Sie sind auf Seite 1von 337

INTERNATIONAL STUDIES

in SCIENCE and ENGINEERING

Roman Weber

COMBUSTION FUNDAMENTALS
with
Elements of Chemical Thermodynamics

Prof. Dr.Ing. Roman Weber


Technische Universitt Clausthal
Institut fr Energieverfahrenstechnik und Brennstofftechnik (IEVB)
Agricolastrasse 4, 38 678 Clausthal-Zellerfeld, Germany
roman.weber@ievb.tu-clausthal.de
Weber, Roman:
Combustion Fundamentals with Elements of Chemical Thermodynamics
Clausthal-Zellerfeld: Papierflieger 2008
ISBN 9783897209213
Bibliografische Information der Deutschen Bibliothek
Die Deutsche Bibliothek verzeichnet diese Publikation in der Deutschen Nationalbibliografie; detaillierte bibliografische Daten sind im Internet ber http://dnb.ddb.de
abrufbar.

INTERNATIONAL STUDIES in SCIENCE and ENGINEERING


Editor in Chief:
Prof. Dr.-Ing. Roman Weber, Clausthal University of Technology (Germany)
Editorial Board:
Dr.-Ing. Rdiger Alt, Clausthal University of Technology (Germany)
Prof. Dr.-Ing. Ryszard Bialecki, Silesian University of Technology (Poland)
Prof. Xu Delong, Xian University of Architecture and Technology (China)
Prof. Dr. Peter v. Dierkes, former President of Berliner Stadtreinigungsbetriebe (Germany)
Dipl.-Math. Marc Muster, Clausthal University of Technology (Germany)
Prof. Dr.-Ing. Andrzej Nowak, Silesian University of Technology (Poland)
Prof. Dr.-Ing. Reinhard Scholz, Clausthal University of Technology (Germany)
First Edition 2008
Copyright by PAPIERFLIEGER, Clausthal-Zellerfeld 2008, Telemannstr. 1, 38678 Clausthal-Zellerfeld, Tel.: 05323/96746, http://www.papierflieger-verlag.de
No part of this publication may be reproduced or transmitted in any form or by any
means, electronic or mechanical, including photocopying, recording, or any information
storage and retrieval system, without the prior permission in writing from the publisher.
ISBN 9783897209213
The cover of this textbook has been designed using photographs provided by Mr. Marc
Muster of TU Clausthal.

To
Professor Stanisaw Jerzy Gdula
who introduced me to the wonderful
world of thermodynamics

iii

International Studies in Science


and Engineering
The Editorial Board encourages its colleagues all over the world to publish in the
INTERNATIONAL STUDIES in SCIENCE and ENGINEERING both text
books which accompany a lecture series for students and other books which
demonstrate how to apply the knowledge acquired in lecture theatres to industrial
practise.

With publishing Combustion Fundamentals with Elements of Chemical Thermodynamics the sixth book in this series was released. At least three further books
are expected to be published in 2008: one concerns Advanced Heat Transfer,
two more on Abfall and Chemie.

Already published:
(the latest editions are listed only)
1. Weber, R.: Lecture Notes in Heat Transfer, 3rd Edition, PapierfliegerVerlag, Clausthal-Zellerfeld 2008, ISBN 3897207028.
2. Jeschar, R.; Kostowski, E.; Alt, R.: Wrmestrahlung in Industriefen,
Papierflieger-Verlag, Clausthal-Zellerfeld 2004, ISBN 3897206862 and
Wydawnictwo Politechniki Slaskiej, Gliwice 2004, ISBN 8373352325.
3. Weber, R.; Alt, R.; Muster, M.: Vorlesungen zur Wrmebertragung, Teil I,
2. Auflage, Papierflieger-Verlag, Clausthal-Zellerfeld 2008, ISBN 389720
7982.
4. Dierkes, P. v.; Bruch, G.: Abfall und Chemie, Teil I, Papierflieger-Verlag,
Clausthal-Zellerfeld 2007, ISBN 3897208792.
5. Dierkes, P. v.; Bruch, G.: Abfall und Chemie, Teil II, Papierflieger-Verlag,
Clausthal-Zellerfeld 2007, ISBN 3897208806.

iv

To the Student
How to get the most from these lecture notes
These lecture notes in Combustion Fundamentals with Elements of Chemical
Thermodynamics have been prepared for undergraduate students attending a
fifteenweek course (one semester) with 1.5 hour of lecture and 45 minutes of
classes per week. This is a typical course at the Clausthal University of Technology. The notes have been prepared not only for Clausthal but also for the Royal
Institute of Technology (Stockholm, Sweden), University of Science and Technology (Beijing, China) and Central South University (Changsha, China). If you
are like most students just opening this textbook, you are enrolled on one of the
few courses in science and engineering at Clausthal that are delivered in English.
Probably English is not your native language but do not worry. This textbook
has been written with you in mind in very simple and plain English. Below I
provide you with few suggestions how Combustion Fundamentals with Elements
of Chemical Thermodynamics can help you to succeed in this course.
Read before a lecture. You will get the most out of your combustion course
if you read each chapter before hearing it. In this way, many of the topics will
already be clear in your mind and you will understand the lecture better.
Study Examples and Problems. Each chapter contains several Examples.
Study them carefully since they illustrate the subject of each particular chapter. In addition to Examples, Problems are formulated for each chapter and they
can be found on the web site (see below). Some Problems will be discussed and
solved in the classroom but the principal idea behind Problems is that they are
your homework. I suggest you solve them, one by one, and in case of difficulties
consult your class instructor.
Study together with your fellow students. Many students find it useful to
form study groups. You can discuss the challenging topics with one another and
have a good time while doing it. Make sure that you do Problems by yourself
since during the exam you will have to prove your skills in problem solving.
Take advantage of the web site. In addition to these lecture notes we have developed a web site to assist you in this course. You can find it by browsing through
the IEVBInstitute website of TU Clausthal: http://www.ievb.tu-clausthal.de/
(follow Science and Lectures).

Acknowledgments
Combustion Fundamentals with Elements of Chemical Thermodynamics has
been written within the scope of a European Community project (CN/ASIALINK/016 (103 187)) of the Asia-Link Programme. The author acknowledges
with thanks the European Community financial support.

I would like to thank my colleagues who carefully scrutinized the manuscript


making it a better textbook:
Dipl.Ing. Stefan Brinker, Dipl.Ing. Sven Gose, Dipl.Ing. Patrick Schwppe,
Dr.Ing. Marco Mancini, Dipl.Math. Marc Muster (all Clausthal University of
Technology, Germany), Dr.Ing. Gabriel Wecel (Silesian University of Technology,
Poland). My special thanks go to Stefan Brinker and Marc Muster who edited
this textbook.
The Vocabulary attached to these lecture notes has been prepared by Dr. Rdiger
Alt. It has been used by students at TU Clausthal attending lectures on Heat
Transfer, Combustion Technology and High Temperature Processes.
Although I have made a concerted effort to make this first edition error free, some
mistakes may have crept in unbidden. I would appreciate hearing from anyone
who finds an error or wishes to comment on the text. You may e-mail or write to
me.
Roman Weber
IEVB TU Clausthal
D-38678 Clausthal-Zellerfeld
Agricolastrasse 4
Germany
roman.weber@ievb.tu-clausthal.de

vi

Contents
1 Stoichiometry
1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.1 Chemical Reactions, Atoms and Molecules in Combustion .
1.2.2 Amount of Substances, Mole and Mass Fractions . . . . . .
1.2.3 Density and Concentration (Molar density) . . . . . . . . .
1.2.4 Equation of State for Gases and Gas Mixtures . . . . . . . .
1.3 Combustion Stoichiometry for Gaseous Fuels . . . . . . . . . . . .
1.3.1 Stoichiometric combustion . . . . . . . . . . . . . . . . . . .
1.3.2 Excess air ratio (air equivalence ratio) and fuel equivalence
ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3.3 Minimum air requirement for a mixture of gaseous fuels . .
1.3.4 Composition of combustion products . . . . . . . . . . . . .
1.4 Combustion stoichiometry for liquid and solid fuels . . . . . . . . .
1.4.1 Minimum oxygen and air requirements and excess air ratio
1.4.2 Combustion products . . . . . . . . . . . . . . . . . . . . .
1.5 Humid Combustion Air . . . . . . . . . . . . . . . . . . . . . . . .
1.5.1 Absolute and relative humidity . . . . . . . . . . . . . . . .
1.5.2 Dew Point Temperature of Combustion Products . . . . . .
1.6 Combustibles burnout for solid fuels . . . . . . . . . . . . . . . . .
1.7 Sub-stoichiometric combustion to carbon dioxide and water vapour
1.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

9
11
11
16
17
17
24
24
29
31
32
34

2 Mass and Energy Balance


2.1 General Formulation of Mass and Energy Balance . . . . . . . . .
2.1.1 Mass and Energy Balance at an Instant . . . . . . . . . .
2.1.2 Mass and Energy Balance over a Time Interval . . . . . .
2.1.3 Mass and Energy Balance under Steady-State Conditions
2.1.4 Example of a Mass Balance of a Furnace . . . . . . . . . .
2.2 The First Law of Thermodynamics . . . . . . . . . . . . . . . . .
2.2.1 System Energy . . . . . . . . . . . . . . . . . . . . . . . .
2.2.2 Energy Entering and Leaving the System . . . . . . . . .

35
35
36
37
38
38
43
46
49

.
.
.
.
.
.
.
.

1
1
4
4
5
6
7
8
8

vii

Contents

2.3

2.4
2.5
2.6

2.2.3 Energy Balance of Thermal Systems (Machines) . . . . . . .


Energy Released in Chemical Reactions . . . . . . . . . . . . . . .
2.3.1 Reaction Enthalpy . . . . . . . . . . . . . . . . . . . . . . .
2.3.2 Standard Enthalpies of Formation . . . . . . . . . . . . . .
2.3.3 Lower Calorific Value (LCV) and Gross Calorific Value (GCV)
2.3.4 Relationships between Calorific Values, Reaction Enthalpies
and Formation Enthalpies . . . . . . . . . . . . . . . . . . .
2.3.5 Dependence of LCV on Temperature . . . . . . . . . . . . .
2.3.6 Example of an Energy Balance of a Furnace . . . . . . . . .
Temperature of Adiabatic Combustion . . . . . . . . . . . . . . . .
Furnace Exit Temperature . . . . . . . . . . . . . . . . . . . . . . .
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

3 Equilibrium Thermodynamics
3.1 Irreversible and Reversible Processes . . . . . . . . . . . . . . . . .
3.2 Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1 Entropy of Liquids and Solids . . . . . . . . . . . . . . . . .
3.2.2 Entropy of Ideal Gases . . . . . . . . . . . . . . . . . . . . .
3.2.3 Entropy of Phase Transition at the Transition Temperature
3.2.4 The Third Law of Thermodynamics . . . . . . . . . . . . .
3.2.5 Absolute Entropy of Pure Substances . . . . . . . . . . . . .
3.3 The Second Law of Thermodynamics . . . . . . . . . . . . . . . . .
3.3.1 The Increase in Entropy Principle . . . . . . . . . . . . . .
3.3.2 Entropy Change for a Continuous Process at Steady-State .
3.3.3 Irreversibility of Processes . . . . . . . . . . . . . . . . . . .
3.4 General Conditions for Thermodynamic Equilibrium . . . . . . . .
3.4.1 Isolated System . . . . . . . . . . . . . . . . . . . . . . . . .
3.4.2 Non-Adiabatic System . . . . . . . . . . . . . . . . . . . . .
3.5 Equilibrium Between Phases . . . . . . . . . . . . . . . . . . . . . .
3.5.1 Single-Component System Consisting of Two Phases . . . .
3.5.2 Phase Transformations of a Pure Substance . . . . . . . . .
3.5.3 Dependence of Gibbs Free Enthalpy on Temperature and
Pressure . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.5.4 Equilibrium in Multi-Component Single-Phase Systems . .
3.5.5 Chemical Potential of Pure Substances . . . . . . . . . . . .
3.5.6 Significance of Chemical Potential . . . . . . . . . . . . . .
3.6 Multi-Component, Multi-Phase Systems . . . . . . . . . . . . . . .
3.6.1 The Phase Rule . . . . . . . . . . . . . . . . . . . . . . . . .
3.7 Thermodynamics of Mixing . . . . . . . . . . . . . . . . . . . . . .

viii

56
60
60
62
65
67
69
72
79
81
86
89
90
96
102
102
103
104
104
107
107
110
111
112
112
113
116
116
119
122
126
130
131
134
138
138

Contents
3.8

Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

4 Chemical Equilibrium
149
4.1 Definition of Chemical Equilibrium . . . . . . . . . . . . . . . . . . 150
4.2 Single Chemical Reaction . . . . . . . . . . . . . . . . . . . . . . . 150
4.2.1 Extent of a Single Reaction . . . . . . . . . . . . . . . . . . 150
4.2.2 Change of Gibbs Enthalpy as a Chemical Reaction Advances153
4.2.3 Gibbs Enthalpy of Selected Reactions . . . . . . . . . . . . 156
4.2.4 Thermodynamic Equilibrium Constant for a Gaseous Reaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
4.2.5 Other Equilibrium Constants . . . . . . . . . . . . . . . . . 165
4.2.6 Effect of Pressure and Temperature on Thermodynamic
Equilibrium Constant . . . . . . . . . . . . . . . . . . . . . 167
4.2.7 Chemical Equilibrium in Presence of a Solid Phase . . . . . 171
4.2.8 Le Chteliers Principle . . . . . . . . . . . . . . . . . . . . 176
4.3 Multiplicity of Chemical Reactions . . . . . . . . . . . . . . . . . . 183
4.3.1 Multi-Component, Multi-Phase Systems with Chemical Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
4.3.2 Choice of Chemical Reactions . . . . . . . . . . . . . . . . . 188
4.3.3 Exact Number of Chemical Reactions Needed for Equilibrium Determination . . . . . . . . . . . . . . . . . . . . . . 189
4.3.4 Linear Dependence of a Reaction Set . . . . . . . . . . . . . 193
4.3.5 The Phase Rule for a System with Chemical Reactions . . . 196
4.4 Equilibrium Composition . . . . . . . . . . . . . . . . . . . . . . . 200
4.4.1 Systems with a one-dimensional reaction basis . . . . . . . . 200
4.4.2 Systems with a two-dimensional reaction basis . . . . . . . 209
4.4.3 Systems with a multi-dimensional reaction basis . . . . . . . 216
4.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
5 Elements of Chemical Kinetics
5.1 Introduction . . . . . . . . . . . . . . . . . . . . . .
5.2 Rate Laws and Reaction Orders . . . . . . . . . . .
5.3 Forward and Reverse Reactions . . . . . . . . . . .
5.4 Elementary Reactions and Reaction Molecularity .
5.5 Rate of Reactions . . . . . . . . . . . . . . . . . . .
5.5.1 Temperature dependence of rate coefficients
5.5.2 Pressure dependence of rate coefficients . .
5.6 Summary . . . . . . . . . . . . . . . . . . . . . . .
6 Mechanisms of Basic Combustion Reactions

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.

221
. 221
. 222
. 226
. 231
. 237
. 238
. 240
. 241
243

ix

Contents
6.1
6.2
6.3
6.4
6.5

6.6

Chain Reactions . . . . . . . . . . . . . .
Combustion of Carbon Monoxide (CO) . .
Combustion of Hydrogen (H2 ) . . . . . . .
6.3.1 Simplified ignition mechanism . . .
Combustion of Methane (CH4 ) . . . . . .
Methods of Solving Chemical Kinetic Rate
6.5.1 Analytical solutions . . . . . . . .
6.5.2 Numerical Solutions . . . . . . . .
Summary . . . . . . . . . . . . . . . . . .

. . . . . .
. . . . . .
. . . . . .
. . . . . .
. . . . . .
Equations
. . . . . .
. . . . . .
. . . . . .

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.

243
245
246
248
251
253
254
264
279

Gaussian Elimination

283

Vocabulary

287

List of Figures
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.8

1.9
1.10
1.11
1.12
1.13
2.1
2.2
2.3
2.4
2.5
2.6
2.7
2.8
2.9
2.10

Primary Energy Sources in Germany per type of fuel . . . . . . . .


Wall Fired Boiler . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Modern Reheating Furnace NKK, Fukuyama Works, Japan . . .
Left pulverised coal flame; Right MILD combustion of natural
gas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
The subsets of hydrocarbons . . . . . . . . . . . . . . . . . . . . . .
Oxygen concentration in dry combustion products as a function of
excess air ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Carbon dioxide concentration as a function of excess air ratio . . .
Saturation pressure of water vapour as a function of temperature.
The plot is obtained using Eq. (1.35) which in the plotted range
provides pressure values that are within 2 % accuracy with the values listed in steam tables [7]. . . . . . . . . . . . . . . . . . . . . .
Cooling of combustion products (or moist air) at a constant temperature. The T -s diagram shows the dew-point temperature. . . .
Effect of sulphur and excess air on acid due point for a crude oil
(adapted from [8]). . . . . . . . . . . . . . . . . . . . . . . . . . . .
Ash and combustibles in a furnace . . . . . . . . . . . . . . . . . .
Burnout of combustibles . . . . . . . . . . . . . . . . . . . . . . . .
Composition of dry combustion products for sub-stoichiometric
combustion of methane . . . . . . . . . . . . . . . . . . . . . . . . .
Control volume for mass and energy balance . . . . . . . . . . . .
Mass balance for the boiler . . . . . . . . . . . . . . . . . . . . .
Illustration of open, closed and isolated systems . . . . . . . . . .
Sankeys diagram for energy balance for an open or closed system
System in translational and rotational motion . . . . . . . . . . .
Definition of the sign of work . . . . . . . . . . . . . . . . . . . .
Work done to the system . . . . . . . . . . . . . . . . . . . . . . .
Specific heats at constant pressure for various molecules . . . . .
Physical enthalpies of various gases as a function of temperature
Illustration of enthalpy of formation of a compound . . . . . . . .

.
.
.
.
.
.
.
.
.
.

2
3
3
4
5
23
23

26
30
30
31
32
34
36
39
44
45
46
51
52
57
58
62

xi

List of Figures
2.11 Illustration of LCV as an amount of heat extracted from a combustion chamber . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.12 Illustration of an energy balance . . . . . . . . . . . . . . . . . .
2.13 Example of an energy balance . . . . . . . . . . . . . . . . . . . .
2.14 Energy balance using enthalpy of formation . . . . . . . . . . . .
2.15 Energy balance using LCV . . . . . . . . . . . . . . . . . . . . . .
2.16 Sankey diagram demonstrating the concept of the available heat .
2.17 Fraction of the available heat as a function of the furnace exit
temperature and excess air ratio for combustion of pure methane
in air. Constant cp values. . . . . . . . . . . . . . . . . . . . . . .
2.18 Fraction of the available heat as a function of the furnace exit
temperature and excess air ratio for combustion of pure methane
in air. JANAF polynomials for cp have been used. . . . . . . . . .
3.1
3.2
3.3
3.4
3.5
3.6
3.7
3.8
3.9
3.10
3.11
3.12
3.13
3.14
3.15
3.16
3.17
3.18
3.19

xii

.
.
.
.
.
.

66
73
75
77
78
82

. 85

. 86

Examples of irreversible processes in closed and open systems . . .


Irreversible process of mixing . . . . . . . . . . . . . . . . . . . . .
Typical thermodynamic processes shown using work-diagram . . .
Reversible and irreversible gas expansion process . . . . . . . . . .
Integration paths from i f . . . . . . . . . . . . . . . . . . . . .
Determination of absolute specific entropy of a pure substance . . .
An open system interacting with other bodies by exchanging mass,
heat and work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Heat source providing heat to the system . . . . . . . . . . . . . .
Mass source providing mass to the system . . . . . . . . . . . . . .
A single component system consisting of two phases maintained at
constant temperature and pressure . . . . . . . . . . . . . . . . . .
Pressure-temperature plot showing the phase-equilibrium curve that
defines stability regions for phases 1 and 2 . . . . . . . . . . . . . .
The experimentally determined phase diagram for water . . . . . .
The variation of the Gibbs enthalpy with temperature for ice, water
and water vapour (at 1 bar . . . . . . . . . . . . . . . . . . . . . . .
The variation of the Gibbs enthalpy with temperature for ice, water
and water vapour at a pressure of 611 Pa . . . . . . . . . . . . . . .
Mixing of two ideal gases . . . . . . . . . . . . . . . . . . . . . . .
Mixture of two ideal gases A and B . . . . . . . . . . . . . . . . . .
Pure species-A at temperature T and pressure p; Species-A in a
mixture with species B at Temperature T and pressure p . . . . . .
Ideal solution of two liquids in equilibrium with its vapour . . . . .
Ideal solution of two liquids A and B; different vapour pressures . .

90
91
94
95
97
106
108
109
110
116
118
120
125
125
139
140
142
143
145

List of Figures
4.1
4.2

Minimisation of Gibbs enthalpy as the reaction advances . . . . .


Thermodynamic equilibrium constant K as a function of temperature for reactions listed in Table 4.2 . . . . . . . . . . . . . . . .
4.3 Equilibrium partial pressure of carbon dioxide in calcination reaction CaCO3 CaO + CO2 . . . . . . . . . . . . . . . . . . . . .
4.4 Accurate and estimated values of the thermodynamic equilibrium
constant for Boudouard reaction . . . . . . . . . . . . . . . . . .
4.5 Variation of the partial pressures of CO2 and CO with temperature
for total pressure of 1 bar and 10 bar for Boudouard reaction . .
4.6 Equilibrium composition of Boudouard reaction at several total
pressures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.7 Illustration of reaction (4.128) . . . . . . . . . . . . . . . . . . . .
4.8 Gibbs enthalpy as a function of extent of the reaction . . . . . . .
4.9 Gibbs enthalpy as a function of the amount of carbon dioxide . .
4.10 Gibbs enthalpy of the considered system as a function of mCH4 and
mH2 O . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
5.1
5.2
5.3
5.4
5.5
5.6
6.1
6.2
6.3
6.4

6.5

6.6

. 155
. 169
. 175
. 180
. 182
.
.
.
.

183
203
207
209

. 215

Concentration change with time for a first-order reaction (t0 = 0) .


Concentration change with time for a second-order reaction . . . .
The approach of concentrations to their equilibrium values for the

reversible reaction A

B that is first order in each direction . .


Activation energy of a chemical reaction . . . . . . . . . . . . . . .
Arrhenius plot for elementary reactions of halogens with molecular
hydrogen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Fall-off curves for the reaction C2 H6 CH3 + CH3 . . . . . . . .

223
225
229
239
240
241

Top rate constant of CO + OH CO2 + H;


Bottom rate constant of H + CO2 CO + OH [19]. . . . . . . . 246
The time behaviour of the chain carriers . . . . . . . . . . . . . . . 251
A simplified scheme of methane oxidation [19] . . . . . . . . . . . . 253
Temporal behaviour of the species concentrations in reactions A1
A2 A3 for k12 = 1 s1 and k23 = 10 s1 (reactive intermediate
A2 ) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
Temporal behaviour of the species concentrations in reactions A1
A2 A3 for k12 = 10 s1 and k23 = 1 s1 (low reactive intermediate A2 ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
Temporal behaviour of the species concentrations for reactions A1
A2 A3 with k12 = k23 = k. . . . . . . . . . . . . . . . . . . . . 259

xiii

List of Figures
6.7

6.8
6.9
6.10
6.11
6.12
6.13
6.14

xiv

Temporal behaviour of the species concentrations for reactions A1


A2 A3 assuming quasi steady state for A2 -species; k12 = 1 s1 ,
k23 = 10 s1 . (to be compared with Fig. 6.4) . . . . . . . . . . . . . 260
Eulers method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
Numerical solutions using the explicit scheme with different time
steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
Numerical solutions using the implicit scheme with different time
steps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
Temporal behaviour of the species concentrations for reactions A1
A2 A3 assuming quasi steady state for A2 -species . . . . . . . . 271
Temporal behaviour of the species concentrations for reactions A1
A2 A3 , k12 = 1 s1 , k23 = 10 s1 , Eulers Explicit Method . . . 275
Temporal behaviour of the species concentrations for reactions A1
A2 A3 , k12 = 1 s1 , k23 = 100 s1 , Eulers Explicit Method . . 276
Temporal behaviour of the species concentrations for reactions A1
A2 A3 , Eulers Implicit Method . . . . . . . . . . . . . . . . . 277

List of Tables
1.1
1.2

Names of aliphatic hydrocarbons . . . . . . . . . . . . . . . . . . . 5


Composition of combustion products . . . . . . . . . . . . . . . . . 12

2.1
2.2
2.3
2.4
2.5
2.6

The incoming and out-coming streams in kmol/h . . . . . . . . .


Composition of the combustion products . . . . . . . . . . . . . .
The incoming and outcoming flow rates in kg/h . . . . . . . . . .
The coefficients for polynomials for T in the range 3001000 K . .
The coefficients for polynomials for T in the range 10005000 K .
Standard enthalpies of formation and standard entropies of some
compounds (JANAF Thermodynamic Tables) . . . . . . . . . . . .
LCV of some selected gaseous fuels . . . . . . . . . . . . . . . . .
Adiabatic flame temperature Tad for stoichiometric combustion in
air. Combustion products contain CO2 and H2 O only. . . . . . .
Calculated adiabatic temperature for stoichiometric combustion of
pure CH4 with air. . . . . . . . . . . . . . . . . . . . . . . . . . .

2.7
2.8
2.9

.
.
.
.
.

41
41
43
57
58

. 64
. 68
. 80
. 85

3.1

Transition temperatures, standard enthalpies and standard entropies


of selected substances . . . . . . . . . . . . . . . . . . . . . . . . . 104

4.1

4.4
4.5

Standard enthalpies of formation, standard entropies and Gibbs


formation enthalpies of some selected compounds . . . . . . . . .
Thermodynamic equilibrium constant K for some reactions important in combustion . . . . . . . . . . . . . . . . . . . . . . . . . .
Thermodynamic equilibrium constant K for some reactions important in combustion . . . . . . . . . . . . . . . . . . . . . . . . . .
Standard Gibbs enthalpies at T = 1300 K . . . . . . . . . . . . .
Standard Gibbs enthalpies at T = 1000 K . . . . . . . . . . . . .

5.1

Elementary reactions in H2 COC1 O2 system . . . . . . . . . . 233

6.1
6.2

Rate coefficients of the chain initiation reactions . . . . . . . . . . 247


Rate coefficients of the chain branching reactions.
The coefficients

Ea
b
are presented in the form k = A T exp R T . . . . . . . . . . . . 248

4.2
4.3

. 158
. 162
. 163
. 204
. 213

xv

List of Tables
6.3

6.6
6.7
6.8

Comparison of the rate coefficients k of the chain initiation and


chain branching reactions. . . . . . . . . . . . . . . . . . . . . . .
Simplified mechanism of ignition of the hydrogen-oxygen system .
Comparison of the rate coefficients k of the chain propagation reactions for methane oxidation. . . . . . . . . . . . . . . . . . . . .
Rate coefficients of the Zeldovich mechanism reactions . . . . . .
Numerical solution using the explicit scheme with h = 0.3 s. . . .
Numerical solution using the implicit scheme with h = 0.3 s. . . .

6.9

Technical vocabulary . . . . . . . . . . . . . . . . . . . . . . . . . . 287

6.4
6.5

xvi

. 249
. 249
.
.
.
.

252
261
266
269

1 Stoichiometry
Contents
1.1
1.2

1.3

1.4

1.5

1.6
1.7
1.8

Introduction
Definitions
1.2.1 Chemical Reactions, Atoms and Molecules in Combustion
1.2.2 Amount of Substances, Mole and Mass Fractions
1.2.3 Density and Concentration (Molar density)
1.2.4 Equation of State for Gases and Gas Mixtures
Combustion Stoichiometry for Gaseous Fuels
1.3.1 Stoichiometric combustion
1.3.2 Excess air ratio (air equivalence ratio) and fuel equivalence ratio
1.3.3 Minimum air requirement for a mixture of gaseous fuels
1.3.4 Composition of combustion products
Combustion stoichiometry for liquid and solid fuels
1.4.1 Minimum oxygen and air requirements and excess air ratio
1.4.2 Combustion products
Humid Combustion Air
1.5.1 Absolute and relative humidity
1.5.2 Dew Point Temperature of Combustion Products
Combustibles burnout for solid fuels
Sub-stoichiometric combustion to carbon dioxide and
water vapour
Summary

1.1 Introduction
Nowadays combustion provides more than 90 % of the energy sources. Despite the
continuing search for alternative energy sources combustion will remain important

1 Stoichiometry
for many decades to come, as shown in Fig. 1.1. For most of high temperature
processes like power generation, glass manufacturing, cement-making and steelmaking, combustion of fossil fuels provides the process energy. Any material that
can be burned to release thermal energy is called a fuel. Most fuels consist primarily of hydrogen and carbon and they are called hydrocarbons. Fig. 1.2 shows
an interior of a combustion chamber of a power plant producing electricity from
a hard coal. A sketch of a modern reheating furnace of Fukuyama steel-work in
Japan [2] is shown in Fig. 1.3. The furnace is fired with a mixture of natural
gas and steel-work gases. Petrol and diesel oil are burned in engines used for
transportation. Propeller-driven airplanes are built with propellers powered by
engines identical to automobile engines. Civilian and military aircrafts are powered on energy generated through combustion processes occurring in gas turbine
combustors. In short, combustion occurs in our every day life.

Fig. 1.1: Primary Energy Sources in Germany per type of fuel in EJ = 1018 J [1]

Industrial flames are essential elements of the processes and flame properties often
affect not only the process efficiency but also product quality. Fig. 1.4 (left) shows
a typical pulverisedcoal flame for boiler application while Fig. 1.4 (right) shows
MILD combustion technology (known also as flameless oxidation) applied to a
reheating furnace. Designing efficient combustion processes in car engines and
gas turbines has been an on-going challenge for combustion engineers.

1.1 Introduction

Fig. 1.2: Wall Fired Boiler

Fig. 1.3: Modern Reheating Furnace NKK, Fukuyama Works, Japan [2]

1 Stoichiometry

Fig. 1.4: Left pulverised coal flame; Right MILD combustion of natural gas

1.2 Definitions
1.2.1 Chemical Reactions, Atoms and Molecules in Combustion
A chemical reaction is an exchange and/or rearrangement of atoms between colliding molecules, for example:
H2 + 21 O2 H2 O
The atoms are conserved (they are not created or destroyed) while molecules are
not conserved. In the above reaction H, O atoms are conserved while molecules
H2 , O2 and H2 O are not. Reactant molecules (H2 and O2 ) are rearranged to
become product (H2 O) molecules. Heat is released in this process.
Atoms relevant in combustion are: C, H, O, N, S, Cl. Compounds of carbon and
hydrogen are called hydrocarbons. Hydrocarbons are classified into: aliphatic
hydrocarbons, alicyclic hydrocarbons and aromatic hydrocarbons. The first ten
members of the unbranched-chain alkane series are:
CH4
C2 H6
C3 H8
C4 H10
C5 H12

methane
ethane
propane
butane
pentane

C6 H14
C7 H16
C8 H18
C9 H20
C10 H22

hexane
heptane
octane
nonane
decane

1.2 Definitions

Fig. 1.5: The subsets of hydrocarbons

Table 1.1: Names of aliphatic hydrocarbons


No.
of C
atoms

alkane

1
2
3
4
5
n

CH4
methane
C2 H6
ethane
C3 H8
propane
C4 H10 butane
C5 H12 pentane
Cn H2n+2

alkene

C2 H4
C3 H6
C4 H8
C5 H10
Cn H2n

ethene
propene
butene
pentene

alkyne

C2 H2
C3 H4
C4 H6
C5 H8
Cn H2n2

alkyl group

ethyne
propyne
butyne
pentyne

CH3
C2 H5
C3 H7
C4 H9
C5 H11
Cn H2n+1

methyl
ethyl
propyl
butyl
pentyl

Other molecules relevant in combustion are:


Haloalkanes
Alcohols
Aldehydes
Amines
Ketones
Carbocyclic Acids

RX
ROH
RCHO
RNH2
RCOR
RCOOH

example:
example:
example:
example:
example:
example:

CH3 Cl
C2 H5 OH
CH3 CHO
CH3 NH2
CH3 COCH3
CH3 COOH

(chloromethane)
(ethanol)
(ethanal)
(methylamin)
(acetone)
(ethanoic acid)

1.2.2 Amount of Substances, Mole and Mass Fractions


Atoms and molecules are counted in amount of substances or moles. 6.023 1023
particles (atoms, molecules) are called one mole of the substance (Avogadro constant is NA = 6.023 1023 atoms or molecules per mole).

1 Stoichiometry
For a mixture of species:
X
n=
ni

n = total number of moles

(1.1)

where n stands for total number of moles, ni is the number of moles of species i,
and the summation extends over all the species.
Mole fraction xi (mole number) of species i is:
xi =

ni
n

(1.2)

The molar mass (molecular weight) in g/mol or kg/kmol is the mass of one mole
of the species (for example: MC = 12 g/mol, MCO2 = 44 g/mol). The mean molar
mass (molecular weight) of a mixture is:
X
Mmean =
xi Mi
(1.3)
i

Frequently mole fractions xi are converted into mass fractions (wi ) and the following relationships hold:
wi =

xi Mi
n i Mi
number of kg of species "i"
=P
=P
n k Mk
xk Mk
total number of kg in the system
k

ni
number of moles of species "i" in 1 kg of mixture
= xi =
n
total number of moles in 1 kg of mixture
wi
wi
wi
Mi
Mi
P
=
Mmean =
=
wk
1
Mi
Mk
M
mean

(1.4)

(1.5)

where the summation extends over all the species and wi stands for mass fraction
of species i.

1.2.3 Density and Concentration (Molar density)


Variables that do not depend on the size (extent) of the system are called intensive
variables (for example density, molar density). These are defined as the ratio of
the corresponding extensive properties and the system volume V ;
density
molar density (concentration)

=
c=

m
V
n
V

(in kg/m3 )
(in kmol/m3 )

1.2 Definitions
Thus,

=
= Mmean
(1.6)
c
n
Note that the molar density c defined above is in chemistry usually denoted in
square brackets [ ] for example cH2 = [H2 ] and chemists prefer to express concentrations in mol/cm3 .

1.2.4 Equation of State for Gases and Gas Mixtures


For gases and gas mixtures an equation of state relates temperature, pressure and
volume:
F (p, T, V ) = 0
(1.7)
There are several equations of state for gases and gas mixtures [3, 4, 5]. The
perfect gas equation of state, called also Clausius-Clapeyron equation, is perhaps
one of the simplest ones and it reads
(1.8)

pV = nRT
or
c=
and
=

p
RT

p Mmean
p
P
=
RT
RT
i

(1.9)

wi
Mi

(1.10)

where p is the pressure (in Pa = N/m2 ), V the volume (in m3 ), n the molar
number (in kmol), T the absolute temperature (in K), and R is the universal
gas constant R = 8314.3 J/(kmol K). An ideal gas is an imaginary substance
that obeys relation (1.8). It has been experimentally observed that relation (1.8)
approximates closely behaviour of real gases at low densities which means at
low pressures and at high temperatures. What are a low pressure and a high
temperature? The pressure or temperature of a substance is high or low relative
to its critical pressure (pcr ) or critical temperature (Tcr ). The useful rules of
thumbs are:
(a) at very low pressures ( ppcr 1), gases behave as an ideal gas regardless of
temperature;
(b) at high temperatures ( TTcr > 2) gases behave also as an ideal gas;
(c) deviation from an ideal gas behavior increases upon approaching the critical
point.

1 Stoichiometry
From Eq. (1.8) follows that one kmol of any (ideal) gas under constant temperature and pressure occupies the same volume. One can easily verify that under so
called normal conditions: a pressure of p0 = 1 bar (105 N/m2 )and a temperature
of T0 = 25 C (298.15 K) a 1 kmol of any ideal gas occupies a volume of 24.79 m3 .
It is important to realisethat normal cubic meter (Nm3 or m3n ) is a unit of mass
(substance) not volume. Thus,
1 kmol of gas = 24.79 m3n
1 mol of gas = 24.79 dm3n
Throughput this lecture course the standard (normal) conditions of p0 = 1 bar
(105 N/m2 ) and T0 = 25 C (298.15 K) are used1 .

1.3 Combustion Stoichiometry for Gaseous


Fuels
1.3.1 Stoichiometric combustion
Combustion is said to be stoichiometric if fuel and oxidiserconsume each other
completely forming only carbon dioxide (CO2 ) and water (H2 O). If there is an
excess of fuel, the system is fuel-rich, and if there is an excess of oxygen, it is
called fuel-lean.
Examples are:
stoichiometric

CH4 + 2 O2 2 H2 O + CO2
CH4 + 3 O2 2 H2 O + CO2 + O2
CH4 + O2 H2 O +

1
2 CO2

1
2 CH4

lean (excess of oxygen)


rich (excess of fuel)

If the reaction describing complete combustion (products are CO2 and H2 O) is


written in such a way that it describes the combustion of 1 kmol of fuel:
1 kmol fuel + O2 products (CO2 + H2 O)
one may easily calculate the mole fraction of fuel in a stoichiometric mixture (with
1

In older books p0 = 760 Tr (1 Tr = 1 mmHg = 133.322 N/m2 ) and T0 = 0 C (273.15 K) are


used so that 1 kmol of ideal gas occupies a volume of 22.41 m3 .

1.3 Combustion Stoichiometry for Gaseous Fuels


oxygen) as follows:
xf uel,stoich in oxygen =

1
number of moles of fuel
=
total number of moles (fuel + oxygen)
1+

(1.11)

For example:
2
1
=
1.5
3
1
xCH4 ,stoich in oxygen =
3
2
1
=
xCO,stoich in oxygen =
1.5
3

H2 + 21 O2 H2 O

xH2 ,stoich in oxygen =

CH4 + 2 O2 CO2 + 2 H2 O
CO + 12 O2 CO2

If dry air is used as an oxidiser it contains only 21 % O2 , 78 % N2 , and 1 % of


noble gases. Thus,
xf uel,stoich in air =

1
1
=
1 + 0.21
1 + 4.762

(1.12)

An example:
Combustion of propane in oxygen:
C3 H8 + 5 O2 3 CO2 + 4 H2 O

and

xC3 H8 ,stoich in oxygen =

1
6

Combustion of propane in air:


C3 H8 + 5 ( O2 + 3.762 N2 ) 3 CO2 + 4 H2 O + 5 3.762 N2
and

xC3 H8 ,stoich in air =

1
= 0.0403
1 + 5 4.762

Note: The higher the hydrocarbon the lower is the fuel mole fraction at stoichiometric conditions.

1.3.2 Excess air ratio (air equivalence ratio) and fuel equivalence
ratio
The excess air ratio is defined as:
=

(wair /wf uel )


(xair /xf uel )
=
(xair /xf uel )stoich
(wair /wf uel )stoich

(1.13)

1 Stoichiometry
whilst the fuel equivalence ratio is defined as: = 1/. The above equation
can be rewritten to allow the calculation of the mole fuel fraction in a mixture of
known or :
1
;
1 + 4.762
xair
;
= 0.21 xair =
4.762

xf uel =

xair = 1 xf uel ;

(1.14)

xO2

xN2 = 3.762 xO2

(1.15)

The combustion process can be divided into:


Rich combustion

>1

<1

Stoichiometric combustion

=1

=1

Lean combustion

<1

>1

Example 1.1
Five hundred
of ethane are combusted with five thousand five hundred
m3n /h of dry air. Calculate the excess air ratio and the fuel equivalence ratio.
Assumptions: the fuel (ethane) is combusted to carbon dioxide and water.
m3n /h

The combustion reaction is then


C2 H6 + 27 O2 2 CO2 + 3 H2 O
kmol air
Thus, (xair /xf uel )stoich = 3.5 4.7619/1 = 16.6667 kmol
fuel
3
500 mn /h of C2 H6 is equivalent to 500/24.79 = 20.169 kmol of C2 H6 /h
5500 m3n /h of dry air is equivalent to 5500/24.79 = 221.864 kmol of air/h
= 221.864/20.169 = 11 and = 11/16.6667 = 0.66 ( = 1.5152).
Thus, xxfair
uel
Comments:

(a) Since the combustion is fuel rich, (1 0.66) 500 = 170 m3n /h of ethane will
not be combusted.
(b) In the above example the conversion from m3n /h into kmol/h is superficial.
End of Example 1.1
Formula (1.11) is very useful and simple when a single fuel is combusted. However, if the gaseous fuel is a mixture of hydrogen, CO, hydrocarbons and other
more complex compounds and the air contains moisture, the calculations of excess
air ratio is more troublesome. Therefore in combustion engineering the excess air
ratio for mixture of technical gases is calculated using the minimum air requirement.

10

1.3 Combustion Stoichiometry for Gaseous Fuels

1.3.3 Minimum air requirement for a mixture of gaseous fuels


Composition of gaseous fuels is usually expressed in mol (volume) fractions. A
chemical analysis of a dry gaseous fuel includes the mol fractions of hydrogen,
carbon monoxide, carbon dioxide, methane, ethane, propane, higher hydrocarbons
as well as oxygen. The following applies:
xH2 + xCO + xCH4 + xCn Hm + xO2 + xCO2 + xN2 = 1

(1.16)

where xi s stand for mole fractions.


The following combustion reactions are applicable:
H2 + 12 O2 H2 O
CO + 21 O2 CO2
CH4 + 2 O2 CO2 + 2 H2 O
C2 H6 + 27 O2 2 CO2 + 3 H2 O
Cn Hm + (n +

m
4 )O2

n CO2 +

m
2 H2 O

Thus, minimum oxygen requirement for a mixture of gases is:


lO2 ,min =

X
1
1
m
xCn Hm xO2
xH2 + xCO + 2 xCH4 +
n+
2
2
4
(lO2 ,min in kmol O2 /kmol dry gaseous fuel)

(1.17)

whilst the minimum dry air requirement is:


ldry

air,min

lO2 ,min
= 4.7619 lO2 ,min
0.21

(1.18)

Thus, the excess air ratio is


=

ldry air
amount of dry air supplied per kmol of fuel
=
minimum dry air requirement per kmol of fuel
ldry air,min

(1.19)

1.3.4 Composition of combustion products


The table below shows the number of moles of combustion products produced in
complete combustion of gaseous fuels: Two types of combustion products are considered in combustion engineering namely wet and dry products. The minimum

11

1 Stoichiometry
Table 1.2: Composition of combustion products

Fuel
Component of
component CO2 H2 O
H2

1
CO
1

CH4
1
2
C2 H6
2
3
Cn Hm
n
m/2
O2

N2

combustion products
N2
O2

amount (at stoichiometric conditions) of wet combustion products is:


Vwet,min = 1 xH2 + 1 xCO + 3 xCH4 + 5 xC2 H6 +
X
m
xCn Hm + 1 xN2 + 0.79 lair,min
n+
2
(Vwet,min is in kmol wet products/kmol fuel)
while for lean combustion:
Vwet = Vwet,min + ( 1) lair,min
The minimum (at stoichiometric conditions) amount of dry combustion products
is then
Vdry,min = 1 xCO + 1 xCH4 + 2 xC2 H6 +

(n xCn Hm ) + 1 xN2 + 0.79 lair,min

(Vdry,min is in kmol dry products/kmol fuel)


while for lean combustion:
Vdry = Vdry,min + ( 1) lair,min

12

1.3 Combustion Stoichiometry for Gaseous Fuels


Composition of combustion products (wet):
1 xCO + 1 xCH4 + 2 xC2 H6 + n xCn Hm
Vwet
1 xH2 + 2 xCH4 + 3 xC2 H6 + 21 m xCn Hm
=
Vwet
1 xN2 + 0.79 lair,min
=
Vwet
0.21 ( 1) lair,min
=
Vwet

xCO2 ,wet =
xH2 O,wet
xN2 ,wet
xO2 ,wet

kmol CO2
kmol wet products
kmol H2 O
kmol wet products
kmol N2
kmol wet products
kmol O2
kmol wet products

Composition of combustion products (dry):


The amount of dry combustion products per 1 kmol of fuel can be calculated as:


Vdry = Vwet 1 xH2 O,wet = Vwet 1 xH2 + 2 xCH4 + 3 xC2 H6 + 12 m xCn Hm
in kmol dry products/kmol of fuel

Thus, composition of dry combustion products can be calculated as follows:


1 xCO + 1 xCH4 + 2 xC2 H6 + n xCn Hm
Vdry
1 xN2 + 0.79 lair,min
=
Vdry
0.21 ( 1) lair,min
=
Vdry

xCO2 ,dry =
xN2 ,dry
xO2 ,dry

kmol CO2
kmol dry products
kmol N2
kmol dry products
kmol O2
kmol dry products

Example 1.2
Calculate composition of wet and dry products of combustion of Dutch (Groningen) natural gas at 20 % excess air [6].
Composition of Dutch natural gas:
CH4
C3 H8

81 (vol%)
1 (vol%)

C2 H6
N2

3 (vol%)
15 (vol%)

Assumptions: the fuel is combusted to carbon dioxide and water.

13

1 Stoichiometry
Thus, the relevant combustion reactions are:
CH4 + 2 O2 CO2 + 2 H2 O
7
2 O2

|| 0.81

2 CO2 + 3 H2 O

|| 0.03

C3 H8 + 5 O2 3 CO2 + 4 H2 O

|| 0.01

C2 H6 +

Minimum oxygen requirement:


lO2 ,min

kmol O2
= 0.81 2 + 0.03 3.5 + 0.01 5 = 1.775
kmol fuel

m3n of O2
m3n of fuel

Minimum air requirement:


lair,min =

kmol dry air


1.775
= 8.452
0.21
kmol fuel

The amount of air for = 1.2;


lair = 1.2 8.452 = 10.14

kmol dry air


kmol fuel

Combustion products:
Fuel
component
CH4
C2 H6
C3 H8
N2

Component of
combustion products
CO2 H2 O N2
O2
1
2

2
3

3
4

molar fraction
in dry fuel
xi
0.81
0.03
0.01
0.15

Minimum amount of wet combustion products (for = 1):


Vwet,min = 3 0.81 + 5 0.03 + 7 0.01 + 1 0.15 + 0.79 8.452 =
kmol wet products
9.477
kmol of fuel
The amount of combustion products for lean combustion (for 1)
Vwet = 9.477 + ( 1) 8.452 and for = 1.2 Vwet = 11.17

14

kmol wet products


kmol of fuel

1.3 Combustion Stoichiometry for Gaseous Fuels


Composition of wet combustion products:
xCO2 ,wet =

xH2 O,wet =

xN2 ,wet =

xO2 ,wet =

1 0.81 + 2 0.03 + 3 0.01


9.477 + ( 1) 8.542
and for = 1.2 one obtains xCO2 ,wet = 0.081
2 0.81 + 3 0.03 + 4 0.01
9.477 + ( 1) 8.542
and for = 1.2 one obtains xH2 O,wet = 0.156

1 0.15 + 8.452 0.79


9.477 + ( 1) 8.542
and for = 1.2 one obtains xN2 ,wet = 0.7309
( 1) 8.452 0.21
9.477 + ( 1) 8.542
and for = 1.2 one obtains xO2 ,wet = 0.0318

The amount of dry combustion products (for 1):


Vdry = 1 0.81 + 2 0.03 + 3 0.01 + 1 0.15 + 0.79 8.452 + ( 1) 8.452 =
7.7271 + ( 1) 8.452
and for = 1.2 one obtains Vdry = 9.4175 kmol products/kmol fuel
Composition of dry combustion products:
xCO2 ,dry =

xN2 ,dry =

1 0.81 + 2 0.03 + 3 0.01


7.727 + ( 1) 8.542
and for = 1.2 one obtains xCO2 ,dry = 0.0956

1 0.15 + 8.452 0.79


7.727 + ( 1) 8.542
and for = 1.2 one obtains xN2 ,dry = 0.8667

15

1 Stoichiometry

xO2 ,dry =

( 1) 8.452 0.21
and for = 1.2 one obtains xO2 ,dry = 0.0377
7.727 + ( 1) 8.542

Comments:
O2 , CO2 and N2 concentrations (dry and wet) are functions of excess air ratio.
Make a graph showing this dependence (see Fig. 1.6 and Fig. 1.7).
End of Example 1.2

1.4 Combustion stoichiometry for liquid and


solid fuels
For solid and liquid fuels the elemental composition (ultimate analysis) is usually
expressed in mass fraction (percentage). A complete fuel analysis includes then
mass fractions of carbon, hydrogen, sulphur, oxygen and nitrogen as elements,
and water (moisture) and ash as compounds.
The following applies:
c + h + o + n + s + moisture + ash = 1

(1.20)

where small letters stand for mass fractions. In the above relationship the mass
fractions of the elements (c, h, s, o, n) are as fired (sometimes called also as
received), it means water and ash are present in the fuel. Fuel composition
is often expressed on a dry basis (without water, but with ash) and then the
following conversion is applicable:
cdry =

cas received
1 moisture

(1.21)

When fuel composition is expressed on a dry ash-free basis the following conversion
is applicable:
cas received
(1.22)
cdry,ashf ree =
1 moisture ash
Obviously, similar conversions can be applied to hydrogen, sulfur, oxygen, and
nitrogen mass fractions.

16

1.4 Combustion stoichiometry for liquid and solid fuels

1.4.1 Minimum oxygen and air requirements and excess air ratio
The following reactions determine the minimum oxygen and air requirements:
C + O2 CO2
S + O2 SO2
H2 + 12 O2 H2 O
N + O2 NO2 2
Thus, the minimum oxygen requirements (in kmol O2 per kg of fuel as received)
is
lO2 ,min = 1

1 h
s
n
o
c
+ + 1
+ 1

12 2 2
32
28 32

kmol O2
kg of fuel as received

(1.23)

or
lO2 ,min



c
1 h
s
n
o
= 24.79 1
+ + 1
+ 1

12 2 2
32
28 32

m3n O2
kg of fuel as received

and the minimum air requirement is:


lair,min =

lO2 ,min
= 4.762 lO2 ,min
0.21

(1.24)

The excess air ratio is then calculable as:


=

ldry air
amount of dry air supplied per kg of fuel
=
minimum dry air requirement per kg of fuel
ldry air,min

(1.25)

1.4.2 Combustion products


Products of complete combustion of fuels contain CO2 , H2 O, SO2 , N2 and excess
O2 . If the temperature of the combustion products is above the dew point the
water vapor does not condense. Combustion products can be considered as dry
or wet:
Vwet = Vdry + VH2 O
(1.26)
2

It has been assumed that sulphur and nitrogen present in the fuel are oxidisedto SO2 andNO2 ,
respectively. In reality, other sulphur and nitrogen oxides also occur in combustion products.
However, since the nitrogen and sulphur contents in the fuel are low, such an assumption is
justifiable for calculating the minimum oxygen and air requirements.

17

1 Stoichiometry
where Vwet is the amount (in kmols or m3n ) of wet combustion products per kg of
fuel while Vdry is the amount of dry products and the VH2 O stands for the amount
of water in the products. The minimum amount (at stoichiometric combustion,
= 1) of wet combustion products is:
Vwet,min = 1

h moisture
s
n
c
+ 1 +
+ 1
+
+ 1 0.79 lair,min
12
2
18
32 28
in kmol wet products/kg of fuel (1.27)

and at lean combustion ( > 1):


h moisture
s
n
c
+ 1 +
+ 1
+
+
12
2
18
32 28
1 0.79 lair,min + ( 1) lO2 ,min + ( 1) 0.79 lair,min
c
h moisture
s
n
=
+ +
+
+
+ ( 1) lO2 ,min + 0.79 lair,min
12 2
18
32 28

Vwet = 1

Vwet

H2 O
N2
CO2
N2
moisture SO2
excess
from s from n oxygen in air The
from c from h in fuel
above expressions for calculating Vwet can be rearranged into a more elegant form
as follows:
c
h moisture
s
n
+ 1 +
+ 1
+
+ 0.79 lair,min + ( 1) lO2 ,min +
12
2
18
32 28
0.79 ( 1) lair,min
c
h moisture
s
n
=
+ +
+
+
+ 0.79 lair,min + 0.21 ( 1) lair,min +
12 2
18
32 28
0.79 ( 1) lair,min
h moisture
s
n
c
+ +
+
+
+ 0.79 lair,min + ( 1) lair,min
=
12 2
18
32 28
= Vwet,min + ( 1) lair,min

Vwet = 1

Thus,
Vwet = Vwet,min + ( 1) lair,min
and

c
h moisture
s
n
+ +
+
+
+ 0.79 lair,min
12 2
18
32 28
is the minimum amount of products (for = 1).

Vwet,min =
where Vwet,min

(1.28a)
(1.28b)

The amount of dry combustion products can be easily calculated by dropping the

18

1.4 Combustion stoichiometry for liquid and solid fuels


terms

h
2

moisture
18

and

Vdry =

from the equations above. Thus,

s
n
c
+
+
+ 0.79 lair,min + ( 1) lO2 ,min + ( 1) 0.79 lair,min
12 32 28
kmol dry products/kg of fuel (1.29a)

or
Vdry = Vdry,min + ( 1) lair,min

kmol dry products/kg of fuel (1.29b)

and
Vdry,min =

c
s
n
+
+
+ 0.79 lair,min
12 32 28
kmol dry products/kg of fuel (1.30)

Composition of the combustion products


The composition of the combustion products can be easily calculated realisingthat
wet products contain H2 O, CO2 , SO2 , N2 and excess O2 . Thus:
xH2 O =

h
2

Vwet

xCO2 ,wet =

c
12

Vwet

xSO2 ,wet =
xO2 ,wet
xN2 ,wet

moisture
18

s
32

Vwet
( 1) lO2 ,min
=
Vwet
n
+
0.79
lair,min
= 28
Vwet

kmol H2 O/kmol wet products


kmol CO2 /kmol wet products
kmol SO2 /kmol wet products
kmol O2 /kmol wet products
in kmol N2 /kmol wet products

Similarly one can derive simple relationships for calculating composition of dry
combustion products (CO2 , SO2 , N2 and excess O2 ):
xCO2 ,dry =
xO2 ,dry

c
12

Vdry
( 1) lO2 ,min
=
Vdry

xSO2 ,dry =
xN2 ,dry

s
32

Vdry
n
+ 0.79 lair,min
= 28
Vdry

where Vdry is the amount of dry combustion products in kmol/kg of fuel.

19

1 Stoichiometry
Example 1.3
A coal of the composition given in the table below is combusted with air at 10 %
excess air. Calculate the composition of the combustion products (wet and dry)
and produce a curve showing the CO2 and O2 mole fractions (dry) as a function
of excess air ratio.
Coal Fettnuss mvb3 (origin Germany) coal analysis obtained from a chemical
laboratory
Basis

Dry ash free


Dry
As fired

H2 O
%-wt

Ash
%-wt
c
89.39

Ultimate Analysis
%-wt
h
n
o
4.68
1.53
3.66

LCV
MJ/kg
s
0.74

4.4

32.87

3.5

Assumptions: the fuel is combusted to carbon dioxide and water.


One begins with calculating the coal composition as fired (with moisture and
ash). Simple conversions result in the following:
Basis

Dry ash free


Dry
As fired

H2 O
%-wt

3.5

Ash
%-wt

4.4
4.246

c
89.39
85.4568
82.4658

Ultimate Analysis
%-wt
h
n
o
4.68
1.53
3.66
4.47
1.4627 3.499
4.3136 1.4115 3.3765

LCV
MJ/kg
s
0.74
0.7074
0.6826

34.38
32.87
31.72

The minimum oxygen requirement ( = 1) is:


lO2 ,min = 1

0.006 826 0.033 765


0.824 658 1 0.043 136
+
+ 1

=
12
2
2
32
32
kmol O2
0.0787
kg fuel as fired

that corresponds to 0.0787 32 = 2.5184 kg


3

mvb - medium volatile bituminous

20

kg O2
of fuel as fired .

1.4 Combustion stoichiometry for liquid and solid fuels


The minimum air requirement ( = 1) is:
0.0787
kmol dry air
lair,min =
= 0.3743
0.21
kg of fuel as fired


2.5184
kg of dry air
m3n dry air
, or
= 10.9496
0.3743 22.79 = 9.279
kg of fuel
0.23
kg of fuel
The amount of combustion products is:
Vwet =

0.824 658 0.043 136 0.035 0.006 826 0.014 115


+
+
+
+
12
2
18
32
28
+ 0.79 0.3743 + ( 1) 0.0787

Vwet = 0.0929 + 0.2957 + 0.0787 ( 1)


in kmol wet products/kg of fuel as fired
and
Vdry =

0.824 658 0.006 826 0.014 115


+
+
+ 0.79 0.3743 + ( 1) 0.0787
12
32
28
= 0.0695 + 0.2957 + ( 1) 0.0787
in kmol dry products/kg of fuel as fired

Composition of wet combustion products:

xH2 O

xCO2 ,wet =

xN2 ,wet

Vwet
0.824 658
12
Vwet
0.006 83
32

0.035
18

0.0235
0.0929 + 0.2957 + 0.0787 ( 1)

0.0687
0.0929 + 0.2957 + 0.0787 ( 1)

0.0002
Vwet
0.0929 + 0.2957 + 0.0787 ( 1)
( 1) lO2 ,min
( 1) 0.0787
=
=
Vwet
0.0929 + 0.2957 + 0.0787 ( 1)
0.79 lair,min
0.79 0.3743
=
=
Vwet
0.0929 + 0.2957 + 0.0787 ( 1)

xSO2 ,wet =
xO2 ,wet

0.043 136
2

All the above mole (volume) fractions are in kmol/kg of wet products.

21

1 Stoichiometry
Composition of dry combustion products:
xCO2 ,dry =
xSO2 ,dry =
xO2 ,dry
xN2 ,dry

0.824 658
12

Vdry
0.006 83
32

0.0687
0.0695 + 0.2957 + 0.0787 ( 1)

0.0002
0.0695 + 0.2957 + 0.0787 ( 1)

Vdry
( 1) lO2 ,min
( 1) 0.0787
=
=
Vdry
0.0695 + 0.2957 + 0.0787 ( 1)
0.79 lair,min
0.79 0.3743
=
=
Vdry
0.0695 + 0.2957 + 0.0787 ( 1)

All the above mole (volume) fractions are in kmol/kg of dry products.
Composition of combustion products for 10 % excess air ratio ( = 1.1)
Using the above relationships one may easily obtain:
Composition of wet products for = 1.1 is:
xH2 O = 0.0552

xCO2 ,wet = 0.1613

xO2 ,wet = 0.0185

xN2 ,wet = 0.7635

xSO2 ,wet = 0.0005

Composition of dry products for = 1.1 is:


xCO2 ,dry = 0.1706

xSO2 ,dry = 0.0005

xO2 ,dry = 0.0195

xN2 ,dry = 0.8079

CO2 and O2 concentrations in dry combustion products as a function of excess


air ratio
The dependence of carbon dioxide and oxygen concentrations as a function of
excess air ratio can easily be derived from the above relationships. Fig. 1.6 and
Fig. 1.7 show the dependence.

22

1.4 Combustion stoichiometry for liquid and solid fuels

Fig. 1.6: Oxygen concentration in dry combustion products as a function of excess air
ratio

Fig. 1.7: Carbon dioxide concentration as a function of excess air ratio

23

1 Stoichiometry
Comments:
(a) For the fuels considered in this example (CH4 , C2 H6 , Groningen natural gas,
coal Fettnuss), the relationship between oxygen content in combustion products and excess air ratio (see Fig. 1.6) is almost linear for < 1.5. This
relationship holds for most of fuels and is used by furnace operators in every
day practice. For example a 2 % (dry) oxygen content in the flue gas indicates
that the furnace is operated at 1.1 excess air ratio.
(b) Carbon dioxide concentration in combustion products reaches maximum at
= 1.0 and decreases almost linearly with excess air ratio as shown in Fig. 1.7.
This maximum carbon dioxide concentration at = 1.0 is a characteristics
of the fuel only.
End of Example 1.3

1.5 Humid Combustion Air


In Section 1.3.3 we have derived Eq. (1.18) for calculating the minimum dry air
requirement for combustion of one kmol of a gaseousfuel. A similar expression,
Eq. (1.24), has been derived to calculate the minimum dry air requirement for
complete combustion of one kilogram of a solid or a liquid fuel. We stress here
again that Eqs. (1.18) and (1.24) are for calculating the dry air requirements.
However, combustion or atmospheric air is seldom completely dry and it contains moisture (water vapour). Accurate calculations of combustion stoichiometry
should account for the presence of water vapour. In this paragraph we show how
to do it. We recall that air that contains no water vapour is called dry air. Atmospheric or combustion air containing water vapour is named here as combustion
air. The temperature of air in combustion applications ranges from about 20 to
about 1300 C. In most of the furnace applications combustion air is supplied at
pressures that are slightly higher than ambient air pressure of 1 bar. However, in
gas turbines and engines the combustion air is typically compressed to 2030 bar
pressure.

1.5.1 Absolute and relative humidity


It is convenient to treat combustion air as a mixture of water vapour and dry air
since the composition of dry air remains constant but the amount of water vapour
changes. It is certainly convenient to treat the water vapour in combustion air as
an ideal gas and for ambient air pressures such an assumption is perfectly valid.

24

1.5 Humid Combustion Air


Even for pressures up to 30 bar and temperatures up to 1300 C the ideal gas
assumption is justifiable and the maximum departure from reality is typically in
the range 0.26 %. Then the combustion air is treated as an ideal-gas mixture
whose pressure is the sum of the partial pressure of the dry air (pdry air ) and that
of the water vapour (pv )
p = pdry air + pv
(1.31)
The partial pressure of water vapour (pv ) is typically referred to as the vapour
pressure. The temperature, however, is uniform throughout the dry-air/watervapour mixture so that
T = Tdry air = Tv
(1.32)
Usually the total pressure (p) is known whereas the partial pressure of water
vapour (pv ) depends on how much moisture is present in the mixture. For an
ideal-gas mixture, the mole fraction of water vapour is then
xv =

pv
p

(1.33)

The amount of water vapour in the combustion air can be specified in various
ways. However, the simplest way is to specify the mass of water vapour present
in a unit of dry air. This is called absolute or specific humidity and in this lecture
series is denoted as and is expressed in kg of water vapour per kg of dry air
since
=

mv
Mv n v
=
mdry air
Mdry air ndry

=
air

18.016
pv
pv

= 0.622
28.97 p pv
p pv

(1.34)

where Mv and Mdry air are the molar masses of water and dry air, respectively
whilst nv and ndry air stand for the number of moles of water vapour and dry air,
respectively.
Dry air contains no water vapour and therefore its specific humidity is zero. Now
let us add some water vapour to this dry air so the specific humidity () increases.
If we add more water vapour the specific humidity keeps increasing until the air
can hold no more moisture. At this point the air is said to be saturated with
moisture vapour and it is called saturated air. The amount of water vapour in
saturated air at a given temperature and total pressure can be determined using
Eq. (1.34) by replacing pv by the saturation pressure psat of water at the given
temperature. The saturation pressure of water vapour is plotted in Fig. 1.8 as a
function of temperature using a relationship4
4

In Chapter 3, Example 3.3, we will derive Eq. (1.35) using Clausius-Clapeyron equation.

25

1 Stoichiometry

psat




1
1
= 611 exp 5304.3

T
273.16

in Pa

(1.35)

Water Vapour

psat in kPa

10

0
280

290

300

310

320

Temperature in K
Fig. 1.8: Saturation pressure of water vapour as a function of temperature. The plot is
obtained using Eq. (1.35) which in the plotted range provides pressure values
that are within 2 % accuracy with the values listed in steam tables [7].

The ratio of the amount of moisture the air holds to the maximum amount of
moisture the air can hold (at the saturation state) at the same temperature is
called the relative humidity
pv
=
(1.36)
pv,sat
where pv,sat stands for the saturation water vapour pressure at the specified temperature. The relative humidity () ranges from 0 for dry air to 1 for saturated
air. The amount of moisture that combustion air can hold depends on its temperature. Thus, the relative humidity of air () changes with temperature even
when its specific humidity () remains constant. Using the relative humidity and
the saturation pressure of water vapour, the specific humidity () can then be
calculated as follows:
psat
= 0.662
(1.37)
p psat

26

1.5 Humid Combustion Air


1.5.1.1 Wet air requirement
In the previous paragraphs we have developed simple formulae for calculating the
dry air requirements (ldry air ); Eq. (1.17) for gaseous fuels while Eq. (1.23) for
liquid and solid fuels. The above considerations on the combustion air humidity
allows for inclusion of water vapour since
lwet

air

= ldry

air

+ ldry

air

= (1 + ) ldry

air

(1.38)

If there is a need to calculate the enthalpy of wet combustion air it can be easily
done since
hwet air = hdry air + hvapour
(1.39)
where h stands for specific enthalpy in J/g (or kJ/kg). The enthalpy of water
vapour at ambient air pressure in the temperature range 10 to 50 C can be
determined approximately using
hvapour (T )
= 2501.3 + 1.82 T

in kJ/kg

where T in C

(1.40)

For higher temperatures and higher pressures values from steam tables should be
used.
Example 1.4
In Example 1.3 we have calculated the air requirement and the composition of dry
and wet combustion products as a function of the excess air ratio for coal Fettnuss.
In Example 1.3 we have ignored the moisture content of the combustion air. The
objective of this example is to include the moisture and by doing so to examine
its effect on the results of the calculations. We assume here that the combustion
air is supplied at 1 bar pressure and at a 20 C temperature. Its relative humidity
is 75 %.
Assumptions: the fuel is combusted to carbon dioxide and water.
One begins with calculating the saturation pressure of water vapour at 20 C using
Eq. (1.35)



1
1
psat = 611 exp 5304.3

= 2298.14 Pa
(1.41)
297.16 273.16
The absolute humidity is then
= 0.662

0.75 2298.14
psat
kg water vapour
= 0.662 5
= 0.0116
p psat
kg dry air
10 0.75 2298.14
(1.42)

27

1 Stoichiometry
so the minimum amount of wet combustion air is
= 1.0116 10.9496 = 11.0766 kg wet air/kg of fuel
(1.43)
The moisture supplied with the combustion air stream occurs in the (wet) combustion products so (see Example 1.3)

lwet

air,min

= (1+) ldry

air,min

Vwet = 0.0929 + 0.2957 + 0.0787 ( 1) +

ldry
18

air,min

(1.44)

and
kmol wet products
kg of fuel as fired
(1.45)
Composition of wet combustion products is therefore as follows:
Vwet = 0.0929 + 0.2957 + 0.0787 ( 1) + 0.007 06

xH2 O =

0.043 136
2

0.035
18

+ 0.007 06

Vwet
0.0235 + 0.007 06
=
0.0929 + 0.2957 + 0.0787 ( 1) + 0.007 06

xCO2 ,wet =
=
xSO2 ,wet =
=
xO2 ,wet =
=
xN2 ,wet =
=

0.824 658
12

Vwet
0.0687
0.0929 + 0.2957 + 0.0787 ( 1) + 0.007 06
0.006 882 6
2

Vwet
0.0002
0.0929 + 0.2957 + 0.0787 ( 1) + 0.007 06
( 1) lO2 ,min
Vwet
( 1) 0.0787
0.0929 + 0.2957 + 0.0787 ( 1) + 0.007 06
0.79 ldry air,min
Vwet
0.79 0.3743
0.0929 + 0.2957 + 0.0787 ( 1) + 0.007 06

All the above mole (volume) fractions are in kmol/kmol of wet products.

28

1.5 Humid Combustion Air


At 10 % excess air ratio ( = 1.1) the above formulae provide:
xH2 O = 0.0721

xCO2 ,wet = 0.1584

xO2 ,wet = 0.018 14

xN2 ,wet = 0.7498

xSO2 ,wet = 0.000 46

Comments:
(a) Taking into account the combustion air moisture has resulted in a 1.2 % correction to the minimum air requirement.
(b) However, the mole fraction of water vapour in (wet) combustion products has
increased by around 31 %.
(c) Obviously neither the amount of the dry combustion products nor its composition is affected by combustion air moisture content.
End of Example 1.4

1.5.2 Dew Point Temperature of Combustion Products


Typically combustion products contain water vapour. When for example a natural gas is combusted the water vapour content in the combustion products may
be as high as 16 % for = 1.2 as shown in Example 1.2. When coal Fettnuss is
combusted the water content of around 7 % (see Example 1.4) is expected. While
designing combustion systems it is required that combustion products are cooled
down to low temperatures before they are released to the atmosphere. Thus,
by cooling down the combustion products we may expect that below a certain
temperature the water vapour begins to condensate. The dew-point temperature
(Tdp ) is defined as the temperature at which condensation begins when the combustion products (or generally moist air) is cooled at a constant pressure. In other
words, Tdp is the saturation temperature of water corresponding to the vapour
pressure, as shown in Fig. 1.9.
As a matter of fact Eq. (1.35) can be used to determine the due-point temperature
if the pressure is given. Let us assume that there is 7 vol% water vapour content in
combustion products of coal Fettnuss combustion. If the combustion products are
at 1 bar pressure, the partial pressure of vapour is 7000 N/m2 and using Eq. (1.35)
we can estimate that the dew-point temperature is around 312.4 K (39.3 C).
Thus, in order to avoid condensation it is desired to keep the combustion products
at temperatures typically 4060 C higher than the due point temperature.
Dew points vary with the amounts of O2 , CO2 , SO2 , NOx ,HCl in combustion
products. In particular sulphur oxides have a pronounced effect on the dew point
temperature as shown in Fig. 1.9 for several excess air ratios.

29

1 Stoichiometry

pv =
con
s

t.

T1

Tdp

s
Fig. 1.9: Cooling of combustion products (or moist air) at a constant temperature.
The T -s diagram shows the dew-point temperature.

The excess air curves are not equally spaced since the extra oxygen tends to
produce more SO3 which exerts a catalytic effect in raising the dew point. For
example the above estimated dew-point temperature of 39.3 C would be almost
doubled if 2 ppm of SO2 was present. This is the reason that operators of coal-fired
power station boilers maintain the flue gases at temperatures above 180200 C.

Acid dew point, C

145

20%

140

r
ss ai
exce
15%
10%

135
130

5%
120
115

1.0

2.0

3.0

4.0

5.0

Weight % sulfur in fuel oil


Fig. 1.10: Effect of sulphur and excess air on acid due point for a crude oil (adapted
from [8]).

30

1.6 Combustibles burnout for solid fuels

1.6 Combustibles burnout for solid fuels


It is not possible to burn 100 % of solid fuels. When firing liquid and solid fuels the
major contribution to the unburns comes usually from the residual oil-coke or
coal char, although the incompletely combusted gases (CO and hydrocarbons)
and soot particles can also make a contribution. In the considerations that follow,
the mass fraction of total combustibles is simply calculated as 1 minus the ash
content as shown in Fig. 1.11.
Total Combustibles = 1 ash

Fig. 1.11: Ash and combustibles in a furnace

Composition of the fuel:


C0 (Total Combustibles) + ash0 = 1
m0 mass flow rate of fuel

Composition of the unburned solids:


C1 + ash1 = 1
m1 mass flow rate of unburned solids

Mass balance of combustibles can be written as:


m0 C0 = m1 C1 + b m0 C0
where m0 C0 (in kg/s) stands for the amount of combustibles entering the furnace,
m1 C1 (in kg/s) represents the amount of combustibles in the unburned solids leaving the furnace while bm0 C0 (in kg/s) stands for the amount of the combustibles
burned whilst b is the fraction of the original combustibles burned. Assuming that
ash is an inert and does not enter into any chemical reactions, its mass balance
can be written as
m0 ash0 = m1 ash1
where the left hand side is the ash input (in kg/s) while the right hand side stands
for ash flow leaving the furnace.
From the above relationships one may obtain:
0
1 ash
C1 ash0
(1 ash1 ) ash0
ash1
b=1
=1
=
C0 ash1
(1 ash0 ) ash1
1 ash0

(1.46)

31

1 Stoichiometry
where b shows the degree of burnout and (1 ash1 ) is often called carbon in
ash (in the USA it is called loss on ignition). Fig. 1.12 shows the relationship
between the combustibles burnout (b) and carbon in ash for several ash contents
of the solid fuel.

Fig. 1.12: Burnout of combustibles

1.7 Sub-stoichiometric combustion to carbon


dioxide and water vapour
Sub-stoichiometric combustion occurs when excess air ratio () is smaller than
1. Obviously, in sub-stoichiometric combustion not all the fuel is combusted
since there is not enough oxidiserpresent and some of the fuel remains unburned.
If the fuel is combusted to carbon dioxide and water vapour, the combustion
calculations leading to the determination of the composition of the combustion
products are rather straight forward as shown in Example 1.5. However if some
species like, for example, carbon monoxide and hydrogen are to be considered
the composition of the combustion products has to be determined using chemical
equilibrium calculations underlined in Chapter 4.

32

1.7 Sub-stoichiometric combustion to carbon dioxide and water vapour


Example 1.5
Produce a curve showing the composition of combustion products obtained in
sub-stoichiometric combustion of methane in air as a function of excess air ratio.
Assumptions: methane is combusted to carbon dioxide and water vapour.
We begin with writing down the oxidation reaction for methane
CH4 + 2 O2 CO2 + 2 H2 O
The excess air ratio is then
=

nair
2
0.21

so that

nair = 9.5238

where nair is the number of moles of air per 1 mole of methane. Thus, when < 1
only moles of CH4 are combusted to CO2 and H2 O while the remaining 1
moles of CH4 appear directly in combustion products as unburned fuel.
Therefore, in the combustion products appear
(1 )

2
0.79 9.5238

moles
moles
moles
moles

of
of
of
of

(unburned) CH4
CO2
H2 O
N2

The composition of combustion products is then


1
1 + 9.5238

=
1 + 9.5238
7.5238
=
1 + 9.5238
2
=
1 + 9.5238

1
1 + 7.5238

=
1 + 7.5238
7.5238
=
1 + 7.5238

xCH4 ,wet =

xCH4 ,dry =

xCO2 ,wet

xCO2 ,dry

xN2 ,wet
xH2 O,wet

xN2 ,dry

Comments:
(a) The maximum value of CO2 volume fraction is obtained for = 1 as shown
in Fig. 1.13. Compare with Fig. 1.7.

33

1 Stoichiometry

Fig. 1.13: Composition of dry combustion products for sub-stoichiometric combustion


of methane

End of Example 1.5

1.8 Summary
Nowadays more than 90 % of the energy used by human beings is generated by
combustion of fossil fuels. Despite the continuing search for alternative energy
sources combustion will remain important for many decades to come.
In the first lecture the basic concepts of chemical reactions, atoms and molecules
have been recalled. The lecture introduces stochiometric, lean and rich combustion as well as the notions of excess air ratio and equivalence ratio. In this lecture
we have assumed that fuels are combusted to carbon dioxide and water vapour.
Students should know how to calculate the amount of air (oxidiser) needed for
complete combustion of a given fuel. Furthermore you should be able to determine composition of wet and dry combustion products for any fuel combusted to
carbon dioxide and water vapour at a given excess air.
Numerous equations on combustion stochiometry have been derived in this chapter. However, the reader should realise that there is no point in memorising them
since they can be easily recreated.

34

2 Mass and Energy Balance


Contents
2.1

2.2

2.3

2.4
2.5
2.6

General Formulation of Mass and Energy Balance


2.1.1 Mass and Energy Balance at an Instant
2.1.2 Mass and Energy Balance over a Time Interval
2.1.3 Mass and Energy Balance under Steady-State Conditions
2.1.4 Example of a Mass Balance of a Furnace
The First Law of Thermodynamics
2.2.1 System Energy
2.2.2 Energy Entering and Leaving the System
2.2.3 Energy Balance of Thermal Systems (Machines)
Energy Released in Chemical Reactions
2.3.1 Reaction Enthalpy
2.3.2 Standard Enthalpies of Formation
2.3.3 Lower Calorific Value (LCV) and Gross Calorific Value (GCV)
2.3.4 Relationships between Calorific Values, Reaction Enthalpies and Formation Enthalpies
2.3.5 Dependence of LCV on Temperature
2.3.6 Example of an Energy Balance of a Furnace
Temperature of Adiabatic Combustion
Furnace Exit Temperature
Summary

2.1 General Formulation of Mass and Energy


Balance
Conservation of mass and energy is the basis for any considerations in combustion
engineering. The formulation presented in this lecture is based on the work of

35

2 Mass and Energy Balance


[9, 10, 11]. In formulating mass and energy balances we need first of all to identify
the control volume. Control volume, Fig. 2.1, is a region of space bounded by a
control surface (boundary) through which energy and matter may be transferred.

Fig. 2.1: Control volume for mass and energy balance

Once the control volume is identified an appropriate time basis must be specified.
There are two options for formulating the balances a formulation at an instant
and a formulation over a time interval.

2.1.1 Mass and Energy Balance at an Instant


Since the mass and energy balances must be satisfied at each and every instant
of time t, one option involves formulating the law on a rate basis. That is, at
any instant there must be a balance between all mass rates expressed in kg/s
and simultaneously, there must be a balance between all energy rates expressed
in J/s (=W).
The mass balance at an instant reads, Fig. 2.1:
m
in + m
g m
out = m
stored

d(m)
dt

(2.1)

where m
in , m
g, m
out , m
stored are the rates with which the mass enters the control
volume, the rate of mass generation inside the control volume, the rate with which
matter leaves the control volume and the rate of accumulation (storage) of matter
within the volume, respectively. All these terms are expressed in kg/s. Symbol
m stands for the mass of the system expressed in kg.
Eq. (2.1) is the most general form of the mass balance at an instant. It contains the
mass generation term (m
g ) that is non-zero only if nuclear reactions are considered

36

2.1 General Formulation of Mass and Energy Balance


during which matter can be converted into energy. In any other systems with nonnuclear reactions present, the mass generation term is zero and therefore during
this combustion course we will use the following equation for the mass balance at
an instance:
dm
(2.2)
m
in m
out =
dt
The energy balance at an instant reads:
dE
E in + E g E out = E stored
dt

(2.3)

where E in , E g , E out , E stored stand for the rates with which energy enters the control volume, the rate of energy generation in the volume, the rate with which the
energy leaves the volume and the rate of accumulation (storage) within the control volume, respectively. All these terms are expressed in J/s (W). The symbol
E represents the system energy expressed in joules.

2.1.2 Mass and Energy Balance over a Time Interval


The second option is to formulate balances over a time interval, thus the mass
balance is formulated in amount of mass (kilograms) whilst the energy balance
is formulated in amount of energy (joules). The mass balance over a time interval
t = t2 t1 reads:
Zt2

m
in dt

t1

Zt2

m
out dt =

t1

Zt2

d(m)
dt = m
dt

(2.4)

t1

or
(2.5)

min mout = m

The terms on the left hand side of Eqs. (2.4) and (2.5) denote the amount of mass
entering the control volume over the time period t, and the amount of mass that
leaves the control volume over this interval, respectively. The right hand side of
Eqs. (2.4) and (2.5) represents the amount of mass stored (accumulated) within
the system in this time period.
The energy balance over a time interval reads:
Zt2

t1

E in dt +

Zt2

t1

E g dt

Zt2

t1

E out dt =

Zt2

d(E)
dt = E
dt

(2.6)

t1

37

2 Mass and Energy Balance


or
Ein + Eg Eout = E

(2.7)

where the terms on the left hand side of Eqs. (2.6) and (2.7) stand for the amount
of energy entering the control volume, the amount of energy generated in the
volume, and the amount of energy leaving the volume in the time period t =
t2 t1 . The right hand side of Eqs. (2.6) and (2.7) represents the energy stored
(accumulated) in the control volume over the time interval t = t2 t1 .

2.1.3 Mass and Energy Balance under Steady-State Conditions


Under steady-state conditions when there is neither mass nor energy accumulation
(storage) in the system we obtain:
from Eq. (2.2) for mass balance:
m
in = m
out

(2.8)

E in + E g = E out

(2.9)

from Eq. (2.3) for energy balance:

2.1.4 Example of a Mass Balance of a Furnace


The purpose of making a mass balance under steady state conditions is
to calculate the out-coming amount of mass and to calculate its make
up (composition). A correct mass balance is a prerequisite to a subsequent
energy balance.
There are two basic ways of formulating a mass balance; by expressing the incoming and outcoming flows in kmol/h (kmol/s) and their composition in mol
(volume) fractions or by expressing the streams in kg/h (kg/s) and their composition in mass fractions. The first way is popular among chemical engineers since
it deals with kmol and the rates of chemical reactions are typically expressed using
concentrations (kmol/m3 /s). The latter way is common in combustion engineering. Both methods have advantages and disadvantages and they are equivalent.
They both lead to the same results. The combustion stoichiometry calculations
(see Chapter 1) are required to calculate composition of combustion products
leaving the system.

38

2.1 General Formulation of Mass and Energy Balance


Example 2.1
A boiler fired with methane is used to generate hot water. The burner of the
boiler is fed with 200 kg/h of methane and 4000 kg/h of dry air and it is operated
under steady state conditions. Make the mass balance for the boiler assuming
that the combustion is complete. Carry out the calculations using the molar flow
rates (kmol/h). Repeat the calculations using mass flow rates (kg/h).

Combustion products

200 kg/h CH4

Boiler

4000 kg/h air


control volume boundary
Fig. 2.2: Mass balance for the boiler

Assumptions:
(a) the methane combustion proceeds to completion resulting in CO2 and H2 O,
(b) we place the control volume over the boiler (see Fig. 2.2),
(c) the boiler operates under steady state and therefore we formulate the mass
balance at an instance with the mass accumulation term equal to zero.
Mass balance in kmol/h
The starting point to any mass balance is an estimation of the incoming (input) streams and writing down proper chemical reactions. Thus, the incoming streams are: m
CH4 = 12.5 kmolCH4 /h and since the molar mass of air is
Mair = 0.21 32 + 0.79 28 = 28.84 kg/kmol we obtain for the incoming stream of
4000kg/h
air m
air = 28.84kg/kmol
= 138.696(kmol air )/h. The oxidation reaction is:
CH4 + 2 O2 CO2 + 2 H2 O
and therefore the minimum air requirement is
excess air ratio is
=

2
0.21

kmol
= 9.5238 kmol

138.696 kmol of air/h


12.5 kmol of CH4 /h

9.5238 kmol of air/kmol of CH4

of air
of fuel .

The

= 1.165

Since > 1 the combustion products contain CO2 , H2 O, N2 and O2 .


The amount of the wet combustion products is 12.5 (3 + 0.79 9.5238 + (1.165
1) 9.5238) = 151.20kmol/h while for the dry combustion products the figure of

39

2 Mass and Energy Balance


12.5 (1 + 0.79 9.5238 + (1.165 1) 9.5238) = 126.20kmol/h is applicable.
The composition of the combustion products is as follows:
xCO2 ,wet =
xH2 O =
xN2 ,wet =
xO2 ,wet =

12.5
151.20
2 12.5
151.20
0.79 9.5238 + 0.79 (1.165 1) 9.5238
109.5654
12.5
=
151.20
151.20
(1.165 1) 9.5238 0.21
4.1250
12.5
=
151.20
151.20

= 0.0827
= 0.1653
= 0.7246
= 0.0273

and the molar mass of the wet combustion products is:


kg
Mwet,products = 0.0827 44 + 0.1653 18 + 0.7246 28 + 0.0273 32 = 27.7766 kmol

The composition of the dry combustion products is as follows:


12.5
= 0.099
126.20
109.5654
=
= 0.8682
126.20
4.1250
= 0.0327
=
126.20

xCO2 ,dry =
xN2 ,dry
xO2 ,dry

and the molar mass of the dry products is:


Mdry,products = 0.099 44 + 0.8682 28 + 0.0327 32 = 29.712kg/kmol
Knowing the molar fractions of the species, one may easily calculate the mass
fractions since:
0.0827 44
= 0.1310
27.7766
0.1653 18
= 0.1071
w H2 O =
27.7766
0.7246 28
wN2 ,wet =
= 0.7304
27.7766

wCO2 ,wet =

40

wCO2,dry =

wN 2,dry =

0.099 44
= 0.1466
29.712

0.8682 28
= 0.8182
29.712

2.1 General Formulation of Mass and Energy Balance


and
wO2 ,wet =

0.0273 32
= 0.0315
27.7766

wO2,dry =

0.0327 32
= 0.0352
29.712

The tables below summarise the mass balance:


Table 2.1: The incoming and out-coming streams in kmol/h

IN
Species

kmol/h

Methane
Nitrogen
Oxygen

12.50
109.57
29.13

151.20

OUT
WET
kmol/h
Carbon dioxide
12.50
Water vapour
25.00
Nitrogen
109.57
Oxygen
4.13
P
151.20
Species

DRY
kmol/h
12.50

109.57
4.13
126.20

Table 2.2: Composition of the combustion products

Species
Carbon dioxide
Water vapour
Nitrogen
Oxygen
P

Molar fraction
Wet
Dry
0.0827 0.0990
0.1653

0.7246 0.8682
0.0273 0.0327
0.9999 0.9999

Mass fraction
Wet
Dry
0.1310 0.1466
0.1071

0.7304 0.8182
0.0315 0.0352
1.0000 1.0000

Mass balance in kg/h

The incoming streams are:


m
CH4 = 200 kg/h and m
air = 4000 kg/h.
The outcoming flow rate (in kg/h) is easily obtained since:
m
products = 200 kg/h + 4000 kg/h = 4200 kg/h.
While making the mass balance in kilograms it is convenient to write the oxidation
reaction
CH4 + 2 O2 CO2 + 2 H2 O

41

2 Mass and Energy Balance


using kilogram rather than kmol:
16 kg CH4 + 64 kg O2 = 44 kg CO2 + 36 kg H2 O
The minimum air requirement is 64/16/0.233 = 17.1674 kg of air/kg of methane
and the excess air ratio is:
=

4000 kg of air/h
200 kg of CH4 /h

17.1674 kg of air/kg of CH4

= 1.165

Since > 1 the combustion products will contain CO2 , H2 O, N2 and O2 .


The combustion products stream contains:
Methane
Carbon dioxide
Water vapour
Nitrogen
Oxygen
TOTAL (wet)

none
550 kg/h
450 kg/h
3068 kg/h
132 kg/h
4200 kg/h

200 44
16 =
200 36
16 =
0.767 4000 =
0.233 (4000 200 17.1674) =

and the amount of combustion products dry is then 3750 kg/h.


The species mass fractions can be easily calculated:
550
= 0.131
4200
450
w H2 O =
= 0.1071
4200
3068
= 0.7305
wN2 ,wet =
4200
132
wO2 ,wet =
= 0.0314
4200
wCO2 ,wet =

wCO2 ,dry = 0.1467

wN2 ,dry = 0.8181


wO2 ,dry = 0.0352

The conversion of the mass fractions into the molar fractions (if required) is also
rather straightforward:
1
Mproducts,wet

0.131 0.1071 0.7305 0.0314


+
+
+
= 0.0360
44
18
28
32
and Mproducts,wet = 27.7795 kg/kmol

42

2.2 The First Law of Thermodynamics


1
0.1467 0.8181 0.0352
+
+
= 0.033 65
=
Mproducts,dry
44
28
32
and Mproducts,dry = 29.716 kg/kmol
and
xCO2 ,wet = 0.131/44/0.0360 = 0.0827 xCO2 ,dry = 0.1467/44/0.033 65 = 0.0991
xH2 O = 0.1071/18/0.0360 = 0.1653
xN2 ,wet = 0.7305/28/0.0360 = 0.7247 xN2 ,dry = 0.8181/28/0.033 65 = 0.8683
xO2 ,wet = 0.0314/32/0.0360 = 0.0273 xO2 ,dry = 0.0352/32/0.033 65 = 0.0327
The table below summarises the mass balance:
Table 2.3: The incoming and outcoming flow rates in kg/h

IN
Species
kg/h

Comments:

Methane
Nitrogen
Oxygen

200
3068
932

IN

4200

OUT
WET
kg/h
Carbon dioxide
550
Water vapour
450
Nitrogen
3068
Oxygen
132
P
OUT
4200
Species

DRY
kg/h
550

3068
132
3750

(a) The incoming and outcoming molar flow rates are equal as shown in Table 2.1.
In the first lecture it has been stressed that in general one cannot balance the
molar flows. The question is what makes the considered example so specific
that the incoming and outcoming molar flows are equal?
(b) Observe that both methods provide the same composition of the wet and dry
combustion products, as one would expect.
End of Example 2.1

2.2 The First Law of Thermodynamics


The first law of thermodynamics is the overall energy balance that is
extended into all possible forms of energy. Historically the first law of thermodynamics was formulated after James Prescott Joule observed in 1840 in his

43

2 Mass and Energy Balance


experiments that heat and mechanical work were equivalent forms of energy. Before Joules experiments, heat was associated with some "fluidic motions" induced
by a temperature difference.
In thermodynamics we consider exchanges of all possible energy forms between
the control volume and the surroundings. The space that is encompassed by
the control volume, in which we have a special interest, is called the system.
The type of system depends on the characteristics of the boundary (surface) of
the control volume, Fig. 2.3. If matter (mass) can be transferred through the
boundary the system is classified as open. If matter cannot pass through the
boundary the system is classified as closed. Both open and closed systems can
exchange energy with the surroundings. An isolated system can exchange neither
matter nor energy with its surroundings.

Fig. 2.3: Left an open system can exchange matter and energy with its surroundings, Middle a closed system can exchange energy with its surroundings,
Right an isolated system can exchange neither energy nor matter with its
surroundings

For an open system that exchanges mass and energy within its surroundings and
with no energy generation within the system, the overall energy balance (Eq. (2.7))
over a time period reads:
Ein = Eout + E
(2.10)
and at an instant

d(E)
E in E out =
dt

(2.11)

E in dt E out dt = dE

(2.12)

dEin = dEout + dE

(2.13)

Eq. (2.11) can be rearranged:

to obtain

44

2.2 The First Law of Thermodynamics


which is a differential form of the first law of thermodynamics. Fig. 2.4 shows a
Sankeys diagram for energy balance expressed by Eqs. (2.10) and (2.13) applicable to both open and closed systems.

Control volume
Ein

DE
Eout

Fig. 2.4: Sankeys diagram for energy balance for an open or closed system

It is instructive to consider the following specific cases of Eqs. (2.10) and (2.13):
(a) For an isolated system we have, Ein = 0 and Eout = 0, and therefore E = 0,
so the energy of an isolated system remains constant. In some (old) thermodynamics text books the reader will find the following formulation of the first
law of thermodynamics:
"The sum of all energies is constant in an isolated system."
(b) If energy of a system remains constant (E = 0) then the energy entering
the system must equal the energy leaving the system, Ein = Eout .
(c) For a system operating in a steady-state (a furnace, a boiler, a turbine) there
is no energy accumulation (dE/dt = 0) and therefore from Eq. (2.11) we
obtain
E in = E out
(2.14)
The above relationship shows that it is not possible to construct a machine operated continuously that would provide mechanical work E out = L 1 without receiving energy (as mechanical energy, electrical, chemical, nuclear, thermal or in any
other form). A device that would operate without a supply of energy is called a
perpetuum mobile of the first kind. Therefore, a popular formulation of the first
law of thermodynamics is:
"It is not possible to construct a perpetuum mobile of the first kind."
1

For the definition of the work L see Section 2.2.2.2.

45

2 Mass and Energy Balance

2.2.1 System Energy


Eqs. (2.10) and (2.13) express the first law of thermodynamics formulated over a
time interval. In order to use the law we need to learn how to calculate the three
terms Ein (dEin ), Eout (dEout ), E (dE) appearing in the equations. We begin
with the third term that is with the system energy E (dE).
If a system is in motion its energy consists of: the internal energy of the system,
the kinetic energy of the motion, and the potential energy:
E = U + m

w2
2
+I
+ mgz
2
2

(2.15)

where U stands for the internal energy, m is the mass of the system, I is the
moment of inertia, w and are the translational and rotational velocities, respectively, g is the gravity while z stands for the height of the center of gravity of the
system. The latter is measured from a reference level (height), see Fig. 2.5. Internal energy U is a sum of kinetic energy, potential energy and electronic energy
of all the molecules and atoms of the system. It includes also energy of chemical
bonds of the molecules. Internal energy is a state variable and therefore its change
from an initial state 1 to a state 2 is:
E = U2 U1 +

I
m
(w22 w12 ) + (22 12 ) + m g (z2 z1 )
2
2

Fig. 2.5: System in translational and rotational motion

46

(2.16)

2.2 The First Law of Thermodynamics


In combustion engineering the kinetic and potential energies of the system can
be neglected since they are much smaller than internal energy. Typically, kinetic
energies are considered when the velocity of the system is larger than 50 m/s
whereas potential energies are taken into account if z2 z1 is larger than 100 m.
Thus, neglecting the kinetic and potential energies we obtain:
(2.17)

E = U2 U1

Internal energy is a function of any two of the three state variables (temperature,
pressure, volume) and is an extensive property (see textbooks on thermodynamics,
for example [3, 5]):
U = m u(T, p) = m u(T, v) = m u(p, v)

(2.18)

and its exact differential is:


dU = d(m u) = m du + u dm

(2.19)

In the above equations U (in joules) stands for internal energy of the whole system
whilst u is specific internal energy expressed in J/kg.
Since specific internal energy is a state variable and its differential is exact, so:


u
u
dT +
dv
(2.20)
du =
T v
v T
or [3, 5]



p
du = cv dT + T
p dv
T v


(2.21)

where v is specific volume (in m3 /kg) and cv (in J/(kg K)) is the specific heat at
constant volume.
It is not possible to calculate an absolute value of specific internal energy. We
can calculate its difference between two states by integrating Eq. (2.21):

u(T, v) = u0 +

ZT,v

du

(2.22)

T0 ,v0

where u0 remains unknown. We can prescribe any value to u0 provided that the
reference state T0 , v0 is properly selected. While balancing physical phenomena
without any phase changes we are completely free in selecting the reference state

47

2 Mass and Energy Balance


T0 , v0 . When phase changes occur, the reference state T0 , v0 should be selected so
as to allow for going through all the phase changes considered while evaluating
the integral in Eq. (2.22). When chemical reactions take place the reference state
T0 , v0 should correspond to the state for which reference substances are specified.
The specific internal energy is interrelated with the specific enthalpy (h) since
h = u + pv

(2.23)

In thermodynamics a reference value for specific enthalpy (h0 ) is usually specified


at a reference state T0 , p0 and therefore, the reference value for specific internal
energy u0 must be calculated using
u0 = h0 p0 v0

(2.24)

where the specific volume v0 is calculated using an equation of state:


v0 = v(T0 , p0 )

(2.25)

Specific Internal Energy of Ideal Gases


For an ideal gas Clausius-Clapeyron equation of state (Eq. (1.8)) is applicable:
v=

RT
M p

(2.26)

where R = 8, 314.3 J/(kmol K) is the universal gas constant and M stands for
the molecular mass of the gas in question. The reference specific enthalpy (h0 )
is usually prescribed zero value at T0 = 298.15 K and p0 = 1 bar. Following
Eq. (2.24) the reference specific internal energy is
u 0 = 0 p0

R T0
R
= T0
M p0
M

For an ideal gas, from Eq. (2.26) we obtain



R
p
=

T v
vM

(2.27)

(2.28)

and after inserting into Eq. (2.21), a simple expression for calculating specific
internal energy emerges:

48

2.2 The First Law of Thermodynamics

u(T ) = u0 +

ZT

R
cv (T ) dT = T0 +
M

T0

ZT

cv (T ) dT

(2.29)

T0

The above expression underlines the fact that internal energy of an ideal gas is a
function of temperature only. For real gases however, internal energy is a function
of temperature and pressure (or volume) [3, 5].

2.2.2 Energy Entering and Leaving the System


In this paragraph we consider the other two terms Ein (dEin ), Eout (dEout ) that
appear in the first law of thermodynamics, Eqs. (2.10) and (2.13). There are
several means of supplying and removing energy from the system. These are:
(a) energy supplied (removed) with a stream of fluid,
(b) energy supplied (removed) as mechanical work,
(c) energy supplied (removed) as heat.
In subsequent paragraphs we deal with all these means of energy supply.
2.2.2.1 Enthalpy as the Energy of a Stream of Fluid
In combustion engineering matter and energy are provided into or removed from
a system (a furnace, a turbine, a boiler, a chemical reactor) using pipe lines.
Streams of fluids carry energy into and from the system. The energy of a fluid
stream consists of the internal energy and the product of static pressure and
specific volume (see Eq. (2.23) for specific enthalpy). This overall energy of the
stream is called enthalpy. Enthalpy is an extensive property
H = m h(T, p) = m h(T, v) = m h(p, v)

(2.30)

and its exact differential is:


dH = m dh + h dm

(2.31)

In the above expression H (in joules) stands for enthalpy of the system whilst h
(in J/kg) is specific enthalpy. Specific enthalpy is a state variable and its exact
differential reads:


h
h
dh =
dT +
dp
(2.32)
T p
p T
49

2 Mass and Energy Balance


or [3, 5]

h
dT +
dh =
T p

!
v
vT
dp
T p

(2.33)

Specific Enthalpy of Ideal Gases


v
|p using the Clausius-Clapeyron equation (Eq. (2.26))
Calculating the derivative T
and after inserting it into Eq. (2.33) we obtain:

h(T ) = h0 +

ZT

cp (T ) dT

(2.34)

T0

After recalling that for the reference state (T0 = 298.15 K, p0 = 1 bar) we have
already chosen h0 = 0 one obtains the final relationship for calculating specific
enthalpy of an ideal gas:
ZT
h(T ) = cp (T ) dT
(2.35)
T0

where cp (T ) in J/(kg K) stands for the specific heat at constant pressure. The
above relationship underlines the fact that specific enthalpy of an ideal gas is a
function of temperature only. For a real gas, for which the equation of state is
different to Clapeyron equation, specific enthalpy is a function of temperature
and pressure (or specific volume) [3, 5].
2.2.2.2 Energy Supplied to the System through Mechanical Work
Work is the fundamental property in thermodynamics. Work is done when an
object is moved against an opposing force. An example of doing work is
an expansion of a gas that pushes out a piston and raises weight. When work is
done to a system, for example by moving a piston and the gas is compressed, the
capacity of the system to do work is increased. Conversely when the system does
work, its energy is reduced because the system can do less work. It is important
to realise that work is not energy. Work is a means of transmitting energy.
Notion of work makes sense only when the work is actually being done. When we
push a piston to compress the gas we are doing work. However, when the gas is
already compressed the work does not exist any more whereas the energy of the
system is increased and it exists.

50

2.2 The First Law of Thermodynamics


It is important to agree upon the sign of work. In this lecture notes work done
to the system is positive whilst work obtained from the system is negative, as
shown in Fig. 2.6. According to this convention work done to compress a gas is
positive. When afterwards the gas expands to a lower pressure, work is obtained
from the system and this work is negative.

Fig. 2.6: Work done to the system is positive; Work done by the system (obtained
from the system) is negative

In thermodynamics [3, 5] you have learned that the infinitesimal work done to
the system is:
dL = p dV
(2.36)
and the amount of work (done to the system) which changes parameters of the
system from a state 1 to a state 2 is:
L=

Z2

p dV

(2.37)

In the above equation L stands for work done to the system expressed in joules,
p is the static pressure (in N/m2 ) while V is the system volume (in m3 ). The
corresponding expressions for the specific work done to the system are:
dl = p dv

(2.38)

Z2

(2.39)

and
l=

p dv

where the specific work is in J/kg and v is the specific volume in m3 /kg.
It is easy to see that work is not a state variable. Consider its differential,

51

2 Mass and Energy Balance


Eq. (2.38), as dl = p dv 0 dp. Thus,
2l

l
v |p

= p and

2l

l
p |v

= 0. Calculating
2

l
l
6= p v
and
the second derivatives v p = 1 and p v = 0 we see that v p
therefore the differential of work is not exact (see Example 2.2). Consequently
the integrals for evaluating the amount of work L (Eq. (2.37)), and the specific
work l (Eq. (2.39)) depend on the integration path. To perform these integrations
the dependence of pressure as a function of volume (or specific volume) must be
known. In other words to evaluate these integrals the thermodynamic process
must be known.

Fig. 2.7: Work done to the system

Example 2.2
In thermodynamics we use frequently functions of two variables. An equation of
state, written in a general form as p = p(T, v) is an example of such a function.
Other frequently used functions are to calculate internal energy u = u(T, p) ,
enthalpy h = h(T, p) and entropy s = s(T, p) . These functions are called state
functions and the variables appearing in them are called state variables. Work
and heat belong to another group of functions that are not state functions. This
example is to recall the criteria for determining whether a function is a state
function or not. A mathematician expresses it by saying: the differential of a
state function is exact. Below you will find what it means.
Consider the purely mathematical problem where F (x, y) is a function of two
independent variables x and y. If one goes to a neighbouring point x + dx and
y + dy, the function F changes by an amount:
dF = F (x + dy, y + dy) F (x, y)

(A1)

dF = A(x, y) dx + B(x, y) dy

(A2)

or

52

2.2 The First Law of Thermodynamics


where
A(x, y) =

F (x, y)
F (x, y)
and B(x, y) =
x
y

(A3)

In Eq. (A1) the infinitesimal quantity dF is here an ordinary differential; it is


also called an exact differential. If one goes from initial point "i" corresponding
to (xi , yi ) to a final point "f" corresponding to (xf , yf ) the corresponding change
in F is:

F =

Zf

dF = Ff Fi

(A4)

The change F can be simply evaluated as F (xf , yf ) F (xi , yi ) and it is dependent only on the value of the function F at (xf , yf ) and (xi , yi ). The change F
may be also written as:

F =

Zf

[A(x, y) dx + B(x, y) dy]

(A5)

Since F does depend only on F (xf , yf ) and F (xi , yi ), the integral (A5) does not
depend on the path along it is evaluated, going from the initial point "i" to the
final point "f".
It can be demonstrated that the necessary and sufficient conditions for the differential (A2) to be an exact differential are:

since

A(x, y)
B(x, y)
=
y
y

(A6)

2 F (x, y)
2 F (x, y)
=
y x
x y

(A7)

From the above considerations it follows:


(a) if the infinitesimal quantity dF is an exact differential, the function F is a
state variable,
(b) if the function F is a state variable, its ordinary differential is exact.
End of Example 2.2

53

2 Mass and Energy Balance


2.2.2.3 Energy Supplied to the System as Heat
Heat is energy in transit due to a temperature difference. The convention
is that heat supplied to the system is positive while heat removed from the system
is negative. Heat is not a state variable and its differential is inexact. Heat
differs from work since work can be fully converted into heat whereas only part
of available heat can be converted into work [3, 5] following the second law of
thermodynamics.
Total heat absorbed by the system consists of heat supplied from the surroundings
and heat generated within the system due to friction, so:
Qt = Q + Qf

(2.40)

where Qt is the total amount of heat absorbed by the system, Q is the amount
of heat supplied from the surroundings, and Qf is the amount of heat generated
within the system due to friction. All these terms are expressed in joules. The
above relationship can be also expressed in the amount of heat per 1 kg of the
matter of the system:
qt = q + qf
(2.41)
Specific Heat Capacities
Temperature of a system changes when heat is transferred into or from the system.
The specific heat capacity describes the temperature change of 1 kg of a system
that results from the heat transfer dqt ;
c=

dqt
dT

(2.42)

where the specific heat (c) is expressed in J/(kg K). Equally well one can define
specific heat per 1 kmol of a system, and this property is going to be denoted in
this lecture series using capital C. Thus,
c denotes specific heat in J/(kg K)
C denotes specific heat in J/(kmol K)
The heat capacity depends on the conditions during the heat addition or removal
(see textbooks on Thermodynamics, for example [3, 5]) and:

54

for v = const process

C = Cv

for p = const process

C = Cp

for T = const process

C=

for dqt = 0 (adiabatic) process

C=0

2.2 The First Law of Thermodynamics


When the system is at a constant pressure, heat addition increases simultaneously
the temperature and the system energy p V through expansion of the system
boundary. Thus, the heat capacity at constant pressure (cp ) is larger than the
heat capacity at constant volume (cv ). For ideal gases the following relationship
applies [3, 5]:
Cp Cv = R
(2.43)
where R is the universal gas constant R = 8, 314.3 J/(kmol K). Obviously the
following is applicable:
R
(2.44)
cp cv =
M
where M is the molecular mass of the matter of the system.
Specific heats of ideal gases can be derived using kinetic theory of gases [3, 5].
The general relationship is as follows:
Cv =

number of degrees of freedom


R
2

(2.45)

For a monatomic gas (H,O,N) there are only three degrees of freedom associated
with translational motion. Even at high temperatures a monatomic gas can move
in three directions only, without experiencing any rotational components. Thus,
specific heats of monatomic gases are Cv = 3 R/2 = 12.47 kJ/(kmol K) and
Cp = 5 R/2 = 20.79 kJ/(kmol K).
For a diatomic gas (H2 ,O2 ,N2 ) there are more degrees of freedom. At low temperatures a diatomic gas experiences only translational motion. However, when the
temperature increases the diatomic system starts rotating around two symmetry
axes and Cv approaches 5 R/2 = 20.78 kJ/(kmol K) (Cp approaches 7 R/2 =
29.09 kJ/(kmol K)). At even higher temperatures the diatomic system begins to
vibrate and Cv approaches 7 R/2 = 29.09 kJ/(kmol K) (Cp approaches 9 R/2 =
37.41 kJ/(kmol K)).
Values of specific heat capacities are calculated using statistical thermodynamics
however for stable molecules the heat capacities are measured. NIST - JANAF2 tables [12] provide the most comprehensive set of physical enthalpy data from which
Cp values can be readlily calculated. The Cp values are expressed as polynomials
of the fourth order in T ,
Cp
= Cp,1 +Cp,2 T +Cp,3 T 2 +Cp,4 T 3 +Cp,5 T 4
R
2

(Cp in kJ/(kmol K)) (2.46)

NIST - National Institute of Standards and Technology (USA)


JANAF - Joint Army Navy Air Force (USA)

55

2 Mass and Energy Balance


In order to improve accuracy, usually two different polynomials are used for low
(T < 1000 K) and high (T > 1000 K) temperatures. Knowing the coefficients of
the polynomials we calculate the specific enthalpy as follows:
h(T ) = h(298.15K)+

ZT

Cp (T ) dT = Cp,6 R+

298.15K

ZT

Cp (T ) dT (in

kJ
)
kmol

298.15K

(2.47)
A new symbol h is introduced to denote specific enthalpies expressed per one kmol
as opposed to h that stands for the specific enthalpy per one kilogram (in JANAF
0
Tables H T is used for h). In order to compute enthalpies an additional constant
h(298.15K) = Cp,6 R (in kJ/kmol) is needed to make sure that at the reference
state conditions (T = 298.15 K, p = 1 bar) the enthalpy is zero.
In numerous text books mean specific heats are given. These are quoted for
the temperature range from zero to T
T
1
C 0 =
T

ZT

C(T ) dT

(2.48)

or for specified temperatures T1 and T2 , so


T
C T21 =

T 2 T1

ZT2

C(T ) dT

(2.49)

T1

The relationship between Eq. (2.48) and Eq. (2.49) is as follows:


T
T
T2
C 0 2 T2 C 0 1 T1
C T1 =
T2 T1

(2.50)

2.2.3 Energy Balance of Thermal Systems (Machines)


In the above paragraphs we have learned how to calculate the individual terms in
the general energy balance over a time period, Eq. (2.10). We have also learned
how to calculate the differentials in the differential form of the energy balance,
Eq. (2.13). If we rearrange Eq. (2.13) so that:
dEin dEout = dE

56

(2.51)

2.2 The First Law of Thermodynamics


Table 2.4: The coefficients for polynomials (2.46) for T in the range 3001000 K

H
O
N
H2
O2
N2
CO
H2 O
CO2
CH4
H
O
N
H2
O2
N2
CO
H2 O
CO2
CH4

Cp,1
2.500 000 00 10+00
3.168 267 10 10+00
0.250 000 00 10+01
2.344 331 12 10+00
3.782 456 36 10+00
0.032 986 77 10+02
3.579 533 47 10+00
4.198 640 56 10+00
2.356 773 52 10+00
5.149 876 13 10+00
Cp,5
9.277 323 32 1022
2.112 659 71 1012
0.000 000 00 10+00
7.376 117 61 1012
3.243 728 37 1012
0.024 448 54 1010
9.044 244 99 1013
1.771 978 17 1012
1.436 995 48 1013
1.666 939 56 1011

Cp,2
7.053 328 19 1013
3.279 318 84 1003
0.000 000 00 10+00
7.980 520 75 1003
2.996 734 16 1003
0.140 824 04 1002
6.103 536 80 1004
2.036 434 10 1003
8.984 596 77 1003
1.367 097 88 1002
Cp,6
2.547 365 99 10+04
2.912 225 92 10+04
0.561 046 37 10+05
9.179 351 73 10+02
1.063 943 56 10+03
0.102 089 99 10+04
1.434 408 60 10+04
3.029 372 67 10+04
4.837 196 97 10+04
1.024 664 76 10+04

Cp,3
1.995 919 64 1015
6.643 063 96 1006
0.000 000 00 10+00
1.947 815 10 1005
9.847 302 01 1006
0.039 632 22 1004
1.016 814 33 1006
6.520 402 11 1006
7.123 562 69 1006
4.918 005 99 1005
Cp,7
4.466 828 53 1001
2.051 933 46 10+00
0.419 390 87 10+01
6.830 102 38 1001
3.657 675 73 10+00
0.039 503 72 10+02
3.508 409 28 10+00
8.490 322 08 1001
9.901 052 22 10+00
4.641 303 76 10+00

Cp,4
2.300 816 32 1018
6.128 066 24 1009
0.000 000 00 10+00
2.015 720 94 1008
9.681 295 09 1009
0.056 415 15 1007
9.070 058 84 1010
5.487 970 62 1009
2.459 190 22 1009
4.847 430 26 1008

Fig. 2.8: Specific heats at constant pressure for various molecules (values obtained using JANAF tables). Black diamonds show values predicted by kinetic theory
of ideal gases.

57

2 Mass and Energy Balance


Table 2.5: The coefficients for polynomials (2.46) for T in the range 10005000 K

H
O
N
H2
O2
N2
CO
H2 O
CO2
CH4
H
O
N
H2
O2
N2
CO
H2 O
CO2
CH4

Cp,1
2.500 000 01 10+00
2.569 420 78 10+00
0.241 594 29 10+01
3.337 279 20 10+00
3.282 537 84 10+00
0.029 266 40 10+02
2.715 185 61 10+00
3.033 992 49 10+00
3.857 460 29 10+00
7.485 149 50 1002
Cp,5
4.981 973 57 1022
1.228 336 91 1015
0.203 609 82 1014
2.002 553 76 1014
2.167 177 94 1014
0.067 533 51 1013
2.036 477 16 1014
1.682 009 92 1014
4.720 841 64 1014
1.018 152 30 1013

Cp,2
2.308 429 73 1011
8.597 411 37 1005
0.174 890 65 1003
4.940 247 31 1005
1.483 087 54 1003
0.148 797 68 1002
2.062 527 43 1003
2.176 918 04 1003
4.414 370 26 1003
1.339 094 67 1002
Cp,6
2.547 365 99 10+04
2.921 757 91 10+04
0.561 337 73 10+05
9.501 589 22 10+02
1.088 457 72 10+03
0.092 279 77 10+04
1.415 187 24 10+04
3.000 429 71 10+04
4.875 916 60 10+04
9.468 344 59 10+03

Cp,3
1.615 619 48 1014
4.194 845 89 1008
0.119 023 69 1006
4.994 567 78 1007
7.579 666 69 1007
0.056 847 60 1005
9.988 257 71 1007
1.640 725 18 1007
2.214 814 04 1006
5.732 858 09 1006
Cp,7
4.466 829 14 1001
4.784 338 64 10+00
0.464 960 96 10+01
3.205 023 31 10+00
5.453 231 29 10+00
0.059 805 28 10+02
7.818 687 72 10+00
4.966 770 10 10+00
2.271 638 06 10+00
1.843 731 80 10+01

Cp,4
4.735 152 35 1018
1.001 777 99 1011
0.302 262 45 1010
1.795 663 94 1010
2.094 705 55 1010
0.100 970 38 1009
2.300 530 08 1010
9.704 198 70 1011
5.234 901 88 1010
1.222 925 35 1009

Fig. 2.9: Physical enthalpies of various gases as a function of temperature

58

2.2 The First Law of Thermodynamics


we realise that the net energy supply (dEin dEout ) into the system increases
the energy of the system. For a closed system, heat and work are the only means
of supplying energy into the system. If we neglect kinetic and potential energies,
internal energy is the only system energy, so
dQt + dL = dU

(2.52)

dQt p dV = dU

(2.53)

or
Introducing the specific properties into Eq. (2.53) we obtain
m dqt p d(m v) = d(m u)

(2.54)

and after some algebra (note that for a closed system dm = 0):
m dqt m p dv = m du

(2.55)

dqt p dv = du

(2.56)

and finally
The reader is requested to compare Eqs. (2.53) and (2.56). Both equations are of
the same form showing that for a closed system the change in the internal energy
equals the sum of the supplied heat and the work done to the system.
Now consider an open system that can exchange matter with the surroundings.
The change of the internal energy of the system is induced through the heat
supplied, the work done to the system and the enthalpy of the stream entering
the system, so
dQt p dV + h dm = dU
(2.57)
and further
m dqt p d(m v) + (u + p v) dm = d(m u)

(2.58)

After some algebra we obtain


dqt p dv = du

(2.59)

The above relationship valid for an open system is identical to Eq. (2.56) that
is valid for a closed system. Thus, for any system (opened or closed) the first
law of thermodynamics states that the increase in the specific internal
energy of the system equals the sum of the heat introduced to the
system and the specific work done to the system (p dv).

59

2 Mass and Energy Balance


We can introduce into Eq. (2.59) the definition of the specific enthalpy h = u+p v
and by doing so we obtain:
dqt p dv = dh v dp p dv

(2.60)

dqt + v dp = dh

(2.61)

and
The above equation leads to another formulation of the first law of thermodynamics. For any system (opened or closed) the increase in the system specific
enthalpy equals the amount of heat provided into the system per unit
of mass plus the product (v dp).
There are two important implications of Eqs. (2.59) and (2.61). In engineering
practices furnaces, boilers and chemical reactors are often operated at a constant
pressure, so
for p = const.

dh = dqt

(2.62)

For autoclaves and chemical reactors operating at constant volumes (batch processes)
for v = const.

du = dqt

(2.63)

2.3 Energy Released in Chemical Reactions


2.3.1 Reaction Enthalpy
A chemical reaction can be written as follows:
1 A1 + 2 A2 + . . . + n An = 0
or shorter as

n
X

i Ai = 0

i=1

where Ai stands for the chemical symbols and i are the stoichiometric coefficients;
i < 0 for reactants (substrates) while i > 0 for products.
For example the reaction
CH4 + 2 O2 CO2 + 2 H2 O

60

2.3 Energy Released in Chemical Reactions


can be re-written as:
CH4 2 O2 + CO2 + 2 H2 O = 0
and then:
A1 = CH4 ,

A2 = O2 ,

A3 = CO2 ,

A4 = H2 O;

1 = 1,

2 = 2,

3 = 1,

4 = 2.

The change of the enthalpy in a chemical reaction is called reaction enthalpy


and it is calculated as the sum of the enthalpies of the compounds participating
in the reaction times the corresponding stoichiometric coefficients:
X
pressure
(2.64)
R Htemp
=
i hi
i

where the overbar indicates that the enthalpy hi 3 is expressed per 1 kmol (or
1 mol) and the reaction enthalpy R H is in joules (or kJ). When the reaction
enthalpy is given for a particular reaction, it applies for the stoichiometric coefficients as written. If each of the stoichiometric coefficients is doubled, the reaction
enthalpy is doubled. For example, the reaction of ammonia synthesis may be
written as
1
2 N2

+ 23 H2 NH3

0
R H298
= 46, 110J

or
N2 + 3H2 2NH3

0
R H298
= 92, 220J

In order to calculate the enthalpies hi (in kJ/kmol), there is a need to define the
reference state of zero enthalpy. The pure elements in their most stable state at
T = 298.15 K and p = 1 bar (the standard state) are prescribed zero enthalpy
(by convention) as shown in Table 2.6. Gases like O2 , H2 , N2 are prescribed
zero enthalpies and so are carbon as graphite and other pure elements. There
is one exception to this general definition of reference state. For phosphorous
the reference state is taken to be white phosphorous despite this allotrope not
being the most stable form but simple the most reproducible form of the element.
3

In textbooks on chemical thermodynamics capital letter H i is typically used to denote specific


(molar) enthalpy. In this lecture we use small letter h so that hi and hi stand for the specific
enthalpies of species i expressed per kg and kmol, respectively. Reaction enthalpy R H is
then expressed in kJ (or joules) and is denoted using a capital letter.

61

2 Mass and Energy Balance


Using this convention of the reference state absolute enthalpies for every chemical
compound can be defined using the concept of standard enthalpy of formation.

2.3.2 Standard Enthalpies of Formation


The standard enthalpy of formation of a substance is the reaction enthalpy R HT00
of its formation reaction from the pure elements in their most stable state at the
temperature T = 298.15 K and the pressure p = 1 bar (indicated by "0"). Again,
0
the overbar in hf,298 denotes molar values given per one kmol (or mol) of the
compound formed whilst subscript f has been added to stress that it concerns
the formation enthalpy. For example for the reaction of formation of H (hydrogen
atoms) written as:
1
H(g)
2 H2 (g)
0

0
R H298.15
= 1 (hf,298.15 )H 21 (hf,298.15 )H2 =

218 kJ/mol 12 0 kJ/mol = 218 kJ/mol

Fig. 2.10: Illustration of enthalpy of formation of a compound


0

Thus, the enthalpy of formation of H at p = 1 bar and T = 298.15 K is (hf,298.15 ) =


218 kJ/mol of H. If one writes the above reaction as
H2 (g) 2 H(g)
the reaction enthalpy is
0

0
R H298.15
= 2 (hf,298.15 )H 0.5 (hf,298.15 )H2 =

2 218kJ/mol 0.5 0 kJ/mol = 436kJ

62

2.3 Energy Released in Chemical Reactions


however, the formation enthalpy of H atoms remains 218 kJ/mol of H. Note that
if the enthalpy of formation of a compound is positive heat has to be supplied to
form the compound and the formation reaction is endothermic. If the formation
enthalpy is negative heat is generated (released) in the formation reaction that
itself is exothermic. Table 2.6 lists standard enthalpies of formation of some
compounds. Direct formation of a compound from its elements is often difficult to
realise in practice. For example the enthalpy of formation of neither the hydroxyl
radical
1
1
OH
2 H2 + 2 O2
nor methane
C + 2 H2 CH4
have been measured directly. However due to the work of Germain Henri Hess
published in 1840 one can determine indirectly the enthalpy of formation. Since
enthalpy is a state function it can be determined using reaction enthalpy of oxidation reactions that are relatively easy to measure. Below we will calculate the
enthalpy of formation of pure methane knowing the reaction enthalpies of other
oxidation reactions. Consider the following reactions:
C(s,graphite) + O2 CO2
H2 (g) + 12 O2 H2 O
CO2 + 2 H2 O(g) CH4 + 2 O2

R1 HT00 = 393.5 kJ/mol

(i)

R2 HT00
R3 HT00

= 241.81 kJ/mol

(ii)

= +802.25 kJ/mol

(iii)

By manipulating the above reactions ((i) + 2 (ii) + (iii)) one obtains the reaction
of formation of methane since:
C + O2 + 2 H2 + O2 + CO2 + 2 H2 O CO2 + 2 H2 O + CH4 + 2 O2
C + 2 H2 CH4
Therefore the reaction enthalpy of the last reaction is:
CH4 HT00 = R1 HT00 + 2 R2 HT00 + R3 HT00
= 393.5 + 2 (241.81) + 802.25
= 74.87 kJ/mol of CH4
Since only one mole of CH4 is formed in the above reaction, the formation enthalpy
0
of methane is (hf,298 )CH4 = 74.87kJ/mol of CH4 . The reader should check
(Table 2.6) that the calculated value is indeed close to the formation enthalpy of
methane.

63

2 Mass and Energy Balance

Table 2.6: Standard enthalpies of formation and standard entropies of some compounds (JANAF Thermodynamic Tables)
h

f,T0
Compound
Aggregation state
kJ/mol
Oxygen
O2 (g)
0.0
Hydrogen
H2 (g)
0.0
Nitrogen
N2 (g)
0.0
Chlorine
Cl(g)
0.0
Carbon
C(s,graphite)
0.0
Aluminium
Al(s)
0.0
Calcium
Ca(s)
0.0
Silicon
Si(s)
0.0
Sulphur
S(s,rhombic)
0.0
Phosphorus
P(s,white)
0.0
Oxygen atoms
O(g)
249.2
Ozone
O3 (g)
142.4
Hydrogen atoms
H(g)
218.0
Water vapour
H2 O(g)
241.81
Water
H2 O(l)
285.83
Hydroxyl radicals
OH(g)
39.3
Nitrogen atoms
N(g)
472.68
Nitrogen monoxide NO(g)
90.29
Nitrogen dioxide
NO2 (g)
33.1
Diamond
C(s,diamond)
1.9
Carbon
C(g)
716.6
Carbon monoxide
CO(g)
110.53
Carbon dioxide
CO2 (g)
393.5
Methane
CH4 (g)
74.85
Ethane
C2 H6 (g)
84.68
Propane
C3 H8 (g)
103.85
Ethylene
C2 H4 (g)
52.10
Benzene
C6 H6 (g)
82.93
Ethanol
C2 H5 OH(g)
235.31
g: gaseous, l: liquid, s: solid

64

s0T
0
J/(mol K)

205.04
130.57
191.50
223.07
5.74
28.33
41.42
18.83
31.80
41.09
160.95
238.8
114.6
188.72
69.95
183.6
153.19
210.66
239.91
2.38
157.99
197.6
213.7
186.10
229.49
269.91
219.45
269.20
282.00

2.3 Energy Released in Chemical Reactions

2.3.3 Lower Calorific Value (LCV) and Gross Calorific


Value (GCV)
In combustion engineering it is common to use calorific values to describe the
chemical enthalpy of the fuel. Both the lower (LCV) and higher (GCV) calorific
values are used. The higher calorific value represents the enthalpy of reaction of
stoichiometric combustion (oxidation) of fuel in oxygen with CO2 and H2 O being
the combustion products and under the condition that water has been condensed.
The lower calorific value is the reaction enthalpy with water remaining in gaseous
state (vapour). The examples below clarify the concepts.
H2 + 12 O2 H2 O(l)

+ 285 830 kJ/kmol of H2


142 915 kJ/kg of H2
11 530 kJ/mn 3 of H2
GCV

H2 + 21 O2 H2 O(g)

(oberer Heizwert oder Brennwert)

+ 241 900 kJ/kmol of H2


119 990 kJ/kg of H2
9758 kJ/mn 3 of H2
LCV

(unterer Heizwert oder Heizwert)

Typically both LCV and GCV are given at temperature T = 298.15 K and pressure
p = 1 bar. In some older text books LCV and GCV are given either at T = 273.15 K
or at T = 293.15 K; the differences are negligible.
For the above reactions: GCVH2 LCVH2 = enthalpy of condensation of water
produced in the oxidation reaction. Thus, GCVH2 LCVH2 = r = 285 830
241 900 = 43 930 kJ/kmol of H2 O
And in general:
GCV (in kJ/kmol of fuel) =
LCV (in kJ/kmol of fuel)+
r x (kmoles of H2 O produced per 1 kmol of fuel)
where r = 43 930 kJ/kmol H2 O or

65

2 Mass and Energy Balance

GCV (in kJ/kg of fuel) =


LCV (in kJ/kg of fuel)+
r x (kg of H2 O produced per 1 kg of fuel)
where r = 2440.56 kJ/kg H2 O
Definitions:
Lower Calorific Value (LCV) is the amount of heat obtained from a complete
combustion of a unit of fuel (one kg or one kmol) after cooling down the
products of combustion to the inlet temperature of the reagents and maintaining the water of the combustion products in vapour state. LCV depends
on temperature of the reagents and it is usually reported either at 0 C,
20 C or 25 C.

Fig. 2.11: Illustration of LCV as an amount of heat extracted from a combustion chamber

Gross Calorific Value (GCV) is the amount of heat obtained from a complete
combustion of a unit of fuel (one kg or one kmol) after condensing the water
of the combustion products and cooling down the products of combustion
to the inlet temperature of the reagents. GCV depends on temperature of
the reagents and it is usually reported either at 0 C, 20 C or 25 C.

66

2.3 Energy Released in Chemical Reactions

2.3.4 Relationships between Calorific Values, Reaction


Enthalpies and Formation Enthalpies
The standard formation enthalpies are specified at standard conditions, e.g. pressure of one bar and temperature of 298.15 K. Consider oxidation reaction of a
fuel at standard conditions. For example taking hydrogen as the fuel:
H2 + 21 O2 H2 O(g)

(p = 1 bar, T = 298.15 K)

the reaction enthalpy of the above reaction can be easily calculated as:
0
R H298.15
=

i h i =
i
0
(hf,298.15 )H2 O

(hf,298.15 )H2 12 (hf,298.15 )O2 = 241.81 kJ/mol

Thus, 241kJ of heat is liberated per each mole of hydrogen combusted. Recalling
the definition of the LCV one observes that the calculated absolute value of the
reaction enthalpy equals the LCV for hydrogen (water vapour remains in the gas
phase). The values of both LCV and GCV are always given as positive, so:


X



0
i hi
LCV (p = 1 bar; T = 298.15 K) = |R H298.15
|=
(2.65)


i

for oxidation reaction4 at standard conditions with water in the gas state,

and




X


0
i hi
GCV (p = 1 bar; T = 298.15 K) = |R H298.15
|=

(2.66)

for oxidation reaction at standard conditions with water in the liquid state.
Thus, LCVs listed in Table 2.7 are absolute values of enthalpies of oxidation reaction of the fuel at standard conditions with water remaining in the gas phase.

The oxidation reaction must be written for 1 unit (1 kmol of 1 kg) of fuel considered

67

2 Mass and Energy Balance


Table 2.7: LCV of some selected gaseous fuels (at p = 1 bar, T = 298.15 K)

Fuel

Formula

Hydrogen
Carbon oxide
Methane
Ethane
Propane
Butane
Pentane
Ethylene
Propylene
Butylene
Acetylene

H2
CO
CH4
C2 H6
C3 H8
C4 H10
C5 H12
C2 H4
C3 H6
C4 H8
C2 H2

Molar mass
g/mol
2.016
28.01
16.03
30.05
44.06
58.08
72.15
28.03
42.05
56.06
26.06

LCV

kJ/mol
241.9
283.0
802.5
1428.0
2044.1
2658.5
3272.9
1323.1
1926.0
2542.8
1255.9

kJ/mn 3
9758
11416
32372
57604
82457
107241
132025
53372
77693
102574
50662

kJ/kg
119990
10104
50062
47521
46394
45773
45362
47203
45803
45359
48193

Example 2.3
Calculate LCV of methane using values of formation enthalpies given in Table 2.6.
Assumptions: LCV is the absolute value of the enthalpy of the reaction of complete methane oxidation under the conditions that water remains in the gas phase.
The oxidation reaction for methane reads:
CH4 + 2 O2 > CO2 + 2 H2 O(g)
or
CH4 2 O2 + CO2 + 2 H2 O(g) = 0
Formula (2.65) is then used to calculate the reaction enthalpy:
X
0
R H298.15
=
i hi = (1) hCH4 + (2) hO2 + 1 hCO2 + 2 hH2 O
i

= (1) (74.85) + (2) 0 + 1 (393.5) + 2 (241.81)


= 802.27 kJ

Since the values of the formation enthalpies are at p = 1 bar and T = 298.15 K,
so the reaction enthalpy R H corresponds also to p = 1 bar and T = 298.15 K.

68

2.3 Energy Released in Chemical Reactions


Note that the negative sign of the reaction enthalpy indicates that the heat is
generated in the oxidation reaction of methane. Thus, the LCV of methane under
p = 1 bar and T = 298.15 K is:
LCVT00 = |802.27| = 802.27 kJ/mol of CH4
Comments: Compare the obtained results with the value listed in Table 2.7.
End of Example 2.3

2.3.5 Dependence of LCV on Temperature


In the previous sections, both Lower and Gross Calorific values are discussed
for standard conditions corresponding to a reference temperature of 298.15 K
and a reference pressure of 1 bar. The question is what is the LCV of a fuel at a
temperature that is different to 298.15 K? To answer the question we will consider
the oxidation reaction of the fuel, however our conceptual reactor is operated this
time at a constant temperature that is different to the standard temperature. The
reactor is fed with one mol of the fuel. Water and carbon dioxide are the only
products of the reaction. The water in the products remains in the gaseous phase.
If we calculate the (oxidation) reaction enthalpy, the LCV would be equal to the
absolute value of the reaction enthalpy. The oxidation reaction may be written
as follows:
1 f uel 2 O2 + 3 CO2 + 4 H2 O = 0
Then, the reaction enthalpy is
R H|0T

4
X

i h i

i=1

and elaborating it further:

R H|0T = hf uel,298.15 +

ZT

298.15

Cp,f uel dT

2 hO2 ,298.15 +

ZT

298.15

Cp,O2 dT +

69

2 Mass and Energy Balance

3 hCO2 ,298.15 +

ZT

Cp,CO2 dT +

298.15

4 hH2 O,298.15 +

ZT

298.15

After rearranging

Cp,H2 O dT = 0



0
0
0
0
R H|0T = hf uel,298.15 2 hO2 ,298.15 + 3 hCO2 ,298.15 + 4 hH2 O,298.15 +
ZT

Cp,f uel dT 2

298.15

ZT

298.15
ZT

Cp,O2 dT +

Cp,CO2 dT + 4

298.15

ZT

Cp,H2 OdT

298.15

and further
R H|0T

R H|0298.15

ZT

ZT

Cp,f uel dT 2

298.15

Cp,O2 dT +

298.15

ZT

Cp,CO2 dT + 4

298.15

ZT

298.15

and since LCVT0 = R H|0T one obtains:


all_molecules

LCVT0

0
LCV298.15

X
i

ZT

Cp,H2 O dT

Cp,i dT

298.15

or when splitting all the molecules taking place in the reaction into subtrates and
products to make it more transparent:

70

2.3 Energy Released in Chemical Reactions

LCVT0

0
LCV298.15

substrates
X

|i |

ZT

Cp,i dT

products
X

|k |

298.15

ZT

Cp,k dT (2.67)

298.15

Relationship (2.67) is known as the integral form of the Kirchhoff s law (note
that Eq. (2.67) is valid for 1 = 1 see the combustion reaction). The differential
form of the Kirchhoff law can be obtained by differentiating the above equations
with respect to temperature:
d LCV
=
dT

all_molecules

i Cp,i

(2.68)

or again making it more explicit:


d LCV
=
dT

substrates
X
i

|i | Cp,i

products
X

|k | Cp,k

(2.69)

Examining the equations describing Kirchhoff law one may observe that the dependence of the LCV with temperature stems from either the changes in the specific heat values with temperature or with the change of the number of mol in the
(oxidation) reaction.
Example 2.4
Calculate LCV of methane at 1298.15 K (1025 C) temperature and 1 bar pressure.
Assume that the molecules taking place in the oxidation reaction can be treated
as ideal gases.
Assumptions:
(a) LCV at T = 298.15 K at p = 1 bar is 802.5 kJ/mol of CH4 (see Table 2.7)
(b) Gases are ideal with constant specific heats as follows:
for CH4 , CO2 and H2 O Cp = 33.3 kJ/(kmol K)
for O2
Cp = 29.1 kJ/(kmol K)
The oxidation reaction of methane is as follows:

CH4 + 2 O2

CO2 + 2 H2 O
or
CH4 O2 + CO2 + 2 H2 O = 0

71

2 Mass and Energy Balance


The LCV at T = 1298.15 K can be calculated using Kirchhoff law:
all_molecules
0
LCV1298.15

0
LCV298.15

0
LCV1298.15
K

= 802.5 (1)

1298.15
Z K

Cp,i dT

298.15

1298.15
Z K

33.3 103 dT

298.15

(2)

(2)

1298.15
Z K

29.1 10

298.15
1298.15
Z K

dT (1)

1298.15
Z K

33.3 103 dT

298.15

33.3 103 dT = 802.5 + 2 (29.1 33.3) =

298.15

802.5 8.4 = 794.1 kJ/mol of CH4


Comments:
(a) Observe that the temperature correction to the LCV is rather small (8.4 kJ/mol)
if compared to the LCV value itself (802.5 kJ/mol). Therefore, in many engineering applications such a correction is unnecessary.
(b) More accurate calculations should incorporate the specific heat dependence
with temperature.
End of Example 2.4

2.3.6 Example of an Energy Balance of a Furnace


A prerequisite to a correct energy balance is an accurate mass balance (see Example 2.1). There are two different approaches to formulating an energy balance. In
the first approach, popular among chemical engineers, enthalpies of each incoming
and out-coming streams are calculated using enthalpies of formation. Typically
molar flow rates (kmol/h) are used in this case. In the second approach, popular
among combustion engineers, calorific values and mass flow rates (kg/h) are used.
Again both approaches are equivalent and lead to the same results. For a furnace
operated under a constant pressure and at a steady state the energy balance reads
(see Fig. 2.12)
H in = H out + Q + L
(2.70)

72

2.3 Energy Released in Chemical Reactions


done by the combustion products is zero; H in , H out stand for
where the work (L)
the enthalpy rate (in W) of the incoming and out-coming streams, respectively,
whilst Q is the rate of heat removal (in W) from the furnace and this includes the
heat transferred to the process (heat sink) as well as the heat losses. The sketch
below clarifies further the meaning of the symbols used in Eq. (2.70).

control volume
.
Hin

Furnace

.
Hout
.
L

Heat Sink (process)

.
Q
Fig. 2.12: Illustration of an energy balance

In order to calculate H in and H out both the amount and composition of the
incoming and out-coming streams have to be known and these are obtainable
from a correctly made mass balance. Then the energy balance reads:
in_species

X
i

ZTin
n i hf,i +
Cp,i dT =

T0

out_species

X
k

n k hf,k +

T
Zout

T0

Cp,k dT + Q (2.71)

where n i stands for the molar flow rate (kmol/s) of each incoming species (type
of molecule), hf,i is the formation enthalpy of the species and Cp,i is the molar
specific heat at constant pressure. For the out-coming streams the summation
extends over all the combustion products including water. Note that the enthalpy
of each species consist of two parts since:
hi = hf,i +

ZT

Cp,i dT

(2.72)

T0

where hf,i represents so called "chemical enthalpy" while the integral represents

73

2 Mass and Energy Balance


"physical enthalpy". Both are calculated using the standard state conditions as
the reference level of zero enthalpy. The formulation (2.72) is consistent with a
statement that any (molecule) species has a chemical energy (positive or negative)
that is either released or taken during chemical reactions. Only the most stable
species (see Table 2.7) under the standard conditions have no (zero) chemical enthalpy as a consequence of the definition of the reference state. For the incoming
streams the integral is insignificant at ambient air conditions but it can be important when either the fuel or oxidiser or both are preheated. For the out-coming
streams the integral represents the physical enthalpy of combustion products.
Combustion engineers calculate the incoming H in enthalpy using Lower Calorific
Value rather than enthalpies of formation. The concept here is that the chemical
energy associated with the combustion reaction is bound with the fuel only. In
other words neither CO2 nor H2 O have chemical enthalpies. After splitting H in
into the oxidiser (air) stream and the fuel stream the energy balance reads:

ZTin
ZTin
m
f uel LCV +
cp dT + m
oxidiser
cp,oxidiser dT =

T0

T0

m
products

T
Zout

cp,products dT + Q (2.73)

T0

and
(2.74)

m
f uel + m
oxidiser = m
products

where m
f uel , m
oxidiser and m
products stand for the mass flow rates (kg/h) of the
fuel, oxidiser and combustion products, respectively. Note that the reference
state (p = 1 bar, T = 298.15 K) for making the energy balance is the same for
Eq. (2.71) and (2.73). In Eq. (2.73) the specific heats are expressed per kilogram.
Relationship (2.73) can be rewritten as:

ZTin
ZTin
oxidiser
cp,oxidiser dT =
cp dT + m
m
f uel LCV +

T0

T0

all_products

X
k

m
k

T
Zout

cp,k dT + Q (2.75)

T0

where m
k stands for the mass flow rate of species k present in the products.

74

2.3 Energy Released in Chemical Reactions


Formulations (2.71) and (2.73) are equivalent. They both have advantages and
disadvantages. The chemical engineering formulation (2.71) is general in the sense
that there is no need to distinguish between the fuel and the oxidiser. They both
take part in the reactions and as long as the reactions are written correctly,
calculating H in and H out enthalpies is relatively straightforward. The generality
of Eq. (2.71) makes it ideal for development of general computer packages for
computer simulations of industrial processes. However, in Eq. (2.71) there is no
explicit term that would show how much energy enters into the system. Thus,
answering the question "what is the thermal input of the system considered?"
requires some further calculations. On the other hand, Eq. (2.73) requires a priori
determination of the fuels and therefore it is less general. The left hand side of
Eq. (2.73) represents explicitly the total (chemical + physical enthalpy) thermal
input, the process variable that is essential in thermal engineering. Furthermore,
the out-coming flow rate of combustion products is rather straight forward to
obtain since the flow rates are in kg/h.
Example 2.5
Consider again the boiler of Example 2.1. Make the energy balance of the boiler
knowing that 2 MW of heat is extracted by the water tube wall (heat sink).
Both the fuel and the air at the burner are at ambient temperatures (Tin =
298.15 K). Calculate the temperature at the boiler exit and the efficiency of the
process. Carry out the energy balance using enthalpies of formation and repeat
the calculations using LCV. For the sake of simplicity treat the gases as ideal with
constant (independent of temperature) specific heats.
Assumptions:
(a) the boiler operates at a steady state,
(b) the control volume encompasses the boiler,

control volume
200 kg/h CH4

Boiler

Combustion
products

4000 kg/h air


Heat Sink

.
Q = 2MW
Fig. 2.13: Example of an energy balance

75

2 Mass and Energy Balance


(c) the gases are treated as ideal with constant specific heats as follows:
for CH4 , CO2 and H2 O Cp = 33.3 kJ/(kmol K)
for O2 , N2
Cp = 29.1 kJ/(kmol K)
Energy balance formulated using enthalpies of formation
The molar flow rates are given in Table 2.1 while the standard enthalpies of
formation are taken from Table 2.6. For the incoming streams (CH4 , O2 and N2 )
the left hand side of Eq. (2.71) is as follows:


RT
12.5
103 + T00 33.3dT
= 259.9 kJ/s
Methane
3600 74.85

R
T0
109.57
=
0 kJ/s
Nitrogen
3600  T0 29.1dT
R
T0
29.13
Oxygen
=
0 kJ/s
3600
T0 29.1dT
TOTAL INPUTS

259.9 kJ/s

For the out-coming streams (CO2 , H2 O, N2 , O2 ) (the right hand side of Eq. (2.71))
the following is applicable:

Heat extracted


12.5
393.5 103 + 33.3 (Tout T0 ) =
3600
1366.32 + 0.115 63 (Tout T0 )

25
241.81 103 + 33.3 (Tout T0 ) =
3600
1679.24 + 0.231 25 (Tout T0 )
109.57
(29.1 (Tout T0 )) = 0.8856 (Tout T0 )
3600
4.13
(29.1 (Tout T0 )) = 0.0334 (Tout T0 )
3600
= 2000.00

TOTAL OUTPUTS

= 1045.56 + 1.2659 (Tout T0 )

Carbon dioxide:

Water vapour:

Nitrogen:
Oxygen:

Thus, the energy balance reads:


259.9 = 1045.56 + 1.2659 (Tout T0 ) (both sides in kJ/s)
and Tout can be readily calculated to be Tout = 918.81 K.

76

2.3 Energy Released in Chemical Reactions

-259.9 kJ/s Chemical


Enthalpy of CH4
Nil Physical Enthalpy
Furnace

control volume
-1045.56 kJ/s Chemical Enthalpy
785.69 kJ/s Physical Enthalpy
of Combustion Products

Tin =298.15K

Tout =918.81K
Heat Sink

(process)

.
Q = 2MW
Fig. 2.14: Energy balance using enthalpy of formation

The thermal input into the furnace consist of the physical enthalpy and the chemical enthalpy. Since the inlet temperature of the substrates (methane and air)
equals the reference temperature the physical enthalpy at the inlet is zero. The
chemical enthalpy, expressed using heats of formation, is: 259.9 kJ
s .
The combustion products enthalpy consists of the physical and chemical enthalpies. The chemical enthalpy includes the carbon dioxide component
(1366.32kJ/s) and the water vapour component (1679.24kJ/s), so its value is
3045.56kJ/s. The physical enthalpy of the combustion products which includes
all the components (CO2 , H2 O, N2 and O2 ) equals 1.2659 (918.81 298.15) =
785.69kJ/s. The efficiency of the boiler can then be calculated as
Ef f iciency =
=

heat extracted by the water tube walls


enthalpy of inlet streams chemical enthalpy of combustion products
2000kJ/s
= 0.718
=
259.9kJ/s (3045.56)kJ/s

Energy balance formulated using LCV


The mass flow rates of the streams are given in Table 2.3 whilst the LCV of CH4
is taken from Table 2.7 to be 50 062 kJ/kg of CH4 . The enthalpy of the incoming
streams is as follows (left hand side of Eq. (2.73)):


R T0
200
33.3dT
/16
= 2781.22 kJ/s

50
062
+
Methane
3600
T0

R
T0
39 068
Nitrogen
=
0 kJ/s
3600  T0 29.1dT /28

R
T0
932
Oxygen
=
0 kJ/s
3600
T0 29.1dT /32
TOTAL INPUTS

2781.22 kJ/s

77

2 Mass and Energy Balance


The figure of 2.78 MW is the total thermal input into the furnace that consists of
the chemical enthalpy of the fuel (2.78 MW) and the physical enthalpy of the fuel
and of the air stream.
For the out-coming streams (CO2 , H2 O, N2 , O2 ) (the right hand side of Eq. (2.73))
the following is applicable:


550 33.3
(Tout 298.15) = 0.115 63 (Tout 298.15)
Carbon dioxide:
3600
44


450 33.3
Water vapour:
(Tout 298.15) = 0.231 25 (Tout 298.15)
3600
18


3068 29.1
Nitrogen:
(Tout 298.15) = 0.8856 (Tout 298.15)
3600
28


132 29.1
(Tout 298.15 = 0.0334 (Tout 298.15)
Oxygen:
3600
32
Heat extracted
= 2000.00
TOTAL OUTPUTS

= 2000.00 + 1.2659 (Tout 298.15)

Thus, the energy balance reads:


2781.22 = 2000.00 + 1.2659 (Tout T0 ) (in kJ/s)
and Tout = 915.28 K. The physical enthalpy of the combustion products can now
be calculated as 1.2659 (915.28 298.15) = 780.87 kJ/s and the energy balance
closes since:
Input = 2.78 MJ/s Output = 2 MJ/s
(heat extracted)
0.78 MJ/s (physical enthalpy of products
at Tout = 915.28 K)
control volume

2.786MW Chemical
Enthalpy of CH4
Nil Physical Enthalpy

0.786MW Physical Enthalpy


of Combustion Products
Furnace
Tout =918.81K

Tin =298.15K
Heat Sink (process)

.
Q = 2MW
Fig. 2.15: Energy balance using LCV

78

2.4 Temperature of Adiabatic Combustion


The efficiency of the boiler is then
Ef f iciency =

2M W
heat extracted by the water tube walls
=
= 0.718
total energy input
2.78M W
(2.76)

Comments:
(a) Observe that when enthalpies of formation are used, the total enthalpy of the
incoming streams can be negative; total enthalpy of the out-coming streams
can also be negative.
(b) When LCV is used enthalpies of the incoming streams and out-coming streams
are always positive.
(c) When enthalpies of formation are used, the thermal input into the furnace is
hidden in the energy balance Eq. (2.71).
(d) When LCV is used, the thermal input into the furnace is explicitly given since
it is just the left hand side of Eq. (2.73).
(e) The furnace exit temperature is calculated to be 918.81 K when the energy
balance is formulated using the enthalpies of formation. When LCV is used,
the furnace exit temperature is 915.28 K. What is the reason for the 3.52 K
difference?
End of Example 2.5

2.4 Temperature of Adiabatic Combustion


In an adiabatic system (Q = 0) the combustion products after complete combustion carry the physical enthalpy equal to the total thermal input, as shown by
Eq. (2.73). If combustion is stoichiometric, the temperature of the combustion
products is the highest achievable for this specific fuel. This temperature is called
the temperature of adiabatic combustion (adiabatic flame temperature) under stoichiometric conditions and is a property of the fuel.
Inserting into Eq. (2.73) the relationships Q = 0 (adiabatic combustion) and
m
oxidiser = lair,min m
f uel (stoichiometric combustion) one obtains:

ZTin
ZTin
m
f uel LCV +
cp dT + lair,min m
f uel
cp,oxidiser dT =

T0

T0

79

2 Mass and Energy Balance


ZTad
cp,products dT
(m
f uel + lair,min m
f uel )
T0

and after dividing both sides by m


f uel
ZTin
ZTin
cp,oxidiser dT =
LCV +
cp dT + lair,min
T0

T0

ZTad
cp,products dT
(1 + lair,min )

(2.77)

T0

Eq. (2.77) can be used to calculate the adiabatic combustion temperature Tad
under stoichiometric conditions. Note that Tad is uniquely defined for any fuel
and therefore is regarded as a fuel property. Table 2.8 lists adiabatic flame temperatures for some fuels which are calculated using Eq. (2.77); the numbers refer
to Tin = 298.15 K, p = 1 bar.
Table 2.8: Adiabatic flame temperature Tad for stoichiometric combustion in air (Tin =
298.15 K, p = 1 bar). Combustion products contain CO2 and H2 O only.

H2
C2 H2
C3 H8

2473 K
2936 K
2400 K

CH4
C2 H6
CO

2285 K
2357 K
2624 K

Table 2.8 shows the calculated values of Tad providing CO2 and H2 O are the only
combustion products. For temperatures in excess of 2273 K dissociation of CO2
and H2 O takes place according to the reactions:

CO2

CO + O

and

H2 O

OH + H

Both reactions are endothermic therefore the real adiabatic flame temperatures
are lower than those listed in Table 2.8. For temperatures in excess of 2273 K,
the combustion products contain not only CO2 and H2 O but also CO, H2 , O, OH
and N radicals. In such high temperatures the combustion products composition
should be determined using chemical equilibrium considerations (see Chapter 4).

80

2.5 Furnace Exit Temperature

2.5 Furnace Exit Temperature


The energy balance for any non-adiabatic combustion system can be written as:

ZTin
ZTin
m
f uel LCV +
cp,f uel dT + lair,min m
f uel
cp,air dT

T0

T0

= (m
f uel + lair,min m
f uel )

T
Zout

cp,products dT + Q (2.78)

T0

In the above equation we assume that the fuel is oxidised completely to CO2 and
H2 O and air is used as oxidiser. However by replacing lair,min with an equivalent variable, one may write such an equation for any oxidiser namely oxygen or
enriched air. The left hand side of Eq. (2.78) can be rewritten as:

TTI = m
f uel LCV +m
f uel


FUEL THERMAL INPUT

ZTin
ZTin
cp,f uel dT + lair,min m
f uel
cp,oxidiser dT

T0

@
I
@
@

T0

PHYSICAL ENTHALPY (PREHEAT)

where TTI stands for total thermal input into the system.
Dividing both sides of Eq. (2.78) by m
f uel LCV and after some algebra one
obtains:
R Tout
(1 + lair,min ) 298.15
cp,products dT
Q
=1
+
m
f uel LCV
LCV
R Tin,air
R Tin,f uel
cp,air dT
lair,min 298.15
298.15 cp,f uel dT
+
LCV
LCV

(2.79)

m
The ratio Q/(
f uel LCV ) is the fraction of the fuel thermal input extracted by
the process (including heat losses). For a given fuel and a given preheat, the ratio
is a function of the excess air () and the furnace exit temperature Tout only, so
that
Q
= f (, Tout )
(2.80)
m
f uel LCV
Combustion engineers use this function to estimate the amount of the energy

81

2 Mass and Energy Balance


available for the process. It is worth observing that for
Q 0

Tout Tad

and if Tin = 298.15 K (no preheat) Eq. (2.79) simplifies further to:
RT
(1 + lair,min ) T0out cp,products dT
Q
=1
X=
m
f uel LCV
LCV

(2.81)

The ratio X = m f uelQ LCV is a fraction of the fuel thermal input that is available
for the process and therefore it is often called in short "percentage available
heat". A Sankey diagram shown in Fig. 2.16 visualises the energy balance of a
furnace operated without air preheat. The figure provides a further explanation
for the "available heat" concept.

Fig. 2.16: Sankey diagram demonstrating the concept of the available heat

82

2.5 Furnace Exit Temperature


Example 2.6
Generate a graph showing relationship (2.81) for combustion of pure methane in
air.
Assumptions:
(a) LCV of methane at T = 298.15 K and at p = 1 bar is 50 062 kJ/kg of CH4 (see
Table 2.7),
(b) the combustion products contain CO2 and H2 O and thermal dissociation is
ignored,
(c) as a first estimation we assume that the gases are ideal with constant specific
heat capacities as follows:
for CH4 , CO2 and H2 O

Cp = 33.3 kJ/(kmol K)

for O2 , N2

Cp = 29.1 kJ/(kmol K)

(d) we repeat the calculations using Cp polynomials to examine the effect of the
assumption (c)
The oxidation reaction of CH4 is

CH4 + 2 O2

CO2 + 2 H2 O
or
16 kg CH4 + 64 kg O2 = 44 kg CO2 + 36 kg H2 O
The minimum air requirement is 64/16/0.233 = 17.1674 kg of air/kg of methane.
For > 1 the combustion products contain CO2 , H2 O, N2 , and O2 . Per 1 kg of
CH4 , the combustion products contain:
Methane
Carbon dioxide
Water vapour
Nitrogen
Oxygen
TOTAL (wet) =

44/16 =
36/16 =
0.767 17.1674 =
0.233 ( 1) 17.1674 =

none
2.75 kg
2.25 kg
13.1674 kg
( 1) 4 kg
1 + 17.1675 kg

83

2 Mass and Energy Balance


Composition of the wet combustion products is:
wCO2 = 2.75/(1 + 17.1675 )
wH2 O = 2.25/(1 + 17.1675 )
wN2 = 13.1675 /(1 + 17.1675 )
wO2 = 4 ( 1)/(1 + 17.1675 )
The above expressions are needed for evaluating the specific heats of combustion
products since
cp,products = wCO2

Cp,CO2
Cp,H2 O
Cp,N2
Cp,O2
+ w H2 O
+ w N2
+ wO 2
44
18
28
32

and the physical enthalpy of the wet combustion products is then


T
Zout

cp,products dT =

298.15

(wCO2

33.3
29.1
29.1
33.3
+ wH2 O
+ w N2
+ wCO2
) (Tout 298.15) (B1)
44
18
28
32

Using Cp polynomials given in Table 2.5, one can evaluate the physical enthalpy
of the combustion products more accurately:
T
Zout

cp,products dT =

298.15

products_species

wi
1
2
{(Cpi,1 (Tout 298.15) + Cpi,2 (Tout
298.152 )+
Mi
2

1
1
3
4
Cpi,3 (Tout
298.153 ) + Cpi,4 (Tout
298.154 )+
3
4
1
5
Cpi,5 (Tout
298.155 )} (B2)
5

Now expression (2.81) can be put together as:


RT
(1 + 17.1674) T0out cp,products dT
Q
X=
=1
m
f uel LCV
50 062

(B3)

where the integral can be evaluated (for any Tout and > 1) using either simplified

84

2.5 Furnace Exit Temperature


expression (B1) or accurate Eq. (B2).
Relationship (B3) for combustion of pure methane is shown in Fig. 2.17 and
Fig. 2.18. Obviously for specific heat capacities taken as independent of temperature, the relationship is linear for any excess air ratio (). For a given furnace exit
temperature the amount (fraction) of the energy extracted is a function of and
therefore any uncontrolled inleakage of air into the furnace should be avoided. The
point of intersection of the line with the X-axis shows the adiabatic combustion
temperature for given (Tin = 298 K). Note for example that for stoichiometric
combustion of pure methane we have calculated the adiabatic combustion temperature around 2800 K when constant cp values were used (Fig. 2.17). When
the specific heat values are evaluated using the JANAF polynomials the relationships become non-linear as shown in Fig. 2.18. More importantly the results
differ substantially from those shown in Fig. 2.17. The adiabatic temperature of
stoichiometric combustion decreases by 525 K as summarised in the table below.
Table 2.9: Calculated adiabatic temperature for stoichiometric combustion of pure
CH4 with air (Tin = 298 K, p = 1 bar). Compare with Table 2.8

Tad

constant cp values
2810 K

cp from JANAF tables


2285 K

Q
Fig. 2.17: Fraction of the available heat (X = m f uel cdotLCV
) as a function of the furnace exit temperature and excess air ratio for combustion of pure methane
in air. Constant cp values.

85

2 Mass and Energy Balance

Fig. 2.18: Fraction of the available heat (X = m f uelQ LCV ) as a function of the furnace
exit temperature and excess air ratio for combustion of pure methane in air.
JANAF polynomials for cp have been used.

Comments:
(a) Note the pronounced effect of using Cp polynomials.
(b) At temperatures of around 2300 K some (not much) CO2 and H2 O may dissociate so precise calculations should be carried out using a chemical equilibrium
procedure (see Chapter 4).
End of Example 2.6

2.6 Summary
Conservation of mass and energy is the basis of combustion engineering. Identification of the control volume and the specification of an appropriate time basis are
essential skills in formulating mass and energy balances. A correct mass balance
of a system is a prerequisite to a subsequent energy balance.
In this lecture the first law of thermodynamics has been recalled in order to identify various forms of energy supplied to and removed from a system. It has been
shown that in combustion engineering, or more general in thermal engineering,
the internal energy is typically the largest energy and the kinetic and potential

86

2.6 Summary
energies can be neglected. The student should realise that the first law of thermodynamics, written for any systems (open or closed) takes the form
dqt p dv = du
expressing the fact that the internal energy (du) can be altered either by the
(total) heat supplied to the system (dqt ) or the work done to the system (p dv)
or both. Equivalently, the above relationship can be written as
dqt + v dp = dh
where h stands for specific enthalpy defined as u + p v = h.
In engineering practises furnaces, boilers and chemical reactors are often operated
at a constant pressure and then
dqt = dh

when

p = const.

For autoclaves and chemical reactors operating at a constant volume (batch processes) one obtains
dqt = du
when
v = const.
It is essential to realise that it is not possible to calculate absolute values of energy.
Instead human beings have defined a reference state for calculating energy as a
state of pure elements in their most stable state at T0 = 298.15K and p0 =
1bar. By convention this state is prescribed zero enthalpy (see Table 2.6). Thus,
enthalpies at any other states are calculated in reference to this reference state as
follows
ZT,p
0
h(T, p) = hf,298 +
dh
T0 ,p0

0
hf,298

where
is the molar formation enthalpy. In this way the total specific enthalpy of a species consists of the chemical enthalpy and the physical enthalpy.
For furnaces and boilers which operate at a steady state and at a constant pressure, Eq. (2.71) formulates the energy balance, or more precisely enthalpy balance,
assuming that the gases are ideal. Formulation (2.71) is general in the sense that
there is no need to distinguish between the fuel and the oxidiser. However in
formulation (2.71) there is no explicit term that would show the thermal input
into the system. For a combustion engineer the thermal input into the system
is of paramount importance. Furthermore, for a combustion engineer all the
chemical enthalpy is associated with the fuel only and carbon dioxide and wa-

87

2 Mass and Energy Balance


ter vapour molecules which are the products of the fuel combustion contain no
chemical enthalpy. Following this concept, the notion of Calorific Values (Lower
and Higher/Gross) has been developed leading to the enthalpy balance equation
(2.73). In this equation the left hand side shows explicitly the thermal input
into the furnace. Both approaches, through Eq.(2.71) or through Eq.(2.73) are
equivalent and the student must command them.
In this lecture we have also introduced the temperature of adiabatic combustion.
If adiabatic combustion is stoichiometric, the temperature of the combustion products is the highest achievable for this specific fuel and therefore this temperature
is a property of the fuel.
While considering the first law of thermodynamics, we have reiterated that both
internal energy and enthalpy are state variables and their differentials are exact.
Work and heat are not state variables and their differentials are inexact.

88

3 Equilibrium Thermodynamics
Contents
3.1
3.2

3.3

3.4

3.5

3.6

Irreversible and Reversible Processes


Entropy
3.2.1

Entropy of Liquids and Solids

3.2.2

Entropy of Ideal Gases

3.2.3

Entropy of Phase Transition at the Transition Temperature

3.2.4

The Third Law of Thermodynamics

3.2.5

Absolute Entropy of Pure Substances

The Second Law of Thermodynamics


3.3.1

The Increase in Entropy Principle

3.3.2

Entropy Change for a Continuous Process at Steady-State

3.3.3

Irreversibility of Processes

General Conditions for Thermodynamic Equilibrium


3.4.1

Isolated System

3.4.2

Non-Adiabatic System

Equilibrium Between Phases


3.5.1

Single-Component System Consisting of Two Phases

3.5.2

Phase Transformations of a Pure Substance

3.5.3

Dependence of Gibbs Free Enthalpy on Temperature


and Pressure

3.5.4

Equilibrium in Multi-Component Single-Phase Systems

3.5.5

Chemical Potential of Pure Substances

3.5.6

Significance of Chemical Potential

Multi-Component, Multi-Phase Systems


3.6.1

The Phase Rule

3.7

Thermodynamics of Mixing

3.8

Summary

89

3 Equilibrium Thermodynamics

3.1 Irreversible and Reversible Processes


In nature some things happen spontaneously (naturally), some do not. For
example a hot body cools down to a temperature of the surroundings in a spontaneous process of giving up heat. After a sufficiently long time equilibrium is
reached and the bodys temperature reaches surroundings temperature. A compressed gas stored in a cylinder expands spontaneously into surroundings as soon
as the interaction with the surroundings is made possible by opening the cylinder
valve. Again after a long enough time equilibrium is reached; the gas pressure
in the cylinder equals the surroundings pressure and the gas outflow from the
cylinder ceases. A cigarettes smoke mixes into the air by a (spontaneous) process of molecular diffusion. All these spontaneous processes proceed naturally, in
one direction, and there is no need to do any work to bring the changes about.
These spontaneous processes are irreversible. By calling a process irreversible
we mean that the initial state, or any other past state, cannot be reached by any
spontaneous process that would proceed in the opposite direction.

surroundings

surroundings
valve

Tsurr
Q
hot
body

T
p
T

Tsurr
psurr

surroundings
cigarette
smoke

compressed
gas

Fig. 3.1: Examples of irreversible processes in closed and open systems.


Left hot body cooling down to surroundings temperature;
Middle compressed gas expanding to surroundings pressure (throttling);
Right cigarette smoke diffusing into surroundings.

Consider again our three examples: the hot body being cooled down, the compressed gas expanding to surroundings pressure (the process is called throttling)
and the cigarette smoke diffusing into the air. The first example is a case of
a closed system while the second and the third examples are open systems, see
Fig. 3.1. During theses spontaneous processes the system (the hot body, the
compressed gas, the cigarette smoke) interacts with surroundings. The irreversibility of open or closed processes means that it is not possible to
get back to any previous states of the system and the surroundings.
After cooling the hot body to surroundings temperature, it is not possible to heat
it up again using surroundings, even though doing so would not violate the first

90

3.1 Irreversible and Reversible Processes


law of thermodynamics. Similarly after expanding the gas to surroundings pressure it is not possible to compress the gas back to any previous states using only
the surroundings. As soon as the cigarette smoke disperses into the air, there
is no way of bringing it back to the cigarette tip using only surroundings. It
does not mean that the original situation can never be restored. It can, provided
that the system plus the surroundings are not kept constant but are allowed to
interact with other systems. Obviously the hot body that has been cooled down
to the surroundings temperature can be heated up back to its initial temperature. However to accomplish this we cannot use the heat that the body has given
up to surroundings but we must employ another source of energy for example a
burner or an electrical heater. Similarly, we can compress the gas to its initial
pressure using a compressor however, it is not possible to achieve it using energy
of surroundings only.

Fig. 3.2: Irreversible process of mixing. An isolated system consisting of a box divided
by a partition. The left hand side is filled with a gas whilst the right hand
side is empty.

As an example of an isolated system consider a box divided by a partition into


two equal parts, each of volume V , Fig. 3.2. The left hand side is filled with gas
while the right hand side is empty. Suppose now that the partition is removed.
As a result of collisions with the walls and with each other, the molecules will
very quickly redistribute themselves over the entire volume of the box. The final
equilibrium state, where the density of molecules is uniform throughout the entire
box is attained rather quickly. The question is what determines the direction of
this spontaneous mixing process? It is certainly not the total energy of this
isolated system since it remains constant through the whole process. However
when a change within the box occurs, the energy is parcelled out (distributed)
differently. This irreversible mixing process proceeds towards the greater chaotic
dispersal of the total energy of the system. The irreversibility of an isolated
system means that it is not possible to reach any previous states of
the system and the process proceeds in one direction only. The spontaneous,

91

3 Equilibrium Thermodynamics
irreversible processes are always associated with a redistribution of energy into a
more disordered form.
There is another observation to be made in conjunction with irreversible processes
taking place in open or closed systems. The more spontaneous and violent the
process is, the more difficult it is to bring the system to the initial state since the
changes ("damages") to surroundings are larger. When the body is hotter or more
precise when the temperature difference between the body and the surroundings is
larger, the cooling process proceeds faster. To bring the body back to the initial
state, we would have to use an energy source (a burner, an electrical heater)
capable of heating the body to a higher temperature.
The second law of thermodynamics governs the direction of irreversible processes.
Historically there have been many formulations of the law. Probably the simplest
formulation is due to Rudolf Clausius who in 1850 stated that
"Heat cannot of itself pass from a colder to a hotter body."
Other more elaborate formulation due to William Thomson (Lord Kelvin) reads:
"A transformation whose only final result is to transform into work
heat extracted from a source which is at the same temperature
throughout is impossible."
Both of these formulations are not applicable in engineering practise. They are
just statements indicating in which direction heat can (or rather cannot) be transferred. After introducing the notion of entropy another formulation can be derived
(see Section 3.3.1) which allows the second law of thermodynamics to be used in
engineering.
Irreversible processes reach a different "degree of irreversibility" depending on
how the processes proceed. Thus, if we alter the way of carrying out the process
we may minimise the changes ("damages") to the surroundings. Perhaps we can
even carry out the process in such a way that the degree of irreversibility would
be zero and the process would be reversible. A reversible process is able to
bring the system to any previous states, and finally to the initial state,
without any changes to surroundings. This is possible only if the net heat
and net work exchange between the system and the surroundings is zero for the
combined (original and reverse) process.
Reversible processes do not occur in nature. However we mentally (conceptually) construct reversible processes in order to compare real, irreversible processes
against these ideal conceptual ones. Reversible processes can be approximated
by actual devices but they can never be achieved. Engineers are interested in

92

3.1 Irreversible and Reversible Processes


reversible processes because devices producing work (engines, turbines) deliver
maximum work when they realise reversible processes. Devises like pumps and
compressors require minimum work when they realise reversible processes. Thus,
reversible processes can be regarded as theoretical limits of the corresponding real processes.
Consider a closed system (of constant mass) that is at an initial state marked in
Fig. 3.3 as state 1. By applying (adding) an infinite number of combinations of
heat and work (see Eq. (2.56)) to the system, the matter of the system can reach
an infinite number of states. Among these infinite number of thermodynamic
processes the following five are typically considered:
(a) isothermal process (either expansion or compression); T = const.,
(b) processes proceeding under constant volume; V = const.,
(c) processes proceeding under constant pressure; p = const.,
(d) adiabatic processes; Q = 0,
(e) polytropic processes, pV = const. with = const. where is called polytropic exponent.
Fig. 3.3 shows these typical processes using a p-V (work) diagram. All of them
might be carried out as reversible processes if additions of the infinitesimal amounts
of heat and work is carried out in a series of quasi-static steps (see below). However
there are two thermodynamic processes that are highly irreversible and cannot be
reversed. These are: throttling (expansion without any work) and mixing. Fig. 3.1
(middle) shows an example of throttling while Fig. 3.1 (right) and Fig. 3.2 show
a highly irreversible process of mixing. Chemical reactions are highly irreversible.
Processes shown in Fig. 3.3, can be carried out reversibly if the infinitesimal
amounts of heat and work are added (or removed) into the system in a process
which is carried out so slowly that the system remains arbitrarily close to equilibrium at all stages of the process. Such a process is said to be "quasi-static",
or "quasi-equilibrium" for the system. A quasi-static process can be viewed as
a sufficiently slow process which allows the system to adjust itself internally so
that properties in one part of the system do not change any faster than in another
part.

93

3 Equilibrium Thermodynamics

V=const.

1 p=const.

T=c
o
2

V1

.
Q

nst.

1-2 =0

2
2

V2

L1-2
.
Q1-2
Fig. 3.3: Typical thermodynamic processes shown using work-diagram; T = const.,
V = const., p = const., Q = 0, (pV = const. not shown).

How slowly one must proceed to make sure that the process is quasi-static depends
on the time (called "relaxation time") that the system requires to attain equilibrium if it is suddenly disturbed. To be slow enough to be quasi-static implies
that the process proceeds slowly compared to the relaxation time. For example,
if the gas in Fig. 3.3 returns to equilibrium within a time of 105 seconds after
the piston is suddenly moved back from position 2 to position 1, then a process
wherein the piston is moved in 0.1 second can be considered quasi-static to a good
approximation.
Consider a piston-cylinder, Fig. 3.4, that contains a gas at initial pressure p1 .
When the piston is moved suddenly to the right the molecules near the face of
the piston move to the space made just available, and therefore the pressure near
the piston decreases. Because of this low pressure, the system (gas) is not longer
in equilibrium since a higher pressure exists at the other (left hand) end of the
cylinder. Such a process is called non quasi-equilibrium. However, if the piston is
moved slowly, the molecules have sufficient time to redistribute and the pressure in
the entire volume of the gas remains always uniform. Since equilibrium within the
gas prevails at all times, this is a quasi-equilibrium or quasi-steady-state process.
In order to carry out a reversible expansion of the compressed gas a series of
infinitesimal reversible steps would have to be realised as shown in Fig. 3.4.

94

3.1 Irreversible and Reversible Processes

Fig. 3.4: Gas expansion process;


Left Irreversible (non-quasi-equilibrium) expansion,
Right Reversible (quasi-equilibrium) expansion.

The work done to the system (gas) during the process is

Lrev =

final
Z state

p dV = Lmax

(3.1)

initial state

The maximum work available from such an expansion from initial to final
states is obtained when the process is carried out reversibly.
The irreversibility of a process is caused by the occurrence of the followings:
(a) friction,
(b) heat transfer through a finite temperature difference,
(c) throttling (gas expansion without any work obtained),
(d) non-quasi equilibrium compression or expansion,
(e) electrical resistance (dissipation of heat),
(f) chemical reactions (destruction of chemical bonds).

95

3 Equilibrium Thermodynamics

3.2 Entropy
In Chapter 2 we have introduced the first law of thermodynamics and one of the
most useful formulations of this law (see Eq. (2.52)) is:
dQt + dL = dU

(3.2)

where U stands for the internal energy and L is the absolute work done to the
system. Symbol Qt (t stands for total) represents the total heat absorbed by the
system:
dQt = dQf + dQ
(3.3)
where Q is the amount of heat supplied from the surroundings or any other heat
source, and Qf is the amount of heat generated within the system due to friction.
All these terms are expressed in joules.
We have already learned that internal energy (U ) is a state variable and its differential is exact while neither heat (Q) nor work (L) are state variables and their
differentials are inexact. However, T1 is an integrating factor for dQt turning it
into an exact differential called entropy (S):
dS =

dQt
T

(3.4)

where dQt is the infinitesimal amount of heat exchanged and T is the temperature
of the quasi-equilibrium, reversible process (see Example 3.1). Entropy is an
extensive property so
S = ms
(3.5)
where S is the total entropy of the system in J/K, m is the mass of the system
in kilograms and s is the specific entropy of the system in J/kg K. Therefore
ds =

dqt
T

(3.6)

Example 3.1
The objective of this example is to demonstrate how an inexact differential can
be transformed into an exact one. We wish to improve the understanding of
relationships (3.4) and (3.6) which transform an inexact differential of heat into
an exact differential of entropy. Consider the infinitesimal quantity:
dG
= dx +

96

x
dy = dx + x d(ln y)
y

3.2 Entropy
Is dG
an exact differential? To check this, we have to calculate the second derivatives:



() = 0
and

=
y
x
y
y
2

G
G
1
since y
x 6= x y the infinitesimal quantity is not an exact differential . Thus,
the integral of dG
should be path dependent. Let "i" denote the initial point
(1, 1) and "f " the final point (2, 2). If one then calculates the integral of dG
along
the path i a f passing through the point "a" with coordinates (2, 1) one
obtains:

iaf

dG
=

Zf

dx +

x
2
dy = (2 1) + 2 ln = + 2 ln 2
y
1

(A1)

Fig. 3.5: Integration paths from i f

Let us calculate the same integral along the path i b f :



Z
Z 
2
x
dG
=
dx + dy = (2 1) + 1 ln = + ln 2 (A2)
y
1
ibf

ibf

We can also calculate the integral along a direct path from (1, 1) to (2, 2). To this
1

The cross-bar over letter d indicates that the differential dG


is inexact. In some textbooks
the inexact differentials of heat and work are marked as dQ
and dL
to distinct them from the
exact differentials of internal energy dU and enthalpy dH [3, 5].

97

3 Equilibrium Thermodynamics
end, we introduce a parametric variable t so:
x=t

1t2

y=t
Z

ilinef


Z 
t
dG
=
dt + dt = ( + ) (2 1) = ( + )
t

(A3)

As it can be clearly seen, we have obtained three different answers (A1), (A2)
and (A3), depending on the integration path. Thus, the integrals are different,
the quantity dG
is not an exact differential, and G is not a state function in the
x, y space.
Let us make a new infinitesimal quantity dF by multiplying dG
by
dF =

1
x

so:

dG

= dx + dy.
x
x
y

The new infinitesimal dF is an exact differential since:


 


=0
and
=0
so
y x
x y

2F
2F
=
xy
yx

Integrating dF along any path between "i" (1, 1) and "f " (2, 2) should give us the
same answer. To confirm this, let us calculate the integral along several paths.

Z
Z 

dF =
dx + dy = ln 2 + ln 2 = ( + ) ln 2
(B1)
x
y
iaf
iaf

Z
Z 

dF =
dx + dy = ln 2 + ln 2 = ( + ) ln 2
(B2)
x
y
ibf

ibf

ilinef

dF =

ilinef

dx + dy
x
y

Z2 
1

= ln 2 + ln 2 = ( + ) ln 2

dt + dt
t
t


(B3)

The above integral does not have to be integrated along a straight line path. The
integral can be evaluated along any path. Let us calculate the integral along a
parabola: y = (x 1)2 that goes through points "i" (1, 1) and "f " (2, 2). In order

98

3.2 Entropy
to perform the integration we introduce a parametric variable t so that:
x=t

1t2
2

y = (t 1) + 1;

dy = 2 (t 1) dt

Thus, the integration gives:


Z
Z
dF =
iparabolaf

iparabolaf

Z2 

dx + dy
x
y

dt +
2 (t 1) dt
t
(t 1)2 + 1

= ln 2 +

Z2

2 (t 1)
dt = ln 2 + ln 2 = ( + ) ln 2
(t 1)2 + 1

(B4)
R
The integral if dF has been evaluated along four different paths and all the
answers are identical ((B1), (B2), (B3), (B4)), as one would expect. The quantity
dF is an exact differential and F is a state function in the x, y space. Since we
know that dF is an exact differential, we can easily calculate the function F (x, y).
Knowing that
F (x, y)

=
x
x

and

F (x, y)

=
y
y

then
F (x, y) =

dx + f (y) = ln x + f (y)
x

and
F (x, y)

f (y)

=
[ ln x + f (y)] =
=
y
y
y
y
so
f (y) =

dy = ln y
y

99

3 Equilibrium Thermodynamics
Thus, the function F (x, y) is of the form:
F (x, y) = ln x + ln y + const.
Knowing the function F (x, y) we can check that:
Zf

dF = F (f ) F (i) = F (2, 2) F (1, 1)

It is so since:
F (f )F (i) = ln 2+ ln 2+const.( ln 1+ ln 1+const.) = (+) ln 2
The reader should note that F (f ) F (i) = ( + ) ln 2 is in full agreement with
the results of the integration along several paths (B1, B2, B3, B4).
There is another important observation to be made in conjunction with this example. The infinitesimal differential dG
is not exact. However, when the differential
1
dG
has been multiplied by x we have obtained a new differential dF that is exact. In mathematics the factor x1 is called the integrating factor. The integrating
factor transforms an inexact differential into an exact differential. In the same
way an inexact differential of heat may be transformed into an exact differential
(entropy) using the integrating factor T1 .
End of Example 3.1
Now our task is to derive appropriate mathematical relationships that would
allow for calculating specific entropy as a function of temperature and pressure
of the system s = s(T, p) and as a function of temperature and specific volume
s = s(T, v). We begin with deriving the exact differential for specific entropy as a
function of temperature and pressure s = s(T, p). To this end we introduce into
the definition of the specific entropy (Eq. (3.6)) the first law of thermodynamics
(Eq. (2.61)):
dh v dp
(3.7)
ds =
T
Introducing relationship (2.33) into the above equation we obtain:

!
1
v
v
1 h
dT + v T
dp dp
(3.8)
ds =


T T p
T
T p
T
100

3.2 Entropy
and after simple algebra:

cp
v
dp
dT
ds =
T
T p

(3.9)

Thus,
knowing the equation of state v = v(T, p) one can evaluate the derivative
v
T p and calculate the exact differential of specific entropy as a function of temperature and pressure.
The procedure for deriving the exact differential of entropy as a function of temperature and specific volume is similar. Into the definition of specific entropy
(Eq. (3.6)) we introduce the first law of thermodynamic (Eq. (2.59)) obtaining:
ds =

du + p dv
T

(3.10)

and making use of Eq. (2.21) and after some algebra one obtains:

cv
p
ds =
dv
dT +
T
T v

(3.11)

Eqs. (3.9) and (3.11) are the basis for calculating the entropy of any substance
as a function of temperature and pressure, or temperature and specific volume,
respectively. Similarly to enthalpy and internal energy we can only calculate the
entropy change with respect to a certain reference state s0 = s(T0 , p0 ) since
ZT,p

s(T, p) = s0 +

T0 ,p0

and
s(T, v) = s0 +

ZT,v

T0 ,v0

ZT,p

ds = s0 +

T0 ,p0

ds = s0 +

ZT,v 

T0 ,v0

!

cp
v
dp
dT
T
T p


cv
p
dv
dT +
T
T v

(3.12)

(3.13)

Thus, to calculate entropy we need to know the specific heats and the equation
of state of the substance. The integration can be carried out along any path from
(T0 , p0 ) to (T, p) or (T, v).

101

3 Equilibrium Thermodynamics

3.2.1 Entropy of Liquids and Solids


Liquids and solids are incompressible and therefore

v
=0
and
cp = cv = c
T p

(3.14)

Eq. (3.12) can be then simplified to

s = s0 +

ZT

c
dT
T

(3.15)

T0

and if the specific heat is independent of temperature we obtain:


s = c ln

T
T0

(3.16)

assuming that s0 = 0.

3.2.2 Entropy of Ideal Gases


Invoking again the equation of state for ideal gases:
pv =

RT
M

(3.17)

from which we can easily obtain:



R
v
=
T p p M

(3.18)

and inserting Eq. (3.18) into Eq. (3.12) we obtain:

s = s0 +

ZT,p 

cp
R
dT
dp
T
M p

T0 ,p0

(3.19)

Since cp of ideal gases is independent of pressure the above integral can be written
as:
Zp
ZT
cp
R
dp
dT

(3.20)
s = s0 +
T
M
p
T0

102

p0

3.2 Entropy
If then cp is independent of temperature we obtain:
s = s0 + cp ln

R
p
T

ln
T0 M
p0

(3.21)

In the above equations s stands for specific entropy in J/kg K, T is the temperature in Kelvin, cp is the specific heat at constant pressure in J/kg K. In many
text books on thermodynamics, the reader can find similar expressions for specific
entropy (s) expressed in J/kmol K written as:
s = s0 + Cp ln

T
p
R ln
T0
p0

(3.22)

where Cp is the specific heat at constant pressure in J/kmol K.


Using Eq. (3.13) one may derive similar expressions for calculating specific entropy
as a function of temperature and specific volume:
s = s0 +

ZT

T0

R
v
cv
dT +
ln
T
M
v0

(3.23)

and if the specific heat at constant volume is independent of temperature the


following is applicable:
s = s0 + cv ln

T
R
v
+
ln
T0 M
v0

(3.24)

3.2.3 Entropy of Phase Transition at the Transition Temperature


We consider a system and its surroundings at a temperature at which two phases
are in equilibrium. Such a temperature is called transition temperature and is
denoted by Ttrs . For example, this temperature is 0 C for ice in equilibrium
with water at 1 bar, and 100 C for water in equilibrium with its vapour at 1 bar.
At the transition temperature any transfer of heat between the system and its
surroundings is reversible since the two phases in the system are in equilibrium.
Because our system is at a constant pressure, the change of the specific entropy
of the system is:
trs s =

trs h
Ttrs

or

trs s =

trs h
Ttrs

(3.25)

103

3 Equilibrium Thermodynamics
where trs h and trs h are the enthalpy of the transition in J/kg and J/kmol,
respectively. Standard (at p = 1 bar) transition enthalpies and transition temperatures are listed in the literature; Table 3.1 is an extract from reference [13].
If the phase transition is exothermic (condensation or freezing) then the entropy
change is negative. This reflects the fact that during condensation or freezing
the system is becoming more ordered. For endothermic transitions the entropy
change is positive and the system becomes more disordered.
Table 3.1: Transition temperatures, standard enthalpies and standard entropies
(p=1bar) of selected substances [13].

Substance

H2 O
O2
N2
CO2
CH4
C2 H6
Cl2

Temperature
K
273.15
54.36
63.15
90.68
89.85
172.1

Melting
trs h

trs s

kJ/mol
6.008
0.444
0.719

J/mol K
22.00
8.17
11.39

0.941
2.86
6.41

10.38
31.83
37.25

Vaporization
Tempetrs h
trs s
rature
K
kJ/mol J/mol K
373.15
40.656
108.95
90.18
6.820
75.63
77.35
5.586
72.22
194.6
25.23
129.65
111.7
8.18
73.23
184.6
14.7
79.63
332.4
29.45
88.60

3.2.4 The Third Law of Thermodynamics


At T = 0 K all energy of thermal motion has been ceased and substances are in
solid state. At T = 0 pure substances are arranged in regular, crystal forms in a
perfect order. The third law of thermodynamics states that entropy of a
pure crystalline substance at absolute zero temperature is zero2 :
s(T = 0) = 0

(3.26)

3.2.5 Absolute Entropy of Pure Substances


The third law of thermodynamics specifies zero entropy at absolute zero temperature for pure substances. By measuring specific heat capacities cp at different
2

For detailed discussions on the third law of thermodynamics the reader is requested to consult
textbooks [3, 5, 13].

104

3.2 Entropy
temperatures and evaluating the integrals in Eq. (3.20) the specific entropy (called
also absolute entropy) of any pure substance can be evaluated:

s(T, p) =

ZT

cp
R
dT

T
M

Zp

dp
p

(3.27)

p0

with the second integral vanishing for liquids and solids. The entropy of phase
transition must be added for each phase transition between T = 0 and the temperature of interest T , thus
s(T, p0 = 1 bar) =

ZTm
0

cp
melt h
dT +
+
T
Tm

ZTb

boiling h
cp
dT +
+
T
Tb

Tm

ZT

cp
dT (3.28)
T

Tb

Fig. 3.6 shows a typical dependence of Tp with temperature as well as specific


c
entropy as a function of temperature. The area under the Tp curve is the specific entropy. Near T = 0 K it is difficult to measure cp and in this region an
extrapolation, called Debye extrapolation [3, 5, 13], is used.
The reader may recall that for enthalpies we have chosen a reference state at
T0 = 298.15 K and p = 1 bar. Furthermore we have agreed that h(T0 , p0 ) = 0
and the formation enthalpies of pure O2 , H2 , N2 , pure graphite and pure sulphur
are given zero values (see Table 2.6). Using Eq. (3.28) we can evaluate specific
entropies of theses substances, and any others, at the reference state T0 = 298.15 K
and p0 = 1 bar. Example 3.2 [13] shows a method of obtaining the specific entropy
for nitrogen at the reference state. Table 2.6 lists specific entropies (s0298 )3 at
standard conditions for several pure substances. These entropies are called thirdlaw entropies. In Paragraph 2.2.2.3 we have recommended the Cp -polynomials
(see Eq. (2.46)) obtained using Joint Army-Navy-Air Force (JANAF) tables for
enthalpies (see Table 2.4 and Table 2.5) of gaseous species. The specific entropy
C
can be easily obtained upon integrating Tp , thus
s0T

s0298

ZT

298

Cp dT = Cp,7 R +

ZT

Cp
dT
T

(3.29)

298

where an additional constant Cp,7 R (in kJ/kmol K) is needed to make sure


that at the reference temperature of 298.15 K the integration results in a value
3

In textbooks on chemical thermodynamics the third-law entropies (in J/mol K) are usually
0
denoted as S298
.

105

3 Equilibrium Thermodynamics
corresponding to the third-law entropy.

Boiling

(b)
S

Dboil s

Solid

Liquid

Debye
approximation

Melting

(a)

cp/T

Dmelt s
Gas
S(0)

Tm

Tb

Tm

Tb

Fig. 3.6: Determination of absolute specific entropy of a pure substance [13].


c
Left the variation of Tp with the temperature;
Right the specific entropy

Example 3.2
Evaluate the standard entropy of nitrogen gas at T0 = 298.15 K and p = 1 bar.
The following data is needed:
Debye extrapolation
Integration from 10 K to 35.61 K
Phase transition at 35.61 K
Integration from 35.61 K to 63.14 K
Melting at 63.14 K
Integration from 63.14 K to 77.32 K
Boiling at 77.32 K
Integration from 77.32 K to 298.15 K
Total

s0T0 in J/mol K
1.92
25.25
6.43
23.38
11.42
11.41
72.13
39.20
191.14

Compare the value obtained with the value listed in Table 2.6.
End of Example 3.2

106

3.3 The Second Law of Thermodynamics

3.3 The Second Law of Thermodynamics


3.3.1 The Increase in Entropy Principle
While considering reversible and irreversible processes (Section 3.1) we have already quoted two formulations of the second law of thermodynamics. However,
none of these formulations have proven general enough for determining direction of
the thermodynamic processes. To this end the notion of entropy is needed. Since
we have just learned how to calculate entropy we can reformulate the second law
of thermodynamics as follows:
In a real irreversible process the sum () of the changes of the entropy
of all bodies participating in the process, including the surroundings,
is positive:
>0
(3.30)
For reversible processes:
(3.31)

=0

However it is not possible to carry out a process for which is less than zero.
Thus,
>0

for irreversible processes,

=0

for reversible processes,

<0

for impossible processes,

where stands for the entropy change in the process. The entropy change of the
phenomena consists of the entropy change of the system and the entropy change
of all bodies, including the surroundings, participating in the phenomena, so
= Ssystem +

all X
bodies

Si

(3.32)

i=1

Application of Eq. (3.32) begins with the identification of a system by defining a


control volume encompassing the system, as shown in Fig. 3.7. The reader will
recognise that Eq. (3.32) has been formulated after observing the system for a
time interval. Everything that interacts with the system is called surroundings.
The system can be divided into a number of subsystems, Fig. 3.7, and therefore:
Ssystem =

all sub-systems
X

Sk

(3.33)

k=1

107

3 Equilibrium Thermodynamics
where the entropy change of a sub-system is the difference between the entropy
of the final state and the initial state:
(3.34)

Sk = Sk,f inal Sk,initial

Fig. 3.7: An open system interacting with other bodies by exchanging mass, heat and
work

The next step in calculating the entropy change of the phenomena is to calculate
all P
bodies
Si . We
the second term of the right-hand side of Eq. (3.32), namely
i=1

begin with identifying all the bodies which interact with the system. These bodies
can interact with the system by:
(a) exchanging heat with the system,
(b) exchanging mass with the system,
(c) exchanging work with the system.

Entropy Change of a Heat Source


The entropy change of a heat source that provides heat to the system can be
calculated directly from the definition of entropy (see Eq. (3.4)) as:

SHS =

f inal
Z state

initial state

108

dQ
THS

(3.35)

3.3 The Second Law of Thermodynamics


For a heat source of constant temperature Eq. (3.35) can be simplified to:
SHS =

Q
THS

(3.36)

In the above equations Q stands for the heat that left the heat source and is
provided to the system, as shown in Fig. 3.8.

Fig. 3.8: Heat source providing heat to the system

Entropy Change of a Mass Source


A body that provides or takes mass from the system is called the mass source,
Fig. 3.9. The change of entropy of the mass source is due to entropy of the stream
leaving or entering the mass source. If an infinitesimal amount of mass dm leaves
the mass source and its specific entropy is sM S then the entropy of the mass
source decreases, so
dSM S = sM S dm
(3.37)
or for the whole process:

SM S =

f inal
Z state

sM S dm

(3.38)

initial state

If the specific entropy of the mass stream is constant throughout the process
Eq. (3.38) can be simplified to
SM S = m sM S

(3.39)

109

3 Equilibrium Thermodynamics

Fig. 3.9: Mass source providing mass to the system

Entropy Change of a Body Providing Work to the System


Entropy of a body that provides work to the system does not change during the
process. Consequently the entropy change for such a body is zero.

3.3.2 Entropy Change for a Continuous Process at Steady-State


In a continuous process at a steady-state there is neither internal energy nor
entropy increase in the system (Ssystem = 0) and therefore Eq. (3.32) simplifies
to:
all X
bodies
=
S i
(3.40)
i=1

and further using Eqs. (3.36) and (3.39)

all heat sources


andX
sinks
i=1

Q i
Ti,HS

all mass sources


andX
sinks

m
k sk

(3.41)

k=1

In the above equation stands for the rate of entropy change in kJ/K s, sk is the
specific entropy of the incoming mass streams in kJ/kg K while Q i and m
k are
the heat provided to the system (in kW) and the mass flow rate of the streams
entering the system (in kg/s), respectively. It is important to note that Q i is
positive for a heat source while Q i is negative for a heat sink. Similarly m
k is
positive for a mass source while it is negative for a mass sink.

110

3.3 The Second Law of Thermodynamics

3.3.3 Irreversibility of Processes


The second law of thermodynamics determines the direction of the process by
stating that the process can proceed only if the entropy change of the system
and all the bodies participating in the process is larger (irreversible process) or
equal (reversible process) zero. The second law also deals with quality of energy. It is concerned with the degradation of energy during a process associated
with the entropy generation, the lost opportunities to do work. We have already
learned that maximum work (Lmax ) is obtainable if the system undergoes a
reversible process between two specific states. However, a real irreversible process
proceeding between the same two states produces less work (L < Lmax ) and in
some extreme cases L can be equal zero. The difference between the maximum
available work (Lmax ) that often is called reversible work and the real work (L)
is called the loss of work or irreversibility (I) of the process:
L = Lmax L = I

(3.42)

For work producing devices, irreversibility is positive. For work consuming devices the minimum work (Lmin ) is needed when the process is carried out reversibly, so
L = L Lmin = I
(3.43)
and irreversibility is again positive.
The second law of thermodynamics allows for quantification of the process irreversibility since [3, 5]
L = I = Tsurr
(3.44)
For continuous processes the loss of power is
N = Tsurr

(3.45)

In the above equations and stand for the entropy change (Eq. (3.32)) and
the rate of entropy change (Eq. (3.40)) in the process, respectively.

111

3 Equilibrium Thermodynamics

3.4 General Conditions for Thermodynamic


Equilibrium
In Section 3.1 we have considered spontaneous processes proceeding towards equilibrium. After each of these processes has been completed a state of equilibrium
is established. The question is how to define equilibrium?
Equilibrium is a thermodynamic state from which no spontaneous (natural) processes can proceed.
In other words, a system is in equilibrium if no changes occur within the system
when no interaction with surroundings takes place. An isolated system is in mechanical equilibrium if no changes occur in pressure, in thermal equilibrium
if no changes occur in temperature, in phase equilibrium if no transformation
between phases occurs. The system is in chemical equilibrium if no changes
occur in the chemical composition of the system. In the forthcoming paragraphs
we will derive criteria for systems to be in equilibrium. Firstly, we consider isolated systems which do not interact with surroundings (Section 3.4.1) and later
on, we consider open and closed systems which interact with surroundings by heat
exchange (Section 3.4.2).

3.4.1 Isolated System


Since an isolated system exchanges neither heat nor mass with surroundings, the
entropy change of any spontaneous process occurring in the system is:
= Ssystem = Sf inal

state

Sinitial

state

(3.46)

Thus, any spontaneous process occurring within the system proceeds in the direction of increasing entropy. Consequently, equilibrium is reached when the system
entropy is maximum and it cannot be increased any further. Therefore the criteria
for an isolated system in equilibrium are
S = Smax

(3.47)

dS = 0

(3.48)

or
The just derived criterion of maximum entropy which is applicable to isolated
systems (or simply to an adiabatic closed system) is not very useful since most
of systems encountered in every-day engineering practise are not isolated and

112

3.4 General Conditions for Thermodynamic Equilibrium


exchange not only heat but also mass with surroundings. If we consider our Universe4 as an isolated system we obtain the well known general statement that the
Universes entropy increases continuously resulting in a continuously increasing
disorder or chaos.

3.4.2 Non-Adiabatic System


When a system involves heat transfer with surroundings, the increase in entropy
principle is (see Eq. (3.32)
= Ssystem
or
Ssystem

Qt
0
Tsurr

(3.49)

Qt
Tsurr

(3.50)

where Qt stands for the total amount of heat transferred to the system (see
Fig. 3.8). The later consists of the heat entering due to friction (Qf ) and the
actual heat (Q) supplied from the surroundings to the system, so
Qt = Qf + Q

(3.51)

Thus, any considerations on reaching equilibrium require calculations of Qt . This


makes the use of inequality (3.50) impractical. Since heat is not a state variable,
there exist an infinite number of Q values when the system undergoes a thermodynamic process from an initial state to an equilibrium state. This observation is
highly disappointing since we cannot derive general equilibrium conditions
for a system interacting with the surroundings to reach equilibrium. Instead we have to restrict ourselves to processes for which we can relate Q to some
state variables of the system. In other words, we have to accept some constraints
(restrictions) on the way the process is carried out and consequently the derived
equilibrium conditions are going to be valid under these constraints only. In this
lecture we restrict ourselves to two types of processes; these which proceed under
a constant temperature and a constant pressure, and these that proceed under a
constant temperature and a constant volume5 .
4

Can we regard our Universe as an isolated system? Perhaps our Universe does interact with
other bodies/Universes?
5
In thermodynamics one develops criteria for thermodynamic equilibrium under many other
constraints; under constant entropy (S) and volume (V), under constant entropy (S) and
pressure (p), and many others.

113

3 Equilibrium Thermodynamics
Before we proceed further, we observe that from Eq. (3.4) one obtains
T dS = dQf + dQ

(3.52)

and since the friction heat (dQf ) is always positive, the following emerges
dQ < T dS

(3.53)

Thus, spontaneous (irreversible) processes proceed towards equilibrium so that


the above inequality is always satisfied.
Systems at constant temperature and pressure (T=const., p=const.)
Consider a system of a fixed mass. The system is at temperature T and pressure
p and interacts with surroundings through exchanging heat and work; however
its temperature and pressure remain constant. For any spontaneous processes
inequality (3.53) is applicable and since the system remains at a constant pressure,
from the first law of thermodynamics, we obtain dQ = dH and therefore:
dH < T dS

(3.54)

dH T dS < 0

(3.55)

dH d(T S) < 0

(3.56)

or
and further
since temperature remains also constant during the spontaneous process proceeding towards equilibrium. Inequity (3.56) can be rearranged into a more elegant
form
d(H T S) < 0
(3.57)
that defines a new variable G:
G = H T S

(3.58)

Variable G is called free enthalpy or Gibbs thermodynamic potential. In


some textbooks G is simple called Gibbs function.
Thus, in a system that remains at constant pressure and temperature spontaneous
processes proceed towards equilibrium in such a way that
dG < 0

(3.59)

and finally, when equilibrium is reached, Gibbs thermodynamic potential G is

114

3.4 General Conditions for Thermodynamic Equilibrium


minimum:
dG = 0

or

G = Gmin

(3.60)

Relationships (3.60) formulate equilibrium criteria for a system when its temperature and pressure are specified.
We finish this paragraph with a remark. If the system under considerations contains only one substance, for example a gas or a solid, specifying the system
temperature and pressure determines its volume through the equation of state.
Consequently such a system is always at equilibrium. Relationships (3.60) are
useful in determining equilibrium for a system containing a number (minimum
two) of species in one or several thermodynamic phases. Such systems rearrange
their composition and some or all of the species undergo phase transitions so as
to satisfy the minimum Gibbs potential conditions.
Systems at constant temperature and volume (T=const., V=const.)
Consider a system of a fixed mass, of volume V and temperature T . The system
interacts with surroundings through exchanging heat and work in such a way
that its temperature and volume remain constant. Inequity (3.53) is applicable
for any spontaneous processes occurring within the system. Since the volume of
the system remains constant, from the first law of thermodynamics, we obtain
dQ = dU and therefore:
dU < T dS
(3.61)
and further
d(U T S) < 0

(3.62)

The above inequity defines a new variable F called free internal energy or
Helmholtz free energy. Thus, in a system that remains at constant volume
and temperature, spontaneous processes proceed towards equilibrium in such a
way that
dF < 0
(3.63)
and finally, when equilibrium is reached, Helmholtz free energy F is minimum:
dF = 0

or

F = Fmin

(3.64)

Relationships (3.64) formulate equilibrium criteria for a system when its temperature and volume are specified.

115

3 Equilibrium Thermodynamics

3.5 Equilibrium Between Phases


3.5.1 Single-Component System Consisting of Two Phases
Consider a single component system which consists of two phases which we shall
denote 1 and 2. These might be solid and liquid, liquid and gas, or gas and solid.
We suppose that the system interacts with surroundings (or other bodies) in such
a way that the systems temperature and pressure remain constant at values T
and p, respectively. But the system can exist in either of its two possible phases
or in a mixture of the two phases. Our task is to find the conditions which must
be satisfied so that the two phases can coexist in equilibrium with each other, see
Fig. 3.10.
We begin with recalling considerations of Section 3.4.2 that led to establishing
the general conditions for a system at constant temperature and pressure to be
at thermodynamic equilibrium. These conditions are:
G = H T S = minimum

or

dG = 0

(3.65)

Fig. 3.10: A single component system consisting of two phases maintained at constant
temperature and pressure

If we denote by mi the number of kilograms of phase i present in the system


and by gi (T, p) the Gibbs free enthalpy per kilogram of phase i at temperature T
and pressure p, then the Gibbs free enthalpy (G) for the whole system G can be
calculated as:
G = m1 g1 + m2 g2
(3.66)

116

3.5 Equilibrium Between Phases


The conservation of matter implies that the total number of kilograms of the
substance remain constant, i. e.,
m1 + m2 = m = const.

(3.67)

By inserting (3.67) into (3.66) we obtain


G = m1 g1 + (m m1 ) g2

(3.68)

Relationship (3.68) shows the dependence of G on m1 , which is the only independent parameter. At equilibrium dG = 0 and therefore:
(g1 g2 ) dm1 = 0

(3.69)

Hence the necessary conditions6 for the thermodynamic equilibrium is


g1 = g2

(3.70)

together with T1 = T2 and p1 = p2 . When these conditions are satisfied, the


transfer of a certain amount of substance from one phase to the other
does not change G in Eq. (3.66) and therefore G is then stationary as
required.
For specified T and p, g1 (T, p) is a well-defined function characteristic of properties
of phase 1 while g2 (T, p) is a well-defined function characteristics of properties of
phase 2. If T and p are such that g1 < g2 , then the minimum of G in Eq. (3.66)
is achieved if the substance exists in phase 1 only, so that G = m g1 , Fig. 3.11.
Phase 1 is then the stable phase. If T and p are such that g1 > g2 then the
minimum value of G is achieved when all the substance transforms into phase 2,
so that G = m g2 . Phase 2 is then the stable one. If T and p are such that
g1 = g2 , then Eq. (3.66) is satisfied and any amount m1 of phase 1 can coexist
with the remaining amount m2 = m m1 of phase 2. Thus, the G value remains
unchanged when m1 is varied.
The locus of points where T and p are such that conditions (3.70) are satisfied is
the phase-equilibrium curve, Fig. 3.11. Along such a curve the two phases can
coexist in equilibrium. The phase-equilibrium curve divides the p, T plane into
two regions: one with g1 < g2 , so that phase 1 is the stable phase, and the other
region where g1 > g2 , so that phase 2 is the stable one. What happens if g1 > g2 ?
Obviously the two phases are not in equilibrium. Since the process must proceed
6

One should demonstrate that G is actually minimum. For this the reader is refered to textbooks on thermodynamics.

117

3 Equilibrium Thermodynamics
so as to satisfy the requirement dG < 0 then (g1 g2 ) dm1 0 ((g1 g2 ) dm1 = 0
at the equilibrium only), so dm1 must be negative, which means that some amount
of the substance in phase 1 must undergo transition to phase 2 until g1 = g2 .
We have just made an important observation. The difference in the Gibbs
specific free enthalpy, namely g1 g2 , is the driving force for phase change
of a single substance, just as the temperature difference is the driving force for
heat transfer [15].

Fig. 3.11: Pressure-temperature plot showing the phase-equilibrium curve that defines
stability regions for phases 1 and 2

It is possible to derive a differential equation of the phase-equilibrium curve. We


begin the derivation with recalling that the conditions along the phase-equilibrium
curve are:
g1 (T, p) = g2 (T, p)
(3.71)
or
dg1 (T, p) = dg2 (T, p)

(3.72)

Invoking the definition of the Gibbs free enthalpy:


g = h T s

(3.73)

dg = dh T ds s dT

(3.74)

we obtain its differential as

Using the first law of thermodynamics (see Eq. (2.61)) we can replace dh with

118

3.5 Equilibrium Between Phases


dqt + v dp so,
dg = dqt + v dp T ds s dT

(3.75)

that simplifies further to


dg = v dp s dT

(3.76)

Since conditions (3.72) must be fulfilled at equilibrium we obtain:


v1 dp s1 dT = v2 dp s2 dT

(3.77)

and after simple algebra


(v1 v2 ) dp = (s1 s2 ) dT

(3.78)

s2 s1
s
dp
=
=
dT
v2 v 1
v

(3.79)

The above equation is called Clausius-Clapeyron equation. It relates the slope


of the phase-equilibrium curve to the entropy change s and (specific) volume
change v of the substance that undergoes a change of phase at a point lying on
the phase-equilibrium curve. Eq. (3.79) is written for one kilogram of phase 1 and
s
phase 2, but since the right hand side ( v
) is a ratio, the equation is also valid for
S
extensive properties V . Since the entropy change due to the phase transition is
s =

htrs
Ttrs

(3.80)

where htrs is the latent heat of the phase change (per kg of the substance)
and Ttrs is the temperature of the phase change. Thus,
htrs
dp
=
dT
Ttrs v

(3.81)

3.5.2 Phase Transformations of a Pure Substance


Fig. 3.12 shows a phase-equilibrium diagram for water with phase-curves separating solid from liquid, liquid from gas and solid from gas. The gas phase is
sometimes called the vapour phase. The three phase-equilibrium curves meet a
common point called the triple point. The triple point of water lies at 273.16 K
and 611 Pa and at this unique temperature and pressure arbitrary amounts of all
three phases can coexist in equilibrium with each other (This property makes the
triple point of water suitable as an easy reproducible standard for temperature).

119

3 Equilibrium Thermodynamics

Fig. 3.12: The experimentally determined phase diagram for water. Note the change
of scale at 2 atm [13]

120

3.5 Equilibrium Between Phases


At the critical point the liquid-gas equilibrium curve ends since the volume change
v between liquid and gas approaches zero. Beyond the critical point there is no
further phase transformation since there exists only one "fluid phase" which can
be characterised by the very dense gas that is indistinguishable from the liquid.
Note that at high pressures there exist several ice forms, all solid phases, which
are named ice II, III, V, and VI. These specific ice phases are formed as a result
of modifications between ice molecules as a result of stress.
During melting processes (going from solid to liquid), the entropy of the substance
almost always increases (degree of disorder increases) and the corresponding latent heat is positive indicating that the phase change is endothermic. In most
cases the solid expands upon melting so v > 0. However, water contracts upon
melting so that v < 0 and therefore the slope of the melting curve for water is
negative.

Example 3.3
In this Example we will reproduce the phase-equilibrium curve for water in the
vicinity of the triple point. We will deal with equilibrium between liquid water
and its vapour. Using Clausius-Clapeyron equation we will derive an approximate expression for the pressure of a vapour in equilibrium with the liquid at a
temperature T . We begin with recalling Clausius-Clapeyron equation:
dp
l
=
dT
T v

(A1)

where l is the latent heat of evaporation (2.45 103 kJ/kg) and v = vgas vliquid
is the change in the specific volume (in m3 /kg). Since the gas (vapour) is much
less dense than liquid, vgas vliquid and we may write v
= vgas . We may
assume that the gas can be treated as an ideal gas, so that its equation of state
is simply
RT
pv =
(A2)
M
Inserting (A2) into (A1) leads to
l M dT
dp
=
2
p
R
T

(A3)

The above equation can be easily integrated:


ln p =

lM
+ constant
RT

(A4)

121

3 Equilibrium Thermodynamics
The integration constant can be evaluated knowing the parameters of the triple
point, so
constant = ln ptr +

lM
R Ttr

(A5)

Thus, the expression for the vapour pressure as a function of temperature is:


1
lM
1
p
=

(A6)
ln
ptr
R
T
Ttr
and finally:
p = ptr e

l RM

1
T

T1

tr

(A7)

For ptr = 611 N/m2 , Ttr = 273.16 K, l = 2.45 103 kJ/kg, M = 18 kg/kmol,
R = 8.314 kJ/kmol K we obtain:



1
1
p = 611 exp 5304.3

(A8)
T
273.16
The reader can easily verify the correctness of this formula for low and medium
temperatures using the phase diagram of water shown in Fig. 3.12. For example
for T = 373.16 K the formula yields p = 1.111 105 Pa that is close to the expected
pressure of 1 atmosphere. The formula breaks down for temperatures approaching
the critical point since the ideal gas equation is not longer valid.
End of Example 3.3

3.5.3 Dependence of Gibbs Free Enthalpy on Temperature and


Pressure
Since Gibbs free enthalpy is such an important property it is instructive to examine how it varies with
temperature
and pressure. Thus, we wish to examine


G
G
the dependence of T p and p T with T and p. We begin with writing down the
general form of the differential of G(T, p) as follows


G
G
dT +
dp
dG =
T p
p T

(3.82)

Differentiating Eq. (3.58), which is the definition of Gibbs free enthalpy, we obtain
dG = dH T dS S dT

122

(3.83)

3.5 Equilibrium Between Phases


and introducing the first law of thermodynamics (dH = dQt +V dp) into Eq. (3.83)
we obtain:
(3.84)

dG = S dT + V dp

The above expression states that the change in G is proportional to the change
in dT and dp with (S) and (V ) being the slopes of the plots of G against T and
p, respectively. When this expression is compared to Eq. (3.82) we observe that:

G
= S
T p

and


G
=V
p T

(3.85)

Similar considerations on specific Gibbs free enthalpy (g) leads to the following
relationships

g
= s
T p

and


g
=v
p T

(3.86)

Relationships (3.85) and (3.86) are important since they determine the dependence of Gibbs free enthalpy with temperature and pressure.
Example 3.4
Make graphs showing the dependence of the specific Gibbs enthalpy as a function of temperature and pressure for ice, water and water vapour. Locate the
regions where each of these phases is stable. Assume that the specific heats
of ice and water are equal Cp = 75.3 kJ/kmol K while for water vapour Cp =
33.58 kJ/kmol K. The table below shows the thermodynamic data.
Thermodynamic data of ice, water and water vapour

Phase
Ice
Water
Water vapour

Standard formation enthalpy


kJ/mol
291.83
285.83
241.81
(see also Table 2.6)

Standard entropy
J/mol K
47.98
69.95
188.72

The molar Gibbs potential (in kJ/kmol) of ice as a function of temperature and

123

3 Equilibrium Thermodynamics
pressure can be calculated as follows:
g ice (T, p) =

0
(hf,298 )ice

ZT

T0



T
0
Cp,ice dT T (s298 )ice + Cp,ice ln
T0

(C1)

where (hf,298 )ice = 291.83 103 kJ/kmol and Cp,ice = 75.3 kJ/kmol K and
(s0298 )ice = 47.98 kJ/kmol K. Similarly for the molar Gibbs potential for water
we obtain:
g water (T, p) =

0
(hf,298 )water +

ZT

T0



T
0
Cp,water dT T (s298 )water + Cp,water ln
T0

0
where (hf,298 )water = 285.83 103 kJ/kmol
and (s0298 )ice = 69.95 kJ/kmol K.

and

Cp,water

(C2)
= 75.3 kJ/kmol K

Assuming that water vapour can be treated as an ideal gas one obtains:
g vapour (T, p) =

0
(hf,298 )vapour +

ZT

Cp,vapour dT

T0



p
T
R ln
(s0298 )vapour + Cp,vapour ln
T0
p0
(C3)

where (hf,298 )vapour = 241.81 103 kJ/kmol and Cp,vapour = 33.58 kJ/kmol K,
(s0298 )vapour = 188.72 kJ/kmol K and R = 8.314 kJ/kmol K. In Eqs. (C1),
(C2) and (C3) the reference state parameters are T0 = 298.15 K and p0 =
1 bar (105 N/m2 ).
Fig. 3.13 shows the dependence of the specific Gibbs enthalpies with temperature
for a pressure of 1 bar. It is interesting to observe that the curves for ice and
water cross at the melting temperature of 273 K while the curves for water and
water vapour cross at the boiling temperature of 373 K. The regions where the
Gibbs potential for ice is the lowest among the three potentials corresponds to
conditions under which ice is the only phase stable. Similarly the region for water
being the only stable phase can be easily identified.

124

3.5 Equilibrium Between Phases

Fig. 3.13: The variation of the Gibbs enthalpy with temperature for ice (function (C1)), water (function (C2)) and water vapour (function (C3)) at 1 bar.

Fig. 3.14: The variation of the Gibbs enthalpy with temperature for ice (function (C1)), water (function (C2)) and water vapour (function (C3)) at a
pressure of 611 Pa.

It is instructive to plot functions (C1), (C2) and (C3) for pressure of 611 Pa that

125

3 Equilibrium Thermodynamics
is the triple point pressure. Such a plot is shown in Fig. 3.14. Not surprisingly the
three curves cross at one point corresponding to the presence of the three phases.
The regions where ice and water vapour are the only stable phases can be easily
identified.
Comments:
(a) Specific Gibbs potential for ice (solid) and water (liquid) is independent of
pressure.
(b) Specific Gibbs potential for water vapour (gas) is strongly dependent on pressure.
End of Example 3.4

3.5.4 Equilibrium in Multi-Component Single-Phase Systems


In this paragraph we consider single-phase systems consisting of minimum two
components that do not react. An example of such a system is a process of
mixing of two (or more) gases. Another example concerning liquid phase is a
process of making solutions like water-alcohol mixtures. For a multi-component
system Gibbs free enthalpy is a function of temperature, pressure and the amounts
of different components so
G = G(T, p, m1 , m2 , . . . , mk , . . . , mN )

(3.87)

where mk stands for the amount (in kg) of k-component while G is the Gibbs free
enthalpy (in joules) for the whole system. We can also write Eq. (3.87) as
g = g(T, p, w1 , w2 , . . . , wk , . . . , wN )

(3.88)

where g stands for specific Gibbs enthalpy (in J/kg) while wk is the mass fraction
of k-component (equally well we may use here molar quantities replacing the
specific Gibbs enthalpy g with molar Gibbs enthalpy g, and replacing the mass
fractions wk with the molar fractions xk , respectively). The exact differential of
Gibbs free energy expressed using Eq. (3.87) can be written as:

dG

126


G
dT
T p,mk


G
dp
p T,mk


G
dm1 +
m1 p,T,mk ,k6=1

3.5 Equilibrium Between Phases




G
G
dm2 + . . . +
dmk +
m2 p,T,mk ,k6=2
mk p,T,ml ,l6=k


G
dmN
... +
mN p,T,mk ,k6=N

and using Eq. (3.85)


N
X
G
dG = S dT + V dp +
dmk
mk p,T,mj
= S dT + V dp +

k=1
N
X

k dmk

j 6= k

(3.89)

(3.90)

k=1

Similarly for specific Gibbs energy we obtain:



N
X
g
dg = s dT + v dp +
dck
ck p,T,mj
k=1

= s dT + v dp +

N
X

k dck

j 6= k

(3.91)

j 6= k

(3.92)

k=1

Eqs. (3.90) and (3.91) introduce a new variable k




G
g
k =
=
mk p,T,mj
ck p,T,wj

which is known as the chemical potential. Expressions (3.90) and (3.91) are the
fundamental equations of equilibrium and chemical thermodynamics.
The chemical potential k which plays an important role in chemical thermodynamics is an intensive variable that is in general a function of the state of the
system, as given by p and T , and the wk s. Even though a species is not present
in a system, its chemical potential nevertheless needs not to be zero since there is
always the possibility of introducing it into the system. In this case G and therefore also g will be altered and the value of the corresponding k must be different
from zero. Gibbs potential G is an extensive property as shown by Eq. (3.87).
Consider a system that is -times larger than that described by Eq. (3.87), that
means that amounts of the species have been increased by a factor of ,
G = G(T, p, m1 , m2 , . . . , mN )

(3.93)

127

3 Equilibrium Thermodynamics
We expect that
G(T, p, m1 , . . . , mk , . . . , mN ) = G(T, p, m1 , . . . , mk , . . . , mN ) (3.94)
Differentiating the above equation with respect to (at constant T and p) gives,
N
X
k=1

( mk )
G

= G(T, p, m1 , . . . , mk , . . . , mN )
( mk )
mk

(3.95)

and using Eq. (3.92) we obtain


N
X

k (T, p) mk = G(T, p, m1 , . . . , mk , . . . , mN )

(3.96)

k=1

and after dividing the above equation by the mass of the system:
N
X

k (T, p) wk = g(T, p)

(3.97)

k=1

The above equation allows to calculate specific Gibbs potential (in J/kg) knowing
the chemical potentials (in J/kg) of the components of the system and their mass
fractions wk . Alternatively, one may calculate molar Gibbs energy g (in J/kmol)
using molar chemical potentials k (in J/kmol) and molar fractions xk :
N
X

k (T, p) xk = g(T, p)

(3.98)

k=1

Gibbs-Duhem Equation
The exact differential of Gibbs potential has been already formulated by Eq. (3.90)
and is here copied for your convenience:
dG = S dT + V dp +

N
X

k dmk

(3.99)

k=1

The differential shows how the Gibbs potential changes upon varying the temperature (T ), pressure (p) and amounts of each component (mk ) of the single-phase
considered. Thus, relationship (3.99) is indeed general. Relationship (3.96) which
is to calculate Gibbs potential of the system knowing the chemical potentials of
all the components and their amounts is also general. It can be differentiated

128

3.5 Equilibrium Between Phases


giving
N
N
N
X
X
X
mk dk
k dmk +
k mk ) =
dG = d(
k=1

k=1

k=1

(3.100)

At equilibrium, in both relationships (3.99) and (3.100), dG = 0 and therefore


S dT + V dp +

N
X

k dmk =

N
X

k dmk +

k=1

k=1

N
X

mk dk

(3.101)

k=1

then finally
S dT + V dp

N
X

mk dk = 0

(3.102)

k=1

The above relationship is referred to as Gibbs-Duhem equation for the system


at equilibrium. It states that the intensive properties (T, p, k ) of the system at equilibrium are not independent but they are interrelated. For a
single-phase system consisting of N components, there are N+2 intensive properties and N+1 are independent only. The Gibbs-Duhem equation can be regarded
as the source of the phase rule (see below). Gibbs-Duhem equation shows also
that in a single-phase equilibrium mixture, the chemical potential of a component
of a mixture cannot change independently of the chemical potential of the other
components since
N
X

mk dk = 0

for

p = const and T = const

(3.103)

k=1

For example, in a binary mixture consisting of species A and B, if chemical


potential of species A is increasing, chemical potential of species B must decrease,
since
mB
dA =
dB
(3.104)
mA
Phase Rule for Single-Phase Multi-Component System
For a single-phase system consisting of N (non-reacting) components there exist N+2 independent intensive variables which uniquely describe the thermodynamic state of the system. At thermodynamic equilibrium there are however only
N+1 independent variables since Gibbs-Duhem equation (3.102) must be satisfied.
Thus, the simplest form of the phase rule is given by N IV = N + 1 for a single
phases system of N-components where N IV stands for a Number of Independent
Variables.

129

3 Equilibrium Thermodynamics

3.5.5 Chemical Potential of Pure Substances


Eq. (3.92) defines the chemical potential k . The chemical potential of a pure
substance shows how the Gibbs energy of a system changes as the substance is
added to it since
(m g)
=g
(3.105)
=
m
Thus, chemical potential of a pure substance is simple its specific Gibbs
enthalpy and it is expressed in kJ/kg or kJ/kmol. In the latter case it is called
molar Gibbs energy.
We have already learned how to calculate the specific Gibbs enthalpy (see Eq. (3.65))
(3.106)

g = h T s

and therefore chemical potential of pure substances can be easily calculated since

g(T, p) =

ZT,p

dh T

T0 ,p0

ZT,p

(3.107)

ds

T0 ,p0

Chemical potential for an ideal gas can be readily calculated as follows:


(T, p) = g(T, p) =

h0f,298 +

ZT

cp dT +T

(s0298 +

T0

ZT

T0

cp
R
p
dT
ln ) (3.108)
T
M
p0

which for a constant specific heat simplifies into


(T, p) = g(T, p) = h0f,298 +cp (T T0 )+T (s0298 +cp ln

p
T R
ln ) (3.109)
T0 M
p0

The above relationship is usually expressed as:


(T, p) = 0 (T ) +

p
RT
ln
M
p0

(3.110)

so the term 0 (T ) is a function of the temperature only. The term 0 (T ) is called


the standard chemical potential at 1 bar expressed in kJ/kg (or alternatively
in kJ/kmol).
In order to calculate the chemical potential of a pure liquid we consider the liquid
to be in equilibrium with its vapour. Under such conditions the chemical potential
of the pure liquid must be equal to the chemical potential of its vapour (see

130

3.5 Equilibrium Between Phases


Section 3.5.1) so:
pure

liquid (T, p)

= 0pure

liquid (T )

pvapour
RT
ln
M
p0

(3.111)

Chemical potential of a solid state substance is a function of temperature only


since
solid (T ) = 0solid (T )
(3.112)
It is important to realise that Eqs. (3.108) and (3.110) are applicable to calculating chemical potential of ideal gases. For real gases the true pressure p is
replaced by an effective pressure called the fugacity. For further considerations
the reader is requested to consult textbooks on physical chemistry [13] or chemical
thermodynamics [5, 14, 4].

3.5.6 Significance of Chemical Potential


While considering non-adiabatic systems that interact with surroundings by exchanging heat (Q), we have already observed (see Section 3.4.2) that it is not
possible to derive general conditions for such systems at thermodynamic equilibrium. It is so because for any non-adiabatic system which is in an initial state,
there is an infinite number of ways for the system to reach the (final) equilibrium
state. Each way (process path) is associated with an amount of heat exchanged
with the surroundings and consequently there exist an infinite number of Q-values
each corresponding to a different path. For each path inequality (3.50) must be
satisfied. Realising that we are not able to derive general conditions for thermodynamic equilibrium we have subjected our considerations to some constraints
(restrictions) accepting that the conclusions from such considerations are valid
under these constraints only. We have already considered non-adiabatic systems
of a given (constant) mass interacting with the surroundings so that both the
pressure and the temperature of the system remain constant throughout the process of approaching equilibrium. Under the constraints of constant temperature
and pressure, we have derived the following conditions for the system to reach
equilibrium: for T = const. and p = const.
G Gmin ;

dG = 0

N
X

k dmk = 0

(3.113)

k=1

where



G
g
k =
=
mk p,T,mj
wk p,T,wj

Consider now another type of constraints; equilibrium at constant temperature

131

3 Equilibrium Thermodynamics
and volume. Under such constraints the system reaches equilibrium when F
Fmin or dF = 0 as shown by Eq. (3.64). Helmholtz potential can be written as
(3.114)

F = U T S = H pV T S = G pV
and its differential is then

(3.115)

dF = dG p dV V dp
Introducing (3.99) into (3.115) we obtain
dF = S dT p dV +

N
X

(3.116)

k dmk

k=1

and by writing down the exact differential of F = F (T, V, m1 , . . . , mk , . . . , mN )


as



N
X
F
F
F
dF =
dT +
dV +
dmk
(3.117)
T p,mk
V T,mk
mk T,V,mj ,j6=k
k=1

we can see that at equilibrium obtained at T = const. and V = const. the


following is applicable
M
X

k dmk = 0

k=1



F
f
with k =
=
mk T,V,mj
wk T,v,wk

j 6= k

(3.118)

where f and wk stand for the specific Helmholtz potential (in kJ/kg) and mass
fraction of k-species, respectively. Summarising, under constraints of constant
temperature and volume we have obtained the following equilibrium conditions:
for T = const. and V = const.
F Fmin ;

dF = 0

N
X

(3.119)

k dmk = 0

k=1

where



f
F
=
k =
mk p,T,mj
wk p,T,wj

j 6= k

We can also consider another type of constraints approaching equilibrium under


constant entropy and volume. By considerations similar to these presented above,
it is possible to demonstrate that the equilibrium conditions for S = const. and

132

3.5 Equilibrium Between Phases


V = const. are
U Umin ;

dU = 0

N
X

(3.120)

k dmk = 0

k=1

where



u
U
=
k =
mk p,T,mj
wk p,T,wj

j 6= k

where U and u are the internal energy (in kJ) and specific internal energy of
the system (in kJ/kg), respectively. Now it is time to examine the equilibrium
conditions obtained. Eqs. (3.113), (3.119) and (3.120) have been derived following
three different ways of reaching equilibrium or in other words by considering three7
different constraints. Remarkably, the same relationship
N
X

k dmk = 0

(3.121)

k=1

appears in each of the equilibrium conditions of Eqs. (3.113), (3.119) and (3.120).
Thus, we may expect that when the equilibrium is reached, relationship (3.121)
must be satisfied, regardless of the path the single-phase system follows while
interacting with surroundings. This is a powerful observation; thermodynamic
equilibrium does not depend on the path the system follows from the
initial to the final (equilibrium) state.
It is remarkable that in Eqs. (3.113),(3.119) and (3.120) the same function, the
chemical potential k appears. We have already learned that the chemical potential shows how the Gibbs function changes when the composition changes
(under constant temperature and pressure), as shown by Eq. (3.92). By deriving
Eqs. (3.119) and (3.120) we have learned that the chemical potential also shows
how Helmholtz potential and internal energy of the system change although again
under different set of conditions. This is why the chemical potential is central to equilibrium thermodynamics and, as we will demonstrate in the next
chapter, it is central to equilibrium chemistry or generally to chemistry.

We may consider many other constraints but relationship (3.121) always emerges.

133

3 Equilibrium Thermodynamics

3.6 Multi-Component, Multi-Phase Systems


without Chemical Reactions
In this paragraph we consider multi-component and multi-phase non-reacting systems. Any property of a component will have two indices; one describing the phase
and the other describing the component according to the convention:
 i-concerns phase i=1,2,...,

(3.122)

k-concerns component k=1,2,...,

We begin with systems at constant temperature T and pressure p, so


T (1) = T (2) = T (3) = . . . = T (i) = . . . = T () = T
p

(1)

=p

(2)

=p

(3)

= ... = p

(i)

= ... = p

()

=p

(3.123)
(3.124)

The methodology of developing equilibrium conditions is going to be rather straightforward. We will formulate a Gibbs function for multi-phase, multi-component
systems and we will find the conditions that minimise this function. To find a
minimum of such a function we will need a special mathematical tool that is called
the method of Lagrangean multipliers. The method, explained in Example 3.5, is
used to find a minimum of a function with constraints.
Example 3.5
Find a minimum of a function G(x, y) = x2 + y 2 under a constraint x + y = 1.
It is easy to see that the global minimum of G, without any constraint, is x = 0
and y = 0. Under the constraint x + y = 1, the minimum is located somewhere
else.
First we rewrite the constraint as
1xy =0

(D1)

Now, we build a new function Y (x, y) as follows:


Y (x, y) = G(x, y) + (1 x y)

(D2)

Y (x, y) = x2 + y 2 + (1 x y)

(D3)

where is called a Lagrangean multiplier.


The necessary conditions for function Y (x, y) to have an extremum (either minimum or maximum) under the specified constraint are (The Method of Lagrangean

134

3.6 Multi-Component, Multi-Phase Systems


multipliers):

Y (x, y)
=0
2x = 0
x=
(D4)
x
2
Y (x, y)

=0
2y = 0
y=
(D5)
y
2

1
1
Y (x, y)
= 0 1 x y = 0 1 = 0; = 1 and x = ; y =
(D6)

2
2
2
2
Thus, the extremum of G(x, y) under constrain x + y = 1 is at x = 21 , y = 12 . One
can easily verify that it is a minimum. It is worth noting that function Y (x, y)
is identical with G(x, y), because the constraints are zero. However, the partial
derivatives of Y and G with respect to x and y are different because function Y
incorporates the constraints.
End of Example 3.5
The Gibbs function for a phase of the system is:
Gi = Gi (T, p, mi1 , . . . , mik , . . . , mi )

(3.125)

and for all phases:


G=

Gi (T, p, mi1 , . . . , mik , . . . , mi )

(3.126)

i=1

Thus, we should find a minimum of G with respect to the amounts of each component (mik ). However, the minimum should be found with constraints imposing
the conservation of mass of each of the components. The constraints reflect the
fact that the system, while approaching equilibrium changes mik , however the
total amount of each component must remain conserved and equal to the initial
amount. So the constraints are mass balances of each species so that:

i=1

mi1 = m01
mi2 = m02

(3.127)

i=1

...

135

3 Equilibrium Thermodynamics

mi = m0

(3.128)

i=1

m01 , m02 , m0k , m0

where
indicate the initial amounts of each component. It is easy
to see that the number of constraints equals the number of species (). Now
we are going to use the Gibbs function (Eq. (3.126)), Lagrangean multipliers
together with constraints to build a new function Y (see Example 3.5):

Y =

Gi (T, p, mi1 , . . . , mik , . . . , mi )+

i=1

mi1

i=1

m01

+ 2

mi2

m02

i=1

mik

i=1

m0k

+ ...+

+ ...+

mi m0

i=1

The necessary conditions for function Y to have a minimum are:



Y
=0
satisfied, since T = const.
T p,mi
k

Y
=0
satisfied, since p = const.
p T,mi
k
Y
G1
+ 1 = 0
11 = 1
=
m11 T,p,mi ,k6=1 m11
k


G2
Y
=
+ 1 = 0
21 = 1

2
m1 T,p,mi ,k6=1 m21

(3.129)

(3.130)
(3.131)
(3.132)
(3.133)

..
.


Y
G
=
+ 1 = 0

m1 T,p,mik ,k6=1 m1

1 = 1

(3.134)

So that 11 = 21 = 31 = = 1 meaning the chemical potential of the first


species in all the phases must be the same.

136

3.6 Multi-Component, Multi-Phase Systems


Differentiating further with the respect to mi2 one obtains

G1
Y
+ 2 = 0
12 = 2
=

1
m2 T,p,mi ,k6=2 m12
k

G2
Y
+ 2 = 0
22 = 2
=
m22 T,p,mi ,k6=2 m22

(3.135)
(3.136)

..
.


Y
G
=
+ 2 = 0

m2 T,p,mik ,k6=2 m2

2 = 2

(3.137)

So that 22 = 22 = 32 = = 2 meaning the chemical potential of the second


species in all the phases must be the same.
Finally, differentiation with respect to mi gives

Y
G1
=
+ = 0
1 =
(3.138)
m1 T,p,mi ,k6= m1
k

G2
Y
=
+ = 0
2 =
(3.139)
m2 T,p,mi ,k6= m2
k

..
.


Y
G
=
+ = 0

m T,p,mik ,k6= m

(3.140)

Thus 1 = 2 = 3 = = the chemical potential of the -species in all the


phases must be the same.
The above equations formulate the general conditions for a multi-component,
multi-phase system to be in equilibrium as:
T (1) = T (2) = T (3) = . . . = T (i) = . . . = T () = T
p(1) = p(2) = p(3) = . . . = p(i) = . . . = p() = p
1k = 2k = 3k = . . . = k

for

k = 1, 2, . . . ,

(3.141)

The general conditions for a multi-component multi-phase system to


be at equilibrium are equality of temperatures and pressures of each
phase as well as equality of the chemical potentials in each phase for
each component.

137

3 Equilibrium Thermodynamics

3.6.1 The Phase Rule


We have already demonstrated in Section 3.5.1 that a single-component twophase system may exist in equilibrium at different temperatures (or pressures).
However, once the temperature is fixed, the system will reach equilibrium state
and all intensive (specific) properties of each phase (except their relative amounts)
will be fixed. Thus, a single-component two-phase system has one degree of
freedom (independent variable) which may be either temperature or pressure.
For a single-phase system consisting of N (non-reacting) components there exist
N+1 independent variables according to the Gibbs-Duhem equation (3.102).
Consider now a system consisting of chemically non-reacting components (species) and phases. The number of possible degrees of freedom is two, for T and
p, and for components so:
NIV

= 2 +

(3.142)

Eq. (3.141) specifies ( 1) conditions8 to be fulfilled at equilibrium. Furthermore for each phase Gibbs-Duhem equation must be satisfied taking up
independent variables. Thus, at thermodynamic equilibrium, the number of independent extensive variables (NIV) is given by:
NIV

= 2 + [ ( 1) + ] = + 2

(3.143)

For a single-component ( = 1) two-phase ( = 2) system discussed in Section 3.5.1, one independent intensive property needs to be specified (N IV = 1).
At the triple point however, = 3 and therefore N IV = 0 that means that none
of the properties of a pure substance at the triple point can be varied.

3.7 Thermodynamics of Mixing


Mixing of substances is an essential process in chemical engineering and combustion. In this paragraph we are going to examine what happens when one mixes
gases and later on liquids. Since the mixing process is spontaneous at thermodynamic equilibrium Gibbs potential must reach a minimum.

Since the overall Gibbs potential of the considered system is kept constant, upon specifying
( 1) chemical potential conditions the chemical potential of each species in the last
phase is automatically determined. Thus, here we have ( 1) conditions and not .

138

3.7 Thermodynamics of Mixing


Mixing of Ideal Gases
Consider two containers containing two ideal gases in amounts nA and nB (n is
expressed in kmol). Both containers are at temperature T and pressure p. The
containers are separated by a partition as shown in Fig. 3.15.

Fig. 3.15: Mixing of two ideal gases

Before mixing, the chemical potential of the two gases have their pure values and
the Gibbs enthalpy of the whole system is




p
p
0
0
G = nA A + nB B = nA A + R T ln
+ nB B + R T ln
p0
p0
(3.144)
where nA and nB are the amounts (in kmol) of gas A and B, respectively whilst
A and B are their molar chemical potentials, respectively. After mixing, the
partial pressure of gases are pA and pB , respectively, however the total pressure
remains the same, since
p = pA + pB
(3.145)
During the mixing process the state variables T and p remain constant. Only partial pressures pA and pB change so as to minimise the value of G. Actually, there
is only one independent variable (either pA or pB ) since the relationship (3.145)
holds. Thus, the Gibbs enthalpy of the mixture can be expressed as a function of
pA (or pB ):




pA
p pA
0
0
+ nB B + R T ln
(3.146)
G(T, p, pA ) = nA A + R T ln
p0
p0
At minimum of G
dG(T, p, pA ) = nA R T

dpA
dpA
nB R T
=0
pA
p pA

(3.147)

139

3 Equilibrium Thermodynamics
and therefore

nB
nA
=
pA
p pA

(3.148)

and further
pA =

nA
p = xA p
nA + nB

and

pB = xB p

(3.149)

Relationship (3.149) allows for calculating the partial pressure of A-component,


pA that minimises G function and therefore corresponds to the equilibrium composition. Eq. (3.149) shows that ideal gases mix in all proportions as shown in
Fig. 3.16. The system of two gases, shown in Fig. 3.15, has got two components
( = 2) and one single phase ( = 1). Thus, N IV = + 2 = 2 1 + 2 = 3
and therefore the pressure, the temperature and the mole fraction of one of the
components can be changed without invoking a phase change.

Fig. 3.16: Mixture of two ideal gases A and B

The relationship pi = xi p valid for mixtures of ideal gases has been used by
students for long. Now it has become apparent that this simple relationship
has got the basis in equilibrium thermodynamics. An ideal gas is a model gas
comprised of imaginary molecules of zero volume. The molecules do not interact.
Each chemical species in a mixture of ideal gases has got its own properties which

140

3.7 Thermodynamics of Mixing


are uninfluenced by the presence of other species. Consequently any partial molar
property (other than the volume) of a consistent species in a mixture of ideal gases
is equal to the corresponding molar property of the species as a pure ideal gas
at the mixture temperature but at a pressure equal to its partial pressure in the
mixture. Therefore for an ideal gas for which its internal energy and its enthalpy
are independent of pressure we have
X
X
umixture =
xi ui
and
hmixture =
xi hi
(3.150)

where umixture and hmixture are specific internal energy and enthalpy of mixtures
of ideal gases (in kJ/kmol), respectively.

ui and hi stand for the specific internal energy and enthalpy of pure species
(in kJ/kmol), respectively while xi is the molar fraction of i-species. However
chemical potential of an ideal gas mixture is a function of the partial pressures of
the components since
mixture =

X
i

xi =

xi (0i + R T ln

pi
)
p0

xi 0i + R T

xi ln

xi p
p0

(3.151)

where p stands for the total pressure.


It is rather easy to relate chemical potential of a species in a mixture to its
potential as a pure substance. Such a relationship is derived in Example 3.6 and
it reads
(i )in_mixture = (i )as_pure_substance + R T ln xi
(3.152)
In chemical thermodynamics, Eq. (3.152) is often used as the definition of
an ideal mixture.
Example 3.6
In this example we consider nA moles of an ideal gas (A) filling up a container
at a temperature T and pressure p, as shown in Fig. 3.17 (left). Its chemical
potential is then
(A )pure_substance = (0A )pure_substance + R T ln

p
p0

(E1)

where we use pA = p since only A-component is present. Now, we add into the
container nB moles of another ideal gas (B), as shown in Fig. 3.17 (right) (since we
wish to keep the total pressure p unaltered, the volume of the container increases

141

3 Equilibrium Thermodynamics
accordingly). We formulate two questions (a) how has the chemical potential of
species A been altered upon addition of species B?, (b) can we relate chemical
potential of species A in the mixture (with species B) to the chemical potential
of pure species A?

Fig. 3.17: Pure species-A at temperature T and pressure p; Species-A in a mixture


with species B at Temperature T and pressure p

Chemical potential of species A in the mixture with species B, at temperature T


and total pressure p, is
(A )in_mixture = (0A )pure_substance + R T ln

pA
p0

= (0A )pure_substance + R T ln

xa p
p0

(E2)

Calculating (0A )pure_substance from (E1) and inserting it into (E2) gives
(A )in_mixture = (A )pure_substance + R T ln xA

(E3)

Comments:
(a) Upon addition of species-B, chemical potential of species-A has decreased, as
shown by Eq. (E3).
(b) Relationship (E3) is to calculate chemical potential of a species in a mixture
of ideal gases.
(c) When xA 1 then (A )in_mixture = (A )pure_substance .
End of Example 3.6

142

3.7 Thermodynamics of Mixing


Ideal Mixtures of Liquids
Mixtures of liquids are called solutions. Consider a solution of two liquids A and
(l)
(l)
B so that xA and xB are molar fractions of species A and B in the mixture
(l)
(l)
(superscript (l) indicates the liquid phase). Obviously xA + xB = 1. By analogy
to gas mixtures, the liquid solution is regarded as ideal when
(l)

(3.153)

(l)

(3.154)

(A )in_solution = (A )pure_substance + R T ln xA
and also

(B )in_solution = (B )pure_substance + R T ln xB

Fig. 3.18: Ideal solution of two liquids, A and B, in equilibrium with its vapour

Consider now this ideal solution of liquids A and B in equilibrium with its vapour
at a given temperature T , as shown in Fig. 3.18. We have two components
( = 2) and two phases ( = 2) in the system. All together there are six variables
describing the system; the temperature, the pressure, the mole fractions of species
(1)
(1)
A and B in the liquid phase, xA and xB , and mole fractions of species A and B in
(v)
(v)
the gas (vapour) phase, xA and xB where superscript (v) indicates the vapour
phase. Among these six variables, only two are independent since, following
the phase rule, N IV = + 2 = 2 2 + 2 = 2. Thus, specifying any
two of the six variables determines uniquely the equilibrium. At equilibrium the
(l)
(v)
chemical potentials of A-species in both phases must be equal, so that A = A .
(l)
(v)
Moreover, for B-species B = B is applicable. Let us first exploit the condition
(l)
(l)
(v)
A = A . Using (3.153) we can calculate A as follows
(l)

(1)

(A )in_solution = (A )pure_substance + R T ln xA =
p
(l)
(0A )pure_substance + R T ln A + R T ln xA (3.155)
p0

143

3 Equilibrium Thermodynamics
where relationship (3.110) has been used und pA stands for the partial pressure
of vapour of A-species at equilibrium with pure A-liquid. The chemical potential
of A-species in the vapour phase is simply
(v)

(A )in_vapour_mixture = (0A )pure_substance + R T ln

pA
p0

(3.156)

where pA is the partial pressure of A-species in the vapour. Equating (3.155) and
(3.156), and after some algebra we obtain
(1)

xA pA = pA

(3.157)

By carrying out similar considerations for B-species, we also obtain


(1)

xB pB = pB

(3.158)

Obviously, the total vapour pressure is


(1)

(1)

p = pA + pB = xA pA + xB pB

(3.159)

Relationships (3.157) and (3.158) express Raoults law. The French chemist
Francois Raoult found experimentally that the ratio of the partial vapour pressure
of each component to its vapour pressure as a pure liquid ( ppA
) is approximately
A

(1)

equal to the mole fraction of A-component in the liquid mixture (xA ). Fig. 3.19
(Top) shows the vapour pressures of an ideal binary solution. We wish to stress
again that Fig. 3.19 is applicable only to an ideal solution for which the Raoults
law applies. Many real solutions depart from this ideal behaviour.

The mole fraction of A-component in the vapour phase is easily calculable since
(1)

xA pA

(v)

xA =

(1)
xA pA

+ (1

(1)
xA ) pB

(1)

1+

1xA
(1)

xA

(3.160)

pB

The above relationship is plotted in Fig. 3.19 (Bottom) for several pB ratios. The
A
essence of the above relationship is that the vapour is richer than the
liquid in the more volatile component.

144

3.7 Thermodynamics of Mixing

Vapour pr essur e

p*B
p = pA+ pB
pA*

(1)

pA = xA pA

0.0

0.2

0.4

0.6

Mole fraction of A,

0.8

1.0

(1)
xA

Mole fr action of A in vapou r, xA

(V)

1.0

0.05

0.8

0.1

0.5

0.6

1
2

0.4

10
20

0.2

0.0
0.0

0.2

0.4

0.6

0.8

1.0

(1)

Mole fraction of A in liquid, xA

Fig. 3.19: Ideal solution of two liquids A and B; (Top) Vapour pressures are proportional to the mole fractions of A and B in the liquid phase, (Bottom) the
mole fraction of A in the vapour as a function of its mole fraction in the
p
liquid, the small labels represent the pressure ratio pB
; the vapour is richer
A
than the liquid in the more volatile component.

145

3 Equilibrium Thermodynamics

3.8 Summary
In this chapter several concepts essential to equilibrium thermodynamics have
been recalled. Most of them should be already known to students from lectures
on chemical thermodynamics and physical chemistry. The key to understanding
equilibrium is the second law of thermodynamics and the associated concepts of
irreversible and reversible processes. The law, expressed as the entropy increase
principle, says that in a real irreversible process the sum of the changes of the
entropy of all bodies participating in the process, including the surroundings, is
positive. Using this formulation it is relatively easy to demonstrate that for an
isolated system thermodynamic equilibrium is reached when the system entropy
reaches a maximum value. If we regard our Universe as an isolated system (is our
Universe the only one?) then its entropy is continuously increasing from the big
bang through the present state until finally, in thermodynamic equilibrium (dead
state), a maximum value of entropy will be reached.
Engineers seldom deal with isolated systems. To the contrary. Furnaces, turbines,
engines interact with the surroundings by exchanging matter and heat. Thus, the
above maximum entropy principle applicable to isolated systems is not of a much
help in engineering. Fortunately both experiments and theory teach that thermodynamic equilibrium does not depend on the path the system follows from the
initial to the final (equilibrium) state. Instead, equilibrium is uniquely determined
by the initial state and any two state variables (p and T ; p and V ; V and T ) at
the final equilibrium state. At equilibrium, reached under constant temperature
and pressure, Gibbs potential of the system defined as G = H T S reaches
a minimum value. For a single-phase multi-component system the fundamental
equation of equilibrium reads
0 = S dT + V dp +

N
X

k dmk

k=1

where k is chemical potential of k-component. When the equilibrium is reached


under a constant temperature (dT
PN= 0) and a constant pressure (dp = 0) the
above relationship simplifies to
k=1 k dmk = 0. When the equilibrium is
reached under other conditions e.g. constant P
temperature and volume, or constant
entropy and volume, the same relationship N
k=1 k dmk = 0 is obtained. This
is why the chemical potential which is the specific Gibbs function, is
essential in thermodynamics of equilibrium states.
In this chapter we have proven that the equilibrium conditions for a multi-

146

3.8 Summary
component (k = 1, 2, . . . , k, . . . , ) multi-phase (i = 1, 2, . . . , i, . . . , ) system are

T (1) = T (2) = T (3) = . . . = T (i) = . . . = T () = T


p(1) = p(2) = p(3) = . . . = p(i) = . . . = p() = p
1k = 2k = 3k = . . . = k

for k = 1, 2, . . . ,

Thus, equality of temperatures and pressures of each phase as well as


equality of the chemical potentials in each phase for each component
are required at equilibrium.
At thermodynamic equilibrium, the number of independent extensive variables is
given by:
N IV = + 2
where N IV stands for the Number of Independent Variables9 , is the number
of chemically non-reacting components and is the number of phases present.
The above equation is a well known phase rule for a non-reacting system.
The principle of Gibbs potential minimisation has been used to analyse mixing of
ideal gases and liquids. It has been demonstrated that gases mix in all proportions
and the relationship pi = xi p ( where p and pi stand for the total pressure and
partial pressure, respectively while xi is i-component mole fraction) has got a
sound basis in equilibrium thermodynamics. The text also contains a derivation
of the Raoults law applicable to mixtures of ideal liquids.

also called: Degrees of Freedom

147

3 Equilibrium Thermodynamics

148

4 Chemical Equilibrium
Contents
4.1

Definition of Chemical Equilibrium

4.2

Single Chemical Reaction

4.3

4.4

4.5

4.2.1

Extent of a Single Reaction

4.2.2

Change of Gibbs Enthalpy as a Chemical Reaction Advances

4.2.3

Gibbs Enthalpy of Selected Reactions

4.2.4

Thermodynamic Equilibrium Constant for a Gaseous


Reaction

4.2.5

Other Equilibrium Constants

4.2.6

Effect of Pressure and Temperature on Thermodynamic


Equilibrium Constant

4.2.7

Chemical Equilibrium in Presence of a Solid Phase

4.2.8

Le Chteliers Principle

Multiplicity of Chemical Reactions


4.3.1

Multi-Component, Multi-Phase Systems with Chemical


Reactions

4.3.2

Choice of Chemical Reactions

4.3.3

Exact Number of Chemical Reactions Needed for Equilibrium Determination

4.3.4

Linear Dependence of a Reaction Set

4.3.5

The Phase Rule for a System with Chemical Reactions

Equilibrium Composition
4.4.1

Systems with a one-dimensional reaction basis

4.4.2

Systems with a two-dimensional reaction basis

4.4.3

Systems with a multi-dimensional reaction basis

Summary

149

4 Chemical Equilibrium

4.1 Definition of Chemical Equilibrium


We begin this chapter with recalling the definition of chemical equilibrium which
we have developed in Section 3.4. The system is in chemical equilibrium
if no changes occur in the chemical composition of the system. The
methodology of this chapter is simple; we begin with the simplest case, namely
with a single-phase system in which only one reaction occurs. Later on we extend
our considerations to a single-phase system of multiplicity of reactions. Finally,
we derive equilibrium criteria for a multi-phase, multi-component system with
multiplicity of reactions.
The goal of this chapter is to provide the knowledge required for calculation
of chemical composition at equilibrium. Most of the combustion processes are
carried out to equilibrium and therefore this topic is essential to combustion technology. For some, not too many, combustion processes which are not carried out
to equilibrium, combustion chamber design is based primarily on reaction rates
(see Chapter 6). However, the choice of the operating conditions may still be
influenced by equilibrium considerations. Furthermore, chemical equilibrium is
essential to understanding pollutants formation and destruction mechanisms.
In Chapter 1 we have learned how to carry out stoichiometric combustion calculations. There the assumption was that the fuel is completely combusted to
products containing carbon dioxide and water vapour. The methods of this chapter allow for calculating equilibrium composition of combustion products which
many contain not only carbon dioxide and water vapour but also unburned fuel,
carbon monoxide or even radicals like H, O, OH or N.

4.2 Single Chemical Reaction


4.2.1 Extent of a Single Reaction
We consider a closed system in which a single chemical reaction
1 A1 + 2 A2 + . . . + 1 A1 + A = 0

(4.1)

occurs. The substrates of this reaction are assigned negative stoichiometric coefficients while the products are given positive coefficients (see Section 2.3.1). The
above reaction can also be written as:

150

4.2 Single Chemical Reaction

(4.2)

k Ak = 0

k=1

where stands for the number of chemical species present in the system. If a
species present in the system does not take part in the reaction, its stoichiometric
coefficient is zero. Let n0k be the number of moles of species k in the initial state
of the system. When the reaction proceeds, the variations of the number of moles
of each species are not independent. Instead they change as determined by the
stoichiometric coefficients of reaction (4.1), so
dn1
dn2
dn1
dn
=
= ... =
=
= d
1
2
1

(4.3)

where nk indicates the number of moles of each species and the variable relates
the changes in the amount of the chemical species present. This variable is called
the extent of reaction1 and is expressed in moles. By integrating (4.3) we
obtain:
nk = n0k + k
(4.4)
The above relationship relates the number of moles of each species to the extent of
the reaction. For the initial state of the system we take usually = 0. Since all the
numbers of moles are always not negative, the extent of the reaction varies from
zero to a value that corresponds to the equilibrium state. If the initial number of
moles n0k for the substrates are sufficiently large so that = 1 is an allowed value,
then this value of corresponds to the conversion of a number of moles equal to
the stoichiometric coefficients. Thus, an increase of by 1 mol is equivalent to
the conversion of number of moles of reactants to number of moles of products
corresponding to the stoichiometric coefficients of the reaction.
Summation of (4.4) over all species yields
n=

X
k

nk =

X
k

n0k +

X
k

k =

(4.5)

where n is the overall number of moles in the system,


of moles at the beginning of the reaction. The term
1

n0k +

n0k is the overall number

The extent of reaction has been given other names such as: reaction coordinate, degree of
reaction and progress variable.

151

4 Chemical Equilibrium

(4.6)

represents the algebraic sum of the stoichiometric coefficients and it can


be either negative, positive or zero. The mole fractions xi of the species present
are related to the extent of the reaction through
xk =

n 0 + k
nk
= P k0
n
k nk +

(4.7)

Example 4.1
Consider a system in which the following reaction occurs
C2 H6 + 3 12 O2 2 CO2 + 3 H2 O
Initially in the system there are present 2 mole C2 H6 , 1 mole O2 , 0.5 mole CO2 .
Derive expressions for the mole fraction of each component as functions of the
extent of the reaction.
Assumptions: none
For the above reaction we obtain
for propane

nC2 H6 = n0C2 H6 + C2 H6 = 2 1

for oxygen

nO2 = n0O2 + O2 = 1 3.5

for carbon dioxide

nCO2 = n0CO2 + CO2 = 0.5 + 2

for water vapour

nH2 O = n0H2 O + H2 O = 0 + 3

for all the species

n = 3.5 + 0.5

Thus, the mole fractions are


2
3.5 + 0.5
0.5 + 2
=
3.5 + 0.5

xC2 H6 =
xCO2

1 3.5
3.5 + 0.5
3
=
3.5 + 0.5

xO2 =
xH2 O

Comments:
P
(a) Since = k k = 0.5 > 0 , the number of moles on the right hand side
of the reaction is larger than the number of moles on the left hand side; the
difference is 0.5 mole.

152

4.2 Single Chemical Reaction


(b) The mole fractions cannot be negative. Therefore, considering propane, one
obtains 2 while considering oxygen 0.2857. Consequently, one expects
that the extent of the reaction at equilibrium will take a value somewhere in
the range 0 < eq 0.2857.
(c) Consider now the same chemical reaction with the following initial number
of moles: n0C H = 5 mol, n0O = 10 mol, n0CO = 1 mol and n0H O = 2 mol.
2 6
2
2
2
The number of moles as functions of the extent of the reaction are then:
nC2 H6 = 5 , nO2 = 10 3.5 , nCO2 = 1 + 2 , nH2 O = 2 + 3 . In
this case an increase of by 1 mol is possible. For = 1 mol one obtains
nC2 H6 = 4 mol, nO2 = 6.5 mol, nCO2 = 3 mol, nH2 O = 5 mol and therefore
1 mol of C2 H6 and 3.5 mol of O2 have been converted producing 2 mol of CO2
and 3 mol of H2 O. Note that the four numbers 1, 3.5, 2 and 3 correspond
to the stoichiometric coefficients of the chemical reaction in question. This
example clarifies the statement ". . . an increase of by 1 mol is equivalent to
the conversion of number of moles of reactants to number of moles of products
corresponding to the stoichiometric coefficients of the reaction".
End of Example 4.1

4.2.2 Change of Gibbs Enthalpy as a Chemical Reaction


Advances
In Section 3.5.4 we have considered equilibrium in multi-component single-phase
systems without chemical reactions. However the methodology of deriving equation (3.90) that is the fundamental equation of equilibrium and chemical thermodynamics, remains unaltered when chemical reactions are considered. This is
bacause after expressing Gibbs free enthalpy as a function of temperature, pressure and amounts of species (see Eq. (3.87))
G = G(T, p, m1 , m2 , . . . , mk , . . . , m )

(4.8)

it is of no relevance whether the amounts of species mk change due to chemical


reactions or mixing. Thus, the general conditions for the chemical equilibrium
are the same as for mixing, namely (see Eq. (3.90))
dG = S dT + V dp +

k dmk = 0

(4.9)

k=1

153

4 Chemical Equilibrium
Therefore, at constant temperature and pressure, the differential of Gibbs enthalpy is:

X
k dmk
(4.10)
dG =
k=1

For a chemically reacting system however, the changes in the amounts of the
reacting species occur according to the stoichiometric reaction scheme (4.1). These
changes can be readily expressed as a function of the extent of the reaction since,
using Eq. (4.4), we obtain
mk = (n0k + k ) Mk

(4.11)

dmk = k Mk d

(4.12)

and
Inserting (4.12) into (4.10) we obtain
dG =

Mk k k d = R GpT d

(4.13)

k=1

where
R GpT =

X
k=1

Mk k k =

X
k=1

k k =

g k k

(4.14)

k=1

is the Gibbs reaction enthalpy named also as free enthalpy of reaction.


In the above equation Mk is the molecular mass of species k, k = gk is the
specific Gibbs energy per kilogram of the species. This variable has already been
introduced as chemical potential; k = g k is the chemical potential expressed per
mol (or kmol). Gibbs reaction enthalpy is an intensive variable and similarly to
chemical potentials it depends on the composition of the system. To obtain a
deeper meaning of Gibbs reaction enthalpy, R GpT , we can express relationship
(4.8) using the extent of reaction as follows:
G = G(T, p, m01 , m02 , . . . , m0k , . . . , m0 , )

(4.15)

where m0k is the initial amount of species k and stands for the extent of reaction.
The exact differential of Gibbs enthalpy can then be written as:

G
dG = S dT + V dp +
d
(4.16)
T,p

154

4.2 Single Chemical Reaction


Comparison of Equations (4.16), (4.9) and (4.13) shows that

X
X
G
=
Mk k k =
k k = R GpT

p,T
k=1

(4.17)

k=1

Gibbs enthalpy, G

Thus, Gibbs reaction enthalpy equals to the derivative of Gibbs potential of the whole system with respect to the extent of the reaction. More
explicitly, the Gibbs reaction enthalpy is the change in the Gibbs potential of the
system when the extent of the reaction () changes by 1 mol (or kmol). Consequently, Gibbs reaction enthalpy is expressed in joules (or related units) per one
mol (or kmol) of the extent of the reaction. Using the above relationship a new
important meaning of Gibbs enthalpy of reaction can be formulated as follows
(see Fig. 4.1):

DRG < 0

DRG > 0

DRG=0
Extent of reaction, x
Fig. 4.1: Minimisation of Gibbs enthalpy as the reaction advances. Note how the slope
of the Gibbs enthalpy changes. Equilibirium corresponds to zero slope.

If R GpT < 0, then the reaction takes place spontaneously from left to right since
dG < 0 implies that d > 0;
If R GpT > 0, then the reaction takes place spontaneously from right to left since
dG < 0 implies that d < 0;
If R GpT = 0, then the reaction reached equilibrium.

155

4 Chemical Equilibrium
Thus, the necessary conditions for a chemical reaction to be at equilibrium at constant temperature and pressure are:
R GpT =

R GpT =

or

Mk k k = 0

k k = 0

(4.18)

k=1

k=1

4.2.3 Gibbs Enthalpy of Selected Reactions


Introducing (into Eq. (4.14)) the definition of the specific Gibbs enthalpy g =
h T s we obtain
R GpT =

k (g pT )k =

k=1

k [(hT )k T (spT )k ]

k=1

k (hT )k T

k (spT )k (4.19)

k=1

k=1

and

R GpT = R HTp T R STp

(4.20)

where
R HTp =

k (hT )k

and

R STp =

k (spT )

(4.21)

k=1

k=1

At this point we reiterate that R HTp is the change of the enthalpy of the system
when the extent of reaction () advances by 1 mol. Thus, R HTp is expressed in
joule (or related units) per 1 mol of the extent of the reaction. Similarly R STp is
the change of the entropy of the system when advances by 1 mol and is therefore
expressed in J/K mol or related units.
Obviously, for standard conditions we have
0
0
R G0298 = R H298
298.15 R S298

(4.22)

and accordingly for a formation reaction at standard conditions one obtains


0
0
f G0298 = f H298
298.15 f S298

(4.23)

In Chapter 2 we have used Table 2.6 which lists both enthalpies of formation
and standard (third law) entropies of some selected substances. Now, using re-

156

4.2 Single Chemical Reaction


lationship (4.23), we can calculate Gibbs formation enthalpies for the formation
reactions. These Gibbs enthalpies of the formation reactions f G0298 are listed
in the fifth column of Table 4.1 (see also Example 4.2). For the reference state
species like for example O2 (g) and H2 (g), the Gibbs enthalpy of the formation
reaction is zero. This is rather obvious, since the formation reaction for O2 can
be written as
O2 (g) O2 (g)
(4.24)
and therefore R GpT = 0 for any temperature and pressure. Thus, for all the
reference state species the Gibbs reaction enthalpy is always zero.
Example 4.2
Calculate the Gibbs formation enthalpy of nitrogen oxide (NO) at standard
conditions.
Assumptions: none
We begin with writing the NO formation reaction
1
2 O2

+ 12 N2 NO

or
12 O2 12 N2 + NO = 0

(A1)

Thus,
0
0
f G0298 = f H298
298.15 f S298
=
h
i
0
0
0
0.5 (h298 )O2 0.5 (h298 )N2 + (h298 )NO


298.15 0.5 (s0298 )O2 0.5 (s0298 )N2 + (s0298 )NO =

90.29 kJ/mol 298.15 [0.5 205.15 0.5 191.61 + 210.76] 103 kJ/mol =
86.60kJ/mol

Comments:
(a) Check the obtained value against the one listed in Table 4.1.
(b) Whenever oxygen and nitrogen molecules react so that the extent of the
reaction is 1 mol, the Gibbs enthalpy changes by 86.60 kJ.
End of Example 4.2

157

4 Chemical Equilibrium

Table 4.1: Standard enthalpies of formation, standard entropies and Gibbs formation
enthalpies of some selected compounds; g: gaseous, l: liquid, s: solid

Compound
Oxygen
Hydrogen
Nitrogen
Carbon
Calcium
Iron

Aggregation
State
O2 (g)
H2 (g)
N2 (g)
C(s g raphite)
Ca (crystal,)
Fe (reference)

hf ,298
(kJ/mol)
0
0
0
0
0
0

s0298
(J/molK)
205.15
130.68
191.61
5.74
41.59
27.32

f G0298
(kJ/mol)
0
0
0
0
0
0

logK
0
0
0
0
0
0

Oxygen atoms
O(g)
249.17
161.06
231.74
-40.60
Ozone
O3 (g)
142.67
238.93
163.18
-28.59
Hydrogen atoms
H(g)
218.00
114.72
203.28
-35.61
Water vapour
H2 O(g)
-241.83
188.83
-228.58
40.05
Water
H2 O(l)
-285.83
69.95
-237.14
41.55
Hydroxyl radicals
OH(g)
39.00
183.71
34.28
-6.01
Nitrogen atoms
N(g)
472.68
153.30
455.54
-79.81
Nitrogen monoxide
NO(g)
90.29
210.76
86.60
-15.17
Nitrogen dioxide
NO2 (g)
33.10
240.03
51.26
-8.98
Carbon
C(g)
716.67
158.10
671.24 -117.60
Calcium oxide
CaO(crystal)
-635.09
38.21
-603.50
105.73
Calcium carbonate+ CaCO3 (calcite)
-1206.90
92.90
-1128.80
197.77
Iron oxide
FeO(crystal)
-272.04
60.75
-251.43
44.05
Carbon monoxide
CO(g)
-110.53
197.65
-137.16
24.03
Carbon dioxide
CO2 (g)
-393.52
213.80
-394.40
69.10
Methane
CH4 (g)
-74.87
186.25
-50.77
8.90
Ethane*
C2 H6 (g)
-83.82
229.12
-31.92
5.59
Propane*
C3 H8 (g)
-104.68
270.20
-24.39
4.27
Ethylene*
C2 H4 (g)
52.51
219.20
68.44
-13.09
Benzene*
C6 H6 (g)
82.88
269.30
129.60
-22.71
Methanol*
CH3 OH(g)
-200.94
239.88
-162.32
28.44
Ethanol*
C2 H5 OH(g)
-234.95
280.64
-167.85
29.41
NIST-JANAF, Thermochemical Tables, American Chemical Society and the American Institute of Physics for the National Institute of Standards and Technology,
1998
* Perrys Chemical Enginering Handbook, 7th Edition, McGraw-Hill, 1997
+ P. Atkins, Physical Chemistry, 6th Edition, W.H. Freeman and Company, 1997

158

4.2 Single Chemical Reaction

4.2.4 Thermodynamic Equilibrium Constant for a Gaseous


Reaction
Consider reaction (4.1) taking place in a closed system at constant temperature
and pressure. All the reaction substrates and products are in gaseous phase.
When the system is at equilibrium, the Gibbs potential of the system reaches
minimum: G Gmin . Equivalently Gibbs reaction enthalpy R GpT is zero; it
means Eq. (4.18) must be satisfied. Since all our species are in gaseous form we
can use Eq. (3.110) to calculate their chemical potential:
k (T, pk ) = 0k + RT ln(

pk
)
p0

(4.25)

where k (T, pk ) is the chemical potential of species k expressed in kJ/kmol,


0k (T, pk ) is the standard chemical potential at 1 bar (in kJ/kmol), R is the gas
constant, T is the temperature and p0 is the standard (reference state) pressure of
1 bar. It is important to note that pk is the partial pressure (in bars) of species k
in the gaseous mixture (Relationship (4.25) is valid for ideal gases; for real gases
the partial pressure pk should be replaced with fugacity [16, 17]).
Inserting relationship (4.25) into Eq. (4.18) we obtain:
R GpT =

k k =

k=1

k (0k + RT ln(

k=1

pk
)) = 0
p0

(4.26)

The above relationship can be easily rearranged into:

k RT ln(

X
pk
k 0k
)=
p0

(4.27)

k=1

k=1

and further
RT ln



Y
pk k

k=1

p0

= R G0T

(4.28)

The above relationship is usually written as


R G0T
RT
where K is known as the standard equilibrium constant defined as:
ln K =

(4.29)

159

4 Chemical Equilibrium

K=

Y
pk k
( )
p0

(4.30)

k=1

The standard equilibrium constant is a dimensionless quantity. For a


given reaction, the value of this constant depends on temperature, on the standard
(reference) state pressure (p0 ) and temperature (T0 ). Using relationship (4.29)
the standard equilibrium constant can be calculated using the standard Gibbs
enthalpies listed in thermodynamic tables. The latter ones can also be calculated
knowing the formation enthalpies and standard entropies of species participating
in the reaction, as shown by Eqs. (4.19) and (4.20). The standard equilibrium
constant is often designated as the thermodynamic equilibrium constant. In
chemical engineering literature Eq. (4.30) is known as the law of mass action.
The thermodynamic equilibrium constant K (see Eq. (4.29)) has been derived
directly from the standard Gibbs reaction enthalpy R G0T and therefore is a
function of temperature only. The constant K is related, through the law of mass
action (Eq. (4.30)), to the partial pressures of the species at the equilibrium.
We have already calculated Gibbs formation enthalpies (in kJ/mol) for several
chemical species and these are listed in the fifth column of Table 4.1. Knowing
these enthalpies one can readily calculate the thermodynamic equilibrium constant for the formation reactions at standard conditions using Eq. (4.29). Take
nitrogen oxide (NO) as an example for which the formation reaction is
1
2 N2

+ 12 O2 = NO

(4.31)

with the Gibbs reaction enthalpy of 86.60 kJ/mol. The thermodynamic equilibrium constant for the above reaction at standard conditions is then
ln K =

86.60 103 kJ/kmol


= 34.936
8.314kJ/kmol/K 298.15K

(4.32)

and therefore K = 6.7219 1016 and log K = 15.17 (compare with Table 4.1).
Since the thermodynamic equilibrium constant is very small indeed, the equilibrium of reaction (4.31) at standard temperature is very much to the left so that
the extent of the reaction is indeed very close to zero. The reader is requested to
practice calculations of the thermodynamic equilibrium constant using Table 4.1
and locate the species for which the equilibrium of their formation reactions at
standard temperature is far to the right.
Table 4.2 lists values of the thermodynamic equilibrium constants (K) at different

160

4.2 Single Chemical Reaction


temperatures for some reactions important in combustion. All but the water gas
shift reaction
CO2 + H2 CO + H2 O
are formation reactions for which values of the Gibbs reaction enthalpy are tabulated [12] for various temperatures. Thus, for these reactions, the thermodynamic
equilibrium constants can be readily evaluated using relationship (4.29). However,
for the water gas shift reaction which is not a formation reaction, relationship
(4.20) is used to evaluate R G0T and then using (4.29) the equilibrium constant
is calculated. Below we show such calculations for the reference temperature of
298.15K. Let rewrite the water gas shift reaction as
CO2 H2 + CO + H2 O = 0

(4.33)

The Gibbs enthalpy for the above reaction is

R G0298 = (g 0f,298 )CO2 (g 0f,298 )H2 + (g 0f,298 )CO + (g 0f,298 )H2 O

(4.34)

and using Table 4.1

R G0298 = (394.40) 0 + (137.16) + (228.58) = 28.66kJ/mol

(4.35)

Then
ln K =

R G0298
28.66 103 kJ/kmol
=
= 11.562
R 298.15
8.314kJ/(kmol K) 298.15K

(4.36)

and K = 9.522 106 so that log K = 5.021. Alternatively, one can obtain the
thermodynamic equilibrium constant by manipulating the K-values of appropriate
formation reactions (see Example 4.3).
Example 4.3
Calculate the thermodynamic equilibrium constant K of the homogeneous water
gas shift reaction
CO2 H2 + CO + H2 O = 0
(B0)
using K-values of other reactions listed in Table 4.2.

161

4 Chemical Equilibrium

162

C(s) + O2
CO2

C(s) + 12 O2
CO

H2 + 12 O2
H2 O(g)

CO2 + H2
CO + H2 O(g)

298
300
400
500
600
700
800
900
1000
1100
1200
1300
1400
1500
1600
1700
1800
1900
2000
2100
2200
2300
2400
2500
2600
2700
2800
2900
3000
3200
3400
3600
3800
4000
4500
5000
5500
6000

C(s) + 2H2
CH4

Temperature
in K

Table 4.2: Thermodynamic equilibrium constant K for some reactions important in


combustion [12]

logK
8.894
8.813
5.492
3.420
1.993
0.943
0.138
-0.500
-1.018
-1.447
-1.807
-2.115
-2.379
-2.609
-2.810
-2.989
-3.147
-3.289
-3.416
-3.531
-3.636
-3.732
-3.819
-3.899
-3.973
-4.042
-4.105
-4.164
-4.219
-4.319
-4.407
-4.485
-4.555
-4.619
-4.753
-4.863
-4.955
-5.034

logK
69.095
68.670
51.539
41.299
34.404
29.505
25.829
22.969
20.679
18.805
17.242
15.919
14.784
13.800
12.939
12.128
11.502
10.896
10.351
9.858
9.406
8.998
8.622
8.275
7.955
7.658
7.383
7.126
6.886
6.450
6.064
5.721
5.413
5.134
4.544
4.068
3.675
3.344

logK
24.030
23.911
19.110
16.236
14.320
12.948
11.916
11.109
10.461
9.928
9.481
9.101
8.774
8.488
8.236
8.013
7.813
7.633
7.470
7.322
7.186
7.062
6.946
6.840
6.741
6.648
6.562
6.481
6.404
6.265
6.140
6.027
5.924
5.830
5.627
5.458
5.313
5.188

logK
40.047
39.785
29.238
22.884
18.631
15.582
13.287
11.496
10.060
8.881
7.897
7.063
6.346
5.724
5.179
4.698
4.269
3.885
3.540
3.227
2.942
2.682
2.443
2.223
2.021
1.833
1.658
1.495
1.344
1.068
0.825
0.608
0.414
0.239
-0.131
-0.428
-0.672
-0.877

logK
-5.018
-4.974
-3.191
-2.179
-1.453
-0.975
-0.626
-0.364
-0.158
-0.004
0.136
0.245
0.336
0.412
0.476
0.583
0.580
0.622
0.659
0.691
0.722
0.746
0.767
0.788
0.807
0.823
0.837
0.850
0.862
0.883
0.901
0.914
0.925
0.935
0.952
0.962
0.966
0.967

4.2 Single Chemical Reaction

298
300
400
500
600
700
800
900
1000
1100
1200
1300
1400
1500
1600
1700
1800
1900
2000
2100
2200
2300
2400
2500
2600
2700
2800
2900
3000
3200
3400
3600
3800
4000
4500
5000
5500
6000

logK
-8.980
-8.944
-7.517
-6.672
-6.114
-5.717
-5.420
-5.188
-5.003
-4.851
-4.724
-4.615
-4.522
-4.441
-4.370
-4.307
-4.251
-4.201
-4.155
-4.114
-4.077
-4.042
-4.011
-3.982
-3.955
-3.930
-3.908
-3.886
-3.866
-3.831
-3.799
-3.772
-3.747
-3.726
-3.681
-3.648
-3.622
-3.602

logK
-79.810
-79.298
-58.710
-46.341
-38.084
-32.180
-27.746
-24.294
-21.530
-19.266
-17.377
-15.778
-14.407
-13.217
-12.175
-11.256
-10.437
-9.705
-9.046
-8.448
-7.905
-7.409
-6.954
-6.535
-6.148
-5.790
-5.457
-5.147
-4.857
-4.332
-3.867
-3.455
-3.085
-2.751
-2.046
-1.480
-1.015
-0.625

logK
-40.599
-40.330
-29.469
-22.936
-18.570
-15.446
-13.098
-11.269
-9.803
-8.603
-7.601
-6.752
-6.024
-5.392
-4.839
-4.350
-3.915
-3.525
-3.175
-2.857
-2.568
-2.304
-2.062
-1.839
-1.633
-1.442
-1.265
-1.100
-0.946
-0.666
-0.420
-0.200
-0.004
0.173
0.547
0.846
1.092
1.296

NO

+ 21 O2

logK
-6.005
-5.963
-4.265
-3.246
-2.569
-2.085
-1.724
-1.444
-1.222
-1.041
-0.981
-0.764
-0.656
-0.563
-0.482
-0.410
-0.347
-0.291
-0.240
-0.195
-0.153
-0.116
-0.082
-0.050
-0.021
0.005
0.030
0.053
0.074
0.112
0.145
0.174
0.200
0.223
0.270
0.307
0.335
0.357

1
2 N2

+ 21 H2
1
2 O2

logK
-35.613
-35.378
-25.876
-20.158
-16.335
-13.597
-11.538
-9.932
-8.644
-7.587
-6.705
-5.956
-5.313
-4.754
-4.264
-3.831
-3.446
-3.100
-2.788
-2.506
-2.249
-2.014
-1.798
-1.599
-1.415
-1.245
-1.087
-0.939
-0.801
-0.551
-0.330
-0.133
0.044
0.203
0.539
0.809
1.029
1.213

OH

H
1
2 H2

O
1
2 O2

N
1
2N2

O2
1
2 N2 +
N O2

Temperature
in K

Table 4.3: Thermodynamic equilibrium constant K for some reactions important in


combustion [12]

logK
-15.172
-15.074
-11.143
-8.784
-7.210
-6.086
-5.243
-4.587
-4.063
-3.633
-3.275
-2.972
-2.712
-2.487
-2.290
-2.116
-1.926
-1.824
-1.699
-1.587
-1.484
-1.391
-1.306
-1.227
-1.154
-1.087
-1.025
-0.967
-0.913
-0.815
-0.729
-0.653
-0.585
-0.524
-0.396
-0.295
-0.213
-0.146

163

4 Chemical Equilibrium
Assumptions: none
Using the following three reactions (see Table 4.2)
CO2 + CO2 = 0

(B1)

H2 21 O2 + H2 O = 0

(B2)

C 21 O2 + CO = 0

(B3)

with the thermodynamic equilibrium constants KB1 , KB2 , KB3 , we can compose
reaction (B0) since
(B1) + (B2) + (B3) =
C + O2 CO2 H2 1/2O2 + H2 O C 1/2O2 + CO =
CO2 H2 + H2 O + CO = 0
Adding up Gibbs enthalpies of the above three reactions, we obtain
GB1 + GB2 + GB3 =
[gC gO2 + gCO2 ] + [gH2 1/2gO2 + gH2 O ] + [gC gO2 + gCO ] =
gCO2 gH2 + gH2 O + gCO = GB0
so that
ln KB0 =

GB0
=
RT
(GB1 + GB2 + GB3 )
GB1 GB2 GB3
=
+
+
RT
RT
RT
RT

Therefore
ln KB0 = ln KB1 + ln KB2 + ln KB3
and
ln KB0 = ln
Thus, finally
KB0 =

164

KB2 KB3
KB1

KB2 KB3
KB1

(B4)

4.2 Single Chemical Reaction


and
log KB0 = log KB2 + log KB3 log KB1

(B5)

Comments:
(a) For T = 298K, using Table 4.2 we obtain log K = 40.047 + 24.030 69.095 =
5.018.
(b) The values listed in column six of Table 4.2 have been obtained using the
logarithms of thermodynamic constants listed in columns three, four and five,
using relationship (B5).
(c) Note that manipulating the logarithms of the thermodynamic equilibrium
constants is equivalent to adding or subtracting Gibbs enthalpies.
(d) Derive relationship (B4) using the law of mass actions, Eq. (4.30).
End of Example 4.3

4.2.5 Other Equilibrium Constants


Other forms of the law of mass action are also used. It is possible to define an
equilibrium constant (Kp ) relative to pressures as follows:
Kp =

pkk

(4.37)

k=1

where by comparison with (4.30)


Kp = K (p0 )v

(4.38)

When partial pressures pk are expressed in bars

Kp = K

(numerically)

(4.39)

since p0 = 1 bar. The equilibrium constant Kp is in general not dimensionless


unless is zero.
Other forms of the law of mass action are encountered when different means of
expressing concentrations are used. Using mol (volume) fractions xk and assuming
ideal gas behaviour

165

4 Chemical Equilibrium

K=



Y
pk k

k=1

p0



Y
xk p k

k=1

p0

p
p0

 Y

(xk )

k=1

p
p0

Kx (4.40)

or

Kx = K

p
p0

v

(4.41)

where p is the total pressure of the system (in bars) and Kx is the equilibrium
constants expressed in molar (volume) fractions

Kx =

(xk )k

(4.42)

k=1

We can also use the following derivation

K=



Y
pk k

k=1

p0



Y
xk p k

k=1

p0



Y
nk p k

k=1

n p0

p
n p0

 Y

(nk )

k=1

p
n p0

Kn (4.43)

or

Kn = K

p
n p0

v

(4.44)

where nk is the number of kmol of species k, n is the total number of kmol in the
system and Kn is defined as follows
Kn =

(nk )k

(4.45)

k=1

Very often there is a need to use concentrations ([Ak ] in mol/m3 ) of species. In


this case the following derivation is applicable assuming the perfect gas law:

166

4.2 Single Chemical Reaction

K=



Y
pk k

k=1

p0








Y
[Ak ] RT k
RT
RT Y
k
Kc
([Ak ]) =
=
=
p0
p0
p0
k=1
k=1
(4.46)

or
Kc = K

 p v
0
RT

(4.47)

where Kc (c- stands for concentration) is defined as

Kc =

([Ak ])k

(4.48)

k=1

It should be emphasised that while the thermodynamic equilibrium constant K is


dimensionless, the other constants (Kp , Kn , Kc ) are generally not dimensionless.
Only when = 0 the other constants are also dimensionless. Values of the
equilibrium constants lie in the range from zero to infinity and the larger the
K-value the further to the right is the equilibrium shifted. In other words, the
larger the K-value the more complete the reaction.

4.2.6 Effect of Pressure and Temperature on Thermodynamic


Equilibrium Constant
As shown by Eq. (4.29) the thermodynamic equilibrium constant (K) depends on
the value of Gibbs reaction enthalpy R G0T which is defined at standard pressure
of 1 bar. The value of R G0T and subsequently K are therefore both independent
of the pressure at which the chemical equilibrium is established. Mathematically
we can express this independence as


K
p

=0

(4.49)

The effect of temperature on the thermodynamic equilibrium constant can be


obtained by differentiating Eq. (4.29):

167

4 Chemical Equilibrium

ln K
T

1
=
R T

R G0T
T

1
=
R T

 P

k=1 k

0k (T )

0
R HT0
1 X h
k k2 =
R
T
RT2

(4.50)

k=1

0 is the enthalpy of formation per kmol of species k while R H 0 is the


where h
T
k
reaction enthalpy. The relationship
R HT0
d ln K
=
dT
RT2

(4.51)

is known as the vant Hoff equation that can be also written in another form:
R HT0
d ln K

=

R
d T1

(4.52)

Eq. (4.52) shows that a graphical representation of ln K as a function of 1/T


allows the determination of the standard enthalpy of reaction. If we assume that
in a temperature range from T1 to T2 the standard enthalpy of reaction is constant
then
R HT0
K(T2 )
ln
=

K(T1 )
R

1
1

T1 T2

(4.53)

For significant changes of temperature the variation of the standard enthalpy with
temperature has to be taken into account when integrating Eq. (4.51), so
K(T2 )
=
ln
K(T1 )

ZT2

R HT0
dT
RT 2

(4.54)

T1

where the dependence of R HT0 with temperature is obtained using Kirchoffs


equation (see Section 2.3.5)
T2

R HT02

R HT01

Z
X
k=1 T

168

k Cp,k (T ) dT

(4.55)

4.2 Single Chemical Reaction


where Cp,k (in kJ/kmol/K) is the specific molar heat at constant pressure of
species k.
Relationship (4.52) implies that a plot of ln K versus the reciprocal of absolute
temperature is a straight line. Fig. 4.2, a plot of log K versus 1/T for the reactions listed in Table 4.2, illustrates this near linearity. Thus, Eq. (4.53) provides
reasonable accurate relation for the extrapolation and interpolation of values of
thermodynamic equilibrium constant. If reaction enthalpy R HT0 is negative , i.e.
if the reaction is exothermic, the thermodynamic equilibrium constant decreases
as the temperature increases. Conversely, K increases with T for an endothermic
reaction. The above explained dependence of the thermodynamic equilibrium
constant on temperature can be explicitly seen by rearranging Eq. (4.29) as
K = exp

R HT0
RT

exp

R ST0
R

= K exp

R HT0
RT

(4.56)

Fig. 4.2: Thermodynamic equilibrium constant K as a function of temperature for


reactions listed in Table 4.2. Note: ln K = 2.302585 log K.

Example 4.4
Calculate thermodynamic equilibrium constant for the reaction

C + O2

CO2

169

4 Chemical Equilibrium
at T = 1000K knowing that at 298K, log K = 69.095 (see Table 4.2).
Assumptions: none
Relationship (4.54) is to be used, so that
K(1000)
ln
=
K(298)

1000
Z

R HT0
dT
RT 2

(C1)

298

In order to perform the above integration the dependence of the reaction enthalpy R HT0 with temperature should be known and it can be obtained using Kirchoffs law (see Eq. (4.55)). In order to simplify the calculations we assume that the reaction enthalpy is independent of temperature and is equal to
0
= 393.52 kJ/mol as listed in Table 4.1. Then, after performing the
R H298
integration we obtain (see Eq. (4.53))



0
K(1000) R H298
1
1
ln

=
=
K(298)
R
298 1000


1
393.52 103 kJ/kmol
1

= 111.5007 (C2)
8.314 kJ/kmol/K
298 1000
Then
K(1000)
= 3.76 1049
K(298)

(C3)

log K(1000) log K(298) = 48.424

(C4)

and

Finding in Table 4.2, log K(298) = 69.095

(K(298) = 1.02 1030 ) we obtain

log K(1000) = 48.424 + 69.095 = 20.671 and K(1000) = 4.688 1020

(C5)

Comments:
(a) The obtained value of log K(1000) = 20.671 is very close to the value of 20.679
listed in Table 4.2.

170

4.2 Single Chemical Reaction


(b) In the temperature range from 298K to 1000K, the dependence of the reaction
enthalpy R HT0 with temperature is indeed weak.
(c) At both temperatures the equilibrium is shifted far to the right meaning that
carbon, producing carbon dioxide, consumes almost all available oxygen .
End of Example 4.4

4.2.7 Chemical Equilibrium in Presence of a Solid Phase


In Section 4.2.4 we have derived the law of mass action, Eq. (4.30), for a chemical
reaction taking place in a gas phase. In this paragraph we consider systems when
gas phase and solid phase are present. Examples of such reactions are:
C(s) + O2 (g) CO2 (g)
CaCO3 (s) CaO(s) + CO2 (g)
FeO(s) + CO(g) Fe(s) + CO2 (g)
Before proceeding further we may recall that chemical potential of a gaseous
species is calculated as given by Eq. (3.110). The chemical potential of solids is
calculable as (see Eq. (3.112)):
0

s0298 + Cs ln

s (T, p) = hf,298 + Cs (T T0 ) + T (

T
)=
0s (T )
T0

(4.57)

where Cs is the molar specific heat of the solid, hf,298 is its molar formation
enthalpy, s0298 is its standard molar entropy and T0 = 298.15 K. Since solids
are incompressible, their chemical potential is a function of temperature only, as
shown in Eq. (4.57), and is identical to their standard potential.
Consider now the first of the reactions given above:
C(s) + O2 (g) CO2 (g)
There are two phases, gaseous and solid, present. There is one species present
in the solid phase (C(s)) and two species (O2 (g) and CO2 (g)) present in the
gaseous phase. Recalling our considerations on equilibrium in multi-phase, multicomponent systems (see Section 3.6) we observe that if the chemical potential
of a species present in two phases is not the same, then the system is not at
equilibrium. Spontaneous transport takes place from the phase where the species

171

4 Chemical Equilibrium
has the highest chemical potential towards a phase where its chemical potential
is lower, thus altering the composition of the system until equilibrium is reached.
At equilibrium the chemical potential of each of the species present in the solid
phase is equal to its chemical potential in the gas phase, so

C(g) =
C(s) =
0C(s)

(4.58)

The equilibrium condition in the gaseous phase is dG = 0 or

R G =

3
X

k
k = 0

(4.59)

k=1

and therefore
0 =
C(s)
0O2(g) RT ln

pCO2
pO2
+
0CO2(g) + RT ln
p0
p0

(4.60)

After some algebra we obtain

where



RT ln K = R G0T =
0CO2(g)
0O2(g)
0C(s)

K=

pCO2
p0
pO2
p0

pCO2
= Kp
pO2

(4.61)

(4.62)

It is important to note that in expression (4.61) for calculating the thermodynamic


equilibrium constant K, the standard Gibbs enthalpy stays the same as for a
gaseous system. However, expression (4.62) that expresses the law of action,
includes only the species that are exclusively present in the gas phase.
The above observations can be generalised and if all gases behave as ideal ones,
we have [17]:

RT ln K =

R G0T

k 0k where the summation extends over

k=1

all the species participating in the reaction, (4.63)

172

4.2 Single Chemical Reaction

Kp = K (p0 ) where =

k and the summation extends over

the species exclusively present in the gas phase, (4.64)

Kp =

pkk where the multiplication extends over

the species exclusively present in the gas phase. (4.65)

Example 4.5
Generate a graph showing the dependence of carbon dioxide partial pressure as
a function of temperature for calcination reaction at equilibrium
CaCO3 CaO + CO2

(D1)

Assumptions: calcium carbonate is chemically pure


We begin with calculating the thermodynamic equilibrium constant for reaction
(D1) using the data listed in Table 4.1. The Gibbs enthalpy for the reaction in
question is



CaCO3 + h
CO2 + h
CaO
G0T = HT0 T ST0 = h
T [
sCaCO3 + sCO2 + sCaO ] =

[ (1206.90) + (393.52) + (635.09)] kJ/mol


T [92.90 + 213.80 + 38.21] 103 kJ/mol =
= 178.29 T 159.15 103 in kJ/mol
So, the thermodynamic equilibrium constant is

ln K =

G0T
(178.29 T 159.15 103 ) 103 kJ/kmol
21, 444.55
=
= 19.142
RT
8.314kJ/(kmol K) T K
T

At the equilibrium K =

pCO2
p0

so that

pCO2 = p0 exp(19.142 21, 444.55/T )

(D2)

173

4 Chemical Equilibrium
Since p0 = 1 bar, the partial pressure of carbon dioxide at equilibrium is
pCO2 = exp(19.142 21, 444.55/T )

in bar

(D3)

Fig. 4.3 shows the dependence of the equilibrium partial pressure of carbon dioxide
with temperature.
Comments:
(a) We have calculated that the reaction enthalpy is 178.29 kJ/mol at 298K. We
have assumed that HT0 does not change with temperature. Using Kirchoffs
law (see Eq. (4.55)) one can easily account for the dependence of HT0 with
temperature. However, such a dependence is often weak. For example, for
the calcination reaction, HT0 at 1200K is 167 kJ/mol.
(b) At ambient temperature (300K), the equilibrium partial pressure of CO2 is
1.95 1023 bar which is a very low value. The partial pressure of CO2 in
the atmosphere is much larger than 1.95 1023 bar so CaO when exposed to
the atmosphere spontaneously absorbs CO2 . This is often used in analytical
chemistry to generate a CO2 -free environment.
(c) It is instructive to consider calcination reaction
CaCO3 CaO + CO2
in terms of the phase equilibrium conditions which we have derived in Section 3.6 (see Eq. (3.141)). We have here two phases; the solid phase and the
gas phase, and three components. The first two conditions (Eq. (3.141) (A)
and (B)) are obviously satisfied. The third condition (Eq. (3.141) (C)) stating
that the chemical potential of each component in the gas phase must be equal
to its chemical potential in the solid phase is rewritten below as
(g)

(s)

(D4)

(g)

(s)

(D5)

(g)

(s)

(D6)

CaCO3 = CaCO3
CaO = CaO
CO2 = CO2

Relationships (D4) and (D5) are satisfied since the temperature of the gas
phase is the same as the temperature of the solid phase. The left hand side
of (D6) is the chemical potential of carbon dioxide in the gas phase which is
a function of the temperature and the CO2 partial pressure. The right hand
side of (D6) stands for the CO2 chemical potential in the solid phase and is a

174

4.2 Single Chemical Reaction

Fig. 4.3: Equilibrium partial pressure of carbon dioxide in calcination reaction


CaCO3 CaO + CO2

175

4 Chemical Equilibrium
function of the temperature only. Thus, for a given temperature there exists
only one CO2 partial pressure that satisfies relationship (D6). Therefore in
Fig. 4.3, on the equilibrium partial pressure line, relationship (D6) is quoted.
When, for a given temperature, the CO2 partial pressure is lower than the
equilibrium pressure then CO2 undergoes a spontaneous phase change from
the solid phase to the gas phase since such a phase change minimises the
(g)
(s)
Gibbs potential of the whole system (CO2 < CO2 ). Conversely, when CO2
partial pressure exceeds the equilibrium pressure, CO2 is absorbed in the solid
(g)
(s)
phase since CO2 > CO2 .
End of Example 4.5

4.2.8 Le Chteliers Principle


Consider a single chemical reaction (4.1) that has reached equilibrium. We name
this equilibrium state as an "old" equilibrium. In this paragraph we examine
what happens to the system when, after reaching this old equilibrium, we impose
some changes to the system by altering for example the temperature, the pressure
or we add some inert components. The question is whether the old equilibrium
is retained or the system moves to a "new" equilibrium state. If the latter is
true, the question is in which direction will the equilibrium be shifted; towards
the products or towards the substrates?
In the previous paragraphs we have learned how to calculate the thermodynamic
equilibrium constant K (see Eq. (4.29)) and how to relate its value to the partial
pressures of the species at equilibrium using the law of mass action (see Eq. (4.30)).
Using this knowledge, we will deduct the information about the system response
to the imposed changes. Anticipating the results, it will be soon apparent, that
the system response can be deducted using Le Chteliers principle which
states that
Any system initially in an equilibrium state when subjected to a change
will shift in composition in such a way so as to minimise the change.
Below we analyse the effect of the temperature change, the pressure change, the
volume change and the effect of adding inert species on the equilibrium state.
Effect of temperature
As a matter of fact, in Section 4.2.6 we have already examined the effect of
temperature on the thermodynamic equilibrium constant K. For convenience we
copy here Eq. (4.56)

176

4.2 Single Chemical Reaction

K = exp

R HT0
RT

exp

R ST0
R

= K exp

R HT0
RT

(4.66)

and we reiterate that for endothermic reactions (R HT0 > 0) K increases with
increasing temperature. Conversely, for exothermic reactions (R HT0 < 0) K
decreases with increasing temperature. Fig. 4.2 illustrates the point further. The
question is how these observations agree with the principle of Le Chtelier. In
an exothermic reaction the temperature increases with the extent of the reaction.
Any removal of heat from the system which is equivalent to lowering the system
temperature will result in shifting the system composition (extent of the reaction)
so as to compensate for the heat removal. Thus, the new equilibrium state will
be shifted to the right; e.g. the extent of the reaction increases. By the same
reasoning we can conclude that an endothermic reaction is favoured by a higher
system temperature.
Effect of pressure
In Section 4.2.6 we have concluded that the thermodynamic equilibrium constant
K is independent of pressure (see Eq. (4.49)). This does not mean that the
equilibrium state (composition) is independent of the pressure. Consider then
a system that is in equilibrium at given a temperature and a pressure. Now,
maintaining the temperature, we are increasing the (total) pressure by decreasing
the volume. We recall here relationship (4.40)

K=

p
p0

 Y

k=1

(xk )k =

p
p0

Kx

(4.67)

and bearing in mind that the left hand side (K) is independent of the pressure
we observe
if = 0 the equilibrium is unaffected,
if < 0 when p the equilibrium is shifted to the right
if > 0 when p the equilibrium is shifted to the left.
It is worthy observing that the algebraic sum of the stoichiometric coefficients
() plays a decisive role. In other words the stoichiometry of the reaction is
important. As an example consider a reaction
C(s) + CO2 2 CO

177

4 Chemical Equilibrium
for which = 1 > 0. The total pressure in the system increases as the reaction
progresses since two moles of a gaseous component (CO) are produced per one
mole of a gaseous substrate (CO2 ). Then, an increase in pressure implies a shift
of the equilibrium to the left since the system attempts to reduce the number of
moles and opposes the invoked pressure increase in agreement with the principle
of Le Chtelier. It is also possible to increase the total pressure by adding a nonreacting (inert) species into the system so that the temperature and the volume
remain unaltered. However, by adding an inert component, the partial pressures
of the reacting species do not change and therefore, following the law of mass
action, Eq. (4.30), the equilibrium composition does not change.
Effect of volume
Now we examine what happens to the system which is in equilibrium at given
a temperature and a pressure and the system volume (V ) changes. The volume
change takes place at a constant temperature. An increase in the system volume
results in a decrease of the total pressure; conversely a volume decrease results
in a corresponding total pressure increase. Since we have just established the
effect of the total pressure on the system, the effect of the volume change on the
equilibrium is as follows:
if = 0 the equilibrium is unaffected,
if < 0 when V the equilibrium is shifted to the right
if > 0 when V the equilibrium is shifted to the left.
where V is the total volume. Again the stoichiometry of the system () controls
the observed shifts in the equilibrium state. If > 0 then a decrease in the
volume implies an increase in the pressure and therefore the system shifts the
equilibrium to the left, opposing the pressure increase, according to the principle
of Le Chtelier. Alternatively, one can derive the observed effects of the volume
change directly from Eq. (4.43) that is here copied for your convenience
K=

p
n p0

Kn =

RT
V p0

Kn

(4.68)

Effect of addition of an inert gaseous species


As we have already observed, an addition of an inert gas, so that the temperature
and the volume remain unaltered, does not change the partial pressures of gaseous
species and therefore does not alter the equilibrium state. However, if the addition
is at a constant pressure and a constant temperature then the system volume

178

4.2 Single Chemical Reaction


changes and the equilibrium shifts as described above under the heading "Effect
of volume".
Example 4.6
Consider the following reaction
CO2 + C 2 CO

(E1)

which in combustion- and process-engineering is known as Boudouard reaction.


Determine the partial pressures of gaseous species at equilibrium. Examine the
effect of pressure on the equilibrium.
Assumptions:
(a) CO2 and CO behave as ideal gases,
(b) to examine the effect of pressure on the equilibrium, we intend to calculate
the equilibrium CO2 and CO partial pressures at total pressures of 0.01, 0.1,
1, 5 and 10 bar.
We begin with calculating the thermodynamic equilibrium constant for reaction
(E1) as a function of temperature. To this end, we use Table 4.2 where the
equilibrium constants of the following two reactions are listed
C + 21 O2 CO

(E2)

C + O2 CO2

(E3)

Since 2 (E2) (E3) = (E1) we have


log KE1 = 2 log KE2 log KE3

(E4)

Alternatively, we can use data listed in Table 4.1 to estimate the equilibrium
constant KE1 . The Gibbs reaction enthalpy is then



CO2 h
C + 2 hCO T [
G0T = h
sCO2 sC + 2 sCO ] =

= [(393.52) 0 + 2 (10.53)] T [213.80 5.74 + 2 197.65] 103 =


= 172.46 T 0.175 76 kJ/mol (E5)

Here again, being a bit lazy, we have ignored the effect of temperature on the
reaction enthalpy HT0 (see again Kirchoffs law, Eq. (4.55)). Then

179

4 Chemical Equilibrium

G0T
172.46 + T 0.17576 103 kJ/kmol
ln KE1
=
=
RT
8.314 kJ/(kmol K) T K
20, 743.33
=
+ 21.14 (E6)
T
Thus, relationship (E4) is exact while (E6) is an estimate of KE1 constant. Fig. 4.4
displays both relationships showing that the approximated relationship (E6) is
indeed very accurate.
The law of mass action provides the link between the equilibrium constant and
the equilibrium composition since

KE1 =

2

pCO
p0
pCO2
p0

(E7)

Fig. 4.4: The accurate values (Eq.(E4)) and the estimated values (Eq.(E6)) of the
thermodynamic equilibrium constant for Boudouard reaction CO2 + C
2CO

180

4.2 Single Chemical Reaction


The above relationship is accompanied by
pT = pCO2 + pCO

(E8)

where pT stands for the total pressure. Simple algebra leads to the following
quadratic equation for pCO2
p2CO2 pCO2 [2 pT + KE1 p0 ] + p2T = 0

(E9)

The roots of (E9) can be easily found as

(pCO2 )1 =

(2 pT + KE1 p0 )

q
2 p2
4 pT KE1 p0 + KE1
0

(E10)

q
2 p2
4 pT KE1 p0 + KE1
0

(E11)

and
(pCO2 )2 =

(2 pT + KE1 p0 ) +

Since the second root, given by (E11), is always larger than the total pressure pT
it is rejected and finally at equilibrium we have
q
2 p2
(2 pT + KE1 p0 ) 4 pT KE1 p0 + KE1
0
pCO2 =
(E12)
2
and
pCO = pT pCO2

(E13)

Thus, for a given total pressure pT , we can produce a curve showing the partial
pressures of carbon dioxide and carbon monoxide as a function of temperature.
Fig. 4.5 shows such curves for pressures of 1 and 10 bar. Observe the shift in the
equilibrium with pressure.
The information presented in Fig. 4.5 can be more elegantly shown using a single
curve as depicted in Fig. 4.6. Here a single variable is plotted against temperature
for total pressures of 0.01, 0.1, 1, 5 and 10 bar. Fig. 4.6 is an equilibrium curve for
CO2 + C 2 CO
reaction which can be found in many textbooks on combustion- or chemicalengineering.

181

4 Chemical Equilibrium

Fig. 4.5: Variation of the partial pressures of CO2 and CO with temperature for total
pressure of 1 bar (top) and 10 bar (bottom) for Boudouard reaction CO2 +
C 2CO

182

4.3 Multiplicity of Chemical Reactions

Fig. 4.6: Equilibrium composition of Boudouard reaction CO2 + C 2CO at several


total pressures

Comments:
(a) Fig. 4.6 shows that at a constant total pressure, higher temperatures favour
production of carbon monoxide it means with an increasing temperature the
equilibrium shifts towards CO. This is in agreement with the principle of Le
Chtelier since reaction
CO2 + C 2 CO
is endothermic.
(b) For a fixed temperature, the equilibrium shifts to the left with an increasing
pressure in accordance with the principle of Le Chtelier.
End of Example 4.6

4.3 Multiplicity of Chemical Reactions


In Section 4.2 we have considered a single chemical reaction while in this paragraph we are going to be more general. We consider here a multi-component
system existing in a several phases with a multiplicity of chemical reactions occurring. Our task is to derive the conditions for such a system to be at equilibrium under a constant temperature and a constant pressure. To some extent

183

4 Chemical Equilibrium
this paragraph is going to be similar to Section 3.6 where we have also considered a multi-component and multi-phase system. However, this time the system
may find its equilibrium by both phase changes and chemical reactions occurring
simultaneously.

4.3.1 Multi-Component, Multi-Phase Systems with Chemical


Reactions
Since we are considering here a multi-component and multi-phase system, any
property of a component will have two indices; one describing the phase and other
the component (species) according to the convention introduced in Section 3.6:
 i-concerns phase i=1,2,...,

k-concerns component k=1,2,...,

(4.69)

Thus, there are components (species) (k = 1, 2, . . . , ) and (i = 1, 2, . . . , )


phases in the system. There are R (r = 1, 2, . . . , R) simultaneous chemical reactions occurring and for each of them a stoichiometric equation can be written
as

X
kr Ak = 0
for
r = 1, 2, . . . , R
(4.70)
k=1

where here we use again the convention that the substrates are assigned negative
coefficients while the products are given positive values.
Since the system temperature and pressure are specified and kept constant we
have
T (1) = T (2) = T (3) = . . . = T (i) = . . . = T () = T

(4.71)

p(1) = p(2) = p(3) = . . . = p(i) = . . . = p() = p

(4.72)

The methodology of developing equilibrium conditions for the system is going to


be the same as in Section 3.6. We will formulate a Gibbs function for multiphase, multi-component systems and we will find the conditions that minimise
this function. To find a minimum of such a function we will use the method
of Lagrangean multipliers. We merely recall that the method which you have
learned in Chapter 3 (see Example 3.5) is to find a minimum of a function with
constraints.

184

4.3 Multiplicity of Chemical Reactions


The Gibbs function for a single-phase of the system is:

G =G

(T, p, mi1 , mi2 , ..., mi )

ik mik

(4.73)

k=1

and for all phases:

G=

Gi =

i=1

ik mik

(4.74)

i=1 k=1

Thus, we should find a minimum of G with respect to the amounts of each component (mik ) under constraints. The equations formulating these constraints are going to be based on the following observations; firstly the overall mass of the system
remains conserved2 and secondly the amount of each species changes in accordance
with the stoichiometric equations (4.70). We formulate as many constraints as
there are species in the system. Each constraint equation should describe how the
amount of the species in question varies when the chemical reactions proceed. For
a single reaction we have already developed relationship (4.11) that says precisely
what we want. This relationship has to be reformulated for a system containing
phases and R chemical reactions. Thus,

i=1

mi1

m01

+ M1

1r r

for the first species

2r r

for the second species

(4.75)

r=1

mi2 = m02 + M2

R
X
r=1

i=1

...

R
X

mi = m0 + M

i=1

R
X

r r

for the -species.

r=1

Now we use Gibbs function (4.74), Lagrangean multipliers (1, 2 , ... ) together
2

Nuclear reactions are excluded.

185

4 Chemical Equilibrium
with -constraints (4.75) to build a new function Y3

Y =

ik

mik

+ 1

mi1

m01

M1

+ 2

1r r

r=1

i=1

i=1 k=1

R
X

mi2 m02 M2

R
X

2r r

r=1

i=1

+ ...+

mi m0 M

R
X

r r

r=1

i=1

(4.76)

The exact differential of Y-function at equilibrium reached under a constant temperature and a constant pressure must be zero, so
dY = 0 =

ik mik + 1

mi1 1 M1

+ 2

mi2

1r dr +

r=1

i=1

i=1 k=1

R
X

2 M2

R
X

2r dr + ....

r=1

i=1

mi

R
X

r dr (4.77)

r=1

i=1

By re-arranging the above equation we obtain


0=

ik

mik
R
X
r=1

mi1

+ 2

i=1

i=1 k=1

1 M1

+ 1

1r dr 2 M2

X
i=1

R
X
r=1

mi2

+ ... +

mi +

i=1

2r dr ... M

R
X

r dr

r=1

If you have forgotten how to build the Y-function see Section 3.6 and Example 3.5.

186

(4.78)

4.3 Multiplicity of Chemical Reactions


and further
0=

ik

mik

i=1 k=1

k=1

1 M1

R
X

mik +

i=1

1r dr 2 M2

R
X

2r dr ... M

r dr

r=1

r=1

r=1

R
X

(4.79)

and after some further algebra we obtain

0=

X
i=1 k=1

X
X

kr dr
k Mk
ik + k mik
k=1

(4.80)

r=1

Thus, the conditions for the system to be at equilibrium are

ik = k

for k = 1, 2, ...,

(4.81)

and

0=

R
X
X

ik

Mk kr dr =

k=1 r=1

R X
X

ik

Mk kr dr =

R
X

r GpT dr (4.82)

r=1

r=1 k=1

or more explicitly
0 = 1 GpT d1 + 2 GpT d2 + ... + R GpT

(4.83)

and therefore
r GpT = 0

for r = 1, 2, ..., R

(4.84)

To summarise, we rewrite here the general conditions for a multi-component,


multi-phase system with a multiplicity of chemical reactions to be in equilibrium

187

4 Chemical Equilibrium
as:
T (1) = T (2) = T (3) = ... = T (i) = ... = T () = T
p

(1)

=p

(2)

=p

(3)

= ... = p

1k = 2k = 3k = ... = k

(k)

= ... = p

()

=p

for k = 1, 2, . . .

and



X
X
G
p

k kr = 0
k Mk kr =
= r G T =
r
k=1

(4.85)
(4.86)
(4.87)
(4.88)
(4.89)

k=1

The general conditions for a multi-component, multi-phase system with


chemical reactions taking place to be at equilibrium are equality of
temperatures and pressures of each phase as well as equality of the
chemical potentials in each phase for each component. Furthermore,
Gibbs enthalpies of all the chemical reactions are zero.
The just derived conditions should be compared with Eq. (3.141) which we have
derived for a non-reacting system. The conclusion is obvious. The same conditions
are needed for phase equilibrium. Additionally Gibbs enthalpies of all chemical
reactions must be zero.

4.3.2 Choice of Chemical Reactions


In Section 4.2 we have considered single chemical reactions. More precisely we
have considered systems that could be described by a single chemical reaction.
While considering an equilibrium in a system containing CO2 , CO and C it has
been "obvious" to us that Boudouard reaction
CO2 + C 2 CO
which contains all the species in question should be used. However, when a large
number of species are present it is not longer obvious what and how many
reactions should be considered.
In Section 4.3.1 we have considered a multiplicity of chemical reactions
(r = 1, 2, . . . , R) proceeding simultaneously. We have just demonstrated that at
equilibrium R-relationships (4.89) hold. We can easily come up with other reactions which can be generated by a linear combination of two or more reactions
already considered by the set of the stoichiometric equations (4.70). Since the
equilibrium condition (4.89) is satisfied for all R reactions, it is also satisfied for

188

4.3 Multiplicity of Chemical Reactions


a linear combination of any of these R reactions. However, such a linear combination does not provide any new relation among the chemical potentials of
the system species. Therefore, it is necessary to determine the exact number of
chemical reactions needed for the determination of the equilibrium state of the
system.

4.3.3 Exact Number of Chemical Reactions Needed for


Equilibrium Determination
We consider a system of chemical species (A1 , A2 , . . . , A ) which are known to
be the only species present. We wish to determine both the number of reactions
(R) as well as the reactions themselves. Any chemical reaction which occurs in
the system can be expressed by a stoichiometric equation of the following form
1 A1 + 2 A2 + ... + k Ak + ... + v A = 0

(4.90)

In the above equation, the stoichiometric coefficients k are unknown and there are
such coefficients to be determined. Each species Ak is build up out of (atoms)
elements like H, C, O and others. Let us assume that there are elements in the
system. For each element we can write down a balance equation so that -balance
equations can be formulated. They can be written as a element balance
matrix with -rows and -columns. If we denote the rank of this matrix as then
the number of chemical reactions to consider is R = where is the number
of species and is the rank of the element balance matrix. Sounds complicated?
Do not worry; it is simple. The following example clarifies the procedure.
The following chemical species H2 , CO, CO2 , CH3 OH, H2 O are the components of
methanol production. Thus, we have identified five ( = 5) chemical species and
we wish to determine the equilibrium state for a given temperature and pressure.
To this end we would like to use the knowledge which we have developed in the
previous paragraphs. At this stage we have no clue as to what and how many
chemical reactions shall we consider. However, we realise that each of the reactions
must be of the following form
1 H2 + 2 CO + 3 CO2 + 4 CH3 OH + 5 H2 O = 0

(4.91)

The five species H2 , CO, CO2 , CH3 OH, H2 O are build out of three ( = 3)
elements H,C and O. For each element we can formulate a balance of atoms as

189

4 Chemical Equilibrium
follows
H

2 1 + 0 2 + 0 3 + 4 4 + 2 5 = 0

(4.92a)

0 1 + 1 2 + 2 3 + 1 4 + 1 5 = 0

(4.92b)

0 1 + 1 2 + 1 3 + 1 4 + 0 5 = 0

(4.92c)

The above set of equations can be written in a matrix notation as


A ~ = ~0
or

2 0 0 4 2

0 1 2 1 1

0 1 1 1 0

(4.93)
1
2
3
4
5

= 0

(4.94)

The important number which is now emerging is the rank of matrix A. There
are several methods for determination of the rank of a matrix and here you should
consult textbooks on linear algebra. Perhaps the most common definition of a
rank of a matrix is the following: "the rank counts the number of independent
rows in the matrix ". Since the rank of the matrix A is = 3 (see below), the
number of chemical reactions is R = = 5 3 = 2. Thus, for the considered
system, consisting of the five species, we have to select two chemical reactions.
The next question is obvious: what reactions?
To answer the question we have to solve the set of three linear equations (4.94)
of five unknowns. To this end we recommend Gauss elimination method (see
Appendix A). The method is constantly used to solve large sets of linear equations.
Say, we have to solve a set of M equations. The idea of Gauss is simple; multiples
of the first equation are subtracted from the other equations, so as to remove the
first unknown from those equations. It leaves a smaller system of M 1 equations.
The process is repeated until there is only one equation and one unknown which
can be solved explicitly. Then, we proceed backwards and find other unknowns
(see Appendix A). Now we will apply Gauss elimination to equation set (4.94).
We begin with a matrix (see Eq. (4.94))

2 0 0 4 2 0
0 1 2 1 1 0
0 1 1 1 0 0

190

(4.95)

4.3 Multiplicity of Chemical Reactions


In order to remove the second number (1) from the third row, we subtract the
second row from the third row to obtain (the first and the second rows remain
unaltered)

2 0 0 4 2 0
0 1 2 1 1 0
(4.96)
0 0 1 0 1 0
In our case, the Gauss elimination happens to be completed by one operation only.
We mention in passing that at the end of the Gauss elimination we have obtained
a matrix which have three non-zero rows4 . Now we begin the process of backward
substitution. Since we have five unknowns 1 , 2 , 3 , 4 , 5 and three equations,
two out of the five unknowns can be freely selected while the remaining three
must be obtained by solving matrix (4.96). To take a full benefit of matrix (4.96),
we select the pair (4 , 5 ) and calculate the remaining three unknowns (1 , 2 , 3 ).
Indeed, we have a full freedom in selecting the pair (4 , 5 ) with one limitation;
we should not take 4 = 0 and 5 = 0 since it would remove both the fourth
species (CH3 OH) and the fifth species (H2 O) from the system. Thus, for any
pair (4 , 5 ) we calculate the triple (1 , 2 , 3 ) and by doing so we determine the
chemical reaction (4.90). Since we need to determine two reactions, we select two
pairs (4 , 5 ). For a free selection of the pair (4 , 5 ) we chose 4 = 1 and 5 = 0.
By examining the third row of matrix (4.96) we obtain
3 5 = 0

so

3 = 0

while examining the second row brings


2 + 2 3 + 4 + 5 = 0

so

2 = 1

so

1 = 2

and finally inspecting the first row provides


2 1 + 4 4 + 2 5 = 0

Thus, the first reaction, determined by the five stoichiometric coefficients taking
values of 1 = 2, 2 = 1, 3 = 0, 4 = 1, 5 = 0, can be written as
2H2 CO + CH3 OH = 0
4

Another definition of the rank of a matrix is "the rank equals the number of non-zero rows
in the final matrix of Gauss elimination process".

191

4 Chemical Equilibrium
or
2H2 + CO = CH3 OH
If we make another selection of the pair (4 , 5 ) so that 4 = 0 and 5 = 1, we
obtain the second set of the stoichiometric coefficients 1 = 1, 2 = 1, 3 =
1, 4 = 0, 5 = 1. This set determines the second chemical reaction
H2 + CO CO2 + H2 O = 0

(4.97)

or
H2 + CO2 = CO + H2 O
We have just determined the first set of two reactions required for considering chemical equilibrium in a system containing the five species H2 , CO, CO2 ,
CH3 OH, H2 O . These are
2H2 + CO = CH3 OH

(4.98a)

H2 + CO2 = CO + H2 O

(4.98b)

The reader can verify that for (4 , 5 )-pairs taking values of (1,0) and (1,1), we
obtain another set of the two reactions
2H2 + CO = CH3 OH

(4.99a)

3H2 + CO2 = CH3 OH + H2 O

(4.99b)

Again, for (0,1) and (1,1) we obtain


H2 + CO2 = CO + H2 O

(4.100a)

3H2 + CO2 = CH3 OH + H2 O

(4.100b)

Indeed, there exist infinitely many pairs of reactions since there are infinitely many
combinations of numbers that can be used for 4 and 5 . It is important to realise
that any plausible chemical reaction that may occur among the five species H2 ,
CO, CO2 , CH3 OH, H2 O can be expressed by a linear combination of set (4.98),
or set (4.99), or set (4.100). In this lecture course each of these sets is called a
reaction basis. In terms of linear algebra, the above sets of two reactions span
the space of all plausible reactions in the five species system. Thus, every reaction
in this reactions space is a linear combination of a reaction basis. Here we stress
again that the two reactions which form a reaction basis are independent. This

192

4.3 Multiplicity of Chemical Reactions


is an essential feature of a reaction basis.
There exist also some other methods for reaction basis determination however they
are more cumbersome. The above described procedure is logical, consistent and
simple. After gaining some experience, you will be able to "guess" an appropriate
reaction basis without following the above procedure. However, you must make
sure that the reactions you have chosen for your reaction basis are independent.
The next paragraph deals with the issue of checking the reaction independence.

4.3.4 Linear Dependence of a Reaction Set


For the sake of arguments, let us assume that somebody, who considers the five
species system H2 , CO, CO2 , CH3 OH, H2 O of methanol production, has come
up with an idea that the following four chemical reactions have to be taken into
account for the determination of the system equilibrium
2H2 + CO = CH3 OH

R1

(4.101a)

H2 + CO2 = CO + H2 O

R2

(4.101b)

H2 + 2CO + H2 O = CH3 OH + CO2

R3

(4.101c)

3H2 + 3CO + H2 O = 2CH3 OH + CO2

R4

(4.101d)

We wish to find out if the above four reactions are independent. We begin with
rewriting the stoichiometric equations using convention (4.70), so that
2H2 CO
+CH3 OH
=
H2 +CO CO2
+H2 O =
H2 2CO +CO2 +CH3 OH H2 O =
3H2 3CO +CO2 +2CH3 OH H2 O =

0
0
0
0

(4.102)

Examining whether the above four reactions are linearly independent means
checking whether there exist four numbers 1 ,2 , 3 ,4 , not all of them equal
to zero, so that
1 (2H2 CO + CH3 OH) + 2 (H2 + CO CO2 + H2 O)+
+ 3 (H2 2CO + CO2 + CH3 OH H2 O)+
+ 4 (3H2 3CO + CO2 + 2CH3 OH H2 O) = 0 (4.103)

193

4 Chemical Equilibrium
or in a matrix form

2
1

0 1 +

0
or more compact

1
1
1
0
1

2 +

1
2
1
1
1

3 +

3
3
1
2
1

4 =

B ~ = ~0

0
0
0
0
0

(4.104)

(4.105)

where B matrix is
~1
R

B =

~2
R

~3
R

~4
R

2 1 1 3
1 1 2 3
0 1 1
1
1
0
1
2
0
1 1 1

(4.106)

Note that the number of columns of B matrix is equal to the number of reactions
so each reaction is represented by a column vector. The number of rows equals
the number of species so that the first row corresponds to H2 , the second one to
CO and so on. If ~ = ~0 is the only solution of equation (4.105) the reactions
are independent. Let us look for the solution of the system (4.105) using Gauss
elimination. The elimination procedure is described in the following steps:

194

4.3 Multiplicity of Chemical Reactions

2 1 1 3
1 1 2 3
0 1 1
1
1
0
1
2
0
1 1 1

0
0
0
0
0

2 1 1 3 0
0 1 1
1 0
0
1 1 1 0
1 1 2 3 0
1
0
1
2 0

2 1
1
3 0
0
1
1
1
0
0
1
1
1 0
0
1.5 1.5 1.5 0
0 0.5 0.5
0.5 0

2 1 1 3 0
0 1 1
1 0

0
0
0
0 0

0
0
0
0 0
0
0
0
0 0

exchanging rows

row(4)-0.5row(1)

and row(5)+0.5row(1)

row(3)+row(2);

row(4)+1.5row(2)

and row(5)-0.5row(2)
(4.107)

As it can be easily seen from (4.107), the rank of matrix B is two. It means
that only two of the four reactions considered are independent. If we wish, we
can determine the relationships between the four reactions. Matrix (4.107) is
equivalent to
2 1 2 3 3 4 = 0

(4.108)

2 + 3 + 4 = 0

(4.109)

We select 3 = 1 and 4 = 0 to obtain 2 = 1 and 1 = 1. Thus,

1
1

for 3 = 1 and 4 = 0, we have ~ =


1 so
0
~2 + 1 R
~3 = 0
~1 + 1 R
1 R

(4.110)

195

4 Chemical Equilibrium
and therefore

~3 = R
~1 R
~2
R

(4.111)

The correctness of (4.111) can be easily verified recalling (4.102) so that

(2H2 CO + CH3 OH) (H2 + CO CO2 + H2 O) =

(4.112)

H2 2CO + CO2 + CH3 OH H2 O

(4.113)

If we wish to establish the relationship between R4 -reaction and R1 and R2 we


need to find another solution to Eqs. (4.108) and (4.109). If we select 3 = 0 and
4 = 1 we obtain 2 = 1 and 1 = 2 so for

2
1
~
~
~

3 = 0 and 4 = 1 we have ~ =
0 so 2 R1 + 1 R2 + 1 R4 = 0 (4.114)
1

and therefore
~4 = 2 R
~1 R
~2
R

(4.115)

Again the above relationship is easily verifiable since

2 (2H2 CO + CH3 OH) (H2 + CO CO2 + H2 O) =


3H2 3CO + CO2 + 2CH3 OH H2 O

4.3.5 The Phase Rule for a System with Chemical Reactions


In Chapter 3 we have derived the phase rule for non-reacting systems of phases
and chemical species as (see Eq. (3.143))
N IV = + 2

(4.116)

The rule must be modified for systems in which chemical reactions occur. As we
have discussed in Section 3.6.1 the number of possible degrees of freedom is 2 +

196

4.3 Multiplicity of Chemical Reactions


. From this number, ( 1) + combinations must be subtracted already
for a non-reacting system (see Section 3.6.1). However, for each independent
chemical reaction, Eqs. (4.89) provide R additional relations that must be satisfied
at equilibrium. Thus, the phase rule for reacting systems is
N IV = + 2 R

(4.117)

where R stands for the number of independent reactions or in other words R is


the dimension of a reaction basis spanned over the chemical species considered.
While deriving Eq. (4.117) we have considered the phase-equilibrium equations
and the chemical-reaction-equilibrium equations which interrelate the variables
appearing in the phase rule. However, in certain circumstances, special additional constraints must be taken into account. If the number of these additional
constraints is denoted by S then the most general form of the phase rule is
N IV = + 2 R S

(4.118)

It is not straight forward to realise when these additional (special) constraints,


marked by S, must be applied. This is for example the case when one of the
reactions is the decomposition of a pure species resulting in more than one gaseous
product. Then, a special relation exists between the partial pressures of the
gaseous products of the decomposition and this relation formulates an additional
constraint. Example 4.7 clarifies further the issue.

Often a new variable = R S is introduced so that the phase rule remains


as

N IV = + 2

(4.119)

and the number is known as the number of independent species.


Example 4.7
We revisit all Examples of this chapter to demonstrate usefulness of the phase
rule. We determine the number of independent variables (N IV ) for the following
systems:
(a) a system consisting of the gases C2 H6 , O2 , CO2 , H2 O in chemical equilibrium
(Example 4.1),
(b) a system consisting of the gases O2 , N2 , NO in chemical equilibrium
(Example 4.2),

197

4 Chemical Equilibrium
(c) a system consisting of the gases CO2 , H2 , CO, H2 O in chemical equilibrium
(Example 4.3),
(d) a system consisting of C(s), O2 and CO2 in chemical equilibrium
(Example 4.4),
(e) a system consisting of CaCO3 (s), CaO(s) and CO2 in chemical equilibrium
(Example 4.5),
(f) a system consisting of C(s), CO2 and CO in chemical equilibrium
(Example 4.6).
Solution (a). The system contains four ( = 4) C2 H6 , O2 , CO2 , H2 O species, all
in a single gas phase ( = 1). The species are built up out of three elements
C, H, O so ( = 3). The rank of A matrix (Eq. (4.93), Paragraph 4.3.4) is
three so that = 3 . Thus, the number of independent chemical reactions is
R = = 4 3 = 1 and one of them can be written as
C2 H6 + 3 21 O2 2 CO2 + 3 H2 O
There are no special constraints, so S = 0. Thus,
N IV = + 2 R S = 4 1 + 2 1 0 = 4
This result means that one is free to specify four phase-rule variables for example
temperature, pressure and two mass (mole) fractions in an equilibrium mixture
of these four chemical species.
Solution (b). The system contains three ( = 3) O2 , N2 , NO species, all in a single
gas phase ( = 1). There are two ( = 2) elements (O and N) in the system and
A matrix is of the rank two , = 2. Since R = = 3 2 = 1 there is only one
chemical reaction (R = 1) that forms the reaction basis. If we take the reaction
1
2 O2

+ 12 N2 = NO

as the reaction basis, we see that if in its initial state, the system was made up
of pure NO, the stoichiometry of the reaction implies that the partial pressure of
oxygen equals the partial pressure of nitrogen. This relation between the partial
pressures of oxygen and nitrogen provides an additional special constraint (S = 1).
Thus,
N IV = + 2 R S = 3 1 + 2 1 1 = 2

198

4.3 Multiplicity of Chemical Reactions


This result means that one is free to specify two phase-rule variables.
Solution (c). N IV = 4 (see Solution (a))
Solution (d). The system contains three ( = 3) C(s), O2 and CO2 species, in
two ( = 2) phases. There are two ( = 2) elements (C and O) and the rank of
A-matrix is two, = 2. Thus, there is only one chemical reaction R = =
3 2 = 1 that forms the reaction basis. We may take the reaction
C + O2 CO2
as the reaction basis. Thus,
N IV = + 2 R S = 3 2 + 2 1 = 2
This result means that one is free to specify two phase-rule variables.
Solution (e). The system contains three ( = 3) CaCO3 (s), CaO(s) and CO2
species, in two solid phases (CaCO3 and CaO) and one gas phase (CO2 ) so that
= 3. There are three ( = 2) elements (Ca, O, C) and the rank of matrix A is
two, = 25 . Therefore R = = 3 2 = 1 and only one reaction forms the
reaction basis. Typically one takes
CaCO3 CaO + CO2
as the reaction basis. There are no specific constraints, so S = 0. Thus,
N IV = + 2 R S = 3 3 + 2 1 = 1
This result means that one is free to specify only one phase-rule variable in an
equilibrium mixture of these three species. We have just demonstrated why calcium carbonate exerts a fixed decomposition pressure at a given temperature, as
shown in Fig. 4.3.
Solution (f). The system contains three ( = 3) C(s), CO2 , CO species, in two
phases ( = 2). There are two ( = 2) elements (C, O) and the rank of matrix A
is two, = 2. Therefore R = = 3 2 = 1 and only one reaction forms the
reaction basis. We take
5

Although the number of elements is three ( = 3), the rank of the element balance matrix A
is two, = 2.

199

4 Chemical Equilibrium

C + CO2 2 CO
as the reaction basis. There are no specific constraints, S = 0. Thus,
N IV = + 2 R S = 3 2 + 2 1 = 2
This result means that one is free to specify two phase-rule variables in an equilibrium mixture of these three species, as shown in Fig. 4.6.
End of Example 4.7

4.4 Equilibrium Composition


In the previous paragraphs of this chapter, we have acquired comprehensive knowledge on chemical equilibrium. In this paragraph we use this knowledge to calculate equilibrium composition of reacting mixtures. We reiterate here that the
equilibrium composition depends on the temperature and pressure at equilibrium
as well as on the initial concentration of the chemical species of the reacting
mixture.
Any analysis of chemical equilibrium begins with identification of the species
which are present in the system. What follows is a determination of the initial
concentrations (amounts, mass or mole fractions, partial pressures ets.) of these
species together with a specification of the temperature and total pressure of the
system. At this stage it is useful to identify the elements (atoms) present. From
this point onwards, there are several methods of calculating the equilibrium composition. We are going to examine most of these methods using examples. We
begin with a system whose reaction basis can be described by a single chemical reaction, it means the reaction basis is one-dimensional. What follows is an example
of a system whose reaction basis is two-dimensional. Finally, we demonstrate how
to determine the equilibrium composition without considering chemical reactions
at all.

4.4.1 Systems with a one-dimensional reaction basis


The mixture of 3 kmol of CO2 and 1 kmol of H2 is heated up to 1300 K at a
pressure of 1 bar and we expect that at equilibrium in addition to CO2 and H2
also H2 O and CO are present. Thus, there are four species ( = 4) CO2 , H2 , CO,

200

4.4 Equilibrium Composition


H2 O in the system. The task is to determine the composition of the equilibrium
mixture.
There are three elements (H, C, O) in the system ( = 3). Following the procedure underlined in Section 4.3.3 we can easily check that the rank of the element balance matrix is three ( = 3) and therefore there is only one reaction
R = = 4 3 = 1 that forms the reaction basis. We chose the following reaction
CO2 + H2 = CO + H2 O

(4.120)

as a reaction basis, so that CO2 = 1, H2 = 1, CO = 1, H2 O = 1 and


= 1 + 1 1 1 = 0. In this way, we have identified the species present and
we have selected an appropriate reaction basis. Furthermore, the initial amounts
of the species have been given as n0CO2 = 3 kmol, n0H2 = 1 kmol, n0CO = 0 kmol,
n0H2O = 0 kmol while T = 1300 K and p = 1 bar. Thus, the preliminary work has
been done and we step into calculating the equilibrium composition.
Method 1 - Based on the extent of the reaction and the equilibrium
constant
We express the concentrations of all four species participating in the water gas
shift reaction (4.120) as a function of the extent of the reaction:
nCO2 () = n0CO2
nH2 () = n0H2
nCO () = n0CO +
nH2 O () = n0H2 O +

(4.121)

and the total number of kmol of the mixture is


nT () = nCO2 ()+nH2 ()+nCO ()+nH2 O () = n0CO2 +n0H2 +n0CO +n0H2 O (4.122)
Since n0H2 = 1 kmol, the extent of the reaction cannot be larger than 1. Thus,
we expect that eq , expressed in kmol, is in the range 0 < eq 1.
From Table 4.2 we find that K = 1.75792361 at T = 1300 K (alternatively we
could calculate K using Eq. (4.29)). Since in this example we have expressed the
concentrations of reacting species in kmol, it is convenient to relate K to Kn . To
this end we use relationships (4.43) to obtain:

201

4 Chemical Equilibrium

K=





H2 O
|CO |
nCO (eq ) nH O (eq )
2




H2
CO
|
2|
nCO (eq ) nH (eq )
2

p
nT (eq ) p0

(4.123)

Substituting (4.121) and (4.122) into (4.123) provides


K=

eq eq
(3 eq ) (1 eq )

(4.124)

By solving Eq. (4.124) for , with K = 1.75792361, we obtain eq = 0.8230 kmol


(since Eq. (4.124) is quadratic there exists another solution, eq = 8.455 kmol,
that is however not applicable to our problem). Thus, the equilibrium composition of the reacting mixture is:
neq
CO = 3 0.8230 = 2.177 kmol
2

(4.125)

neq
H2 = 1 0.8230 = 0.1770 kmol
neq
CO = 0 + 0.8230 = 0.8230 kmol
neq
H2 O = 0 + 0.8230 = 0.8230 kmol
If needed, the mole fractions at the equilibrium can be easily calculated

2.177
= 0.54425
4
0.8230
= 0.20575
=
4

xeq
CO2 =
xeq
CO

0.1770
= 0.04425
4
0.8230
=
= 0.20575
4

xeq
H2 =
xeq
H2O

(4.126)
(4.127)

It is important to note that the equilibrium composition for the water gas shift
reaction is independent of the total pressure since v = 0.
Method 2 - Based on the balance of elements and the equilibrium constant
The stoichiometric equation (4.120) is obviously valid. However, we can write the
actual reaction that reaches equilibrium as:
3 CO2 + 1 H2 = xCO2 + yH2 + zCO + wH2 O

202

(4.128)

4.4 Equilibrium Composition


where xCO2 +yH2 indicates the leftover (un-reacted) reactants whilst zCO+wH2 O
represents the reaction products, as shown in Fig. 4.7. The task is to determine
x, y, z, and w that correspond to equilibrium at T = 1300K and p = 1bar.
There are three elements C, H and O in the system and for each of them we can
formulate a balance equation as follows:

C balance

3=x+z

or

z =3x

(4.129a)

O balance

6 = 2x + z + w

or

w =3x

(4.129b)

H balance

2 = 2y + 2w

or

y =x2

(4.129c)

Before reaction

After reaction

Initial amounts

At equilibrium
x kmol CO2
y kmol H2

3 kmol CO2
1 kmol H2

z kmol CO
w kmol H2O

} leftover reactants
}leftover products

Fig. 4.7: Illustration of reaction (4.128)

The total number of kmol is then


(4.130)

nT = x + y + z + w
Using again Eq. (4.123) we obtain
|

K=

CO
nCO
(eq ) nHH2O
(eq )
2O

CO2
nCO2
(eq ) nHH2
(eq )
2

p
nT (eq ) p0

zw

=
xy

p
(x + y + z + w) p0

0

(4.131)

Eqs. (4.129a), (4.129b), (4.129c) and (4.131) formulate a system of four equations
with four unknowns (x, y, z, w). Inserting (4.129a), (4.129b) and (4.129c) into

203

4 Chemical Equilibrium
(4.131) we obtain:
1.75792361 =

(3 x) (3 x)
x (x 2)

(4.132)

There are two solutions to the above equation; x = 5.454568 and x = 2.17699.
Obviously only the second solution is applicable to our problem. Thus, at the
equilibrium:
neq
CO2 = x = 2.177 kmol

(4.133)

neq
H2 = y = 0.1770 kmol
neq
CO = z = 0.8230 kmol
neq
H2 O = w = 0.8230 kmol
The above solution is identical to the equilibrium composition given by (4.125).
Method 3 - Minimization of Gibbs enthalpy of the system using the
extent of the reaction
Consider, for the third time, the mixture of three kmol of carbon dioxide and
one kmol of hydrogen that is heated up to 1300K. Our task is to determine the
equilibrium composition of the reacting mixture by minimising Gibbs enthalpy of
the reacting mixture consisting of CO2 , H2 , CO and H2 O. We assume that the
gases behave as ideal ones. We begin with finding the Gibbs standard enthalpies
at T = 1300K for the species considered. Table 4.4 lists the values taken from
the JANAF tables [12]6 .
Table 4.4: Standard Gibbs enthalpies at T = 1300 K [12]

Species
CO2
H2
CO
H2 O(g)

f G01300
kJ/mol
-396.177
0
-226.509
-175.774

We pause here to check whether the data listed in Table 4.4 are consistent with
the previously used value of the equilibrium constant K = 1.75792361. Using
Table 4.4 we can calculate the Gibbs energy for reaction (4.120) as follows:
6

In Table 4.1 Gibbs enthalpies at 298K are listed only. To find the values at T = 1300K one
needs either the whole JANAF tables or one calculates R G01300 using R G0298 .

204

4.4 Equilibrium Composition

R G01300 = 226.509 175.774 + (1) (396.177) + (1) 0 = 6.106 kJ/mol


(4.134)
and therefore

K(1300 K) = exp(

6.106 103 kJ/kmol


R G01300
) = exp(
) = 1.757926
RT
8.314kJ/(kmol K) 1300K
(4.135)

which is close enough to the value of K = 1.75792361 previously used. As in


Method 1, we express the number of kmol of the reacting species as a function of
the extent of the reaction using relationships (4.121) which we will copy here for
the sake of completeness:
nCO2 () = n0CO2
nH2 () = n0H2
nCO () = n0CO +
nH2 O () = n0H2 O +

(4.136)

and the total number of kmol of the mixture is


nT () = nCO2 ()+nH2 ()+nCO ()+nH2 O () = n0CO2 +n0H2 +n0CO +n0H2 O (4.137)
where n0CO2 = 3 kmol, n0H2 = 1 kmol, n0CO = 0 kmol , n0H2 O = 0 kmol and
n0T = 4 kmol.
The Gibbs enthalpy of the system can be readily calculated as follows:
G() = GCO2 () + GH2 () + GCO () + GH2 O ()

(4.138)

and the above equation is the heart of this method (compare with Eq. (4.15)).
The four terms appearing in the above equations are calculated using Eq. (3.110)
applicable to ideal gases



pCO2
GCO2 () = (3 ) 10 396.177 + RT ln
=
p0



(3 ) p
3
(3 ) 10 396.177 + RT ln
(4.139)
4 p0
3

205

4 Chemical Equilibrium




pH2
GH2 () = (1 ) 10 0 + RT ln
=
p0



(1 ) p
3
(4.140)
(1 ) 10 RT ln
4 p0
3




pCO
GCO () = 10 226.509 + RT ln
=
p0



p
103 226.509 + RT ln
(4.141)
4 p0
3




pH2 O
GH2 O () = 10 175.774 + RT ln
=
p0



p
3
10 175.774 + RT ln
(4.142)
4 p0
3

In the above equations p0 is the reference pressure (1bar) and p is the total pressure
(in bars) at equilibrium while the universal gas constant is R = 8.314 103 molkJ K .
Fig. 4.8 shows the plot of the Gibbs enthalpy (4.138) as a function of the extent
of the reaction for T = 1300K and p = 1bar. The minimum of the Gibbs function
is at eq = 0.823, as expected. And the equilibrium composition of the mixture is
then given by (4.125) or (4.133).
Method 4 - Without considering chemical reactions
The starting point to this method is the expression for the Gibbs enthalpy of the
system
G = GCO2 + GH2 + GCO + GH2O =
mCO2 CO2 + mH2 H2 + mCO CO + mH2O H2O (4.143)
with
GCO2 = mCO2

206

RT
9.004 +
ln
MCO2

mCO2 M p

m MCO2 p0



103

(4.144)

4.4 Equilibrium Composition


-1.210

G ib b s en th alp y (G J )

-1.215

CO2/H2/CO/H2O mixture
-1.220

T = 1300 K

-1.225

n CO = 3 kmol

n H = 1 kmol
2

-1.230
-1.235
-1.240
-1.245
0.0

0.2

0.4

0.6

0.8

1.0

Extent of the reaction (kmol)


Fig. 4.8: Gibbs enthalpy as a function of extent of the reaction for a CO2 /H2 /CO/H2 O
mixture at T = 1300K, p = 1bar; initial conditions: n0CO2 = 3 kmol, n0H2 =
1 kmol




RT
mH2 M p
= mH2 0 +
ln

103
MH2
m MH2 p0

(4.145)




mCO M p
RT
ln

103
= mCO 8.090 +
MCO
m MCO p0

(4.146)

GH2

GCO

GH2O = mH2O




mH2O M p
RT
ln

103
9.765 +
MH2O
m MH2O p0

(4.147)

where mCO2 , mH2 , mCO , mH2O stand for the amounts (in kg) of the species and
MCO2 , MH2 , MCO , MH2O are the molecular masses in kg/kmol. While approaching the equilibrium, the system changes the four variables mCO2 , mH2 , mCO , mH2O
and at the equilibrium the function (4.143) reaches its minimum. However, the
changes in the four variables mCO2 , mH2 , mCO , mH2O occur so that the overall
mass (m) remains constant
m = m0CO2 + m0H2 + m0CO + m0H2O = 3 44 + 1 2 = 134 kg

(4.148)

207

4 Chemical Equilibrium
and
m = mCO2 + mH2 + mCO + mH2O

(4.149)

In Eqs. (4.144) - (4.147), M stands for the molecular mass of the mixture which
is related to the four unknowns through the relationship
mCO2
mH2
mCO
mH2O
1
=
+
+
+
M
m MCO2 m MH2 m MCO
m MH2O

(4.150)

Mathematically speaking we should find values of the four variables mCO2 , mH2 ,
mCO , mH2O which minimise function (4.143). However, while looking for the
minimum we are not allowed to vary the unknowns freely; we have to make sure
that the number of C atoms, the number of O atoms and the number of H atoms is
the same as in the initial mixture. Thus, the following equations which formulate
the balance of elements must be satisfied

balance of C
balance of O
balance of H

12
12
m0CO2 =
mCO2 +
44
44
32
32
m0CO2 =
mCO2 +
44
44
2
m0H2 = mH2 +
mH2O
18

12
mCO
28
16
16
mCO +
mH2O
28
18

(4.151)
(4.152)
(4.153)

We mention in passing that summing up Eqs. (4.151), (4.152) and (4.153) provides
Eq. (4.149), as one would expect.
Now, using again the language of mathematics, we are facing a problem of finding a minimum of function (4.143) with the three constraints formulated by
Eqs. (4.151), (4.152) and (4.153). The Lagrangean multiplier method is suitable
for such problems and it can be applied here, as pointed out earlier. However,
our current problem is rather simple and it can also be solved graphically. Using
Eqs. (4.151) - (4.153) we can relate mH2 , mCO , mH2O to mCO2
mCO = 84 0.63636 mCO2
mH2O = 54 0.4091 mCO2
mH2 = 4 + 0.04546 mCO2

(4.154)
(4.155)
(4.156)

We plot relation (4.143) as a function of a single variable mCO2 and we find that
the minimum is at mCO2 = 95.788kg, as shown in Fig. 4.9. It is easy to see

208

4.4 Equilibrium Composition


that the minimization of the Gibbs enthalpy of the system with respect to CO2 amount with constraints (4.154), (4.155) and (4.156) has provided the equilibrium
mixture composition which is in full agreement with previously obtained results
using Methods 1,2 and 3.
We complete Method-4 with a remark. The reader should notice that for the first
time in this textbook, we have obtained an equilibrium composition of a system
within which a chemical reaction occurs, without considering the reaction itself.
This is an important observation since it invokes immediately a question "Do we
need to consider at all chemical reactions for the determination of the
equilibrium of a chemically reactive mixture?". We will elaborate on this
important issue in the forthcoming paragraphs.
-1.210

G ib b s en th alp y (G J )

-1.215

CO2/H2/CO/H2O mixture
-1.220

T = 1300 K

-1.225

n CO = 3 kmol (132 kg)

n H = 1 kmol (2 kg)
2

-1.230
-1.235
-1.240
-1.245
80

90

100

110

120

130

140

mCO (kg)
2

Fig. 4.9: Gibbs enthalpy as a function of the amount of carbon dioxide for a
CO2 /H2 /CO/H2 O mixture at T = 1300K and p = 1bar (compare with
Fig. 4.8)

4.4.2 Systems with a two-dimensional reaction basis


As an example of a system with a two-dimensional reaction basis, we consider
a gas phase system containing the species: CH4 , H2 O, CO, CO2 , and H2 . The
task is to calculate the equilibrium composition at T = 1000K and p = 1bar; the
initial unreacted mixture contains 2kmol of CH4 and 3kmol of H2 O.

209

4 Chemical Equilibrium
There are five ( = 5) species and three elements ( = 3) in the system. All
possible reactions can be written as
1 CH4 + 2 H2 O + 3 CO + 4 CO2 + 5 H2 = 0

(4.157)

The balance of elements provides a matrix

1 0 1 1 0 0
0 1 1 2 0 0
4 2 0 0 2 0

(4.158)

where the first, second and the third rows show the balance of carbon, oxygen
and hydrogen, respectively. Gauss elimination process applied to matrix (4.158)
provides

1 0 1
1 0 0
1 0 1
1 0 0
0 1 1
2 0 0
2 0 0 0 1 1
0 0 6 8 2 0
0 2 4 4 2 0

(4.159)

Since the rank of matrix (4.159) is three ( = 3) then the reaction basis is two
dimensional (R = = 5 3 = 2) it means two independent reactions form
a basis. Selecting 4 = 1, 5 = 0 we obtain 3 = 4/3, 2 = 2/3, 1 = 1/3, so
the first reaction is
1
2
4
3 CH4 3 H2 O 3 CO

+ CO2 = 0

or
CH4 + 3 CO2 = 4 CO + 2 H2 O

(4.160)

Selecting 4 = 0, 5 = 1 and obtaining from matrix (4.159) 3 = 1/3, 2 =


1/3, 1 = 1/3 so that
13 CH4 13 H2 O + 31 CO + H2 = 0
or
CH4 + H2 O = CO + 3 H2

210

(4.161)

4.4 Equilibrium Composition


Thus, we have selected reactions (4.160) and (4.161) as a reaction basis. Below
we present two methods of finding the equilibrium composition. Only in the first
methods we use the selected reaction basis.
Method 1 - Based on the extent of the reactions and the equilibrium
constants
We begin with calculating the thermodynamic equilibrium constants for the selected reaction basis. To this end we use the first four reactions listed in Table 4.2
which are here copied for completeness
C2 H2 + CH4 = 0

(R1)

CO2 + CO2 = 0

(R2)

C 21 O2 + CO = 0

(R3)

H2 12 O2 + H2 O = 0

(R4)

Since reaction (4.160) can be composed using reactions R1-R4 as follows


R(4.160) = R1 3 R2 + 4 R3 + 2 R4
log K(4.160) = (1.018) 3 20.679 + 4 10.461 + 2 10.060 = 0.945
and K(4.160) = 8.8105.
Reaction (4.161) can be created using reactions R1-R4 as follows
R(4.161) = R1 + R3 R4
and therefore
log K(4.161) = (1.018) + 10.461 10.060 = 1.419
and K(4.161) = 26.2422.
We pause here to summarise. We have the following two reactions into which we
prescribe two extents, 1 and 2 , so that
CH4 + 3CO2 = 4CO + 2H2 O

= 6 4 = 2

K = 8.8105

CH4 + H2 O = CO + 3H2

= 4 2 = 2

K = 26.2422

211

4 Chemical Equilibrium
with the initial number of kmol: n0CH4 = 2kmol and n0H2O = 3kmol. Now we
express the number of kmol of each species as a function of 1 and 2 so that
nCH4 = n0CH4 1 2 = 2 1 2
nH2O = n0H2O + 2 1 2 = 3 + 2 1 2
nCO2 = 3 1
nCO = 4 1 + 2
nH2 = 3 2

nT = n0CH4 + n0H2O + 2 1 + 2 2 = 5 + 2 1 + 2 2

Using relationship (4.44) to calculate equilibrium constant Kn and after some


algebra we obtain for the first reaction
(4 1 + 2 )4 (3 + 2 1 2 )2
8.8105 (5 + 2 1 + 2 2 )2 = 0
(2 1 2 ) (3 1 )3

(4.162)

and for the second reaction


(4 1 + 2 ) (3 2 )3
26.2422 (5 + 2 1 + 2 2 )2 = 0
(2 1 2 ) (3 + 2 1 2 )

(4.163)

It is not a trivial matter to find the solution of Eqs. (4.162) and (4.163) since
they are highly non-linear. There exist several pairs 1 and 2 which satisfy the
equations and therefore the more we know about the solution the better. After
examining the interrelationships between the number of kmol and the reaction
extends 1 and 2 , we expect 1eq < 0 and 2eq > 0 at the equilibrium. One of
the easiest methods for finding a solution of Eqs. (4.162) and (4.163) is to frame
the problem in terms of residuals. If we mark the left hand side of Eq. (4.162) as
F1 (1 , 2 ) and the left hand side of Eq. (4.163) as F2 (1 , 2 ), finding the solution
is equivalent to minimising the residuals of a function F12 (1 , 2 ) + F22 (1 , 2 ). To
minimise this function we have used a general gradient method which is easily
accessible since it is incorporated into the Solver option of the Excel program. The

212

4.4 Equilibrium Composition


Solver needs an initial guess of 1 and 2 . For the initial values of 1 = 0.5 and
2 = 2.0 the solution has been found to be 1eq = 0.1065364 and 2eq = 1.931939.
eq
Consequently, the equilibrium mixture contains: neq
CH4 = 0.17460 kmol, nH2O =
eq
eq
eq
0.85499 kmol, nCO2 = 0.31961 kmol, nCO = 1.50579 kmol, nH2 = 5.79582 kmol
eq
and neq
T = 8.65081kmol. Thus, the mole fractions at the equilibrium are xCH4 =
eq
eq
eq
eq
0.02018, xH2O = 0.09883, xCO2 = 0.03695, xCO = 0.17406, xH2 = 0.66997.
Method-2 Minimization of Gibbs enthalpy of the system
The method presented in this paragraph is almost identical to Method-4 of the
previous paragraph. The main difference is in that the minimization procedure
is going to be performed with respect to two variables. The Gibbs enthalpy at
T = 1000K for the five species present in the system are given in Table 4.5.
Table 4.5: Standard Gibbs enthalpies at T = 1000 K [12]

Species
CH4
H2 O(g)
CO
CO2
H2

f G01000
kJ/mol
19.492
-192.590
-200.275
-395.886
0

The starting point is the expression for the Gibbs enthalpy of the whole system
G = GCH4 + GH2O + GCO + GCO2 + GH2

(4.164)




RT
mCH4 M p
= mCH4 1.218 +
ln

103
MCH4
m MCH4 p0

(4.165)




mH2O M p
RT
ln

103
= mH2O 10.699 +
MH2O
m MH2O p0

(4.166)




RT
mCO M p
= mCO 7.153 +
ln

103
MCO
m MCO p0

(4.167)

with

GCH4

GH2O

GCO

213

4 Chemical Equilibrium

GCO2 = mCO2

GH2

RT
8.997 +
ln
MCO2

mCO2 M p

m MCO2 p0



103




mH2 M p
RT
ln

103
= mH2 0 +
MH2
m MH2 p0

(4.168)

(4.169)

where mCH4 , mCO2 , mH2 , mCO , mH2O stand for the amounts (in kg) of the species
and MCH4 , MCO2 , MH2 , MCO , MH2O are the molecular masses in kg/kmol and
M stands for the molecular mass of the mixture which is related to the species
amounts through the relationship

1
mCH4
mH2O
mCO
mCO2
mH2
=
+
+
+
+
M
m MCH4 m MH2O
m MCO
m MCO2 m MH2

(4.170)

The overall mass of the system is


m = m0CH4 + m0H2O = 2 16 + 3 18 = 86 kg

(4.171)

Now the task is to find values of the five variables mCH4 , mH2 O , mCO , mCO2 ,
mH2 which minimise function (4.164). The following equations formulate the
constraints of the minimization problem:

balance of C
balance of O
balance of H

12
12
m0CH4 =
mCH4 +
16
16
16
16
m0H2O =
mH2O +
18
18
4
2
m0CH4 +
m0H2O =
16
18

12
12
mCO +
mCO2
(4.172a)
28
44
16
32
mCO +
mCO2
(4.172b)
28
44
4
2
mCH4 +
mH2O + mH2
16
18
(4.172c)

Since we have five variables mCH4 , mH2 O , mCO , mCO2 , mH2 and three constraints,
we express7 mCO , mCO2 , mH2 as a function of mCH4 and mH2 O so that
mCO = 28 3.5 mCH4 + 1.5556 mH2O
7

(4.173)

You can practice Gauss elimination to obtain relationships (4.173), (4.174) and (4.175).

214

4.4 Equilibrium Composition

mCO2 = 44 + 2.75 mCH4 2.4445 mH2O

(4.174)

mH2 = 14 0.25 mCH4 0.1111 mH2O

(4.175)

Fig. 4.10: Gibbs enthalpy of the considered system as a function of mCH4 and mH2 O .
The mixture contains five species CH4 , H2 O, CO, CO2 , H2 at T = 1000K
and p = 1bar. Initial non-reacted mixture of 2 kmol CH4 and 3 kmol H2 O.

Fig. 4.10 shows a plot of function (4.164) in the vicinity of the minimum for
the system in question. It is to observe that the minimum of function (4.164) is
somewhere around mCH4 2.5 kg and mH2O 15 kg. To find the exact values
of mCH4 which minimise function (4.164) we have used a general gradient method
which is incorporated into the Solver option of the Excel program. For an initial
guess of mCH4 = 2.5 kg and mH2O = 15 kg, the solver finds the exact solution as
eq
meq
CH4 = 2.7901 kg and mH2O = 15.3835 kg. The remaining species are present
at the equilibrium in amounts determined by Eqs. (4.173), (4.174) and (4.175)
eq
eq
so meq
CO = 42.165 kg, mCO2 = 14.068 kg and mH2 = 11.593 kg. Recalculating

215

4 Chemical Equilibrium
these values into mole fractions we obtain the following equilibrium composition:
eq
eq
eq
eq
xeq
CH4 = 0.02016, xH2O = 0.09879, xCO = 0.17407, xCO2 = 0.03696, xH2 = 0.670.
Obviously the just calculated values are in close agreement with the solution
obtained using Method-1.

4.4.3 Systems with a multi-dimensional reaction basis


In Section 4.4.1 we have considered systems with a one-dimensional reaction basis
(R = = 1). We have observed that when a single chemical reaction is
sufficient to describe the system (to form reaction basis) we could determine the
equilibrium composition by using several methods namely Method 1, Method 2
and Method 3. All these three methods have used reaction (4.120) and they
are rather simple. Indeed, it is likely that you have already used Method 1 (or
Method 2) in your high school chemistry course. However, Method 4, based
on the minimization of Gibbs enthalpy, is perhaps the most interesting since we
have succeeded in calculating the equilibrium composition without considering
any chemical reactions whatsoever.
In Section 4.4.2 we have considered systems with a two-dimensional reaction basis
(R = = 2). Method-1, based on extents of the reaction basis and the
equilibrium constants, has already become a bit more cumbersome. Firstly, we
had to identify the two reactions forming the reaction basis. Secondly we had
to calculate the equilibrium constants of the reactions and thirdly the resulting
Eqs. (4.162) and (4.163) were highly non-linear. All together, the procedure
underlined in Method-1 was doable but tedious and perhaps it was at the limit
of our patience. Imagine what would happen when a system of, let say, twenty
species would have to be considered. In short, Method-1 does not seem to be
easy to generalise to tackle systems of many reacting species. To the contrary,
Method-2 based on Gibbs enthalpy minimization retains its simplicity even if
systems of many species are considered. Most importantly, we are not bothered
with the chemical reactions so the question of what and how many reactions
should we chose is irrelevant. We just have to make sure that both Gibbs system
enthalpy and constraints are properly formulated. Having done that, we need a
mathematical procedure for finding a minimum of the Gibbs enthalpy with all
constraints satisfied. As a matter of fact, we have already used in Section 3.6 and
Section 4.3.1 the Lagrangean multipliers method which does just this.
Method - Minimization of Gibbs system enthalpy using Lagrangean
multipliers
Although the theoretical background of the method has already been provided in

216

4.4 Equilibrium Composition


Section 3.6 (see Example 3.5) we demonstrate here its applicability for calculating
the equilibrium composition of the system containing the five species: CH4 , H2 O,
CO, CO2 and H2 . Again we calculate the equilibrium composition at T = 1000K
and p = 1bar; the initial unreacted mixture contains 2kmol of CH4 and 3kmol of
H2 O.
The starting point is again relationship (4.164) which we copy for the sake of
completeness
G = GCH4 + GH2O + GCO + GCO2 + GH2
(4.176)
with the constraints (4.172a), (4.172b), (4.172c)
24 0.75000 mCH4 0.42857 mCO 0.27273 mCO2 = 0

(4.177)

48 0.88889 mH2O 0.57143 mCO 0.72727 mCO2 = 0

(4.178)

14 0.25000 mCH4 0.11111 mH2O mH2 = 0

(4.179)

The new function Y (see Example 3.5) is formulated as


Y = GCH4 + GH2O + GCO + GCO2 + GH2 +
+ 1 [24 0.75000 mCH4 0.42857 mCO 0.27273 mCO2 ] +
+ 2 [48 0.88889 mH2O 0.57143 mCO 0.72727 mCO2 ] +
+ 3 [14 0.25000 mCH4 0.11111 mH2O mH2 ] = 0 (4.180)
The necessary conditions for a minimum of Y-function are:
Y
= CH4 0.75000 1 0.25000 3 =
mCH4



mCH4 M p
RT
ln

0.75000 1 0.25000 3 = 0
1.218 +
MCH4
m MCH4 p0
(4.181)



mH2O M p
RT
Y
= 10.699 +
ln

mH2O
MH2O
m MH2O p0
0.88889 2 0.11111 3 = 0 (4.182)
Y
=
mCO



mCO M p
RT
ln

0.42857 1 0.57143 2 = 0
7.153 +
MCO
m MCO p0
(4.183)

217

4 Chemical Equilibrium
Y
=
mCO2



RT
mCO2 M p
8.997 +
0.27273 1 0.72727 2 = 0
ln

MCO2
m MCO2 p0
(4.184)

Y
mH2




mH2 M p
RT
0+
ln

3
MH2
m MH2 p0

0 (4.185)

Y
= 24 0.75000 mCH4 0.42857 mCO 0.27273 mCO2 = 0 (4.186)
1
Y
= 48 0.88889 mH2O 0.57143 mCO 0.72727 mCO2 = 0 (4.187)
2
Y
3

14 0.25000 mCH4 0.11111 mH2O mH2

0 (4.188)

Inserting into (4.181) - (4.188) T = 1000K, p = 1bar, m = 86kg and appropriate


values for the molecular masses we obtain


mCH4 M
2.3438 + ln
1.4432 1 0.4811 3 = 0
(4.189)
1.376 103


mH2O M
23.1616 + ln
1.9243 2 0.2405 3 = 0
(4.190)
1.548 103


mCO M
24.0869 + ln
1.4432 1 1.9243 2 = 0
(4.191)
2.408 103


mCO2 M
1.4432 1 3.8486 2 = 0
(4.192)
47.6128 + ln
3.784 103


mH2 M
0 + ln
0.2405 3 = 0
(4.193)
172
24 0.75000 mCH4 0.42857 mCO 0.27273 mCO2 = 0
(4.194)
48 0.88889 mH2O 0.57143 mCO 0.72727 mCO2 = 0

(4.195)

14 0.25000 mCH4 0.11111 mH2O mH2 = 0

(4.196)

The above equations are accompanied by relationship (4.170) which is here written

218

4.5 Summary
as
mCH4
mH2O
mCO
mCO2
mH2
1
=
+
+
+
+
3
3
3
3
M
1.376 10
1.548 10
2.408 10
3.784 10
172

(4.197)

Perhaps it is time to pause and summarise. We have formulated eight equations


(4.189) - (4.196); five equations (4.189) - (4.193) stem from the differentiating
of the Gibbs enthalpy with respect to the amounts of the five species while the
remaining three equations are the constraints. Thus, all together we have eight
equations and eight unknowns (five unknown amounts of the species mCH4 , mH2O ,
mCO , mCO2 , mH2 and three multipliers 1 , 2 , 3 ). Finding a solution of the eight
equations is equivalent to minimising the residuals of a function F12 + F22 + ... + F82
which at the solution should take near zero value. To minimise this function
we used again the Solver option of the Excel program. For an initial guess of
mCH4 = 10 kg and mH2O = 20 kg, the solver finds the exact solution as meq
CH4 =
eq
eq
eq
2.7906 kg and mH2O = 15.3855 kg, mCO = 42.166 kg, mCO2 = 14.065 kg and
meq
H2 = 11.593 kg. Obviously the just calculated values are in close agreement
with the solutions obtained previously.

4.5 Summary
This chapter is concerned with chemical equilibrium which has been defined as
the state of the system if no changes occur to its chemical composition. The
ultimate goal of this chapter is to provide the knowledge required for calculation
of chemical composition at equilibrium.
Extent of a reaction (), expressed in mol or kmol, is a single variable which relates
the changes in the amount of species participating in the chemical reaction. It
varies from zero to a value that corresponds to the equilibrium state. At constant
a temperature and a pressure, when a reaction proceeds towards equilibrium,
Gibbs enthalpy of the system approaches a minimum. The derivative of Gibbs
enthalpy of the system with respect to the extent of the reaction is the Gibbs
reaction enthalpy (R GpT ). More explicitly, the Gibbs reaction enthalpy is the
change in Gibbs enthalpy of the system when the extent of the reaction changes
by 1 mol (or kmol). At the equilibrium the Gibbs reaction enthalpy is zero. On
the basis of this statement the thermodynamic equilibrium constant is derived
which is a dimensionless quantity. Several other equilibrium constants have been
derived from the thermodynamic constant.
The general conditions for a multi-component, multi-phase system with chemical

219

4 Chemical Equilibrium
reactions to be at equilibrium are:
T (1) = T (2) = T (3) = . . . = T (i) = . . . = T () = T
p(1) = p(2) = p(3) = . . . = p(i) = . . . = p() = p
1k = 2k = 3k = . . . = k
and


G
r

= r GpT =

X
k=1

for k = 1, 2, . . . ,

k Mk kr =

k kr = 0

k=1

Thus, equality of temperatures and pressures of each phase as well as equality


of the chemical potentials in each phase for each component are required. Furthermore, Gibbs enthalpies of all the chemical reactions must be zero at chemical
equilibrium.
A considerable part of this chapter has been dedicated to methods used for determination of equilibrium composition for given both a pressure and a temperature
and when the composition of the initial (un-reacted) mixture is given. This is
a classical problem in combustion-, process- and chemical-engineering and must
be well understood. In high-schools you have learned how to determine the equilibrium composition for a single chemical reaction using both the extent of the
reaction and the equilibrium constant. Here you have extended this knowledge
into more complex systems supporting a multiplicity of chemical reactions. A
procedure for the determination of both the number of reactions and the reactions themselves has been underlined in this chapter. However, if the number
of reaction exceeds two, the calculation methods have become tedious and they
exceed the upper limit of patience for a reasonably human being.
Perhaps the most important message of this chapter is that there is no need to
consider chemical reactions at all to calculate the chemical equilibrium. This
should not be too surprising since from the thermodynamic point of view the
equilibrium state is uniquely determined by the temperature, pressure and the
initial (un-reacted) mixture. The procedure of Gibbs enthalpy minimization is
highly recommended. In this procedure, the question of what chemical reactions
are involved never enters directly into any of the equations. However, the choice
of a set of species is entirely equivalent to the choice of a reaction basis among
the species. In any case, a set of species must always be assumed and different
assumptions produce different results. Students should master the method of Lagrangean multipliers used to determine a minimum of functions describing Gibbs
enthalpy of the system.

220

5 Elements of Chemical Kinetics


Contents
5.1

Introduction

5.2

Rate Laws and Reaction Orders

5.3

Forward and Reverse Reactions

5.4

Elementary Reactions and Reaction Molecularity

5.5

Rate of Reactions

5.6

5.5.1

Temperature dependence of rate coefficients

5.5.2

Pressure dependence of rate coefficients

Summary

5.1 Introduction
In Chapter 4 we have learned that equilibrium composition can be found without
any considerations as to the mechanisms of chemical reactions. This is, because
the equilibrium composition for a given temperature and a pressure is uniquely
determined by the initial conditions. The calculations carried out in Chapter 4 do
not provide any hint as to how long it takes to reach the equilibrium. A discipline
called Chemical Kinetics aims both at providing insights into the reaction
mechanisms and at determining the rates of individual reactions. Thus, Chemical
Kinetics provides means of estimating the time needed for reactions to reach a
certain extent. Essential features of chemical kinetics, which occur frequently in
combustion phenomena, are reviewed in this chapter. For a thorough coverage of
the subject one should refer to books on chemical kinetics [18, 19].

221

5 Elements of Chemical Kinetics

5.2 Rate Laws and Reaction Orders


Lets write a chemical reaction as:
A A + B B + C C + . . . D D + E E + F F + . . .

(5.1)

where A, B, C, . . . and D, E, F, . . . denote chemical species involved in the reaction.


A rate law is an empirical formula for the reaction rate; it means for the rate of
formation (for example in mol/s) or consumption of a species. In chemical kinetics,
the rate of consumption of species A is expressed as:
d[A]
= k [A]a [B]b [C]c . . .
dt

(5.2)

In the above expression a, b, c, . . . are reaction orders with respect to species


A, B, C, . . . and k is the rate coefficient of the reaction. The sum of the all
exponents a, b, c, . . . is the overall reaction order. The convention used in
this chapter is that square parenthesis around a chemical symbol denotes the
concentration of that species in moles (or grams) per cubic centimeter. If for
example concentrations of species B and C remain nearly constant during the
course of the reaction (species B and C are in excess), one may obtain a simplified
version of expression (5.2)
d[A]
= kexper [A]a
dt

(5.3)

where kexper denotes the rate constant for a particular experiment:


kexper = k [B]b [C]c

(5.4)

Example 5.1
When exponent "a" in Eq. (5.3) equals one (a = 1) the reaction is called a firstorder reaction. Calculate the rate of change of the A-species concentration with
time for a first-order reaction.
Assumptions: none
For a = 1 the rate of the consumption of A-species is:
d[A]
= kexper [A]
dt

222

5.2 Rate Laws and Reaction Orders


A separation of the variables leads to:
d[A]
= kexper dt
[A]
and a subsequent integration gives:
ln

[A](t)
= kexper (t t0 )
[A]0

Fig. 5.1: Concentration change with time for a first-order reaction (t0 = 0)

Comments:
(a) For first-order reactions, logarithm of normalised concentration of the reagent
[A]
[A]0 is a linear function of time.
(b) The slope is determined by the rate constant k.
End of Example 5.1
It is informative to compare how quickly the concentration of the reacting species
(A) decays with time, depending on the reaction order and the initial concentra-

223

5 Elements of Chemical Kinetics


tion of [A]0 . The equations derived in Examples 5.1 and 5.2 yield:
[A](t) = [A]0 exp(kexper (t t0 ))
1
[A](t) = 1
[A]0 + kexper (t t0 )

for a first-order reaction


for a second-order reaction

while
[A](t) =

1
1
[A]20

+ 2 kexper (t t0 )

for a third-order reaction.

As for any time-dependent process one can calculate the time constant for any
single reaction. The time constant is the time required for the initial concentration
[A]0 to drop to [A]e 0 , where e = 2.718 281 83 is the base of natural logarithm. For
example, for a first-order reaction:
[A](t) = [A]0 exp(kexper (t t0 ))
1
[A](tc )
= = exp(kexper (tc t0 ))
[A]0
e

(5.5)
(5.6)

and
tc = t0 +

1
kexper

(5.7)

It is easy to verify that time constants for the second- and third-order reactions
are given by:
(e 1)
kexper [A]0

e2 1
tc = t0 +
kexper [A]20
tc = t0 +

for a second-order reaction

(5.8)

for a third-order reaction

(5.9)

Example 5.2
When exponent "a" in Eq. (5.3) equals two (a = 2) the reaction is called a
second-order reaction while for a = 3 the reaction is third-order. Calculate
the rate of change of the A-species concentration with time for such reactions.
Assumptions: none

224

5.2 Rate Laws and Reaction Orders


For a = 2 the rate of the consumption of A-species is:
d[A]
= kexper [A]2
dt
and after separation of variables:
d[A]
= kexper dt
[A]2
and integration, one obtains:
1
1

= kexper (t t0 )
[A](t) [A]0

Fig. 5.2: Concentration change with time for a second-order reaction

Comments:
(a) Fig. 5.1 and Fig. 5.2 show a good method of determining the reaction order
using measured data.
End of Example 5.2

225

5 Elements of Chemical Kinetics

5.3 Forward and Reverse Reactions


At chemical equilibrium the rate of consumption equals the rate of formation of
chemical species and the forward and backward reactions are of the same rate.
While expression (5.2) holds for the forward reaction, the following is applicable
for the backward reaction:
d[A]
= kb [D]d [E]e [F ]f . . .
dt

(5.10)

At chemical equilibrium one has


kf [A]a [B]b [C]c . . . = kb [D]d [E]e [F ]f . . .

(5.11)

where kf and kb are the rate constants for the forward and backward reactions,
respectively.
Rearranging Eq. (5.11) one obtains:
kf
[D]d [E]e [F ]f . . .
=
a
b
c
kb
[A] [B] [C] . . .

(5.12)

The left hand side of the above equation is the equilibrium constant which we
have already derived in Chapter 4, thus
Kc =

kf
kb

(5.13)

Example 5.3
Consider a first-order reaction

B
close to equilibrium. Derive a formula for calculating A- and B- species concentrations with time. Assume that initially (at t = 0) only A-species is present
and its initial concentration is [A]0 . Calculate both the equilibrium composition
and the time required to reach the equilibrium for kf = 2 s1 , kb = 2 s1 and
kf = 3 s1 , kb = 1 s1 .
Assumptions: Both forward and backward reactions are first order.
We begin with noting that the concentration of A-species is reduced by the forward
reaction and it is increased by the backward reaction. The net rate change is
therefore:
d[A]
= kf [A] + kb [B]
dt

226

5.3 Forward and Reverse Reactions


Since the initial concentration of A is [A]0 and no B-species is present, then at
any instant [A] + [B] = [A]0 and
d[A]
= kf [A] + kb ([A]0 [A]) = (kf + kb ) [A] + kb [A]0
dt
or
d[A]
+ (kf + kb ) [A] = kb [A]0
dt
The above equation is a first-order linear differential equation of a general form:
y + p(x) y = r(x)
where p(x) is a constant equal (kf + kb ) and r(x) is also a constant equal kb [A]0 .
The solution of such an equation can
R be found upon multiplying both sides of the
equation by an integrating factor e p(x) dx obtaining:
e

p(x) dx

y + p(x) y e

p(x) dx

= r(x) e

p(x) dx

Observing the product rule of differentiation one obtains:




R
R
y e p(x) dx = r(x) e p(x) dx

and now integrating with respect to x, we obtain the general form of the solution:
Z
R
R
p(x) dx
= r(x) e p(x) dx dx + C
ye

Now we apply this general solution to our chemical reaction rate equation. We
recall that: p(x) = (kf + kb ) and r(x) = kb [A]0 while y-variable represents [A](t)
and x-variable represents the time t, so
e
Z

r(x) e

p(x) dx

p(x) dt

dx =

=e

(kf +kb ) dt

= e(kf +kb ) t

kb [A]0 e(kf +kb ) t dt =

kb [A]0
e(kf +kb ) t
(kf + kb )

227

5 Elements of Chemical Kinetics


and
[A](t) =

C
kb [A]0
+ (k +k ) t
(kf + kb ) e f b

We can easily calculate the integration constant C using the initial conditions:
[A](0) = [A]0 so
kb [A]0
+C
[A]0 =
(kf + kb )
and
C = [A]0

kf
(kf + kb )

Thus, the final form of the solution of the chemical rate equation is:
[A](t) =

kb + kf e(kf +kb ) t
[A]0
(kf + kb )

for A-species (C1)

and
[B](t) = [A]0 [A](t) = [A]0

kb + kf e(kf +kb ) t
1
(kf + kb )

for B-species (C2)

Fig. 5.3 shows the plot of the A-species and B-species concentrations with time
for different values of kf and kb rate constants.

The above equations can be used to calculate the equilibrium composition of the
reacting species. This is obtained upon examining what happens to [A](t) and
[B](t) when t . Such an analysis leads to the following:
[A]() = lim [A](t) = [A]0 lim
t

[B]() = lim [B](t) = [A]0 lim


t

kb + kf e(kf +kb ) t
kb
= [A]0
(kf + kb )
kf + kb
!
(k
+k
)

t
kf
kb + kf e f b
= [A]0
1
(kf + kb )
kf + kb

For kf = 2 s1 and kb = 2 s1 , the equilibrium composition is


[B]eq
[A]0
[A]eq
[A]0

228

1
2

=
=

1
4

[A]eq
[A]0

1
2

and

while for kf = 3 s1 and kb = 1 s1 the equilibrium composition is


and

[B]eq
[A]0

3
4

as shown in Fig. 5.3.

5.3 Forward and Reverse Reactions

Fig. 5.3: The approach of concentrations to their equilibrium values for the reversible

reaction A

B that is first order in each direction. The solid line: kf =


2 s1 and kb = 2 s1 ; the dash dot line: kf = 3 s1 and kb = 1 s1 .

We will now calculate the time required to reach the equilibrium. Strictly speaking an infinitely long time is needed to reach the equilibrium as shown in the
above analysis. However, as shown in Fig. 5.3, practically after around 1.3 s
the concentrations of the reacting species do not vary any more with time for
kf = kb = 2 s1 . Thus, we will calculate the time for the A-species concentration
to be within 1 % from the equilibrium concentration. To this end we begin with
the relationship describing the dependence of the A-species concentration with
time:
kb + kf e(kf +kb ) t
[A](t) =
[A]0
(kf + kb )
and
1.01

kb + kf e(kf +kb ) t0.01


kb
[A]0
[A]0 =
kf + kb
(kf + kb )

and
0.01 kb = kf e

(kf +kb ) t0.01

t0.01 =

kb
ln 0.01
kf

(kf + kb )

Thus, for kf = 2 s1 and kb = 2 s1 the time required for A-species to reach a

229

5 Elements of Chemical Kinetics


concentration that is 1 % different from the equilibrium concentration is 1.15 s. For
kf = 3 s1 and kb = 1 s1 a value of 1.43 s is applicable. To complete our exercise
we will carry out simple thermodynamic calculations to obtain the equilibrium

composition. The equilibrium constant for the reaction A

B reads:
Kc =

kf
[B]eq
=
[A]eq
kb

The equilibrium concentrations can be expressed in terms of the extent of the


reaction (see Section 4.2.1) by drawing up the following table:

Initial concentration
Change to reach equilibrium
Concentration at equilibrium

A-species
[A]0
[A]0
[A]0 (1 )

B-species
0
[A]0
[A]0

The equilibrium constant is therefore:


Kc =

kf
eq [A]0
eq
=
=
kc
(1 eq ) [A]0
1 eq

For kf = 2 s1 and kb = 2 s1 , the equilibrium constant Kc equals 1 and eq


equals 12 , so the equilibrium composition is [A]eq = 12 [A]0 and [B]eq = 12 [A]0
while for kf = 3 s1 and kb = 1 s1 the equilibrium composition is [A]eq = 41 [A]0
and [B]eq = 43 [A]0 . These values are in agreement with our kinetic calculations
shown in Fig. 5.3.
Comments:
(a) Mathematically speaking an infinitely long time is needed to reach equilibrium.
(b) Relationship (C1) which has been derived on the basis of chemical kinetics,
provides the equilibrium concentration of A-species upon invoking the condition t ; similarly relationship (C2) gives the equilibrium concentration
of B-species upon t .
(c) Generally, when t chemical kinetics meets equilibrium thermodynamics.
(d) Note that relationship (5.13), namely Kc =
only.

kf
kb ,

End of Example 5.3

230

is valid at the equilibrium

5.4 Elementary Reactions and Reaction Molecularity

5.4 Elementary Reactions and Reaction


Molecularity
If a reaction occurs on a molecular level exactly as it is described by the reaction
equation (Eq. (5.1)) then it is called an elementary reaction. The reaction of
hydroxyl radicals (OH) with molecular hydrogen (H2 ) forming water and hydrogen
atoms is an elementary reaction,
OH + H2 H2 O + H

(5.14)

Hydroxyl radical molecules collide with hydrogen molecules and in the case of
reactive collisions water molecules and hydrogen molecules are formed. In the
case of non-reactive collisions the molecules bounce away.
The reaction
2 H2 + O2 2 H2 O

(5.15)

is not an elementary reaction. According to this reaction two molecules of hydrogen should collide with one molecule of oxygen to produce two molecules of
water. Detailed studies show that many reactive intermediates like H, O and
OH are formed. Thus, reaction (5.15) describes an overall effect of numerous
elementary reactions; reactions (5.15) are called overall reactions. Often these
overall reactions have complicated rate laws with the reactions orders a, b, c, . . .
(see Eq. (5.2)) being usually not integers that can be negative. Table 5.1 (adapted
from [20]) shows some elementary reactions in H2 COC1 C2 O2 system.
There are several advantages in using elementary reactions. The reaction order of
these reactions is constant (not dependent on time and experimental conditions)
and it can be easily determined knowing the reaction equation. In order to be
able to use the elementary reactions one has to establish the molecularity of
the reaction. The molecularity is the number of species that form the reaction
complex that is the transition state between reactants and products. Only three
values of the reaction molecularity have been observed experimentally. These are
unimolecularity, bimolecularity and trimolecularity.
Unimolecular reactions proceed according to the equation:
A products

(5.16)

and they describe dissociation of a molecule. The rates of these reactions are
described by a first-order equation (see Example 5.1).

231

5 Elements of Chemical Kinetics


Bimolecular reactions proceed according to the equations:
A + A products

(5.17)

A + B products

(5.18)

or

and they have a second-order rate law.


Trimolecular reactions are proceeding according to the equations:
A + A + A products
A + A + B products

(5.19)

A + B + C products
and these are usually recombination reactions.
It is important to distinguish molecularity from order: reaction order is an empirical quantity obtained from the experimental rate law, while the molecularity
refers to an elementary reaction proposed as an individual step in a mechanism.
In general the molecularity equals the order for elementary reactions.
Consider a system containing N species (n = 1, 2, . . . , N ) taking part in R elementary reactions (r = 1, 2, . . . , R). Each of these elementary reactions r can be
written as:
N
X

n=1

r
(s)
r,n
An

N
X

(p)
r,n
An

for r = 1, 2, . . . , R

n=1

(5.20)

where the superscripts (s) and (p) represent substrates and products, respectively.
The rate of formation of the nth -species in reaction r is given by:

N


Y
(e)
[An ]
r,n
(p)
(e)
[A
]
=

n
r
r,n
r,n
t r-reaction

for n = 1, 2, . . . , R

n=1

232

(5.21)

5.4 Elementary Reactions and Reaction Molecularity


The summation for all the elementary reactions gives the overall rate of formation
of nth -species as:

R
X
[An ]
t
r=1

=
r-reaction

R 
X
r=1

N

Y
(e)
(p)
(e)
[An ]r,n
r,n
r,n
kr

for n = 1, 2, . . . , N

n=1

(5.22)

Table 5.1: (adapted from [20]).


Elementary reactions in H2 COC1 O2
system at p = 1 bar for high temperature (T > 1200 K); rate

coefficients are presented in the form k
=
A T b exp RET ,
[M ] = [H2 ] + 13
[ H2 O] + 52 [ O2 ] + 25 [ N2 ] + 34 [ CO] + 32 [ CO2 ] + 3 [ CH4 ];
2
only forward reaction is considered; reverse reaction to be calculated using
(5.13).
Reaction
01.04. H2 CO Oxidation
01. H2 O2 Reactions (HO2 , H2 O2 not included)
O2
+H
=OH
+O
H2
+O
=OH
+H
H2
+OH
=H2 O
+H
OH
+OH
=H2 O
+O
H
+H +M
=H2
+M

O
+O +M
=O2
+M

H
+OH +M
=H2 O
+M
02. HO2 Formation/Consumption
H
+O2 +M
=OH
+O +M
HO2
+H
=OH
+OH
HO2
+H
=H2
+O2
HO2
+H
=H2 O
+O
HO2
+O
=OH
+O2
HO2
+OH
=H2 O
+O2
03. H2 O2 Formation/Consumption
HO2
+HO2
=H2 O2
+O2
OH
+OH +M
=H2 O2
+M
H2 O2
+H
=H2
+HO2
H2 O2
+H
=H2 O
+OH
H2 O2
+O
=OH
+H2 O
H2 O2
+OH
=H2 O
+HO2
04. CO Reactions
CO
+OH
=CO2
+H
CO
+HO2
=CO2
+OH
CO
+O +M
=CO2
+M
CO
+O2
=CO2
+O
10.19. C1 -Hydrocarbons Oxidation

A
[cm,mol,s]

E
[kJ/mol]

2.00 1014
5.06 104
1.00 108
1.50 109
1.80 1018
2.90 1017
2.20 1022

0.00
2.67
1.60
1.14
1.00
1.00
2.00

70.30
26.30
13.80
0.42
0.00
0.00
0.00

2.30 1018
1.50 1014
2.50 1013
3.00 1013
1.80 1013
6.00 1013

0.80
0.00
0.00
0.00
0.00
0.00

0.00
4.20
2.90
7.20
1.70
0.00

2.50 1011
3.25 1022
1.70 1012
1.00 1013
2.80 1013
5.40 1012

0.00
2.00
0.00
0.00
0.00
0.00

5.20
0.00
15.70
15.00
26.80
4.20

6.00 106
1.50 1014
7.10 1013
2.50 1012

1.50
0.00
0.00
0.00

3.10
98.70
19.00
200.00

233

5 Elements of Chemical Kinetics


Table 5.1: (continued)
Reaction
10.
CH
CH
CH
CH
CH
CH
11.
CHO
CHO
CHO
CHO
CHO
CHO
CHO
12.
CH2
CH2
CH2
CH2
CH2
CH2
CH2
CH2
CH2
CH2
CH2
13.
CH2 O
CH2 O
CH2 O
CH2 O
CH2 O
CH2 O
CH2 O
14.
CH3
CH3
CH3
CH3
CH3
CH3 O
CH3
CH3
CH3

234

CH Reactions
+O
+O2 +M
+CO2
+H2 O
+H2 O +M
+OH
CHO Reactions
+M
+H
+O
+O
+OH
+O2
+CHO
CH2 Reactions
+H
+O
+CH2
+CH2
+CH3
+O2
+O2
+M
+O2
+H2
+ CH3
CH2 O Reactions
+M
+H
+O
+OH
+HO2
+CH3
+O2
CH3 Reactions
+M
+M
+O
+H
+OH
+H
+O2
+HO2
+HO2

A
[cm,mol,s]

E
[kJ/mol]

=CO
=CHO
=CHO
=CH2 O
=CHO
=CHO

+H
+O
+CO
+H
+O
+CO

4.00 1013
6.00 1013
3.40 1012
3.80 1012
6.00 1012
3.40 1013

0.00
0.00
0.00
0.00
0.00
0.00

0.00
0.00
2.90
3.20
3.20
0.00

=CO
=CO
=CO
=CO2
=CO
=CO
=CH2 O

+H +M
+O
+CO
+H
+O
+CO
+CO

7.10 1012
9.00 1013
3.00 1013
3.00 1013
1.00 1014
3.00 1013
3.00 1013

0.00
0.00
0.00
0.00
0.00
0.00
0.00

0.00
0.00
2.90
3.20
3.20
0.00
0.00

=CH
=CO
=C2 H2
=C2 H2
=C2 H4
=CO
=CO2
=3 CH2
=CO
=CH3
=C2 H4

+H2
+H +H
+H2
+H +H
+H
+OH +H
+H2
+M
+OH +H
+H
+H

6.00 1012
8.40 1012
1.20 1013
1.10 1014
4.20 1013
1.30 1013
1.20 1013
1.20 1013
3.10 1013
7.20 1013
1.60 1013

0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00

7.50
0.00
3.40
3.40
0.00
6.20
6.20
0.00
0.00
0.00
2.38

=CHO
=CHO
=CHO
=CHO
=CHO
=CHO
=CHO

+H +M
+H2
+OH
+H2 O
+H2 O2
+CH4
+HO2

5.00 1016
1.60 1013
1.60 1013
1.60 1013
1.60 1013
1.60 1013
1.60 1013

0.00
1.05
0.57
1.20
0.00
0.00
0.00

320.00
13.70
11.60
1.90
54.70
25.50
171.00

=CH2
=CH2
=CH2 O
=CH4
=CH3 O
=CH3
=CH2 O
=CH3 O
=CH4

+H +M
+H +M
+H

6.90 1014
1.00 1016
8.43 1013
1.93 1036
2.26 1014
4.75 1016
3.30 1011
1.80 1013
3.60 1012

0.00
0.00
0.00
7.00
0.00
0.13
0.00
0.00
0.00

345.00
379.00
0.00
38.00
64.80
88.00
37.40
0.00
0.00

+H
+OH
+OH
+OH
+O2

5.4 Elementary Reactions and Reaction Molecularity


Table 5.1: (continued)
Reaction
CH3
+CH3
CH3
+CH3
15a. CH3 O Reactions
CH3 O
+M
CH3 O
+H
CH3 O
+O2
CH2 O
+CH3 O
CH3 OH
+CHO
CH3 O
+O
CH3 O
+O
15b. CH2 OH Reactions
CH2 OH
+M
CH2 OH
+H
CH2 OH
+O2
16. CH3 O2 Reactions
CH3 O2
+M
CH3
+O2 +M
CH3 O2
+CH2 O
CH3 O2 H
+CHO
CH3 O2
+CH3 O
CH3 O
+CH3 O
CH3 O2
+HO2
CH3 O2 H
+O2
CH3 O2
+CH3 O2
CH3 O2
+CH3 O2
17. CH4 Reactions
CH4
+H
CH4
+O
CH4
+OH
CH4
+HO2
CH4
+CH
CH4
+CH2
18. CH3 OH Reactions
CH3 OH
CH3 OH
+H
CH3 OH
+O
CH3 OH
+OH
CH3 OH
+HO2
CH2 OH
+H2 O2
CH3 OH
+CH3
CH3 O
+CH3 OH
CH2 OH
+CH3 OH
CH3 OH
+CH2 O
CH3 O
+CH3 O

A
b
[cm,mol,s]
1.00 1016
0.00
1.69 1053 12.00

E
[kJ/mol]
134.00
81.20

=C2 H4
=C2 H6

+H2

=CH2 O
=C2 H6
=C2 H6
=CHO
=CH2 O
=O2
=OH

+H +M
+H2 OH
+HO2
+CH3 OH
+CH3 O
+CH3
+CH2 O

5.00 1013
1.80 1013
4.00 1010
6.00 1011
6.50 1009
1.10 1013
1.40 1012

0.00
0.00
0.00
0.00
0.00
0.00
0.00

105.00
0.00
8.90
13.80
57.20
0.00
0.00

=CH2 O
=CH2 O
=CH2 O

+H +M
+H2
+HO2

5.00 1013
3.00 1013
1.00 1013

0.00
0.00
0.00

105.00
0.00
30.00

=CH3
=CH3 O2
=CHO
=CH3 O2
=CH3 O
=CH3 O2
=CH3 O2 H
=CH3 O2
=CH2 O
=CH3 O

+O2 +M
+M
+CH3 O2 H
+CH2 O
+CH3 O
+CH3
+O2
+HO2
+O2 +CH3 OH
+CH3 O +O2

7.24 1016
1.41 1016
1.30 1011
2.50 1010
3.80 1012
2.00 1010
4.60 1010
3.00 1012
1.80 1012
3.70 1012

0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00

111.00
30.00
37.70
42.30
5.00
0.00
10.90
163.00
0.00
9.20

=H2
=OH
=H2 O
=H2 O2
=C2 H4
=CH3

+CH3
+CH3
+CH3
+CH3
+H
+CH3

1.30 1004
6.92 1008
1.60 1007
1.10 1013
3.00 1013
1.30 1013

3.00
1.56
1.83
0.00
0.00
0.00

33.60
35.50
11.60
103.00
1.70
39.90

=CH3
=CH2 OH
=CH2 OH
=CH2 OH
=CH2 OH
=HO2
=CH4
=CH2 OH
=CH3 O
=CH3 O
=CH3 OH

+OH
+H2
+OH
+H2 O
+H2 O2
+CH3 OH
+CH2 OH
+CH3 OH
+CH3 OH
+CH3 O
+CH2 O

9.51 1029
4.00 1013
1.00 1013
1.00 1013
6.20 1012
1.00 1007
9.00 1012
2.00 1011
2.20 1004
1.53 1012
3.00 1013

4.30
0.00
0.00
0.00
0.00
1.70
0.00
0.00
1.70
0.00
0.00

404.00
25.50
19.60
7.10
81.10
47.90
41.10
29.30
45.40
333.00
0.00

235

5 Elements of Chemical Kinetics


Table 5.1: (continued)
Reaction
19. CH3 O2 H Reaction
CH3 O2 H
=CH3 O
+OH
OH
+CH3 O2 H
=H2 O
+CH3 O2
20. 29. C2 -Hydrocarbons Oxidation
20. C2 H Reactions
C2 H
+O
=CO
+CH
C2 H
+O2
=HCCO
+O
21. HCCO Reactions
HCCO
+H
=C2
+CO
HCCO
+O
=CO
+H +CO
HCCO
+CH2
=C2 H3
+CO
22. C2 H2 Reactions
C2 H2
+M
=C2 H
+H +M
C2 H2
+O2
=HCCO
+OH
C2 H2
+H
=C2 H
+H2
C2 H2
+O
=CH2
+CO
C2 H2
+O
=HCCO
+H
C2 H2
+OH
=H2 O
+C2 H
C2 H2
+C2 H
=C4 H2
+H
23. CH2 CO Reaction
CH2 CO
+M
=CH2
+CO +M
CH2 CO
+H
=CH3
+CO
CH2 CO
+O
=CHO
+CHO
CH2 CO
+OH
=CH2 O
+CHO
24. C2 H3 Reactions
C2 H3
=C2 H2
+H
C2 H3
+OH
=C2 H2
+H2 O
C2 H3
+H
=C2 H2
+H2
C2 H3
+O
=C2 H2
+OH
C2 H3
+O
=CH3
+CO
C2 H3
+O
=CHO
+CH2
C2 H3
+O2
=CHO
+CH2 O
25a. CH3 CO Reactions
CH3 CO
=CH3
+CO
CH3 CO
+H
=CH2 CO
+H2
25b. CH3 CHO Reactions
CH2 CHO +H
=CH2 CO
+H2
26. C2 H4 Reactions
C2 H4
+M
=C2 H2
+H2 +M

C2 H4
+M
=C2 H3
+H +M
C2 H4
+H
=C2 H3
+H2
C2 H4
+O
=H
+CH2 CHO
C2 H4
+O
=CHO
+CH3
C2 H4
+OH
=C2 H3
+H2 O

236

A
[cm,mol,s]

E
[kJ/mol]

4.00 1015
2.60 1012

0.00
0.00

180.00
0.00

1.00 1013
3.00 1012

0.00
0.00

0.00
0.00

1.50 1014
9.60 1013
3.00 1013

0.00
0.00
0.00

0.00
0.00
0.00

3.60 1016
2.00 1008
6.02 1013
1.72 1004
1.72 1004
6.00 1013
3.00 1013

0.00
1.50
0.00
2.80
2.80
0.00
0.00

446.00
126.00
116.00
2.10
2.10
54.20
0.00

1.00 1016
3.60 1013
2.30 1012
1.00 1013

0.00
0.00
0.00
0.00

248.00
14.10
5.70
0.00

4.73 1040
5.00 1013
1.20 1013
1.00 1013
1.00 1013
1.00 1013
5.40 1012

8.80
0.00
0.00
0.00
0.00
0.00
0.00

194.00
0.00
0.00
0.00
0.00
0.00
0.00

2.32 1026
2.00 1013

5.00
0.00

75.10
0.00

2.00 1013

0.00

0.00

7.50 1017
8.50 1017
5.67 1014
1.40 1006
2.42 1006
2.11 1013

0.00
0.00
0.00
2.08
2.08
0.00

320.00
404.00
62.90
0.00
0.00
24.90

5.5 Rate of Reactions


Table 5.1: (continued)
Reaction
27. CH3 CHO Reactions
CH3 CHO +M
CH3 CHO +H
CH3 CHO +H
CH3 CHO +O
CH3 CHO +O
CH3 CHO +O2
CH3 CHO +OH
CH3 CHO +HO2
CH3 CHO +CH2
CH3 CHO +CH3
28. C2 H5 Reactions
C2 H5
C2 H5
H
C2 H5
O
C2 H5
O
C2 H5
O2
C2 H5
CH3
C2 H5
C2 H5
29. C2 H6 Reactions
C2 H6
H
C2 H6
O
C2 H6
OH
C2 H6
HO2
C2 H6
O2
C2 H6
CH2
C2 H6
CH3

A
[cm,mol,s]

E
[kJ/mol]

=CH3
=CH3 CO
=H2
=CH3 CO
=OH
=CH3 CO
=CH3 CO
=CH3 CO
=CH3 CO
=CH3 CO

+M
+H2
+CH2 CHO
+OH
+CH2 CHO
+HO2
+H2 O
+H2 O2
+CH3
+CH4

7.00 1015
2.10 1009
2.00 1009
5.00 1012
8.00 1011
4.00 1013
2.30 1010
3.00 1012
2.50 1012
2.00 1006

0.00
1.16
1.16
0.00
0.00
0.00
0.73
0.00
0.00
5.64

343.00
10.10
10.10
7.60
7.60
164.00
4.70
50.00
15.90
10.30

=C2 H4
=CH3
=H
=CH2 O
=C2 H4
=C2 H4
=C2 H4

+H
+CH3
+CH3 CHO
+CH3
+HO2
+CH4
+C2 H6

1.02 1043
3.00 1013
5.00 1013
1.00 1013
1.10 1010
1.14 1012
1.40 1012

9.10
0.00
0.00
0.00
0.00
0.00
0.00

224.00
0.00
0.00
0.00
6.30
0.00
0.00

=C2 H5
=C2 H5
=C2 H5
=C2 H5
=C2 H5
=C2 H5
=C2 H5

+H2
+OH
+H2 O
+H2 O2
+HO2
+CH3
+CH4

1.40 1009
1.00 1009
7.20 1006
1.70 1013
6.00 1013
2.20 1013
1.50 1007

1.50
1.50
2.00
0.00
0.00
0.00
6.00

31.10
24.40
3.60
85.90
217.00
36.30
25.40

5.5 Rate of Reactions


Consider a reaction
A + 2 B 3 C + D

(5.23)

for which, at any instant, the rate of consumption or formation of the reacting
species fulfil the following relationship:
d[D]
1 d[C]
d[A]
1 d[B]
=
=
=
dt
3 dt
dt
2 dt

(5.24)

so there are several rates connected with the reaction. In order to avoid the
problem of having several possible different rates to describe the same reaction

237

5 Elements of Chemical Kinetics


the unique rate of reaction is defined as:
r=

1 d[J]

j dt

(5.25)

where j is the stoichiometric number of j-species with j negative for reactants


and positive for products.

5.5.1 Temperature dependence of rate coefficients


Rate coefficients (k) depend strongly on temperature. This temperature dependence is described by a simple law, named Arrhenius law,


Ea
(5.26)
k = A exp
RT
According to some more recent measurements the dependence of k on temperature
is described with a better accuracy using the following function:


Ea
b
(5.27)
k = A T exp
RT
Table 5.1 lists values of A, b and Ea for many elementary reactions. The coefficient A T b is called the pre-exponential factor. The activation energy Ea
corresponds to an energy barrier which has to be overcome during the reaction,
see Fig. 5.4. The maximum of the activation energy corresponds to the bond
energies in the molecule. The activation energy can be very small, or even zero if
new bonds are formed simultaneously with the breaking of the old bonds.
Fig. 5.5 shows the temperature dependence of some elementary reactions of halogen atoms with molecular hydrogen. This is a typical Arrhenius plot with the
logarithm of the rate coefficient plotted against the reciprocal
of temperature. A

linear dependence is obtained since log k = log A T b RETa and a weak dependence of the term A T b is often obscured by the experimental error. As shown by
Eq. (5.27) for activation energies very low or for very high temperatures, the exponent term goes to one, and the rate constant k goes to A T b . In chemical kinetics
linear extrapolations of Arrhenius plots are common methods used to estimate
rate coefficients at temperatures where they have not been measured directly.
However, Arrhenius plots may also be curved and then both pre-exponential factor and activation energy increase with temperature.

238

(f)

Ea

5.5 Rate of Reactions

(b)

Ea

Potential energy

Ureactants

Uproducts

Extent of reaction
Fig. 5.4: Activation energy of a chemical reaction

In trimolecular reactions listed in Table 5.1, M represents an inert collision


partner called sometimes a third body factor. Although M appears on both
sides of the chemical equation it cannot be left out. Its appearance in the chemical
equation is associated with the role of other (third body) molecules as collisional
energisers and deenergisers. Consider for example the 6th reaction in Table 5.1
O + O + M = O2 + M

(5.28)

If two oxygen atoms collide and form an oxygen molecule it will immediately
dissociate because of the energy excess of the newly formed O2 . The third collision
partner [M ] takes away the energy excess stabilising the O2 molecule. The
concentration of this collision partner is calculated as

[M ] = [H2 ] + 6.5[H2 O] + 0.4[O2 ] + 0.4[N2 ] + 0.75[CO] + 1.5[CO2 ] + 3[CH4 ] (5.29)


showing that each of the stable molecules play a role in the energy removing
process but with a different efficiency (weighting factor).

239

5 Elements of Chemical Kinetics

Fig. 5.5: Arrhenius plot for elementary reactions of halogens with molecular hydrogen
[20]

5.5.2 Pressure dependence of rate coefficients


For a unimolecular reaction, A products, the rate of formation of products
can be written as:
d[products]
= k [A]
(5.30)
dt
and the rate coefficient k depends on pressure and temperature. The theory of
kinetics of unimolecular reactions yields fall-off curves. An example of such a
curve is given in Fig. 5.6.
For bimolecular and trimolecular reactions the dependence of the rate constant
k on temperature is rather complex and the reader is referred to textbooks on
chemical kinetics [18, 16, 19].

240

5.6 Summary

Fig. 5.6: Fall-off curves for the reaction C2 H6 CH3 + CH3 [20]

5.6 Summary
In this chapter a discipline of Chemical Kinetics has been introduced. Principally, Chemical Kinetics is concerned with two issues: with the rate of chemical
reactions and with mechanisms of chemical reactions. Chapter 5 underlines the
basis for calculating the rates of chemical reactions while in the forthcoming Chapter 6 we will deal with combustion mechanisms.
One should become familiar with the equation
d[A]
= k [A]a [B]b [C]c . . .
dt
describing the rate of consumption of A-species in a chemical reaction. The rate
constant k which is strongly temperature dependent, is calculable using Arrhenius
equation
k = A T b exp(

Ea
)
RT

241

5 Elements of Chemical Kinetics


Reaction order and reaction molecularity are inevitable associated with chemical
kinetics and so is the concept of an elementary reaction. First-order, second-order
and third-order reactions have been considered in this chapter.
The concept of the forward and reversible reactions have been introduced. It has
been emphasised that a correctly formulated system of kinetic equations describing the rates of both forward and reversible reactions must lead to the chemical
composition at equilibrium when t . In short, when t chemical kinetics
meets equilibrium thermodynamics and then and only then
Kc =

242

kf
kb

6 Mechanisms of Basic Combustion


Reactions
Contents
6.1
6.2
6.3

Chain Reactions
Combustion of Carbon Monoxide (CO)
Combustion of Hydrogen (H2 )
6.3.1 Simplified ignition mechanism
6.4 Combustion of Methane (CH4 )
6.5 Methods of Solving Chemical Kinetic Rate Equations
6.5.1 Analytical solutions
6.5.2 Numerical Solutions
6.6 Summary

6.1 Chain Reactions


In our calculations of combustion stoichiometry (see Chapter 1) we used an overall
reaction for hydrogen burning
2 H2 + O2 2 H2 O

(6.1)

Even such a simple overall reaction is a chain process, in which O, H and OH


radicals are active centres. Table 5.1 lists a number of elementary reactions taking
place in combustion of hydrogen (H2 ) carbon monoxide (CO), methane (CH4 )
and C2 -hydrocarbons (ethane C2 H6 ; ethene C2 H4 ; ethyne C2 H2 ). In most
instances, two reacting molecules do not react directly but one of them dissociates
first to form radicals. These radicals, that are very reactive, initiate a chain of
steps (chain of elementary reactions).
Consider reactions taking place in H2 /O2 system (reactions listed in Table 5.1
in groups 01 to 03). Reaction (6.1) appears nowhere in the table since this is

243

6 Mechanisms of Basic Combustion Reactions


an overall reaction and not an elementary reaction. Two hydrogen molecules do
not react directly with an oxygen molecule to form two water molecules. Instead
either hydrogen molecule or oxygen molecule dissociate forming radicals, so
O2 + M O + O + M

(6.2)

H2 + M H + H + M

(6.3)

or
Reactions (6.2) and (6.3) are chain initiation steps (reactions). In chain
initiation reactions radicals (reactive species) are formed from stable species. The
newly formed O and H radicals attack the oxygen molecule and the hydrogen
molecule according to the reactions:
O + H2 OH + H

(6.4)

H + O2 OH + O

(6.5)

and
Both above reactions are chain branching steps (reactions). In chain branching reactions a reactive species (radical) react with a stable species (molecule)
forming two reactive species e.g. two radicals. Table 5.1 shows that radicals H,
O and OH take part in so called chain propagation reactions, for example:
OH + H2 H2 O + H

(6.6)

CO + OH CO2 + H

(6.7)

H2 O2 + H OH + H2 O

(6.8)

or
or
In chain propagation reaction a reactive species (a radical) reacts with a stable
species forming another reactive species as exemplified in reactions (6.6), (6.7)
and (6.8). In chain termination reactions, reactive species react to stable
species and the chain is terminated:
H + O2 + M HO2 + M

(6.9)

HO2 + OH H2 O + O2

(6.10)

or

244

6.2 Combustion of Carbon Monoxide (CO)


or
O + O + M O2 + M

(6.11)

6.2 Combustion of Carbon Monoxide (CO)


Consider oxidation of carbon monoxide that can be described using the "overall"
equation
2 CO + O2 2 CO2
(6.12)
In a reacting system containing only CO and O2 , without presence of any water
molecules, the only elementary reactions leading to CO2 are (see Table 5.1):
(6.13)

CO + O2 CO2 + O

CO + O + M CO2 + M

(6.14)

where reaction (6.13) is chain initiation step and reaction (6.14) is chain termination.
Using the above mechanism (reactions (6.13) and (6.14)), one may write appropriate rate equations and calculate the concentrations of all the species taking
part in the reaction as a function of time, until equilibrium is reached. However,
the rate of CO oxidation is substantially affected by the presence of OH radicals.
Thus, in the presence of hydrogen containing molecules another reaction
CO + OH CO2 + H

(6.15)

opens up a so called "wet" route of CO oxidation. As it is easy to see, any


elementary reactions leading to formation and destruction of hydroxyl radicals
(OH) may be important. In the presence of water vapour the reaction
H2 O + M H + OH + M

(6.16)

is important since, by producing H and OH radicals, it opens up new possibilities


of OH radicals formation and destruction (see the first three groups of reactions
in Table 5.1). In combustion of hydrocarbons almost all CO is oxidised by OH
radicals. Fig. 6.1 shows the rate constants of reaction (6.15) and its backward
counterpart.

245

6 Mechanisms of Basic Combustion Reactions

Fig. 6.1: Top rate constant of CO + OH CO2 + H;


Bottom rate constant of H + CO2 CO + OH [19].

6.3 Combustion of Hydrogen (H2)


A complete mechanism of combustion of hydrogen with oxygen is listed in Table 5.1 in groups 0103. The mechanism is relatively simple [22]. There are only
eight chemical species (O2 , H2 , H, OH, O, H2 O, HO2 , H2 O2 ) and nineteen elementary reactions involved. It is possible to write down rate equations (5.22)

246

6.3 Combustion of Hydrogen (H2 )


for each species. Using special numerical procedures it is possible to solve the
equations and by doing so obtaining information about the change of the concentration of each species in time. In this paragraph we intend to analyse only a few
steps of the mechanism indicating the most important reactions.
The chain initiation reactions are:
O2 + M O + O + M

(6.17)

with its backwards counterpart:


O + O + M O2 + M

(6.18)

H2 + M H + H + M

(6.19)

H + H + M H2 + M

(6.20)

and
with
The rate constants and the activation energies for these chain initialization reactions are given in Table 6.1. The rate constants and the activation energies
of O2 and H2 dissociation reactions are similar and so are the rate constants for
the O-radicals and H-radicals recombination reactions. Thus, one may anticipate
that for thesame initial concentrations of H2 and O2 both reactions will produce
H- and O-radicals in proximately equal amounts and with equal rates.

Table 6.1: Coefficients of the rate constant equation k = A T b exp REaT for the chain
initiation reactions.
[M ] = [H2 ] + 13
[ H2 O] + 52 [ O2 ] + 25 [ N2 ] + 43 [ CO] + 32 [ CO2 ] + 3 [ CH4 ]
2

Reaction
O2 + M O + O + M
O + O + M O2 + M
H2 + M H + H + M
H + H + M H2 + M

A [cm,mol,s]
3.48 1014
2.9 1017
6.19 1014
1.8 1018

b
0
-1
0
-1

Ea [kJ/mol]
451
0.0
402
0.0

After H- and O-radicals are formed through the chain initialization reactions they
attack oxygen molecules and hydrogen molecules, respectively through the chain
branching reactions:
O2 + H OH + O
(6.21)
and
H2 + O OH + H

(6.22)

247

6 Mechanisms of Basic Combustion Reactions


Table 6.2 lists the rate coefficients for the chain branching reactions and their
backward counterparts. It is important to notice that the activation energies for
the chain branching reactions are substantially lower than these for the chain
initialization reactions of dissociation of hydrogen and oxygen molecules. There
are no direct measurements of the fourth reaction listed in Table 6.2. However, it
is possible to estimate the rate of this reaction knowing the equilibrium constant
for the third reaction in Table 6.2.
Table 6.2: Rate coefficients of the chain branchingreactions. The coefficients are presented in the form k = A T b exp REaT .

Reaction
O2 + H OH + O
OH + O H + O2
H2 + O OH + H
OH + H O + H2

A [cm,mol,s]
2.0 1014
1.8 1013
5.06 1014
?

b
0
0
2.67
?

Ea [kJ/mol]
70.3
0.0
26.3
?

It is instructive to compare the rate coefficients of the chain initialization reactions (Table 6.1) with the rates of the chain branching reactions(Table 6.2).
Table 6.3 shows such a comparison for several temperatures. It is easy to see
that the chain branching reactions are overwhelmingly faster than the dissociation reactions. Furthermore H-radical consumed in reaction (6.21) is produced in
reaction (6.22) and vice versa O-radical consumed in reaction (6.22) is regenerated
in reaction (6.21). Consequently, the most important radical in the initial stages
of combustion of hydrogen molecules is the hydroxyl (OH) radical which is produced in reactions (6.21) and (6.22). Therefore instead using reactions (6.17) and
(6.19) as the chain initiation reactions, the sum of reactions (6.21) and (6.22)
which can be written as:
O2 + H2 2 OH
(6.23)
is often used, in the literature on gas reactions kinetics, as the chain initiation
reaction. Reaction (6.23) cannot be found in Table 5.1 since it is an overall
reaction and not an elementary reaction. Reaction (6.23) allows to simplify the
mechanism and ignore unimportant hydrogen and oxygen dissociation reactions
(they might be of some importance at very high temperatures).

6.3.1 Simplified ignition mechanism


Following the considerations of the previous paragraph and using Table 5.1, we
may derive a simplified chain mechanism for ignition of the hydrogen-oxygen
system. The most important reactions are shown in Table 6.4 [22, 23, 28].

248

6.3 Combustion of Hydrogen (H2 )


Table 6.3: Comparison of the rate coefficients k of the chain initiation and chain
3
branching reactions. The
 k values, given in cm /mol s, are calculated
from k = A T b exp REaT for specified temperatures.

Reaction
O2 + M O + O + M
H2 + M H + H + M
O2 + H OH + O
H2 + O OH + H

500 K
2.7 1033
6.2 1028
9.0 106
1.5 109

1000 K
9.6 1010
6.2 107
4.3 1010
2.2 1011

2000 K
5.8 102
2.0 104
2.9 1012
6.8 1012

3000 K
4.9 106
6.2 107
1.2 1013
3.4 1013

Table 6.4: Simplified mechanism of ignition of the hydrogen-oxygen system

(0)

H 2 + O2

2 OH

(1)
(2)
(3)
(4)
(1+2+3)

OH + H2
H + O2
O + H2
H + O2 + M
2 H 2 + O2

=
=
=
=
=

H2 O + H
OH + O
OH + H
HO2 + M
H2 O + OH + H

chain
initiation
(reaction rate R)
chain propagation
chain branching
chain branching
chain termination

If one sums up the chain propagation and chain branching reactions (1+2+3), one
can see that H and OH radicals are produced from the reactants. The chain reactions go on until no radicals are produced any more. The radicals are completely
consumed in chain termination reactions. We can write down the rate laws for
the three radicals:
d[H]
= k1 [OH] [H2 ] k2 [H] [O2 ] + k3 [O] [H2 ] k4 [H] [O2 ] [M ]
dt
d[OH]
= 2 R k1 [OH] [H2 ] + k2 [O2 ] [H] + k3 [O] [H2 ]
dt
d[O]
= k2 [H] [O2 ] k3 [O] [H2 ]
dt

(6.24)
(6.25)
(6.26)

These three radicals, H, OH and O are chain carriers and the term 2 R is called
the rate of the chain initiation. By summing up the three equations we obtain
the rate of formation of the free valences (H, OH, and O). Oxygen atoms however
should be counted twice since there are two free valences per oxygen radical,
d([H] + [OH] + 2 [O])
= 2 R + (2 k2 [O2 ] k4 [O2 ] [M ]) [H]
dt

(6.27)

249

6 Mechanisms of Basic Combustion Reactions


To simplify the notation, we mark the concentration of the free valences as [n], so
(6.28)

[n] [H] + [OH] + 2 [O]

As a crude approximation we may replace the concentration of the hydrogen


radicals as one fourth of the concentration of the free valences, so
1
[H] [n]
4

(6.29)

With the above introduced simplifications, Eq. (6.27) can be written as


1
d[n]
= 2 R + (2 k2 [O2 ] k4 [O2 ] [M ]) [n]
dt
4

(6.30)

and the dependence of [n] with time can be easily obtained as


[n] = 2 R t

for g = f

(6.31)




2R
1
[n] = 1
(f g) t
1 exp
4
4 (g f )

for g 6= f

(6.32)

where f = 2 k2 [O2 ] and g = k4 [O2 ] [M ]. The solution (Eqs. (6.31) and (6.32))
is shown schematically in Fig. 6.2.
For g > f the exponential term in Eq. (6.32) tends to zero with increasing time.
For time being large enough, a time-independent solution is obtained for the chain
carriers
8R
(6.33)
[n] =
(g f )
When g = f a linear increase of the chain carriers concentration is obtained. For
g < f , after a short time, the exponential term in Eq. (6.32) is larger than one
and the result is an exponential growth of the carriers concentration


1
8R
exp
(f g)
(6.34)
[n] =
(g f )
4
thus an explosion is observed.
It should be realised that the ignition scheme shown in Table 6.4 is a simplification
of the complete mechanism listed in Table 5.1. The scheme has been used to
illustrate the importance of H, O, OH radicals in the initiation of the hydrogen
combustion. Having at the disposal the complete kinetic scheme (Table 5.1) and
a general numerical procedure for solving ordinary (stiff) differential equations,

250

6.4 Combustion of Methane (CH4 )


it is possible to examine whether the combustion is extinct, propagates through
the mixture or an explosion takes place.

Fig. 6.2: The time behaviour of the chain carriers (Eqs. (6.31) and (6.32))

6.4 Combustion of Methane (CH4)


Combustion of hydrocarbons is a complicated chemical process. Methane is the
most common hydrocarbon and it is the basic component of natural gas. Energies
of the CH bonds in methane exceed those of other hydrocarbons and therefore
methane is the least reactive of all hydrocarbons. Table 5.1, groups 1019, lists
the most important elementary reactions for methane oxidation. Since a CH4 molecule contains only one atom of carbon, no aldehyde groups (RCHO) are
formed in contrast to elementary reactions for oxidation of C2 -hydrocarbons (see
Table 5.1 groups 2029).
Methane molecules are first attacked by H, O and OH radicals [19, 23]
CH4 + H CH3 + H2

(6.35)

CH4 + O CH3 + OH

(6.36)

CH4 + OH CH3 + H2 O

(6.37)

Table 6.5 lists the rate constants of these reactions for several temperatures. All
these reactions proceed with similar rates and therefore they are of equal importance. The methyl radical (CH3 ) is formed as a result of these chain propagation

251

6 Mechanisms of Basic Combustion Reactions


reactions. The methyl radical is almost exclusively removed by a reaction with
oxygen atoms
CH3 + O CH2 O + H
(6.38)

Table 6.5: Comparison of the rate coefficients k of the chain propagation reactions for
methane oxidation. The k values, given in cm3 /mol s, are calculated from
k = A T b exp REaT for specified temperatures.

Reaction
CH4 + H CH3 + H2
CH4 + O CH3 + OH
CH4 + OH CH3 + H2 O

500 K
5.0 108
3.6 109
8.5 1010

1000 K
2.28 1011
5.9 1011
1.2 1012

2000 K
1.4 1013
1.2 1013
8.8 1012

3000 K
9.1 1013
4.4 1013
2.3 1013

The formaldehyde (CH2 O) is destroyed again by reactions with the radicals:


CH2 O + OH CHO + H2 O

(6.39)

CH2 O + H CHO + H2

(6.40)

CH2 O + O CHO + OH

(6.41)

The CHO radical is converted to CO in the following reactions:


(6.42)

CHO + O2 CO + HO2

CHO + M CO + H + M
CHO + H CO + H2

(6.43)
(6.44)

The conversion of methane into CH3 via reaction (6.37) competes with the basic
reaction of CO oxidation (reaction (6.15)) as soon as carbon monoxide is produced.
Reaction (6.37) is around one order faster than reaction (6.15) and combustion
of methane, as with combustion of other hydrocarbons, proceeds in two distinct
steps. Firstly a rapid conversion of CH4 into CO, H2 , H2 O and free radicals
take place. Secondly, a relatively slow after-burning of CO takes place. For this
second step, the mechanisms of CO and H2 oxidations discussed previously are
valid. Fig. 6.3 shows a simplified scheme of methane oxidation. In Chapter 1 on
combustion stoichiometry calculations we used an overall reaction for combustion
of methane
CH4 + 2 O2 CO2 + 2 H2 O
(6.45)
In light of the above considerations a better overall reaction scheme for methane

252

6.5 Methods of Solving Chemical Kinetic Rate Equations


oxidation would be
CH4 + O2 CO + H2 + H2 O
CO +
H2 +

1
2 O2
1
2 O2

(6.46)

CO2

(6.47)

H2 O

(6.48)

or perhaps reaction (6.46) could be combined with a radical mechanism of COH2


oxidation.

Fig. 6.3: A simplified scheme of methane oxidation [19]

6.5 Methods of Solving Chemical Kinetic Rate


Equations
Complete mechanisms for oxidation of hydrocarbons may consist of several hundred elementary reactions with as many as a hundred species. On the other

253

6 Mechanisms of Basic Combustion Reactions


hand, a simple mechanism of CO oxidation, in water-free environment, consists
of two reactions with only four species involved. In order to determine how the
concentration of each species changes with time we formulate a set of ordinary
differential equations of the type
d[Ai ]
= fi ([A1 ], [A2 ], . . . , [AN ]; k1 , k2 , . . . , kR )
dt

for i = 1, 2, . . . , N

(6.49)

with initial conditions


[Ai ](t = t0 ) = [Ai ]0

(6.50)

We write down the rate law (Eq. (6.49)) for each species of the mechanism considered, thus N stands for a number of species while R represents a number of
elementary reactions. In above equations the time is the independent variable,
the species concentrations are the dependent variables while ki are parameters of
the system and [Ai ]0 denote the initial conditions at time t0 . Eqs. (6.24), (6.25)
and (6.26) are such rate laws describing the concentrations of the three radicals
for the simplified ignition mechanism of hydrogen. We assume that the temperature, needed to calculate the rate coefficients ki is given, thus Eqs. (6.49) and
(6.50) are valid for an isothermal system (we can easily include an energy balance
equation to calculate the temperature as a function of time, if desired).
We wish to possess a general mathematical procedure to solve the system (6.49)
with the initial conditions (6.50) for any temperature and pressure. In some
specific cases, when a number of species is perhaps not larger than three or four,
we may attempt to solve the system analytically. However, we expect that the
general solver is numerical.

6.5.1 Analytical solutions


In Chapter 5 (Examples 5.1 and 5.2) we have derived solutions to the rate law
equations for a first-order, a second-order or generally a nth -order (n-integer) irreversible reaction (the solution for a zero-order reaction may be easily found).
Consider a simple reaction chain consisting of two elementary steps; a chain initiation and chain termination
k

A1 12
A2 23
A3
The rate laws for the three species A1 , A2 and A3 are

254

(6.51)

6.5 Methods of Solving Chemical Kinetic Rate Equations

d[A1 ]
= k12 [A1 ]
dt
d[A2 ]
= k12 [A1 ] k23 [A2 ]
dt
d[A3 ]
= k23 [A2 ]
dt

(6.52)
(6.53)
(6.54)

At time zero the only component is A1 so the initial conditions are: [A1 ](0) =
[A1 ]0 ; [A2 ](0) = 0 and [A3 ](0) = 0. The solution of Eq. (6.52) is easy to find
(6.55)

[A1 ](t) = [A1 ]0 exp(k12 t)


Inserting (6.55) into (6.53) leads to the following equation for [A2 ]
d[A2 ]
+ k23 [A2 ] = k12 [A1 ]0 exp(k12 t)
dt

(6.56)

Eq. (6.56) is an ordinary differential equation of a type


y + k23 y = r(t)

(6.57)

which we have considered in ChapterR 5. The solution is found upon multiplying


Eq. (6.56) by the integrating factor e k23 dt = ek23 t so
d[A2 ] k23 t
e
+ k23 ek23 t [A2 ] = k12 [A1 ]0 exp(k12 t) ek23 t
dt

d 
[A2 ] ek23 t = k12 [A1 ]0 exp(k12 t) ek23 t
dt

Integration of the above equation leads to


Z
k23 t
[A2 ] e
= k12 [A1 ]0 exp(k23 t) ek23 t dt + C1

(6.58)
(6.59)

(6.60)

and after some algebra


[A2 ](t) =

k12 [A1 ]0

and
[A2 ](t) =

e(k12 +k23 )t dt

ek23 t

C1
ek23 t

k12 [A1 ]0 e(k23 k12 )t


C1
+
k
t
(k23 k12 ) e 23
ek23 t

(6.61)

(6.62)

255

6 Mechanisms of Basic Combustion Reactions


Using the initial condition [A2 ](0) = 0 we can calculate the integration constant
C1 as
k12 [A1 ]0
(6.63)
C1 =
(k23 k12 )
so finally, the solution for the species A2 reads:
[A2 ](t) =

i
k12 [A1 ]0 h k23 t
e
ek12 t
k12 k23

(6.64)

Knowing the dependence of A2 -species concentration with time, Eq. (6.64), we


may easily calculate the dependence of A3 -species with time. To this end we
insert Eq. (6.64) into (6.54), and we perform the integration obtaining
Z
i
k12 k23 [A1 ]0 h k23 t
[A3 ](t) =
e
ek12 t dt + C2
(6.65)
(k12 k23 )
 k12 t

e
ek23 t
k12 k23 [A1 ]0

[A3 ](t) =
+ C2
(6.66)
(k12 k23 )
k12
k23
We obtain C2 integration constant using the initial condition [A3 ](0) = 0, so
(6.67)

C2 = [A1 ]0
and after some algebra

[A3 ](t) = [A1 ]0 1 +

k12
k23
ek12 t
ek23 t
k12 k23
k12 k23

(6.68)

Thus, the solution to the problem formulated by Eqs. (6.51)(6.54) is given by


formula (6.55) for A1 -species, formula (6.64) for A2 -species, and formula (6.68)
for A3 -species.
Fig. 6.4 shows the temporal behaviour of the concentrations of species A1 , A2
and A3 in reaction (6.51) for k12 = 1 s1 and k23 = 10 s1 . In this case, when
the k23 k12 the concentration of the intermediate A2 -species remains low. This
situation corresponds to A2 -species being very reactive and therefore its rate of
consumption is approximately equal to its rate of formation. Fig. 6.5 shows the
temporal behaviour of the concentration of the three species for k12 = 10 s1
and k23 = 1 s1 . This situation corresponds to a low reactive intermediate A2 species which, as shown in Fig. 6.5, is present in rather large quantities and its
concentration varies with time.

256

6.5 Methods of Solving Chemical Kinetic Rate Equations

Fig. 6.4: Temporal behaviour of the species concentrations in reactions A1 A2


A3 for k12 = 1 s1 and k23 = 10 s1 (reactive intermediate A2 )

Fig. 6.5: Temporal behaviour of the species concentrations in reactions A1 A2


A3 for k12 = 10 s1 and k23 = 1 s1 (low reactive intermediate A2 ).

257

6 Mechanisms of Basic Combustion Reactions


Example 6.1
Consider again a simple reaction system given be Eqs. (6.51)(6.54). Derive
appropriate formulae for calculating the concentrations of the A1 , A2 and A3
species as a function of time for k12 = k23 = k. The initial conditions are
again: [A1 ](0) = [A1 ]0 ; [A2 ](0) = 0; [A3 ](0) = 0. We have already derived
formulae (6.55), (6.64) and (6.68) for calculation the species concentration as a
function of time. Relationship (6.55) can be used putting k12 = k, so
[A1 ](t) = [A1 ]0 ekt

(E1)

However in formulae (6.64) and (6.68) singularity appears for k12 = k23 = k.
Thus, we have to step back to Eq. (6.61) that for k12 = k23 = k gives:
[A2 ](t) =

k [A1 ]0 t C1
+ kt
ekt
e

Since [A2 ](0) = 0, C1 -constant is also equal zero and the formula for the concentration of A2 -species is
[A2 ](t) = k [A1 ]0 t ekt
(E2)
Inserting the above relationship into Eq. (6.54) gives:
d[A3 ]
= k 2 [A1 ]0 t ekt
dt
and
2

[A3 ](t) = k [A1 ]0

tekt dt + C2

While performing the integration we use the integration by parts method


Z
Z
u(x)v (x) dx = uv u (x)v(x) dx
to obtain

Z kt 
e
tekt
+
dt + C2 =
[A3 ](t) = k [A1 ]0
k
k
2

k 2 [A1 ]0



(tk + 1) ekt
+ C2

k2

Using the initial condition [A3 ](0) = 0 we calculate the integration constant to
be C2 = [A1 ]0 and therefore the formula for calculating the concentration of

258

6.5 Methods of Solving Chemical Kinetic Rate Equations


A3 -species reads:

[A3 ](t) = [A1 ]0 1 (kt + 1) e

kt

(E3)

Fig. 6.6 shows the temporal behaviour of the concentrations of species A1 , A2 and
A3 as predicted using Eqs. (E1), (E2) and (E3) valid for k12 = k23 = k.

Fig. 6.6: Temporal behaviour of the species concentrations for reactions A1


A2 A3 with k12 = k23 = k.

End of Example 6.1


6.5.1.1 Quasi-Steady State Assumption
In engineering practise we are often interested in the change of the substrates
and products with time. The intermediate species are of lesser importance and
if possible we would like to omit them in our considerations. Consider again
the reaction sequence (6.51) where A1 represents substrates, A3 products and A2
stands for intermediates. Fig. 6.4 and Fig. 6.5 show the temporal behaviour of the
three components in the reaction sequence (6.51) for k23 k12 and k12 k23 ,
respectively. In the later case (Fig. 6.5), the concentration of the A2 -intermediate
reaches relatively high values and what is even more important, it varies with time
as rapidly as the concentration of substrates A1 or products A3 . Mathematically,
2]
the derivative d[A
dt is far from zero throughout the course of the reaction and it is

259

6 Mechanisms of Basic Combustion Reactions


not possible to omit this species. In the former case (Fig. 6.4) the concentration
of A2 -intermediates is low (since it is a reactive intermediate) and it does not
2]
vary too much with time. Mathematically, we can say that the derivative d[A
dt is
almost zero, thus
d[A2 ]
= k12 [A1 ] k23 [A2 ] 0
(6.69)
dt
With this simplification the rate equations describing our system read:
[A1 ](t) = [A1 ]0 exp(k12 t)

(6.70)

k23 [A2 ] = k12 [A1 ]

(6.71)

[A3 ](t) = k23 [A2 ] = k12 [A1 ] = k12 [A1 ]0 exp(k12 t)

(6.72)

or after integration of Eq. (6.72)


[A3 ](t) = k12 [A1 ]0 (1 exp(k12 t))

(6.73)

Fig. 6.7: Temporal behaviour of the species concentrations for reactions A1


A2 A3 assuming quasi steady state for A2 -species; k12 = 1 s1 , k23 =
10 s1 . (to be compared with Fig. 6.4)

Fig. 6.7 shows the temporal behaviour (Eqs. (6.70), (6.71) and (6.73)) obtained
using the quasi-steady state assumption for A2 -intermediate for k12 = 1 s1 and
k23 = 10 s1 . It is easy to see that the simplified solution (Eqs. (6.70), (6.71) and

260

6.5 Methods of Solving Chemical Kinetic Rate Equations


(6.73)) is only marginally different from the exact solution (Eqs. (6.55), (6.64),
(6.68)) shown in Fig. 6.4. This remains to be so as long as k23 k12 .
We consider formation of nitric oxide (NO) in order to provide an example of a
steady-state approximation. In 1946 Zeldovich postulated the mechanism of NO
formation by three elementary reactions [22, 23, 20]:
O + N2 NO + N

(6.74)

N + O2 NO + O

(6.75)

N + OH NO + H

(6.76)

The first two reactions (6.74) and (6.75) are called Zeldovich mechanism and
here the role of hydroxyl radicals is neglected. All the three reactions are named
extended Zeldovich mechanism. Table 6.6 lists the rate constants and the
rates of the reactions at several temperatures. Reaction (6.74) is substantially
slower than reactions (6.75) and (6.76). Its activation energy of 316 kJ/mol is
very high since there is a strong triple bond in the N2 -molecule to break. Thus,
this reaction may be of some importance at high temperatures only and this is
why the mechanism is named as thermal mechanism of NO formation. The
rate of NO formation for the extended Zeldovich mechanism is
d[NO]
= k1 [O] [N2 ] + k2 [N] [O2 ] + k3 [N] [OH]
dt

(6.77)

where k1 , k2 and k3 are the rate constants of reactions (6.74), (6.75), (6.76),
respectively. These are given in Table 6.6.
Table 6.6: Rate coefficients of the Zeldovich mechanism
 reactions [25]. The rate coefficients are in the form k = A T b exp REaT .

Reaction

O + N2 NO + N
N + O2 NO + O
N + OH NO + H

A
cm3 /mol/s
7.6 1013
6.4 109
1.0 1014

b
0
1
0

E
kJ/mol
316
26.2
0

500 K

1000 K

2000 K

7.4 1020 2.4 1003 4.2 1005


5.9 1009 2.7 1011 2.6 1012
1.0 1014 1.0 1014 1.0 1014

3000 K

2.4 108
6.7 1012
1.0 1014

Table 6.6 shows that reaction (6.74) is the rate controlling step of the thermal
NO-formation mechanism. Consider the nitrogen radicals. They are formed in
reaction (6.74) with a relatively low rate. However, they are rapidly removed
in reactions (6.75) and (6.76). Here we come across a situation analogue to the
reaction sequence A1 A2 A3 with k23 k12 . We can assume the steady
state approximation for the N-radicals. The rate formation of N-radicals for the

261

6 Mechanisms of Basic Combustion Reactions


extended Zeldovich mechanism reads:
d[N]
= k1 [O] [N2 ] k2 [N] [O2 ] k3 [N] [OH] 0
dt

(6.78)

]
and the derivative d[N
dt is almost zero. Inserting Eq. (6.78) into (6.77) results in
a simple expression for the thermal NO-formation rate:

d[NO]
= 2 k1 [O] [N2 ]
dt

(6.79)

The equation shows that the important factors affecting the NO-formation rate
are: the temperature (through k1 ), the O-radicals concentration and N2 -concentration.
6.5.1.2 Partial Equilibrium
Relationship (6.79) can be used to calculate the NO concentration as a function
of time. However, in order to proceed with the integration we must know (in addition to given initial conditions) the temperature (to calculate k1 ), the O-radical
concentration and the N2 -concentration at any instant. Alternatively we may
write down a reaction mechanism for O-radicals formation and destruction and
formulate appropriate rate equations. However, it is clear that numerical integration would be needed to solve such a system. Partial equilibrium assumption
offers a mathematically simpler method.
We will apply the partial equilibrium concept to solve Eq. (6.79). In order to
estimate the O-radicals concentration we assume that they are in equilibrium
with O2 -molecules

O2 + M
(6.80)

O+O+M
so
Kc =
and therefore

[O]2
[O2 ]

(6.81)

q
d[NO]
(6.82)
= 2 k1 Kc [O2 ] [N2 ]
dt
Using the rate coefficients of the O2 -dissociation reactions (Table 6.1) and the
rate coefficient of reaction (6.75) for NO-formations (Table 6.6) it is easy to verify
that at high temperatures the equilibrium between the molecular oxygen and oxygen atoms is reached quickly if compared to the progress of the rate-controlling
reaction (6.74). Thus, the assumption of the partial equilibrium in reaction (6.80)

262

6.5 Methods of Solving Chemical Kinetic Rate Equations


seems to be justified. However, it has been observed that the equilibrium, particularly at low pressures, under-predicts [O] by a factor of up to five. This is
particularly true at flame fronts where oxygen atoms are generated by many other
routes and are not at all at equilibrium.
A better approximation for [O] radicals can be made assuming the equilibrium of
the following elementary reactions:

H + O2

OH + O

O + H2

OH + H

(A)

(6.83)

(B)

(6.84)

OH + H2

H2 O + H

(C)

(6.85)

Thus, at equilibrium the following is applicable:


[OH] [O]
[H] [O2 ]
[OH] [H]
=
[O] [H2 ]
[H O] [H]
= 2
[OH] [H2 ]

Kc,A =

(6.86)

Kc,B

(6.87)

Kc,C

(6.88)

where Kc,A , Kc,B and Kc,C are the equilibrium constants of reactions (6.83), (6.84)
and (6.85), respectively. Using these equations one can express the concentrations
of the intermediates (O, H, OH) using the concentrations of the stable species:
[O] = Kc,A Kc,C

[O2 ] [H2 ]
[H2 O]

(6.89)
1

[O ] 2 [H2 ] 2
[H] = Kc,A Kc,B Kc,C 2
[H2 O]
1
2

1
2

1
2

2
2
Kc,B
[O2 ] 2 [H2 ] 2
[OH] = Kc,A

(6.90)
(6.91)

If we now insert Eq. (6.89) into (6.79) we obtain


[O ] [H2 ] [N2 ]
d[NO]
= 2 k1 Kc,A Kc,C 2
dt
[H2 O]

(6.92)

One should note that k1 -rate constant corresponds to reaction (6.74) while the
equilibrium constants Kc,A and Kc,C refer to reactions (6.83) and (6.85). It is
important to stress that the partial equilibrium assumption provides satisfactory results only at high temperatures (typically > 1800 K).

263

6 Mechanisms of Basic Combustion Reactions


Most combustion systems entail oxidation mechanisms containing numerous individual reaction steps. Under certain circumstances a group of reactions may
proceed rapidly and reach a quasi-equilibrium whilst other reactions may proceed
slowly. If the rate constant of this slow reaction is to be determined, the concentrations of these reactive intermediates (radicals) can be estimated using the
partial equilibrium assumption. Thus, the partial equilibrium assumption looks
like the steady-state approximation we considered earlier. There is however one
essential difference between these approaches; in the steady state approximation
one is concerned with a particular species. In the partial equilibrium approach
one is concerned with particular reactions [23].

6.5.2 Numerical Solutions


We are concerned with solving N coupled ordinary differential Eqs. (6.49) with
initial conditions (6.50). For a number of specific cases we have succeeded in
finding analytical solutions when only a few chemical species are present in the
system. However, we realise that a general solver for the system of Eqs. (6.49)
has to be a numerical one.
6.5.2.1 One Ordinary Differential Equation
Prior to considering a set of coupled ordinary differential equations we will analyse
just one equation:
dy
= f (t, y)
(6.93)
dt
with the initial condition
y(0) = y0
(6.94)
where the function on the right-hand side of Eq. (6.93) is known. Mathematicians
call this problem an initial value problem since y-value is given at a starting
point t0 , and it is desired to find the ys at some final point tf inal or at some
discrete list of points.
The underlying idea of any numerical procedure for solving the initial value problem is always the same. Write the dy and dt in Eq. (6.93) as finite steps y
and t, and multiply Eq. (6.93) by t. In this way we obtain an algebraic formula for the change in the function when the independent variable t is increased
by one time-step t. In the limit of making the time-step very small, a good
approximation to the underlying differential equation is achieved. Literal implementation of this procedure results in Eulers method (see below). However we

264

6.5 Methods of Solving Chemical Kinetic Rate Equations


do not recommend this method for any practical use. The method is conceptually
important since all practical methods originate from this idea.
Eulers explicit and implicit methods
The formula for the Euler method applied to Eq. (6.93) is
yn+1 = yn + h f (yn , tn )

(6.95)

which advances a solution from tn to tn+1 = tn + h, see Fig. 6.8. This method is a
step-by-step method that is, we start from the given y0 and proceed stepwise
computing approximate values of the solution y(t) at the discrete points tn+h .
Formula (6.95) is unsymmetrical since it advances the solution through the timestep h using derivative information (f (yn , tn )) only at the beginning of that time
interval. The method is called explicit (or forward) because the new value
yn+1 is given explicitly in terms of the old value yn . One may wish to evaluate
the derivative at the new value yn+1 obtaining a formula
yn+1 = yn + h f (yn+1 , tn+1 )

(6.96)

for the implicit (or backward) Eulers method. This method requires solving
Eq. (6.96) for yn+1 .

Fig. 6.8: Eulers method

The main reason that Eulers methods are not recommended for practical use
is that there are more accurate and faster methods available that run at the

265

6 Mechanisms of Basic Combustion Reactions


equivalent time-step size. Furthermore, the explicit Eulers scheme may become
unstable if the size of the time step is not small enough (see Example 6.2).
Example 6.2
Consider the rate law for a first order irreversible chemical reaction (see Example 5.1)
d[A1 ]
= k12 [A1 ]
(F1)
dt
and the initial condition [A1 ](0) = [A1 ]0 . Solve the equation using the explicit
Eulers scheme. Vary the size of the time-step h. For the rate constant use
a value k12 = 1 s1 . The explicit Euler scheme for integrating Eq. (F1) with
time-step h is
yn+1 = yn + h f (tn , yn ) = yn k h yn = (1 h k) yn

(F2)

We expect that with n the solution yn+1 = (1 h k)n+1 y0 0, so


< 1 and therefore the size of the time step should be chosen so
1 < (1 hk)
that
2
h<
(F3)
k
The solution of Eq. (F1) using the explicit scheme (F2) for h = 0.3 s is shown in
Table 6.7. Fig. 6.9 shows both the solutions for several values of h and the exact
solution. The numerical solution approaches the exact one when the time-step
h decreases. The numerical solution for h = 1.5 s oscillates around the correct
solution and for t it converges to the correct answer. For h = 2.14 s the
numerical procedure is unstable as predicted by requirement (F3).
Table 6.7: Numerical solution of Eq. (F1) using the explicit scheme (F2) with h = 0.3 s.

n
0
1
2
3
4
5
6
7
8
9

266

tn
0.0
0.6
1.2
1.8
2.4
3.0
3.6
4.2
4.8
5.4

yn
1
0.4
0.16
0.064
0.0256
0.01024
0.004096
0.0016384
0.00065536
0.00026214

6.5 Methods of Solving Chemical Kinetic Rate Equations

Fig. 6.9: Numerical solutions of Eq. (F1) using the explicit scheme (F2) with different
time steps. An instability has been encountered in integrating using a too
large time-step.

Comments:
(a) The method provides an accurate solution only when the time step is sufficiently small.
(b) For a too large time step, the method is unstable (see Eq. (F3)).
(c) The faster the reaction, the smaller the time step required for obtaining an
accurate numerical solution. If this method were applied to reactions listed
2 14
) sec.
in Table 6.5, the time step would have to be smaller than ( 10
End of Example 6.2
Stability and accuracy are important features of numerical schemes. In Examples 6.2 and 6.3 we consider a simple linear equation for which the Eulers
explicit scheme becomes unstable when a too large time-step is used while the
implicit scheme is unconditionally stable. This nice feature of implicit methods
holds only for linear systems but even in the general case implicit methods give
better stability.
Now, we will consider accuracy of the Eulers scheme (6.95). We may expand our
unknown function y(t) around a point tn as suggested by the Taylor series:
y(tn + h) = y(tn ) + h y (tn ) + h2 y (tn ) + . . .

(6.97)

267

6 Mechanisms of Basic Combustion Reactions


For a small value of h, the higher powers of h2 , h3 , . . . are very small so a crude
approximation is
(6.98)
y(tn + h)
= y(tn ) + h y (tn )
which is exactly the same expression as the Eulers explicit formula (6.95). The
Eulers method is a first-order method since in Eq. (6.98) we take only the
constant terms and the term containing the first power of h. The omission of
the further terms causes an error which is called the truncation error of the
method. For small h, the third and higher powers of h are small compared with
h2 in the first neglected term in Eq. (6.97). Therefore we say that the truncation
error per step (or local truncation error) is of the order h2 .
Any arithmetic operation among floating numbers introduces an additional error
called roundoff error that is computer (hardware) dependent (More precisely,
the error depends on how many bits are in the mantissa for the floating point
representation of numbers). Roundoff errors accumulate with increasing amounts
of calculations. As a general rule there is not much that a programmer can do
about roundoff errors, other than to choose algorithms that do not magnify them
unnecessarily.
Example 6.3
Consider again the rate law for a first order irreversible chemical reaction (see
Example 6.2)
d[A1 ]
= k12 [A1 ]
(G1)
dt
and the initial condition [A1 ](0) = [A1 ]0 . Solve the equation using the implicit
Eulers scheme. Vary the size of the time-step h. For the rate constant use a
value k12 = k = 1 s1 .
Assumptions: None
In the implicit Eulers scheme we evaluate the right hand side of Eq. (G1) at the
new yn+1 corresponding to the new time tn+1 so the scheme reads (see Eq. (6.96))
yn+1 = yn + h f (yn+1 , tn+1 ) = yn k h yn+1
and
yn+1 =

yn
(1 + h k)

(G2)
(G3)

It is easy to see that the method is absolutely stable since even at n ,


yn+1 = (1+h y0k)n+1 0 which is in fact the correct solution of Eq. (G1). Thus,
for late times the implicit method converges to the true equilibrium solution even
for large time-steps. The solution of Eq. (G1) using the implicit scheme (G2) for

268

6.5 Methods of Solving Chemical Kinetic Rate Equations


h = 0.3 s is shown in Table 6.8. Fig. 6.10 shows both the solutions for several
values of h and the exact solution. No instability has been encountered in the numerical integration and the numerical solution converges nicely towards the exact
solution. Of course if the time step is large in following the evaluation towards
equilibrium we give up accuracy but we maintain stability.
Table 6.8: Numerical solution of Eq. (G1) using the implicit scheme (G2) with h =
0.3 s.

n
0
1
2
3
4
5

tn
0.0
0.6
1.2
1.8
2.4
3.0

yn
1
0.625
0.390625
0.24414063
0.15258789
0.09536743

n
6
7
8
9
10

tn
3.6
4.2
4.8
5.4
6.0
0

yn
0.05960464
0.0372529
0.02328306
0.01455192
0.00909495

Fig. 6.10: Numerical solutions of Eq. (G1) using the implicit scheme (G2) with different time steps. No instability has been encountered.

Comments:
(a) The method is unconditionally stable.
(b) We control the accuracy of the method by varying the time step.
End of Example 6.3

269

6 Mechanisms of Basic Combustion Reactions


Runge-Kutta Method (of Fourth Order)
In the Eulers schemes (see Eq. (6.95) and (6.96)), the derivative of the unknown
function y(t) is evaluated either at the beginning (explicit scheme) of the time
interval or at the end (implicit scheme). There are many other ways of evaluating
the derivative and therefore there exist a number of methods for integration of
the ordinary differential equations. For details the reader should consult textbooks on numerical methods [26, 27]. By far the most popular and arguable even
most useful is the fourth-order Runge-Kutta method. The algorithm of the
method is as follows:
l1 = h f (tn , yn )


l1
h
l2 = h f tn + , yn +
2
2


h
l2
l3 = h f tn + , yn +
2
2
l4 = h f (tn + h, yn + l3 )
l1 l2 l3 l4
yn+1 = yn + + + + + O(h5 )
6
3
3
6

(6.99)

The Runge-Kutta method treats every time-step in a sequence of five sub-steps


described in the algorithm (6.99). For serious computing an adaptive time stepsize control is essential [26]. The purpose of this adaptive step-size control is to
achieve some predetermined accuracy in the solution with minimum computational effort.
6.5.2.2 Two consecutive elementary reactions
We have already considered (see Section 6.5.1) a simple chain consisting of two
elementary steps; a chain initiation and chain termination
k

A1 12
A2 23
A3

(6.100)

The rate laws for A1 , A2 and A3 -species have been given by Eqs. (6.52), (6.53)
and (6.54). The analytical solutions showing the temporal behaviour of species
A1 , A2 and A3 have been obtained in Section 6.5.1 (see Eq. (6.55), (6.64) and
(6.68)) and they are quoted here for compactness:
[A1 ](t) = [A1 ]0 exp(k12 t)
i
k12 [A1 ]0 h k23 t
k12 t
e
e
[A2 ](t) =
k12 k23
270

(6.101)
(6.102)

6.5 Methods of Solving Chemical Kinetic Rate Equations



[A3 ](t) = [A1 ]0 1 +

k23
k12
ek12 t
ek23 t
k12 k23
k12 k23

(6.103)

Fig. 6.11 shows the temporal behaviour of the concentrations of species A1 , A2 and
A3 for k12 = 1 s1 and k23 = 10 s1 . Fig. 6.11 shows the behaviour for k12 = 1 s1
and k23 = 100 s1 . Both figures correspond to A2 -species being very reactive if
compared to the reactivity of A1 -species. The concentration of A2 -intermediate
decreases rapidly with increasing k23 and in Eqs. (6.102) and (6.103) the terms
ek23 t can be neglected.

Fig. 6.11: Temporal behaviour of the species concentrations for reactions A1


A2 A3 assuming quasi steady state for A2 -species; k12 = 1 s1 ,
k23 = 100 s1 (compare with Fig. 6.4).

Eulers Explicit Method


In engineering practise, we are often interested in temporal behaviour of substrates
and products while intermediates are of lesser importance. Consequently we would
be happy with an integration procedure that would allow for calculating accurately
only the substrates and products while we might accept some inaccuracy on the
intermediates. Fig. 6.12 shows the numerical integration of Eqs. (6.52), (6.53)
and (6.54) using Eulers explicit method (see Example 6.2) for k12 = 1 s1 , k23 =
10 s1 . Since we are mainly interested in concentrations of A1 - and A3 -species, we
perform the integration using a 0.2 s time-step that on the basis of Example 6.2
is small enough for achieving both a good accuracy and stability. Fig. 6.12(top) shows however that while the A1 -species is calculated with good enough an
accuracy, the A3 -product concentration shows an unstable behaviour. The A2 intermediate concentration exhibits also oscillations. When the time-step of the
integration is decreased to 0.1 s, accurate and stable numerical solutions for all the

271

6 Mechanisms of Basic Combustion Reactions


species participating in the reactions are obtained as shown in Fig. 6.12-(bottom).
Consider now a case when k12 = 1 s1 and k23 = 100 s1 , so the reactivity of the
intermediate species (A2 ) is increased by a factor of ten. As shown in Fig. 6.11,
the A2 -species concentration is indeed very small and it can be certainly neglected
while considering A1 - and A3 -species. Fig. 6.13, top and bottom, show the results
of the numerical integration when the Euler method is used with an integration
time-step of 0.02 s and 0.01 s, respectively. Also in this case we had to reduce
the integration step to a value of 0.01 s to obtain accurate and stable results.
Integration using a time-step larger than 0.02 s would lead to instability.
So far we have considered two situations; in the first case k12 = 1 s1 and
k23 = 10 s1 (Fig. 6.12) while in the second case k12 = 1 s1 and k23 = 100 s1
(Fig. 6.13). In both cases the size of the integration time-step, needed for obtaining a stable numerical solution, has been determined by the second rate constant,
namely by k23 . In order to obtain a stable numerical solution using the Eulers
explicit scheme we used a time-step equal to k123 , as predicted by Eq. (F3) of Example 6.2. When the integration step is larger than k223 the Eulers explicit method
"explodes". We observe that the size of the time-step is determined by
the fastest reaction even though the contribution of the intermediate
component A2 to the concentration of the A1 - and A2 -species is negligible. This is a frustrating observation. In Table 5.1 there are reactions that differ
in their rate constants, and so in their rates, by as much as 1015 . Thus, if we were
to develop a general numerical procedure for solving a set of Eqs. (6.49) and the
solver would be based on the Euler explicit scheme, the size of the integration
time-step would be based on the fastest reaction even though they might be of
little importance in formation of the (final) stable products. The question is do
we really have to resolve all these small scales even if the intermediates are of no
interest to us?
Example 6.4
Develop a numerical scheme for integration of Eqs. (6.52), (6.53) and (6.54) using
Eulers explicit method. The initial conditions are [A1 ](0) = [A1 ]0 , [A2 ](0) = 0
and [A3 ](0) = 0.
Assumptions: None
In order to simplify the notation we introduce y1 , y2 and y3 in place of [A1 ], [A2 ]
and [A3 ] so Eqs. (6.52), (6.53) and (6.54) read:
dy1
= k12 y1
dt
dy2
= k12 y1 k23 y2
dt

272

(H1)
(H2)

6.5 Methods of Solving Chemical Kinetic Rate Equations


dy3
= k23 y2
dt

(H3)

and the initial conditions are y1 (0) = y10 , y2 (0) = 0 and y3 (0) = 0.
Here we recall that Eulers explicit method is based on (see Eq. (6.95)) the following relationship
yn+1 = yn + h f (yn , tn )
(H4)
where yn+1 and yn are values of the independent variable at time-steps tn+1 and
tn , respectively.
Therefore, for y1 we obtain:
y1,n+1 = y1,n + h (k12 y1,n ) = y1,n (1 k12 h)

(H5)

for y2
y2,n+1 = y2,n + h (k12 y1,n k23 y2,n ) = y2,n (1 k23 h) + h k12 y1,n (H6)
and finally for y3
y3,n+1 = y3,n + h k23 y2,n

(H7)

The formulae (H5), (H6) and (H7) constitute the Eulers explicit method for the
linear ordinary differential Eqs. (H1), (H2) and (H3). This scheme has been used
to produce numerical solutions shown in Fig. 6.12 and Fig. 6.13.
End of Example 6.4
Eulers Implicit Method
So far, the numerical solutions describing the temporal behaviour of the two
consecutive reactions (6.100) have been obtained using Eulers explicit scheme.
Fig. 6.14-top shows the solution obtained using Eulers implicit scheme (see Example 6.5) for k12 = 1 s1 and k23 = 10 s1 .
Note, that even for such a large time-step as 1 s, a stable solution has been obtained. Fig. 6.14-bottom shows that when k23 is increased to 100 s1 , Eulers
implicit scheme is stable for all the used time steps. Already in Example 6.3 we
have predicted that Eulers implicit scheme is unconditionally stable when one ordinary, linear differential equation is considered. Fig. 6.14 demonstrates the same
point for the reaction system (6.100) described by three Eqs. ((6.52)(6.54)). In
this way we have demonstrated superiority of the implicit scheme over the explicit one. It is generally true that for any system of linear ordinary differential
equations the implicit schemes are unconditionally stable and therefore they are
superior over explicit schemes.

273

6 Mechanisms of Basic Combustion Reactions


Example 6.5
Develop a numerical scheme for integration of Eqs. (6.52), (6.53) and (6.54) using
Eulers implicit method. The initial conditions are [A1 ](0) = [A1 ]0 , [A2 ](0) = 0
and [A3 ](0) = 0.
Assumptions: None In order to simplify the notation we introduce y1 , y2 and y3
in place of [A1 ], [A2 ] and [A3 ] so Eqs. (6.52), (6.53) and (6.54) read:
dy1
= k12 y1
dt
dy2
= k12 y1 k23 y2
dt
dy3
= k23 y2
dt

(I1)
(I2)
(I3)

and the initial conditions are y1 (0) = y10 , y2 (0) = 0 and y3 (0) = 0.
Here we recall that Eulers implicit method is based on (see Eq. (6.96)) the following relationship
yn+1 = yn + h f (yn+1 , tn+1 )
(I4)
where yn+1 and yn are values of the independent variable at time-steps tn+1 and
tn , respectively.
Therefore, for y1 we obtain
y1,n
(I5)
(1 + h k12 )
y2,n + h k12 y1,n+1
=
(1 + h k12 )
(I6)

y1,n+1 = y1,n + h (k12 y1,n+1 )

so y1,n+1 =

y2,n+1 = y2,n + h (k12 y1,n+1 k23 y2,n+1 )

so y2,n+1

y3,n+1 = y3,n + h k23 y2,n+1

(I7)

The formulae (I5), (I6) and (I7) constitute the Eulers implicit method for the
linear ordinary differential Eqs. (I1), (I2) and (I3). This scheme has been used to
produce numerical solutions shown in Fig. 6.14.
End of Example 6.5

274

6.5 Methods of Solving Chemical Kinetic Rate Equations

Fig. 6.12: Temporal behaviour of the species concentrations for reactions A1


A2 A3 , k12 = 1 s1 , k23 = 10 s1 . Numerical solutions obtained using Eulers Explicit Method.
Top h = 0.2 s integration time step,
Bottom h = 0.1 s time-step. (see Example 6.4).

275

6 Mechanisms of Basic Combustion Reactions

Fig. 6.13: Temporal behaviour of the species concentrations for reactions A1


A2 A3 , k12 = 1 s1 , k23 = 100 s1 . Numerical solutions obtained
using Eulers Explicit Method.
Top h = 0.02 s;
Bottom h = 0.01 s. (see Example 6.4).

276

6.5 Methods of Solving Chemical Kinetic Rate Equations

Fig. 6.14: Temporal behaviour of the species concentrations for reactions A1


A2 A3 ;
Top curves k12 = 1 s1 , k23 = 10 s1 ;
Bottom curves k12 = 1 s1 , k23 = 100 s1 Numerical solutions obtained
using Eulers Implicit Method.
(see Example 6.5).

277

6 Mechanisms of Basic Combustion Reactions


6.5.2.3 Set of Non-Linear Equations
In previous paragraphs we have considered solutions of either one
(Paragraph 6.5.2.1) or three (Paragraph 6.5.2.2) ordinary differential equations.
We have observed the superiority of the implicit numerical scheme over an explicit
scheme.
In order to simplify the notation we rewrite Eqs. (6.49) using a matrix notation:
y = F(y)

(6.104)

for which the initial conditions-vector y(0) is given.


The application of Eulers implicit scheme to (6.104) gives:
yn+1 = yn + h F(yn+1 )

(6.105)

In general this is some nasty set of nonlinear equations that has to be solved
iteratively at each time-step. Usually we can get away with linearization the
equation:
"
#

F
yn+1 = yn + h F(yn ) +
(yn+1 yn )
(6.106)
y yn
Here F
y is the matrix of the partial derivatives of the right-hand side, and so at
each time step we have to invert the matrix 1 h F
y to find yn+1 . As we have
previously noted, elementary reactions have greatly different reaction rates and
reaction time constants (scales). The ratio between the largest and the smallest
time constants of the reactions may be used to determine a degree of stiffness
of the reaction system. The F
y matrix is called Jacobian matrix of the system
under considerations. Eigenvalues and eigenvectors of the Jacobian matrix reveal
information about the time scales of the chemical reactions. These widely different
time scales present classical methods (such as Runge-Kutta method) with the
following difficulty; to ensure stability of the numerical solution, these methods
are restricted to using very short time steps that are determined by the fastest
reactions (the smallest time constant). However, the time for all chemical species
to reach near-equilibrium values is determined by the slowest reactions (the largest
time constant). Stiffness is an efficiency issue. If we were not concerned with how
much time a computation takes, we wouldnt be concerned about stiffness. Nonstiff methods can solve stiff problems. They just take a long time to do it. Thus,
for a fast and robust solver, special numerical procedures suitable for solving stiff
equations are required [27].

278

6.6 Summary

6.6 Summary
In this chapter an insight into oxidation mechanisms of carbon monoxide, hydrogen and methane has been given. One should become familiar with chain
initiation, chain branching and chain termination reactions. Using hydrogen as
an example, the role of O, H and OH radicals not only in initiating but also in
propagating combustion process has been underlined. Indeed, oxidation of carbon monoxide proceeds mainly though the "wet" route due to the reaction with
OH-radical. Oxidation of methane which is the most stable hydrocarbon, is initiated by the attack of O, H and OH radicals which produces reactive methyl
radical (CH3 ). Then, through several intermediate reactions, carbon monoxide is
produced which is further oxidised by OH to carbon dioxide. It has been stressed
that carbon monoxide is an important intermediate in methane oxidation and the
overall methane oxidation reaction
CH4 + 2 O2 CO2 + 2 H2 O
is an over-simplification. The following overall scheme
CH4 + O2 CO + H2 + H2 O
CO + 12 O2 CO2
H2 + 21 O2 H2 O
is better since it accounts for CO-intermediate.
Studying mechanisms of combustion reactions involves both formulation and solving chemical kinetic rate equations which form a set of ordinary non-linear differential equations of the type
d[Ai ]
= fi ([A1 ], [A2 ], . . . , [AN ]; k1 , k2 , . . . , kR )
dt

for i = 1, 2, . . . , N

with initial conditions


[Ai ](t = t0 ) = [Ai ]0
As a matter of fact in Chapter 5 we have already found analytical solutions to
simple ordinary differential equations that describe first, second and third order
reactions (see Section 5.2). In this chapter we have developed analytical solutions
to a system
k
k
A1 12
A2 23
A3

279

6 Mechanisms of Basic Combustion Reactions


which contains not only substrates (A1) and products (A3) but also intermediates
(A2). For first-order reactions we have developed analytical formulae for calculating the variation of the concentrations of species A1,A2 and A3 with time. Then,
we have analysed how the species concentrations vary with time for three cases:
k12 k23 , k12 k23 and k12 = k23 . We have observed that when k12 k23 we
could have simplified the solution (without compromising the accuracy too much)
by invoking a quasi-steady state assumption for A2-intermediates. The quasisteady state approximation has been applied to Zeldovich mechanism of nitric
oxide (NO) formation. The assumption of partial equilibrium for some reactions
has been invoked to estimate the concentration of O-radicals. Students should
understand both concepts important in kinetics; quasi-steady state assumption
and partial equilibrium.
A comprehensive mechanism of methane combustion may involve up to a hundred
species with perhaps as many as two hundred radical reactions. Thus, a general
solver for solving chemical kinetic rate equations must be numerical. Eulers
explicit and implicit methods are presented in the lecture however we do not
recommend them for any practical use. The methods are conceptually important
since all practical methods originate from this idea. Using these methods we have
introduced the notions of stability and accuracy which are two important features
of numerical schemes. It has been underlined that the major problem associated
with the simultaneous numerical integration of large sets of chemical kinetics rate
equations is that of stiffness. For an efficient and robust solver special numerical
schemes applicable to stiff equations are needed.

280

Bibliography
[1] E. Thne, U. Fahl. Energiewirtschaftliche Gesamtsituation. BWK: das Energie - Fachmagazin,Bd. 59 (2007), 4, S.34-50.
[2] H.Tsuji, A.K. Gupta, T. Hasegawa, M. Katsuki, K. Kishimoto and M. Morita.
High Temperature Air Combustion, CRS Press, 2003.
[3] Y.A. Cengel and M.A. Boles. Thermodynamics - An Engineering Approach,
4th Edition, McGraw-Hill, 2002.
[4] Ch. Ldecke and D. Ldecke. Thermodynamik, Springer, 2000.
[5] H.D. Baehr. Thermodynamik, 9th Edition, Springer, 1996.
[6] W. Leuckel. Technische Verbrennung, Vorlesungsskript, TU Karlsruhe, 1992.
[7] G.J. Van Wylen and R.E. Sonntag. Fundamentals of Classical Thermodynamics, EnglishSI version, 3rd Edition, John Wiley & Sons, New York, 1965.
[8] North American Combustion Handbook, North American Mfg. Co., Cleveland, Ohio, USA, Volume II, 3rd Edition, 1995.
[9] F.P. Incropera and D.P. DeWit, Fundamentals of Heat and Mass Transfer,
4th Edition, John Wiley & Sons, New York, 1996.
[10] R. Weber, Lecture Notes in Heat Transfer, 2nd Edition, Papierflieger,
Clausthal-Zellerfeld, 2004.
[11] R. Weber, R. Alt and M. Muster Vorlesungen zur Wrmebertragung, Teil
I: Grundlagen, 3rd Edition, Papierflieger, Clausthal-Zellerfeld, 2007.
[12] NIST-JANAF Thermochemical Tables. 4th Edition. Journal of Physical and
Chemical Reference Data. Monograph No. 9. Published by the American
Chemical Society and the American Institute of Physics for the National
Institute of Standards and Technology, 1998.
[13] P. Atkins. Physical Chemistry, 6th Edition, W.H. Freeman and Company,
New York, 1998.
[14] J.M. Smith, H.C. Van Ness, M.M. Abbott. Introduction to Chemical Engineering Thermodynamics, 6th Edition, McGraw-Hill, 1996.

281

Bibliography
[15] F. Reif. Fundamentals of Statistical and Thermal Physics. McGraw-Hill,
1965.
[16] J. M. Smith, H.C. Van Ness, M.M. Abbott, Introduction to Chemical Engineering Thermodynamics , 6th Edition, McGraw-Hill, 1996.
[17] M. Graetzel, P. Infelta The Bases of Chemical Thermodynamics , Universal
Publishers, Parkland, Florida, 2002.
[18] S.W. Benson. The Foundations of Chemical Kinetics, McGraw-Hill, New ork,
1960.
[19] W.C. Gardiner. Combustion Chemistry, Springer-Verlag, 1984.
[20] J. Warnatz, U. Maas, R.W. Dibble. Combustion. 2nd Edition, SpringerVerlag, 1999.
[21] D.L. Baulch, D.D. Drysdale, J. Duxbury and S.J. Grant, Evaluated Kinetic
Data for High Temperature Reactions , Vol.1: Homogeneous Gas Phase Reactions of the H2-O2 System , Butterworths, London, 1976.
[22] J. Chomiak. Combustion; A Study in Theory, Fact and Application , Abacus
Press, 1990.
[23] I. Glassman. Combustion, 2nd -Edition, Academic Press, 1987.
[24] R. K. Wilk, Low-Emission Combustion , Wydawnictwo Politechniki Slaskiej,
Gliwice, 2002.
[25] Miller I.A. and T. Bowmann. Mechanism and Modeling of Nitrogen Chemistry in Combustion. Progress in Energy and Combustion Science, 15, 287338, 1989.
[26] W.H. Press, B.P. Flannery, S.P. Teukolsky, W.T. Vetterling. Numerical
Recipes, Cambridge University Press, 1988.
[27] E. Hairer and G. Wanner, Solving Ordinary Differential Equations II. Stiff
and Differential-Algebraic Problems, 2nd Revised Edition, Springer-Verlag,
2002.
[28] C.K. Law, Combustion Physics, Cambridge University Press, 2006.

282

Gaussian Elimination
There are two well-established ways to solve linear equations. The first one, which
is a rather sophisticated way, introduces the idea of determinants. There is an
exact formula called Cramers rule, which gives the correct values of the unknowns
as a ratio of two M by M determinants. However for a reasonably human being
M = 3 is about the upper limit of patience.
The second method is called Gaussian elimination algorithm, or in short
Gaussian elimination, which we highly recommend. The algorithm is constantly used to solve large systems of linear equations. The idea of Gauss is
extremely simple: multiples of the first equation are subtracted from the other
equations, so as to remove the first unknown from those equations. It leaves a
smaller system of M 1 equations in M 1 unknowns. The process is repeated
until there is only one equation and one unknown, which can be solved explicitly. Then, we proceed backwards and find other unknowns in reverse order. The
example below will clarify the algorithm. We begin with a system of M = 3
equations:

2x + 3y + z = 5
x y = 1
4x + 2y + z = 3

(Ap1)

which can be written in matrix notation


5
x
2
3 1
1 1 0 y = 1
3
z
4 2 1

(Ap2)

The task is to find the unknowns x, y, z using Gauss elimination. We subtract


multiples of the first equation from the others, so as to eliminate x from the last
two equations. To this end, we
(a) subtract 0.5 times the first equation from the second equation,

283

Gaussian Elimination
(b) subtract (-2) times the first equation from the third equation.
The result is the following system of equations:

2x + 3y + z = 5
1
7
5
y z=
2
2
2
8y + 3z = 7
or


5
x
2
3
1
0 2.5 0.5 y = 3.5
7
z
0
8
3

(Ap3)

(Ap4)

The coefficient 2, which appears in front of the first unknown x, in the fist equation, is called the first pivot. Elimination is constantly dividing the pivot into
the numbers underneath it, to find out the right multipliers.
At the second stage of elimination; we ignore the first equation. The other two
equations contain only y and z unknowns and the same elimination process can be
applied to them. The pivot for this elimination stage is (5/2). A multiplicity of
the second equation will be subtracted from the remaining equations (in this case
only the third equation remains). We multiply the second equation by 8 25 = 16
5
and add it to the third one. In other words, we
(c) subtract (16/5) times the second equation from the third one:

2x + 3y + z = 5
1
7
5
y z=
2
2
2
7
21
z=
5
5

(Ap5)

or


5
x
2
3
1
0 2.5 0.5 y = 3.5
4.2
z
0
0
1.4

284

(Ap6)

The elimination process is now complete. Observe that the final matrix of
the elimination process contains zeros in all places located below the
diagonal. The order in which to solve the above system (Ap6) is obvious. The
last equation gives z = 3. Substituting into the second equation, we find y =
2. Then the first equation provides x = 1. The last process is called backsubstitution. Note that under completion of the elimination process we obtain
a triangular system which is easy to solve by back-substitution.
The elimination process, which in our case includes three steps, has produced
three pivots; 2 the first stage pivots, (5/2) - the second stage pivot and (7/5)
the third stage pivot. The reader will notice that pivots cannot be zero. We
need to divide by them.
An important number which we use is the rank of a matrix; say matrix A of
Eq. (Ap(2)). There are several definitions of the rank and they are equivalent.
Three most important definitions are quoted here for completeness:
the rank counts the number of independent rows in the matrix,
the rank is the number of pivots in the Gauss elimination process,
the rank is the number of non-zero rows in the final matrix of the Gauss elimination process.

285

Gaussian Elimination

286

Vocabulary
Table 6.9: Technical vocabulary. The vocabulary has been prepared for students of
TU Clausthal attending lectures on Combustion Technology, Heat Transfer,
Advanced Heat Transfer and High Temperature Processes.

English
Ability
absorb
absorptance
absorption capacity
absorptivity
acceleration
accordingly
accumulation
accuracy
accurate
achieve
acid
acid rain
adiabatic
adjacent
advanced
advantage
affect
ageing
agreement
aggregation state
aim (at)
air blower
air equivalence ratio
airflow system
air stream

Deutsch
Fhigkeit
absorbieren
Absorptionsgrad
Absorptionsvermgen
Absorptionsgrad, -vermgen
Beschleunigung
demgem, folglich
Ansammlung, Anhufung, Speicherung
Genauigkeit
genau
vollenden, erhalten, erreichen
Sure; sauer
saurer Regen
adiabat
angrenzend, benachbart
erweitert
Nutzen, Vorteil
beeinflussen, sich auswirken
Alterung
Abkommen, Zustimmung, Einigung
Aggregatzustand
zielen (auf), anstreben, beabsichtigen
Luftgeblse, Ventilator
Luft(verhltnis)zahl (z.B. 1,1 ... 1,15)
Belftungssystem
Luftstrom

287

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
aligned
alloy
alloy steel
aluminum
ambient
ambiguity
ammonia
amount
amount of heat "Q" (in J)
anodized
annular
annulus
aperture
application
appreciate
approach
appropriate
approximation
a priori determination
arbitrary
area
ash
assess
assessment
associated
assumption
atomization
attenuation
audit
augmented
available
average
averaged
avoid

Deutsch
ausgerichtet
Legierung
legierter Stahl
Aluminium
umgebend, UmgebungsZweideutigkeit, Mehrdeutigkeit
Ammoniak
Menge
Wrmemenge (in J)
eloxiert
ringfrmig
Ringspalt
ffnung
Anwendung, Einsatz
abschtzen
Annherung
passend, geeignet
Nherung
Vorherbestimmung
willkrlich, beliebig, allgemein
Flche
Asche
abschtzen, bewerten, ansetzen, festlegen
Wertung, Abschtzung
zugeordnet
Annahme
Zerstubung
Abschwchung, Verminderung
Prfung
vermehrt
verfgbar
Durchschnitt
gemittelt
vermeiden

Baffle

Umlenkung, Blende, Trennwand

288

Table 6.9: Technical vocabulary (continued)

English
balance
bank of tubes
bar
basis
basement
beam
bell mouth
benchmark
benefit
bent
blackbody
blast furnace
blast furnace gas
blower
body force
boiler
boiling
bond
booster
bottom
bottom line
boundary
boundary condition
boundary layer
bounding curve
branching
brass
brick
bright
bubble
buffer
bulb
bulk
bulk flow
bulk motion

Deutsch
Bilanz
Rohrbndel
Balken
Grundflche, Basis
Keller
Strahl
Schalltrichter
Bewertung, Mastab, Richtwert
Nutzen; ntzen
gebogen
Schwarzer Krper
Hochofen
Hochofengas, Gichtgas
Lfter, Geblse
Volumenkraft
Kessel
Sieden
Bindung (chemisch)
Zusatzantrieb
Boden
das Entscheidende
Grenze, Rand, Begrenzung
Randbedingung
Grenzschicht
Grenzkurve
Verzweigung
Messing
Ziegelstein
blank, glnzend
Blase
Zwischenspeicher, Dmpfer
Glhbirne
Volumen, Menge, Masse, Schttung,
Klumpen
Massenstrmung, Kollektivstrmung
(von Moleklen)
Massenbewegung

289

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
bulk temperature
bumpy
bundle
buoyancy
burner
burnout, degree of combustion, loss
on ignition (LOI)

Calculation

calibration curve
capacitance
capacity
capsule
carbon
carbon dioxide
carbonic acid
carbonization
cartoon representation
cavity
cement kiln
center of gravity
ceramic fiber
chain
chain reaction
channel
char
chemical engineering
chemistry
chimney
chlorine
chromium
circuit
circumference
clarify
classes

290

Deutsch
mittlere Temperatur, kalorische Temperatur
uneben
Bndel
Auftrieb
Brenner
Ausbrand

Berechnung
Eichkurve
Kapazitt, Belastbarkeit
Vermgen, Fhigkeit, Leistung
Kapsel
Kohlenstoff
Kohlendioxid
Kohlensure
Kohlenstoffanreicherung, Verkokung
Prinzipdarstellung
Holhlraum
Zementofen
Schwerpunkt
Keramikfaser
Ablauf, Kette
Kettenreaktion
Kanal
Holzkohle
Chemieingenieurwesen
Chemie
Schornstein
Chlor
Chrom
Kreis, Anlage, Schaltung, Schaltkreis,
Schaltbild
Umfang
klren
bungsstunden

Table 6.9: Technical vocabulary (continued)

English
clay
clay brick
closed system
cluster
coal
coarse
coat
coating
coflow, co-current flow, (parallel flow)
coke
coke-oven gas
cold blast
collapse
collide
collision
collision factor
colloidal
column
combustible
combustion
combustion chamber
combustion engineering
combustion product
compact
comparison
compass
composite
composition
compound
comprehensive
compression
compression work
comprise
concentration
concern

Deutsch
Lehm, Ton
Lehmziegel
geschlossenes System
Anhufung, Ballen, der Cluster
Kohle
grob
beschichten
Beschichtung, Belag
Gleichstrom
Koks
Koksofengas, Kokereigas
kaltgeblasen
Zusammenbruch, Einsturz; zusammenfallen
kollidieren, zusammenstoen
Zusammensto
Stofaktor
fein verteilt
Sule, Spalte
Brennbares, Brennmaterial
Verbrennung
Brennkammer, Brennraum
Verbrennungstechnik
Verbrennungsprodukt
dicht
Vergleich, Abgleich, Gegenberstellung
Zirkel
zusammengesetzt; Verbundwerkstoff
Zusammensetzung
Verbindung, Mischung
umfassend, ausgedehnt
Kompression, Verdichtung
Drucknderungsarbeit
zusammenfassen
Konzentration
sich befassen mit

291

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
concrete
condensation
condense
condenser
condition
conduction
conductivity
conductor (electric)
cone
confidence intervall
configuration factor
conjecture
conservation
constrain
constrained
constraint
consumption
contact surface
contaminated soils
continuity equation
contribute
contribution
control volume
convection
(convective) heat transfer coefficient
"h" (in W/m2 K)
convenient
conversion
cool
cooling
cooling rate
coordinate
copper
correlation
corrugated

292

Deutsch
Beton
Kondensation, Verdichtung
kondensieren, niederschlagen
Kondenser, Verflssiger
Bedingung
Leitung
Leitfhigkeit
Leiter (elektrischer)
Kegel
Vertrauensbereich
Einstrahlzahl,
Winkelverhltnis,
Formfaktor
Vermutung
Erhaltung
einschrnken
zwangslufig
Nebenbedingung, Einschrnkung
Verbrauch
Berhrungsflche
Altlasten
Kontinuittsgleichung
beitragen, beisteuern, liefern
Beitrag
Kontrollvolumen
Konvektion, "Mitfhrung" von Energie durch kleinste Strmungsteilchen
(konvektiver)Wrmebergangskoeffizient
"" (in W/m2 K)
passend, geeignet, bequem
Umformung, Umwandlung
khlen, abkhlen
Khlung
Khlrate
Koordinate
Kupfer
Wechselbeziehung
gewellt

Table 6.9: Technical vocabulary (continued)

English
corrugated iron
corrugation
counter balance
counterflow, countercurrent flow
coupled differential equations
courtesy of
cross flow
cross section
crucial
cumulative
cure
current
curvature
cyclic integral

Dashed line

data gathering
decay
decline
decrease
deduce
deep freezer
deficient
deficiency
defrost
degree of freedom
delivery
denominator
dense
density
departure
dependence
depletion
derivation
desalination

Deutsch
Wellblech
Welle, Riffelung
Gegengewicht; ausgleichen
Gegenstrom
gekoppelte Differentialgleichungen
freundlicherweise zur Verfgung
gestellt
Kreuzstrom
Querschnitt
ausschlaggebend
anwachsend
trocknen, aushrten
(elektrischer) Strom; laufend,
Krmmung
Umlaufintegral
gestrichelte Linie
Datenerfassung
Zerfall, Abnahme
abnehmen, neigen
abnehmen, vermindern
herleiten, folgern
Tiefkhltruhe
unzureichend
Mangel
auftauen
Freiheitsgrad
Zufhrung, Lieferung
Nenner
dicht
Dichte
Abweichung
Abhngigkeit
Abreicherung, Substanzverringerung
Herleitung, Ableitung
Entsalzung

293

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
design
determination
determine
development
deviate
device
devolatilization
dew point
diameter
diatomic
diffraction
diffuse
dihedral
dilute
dimensionless
direction
disadvantage
discovery
dispatch
displacement
disposal
dissipate
dissociation
distinct
distinguish
distribution
disturb
dot ("e dot")
double pipe heat exchanger
downstream
draft, draught
drag coefficient
drag force

294

Deutsch
Plan, Konstruktion, Entwurf, Auslegung konstruieren, auslegen
Bestimmung
bestimmen
Entwicklung
Abweichen
Apparat, Einrichtung, Vorrichtung,
Gert
Entgasung, Entfernung flchtiger Bestandteile
Taupunkt
Durchmesser
zweiatomig
Beugung, Ablenkung
diffus
zweiflchig
abschwchen, verdnnen
dimensionslos
Richtung
Nachteil
Entdeckung
Beseitigung, Bericht
Verschiebung
Beseitigung, Deponie, Entsorgung,
Verwendung, Verfgung
(Wrme) abfhren, zerstreuen
Dissoziation
verschieden
unterscheiden
Verteilung
stren
Punkt ("e ")
Doppelrohrwrmebertrager
stromabwrts, nachgeschaltet
Zug, Schornstein-, Kaminzug
Widerstandsbeiwert
(hydrodynamische) Widerstandskraft

Table 6.9: Technical vocabulary (continued)

English
drawback
driving force
droplet
dry
duct
due to
dummy variable
durability
duration
dust

Deutsch
Nachteil, Hindernis
Antriebskraft, treibende Kraft
Trpfchen
trocken
Kanal, Rhre, Schacht
beruhend auf
Integrationsvariable
Haltbarkeit
Zeitdauer
Staub

Edge

Ecke, Rand
Strahlungstemperatur
Wirkungsgrad, Wirtschaftlichkeit
Abflu, Ausflu
Anstrengung, Leistung
z. B.
Glhbirne
elektromagnetische Wellen
Emission, Aussendung
Energiestromdichte, spezifische Ausstrahlung
Emissionsgrad, -vermgen
aussenden
betonen
anwenden, einsetzen
einkapseln
Hohlraum, Umhllung, Einschlu,
abgeschlossener Raum
einschlieen, enthalten
endotherm, wrmeaufnehmend, -verbrauchend
Energie
Energiebilanz
Maschine
Technik, Technologie
erhhen

effective temperature
efficiency
efflux
effort
e.g. (=exempli gratia, for instance)
electrical bulb
electromagnetic waves
emission
emissive power
emissivity
emit
emphasise
employ
encapsulate
enclosure
encompass
endothermic
energy
energy balance
engine
engineering
enhance

295

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
enriched
ensure
enthalpy
enthalpy of formation
entrain
entrainment
entropy
environment
equalise
equally spaced
equation
equation of state
equilibrium
equipment
equivalence
estimate
estimation
ethene
ethyne
evacuate
evaluate
evaporation
exceed
excess
excess air ratio, excess air factor
excess temperature
exchange
exhaust
exhaust gas
exothermic
expand
expansion
experimenter
explanation
expose

296

Deutsch
angereichert
sicherstellen, ergeben
Enthalpie
Bildungsenthalpie
mitfhren
Mitfhren, Mitreien, Eintrag
Entropie
Umgebung
ausgleichen
quidistant
Gleichung
Zustandsgleichung
Gleichgewicht
Ausrstung, Einrichtung
Gleichwertigkeit
schtzen, berechnen, bestimmen
Bewertung
then
thin, Azetylen
evakuieren
berechnen
Verdampfung
bersteigen
berschu
Luftzahl, -berschu (z.B. 0,1 ...
0,15)
bertemperatur
Austausch
absaugen
Abgas
exotherm, wrmeabgebend
ausdehnen
Expansion, Ausdehnung, Entspannung
Experimentator
Erklrung
aussetzen

Table 6.9: Technical vocabulary (continued)

English
exposure
extended
extent
external flow

Deutsch
Bestrahlung
erweitert
Ausdehnung
berstrmt

Factory

Betrieb, Fabrik
Ausfall
vertraut
Fhn, Ventilator
Geblseleistung
Fehler, Defekt
Durchfhrbarkeit,
Wahrscheinlichkeit
Eigenschaft, Merkmal
Rohstoff, Ausgangsmaterial
Speisewasservorwrmer
Faser
Abbildung, Zeichen, Ziffer
Glhfaden, Wendel
Rippe, Khlrippe
gerippt
Schamottestein
einfgen, befestigen, anpassen
Ausgleichskurve
flockenartig
eben, flach
Mehl
Bewegung, Durchflu, Strmung;
strmen
Durchsatz, Volumenstrom
Schwankung
Strmungsgleichungen
Wirbelschicht, Fliebett
Flu, Stromdichte
konzentrieren auf
Folie

failure
familiar
fan
fan capacity
fault
feasibility
feature
feedstock
feedwater
fiber
figure
filament
fin
finned
fire clay brick
fit
fitting curve
flake-type
flat
flour
flow
flow rate
fluctuation
fluid flow equations
fluidised bed
flux
focus on
foil

297

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
forced convection
forced flow
forecast
fouling
fraction
fractional emissive power
free convection
free jet
free stream
frequency
friction
frustum, frustrum
fuel
furnace

Gage, gauge

gaseous
gas phase, gaseous phase
gathering
general
generate
generation
globe
goal
goggles
govern
governing equation
gradient
graph
graphite
gravitational force
gravity
gray (grey) body
grid
grip

298

Deutsch
erzwungene Konvektion, Zwangskonvektion
erzwungene Strmung
Vorhersage; vorausberechnen
Beschlagen, Verschmutzung
Bruchteil
Bruchteilfunktion
freie Konvektion
Freistrahl
freie, unbehinderte Strmung
Frequenz, Hufigkeit
Reibung
Stumpf, Kegelstumpf
Brennstoff
Ofen
Meinstrument
gasfrmig
Gasphase
Sammeln, Erfassung
allgemein
erzeugen
Erzeugung
Erdkugel
Zielpunkt, -setzung
Brille
beschreiben, beeinflussen, bestimmen, herrschen
prozebestimmende Gleichung
Gradient, Neigung, Steigung
Graphik
Graphit
Gravitationskraft, Schwerkraft
Schwerkraft
Grauer Krper
Gitter, Raster
Druckwalze

Table 6.9: Technical vocabulary (continued)

English
gross calorific value "GCV"
gypsum plaster

Deutsch
Brennwert, oberer Heizwert "ho "
Gipskartonplatte

Half-life time

Halbwertzeit
Halbwertsbreite
unbedenklich, unschdlich
schraffiert
Wrmekapazitt
Wrmeleitfhigkeit "" (in W/m K)
Wrmeleitungsgleichung

half width
harmless
hatched
heat capacity
heat conductivity "k" (in W/m K)
heat conduction (or: diffusion) equation
heat diffusion
heater
heat exchanger
(in J/s= W)
heat flow "Q"
heat flux "q"
(in W/m2 )
heat generation
heating rate
heating time
heating zone
heat loss
heat recovery
heat removal
heat sink
heat source
heat transfer
heat transfer coefficient "h" (in
W/m2 K)
(in J/s= W)
heat (transfer) rate "Q"
hemispherical
heterogeneous
high alloy steel
hollow cylinder
homogeneous
honeycomb

Wrmeausbreitung
Heizkrper
Wrmebertrager
(in J/s= W)
Wrmestrom "Q"
Wrmestromdichte "q"
(in W/m2 ),
Wrmeflu
Wrmequelle
Aufheizgeschwindigkeit
Erwrmungszeit
Heizzone, Wrmzone
Wrmeverlust
Wrmerckgewinnung
Wrmeabfuhr
Wrmesenke
Wrmequelle
Wrmebertragung
Wrmebergangskoeffizient " " (in
W/m2 K)
(in J/s= W)
Wrmestrom "Q"
hemisphrisch, Halbkugelheterogen, ungleichartig
hochlegierter Stahl
Hohlzylinder
homogen
Wabe

299

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
hot blast
hot blast stove
hot-wire anemometer
humidity
hydrocarbon
hydrochlorid acid
hydrodynamic
hydrodynamic boundary layer
hydrodynamic entrylength
hydrogen

Deutsch
heigeblasen
Winderhitzer
Hitzdrahtanemometer
Feuchtigkeit, Luftfeuchte
Kohlenwasserstoff
Salzsure
hydrodynamisch
Strmungsgrenzschicht
hydrodynamische Anlaufstrecke
Wasserstoff

Ideal gas

Ideales Gas
Entznden, Zndung
eintauchen
Aufprall, Auswirkung
Prall, Zusammensto
wichtig
Verbesserung, Fortschritt
Verunreinigung
weiglhend
(Strahlungs-, Licht-) Einfall, Einflu
einfallend, auftreffend
Veraschung, Veraschen
eintretend
bercksichtigen
wachsen, zunehmen
(temperatur-)unabhngig
hervorrufen, veranlassen
Saugzuggeblse
Ungleichung
Trgheitskraft
infinitesimal, unendlich klein
Ablenkung
Wendepunkt, Knickpunkt
Einflieen
Block
Anfangs-

ignition
immerse
impact
impingement
important
improvement
impurity
incandescent
incidence
incident
incineration
incoming
incorporate
increase
independent (of temperature)
induce
induced draft fan
inequality
inertial force
infinitesimal
inflection
inflection point
inflow
ingot
initial

300

Table 6.9: Technical vocabulary (continued)

English
initial condition
inleakage
inlet temperature
insignificant
instant (of time)
insulating plate
insulation
integer
intensity
intercept

inviscid
irradiation, irradiance
irreversible
isolated system
isolation
isotherm
isotropic
issue

Deutsch
Anfangsbedingung
Leckluft
Eintrittstemperatur
unwichtig, bedeutungslos
Zeitpunkt, Augenblick
Isolierplatte
Isolation
ganze Zahl
Intensitt, Strahldichte
Achsenabschnitt; auffangen, empfangen
Schnittstelle, Grenzschicht
dazwischenliegend
Innere Energie
durchstrmt
untereinander zusammenhngen
Wechselbeziehung
Schnittpunkt, Schnittmenge, Durchschnitt
reibungsfrei
Bestrahlung
nicht umkehrbar
isoliertes ("abgeschlossenes") System
Isolation
Isotherme; isotherm
isotrop, richtungsunabhngig
Ausgabe, Ergebnis, Thema, Problem

Jet

Strahl

Kernel of integral equation

Kern der Integralgleichung


Trockenofen, Brennofen
kinetische Energie

Laminar flow

laminare Strmung
Lampe

interface
intermediary
internal energy
internal flow
interrelate
interrelation
intersection

kiln
kinetic energy

lamp

301

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
landfill
lattice
layer
lead
leading edge
leakage
lean combustion
least square method
leave
lignin
limestone
linearization
line graph
liquid
localised
log mean temperature difference
"LMTD"
low alloy steel
lower calorific value "LCV"
lubricating oil
lumped
lumped capacitance method
lumped thermal capacity

Magnification
maintain
make up
malfunction
manufacturing
manure
margin
masonry wall
mass balance
mass flow
mass flow rate
mass fraction

302

Deutsch
Ablagerung, Deponie
Gitter
Schicht
Blei
Stirnseite, Anstrmkante
Leck, Undichtigkeit, Falschluft
berstchiometrische Verbrennung
Methode der kleinsten (Fehler-)
Quadrate
verlassen
Lignin, Holzstoff
Kalkstein
Linearisierung
Liniendiagramm
Flssigkeit; flssig
punktfrmig
mittlere logarithmische Temperaturdifferenz
niedriglegierter Stahl
(unterer) Heizwert "hu "
Schmierl
punktfrmig verteilt, konzentriert
"Newtonsches Abkhlungsgesetz"
thermische Blockkapazitt
Vergrerung
(aufrecht)erhalten
Zusammensetzung
Fehlfunktion, Strung
Erzeugung, Produktion
Dnger
Rand
gemauerte Wand
Massenbilanz
Massenstrom
Massendurchsatz
Massenbruch, Massenanteil

Table 6.9: Technical vocabulary (continued)

English
mat
matching
material property
matter
mean free path
measure
measurement
mechanism
merger
merit function
methane
mill
mineral fibre
minimum air requirement
moisture
molar mass
molar weight
mole fraction, molar fraction
molecularity
molten
moment of inertia
momentum
momentum balance equation
monatomic

Deutsch
matt
Abgleich, Anpassung
Materialeigenschaft
Materie
mittlere freie Weglnge
Ma; messen
Messung
Mechanismus
Verschmelzung
Gtefunktional, Merit-Funktion
Methan
Walze
Mineralfaser
minimaler Luftbedarf
Feuchte, Feuchtigkeit
molare Masse (in kg/kmol)
Molekulargewicht
Molenbruch, Molanteil
Molekularitt
geschmolzen
Trgheitsmoment
Impuls
Impulsbilanzgleichung
einatomig

Natural convection

natrliche (freie) Konvektion


Erdgas
notwendig
vernachlssigen
vernachlssigbar
netto
Stickstoff
Edelgas
senkrecht (zu)
Norm(al)bedingung, Standardbedingung
Haftbedingung

natural gas
necessary
neglect
negligible
net
nitrogen
noble gas
normal (to)
normal condition
no-slip condition

303

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
nozzle
nucleate boiling
nucleation site
numerator
numerically

Deutsch
Dse
Blasensieden
Keimstelle
Zhler
numerisch, in Zahlen

Objective

Ziel
beobachtbar
erhalten
offensichtlich
vorkommen, stattfinden, sich ereignen, auftreten
Knick, rumliche Verschiebung
Anfang, Umschlag zu
Undurchsichtigkeit
opak, undurchsichtig
offenes System
Betriebsbedingung
optische Dicke
Grenordnung
Erz
ausgerichtet
entstehen, ausgehen von
Oszillator, Schwingungserzeuger
Ergebnis
austretend
Abfluss
Ausgangsspannung
Ofen
geteilt durch: z geteilt durch n ( nz )
Gesamtwrmeverlust
Wrmedurchgangskoeffizient "k" (in
W/m2 K)
berdruck
Oxidierung
Oxidationsmittel (Sauerstoff, Luft)
Verbrennung mit reinem Sauerstoff

observable
obtain
obvious
occur

offset
onset
opacity
opaque
open system
operating condition
optical thickness
order of magnitude
ore
oriented
originate
oscillator
outcome
outcoming
outflow
output voltage
oven
over: z over n
overall heat loss
overall heat transfer coefficient "U"
(in W/m2 K)
over-pressure
oxidation
oxidiser
oxy fuel combustion

304

Table 6.9: Technical vocabulary (continued)

English
oxygen

Deutsch
Sauerstoff

Parallelepiped

Quader
Ballen, Bndel
Partialdruck
Muster
Pellet
Pendel
vorletzte
prozentualer Anteil
Ideales Gas
Ideales-Gas-Gesetz,
Zustandsgleichung fr Ideale Gase
leisten, erfllen, durchfhren
Verhalten, Durchfhrung, Leistung
Umfang
Permeabilitt, Durchlssigkeit
Normale, Senkrechte; senkrecht (zu)
gehren zu
sachdienlich, passend
Phasenwechsel
Roheisen
Bolzen
Rohr
Rohrleitung
Kolben
Neigung, Steigung, Abstand
eben
ebene Wand
Anlage
Platte
Plattenwrmebertrager
Rauchfahne, Abwasserfahne
betonen, hervorheben, auf etwas hinweisen
Schadstoff; verschmutzt
Verunreinigung

parcel
partial pressure
pattern
pellet
pendulum
penultimate
percentage
perfect gas
perfect gas equation of state
perform
performance
perimeter
permeability
perpendicular
pertain to
pertinent
phase change
pig iron
pin
pipe
pipeline
piston
pitch
planar, plane
plane wall
plant
plate
plate heat exchanger
plume
point out
pollutant
pollution

305

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
polynomial
polystyrene
pool boiling
porous
potential energy
power
power engineering
power generation
power plant
power station
precisely
preferential
preheat
preheating zone
preparation
prerequisite
pressure
pressure drop
prevail
previous
primitive
printed board
product
profile
propagation
property
proposal
propulsion
protective atmosphere
provide
provide for
pulverised coal
pump
purchasing
purify
purpose

306

Deutsch
Polynom
Polystyrol
Behltersieden
pors
potentielle Energie
Leistung
Energietechnik
Stromerzeugung
Kraftwerk, Triebwerk
Kraftwerk, Elektrizittswerk
genau
bevorzugt
vorwrmen
Vorwrmzone
Bereitstellung, Aufbereitung
Vorbedingung, Voraussetzung
Druck
Druckverlust, -abfall, -absenkung
vorherrschen
vorhergehend
Stammfunktion, (unbestimmtes) Integral
Leiterplatine
Produkt, Erzeugnis
Profil
Wachstum
Eigenschaft
Ansatz, Vorschlag
Antrieb
Schutzatmosphre, Schutzgas
beschaffen, zufhren, liefern
gewhrleisten
pulverisierte Kohle, Kohlenstaub
Pumpe
Beschaffung, Erwerb
reinigen
Zweck

Table 6.9: Technical vocabulary (continued)

English

Deutsch

Quantum number

Quantenzahl
Wirkungsquantum
abschrecken
ruhend
angeben, zitieren

Radiance

Strahlstrke
Strahlung
Strahler, Heizkrper
Helligkeit
regellose, zufllige Bewegung
Bereich
schnell
Strom, Menge je Zeiteinheit, pro Zeit
Geschwindigkeitskoeffizient
Geschwindigkeitskonstante
Transportansatz
Geschwindigkeitsgesetz
(in J/s=W)
Wrmestrom "Q"
Umsetzungsgeschwindigkeit
Rohmaterial
Reaktant, Reaktionspartner, Edukt
Reaktion
Reaktionsenthalpie
Reagens
Heck
umstellen, umordnen
vernnftig
Argumentation, Schlufolgerung
empfohlen, vorgeschlagen
in Einklang bringen
beziehen auf
Raffinierung, Verbesserung
reflektieren, spiegeln
Reflexionsgrad, -vermgen

quantum of action
quench
quiescent
quote

radiation
radiator
radiosity
random motion
range
rapid
rate
rate coefficient
rate constant
rate equation
rate law
(in J/s=W)
rate of heat transfer "Q"
rate of consumption
raw material
reactant
reaction
reaction enthalpy
reagent
rear
rearrange
reasonable
reasoning
recommended
reconcile
refer to
refinement
reflect
reflectivity

307

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
refractory
refractory brick
refractive index
refrigerant
refuse derived fuel (RDF)
regenerative burner
regenerator
region
reheat
reheating furnace
relation
relationship
release
reliability
removal
replacement
require
rescale
resistance
responsible
resource
result
retardation
reversible
rich combustion
right circular cone
rocket booster
rod
rolling mill
roughly
roughness
roundoff error, rounding error
rubbish

Sample

saturated

308

Deutsch
feuerfestes Material; feuerfest
feuerfester Stein, Schamottestein
Brechungskoeffizient,
Brechungsindex, Brechzahl
Kltemittel, Khlflssigkeit
Sekundrbrennstoff aus Abfall
Regenerativbrenner
Regenerator
Bereich
wiedererwrmen
Nachwrmofen
Beziehung
Verhltnis, Beziehung
freisetzen, abgeben
Zuverlssigkeit, Sicherheit
Abtransport
Ersetzung
fordern, anfordern
neuen Mastab festsetzen
Widerstand
verantwortlich
Ressource, Quelle, Hilfsmittel
Resultat, Ergebnis
Verzgerung
umkehrbar
unterstchiometrische Verbrennung
gerader Kreiszylinder
Zusatztriebwerk
Stab
Walzwerk
grob gesagt
Rauhigkeit
Rundungsfehler
Abfall, Schutt
Probe, Exemplar, Muster
gesttigt

Table 6.9: Technical vocabulary (continued)

English
saturation temperature
saving
scale
scatter
scope
screen
sealing
separation
series
series expansion
sewage sludge
shape
shape factor
shear stress
shell
shell-and-tube-heat exchanger
shield
shrouding
side wall
silicium carbide
silver
simultaneous
sink
sinter
site
sketch
skilfully
slab
slag
slide
slope
sludge
slurry
smelt
smooth

Deutsch
Sttigungstemperatur
Einsparung
Mastab
streuen
Bereich
Bildschirm, Blende, Sieb; durchsieben, berprfen
Dichtung
Ablsung
Reihe, Folge
Reihenentwicklung
Klrschlamm
Form, Gestalt
Formfaktor, Einstrahlzahl, Winkelverhltnis
Schubspannung, Scherspannung
Hlle, Mantelflche, Ofenwanne
Rohrbndelwrmebertrager
Schirm
Abdeckung
Seitenwand
Siliziumkarbid
Silber
gleichzeitig
Senke
Sinter
Ort, Lage, Gebiet
Skizze
geschickt
Platte
Schlacke
Dia, Folie
Steigung, Geflle
Klrschlamm
Schlamm, wssrige Masse
schmelzen
flach, gleichmig; gltten

309

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
soak
soaking zone
soften
soils (contaminated soils)
solid
solid angle
solidification
solid material
solid state
solution
soot
sought
source
space shuttle
sparse
specific heat capacity (in J/kg K)
species
sphere
sponge
spring
square (r square = r2 )
stability
stable
stack
staged air
staggered
stagnant air
stagnation point
stainless steel
standard enthalpy of formation
state equation
state variable
static pressure
stationary
steady state

310

Deutsch
ausgleichen
Ausgleichszone
weich machen
Bden (Altlasten)
fest
Raumwinkel
Erstarrung
Feststoff, Schleudergut
Festkrper
Lsung
Ru
gefragt, gesucht
Quelle
Raumfhre
sprlich, wenig
spezifische
Wrmekapazitt
(in
J/kg K)
Spezies, Art, Substanz, Stoff
Kugel
Schwamm
Feder
Quadrat
Stabilitt
stabil, bestndig
Speicher, Stapel
gestufte Luft
versetzt, gestaffelt
ruhende Luft
Staupunkt
rostfreier Stahl
Standardbildungsenthalpie
Zustandsgleichung
Zustandsvariable, -gre
statischer Druck
stationr, ortsfest
stationrer Zustand; stationr, zeitunabhngig

Table 6.9: Technical vocabulary (continued)

English
steam
steel
steelmaking
steelwork
steep
stoichiometry
storage
store
straw
stream
stress-strain relation
subscript
substance
substantial
substrate
subtend
sufficient
suggest
suitable
sulphur, sulfur
superficial
supply
support
surface
surrounding
suspension
swirl
symmetry axis
symmetric plane

Deutsch
Wasserdampf
Stahl
Stahlerzeugung
Stahlwerk
steil
Stchiometrie
Speicherung
speichern
Stroh
Strom
Spannungs-Dehnungs-Beziehung
Index
Stoff
wesentlich
Substrat
gegenberliegen
hinreichend
vorschlagen
brauchbar, geeignet
Schwefel
oberflchlich
Versorgung; zufhren
Halterung, Fassung, Lager
Oberflche
Umgebung
Aufschwemmung
Wirbel
Symmetrieachse
Symmetrieebene

Tackle

bewltigen, in Angriff nehmen


Teerpappe
Temperatur
Temperaturdifferenz
Temperaturverteilung
Temperaturabfall, -abnahme
Tetraeder

tar paper
temperature
temperature difference
temperature distribution
temperature drop
tetrahedron

311

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
thermal
thermal boundary layer
thermal capacity
thermal
conductivity
"k"
(in
W/m K)
thermal diffusivity " "(in m2 /s)
thermal expansion coefficient
thermal radiation
thickness
thin walled
throttling
time constant
tin
torque
total thermal input "TTI"
total thermal resistance "TTR" (in
W/K)
to the power: T to the power 4
trailing edge
transfer
transient
transition
transmission
transmissivity
transmit
transparency
transparent
transverse
transverse pitch
trap
treat as
treatment
trickle

312

Deutsch
thermisch, WrmeTemperaturgrenzschicht
Wrmekapazitt
Wrmeleitfhigkeit, -koeffizient ""
(in W/m K)
Temperaturleitfhigkeit, -koeffizient
"" (in m2 /s)
Wrmeausdehnungskoeffizient
Temperatur-, Wrmestrahlung
Dicke
dnnwandig
Drosselung
Zeitkonstante
Zinn
Moment, Drehmoment
insgesamt zugefhrte thermische Energie
Gesamtwrmewiderstand
hoch: T hoch 4 (T4 )
Hinterkante
bertragen
zeitlich vernderlich
bergang
(Wrme-)Durchgang
(Durchla-),
Transmissionsgrad,
-vermgen
transmittieren, durchlassen
Transparentbild, Diapositiv, Folie
durchsichtig
quer
Querteilung
auffangen, einschlieen
behandeln als
Aufbereitung, Bearbeitung, Verfahren
trpfeln, rieseln

Table 6.9: Technical vocabulary (continued)

English
trigger
tripping
trip wire
tube
tube bank
tubular heat exchanger
tundish
tungsten
turbulent flow

Deutsch
auslsen
Auslsung
Reileine
Rohr
Rohrbndel
Rohrbndel-Wrmebertrager
Giewanne
Wolfram
turbulente Strmung

Unaltered

unverndert
nichtbrennbare Anteile
Bildunterschrift; unterstreichen, betonen
gleichverteilt
Vereinigung (von Mengen)
einheitlich, eindeutig, einzig
Einheit
Einheitsvektor
sperrig, unhandlich
stromaufwrts, vorgeschaltet

unburnts
underline

uniform
union (of sets)
uniquely
unit
unit vector
unwildy
upstream

Valid

validate
valve
vapor, vapour
vaporization
velocity
vibration
vicinity
view factor
viscous
viscous force
voidage

gltig
besttigen, fr gltig erklren
Ventil
Dampf
Verdampfung
Geschwindigkeit
Schwingung
Umgebung
Einstrahlzahl,
Winkelverhltnis,
Formfaktor
viskos, zh, dickflssig
Zhigkeitskraft
Leerraum, Porositt, relatives Porenvolumen

313

Vocabulary
Table 6.9: Technical vocabulary (continued)

English
volatile
voltage
volumetric flow rate
volumetric heat capacity: "cp "
voluntary
vortex

Wake
walking beam
waste
waste water
water
wavelength
wave number
wedge
wet
width
wire
wooden lagging
work

Deutsch
Flchtige; flchtig
(elektrische) Spannung
Volumenstrom
volumenbezogene Wrmekapazitt:
"cp "
willkrlich
Wirbel
Nachlauf, Kielwasser, Spur, Wirbelschleppe
Hubbalken
Abfall
Abwasser
Wasser
Wellenlnge
Wellenzahl
Keil
benetzen; feucht
Breite
Draht
Holzverschalung
Arbeit

X-axis
X-rays

x-Achse, Abszisse
Rntgenstrahlen

Y-axis

y-Achse, Ordinate

Zinc

Zink

314

Acronyms
NIST

National Institute of Standards and Technology (USA)

JANAF

Joint Army-Navy-Air Force

GCV

Gross Calorific Value (oberer Heizwert oder Brennwert)

LCV

Lower Calorific Value (unterer Heizwert oder Heizwert)

TTI

Total Thermal Input

NIV

Number of Independent extensive Variables

315

Acronyms

316

Nomenclature
Latin Letters
C

specific heat capacity

concentration

specific heat capacity

J
kmol K
mol
m3
J
kg K
J
mol K
J
kg K
J
mol K
J
kg K

Cp

specific heat at constant pressure

cp

specific heat at constant pressure

Cv

specific heat at constant volume

cv

specific heat at constant volume

d
E

inexact differential
energy rate

energy amount

Ea

activation energy

Helmholtz free energy or free internal en-

J
mol

ergy
g

specific free enthalpy

Gibbs thermodynamic potential or free

J
mol

enthalpy
J
kg

Gibbs specific free enthalpy

g
H

earths gravity

9.806 65 m/s2

enthalpy rate

specific enthalpy

J
mol

enthalpy

specific enthalpy

J
kg

Irreversibility

317

Nomenclature
moment of inertia

standard equilibrium constant


rate constant for a

k
k, kf , kb
L

ldry

kg m2

1st

order reaction

rate constants for a reaction

1
s

depends on reaction order

mechanical power

work

specific work

air,min

minimum amount of air required per unit

J
kg
kg air
kmol fuel

of fuel
ldry

air

mass rate

mass

Mi

molar mass of species i

Mmean

average molar weight of a mixture

kg air
kmol fuel
kg
s

kg
g
mol
g
mol
kmol
s
6.023 1023 mol1

n i

molar flow rate

NA

Avogadro constant

Loss of power

amount of substance

mol

ni

amount of substance i

mol

pressure

Pa

pi

partial pressure of component i

Pa

saturation pressure

Pa

rate of heat removal

heat

specific heat

J
kg

universal gas constant

enthalpy of condensation of water

specific entropy

entropy

specific entropy

temperature

time

psat
Q

318

amount of air supplied per unit of fuel

8314.3 J/(kmol K)
43 940 kmol kJ
of H

2O

J
mol K
J
K
J
kg K

Nomenclature
umixture

specific internal energy

internal energy

specific internal energy

volume

specific volume

Vdry,min

minimum dry amount of combustion

kJ
kmol

J
J
kg
m3
m3
kg
kmol dry products
kmol fuel

products
Vdry
Vwet,min

dry amount of combustion products


minimum wet amount of combustion

kmol dry products


kmol fuel
kmol wet products
kmol fuel

products
Vwet

wet amount of combustion products

kmol wet products


kmol fuel
m
s

translational velocity

wi

mass fraction of species i

xi

mole fraction of species i

height of center of gravity

Greek Letters
R GpT

Gibbs reaction enthalpy

kJ
mol

relative humidity

excess air ratio

chemical potential

chemical potential

rate of entropy change

fuel equivalence ratio

sum of the entropy change in process

specific humidity

density

stoichiometric factor

rotational velocity

s1

extent of reaction

mol

1
J
kg
J
mol
J
Ks

1
J
K
kg water vapour
kg dry air
kg
m3

319

Nomenclature

Superscripts
0

denotes standard pressure of 1 bar

Subscripts
0

starting point, initial condition

cr

critical

constant

eq

equilibrium

formation

friction

generated

species

species

normal or standard condition

total

trs

320

transition

Nomenclature

321

Das könnte Ihnen auch gefallen