Vous êtes sur la page 1sur 124

Engineering Science Prelims

THERMODYNAMICS

Dr P D McFadden
Hilary Term 2008


with minor updates for Hilary 2014 by
Professor Peter Ireland

CONTENTS 1
Contents
1 Introduction 6
1.1 What is Thermodynamics? . . . . . . . . . . . . . . . . . . . . . . . . 6
1.2 Books . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1.3 Syllabus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.4 Example Sheets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
1.5 Lecture Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.6 Learning Outcomes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.7 Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2 Revision 10
2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.2 Atoms and Atomic Mass . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3 Compounds, Molecules and Molecular Mass . . . . . . . . . . . . . . 12
2.4 Avogadros Number and the kmol . . . . . . . . . . . . . . . . . . . . . 12
2.5 Boyles Law and Charles Law . . . . . . . . . . . . . . . . . . . . . . . 13
2.6 State and Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
3 Properties of Solids and Liquids 16
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2 Compressibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
3.2.1 Compressibility of Solids . . . . . . . . . . . . . . . . . . . . . 16
3.2.2 Compressibility of Liquids . . . . . . . . . . . . . . . . . . . . . 18
3.3 Thermal Expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
3.3.1 Thermal Expansion of Solids . . . . . . . . . . . . . . . . . . . 18
3.3.2 Thermal Expansion of Liquids . . . . . . . . . . . . . . . . . . . 19
3.3.3 Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
3.3.4 Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
3.4 Specic Heat Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.4.1 Specic Heat Capacity and Thermal Capacity . . . . . . . . . . 22
2 CONTENTS
3.4.2 Example 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
3.4.3 Example 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.4.4 Example 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.5 Transport Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.5.1 One-Dimensional Conduction . . . . . . . . . . . . . . . . . . . 24
3.5.2 Example 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.5.3 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
3.5.4 Example 7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
3.5.5 Diusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
3.5.6 Example 8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4 Behaviour of Gases 29
4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2 Ideal Gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2.1 Ideal Gas Law . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
4.2.2 Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.2.3 Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.2.4 Specic Heat Capacities . . . . . . . . . . . . . . . . . . . . . . 30
4.2.5 Partial Pressures . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.2.6 Example 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.2.7 Gas Mixtures . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
4.2.8 Example 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.3 Non-Ideal Gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.3.1 Reasons for Non-Ideal Behaviour . . . . . . . . . . . . . . . . . 33
4.3.2 Generalised Compressibility Chart . . . . . . . . . . . . . . . . 33
4.3.3 Example 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.4 Real Gases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.4.1 Phase Changes . . . . . . . . . . . . . . . . . . . . . . . . . . 35
4.4.2 Critical Point . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
4.4.3 p v Diagram . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.4.4 p v T Surface . . . . . . . . . . . . . . . . . . . . . . . . . . 37
4.4.5 Phase Rule . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
5 Behaviour of Steam-Water Mixtures 40
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
CONTENTS 3
5.2 Saturated Region . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.2.1 Saturated Table in HLT . . . . . . . . . . . . . . . . . . . . . . 40
5.2.2 Dryness Fraction . . . . . . . . . . . . . . . . . . . . . . . . . . 41
5.2.3 Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.2.4 Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
5.3 Superheated Steam . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.3.1 Superheated Table in HLT . . . . . . . . . . . . . . . . . . . . . 43
5.3.2 Example 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
5.4 Supercritical Steam Properties . . . . . . . . . . . . . . . . . . . . . . 44
5.5 Other Tables with a Phase Change . . . . . . . . . . . . . . . . . . . . 44
6 First Law for Closed Systems 45
6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.2 Systems, Processes, Paths and Cycles . . . . . . . . . . . . . . . . . 45
6.2.1 Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
6.2.2 Processes and Paths . . . . . . . . . . . . . . . . . . . . . . . 46
6.2.3 Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
6.3 First Law for Closed System . . . . . . . . . . . . . . . . . . . . . . . . 48
6.3.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
6.3.2 Internal Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
6.3.3 Specic Heat Capacity at Constant Volume . . . . . . . . . . . 49
6.3.4 Reversible Work in a Closed System . . . . . . . . . . . . . . . 50
6.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.4.1 Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.4.2 Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
6.4.3 Example 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
6.4.4 Example 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
6.4.5 Example 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
6.5 Polytropic Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
6.5.1 Polytropic Equation . . . . . . . . . . . . . . . . . . . . . . . . 57
6.5.2 Example 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
6.5.3 Eect of Polytropic Index . . . . . . . . . . . . . . . . . . . . . 59
6.5.4 Polytropic Processes in Cycles . . . . . . . . . . . . . . . . . . 60
7 First Law for Open Systems 62
4 CONTENTS
7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
7.2 Steady Flow in Open Systems . . . . . . . . . . . . . . . . . . . . . . 62
7.2.1 Steady Flow Energy Equation . . . . . . . . . . . . . . . . . . . 62
7.2.2 Enthalpy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
7.2.3 Specic Heat at Constant Pressure . . . . . . . . . . . . . . . 65
7.2.4 Relationship Between Specic Heats . . . . . . . . . . . . . . 66
7.2.5 Negligible Terms in the SFEE . . . . . . . . . . . . . . . . . . . 67
7.2.6 Work in a Reversible Steady Flow Process . . . . . . . . . . . 67
7.3 Steady Flow Devices . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
7.3.1 Compressor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
7.3.2 Turbine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
7.3.3 Throttle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
7.3.4 Nozzle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
7.3.5 Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
7.4 Enthalpy and First Law for Closed Systems . . . . . . . . . . . . . . . 73
7.4.1 Closed Systems with Moving Boundaries . . . . . . . . . . . . 73
7.4.2 Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
7.5 SFEE Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.5.1 Example 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
7.5.2 Example 4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
7.5.3 Example 5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
7.5.4 Example 6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
7.6 Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.6.1 Rankine Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
7.6.2 Heat Engines and Heat Pumps . . . . . . . . . . . . . . . . . . 81
8 Combustion 83
8.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
8.2 Chemical Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
8.2.1 Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
8.2.2 Combustion in Air . . . . . . . . . . . . . . . . . . . . . . . . . 84
8.2.3 Hydrocarbon Fuels . . . . . . . . . . . . . . . . . . . . . . . . . 86
8.2.4 Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
8.2.5 Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
CONTENTS 5
8.3 Energy of Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
8.3.1 Molar Specic Heats . . . . . . . . . . . . . . . . . . . . . . . . 89
8.3.2 Open Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
8.3.3 Gross and Net Caloric Values . . . . . . . . . . . . . . . . . . 93
8.3.4 Closed Systems . . . . . . . . . . . . . . . . . . . . . . . . . . 93
8.3.5 Example 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
9 Internal Combustion Engines 97
9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
9.2 Idealised Cycles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
9.2.1 Air Standard Cycles . . . . . . . . . . . . . . . . . . . . . . . . 97
9.2.2 Otto Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
9.2.3 Diesel Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
9.3 Real Internal Combustion Engines . . . . . . . . . . . . . . . . . . . . 102
9.3.1 Spark-Ignition Engine . . . . . . . . . . . . . . . . . . . . . . . 102
9.3.2 Compression-Ignition Engine . . . . . . . . . . . . . . . . . . . 105
9.3.3 Swept Volume . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
9.3.4 Mean Eective Pressure . . . . . . . . . . . . . . . . . . . . . . 108
9.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
9.4.1 Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
9.4.2 Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
10 Gas Turbine Engines 112
10.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
10.2 Idealised Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
10.2.1 Air Standard Cycles . . . . . . . . . . . . . . . . . . . . . . . . 112
10.2.2 Joule Cycle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 112
10.3 Real Gas Turbine Engines . . . . . . . . . . . . . . . . . . . . . . . . . 115
10.4 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
10.4.1 Example 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
10.4.2 Example 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117
6 TOPIC 1. INTRODUCTION
Topic 1
Introduction
1.1 What is Thermodynamics?
Thermodynamics is about heat and work, and their inter-conversion. Although you
have not yet studied this subject formally, you will already be familiar with many
examples, such as the internal combustion engine, which converts into work some
of the heat generated by burning liquid hydrocarbon fossil fuel.
Heat is Energy in Transport. Objects do not contain heat!
Internal Energy causes the atoms or molecules of a substance to move faster.
In a gas or liquid they move freely along random paths, and in a solid they
vibrate about their mean position. In both cases the motion is random.
Work is done when a force moves its point of application through a distance.
This movement is not random but directed.
It is this ability to convert the random motion associated with internal energy into
the directed motion needed to do work that makes thermodynamics so important.
It was thermodynamics, initially in the form of steam power, that rst drove the
industrial revolution. Later developments, such as the internal combustion engine
and the jet engine, have had similarly massive impacts on our daily lives, and have
given us the highly technological lifestyle we enjoy today.
1.2 Books
Favourites are
WHALLEY Basic Engineering Thermodynamics (OUP). Well suited to the
course.
ROGERS AND MAYHEW Engineering Thermodynamics Work and Heat Transfer
(Longman, 4th edition). The most complete. Beware opposite sign convention
1.3. SYLLABUS 7
for work.
EASTOP AND MCCONKEY Applied Thermodynamics for Engineering
Technologists (Longman, 5th edition). An easy introduction. Beware opposite
sign convention for work.
Other possibilities include
TABOR Gases, Liquids and Solids (CUP, 3rd edition).
SPALDING AND COLE Engineering Thermodynamics (Arnold). An easy
introduction.
SEARS AND SALINGER Thermodynamics Kinetic Theory and Statistical
Thermodynamics (Addison-Wesley). Good for kinetic theory.
WARK Thermodynamics (McGraw-Hill). Beware mixture of SI and imperial
units.
STONE Introduction to Internal Combustion Engines (Palgrave, 3rd edition).
Excellent.
FENN Engines Energy and Entropy (Freeman). Intended for non-engineers.
MORAN, SHAPIRO, MUNSON AND DEWITT Introduction to Thermal Systems
Engineering (Wiley, 2003).
SONNTAG, BORGNAKKE AND VAN WYLEN Fundamentals of Thermodynamics
(Wiley, 6th edition).
MORAN AND SHAPIRO Fundamentals of Engineering Thermodynamics (Wiley,
5th edition, SI units, 2006).
1.3 Syllabus
The expanded syllabus for this lecture series, shown in the Course Handbook, is
Thermodynamic Properties of Matter and Conduction
Compressibility, temperature coecient of expansion, specic heat,
conductivity, viscosity, diusion. One-dimensional heat conduction, use of heat
transfer coecients. Properties of ideal gases. Properties of mixtures. Use of
tabulated data for steam and other uids involving liquid and vapour phases.
Properties of real gases. Use of generalised compressibility chart.
8 TOPIC 1. INTRODUCTION
First Law of Thermodynamics and Combustion
Basic concepts and terminology of thermodynamics. Heat, work and the First
law. Denition of internal energy. Applications and examples. First law applied
to open systems. Denition of enthalpy. First Law based combustion
calculations. Applications and examples.
The twelve lectures will take us through
Introduction.
Revision of atoms and molecules, atomic and molecular mass, Avogadros
number and the kmol, Boyles and Charles laws.
Compressibility, thermal expansion, specic heats, viscosity, conduction and
diusion in solids and liquids.
Denitions of system, state, properties, processes, cycles, work and shaft
work.
Behaviour of ideal, non-ideal and real gases, phase rule, p - v - T diagrams,
water-steam properties.
First law of thermodynamics for closed systems, internal energy, polytropic
processes.
First law of thermodynamics for open systems, steady ow energy equation,
enthalpy.
Combustion.
Applications including Rankine, Otto, Diesel and Joule cycles.
1.4 Example Sheets
There are three example sheets, one for each four lectures on average, entitled
1P4E Thermodynamic Properties of Materials
1P4F First Law of Thermodynamics
1P4G First Law of Thermodynamics and Combustion
1.5. LECTURE NOTES 9
1.5 Lecture Notes
A colour version of the lecture notes, in PDF format with hyperlinks, will be placed
on WebLearn near the end of the lecture series. Students with special needs are
welcome to raise these with the lecturer.
1.6 Learning Outcomes
After the completion of this course and the accompanying example sheets, you
should be able
to analyse simple problems involving thermodynamic properties such as the
compressibility, specic heat, thermal conductivity, viscosity, the diusion
coecient, and the coecient of thermal expansion
to determine thermodynamic properties using the ideal gas law, generalised
compressibility chart, and tabulated data for steam and other uids
to apply the First Law of thermodynamics to the analysis of closed systems
to apply the steady ow energy equation to the analysis of open systems
to analyse simple combustion reactions
to apply the above to the analysis of simple thermodynamic machines.
1.7 Acknowledgements
These lecture notes and the associated example sheets were originally prepared
by Dr P D McFadden, drawing on material prepared in the past for this and related
courses by Dr David Kenning, Dr Colin Wood, Professor Richard Stone and the
late Dr Peter Whalley. Their contributions are most gratefully acknowledged. Minor
updates have been made for Hilary term 2010.
10 TOPIC 2. REVISION
Topic 2
Revision
2.1 Introduction
In this chapter we will review some of the basic chemistry associated with atoms,
molecules, atomic and molecular mass, the mole concept, and the gas laws. We
will also introduce the concepts of the thermodynamic state and thermodynamic
properties.
2.2 Atoms and Atomic Mass
Matter consists of atoms. In simple terms, an atom can be thought of as a nucleus,
containing protons and neutrons, orbited by electrons. Each atom has N
p
protons,
each with one positive charge, N
n
neutrons, with no charge, and N
e
electrons,
each with one negative charge. The protons and neutrons are relatively heavy
particles, of approximately equal mass. The electrons are much lighter. In practice
it is found that
N
n
N
p
for N
p
= 1
There is no net charge on an atom, so
N
p
= N
e
= Z
where Z is the atomic number . The atomic number denes the chemical species.
Also
N
n
+ N
p
A
where A is the atomic mass. The proper denition of atomic mass is
atomic mass =
mass of 1 atom
mass of 1 atom of carbon 12
12
In physics or chemistry, the units or dimensions of atomic mass are given as
atomic mass units (amu). In engineering it is more appropriate for us to use units
of kg kmol
1
. The masses of the fundamental particles are
2.2. ATOMS AND ATOMIC MASS 11
particle mass (amu)
proton 1.007276
neutron 1.008665
electron 0.000549
Note that carbon, like many other substances, exists in nature in a number of
isotopes. Isotopes have the same Z, so are the same chemical species, but have
dierent A, due to diering numbers of neutrons. The dierent isotopes of a
species X can be written as X
A
Z
. For example, the dierent isotopes of carbon can
be written as
isotope protons neutrons
C
12
6
6 6
C
13
6
6 7
C
14
6
6 8
The atomic numbers and masses for the rst ten elements of the periodic table are
shown in table 2.1.
Table 2.1: Atomic numbers and masses
species symbol Z A
exact
A
approx
hydrogen H 1 1.008 1
helium He 2 4.003 4
lithium Li 3 6.939
beryllium Be 4 9.012
boron B 5 10.811
carbon C 6 12.011 12
nitrogen N 7 14.007 14
oxygen O 8 15.999 16
uorine F 9 18.998
neon Ne 10 21.183
H, C, N and O are the species we are most likely to deal with this year in
thermodynamics, and you can see that the A
approx
values above are good enough
for engineering calculations.
12 TOPIC 2. REVISION
2.3 Compounds, Molecules and Molecular Mass
Elements consist of atoms, either individually or as groups of atoms called
molecules. Compounds consist of dierent atoms and these can only exist as
molecules. Common examples of molecules are the diatomic gases such as H
2
,
O
2
and N
2
. These gases normally exist as molecules, not as atoms of H, O or N. In
contrast, the gases He and Ne exist as atoms. Examples of compounds include
water H
2
O, carbon dioxide CO
2
, and carbon monoxide CO.
The molecular mass M is dened as
molecular mass =
mass of 1 molecule
mass of 1 atom of carbon 12
12
Molecular masses can be found by addition of the atomic masses, as illustrated in
table 2.2.
Table 2.2: Some common molecular masses
species molecular mass kgkmol
1
O
2
15.999 + 15.999 = 31.998 32
N
2
14.007 + 14.007 = 28.014 28
H
2
1.008 + 1.008 = 2.016 2
H
2
O 1.008 + 1.008 + 15.999 = 18.015 18
CO
2
12.011 + 15.999 + 15.999 = 44.009 44
CO 12.011 + 15.999 = 28.010 28
The appropriate engineering units for molecular mass are kg kmol
1
, as for atomic
mass. Atomic (or molecular) mass is also referred to as atomic (or molecular)
weight, or relative molar mass.
2.4 Avogadros Number and the kmol
Suppose a substance has an atomic or molecular mass of M. If we take M kg of
that substance, for example 18.015 kg of H
2
O or 4.003 kg of He, then how many
atoms or molecules does this mass contain? From the denition of atomic or
molecular mass, it must be the same number as the number of C
12
atoms in 12 kg.
This number is known as Avogadros number N, given in HLT on page 42. This
2.5. BOYLES LAW AND CHARLES LAW 13
number of atoms or molecules is referred to as one kilomole or more simply one
kmol , so that
N = 6.022 10
26
kmol
1
Thus 1 kmol of N
2
is 28 kg, 1 kmol of CO
2
is 44 kg, and so on. In chemistry you
may have used gram mole or gmol, with an Avogadros number a factor of 10
3
smaller. Watch out! Use kmol from now on! Note also that some texts use the term
mol , kgmol or kgmole.
A mass of m kg of a substance can be described as m kg or as n kmol where
n =
m
M
where M is the atomic or molar mass with units of kgkmol
1
. For example, if we
have 100 kg of O
2
then m = 100 kg and M = 32 kg kmol
1
giving
n =
m
M
=
100
32
= 3.125 kmol
2.5 Boyles Law and Charles Law
Gases at relatively high temperature and low pressure behave predictably.
Boyle showed that at constant temperature
pressure volume = constant or pV = constant
Charles showed that at constant volume
pressure
temperature
= constant or
p
T
= constant
provided that the temperature is measured on the absolute scale.
The combination of Boyles and Charles laws gives the ideal gas law
pV
T
= constant
which we shall come back to later in section 4.2.1.
14 TOPIC 2. REVISION
solid liquid gas
Figure 2.1: Solid, liquid and gas
2.6 State and Properties
The condition or state of a thermodynamic system is identied by a number of
observable characteristics or properties of state, typically pressure, temperature
and volume. We shall consider later how many properties are required to dene
fully the state of the system. An equilibrium state is one in which there is no
detectable change in the properties with time.
Initially we shall consider only systems containing pure substances which are
homogeneous or uniform in composition, and involve no reactions, so are
unchanging in chemical state. Mixtures are allowed, provided the ratio of the
components is the same throughout the system.
As illustrated in gure 2.1, the state of a pure substance is rst divided into three
basic categories
Solids Molecules are closely packed in a static array, perhaps vibrating,
which displays long-range molecular order.
Liquids Molecules are less closely packed, and are free to move but display a
short-range molecular order, restrained by lack of space and by mutual
attraction or repulsion.
Gases Molecules are widely separated and randomly arranged, move
independently and randomly, and collide with exchange of momentum.
but other states can also be important. A phase is a region of uniform composition
and state. For the present we shall only consider the bulk phases solid, liquid and
gas. Phases in equilibrium with each other are said to be mutually saturated. In
these rst lectures, we will occasionally refer to molecular motion and interaction in
order to understand and explain properties. Later we shall see that
thermodynamics has no need to consider matter on a molecular scale.
2.6. STATE AND PROPERTIES 15
Note that uid describes a gas or liquid.
Properties are of two types.
An extensive property property is one which depends on the mass of the
substance, for example the volume V m
3
, and the internal energy U kJ.
An intensive property property is one which does not depend on the mass of
the substance, for example the temperature T K (or
o
C), the pressure p Nm
2
(or Pa), the specic volume v m
3
kg
1
, and the specic internal energy u
kJ kg
1
.
By convention, we will try to represent extensive properties by upper case letters,
such as V , U, and intensive properties by lower case letters, such as p, v , u. The
exceptions to this convention are the temperature T and the specic gas constant
R.
16 TOPIC 3. PROPERTIES OF SOLIDS AND LIQUIDS
Topic 3
Properties of Solids and Liquids
3.1 Introduction
In this chapter we will look at various properties of solids and liquids, including
compressibility, thermal expansion, specic heat capacity, conductivity, viscosity
and diusion. These properties are important for understanding the
thermodynamic behaviour of solids and liquids.
3.2 Compressibility
3.2.1 Compressibility of Solids
Stresses and strains in elastic solids in three dimensions are related by Youngs
modulus E and Poissons ratio by the equations

x
=
1
E
(
x

y

z
)

y
=
1
E
(
y

z

x
)

z
=
1
E
(
z

x

y
)
In thermodynamics, the relevant stress is the pressure p. Pressure acts uniformly
in all directions, and is a compressive stress, so when we convert between
pressure and stress we must remember that = p.
Consider a solid subjected to a small change in pressure p. The pressure acts
uniformly in all directions so

x
=
y
=
z
= p
giving equal strains of

x
=
y
=
z
=
p
E
(1 2)
3.2. COMPRESSIBILITY 17
If we imagine a cube, initially of length L each side and volume V = L
3
, then the
volume V

of the cube after the change in pressure can be related to the change
V in the volume by
V

= V + V = L(1 +
x
) L(1 +
y
) L(1 +
z
) = L
3
_
1
p
E
(1 2)
_
3
which can be simplied using the binomial expansion to
V 3V
p
E
(1 2)
Taking the limit as p 0, we can write the bulk modulus K as
K = V
dp
dV
from which we see that
K =
E
3(1 2)
(3.1)
Alternatively, the reciprocal of the coecient of compressibility is sometimes
useful
=
1
V
dV
dp
=
1
K
(3.2)
Table 3.1 lists values of E, and K for some common solids. Try some of these
values in equation 3.1. Agreement is not brilliant, is it? Rounding errors,
particularly on , would account for the discrepancies. Values of E, and K for
other solids are given in HLT on pages 4748.
Table 3.1: Comparison of solid properties
substance Youngs modulus E Poissons ratio bulk modulus K
(Pa) (Pa)
aluminium 70 10
9
0.34 75 10
9
copper 130 10
9
0.34 138 10
9
lead 16 10
9
0.44 46 10
9
mild steel 210 10
9
0.30 170 10
9
glass 74 10
9
0.23
mercury 25 10
9
water 2.3 10
9
18 TOPIC 3. PROPERTIES OF SOLIDS AND LIQUIDS
What about the eect of temperature? If the temperature were to increase, most
substances would expand and that would confuse the discussion of
compressibility. In elasticity we neglect to specify that T should remain constant,
but in thermodynamics we have to be more careful. Therefore the denitions are
always written as partial derivatives with a subscript to remind us that T must be
constant. Also, we often consider unit mass of the substance and use specic
volume v instead of absolute volume V .
So the bulk modulus can be dened as
K = v
_
p
v
_
T
and the coecient of compressibility as
=
1
v
_
v
p
_
T
3.2.2 Compressibility of Liquids
Unlike solids, there is no relationship between compressibility and linear elasticity,
as E and do not exist for liquids, as shown in table 3.1. We simply apply the
above denitions of bulk modulus or coecient of compressibility directly. More
data for liquids may be found in HLT page 53.
3.3 Thermal Expansion
3.3.1 Thermal Expansion of Solids
The coecient of linear expansion, , is written in terms of the length L by
=
1
L
dL
dT
so that the change in length L due to a temperature change T is
L = LT
We can also write the coecient of cubical expansion, , in terms of the volume V
by
=
1
V
dV
dT
3.3. THERMAL EXPANSION 19
so that the change in volume V due to a temperature change T is
V = V T
How are and related for a solid? Consider a cuboid of volume V with sides of
length L
x
, L
y
and L
z
, undergoing a temperature change T . The new volume of
the cuboid will be
L
x
L
y
L
z
(1 + T ) = L
x
(1 + T ) L
y
(1 + T ) L
z
(1 + T )
which expands to give
(1 + T ) = (1 + T )
3
which can be simplied further for T 1 to
= 3
What about the eect of pressure? When we discuss thermal expansion it is
usually assumed that the measurement is made at atmospheric pressure.
However in thermodynamics, where temperature and pressure may both vary
simultaneously, we need a more careful denition. Therefore the denition is
always written as a partial derivative with a subscript to remind us that the
pressure p must remain constant.
So the coecient of linear expansion can be dened as
=
1
L
_
L
T
_
p
and the coecient of cubical expansion as
=
1
V
_
V
T
_
p
3.3.2 Thermal Expansion of Liquids
The concept of linear expansion is not applicable to liquids, so we only need
consider the coecient of cubical expansion dened by
=
1
v
_
v
T
_
p
20 TOPIC 3. PROPERTIES OF SOLIDS AND LIQUIDS
Table 3.2 compares the expansion coecients for several solids with those for
water and mercury. Note how the liquids have much higher expansion coecients.
Table 3.2: Comparison of thermal expansion coecients
substance coecient of linear coecient of volumetric
expansion expansion
(K
1
) (K
1
)
aluminium 23 10
6
69 10
6
copper 17 10
6
51 10
6
lead 29 10
6
87 10
6
mild steel 11 10
6
33 10
6
glass 4 10
6
12 10
6
mercury 180 10
6
water 210 10
6
More data for other solids are given in HLT pages 4546. Data for liquids are in
HLT page 53.
3.3.3 Example 1
Problem: A cube of aluminium of side 100 mm is heated from 10
o
C to 310
o
C at
constant pressure.
(a) Find the new length of each side.
(b) Hence nd the increase in volume.
(c) Compare the result (b) with the volume increase calculated using the volume
coecient of expansion.
(d) Calculate the volume change if, in addition to the temperature rise, the
pressure also increases by 1000 bar (10
8
Pa).
Solution: (a) Here L = 100 mm, T = 310 10 = 300
o
C, and table 3.2 lists
= 23 10
6
K
1
. Hence the new length of each side is
L + L = 100 (1 + 23 10
6
300) = 100.69 mm
3.3. THERMAL EXPANSION 21
(b) We can nd the new volume directly from the increased side length by
V + V = (100 (1 + 23 10
6
300))
3
100
3
(1 + 3 23 10
6
300) mm
3
by the binomial expansion
and hence the change in volume
V = 100
3
3 23 10
6
300 = 20700 mm
3
(c) Table 3.2 lists = 69 10
6
K
1
hence we can calculate the change in volume
directly as
V = 10
6
69 10
6
300 = 20700 mm
3
(d) Table 3.1 gives the bulk modulus K = 75 10
9
Pa so the change in volume
after the pressure rise will be
V = 20700
10
6
10
8
75 10
9
= 19367 mm
3
3.3.4 Example 2
Problem: A mercury-in-glass thermometer has a bulb volume of 20 mm
3
and a
tube diameter of 0.08 mm. If the mercury occupies the bulb and 10 mm of tube at
0
o
C, what will be the additional length of the mercury column at 100
o
C? Ignore
eects of changing pressure in the vapour space above the mercury.
Solution: Here we have diameter D = 0.08 mm, a temperature rise T = 100
o
C,
and from table 3.2 we see that = 180 10
6
K
1
. Hence
initial tube cross-section area = 0.08
2
/4 = 0.005027 mm
2
nal tube cross-section area = 0.005027 (1 + 2 4 10
6
100)
= 0.005031 mm
2
initial mercury volume = 20 + 0.005027 10 = 20.05027 mm
3
initial glass volume to 0
o
C mark = 20.05027 mm
3
increase in mercury volume = 20.05027 180 10
6
100
= 0.360905 mm
3
increase in volume to 0
o
C mark = 20.05027 3 4 10
6
100
= 0.02406 mm
3
22 TOPIC 3. PROPERTIES OF SOLIDS AND LIQUIDS
dierence between increases in volume = 0.360905 0.02406 = 0.336844 mm
3
extra column length at 100
o
C = 0.336844/0.005031 = 67 mm
3.4 Specic Heat Capacity
3.4.1 Specic Heat Capacity and Thermal Capacity
If heat is added to solid or liquid of mass m to raise its temperature by T , then
the amount of heat required Q is given by
Q = mcT
where the constant c is a property of the material called the specic heat capacity.
The product mc, which obviously depends not only on the nature of the material
but also the amount of the material, is called the thermal capacity.
Table 3.3 lists the specic heats, in units of J kg
1
K
1
, for some common
materials, along with some other properties that will be used in the following
sections. For more data see also HLT pages 45, 53 and 75.
Table 3.3: Comparison of thermal properties
substance density specic heat conductivity k viscosity
capacity c
(kg m
3
) (J kg
1
K
1
) (Wm
1
K
1
) (kg m
1
s
1
)
aluminium 2700 880 205
copper 8960 380 390
lead 11300 126 126
mild steel 7850 450 50
glass 2480 139 1
mercury 13500 139 8 1.55 10
3
water 1000 4180 0.61 1.00 10
3
3.4.2 Example 3
Problem: The 100 mm aluminium cube of the example in section 3.3.3 is heated
from 10
o
C to 310
o
C. Find the amount of heat required.
Solution: From table 3.3 (or HLT), the density of aluminium is = 2700 kgm
3
and
the specic heat capacity is c = 880 J kg
1
K
1
therefore the mass is
3.4. SPECIFIC HEAT CAPACITY 23
m = 0.1
3
2700 = 2.7 kg. Given that the temperature rise is T = 310 10 = 300
K, we can calculate the required heat from
Q = mcT = 2.7 880 300 = 712800 J = 712.8 kJ
3.4.3 Example 4
Problem: In the previous example, is any energy required to expand the aluminium
block against the external pressure? What if the surrounding pressure were 1000
bar instead of 1 bar?
Solution: Try a calculation at a pressure of 1000 bar.
force due to pressure on one face = 1000 10
5
0.1
2
= 10
6
N
increase in distance between faces = 0.1 23 10
6
300 = 6.9 10
4
m
work to move two faces apart = 10
6
6.9 10
4
= 690 J
total work for all faces = 3 690 = 2070 J
Alternatively we can say
pressure increase in volume = 1000 10
5
0.1
3
69 10
6
300 = 2070 J
In this case the result of 2070 J = 2.07 kJ is negligible compared with 712.8 kJ, but
we will see later the result is very dierent when we heat a gas at constant
pressure.
3.4.4 Example 5
Problem: 0.2 kg of copper at 95
o
C is lowered into 0.5 kg of water at 5
o
C. What is
the nal temperature?
Solution: Let T
o
C be the nal temperature of the copper and the water. The heat
energy which leaves the copper all goes to the water, so we can write
m
copper
c
copper
|T
copper
| = m
water
c
water
|T
water
|
0.2 380 (95 T ) = 0.5 4180 (T 5)
T =
76 95 + 2090 5
2090 + 76
= 8.16
o
C
Note that the values 76 and 2090, with units of J K
1
, represent thermal capacities
(mass specic heat capacity).
24 TOPIC 3. PROPERTIES OF SOLIDS AND LIQUIDS
3.5 Transport Properties
3.5.1 One-Dimensional Conduction
In a solid (or in a liquid or gas that is not stirred), heat ows in any direction when
there is a gradient of temperature (decreasing) in that direction. The rate of heat
transfer

Q may be related to the cross-sectional area A, the temperature gradient
dT/dx and the thermal conductivity k of the material by

Q = k A
dT
dx
The heat ow per unit area is the heat ux, q, so
q = k
dT
dx
3.5.2 Example 6
Problem: A wall is constructed of 20 mm thickness of copper (k = 390 Wm
1
K
1
)
overlaid with 30 mm thickness of asbestos (k = 20 Wm
1
K
1
). If the exposed
copper surface is at 500
o
C and the exposed asbestos surface is at 50
o
C as
shown in gure 3.1, nd the temperature at the copper-asbestos interface and
calculate the steady heat ow per unit area across the composite wall.
Solution: When the temperatures are steady, the rate of heat transfer

Q is a
constant, so dT/dx is constant and the temperature varies linearly with x across
the thickness of each material. Let the unknown intermediate temperature at the
copper-asbestos interface be T
o
C. The copper and asbestos share the same
asbestos copper
T C
o
50 C
o
500 C
o
20 30 mm mm
Q
.
Figure 3.1: One-dimensional conduction example
3.5. TRANSPORT PROPERTIES 25
cross-sectional area A, and the same heat ow rate

Q goes through both parts in
series.
Hence
for the copper

Q
A
= 390
T 500
0.02
for the asbestos

Q
A
= 20
50 T
0.03
Solving these simultaneous equations gives
T = 485.12
o
C and

Q
A
= 290 kWm
2
3.5.3 Viscosity
Shear stresses arise in a uid when there is a gradient of velocity u, as illustrated
in gure 3.2. The shear stress is a force per unit area, dened by
=
_
u
y
_
where y is the distance perpendicular to the ow. The partial derivative is used to
remind us that the derivative may also depend on the position x along the ow of
the stream, and perhaps on time t too. The constant is called the coecient of
viscosity.
There are in fact two ways of dening the viscosity. The dynamic viscosity and
the kinematic viscosity are related by the density of the uid such that
=
u
y
fluid flow

Figure 3.2: Shear stresses and velocity gradient


26 TOPIC 3. PROPERTIES OF SOLIDS AND LIQUIDS
These lecture notes will use only the dynamic viscosity, but you should beware
when looking at thermodynamics and uid mechanics textbooks that some may
use the kinematic viscosity.
3.5.4 Example 7
Problem: A circular shaft of diameter 50 mm rotates at 3000 revolutions per minute
inside a circular bearing of diameter 50.2 mm and axial length 50 mm as shown in
gure 3.3. The space between the two is lled with lubricating oil of viscosity 10
mPa s (0.01 Ns m
2
). Calculate the power dissipated in viscous friction.
Solution: We will begin by making the (very good) approximation that the space
between the two cylindrical bearing surfaces is so small that the problem may be
treated as if the two surfaces were at. You will nd this approximation is very
common in such problems.
relative speed of surfaces U =
0.05
2

2 3000
60
= 7.854 ms
1
radial thickness of oil lm h =
0.0502 0.0500
2
= 0.0001 m
We will now make an addition assumption, without proof, that the velocity gradient
is uniform across the thickness of the oil lm so that
dU
dy
=
U
h
This assumption is reasonable because there are no other forces acting on the
3000 rpm
0.1 mm
Figure 3.3: Rotating shaft with oil lm
3.5. TRANSPORT PROPERTIES 27
uid in the x direction.
velocity gradient in oil
dU
dy
=
7.854
0.0001
= 78540 s
1
shear stress = 0.01 78540 = 785.4 Nm
2
bearing surface area A = 0.05 0.05 = 0.00785 m
2
force on shaft surface A = 785.4 0.00785 = 6.168 N
power = 6.168 7.854 = 48.4 W
3.5.5 Diusion
Diusion in liquids and gases occurs when molecules of one species migrate
among molecules of another. It is a process by which spatially non-uniform
distributions of species tend to become uniform by random motion and collisions
between molecules, as illustrated in gure 3.4. If the number-density of molecules
per unit volume n
1
for the migrating species 1 is very much less than the
number-density n
2
of the host species 2, then the host density remains
approximately uniform and the migration is governed by a very simple equation in
terms of n
1
only. The number ow rate

N
1
of species 1 crossing an x z plane of
area A is related to the local number-density gradient dn
1
/dy by

N
1
= AD
12
dn
1
dy
More conveniently, multiplying both sides by the mass of a molecule, the equation
can be written to relate the mass ow rate to the mass density, or concentration, of
before after
Figure 3.4: Diusion of one species into another
28 TOPIC 3. PROPERTIES OF SOLIDS AND LIQUIDS
the migrating species by
m
1
= AD
12
d
1
dy
where D
12
is the diusion coecient specic to species 1 in the given
number-density of species 2.
3.5.6 Example 8
Problem: A cylindrical jar of Silica Gel drying agent 80 mm in diameter is opened
and a wet cloth is laid across the top so that, adjacent to the cloth, the density of
water vapour is 0.018 kg m
3
while at the dry surface of the powder 50 mm below,
the vapour density is zero. The diusion coecient for water vapour in air at
atmospheric pressure and temperature is 2.6 10
5
m
2
s
1
. Calculate the rate at
which water is transferred from the cloth to the Silica Gel.
Solution: Here the radius r = 80/2 = 40 mm = 0.04 m so the magnitude of the
diusion rate is
m = 2.6 10
5
0.04
2

0.018
0.05
= 4.7 10
8
kg s
1
TOPIC 4. BEHAVIOUR OF GASES 29
Topic 4
Behaviour of Gases
4.1 Introduction
In this chapter we will look at some of the ways we can relate the properties of
gases, beginning with ideal gases and the ideal gas law, then exploring why gases
may not behave ideally and how we can account for this, before looking at what
happens to real gases at extremes of temperature and pressure. We will nish by
considering the phase rule.
4.2 Ideal Gases
4.2.1 Ideal Gas Law
An equation of state links the properties of state, usually p, V and T . The equation
of state for an ideal gas, known as the ideal gas law, is
pV = n R
0
T or pV =
m
M
R
0
T or pV = mRT or pv = RT
where
p = pressure Nm
2
(or Pa)
V = volume m
3
v = specic volume m
3
kg
1
T = absolute temperature K
m = mass kg
M = molar mass kg kmol
1
n = number of kmol
R
0
= molar gas constant = 8315 J kmol
1
K
1
R = specic gas constant J kg
1
K
1
30 TOPIC 4. BEHAVIOUR OF GASES
Note the relationships between the number of moles, the mass, the molar mass,
and the molar and specic gas constants
n =
m
M
R =
R
0
M
4.2.2 Example 1
Problem: Calculate the mass of air in a tank of volume 0.1 m
3
if the pressure is 1
bar and the temperature 20
o
C. For air, the mean molar mass may be taken as
29 kg kmol
1
.
Solution: Given V = 0.1 m
3
, p = 1 bar = 10
5
Pa, T = 20
o
C = 293.15 K, M = 29
kg kmol
1
. Hence calculate the specic gas constant R = R
0
/M = 8315/29 = 287
J kg
1
K
1
then
m =
pV
RT
=
10
5
0.1
287 293.15
= 0.119 kg
There are two important steps to note here. First, we must convert the pressure p
from bar to Pa, and the temperature T from
o
C to K. Secondly, we must express
the specic gas constant R in units of J kg
1
K
1
not kJ kg
1
K
1
.
4.2.3 Example 2
Problem: Calculate the pressure in a tank of volume 5 m
3
containing 2 kmol of gas
at a temperature of 100 K.
Solution: We are given V = 5 m
3
, n = 2 kmol, T = 100 K, hence
p =
n R
0
T
V
=
2 8315 100
5
= 3.32 10
5
Pa = 3.32 bar
Note that because we are given the number of moles, we do not need to know
what gas is in the tank.
4.2.4 Specic Heat Capacities
In section 3.4.1 we dened the specic heat capacity c of a solid or liquid in terms
of the amount of heat Q added to a mass m to raise its temperature by T
Q = mcT
We also saw in section 3.4.3 that the amount of energy required to expand the
solid or liquid against a constant pressure was negligible.
4.2. IDEAL GASES 31
For gases, we nd that the amount energy Q which must be added to a mass m
of gas to raise its temperature by T is signicantly higher if we keep the gas at
constant pressure than if we keep it at constant volume. This is because additional
energy is required as the gas expands and does work against the surroundings.
We must therefore dene two specic heats for gases.
The specic heat at constant volume is denoted by c
v
and the specic heat at
constant pressure by c
p
. Both have units of J kg
1
K
1
(or kJ kg
1
K
1
). They can
be dened in terms of the amount of heat Q which must be added to m kg of gas
to raise its temperature by T as
Q = mc
v
T at constant volume
Q = mc
p
T at constant pressure
We will prove later that for ideal gases the two specic heats are related by
c
p
c
v
= R
where R is the specic gas constant. We will also meet the ratio of specic heats
dened by
=
c
p
c
v
from which it can be shown for ideal gases that
c
v
=
R
1
and c
p
=
R
1
The kinetic theory of gases predicts that for monatomic gases = 5/3 = 1.67, for
diatomic gases = 7/5 = 1.40, and for gases with three or more atoms per
molecule = 4/3 = 1.33. For air, which is composed principally of the diatomic
gases N
2
and O
2
, = 1.4. Values of c
v
, c
p
and are tabulated for some typical
gases on HLT page 75.
4.2.5 Partial Pressures
Daltons law of partial pressures states that in a gas mixture occupying a volume V
at absolute temperature T each constituent gas of mass m
i
exerts a partial
32 TOPIC 4. BEHAVIOUR OF GASES
pressure p
i
equal to the pressure that it would have exerted if it had occupied the
entire volume at the same temperature without any other gas present.
Each gas in the mixture will obey the ideal gas law in the form
p
i
V = n
i
R
0
T or p
i
V = m
i
R
i
T
where n
i
is the number of moles of gas i and R
i
is the specic gas constant for gas
i . Thus the total pressure is the sum of the partial pressures
p =

i
p
i
4.2.6 Example 3
Problem: Calculate the pressure in a tank of volume 5 m
3
containing 0.4 kmol of
oxygen and 1.6 kmol of nitrogen at a temperature of 200 K.
Solution: We are given V = 5 m
3
, n
O
2
= 0.4 kmol, n
N
2
= 1.6 kmol, T = 200 K,
hence
p =

i
p
i
=

i
n
i
R
0
T
V
=
R
0
T
V

i
n
i
=
8315 200
5
(0.4 + 1.6)
= 6.64 10
5
Pa
= 6.64 bar
4.2.7 Gas Mixtures
It follows from the law of partial pressures that for a mixture of ideal gases the
equivalent gas constant R is determined by the mass fraction of each gas. Thus if
gas i has mass m
i
and specic gas constant R
i
, the equivalent gas constant for
mixture of total mass m is
R =
1
m

i
m
i
R
i
Similar equations can be written for the specic heats c
v
and c
p
.
4.3. NON-IDEAL GASES 33
4.2.8 Example 4
Problem: Assuming air consists of 23 % oxygen and 77% nitrogen by mass,
estimate the specic gas constant for air.
Solution: We are given m
O
2
/m = 0.23, m
N
2
/m = 0.77. Using M
O
2
= 32 kgkmol
1
,
M
N
2
= 28 kg kmol
1
and R
0
= 8315 J kmol
1
K
1
, we get
R =
1
m

i
m
i
R
i
=

i
m
i
m
R
0
M
i
= R
0

i
m
i
/m
M
i
= 8315
_
0.23
32
+
0.77
28
_
= 288 J kg
1
K
1
which is in good agreement with the tabulated value of 287 J kg
1
K
1
.
4.3 Non-Ideal Gases
4.3.1 Reasons for Non-Ideal Behaviour
Gases approximate ideal behaviour at high temperature and low pressure. Both
these eects contribute to give a low density where the molecules are far apart. In
the kinetic theory of gases, molecules move randomly and there are no forces
between the molecules. Non-ideal behaviour arises as we reduce the temperature
and increase the pressure, forcing the molecules closer together. Firstly, a
molecule cannot move to a volume where there is another molecule. Secondly,
there actually are intermolecular forces.
4.3.2 Generalised Compressibility Chart
Instead of using the ideal gas law pv = RT , we can adjust for non-ideal behaviour
by
pv = ZRT
where Z is the compressibility factor . How do we obtain values of Z? First we
dene the dimensionless variables reduced pressure p
r
and reduced temperature
T
r
as
p
r
=
p
p
c
and T
r
=
T
T
c
34 TOPIC 4. BEHAVIOUR OF GASES
Figure 4.1: Generalised compressibility chart, HLT page 88
4.4. REAL GASES 35
p
Z
T
r
r
Figure 4.2: Example 5
where p
c
and T
c
are the pressure and temperature at the critical point , explained in
section 4.4.2. Values of Z obtained by experiment can then be read from the
generalised compressibility chart in gure 4.1 showing Z as a function of p
r
for
values of constant T
r
.
4.3.3 Example 5
Problem: Given p and T , calculate v for a certain gas.
Solution: First we nd values of p
c
and T
c
from HLT page 75, then we calculate the
reduced pressure and temperature p
r
and T
r
. From the p
r
axis, rule a line vertically
up to the correct T
r
value, then a line horizontally across and read o the Z axis,
as illustrated in gure 4.2. Finally, we calculate v from pv = ZRT .
4.4 Real Gases
4.4.1 Phase Changes
Real gases may behave similarly to ideal gases at high temperature and low
pressure, but at low temperature and high pressure they can be liqueed and even
solidied! Consider a sealed vessel containing water (liquid) and water (vapour) in
thermal equilibrium. If the temperature of the system is 100
o
C, then the pressure
is approximately 1 bar. At other temperatures, the pressure will be dierent. We
can plot the pressure p as a function of the temperature T to give the vapour
pressure curve in gure 4.3. Along the line are liquid and vapour, both saturated.
The liquid is at its boiling point and the vapour is at the same temperature. O the
line we have either liquid below its boiling point, called subcooled liquid, or vapour
above the saturation temperature, called superheated vapour.
36 TOPIC 4. BEHAVIOUR OF GASES
superheated
vapour
subcooled
liquid
C
100 C
1.013 bar
T
p
Figure 4.3: Variation of p with T for real substance
4.4.2 Critical Point
Consider now the variation of specic volume with temperature. If in the above
experiment the specic volumes are plotted at each temperature as in gure 4.4,
then it is found that at each temperature there are two values of specic volume,
one for liquid alone, v
f
, and one for vapour alone, v
g
. Eventually they become
equal at C, where the saturation vapour pressure curve ends. At this point, known
as the critical point , all the liquid and vapour properties become equal. There is
now no dierence between liquid and vapour. The surface tension is zero. Even
the interface cannot be seen as the refractive indices are equal.
C
v
p or T
vapour
liquid
Figure 4.4: Variation of p with v for real substance
Examples of critical points are shown in table 4.1. More can be found on HLT page
75.
Table 4.1: Critical points
substance p
c
bar T
c
o
C v
c
m
3
kg
1
water 221.2 374.15 3.17 10
3
nitrogen 33.9 147 3.2 10
3
carbon dioxide 74.0 31 2.2 10
3
4.4. REAL GASES 37
p
v
A B
C
1
2
3
4
increasing
T
Figure 4.5: p v diagram with isotherms for a real substance
Sometimes the term gas is used if the temperature is above the critical
temperature, and vapour is used if the temperature is below the critical
temperature. Thus air, mostly nitrogen, is normally a gas, whereas steam is
normally a vapour.
4.4.3 p v Diagram
The p v diagram with isotherms for a real substance is shown in gure 4.5 where
C is the critical point
AC is the saturated liquid line
BC is the saturated vapour line
1 is the subcooled liquid region
2 is the saturated liquid and vapour region
3 is the superheated vapour region
4 is gas
Note that at the critical point C the gradient and curvature of the critical
temperature isotherm on the p v diagram are both zero.
_
p
v
_
c
= 0 and
_

2
p
v
2
_
c
= 0
4.4.4 p v T Surface
Consider the behaviour of an ideal gas represented as a p v T surface as in
gure 4.6. The projection of the surface on the p v plane gives Boyles law pv =
constant when T = constant, and the projection on the p T plane gives Charles
law p/T = constant when v = constant.
38 TOPIC 4. BEHAVIOUR OF GASES
p
v
p
T
pv =
p/ T =
constant
constant
Figure 4.6: p v T surface and its projections for ideal gas
Figure 4.7: p v T surface on freezing for real substance
For a real substance (solid liquid gas) which contracts on freezing, the
p v T surface looks like the surface on the right of gure 4.7. For a substance
which expands on freezing (such as water), it looks like the surface on the left.
Such behaviour is less common.
4.4.5 Phase Rule
A phase is a region of uniform composition and state. We shall only consider the
bulk phases, such as solid, liquid and gas, and dierent crystalline solid forms, but
will not consider electrical elds and interfacial eects. The phase rule states that
F + P = C + 2
4.4. REAL GASES 39
where
F is the number of degrees of freedom, that is, the number of thermodynamic
properties which can be varied independently.
P is the number of phases, that is, solid, liquid or gas, or dierent crystalline
solid forms.
C is the number of chemical species. We are usually concerned with C = 1.
This is true even for air because the composition of air is xed.
If there is one phase, such as liquid or gas, then P = 1 F = 2. Hence only two
independent intensive properties are needed to dene the state and all other
dependent properties. To dene extensive properties, the mass m of the system is
needed. For example, the pressure and the temperature can be specied to
determine the system, and then the equation of state will calculate the volume, or
( p,T ) v . Alternatively, ( p,v ) T or ( T ,v ) p.
If there are two phases, such as liquid + vapour together, then P = 2 F = 1 so
only one independent variable can be specied. If the pressure is given the
temperature is then xed, and vice versa. That is, p T , or p v
f
,v
g
. For
example, for water, if p = 1.013 bar, then T = 100
o
C as long as water and steam
are both present. One independent intensive property denes the dependent
intensive properties of the phases. Note that v is not xed, only its limits
v
f
v v
g
. The extra information needed to specify v is given by the dryness
fraction x described in section 5.2.2.
If there are three phases, solid + liquid + vapour together, then P = 3 F = 0.
There is no freedom and so the point called the triple point is dened. For water,
the triple point is at T = 273.16 K = 0.01
o
C and p = 0.00611 bar, at which water,
ice and steam co-exist in equilibrium.
40 TOPIC 5. BEHAVIOUR OF STEAM-WATER MIXTURES
Topic 5
Behaviour of Steam-Water Mixtures
5.1 Introduction
The thermodynamic behaviour of steamwater mixtures is of great engineering
importance. Properties are found from a steam chart or from steam tables like
those in HLT on pages 5471. There are dierent tables for the saturated,
superheated and supercritical regions.
5.2 Saturated Region
5.2.1 Saturated Table in HLT
When water and steam are in equilibrium with each other, we say that the steam is
saturated. To nd the properties of such steamwater mixtures we use the
saturated table in HLT on pages 5465. Pages 5455 are tabulated as a function
of temperature and pages 5665 as a function of pressure.
Pages 5455 t
s
o
C p bar v
f
L kg
1
v
g
L kg
1
50 0.123350 1.0121 12046

saturation corresponding specic volumes
temperature pressure f = liquid g = vapour
multiply by 10
3
to
convert to m
3
kg
1
Pages 5665 p bar t
s
o
C v
f
L kg
1
v
g
L kg
1
200 365.71 2.0374 5.87

saturation corresponding specic volumes
pressure temperature f = liquid g = vapour
5.2. SATURATED REGION 41
Note that the tables tell us the pressure and temperature of the mixture with
certainty, and the values of other properties, such as specic volume, were the
mixture to be all liquid or all vapour. For example, at 50
o
C, we can see from the
table on HLT page 54 that the specic volume of the vapour present in a mixture
will 12046 L kg
1
, and that at a pressure of 200 bar, HLT page 64 shows that the
liquid present will have a specic volume of 2.0374 Lkg
1
.
However, the tables cannot tell us what fraction of the mass of any mixture will be
liquid and what fraction of the mass will be vapour without further information. For
example, if we knew the specic volume v of the mixture as well as the pressure
and temperature, we could then use the values v
f
and v
g
to estimate the mass
fractions of liquid and vapour present.
Note that in HLT the specic volume v is given in units of 10
3
m
3
kg
1
, or L kg
1
.
Remember to convert from L kg
1
back to m
3
kg
1
when necessary.
5.2.2 Dryness Fraction
The mass fractions of liquid and vapour present in the saturated region are
conveniently described by the dryness fraction, also called the quality of the
steam. If a mixture of total mass m kg contains a mass of liquid m
f
kg and a mass
of vapour m
g
kg, then the dryness fraction x is dened as
x =
m
g
m
f
+ m
g
=
m
g
m
= 1
m
f
m
Thus
if x = 0 the mixture is saturated liquid
if x = 1 the mixture is saturated vapour also called dry saturated vapour or dry
steam
if 0 < x < 1 the mixture is saturated liquid + vapour also called wet steam
x < 0 and x > 1 are impossible.
The masses of liquid and vapour may be determined from the total mass m and
the dryness fraction x by
m
f
= (1 x)m m
g
= x m
42 TOPIC 5. BEHAVIOUR OF STEAM-WATER MIXTURES
The volume of the mixture can be calculated from the dryness fraction by
V = mv = m
f
v
f
+ m
g
v
g
= (1 x)m v
f
+ x m v
g
= m((1 x) v
f
+ x v
g
)
and the specic volume by
v = v
f
+ x (v
g
v
f
)
We use a similar approach to calculate other properties, such as the internal
energy U which we will meet in section 6.3.2 and the enthalpy H which is
introduced in section 7.2.2. For example, the internal energy of the mixture can be
calculated from the dryness fraction by
U = mu = m
f
u
f
+ m
g
u
g
= (1 x)m u
f
+ xm u
g
= m(u
f
+ x(u
g
u
f
))
HLT also tabulates the dierence u
fg
= u
g
u
f
. Hence we calculate the specic
internal energy simply by
u = u
f
+ x u
fg
and similarly for other properties such as h and the specic entropy s.
5.2.3 Example 1
Problem: Calculate the specic volume and specic internal energy of a
steamwater mixture at 200 bar if x = 0.3.
Solution: From HLT page 64, at 200 bar v
f
= 2.0374 L kg
1
= 2.0374 10
3
m
3
kg
1
, v
g
= 5.87 L kg
1
= 5.87 10
3
m
3
kg
1
, u
f
= 1785.9 kJ kg
1
and
u
fg
= 514.9 kJ kg
1
.
v = v
f
+ x (v
g
v
f
)
= 2.0374 10
3
+ 0.3 (5.87 2.0374) 10
3
= 3.1872 10
3
m
3
kg
1
u = u
f
+ x u
fg
= 1785.9 + 0.3 514.9
= 1940.4 kJ kg
1
5.2.4 Example 2
Problem: A steamwater mixture of mass 1 kg is contained in a vessel of volume 2
m
3
at 90
o
C. Calculate the mass of water.
5.3. SUPERHEATED STEAM 43
Solution: From HLT page 54, at 90
o
C, v
f
= 1.0361 L kg
1
= 1.0361 10
3
m
3
kg
1
and v
g
= 2361 L kg
1
= 2.361 m
3
kg
1
. The mass is m = 1 kg and the volume is
V = 2 m
3
, so the specic volume is v = 2 m
3
kg
1
.
v = v
f
+ x (v
g
v
f
)
x = (v v
f
)/(v
g
v
f
) = (2 1.0361 10
3
)/(2.361 1.0361 10
3
) = 0.847
m
f
= (1 x) m = (1 0.847) 1 = 0.153 kg
5.3 Superheated Steam
5.3.1 Superheated Table in HLT
If steam is at a temperature above the saturation temperature for the particular
pressure, there cannot be any liquid water in equilibrium with it. We call this
superheated steam. To nd the properties of superheated steam, we use the
superheated table in HLT on pages 6670.
p t: t
s
400 500
(t
s
)
200 v 5.8745 9.9470 14.771
(365.7)
values of specic volume v L kg
1
for steam
pressure and at t
s
= 365.7
o
C, 400
o
C and 500
o
C,
corresponding all at p = 200 bar
saturation temperature
For example, at 200 bar and 400
o
C, the steam will have a specic volume of
9.9470 L kg
1
.
5.3.2 Example 3
Problem: Calculate the specic volume of steam at 10 bar, 350
o
C.
Solution: From HLT pages 6869, at p = 10 bar, t
s
= 179.9
o
C. We are given
T = 350
o
C, so T > t
s
, conrming that the steam is superheated. If T < t
s
, we
would have to go back to the saturated tables. For 300
o
C, v
300
= 257.98 L kg
1
and for 400
o
C, v
400
= 306.49 L kg
1
.
44 TOPIC 5. BEHAVIOUR OF STEAM-WATER MIXTURES
Hence for T = 350
o
C by linear interpolation
v
350
= v
300
+ (v
400
v
300
)
350 300
400 300
= 257.98 + (306.49 257.98) (50/100) = 282.2 L kg
1
= 0.2822 m
3
kg
1
5.4 Supercritical Steam Properties
The supercritical table in HLT on page 71 gives steam properties in the range
400
o
C T 800
o
C and 240 bar p 1000 bar, but you will probably not need
these this year.
5.5 Other Tables with a Phase Change
HLT pages 7273 has data for ammonia (NH
3
), tetrauoroethane (CH
2
FCF
3
) and
carbon dioxide (CO
2
) at saturation and slightly superheated conditions.
TOPIC 6. FIRST LAW FOR CLOSED SYSTEMS 45
Topic 6
First Law for Closed Systems
6.1 Introduction
You have already met the First Law of thermodynamics in the form of conservation
of energy. In this chapter we will recast it into the form we need to understand and
solve thermodynamic problems. We must begin by mastering some new
terminology, and discuss internal energy.
6.2 Systems, Processes, Paths and Cycles
6.2.1 Systems
The system is that part of the universe which interests us. The rest of the universe
is the surroundings, therefore
system+ surroundings = universe
The system is separated from the rest of the universe by the system boundary or
control surface. A typical system might be a mass of gas, or a mixture of liquid and
gas, such as steam and water. We shall consider three types of system, illustrated
in gure 6.1, as follows
an isolated system in which heat, work and matter cannot cross the system
boundary the system will always consist of the same mass, and the
boundary will not move or deform
a closed system in which heat and work can cross the system boundary, but
matter cannot the system will always consist of the same mass, but the
boundary may move or deform
an open system in which heat, work and matter can cross the system
boundary the total mass in the system may vary, and the boundary may
move or deform but is usually xed.
46 TOPIC 6. FIRST LAW FOR CLOSED SYSTEMS
Q
W
W
isolated
Q
closed
Q
W
m
m
open
Figure 6.1: Types of system
The convention we shall use for positive Q and W, as illustrated in gure 6.2, is
heat transfer to the system is positive
work done by the system is positive
but this convention is not universal. Recent editions of the books by ROGERS AND
MAYHEW and EASTOP AND MCCONKEY use the convention that work done on the
system is positive, so take care if you consult these.
Today we recognise that Q and W are dierent forms of energy, so we use the
same system of units, J, kJ or MJ, for both. In the past, heat and work were often
specied in dierent units, such as mechanical units of ft lbf for work, and heat
units such as British Thermal Units or BTU for heat. It was not until 1850 that
Joule accurately measured the conversion factor for BTU and ft lbf.
6.2.2 Processes and Paths
A process is a change in the system from one state to another along a path that
may be dened by a succession of intermediate states and/or by the interaction of
system
Q
W
Figure 6.2: Sign convention for heat and work
6.2. SYSTEMS, PROCESSES, PATHS AND CYCLES 47
F
2
F
1
cylinder
piston
T
1
T
2
gas
Figure 6.3: Example of closed system
the system with its surroundings. Processes are of two general types
a reversible process is one which can be reversed so that the system and the
surroundings are unchanged at the end of the process and the reverse process
an irreversible process is one which cannot be reversed in this way.
As well as being reversible or irreversible, processes can also be described as
follows
an adiabatic process is one in which the heat transfer across the boundary of
the system is zero
a diabatic process is one in which the heat transfer across the boundary of the
system is non-zero
an isothermal process is one which takes place at constant temperature
an isobaric process is one which takes place at constant pressure
an isochoric process is one which takes place at constant volume.
As an example of reversible and irreversible processes, consider a closed system
of gas contained in a cylinder and piston shown in gure 6.3. The gas may be
allowed to expand reversibly, or irreversibly. To expand reversibly, it must be done
slowly, so that force F
1
is only slightly greater than F
2
and so that there is sucient
time for heat to be transferred to or from the surroundings so that T
1
T
2
. Also,
the piston must slide without friction. To expand irreversibly, remove the restraining
force F
2
. The piston will now move quickly, and T
1
= T
2
, F
1
= F
2
.
6.2.3 Cycles
A cycle is a series of processes after which the nal thermodynamic state of the
system is the same as the initial state. Later we shall see the thermodynamic,
Rankine and refrigeration or heat pump cycles. On a p V or other diagram, the
lines representing the series of processes in a cycle will form a closed loop as in
48 TOPIC 6. FIRST LAW FOR CLOSED SYSTEMS
p
V
1
2
3
4
Figure 6.4: Cycle on a p V diagram
gure 6.4. The arrows show the direction of the processes. In power cycles, such
as the Rankine, Otto, Diesel and Joule cycles, the arrows go clockwise. In heat
pump and refrigeration cycles, the arrows go anti-clockwise.
6.3 First Law for Closed System
6.3.1 Denitions
The First Law of thermodynamics for a closed system may be dened in several
ways. One possible verbal denition is
If a closed system is taken through a cycle, then the net work delivered to the
surroundings is equal to the net heat taken from the surroundings.
This can be written mathematically as

cycle
Q =

cycle
W As the thermodynamic state is the
same at the start and at the end
where

cycle
means the summation around the cycle.
A more useful denition, applicable to individual processes in a closed system,
relates the heat in and work out to the change in internal energy of the system by
Q W = U or q w = u
For a process from state 1 to state 2, the First Law can be written as
Q W = U
2
U
1
or q w = u
2
u
1
The First Law can also be written in dierential form as
dQ dW = dU or dq dw = du
6.3. FIRST LAW FOR CLOSED SYSTEM 49
1
2
A
B
C
p
V
Figure 6.5: Cycle on a p V diagram
6.3.2 Internal Energy
The phase rule tells us that the state of a single-phase single-component system,
such as a gas, can be dened by two properties, for example pressure and
volume. It is often useful to represent a cycle on a diagram of pressure against
volume, a p V diagram, as illustrated in gure 6.5.
Consider two cycles. The rst is process A from state 1 to state 2 and process B
from state 2 back to state 1. The second is process A and process C from state 2
back to state 1. The First Law gives
First Cycle Q
A
+ Q
B
= W
A
+ W
B
Second Cycle Q
A
+ Q
C
= W
A
+ W
C
therefore
Q
B
W
B
= Q
C
W
C
= Q W
for any path between points 1 and 2. Hence Q W is independent of the process,
or path from 1 to 2, therefore it is a property of state. This property of state is
called the internal energy. It represents the energy of the molecular motion, and it
increases with temperature. For a given mass of substance, the total internal
energy is denoted by U J (or kJ), and the specic internal energy by u J kg
1
(or
kJ kg
1
).
6.3.3 Specic Heat Capacity at Constant Volume
In the previous section, we chose pressure and volume to dene the state of the
gas in our system, but we could have chosen any two other properties. If we
choose T and v , we can write an equation to represent the fact that the internal
energy u of the gas in our system is a function of T and v in the form
u = u(T ,v ) (6.1)
50 TOPIC 6. FIRST LAW FOR CLOSED SYSTEMS
although we dont know yet what the function is. Dierentiating gives
du =
_
u
T
_
v
dT +
_
u
v
_
T
dv (6.2)
The First Law can be written in dierential form as dq dw = du. If the volume of
the system is held constant, dv = 0. This means the boundary of the system
cannot move and the work done is dw = p dv = 0, and therefore dq = du.
Substituting dq = du and dv = 0 into equation 6.2 gives
dq =
_
u
T
_
v
dT
or
_
q
T
_
v
=
_
u
T
_
v
But (q/T )
v
is a specic heat , the amount of heat which must be added to 1 kg
of mass to raise the temperature by 1
o
C. Because v is constant, this is c
v
, the
specic heat at constant volume, with units of J kg
1
K
1
(or kJ kg
1
K
1
), so u and
c
v
are related by
c
v
=
_
u
T
_
v
We will prove next year that for an ideal gas (u/v )
T
= 0 so equation 6.1
simplies to u = u(T ). That is, for an ideal gas the internal energy is a function of
temperature only. The partial derivative above becomes a total derivative
c
v
=
du
dT
and hence for an ideal gas we get the very important result that
u = c
v
T
whether v is constant or not .
6.3.4 Reversible Work in a Closed System
Consider a closed system of volume V containing a substance at pressure p. If
the system boundary moves slowly and without friction, allowing the volume to
increase by some small amount dV , then the system must do work against the
surroundings given by
dW = p dV
6.4. EXAMPLES 51
If the system changes from state 1 to state 2, then the reversible work done by the
system is
W =
2
_
1
p dV or w =
2
_
1
p dv
This corresponds to the area under a p V curve, between the curve and the V
axis.
6.4 Examples
6.4.1 Example 1
initial final
air
T
2
vacuum air
T
1
Figure 6.6: System for example 1
Problem: An insulated rigid box has two compartments as shown in gure 6.6.
One contains air at 298 K, the other a vacuum. The partition is removed. Calculate
the air temperature after expansion and mixing. Assume air is an ideal gas with
c
v
= 718 J kg
1
K
1
.
Solution: Apply the First Law to the system whose boundary is the large box.
Q = 0 because the box is insulated, and W = 0 because there is no volume
change, hence U
1
= U
2
. We saw in the previous section that for an ideal gas
u = u(T ) only, therefore T
1
= T
2
= 298 K.
6.4.2 Example 2
Problem: An insulated evacuated box, illustrated in 6.7, is suddenly punctured and
air enters from the atmosphere which is at 298 K. Calculate the nal temperature
of the air in the box.
Solution: Draw around the system an imaginary boundary across which no mass
passes. You can imagine the air contained in a large rubber balloon which
collapses as air enters the box. Now we can apply the First Law for a closed
52 TOPIC 6. FIRST LAW FOR CLOSED SYSTEMS
T
2
T
1
V
1
p
0
initial final
Figure 6.7: System for example 2
system. The box is insulated, so Q = 0, but the system boundary is moving, so
W = 0. The atmosphere does work on the system as it pushes the air into the box,
given by
W = p
0
V
1
therefore Q W = U gives
p
0
V
1
= U
2
U
1
Assuming air behaves as an ideal gas, c
v
= du/dT and c
v
is constant, hence
u = c
v
T . For a mass m of gas U = mc
v
T so
p
0
V
1
= mc
v
(T
2
T
1
)
Applying the ideal gas equation to the original air gives
p
0
V
1
= mRT
1
therefore
RT
1
= c
v
(T
2
T
1
)
and
T
2
= T
1
_
R
c
v
+ 1
_
= T
1
_
R
0
Mc
v
+ 1
_
= 298
_
8315
29 718
+ 1
_
= 417 K
The shaded area under the line on the p V diagram in gure 6.8 is the work done
on the system. Drawing a p V (or p v ) diagram can be extremely useful in
many examples.
6.4. EXAMPLES 53
2 1
V
p
p
0
V
box
V +V
box 1
shaded area is work done
on the system
Figure 6.8: p V diagram for example 2
6.4.3 Example 3
Problem: Air at temperature T
1
and volume V
1
is contained in a cylinder by a
frictionless piston of area A under the action of a spring of stiness k as in gure
6.9. The pressure is atmospheric, so there is no force in the spring. Now the air in
the cylinder is heated. Calculate the heat supply Q to raise the pressure in the
cylinder to p
2
.
Solution: First calculate the p V relationship. The piston movement x is given by
x =
V V
1
A
Equilibrium of the spring force and pressure forces on the piston gives
k
V V
1
A
= (p p
0
) A
That is, the spring constant the piston displacement equals the pressure
dierence across the piston the piston area. Rearranging gives
p = p
0
+ (V V
1
)
k
A
2
p
0
k
Q
A
air
V
1
Figure 6.9: System for example 3
Draw the p V diagram of gure 6.10. The work done by the system on the spring
and the atmosphere is the shaded area under the line, given by
W = p
0
(V
2
V
1
) +
1
2
(p
2
p
0
)(V
2
V
1
)
54 TOPIC 6. FIRST LAW FOR CLOSED SYSTEMS
W
2
1
V
p
p
0
V
1
V
2
p
2
Figure 6.10: p V diagram for example 3
The ideal gas law can be used to nd the mass of gas from
p
0
V
1
= mRT
1
and the nal temperature T
2
from
p
2
V
2
= mRT
2
The change in internal energy is then found from
U
2
U
1
= mc
v
(T
2
T
1
)
and then from the First Law
Q W = U
2
U
1
giving
Q = W + U
2
U
1
so Q can be solved in terms of p
2
.
6.4.4 Example 4
Problem: Consider a rigid insulated vessel of total volume 10 m
3
. Initially the
vessel is partitioned into two compartments, the rst of volume 1 m
3
, containing
saturated water and steam at 10 bar with dryness fraction 0.5, and the second
being evacuated as in gure 6.11. Calculate the nal conditions in the vessel if the
partition is removed.
Solution: With a steam-water mixture, the principles are the same but the
calculations are a little more involved. Instead of using the ideal gas law,
properties must be found from HLT.
The vessel is insulated, so Q = 0, and rigid, so W = 0, therefore from the First
Law the internal energy is constant, so U
1
= U
2
, and also u
1
= u
2
because the
6.4. EXAMPLES 55
V
2
vacuum
initial final
saturated water and steam V
1
Figure 6.11: System for example 4
mass is constant. First nd the initial specic volume taking values of v
f
and v
g
for
p = 10 bar from HLT page 60.
v
1
= v
f1
+ x
1
(v
g1
v
f1
)
= 1.1274 10
3
+ 0.5 (194.30 10
3
1.1274 10
3
)
= 0.0977 m
3
kg
1
The total mass of water + steam is
m =
V
1
v
1
=
1
0.0977
= 10.23 kg
and the nal specic volume is
v
2
=
V
2
m
=
10
10.23
= 0.977 m
3
kg
1
The internal energy, using values of u
f
and u
g
from HLT page 60 is
u
2
= u
1
= u
f1
+ x
1
u
fg1
= 761.5 + 0.5 1820.4 = 1671.7 kJ kg
1
Thus the nal state is dened by v
2
= 0.977 m
3
kg
1
and u
2
= 1671.7 kJ kg
1
.
Assuming the mixture is still saturated, use the saturation table and a process of
trial and error, or iteration, to nd x
2
. First try p
2
= 1 bar. This seems a sensible
initial guess because we know from the volume changes that v
2
= 10v
1
. For the
known value of u
2
and HLT page 58 calculate the corresponding dryness fraction
x
2
using
u
2
= u
f2
+ x
2
u
fg2
x
2
=
u
2
u
f2
u
fg2
=
1671.7 417.4
2088.6
= 0.6005
56 TOPIC 6. FIRST LAW FOR CLOSED SYSTEMS
and then calculate the corresponding v
2
as
v
2
= v
f2
+ x
2
(v
g2
v
f2
)
= 1.0434 10
3
+ 0.6005 (1693.7 10
3
1.0434 10
3
)
= 1.017 m
3
kg
1
which is too high. If we try again using p
2
= 1.1 bar, we get x
2
= 0.5974 and
v
2
= 0.926 m
3
kg
1
, which is too low. Interpolating linearly between 1.0 bar and 1.1
bar gives p
2
= 1.044 bar and x
2
= 0.599 and the temperature has dropped from
the saturation temperature of 179.88
o
C at 10 bar to the saturation temperature of
100.8
o
C at 1.044 bar.
6.4.5 Example 5
initial final
p
1
V
1
m
M
p
V
2
2
B
C

Figure 6.12: System for example 5
Problem: An ideal gas of mass m initially at volume V
1
and pressure p
1
, equal to
atmospheric pressure p
0
, is contained in a frictionless cylinder. The piston of area
A is linked by a exible cord to a drum of radius B with a mass M on the end of an
arm of length C as shown in gure 6.12. What heat must be removed from the gas
to raise the arm from the vertical to the horizontal?
Solution: First consider the gas to be the system. The First Law gives
Q W = U = m(u
2
u
1
)
Q =
2
_
1
p dV + mc
v
(T
2
T
1
)
and the ideal gas law pV = mRT gives
mT =
pV
R
6.5. POLYTROPIC PROCESSES 57
1
2
p
V
Figure 6.13: p V diagram for example 5
so that
Q =
2
_
1
p dV +
c
v
R
(p
2
V
2
p
1
V
1
)
For a rotation of the arm the volume V is
V = V
1
AB or V
2
= V
1
AB

2
and dV = AB d
and by statics the pressure p is
p = p
0

C
AB
Mg sin and p
2
= p
0

C
AB
Mg for =

2
so that, from gure 6.13 we see
2
_
1
p dV =

_
0
(p
0

C
AB
Mg sin) AB d = p
0
AB CMg(cos 1)
= p
0
AB

2
+ CMg for = /2
Hence
Q = p
0
AB

2
+ CMg +
c
v
R
_
(p
0

C
AB
Mg)(V
1
AB

2
) p
0
V
1
_
=
_c
v
R
+ 1
_
p
0
AB

2

c
v
R
CMg
_ V
1
AB


2
_
The work W is negative because work is done on the system. The heat transfer Q
is negative because heat is removed from the system.
6.5 Polytropic Processes
6.5.1 Polytropic Equation
A polytropic process is a process obeying the law
pV
n
= constant
58 TOPIC 6. FIRST LAW FOR CLOSED SYSTEMS
p
V
pV =
n
constant
pV = constant
n >1
Figure 6.14: Polytropic processes on p V diagram
where n is called the polytropic index. Figure 6.14 shows the form of this
relationship on a p V diagram. Many processes approximate to a reversible law
of this sort in practice. Usually n 1.
If the uid is an ideal gas then
pV
T
= constant
and by eliminating p or V we can get other useful forms
p
n1
T
n
= constant and T V
n1
= constant
where T is the absolute temperature.
If the change is reversible, then the work done for a polytropic process is
W =
2
_
1
p dV =
2
_
1
C
V
n
dV =

CV
1n
1 n
_
2
1
=
[pV ]
2
1
1 n
so
W =
_
_
_
_
_
_
_
p
2
V
2
p
1
V
1
1 n
for n = 1
p
1
V
1
ln
V
2
V
1
for n = 1
These apply to gases, and to such uids as steam-water mixtures. For an ideal
gas we can substitute pV = mRT and these become
W =
_
_
_
_
_
_
_
mR(T
2
T
1
)
1 n
for n = 1
mRT
1
ln
V
2
V
1
for n = 1
All these and other related equations are in HLT page 86.
6.5. POLYTROPIC PROCESSES 59
6.5.2 Example 6
Q
W
Figure 6.15: System for example 6
Problem: 1 kg of air expands from 2 bar, 25
o
C to 1 bar with pV
1.1
= constant as
shown in gure 6.15. Calculate W, Q and U. Take R = 287 J kg
1
K
1
.
Solution: Given p
1
= 2 bar, p
2
= 1 bar, T
1
= 298 K. Polytropic expansion gives
p
n1
1
T
n
1
=
p
n1
2
T
n
2
hence
T
2
= T
1
_
p
2
p
1
_
(n1)/n
= 298
_
1
2
_
(1.11)/1.1
= 280 K.
Note we are using the absolute temperature. Assuming air behaves as an ideal
gas we get
W =
mR
1 n
(T
2
T
1
) =
1 287
1 1.1
(280 298) = +52000 J
U = mc
v
(T
2
T
1
) = 1 718 (280 298) = 13000 J
Q = U + W = 13000 + 52000 = +39000 J
Hence W = +52 kJ, Q = +39 kJ and U = 13 kJ. Note that work W is done by
the system so is positive, and Q is heat transfer to the system so is positive.
6.5.3 Eect of Polytropic Index
How do the results vary with the polytropic index n? Repeating the calculation
with dierent values of n gives
60 TOPIC 6. FIRST LAW FOR CLOSED SYSTEMS
n T
2
W U Q
K kJ kJ kJ
1.0 298 +59 0

+59
1.1 280 +52 13 +39
1.2 265 +47 24 +23
1.3 254 +42 32 +10
1.4 244 +38 38 0

1.5 237 +35 44 9


The values marked

are clearly special cases.
n = 1.0 is called an isothermal process because T = 0
n = 1.4 is called an adiabatic process because Q = 0.
These denitions and the signicance of n = 1.4 for air will be discussed later.
Note that W is a maximum for n = 1. Thus if we want the maximum work out of the
expansion we want an isothermal process. However, this can be dicult to achieve
in practice as a relatively large amount of heat must be transferred to the gas as it
expands to maintain the temperature constant, and heat transfer takes time.
6.5.4 Polytropic Processes in Cycles
A
B
C
D
p
V
Figure 6.16: Cycle on a p V diagram
A cycle can be made up of a number of dierent polytropic processes. Consider
such a cycle represented on the p V diagram of gure 6.16. For example, if the
following cycle goes clockwise A-B-C-D, the area enclosed on the p V diagram is
equal to the net work output, that is the work out minus the work in. This cycle
converts heat into work.
If we let processes A B and C D be isothermal, and processes B C and
6.5. POLYTROPIC PROCESSES 61
D A be adiabatic, then for the cycle we nd
Process W Q
A B ve ve
B C ve 0
C D +ve +ve
D A +ve 0
Total +ve +ve
p V diagrams like this can be drawn for a real internal combustion engine, as in
gure 6.17. We shall examine these diagrams in more detail later in section 9.2.2.
The area enclosed by the graph is equal to the net work output. If the cycle is
reversed, and goes anti-clockwise, then the cycle converts work into heat.
p
V
Figure 6.17: Internal combustion engine cycle
62 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
Topic 7
First Law for Open Systems
7.1 Introduction
Many thermodynamic machines may be modelled as open systems. The First Law
of thermodynamics that we met in the previous chapter for closed systems is not
readily applicable to such machines, so we must recast it into a new form, the
steady ow energy equation. We will also meet a new property, enthalpy, which is
particularly useful for these calculations.
7.2 Steady Flow in Open Systems
7.2.1 Steady Flow Energy Equation
For a closed system the First Law states
Q W = U
where U is the internal energy of the system due to thermal motion of the
molecules. We can extend this concept to include the total energy E of the
system, comprising internal, gravitational and gross kinetic energy, so that
E = U + mgz +
1
2
mc
2
where
z = height above an arbitrary level m
c = bulk velocity of the uid ms
1
.
Thus gz J kg
1
(or kJ kg
1
) is the potential energy per unit mass, and
1
2
c
2
J kg
1
(or kJ kg
1
) is the gross kinetic energy per unit mass. Usually in closed systems
the initial and nal values of z are equal, and the values of c are zero, so that the
total energy E is the same as the internal energy U.
Consider an open system with a steady mass ow m kg s
1
into and out of the
system boundary, as illustrated in gure 7.1. The heat ow rate into the system is
7.2. STEADY FLOW IN OPEN SYSTEMS 63
Q
c
1
p v
1 1
,
W
c
2
p
2
v ,
2
z z
1 2
s
. .
m
.
m
.
Figure 7.1: Steady ow system

Q J s
1
(or W) (or kW) and the shaft work out of the system is

W
s
J s
1
(or W) (or
kW).
Now consider what happens during one second. An element of uid A is pushed
into the system, and an element of uid B is expelled as shown in gure 7.2. The
system boundary can be analysed as a closed system but with a moving
boundary.
The First Law for a closed system gives

Q

W =

E = me
Note that

W is the total mechanical power consisting of
the shaft work

W
s
done by the system
the work done on the system by the surroundings to push uid A into the
system, given by mv
1
p
1
where the sign occurs because it is work done on
the system, mv
1
is the volume of uid pushed into the system in that one
A
B
initial final
system boundary
Figure 7.2: Initial and nal boundaries
64 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
second, and p
1
is the constant pressure pushing on the system boundary
the work done by the system on the surroundings to push uid B out of the
system, given by + mv
2
p
2
.
Hence

Q

W =

Q (

W
s
mp
1
v
1
+ mp
2
v
2
) = me
where the pv terms are referred to as the ow work.
Therefore the First Law equation can be written as the steady ow energy
equation, or SFEE

Q

W
s
= m
_
u + pv + gz +
1
2
c
2
_
2
1
or more simply as

Q

W
s
= m
_
u + pv + gz +
1
2
c
2
_
where u is the internal energy, pv can be thought of as a pressure energy, gz is
the gravitational energy, and
1
2
c
2
the kinetic energy, all per unit mass. If we divide
by the mass ow rate m, we get the specic form of the SFEE
q w
s
=
_
u + pv + gz +
1
2
c
2
_
Note that the work terms in the SFEE are shaft work w
s
and W
s
.
7.2.2 Enthalpy
The group of terms u + pv often occurs in thermodynamic problems. We now
dene the enthalpy H J (or kJ) or specic enthalpy h J kg
1
(or kJ kg
1
) as
H = U + pV or h = u + pv
Hence we can rewrite the SFEE in even more compact and useful form as

Q

W
s
= m
_
h + gz +
1
2
c
2
_
or
q w
s
=
_
h + gz +
1
2
c
2
_
The enthalpy, like the internal energy, is a property of state, because u, p and v
are properties.
7.2. STEADY FLOW IN OPEN SYSTEMS 65
7.2.3 Specic Heat at Constant Pressure
In section 6.3.3 we saw that c
v
is related to u by
c
v
=
_
u
T
_
v
Following the same procedure, we can represent the enthalpy of the gas in our
system as a function of two properties, in this case T and p, in the form
h = h(T ,p) (7.1)
although we dont know yet what the function is. Dierentiating gives
dh =
_
h
T
_
p
dT +
_
h
p
_
T
dp (7.2)
p
0
dQ
m
p
Figure 7.3: Heating closed system at constant p
Consider a closed system being heated at constant pressure as shown in gure
7.3. If we add heat dQ = mdq, then work of dW = p dV or dw = p dv will be
done by the system on the surroundings as the system expands to maintain the
pressure constant, while the temperature will rise by dT .
The First Law for a closed system can be written in dierential form as
dq dw = du where dw = p dv , hence
dq = du + p dv
By denition h = u + pv . Dierentiating gives
dh = du + d( pv ) = du + p dv + v dp
but for constant pressure dp = 0 and this simplies to
dh = du + p dv
66 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
from which we see that for heating at constant pressure dq = dh. Substituting
dq = dh and dp = 0 into equation 7.2 gives
dq =
_
h
T
_
p
dT
or
_
q
T
_
p
=
_
h
T
_
p
But (q/T )
p
is a specic heat , the amount of heat which must be added to 1 kg
of mass to raise the temperature by 1
o
C. Because p is constant, this is c
p
, the
specic heat at constant pressure, with units of J kg
1
K
1
(or kJ kg
1
K
1
), so h
and c
p
are related by
c
p
=
_
h
T
_
p
We will prove next year that for an ideal gas (h/p)
T
= 0 so equation 7.1
reduces to h = h(T ). That is, for an ideal gas the enthalpy is a function of
temperature only. The partial derivative above becomes a total derivative
c
p
=
dh
dT
and hence for an ideal gas we get the very important result that
h = c
p
T
whether p is constant or not .
7.2.4 Relationship Between Specic Heats
If we substitute the ideal gas law pv = RT into the denition of enthalpy
h = u + pv we get
h = u + RT
Dierentiating gives
dh = du + R dT
We know from sections 6.3.3 and 7.2.3 that for an ideal gas
du = c
v
dT and dh = c
p
dT
hence
c
p
dT = c
v
dT + R dT
7.2. STEADY FLOW IN OPEN SYSTEMS 67
from which we see that for an ideal gas the two specic heats are related by
c
p
c
v
= R
It is sometime convenient to use the ratio of specic heats dened by
=
c
p
c
v
from which it can be shown for ideal gases that
c
v
=
R
1
and c
p
=
R
1
7.2.5 Negligible Terms in the SFEE
In the SFEE
q w
s
=
_
h + gz +
1
2
c
2
_
the changes in the gravitational and kinetic energy terms are often small. If no
information is given, or required, about heights or velocities, these terms may be
neglected so that the SFEE becomes simply
q w
s
= h
We will see examples of the use of this simplication later in this chapter.
7.2.6 Work in a Reversible Steady Flow Process
In a closed system the reversible work is W =
_
p dV , but what is the
corresponding expression for a steady ow process in an open system? When we
derived the SFEE
q w
s
=
_
h + gz +
1
2
c
2
_
we used
w = w
s
p
1
v
1
+ p
2
v
2
where w was the work for a closed system, w
s
was the shaft work, and
p
1
v
1
+ p
2
v
2
was the pv or ow work required to push uid into the system and
work done when the system pushed uid out, or
w
s
= w + p
1
v
1
p
2
v
2
or p
2
v
2
p
1
v
1
= w + (w
s
)
68 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
p
v
1
2
p
v
1
2
p
v
1
2
w
p v
1 1
p v
2 2
-w
s
(a) (b) (c)
Figure 7.4: Areas on p v diagram
This can be interpreted graphically if we recognise that
w =
2
_
1
p dv
is the area under a p v curve, between the curve and the v axis of gure 7.4(a).
Now p
1
v
1
and p
2
v
2
are the two rectangles in gure 7.4(b), so that the shaft work w
s
is given by minus the area between the curve and the p axis of gure 7.4(c), or
analytically by
w
s
=
2
_
1
v dp
7.3 Steady Flow Devices
7.3.1 Compressor
p > p
2 1
C
1
2
W
s
m
.
.
Figure 7.5: Compressor
A compressor or pump is a device for increasing the pressure of a uid, illustrated
schematically in gure 7.5. The term compressor is generally used for gases, and
the term pump is used for liquids. In a centrifugal compressor a rotating impeller
ings the uid radially outward, thus increasing its pressure. In an axial
7.3. STEADY FLOW DEVICES 69
compressor, the uid is pushed by blades on a rotating wheel through stationary
blades into a smaller space. A reciprocating compressor operates in a dierent
fashion.
Consider a compressor which takes uid at pressure p
1
and delivers it at pressure
p
2
where p
2
> p
1
. If we assume that
the compressor is adiabatic
kinetic and potential energy changes are negligible
then

Q = 0, z = 0 and c = 0 so that the SFEE gives
0

W
s
= m(h
2
h
1
+ 0 + 0)
hence

W
s
= m(h
2
h
1
) or w
s
= (h
2
h
1
)
The rst assumption is reasonable because in many compressors with a high ow
rate there is little time for signicant heat transfer. The second assumption is
reasonable because in practice the changes are often negligible compared with
the change in enthalpy h.
If the uid is an ideal gas, then we shall see next year that h = c
p
T so

W
s
= mc
p
(T
2
T
1
) or w
s
= c
p
(T
2
T
1
)
Note that the compressor work is negative as it is done by the surroundings on the
system.
7.3.2 Turbine
1
2
W
s
m
.
2
p > p
1
T
.
Figure 7.6: Turbine
A turbine is a device for extracting work from a high pressure uid, illustrated
schematically in gure 7.6. The basic principle is that the uid expands through
70 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
stationary nozzles or blades and impacts against blades mounted on a rotating
wheel. This process can be repeated many times in stages. At each stage the uid
loses some of its original pressure. Common examples are the steam turbine and
the gas turbine.
Consider a turbine which takes uid at pressure p
1
and exhausts it at pressure p
2
where p
1
> p
2
. If we assume that
the turbine is adiabatic
kinetic and potential energy changes are negligible
then

Q = 0, z = 0 and c = 0 so that the SFEE gives
0

W
s
= m(h
2
h
1
+ 0 + 0)
hence

W
s
= m(h
1
h
2
) or w
s
= h
1
h
2
If the uid is an ideal gas, then h = c
p
T so

W
s
= mc
p
(T
1
T
2
) or w
s
= c
p
(T
1
T
2
)
Note that the turbine work is positive as it is done by the system on the
surroundings.
7.3.3 Throttle
1 2 p
1
> p
2
Figure 7.7: Throttle
A throttle is a deliberate restriction placed in a owing uid to decrease its
pressure or restrict its ow rate, illustrated schematically in gure 7.7.
Consider a gas owing steadily through a throttle such as a porous plug or an
adjustable gate valve or buttery valve. The pressure falls across the throttle from
p
1
to p
2
. There is no shaft so no shaft work. If we assume that
the throttle is adiabatic
kinetic and potential energy changes are negligible
7.3. STEADY FLOW DEVICES 71
then

Q = 0 z = 0 and c = 0 and

W
s
= 0 so that the SFEE gives
h = constant or h
1
= h
2
For an ideal gas, u = u(T ) only, but by denition h = u + pv , so for an ideal gas
h = u +RT , hence h = h(T ) only, so if h = constant then T = constant, or T
1
= T
2
.
For real gases T
1
= T
2
, and this experiment turns out to be a very sensitive test for
an ideal gas. At room temperature
for most gases T
2
< T
1
and the gas is cooled
for hydrogen and helium T
2
> T
1
and the gas is heated.
However if hydrogen or helium is cooled to below its inversion temperature, then
cooling will take place. This is known as the Joule-Thomson experiment. The
Joule-Thomson coecient Kbar
1
(or KPa
1
) given by
=
_
T
p
_
h
is therefore positive for most gases. For example, for nitrogen at 1 bar 25
o
C,
= +0.22 Kbar
1
. The inversion temperature for nitrogen is 620 K.
The Joule-Thomson or Joule-Kelvin eect (so-called because W.H.Thomson later
became Lord Kelvin) can be used to liquefy gases.
7.3.4 Nozzle
1 2
1 2
p > p
c
2
> c
1
Figure 7.8: Nozzle
A nozzle is a duct of varying cross-sectional area designed so that a drop in
pressure from inlet to outlet accelerates the ow, illustrated schematically in gure
7.8.
Consider a nozzle which takes uid at pressure p
1
and discharges it at pressure p
2
where p
1
> p
2
. There is no shaft work. If we assume that
the nozzle is adiabatic
potential energy changes are negligible
72 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
then

Q = 0, z = 0 and

W
s
= 0 so that the SFEE gives
0 0 = m(h
2
+
1
2
c
2
2
h
1

1
2
c
2
1
)
h
1
h
2
=
1
2
(c
2
2
c
2
1
)
enabling the conversion of enthalpy into kinetic energy. If the uid is an ideal gas,
then h = h(T ) only, therefore h = c
p
T and hence
c
p
(T
1
T
2
) =
1
2
(c
2
2
c
2
1
)
7.3.5 Example 1
T
1
c
2
T
2
air
p
0
Figure 7.9: System for example 1
Problem: Air from a large reservoir at 298 K, 2 bar expands adiabatically through a
nozzle to 1 bar, as shown in gure 7.9. The exit temperature is 255 K. Calculate
the air velocity at exit. Assume air is an ideal gas. Use c
p
= 1005 J kg
1
K
1
.
Solution: We are given that the expansion is adiabatic so q = 0. There is no shaft
work so w
s
= 0. Assume no height changes so gz = 0. Hence the SFEE gives

_
h +
1
2
c
2
_
= 0 or h
1
+
1
2
c
2
1
= h
2
+
1
2
c
2
2
The reservoir is large so the initial velocity will be eectively c
1
= 0, so
c
2
=
_
2(h
1
h
2
)
If c
p
= constant, then for an ideal gas h = h(T ) only and so h = c
p
T giving
c
2
=
_
2c
p
(T
1
T
2
) =
_
2 1005 (298 255) = 294 ms
1
Next year we will be able to show that this is not a reversible process. However the
SFEE is applicable to any steady ow process.
7.4. ENTHALPY AND FIRST LAW FOR CLOSED SYSTEMS 73
7.4 Enthalpy and First Law for Closed Systems
7.4.1 Closed Systems with Moving Boundaries
When we began to use the First Law for closed systems, we encountered the
internal energy. It was only when we started dealing with open or steady ow
systems that we met enthalpy. However, there are many examples of a closed
system with a moving boundary where we will nd enthalpy terms appearing in the
equations we derive from the First Law because of the work done when the
boundary moves against uid pressure. Although the system is technically a
closed system, we will often nd it easier to work with h terms than with a larger
number of u and pv terms.
7.4.2 Example 2
100 mm
p
1
T
1
p
2
T
2
initial final
1 kW
p
0
Figure 7.10: System for example 2
Problem: An insulated cylinder of cross-section area 0.2 m
2
contains a frictionless
piston free to rise vertically, resting on 1 kg of water at 20
o
C, as shown in gure
7.10. The piston exerts a constant pressure of 5 bar. An electrical heater rated at 1
kW at the bottom of the cylinder is used to raise the piston by 100 mm. How long
will it take?
Solution: The system is closed, and consists of the uid only. The First Law for a
closed system gives
Q W = U = U
2
U
1
There is no friction and the expansion is slow, so the work done by the system is
reversible. Also the pressure p is constant so
74 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
p
v
5 bar
20 C
we are here
use and here v h
o
Figure 7.11: Obtaining properties for subcooled liquid
W =
2
_
1
p dV = p(V
2
V
1
)
so that
Q = p(V
2
V
1
) + U
2
U
1
= (U
2
+ pV
2
) (U
1
+ pV
1
) = H
2
H
1
= m(h
2
h
1
)
Note that enthalpy terms appear even though the system is closed.
State 1 is compressed liquid water 20
o
C, 5 bar. This is in the subcooled region,
which isnt tabulated in HLT. However, properties for subcooled liquid are
determined mainly by the temperature, not the pressure, therefore we can get
good approximations by using the saturated liquid values at 20
o
C, not the values
at 5 bar, as illustrated in gure 7.11. Hence from HLT pages 4849 we get
h
1
h
f20
= 83.9 kJ kg
1
and v
1
v
f20
= 1.0017 l kg
1
= 1.0017 10
3
m
3
kg
1
.
The initial volume is
V
1
= mv
1
= 1 1.0017 10
3
= 1.0017 10
3
m
3
If A is the piston area and d is the distance the piston is raised, then the nal
volume at state 2, as shown in gure 7.12, is
V
2
= V
1
+ Ad = 1.0017 10
3
+ 0.2 0.1 = 21.002 10
3
m
3
p
V
1 2
Figure 7.12: p V diagram for example 2
7.5. SFEE EXAMPLES 75
Hence v
2
= V
2
/m = 21.002 10
3
m
3
kg
1
= 21.002 l kg
1
.
At 5 bar we get from HLT pages 5253 v
f
= 1.0928 l kg
1
and v
g
= 374.66 l kg
1
.
Note that v
f
< v
2
< v
g
conrming that state 2 is wet steam, hence T
2
= 151.85
o
C.
Now nd x
2
from
v
2
= v
f2
+ x
2
(v
g2
v
f2
) = 1.0928 + x
2
(374.66 1.0928) = 21.002 l kg
1
x
2
= 0.0533
hence
h
2
= h
f2
+ x
2
h
fg2
= 640.1 + 0.0533 2107.4 = 752.4 kJ kg
1
Q = m(h
2
h
1
) = 1 (752.4 83.9) = 668 kJ
The heating time at 1 kW is therefore 668 seconds or 11.1 minutes.
7.5 SFEE Examples
7.5.1 Example 3
1
2
pump
w
in
Figure 7.13: System for example 3
Problem: A steady ow pump compresses liquid water at 20
o
C from 0.0233 bar to
100 bar, as shown schematically in gure 7.13. The pump is reversible. Calculate
the work required per kg of water. If the pump is also adiabatic, calculate the
change in specic enthalpy of the water.
Solution: Liquid water is almost incompressible, so on the p v diagram of gure
7.14 the line from 1 to 2 is almost vertical. At low temperatures the specic volume
of water is eectively constant and equal to v = 1 l kg
1
= 10
3
m
3
kg
1
. If the
process is reversible, then the shaft work is
w
s
=
2
_
1
v dp v ( p
2
p
1
)
We must be careful of the units. If v is in m
3
kg
1
and p is in Nm
2
or Pa, then w
s
will be in J kg
1
. HLT gives v in l kg
1
and p in bar. We must remember to convert
76 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
2
1 v
p
Figure 7.14: p v diagram for example 3
v from l kg
1
into m
3
kg
1
, and p from bar into Pa. Here p
1
= 0.0233 bar = 2330
Pa, and p
2
= 100 bar = 10
7
Pa. Hence
w
s
10
3
(10
7
2330) = 9997.67 J kg
1
10 kJ kg
1
This is the shaft work done by the system. The shaft work done on the system is
w
in
= +9997.67 J kg
1
+10 kJ kg
1
For an adiabatic pump, we can calculate h using the SFEE. Ignoring velocity and
height changes
q w
s
= h
Here q = 0 and w
s
= 10 kJ kg
1
so
h = h
2
h
1
= 0 (10) = +10 kJ kg
1
Also the actual enthalpies can be found. Water at 20
o
C, 0.0233 bar is saturated,
and from the saturation table in HLT pages 4849 we get h
1
= 83.9 kJ kg
1
,
therefore h
2
= 83.9 + 10 = 93.9 kJ kg
1
.
Note that for liquid water, like any substance, the enthalpy is h = h(T ,p) but the
enthalpy, like the specic volume, is almost independent of the pressure. Hence a
large change in pressure, a factor of > 4000 in this example, produces only a small
change in enthalpy.
7.5.2 Example 4
Problem: A boiler takes in water at approximately 20
o
C and 100 bar, heats it to
boiling temperature, vaporises it, and superheats the steam to 500
o
C, as shown in
gure 7.15. The boiler operates at constant pressure, so on the p v diagram of
gure 7.16 the line 2 to 3 is horizontal. Calculate the heat input per kg of uid.
7.5. SFEE EXAMPLES 77
2 3
water steam
boiler
q
in
Figure 7.15: System for example 4
Solution: The boiler does no work, so w
s
= 0, and the SFEE gives
q = h = h
3
h
2
From the previous example, h
2
= 93.9 kJ kg
1
, and from the superheat table in HLT
page 64 at 100 bar, 500
o
C we get h
3
= 3374.6 kJ kg
1
, so the heat input is
q
in
= 3374.6 93.9 = 3280.7 kJ kg
1
3
2
v
p
Figure 7.16: p v diagram for example 4
7.5.3 Example 5
Problem: Steam at 500
o
C and 100 bar is passed through an adiabatic turbine and
exhausts at a pressure of 0.0233 bar and a dryness fraction of 0.8, as shown in
gure 7.17. Calculate the shaft work produced by the turbine per kg of uid.
Solution: The turbine is adiabatic, so q = 0, and the SFEE gives
w
s
= h
4
h
3
or w
s
= h
3
h
4
3
4
w
turbine
out
Figure 7.17: System for example 5
78 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
3
4
v
p
Figure 7.18: p v diagram for example 5
On the p v diagram of gure 7.18, the line forms a gentle curve from 3 to 4. From
the previous example, h
3
= 3374.6 kJ kg
1
. We can write h
4
as
h
4
= h
f4
+ x
4
h
fg4
where h
f4
is the saturation enthalpy of liquid, h
fg4
is the enthalpy change on
vaporisation and x
4
is the dryness fraction. From HLT pages 4849 at 0.0233 bar
we nd h
f4
= 83.9 kJ kg
1
and h
fg4
= 2454.3 kJ kg
1
so that
h
4
= 83.9 + 0.8 2454.3 = 2047.3 kJ kg
1
hence
w
s
= h
3
h
4
= 3374.6 2047.3 = 1327.3 kJ kg
1
and so the work done by the system is
w
out
= 1327.3 kJ kg
1
7.5.4 Example 6
5
steam water
4
q
condenser
out
Figure 7.19: System for example 6
Problem: Wet steam at 0.0233 bar and dryness 0.8 is condensed to give saturated
water at 20
o
C, as illustrated in gure 7.19. Calculate the heat released by the
condensation per kg of uid.
Solution: The wet steam is condensed in a condenser . Typically the wet steam
impacts against a surface cooled by cold water. No shaft work is done, so w
s
= 0,
7.6. CYCLES 79
4
5 v
p
Figure 7.20: p v diagram for example 6
and the SFEE gives
q = h
5
h
4
On the p v diagram of gure 7.20, the line from 3 to 4 is horizontal. From the
previous examples, h
4
= 2047.3 kJ kg
1
and h
5
= 83.9 kJ kg
1
. Note that the
saturation pressure at 20
o
C is 0.0233 bar. Hence
q = h
5
h
4
= 83.9 2047.3 = 1963.4 kJ kg
1
This is the heat transfer to the system. Thus the heat released by the
condensation and taken out of the system is
q
out
= +1963.4 kJ kg
1
7.6 Cycles
7.6.1 Rankine Cycle
It will be obvious that the last four examples describe processes which can be
joined to form a cycle. This cycle, called the Rankine cycle, is illustrated in gure
7.21, and is the basis of most electrical generating stations, such as the Didcot
power station. It converts heat into work.
Note the net values
net work output w
net
= w
out
w
in
= 1327.3 10.0 = 1317.3 kJ kg
1
net heat input q
net
= q
in
q
out
= 3280.7 1963.4 = 1317.3 kJ kg
1
and for the whole cycle, w
net
= q
net
as h
cycle
= 0 so
q
in
q
out
= w
out
w
in
or q
in
+ w
in
= q
out
+ w
out
80 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
feed
pump
boiler
condenser
turbine
4
3 2
1
w
out
q
in
w
in
q
out
Figure 7.21: Flow diagram for Rankine cycle
The First Law of thermodynamics implies that the maximum possible eciency of
converting heat into work is 100 %. Actually, in practical terms, the eciency is
much less, because the heat rejected to the condenser q
out
is not useful heat as it
is rejected at the low temperature of 20
o
C. On the other hand, the heat to the
boiler q
in
must be supplied at high temperatures, up to 500
o
C. Thus the practical
thermodynamic eciency is dened as
=
net work out
heat in
=
w
net
q
in
=
w
out
w
in
q
in
=
q
in
q
out
q
in
= 1
q
out
q
in
which in this example gives
=
w
net
q
in
=
1327.3 10.0
3280.7
= 0.401 = 40.1 %
Typical values for the thermodynamic eciency of a large modern power station
are around 42 or 43 %.
3
4
2
1
v
p
Figure 7.22: Rankine cycle on p v diagram
It is useful to represent the Rankine cycle on a p v diagram as in gure 7.22.
Label the end points of each process clearly and mark arrows to show the direction
of the cycle. Note the arrows go clockwise around the cycle. The processes are
1 2 feed pump, 2 3 boiler, 3 4 turbine and 4 1 condenser. Note that in
7.6. CYCLES 81
the boiler and the condenser, the pressure remains constant. The feed pump line
is almost vertical as the specic volume v decreases only slightly from 1 2.
Point 1 lies on the saturated liquid line, point 2 is in the subcooled region, point 3 in
the superheat region, and point 4 inside the saturated liquid and vapour region.
7.6.2 Heat Engines and Heat Pumps
T
H
T
C
Q
H
Q
C
W
hot reservoir
cold reservoir
heat engine
Figure 7.23: Schematic diagram for heat engine
A machine or cycle which converts heat into work is called a heat engine and is an
important concept in thermodynamics. A heat engine must operate between a
heat source or hot reservoir at a high temperature T
H
from which it removes a
quantity of heat Q
H
, rejecting a quantity of heat Q
C
to a sink or cold reservoir at a
low temperature T
C
and performing net work W. This is illustrated schematically in
gure 7.23. The thermodynamic eciency of the heat engine is
=
W
Q
H
=
Q
H
Q
C
Q
H
= 1
Q
C
Q
H
= 1
heat out
heat in
The heat and work Q and W could of course be written in terms of specic
quantities of heat and work q and w. For example, in the Rankine cycle
considered earlier, we saw
q
H
= boiler heat input = 3280.7 kJ kg
1
q
C
= condenser heat output = 1963.4 kJ kg
1
w = net work output = turbine work output feed pump work input
= 1327.3 10.0 = 1317.3 kJ kg
1
The hot reservoir temperature T
H
would be the temperature of, for example, the
burning coal. To heat the steam to 500
o
C would require T
H
500
o
C. Similarly, the
82 TOPIC 7. FIRST LAW FOR OPEN SYSTEMS
T
H
T
C
Q
H
Q
C
W
hot reservoir
cold reservoir
heat pump
Figure 7.24: Schematic diagram for heat pump
cold reservoir temperature would be the temperature of, for example, the water in
the evaporative cooling tower. To condense the wet steam leaving the turbine to
saturated water at 20
o
C would require T
C
20
o
C.
A reversible heat engine is one where all the heat and work ows can be reversed
in direction without changing their magnitude. It can be shown, after we have
studied the Second Law next year, that the turbine in the example of the Rankine
cycle in section 7.6.1 is not reversible, and so the whole cycle is irreversible. You
will also study the Carnot cycle, which delivers the maximum possible eciency.
A reversed heat engine is called a heat pump or a refrigerator , illustrated
schematically in gure 7.24. We would generally use the name heat pump when
we want to deliver heat Q
H
to a high temperature reservoir T
H
, and refrigerator
when we want to remove heat Q
C
from a low temperature reservoir T
C
.
The analogue of the eciency for a heat pump is the coecient of performance or
COP. The COP for a heat pump and for a refrigerator are dened as
COP
heat pump
=
Q
H
W
COP
refrigerator
=
Q
C
W
Both eciency and coecient of performance can be thought of as
what you want
what you pay
However, the eciency is always < 1, whereas it is quite usual to nd COP > 1.
TOPIC 8. COMBUSTION 83
Topic 8
Combustion
8.1 Introduction
In this chapter we examine some of the chemical reactions involved in the
combustion of fuels in air. Air is composed primarily of nitrogen, oxygen and
carbon dioxide. The nitrogen and carbon dioxide do not take part in the
combustion reaction itself, but their presence does aect the nal temperature of
the combustion products. We will meet molar specic heats, and the enthalpy of
reaction.
8.2 Chemical Reactions
8.2.1 Equations
For reacting mixtures the use of molar quantities is invaluable since reactions
occur between integral numbers of molecules, and the mole is a unit quantity of
molecules. The kmol, as introduced in section 2.4, is the amount of substance in a
system that contains as many elementary entities as there are atoms in 12 kg of
carbon 12. Recall that the number of atoms or molecules in 1 kmol is given by
Avogadros number
N = 6.022 10
26
kmol
1
Consider the reaction between two molecules of carbon monoxide CO and one
molecule of oxygen O
2
to produce two molecules of carbon dioxide CO
2
2CO+ O
2
2CO
2
There is conservation of mass and also conservation of the number of atoms.
It is often convenient to consider a unit quantity of fuel, for instance 1 kmol, so the
above reaction can be written in terms of kmol as
CO+
1
2
O
2
CO
2
84 TOPIC 8. COMBUSTION
A further advantage of using kmol is that, for the same conditions of temperature
and pressure, equal volumes of gas contain the same number of kmol. Thus the
volumetric composition of a gas mixture is also the molar composition of the gas
mixture. This is obviously not the case if liquid or solids are also present.
Although mass is conserved, it is not possible merely by inspection to write down
the masses of each constituent, but it is necessary rst to write down the reaction
in terms of kmol. If we take the molar masses of C and O to be 12 and 16
kg kmol
1
respectively, we get
CO +
1
2
O
2
CO
2
kmol 1
1
2
1
kg 12 + 16
1
2
(2 16) 12 + 16 + 16
As well as the forward reaction of carbon monoxide oxidising to carbon dioxide,
there will be a reverse reaction of carbon dioxide dissociating to carbon monoxide
and oxygen
CO+
1
2
O
2
CO
2
When equilibrium is attained, the mixture will contain all possible species from the
reaction, but in addition the oxygen can also dissociate
O
2
2O
However, equilibrium calculations (which derive from the Second Law of
thermodynamics) will not be developed here. Furthermore, there is not always
sucient time for equilibrium to be attained, since the rates of reaction, which
depend on the chemical kinetics, may be too slow compared with the time
available for the reaction. However, in all our calculations in this course we will
assume complete combustion.
8.2.2 Combustion in Air
Fuels are usually burnt in air, which has the following composition:
molar composition O
2
= 21.0 % N

2
= 79.0 %
gravimetric composition O
2
= 23.2 % N

2
= 76.8 %
Note that the term gravimetric means by weight . The term N

2
shown above is used
to denote atmospheric nitrogen which represents all the constituents of air except
oxygen and water vapour, notably carbon dioxide, argon and neon. Thus the true
8.2. CHEMICAL REACTIONS 85
molar composition of air is more like
N
2
= 78.08 % O
2
= 20.95 % Ar = 0.93 % CO
2
= 0.03 % Ne = 0.002 %
However, the thermodynamic properties of N

2
are usually taken to be those of
pure N
2
.
The molar masses (or molecular masses) for the key constituents are as follows,
with the usual approximations in brackets.
O
2
= 31.998(32) N
2
= 28.014(28) N

2
= 28.15 Air = 28.96(29)
When 1 kmol of CO is burnt in air the reaction in kmol is
CO+
1
2
(O
2
+
79
21
N
2
) CO
2
+
1
2
79
21
N
2
The N
2
must be kept in the equation even though it does not take part in the
reaction. It aects the volumetric composition of the products and, as will be seen
later, the combustion temperature. The molar or volumetric air/fuel ratio is
1 :
1
2
(1 +
79
21
) or 1 : 2.38
The gravimetric air/fuel ratio is found by multiplying the number of kmol by the
respective molar masses, 12 + 16 kg kmol
1
for CO and 29 kgkmol
1
for air to give
12 + 16 : 2.38 29 or 1 : 2.47
So far the reacting mixtures have been assumed to be chemically correct for
complete combustion. That is, the mixtures have been stoichiometric. In general,
reacting mixtures are non-stoichiometric. If the quantity of oxygen was below
stoichiometric, then there would be products of partial combustion present, notably
carbon monoxide CO and hydrogen H
2
. Excess air is often used in practice, so as
to ensure that combustion is as complete as possible.
Consider the same reaction as before with 25% excess air .
CO+ 1.25
1
2
(O
2
+
79
21
N
2
) CO
2
+ a O
2
+ 1.25
1
2
79
21
N
2
A temporary variable a has been introduced since it may not be immediately
apparent what quantity of oxygen will be left, but this can easily be found from an
oxygen balance. The carbon and nitrogen have already been balanced, and the
only pitfall is to mix atoms and molecules, which is less likely to occur if we use
86 TOPIC 8. COMBUSTION
chemical symbols. So, for an atomic oxygen balance
O balance 1 + 1.25
1
2
2 = 2 + 2a
which gives a = 1/8 so that the balanced equation is
CO+ 1.25
1
2
(O
2
+
79
21
N
2
) CO
2
+
1
8
O
2
+ 1.25
1
2
79
21
N
2
Reacting mixtures can also be dened in terms of the the equivalence ratio
dened as
=
stoichiometric air/fuel ratio
actual air/fuel ratio
which in the above example is = 1/1.25 = 0.8. Another way of describing the
mixture is the non-dimensional air/fuel ratio lambda or dened by
=
1

The air/fuel ratio can be either gravimetric or molar. The usual form is gravimetric
and this is often implicit. An exception is for gaseous fuels, for which it is common
to use a volumetric or molar air/fuel ratio.
8.2.3 Hydrocarbon Fuels
Most practical fuels are hydrocarbons that comprise hydrogen and carbon, and
these are the only ones to be analysed here. They occur as gases (for example,
methane CH
4
, ethane C
2
H
6
and ethene C
2
H
4
), liquids and solids (for example,
coal and wax). Some fuels contain oxygen, and these are known as oxygenates,
the commonest being methanol (CH
3
OH) and ethanol (C
2
H
5
OH).
Fuels are often mixtures of hydrocarbons, with bonds between carbon atoms, and
between hydrogen and carbon atoms. During combustion these bonds are broken,
and new bonds are formed with oxygen atoms, accompanied by a release of
chemical energy. The principal products are carbon dioxide and water. The
dierent compounds in fuels are classied according to the number of carbon
atoms in the molecules. The size and geometry of the molecule have a profound
eect on the chemical properties. Each carbon atom requires four bonds. These
can be single bonds or combinations of single, double and triple bonds. Hydrogen
atoms require a single bond.
8.2. CHEMICAL REACTIONS 87
8.2.4 Example 1
Problem: If a fuel mixture can be represented by the general formula C
x
H
2x
show
that the stoichiometric gravimetric air/fuel ratio is 14.8 : 1.
Solution: The molar (or volumetric) composition of air can be assumed to be 21%
oxygen and 79 % atmospheric nitrogen. Thus for every kmol of oxygen there are
79/21 = 3.76 kmol of atmospheric nitrogen. Our reaction, using a temporary
variable z for the amount of oxygen, is thus
C
x
H
2x
+ z (O
2
+ 3.76 N

2
) a CO
2
+ b H
2
O+ 3.76z N

2
We have already applied an atomic balance for nitrogen, and we continue with the
atomic balances for carbon, hydrogen and oxygen.
C balance x = a
H balance 2x = 2b
O balance 2z = 2a + b
Substituting the C and H balances into the O balance gives
2z = 2x + x z = 1.5x
The mass of 1 kmol of the fuel C
x
H
2x
is
x(12) + 2x(
1
2
2) = 14x kg
1 kmol of this fuel reacts with z kmol of oxygen, contained in air of mass
z (32 + 3.76 28.15) = 137.88z = 206.82x kg
Dividing the mass of air 206.82x kg by the mass of fuel 14x kg gives the
stoichiometric gravimetric air/fuel ratio of 206.82x/14x = 14.77.
8.2.5 Example 2
Problem: The dry exhaust gas analysis from an engine burning a hydrocarbon fuel
is as follows: CO
2
0.121, O
2
0.037, N

2
0.842. Determine the gravimetric
composition of the fuel, the equivalence ratio of the air/fuel mixture, and the
stoichiometric air/fuel ratio.
Solution: Assume that the hydrocarbon C
x
H
y
reacts with z kmol of oxygen to
produce 100 kmol of dry products of combustion. There will of course be water
88 TOPIC 8. COMBUSTION
present in the exhaust, so that our specied quantity of reactants produces more
than 100 kmol of products.
The molar or volumetric composition of air can be assumed to be 21 % oxygen and
79 % atmosepheric nitrogen. Thus for every kmol of oxygen there are
79/21 = 3.76 kmol of atmospheric nitrogen. Our reaction is thus
C
x
H
y
+ z (O
2
+ 3.76 N

2
) 12.1 CO
2
+ 3.7 O
2
+ 84.2 N

2
+ w H
2
O
As a cross-check, 12.1 + 3.7 + 84.2 = 100.0. Note that the w kmol of H
2
O do not
contribute to the 100 kmol of dry products. The oxygen level implies a weak
mixture, which is also consistent with there being no measurement of carbon
monoxide.
The four atomic balances are
N balance 3.76 2z = 2 84.2 z = 22.39
C balance x = 12.1
O balance 2z = 12.1 2 + 3.7 2 + w = 31.6 + w
H balance y = 2w
Substituting the value of z from the N balance into the O balance gives
2 22.39 = 31.6 + w w = 13.18
Substituting for the value of w from the H balance gives
y = 26.36
The molar composition of the fuel is thus C
12.1
H
26.36
. However it is only the C:H
ratio that has any meaning as we know nothing about the mean molar mass, since
these values are a consequence of assuming 100 kmol of dry products.
To nd the gravimetric composition of the fuel, we need to nd separately the
mass of the carbon and hydrogen. The mass of carbon is 12x = 145.2 kg, and the
mass of hydrogen is 1y = 26.36 kg, giving a total mass for the fuel of
145.2 + 26.36 = 171.56 kg. The mass fractions of C and H in the fuel are therefore
145.2/171.56 = 0.846 and 26.36/171.56 = 0.154 respectively.
To nd the equivalence ratio, dene z
s
as the quantity of air required for
stoichiometric combustion. Hence
C
12.1
H
26.36
+ z
s
(O
2
+ 3.76 N

2
) a CO
2
+ b H
2
O+ c N

2
8.3. ENERGY OF REACTIONS 89
The atomic balances are
C balance 12.1 = a
H balance 26.36 = 2b
O balance 2z
s
= 2a + b
from which we obtain z
s
= 18.69.
The equivalence ratio is the ratio of the quantity of air required for stoichiometric
combustion z
s
compared to that actually used z. Thus
=
z
s
z
=
18.69
22.39
= 0.835
The stoichiometric air/fuel ratio can be found from the mass of fuel, 171.56 kg,
and the mass of the stoichiometric quantity of air, given by
18.69 (32 + 3.76 28.15) = 2576.3 kg. Hence
AFR
s
=
2576.3
171.56
= 15.02
8.3 Energy of Reactions
8.3.1 Molar Specic Heats
For a gas occupying volume V m
3
at pressure p kPa, the enthalpy H kJ and
internal energy U kJ are related by
H = U + pV
If the gas is ideal, we can write the ideal gas law in terms of the number of kmol n
as
pV = n R
0
T
where R
0
kJ kmol
1
K
1
is the universal gas constant. Hence
H = U + n R
0
T
Until now we have been using the specic enthalpy h kJ kg
1
and specic internal
energy u kJ kg
1
. More accurately, these are mass specic, meaning that they tell
us the enthalpy and internal energy respectively per kg. Let us now dene the
molar specic enthalpy

h kJ kmol
1
and molar specic internal energy u kJ kmol
1
90 TOPIC 8. COMBUSTION
as

h =
H
n
and u =
U
n
which tell us the enthalpy and internal energy respectively per kmol of the gas. We
can therefore relate the enthalpy and internal energy by

h = u + R
0
T
Since we are restricting ourselves to ideal gases, in which the specic heats are
constant, then with a datum of zero internal energy or enthalpy at 0 K, we can also
dene molar specic heats as

h = c
p
T and u = c
v
T
so that
c
p
c
v
= R
0
Note the similarities between these molar specic quantities and their mass
specic counterparts.
As reactions involve mixtures of gases, then obviously it is necessary to be able to
calculate the enthalpy or internal energy of mixtures. We will only deal with ideal
mixtures for which the enthalpy or internal energy are calculated by summing the
contributions from each of the original species. This can be done on either a mass
(gravimetric) or a molar basis
H =

m
i
h
i
and U =

m
i
u
i
H =

n
i

h
i
and U =

n
i
u
For ideal gases these become
H = T

m
i
c
pi
and U = T

m
i
c
vi
H = T

n
i
c
pi
and U = T

n
i
c
vi
Finally, inspection of the above equations shows that for a mixture of ideal gases
the heat capacities are given by
C
p
=

m
i
c
pi
=

n
i
c
pi
and C
v
=

m
i
c
vi
=

n
i
c
vi
8.3. ENERGY OF REACTIONS 91
Since reactions occur between molecules it is invariably easier to use the molar
equations than the gravimetric equations, even if it is a matter of converting the
gravimetric quantities to molar quantities by means of the molar mass M using
c
p
= M c
p
and c
v
= M c
v
To solve combustion problems, the additional information that is needed is how the
enthalpy or internal energy of the reactants and products varies with temperature.
For real gases, the specic heat capacities are not constant. However, we will
assume constant specic heats. Table 8.1 shows molar masses and molar specic
heats to be used in combustion calculations. The specic heats have been
evaluated from the change in enthalpy between 25
o
C and 2000 K.
Table 8.1: Molar specic heats
species M c
p
c
v
(kg kmol
1
) (kJ kmol
1
K
1
) (kJ kmol
1
K
1
)
N
2
28 33.0 24.7
N

2
28.15 32.8 24.5
O
2
32 34.8 26.5
air 29 33.2 24.9
H
2
2 31.1 22.8
CO 28 33.3 25.0
CO
2
44 53.7 45.4
H
2
O 18 42.7 34.4
CH
4
16 65.3 57.0
C
3
H
8
44 158 150
C
8
H
18
114 392 384
8.3.2 Open Systems
Combustion at constant pressure in an open system is of great practical
importance as it is a good approximation for the behaviour in the combustion
chamber of a gas turbine engine or furnace, as illustrated in gure 8.1. We can
understand how to calculate the amount of heat released in the reaction if we use
a graph of the temperature T of the reactants and products against their enthalpy
H. The two solid lines in gure 8.2 are isobars representing the enthalpy of the
92 TOPIC 8. COMBUSTION
fuel
air exhaust
Figure 8.1: Combustion at constant pressure
reactants and the products as a function of temperature. If the reactants and
products were ideal gases, these lines would be straight because for an ideal gas
c
p
= constant, but the slopes would probably be dierent as the values of c
p
of the
reactants and products would dier.
With an adiabatic combustion process from reactants R to products P the enthalpy
is constant, shown by a vertical line from point 1R to point 2P on the graph, but
there is a substantial rise in temperature. In a steady ow system the temperature
T
2
of the products is sometimes called the adiabatic ame temperature. Thus
H
1R
= H
2P
and T
2
> T
1
(8.1)
In the case of an isothermal combustion process, shown by a horizontal line from
point 1R to point 1P, the temperature obviously remains constant, and the
dierence in enthalpy corresponds to the standard enthalpy of reaction H
T
, also
called the isobaric caloric value, where
H
T
= H
PT
H
RT
Note that the value H
T
will be negative because, to maintain the products at the
same temperature as the reactants, enthalpy must be removed from the system.
Values of standard enthalpy of reaction are usually tabulated for a temperature of
T
H
adiabatic
isothermal enthalpy of
reactants H
enthalpy of
products H
1R
2P
1P
0R 0P
T
2
T
1
T
0
R
P
Figure 8.2: Temperatureenthalpy diagram for reactants and products
8.3. ENERGY OF REACTIONS 93
25
o
C = 298 K. HLT page 70 lists H
T
for a number of simple reactions at 298 K
and other higher temperatures. Other symbols such as H
0
T
and H
0
are also
used to represent the enthalpy of reaction.
We can now reconsider, as a simple example, the adiabatic combustion at
constant pressure from temperature T
1
to T
2
, as shown in gure 8.2. Equation 8.1
tells us that the enthalpies are equal, but to be able to determine the adiabatic
ame temperature T
2
given the caloric value of the fuel which is for a constant
temperature, we need to recall that enthalpy is a property so is independent of the
path by which it was reached. In other words, the process 1R 2P is equivalent
to the three chained processes 1R 0R, 0R 0P, 0P 2P. Rewriting Equation
8.1 gives
H
1R
H
2P
= 0 = (H
1R
H
0R
) + (H
0R
H
0P
) + (H
0P
H
2P
)
Introducing the heat capacities at constant pressure, dened in section 8.3.1, for
the reactants and products, this becomes
0 = C
pR
(T
1R
T
0R
) H
0
+ C
pP
(T
0P
T
2P
) (8.2)
where
H
0
= (H
0R
H
0P
) = CV
p
and CV
p
is the caloric value at constant pressure.
8.3.3 Gross and Net Caloric Values
Caloric values of fuels are usually tabulated for all reactants and products in the
vapour phase. This is referred to as the net or lower caloric value LCV. When the
water vapour in the products has been condensed to the liquid state, the caloric
value of the fuel is known as the gross or higher caloric value HCV. LCV and
HCV are related by
HCV = LCV + n
H
2
O

h
fg
where n
H
2
O
is the number of moles of H
2
O in the products per mole of fuel, and

h
fg
kJ kmol
1
is the molar enthalpy of condensation of H
2
O.
8.3.4 Closed Systems
Combustion at constant volume in a closed system is of interest as it is of practical
use in measuring the caloric value of a fuel with a bomb calorimeter , as
94 TOPIC 8. COMBUSTION
bomb
calorimeter
Figure 8.3: Combustion at constant volume
illustrated in gure 8.3. We can plot temperature T of the reactants and products
against their internal energy U. The two solid lines in gure 8.4 are lines of
constant volume or isochores representing the internal energy of the reactants and
the products as a function of temperature. If the reactants and products were ideal
gases, these lines would be straight because for an ideal gas c
v
= constant, but
the slopes would probably be dierent as the values of c
v
of the reactants and
products would dier.
For the constant volume combustion system then the same arguments as in the
previous section can be repeated but using internal energy instead of enthalpy.
With an adiabatic combustion process from reactants R to products P the internal
energy is constant, but there is a substantial rise in temperature. In a closed
system the temperature T
2
of the products is sometimes called the adiabatic
combustion temperature. Thus
U
1R
= U
2P
and T
2
> T
1
In the case of an isothermal combustion process, shown by a horizontal line from
point 1R to point 1P, the temperature obviously remains constant, and the
T
U
adiabatic
isothermal internal energy of
reactants U
internal energy of
products U
1R
2P
1P
T
2
T
1
R
P
Figure 8.4: Temperatureinternal energy diagram for reactants and products
8.3. ENERGY OF REACTIONS 95
dierence in internal energy corresponds to the isochoric caloric value U
T
,
where
U
T
= U
PT
U
RT
As in the previous section, we can use the heat capacities at constant volume to
get
0 = C
vR
(T
1R
T
0R
) U
0
+ C
vP
(T
0P
T
2P
) (8.3)
where
U
0
= (U
0R
U
0P
) = CV
v
and CV
v
is the caloric value at constant volume.
8.3.5 Example 3
Problem: A stoichiometric mixture of propane (C
3
H
8
) and air at 100
o
C is burnt in a
constant pressure system. What is the adiabatic ame temperature? The caloric
value for propane (gaseous reactants and products) is H
0
= 2043.15 MJ kmol
1
at 25
o
C.
Solution: First we need the stoichiometric combustion equation for 1 kmol of fuel.
Using a temporary variable z for the amount of oxygen we get
C
3
H
8
+ z (O
2
+ 3.76 N

2
) a CO
2
+ b H
2
O+ 3.76z N

2
We have already applied an atomic balance for nitrogen, and we continue with the
atomic balances for carbon and hydrogen
C balance 3 = a
O balance 2z = 2a + b
H balance 8 = 2b
so z = 5 and the stoichiometric combustion equation is
C
3
H
8
+ 5 (O
2
+ 3.76 N

2
) 3 CO
2
+ 4 H
2
O+ 18.8 N

2
We can now use an energy balance, and since it is constant pressure combustion
we need to use enthalpies, or the heat capacity at constant pressure. For the
reactants, with 1 kmol of fuel, using the data from the table in section 8.3.1, we get
96 TOPIC 8. COMBUSTION
C
pR
=

R
n
i
c
pi
= n
C
3
H
8
c
pC
3
H
8
+ n
O
2
c
pO
2
+ n
N

2
c
pN

2
= 1 158 + 5 34.8 + 18.8 32.8
= 948.6 kJ kmol
1
C
3
H
8
K
1
For the products, with 1 kmol of fuel
C
pP
=

P
n
i
c
pi
= n
CO
2
c
pCO
2
+ n
H
2
O
c
pH
2
O
+ n
N

2
c
pN

2
= 3 53.7 + 4 42.7 + 18.8 32.8
= 948.5 kJ kmol
1
C
3
H
8
K
1
Using equation 8.2 we get
0 = C
pR
(T
1R
T
0R
) H
0
+ C
pP
(T
0P
T
2P
)
which can be rewritten as
T
2P
= (C
pR
(T
1R
T
0R
) H
0
)/C
pP
+ T
0P
Substitution of numerical values gives
T
2P
=
948.6(100 25) + 2043150
948.5
+ 25 = 2254
o
C
TOPIC 9. INTERNAL COMBUSTION ENGINES 97
Topic 9
Internal Combustion Engines
9.1 Introduction
The internal combustion engine, in its two common forms, the spark-ignition
engine and the compression-ignition or Diesel engine, must rank as one of the
most important inventions ever made, providing controllable power from a truly
portable unit. In this chapter we will look at the theoretical cycles behind these
machines, using our knowledge of the First Law for a closed system, and also at
how real engines work.
9.2 Idealised Cycles
9.2.1 Air Standard Cycles
When we considered the Rankine cycle, the working uid at every point in the
cycle was steam, or water, a pure substance. The same water simply went around
and around the cycle, undergoing each process in turn, in a closed system. The
internal combustion engine is rather dierent.
The spark-ignition engine draws air from the atmosphere into a chamber, mixes it
with fuel, compresses the fuel-air mixture, ignites it with a spark, allows the
resulting hot gases to expand and do work, then expels the gases into the
atmosphere.
The compression-ignition engine is very similar except that the air alone is
compressed, after which the fuel is injected, igniting spontaneously. Thus there
are two features which are immediately apparent.
The internal combustion engine is not strictly a closed system, as air is drawn
in and later expelled, along with combustion products. True, that same air
could eventually be drawn into the engine again, making it a cycle as
illustrated in gure 9.1, but this is stretching the denition somewhat.
98 TOPIC 9. INTERNAL COMBUSTION ENGINES
engine
atmosphere
compression
combustion
expansion
cooling
heat rejection
heat addition
Figure 9.1: Internal combustion engine
The working uid is not a pure substance, and in fact changes considerably in
its composition and thermal properties, from process to process. Furthermore,
the mass of working uid contained within the engine varies with time. The
combustion process in a real engine is not the simple stoichiometric equation
you have seen in the last chapter, but is very complicated, with the formation of
many other products, such as oxides of nitrogen due to the high temperatures,
and hydrocarbons due to premature quenching of the combustion on the
relatively cold chamber walls.
To simplify the analysis of the internal combustion engine, we use the air standard
cycle, in which the working uid is assumed to be air, which circulates within the
engine as in a true closed system. This gives a good rst approximation, because
the principal constituent of air is N
2
, which remains unchanged during the cycle
(except for the formation of small amounts of NO and NO
2
).
9.2.2 Otto Cycle
The Otto cycle is an air standard cycle approximating the spark-ignition engine
cycle. There are four processes, as illustrated in gure 9.2.
1 2 represents reversible adiabatic compression of the air, requiring work
W
in
, from pressure p
1
and volume V
1
to p
2
and V
2
2 3 represents the addition of heat Q
in
to the air at constant volume V
2
= V
3
,
raising the pressure to p
3
3 4 represents reversible adiabatic expansion of the air, producing work
9.2. IDEALISED CYCLES 99
1
2
3
4
Q
out
Q
in
V
p
Figure 9.2: Otto cycle on a p V diagram
W
out
, to pressure p
4
and volume V
4
= V
1
4 1 represents the rejection of heat Q
out
from the air at constant volume
V
4
= V
1
, reducing the pressure to p
1
.
From the First Law, we can relate the work and heat terms by
Q
in
Q
out
= W
out
W
in
The thermal eciency
th
is given by

th
=
W
net
Q
in
=
W
out
W
in
Q
in
=
Q
in
Q
out
Q
in
= 1
Q
out
Q
in
Each of the heat transfer processes, 2 3 and 4 1, occurs at constant volume,
so the amounts of heat transfer are simply
Q
in
= Q
23
= mc
v
(T
3
T
2
) and Q
out
= Q
41
= mc
v
(T
4
T
1
)
where m is the mass inside the engine.
Hence the thermal eciency is

th
= 1
T
4
T
1
T
3
T
2
But the compression and expansion processes, 1 2 and 3 4, are both
reversible and adiabatic, so if the air behaves as an ideal gas, can be represented
as we saw in section 6.5 by
T
2
T
1
= r
1
v
and
T
3
T
4
= r
1
v
where r
v
is the volumetric compression ratio of the engine given by
r
v
=
V
1
V
2
=
V
4
V
3
100 TOPIC 9. INTERNAL COMBUSTION ENGINES
Thus the thermal eciency of the air standard Otto cycle is determined solely by
the compression ratio

th
= 1
1
r
1
v
To increase the thermal eciency, we should increase the compression ratio r
v
.
However there is a practical limit to r
v
of about 9 to 10, because as we compress
the air-fuel mixture from 1 2, it becomes hot, and may ignite spontaneously
before the end of the compression process. This pre-ignition reduces performance
and can damage the engine. The addition of tetra-ethyl lead to the fuel delays this
spontaneous ignition allowing a greater compression ratio and hence higher
thermal eciency, but discharges lead into the atmosphere. Lead-free petrol has
to be more highly rened to perform satisfactorily.
9.2.3 Diesel Cycle
The Diesel cycle is an air standard cycle approximating the compression-ignition
engine cycle. There are four processes, as illustrated in gure 9.3.
1 2 represents reversible adiabatic compression of the air, requiring work
W
in
, from pressure p
1
and volume V
1
to p
2
and V
2
2 3 represents the addition of heat Q
in
to the air at constant pressure
p
2
= p
3
, while the volume expands to V
3
, producing work
3 4 represents reversible adiabatic expansion of the air, producing work, to
pressure p
4
and volume V
4
= V
1
4 1 represents the rejection of heat Q
out
from the air at constant volume
V
4
= V
1
, reducing the pressure to p
1
.
1
2 3
4
Q
out
Q
in
V
p
Figure 9.3: Diesel cycle on a p V diagram
9.2. IDEALISED CYCLES 101
As for the Otto cycle, the thermal eciency
th
is given by

th
= 1
Q
out
Q
in
and the heat rejected from 4 1 is
Q
out
= Q
41
= mc
v
(T
4
T
1
)
However, the heat addition from 2 3 is at constant pressure, and so by denition
Q
in
= Q
23
= mc
p
(T
3
T
2
)
Note the use of c
p
in place of c
v
.
The compression and expansion processes, 1 2 and 3 4, are both reversible
and adiabatic, so if the air behaves as an ideal gas, the temperatures are related
by
T
2
T
1
= r
1
v
and
T
3
T
4
=
_
V
4
V
3
_
1
During heat addition from 2 3, the ideal gas law gives
p
2
V
2
T
2
=
p
3
V
3
T
3
but p
2
= p
3
so
V
2
T
2
=
V
3
T
3
giving
T
3
T
2
= r
c
where r
c
is the cut-o ratio dened by
r
c
=
V
3
V
2
From these we eventually get

th
= 1
1
r
1
v
r

c
1
(r
c
1)
How do we achieve a constant pressure heat addition from 2 3? In the Diesel
engine we compress the air alone, then spray ne droplets of fuel into the cylinder.
102 TOPIC 9. INTERNAL COMBUSTION ENGINES
diffusion of fuel vapour
diffusion of oxygen
fuel droplet
flame front
air
Figure 9.4: Combustion of droplets in Diesel engine
These droplets vaporise and burn relatively slowly as shown in gure 9.4, while the
piston moves down. This combination of increasing pressure due to combustion
and decreasing pressure due to expansion means that the pressure remains
approximately constant during the combustion process.
For a given compression ratio r
v
, the thermal eciency of the Diesel cycle is lower
than that of the Otto cycle. However, in practice, we can operate a Diesel cycle at
a much higher compression ratio than an Otto cycle, so achieve a higher thermal
eciency. In fact, we have to use a higher compression ratio, at least 12, in order
to make the air hot enough to ignite the fuel droplets! Compression ratios up to 20
are typical.
9.3 Real Internal Combustion Engines
9.3.1 Spark-Ignition Engine
Most spark-ignition engines are four-stroke engines, meaning that there are four
distinct processes in each cycle taking a total of two engine revolutions to
complete. A stroke is a single up or down movement of the piston, hence there are
four strokes per cycle. The four strokes, shown in gure 9.5, are
When the piston is at top-dead-centre (TDC) the inlet valve opens. The piston
moves down, drawing air-fuel mixture into the cylinder. This is the intake or
induction stroke.
At bottom-dead-centre (BDC) the inlet valve closes, trapping the air-fuel
mixture in the cylinder. The piston now moves up, compressing the mixture.
This is the compression stroke.
At TDC, an electric spark ignites the air-fuel mixture, which burns rapidly. The
9.3. REAL INTERNAL COMBUSTION ENGINES 103
piston
inlet valve
air-fuel
mixture
spark
combustion
products
exhaust valve
induction compression expansion exhaust
Figure 9.5: Four-stroke spark-ignition engine
hot combustion products then push the piston downward as they expand. This
is the expansion stroke.
At BDC, the exhaust valve opens, allowing some of the combustion products to
escape immediately. The piston then moves up, forcing the combustion
products out the exhaust valve. This is the exhaust stroke. At TDC, the
exhaust valve closes.
Where does the air-fuel mixture come from? Air from the atmosphere is drawn
through a lter to remove dust particles. In most engines the air then enters a
carburettor , illustrated in gure 9.6. Inside the carburettor is a venturi , a local
narrowing of the tube, which causes the velocity to increase and the pressure to
decrease. Liquid fuel is drawn by this low pressure through a ne nozzle or jet to
mix with the air and vaporise. The carburettor is carefully designed so that the
air-fuel ratio remains constant over a wide range of air ows to ensure proper
operation of the engine. The air-fuel mixture then passes through an adjustable
throttle, usually a buttery valve connected by a cable to the accelerator pedal.
fuel
venturi
jet
air air-fuel mixture
throttle
Figure 9.6: Carburettor for spark-ignition engine
104 TOPIC 9. INTERNAL COMBUSTION ENGINES
throttle
flow
meter
controller injector
fuel
air-fuel mixture air
temperature
sensor
Figure 9.7: Fuel injection for spark-ignition engine
When the buttery valve is partially closed, the amount of mixture entering the
engine is reduced, so controlling the power output. The air-fuel mixture from the
carburettor then passes into the inlet manifold, a multi-branched pipe which
delivers the mixture to the cylinders of the engine.
Many modern spark-ignition engines are tted with fuel injection, shown in gure
9.7, instead of a carburettor. In these engines, air is drawn in through a lter, past
a temperature sensor, through a ow-measuring sensor, and then into the inlet
manifold. A pump delivers liquid fuel to an electrically-operated injector in the
manifold. An electronic circuit controls the amount of fuel squirted into the
manifold, where it quickly vaporises, for a given air ow rate and other operating
conditions in order to produce the best air-fuel ratio. Some high-performance
engines have several injectors, often one just before each inlet valve. Note that
fuel injection in a spark-ignition engine operates at relatively low pressure into the
inlet manifold, unlike the compression-ignition engine, in which fuel injection
requires a high pressure to deliver the fuel directly into the cylinder near the end of
the compression stroke.
Because the valve opening and closing in a real engine require a nite time, it is
usual for the valves to open slightly before the TDC or BDC position and to close
slightly after. Similarly, the combustion requires a nite time, so it is usual for the
spark to occur before TDC. The amount of spark advance required for best
performance depends on the engine speed and the throttle opening, so the
ignition system usually includes a centrifugal mechanism to sense the speed and a
bellows to sense the pressure in the inlet system.
On a p V diagram like that in gure 9.8, the real spark-ignition engine shows
several dierences from the Otto cycle.
9.3. REAL INTERNAL COMBUSTION ENGINES 105
V
p
A
B
positive work
negative work
Figure 9.8: p V diagram for four-stroke spark-ignition engine
During the compression stroke, heat is transferred between the relatively hot
cylinder walls and the air-fuel mixture, so the compression is not adiabatic, but
can usually be considered as polytropic.
Combustion begins before TDC (point A) and continues until after TDC (point
B). Thus the combustion is not a constant volume process.
During the expansion stroke, heat is transferred from the hot combustion
products to the relatively cool cylinder walls. Thus the expansion is not
adiabatic, but may be considered as polytropic.
During the exhaust stroke, the pressure in the cylinder may remain a little
above atmospheric as the piston pushes the combustion products through the
restriction caused by the exhaust valve.
During the intake stroke, the pressure in the cylinder may fall a little below
atmospheric as the air is drawn into the carburettor through the throttle valve
and then through the restriction caused by the inlet valve.
The positive work done by the engine is represented by the area enclosed by the
compression and expansion strokes. The negative work, or induction work, is
represented by the area enclosed by the induction and exhaust strokes. At
part-throttle, the positive work decreases, and there is less air-fuel mixture to burn,
and the negative work increases, as the pressure drop across the throttle valve
increases. Thus the thermal eciency of the spark-ignition engine is reduced at
part-throttle.
9.3.2 Compression-Ignition Engine
Most compression-ignition engines are four-stroke engines, meaning that there are
four distinct processes in each cycle taking a total of two engine revolutions to
106 TOPIC 9. INTERNAL COMBUSTION ENGINES
piston
inlet valve exhaust valve
induction compression expansion exhaust
air
fuel
droplets
Figure 9.9: Four-stroke compression-ignition engine
complete. The four strokes, illustrated in gure 9.9, are
When the piston is at top-dead-centre (TDC) the inlet valveindexinlet valve
opens. The piston moves down, drawing air into the cylinder. This is the intake
or induction stroke.
At bottom-dead-centre (BDC) the inlet valve closes, trapping the air in the
cylinder. The piston now moves up, compressing the air. This is the
compression stroke.
At TDC, liquid fuel is injected as a mist of ne droplets into the hot compressed
air. The droplets vaporise and burn gradually. The hot combustion products
then push the piston downward as they expand. Because the combustion is
gradual, it continues during the rst part of the expansion. This is the
expansion stroke.
At BDC, the exhaust valve opens, allowing some of the combustion products to
escape immediately. The piston then moves up, forcing the combustion
products out the exhaust valve. This is the exhaust stroke. At TDC, the
exhaust valve closes.
Air from the atmosphere is drawn through a lter to remove dust particles, and
then into the inlet manifold and then into the cylinder. Unlike the spark-ignition
engine, the compression-ignition engine has no throttle. A high-pressure fuel
pump delivers liquid fuel to an injector which squirts the fuel directly into the
cylinder near TDC on the compression stroke. The amount of fuel delivered
9.3. REAL INTERNAL COMBUSTION ENGINES 107
A
B
positive work
negative work
p
V
Figure 9.10: p V diagram for four-stroke compression-ignition
depends on the position of the accelerator pedal. Note that the amount of air in the
cylinder is independent of the accelerator position.
How can the compression-ignition engine operate without a throttle? As each fuel
droplet vaporises, the gaseous fuel diuses away from the droplet creating a
concentration gradient. At some small distance from each droplet, the air-fuel ratio
will be just right for combustion to occur.
On a p V diagram like that in gure 9.10, the real compression-ignition engine
shows several dierences from the Diesel cycle.
During the compression stroke, heat is transferred between the relatively hot
cylinder walls and the air, so the compression is not adiabatic, but can usually
be considered as polytropic.
Combustion begins before TDC (point A) and continues until well after TDC
(point B).
During the expansion stroke, heat is transferred from the hot combustion
products to the relatively cool cylinder walls. Thus the expansion is not
adiabatic, but may be considered as polytropic.
During the exhaust stroke, the pressure in the cylinder may remain a little
above atmospheric as the piston pushes the combustion products through the
restriction caused by the exhaust valve.
During the intake stroke, the pressure in the cylinder may fall a little below
atmospheric as the air is drawn through the restriction caused by the inlet
valve.
The positive work done by the engine is represented by the area enclosed by the
compression and expansion strokes. The negative work, or induction work, is
108 TOPIC 9. INTERNAL COMBUSTION ENGINES
1
2
3
4
V
p
V
swept
Figure 9.11: Swept volume
represented by the area enclosed by the induction and exhaust strokes. Because
there is no throttle to increase the negative work, the thermal eciency of the
compression-ignition engine at part-load is higher than that of the spark-ignition
engine at part-throttle.
9.3.3 Swept Volume
The swept volume of an internal combustion engine is the volume swept out by the
piston moving from bottom-dead-centre (BDC) to top-dead-centre (TDC). On a
p V diagram such as gure 9.11, the swept volume is given by
V
swept
= V
1
V
2
= V
4
V
3
In a multi-cylinder engine, the swept volume for the engine is taken to be the total
swept volume for all of the cylinders.
9.3.4 Mean Eective Pressure
The mean eective pressure of an internal combustion engine is the height of a
rectangle having the same length and area as the cycle plotted on a p V diagram
such as gure 9.12. This area represents the net work W
net
per cycle. The length
of the rectangle represents the swept volume. Hence the mean eective pressure
represents that constant pressure which would need to be applied to the piston
during the expansion stroke to produce the same net work.
p
mean
=
W
net
V
swept
The mean eective pressure provides a useful way of comparing the performance
of similar engines of dierent size. Strictly speaking, this is the indicated mean
9.4. EXAMPLES 109
1
2
3
4
V
p
p
mean
W
net
V
swept
W
net
Figure 9.12: Mean eective pressure
eective pressure, or imep, because it is derived from the p V diagram,
sometimes called the indicator diagram. There is also the brake mean eective
pressure, or bmep, which is calculated from the brake work, the work produced at
the output shaft. Note that in a real engine bmep < imep because of friction and
other losses.
9.4 Examples
9.4.1 Example 1
Problem: In an air standard Otto cycle the minimum and maximum temperatures
are 15 and 1550
o
C. The heat supplied is 800 kJ per kg of air. Calculate (a) the
compression ratio (b) the cycle eciency (c) the ratio of maximum to minimum
pressures (d) the temperature at the end of the expansion stroke.
Solution: We are given T
1
= 15
o
C = 288 K, T
3
= 1550
o
C = 1823 K, and q
in
= 800
kJ kg
1
. Combustion is at constant volume, hence
q
in
= c
v
(T
3
T
2
) = 800 kJ kg
1
and
T
2
= T
3
q
in
/c
v
= 1823 800/0.718 = 1823 1113 = 708.8 K
For reversible adiabatic compression 1 2
T
2
T
1
= r
1
v
so with = 1.4
r
v
=
_
T
2
T
1
_
1/(1)
=
_
708.8
288
_
1/0.4
= 9.50
110 TOPIC 9. INTERNAL COMBUSTION ENGINES
The cycle eciency is

th
= 1
1
r
1
v
= 1
1
9.50
0.4
= 0.5937
For reversible adiabatic compression
p
2
= p
1
_
v
1
v
2
_

= p
1
r

v
= 9.50
1.4
p
1
= 23.39p
1
If the air behaves as an ideal gas then during combustion 2 3
p
2
v
2
T
2
=
p
3
v
3
T
3
but v
2
= v
3
because the combustion is at constant volume, so
p
3
= p
2
_
T
3
T
2
_
= 23.39p
1
_
1823
708.8
_
= 60.15p
1
For reversible adiabatic expansion 3 4
T
3
T
4
= r
1
v
hence
T
2
T
1
=
T
3
T
4
giving
T
4
= T
3
T
1
T
2
= 1823
288
708.8
= 740.7 K
9.4.2 Example 2
Problem: An air standard Diesel engine has a compression ratio of 15, minimum
and maximum temperatures of 15 and 1650
o
C, and a maximum pressure of 45
bar. Calculate (a) the cycle eciency (b) the mean eective pressure.
Solution: We are given T
1
= 15
o
C = 288 K, T
3
= 1650
o
C = 1923 K, p
2
= p
3
= 45
bar = 4.5 MPa, and r
v
= 15. During reversible adiabatic compression 1 2
T
2
T
1
= r
1
v
so
T
2
= T
1
r
1
v
= 288 15
0.4
= 850.8 K
If the air behaves as an ideal gas then during combustion 2 3
p
2
V
2
T
2
=
p
3
V
3
T
3
9.4. EXAMPLES 111
but p
2
= p
3
so
r
c
=
V
3
V
2
=
T
3
T
2
=
1923
850.8
= 2.26
During expansion 3 4
V
4
V
3
=
V
4
V
2
V
2
V
3
=
V
1
V
2
V
2
V
3
= r
v

V
2
V
3
=
15
2.26
= 6.64
The expansion is reversible and adiabatic hence
T
3
T
4
=
_
V
4
V
3
_
1
and so
T
4
= T
3
_
V
3
V
4
_
1
= 1923
_
1
6.64
_
0.4
= 902 K
The heat in is
q
in
= c
p
(T
3
T
2
) = 1.005(1923 850.8) = 1077.6 kJ kg
1
and the heat out is
q
out
= c
v
(T
4
T
1
) = 0.718(902 288) = 440.9 kJ kg
1
From the First Law
w
net
= w
out
w
in
= q
in
q
out
= 1077.6 440.9 = 636.7 kJ kg
1
so the eciency is

th
=
w
net
q
in
=
636.7
1077.6
= 0.591
Alternatively we can use

th
= 1
1
r
1
v
r

c
1
(r
c
1)
= 1
1
15
0.4
2.26
1.4
1
1.4(2.26 1)
= 0.591
If the air behaves as an ideal gas then
v
2
=
RT
2
p
2
=
287 850.8
4.5 10
6
= 0.0543 m
3
kg
1
so
v
1
= 15v
2
= 0.8139 m
3
kg
1
and the mean eective pressure is
p
mean
=
w
net
v
1
v
2
=
636.7 1000
0.8139 0.0543
= 8.38 10
5
Pa = 8.38 bar
112 TOPIC 10. GAS TURBINE ENGINES
Topic 10
Gas Turbine Engines
10.1 Introduction
The gas turbine engine has revolutionised air travel with its amazing power to
weight ratio. In this chapter we will look at the theoretical cycle behind the gas
turbine, and also how a real gas turbine engine works.
10.2 Idealised Cycle
10.2.1 Air Standard Cycles
As for the discussion of internal combustion engine cycles, we use air standard
cycles when we rst consider gas turbine engines. That is, we assume that the
working uid is air, which circulates within the engine as in a true closed system.
This gives a good approximation because the principal constituent of air is N
2
.
10.2.2 Joule Cycle
The Joule cycle is an air standard cycle approximating the gas turbine engine.
There are four processes as illustrated in gure 10.1.
1 2 represents reversible adiabatic compression of the air, requiring work
w
in
, from pressure p
1
and temperature T
1
to p
2
and T
2
p
v
1
2 3
4
p
2
p
1
Figure 10.1: Joule cycle on a p v diagram
10.2. IDEALISED CYCLE 113
2 3 represents reversible addition of heat q
in
to the air at constant pressure
p
2
= p
3
, raising the temperature to T
3
3 4 represents reversible adiabatic expansion of the air, producing work
w
out
, to pressure p
4
and temperature T
4
4 1 represents reversible rejection of heat q
out
from the air at constant
pressure p
4
= p
1
, reducing the temperature to T
1
.
From the First Law, we can relate the work and heat terms by
q
in
q
out
= w
out
w
in
The thermal eciency
th
is given by

th
=
w
net
q
in
=
w
out
w
in
q
in
Both of the heat transfer processes, 2 3 and 4 1, occur at constant pressure,
so the amounts of heat transfer are simply
q
in
= q
23
= c
p
(T
3
T
2
) and q
out
= q
41
= c
p
(T
4
T
1
)
Similarly, the compressor and turbine work are given by
w
in
= w
12
= c
p
(T
2
T
1
) and w
out
= w
34
= c
p
(T
3
T
4
)
Hence the thermal eciency is

th
=
(T
3
T
4
) (T
2
T
1
)
T
3
T
2
= 1
T
4
T
1
T
3
T
2
The compression and expansion processes, 1 2 and 3 4, are both reversible
and adiabatic, so if the air behaves as an ideal gas, they can be represented using
the polytropic law from section 6.5 by
T
2
T
1
= r
(1)/
p
and
T
3
T
4
= r
(1)/
p
where r
p
is the pressure ratio of the gas turbine engine given by
r
p
=
p
2
p
1
=
p
3
p
4
Substituting these temperatures and simplifying gives

th
= 1
1
r
(1)/
p
114 TOPIC 10. GAS TURBINE ENGINES
th
r
p
1
0

Figure 10.2: Joule cycle eciency versus pressure ratio


Thus the thermal eciency of the air standard Joule cycle is determined solely by
the pressure ratio. Increasing the pressure ratio r
p
increases the thermal eciency.
This is illustrated for = 1.4 in gure 10.2. Note that the eciency falls to zero at
r
p
= 1.
The work ratio r
w
is dened as the ratio of the net work out to the turbine work
r
w
=
w
net
w
out
=
w
out
w
in
w
out
hence
r
w
=
(T
3
T
4
) (T
2
T
1
)
T
3
T
4
Substituting and simplifying gives
r
w
= 1
_
T
1
T
3
_
r
(1)/
p
The work ratio indicates what fraction of the turbine work is available to drive an
external load such as a propeller, fan or generator. For given values of minimum
and maximum temperatures, increasing the pressure ratio decreases the work
ratio. To achieve a given power output at high thermal eciency requires high
pressure ratio, but a large turbine and compressor.
The work ratio falls to zero when
r
p
=
_
T
3
T
1
_
/(1)
That is, the cycle is producing no net work, and all the work produced by the
turbine is needed to drive the compressor. For = 1.4, minimum temperature
T
1
= 300 K and maximum temperature T
3
= 1000 K, this occurs at pressure ratio
r
p
68. This represents a likely upper limit to the pressure ratio which might be
achieved in practice.
10.3. REAL GAS TURBINE ENGINES 115
combustion
chamber
compressor turbine
fuel
air exhaust
w
net
w
in
Figure 10.3: Schematic diagram of gas turbine engine
10.3 Real Gas Turbine Engines
A real gas turbine engine has a compressor, combustion chamber and turbine as
shown in gures 10.3 and 10.4.
Air is drawn into the intake and compressed by one or more stages of rotating
blades. In a radial compressor the pressure increase is caused by the
centrifugal force as the air is ung outwards. In an axial compressor, the
pressure increase is caused by forcing the air through progressively narrower
passages. Adjustable stator blades attached to the engine casing are used to
control the direction of the air ow over the rotor blades attached to the
rotating hub.
Fuel is then squirted under pressure into the compressed air inside a
combustion chamber . An igniter is used to start the combustion, after which a
steady ame continues.
The hot combustion products then pass at high speed through the turbine,
compressor turbine
combustion
chamber
rotor
inlet exhaust
stator
Figure 10.4: Cross-section of gas turbine engine
116 TOPIC 10. GAS TURBINE ENGINES
usually an axial turbine, which consists of one or more stages of rotating
blades. A shaft transmits power from the turbine to drive the compressor.
The exhaust gases may be used to drive another turbine stage, called the
power turbine, which can be connected to an external load, such as a gearbox
to drive a propellor, helicopter rotor, or a generator. In the jet engine, the hot
gases are simply passed through a nozzle to produce a very high speed jet of
hot gases which provides the forward thrust.
10.4 Examples
10.4.1 Example 1
Problem: Calculate the specic work output for an engine operating on the Joule
cycle, and nd the pressure ratio for which it is a maximum.
Solution: The specic work output is
w
net
= w
out
w
in
= c
p
((T
3
T
4
) (T
2
T
1
))
where
T
2
T
1
= r
(1)/
p
and
T
3
T
4
= r
(1)/
p
so that
w
net
= c
p
_
T
3
_
1
T
4
T
3
_
T
1
_
T
2
T
1
1
__
= c
p
_
T
3
_
1 r
(1)/
p
_
T
1
_
r
(1)/
p
1
__
Dierentiating with respect to r
p
at constant T
1
and T
3
gives
dw
net
dr
p
= c
p
_
T
3
_
1

_
r
(1)/
p
T
1
_
1

_
r
(1)/
p
_
This derivative is zero when
T
3
_
1

_
r
(1)/
p
= T
1
_
1

_
r
(1)/
p
giving
r
p
=
_
T
3
T
1
_
/2(1)
10.4. EXAMPLES 117
10.4.2 Example 2
Problem: A gas turbine engine has minimum and maximum temperatures of 20
and 760
o
C and a pressure ratio of 7. Calculate (a) the cycle eciency (b) the work
ratio.
Solution: We are given r
p
= 7, T
1
= 20
o
C = 293 K, and T
3
= 760
o
C = 1033 K. The
compression from 1 2 is reversible and adiabatic so
T
2
T
1
= r
(1)/
p
hence
T
2
= T
1
r
(1)/
p
= 293 (7)
0.4/1.4
= 510.9 K
The heat input is
q
in
= c
p
(T
3
T
2
) = 1.005 (1033 510.9) = 524.7 kJ kg
1
The expansion from 3 4 is reversible and adiabatic so
T
4
= T
3
/r
(1)/
p
= 1033/(7)
0.4/1.4
= 592.4 K
The heat output is
q
out
= c
p
(T
4
T
1
) = 1.005 (592.4 293) = 300.9 kJ kg
1
so from the First Law the net work is
w
net
= w
out
w
in
= q
in
q
out
= 524.7 300.9 = 223.8 kJ kg
1
and the cycle eciency is

th
=
w
net
q
in
=
223.8
524.7
= 0.427
The turbine work is
w
out
= c
p
(T
3
T
4
) = 1.005 (1033 592.4) = 442.8 kJ kg
1
and the work ratio is
r
w
=
w
net
w
out
=
223.8
442.8
= 0.505
118 INDEX
Index
adiabatic
compression, 75
compressor, 69
expansion, 77
ame temperature, 92
nozzle, 71
process, 47, 60
throttle, 70
turbine, 70
air
combustion in, 84
composition, 84
excess, 85
standard cycle, 97, 112
air-fuel ratio, 85, 103
atmospheric nitrogen, 84
atomic
balance, 87
mass, 10
number, 10
Avogadros number, 12
axial
compressor, 115
turbine, 116
BDC, 102, 106
bmep, 109
boiler, 76
bottom-dead-centre, 102, 106
boundary
moving, 52
system, 45
Boyles law, 13, 37
brake
mean eective pressure, 108
work, 109
bulk modulus, 17
caloric value
gross and net, 93
higher and lower, 93
isobaric, 92
carbon monoxide, 85
carburettor, 103
centrifugal compressor, 68
Charles law, 13, 37
chemical reactions, 83
closed system, 45, 48
coecient
compressibility, 16
cubical expansion, 18, 19
diusion, 28
Joule-Thomson, 71
linear expansion, 18
of performance
heat pump, 82
refrigerator, 82
thermal
conductivity, 24
expansion, 20
viscosity, 25
combustion
INDEX 119
chemical reactions, 83
constant pressure, 91
constant volume, 93
stoichiometric, 85
compressibility
coecient, 16
factor, 33
solids, 16
compression
ratio, 99, 102
stroke, 102, 106
compression-ignition engine, 97, 100
compressor
adiabatic, 69
axial, 115
centrifugal, 68
radial, 115
reciprocating, 69
work, 68
condenser, 78
conduction, thermal, 24
conductivity, thermal, 24
constant
pressure
combustion, 91
specic heat, 65
volume
combustion, 93
specic heat, 49
control surface, 45
COP, 82
critical
point, 36, 37
temperature, 37
cubical expansion, 18, 19
cut-o ratio, 101
cycle
air standard, 97, 112
denition, 47
Diesel, 100
gas turbine, 112
Joule, 112
Otto, 98
Rankine, 79
diabatic process, 47
Diesel cycle
denition, 100
eciency, 101
diusion, 27
dryness fraction, 41
eciency
Diesel cycle, 102
Joule cycle, 114
Otto cycle, 100
Rankine cycle, 79
thermodynamic, 80
energy
gravitational, 62
internal, 49
kinetic, 62
total, 62
engine
compression-ignition, 97, 100
Diesel, 100
four-stroke, 102
gas turbine, 112
internal combustion, 97
120 INDEX
spark-ignition, 97, 98
enthalpy
closed system, 73
compressor, 68
denition, 64
nozzle, 71
throttle, 70
turbine, 69
equation
ideal gas, 29
steady ow energy, 62
equivalence ratio, 86, 89
ethane, 86
ethanol, 86
ethene, 86
excess air, 85
exhaust
stroke, 103, 106
valve, 103, 106
expansion stroke, 103, 106
feed pump, 75
rst law
closed system, 48, 73
denition, 48
enthalpy, 73
open system, 62
four-stroke engine, 102
fuel-injection, 104
gas
constant, molar, 30
constant, specic, 30
ideal, 29
non-ideal, 33
real, 35
turbine engine, 112, 115
gravimetric composition, 84
gravitational energy, 62
gross caloric value, 93
heat
engine, 81
pump, 81
sign convention, 46
higher caloric value, 93
hydrocarbon fuels, 86
ideal gas
equation, 29
law, 29
p-v-T surface, 37
imep, 109
index, polytropic, 58, 59
induction
stroke, 102, 106
work, 105, 107
inlet valve, 102
internal
combustion engine, 97
energy, 49
inversion temperature, 71
irreversible
expansion, 47
process, 47
isobaric
caloric value, 92
heating, 76
process, 47
isochoric process, 47
INDEX 121
isolated system, 45
isothermal process, 47, 60
isotopes, 11
Joule cycle
denition, 112
eciency, 113
work ratio, 114
Joule-Thomson coecient, 71
kinetic energy, 62
kmol, 13
law
Boyles, 13, 37
Charles, 13, 37
rst, 48, 62
ideal gas, 29
polytropic, 57
lead, tetra-ethyl, 100
liquid
saturated, 37
subcooled, 35, 37, 74
thermal expansion, 19
lower caloric value, 93
mass
atomic, 10
molecular, 12
mean eective pressure, 108
methane, 86
methanol, 86
mixture, stoichiometric, 85
modulus
bulk, 17
Youngs, 16
molar
composition, 84
specic heat, 89
mole, 30
molecular mass, 12
moving boundary, 52, 73
net caloric value, 93
nitrogen
atmospheric, 84
balance, 87
non-ideal gas, 33
nozzle, 71, 116
number
atomic, 10
Avogadros, 12
open system, 45, 62
Otto cycle
denition, 98
eciency, 100
oxygen balance, 85
oxygenates, 86
p-v diagram
Diesel cycle, 100
Otto cycle, 98
Rankine cycle, 80
real substance, 37
reversible work, 51, 68
p-v-T surface, 37
path, 46
phase
bulk, 14, 38
rule, 38
Poissions ratio, 16
122 INDEX
polytropic
index, 58, 59
law, 57
process, 57
work, 58
pre-ignition, 100
pressure
mean eective, 108
ratio, 113
reduced, 33
vapour, 35
process
adiabatic, 47, 60
diabatic, 47
irreversible, 47
isobaric, 47
isochoric, 47
isothermal, 47, 60
polytropic, 57
reversible, 47
property
extensive, 15
intensive, 15
pump
feed, 75
reversible adiabatic, 75
quality, 41
radial compressor, 115
Rankine cycle, 79
ratio
air-fuel, 85, 103
compression, 99, 102
cut-o, 101
equivalence, 86
pressure, 113
reaction, stoichiometric, 85
real gas, 35
reciprocating compressor, 69
reduced
pressure, 33
temperature, 33
refrigerator, 82
reversible
expansion, 47
process, 47
work, 50, 67
rotor blades, 115
saturated
liquid, 37
vapour, 37
SFEE, 62
shaft work, 63, 68
solids
compressibility, 16
thermal expansion, 18
spark advance, 104
spark-ignition engine, 97, 98
specic
enthalpy, 64
heat
constant pressure, 65
constant volume, 49, 50
gases, 30
liquids, 22
molar, 89
relationships, 66
solids, 22
INDEX 123
internal energy, 15, 49
volume, 15
state, 14
stator blades, 115
steady ow energy equation, 62
steam
dryness fraction, 41
quality, 41
tables, 40
stoichiometric, 85
strain, 16
stress, 16
subcooled liquid, 35, 37, 74
superheated vapour, 35, 37
surroundings, 45
swept volume, 108
system
boundary, 45
closed, 45, 48
denition, 45
isolated, 45
open, 45
TDC, 102, 106
temperature
critical, 37
inversion, 71
reduced, 33
tetra-ethyl lead, 100
thermal
capacity, 22
conductivity, 24
expansion
coecient, 20
liquids, 19
solids, 18
thermodynamic eciency, 80
throttle, 70, 103
top-dead-centre, 102, 106
total energy, 62
turbine
adiabatic, 70
axial, 116
gas, 112
work, 69
vapour
pressure, 35
saturated, 37
superheated, 35, 37
venturi, 103
viscosity, 25
volume, swept, 108
work
brake, 109
closed system, 50
compressor, 68
induction, 105, 107
open system, 67
p-v, 51, 68
polytropic, 58
reversible, 50, 67
shaft, 63, 68
sign convention, 46
steady ow, 67
turbine, 69
Youngs modulus, 16

Vous aimerez peut-être aussi