure by diffuser action (Sec. 17-7). For exhaust pres-


sures in the lower part of the range between p<j and pb the normal
shock does not occur; the flow follows the path o or 6 to the vicinity
of the exit, where the pressure rises to exhaust pressure by an irre-
versible process called an oblique shock process. In this case the
exit velocity is supersonic but it is not as great as for a reversible
expansion to the existing exhaust pressure.
For exhaust pressures such as between pd and pi, the minimum
p«,
pressure in the nozzle is reached at the throat and is greater than the
pressure for the critical ratio. The flow throughout the nozzle is
subsonic, the velocity increasing to a maximum at the throat and
then decreasing to the exit velocity. Under such conditions the di-
verging portion of the nozzle is not acting as a nozzle, but as a diffuser
(Sec. 17-7).
The flow rate per unit of minimum area for a converging-diverging
nozzle varies as a function of the exhaust pressure ratio in the manner
shown in Fig. 17-10.The variation is similar to that for the converg-
ing nozzle, but the maximum flow is reached when the throat pressure
17-6] FLUID FLOW 307

ratio reaches the critical (curve d in Fig. 17-9), which occurs when
the exhaust pressure ratio is greater than the critical.
At best, the velocity coefficient of a converging-diverging nozzle
is smaller than that of an equally well-made converging nozzle for
two reasons: (1) the greater surface for friction forces to act upon;

Fig. 17-10. Rate of flow through a converging-diverging nozzle.

(2) the higher velocities, friction forces being roughly proportional to


the square of the velocity. Separation of the stream from the diverg-
ing walls may also reduce the velocity coefficient. For correct oper-
ating conditions the velocity coefficient of a well-made converging-
diverging nozzle may be of the order of 0.95. As the exhaust pressure
departs in either direction from the correct value, the velocity coeffi-
cient drops off. For exhaust pressures between pi, and pd (Fig. 17-9),
the velocity coefficient falls to a low value and then rises again toward
unity as the exhaust pressure approaches pi.
The discharge coefficient of a well-made converging-diverging
nozzle will be close to unity for exhaust pressure ratios below the
critical. For exhaust pressure ratios above the critical the reversible
reference nozzle is a converging nozzle having exhaust pressure at its
minimum section; the real converging-diverging nozzle, however, may
have a lower pressure than exhaust pressure at its throat.
in this case
Therefore the flow in the real nozzle may be greater than in the
308 FLUID FLOW [17-6

reference nozzle, and the coefficient of discharge may rise well above
unity.*
17-6 Real nozzle computations. A properly designed real noz-
zle has cross-sections proportioned to agree with the area of the stream
as in the case of the reversible nozzle, but the areas are modified by
the effects of friction. In an adiabatic nozzle any loss of kinetic
energy through friction must be balanced by an equal increase in the
enthalpy of the fluid; therefore in a real nozzle the enthalpy differ-

ence between inlet and exhaust states will be less than in a reversible
nozzle by the amount of the friction loss. This is expressed by the
following equations;

H nn 2go(^ - hu)

ho — hi = tin{ho — hi,) (14)

The states involved are plotted in Fig. 17-11,

Fig. 17-11. Flow state paths, reversible and irreversible.

• J. H. Keenan, ‘^Reaction Tests of Turbine Nozzles for Supersonic Velocities.”


Trana. ASME^ Vol. 71, October 1949, p. 773.
17-6] FLUID FLOW 309

Some examples of the manner of using nozzle coefficients in com-


putations will now be given.
Example 2. Find the throat and exit areas for a nozzle to pass 10,000

Ib/hr of steam from an initial state of 250 psia, 500®F, negligible velocity,

to a final pressure of 1 psia, if the velocity coefficient is 0.949 and the dis-
charge coefficient is unity. Find the exit jet velocity.

Solution: ho = 1263.4 Btu/lb; sq = 1.5949 Btu/lb®R. The throat


pressure pt is obtained from the critical pressure ratio as in Example 1 of
this chapter.

pt = 0.55po = 137.5 psia

At Pt and so, from the tables, ht^ = 1208.2 Btu/lb and i;*, = 3.415 cu ft/lb.
At p 2 and So find = 891.0 Btu/lb. The throat area is fixed by the desired
flow and the coefficient of discharge.

c„ Vt.

Vt. = V2ffo 778(1263.4 - 1208.2) = 223.8\/55 = 1660 fps

10,000
w = = 2.778 Ib/sec
3,600

at = 0.00572 ft*

The exit area is fixed by the desired flow and the actual volume and velocity
(not by the discharge coefficient, which applies solely to the minimum section
of the nozzle). The actual exit velocity is given by

V2 = Cv V 2go(fio — Ihs) = 4,094 fps

The actual exit volume can be found only through the actual enthalpy which
is given by Eq. (14);
A2 = ^ 7ln{ho ““ hu)

Vn = (CvY = 0.90

A2 = 928 Btu/lb

From the tables, at pz and A 2 ,

1^2 = 276 cu ft/lb

Oj = = 0.187 ft*

For the frictionless noezle, for comparison,

ou - 0.170 ft*
310 FLUID FLOW [17-6

If the discharge coefficient had been 0.98 what would the throat and exit
areas have been?

o, . ^ * 0.00584 ft*
Cw

The exit area is unchanged because the actual flow rate and the nozzle effi-

ciency, which fix the exit area, are specified conditions independent of the
discharge coefficient.

Example 3. Find the exit velocity and the throat and exit areas for a
nozzle to pass 1 Ib/sec of air from an inlet state of 150 psia, 340^F, negligible
velocity, to an exhaust pressure of 15 psia if the nozzle efficiency is 88 percent
and the discharge coefficient is unity.
Solution: (a) By the perfect gas rules.
For air, fc = 1,40; pf/po = 0.53

""
'
(0.53)(i -»-*>/i-^ = 0.833
T (p )
Tt. = (800) (0.833) =
2go{hQ — hu) = ^ ^oCpiTo — Tu)

Vu - V2go 778(0.24)(134) = 1270 fps

Vt, = ^ Pi
= 3.12 cu ft/lb

at = ^^ = V,.
0.00246 ft*

*~*>'‘ *
la / 15
Y* = 0.518
T, Visoj
Tu = (0.518) (800) = 414‘’R

From (14), since Ah = Cp AT,

CpiTo — T2) = VnCpiTo — Tit)

Ti^ To- riniTo - Ti.) = 460®R

Vi = ^2^00,(70 - Ti) = 2030 fps

», = ^* = 11.4 cu ft/lb
Pt

0, = ^* = 0.00560 ft*
Vi

(b) By the Gas Tables of Keenan and Kaye. From Table 1 at Tu,
* 191.81 Btu/lb; pr« *= 5.526. From Tabic 2, in the temperature range
17-7] FLUID FLOW 311

involved, k » 1.39 to 1.40. From Table 24 (or from Eq. (9) of this chapter),

P(/po 0.53. By Hie method of the tables, for an isentropic,

= £-‘ = 0.53
Pro PO

Pn * (0.53)(5.526) = 2.929

From Table 1 at pr = 2.929, Tu = 668®R, hu = 159.9 Btu/lb.

Vu = V2po 778(191.81 - 159.9) = 1264 fps

= M!l* = 3.11 cu ft/lb


= = 0.00246 ft*
C-F,.

The throat area for reversible flow can also be obtained directly from Table 2
of Keenan and Kaye, which gives values of the maximum flow rate per unit
area per unit initial pressure for air expanding from any given initial tem-
perature.

For the exit section

Prt = ^p.0 = 0.5526

Ts. = 415'’R

/ti. = 99.13 Btu/lb

hi — htt — iiniho — hu) = 110.25 Btu/lb

Corresponding to hi,

Ti = 462“R
Vi = V2go{h, - hi) = 2020 fps

vi = = 11.4 cu ft/lb
Pi

ai:=w^ = 0.00565 ft*


Vi

For converging nozzles the computations are made in a similar way but
the minimum area is also the exit area. Hence in this case the exit area is
not computed from exit velocity and volume but comes from the discharge
coefficient, which applies to the minimum area, wherever situated.

17-7 Diffusers. If a nozzle process is reversed, the result will


the compression of a fluid at the expense of its kinetic energJ^
duct in which this effect is accomplished is called a diffuser. The
312 FLUID FLOW [17-7

6000 NOZZLE POOR NOZZLE


POOR DIFFUSER GOOD DIFFUSER

Fig. 17-12. Subsonic nozzle and diffuser.

same equations apply to diffusers as to nozzles, and in the reversible


case the only difference in operation between the nozzle and the
diffuser is the reversal in direction of all velocities.
In the real case, for subsonic flow the diffuser shape must differ
from the nozzle shape because of the difficulty of causing the fluid to
fill a diverging passage. The pressure in a diffuser is rising in the
direction of flow; consequently the establishment of reverse flow and
separation from the wall in the slower-moving fluid near the wall
(boundary layer) is aided by the pressure gradient. Separation re-
sults in the formation of large eddies in the fluid, with consequent
friction losses. For this reason the angle of divergence of an efficient

diffuser must be small ;


tests indicate minimum losses when the angle
between the walls is of the order of 7 degrees.
Figure 17-12 indicates the general difference in shape between a
good nozzle and a good diffuser for a pressure ratio greater than criti-
cal. For a pressure ratio smaller than critical the diffuser must con-
verge and then diverge as indicated in Fig. 17-13. The relation of
supersonic diffusers to supersonic nozzles is not as simple as that
between subsonic diffusers and nozzles, because of shock effects; this
matter is beyond the scope of this book.
The efficiency ofa diffuser is defined as the ratio of the enthalpy
increase in a reversible adiabatic diffuser to the enthalpy increase in
the real diffuser, when both diffusers receive fluid at the same state,

NOZZLE DIFFUSER
FLOW FLOW

NOZZLE DIFFUSER

Fig. 17-13. Supersonic nozzle and diffuser.


17-8] FLUID FLOW 313

and discharge to a region at the same pressure. Diffuser efficiency


is related to nozzle efficiency as pump efficiency is related to engine

efficiency.
Diffusers are used as adjuncts to, or integral parts of,pumps,
fans, jet pumps (ejectors and injectors), venturi meters, and other
apparatus where kinetic energy is available and increased pressure is
desired. Except for the development of the flow equations, which is
done as for the nozzle, the analysis of such devices belongs in the fleld

of fluid mechanics.

17-8 Supersaturation in nozzle flow. When a nozzle process


starts with superheated vapor and ends with a wet mixture, the states
of the fluid in the process do not always correspond to the states
shown on the ordinary diagrams of properties. As the fluid passes
the saturation line at pa in Fig. 17-15 it might be expected that con-
densation would begin; however, the velocity of the fluid may be so
great that before any drops of liquid have time to form the fluid has
passed well below the saturation state. During this interval the
fluid is not in a state of stable equilibrium, for it is a homogeneous
vapor at a temperature below the saturation temperature for the
existing pressure; it is said to be in a metastable state. The name
17- 9 ] FLUID FLOW 317

Now for compressible fluids it is possible to write

where the coefficient Y is defined by setting (19) equal to (17). Y is


a function of the dimensionless ratios k, 02 / 01 , and Pi/pi] it has been
evaluated and tabulated for certain frequently encountered cases
and values may be found in the references at the end of this chapter.
For the special case of critical flow (exhaust pressure ratio less
than critical pressure ratio) Eq. (17) is greatly simplified by substitu-
tion from which gives a fixed value of the pressure ratio at the
(9),
minimum any given value of k. For example, for critical
section, for
flow with k equal to 1.4 Eq. (17) reduces to

The relative simplicity of this type of equation is apparent; the fact


that under critical flow conditions the coefficient of a well-made
rounded entrance nozzle can be relied upon to be close to unity
(approximately 0.99) recommends the critical flow meter for applica-
tions in which the large pressure drop is acceptable.
The is a form of flow meter which uses a sharp
thin plate orifice
edged hole in a plate, as shown in Fig. 17-17, for a metering passage.
The flow through the orifice is similar to that through a flow nozzle
but the fluid passing the orifice has not been guided into a parallel
stream; it is converging on the hole from and in the absence
all sides,

of infinite forces cannot turn immediately around the sharp comer


Hence the fluid completes its convergence at
into a parallel stream.
some distance downstream from the edge of the orifice plate, where
the area of the stream is less than the area of the orifice. Since the
orifice a physical dimension of the apparatus, it is used for Ot
area is

in the flow equations, and the difference between the fluid stream
area at section 2 and the orifice area is accounted for in the expert-
318 FLUID FLOW

Fig. 17-17. Thin-plate orifice flow meter.

mental coefficient C. The value of C will then depend upon the


location of the pressure taps (measuring holes) as well as upon the
ratio of orifice area to pipe area, and other factors.

The equations used for the thin plate orifice are identical with
those used for the flow nozzle but the values of the coefficient C are
quite different, orifice coefficients being of the order of O.fi while nozzle
coefficients are of the order of unity. The coefficient Y for com-
pressible flow is also different for orifices from that for flow nozzles
and venturi meters. Values of coefficients for standardized designs
of flow nozzles and orifices may be found in the ASME references.

PROBLEMS
17-1. (a) In reversible adiabatic nozzle flow under the assumptions made
in this chapter, what will happen to the stagnation temperature and stagnation
pressure as the static pressure falls? (b) In irreversible adiabatic nozzle flow of
a p«>rfect gas what will happen to the stagnation temperature and stagnation
pressure as the static pressure falls?
17-2. (a) Steam is flowing in a duct at 50 psia, SOO^F, 1000 fps; find the
FLUID FLOW 319

stagnation enthalpy, stagnation pressure, and stagnation temperature of the


steam, (b) Repeat (a) for a velocity of 200 fps. (c) At low velocities the static

and impact properties differ negligibly. For air at total temperature of 140^F
find the velocity at which the static and impact pressures differ by 1 percent,
(d) Repeat (c) for a difference of 1 percent in static and impact temperatures
(1 percent of absolute temperature).
17-3. Make a plot like Fig. 17-3 for the reversible adiabatic expansion of
steam in a nozzle from a stagnation state at 160 psia, 600°F, to atmospheric
pressure, taking property values from the steam tables. Estimate the critical
pressure ratio from your plot.
17-4. A nozzle is to be designed to pass 10,000 Ib/hr of liquid water from a
pipe to a tank. The water 350 psia, 70°F, and 10 fps velocity;
in the pipe is at

the static pressure in the tank is Consider the water incompressible


15 psia.
{v = constant). Sketch a suitable nozzle shape and calculate the minimum
diameter of the nozzle.
17-5. A turbine nozzle is to pass 1 Ib/sec of hydrogen from 100 psia, 70®F,
to a region at 10 psia. If the process is frictionless and adiabatic find the nozzle
throat and exit areas, the gas exit velocity, and the gas temperature at exit.
State any assumptions made.
17-6. A steam turbine nozzle is to receive steam at 800 psia, 800®F, 450 fps,
and to discharge to a region at 700 psia. The coefficient of discharge is 1.01,
and the nozzle efficiency is 0.98. Find the minimum area, and the exit area, of
the nozzle if the flow rate is 1 Ib/sec.
17-7. Gaseous combustion products having properties, Cp = 0.26 Btu/lb F,
k = 1.36, and R = 53.2 ft Ib/lb R, are to flow at the rate of 100 Ib/sec through
a nozzle having inlet conditions of 15 psia, 1500°F, 600 fps, and discharge pressure
of 5 psia. The discharge coefficient is 0.96, and the nozzle efficiency is 0.88.
Find the proper throat and exit areas and the exit temperature.
17-8. Solve Problem 17-7 if the gas is air having the properties given in the
Gas Tables of Keenan and Kaye.
17-9. A nozzle receives 10,000 Ib/hr of steam at 250 psia, 500°F, 200 fps,
and discharges to a region at 100 psia. The average exit velocity is measured
as 1850 fps. The nozzle exit area is 1.00 sq in. Find: (a) the nozzle efficiency;
(b) the throat area,assuming reversible flow in the nozzle as far as the throat;
(c) whether the stream is over-expanded or under-expanded.
17-10. Steam enters a diffuser at 1000 fps, 20 psia, 300°F, and is diffused to
50 fps. (a) If the diffuser is reversible find the exit pressure and the ratio of
exit area to inlet area, (b) If the diffuser efficiency is 80 percent find the exit
pressure and the ratio of exit area to inlet area.
17-11. An adiabatic diffuser is to be used to recover kinetic energy from the
discharge of an exhaust fan. The gas flowing is a mixture of combustion products
having properties Cp — 0.25 Btu/lb°F, k = 1.38, and 7? =» 54 ft lb/lb®R. The
gas leaves the fan at 400 fps, 540®F, and flows through the diffuser which reduces
the velocity to 100 fps at the discharge to atmosphere. If the diffuser efficiency
is90 percent how much below the atmaspheric pressure of 15 psia will the diffuser
inlet pressure be? What will be the ratio of diffuser exit area to inlet area?
17-12. An ejector, Fig. 17-14, has a primary flow, lUi, of 1.0 Ib/sec of air at
320 FLUID FLOW

100 peia, 340°F; the secondary air flow, Wzy enters at 5 psia, 340*^F. The dis-
charge at section 4 is to the atmosphere at 15 psia. The efficiency of the ejector
as a pump is defined by
_ ^^(^4 — ht),
’ m>i(Ai - fc«).

where the enthalpy differences are taken on isentropic paths from the respective
initial states to the pressure at 4. (a) If the efficiency of the ejector is 20 percent
find the flow rate, W2 . (b) If the entrance and exit velocities are all 80 fps, and
the efficiency is 20 percjent, find the diameter at section 4.

17-13. Steam from a large chamber at 180 psia, 400^F, expands to atmos-
pheric pressure through an adiabatic nozzle of 0.00270 sq ft throat area; the
flow rate is 1.00 Ib/sec. Find the based upon each of
coefficient of discharge
two an equilibrium isentropic path (steam
ideal state ][.ath6: (a) the fluid follows
table data); (b) the fluid follows a metastable isentropic path in the mixture
region. (Supersaturated properties obtained by using pv^ - constant; from
Fig. 8 of Keenan and Keyes, k is about 1.30 for the given conditions.)
17-14. When a hot liquid passes below ithc saturation pressure during a nozzle
process some of the liquid may ‘flash' into vapor. Such ^flashing flow" requires
* ^
^

much more area than liquid flow. Assuming reversible adiabatic flow, find from
the steam table data the nozzle exit area required for the conditions of Problem
17-4, if the initial temperature is 300®F. (Actually, the vaporization will be
delayed, and the flow rate per unit area may be much greater than calculated.
This is analogous to the supersaturation phenomenon with vapor.)
17-15, A flow nozzle of 6.00 in. diameter in an 8.00 in. diameter pipe is used
to meter steam which arrives at the upstream pressure tap at 40 psia, 300®F.
If the differential pressure (pi — pa) is 80 in. of water, the nozzle coefficient, C,
is 0.985, and the expansion factor, F, is 0.94, find the flow rate in Ib/hr.

17-16. A sharp-edged orifice, 0.75 in, diameter in a 3.00 in. pipe, is calibrated
with air at 14.8 psia and 75®!^ when the air flow rate (measured by a direct
volumetric method) is 7.75 cu ft/min, the differential pressure is 1.00 in. of water.
This is so small (about J pt^rcent of the absolute pressure) that the flow may be
considered incompressible. What is the coefficient, (7, for the orifice?
17-17. A well-known empirical formula for critical flow of saturated steam
through a rounded-entrance nozzle with negligible approach velocity is Napier’s
formula,

where w is flow rate, Ib/sec; a is nozzle throat area, sq in.; and pi is upstream
pressure (stagnation), psia.
Compute the rate of flow through a nozzle of 1.00 sq in. area, with steam
supplied at 100 psia, saturated, for critical flow, using: (a) Napier's formula;
(b) the thermodynamic nozzle equations, taking the flow to be isentropic. (Note:
Although good agreement is found between Napier's formula and the thermo-
d3mamic equations for^a considerable range of pressures, using saturated steam,
the formula does not apply for other conditions. Empirical formulas are useful,
but care should be taken not to mis-use them.)
FLUID FLOW 321

17-18. Fliegner’s empirica] formula for critical flow of air through a rounded-
entrance nozzle with negligible approach velocity is

where w is flow rate, Ib/sec; a is nozzle throat area, sq in.; pi is upstream pressure
(stagnation), psia; and Ti is upstream temperature (stagnation), Fabe.

Taking air as a perfect gas of specificand gas constant 53.34


heat ratio 1.4,

ft lb/lb®R, derive from the thermodynamic flow equation an equation in the

form of Fliegner’s formula and compare the values of the numerical coefficients
in the derived equation and in the empirical formula.
Chapter 18

TURBINES

Turbines, driven either by steam or by water, have for many years


dominated the field of large stationary prime movers. The gas tur-
now replacing large aircraft reciprocating engines, and in future
bine is
may well replace many diesel engines in land and marine plants.
Hence the turbine is a machine of the greatest practical importance
in the field of power engineering.
18«1 Turbines. A turbine is a machine in which the acceleration
of a stream of fluid produces a turning moment on a rotating shaft.
The acceleration may be a change in either magnitude or direction
of the fluid velocity. Thus in the classical turbine of Hero, Fig. 18-1,

the fluid entering through the axis of the sphere changes direction to

ESCAPING FLUID

Fig. 18-1. Hero’s turbine.

322
18-1] TURBINES 323

escape through the tangential exhaust nozzles. If the nozzles are


smaller in cross-section than the axial inlet pipes, there may also be
an increase in magnitude of the velocity. In any case, by Newton's
laws of motion the force required to accelerate the fluid must have
an equal and opposite reaction; the reaction constitutes a tangential
force on the sphere. If the sphere rotates, this force does work and

the device becomes a turbine.

Fig. 18-2. Simple impulse turbine. (Courtesy De Laval Steam Turbine Co.)

In modern turbines the same principle applies as in Hero's tur-


bine; a tangential force on a rotating body is produced by a change
in the tangential component of velocity of a fluid stream. In Fig.
18-2 is shown a simple type of turbine named for its inventor, DeLaval.
In this turbine a jet of fluid from the fixed nozzle strikes the buckets
on the wheel in nearly the tangential direction. The curvature of

the buckets is such that the stream is turned, relative to the buckets,
almost 180 degrees, reversing the tangential component of its velocity
relative to the buckets. The force required to accelerate the fluid
in this manner comes from the buckets; consequently the reaction
force is a tangential force upon the buckets.
324 TURBINES [ 18-2

18-2 Force relations for a fluid stream in steady flow. The


analysis of the turbine process for compressible fluids requires the
use of both mechanics and thermodynamics. A simple presentation
of the basic mechanics follows. In this development a one-dimen-
sional flow is assumed; this means that velocity does not vary over
any cross-section taken normal to the stream. Also, in this chapter
gravity forces are neglected.
By Newton’s laws of motion, the force acting on a mass to change
its velocity is proportional to the rate of change of momentum of the
mass.

F.C^ 1
( )

Consider a stream flowing steadily at rate w Ib/sec through a curved


passage in the xy plane, Fig. 18-3; take a control volume bounded by

the .walls of the passage and the sections 1 and 2, normal to the flow
axis. It may be shown from fluid mechanics that the application of
Eq. (1) to the fluid within the control volume gives the following
result:

F= ^ (T, - V,) (2)


go

Equation (2) says that the accelerating force upon the fluid within
the control volume is equal to the difference between the momentum
volume per unit time and the momen-
of the fluid leaving the control
tum of the fluid entering the control volume per unit time.
Equation (2) is a vector equation. For many purposes it is most
conveniently written in terms of components, thus:
18-2 ] TURBINES 325

= 7 (^2, - Vu)
I (2a)

The forces acting upon the volume are


fluid within the control
represented in Fig. 18-4. piOi and ptOi are pressure forces from the
adjacent fluid, upon the sections 1 and 2; siOi and 8202 are shear forces

on these sections; R is the resultant of pressure and shear forces from


the passage walls upon the fluid; F is the overall resultant of all forces
upon the fluid, and — F is the resultant reaction of the fluid. The
relations of these forces may be expressed by

F~ piOi + piOi + Sitti -t- SiOi +R (3)

or 0 = (—F) + piOi + p2a2 -f SiOi -|- ^202 -f- R


In these equations each term is a vector and the addition is by the
laws of vectors. Equation (3) may also be written in terms of com-
ponents, for example,

Fx - (piOi), + {piCh)x + (siffli), f {siOi), -I- Rx (3a)

In this equation, forces are positive when they act in the positive x
direction ; if a: is positive in the direction of flow, (piCi)* is a positive
force, while (p202)* is negative.

Under the assumptions made here, the velocities at sections 1 and


2 are uniform across the section, and there will not be any shear
forces on these sections. For the real flow patterns ordinarily en-
countered the net shear forces on sections 1 and 2 are negligible, first
326 TURBINES [IM

because the friction forces are small, and second because s}miimetry
in the flow tends to cause shear forces to add to zero at each section.

Fig. 18-5. Turbine wheel and linear bucket row.

18-3 Force and work on a turbine bucket. In this section it

is assumed that the row of turbine buckets on the rim of a wheel is

equivalent to a straight line of buckets moving in the tangential direc-


tion as indicated in Fig. 18-5. Looking radially in toward the axis of
rotation, the cross-section of the flow path is as shown in Fig. 18-6.
The velocities of the fluid at the entrance and exit from the buck-
ets are shown by a velocity diagram, Fig. 18-7. The velocity of the

Fig. 18-6. Flow path nomenclature.


18-3] TURBINES 327

Fig. 18-7. Velocity diagram.

nozzle jet Vi is the absolute velocity entering the bucket; the angle a
between Vi and the tangential direction is the nozzle angle. Since
the bucket Js moving with velocity Ft the^rcZofiv^elocity entering
the bucket Fir is obtained by subtracting F6 from Fi; this is a vector
subtraction as shown in the figure.
The relative velocity of the stream leaving the bucket Fi« is fixed
in direction by the hvxket exit angle The magnitude of F 2 jt is

determined by considering the bucket passage as a nozzle. Applying


Eq. (17-10) with a velocity coefficient Ct,

ViR = ^2g,{}i, - h.) + (4)

The absolute velocity leaving the bucket F2 is obtained by adding


F6 to F2« vectorially as shown in Fig. 18-7.
The tangential and axial components of the several velocities are
designated respectively by y and z with appropriate subscripts. The
two velocity triangles of Fig. 18-7 are shown complete with compo-
nents in Fig. 18-8.
From Eq. (3a), neglecting shear forces on sections 1 and 2, the

y component of force upon the fluid from the bucket passage walls is

Ry — — {piai)y — {pjfli)y (5)

Now the tangential components of pressure force on the surfaces


1 and 2 are zero because of the direction of the surfaces. Therefore
328 TURBINES [ 18 ^

Fig. 18-8. Velocity components.

if the tangential component of force upon the bucket from the fluid
is Ft,

Ft = -Ry = -Fy ( 6)

where y is in the tangential direction. Then from (2a) and (6)

F, = j{yi-y2) (7)
9o

The axial direction is designated the z direction; then the axial


force on the bucket Fa is the negative of Ry, the z component of
bucket force on the fluid. The surfaces 1 and 2 are normal to the z
direction and the pressure on ai is in the positive sense, while the
pressure on Ot is in the negative sense. Then from (3a), since the
shear forces have no axial component,

F, = PlOl - P20i! + Ry
and Fa = —Ry = —F, + piOi — pjo*

w -
or =7 (*» 2*) + PiOi ^ (7a)
9o

In actual turbines the effective areas against which the pressures


Pi and pt act are not limi ted to the bucket passage areas but include
all the areas of the buckets and their supporting disc or drum which
are subjected to those pressures.
18^] TURBINES 329

The work done per second on a turbine bucket is given by the


product of tangential force on the bucket and tangential velocity of
the bucket, or

,Tr W
=-
wW, = FtV, (j/1 - ydVt ( 8)
go

where Wg is the work done per pound of fluid. Then


jrr
^ ivi - 1h)Vh ft Ibf
(9)
go Ibm
By substituting for j/i its equal {yiR + Vi), and for yt its equal

{ytR + Vi), (7) becomes


w
=~
Ft (yiR — yoR) ( 10 )
9o

and (9) becomes F. = ( 11 )

Equations (9) and (11) may be written in different forms by


geometric reasoning. Referring to Fig. 18-8, by the properties of
right triangles,
y\ =V? +zj
y\R — ViB "b
I (a)
yl =Vl +zl
vIr — vis + ziji

Also, by the properties of relative and absolute velocities,

Vo = yi- yiR = yi- ytR (b)

Substituting from (b) in (9) and (11) and adding (9) to (11) gives

TT yi - yiR + iAr - yl
( 12 )
2go

Substituting from (a) in (12) gives

V? - v!« + vi« - Vi
Wg = (13)
2go

Equation (13) may also be derived by thermodynamic reasoning,


as follows. Consider the turbine bucket passage in Fig. 18-7. Ap-
plying the steady flow energy equation to a stationary control volume
330 TURBINES [184

extending through the bucket region from section 1 to section 2 gives,


for an adiabatic turbine,

(c)

The steady flow energy equation can also be applied to a control


volume moving with the buckets at constant speed; in this case rela-
tive velocitiesmust be used. Since no mechanism moves through
the control surface there is no shaft work (on the basis of a control
volume moving with the buckets the flow through the buckets is
simply duct flow). The enthalpies at 1 and 2, however, are inde-
pendent of the basis of velocities. Therefore on the relative basis

VlK
'
_
“ , VIh
4.yk
i,
"I +
_i_

2go
.

2go

Eliminating hi and hi between (c) and (d) gives (13).


The relations derived in this section are based on the assumption
of uniform velocities across the fluid stream at all sections and uni-
form bucket velocity. In real turbines the fluid velocity will vary
across each section, the bucket velocity will vary in the radial direc-
tion, and the mean radius of the bucket (and its mean velocity) often
increase in the direction of flow. Some turbines are arranged for
radial flow instead of the axial flow considered here. These varia-
tions are accounted for by more detailed methods of analysis, which
do not fall within the scope of this book.
18-4 Nozzle-bucket efficiency. The effectiveness of a nozzle
and bucket combination in obtaining shaft work from a fluid stream
is stated in terms of the nozde-bucket effieimcy, i}nb, defined by

Vnb = (14)

where Wx is the work done by the fluid on the buckets, ha is the


enthalpy of the fluid approaching the nozzle, and hu is the enthalpy
at a state having the exit pressure and the entropy of the fluid ap-
proaching the nozzle. The nozzle-bucket efficiency compares the
work of the actual combination to the shaft work of a reversible adia-
batic expansion in steady flow, beginning at the nozzle inlet state and
ending at the bucket exit pressure. In Fig. 18-9 are plotted the state
184 ] TURBINES 331

Fig. 18-9. (a) Turbine stage flow diagram, (b) State path plot.

paths for a nozzle-bucket process and the corresponding reversible


adiabatic process.
It is sometimes convenient to substitute for the isenlropk enthalpy
drop (ha — hi) an equivalent kinetic energy in terms of a characterislic
velocity 7c defined by

7c = V2g,(ha - M (15)

Wa
Then rini = « (16)
7?/20b

The nozzle-bucket efficiency depends upon the geometry of the


passages, and upon irreversibility due to friction and shock effects.

The nozzle-bucket eflBciency for certain hypothetical reversible com-


binations will be determined in the following sections in order to
compare the various arrangements.
18>5 Turbine stages —^passage area. Many turbines have sev*
eral successive rows of nozzles and buckets through which the fluid

passes. Each row of nozzles together with the buckets and guide
vanes through which the fluid passes before reaching the next row of
nozzles is called a stage of the turbine.

At each stage of a turbine it is necessary to provide sufficient area


for the flow. Since the pressure of the fluid is progressively lower as
it passes through the stages of a turbine, the volume is larger and the
332 TURBINES [18-8

cross-section of the flow path must increase to acconunodate the flow.


This increase may take place both by increasing the mean radius of
the stage and by increasing the radial length of the passages as shown
in Figs. 18-24 and 18-25.
In small turbines the nozzles may be of circular cross-section but
these do not utilize effectively the full annular area of the bucket
passages; therefore in larger turbines the nozzles are made by arrang-
ing partitions in an annular ring, just as the buckets are arranged
around the wheel. This is illustrated in Fig. 18-10. For either noz-
zles or buckets the net area of the passages normal to the flow is

Fig. 18-10. Passage area.


18-6 ] TURBINES 333

obtained in the manner illustrated in Fig. 18-10 for the particular


case of the nozzle exit area. In this figure the total annular area
occupied by the exit plane of the nozzle (assuming I small relative
to D) is ttDI; but the partitions have a certain edge thickness b,

leaving only a length o open for flow. The ratio of o to (a + b)


may be called the area factw The area normal to the flow is re-
f.

duced further by the angularity of the flow, so the net area is given by

A = irDlf sin a (17)

Corresponding relations apply to the bucket inlet and exit areas.


For any given velocity leaving the nozzle the flow area required
can be determined from the continuity equation if the state of the
fluid is known; this was discussed in Sec. 17-6. The same computa-
tion can be made for the bucket, taking care to use relative velocities
in this case. Once the required area is known the values of D, I, f,

and the angle may be chosen to give that area. As in all such design
problems the actual values chosen for these variables are governed
by considerations of manufacturing practice as well as by the effects

of the various possible choices on the characteristics of the machine.


Experience and judgment are necessary to determine approximate
values which must then be adjusted to satisfy best the various re-
quirements.
18-6 The simple impulse stage. The classical distinction among
turbine stages is that between impulse and reaction stages. As a
matter of convenience this distinction will be followed here, but it

should be understood that in practice the two types differ more in


degree than in kind. In the simple impulse stage the entire pressure
drop occurs in the nozzle; this can be made to happen simply by
providing free passages (other than the bucket passages) from one
side of the buckets to the other so that the pressure is equalized.
The changes of pressure and absolute velocity during flow through
such a stage are indicated in Fig. 18-11.
A velocity diagram for an impulse stage is shown in Fig. 18-12.
The forces and the work on the bucket are determined by the rela-
tions derived in Sec. 18-3. The various velocities are determined as
follows. The nozzle jet velocity Vi is fixed by the fluid state before

the nozzle, the exit pressure, and the nozzle velocity coefficient, as
shown in Chap. 17. The bucket velocity is fixed by the speed
18-6] TURBINES 335

and size of the wheel. Then the other velocities can be calculated
if the angles a and and the bucket velocity coefficient are known.
The relations used are summarized below.

j/i = Vi cos a

yiR = yi- Vi,

ViK = Vzjf + fig

V,R = CiVin
yiR = ViR cos j8

2/j = ytR +
22 = ViR sin /3

It is to be noted that all additions are algebraic and that yiR is

always in the negative direction; hence yt may be either positive or


negative depending upon the relative magnitudes of ytR and 76. It
is also to be noted that there is no nozzle action in the bucket; the

difference in magnitude ofViR from Vir is due to friction. (In real


impulse stages there may be some pressure drop in the buckets.)
The bucket entrance angle does not enter the above computations
at but from physical considerations it is apparent that the bucket
all,

entrance should be aligned with the relative velocity 7i«. There is,
however, nothing to prevent the operation of a turbine at such condi-
tions that the angle of Fia is not equal to the bucket entrance angle,
in which case there willbe some shock loss due to the incorrect angle.
In the consideration of hypothetical stages under different design
conditions it is assumed in each case that the correct entrance angle
is used. In the consideration of different operating conditions for a
given design, however, the reduction of efficiency due to shock at the
bucket entrance must be taken into account for conditions other than
the design condition. The distinction between design characteristics
and operating characteristics is often overlooked by students, but it
is of fundamental importance in the study of machinery of any type.
It is possible to show the general nature of the design character-
istics of an impulse stage by means of a hypothetical reversible stage
having variable bucket angles such that the bucket entrance angle is

always kept equal to the angle of the entering relative velocity Fia
336 TURBINES

and the bucket e?^ angl^ is kept equal to the bucket entrance angle.
For such a case 72* = Fiji (no friction.) and Zt = Zi. ptg = ym in
magnitude (similar triangles), but algebraically y2it = —yis. Then

^2 = 7j + ym = 2Vt — yi (a)

The tangential force is

7j = ^(j/l- J/2) =^(j/i- F») (b)


ffo ffo

The work per pound of fluid is

17. = (yj, - VI) (c)


w ^
go

From (15) yi = 7i cos a = 7* cos a (d)

From (16), (o) and (d) the nozzle-bucket efficiency for the hypotheti-
cal stage is

-
1
J -6 = ^ (7»7.
ye
cos a 71)

Plotting efficiency as a function of VolV^ for nozzle angles of zero


and 20 degrees gives the parabolas of Fig. 18-13. By differentiation
it can be shown that the maximum efficiency for the hypothetical

stage occurs at 7fc/7c = 0.5 cos a and is equal to cos* a. Thus, to


obtain maximum efficiency, the nozzle angle should be as small as
possible. Since zero angle would give zero area for flow, the nozzle
angle must have some finite value; it may be of the order of 15 or 20
degrees in actual turbines.
The relations for the hypothetical reversible impulse stage with
equal entrance and exit angles show qualitatively what happens in
real impulse stages as the design conditions are altered. For maxi-
mum nozzle-bucket efficiency the bucket velocity must be in the
vicinity of half the nozzle jet velocity, but the exact point of best
efficiency depends upon the angles and the friction and shock losses.

Example 1. Given the following data for a simple impulse turbine stage.
Fig. 18-11, find the work done per pound of fluid flow, the axial thrust per
pound of fluid flow per second, and the nozzle-bucket efficiency.
18-6] TURBINES 337

0.5

Fig. 18-13. Impulse stage eflBciency; for reversible flow


with symmetrical buckets.

Isentropic enthalpy drop : 50 Btu/lb


Approach velocity: negligible

Cv = 0.95

^6 - 700 fps

a = 20 deg

/3 = 30 deg

Cb = 0.95
Solution:

Vi = CW2go (M)s = 0.95V'(64.4)(778)(50)

= 1500 fps
yi = ?» cos a = (1500) (0.940) * 1410 fps

«i = ?i sin a = (1500) (0.342) - 513 fps


yui = yi - “ 1410 - 700 = 710 fps
338 TURBINES tlS-7

* ^ViR "f* *? ” 876 fpB


VtB = CiViB = (0.95) (876) = 832 fps (magnitude)
yi* = cos /3 = -(832) (0.866) = -720 fps

Note the negative sign; inspection of the diagram shows that yis is in the
negative tangential direction.

ft = sin /3 = (832) (0.500) = 416 fps

From (11) JF.

= ^ (710 + 720)(700)
= 31,080 ft Ibf/lbm

From (7a) Fa = - (ft — ft) + piOi - ptot


go

In the simple impulse turbine, pi = pt and oi and Oj are usually made equal
in the design; then

W, 31^080
0.80
F?/2jo 38,900

The nozzle efficiency in the above example is or 0.903; there-


fore the efficiency of the buckets in utilizing the kinetic energy actu-
ally supplied by the nozzle is 0.80/0.903 or 0.887. Thus despite fric-
tion in the buckets the efficiency of the buckets is greater than for
the hypothetical reversible stage of equal angles, which has efficiency
of cos‘ 20 deg or 0.883. This is made possible by the fact that the
exit angle is smaller than the bucket entrance angle and the work
of the stage is thereby increased over the work for the hypothetical
stage.
18-7 The velocity-compounded stage. In a two-row velocity-
compounded stage, Fig. 18-14, the nozzle jet works on two rows of
moving buckets in succession, and passes through an intermediate
row of fixed guide passages between the two bucket rows. The inter-
mediate guide passages are not nozzles because they are not designed
to expand the fluid but simply to change its direction (in actual cases
some small expansion may occur in these passages). The changes of
pressure and absolute velocity in a two-row stage are indicated in
Fig. 18-15.
18-7] TURBINES 339

Fig. 18-14. Reaction stage and velocity-compounded stage. Note sealing


strips to minimize leakage around ends of reaction blading. (Courtesy Westing-
house Electric Corporation.)

Fig. 18-15. Velocity-compounded stage.


340 TURBINES [ 18-7

A velocity diagram for a two-row stage is shown in Fig. 18-16.


Considering the hypothetical case of a reversible stage with equal
bucket entrance and exit angles in each row (but different angles in
the different rows), the following relations will hold:

yiR = 2/1
- Ffc

ViR — —yut

2/j = 2/2fi + ^6 = — 2/i 2 7b

2/8
= -2/s = yi- 2Vb
ytR = yz- Vb
yzR = —yzR
2/4 = yzR + 1^6 = —2/8 + 2Vb — — 2/l + 4y6
The work on the first row is

F.1 = - (2/1 - y2)Vb =- (22/1 - 2 Vb)Vb (a)


go go

The work on the second row is

W, - (2/8 - y<)Vb =- (22/1 - 678)78 (b)


go go

The total work of the stage is

T7x ^W,i + Wa = - (42/1 - 878)78 (c)


go

Then PTx =- (47c cos a - 878) 78 (d)


ffo
18-7] TURBINES 341

The nozzle-bucket efficiency for the hypothetical stage will be

Plotting efficiency as a function of Vt/V, for nozzle angles of zero


and 20 degrees gives the parabolas of Fig. 18-17. By differentiation
it can be shown tlmt the maximum efficiency for the hypothetical

stage occurs at Vb/Vc = 0.25 cos bt, and is equal to cos' a.

0.Z5

Fig. 18-17. Two-row velocity-compounded stage efficiency; for reversible flow


with symmetrical buckets.

Comparing Eqs. (c) and (e) of this section with the corresponding
equations of the preceding section it is seen that for the reversible
two-row stage with equal bucket angles the maximum efficiency and
work output are the same as for the reversible simple impulse stage
with equal angles. The bucket speed for maximum efficiency, how-
ever, is only half as great for the two-row stage as for the simple
impulse stage. This enables the two-row stage to utilize at maximum
efficiency four times the enthalpy drop that a single-row stage can

M
342 TURBINES [ 18-8

use at wiaYimiiTn efficiency if the bucket speed has the same limit in
both cases; or if the same enthalpy drop is used, the two-row stage

can run at lower bucket speed and lower stresses. In actual ma-
chines, because of the greater complexity of the flow path and the
greater amount of surface involved, the friction losses in the two-row
stage are greater than in the single-row stage; hence the advantage
of the two-row stage is obtained at the cost of a loss of efficiency.
The balancing of efficiency loss against the advantages determines
when the two-row stage should be used. See Sec. 18-10 in this
connection.
18-8 The reaction stage. It is often desirable to operate turbine

Fig. 18-18. lioactioii stages.


1S4] TURBINES 343

stages in such a manner that there is a pressure drop in the bucket


passages as well as in the nozzles; in this case the bucket passages
becoming moving nozzles. A turbine stage operating in this way is
called a reaction stage. The overall pressure drop through a reaction
stage may be divided in various proportions between nozzle pressure
drop and bucket pressure drop. Hero’s turbine, Fig. 18-1, is one
limiting case in which the total pressure drop occurs in the moving
nozzles; the simple impulse turbine may be considered as the other

Fig. 18-19. Reaction stage velocity diagram.

limiting case in which all the pressure drop occurs in the fixed nozzles.
In many modern reaction turbines approximately half the pressure
drop occurs in the moving nozzles or buckets.
The changes of pressure and absolute velocity of the fluid during
flow through a series of reaction stages are indicated in Fig. 18-18.
A velocity diagram for a reaction stage is shown in Fig. 18-19.
In the reaction stage the nozzle jet velocity is determined by the
fluid state before the stage, the exit pressure, the division of pressure
drop between fixed and moving elements, and the velocity coefficient
of the nozzle. The bucket velocity is fixed by the speed and size of

the wheel. Then the other velocities can be calculated if the angles
o and /3 and the bucket velocity coefficient are known.
For example, in a so-called pure reaction stage it is specified that
the isentropic enthalpy drop for the whole stage shall be divided
344 TURBINES [1&-8

(a) (b)

Fig. 18-20. Reaction atage: (a) flow diagram, and (b) state path plot.

equally between the fixed and moving elements. Referring to Fig.


18'20, the inlet statea and the exit pressure p2 are given; the inter-

mediate pressure pi is then fixed by the relation

h* — hi, — ~~
/*j») (a)

Vi = Cv^^goiha - - hu) + r.

yi = Vi cos a

yiR = J/I
- V*

Zl = Fi sin a

ViH

Vuc *C6^2so(hi-hu) + V\^

ViR = - Fjs cos |3

yt = yiR + F»

it = sin fi

Vt

It is possible to show the general derign characteristics of a pure


reaction stage by means of a hypothetical reversible stage having
18^] TURBINES 345

equal angles a and /3, and a negligible approach velocity Fa- For
such a stage

v
^ _ /2go(A.B ht,) Vt
"V 2 ^V2

ViR = yi ——
Vb —
Vt
cos a ——
1^6

By the law of cosines

-
= y? + V? 2ViVi cosa = Y + y?-\^ VtVi cos «

= Vy + y?« = yjn + Vl-V2VtVbCOSa


ViB = - Fji cos i3
= - ViB cos a
The work is

The nozzle-bucket efficiency is

Wt

Plotting efficiency as a function of Vb/Ve for nozzle angles of zero


and 20 degrees gives the curves of Fig, 18-21. The curves show
qualitatively the characteristics of real reaction stages of equal en-
thalpy drop as the design conditions are altered. The maximum
efficiency occurs for a =
0 at Vt/Ve = l/v^, and for useful values
of a at magnitude approaches unity as the
slightly lower speeds; its
nozzle angle and bucket angle approach zero. For any fraction of
isentropic enthalpy drop in the bucket other than five-tenths, ^e
curves are similar but the maximum is at a different value of Vb/Vt,
346 TURBINES [lS-9

Fig. 18-21. Reaction stage efficiency; for reversible flow with equal angles and
equal enthalpy drops in nozzles and buckets.

For any given nozzle angle the maximum efficiency of the hypotheti-
cal reaction stage is greater than that of the hypothetical impulse
stage discussed in the preceding sections. In practice, however, either
type of stage may be better, depending upon the operating conditions,
and each type is used where it is better. The reaction stage in gen-
eral has lower fluid velocities than the impulse stage, resulting in
lower friction losses. The reaction stage has the disadvantage of a
higher ratio of bucket velocity to characteristic velocity than the
impulse stage. Further comparisons are made in Sec. 18-10.
18-9 Multistage turbines. In the preceding sections it has been
shown that for any type a certain bucket
of turbine stage there is
velocity which gives best efficiency for a given enthalpy drop. It
does not require a very large enthalpy drop to make tliis bucket
velocity impractically large. For example, let steam at 150 psia,
saturated, expand to 1 psia; this gives an isentropic enthalpy drop
of about 320 Btu/lb, which gives a characteristic velocity of 4000 fps.
In a simple impulse stage the corresponding bucket velocity for maxi-
mum efficiency would be about 1800 fps. A bucket of uniform sec-
tion, two inches long, travelling 1800 feet per second at a radius of
one foot would be subject to tensile stress of the order of 50,000 psi
from centrifugal force alone; bending stresses would also exist, to say
nothing of vibration effects. No ordinary machine can be operated
under such stresses. If the bucket velocity is kept low to make the
stresses acceptable, the nozzle-bucket efficiency will fall off rapidly
as the enthalpy drop increases. Since the more efficient power cycles
18-»] TURBINES 347

involve large enthalpy drops, the single stage is unsuitable for efficient
power plants.
One way of obtaining a lower bucket velocity at maximum effi-

ciency is to use a two-row stage, but the two-row stage is relatively


inefficient and in any case will itself reach a limit of velocity.
If a number of stages are arranged in series, the enthalpy drop
may be divided among them so that the velocity for each stage is not
excessive. If equal enthalpy drops are taken in all stages, the single-
stage characteristic velocity will be divided by the square root of the
number of stages used; thus to reduce the velocity by half requires
four stages.
The use of multiple stages not only permits operation at the speed
of maximum efficiency but also permits the attainment of higher effi-

ciency than a single stage would give. First, in the individual stages

the reduction of velocity results in smaller friction losses, since the


friction losses are in general proportional to the square of the velocity.
Second, by reducing the general level of velocity in the turbine the
wasted kinetic energy at the exhaust end of the turbine (leaving loss)

is reduced. Third, with small pressure differences between stages the


leakage of fluid from stage to stage without doing work is reduced.
Fourth, the efficiency of a multistage turbine is higher than the effi-

ciency of its stages because of the reheat effect described below.


The efficiency of a stage, accounting for all losses, may be de-
fined by
Ws
vs (18)
ha /l2»

where W's is the work done on the shaft per pound of fluid flowing
through the stage and ha — fh, is the isentropic enthalpy drop for
the stage. work done on the buckets both be-
H's differs from the
fluid, and because
cause of friction between the rotating parts and the
of the leakage flow which does no work. Applying the steady flow
energy equation to an adiabatic stage
1'2
* a
= ha — hi
j_2
2gv 2go

In general, however, the velocities entering and leaving a stage will


be substantially equal, so it is permissible to write

IF5 = Ao — hi
348 TURBINES [18-9

Substituting from (18),

ho ~ hi = ris(ha — ht,)

where the states referred to are


plotted in Fig. 18-22.
When several stages are
arranged in series the expansion
of the fluid may be plotted on
the enthalpy-entropy plane as
shown in Fig. 18-23 for a three-
stage turbine. The expansion
in the first stage is from the
initial state a to the state 1,

Fig. 18-22. at pressure pi. State 1 is

located by the relation

hi “ ha vs(ha — hu)

The expansion in the second stage starts at 1 and goes to 2. State 2


is located by the relation

Fig. 18-23. Three-Rtage atate path.

hi = hi — Tfsihi — hi)

where state 2' is at the pressure of state 2 and the entropy of state 1.

The expansion in the third stage goes from 2 to 3, and state 3 is

located in similar manner to state 2.


T )

18-9] TURBINES 349

The work of the turbine Wtib given by the sum of the works of
the stages. Assuming uniform stage efficiency,

Wt = fjsiha hu + "" ^2
+ h2 — A3)

The turbine work may also be written in terms of the turbine internal
ijt, and the
efficiency y* overall isentropic enthalpy drop thus:

TV ” Aa As — 7JT (Jla Ast

Then the ratio of turbine efficiency to stage efficiency is

ij 7» hg — hu "V hi — A2 ~~h A2 — A3

rfs ha — hu
Now it is a property of vapors that the isentropic enthalpy drop
between two given pressures becomes greater as the entropy increases;
that is, the constant-pressure lines on the h-s diagram diverge toward
the right. Consequently Ai^A^ is greater than Ai,“A2 *, and A^-AJ is

greater than Jhrhza] then the sum of the individual isentropic enthalpy
drops for the stages is greater than the overall isentropic enthalpy
drop for the turbine. The ratio of these quantities, which is equal
to ftr/rts, is called the reheat factor. For efficient turbines of many
stages as commonly used in power plants, the reheat factor may be
of the order of 1.05; for less efficient turbines it may be greater or less
depending on the number of stages.

Example 2. A three-stage impulse turbine is to operate between an


initial state of 100 psia, saturated steam, and an exhaust pressure of 1.0 psia.
The intermediate pressures, pi, (Fig. 18-23), are fixed by dividing the
overall isentropic enthalpy drop into equal steps. The stage efficiency is

65 percent for each stage. The approach velocity is negligible. Find the
internal efficiency and the reheat factor for the turbine.

Solution: Find the intermediate pressures.

ha = 1187.2 Btu/lb from tables

«a = 1.6026 Btu/lb®R

At 1.0 psia hu = 895 Btu/lb

Dividing the enthalpy drop into three parts

hu = 1090 Btu/lb Pi == 28.0 psia

A2. == 993 Btu/lb p2 = 6.2 psia


• The internal efficiency is based on work delivered to the turbine shaft; it

does not account for bearing losses or work to drive oil pumps or other auxiliaries.
350 TURBINES [18-10

Then h\ 1187 - (.65)(1187 - 1090) = 1124 Btu/Ib

9i 1.650 Btu/lb®R

hi 1024 Btu/lb

ht 1124 - (.65) (1124 - 1024) = 1059 Btu/lb

82 1.706 Btu/lb^R

hi 953 Btu/lb

hz 1059 - (.65) (1059 - 953) = 990 Btu/lb

Wt ^ 1187 - 990
^ Q g-.
Vt -
ha - hz. 1187 895

0.675
Reheat factor == 1.04
0.65

18-10 Comparison of turbine types. The relations presented


above for the efficiency of three types of stages are design relations,
not operating relations. The operating characteristics of a turbine
depend upon the load characteristics. A load such as a centrifugal
pump or a fan has torque-speed characteristics which match those of
a turbine running at a fixed ratio of Vb/Vc] thus if the turbine is
properly chosen to suit the load it may always run close to its best
value of Vh/Ve (if not properly chosen it may always run at very
poor efficiency). An electric generator load requires that the turbine
run at constant speed for all loads; this necessitates some variation
in the ratio Vt/Vc and a consequent reduction in efficiency at loads
higher or lower than the design load. Consideration of Figs. 18-1^,
17, and 21 appears to show that the effect of varying the ratio Vb/Ve
is less harmful in the reaction stage than in the impulse stages.

However, in turbines as actually built there does not appear to be


any decisive difference between impulse and reaction machines with
respect to efficiency characteristics.
The various types of stages have certain characteristics which
determine their applications. The reaction stage, having large leak-
age areas, is best used where the fluid passages are large and pressure
differences small so that the leakage will not be an excessive portion
of the total flow. Hence reaction stages are used in the low-pressure
sections of large turbines. Impulse stages, on the other hand, have
small leakage areas and are therefore better adapted to small turbines
and to the high-pressure sections of large turbines. Two-row stages
are superior to simple impulse or reaction stages in one respect, the
ability to utilize large enthalpy drops at a reasonable efficiency. For
18-10] TURBINES 351

Fig. 18-24. Cross section of a fourteen-stage impulse turbine. Between the


eighth and ninth stages is an extraction opening through which steam may flow
to supply a heating load. Steam not required for heating flows through the last
six stages to the condenser. (Courtesy General Electric Co.)

this reason two-row stages have two fields of application, (1) in single-

stage turbines and (2) as the first stage of multistage turbines. In


small turbines it is undesirable to go to the complication of multiple
stages because if stages Avere built small enough to handle the small
flow at high pressure, they would be inefficient because of the small
flow passages. Also multiple disc construction would involve larger
disc friction losses than single disc construction. Consequently the
single two-row stage is preferred over the equivalent four single-row
stages for small machines. In larger multistage turbines use of a
two-row stage as a first stage permits elimination of several relatively
inefficient stages from the high-pressure end of the turbine, thereby
reducing the shaft length, which may be of great benefit to the
mechanical design. Moreover, the use of the two-row stage may
make possible appreciable reduction of the maximum pressure and
temperature to which the turbine casing is exposed, a matter of spe-
cial importance in modern high-temperature matdiines. The effi-
ciency of a two-row stage is inherently lower than that of a single-row
stage because of the greater surface exposed to fluid friction effects,

354 TURBINES [18-11

and because of the losses connected with transferring the fluid three
times between stationary and moving passages. The lower efficiency
can be accepted in a first stage followed by 15 or 20 efficient single-row
stages, even though it would be unacceptable for all the stages oS a
turbine designed for high efficiency.
18-11 Turbine characteristics. The efficiency characteristics

of turbines may be expressed in several ways; as fluid flow vs. output,


fluid rate vs. output, or efficiency vs. output.

FRACTION OF RATED LOAD


Fig. 18-27. Steam flow vs. load for typical turbine generators. (From data of
publication GEA-3277B, General Electric Co.)

The fluid flow vs. output relation is approximately a straight line,


as shown in Fig. 18-27, for ranges of output in which the flow path
is substantially the same. When overload valves are opened, by-
passing some stages of a multistage turbine (or opening more nozzles
18-11] TURBINES 355

in a single-stage turbine), a different straight line is followed, as indi-


cated. The name WiUans line is applied to the straight line plot.
The fluid rate of a turbine is the ratio of fluid flow to the turbine
divided by the output, fpr example pounds per kilowatt-hour. The
efficiency of a turbine is the ratio of work output per pound of fluid
flow divided by the overall isentropic enthalpy drop for the turbine.
Reference to Chaps. 14 and 15 will show how the turbine eflSciency
is utilized in cycle calculations. Some examples of turbine steam
rate curves and efficiency curves are given in Fig. 18-28.
General levels of efficiency for modem steam turbines operating
at best efficiency are shown for several sets of conditions in Fig. 18-29.
In general, large turbines are more efficient than small turbines.

Turbine stages operating with wet steam suffer a loss in efficiency


of more than one percent for each percent of moisture present. This
fact and the erosion of blading by excessive moisture account for the
use of the reheat cycle as explained in Chap. 15.

Turbines as a class of machines are best adapted to large power

FRACTION OF RATED LOAD


Fig. 18-28. Typical turbine generator characteristics. (From data of publica-
tion GEA-3277B, General Electric Co.)
356 TURBINES [18-11

10 20 25 30 ^5 40 45 50 60
RATING 1000 KW
Fig. 18-29. Turbine generator efficiency as a function of rated capacity and
steam conditions. Turbine efficiency is 1 or 2 percent higher than turbine gen-
erator efficiency. (Courtesy Westinghouse Electric Corp.)

and high-speed applications; thus they are universally used for driv-
ing large electric generators. The problem of driving low-speed ma-
chinery by means of turbines has been solved by the development of
reduction gears of high efficiency such as used with ship propulsion
turbines. The turbine is well adapted for use with the regenerative
steam cycle because it permits convenient extraction of steam for feed
water heating at numerous temperature levels. For small power the
turbine is less efficient than a reciprocating engine but small turbines
are frequently used because high speed is desired or because of their
and ease of operation and upkeep.
simplicity, small size,
Steam turbines have reached a highly satisfactory state of devel-
opment. Gas turbines are not in such a satisfactory state because,
as was shown in Chap. 16, they must work at extremely high tem-
peratures and efficiencies in order to be able to compete in plant
efficiency with modern steam or diesel plants. However, despite
lower efficiency, the gas turbine plant will probably replace many
diesel engines because of the inherent advantages of rotating machines
over reciprocating machines.
TURBINES 367

PROBLEMS
18-1. In a rocket the driving fluid starts from rest, relative to the rocket,
and is accelerated out through the tail pipe or nozzle. In a certain rocket
2 Ib/sec of combustion products ore discharged at a relative velocity of 4000 fps.
Find the thrust on the rocket.
18-*2. A flexible fire hose 2.5 inches in diameter runs through a 90-degree

bend to a nozzle having an exit diameter of 1.00 in. Water flows through the
hose at the rate of 0.341 cfs. How much force must the fireman exert, parallel
to the axis of the nozzle, to hold the nozzle? In which direction must he exert
the force?
18-3. Referring to Problem 18-2, for each case listed below, what is the mag-
nitude of the total stress on the connection between the nozzle and the hose,
and is it tension or compression? Case (a), the fireman holds the nozzle itself.
Case (b), the fireman holds the hose adjacent to the nozzle. Neglect gravity
forces.
18-4. Steam enters a turbine bucket passage at 1400 fps absolute, with a
nozzle angle of 14 degrees. The bucket velocity
is 600 fps. The bucket velocity
coefficient is unity, no pressure change through the bucket passage;
and there is

the bucket exit angle is 32 deg. Find, per pound of fluid flow per second: (a) the
power supplied to the buckets; (b) the tangential force on the buckets; (c) the
axial force on the buckets.
18-5. Air enters a turbine bucket passage at 20.0 psia, lOOO^’F, 850 fps abso-
lute velocity, with a nozzle angle of 20 deg. The pressure drop through the
bucket passage is 2.00 psi, the bucket velocity is 650 fps, and the bucket exit
angle is 20 deg. (a) If the bucket velocity coefficient is unity find the power
supplied to the buckets per pound of air flow per second, (b) If the area factor
at the bucket exit is 0.85, and the bucket annulus area is the same at inlet and
exit, find the axial thrust on a section of bucket annulus which would pass

1.0 Ib/sec of air.


18-6. In an impulse turbine stage the nozzle exit velocity is 1200 fps and

the nozzle angle is 14 deg. The bucket exit angle is25 deg and the bucket
velocity coefficient is 0,90. If the bucket velocity is 600 fps find: (a) the work
done per pound of fiuid flow; (b) the kinetic energy per pound of fluid in the
stream leaving the buckets; (c) the angle between the absolute velocity leaving
the buckets and the tangential direction. Sketch the velocity diagram.
18-7. For the conditions of Problem 18-6, assuming the bucket velocity
coefficient stays constant, find the bucket velocity for maximum work (assume
bucket velocities, find the work, and make a plot). At this bucket velocity find
the angle of the relative fluid velocity entering the bucket.
18-8. In an impulse air-turbine stage the air supply is at 105 psia, 80^F,
negligible velocity, and the exhaust pressure is 15 psia. The nozzle angle is
15 deg and the bucket exit angle is 30 deg. The pitch diameter of the wheel
is 8 m. The nozzle velocity coefficient is 0.96, and the bucket velocity coefficient
is 0.88; the bucket velocity is half the tangential component of the nozzle exit

velocity. Find: (a) the turbine speed (rpm); (b) the nozzle-bucket efficiency;
(c) the nozzle exit area required to do work on the buckets at the rate of 100 hp;
(d) the temperature of the air leaving the buckets.
358 TURBINES
18-9. (a) In the hypothetical reversible two-row stage with equal angles^ at
the speed of maximum efficiency, what fraction of the total work is done on each
row? a similar three-row stage, at the speed of maximum efficiency, what
(b) In
fraction of the total work would be done on each row?
18-10. Steam is supplied to a turbine at 1300 psia, 900°F. The first stage
wheel is 3 ft* in pitch diameter and the speed is 3600 rpm. Taking the nozzle
angle as 14 deg and the nozzle velocity coefficient os 0.95, estimate the pressure
and temperature of the steam in the first stage bucket passages for two cases:
(l) a single-row impulse stage; (b) a two-row impulse stage. Each stage is de-
signed to operate close to maximum efficiency.
18-11. In a two-row stage the nozzle angle is 14 deg, the nozzle exit velocity
is2500 fps, and the bucket velocity is 600 fps. The buckets and the intermediate
guide vanes are symmetrical; the angle in each row is equal to the angle of the
relative velocity entering the row. If the velocity coefficient of each row (mov-
ing and stationary) is 0.92 find: (a) the bucket and guide vane angles; (b) the
power supplied to the buckets at a steam flow rate of 100,000 Ib/hr.
18-12. In a hypothetical reversible reaction turbine stage with zero angles
the enthalpy drop is f in the nozzle and i in the bucket passage. Find the
ratio of bucket velocity to characteristic velocity for maximum efficiency.
18-13. A reversible reaction stage has equal enthalpy drops in the moving
and stationary rows. Air enters the stage at 1500®F, 30 psia, 500 fps, and leaves
at 20 psia. The nozzle and bucket passages have the same exit angle, 25 deg,
and the same area factor at exit. The speed is such that the relative velocity
entering the buckets is at 90 deg to the bucket velocity.
Find: (a) the pressure at the nozzle exit; (b) the bucket velocity; (c) the abso-
lute velocity leaving the buckets ;
(d) the nozzk^-bucket efficiency (can you explain
why this is higher than for the hypothetical stage of Sec. 18-8?); (e) the ratio
of bucket velocity to characteristic velocity for the stage; (f) the power supplied
to the buckets per pound per second of air flow; (g) the ratio of bucket exit area
to nozzle exit area (this indicates how much the radial length of the bucket exit
must exceed that of the nozzle exit) ;
on the buckets, assum-
(h) the axial thrust
ing that the passage area is a flow rate of 100 Ib/sec, the area factor
sufficient for
at the bucket exit is 0.90, and the total area exposed to pressure on each side
of the bucket row is 1.1 times the bucket annulus area at the exit.
18-14. A steam turbine is to operate between 600 psia, 850®F, and 1.5 in.
HgabB. The bucket velocity is to be 600 fps and the .average nozzle efficiency is
expected to be 95 percent, except for a tw’o-row stage, for which it will be 90
percent. Nozzle angles will be assumed as 15 deg for impulse stages and 25 deg
for reaction^tages. All stages operate close to the speed of maximum efficiency.
Estimate the number of stages required for each of the following arrangements:
(a) all simple impulse stages; (b) all 50 percent reaction stages (50 percent of Ah
in the nozzle); (c) a two-row first stage followed by simple impulse stages; (d) a
two-row first stage followed by 50 percent reaction stages,
18^15. A steam turbine working between 250 psia, 700^F, and 30
five-stage
psia, has equal enthalpy dropsin all stages, and the stage efficiency is 75 percent
in all stages. Plot on the mollier chart the condition line^ i.e. the line through
the state points for the steam entering and leaving the stages; find the interstage
TURBINES 359

pressures, the reheat factor, and the turbme intemai efficiency, using properties

from the chart.


Chapter i9

RECIPROCATING EXPANDERS
AND COMPRESSORS

Reciprocating machines, or cylinder and piston machines, are widely


used in heat engineering processes as expansion engines and as pumps
and compressors. Their particular field of usefulness, in contrast to
turbine machineiy, is in the range of small capacities and high pres-
sures. Cylinder and piston machines belong to the class of positive
displacement machines, which also includes gear pumps, vane pumps
and similar machines. Positive displacement machines are classified
as engines when the net work is positive, and as pumps or compressors
when the net work is negative.
19*1 Cylinder and piston arrangements. In a simple recipro-
cating machine. Fig. 19-1 the working Buid
,
is enclosed in a cylindrical

CYLINDER
HEAD-
''PISTON ROD

CYLINDER
VOLUME CYLINDER WALL

Fig. 19-1. Reciprocating machine cylinder.

space by the cylinder walls, a stationary cyhnder head, and a moving


piston, often sealed by rings fitted closely to the cylinder walls. The
piston is usually linked to a rotating crank-shaft to cause periodic
travel; tlie distance travelled is called the stroke (Fig. 19-2). The
360
19-2] EXPANDERS AND COMPRESSORS 361

Fig. 19-2. Cylinder nomenclature.

size of such a machine is given in terms of its bore (cylinder diameter)


and its stroke. The minimum cylinder volume is called the clearance
volume,and the volume change during the stroke of the piston is called
the 'pisUm dieplacemerU. There must in general be some clearance
volume in a machine to avoid mechanical interference. The thermo-
dynamic aspects of clearance will be discussed later in relation to the
particular types of machines.
Reciprocating machines are said to be either single-acting or double-
acting, according as one or both sides of the piston form the boundary
of a working cylinder. In Fig. 19-3 are shown typical arrangements.
The trunk piston construction, with the connecting rod pinned di-
rectly to the piston, is almost always used for single acting machines,
while the piston rod and cross-head construction is used with double-
acting machines to permit a pressure-tight enclosure of the crank end
of the cylinder.
Some common cylinder arrangements other than the horizontal
arrangement of Fig. 19-3 are sketched in Fig. 19-4. The vertical and
the angle arrangements are used with double-acting cylinders as well
as single-acting. All arrangements shown may have several units on
a single shaft. All are used for internal combustion engines. The
horizontal, vertical,and angle arrangements are used for steam en-
gines and for air compressors and refrigeration compressors.
19-2 Ports and valves. Almost without exception reciprocating
machines are used in connection with flow processes, so means must
be provided for permitting flow into and out of the cylinder as desired.
362 EXPANDERS AND COMPRESSORS [19-2

SINGLE-ACTING MACHINE

HEAD -END CYLINDER VOLUME

DOUBLE-ACTING MACHINE
Fig. 19-3. Machine nomenclature.

The passages through which the fluid enters and leaves, called porfe,
may be opened and closed by vdves or by the passage of the pistoq
over the port openings.
In engines the valves must in general be operated by a mechanism
in order to connect the cylinder to the inlet and exhaust pipes at the
proper times. In compressors and pumps it is often possible to use
autohiatic valves which open by fluid pressure to permit flow in the
desired direction, and close to prevent reverse flow. Illustrations of
types of ports and valves may be seen in Chaps. 20 and 21.
IM] EXPANDERS AND COMPRESSORS 363

Fig. 19-4. Cylinder arrangements.

19-3 Work in a reciprocating expansion engine or com-


pressor: indicator card analysts. The expansion engine takes fluid
from a high-pressure region, expands it to a low pressure and then
discharges to a low-pressure region; the compressor carries out the
reverse of this process. Machines like the internal combustion engine
operate alternately as compressors and engines, and are not included
in the categories discussed in this section.
The following indicator card analysisis based upon the idealized

reversible adiabaticmachine in which the states of the fluid entering


and leaving the machine are constant with time and the kinetic and
potential energy entering and leaving are negligible. Let the ideal-
ized machine of Fig. 19-5 operate as an engine. The fluid is available
in the supply pipe at state 1, for which the pressure is pi. The
machine cycle (not a thermodynamic cycle, be it noted) starts at
point a on the indicator diagram, with the cylinder volume at mini-
mum and the cylinder pressure at pi. The following machine proc-
esses are then carried out.

(1) Admission. The inlet valve opens at a and, as the piston


moves out to point 6, fluid at constant pressure flows into the cyl-
inder. This is a simple displacement in space with no change in the
state of the fluid.

(2) Expansion. The inlet valve closes at b and as the piston


completes its stroke to point c, the flmd in the cylinder expands
reversibly and adiabatically to the exhaust pressure ps.

(3) Exhaust. The exhaust valve opens at c and, as the piston


364 EXPANDERS AND COMPRESSORS [IM

INLET VALVE

^ PISTON TRAVEL OR
CYLINDER VOLUME
(b)

Fig. 19-5 . Idealized machine cycle: (a) flow diagram; (b) indicator diagram.

moves back to point d, fluid flows out of the cylinder. This is a


simple displacement in space with no change in the state of the fluid.

(4) Compression. The exhaust valve closes at d and, as the pis-


ton completes its stroke to point a, the fluid in the cylinder is com-
pressed reversibly and adiabatically to the inlet pressure pi.

The mass of fluid in the clearance space at point a is called the


clearance fluid me. The mass of fluid taken in during the process
a-b and expelled during the process c-d is called the flow fluid m/.
Then during the expansion mass
process b-c the cylinder contains a
of fluid (me + m/). During the admission and exhaust processes the
mass in the cylinder is varying.
The work of the ideal engine is given by the area of the indicator
19-4] EXPANDERS AND COMPRESSORS 365

diagram, which may be written in terms of pressure and cylinder


volume V as the integral of p dV around the machine cycle. This
integral will be evaluated in terms of p and V in Chap. 21, for the
gas compressor. At this point, however, a more general approach
will be used. The work done by the fluid in one machine cycle is
given by the sum of the work quantities for the processes:

W=W + We-d + (a)

Since the pressure is constant in process a-b,

Wa-i = Pl(n - Va)

but Vt = (wi/ + m,)vi and Va =


so Wa-i = PlViTTlf (b)

This is the flow work done in Ailing the cylinder. Since the process
b-c is adiabatic,

but Ui = (m/ + TOe)ui and Ve — im; + inc)'Ui


so WiH-, = (m/ + jnc)(Mi - Ui) (c)

This is the expansion work of the total fluid in the cylinder. Similar
reasoning applies to processes c-d and d-a.

Wc~d = -ptVtmf (d)

this is the flow work done in emptying the cylinder.

= -mdui - U2) (e)

This is the compression work of the clearance fluid. Then substi-


tuting in (a),
W = m/ipiVi - + ui- ptVt v^)

or W — mf(hi — h) (1)

Equation (1) applies not only to the reversible adiabatic engine


described, but also to a compressor in which the reversed operation
is carried out.
The mean effective pressure (pm or mep) of a reciprocating ma-
chine is by the ratio of indicated work to piston displacement.
given
19-4 Steady flow analysis. The work of the reciprocating ma-
366 EXPANDERS AND COMPRESSORS [19-6

EXHAUST OR
DISCHARGE
Fig. 19-6. Steady flow analysis.

chine under steady flow conditions may be obtained from the steady
flow energy equation. Referring to Fig. 19-6, assuming negligible
changes of kinetic and potential energy, •

+Q= ^2 + Wx
Then in general
Wz ^ hi — ^2 4“ Q (2)

and for an adiabatic machine

H
II 1 (3)

Noting that m/ in Eq. (1) is the fluid flowing through the machine
per machine cycle, and W in Eq. (1) is the work per machine cycle,
it is clear that (1) and (3) are equivalent. Thus the steady flow
analysis shows that under steady flow conditions Eq. (1) is true not
only for reversible adiabatic machines but for any adiabatic machine.
From (3) the an adiabatic machine may be measured
work of
solely in terms of the states of the fluid entering and leaving.
19-5 Machine efficiency. The efficiency of a machine from the
thermodynamic viewpoint serves to compare the actual machine to
a specifled idealized machine. This is not the same concept as that
of mechanical efficiency, as is shoTvn below.
The definition of the efficiency of a reciprocating machine depends
upon whether the machine is an engine or a pump.* For an engine
* These definitions were given in Chap. 1 4 in connection with the study of
power cycles. See Eqs. (14-7) and (14-13).
19-6] EXPANDERS AND COMPRESSORS 367

the engine efficiency is the ratio of the actual engine work to the work
of a reversible adiabatic engine when both engines are supplied with
fluid at the same state and both exhaust to a space at the same
pressure.

(4)

where tie is engine efficiency, is the shaft work per pound of flow
fluid, hi is the enthalpy of the fluid supplied, and hi, is the enthalpy
of the fluid after an isentropic process from the initial state to the
final pressure. For a pump or compressor the work is an input rather
than an output, therefore the definition of pump efficiency (or com-
pressor efficiency) is the inverse of the definition of engine efficiency.
For the pump or compressor

rjp = ( 5)

It may be observed that both the numerator and denominator in


this case are negative.
If an actual machine is adiabatic the shaft work will be given for
steady flow conditions by

W, = hi- hi

assuming potential and kinetic energy changes negligible. Then for


adiabatic machines only the efficiency of an engine is given by

hi — hi
" Ai - hi.

and of a pump by
hi — hu
hi — hi

Thus in the cftse of adiabatic machines the efficiency is solely a func-


tion of the fluid properties enteringand leaving the machine.
The efficiency may be obtained on more than one basis depending
upon where the control surface is located with reference to the meas-
urement of Wx- If Wx is measured as the work transferred between
the fluid and the piston of the machine, Wx is called indicated work*
and the efficiency is the indicated efficiency. If Wx is measured at

* See discuseion of the indicator disLgram, Sec. 2-5.


368 EXPANDERS AND COMPRESSORS [m
the shaft coupling of the machine, TT, is called brake work and the
efficiency is the brake efficiency.
The difference between brake work and indicated work is work
used in overcoming mechanical friction of the machine parts, The
mechanical efficiency of an engine is defined by

_ brake work brake efficiency


’Jmech (engina) — "" ( 6)
jQ^Qg^ted work indicated efficiency

For a piunp or compressor the inverse ratio applies since the flow of
work through the machine is reversed.
i ndicated work _ indicated efficiency
^mech (pump)
brake work ~ brake efficiency ( 7)

19-6 Flow capacity of reciprocating machines. The flow ca-


pacity of a machine is the rate at which it will pass fluid under some
given conditions.* For the idealized machine of Fig. 19-5 the TnaM
of fluid passed per unit time is ni/N, where N
is the number of ma-

chine cycles per unit time. The mass nt/ which the idealized machine
can pass per machine cycle is determined by the specific volume of
the fluid and the volume characteristics of the machine, namely the
piston displacement and the clearance volume (Fig. 19-2). Eeferring
to Fig. 19-7, the piston displacement is given by

Fig. 19-7.

* Engines have a power capacity which is a rate of work output; the power
capacity is related to the flow capacity through Eq. (1).
l»-6] EXPANDERS AND COMPRESSORS 369

(PD) = F. - y,
the clearance is given as a fraction of the piston displacement by

Cl--^
“ (PD)
For an engine, state 2 is at the lower pressure and the mass of flow
fluid in the cylinder is given by

m/
Vc-Vi
Vi

but Fe (1 + C1)(PD)

and Va F»- = Cl(PD)-


Va Vi

Then for an engine

m.f ( 8)

For a compressor the inlet and outlet states are interchanged so

Thus it is apparent that the flow capacity machine


of the idealized
is directly proportional to the piston displacement, and inversely pro-
portional to the maximum specific volume of the fluid, if the fractional
clearance of the machine and the ratio of specific volumes remain
constant. Since the specific volume ratio in both (8) and (8a) is

greater than unity, the flow capacity will increase with decreasing
clearance, other conditions being constant. Also the flow capacity
will increase with decreasing volume ratio.
specific ,

It should be noted that Eqs. (8) and (8a) are valid not only for a
reversible adiabatic machine but for any reciprocating machine in
which the state of the fluid is constant during admission and exhaust
and the compression and expansion processes have the same state
path. Although the capacity of an actiuil machine will be different
from the capacity of an idealized machine, the equations indicate the
manner of variation of the actual flow capacity with changing condi-
tions. Thus to obtain the maximum capacity with a given size of
machine the clearance should be kept to a minimum This is par- -

ticularly important for machines operating through a large pressure


ratio since that implies also a large specific volume ratio.
;

370 EXPANDERS AND COMPRESSORS [IM

In a real engine the intake process is often carried beyond the


ideal end point b, Fig. 19-7. The expansion process is at a higher
pressure than fr-c, and is followed by an abrupt drop to the exhaust
pressure at the end of the stroke. In this way the work and the
mean effective pressure may be increased, but the efficiency decreases.

Example 1. A reversible adiabatic steam engine is to receive saturated


steam at 90 psia and to exhaust at 15 psia. The machine is double-acting
and has piston displacement of 1.0 cu ft at the head end and 0.98 cu ft at
the crank end (the piston rod occupies some volume at the crank end). The
fractional clearance at each end is 0.05. The speed is 300 revolutions per

minute (one machine cycle per revolution). Determine the horsepower out-
put and the mean effective pressure.

Solution: The work per pound of steam is

Wx — h\ —
From the tables, noting that this is a reversible adiabatic machine,

hi = 1185.3 Btu/lb Vx = 4.896 cu ft/lb

hi = 1054.3 Btu/lb vi = 23.66 cu ft/lb

Wx = 131.0 Btu/lb

The flow fluid is obtained by substitution in (8)

f
fi nnr/23.66

= 0.067 Ib/machine cycle

The flow rate is then

m/N = (3(K)) (0.067) = 20.1 Ib/min

= W. (Ib/min) _ (131.0)(20.1)
62.2 hp
2545/60 42.42

mep
W 62.3(33,000)
= 3460 psf = 24.0 psi
(PD) 1.98(300)

In an actual engine the i)ower output and mep would be somewhat greater
than here calculated, but the steam flow rate w’ould be much greater due to
condensation of steam and incomplete expansion as explained in the next
chapter.

Example 2. In a certain air compressor the air enters at 15 psia 80‘*F


and leaves at 60 psia 260®F. The clearance of the compressor is 3 percent
of the piston displacement. If the dimensions and speed of the machine
19-7] EXPANDERS AND COMPRESSORS 371

are such that the piston displaces 150 cu ft per minute, how much air would
the corresponding idealized reversible adiabatic machine pass per minute?
If the clearance were increased to 6 percent how much air would be passed?
SoLtniON:

RTi (53.34) (540)


13.35 cu ft/lb
Pi
“ (15)(144)

Si _ ^ £s (540) (60)
= 3.00
Pi Ti (15)(720)

JVm/ = - 0.03(3.0 - 1)] = 10.6 Ib/min

For clearance of 6 percent


1

[1 - 0-06(3.0 - l)j = 9.9 Ib/min


ld.35

19-7 Liquid pumps. When a reciprocating machine is used to


pump a liquid, assuming the liquid incompressible, *>2 = Vi, and the
flow capacity is

iV(PD)
Nnif = (9)
V

independent of clearance. For this reason no particular necessity


exists for minimizing the clearance of a liquid pump, and construc-
tions which would be entirely impractical for compressible fluid pump-
ing are quite satisfactory for liquids. On the other hand the greater
density of liquids makes it more difficult to accelerate them so as to
follow the motion of the piston. The pressure difference necessary
to accelerate the fluid from the suction pipe as it enters the cylinder
plus the pressure necessary to overcome friction may exceed the abso-
lute pressure in the supply pipe if the piston moves too fast. Then
the liquid will not follow the piston closely, but as the pressure drops
below the saturation pressure for the e-xisting temperature, vapor
pockets or cavities will form in the liquid. Later, as the piston
acceleration is reduced, the liquid Avill overtake the piston, increasing
the pressure, and the cavities wll collapse as the vapor condenses.
This formation and collapse of vapor pockets, called caviUUum, results
in serious shocks to the pump parts, and also reduces the capacity.
In order to reduce the probability of cavitation, reciprocating liquid
372 EXPANDERS AND COMPRESSORS
pumps require large ports and valve openings to reduce friction, and
they must be operated at limited speeds to enable the fluid to follow
the motion of the piston.

PROBLEMS
19*1 . A
double-acting reciprocating expansion engine has a 6 in. bore and
9 in. stroke; the piston rod, of 1 in. diameter, passes through one end of the
cylinder. The clearance at each end is 5 percent of the piston displacement at
that end. Find: (a) the maximum cylinder volume at each end; (b) the mini-
muni cylinder volume at each end; (c) the total piston displacement per minute
at 400 rpm.
19-2. An engine receives air at 105 psia, 140^F, and exhausts at 15 psia.
The air flow rate is 8.5 Ib/min and the brake output is 6.0 hp. If the mechanical
efficiency 82 percent find the engine efficiency based upon indicated output.
is

19-3. The engine


of Problem 19-1 is to be supplied with air at 120 psia,
400'’F, and to exhaust at 20 psia. On the basis of the idealized reversible-
adiabatic engine process calculate the air flow rate and the indicated horsepower
for an engine speed of 400 rpm.
19-4. Solve Problem 19-3 on the basis of an idealized engine process, Fig. 19-7,
having identical state-paths for expansion and compression, according to the
relation » constant. (Note that the indicated power must be obtained
from the integral of p dv on the indicator diagram, not from Ah, since the process
is not stated to be adiabatic.)

19-5. A refrigeration compressor has two single-acting cylinders of 3.00 in.


bore and 3.00 in. stroke. The clearance is 3 percent of the piston displacement.
It desired to compress 3 lb per minute of Freon-12 from 14 psia, 0®F, to 140
is

psia. On the basis of an idealized reversible-adiabatic compressor process find


the speed (rpm) required.
19-6. The compressor of Problem 19-5 is to be used to compress helium from
14 psia, 80"F, to 140 psia. On the basis of a reversible adiabatic process how
many pounds per hour would be compressed at a compressor speed of 1000 rpm?
Find the indicated horsepower.
19-7. Solve Problem 19-6 if the expansion and compression processes are
reversible polytropics according to the relation pi^'^ ^ constant. The com-
pressor is water-cooled.
19-8. Solve Problem 19-6 if the fluid is nitrogen.
Chapter 20

THE STEAM ENGINE

The steam engine, having been the basis of the industrial revolution
and the main inspiration for the original development of thermo-
dynamics, is of considerable historical interest. Moreover it may be
a useful academic vehicle for the illustration of some practical appli-
cations of thermodynamics. As a source of power, however, the
steam engine is now of distinctly minor importance.
20-1 The slide-valve steam engine. The slide-valve steam en-
gine, Fig. 20-1, is a double-acting engine with a single valve for
controlling the inlet and exhaust of steam at both ends of the cylinder.

ECCENTRIC (VALVE CRANK)

Fig. 20-1. Slide valve engine.

The valve motion comes from an eccentric (a crank of short radius)


on the engine shaft, displaced angularly from the piston crank as
shown. At the position showm, the piston is at one extreme of its
travel, the head-end dead point, and the valve is slightly open to admit
steam to the head end of the cylinder. When the shaft turns in the
373
374 THE STEAM ENGINE [20-1

indicated direction evident that the valve will open further to


it is

admit steam more freely as the piston moves on its stroke.


When the crank-shaft has turned about 75 degrees from the posi-
tion shown, the valve will be at its extreme open position and the
piston will be nearly at the middle of its stroke. Before the piston
reaches the end of its stroke the valve will have returned and closed
off the steam supply to the head end. Just before the piston reaches
the crank-end dead point the valve will open to the exhaust pipe
from the head end of the cylinder. During most of the return stroke
the head end will be connected to the exhaust, but near the end of
the stroke the exhaust will be closed; when the piston reaches the

head-end dead point the steam connection will again be open. A


end of the cylinder.
similar series of events occurs at the crank
The indicator diagram for the head end of the cylinder would
look like Fig. 20-2. The plot may be divided into four segments
corresponding to the four processes of admisaion, expansion, exhaust,
and compression, as in the case of the idealized engine of the preceding
chapter.
The valve events which mark the limiis of the four processes are:
admission. A, at which the steam valve opens; cutroff, C, at which
the steam valve closes; release, R, at which the exhaust valve opens;
and comp^omon, K, at which the exhaust valve closes.
The work done on one face of the piston during one revolution of
20-2 ] THE STEAM ENGINE 375

the engine is given to scale by the area of the diagram as explained


in Sec. 2-5. The total work for one revolution is the sum of the
works for both ends of the cylinder.
20>2 Comparison of real and idealized indicator diagrams.
In Fig. 20-3 an indicator diagram for a real engine is superimposed
upon a diagram for the corresponding idealized engine. Some of the
reasons for the differences between the diagrams are given below.

Throttling. The pressure during the admission period drops be-


low the initial steam pressure and the cylinder pressure during exhaust
is higher than the exhaust pressure because of throttling or friction

effects in flow through the ports and valves. Both of these pressure
differences reduce the work of the real engine.
Incomplete Expansion. The stroke of the real engine is made
shorter tWn that of the ideal engine, stopping the expansion process
before the cylinder pressure reaches the exhaust pressure. This
causes a reduction in the work done by the steam, as shown by the
loss of the tip of the ideal diagram; however this need not mean a
reduction in the real engine output. The real engine has mechanical
frictionwhich demands a certain area on the indicator diagram before
any useful work is available at the shaft. At the low-pressure end
of the ideal diagram not enough work is indicated to oWM'come the
friction losses corresponding to that portion of the stroke in the real
376 THE STEAM ENGINE [2(W

engine; hence the stroke is cut short at a point where there is still

sufficient pressure to overcome the engine friction. There is a slight


loss of work due to the opening of the exhaust valve before the end
of the stroke.
Clearance and Compression. The real compression process starts
at a higher pressure than the ideal process because of throttling as
the exhaust valve closes. Because of heat transfer to the cylinder
head and walls, however, the real compression ends at a pressure
Then, when the steam
considerably below the initial steam pressure.
valve opens, there is an irreversible expansion of steam into the clear-
ance space to raise the cylinder pressure to the steam pressure.
initial

The indicated work is not much affected by the between


difference
the real and ideal compression processes, but the amount of steam
required by the engine is increased and the efficiency is thus reduced.
20-3 Cylinder condensation. The most important cause of re-
duced efficiency of the real engine, compared to the idealized engine,
is not directly evident on the indicator diagram. It is the condensa-
tion of steam in the cylinder during the admission process and the
subsequent re-evaporation during the expansion and exhaust proc-
esses. Because the cylinder is exposed to steam alternately at the
initial temperature and at the exhaust temperature, the average tem-

perature of the interior metal surfaces of the cylinder will lie between
these limits. Also in the slide-valve engine. Fig. 20-1, the surfaces
of the ports and the cylinder heads are swept by exhaust steam and
brought close to exhaust temperature during the exhaust process.
During the compression process there may be a slight rise in metal
surface temperature but there remains a large temperature difference
between the metal surfaces and the incoming steam so that a large
fraction of the steam may condense on the walls. The resulting
liquid has negligible volume and therefore adds nothing to the work
during the admission process, but does add to the steam flow. As
the pressure drops during the expansion process some of the liquid
evaporates and contributes some work, but the main effect of re-
evaporation is to increase the pressure drop through the exhaust port
and to cool the cylinder head and the exhaust port walls. The frac-
tion of the steam which condenses is in effect wasted, so the steam
consumption of the engine is increased without a corresponding in-
crease of work.
It should be observed that cylinder condensation does not depend
20-4 ] THE STEAM ENGINE 377

primarily upon heat loss from the cylinder, since most of the heat
transferred from the steam to the walls by condensation is transferred
back to the steam when re-evaporation occurs. Thus the work loss
is not an energy loss but is a loss of availability of energy due to the

irreversible heat transfer from the steam to the cylinder and back to
the steam. The effect of the irreversible heat transfer is to increase
the entropy of the steam as explained in Sec. 8-6 and 8-7. In the
expansion with no net heat transfer and with increasing entropy, the
enthalpy change, and therefore the work, is less than in an isentropic
expansion.
20-4 Methods of reducing cylinder condensation. The im-
portance of the condensation and re-evaporation processes in the
engine cylinder stems from the high rates of heat transfer associated
with such processes. The methods of reducing condensation are es-
sentially methods of reducing heat transfer between steam and walls.
Superheat. For a given temperature difference between wall and
fluid, the rate of heat transfer from condensing steam may be several

hundred times as great as the rate of heat transfer between super-


heated steam and dry cylinder walls. Therefore one of the most
effective methods of minimizing heat transfer from the entering steam
is to superheat the steam. By superheating sufficiently to insure no
condensation before the cut-off point, the efficiency of a small engine
has been increased as much as 40 percent, and there are numerous
cases in the literature that show efficiency gains of 10 to 20 percent
by superheating.
Jacketing. Engine cylinders may be surrounded by steam-filled
jackets which keep the outer surfaces of the cylinder walls at the
inlet steam temperature. This results in a higher average tempera-
ture of the cylinder walls, and thereby reduces heat transfer and
condensation. Since some steam is used in the jackets the net gain
may be small or even negative, particularly when other methods of
reducing condensation are used simultaneously with the jackets.
Compounding. If the expansion of the steam is carried out in
two cylinders in succession, the temperature difference between inlet
and exhaust for each cylinder may be reduced to half the overall
temperature difference, thereby reducing the heat transfer rates in
about the same proportion. A compound engine, shown schematically
in Fig, 2(M, has a small high-pressure cylinder and a large low-
pressure cylinder, usually so proportioned that the work is divided
378 THE STEAM ENGINE [20-4

about equally between them. The receiver between the cylinders


provides for the time difference between the exhaust from the high-
pressure cylinder and the admission
HIGH PRESSURE
STEAM CYLINDER to the low-pressure cylinder. The
indicator diagrams for the
two
cylinders of an idealized compound
engine, plotted to common scales
of pressure and cylinder volume,
are shown in Fig. 20-5. In the
RECEIVER
past the compound engine and the
triple expansion engine (using three
cylinders in series) were much used
for factory drives, for marine pro-
pulsion, and for water-works pump-
ing service where high efficiency
desired and considerable bulk
LOW PRESSURE
EXHAUST CYLINDER weight could be tolerated.
^
rig. 20-4.
Fig.
^
Compound
_
now
engine flow »>
Separate
. r.
Inlet
..
and
^
Exhaust

diagram.
Ports. Separation of tne steam
and exhaust ports aids in reducing
cylinder condensation, and has other advantages. In four-valve
engines like the Corliss engine. Fig. 20-6, the valves are close to the
cylinder, thereby reducing the port surface on which condensation
may occur, and the steam ports are kept near the steam temperature
because they are not traversed by exhaust steam. Also the steam

Fig. 20-5. Compound engine indicator diagram.


2(M] the steam engine 379

and exhaust chests are remote from each other, reducing the direct
transfer of heat from inlet steam to exhaust steam which occurs in
slide-valve engines.
In addition to reduced condensation the four-valve engine offers
the advantages of more freedom in the choice of valve timing, and
no possibility of direct steam leakage from the steam chest to the
exhaust chest.
Unijlow Cylinder. The separation of steam and exhaust ports is

carried to its logical conclusion in the uniflow cylinder. Fig. 20-7.


In this arrangement the steam enters the cylinder at the end, as in a
four valve engine, but the exhaust ports are as far as possible from
the end of the cylinder. The exhaust ports are closed by the piston,
release of steam to exhaust occurring when the piston runs past the
edge of the ports. Compression must occur correspondingly early in
the return stroke, when the piston covers the exhaust ports, as shown
in the indicator diagram in Fig. 20-7.
The basic advantage of the uniflow engine is that the exhaust
steam never sweeps over the surfaces with which steam comes in
contact during the period of admission. Consequently the average
temperature of these surfaces remains close to the steam temperature
and condensation is minimized. It is common practice to use steam
jackets on the cylinder heads in uniflow engines. The uniflow engine
380 THE STEAM ENGINE [20-J

Fig. 20-7. Uniflow engine.

is well adapted to the use of high vacuum (low absolute exhaust


pressure) because the exhaust port arrangement provides the great
flow area necessary to pass large volumes of exhaust steam without
excessive pressure drop. It is not so well adapted to atmospheric
exhaust pressure because the early closing of the exhaust port then
causes excessive compression pressure unless an auxiliary exhaust
valve near the cylinder head, or a large clearance volume, is provided.
Either of these expedients tends to nullify the advantage of the uni-
flow principle.
20-5 Conclusion. The steam engine may be superior in effi-

ciency to the turbine in the smaller sizes (up to several hundred


horsepower) and in the higher exhaust pressure ranges. In larger
sizes,and when low exhaust pressure is used, the turbine may be
superior in efficiency to the engine. For units of more than several
thousand horsepower mechanical problems and great bulk render
engines impractical. In the size range where engines and turbines
might be competitive with respect to efficiency, turbines have largely
displaced engines because the turbine has the advantages of less
weight and bulk, less operating attention and maintenance, simpler
foundations, and better speeds for driving electric generators and
centrifugal pumps. Often these advantages outweigh lower efficiency,
THE STEAM ENGINE 381

and the turbine is used in small


wheie an engine would be more
sizes
efficient. The disappearance of small steam engines, however,
is not

due so much to the turbine as to the electric motor and the internal
combustion engine.

PROBLEMS
20-1. A simple steam engine will usually have a lower engine eflSciency and
a lower steam rate (Ib/hphr) when exhausting to a condenser under vacuum
than when exhausting to atmosphere, the steam being supplied at the same state
in both cases. Give a reasonable explanation of these facts.
20-2. A double-acting slide-valve engine of 18in. bore and 48 in. stroke was

tested both condensing and non-condensing, with the following results;

Condensing Non-condensing
Steam pressure (saturated), psia 110 110
Exhaust pressure, psia 4.0 15
Indicated horsepower 144 122
Steam rate, Ib/hphr 20.5 26.0
Engine speed, rpm 75 75

Find, for both tests: (a) the mean effective pressure (neglect piston-rod vol-
ume) (b) the engine efficiency. Explain the differences between the answers for
;

the two tests.


20-3. A small steam engine was tested with saturated steam and with super-
heated steam: the results were as follows:

Steam Steam Exhaust Indicated Steam


pressure temperature pressure horsepower rate

102 psia saturated 15 psia 13.3 39.6 IbApbr


100 650°F 15 13.5 20.1

For each test find: (a) the engine efficiency; (b) the heat rate for the engine.
(The heat rate is defined as the product of steam rate X {hi — ^2/), where hi is
the enthalpy of the steam supplied and ^2/ is the enthalpy of saturated liquid at
the exhaust pressure.) Explain the differences between the answers for the two
tests.

20-4. A simple slide-valve engine has 10 percent clearance. When supplied


with steam at 120 psia, saturated, and exhausting to the atmosphere, the engine
has a mean effective pressure of 50 psi as determined from the actual indicator
diagram. Find the ratio of this to the mean effective pressure that would be
obtained from a reversible adiabatic engine having 10 percent clearance.
20-5. The engine of Problem 20-4 has an actual steam rate of 40 Ib/hphr,
based upon indicated output. Find the engine efficiency; compare this value
(a work ratio) to the ratio of mean found in Problem 20-4;
effective pressures
explain with sketches of indicator diagrams how the actual engine can have a
much higher mean effective pressure than the reversible adiabatic engine, even
though the work per pound of steam is much smaller.

N*
382 THE STEAM ENGINE
20-6. For a certain uniflow engine the clearance is 2 percent; steam is sup-
plied at 150 psia, and the exhaust pressure is 4 in. Hg^bs. By test at rat^
load the mean effective pressure is 46 psi and the steam rate is 12.0 Ib/hphr.
Compare these test results to the values computed for a reversible adiabatic
engine with 2 percent clearance. Compare with the results of Problems 20-4
and 20-5 and explain the differences. These results show that the actual engine
can be far superior in work capacity to the reversible adiabatic engine, even
though its efficiency is smaller. «
Chapter 21

GAS COMPRESSION

The compression of gases is an important process in many power


plants, refrigeration plants, and industrial plants of which power
plants and refrigeration plants are discussed elsewhere in this book.
Industrial uses of gas compression occur in connection with com-
pressed air motors for tools, air brakes for vehicles, servo-mechanisms,
metallurgical and chemical processes, conveying of materials through
ducts, transportation of natural gas, and production of bottled gases.

The term gas compression is usually applied only to processes involv-


ing appreciable change of gas density; this excludes ordinary ventila-
tion and furnace draft processes.
The machinery used in gas compression may be turbine t}rpe,

such as centrifugal and axial flow machines; or positive displacement


type, such as reciprocating machines, meshing rotor or gear machines,
and vane-sealed machines. Insofar as it operates under steady flow
conditions, any of these types of machine may have its energy analysis
written in the form of the steady flow energy equation; in this chapter
some general deductions will be made upon this basis. A more de-
tailed study will be made of the reciprocating compressor, but not of
the other machines.
21-1 Compression processes. A gas compression process may
be designed either to be adiabatic or to involve heat transfer, depend-
ing upon the purpose for which the gas is compressed. If the com-
pressed gas is to be used promptly in an engine or in a combustion
process, adiabatic compression may be desirable in order to obtain
the mayimiiTn available energy in the gas at the end of compression.
In many applications, however, the gas is not used promptly but is

stored in a tank for use as needed. The gas in the tank transfers
heat to the surroimdings, so that when finally used it is at room
temperature. In these oases the overall effect of the compressioD
Fig. 21-1. Rotaiy compressors: (a) 4-stage centrifugal; (b) 20-stage axial
flow; (c) and (d) vano-type, positive displacement compressors,
lobe-type;
[(a), (c) from Compressed Air Power, Compressed Air and Gas Institute; (b)

courtesy Allis-Chalmers Mfg. Co.; (d) courtesy Socony-Vacuum Oil Co.]


21 -2 ] GAS COMPRESSION 385

Fig. 21-2. Compression processes.

and storage process is simply to increase the pressure of the gas


without change of temperature. Now it will be shown below that
if the gas is cooled during compression, instead of after that process,
the work required will be less than for adiabatic compression. A
further advantage of cooling is the reduction of volume and the con-
sequent reduction of pipe line friction losses. For this reason, since
cooling during the compression process is not very effective, after-
coolers are often used to cool the gas leaving a compressor.
Because of the situation described above it is of interest to con-
sider the effect of cooling upon
a compression process. It is customary
to investigate two particular idealized ca«“® nomoKr raTrnretKIa orlia.

batic and reversible isothermal, as well


as a general case of a reversible poly-
tropic process {jw^ = constant). The
paths of such processes are plotted
in Fig. 21-2 for a perfect gas compres-
j Wr
sion process from state
COMPRESSOR
1 to pressure
Pi-

Work of compression in
21-2
steady flow. The steady jflow energy t'
^FLOW
equation for a compression process
Fig. 21-3.
may be written, referring to Fig. 21-3,
as

Ai + Q = ^ + Wx (1)

assuming that changes of potential and kinetic energy are negligible.


It is often convenient to write the equation in terms of pressure and
386 GAS COMPRESSION [21-2

volume rather than enthalpy; this is done by using the property


rdation
T da — dh — V dp (9-6)

For a reversible process this gives

Q = — Jvdp (2)

then for any of the idealized cases of Fig. 21-2, from (1) and (2)

F* = -JV dp (3)

It is assumed that for any gas a compression process may be rep^


resented with sufficient accuracy by an equation of the form pv" =
constant. Then

V pl/n

and F.= p*'"


Ji

= ^ Zi/n ~

Now the work of compression, or the steady flow work input to the
gas, is the negative of the shaft work Wx- Designating the work of
reversible polytropic compression by F„, and the work of reversible
adiabatic compression by Wk,

>nd (5)

For isothermal compression of a perfect gas, pv => constant; then the


work of reversible isothermal compression for a perfect gas is

F, = -F. = p^v, = pkVy In


^ (6)
Ji P Pi

In the Tp-v plot of Fig. 21-2 the work of compression for each type
of process is represented by the area between the path of that process
and the axis of pressures. It is evident that the work of reversible
21 - 2 ] GAS COMPRESSION 387

isothermal compression than the work of reversible adiabatic


is less

compression; the work of reversible poly tropic compression is inter-


mediate between the others if n lies between k and unity. This will

be the case if the polytropic process involves some cooling but not
enough to obtain isothermal compression; such conditions obtain in
actual reciprocating compressors. In a real compressor the work will
be greater than the work of the reversible compression process because
of friction. In such cases the path of compression may be represented
by pi>“ = constant, but the work of compression is not given by
y*t> dp; the shaft work cannot be determined solely from the proper-
ties of the fluid. The friction effects in a reciprocating compressor
are often small so that the work may be computed by the int^al
of V dp without great error.

Example 1. Tests on reciprocating air compressors with water-cooled


cylinders show that it is practical to cool the air sufficiently during compres-
sion to correspond to a polytropic exponent n in the vicinity of 1.3. Com-
pare the work per pound of air compressed from 15 psia, SO^F to 90 psia,
according to three processes: reversible adiabatic, reversible isothermal, and
reversible pr* * = constant. Find the heat transferred from the air in each
case.

Solution: For air, considered as a perfect gas,

pivi = RTi = (53.34) (540)

Then ^* = (53.34)(540)

- l]
51 [(
= (101,000) (1.67 - 1)

= 67,700 ft Ib/lb

= (125,000) (1.52 - 1)

= 65,000 ft Ib/lb

W, = (53.34) (540) In??


15

= 51,900 ft Ib/lb

The heat transferred in the adiabatic process is zero. In the polytropic


process, using Eq. (11-24),

Q»c.f^(r,-r.)
1 —
388 GAS COMPRESSION [21-3

= (0.171)(778) (540)(1.52 - 1)

= -16.0Btu/lb

Then the heat transferred from the air is 16.0 Btu/lb. In the isothermal
process with a perfect gas the heat transfer is equal to the work; then the
heat transferred from the air is 51,900 ft Ib/lb or 66.7 Btu/lb.
The usefulness of cooling for work reduction in the idealized compression
process is clearly shown by Example 1. In real compression processes the
desired cooling can only be approximated because it is impractical to build
a compressor with suflScient heat transfer capacity without sacrificing other
desired characteristics.

21-3 Efficiency of a compressor* As already shown in the chap-


ters on power plant cycles and on reciprocating machines, the effi-

ciency of a compressor working in a steady flow process may be


defined as

(7)
Wc Wc
where is the shaft work supplied to the actual compressor per
pound of gas passing through, and TTt is the shaft work supplied to
a reversible adiabatic compressor per pound of gas compressed from
the same initial state to the same final pressure as in the actual
compressor.
In the case of gas compression, as explained above, the desirable
idealized process may sometimes be a reversible isothermal process
rather than a reversible adiabatic process. Therefore the compressor
efficiency is sometimes defined as

_ El ( 8)
Wc
where Wt is the work of reversible isothermal compression from the
actual initial state to the actual final pressure.
The two efficiencies of Eqs. (7) and (8) are called respectively the
adiabaUc efficiency and the isothermal efficiency. For general thermo-
d}mamic purposes it is well to use the reversible adiabatic basis, but
21*3] GAS COMPRESSION 389

the existence in practice of the other concept should be recognized


and if the possibility of confusion exists, the basis of any stated value
of efficiency should be specified.

Because of the effects of cooling, the adiabatic efficiency of a real


compressor may be greater than unity. This should cause no logical
difficulty if the definition of the machine efficiency as an arbitrary
ratio is understood.
Many turbine-type compressors are essentially adiabatic ma-
chines. For an adiabatic machine the work of compression is equal
to the enthalpy rise of the gas;

Wc^fh- hi

Then for an adiabatic compressor the efficiency is

hu — hi
(9)
hi

Observe the distinction between the adiabatic efficiency of any ma-


chine, Eq. (7), and the efficiency of an adiabatic machine, Eq. (9).
For an adiabatic compressor working with a perfect gas the efficiency
may be obtained from (9) if only the pressures and temperatures at
inlet and outlet are known. The actual temperature rise gives the
actual enthalpy change per pound, and the pressure ratio, with the
initial temperature, gives the isentropic enthalpy change per pound.
As explained in Sec. 19-5, compressor efficiency of a reciprocating
machine may be on an indicated work basis or a brake work basis,
depending upon where the work input is measured.

Example 2. For the conditions of Example 1 find the adiabatic efiSr


ciency and the isothermal efficiency of the reversible polytropic compressor.
Solution: For the given conditions TFc is the TFn of Example 1.

T 1
Isothermal: ’Jc = ^= ^
51-900
= non
0.80

Aj- V.
Adiabatic:

Example 3. For the same


^67,700

initial state and final pressure as in Exam-


ple 1, a real compressor has an efficiency of 95 percent on the adiabatic basis.
The initial and final states correspond to a polytropic compression with
n = 1.3. Find the heat transferred per pound of air.

SoLTjriON: Since this is not a reversible compression process the heat


390 GAS COMPRESSION [21-3

traDsferred cannot be obtained from the polytropic equation. By the steady


flow equation

+Q = + TT*
Aj

Q = A2 — Ai + TT,
From Example 1, Wk = 67,700 ft Ib/lb. Then

IF. = = -71,200 ft Ib/lb


ije 0.95

hi - hi = c,{Ti — Ti) = CpTi


(S-)
From Example 1, for the process between the same end states,

540(1.52 - 1)

then hi - /t, = (0.240) (540) (1.52 - 1) = 67.7 Btu/lb.

Q = 67.7 - = -24.0 Btu/lb.

Compare the work, the heat, and the enthalpy change in this example with
the corresponding quantities for the reversible polytropic process in Exam-
ple 1.

Fig. 21-4. lleciprocatiug compressor. Note the cooling-water jackets. The


valve lifter isan unloading device which stops the pumping action by holding
the inlet valves open when the air reservoir is filled to capacity. (Courtesy
Socony- Vacuum Oil Co.)
21-4] GAS COMPRESSION 391

21-4 Reciprocating compressors —^work of compression.


The general characteristics of reciprocating machines have already
been presented in Chap. 19. A reciprocating compressor shown
is

in section in Fig. 21-4, and a typical indicator card from such a ma-
chine is shown in Fig. 21-5. The sequence of operations in the cyl-

inder is as follows:

Fig. 21-5. Compressor indicator diagram.

(1) Compression: Starting at maximum cylinder volume, point


o, slightly below the inlet pressure pi, as the volume decreases the
pressure rises until reaches pz at 5; the discharge valve does not
it

open until the pressure in the cylinder exceeds pt by enough to over-


come the valve spring force.
(2) Discharge: Between h and c gas flows out at a pressure higher
than Pi by the amount of the pressure loss through the valves; at c,
the point of minimum volume, the discharge valve is closed by its

spring.

(3) Expansion: From c to d, as the volume increases, the gas re-


maining in the clearance volume expands and its pressure falls; the
suction valve does not open until the pressure falls sufficiently below
Pi to overcome the spring force.
(4) Intake: Between d and a gas flows into the cylinder at a
pressure lower than pi by the amount of the pressure loss through
the valve.
The total area of the diagram represents the actual work of the
392 GAS COMPRESSION [21-4

compressor on the gas. The cross-hatched areas of the diagram


above p2 and below pi represent work done solely because of pressure
drop through the valves and port passages; this work is called the
valve loss.

The machine to which the actual machine is compared


idealized
has an indicator diagram like Fig. 21-6, in which there are no pressure

CYLINDER VOLUME

Fig. 21-0.

loss effects,and the processes a-b and c-d are reversible polytropic
processes. Assuming no state change in the intake and discharge
processes, d-a and b-c, and assuming equal values of the exponent n
in the compression and expansion processes, a-b and c-d, the ideal
work of compression can be found by taking the integral of p dV
around the diagram. Let the flow fluid mass be ms (this is the mass
of fluid taken in and discharged per machine cycle). Then using the
known work values for the constant-pressure and polytropic processes,
W = -f- Wi-c + Wc-a + Wa-a
bV h ~
+ MVc -
VttVg
Ffc) + PiiVa - Va)
1—71 -f-
1 — 71

o^(Va - Va)
1—71
- MVb - Vc) -b
1—71
- Pi(.Va - Va)

[pj(n - Vc) + PliVa - Va)]


1 — 77

— [jhmsVt - pmA>\]
1 n

«= 771/
_ [ptVz - Plt»i]
J ^
21 -6 ] GAS COMPRESSION 393

r p2t>2
LpiV] 0
Since pr" = constant

7hVi _ /pgV""”'*
PiVi \pj
Thus it is seen that the work per pound of fluid flow is the same as
obtained from the steady flow analysis, and given in Eq. (4). It is
therefore unnecessary to make any further analysis of the work of
the idealized reciprocating compressor since all desired results have
already been obtained by the steady flow analysis.
One point to be noted here is that the mass of clearance fluid does
not appear in the work equation that ; is to say the work per pound
of flow fluid in the idealized compressor is independent of the clearance
volume of the compressor. This is only approximately true for real
compressors.
21-5 Volumetric efficiency of reciprocating compressors*
The flow capacity of positive displacement compressors (and internal
combustion engines) is expressed in terms of a quantity called the
volumetric efficiency riv This is defined as the ratio of the actual
volume of fluid taken into the compressor per machine cycle, to the
displacement volume of the machine. For a reciprocating com-
pressor

_ m/Vi
Vv
- ( 10 )

where m/ is the flow fluid mass per machine cycle, Vi is the specific
volume of the fluid approaching the inlet valve, and (PD) is the pis-
ton displacement per machine cycle.
The true volumetric eflficiency can be determined only by meas-
uring the flow through the machine. An approximate or apparent
volumetric efficiency may be obtained from the indicator diagram as
shown in Fig. 21-5. Here the volume Vi is the volume between the
point where the cylinder pressure reaches pi during the expansion
process and the point where it reaches pi during the compression
process. If the gas remained at constant temperature during the
intake process, the volume Vt would be the actual volume taken in
at state 1 then the ratio F//(PD) would be the volumetric efficiency.
;

In an actual compressor, because of heat transfer from the cylinder


394 GAS COMPRESSION [2M

walls, the gas is at higher temperature after entering the cylinder


than at state 1. Consequently the volume Vi is greater than the
volume taken in from the supply pipe, and the ratio F//(PD) is
larger than the true volumetric efficiency; hence the name apparent
volumetric efficiency.
21-6 Volumetric efficiency and clearance. The volumetric
efficiency of an idealized compressor having an indicator diagram
like Fig. 21-6 can be written directly from Eq. (10) and the relation
for flow capacity of an idealized compressor, Eq. (19-8a).

Since it is usually convenient to deal with pressure ratio rather than


volume ratio this equation may be rewritten as follows for a com-
pressor working on an ideal polytropic process;

Equation (11) is plotted in Fig. 21-7. Noting that (pj/pi)^'" is


always greater than unity, it is evident that the volumetric effidency

Fig. 21-7. Effect of clearance on volumetric efficiency.

of the idealized compressor decreases as the clearance increases and


as the pressure ratio increases.
In order to get maximum flow capacity, compressors are built
with the minimum practical clearance. Sometimes, however, the
clearance is deliberately increased as a means of controlling the flow
21-7] CAS COMPRESSION 395

through a compressor driven by a constant speed motor. The com-


pressor cylinder of Fig. 21-8 is fitted with a clearance pocket which
can be opened or closed at will

by a valve. Suppose this ma-


chine is operating at conditions
corresponding to line e in Fig.
21-7; if the clearance volume
is at minimum value a the vol-
umetric efficiency and the flow
through the machine are maxi-
mum. If the clearance pocket
is then opened to increase the
clearance to the value b, the Fig. 21-8. Clearance pocket for capacity
volumetric efficiency and the control.

flow are reduced. By increas-


ing the clearance in steps, as indicated by points c and d, the flow
may be reduced in steps to zero. As pointed out in the preceding
section the work per pound of gas compressed in an idealized com-
pressor is not affected by the clearance volume. In a real compres-
sor all the mechanical losses of the machine continue during opera-
tion at reduced flow rate so the work input to the machine per pound
of fluid increases as the flow rate decreases.

Example 4. A reciprocating air compressor operates between 15 psia


and 75 psia with a polytropic exponent of 1.3. IIow much clearance would
have to be provided, in the ideal case, to make the volumetric efficiency 50
percent? To make it zero?

£->y'" = 5>« > = 3.45


PiJ

n, = 0.50 = 1 - Cl(3.45 - 1)

Cl = ^
2.4o
= ().2(M

= 0 = 1 - Cl(3.45 - 1)

Cl = ^
ZAi)
= 0.408

21-7 Volumetric efficiency and pressure ratio multistage —


compression. It is evident from Fig. 21-7 that as the pressure ratio
is increased the volumetric efficiency of a compressor of fixed clearance
396 GAS COMPRESSION [21->

decreases, eventually becoming zero. This can also be seen in an


indicator diagram, Fig. 21-9. As the discharge pressure is increased,
the volume Vi, taken in at At some pressure pje the
pi, decreases.

compression line intersects the line of clearance volume and there is

Fig. 21-9. Effect of pressure-ratio on capacity.

no discharge of gas. An attempt to pump to any higher


p2c (or
pressure) would result in compression and re-expansion of the same
gas repeatedly, with no flow in or out.
The maximum pressure ratio attainable with a reciprocating com-
pressor cylinder is then seen to be limited by the clearance. There are
practical and economic limits to the reduction of clearance; when
these limits interfere with the attainment of the desired discharge
pressure, it is necessary to use multistage compression. In a multi-
stage compressor the gas is passed in series through two or more
compressors, or stages, each of which operates on a small pressure
ratio. Disregarding pressure losses between stages, the overall pres-
sure ratio is the product of the pressure ratios of the stages. A two-
stage compressor is illustrated in Fig. 21-10.
Figure 21-11 shows the comparative idealized indicator diagrams
for compression of a gas from pi to p2 by a two-stage machine or by
21-7] GAS COMPRESSION 397

F0«CE-fEE0
LUBRICATOR

>Di

.FIRST- 9TA6E
INLET

Fig. 21-10. Two-stage reciprocating compressor with intercooler. (Courtesy


Socoiiy-Vacuum Oil Co.)

a single-stage machine of the same piston displacement and clearance


as the first stage of the two-stage machine. The single-stage machine
compresses gas from a\ to 62, discharges at p2 from 62 to c', expands
from c' to d', and takes in gas from d' to ai. Thus the capacity per
machine cycle is Vai — V^.

SECOND
STAGE

RECEIVER

TWO STAGE CL(PD),-1


FLOW DIAGRAM

Fig. 21-11. Two-stage compression


398 GAS COMPRESSION [ 21-8

The first stage of the two-stage machine compresses gas from


Oi to bi, discharges at px from bi to Ci, expands from Ci to di, and
takes in gas at pi from di to Oi. The capacity per machine cycle is

Vai — Vdi, which is appreciably larger than the capacity of the single-
stage machine.
The second stage takes in gas from diiooi (which coincides with
bi~Ci), compresses from <h to 62 , discharges at p^ from bi to C3 , and
expands from Cj to di. The flow capacity of the two-stage machine
is the capacity of the first stage, since all the gas is taken in by the
first stage.
The two-stage compressor has greater capacity than the single?
stage compressor of the same clearance, at the same pressure ratio
Pi/Pi- This advantage is greater at larger pressure ratios, and at
sufficiently large pressure ratios the single-stage compressor becomes
uneconomical because of low volumetric efficiency (i.e. low flow
capacity).

Example 5. A gas is to be compressed from 5 psia to 83.5 psia. It is


known that cooling corresponding to a polytropic exponent of 1.25 is prac-
tical and the clearance of the available compressors is 3 percent. Compare
the volumetric efficiencies to be anticipated for (a) single-stage compression,
and (b) two-stage compression with equal pressure ratios in the stages.
(A reasonable comparison can be made on the idealized basis even though
the actual volumetric efficiencies may be lower than the ideal.)

Solution: For the single-stage machine

= 1 - 0.03[(16.7)W‘-“ - 1] = 0.744

For the two-stage machine the pressure ratio in each stage is ^ 16.7 and the
volumetric efficiency is that of the first stage. .

Vv = l - 0.03[(4.09)»« “ - 1] = 0.934

The relative advantages of single- and multiple-stage compression


will be discussed after the influence of intercooling has' been considered.
21-8 Intercooling. The advantage of multistage compression
in itself is primarily that of increased flow capacity or volumetric
efficiency for a given pressure ratio. In the idealized case no reduc-
tion is obtained in the work of compression per pound of gas passed
through the machine; in actual cases some work may be saved, but
considering the added mechanical complication of the two-stage ma-
chine such a saving is likely to be small or even negative. In both
21 -8 ] GAS COMPRESSION 399

ideal and actual cases, however, an appreciable saving of work may


be obtained by utiUzing the opportunity for effectively cooling the
gas between the stages of compression.The cooling is usually done
by a water-cooled tubular heat exchanger which also serves as a
receiver or reservoir between the stages.
The work saved by intercooling in the idealized two-stage recipro-
cating compressor is illustrated on the indicator diagram of Fig. 21-12.

FLOW DIAGRAM
Fig. 21-12. Two-stage compression with intercooling.

Cooling by the cylinder water jackets is never very effective; the


compression curve is always closer to adiabatic than to isothermal.
Therefore the gas discharged from the first stage at state x is at a
higher temperature than the inlet temperature Ti\ if the gas is then
cooled to state y at temperature Ti, the volume entering the second
stage will be less than the volume leaving the first stage. The com-
pression in the second stage will then proceed alonga new polytropic
curve at smaller volume. The cross-hatched area between the two
polytropic curves in Fig. 21-12 represents work saved by interstage
cooling to the initial temperature. Actual cooling might be to some
other temperature, but it is conventional to discuss cooling to Ti.
The saving of work by two-stage compression with intercooling
400 GAS COMPRESSION [21-9

will depend upon the interstage pressure p* chosen. Obviously, as


p« approaches either pi or pi the process approaches single-stage com-
pression. Any saving of work must increase from zero to a maximum
and return to zero as p, varies from pi to p2 .

21-9 Minimum work in two-stage compression with inter-


cooling. The conditions affecting the work of compression may be
studied by use of the steady flow system and T-s diagram of Fig.
21-13. As shown, a perfect gas is compressed from the initial state

Fig. 21-13. T-s plot of two-stage compression process.

Pi, Ti to p*; it is then cooled at constant pressure to Ty, and then


compressed from px, Ty to p2 . Given pi, Ti, Ty, and p2 ,
it is desired
to find the value of p* which gives minimum work. Let the adiabatic
compression efficiencies of the two stages be respectively vci and 1/C2
;

then the work of compression is

(k-DIk

0
21-9] GAS COMPRESSION 401

and
/gaV*-!)/* _
~ /£2 1 II
\Py) VPx

Then ^ rT,/n
- - UAr. •)]
It 1 )^nc2\n
Taking the derivative with respect to T'x and setting it equal to zero
(noting that Ti, T'2, and are constant),

^
dT'x
= A
"
= _^ri-
fc- lUci
-u M V(WJM
Then {T^y = ^ TJ'i
jvci Ty Tj
and
\17c2 Ti Ti

for minimum work. Now

therefore for minimum work in two-stage compression of a perfect


gas with intercooling to a fixed temperature Ty,

For the special case of Ty


= Ti and)K i = i?c2, which is
7

often taken as a standard of


comparison, the requirement
for minimum work is

? = ( 13 )
Pi \pi
Also for this special case the
condition of mini mum work is
the condition of equal work in
the two stages.
When three stages of equal Pig_ 21-14. Throe-stage compression with
efficiency are used, with inter- intercooling to initial temperature.
402 GAS COMPRESSION [21-10

cooling to the initial temperature at two points as shown in Fig.


21-14, the condition of minimum work, and of equal division of
work among stages, is

EiJ = ?/B n4)


^
Pi Pxl Px» \ Pi

21-10 Ckimparison of compressor types. The choice of single-


or multistage compression for a given processis generally based on

an economic comparison; the physical characteristics which govern


the economic comparison are outlined below.
The single-stage compressor is simpler and may have less total
piston displacement than the two-stage compressor of the same flow
capacity. On the other hand, the single-stage cylinder must be larger
than the first stage of the two-stage machine and must at the same
time be built for the maximum pressure. In the two-stage machine
the higher pressures occur only in a small cylinder, which is more
easily built for high pressure. The two-stage compressor with inter-
cooler, which is rarely absent, requires appreciably less work and may
require less total piston displacement than the single-stage compressor
for the same Moreover the maximum temperature reached
service.
during the compression process may be greatly reduced by inter-
cooling; this reduces lubrication difficulties and explosion hazards.
The same considerations apply to larger numbers of stages when the
pressure ratio is large.

Practice appears to indicate that the economical value of pressure


3 to 5 for reciprocating compressors.
ratio per stage is in the range of
For compression of atmospheric air, single-stage machines are often
used up to 80 psi discharge pressure, two-stage from 50 to 300 psi,
three-stage from 300 to 1000 psi, and four or more stages for higher
pressures.
Turbine-type compressors are best adapted to large flow rates
and low pressure ratios. In general, the larger the flow rate the
higher the pressure ratio for which such machines are suited. For
industrial processes requiring large volumes of gas at pressures of the
order of one or two atmospheres, centrifugal compressors have long
been used. Axial flow compressors are now available for such appli-
cations. By using many stages they may be built conveniently for
high pressure ratios, although the pressure ratio per stage is not far
from unity. Such compressors do not have intercooling after every
GAS COMPRESSION 403

stage, but they may have the stages divided into two or three
groups with intercooling between the groups.
Rotary positive displacement compressors are widely used for
certain special applications such as vacuum pumps and domestic
refrigeration compressors, and for low pressure-ratio service where
the positive displacement characteristic of fairly constant flow against
variable pressure is desired.

PROBLEMS
21-1. Compare the work of steady flow compression of helium from 15 psia,
80®F, to 1000 psia by two processes: (a) reversible adiabatic; (b) isothermal.
21-2. Solve Problem 21-1 for carbon dioxide.
21-3. For the processes of Problem 21-1, find for (a) the temperature rise,

and for (b) the heat transfer per pound.


21-4. Solve Problem 21-3 for the processes of Problem 21-2.
21-5. A refrigeration compressor requires an input of 4.5 hp when com-
pressing 90 Ib/hr of ammonia in steady flow from 35 psia, saturated vapor, to
170 psia. Find the compressor adiabatic efficiency. If the heat transferred from
the compressor to cooling water and the surroundings is 5000 Btu/hr, find the
ammonia.
final state of the
21-6. Assume that the compression process of Problem 21-5 can be repre-
sented by the equation = constant. Find the value of n for the process in
the problem; find n for a reversible adiabatic process,
21-7. A well-insulated centrifugal compressor takes in ail* at 14 psia, 70®F,
and discharges it at 28 psia, 230®F. Find the adiabatic efficiency of the machine;
find the value of the exponent n if the process is represented by the equation
= constant.
21-8. An compressor cylinder has 8 in. diameter and 8 in. stroke, and
air
the clearance 5 percent. The machine operates between 14.5 psia, 80°F, and
is

95 psia; the poly tropic exponent is 1.32. Sketch the idealized indicator diagram
and find: (a) the total cylinder volume at each corner of the diagram; (b) the
flow mass and the clearance mass of air; (c) the capacity, cu ft/min, at 300 rpm,
based upon the idealized diagram.
21-9. For the compre.s8or of Problem 21-8 find: (a) the ideal volumetric effi-

ciency; (b) the mean effective pressure; (c) the heat transferred, stated as a frac-
tion of the indicated work.
21-10. Find the necessary piston displacement per minute for the compressor
of Problem 21-5 if the actual volumetric efficiency is 0.9 of the volumetric effi-

ciency based on the idealized indicator diagram for the process; the exponent n
is 1.20, and the compressor clearance is 3 percent.

21-11. An air compressor has a volumetric efficiency of 70 percent when


tested; the discharge state is 75 psia, 300® F, and the inlet state is 15 psia, 60®F.
If the clearance is 4 percent, predict the new volumetric efficiency when the
discharge pressure is increased to 100 psia. (Assume the ratio of real to ideal
volumetric efficiency stays constant, and the exponent n stays constant.)
404 GAS COMPRESSION

21 - 12 . Find the percentage saving in work by compressing nitrogen in two


stages, compared to single-stage compression, for the following conditions: initial
400 psia. Compression in each stage is reversible
state 25 psia, O'^F; final pressure
and adiabatic; in the two-stage process the pressure ratios of the stages are equal,
and there is intercooling to 90°F.
21-13. Air is compressed in three stages of 2.5 compression ratio per stage.
Each stage is adiabatic and has an efiiciency of 75 percent. A cooler after each
of the stages reduces the air temperature to lOOT; there is 2 psi pressure drop
through each cooler. The air enters the first stage at 14 psia, 70®F. Find:
(a) the pressure, temperature, ard specific volume at the entrance and exit of
each of the three coolers; (b) the heat transferred, per pound of air, in each cooler.
21-14. In Problem 21-13 the air flow rate is 50 Ib/sec. Find: (a) the power
required for each stage, and the total power; (b) the power required to reach
the same discharge pressure in a single adiabatic stage of 75 percent efficiency.
21-15. Carbon dioxide is to be compressed in two stages from 1 atm, — 60*'F,

to 16 atm. The compression in the first stage will be at 80 percent efficiency,


and in the second stage at 85 percent. Intercooling to 80°F will be obtained.
Find the interstage pressure for minimum work.
21-16. For the conditions of Problem 21-15 assume an interstage pressure
of 100 psia; find the work of each rtage and the heat transferred in the intercooler
per pound of gas.
21-17. Derive an expression for the interstage pressure for minimum work
m two-stage compression of a perfect gas when there is a pressure drop in the
intercooler, the other conditions being the same as in the derivation given in
The an expression Eq. - 12 but including a
Sec. 21-9. result should be like ( 21 ),

pressure-drop factor F defined by Py = FP*.


21-18. Derive an expression for the interstage pressure for minimum work in
two-stage compression of a perfect gas under the following conditions: process
as in Fig. 21-13; adiabatic stages of fixed efficiencies, rf^i and 97 (72 ; no pressure drop
in the intercooler; a fixed intercooler effectiveness vx defined by

The resulting expression should be in terms of pi, jht, fc, ijc-i, 17 (72 , and n*.
Chapter 22

COMBUSTION PROCESSES-
FIRST LAW ANALYSIS

Combustion processes are important in heat engineering because

they are the means of obtaining the high temperatures necessary for
efficient power plants. A combustion process is essentially the chem-
ical reaction of combining a fuel with oxygen. Such processes are
subject to the laws of thermodynamics, but before these laws can be
applied it is necessary to know something of the properties of the
systems involved and of the nature of the processes of combustion.
The discussion of the application of the First Law to combustion
processes is therefore preceded by an outline of the elementary chem-
istry of combustion.
22-1 Fuels. Fuels used in engineering practice may be solid,

liquid, or gaseous; they may range from pure substances to hetero-


geneous mixtures of combustible materials with rock or other min-
erals. In all ordinary cases the essential chemical elements in the
fuel are The other elements present may be
carbon and hydrogen.
considered impurities; they often have a profound effect upon the
burning process, but assuming that the fuel does bum, they have
only a minor effect upon the thermodynamic analysis of the process.
The ultimate analysis of a fuel is given in terms of the mass frac-

tions of carbon, C; hydrogen, H; oxygen, 0; sulfur, S; nitrogen, N ;

ash, A ;
and moisture, M. Except for ash and moisture the analysis
is in terms of the chemical elements, with no indication of the com-
pounds in which they may be combined. Moisture represents water
which is simply mixed with the fuel there ;
may also be water chem-

ically combined as an intrinsic part of the fuel structure and which


is accounted for by part of the H and 0 of the analysis. Ash is the
solid residue after all combustible and volatile components of the fuel

40-5 o
406 COMBUSTION PROCESSES [22-1

Table 22-1

Analyses of Typical Fuels*

Higher Heating
Ultimate Analysis
Value
(percent by mass)
(— AH, Btu/lb,
Fuel H 0 8 N dry basis)

Coals:
Anthracite 84.2 2.8 2.2 0.6 0.8 9.4| 13,810
Bituminous 86.4 4.9 3.6 0.6 1.6 2 .g| 15,178
Bituminous 73.1 4.8 8.9 2.6 1.5 9.1 13,469
Lignite 64.1 4.6 18.3 0.8 1.2 ll.O! 11,084

Oils:
Heavy Fuel Oil. 84.0 11.3 3.0 0.9 2.1 18,370
Light Fuel Oil. . 85.4 13.1 1.1 0.2 19,230
Gasoline 85.5 14.4 0.1 20,160

Gas:
Natural Gas. . . . CH4 , 96; Nj, 3.2; COj, 0.8 percent by volume
Natural Gas CH 4, 75.5; CsHt, 22 Ni,
;
1. 2 COj,
; 1.3 percentby volume
Coal Giis Hs, 46.5; CH«, 32.1; CO, 6.3; Ni, 8 . 1 ;
COi, 2 2 ; other
.

g!isea, 4.8 percent by volume.

• Actual analyses vary widely.


(From O. de Lorenzi, ed., Combustion Engi-
neering. New York:
Combustion Engineering-Superheater, lnc>, 1948; and John
Griswold, Fuds, Furnaces and Combustion. New York: McGraw-Hill, 1946.)

have been removed by burning. The ultimate analysis is used in


making the material balance for a combustion process.
Coals are mixtures of carbon, hydrocarbon compounds, sulfur
compounds, and non-combustible substances such as water, nitrogen,
and minerals (ash) The great variety of compositions in which coals
.

occur complicates the practical problems of furnace design and opera-


tion, but does not affect the energy analysis.
Liquid 'petroleum fuels such as gasoline, kerosine, and fuel oils are
almost entirely mixtures of hydrocarbon compounds. Fuel oils usu-
ally contain small amounts of impurities such as sulfur compounds,
nitrc^en, oxygen, water, and sediment (ash). When the ultimate
analysis is not known, a good approximation for the purpose of com-
bustion calculations is to assume that gasoline or kerosine has the
22-2 ] COMBUSTION PROCESSES 407

compodtion of octane, CsHu, diesel fuel oil has the compositiou CuHm,
and furnace fuel oil has the composition CieHse.
Oas fveU may be hydrogen, carbon monoxide, hydrocarbon com-
pounds. or mixtures thereof.
22-2 Combustion reactions —^material balance. Combustion
is the oxidation of a fuel at such a rate as to cause an appreciable
temperature hundred degrees) in the substances in-
rise (of several

volved in the process. The details of combustion reactions may be


exceedingly complex, and perhaps not fully understood, but the over-
all results can be described by simple chemical equations. The basic
equations describe the reaction between fuel and oxygen, and indicate
the amount of oxygen necessary to bum the fuel. Since the usual
source of oxygen is atmospheric air it is desirable also to set up equa-
tions for the air required (theoretical air).
Assumptions. Certain assumptions will be made as to the proper-
ties of the substances involved in the reactions.
Dry air is assumed to have the composition 21 percent oxygen
and 79 percent nitrogen by volume. (For convenience the argon and
other inert constituents, which make up about one percent of dry air,
are taken as nitrogen.) The fractions by mass are then, oxygen
0.232, and nitrogen 0.768.
The molecular weights of the elements involved in combustion
may be taken as follows: O2 32; N2 28; H 2 2; C, 12; S, 32. Then
, , ,

the equivalent molecular weight of air becomes

(0.21)(32) -I- (0.79)(28) = 28.8

which is often rounded off to 29. The molecular weights of combus-


tion products become: CO2 , 44; CO, 28; H 2O, 18; SO2 64. ,

It is assumed that all gases involved are perfect gases.


Combustion in Oxygen. The oxygen required to bum the ele-
ments, carbon, hydrogen, and sulfur, will be determined. The
equation for the combustion of carbon to carbon dioxide (complete
combustion) is

C-fOj-^COj
This equation is on a mol basis; that is, it states that one mol (molec-
ular wei^t) of carbon combines with one mol of oxygen to yield one
mol of carbon dioxide. By substituting the molecular weights of the
substances, a mass balance may be written:
*

408 COMBUSTION PROCESSES [22-2

12 lb C+ 32 lb 44 lb CO 2

This may be reduced to the basis of one pound of fuel as follows:

1 lb + 2.67 lb 3.67 lb

In the same way, for the burning of hydrogen,

H2 + i O2 H2O
1 mol + ^ mol —* 1 mol

2 lb + 16 lb 18 lb

1 lb + 8 lb 9 lb
For sulfur
S +O — SO2
2

1 mol + 1 mol — 1 mol >

1 lb + 1 lb 2 lb

Carbon does not always burn to carbon dioxide; it may burn to


carbon monoxide, a result sometimes called incomplete combustion.
For the combustion of carbon to carbon monoxide the material bal-
ance is:

C’ -t- ^ O2 CO
] lb + 1.331b ->2.33 lb

If the carbon monoxide is then burned, the balance is

CO -f- ^ O 2 CO 2

1 lb -I- 0.57 lb -> 1.57 lb

Observe that one pound of carbon will require the same amount of
oxygen to bum to carbon dioxide whether it burns directly to carbon
dioxide or indirectly by burning first to carbon monoxide; 2.33 pounds
of CO, which from burning one pound of carbon, will require
result
1.33 lb of oxygen and will yield 3.67 lb of CO 2 .

In all the reactions the mass of products must equal the mass of
reactants and the mass of each individual element must be the same
on both sides of the equation.
Combustion in Air. When the oxygen is supplied from air the
nitrogen of the air may be added to both sides of the equation since
22- 2 ] COMBUSTION PROCESSES 409

it passes through the reaction unchanged. As shown in Sec. 12-3,


the fractions by volume in a perfect gas mixture are equal to the
fractions by mols. Therefore, since air is 21 percent oxygen and
79 percent nitrogen by volume, each mol of oxygen is accompanied
by 79/21 mols of nitrogen. Taking the combustion of carbon to car-
bon dioxide as an example,

C+ + ^1 Ns
02 ^C02 + PrN2
12 lb + 32 lb + CA) 28 lb 44 lb + iU) 28 lb

1 lb + 2.G7 lb + 8.78 lb 3.07 lb + 8.78 lb

Then each pound of carbon requires 2.()7 + 8.78 or 11.15 lb of air.


This is called the theoretical air for burning carbon to carbon dioxide.
The theoretical air for several other reactions may be found in Ap-
pendix Table A-2. If the ultimate analysis of a fuel is known, the
theoretical air required to burn it may be calculated from the data
in the table.

Example 1. Find the thooroticnl air for combustion of octane (CgHig)


to CO 2 and H O.2

Solution: Each C rccjuircs 1 O2 ,


and each H 2 requires J O 2 then 8 C ]

and y H 2 require 12J mols of O2 ,


accompanied by (12.i)(H) inols of N 2 .

CJI 18 + 12.5 O2 + ( 125 )(H)N 2 8 CO + 2 H O f 125 )(lf)N


9 2 ( 2

114 lb + 400 lb + 1315 lb - 352 lb + 162 lb + 1315 lb

1 lb + 3,51 lb + 11,6 lb 3.09 lb + 1.42 lb + 11.6 lb

The theoretical air is 3.51 lb + 11.6 lb = 15.1 lb air per lb CsHi®.

Example 2. Find the theoretical air for a fuel of ultimate analysis


C, II. 0, S, N, A, and M.

Solution: The fuel contains three combustible elements, carbon, hj^dro-


gen, and sulfur, each of which requires a certain amount of air as indicated
in Table A-2. However, there is some oxygen already in the fuel, which
need not be supplied from the air. It is assumed for convenience that all of
this oxygen will eventually be combined with hydrogen; then the amount
of hydrogen which is finally combined with oxygen from the fuel is, per pound
of fuel, 0/8, where 0 is the fraction by mass of oxygen in the fuel. The

remaining hydrogen called free hydrogen^ requires oxygen from

the air. The theoretical air per pound of fuel is then given by
410 COMBUSTION PROCESSES [22-2

theo. air 11.5 C+ 34.2 + 4.31 S


where the letters are the fuel analysis fractions, and the numerical factors
are the air requirements per pound of the respective elements.

Example 3. Fuel oil containing 86 percent carbon and 14 percent hy-


drogen by mass is to be burned with 10 percent excess air (110 percent
theoretical air).
(a) If the air is dry, and combustion is complete, what will be the com-
position by mass of the products? (b) If the pressure is 15 psia what is

the dew point of the products? (c) If the products are cooled to 100®F
at 15 psia how much liquid water will condense per pound of fuel burned?

Solution: (a) To bum 1 lb of carbon with theoretical air,

1 lb C + 2.67 lb O + 8.78 lb N 2 2 -v 3.67 lb CO2 + 8.78 lb N*


With 110 percent theoretical air

1 lb C+ 2.94 lb O2 + 9.66 lb N 2 3.67 lb CO 2 + 0.267 lb 0* + 9.66 lb N 2

One pound of fuel contains 0.86 lb of carbon; the products of combustion


of this carbon with 110 percent theoretical air will be

0.86 (3.67 lb CO 2 + 0.267 lb O2 + 9.66 lb N) 2

or 3.16 lb CO 2 + 0.23 lb O + 8.30 lb Ns 2

By similar procedure, burning 0.14 lb of hydrogen with 110 percent theoreti-


cal air is found to give products as follows:

1.26 lb H O + 0.112 lb O + 4.09 lb N


2 2 2

Then for one pound of fuel the products will be

3.16 lb CO 2 + 1.26 lb HO 2 -t- 0.34 lb O2 + 12.39 lb Ns


or 17.15 pounds of products per pound of fuel. In fractions by mass the
composition is 0.184 CO2 ,
0.074 H O,2 0.020 O2 ,
0.722 N 2.

(b) The dew point of the products is the saturation temperature corre-
sponding to the partial pressure of the water vapor (Sec. 12-6). The partial
pressure of the water vapor is given by

where pw is the partial pressure of the water, p is the total pressure of the
gas, and isthe mol fraction of water vapor.

mols water vapor per pound of mixture


X,
mols mixture per pound of mixture
2M] COMBUSTION PROCESSES 411

0.074
18 0.00411
0.184 0^ 0^ 0^ ,
0.0347
44 18 32 28

Then py, = (0.118)(15) = 1.77 psia

From the steam tables the saturation temperature for this pressure is 121.6®r,
which is the dew point required.
(c) If the products are cooled to 100°F the partial pressure of the water
vapor will be the saturation pressure for 100°F; this is 0.9492 psia, from the
tables. Then neglecting the volume of am' liquid water pmsent, the mol
fraction of water vapor in the gas is

p^ 0.9492 ^
p 15

Let y be the pounds of water vapor per pound of mixture at 100®F; then the
mols of water vapor per pound of mixture is 2//I8. Similarly the mols of
each component w'ill be the mass of that component divided by its molecular
weight; then the total mols of all gaseous components of the mixture, per
pound of mixture, is

CO2 H2O O2 N2

44
+
^
-2^-

18
“ 5:^
+ ^
32 ^
2:122
28
^ o.0346 +
^18
The mol fraction of water vapor may then be written

?//18
=
0.0346 + (y/18)

Setting equal the two expressions for Xy,y

lb water vapor
y
lb mixture

The liquid water is the difference between the total water and the vapor;
0.074 — 0.042 = 0.032 lb of liquid water per pound of products mixture.
Since there are 17.15 lb of products per pound of fuel, the liquid water
formed per pound of fuel is 17.15(0.032) = 0.53 lb.

22-3 Products analysis. In an actual combustion process if no


more than the theoretical air is supplied, the reaction will not be
complete; the products will contain some carbon monoxide, unbumed
hydrocarbons or hydrogen. This is wasteful of fuel since, to obtain
a given amount of useful energy from a combustion process, more
fuel must be burned if the reactions are incomplete. Complete com-
412 COMBUSTION PROCESSES [22-4

bustion of fuel can be obtained only by supplying excess air; the


necessary excess depends upon the fuel and the apparatus used for
combustion, and may vary in practice from a few percent to 50
percent or more of the theoretical air. It is undesirable to supply
more air than necessai^ because the excess air, being taken in at a
low temperature and discharged at a high temperature, represents a
direct waste of energy.*
In order to determine the completeness of combustion and the
amount of excess air actually used in a combustion process the actual
products analysis is In Fig. 22-1 a general steady flow com-
used.
bustion process is represented diagrammatically. The products ap-
pear as flue gas and, in some cases, solid refuse. The flue gas com-

•FLUE GAS

COMBUSTION
CHAMBER
r SOLID REFUSE

Fig. 22-1. (’oinl)UHtion process; st(‘ady flow diagrum.

position may often be obtained by means of an Orsat gas analysis


apparatus, in which the components of the gas are selectively ab-
sorbed by chemical solutions. The usual Orsat analysis gives the
CO2 CO, O 2 and N 2
composition of the gas as fractions by volume of , ,
.

The analysison a dry gas basis, moisture being accounted for by


is

other methods. Small amounts of SO 2 are measured as CO 2 and are


ignored in the calculations.
When solid' refuse is appreciable as in coal burning furnaces, it is

analyzed by ^^igniting” a sample, i.e. heating it above the ignition


temperature in the presence of air for a long period. The loss of
weight of the sample is measured, and is taken to be carbon in the
refuse, while the residue is the ash in the refuse.
22-4 Material balance for a combustion process. The pro-
cedure for making the material balance is now given. Assume that
• It is sometimes necessary to use large amounts of excess air in order to
limit the temperature of the products to a desired value; this happens in gas
turbine power plants. On the other hand it issometimes desirable to use less
than theoretical air; this is done in internal combustion engines to obtain maxi-
mum. power output (but not maximum efficienc>0-
22- 4 ] COMBUSTION PROCESSES 413

the compositions of the four streams of Fig. 22-1 are as follows.


Fuel: C, H, O, N, S, A, M, fractions by ma.s.s;

Dry air plus water vapor, yA lb of wafer per lb of dry air.


Air:
Flue Gas: i'Oi, (D, 0«, 2 N
fractions by volume of clrj'^ gas, plus
,

(HjO)^, water vapor, lb water in flue gas per lb of fuel.


Refuse: Ar = ash, and Cn = carbon in refuse, fractions bj^ mass.
The basis of quantities for the calculation is one pound of fuel
burned.
Mass of Solid Refuse Per Pound of Fuel. The mass of refu.se is
obtained by an ash balance.

ash in refu.se = ash in fuel


niRAit = .1

where tur = mass of refuse, lb per lb of fuel. Then the mass of


refuse per pound of fuel is

The ma.ss of carbon in the refuse per pound of fuel is then ('ff(wK),
or Cr{A/Ar).
Mass of Dry Flue Gas Per Pound of Fuel. The ma.s.s of dry flue
gas is obtained by a carbon balance. The carbon in flic flue gas
per pound of fuel is the carbon in the fuel minus the carbon in the
refuse, or

Cg = C-Cht-
Ar
(')

where Vq is the mass of carbon in the flue gas per pound of fuel.
The gas analysis by mols is the same as the analysis by volume;
therefore the mass of each component per mol of gas is its fraction

by volume multiplied by its molecular weight. The mass of a mol


of dry flue gas is then

44 C()2 + 28 CO + 28 No + 32 O2
Also the mass of carbon in a mol of dry flue gas is

12 (COo + CO)

since each mol of CO 2 and of C'O contains 12 lb of carbon. Then


dividing the mass of a mol of gas by the mass of the carbon in a
mol of gas gives the mass of gas per pound of carbon in the gas.
414 COMBUSTION PROCESSES [m
Multiplying by the mass of carbon in the gas per pound of fuel gives
the mass of gas per pound of fuel.

ma p 3.67 CO2 + 2.33 CO + 2.33 Na + 2.67 O5


(2 )
CO2 + CO
where ma is the pounds of dry flue gas per pound of fuel.
Mass of Dry Air Per Pound of Fuel, The mass of dry air is ob-
tained by a nitrogen balance, equating the nitrogen entering to the
nitrogen leaving. Nitrogen in the air per pound of fuel is equal to
nitrogen in the flue gas per pound of fuel, minus nitrogen in the fuel.

Since the fraction by mass of nitrogen in the air is 0.768,

= ma ^
+ 32 O “
0.768m4
^ 28 CO +*28 Nj 2

where m^ is the mass of dry air per lb of fuel, and the fraction on
the right side is the fraction by mass of nitrogen in the dry gas.
Substituting for ma from (2)

28 NtCg
(0.768)(12)(C02 + CO) 0.768
^^ 3.04
CO2
NjCg
+ CO
_
0.768
...

Nitrogen in the fuel may often be negligible.


Mass of Waier in Flue Gas. The water in the flue gas per pound
of fuel is the sum of the moisture in the fuel, the water produced by
burning the hydrogen in the fuel, and the water in the air.

(H20)o = Af + 9 H + mAi/A (4)

where (H20)o is the mass of water in the flue gas per lb of fuel, and
7a is the speciflc humidity of the air, pounds water per pound of
dry air. The water from the air is often negligible.
The total mass of flue gas per pound of fuel is the sum of the dry
gas and the water in the gas.
Theoretical Air, The theoretical air is given by

theo. air = 11.5 C+ 34.2 (h - + 4.31 S ( 5)


f)
as shown in Example 2 of Sec. 22-2.
Excess Air, The excess air is the difference between ^e actual
air per pound of fuel and the theoretical air, expressed as a fraction
theoretical air.
22-4] COMBUSTION PROCESSES 415

excess air
ttiA — theo. air
(6)
theo. air

An approximate value for excess air, based on the flue gas analysis
alone, may be obtained if it is assumed that combustion is complete
except for CO in the flue gas, and that the nitrogen in the fuel is

negligible. On this basis the excess oxygen in the flue gas is that
which would remain after using sufficient oxygen from the flue gas
to bum all CO in the gas to COj. Each mol (or volume) of CO
requires one half mol
volume) of Os; then the oxygen needed to
(or
complete the combustion is 0.5 CO, and

excess Os = 0* — 0.5 CO
Since each mol of Os in air is accompanied by 3.76 mols of Ns,

excess Ns = 3.76 (Os - 0.5 CO)

Now by a nitrogen balance, if all the nitrogen came from air,

actual air _ actual Ns Ns


theo. air
~ theo. Ns Ns — 3.76 (Os — 0.5 CO)
The fraction of excess air is then, from (6),

actual air Os - 0.5 CO


excess air = ( 7)
theo. air
(Os -0.5 CO)

This equation, involving only data from the gas analysis, is com-
monly used for determining excess air in experimental checks on
steam-boiler furnace operation.

Example 4. From the data given below for a combustion process in a


boiler burning coal on a stoker, calculate the excess air by a complete mate-
rial balance and by the approximate method based on the gas analysis alone.
Fuel analysis, weight percent: C 76, H 4, N 1, S 1, A 12, M 2.

Flue gas analysis, volume percent: CO 2 14, CO 0.2, O2 4.8, Nj 81.


Refuse analysis, weight percent: C* 25, Ar 75.

Solution: Carbon in refuse;

Cb(A/Ar) = 0.25(0.12/0.75) = 0.04 Ib/lb fuel

Carbon in flue gas:

C -C-
(7 0.04 = 0.72 Ib/lb fuel
Mass of dry air:
416 COMBUSTION PROCESSES [22-5

3.04 NaCo _ N (3.04) (0.81) (0.72) _ 0^


COs + CO 0.768 0.14 + 0.002 0.768

12.44 Ib/Ib fuel


Theoretical air:

=
11.5C + 34.2<-l) + 4.31 S 9.95 Ib/lb fuel

Excess air:

- 0.251 or 25 percent
theo. air

By the approximate method,

= Go - 0.5 CO 0.048 - 0.001


Excess air = 0.28

^ - (O2 - 0.5 CO) - (0.048 - 0.001)

or 28 percent. The difference is due mainly to neglecting the combustible


in the refuse.

22-5 Energy equation for a chemical reaction. When a sys-


tem undergoes a process involving a chemical reaction the First Law
applies as for any other process. From the viewpoint of thermo-
dynamics a chemical reaction is a change of phase in a system
analogous to the physical phase change of vaporization. When one
pound of carbon combines with 2.G7 pounds of oxygen to form 3.G7
pounds of carbon dioxide, the material involved is still the same sys-
tem even though it has become a different chemical substance. It
would be possible to construct tables of properties for such a system
so that the internal energy change could be taken from the tables
for a process starting with the reactants at any given pressure and
temperature and ending with the products at any given pressure and
temperature. This would be analogous to the tables for liquid and
for superheated vapor in the steam tables. In general, however, such
highly organized data are not available, and a step by step procedure
is used as described below.
Consider a constant-volume reaction in the absence of motion;
the First Law for this process may be written

Q = E2-E, ( 8)

where E 2 is the internal energy of the products and E\ is the internal


energy of the rcacjtants.* It is clear, however, that values of E\ and
* Under the conditions here imposed E is the Ev of Sec. 5-5, but for simplicity
the .suhscrij)t T will be omitted.
22-5] COMBUSTION PROCESSES 417

E 2 taken from any arbitrary tables of data for the reactants and prod-
ucts would not necessarily satisfy the equation because the state of
zero E might have been chosen arbitrarily for each substance and
E ”2 El could have any value whatsoever. In order to be used di-
rectly in this equation, the data for products and reactants must
have a common base, just as the liquid and vapor tables for water
are referred to the same base state. The relation between data on
arbitrary bases and on a common base is illustrated in Fig. 22-2.

E vs t at constant volume E vst ot constant volume


(a) (b)

Fig. 22-2. E-t diagrams for a (*on.staiit-volume process involving a chemical


reaction.

At (a) is a plot of E vs, t at constant volume with the zero of E


taken arbitrarily at zero t for both phases. For all processes which

involve the products or the reactants individually this plot would be


correct, but for any process involving a change from one phase to the
other only one of the curves has an arbitrary zero point ;
the other

must be shifted vertically to account for the properly relationship


between the two phases. The proper relative position of the two
curves can be determined by an experiment in whi(*h the heat (pian-
tity of Eq. (8) is measured. As shown at (b) in Fig. 22-2, if states

1 and 2, and AEn can be measured, the relative positions of the


curves can be fixed.
In practice the experimental determination of the relation be-
tween the internal energies of products and reactants is carried out

so that the final temperature is close to the initial temperature. The


internal energy change AEb at a fixed base temperature tsy Fig. 22-2,
Fig. 22-3. Bomb calorimeter. A weighed sample of fuel is burned completely
inoxygen and the small temperature rise of the water and the bomb is measured.
From the measured temperature rise and the known heat capacity of the appa-
ratus the internal energy of combustion may be calculated. (Courtesy Emerson
Apparatus Company.)
418
22-6 ] COMBUSTION PROCESSES 419

is called the irUernal energy of reaction at The determination is

made by carrying out the reaction in a bomb calorimeter, Fig. 22-3,


and measuring the heat transferred out.*
Given the internal energy of reaction at some fixed temperature,
and the property relations for the reactants and products individu-
ally, it is easy to calculate the change of internal energy between any
two and 2, because the internal energy change is independent
states, 1
of the path. Taking a path composed of three steps, 1-1 b, 1f-2b,
2b-2, it follows that

E2 ~~ El = {E2 — Ezb) 4* {E2B ~~ Eib) + {Eib “ El) ( 9)

(E2 — E2b) can be found from the properties of the products,


(E2 B — Eib) is the internal energy of reaction, and (Eib — Ei) can
be found from the properties of the reactants.

Example 5. The internal energy of reaction for burning carbon to car-


bon dioxide at 68°F is —14,087 Btu/lb of carbon. Find the heat trans-
ferred when a system composed of 1 lb of carbon and 4 lb of oxygen at
300°F bums at constant volume to a mixture of CO 2 and O 2 at a final tem-
perature of lOOO^F. The specific heat of carbon is 0.17 Btu/lb F.

Solution: For a constant-volume process with no work the energy equa-


tion is

Q = E2 ~~ El = (E 2 — E2b) + {E2ji *" E\b) + {E\b — El)

The material balance gives for the burning of carbon with oxygen as specified

1 lb C + 4 lb O 2 3.67 lb CO + 2 1.33 lb O 2

Taking the separate terms of the ejiergy equation,

E — Ew =
2 (f/2 — 1 /25) for O 2 + (U 2 "" for CO 2

= ^,(Cv)Oay2 — Ib) + WCOi(Cv)cOs(^2 — Ib)

From the material balance mot = 1.33 lb and wicoa = 3.67 lb; taking the
specific heats as (cv)oj = 0.155 Btu/lb ®F and (Ct)coa = 0.165 Btu/lb °F,
it follows that

E - 2 Eib = (1.33)(0.155)(1000 - 68) + (3.67) (0.166) (1000 - 68)

= 755 Btu

*
In a constant-temperature process the internal energy of substances that
do not react chemically is almost constant, assuming no phase changes occur.

Therefore the inU^rnal energy of reaction depends only upon the reacting sub-
stances in the system, and is independent of the amount of non-reacting sub-
stances, such as excess air, in the system.
420 COMliUSTION PROCESSES [22-5

EB —
2 E\b is tlic given internal energy of reaction, —14,087 Btu/lb of
carbon, multiplied by the pounds of carbon, or — 14,087 Btu.
Eib — El = {ViB — Ui) for carbon + (L'l^ — L^) for O 2

= mc(c)c(^B ^i) + h)

= (1)(0.170)(GS - 300) + (4)(0.155)(68 - 300)

= -183 Btu
Then Q = 7o5 - 14,087 - 183 = -13,315 Btu
This is a heat transfer from the system.

The relations between enthalpy of rear‘tants and enthalpy of prod-


ucts are similar to the internal energy relations. The enthalpy change
for a process involving a reaction may he written

H -
2 Hi = (II 2 - + (H.,^ - Ihn) + (Hib - Hi) (10)

where the terms are ana-


logous to those of Eq. (9).

The enthalpy relation is

illustrated in Fig. 22-4.


The eniluilpy of reaction

at Ib (H 2 /?
— Hib) may, in
principle, be found by
carrying out a steady flow
reaction at tn, making shaft
work, kinetic energy change,
and potential energy change
negligible, or by carrying
out a constant-pressure re-
Hg. 22-4. H-t diagram for a constant-pres-
action in a stationary sys-
sure proeoss involving a chemical reaction.
tem in which all work is

p dv work. In either case a measurement of the heat transfer will


give the enthalpy change:

Q = H2-Hi (11)

is not measured
In most cases in practice the enthalpy of reaction
directly but computed from the measured internal energy of reac-
is

tion. Gases are an exception; their enthalpies of reaction are often


•measured by a steady flow calorimeter process. The computation
of enthalpy of reaction from internal energy of reaction is based upon
22-5] COMBUSTION PROCESSES 421

the definition of enthalpy, Eq. (5-13), and the assumption that the
gaseous reactants and products follow the perfect gas laws, while
the solids and liquids involved have negligible volume. From (5-13)
(noting that the E of this chapter is the Eu of Chap. 5), for any
reaction,
H ”
2 Hi = E —2 El P 2 V2 — piVi (12)

Let states 1 and 2 be the initial and final states of a reaction at Ib


and at some constant pressure p; then

H2 — Hi = ^Ilji

Now since the products follow the perfect gas law, (or are solid or
liquid) the internal energy of the products depends only upon tem-
perature. Eo is therefore the same at the end of the given constant-
pressure reaction as it would have been at the end of a constant-
volume reaction at the same temperature, and

E2 — El = AEb
This is true even though state 2 is not the state 2 of the constant-
volume reaction by which AEb was determined. To find a value for

(p^Vz — piVi) observe that the perfect gas equation on the mol basis
is independent of the chemical composition of the gas; thus for both
the gaseous products and the gaseous reactants

pV = uRT
At the base temperature Tbj if only the gaseous components have
appreciable volume,

(P2F2 - PiFi)^ = RTjiinz - ni)

where Ui is the number of mols of gaseous reactants, and tiz is the


number of mols of gaseous products. Substituting in (J2),

Alls ~ AEb RTb{u2 — Til) (^ 3 )

The use of Eq. (13) is illustrated in Example 7 in the next section.


The terms internal energy of reaction and enthalpy of reaction have
general application to reactions ofany kind, but for special kinds of
reaction special names are often used. Enthalpy of formation and
enthalpy of combustion are used for the case of formation of a com-
pound from its elements, and combustion of a fuel, respectively.
;

422 COMBUSTION PROCESSES [ 22-6

The terms constanirpressure heat of reaction and constant-pressure heat-


ing value are used to signify the negative of the enthalpy of reaction.
For example, the same information is conveyed by any one of the
following statements:
(a) C+O CO2 2 ;
A// = —94,052 cal/gm formula weight.
(b) C+ O —> CO 2 2 ;
constant-pressure heat of reaction is 94,052
cal/gm formula weight.
(c) The enthalpy of formation of CO 2 is —94,052 cal/gm mol.
(d) The constant-pressure heat of combustion (or heating value)
of carbon, burned to CO 2 , is 14,087 Btu/lb.
The units in the statements above are related as follows: 1 gm
formula weight of carbon in the equations above is 12.010 gm, or
is the mass of 1 gm mol, since the formula contains
0.02648 lb; this
1 mol of carbon. Since 94,052 cal/gm = 169,180 Btu/lb, 94,052
cal/gm formula weight = 14,087 Btu/lb.
Values of heat of combustion for some important substances are
given in Table A-2.
To illustrate thejabove discussion an example will be given of the
computation of enthalpy of reaction for a given reaction from data
on other reactions involving the same substances.

Example 6. Find the enthalpy of combustion for methane CH4 at 77®F


and 1 atm, given the following enthalpies of formation at 77®F and 1 atm:

H2O —68,317 cal/gm formula wt

CO 2 —94,052 cal/gm formula wt

CH4 — 17,889 cal/gm formula wt

(Enthalpies of formation are found in such sources as the International


Critical Tables.)

Solution: The basis of the solution is the fact that the enthalpy change
is independent of the path of a process. Instead of burning methane di-
rectly, substitute a series of reactions between the same end states, as
follows:

(1) Break up methane into its elements (the enthalpy change is the
negative of the enthalpy change in the formation of methane)

CHi C + 2 Ha; A// = 17,889 cal/gm formula wt

(2) Combine the resulting elements with oxygen (since 2 H O is involved,


2

A/f is twice the enthalpy of formation for H O);


2

C -f On — CO 2 ;
Aff = —94,052 cal/gm formula wt
23-6] COMBUSTION PROCESSES 423

2 + Oj - 2 H,0; Afl = 2(-68,317) cal/gm formula wt

The overall effect of these processes is the same as the effect of the combus-
tion of methane,

CH« -b 2 Os -» CO, -1- 2 HA AH = x


but in the substitute process the enthalpy change for each step is known.
By adding the enthalpy changes algebraically the overall change is obtained.
I = AH = 17,889 - 94,052 - 2(68,317)

= —212,797 cal/gm formula wt

Then the constant-pressure heating value of methane is 212,797 cal/gm


formula wt, or 23,860 Btu/lb.

22-6 Heating values of fuels. The heat of combustion or heat-


ing value of a fuel is generally given for a reaction starting at 1 atm
and 20‘’C or 25‘’C. In tables of scientific data for reactions between
definite chemical substances the temperature and pressure are always
specified, but in engineering data on fuel samples these details are
not usually given. Since heating values do not change much with
pressure, nor with temperature in the range of room temperatures,
the heating value is assumed to apply to any normal room condition.
It is necessary, however, to distinguish between different heating
values which may be obtained for a single fuel, depending upon the

conditions of the reaction. For a fuel containing no hydrogen there


may be at least two heating values, the constant-pressure heating
value (— AHb), and the constant-volume heating value (— A£b).
These may sometimes be equal, as in the case of carbon. For the
reaction C
-b O,
—» CO, there is one mol of gas on each side of
the equation; hence by Eq. (13) AHb = AEb and the constant-volume
heating value is equal to the constant-pressure heating value. For
the combustion of carbon monoxide to carbon dioxide the two heating
values are different.

Example 7. Given the reaction 2 CO -f 0, - 2 CO, and the constant-


pressure heating value, 4,344 Btu/lb of CO, find the constant-volume heating
value.

Solution: The mol baas, using Eq. (13), is most convenient. The two
mole of CO contain 56 lb, so the heat of reaction is (56) (4,344) =*> 243,300
Btu/lb formula weight. There are 3 mols of ga^us reactants and 2 mols
of gaseous products; then by Eq. (13), taking R = 1.986 Btu/lb mol^R,
and taking Tb as 530°R,
424 COMBUSTION PROCESSES L22-6

^EB = -243,300 - (1.986) (530) (2 - 3)

= -243,300 + 1050

= —242,200 Btu/lb formula wt

Then the constant-volume heating value is 4,325 13tu/lb of CO.

For fuels containing hydrogen another variable is encountered in

the fact that the water in the products may be either liquid or^vapor;
this may have an appreciable effect upon the final properties of the
system. Suppose hydrogen is burned to water under such conditions
that all the water vapor formed is condensed to liquid.* The chem-
ist writes for this reaction at 77°F

Ha + \ Oa — > HaO(0 ;
A// = “68,317 cal/gm formula wt

where the {1) signifies that the water is finally all liquid. The engi-
neer computes from this that the heat of combustion of hydrogen to
liquid water is 60,958 Btu/lb of hydrogen. This is called the higher
heating value (HHV) because it includes the heat transferred out from
the system as the result of the condensation of the HaO, and is

therefore larger than if the final state had been vapor. The difference
is appreciable; each pound of hydrogen produces 9 pounds of water
and at 77°F the enthalpy of vaporization is 1050.4 Btu/lb. There-
fore if the reaction had ended with water vapor at 77°F, instead of
liquid, the heat transferred out would have been reduced by 9 X 1050.4

or 9454 Btu/lb of hydrogen. The lower heating value (LHV) is given


by
LHV = HHV - mu^hfa (14)

where mHjO is the mass of HO


2 produced per pound of fued, and h/g

is taken at the base temperature. t Then for this case

LHV = 60,958 - 9(1050.4) - 51,504 Btu/lb hydrogen

It should be observed that the higher and lower heating values


obtained above were both at constant pressure. It is possible also

to determine higher and lower heating values at constant volume;


for this case

* This is done in calorimetric practice by putting a drop of liquid water in


the calorimeter bomb so that the space is already saturated with water vapor
before the reaction, and any additional vapor formed must condeiKst*.
tin practice it is usual to take h/g aa some fixed arbitrary value such as
1060 Btu/lb.
22-7] COMBUSTION PROCESSES 425

(LHV). = (HHV), - (15)

The higher heating value gives the maximum heat which could
be transferred to constant-temperature surroundings from a given
reaction. In practice, however, the heat of vaporization of the water
in the products is not utilized, first, because serious corrosion problems
usually arise if moisture is permitted to condense from flue gas and,
second, because low temperature heat is of small thermodynamic
value. Since in most actual processes the moisture will be vapor, it

is convenient to use the lower heating value in computations involv-


ing the properties of the products. In experimental measurement of
heating value, however, it is easier to insure complete condensation
of the water than complete vaporization. Therefore the higher heat-
ing value measured in the calorimeter, and the lower heating value
is

is computed.

The relation of higher and lower heating values to the efficiency


of combustion processes is discussed in Sec. 22-8.
Some fuels may be supplied to a combustion process either as
liquid or as gas; in such cases the heating value of the fuel is taken
for liquid or gaseous fuel as the case may be. The two values will
differ by the latent heat of vaporization For any ordinary
of the fuel.
fuel this is a smaller difference than that resulting from the latent
heat of the water in the products. For example, Keenan and Kaye
give for the enthalpy of combustion at 25‘’C of octane CgHis to gaseous
products, — 19,100 Btu/lb for liquid octane, and —19,256 Btu/lb for
gaseous octane. These numbers (taken positive) are both lower heat-
ing values; the corresponding higher heating values are 1,490 Btu/lb
greater, or 20,590 and 20,746 Btu/lb re.spectively.

Heating values for some typical fuels are given in Table A-2.
22-7 Temperature of products. The temperature of the prod-
ucts of combustion is important in the design and operation of
combustion apparatus. In some cases the maximum possible tem-
perature may be desired, but for power-plant purposes the problem
is usually to limit the temperature of the products to some desirable
range, neither too high nor too low. From the thermodynamic view-
point high temperature is desirable, but in practice, temperature must
be limited to avoid such effects as slag deposits in steam boilers or
overheating of metal parts of gas turbines.
The temperature of the products of a combustion process in which
426 COMBUSTION PROCESSES [22-7

there is no work other than p do work is a function of the reactions

involved and the heat transferred. In Fig. 22-5 the temperature


change and enthalpy change
are plotted for a given
reaction with three different
amounts of heat transfer.
For the process 1-^, the
temperature is constant;
this is the calorimeter type
of process in which all the
heat of combustion is trans-
ferred from the system. In
the process 1-3, the heat
H vt 1 ot constant pressure t transferred is less than the

Fig. 22-5. Temperature change in a combus- heat of combustion and


tion process. there is a temperature rise
in the system; this is the
usual type of combustion process. In the process 1-4, there is no
heat transfer and H 4 is equal to Hi; the temperature h is called the
temperature of adiabatic combustion. may be
This temperature
approached in uncooled combustion chambers for producing hot gas
streams, such as the combustion chambers of gas turbine power plants
or rockets.
The factors which influence the products temperature may be
illustrated by an example.

Example 8. Assume carbon burns with air to carbon dioxide in a steady


flow process, Fig. 22-6. If theoretical air is used, calculate the
products
temperature for adiabatic combustion, assuming the products have constant
specific heats of room temperature magnitude.
22-7] COMBUSTION PROCESSES 427

Solution: + 3.76 N CO + 3.76 N


C+O 2 2 2 2

1 lb + 2.67 lb + 8.78 lb 3.67 lb + 8.78 lb

Q — H — Hi = H — H B II B — HtB Hib
2 2 2 *f* 2 -f" Hi
Assume that h = (b = 77®F; then Hib — Hi is zero and Hzb - Hib is
— 14,087 Btu/lb of carbon.

Then , Q = 0 = H2 - Haa - 14,087

H - 2 Hsb = 14,087

The enthalpy change of the products may also be written in terms of the
specific heats and temperature change:

H2 - H2B = [(mCp)cOa + (wiCp)Nj(f2 - Ib)

14,087
Then I2 “ fa = 4,900®F
(3.67) (0.196) + (8.78) (0.248)

Assuming is approximately 100°F, ^2 = 5000®F.


The temperature found above would never be reached in an actual process
for several reasons.
Excess Air. In an actual process complete combustion will not be ob-
tained without some excess air. This will increase the mass of products,
thereby reducing the temperature rise for a given enthalpy of combustion.
Heat Transfer. In an actual process the heat transfer will not be zero.
Variation of Specific Heats. The specific heats of the gases are greater
at higher temperatures, giving a smaller temperature rise than calculated
above. This may be shown by taking average specific heat values for the
conditions of Example 8 from Fig. A-4, p. 507; taking for carbon dioxide
Cp = 0.300, and for nitrogen 0.285,

4 ^ 4 _ 14,087 __ o Qin®'!?
'
(3.67) (0.300) + 8.78(0.285) ^

Then <2 = 4000‘’F

Chemical Equilibrium or Dissociation. The combustion reaction does not


go to completion at high temperatures. The composition of the products tends
to approach a certain equilibrium state which depends upon the relative
quantities of the various chemical elements present and upon the tempera-
ture and pressure. At low temperatures the equilibrium condition for carbon
and oxygen, if suflScient oxygen is supplied, is carbon dioxide. At high tem-
peratures the equilibrium state is a mixture of carbon dioxide, carbon mon-
oxide, and oxygen. The higher the temperature (and the lower the pressure)
the greater is the fraction of carbon monoxide for equilibrium. If sufficient

oxygen is supplied to form carbon dioxide, there will be negligible carbon


monoxide at equilibrium for temperatures below 3500®F. At 5000®F and
428 COMBUSTION PROCESSES [22-7

1 atm there will be approximately as many mols of carbon monoxide as of


carbon dioxide. TKe heat of reaction for carbon burned to carbon monoxide
is only about 28 percent of that for carbon burned to carbon dioxide; hence
for temperatures above 3500°F the temperature actually reached in the
combustion of carbon will be lower than that calculated on the assumption

of complete combustion. At temperatures above 4500°r the combustion of


hydrogen begins to be appreciably incomplete.
The term dissociation is often used in connection with equilibrium prob-
lems, but it has a rather restricted implication which may be misleading.

Chemical Equilibrium. Determination of equilibrium conditions is one of the


important problems of chemical thermodynamics; its solution is necessary to
define the conditions under which a proposed reaction is possible, and the extent
to which it can proceed. Like other consequences of the Second Law the condi-
tions of equilibrium are limiting or bounding conditions. A system need not
always be in its equilibrium state; for example carbon can exist in contact with
oxygen at room temperature without forming carbon dioxide. However, if a
change is started (as by introducing a catalyst, or by raising the temperature to
the ignition temperature) the direction of the change will be toward the equilib-
rium state.

At high temperatures the combustion reactions can proceed toward equilib-


rium as fast as the various atoms and molecules can come in contact with the
other atoms and molecules with which they must react As the temperature
is lowered the speed of the reactions decreases, becoming effectively zero at some

temperature. Therefore if the products of combustion are cooled rapidly, the


reaction may stop before the combustion is complete; the tendency toward com-
pletion still exists, but the rate of reaction is zero. Such a condition is sometimes
designated by the term frozen equilibrium; a more rational name would be frozen
disequilibrium.

In steady flow burning of carbon in air, Example 8, the final tem-


perature will not be so high that equilibrium will play an important
part in limiting the temperature; the limiting factors will be the frac-
and the amount of heat transfer. If insufficient
tion of theoretical air
some carbon monoxide will be formed even at low
air is supplied,
temperatures and the products temperature will be correspondingly
reduced.
In constant-volume burning the calculated temperature rise will

be greater than for steady flow. For burning carbon completely in


theoretical air at constant volume,

-AEb = 14,087 = [(mc„)co, + (mc,.)N ,](/2 - W


Using average values of if the reaction goes to completion.
22- 8 ] COMBUSTION PROCESSES 429

^ (3.67)(0.255) + 8.78(0.215)

Then <2 = 5100°F

Thus, in constant-volume burning, the temperature rise may well be


sufficient for equilibrium to have the controlling influence on the
maximum temperature attainable. This is the case in many internal
combustion engines.
When the products temperature is limited by equilibrium, a
cut-and-try procedure is necessary to find the final temperature be-
cause the temperature determines the products composition and the
composition determines the temperature. The necessary data for
such computations may be found in treatises on physical chemistry
or chemical thermodynamics. For reactions of great technical im-
portance, specifically, combustion in the internal combustion engine,
the results of such computations are available in the literature.
Products of combustion that are dissociated at a high temperature
may recombine as the temperature drops, provided sufficient time is

available before the temperature becomes too low for the maintenance
of combustion. The time needed depends upon the amount of excess
oxygen present and upon the degree of mixing of the gases, as well
as upon the pressure, since all these factors influence the probability
of the molecules of oxygen meeting the molecules of carbon monoxide,
for example.
In steam boiler furnaces it is usually practical to provide sufficient
time and oxygen to obtain substantially complete combustion before
the gases leave the furnace and enter the relatively cold tube banks
of the boiler. Provided combustion is finally complete, the existence
of dissociation in the hot zones of a boiler furnace does not reduce the
available energy from the combustion, since the energy is later to
be transferred as heat to relatively low-temperature steam. The
thermodynamic advantage of the high temperatures reached in the
furnace would be wasted by the irreversible heat transfer, regardless
of dissociation. In an internal combustion engine, however, some
work would be lost through dissociation even if combustion were
finally complete, because the reduction of the maximum temperature
of the gases would reduce the work available from their expansion.
22-8 Efficiency of processes involving combustion. The effi-
430 COMBUSTION PROCESSES [22-8

ciency of processes involving combustion may be defined in many


different ways, depending upon the purposes of the processes and the
arbitrary definitions of input and output.
Considering the combustion process alone, without regard to heat
or work effects, an efficiency may be based upon the fraction of com-

plete combustion obtained. One way to reduce this to numbers is


to take for the numerator of the efficiency ratio the heat of reaction
for the reaction that actually occurs and, for the denominator, the
heat of reaction for complete combustion.Another way to state this
is to take for the numerator the mass of
would be required
fuel that
to produce the actual products temperature with complete combus-
tion and, for the denominator, the actualmass of fuel. This type of
efficiency may
be used to describe the performance of adiabatic com-
bustion chambers for the production of hot gas streams, as used in
internal combustion turbine power plants.

Example9. Suppose one pound of carbon bums at constant pressure


so that 0.9pound goes into COj, 0.05 lb goes into CO, and 0.05 lb emerges
as unburaed carbon (these fractions are determined by a products analysis);
find the efficiency of the combustion process.

Solution; For complete combustion of the pound of carbon the heat of


reaction is 14,087 Btu; for combustion of 0.9 Ib of carbon to COt the heat
of reaction is 0.9(14,087) or 12,678 Btu; for combustion of 0.05 lb of carbon
to CO the heat of reaction is 0.05(3952) or 198 Btu. Then the efficiency on
the basis here considered is (12,678 -f 198)/14,087 or 0.91.

In reciprocating internal combustion engines the purpose of com-


bustion is the production of work by a process which is not easily

broken down into separate parts; therefore it is convenient to define


an efficiency for the process as a whole. In this case the numerator
of the efficiency ratio may be taken as the work output of the engine
per unit of fuel supplied, while the denominator may be taken as
the heat of reaction for complete combustion of thefuel. For exam-
an engine consumes 0.5 lb of fuel per horsepower-hour of work
ple, if

output, and the heating value of the fuel is 19,000 Btu/lb, the effi-
ciency is 2545/(0.5)(19,000), or 0.257.
In steam boiler furnaces the purpose of the combustion process is
to supply heat to water; the efficiency of the overall process in a steam
generator is defined as the ratio of the heat transferred to the water
divided by the heating value of the fuel.
2M] COMBUSTION PROCESSES 431

Example 10. A steam generator produces 10,000 Ib/hr of steam at


150 psia, saturated vapor. In addition there is drawn off from the boiler
600 Ib/hr of saturated liquid at 150 psia as “blow down” (this water carries
with it the dissolved solids which would otherwise accumulate and precipitate
as scale in the boiler). The feed water is supplied to the boiler at 210^F.
Oil fuel having a heating value of 19,500 Btu per pound is burned in the
furnace at the rate of 700 Ib/hr. Find the efiBciency of the steam generator,
as defined above.

Solution: The heat transferred to the water is found by applying the


steady flow energy equation and mass equation; referring to the sketch,

Wi W2 + Wb

w\h\ +Q = W2h% + witib


Substituting values from the data and the steam tables

Q= 10,240,000 Btu/hr

The heating value of the fuel burned, on an hourly basis, is

(700) (19,500) = 13,650,000 Btu/hr

efficiency *= 1024/1365 = 0,75

In the discussion above no consideration was given to the ques-


tion of which of the various heating values should be used as a basis
for input in efficiency computations. This choice is arbitrary, since
an efficiency is simply a comparison to an arbitrary standard. Since
most technical combustion processes are associated with operations
which are essentially of the steady flow type, it is customary to use
constant-pressure heating values, which represent enthalpy changes.
432 COMBUSTION PROCESSES
This applies even to engines in which the combustion process proper
is a constant-volume process, since from the external viewpoint such
engines operate on the steady flow basis. Another choice remains,
in the case of fuels containing hydrogen, of using either the higher
or lower heating value; thus two different efficiency values may be
obtained for the same process.
In steam power plants in the United States the higher heating
value is used as a basis for efficiency because it is the experimentally

measured heating value and it gives the lower value of efficiency.


Internal combustion plant efficiencies and European steam plant effi-
ciencies arc often based on lower heating values, which give larger
numbers for the same performance. Because of such variations in
practice, quoted efficiency values should not be compared indiscrimi-
nal ely.

rnOBLEMS*
22-1. (a) Find the '‘thoorotiral” quantity of oxygon for buniing 1 Ib of
aeotylene, C2H2. (b) Find the analysis by mass of the products of complete
combustion of acetylene with 125 p(*rcent theoretical oxygen.
22-2. A gaseous fuel consists of a mixture of methane, CH4, and carbon
monoxide, CO, in equal parts by volume. Find the th(‘orotical air, lb per lb of
fuel, and the analysis by mass of the products of complete combustion of the
fuel with 110 percent theorcitical air.
22-3. A
power plant burns coal having an analysis as follows: C, 71.04;
H, 5.28; 0, 6.72; S, 4.96; N, 1.34; A, 10.66; in percent by mass. The combus-
tion equipment will operate cfTiciently with 20 percent excess air. For this case
find the pounds of air per pound of fuel, and the analysis by mass of the flue
gas, assuming combustion is complete and everything but the ash appears in
the flue gas.
22-4. For the combustion process of Problem 22-3 find the saturation tem-
perature of the water vapor in the products for a total pressure of 15 psia. (The
actual dew point of the flue gas will be higher than determined hero; because of
sulfur compounds in the gas the condensate will not be pure water but will be
acid, wdiich has a higher saturation temperature.)
22-5. A coal had the following analysis: C, 62.40; II, 5.05; O, 16.16; S, 1.20;
N, 1.12; A, 14.07 percent by mass. When this coal was burned in a boiler fur-
nace the gaseous products analysis (dry basis) was as follows: CO2 12.10, O2 6.93,
CO 0.19, No 80.78 percent by volume. The solid njfuse was, by mass, 22,1 per-
cent combustible (assumed carbon). The air supply had 60 percent relative
humidity at 75®F. Find: (a) the percentage excess air used; (b) the pounds of
dry flue gas per pound of fuel (c) the total pounds of flue gas per pound of fuel.
;

22-6. Carbon monoxide is burned completely with theoretical ox^sgen in

*
Heats of reaction at 25®C, 1 atm, are given in Table A-2, p. 504.
COMBUSTION PROCESSES 433

steady flow. The reactants are supplied at 400®F, 20 psia, and the products
leave at 600°F, 16 pwsia. Velocities are negligible and there is no shaft work.
Assuming all the gases are perfect gases find the heat transfer, Btu/Ib of CO,
for the process.
22-7. Hydrogen is burned with 50 percent excess air in steady flow. Dry
air and hydrogen are supplied at 77®F; the products leave at 77°F. Velocities
are negligible and there is no shaft work. Find; (a) the pounds of liquid water
in the products per pound of hydrogen; (b) the heat transfer per pound of hy-
drogen. The process occurs at atmospheric pressure.
22-8. For the fuel of Problem 22-2 find the constant-pressure heat of com-
bustion, Btu/Ib of fuel, at 77°F, (a) if the water in the products is liquid, and
(b) if the water in the products is vapor.
Hydrogen peroxide, H2O2, is to be used as a source of oxygen for a
22-9.
power plant which is to be independent of an air supply. The following reaction
will take place:
C + 2 H2O2 (liq) 2 H2O -h CO2
Enthalpies of formation, calories per gram formula weight, at 68°F:

H2O2 (liq) -44,500; H2O (liq) -68,320; CO2 (gas) -94,050.

Find (a) the pounds of hydrogen peroxide needed per pound of carbon (b) the
: ;

enthalpy of reaction, Btu per lb of carbon, at 68'’F, for the reaction, if the water
in the products is all liquid.
22-10. The coal of Problem 22-3 has a higher heating value of 13,025 Btu/lb.
Find its lower heating value.
22-11. According to US Bureau of Standards Circular r461, methane, CH4,
has a constant-pressure higher heating value at 25°C of 23,861 Btu/lb. Assum-
ing that all gases involved are perfect gases, and that liquid water has negligible
volume, find for methane at 25®C: (a) the conbtant-pressure lower heating value;
(b) the constant-volume higher heating value; (c) the constant-volume lower
heating value.
22-12. The enthalpy of combustion of octane vapor, CsHis, at 25'’C, is

—20,747 Btu/Ib when all the water in the products is liquid. The enthalpy of
vaporization of octane at 25°Cis 156 Btu/lb. In a certain process octane, sup-
plied as liquid at 25°C, burned completely with 400 percent theoretical air
is

supplied at 25°C. There is no heat transfer, but work is transferred out. If


the gaseous products of combustion are discharged at 500°F, how much work
is done per pound of octane burned?

22-13. A certain fuel oil is 84 percent carbon and 16 percent hydrogen by


mass, and the constant-pressure heating value (to gaseous products) is 19,000
Btu/Ib at 70®F. The oil is burned in an adiabatic gas-turbinc combustion
chamber with sufficient air so that the products temperature is 1450°F. Air
and oil are both supplied at 70°F. Using average specific heats from the plots
in the appendix, find the air-fuel ratio required if combustion is complete.

22-14. The fuel of Problem 22-13 is burned completely with 125 percent
theoretical air, and liquid water at 70°F is then injected into the combustion
products so that the final temperature of the mixture, after the water evaporates,
is HSO^F. How mueh water must be injected per pound of fuel?
434 COMBUSTION PROCESSES
22-15. A diesel engine uses 29 lb of fuel per hour when the brake output is

75 hp. If the heating value of the fuel is 19,700 Btu/lb what is the brake thermal
efficiency of the engine?
22-16. On the basis of reduction of heating value due to mcomplete combus-
tion, find the efficiency of combustion for the process described in Problem 22-5.
The heating value of the fuel is 10,800 Btu/lb.
22-17. In the process of Problem 22-5 the fuel consumed was 140,000 lb {n
24 hours; the output of the boiler in the same period was 900,000 lb of steam
at 260 psia, 550°F, from feed water at 240'’F. The heating value of the fuel is

10,800 Btu/lb. Find the efficiency of the complete steam generator.


Chapter 23

INTERNAL COMBUSTION
POWER PLANTS

Internal combustion power plants include both reciprocating engines


and turbine plants. An important difference between these types of
plant, from the viewpoint of thermod 3mamic analysis, is that the
reciprocating engine carries out in one complex process compression,
combustion, and expansion of the working fluid, whereas the turbine
plant uses physically distinct processes.* In this chapter examples
of the analysis of both types of plant will be given. Knowledge of

the material in Chaps. 16 and 22 is assumed in this chapter.

23-1 Spark-ignition engine process. In Chap. 16 several air-

standard cycles for reciprocating engines were described. These


cycles are useful only as a rough approximation; for more exact
analysis ideal fuel-air processes are used, as illustrated below. Con-
sider a four-stroke engine using a pre-mixed charge of fuel and air
with spark ignition. The ideal plant process on which the analysis
is based is the following (Fig. 23-1):
Intake. Beginning at 6, with the minimum cylinder volume full

of combustion products at the exhaust pressure p, the products ex-


pand reversibly and adiabatically until the pressure drops to the
intake pressure p< at 6' pi than p« because of pressure loss in
;
is less

the piping and the engine throttle valve. During the remainder of
the stroke, 6' to 1, the fresh charge of fuel and air comes in at con-
stant pressure. There is no heat transfer with the walls.
Compression. From 1 to 2 the mixture of air, fuel, and combus-
tion products is compressed reversibly and adiabatically.

* A plant could be built to use separate reciprocating machines for compres-


sion and expansion, and a separate combustion chamber, but this would not be
an internal combustion engine in the usual sense of the term. Such a plant
would be analyzed in the same way as the turbine plant.
435
436 INTERNAL COMBUSTION [23-1

Combustion. From 2 to 3 the mixture burns adiabatically at


constant volume.
Expansion. I'rbm 3 to 4 the products expand reversibly and
adiabatically to the maximum cylinder volume.

PISTON TRAVEL

Fig. 23-1. Idealized indieator diagram; spark ignition engine with intake
throttled.

Blowdown. iTom 4 to 5 the products blow out into the exhaust


pipe, the part remaining in the cylinder being assumed to expand
reversibly and adiabatically. The pcint 5' indicates
the end point
of the expansion if the whole ma.ss of gas had gone to the same end
statue as the residual gas.
Exhaust. From 5 to 6 the products are discharged at constant
pressure without state (diange of the part remaining in the cylinder.
The work of the plant process is given by

W =
W = mrX'ih “ WeO + pi{Vi - FeO

+ m(ui - U2) + m{uz - Ui) + pe{V^ - 7 ),


b ( 1)

where Mr is the mass of combustion products in the clearance volume


at state 6 (residual gas), and m is the sum of me and the mass of

fresh charge m,-. The properties at the various states may be evalu-
ated by detailed computations based upon the composition of the
fuel-air mixture, and equilibrium data, but this is an exceedingly
tedious procedure. For certain fuel-air mixtures, however, charts of
properties have been published an example is Chart E-5 in the back
;
23-1] INTERNAL COMDUSTION 437

cover envelope.* These charts are on coordinates of internal energy


E vs. entropy 5. The small chart is for the fuel-air mixture, or react-
ants, and the It is assumed that the prop-
large chart for products.
erties of the fuel-air mixture are correct to sufficient accuracy for
any fraction of residual combustion products that may be mixed with
the fresh charge. The '^sensible'^ internal energy U,, and enthalpy
Hm, refer to undissociated mixtures, while the ‘'total'^ internal energy
E, and enthalpy 77, refer to mixtures dissociated according to the
equilibrium conditions. The charts give properties for the total mass
of gas mixture, but on the basis of one pound of mixture substance
which was air when it entered the engine. Thus, if the mixture is
10 percent fuel, the enthalpy which corresponds to 100 Btu per lb of
mixture would be plotted as 100/0.9 or 111.1 Btu per lb of air.
The following example is given to show the use of the charts.

Example 1. Given an ideal plant process such as shown in Fig. 23-1,


with piston displacement Vi — Vi — 0.120 cu ft, compression ratio r = 6,
and a fuel-air mixture of octane Cgllig with ICO jxircent theoretical air; the
exhaust pressure p* ~ 15 psia, and the intake pressure p< = 12 psia, due to
throttling of the engine; the intake temperature Tx = 600°R.

Solution: In order to fix the states of the f.uid it is necessary to proceed


through the process from inlet to exhaust. A complication is encountered
at the beginning because the mass and temperature of the residual gas in
the clearance volume will not be known until the end of the computation,

yet they influence the whole process; therefore a cutr-and-try procedure is

necessary. Assume, subject to check at the end, that the ratio of clearance
gas mass to total gas mass, rric/m = 0.05, and that the temperature, Tty is
2600°R.
In the process 0-G' the clearance gas expands iscn tropically from 15 psia,
2600°R, to 12 p.sia. On the chart for products this gives = 2480®R,
= 82 cu ft, 4G5 Btu, and = 655 Btu. {E = jB„ and H= //.,

at this low temperature.)


Since the intake process 6-1 is a frictionless adiabatic constant- pressure

process, the enthalpy is constant for a system comprising the total mass
finally in the cylinder. Then
mill = mjl^^ + milli

or, since tne/m —/


/A =/i7fl,-h(l-/)//i
*
For more complete charts see Hottel, H. C., and others, Thermodynamic
Charts for Combustion Processes. New York: Wiley, 1949.
P
438 INTERNAL COMBUSTION [23-1

Hi is taken from the chart for reactants, where it appears as » 58 Btu


at 600®R. Then
Ih = 0.05(655) + 0.95(58) 87.7 Btu
From the chart for reactants at //. = 87.7 and p = 12 psia, * 37
Btu, Fi = 22 cu ft, Ti = 705^R.
In the isentropic compression process 1-2 the volume decreases to Fi/6,
or 3.67 cu ft per lb of air in the mixture; from the chart for reactants E2 » 155

Btu, T2 = 1250®R, p 2 = 125 psia.


For the combustion process the absolute value of the internal energy of
reaction must be added to E2 to get Ei of the products. The E of combus-
tion as given on the chart of reactants* is

1280(1 -/) « 1216 Btu

thep Ei = 1216 + 155 = 1371 Btu.

From the products chart at this value of E, and at F= 3.67 cu ft,

Ti = 4895®R, Pa = 560 psia, and Sa == 0.555 Btu/®R


After isentropic expansion to Fa = Fi = 22 cu ft,

Ti — 3610®R, Ei « 798 Btu, pa = 69 psia

After further isentropic expansion to pa - 15 psia,

Ti = 2640°R, Ei = 520 Btu, and Fg^ = 73 cu ft

Fg' is the volume that the entire charge of gas would have occupied if it had
expanded to ps, Ti.

In the exhaust process 5-6, the gas is displaced without state change;
at 6 there remains in the cylinder Fe = F = 2 3.67 cu ft of products at Ta, pa.
The mass of this clearance gas, expressed as a fraction of the total charge,
is given by / = Fa/Fa* = 0.0502. This is close enough to the assumed value
of 0.05, and T* is close enough to the assumed value of 2600°R, so that no
recalculation is necessary' in this case.
The property values obtained above may be substituted in Eq. (1) tc
obtain the work output of the plant process. The values of E and F ob-
tained from the charts apply to a system consisting of one pound of air plus
the specified quantity of fuel. Thus substitution in (1) will give the work
per pound of substance which came in as air (including that part of the
residual gas which came in as air). On this basis

* The E
of combustion is not based on a unit mass of fuel, but on a unit
mass of the gas substance that entered the engine as air. This is why its
all

value changes with the mixture composition, while the internal energy of reaction
as normally defined is independent of xxuxture composition; For mixtures con-
taining less than theoretical air the £
of combustion is also affected by the influ-
ence of the residua] gas fraction upon the equilibrium after combustiom
23-2] INTERNAL COMBUSTION 439

- f(Et - Er) “ 0.05(510 - 465) * 2.25 Btu

(Note that £e and the maas present is / X total mass.)


Wv-i = p<(Fi - fVv) = +4t 12[22 - 0.05(82)] = 39.7 Btu

(at 6' only / times the total mass is present; at 1 the total mass is present).

= £, - Ej = 37 - 155 = -118 Btu


Wt^ = B, - £4 = 1371 - 798 = 573 Btu

= p.(F, - 7.)

but 7, = 7j. and 7. = 7,

so TFi., - p.(7, - 7i) = HI 15(3.67 - 22) = -51.0 Btu

Then 17 « 446 Btu per pound of air substance present. The piston dis-
placement per pound of air substance is

7i - 7j = 22 - 3.67 = 18.33 cu ft,

but the piston displacement of the given engine is 0.120 cu ft. Therefore
the work of the engine per machine cycle is

446(0.120/18.33) - 2.92 Btu


The mean effective pressure is

Pm
JL (2.92) (778)
= 131 psi
Ar (0.12)(144)

The efficiency of an internal combustion engine is commonly based on


the lower heating value of the fuel consumed. In the present case the fuel
is octane, LHV = 19,256 Btu/lb, and the fuel-air ratio is 0.0665 Ib/lb. The
air consumed is 1 — / multiplied by the air substance present in the process;
therefore
446
~
(1 - 0.05)(0.0666)(19,266)
For comparison, the efficiency of the air-standard cycle is slightly greater
than 0.5 at compression ratio of 6. Comparisons of fuel-air cycles with
air-standard cycles for various conditions may be found in standard texts
on internal combustion engines.

23-2 Compression-ignition engine process. The compression-


ignition engine process differs from the spark-ignition engine process
in that only air and residual products are compressed, the fuel being
injected after compression. The air supplied is normally well above
the theoretical quantity because, in the absence of time for complete
mixing, it is impossible to bum the fuel satisfactorily without con-
440 INTERNAL COMBUSTION [23-2

siderable excess air. Because of the excess air the properties of the

combustion products can be obtained with satisfactory accuracy from


gas tables of the Keenan and Kaye type.*
‘ The ideal plant process may correspond to the Diesel cycle or to
the mixed cycle as appropriate. Taking the Diesel type, without
throttling, as a simple example, the idealized process would be as
shown in Fig. 23-2. Beginning with residual gas at 6, the air charge

2 3

Fig. 23-2 I(loa1iz(Hl indic'utor diagram; comjjroRsion ignition engine without


throttling.

is taken in by a frictionless adiabatic constant-pressure process, G-1,


and the mixture is compressed in a reversible adiabatic process, 1-2.
From 2 to 3 fuel is injected and burns at constant pressure. The
lemaining processes from 3 to 6 are as described for the spark-ignition
engine.
Given the mass of fuel m/, the mass of air ma, and the mass of
residual gas rrir, the work can be obtained as follows. Since there is

no throttling We-i will cancel otherwise they would be com-


puted as before.
iri_2 = ?ni{ui - t/ 2 )

* In engines a pre-mixed charge of gas and air is


.sonH‘ compr(*s.sion-ignition
<*oinpres8ed, but du(* to tlu* high ignition temperature of gas fuels ignition does
not occur until a small ‘‘pilot" charge of oil is injected. In such a case the
excess air requirement is not so great, and the process is more like the spark-
ignition process than the compre^asion-ignition process. A rational classification
of engines, from the procc‘K8 vi<*wpoint, would be on the basis of pre-mixed charge
or injected fuel, rather than on the b.'isia of spark ignition or compression ignition.
23-2] INTERNAL COMBUSTION 441

where mi = nta + me

= P2{Vz - V 2)

Wa-4 = miuz - Ui)

where m = ma +
.3 me + m/

The work of pumping the liquid fuel is assumed negligible.

Example 2 . An idealized Diesel engine process, Fig. 23-2, operates with


a compression ratio of 15, receiving air at 15 psia, 550®R, and burning oil

represented by the formula Ciellao, at a rate corresponding to 50 percent


theoretical fuel. Find the efficiency of the plant process if the lower heating
value of the oil, at constant pressure, is 18,500 Btu/lb when supplied as
liquid.

Solution: The properties must first be fixed as in the preceding example.


Assume that the ratio of residual gas to air / is 0.03, and that its temperature
Te is 1500°R. For process 0-1,

hi = fhi + (1 f)hi

where / = mc/Ma
From the air tables K= 131.46 Btu/lb at 550®R. If the gas tables are
available, can be obtained from the table for 200 percent theoretical air;

in the absence of the tables an approximate value can be obtained from the
plot, Fig. 23-3. At 1500®R, for products of burning Ciellao at 50 percent of
theoretical fuel, h/T = 0.254; then = 381 Btu/lb. Using these values,
hi = 0.03(381) + 0.97(131.46) = 138.6 Btu/lb

Now, since the final mixture closely approximates air, use the air table. At
h = 138.6 Btu/lb, Ti = 580°R, ui = 98.90 Btu/lb, and vu = 120.70 cu ft/lb.
Using the air tables for process 1
-2
,

Vj^ 1^2 1

Vlr Vi 15

120.70
ihr = = 8.04
15

Then = 1615®R, = 289.05 Btu/lb

and P* = S - ? = iSr ( 1 ^)
= ^’26 psia
ll Vi OoU

In the process 2-3 the fuel is and burns at constant pressure. The
injected
theoretical fuel in this case can be found by the methods of Chap. 22 to be
0.069 Ib/lb of air. Then the fuel injected will he
INTERNAL COMBUSTION

IPKaineaiiMni
mmmmmKicmmKfi mmmmmm
saiBesaiHnppsa
RKiiBRSi'aiBnatiffiiu
it
lamg

mmmmmumuummmmmyjni^%%mr
JIIK^9BS^aaPSii9HHBBaniBI
wmmKnavmmKismsitvmKmmmma'iismuummmmmn
BBBBBaaiSSiSaiSiSaPRaBBRBBBBaBBBBBBaBR
mmmmmmKMKXMmKicwmmmmmmmmmmmmmmmmmma
BBBBB^PBQW&iBBPffaBBMr

l aaaaa BaBBaPBBBBBBBBaBaaaaBBBaaBBBBR
BBBBPBBBBaBBBBRBBBRl
liniBniaiBpBaBaMBBBBBBaBaBHBBBBaBB I

TEMPERATURE, •F ABS

Fig. 23-3. h/T vs. T for products of complete combustion of certain hydro-
carbons. (Based on data from the Gas Tables of Keenan and Kaye.)

iBBaBBiBBBBBBBBBBBBBBBBBBBBBBBBBBBBlI
BiaaBsaBinBaar
iBBBBBSBSaSRBBBBBaai
IBIl
IBlI

gPBBfiSSBBBBBBUESSi
nBfiSS9 BBBBlSS&Z«
HBBSSa^^tlBi
IBBBBBBBRI
2500 3000 3500 4000
TEMPERATURE, •F ABS

Fig. 23-4. Specific heat ratio vs. T for products of complete combustion of cer-
tain hydrocarbons. (Based on data from the Gas Tables of Keenan and Kaye.)
2S-2] INTERNAL COMBUSTION 443

(0.069)(0.5) = 0.0345 lb fuel/lb air

since there is 50 percent theoretical fuel.


For the combustion process, by the method of Chap. 22,

Q = 0 = tnziht — hta) *|- wi/ ^Hb "h — h/) + tntihtB — Ax)


^en
-m/jAHa 4- h/a - hi) - mtjhtB - Ag)
^„ fn%
^
Assume Tb = 520®R. The properties of products may be taken from the
Gas Tables for 200 percent theoretical air; if tables are not available, Fig.
23-3 gives
hiB = 0.242(520) = 126 Btu/lb

On a basis of one pound of air supplied,

m,^HB = (0.0346)(- 18,500) = -638 Btu/lb of air

Assuming that the fuel is supplied at the base temperature Tb,

tn/ih/B ^ h/) — 0

If the fuel were not supplied at base temperature, the enthalpy change could
be computed with sufficient accuracy for liquid petroleum fuels by taking
the specific heat as 0.5 Btu/lb ®F. Per pound of air, m =
2 1.03 lb; from the
air tables h 2B = 124.27 Btu/lb and *= 399.76 Btu/lb. Then
— W =* —284 Btu/lb of air

Finally m, = 1.00 + 0.03 + 0.0345 = 1.065 Ib/lb of air

Then A, = + 126 = 992 Btu/lb


l.Uoo

If the gas tables are available the expansion process, 3-4 can be com-

puted by their aid; otherwise use the plots, Figs. 23-3 and 23-4 with the
perfect gas rules. The temperature corresponding to h can be found by a
simple cut-and-try operation in Fig. 23-3 to be Ta = 351 0®R. Since the gas
constant for the products mixture is 53.4 ft Ib/lb ®R.

(53.4) (3520)
2.08 cu ft/lb
(626) (144)

Then per pound of air consumed

- 2.08(1.065) = 2.22 cu ft/lb of air

Taking the gas at 1 as having the gas constant of air, 53.35 ft-lb/lb

(53.35) (580)
14.3 cu ft/lb.
(15)(144)
444 INTERNAL COMBUSTION [23-2

Then since there are 1.03 lb of gas per lb of air at 1,

Vi = (14.3) (1.03) = 14.7 cu ft/lb of air

Then for the isentropic process 3-4, since V\ and ps = Jh,

” =
= = 626 54.4 psia
P*
{yj ( fHy
where the value of k is estimated from Fig. 23-4 for the approximate tem-
perature range involved, which is to be checked after T 4 is computed.

=
7’4
=
pi V
p r, =
3 (026) ( 2 22 ) .
3520 2030‘’R

The assumed value of k checks as a reasonable value for the temperature


range from 3 to 4. From Fig. 23-3, hi - 531 Btu/lb at 2030°R.
l^etting the residual gas expand isentropically to 15 psia, the volume
corresponding to the total mass at the resulting state is
/54
y
where k is taken from Fig. 23-4 at a reasonable average temj)erature.

Tj = 2 .‘
T, = /^/]|39.2) 2030 ^ ^ 49501^
pi Vi (54.4) (14.7)

The mass of residual gas is given by

vie = (wc + ma + Vlf)



Fa

then — = 1.065^^^ = 0.0267


Via 39.2

The check on mjma and on Te = Tb is satisfa(;tory, since vie/ via has only a
™all effect on the properties.
With the properties available, the work for the process may be computed.

TF1-.2 = miiui - W2) = 1.03(98.90 - 289.05) = -196 Btu/lb of air

= pj(y, - VO = 626^^2.22 “ = 1^4 Btu/lb of air

IV,-, = m,(u, -Ui) = m,[(h, - fu) - a/77H){RT, - RT*)]

= 1.065[(992 - 531) - (53.4/77S)(3520 - 2030)|

= 382 Btu/lb of air

W = - 196 + 144 + 382 = 330 Btu/lb of air.

The efficiency is then


23-3] INTERNAL COMBUSTION 445

^ (0.0345)(18,500)

For comparison, the same problem worked by use of the gas tables for
the products gives an efficiency of 0.51. The corresponding air-standard
cycle, worked on the variable specific heat basis using the air tables, gives
an efficiency of 0.527. For engines using less than 50 percent theoretical
fuel the variable specific heat air cycle gives a good approximation to the
fuel-air process analysis.

23-3 Supercharged engines. When an engine is supplied with


air or fuel-air mixture at a pressure higher than the surrounding
atmospheric pressure it is said to be supercharged. The basic reason
for supercharging is to increase the mass of gas in the engine process
by increasing the density; this gives added power output. Any of
the usual types of compressors may be used for supercharging, but
centrifugal machines and rotary positive displacement machines seem
to be favored. They may be driven by mechanical connection to
the engine shaft, or by an exhaust gas turbine, in which case the
engine exhaust must be at a pressure which will give sufficient pressure
drop through the turbine to obtain the necessary power.

PISTON TRAVEL

Fig. 23-5. Idealizod indicator diagram; supcrchargetl engine.

An idealized indicator diagram for a supercharged engine is shown


in Fig. 23-5. The residual gas left in the cylinder at -6 is compressed
from p, to Pi by the incoming gas charge during the process 6-7.
The resulting mixture is further mixed with additional fresh charge
which enters at constant pressure p,- in process 7-1. The remainder
of the plant process is the same as for an unsupercharged engine.
446 INTERNAL COMBUSTION [23-4

Assuming the processes 6-7 and 7-1 are adiabatic, the energy
balance can be written in terms of m, (the residual gas) and mi (the
total gas mass to be compressed).

nicUt 4- (mi — mc)hi — miU\ -\-W = miUi + p<(Vi — Ft)

If the compression ratio is r,

I'. - r,. r,(i -i).m,„.(i-i)

Then —
mi
m« + ( I
- — )
mi/
h, = «i -|- piVi
^
\ \ Tf

It is necessary to assume values for and mtlm\ then by simple


;

cut-and-try in tables or charts, or by simultaneous solution of the


equations above with the perfect gas equation of state, the values
of Ti and U\ (or which satisfy the conditions may be found.
From here the procedure is as in the preceding sections, including a
check on the assumed temperature and mass of residual gas, and
recalculation if necessary.
In the supercharged engine process the work calculated from the
diagram of Fig. 23-5 is all positive work in both the upper and lower
loops of the diagram; however, it is not the net work of the plant
since some work was used in the supercharger to compress the fresh
charge. For an idealized case this work could be computed as shown
in Chap. 20. The work lost in power transmission, compressor losses,
and duct friction makes the supercharger a much less efficient pump-
ing device than the engine itself, so in general, supercharging does not
improve the overall plant efficiency. It does, however, increase plant
capacity by supplying more dense air; this is practically indispensable
for modem aircraft plants and is often desirable for other applications
such as locomotives. When, as in the case of the two-stroke engine,
a separate air pump is otherwise necessary, it is only reasonable to
take advantage of its presence and do some supercharging. This is
standard practice with two-stroke Diesel engines.
23^ The simple gas turbine plant. In Chap. 16, cyclic gas-
turbine plants were discussed. Their general characteristics apply
to non-cyclic internal combustion plants such as shown in Fig. 23-(),

but the analysis of the combustion-plant process ditfers in detail from


23-4] INTERNAL COMBUSTION 447

(b)

Fig. 23-6. Internal combustion twhine plant, (a) Flow diskgram;


(b) state plot.

that of the cycle because of the introduction of the combustion process


in place of a heat exchanger. The fact that the plant process is
*‘open/’ that it takes in air from atmosphere and discharges to atmos-
phere, would not in itself affect the analysis.
In the ample plant of Fig. 23-6 air is taken into the compressor C
from atmosphere, state 1, at flow rate Wa Ib/sec. After compression
it goes to the burner B where it bums with fuel supplied at flow rate
tO/ The products stream then expands through the turbine T
Ib/sec.
which exhausts to atmosphere. The burner outlet temperatme Tt is
controlled to suit the turbine requirements by regulating the fuel-air
ratio. As brought out in Chap. 22, adiabatic combustion of a chem-
ically correct mixture of an ordinary fuel will give a temperature of
the order of SSOO^F, so only a fraction of theoretical fuel is needed

to reach permissible turbine inlet temperatures.


448 INTERNAL COMBUSTION [23-4

The energy analysis for the simple plant is as follows, based upon
the steady flow energy equation.
Compressor. Assuming an adiabatic machine of efliciency tic,

Wc = W,{h2 - Ai) = ^
Ve
(^2, - hi)

Burner. For an adiabatic combustion process

0” 0 = {Wa + Wf){hi — hzB) 4" U^aQhB h^ + W/QlfB ” hf) + Wfi^ihs)


_ fea/? ~~ hz ^ IfhB

Wa hz — hzB 4" h/B “ /l-/ 4“ i^hs)

where the enthalpy differences are defined as follows:

{hi — hzB) = enthalpy change of products between Tb and Tz


{thB — h) = enthalpy change of air between and Tb
{h/B ~ hf) = enthalpy change of fuel between its inlet state /
and its state at Tb] this may sometimes include
a latent heat.
(Ahn) = enthalpy of combustion of fuel at Tb, Btu/lb

Turbine. Assuming an adiabatic turbine of efficiency rjr,

Wr = {Wa + Wf){hz - hi) = {Wa + Wf)rjT{hz - hu)

The same analysis applies if the pressure drops in a real piping


system and a real burner are taken into account as indicated in
Fig. 23-7.
The properties of working fluids for gas turbine plants may be
obtained with good accuracy from the gas tables. For products mix-
tures the products tables may be suitable in some cases, but often
the fuel-air ratio is so low that the air tables are suitable, and in any
practical case the air tables will give a good approximation.

Example 3. Find the efficiency, air rate (Ib/hphr), and back work
ratio Wc/Wj^e% for a gas turbine plant of the following description: flow dia-
gram Fig. 23-C; inlet air temperature 60®F; inlet pressure, pi = p4 = 15
psia; pressure ratio pa/pi = 6; machine efficiencies of compressor and turbine
both 80 percent; fuel, CieHao; lower heating value 18,500 Btu/lb; turbine
inlet temperature 1450®F; heat losses and pressure losses negligible.

Solution: Use air tables; for the compressor at 60®F = 520®R,

Ai = 124.27; = 1.2147
23-4 ] INTERNAL COMBUSTION 449

Hf?. 23- 7 . SUU<' i>lot Hhowiiig prc^ssuri' losm-h.

Pr2
= ^Pri = fi.073
Pi
At this value of p,, hu — 197.5

hi = hi + hitnJh = 215.8 lltu/lb

For the combustion process,

{Mh) = - 18,500 Btu/lb

Assuming that the fuel is supplied at Tbj and that is 520®R,

hfB — A/ = 0.

For the air supply, since Th = 520®R,


h2B-h2= 124.3 - 215.8 = -91.5
Assuming the products properties are the same as those of air, at

Tz = 1910'^R, hs = 479.85, pr, = 144.53

hi - hiB = 479.85 - 124.3 = 355.5

^ = ^ 18,500 = ^> 0146 lb fuel/ll) air


Wa
o-r
3o5.5 + 0 — !i'rn»
This is about 470 percent theoretical air, so the products differ little from air.

For the turbine

Pr<
= ^PrS = 28.91

at this value of p,, fk, = 30G.9

*
ht = hi — iiT{ht — fk,) = 341.5 Btu/lb

TTr = (1 -t- 0.0146) (479.86 - 341.5) = 140.4 Btu/lb air


450 INTERNAL COMBUSTION [2»^

IT. - (1)(216.8 - 124.3) = 91.6 Btu/lb dr


TTnit 48.9 Btu/lb air

48.9 « .g*
~ ^
{wf/Wc){MB) (0.0146) (18,500)

air rate 52.1 lb air/hphr


4o.y
Back work ratio

= 91J 1 87
Tr„.t 48.9

23-5 More complex gas turbine plants—jet propulsion. In-


ternal combustion plants are built with intercooling and reheat as
described in Sec. 16-8 for cyclic plants; the analysis of such plants
follows the same pattern as that of the simple plant. In the case of
the reheat plant the total fuel burned may approach much closer to
the theoretical fuel than in the simple plant, so the air tables may not
be as suitable for the properties of the products of the reheat com-
bustion as for the properties in the simple process.

PROPELLER
(IF USED)

Fig. 23-8. Jet-propulsion plant flow diagram.

Another ts^pe of plant, the jet propulsion plant for aircraft, is


shown in the flow diagram, Fig. 23-8. Since the jet plant is ordi-
narily moving with respect to the atmosphere it is necessary to use a
from zero absolute velocity to approximately
diffuser to bring the air
zero velocity relative to the plant. The air taken in through the
diffuser passes through a compressor, combustion chamber, and tur-
bine, as in the simple gas-turbine plant, and then expands through
23-6] INTERNAL COMBUSTION 451

an exhaust nozzle or tail pipe to form the driving jet for the aircraft.
In the jet plant the turbine exhaiist pressure is higher than atmos-
phere, in order to provide the pressure required for the exhaust
nozzle. There is usually no shaft work taken out, the compressor
and auxiliaries absorbing all the turbine work. The propeller shown
in Fig. 23-8 is absent in the jet engine, but is used with the so-called
turbo-prop engine.
In analyzing the jet engine process a suitable set of coordinates for
velocity measurement must be used. The most convenient method
for unaccelerated operation is to measure velocities relative to the
plant. Since only steady flow conditions will be considered here the
velocities will be taken relative to the plant. The example below is

based upon the analysis of diffusers and nozzles in Chap. 17, and the
analyffls of the simple gas-turbine plant in this cliapter.

Example 4. A jet engine is operating at 400 miles per hour in air at

10 psia and 480°R. In the diffuser the air is brought to rest relative to the
plant; then compression through a pressure ratio of 6 to 1 is followed by
burning of octane to produce a turbine inlet temperature of ITOO'F. After
expansion through the turbine the gas expands through a nozzle to atmos-
phere at 10 psia. The diffuser efficiency (ratio of isentropic enthalpy rise
to actual enthalpy rise for the same pressure rise) is 90 percent, and the
nozzle efficiency is 90 piercent. The efficiencies of compressor and turbine
are each 80 percent and the pressure drop in the burner is 2 psi; combustion

is complete. Find the fuel-air ratio and the specific impulse of the engine,
Ibf/lbm of air taken in per second. If the effective cross-section of the

stream of air at inlet conditions is one square foot what is the total thrust?

Fig. 23-9. Jet-propulsion plant state plot.


462 INTERNAL COMBUSTION [23-5

Solution: Assume for convenience that the air tables are satisfactory
for both air and products of combustion. Take velocities relative to the
engine. Refer to Fig. 23-8 and the state diagram, Fig. 23-9.
DiffiLser. By the steady flow equation and the definition of diffuser
efficiency
~~ h\)2go
--

Since F2 = 0, _
Aj. = + A.
2go

Vi = 400 mph = 587 fps

(0.90)(587)«
hu + 114.69 = 120.89
778(64.4)

At this value of h, Vtu = 1.104

At state 1, = 0.9182

p, « Pi = 12.05 psia
Pn

At this value of hj Ti = SOO^R, prt = 1.127.

Compressor,

PrU = Pr2^ = 6.76


P2

At this value of pr, A,. = 203.2

- Aa ,
= 203.2 - 121.57
h ^8,
+ A2
.

0.80
+ 121.J

At Aj, Ta = 930°R
ps = 6p2 = 72.3 psia
Burner,
A4fl ~ A + Aa — hzB
4

Wa Ai — A4B + h/B — A/ + (AAb)


For liquid octane (AAb) at Tb = 530®R is —19,100 Btu/lb (products gase-
ous). At
Ta = 1700°F = 2160®R, A4 = 549.35, pri = 238.0

Assuming products have the same properties as air and Tb - 530°F,

= hiB 126.66

hi-hAB = 422.7
23-51 INTERNAL COMBUSTION 453

*3 - hw = 223.6 - 126.66 = 96.9

Assuming the liquid fuel is supplied at atmospheric temperature, 480°R,


and that its specific heat is 0.5 Btu/lb °R,
h/B — h/ = 0.5(530 — 480) = 25 Btu/lb
Then
^ -422.7 + 96.9 0.0174 lb fuel/lb air
422.7 + 25 - 19,100
This is about 380 percent theoretical air.

Pi = p3 — 2 = 70.3 psia

Turbine, The turbine problem in this case is to find the necessary pres-
sure drop to supply the work needed by the compressor. From the com-
pressor solution this work is hz — = 101.1 Btu/lb of air. On the basis
of the gas flow through the turbine this becomes

101.1 —+
Wa Wf
= 99.3 Btu/lb of gas

Then hi — h == 99.3 = rjT{hi — hs)

or /i6, = 549.4 - (99.3/0.80) = 425.3

At this A, prbs = 93.1

Ph = Piiprha/pri) = 70.3(93.1/238) = 27.6 psia

hs = 549.4 - 99.3 = 450

Ti = ISOrR
Pr, = 114.28
Nozzle.

Then at this Pn - 340


ht = hi — 7}n{hi — ho^) = 3i)l

At this h, T, = I43ini

Vi - V2g,{hi - hi) = 2220 fps

Specific Impulse. The s]3ecific impulse is given by the rate of change of


momentum of the fluid passing through the process, f)er pound of fluid flow

per second. Using the momentum equation from Sec. 18-2,

Specific impulse = ibf/lOm air i)er sec

(The mass leaving at 6 includes the mass of fuel wj which stai*ted at zero
relative velocity.) The difference in pressure foices between inlet and outlet
454 INTERNAL COMBUSTION
is not considered here, but is assumed to be taken into account as part of
the drag of the aircraft.

(1-0174) (2220) 587


Specific impulse = _
= g 2 5 Ib/lb
52.5 ib/lb per sec of air

Total Thrust The total thrust is obtained by multiplying the specific


impulse by the air flow rate. By continuity,

Pi (10)(144)

mgZ1.33lb/s«

thrust = 33(52.5) = 1730 lb

At 400 mph (587 fps) this is equivalent to 1840 hp. The fuel rate is

33(0.0174) = 0.574 Ib/sec or 2070 Ib/hr. Specific fuel consumption is then

—2070
U
= ,,oiu/uu
21h/hphr

and the efficiency based on lower heating value is

= 0.119 or 11.9 percent.


1.12(19,100)

It may be observed that this is an overall efficiency, comparable to the


product of brake thermal efficiency and propeller efficiency for an ordinary
engine; it is also low because the speed of 400 mph is low for efficient jet

propulsion.

If the plant of Example 4 had been a turbo-prop plant, the turbine


would have had to supply the power required by the propeller as well
as the compressor power. The analysis would have been unchanged
in other respects, except for adding the propeller thrust at the end.

PROBLEMS
23-1. In a gasoline engine operating with a mixture of octane and theoretical
air the indicatordiagram shows at the beginning of expansion (after combustion)
a pressure of 600 psia and a volume of 0.02 cu ft; at the end of expansion (before
the exhaust valve opens) the pressure is 80 psia and the volume is 0.10 cu ft.
The work done on the piston during the expansion process is shown by the
diagram to be 3,200 ft lb. From other measurements it is known that the gas
mass is 0.0066 lb of air-fuel mixture. Taking gas properties from the charts
(Chart £-5), find the heat transferred during the process.
INTERNAL COMBUSTION 455

23-2. Solve P)x>blem 23-1 taking the combustion products to be a perfect


gas of gas constant 53.8 ft Ib/lb and specific heat ratio 1.28.
23^. A gasoline engine gives an indicator diagram similar to Fig. 23-1.
The pressure at point 6 is 16 psia and at point 1 is 14 psia; the work on the
piston in the process fi-1 is 186 ft lb. The temperature of the residual gas at
6 is 2100"F; the mixture of octane vapor with theoretical air in the intake pipe
isat 14.5 psiai lOO^’F. The engine has a piston displacement of 0.1 ou ft, and
a compression ratio of 6.5. If the process 6-1 is adiabatic find the temperature
at 1 and the ratio of residual gas mass to total gas mass. Use the charts
(El-5) for gas properties.
23-4. A gasoline engine bums octane with theoretical air.
Constant-volume
combustion starts at 600"F, 110 Assuming the combustion process adia-
psia.
batic compare the final temperatures and pressures obtained by two methods of
computation: (a) accounting for chemical equilibrium by using Hershey, Eber-
hardt and Hottel’s charts (Chart E-5); (b) assuming complete combustion,
and using either the gas tables or average specific heats from data given in the
appendix. (Note that this is constant-volume combustion, not steady flow.)
23-5, Solve Example 1, Sec. 23-1, with no throttling, that is, with inlet
pressure of 15 psia.
23-6. Solve Example 1, Sec. 23-1, for a compression ratio of 8. Compare
the percentage effect on thermal efficiency, for the change in compression ratio,
with the percentage effect of the same change in the case of the air-standard
Otto cycle.
23-7. Solve Example 2, Sec. 23-2, for 25 percent theoretical fuel, using air
tables for the properties of products of combustion. Compare the efficiency for
this case with the efficiencies for 50 percent theoretical fuel (Example 23-2) and
zero fuel (Otto cycle with same compression ratio). How much work output
would be obtained with zero fuel?
23-8. A compression-ignition engine operates on the mixed cycle, using
50 percent theoretical fuel, assumed to be CieHso. The lower heating values of
the fuel are: at constant pressure, 18,500 Btu/lb, and at constant volume, 18,450
Btu/lb. The air supply is at 15 psia, 550°R, the compression ratio is 15, and
the maximum pressure in the process is 900 Find the efficiency of the
psia.

idealized process on the variable specific heat basis (air tables), and compare
with the result of Example 2, Sec. 23-2.
23-9. Solve Example 2, Sec. 23-2, assuming the engine is supercharged with
air at 30 psia, 600°R. The supercharger is driven from the engine shaft, so the
exhaust pressure remains at 15 psia. Make calculations on the variable specific
heat basis, using air tables. In computing net work and efficiency, charge against
the engine 1.3 times the reversible adiabatic work of compression, for driving the
supercharger. Compare mean effective pressuie with that for Example 2.
23-10. In a gas turbine plant based upon the Brayton cycle the inlet air
pressure and temperature are respectively 15 psia and 80®F. The pressure ratio
is 6.25, and the temperature entering the turbine is 1440®F. The fuel is liquid
octane, CsHu, at 80®F. The compressor and turbine efficiencies are each 80 per-
cent; heat losses and pressure losses are negligible. Find: (a) the air-fuel ratio;
fuel
(b) the back-work ratio, Wt/Wj^\ fc) the air rate. Ib/hphr; (d) the specific
466 INTERNAL COMBUSTION

consumption, Ib/hphr; (e) the thermal efficiency based on lower heating value.
Compare results with Problem 1&-17.

23-11. Solve Problem 23-10 if a regenerator of 75 perpent effectiveness is

added to the plant. Compare with results of Problem 16-18.

23-12. Solve Problem 23-10 if a reheat cycle is used; the turbine expansion
is divided into two steps, (lach of pressure ratio 2.5 and efficiency 80 percent, with
reheat to 1440°F. Compare with results of Problem 16-21.

23-13. Solve Problem 23-12 if a regenerator of 75 percent effectiveness is

added to the plant. Compare with results of Problem 16-22.


23-14. A gas turbine plant operates under the same conditions as in Problem
23-10 except that the air flow is limited to 150 percent theoretical air for the fuel
burned, and liquid water at 80°F is injected into the combustion chamber where
it evaporates. The mixture of gases and water vapor enters the turbine at
1440°F, the amount of water injected being regulated to obtain this temperature.
Find all the results called for in Problem 23-10, and also the ratio of water injected
to fuel burned. Compare results with Problem 23-10.
23-15. A turbo-prop plant operates under conditions the same as those of
Example 4, Sec. 23-5, exc^ept that work corresponding to 25 Btu per pound of
gas is taken from the turbine shaft, and supplied to a propeller. The propeller
has an efficiency (based on power) of 75 percent. Find: (a) the thrust con-
tributed by the propeller; (b) the thrust contributed by the jet; (c) the specific
fuel consumption, Ib/hphr; (d) the thermal efficiency based on lower heating
value. Compare results with Example 4.

REFERENCES
See the references at the end of Chap. 16, p. 292.
Chapter 24

REFRIGERATION

Refrigeration is the cooling of a system below the temperature of its

surroundings. This may be accomplished by non-cyclic processes


such as the melting of ice or the sublimation (vaporization) of solid
carbon dioxide. Of greater interest, however, are methods in which
the cooling substance is not consumed and discarded but used in a
thermodynamic cycle. Such methods, called mechanical refrigeration
processes, will be discussed in this chapter.
24-1 Reversed heat engine cycles. In Sec. 7-8 it was pointed
out that a reversible heat engine cycle might be visualized as operat-
ing in reverse, receiving heat from a low-
temperature region, discharging heat to a
high-temperature region, and receiving a net
inflow of work. Under such conditions the
cycle is called a heat pump cycle, or a refrig-
eration cycle.* Figure 24-1 indicates, sche-
matically, a reversed heat engine A operat-
ing as a refrigerator or heat pump. By the
First Law, if the heat engine works in a
cycle, Qi must equal the sum of Qt and W’.

By the Second Law and the definition of the


Fig. 24-1 . Ucfrigi'iiitor
temperature scale, if A is a reversible heat or liwit pumj).
engine such as a Carnot engine, working in
reverse between the fixed temperatures T2 and Tu the relationship
between the work and heat quantities is

*The conventional distinction between a beat pump and a refngiTator is


entirely arbitrary. If the primary purpose is to discharge heat to a certain high-
teraperature region, the system is called a heat pump. If the purpose is to
absorb heot/rom a certain low-temperature region, the system is called a refrig-
erator.

457
458 REFRIGERATION [24.2

Tx - T2 Ti-Tt
W= Ti
Qi
Ti
Qt (1)

The work a reversible heat pump as given by Eq. (1) is the mini-
of
mum work for any heat pump working in a cycle between the fixed
temperatures T2 and Ti. This may be shown by the method used
in Chaps. 7 and 8, that is, assume the statement tmtrue and show
that a violation of the Second Law follows.
As an index of the perfection of a heat pump or refrigerator a
quantity called the coefficient of performance (CP) is defined as
follows:
For a heat pump,

(CP) = A
rr net
(2)

For a refrigerator,

(CP) - 5^ (3)
Fr net

These coefficients correspond to the efficiency of a heat engine; in


each case the numerator is the measure of the desired effect, or
output, of the apparatus, while the denominator is the input. In the
case of a heat pump the desired effect is the heat transferred from
the system, while in the case of the refrigerator it is tlie heat trans-
ferred to the system. In both cases the input is the work.
The maximum coefficient of performance with fixed limits of tem-
perature Ti and Tj is that of a Carnot cycle.
For a heat pump.

(CP)„„ = (4)

For a refrigerator.

(5)

maximum coefficients may be greater


It will be observed that these
than unity, and that they become greater as the temperature differ-
ence decreases.
24-2 Vapor compression refrigeration cycles. Practical re-
frigeration cycles, like practical power cycles, are designed to repre-
sent with reasonable fidelity the operations carried out in an actual
plant. The basic operations involved in a vapor compression refrig-
24-2] REFRIGERATION 459

Fig. 24-2. Vapor-compression refrigerator flow diagram.

eration plant are illustrated in the flow diagram, Fig. 24-2, and
property diagrams. Fig. 24-3. The operations represented are as
follows for an idealized plant:
Compression. A reversible adiabatic process 1-2, starting with
saturated vapor, and ending with superheated vapor.
Cooling and Condensing. A frictionless constant-pressure process,
2-3, ending with saturated liquid. Heat is transferred out.
Expansion. A throttling process 3-4, for which the enthalpy is

unchanged. (The throttling process is not a constant enthalpy proc-


ess as shown in the diagram, but it has the same flnal effect and is
commonly represented as a constant-enthalpy process.) There is no
heat transfer.

p»c

fig. 24-3. Vspor-compreasion refrigerator cycle.


460 REFRIGERATION [24-2

Evaporation. A constant-pressure process 4-1, which completes


the cycle. This is the process in which the refrigerating effect occurs,
as heat is transferred to the evaporating fluid.
If a real plant is to be analyzed, the irreversibilities of the various
processes and the undesired transfers of heat would be accounted for
in the same way as in the case of power cycles; the manner of doing
this has been amply demonstrated in previous chapters.
The compressor of a refrigerating plant is often a reciprocating
machine as sketched, but it may be a rotary positive displacement
machine or a turbine type machine if such is suitable for the condi-
tions involved. The condenser
is usually a tubular heat exchanger

which may from the system to the atmosphere or to


transfer heat
cooling water. The expansion valve may be a manually operated
valve as sketched, but is usually some form of automatic regulating
valve, such as a pressure regulator, a temperature regulator, or a
liquid-level regulator, according to the particular control scheme in
use. The evaporator is usually a tubular heat exchanger which is
in contact with the substance being cooled by the refrigerator. For
example in a water cooler the tubes might be submerged in water,
while in a cold storage room thp evaporator tubes could be suspended
in the air of the room, receiving heat directly from the air.
The choice of a working fluid, or refrigerant, for a given cycle
depends upon considerations similar to those which determine the
desirability of a fluid for a power cycle, as explained in Sec. 15-4.
For given condenser and evaporator temperatures the fluid should
not require an extremely large pressure range; it is desirable that
the pressure should always be above one atmosphere to avoid leakage
of air into the apparatus. The latent heat should be large to mini-
mize the quantity of fluid circulated. The fluid should, if possible,
be low in cost, non-toxic, stable, and inert with respect to materials
of construction. It must not freeze at the lowest temperature of the
cycle. Several other characteristics may be of importance to the
refrigeration engineer, but those mentioned should suffice to show
why no single fluid is suitable for all installations. A few of the
refrigerants commonly used are ammonia, NHj; Freon-12, CCljFj;
methyl chloride, CH3CI; sulfur dioxide, SO2 and water, H20. Brief
;

tables of properties of some of these substances will be found in the


Appendix; charts for some of them are also provided in the back
cover envelope.
24-2] REFRIGERATION 461

Example 1. A refrigeration plant is to operate with an evaporator satu-


ration temperature of 0°F while removing 10,000 Btu/hr from a cold room.
The condenser is to be cooled by water so that the saturation temperature
can be kept at 76®F. The
ammonia. refrigerant is
(a) Assuming the plant works on a cycle like that of Fig. 24-3, find its
coefficient of performance and compare this with the coefficient of perform-
ance for a Carnot refrigerating machine to do the same job.
(b) If the volumetric efficiency of the compressor is 70 percent how much
piston displacement per minute will be needed?

Solution: (a) Referring to Fig. 24-3, on a basis of one pound of fluid


the heat transferred in the evaporator is

Qe — hi — hi Btu/lh

The net work to the cycle is the compressor work, given by

ITc = /i 2 — hi Btu/lb

The enthalpy values are obtained from the ammonia tables:

hi = hg at 0°F = 611.8 Btu/lb

/i 2 = at 8i and p 2

8i = 5 , at O^’F == 1.3352 Btu/lb °R

P2 = saturation pressure at 76®F = 143.0 psia

hi = 704.4 Btu/lb

hz = hf at 76°F = 127.4 Btu/lb

hi = h = 127.4 Btu/11)

Qt’ _ hi hi 484.4 fjQ

For a Carnot refrigerator


4«)
6.05
76

(b) The piston displacement required is given by

(PD) = ^
Vv

where w is the refrigerant flow rate, Ib/min. The flow rate is found from
the time rate of heat flow to the refrigerant, and the value of Qe]

10,000/60
w = 0.344 Ib/min
484.4

vi = Vg at 0°F = 9.116 cu ft/lb


462 REFRIGERATION [24-3

(0.344)(9.116)
(PD) 4.48 cu ft/min
0.70

If the cycle involved a real (irreversible) compressor the calculation would


be handled as in Chap. 21; also, as explained in that chapter, for cycles

requiring a wide pressure range multistage compression might be desirable.


Another refinement sometimes used for refrigeration through a wide range
of temperature is a cascade arrangement in which the condenser of one plant
is the evaporator of another, so the heat-pumping process takes place in two
stages. This is the reverse of the binary vapor power cycle process.

24-3 Capacity of a vapor


compression plant. The ca-
pacity of a refrigerating plant, or
the rate at which it can absorb
heat from some region, is ex-
pressed in tons. One refrigera-
tion ton is defined as the transfer
of heat to the plant at the rate
of 200 Btu per minute. This
is approximately the rate of
cooling obtained by melting ice

Pig. 24-4.
at the rate of one ton (2000 lb)
per day.
Consider the cycle of Fig. 24-4; the heat received from the cold
region is

Qs — hi — hi Btu/lb of refrigerant circulated

Converting tliis into terms of tonnage

(IteO (6)

where (Ref) is refrigeration rate in tons, and w is refrigerant flow


ra^e in Ib/min. Then for a plant using a reciprocating compressor

= y^(hi - hi)
_ (hi - hi)
(7 )
200 200

where N
is compressor speed in rpm, (PD) is compressor piston dis-

placement in cu ft, i;« is compressor volumetric efficiency, Vi is refrig-


erant volume flow rate entering the compressor in cu ft/min, Vids
refrigerant specific volume entering the compressor in cy ft/lb.
24^] REFRIGERATION 463

From Eq. (7) it appears that the capacity of a plant naing a


reciprocating compressor is dependent upon the following factora:

cycle temperatmes; these afifect hi, ht, Vi, Ht.

refrigerant used; this affects hi, hi, Vi, n,.

compressor size; this affects (PD).


compressor speed; this affects N, iiv.

compressor clearance; this affects 17 ,.

heat transfer and fluid friction in the compressor; these affect i},.

In the above reasoning no account is taken of the heat transfer


capacities of the evaporator and condenser; these capacities will affect
the saturation temperatures actually attainable in any given case, as
explained in Sec. 24-5.
The capacity a plant depends upon so many factors determined
of
by cycle conditions that it is meaningless to specify a plant rating of
a certain number of tons unless the cycle conditions at which this
rating can be obtained are also specified. Since, for many years, the
most important uses of mechanical refrigeration were in ice manu-
facturing and food storage, a set of standard cycle conditions suited
to these applications was established as follows: evaporator satura-
tion temperature 5“F; condenser saturation temperature 86 °F; com-
pressor inlet state, vapor superheated 9°F at evaporator pressure;
expansion valve inlet state, liquid subcooled 9°F at condenser pressure
(for practical purposes this is equivalent to saturated liquid at 77‘’F).
For each refrigerant these conditions will give certain definite values

ofhi, hi, and vi.

Many modem applications of refrigeration have requirements


quite different from those of ice making or food storage plants. Air
conditioning often requires much higher evaporator temperatures,
while many industrial processes require cooling to much lower tem-
peratures. When a plant is intended to operate at conditions which
differ greatly from those specified in the standard rating cycle, the

plant should be rated for its designed operating conditions, and these
conditions must then be specified in each individual case.
24^ Powereonsumptionof a vapor compression plant. The
power required to operate a vapor compression plant of unit capacity
is generally expressed in one or the other of two forms, as a coefficient

of performance, or as honepower per refrigeration ton.


464 REFRIGERATION [24-6

The coefficient of performance was defined above for a cycle; for


a refrigeration plant the same general definition holds,

Qe
(CP)
w
The work may be measured as compreiSsor indicated work, compressor
brake work, or electrical input to the motor, according to the purpose
for which the coefficient of performance is to be used. Similarly the
heat absorl>ed by the plant may be measured by the enthalpy change
of the refrigerant between expansion valve inlet and compressor inlet,

or it may be measured as the heat transferred in the evaporator alone.


The use of different bases for computing performance coefficient is
analogous to the use of different bases for the efficiency of power
plants.
Horsepower per refrigeration ton is a ratio of a work flow rate to
a heat flow rate; its relation to coefficient of performance may be seen
from the following equations.

hp input {W Btu /lb X w lh/min)/(42.42 Btu/hp min )

refrig, tons (Qe Btu/ll) X w lb/min)/(200 Btu/ton min)

hp/ton = W 200 4.175


8)
(
Qe 42.42 ((;P)

For ordinary food storage purposes actual refrigeration plants require


approximately 1 hp/ton,
24-5 Effect of irreversible heat transfer on plant perform-
ance. Up to this point it has been assumed that the saturation
temperatures in the evaporator and condenser are fixed, for a given
application, by the surrounding conditions. Actually, however, it is

the desired temperature in the region being cooled, and the available
temperature of condenser cooling water (or air) that are fixed. The
evaporator saturation temperature must be lower than the desired
cold-region temperature and the condenser saturation temperature
must be higher than the available cooling water temperature by suffi-
cient amounts to obtain the necessary rates of heat transfer. This
situation is illustrated in Fig. 24-6, in which tew is cooling water tem-
perature, tnf is the cold-region temperature, Me is the temperature
difference in the condenser, and M^ is the temperature difference in
the evaporator.
24- 5 ] REFRIGERATION 465

Fig. 24-5. Cycle with temperature differences in condenser and evaporator.

The temperature differences Ate and Ate are fixed by economic


considerations, not by thermodynamics; for if there were no limits
to the size of the heat transfer apparatus and the quantity of cooling
water used, the temperature differences might be made to approach
zero. Considering only the performance coefficient and the capacity
of the plant it would be desirable to make both temperature differ-

ences as small as possible for this would reduce the tempera:ture range
of the cycle and increase the performance coefficient and capacity.
This would mean reduced expense both for initial cost of the com-
pressor and for power supply. The reduction of temperature differ-

ence, however, can only be obtained at an increased expense for the


heat transfer apparatus and for the cooling water supply; these
expenses are small when At is large, but go toward infinity as At
approaches zero. At some value for each Aty depending upon the
relative magnitudes of the various elements of cost, the total cost
for compressor power, heat exchanger installation, and cooling water
supply will be a minimum; these particular values of At are what
the designer tries to use.
Practice seems to indicate that temperature differences of the
order of 5 to 25°F are economical under various conditions. It is
clear that such differences may have an appreciable effect on per-
formance. Consider a case of a cold region at 32°F, with cooling
water available at 72®F; if a (^arnot refrigerator worked between

these limits would have a performance coefBcient of 12.3. Now if


it

the temperature difference in both the evaporator and the condenser


is taken to be 10°F, since the external conditions remain the same,

the cycle must work between 22®F and 82® and have a performance
466 REFRIGERATION [244

coefficient cS 8.03. Although the effect on a real cycle would not be


identical to the effect on the Carnot cycle, the order of magnitude
would be the same. It is obvious that irreversibility in heat transfer

can be of primary importance in r^iigeration plants. -

24-6 Eaqiansion engines—gas f»oIing. The common vapor


compression refrigeration cycle described in the preceding sections is

not, even in the ideal case, a reversible cycle because the process in
the expansion valve is irreversible. It is possible, however, to sub-
stitute for the expansion valve an expansion engine, which in the ideal

Fig. 24-6. Vapor-compression plant with expansion engine.

case can effect a reversible adiabatic expansion of the fluid. The


resulting flow diagram and state plot are shown in Fig. 24-6.
In this cycle it is clear that the net work input is less than the
work of the compressor by the work of the expansion engine; more-
over, theend state 4 of the expansion in the engine is a state of lower
enthalpy than the end state 4' of the expansion in a throttle valve.
Thus by use of the engine it is possible to increase the refrigerating
effect hi —and at the same time decrease the net work required.
hi
The magnitude of the improvement in capacity and performance
obtained by use of the expansion engine depoxds upon the properties
of the particular refrigerant involved. The enthalpy change associ-

ated with the isentropic expansion of a saturated liquid is small, so


24-6] REFRIGERATION 467

the advantage of the expansion engine with vapor cycles is not great,

even in the ideal In actual cases, after allowing for the irre-
case.
versibility of the real engine process, the gain by use of the expansion
engine is usually negligible and such machines are not used in modem
vapor refrigerating plants. When gas is used as a refrigerant little if

any cooling can be obtained by throttling since, for a perfect gas, the
temperature remains constant in a constant-enthalpy process. The
expansion engine is often used, therefore, in plants involving gaseous
refrigerants and in gas liquefaction plants.
An ideal gas refrigeration cycle using an expansion engine is

plotted in Fig. 24-7. The gas is compressed reversibly and adia-

Fig. 24-7. Gas compression cycle; reversed Brayton cycle.

batically from 1 to 2, cooled at constant pressure to the cooling water


temperature at 3, expanded reversibly and adiabatically to the
initial pressure at 4, and heated at constant pressure to the tempera-
ture of the cold region at 1. This cycle is seen to be a reversed
Brayton cycle (Sec. 16-6).
Gas refrigeration cycles are less efficient than vapor cycles, be-

cause of ths wide overall temperature range relative to the useful


range Uv, — tnt] this is evident in Fig. 24-7. Gas cycles have been
used in the past in place of « vapor cycles for applications where a
harmless gas such as air was considered safer in case of leakage than
an obnoxious and toxic substance like ammonia. The principal ap-
plication was on board ship. With the development of less noxious
vapor refrigerants such as the various Freon substances, this t3T)e of
application has become obsolete. Gas refrigeration is still xised, how-
468 REFRIGERATION [24-6

ever, for two types of application, air cooling, and gas liquefaction.
Air cooling by expansion in a turbine is particularly useful in the
cooling of aircraft cabins. Here the wei^t and bulk of a vapor plant
with its heat exchangers are extremely costly because they deprive
the aircraft of equivalent load-carrying capacity and its correspond-
ing revenue. Plants in which the air supply itself serves as the
refrigerant can save as much as half the weight of vapor compression
plants for the same cooling load. A flow diagram and state diagram
for an air expansion cooling plant are shown in Fig. 24-8 for an

COOLING AIR FROM


ATMOSPHERE

Fig. 24-8. Air expansion cooling plant.

idealized case. Air is supplied at I from the engine-driven super-


charger; this air would be necessary for ventilation in any case, but
it is supplied here at pi, a somewhat higher pressure than the cabin
pressure Pi and at a relatively high temperature. The air is com-
pressed from 1 to 2, in an adiabatic process; the temperature at 2 is

then higher than the atmospheric temperature so the air can be cooled
in a heat exchanger by means of cooling air taken in from the atmos-
phere. After cooling to state 3 the air is expanded in the turbine,
the work from which is used to drive the compressor. The cool air

leaving the turbine goes into the cabin, providing cooling and venti-
lation simultaneously. In order to supply the energy to overcome
friction the pressure drop in the turbine is greater than the pressure
rise in the compressor, that is, p4 is less than pi. This, in effect, is

equivalent to taking a net work input from the supercharger to drive


the cooling plant.
24-7] REFRIGERATION 469

24-7 liquefaction of gases. An important use of gas refriger-


ating processes is An elementary flow
in the liquefaction of gases.
diagram and state diagram for such a process are shown in Fig. 24-9.
In the operation of the liquefaction plant gas is compressed to a
high pressure and cooled to room temperature, process 1-2-3, Fig.
24-9. From the cooler the gas passes in process 3-4 through a heat
exchanger .4 cooled by a stream of gas returning from the liquefaction
process (only a small fraction of the compressed gas is liquefied, and
the remainder becomes the cooling fluid for the heat exchanger). At

COOLER
(WATER COOLED) M COMPRESSOR

.MAKE-UP
GAS IN

INSULATED
HEAT EXCHANGER I ?? I

fi '

EXPANSION
ENGINEv

EXPAt^SION VALVE 4^'


LIQUID
SEPARATOR

Fig. 24-9. Gas liquefaction plant.


470 REFRIGERATION [24-7

4 the gas is divided into two streams. The larger stream passes
through the expansion engine, coming out at state 6', while the
smaller stream passes through the second heat exchanger B, coming
out at state 5. The cooling fluid iii the exchanger is the same stream
of gas that later passes through heat exchanger A, and is for the
most part the cold exhaust gas from the engine. In exchanger B
the smaller stream of gas is partially condensed because the tempera-
ture of the exhaust gas at 6' is lower than the saturation temperature
for the pressure of states 4 and 5. The liquid-vapor mixture at
state 5 enters a valve in which it is throttled to a low pressure and
temperature at state 6. The mixture at this state flows to a separator
and liquid reservoir from which the vapor portion, at state 6„ returns
to the main stream of exhaust gas, and passes into the heat exchanger.
The liquid, at state 6/, may be drawn off from the separator as desired.
The heat exchangers are arranged with counterflow paths so that
the maximum cooling can be obtained from the coldest available
gas. The divided flow scheme is used to avoid expanding into the
wet mixture region in the engine, since condensation of liquid in the
engine would cause operating difficulties.
When a plant such as described above is first started there will
be no cooling in the insulated heat exchanger, but as soon as the
expansion engine has cooled a quantity of gas through a few degrees
this gas returning through the exchanger will cool the stream ap-
proaching the expansion engine through a corresponding temperature
interval. This effect will be cumulative, so an engine which is inher-
ently capable of cooling through only a few degrees may eventually
reduce the temperature to a very low level. A plant of this kind
can liquefy only a small fraction (perhaps 10 percent) of the gas
circulated through the engine because most of the gas must be re-
turned through the heat exchanger to obtain the necessary cooling
effect before the e.xpansion.
The liquid produced represents a steady loss of substance from
the circulating system, so make-up gas is introduced steadily as
indicated.
sometimes possible to use an expansion valve in place of an
It is
expansion engine for gas hquefaction. Although a perfect gas does
not change temperature in a throttling process, real gases may have
either an increase or a decrease of temperature in such a process,
depending upon the initial state. The cooling or warming charac-
24-7] REFRIGERATION 471

teristic is indicated by the Joule-Thomson coefficient (dT/dp)h which


is the rate of change of temperature with pressure change at constant
enthalpy. The Joule-Thomson coefficient for a gas is positive at low
pressures and temperatures (at states not too far from the vapor
dome). This means that a pressure drop results in a temperature
drop, and the expansion can be used for cooling. At states of high
pressure and temperature an inversion occurs and the coefficient be-
comes negative, so the expansion results in warming. The inversion
temperature is so high for most gases as to have no influence on refrig-
eration possibilities: for example, nitrogen has an inversion tempera-
ture above 600“F at 1 atm. For helium and hydrogen, however, the
inversion temperature is in the neighborhood of — 400®F and — 100®F,
respectively, so these gases cannot be cooled by throttling without a
great deal of precooling by other means. Cooling by a throttling
process is indicated in Fig. 24-9 by the process 4'-6.
Comparing the two cooling methods, the engine process can give
greater cooling for a given pressure difference, but it requires opera-
tion of the engine at the lowest temperature of the process. Until
recently, such operation was and exceedingly troublesome
inefficient

because of lubrication difficulties, leakage, and heat transfer, so the


higher pressure of the throttling process has often been accepted in
order to avoid the use of engines. Modern materials and improved
design have made possible the building of efficient, trouble-free en-
gines for this service. In the past the liquefaction of helium has been
achieved by a cascade process in which evaporating liquid air precools
hydrogen so it can be liquefied by the throttling process. The liquid
hydrogen then evaporates to precool the helium sufficiently to obtain
a positive Joule-Thomson coefficient, after which the helium is lique-
fied by throttling. A modern liquid helium plant perfected by Pro-
fessor S. C. Collins uses expansion engines to accomplish the lique-
faction without auxiliary liquid gas cooling.
Actual gas liquefying plants are more complicated than indicated
in Fig. 24-9, particularly when air is to be liquefied. Impurities such
as water and carbon dioxide freeze out in the plant and clog the
passages, making it necessary to use duplicate heat exchangers so
one can be flushed out with warm gas while the other is in use.
Other refinements may include multistage compression and expan-
sion, and distillation systems for separating mixtures.
Solid carbon dioxide (dry ice) is made by a process of the throt-
472 REFRIGERATION [24-8

tling type in which, by operating at the proper conditions, “snow” is

formed, instead of liquid.


24-8 Absorption refrigeration. The function of the compres-
sion process in a vapor-compression refrigeration plant can be per-
formed by other types of processes in which little or no work input
is required. Any process which will receive low-pressure vapor and
discharge high-pressure vapor may be combined with a condenser,
expansion valve, and evaporator to make a refrigeration cycle. Some
examples of such processes, which operate by absorption of vapor,
will be described below.
The ammonia absorption cycle is based upon the fact that the
solubility of ammonia in water is a function of the temperature of

the solution and the pressure of the vapor in contact with the solu-
tion. A simple absorption plant flow diagram is shown in Fig. 24-10.

Fig. 24-10. Ammonia absorption refrigerating plant; simple flow diagram.


24-8 ] REFRIGERATION 473

Heat Qa supplied to the generator causes a vapor mixture of ammonia


and water to be boiled off at high pressure (for instance, 150 psia)
from a liquid solution of ammonia in water. The vapor mixture
phases up through the analyzer (or rectifying column) where it comes
in contact with a stream of strong liquid solution coming from the
absorber. Since the vapor is not rich enough in ammonia to be in
equilibrium with the strong solution, most of the water in the vapor
mixture condenses into the liquid stream and drains back to the gen-
erator. Vapor containing only a small amount of water rises past
the strong solution inlet and meets a stream of liquid ammonia com-
ing from the condenser. This reflux stream is a small part of the
total liquid formed in the condenser, but it is sufficient to absorb
the remaining water from the vapor mixture and carry it down to
the generator. Thus anunonia vapor, free of water, leaves the top
of the column and goes to the condenser. The condenser, expansion
valve, and evaporator operate the same as in a compression plant.
The low-pressure ammonia vapor from the evaporator (at perhaps
20 psia) goes to the absorber where it comes in contact with a stream
of cool, weak solution returning from the generator. The absorber
is cooled by circulating water to absorb the heat of solution, which

is an enthalpy decrease (or a latent heat transferred out) when the


ammonia dissolves in water at constant temperature. The strong,
cold solution formed in the absorber is pumped to the high pressure
of the generator, and returned to the generator by way of a heat
exchanger. The heat exchanger preheats the strong solution going
to the generator at the same time that it precools the weak solution

returning to the absorber. Thus the heat transferred in the exchanger


gives a corresponding reduction in both the heat Qa supplied to the
generator and the heat Qa that must be transferred from the absorber.
In the operation of the absorption plant very little work input is
required for the liquid pump compared to the work input for a vapor
compression cycle. This does not mean, however, that the absorp-
tion cycle can circumvent the Second Law; the operation of the ab-
sorption cycle requires a supply of heat Qa at a high temperature and
a rejection of heat Qa at the temperature of the surroundings. This
heat input at high temperatmre and rejection at the temperature of
the surroundings could have been used for work production in a heat
engine cycle, therefore its use in the absorption cycle is entirely
equivalent to an input of work. Figure 24-11 shows the heat engine
476 REFRIGERATION [24-9

gen, so the partial pressure of the ammonia is low, while the total
pressure is the same as on the "high-pressure” side of the plant. No
expansion valve is needed; the liquid passing through the seal loop
between the condenser and the evaporator has equal total pressures
on both ends so the seal does not blow out, but when the liquid
reaches the evaporator it behaves with respect to evaporation as if
the total pressure upon it were the partial pressure of the ammonia
vapor. Evaporation thus proceeds at the low temperature corre-
sponding to the partial pressure.
The absorber must dissolve ammonia in water as fast as it is

evaporated, in order to maintain the low evaporator partial pressure.


This process is aided by a closed-loop circulating system in which a
gravity circulation of the mixture of hydrogen and ammonia is set
up. The path of flow is from the top of the evaporator down through
a heat exchanger to the bottom of the absorber, up through the ab-
sorber, the heat exchanger, and the evaporator. The hydrogen cir-
culates continuously around the loop, picking up ammonia in the
evaporator, and losing most of it in the absorber where it meets the
weak solution.
The actual Servel refrigerator involves several refinements which
make the process more complex in order to improve the performance.
Thermodynamic analysis of absorption plants depends upon a
knowledge of the properties of the mixtures involved. For ammonia-
water mixtures considerable data are available in the references at
For other mixtures the data can be worked
the end of this chapter.
up by methods of chemical thermodynamics which are beyond the
scope of this book.
The principal advantages of an absorption refrigeration plant
compared to a compression plant are (1) the use of little or no moving
machinery, (2) the ability to use relatively low-temperature waste
heat as an energy source, (3) greater flexibility of operating condi-
tions without great loss of capacity or performance coefficient. The
choice of one or the other type of plant is influenced greatly by the
relative costs of heatand power.
24*9 Heat pumps. Every cyclic refrigeration plant is a heat
pump, but this name has come to be applied specifically to plants in
which the heat flow from the plant is the desired effect. Much effort
has been expended to popularize heat pumps for building heating.
In such installations the evaporator takes heat from the atmosphere.
REFRIGERATION 477

a natural body of water, or the earth, and the condenser discharges


heat at a higher temperature for the heating application. The opera-
tion is analyzed in the same way as is the refrigeration plant; in fact
many installations use the same plant to cool a building in summer
and heat it in winter. Technically, the heat pump is a perfectly
practical device, but up to the present its economic justification has
depended upon special circumstances, for example, a large cooling
system that was to be installed in any event and would be available
for use in winter as a heat pump. In other cases particularly favor-
able power costs might exist. If the relative cost of heating fuel
compared to electric power should increase appreciably, the heat
pump might become economically attractive, particularly in regions
of relatively mild winter climate.
An application to which the heat pump appears well adapted is

as a source of heat for the operation of a distilling plant, that is, the
compression distillation process. In this process water is evaporated
by heat from condensing steam, the vapor is compressed so that its

saturation temperature is raised, and it then becomes the heating


steam which condenses to evaporate more water. When a cheap
supply of waste heat is not available, and water must be distilled,
this process may be of great value. It has been used primarily for
military applications, to supply potable water for submarines and for
island outposts. At least one installation has been made in a power
plant to supply make-up feed water for the boilers.

PROBLEMS
24-1. Plot the coefficient of performance of Carnot refrigerating cycles vs. the
refrigerator temperature, for a hot-region temperature of 90°F; cover the range
of refrigerator temperatures from 70°F to — 100®F.
24-2. Plot the coefficient of performance of Carnot heat pump cycles vs. the
hot-region temperature, for a heat source temperature of 30°F; cover the range
of hot-region temperatures from 60°F to 200°F.
24-3. An ideal refrigerating plant. Fig. 24-2, operates between a cold region
at lO^F and a hot region at 80°F; saturated vapor enters the compressor, and
saturated liquid enters the expansion valve. The plant operates at the rate of
10 tons (2000 Btu/min). The refrigerant is ammonia. Find: (a) the coefficient
of performance; (b) the refrigerant flow rate, Ib/min; (c) the volume flow rate
entering the compressor, cu ft/min; (d) the maximum and minimum pressures
of the cycle.
24-4. If the compressor for the plant of Problem 24-3 has a constant volu-
metric efficiency, what refrigeration capacity (tons) can the compressor serve
Chapter 25

AIR-WATER VAPOR MIXTURES

In Chap. 12 the properties of mixtures of perfect gases were presented


and the properties of mixtures containing water vapor were consid-
ered briefly. The properties of moist air are imp>ortant in such indus-
trial processes as drying and evaporative cooling; also, control of the
moisture content of the atmosphere is essential to the satisfactory
operation of many processes involving hygroscopic materials like
paper or textiles and is a major factor in comfort air conditioning.
For these reasons mixtures of air and water vapor have been studied
in great detail; the name psychrometrics is given to this i^bject.
Before reading this chapter it would be well to review Sec. 12-6.
25-1 Definitions. The field of psychrometrics has a special vo-
cabulary of which some terms will be defined below.
Specific humidity or humidity ratio is the mass of water per unit
mass of dry air in a mixture of air and water vapor.
Rehim humidity is the ratio of the partial pressure of the water
vapor in a mixture to the saturation pressure of pure water at the
same temperature. If the perfect gas rules are assumed to hold,
the relative humidity may be defined alternatively as the ratio of the
mass of water vapor in a given volume of mixture to the mass of
saturated water vapor which would fill the same volume at the same

temperature. For real mixtures the two definitions give different


values, but at normal atmospheric pressures and temperatures the
difference is small.*
Dry air is the mixture of all the normal components of atmospheric
air except water vapor. An approximate analysis by volume is 78
percent nitrogen, 21 percent oxygen, 1 percent argon; the argon is

often taken as nitrogen.

* In air conditioning practice the humidity defined by the masa ratio ia

aometimea called percent caturaticn.


480
25-2] AIR-WATER VAPOR 481

Saturated a mixture of dry air and saturated water vapor, or


is

alternatively, a mixture having relative humidity equal to 1.00.


Dew-point temperatme is the temperature at which the mixture
becomes saturated (or condensation of vapor begins) when a mixture
of air and water vapor is cooled at constant pressure from an unsatu-
rated state.
Dry-bvib temperature is the actual temperature of a gas; the desig-
nation “dry bulb” is used merely to avoid ambiguity.
Wdftyutb temperature is the temperature indicated by a thermom-
eter having its bulb covered by a film of water, when the thermometer
is exposed to a stream of air in turbulent flow.
25-2 Humidity relations. Assuming that the mixtures follow
the perfect gas rules and Dalton’s law, the relations between specific
humidity and relative humidity may be developed. Specific humid-
ity 7 is given by

7 *

= —
Vy,
/ i
(1)
\

where is the mass of water vapor in a given space, ma is the mass


of dry air in the same space, is the specific volume of the dry air
in the mixture, is the specific volume of the water vapor in the

mixture.
Relative humidity is given by

2s ( 2)
Pt

where the subscript w refers to the water vapor in the actual mixture
and the subscript g refers to the properties of saturated water vapor
at the same temperature. Figure 25-1 is a temperature-entropy dia-
gram for the water vapor alone. Combining Eqs. (1) and (2),

y = <t>~~ ( 3)

By the perfect gas rules and Eq. (1),

RgT/Pa _ RaPv _
0.622 2s (4 )
RuT/pa R’wPa

Combining (2) and (4),

^ ( 5)
0.622p,
482 AIR-WATER VAPOR [26-3

Fig. 25-1. States of water vapor in mixture.

By Dalton’s law, p = Po + Pw (6)

where p is the total pressure of the mixture.


In humidity computations the saturation pressure p^ is taken
from the steam tables; the saturation volume may be taken from the
tables or computed by the gas laws.

Example 1. A mixture of air and water vapor at ISf’Y and 14.7 psia
has relative humidity 0.50; find its specific humidity and its dew-point
temperature.

Solution: From (4), 0.622 ^



From (2) p. = = 0.5(0.4298) = 0.2149 pei

From (6) Pa = p — Pw = 14.49 psi

n 914.Q
Then 0.622 ’
0.00925 Ib water/lb dry air
14.49

The dew-point temperature is the saturation temperature corresponding


to p.,; this is found in the tables to be fib^F. The dew point is indicated as
point A in Fig. 25-1.

25-3 Wet-bulb temperature. The usual method of determin-


ing experimentally the humidity of a mixture of air and water vapor
is to obtain dry-bulb and wet-bulb temperatures. Experimental cor-
rdations of these temperatures with humidity are available as for-
mulas, tables, and charts. A typical psychrometric chart is folded in
the back cover envdope of this book.
In the usual wet-bulb thermometer a water film is maintained
25-3] AIR. WATER VAPOR 483

around the bulb by a cotton wick saturated with water; for inter-
mittent use the wick may be dipped in water before taking the
reading, while for continuous use the wick may extend to a resrawoir
of water. The sling psychrometer, a common instrument for inter-
mittent readings, consists of a pair of thermometers moimted on a
holder which can be whirled through the air manually. One of the
thermometers has a wet bulb, the other a dry bulb. A continuous
psychrometer with a fan for drawing air over the thermometer bulbs
is shown in Fig. 25-2 such devices are used in permanent installations.
;

Fig. 25-2. Continuous wet- and dry-bulb psychrometer.

In a psychrometer the dry-bulb thermometer gives the actual air


temperature. At the wet bulb, if the air is not saturated, evaporation
will occur from the liquid film on the bulb. The immediate effect is
to lower the liquid temperature, but as soon as the liquid temperature
falls below the air temperature there will be heat transfer from the

air to the liquid film. When steady state conditions are reached
there will be a balance between energy removed from the liquid film
by vaporization, and energy supplied to the liquid film by heat trans-
fer. If the conditions are such that the heat transfer to the film is

solely by convection from air in turbulent flow the temperature of


the Uquid film is the wet-bulb temperature.
It is clear that the wet-bulb temperature is not a property of the
gas mixture since depends upon heat transfer rates and mass trans-
it

fer ratesbetween the liquid film and the air. These rates depend
upon the geometry of the bulb and the air velocity, neither of which
is a thermodsmamic property of the mixture. In practice the wet-
bulb reading is subject to all the errors of dry-bulb thermometry.
484 AIR-WATER VAPOR [26-4

particularly errorsdue to radiant heat transfer, since radiation dis-


turbs the deared balance between convection and evaporation at the
bulb. In order to make convection large with respect to ordinary
radiation it is desirable to have a gas velocity of about 1000 fpm
past the bulb. If the location is directly exposed to the sun or to
hot industrial plant surroundings, radiation shields should be used
around the bulbs, but care should be taken to have the same condi-
tions of air velocity and shielding around both the dry-bulb and the
wet-bulb thermometers.
Since the humidity is determined as a function of the “wet-bulb
depression” or the temperature difference between the dry and wet
bulbs, which is often only a few degrees, it is customary to use finely-
graduated thermometers in matched pairs. The water for the wet
bulb should be distilled, if possible, and the wick should be clean,
since dissolved solids can change the evaporating temperature appre-
ciably.
The wet-bulb temperature always lies between the dry-bulb and
the dew-point temperatures, since, as the liquid film approaches the
dew point, the rate of evaporation (and the cooling effect) approaches
zero. If the air is initially saturated, the dry-bulb, wet-bulb, and
dew-point temperatures will be identical.
25-4 Adiabatic saturation process. A process which is similar
in some respects to the process at the wet bulb, but in which the air
and the liquid water come to a true equilibrium state, may be accom-
plished as follows. Assume that a mixture of air and water vapor

V ///// ////// SATURATED


AIR
AIR

Si
WATER
Fig. 25-3. Adiabatic saturation process.

passes in steady flow through an adiabatic saturating chamber. Fig.


25-3. This is a long duct in which the air is in contact with a body
of water, the whole apparatus being well insulated. The mixture
entering is not saturated, but if the duct is long enough the mixture

leaving is saturated. It is assumed that the water level is kept con-


26-6] AIR-WATER VAPOR 485

stant by make-up water supplied at the temperature of the saturated


air leaving the chamber. Such a process is called an adiabatic saturor-
tion process and the temperature of the air leaving is the adiabatic
saturation temperature.
The adiabatic saturation process differs from the wet-bulb process
in that the liquidand the mixture of air and vapor are kept in contact
until they come to equilibrium and the final state is not dependent
in any way upon the rates of heat transfer and evaporation. It hap-
pens, however, that for water-air mixtures in the range of ordinary
atmospheric pressures and temperatures the wet-bulb temperature
and the adiabatic saturation temperature are practically identical.*
This does not hold true at other states, or for substances other than
air and water.
In the adiabatic saturation process the enthalpy decrease of the
original mixture is equal to the latent heat of vaporization of the
water added to the mixture. Thus

Ai — /i2 = (72 ~ 7 i)A/g2 (7 )

where Aa is the enthalpy per pound of dry air of a mixture at the


specific humidity of state 1, but at the temperature of state 2. By
substituting in Eq. (7) expressions for the three enthalpy terms as
functions of temperature it is possible to obtain an equation relating
the specific humidities to the dry-bulb temperature and the tempera-
ture of adiabatic saturation. Such equations are used as the basis
for the construction of psychrometric charts, as explained in the refer-
ences at the end of this chapter.
25-5 The psychrometric chart. Chart E-6 (see envelope inside
back cover) is an example of a psychrometric chart suitable for air
conditioning computations. This chart is based on the energy equa-
tion for the adiabatic saturation process at a total pressure of 1 atmos-
phere (29.92 in Hg); corrections for other barometric pressures are
tabulated on the chart.! The principal coordinates of the chart
are specific humidity and dry-bulb temperature. Specific humidity
is plotted in grains per pound of dry air (1.0 grain = 1/7000 lb), but
a scale of poimds per pound of dry air is also given.
The curved line at the left of the plot is the saturation line; pro-

* Carrier and lindaay, “The Temperature of Evaporation of Water into


Air,” Trans. ASME, Vol. 44, p. 326, 1922.
t Refermoea to charts for other pressures are given at the end of the chapter.
486 AIR-WATER VAPOR [25-6

jection of dty-bulb temperatures on this line gives both dew-point


and wet-bulb temperatures at saturation. Lines of constant dew-
point, being lines of constant specific humidity, are horizontal, lines
of constant wet-bulb temperature slope downward to the right be-
cause, for unsaturated mixtures, the wet-bulb temperature is higher
than the dew-point.
The chart gives the enthalpy of the saturated mixture (i.e. both
the air and the water) on the basis of Btu per pound of dry air. For
unsatiuated mixtures the enthalpy at constant wet-bulb temperature
is nearly constant, and rough computations may be made assuming
that it is constant; however, the chart provides curves of enthalpy,
deviation for greater accuracy. The deviation is to be added alge-
braically to the saturation enthalpy corresponding to the wet-bulb
temperatme. Enthalpies in the chart are given above a base of zero
enthalpy at zero ®F for the dry and zero enthalpy at 32‘’F satu-
air,

rated liquid for the water vapor (the same base as the steam tables).
Mixture enthalpies may be computed as follows:

h — Cpat^b ( 8)

where h is in Btu/lb of dry air, Cpa is 0.240 Btu/lb ®F, h„ is the


enthalpy of the vapor.
Since the tables for superheated steam do not extend to low
enough temperatures for ordinary psychrometric work, hg, may be
obtained from one of the following equations:

h|p “ hgdp "I**


*“
^p) ” “1“ 0.45(^Jb ““ ^p) C^®*)

or ht, = hf at O^F + Cpy/ta = 1061.8 + 0.44^6 (9b)

or h„ = hgdt (perfect gas approximation) (9c)

The subscripts dp and db indicate dew point and dry bulb, respec-
tively. The specific heat for water vapor is chosen in each case as
a good average for usual conditions.

Exaupu: 2. Atmospheric air has dry-bulb temperature 72*’F and wet-


bulb 58°F. Find the relative humidity, specific humidity, dew-point tem-
perature, and enthalpy.
SoiiTmoM: Referring to tixe psychrometric chart, at the intersection of
the 72°F dry-bulb line with the SS^F wet-bulb line the relative humidity is

read as 42 percent or 0.42. Following the horizontal to the right read spe-
cific humidity as 0.0071 lb water vapor per lb dry air. Following tire hori-
25-6] AIR. WATER VAPOR 487

zontal to the left to saturation, read the dew-point tempeiatuie as 48°F.


The saturation enthalpy corresponding to SS'T wet bulb is 26.15 Btu/lb of
dry air, the deviation at the given state is —0.08, and the enthalpy is 25.07
Btu/lb of dry air. The enthalpy may also be computed from
h = CfJkb + yh„
K “ 1061.8 -I- 0.44(72) = 1093.5

h = 0.240(72) + 0.0071(1093.5) = 25.07 Btu/lb dry air

25-6 Humidifying processes. Processes with mixtures of air


and water vapor may involve no change of mixture composition, in
which case the process is analyzed by the methods used with pure

1 HEATING COIL WATER SPRAY 2 |

1 r
r r”

Fig. 25-4. Humidifying process; spray cooler with heating coil.

substances. Processes involving an increase of specific humidity,


called humidifying processes, are discussed in this section. Processes
in which the hmnidity decreases, called dehumidifying proc-
specific

esses, and processes involving mixing are discussed in later sections.


A simple humidifying process is often used with a heating process
in comfort air conditioning; Fig. 25-4 is a di^am of such a process.

For this process the steady flow equations may be written as follows;
tVat = Wat = Wa

Wahl -t" W/hl "I" 0= Wahi

where fh and ht are in.Btu/lb of dry air.


488 AIR. WATER VAPOR [26-7

Example 3. In the process shown in Fig. 2&-4 the air is recdved at


1 atm, 40°?, relative humidity 60 percent, and it is dedred tq discharge it

at TOT, relative humidity 50 percent. How much heat and how much water
at 45'’F must be supplied per pound of dry air passing through the apparatus?
SoLimoN: From the chart, at state 1 the enthalpy is 12.9 Btu/lb of dry
air and specific humidity is 0.0032 lb water/lb dry air. At state 2 the
enthalpy is 25.33 Btu/lb of dry air and the specific humidity is 0.0078 lb
water/lb dry air. The enthalpy of the liquid water hi is h/ at 45T or 13
Btu/lb from the steam tables. The water to be supplied is the difference
between the specific humidities of the air:

— 7i = 0.0046 Ib/lb of dry air

Then the energy equation on the basis of one pound of air is

12.9 + 0.0046(13) Q= 25.33

Q= 12.4 Btu/lb of dry air

The probl^ may also be solved by the mixture rules as follows.

7 =
, 0.622 2*^
Pa

From the steam tables at 40®F, pg = 0,1217 psia. Since Po =p— Pt

7i = 0.622 (0.60) = 0.00311 Ib/lb of dry air.

By the same method yt = 0.00788 Ib/Ib of dry air

Then wi = 0.00477 Ib/lb of dry air

The energy equation is

CfJi + 7iA»i + lOjhj + 0 = Cfjti 7thwi

= 1061.8 + 0.44<, . 1079.6 Btu/lb

* 1092.6 Btu/lb

hi = 13 Btu/lb

Q- 0.240(70 - 40) + 0.00788(1092.6) - 0.00311(1079.6) - 0.00477(13)


Q -a 12.39 Btu/lb of dry air

25>7 Mixing processes. In air conditioning processes it is often


desired to introduce a certain amount of fresh air into the process
continuously, while discarding an equal amount of stale air. This
may sometimes be done in such a way that the desired final tempera-
ture and humidity are obtained by properly conditioning the fresh
25-7] AIR-WATER VAPOR 489

air before mixing. Figure 25-5 shows a simple mixing process for
which the steady flow equations may be written as follows;

Wa\ + U)«2 = tta3

= tOwi

tOolfcl -|- WiJit = WaJh

where enthalpies are per pound of dry air.

ExAifPLii 4. It is desired to supply to a room 1000 cfm of air at 1 atm,


68**F, 50 percent relative humidity, by mixing recirculated air from the room
with fresh air as shown in Fig. 25-5. The recirculated air from the room is

at TS^F, 70 percent relative humidity. If it is desired that 50 percent of


the dry air supplied to the room be fresh, bow much air must be supplied
at 2 and what must be its temperature and humidity?

Solution: It is first necessary to determine the properties at states 1


and 3. From the chart, at 75°F, 70 percent relative humidity

7 *
i 0.0131 lb water/lb dry air

hx = 32.36 Btu/lb dry air

At 68'*F, 50 percent relative humidity

7, = 0.0073

h 24.26

The flow rates of air and vapor are now found.

Pat “ = 0.50(0.3300) 0.1606 pd


490 AIR-WATER VAPOR [26-7

Wai = —=7s
73.8 Ib/min

From the given conditions,

Wai = 0.5woz = 36.9 Ib/inin

Wa2 = O.Bwai = 36.9 Ib/min

Then w^i = yiWai = 0.484 Ib/min

By the mass equation

Wv2 - Wta — Wwi = 0.055 Ib/min

Then 72 —
Wa2
0.00149 lb water/Ib dry air

By the energy equation

h2 16.2 Btu/lb of dry air


Wa2

From the chart, at A *= 16.2 and 7 = 0.00149

tdb = 61°F approximately

Mixing problems may also be solved directly on the psychrometric


chart. The specific humidity of a mixture, if no condensation occurs,
must be the average, weighted according to the mass of dry air, of
the specific humidities of the components. Also the temperature
must be very nearly the average, weighted in the same way, of the
temperatures of the components (this is not an exact relation because

DRY-BULB TEMPERATURE
Fig. 25-6. Mixing process on the psychrometric chart
25^] AIR. WATER VAPOR 491

the specific heats of components of different specific humidities are


slightly different). From these facts it follows that the mixture state
will be located on a straight line through the two component state
points and at a distance between these two points proportional to
the mass of dry air in each component. In Fig. 25-6, if A and B
represent the states of two streams of air and water vapco* which are
mixed, the state of the mixture C will lie on line A-B. If the pro-

portions by pounds of dry air are fB and fA, then A-C will be f of
A-B. The reader may easily check the solution of Example 4 by
this method.
25-8 Dehumidifying processes. Dehumidifying, or the removal
of moisture from air, is often required in comfort air conditioning

Fig. 25-7. State paths of water in a dehumidifying process.

and in industrial processes. Moisture may be removed by absorption


in liquids or solids (so-called chemical dehumidifying), or by cooling
below the dew point. Only the latter method will be considered
here. The removal of moisture by cooling is shown on the tem-
perature-entropy diagram. Fig. 25-7, which shows the states of the
vapor only. Figure 25-8 a flow diagram for the process.
is

If a mixture originally containing water vapor at state 1 is to


have its specific humidity reduced to that of state 2 by cooling, the
mixture must be cooled to the dew-point temperature
first corre-
sponding to state The cooling will take place in two steps, first,
2.

cooling of the original mixture to its dew point ta and then cooling
of the saturated mixture, with condensation of water as the tempera-
492 AIR-WATER VAPOR [26-8

ture falls. If the condensed liquid is separated from the mixture,


the saturated air at state d2 may then be heated to the desired final

state, 2. In Fig. 25-9 the path of the process is shown on the psychro-
metric chart.

DRY -BULB TEMPERATURE

Fig. 25-B. Dehumidifying process on the psychrometric chart.

It will he observed that the nd, heat transfer for process 1-2 is
two
of little significance, since to accomplish the desired effect the
individual transfers Qc and Qg are both essential. Therefore the
mergy equation, to be useful, must be written for the two separate
processes of cooling and heating. The twum and enei^ equations
are
2W] AIR-WATER VAPOR 493

+ Wufl = + Wv2 + Wi
Wa ti?fl

wjii + Qc - wjidi + wihf

Wahoi + Qh == Wahl

where the mixture enthalpies are per pound of dry air, and both heat
quantities are taken positive for heat flow to the mixture.

Example 5. Air at 1 atm, 75^F, 70 percent relative humidity is to be


brought to 70°F, 60 percent relative humidity by the process of Fig. 25-8.
To what temperature must the mixture be cooled? How much heat must
be removed by the cooling coil and how much must be supplied by the
heating coil per pound of dry air? What fraction of the heat removed in
the cooling coil is required to cool and condense the water removed?

Solution: Using the psychrometric chart, and referring to Fig. 25-8


and 9,

71 = 0.0131 lb water/lb dry air

72 = 0.0093

hi = 32.36 Btu/lb dry air

hi = 27.03

The mixture must be cooled to the dew-point temperature corresponding to


state 2; this is found on the chart to be 55.3^F. Then from the chart and
the steam tables

hdi = 23.40 Btu/lb of dry air

hf = 23.4 Btu/lb of liquid

From the mass equation,

Hi = -yi
— -yj = 0.0038 lb water/lb dry air
Wa

By the energy equations,

Qc = ~ A/ * -8.87 Btu/lb dry air


Wa

Qh = hi — hdi - 3.63 Btu/lb dry air

The heat transfer required to cool and condense the water removed is given by

tOa

taking h»i as hg at TST, the beat transfer is

0.0038(1094.5 - 23.4) » 4.07 Btu/lb dry Ail


494 AIR-WATER VAPOR [26-9

Then the fraction of tiie heat removed is

4.07
0.46
8.87

It is seen that in this instance more than half tlie cooling load is
used, not for the primary purpose of removing water, but for the
unavoidable cooling of the mass of gas and vapor, which must then
be reheated. The cost of such incidental cooling and reheating be-
comes relatively great as the desired final moisture content becomes
very low. Therefore for industrial processes requiring very low
humidity it is not unusual to use either chemical dehumidif3dng or a
regenerative type of cooling and heating system with a coimterflow
heat exchanger in which a large part of the cooling and heating are
accomplished.
25-9 Cooling towers. An important application of the humidi-
fying process is in cooling towers. A cooling tower is an apparatus
for reducing the temperature of water by evaporating a portion of
the water into the atmosphere. Cooling towers are used for cooling
the circulating water for power plants, refrigerating plants, and indus-
trial processes where it is impractical to transfer heat to a natural
body of water. Evaporative cooling has several advantages over
cooling by simple heat transfer to the atmosphere; more energy can
be transferred to a given amount of air, smaller apparatus is needed,
and the lowest available temperature is the wet bulb instead of the
dry bulb. A disadvantage is the loss of the evaporated water.
Figure 25-10 is a diagrammatic sketch of one type of cooling tower,
an induced draft tower with wood-slat packing. The packing is a
net-work of wood arranged to present a large, well-distributed wetted
surface to the air passing up through the tower. The water drips
over the packing and collects in the sump.
The Internal analysis of a cooling tower is a problem in heat
transfer and mass transfer; these phenomena determine the size of
tower needed. Thermodynamic analysis from the external viewpoint
cannot determine tower but can determine linxiting tempera-
size,

tures and the relative flow rates of air and water to satisfy given
conditions. For the tower of Fig. 25-10 the steady flow equations
are
Wa + W„i + Wx = Wa + Wya + V>v
rejii -f wjix = Wahi + Wyhy
26-0] AIR-WATER VAPOR 495

Fig. 25-10. Induced-draft cooling tower; flow diagram.

where hi and ht are mixture enthalpies per pound of dry air. If the
fan were included within the control volume, a shaft work term would
appear in the energy equation, but this may often be considered to

be negligible or to balance external heat losses.


The temperatures at which a cooling tower operates depend upon
the purpose and the economics of the installation. The minimum
temperature to which the water can be cooled is the adiabatic satura-
tion temperature, or in effect the wet-bulb temperature of the
almos-

phere. This minitnuTn cannot be reached with a reasonable size of


the
tower, but an “approach” (difference between the temperature of
wet-bulb temperature of the entering air) of
leaving water and the
requiring a larger tower.
5 to 20“F is reasonable, the smaller value
air leaving a tower may be at 90 to 100 percent
relative humidity.
The
The temperature of the warm water coming to the tower may be
comes. For a
limited by the conditions of the process from which it
496 AIR-WATER VAPOR
given amount of heat to be dissipated the tower will be smaller as the
water inlet temperature is higher.
Since the operation of a cooling tower depends upon atmospheric
conditions, designs must be based upon records of temperature and
humidity for the location of the tower. Values of dry-bulb and wet-
bulb temperatures suitable for design purposes have been compiled
for numerous localities and may be found in such references as the
ASHVE Guide and the ASRE Data Book.
PROBLEMS
25-1. Atmospheric air is at a pressure of 14.7 psia and temperature 76^F.
Plot the mixture density vs. relative humidity as the latter varies from 0 to 1.00.
What percentage error would be made in using dry-air specific volume for a cal-
culation involving air of 80 percent relative humidity at 75'’F7
25-2. Condensation on cold water pipes often occurs in warm humid rooms.
If the water temperature may reach a minimum of 45°Fand the room tempera-
ture is kept at 75^F, what is the limit of relative humidity in the room to avoid
condensation at normal atmospheric pressure? What is the limit at a barometer
of 26 in. Hg?
25-3. Air at 1 atm has a dry-bulb temperature of SO^’F and a wet-bulb
temperature of 70T. Find: (a) the specific humidity; (b) the relative humidity;
(c) the enthalpy, Btu/lb of dry air.

25-4. The air of Problem 25-3 is heated in steady flow to 135"F, 1 atm.
Find: (a) the final specific humidity; (b) the final relative humidity; (c) the heat
transferred, Btu/lb of dry air; (d) the heat transferred, Btu/lb of mixture;
(e) the heat transferred, Btu/1000 cu ft of mixture.
25-5. The Carrier psychrometric equation, used in constructing psychro-
metric charts, may be written
« (p PgwbXUb ty,t)

2830 - 1.44^„6

where is the partial pressure of the water vapor, is the saturation pressure
at the wet-bulb temperature, p is the total pressure of the mixture, tab is the dry-
bulb temperature, and twb is the wet-bulb temperature.
On Uie basis that the wet-bulb temperature is the same as the temperature
of adiabatic saturation, derive the Carrier equation from £qs. 25-7, 8 and 9b.
25-6. Three thousand cfm of air at
atm, 69°F, 60 percent relative humidity,
1

are to be humidified by passing through a water spray chamber. The water in


the spray chamber is continuously recirculated through the sprays and back to
a reservoir; heat is supplied to keep the reservoir temperature constant, and
make-up water is supplied at 60°F to compensate for the evaporation. Assum-
ing that external heat losses and the work of the spray pump are negligible, if

the air comes out at fiO^’F, 80 percent relative humidity, find: (a) the rate of
make-up water supply; (b) the rate of heat supply.
25-7. Solve I^blem 25-6 if the air comes out at 60T, saturated.
A1R-WAT£R VAPOR 497

25-8. Solve Problem 25-6 if the air comes out at 55**F, saturated.
25-9. What will be the final temperature, relative humidity, and specific
humidity of a stream of air obtained by mixing adiabatically at 1 atm, 1500 ofm
of air at 78®F, 80 percent relative humi^ty, and 500 cfm of air at 65®F, 50 percent
relative humidity?
25-10. In a ventilating system fresh air is to be taken in at 40^F, 60 percent
relative humidity. The air will be split into two streams, one of which will be
heated and humidified to saturation, while the other will simply be heated. The
two streams will then be mixed, to obtain the desired final state of 70^F, 65 per-
cent relative humidity. Find: (a) the temperature to which the saturated stream
is heated, if the other stream is heated to SS'^F; (b) the fractions into which the
total stream is split.

25-11. Four thousand cfm of air at 75**F, 70 percent relative humidity, are
to be cooled and dehumidified to 68T, 60 percent relative humidity by cooling
and reheating the entire fiow at 1 atm pressure. Find: (a) the temperature to
which the air must be cooled; (b) the tons of refrigeration required; (c) the
Btu/hr of heat required.
25-12. The requirements of Problem 25-11 can also be satisfied by a system
in which only a portion of the total air flow is cooled and dehumidified, and is
then mixed with the remainder of the air to obtaio the final state desired. In
such a process, if the cooling is carried to 45^F find: (a) the fraction of the entire
flow which passes through the cooler; (b) the tons of refrigeration required;
(c) the Btu/hr of heat required. Compare with results of Problem 25-11.
25-13. Air enters a cooling tower at 1 atm, 95®F dry-bulb, 78®F wetrbulb
temperature; the air leaves the tower at 90*’F, and 95 percent relative humidity.
Water is cooled from 100®F to 85®F in the tower; the water flow rate to the
tower is 100 gpm. (a) What air flow rate (cfm) is necessary? (b) What frac-
tion of the water evaporates? (c) Could the desired cooling have been accom-

plishedby simple heat transfer to the air in a heat exchanger? Explain.


The cooling water for an internal combustion engine is to be cooled
25-14.
from 150®F to 110®F by one of the following methods: (a) heat transfer to city
water which from 70®F to 130®F; (b) heat transfer to atmospheric air
will rise
which from 90®F to 120®F; (c) use of a cooling tower which will receive
will rise
air at 90°F diy bulb, 75®F wet bulb, and discharge air at 95®F, 90 percent relative
humidity. Evaporated water is made up by city water at 70®F, Compare the
three methods with respect to air required, and city water required, per 1000
gallons of cooling water cooled.

REFERENCES
American Society of Heating and Ventilating Engineers, Healing and Ventilating
Guide. Published annually.
American Society of Refrigerating Engineers, Refrigerating Data Book. Published
periodically.
PaJmatier and Wile, ''A New Psychrometric Chart” R^Hgerating Engineering^
V. 52, no. 1, 1946, p. 31.
Karig, H. E., /Tsychrometric Charts for High Altitude Calculations.” Refriger*
oHng Engineering, v. 52, no. 5, 1946, p. 434.
-

498 AIR. WATER VAPOR

Rohsenow, W. M., "Psychroinetric Determination of Abeolute Homidity at


Elevated IVeaBures," Refrigerating Engineering, v. 61, na 6, 1946, p. 423.
Carrier, W. H. and othos. Modem Air Conditioning, Heating, and Ventilating,
New York: Pitman, 1960.
Jordan, R.C. and G.B.PrieBter, Aj/nperaltbn (tad Aw Conditiionmp. New York:
Pnnti(»-HaU, 1948.
APPENDIX

Definitions of Symbols 500

Table A1: Propebties of Gases 502

Table A2: Combustion Data 504

Table A3: Conversion of Units 506

Table A4: Specific Heat at Constant Pressure, and Spe-


cinc Heat Ratio, for Gases at Low Pressure 507

Table A5: h/T for Gases at Low Pressure .... 508

Table A6: Specific Heat at Constant Pressure, h/T, and


Specific Heat Ratio 509

Table A7: Generalized Compressibility Charts for


Gases 510

Table A8: Air Tables 512

Table A9: Steam Tables 514

Table AlO: Ammonia Tables 522

Table All: Mercury Table 526

499
Definitions of Symbols

Symbols of general interest are defined below, while symbols of limited interest
have been defined where they first occur m the text. The units given are as
customarily used, but their exclusive use is not implied.

a or A area, sq ft
c heat capacity, Btu/lb ®F
Cp specific heat at constant pressure, Btu/lb
Cv specific heat at constant volume, Btu/lb
Cf coefEicient of velocity, dimensionless

Cw coefficient of discharge, dimensionless


C centigrade temperature, degrees
E internal energy of a systm.^ Btu or ft Ibf

F Fahrenheit temperature, degrees


F force, Ibf

g acceleration of gravity, ft/sec*

Qq constant in Newton’s second law; 32.17 Ibm ft/lbf sec^


h specific enthalpy, Btu/lbm or ft Ibf/lbm
H enthalpy, Btu or ft Ibf

J ratio of work unit to heat unit; 778.16 ft Ib/Btu


k ratio of specific heats, Cp/cv
K Kelvin temperature (absolute centigrade)
I length or distance, ft
In natural logarithm
m mass, Ibm
M molecular weight, Ibm/lb mol
n number of mols
n exponent in » constant
p pressure, Ibf^q ft

Q heat transferred, Btu


R Rankine temperature (Fahrenheit absolute)
R gas constant, ft Ibf/lbm ^R
S universal gas constant, 1545.3 ft Ibf /lb mol °R
8 specific entropy, Btu/lbm ^R
S entropy, Btu/®R
t temperature, F or C
T absolute temperature, R or K
u specific internal energy of a svbatancey Btu/lbm
U internal energy of a svbatancef Btu
V specific volume, cu ft/lbm
V volume, cu ft

V velocity, ft/sec
10 mass flow rate, Ibm/sec
IP power plant fluid rate, Ibm/hphr or Ibm/kwhr
W work, ft lb or Btu
500
Wg shaft work, ft lb or Btu
X vapor fraction of liquid-vapor mixture, fraction by mass
X mol fraction of a gas component of a mixture
z elevation above a datum level, ft

Greek Letters
a alpha nozzle angle, degrees
A delta mathematical symbol for “change of’
y gamma specific humidity
ly eta efficiency

^ phi relative humidity

Subscripts

A, Bf etc., or 1, 2, etc., identify a quantity with a certain point or path in a


process
/ saturated liquid state
fg difference between saturated liquid and saturated vapor
g saturated vapor state
h constant enthalpy
0 stagnation state
see also go

p constant pressure
s constant entropy
tor T constant temperature
u constant internal energy
X mixture state of quality x
see also Wx

501
..

Table Al; Properties of Gases}

Specific heats,
Btu/lbm “F Gas constant, R,
at 1 atm, ft Ibf/lbm ‘RS
Molec- ordinary room
For- ular temperatures ^/M^R pv/T-R
Ou mula weight Cp Ct e./c. at Op latm32*F

Airt 28.97 0.240 0.171 1.40 53.35 53.34


Monatomic
Gases
Argon A 39.94 0.123 0.074 1.67 38.68 38.65
Helium He 4.003 1.25 0.75 1.66 386.2 386.3
Diatomic Gases
Carbon monox-
ide CO 28.01 0.249 0.178 1.40 55.18 55.13
Hydrogen H, 2.016 3.42 2.43 1.41 766.6 767.0
Nitrogen N, 28.02 0.248 0.177 1.40 55.16 55.13
Oxygen 0, 32.00 0.219 0.156 1.40 48.29 48.24

Triatobuc Gases
Carbon dioxide. CO, 44.01 0.202 0.156 1.30 35.12 34.88
Sulfur dioxide. so, 64.07 0.154 0.122 1.26 24.12 23.65
Water vapor. . H,0 18.016 0.446* 0.336* 1.33 85.78 85.58*

Htdrocabbons
Acetylene C.H, 26.04 0.383 0.303 1.26 59.35 58.77
Methane CH. 16.04 0.532 0.403 1.32 96.35 96,07
Ethane C,H. 30,07 0.419 0.342 1.22 51.40 50.82
Iso-butane C.H„ 58.12 0.398 0.358 1.11 26.59 25.79

}Data mainly from U.S. Department of Commerce, Bureau of Standards


Circular No. C 461 ;
and from Eshbach, Handbook of Engineering Fundamentdla.
New York: Wiley, 1936.
*pvlT for water vapor at 1 psia, 300°F, data from Keenan and Keyes.
Cp and for water vapor at pressures below 1 psia.
§ The universal gas constant is taken as R
^ 1.986 Btu/lb mol deg R - 1545.3
ft Ibf/lb mol deg R.

t The composition of air, percent by volume, is taken as Ns, 78.03; Os, 20.99;
A, 0.98; following Keenan and Kaye.

602
.

Table Als Concluded

Mol Van de Waala


volume Critical constants constants
cu ft/lb (a) (b)

mol,
at 1 atm, Temp. Pressure atm ft* ft*

Gas 32T atm (Ib mol)* lb mol

Airf 359.0 -221.3 37.2 343.5 0.686


Monatomic Gases
Argon 359.6 -187.7 48.0 346 0.617
Helium 359.2 -450.2 2.26 8.57 0.372
DuToiac Gases
Carbon monoxide. 358.8 -218.2 35.0 381 0.639
Hydrogen 359.3 -399.8 12.8 62.8 0.426
Nitrogen 358.9 -232.8 33.5 346 0.618
Oxygen 358.7 -181.8 49.7 349.5 0.510

Triatomic Gases
Carbon dioxide 356.6 88.0 73 926 0.686
Sulfur dioxide 350.6 315.0 77.7 1737 0.910
Water vapor 705.6 218.5 1400 0.488

Htdbocarbons
Acetylene 355.6 103.6 62.0 1129 0.8232
Methane 358.0 -116.6 45.8 581.2 0.6856
Ethane 355.1 90.1 48.2 1391 1.028
Iso-butane 348.2 273.2 36.9 3265 1.807
.

Table A2: Combustion Dataf

Molecular
Substance Formula weight Combustion reaction

Hydrogen H,(gas) 2.016 2H, + 0, -» 2H,0


Sulfur S (solid) 32.07 S + Oj — SOi
Carbon C (solid) 12.01 C -|- 0* COf
C + §0, CO
Carbon monoxide. CO (gas) 28.01 CO + io, COi
Methane CH4 (gas) 16.04 CH. + 20» -+ COi + 2H,0
Ethane C,H. (gas) 30.07 CsH, + 3iO, - 2CO, + 3H,0
Propane C,H, (gas) 44.09 C,H, + 60» - SCOt + 4H,0

Butane CiHio (gas) 68.12 CAo + 6iO, 4CO, + 6H,0


Octane CJI„ (liq) 114.2 C^Hia -J- 12i0i —* 8C0i 9HjO
Benzene C,H, (liq) 78.11 C.H, + 7JOi - 6COs + 3H,0
Acetylene C,H. (gas) 26.04 C,H, + 210, 2CO, + H,0

Material balance Heatof combustion ( — A/i)


(Ib/Jb of fuel) (Btu/lb of fuel, at 77^F)
Required Produced Liquid H2O All gaseous
Substance 0, Air CO, H,0 in products products

Hydrogen 7.94 34.2 8.94 60,958 51,571


Sulfur 1.00 4.31 2.00(S02) 3,895
Carbon 2.66 11.5 3.66 14,087
1.33 6.75 2.33(CO) 3,952

Carbon monoxide 0.571 2.46 1.57 4,344


Methane 3.99 17.2 2.74 2.25 23,861 21,602
Ethane 3.73 16.1 2.93 1.80 22,304 20,416
Propane 3.63 15.7 3.00 1.63 21,646 19,929

Butane 3.58 15.5 3.03 1.65 21,293 19,665


Octane 3.60 15.1 3.08 1.42 20,591 19,100
Benzene 3.07 13.3 3.38 0.69 17,986 17,269
Acetylene 3.07 13.3 3.38 0.69 21,460 20,734

* Small differences between the material quantities in this table and those
computed in the examples in the text are due to rounding off the values of the
molecular weights in the examples.
t Data mainly from U.S. Department of Commerce, Bureau of Standards
Circular No. C 461.

S04
Table A2x (Concluded)

haknl Heats of Vaporization^ h/g^ Btu/lb, at 77^F


and Boiling Point at 1 atm (NBP) for Some Hydrocarbons
Compound hfgj Btu/lb NBP,T
C,H, 147.07 -44
CiH,, 155.83 31
CJIu 156.14 258
C,H. 186.31 176

Specific Heats of Certain Fuel Svbttanca Btu/lb


,
T
Substance Cp at TTT c, avg 77*F to
Carbon 0.172 0.205
Sulfur 0.168 0.178
Average for
petroleum liquid fueb 0.45 0.50

505
t

Table A3: Convenion of Units

CJass of Units To obtain multiply by

liOngth
cu
cm
in.
mmgmm 30.480
1,728
Volume cu cm 28,320
cu in. gallons 231
Mass grams Ibm 453.59
Ibm slugs 32.17
Specific volume cu cm/gm cu ft/lbm 62.428
psf psi 144
in. Hg psi 2.036
Pressure ft HO 2 psi 2.309
psi atm 14.696
psi kg/sq cm 14.223
IT calorie* Btu 251.996
ftlbf Btu 778.16
Energy hphr Btu 3.93010 X 10“^

kwhr Btu 2.93018 X 10-*

Btu hphr 2544.46


Btu kwhr 3412.76
Specific energy IT calorie/gm Btu/lbm 0.555556
Specific energy /deg IT cal/gm ®C Btu/lbm ®P 1

Specific entropy IT cal/gm “K Btu/lbm ®R 1

'
IT » International steam tables.

Temperature Conversimt
•P 1.8"C + 32
“R - ‘P + 459.69
•K - *C + 273.16
"R - l.g'K

t Prom U.S. Department of Commerce, Bureau of Standards Circular No.


C461.

506
Table A4i Specific Heat at Constant Pressure, and Specific Heat Ratio,
for Gases at Low Pressure* (as functions of temperature)
0.340 rLL Li_liJ_J_L_l T M -r r T-flT 7TTt T1 1 I 1 I I I ri T ni I' l I iTI -
TT I [ I I

0.330

0.3201

0.310

O.^Odi

0.290

0.280
»qoj*q

0.270

nil 0.260

1000 2000 3000 4000 5000


TEMPERATURE-*F obs

507
Table A5: b/T for Gasee at Low PreMure*

-BTU/lb*Fab«

h/T

0 100 0 2000 3000 4000. SQOO


TEMPERATURE-*Fobt
* Based on data from the Gas Tables of Keenan and Kaye.

508
;
F+rLR-i-FH-rr t.i H4t-K'^1Tm4-l-j-r!-H I Hff44-1-j f t >-H-R
0 1000 2000 3000 4000 5000
TEMPERATURE-*Fobs
Based on data from the Gas Tables of Keenan and Kaye.
509
Table A7 s Generalized CompreMiblUty Charta for Gaaea*

Wiley, 1939, pp. 108-109.

610
611
Table A8: Air Tables (for one pound)
(The properties given here are condensed by permission of authors and publisher from Qat Tohlss, by J. H. Keenan
and J. Kaye, published by John Wiley and Sons, 1948.)

I
I
I SfSSS
I
§§§1S I
I
I
I SSS§S

gssill I
I
I
I
ISSII I ssass

I
'
'
'
lllli ' s?s“a

ISIIS
?S8iS

SISIS
§§ll§

Igsgg

SISI5
SSiSS

ISISS
gSSii

SSS§§

iiiii

i§|
IIS

512
(For One Pound)
niie properties given here are condensed by penninion of authors and publisher from Qm Foblsi. by J. H. Kecnao
and J. Kaye, published by John Wiley and Sons, 1948.)

r.
"F
abs
m Pr
'

Btu/
u,

lb
B Btu/
lb"F
T,
•F
abs
t,

“F
A.
Btu/
lb
Pr Btu/
lb
Vr
0.
Btu/
lb*F

2300 1840 588.8 308 431.2 2.76 .9712 3000 2540 IBfll 941 Btfll npi 1.0478
2320. 1860 594.5 319 435.5 2.69 .9737 3020 2560 796.5 %9 589.5 1.155 1.0497
2340 1880 600.2 331 439.8 2.62 .9761 3040 2580 802.4 996 594.0 1.130 1.0517
2360 1900 605.8 343 444.1 2.55 .9785 3060 808.3 ohH 598.5 1.106 1.0536
2380 1920 611.5 355 448.4 2.48 .9809 3080 2620 814.2 1054 1.083 1.0555

2400 1940 617.2 368 452.7 2.42 .9833 2640 lllfl 1.0574
2420 1960 622.9 EVrKm 2.36 .9857 3120 2660 825.9 1114 612.0 1.038 1.0593
2440 1980 628.6 394 461.4 wEm 3140 2680 831.8 1145 616.6 1.016 1.0612
2460 2000 634.3 407 d65.7 2.24 3160 2700 837.7 1176 621.1 .995 1.0630
2480 2020 640.0 421 2.18 .9927 3180 2720 843.6 1209 625.6 .975 1.0649

2500 2040 645.8 436 474.4 2.12 3200 2740 849.5 1242 .955 1.0668
2520 2060 651.5 478.8 2.07 .9972 3220 2760 855.4 1276 634.6 .935 1.0686
2540 2080 657.2 466 483.1 mMSm .9995 3240 2780 861.3 MtH 639.2 .916 1.0704
2560 2100 663.0 481 487.5 1.971 lUiMM 3260 2800 867.2 1345 643.7 kOTO 1.0722
2580 2120 668.7 497 491.9 1.922 1.0040 3280 2820 873.1 1381 648.3 .880 1.0740

2600 2140 674.5 514 4%. 3 1.876 1.0062 3300 2840 879.0 1418 652.8 .862 1.0758
2620 2160 680.2 530 ULtltl 3320 2860 884.9 1455 657.4 .845 1.0776
2640 2180 686.0 548 1.786 3440 2880 1494 661.9 .828 1.0794
2660 2200 691.8 565 509.4 1.743 1.0128 3360 2900 896.8 1533 666.5 .812 1.0812
2680 2220 697.6 583 513.8 3380 2920 902.7 1573 671.0 .796 1.0830

2700 2240 703.4 602 518.3 1.662 1.0171 3400 2940 908.7 1613 675.6 .781 1.0847
2720 2260 709.1 621 522.7 1.623 1.0193 3420 2960 914.6 1655 .766 1.0864
2740 2280 714.9 527.1 1.585 1.0214 3440 2980 920.6 1697 684.8 .751 1.0882
2760
2780
2300
2320
720.7
726.5
660
681
531.5
536.0
1.548
1.512
1.0235
1.0256
3460
3480 3020
926.5
932.4
Ed 1784
689.3
693.9
.736
.722
1.0899
1.0916

2800 2340 732.3 702 540.4 1.478 1.0277 3500 3040 938.4 1829 698.5 1.0933
2820 2360 738.2 724 544.8 1.444 1.0297 3520 3060 944.4 1875 703.1 .695
2840 2380 744.0 746 549.3 1.411 1.0318 3540 3080 1922 .682 1.0967
2860 2400 749.8 768 553.7 1.379 1.0338 3560 956.3 712.2 1.0984
2880 2420 755.6 791 558.2 1.348 1.0359 3580 3120 962.2 716.8 .657
1

2900 2440 761.4 815 562.7 1.318 1.0379 3140 968.2 2068 721.4 .645
2920 2460 767.3 839 567.1 1.289 1.0399 3620 3160 974.2 2118 .633 1.1034
864 571.6 1.0419 3640 3180 mMM 730.6 1.1050
2940
2960
2980
2480
2500
2520
773.1
779.0
784.8
889
915
576.1
1.261
1.233 1.0439
1.0458
3660
3680
KFiTil 986.1
3220 992.1
2222
2276
735.3
739.9
mi
.621

.599
1.1066

ExAMPiiB 1. CoMPRBBBioN OF AiB IN Stbadt Fz<ow. Air at a pressure of 1 atm abs and a tem-
perature of 520 F abs is compressed in steady flow to a pressure of 6 atm abs. Find the work of com-
pression and the temperature after compression for (1) 100% efficiency of compression and (2) 60%
efficiency of compression. The efficiency of compression is here defined as the ratio of the isentropic
work of compression to the actual work of compression.
Solution. (1) From Table 2 we get for Ti — 520 F abs,
Pri - 1.215, hi - 124.3 Btu/lb
where subscript 1 refers to the state at the To determine the properties at the
compressor inlet.
compressor outlet for isentropic compression we compute the relative pressure there
Pr2 - 6/1 X 1.215 - 7.29

Interpolating in Table 2 with this value of pr. wo find, for h 2« and Tu, the enthalpy and temperature
at the compressor outlet for isentropic compression
h 2B - 207.6 Btu/lb, Tu « 864.6 F abs
The work of compression for 100% efficiency is then

h 2g “* h\ * 83.3 Btu/lb

(2) Since the efficiency of compression is defined by the equation q » (hu hi)/work per potud,
we have for 60% efficiency
83 3
Work per pound — j-
0.60
188.8 Btu/lb

513
^.mu Mus sssss ISUi sau;

1141111
^ mMcietm
mi iiiil iiiiriiiii ilili
c«eie4e9*^

{ |.4Pi
eiNOlMCl
iiiii iiiil iiiii iiiil liiH

ooooo ooooo ooooo ooooo ooooo ooooo

a
I ^ ^ d ooMco«<D ctmtooM ««oo^ea c4c(>e««40k ammcsm M»r*>cot*io

Table
I I iiiii iiiil iiiig siiii HiH
Steam
^„ ^S838& 838S8 SSSSS SS8SS S&S83 9S38S
-g-Bg |:feS|S gggll gSgll S||§|

A9:

8S8S8 sssii sSsSS


Table

li'— ,52 83!58& £:8S$S sisdss SSS”*^


* iiiil
<3
s
I i
P t
5
H
1
«Dao^ec9 »88S8 t;S53SS
{:8SS8
sssiS
K<»oeo a»b
nm
m^obt*
p§l§ iiSSS

1703.
3306 2Q47 2444 2030

e>K»<^r»co ONh-^ei
§§§!§ 3SSSS
9«0cp(0 sSilS ssssi OCPC0»S
ooobb bbobb bbooo booob bbbbb obobo
eooee bbbbb bbbbb bbbbb bbbbb bbbbb
«iOOC«Pi«
iOAr»iOi-<
SS«i«»»ao 8S8|»
iiiii SisqlR
bbbbb bbobb ririddii
sips nUH
bbr»*b^ bbfe^bb
§§§13
xt^MC^eM 88888

h 8883S 888S§ 8SS3S 3SSSS SS8S2 S3S8f


514
tSHi UHi n»i usH a«i nin a JVederiek

and

H.Keeiiio

iiiii ggisi Sc Joseph

40 ooeoo ooooo bbo'do'


ooooo ooooo
bbbbb o
by
1937,

ISlil iiiii iiiii iiiii Iiiii iilli if


bbbbb bbbbb bbbbb bbbbb bbbbb bbbbb bi^
Copyright,

co^eo«i40 o«i4^ooe» oomkoo

iiiii iiiii iiiii iiiii' iiiii iiiii si Keyes.

Q.
•NOOOO r*C«iO<D'«> OeomO‘«i* lOlDOMek ^
iim iim mii Mii iuii iim r IMerfok

SS^SS fH^oo^o eoooevo okomco coi*


asd

mni imi iiiii iiiii iim ssisg ii Swnui

H.

Sipg s§pi iiiii iiiii iiiii iiiii 11


bbobb bbbbb bbbbb bb
Joseph

bbbbb bbbbb
by

iisig mn liip Iiiii iiiii iiiii i<


bbbbb bbbbb bbbbb bbbbb bbbbb bbbbb b
Tock,

Now

**HiermodynuiuoProperti«BOf'StMm'*

1m.»

iiiii gill iiiii iii,, ,,, Bone,

bbbbb bbbbb bbbbb bbbbb bbbbb bbbbb bb 4


WXhqr

John

hgr

^Abridfedfrom

Fi^lbod

mil iim mil mil iim mil ii a


615
*

5sg8S SSSSS SSSiSS 8SS8S



» s
I
b MO»r«ci^ ^«tf*c9Q0«« oaooif-ioo wotocor* 0»<-«e«c9>i4 »t.io«>aio

§§§ig SSs'iS §l§il li§§§ §§S§§

•*5 P 5 *^!S^ cieo^co qommc^


Id -3 fr-.o>e4
ao «-4 a>c^c^*-4 tH oo^cot* =$;8Si
?acoq»oo
<oa»OMra
b^bboo r«ioeoao NOO’u*

^ 88 s;s;s;

II

Table

u oe««eo*^ ookco^co ^aoec«D^ cotHtooi^ cieoctMOk


Steam

^
^8873 SSSSS {zgisss sssss
liSlSSSZS IZSIhI^m^ snzjziS

A9: o, €Oe*Ct^O C9^tOC9^ C0 ^-P^«-4 C0 C4 r*«DO (0 0 t00>i0 »-*


I

*3 O^COOO^Ok IA> 1-4 (nJ f4


ja coconcsi^ •“"-•OOO
Table
U pc bbooo O) O) O)

•~|Q(00>0
oocooco 9C
ox o»co«oo<p
CO *010^0
^3 00^00 !9
<S.?^SSS8S
iJ
S3SSS S;8$SS
!o7SSS «-4i-4i««^04 C404CMC<(C<I
oooooo
<

MCVIC4 MCO

oooo>-4 U)r« •OlOOtO


•Olootoe* 00 OC4 C4 O
•;2 r;ss S 3 ^?$
oo>o^ao
O0 ^'«»'i«t>»
to > 00 ^
»-<oc5 oo
^‘ootobr^ t^bobio
^S3 SSgJ5 ssftsis

eo^'CiiiSxs
«; 3 tooocob folotototo totototofo tototofoto
!
li

I
|g K ^
^
00000 00000 00000 00000 00000 00000 1-4 ro fN 1 i-«

odobd ddddd dddo'd ddddd ddddd doddd

y
Cl
^QOQOt^^
k* 0 '««i 0>c« 83 S 8S 88 S&S S 83 S& >:&SSS SSiSSK
I'
'"''•sssss
Sf-«<-ii-iM
ssasg assss
C4C^CMCMe>4
sgRss a&ssa
CstC^C^CMOl
sssgs
NC^COCOCQ COCOCOCOCO

90000
•.*« COOOlOO lOOlOOlO OlOOlOO
-^-lomco MOi^Sio oor-fe-o
ooooo
oo>oo«o
i jo 5 •

I -a
516
1 3

'5
SsSSS §!§!§ !§§§§ §§§§1 iilii §§§i
£
*s
ototoioei e)iOiHb.Q0 **«etOhotO <«CO<-iQ0^ «4tO'«<«e9 ««>t^Ok 2
^ M Q> o
ooOi-4i-4
1-4 ^cicr>co*A
th wiH i-< t-t »-i
r-^ 00 00 00 00 oo t-T »J jo u> ^ne^og* ^nai»tSfi
00000 lo o' 04 oi
gpcoi^r—
S
g
r4«-«1>4i-4i»l •-4f-ii>4T-tr’1 i-if^t-li-lr-l pim-*®*®® ^
MOQio-Ho esiMirtaO''^ o« ^ ”
Oeocooco ro-^OKdO oo^^ooo cdoi^o^ ot-Ofc«i-i e4cQio*o •£!

^^•a*o»«- ooc4ro^ *o«or^ooqj q-hcmcoco ddddd


cioo^dio
M^ciacoco
ddo^MO ciooddisj oooddoo oncodoo
uocdoooo
dh-'rocsi S'
coeoeoc'^ cococdcoco co<«^io<’4* ^ >0 kQioiotOkO tOiOkOioo
qo>-<coi^
ot-t^oo £9

tMr-poM ^to^4t«4«i eomor'-t'- t^O«11«4


.g2SS5
«0 001'- to CO
S ‘^010(0
iQ<Q4t<^eM
Oto^eoc^
ftoaoi'-to
i^kdr-oo) tOOOC4kOO
1-tOOOOO
g CDtOkOkO
r-coto^co
lOtOiOOtO UOi^tOtOtO CO CO CO CO CO cO

m 2222 iCOOOOh-
Cll^C4tpcO
CM^eor-4*t COCO^ft-O
00 -1 * ta^iiiioar^
3<o 4«OtO
k0tOOi-*O
Oi0i.4OtD
ssigi SooSo 088800 or - >0 CO 04
r'.r-r»t'-t»
Ooot^to3«
r- CD CD CO CO
C4 or-3iC4i
CO coco coco
4230
.3197 .1885

00000
«DIOOQO^ oto^toh- to-iittoua W^UOIOO io*HOcoe4
S
ssiss CMCOCO^tO t-kOi-llOCO
OOOCMra^
C'lOOCOt'-
2SRw3 £or:s$s
lOiOtCtOdd lOCOtOtOtO r«r^t-r-ao
do odd odddd

^I^O«-i«^ OOeO^f-4 000*00^ OOCOCIO Ofi^^l^OD OOtOccD^O f-tv-icor* *0


Q CO 9 d d to 00 2 d CO d d £0 £5 sj^d 00 i-’d cod dddcod lod dd
>-t I;;-'
; S
11 2222
S 11 11 1] 11
22222
1 1 1 1 11 22222
11 11 11 11 11 2
1m1 22
11§ 22222
1 1 22222
1 1 1 1 11 11 gi§
1111 2 1( 1( a

877.9 872.9 868.2 863.6 859.2 854.9 850.8 846.8 843.0 826.1 809.0 794.2 780.5 767,4 765.0 743.1 731.6 720.5 709.7 699.2 688.9 678.8 668.8 659.1 649.4 630.4 611.7 593.2 574.7 556.3 463.4 360.5 217.8

4*»iMCM*HCO
^OOGOiOOi 8322S S3o04 4«I 00 40 00 to 00 »-eototo^ t-COtOI-.
e4eo-««oio rS'S^S'S^ eoo'**'*^2
0004 00^
CO <«
8?:SSS 822^53
i^i >«• iSiOlOiOtO
411
iot«o6o>'4
lOiOiOiOtO

4**ooco4»teo
04000100
r^ii-oc-to
00e40.-ii0
^OCOWIO
ootooKS-i**
e^iAoii-ieo
t«4f«c4oa6
•'^S22C2SS2
£i<222?21
toto^cvtoo
coo-^tNr-.
^cmcocooj
ioco-'«oo
tooioo
S r-.c-.t0to tocoor-^
i0i0»0-o«'^
8o2o2
CO CO CO C4
4*1

coco CO cod dojcid*-* ddi-idd odddd ddddd

o^coto^ O-4CO«0t;
OOOOOO ooooo ooe»^<0 Oeor-.-i^to r.-h-oM
toco^o
ooooo iliil
iilii Oi-4«^i-ii-t oi 04 coco S'*

ooooo OOOOO -MCiKMCVlOl


ooooo C4C4C4C'4C'l
ooooo 33838 ooo8
ddddd ddddo ddddd oodod doddd ddddd odod

^1-4000
0C4O«-«O
eotoooeo^
<NC'«O'O*t0
h-v-t^tor- do4«<'dd QDic—too^ft ^I'.t-t'-to loooioio
s&sss
COCOCOeOCO
S{;^ 3 S
C0C0C0C04I
r-iOoOi-4 -icMeoto'^
iptOiOtlStO
«otpr-ooo
kOiOiOtOiO
<0 4000
tDtOfOt'-

SSSSS 3g§§§ §§§§§ l§§ll §§!§§ §1111 §§!

517
III

Table

Steam

A9:

Table
519
W4 *00

S
Si!
0>nO *4000 MCICD ^(00 0^10 <oe»c9

SSSS BUS S(0^


0»MI^
esit« •
-t* ‘1^ • ‘h* "i^ • • -(s. .
'

S’^.s 8^-1 S"".? S*^.! S ^.2 s*’?: 5 *?^ 2 ’^.g fe^S


i-.SS
C4<D
S? 5S
-lO
1:82
-<0

S 82
® 882 882 '5 ' • • •
fe 22 8 :2 J

5522
-O •
222
•« •
P4>H|-I ^*Hi-« Ok^M 0*^«^

c 8 ®S OACO ^OOM
5
2 "^.g 8*.2 2 ®S
Oe<^ OCOO> ^I-IOO
g"".g
NOOtp C4«iO
S’^.S
Sc>l^
§«S
»»C4
8 aoM»«
^ .lO .wo• -lO
WfMto 6r?<o
.Ift -lO • • . -ift . •
o><5<5
« •
ooo®
.»o * •!!» •
o8«
-^ •

2*^2 UdOC4
m-* -00
txcoao COION
•«
NOW W'C*^
W'C* oaoto ^NfH OlOlo or*
ONC^
§ 2^2 iS^S CO <0(0 2go ^gS ^(O b>^<0 <Oi«iiO <0
*H*ii4*^ r>4i^v^ 0^»^ <•« 0*-iv4 o»^^

^WO
i^ .A
lO^*H OOON Ot^W WNto
o -r* w ‘W o o •'«
(0<0N
CO
(OO^ OONO ^COW t»WQ
-r* o -w -C

S ssi
^ •
iss 2^2 322
-^ • •
5522
W •
222
-w
282
-W
322
TO
• • .
22:2
-w
*
CnOiO
S -3
(Continued)

Steam
%m ,
^10(0
SfSS
WiflcO
»H«-ll-<
2®o
:S2 232 sss
eo
•-I.^.H

Ort
<o


CO <0(0
oooo'jc

pHIfC
282
O^*^
OiOf^
(DNiO
w
0>-«*H

2 ”S S
2SCO

2 2
0*^ri4
• •(

0^*4

a 2*".2 lOO ^ww OOOO lO <^.2 ooo»« N'<'»-l


w4 •
N •-•h-N
CC -(O
N»«0
lO •*'4
III
N Q0*H 222 2g2
Soo'S NQOO I^IOIN*
Superheated
I 82 Or^lO
^ w •
2=S ^w 'ST
O' 0»H*^
I
Table

w o> oww
WOw
oONio O)aoi<» o>wo eooo
i IN,
•cr5!q»
•<< »o -00
NOc* S?
nin.o *2
oo:t
'«'
w K
i»Inn
N §:i
B NW>0
•w •
-«COIO
'W •
ON*0
-w •
»Ni
'W
6 tN„mo ssg
w •
N ri^r
of O
(fj
^,4

0*^»-c o;4^ dSJ^ o^^
Steam

«0^0> S®*°® ^•OW <OOI>- 0>'#iCO o^co


“-®
§iS3
« .CO
SSS
-W
SSS
-w
S&S
SSiS
-w-W
gss
-N ^
3S3 Sdg -N
9g$
*N - • • •
333
-N •

o«^
. • •

A9:
Pboferties

0«H^ OfH^ 0>^ tm.


©»
NCOlO Wl^^ 0»l»W t^coN er>o»o ooo WW NOM
Table

§S& OIO-S raWO NWO


§§31
•r^
is! 3 sl
•TO

S 33 is <or^<« lOo^ 4Ci25^ • •
*01 • •

^Pi4vi4 O**^^ oa-; dS-

ooio F4C02 NIvfN OlOO


2"-S »o_ S*!® .. -
ONIn. Sod^ •hO^ ^<0^
§ woo 822 882 ?sr
,-

“S'*.
ssis
Oi^F4 O^fH O^F^ d^^ O^fC Omm
OOOC miOO t^NN NrQOt^
NW? 5d2 S^2

NION
§”5
tdaoat
^ «esi >
§8S (ON'« N«iO>W
©F^F^

FC^IO t^W0» Nt.F^ ('•I'nCC

NodS O'-*© ©iOMI


§ '^-8^. *.8T *^.3'"
*a<FCF4 O^fC Of4^ Of^F

•»«•• •»< «• «»< «v «•<«•» •(£ i» <>< W «»*£• *»•«••

|i! d i! s! ij ?| i| i! i| sf s!
Ja* I s s I I I s "i "I "s
520
I-IOON IOU3M 0300^ ooioeo ecooM giog
ci .ua g^g wo>c
g«j5
iooo*o ^00*0 5$5I ?5SS
f-l®^ S$SS
•CO •
5.^.^ 0,^-4 O^iH OvXf^ O-HiM Of^f-i Or*»^ Oi-i*-i

Mt^K «OMSO» «DiO«^ Qi^c


00 -S
•DiD^
00 -o CO -CO CD N to
r^O»^
M M
^^00 N«iO-N
sss ^N*0 g2S coOto
^ §sP
dZ^ d2*^‘

•1 CO -CO
e^c^^•5
O -CO
^lOtO
r« t^
OOO
to -a*
MiOaO
00-0
^OSiH
CD
C400r<- OOO^
o» o s®s
oooa “it* *5 0»0*D b.r«cD looua co'«c^l r«cooi
«r<oio O^iD-O* eo^'4< CO CO
CO
-O'

n r-»CDC0 t^iDCO •-<eMC0 rH r«. CSi 0<-«N 3SS
d2t^‘ o2*4 o 2 >4
2*^5; OiOcO COOOW . ouao i^ooh- cocDtD
2”S
SidS 00 -00
<«0»^
00
«fiDO
00
AOa>
,-co .
.
O
mo>o 00
otcoco
»>. CO
frjcoc'*
rgr W’^'CO IHIXCO OcP'-t • OQOO> 0(0c» 0>00>

e«ioo» vhcok
^ ^ .r- S“S S«S5
222 22 •H l^OOOO co^ ^esi 0
^sr

?8 **S
<M|oC9 KOOO ^a>Q
ri ID en -CO ‘Ofio
•rocM fCOOO »o^®
N^eo ciaoeo
eoeoeo “
©2^' d2-i 62^
?^o» coco ID
CO
oo^ aioV-
s^g
0«i4p4 0*4*4 0*-C*4

eoooa
Q’h»
cQoooo
CO "CO
t-r-^
«j4co wfloeo
e>4 ^ o 2*4
2»oa - • •

jjcoS • • •
cO0»e4 • • •

d
Table AID: Ammonia Tables—Saturated Stateaf

Temp. Pressure Volume Density Enthalpy from Entropy Uom


-40 F

Abs. Gage Vapor Vapor liquid


Liquid Vapor
Latent
F Ib/in.* Ib/in.* ft«/lb IbM* Btu/lb Btu/lb
Bt^lb Btu/lb

t P Pd !/»• hf hfc «/ *0

-60 6.65 44.73 -21.2 589.6 610.8 -0.0517 1.4769


-66 6.54 38.38 -15.9 591.6 1.4631
-60 7.67 •14.3 33.08 -10.6 593.7 604.3 -0.0256 1.4497
-46 8.06 •11.7 28.62 0.03494 - 5.3 595.6 600.9 -0.0127 1.4368
-40
-38
•8.7
•7.4
24.86
23.53
cm
.04251
0.0
2.1
597.6
598.3
597.6
596.2
1.4242
.4193
-36 11.71 •6.1 22.27 .04489 4.3 699.1 594.8 .0101 .4144
-34 12.41 •4.7 21.10 6.4 599.9 593.5 .0151 .4006
-32 13.14 *3.2 20.00 .04999 8.5 600.6 592.1 .0201 .4048
-30 KJLIL •1.6 18.97 10.7 601.4 590.7 1.4001
-28 14.71 0.0 18.00 .05555 12.8 602.1 589.3 .3955
-26 16.55 0.8 17.09 14.9 602.8 587.9 .3909
-24 16.42 1.7 16.24 .06158 17.1 603.6 586.5 .0399
-22 17.34 2.6 15.43 .06479 19.2 604.3 585.1 .0448 .3818
-20 18.30 3.6 14.68 0.06813 21.4 605.0 583.6 1.3774
-18 19.30 4.6 13.97 .07161 23.6 605.7 582.2 .0545 .3729
-16 5.6 13.29 .07522 25.6 606.4 580.8 .0594 .3686
-14 21.43 6.7 12.66 .07898 27.8 607.1 579.3 .0642 .3643
-12 22.56 7.9 12.06 .08289 30.0 607.8 577.8 .3600
-10 23.74 9.0 11.50 0.08695 32.1 608.5 576.4 1.3558
- 8 24.97 10.3 10.97 .09117 34.3 609.2 574.9 .0786 .3516
- 6 26.26 11,6 10.47 .09555 36.4 609.8 573.4 .0833 .3474
- 4 27.59 12.9 9.991 38.6 610.5 571.9 .3433
- 2 28.08 14.3 9.541 40.7 611.1 570.4 .0928 .3393
0 15.7 9.116 42.9 611.8 568.9 WjliTO 1.3352
2 31.92 17.2 8.714 .1148 45.1 612.4 567.3 .3312
4 33.47 18.8 8.333 mssm 47.2 613.0 565.8 .3273
6 20.4 1
7.971 .1254 49.4 613.6 564.2 .1115
8 36.77 22.1 7.629 .1311 51.6 614.3 562.7 .1162 .3195
10 38.51 23.8 53.8 614.9 561.1 1.3157
12 KliXli] 25.6 6.996 .1429 56.0 615.5 559.5 .1254 .3118
14 42.18 27.5 6.703 .1492 58.2 616.1 557.9 .3081
16 44.12 29.4 6.425 .1556 60.3 616.6 556.3 .1346 .3043
18 46.13 31.4 6.161 .1623 62.5 617.2 554.7 .1302 FOTl
20
22
48.21 33.5
35.7 5.671 .1763
64.7
66.0
617.8
618.3
553.1
551.4
mgm.1483
1.2969
.2933
24 52.59 37.9 5.443 .1837 69.1 618.9 549.8 .1528 .2897
26 40.2 6.227 msmm 71.3 619.4 548.1 .1573 .2891
28 57.28 42.6 .1992 73.5 619.9 546.4 .1618 .^5
80 59.74 45.0 4.825 0-.2073 75.7 620.5 544.8 1.2790
32 62.29 47.6 4,637 .2156 77.9 621.0 543.1 .2755
84 64.91 mfjwm 4.459 .2248 80.1 621.5 541.4 .1753 .2721
86 67.63 52.9 4.289 .2332 82.3 622.0 539.7 .1797 .2686
88 70.43 56.7 4.126 .2423 84.6 622.5 537.9 .1841 .2652

* Inches of mercury below one atmosphere.


t Abstracted.
Ammonia," U. 8.
^permission, from Tables of Thermodsmamic Properties of
Dep^ment of Commerce, Bureau of Standards Circular No. 142,

522
Table AlO: Ammon ta Tables —Superheated Ammonia
Absolute Pressure in Ib/in.* (Saturation Temperature in italics)

Temp. 60 60 70 80
F 21.67 30.21 37.70 U.40

t V h 8 V h 8 h a V h a

Sai. 5 710 618 S 1 . 2939 A. 805 620 B 1 2787 4 IBl 622 4 1 2658 3 655 G2A 0 1.2645

80
40
5 838 623 4 1.3046
5 988 629 5 3169 4 933 Hi 1.2913 4 177 623 9 1 2688

50 6 136 635 4 1.3286


641.2 3399 5 184
IKS 639 0
1 3035 4 290 630 4 1.2816 3 712
3152 4 401 6.^6 6 2937 3 812
627
634
7
3
1 2619
2746
60
70 6 423 646 0 3508 5 307 644 9 .3265 4 500 642.7 .3054 3 909 640 6 2806
80 6 564 662 6 .3613 6 428 650 7 .3373 4 615 648 7 .3160 4 005 046 7 20 s 1
90 6 704 668 2 .3716 5 547 656 4 .3479 4 719 654 6 .3274 4.098 652 8 30VI2

100 0.843 663 7 1.3816 662.1 1.3581


5 065 4 822 660 4 1.337S 4 100 658 7 ] 3199
no 6.080 WMm
674 7
3914
.4009
667 7 .3681
5 781
673 3
5 897 3778
4 924
5 025
666 1 3480 4 281 KliIXE
671 8 .3579 4 371 670 4
3303
3404
120 7 117
130 7 252 680,2 4103 6 012 678.9 .3873 5 125 677 5 3676 4 460 676 1 .3602
140 7 387 686 7 .4195 6.126 684.4 .3966 5 224 683 1 .3770 4 548 681 8

160 7 521 691 1 1.4286 6 239 689 9 1.40.58 5.323 688.7 1.3863 4 635 687.5 1 3692
160 7 656 4374 BKisiy 695 6 .4148 5 420 604 3 3054 4 722 693 2 .3784
170 7 788 702.1 .4462 6 464 EZiTWil .42.36 5 518 690 9 .4043 4 808 698 8 .3874
180 7 921 wuiijM .4548 6 570 706 5 4323 5 615 705 6 .4131 4 893 704 4
100 8.053 .4033 |rJ^
4409 5 711 711.0 .4217 4 978 710,0 .4060

8 185 718 5 1 4716 6 798 717 5 1.4493 5 807 716 6 1 . 4302 5 063 715 6 1.4136
8 317 724 0 . 4799 6 909 723 1 .4576 5 002 722 2 4386 5 147 721 3 4220
8 448 720 4 4380 7,019 728 6 4658 5 098 727 7 .4-i60 5 231 726 9 ,4304
8 710 740.5 5040 7 238 739 7 .4819 6 187 738 9 .4631 5.398 738 1 .4467
8 970 751.6 .6197 7 457 750 9 4970 6 376 750 1 .4789 5 565 749 4 .4626

880 9.230 1 5350 7 675 762 1 1 5130 6.563 761 4 1 4943 5 730 760 7 1.4781
300 9 489 774.0 6500 7 892 773 3 .52816 750 772 7 5095 5 894 772 1 4933

Temp. 120 140


F BO A7 rtO.OB 00.02 74.79

Sat. 3 200 625 S 1 2M5 •’


952 626 5 1 235U S A7o\ 02S./t 1 0201 620 9 1.2068
1

60
60 3 353 631 8 1 2671 2 985 629 3 1.2409
70 3 442 638 3 2696 3.068 636 0 .2530 2 605 631 3 1 .2255
80 3 620 C44 7 .2814 3 149 2661 2.576 638 3 1 .2386 2 166 633 8 1 2140
90 3.614 660 9 .2928 3.227 649 0 .2778 2 645 645 0 2510 2.228 640 9 2272

100 3 608 657.0 1 3038 3 304 655 2 1.2891 2 712 651 6 1 . 2628 2 288 647.8 1 2396
no 3 780 663.0 .3144 3 380 601 3 .2999 2.778 658 0 .2741 2 347 654 5 .2515
120 668 0 3247 3 4.54 .3104 2 842 664 2 .2850 2.404 661.1 .2628
130 3 042 674.7 3347 3.527 673 3 .3206 2 905 670,4 .2956 2 460 667 4 273H
140 4.021 680.5 .3444 KKjljy 679.2 .3305 2.907 676.5 13058 2.515 673 7 .2843

160 4.100 ISdKJ 1.3539 3.672 685 0 1.3401 3 020 082 5 1.3157 2 569 679 9 1 . 2945
160 4.178 692 0 .3633 3 743 690 8 .3495 3 08! 688 4 .3254 2 622
> 686 0 3045
170 4.255 697.7 .3724 3.813 .3588 3 149 694 3 .3348 2 675 692.0 3141
180 4 332 703.4 .3813 3.883 702 3 .3678 3.209 700 2 .3441 2 727 698 0 .3236
100 709.0 .3901 3.952 .3767 3.268 706.0 .3531 2.779 704.0 .3328

4.484 714,7 1.3988 4.021 713.7 1 3854 3.326 711.8 1,3620 2 830 709 9 1.3418
4.660 720.4 .4073 4.090 719.4 .3940 3 385 717.6 .3707 2 880 716 8 3507
4,635 .4167 4.158 725.1 .4024 3.442 723 4 3793 2 931 721 6 3594
4.710 731.7 .4230 4 226 730.8 .4108 3 500 729.2 .3877 2 981 727 5 .3679 !

4.785 737.3 .4321 4.294 736.5 .4190 3 557 734.9 .3960 3 030 733 3 .3763

4 850 743.0 1.4401 4.361 742.2 1.4271 3 614 740 7 1.4042 3 080 739 2 1.3846
4 933 748 7 .4481 4 428 747.9 .4350 3 671 746 5 .4123 3.129 I
746 0 3928
5.081 .4637 4.662 759 4 .4507 3.78^1758 0 4281 3 227 1
756 7 .4088
5 228 771.5 4789 770 8 3 895 769.6 4435 3.323 768 3 4243
1

524
Absolute Pressure in Ib/in.* (Saturation Temperatures in italics)

1 261 635 3 1.1621


1 302 643 5 .1764
1 342 651 3 .1898
1 380 658 8 .2025

1.416 666 1 1.2145


1 452 673 1 2250
1.487 680 0 .2369
1.521 686 7 .2475
1.554 693.3 .2577

1 587 699 8 1,2677


1.619 706 2 2773
1.651 712 6 2867, !

1 683 718 9 2950


1 714 725 1 .3040

1 745 731 3 1.3137


1 775 737 5 ,3224
1 805 743 6 .3308
1 835 749 8 .3392
1 865 765.9 3474

1 895 762 0 1 3554


1 954 774 1 3712
2 012 786 3 .3866
2 069 798 4 4016
2 126 810 6 .4163
Mei

of

Properties

Alls

Table

526
Co.
SSS3
NWNri r:3s;ss ss ssssss
ooooo ooooo oooo oooo ooooo oS
W»HOJOOt» t«0<0l0>0 vHf-4?-ii-4^
•HfHOOO oo Electric

General

DdO C^KIOOO
COQO^QI» WiOOO^ of
iST^
W c5 N c5 ^ 0>00h«-h«»
permission

^t;QOa» aoOQOep <©»ft


OONCSO
0i0o2>»0
tOCOOOCO
'^d.nSa dM by

Reprinted

Or«U3Q0i<
SSSSJ, d«I^O
.. cortco-^^
<N^<0_
38S
-
SiSI
row wrororow •^S|o
ggggg g oo ooooo ooc sssss ss
Vapor.’*

Mercury

ooKr-cod ^fHIAWO o>o»wd


2SS t^wwd SSoS^ ss&ss
^o woco
Wdo SS(Oi-4FHt>.
lOO^ SSSSIS wr»o CO w t*eo of
^dd
WWW sssfssi
woq»^ Kooaiflb
sssisiii WWW w w w ww SSSSS SS
^ fH fH fH iH lH *H ^ ^ FH ^ iM fH »H F^ fH ^ 1H lH ^
iH fH F.1 f-4 1-1

Properties

wdco'^t^
O^dO-^
O5fG0c«Ni
TtjOi-igOO
d»>-roW'^
OiOOC
5rHC2»i ssss 8 1^ opo w
d W»Hh. D
.hoowI'- Thwwdd I^OC QOt«OU '^doot^ S ul

rH fH 1^ 1—4 1-4 ^ f—t fH ^ fH tH ^ fH 1-H fM ?-t i^ fH i-4 fH i^


snssn
fH fH ^ fH ^
Thermodynamic

dOlOOOO 3iOTf«OOC
Sd S532
lOOQO'^ 5i-idiHC *'

r-nOO W^- 9'^t»OC *0 r^w coi^ww isip ooo


dcSSdd SdSdd SwSc F-Jdd
5e8SS5 dd
Sheldon,

A.

F^N.^oq»o oqqrooro -^qqww dociw d^^d qwwi-;d qq


d wq;5<^ q o CP SSSS SS Lucian

ooSSwQor S8S5 r^CiwiC ww


from

Data

*
sssss §3§§s ssgss |g|g Ills Hill IS
iHiH

627
INDEX

absolute temperature, 27 approach, temperature, 495


absolute thermodynamic tempc^raturc, area:
115 annular, 332
absolute velocity, 327 passage, turbine stage, 331
absorber, refrigeration, 473, 475 factor, 333
absorption refrigeration, 472 argon, properties, 502, 503
acceleration force on a fluid stream, 324 atoms and molecules, 44, 54
acetylene, properties, 502-504 availability,. 124
adiabatic compression, 382, 385 loss of, 127, 376
adiabatic machine, efficiency, 367 Avogadro’s law, 171
adiabatic process, 32 axial flow compressor, 383, 384
mixing, 197 axial flow turbine, 330
reversible, 103
back work, 233
calculations, 208 Bray ton cycle or gas turbine plant,
in
with perfect gas, 179
283
on is plot, 123
in Rankine and Carnot cycles, 235
adiabatic saturation, 484
Barnard, W. N., 238
admission, 363, 374
base temperature, combustion process,
air:
417
dry, 480
benzene, properties, 504, 505
excess, 412, 414
binary vapor cycle, 259
expansion cooling plant, 468
Binder, R. C., 321
properties, 502, 503, 507, 508, 512,
boiler, steam, 215
513, back cover chart E2 bomb calorimeter, 418
table, 512, 513
bore, 361
air-standard cycles, 267
brake thermal efiiciency, 270
air-water vapor mixtures, 198, 480
Brayton cycle, 277
American Institute of Physic^s, 40
effect of pressure ratio, 284
American Society of Heatini; and Ven- effect of regenerator, 286
tilating Engineers, 168, 497 efficiency, 278
American Society of Mechanical En- efficiency and capacity characteris-
gineers, 321
tics, 284
American Society of Refrigerating En- examples, 283
gineers, 162, 168, 479, 497 reversed, 467
ammonia tables, 522-525
Bra3don, George B., 277
analysis, ultimate, 405 bucket row, linear, 326
analyzer, 473 bucket velocity, limit on efficiency of
annulus area, 332 turbine stage, 346

529
530 INDEX
butane, properties, 502-505 coefficient:

by-product power, 256 flow meter, 815


discharge, of a noszle, 801
calculations, processes in stationary performance, 458
systems, 203 velocity, of a nozzle, 300
calorimeter: Collins, S. C., 471
bomb, or fuel, 418 combustion:
electric, 163 data, 504, 505, back cover chart E5
separating, 162 energy equation, 416
steam, use of, 166 heat of, 422
throttling, 1^ in air, 408
capacity of a cycle, 215 in oxygen, 407
carbon, 405, 407, 408, 504, 505 incomplete, 408, 412
carbon dioxide, properties, 502, 503, material balance, 407, 418
507, 508, back cover chart El processes, 405
carbon monoxide, properties, 502-504, efficiency, 430
507,508 reaction, see reaction
Carlson, J. R., 350 turbine plant, 446-450
Carnot, N. L. 95S., components of velocity, 327
Carnot cycle, 10^104, 123 compound engine, 377
comparison with Rankine cycle, 232 Compressed Air and Gas Institute, 384,
Carrier, W. H., 485, 498 404
Carrier psychrometric equation, 496 compressed liquid tables, 160
cascade refrigerating plant, 462, 471 compressibility charts, 187, 610, 511
cavitation, 371 compression, 364, 874
centrifugal compressor, 383, 384 distilling plant, 477
centrifugal forceon turbine bucket, 346 -ignition engine, 273, 439
characteristic velocity for a turbine processes, 383, 385
stage, 331 ratio:
charts of properties, 161, Appendix, Diesel cycle, 274
back cover envelope Otto cycle, 271
chemical equilibrium, 427-429 three-stage, 401, 402
chemical process, see reaction two-stage, work analysis, 400
chemical reaction, see reaction work of, ^6
choking flow, 804 compressor, 363, 883, 884
Church, A. H., 404 adiabatic, efficiency of, 889
Clausius, R.: efficiency, 367, 388
inequality of, 121 indicator diagram, 391
Second Law according to, 106 reciprocating, illustration, 390, 397
clearance, 87, 391 work 392
analysis,
and compression, 376 refrigeration, 459,460
364
fluid, types, comparison, 401
pocket for capacity control, 395 condensate, 89, 216
volume, 361 condensation in the steam engine, 376
closed cycle gas turbine plant, 279 condenser, 89, 216
closed beater, 253 refrig^tion, 459, 460
coal, 406 condensing steam power plant, 216
INDEX 531

OQDditilon line, 357; aee multistage state DeLaval nozzle, see converging-diverg-
path ing nozzle
CQDBtant, gas, see gae constant DeLaval turbine, 323
constant-pressure process, calculations, density, 5
205 design characteristicB and operating
constant-pressure reaction, 420 characteristics, 335
constant-temperature (isothermal) pro- detonation, 272
cess, calculations, 206 deviation, enthalpy, 486
constant-volume process, calculations, dew-point, 199, 481
203 of flue gas, example, 410
continuity, equation of, 73 Diesel cycle, 272
control surface, 68 efficiency, 274
control volume, 68 variable specific heat, 275
moving, 330 diffuser, 306, 311, 450, 452
converging nozzle, 297, 301 efficiency, 312
converging-diverging nozzle, 297, 304 discharge coefficient of a nozzle, 301
conversion of units, 506 {table) dissociation, see chemical equilibrium
cooling tower, 494-496 distilling plant, compression, 477
corliss engine, 379 divergence, angle, of nozzle, 304
crank, 362 Dorsey, Noah E., 168
crank-end, 362 double-acting, 361
critical pressure, 142 drip, heater, 253
critical pressure ratio in nozzle flow, 299 drip pump, 253
critical state, 142; data, 503 drop, enthalpy, 331, 349, 351, 355
critical temperature, 142 dry bulb, 481
cross-head, 361 dry ice, 144, 471
cut-off, steam engine, 374
cut-off ratio, Diesel cycle, 274 Eberhardt, J. E., back cover chart E5
cycle, 3 eccentric,373
Carnot, 102-104 economizer, 260

on t-s plot, 123 effectiveness:

heat engine, 92-94 heat exchanger or regenerator, 281,


gas, 265-292
456
refrigeration, 457-479 intercooler, 404
reversible, efficiency of. 111 efficiency, see cycle or apparatus involved

vapor, 214-264 adiabatic, of a compressor, 388


see Bra3rton, Diesel, Ericsson, Otto, brake, 270, 368

Euikine, regei^erative, reheat, combustion process, 430


engine, 226, 367
and Stirling; also fluid involved,
heat engine or cycle, 94
cylinder, 360
indicated, 367
condensation, 376
isothermal, of a compressor, 388
head, 360
mechanical, 368
pump, 229
Dalton’s law, 192 nozzle, 300
dead-point, 373 nozzle-bucket, 330
de-humidifying inrocess, 491-494 stage, of a turbine, 347
2

532 INDEX
efficiency (conVd): equation {coni' d)\

thermal, 94 of state, 137, 170, 172, 183, 187, 193


volumetric, 393 equilibrium;
ejector, 313 between phases, 142
Ellenwood, F. 0., 162, 168, 238 chemical, 427-429
Elston, C. W., 359 diagram for water, 143, 144
end state of actual engine or turbine frozen, 428
process, 227 lack of, 99
energy, 1, 47 Ericsson cycle, 226
equation for chemical reaction, 416 approximation by multistage com
equation, steady flow, 74-80 pressiori and expansion, 290
hypothesis of, 47 Eshbach, 0. W., 502
internal, 50; see internal energy ethane, properties, 502-504
kinetic, 52, 73, 76, 85, 99 evaporator, refrigeration, 459, 460
molecular and atomic, 44, 54 excess air, 412, 414, 415
potential, 53, 73, 85, 99 exhaust, 347, 363, 373-380. 436, 445
due to surface tension, electricity, -gas turbine, 445
and magnetism, 54 steam, 215, 256
units, 49 exit angle, bucket, 327
engine, see expansion cugine, hot-air expansion, 363, 374
engine, heat engine, internal engine, 363
combustion engine, reciprocat- in refrigeration cycle, 466-470
ing engine, steam engine factor Y for flow meter, 317
engine, actual, 226 incomplete, 375
engine efficiency, 226, 367 over-, 306
enthalpy, 57 process in actual steam cycle, 227
deviation, 486 ratio, Diesel cycle, 274
drop, 331, 349, 351, 355 ratio of a nozzle, 304
of mixtures, 197 under-, 306
of a perfect gas, 173 valve in gas (looling, 467, 470
of a pure substance, 146 valve, refrigeration, 459, 460
or reaction, 420, 421; see heating extensive proper ties, 156
value and heat of combustion extraction cyck*, 249
-entropy chart, 162, back cover
charts E3, E5 feed water, 215
entropy, 42, 116, 119 heater, 249
change of, from the Gas Tables, 184 First Law of thermodynamics, 49, 4f>~

of mixture, 197 66
of a perfect gas, 177 Fliegner’s formula for air flow, 321
of a pure substance, 148 flow, 67-91, 293-321
principle of increase of, 129 fluid,364
relation to J* dQ/T, 120 meters, 315-318
equation, see name
of equation nozzle, 315
chemical, 407-409 processes, 67
combustion, 407-409 work, 73
continuity, 73 flue gas, 41
energy, steady flow, 74 fluid rate, turbine, 354
INDEX 533

Foote, W. R., 359 Hawkins, L. A., 4


force and mass, 4 head-end, 362
force on fluid in a curved passage, 325 heat, 30-32, 44
force on turbine bucket, 328 of combustion, 406, 422, 505
force to accelerate a fluid stream, 324 engine cycles, 92
formation, see reaction maximum eflicieiicy of, 111-113
four-stroke cycle, 270 engine, limit of eflicicncy in terms ol

four-valve engine, 379 temperature, 116


free expansion, example, 100 engine, reversed, 104, 457, 458. 467,
Freon- 12, properties, back rover chart 476, 477
E4 exchanger, in Brayton cycle, 280
friction, 90-99, 102 latent, scr latent heat
factor in pipe flow, 90 pump, 105, 457, 476
fuels, 405, 406, 504, 505 rate, 219
fuel-air mixture charts, hack cover of reaction, 422
chart E5 spe(‘ifk*, see specific heat
example of use, 437-439 heater:
fusion, latent heat of, 35, 59, 146 closed, 253
drip, 253
Gaffert, G. A., 238, 359 feed water, 249

gas, 140, 170-191 open 0 ^ contact, 253


compression, 383-404 heating of a substance at constant

(‘oinpressioii cooling cycle, 4G7 pressure*, 138-142

constant, 170 heating value, 423

universal, 171 choice of, for eflicicncy bjisis, 431


cooling, 466 higher and lower, 424

cycles, 265-292 data for fuels, 406, 504


fuels, 406 helium, properties, 502, 503
Heroes turbine, 322
laws, 170
liquefaction, 469-471 Hershey, R. L., back cover chart E5
Hirshfeld, C. F., 238
perfect, 170-191
properties, 170-191, 502-505, 507- horsepower per ton, refrigeration, 463,
back cover charts 464
513,
Tables, 183-186 hot-air engine, 267

entropy change from, 184 Hottel, II. C., 168, 437, back cover
reversible adiabatic process, 184 chart E5
abridged air table, 512, 513 humidifying process, 487, 488
gas turbine plant, 278, 281, 446-450 humidity:

gas-water vapor mixtures, 198-201, chart, back cover chart E6


480-498 ratio, 480
gasoline, 406 relative, 200, 480

generator, absorption refrigerator, 473, relations, 200, 201, 481, 482


475 specific, 200, 480
Gibbs-Dalton law, 193 Hunsaker, J. C., 321

Gibbs theorem, 193 Hunt, F. B., back cover chart El


Griswold, John, 406 hydraulic equation for flow m(4ers, 315
534 INDEX
hydrocarbons, properties for combus- irreversibility:

tion calculations, 442, 604, 505 due to free expansion, example, 100
hydrogen, 405, 408, 502-504, 509 due to friction, example, 96
hypothetical reversible impulse stage, due to temperature difference, exam-
385 ple, 99

hypothetical reversible reaction stage, mechanical, 102


344 thermal, icb
irreversible heat transfer, 127, 128,

ideal cycles, 214 131-133


impulse, 453 effect on refrigerating plant, 464
impulse turbine stage, 333 irreversible processes, 98-102
diagrams, 334 isentropic enthalpy drop, 331, 349, 355

hypothetical reversible, 335 isentropic process, 120, see reversible

reversible, efficiency, 336 adiabatic process

incomplete expansion, 370, 375 isolated system, entropy of, 128

indicator diagram, 17 internal energy of, 129

compressor, 391 isothermal process, 103

reciprocating machine, analysis, 363- calculations, 206


365 compression, 383, 385

intensive properties, 156 reversible, on t-s plot, 123


intercooler effectiveness, 404
intercooling in compressors, 398 jackets, steam engine, 377
intercooling in gas turbine plant, 288 jackets, cooling water, 390, 399
internal combustion engine, approxi- jet propulsion plant, 450-454
mation by cycle, 267 jet pump, 313
internal combustion engine efficiency, jet velocity, 333
439 Jordan, R. C., 479, 498
internal combustion engine working Joule, J. P., experiments on energy con-
fluid, properties, back cover servation, 48
chart £5 Joule’s law, 173
internal combustion power plants, 435 Joule-Kelvin, see Joule-Thomson
internal combustion turbine plant, 446- Joule-Thomson coefficient, 471
450 Joule-Thomson experiment, 147
internal energy, 42, 50-55; see energy,
and substances involved
Karig, H. £., 497
kinetic, 52
Kaye, J., 55, 153, 168, 183, 292, 442,
of a mixture, 197
502, 507, 508, 509, 512, 513
molecular and atomic, 44, 54
Keenan, J. H., 40, 55, 116, 153, 168,
of a perfect gas, 173
183, 292, 308, 321, 442, 502, 507,
potential, 53
508, 509, 512, 513, 615, 517, 519,
of a pure substance, 55, 148
521
of reaction, 419, 421 292
Keller, C.,
International Critical Tables, 168 Kelvin, Lord (Sir William Thomson),
inversion, Joule-Thomson coefficient, 95, 115
471 Kelvin temperature scale, 30, 115
invernon temperature, 471 kerosine, 406
INDEX 535

Keyes, F. G., 55, 153, 168, 515, 517, nuxture


519, 521 internal energy, 197
kinetic energy, 52, 75, 76, 85, 99, 301, liquid-vapor, use of tables, 156-159
311, 313, 331, 347 perfect gas, p, 7, T relations, 193-^
knocking, 272 196
Knowltpn, P. H., 359 properties, 192
Kraft, H., 304 saturated, 198
specific heats, 197
latent heat, 35, 59
moisture fraction, 158
of fusion, 35, 146
mol, 171
of vaporization, 35, 147, 505
mol fraction, 195
Laval, see De Laval
mol volume, 172
leaving loss, turbine, 347
molecular weights, doto, 502-504
Lieb, E. F., back cover chart E3
molecules, 2, 44, 54, 194
Lichty, L. C., 292
Mollier chart, 162, for steam, back
limited-pressure cycle, 275
cover chart, E3
Lindsay, D. C., 485
Morse, F. T., 238
liquefaction of gases, 469-471
Moss, S. A., 4, 359
liquid, compressed, tables, 160
multistage compression, 396-401, 462
liquid-vapor mixtures, use of tables,
multistage state path, 348
156 see substance involved
;
multistage turbines, 346
Lorenzi, dc, 0., 406
Napier’s formula for steam flow, 320
Macintire, H. 479 J., nitrogen, properties, 502, 503, 507, 508
Mackey, C. 0., 162 non-condensing steam power plant, 215
makeup water, 215 nozzles, 293-321
material balance, steady flow, 72, 75 nozzle:
material balance, combustion, 407-411 analysis, assumptions, 293
mean effective pressure, 20, 365, 439 angle, turbine, 327
mercury, properties, 526, 527 coefficients, 300
mercury vapor cycle, 260-262 computations, 308
Messersmith, C. W., 40 converging, 301
metastable state, 313 pressure variation in, 302
meter, flow, 315 converging-diverging, 304
methane, properties, 502-504 pressure variation in, 305
mixed cycle, 275 efficiency, 300
mixing process, adiabatic, 197 flow meter, 315
mixing process, humidity calculation, reversible, 296
488-490 cross-sectional area, 296
mixture: nozzle-bucket efficiency, 330
enthalpy, 197
air-water vapor, 486
octane, 407, 409, 425, 442, back cover
entropy, 197
chart E5
equivalent molecular weight, 194
oil fuel, 406
gas constant, 194
one-dimensional flow, 324
gaseous, 192-202
gases and water vapor, 198-201, 480- open cycle, 447
498 open heater, 253
536 INDEX
operatiDg characteristics and design potential energy, 53, 73, 85, 99
characteristics, 335 due to surface tension, electricity,
orifice flow meter, 317 and magnetism, 54
Orsat gas analysis, 412 pre-mixed fuel-air charge, 435, 440
Otto cycle,268 pressure, 5, 12
efficiency, 270 mean effective, 20, 365, 439
variable specific heat, 271 partial, 193, 195
over-expansion, 306 relative, Gas Tables, 185
oxygen, properties, 502, 503, 507, 508 pressure-enthalpy chart, Freon-12, back
cover chart E4
jw" = constant, irreversible process, Priester, G. B., 479, 498
calculations, 212 process, 3; see type of process
Palmatier, E. P., 497 products analysis, combustion, 411, 504
part by mass, 195 products of combustion, properties,
part by volume, 195 442, back cover chart E5
partial pressure, 193, 195 products temperature, combustion,
partial volume, 194, 195 425-429
passage area, turbine stage, 331 propane, 504, 505
path function, 42 properties, 3, 41-45; see substance in-
perfect gas, 170; see gas volved
perfect gas mixtures, 193; see mixture charts of, 161, 442, Appendix, back
perpetual motion machine of first kind, cover envelope
96 gases, 502, 503, 507-511
pKirpetual motion machine of second property relations, general, pure sub-
kind, 95 stance, 150-152
phases, 138, 140, 416 psychrometer, 483
phase diagram, pressure- volume, 145 psychrometric chart, 485, back cover
phase diagram, temperature-entropy, chart E6
150 psychrometrics, 480
pipe friction, 79, 90 pump:
pipe lines, actual, in steam cycle, 229 actual, in steam cycle, 229
piston, 11, 17, 360 efficiency, 229, 367
piston displacement, 361, 368, 369, 391, liquid, 371
439 work in Rankine cycle, 220
piston ring, 360 pure substance, 54-56, 137-153, 192
piston rod, 360
Planck, Max, 95, 116
quality, 157-159
point functions and path functions, 42
of steam, measurement^ 162-167
polytropic compression, 383, 385
polytropic process, perfect gas, 181
polytropic process, reversible, calcular radial flow turbine, 330
lions, 210 radiation shield for thermometer, 484
polytropic specific heat, 181 Rankine cycle, 214-238
porous plug, 147 comparison with Carnot cycle, 232-
ports, 362, 379, 380 236
positive displacement machines, 360, efficiency, 219, 220
383, 384, 403 influence of exhaust pressure, 224
INDEX 537

Rankine cycle (cmVd): regenerative vapor (steam) cycle, 245-


influence of initial temperature, 223 255
influence of pressure, 222 comparison with Carnot cycle, 248
initial pressure limited by tempera- energy and flow analysis, 251, 252
ture, 241, 242 heat balance, 251, 252
Rankine temperature, 29, 115 practical process, 249
ratio of specific heats, see specific heat reduction in heat consumption by
ratio use of, 254, 255
reaction (chemical): regenerator in Brayton cycle (gM tur-
constant pressure, 420-422 bine plant), 280
constant volume, 416-420 reheat cycle, gas, 288-290, 450
energy equation, 416 reheat cycle, vapor, 239-245
heat of, 422 comparison with Rankine cycle, 243-
internal energy of, 419, 421 245
enthalpy of, 420, 421 efficiency and steam rate, 241
material balance, 407-411 reheat effect in turbines, 347-350
reaction stage (turbine), 342-346 reheat factor, 349
hypothetical reversible, 344 reheater, 240
efficiency, 345 relative humidity, 200, 480
velocity diagram, 343 relative pressure, Gas Tables, 185
reciprocating engine, steady flow anal- relative vtilocity, 327-330, 333, 335
ysis, 366 relative volume. Gas Tables, 185
reciprocating engine, work analysis, release, 374
365 resuperheating, 239
reciprocating machine, 360-372 reversed heat engine, 104, 457
adiabatic, 366 reversibility, 97-102, 128, 129
efficiency, 366 reversible adiabatic prociess, 120; see
flow capacity, 368 cycle or apparatus involved
rectifying column, 473 calculations, 208
reduced properties, 188 Gas Tables, 184
Reese, H. R., 359 perfect gas, 179
reflux, 473 reversible cycles, 111-113; see name of
refrigerant, desirable properties, 460 cycle
refrigeration, 457-479 reversible nozzle, 296-300
rcfri^reration cycle: reversible polytropic process, calcula-
absorption, 472-476 tions, 210
capacity, 462, 463 reversible system, 102
effect of irreversible heat transfer, Kightmirc, B. G., 321
464-466 Rohsenow, W. M., 498
468
gas, 467,
power consumption, 463, 464 Salisbury, J. K., 359
standard conditions, 463 saturated air, 481
vapor compression, 458-462 saturated mixture, 198
refrigeration ton, 462 saturation:
refrigerator,457 adiabatic, 484
refuse, combustion product, 412 curve, 141
regenerative gas cycle, 266 percent, 480
538 INDEX
saturation (cont*d): stage efficiency, turbine, 347
states, staged compression, gas turbine plant,
in tables, 164 288
Scatchard, G., 479 staged feed heating, 249
Schiller station, 261 effect of number of stages on heat
Sears, F. W., 40, 153 consumption, 254, 255
Second Law, 92-110, 152, 221, 228, 239, stagnation enthalpy, 294
428, 473 stagnation pressure and temperature,
applications, 111-136 295
Clausius form of, 106 standard refrigeration cycle conditions,
Kelvin-Planck form of, 95 463
problems, examples, 130 state, equation of, see equation of state
separation, flow, in divergmg duct, 307, static pressure and temperature, 295
312 stationary system, 12
Servel refrigerator, 474 applications of the First Law, 59
shaft work, 71 process calculations, 203-213
Sheldon, L. A., 527 st^y flow, 70
shock loss in turbine, 335 energy equation, 74
shock, normal, 306 equations, applications, 76-80
shock, oblique, 306 steam chart, back cover chart E3
single-acting, 361 steam engine, 373-382
sink, 94 comparison of real and idealized indi-
Skrotzki, B. G. A., 238 cator diagrams, 375
slide valve, 373 steam rate, 219, 355
Smith, C. W., 359 actual, 226
Smithsonian Physical Tables, 168 steam tables, 154-169, 614-621
solution, heat of, 473 Stirlmg cycle, 265
sound, velocity of, 299, 303 Stodola, Am ^9
spark-ignition engine, 268, 435, 436, stroke, 360
440 subcooled vapor, 314
specific heat, 32-35, 38, 56, 58; data, sublimation, 35, 144
502, 503, 505, 507, 509 8ubHM>nic, 299, 306, 312
at constant pressure, 58 substance, pure, see pure substance
at constant volume, 56 sulfur, 36, 405, 408, 504, 505
molal, 177 sulfur dioxide, 502, 503
of mixtures, 197 supercharged engine, 445, 446
of perfect gas, 173 superheat, 159
polytropic, 181 effecton Rankine cycle, 223, 236, 242
variable, 183 effect on steam engine, 377
specific heat ratio of a perfect gas, 176 effect on steam turbine, 237, 241,
specific heat ratio data, 502, 503, 507, 355,356
509 superheated vapor, 142
specific humidity, 200, 480 tables, 159; see substance inocived
Spiers, H. M., 169 supersaturation, 313-315
souroe, 94 super-sonic, 299, 306, 312
stage, compressor, 396 surroundings, temperature of, 125
stage, turbine, 331 symbols, 500, 501
INDEX 539

system, 2 turbine {canfd):


reversible, 102 multistage, 346
stationary, 12; see stationary system impulse, illustration, 351
reaction, illustration, 352
operating characteristics, 350
tail pipe, 451
reheat factor, 349
taps, pressure, 317
stage, 331
Taylor, C. F., 292, back cover chart E5
types, comparison, 350-354
Taylor, E. S., 292, back cover chart E5
turbo-prop engine, 451, 454
temperature, 24
two-row stage, see velocity-compounded
absolute, 27
stage
thermod 3mamic, 115
two-stage compression;^ work analysis,
constant, see isothermal
400,401
difference in condenser and evapora-
tor, 465 ultimate analysis, 405
of combustion products, 425-429 unavailable energy, 124
scales,26 under-expansion, 306
temperature-entropy diagram: uniflow engine, 379
air, back cover chart E2 units, 3
carbon dioxide, back cover chart El conversion table, 506
temperature-entropy plot, 122-124 U.S. Dept, of Commerce, Bureau of
theoretical air, 407, 409, 414 Standards, 162, 169, 502, 504,
thermal power plants and ideal cycles, 506, 522
214, 215
valves, 362
thermometer, 25
valve loss, 302
gas, 27
Van de Waals equation, 187; data, 503
thin-plate orifice flow meter, 317
vapor, 140
three-stage compression, 401, 402
vapor cycles, 214-264
throat, nozzle, properties at, 300
actual processes, 225
throttle flow, 252
more efficient cycles, 239
throttling process, 147, 375, 435, 459, working fluid, desirable properties,
467, 470, 471 259
thrust, 454 vapor dome, 142
ton, refrigeration, 462 vapor, subcooled or supersaturated, 314
total pressure and temperature, 295 vaporization, latent heat of, 35, 147,

trap, 253 505


variable flow processes, 80, 83
triple point, 144
velocity:
trunk piston, 361
absolute, 327
turbine, 322-359; see bucket, nozzle,
characteristic, for a turbine stage,
impulse, reaction
331
efficiencies and steam rates, 355, 356
coefficient, bucket, 327
expansion, for refrigeration, 468 coefficient, nozzle, 300
extraction, illustration, 351 components, 327
flow path nomenclature, 326 diagram, tuii>ine, 327
for gas turbine plant, illustration, 353 in a nozzle, 295
internal efficiency, 349 relative, 327
540 INDEX

velocity-compounded stage, 338 Wile, D. D., 497


efficiency, 341 Willans line, 354, 355
velocity diagram, 340 work, 8-23; see process, cycle, or oppa-
venturi meter, 315 ralus involved
volume, partial, 194, 195 back, see back work
volume, relative, Gas Tables, 185 of compression, 386
volume, specific, 5 see substance involved
;
engine, 93
volumetric efficiency, 393 flow, 73
effect of clearance, 394 graphical representation, 13
effect of pressure ratio, 395 indicator diagram, 18-21
Vopat, W. A., 238 lost, 128, 230-232
on a turbine bucket, 329, 330
Warner, C. F., 4C
paddle-wheel. 13, 101
Warren, G. B., 359
pump, 93
water, fecc^, 215
working fluid, desirable properties, va-
water, liquid and vapor in combustion
por cycle, 259, 260
products, 410, 414
refrigeration cycle, 460
water, properties, 138, 142-144, 502,
503, 509, 514-521, back cover Y coefficient, flow meter, 317
chart E3; see steam
Weber, II. C., 188, 510 Zemansky, M. W., 2, 40, 95, 102, 116
wet bulb, 481-486 153, 193

You might also like